Introduction To Optimization For
Introduction To Optimization For
By
Louis Theodore and Kelly Behan
CRC Press
Taylor & Francis Group
6000 Broken Sound Parkway NW, Suite 300
Boca Raton, FL 33487-2742
This book contains information obtained from authentic and highly regarded sources. Reasonable efforts have been
made to publish reliable data and information, but the author and publisher cannot assume responsibility for the
validity of all materials or the consequences of their use. The authors and publishers have attempted to trace the copy-
right holders of all material reproduced in this publication and apologize to copyright holders if permission to publish
in this form has not been obtained. If any copyright material has not been acknowledged please write and let us know
so we may rectify in any future reprint.
Except as permitted under U.S. Copyright Law, no part of this book may be reprinted, reproduced, transmitted, or
utilized in any form by any electronic, mechanical, or other means, now known or hereafter invented, including pho-
tocopying, microfilming, and recording, or in any information storage or retrieval system, without written permission
from the publishers.
For permission to photocopy or use material electronically from this work, please access www.copyright.com (http://
www.copyright.com/) or contact the Copyright Clearance Center, Inc. (CCC), 222 Rosewood Drive, Danvers, MA
01923, 978-750-8400. CCC is a not-for-profit organization that provides licenses and registration for a variety of users.
For organizations that have been granted a photocopy license by the CCC, a separate system of payment has been
arranged.
Trademark Notice: Product or corporate names may be trademarks or registered trademarks, and are used only for
identification and explanation without intent to infringe.
Donald J. Trump
Preface............................................................................................................................................ xiii
About the Book............................................................................................................................ xvii
Authors.......................................................................................................................................... xix
1 Optimization Overview.........................................................................................................3
1.1 History of Optimization...............................................................................................4
1.2 The Computer Age.........................................................................................................6
1.3 The Scope of Optimization...........................................................................................7
1.4 Conventional/Established Optimization Procedures..............................................7
1.5 Contemporary Optimization: Linear Programming................................................ 8
References................................................................................................................................ 11
2 Mathematical Operations.................................................................................................... 13
2.1 The Quadratic Equation.............................................................................................. 15
2.2 Interpolation and Extrapolation................................................................................ 17
2.3 Significant Figures and Approximate Numbers..................................................... 20
2.4 Errors.............................................................................................................................22
2.5 Differentiation.............................................................................................................. 23
2.6 Numerical Integration................................................................................................. 26
2.6.1 Trapezoidal Rule............................................................................................. 27
2.6.2 Simpson’s Rule................................................................................................ 28
2.7 Simultaneous Linear Algebraic Equations............................................................... 29
2.7.1 Gauss–Jordan Reduction...............................................................................30
2.7.2 Gauss Elimination..........................................................................................30
2.7.3 Gauss–Seidel Approach................................................................................. 31
2.8 Nonlinear Algebraic Equations................................................................................. 31
2.9 Ordinary Differential Equations............................................................................... 32
2.10 Partial Differential Equations.................................................................................... 37
2.10.1 Parabolic PDE.................................................................................................. 37
2.10.2 Elliptical PDE................................................................................................... 39
References................................................................................................................................ 40
3 Perturbation Techniques...................................................................................................... 41
3.1 One Independent Variable..........................................................................................42
3.2 Two Independent Variables........................................................................................ 45
3.3 Three Independent Variables..................................................................................... 48
References................................................................................................................................ 49
4 Search Methods..................................................................................................................... 51
4.1 Interval Halving Method............................................................................................ 53
4.2 The Bisection Method.................................................................................................. 56
vii
viii Contents
5 Graphical Approaches..........................................................................................................65
5.1 Rectangular Coordinates............................................................................................ 66
5.1.1 Quadrants........................................................................................................ 67
5.1.2 Linear versus Non-Linear.............................................................................. 67
5.1.3 Intercepts.......................................................................................................... 68
5.2 Logarithmic-Logarithmic (Log-Log) Coordinates.................................................. 68
5.3 Semi-Logarithmic (Semi-Log) Coordinates............................................................. 73
5.3.1 Plotting Straight Lines on Semi-Logarithmic Paper.................................. 73
5.3.2 Other Graphical Coordinates........................................................................ 76
5.4 Methods of Plotting Data............................................................................................ 76
5.5 Applications..................................................................................................................77
References................................................................................................................................ 81
6 Analytical Methods...............................................................................................................83
6.1 Breakeven Considerations..........................................................................................83
6.2 One Independent Variable..........................................................................................85
6.3 General Analytical Formulation of the Optimum.................................................. 88
6.4 Two Independent Variables........................................................................................ 89
6.5 Three Independent Variables..................................................................................... 93
References................................................................................................................................ 94
7 Linear Programming............................................................................................................. 95
7.1 Definitions..................................................................................................................... 96
7.2 Basic Concepts of Optimization................................................................................. 97
7.3 Applied Mathematical Concepts in Linear Programming.................................... 99
7.4 Applied Engineering Concepts in Linear Programming.................................... 102
7.5 Other Real-World Applied Concepts in Linear Programming........................... 105
References.............................................................................................................................. 107
It is no secret that in recent years, the number of people entering the engineering field has
increased. Most are beginning college students and some had earlier chosen a non-tech-
nical major/career path. A large number of these individuals are today seeking degrees in
environmental and/or chemical engineering. These prospective students will require an
understanding and appreciation of the numerous mathematical and optimization meth-
ods that are routinely employed in practice. This technical stepping stone to a successful
career is rarely provided at institutions that award engineering degrees. This introductory
text on optimization attempts to supplement existing curricula with a sorely needed tool
to eliminate this void.
More on the need for this book. The question often arises as to the educational back-
ground required for engineering students to possess meaningful analysis capabilities since
technology has changed the emphasis that is placed on certain technical subjects. Before
computer usage became popular, instruction in engineering analysis was (and still is in
many places) restricted to simple systems, and most of the effort was devoted to solving a
few elementary equations that were derived. These cases were mostly of academic inter-
est, and because of their simplicity, were of little practical value. To this end, a considerable
amount of time is now required to acquire skills in mathematics, especially in numerical
methods, statistics, and optimization. In fact, most engineers are given courses in classi-
cal mathematics, but experience shows that very little of this knowledge is retained after
graduation for the simple reason that these mathematical methods are not adequate for
solving most systems of equations encountered in industry. In addition, advanced math-
ematic skills are either not provided in courses or are forgotten through sheer disuse since
they are not related to engineering practice.
The material in this book was prepared primarily for environmental and chemical
engineering students and, to a lesser extent, for engineering professionals who wish to
obtain a better understanding of the various optimization methods that can be employed
in solving technical problems. In presenting the text material, the authors have stressed
the pragmatic approach in the application of mathematical tools to later assist the reader in
grasping the role of optimization in engineering problem-solving situations.
As noted previously, this book serves two purposes. It may be used as a textbook for
beginning environmental and chemical students or as a “reference” book for practicing
engineers involved with optimization applications. For both audiences, it is assumed that
the reader has already taken basic courses in physics and chemistry, and should have a
minimum background in mathematics through elementary calculus. The authors’ ulti-
mate aim is to offer the reader the fundamentals of several optimization methods with
accompanying practical engineering applications. The reader is encouraged through the
references to continue his or her own development beyond the scope of the presented
material.
As is usually the case in preparing any text, questions of what to include and what to
omit have been particularly difficult. As noted above, the material in this book attempts
to address optimization calculations common to both the environmental and chemi-
cal engineering disciplines. The book provides the reader with numerous solved illus-
trative examples so that the interrelationship between both theory and application
is emphasized in nearly all of the chapters. One key feature of this book is that the
xiii
xiv Preface
solutions to the problems are presented in a stand-alone manner. Throughout the book,
the illustrative examples are laid out in such a way as to develop the reader’s technical
understanding of optimization, with more difficult examples located at or near the end
of each chapter.
The book is divided into four (IV) parts (see also Table of Contents):
It should be noted that Part IV was an afterthought. After completing the manuscript,
the authors felt that it lacked some real-world applications drawn from both the envi-
ronmental and chemical engineering industries. As such, three chapters were added to
this work—one concerned with environmental engineering, one concerned with chemical
engineering, and one concerned with term projects. Chapter 18, titled Select Environmental
Engineering Applications, is an extension of Part II and contains five sections—air man-
agement, water management, solid waste management, health risk assessment, and haz-
ard risk assessment. Chapter 19, titled Select Chemical Engineering Applications, is on
the other hand an extension of Part III and also contains five sections—fluid flow, chemi-
cal reactors, mass transfer applications, heat transfer applications, and plant design.
Chapter 20 contains four sections that address more advanced optimization material.
Most chapters contain a short introduction to the topic in question, which is followed by
developmental material, which, in turn, is followed by several illustrative examples. Since
this book offers material not only to individuals with limited technical background but
also to those with extensive engineering industrial experience, this book can be used as a
text in either a general introductory optimization course or (perhaps) as a training tool in
industry for challenged engineering professionals.
Hopefully, the text is simple, clear, to the point, and imparts a basic understanding of the
theory and application of some of the important optimization methods employed in prac-
tice. It should also assist the reader in helping master the difficult task of explaining what
was once a very complicated subject matter in a way that is easily understood. The authors
feel that this distinguishes this text from the numerous others in this field.
Is should also be noted that the authors have long advocated that basic science
courses—particularly those concerned with mathematics—should be taught to engi-
neers by an engineer; and the books adopted for use in these courses should also be
written by an engineer. For example, a mathematician may lecture on differentiation—
say dy /dx —not realizing that in a real-world application involving an estuary, y could
refer to concentration while x could refer to time. The readers of this book will not
encounter this problem.
The reader should also note that parts of the material in the book were drawn from one
of the author’s notes of yesteryear. In a few instances, the original source was not available
for referencing purposes. Any oversight will be corrected in a later printing/edition.
The authors wish to express appreciation to those who have contributed suggestions for
material covered in this book. Their comments have been very helpful in the selection and
presentation of the subject matter. Special appreciation is extended to Drs. Walter Matystik
and Francesco Ricci.
Preface xv
Finally, the authors are especially interested in learning the opinions of those who read
this book concerning its utility and serviceability in meeting the needs for which it was
written. Corrections, improvements, and suggestions will be considered for inclusion in
any later editions.
Lou Theodore
East Williston, NY
Kelly Behan
Mineola, NY
About the Book
As noted in both the Preface and Table of Contents, this optimization book is divided into
four parts:
xvii
xviii About the Book
1. R. Perry and D. Green, Perry’s Chemical Engineers’ Handbook , 8th edition, McGraw-
Hill, New York City, NY, 2008.
2. L. Theodore, Chemical Engineering: The Essential Reference , McGraw-Hill, New York
City, NY, 2014.
3. J. Reynolds, J. Jeris, and L. Theodore, Handbook of Chemical and Environmental
Engineering Calculations , John Wiley & Sons, Hoboken, NJ, 2004.
4. L. Theodore, R. Dupont, and K. Ganesan, Unit Operations in Environmental
Engineering , Wiley‑Scrivener, Salem, MA, 2018.
5. M. K. Theodore and L. Theodore, Introduction to Environmental Management , CRC
Press/Taylor & Francis, Boca Raton, FL, 2000.
6. G. Burke, B. Singh, and L. Theodore, Handbook of Environmental Management and
Technology , 2nd edition, John Wiley & Sons, Hoboken, NJ, 2000.
The beginning of each chapter (not including Part IV) contains additional references that
are topic specific. The reader, particularly the beginning student, should consider review-
ing (and perhaps reading) these references before attempting to grasp the fundamentals,
principals, concepts, and so on, associated with each chapter.
Finally, an important point needs to be made. There have been numerous occasions dur-
ing one of the author’s 50-year tenure as an educator when students solved a problem
using a packaged program such as Excel, MathCAD, and so on. For example, the problem
could have involved the solution to a differential equation or the regression of some data.
On being questioned how the packaged program performed the calculation, the student
almost always responded with something to the effect of, “I don’t know, and I don’t care.”
For this reason, the reader should note that no detailed attempt was made to introduce and
explain the packaged computer programs presently available that provide easy access to
trouble-free solutions of (even) complex mathematical problems, including optimization.
The emphasis in this introductory text on optimization was to provide the reader with
an understanding of fundamental principles in order to learn how these methods can be
used to obtain answers to questions within one’s own subject matter specialty; becoming
“computer literate” in this field was not the objective. Thus, anyone privy to applicable
software can obtain detailed solutions to large and complex optimization problems by
following relatively simple instructions associated with the software/computer. It should
also be noted that significant material of an example/problem nature was drawn from
lecture notes prepared for several courses offered by one of the authors during his career
as an educator; the source of some of this material is no longer known.
Authors
Over the past 50 years, Dr. Louis Theodore was a successful educator at Manhattan
College (holding the rank of full professor of chemical engineering), graduate program
director, researcher, professional innovator, and communicator in the engineering field.
During this period, he was primarily responsible for his program achieving a number two
ranking by US News & World Report. He has authored nearly 100 text/reference books
and over 100 technical papers. He currently serves as a part-time consultant to the US
Environmental Protection Agency (EPA) and runs Theodore Tutorials. He is a member of
Phi Lambda Upsilon, Sigma Xi, Tau Beta Pi, American Chemical Society, American Society
of Engineering Education, Royal Hellenic Society, and a fellow of the International Air &
Waste Management Association (AWMA). Dr. Theodore is the recipient of the AWMA’s
prestigious Ripperton award that is “presented to an outstanding educator, who, through
example, dedication, and innovation has so inspired students to achieve excellence in their
professional endeavors.” He was also the recipient of the American Society for Engineering
Education’s (ASEE) AT&T Foundation award for “excellence in the instruction of engineer-
ing students.”
Ms. Behan was employed in the utility industry and worked in the area of develop-
ing growth jobs and analyzing site plans. She also worked for the New York City Fire
Department and was primarily involved in physiological research on World Trade Center
victims. Ms. Behan previously served as Editorial Manager for a book on basketball coach-
ing. She is currently involved with structural and civil engineering activities. In addition,
her recent activities have centered on structural engineering with Turner Construction.
xix
Part I
Optimization Fundamentals
and Principles
As one might suspect, the term optimization has come to mean different things to differ-
ent people. It has also come to mean different things for different applications; for example,
it could involve a simple two-step calculation or one that requires the use of a detailed
numerical method. To take this a step further, the authors were initially undecided on how
to include optimization subject matter in this text. After much deliberation and medita-
tion, it was decided to present introductory material in the first Part of the book. This deci-
sion was primarily influenced by the desire for this book to be a book that both introduces
optimization mathematical methods and addresses applications that appear in Parts II
and III for environmental and chemical engineers, respectively, as well as in Part IV.
The optimization problem has been described succinctly by Aris1 as “getting the best
you can out of a given situation.” Problems amenable to solution by mathematical optimi-
zation techniques generally have one or more independent variables whose values must be
chosen to yield a viable solution to distinguish between the many viable solutions gener-
ated by the different choices of these variables. Mathematical optimization techniques are
generally used for guiding the problem solver to the choice of variables that maximizes the
“goodness” measure (e.g., profit) or that minimizes some “badness” measure (e.g., cost).
In addition to the optimization definition by Aris presented previously, one might offer the
following generic definition for many engineers: “Optimization is concerned with determin-
ing the ‘best’ solution to a given problem.”2 Alternatively, a dictionary would offer something
to this effect: “to make the most of … develop or realize to the utmost extent … often the most
efficient or optimum use of.” The process of optimization in practice is required in the solu-
tion of many problems and can involve the maximization or minimization of a mathematical
function. As one might suppose, many of the applications involve economic considerations.
One of the most important areas for the application of mathematical optimization tech-
niques is in chemical engineering design. Topics can include the following:
This first chapter in Part I serves to provide a broad overview of the general subject of
optimization. Section headings and titles are listed here:
3
4 Introduction to Optimization for Chemical and Environmental Engineers
The first three sections need no introduction. The fourth section addresses standard
(some might say traditional) optimization procedures employed by engineers today and in
the past. These include such topics as brute force methods, perturbation schemes, elemen-
tary search methods, graphical approaches, and analytical methods. A separate chapter is
devoted to each of these topics later in this Part. The last section is called Contemporary
Optimization (some might replace contemporary with complex) and primarily keys on
linear programming. These approaches will be emphasized in the solution to the environ-
mental and chemical engineering illustrative examples in Parts II and III, respectively, as
well as in Part IV.
SOLUTION:
Optimization is viewed by many as a tool in decision-making. It often aids in the selec-
tion of values that allow the practicing environmental and chemical engineer to better
solve a problem. This brief answer provides a qualitative look at optimization.
As noted previously, in its most elementary and basic form, one may say that optimi-
zation is concerned with the determination of the “best” solution to a given problem.
This process is required in the solution of many general problems in engineering and
applied science – in the maximization (or minimization) of a given function(s), in the
selection of a control variable to facilitate the realization of a desired condition, in the
scheduling of a series of operations or events to control completion dates of a given proj-
ect, in the development of optimal layouts of organizational units within a given design
space, and so on. In engineering design, once a particular subject or process scheme
has been selected for study, it is common practice to optimize the process from both a
capital cost and a O&M (operation and maintenance) standpoint.
There are many optimization procedures available, most of them too detailed for
meaningful application in a text of this nature. These sophisticated optimization tech-
niques, some of which are routinely used in the design of conventional chemical and
petrochemical plants, invariably involve computer calculations. Although the use of
these techniques in the majority of environmental and chemical engineering applica-
tions was not warranted in the past, more and more real-world problems are requiring
the use of optimization techniques.
1.1 History of Optimization
The subject of mathematics encompasses the study of relationships among quantities,
magnitudes, and properties and of logical operations by which unknown quantities,
magnitudes, and properties may be deduced. In the past, mathematics was essentially
regarded as a science of magnitudes and numbers. Toward the middle of the 19th century,
however, mathematics came to be regarded increasingly as the science of relations, or as
Optimization Overview 5
the science that draws necessary conclusions. This latter view encompasses mathemati-
cal or symbolic logic, the science of using symbols to provide an exact theory of logical
deduction and references based on definitions, axioms, postulates, and rules for com-
bining and transforming primitive elements into more complex relations and theorems.
Enter optimization.
This brief survey of the history of optimization traces the evolution of mathematical
ideas and concepts, beginning in prehistory. Indeed, mathematics is nearly as old as
humanity itself; evidence of a sense of geometry and interest in geometrical patterns
has been found in the designs of prehistoric pottery and textiles, and in cave paintings.
Primitive counting systems were almost certainly based on using the fingers of both
hands, as evidenced by the predominance of the number 10 as the base for many number
systems employed today.
Interestingly, mathematics of the late 19th century and the 20th century is character-
ized by an interest in unifying elements across numerous fields of mathematical endeavor,
especially in logic. For example, group theory has proved a particularly effective unifier.
The amount of new math and the particular topics arising at that time were numerous and
varied: unified set theory, intuitive geometry, the development of the number systems –
including methods of numeration, binary and other bases of notation, and modular arith-
metic and measurement, with attention to accuracy, precision, and error. It also included
studies of algebraic systems, linear algebra, modern algebra, vectors, matrices, logic, truth
tables, the nature of proofs, Venn and Euler diagrams, relations, functions, probability and
statistics, linear programming, and (of course) computer programming.
As to the origin of optimization, it depends on who provides the response, because
there are so many aspects of optimization of interest to the practitioner. For example, some
claim it was Thomas Edison when he developed a long-lasting, high-quality light bulb in
the 1870s. His success was primarily the result of an extensive trial-and-error search for
the optimum filament material. A few now refer to it as the Edisonian approach. One of the
authors refers to it as the perturbation approach; on occasion, he has modestly termed it
the Theodore approach.
Advances in computing have not only reduced the cost and time required to perform any
iteration calculation but also provide a better understanding of how this method works.
However, the success of this approach still relies heavily on the limitations of the user’s
intuition and experience since one often cannot predict or even comprehend the effects of
changing numerous variables in a complex system. Despite this barrier, the computer has
expanded and improved the employment of the classical methods of optimization.
With the promise of reducing design time and cost while improving product quality,
automated design optimization held tremendous potential. Starting with a sub-optimal
design, a numerical optimization algorithm could be used to iteratively adjust a set of pre-
selected design parameters in an attempt to achieve a set of design targets. This new class
of optimization technology enables broader, more comprehensive, and faster searches for
innovative designs than was possible using previous generations of tools. Moreover, it
requires no expertise in optimization theory, so it is easier to use for non-experts and
experts alike. By leveraging an engineer’s potential to discover new design concepts,
this new class of optimization technology overcomes the limits of human intuition and
extends the designer’s professional capability to achieve breakthrough designs and accel-
erated innovation.
As noted earlier, today’s computers now allow an optimization procedure (or math-
ematics, computer science, operations research, mathematical optimization, or math-
ematical programming) to select the best element (with regard to some criteria) from
6 Introduction to Optimization for Chemical and Environmental Engineers
some set of available alternatives. In the simplest case, an optimization problem con-
sists of optimizing (maximizing or minimizing) a real function by systematically
choosing input values from within an allowed set of variables in order to compute
the value of the function. Mathematically speaking, optimization allows one to find
the best available value(s) of some (often referred to as the objective) function given a
defined domain (or a set of constraints), including a variety of different types of objec-
tive functions and different types of domains. These various options are introduced
in the last section of this chapter and treated in more detail in Chapter 7. A feasible
solution to the objective function that minimizes (or maximizes), if that is the goal, is
called an optimal solution.
1. Perturbation methods
2. Search methods
3. Graphical approaches
4. Analytical methods
Introductory details on these topics follow, noting that linear programming briefly receives
an introduction in Section 1.5.
Topic (1) basically involves a guessing game when attempting to solve an optimiza-
tion problem. High-powered mathematics is usually not involved. The same can be said
for any trial-and-error method whether it be systematic or not. Numerous elementary
search methods (2) are available but only two will be discussed in Chapter 4. Graphical
approaches (3), as one might suppose, involve graphing available data and calculations in
an attempt to arrive at an “optimization” solution. Analytical methods (4) often employ
ordinary and partial derivatives in obtaining a solution.
Topics (1)–(4) have been defined by most mathematics as direct approaches, while linear
programming is referred to as an indirect approach. As one might assume, indirect meth-
ods are generally preferred. There are, however, numerous other methods – in addition
to those mentioned previously – available for solving optimization problems. These can
8 Introduction to Optimization for Chemical and Environmental Engineers
involve the aforementioned procedures that simply require a yes or no answer, select-
ing the best option of the two alternatives, selecting the best of more than two options,
and so on.
The selection of the independent variable(s) is often set in academic problems involv-
ing standard optimization calculations. The choice of these variables is usually based
on past experience or sound engineering judgment. For example, the performance of an
absorber – for control of a pollutant in an environmental engineering application or for
recovery in a chemical engineering application – is usually assumed to be a function of
temperature and pressure. Other effects, such as the time of day, or lunch details, or … are
usually and understandably neglected. This often reduces the problem to one that does not
involve the need to resort to any sophisticated optimization mathematical methods.
One should also note that these optimization problems, like many engineering prob-
lems, usually involve a simple two-step procedure:
Step (1) falls in the domain of the engineer. Step (2) traditionally fell in the domain of
both the engineer and mathematician. However, the second step is essentially no longer a
concern to many engineers with the advent of computers, particularly as it applies to con-
temporary optimization – a topic that is addressed in the next section.
1. Perturbation methods
2. search methods
3. graphical approaches
4. analytical methods
SOLUTION:
This is a difficult question to answer since it is most probably a function of the experi-
ence, interest, and capabilities of the individual being questioned. The answer for one of
the authors (2) earlier in his career would probably be (1) and (3). In more recent times,
both authors would definitely select (4).
variables must assume integer values. If the assumption of linearity cannot be made, as
in the case of linear programming, there exist some general procedures for nonlinear
problems. The collection of techniques developed for these problems is called nonlinear
programming.
The application of linear programming usually involves a constrained optimization
problem. The objective is to locate the values of the decision variable(s) that either maxi-
mize or minimize a linear objective function, where the decision variables are subject to
linear constraints. These mathematical exercises involve locating a value (the feasibility
value) that minimizes the objective function while satisfying the constraints. Thus, linear
programming provides a method of obtaining the optimum feasible value from a near
infinite number of values.
Some optimization problems can be divided into parts, for which each part is then opti-
mized. In some instances, it is possible to attain the optimum for the original problem by
simply realizing how to optimize these constituent parts. This process is very powerful, as
it allows one to solve a series of smaller, easier problems rather than one large one. One of
the best-known techniques to attack such problems is dynamic programming. This approach
is characterized by a process that is performed in stages, such as manufacturing processes.
The possible states of any stage of the process must include enough information regarding
conclusions and/or decisions made at the preceding stage (or stages) to allow an informed
decision at the present stage. Rather than solving the problem as a whole, dynamic pro-
gramming thus optimizes one stage at a time to produce an optimal set of decisions for
the whole process–in effect reformatting the optimization problem as a multistage one.
Although dynamic programming has applicability in some environmental and chemical
engineering systems/processes, it is not reviewed in this book.
One of the major responsibilities in optimization is to construct the correct objective
function to be maximized (or minimized). For example, in a simple environmental prob-
lem, the objective function might be the profit associated with the operation of a landfill
that accepts two categories of waste: “inert” and “non-inert.” One optimization problem
might be to maximize the profit (from handling and treating the waste) by employing an
operating treatment schedule subject to the landfill’s capacity and environmental regula-
tory requirements. This problem might also be rephrased in the following manner: What
combination of the quantities of the two wastes will produce the maximum profit subject
to the constraints connected with the landfill’s processing capabilities and the different
regulatory requirements imposed on each waste?
The objective function may be quite simple and easy to calculate in some environmental
and chemical engineering applications, or it may be complicated and difficult to not only
calculate but also specify and/or describe. The objective function may also be very elusive
due to the presence of conflicting or dimensionally incompatible objectives; for example,
one might be asked to optimize the profit for the aforementioned landfill that not only
minimizes air and water contamination and/or emissions but also is aesthetically appeal-
ing. Thus, it may not be always possible to quantify an objective function and hence be able
to use any of the mathematical optimization procedures available.
SOLUTION:
The (objective) function is
y = 4 x1 − 3 x2
0 ≤ x1 + x2 ≤ 10
and
0 ≤ x1x2 ≤ 36
plus
0 ≤ x1 and 0 ≤ x2
The remainder of Part I addresses six additional topics related to optimization. Chapter
numbers and subject matter are again presented here:
This material provides optimization fundamentals and principles ranging from the
optimization (or minimization) of a single variable function to more detailed nonlinear
constrained problems. Although the optimization methods are provided with comput-
ers in mind, the reader is again reminded that programming and coding details are not
addressed. Also note that the performance criteria employed can vary from classical eco-
nomic choices to technological factors, such as energy conservation, the degree of purity,
and so on.
Finally, most engineering optimization applications involve economics: maximizing
profit, minimizing cost, or both. In addition to the traditional engineering applications
discussed earlier, environmental applications can involve minimizing toxic emissions,
maximizing energy usage and conservation, minimizing health and safety risks, standard
economic concerns, and so on. The reader should not lose sight of the fact that most real-
world industry applications involving optimization usually require simple solutions. Here
is a comment from a retired engineer:5 “Generally, the consumer goods industry where
I worked does not rely on sophisticated models. Basically, optimization was conducted
to determine overall consumer preference or liking, to determine how to use minimum
resources, and lastly, to determine how to minimize costs subject to constraints. In the
determination of consumer preference, we typically used design of experiments6 to exam-
ine the broad range of attributes that led to maximum product liking or decision to pur-
chase. Occasionally, we modeled processes and used regression models.6 In that case, we
could search or compute the desired maximum. The last set of techniques used is what the
Optimization Overview 11
business world calls ‘Operations Research.’ Typically, linear or dynamic programming was
used to evaluate an ‘objective function subject to constraints.’ For example, we used linear
programming in hot dog manufacturing. The goal was to come up with a mixture of meat
cuts that minimized costs while meeting governmental requirements for protein, fat, and
water content.”
References
1. R. Aris, Discrete Dynamic Programming, Blaisdell, Boston, MA, 1964.
2. L. Theodore, personal notes, East Williston, NY, 2011.
3. J. Famularo, personal communication to L. Theodore (with permission), Englewood Cliffs, NJ,
1973.
4. L. Theodore, personal notes, East Williston, NY, 2005.
5. R. Altomare, communication to L. Theodore, Bronx, NY, 2016.
6. S. Shaefer and L. Theodore, Probability and Statistics Applications for Environmental Science, CRC
Press/Taylor & Francis Group, Boca Raton, FL, 2007.
2
Mathematical Operations
The natural numbers, or so-called counting numbers, are the positive integers: 1, 2, 3, …
and the negative integers: −1, −2, −3, … The following applies to real numbers:
1 1
If a < b and ab > 0, then > (2.5)
a b
a + b = b + a; a ⋅ b = b ⋅ a = ( a )(b ) (2.6)
a + (b + c ) = ( a + b ) + c; a ⋅ (bc ) = ( ab ) ⋅ c (2.7)
a (b + c ) = a ⋅ b + a ⋅ c; with a + 0 = a; a ⋅ 1 = a (2.8)
1
a− n = , a ≠ 0 (2.11)
an
( ab )n = anb n (2.12)
(a )
n m
= a nm , a n a m = a n+ m (2.13)
a0 = 1, a ≠ 0 (2.14)
a
log = log a − log b (2.17)
b
13
14 Introduction to Optimization for Chemical and Environmental Engineers
1
log n a = log a (2.18)
n
ln a = (log e 10 ) common log a = ( 2.3026 ) common log a = 2.3026 log a (2.19)
( 4)(9) = (3.60)(10)1
(6)3 = (2.16)(10)2
3
3375 = (1.50 )(10 ) =
1
( 3
3.375 )( 3
1000 )
log ( 245) = 2.3892
ln ( 24.5) = 3.199
ln (0.245) = −11.4065
The authors have assumed that the reader is not only familiar with that described previ-
ously but also other more advanced topics that can include simultaneous linear algebraic
equations, nonlinear equations, differentiation, integration, the solution of both ordinary
differential equations and partial differential equations, and the Monte Carlo method.
Only the numerical solution approaches for the previously mentioned are reviewed in
this chapter. Elementary statistics and least squares approximation are not reviewed, but
details are available in the literature.1
Ten sections complement this mathematical operations chapter:
The first four sections may be viewed as elementary material. The last six sections are
more quantitative and can be viewed as fitting in the domain of numerical methods. Since
a detailed treatment of each of these topics is beyond the scope of this text, the reader is
referred to the literature2–4 for more extensive analysis and additional information. The
remainder of this chapter briefly examines the topics listed previously.
ax 2 + bx + c = 0 (2.20)
a number of methods of solution are possible depending on the specific nature of the equa-
tion in question. If the equation can be factored, then the solution is straightforward. For
instance, consider
x 2 − 3 x = 10 (2.21)
x 2 − 3 x − 10 = 0 (2.22)
( x − 5) ( x + 2) = 0 (2.23)
This condition can be met, however, only when the individual factors in parentheses are
zero, i.e., when x = 5 and/or x = −2. That these are indeed the solutions to the equation may
be verified by substitution.
If, upon inspection, no obvious means of factoring the above equation can be found, an
alternative approach may exist. For example, in the equation
4 x 2 + 12x = 7 (2.24)
the expression
4 x 2 + 12x (2.25)
4 x 2 + 12x + 9 (2.26)
which equals
This can easily be achieved by adding 9 to the left side of the equation. The same amount
must then, of course, be added to the right side as well, resulting in:
4 x 2 + 12x + 9 = 7 + 9 (2.28)
(2x + 3 ) = 16 (2.30)
or
2x + 3 = +4 (2.31)
and
2x + 3 = −4 (2.32)
Since 16 has two solutions, the first equation leads to the solution x = 1/2 or x = 0.5, while
the second equation leads to the solution x = −7/2, or x = −3.5.
If the methods of factoring or completing the square are not possible, any quadratic
equation can always be solved by the quadratic formula. This provides a method for deter-
mining the solution of the equation if it is once again in the form
ax 2 + bx + c = 0 (2.33)
−b ± b 2 − 4 ac
x= (2.34)
2a
For example, to find the roots of
x 2 − 4 x = −3 (2.35)
x 2 − 4 x + 3 = 0 (2.36)
As a result, a = 1, b = −4 and c = 3. These terms are then substituted into the quadratic for-
mula presented in Equation 2.34.
− ( −4 ) ± ( −4)2 − 4 (1)(3) 4 ± 16 − 12
x= = (2.37)
2 (1) 2
4± 4 4±2
= = = 3, and 1 (2.38)
2 2
Mathematical Operations 17
The practicing environmental and chemical engineer occasionally has to solve not just a
single equation but several at the same time. The problem is to find the set of all solutions
satisfying all equations. These are called simultaneous equations, and specific algebraic
techniques may be used to solve them. For example, a simple solution exists for the follow-
ing two linear equations with two unknowns:
3 x + 4 y = 10 (2.39)
2x + y = 5 (2.40)
The variable y in Equation 2.40 is isolated (y = 5 − 2x), and then this value of y is substituted
into Equation 2.39. The result is Equation 2.41.
3 x + 4 ( 5 − 2x ) = 10 (2.41)
This reduces the problem to one involving the single unknown x, and it follows that
so that
x = 2 (2.43)
When this value is substituted into either previous equation, it follows that
y = 1 (2.44)
8 x + 4 y = 20 (2.45)
3 x + 4 y = 10 (2.39)
If Equation 2.39 is subtracted from Equation 2.45, 5x = 10 results, or x = 2. This procedure
leads to another development in mathematics, matrices, which can help to produce solu-
tions for any set of linear equations with a corresponding number of unknowns (see also
Section 2.7).
y = a + bx (2.46)
and then employ this equation to calculate y for some value of x lying between x1 and
x 2. Most practitioners do this mentally when reading values from the (steam) tables.
If a number of points are used, a polynomial of a correspondingly higher degree may
also be employed.1 Interpolation is necessary to find y when x is some value not given
in the table. Thus, interpolation may be viewed as the process of finding the value of
a function at some arbitrary point where the function is not known but is represented
over a given range as a table of discrete points. (See also Table 2.1 where y represents a
reservoir’s height as a function of time in days during a rainy season.) For instance, one
may be interested in finding y when x = 7. (The process of finding x when y is known
is referred to as inverse interpolation). Given a table with (x, y) data (Table 2.1), one can
draw a picture (Figure 2.1) and write the equation of the straight line through the points
(x1, y1) and (x 2, y2) for y.
TABLE 2.1
Reservoir Height
x, Time y, Height
0 30
3 31
6 33
9 35
12 39
15 46
18 52
Mathematical Operations 19
35
30
25
20
y2
y
15
y1
10
5 x1 x2
0
1 2 3 4 5 6 7 8 9 10
x
FIGURE 2.1
Interpolation procedure.
y − y1
y − y1 = 2
x2 − x1
( x − x1 ) (2.47)
Equation 2.47 can be solved for y in terms of x
( y 2 − y1 ) ( x − x1 )
y = y1 + (2.48)
x2 − x1
or (for points 0 and 1)
y 0 ( x − x 0 ) + ( y1 − y 0 ) ( x − x 0 )
y= (2.49)
x1 − x0
SOLUTION:
Proceed as follows, keying (obviously) on the values for x = 9 and x = 12.
y = 35 +
( 4)( 2) = 35 + 8 = 37.67
3 3
Inverse interpolation involves estimating x which corresponds to a given value of y, and
(as noted previously) extrapolation involves estimating values of y outside the interval in
which the data x0, … , xn fall. It is generally unwise to extrapolate any empirical relation
significantly beyond the last data point or before the first point. If, however, a certain form
of equation is predicted by theory and substantiated by (other) available data, reasonable
extrapolation is ordinarily justified.
20 Introduction to Optimization for Chemical and Environmental Engineers
1. the last digit (whether it is nonzero or zero) on the right, if there is a decimal point
present, or
2. the last nonzero digit on the right of the number, if there is no decimal point
present.
For example,
0.000002807 = 2.807 × 10 −6 = ( 2.807 )(10 )
−6
A positive feature of using scientific notation is that only the significant figures need to
appear in the number.
When approximate numbers are added, or subtracted, the results are also in terms of the
least precise number. Since this is a relatively simple rule to master, note that the answer
follows the aforementioned rule of precision. For example,
(The complete result is 4.667 L.) The terms in Equation 2.50 have two, one, and three deci-
mal places, respectively. The least precise previous number (least decimal places) is 2.8, a
value carried only to the tenths position. Therefore, the answer must be reported to the
tenths position only. Thus, the correct answer is 4.7 L, with the last 6 and the 7 dropped
from the 4.667 L, and the first 6 is rounded up to provide 4.7 L.
In multiplication and division of approximate numbers, finding the number of signifi-
cant digits is used to determine how many digits to keep (i.e., where to truncate). One must
first understand significant digits in order to determine the correct number of digits to keep
or remove in multiplication and division problems. As noted earlier in this section, the dig-
its 1 through 9 are considered to be significant. Thus, the numbers 123, 53, 7492, and 5 contain
three, two, four and one significant digits, respectively. The digit zero must be considered
separately.
Zeroes are significant when they occur between significant digits. In the following examples,
all zeroes are significant: 10001, 402, 1.1001, 500.09 has five, three, five, and four signifi-
cant figures, respectively. Zeroes are not significant when they are used as placeholders. When
used as a placeholder, a zero simply identifies where a decimal is located. For example,
each of the following numbers has only one significant digit: 1000, 500, 60, 0.09, 0.0002.
In the number 1200, 540, and 0.0032 there are two significant digits, and the zeroes are
not significant. When zeroes follow a decimal and are preceded by a significant digit, the zeroes
are significant. In the following examples, all zeroes are significant: 1.00, 15.0, 4.100, 1.90,
10.002, 10.0400. For 10.002, the zeroes are significant because they fall between two sig-
nificant digits. For 10.0400, the first two zeroes are also significant because they fall
between two significant digits; the last two zeroes are significant because they follow a
decimal and are preceded by a significant digit. Thus, when approximate numbers are
multiplied or divided in a problem, the result is expressed as a number having the same
number of significant digits as the expression in the problem having the least number of
significant digits.
When truncating (removing final, unwanted digits), rounding is normally applied to
the last digit to be kept. Thus, if the value of the first digit to be discarded is less than 5,
one should retain the last kept digit with no change. If the value of the first digit to be
discarded is 5 or greater, one should increase the last kept digit’s value by one. Assume,
for example, only the first two decimal places are to be kept for 25.0847 (the 4 and 7 are to
be dropped). The number is then 25.08. Since the first digit to be discarded (4) is less than
5, the 8 is not rounded up. If only the first two decimal places are to be kept for 25.0867
(the 6 and the 7 are to be dropped), it should be rounded to 25.09. Since the first digit to be
discarded (6) is 5 or more, the 8 is rounded up to 9.
22 Introduction to Optimization for Chemical and Environmental Engineers
When adding or subtracting approximate numbers, a rule based upon precision deter-
mines how many digits are kept. In general, precision relates to the decimal significance
of a number. When a measurement is reported as 1.005 cm, one can say that the number is
precise to the thousandth of a centimeter. If the decimal is removed (1005 cm), the number
is precise to thousands of centimeters.
For example, in some water pollution studies, a measurement in gallons or liters may
be required. Although a gallon or liter may represent an exact quantity, the measuring
instruments that are used are only capable of producing approximations. Using a stan-
dard graduated flask in liters as an example, can one determine whether there is exactly
one liter? Not likely. In fact, one would be pressed to verify that there was a liter to within
±1/10 of a liter. Therefore, depending upon the instruments used, the precision of a given
measurement may vary.
If a measurement is given as 16.0 L, the zero after the decimal indicates that the measure-
ment is precise to within 1/10 L, that is, 0.1 L. Given a measurement of 16.00 L, one has
precision to the 1/100 L. As noted, the digits following the decimal indicate how precise
the measurement is. Thus, precision is used to determine where to truncate when approxi-
mate numbers are added or subtracted.
2.4 Errors
This section introduces the various classes of errors that arise in technical calculations
while the next section demonstrates the propagation of some of these errors. As one might
suppose, numerous books have been written on the general subject of “errors.” Different
definitions for errors appear in the literature, but what follows is the authors’ attempt to
clarify the problem.1
Any discussion of errors would be incomplete without providing a clear and concise
definition of two terms: the aforementioned precision and accuracy. The term precision is
used to describe a state or system or measurement for which the word precise implies
little to no variation; some refer to this as reliability. Alternatively, accuracy is used to
describe something free from the matter of errors. The accuracy of a value, which may be
represented in either absolute or relative terms, is the degree of agreement between the
measured value and the true value.
All measurements and calculations are subject to two broad classes of errors: determi-
nate and indeterminate. The error is known as a “determinate error” if an error’s mag-
nitude and sign are discovered and accounted for in the form of a correction. All errors
that either cannot be or are not properly allowed for in magnitude and sign are known as
“indeterminate errors.”
A particularly important class of indeterminate errors is that of accidental errors. To
illustrate the nature of these, consider the very simple and direct measurement of tem-
perature. Suppose that several independent readings are made and that temperatures are
read to 0.1°F. When the results of the different readings are compared, it may be found that
even though they have been performed very carefully, they may differ from each other by
several tenths of a degree. Experience has shown that such deviations are inevitable in all
measurements and that these result from small unavoidable errors of observation due to
the sensitivity of measuring instruments and the keenness of the sense of perception. Such
errors are due to the combined effect of a large number of undetermined causes, and they
can be defined as “accidental errors.”
Mathematical Operations 23
Regarding the words precision and accuracy, it is also important to note that a result
may be extremely precise and at the same time inaccurate. For instance, the temperature
readings just mentioned might all agree within 1°F. From this, it would not be permis-
sible to conclude that the temperature is accurate to 1°F until it can be definitively shown
that the combined effects of uncorrected constant errors and known errors are negligible
compared with 1°F. It is quite conceivable that the calibration of the thermometer might be
grossly incorrect. Errors such as these are almost always present and can never be detected
individually. Such errors can be detected only by obtaining the readings with several dif-
ferent thermometers and, if possible, several independent methods and observers.
It should also be understood once again that most numerical calculations are by their
very nature inexact. The errors are primarily due to one of three sources: inaccuracies in
the original data, lack of precision in carrying out calculations, or inaccuracies introduced
by approximate or incorrect methods of solution. Of particular significance are the afore-
mentioned errors due to “round-off” and the inability to carry more than a certain number
of significant figures.
The errors associated with the method of solution are usually the area of greatest con-
cern.3 These usually arise as a result of approximations and assumptions made in the
development of an equation used to calculate a desired result and should not be neglected
in any error analysis.
Finally, many list the following three errors associated with a computer (calculator):
1.
Truncation error: With the truncation of a series after only a few terms, one is com-
mitting a generally known error. This error is not machine-caused but is due to
the method.
2.
Round-off error: This is machine-made and is caused by the limitations of the par-
ticular computer. If one computes using eight digits in multiplication, one would
expect 16 digits by hand calculation. The computer retains only the most signifi-
cant eight of these. If the number was rounded, this error would be minimized,
but rounding is generally not done.
3.
Propagation or inherited error: This is caused by sequential calculations that include
points previously calculated by the computer that already are erroneous owing
to the two previously mentioned errors. Since the result is already off the solu-
tion curve, one cannot expect any new points computed to be on the correct solu-
tion curve. Adding the round-off errors and truncation errors into the calculation
causes further errors to propagate adding more error at each step.
2.5 Differentiation
Several differentiation methods are available for generating expressions for a derivative.
Consider the problem of determining the benzene concentration – time gradient dC/dt5 at
t = 4.0 s; refer to Table 2.2. Six differentiating methods are presented subsequently.5
Method 1: This method involves the selection of any three data points and calculating
the slope m of the two extreme points. This calculated slope is approximately equal to the
slope at the point lying in the middle. The value obtained is the “equivalent” of the deriva-
tive at that point 4. Using data points from 3.0 to 5.0, one obtains
24 Introduction to Optimization for Chemical and Environmental Engineers
TABLE 2.2
Time-concentration Data
Time (s) Concentration of Benzene (mg/L), C
0.0 7.46
1.0 5.41
2.0 3.80
3.0 2.70
4.0 2.01
5.0 1.63
6.0 1.34
7.0 1.17
C5 − C3
Slope = m =
t5 − t3
1.63 − 2.70
= = −0.535
5.0 − 3.0
Method 2: This method involves determining the average of two slopes. Using the same
points chosen previously, two slopes can be calculated, one for points 3 and 4 and the other
for points 4 and 5. Adding the two results and dividing them by 2 will provide an approxi-
mation of the derivative at point 4. For the points used in this method, the results are:
C 4 − C3
m1 = slope1 =
t 4 − t3
2.01 − 2.70
= = −0.69
4.0 − 3.0
C − C4
m2 = slope2 = 5
t5 − t 4
1.63 − 2.01
= = −0.38
5.0 − 4.0
−0.69 + ( −0.38 )
mavg = slopeavg = = −0.535
2
Method 3: This method consists of selecting any three data points (in this case the same
points chosen before) and fitting a curve to it. The equation for the curve is obtained by
employing a second-order equation and solving it with the three data points. The follow-
ing equation results:
The analytical derivative of the equation is calculated and may then be used to evaluate
the slope at any point. Here, point 4 is used:
dC
= 0.31t − 1.775
dt
Mathematical Operations 25
Evaluated at t = 4.0 s
dC
= 0.31 ( 4.0 ) − 1.775 = −0.535
dt
Method 4: This method uses the method of least squares.1 In this case, all the data points
are used to generate a second-order polynomial equation. This equation is then differenti-
ated and evaluated at the point where the value of the derivative is required. For example,
Microsoft Excel can be employed to generate the regression equation. Once all the coef-
ficients are known, the equation is once again analytically differentiated.
dC
= 0.3252t − 1.9905
dt
Evaluated at t = 4.0 s:
dC
= 0.3252 ( 4.0 ) − 1.9905 = −0.6897
dt
Methods 5 and 6: These two methods are somewhat similar. They are based on five data
points used to generate coefficients. For this development, represent C and t by f and x (as
it appeared in the literature6), respectively. Method 5 employs five data points to generate a
five coefficient (fourth-order) model using an equation of the form f = A + Bx+ Cx2+ Dx3+ Ex4.
This method is known as interpolating. A set of equations is used to evaluate numerical
derivatives from the interpolating polynomial. The describing equations are listed here:
−25 f 0 + 48 f1 − 36 f 2 + 16 f 3 − 3 f 4
mo = f ′ ( x0 ) = (2.51)
12h
−3 f 0 − 10 f1 + 18 f 2 − 6 f 3 + 3 f 4
m1 = f ′ ( x1 ) = (2.52)
12h
f i − 2 − 8 f i −1 + 8 f i + 1 − f i − 2
mi = f ′ ( xi ) = (2.53)
12h
For example, the equation obtained for “the five-data set” from 1.0 to 5.0 s, that is, t = 1.0,
2.0, 3.0, 4.0, and 5.0 s, using the equations given previously is
All these equations are evaluated for each value of x and f(x). The value obtained for point
4.0 is −0.5448. Method 6 also employs five data points, but only three coefficients are gen-
erated for a second-order polynomial equation of the form f = A + Bx+ Cx2. Another set of
equations is used to evaluate the derivative at each point using this method. The describ-
ing equations are provided here:6
f ′ ( x0 ) =
( −54 f0 + 13 f1 − 40 f2 + 27 f3 − 26 f 4 ) (2.56)
70 h
f ′ ( x1 ) =
( −34 f0 + 3 f1 + 20 f2 + 17 f3 − 6 f 4 ) (2.57)
70 h
f ′ ( xi ) =
( −2 fi−2 − fi−1 + fi+1 + 2 fi−2 ) (2.58)
70 h
f ′ ( x n −1 ) =
(6 fn−4 − 17 fn−3 − 20 fn−2 − 3 fn−1 + 34 f3 ) (2.59)
70 h
f ′ ( xn ) =
(26 fn−4 − 27 fn−3 − 40 fn−2 − 13 fn−1 + 54 f3 ) (2.60)
70 h
At point 4.0, the solution for the derivative using the method is −0.6897.
