11  Projects

The projects on the following pages are intended to guide you in the exploration of various topics in mathematics. While each project provides steps that guide you toward proofs of the theorems in each section, these projects can be quite demanding. At the start of each project we indicate the background that is needed in order to tackle the topic being presented.

11.1.  The Pythagorean Theorem

The Pythagorean Theorem is often referred to as the “first great theorem,” so it is very suitable to examine it, even if you have only read Chapter 1. We begin by presenting one proof of the theorem, and then asking you to write up additional proofs and consider ways to extend the result.

THEOREM 11.1.   Let a and b be the lengths of the two shortest sides of a triangle image, and let c be the length of the longest side. Then image is a right triangle if and only if

image

The history of this theorem is long and convoluted. It may or may not have been proven by Pythagoras of Samos (circa 500 BCE); it is a highlight of the first book of Euclid’s Elements; and it certainly was known in ancient China. In fact, Figure 1 appears in Arithmetic Classic of the Gnomon and the Circular Paths of Heaven (China, circa 600 BCE). The accompanying argument is written using the specific values of 3, 4, and 5, but it is easy to use modern notation and convert this ancient argument-by-example into what would be accepted as a modern mathematical proof.

image

Figure 1. The illustration accompanying an ancient Chinese proof of the Pythagorean Theorem.

A PROOF OF THE PYTHAGOREAN THEOREM.   Let image be a right triangle, with sides of length a, b, and c, where for convenience1 we have image. Then you can arrange four copies of image as in Figure 1, where the right angles are in the interior of the figure, and the four hypotenuses form a bounding quadrilateral. Call this quadrilateral image, since it has imageour sides.

Because the sum of the interior angles in any triangle is 180image, the sum of the two non-right angles of image is 90image. At each corner of the boundary quadrilateral image there is a copy of each of these non-right interior angles, hence at each corner of image there is a right angle. Thus image is a rectangle. Furthermore, each side of image comes from the hypotenuse of a copy of image, so image is a rectangle with four congruent sides, hence image is a square. The area of image is then image. Thinking of image as being composed of four triangles and a smaller interior square, we know the area of image can also be expressed as image. This implies

image

Notice, however, that we have not yet established the full Pythagorean Theorem. This argument shows only half of it: that the lengths of the three sides of a right triangle have a certain arithmetic property. The other fact claimed in the Pythagorean Theorem, the other half of the “if and only if” statement, is equally important: triangles with this arithmetic property are right triangles. This fact was used by the ancient Egyptians in surveying and construction, and it is still used by carpenters to construct right angles.

To prove the other direction, assume we have a triangle image where image. Let image be a right triangle with sides a and b and hypotenuse h. By the argument above we know image, so image. Thus image and image are congruent, and therefore image is a right triangle.

image

You do not need to write multiple proofs to establish the validity of a mathematical theorem. But multiple proofs can provide additional insights, and there are many fun and insightful proofs of the Pythagorean Theorem. In the following example and exercise, we are proving only that the sides of right triangles satisfy the arithmetic condition.

President James Garfield constructed a proof of the Pythagorean Theorem, long before he became President of the United States. His argument begins by forming a trapezoid using two copies of a right triangle, arranged so that side a of one copy and side b of the other copy lie on a line, as in Figure 2.

image

Figure 2. The trapezoid used in President Garfield’s proof of the Pythagorean Theorem.

Exercise 11.1   Verify that Garfield’s trapezoid is composed of three right triangles. Then establish the arithmetic identity image by comparing the area formula for a trapezoid with the sum of the areas of the three triangles that comprise Garfield’s trapezoid.

Exercise 11.2   Figure 3 is commonly used in proving the Pythagorean Theorem. Use this figure as inspiration to construct yet another proof of the Pythagorean Theorem.

image

Figure 3. Another arrangement that leads to a proof of the Pythagorean Theorem.

Exercise 11.3   Find one additional proof of the Pythagorean Theorem, so that you will have seen at least four proofs of this result. You can try to find this proof on your own, but we are really only asking you to consult available resources and read a fourth proof. Which of these proofs do you think give the most insight into why the theorem is true? Explain!

Exercise 11.4   Let A, B, and C be the three vertices of a triangle Δ, where the side opposite to A is a, the side opposite to B is b, and the side opposite to C is c. If the angle at C is less than 90image, what can you conclude, if anything, about the two values image and image? Why? What if the angle at C is greater than 90image?

Exercise 11.5   Let image, and z be three positive real numbers, and let Δ be the tetrahedron in image formed by the points

image

Because one corner of Δ corresponds to a corner of an octant in image, Δ is a three-dimensional analogue of a right triangle. In fact, a Pythagorean-style formula holds, connecting the areas of the four faces of Δ. Let A, B, and C be the areas of the faces that touch the origin, and let D be the area of the triangle formed by the points X, Y, and Z. Prove that

image

11.2.  Chomp and the Divisor Game

Chomp is a game that is easy to describe and play, and which is amenable to mathematical analysis. You should be able to work on this project after having worked through the first two chapters of this book. One of the questions makes use of a result from Chapter 4, but it is not an essential question that is necessary for the successful completion of this project. Chomp appears in Exercise 3.86, which is a nice problem, but not one that is necessary for this project.

Chomp is a fun game for two players, who we’ll call Alice and Bob. They start with an image rectangular “cake,” such as the image cake shown in Figure 4.

image

Figure 4. Two plays into the game Chomp. The nine unshaded squares were eaten first; the four lightly shaded squares were eaten second.

The players alternate turns, with Alice going first. On a turn, the player takes a right-angled “bite” from the northeast: a square is removed along with all other squares to the north and east and anywhere in between these directions. In Figure 4 the darkest shading shows the cake remaining after Alice first takes a bite (no shading), followed by a bite by Bob (lighter shading).

The loser of this game is the person who is forced to take the last bite. In the mathematical study of two player strategy games, here’s an important question: If both players use optimal strategy, who wins?

Exercise 11.6   Play Chomp with a friend on boards of varying sizes, such as image, image, image, and image. Play several games of each size, and let each player be Alice half of the time. Make conjectures about which board sizes favor Alice, and which favor Bob. Are there certain boards that you think are guaranteed wins for Alice if she plays optimally? Guaranteed wins for Bob?

