As we mentioned in the Preface, solving the in-text exercises as you encounter them in the text is an essential task. Make earnest attempts on your own to solve each problem that shows up in the text before reading any further or consulting the material in this appendix. Use the solutions, answers, and comments below to check your work or assist your progress if you get stuck.
Chapter 1
Exercise 1.1 Here’s one way to show that if a and b are two non-negative real numbers, then
First,
because the square of any real number is non-negative. Expanding the left-hand side yields
Adding to both sides gives
Since ab is non-negative, we know further that
This can be rewritten as
The conclusion we desire now follows from taking square roots.
Take a look at your scratch work and attempts on this problem. You might have started with , squared both sides, and then followed a series of calculations and deductions that are similar to the steps just given, but in reverse. That is certainly a correct approach to first understand the problem. However, the proof should end with
as the conclusion, not the first step. “If …, then …” statements and their proofs are studied in Chapter 2.
Exercise 1.2 In the sixth month, the original pair and their first and second offspring each create offspring, bringing the total number of pairs to 8.
In the seventh month, the original pair; their first, second, and third offspring; and their first offspring’s offspring each create offspring, bringing the total number of pairs to 13.
Now things are getting quite complicated – there must be a better way to keep track of things …
Exercise 1.3 They are
Exercise 1.4 The key here is to make sure that counts all of the rabbit pairs that exist in month
in a way that doesn’t count any pair twice. Every rabbit pair in month
is exactly one of two types:
(a) It already existed in month n.
(b) It did not exist in month n.
How many pairs are in case (a)? That’s . How many pairs are in case (b)? That’s
, because we can count their parents, pairs that must have existed two months prior to month
. Add these two numbers together to get
.
Notice that, in order to justify that is the correct count for the number of rabbit pairs in month
, we are tacitly assuming that
is the correct count for month n and
is the correct count for month
. This type of reasoning lies at the heart of proofs by mathematical induction, a topic we preview in this chapter and study in Section 2.6.
Exercise 1.5 Let give the number of different ways to make sums of 1s and 2s that equal n. Then the displayed equalities show that
,
,
, and
. You might now guess that
for
as well. How could you make yourself feel more confident about this guess? And if the equality indeed holds in general, how might you prove it?
Exercise 1.6 The correct value for P should certainly be greater than . If
, then
, while
causes
.
Exercise 1.7 The calculations are
and
Exercise 1.8 The first verification is the calculation
For the second one, first rewrite
as
While it would be easy enough to continue from here by substituting the numerical formulas for ϕ and , let’s instead foreshadow the proof technique of Proposition 1.4 by rearranging the terms in the numerator to give
Since you already know that
and
it must be the case that
Exercise 1.9 First use Lemma 1.5 to rewrite
as
Then rearrange the terms in the numerator to give
Exercise 1.10 (a) If ,
, and
, then
and
, but
and so
. Thus, the statement is false.
This is only one of many different counterexample sets of values for a, b, and c; another one is ,
, and
. You only need one counterexample to prove that the statement is false.
(b) All of your experimentation should have led you to guess that the statement must be true. For instance, if you tried ,
, and
, then that particular instance is true because
.
Here’s a proof that the statement is true in general.
PROOF. Since you are assuming that , you know that there is a natural number m such that
. Multiplying both sides of this equation by c gives
But now look: bc is a times the natural number mc. This means that by the definition of divisibility.
(c) This statement is true and is called the “transitive” property of divisibility. In Chapter 6 you will study the transitive law in several different contexts.
PROOF. Since , there is some natural number m such that
. Since
, there is some natural number n such that
. Since
and mn is a natural number, we conclude that
.
(d) This proof will follow the proof of Proposition 1.7 very closely. Be sure that you understand how the numbers 2 and 3 can be inserted into that proof with little difficulty.
PROOF. Since , there is a natural number m such that
. Similarly, since
, there is a natural number n such that
. This means that
where is a natural number, which shows that
.
(e) Look ahead to Lemma 1.10!
Exercise 1.11 (a) Here’s one way to prove that this statement is true.
PROOF. If , then there exists a natural number c such that
. Multiplying both sides of this equation by n gives
which shows that (because nc is a natural number).
Another way is to use the result of part (b) in Exercise 1.10. Can you see how to do this?
(b) This statement is not the same as the one in (a) because the hypothesis and conclusion have switched places, so don’t expect a proof of (b) to be similar to a proof of (a).
PROOF. Natural numbers are either odd or even. If n were odd, then would also be odd, which would imply that
. So the only cases when
must have n even, which means that
.
(c) This statement is false: for a counterexample, simply let .
Exercise 1.12 Since , there is a natural number c such that
. Since
and both a and are non-negative, we know that
. Thus
.
Exercise 1.13 The only way to factor 1 is , and those two factors aren’t distinct.
Exercise 1.14 There are 25 primes less than 100:
Notice that the proportion of natural numbers between 1 and n that are prime seems to decrease as n gets larger. If this fascinates you, research the “prime number theorem.”
Exercise 1.15 They are
Do you see the systematic way we enumerated them? Maybe you found a better way to do this.