Comparing all six values obtained for the derivative at t = 4.0 s, one can conclude that the
answers are in close proximity to each other. It is important to note that these are approxi-
mate values and that they vary depending on the approach and the number of data points
used to generate the equations.
Some useful analytical derivatives in engineering calculations are provided in the
literature.4
2.6 Numerical Integration
Numerous engineering and science problems require the solution of integral equations.
In a general sense, the problem is to evaluate the function on the right-hand side (RHS) of
Equation 2.61.
∫
I = f ( x ) dx (2.61)
a
where I is the value of the integral. There are two key methods employed in their solution:
analytical and numerical. If f(x) is a simple function, it may be integrated analytically. For
example, if f(x) = x2, then
∫
I = x 2dx = (
1 3
3
)
b − a 3 (2.62)
a
Mathematical Operations 27
If, however, f(x) is a function too complex to integrate analytically, for example,
I = log tanh e x
( ) (2.63)
2
−2
one may resort to any of the many numerical methods available. Two simple numerical inte-
gration methods that are commonly employed in environmental and chemical engineer-
ing practice are the trapezoidal rule and Simpson’s rule. These are described subsequently.
2.6.1 Trapezoidal Rule
In order to use the trapezoidal rule to evaluate the integral I given by Equation 2.61 as
b
∫
I = f ( x ) dx (2.61)
a
h
I= y0 + 2 y1 + 2 y 2 + + 2 y n−1 + y n (2.64)
2
where:
h is the incremental change in x; that is, ∆x
yi are the values of f(x) at xi, that is, f(xi)
In addition,
y0 = f ( x0 ) = f ( x = a ) (2.65)
y n = f ( xn ) = f ( x = b ) (2.66)
b a
h=
(2.67)
n
This method is known as the trapezoidal rule because it approximates the area under
the function f(x) – which is generally curved – with a two-point trapezoidal rule calcula-
tion. The error associated with this rule is illustrated in Figure 2.2.
f(x)
f(x)
f(a) f(b)
x
= Error = Calculated value
FIGURE 2.2
Trapezoidal rule error.
28 Introduction to Optimization for Chemical and Environmental Engineers
There is an alternative for improving the accuracy of this calculation – the interval (a, b)
can be subdivided into smaller intervals. The trapezoidal rule can then be applied repeat-
edly in turn over each subdivision.
2.6.2 Simpson’s Rule
A higher-degree interpolating polynomial scheme can be employed for more accurate
results. One of the more popular integration approaches is Simpson’s rule. For Simpson’s
3-point (or one-third) rule, one may use the equation
h
I= y a + 4 yb+ a/2 + yb (2.68)
3
The equation for the general form of Simpson’s rule (where n is an even integer) is
h
I=
3
( y0 + 4 y1 + 4 y2 + + 4 yn −1 + yn ) (2.69)
This method also generates an error, although it is usually smaller than that associated
with the trapezoidal rule. A diagrammatic representation of the error for a 3-point calcula-
tion is provided in Figure 2.3.
The reader should note that the trapezoid rule is often the quickest but least accurate
way to perform a numerical integration by hand. However, if the step size is decreased,
the answer should converge – subject to round-off error – to the analytical solution. The
results of each numerical integration must be added together to obtain the final answer for
smaller step sizes.
Some useful analytical integrals in engineering calculations are provided in the literature.4
f(x)
f(x)
Error
Simpson
x
x0 x1 x2
= Error
FIGURE 2.3
Simpson’s rule error.
Mathematical Operations 29
a21x1 + a22 x2 + … a2 n xn = y 2
(2.70)
an1x1 + an 2 x2 + … ann xn = y n
where:
a is the coefficient of the variable x
y is a constant
This set is considered to be linear as long as none of the x terms are nonlinear, for exam-
2
ple, x2 or ln x1; thus, a linear system requires that all terms in x be linear. The previously
mentioned system of linear algebraic equations may be set in matrix forms:
1. Gauss–Jordan reduction
2. Gauss elimination
3. Gauss–Seidel approach
4. Cramer’s rule
5. Cholesky’s methods
2.7.1 Gauss–Jordan Reduction
Carnahan and Wilkes2 provide an example that solved the following two simultaneous
equations using the Gauss–Jordan reduction method:
3 x1 + 4 x2 = 29 (2.73)
6 x1 + 10 x2 = 68 (2.74)
4 29
x1 + x2 = (2.75)
3 3
2. Subtract a suitable multiple, in this case 6, of Equation 2.74 from Equation 2.73, so
that x1 is eliminated. Equation 2.75 remains untouched, leaving
34 5 4 29
− x2 + x2 = (2.76)
3 3 3 3
with the corresponding solution
2x2 = 10 (2.77)
3. Divide Equation 2.77 by the coefficient of x2, that is, solve Equation 2.77.
x2 = 5 (2.78)
4. Subtract a suitable factor of Equation 2.78 from Equation 2.75 so that x2 is elimi-
nated. When (4/3)x2 = 20/3 is subtracted from Equation 2.75, one obtains
x1 = 3 (2.79)
2.7.2 Gauss Elimination
Gauss elimination is another method used to solve linear sets of equations. This
method utilizes the augmented matrix as presented in Equation 2.72. The goal with
Gauss elimination is to rearrange the augmented matrix into a triangle form, where all
the elements below the diagonal are zero. This is accomplished in much the same way
as in Gauss–Jordan reduction. The procedure employed follows. Start with the first
equation in the set. This is known as the pivot equation and will not change through-
out the procedure. Once the matrix is in triangle form, back substitution can be used
to solve for variables.7
Gauss elimination is useful for systems that contain fewer than 30 equations. Systems
larger than 30 equations become subject to round-off error where numbers are truncated by
computers performing the calculations. This method is also effective to calculate the rank,
determinant, and inverse of an invertible matrix.
Mathematical Operations 31
2.7.3 Gauss–Seidel Approach
Another approach to solving an equation or series/sets of equations is to make an informed
or educated guess. If the first assumed value(s) does not work, the value is updated. By
carefully noting the influence of these guesses on each variable, one can approach these
answers or correct set of values for a system of equations.7 The reader should note that
when this type of iterative procedure is employed, a poor initial guess does not prevent the
correct solution from ultimately being obtained.
Ketter and Prawler3 provide several excellent illustrative examples.
f ( x ) = 0 (2.80)
is obtained by guessing a value for each x, for example (xold), that will satisfy this equa-
tion. This value is continuously updated (xnew) using the equation (the prime represents a
derivative)
f ( xold )
xnew = xold − (2.81)
f ′ ( xnew )
until either little or no change in (xnew − xold) is obtained. One can also express this opera-
tion graphically (see Figure 2.4). Noting that
f(x)
df (x)
Slope – – f ¢ (xold)
dx x = xold
f(xold)
xnew xold
Exact solution
since f(x) = 0
FIGURE 2.4
NR method.
32 Introduction to Optimization for Chemical and Environmental Engineers
df ( x ) ∆f ( x ) f ( xold ) − 0
f ′ ( xold ) = ≈ = (2.82)
dx ∆x xold − xnew
one may rearrange Equation 2.82 to yield Equation 2.83 below. The xnew then becomes xold
in the next calculation.
This method is also referred to as Newton’s method of tangents (NMT) and is a widely used
method for improving a first approximation to a root to the aforementioned equation of
the form f(x) = 0. This development can be rewritten in subscripted form to (perhaps) better
accommodate a computer calculation. Thus
f ( xn )
f ′ ( xn ) = (2.83)
x n − x n +1
from which
f ( xn )
x n +1 = x n − (2.84)
f ′ ( xn )
The term xn + 1is again the improved estimate of xn – the previous guess for a solution
to the equation f(x) = 0. The value of the function and the value of the derivative of the
function are determined at x = x n, and using the proposed procedure, and the new
approximation to the root xn + 1 is obtained. The same procedure is repeated, with the
new value, to obtain a still better approximation of the root. This continues until suc-
cessive values for the approximate root differ by less than a prescribed small value
ε, which controls the allowable error (or tolerance) in the root. Relative to the previous
estimate, ε is given by
x n +1 − x n
ε= (2.85)
xn
Despite its popularity, the method suffers for two reasons. First, an analytical expres-
sion for the derivative, specifically, f′(xn), is required. In addition to the problem of having
to compute an analytical derivative value at each iteration, one would expect Newton’s
method to converge fairly rapidly to a root in all cases. However, as is common with some
numerical methods, it may fail occasionally in certain instances. A possible initial oscil-
lation followed by a displacement away from a root can occur. Note, however, that the
method would have converged if the initial guess had been somewhat closer to the exact
root. Thus, the first guess may be critical to the success of the calculation.8,9
dy
= f ( x , y ) (2.86)
dx
Mathematical Operations 33
h
y n +1 = y n +
6
(D1 + 2D2 + 2D3 + D4 ) (2.87)
where
D1 = hf ( x , y )
h D
D2 = hf xn + , y n + 1
2 2 (2.88)
h D
D3 = hf xn + , y n + 2
2 2
D4 = hf ( xn + h, y n + D3 )
The term h represents the increment in x. The term yn is the solution to the equa-
tion at yn + 1, and yn + 1 is the solution to the equation at xn + 1 where xn + 1 = xn + h. Thus,
the RK method provides a straightforward means for developing expressions for ∆y,
namely, xn + 1 −yn, in terms of the function f(x,y) at various “locations” along the interval
in question.
Consider a simple equation of the form
dC
= a + bC (2.89)
dt
where:
t = 0
C = C0
h
C1 = C0 +
6
(D1 + 2D2 + 2D3 + D4 ) (2.90)
where
D1 = hf ( x , y ) = h ( a + bC0 )
h D D
D2 = hf xn + , y n + 1 = h a + b C0 + 1
2 2 2
h D C + D2 (2.91)
D3 = hf xn + , y n + 2 = h a + b 0
2 2 2
D4 = hf ( xn + h, y n + D3 ) = h a + b (C0 + D3 )
The same procedure is repeated to obtain values for C2 at t = 2h, C3 at t = 3h, and so on.
34 Introduction to Optimization for Chemical and Environmental Engineers
The RK method can also be used if the function in question also contains the indepen-
dent variable.8,9 Consider the following equation:
dC
= f (C , t ) (2.92)
dt
For this situation, one obtains
h
C1 = C0 +
6
(D1 + 2D2 + 2D3 + D4 ) (2.93)
with
D1 = hf (C , t )
D h
D2 = hf C0 + 1 , t0 + (2.94)
2 2
D h
D3 = hf C0 + 2 , t0 +
2 2
D4 = hf (C0 + D3 , t0 + h )
For example, if
dC
= 5C − e − Ct (2.95)
dt
then
D1 h
− C0 + 2 t0 + 2
dC D1
= h 5 C0 + − e
(2.96)
dt 2
Situations may arise when there is a need to simultaneously solve more than one ordi-
nary differential (ODE). In a more general case, one could have n dependent variables
y1, y2, … yn with each related to a single-independent variable x by the following system
of n simultaneous first-order ODEs:
dy1
= f 1 ( x , y1 , y 2 , … , y n )
dx
dy 2
= f 2 ( x , y1 , y 2 , … , y n )
dx (2.97)
dy n
= f n ( x , y1 , y 2 , … , y n )
dx
Mathematical Operations 35
Note that the equations in Equation 2.97 are interrelated, that is, they are dependent on
each other. This is illustrated in the following two equations:8
dC
= Ae − E/RT C = f (C , t ) (2.98)
dt
dT ∆H
= − kC = g (C , t ) (2.99)
dt ρCp
or in a more general sense
dy
= f (x, y , z) ; (e.g., xyz ) (2.100)
dx
dz
dt
= g (x, y , z) ; (e.g., x y e ) (2.101)
2 2 −z
1
y1 = y 0 +
6
(RY1 + 2RY2 + 3RY3 + RY4 ) (2.102)
1
z1 = z0 +
6
(RZ1 + 2RZ2 + 3RZ3 + RZ4 ) (2.103)
where y1 − y0 = ∆y , z1 − z0 = ∆z , h = ∆x and
RY1 = h × f ( x0 , y0 , z0 )
RZ1 = h × f ( x0 , y0 , z0 )
h RY1 RZ1
RY2 = h × f x0 + , y0 + , z0 +
2 2 2
h RY1 RZ1
RZ2 = h × f x0 + , y0 + , z0 +
2 2 2
(2.104)
h RY2 RZ2
RY3 = h × f x0 + , y0 + , z0 +
2 2 2
h RY2 RZ2
RZ3 = h × f x0 + , y0 + , z0 +
2 2 2
Although the RK approach (and other similar methods) has traditionally been employed
to solve first-order ODEs, it can also treat higher ODEs. The procedure requires reducing
an nth-order ODE to n first-order ODE. For example, if the equation is of the form8
d2 y
= f ( y , x ) (2.105)
dx 2
set
dy
z= (2.106)
dx
so that
dz d 2 y
= (2.107)
dx dx 2
The second-order equation in Equation 2.107 has now been reduced to the two first-order
ODEs provided in Equation 2.108:
d 2 y dz
= = f ( y , x ) (2.108)
dx 2 dx
dy
= z
dx
The procedure expressed in Equation 2.102 and 2.103 can be applied to generate a solu-
tion to Equation 2.105. Note, however, that the first derivative (i.e., dy/dx or its estimate) is
required at the start of the calculation. Extending the procedure to higher-order equations
is left as an exercise for the reader.
The selection of increment size remains a variable to the practicing engineer. Few numer-
ical analysis methods provided in the literature are concerned with error analysis. In gen-
eral, round-off and numerical errors appear as demonstrated earlier in Figure 2.5. In the limit,
when the increment nears 0, one approaches an analytical solution. However, the number
of calculations correspondingly increases the error Ɛ, which increases exponentially as the
increment nears 0. Note that selecting the increment size that will minimize the error is
rarely a problem in practice; in addition, computing it is also rarely a concern.
Roundoff
ε
Numerical
Increment
FIGURE 2.5
Roundoff and numerical error.
Mathematical Operations 37
∂T ∂ 2T
= α 2 (2.109)
∂t ∂z
The elliptical equation:
∂ 2T ∂ 2T
+ = 0 (2.110)
∂x 2 ∂y 2
The hyperbolic equation:
∂ 2T ∂ 2T
= α (2.111)
∂t 2 ∂x 2
The preferred numerical method of solution involves finite differencing. Only the para-
bolic and elliptical equations are considered subsequently.
2.10.1 Parabolic PDE
Examples of parabolic PDEs include
∂T ∂ 2T
= α 2 (2.112)
∂t ∂x
and (the two-dimensional)
∂T ∂ 2T ∂ 2T
= α 2 + 2 (2.113)
∂t ∂x ∂y
Ketter and Prawler,3 as well as many others, have reviewed the finite difference approach
to solving Equation 2.112. This is detailed subsequently.
38 Introduction to Optimization for Chemical and Environmental Engineers
Consider the (t, x) grid provided in Figure 2.6. The partial derivatives may be replaced by
∂T ∆T −T4 + T2 −T4 + T2
≅ = = ; ∆t = k (2.114)
∂t ∆t 2 ( ∆t ) 2k
and
∂T 2 ∆ ∆T T3 − 2T0 − T1
2 ≅ = ; ∆x = h (2.115)
∂x ∆x ∆x h2
−T4 + T2 T3 − 2T0 − T1
= (2.116)
2k h2
Solving for T2:
where r = k/h2.
Thus, one may calculate T2 if T0, T1, T3, and T4 are known. Unfortunately, stability and error
problems arise in employing the previously mentioned approach. These can be removed
by replacing the central difference term in Equation 2.114 by a forward difference term:
∂T ∆T −T0 + T2 −T0 + T2
≅ = = (2.118)
∂t ∆t ( ∆t ) k
It can be shown that the problem associated with the central difference derivative is
removed if r ≤ 0.5.
T2
t
∆t = k
T3 T0 T1 x
T4
∆x = h
FIGURE 2.6
Two variable grid.
Mathematical Operations 39
2.10.2 Elliptical PDE
For this equation, examine the grid in Figure 2.7. Using finite differences to replace the
derivatives in Equation 2.110 ultimately leads to
1
4
(T1 + T2 + T3 + T4 ) ; ∆x = ∆y (2.120)
T0 =
In effect, each T value calculated reduces to the average of its four nearest neighbors in
the square grid. This difference equation may then be written at each interior grid point,
resulting in a linear system of N equations, where N is the number of grid points. The sys-
tem can then be solved by one of several methods provided in the literature.9,10
Another solution method involves applying the Monte Carlo approach, requiring the use
of random numbers.9,10 Consider the squares shown in Figure 2.8. If the describing equa-
tion for the variation of T within the grid structure is
∂ 2T ∂ 2T
+ = 0 (2.110)
∂x 2 ∂y 2
T2
y
∆y
T3 T0 T1 x
T4
∆x
FIGURE 2.7
Elliptical grid.
y= a
7 8 9
6 5 4
1 2 3
y= 0 x
x= 0 x =a
FIGURE 2.8
Monte Carlo approach.
40 Introduction to Optimization for Chemical and Environmental Engineers
with specified boundary conditions (BCs) for T(x, y) of T(0, y), T(a, y), T(x, 0), and T(x, a), one
may employ the following approach:
This method of solution is not limited to square systems. In addition, one of the authors11
has applied this method of solution to numerous real-world applications.
References
1. S. Schaefer and L. Theodore, Probability and Statistics Applications for Environmental Science, CRC
press/Taylor & Francis Group, Boca Raton, FL, 2007.
2. B. Carnahan and J. Wilkes, Digital Computing and Numerical Methods, John Wiley & Sons,
Hoboken, NJ, 1973.
3. R. Ketter and S. Prawler, Modern Methods of Engineering Computations, McGraw‑Hill, New York
City, NY, 1973.
4. J. Reynolds, J. Jeris, and L. Theodore, Handbook of Chemical and Environmental Engineering
Calculations, John Wiley & Sons, Hoboken, NJ, 2004.
5. L. Perez, homework assignment submitted to L. Theodore, Manhattan College, Bronx, NY,
2003.
6. F. Lavery, The Perils of Differentiating Engineering Data Numerically, Chem Eng, New York
City, NY, 1979
7. L. Theodore and C. Prochaska, Introduction to Mathematical Methods for Environmental Engineers
and Scientists, Wiley‑Scrivener, Salem, MA, 2018.
8. L. Theodore, class notes, Manhattan College, Bronx, NY, 1971.
9. L. Theodore, personal notes, East Williston, NY, 1965.
10. L. Theodore, Heat Transfer for the Practicing Engineer, John Wiley & Sons, Hoboken, NJ, 2011.
11. L. Theodore, personal notes, East Williston, NY, 1985.
3
Perturbation Techniques
A significant number of optimization problems face the environmental and chemical engi-
neer. The optimal design of industrial processes, as well as process equipment, has long
been of concern to the practicing engineer, and indeed, for some, might be regarded as a
definition of the function and goal of applied engineering. The attainment of an optimum
design is generally a result of factors that include mathematical analysis, empirical infor-
mation, and both the subjective and objective experience of the environmental and/or
chemical engineer.
Regarding design, formal optimization techniques have as their goal the development of
procedures for the attainment of an optimum in a system that can be characterized math-
ematically. The mathematical characterization may be (1) partial or complete, (2) approxi-
mate or exact, and/or (3) empirical or theoretical. The resulting optimum may be a final
implementable design or a guide to practical design and a criterion for judging practical
designs. In either case, the optimization techniques should serve as an important part of
the total effort in the design of the units, structures, and control of not only equipment but
also industrial system processes.
One simple procedure for solving optimization problems that is recommended by the
authors is a perturbation scheme. This involves a systematic change of variables, one by one,
in an attempt to locate an optimum. To be practical, this can mean that one must limit
the number of variables by assuming values to those (process) variables that are known
beforehand to play an insignificant role. Reasonable guesses or simple or shortcut math-
ematical methods can further simplify the procedure. Much information can be gathered
from this type of study since it usually identifies those variables that significantly impact
the solution, such as on the overall performance of equipment or process, and it helps iden-
tify the major contributors affecting the optimization calculations.1
What does the previous paragraph mean? If the relationship between the indepen-
dent variables and the dependent variable cannot be solved graphically or analytically,
some form of trial-and-error calculation may be employed to arrive at the optimum. For
example, a sequence of values over a range of the independent variables may be calcu-
lated to completely map out the region of interest. Another scheme is to vary one factor
at a time, holding the other(s) constant. These two schemes are outlined in this chapter
in the three sections that follow. Finally, some in the technical community might malign
this approach of solving optimization problems. Certainly, linear programming, which
is addressed in the last two chapters in this Part, has its place in mathematical methods,
but so does the perturbation approach. The readers should keep this in mind when
there is a need to solve an engineering problem that involves optimization. In fact,
one of the illustrative example applications in Part III involves the application of the
Monte Carlo method,1 which some might argue requires a trial-and-error perturbation
approach.
41
42 Introduction to Optimization for Chemical and Environmental Engineers
This chapter contains three sections that address an optimization problem using the
perturbation approach when there is
y = f ( x ) (3.1)
y = 100 (1 − x ) (3.2)
2
0 ≤ x ≤ 1 (3.3)
SOLUTION:
To apply the perturbation method of solution, one could calculate y at 0.1 intervals of x
between 0 and 1. The results are provided in Table 3.1.
TABLE 3.1
Perturbation Method for
y = 100(1 −x)2
y x
100 0.0
81 0.1
64 0.2
49 0.3
36 0.4
25 0.5
16 0.6
9 0.7
4 0.8
1 0.9
0.0 1.0
Perturbation Techniques 43
There is both a maximum and a minimum over the 0.0–0.1 range of x. As expected,
the maximum occurs at
x = 0 with y = 100
f ( x ) = x 3 − 2x 2 − 5x + 6
Calculate the minimum of the function over the range 0 ≤ x ≤ 4 using a perturbation
technique.
SOLUTION:
Calculate f(x) in the range (0, 4) in increments of 0.4 for x. Results are presented in Table 3.2.
Table 3.2 suggests the minimum is located approximately at x = 2.0 with a value of 4.0.
TABLE 3.2
Calculations for Illustrative Example 3.2.
f(x) x
6.0 0.0
3.74 0.4
1.23 0.8
−1.15 1.2
−3.02 1.6
−4.0 2.0
−3.7 2.4
−1.73 2.8
2.29 3.2
8.74 3.6
18.0 4.0
11, 900
y = 162 − 2.33 x −
x
Locate the value of x that corresponds to the minimum value of y. Once again employ
the perturbation technique.
44 Introduction to Optimization for Chemical and Environmental Engineers
TABLE 3.3
Calculations for Illustrative Example 3.3
y x
60 −176.1
62 −174.4
64 −173.1
66 −172.1
68 −171.4
70 −171.1
72 −171.1
74 −171.2
76 −171.7
78 −172.3
80 −173.2
SOLUTION:
The following results if one were to neglect the last term in y:
y = 162 − 2.33 x
162
x= ≈ 70
2.33
At this value of x,
11, 900
y = 162 − 2.33 x −
70
y = 171.1
Generate values for x that straddle the value of 70 over the range 60–80 in increments
of 2. Results are provided in Table 3.3.
It appears the minimum is indeed around 70. Additional calculations could further
refine these results. For example, for x = 71, y = −171.0. This example will be revisited in
Chapter 6.
k −1
k −1
x2 k x3 k
y = f ( x2 ) = + − 2 ;
x1 x2
TABLE 3.4
Results of Illustrative Example 3.4
y x2
2 0.800
3 0.777
4 0.780
5 0.800
6 0.814
7 0.814
8 0.875
9 0.901
SOLUTION:
Set x1 = 1, x3 = 10, and k = 1.4 in the previous equation.
x 0.285 10 0.285
y ( x2 ) = 2 + − 2
1 x2
Since 1 ≤ x ≤ 10, select value of x2 in the interval (2, 9) in increments of 1.0 and calculate y.
The results are provided in Table 3.4.
The results of Table 3.4 suggest that the minimum of y is approximately 0.777 and
occurs at x2 ≅ 3.0. The reader may check this for x2 = 3.5 where y = 0.777.
y = 3 x1 + 5x2
0 ≤ x1 ≤ 3, 0 ≤ x2 ≤ 4
SOLUTION:
Proceed as illustrated below. The function can be calculated at the 20 points shown
in parenthesis in Table 3.5. The value of the Table 3.5 point locations for Illustrative
Example 3.5 are shown immediately to the right of the point in parenthesis in
Table 3.6.
As expected, the optimum y is 29 and occurs at x1 = 3 and x2 = 4, and the minimum y is
located at x1 = 0 and x2 = 0 with a minimum value of 0.
46 Introduction to Optimization for Chemical and Environmental Engineers
TABLE 3.5
Optimization Calculations for Illustrative Example 3.5
(0, 0) (1, 0) (2, 0) (3, 0)
(0, 1) (1, 1) (2, 1) (3, 1)
(0, 2) (1, 2) (2, 2) (3, 2)
(0, 3) (1, 3) (2, 3) (3, 3)
(0, 4) (1, 4) (2, 4) (3, 4)
TABLE 3.6
Function Values for Illustrative Example 3.5
(0, 0) 0 (1, 0) 3 (2, 0) 6 (3, 0) 9
(0, 1) 5 (1, 1) 8 (2, 1) 11 (3, 1) 14
(0, 2) 10 (1, 2) 13 (2, 2) 16 (3, 2) 19
(0, 3) 15 (1, 3) 18 (2, 3) 21 (3, 3) 24
(0, 4) 20 (1, 4) 23 (2, 4) 26 (3, 4) 29
SOLUTION:
The values for 20 points appear in Table 3.7.
The above function exhibits both a maximum and a minimum over the specified
range for x1 and x2. For this example, the minimum is located at (0, 0) with a value of 0,
and the maximum is located at the point (0, 4) with a value of 52.
This problem could also have been solved analytically; the solution is provided in
Chapter 6.
11, 900
y = 162 − 2.33 x1 − − 1.86 x1
x1x2
TABLE 3.7
Calculations for Illustrative Example 3.6
(0, 0) 0 (1, 0) 4 (2, 0) 10 (3, 0) 18
(0, 1) 7 (1, 1) 6 (2, 1) 7 (3, 1) 10
(0, 2) 18 (1, 2) 12 (2, 2) 8 (3, 2) 6
(0, 3) 33 (1, 3) 22 (2, 3) 13 (3, 3) 6
(0, 4) 52 (1, 4) 36 (2, 4) 22 (3, 4) 10
Perturbation Techniques 47
0 ≤ x1 ≤ 30, 0 ≤ x2 ≤ 30
SOLUTION:
The calculations for this example are presented in Table 3.8.
It appears that the maximum value of y is approximately 80.4 and is located at x1 = 10
and x2 = 30. This problem will also be revisited in Chapter 6.
SOLUTION:
Since x2 > x1, initially assume increments of 1 for x1 and x2 in the range of (2, 5) and (4, 9),
respectively. Calculated values for y for various combinations of x1 and x2 are provided
in Table 3.9.
The results in Table 3.9 suggest the minimum occurs in the neighborhood of x1 = 2.0
and x2 = 4.0. What further calculation would the reader suggest? One of the authors fol-
lowed the previous calculations with two additional calculations, the results of which
are presented here:
These two latter values for y further implies that the minimum is indeed in the range of
x1 = 2.0 and x2 = 4.0. The reader is referred to both Chapters 4, 5, 6, and 7 and Part III – Fluid
Flow, Chapters 13. The reader will soon discover that the analytical answer is
TABLE 3.8
Calculation Chart for Illustrative Example 3.7
x2 x1
0 5 10 15 20 25 30
0 −∞ −∞ −∞ −∞ −∞ −∞ −∞
5 −∞ −335.0 −117.9 −59.5 −40.8 −38.0 −43.0
10 −∞ −97.0 +1.1 +19.8 +18.7 +9.7 −3.4
15 −∞ −17.6 +40.8 +46.3 +38.5 +25.5 +9.9
20 −∞ +22.1 +60.6 +59.5 +48.5 +33.5 +16.5
25 −∞ +45.9 +72.5 +67.4 +54.4 +38.2 +20.4
30 −∞ +61.7 +80.4 +72.7 +58.4 +41.4 +23.1
48 Introduction to Optimization for Chemical and Environmental Engineers
TABLE 3.9
Calculation Chart for y(x1, x2) for Illustrative Example 3.8
x2 x1
4 5 6 7 8 9
2 0.654 0.734 0.743 0.754 0.810 0.784
3 0.671 0.742 0.742 0.748 0.798 0.766
4 − 0.768 0.762 0.765 0.810 0.775
5 − − 0.792 0.790 0.791 0.794
y = 3 x1 − 5x2 − 8 x3
Assume the same constraints as the previous two examples for x1 and x2; the constraint
for x3 is
1 ≤ x3 ≤ 2
SOLUTION:
Apply the perturbation method again as shown below. See Table 3.10 for
x3 = 1:
TABLE 3.10
Calculations for x3 = 1 for Illustrative Example 3.9
(0, 0) −8 (1, 0) −5 (2, 0) −2 (3, 0) 1
(0, 1) −13 (1, 1) −10 (2, 1) −7 (3, 1) −4
(0, 2) −18 (1, 2) −15 (2, 2) −12 (3, 2) −9
(0, 3) −23 (1, 3) −20 (2, 3) −17 (3, 3) −14
(0, 4) −28 (1, 4) −25 (2, 4) −22 (3, 4) −19
TABLE 3.11
Calculations for x3 = 2 for Illustrative Example 3.9
(0, 0) −16 (1, 0) −13 (2, 0) −10 (3, 0) −1
(0, 1) −21 (1, 1) −18 (2, 1) −15 (3, 1) −6
(0, 2) −26 (1, 2) −23 (2, 2) −20 (3, 2) −11
(0, 3) −31 (1, 3) −28 (2, 3) −25 (3, 3) −16
(0, 4) −36 (1, 4) −33 (2, 4) −30 (3, 4) −21
Perturbation Techniques 49
Based on the these two trials for x3, the maximum is located at (3, 0) and the minimum
value is located at (0, 4) with corresponding y values of 1 and −36, respectively. The
previous solution could be further refined by performing calculations at x3 = 1.25, 1.5,
and 1.75.
References
1. L. Theodore, Chemical Engineering: The Essential Reference, McGraw-Hill, New York City, NY,
2014.
2. L. Theodore, personal notes, Theodore Tutorials, East Williston, NY, 2015.
4
Search Methods
The mathematicians’ literature abounds with search methods, that is, methods that allow
one to search for either a solution to an equation or the optimum value for an equation.
There are both indirect and direct approaches. However, direct methods – those that pro-
duce an exact answer – are generally preferred by environmental and chemical engineers.
These methods have also been defined in the literature as seek methods, and usually apply
to single-independent variable unimodal equations; they provide results in intervals (or
limits) of uncertainty for the variable in a (minimum) number of trials during the search
process. The unimodal function referred to for single-independent variable equations con-
tains a single optimization (maximum or minimum) point over a specified interval/range.
Two of the procedures that are reviewed in this chapter are the interval halving (or bisec-
tion) and golden section search methods. These two methods can also be employed to
obtain the solution to some equations. For example, the interval halving method depends
on finding an approximation to a solution of the form f(x) = 0. Initial guesses x0 and x1 are
sequenced provided that f(x) is continuous for x0 ≤ x ≤ x1 and the product [( f(x0))( f(x1))] < 0.
This guarantees that the curve y = f(x) crosses the x-axis between x0 and x1. A new approxi-
mation of f(x) is then calculated at
x2 =
( x0 + x1 ) (4.1)
2
that is, f(x2). If [(f(x1))(f(x2))] < 0 then f(x) is calculated at
x3 =
( x1 + x2 ) (4.2)
2
If, however, [(f(x0))(f(x2))] < 0 then f(x) is calculated at
x3 =
( x0 + x2 ) (4.3)
2
x 3 − 9x + 1 = 0
SOLUTION:
First, calculate
f ( 2 ) = −9
51
52 Introduction to Optimization for Chemical and Environmental Engineers
f ( 4 ) = 29
The results assure that a root exists between x = 2 and x = 4. Now, set
x2 =
( x1 + x0 ) = 2 + 4 = 3
2 2
Since f(x = 3) = 1, the next approximation is determined by noting that [f(2) · f(3)] = (−9)
(1) = −9 < 0. Therefore, set
2+3
x3 = = 2.5
2
Perhaps the simplest of optimization methods that have been employed in the past by
practitioners are the interval halving and bisection methods. Unfortunately, both have
come to mean different things to different people, and to add to the confusion, both terms
have been used interchangeably. The calculations can be performed with or without the
need to evaluate the derivative of the function in question. For this text, the method involv-
ing derivatives will be referred to as the bisection method.
Some complications can arise for search methods involving multi-independent variable
equations. The most popular of these methods is the steepest ascent/descent method. This
topic is also reviewed in this chapter although there are numerous other methods that can
be employed.
Finally, it should be noted that nonlinear algebraic equations are often solved using the
Newton–Rhapson method, as described earlier and in the literature.1 However, the method
requires that the derivative of the function be available.
Several of the search methods do not require the derivative of the function in generating
a solution. It should be noted that an expression for the derivative f′(x) from f(x) can nor-
mally be derived for many problems involving simple functions. When f(x) is such that its
derivative cannot be obtained or is such that it is very complex, an approximation for f(x)
may be used. By definition,
f ( x + ∆x ) − f ( x )
f ( x ) = lim (4.4)
∆x → 0 ∆x
or
f ( x + ∆x ) − f ( x − ∆x )
f (x) ≈ (4.5)
2∆x
Four sections complement the presentation in this chapter. Search procedures are pro-
vided for
1. Compute f(x) at the midpoint x = xm of the interval, L, that is, f(xm).
L = b − a (4.6)
xm =
( a + b ) (4.7)
2
2. Set
L
x1 = a + (4.8)
4
L
x2 = b + (4.9)
4
3. Compute f(x1) and f(x2).
4. Compare f(x1) and f(xm). If (x1) ≤ f(xm), eliminate the interval xm to b from consider-
ation so that the remaining interval is a to xm. The midpoint of the new interval is
previously specified x1. Update xm in this new interval by setting
x1 = xm (4.10)
5. Calculate an updated L,
L = xm − a = x1 − a (4.11)
54 Introduction to Optimization for Chemical and Environmental Engineers
The solution for the minimum is in the new interval for L in Example 4.11. Unless
L has achieved an acceptable small value, return to step (2) and repeat the calcula-
tion for further refinement.
6. If (x2) ≤ f(xm), eliminate the interval for a to xm from consideration. The remaining
interval is xm to b. The midpoint of the new interval is x2. Update xm in this new
interval by setting
x2 = xm (4.12)
7. Calculate an updated L,
L = b − xm = b − x2 (4.13)
The solution for the minimum is in the new interval of L. Unless L has achieved
an acceptable small value, return to step (2) and repeat the calculation for further
refinement.
8. If by chance both f(x1) ≥ f(xm) and f(x2) ≤ f(xm), eliminate the interval (a, x1) and (b, x2).
The remaining (new) interval has therefore changed from (a, b) to (x1, x2) with the
same midpoint xm. The calculation then proceeds as outlined previously.
9. If by chance both f(x1) ≤ f(xm) and f(x2) ≤ f(xm), there is a fault in the mathematical
description of the minimization system since only a maximum is possible.
f ( x ) = x 3 − 3x 2 + 2
using the interval halving method provided in steps (1–9) above. Assume that 2 ≤ x ≤ 6.
SOLUTION:
Apply the algorithm provided in steps (1)–(9).
Step 1:
L =6 − 2 = 4
xm =
(2 + 6) = 4
2
Since:
f ( x ) = x 3 − 3x 2 + 2
f ( xm ) = f ( 4 ) = 64 − 48 + 2 = 18
Step 2:
2
x1 = 2 + =3
2
Search Methods 55
2
x2 = 6 − =5
2
Step 3:
f (3) = 2
f ( 5 ) = 52
Step 4: Since f(3) ≤ f(4), the remaining (surviving) interval is (2, 4). The midpoint of the
new interval is 3. The process is continued as per the suggested algorithm. The solution
will ultimately approach a value of x = 2.0 with f(x) = −2.0.
f ( x ) = x 3 − 2x 2 − 5x + 6
Obtain the minimum value of this function over the range (1, 3) by employing the inter-
val halving approach.
SOLUTION:
For this example, a = 1, b = 3, and
xm = 2.0
Set
x1 = 1.5
and
x2 = 2.5
f (1.5 ) = −2.63
f ( 2.0 ) = −4.0
f ( 2.5 ) = −3.38
Since f(1.5) ≥ f(2.0) and f(2.5) ≥ f(2.0), eliminate the interval (1, 1.5) and (2.5, 3). The
remaining interval is then (1.5, 2.5) with the same midpoint xm = 2.0. Proceed to cal-
culate f(1.75) and f(2.25). The remainder of the problem is left as an exercise for the
reader. The reduced interval will ultimately center around x = 2.1 with f(x) approxi-
mately 4.0.
56 Introduction to Optimization for Chemical and Environmental Engineers
SOLUTION:
Note once again that this method requires the calculation of the derivative of the func-
tion. For this example,
f ′ ( x ) = 3x 2 − 4x − 5
Set
1+ 3
xm = x3 = = 2.0
2
and
f ′ ( x ) = f ( x = 2 ) = −1
Since f′(x = 2.0) ≤ 0, the minimum lies to the right of x3 and the uncertainty interval is
now (2, 3). Continuing, set
2+3
x3 = = 2.5
2
and
f ′ ( 2.5 ) = 3.75
The reader is left the exercise of continuing the calculation – employing the new interval
(2, 2.5) – and showing that the minimum approaches an interval centered about x = 2.1
with f(2.1) = −4.06.
Search Methods 57
α α
= (4.14)
1 1− α
0.236 − 0
IRR = = 0.618
0.618 − 0.236
α 1−α
1 − α = α2 2α − 1 = α3 1−α
x1 = 0 x4 x3 x2 = 1.0
FIGURE 4.1
Golden section diagram – first trial.
58 Introduction to Optimization for Chemical and Environmental Engineers
α 1−α
1 − α = α2 2α − 1 = α3 1−α
α
4=
α
2α − 1 = α3 α2 − α3 2α − 1
= 2 − 3α
1 − α = α2 1 − α = α2
First interval
reduction = 1− ∝
x1 = 0 x5 x4 x3 x2 = 1.0
FIGURE 4.2
Golden section diagram – second trial.
0.236 − 0.146
IRR = = 0.618
0.146 − 0
Finally, it should be clear that the golden section search method outperforms the previous
two methods by requiring the least number of calculations for the same reduction of the
interval range.
f ( x ) = x 3 − 3x 2 + 2
SOLUTION:
First note for this example the interval is
L = 3 −1= 2
Search Methods 59
( )
x4 = 1 + α 2 L
( )
x 4 = 1 + 0.618 2 L = 1 + ( 0.382 )( 2 )
x 4 = 1.764
f ( x 4 ) = f (1.764 )
f ( x 4 ) = −1.846
Set x3.
x3 = 3 − ( 0.382 )( 2 ) = 2.236
f ( x3 ) = f ( 2.236 )
= −1.819
1
α 1−α
1−α= α2 2α − 1 = α3 1−α
α
α4 =
2α − 1 = α3 2α − 1 (First interval
α2 − α3
reduction)
= 2 − 3α
1 − α = α2 1 − α = α2
1 − α = α2 (Second interval
reduction)
α4 α5 α4 = (α2)2
x1 = 0 x6 x5 x4 x3 x2 = 1.0
FIGURE 4.3
Golden selection; subsequent interval – third trial.
60 Introduction to Optimization for Chemical and Environmental Engineers
Since f(1.764) < f(2.236), the interval of (2.236, 3) is eliminated from consideration, and the
new limits are 1 ≤ x* ≤ 2.236. Set x5 where
x5 = 1 + ( 0.618 )
( ) ( )
x5 = 1 + α 3 L = 1 + 0.236 3 ( 2) = 1.472
and proceed as noted previously. As one would suppose, the minimum will ultimately
converge at x = 2.0.
∂f ( x ) ∂f ( x ) ∂f ( x )
∇ f ( x ) = i1 + i2 + i3 +… (4.15)
∂x1 ∂x2 ∂x3
where i1, i2, i3… are unit vectors in the direction of coordinate axes representing the vari-
ables, each variable being associated with one dimension in a multidimensional space.5
The gradient ∇[f(x)] is a vector that has the property of being directed along the path
on which the function increases most rapidly (i.e., the path of “steepest ascent”). The
direction is along a path proportional to the rate of increase for a unit change of each
variable.
Summarizing, the method of steepest descent is a gradient operation where a step
size is chosen to achieve the maximum amount of decrease of the objective function
for each individual step. The method of steepest descent is an example of an iterative
algorithm in which the operation generates a sequence of points, each calculated on
the basis of the points preceding it. The operation is a decent method because the cor-
responding value of the objective value decreases in value as each new point is gen-
erated by the algorithm. The steepest descent algorithm proceeds as follows: at each
step, starting from the point. x1, a line search until a minimizer, x2, is located. A typi-
cal sequence is provided later in Figure 4.4. Note that the method of steepest descent
moves in orthogonal steps.
The following example, similar to Illustrative Examples 3.4 and 3.8, illustrates the way
the procedure works. By applying it in a case where the result is known, it is possible to
readily note its limitations. Technically, it is correct only at the base chosen for each ascent.
Additional details on this search method are presented later in the text.
k −1
k −1
k −1
x1 k x2 k x3 k
y = f ( x1 , x2 ) = + + − 3 ; x0 = 1, x3 = 10, k = 1.4
x0 x1 x2
SOLUTION:
Happel’s graphical/analytical solution4 follows where f(x1, x2) has been multiplied by 4.0.
Contours representing other values of y for values of x2 and x3 are shown in Figure 4.4.
The optimum values of x1 and x2 that minimize y will later be shown to be
x1 = 2.154
x2 = 4.642
62 Introduction to Optimization for Chemical and Environmental Engineers
10
x2 = 2.70
8 2.65
2.60
6
x2
2.59
2.58
5
2
0 1 2 3 4 5
x1
FIGURE 4.4
Contours of y; Note: The path of steepest descent is shown by a dotted line.
Assume at the start of the calculation for the steepest ascend method that x1 = 4 and
x2 = 7. The direction of steepest ascent at x1 = 4 and x2 = 7 is obtained in the following
manner:4
1− k 1 1− k 1− 2 k
∂y −
= x0 k
x
1
k
− x2 k k
x1 = 0..079
∂x1
Similarly,
∂y
= 0.009
∂x2
To obtain the path of steepest descent, increments of x must be taken proportionally to
these rates of change; thus,
∆x1 0.079
= = 8.8
∆x2 0.009
This represents the values that will produce the maximum change in the objec-
tive function. If increments of x2 are taken equal to 0.05, then the increments of
∆x1 = (0.05) ⋅ (8.8) = 0.44. In this way, the first descent is established starting at x1 = 4 and
x2 = 7, where x1 = 4.00 − 0.44 = 3.56 and x2 = 7.00 − 0.05 = 6.95 but also (Table 4.1).
Search Methods 63
TABLE 4.1
Results of Illustrative Example 4.4
x1 x2 x
4.00 7.00 2.681
3.56 6.95 2.653
3.12 6.90 2.629
2.68 6.85 2.615 (min)
2.24 6.80 2.622
References
1. R. Ketter and S. Prawel, Modern Methods of Engineering Computation, McGraw-Hill, New York
City, NY, 1969.
2. G. Reklaitis, A. Ravindran, and K. Ragsdell, Engineering Optimization, Wiley Interscience,
Hoboken, NJ, 1983.
3. H. Mickley, T. Sherwood, and C. Reed, Applied Mathematics in Chemical Engineering, McGraw-
Hill, New York City, NY, 1957.
4. J. Happel, Chemical Process Economics, John Wiley & Sons, Hoboken, NJ, 1958.
5. L. Theodore, Transport Phenomena for Engineers, Theodore Tutorials, East Williston, NY, origi-
nally published by the International Textbook Company, Scranton, PA, 1971.
5
Graphical Approaches
Engineering data is often best understood when presented in the form of graphs or math-
ematical equations. The earlier preferred method of obtaining the graphical relations
between variables was to plot the data as straight lines and use the slope-intercept method
to obtain coefficients and exponents. Therefore, it behooves the environmental and chemi-
cal engineer to be aware of the methods of obtaining straight lines on the various types of
graph paper, and of determining the equations of such lines.1 One always strives to plot
data as straight lines because of the simplicity of the curve, ease of interpolation, and ease
of extrapolation. Graphical methods proved invaluable in the past in the analysis of the
relatively complex processes of that era. Much of the basic physical and chemical data are
still best represented graphically.
One or more of the many types of graphical representations may be employed for the
following purposes:
The relation between two quantities, y – the dependent variable and x – the independent
variable, is commonly obtained as a tabulation of values of y for a number of different val-
ues of x. The relation between y and x may not be easy to visualize by studying tabulated
results and is often best seen by plotting y vs. x. If the conditions are such that y is known
to be a function of x only, the functional relation will be indicated by the fact that the
points may be represented graphically by a smooth curve. Deviations of the points from a
smooth curve can provide a measure of the reliability of the data. If y is a function of two
variables, x1 and x2, a series of results of y in terms of x1 may be obtained for a (definite)
constant x2. When plotted, the data will be represented by a family of curves, each curve
representing the relation between x1 and y for a constant value of x2. If another variable x3
is involved, one may have separate graphs for constant values of x3, each showing another
family of curves of y vs. x1. One may expand this method of representing data to more than
three independent variables, although this rarely occurs in practice. In addition, graphs
may be employed to perform common mathematical manipulations, such as integration
and differentiation, operations that usually arise in analytical analyses.
As noted previously, graphs also can be used for interpolation and extrapolation, and
to perform common mathematical manipulations, such as integration and differentiation.
The data may be of almost any type encountered in engineering and science practice.
65
66 Introduction to Optimization for Chemical and Environmental Engineers
It may be physical property data, such as the variation in density and viscosity with tem-
perature along a tube in a heat exchanger. One of the things that should be kept in mind
is the loss of accuracy arising due to the use of graphs whenever one works with plotted
data. In general, the number of significant figures may be only as great as the size of the
divisions on the graph.
Five sections complement this presentation. Section numbers and subject titles follow:
5.1 Rectangular Coordinates
Rectangular (sometimes referred to as Cartesian) coordinate graph paper is most generally
used to represent equations of the form
y = mx + b (5.1)
where:
y = variable represented on the ordinate
x = variable represented on the abscissa
m = slope of the line
b = y-intercept at x = 0
For engineering use, the most common form of graph paper is the 8 1/2 × 11-inch sheet,
having 20 lines per inch, with every fifth line accented, and every tenth line heavily
accented. This type of coordinate graph may be obtained on drawing or tracing paper,
with lines in black, orange, or green, and with or without accented and heavy lines; it is
also available in other sizes. Also note that once again the bulk of the development in the
last section of this chapter will key on rectangular coordinates.
All graphs are created by drawing a line or lines, about one or more axes. For the pur-
pose of this chapter, the presentation will be primarily concerned with graphs built around
two axes, the x-axis and the y-axis. The graph data may be presented as coordinates of the
x and y-axis in the form (x, y), or they may be presented as the abscissa (distance from the
y-axis or the x coordinate) and the ordinate (distance from the x-axis or the y coordinate).
Referring to Figure 5.1, point L5 has an abscissa (x coordinate) of +60 and an ordinate of
+40 (y coordinate). Using coordinate notation, the point can be described as (60, 40) where
the points are listed as (x, y). Point L2 can be described as (−6.7, 0).
All coordinates are produced as the result of solving an equation by assigning different
values to x or y, and solving for the value of the other. By convention, equations are usually
stated in the following form: y equals some value(s) of x and other constant(s), for example,
y = mx + b.