Exercise 11.7   Prove that Alice has a winning strategy for any image or image board, when image.

Exercise 11.8   For this exercise you should focus on the case of a image game of Chomp. Label each of the six squares with an “A” or a “B” depending on who has a winning strategy if Alice’s first bite is based at that square. We have begun this process in Figure 5. Verify that our two labels are correct, and then label the remaining four squares.

image

Figure 5. Beginning to label a image game of Chomp. The “A” indicates that if Alice bites this position, then she has a winning strategy. The “B” indicates that if Alice bites in this position, then Bob has a winning strategy.

Exercise 11.9   Prove that, for any image board, Alice has a winning strategy so long as image. One way to do this is via a proof by contradiction, starting with the assumption that Bob has a winning strategy, and then showing that Alice could always make use of that strategy herself.

Exercise 11.10   Here’s a related game called the Divisor Game, and you will need the Fundamental Theorem of Arithmetic (Theorem 4.11) in order to fully analyze this game. Alice and Bob start with a natural number n. They alternate turns in which a player calls out a natural number that divides n, with the following restriction: each number called out must divide n, but it can’t be a multiple of any number that has already been called out. The loser is the person forced to say 1.

Here’s an example game for image:

Alice starts by saying 10; then Bob says 2; then Alice says 15; then Bob says 3; then Alice says 5; and then Bob is forced to say 1. Alice wins!

Who wins for image? For image? For any n?

The game of Chomp can be generalized in many ways. One generalization is to remove the assumption that you start with a rectangular cake, so that you allow Alice and Bob to play Chomp on a cake of any shape made up of n image squares in the square lattice. An example of the sort of cake we are allowing is shown in Figure 6. As in the original version, a bite removes some remaining square and all other squares in the northeast quadrant from it, and the player forced to remove the last square loses. In Figure 6, Alice can win by eating the square adjacent to the southwest corner square on her first turn.

image

Figure 6. A “cake” made of seven squares.

Exercise 11.11   Consider the image case and the image case.

(a)   When image, Alice has two choices of which square to eat, and she does not want to eat a square that would remove both squares, as then Bob would win. By examining the possible configurations of two squares, show that Alice always has a choice that will leave one remaining square, meaning that Alice has a winning strategy.

(b)   Create a configuration for image where Alice has a winning strategy and a different configuration for image where Bob has a winning strategy.

In this more general situation we cannot give a simple statement of which player has a winning strategy. We do, however, know that one of them does have a winning strategy.

THEOREM 11.2.   For any image and any cake made up of n image squares in the square lattice, one of the two players has a winning strategy for the game of Chomp beginning with that configuration.

This is an example of a more general result about certain types of two-player games: you may not know who has the winning strategy, but you know that someone does!

Exercise 11.12   Here we outline a proof by induction for Theorem 11.2. Start by establishing the base case when image.

For the inductive step, assume that for any value of k such that image, someone has a winning strategy for each possible cake shape made up of k little image squares.

Now, consider Alice eating first from a cake made up of image little image squares. Since after any bite she makes there will be no more than n little image squares left, by the inductive assumption we know that either Alice or Bob has a winning strategy for the cake that remains after her first bite. Use this insight to prove that either Alice or Bob must have a winning strategy for the original cake.

GOING BEYOND THIS BOOK.   If you would like to know more about Chomp, or the mathematics underlying many other games, we recommend Winning Ways for Your Mathematical Plays by Berlekamp, Conway, and Guy [BCG01]. Over the years there have been different editions and volumes of this work – we have listed some in the bibliography – and they are all excellent.

11.3.  Arithmetic–Geometric Mean Inequality

The arithmetic–geometric mean inequality is a fundamental result in analysis. The two-variable version was stated in Exercise 1.1: If a and b are two non-negative real numbers, then

image

In this project we will guide you through a proof of this result as well as its natural generalization to any finite number of non-negative real numbers. The argument is a somewhat complicated application of mathematical induction, and so we recommend having worked though Chapters 1 and 2 before beginning this project.

There are a number of ways to compute an average for a given a set of positive numbers, image. The most common method is to add them up and divide:

image

This is the arithmetic mean of image. Another approach is to multiply and then take a root:

image

This is the geometric mean. We note that the arithmetic mean can easily be applied to a mix of positive and negative numbers, while the geometric mean cannot. Hence for the remainder of this project we will be limiting ourselves to non-negative real numbers.

Exercise 11.13   Gather five or more friends. Give each of them n randomly chosen, non-negative real numbers, where n is some integer greater than 2. Your friends’ job is to compute the arithmetic and geometric means of the numbers you gave them. We will wager that in all cases image. If you can’t find five friends willing to help you with these computations, perhaps you could have a computer assist you in collecting evidence.

You should have found that the arithmetic means were always at least as large as the geometric means of your random sets of real numbers chosen for Exercise 11.13, and that the arithmetic mean was usually strictly larger than the geometric mean. The arithmetic–geometric mean inequality says that the results of Exercise 11.13 were not random. If you didn’t get such results, some of your friends probably need a refresher course on arithmetic.

THEOREM 11.3.   For any n non-negative real numbers image,

image

(11.1)

Further, the inequality is strict unless image.

The second statement, that the inequality is strict, means that

image

whenever there is an image that is not equal to some image in image.

You will establish the arithmetic–geometric mean inequality via two lemmas.

LEMMA 11.4.   Inequality 11.1 holds for all image, image. Further, when image the inequality is strict whenever image for some image.

The proof of Lemma 11.4 is by induction on the exponent k.

Exercise 11.14   Begin by establishing the base case, when image, of the inequality. Note that in this case image and image for some image. Explain why it suffices to establish

image

for all image. And then use algebra to show that this is equivalent to showing that

image

Since squares of real numbers are non-negative, the result follows.

Exercise 11.15   Explain how the second claim, that the inequality is usually strict, follows from the fact that when image, image.

For the inductive step, assume the inequality holds for image up to some image, and let image. We need to establish for any list of image non-negative real numbers, the inequality

image

holds.

