UMUC's "Commentary" for CMIS 160

Chapter 1

Propositional Logic

A. Statements and Logical Expressions Read Epp: pp. 1-8

Don't rush through the topics in this section, because most of the remaining material in the module is made up from the pieces presented here. Study these topics until you can do the exercises with confidence.

Try to get a clear understanding of the difference between a statement (defined on page 2) and a sentence; it will help you throughout the module. The two terms are often used interchangeably in an everyday sense when we speak or write, but logic requires greater precision. For example, "Discrete Mathematics with Applications is a good textbook" is a sentence but not a statement, while "Discrete Mathematics with Applications is a textbook" is both a sentence and a statement. What is the difference between a statement and a sentence?

A sentence that is not a statement could be true sometimes (or under some conditions) and false at other times (or under different conditions). A sentence that is a statement must have only one of the values true or false at all times (or under all conditions).

We can assign a variable (a letter) such as p to a statement, then say that p can be T (true) or F (false). Once we do that, it is possible to talk about statements in the abstract as p or q or r without using English to express them. They are variables with values and the values are true or false. Try exercise 5 on page 15 to get an idea of what is involved.

If p is a statement, then not p (~p) is also a statement. Its truth values are the opposite of p's truth values. If p and q are statements, then their conjunction (p and q) and disjunction (p or q) are also statements. Truth tables for not, and, and or are representations of the definitions of negation, conjunction, and disjunction. In general, truth tables are convenient representations of the truth values of statements. Truth tables give all possible true-false values of a statement.

Any statement that combines simple statements (p, q, r) with logical operators (ands, ors, nots) is called a compound statement. The representation of a compound statement in symbols is called a logical expression, or statement form, and is denoted by an upper-case letter such as P, Q, or R. Such an expression or form also has a truth table to represent all possible true-false values.

Once the basic rules and definitions are known, the construction of a truth table can be a mechanical process. To begin with, a truth table for a simple statement p has two rows, because p has only two possible values: true and false. A truth table for compound statements made up from p and q has four rows, because for each value of p there are two possible values of q. A truth table for compound statements made up from p, q, and r has 2 x 2 x 2 = 8 rows. Why? A truth table for a compound statement can be constructed in steps.

After studying the definitions and working through the examples on pages 4-8, try exercises 6-10 and 12-16 on page 15. Begin the construction of each truth table by completing columns for p, q, and r separately before working on the ands, ors, and nots. Then, allow a column for each logical operation (and, or, not) in the order of its evaluation, as shown in the example. Note that parentheses have the same effect as in algebra: operations inside parentheses are performed first.

B. Logical Equivalence, Tautologies, and Contradictions Read Epp: pp. 8-14

Truth tables can help clarify the three main topics in this section. Construct the truth tables for two statement forms P and Q, then examine the results. If the truth tables are identical, then the statement forms are logically equivalent. In that case, one statement form can be substituted for another without affecting the results. This substitution can simplify complicated expressions, making them easier to understand. Logical equivalence of expressions can also be used to simplify a computer program or a digital logic circuit, discussed later in this module.

If the same row of the two truth tables has different values (one T and the other F), then the two statement forms are not logically equivalent. Examples 1.1.7 and 1.1.8 illustrate non-equivalence and equivalence. You should be able to establish the equivalencies in Theorem 1.1.1 by using truth tables. To test yourself, try exercises 17-32 on pages 15 and 16.

A statement form is a tautology when its truth table contains T (or 1) in every row of the column showing the final result, meaning that the statement is always true regardless of the values of any of the simple statements from which it is composed. A statement form is a contradiction when its truth table contains F (or 0) in every row of the column showing the final result. By definition, a contradiction can never be true. To test your understanding of these concepts, first work through our examples illustrated by the hot-links for tautology and contradiction. Then work through the examples on pages 12 and 13, and finally, try exercises 37-40 on page 16.

Exercise 50 on page 16 provides an example of a complicated set of criteria that can be simplified, or at least clarified somewhat, by using propositional logic. There are five simple statements in the exercise. How many rows will be in the truth table, assuming all five statements are used in a compound condition for terminating membership?

C. Conditional Statements Read Epp: pp. 17-27

The if-then construction is frequently used in writing, programming, and conversation. It is a particularly troublesome construction in logic because its definition as depicted by a truth table doesn't conform to our usual way of thinking about the statement.

For example, we don't like to think of the statement "If it is raining, then I will get wet" as true when, in fact, it is not raining. But in logic, the conditional statement "If it is raining, then I will get wet" is defined to have the value T when the statement "It is raining" has the value F. The key ingredient here is "defined." There is no real explanation necessary (or, perhaps, possible) for a definition. Note that the author tries to convince you by giving an example at the bottom of page 17, but the definition (and truth table) on page 18 is all that matters, whether you are convinced or not. Before proceeding further, carefully work through the examples on pages 18-20 and try exercises 1-8 on page 27.

Study the definitions of contrapositive, converse, inverse, only if, and biconditional on pages 21-24 and use the truth tables to distinguish among the terms. Try exercises 8-19 on page 27 to help fix the concepts in your mind.

You will often encounter the terms necessary condition, sufficient condition, and necessary and sufficient condition in mathematics and in technical writing. The terms are alternatives to the if-then forms you have already studied. Study the definitions of these terms given on page 25. Work through the examples on that page, then try exercises 38-40 on page 28. You should also read the "Remarks" section on pages 26 and 27 to better understand some of the distinctions between logic and informal-language usage.

D. Arguments Read Epp: pp. 28-39

Before you begin this section, try to clear your mind of the usual thoughts that occur when you hear or use the word "argument." Then read Epp's definition of argument. In logic, the main concern is whether an argument is valid (or invalid) as Epp defines it. This use is quite different from what "argument" usually implies in informal language. One example of a valid argument is a proof of a theorem in mathematics.

Although there are many terms and issues in this section, you will be expected to know only how to determine whether an argument is valid by using truth tables. The steps to follow are listed at the bottom of page 29 and two examples are given on page 30.

One way to save time in determining the validity of an argument is to know a collection of valid argument forms, such as those on pages 31-33 (also see the summary on page 39). You will be expected to recognize and use only the two most common forms, modus ponens and modus tollens, defined on pages 31 and 32. You will also be expected to recognize the contradiction rule argument form on pages 37 and 38. Try exercises 1-11 on pages 39 and 40 to assess your understanding.

You will not be expected to know the material on pages 34-39. There are, however, some interesting examples on these pages. You may have encountered in puzzle books the type of logical problem in examples 1.3.9 and 1.3.15.

Also, note the definition of fallacy on page 35 and the CAUTION! on pages 36 and 37 that includes the examples of a valid argument having a false conclusion and an invalid argument having a true conclusion. Examples of logical misuse are all too common in discussions, in the media, or in advertisements.