Graphical analysis 67
y-axis
60
50
Quadrant II Quadrant I
L5
40
(60, 40)
30
x +4
0.6
y=
20
L4
10 L3 (10, 10)
L2
x-axis
(0, 4)
0
(–6.7, 0)
–10
L1
–20 (–40, –20)
–30
Quadrant III Quadrant IV
–40
–40 –30 –20 –10 0 10 20 30 40 50 60
FIGURE 5.1
Linear coordinates.
5.1.1 Quadrants
The x and y-axes divide a graph into four quadrants. The quadrants are usually num-
bered in a counter-clockwise sequence, beginning with the upper right quadrant (see
Figure 5.1). Note that in quadrants I and IV, the x values are positive while in quadrants
II and III, they are negative. The y values are positive in quadrants I and II and negative
in quadrants III and IV. The quadrant identifications are only of value when information
is provided that a point or result lies in a certain quadrant. Without knowledge of the
precise values of x or y one can still determine whether the x and y values are positive or
negative.
y-axis
90
80 N5 (5, 78.5)
70
60
N4
50 y = π x2
(4, 50.3)
40
30 N3
(3, 28.3)
20
N2
(2, 12.6)
10
N1
N0
x-axis
(1, 3.1)
0
(0, –0)
–10
–3 –2 –1 0 1 2 3 4 5 6 7
FIGURE 5.2
Nonlinear equation.
5.1.3 Intercepts
Reference to the x and y intercepts are often made when working with graphs. An inter-
cept is the point at which a line crosses the x- or y-axis. The x-intercept is the point where a
line crosses the x-axis (y = 0), and the y-intercept is the point where a line crosses the y-axis
(x = 0). In Figure 5.1, the x-intercept occurs are (−6.7, 0) and the y-intercept occurs at (0, 4).
There can be only one x- and one y-intercept in a linear equation. With nonlinear equa-
tions, there may be no x- or y-intercept, or there may be multiple intercepts. Note that in
Figure 5.3 there are two x-intercepts [(−2, 0) and (0, 0)] and one y-intercept (0, 0).
y-axis
16
(–5, 15) (3, 15)
14
12
x + x2
10
y=2
(–4, 8) (2, 8)
8
4
(–3, 3)
(1, 3)
2
(–2, 0)
x-axis
0
(0, 0)
(–1, –1)
–2
–4
–10 –8 –6 –4 –2 0 2 4 6 8 10
x
FIGURE 5.3
Intercept analysis.
1 2 3 4 5 6 7 8 9 10
10
7
Dependent variable, y
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Log of a dependent variable, x
FIGURE 5.4
The construction of a logarithmic scale.
70 Introduction to Optimization for Chemical and Environmental Engineers
of the ordinate and abscissa; for example, logarithmic 2 × 3 cycles means two cycles on the
ordinate and three on the abscissa. Cycle lengths are generally the same on both axes of a
particular graph. In addition, the distance between numbers differing by a factor of 10 is
constant on a logarithmic scale.
Equations of the general form
y = bx m (5.2)
where:
y = a variable
x = a variable
b = a constant
m = a constant
will plot as straight lines on (full) logarithmic paper. A form of the equation analogous
to the slope-intercept equation for a straight line is obtained by writing and then subse-
quently plotting the equation in logarithmic form, as represented in Equation 5.3.
If log y versus log x were plotted on rectangular coordinates, a straight line of slope m and
y-intercept log b would be obtained. An expression for the slope is obtained by differenti-
ating Equation 5.3.
d (log y )
= m (5.4)
d (log x)
y = 2x1.3 (5.6)
on rectangular coordinate paper is given in Figure 5.5. It should be noted that although
the slope was obtained from a ratio of logarithmic differences, the same result could have
been obtained using measured differences if the scale of ordinate and abscissa were the
same (in this case they were).
Inspection of Figure 5.5 reveals that equal measured distances on the logarithmic scale
are equivalent to equal logarithmic differences on the abscissa of that plot. It therefore fol-
lows that since the slope of Equation 5.3 is a ratio of coordinate differences, the slope on
full logarithmic paper could be obtained using measured distances. The plot of Equation
5.6 on 2 × 2 cycle logarithmic paper is given in Figure 5.6. The slope using logarithmic dif-
ferences is
1.08 − 0.43
m= = 1.3
0.60 − 0.10
Graphical analysis 71
1.08
1.0
log10 y 0.43
0.65
m= = 1.3
0.5
0.43
0.5
0.302
0.1 0.6
0
log10 x
FIGURE 5.5
A linear plot of y = 2x1.3 on rectilinear graph paper.
100
80
60
40
20
a
10
y
8
6
4
b
1
1 2 4 6 8 10 20 40 60 80 100
x
FIGURE 5.6
The equation y = 2x1.3 plotted on logarithmic graph paper.
To further illustrate the use of full logarithmic paper, consider the determination of a
mathematical equation to represent data that are known to plot as a straight line on log-log
paper. If data, as provided in Table 5.1, are to be plotted, it is evident that more than one
cycle will be needed in both directions. For x, the data covers the range from
TABLE 5.1
Log-Log x-y Data
x y
2.0 3.6
3.9 6.0
7.0 9.4
14.0 16.0
26.5 26.0
62.0 50.0
In both cases, two cycles are needed to cover the ranges from 100 to 101 and from 101 to 102.
A general rule would be that the number of cycles is equal to the number of exponents of
ten involved in the range of data, in this case two for each coordinate. The data is plotted
in Figure 5.7. The slope may once again be calculated by:
100
80
60
40
20
c
10
y
8
6
d
2
1
1 2 4 6 8 10 20 40 60 80 100
x
FIGURE 5.7
Determination of a mathematical equation from raw data.
Graphical analysis 73
In cases where the line x = 1 is several cycles removed from the data range, it may be incon-
venient to extrapolate; the intercept can then be determined by substituting the value of
the slope and one data point and solving for b.
y = ne mx (5.10)
or
y = n10 mx (5.11)
where:
y = dependent variable
x = independent variable
n = a constant
m = a constant
e = the natural logarithm base
Here again, a form of the equation analogous to the slope-intercept equation can be
obtained by placing the previous equations in logarithmic form. For example, Equation
5.10 may be written as
log y = b + mx (5.12)
where b = log n
74 Introduction to Optimization for Chemical and Environmental Engineers
An expression for slope is obtained from the differential form of Equation 5.12:
d (log y )
= m (5.13)
dx
x
log y = 2.2 + (5.14)
46.3
The line may be constructed (see also Figure 5.8) by choosing values of x, solving
for y, and plotting the results. Note that values of y and not log y are plotted on the
ordinate. The slope is a logarithmic difference divided by an arithmetic difference
and in this case may be calculated from the two indicated points (see Figure 5.8) as
follows:
The slope of a line on semi-logarithmic coordinates may be calculated from any two
points, but the simplest method is the one used previously in which the slope is the
5000
y
1000
500
100
0 10 20 30 40 50 60 70 80
x
FIGURE 5.8
x
The equation log y = 2.2 + plotted on semi-logarithmic paper.
46.3
Graphical analysis 75
logarithmic difference on one cycle (equal to log 10 or 1.0) divided by the arithmetic dif-
ference of one cycle. Note that if the cycle had been taken between y = 300 and y = 3000,
the slope would be
It should also be noted that the y-intercept is 2.2, that is, the of log 159.
The method of obtaining a mathematical equation from data that are known to plot a
straight line on semi-logarithmic paper may be illustrated by the use of the data presented
in Table 5.2. The data is plotted in Figure 5.9. These cycles on the logarithmic scale were
necessary because values of y ranged from
TABLE 5.2
Semi-Log-Log x-y Data
x y
0.93 432
1.52 886
2.56 2,830
3.20 6,000
3.80 12,000
4.50 27,000
10,000
y
1,000
100
0 1 2 3 4 5
x
FIGURE 5.9
The determination of a mathematical equation from raw data.
76 Introduction to Optimization for Chemical and Environmental Engineers
and three exponents of ten are involved, for example, 2, 3, and 4. The slope may be calcu-
lated from the middle cycle as follows:
1 1
m= =
3.65 − 1.65 2
The line is extrapolated to x = 0 to give a y-intercept of 150; hence, the equation represent-
ing the data is
x
log y = log 150 + (5.16)
2
or
y = 150 ( 10 )
x /2
(5.17)
5.5 Applications
This last section contains four illustrative examples. Some are extensions of problems dis-
cussed earlier with the material keying on solutions employing graphical approaches.
where:
0.429x
A(x) =
x + 15
and
B ( x ) = 0.006 x
SOLUTION:
The describing equation for f(x) is
0.429x
y (x) = f (x) = − 0.006 x
x + 15
Refer to Figure 5.10. It can be seen that f(x) = 0 at approximately x = 56.6.
0.5
0.4
CP
0.3
RV
0.2
y
Profit
0.1
Break-even
0
–0.1
–0.2
0 10 20 30 40 50 60 70 80
x
FIGURE 5.10
x-y plot from Illustrative Example 5.1.
78 Introduction to Optimization for Chemical and Environmental Engineers
SOLUTION:
Refer to Figure 5.10 once again. The maximum value of f(x) is 0.75 and occurs at x
approximately equal to x = 15.0. This problem will be revised in next chapter and an
exact analytical solution to the problem will be obtained.
11, 900
y = f ( x ) = 162 − 2.33 x −
x
was solved earlier in Chapter 3 (see also Illustrative Example 3.3) using a perturbation
approach. Recalculate the minimum value of f(x) using a graphical approach.
SOLUTION:
The following y vs. x values were provided in Illustrative Example 3.3. These calcula-
tions in Table 5.4 can be plotted. That is left as an exercise for the reader. Hint: The
solution appears here:
x = 71
f ( 71) = −171.0
As expected, the results compare favorably with the answer provided earlier. The reader
should also note that this problem will be solved analytically in the next chapter.
11, 900
y = 2.33 x1 + + 1.86 x2 + 10
x1x2
Determine the values of x1 and x2, which will produce the minimum value of y in the
above equation.
TABLE 5.4
y vs. x value for Illustrative Example 5.3
y x
60 −176.1
62 −174.4
64 −173.1
66 −172.1
68 −171.4
70 −171.1
72 −171.1
74 −171.2
76 −171.7
78 −172.3
80 −173.3
Graphical analysis 79
SOLUTION:
When two or more independent variables are to be optimized, the procedure for deter-
mining the optimum conditions may be more difficult than when one independent
variable is involved. For this case, the relationship among y, x1, and x2, could be shown
as a curved surface in a three-dimensional plot, with a minimum value of y occurring
at the optimum values of x1 and x2 However, the use of a three-dimensional plot may
not be practical for some engineering calculations. The optimum values of x1 and x2
can be found graphically on a two-dimensional plot by the method that is provided in
Figure 5.11.
For the equation in question, f(x) is plotted against one of the independent variables x1,
with the second variable x2 held at a constant value. A series of such plots can be made
with each dashed curve representing a different constant value of the second variables.
As shown in Figure 5.11, each of the curves (a, b, c, d, and e) gives one value of the vari-
able x1 at the point where the f(x) or f(x1, x2) is a minimum. The curve AB represents the
locus of all these minimum points and the optimum value of x1 and x2 occurs at the
minimum point on the dashed curve AB – noted as (x1* , x2* ) at y*.
The following constant values of x2 are chosen arbitrarily for the example at hand:
= =
xI 32 =
xII 26 =
xIII 20 xIV 15 xV = 12
Plots of y vs. x1 at each constant value of x2, that is, xI, xII, xIII, xIV, and xV are generated.
These plots are presented as solid lines in Figure 5.12 as curves a, b, c, d, and e. A sum-
mary of these results is presented in Table 5.5 with x2*, and x1* representing the mini-
mum values of x2 for the solid curves a, b, c, d, and e, respectively. The dashed curve
AB represents the relationship for the minimum five x1*’s as a function of y. The point
x1* represents the minimum y, which occurs at approximately x2 III and x1*III ; the corre-
sponding value of y is 121.6.
A B
y = f(x)
FIGURE 5.11
Optimum value of the function of (y) in terms of two independent variables x1 and x2.
80 Introduction to Optimization for Chemical and Environmental Engineers
A B
y = f(x)
x2II = 26 x2IV = 15
(b) (d)
FIGURE 5.12
Optimum value function of (y) in terms of two independent variables (x1, x2).
TABLE 5.5
Optimal x2, and x1* Results for Illustrative Example 5.4
The reader should also note that similar graphical procedures can be used when
there are more than two independent variables, for example, x1, x 2, and, x3, where
y = f(x1, x 2, x3). For this case, and as noted earlier, the first step would be to make
a plot similar to Figure 5.12 at one constant value of x3. Similar plots would then
be made at other constant values of x3. Each plot would give an optimum value of
x1, x 2, and y for a particular x3. Finally, the overall optimum value of x1, x 2, and y
could be obtained by plotting of x3 versus the individual optimum value of y, where
y = f(x) = f(x1, x 2, x3).
Graphical analysis 81
References
1. L. Theodore and F. Ricci, Mass Transfer Operations for the Practicing Engineer, John Wiley & Sons,
Hoboken, NJ, 2010.
2. L. Theodore, Air Pollution Control Equipment Calculations, John Wiley & Sons, Hoboken, NJ,
2008.
3. S. Shaefer and L. Theodore, Probability and Statistics Applications for Environmental Science, CRC
Press/Taylor & Francis Group, Boca Raton, FL, 2007.
6
Analytical Methods
The subject of analytical methods serves as an excellent starting point before preceding to
the last subject in this first Part—linear programming. In terms of introduction, once again
consider the following. Since the derivative dx/dy or f′(x), represents the rate of change of
y with respect to the change in x, it is evident that the derivative will be zero if the func-
tion passes through a maximum or minimum. If the occurrence of such a maximum or
minimum is to be determined and located, the derivative must be equated to zero, thus
providing the condition for which the maximum or minimum exists. This analytical pro-
cedure is of considerable value in environmental and chemical engineering calculations
since the location of a maximum or minimum is frequently of practical importance. As one
might suppose, many of the applications are related to economics, for example, maximiz-
ing profit and/or minimizing cost.
Five sections complement this presentation. Section numbers and titles follow:
6.1 Breakeven Considerations
In solving many environmental and chemical engineering optimization problems, one
often selects optimum conditions from the standpoint of economics. Consider, a simple
case in which there is only one independent variable affecting profit P, that is, income I
minus cost C. For example, refer to Figure 6.1. The term y is the dependent variable that
will later be referred to as the objective function and x is the independent variable. The
former term could represent the aforementioned profit and x could be the operating tem-
perature. As can be seen from Figure 6.1, the “system” operates at a loss when operating
below x* and operates at a profit when above x*. The term x* is located at the breakeven point,
that is where the profit is zero.
The above breakeven problem can be solved by three methods presented below:1
Method 1: Solve the problem analytically as follows. Assume I and C can be repre-
sented in equation form, that is,
I = I ( x ) (6.1)
83
84 Introduction to Optimization for Chemical and Environmental Engineers
C = C ( x ) (6.2)
Since
P = I – C (6.3)
the solution for x, when P = 0, provides the breakeven point (of operation).
Method 2: Select various values of x and calculate both the income and cost, and then
the profit. The breakeven point arises when the profit is zero.
Method 3: Graph the results.2 A typical solution is presented in Figure 6.1. See also the
previous chapter.
Economic studies often reveal that certain cost factors increase while others decrease
relative to an independent variable. Now relate this concept to an LNG (liquefied nitro-
gen gas) plant. Liquefied nitrogen gas is used by gas utility companies around the coun-
try to heat residential and commercial establishments. This gas has an extremely low
boiling point and must be kept at a temperature of −260°F to be maintained in a liqui-
dated state. LNG is most easily and safely transported in its liquid form. It is odorless
and colorless in its gaseous state, making it extremely difficult to spot a leak if stored as
a gas. The gas companies transport this liquefied LNG gas from trucks and other sources
and store it in large tanks until it is converted to a gas for customer use. Consider the
following LNG example:
The safety engineer employed at Theohan Gas Company was inspecting the LNG
transport pipes when she noticed the insulation on one of the pipes looked damaged.
After removing a section of the insulation, it was clear it needed to be replaced and the
engineer proposed replacing it with a thicker insulating cover to prevent/reduce leakage.
Looking at this from an economic standpoint, as the thickness of the insulating cover
(x) on an LNG pipe increases, the cost increases, but correspondingly the LNG lost in
travel (the LNG transportation cost) decreases. Here the thickness of the cover serves as
the independent variable, and if the installed cost of the insulation and the LNG transpor-
tation leakage losses could both be determined as functions of the independent variable,
(I, x)
Income, I
(P, x)
Profit, P x
x*
Cost, C
(C, x)
FIGURE 6.1
Breakeven figure.
Analytical Methods 85
a condition would exist for which the total cost is a minimum. This solution is pictured in
Figure 6.2 and the income optimum is located at the square point, x*. Here, the total cost
is plotted against the insulating cover thickness. This problem will be revisited in Part III,
Section 16.2, Illustrative Example 16.2.
y = f ( x ) (6.4)
dy
= 0 (6.5)
dx
Fixed charges
Total cost
Economic
optimum
x*
Insulation thickness, x
FIGURE 6.2
Total cost vs. insulating cover thickness.
86 Introduction to Optimization for Chemical and Environmental Engineers
The optimum y is obtained by substituting the value of x from Equation 6.5 with Equation
6.6. And, y is a maximum if
d2 y
< 0 (6.6)
dx 2
When x from Equation 6.5 is substituted with Equation 6.4 the calculated y is a minimum if
d2 y
> 0 (6.7)
dx 2
11, 900
y = 162 – 2.33x +
x
Determine the maximum (or minimum) of the above function by analytical means.
SOLUTION:
Apply Equation 6.5.
dy 11, 900
= −2.33 − =0
dx x2
Solving for x,
11, 900
x2 = = 5, 107
2.33
x = 71.5
The second derivative d2y/dx2 is negative since
dy 2 23, 800
2
=−
dx x3
The calculated value of y at x = 71.5 is therefore a maximum with a value of
11, 900
y = 162 – ( 2.33 )( 71.5 ) +
71.5
= 162 – 166.6 + 166.4
= 161.8
SOLUTION:
Once again, the describing equation is
0.429x
y= − 0.006 x
x + 15.0
To proceed analytically, calculate dy/dx and set it equal to zero.
dy 0.429 0.429x
= − − 0.006 = 0
dx x + 15.0 ( x + 15.0)2
Solve for x by any suitable method.
0.429x
0.429 − − 0.006 ( x + 15.0 ) = 0
x + 15.0
x = 17.5 psf
Since dy2/dx2 < 0, the corresponding maximum profit is
( 0.429 )(17.5 ) −
y= ( 0.006 )(17.5 )
17.5 + 15.0
= 0.231 − 0.105
= 0.126
The above analytical result compares favorably with the results provided earlier in
Illustrative Example 5.1.
SOLUTION:
As noted in Illustrative Example 3.2 the total work is given by
k −1
k −1
x2 k x3 k
y = f ( x2 ) + − 2
x1 x2
If this quantity is to be a minimum, then the derivative must be zero, that is,
dy k − 1 k − k k − 1 k k
1− k 1 1− k 1− 2 k
= x
1 x 2 − x
3 x2 =0
dx2 k k
Solving for x2 yields
2k − 2 1− k 1− k
x2 2k
= x1 k
x3 k
or
k −1 1− k 1− k
2
x2 k
= x1 k
x3 k
88 Introduction to Optimization for Chemical and Environmental Engineers
Thus,
x22 = x1x3
x2 = ( x1x3 )
1/2
1/2
x2 = (1)(10 ) = 3.16
dy 2
Since < 0 the corresponding minimum value of y is
dx22
y = f ( x2 = 3.162 ) = 2.726
This result compares favorably with the answers provided earlier in Illustrative
Examples 3.4.
Interestingly, Happel2 addressed the problem of maximizing the work requirement
for a three-stage compressor operating between 1 and 10 atm. This problem was solved
using an analytical approach along with the method of steepest ascent2 (see also the
previous development) in Chapter 4, Illustrative Example 4.6, and in Part III, Chapter
13, Illustrative Example 13.6. The analytical method for the two intermediate pressures
P2 and P3 yielded the equations (see Illustrative Example 6.7 for details):
( )
1/3
P2 = P12 P4
(
P3 = P42 P1 )
and
E = 2.575 ft ⋅ lbf
∂y ∂y
= = 0 (6.9)
∂x1 ∂x2
with
∂2y ∂2y
2
< 0; 2 < 0 (6.10)
∂x1 ∂x2
and
2
∂ 2 y d2 y ∂ 2 y
> (6.11)
∂x12 dx22 ∂x1∂x2
Similarly, if ∂ 2 y / ∂x12 and ∂ 2 y / ∂x22 are positive and Equation 6.11 holds, there will be a
minimum. But, if
2
∂2y ∂2y ∂2y
> (6.12)
∂x12 ∂x22 ∂x1∂x2
then the point concerned will be a saddle point. The case where
2
∂2y ∂2y ∂2y
= (6.13)
∂x12 ∂x22 ∂x1∂x2
is open and the value may be a maximum, a minimum, or neither.
Determine if the function has a maximum or a minimum. Also, calculate that value.
SOLUTION:
For the above equation
∂f
= 2x1 − 2 + x2 = 0
∂x1
∂f
= 2x2 + x1 = 0
∂x2
The following solution results, when the above two equations are set equal to zero.
4
x1 = −
3
2
x2 =
3
Since
∂2 f
=2
∂x12
∂2 f
=2
∂x22
the solution provided for x1 and x2 represent a minimum with
y = f ( x1 , x2 ) = 5
11, 900
P = 162 − 2.33 x1 − 1.86 x2 − =0
x12 x2
Calculate values of x1 and x2 that would either maximize or minimize P.
SOLUTION:
Begin by generating the first derivative and setting them equal to zero.
∂P 11, 900
= −2.33 + 2 =0
∂x1 x1 x2
∂P 11, 900
= −1.86 + =0
∂x2 x1x2 2
Solving simultaneously yields
x1 = 16
x2 = 20
Analytical Methods 91
The corresponding P (16,20) is a maximum since both second derivatives are negative.
Thus,
P (16, 20 ) = 50.3
11, 900
E = 2.33T + + 1.86P + 10
TP
SOLUTION:
Begin once again by presenting the first derivatives and setting them equal to zero.
∂E 11, 900
= 2.33 − =0
∂T T 2P
∂E 11, 900
= 1.86 − =0
∂P TP 2
For this example, the simultaneous solution to the above three equations yields
T = 16
T = 20
Furthermore,
E = E (16, 20 ) = 121.6
Since
∂ 2E ( 2 )(11, 900 )
= =+
∂T 2 (16)2 ( 20 )
∂ 2E ( 2 )(11, 900 )
= =+
∂P 2 (16 ) (20)2
the optimum conditions represent a minimum emission. The interested reader should
compare the results of this illustrative example with that in the previous illustrative
example.
SOLUTION:
The describing equation is once again
k −1
k −1
k −1
x2 k x3 k x 4 k
y = + + − 3 ; x1 = 1, x 4 = 10, k = 1.4
x1 x2 x3
1− k 1 1− k 1− 2 k
∂y −
= x1 k
x2 k − x3 k
x2 k
∂x 2
1− k 1 1− k 1− 2 k
∂y −
= x2 k
x3 k − x 4 k
x3 k
∂x 3
Setting ∂y/∂x2 and ∂y/∂x3 = 0, ultimately leads to
so that
( ) ( )
1/2 1/2
x2 = x12 x 4 x2 = x 4 2 x1
x2 = 2.154
x3 = 4.642
and
One can further show that (see also Part III Chapter 13)
∂2y
=+
∂x2 2
∂2y
=+
∂x3 2
∂2y
=+
∂x2∂x3
and
P = 8 x + 6 y + 8 z + xy + xz + yz − 2x 2 − 3 y 2 − 4 z 2
SOLUTION:
∂P
= 8 + y + z − 4x = 0
∂x
∂P
= 6 + x + z − 6y = 0
∂y
∂P
= 8 + y + x − 8z = 0
∂z
for which
P = 78.07
∂ 2P
= −4
∂x 2
∂ 2P
= −6
∂y 2
∂P
= −8
∂z 2
Since the three derivatives are negative, the value of P of 78.07 is a maximum; that is,
any other calculated P for a different x, y, and z would produce a value of P less than
78.07. The “proof” is left as an exercise for the reader. However, if one were to select
values of x, y, and z straddling the maximum set of values, P should be below 78.07.
For example, the value of P for x = 2, y = 2, and z = 2 is −20. For x = y = z = 10, the value of
P is −380.
94 Introduction to Optimization for Chemical and Environmental Engineers
References
1. L. Theodore, Chemical Engineering: The Essential Reference, McGraw-Hill, New York City, NY,
2014.
2. J. Happel, Chemical Process Economics, John Wiley & Sons, Hoboken, NJ, 1958.
7
Linear Programming
Linear programming is one of several techniques for optimization that has increased
in usefulness because of the availability of digital computers. The title results from the
assumptions in the method that linear relationships describe the problem/system/rela-
tionship under consideration. For example, each production element increase per day in a
process could produce a corresponding linear profit increase per day. However, there are
many instances where a linear model does not accurately describe the totality of the prob-
lem/system/relationship. It may then be necessary to resort to other procedures such as
nonlinear programming; some brief details will be presented for handling the nonlinear
situation. The definitions associated with some of these terms appear earlier in this Part.
The reader should note that the terms linear programming and nonlinear programming
are essentially similar from an applications perspective. There were problems around the
middle of the last century because of some difficulties that arose in attempting to solve
nonlinear programming problems. However, the arrival of the modern-day computer (see
Chapter 1) and sophisticated software (e.g., Excel) have removed these problems. Unless a
solution is presented graphically or algebraically, it will be obtained directly from Excel so
that the reader need not be concerned with whether a system’s pertinent describing equa-
tions and constraints are linear or nonlinear. Thus, details regarding the solution method-
ology for both linear programming and nonlinear programming problems as well as for
some of the illustrative examples will not be presented. They are simply not necessary and
beyond the scope of this introductory optimization text. However, illustrative examples
involving traditional optimization will address both the development of the describing
equations, that is, the model, and the solution to the problem.
Other than linear programming, dynamic programming is another often used optimiza-
tion method by the practicing environmental and chemical engineer. As described earlier,
it is a process involving decisions made at multiple stages. Just as an optimization model is
defined by its variables, constraints, and objective function, a multistage decision process
is defined by its stages and decisions. At any stage, there is often enough information about
decisions made at preceding stages to make informed decisions at the present stage. Thus,
some optimization problems can be reformulated as multistage decision process problems.
One then proceeds through subsequent stages until the final stage of the process is reached.
Some additional details are provided in the next section.
As one might suppose, there are numerous other optimization methods available to the prac-
ticing environmental engineer and scientist. Several of these are detailed in Perry’s Handbook:1
1. Mixed integer
2. Mixed integer linear programming
3. Mixed integer nonlinear programming
4. Quadrative programming
5. Differential/nondifferentiable systems
6. Convex/nonconvex
95
96 Introduction to Optimization for Chemical and Environmental Engineers
The previously mentioned topics are not reviewed in this text. The reader is referred to
the optimization literature2–6 for details.
Five sections compliment the presentation for this chapter:
7.1 Definitions
7.2 Basic concepts of optimization
7.3 Applied mathematical concepts in linear programming
7.4 Applied engineering concepts in linear programming
7.5 Other real-world applied concepts in linear programming
7.1 Definitions
Seven key definitions – as they apply to optimization and linear programming are pro-
vided below (in alphabetical order).
1.
Algebra: A branch of mathematics in which letters are used to represent basic
arithmetic relations. As in arithmetic, the basic operations of algebra are addition,
subtraction, multiplication, and division. Arithmetic, however, cannot generalize
such mathematical relations as the Pythagorean theorem (which states that the
sum of the squares of the sides of any right triangle is also a square). Arithmetic
can produce specific instances of this relations (e.g., 3, 4, and 5, where 32 + 42 = 52).
Algebra can make a purely general statement that fulfills the conditions of the
theorem, that is, a2 + b2 = c2. Any number multiplied by itself is termed squared
and is indicated by a superscript number 2. For example, 3 × 3 is noted as 32 and is
termed three squared; similarly, a × a is equivalent to a2. Classical algebra, which is
concerned with solving equations, uses symbols instead of specific numbers and
employs arithmetic operations to establish procedures for manipulating symbols,
while modern algebra has evolved from classical algebra by placing more atten-
tion to the structures within mathematics.
2.
Constraint: “A constraining or being constrained…confinement or restrictions…
forced…something that constrains.” The term constraint is also defined as “to bond
together, draw together…to force into or hold in, close bounds, confine…restrain.”
3.
Dynamic programming: The dimensionality of a particular operation is reduced
from (m)(n) variables to m repeated n times. Thus, if it is not practical or possible
to solve a large complex system of equations, the system is reduced in complexity
in a manner that requires the solution of a smaller subset of the original problem,
noting that for each n (e.g., each stage in a mass transfer operation)7, 8 only m vari-
ables are involved in the optimization process. One can then work backward and
combine the solutions generated at each n.
4.
Linear algebra: A branch of mathematics that is concerned with systems of linear
equations, linear transformations, vectors, vector spaces, and related topics. The
major interest is with linear equations.
5.
Linear programming: A mathematical and operations research technique used
in engineering, science, administrative engineering, and economic planning to
Linear Programming 97
f ( x ) = x 6 − 4 x + 6; − 3 ≤ x ≤ 6 (7.2)
is a constrained function. When constraints are present, the function being optimized is
usually referred to as the objective function to distinguish it from any other functions that
may be used to define the constraints or the problem. Although the emphasis to follow is
on linear constraints, many real-world applications also involve nonlinear constraints and
nonlinear objective functions. To further complicate the problem, the constraints can be of
an equality or monoequality/inequality nature.
From a basic perspective, the problem could be to optimize the objective function
y = f ( x1 , x2 ) (7.3)
g ( x1 , x2 ) < 0 (7.4)
x1 ≥ 0 (7.5)
x2 ≥ 0 (7.6)
This is a simple yet typical problem in optimization – the maximization or minimization of
a real-value function of a number of real variables (occasionally just a single variable) subject
to a number of constraints (occasionally the number is zero). However, the mathematical
model normally contains several (n, e.g.) variables known as the independent design variables
denoted as x1, x2, x3, … xn. This set of variables has been described by some as a vector, x. One
may view each x as one dimension in the set (space) of variables, and any particular set of
variables, and any particular set of values for these variables in this system as a solution.
One of the more important recent areas for the application of mathematical optimiza-
tion techniques is in engineering and science. Many problems require an iterative and/or
trial-and-error solution and the problem of concern is often stated in an ambiguous and
open-ended way. The first task is to decide what the problem is and to identify the require-
ments or specifications for any proposed and/or feasible solution. One then generates a
concept, perhaps in the form of a rough configuration, for the object, system, or process to
be optimized. This phase of the problem is usually the most difficult and can require the
greatest ingenuity and engineering judgment.
The next step attempts to model mathematically the object, system, or process. This model
may assume a wide variety of forms, but the usually preferred ones are mathematical in
nature. One may then use many mathematical, scientific, and engineering tools. These can
involve not only the most empirical of data correlation but also the most rigorous math-
ematical procedures to the least formal cut-and-try methods for assisting in the analysis of
this aspect of the solution.
As noted above, it may happen that certain functional equality constraints also apply, for
example,
f1 ( x1 , x2 , x3 ,… , xn ) = 0
(7.7)
f k ( x1 , x2 , x3 ,… , xn ) = 0
There may also be certain inequality constraints such as
Linear Programming 99
g1 ( x1 , x2 , x3 ,… , xn ) < 0
(7.8)
g m ( x1 , x2 , x3 ,… , xn ) < 0
that are specified on the space so that only points in a portion of the total space may be
considered as acceptable values for the solution of the problem. This region is defined as
the feasible space and points in the region are defined as feasible solutions and/or values.
The previously mentioned equality constraints – Equation 7.7 – come from functional
relationships that must be satisfied among the aforementioned n variables. If there is to
be a choice in the values for which at least some of the variables may be assumed, then k
must be smaller than n. The inequality constraints in Equation 7.8 usually arise from some
specified limitations (maximum permissible stress, minimum emissions, minimum allow-
able temperatures, etc.). Note that there is no upper limit on the value of m. Additional
information is provided in Section 7.4.
Summarizing, the optimization problem to be considered is that of minimizing (mini-
mizing should be italicized) f(x) subject to certain constraints. The term f(x), is a real-valued
function called the objective function (or cost function). A vector x is an n-vector of indepen-
dent variables x1, x2, …, xn. These variables are often referred to as decision variables. The
problems require finding the “best” values of the vector x over all the possible values of x.
The “best” value refers to the one that results in the smallest value of the aforementioned
objective function. Optimization problems can also involve maximization of the objective
function. This can be represented similarly to the minimization presentation above since
maximizing f(x) is equivalent to minimizing f(x). Therefore, one can apply the presented to
the minimization problem without concern.
Finally, the following comments are provided on the objective function. It may be quite
simple and easy to determine, or complicated and difficult to calculate. The objective func-
tion may also be very elusive due to the presence of conflicting or dimensionally incompat-
ible sub-objectives; for example, one might be asked to operate an absorber7 at minimum
cost and that it is aesthetically appealing. In some cases, it may not be possible to find a
quantitative objective function. In this case, one can assume that an alternative suitable
objective function can be formulated.
E = a1x1 + a2 x2 (7.9)
100 Introduction to Optimization for Chemical and Environmental Engineers
C ≤ c1x1 + c2 x2 (7.11)
In addition,
x1 ≥ 0 (7.12)
x2 ≥ 0 (7.13)
In terms of a solution, first examine the previous system of equations analytically. The
problem can be viewed in Figures (7.1) and (7.2). First, note that only the first quadrant need
be considered since Equations 7.12 and 7.13 must hold. Equations 7.10 and 7.11 are rewritten
in the following form:
B b1
x2 ≤ − x1 (7.14)
b2 b2
and
C c1
x2 ≤ − x1 (7.15)
c2 c2
x2 x2
Equation Equation
(7.14) (7.15)
x2
x1 x3
(a) (b)
x1
x2
(c)
FIGURE 7.1
Two-variable minimization problems.
Linear Programming 101
x2
E 1 =a 1
E*
=
x2
E2
a 1 x 1+
=
E
x1
=
+
3
a
a2 x2
1x
x2
a2
1 +
a
2x
2
x2
x1
(a)
x2
E1
E 2 E*
E3
=
x2
x1
(b)
FIGURE 7.2
Solution to a two-variable minimization problem.
It is apparent that the only values of x1 and x 2 that will satisfy Equations 7.14 and 7.15
lie above the solid lines (the cross-hashed area) in Figure 7.1. This shaded area is termed
the constraint set. Now place in Figure 7.2 a series of straight lines (or contours) corre-
sponding to Equation 7.9 with a different value of E used for each line (see Figure 7.2a).
Of all these lines, the smallest E for which x1 and x 2 remain in the specified region will
be that minimum E (see Figure 7.2b), i.e., E3. This will yield that proportion of x1 and
x 2 that minimizes the objective function. Stated in other words, one desires the lowest-
value contour having some point in common with the constraint set. For the particular
value of slope chosen for Equation 7.9, the minimum point is seen to correspond to E
where
E3 = Es* (7.16)
The point that represents the minimum is therefore determined by the constraints a1
and a2 of the problem. Further, if the constants c1 and c2 in Equation 7.11 are changed, it
is possible that the minimum point would occur elsewhere. The situation is essentially
reserved for a maximization problem, that is, for a profit P. This is demonstrated in
Figure (7.3).
It should be pointed out that if the objective function – Equation 7.9 – was nonlinear,
the contour lines would be curved, and if the constraint equations were nonlinear, the
constraint lines would not be straight. It should be intuitively obvious that locating the
minimum point graphically is more complicated in this case.
102 Introduction to Optimization for Chemical and Environmental Engineers
x2
P3
=P
*
x2
P3
*
P2
P1
x1
x1
FIGURE 7.3
Possible solution to a two-variable maximization problem.
The previously mentioned methodology can be extended to more than two variables
and two constraints. The following would apply for three variables and four constraints:
y = f ( x1 , x2 , x3 ) = a1x1 + a2 x2 + a3 x3 (7.17)
and
B = b1x1 + b2 x2 + b3 x3 (7.18)
C = c1x1 + c2 x2 + c3 x3 (7.19)
D = d1x1 + d2 x2 + d3 x3 (7.20)
E = e1x1 + e2 x2 + e3 x3 (7.21)
and
x1 ≥ 0 (7.22)
x2 ≥ 0 (7.23)
x3 ≥ 0 (7.24)
Naturally, the previous can be rewritten for n objective functions and m constraints.
to be optimized. As noted previously, this phase of the problem is often the most difficult
and requires the greatest ingenuity and engineering judgment.
The next step attempts to model the object, system, or process. This model may assume
a wide variety of forms, but the usually preferred ones are mathematical in nature. One
may then use many mathematical, scientific, and engineering tools. As noted earlier, these
can involve not only the most empirical of data correlations but also the most rigorous
mathematical procedures or formal cut-and-try methods for assisting in the analysis of
this aspect of the solution.
Regarding constraints, Figure (7.4) illustrates a simple situation involving the operation
of a control equipment absorber7–10 with temperature and pressure limits for which each
depends on the magnitude of the other. The former might come from limitations due to
equilibrium considerations, since “the higher the temperature in the absorber the lower
the recovery absorptive capacity of the absorbing liquid while the higher the pressure in
the absorber increases the absorption.”8–10 The region of feasible solutions for the absorber
problem is the region enclosed by the constraints. The optimization problem is to find the
one set of variables or point in the feasible solution space that corresponds to the best solu-
tion. In order to locate the best feasible solution of the potentially infinite number of feasi-
ble solutions using a mathematical optimization technique, it is essential that a meaningful
objective function f(x) of the independent design variables be specified for this operation
concerned with the control and/or recovery of a particular component of gaseous emis-
sion. For simplicity, assume that one must choose for the aforementioned two independent
variables – temperature T and pressure P. The space is, therefore, two-dimensional, with
temperature and pressure as the coordinates. In practice, there usually are certain inequal-
ity constraints placed upon the range on both T and P. As shown in Figure 7.4, there could
be lower limits for the temperature and pressure, and an upper limit for each that depends
on the magnitude of the other.
Once again, the optimization problem is to find the one set of variables or point, x, in
the feasible solution space that corresponds to the best solution. In order to locate the best
Upper constraint on
Lower both T and P
Pressure, P
constraint Feasible
on T solution
Lower constraint on P
Temperature, T
FIGURE 7.4
Feasible solution for operation of an absorber.
104 Introduction to Optimization for Chemical and Environmental Engineers
where the sum of the coefficient aij times the value of each variable xi equals a requirement
bi in the ith equation. The number of variables must exceed the number of equations, so
that an infinite number of solutions is possible. The best one is selected. The relation-
ships in Equation 7.25 are often expressed as inequalities. “Dummy” additional variables
can then be introduced in order to convert these inequalities into appropriate equations.
Another requirement is that no quantity can be negative. Thus, mathematically
xi ≥ 0 and bi ≥ 0 (7.26)
This requirement will ensure that only useful answers are obtained.
The solution of a problem by linear programming also requires a profit function or
objective to be maximized (or minimized). This requirement is stated as
where ci is the profit (or cost) per unit of xi used. The solution is carried out by an iterative
process, which at each stage of calculation assigns either zero or positive values to all xi
variables. As the calculation proceeds, the values selected, while meeting the constraints
imposed, will tend to increase the summation to be maximized at each stage. If a solution
exists, a maximum profit value will finally be realized.”
Linear Programming 105
Constraints:
x2 ≤ −0.5x1 + 3 (2)
x2 ≤ −3 x1 + 6 (3)
x1 ≤ 3 (4)
x2 ≤ 4 (5)
Once again, the solution is first provided in graphical form. Refer to Figure 7.5.
Conditions (4) and (5) require that the solution, that is, the values x1 and x2 that will max-
imize P, lie within the rectangle provided in Figure 7.5a. Conditions (2) and (3) require
that the solution be within the area ABCD in Figure 7.5b. When the objective function
equation is superimposed on Figure 7.5b, Figure 7.5c results. The maximum P is located
at x1 = 1.2 and x2 = 2.4 for which P = 15.6. Excel provides the same solution.
SOLUTION:
As noted, xA and xB are the daily production rates of A and B respectively. Therefore,
P = x A PA + xB PB − x A ac − xBbd
x A + xB ≥ N
x A , xB ≥ 0
(5)
4.0
(3)
B
3.0
x2 x2 C
2.0 (4)
(2)
1.0
1.0 x1 3.0 A D x1
(a) (b)
5.00
4.80
3.12
Objective function
P*
x2 = 2.4
FIGURE 7.5
Graphical solution for Illustrative Example 7.1.
SOLUTION:
Based on the problem statement one concludes:
P = PA x A + PB xB + PC xC
M A x A + M B x B + MC x C ≤ Y
EA x A + EB xB + EC xC ≤ Z
x A , x B , xC ≥ 0
References
1. D. Green and R. Perry (editors), Perry’s Chemical Engineers’ Handbook, 8th edn, McGraw-Hill,
New York City, NY, 2007.
2. R. Ketter and S. Prawel, Modern Methods of Engineering Computation, adapted from, McGraw-
Hill, New York City, NY, 1969.
3. L. Lapidus, Digital Computations for Chemical Engineers, adapted from, McGraw-Hill, New York
City, NY, 1962.
4. E. Chong and S. Zak, An Introduction to Optimization, 4th edn, John Wiley & Sons, Hoboken, NJ,
2013.
5. G. Reklaitis, A. Ravindran, and K. Ragsdell, Engineering Optimization: Methods and Applications,
John Wiley & Sons, Hoboken, NJ, 1983.
6. A. Levy, The Basics of Practical Optimization, Society for Industrial and Applied Mathematics,
Philadelphia PA, 2009.
7. H. Mickley, T. Sherwood, and C. Reed, Applied Mathematics in Chemical Engineering, McGraw-
Hill, New York City, NY, 1957.
8. L. Theodore and F. Ricci, Mass Transfer Operations for the Practicing Engineer, John Wiley & Sons,
Hoboken, NJ, 2010.
9. L. Theodore, Chemical Engineering: The Essential Reference, McGraw-Hill, New York City, NY,
2014.
10. L. Theodore, F. Ricci, and T. VanVilet, Thermodynamics for the Practicing Engineer, John Wiley &
Sons, Hoboken, NJ, 2009.
11. J. Happel, Chemical Process Economics, John Wiley & Sons, Hoboken, NJ, 1958.
12. L. Theodore, Chemical Rector Analysis and Application for the Practicing Engineer, John Wiley &
Sons, Hoboken, NJ, 2008.
Part II
Environmental Engineering
Applications
Five chapters compliment the presentation in Part II. The titles follow:
The first chapter in Part II contains six illustrative examples. Example titles are listed here:
Seven select references for suggested reading for Chapter 8 are provided below:
Gaseous and particulate pollutants discharged into the atmosphere can be controlled.
The five generic devices available for particulate control include gravity settlers, cyclones
(centrifugal separators), electrostatic precipitators, wet scrubbers, and baghouses (fabric
filtration). The four generic devices for gases include condensers, absorbers, adsorbers, and
incinerators. Detailed information on these control devices are available in the literature.1
There are a number of factors to be considered prior to selecting a particular piece of air
pollution control hardware. In general, they can be grouped into three categories: environ-
mental, engineering, and economic. The latter category is discussed later since economics
are considered in the applications to follow. Key economic factors include1
i (1 + i )
n
CRF = (8.1)
(1 + i) n
−1
where:
i = interest, fractional value
n = equipment lifetime, years
SOLUTION:
The annualized cost for a new process is determined based on the following input data:
Interest , i = 7%
Term , n = 5 yr
With i = 7% and n = 5 yr, the CRF from Equation 8.1 is:
i (1 + i ) 0.07 (1 + 0.07 )
n 5
CRF = = = 0.2439
(1 + i)n − 1 (1 + 0.07 )5 − 1
The total annualized cost for the new process is then calculated by Equation 8.2:
Annualized cost = (CRF)(AIC) + AOC (8.2)
= Annualized installation + Annualized operation cost
Annualized cost = (0.2439)($150, 000 ) + $15, 000 = $36, 585 + $15, 000 = $51, 585
Air Pollution Management 113
Since this annual cost is lower than the original annual cost of $75,000 for the old pro-
cess, select the new process.
Coal combustion products (CCPs) can play an important role in its usage. CCPs are the
by-products generated from burning coal in coal-fired power plants. These by-products
include2
1. Fly ash
2. Bottom ash
3. Boiler slag
4. Flue gas desulfurization gypsum
5. Other types of material, such as fluidized-bed combustion ash, cenospheres, and
scrubber residues
The key air pollution control option for coal-fired boilers today is the baghouse, details
of which will be provided in the next section (Section 8.3).
TABLE 8.1
Data for Coal-Fired Power Plant Options
Time to Annual Salvage Value
Company Initial Cost Failure (yr) Operation Cost in Year 20
A $15,000 10 $2,300 0
B $22,000 45 $1,100 0
114 Introduction to Optimization for Chemical and Environmental Engineers
SOLUTION:
The annual probability of failure of the baghouse provided by Company A is 1/10 = 0.10,
while that of Company B is 1/45 = 0.022. With these results, the estimated annual cost of
each of the baghouse options can be determined as follows:
Annual cost = Annualized capital cost + Reinvestment cost + Annual O/M cost (8.3)
The value of the CRF for 20 years at 4% is 0.0736 (see Section 8.1). For Company A, the
estimated annual cost is:
Annual cost = ($15, 000 )(0.0736 ) + $15, 000 (0.45)(0.10 ) + $2, 300
= $4, 079 / yr
For Company B, the estimated annual cost is
Annual cost = ($22, 000 )(0.0736 ) + $22, 000 (0.45)(0.022) + $1, 100
= $2, 938 / yr
From this present worth evaluation, Company B’s bid for the baghouse installation
should be selected since the annual cost is minimized.
matter and gases from effluent gas streams. The evaluation and selection of a scrubber
system is somewhat simplified by the observation that collection efficiency is a function
of the amount of energy required to operate the scrubber. This means, that independent
of design and generally speaking, the more power put into the system, the greater the col-
lection efficiency. Therefore, for systems of nearly equivalent power inputs, the collection
efficiencies should be nearly equivalent. The selection or evaluation procedure between
two systems could then concentrate on ease of operation, potential maintenance problems,
and comparative costs.1
One of the oldest, simplest, and most efficient methods for removing solid particulate
containments from gas streams is by filtration through fabric media. The fabric filter is
capable of providing high collection efficiencies for particles as small as 0.1 µm and will
remove a substantial quantity of those particles as small as 0.01 µm. In its simplest form,
the industrial fabric filter consists of a woven or felted fabric through which dust-laden
gases are forced. A combination of factors results in the collection of particles on the fab-
ric filters. When woven fabrics are used, a dust cake eventually forms; this, in turn, acts
predominantly as a sieving mechanism. When felted fabrics are used, this dust cake is
minimal or almost nonexistent, and the primary filtering mechanisms are a combination
of inertial forces, impingement, and so on, on the fabric. These are essentially the same
mechanisms that are applied to particle collection on wet scrubbers where the collection
media is in the form of liquid droplets rather than solid fibers.1
Proper selection of a particular system for a specific application can be extremely diffi-
cult and complicated. In view of the multitude of complex and often ambiguous pollution
control regulations, it is in the best interest of the prospective user (as noted above) to work
closely with regulatory officials as early as possible in the process.