Exercise 11.16   Define image and more generally, image. Since each image, our induction hypothesis says

image

Use the induction hypothesis and the fact that image to establish the inequality.

Exercise 11.17   Prove that the inequality will be strict should image for some image.

The second lemma in the proof of Theorem 11.3 works backwards, and can be used to fill in the gaps between image and image.

LEMMA 11.5.   If Inequality 11.1 holds for a natural number n, then it also holds for image.

Exercise 11.18   We want to show that the inequality holds for a list of non-negative numbers image to image, assuming it holds for lists with one additional number. Define

image

the arithmetic mean of the original list of numbers. Then start from the assumption that

image

to establish Inequality 11.1 for the original list of image non-negative real numbers.

Exercise 11.19   Verify that you continue to get a strict inequality as long as some image.

We can now establish the arithmetic–geometric mean inequality.

PROOF OF THEOREM 11.3.   Given any image there is a image such that image. Lemma 11.4 shows that Inequality 11.1 holds for lists of length image; Lemma 11.5 then shows that the inequality holds for lists of length n. See Figure 7.

image

image

Figure 7. The structure of the proof of the arithmetic–geometric mean inequality. The arrows pointing right are applications of Lemma 11.4, while the ones pointing left are Lemma 11.5.

GOING BEYOND THIS BOOK.   This proof of the arithmetic–geometric mean inequality is essentially due to Augustin Cauchy. Our presentation in this project is heavily inspired by An Introduction to Inequalities by Beckenbach and Bellman [BB61]. If this section strikes you as interesting, we strongly recommend you take a look at that book.

11.4.  Complex Numbers and the Gaussian Integers

This project introduces the complex numbers, image, and focuses on the Gaussian integers, image. Anyone who has worked through the first four chapters should be able to work on this project. We do use some of the language of functions, which is covered in Chapter 5, but our use of this language is minimal and we expect that it is not necessary to have completed Chapter 5 before taking up this project.

The complex numbers image are formed by adding image to the real numbers. Every element of image can be expressed as image, where image. Addition, subtraction, and multiplication in image are defined keeping in mind that image and image. Thus

image

The conjugate of a complex number changes the sign of the coefficient of image. That is, the conjugate of image is image.2 Conjugation helps explain division by non-zero complex numbers. In general, as long as image,

image

For example,

image

The complex numbers are usually represented by points in the plane, where image corresponds to the point image. The distance from the origin to image is image, and so the product of image with its conjugate image is the square of the distance to the origin. As we shall see, this norm

image

is a very useful function from image to image.

Exercise 11.20   The Gaussian integersare the subset of the complex numbers consisting of all integer combinations of 1 and image. That is,

image

Show that the Gaussian integers are closed under addition, subtraction, and multiplication.

Exercise 11.21   If image, we say α divides β if there is a image with image. As you might expect, we denote this image, and when α does not divide β we write image.

(a)   Show that image.

(b)   Show that the only divisors of image are image, and image.

(c)   Show that image, and conclude that image does not deserve to be called a prime number (in the context of Gaussian integers).

Exercise 11.22   Let image be the norm described above.

(a)   Prove that, for every α and image, image.

(b)   Show that the restriction of the norm to image produces a function from image to image.

(c)   Is there any image where image?

(d)   There are 12 Gaussian integers whose norm is 25. Find them, and plot them in the complex plane.

The Gaussian integers seem like a reasonable analog of the integers, just with the addition of the element image. It is reasonable to see if Gaussian integers have similar divisibility properties, including unique factorization.

As a cautionary example, we note that

image

which might temper your hopes for unique factorization. The situation might seem less severe if one were to notice:

image

Thus the two expressions given above for image are connected by

image

Define the units of image to be those elements image where image for some image.

Exercise 11.23   Use the norm to show that if α is a unit in image, then image. After establishing this lemma, verify that the units of image are image.

The norm can help in finding factorizations. For example, image. Since image, there is a chance that image where image and image. In fact,

image

Exercise 11.24   Use the norm to assist you with the following questions.

(a)   Find a factorization of image where neither factor is a unit.

(b)   Show that if image, then either α or β must be a unit.

(c)   Just to keep your enthusiasm in check, verify that image divides image but image in image.

We are now ready to define the notion of a prime number for the Gaussian integers.

DEFINITION 11.6.   An element image is prime if whenever image, then exactly one of α or β is a unit of image.

Exercise 11.25   Determine if the following Gaussian integers are prime or not. They are arranged in order of increasing norm. While there are quite a few of them, you will gain some critical familiarity with image by checking them all.

(a)   image

(b)   2

(c)   image

(d)   image

(e)   3

(f)   image

(g)   image

And, given your work above, you should now be able to do the following quite quickly.

(h)   Prove that if image is a prime number in image, then α is a prime in image.

(i)   Show by example that there are integers that are prime in image which are not prime in image.

Exercise 11.26   Prove that every Gaussian integer is a Gaussian prime or can be expressed as a product of Gaussian primes. We recommend a proof by strong induction where you induct on the norm of the Gaussian integer.

Exercise 11.27   At this point we have finished guiding you! Review the proof of unique factorization of the integers, and following that general approach, you should now be able to state and prove an analogous result for the Gaussian integers. We aren’t saying this will be easy; we are merely claiming that we have great confidence in your mathematical abilities.

GOING BEYOND THIS BOOK.   If you would like to know more about the Gaussian integers and similar number systems, you should read On Quaternions and Octonions [CS03].

Exercise 1.36 asks you to prove that there are arbitrarily large gaps between consecutive primes in image, so it is impossible for someone to walk from 0 to infinity along the real line, where the steps are of bounded length and the person is only allowed to step on prime integers. A currently unsolved problem asks the same question in the context of the Gaussian integers, thought of as lattice points in the plane. Namely, is it possible for someone to walk from image to infinity when the steps are of bounded length and the person must only step on points image where image is a Gaussian prime? Figure 8 displays Gaussian primes near the origin. You can read more about this problem in [GWW98], [Loh07], and the references cited by these papers.

image

Figure 8. The Gaussian integers inside of the complex plane. The filled circles correspond to the Gaussian primes.