E. Application Read Epp: pp. 41-54

This section connects logical expressions, truth tables, and digital logic circuits, such as those in computer systems. Pages 41-44 provide background information related to the topics in the rest of the section.

The input and output of digital logic circuits, also called combinational circuits, are represented by the binary digits 1 and 0 in I/O tables (instead of the T and F in truth tables of logical expressions). P, Q, and R stand for inputs with values 1 (T) and 0 (F) in some part of a digital logic circuit. These inputs pass through and, or, and not gates. The output (1 or 0) at each gate is determined by the values entering the gate and the type of gate.

For example, if P and Q are inputs to an and gate, and P = 1 and Q = 0, then the output P and Q is 0 (just as the logical expression p q is F when p is T and q is F). Figure 1.4.3 on page 45 shows types of gates, their symbolic representation, and their I/O table. Exercises 1-4 on page 55 will test your ability to determine outputs for specific inputs in a circuit.

An I/O table for a digital logic circuit is constructed by evaluating outputs for all possible inputs of 1 and 0 as they pass through a sequence of and, or, and not gates, just as constructing a truth table of a compound logical form is obtained by evaluating each simple form in sequence for all possible true-false values.

Click here to see a digital logic circuit with 1s and 0s passing through gates, resulting in an I/O table.

The term Boolean expression applies to any expression consisting of (1) variables that take binary values and (2) and, or, and not operators. A digital logic circuit can be represented by a Boolean expression as shown in example 1.4.3. A circuit, as you have already learned, can also be represented by an I/O table. It seems reasonable, therefore, that a Boolean expression can be represented by an I/O table. Finally, it follows that given any one of the three entities--digital logic circuit, a Boolean expression, or an I/O table--the other two entities can be determined. Work through examples 1.4.4 and 1.4.5, then try exercises 5-21 on pages 55 and 56.

The comparison of logic and circuits can be continued. Equivalence plays an important role. When the truth tables of two logical forms are identical, the forms are equivalent and either form can replace the other. When two digital logic circuits have identical outputs in their I/O tables, then the circuits are equivalent and the simpler one can replace the more complicated one, resulting in an increase in speed in any system using the circuit. See pages 52 and 53 for an example, then try any of the exercises 26-30 on pages 56 and 57, according to your own interest.

***************************************************************

Chapter 2

Predicate Logic

A. Basic Concepts Read Epp: pp. 75-77

Predicate logic combines the notion of predicate from natural language with set and functional notation from mathematics. Here is an example: Assume the predicate P represents the phrase "is currently a student in CMIS 160." For what values of x is the statement P(x): "x is currently a student in CMIS 160" true? Without thinking much about it, your answer is probably something like "all of the names of students on the CMIS 160 class list." For example, your name is one value of x for which the statement is true. More specifically, if Mary is a current student in CMIS 160, then P(Mary): Mary is currently a student in CMIS has the value T.

Let's formalize the above ideas and add notation. Let P stand for the words "is currently a student in CMIS 160" Let S be the set of all current students in UMUC Let C be the set of all current students on the CMIS 160 class list

Then P(x) is "x is currently a student in CMIS 160." If a student x is in S but not in C, then the value of P(x) is false. For every student x who is in C, P(x) has the value T. Another way of saying this is: "the truth set of P(x) is C." In notation, the truth set of P(x) is

{x S | P(x)} = C

where the symbol | is read "such that."

This notation appears in the definition on page 77, but first study the definitions of predicate and domain on page 76. Note that truth tables are not involved here. Instead you are dealing with sets of elements for which a predicate is true.

Study the notation in the blue-shaded box on page 77. Note the similarity in concept between the arrow and subset. In particular, logic uses the single-direction arrow to indicate that one truth set is a proper subset of another, and the double-direction arrow to indicate identical truth sets.

B. Quantifiers Read Epp: pp. 77-80

"For all" ( ) and "there exists" ( ) provide additional ways to distinguish elements of truth sets. On pages 78 and 79, note the different ways to express the universal quantifier and the existential quantifier , then study the definitions and read through the two examples. Although the examples use infinite sets, you should be able to extract the ideas involved in establishing the truth and falsity of quantified statements.

As in your work with sets, finding a counterexample is a standard and effective way to establish the falsity of a universal statement. For finite sets, the method of exhaustion is also a standard and effective way to establish the truth of a universal statement. In comparison, for statements using the existential quantifier, you only need to find one element to establish truth, but all elements to establish falsity.

Using the example above, Let P stand for the words "is currently a student in CMIS 160" Let S be the set of all current students in UMUC Let C be the set of all current students on the CMIS 160 class list

To show the truth of

an x in S such that P(x),

all you need to do is identify one student who is currently in CMIS 160.

To show the truth of

x in C, P(x),

you need to find all students in C on the class list for CMIS 160.

However, to show the falsity of

x in S, P(x),

you only need to find one student in S who is not on the CMIS 160 class list.

After working through examples 2.1.4 and 2.1.5, you should begin to appreciate the use of notation in logic. Such notation gives precise form and meaning to sometimes confusing informal language sentences. Try exercises 8-10 on page 87 to acquire some facility with translating between formal and informal language.

C. Universal Conditional Statements and Negations Read Epp: pp. 80, 81, 83-87

Combining the universal quantifier with a conditional statement produces a form that is commonly encountered in mathematical writing. Examples 2.1.6 and 2.1.7 illustrate the informal and formal versions. Try exercises 13-16 on page 88 to test your understanding and to gain some practice in writing formal and informal versions of the universal conditional statement.

Negations of quantified statements, especially the universal conditional statement, also are important in informal and formal discourse. On pages 83 and 84, study the two theorems and Epp's informal statements describing them (in bold). Work through examples 2.1.10-2.1.12. Pay particular attention to the CAUTION! on page 85.

You can find in the media examples of the type of misuse of negation described in the CAUTION! Look for statements such as "All _____ not ______" (e.g., "All is not well in this country" or "All humans are not equal"). The misuse is much more frequent than you might expect.

Negations of universal conditional statements are a bit more complicated. The symbolic form of a negation is given on page 86. After working through example 2.1.13, try exercises 19, 24-28, 30, and 34. The "Vacuous Truth of Universal Statements" section provides some interesting reading.

D. Multiple Quantifiers, Relationships among Quantifiers, Universal Conditional Statement Variants, and Prolog Read Epp: pp. 89-97

Here we expand predicate logic topics and concepts introduced previously. Pages 89-92 acquaint you with writing statements containing two quantifiers: combinations of (there exists) and (for every). Study examples 2.2.1 and 2.2.2 to learn how to write statements with more than one quantifier both formally and informally. Then try exercises 3-5, 9, and 10, and part (a) of 12, 15, and 16 on page 97 to test your understanding.