The final choice in equipment selection is usually dictated by that piece of equipment
capable of achieving compliance with regulatory codes at the lowest uniform annual cost
(amortized capital investment plus operation and maintenance costs). More recently, there
have been attempts to include liability problems, neighbor/consumer goodwill, employee
concerns, and so on, in the economic analysis, but these effects – although important – are
often extremely difficult to quantify.1
In order to compare specific control equipment alternatives, knowledge of the particular
application and site is also essential. A preliminary screening, however, may be performed
by reviewing the advantages and disadvantages of each type of air pollution control equip-
ment. For example, if water or a waste treatment system is not available at the site, this may
preclude the use of a wet scrubber system and instead focus particulate removal on dry
systems, such as cyclones, baghouses, and/or electrostatic precipitators. If auxiliary fuel is
unavailable on a continuous basis, it may not be possible to combust the organic pollutant
vapors in an incineration system. If the particulate-size distribution in the gas stream is
relatively fine, gravity settlers and cyclone collectors most probably would not be consid-
ered. If the pollutant vapors can be reused in the process, control efforts may be directed to
adsorption systems. There are many other situations where knowledge of the capabilities
of the various control options, combined with common sense, will simplify the selection
process.1
Environmental Inc. – has provided information on the cost of the equipment, as well as
installation, operating and maintenance costs, according to the conditions specified for
it to be treated. Given in Table 8.2 is all the collected data. Determine what equipment
you would select for this work in order to minimize costs on an annualized basis.
Costs are based on comparable overall collection efficiencies. The interest rate is 10%
and there is zero salvage value.
SOLUTION:
The first step is to convert the equipment, installation, and operating costs to total costs
by multiplying each by the total gas flow, 100,000 acfm. Hence, for the electrostatic pre-
cipitator, the total costs are
$3.1
Equipment cost = 100, 000 acfm = $310, 000
acfm
$0.80
Installation cost = 100, 000 acfm = $80, 000
acfm
$0.06
Operating cost = 100, 000 acfm = $6, 000/yr
acfm − yr
Note that the operating costs are on an annualized basis. The equipment cost and the
installation cost must then be converted to an annual basis using Equation 8.1 with
n = 20 and i − 0.10.3
CRF =
(0.1)(1 + .10)20 = 0.1175
(1 + .10 )20 − 1
The annual costs for the equipment and the installation is the product of the CRF and
the total costs of each.
TABLE 8.2
Precipitator, Scrubber, and Baghouse Cost; Illustrative Example 8.3
Electrostatic Precipitator Venturi Scrubber Baghouse
Equipment cost $3.1/acfm $1.9/acfm $2.5/acfm
Installation cost $0.80/acfm $1.4/acfm $1.0/acfm
Operating cost $0.06/acfm-yr $0.06/acfm-yr $0.095/acfm-yr
Maintenance cost $14,000/yr $28,000/yr $9,500/yr
Lifetime of equipment 20 yr 15 yr 20 yr
Air Pollution Management 117
TABLE 8.3
Calculation for Air Pollution Equipment; Illustrative Example 8.3
Electrostatic Precipitator Venturi Scrubber Baghouse
Equipment cost $310,000 $190,000 $250,000
Installation cost $80,000 $140,000 $100,000
CRF 0.1175 0.1315 0.1175
Annual equipment $36,412 $24,985 $29,375
Annual installation $9,400 $18,410 $11,750
Annual operating $6,000 $6,000 $9,500
Annual maintenance $14,000 $28,000 $9,500
Total annual cost $65,812 $77,395 $60,125
According to the analysis, the baghouse is the most economically attractive, that is, the
optimum device, since the annual cost is the lowest.
8.4 Recovering Dust
From an economic point of view, the break-even point of a process operation is defined as
that condition when the costs (C) exactly balance the income (I). The Profit (P) is, therefore,
P = I − C (8.4)
∆P
E= (8.5)
∆P + 15.0
where:
E = fractional collection efficiency
∆P = pressure drop, lbf/ft2
If the overall fan is 55% efficient (overall) and electric power costs $0.18/kWh, at what
collection efficiency is the cost of power equal to the value of the recovered material?
What is the pressure drop in inches of water at this condition?
SOLUTION:
The value of the recovered material (RV) may be expressed in terms of the collection effi-
ciency E, the volumetric flow rate q, the inlet dust loading w, and the value of the dust (DV):
RV = ( q) ( w )( DV )(E) (8.6)
118 Introduction to Optimization for Chemical and Environmental Engineers
The recovered value can be expressed in terms of pressure drop, that is, replace E by
Equation 8.5:
RV =
(0.429)( ∆P) $/min (8.7)
∆P + 15.0
The cost of power (CP) in terms of ∆P, q, the cost of electricity (CE) and the fan efficiency,
Ef, is
( )
CP = ( q) ( ∆P )( CE) / E f (8.8)
Substitution yields
The pressure drop at which the cost of power is equal to the value of the recovered
material is found by equating RV with CP:
RV = CP (8.9)
Substituting Equations 8.7 and 8.8 for RV and CP, respectively ultimately leads to
Figure 8.1 shows the variation of RV, CP, and profit with pressure drop. The collection
efficiency corresponding to the previously calculated ∆P from Equation 8.5 is:
∆P (8.5)
E=
∆P + 15.0
56.5
=
56.5 + 15.0
= 0.79
= 79%
The reader should note that operating below this efficiency (or the corresponding pres-
sure drop) will produce a profit; operating above the value leads to a loss.
The operating condition for maximum profit can also be estimated from the graph.
Calculating this value is left as an exercise for the reader. [Hint: Set the first derivative
of the profit (i.e., RV-CP) with respect to ∆P equal to zero. The answer is 13.9 lbf/ft2.]
However, this application will be revisited later in the book.
The reader should realize that this problem was addressed earlier in the text. One is
left the option of revisiting Illustrative Example 5.1 and 6.2.
Air Pollution Management 119
0.5
0.4
CP
0.3
RV
Value, $/min
0.2
Profit
0.1
Break-even
0
–0.1
–0.2
0 10 20 30 40 50 60 70 80
Pressure drop, lbf /ft2
FIGURE 8.1
Profit as a function of pressure drop; illustrative example 8.4.
a fuel or waste is oxidized occurs in a series of complex, free radical chain reactions. The
precise set of reactions by which combustion occurs is termed the mechanism of combus-
tion. By analyzing the mechanism of combustion, the rate at which the reaction proceeds
and the variables affecting the rate can be estimated. For most combustion devices, the rate
of reaction proceeds extremely fast compared to the mechanical operation of the device.
Maintaining efficient and complete combustion is somewhat of an art rather than a sci-
ence. Therefore, this section focuses on the factors that influence the completeness of com-
bustion rather than analyzing the mechanism involved.4,5
To achieve complete combustion once the air (oxygen), waste, and fuel have been brought
into contact, the following conditions must be provided: a temperature high enough to ignite
the waste/fuel mixture, turbulent mixing of the air and waste/fuel; and, sufficient residence
time for the reaction to occur. These three conditions are referred to as the “three T’s of com-
bustion.” Time, temperature, and turbulence govern the speed and completeness of reaction.
They are not independent variables since changing one can affect the other two.4
For combustion processes, ignition is accomplished by adding heat to speed up the oxi-
dation process. Heat is needed to combust any mixture of air and fuel until the ignition
temperature of the mixture is reached. By gradually heating a mixture of fuel and air, the
rate of reaction and energy released will gradually increase until the reaction no longer
depends on the outside heat source. In effect, more heat is being generated than is lost to
surroundings. The ignition temperature must be reached or exceeded to ensure complete
combustion. To maintain combustion of the waste, the amount of energy released by the
combusted waste must be sufficient to heat the incoming waste (and air) up to its ignition
temperature; otherwise, a fuel must be added. The ignition temperature of various fuels
and compounds can be found in the literature.4,5
The dangers of human exposure to many metallic chemical elements have long been
recognized. Metals such as mercury, lead, cadmium, and arsenic are toxic. Health effects
range from retardation and brain damage, especially in children from lead poisoning,
to the impairment of the central nervous system as a result of mercury exposure. Since
these toxic metals are chemical elements, they cannot be broken down by any chemical
or biological process. As a result of this, it is not uncommon to have metallic buildup or
bioaccumulation.6
Mercury and most of its compounds are toxic substances. Nineteenth-century hat mak-
ers, for example, developed a characteristic shaking and slurring of speech from occu-
pational exposure to large quantities of inorganic mercury during the manufacturing
process – symptoms that gave rise to the phrase “mad as a hatter.” Such problems result
from impairment of the central nervous system. In addition, high levels of mercury can
cause kidney damage, birth defects, and in extreme cases, death. The most common source
of the silvery liquid can be found in the bulbs of thermometers and thermostats. Large
quantities of mercury can also be found in one type of barometer or another. Small quanti-
ties are also found in fluorescent lights, mercury switches, batteries, hearing aids, smoke
detectors, watches, and cameras.6 Consider the following application.
The incinerator can operate with a heat rate between 25 and 40 million Btu/h. These
specifications provide two additional constraints on the operation of the incinerator.
3890
msludge = 3890 − mplate ; < 40 MMBtu/h (8.10)
5000
2431
msludge = 2431 − mplate ; > 25 MMBtu/h (8.11)
3125
Determine the maximum amount of mercury-contaminated plating waste that can be
incinerated employing a graphical approach involving the plotting of the two principal
variables, msludge and mplate (plating waste).
SOLUTION:
Let msludge and mplate represent the flowrates of the two wastes to the waste incinerator
in lb/h. The restrictions on the available blending options are provided in Table 8.4 and
shown graphically in Figure 8.2 using numbered lines for the constraints. The permissi-
ble area in Figure 8.2 is the quadrilateral area (ABCD) bounded by the lines 2, 4, 5, and 6.
The point farthest to the right (point A), represents the maximum rate at which plating
waste can be incinerated. This occurs where constraints 2 and 5 intersect. The describ-
ing equation for a minimum sludge rate of 1000 lb/h is then (employing Equation 8.12)
lb lb 3890 lb/h
1000 = 3890 − mplate
h h 5000 lb/h
so that
Can the reader determine the maximum rate at which the sludge can be incinerated?
TABLE 8.4
Flowrate Restrictions Related to Illustrative Example 8.5
Line 1 Maximum of 5000 lb/h of sludge msludge < 5000
Line 2 Minimum of 1000 lb/h of sludge msludge > 1000
Line 3 Maximum of 5000 lb/h of plating waste mplate < 5000
Line 4 Minimum of 1000 lb/h of plating waste mplate > 1000
Line 5 Maximum heat rate of 40 MMBtu/h5 3890
msludge = 3890 − mplate
5000
Line 6 Minimum heat rate of 25 MMBtu/h5 2431
msludge = 2431 − mplate
3125
122 Introduction to Optimization for Chemical and Environmental Engineers
6000
Line 4 Line 3
5000 Line 1
3000 B
Line 5
2000
C D
1000 Line 2
A
Line 6
0
0 1000 2000 3000 4000 5000 6000
mplate lb/h
FIGURE 8.2
Diagram showing six constraints for mercury incinerator application.
12, 000
E (lb/h ) = 0.23T + 0.19P + + 4.31; T = °C, P = psi (8.12)
PT
Concerned about the worst-case scenario of emissions, you have been hired to determine
the maximum emission rate from the previously mentioned incinerator. Earlier tests sug-
gest that the allowable operating range of the unit is in the 25°C–50°C temperature and
14.7–50.0 psi pressure ranges. Obtain the solution to this application employing a brute
force perturbation approaches (The reader may choose to revisit Illustrative Example 6.6).
SOLUTION:
Generate emission rates from Equation 8.12 for a number of combinations of pressure
and temperature. Some of the calculated results are provided in Table 8.5.
TABLE 8.5
Incineration Emission Rates as a Function of Temperature and
Pressure; Illustrative Example 8.6
P, psi T,°C
25.0 30.0 35.0 40.0 45.0 50.0
14.7 45.51 41.21 38.27 36.71 33.59 34.93
20.0 37.86 35.01 33.30 32.31 31.79 31.61
25.0 34.01 31.96 30.82 30.26 30.08 30.16
30.0 31.76 30.24 29.49 29.21 29.25 29.51
40.0 29.66 28.81 28.53 28.61 28.93 29.41
50.0 29.16 28.71 28.72 29.01 29.49 30.11
Air Pollution Management 123
The reader is left the exercise of refining (expanding) the table. The results suggest that
the minimum emission rate occurs with T ≈ 35–40°C and P ≈ 40 psi with E ≈ 28.5 lb/h.
The reader should also realize that the application could have been solved analyti-
cally by employing an approach by requiring partial derivatives.7,8
∂E 12, 000
= 0.23 − (8.13)
∂T T 2P
∂E 12, 000
= 0.19 − (8.14)
∂P TP 2
Setting both Equations 8.13 and 8.14 equal to zero and solving simultaneously leads to
T = 35.06°C
P = 42.44 psi
12, 000
E = 0.23 ( 35.06 ) + 0.19 ( 42.44 ) + + 4.31
( 42.44)( 35.06)
= 8.06 + 8.06 + 8.06 + 4.31
= 28.49 lb/h
∂ 2E
=+
(12, 000)( 2)
∂T 2 T 3P
∂ 2E
=+
(12, 000)( 2)
∂P 2
TP 3
Since both second partial derivatives are positive, the calculated value for the previously
mentioned E is indeed a minimum.
References
1. L. Theodore, Air Pollution Control Equipment Calculations, John Wiley & Sons, Hoboken, NJ,
2008.
2. K. Skipka and L. Theodore, Energy Resources: Availability, Management, and Environmental
Impacts, CRC Press/Taylor & Francis Group, Boca Raton, FL, 2016.
3. S. Shaefer and L. Theodore, Probability and Statistics for Environmental Science, CRC Press/Taylor
& Francis Group, Boca Raton, FL, 2007.
4. J. Santoleri, J. Reynolds, and L. Theodore, Introduction to Hazardous Waste Incineration, 2nd edn,
John Wiley & Sons, Hoboken, NJ, 2004.
5. L. Theodore, “Chemical Reactor Analysis and Application for the Practicing Engineer”, John
Wiley & Son, Hoboken, NJ, 2012.
124 Introduction to Optimization for Chemical and Environmental Engineers
6. L. Theodore and R. Dupont, “Environmental Health and Hazard Risk Assessment: Principles
and Calculations”, CRC Press/Taylor & Francis Group, Boca Raton, FL, 2012.
7. L. Theodore, personal notes, East Williston, NY, 2000.
8. L. Theodore and C. Prochaska, Introduction to Mathematical Methods for Environmental Engineers
and Scientists, Wiley‑Scrivener, Salem, MA, 2018.
9
Water Pollution Management
The second chapter in Part II contains six illustrative examples. Example titles are listed
here:
125
126 Introduction to Optimization for Chemical and Environmental Engineers
SOLUTION:
Write the equation relating C to I. Note that at break-even operation, P = 0.
I =C
N ≈ 1, 384, 600
≈ $89, 077
The reader should note that as N decreases below 1,384,600 items, P is negative (there is
a cost). Higher values of N lead to profits.
SOLUTION:
To select the most economical control device, a comparison can be performed among
the three units based on the total annualized cost (TAC). Table 9.2 can be used to sim-
plify these calculations:
A comparison among units A, B, and C indicates that unit C has the lowest TAC and
should be selected as the most economical unit of the three being evaluated. A similar result
would be obtained if a return on investment (ROI) method of analysis were employed.
TABLE 9.1
Economic Cost Data for Waste Pollution Control Options
Device Initial Cost Annual Operating Cost Salvage Value in Year 10
A $300,000 $50,000 0
B $400,000 $35,000 0
C $450,000 $25,000 0
TABLE 9.2
Total Annual Cost Calculations
Equipment A B C
Capital investment $300,000 $400,000 $450,000
Depreciation $30,000 $40,000 $45,000
Operating costs $50,000 $35,000 $25,000
Total annual costs $80,000 $75,000 $70,000
128 Introduction to Optimization for Chemical and Environmental Engineers
3. The problem is still quite relevant today and is manifested in the continuing high
levels of waterborne diseases in some countries, closing of bathing beaches, and
restrictions on water consumption.
There are three broad categories for control of pathogenic bacteria, viruses, and para-
sites: control at the input source of the microorganism, control at the area of water use, and
control of the product that is affected by contamination. Point sources of municipal wastes
are the usual principal inputs of communicable disease organisms, and such inputs can
be reduced by treatment of wastes without disinfection, and disinfection by chlorination,
ultraviolet (UV) radiation, ozonation, and chlorine dioxide. Controls of the area of water
use would include bathing restrictions on a transient basis (i.e., during and after storms,
resulting in combined sewer overflows), construction of dikes, and diversion structures to
protect a given area. Controls of the product would include chlorination or other forms of
disinfection of water used for municipal water supply, treatment plants for contaminated
shellfish to allow depuration of bacteria prior to marketing, and distribution of high-qual-
ity bottled water during emergencies.1
ADDITIONAL INFORMATION:
Conversion factor: A dose of 1 mg/L = 8.34 lb/106 gal (MG)
Straight-line depreciation equation:
SOLUTION:
Calculate the daily Cl2 and SO2 dosage rates as follows:
= 1, 278, 595 lb yr
= 426, 320 lb yr
The annual costs for Cl2 and SO2 addition are calculated as follows:
= $511, 584 yr
130 Introduction to Optimization for Chemical and Environmental Engineers
= $600, 000 yr
The annual cost savings for switching to the UV system is determined as follows:
Annual cost savings = Scenario A operating cost − Scenario B operating cost (9.8)
where scenario A is the total annual Cl2 cost and scenario B is the total annual UV
cost.
= $690, 000 yr
= 5.8%
This is more than the 3.0% simple interest the utility could make from cash in its bank
account (at the time of preparation of this problem). Thus, based on operating cost con-
siderations alone, the utility should invest in the UV system. Other considerations may
make UV even more attractive. There may be hidden costs to continued Cl2 and SO2
usage, such as maintaining safety equipment and training, hazardous materials plan-
ning, worker and public concerns, stockholder support, and, perhaps, capital costs to
maintain or upgrade the Cl2 and SO2 storage and dosing systems. Additional calcula-
tions should be made that include these as well as potentially other hidden costs before
a final decision is made.
Water Pollution Management 131
SOLUTION:
Set x1 and x2 equal to the daily delivery schedule of bottled water for Edmonton1 and
Winnipeg2, respectively. The profit objective function is then
x1 ≥ 0 (9.12)
x2 ≥ 0 (9.13)
Employing the previously mentioned three equations, Excel provides the following
solution:
SOLUTION:
Set x1, x2, and x3 as the tons of waste A, B, and C, respectively. The objective function for
the cost is given by
$ = 50 x1 + 20 x2 + 35x3 (9.14)
and
x1 ≥ 0 (9.18)
Water Pollution Management 133
x2 ≥ 0 (9.19)
x3 ≥ 0 (9.20)
To minimize the waste cost, 29.0 tons of Waste A, 9.4 tons of Waste B and 15.6 tons of
Waste C should be purchased. This will produce a minimum cost of $2,181. The calcula-
tions are left as an exercise for the reader. Hint: Employ Excel.
The four processes have their differences. The main difference between reverse osmosis
(RO) and ultrafiltration (UF) is that the size/diameter of the particles or molecules in solu-
tion to be separated/ concentrated are generally solids or colloids rather than molecules
in solution. In microfiltration (MF), the particles to be separated/concentrated are gener-
ally solids or colloids rather than molecules in solution. Gas permeation (GP) is another
membrane process that employs a nonporous semipermeable membrane to “fractionate”
a gaseous stream.6
The heart of the membrane process is the membrane itself. A membrane is an ultra-thin
semipermeable barrier separating two fluids that permits the transport of certain species
through the barrier from one fluid to the other. The membrane is typically made from
various polymers, such as cellulose acetate or polysulfone, but ceramic and metallic mem-
branes are also used in some applications. The membrane is selective since it permits the
transport of certain species while rejecting others. The term semipermeable is frequently
used to describe the selective action.6
134 Introduction to Optimization for Chemical and Environmental Engineers
TABLE 9.3
Membrane Data; Illustrative Example 9.6
Membrane
A B C D Maximum Input
Individual days 3 3 1 2 40
Lbs of raw membrane 5 5 7 6 80
Lbs of preferred membrane 3 3 6 10 65
Labor costs 2 5 4 3 50
provided in Table 9.3. If the A, B, C, and D membranes produced sell for $5, $5, $8, and
$20, respectively, determine the maximum production revenue subject to the constraints
listed in Table 9.3.
SOLUTION:
Set x1, x2, x3, and x4 as the production numbers for membranes A, B, C, and D. The objec-
tive function describing the production membrane number for maximizing the total
revenue (R) is presented here:
3 x1 + 3 x2 + x3 + 2x 4 ≤ 40 (9.22)
3 x1 + 3 x2 + 6 x3 + 10 x 4 ≤ 65 (9.24)
with
x1 ≥ 0 (9.26)
x2 ≥ 0 (9.27)
x3 ≥ 0 (9.28)
x 4 ≥ 0 (9.29)
The solution to this problem, subject to the information provided is from the previous
equation,
Water Pollution Management 135
x1 = 6.71
x2 = 3.71
x3 = 1.47
x 4 = 1.82
P = $100
The calculations are left as an exercise for the reader. Hint: employ Excel.
References
1. G. Burke, B. Singh, and L. Theodore, Handbook of Environmental Management and Technology,
2nd edn, John Wiley & Sons, Hoboken, NJ, 2000.
2. M. K. Theodore and L. Theodore, Introduction to Environmental Management, CRC Press/Taylor
& Francis Group, Boca Raton, FL.
3. L. Theodore and F. Ricci, Mass Transfer Operations for the Practicing Engineer, John Wiley & Sons,
Hoboken, NJ, 2010.
4. R. Dupont, K. Ganesan, and L. Theodore, Pollution Prevention: The Waste Management Approach
for the 21st Century, CRC Press/Taylor & Francis Group, Boca Raton, FL, 2000.
5. R. Dupont, K. Ganesan, and L. Theodore, Pollution Prevention, Sustainability, Industrial
Engineering, CRC Press/Taylor & Francis Group, Boca Raton, FL, 2017.
6. S. Slater, Membrane Technology, NSF Workshop Notes, Manhattan College, Bronx, NY, 1991
(adapted with permission).
10
Solid Waste Management
The third chapter in Part II contains six illustrative examples. Example titles are listed
below:
137
138 Introduction to Optimization for Chemical and Environmental Engineers
contain more than one type (e.g., a mixture of heavy metals and organics), more than one
method of treatment may be needed.
Waste treatment may be divided into five categories: physical, chemical, biological, ther-
mal (incineration), and solidification or encapsulation (a subcategory of physical treatment).
Physical treatments may be used alone or with the other types of treatments; their function
is to concentrate wastes, reduce waste volume, and separate different waste components for
continued treatment or disposal. Chemical treatments are used to convert hazardous wastes
into other less hazardous forms; they are generally used on one substance or similar sub-
stances because the reactions involved are specific. Biological treatments are more specific
than the other treatment types and are used mainly on organic wastes. Thermal treatment is
usually concerned with incineration (see later paragraph). Solidification and encapsulation are
also treatments for specific types of waste; these methods work well on inorganic wastes, but
physical and/or chemical pretreatment is usually necessary. Incineration is one of the most
popular forms of thermal waste treatment, especially when organic wastes are involved. It
destroys most organic components in the waste, and the resulting sludge and ash are almost
always less hazardous and easier to dispose of than the original waste. It is difficult to obtain
a permit for incineration, however, and many stringent regulations govern its use.1,2
Processes to perform various waste treatment tasks include:
Not all of these processes are currently feasible for hazardous wastes.
The objective of the treatment is another important factor that influences treatment selec-
tion. It may be desired, for example, to destroy the hazardous components in the waste,
reduce the hazardous components, and then isolate them, or merely to separate the haz-
ardous components. The end product of the waste treatment must be compatible with the
type of ultimate disposal to be used for the waste: sewage discharge, landfilling, deep-well
injection, and so on. For example, if the waste is to be disposed of by landfarming, the end
product of the waste treatment should consist of biodegradable organic components.
Economic feasibility and environmental factors are obviously also important factors to
be considered in treatment selection. Cost is a strong function of the amount of waste to
be treated and the simplicity and degree of commercial usage of the treatment process.
The various costs include capital investment, operating costs, utility costs, and expenses
associated with the final disposal of the waste product. Environmental factors, in the form
of state and local environmental regulations, also influence waste treatment selection.
The Environmental Protection Agency (EPA) Resource Conservation and Recovery Act
(RCRA) regulations especially control the type of waste treatment that may be used and
the characteristics of the waste that is to finally be disposed.1
producing the three copper wastes, A, B, and C, is 2.20, 1.94, and 3.10, respectively,
$/lb. Copper Waste A generates 43%, 14%, and 43% of pure, medium, and poor grade
copper, respectively. Copper Waste B generates 35%, 35%, and 30% of pure, medium,
and poor grade copper, respectively. Copper Waste C generates 75%, 7%, and 18% of
pure, medium, and poor grade copper, respectively. However, the maximum produc-
tion of pure, medium, and poor grade copper is limited to 340,000, 66,000, and 145,000
lb/mo. Calculate the daily processing rate of the three wastes to maximize the profits.
SOLUTION:
Set x1, x2, and x3, as the daily processing rate of wastes A, B, and C, respectively. The
objective function for the income I is then
with
x1 ≥ 0 (10.5)
x2 ≥ 0 (10.6)
x3 ≥ 0 (10.7)
x1 = 154, 838
x2 = 59, 255
x3 = 336, 907
Waste liquids in the process industries are numerous in kind and therefore defy easy
definition. Disposal of these waste liquids has become a serious problem for plant opera-
tors. Regulations today cover wastes once considered wastewater streams, as well as the
organic streams known to be made up of hazardous compounds. LI incinerators are avail-
able today to handle the various liquid streams that are being generated by the process
industries. These units may be on-site (at the location of the generator) or off-site (at a sister
facility or at a commercial disposal operation).
LI incinerators are usually refractory-lined chambers (horizontal, vertical with orienta-
tion up or down), generally cylindrical, and equipped with a primary combustor (waste
and auxiliary fuel-fired) and often secondary combustors, or injection nozzle for low-cal-
orific-value materials (aqueous wastes containing either organic or inorganic compounds,
or both). These units operate at temperature levels from 950°C (1742°F) to 1700°C (3092°F).
Residence time in the chamber may vary from milliseconds to 2.5 s. The viscosity deter-
mines whether the material being incinerated is considered a liquid, slurry, or sludge.
Liquid units today are capable of burning high viscosity materials of 4500 Saybolt seconds
universal (SSU) or less.1,2 Critical to the operation of the unit would be the atomizing noz-
zle used to convert the liquid stream into finely atomized droplets.2 When considering the
LI incinerator, one must be concerned about the ability of the system, that is, the design,
components, controls, and air pollution control system, to meet the requirements of the
federal and state regulations.
The physical, chemical, and thermodynamic properties of the waste must be consid-
ered in the basic design requirements of the total incinerator system. These include stor-
age tanks, mixers, pumps, control valves, piping, atomizers, combustors, refractory, heat
recovery, quench system(s), and the air pollution equipment. The types of data needed
by the designer to properly engineer the total system and the individual components are
available in the literature.1
Rotary kiln (RK) process incinerators were originally designed for lime processing.
Recently the RK has been utilized to process hazardous waste because it has the unique
capability to stabilize the combustion process for many different materials that may be
fed simultaneously. These kilns have demonstrated the ability to handle a wide variety of
materials with minimal impact on their performance. The result is minimal preprocessing
of the wastes and is one of the main reasons for the RK popularity.
The processing of raw materials has been a historic application of the RK. The basic
building materials of a community – cement, lime, coal, iron ore, aggregates, and other
bulk materials – require heating to high temperatures. These materials historically have
been produced in an RK. This high-temperature treatment and the continual mixing and
blending of the raw materials for cement are constantly exposed to the heat of the kiln.
Lime, lightweight aggregate, cement, and other process kilns may operate at temperatures
up to 1925°C (1350°F). Hazardous waste incinerators operate in the range of 700−1315°C
(1300−2400°F) depending on the waste types and shapes fed.
The RK is a cylindrical refractory-lined shell that is mounted at a slight incline from the
horizontal plane to facilitate mixing the waste materials and exposing the surface to the
auxiliary burner flame plus the waste fuel flames and flames generated over the burning
surface of solids. The kiln accepts all types of solid waste materials with heating values
between 555 and 8333 kcal/kg (1000 and 15,000 Btu/lb), and even higher. Solid wastes and
drummed wastes are usually fed by a pack-and-drum feed system, which may consist of
a bucket elevator for loose solids and a conveyor system for drummed wastes. Pumpable
sludges and slurries are injected into the kiln through nozzles. Temperatures for burning
vary from 700°C to 1315°C (1300°F to 2400°F) depending upon the waste types and shape.
Solid Waste Management 141
RKs are classed as slagging [>1000°C (1832°F)] or non-slagging [<1000°C (1832°F)]. The
non-slagging rotary kiln for hazardous waste destruction is not a total incineration sys-
tem. The kiln does not attain high destruction and removal efficiencies (DRE) required
of an incinerator due to the basic manner that it operates. It will volatilize the organic
content of the materials placed into it. The volatilized organics are carried in the com-
bustion products exit gas stream. A secondary combustion chamber (SCC) is located
immediately downstream of the kiln to perform the final combustion (destruction) pro-
cess on the kiln exit combustion products. The non-volatile part of the material fed into
the kiln, such as metals, stones, sand, and other solid inorganics exit as ash. The ash and
residue exiting the kiln should contain minimum to zero organics and may be treated
and disposed of in a hazardous waste landfill. Some metals and other elements, due to
the operating temperature will vaporize within the kiln and be carried in the exhaust
gas stream.
The RK combines three very important aspects of waste material processing into one
piece of equipment: mixing, temperature, and time. The mixing process is well known
in the production of cement where lime, iron, and sand or clay are mixed while being
heated. Good mixing is required to obtain a quality product. Therefore, the processing
equipment must allow for the wide variations in the material while achieving a uniform
bottom ash that meets regulatory requirements. The waste material that is placed into
the kiln must have sufficient time for its organic components to volatilize. With proper
time, the heat source within the kiln can raise the waste material to a desired tempera-
ture. The RK provides excellent mixing through the tumbling action to distribute the
heat evenly to all of the material being processed and to expose the material to the heat
source. Interestingly, the original kilns used for waste incineration were called “tumble
burners.”
SOLUTION:
For both units:
= 0.1614
Annual capital and installation costs for the cement kiln (CK) are
TABLE 10.1
Economic and Financial Data Based on Illustrative Example 10.2
Costs/Credits Cement Kiln Rotary Kiln
Capital ($) 2,625,000 2,975,000
Installation ($) 1,575,000 1,700,000
Operation ($/yr) 400,000 550,000
Maintenance ($/yr) 650,000 775,000
Income ($/yr) 2,000,000 2,500,000
TABLE 10.2
Incinerator Cost Comparisons for Illustrative Example 10.2
Cement Kiln Rotary Kiln
Total installed ($/yr) 678,000 755,000
Operation ($/yr) 400,000 550,000
Maintenance ($/yr) 650,000 775,000
Total annual cost ($/yr) 1,728,000 2,080,000
Income credit ($/yr) 2,000,000 2,500,000
Profit ($/yr) 272,000 420,000
Solid Waste Management 143
P = 0.023
SOLUTION:
For this condition, the probability that the lead (Pb) concentration in the receiving tanks
exceeds 15 ppm is
The value of z0 from the standard normal table3 is then 2.05, which corresponds to the
number of standard deviations above the mean tank concentration. According to the
aforementioned central limit theorem, the standard deviation of the mean lead concen-
tration is given by
σ 10 ppm
= (10.10)
n n
where n is the number of drums. Therefore, for a 2% probability that the mean concen-
tration in the tank exceeds 15 ppm, the number of drums can be found by solving
15 − 11
2.05 =
10
n
or
n = 26.3
IT = 0.5 ( TI ) (10.11)
The TI is obtained by subtracting the AOC and the depreciation of the plant (D) from
the revenues generated (R) or
TI = R − AOC − D (10.12)
If straight-line depreciation is assumed, the plant will depreciate uniformly over the
life of the plant. Thus, for a 10-year lifetime, the facility will depreciate 10% each year.
A = R − AOC − IT (10.14)
This procedure involves a trial-and-error solution. There are both positive and negative
cash flows. The positive cash flows consist of A and the recoverable working capital in
year 10. Both should be discounted backward to time = 0, the year the facility begins
operation. The negative cash flows consist of the TCC and the initial working capital
(WC). In actuality, the TCC is assumed to be spent evenly over the two-year construc-
tion period. Therefore, one-half of this flow is adjusted forward from after the first con-
struction year, (time = −1 yr) to the year the facility begins operating (time = 0). The other
half, plus the WC, is assumed to be expended at time = 0. Forward adjustment of the 50%
TCC is accomplished by multiplying by an economic parameter known as the single-
payment compound amount factor F/P, given by
F
= (1 + i ) (10.15)
m
P
Solid Waste Management 145
where:
i = rate of return (fraction)
n = the number of years (in this case, 1 yr)
For the positive cash flows, the annual after-tax cash flow (A) is discounted backward
by using a parameter known as capital recovering factor (CRF) or the uniform series
present worth factor (P/A). This factor is dependent on both interest rate (rate of return)
and the lifetime of the facility and is defined by
(1 + i ) − 1 (10.16)
n
P/A =
i (1 + i )
n
1
P/F = (10.17)
(1 + i )
n
where:
(1 + i )10 − 1
Term 1 = A; worth at year 0 of annual after − tax cash flows (10.19)
i (1 + i )
10
1
Term 2 =
10 WC ; worth at year 0 of recoverable WC after 10 year (10.20)
(1 + i )
1
Term 3 = WC + TCC ; assumed exoenditures at year = 0 (10.21)
2
1
( TCC ) (1 + i ) ; worth at year = 0 of assumed expenditures at year = −1 (10.22)
1
Term 4 =
2
SOLUTION:
For the fixation system, calculate, D, WC, TI, IT, and A. The depreciation is
D = 0.1( TCC )
= $250, 000
The WC is set at 10% of the TCC.
146 Introduction to Optimization for Chemical and Environmental Engineers
WC = 0.1( TCC )
= $250, 000
In addition,
TI = R − AOC − D
IT = ( 0.5 ) TI
A = R − AOC − IT
= $3, 600, 000 − $1, 200, 000 − $1, 075, 000 (10.27)
The rate of return, i, for the fixation unit is also calculated. The rate of return can be
computed by solving the following equation:
(1 + i )10 − 1 1
10 WC = WC + ( 0.5 ) TCC + ( 0.5 ) TCC ( 1 + i ) (10.28)
1
A+
I (1 + i ) (1 + i )
10
or (on substituting)
(1 + i )10 − 1 1
I (1 + i )
10 ( )
1.325 × 10 6 +
(1 + i )
10 ( )
0.250 × 10 6
( ) (
) ( (
= 0.250 × 10 6 + ( 0.5 ) 1.250 × 10 6 + ( 0.5 ) 1.250 × 10 6 )) (1 + i )
1
By trial-and-error,
i = 39.6%
WC = D (10.29)
= $350, 000
Solid Waste Management 147
(1 + i )10 − 1 1
I (1 + i )
10 ( )
2.125 × 10 6 + (
(1 + i )
10
0.3650 × 10 6 )
( ( ) )
= 0.350 × 10 6 + ( 0.5 ) 1.750 × 10 6 + ( 0.5 ) 1.750 × 10 6 (1 + i )
1
By trial-and-error,
i = 44.8%
Hence, by the discounted cash flow method, the rate of return on the initial capital
investment is approximately 5% greater for the encapsulation system than the fixa-
tion system. From a purely financial standpoint, the encapsulation system is more
attractive.
bioconcentration becomes staggering. Fish at the top of a food chain may have levels of
methyl mercury a million times the level in the surrounding water.6,7
SOLUTION:
A seventh constraint is added to the earlier application (See also Table 8.4 and Figure 8.1).
It is required to keep the Hg < 150 lb/h. This may be represented by an equation in the
form:
The constraint is shown as line 7 in Figure 10.1. The equation of this line (providing the
upper limit of msludge) is
150 0.08
msludge = − mplate
0.02 0.02 (10.31)
= 7500 − 4.0mplate
Given all seven constraints, the permissible area is then the quadrilateral ABCD
bounded by lines 2, 5, 6, and 7. The farthest point to the right in this area (point C) on
8000
7000 Line 7
5000 Line 1
msludge lb/h
4000
3000 A
Line 5
B
2000
D C
1000 Line 2
Line 6
0
0 1000 2000 3000 4000 5000 6000
mplate lb/h
FIGURE 10.1
Diagram showing seven constraints for mercury application.
Solid Waste Management 149
the figure is the maximum incineration rate of the plating waste. This occurs where
constraints 6 and 7 interest, that is,
2431 lb/h
7500 lb/h − 4.0mplate = 2431 lb/h − mplate
3125 lb/h
mplate = 1573 lb/h
Interestingly, the shaded area on the figure represents the permissible operating range
for this unit given all seven constraints. The operating range does not allow much flex-
ibility. The solution is deterministic, and variation in waste, operator error, waste analy-
sis uncertainty, and so on, are unaccounted for. This graphical technique is adequate for
the two variables investigated in this problem.
This problem can also be solved using the optimization techniques such as linear
programming, a technique that can handle numerous constraints and variables and
can be used, for example, to minimize supplemental fuel costs while operating within
given system constraints. This approach will be employed to solve numerous applica-
tions later in this book.
materials are usually composed of sand (fine aggregate) and gravel, crushed stone, and
slag (coarse aggregate).7
Under normal conditions, concrete grows stronger as it grows older. The chemical reac-
tions between cement and water that cause the paste to harden and bind the aggregates
together require time. The reactions take place very rapidly at first and then more slowly
over a long period of time. In the presence of moisture, concrete continues to gain strength
for years.7
Concrete mixtures are usually specified in terms of the dry-volume ratios of cement,
sand, and coarse aggregates used. A 1:2:3 mixture, for instance, consists of one part by vol-
ume of cement, two parts of sand, and three parts of coarse aggregate. Depending on the
applications, the proportions of the ingredients in the concrete can be altered to produce
specific changes in its properties, particularly strength and durability. The ratios can vary
from 1:2:3 to 1:2:4 and 1:3:5; these mixtures are about 1–1.5 times the volume of the cement.
For high-strength concrete, the water content is kept low, with just enough water added to
preserve the ductility of the mix.7
SOLUTION:
The objective function is:
C = f ( x1 , x2 ) = 5x1 + x2 (10.32)
where x1 and x2 represent the mass of concrete (1) and (2), respectively.
The constraints are:
with
x1 ≥ 0 (10.36)
x2 ≥ 0 (10.37)
The calculation is left as an exercise for the reader. (Hint: x1 = 0, and x2 = 50.)
Solid Waste Management 151
References
1. T. Shen, Y. Choi, and L. Theodore, EPA Manual: Hazardous Waste Incineration, USEPA/APTI,
RTP, NC, 1985.
2. J. Santoleri, J. Reynolds, and L. Theodore, Introduction to Hazardous Waste Incineration, 2nd edn,
John Wiley & Sons, Hoboken, NJ, 2000.
3. F. Taylor and L. Theodore, Probability and Statistics, A Theodore Tutorial, Theodore Tutorials,
East Williston, NY, originally published by USEPA/APTI, RTP, NC, 1999.
4. S. Shaefer and L. Theodore, Probability and Statistics for Environmental Science, CRC Press/Taylor
& Francis Group, Boca Raton, FL, 2007.
5. L. Theodore and R. Dupont, "Environmental Health and Hazard Risk Assessment: Principles
and Calculations", CRC Press/Taylor & Francis Group, Boca Raton, FL, 2014.
6. R. Dupont, K. Ganesan, and L. Theodore, Pollution Prevention, Sustainability, Industrial Ecology,
and Green Science and Engineering, CRC Press/Taylor & Francis Group, Boca Raton, FL, 2017.
7. K. Behan, personal notes, Buffalo, NY, 2017.
11
Health Risk Assessment
The next to last chapter in Part II contains six illustrative examples. Example titles are
listed here:
Seven select references concerned with health risk assessment (HRA) are provided here:
153
154 Introduction to Optimization for Chemical and Environmental Engineers
plant and purged with air to provide a breathable atmosphere. Alternatively, the worker
may wear an airpack or other breathing apparatus.1
SOLUTION:
Set up an unsteady-state material balance for oxygen in the vessel. Let
C = oxygen concentration, mole fraction
V = vessel volume, ft3
F = air flow rate, ft3/min
t = time, minutes
CA = oxygen concentration in purge air
C0 = oxygen concentration in vessel at time zero
(V )
dC
= ( F )(C A ) − ( F )(C ) (11.2)
dt
Solve the resulting differential equation. This is a simple, first-order equation. The
result is
t
C = C A − ( C A − C0 ) 1 − exp − (11.3)
V F
Rearranging to solve for t:
V (C A − C0 ) (11.4)
t = ln
F (C A − C )
Calculate the time required for the oxygen concentration to reach 18% for C0 = 0:
V (C A ) (11.5)
t = ln
F (C A − C )
Health Risk Assessment 155
1200 0.21
t= ln = (12)(2.64)
100 0.21 − 0.195
= 31.6 min
= 0.53 h
In practice, some extra time would be allowed as a safety factor before entering and
an assistant should be standing by outside the vessel.
1.
Cancer: A tumor formed by mutated cells
2.
Carcinogen: A cancer-causing chemical
where:
RiskLD is a unitless probability (e.g., 2 × 10−5) of an individual developing cancer at low-
dose conditions
CDI is the chronic daily intake averaged over 70 years, mg/kg‑day
CSF is the carcinogenic SF, expressed in (mg/kg‑day)−1
The risk obtained from Equation 11.6 may be viewed as the product of an exposure term
(CDI) and a cancer “potency” factor, SF. However, this linear equation is valid only at low-
risk levels (i.e., below estimated risks of 0.01). For situations where chemical intakes might
be high (i.e., risks above 0.01), an alternative equation should be used. The one-hit equation,
156 Introduction to Optimization for Chemical and Environmental Engineers
which is consistent with the linear low-dose model given previously and described in the
following equation, should be used instead:
Risk HD = 1 − e (
− CDI×SF )
(11.7)
SOLUTION:
Denote xj as the number of a specific drug required for remission. The total cost TC
objective function to be minimized is
The nutritional constraints associated with each of the m ingredients are given by
xn ≥ 0
xn ≥ 0
(11.10)
xn ≥ 0
Health Risk Assessment 157
Bx
IR = A + − Dx ; $ yr (11.11)
C+x
( 1) ( 2 ) ( 3 )
where:
IR is the annual insurance reduction
Term1 is the savings factor concerned with making the employees aware of the
problem
Term2 is the savings factor concerned with the installation of x fans
Term3 is the savings factor (a cost) concerned with the purchase of x fans
Determine the number of fans that should be installed to maximize the annual insur-
ance reduction.
SOLUTION:
In order to reduce insurance costs (a maximization operation), the intern decided to
solve the insurance reduction (IR) model analytically. The calculation involved generat-
ing d(IR)/dx and setting the result equal to zero, that is,
d ( IR )
= 0 (11.12)
dx
For the equation provided in the problem statement,
d ( IR ) B Bx
= − − D = 0 (11.13)
dx C + x ( C + x )2
Equation 11.13 is a form similar to that for a quadratic equation, that is,
( )
B ( C + x ) − Bx − D C 2 + 2Cx + x 2 = 0 (11.14)
or
Dx 2 + 2CDx + C ( CD − B ) = 0 (11.15)
−b ± b 2 − 4 ac
x= (11.16)
2a
where:
a = D
b = 2CD
c = C(CD−B)
Substitution yields
hazardous agent. Health risk is therefore a function of exposure and dose. Consequently,
HRA is defined as the process or procedure used to estimate the likelihood that humans
or ecological systems will be adversely affected by a chemical or physical agent under a
specific set of conditions.
Corporations that manufacture, sell, and purchase chemicals now realize that the chem-
icals they handle present health risks to their employees, their customers, and/or the pub-
lic. The assessment process is often an enormous task since the health risks of scores of
chemicals may have to be assessed from a risk perspective. Each individual HRA is a
multistep process consisting of
DA = 3 x1 + 4 x2 − x3 ; 10 −4 $ (11.18)
(Salvaging the existing boilers resulted in a credit.) The constraints, primarily due to
emission rules for NOX, SOX, and ROX were:
with
x1 < x2 < x3 ≤ 50 (11.22)
The state-run facility has been informed that capital cost expenditures would have to
be below $500,000 for this cost minimization project.
SOLUTION:
Excel provides the following solution:
x1 = 4
160 Introduction to Optimization for Chemical and Environmental Engineers
x2 = 10
x3 = 26
with
Thus, this minimization project should go forward. The solution requires the pur-
chase of four new boilers that burn oil (1), 10 new boilers that will burn oil (2), and the
reassignment of 26 of the old coal-fired boilers.
50 x1 + 40 x2 ≥ 300 (11.24)
20 x1 + 10 x2 ≥ 120 (11.25)
and
x1 ≤ 6 (11.26)
x2 ≤ 6 (11.27)
Health Risk Assessment 161
SOLUTION:
Excel provides the following solution for the number of vitamins:
x1 = 6
x2 = 0
and
C = $1.20
TABLE 11.1
Drug Laboratory Production Time Expenditure (Hours)
for Illustrative Example 11.6
Lab Drug A Drug B Drug C Drug D
1 10 4 8 9
2 6 5 3 9
3 3 9 7 12
is required in each of the three research labs to make one unit of each of the four
products.
Laboratory 1 has no more than 300 man-hours available per day. Laboratory 2 is lim-
ited to no more than 400 man-hours per day. Similarly, Laboratory 3 is limited to 425
man-hours per day.
As a recent biomedical engineering graduate, you have been hired to determine the
number of A, B, C, and D units that should be produced to maximize daily profit, P. The
following information is provided
where x1, x2, x3, and x4 represent the number of drugs A, B, C, and D, respectively, pro-
duced per hour. The man-hour constraints are
10 x1 + 4 x2 + 8 x3 + 9 x 4 ≤ 300 (11.29)
and
x1 ≥ 0 (11.32)
x2 ≥ 0 (11.33)
x3 ≥ 0 (11.34)
x 4 ≥ 0 (11.35)
SOLUTION:
Excel provides the following solution:
x1 = 0
x2 = 29
x3 = 23
x4 = 0
Health Risk Assessment 163
with a profit of
P = $8, 950 h
= $214, 900 day ( round the clock operation )
References
1. L. Theodore and R. Dupont: Environmental Health and Hazard Risk Assessment: Principles and
Calculations, CRC Press/Taylor & Francis Group, Boca Raton, FL, 2012.
2. J. Reynolds, J. Jeris, and L. Theodore, Handbook of Chemical and Environmental Engineering
Calculations, John Wiley & Sons, Hoboken, NJ, 2004.
3. L. Theodore, R. Dupont, and K. Ganesan, Unit Operations in Environmental Engineering,
Wiley‑Scrivener, Salem, MA, 2018.
4. L. Theodore, J. Reynolds, and K. Morris, Concise Dictionary of Environmental Terms, Gorden and
Breach Science Publishers/CRC Press/Taylor & Francis Group, Boca Raton, FL, 1997.
5. USEPA, Guidelines for Carcinogen Risk Assessment, EPA/630/R-98/002, Risk Assessment Forum,
Washington, DC, 2005.
6. K. Skipka and L. Theodore, “Energy Resources: Availability Management, and Environmental
Impacts”, CRC Press/Taylor & Francis Group, Boca Raton, FL, 2014.
12
Hazard Risk Assessment
The last chapter in Part II contains six illustrative examples. Example titles are listed here:
165
166 Introduction to Optimization for Chemical and Environmental Engineers
hazards, and potential exposures will determine the amount, kind, and location of this
equipment.
Regarding fires, water is the primary extinguishing agent, and it should be available
in adequate supply and at adequate pressure at all locations in the plant. The layout for
various types of fire protection installations and the appropriate recommendations for
their design are found in the standards of the National Fire Protection Association. Fire
hydrants, hose lines, and automatic sprinklers and water spray systems should all be a
part of the permanent equipment facilities of the plant.
Fire extinguishing systems can include foam, carbon dioxide, and dry chemicals. Wetting
agents and high-expansion foam have been used in some plant protection systems. All fire
extinguishing systems should be evaluated for their potential health risks as well as their
overall effectiveness before incorporation into a plant.