11.5.  Pigeons!

In this project we explore …

THE PIGEONHOLE PRINCIPLE (Lemma 5.21). Let image, where A and B are finite, non-empty sets. If image, then f is not injective.

We proved this result at the end of Section 5.5. Here we take up some of its consequences and ask you to state and prove a generalized Pigeonhole Principle. We expect that to be successful in this project you will need to have finished the first three chapters and Chapter 5, through the proof of the Pigeonhole Principle in Section 5.5. It will be even easier if in fact you have worked through all of Chapter 5.

11.5.1. Lots of Pigeons!   The Pigeonhole Principle can be generalized. For example, if you have 25 pigeons fly into 12 coops, it is certainly the case that at least one coop has more than one pigeon. In fact, it must be the case that at least one coop has more than two pigeons.

Exercise 11.28   In this exercise you will discover and prove a generalization of the Pigeonhole Principle.

(a)   Let A and B be two finite sets, with image. How large must A be to ensure that given any image there is some image where at least three elements of A are mapped to b?

(b)   Let A and B be two finite sets, with image. How large must A be to ensure that given any image there is some image where at least k elements of A are mapped to b?

(c)   State a conjectured generalization of the Pigeonhole Principle.

(d)   Prove your conjecture.

11.5.2. Pigeonhole Projects.   The exercises in this subsection are independent of each other, so you may pick and choose those you find most interesting. Their diversity illustrates the broad applicability of the Pigeonhole Principle, and each is written in a manner that requires you to do some creative exploration of the application. Our recommendation is that you write up a presentation and proof of your generalized Pigeonhole Principle, and illustrate this result by presenting your work on one of the following exercises.

Exercise 11.29   Define a rectangular game board to be an image rectangle, with image and both at least 2, which is divided into colored unit squares. A standard chessboard is one example; another example is shown in Figure 9.

(a)   A image board is divided into 21 unit squares, each colored black or white. Prove that there must be a rectangular game board contained in this board whose four corner squares all have the same color.

(b)   Find examples that illustrate that the claim in part (a) is not true on a image game board for any image.

(c)   Find another example that illustrates that the claim in part (a) is not always true for image game boards.

(d)   Consider image game boards where the squares can be colored by three different colors. Can you find values of m and n that guarantee there is a game board contained in the image board where all four corners have the same color?

(e)   Can you state and prove a result that works for c colors, where c is an integer that is greater than or equal to 2?

(f)   How might you frame questions like these for rectangular bricks which have been divided into unit cubes? What sort of results can you find in this context?

image

Figure 9. A image game board colored by two colors. The thick boundaries highlight a image game board, and a image game board, where in each case the four corners have the same coloring.

Exercise 11.30   The positive difference of a pair of distinct integers image is image. In this exercise you will consider subsets image and prove results about the positive differences of pairs chosen from S. You may assume that the number of pairs of elements contained in a set with k elements is image. More information about this and related formulas is presented in Section 3.8.

(a)   Consider subsets image. Prove that if image, then there are at least two pairs of elements in S with the same positive difference.

(b)   Give an example that shows the claim in (a) is false if image.

(c)   Consider subsets image. Find a value of k such that when image, then there are at least two pairs of elements in S with the same positive difference.

(d)   Conjecture a result along the lines of

image

Here you should give an explicit formula for f in terms of the variable n. And of course you should prove that your conjecture is correct.

(e)   How might you extend your ideas to the situation where S is sufficiently large as to guarantee that there are at least three pairs of elements in S with the same positive difference?

Exercise 11.31   A lattice point in image is any point image where x and y are both integers.

(a)   Prove that if five lattice points are chosen in the plane, then there must be a pair where the midpoint between them is itself a lattice point.

(b)   How many lattice points must you choose to guarantee that there are two pairs where the midpoint between each pair is itself a lattice point.

(c)   State and prove a result about the number of lattice points that will guarantee that there are n pairs where the midpoint between each pair is a lattice point.

(d)   Extend your ideas to lattice points in image.

Exercise 11.32   In this problem you will explore the clustering of points chosen inside of a unit square.

(a)   Prove that given any 9 points in the unit square, there must be three that are contained in a circle of radius image.

(b)   Prove that given any 19 points in the unit square, there must be three that are contained in a circle of radius image.

(c)   Prove that given any 28 points in the unit square, there must be four that are contained in a circle of radius image.

(d)   Conjecture a result along the lines of the previous parts of this exercise, and prove that your conjecture is correct.

(e)   Conjecture a result along the lines of

image

Here you should give an explicit formula for f in terms of the variable n. And of course you should prove that your conjecture is correct.

GOING BEYOND THIS BOOK.   Like any fundamental result in mathematics, the history of the Pigeonhole Principle is fascinating and complicated. If you are interested in this history, we recommend the article by Rittaud and Heeffer [RH14] as a good entry point to the literature.

11.6.  Mirsky’s Theorem

Mirsky’s Theorem is a general result about partially ordered sets, so you should have completed Section 6.2 before working on this project, including the definition and exercises on chains and maximal chains in the end-of-chapter exercises for Chapter 6.

Consider the following sequence of nine integers:

image

By focusing on particular entries, you can find subsequences that are increasing like the bold entries in

image

and subsequences that are decreasing like the bold entries in

image

Call such a subsequence a monotone subsequence. We designed our sequence of nine numbers so that the longest monotone subsequences have three terms.

Exercise 11.33   Verify that our sequence contains no monotone subsequences with four terms.

Something interesting happens when a tenth integer is added into this sequence. We randomly selected a new integer and inserted it into a random location in our original sequence to produce

image

This sequence now contains monotone subsequences with four terms, for example,

image

and

image

Exercise 11.34   Do some experimenting to see the emergence of a monotone subsequence with four terms was an accident or if it might be an example of a larger pattern.

(a)   Find your own additional integer and insert it into the sequence

image

at a random location. Can you find a monotone subsequence with four terms contained in your new sequence?

(b)   Repeat part (a).

(c)   Repeat part (a) again.