Exercise 1.16 If a is any integer, then . This means that
.
If a is any non-zero integer, then there’s no such that
. This means that
. However, given our definition of divisibility,
seems to be true, since any integer c satisfies
. But you’ve likely heard that it’s never good to divide by 0, so we should really modify our definition of divisibility for
to exclude this possibility. If you need convincing, see Zero: The Biography of a Dangerous Idea by Charles Seife [Sei00].
Exercise 1.17
Proposition 1.7 If a, b, and c are any integers such that both and
, then
.
PROOF. Since , there is an integer m such that
. Similarly, since
, there is an integer n such that
. This means that
which shows that is an integer multiple of a. In other words,
.
Lemma 1.10 If a, b, and c are any integers such that both and
, then
.
PROOF. Since , there is an integer m such that
. Similarly, since
, there is an integer n such that
. Thus
which shows that c is an integer multiple of a, so we conclude that .
Exercise 1.18 In view of
we see that .
Similarly,
shows that .
Exercise 1.19 The Division Algorithm says that, given a natural number b, every integer can be written in the form , where r is an integer between 0 and
(inclusive). So if
, every number can be written as
, which is just a small variation on the statement that every integer can be expressed as
, or
.
Exercise 1.20
(a) is not true. Natural numbers are positive.
(b) is true from the definition of
.
(c) is not true, as was proved in Proposition 1.14.
(d) Since , the statement is true.
(e) is not true. We can prove this with a short proof by contradiction.
PROOF. Assume for the moment that the statement is true. Then
, where
is a rational number. That implies that
, where
is another rational number, namely
. This contradicts the fact that
is not a rational number.
(f) is also not true. Try a short proof by contradiction, like the one just given above.
(g) , so the previous result applies.
(h) is true, because
.
One of the take-aways from this exercise is that while the sum (or product, or difference) of two rational numbers is rational, the same can’t be said for irrational numbers.
Chapter 2
Exercise 2.1 As in Example 2.1, let C be the proposition “Cecelia is a knight” and let D be “Desmond is a knight.” Then the truth of Cecelia’s statement is determined by the fourth column of the table at the top of the next page.
Since Cecelia is making the statement, we are looking for any rows where the first and final entries agree. This only happens in the final row, so both Cecelia and Desmond are knaves.
C | D | ![]() |
![]() |
T | T | F | F |
T | F | F | F |
F | T | T | T |
F | F | T | F |
Exercise 2.2 P is true and Q is false, hence is true. This means
is true. One way to express this statement is “4 is a divisor of
or 4 is not a divisor of 6.”
Exercise 2.3 They are
P | Q | ![]() |
T | T | T |
T | F | T |
F | T | T |
F | F | F |
and
P | Q | ![]() |
![]() |
![]() ![]() |
![]() |
T | T | F | F | F | T |
T | F | F | T | F | T |
F | T | T | F | F | T |
F | F | T | T | T | F |
Notice that the first, second, and last columns of the two tables agree. This means that whatever statements you might use to replace P and Q, the truth values of and
will be the same. Later we will say that these two statements are “logically equivalent.”
Exercise 2.4 There will be rows for each table because there are 3 statements (P, Q, and R) and 2 possible values for each variable (T and F). Here are the tables:
P | Q | R | ![]() |
![]() |
T | T | T | T | T |
T | T | F | T | F |
T | F | T | F | T |
T | F | F | F | T |
F | T | T | F | T |
F | T | F | F | T |
F | F | T | F | T |
F | F | F | F | T |
and
P | Q | R | ![]() |
![]() |
T | T | T | T | T |
T | T | F | T | F |
T | F | T | T | T |
T | F | F | T | F |
F | T | T | T | T |
F | T | F | T | F |
F | F | T | F | F |
F | F | F | F | T |
Exercise 2.5 Observe that the final column of
P | Q | ![]() |
![]() |
T | T | T | T |
T | F | T | T |
F | T | F | T |
F | F | T | T |
is all Ts.
Exercise 2.6 Truth tables show that the statements in Exercise 2.3 are logically equivalent, while the statements in Exercise 2.4 are not.
Exercise 2.7 The third and sixth columns are the same in this truth table:
P | Q | ![]() |
![]() |
![]() |
![]() |
T | T | T | F | F | T |
T | F | F | T | F | F |
F | T | T | F | T | T |
F | F | T | T | T | T |
Exercise 2.8 You can verify this claim by comparing the truth tables for and
.
Exercise 2.9 You can visualize the sum as a sequence of stacked squares, the first stack being of height 1, the second of height 2, and so on. Add n squares to the top of the first stack, and then
squares to the second stack, and so forth. The result is a rectangular array of boxes that is
. Thus, as before,
. Can you think of a way to draw a diagram illustrating the general case?
Exercise 2.10
(a)
(b) This might require some insight. Since
and
we have
Thus .
Exercise 2.11 Let a be a natural number and let x be an irrational number. Assume to the contrary that the product ax is rational, so that for some integer m and some natural number n. Then
, where an is a natural number, contradicting the fact that x is irrational.
What happens if you try the argument above with ?
Exercise 2.12 No, because has a rational solution of
.