Since many chemical plants have severe potential health and hazard problems, it is
essential to provide medical facilities and first aid stations in case rapid medical response
is required. In addition, showers and eyewash stations are necessary in certain hazard-
ous areas. Also, guards and covers should be provided for all moving equipment. Ten of
the key protective equipment elements in an environmental or chemical process plant or
manufacturing facility are listed here:1
x ( 2.56 )
V = 25.6 + − 0.015x ; MM$ (12.1)
x + 3.5
SOLUTION:
This illustrative example requires the optimization of V. To employ an analytical solu-
tion, proceed as follows:
dV 2.56 x ( 2.56 )
= − − 0.015 (12.2)
dx x + 3.5 x + 3.52
Setting this equation equal to zero, that is, dV/dx = 0 and solving for x (positive/negative)
yield x ≅ 20. Therefore, 20 quake-buster columns should be employed in the new high-
rise to maximize the value of the structure, increasing it from 25.6 to 27.5 MM$.
This question requires answers in today’s high-tech environment. Failure to properly and
realistically address and answer this question can result/lead to an
ACCIDENT !
Once an industrial accident has occurred, the responsible individual (usually the oper-
ator or engineer) should react as quickly as possible in a responsible manner (exercise
good judgment) to detect and correct the situation, and hopefully reduce and/or eliminate
potential danger(s).
Industry now knows that equipment sometimes partially fails and sometimes cata-
strophically fails. In any event, it behooves the engineer to carefully and thoroughly exam-
ine the various pieces of plant equipment that can fail. A partial list of some of the more
common equipment that can be expected to fail in a plant with some regularity is pro-
vided subsequently. Details are available in the literature.1,2
1. Foundations
2. Structural steel
3. Vessels and tanks
4. Pumps
5. Compressors
6. Fans
7. Heat exchangers
8. Turbines
9. Electrical systems
10. Instrumentation and controls
168 Introduction to Optimization for Chemical and Environmental Engineers
Deviations, that is, abnormal conditions that can lead to a failure and/or an accident gener-
ally occur because of
1. Abnormal temperatures
2. Abnormal pressures
3. Material flow stoppage
4. Equipment leaks
5. Equipment spills
6. Materials failure due to wear
7. Materials failure due to imperfections
8. Materials failure due to poor maintenance
9. Materials failure due to corrosion
1. Blocked outlets
2. Opening/closing valves
3. Cooling water failure
4. Power failure
5. Instrument air failure
6. Thermal expansion
7. Vacuum problems
SOLUTION:
The total annualized cost calculations (TAC) are provided in Table 12.2.
Comparing projects 1, 2, 3, and 4 indicates that unit 3 has the lowest TAC and should
be selected as the most economical project of the four being evaluated.
Hazard Risk Assessment 169
TABLE 12.1
Cost Data for Four Safety Purposes; Illustrative Example 12.1
Project Initial Cost, $ Operating Cost, $/yr
1 2,600,000 200,000
2 2,400,000 250,000
3 2,450,000 275,000
4 1,850,000 350,000
TABLE 12.2
Cost Analysis for Four Safety Purposes; Illustrative Example 12.1
Project 1 2 3 4
Capital cost ($) 2,600,000 2,400,000 2,450,000 1,850,000
Depreciation ($) 130,000 120,000 125,000 92,250
Operating costs ($) 200,000 150,000 275,000 350,000
TAC ($) 330,000 270,000 400,000 442,250
2. f ( x ) ≥ 0 (12.4)
3. ∑ f (x) = 1 (12.5)
x
Property 1 indicates that the pdf of a discrete random variable generates probability by
substitution. Property 2 and Property 3 restrict values of f(x) to non-negative real numbers
and numbers whose sum is 1, respectively.
The pdf of a continuous random variable x has the following properties:
∫ f ( x ) dx = p(a < x < b) (12.6)
a
170 Introduction to Optimization for Chemical and Environmental Engineers
f ( x ) ≥ 0 (12.7)
∫ f ( x ) dx = 1 (12.8)
−∞
Equation 12.6 indicates that the pdf of a continuous random variable generates probability
by integration of the pdf over the interval whose probability is required. When this inter-
val contracts to a single value, the integral over the interval becomes zero. Therefore, the
probability associated with any particular value of a continuous random variable is zero.
Consequently, if x is continuous
P ( a ≤ x ≤ b ) = P( a < x ≤ b)
= P( a ≤ x < b)
Equation 12.7 restricts the values of f(x) to non-negative numbers. Equation 12.8 follows
from the fact that
The following properties of the cdf of a random variable, x, can be deduced directly from
the definition of F(x).
F ( b ) − F ( a ) = P( a < x ≤ b) (12.11)
F ( +∞ ) = 1 (12.12)
F ( −∞ ) = 1 (12.13)
These four properties apply to the cases of both discrete and continuous random variables.
0.015
0.0125
0.010
f(t) 0.075
0.005
0.025
0.0
8 16 32 64 128
Time (d)
FIGURE 12.1
Failure rate, f(t), as a function of time for Device A; Illustrative Example 12.3.
0.015
0.010
f(t)= 0.008
f(t)
0.005
0.0
8 16 32 64 128
Time (d)
FIGURE 12.2
Failure rate, f(t) as a function of time for Device B; Illustrative Example 12.3.
SOLUTION:
1. Calculate the probability of the device lasting at least 80 days using the follow-
ing equation:
80 8 80
∫
F ( t ) = f ( t ) dt =
0
∫ (0.0125) dt + ∫ (0.005) dt
0 8
80
F (t ) =
∫ (0.008 ) dt = 0.64
0
172 Introduction to Optimization for Chemical and Environmental Engineers
$ 400
Device A: 21.7 = $8, 680
device
$500
Device B: 15.6 = $7 , 800
device
So, the conclusion remains the same.
1737 1.875
P ( psi ) = + − 0.01556 (12.15)
D2 D
with D in feet, for the range from 3 to 400 ft. The distance for equal overpressures then
scales as the 1/3 power of the energy of the explosion (E):
1
D1 E1 3
D = E (12.16)
2 2
1. Estimate the maximum possible energy release (Emax) and the TNT equivalency.
2. Estimate the radius from the center of the explosion at which overpressures of
1 psi (sufficient to injure many people and cause damage to buildings), and 5
psi (sufficient to break wooden telephone poles) would occur.
3. Based on the results of (1) and (2), discuss how the hazard risks from the cat-
cracker can be minimized.
The cracker has a volume of 800 ft3. The vapors at the time of release consist of light
hydrocarbons with an average molecular weight of 55, at 25 psi and 350°F. The average
heat of combustion in air of the hydrocarbons is 19,000 Btu/lb.
SOLUTION:
PV
N hc =
RT
=
( 25)(8000) = 2.3 lbmol
(10.73)(810)
Mhc = ( 2.30 )( 55 ) = 127 lbs
2. Calculate the distance for an overpressure of 5 psi. Employ Equation 12.15. For
5 psi overpressure with 1 kg TNT:
1737 1.875
5= + − 0.01156
D2 D
Solving for D:
D = 18.8 ft
1210 lb TNT
E= = 550 kg TNT
2.2 lb / kg
Scale the distance for 550 kg TNT. Employ equation 12.16:
1 1
E 3 550 3
D1 = D2 1 = (18.8 ) = 154 ft
E
2 1
Finally, calculate the distance for an overpressure of 1 psi. The distance for 1
psi overpressure with 1 kg TNT is given by:
1737 1.875
1= + − 0.01156
D2 D
Solving for D:
D = 42.1 ft
1 1
E 3 550 3
D1 = ( D2 ) 1 = ( 42.1) = 345 ft
E
1 1
3. It has been reported that in most hydrocarbon vapor explosions, only 10 to 25%
of the maximum possible combustion energy is released. This results from the
vapor/air mixture being ignited when the mixture is still “rich”, that is, before
the amount of air for complete combustion has had a chance to mix with the
hydrocarbon vapors. Thus, efforts should be directed to reducing the amount
of energy released. In addition, consideration should be given to surrounding
the catcracker with barriers.
1. Prevent leaks.
2. Use appropriate sealings (cement finish or aluminum foil).
3. Use special insulation materials such as foam glass or crimped aluminum sheeting.
The potential for fire hazards is rather high in the chemical industry. However, this
potential is generally judged to be less than that of an explosion or toxic release. The scale
of a fire hazard can be determined by assessing the following factors: 1,2
1.
Inventory. The larger the inventory of material, the greater the loss potential.
2.
Energy. The more energy available for a release, such as the stored energy in a
material state or chemical reaction, the greater the potential.
3.
Time. A higher rate of release of a hazard and the distance over which it may cause
injury or damage also directly affect the potential.
4.
Exposure. The intensity of the hazard and the distance over which it may cause
injury or damage also directly affect the potential.
Most fire accidents involve a large loss of life and extensive property damage. Frequently,
a fire is preceded by an explosion. Explosions are usually more lethal than fires.1 Some
major fires in industrial plants have been described in detail in the literature.1,2 The ther-
mal radiation intensity and the time duration of fires often are used to estimate injury and
damage due to a fire.
The best way to fight a fire is to remove any one of three essential conditions required
to sustain the fire:
3. The supply of oxygen (e.g., by applying foams or inert gases). Various fire extin-
guishing systems can be used. These include
a. Water systems
i. Automatic sprinklers
ii. Fire hoses
b. Foam systems
i. Chemical foams: formed by the chemical reaction in which bubbles of CO2
gas and a foaming agent combine to produce and expand froth.
ii. Mechanical foams: bubbles of air, which are produced when air and water
are mechanically agitated with a foam-making agent.
iii. High-expansion foams: tiny foam bubbles filled with air, created by a fan
blowing the air through a mesh over which a detergent base solution is
flowing. High-expansion foam is used where water damage is a problem
or access to an area is not feasible. For example, such a foam would be
suitable if a flammable liquid held in an atmospheric storage tank were to
spill as a result of overfilling, causing the formation of an ignitable vapor
cloud.
c. Carbon dioxide systems.
d. Dry chemical systems (e.g., sodium bicarbonate). These are not toxic and do
not conduct electricity or freeze.
e. Water spray systems. These are used for explosion protection of buildings,
tanks, and control of flammable liquid fires.
f. Steam jet systems. These are used to smother some fires in closed containers or
in confined spaces.
2x1 + 3 x2 ≤ 30 (2)
x1 + 10 x2 ≤ 25 (3)
with
x1 ≥ 0 (4)
x2 ≥ 0 (5)
Qualitatively explain the previously mentioned profit mathematical model; that is, what
number of the two devices will maximize the profits.
Hazard Risk Assessment 177
SOLUTION:
Equation (1) is the objective function and describes the profit that will be produced from
the installation of the safety devices. Equations (2) and (3) are part of the constraints
on the limitation of the number of devices that can be installed because of space and
regulatory issues. Equation (4) and (5) require that the number of devices (naturally) be
a positive number.
The calculations are left as an exercise for the reader. (Hint x1 = 14 devices and x2 = 2
devices)
= 0; elsewhere (12.18)
y
0°C
0°C 0°C 0°C 0°C
y = b 0°C
1 2 3
0°C 0°C
4 5 6
0°C 0°C
7 8 9
0°C 0°C
y= 0 x
100°C 100°C 100°C 100°C 100°C
x= 0 100°C x =a
a =b
FIGURE 12.3
Monte Carlo gird approach.
With specified boundary conditions (BCs) for T(x, y) of T(0, y), T(a, y), T(x, 0), and T(x, a),
one may employ the following approach:
The results?
T1 = 7.07°C
T2 = 9.80°C
Hazard Risk Assessment 179
T3 = 7.16°C
T4 = 18.67°C
T5 = 25.12°C
T6 = 18.77°C
T7 = 42.93°C
T8 = 52.57°C
T9 = 42.80°C
x −x
− x6 tanh 3 6
( )( )
MC = x13 + x23 − x33 + x12 x22 x32 − x 42 x52 x62 + ln ( x1 ) ln ( x2 x3 x 4 ) + e x1
; $ (12.19)
The independent variables, x1, x2, and x3 relate to operating conditions while x4, x5, and
x6 relate to health and safety concerns.
The constraints are
4 ≤ x1 ≤ 99
4 ≤ x2 ≤ 99
4 ≤ x3 ≤ 99
4 ≤ x 4 ≤ 99
4 ≤ x5 ≤ 99
4 ≤ x6 ≤ 99
Note that there are (96)6 MC integer values to be potentially checked. Outline a method
of solution.
SOLUTION:
Apply the Monte Carlo method in the solution to the problem. Here is one of the author’s
approaches.7 It involves the reading/sampling of 10,000 (it could be more or less) pos-
sible solutions and selecting the maximum value derived from the objective function
calculations. First, set MC = 0.
180 Introduction to Optimization for Chemical and Environmental Engineers
The answer in (11) represents the maximum value of the sample of the 10,000 solu-
tions. The “actual” or “true” maximum of the entire population of (96)6 MC’s is prob-
ably slightly higher than the value in (11). Naturally, the more solutions attempted will
improve the estimate provided in (11).
The reader should consider what other approaches could be employed to improve on
the accuracy of the answer provided by the Monte Carlo method.
References
1. L. Theodore and R. Dupont, Environmental Health and Hazard Risk Assessment: Principles and
Calculations, CRC Press/Taylor & Francis Group, Boca Raton, FL, 2012.
2. L. Theodore, Chemical Engineering: The Essential Reference, McGraw-Hill, New York City, NY,
2014.
3. J. Reynolds, J. Jeris and L. Theodore, Handbook of Chemical and Environmental Engineering
Calculations, John Wiley & Sons, Hoboken, NJ, 2004.
4. D. Kaufmann, Process Synthesis and Design, A Theodore Tutorial, Theodore Tutorials, East
Williston, NY, originally published in the USEPA/APTI, RTP, NC, 1996.
5. L. Theodore and C. Porchaska, Introduction to Mathematical Methods for Environmental Engineers
and Scientists, Wiley‑Scrivener, Salem, MA, 2018.
6. S. Shaefer and L. Theodore, Probability and Statistics Applications for Environmental Science, CRC
Press/Taylor & Francis Group, Boca Raton, FL, 2007.
7. L. Theodore, Personal Notes, East Williston, NY, 2017.
Part III
Chemical Engineering
Applications
Five chapters compliment the presentation in Part III. The titles follow:
Note that the examples in Chapter 17 apply to both environmental and chemical
engineering.
Optimization provides important methods for decision-making in chemical engineer-
ing. It has evolved from a methodology of academic interest into a technology that contin-
ues to have a significant impact on both engineering research and practice. Optimization
algorithms include tools for experimental design, model development, and statistical anal-
ysis; process synthesis analysis, design, and retrofit; model predictive control and real-
time optimization; and planning, scheduling, and the integration of process operations.
Part III keys on optimization applications in chemical engineering. The five top-
ics addressed are: fluid flow (Chapter 13), chemical reactors (Chapter 14), mass transfer
(Chapter 15), heat transfer (Chapter 16), and plant design (Chapter 17). Each chapter con-
tains a brief introduction to the subject in question, which is followed by a host of illus-
trative examples keying on industrial applications that employ some— if not all— of the
optimization methods introduced in Part I. As noted in the previous paragraph, the per-
formance criteria can take many forms. For chemical engineering applications, empha-
sis has been placed on process variables, for example, heat exchange area, not equipment
details, for example, baffle spacing, and so on.
182 Introduction to Optimization for Chemical and Environmental Engineers
The reader should note that illustrative examples involving traditional optimization
will address both the development of the describing equation(s), that is, the model, and
the solution to the problem. Linear programming illustrative examples will key solely on
the modeling aspect of the problem and not the solution, which more often than not will
be “ obtained” from Excel. This statement particularly applies to some of the illustrative
examples at the end of the chapters.
13
Fluid Flow Applications
The first chapter in Part III contains six illustrative examples concerned with the general
subject of fluid flow. The title and one-line descriptions of each example are provided here:
1. R. Perry and D. Green, Perry’s Chemical Engineers’ Handbook, 8th edn, McGraw‑Hill,
New York City, NY, 2008.
2. L. Theodore, Chemical Engineering: The Essential Reference, McGraw‑Hill, New York
City, NY, 2014.
3. J. Reynolds, J. Jeris, and L. Theodore, Handbook of Chemical and Environmental
Engineering Calculations, John Wiley & Sons, Hoboken, NJ, 2004.
4. L. Theodore, R. Dupont, and K. Ganesan, Unit Operations in Environmental
Engineering, Wiley‑Scrivener, Salem, MA, 2018.
5. P. Abuleucia and L. Theodore, Fluid Flow for the Practicing Chemical Engineer, John
Wiley & Sons, Hoboken, NJ, 2009.
13.1 Fan Selection
If a pressure difference is required between two points in a system, a prime mover such
as a fan, pump, or compressor is usually used to provide the necessary pressure and/or
flow impetus. Environmental and chemical engineers are often called on to specify prime
movers more frequently than any other piece of processing equipment, particularly in the
chemical industry. In a general sense, these prime movers are to a process plant what the
engine is to one’s automobile. Whether one is processing petrochemicals, caustic soda,
acids, and so on, the fluid must usually be transferred from one point to another some-
where in the process. At a chemical plant, chemicals must be loaded or unloaded, sent to
heat exchangers or cooling coils, transferred from one processing unit to another, or pack-
aged for shipment.1
183
184 Introduction to Optimization for Chemical and Environmental Engineers
To move material (either a fluid or slurry) through the various pieces of equipment
at a facility (including piping and duct work) requires mechanical energy, not only to
impart an initial velocity to the material, but more importantly, to overcome pressure
losses that occur throughout the flow path. This energy may be imparted to the mov-
ing stream in one or all of three modes: an increase in the stream’s velocity, an increase
in the stream pressure, or an increase in stream height. In the first case, the additional
energy takes the form of increased external kinetic energy as the bulk stream veloc-
ity increases. In the second, the internal energy (mainly potential energy, but usually
some kinetic as well) of the stream increases. This pressure increase may also cause a
stream temperature rise. The term fans and blowers are often used interchangeably, and
no distinction will be made between the two in the following discussions. Whatever
is stated about fans applies equally to blowers. Generally speaking, however, fans are
used for low-pressure (drop) operation, usually below 2 psi. Blowers are employed
when generating pressure heads in the 2.0–14.7 psi range. High-pressure operations
require compressors.
Fans are usually classified as centrifugal or axial-flow types. In centrifugal fans, the gas
is introduced into the center of the revolving wheel (the eye) and discharges at right angles
to the rotating blades. In axial-flow fans, the gas moves directly (forward) through the axis
of rotation of the fan blades. Both types are used in industry, but it is the centrifugal fan
that is employed at most facilities.
The gas in a centrifugal fan is subjected to centrifugal forces. These forces compress the
gas giving it additional static pressure. Centrifugal fans are enclosed in a scroll-shaped
housing that helps convert kinetic energy to static pressure. Gas rotating between the fan
blades is compressed in the fan scroll, which increases the static pressure. Centrifugal fans
are classified by blade configuration as not only radial, forward curved, backward curved,
but also as air foil and radial tip-forward curved heel.
Radial or straight blade fans physically resemble a paddle wheel with long radial blades
attached to the rotor and are the simplest design of all centrifugal fans. This enables most
radial blade fans to be built with great mechanical strength and to be easily repaired.
These fans can be used in a variety of situations, especially heavy-duty applications. This
type of fan can handle erosive and corrosive gases as well as very viscous gases. It is
particularly well-suited for high static pressure operations and can generate pressures in
excess of 50 in H2O. When operated properly, the horsepower efficiency range is 55–69%,
with 65% as a typical value. Forward curved fans are the most popular for general ventila-
tion purposes (high flow rates and low static pressures). These fans have both the heel
and the tip of the blade curved forward in the direction of rotation. Blades are smaller
and spaced much closer together than in other blade designs. They are generally not used
with dirty gases when dust or sticky materials are present because contaminants easily
accumulate on the blades and cause imbalance. Efficiencies range from 52% to 71%, with
65% being typical. Backward curved or backward inclined fans have blades inclined in a
direction opposite to that of the direction of rotation. This feature causes the gas to leave
the tip of the blade at a lower velocity than the wheel-tip speed, a factor that improves the
mechanical efficiency. These types of fans are not suitable for a heavily particulate-laden
gas, sticky material, or abrasive dust. Centrifugal forces tend to build up particulate mat-
ter on the backside of the fan blades. The airfoil is similar to the backward curved fan,
except that the blade has been contoured to increase stability and operating efficiency.
These fans are more expensive to construct than backward curved fans, but they have
lower power requirements. They are rarely used in air pollution control where the gas
must be clean and noncorrosive.1
Fluid Flow Applications 185
i (1 + i )
n
A
CRF = = (13.1)
P i ,n ( 1 + i ) − 1
n
SOLUTION:
The annualized cost for the new fans is determined based on the following input data:
Capital cost = $94,000
Interest, i = 4.9%
Life, n = 6 yrs
For i = 0.049 and n = 6, the CRF is:
0.049 (1 + 0.049 )
6
CRF =
(1 + 0.049 )
6
−1
= 0.1958
The total annualized cost for the fans is therefore
13.2 Pump Selection
Pumps are required to transport liquids, liquid-solid mixtures such as slurries and slud-
ges, auxiliary fuel, and so on. Pumps are also needed to transport water to and/or from
such peripheral devices as boilers, quenchers, scrubbers, heat exchangers, distillation col-
umns, and so on.
Pumps may be classified as reciprocating, rotary, or centrifugal. The reciprocating and
rotary types are referred to as positive displacement pumps because unlike the centrifugal
type, the liquid or semiliquid flow is broken up into small portions as it passes through the
pump. These three classes of units are described below.1,2
Reciprocating pumps operate by the direct action of a piston on the liquid contained in a
cylinder. As the liquid is compressed by the piston, the higher pressure forces it through
discharge valves to the pump outlet. As the piston retracts, the next batch of low-pres-
sure liquid is drawn into the cylinder and the cycle is repeated. The piston may be either
186 Introduction to Optimization for Chemical and Environmental Engineers
directly steam driven or moved by a rotating crankshaft through a crosshead. The rate
of liquid delivery is a function of the volume swept out by the piston and the number of
strokes per unit time. A fixed volume is delivered for each stroke, but the actual delivery
may be less because of both leakage past the piston and failure to fill the cylinder when
the piston retracts. The volumetric efficiency of the pump is defined as the ratio of the
actual volumetric discharge to the pump displacement. For well-maintained pumps, the
volumetric efficiency is at least 95%. Reciprocating pumps are used for some applications.
Reciprocating pumps can deliver the highest pressure of any type of pump (20,000
psig); however, their capacities are relatively small compared to the centrifugal pump.
Also, because of the nature of the operation of the reciprocating pump, the discharge flow
rate tends to be somewhat pulsating. Liquids containing abrasive solids can damage the
machine surfaces of the piston and cylinder. Because of its positive displacement opera-
tion, reciprocating pumps can also be used to measure liquid volumetric flow rates.
The rotary pump combines rotation of the liquid with positive displacement. The rotat-
ing elements mesh with elements of the stationary casing in much the same way that two
gears mesh. As the rotating elements come together, a pocket is created that first enlarges,
drawing in liquid from the inlet or suction line. As rotation continues, the pocket of liquid
is trapped, reduced in volume, and then forced into the discharge line at a higher pressure.
Centrifugal pumps are the most widely used in the process industry because of simplic-
ity of design, low initial cost, low maintenance, and flexibility of application. Centrifugal
pumps have been built to move as little as a few gallons per minute against a small head,
and as much as several thousand gallons per minute against a pressure of several hun-
dred pounds force per square inch (psi). In its simplest form, this type of pump consists
of an impeller rotating within a casing. Fluid enters the pump near the center of the rotat-
ing impeller and is thrown outward by centrifugal force. The kinetic energy of the fluid
increases from the center of the impeller to the tips of the impeller vanes. This high veloc-
ity is converted to a high pressure as the fast-moving fluid leaves the impeller and is driven
into slower moving fluid in the volute or diffuser.
SOLUTION:
The hydraulic horsepower (HHP) required is1,2
TABLE 13.1
Data for Illustrative Example 13.2
Pump Specifications Pump A Pump B
Cost delivered to plant $4,500 $6,000
Efficiency at desired flow conditions 55% 60%
Fluid Flow Applications 187
= 9.79 HP
9.79
BHPA = = 17.8 HP
0.55
9.79
BHPB = = 16.3 HP
0.60
Converting to kW,
The total electrical energy in kWh required for three years becomes
The total three-year cost for each pump can now be calculated (see also Table).
Pump B is the more economical choice. Note, however, that the difference is only
about 2%. Small changes in the cost of electricity, the actual pump efficiencies, or the
stream factor could shift the choice to pump A.
In the final design and purchase of equipment for a new plant, hundreds of decisions
such as this are made after the overall process flow design is completed.
TABLE 13.2
Pump Selection Criteria
Cost Pump A Pump B
Purchase cost $4,500 $6,000
Electricity cost $26,700 $24,400
Total $31,200 $30,400
188 Introduction to Optimization for Chemical and Environmental Engineers
SOLUTION:
The usual calculational procedure is as follows:
The only variable that will appear in the resulting total-cost expressions is the pipe
diameter. The optimum economic pipe diameter can be generated analytically by taking
the derivative of the total annual cost with respect to pipe diameter, setting the result
equal to zero, and solving for the diameter. The derivative operation can be replaced by
a trial-and-error procedure that involves calculating the total cost for various diameters
and simply selecting the minimum. These calculations can be employed to generate
a graph of the form provided in Figure 13.1. In any event, computerized optimization
Total cost
Cost $/yr
Operating cost
Capital cost
Pipe diameter
FIGURE 13.1
Optimum Pipe Diameter.
Fluid Flow Applications 189
procedures and generalized charts are available that enable the practicing engineer to
optimize pipe diameter for numerous applications.
The reader should note that this topic will be revisited in Part IV, Chapter 19.
The reader should also note that tube/pipe size comes into play with heat e xchangers –
a topic that is reviewed in Chapter 15. For heat exchangers, as the tube/pipe diameter is
decreased, the heat transferal increases (since the coefficient of heat transfer increases)
somewhat in many cases; hence, the required area decreases. However, the pressure
drop increases, and consequently there is an optimum diameter corresponding to the
minimum sum of fixed/capital charges and power/operating costs. Note that the dif-
ficulty of cleaning the tubes increases with a decrease in diameter, which may result in
higher maintenance costs.
13.4 Two-Stage Compressor
Compressors operate in a similar fashion to pumps and have the same classifications:
rotary, reciprocating, and centrifugal. An obvious difference between the two opera-
tions is the large decrease in volume resulting from the compression of a gaseous stream
compared to the negligible change in volume caused by the pumping of a liquid stream.
Details on the three compressor classifications follow.
Centrifugal compressors are employed when large volumes of gases are to be handled at
low to moderate pressure increases (0.5–50 psi). Rotary compressors have smaller capaci-
ties and can achieve discharge pressures up to 100 psi. Reciprocating compressors are the
most common type used in industry and are capable of compressing small gas flows to
as much as 3500 psig. With specially designed compressors, discharge pressures as high
as 25,000 psig can be reached, but these devices are capable of handling only very small
volumes and do not work well for all gases. For the environmental applications mentioned
earlier – atomizing of liquids for combustion or of venturi scrubber water for gas c leaning –
reciprocating compressors are normally used.
The following equation may be used to calculate compressor power requirements when
the compressor operation is adiabatic and the gas (usually air) follows ideal gas behavior:3,4
( γ − 1)/γ
γRT P2
Ws = − 1 (13.4)
γ − 1 P1
where:
Ws = compressor work required per lbmol of air
R = 1.987 Btu/(lbmol · °R)
T = air temperature at compressor inlet conditions (°R)
P1, P2 = air inlet and discharge pressures
γ = ratio of the heat capacity at constant pressure to that at constant volume – typi-
cally 1.3 for air
( k − 1)/k
T2 P2
= ; adiabatic operation (13.5)
T1 P1
190 Introduction to Optimization for Chemical and Environmental Engineers
( N − 1)/N
T2 P2
= ; polytropic operation (13.6)
T1 P1
where:
k = adiabatic exponent, Cp/Cv
N −1 K −1
N = polytropic exponent, =
N KEp
P1, P2 = suction, discharge pressures, respectively psia
T1, T2 = suction, discharge temperatures, respectively °R
Ep = polytropic efficiency, fraction
k −1
k −1
k P2 k P3 k (13.7)
E = NRT + − 2
k − 1 P1 P
1
where:
P = pressure, psia
N = lb mol of gas compressed
R = ideal gas constant, (consistent units)
k = ratio of heat capacity at constant pressure to the heat capacity at constant vol-
ume for gas compressed
T = inert gas temperature, °R
E = total work, ft · lbf
Calculate the inter-stage pressure P2 that will minimize the work if P1 = 14.6 psia and
P2 = 79.2 psia.
SOLUTION:
If the dependent variable E is to be a minimum, then the derivative must be zero, that is,
d ( E) k k k − k k − 1 k k
1− k 1 k −1 1− 2 k
= NRT1 P P − P P = 0 (13.8)
k − 1 k − 1
1 2 3 2
dP2 k
Fluid Flow Applications 191
2 k −2 k −1 k −1
P2 k
= P1 k
P3 k
(13.9)
or
2( k −1) k −1 k −1
k
P2 = P1 k
P3 k
(13.10)
P2 = P1P3 (13.11)
13.5 Ventilation Models
Indoor air pollution is rapidly becoming a major health issue in the United States. Indoor
pollutant levels are quite often higher than outdoors, particularly where buildings are
tightly constructed to save energy. Since most people spend nearly 90% of their time
indoors, exposure to unhealthy concentrations of indoor air pollutants is often inevitable.
The degree of risk associated with exposure to indoor pollutants depends on how well
buildings are ventilated and the type, mixture, and amounts of pollutants in the building.1,2,5
Alternatively, industrial ventilation is the field of applied science concerned with con-
trolling airborne contaminants to produce healthy conditions for workers and a clean
environment for the manufacture of products. To claim that industrial ventilation will
prevent contaminants from entering the workplace is naïve and unachievable. More to
the point, and within the realm of achievement, is the goal of controlling contaminant
exposure within prescribed limits. To accomplish this goal, one must be able to describe
192 Introduction to Optimization for Chemical and Environmental Engineers
the movement of gaseous and particle contaminants in quantitative terms that take into
account
As an authority in the field (having taken the Theodore course on Health, Safety, and
Accident Management), you have been requested to
SOLUTION:
Use the laboratory room as the control volume. Apply the concentration law for mass to
the nanochemical:
rate of mass
= v0 c0
in
rate of mass
= v0 c
out
rate of mass
= rV
generated
d(cV ) (13.13)
v0 c0 − v0 c + rV =
dt
Since the laboratory room is constant, V may be taken out of the derivative term. This
leads to
v0 dc
( c0 − c ) + r = (13.14)
V dt
The term V/v0 represents the average residence time the nanochemicals reside in the
room and is usually designated as τ. The previous equation may then be rewritten as
dc ( c0 − c )
= + r (13.15)
dt τ
1. If r = 0,
dc ( c0 − c )
= (13.16)
dt τ
Separating variables
dc dt
=
( c0 − c ) τ
194 Introduction to Optimization for Chemical and Environmental Engineers
c t
dc dt
∫
ci
( c0 − c )
=
∫τ
0
c −c t
ln 0 = τ
c0 − ci
t
c0 − c − τ
c −c = e
0 i
t t
− −
c = c0 − ( c0 − ci ) e τ = c0 + ( ci − c0 ) e τ (13.17)
2. If r = −k,
dc ( c0 − c ) c c c − kτ c
= −k = 0 −k− = 0 − (13.18)
dt τ τ τ τ τ
Separating variables,
dc
= dt
c0 − k τ c
τ τ −
dc dt
=
( c0 − k τ ) − c τ
c t
dc dt
∫
ci
( c0 − k τ ) − c
=
∫τ
0
( c0 − k τ ) − c dt
− ln =
( c0 − k τ ) − ci τ
( c0 − kτ ) − c =e
t
−
τ
( c0 − kτ ) − ci
which ultimately leads to
−
t −
t
c = ci e τ + ( c0 − k τ ) 1 − e τ (13.19)
Fluid Flow Applications 195
13.6 Three-Stage Compressor
k −1
k −1
k −1
k P2 k P3 k P4 k
E ( P2 , P3 ) = NRT + + − 3 (13.20)
k − 1 P1 P2 P3
Discuss the approach that was employed in Part I, to solve this problem if NRT = 1,
k = 1.4, P1 = 1 atm and P4 = 10 atm.
SOLUTION:
The reader is first referred to Illustrative Example 4.6 in Section 4.4. The method of
steepest ascent/descent was employed. The results provided were presented as 3
P2 = 2.154 atm
P3 = 4.642 atm
with
The problem was also solved in Illustrative Example 6.7 using analytical means. The
results were essentially identical to those provided in Illustrative Example 4.6. Another
approach that could have been employed to solve the problem was to use any brute
force trial-and-error methods that could have included graphical analysis.
Regarding the analytical solution, the reader should note that for the variable E to be
a minimum, first ∂E/∂P2 and ∂E/∂P3 must be zero. The derivatives are given by
∂E 1−k −1 k −1 1− 2 k
= NRT1 P1 k P2 k − P3 k P2 k (13.21)
∂P2
∂E 1−k −1 k −1 1− 2 k
= NRT1 P2 k P3 k − P4 k P3 k (13.22)
∂P3
196 Introduction to Optimization for Chemical and Environmental Engineers
P2 2 = P1P3 (13.23)
and
P3 2 = P2 P4 (13.24)
Therefore
( )
1/3
P2 = P12 P4 (13.25)
1
(
P3 = P4 2 P1 ) 3 (13.26)
If one now differentiates ∂E/∂P2 with respect to P2 and substitutes the values for P2
and P3 from the previous, the following is obtained:
k − 1 3 k − 3 k
1− 5 k 1+ k
∂ 2E
∂P = 2 NRT1 P1 P4 (13.27)
2
2
k
0
The subscript 0 indicates that the value of ∂ 2E / ∂P2 2 is at the possible minimum that one
is attempting to verify. Since the powers of P1 and P4 must be positive, and k > 1, it is clear
∂ 2E
that, is greater than zero, a positive quantity. Similarly, by differentiation of
∂P2 2 0
Equation 13.21 with respect to P3 and substitution of values from Equation 13.24 and 13.25
k − 1 3 k − 3 k
1− 3 k 1+ 3 k
∂ 2E
∂P 2 = 2 NRT1 P1 P4 (13.28)
3 k
0
This, by similar reasoning, is positive. The product of the second derivatives lead to
NRT1 ( k − 1)
2 2−8k 2+ 4k
∂ 2E ∂ 2E 3k
−
3 k
∂P 2 ∂P 2 = 4 P1 P4 (13.29)
2 0 3 0 k
Similarly, the cross derivative is given by
NRT1 ( k − 1)
2 2−8k 2+ 4k
∂ 2E 3k
−
3 k
∂P ∂P = P1 P4 (13.30)
2 3 0 k
Thus, it has also been demonstrated that
2
∂ 2E ∂ 2E ∂ 2E
∂P 2 ∂P 2 > ∂P ∂P (13.31)
2 3 2 3
Finally, it should be noted that considerable accuracy in computation is often required
to establish the optimum. This frequently occurs in applying other methods so that it is
necessary to use an analytical expression to establish the optimum accurately. Another
limitation of the procedure, using the method of steepest descent/ascent previously
discussed and the general analytical differentiation procedure, is that they do not take
into account external constraints imposed on the system. For example, the gradient may
lead into an area in which operations are not possible.
Fluid Flow Applications 197
References
1. I. Farag, Fluid Flow, A Theodore Tutorial, Theodore Tutorials, East Williston, NY, originally
published by the USEPA/APTI, RTP, NC, 1995.
2. P. Abulencia and L. Theodore, Fluid Flow for the Practicing Chemical Engineer, John Wiley & Sons,
Hoboken, NJ, 2009.
3. L. Theodore, Heat Transfer Applications for the Practicing Engineer, John Wiley & Sons, Hoboken,
NJ, 2011.
4. L. Theodore, F. Ricci, and T. VanViliet, Thermodynamics for the Practicing Engineer, John Wiley &
Sons, Hoboken, NJ, 2009.
5. R. Heinsohn, Industrial Ventilation: Engineering Principle, John Wiley & Sons, Hoboken, NJ, 1991.
6. R. Dupont, T. Baxter, and L. Theodore, Environmental Management: Problems and Solutions, CRC-
Lewis Publishers, Boca Raton, 1998.
7. K. Ganesan, L. Theodore, and R. Dupont, Air Toxins: Problems and Solutions, Gordon and Breach,
New York City, NY, 1996.
8. L. Theodore, “Nanotechnology: Basic Calculations for Engineers and Scientists”, John Wiley &
Sons, Hoboken, NJ, 2006.
14
Chemical Reactor Applications
The second chapter in Part III contains six illustrative examples concerned with the gen-
eral subject of chemical reactors (CR). The title and one-line descriptions of each example
are provided here:
14.1 Comparing continuous stirred tank reactor (CSTR) to tubular flow (TF) reactor
performance
14.2 Optimizing fluidized-bed reactor performance
14.3 Two reactors in series: achieving maximum conversion
14.4 Maximizing selectivity
14.5 Optimizing batch reactor performance
14.6 Optimizing operating schedules
The six references that follow are suggested reading for chemical reactors:
1. R. Perry and D. Green, Perry’s Chemical Engineers’ Handbook, 8th edn, McGraw‑Hill,
New York City, NY, 2008.
2. L. Theodore, Chemical Engineering: The Essential Reference, McGraw‑Hill, New York
City, NY, 2014.
3. J. Reynolds, J. Jeris, and L. Theodore, Handbook of Chemical and Environmental
Engineering Calculations, John Wiley & Sons, Hoboken, NJ, 2004.
4. L. Theodore, R. Dupont, and K. Ganesan, Unit Operations in Environmental
Engineering, Wiley‑Scrivener, Salem, MA, 2018.
5. L. Theodore, Chemical Reactor Analysis and Applications for the Practicing Engineer,
John Wiley & Sons, Hoboken, NJ, 2012.
6. L. Theodore, Chemical Reaction Kinetics, A Theodore Tutorial, Theodore Tutorials,
East Williston, NY, originally published by USEPA/APTI, RTP, NC, 1992.
199
200 Introduction to Optimization for Chemical and Environmental Engineers
A reaction where mixing is important is the tank flow or (CSTR); it is also referred to as a
“back-mix” reactor. This type of reactor, like the batch reactor discussed in a later section,
essentially consists of a tank or kettle equipped with an agitator. As with the batch reac-
tor, perfect mixing is normally assumed so that both the concentration and temperature
within the reactor is the same, that is, there is no spatial variation so that both terms are
not a function of position within the reactor. This perfect mixing assumption also requires
that both the concentration and temperature at the discharge point (the outlet pipe) is
identical to that in the reactor. Fogler3 notes that for perfect mixing: “The idea that the com-
position is identical everywhere in the reactor and in the exit pipe requires some thought.
It might seem that, since the concentration changes instantly at the entrance where mixing
occurs, reaction occurs there and nothing else happens in the reactor because nothing is
changing. However, reaction occurs throughout the reactor, but mixing is so rapid that
nothing appears to change with time and position.” One can extend this reasoning to also
conclude the same within the reactor and the discharge port.
This class of reactor finds application in industry when
The bottom line is that in addition to producing greater quantities with smaller equip-
ment, most industrial reactors are operated in a continuous rather than batch mode. They
also require less labor and maintenance; however, flow processes are more difficult to start
and stop than their batch counterparts.
As noted previously, the CSTR essentially consists of a tank or kettle equipped with an
agitator. It may be operated under steady or transient conditions. Reactants are fed contin-
uously, and the product(s) are withdrawn continuously. The reactants and products may
be liquid, gas, or solid, or a combination of them all. If the contents are perfectly mixed,
the reactor design problem is greatly simplified for steady conditions because the mixing
results in the aforementioned uniform concentration, temperature, and so on, throughout
the reactor. This means that the rate of reaction is also constant and the describing equa-
tions are not differential and, therefore, do not require integration. In general, CSTRs are
used for liquid phase reactions. High reactant concentrations can be employed with low
flow rates so that conversions approaching 100% can be achieved. However, the overall
economics of the system may be reduced because of low throughput rates.
CSTRs (as well as TF reactors to be addressed shortly) are often connected in series, in
such a manner that the exit stream of one reactor is the feed stream for another reactor.
Under these conditions, it is convenient to define the conversion at any point downstream
in the battery of CSTR reactors in terms of inlet conditions, rather than with respect to any
one of the reactors in the series. The conversion X (or X A) is then the total moles of A that
have reacted up to that point per mole of A fed to the first reactor. However, this definition
should only be employed if there are no side stream withdrawals and the only feed stream
enters the first reactor in the series. The conversion from reactors 1, 2, 3,… in the series are
usually defined as X1, X2, X3, …, respectively, and effectively represent the overall conver-
sion for that reactor relative to the feed stream to the first reactor.
As noted, the CSTR has certain advantages because of the near uniform temperature,
concentration, and so on, that results because of mixing. However, for perfect mixing,
the reaction occurs at a rate determined by the concentration of the discharge (or exit)
stream from the reactor. Since the rate usually decreases with the extent of reaction (con-
version), the CSTR operates in the range between the high reaction rate corresponding to
Chemical Reactor Applications 201
the concentration in the feed stream and the (normally) lower rate corresponding to the
concentration leaving the reactor. The CSTR therefore requires a larger volume than a
tubular reactor to accomplish a given degree of conversion. This statement applies to all
positive reaction orders. The reaction/residence time ratio for a zero order reaction is one;
that is, both perform identically. However, the previous analysis occasionally does not
apply, particularly if the reaction mechanism is complex and/or non-elementary.
The second reactor to be examined is the TF reactor. The most common type involves
the use of the single-pass cylindrical tube; another type is one that consists of a number of
tubes in parallel. Both will be reviewed in more detail subsequently.
The TF reactor can be conceptualized by two different views. The first is to consider a
CSTR (see previous development) of given volume V, and feed rate q (or F) and divide the
volume into a number of much smaller CSTRs in series. In the limit, an infinite number
of these CSTRs with a total volume V reduces to a TF reactor. In other words, the time of
reaction in a flow system may be visualized as the time required for an infinitesimal parcel
of mass to pass through the reactor in an imaginary compartment (a CSTR of differential
volume) that moves with the flowing stream and reacts, producing changes to the tem-
perature, pressure, and composition of the stream. The second approach is to envision a
pipe (or tube) of uniform diameter with reactants and products flowing axially along the
pipe. This simpler perspective is also a TF reactor.
In terms of kinetic behavior, the TF reactor may be viewed as “identical” to a batch reac-
tor. The time t or reaction that can be defined for a batch reactor is now expressed in terms
of τ (the residence time) and represents the length of time necessary for any given incre-
ment of feed to travel the entire length of the reactor.
TF reactors are characterized by a continuous and decreasing concentration of reactants in
the direction of flow. Either horizontal or vertical orientation is common. The reactants are
charged continuously at one end, and the products are removed continuously at the other
end. The unit almost always operates in a steady-state mode. This greatly simplifies design
and predictive calculations. It is a unit that is amendable to automatic control and to experi-
mental work. As noted, the most common type of TF reactor is the single-pass cylindrical
tube. Another type is one that consists of a number of tubes in parallel. These reactor(s)
may be vertical or horizontal. The feed is charged continuously at the inlet of the tube, and
the products are continuously removed at the outlet. If heat exchange with surroundings is
required, the reactor setup includes a jacketed tube and effectively acts as either a double
pipe or a single pass heat exchanger.1 If the reactor is “empty,” a homogeneous reaction –
one phase present – usually occurs. If the reactor contains catalyst particles, the reaction is
said to be heterogeneous. This type will be considered later in this chapter.
TF reactors are usually operated under steady conditions so that physical and chemical
properties do not vary with time. Unlike the batch and tank flow reactors, there is no mechan-
ical mixing. Thus, and as indicated earlier, the state of the reacting fluid will vary from point
to point in the system, and this variation may be in both the radial r and axial z-direction. The
describing equations are then differential, with position as the independent variable.
In the describing equations to follow, the reacting system is assumed to move through
the reactor in plug flow. It is further assumed that there is no mixing in the axial direc-
tion and complete mixing in the radial direction so that the concentration, temperature,
and so on, do not vary through the cross section of the tube. Thus, the reacting fluid flows
through the reactor in an undisturbed plug of mass. Note once again that time for this
hypothetical plug to flow through this type of reactor is the same as the contact or reac-
tion time in a batch reactor. Under these conditions, the form of the describing equations
for batch reactors will also apply to TF reactors.1,2 In a plug flow reactor, the entire feed
202 Introduction to Optimization for Chemical and Environmental Engineers
stream moves with the same radially uniform axial velocity along parallel streamlines.
The entire feed stream, therefore, has the same residence time since there is no mixing
in the axial direction but (effectively) complete mixing radially. From a qualitative point
of view, as the length of the reactor approaches infinity, the concentration of a (single)
reactant approaches zero for irreversible reactions (except zero order) and the equilibrium
concentration for reversible reactions.
In actual practice, TF reactors deviate from the plug flow model because of the veloc-
ity variations in the radial direction. For any of these conditions, the residence time for
annular elements of fluid within the reactor will vary from some minimal value at a point
where the velocity is a maximum, to a maximum value near the wall where the velocity
approaches zero. The concentration and temperature profiles, as well as the velocity pro-
file, are therefore also not constant across the reactor. The describing equations based on
the plug flow assumption are then not applicable.
The design equation for tubular reactors operating in a steady-state mode will be dif-
ferential in form, unlike the algebraic design equation for CSTRs. The describing equation
is of an ordinary differential form for variations along the axial (longitudinal) length of the
reactor. A partial differential equation arises if there are variations along both the axial and
radial directions.
The key assumption with a CTSR is that it is completely mixed. The key assumption
with a TF unit is that plug flow is present, that is, the fluid is completely unmixed along the
reactor length and flows down the tube as a plug since it is perfectly mixed radially. Even
though the plug-flow approximation may not apply rigorously in some applications, the
simplicity of solutions in the limit of perfect plug flow makes it a very useful model from
both an academic and industrial perspective.
Ultimately, the performance of the three different classes of reactor is based on a host of
considerations. They include
1. Conversion
2. Production rate
3. Volume requirement(s)
4. Selectivity (and the effect of unwanted by-products)
Energy considerations, space, power requirements, and so on, also play a role but super-
imposed on all of the previously mentioned is the economics associated with the choice.
Walas4, in an earlier article, provides some general “rules of thumb” regarding reactor
selection. Some of his (unedited) suggestions are noted here:
1. The rate of reaction in every instance must be established in the laboratory, and
the residence time or space velocity and product distribution must be found in a
pilot plant.
2. The optimum proportions of stirred tank reactors are a liquid level equal to tank
diameter; at high pressures, slimmer proportions are more economical.
3. Power input to a homogenous-reaction stirred tank is 0.5–1.5 hp/1000 gal, but
input is three times this amount when heat is to be transferred.
4. CSTR behavior is approached when the mean residence time is 5–10 times the time
needed to achieve homogeneity, which is accomplished by 500–2000 revolutions of
a properly designed stirrer.
Chemical Reactor Applications 203
5. Batch reactions are conducted in stirred tanks for small, daily production rates,
or when the reaction times are long, or when some condition, such as feed rate or
temperature, must be programmed in some way.
6. Relatively slow reactions of liquids and slurries are conducted in continuous
stirred tanks. A battery of four or five units in series is most economical.
7. TF reactors are suited to high production rates at short residence times (seconds
or minutes) and when substantial heat transfer is needed. Jacketed tubes or shell-
and-tube construction are used.
8. For conversions under about 95% of equilibrium, the performance of a five-stage
CSTR unit approaches plug flow.