It is a fact that any sequence of ten distinct integers will have a monotone subsequence with four (or possibly more) terms. This is a special case, an example really, of a result first proved by Erdös and Szekeres in the 1930s. After we explore Mirsky’s Theorem, we will return to the Erdös–Szekeres Theorem and use Mirsky’s Theorem to prove it.

Mirsky’s Theorem connects the height of a poset to its antichains. Let image be a partially ordered set. An antichain is a subset of S where no two elements are comparable. For example, if image is ordered by containment, then

image

is an antichain.

Exercise 11.35   In this problem we will continue to work with the poset of subsets of image ordered by containment.

(a)   The poset of subsets of image, ordered by containment, contains 168 antichains. Find five of them.

(b)   The largest antichains in the poset of subsets of image contain six elements. Find such an antichain.

(c)   Find another antichain with six elements, or prove that the antichain with six elements is unique.

A maximal antichain is one that is not properly contained in any other antichain.

Exercise 11.36   For image, let image be a zigzag poset with n elements. Find a formula for the size of the largest antichain in image, and prove that your formula is correct. (Zigzag posets are defined in Exercise 6.49.)

DEFINITION 11.7.   An antichain cover of a poset is a collection of antichains, whose union contains all the elements of the partially ordered set. If the antichain cover consists of the antichains image, then the cover has size k.

Exercise 11.37   The following questions provide a bit of practice with antichain covers, in the context of the poset of subsets of image ordered by containment.

(a)   Prove that for any fixed integer k, image, the collection of k-element subsets is a maximal antichain.

(b)   Prove that the poset image can be covered by image antichains.

THEOREM 11.8 (Mirsky’s Theorem).   Let S be a finite set, partially ordered by image. Then the height of the poset image equals the size of a minimal antichain cover of S.

Below we outline two approaches to proving Mirsky’s Theorem. The first proof connects the height directly to antichains, and the second is a proof by induction.

Exercise 11.38   Let image be a finite poset.

(a)   Prove that the intersection of any antichain with a chain can contain at most one element.

(b)   Prove that if the height of image is H, then any antichain cover must contain at least H antichains.

Exercise 11.39   Let image be a finite poset, and define the height of an element image to be the maximum number of terms that occur in any chain with s as its largest element. These heights give us a function image, where image is the height of the element s.

(a)   If H is the height of the poset image, prove that h gives a surjection onto image.

(b)   Prove that if image, then image is an antichain.

(c)   Prove there is an antichain cover of image consisting of H antichains.

Exercises 11.38 and 11.39 immediately imply Mirsky’s Theorem:

PROOF OF MIRSKY’S THEOREM.   Let image be a finite poset of height H, where a minimal antichain cover of image contains A antichains. Then Exercise 11.38 shows that image, and Exercise 11.39 shows that image. Thus image.

image

Mirsky’s Theorem can also be proven by induction on the height of the poset.

Exercise 11.40   When you complete the following steps you will have outlined an inductive proof for Mirsky’s Theorem.

(a)   Base Case: Prove that Mirsky’s Theorem holds for finite posets of height 1.

(b)   Lemma: Prove that the set of all maximal elements of a finite poset forms an antichain.

(c)   Lemma: Let M be the set of all maximal elements of the poset image. Prove that image is a poset whose height is smaller than the height of image.

(d)   Use the two lemmas above to establish the inductive step.

Having proven Mirsky’s Theorem (twice!), we are now able to head toward a proof of the Erdös–Szekeres Theorem.

COROLLARY 11.9.   Let m and n be natural numbers. Then in any partially ordered set image with image or more elements, there is a chain with image elements or an antichain with image elements.

PROOF.   Using the argument from Exercise 11.39 we see that if the height of the poset is less than image then there is an antichain cover consisting of fewer than image antichains, where no two antichains overlap. Denote these antichains by image (image). Thus image, and so

image

Thus image, and so there must be an antichain with at least image elements.

image

Exercise 11.41   Use Corollary 11.9 to prove the Erdös–Szekeres Theorem: Each sequence of image distinct real numbers possesses a monotone subsequence with image terms. As a hint, define an ordering on the terms of the sequence by saying image if image and image.

For a graphical interpretation of this result, see Figure 10.

image

Figure 10. Mirsky’s Theorem guarantees that you can find an increasing or decreasing constellation of length image in the graph of a sequence with image distinct terms.

Exercise 11.42   In Exercise 6.8 we described a partial ordering on closed and bounded intervals image. Use Mirsky’s Theorem to prove that for any collection of image closed and bounded intervals there must be either image that are disjoint or image that share a common point.

GOING BEYOND THIS BOOK.   Mirsky first published his now eponymous theorem in [Mir71]. That paper is quite readable and is recommended, particularly for the way it discusses the connection between the result presented in this project and a related result, Dilworth’s Theorem. The original paper by Erdös and Szekeres [ES35] is also worth reading, particularly with an eye toward how these ideas have evolved over the decades since their first presentation.

11.7.  Euler’s Totient Function

We have seen that knowing the order of an Abelian group can be useful. It is immediate that image, but the order of image is less clear. Since the elements of image correspond to the integers k where image and image, we are simply wondering how many natural numbers less than or equal to m are relatively prime to m. The function

image

is known as Euler’s totient function. Our goal in this section is to find a nice formula for ϕ. As part of the motivation for studying ϕ comes from the units arising in modular arithmetic, this project is appropriate for students who have worked through Section 6.6.

Exercise 11.43   Explain why image for all primes p.

Here’s one that is a bit more difficult.

Exercise 11.44   Prove that if p is prime, then image.

DEFINITION 11.10.   Let image, where both are at least 2. Define the remainder map image by

image

For example, if you are considering image then

image

We do need to verify that this remainder map is well defined.3 To do this, consider any integer b where image. Thus image. So

image

Thus our definition does not depend on the representative used in computing the answer.

We now get to play with this remainder map. Let’s first look at image.

image

The remainder map has given us a bijection between the six elements in image and the six elements in image. Before we make any hasty conjectures, let’s look at another example, this time considering image:

image

In this case we do not get a bijection. As always, this ought just to make us more curious …

LEMMA 11.11.   If m and n are relatively prime, then image is a bijection.