Exercise 2.13
(a) Original implication: “If is even, then n is even.”
Contrapositive: “If n is not even, then is not even.” Because the contrapositive is logically equivalent to the original implication, either both statements are true or both statements are false. In this case, the contrapositive is probably in the best form to prove that they are true: if n is not even, then
for some integer k, and so
, which is certainly not even.
Converse: “If n is even, then is even.” This is also true, but for a reason that is completely independent of the other two cases (see Exercise 1.11).
(b) Original implication: “If is rational, then x is rational.” This is false: a counterexample is
.
Contrapositive: “If x is not rational, then is not rational.”
Converse: “If x is rational, then is rational.” This is true. The proof is direct: if x is rational, then
for some
and
, and so
, which is also a rational number.
(c) Original implication: “If , then
.” This is false: a counterexample has
and
.
Contrapositive: “If , then
.” Notice how negating
leads to
.
Converse: “If , then
.” This is true. Can you prove it?
(d) Original implication: “If the professor is 5 minutes late, then class is cancelled.” This is likely false; consult your instructor for the correct answer.
Contrapositive: “If class is not cancelled, then the professor is not 5 minutes late.” Be sure you understand why this statement must have the same truth status as the original implication.
Converse: “If class is cancelled, then the professor is 5 minutes late.” This is almost certainly false. Class could be cancelled for lots of reasons that have nothing to do with your instructor being late. For example, if she happens to walk by some pygmy marmosets just before class, she will probably not make it to class that day: she’ll just keep petting them.
Exercise 2.14 We hope you can do these types of problems quite well by now …
Exercise 2.15 …including this one!
Exercise 2.16 After rewriting the inequality as , we can factor the right-hand side as
. If n is greater than 1, both n and
are positive, so
. Also, if n is less than 0, both n and
are negative, so
. The only integer values of n we haven’t yet considered are
and
, and each one makes
. Thus,
for all integers n.
Exercise 2.17 Replace n with q in the preceding proof, and all is fine until you get to the sentence that starts with “The only …”. For a specific counterexample, try .
Exercise 2.18 Two of the four statements are true – which ones?
Exercise 2.19 Part (a) shows that a solution exists, while part (b) shows that the solution is unique. Proving the truth of a statement usually requires two arguments like these.
Exercise 2.20
(a) , h is mortal.
(b) , h scored higher than me on the test.
(c) , t is better than apple pie).
For practice, rewrite part (c) starting with .
Exercise 2.21 The first one can be “For every real number y there is a real number x such that .” The second can be “There is a real number x such that, for all real numbers y,
.”
Exercise 2.22 She said persons,
true soul mate; in this order, the soul mate can depend on the person. Dilbert stupidly thought
true soul mate for all persons. His question about the monkey is a good one.
Exercise 2.23 Only one of the statements is false – which one?
Exercise 2.24
(a)
(b) Now try to write this with an
but no ~.
(c) , t can’t stop the Sandman!
(d) , d will not open. One wonders why they bother to make the stop! What is a more accurate statement as the train approaches the station?
Notice that in part (a), the statement accommodates either or
. Another possible statement would be
(a)
The reason this statement also works is that for any two distinct real numbers, one of them must be less than the other, and so we might as well call the smaller one r.
Exercise 2.25 and
.
Exercise 2.26 The inductive assumption lets us know that expression B is a multiple of 6.
Exercise 2.27 The only modification necessary to the template is replacing “Assuming that is true for all values of k with
” with “Assuming that
is true.” Now revisit Theorem 2.12 with 9 and 10 in place of 6 and 7.
Exercise 2.28
(a)
(b)
(c)
(d)
It is pretty clear that simple induction will not be helpful, as there is no apparent way to relate the factorization of to the factorization of n.
Exercise 2.29 In the final presentation of your proof, five base cases might be acceptable but are certainly unnecessary and possibly distracting: only two bases are needed. In your initial work on the problem you might have directly verified that the statement holds for five cases like these, and that’s totally fine.
Exercise 2.30 The proof uses strong induction because both and
, which are the respective formulas
and
are assumed to be true to prove that is true.
Exercise 2.31 Despite the fact that Exercises 1.45 and 1.46 involve the Fibonacci numbers, neither requires consideration of only the two previous cases: Exercise 1.45 can be proved using simple induction and a single base case, while Exercise 1.46 is probably best proved with a single base case and a strong induction step that assumes the truth of the statement over a wide range of previous cases.
Exercise 1.47 might best be proved directly: consider the expression
after writing n in terms of powers of 10.
Exercise 1.48 is probably best proved using strong induction, by letting be the highest power of 2 less than or equal to
and using the inductive assumption to tell you something about
. The argument is similar to one you might use for Exercise 1.46.
Exercise 2.32 Start by declaring at the outset that x is a real number such that . The proof uses simple induction on the natural number n.
Base Case: If ,
and
both equal
, so the inequality holds (as an equality).
Inductive Step: We assume that the inequality is true for . Then to prove that the inequality is true for
, consider
where the first inequality in the derivation uses the inductive assumption, and the final inequality holds because can’t be negative. Once you have also said where it is important in the derivation that
, you will have a complete proof.