Example 2.2.4 illustrates a technique for finding truth values of multiply quantified statements. Note that the example does not generate a truth table. Exercise 20 on pages 97 and 98 continues this example.

Study the discussion of negating multiply quantified statements on pages 90-92, including example 2.2.5. Sometimes the negation of a statement can be better understood than the statement itself. Try part (b) of exercises 12, 15, and 16 on page 97 to test your understanding.

You will not be responsible for the rest of the material of pages 92-97, but read it to gain some appreciation for the complexity involved. In particular, note the brief Prolog section; it suggests how this computer language might be applied to logical problems.

***************************************************************

Chapter 3

Methods of Proof

Direct Proof and Counterexample Read Epp: pp. 112-121

The topics in this section apply more to mathematics than to computing. Although these topics are beyond the scope of this course, you should read the pages to become acquainted with the ideas and discussions that are given. In particular, skim the following topics:

the introductory material on pages 112 and 113 the notion of constructive proofs of existence on pages 115 and 116 the limitations of proof by exhaustion on page 116 the steps involved in the method of direct proof the discussion in the solution to example 3.1.5 on pages 117 and 118 the directions for writing proofs of universal statements on pages 119 and 120 the common mistakes identified on pages 120 and 121

When you gain experience in writing a fairly long computer program from specifications, given to you or determined by you, then you will be able to relate the programming process you go through to that of writing a proof given a set of conditions or axioms. (Note the reference in Knuth discussing this relationship, given at the bottom of page 118.)

***************************************************************

Chapter 4

Proof by Mathematical Induction Read Epp: pp. 194-201

Mathematical induction is a proof technique that can be applied to computing, primarily in proving program segments correct and in determining algorithm or program efficiency (for example, in determining how many times an operation is performed in an algorithm). Induction is essentially an inverse process of recursion, a standard process in computer algorithms and programming. Recursion is introduced in the next module.

Carefully work through the explanations on pages 194-197. Study the "Principle of Mathematical Induction" given on page 196, which should give you an intuitive idea of the process of a proof by mathematical induction. Then study example 4.2.1 to get a better idea of what constitutes a proof by mathematical induction. You may want to review some algebraic rules before you begin.

Example 4.2.1 illustrates the typical amount of work involved in the proof and typifies the two-step structure of the proof. In general, step one, establishing the truth of the proposition for the lowest value of an integer, does not require much work; a mere statement or simple equation might be enough. Step two, showing that the proposition is true for the integer value n + 1 (assuming it is true for all n between the lowest integer value and n) usually requires the application of algebraic rules, such as factoring or exponentiation.

In trying to explain the process of proof by mathematical induction, Epp includes explanations within the proof. They are not part of the formal proof and might lead to some confusion. They also make the proof seem longer and more complex than it really is. Rewrite the proof of Theorem 4.2.2 on pages 199 and 200 without the italicized statements. Doing this might help clarify what is required in the proof. Then try exercises 3-5, 7-9, and 12 on pages 204 and 205 to test your ability to apply the proof process.

Only one form of proof by mathematical induction is considered in this course. There are several other forms that apply to different types of propositions. The remaining parts of the chapter in Epp discuss two additional forms, give examples of each, and provide an application in proving a program segment correct. These topics are beyond the scope of this course, but you may want to read about them for your own interest.

***************************************************************

Chapter 5

Set Theory

A. Elements of a Set Read Epp: pp. 231 and 232

It might seem strange that the key terms, set and element, are undefined, but there really are no simpler words for defining these terms. Be sure you understand what a set and an element of a set are. Naming a set immediately establishes a binary condition: either an item is an element of that set or it isn't. There is no other possibility.

Epp uses the upper case letters R, Z, and Z+ consistently in the text to represent, as shown below, the set of all real numbers, the set of all integers, and the set of all positive integers: R the set of all real numbers Z the set of all integers Z+ the set of all positive integers

Although these sets, or subsets of them, appear in various places throughout the text, in this course we emphasize only the finite subsets of these infinite sets, rather than the infinite sets themselves or their infinite subsets.

B. Subsets Read Epp: pp. 232 and 233

A is a subset of B if, and only if, every element of A is an element of B. Click the Animate button to see this concept illustrated.

In the example below, C is not a subset of B because not every element of C is an element of B.

A is a proper subset of B if, and only if, every element of A is an element of B, but there is at least one element of B that is not in A. In the first animated example above, A is a proper subset of B.

Learn the definitions of subset and proper subset, given also in the highlighted portions of pages 232 and 233. If you are unfamiliar with set notation, refer to the inside front cover of Epp's text.

Use Venn diagrams to depict the sets in example 5.1.3, and verify the solutions by drawing Venn diagrams. Then try exercises 4-6 on page 242 to check your understanding of subsets, using Venn diagrams as an aid in seeing the solutions to these problems as well.

The distinction between element of a set and subset of a set as illustrated in example 5.1.5 is worth noting. Test your understanding by doing exercise 2 on page 242.

C. Set Equality Read Epp: pp. 234 and 235

Whether or not sets are equal might seem obvious, but set equality can be difficult to prove. Learn the definition and test your understanding by doing exercises 1 and 3 on page 242.

If you have time, browse through example 5.1.6, but don't get discouraged by it. The solution contains some formal proof methods that you will not need to know at this time. You will become more adept at argumentation of this kind as you proceed through the course.

D. Operations on Sets Read Epp: pp. 236 and 237

The definitions of the operations intersection, union, difference, and complement are given on page 236. To see animations illustrating these operations, click on the buttons below the picture of sets A and B. <../../images/blank.gif>

Learn these definitions of operations on sets. The result of each operation is always a set. Note the role played by the universal set.

Venn diagrams are especially helpful in visualizing results of set operations. Test your understanding of the operations by trying the examples and exercises 8-10 on pages 242 and 243. Exercises 12-16 are more complex: try them after you finish the others.

E. Cartesian Products of Sets Read Epp: pp. 237-239

The definitions and notation for ordered n-tuples and Cartesian products were used in the Application Context discussion. Study the definitions and examples on pages 237-239, then try exercises 17 and 18 on page 243.

F. Formal Languages Read Epp: pp. 239-242 (optional)

We will study the material on formal languages in module 5 of this course. You do not need to study it now. (If you have algorithm development and programming experience, you might be interested in the discussion of testing for a subset on pages 241 and 242, but it is optional for this course.)