PROOF.   Since image, it suffices to show that ρ is injective.

If image, then n and m divide image. Hence the least common multiple of m and n divides image (see Exercise 4.42). Because m and n are relatively prime, image divides image. So image.

image

Exercise 11.45   Show that the remainder map gives a bijection:

image

when m and n are relatively prime.

Exercise 11.46   Let the prime decomposition of n be image, with each image. Prove that

image

At this point we have an effective method for computing Euler’s totient function. For example, image.

Exercise 11.47   Compute image.

Finally, we are in a position to present an alternate formula for ϕ that is often preferable:

THEOREM 11.12.   Let image. Euler’s totient function ϕ is given by

image

where image stands for the set of primes p that divide n.

For example,

image

Exercise 11.48   Prove Theorem 11.12.

If you are ambitious, we recommend that you explore bounds on Euler’s totient function in the following exercise.

Exercise 11.49   Figure 11 shows the graph of Euler’s totient function image, up to image.

(a)   Prove that this graph is bounded above by the straight line image, as Figure 11 suggests.

(b)   Prove that no line of slope image could be an upper bound on the graph of ϕ.

(c)   In Figure 11 we have drawn in the line connecting the origin to image. It appears that this line gives a lower bound on ϕ. Show that this is in fact false.

(d)   Prove that image for all integers image.

image

Figure 11. The Euler totient function for image to image.

11.8.  Proving the Schröder–Bernstein Theorem

In Section 7.6 we demonstrated the tremendous utility of the Schröder–Bernstein Theorem, but we did not provide a proof of this theorem. The goal of this project is for you to construct a proof, and before working on this project you should review Exercise 7.18.

THEOREM 11.13 (The Schröder–Bernstein Theorem).    Let A and B be sets, and assume there are injections image and image. Then A and B have the same cardinality.

Every element image determines a sequence of elements, alternating between the sets A and B. The sequence starts with a and then alternately applies f and g:

image

It is cleaner to visualize this sequence as the filling in of blanks in the following diagram:

image

You can also attempt to build out this sequence to the left, essentially applying the inverse of g and f. However, this is only possible if you are working with elements in the range of g and f, respectively. Thus, while the sequence will continue to the right indefinitely, it may not extend indefinitely to the left.

image

We will call these sequences back-and-forth sequences, as they bounce between A and B. When discussing a back-and-forth sequence, we always assume that it has been extended as far as possible to the left.

EXAMPLE 11.14.   The Schröder–Bernstein Theorem can be easily applied to show that image. For example, there are injections image and image, where image and image. These functions lead to the following sequences:

image

The top back-and-forth sequence cannot be extended to the left, because the value 0 is not in the range of g. The middle sequence can be extended one step to the left, since image. However, image, so this sequence extends no further to the left. Finally, it is clear that the bottom sequence extends indefinitely to the left. Thus we have:

image

Exercise 11.50   Let image and image be injections. Define a relation ~ on the elements of A by saying image when a and image are both in a shared back-and-forth sequence.

(a)   Prove that ~ is an equivalence relation, and therefore the equivalence classes partition A.

(b)   State and prove the same result for the elements of B.

DEFINITION 11.15.   A back-and-forth sequence that extends indefinitely to the left is a bi-infinite sequence. Those that cannot be extended indefinitely to the left are called A-sequences or B-sequences, depending on whether the left-most element is in A or B.

Exercise 11.51   This exercise refers back to the situation described in Example 11.14.

(a)   Show that there are only two A-sequences for this example.

(b)   Show that there are infinitely many B-sequences.

Exercise 11.52   Let image and image be injections.

(a)   Prove that f is a bijection when restricted to the elements occurring in an A-sequence.

(b)   Prove that image is a bijection when restricted to the elements occurring in a B-sequence.

(c)   Prove that f and image are both bijections when restricted to the elements in a bi-infinite sequence.

Exercise 11.53   Prove the Schröder–Bernstein Theorem.

Exercise 11.54   Apply your proof of the Schröder–Bernstein Theorem to construct an explicit bijection between image and image.

Exercise 11.55   Show that the bijection h given in Exercise 7.18 comes directly from the injections f and g discussed earlier in Section 7.6.

11.9.  Cauchy Sequences and the Real Numbers

This project is in fact two projects. We begin with an elementary discussion of Cauchy sequences. The first sub-project proves that a real-valued sequence is Cauchy if and only if it converges. The second sub-project, which can be done independently of the first, outlines the construction of the real numbers via Cauchy sequences of rational numbers. We make frequent use of material in Chapter 8, up to and including Section 8.5, throughout this project.

DEFINITION 11.16.   A sequence image of real numbers is a Cauchy sequence if, given any real number image, there is an image such that

image

whenever m and n are both greater than N.

The first exercise provides a good test case for practicing with this definition.

Exercise 11.56   Let image be a decimal expansion of some real number r. Here image and each image. For each image define

image

Prove that the sequence of rational numbers image is a Cauchy sequence.

Exercise 11.57   Construct an example of a Cauchy sequence that is not monotone.

11.9.1. Real-valued Cauchy Sequences.   The goal of this sub-project is to prove that a real-valued sequence image converges if and only if it is Cauchy.

Exercise 11.58   Let image be a real-valued Cauchy sequence. Because image is a Cauchy sequence, by its definition there is an image such that image for all m and n greater than N. Consider then the set of initial terms image and the set of values that appear in the tail image.

(a)   Explain why the set of initial terms, image, is bounded.

(b)   Show that the terms in the tail, image, are all contained in the closed interval image.

(c)   Combine the facts above to prove that Cauchy sequences are bounded.

Because Cauchy sequences are bounded, given a Cauchy sequence image we may define an associated sequence of upper bounds image. We let image be any upper bound for all the terms in image. We define image to be the average of image and image, if this average is an upper bound for image, and otherwise we set image. And in general

image

This sequence of upper bounds is illustrated in Figure 12, where the original Cauchy sequence is graphed using black circles, and the associated sequence of upper bounds (with image chosen to be 4) is displayed with crosses. The figure indicates, correctly, that both sequences converge to a limiting value of 2.

image

Figure 12. A Cauchy sequence is shown with black circles, and an associated sequence of upper bounds is displayed with crosses.