Exercise 2.33 Look for them!
Exercise 2.34 There is no absolute minimum number of examples that are needed to provide the basis for a conjecture, and it is sometimes the case that more examples are better. Using data for only the first four positive integers might not be enough for this problem. On the other hand, you can waste a lot of time checking cases when an attempted proof might highlight the sort of cases that could be problematic.
Exercise 2.35 Before you play, be kind and wish your friend good luck.
Exercise 2.36 Among other things, be sure that you explain why the second player always has a move to make.
Exercise 2.37 The case of is handled in the paragraph that follows this exercise.
Exercise 2.38 If four cases aren’t enough, try eight or more.
Exercise 2.39 For example,
Exercise 2.40 Here’s an easy question to ask: “Can you always avoid using ?”
Exercise 2.41 In part (b), you need to point out that so that you avoid dividing by 0.
Exercise 2.42 Using no lines leaves the plane whole, so seems best.
Chapter 3
Exercise 3.1 Try “” or “
.”
Exercise 3.2 When , the element
is an integer, so it’s already in the set in the form
.
Exercise 3.3 We know that because setting
and
in the definition of
gives
.
We also know that because taking
and
in the definition of
gives
. Note that there is no requirement in the definition of
that
be in lowest terms when
is an integer.
Now let’s show that , using a proof by contradiction. Suppose for the moment that
. Then Exercise 3.2 implies that
Cross-multiplying gives , but the right-hand side is divisible by 3 while the left-hand side is not. This is a contradiction, so we conclude that
.
Exercise 3.4 Regarding the existential quantifier, the assertion that is non-empty means that there is at least one element x in that set; and, by the definition of the set,
is true, so
is true as well. Conversely, if
is true, then
is true for some
. This means that x is in the set
, so the set is non-empty.
Exercise 3.5 If , then
for integers x and 0, so
.
Here’s one reason is a proper subset of
if
:
, but
.
Exercise 3.6
(a) 4 is not equal to 1, 2, 3, or , so
.
(b) Since ,
.
(c) is not equal to 1, 2, 3, or
, so
.
(d) , but
.
Exercise 3.7 We can put them in a line: . They are proper subsets owing to numbers like
,
, and
, respectively.
Exercise 3.8 They are ,
,
,
,
, and
.
Exercise 3.9 This requires two arguments: one to show that , and one to show that
.
For the first, let . Then
for some
, so that
with
. This means that
, so
.
The second argument is easy: , but
.
Exercise 3.10 The sets are as follows.
(a)
(b)
(c)
(d)
(e)
Exercise 3.11 Both statements are logically equivalent to .
For the first statement, implies that B contains no elements of A, and
implies that A contains no elements of B; thus A and B have no common elements. Conversely, if A and B have no common elements, then we must also have
and
.
Exercise 3.12 Remember that .
Exercise 3.13 (a) In Figure 2 on page 71, both and
can be seen as the crescent-shaped region on the left.
(b) First, let . This means that
and
. Since x is not an element of B, it can’t be an element of both A and B, so
. Thus,
and
, so
.
Second, let . This means that
and
. Since
, the only way that x is not in
is if
. Since
and
,
.
Exercise 3.14 The proof is similar to the first half. Let . By the definition of the union of two sets, this implies that
or
, which (again by the definition of union) means that
or
or
. This in turn means that
or
, which implies that
. Thus, we have shown that
.
Exercise 3.15 Let . Then it must be the case that
or
. Either way, we know
. We also know that
or
, and so
. Thus,
.
Exercise 3.16 Begin by assuming that , and use the assumption that
to prove that
.
Exercise 3.17 Both expressions describe the shaded region in Figure 1. This should be a sufficient hint for leading you to a proof (or a radioactive safe room).
Exercise 3.18 Looking at a 2-set Venn diagram, we see that enumerating all of the elements of A plus all of the elements of B double-counts the elements in . So, to get the correct size of
, we must subtract the number of elements in
from
.
A similar formula for accounts for double- and triple-counting certain elements:
Exercise 3.19 There are many good ways to do this. Our attempt at a picture is shown in Figure 2. The intervals are nested, as each sits inside the previous interval, like a matryoshka doll.
Exercise 3.20 The key is that “for some” expresses the same concept as “or,” and “for all” corresponds to “and.”
Exercise 3.21 The union consists of all of the points on the curves in Figure 3. The intersection contains four points:
Exercise 3.22 There are no ordered pairs because there is no element available for the first entry, so .
Exercise 3.23 Suppose that A and B are both finite. Since an ordered pair is just a short sequence of length 2, Proposition 2.13 tells us that
because there are choices for the first term and
choices for the second term. If at least one of A and B is empty, then either Exercise 3.22 or Proposition 2.13 shows that the formula still holds. The only other case is if at least one of A and B is infinite and the other is non-empty, and it shouldn’t be too hard to see that
must be infinite in this case.
Exercise 3.24 , whereas
.
Exercise 3.25
(a) Any set A with can be expressed as
for some element a, so
.
(b) Similarly, if , then
for some distinct elements a and b, so
.
(c) if
. Can you write down all eight elements of
?