G. Proving or Disproving Set Properties Read Epp: pp. 244-256

This section on properties of sets may be difficult to read, especially if this is your first encounter with proofs in mathematics. Work on one part of a page or, at most, a few pages at a time. Draw Venn diagrams as visual aids whenever you can. (Epp has highlighted set properties, theorems, and proof methodology.) The use of a counterexample to establish that an alleged property is false is an important proof methodology.

A proof is a precise sequence of true statements that flow logically from one or more assumptions to a conclusion. The proof itself requires no explanatory remarks, but it is good practice to include with each statement a reference to a definition, property, or theorem so that the reader can follow the argument more readily. Epp explains the process to follow in great detail, then shows the actual proof. Note the difference between the explanation and the proof by studying example 5.2.2. The proof is considerably shorter than the explanation.

When confronted with an alleged theorem, you have two choices: prove it or disprove it. Disproving an alleged theorem (i.e., proving that a property is false) requires finding one counterexample (a case for which the property is not true). For a theorem or property to be true, it must apply to all cases, not just some. Thus, if you can find just one instance for which the alleged theorem or property is not true (a Venn diagram might help in this regard), then you have completed the task. If you cannot find a counterexample, then try to construct a proof.

Before proving a statement, it is necessary to understand what the statement says. Separate the assumptions from what is to be proved. After doing this, proceed to the proof itself. A proof of a set property or theorem usually consists of the following steps:

Write down what is to be proved. Write definitions and/or already proven theorems that apply to the situation. Assume x is an element of the set on the left side of a relation. Use properties and/or theorems to deduce that x is an element of the right side. When appropriate, repeat steps 1-4 by reversing the left and right sides.

The box outlined in blue on page 249 illustrates this process. Note that step 5 is appropriate only when the theorem to be proved involves set equality.

H. The Empty Set, Partitions of Sets, and the Power Set of a Set Read Epp: pp. 258-265

Study the definitions and properties, and theorems pertaining to the empty set. The empty set plays an important role in set theory. For example, it provides a handy way to define two or more sets as disjoint, i.e., not having any elements in common. (See page 262.) Also, remember that the empty set is a subset of every set.

The definition of disjoint sets leads to the definition of partitions of sets. A given set can be partitioned in a number of ways. For example, the set of all students in this course can be partitioned into the two subsets, all males and all females. Another partition is all students younger than 25, all students aged 25 through 40, and all students over 40.

Partitioning a Set Two Ways PARTITION 1

Assume: D is the set of all students in this section of CMIS 160 M is the set of all male students in D F is the set of all female students in D

Then {M,F} is a partition of D

PARTITION 2

Assume: D is the set of all students in this section of CMIS 160 A is the set of all students in D whose age is less than 25 B is the set of all students in D whose age is greater than or equal to 25 and less than or equal to 40 C is the set of all students in D whose age is greater than 40

Then {A,B,C} is a partition of D

How can you apply the idea of set partition to the process of putting together a jigsaw puzzle?

Note that the elements of a power set are all sets, including the empty set and the set itself. You should be able to list all elements in the power set of a given set and know how many elements there are, but you will not be expected to prove statements about power sets.

Read Epp: pp. 239-242 and 588-589 Formal Languages

The material on these pages uses the basic notions of sets, so you may need to review the definitions of set theory in module 1 and on pages 231-236 of Epp. Using the English language as an analogy might help you to grasp the concepts and techniques related to formal languages. The English language consists of 26 characters (letters) and words made up of the characters. The rules for forming words are complex and far too numerous to list here. The 26 characters form the alphabet of the language and the words are strings of characters in the alphabet. More formally, let the alphabet be

S = {a, b, c, d, ..., y, z},

where S is the Greek letter sigma. Then strings are ordered n-tuples of characters of S (where n is a finite, positive integer) that follow the rules for constructing English words. The English language over S is the set of strings over S. Note that the language could be infinite even though the alphabet is finite and each string is of finite length.

In formal languages, there is no guarantee that the strings over an alphabet make any sense. You can specify an alphabet any way you want, as long as the number of distinct characters is finite. You can impose any rules you want to limit the strings that are in the language over the alphabet. The set of strings that results from the alphabet and rules specifications becomes the formal language over the alphabet.

Study the summary of definitions and the notation in the shaded boxes on page 240. Work through the three examples on pages 240 and 241, but omit Algorithm 5.1.1 and example 5.1.14. Try exercises 19 and 20 on page 243 to assess your ability to apply the definitions.

In module 3, we indicated that the material on lexicographic order would be revisited in module 5. The discussion of formal languages we have just completed, combined with your understanding of partial order relations from module 3, should prepare you for the theorem, definition, and example of lexicographic order on pages 588 and 589.

If the notation in the first two parts of the theorem seems confusing, consider a different pair of words for each part. For example, for part 1, consider the two words compute and computer. Let the a's in the theorem refer to the characters in compute, and the b's refer to the characters in computer. Then m = 7 (the number of characters in compute) and n = 8 (the number of characters in computer). Note that the first m characters are the same in both words.

Applying the theorem gives the result that compute precedes computer in lexicographic order. For part 2 of the theorem, try the two words computer and computing. After working through example 10.5.6, try exercise 11 on page 600 to assess your understanding.

***************************************************************

Chapter 7

Functions

A. Basic Definitions Read Epp: pp. 344 and 345

The definition in the shaded box on page 345 is critical to an understanding of functions. Keep it handy as you work through this chapter. Although the definition includes all functions, in discrete mathematics it is applied to functions on sets--more specifically, sets with a finite or countable number of elements. The examples and exercises that pertain to sets with an infinite, noncountable number of elements will not be included in this course.

B. Function Representations Read Epp: pp. 345-354

Several convenient ways to visualize a function include an arrow diagram, function machine, and table. The arrow diagram works well for sets with very few elements. As examples 7.1.1 and 7.1.2 show, you can readily identify whether a function or nonfunction is represented.

The function machine might appeal to your interest in computing. The function machine requires that you are given some input, then you perform the operations specified, and that produces output. This input/output may imply that a computer program can be considered a function. Can all programs be considered functions? If your answer is no, then you should be able to describe conditions for which a given program is not a function. Of course, you could try to find a counterexample as you did in the section on sets.

A table is another way to specify a function. A table was described in the Application Context for this module, and an example of a function was included.

Examples 7.1.10 and 7.1.11 give insights into some interesting applications in computing. Note that Figure 7.1.6 on page 352 shows a table and an arrow diagram for the same function. These examples, and example 7.1.13, will not be pursued further in this course.

Try exercises 1-5, 12, 20, 21, 25, and 26 on pages 354-356 to assess your understanding of the topics in Section 7.1.