Exercise 11.59   Let image be a Cauchy sequence and let image be any upper bound on the set of terms image.

(a)   Prove that the associated sequence of upper bounds image defined as above is a bounded, monotone decreasing sequence. Conclude that it must converge to some real number μ.

(b)   Prove that the Cauchy sequence image converges to μ as well.

Exercise 11.59 gives a very important method of demonstrating that a real-valued sequence converges. Constructing a proof of convergence using the original definition requires that you also already know the limit of the sequence. Proving that a sequence is Cauchy, however, requires only an argument that uses the terms of the sequence.

You are now in a position to establish the main result of this subsection. We anticipate that you will want to use the following approach: If image converges to a limiting value α, then for any image you know that

image

This is an application of the triangle inequality: image for all image.

Exercise 11.60   Prove that a real-valued sequence image is Cauchy if and only if it converges.

11.9.2. Constructing the Real Numbers.   Given the natural numbers, image, you can construct the set of integers, image, by adding 0 and defining negative numbers. Given the integers, you can construct the rationals, image; one method for doing this is presented in Exercise 6.75. Here we describe a way to construct the real numbers, image, from the rationals.

There are various definitions that lead to the real number system, and the following set of axioms is one of the most common. This definition begins by assuming there is a set of elements denoted by image, two binary operations image and image that map image to image, and a relation image on image. This collection defines the real numbers if it satisfies the following 12 axioms.

(1)   For every a, b, and image, image and image.

(2)   For every a and image, image and image.

(3)   For every a, b, and image, image.

(4)   There is an element called 0 such that image for all image.

(5)   Given any image, there is an element image such that image.

(6)   There is an element called 1 such that image for all image.

(7)   For every image, there is a image such that image.

(8)   For every image and image, if image then image.

(9)   For every image and image, if image and image then image.

(10)   For every a and image, exactly one of the following is true: image, image, or image.

(11)   For every a and image and every image, if image then image.

(12)   If S is a non-empty subset of image that is bounded above, then S has a least upper bound.

Our goal then is to start with the rational numbers, construct a set image along with image and image, and verify that the axioms above hold. In order to do so, we limit our discussion in this section to rational sequences, that is, sequences image where every image. If q is a rational number, we let image be the constant sequence,

image

DEFINITION 11.17.   Two Cauchy sequences image and image are defined to be equivalent if the sequence of differences image converges to 0.

Exercise 11.61   Prove that the sequence image and the sequence image are equivalent Cauchy sequences.

Exercise 11.62   Prove that the “equivalent” relation defined above lives up to its name and is indeed an equivalence relation on the set of Cauchy sequences.

We define image to be the set of equivalence classes of rational Cauchy sequences. We define addition and multiplication of rational Cauchy sequences via

image

For example,

image

Exercise 11.63   Addition and multiplication are defined for specific sequences, not for equivalence classes. And our definition of image states the the elements of image are not sequences, but rather equivalence classes of sequences. This means you have some work to do! We need to establish that the sum and product of rational, Cauchy sequences are rational Cauchy sequences. We also need to verify that these operations are well defined on equivalence classes.

(a)   Prove that if image and image are rational Cauchy sequences, then so is image.

(b)   Prove that if image and image are rational Cauchy sequences, then so is image.

(c)   Let image and image be equivalent sequences. Prove that

image

(d)   Let image and image be equivalent sequences. Prove that

image

Exercise 11.64   Verify that the Algebraic Axioms hold, with the constant sequence image functioning as the additive identity 0 and image functioning as the multiplicative identity 1.

DEFINITION 11.18.   For two rational Cauchy sequences image and image, define image if they are not equivalent and if there is an image such that image for all image. We say image if either the two sequences are equivalent or if image.

Exercise 11.65   We have defined image for specific sequences, not on equivalence classes of sequences. We need to verify that image makes sense on equivalence classes, and then verify that the order axioms hold.

(a)   Let image and image be equivalent sequences. Prove that

image

Thus our definition is well defined on equivalence classes.

(b)   Verify that the order axioms hold.

Exercise 11.66   Let S be a set of rational Cauchy sequences which is bounded above. Prove that S has a least upper bound by mimicking the “sequence of upper bounds” approach from the previous subsection. If you are working with representatives from the equivalence classes, be sure to verify that your arguments hold regardless of the chosen representatives.

Finally, having constructed the real numbers, it is a good idea for us to verify that the usual description in terms of decimal expansions is indeed accurate. Define a standard decimal expansion sequence to be a Cauchy sequence of the form discussed in Exercise 11.56, with the added requirement that given any image there is an image such that image, so that decimal expansions are not allowed to have a tail of repeating 9s. This final exercise justifies the use of decimal expressions to represent the real numbers.

Exercise 11.67   Prove that every rational Cauchy sequence is equivalent to a standard decimal expansion sequence.

GOING BEYOND THIS BOOK.   There are other ways of constructing the real numbers that do not involve equivalence classes of Cauchy sequences. The method of Dedekind cuts is nicely presented in Abbott’s Understanding Analysis [Abb15]. Both the approach via Cauchy sequences and the approach of Dedekind date to the early 1870s. It should be noted, however, that the history of constructions of image is complicated, with credit certainly being due to Cantor, Heine, Dedekind, Méray, and Weierstrass (at the least!). Some would also include work of Simon Steven on numbers with unbounded decimal expansions in the 1600s, and perhaps even work of Eudoxus in ancient Greece. While it is not particularly focused on constructions of image, Grabiner’s The Origins of Cauchy’s Rigorous Calculus [Gra81] is an excellent history of the development of analysis.

11.10.  The Cantor Set

In this project we introduce the Cantor set, a subset of the closed unit interval image that is formed by repeatedly removing open intervals. The Cantor set has a number of interesting and often counterintuitive properties, the most accessible being the fact that it is both a small subset of image and a large collection of points in image. In this project we assume you have worked through Chapters 7 and 8, or at the least have a good intuition about the real numbers.