Exercise 3.26 Since ,
is equal to the 4-element set
contains 16 elements – can you find all of them?
Exercise 3.27 The first case requires , so the eight possible subsets B are
The second case requires . The subsets B are listed below, and the corresponding sets
are obtained by simply removing the element d:
Exercise 3.28 In no particular order, they are
Exercise 3.29 For the partition into three blocks, note that
Exercise 3.30 Since there are n options for the ith term of the sample, the total number of samples is
Exercise 3.31 counts the number of ordered samples of size 2, without replacement:
Exercise 3.32 The empty set is the first set in any path. Regardless of the choices made along the way, it is clear from the number of edges directly above each subset in the diagram that there are 4 options for the second set on the path, 3 options for the third, 2 options for the fourth, and the path ends with . By Proposition 2.13, there are
such paths.
Now see if you can arrive at in a different way, by having a permutation of
determine the four edges of the path.
Exercise 3.33 Notice that each 2-combination of is represented by exactly
of the
ordered samples in Exercise 3.31, so the number of 2-combinations is
Exercise 3.34 Every subset of an n-element set contains either 0 elements, or 1 element, or 2 elements, …
Exercise 3.35 You will get the same partition by choosing to be a subset of S or its complement in S. And you had better not let
equal either S or its complement,
.
Exercise 3.36 The number of blocks can be only if all of the blocks contain a single element of S, except for one block containing two elements. In how many ways can you choose these two elements?
Chapter 4
Exercise 4.1 No. If you thought q was it, consider .
Exercise 4.2 After some experimentation, and deciding that “0 cents postage” doesn’t make much sense, we know that we can create the following values using only 10 and 4 cent stamps.
We can prove that all of these values are possible by considering numbers of the form and
.
We conjecture that this is a complete list. To prove this conjecture, assume to the contrary that there are counterexamples. By the Well-Ordering Principle there must be a minimal counterexample, m, which we know by our earlier computations must be at least 20. We also know that m must be even, since 4 and 10 are even. Because , it must be the case that
is an even integer that is at least 16. Thus
for some non-negative integers a and b. But then
, which contradicts the claim that m is the minimal counterexample.
Exercise 4.3 Finding an integer combination that equals 1 isn’t hard to do: is one, as is
. How many different possibilities are there? Finding a combination of 8 and 3 that equals 0 might help.
On the other hand, 8 and 6 are both even, so any integer combination of them must be even, too.
Exercise 4.4 The pairs are ,
,
,
,
,
,
,
; no pair shares a common integer divisor greater than 1. But
and
, as you might also recall from Section 2.9.2.
Exercise 4.5 There are only two integers relatively prime to 0, and 0 isn’t one of them.
Exercise 4.6
(a) False: For example, let and
.
(b) True: A proof very similar to the one given in Lemma 4.7 works here, although you’ll need to tweak Exercise 1.17.
(c) True: Again, a proof with the same overall logic as the one given in Lemma 4.7 works here. A few tweaks will be needed; and remember that if , then
for any integer c.
(d) False: For example, let ,
, and
.
Exercise 4.7 Yes. The “only if” part of the theorem doesn’t depend on whether a and b are positive, negative, or 0. And by Exercise 4.5, 0 is only relatively prime with 1 and . Since
and
both equal 1, the “if” part of the theorem is true as well.
Exercise 4.8 Yes. Again, the “only if” part of the theorem doesn’t depend on whether a and b are positive, negative, or 0; and Exercise 4.7 handles the situations when at least one of them equals 0.
We are thus left with two cases not yet handled by Theorem 4.8:
One possible way to handle each of these cases is to use induction on the quantity (which of course equals
when both a and b are positive). You will need to rearrange and alter some of the inequalities given in the proof, and refer to
and
instead of a and b; and for one of the cases, you may need to use
instead of
.
Another way to handle these cases is just to use Theorem 4.8 directly applied to and
, which are both positive. In any resulting linear combination, you can replace
with a and
with b, so long as you negate certain coefficients. You would then need only a lemma that says a and b are relatively prime if and only if
and
are relatively prime, regardless of their signs.
Good luck putting the pieces together!
Exercise 4.9 For certain values of a and b, it may be challenging to find integer combinations that equal 1 by trial and error. There is a method, called the Euclidean Algorithm, that works for any pair of integers; we hope that these exercises motivate you to research it!
(a) Since , we have
.
(b) Since , we have
.
(c) Notice that . Do you think it’s a coincidence that 13 is another Fibonacci number?
Exercise 4.10 Recall the structure of the proof of Proposition 2.13.
Exercise 4.11 There will be a quiz later!
Exercise 4.12 For the first part, you can simply let one of a and b be equal to 15 and the other 225; there are other possibilities as well. For the second part, notice that ; explain why this is problematic.
Exercise 4.13 For one approach, note that the product ab is a multiple of both a and b, so the smallest common multiple of a and b is in the finite set of natural numbers .
For a different approach, consider the set M of all common multiples of a and b, and notice that means that it is a non-empty subset of
.