C. Function Properties Read Epp: pp. 369-384

Read each definition a number of times, and try to provide your own examples with the help of an arrow diagram. Work through Epp's examples for one-to-one, onto, and inverse functions, as well as one-to-one correspondence (bijection). (Omit the hash function application on pages 373 and 374 for now. Hashing will be referred to in part III, section E of the Commentary.)

Take your time with this material; it can be confusing. For your own interest, you may want to look briefly at the examples involving infinite sets, but concentrate on the finite set examples. Then try exercises 1b, 1c, 2-6, 31, and 32 on pages 384-387.

Section 7.4 on the pigeonhole principle will not be covered in this course. Essentially, the principle states that if you try to put objects into containers and you have more objects than containers, then at least one container will get more then one object. Yes, it is as simple as it sounds, but the principle has a number of fascinating and unexpected applications. You can get some idea of this by browsing through the examples and exercises, if you wish. B. Finite-State Automata Read Epp: pp. 357-369

Review the section on combinational circuits, and then note the difference between combinational and sequential circuits indicated on page 357. The work on finite-state automata you are about to undertake has direct application to these types of circuits, hence to computers themselves.

On page 358, Epp provides some historical background appropriate to the topic of finite-state automata. Although mathematical in origin and intent, the topic has had a great impact in computing. For a more detailed discussion of the work of Alan Turing and his Turing machine, see Gersting's book, pages 597-609. Additional material on pages 610-617 takes you into the realm of theoretical computer science, but it gives you more background on why we study formal languages and finite-state automata in computing.

The material in Gersting also includes a very interesting result: there are unsolvable problems in the sense that an algorithm cannot be specified to solve the problem (or halt) after a finite number of steps. All of this material in Gersting is for your own information and will not be tested in this course. (Information about Gersting's book is given in the syllabus under "Course Materials.")

Work through example 7.2.1, then try exercise 1 on page 367 to test your understanding. Study the definition of a finite-state automaton on page 360 and the concepts of a (state-) transition diagram and next-state table. Examples 7.2.2 and 7.2.3 illustrate the ideas and definitions. Exercises 2-13 on pages 367-369 pertain to these topics. Work enough of them to convince yourself that you understand the processes before approaching the next topics.

Pages 362-367 introduce four issues. We will study two of them:

1. given a finite-state automaton, what language does it accept

2. given a language, what finite-state automaton accepts it

The other two issues are concerned with (1) finding the eventual state of an automaton that is already in a state when an input string is given and (2) writing a program to simulate a specific finite-state automaton. They will not be included in this course.

Study the definitions in the box on page 362 and work through example 7.2.4. The definitions specify what it takes for a string and a language to be accepted by a finite-state automaton. The example includes only one accepting state shown by the double circle. In general, a finite-state automaton can have more than one accepting state. Exercises 14-19 on page 369 will test your understanding of the process of finding a language accepted by a finite-state automaton.

Examples 7.2.6 and 7.2.7 show how to design a finite-state automaton given a specific language. Carefully work through these examples. Then try exercises 20 and 22-29 on page 369 to assess your ability to design a finite-state automaton for a specified language. Note that exercise 30 suggests a practical application of the theory.

D. Finite-State Automata Read Epp: pp. 357-369

Review the section on combinational circuits, and then note the difference between combinational and sequential circuits indicated on page 357. The work on finite-state automata you are about to undertake has direct application to these types of circuits, hence to computers themselves.

On page 358, Epp provides some historical background appropriate to the topic of finite-state automata. Although mathematical in origin and intent, the topic has had a great impact in computing. For a more detailed discussion of the work of Alan Turing and his Turing machine, see Gersting's book, pages 597-609. Additional material on pages 610-617 takes you into the realm of theoretical computer science, but it gives you more background on why we study formal languages and finite-state automata in computing.

The material in Gersting also includes a very interesting result: there are unsolvable problems in the sense that an algorithm cannot be specified to solve the problem (or halt) after a finite number of steps. All of this material in Gersting is for your own information and will not be tested in this course. (Information about Gersting's book is given in the syllabus under "Course Materials.")

Work through example 7.2.1, then try exercise 1 on page 367 to test your understanding. Study the definition of a finite-state automaton on page 360 and the concepts of a (state-) transition diagram and next-state table. Examples 7.2.2 and 7.2.3 illustrate the ideas and definitions. Exercises 2-13 on pages 367-369 pertain to these topics. Work enough of them to convince yourself that you understand the processes before approaching the next topics.

Pages 362-367 introduce four issues. We will study two of them:

1. given a finite-state automaton, what language does it accept

2. given a language, what finite-state automaton accepts it

The other two issues are concerned with (1) finding the eventual state of an automaton that is already in a state when an input string is given and (2) writing a program to simulate a specific finite-state automaton. They will not be included in this course.

Study the definitions in the box on page 362 and work through example 7.2.4. The definitions specify what it takes for a string and a language to be accepted by a finite-state automaton. The example includes only one accepting state shown by the double circle. In general, a finite-state automaton can have more than one accepting state. Exercises 14-19 on page 369 will test your understanding of the process of finding a language accepted by a finite-state automaton.

Examples 7.2.6 and 7.2.7 show how to design a finite-state automaton given a specific language. Carefully work through these examples. Then try exercises 20 and 22-29 on page 369 to assess your ability to design a finite-state automaton for a specified language. Note that exercise 30 suggests a practical application of the theory.

E. Function Composition Read Epp: pp. 401-410

Composition of functions extends the idea of functions. The notation might be difficult, but the basic concept is relatively simple: my sending a message to person A, who adds a note and forwards the composite message to person B, produces the same result (from person B's point of view) as if I had sent the composite message to person B directly.

Work through the definitions and theorems, but not the proofs. Then try exercises 1, 2, and 20 on pages 410 and 411. The application of composite functions to tables in a relational database is straightforward.

***************************************************************

Chapter 8

Recursion Read Epp: pp. 424-435

A. Recursively Defined Sequences

Before starting this section, read the material on sequences on pages 180-182, including examples 4.1.1 and 4.1.2. This material provides some basic information about sequences and gives examples of two of the three ways to define a sequence, as indicated on pages 424 and 425. The third way to define a sequence, using recursion, is the topic of this section. Note the paragraph on page 425 that compares proving theorems by mathematical induction to defining sequences by recursion.

Study the definitions on page 425. Initial conditions play an essential role in recursion. They provide the starting point for determining the rest of the values of a sequence. Work through examples 8.1.1-8.1.3, then try exercises 1-8 on page 438 to test your ability to find terms of a recursively defined sequence.

Example 8.1.4 involves a proof process for a recurrence relation. The object is to show

Sk = -kSk-1

As you work through the proof, note that the process hinges on multiplying both sides of the line numbered 8.1.2 by the quantity -k. With this multiplication and some algebraic manipulation, we can show Sk = -kSk-1. Try exercises 9-11 on page 438 to see if you can follow the process illustrated in the example.