A → products
k1C1A/2
−rA = (14.1)
1 + k 2CA2
where:
k1 = 5 (gmol/L)1/2/h
k2 = 10(L/gmol)2
SOLUTION:
1. The rate of reaction, −rA, in terms of the conversion variable X is
C A = C A 0 (1 − X ) ; X = X A (14.2)
so that
k1C A1/02 (1 − X )
1/2
−rA = (14.3)
1 + k 2C A2 0 (1 − X )
2
FA 0 X
V= ; X = 0.6
− rA
(14.4)
1 + (10 )(0.5)2 (1 − 0.6)2
= FA 0 X
( 5)(0.5) (1 − 0.6)1/2
1/2
204 Introduction to Optimization for Chemical and Environmental Engineers
1 + (10 )( 0.5 ) (1 − 0.6)
2 2
V = ( 200 )( 0.6 ) 1 1 ; X = 0.6
( 5 )( 0.5 ) 2 (1 − 0.6) 2
= 75.1 liters
1 + k 2C A2 0 (1 − X )
0.6 2
I=
∫
0
(1 − X )1/2
dX
(14.6)
0.6
=
∫ f (X ) dX
0
h
I = f (0.0 ) + ( 4 ) f (0.3 ) + f (0.6 ) h = 0.3
3
0.3
= ( 3.5) + ( 4 )( 2.659) + 2.214
3
= 1.635
The volume is then
FA 0 I
V=
k1C A1/02 (14.7)
= 92.5 liters
Note that the integral could have been evaluated by any one of several methods.
3. Surprisingly, the CSTR, is the optimum choice since it requires a smaller vol-
ume. This result arises because of the unique nature of the rate expression.
bottom of the vessel and forced through the bed. At a low flow rate, the fluid (liquid or
gas) moves through the void spaces between the stationary and solid catalyst particles
and the bed is referred to as fixed. As the flow rate increases, the catalysts begin to vibrate
and move about slightly, resulting in the onset of an expanded bed. When the flow of fluid
reaches a certain velocity, the solid catalysts become suspended because the upward fric-
tional force between the catalyst and the fluid balances the gravity force associated with
the weight of the catalyst. This point is termed minimum fluidization or incipient fluidization
and the velocity at this point is defined as the minimum or incipient fluidization velocity.
Beyond this stage, the bed enters the fluidization state where bubbles of fluid rise through
the solid catalysts, thereby producing a circulatory and/or mixing pattern.5
From a force balance perspective, as the flow rate upward through a packed bed is
increased, a point is reached at which the frictional drag and buoyant force is enough to
overcome the downward force exerted on the bed by gravity. Although the bed is usually
supported at the bottom by a screen, it is free to expand upward, as it will if the velocity is
increased above the aforementioned minimum fluidization velocity. At this point, the cata-
lysts are no longer supported by the screen, but rather are suspended in the fluid and act
and behave as the fluid. The bed is then said to be fluidized. From a momentum or force bal-
ance perspective, the sum of the drag, buoyancy, and gravity forces must be equal to zero.
The terminal settling velocity can be evaluated for the case of flow past one catalyst par-
ticle in the bed. By superimposition, this case is equivalent to that of the terminal velocity
that a catalyst particle would attain flowing through a fluid. Once again, a force balance
can be applied, and empirical data can be used to evaluate a friction (drag) coefficient.1,2
At intermediate velocities between the minimum fluidization velocity and the terminal
velocity, the bed is expanded above the volume that it would occupy at the minimum
value. Note also that above the minimum fluidization velocity, the pressure drop stays
essentially constant.
One of the novel characteristics of fluidized beds is the uniformity of temperature found
throughout the system. Essentially constant conditions are known to exist in both the hori-
zontal and vertical directions in both short and long beds. This homogeneity is due to the
turbulent motion and rapid circulation rate of the solid catalyst particles within the fluid
stream described previously. In effect, excellent fluid-particle contact results. Temperature
variations can occur in some beds in regions where quantities of relatively hot or cold
catalyst particles are present but these effects can generally be neglected. Consequently,
fluidized beds find wide application in industry, that is, oil cracking, zinc coating, coal
combustion, gas desulfurization, heat exchangers, plastics cooling, and fine powder
granulation.
k
A↔R A = waste
k′
Calculate the bed reaction operating temperature that will minimize the volume of
the reactor and achieve the optimum desired degree of waste conversion.
SOLUTION:
A fluid-bed reactor is best described by a CSTR model.1,2 The rate of reaction, −rA, in
terms of the concentration of A, CA, is1,2
V C A1 − C A1
= ; q = volumetric flow rate (14.9)
q −rA
C A 0 − C A1
= (14.10)
kC A1 − k′CR1
where:
V = reactor volume
q = volumetric flowrate of the waste
CA1 and CR1, the outlet concentrations of A and R, respectively, can now be expressed in
terms of CA0. Note that for 99.99% destruction of the waste, A, the conversion variable
X A becomes
X A = 0.9999
Thus,
C A1 = 0.0001C A 0
CR1 = 0.9999C A 0
The design equation for V may now be rewritten in terms of CA0 and the k’s
V C A 0 − 0.0001(C A 0 )
=
q ( k )( 0.0001) C A 0 − ( k′ )( 0.9999 ) CA 0
0.9999(CA 0 ) 0.9999
= =
( 0.0001k − 0.9999k )( CA0 ) 0.0002k − 0.9999k
′
The reaction velocity constants, k and k′, are described by the Arrhenius equation given
in the problem statement. Thus,
k = Ae − E/RT (14.11)
A = 1.0
Chemical Reactor Applications 207
A’ = 9.89
E
= −10, 000
R
E
= −35, 000
R
To calculate the operating temperature that will require the minimum volume to accom-
plish a conversion of 99.99%, minimize the volume by setting dV/dT = 0 and solving for
the temperature:
dk dk′
d (V / q ) dT − ( 9999 ) dT
= ( 0.9999 ) (14.13)
( k − 9999k )
2
dT ′
Setting
d (V / q )
= 0 (14.14)
dT
dk dk′
= ( 9999 )
dT dT
dk′
( )
≈ 10 4
dT
Since
dk AE − E/RT
= e (14.15)
dT RT 2
E′ − E
A E
10 4 = e RT
A ′ E′
−35 , 000 ′ − −10 , 000
( ) ( )
1.0 −10, 000
4
10 = e T
9.89 −35, 000
25 , 000
−
346, 150 = e T
T = 1960 °R
= 1500°F
This represents the optimum operating temperature to minimize the volume of the
reactor.
208 Introduction to Optimization for Chemical and Environmental Engineers
SOLUTION:
Write the design equation for a CSTR in terms of the concentration (not the conversion)
and the residence time τ.1,2
= qC A0 ( X − 0 ) / kC A1 (14.18)
τ = (C A0 − C A1 )/ kC A1 (14.19)
C A1 = C A0 / (1 + k τ ) (14.20)
Write the design equation for a TF reactor in terms of the concentration (not the conver-
sion) and the residence time τ.
X
1
∫
V = FA0 −
0
rA
dX (14.21)
X
1
0
∫
V/q = C A0 − dX (14.22)
rA
(a) q
or
(b) q
FIGURE 14.1
Physical arrangements of CSTR and TF.
Chemical Reactor Applications 209
Since C A0 dX = −dC A
C A1
1 1
τ=−
k ∫ C
CA0
dC A
A
(14.23)
C
−k τ = ln A1
CA0
CA1
= exp ( −k τ )
CA0
C A1 = C A0 exp ( −k τ ) (14.24)
Consider the first case (a) – a CSTR followed by a TF reactor – and determine the outlet
concentration C A1 from the CSTR in terms of τ1.
C A1 = C A0 / (1 + k τ1 ) (14.25)
Determine the outlet concentration, C A2 , for the TF reactor. Set τ = τ2 for the TF reactor
C A2 = C A1 exp ( − k τ 2 )
(14.26)
= C A0 exp ( − k τ 2 ) / (1 + τ 2 )
Since τ = τ1 = τ2,
C A2 = C A0 exp ( −k τ ) / (1 + k τ ) (14.27)
Consider the second case (b) – a TF reactor followed by a CSTR – and determine C A1
and C A2.
C A1 = C A0 exp ( −k τ )
C A2 = C A0 / (1 + k τ ) (14.28)
= C A0 exp ( −k τ ) / (1 + k τ )
The previous results are identical; that is, both cases provided the same outlet
conversion.
How would the analysis in the previous example be affected if the volumes of the two
reactors are not equal? One would repeat the previous calculation if the two volumes
are not equal by employing τT and τc, rather than τ. For the first case,
C A2 = C A0 exp ( −k τT ) / (1 + k τC ) (14.29)
C A2 = C A0 exp ( −k τT ) / (1 + k τC ) (14.29)
may affect the performance. The reader is left the exercise of verifying this statement;
however, for a zero order or ∞ order reaction, that is, for n equal to either zero or ∞, the
“performance” or degree of conversion would still be identical.
14.4 Maximizing Selectivity
When multiple reactions (either parallel or consecutive) occur, the side or undesired reac-
tions compete with the main or desired reaction; the less predominant the main reaction
is, the smaller the amount of desired product is for a given amount of reactant. Suppose
the following hypothetical parallel reactions occur:
A→B
A → C (14.30)
A→D
Suppose also that, with a starting amount of 10 mol of A, four form the desired product B,
two form the undesired product C, one forms the undesired product D, and three remain
unreacted. While this process produces 4 mol of valuable product, it could have produced
10 if everything went the way a practicing engineer would want it to; that is, if all 10 mol
of A reacted to form 10 mol of B. The ratio of the 4 mol of B actually produced, to the 10 it
potentially could have produced, is called the yield (0.40 or 40%). By definition, the yield of
a reaction is a measure of how much of the desired product is produced relative to how
much would have been produced if only the desired reaction occurred and if that reaction
went to completion. Obviously, the fractional yield must be a number between zero and
infinity. Another definition that has been used by industry is the “ratio of product formed
to the initial (inlet) reactant.” An additional term used in conjunction with multiple reac-
tions is selectivity. Selectivity is a measure of how predominant the desired reaction is rela-
tive to one of the undesired products produced by a side reaction. Obviously, selectivity
becomes more important than conversion if “undesired” side reaction(s) take precedence
over “desired” reaction(s). Other definitions of selectivity that have appeared in the litera-
ture include the popular “ratio of the amount of desired products formed to the amount
of all products formed.” Definitions have also included “rate of production of one product
relative to another product (point selectivity)” and “amount of one product formed relative
to another product (overall selectivity).” In the previously mentioned example, the selec-
tivity of B over C is 2.0 and that of B over D is 4.0. These definitions are employed in the
illustrative example to follow.
kD
A → D ( desirable ) (14.31)
kU
A → U (undesirable) (14.32)
Chemical Reactor Applications 211
with
rU = kU C Aα2 (14.34)
and
−rA = rD + rU (14.35)
Suggest procedures that can be employed to maximize the point selected, S, where
S = rD/rU.
SOLUTION:
If one desires rD ≫ rU, then the point (or local) selectivity is
rD kD α1 −α2
S= = CA (14.37)
rU kU
If α1 − α2 = a, then
k a
S= D CA (14.38)
kU
To maximize S, the need to maintain the product of both terms in Equation 14.38 is
high. If a is positive, one needs to maximize CA. This generally means no inert(s), no
product(s), and for gas phase reactions to operate at low temperature and high pressure.
If a is negative, the reverse applies.
The effect of temperature is handled through the Arrhenius equation:
1
kD AD − RT ( ED −EU )
= e (14.39)
kU AU
If ED > EU, the reaction should be run at the highest temperature economically possible.
If ED < EU, it should be run at the lowest temperature possible (while maintaining a rea-
sonable rate).
The reader should also attempt to analyze the following application:
k1
A + B → D (14.40)
k2
A + B → D (14.41)
where:
rU = k 2C Aα 2CBβ 2 (14.43)
212 Introduction to Optimization for Chemical and Environmental Engineers
rD k1 α1 −α2
S= = (
rU k 2
CA)( )
CBβ1 −β2 (14.44)
k2
A + D ↔ U (14.46)
k3
where D is the desired product and U is an undesired by-product. This problem is also
left as an exercise for the reader.
1. Isothermally
2. Non-isothermally
3. Adiabatically
4. Any combination of the above
Chemical Reactor Applications 213
In (1), the temperature is maintained constant during the course of the reaction. This con-
dition can be approached by providing sufficient heat exchanger equipment to account
for enthalpy effects arising during reaction. This mode of operation finds its major appli-
cation in laboratory kinetic studies. Most industrial reactors are described by (2). Some
energy in the form of heat is added to or removed from the reactor, but isothermal condi-
tions are not satisfied. For (3), the reactor is insulated to minimize heat transfer between
the reactor contents and the surroundings. Many industrial reactors attempt to operate
in this manner.
As noted previously, batch reactors are commonly used in experimental studies. Their
industrial applications are somewhat limited. They are seldom used for gas phase reac-
tions, for example, combustion, since small quantities (mass) of product are produced with
even a very large-sized reactor. It is primarily used for liquid phase reactions when small
quantities of reactants are to be processed and thus finds limited application in industry.
As a rule, batch reactors are less expensive to purchase but more expensive to operate than
either CSTRs or TF reactors.
= t + td ; t = treact
Expand the previously mentioned development in terms of maximizing the production
rate in a batch reactor.6.
SOLUTION:
For the reactor
aA → pP
N A0 X A
PR ( A ) = (14.48)
t + td
and is a function of both X and t. Normally the production rate refers to one of the
products formed. This production rate, using the same conversion variable, is then (for
product P)
pN X
PR ( P ) = PR = A0 A (14.49)
a t + td
where p and a are the stoichiometric coefficients in the reaction equation for species P
and A, respectively.
To maximize production, one needs to express PR in terms of X (replace t by the
design equation that contains X) and solve the equation
214 Introduction to Optimization for Chemical and Environmental Engineers
d ( PR )
= 0 (14.50)
dX A
or alternatively
d ( PR )
= 0 (14.51)
dt
For example, if the stoichiometric reaction equation is given by
kA
A → products
then
X A = 1 − e − k At
or
ln (1 − X A )
t=− (14.52)
kA
To maximize P, one can either set
d ( PR )
; X A = 1 − e − k At (14.53)
dt
or
d ( PR ) ln (1 − X A )
; t=− (14.54)
dX A kA
equal to zero and then solve for either X A or t.
One should check to see that the second-order derivative is negative, indicating that
the solution is a maximum and not a minimum.
1. There are three major costs associated with each reactor: labor (engineers),
labor (operators), and raw materials.
2. Reactors A, B, and C require 3, 1, 1 engineers per tank of product per week
respectively, but are limited to a maximum of 30 engineers.
Chemical Reactor Applications 215
3. Weekly labor requirements for operators for reactors A, B, and C are 3, 2, and 4
operators per tank of product respectively, but are limited to a maximum of 82
operators.
4. Only 110 tons of raw material is available on a weekly basis that serves as feed
to the reactors. Each tank of product A, B, and C produced require 2, 4, and 5
tons of raw material, respectively.
5. The income, I, of this chemical operation for products A, B, C are 400, 500, and
800 $/tank, respectively. Develop a model that will maximize income.
SOLUTION:
The objective function I is given by:
3 x1 + x2 + x3 ≤ 30 ( engineers ) (14.56)
with
x1 ≥ 0 (14.59)
x2 ≥ 0 (14.60)
x3 ≥ 0 (14.61)
This problem can be solved using Excel, which gives the weekly outcome of x1 = 3, x2 = 8,
x3 = 15, and I = $17,200.
References
1. L. Theodore, Chemical Reaction Kinetics, A Theodore Tutorial, Theodore Tutorials, East Williston,
NY, originally published by the USEPA/APTI, RTP, NC, 1992.
2. L. Theodore, Chemical Reactor Analysis and Applications for the Practicing Engineer, John Wiley &
Sons, Hoboken, NJ, 2012.
3. S. Fogler, Elements of Chemical Reaction Engineering, 4th edn, Prentice-Hall, Upper Saddle River,
NJ, 2006.
4. S. Walas, Rules of Thumb: Selecting and Diagnosing Equipment, Chem. Eng., New York City, NY,
1987.
5. D. Kunii and O. Levenspieal, Fluidization Engineering, John Wiley & Sons, Hoboken, NJ, 1969.
6. L. Theodore, personal notes, Manhattan College, Bronx, NY, 1967.
15
Mass Transfer Applications
The third chapter in Part III contains six illustrative examples concerned with the general
subject of mass transfer. The titled and one-line description of each example is provided
here:
Seven references are provided for suggested reading for this chapter:
1. R. Perry and D. Green, Perry’s Chemical Engineers’ Handbook, 8th edn, McGraw‑Hill,
New York City, NY, 2008.
2. L. Theodore, Chemical Engineering: The Essential Reference, McGraw-Hill, New York
City, NY, 2014.
3. J. Reynolds, J. Jeris, and L. Theodore, Handbook of Chemical and Environmental
Engineering Calculations, John Wiley & Sons, Hoboken, NJ, 2004.
4. L. Theodore, R. Dupont, and K. Ganesan, Unit Operations in Environmental
Engineering, Wiley‑Scrivener, Salem, MA, 2018.
5. L. Theodore, F. Ricci, and T. VanVliet, Thermodynamics for the Practicing Engineer,
John Wiley & Sons, Hoboken, NJ, 2009.
6. L. Theodore and F. Ricci, Mass Transfer Operation for the Practicing Engineer, John
Wiley & Sons, Hoboken, NJ, 2020.
7. R. Treybal, Mass Transfer Operations, McGraw-Hill, New York City, NY, 1955.
217
218 Introduction to Optimization for Chemical and Environmental Engineers
Distillation may be defined as the separation of the components of a liquid feed mixture
by a process involving partial vaporization through the application of heat. In general,
the vapor evolved is recovered in liquid form by condensation. The more volatile (lighter)
components of the liquid mixture are obtained in the vapor discharge at a higher concen-
tration. The extent of the separation is governed by two important factors: the properties of
the components involved and the physical arrangement of the unit for distillation.
In continuous distillation, a feed mixture is introduced to a column where vapor rising
up the column is contacted with liquid flowing downward (which is provided by condens-
ing the vapor at the top of the column). This process removes or absorbs the less volatile
(heavier) components from the vapor, thus effectively enriching the vapor with the more
volatile (lighter) components. This occurs in the section above the feed stream, which is
referred to as the enriching or rectification section of the column. The product (liquid or
vapor) removed from the top of the column is rich in the more volatile components and
is defined as the distillate. The section below the feed stream is referred to as the stripping
section of the column. In this section, the liquid is stripped of the lighter components by
the vapor produced in a reboiler at the bottom of the column, called the bottoms, which is
richer in the heavier components.
Distillation columns are used throughout industry when mixtures (primarily in liquid
form) must be separated. One such example is the petroleum industry. In such an applica-
tion, crude oil is fed into a large distillation column and different fractions (oil mixtures of
varying composition and volatility) are withdrawn at different heights in the column. Each
fraction, such as jet fuel, home heating oil, gasoline, and so on, is used by both industry and
the consumer in a variety of ways. The separation achieved in a distillation column depends
primarily on the relative volatilities1,2 of the components to be separated, the number of con-
tacting trays (plates) or packing height, and the ratio of liquid and vapor flow rates.
Distillation columns are rarely designed with packing in large-scale production because
of the liquid distribution problems that arise with large diameter units and the enormity
of the height of many columns. However, where applicable, towers filled with packing are
competitive in cost, and are particularly useful in cases where the pressure drop must
be low, and/or the liquid holdup must be small. Packed towers are occasionally used for
bench-scale or pilot plant work. In contrast, use of trayed towers extends to many areas of
the chemical industry.
SOLUTION:
As indicated in Figure 15.1, the optimum reflux ratio occurs at the point where the sum
of fixed charges and operating costs is a minimum. As a rough approximation, the opti-
mum reflux ratio usually falls in the range of 1.1–1.5 times the minimum reflux ratio.
This figure illustrates the general method for determining the optimum reflux ratio in
distillation operations.
Mass Transfer Applications 219
100,000
60,000
Annual cost
FIGURE 15.1
Optimum reflux ratio in distillation operation.
( )
V f2 = 2.25 × 10 4 ( t f + 0.11) (15.1)
The cake formed in each cycle must be washed with an amount of water equal to
one-sixteenth times the volume of filtrate delivered per cycle. The washing rate remains
constant and is equal to one-fourth of the filtrate delivery rate at the end of the filtra-
tion. The time required per cycle for dismantling, dumping, and reassembling is six
hours. Under the conditions where the preceding information applies, determine the
total cycle time necessary to permit the maximum output of filtrate during each 24-hour
cycle.
SOLUTION:
Let tf = hours of filtering time per cycle and filtrate delivered per cycle = Vf,cycle.
( )
1/2
V f , cycle = 150 t f + 0.11 ; ft 3 (15.2)
Mass Transfer Applications 221
1
dV f 150
Vw × 4 =
dt f
=
2
( t f + 0.11) 2 ; ft 3 (15.3)
t f + 0.11
= ; h (15.5)
2
24
Cycles per 24 h = (15.7)
1.5t f + 6.06
1
24
( )
V f , cycle ( cycles per 24 h ) = 150 t f + 0.11 2 (15.8)
1.5t f + 6.06
At the optimum cycle time giving the maximum output of filtrate per 24-hour,
Performing the differentiation4 (left as an exercise) and solving for tf ultimately yields
t f ,opt = 3.8 h
Design engineers create process designs based on the production capacity for
the plants. However, the optimum rate of production may vary from the “designed
capacity” once the plant is put in operation. For instance, suppose a process has been
designed with a designed capacity of one batch of production every six hours. After
the plant has been put into operation, it has been determined that the total produc-
tion per month can be increased if the time of production of each batch is reduced.
However, if the production time per batch is decreased, more labor hours are required,
per cent conversion of raw materials is reduced, and steam and power costs increase.
While the engineer may have maximized the total production per month at the time
of design, this obvious example demonstrates how price and market conditions con-
stantly change and the process operation can often be improved over time. The follow-
ing analysis indicates the general method for determining economic production rates
and lot sizes.
The total product cost per unit of time is generally divided into the two classifications of
operating costs and organization costs. Operating costs depend on the rate of production and
include expenses for direct labor, raw materials, power, heat, supplies, and similar items,
which are a function of the amount of material produced. Organization costs are due
to expenses for directive personnel, physical equipment, and other services or facilities,
which must be maintained irrespective of the amount of material produced. Organization
costs are independent of the rate of production.4
It is convenient to consider operating costs on the basis of one unit of production. When
this is done, the operating costs can be divided into two types of expenses as follows:
1. Minimum expenses for raw materials, labor, power, and so on that remains con-
stant and must be paid for each unit of production as long as any amount of mate-
rial is produced
2. Extra expenses due to increasing the rate of production
These extra expenses4 are known by some as superproduction costs. They can become par-
ticularly important at high rates of production. Examples of superproduction costs are
extra expenses caused by overload on power facilities, additional labor requirements, or
decreased efficiency of conversion.
SOLUTION:
1. The total cost per membrane is given by
1750 + 7325
cTm = 47.73 + 0.1P1.2 + ; $ membrane (15.10)
P
Mass Transfer Applications 223
dcTm 9075
= 0 = 0.12P 0.2 − 2 ; number of membranes day (15.11)
dP P
Solution for P gives
9075
P 2.2 =
0.12
P = 165 membranes day
The daily profit at a production schedule for minimum cost per membrane is DP ($/day),
where DP is
Therefore, the daily profit for the minimum cost per member is (with P = 165)
9075
DP (min ) = 173 − 47.73 − 0.1(165) −
1.2
165
165
= $4035 $ day
d ( DP )
= 0 = 125.27 − 0.22P1.2 (15.13)
dP
Solving Equation 15.13 yields
The daily profit at production schedule for maximum daily profit is therefore
9075
DP (max ) = 173 − 47.73 − 0.1(198 ) −
1.2
198
198
= $4439 $ day
The life of a cleaning bath can be extended by proper and regular bath maintenance (i.e.,
regular replenishing, pH measurement, metal content measurement, etc.). The life of a plating
bath can also be extended through bath treatment that removes metal contaminants. Copper
is a common metal contaminant that builds up in plating baths. Copper can be removed from
a bath by “dummying.” The dummying process is based on the principle that the copper can
be plated at a low electrical current. When the copper content of a process bath becomes too
high, an electrolyte panel is placed in the bath and a low current (1–2 amperes per square foot)
is then run through the system. At this low current, only the copper in the bath will be plated
out on the panel and other bath additives will remain unaffected by the current.
Process chemical loss as a result of drag-out is the most significant source of chemicals pres-
ent in the wastewater. The factors that contribute to drag-out are the workplace size and shape,
bath viscosity, bath concentration, surface tension, and temperature. The available techniques
of reducing process chemical drag-out are summarized in the following guidelines:
finF f1F
1
r1R rinR
FIGURE 15.2
Flow diagram for one-stage operation.
Mass Transfer Applications 225
SOLUTION:
1. Using the flow diagram provided in Figure 15.2, a material balance for the
residue may be written
This equation may be rearranged in terms of λ. For the single-stage operation, set i = 1:
R ( f in − f1 ) λ ( f in − f1 )
= = (15.16)
F r1 f1
x=
( fin − f1 ) (15.17)
f in
R x
=λ (15.18)
F 1 − x
The flow diagram for a two-stage (stages 1 and 2) countercurrent operation is provided
in Figure 15.3. Material balances on the residue for each stage are
Each of these equations may be solved for R/F/ in terms of λ and f1 using the defining
equation for λ:
R λ ( f in − f1 )
stage 1: = (15.21)
F f1 − f 2
R λ ( f1 − f 2 )
stage 2: = (15.22)
F f2
If one substitutes x into the mass balance expressions,
R f in − f1
stage 1: = (15.23)
F f1 − fin (1 − x )
finF f1F f 2F
FIGURE 15.3
Flow diagram for two-stage countercurrent operation.
226 Introduction to Optimization for Chemical and Environmental Engineers
R f1 − f in (1 − x )
stage 2 : = (15.24)
F f in (1 − x )
The right-hand sides (RHS) of stages 1 and 2 equations may be set equal to each other:
f in − f1 f1 − f in (1 − x )
= (15.25)
f1 − f in (1 − x ) f in (1 − x )
This equation may be rearranged to obtain a quadratic equation in f1:
{ }
0.5
2
f in (1 − x ) ± − f in (1 − x ) − 4 (1) x (1 − x ) f in
2
f1 = (15.27)
2
This can be rewritten as
f
f1 = in 1 − x ± ( 3 x + 1) (1 − x ) (15.28)
0.5
2
The following expression results if (1 − x) is factored out:
f in (1 − x ) 3 x + 1 0.5
f1 = 1 ± 1 − x (15.29)
2
f in (1 − x ) 3 x + 1 0.5
f1 = 1 + 1 − x (15.30)
2
The term f1 may be substituted into the stage 2 mass balance equation. Equation 15.31
results:
f in (1 − x ) 3 x + 1 0.5
λ 1 + − f2
R 2 1 − x
= (15.31)
F f2
f 2 = f in (1 − x ) (15.32)
f in (1 − x ) 3 x + 1 0.5
λ 1 +
− f in (1 − x )
R 2 1− x
= (15.33)
F f in (1 − x )
Mass Transfer Applications 227
which reduces to
R λ 3 x + 1
0.5
= − 1 (15.34)
F 2 1 − x
The rinse water requirements for both a single and two-stage countercurrent unit
becomes
0.99
Single − stage : R = λ F = 99.0lambaF (15.35)
1 − 0.99
λ 3 (0.99) + 1
0.5
99.0λF − 9.46λF
= 0.904 = 90.4% reduction (15.37)
( 99.0λF )
2. For fixed residue removal, the total mass of metals in the rinse water will be the
same; with reduced water duty, the metals concentration increases. A smaller
water volume now needs to be processed (or sewered). Alternatively, if the
metals concentration is sufficiently high, it can be returned to the plating bath
(for reuse or can be recovered) to further optimize performance.
have to be met. For example, if the pressure drop through the column is large enough such
that horsepower costs become significant, a packed column may be preferable to a plate-
type column because of the lower pressure drop. Again, primary emphasis in this section
is placed on packed and plate columns.1–4
In most processes involving the absorption of gaseous constituents from a gas stream,
the gas stream is the process fluid; hence, its inlet conditions (flow rate, composition, and
temperature) are usually known. The objectives, then, in the design of an absorption col-
umn, are the determination of the solvent (liquid) flow rate and the calculation of the prin-
cipal dimensions of the equipment (column diameter and height).
The general design procedure consists of a number of steps that have to be taken into
consideration (details of which follow shortly). These include1,2
1. Solvent selection.
2. Equilibrium data evaluation.
3. Estimation of operating data (usually obtained from a mass and energy balance,
where the energy balance determines whether the absorption process can be con-
sidered isothermal or adiabatic).
4. Column selection (should the column selection not be obvious or specified, calcu-
lations must be carried out for the different types of columns, and the final selec-
tion based on economic considerations).
5. Calculation of column diameter (for packed columns that is usually based on
flooding conditions, and for plate columns it is based on the optimum gas velocity
or the liquid handling capacity of the plate).
6. Estimation of the column height or the number of plates (for a packed column, the
column height is obtained by multiplying the number of transfer units, obtained
from a knowledge of equilibrium and operating data by the height of a transfer
unit; for plate columns, the number of theoretical plates, often determined from
the plot of equilibrium and operating lines, is divided by the estimated overall
efficiency to give the number of actual plates, which in turn allows the column
height to be estimated from the plate spacing).
7. Determination of pressure drop through the column (for packed columns, correla-
tions dependent on packing type, column operating data, and physical properties
of the constituents involved need to be available to estimate the pressure drop
through the packing; for plate columns, the pressure drop per plate is obtained
(often estimated) and multiplied by the number of plates).
1.
Gas solubility: A high gas solubility is desired since that increases the absorption
rate and minimizes the quantity of solvent necessary; generally, a solvent of a
chemical nature similar to that of the solute to be absorbed will provide good
solubility.
2.
Volatility: Low solvent vapor pressure is desired since the gas leaving an absorp-
tion unit is ordinarily saturated with the solvent and much will therefore be lost.
3.
Corrosiveness.
Mass Transfer Applications 229
4.
Cost (particularly for solvents other than water).
5.
Viscosity: Low viscosity is preferred for reasons of rapid absorption rates,
improved flooding characteristics, lower pressure drops, and good heat transfer
characteristics.1,2
6.
Chemical stability: The solvent should be chemically stable and, if possible,
non-flammable.
7.
Toxicity.
8.
Low freezing point: If possible, a low freezing point is favored since any solidifica-
tion of the solvent in the column could prove disastrous.
Once the solvent is specified, the choice (and design) of the absorption system may be
determined.
Extraction is a term that is used for any operation in which a constituent of a liquid or a
solid is transferred to another liquid (the solvent). The term liquid–liquid extraction describes
the processes in which both phases in the mass transfer process are liquids. The terms
solid–liquid extraction is restricted to those situations in which a solid phase is present and
includes those operations frequently referred to as leaching, lixiviation, and washing. These
terms are subsequently used interchangeably.
Extraction involves the following two steps: contact of the solvent with the liquid or solid
to be treated so as to transfer the soluble component (solute) to the solvent, and separation
or washing of the resulting solution. The complete process may also include a separate
recovery procedure involving the solute and solvent; this is normally accomplished by
another operation, such as evaporation, distillation, or stripping. Thus, the streams leav-
ing the extraction system usually undergo a series of further operations before the fin-
ished product is obtained; either one or both solutions may contain the desired material.
In addition to the recovery of the desired product or products, recovery of the solvent for
recycling is also often an important consideration.
In practice, the manner and the equipment in which these operations are carried out are
based on the difference in physical states. Because solids are more difficult to handle and
do not readily lend themselves to continuous processing, leaching is commonly accom-
plished in a batch-wise fashion by agitating the crude mixture with the leaching agent and
then separating the residual insoluble from the resultant solution. Liquid–liquid extraction
may also be carried out in a batch operation. The ease of moving liquids, however, makes
liquid extraction more amenable to continuous flow in various types of columns and/or
stages.
Liquid–liquid extraction is used for the removal and recovery of primarily organic sol-
utes from aqueous and non-aqueous streams. Concentrations of solute in these streams
range from either a few hundred parts per million to several mole mass percent. Most
organic solutes may be removed by this process. Extraction has been specifically used
in removal and recovery of phenols, oils, and acetic acid from aqueous streams, and in
removing and recovering freons and chlorinated hydrocarbons from organic streams.
Treybal6 has described the liquid–liquid extraction process in the following manner.
“If an aqueous solution of acetic acid is agitated with a liquid such as ethyl acetate, some
of the acid but relatively little water will enter the ester phase. Since the densities of the
aqueous and ester layers are different at equilibrium, they will settle on cessation of
agitation and may be decanted from each other. Since the ratio of acid to water in the
ester layer is now different from that in the original solution and also different from that
in the residual water solution, a certain degree of separation has occurred. This is an
230 Introduction to Optimization for Chemical and Environmental Engineers
example of stage-wise contact and it may be carried out either in a batch or continuous
fashion. The residual water may be repeatedly extracted with more ester to additionally
reduce the acid content.” As will be discussed shortly, one may arrange a countercurrent
cascade of stages to accomplish the separation. Another possibility is to use some sort
of countercurrent or crosscurrent continuous-contact device where discrete stages are
involved.
The solution whose components are to be separated is the feed to the process. The
feed is composed of a dilutant and solute. The liquid contacting the feed for purposes of
extraction is referred to as the solvent. If the solvent consists primarily of one substance
(aside from small amounts of residual feed material that may be present in a recycled
or recovered solvent), it is called a single solvent. A solvent consisting of a solution of one
or more substances chosen to provide special properties is a mixed solvent. The solvent-
lean, residual feed solution, with one or more constituents removed by extraction, is
referred to as the raffinate. The solvent-rich solution containing the extracted solute(s) is
the extract.
The degree of separation that arises because of the aforementioned solubility difference
of the solute in the two phases may be obtained by providing multiple-stage countercur-
rent contacting and subsequent separation of the phases, similar to a distillation operation.
In distillation, large density differences between the gas–liquid phases are sufficient to
permit adequate dispersion of one fluid in the other as each phase moves through the col-
umn. However, in liquid extraction, the density differences are significantly smaller and
mechanical agitation of the liquids is frequently employed at each stage to increase contact
and to increase the mass transfer rates.
The minimum requirement of a liquid extraction unit is to provide intimate contact
between two relatively immiscible liquids for the purposes of mass transfer of con-
stituents from one liquid phase to the other, followed by the aforementioned physi-
cal separation of the two immiscible liquids. Any device or combination of devices
that accomplishes this is defined in this text as a stage. If the effluent liquids are in
equilibrium, so that no further change in concentration would have occurred within
them after longer contact time, the stage is considered a theoretical or ideal stage. The
approach to equilibrium stage provides a mechanism by which two immiscible phases
intimately mix until equilibrium concentrations are reached and then physically sep-
arated into clear layers. A multi-stage cascade is a group of stages usually arranged
in a countercurrent flow between stages for the purpose of enhancing the extent of
separation.
TABLE 15.1
Economic Data Extraction vs. Absorber Comparison
Costs/Credits Extraction Absorber
Mass transfer unit $750,000 $800,000
Peripherals $1,875,000 $2,175,000
Total capital $2,625,000 $2,975,000
Installation $1,575,000 $1,700,000
Operation $400,000/yr $550,000/yr
Maintenance $650,000/yr $775,000/yr
Income $2,000,000/yr $2,500,000/yr
SOLUTION:
Calculate the capital recovery factor (CRF):
( 0.12 ) (1 + 0.12 )
12
CRF =
(1 + 0.12 ) − 1
12
(15.38)
= 0.1614
Determine the annual capital and installation costs for the extraction (EXT) unit:
TABLE 15.2
Cost Analysis for Illustrative Example 15.5
Extraction Absorber
Total installed ($/yr) 678,000 755,000
Operation ($/yr) 400,000 550,000
Maintenance ($/yr) 650,000 775,000
Total annual cost ($/yr) 1,728,000 2,080,000
Income credit ($/yr) 2,000,000 2,500,000
232 Introduction to Optimization for Chemical and Environmental Engineers
The reader should note that detailed cost estimates are beyond the scope of this text.
Such procedures are capable of producing accuracies in the neighborhood of ±5%.
However, such estimates generally require many months of engineering work. This type
of analysis is designed to give the reader a basis for a preliminary cost analysis only.
TABLE 15.3
Crude Oil Usage and Production Information
North Texas (%) Venezuelan (%) Maximum Production Rate (gal/day)
Diesel fuel 8 11 1,500
Home heating oil 29 54 5,500
Gasoline 63 35 11,000
SOLUTION:
Set NA and NB equal to the daily usage rate of North Texas and Venezuelan crude. One
may then write the following based on the problem statement:
Constraints
N A ≥ 0 (15.43)
N B ≥ 0 (15.44)
References
1. L. Theodore and J. Barden, Mass Transfer Operations, A Theodore Tutorial, Theodore Tutorials,
East Williston, NY, originally published by USEPA/APTI, RTP, NC, 1995.
2. L. Theodore and F. Ricci, Mass Transfer Operations for the Practicing Engineer, John Wiley & Sons,
Hoboken, NJ, 2010.
3. R. Treybal, Mass Transfer Operations, McGraw-Hill, New York City, NY, 1955.
4. M. Peters, Plant Design and Economics for Chemical Engineers, McGraw-Hill, New York City, NY,
1958.
5. L. Theodore and R. Allen, Pollution Prevention, A Theodore Tutorial, Theodore Tutorials, East
Williston, NY, originally published by USEPA/APTI, RTP, NC, 1956.
6. R. Treybal, Liquid Extraction, McGraw-Hill, New York City, NY, 1956.
7. K. Skipka and L. Theodore, Energy Resources: Availability, Management, and Environmental
Impacts, CRC Press/Taylor & Francis Group, Boca Raton, FL, 2014.
16
Heat Transfer Applications
The fourth chapter in Part III contains six illustrative examples concerned with the gen-
eral subject of heat transfer (HT). The title and one-line description of each example are
provided here:
Six references are provided below for suggest reading for this chapter:
1. R. Perry and D. Green, Perry’s Chemical Engineers’ Handbook, 8th edn, McGraw‑Hill,
New York City, NY, 2008.
2. L. Theodore, Chemical Engineering: The Essential Reference, McGraw-Hill, New York
City, NY, 2014.
3. J. Reynolds, J. Jeris, and L. Theodore, Handbook of Chemical and Environmental
Engineering Calculations, John Wiley & Sons, Hoboken, NJ, 2004.
4. L. Theodore, R. Dupont, and K. Ganesan, Unit Operations in Environmental
Engineering, Wiley‑Scrivener, Salem, MA, 2018.
5. L. Theodore, Heat Transfer Applications for the Practicing Engineer, John Wiley &
Sons, Hoboken, NJ, 2011.
6. A. M. Flynn, T. Akashige, and L. Theodore, Kern’s Process Heat Transfer, 2nd edn,
Wiley‑Scrivener, Salem, MA, 2018.
16.1 Steam Options
The term phase for a pure substance refers to a state of matter that is gas, liquid, or solid.
Latent enthalpy effects are associated with phase changes. These phase changes involve
no change in temperature, but there is a transfer of energy to/from the substance. There
are three possible latent effects, as detailed here:
1. Vapor–liquid
2. Liquid–solid
3. Vapor–solid
235
236 Introduction to Optimization for Chemical and Environmental Engineers
Vapor–liquid changes are referred to as condensation when the vapor is condensing and
vaporization when liquid is vaporizing. Liquid–solid changes are referred to as melting
when the solid melts to liquid and freezing when a liquid solidifies. Vapor–solid changes
are referred to as sublimation. One should also note that there are enthalpy effects associ-
ated with a phase change of a solid to another solid form; however, this enthalpy effect is
small compared to the other effects mentioned previously. Finally, vapor–liquid changes
are of primary interest to the practicing environmental and chemical engineer. This leads
to the subject of steam tables.
Water is one of the most abundant substances in the earth’s crust and its properties
have been determined with considerable accuracy. Since water is used in most chemical
operations as a coolant in the liquid form or as a heating medium in the form of steam, the
environmental and chemical engineer should be familiar with the compilation of data on
water known as the Steam Tables. These tables are available in the literature.1–3
• 1998 = 230
• 2016 = 360
• Capital cost (C) = (BC)(IF)(PF)(CI).
SOLUTION:
Calculate the overall heat duty:
Calculated log-mean temperature differences for each case are provided in Table 16.2.
5, 400, 000
A150 = = 436 ft 2
(138 )( 89.8 )
5, 400, 000
A 300 = = 244 ft 2
(138 )(160.5 )
Determine the capital cost for each case:
360
Cost 150 = (117 )( 436 ) ( 3.29)(1.15) = $36, 000
0.65
230
360
Cost 300 = (117 )( 244 ) ( 3.29)(1.20 ) = $25, 800
0.65
230
Obtain the steam requirement for each case in lb/yr:
St150 =
( 5, 400, 000 Btu/h )(8760 × 0.9 h/yr ) = 49.3 million lb/yr
863.6 Btu/lb
St 300 =
( 5, 400, 000 Btu / h )(8760 × 0.9 hr / yr) = 52.6 million lb / yr
809.0 Btu / lb
Use the previous calculations to determine the annual steam cost for each case:
( )
StCost 150 = 49.3 × 10 6 lb / yr ( 0.00520 $ / lb ) = $256, 000 / yr
( )
StCost 300 = 52.6 × 10 6 lb / yr ( 0.00575 $ / lb ) = $303, 000 / yr
The 300-psia exchanger costs $10,200 less to purchase and install, but it costs $47,000
per year more to operate. Choosing the more expensive 150-psia exchanger is the obvi-
ous choice.
TABLE 16.1
Steam Data
150 psia 300 psia
Saturation temperature, °F 358.0 417.0
Latent heat (enthalpy), Btu/lb 863.6 809.0
Cost, $/1000 lb 5.20 5.75
TABLE 16.2
Log-Mean Temperature Differences Calculations
150 psia case 300 psia case
ΔT1 358 − 150 = 208 417 − 150 = 267
ΔT2 358 − 330 = 28 417 − 330 = 87
ΔtLM = LMTD 89.8°F 160.5°F
238 Introduction to Optimization for Chemical and Environmental Engineers
1. Ruggedness
2. Convenience
3. Volume and weight
4. Ease of handling
5. Thermal effectiveness
6. Cost
1. Vacuum insulation, which employs an evacuated space that reduces radiant heat
transfer
2. Multilayer insulation, referred to by some as superinsulation, which consists of
alternating layers of highly reflective material and low conductivity insulation in
a high vacuum
3. Powder insulation, which utilizes finely divided particulate material packed
between surfaces
4. Foam insulation, which employs non-homogeneous foam whose thermal conduc-
tivity depends on the amount of insulation
5. Special insulation, which includes composite insulation that incorporates many of
the advantageous qualities of the other types of insulation
Heat Transfer Applications 239
SOLUTION:
The cost of heat loss is given as
= $0.922
The insulation that should be chosen, based on Table 16.4, is in the two-inches range.
A plot of these values may also be performed; it would result in a graph similar to that
seen in Figure 16.1.
TABLE 16.4
Cost Data; Illustrative Example 16.3
Insulation Initial Annual Cost of Total
Thickness Installed Fixed Annual Heat Loss Annual Heat Annual
(inches) Cost ($) Cost ($) (Million Btu)* Loss ($) Cost ($)
0 0 0 5,260 4,860 4,860
1 6,000 1,200 2,620 2,420 3,620
2 12,000 2,400 1,050 970 3,370
3 18,000 3,600 524 484 4,084
4 24,000 4,800 175 162 4,962
6 36,000 7,200 61 56 7,256
* There are 8760 h per yr.
Total cost
Total annual cost $
Fixed costs
Insulation thickness
FIGURE 16.1
Optimum thickness of insulation; consistent units.
compact design, greater ease of cleaning, and less possibility of leakage, the double pipe
heat exchanger still finds use in practice.
The double pipe unit consists of two concentric pipes. Each of the two fluids – hot and
cold – flow either through the inside of the inner pipe or through the annulus formed
between the outside of the inner pipe and the inside of the outer pipe. Generally, it is more
economical (from a heat efficiency perspective) for the hot fluid to flow through the inner
pipe and the cold fluid through the annulus, thereby reducing heat losses to the surround-
ings. In order to ensure sufficient contacting time, pipes longer than approximately 20 ft
are extended by connecting them to return bends. The length of pipe is generally kept to a
maximum of 20 ft because the weight of the piping may cause the pipe(s) to sag. Sagging
may allow the inner pipe to touch the outer pipe, distorting the annulus flow region and
disturbing proper operation. When two pipes are connected in a “U” configuration by a
return bend, the bend is referred to as a hairpin. In some instances, several hairpins may
be connected in series.
Double pipe heat exchangers have been used in the chemical process industry for over
90 years. The first patent on this unit appeared in 1923. The unique (at that time) design
provided “the liquid to be heated or cooled and flow longitudinally and transversely
around a tube containing the cooling or heating liquid, etc.” The original design has not
changed significantly since that time. Although this unit is not extensively employed in
Heat Transfer Applications 241
industry (the heat transfer area is small relative to other heat exchangers), it serves as an
excellent starting point from an academic and/or training perspective in the treatment of
all the various heat exchangers.
Shell-and-tube (also referred to as tube and bundle) heat exchangers provide a large heat
transfer area economically and practically. The tubes are placed in a bundle and the ends
of the tubes are mounted in tube sheets. The tube bundle is enclosed in a cylindrical shell,
through which the second fluid flows. Most shell-and-tube exchangers used in practice are
of welded construction. The shells are built as a piece of pipe with flanged ends and neces-
sary branch connections. The shells are made of seamless pipe up to 24 inches in diameter;
they are made of bent and welded steel plates if above 24 inches. Channel sections are
usually of built-up construction, with welding-neck forged-steel flanges, rolled-steel bar-
rels, and welded-in pass partitions. Shell covers are either welded directly to the shell, or
they are built-up constructions of flanged and dished heads and welding-neck forged-steel
flanges. The tube sheets are usually non-ferrous castings in which the holes for inserting
the tubes have been drilled and reamed before assembly. Baffles can be employed to both
control the flow of the fluid outside the tubes and provide turbulence.
There are vast industrial uses of shell-and-tube heat exchangers. These units are used
to heat or cool process fluids, either through a single-phase heat exchanger or a two-phase
heat exchanger. In single-phase exchangers, both the tube side and shell side fluids remain
in the same phase that they enter. In two-phase exchangers (examples include condensers
and boilers), the shell side fluid is usually condensed to a liquid or heated to a gas, while
the tube-side fluid remains in the same phase.
Generally, shell-and-tube exchangers are employed when double pipe exchangers do not
provide sufficient area for heat transfer. Shell-and-tube exchangers usually require fewer
materials of construction and are usually more economical when compared to double pipe
and/or multiple double pipe heat exchangers in parallel.
TABLE 16.5
Costs/Credits Data; Illustrative Example 11.3
Costs/Credits Double Pipe (DP) Shell-and-tube (ST) Extended Surface (ES)
Capital ($) 26,250 29,750 34,350
Installation ($) 15,750 17,000 17,550
Operation ($/yr) 4,000 5,500 5,900
Maintenance ($/yr) 6,500 7,750 8,450
Income ($/yr) 20,000 25,000 25,100
242 Introduction to Optimization for Chemical and Environmental Engineers
SOLUTION:
Calculate the capital recovery factor CRF:
i (16.3)
CRF =
1 − (1 + i )− n
0.05
=
1 − (1 + 0.05)−20
= 0.080
Determine the annual capital and installation costs for the DP unit:
= $3, 360 / yr
Determine the annual capital and installation costs for the ST unit:
= $3, 740 / yr
Determine the annual capital and installation costs for the ES unit:
= $4, 150 / yr
The profit is the difference between the income credit and the total annual cost:
The information in Table 16.6 provides a comparison of the costs and credits for both
exchangers.
A shell-and-tube heat exchanger should therefore be selected based on the previously
mentioned economic analysis.