Exercise 4.14 Let . Since d must divide both a and b, its prime divisors can only be among the primes
, so
, where each
is a non-negative integer. If
for each i, then certainly
and
. On the other hand, if some
, then at least one of
and
is true because
would be greater than one of
and
.
The argument for is similar.
Exercise 4.15 Using the notation in the preceding paragraph, the formulas for and
imply
Once you have explained why , you will know that
Exercise 4.16 Since , we have
Exercise 4.17 Good luck!
Exercise 4.18 There’s only one such set.
Exercise 4.19 Computations like
show that is closed under addition, subtraction, and multiplication. For division, consider
, where p is a prime that doesn’t divide n.
Exercise 4.20 The hardest step involves the computation
Chapter 5
Exercise 5.1 There are a lot of choices, including the floor and ceiling functions, which are defined later in this section if you have never seen them before. Or you could be boring and just do something like set for all
.
Exercise 5.2
(a) The number of characters in ω is the sum of the numbers of characters in and
.
(b) The alphabet for consists of only the characters 0 and 1, so
is the number of 0s plus the number of 1s.
Exercise 5.3 One possibility for (d) would be
As examples, and
. What is important for your function is that every ordered pair in
is sent to a unique and unambiguous polynomial in
. Thus the following would not define a function for (d), for a couple of reasons:
Exercise 5.4 The range of is
, and the range of ϕ is
. As for
, does any polynomial map to the string
consisting of n zeros for any
?
Exercise 5.5 Just two of the seven are not functions. In those cases, there exist elements of the domain whose images are not in the given codomains.
Exercise 5.6
(a) Let and
be any elements of (the domain)
. If
, then
, and squaring both sides gives
. Thus, f is injective.
(b) The proof of injectivity is essentially the same as the one given in part (a), with reciprocating in place of squaring.
(c) Note that , so h is not injective.
(d) Note that , so m is not injective.
Exercise 5.7
(a) If a is any element in (the codomain) A, then that same a (which is also in the domain A) satisfies , and thus
is surjective.
If and
are any elements of (the domain) A such that
, then
so is injective.
(b) The floor function is surjective because if n is an integer, then . It is not an injective function because
. Similar reasonings imply that the same results hold for the ceiling function.
(c) If and
are any elements of A such that the ordered pairs
and
are equal, then their first coordinates,
and
, must be equal, so f is injective.
If , then there are distinct elements
in A, and
is not satisfied for any
, so f is not surjective. However, if
or
… you can take it from here!
Exercise 5.8 For (b), perhaps show that is both injective and surjective. Does this example suggest a formula for a bijection from
to
for any real numbers
?
The same type of formula probably won’t work for (c). Maybe incorporate ?
Exercise 5.9 Since f is an injection, so is :
And , so
is surjective.
Exercise 5.10
(a) Each of the 3 elements in A can be paired with one of the 5 elements in B, so you can think of this as creating a sequence of length 3 with 5 options for each term. Perhaps Proposition 2.13 would be helpful?
(b) After an element of B has been paired with an element of A, it can not be paired with another element of A. Proposition 2.13 still seems like a good approach, or consult Section 3.8.
(c) None. Why?
(d) The answer to part (a) simply changes in value; the answer to part (b) changes entirely!
Exercise 5.11 It is meaningless, because the codomain of g and the domain of f are different.
Exercise 5.12
(a) g is not injective because, for example, . But
is injective: if
, then
and
both being positive means that
, which implies
. Plotting a graph of
will confirm this result.
(b) An example is shown in Figure 4.
Exercise 5.13 Continuing the hints given in the text …
(a) …there is a such that
. Since f is surjective, there is an
such that
. Thus,
, so
is surjective.
(b) …we know that . Since f is injective, we know that
. Thus,
is injective.
(c) There’s at most one sentence to write here!
Exercise 5.14 Yes: f must contain an ordered pair of the form , and nothing else.
Exercise 5.15 The function contains eight such elements: ω can be any element of the set
Exercise 5.16 A function is injective if and
imply that
. A function is surjective if for each
there is an
such that
.
Exercise 5.17 For the second step, if is not injective, there exist distinct elements
and an element
such that both
and
are in
. But
is a subset of f, so
and
are in f, too.
Exercise 5.18 To find all points of intersection, solve . This leads to
so the x-values are , 0, and 1; and the y-values are the same because these points must lie on the line
. Be sure that your two graphs are reflections of each other across this line.
Exercise 5.19 Using the logical equivalences discussed in Exercises 2.7 and 2.8 and the paragraph that follows them, we just need to show that Q (
) is logically equivalent to
. Try a truth table!
Exercise 5.20 To show that f is injective, suppose that for elements
. Then
, and so
, implying that
. To show that f is surjective, consider any
. Since
for some
, we have
A similar argument implies that g is bijective as well: just swap the symbols f and g, A and B, and a and b!
Finally, since f is bijective and thus the function exists, let’s show that
. Let
, and suppose that
and
. Then
from the given properties of f and g, and
from the definition of . Since f is injective, we must have
. Thus, g and
agree for all b in their common domain, B.
Exercise 5.21
(a) Since exactly when
, the pre-image of 0 is
.
(b) From the graph (or calculus) it is clear that the pre-image of 6 contains exactly one element. Since , the pre-image of 6 is
.