Three common examples of the use of recursively defined sequences are the towers of hanoi, Fibonacci numbers, and compound interest. Read pages 427-433, working through examples 8.1.5-8.1.7. Try exercises 17, 18, 22, 23, 27, and 29 on pages 439 and 440 to practice applying the concepts and techniques related to recursively defined sequences. The additional examples on pages 433-438 will not be included in this course, nor will the extensions of the topic on pages 441-471. Our only purpose is to introduce you to the methodology for handling recursively defined sequences.

B. Recursive Functions Read Epp: pp. 471-473

These pages introduce another aspect of recursion, the notion of functions that are defined recursively. Work through the three examples that are given. Examples 8.4.7 and 8.4.8 are well known in the area of theory of computing. Example 8.4.8 deals with one of the difficulties in the area of recursive functions, determining whether they are well defined. Try exercises 19, 21, and 23 on page 475 to get an idea of how values of recursive functions are computed. Note how quickly the complexity of the calculations increases when higher values of the functions are sought.

***************************************************************

Chapter 10

Relations

A. Basic Definitions Read Epp: pp. 533-536

In chapter 7, you were introduced to a special kind of relation, a function. There, each element a of a set A was related to one and only one element b of a set B. In other notation, the pair (a,b) is an element of the set AxB. A relation can be more general, however. It allows an element of A to be related to more than one element of B. That is, the pairs (a,b), (a,c), and (a,d) could all be a relation (but not a function), provided of course, that b, c, and d were all elements of B.

The elements depicting a relation are ordered; the first element is in A and the second element is in B. If the pairs were written (b,a), (c,a), and (d,a), then they would be elements of the set BxA, where the elements of B are related to elements of A. Pairs signify a binary relation, which is the type of relation that appears in this chapter.

Some examples of relations were identified in the database application in this module. As another example, consider the set B of book names and the set I of ISBNs. The binary relation of book names to ISBNs is a function because each book name has a unique ISBN. The same is true of the relation of book names to publisher names. However, if the binary relation of publisher names to book names were a function, then the publishers would soon be out of business. Why?

Study section 10.1 in the same way you studied Section 7.1, again concentrating on finite sets. Try Exercises 1, 2, 9, 12, 13, 17, and 18 on pages 544 and 545.

B. Relation Representations Read Epp: pp. 537-542

Binary relations are commonly represented by arrow diagrams and tables. If A = {0, 2, 3, 5} and B = {1, 4, 7}, then the representation R = {(0,1), (0, 7), (2,4), (3,7), (5,1)} identifies the same relation as the table: A 0 0 2 3 5 B 1 7 4 7 1

and the arrow diagram below. Click "Relation" to see how the relation is formed.

What is the inverse of this relation? Click "Inverse" to see how the inverse of this relation is formed.

The definition on page 543 extends the notion of a binary relation to an n-ary relation. Example 10.1.11 illustrates the concept.

C. Relations as Directed Graphs Read Epp: pp. 542-544

When a relation is defined from a set to itself, the corresponding arrow diagram is also called a directed graph. Suppose we change the previous example to put all elementsq of both sets into just one set, A, but keep R the same. Then A = {0, 1, 2, 3, 4, 5, 7} and the arrow diagram is now a directed graph.

A directed graph can have arrows from an element to itself (if, for example, (3,3) were in the relation) and from an element to another element then back to the first element (if, for example, both (2,3) and (3,2) were in the relation). Directed Graph

Directed graphs are helpful representations in understanding the properties defined in the next section, and they will be discussed in more detail in module 2.

D. Relation Properties Read Epp: pp. 546-553

Three important properties of relations--reflexive, symmetric, and transitive--are introduced in section 10.2. As an example, consider the set of people in a movie theater and the relation defined as "in the same row as." This relation is reflexive because each person is in the same row as him- or herself. The relation is symmetric because, for each person in the same row as a second person, that second person is in the same row as the first person. Finally, the relation is transitive because for each person in the same row as a second person who is also in the same row as a third person, the first person is in the same row as the third person.

Showing that a relation is reflexive, symmetric, and transitive is necessary to establish an equivalence relationship, the topic of the next section. Working through the examples and exercises 1-8, 15, 16, 18, and 37 on pages 554 and 555 is essential to your grasping the three properties of relations.

As before, work only with finite sets--with one exception, the relation of congruence modulo n on pages 552 and 553. Before studying example 10.2.5, you might need to refer back to examples 10.1.1 and 10.1.2 on page 535. The concept of congruence modulo n is particularly useful in computing. Computer programs that incorporate methods such as hashing and encryption use congruence modulo p, where p is a prime number.

E. Partitions, Equivalence Classes, and Equivalence Relations Read Epp: pp. 555-570

In this section, you may omit the proofs. The terms partition, equivalence classes, and equivalence relation are interrelated. Working exercises 1-5, 10, 11, 14, 16, 18, and 23 on pages 570 and 571 will give you a grasp of the three concepts.

An equivalence relation on a set partitions the set into equivalence classes. Understanding this statement gives you an immediate grasp of the computing method known as hashing, an alternative to sorting when you need to store items in a particular order.

Hash functions are initially discussed on pages 373 and 374 as a computer application for the topics in Epp's chapter 5. If you are familiar with the notion of hash functions through other computing experience then you should read this material. Otherwise, it might not mean anything to you at this time. When you do encounter hash functions in a different computer course, you will understand them better because of your study of functions. Hashing involves partitioning a set of items into equivalence classes.

If you have not encountered the term (or the idea of) hash functions, and you are interested in finding out more about them, look in a programming text that covers the second semester of a two-semester course sequence.

Partial Order Relations

A. Basic Definitions Read Epp: pp. 585-588

Compare the definition of antisymmetric to that of symmetric presented in the previous module (Epp, page 547). Antisymmetric is a stronger (more restrictive) property; it requires the extra condition that a and b be the same element whenever a is in relation to b and b is in relation to a.

Arrow diagrams clearly distinguish the two concepts. Any relation that has in its arrow-diagram representation an arrow from a to b and an arrow from b to a cannot be antisymmetric. Example 10.5.1 illustrates this statement, and exercises 1, 2, and 4 on pages 599 and 600 will test your understanding.

Note that in example 10.5.2 actual numbers (elements of the set) are used to provide a counterexample. (Remember that one counterexample is all that is needed to disprove a statement.) Using specific elements in a set is a good technique; a concrete example makes it easier to come up with a general solution.