TABLE 16.6
Comparison of Economic Exchanger Results; Illustrative Example 16.3
Double pipe Shell-and-tube Extended Surface
Total installed ($/yr) 3,360 3,740 4,152
Operation ($/yr) 4,000 5,500 5,900
Maintenance ($/yr) 6,500 7,750 8,450
Total annual cost ($/yr) 13,860 16,990 18,500
Income credit ($/yr) 2,000 25,000 25,100
Profit ($/yr) 6,140 8,010 6,600
greater heat loss due to an increase in insulation. This situation arises from “small” diam-
eter pipes when the increase in area increases more rapidly than the resistance arising
from the thicker insulation; in effect, the heat loss can initially increase with increasing
insulation.
Consider the system shown in Figure 16.2. The area terms for the heat transfer equation
in rectangular coordinates are no longer valid in cylindrical coordinates equal in rectan-
gular coordinates (for the inside surface, the heat transfer area is given by 2πrL).3,4 Applying
the development to a pipe/cylinder system provided by Theodore2,3 leads to
T1 − T0
Q. =
1 1 xw 1 1
+ +
2πri L hi k w 2πLrim , i 2πrL h0
(16.8)
2πL (T1 − T0 ) 2πL (T1 − T0 )
= =
1 ln r0
+
r
+
1 ( ) f (r )
ri hi kw rh0
where:
f (r) =
1 ln r0
+
r
+
( )
1
(16.9)
ri hi kw rh0
h0
ki
T0
kw ∆Xi
∆Xw 1
h1 T1
r ri r0
FIGURE 16.2
Critical insulation thickness for a pipe.
244 Introduction to Optimization for Chemical and Environmental Engineers
Assuming that goes through a maximum or minimum as r is varied, L’Hopital’s rule can
be applied to Equation (16.9):
df ( r )
−
2πL ( T1 − T0 ) dr = 0 (16.10)
f (r)
2
with
df ( r ) 1 1
− =− + (16.11)
dr rki r 2 h0
One may therefore write,
df ( r ) 1 1
− =− + 2 = 0 (16.12)
dr rki r h0
For this maximum/minimum condition, set r = rc and solve for rc
ki
rc = (16.13)
h0
The second derivate of Equation (16.9) provides information as to whether there is a maxi-
mum or minimum at rc.
d2 f ( r ) 1 2 h2 h2
2
= 2 − 3 = 03 = 03 ( 1 − 2 ) (16.14)
dr r ki r h0 k k
Clearly, the second derivative is a negative number; therefore, a maximum is located at r = rc. It
then decreases monotonously as r is increased beyond rc. However, one should exercise care in
interpreting the implications of the previously mentioned development. This result applies only
if r0 is less than rc; that is, it generally applies to “small” diameter pipes/tubes. Thus, rc repre-
sents the outer radius (not the thickness) of the insulation that will maximize the heat loss, and
at which point any further increase in insulation thickness will result in a decrease in heat loss.
A graphical plot of the resistance R2,3 versus r is provided in Figure 16.3. (The curve is
inverted for the plot of Q vs. r.) In other words, the maximum heat loss from a pipe occurs
when the critical radius equals the ratio of the thermal conductivity of the insulation to the
surface coefficient of heat transfer (see also Equation 16.13). This ratio has the dimension
of length (e.g., ft). Note that once again the equation for ri can be rewritten in terms of a
dimensionless number defined as the Biot number, Bi:
ki
= Bi−1 (16.15)
h0 rc
To apply the following equation for a base wall (r = rc), r must be greater than rc, that is,
r > rc. The radius at which this occurs is denoted at r*. The term r* may be obtained by solv-
ing the equation
Resistance, R
ri r0 rc r*
Radius, r
FIGURE 16.3
Resistance associated with the critical insulation thickness for a bare surface.
The previously mentioned development applies where r is less than rc. If r0 is larger
than rc, the previous analysis again applies, but only to the results presented for r > rc; that
is, it will decrease indefinitely as r increases. Note that there is no maximum/minimum
(inflection) for this case since values of ln(r/r0) are indeterminate for r > r0. Once again, it
approaches zero in the limit as r approaches infinity.
As noted earlier, as the thickness of the insulation is increased, the cost associated with
heat loss decreases, but the insulation cost increases. The optimum thickness is determined
by the minimum of the total costs. Thus, as the thickness of the insulation is increased, the
heat loss reaches a maximum value and then decreases with further increases in insula-
tion. Reducing this effect can be accomplished by using an insulation of low conductivity.
SOLUTION:
Assume the resistances are approximately equal, i.e., ki = horc and employ Equation
(16.13). Since BL ≈ 1.0, set
ki
rc =
h0
0.44
= = 0.333 ft
1.32
= 4.0 inches
The critical insulation thickness is therefore
2.0
L1 = ∆xi = 4.0 − = 3.0 inches
2
Since r0 = 1.0 inches, r0 < rc, so that the heat loss will increase as insulation is added, but it
will start to decrease when the radius of the insulation increases above rc, that is, when r > rc.
246 Introduction to Optimization for Chemical and Environmental Engineers
SOLUTION:
The first law of thermodynamics is a conservation law concerned with energy transfor-
mation. Regardless of the types of energy involved in processes – thermal, mechanical,
electrical, elastic, magnetic, and so on – the change in the energy of a system also allows
free convertibility from one form of energy to another, as long as the overall quantity is
conserved. Thus, this law places no restriction on the conversion of work into heat, or on its
counterpart – the conversion of heat into work. However, the second law is another matter.
One of the areas where the meaningful energy conservation measures can be realized
is in the design and specification of process (operating) conditions for heat exchangers.
This can be best accomplished by the inclusion of second law principles in the analysis.
The quantity of heat recovered in an exchanger is not alone in influencing size and cost.
As the temperature difference driving force (LMTD) in the exchanger approaches zero,
the “quality” heat recovered increases.5,6
Any discussion of energy conservation regarding heat exchangers leads to an impor-
tant second law consideration – energy has quality as well as quantity. Because work is
100% convertible to heat, whereas the reverse situation is not true, work is a more valu-
able form of energy than heat. Although it is not obvious, it can also be shown through
second law arguments that heat also has “quality” in terms of the temperature at which
it is discharged from a system. The higher the temperature, the greater the possible
energy transformation into work. Thus, thermal energy stored at high temperatures is
generally more useful to society than that available at lower temperatures. While there
is an immense quantity of energy stored in rivers and oceans for example, its present
availability to society for performing useful tasks is its “quality” since it is degraded
when it is transferred by means of heat transfer from one temperature to a lower one.
Other forms of energy degradation include energy transformation due to frictional
effects and electrical resistance. Such effects are highly undesirable if the use of energy
for practical purposes is to be maximized.5,6
The second law provides the means of measuring this energy degradation through a
thermodynamic term referred to as entropy, and it is the second law (of thermodynamics)
that serves to define this important property. It is normally designated as S with units of
energy per absolute temperature (e.g., Btu/R or cal/K). Furthermore, entropy calculations
can provide quantitative information on the “quality” of energy and energy degradation.5,6
Heat Transfer Applications 247
There are a number of other phenomena that cannot be explained by the first law of
conservation of energy. It is the second law of thermodynamics that provides an under-
standing and analysis of these adverse effects. However, among these considerations,
it is the second law that can allow the measuring of the aforementioned “quality” of
energy, including its effect on the design and performance of heat exchange.
metric) flowrates from A and B, that is, qA and qB, and that will maximize the profit P for
the proposed processes.
SOLUTION:
The line diagram of the system is provided in Figure 16.4.
The describing equations are as follows. The objective function for the profit is
0.75 qB
qST ≤ 12,000
qA ≤ 8000 Shell and tube (ST)
A
0.40 qA
0.60 qA
qDP ≤ 6000
Double pipe (DP)
0.25 qB
qB ≤ 6000
FIGURE 16.4
Shell-and-tube double pipe exchangers; Illustrative Example 16.6.
248 Introduction to Optimization for Chemical and Environmental Engineers
qA ≤ 8, 000 (16.18)
qB ≤ 6, 000 (16.19)
0.75qA + 0.40qB = qST ≤ 12, 000 (16.20)
0.60qA + 0.25qB = qDP ≤ 6, 000 (16.21)
qA ≥ 0; qB ≥ 0 (16.22)
SOLUTION:
One may also calculate the annual (365-day basis) profit based on this condition.
Employing a suitable optimization program gives
qA = 7 , 500 ft 3 /day
qB = 6, 000 ft 3 /day
P = 24, 750 $/day
= $8, 910, 000 $/year
Note that operating, maintenance, depreciation, and so on, costs have not been included
in the analysis.
References
1. R. Perry and D. Green, Perry’s Chemical Engineers’ Handbook, 8th edn, McGraw-Hill, New York
City, NY, 2008.
2. L. Theodore, Chemical Engineering: The Essential Reference, McGraw-Hill, New York City, NY,
2014.
3. L. Theodore, Heat Transfer Applications for the Practicing Engineer, John Wiley & Sons, Hoboken,
NJ, 2011.
4. Webster’s New World Dictionary, Second College Edition, Prentice-Hill, Upper Saddle River, NJ,
1971.
5. L. Theodore, F. Ricci, and T. VanVliet, Thermodynamics for the Practicing Engineer, John Wiley &
Sons, Hoboken, NJ, 2009.
6. L. Theodore and J. Reynolds, Thermodynamics, A Theodore Tutorial, Theodore Tutorials, East
Williston, NY, originally published by the USEPA/APTI, RTP, NC, 1991.
17
Plant Design Applications
The last chapter in Part III contains six illustrative examples concerned with the general
subject of plant design (PD). The title and one-line description of each example are pro-
vided here:
The seven references that follow are for suggested reading for plant design applications:
1. R. Perry and D. Green, Perry’s Chemical Engineers’ Handbook, 8th edn, McGraw‑Hill,
New York City, NY, 2008.
2. L. Theodore, Chemical Engineering: The Essential Reference, McGraw‑Hill, New York
City, NY, 2014.
3. J. Reynolds, J. Jeris, and L. Theodore, Handbook of Chemical and Environmental
Engineering Calculations, John Wiley & Sons, Hoboken, NJ, 2004.
4. L. Theodore, R. Dupont, and K. Ganesan, Unit Operations in Environmental
Engineering, Wiley‑Scrivener, Salem, MA, 2018.
5. J. Happel, Chemical Process Economics, John Wiley & Sons, Hoboken, NJ, 1958.
6. F. Vilbrandt and C. Dryden, Chemical Engineering Plant Design, McGraw-Hill, New
York City, NY, 1959.
7. M. Peters, Plant Design and Economics for Chemical Engineers, McGraw-Hill, New
York City, NY, 1958.
17.1 Shipping Facilities
A large chemical plant complex is likely to have facilities for shipping by pipeline, sea,
inland waterway, rail, and road. Smaller plants may have facilities for only one or two of
these modes. In any case, they must be designed for safe loading and unloading. In addi-
tion, there is often a major economic incentive to load and unload ships, barges, rail cars,
and trucks as quickly as possible.
249
250 Introduction to Optimization for Chemical and Environmental Engineers
The cost of shipping materials, both raw materials and products, is often a very significant
factor in the chemical industry. The energy cost associated with shipping can range from
0.001 gallons of fuel per ton-mile of cargo moved for supertankers to 0.2 for cargo jet aircraft.
SOLUTION:
The number of tons per day in transit is given by the volume per day times the density. The
density of water is 8.33 (specific gravity = 1.00) in lb/gal, and 2000 converts from lbs to tons.
To calculate the shipping cost, the tons are multiplied by the number of miles shipped and
the ton-mile shipping cost. Liquid products are calculated in the same manner. Solid prod-
ucts are calculated similarly, except that the number of pounds per day is already given.
Calculate the shipping cost of the feed by barge and rail:
Barge
Rail
TABLE 17.1
Shipping Distance Data; Illustrative Example 17.1
Distance
Shipped
Product (miles)
Feed 800
Liquid product A 250
Liquid product B 200
Solid product C 40
Solid product D 300
Solid product E 45
Solid product F 120
TABLE 17.2
Shipping Cost Data; Illustrative Example 17.1
Transportation Type Distance (miles) Cost ($/ton-mile)
Barge 20–80 0.03
80–300 0.02
300–1000 0.015
Rail 10–100 0.08
100–1000 0.04
Plant Design Applications 251
Barge, Product B:
Rail, Product A:
Rail, Product B:
500, 000
( 40 )(0.03 ) = 3, 000 $ day
2000
Barge, Product D:
500, 000
( 300 )(0.02) = 1, 500 $ day
2000
Barge, Product E:
500, 000
( 45)(0.03 ) = 340 $ day
2000
Barge, Product F:
500, 000
(120 )(0.02) = 600 $ day
2000
Rail, Product C:
500, 000
( 40 )( 0.03 ) = 800 $ day
2000
Rail, Product D:
500, 000
( 300 )(0.04 ) = 3, 000 $ day
2000
Rail, Product E:
500, 000
( 45)(0.08 ) = 900 $ day
2000
252 Introduction to Optimization for Chemical and Environmental Engineers
Rail, Product F:
500, 000
(120 )(0.04 ) = 1, 200 $ day
2000
The sum of the costs for barge transport are shown in Table 17.3. The sum of the costs for
rail transport are shown in Table 17.4.
Conclusion:
Transport by ship and/or barge is usually much cheaper than transport by rail or truck.
It often requires longer travel times, however, and thus more inventory can be tied up
in shipping.
17.2 Tank Farms
Liquid feeds, products, intermediates, and fuels at chemical and petrochemical plants are
stored in tanks, which are usually located in a “tank farm” adjacent to the process plant
area. The tank capacities are most often expressed in gallons or barrels (one barrel = 42
gallons). The individual tanks may range in size from a thousand gallons or less to several
million gallons.
TABLE 17.3
Costs for Barge Transport
Product Cost ($)
Feed 25,000
Product A 2,080
Product B 1,670
Product C 300
Product D 1,500
Product E 340
Product F 1,200
Total: 31,490
TABLE 17.4
Costs for Rail Transport
Product Cost ($)
Feed 133,300
Product A 4,170
Product B 3,330
Product C 800
Product D 3,000
Product E 900
Product F 600
Total: 146,100
Plant Design Applications 253
Whenever possible, liquids are stored at ambient temperature and pressure. Volatile liq-
uids may have to be stored under pressure. In addition, they often require vapor recovery
systems2 or other devices to prevent releases to the atmosphere as the tanks are filled and
emptied. Liquids that are very viscous at ambient temperature or that would solidify at
ambient temperature are kept in heated tanks.
Standard practice calls for each tank to be surrounded by a dike or berm sufficiently
high to contain all the liquid stored in a tank in case it should rupture. Firefighting equip-
ment are normally located near tanks containing flammable liquids.
SOLUTION:
Calculate the storage requirement for feed and select an appropriate number of tanks:
Barge shipping:
Rail shipping:
TABLE 17.5
Available Standard Tank Sizes; Illustrative Example 17.2
Capacity (gallons) Height (ft) Diameter (ft)
216,000 30 35
429,000 36 45
1,040,000 36 70
2,110,000 36 100
4,060,000 48 120
254 Introduction to Optimization for Chemical and Environmental Engineers
Rail shipping
Rail shipping 45 ft × 36 ft tanks would also be satisfactory. Using the 35 ft × 30 ft tanks
allows a little more flexibility since there would be a total of eight tanks the of the same
size.
Specify a single set:
Barge shipping
Rail shipping
gas stream is forced to turn, but the larger particles have more momentum and cannot turn
with the gas. These larger particles impact and, under the influence of gravity, fall down
the cyclone wall and are collected in a hopper. The gas stream actually turns a number of
times in a helical pattern, much like the funnel of a tornado. The repeated turnings provide
many opportunities for particles to pass through the streamlines, thus hitting the cyclone
wall.
The removal efficiency of a cyclone for a given size particle is very dependent on the
cyclone dimensions. The efficiency at a given volumetric flow rate is most affected by the
diameter. The overall length determines the number of turns of the vortex. The greater
the number of turns, the greater the efficiency. The length and width of the inlet are also
important, since the smaller the inlet, the greater the inlet velocity becomes. A greater
inlet velocity gives greater efficiency but also increases the pressure drop. A number of
equations have been developed for determining the fractional cyclone efficiency, Ei, for
a given size particle. As noted earlier, fractional efficiency is defined as the fraction of
particles of a given size collected in the cyclone, compared to those of that size going into
the cyclone.
Multiple-cyclone collectors (multiclones) are high-efficiency devices that consist of a
number of small-diameter cyclones operating in parallel with a common gas inlet and out-
let. The flow pattern differs from a conventional cyclone in that instead of bringing the gas
in at the side to initiate the swirling action, it is imparted by a stationary vane positioned
in the path of the incoming gas. The diameters of the collecting tubes usually range from
6 to 24 in. with pressure drops in the 2–6 in range.2
Properly designed units can be constructed and operated with collection efficiency as
high as 90% for particulates in the 5–10 µm range. The most serious problems encountered
with these systems involve plugging and flow equalization. Since the gas flow to a multi-
cone is axial (usually from the top), the cross-sectional area available for flow inlet condi-
tions is given by the annual area between the outlet tubes and cyclone body. The outlet
tube diameter is usually one-half the body diameter.2
The pressure drop across a cyclone is an important parameter to the purchaser of such
equipment. Increased pressure drop means greater costs for power to move an exhaust
gas through the recovery/control device. With cyclones, an increase in pressure drop usu-
ally means that there will be an improvement in collection efficiency (one exception to
this is the use of pressure recovery devices attached to the exit tube; these reduce the
pressure drop but do not adversely affect collection efficiency). For these reasons, there
have been many attempts to predict pressure drops from design variables. The idea is
that with having such an equation, one could work back and optimize the design of new
cyclones.2
TABLE 17.6
Particle Size Distribution; Illustrative Example 17.3
Particle Diameter
Range (µm) Weight Fraction (w i)
5–35 0.05
35–50 0.05
50–70 0.10
70–110 0.20
110–150 0.20
150–200 0.20
200–400 0.10
400–700 0.10
100
6" diameter
24" diameter
80
Collection efficiency, %
12" diameter
60
40
20
0
0 20 40 60 80 100
Particle diameter, microns
FIGURE 17.1
Fractional efficiency data; Illustrative Example 17.3.
SOLUTION:
Calculate the required collection efficiency, ER:
3.1 − 0.06
ER =
3.1
(100)
= 98%
= 0.98
Calculate the average particle size associated with each size range (see Table 17.7).
Table 17.8 provides the overall efficiency, E6, for the 6-in tubes. Since E6 > ER, the 6-in
tubes will do the job. Table 17.9 is generated for the 12-in tubes. Since the overall effi-
ciency, E12 < ER, the 12-in tubes will not do the job. Thus, it will be necessary to use the
6-in tubes for a conservative design.2
Since the outlet tube diameter is one-half the body diameter, the inlet cross-sectional
area (for axial flow) for each 6-in (0.5 ft) tube will be
(
A = 0.785 0.52 − 0.252 )
= 0.147 ft 2
Plant Design Applications 257
TABLE 17.7
Average Particle Size Calculations
Particle Diameter Average Particle
Range (µm) Diameter (µm)
5–35 20
35–50 42.5
50–70 60
70–110 90
110–150 130
150–200 175
200–400 300
400–700 550
TABLE 17.8
Overall Efficiency Calculations for 6-in Tubes
Average Particle Weight Efficiency for Ei (w i) for 6-in
Diameter (µm) Fraction (w i) 6-in Tubes (%) Tubes (%)
20 0.05 89 4.45
42.5 0.05 97 4.85
60 0.10 98.5 9.85
90 0.20 99 19.8
130 0.20 100 20
175 0.20 100 20
300 0.10 100 10
550 0.10 100 10
E6 = 98.95
TABLE 17.9
Overall Efficiency Calculations for 12-in Tubes
Average Particle Weight Efficiency for Ei (w i) for 12-in
Diameter (µm) Fraction (w i) 12-in Tubes (%) Tubes (%)
20 0.05 82 4.1
42.5 0.05 93.5 4.67
60 0.10 96 9.6
90 0.20 98 19.6
130 0.20 100 20
175 0.20 100 20
300 0.10 100 10
550 0.10 100 10
E12 = 98.95
Since the velocity in each tube is 60 ft/s, the number of tubes, n, is given by
n = 190 tubes
258 Introduction to Optimization for Chemical and Environmental Engineers
Thus, 190 tubes are required in this multiple-cyclone unit. A 15 × 15, 14 × 14, or 12 × 16
design is recommended.
Total annual cost (TAC ) = (operating and shutdowncosts cycle) (cycles year )
(17.1)
+ annualized fixed costs
SOLUTION:
(
Total annual costs (TAC ) = 28 (PPB )
0.33
) 1, 200, 000
+ 21
PPB
(17.9)
+300 (PPB )
0.75
+ 300, 000; $
The total annual cost is a minimum where d(TAC)/d(PPB) = 0. By performing the dif-
ferentiation, one obtains
This result could also have been obtained by plotting total annual cost versus PPB and
determining the value of PPB at the point of minimum annual cost. For conditions of
minimum annual cost and 1 million lb/year production,
In addition,
1, 200, 000
Total time used per year = (11.6 ) = 5872 hr
2432
Thus, for conditions of minimum annual cost and a production of 1.2 million lb per
year, the available operating and shutdown time would be sufficient.
1. If r = −kc
dc c0 − c c c
= − kc = 0 − − kc (17.11)
dt τ τ τ
260 Introduction to Optimization for Chemical and Environmental Engineers
c0 kτ + 1
= − c
τ τ
dc
= dt
c0 − c ( k τ + 1)
c t
dc dt
∫ci
c0 − c ( k τ + 1)
=
∫τ
0
1 c0 − (1 + k τ ) c t
− ln =
1 + k τ c0 (1 + k τ ) ci τ
c0 − (1 + k τ ) c t
− (1+ k τ )
τ
c 1 + kτ c = e
0( )i
Rearranging yields
c − (1+ k τ )
t t
− (1+ k τ )
τ
c = ci e + 0 1 − e τ (17.12)
1 + kτ
t
− −
t
c = ci e τ
+ (c0 − k τ ) 1 − e τ
t t
− −
τ τ
= ci e + c0 − c0 e
Rearranging yields
t
−
= c0 + (ci − c0 ) e τ
(17.13)
3. If k ≈ 0 in (1),
c − (1+ k τ )
t t
− (1+ k τ )
τ
c = ci e + 0 1 − e τ
1 + kτ
t t
− −
τ τ
= ci e + c0 − c0 e (17.14)
t
−
= c0 + (ci − c0 ) e τ
4. If v0 varies and τ varies, solving the equation becomes more complex. Variations
need to be included in the describing equation
dc c0 − c
= + r (17.15)
dt τ
This may require a numerical solution.
5. If both c0 = c0(t) and v0 = vd(t), the solution in part (4) again applies.
SOLUTION:
As shown in Figure 17.2, x1 = depth (ft), x2 = width (ft), and x3 = height (ft). The material
cost is
x3 = height
x1 = depth
x2 = width
FIGURE 17.2
Rectangular structure; Illustrative Example 17.6.
Width
x2 ≤ 3 x1 (17.22)
Height
2
x3 ≤ x2 (17.23)
3
Variable limits
0 ≤ x1 ≤ 60 ft (17.24)
0 ≤ x2 ≤ 80 ft (17.25)
The optimum design solution is left as an exercise for the reader. (Hint: the values gen-
erated from Excel were [in ft] x1 = 13.9, x2 = 41.6, and x3 = 27.9).
References
1. D. Kauffman, Process Synthesis and Design, A Theodore Tutorial, Theodore Tutorials, East
Williston, NY, originally published by USEPA/APTI, RTP, NC 1992.
2. L. Theodore, Air Pollution Control Equipment Calculation, John Wiley & Sons, Hoboken, NJ, 2008.
3. M. Peters, Plant Design and Economics for Chemical Engineers, McGraw-Hill, New York City, NY,
1958.
4. L. Theodore, Chemical Engineering: The Essential Reference, McGraw-Hill, New York City, NY,
2014.
5. K. Behan, personal notes, University of Buffalo, Buffalo, NY, 2017.
Part IV
As one would conclude from the title of this Part, it is concerned with applications. Two
chapters complementing the presentation for this Part contain five sections that are loosely
based on the contents of Parts II and III; chapter and section titles are listed below. The
third (and last) chapter is concerned with advanced optimization term projects; four sec-
tions compliment this presentation.
Chapter 18: Select Environmental Engineering Applications
It should be noted that last two sections of Chapter 20 were drawn (with permission)
from the classic work of Happel’ s 1959 John Wiley & Sons text, Chemical Process Economics .
Happel’ s original units and notation were retained in these presentations.
18
Select Environmental Engineering Applications
18.1 Air Management
Gaseous (and particulate) pollutants discharged into the atmosphere can be controlled.
The four generic devices for gases include absorbers, adsorbers, scrubbers, and incin-
erators. Gas absorption, as applied to the control of air pollution, is concerned with the
removal of one or more pollutants from a contaminated gas stream by treatment with a
liquid. The necessary condition is the solubility of these pollutants in the absorbing liquid.
Design and performance characteristics of any particular control system are important
considerations. Optimizing the performance of an air pollution control device is often a
standard operating practice.
SOLUTION:
Theodore1 proposed a host of possible solutions. Many of them are listed here:
1. Use a smaller packing size. A different size may produce a better packing
factor.
2. Use a different type of packing. A different packing type may also produce a
better packing factor.
3. Make sure there is no channeling inside the column. The packing should be
randomly distributed and there should be no open spots where water will
accumulate.
265
266 Introduction to Optimization for Chemical and Environmental Engineers
50 ppm (design)
90°F water
1 atm
rasching rings)
5 ft (one inch
1000 acfm gas
1000 ppm (solute)
90°F
FIGURE 18.1
Faulty design?
4. Increase the water flow entering the column. In effect, the slope of the operat-
ing line will increase1 and provide a greater driving force for mass transfer.
Thus, there will be a greater mass transfer between the liquid and gas phase
and the efficiency of the column will increase.
5. Use a liquid other than water to scrub the gas. Another liquid may cause better
mass transfer between the gas and liquid phase.
6. Check to see if flooding is occurring. If the column is near the flooding point,2
there will not be much mass transfer. The water will be pushed up the column
by the gas rather than the water coming down the column.
7. Lower the initial concentration in the carrier gas (if possible).
8. Increase the height of the bed. This will provide more surface area for mixing
and more time in the absorber.
9. Increase the pressure. This will reduce the slope of the operating line1 and lead
to a greater driving force.
10. Decrease the temperature. This will reduce the slope of the operating line and
lead to a greater driving force.
11. Increase the pressure and decrease the temperature.
12. Additional packing height could be added at the top of the column. This would
increase the amount of the solute removed from the gaseous phase.
13. Add sprays at the bottom of the column.
14. Modify the process producing the problem. Since the details of this process
are not included in the problem statement, no specific recommendations can
be mentioned in this discussion.
15. Finally, before considering changes to the system, one should undertake a thor-
ough inspection of the unit and surrounding components. The emission moni-
toring system should be recalibrated. All valves, fittings, and pipes should be
checked for plugging or leaks. The liquid distributor should be checked to
make sure it is functioning properly. The distribution of the packing should be
inspected to make sure it is as uniform as possible. Any problems encountered
during this inspection should be corrected immediately. Following the main-
tenance check, the performance of the unit should be reevaluated.
Select Environmental Engineering Applications 267
The reader should note that some of the previously mentioned recommendations could
lead to higher pressure drops and potential problems with the flow. An example of this
problem would be the implementation of suggestion (1). The reader is left the exercise of
determining what other steps could lead to flow/pressure drop problems. Hint: There
are at least six suggestions that fall into the proposed solution, including firing the envi-
ronmental engineer who designed the absorber.
A similar situation could exist with a gaseous adsorption unit. Once again, there are
options available to bring the unit into compliance with the specified design concen-
tration. Some of Theodore’s1 recommendations are listed below. Assume the adsorber
packing is activated carbon and there is no liquid flow.
1. Check/increase the depth of the adsorption bed. Increasing the depth of the
bed will increase the removal capacity of the column.
2. Change the adsorbent type of size. Use an adsorbent with a higher saturation
capacity.
3. Reducing the flow will increase the residence time and should allow for more
adsorption.
4. Decrease the inlet temperature. This will favor the adsorption equilibrium
process.
5. Increase the system pressure. This will also favor the adsorption equilibrium
process.
6. Regenerate the adsorbent for a longer period.
7. Replace the adsorbent.
8. Make up any adsorbent lost to carryover (as necessary).
9. Consider changing the flow direction.
1. The first thing that needs to be investigated is what is the current state of the
column? How was it originally designed? Were safety margins included in the
design that allow for some “tweaking” of the actual structure of the column?
How much change in pressure (drop) is allowed? What will the column mate-
rials stand up to?
2. A possibility for the column is changing the flow rate of gas through the col-
umn. If the gas is made to pass through the column at a higher velocity, it
should result in more foaming of the liquid on the trays. This will increase the
mass transfer area. This will also help to separate the products. Unfortunately,
this could create new problems with flooding and entrainment. This needs to
be examined very carefully before it is implemented.
3. It also might be possible to change the reboiler and condenser2 conditions. It is
possible after 24 years of service that these two devices are not working at their
optimum settings. An examination should be made of their contribution to the
failure to perform the desired separation. Is the column failing because these
devices are no longer functioning? If so, then they need to be either retrofitted
or replaced with newer models.
4. The composition and condition of the feed stream also need to be examined.
Is it the same as what the column was originally designed to separate? Has the
process been altered upstream in such a way that the separation is no longer
268 Introduction to Optimization for Chemical and Environmental Engineers
feasible? Perhaps it has changed through the years in such a manner that the
column is no longer designed to deal with the stream; if so, the feed stream
could be altered to adhere to the original conditions to make the column oper-
ate correctly again. Also, the feed tray (location) could be changed. Finally, the
inlet feed quality might also be changed.2
5. Adding trays is the most difficult thing to implement but would probably have
the largest impact. The trays could be added near the top of the column if a
safety margin of headspace had been designed into the column. A possibility
would be the complete removal of all the trays and reinstalling them with a
smaller tray spacing between them. If that is done, the environmental engineer
should consider replacing the trays with a better tray design. Significant work
has gone into tray design in the past half-century; improvements have been
made to optimize their efficiency and capabilities.
6. Finally, it is possible that the column needs cleaning and that too could opti-
mize/improve the column’s performance.
18.2 Water Management
Shipping and storage facilities often require a majority of the total land area at chemical
and petrochemical plants and also represent a very large total capital investment. Though
their design is not usually so sophisticated as that of the actual process units, they often
represent the largest potential source of safety and environmental problems because of the
large inventories involved.
Gaseous feeds, products, and fuels may be shipped to and from a plant by pipeline, pres-
surized truck, or tank car, or in individual cylinders. Pipelines are by far the least expen-
sive means of shipment of gases, though they have very high initial capital costs. Gases
are the least practical materials to store at a plant site. They may be stored in large storage
tanks or in trucks, tank cars, and individual cylinders. Underground, hollowed-out salt
domes are sometimes used where very large amounts of storage are needed, such as for
natural gas, ethane, and ethylene.
Liquid feeds, products, and fuels may be shipped by pipeline, tanker ship barge, tank
car, truck, or, occasionally, individual small containers. Pipelines and very large ocean-
going tankers are the least expensive shipping means. Barges are the next most expensive,
followed by rail cars and trucks. Liquid feeds, products, intermediates, and by-products
are commonly stored in large tanks in a “tank farm” adjacent to the process units. Tanks
must be surrounded by dikes or other barriers to prevent loss of the liquid if a tank should
rupture. Vent condensers, floating roofs, and other features are used with volatile liquids
to prevent release of their vapors to the atmosphere.
Solid feeds, products, and fuels are commonly shipped by ocean freighter, barge, rail car,
and truck. Storage may be in silos or covered hoppers, in sacks, or special containers, or
simply in piles out in the open.
As noted in the previous chapter, feeds, products, intermediates, and fuels at chemical and
petrochemical plants are stored in tanks, which are usually located in a “tank farm” adjacent to
the process plant area. The tank capacities are most often expressed in gallons or barrels (one
barrel equaling 42 gallons). The individual tanks may range in size from a thousand gallons or
less to several million gallons. Whenever possible, liquids are stored at ambient temperature
and pressure. Volatile, hazardous, and toxic liquids may have to be stored under pressure. In
addition, they often require vapor recovery systems or other devices to prevent releases to the
Select Environmental Engineering Applications 269
atmosphere as the tanks are filled and emptied (defined as working losses). Liquids that are
very viscous at ambient temperature or that would solidify at ambient temperature are kept
in heated tanks. Standard practice calls for each tank to be surrounded by a dike or berm suf-
ficiently high to contain all the liquid stored in a tank in case it should rupture. Firefighting
equipment is usually permanently located near tanks containing flammable liquids.3
Storage tanks can present major risks both in terms of fire and environmental pollution. In
both cases, this is primarily due to the large inventories of materials that could be involved
in any accident. There has been a major effort in industry in the last century to reduce the
amount of storage at process plants, especially for flammable and/or toxic materials.3
From a water pollution perspective, the major problem with these tank farms falls under
the heading of operation, storage, and transportation. The objective of any water quality
management program is to control the discharge of any liquid pollutants so that water
quality is not degraded to an unacceptable extent below the natural background level.
The impact of the pollutant on water quality must be predicted and compared to the back-
ground water quality that would be present without human intervention.
TABLE 18.1
Feed Data
Feed 1 110,000 lb/day ρ = 49 lb/ft3
Feed 2 50,000 lb/day ρ = 68 lb/ft3
TABLE 18.2
Product Data
Product A 40,000 lb/day ρ = 52 lb/ft3
Product B 25,000 lb/day ρ = 62 lb/ft3
Product C 10,000 lb/day ρ = 52 lb/ft3
Product D 95,000 lb/day ρ = 47 lb/ft3
TABLE 18.3
Tank Data
Standard Tank Sizes (gallons)
2,800 11,200 56,100 561,000
5,600 16,800 140,000 1,123,000
8,400 28,100 281,000
270 Introduction to Optimization for Chemical and Environmental Engineers
TABLE 18.4
Storage Requirements
Material Lb/day Gal/day Days Gals Storage Required
Feed 1 110,000 16,800 28–35 470,000–587,800
Feed 2 50,000 5,500 28–35 154,000–192,500
Product A 40,000 5,750 28–42 161,100–241,700
Product B 25,000 3,020 28–42 84,500–126,700
Product C 10,000 1,440 28–42 40,300–60,400
Product D 95,000 15,120 7–14 105,800–211,700
SOLUTION:
Unfortunately, there is no single, simple method for determining the optimum mix of
storage tanks for a chemical plant. Most often, estimates are made of the minimum and
maximum amounts of feeds, intermediates, and products that must be kept on hand.
Some additional allowance is then made to permit periodic cleaning and maintenance
of the tanks. One solution and some calculation details follow.3
First, determine the maximum and minimum amounts of each material to be stored
(see Table 18.4). The conversion to gallons requires dividing the rate in lb/day by the
density in lb/ft3, and then multiplying by 7.481 gal/ft3.
For each material, select a set of tanks:
This set is 5% short on the maximum of Feed 1, 11% short on Product B, and 7% short on
Product C. For most situations, this would be acceptable. Select spare and/or “swing”
tanks to provide for maintenance:
The number of tanks required will be quite large in this application. If market forces,
such as environmental regulations and fluctuating demand require this much storage,
they may all be necessary. More modern commercial operations, such as “just-in-time”
or “on demand” manufacturing, call for reducing in-plant inventory to the absolute
minimum possible, particularly with toxic, flammable, and explosive3 liquid waste. To
Select Environmental Engineering Applications 271
further complicate matters, the minimum number of tanks may not always be optimum
if the tanks are extremely large. The reader should note that several smaller tanks may
cost somewhat more initially, but they offer more flexibility in use.
Assume 8550 hr of operation/yr, five coal-fired boilers, and (1 boiler hp = 33,500 Btu/hr
for R = 100).
CA =
( 5)(14, 000 + 0.04R1 ) = 8.18 + 0.0000234R2 (18.3)
8, 550
The total output, T, in actual boiler hp developed is given by
R
T = 5 ( 300 ) = 15R (18.4)
100
Optimum results will be obtained when the sum of the previously mentioned variable
costs, divided by the output, is a minimum. Refer to Table 18.6. As shown in Table 18.6
and in Figure 18.2, the optimum costs per unit output are found at approximately
200% of nominal rating, corresponding to a cost of $14.30 per 1000 boiler hp devel-
oped (33.5 million Btu), or $0.428 per million Btu, exclusive of cost of labor. This
cost is higher than that in a larger power plant in which the fixed charges per boiler
horsepower developed would be less than with these small boilers, and the efficiency
would be higher.
TABLE 18.5
Boiler-Efficiency Data
R 1, % of Nominal Rating E, % Overall Thermal Efficiency
100 75
150 76
175 75
200 74
225 72
250 69
275 65
300 61
Select Environmental Engineering Applications 273
TABLE 18.6
Cost Calculations; Illustrative Example 18.3
CF + CA
(1,000 )
R E R/E CF CA C F+C A T T
100 75 1.33 16.7 8.4 25.1 1500 16.7
150 76 1.97 24.8 8.7 33.5 2250 14.9
175 75 2.33 29.3 8.9 38.2 2625 14.6
200 74 2.70 33.9 9.0 42.9 3000 14.3
225 72 3.12 39.2 9.4 48.6 3375 14.4
250 69 3.62 46.5 9.6 56.1 3750 15.0
275 65 4.23 53.2 10.0 63.2 4125 15.3
300 61 4.92 61.8 10.3 72.1 4500 16.0
20
Dollars per 1000 BHP developed
16
Total, T
12 Coal, CF
4
Fixed charges, CA
0
100 140 180 220 250 300
R, % of nominal rating
FIGURE 18.2
Optimum cost for boiler operation; Illustrative Example 18.3.
The term risk assessment is not only used to describe the likelihood of an adverse
response to a chemical or physical agent but has also been used to describe the likeli-
hood of any unwanted event. These include risks, such as explosions or injuries in the
workplace; natural catastrophes; injury or death due to various voluntary activities such
as skiing, skydiving, flying, and bungee jumping; diseases, death due to natural causes;
and, many others. This risk scenario also receives treatment in the industrial application
that now follows.
Two weather conditions are envisioned, namely, a northeast wind and a southwest
wind (6.0 mph) with Stability Class B.1 Associated with these two wind directions are
Events IIA and IIB, respectively, defined as follows:
P (I ) = 10 −6
P (IIA|II ) = 0.33
P (IIB|II ) = 0.67
Note that P(IIA|II) represents the probability that Event IIA occurs given that Event
II has occurred.9 The consequences of Events I, IIA, and IIB, in terms of the number of
people killed, are estimated as follows:
I – All persons within 200 m of the explosion center are killed; all persons beyond
this distance are unaffected.
IIA – All persons in a pie-shaped segment of 22.5° width (downwind of the source)
are killed if the concentration of the toxic gas is above 0.33 μg/L; all persons
outside this area are unaffected.
IIB – Same as IIA.
Thirteen people are located within 200 m of the explosion center but not in the pie-
shaped segment described previously. Eight people are located within the pie-shaped
Select Environmental Engineering Applications 275
segment southwest of the discharge center; five are 350 m downwind, and three are
600 m away at the plant fence (boundary). Another six people are located 500 m away
outside the pie-shaped segment but within the plant boundary. All individuals are at
ground level.
Theodore Associates have been specifically requested to calculate the average
annual individual risk (AAIR) based on the number of individuals potentially
affected as well as the average risk based on all the individuals within the plant
boundary. Their solution is presented subsequently. Hint: Perform atmospheric dis-
persion calculations at various distances from the emission source and combine
these predicted concentrations with consequence information from the problem
statement.
SOLUTION:
Draw a line diagram of the plant layout and insert all pertinent data and information
(see Figure 18.3). An event tree9 for the process is presented in Figure 18.4. First, calcu-
late the probability of Event IIA occurring; also calculate the probability of Event IIB
occurring9 as follows:
1 1
( )
P (IIA ) = P (II ) P IIA II =
33, 333
(0.33) = 100, 000 = 10 −5 (18.5)
1 2
( )
P (IIB ) = P (II ) P IIB II =
33, 333
(0.67 ) = 100, 000 = 2 × 10 −5 (18.6)
Perform a dispersion calculation to determine the zones where the concentration of
the chemical exceeds 0.33 μg/L. Assume a continuous emission for a point source.1 To
maintain consistent units, convert wind speed from mph to m/s and concentration from
μg/L to g/m3 as follows:
miles ft 1h m m
u = 6.0
5280 0.3048 = 2.68
h mile 3600s ft s
13 N
≤200 m
100 m
Number of
5
X
350 m
Individuals at location
3 6
600 m 500 m
FIGURE 18.3
Plant layout and pertinent data; Illustrative Example 18.4.
276 Introduction to Optimization for Chemical and Environmental Engineers
Unstable
chemical
Flammable
Detonation toxic gas
release
NE wind SW wind
I IIA IIB
FIGURE 18.4
Event tree for process emission and accident; Illustrative Example 18.4.
0.33 µg g L g
c= 1 6 10 3 3 = 3.3 × 10 −4 3
L 10 µg m m
These values are then used as input to the Pasquill-Gifford model for centerline, ground-
level concentrations of a continuous source of pollutant at an elevated emission height
of H*, as shown in the following equation:8–12
1 H * 2
q
c ( x , 0, 0; 125) = exp −
πσ y σ z u 2 σ z
(18.7)
240 g/s 1 125 m 2
= exp −
( )
π σ y (σ z ) ( 2.68 m/s ) 2 σ z
The downwind concentrations can be calculated based on the previous equations. The
concentration results for select downwind distances are provided in Table 18.7. A linear
interpolation of these calculations indicates that the maximum ground-level concentra-
tion (GLC) is approximately 1.01 × 10−3 g/m3 and is located at a downwind distance of
about 800m. In addition, the “critical” zone, that is, where the concentration is above
3.3 × 10−3 g/m3, is located between 475 and 1800 m.
It should be noted that only one “average” weather condition was considered in this
example. However, one often selects the worst-case weather condition that corresponds
with a reasonable probability of occurrence in the location of the site being evaluated.
Employing this worst-case condition produces risk results on the conservative side. An
analysis that includes a full spectrum of wind speeds, directions, and stability classes
would obviously provide a more complete set of risk assessment calculations than is
provided here.
Select Environmental Engineering Applications 277
TABLE 18.7
Downtown Concentration Profile for this Site
x (m) σy (m) σz (m) c (g/m3)
300 47 30 3.43 × 10−6
400 60 41 1.11 × 10−4
500 75 2 4.07 × 10−4
550 80 60 6.78 × 10−4
600 90 65 7.67 × 10−4
700 105 77 9.44 × 10−4
800 120 90 1.01 × 10−3
900 150 110 9.06 × 10−4
1000 170 140 8.04 × 10−4
1500 250 240 4.15 × 10−4
1700 275 275 3.40 × 10−4
2000 300 380 2.37 × 10−4
Now, determine which individuals within the pie-shaped segment downwind from
the source will be killed if either accident (I or II) occurs. Referring to Figure 18.3, 13
individuals within the 200 m radius will die from Accident I. Three individuals located
in the pie-shaped segment and 600 m southwest of the emission source will die from
Accident II. The five individuals located 350 m southwest of the emission source are in
the path of the dispersing plume but are all outside the critical zone. The six individuals
located outside of the pie-shaped impact segment are within the plant boundary but
are not potentially affected by either the explosion or the dispersing plume. The plant
health risk and hazard risk assessment line diagrams7,10 are presented in Figures 18.5
and 18.6, respectively.
The total annual deaths (TAD) for the process if the accident occurs are therefore
TAD = 13 + 3 = 16 deaths / yr
The total annual risk (TAR) is obtained by multiplying the number of people in each
impact zone by the probability of the event affecting that zone and summing the result.
Thus,
( ) ( )
TAR = (13 ) P (1) + ( 3 ) P (IIA ) = (13 ) 10 −6 + ( 3 ) 10 −5 = 4.3 × 10 −5
The average annual individual risk (AAIR) is obtained by dividing this result by
the number of people in the impacted zone. The average annual risk (AAR) is calcu-
lated based only on the “potentially affected” people. Since 21 people are “potentially
affected,” that is, are within the impact area of the explosion or are in the path of the
dispersing plume, the AAR is determined to be:
4.3 × 10 −5
AAR = = 2.05 × 10 −6
21
The AAIR is based on all the individuals within the plant boundary. For this case, this
AAIR is now based on 27 rather than 21 individuals. Thus,
4.3 × 10 −5
AAIR = = 1.6 × 10 −6
27
278 Introduction to Optimization for Chemical and Environmental Engineers
Health problem
identification
Risk
characterization
FIGURE 18.5
Plant health risk assessment.
Chemical explosion
Hazard
identification
Explosion of chemical
Risk
determination
FIGURE 18.6
Plant hazard risk assessment.
Select Environmental Engineering Applications 279
FIGURE 18.7
Bathtub curve.
280 Introduction to Optimization for Chemical and Environmental Engineers
is constant and equal to α. For β > 1, the failure rate increases with time. Using Equation
18.8 to translate the assumption about failure rate into a corresponding assumption about
the probability distribution function (pdf)12 of t, time of failure, one obtains Equation 18.9,
which primarily defines the pdf of the Weibull distribution. (The exponential distribution
is a special case of the Weibull distribution with β = 1.)
t
0
∫
f ( t ) = αβtβ−1 exp αβtβ−1dt
(18.9)
( )
= αβtβ−1 exp −αtβ ; t > 0; α > 0,β
β>0
The variety of assumptions about failure rate and the probability distribution of time to
failure that can be accommodated by the Weibull distribution make it especially attrac-
tive in describing failure-time distributions in industrial and process plant applications.
Estimating the coefficients in the Weibull distribution can be accomplished using a graph-
ing procedure developed by Bury.13
The Weibull distribution has played an important role in helping to reduce and/or elimi-
nate accidents and other hazards from occurring in both the workplace and in everyday
life. Shaefer and Theodore11, and Theodore and Dupont7 have provided both development
material and numerous illustrative examples that are concerned with this distribution. It
has also served as the basis for a host of technical papers prepared by one of the authors.
In recent years, the Weibull distribution has come under fire, due primarily to the
efforts of one of the authors. The last 6 years has provided an opportunity for Dupont and
Theodore14, along with Ricci and others14–17, to carefully analyze the merits and limitations
of the Weibull distribution. This has led to the development of the DRaT models. Details
are given subsequently.
As noted in the previous section, the general two-coefficient Weibull model is repre-
sented by an equation that can be applied to three failure rate periods representing three
failure mode stages. As such, the model consists of six coefficients – two for each of the
three failure mode stages. Theodore and Dupont14 viewed this six-coefficient relationship
as both cumbersome and unnecessary. After some deliberation, it was decided to employ
a new, simpler approach to represent failure behavior – specifically the failure-time (as
opposed to failure rate-time) relationship depicted in Figure 18.8. After even more delib-
eration and analysis, these authors settled on four models that are based on the melding of
either a power function (P) or an exponential relationship (E) for the BI period with either
a power function or exponential relationship for the WO period – the sum of which results
in a curve as shown in Figure 18.8. There are therefore four combinations of the power
function and exponential term equations: BI(P) – WO(P), BI(E) – WO(E), BI(E) – WO(P), and
BI(P) – WO(E), and the corresponding equations resulting from these combinations are
defined as the DRaT II, DRaT III, DRaT IV, and DRaT V models, respectively.
Note that the general relationship for the power function for either the BI or WO period
takes the form
The general form of the exponential function takes the form of either
( )
E ( t ) = A1 1 − e − A2t (18.11)
Select Environmental Engineering Applications 281
Number of failures, N
Time, t
FIGURE 18.8
Failure-time relationship for the Weibull distribution.