(c) We have when
; these are the x-values of the points where the local minima and maxima occur. Since
and
, each of the pre-images of
and
are 2-element sets.
Exercise 5.22
(a) Use the fact that , so
.
(b)
(c)
(d)
(e) Since the range of can only consist of non-negative integers, and since
for
,
as well.
Exercise 5.23 , and
.
Exercise 5.24 The graphs are best suggested using dotted lines with very fine spacing. We will see in Chapter 7 that some of the lines should have much finer spacing than others!
(a) and
.
(b) .
(c) , because there are rational and irrational numbers in the interval
.
(d) .
(e) Notice that , because
for each
. Thus,
.
(f) Since , we know that
, so
. Since S contains neither 0 nor 1, the answer is
.
Exercise 5.25
(a) Try finding a counterexample with ,
, and
.
(b) Let . Since f is surjective, there is an
with
. So we know that
, and this implies that
. Thus
, and with part (b) of Proposition 5.24 we conclude that
.
(c) Which is most helpful: part (a) or part (b) ?
Exercise 5.26 Try letting X and Y be disjoint.
Chapter 6
Exercise 6.1 The graph of A is a line, the graph of B is a closed half-plane, and the graph of C is shown in Figure 5.
Exercise 6.2 C is reflexive since for all
, and it is symmetric because
for all
. But it is not transitive: for example,
and
, but
.
B is reflexive and transitive; A is all three.
Exercise 6.3 Our coin flips asked us to find a relation that is reflexive and symmetric but not transitive. An example of such a relation is C of Example 6.2 .
Exercise 6.4 By the discussion in Section 5.5, any function f is a subset of , so it’s a relation! One such g is
.
Exercise 6.5 If and
, then we know there are natural numbers m and n such that
and
. By substitution this implies
and so
. Thus
, so
.
Exercise 6.6 Yes – for example, move 5 and 10 to the left of the rows they are on.
Exercise 6.7 It is a poset because is reflexive, antisymmetric, and transitive. For example, it is antisymmetric since any two rabbit pairs can’t both be descendants of each other, so
and
implies that x and y are the same pair.
The Hasse diagram for contains five rabbit pairs, with three minimal elements and the original pair at the top as the maximal element.
Exercise 6.8 This is an interesting infinite poset.
(a) , but
.
(b) The most complicated argument should be the one establishing antisymmetry. Try a proof by contradiction, assuming that and considering what is implied by
and
.
(c) Assume to the contrary that is a maximal element with
. Then consider
.
(d) See the suggestion for part (c) .
(e) is both maximal and minimal.
Exercise 6.9 The only one that isn’t is (c).
Exercise 6.10 The posets (b) and (d) are well ordered.
Exercise 6.11 We need to establish that has three properties.
Reflexive: Since is even for all
,
is reflexive.
Symmetric: If , then
for some
. But then
also equals
, so
.
Transitive: If and
, then
for some . Thus
, which is even. Therefore
as well.
Exercise 6.12 The proof of Proposition 6.11 can be adapted to prove this is an equivalence relation, and again we point you to the more general result in Exercise 6.57.
Exercise 6.13 To answer part (c), notice that an equivalence class corresponds to the set of x-values where a horizontal line intersects the graph of sine. If two x-values correspond to the same horizontal line, their equivalence classes are the same; and if they correspond to different horizontal lines, the two equivalence classes have no x-values in common (because sine is a function). The dotted horizontal line in Figure 6 intersects the graph of sine in the equivalence class of , which may be helpful for part (b).
Exercise 6.14 It’s equality.
Exercise 6.15 Look at Theorem 6.14 and Exercise 3.28.
Exercise 6.16 The equivalence classes modulo 2 are the evens and odds. The equivalence classes modulo 3 are:
Exercise 6.17 We ask you …
Exercise 6.18 …to complete these three exercises …
Exercise 6.19 …without any additional hints from us.
Exercise 6.20 Each element of is an equivalence class that must be mapped to a unique element of
.
Exercise 6.21 The formulas follow from the fact that and
in
.
Exercise 6.22 The multiplication table for is shown in Figure 7. Notice that we have
in
, which is directly related to the factorization
.
Exercise 6.23 Recall the proof of Proposition 2.16 and the discussion preceding it.
Exercise 6.24 The last digit is a 9. Here’s a slightly more difficult computation: What’s the last digit of ?
Exercise 6.25 The argument given for would fail when applied to the proposed squaring function from
to
. If
, then
for some
. So
Since and
, we can’t reduce
to zero as we could when we were defining the function from
to
. In particular, we cannot make a statement similar to “Since 3 divides 12 and 36 we know
and
.”
Exercise 6.26 Here is a proof of part (d).
If in
, then
for some
. Therefore
Thus , and so
if
.
If you would like an additional challenge, find values of m and n where this squaring function from to
is not surjective.
Exercise 6.27 You can prove this by mimicking the proof of Theorem 6.19, or you can use an argument inspired by .
Exercise 6.28 , so 4 plays the role of “1/4.”