Also compare the definitions of partial order relation and equivalence relation (page 559). Use actual sets and numbers when working through examples 10.5.3-10.5.5 to get some idea of what to expect in a partial order relation. Then try exercise 6 on page 600. Also, go back to exercises 1, 2, and 4 to determine whether each relation is a partial order relation.

For now, omit the section on lexicographic order (pages 588 and 589) and the corresponding exercises on pages 599 and 600. We will return to this topic in module 5.

B. Hasse Diagrams Read Epp: pp. 590-592

A Hasse diagram is a means of graphically representing a partial order relation. It is (1) a graph and (2) the simplest representation from which all relations in the partial order can be derived.

To construct a Hasse diagram, given a partial order relation, first construct its arrow diagram with all arrows pointing upward. Then, systematically remove edges that are unnecessary because of the reflexive and transitive properties of the relation.

Look at our example illustrating how a Hasse diagram is constructed.

If you add times for completing jobs to the edges, the Hasse diagram becomes a PERT chart controlling job scheduling (recall the Application Context).

Try to construct the diagram without first doing the digraph. Also, reverse the steps to get the digraph from the Hasse diagram. Examples 10.5.7 and 10.5.8 illustrate this process. Try exercises 16, 17, 19, and 21b on page 600.

C. Sets that are Partially or Totally Ordered Read Epp: pp. 592-594

In the previous example, elements such as 3 and 12 are called comparable and elements such as 2 and 3 are noncomparable, according to the definitions on page 592. The set A in the previous example is a partially ordered set (poset) with respect to the relation "divides," because "divides" is a partially ordered relation on A. However, A is not a totally ordered set, because it contains noncomparable elements. A set such as {1, 2, 4, 8, 16} is a totally ordered set. Note that its Hasse diagram is the single chain

1 ------ 2 ------ 4 ------ 8 ------ 16

All totally ordered sets have this kind of Hasse diagram.

Study the definitions on page 594 and work through example 10.5.10. Note that a partially ordered set can have more than one maximal or minimal element, but at most one greatest or least element. The existence of a minimal element in a finite, nonempty, partially ordered set is the basis for the next topic, topological sorting. Try exercises 22-29, 34, 39, and 40 on pages 600 and 601.

D. Topological Sorting and Applications Read Epp: pp. 596-599

Note that the discussion of topological sorting is restricted to finite, nonempty, partially ordered sets. A topological sorting on such a set produces a listing of the set's elements in the form of a chain; that is, it creates a total order on the set. There can be more than one topological sorting of a given set. Work through example 10.5.11 and the first application on pages 595-597, then try exercises 42, 43, and 46a on page 601.

Reread the Application Context. The scenario is a job-scheduling problem similar to the application in example 10.5.12. The PERT chart in the example is a Hasse diagram. The chart portrays the relations among components. When weights (time, in the example) are added to the edges, then a critical path can be determined. This path indicates the minimum amount of time required to complete the entire job. Try exercises 46b and 47 on page 601 to test your understanding of the application in example 10.5.12.

***************************************************************

Chapter 11

Graphs

A. Basic Definitions and Terminology Read Epp: pp. 602-607

The blue-shaded box on page 603 contains a number of definitions and terms that are essential to the study of graphs. Refer to the box frequently until you are clear about the meaning of each term. The notion of an empty graph might seem strange, but it is a useful one when studying algorithms on graphs.

A standard way to illustrate the definition of a graph is to use a table of edges and endpoints, as in qexample 11.1.1. A table can show that two different drawings of a graph either represent different graphs or the same graph (as illustrated in examples 11.1.2 and 11.1.3). Do exercises 1-7 on pages 616 and 617 to test your understanding of some of the basic graph definitions.

Note the difference between a directed graph (digraph) as defined on page 607, and a graph. In a digraph, ordered pairs of vertices indicate both the edges and the directions of edqges. In a graph, there are no directions from one vertex to another, so a pair of vertices (irrespective of order) represents an edge, but in a digraph, order is important. For example, if a, b, c, d represent vertices, then (a,b) and (b,a) both represent the same edge in a graph, but represent different edges in a digraph.

B. Applications Read Epp: pp. 607-609

Read examples 11.1.4-11.1.6 carefully. They illustrate applications of graphs. To test your understanding of example 11.1.5, answer the following questions: Is Poetry Magazine a suburban weekly? Does it contain printed writing? What kind of paper finish does it have? Also try exercises 10 and 11 on page 617.

C. Other Types of Graphs Read Epp: pp. 609-612

The definitions and examples on pages 609-612 identify additional types of graphs. Note the similarity between the concepts of subgraphs and subsets. Try exercises 24 and 36-39 on pages 617-619 to test your understanding of the definitions in this section. Note that a few of the exercises also contain definitions.

Paths and Circuits

A. Definitions Read Epp: pp. 602-604, 612, 619-625

Review the information on pages 602-604, and be sure you understand the definition of degree, given on page 612.

The Bridges of Konigsberg problem in Section 11.2 exemplifies a useful aspect of mathematics: that of changing the representation of a problem either to apply known results about the new representation or to understand the problem better. The solution to the Konigsberg problem should become clear after you study the material in this section.

A number of new terms are defined: walk, closed walk, path, simple path, circuit, simple circuit, connected vertices and graphs, and connected component of a graph. Try to get an intuitive idea about each term by thinking about it without reading the definition.

For example, consider the possible differences between taking a meandering walk and following a path. Then think of the differences between ending up where you started or at another destination. Work through the examples given in this section of Epp and try exercises 1, 2, 4, and 5 on page 636 to get a more concrete understanding of each term's meaning. The table on page 622 may help you remember characteristics and differences.

Item a. in Lemma 11.2.1 on page 625 presents an important characteristic of trees. After you do the examples of connected graphs and components of graphs, try exercises 6-8 on pages 636 and 637. The concept of a bridge, as defined in exercise 6, can be applied to networking.

By the time you finish all of the exercises and examples, you should realize that the definitions of the terms in this section are little more than formalized expressions of intuitive notions about the terms. In other words, you should have little or no difficulty in explaining to someone else what each term means, even though you might not state the exact definition. If you cannot do this, then review the definitions and rework the exercises and examples, or try similar exercises in a different discrete mathematics text.

B. Euler and Hamiltonian Circuits Read Epp: pp. 625-635

The topics in this part of section 11.2 are a little more complicated than the preceding topics. Study the definitions of Euler and Hamiltonian circuits. Note their differences: an Euler circuit can visit the same vertex more than once and must include all edges, whereas a Hamiltonian circuit must include every vertex exactly once and might not include all edges.