( )
E ( t ) = B1 e B2t − 1 (18.12)
f ( t ) = g ( t ) + h ( t ) (18.13)
For example, if both the BI and WO periods are represented by an exponential function
(see also Figure 18.9) it would appear as
( )
g ( t ) = A1 1 − e − A2t (18.14)
and
( )
h ( t ) = B1 e B2t − 1 (18.15)
The sum of Equations 18.14 and 18.15, presented in Figure 18.9, represents the failure model
previously defined as the DRaT III model, that is,
( ) ( )
N ( III ) = g ( t ) + h ( t ) = A1 1 − e − A2t + B1 e B2t − 1 ; N = number of failures (18.16)
Specific details on each of the remaining DRaT models are provided in the next three
subsections.
18.5.1 DRaT II Model
This model assumes that BI(P) and WO(P) relationships apply. Thus,
N ( III ) = N ( BI ) + N ( WO ) (18.17)
282 Introduction to Optimization for Chemical and Environmental Engineers
(a)
h(t) = B1(eB2t – 1)
(b)
g(t)
h(t)
(c)
FIGURE 18.9
DRaT model II representation. (a) BI period, (b) WO period, and (c) combined BI and WO period.
and
18.5.2 DRaT IV Model
This model assumes that BI(E) and WO(P) relationships apply; thus,
( )
N ( BI ) = A1 1 − e − A2t (18.19)
and
so that
( )
N ( IV ) = A1 1 − e − A2t + B2t + B3t 2 (18.21)
18.5.3 DRaT V Model
The last model assumes BI(P) and WO(E) relationships apply. Thus,
and
( )
N ( WO ) = B e B2t − 1 (18.23)
so that
( )
N ( V ) = A2t + A3t 2 + B e B2t − 1 (18.24)
Similar to the last two models, this model also contains four coefficients.
Three improvements in the proposed DRaT models relative to the Weibull model become
immediately apparent:
1. The DRaT model requires either three of four (not six) coefficients to be estimated.
2. The DRaT model is continuous over the entire time range, as opposed to the
Weibull model that is evaluated separately over three compartmentalized failure
stages.
3. The requirement of a constant failure rate period in the Weibull model has been
replaced by a more realistic failure rate that can continue to increase slightly with
time during the supposed constant failure rate period.
The reader should note once again that the failure rate for most applications is a calcu-
lated quantity obtained from the number of failures (N) versus time (t) data. The failure
rate at a specified time is thus approximately equal to the slope (or derivative) of N versus
t at the time point in question. Theodore and Ricci18 provide six different numerical dif-
ferentiation procedures to calculate this derivative, that is the failure rate, and the reader is
directed to that reference for more details.
References
1. L. Theodore, Air Pollution Control Equipment Calculations, John Wiley & Sons, Hoboken, NJ,
2009.
2. L. Theodore and F. Ricci, Mass Transfer Operations for the Practicing Engineer, John Wiley & Sons,
Hoboken, NJ, 2010.
3. D. Kaufmann, Plant Design, A Theodore Tutorial, Theodore Tutorials, East Williston, NY, origi-
nally published by the USEPA/APTI, RTP, NC, 1997.
284 Introduction to Optimization for Chemical and Environmental Engineers
19.1 Fluid Flow
Fluids are usually transported in pipes or tubes. Generally speaking, pipes are heavy-
walled and have a relatively large diameter. Tubes are thin-walled and often come in coils.
Pipes are specified in terms of their diameter and wall thickness. The nominal diameters
range from 1/8 to 30 inches for steel pipe. Standard dimensions of steel pipe are provided
in the literature and are known as IPS (iron pipe size) or NPS (nominal pipe size). The wall
thickness of the pipe is indicated by the schedule number, which can be approximated
from 1000(P/S) where P is the maximum internal service pressure (psi) and S is the allow-
able bursting stress in the pipe material (psi). (The S value varies by material, grade of
material and temperature; allowable S values may be found in piping handbooks).1–3
Tube sizes are indicated by the outside diameter. The wall thickness is usually given
a BWG (Birmingham Wire Gauge) number. The smaller the BWG, the heavier the tube.
The literature lists the sizes and wall thicknesses of condenser and heat exchanger tubes.
For example, a ¾ inch 16 BWG tube has an outside diameter (OD) of 0.75 inches, an inside
diameter (ID) of 0.62 inches, a wall thickness of 0.065 inches, and a weight of 0.476 lb/ft.
When designing a piping system for an industrial process, savings can be effected by
determining the optimum pipe diameter and hence the economic optimum condition by
means of an economic balance. As mentioned previously, to obtain an economic balance,
one type of cost must increase while the other decreases with an increase in the indepen-
dent variable. For the case under consideration, the primary independent variable is usu-
ally the diameter of the pipe. One set of costs, namely, the fixed charges (cost of the pipe,
fittings, and installation) will increase with an increase in the pipe diameter. On the other
hand, for a given flow rate, the power costs due to pressure drop decrease as the pipe size
increases. In other words, if the pipe is oversized, the fixed charges will be too large, and if
the pipe diameter is too small, the pressure drop will be excessive, resulting in large power
costs. The economic balance therefore is made between the fixed costs and the power costs.
285
286 Introduction to Optimization for Chemical and Environmental Engineers
TABLE 19.1
Pipe Cost Data; Illustrative Example 19.1
Pipe Size, I.D. (in) Installed Cost ($/ft)
1.0 1.80
1.5 1.89
2.0 2.19
2.5 3.18
3.0 3.99
SOLUTION:
First, perform the calculations for the 1.0-in pipe assuming turbulent flow and a friction
factor of approximately 0.005.1,2
Annual power costs will depend on the pressure drop. First, calculate the velocity:
v=
( 30 )( 4 )
( 60 )( 7.48 )(1 / 12 )
2
= 38.5 ft / s
Apply Fanning’s equation. 1–3
4fLv 2 (19.1)
∆P =
2g c D
∆P =
( 2 )( 32.2 )(1 / 12 )
∆P = 5, 523 ft H 2 O
PC =
( 28, 732 )( 30 )( 60 )( 8700 )( 0.06 )
( 7.48 )( 778 )( 3412 )
Annual power costs = $1359
Table 19.2 provides the calculated results for the remaining pipe sizes.
Plotting both the annual fixed costs and power costs versus the pipe size will result
in a graph similar to that in Figure 19.1, as presented earlier in Part III, Figure 13.1. The
optimum pipe diameter is approximately 1.75 inches. Performing the calculations for
pipe sizes of 1.25 inches and 1.75 inches is left as an exercise for the reader.
The reader should note that this topic will be revisited in Part IV, Chapter 19.
19.2 Chemical Reactors
A reactor where mixing is important is the tank flow or continuous stirred tank reactor
(CSTR). It is also referred to as a “back-mix” reactor. This type of reactor, like the batch
reactor,4,5 essentially consists of a tank or kettle equipped with an agitator. As with the
batch reactor discussed previously, perfect mixing is normally assumed so that both the
concentration and temperature within the reactor is the same, that is, there is no spatial
variation so that both terms are not a function of position within the reactor. This per-
fect mixing assumption also requires that both the concentration and temperature at the
discharge point (the outlet pipe) is identical to that in the reactor. As noted in Chapter 16
in Part III, Fogler6 noted that for perfect mixing: “The idea that the composition is identi-
cal everywhere in the reactor and in the exit pipes requires some thought. It might seem
that, since the concentration changes instantly at the entrance where mixing occurs, the
reaction occurs there and nothing else happens in the reactor because nothing is chang-
ing. However, reaction occurs throughout the reactor, but mixing is sufficiently rapid that
TABLE 19.2
Summary of Pipe Diameter Calculations; Illustrative Example 19.1
Pipe Size (in) Initial Charges ($/yr) Power Cost ($/yr) Total Costs ($/yr)
1.0 540 1359 1899
1.5 567 179 746
2.0 657 42 700
2.5 954 14 968
Total cost
Cost $/yr
Operating cost
Capital cost
Pipe diameter
FIGURE 19.1
Economic balance plot.
288 Introduction to Optimization for Chemical and Environmental Engineers
nothing appears to change with time or position.” One can extend this reasoning to also
conclude that the rate of reaction within the reactor is also uniform so that the extent of
conversion of reactant is the same within the reactor and the discharge port.
The CSTR essentially consists of a tank or kettle equipped with an agitator (see Figure 19.2)
and it may be operated under steady or transient conditions. Reactants are fed continu-
ously, and the product(s) are withdrawn continuously. The reactant(s) and product(s) may
be liquid, gas, or solid, or a combination of them all. If the contents are perfectly mixed,
the reactor design problem is greatly simplified for steady conditions because the mixing
results in the aforementioned uniform concentration, temperature, and so on, throughout
the reactor. This means that the rate of reaction is also constant and the describing equa-
tions are not differential and, therefore, do not require integration.
Consider the elementary irreversible reaction between benzoquinone (A) and cyclopen-
tadiene (B):
A + B → products (19.2)
If one employs a feed containing equimolar concentrations of reactants, the reaction rate
expression can be written as
−rB = kC A C B = kC 2B ; k = k B (19.3)
Proceed to calculate the reactor size requirements for one CSTR. Also, calculate the vol-
ume requirements for a cascade composed of two identical CSTRs. Assume isothermal
operation at 25°C where the reaction rate constant is equal to 9.92 m3/(kgmol/ks). Reactant
concentrations in the feed are each equal to 0.08 kgmol/m3, and the liquid feed rate is equal
to 0.278 m3/ks. The desired degree of conversion is 87.5%.
The rate equation, −rB, in terms of conversion variable X is4,5
−rB = −r = kC B2 (19.4)
Motor
Feed line
Impeller
Cooling or
heating jacket
Product
FIGURE 19.2
Typical continuous stirred tank flow reactor.
Select Chemical Engineering Applications 289
C B = C B0 ( 1 − X ) (19.5)
−r = kC 2B ( 1 − X ) (19.6)
2
FB0 X
V=
−r
(19.7)
FB0 X
=
kC B2 0 ( 1 − X )
2
Since
FB0 = C B0 q
= (0.08 )(0.278 )
(19.8)
= 0.02224 kgmol / ks
= 0.02224 gmol / s
The volume may be calculated employing Equation (19.7).
3
= 19.6 m
The previous design equation may also be applied to two reactors of equal volume in
series, noting that X1 − 0 = X1.3
FB0 X1
V1 = (19.9)
kC 2B0 ( 1 − X1 )2
FB0 X1
V2 = ; ∆X = X 2 − X1 (19.10)
kC 2B0 ( 1 − X 2 )2
For reactor 1,
kC 2B0 V1 X1
= 2
FB0 ( X1 )
1 −
For reactor 2,
kC 2B0 V2 X − X1
= 2 2
FB0 (1 − X 2 )
Since the LHS of both of the previous equations are equal, (V1 = V2),
X1 X 2 − X1
=
(1 − X1 ) (1 − X 2 )
2 2
290 Introduction to Optimization for Chemical and Environmental Engineers
X1 = 0.7251
Thus,
V1 =
( 0.278 )( 0.08 )( 0.7251)
( 9.92 )( 0.08 ) (1 − 0.7251)
2 2
= 3.36 m 3
In addition,
VT = V1 + V2
= ( 2 )( 3.36 ) (19.11)
= 6.72 m 3
The previously mentioned results – in terms of the volume requirements – for the two
cases may be compared. Using one reactor requires a volume of
V1 = 19.6 m 3
VT = V1 + V2 ; V1 ≠ V2 (19.12)
F X1 X 2 − X1
VT = B02 2 + 2 ; X 2 = 0.875
(19.13)
kC B0 (1 − X 1 ) (1 − X 2 )
To solve the previous equation for the intermediate conversion X1 under minimum volume
conditions, one needs to minimize VT, by setting
dVT
= 0; VT = VT ( X 1 ) (19.14)
dX 1
Select Chemical Engineering Applications 291
dVT 1 2X 1
= + − 64 = 0 (19.15)
dX1 ( 1 − X1 )2 ( 1 − X 2 )2
Solving by trial-and-error gives
X1 = 0.702
The volumes of the two individual reactors and the total minimum volume requirement
can now be calculated:
V1 =
( 0.02224 gmol / s ) ( 0.702 )
( 9.92 L / gmol ⋅ s ) (0.08 gmol / L)2 (1 − 0.875 )2
= 2.77 L
V2 =
( 0.02224 gmol / s ) ( 0.875 − 0.702 )
( 9.92 L / gmol ⋅ s ) (0.08 gmol / L)2 (1 − 0.875 )2 (19.16)
= 3.88 L
VT = V1 + V2
= 2.77 L + 3.88 L
= 6.65 L
The previous result is slightly lower than that calculated earlier (VT = 6.72 L).
The effect of using a cascade of three CSTRs that differ in size on the volume require-
ments for the reactor network may also be investigated. If the same feed composition and
flow are employed and the reactors are operated isothermally at 25°C, the minimum total
volume required and the manner in which the volume should be distributed between
the three reactors is another optimization problem. An overall conversion of 0.875 is to be
achieved again.
Proceeding as before, write the equation for the volume of each reactor and total volume
requirement for three reactor volumes that are not equal.
For each reactor
F X1
V1 = B02 2
(19.17)
kC B0 (1 − X1 )
F X − X1
V2 = B02 2 2
(19.18)
kC B0 (1 − X 2 )
with
F X − X2
V3 = B02 3 2
(19.19)
kC B0 (1 − X 3 )
292 Introduction to Optimization for Chemical and Environmental Engineers
F X 1 FB0 X 2 − X 1 FB0 X 3 − X 2
VT = B02 2 + 2 + 2
(19.21)
kC B0 (1 − X 1 ) kC 2B0 (1 − X 2 ) kC 2B0 (1 − X 3 )
Take the partial derivative of the equation obtained previously with respect to X1. Note that
the partial derivative must now be employed since the total volume, VT, depends on two
variables, X1 and X2, that is, VT = VT(X1, X2).7
∂VT FB0 1 2X 1 1
= 2 2 + 3 − 2
(19.22)
∂X1 kC B0 (1 − X 1 ) (1 − X 2 ) (1 − X 3 )
Set this derivative equal to zero.
∂VT FB0 1 2X 1 1
= 2 2 + 3 − 2 = 0
(19.23)
∂X 1 kC B0 (1 − X 1 ) (1 − X 2 ) (1 − X 3 )
Take the partial derivative for VT with respect to X2.
F X1
V1 = B02 2
kC B0 (1 − X1 )
= 1.19 L
F X − X1
V2 = B02 2 2
kC B0 (1 − X 2 )
= 1.58 L
Select Chemical Engineering Applications 293
TABLE 19.3
Trial-and-Error Calculation for Three
Reactor Set
X1 X2 Residual
0.6 0.8 11
0.59 0.7918 4
0.58 0.7835 −2.5
0.585 0.7879 0.8
F X − X2
V3 = B02 3 2
kC B0 (1 − X 3 )
= 1.95 L
VT = V1 + V2 + V3
= 1.19 L + 1.58 L + 1.95 L (19.29)
= 4.72 L
Note the reduction in the volume requirement when three CSTRs in series are employed.
The reader is left the exercise of showing that the total volume requirement for three
CSTRs of equal volume is
V1 = V2 = V3 = 2.17 m 3
VT = V1 + V2 + V3 = 6.51 m 3
with
X1 = 0.6710
X 2 = 0.7782
Theodore4,5 also calculated the reactor size requirement for a cascade composed of two
CSTRs with the first inlet reactor THRICE the size the second, the reactor size require-
ments for a cascade composed of two CSTRs with the first inlet reactor ONE-THIRD the
size of the second, if a very large number (approaching infinity) of CSTRs are employed,
and if the reaction is reversible for the previously mentioned conclusions.
or by a local exhaust system. In dilution ventilation, air is brought into the work area to
dilute the contaminant sufficiently to minimize its concentration and subsequently reduce
worker exposure. In a local exhaust system, the contaminant itself is removed from the
source through hoods.
A local exhaust is generally preferred over a dilution ventilation system for health prob-
lems because a local exhaust system removes the contaminants directly from the source,
whereas dilution ventilation merely mixes the contaminant with uncontaminated air to
reduce the contaminant concentration. Dilution ventilation may be acceptable when the
contaminant concentration has a low toxicity, and the rate of contaminant emission is con-
stant and low enough that the quantity of required dilution air is not prohibitively large.
However, dilution ventilation generally is not acceptable when the concentration is less
than 100 ppm.
In determining the quantity of dilution air required, one must also consider mixing char-
acteristics of the work area in addition to the quantity (mass or volume) of contaminant
to be diluted. Thus, the amount of air required in a dilution ventilation system is much
higher than the amount required in a local exhaust system. In addition, if the replacement
air requires heating or cooling to maintain an acceptable workplace temperature, then
the operating cost of a dilution ventilation system may further exceed the cost of a local
exhaust system.
SOLUTION:
The major components of an industrial ventilation system include the following:
1. Exhaust hood
2. Ductwork
3. Contaminant control device
4. Exhaust fan
5. Exhaust vent or stack
Several types of hoods are available. One must select the appropriate hood for a specific
operation to effectively remove contaminants from a work area and transport them into
the ductwork. The ductwork must be sized such that the contaminant is transported
without being deposited within the duct; adequate velocity must be maintained in the
duct to accomplish this. Selecting a control device that is appropriate for the contami-
nant removal is important to meet certain pollution control removal efficiency require-
ments. The exhaust fan is the workhorse of the ventilation system. The fan must provide
the volumetric flow at the required static pressure, and it must be capable of handling
contaminated air characteristics such as dustiness, corrosiveness, and moisture in the
air stream. Properly venting the exhaust out of the building is equally necessary to
avoid contaminant recirculation into the air intake or into the building through other
openings. Such problems can be minimized by properly locating the vent pipe in rela-
tion to the aerodynamic characteristics of the building. In addition, all or a portion
of the cleaned air may be recirculated to the workplace. Primary (outside) air may be
added to the workplace and is referred to as makeup air; the temperature and humid-
ity of the makeup air may have to be controlled. It also may be necessary to exhaust a
portion of the room air. Each of these variables needs careful analysis if a ventilation
system’s operation and performance is to be maximized.
Select Chemical Engineering Applications 295
To
stack
Duckwork
Particulate control
equipment
Makeup Return/
air recirculation air
Discharge/
Room/workplace exhaust air
Source
FIGURE 19.3
Industrial ventilation systems.
19.4 Heat Exchangers
In many environmental, chemical, and petrochemical plants there are cold streams that
must be heated and hot streams that must be cooled. Rather than use steam to do all the
heating and cooling water to do all the cooling, it is often advantageous to have some of the
hot streams heat the cold ones. In addition, the problem of optimizing the heat exchanger
networks that would be required has been extensively studied, but it remains a daunt-
ing exercise for practicing engineers.8–9 What follows is one simple illustration of a heat
exchanger network.
Consider the following application.10 A plant has three streams to be heated and three
streams to be cooled. Cooling water (90°F supply, 155°F return) and steam (saturated at 250
psia) are available. Devise a network of heat exchangers that will make full use of heat-
ing and cooling streams against each other, using utilities only if necessary. Information
on the three streams to be heated are presented in Table 19.4. Information on the three
streams to be cooled are presented in Table 19.5. Saturated stream at 250 psia has a tem-
perature of 401°F.
Here is the solution represented by Kaufman.10 The heating duties for all streams are
first calculated. The results are tabulated in Table 19.6.
The total heating and cooling duties are next compared:
Heating : 7 , 745, 000 + 6, 612, 000 + 9, 984, 000 = 24, 071, 000 Btu / h
Cooling : 12, 600, 000 + 4, 160, 000 + 3, 150, 000 = 19, 910, 000 Btu / h
TABLE 19.4
Heated Network Streams
Stream Flowrate (lb/h) cp [Btu/(lb⋅°F)] Tin(°F) Tout (°F)
1 50,000 0.65 70 300
2 60,000 0.58 120 310
3 80,000 0.78 90 250
TABLE 19.5
Cooled Network Streams
Stream Flowrate (lb/h) cp [Btu/(lb⋅°F)] Tin(°F) Tout (°F)
4 60,000 0.70 420 120
5 45,000 0.52 300 100
6 35,000 0.60 240 90
TABLE 19.6
Network Heating and Cooling Duties
Stream Number Stream Type Duty
1 Heating 7,745,000
2 Heating 6,612,000
3 Heating 9,984,000
4 Cooling 12,600,000
5 Cooling 4,160,000
6 Cooling 3,150,000
Figure 19.4 represents a system of heat exchangers that will transfer heat from hot
streams to the cold ones in the amounts desired. It is important to note that this is but
one of many possible schemes. The optimum system would require a trial-and-error
procedure that would examine a host of different schemes. Obviously, the econom-
ics would also come into play. Theodore9, and Abulencia and Theodore,6 discuss the
2 2
120° F 310° F
6,612,000 4
420° F
4 Steam
120° F 158.1° F
1,600,000 4,388,000 1,496,000
3 3
90° F 115.6° F 250° F
2,500,000
6
240° F
6 Steam
90° F 121° F
650,000 2,665,000
1 1
70° F 90° F 218° F 300° F
4,160,000
5 5
100° F 300° F
FIGURE 19.4
Flow diagram for a heat exchanger network.
Select Chemical Engineering Applications 297
ramifications that arise if the heat exchanger cost is a function of the heat exchanger
area; that is, the network should be such that the total cost of the resulting network is
minimized. They also discuss how to solve the problem so that the network’s entropy
increase11,12 and the network’s cost are minimized. The practicing engineer or scien-
tist should consider how to arrive at a reasonable solution to this energy conservation
problem.
Finally, highly interconnected networks of exchangers can save a great deal of energy
in a chemical plant. The more interconnected they are, however, the harder the plant
is to control, startup, and shut down. Often auxiliary heat sources and cooling sources
must be included in the plant design in order to ensure that the plant can operate
smoothly. The previously mentioned problem and considerations indicate how dif-
ficult it can be to solve some real-world industrial applications. The interested reader
is encouraged to contact the authors if other methods of solution are available and
possible.
19.5 Plant Design
Plant design requires the use of engineering principles and theories combined with a
practical realization of the limits imposed by industrial conditions. The development of a
plant-design project involves a wide variety of talents. Research, market analyses, design
of individual pieces of equipment, plant-location surveys, and many other problems are
involved in the design of industrial plants.10
Although the environmental and chemical engineer must understand the principles
related to the various design phases, the individual parts of the design are of little benefit
until they are combined to give one complete unit. The complete plant, i.e., the complete
unit – that is the design engineer’s goal. However, one major qualification must be added
before the goal is attained. This is the qualification that the operation must be able to pro-
ceed economically, and this often requires that energy be conserved. Concerns and details
regarding the conservation of energy follow.
One of the areas where meaningful energy conservation measures can be realized is in
the design and specification of process (operating) conditions for heat exchangers. This
can be best accomplished by the inclusion of second law principles in the analysis. The
quantity of heat recovered in an exchanger is not alone in influencing size and cost. As the
temperature difference driving force (LMTD) in the exchanger approaches zero, the “qual-
ity” heat recovered increases.5
Most heat exchangers are designed with the requirements/specification that the tem-
perature difference between the hot and cold fluid be at all times positive and be at least
20°F. This temperature difference or driving force is referred to by some as the approach
temperature. However, the corresponding entropy change is also related to the driving
force, with large temperature difference driving forces resulting in large irreversibilities
and the associated large entropy changes.11,12
The individual designing a heat exchanger is faced with two choices. He/she may decide
to design with a large LMTD that results in both a more compact (smaller area) design and
a large entropy increase that is accompanied by the loss of “quality” energy. Alternatively,
a design with a small driving force results in both a large heat exchanger and a small
entropy change with a larger recovery of “quality” energy.9,11
298 Introduction to Optimization for Chemical and Environmental Engineers
She also notes that the higher the discharge temperature of the heated air, t, the smaller
will be the temperature difference driving force, and the higher the area requirement of
the exchanger and, correspondingly the higher the equipment cost. Alternatively, with
a higher t, the “quality” of the recovered energy is higher, thus leading to an increase in
recovered profits (by reducing fuel costs).
Based on similar system designs, Francesco Ricci Consultants (FRC) has provided the
following annual economic models for the heat exchanger,
For the previously mentioned system, FRC suggests a value for the coefficients in the
cost model be set at
A = 10
B = 100,000
Employing the information provided, Shannon has been asked to calculate a t that will
TH = 500°F
Hot stream
T = 120°F
FIGURE 19.5
Proposed countercurrent exchanger; Illustrative Example 19.3.
Select Chemical Engineering Applications 299
She is also required to perform the calculation if A = 10, B = 4000, as well as A = 10,
B = 400,000. Finally, an analysis of the results is requested.
SOLUTION:
1. Since there are two contributing factors to the cost model, one may write the
following equation for the profit, P
B
P = A ( t − tc ) − ; TH = 500 and tc = 100 (19.32)
( TH − t )
For breakeven (BE) operation, set P = 0 so that
B
( t − tc ) ( TH − t ) = (19.33)
A
This may be rewritten as
B
t 2 − ( TH + tc ) T + + TH tc = 0 (19.34)
A
The solution to this quadratic equation for A = 10 and B = 100,000 is
dP B
= A− 2 = 0 (19.35)
dt ( TH − t )
Solving,
B
( TH − t )
2
=
A
B
TH − t =
A
t = 400°F
300 Introduction to Optimization for Chemical and Environmental Engineers
B = 4,000
B = 100,000
Profit t
100°F 500°F
B = 400,000
FIGURE 19.6
Heat exchanger results.
Upon analyzing the first derivative with t values greater than and less than 400°F,
Shannon observes that the derivative changes sign from + → − about t = 400, indicating
a relative maximum.
Similarly, for A = 10, B = 4000,
tBE = 300°F
tmax = 300°F
Graphical results for the three scenarios are shown in Figure 19.6.
Finally, there are a host of problems peripheral to this “quality energy” concern asso-
ciated with heat exchangers that the reader may choose to consider. This includes topics
that involve minimizing capital costs, minimizing operating costs, minimizing entropy
change(s), minimizing area requirements, minimizing energy requirements, and so on,
and/or any combination of these.
References
1. P. Abulencia and L. Theodore, Open-Ended Problems: A Future Chemical Engineering Education
Approach, Wiley‑Scrivener, Salem, MA, 2015.
2. W. Baasel, Preliminary Chemical Engineering Plant Design, 2nd edn, Van Nostrand Reinhold,
New York City, NY, 1990.
3. R. Huntington, High Pressure Pipe Line Research, Clark Bros., Olean, NY, 1947.
4. L. Theodore, Chemical Reactor Analysis and Applications for the Practicing Engineer, John Wiley &
Sons, Hoboken, NJ, 2012.
Select Chemical Engineering Applications 301
5. L. Theodore, Chemical Reactor Kinetics, A Theodore Tutorial, Theodore Tutorials, East Williston,
NY, originally published by the USEPA/APTI, RTP, NC, 1995.
6. H. Fogler, Elements of Chemical Reaction Engineering, 4th edn, Prentice-Hall, Upper Saddle River,
NJ, 2006.
7. H. Mickley, T. Sherwood, and C. Reed, Applied Mathematics in Chemical Engineering, McGraw-
Hill, New York City, NY, 1957.
8. J. Reynolds, Heat Transfer, A Theodore Tutorial, Theodore Tutorials, East Williston, NY, origi-
nally published by the USEPA/APTI, RTP, NC, 1995.
9. L. Theodore, Heat Transfer Applications for the Practicing Engineer, John Wiley & Sons, Hoboken,
NJ, 2011.
10. D. Kauffman, Process Synthesis and Design, A Theodore Tutorial, Theodore Tutorials, East
Williston, NY, originally published by the USEPA/APTI, RTP, NC, 1992.
11. L. Theodore, F. Ricci, and T. VanVliet, Thermodynamics for the Practicing Engineer, John Wiley &
Sons, Hoboken, NJ, 2009.
12. L. Theodore, Thermodynamics, A Theodore Tutorial, Theodore Tutorials, East Williston, NY,
originally published by the USEPA/APTI, RTP, NC, 1995.
20
Advanced Optimization Term Projects
The final chapter (20) in Part IV contains four illustrative examples that could best be
described as advanced optimization term projects. Illustrative Examples 20.3 and 20.4
were drawn, with permission, from the classic work of Happel1 titled, Chemical Process
Economics, John Wiley & Sons, Hoboken, NJ, 1957; one example is concerned with flow/
piping optimization and the other with a distillation process. The first example in this
chapter is concerned with a sulfuric acid process and the second with optimizing indoor
ventilation flowrates. Note that no introductory material for each example is provided in
this final chapter.
303
304 Introduction to Optimization for Chemical and Environmental Engineers
Rufus Hampton Jr., the plant engineer recently requested that the authors calculate
the maximum and minimum hourly flow rate of each stream that can be processed that
satisfies all of the aforementioned constraints and maximizes profit.
SOLUTION:
Employ Equations (20.1) and (20.2) in conjunction with the following two equations:
50 ≤ x1 ≤ 250 (20.4)
50 ≤ x2 ≤ 250 (20.5)
SOLUTION:
1. The applicable model for this case is (see also Part III, Illustrative Example 13.5):
dc
v0 (c0 − c ) + rV = V (20.6)
dt
Under steady-state, dc/dt = 0. Pertinent information includes
rV = 250 ng / min
c0 = 10 ng / m 3
c = 35 ng / m 3
Substituting gives
− rV
v0 =
c0 − c
rv
=
c0 − c (20.7)
250
=
35 − 10
= 10 m 3 / min = 353 ft 3 / min
dc
v0 (c0 − c ) + rV = V (20.8)
dt
For steady-state conditions dc/dt = 0. In addition, based on the information pro-
vided, r = 0 and c0 = 0. Therefore, and as expected,
c=0
306 Introduction to Optimization for Chemical and Environmental Engineers
t
− −
t
c = ci e τ
+ (c0 + k τ ) 1 − e τ (20.9)
Substituting gives
t
− t
−
11.73
20.7 = 85e + (10 + 2.48 ) 1 − e 11.73
Solving by trial-and-error gives (approximately)
t = 29 min
4. First, calculate the steady-state concentration for this condition. The applicable
model is obtained from part 2, after setting dc/dt = 0
c0 + k τ c
− = 0 (20.10)
τ τ
Solving and substituting gives
c = c0 + k τ
= 10 + (0.211)(11.73 )
= 12.48 ng / m 3
Since this is the steady-state concentration, it will take an infinite period of times to
reach this value. However, the steady-state concentration represents the minimum con-
centration achievable in the laboratory based on the conditions specified. Therefore, the
concentration will never reach a value of 12.2 ng/m3.
perfect-gas laws at the suction pressure, and a is the reference brake horsepower per mil-
lion cubic feet of gas per day. The investment in compressors is then
aQZ2
X ; dollars / mile (20.11)
L106
The fixed charge on compressors, α, includes maintenance, depreciation, and allowance
for profit on investment. The yearly fixed charge on the compressors is then
αXQZ2
; dollars / ( yr ) (mile) (20.12)
L106
Pipe costs Y dollars/ton. If W tons of pipe per mile are necessary for construction of the
line, the investment in pipe is YW tons/mile. For steel pipe W = 28.2(D + t)t where D is the
inside diameter of the pipe in inches, and t is the pipe wall thickness in inches. Barlow’s
formula for the wall thickness of pipe in terms of maximum allowable pressure, P lbf/sq
in., and maximum allowable stress, S lbf/sq in., is
t = P1A / 2 (S − P1 ) (20.13)
P1D P1D
YW = Y 28.2 D + ; dollars / mile (20.14)
2 (S − P1 ) 2 (S − P1 )
A good designation of laying cost is H dollars/mile for right of way, surveying, clear-
ing, supervision, engineering; G dollars/ton for unloading, hauling, aligning, placing, and
welding or coupling; and, N dollars/(mile)(in. diam.) for ditching, painting, back filling,
and laying equipment. The total pipe cost is then
B is the fractional fixed charge per year on the installed pipe. The yearly charges on the
pipe will be
P1D P1D
B ( Y + G)( 28.2) D + + BND + BH dollars / ( yr ) (mile) (20.16)
2 (S − P1 ) 2 (S − P1 )
Next, operating expenses must be considered. Labor, supervision, and salaries are taken at
$3850/month for a 4000-hp station. If direct proportionality is assumed for other stations
for approximately the same size, this expense is
QZ2
3.40 ; dollars / ( yr ) (mile) (20.18)
L106
308 Introduction to Optimization for Chemical and Environmental Engineers
About 10 std cu ft of fuel/bhp are required at a cost of Cf dollars/M std cu ft. The yearly
cost of fuel is
KQZ2 P1 P12
Cy =
6
kln ( P / P ) + b
+ K D 2
+ 2 + BND + BH + 100 (20.21)
2 (S − P1 ) 4 (S − P1 )
1 2 1
L10
where:
K 1 = 28.2B ( Y + G) (20.23)
The following values of various constants mentioned previously are believed to be repre-
sentative of 1957 costs:
With these constants provided, the cost Cy of operating a pipeline system for a specified
capacity is a function only of the pressures P1 and P2, the distance between stations L, and
diameter of pipe D. It is desired to compute the optimum values for these parameters for
the economic transportation of 150 MM std ft3/day of natural gas. Additional data from
various standard sources is provided by Huntington.2 The flow equation relating Q to the
independent variables involved is
Advanced Optimization Term Projects 309
( )
0.541
D 2.622 P12 − P2 2
Q = K2 (20.24)
( )
0.541
L0.541 Zavg
where:
D = inside diameter of pipe, in.
P1 = discharge pressure from compressors, lb/sq in. abs
P2 = suction pressure to compressors, lb/sq in. abs
L = distance between compressor stations, miles
Zavg = average deviation from perfect-gas laws for terminal pressures P1 and P2
with
0.541
80.8 M
K2 = (20.25)
ρ0µ 0.0814 T
where for methane
ρ0 = density, lb/cu ft at 60°F = 0.0422
μ = viscosity in ln/(ft)(sec) at 60°F = 7 × 10−6
M = molecular weight of methane = 16
T = absolute temperature, °F abs = 520°R
∂C y ∂ψ
+λ = 0 (20.26)
∂P2 ∂P2
∂C y ∂ψ
+λ = 0 (20.27)
∂L ∂L
∂C y ∂ψ
+λ = 0 (20.28)
∂D ∂D
whence
KQZ2a Q
− − λ 0.541 = 0 (20.30)
L2 106 L
P P12 Q
K 1D 1 + 2 + NB + λ 2.622 = 0 (20.31)
S − P1 2 (S − P1 ) D
The previous three equations and Equation (20.24) constitute four independent equations
with unknown P1, L, D, and λ. They are solved by elimination and on the assumption that
Z2 and Zavg are sufficiently close to being constant so that their derivatives will equal zero.
With these assumptions it is found that
P1
= 1.33 (20.32)
P2
and
NB
D 6.85 + D 5.85
K 1 [ P1 /(S − P1 ) + P12 / 2 (S − P1 ) ]
2
(20.33)
4.85KQ 2.85Z2Zavg a
=
(
K 12.85 P12 − P2 2 ) K 1 P1 / (S − P1 ) + P12 / 2 (S − P1 ) 106
2
The method of solution then involves calculations of P2 from an assumed value of P1.
D is then computed from Equation (20.31), and finally the flow Equation (20.24) is used to
solve for L.
In the present instance, a series of trials were made for a P1 variation between 300 and
800 lb/sq in. abs. Thus, assuming
P1 = 600
P2 = 451
Z2 = 0.933
Zavg = 0.923
P1 P2
+ = 0.0312
S − P1 2 (S − P1 )2
N = 500
B = 0.27
K1 = 28.2 × 0.27(240 + 30) = 2,060
K = 0.25(250) + 26.4(0.1) + 15 = 86.1
D 6.85 +
(500)(0.27 ) D 5.85 = D 6.85 + 2.10D 5.85
(2060)(0.0312)
The value of D is obtained by solving the equation
The solution of the previous equation by trial-and-error gives D = 19.0 in. The value of L
from the flow equation is
L=
(2.15 × 10 )(1.75 × 10 )(3.6 × 10 ) × 0.435 = 48.7 miles
5 6 5
(1.31 × 10 ) (0.923)
15
whence the yearly cost, Cy corresponding to the assumed 600 lb/sq in. abs discharge pres-
sure is
C y =
86.1 × 1.5 × 108 × 0.933 × 22.3
+ 2060 +
(19) × 0.0312 + 0.27 × 500 × 19 + 0.27 × 5900 + 100
2
48.7 × 106 2
C y = $21, 370
The results of the calculations at other discharge pressures for Q = 150 MM ft3/day are
given in Table 20.1. It appears that the optimum lies close to 400 lb/sq in. abs discharge
pressure for the conditions assumed. One notes that spacing between stations does not
vary a great deal.
TABLE 20.1
Calculation for Optimum Compressor Installation
Discharge Pressure for Inside
Compressor lb/sq in. Diameter of Distance between Annual
abs, P1 Pipe, in., D Stations, miles, L Cost, $/yr, Cy
300 26 51.6 $21,640
400 23 51.9 $21,180
500 21 53.6 $21,250
600 19 48.7 $21,370
800 17 53.1 $21,560
312 Introduction to Optimization for Chemical and Environmental Engineers
of the components involved, and the physical arrangement of the unit used for distilla-
tion. A common technique used for this separation is continuous distillation. This process
removes and absorbs the less volatile (heavier) components from the vapor, thus efficiently
enriching the vapor with the more volatile (lighter) components.6,7
It is often convenient to express the profitability equation for a mass transfer opera-
tion in terms of some dependent process variable other than the venture worth (W)
itself. Simplification that is possible is based on the observation that unit costs of each
process variable are usually expressible as simple functions of the independent vari-
ables. The modified equation to be differentiated graphically may then be expressed
as follows:
W = Aa + Uu + Vv + + Zz (20.34)
where unit costs of each process variable A, U, V,⋯, Z are expressed as simple functions of
the independent variables. The dependent variable a is taken as a desired process result,
and the independent variables u, v,⋯, z as the operating conditions whereby it is obtained.
Upon differentiation with respect to each of the independent variables, the following
equations result:
∂W ∂A ∂a ∂U ∂V ∂Z
= a +A +u +U+v ++ z =0
∂u ∂u ∂u ∂u ∂u ∂u
∂W ∂A ∂a ∂U ∂V ∂Z
= a +A +u +v + V ++ z =0
∂v ∂v ∂v ∂v ∂v ∂v (20.35)
∂W ∂A ∂a ∂U ∂V ∂Z
= a +A +u +v ++ z +Z=0
∂z ∂z ∂z ∂z ∂z ∂z
A system for employing Equation (20.35), using the plan of variation of one variable at a
time is provided in the following mass transfer application1 developed by Harbert5 that
illustrates its employment in a mass transfer problem involving two independent variables
where process correlations are available for the full range of interest of the variables.
Consider the following application. A chemical by-product is obtained at a constant
rate of 28 mole % solution in water. It is desired to recover this by-product at a purity
corresponding to 99.2 mole % as the overhead from a distillation column. One percent
recovery of this chemical for one day’s operation is worth $4.20. Sufficient heat pickup is
available to bring the feed to its boiling point. The relative volatility for dilute water in a
chemical (top of column) is 1.58, and for a dilute chemical in water (bottom of column)
is 1.75.
The cost of one additional MM (million) Btu of reboiler heat to the column is $0.60. This
figure includes the cost of heat itself, the cost of the condenser cooling to remove the heat,
and the cost of additional reboiler and condenser surfaces, allowing for an appropriate
profit for the investment involved for the additional surface.
For the completely installed column, the cost of one additional plate together with the
section of tower it occupies is estimated to be $96 plus $44/sq ft of area. With 18-in. plate
spacing, 1 sq ft of plate area will handle a vapor load corresponding to 20 MM Btu of
reboiler heat per day. Plate efficiency for these conditions is taken at 90%. It is estimated
that maintenance will amount to 2%/yr of installed cost and that operating time will be
300 days/yr. Management expects a payout time at most of two years.
Advanced Optimization Term Projects 313
It is desired to determine the conditions for the most economical design. The operating
cost may be expressed in terms of the variables,
The interrelations between these variables as calculated for given conditions are shown
in Figure 19.3, which provides the product recovery versus the reboiler heat input for tow-
ers of various numbers of plates (a vs. u at constant v). These curves were obtained by a
trial-and-error computation, using McCabe–Thiele diagrams.6,8
The first step in determining the required optimum is to express the costs of additional
units of the variables (values of A, U, and V) on the basis of dollars per operating day. For
the product recovery, the value given for one additional percent for one day is
A = 4.20
For the reboiler heat variable, the cost per day for 1 MM Btu is $0.60 plus the cost of the
additional plate area required. From the data given, the cost of this ($1/20 sq ft) is
The negative sign denotes a cost instead of profit item. For the number of plates variable,
the total area required per plate is (u/20) sq ft, so that the original cost of one additional
plate is 96.0 + 44.0(u/20). The cost per day of an equilibrium plate then becomes
∂U
= −0.00423 (20.38)
∂v
∂V
= −0.00423 (20.39)
∂u
Therefore, from Equation (20.35)
∂a
4.20 − (0.60 + 0.00423 v + 0.00423 v ) = 0
∂u
(20.40)
∂a
4.20 − (0.60 + 0.00846 v ) = 0
∂u
314 Introduction to Optimization for Chemical and Environmental Engineers
where ∂a/∂u is the slope of the curves of Figure 20.1. Figure 20.2, the cross plot of the data
of Figure 20.1, as percent recovery versus number of plates for constant amounts of reboiler
heat (a vs. v at constant u) gives, as the slope of its curves, values for ∂a/∂v.
Similarly,
∂a
4.20 − (0.185 + 0.00846 v ) = 0 (20.41)
∂u
A set of reasonable values for the independent variables is next assumed: u = 180, v = 50.
Substitution of v = 50 into Equation (20.40) gives
∂a 1.023
= = 0.244 (20.42)
∂u 4.20
The corresponding value of u is 165 MM Btu/day from Figure 20.2. Substitution of u = 180
into Equation (20.40) gives
∂a 1.705
= = 0.406
∂v 4.20
The slope ∂a/∂v = 4.06 corresponding to the line u = 180 on Figure 20.2 gives v = 39. As a
second trial, u = 165 and v = 39 are assumed. This gives, first ∂a/∂v = 0.222 corresponding to
u = 164, and second, ∂a/∂v = 0.382 corresponding to v = 38.
Alternatively, one might have selected a series of values of v and plotted Equation
(20.50) on Figure 20.1 (dashed line). Similarly, Equation (20.40) can be plotted on Figure 20.1
(dot-dash line). Values for Equation (20.35) could be cross-plotted to Figure 20.1 (dashed
line). The two curves on Figure 20.2 would intersect as shown at approximately the same
value as obtained by the two successive trails; namely u = 164, v = 38.
100
∞ ys
mt f
ra
ilib o. o
equ υ = n
riu
60
55
a, product recovery, % of total
98
50
45
96
40
35
94
Eq.
(3.15)
30
92
160 170 180 190 200 210 220
u, reboiler heat duty, MM Btu/day
FIGURE 20.1
Recovery versus number of plates.
Advanced Optimization Term Projects 315
100
day
ler u/
b oi Bt
re M
u, y, M
a, product recovery, % of total
98 t
du
at 0
he 22 0
1
2 Eq. (3.16)
0
Eq. (3.15)
20
0
19
96
0
18
0
17
94
0
16
92
30 35 40 45 50 55 60
υ, number of plates
FIGURE 20.2
Recovery versus number of plates.
The economic design thus corresponds to 38 equilibrium plates with a reboiler duty of
164 MM Btu/day. Resultant chemical recovery is 93.4%. The tower would have 38/0.9 or 42
actual trays with a vapor area of 164/20 = 8.4 sq ft. This design would result in a maximum
overall profit for the operation.
References
1. J. Happel, Chemical Process Economics, John Wiley & Sons, Hoboken, NJ, 1958.
2. R. Huntington et al., High Pressure Pipe Line Research, Clark Bros., Olean, NY, 1947.
3. R. Ketter and S. Prawler, Modern Methods of Engineering Computations, McGraw-Hill, New York
City, NY, 1969.
4. H. Mickley, T. Sherwood, and C. Reed, Applied Mathematics in Chemical Engineering, McGraw-
Hill, New York City, NY, 1957.
5. W. Harbert, title unknown, Petroleum Refinery, 27, 185, location unknown, 1948.
6. L. Theodore and J. Barden, Mass Transfer, A Theodore Tutorial, Theodore Tutorials, East
Williston, NY, originally published the USEPA/APTI, RTP, NC, 1995.
7. L. Theodore and F. Ricci, Mass Transfer Operations for the Practicing Engineer”, John Wiley &
Sons, Hoboken, NJ, 2010.
8. L. Theodore, Chemical Engineering: The Essential Reference, McGraw-Hill, New York City, NY,
originally published by USEPA/APTI, RTP, NC, 2014.
Index
317
318 Index
Rotary vacuums filters, 220 ST, see Shell-and-tube (ST) heat exchangers
Round-off error, 23, 30, 36 Statistical methods, 76
Runge–Kutta (RK) method, 32–4 Steam jet systems, 176
Steam options, 235–237
Safety expenditures vs. return of safety Steam Tables, 236
investments, 165–167 Steepest ascent/descent method, 60–63
SCC, see Secondary combustion chamber (SCC) Straight blade fans, see Radial blade fans
Scientific notation, 20 Structure design, plant, 261–262
Search methods, 51–63 Sulfuric acid process, 303–304
bisection, 56 Superinsulation, 238
golden section, 57–60 Superproduction costs, 222
interval halving, 53–55
steepest ascent/descent, 60–63 Tank farms, 252–254
Secondary combustion chamber (SCC), 141 TF, see Tubular flow (TF) reactor, continuous
Second law of thermodynamics, 246, 247 stirred tank reactor (CSTR) vs.
Second-order polynomial equation, 25 Theodore, Lou, 157, 243, 280, 283, 296, 298
Seek methods, see Direct approach Theodore approach, 5
Selectivity, 210–212 Theohan Gas Company, 84
Semi-logarithmic (semi-log) coordinates, 73–76 Thermal conductivity, 238
plotting straight lines, 73–76 Thermal insulation, 238–239
triangular (trilinear) coordinates, 76 Thermal waste treatment, 138
SF, see Slope factor (SF) Three independent variables, 48–49, 93
Shaefer, S, 280 Three-point equal-interval search method, 53
Shell-and-tube (ST) heat exchangers, 239, 241 Three-stage compressor, 195–196
Shipping facilities, 249–252 “Three T’s of combustion,” 120
Significant figures, 20–22 TNT, 172–174
Simpson’s rule, 28 Toxicity, 229
Simultaneous linear algebraic equations, 29–31 Transient operation, 212
Gauss elimination, 30 Trapezoidal rule, 27–28
Gauss–Jordan reduction, 30 Treybal, R., 229
Gauss–Seidel approach, 31 Trial-and-error method, 5, 41, 60, 98, 99, 104, 296,
Single-pass cylindrical tube, 201 313
Single-payment compound amount factor, 144 Triangular coordinates, 76
Single-phase heat exchanger, 241 Truncation error, 23
Single solvent, 230 Tubular flow (TF) reactor, continuous stirred
Single-variable optimization, 97 tank reactor (CSTR) vs., 199–204,
Slope factor (SF), 155, 156 208–209
Slope-intercept method, 65, 70, 73 Tumble burners, 141
Slurry flows, 219 Two independent variables, 45–48, 89–92
Solidification, 138 Two-phase heat exchanger, 241
Solid–liquid extraction, 229 Two-stage compressor, 189–191
Solid waste management, 137–150, 271–272
cement kiln vs. rotary kiln incinerator, Ultrafiltration (UF), 133
139–141 Unconfined vapor cloud explosions, 172
discounted cash flow application, 144–147 Unimodal function, 51, 60
maximum mercury waste emission Unsteady-state operation, see Transient
constraints, 147–149 operation
sludge waste receiving tank size, 142–143 US Department of Agriculture (USDA), 160
solid waste treatment, 137–139
structural considerations of concrete mix, Vacuum insulation, 238
149–150 Vapor–liquid changes, 236
Solid waste treatment, 137–139 Vapor–solid changes, 236
Solvent, 230 Ventilation models, 191–194, 259–261, 293
Index 323