The process for defining the Cantor set image begins with the interval image, and then creates

image

The subset image consists of two closed intervals, and we remove the “middle thirds” from both of them to form image. That is, we remove image and image from image to form image:

image

The subset image consists of four closed intervals. We then remove the middle thirds of each of them to form image, consisting of eight closed intervals, and we continue this process to create a subset image for each image. The process is visualized in Figure 13.

image

Figure 13. The first steps in constructing the Cantor set by removing middle thirds.

The Cantor set image is the end result of this process; that is, image is the set of points in image that are never removed in forming any of the sets image.

You can also develop the Cantor set using rescaling and translating. If image and image, define

image

For example, if image then image is the interval image. And, directly related to our discussion of the Cantor set,

image

Exercise 11.68   Verify that the next step of forming image can be expressed in the same manner:

image

DEFINITION 11.19.   Let image and image be as above. Define the set image to be

image

and the Cantor set image is the intersection

image

of all of these sets of closed intervals.

Exercise 11.69   Prove that the set image is composed of image closed intervals, each of length image, for a total combined length of image. Then prove that the open intervals removed in forming the Cantor set are pairwise disjoint and have combined length equal to 1.

Exercise 11.69 indicates that image is small, but it is not empty. The numbers 0 and 1 are in every image, hence 0 and 1 are elements in the Cantor set image. And there are quite a few more points in image in addition to these two.

Exercise 11.70   We pointed out that 0 and 1 must be in image. Find two other real numbers that are in image and explain how you know that they must be in the Cantor set. Then describe an infinite set of real numbers in image.

Our discussion of the cardinality of the Cantor set will use the Nested Intervals Property of the real numbers.

THEOREM 11.20 (Nested Intervals Property).    Let image be a collection of closed, bounded intervals that are nested; that is,

image

Then the intersection of all of these intervals is non-empty:

image

Further, if the widths of the intervals decrease to zero, that is

image

then the intersection is a set containing a single real number.

Exercise 11.71   Exercise 8.39 asked you to prove the main part of this result: the intersection is non-empty. Please complete this exercise if you haven’t done so already, and then prove that the additional claim that having the widths go to 0 implies that the intersection is a single real number.

In order to apply the Nested Intervals Property to the Cantor set we need a bit of additional notation. The first approximation to the Cantor set, image, consists of two intervals,

image

We will refer to image as the left interval, which we denote image, and image as the right interval, image. Thus image.

In passing to image we need to divide image into its left and right sub-intervals, image and image, and we need to divide image into its left and right sub-intervals, image and image. Thus

image

Similarly listing the eight intervals that comprise image from left to right, we have

image

Define a properly nested sequence of Cantor sub-intervals to be a nested sequence where the first interval comes from image, the second from image, and so on. One example would be

image

Exercise 11.72   Find a one-to-one correspondence between the set of properly nested sequences of Cantor sub-intervals and the Cantor set, and prove that your function is a bijection.

Exercise 11.73   Prove that the Cantor set is uncountable. (We recommend using the idea behind Cantor’s Diagonal Argument, applied to properly nested sequences of Cantor sub-intervals.)

In Exercise 11.69 you showed that the Cantor set is formed by removing open intervals of total length 1, starting from the closed interval image. This indicates that image must be rather small. Yet in Exercise 11.73 you proved that image has the same cardinality as image. That is a good stopping place. However, if you feel like pushing forward just a bit more, feel free to work on a final exercise.

Exercise 11.74   If A and B are two sets of real numbers, we can define the sum of A and B as

image

For instance, if image and image, then

image

As another example, if A and B are both equal to the closed interval image, then image. What’s remarkable is that, despite how small image seems to be as a geometric subset of image, you can show that image is equal to the entire closed interval image as well.

(a)   Recall that image. Thus image consists of four solid squares in the first quadrant, as shown in Figure 14. Prove that image consists of those image such that the line image intersects image. For example, Figure 14 shows that image.

(b)   Conclude that image.

(c)   Use induction to prove that image.

(d)   Prove that image. Warning: This does not follow immediately from the previous statement!

image

Figure 14. The Cartesian product image for Exercise 11.74.

11.11.  Five Groups of Order 8

In Section 10.8 we explored ways you can distinguish some groups of order 12. In this project you will come up with arguments proving that five groups of order 8 are indeed distinct, that is non-isomorphic. It is also the case that these five groups are the only groups of order 8, meaning that any group of order 8 must be isomorphic to one of these five groups. Proving that they are the only groups of order 8 is, however, beyond what we can do given our limited introduction to group theory.

Exercise 11.75   Prove that the three Abelian groups

image

are distinct. That is, no two are isomorphic.

Exercise 11.76   Show that the symmetries of a square – image from Section 10.2 – is a non-Abelian group of order 8. What result proves that this group cannot be isomorphic to the three Abelian groups of order 8 in Exercise 11.75?

Let’s look at one more group of order 8, the unit quaternions. Let

image

where all products can be determined from the following equalities: image, image, image and image. These rules can be remembered by using the circle shown in Figure 15: the product of two consecutive elements is the third one, going clockwise. The same is true going counterclockwise, except that you get the negative of the third element.

image

Figure 15. A mnemonic to help remember multiplication in image.

Exercise 11.77   Make a Cayley table for image, and illustrate via examples that image is not Abelian.

Exercise 11.78   If you have had exposure to matrices and matrix multiplication, then you can also view image as a group of eight matrices. As examples, the elements i and j are described by the following image matrices:

image

Write down all eight matrices that form image.

Exercise 11.79   In order to prove that we now have five non-isomorphic groups of order 8, you need to show that image and image are not isomorphic. One approach would be to determine how many elements in each group have order 2.

If you would prefer a bigger challenge, where we have not given as many guiding hints, then you might explore groups of order 18 as an alternative to this project. There are five non-isomorphic groups of order 18. Find them, describe them, and prove that it is the case that no two of them are isomorphic.


1   Why is it important for us to say “for convenience” when making this assumption? Why is it also a legitimate step?

2   The conjugate of a complex number α is usually denoted by placing a bar over α, as in image.

3   Remember ‘A Cautionary Tale’ in Section 6.5?