You may omit the proofs, but take a common sense approach to understanding the theorems and the corollary in this section of Epp. For example, consider Theorem 11.2.2 on page 626. For an Euler circuit to exist, each vertex must have the characteristic that any time you arrive at the vertex along an edge, you must also depart the vertex by a different edge (otherwise you would traverse the same edge more than once). This means that each vertex has degree 2 if it is counted only once in the circuit, degree 4 if counted twice, and so on. In short, each vertex must have even degree.

Apply the same type of reasoning to the contrapositive version of Theorem 11.2.2 on page 627. If any vertex of a graph has odd degree, then either you arrive at that vertex along an edge and do not depart, or you depart, but do not arrive. In either case, you cannot complete a circuit by traversing all edges, hence there can be no Euler circuit in the graph.

After working through the examples on pages 627-632, try exercises 12-22 on pages 637 and 638 to test your understanding of Euler paths and circuits. Work through example 11.2.8, then try exercises 23-31 on pages 638 and 639 to assess your understanding of Hamiltonian circuits. Finally, do example 11.2.9, then test your ability to solve a traveling salesman problem by completing exercise 36 on page 639.

D. Matrix Representations Read Epp: pp. 640-656

Any experience you might have had with two-dimensional arrays, or matrices, in a programming language should help you with the material on pages 640-656. In particular, the representation and storage of matrices, and operations on them, play an important role in working with graphs. This is another instance of the utility of knowing about one set of concepts--in this case, aspects of matrices--as an aid in understanding other concepts, properties of graphs.

As indicated on pages 644-646, the adjacency matrix of a directed or undirected graph tells quite a bit about the graph, especially if the adjacency matrix is symmetric. By examining the result of multiplying an adjacency matrix by itself, you can determine the number of connected edges there are between certain pairs of vertices in a graph. For example, if a graph represents an airline routing among numerous cities, then the graph's adjacency matrix and powers of the matrix would enable you to determine the number of stops required to get from any one city to any other. Adjacency Matrix Example Let A, B, C, D be four cities served by an airline as shown in this digraph: The corresponding adjacency matrix M is: M A B C D A B C D

In the matrix, vertices of the digraph are shown as row and column headers. The matrix shows that there is exactly one nonstop flight from A to B (a 1 in Row A, column B), from D to C (a 1 in row D, column C), and so on. Confirm this information by looking at the digraph.

Let's look at M2 . M2=M·M= M · M = M2 A B C D A B C D

M2 gives information about paths of length 2, which require exactly one stopover as you proceed from origin to destination. (You may need to review how to multiply M by M to get M2.)

The entries in M2 provide information such as the following:

There is one path from A to B that requires exactly one stopover (A to C to B). (The 1 in row A, column B indicates one path of length 2 from A to B.)

There is one path from A to D that requires exactly one stopover.

There are no paths from A to C requiring exactly one stopover.

There are two paths from D to B that require exactly one stopover.

To continue the example, consider M3 = M · M2 M3=M·M2= M · M2 = M3 A B C D A B C D

M3 gives information about paths of length 3, which require exactly two stopovers as you proceed from origin to destination. The matrix M3 shows:

There is one path of length 3 from A back to A, from A to C, and from A to D.

There are 2 paths of length 3 from B back to B, and so on.

Try exercises 2-7, 19, and 22 on pages 654 and 655 to test your understanding of the definitions in this section. If you have little or no experience with matrix operations, then try exercises 1, 8-10, and 12. II. Trees

A. Basic Definitions Read Epp: pp. 664 and 665

A tree is a graph that has a number of restrictions, as given in the definition on page 664. The definition includes the word circuit, a term that will be discussed in detail in module 3. You can find its definition on page 622.

For now, you need to know that a circuit (in a graph sense, not in an electronic sense) is a sequence of vertices and edges that begin and end at the same vertex without repeating an edge. For example, in the graph below, the sequence of edges from B to C to E to D to B is a circuit

but the sequence (below) from A to C to E to D is not a circuit.

The sequence of edges from C to B to D to E to B to C in the graph below also is not a circuit. Why?

B. Examples Read Epp: pp. 665-668

Decision trees are helpful structures in determining outcomes of certain kinds of games and as problem-solving tools.

For example, you can construct a tree depicting all possible moves in the game of Tic-Tac-Toe. Start with an empty diagram. Draw edges to diagrams containing each possible first insertion of an X (or O). From each of those diagrams, draw edges to each possible insertion of an O (or X). Continue this process until the entire tree is constructed.

You can reduce your effort considerably by noting, for example, that an initial insertion into one corner has the same effect as an insertion into any other corner. Thus, you need to show only one, not four diagrams for that insertion. Try this yourself, then look at our partial solution to the TIC-TAC-TOE example.

How would you use the information shown in the completed tree to play the game? Could you construct a tree for the game of chess?

Syntactic derivation, or parse trees, not only have applications in natural languages, but in computing as well. Numerous compilers used tree structures to store a computer program's identifiers and to verify proper syntax. A tree, rather than a graph, is an appropriate structure for syntactic verification because there is a unique path to a vertex in a tree. This prevents ambiguity, i.e., more than one interpretation.

Try exercises 1 and 2 on page 681 to assess your understanding of the types of trees presented in this section.

C. Terminology Read Epp: pp. 668-676

For this course, you should become familiar with the terms defined on pages 669 and 675, as well as Theorem 11.5.2 on page 670 (but not the proof). Theorem 11.5.4 on page 674 specifies a condition for a graph to be a tree. You can read the rest of the material if you are interested, but we will not use it in this course. On pages 681 and 682, exercises 7, 32, and 33 will help you develop familiarity with the relevant terms and concepts.

D. Binary Trees Read Epp: pp. 676-681

A binary tree is a key data structure in computing. For example, one application of binary trees involves sorting and searching.

Learn the definitions on page 676 and the theorems on pages 678 and 679 (but not the proofs). Try exercises 34-50 on pages 682 and 683 to test your understanding. III. Spanning Trees

A. Basic Concepts Read Epp: pp. 683-685

Learn the definitions and the proposition (but not the proof) on pages 684 and 685. They are essential to following the algorithms that come next. Work through example 11.6.1 on these pages, then try exercises 1-4 on page 692.

B. Finding Minimal Spanning Trees Read Epp: pp. 686-692

Spend time working through the Kruskal and Prim algorithms on pages 686 and 689. Study the examples that come immediately after the algorithms. Each step of Kruskal's algorithm, as shown in our example, looks at all remaining edges of the graph (those edges that have not been deleted in previous steps), then adds the one that has the least weight, but does not create a circuit.

Each step of Prim's algorithm, as shown in our example, looks only at adjacent edges before adding an edge.

Note that the theorems on pages 688 and 691 guarantee that a minimal spanning tree results from applying either algorithm. You can omit the proofs of the theorems. Carefully work through example 11.6.4, then try exercises 5-11 on pages 692 and 693.