Exercise 6.29 Construct the rows corresponding to multiplication by 2 and 10 in order to verify that [2] and [10] do not have inverses. On the other hand, , so
is the multiplicative inverse of
and
is the multiplicative inverse of
.
Exercise 6.30 If and
, then
. This means that
, with
. What can you conclude about m and
?
Exercise 6.31 Figure 8 should get you started, if you are truly stuck.
Exercise 6.32 Search for hints among the various facts we have proven about relatively prime integers.
Chapter 7
Exercise 7.1 Suppose that n new guests arrive. Hilbert should announce over the intercom to everyone who already has a room: “I again apologize for the inconvenience, but in order for us to accommodate some new guests, we need you to change rooms. If you are in room #k, please move to room # for the night. Thank you!” He can then send the new arrivals to rooms #1 through #n.
Exercise 7.2 Hilbert can first move every current guest into an even-numbered room by announcing: “I once again apologize for the inconvenience, but in order for us to accommodate some new guests, we need you to change rooms. If you are in room #k, please move to room # for the night. Thank you!” He then tells the new guests to move into the odd-numbered rooms, with new guest #n sent to room #
.
Exercise 7.3 None! Cookie #n will be eaten in round n, and that accounts for all of them.
Exercise 7.4 The answers are:
(a) Cookies #1 through #9 will be on his plate.
(b) The even-numbered cookies will be on his plate.
If you found this to be an interesting exercise, or if you found it difficult, then you should also work on Exercise 7.22 at the end of the chapter.
Exercise 7.5 If you have worked earnestly on this question for more than two hours and are stuck, then perhaps you should stop and begin reading the remainder of this section and the next, where almost all is revealed.
Exercise 7.6 We recommend using the fact that all of these functions have inverses: the natural logarithm ,
, and
.
Exercise 7.7 Constructing an explicit function might help you prove that this is a bijection. Using in your formula will allow you to change signs back and forth between positive and negative, and the function
is also handy, as for example
.
Exercise 7.8 We already gave you a nice hint!
Exercise 7.9 For part (a), by shifting his current guests from room #n to room #, Hilbert has opened up room #1 for a frog. It’s quite generous of him to go to all of this trouble for a frog.
For part (b), you might wish to view the partial table presented in this section as one corner of a vast parking lot.
Exercise 7.10 Again, we already gave you a nice hint!
Exercise 7.11 Since is an infinite subset of
, use Lemma 7.7.
Exercise 7.12 because
. There are surely simpler examples than this one.
Exercise 7.13 Randomly selecting the image of each , we created the injective function
defined by
The resulting set is , which is not in the range of f.
Exercise 7.14 Just change “ is countable” to “
,” and change all occurrences of
to A. And change n to x for style points, if you think n should only represent integers.
Exercise 7.15 We are not uncomfortable with the original argument, so we skipped this exercise.
Exercise 7.16 We know that . If we assume to the contrary that
is a countable set, then we have expressed
as the union of two countable sets. Lemma 7.10 would then imply that
must be countable, which contradicts Theorem 7.15.
Exercise 7.17 The function defined by
is a bijection, so . In the previous paragraph we applied Schröder–Bernstein to prove
. Thus, by the transitivity of
we will be finished if we can show that
. We can do this by considering containment as the injection
and
as the injection
, and then appealing to Schröder–Bernstein.
Exercise 7.18 First prove that h is a surjection. Then argue that it is sufficient to prove that h is injective when restricted to X and when restricted to .
Exercise 7.19 You could try to do this by mimicking the proof of Lemma 7.20, or you could instead quote results we have already established: has the same cardinality as the open interval
, which leads to
having the same cardinality as
, and then you can apply Lemma 7.20.
Exercise 7.20 In fact we only need Lemmas 7.21 and 7.22 to prove
Chapter 8
Exercise 8.1 One lower bound is . Any
is also a lower bound, because if
, then
. Similarly, an upper bound is 2, and any
will serve as an upper bound. Figure 9 might help to clarify things.
Exercise 8.2 Every real number r is an upper bound for the empty set, since for all
: there are no such s! Similarly, every real number is a lower bound.
Since every element of is an upper bound, there is certainly no least one.
Exercise 8.3 A lower bound is 2 and an upper bound is 3. The greatest lower bound is 2.7, and the least upper bound is e. For a proof that e is irrational, see Section 8.6.2.
Exercise 8.4 The set of all lower bounds for A is , so the greatest lower bound is
. The set of all upper bounds for A is
, so the least upper bound is 1.
Exercise 8.5 Corollary 8.4 guarantees the existence of an such that
. There is a largest power of 10 within the non-empty finite set
; suppose it is
for some
. Then
, so
.
Exercise 8.6 Assume to the contrary that is a lower bound for
. Since r is negative and
, we know that
for all . Since r is negative, dividing by r changes the order of the inequalities, giving
for all . Thus
for all . Since
, we know that
and thus
and
are positive real numbers, so we have a contradiction to Corollary 8.3.
Exercise 8.7 The terms of the first two sequences are heading toward a limit of 0. The terms of the third sequence aren’t heading toward any finite value, so the limit doesn’t exist, or you might say that it’s “”; see Exercise 8.27. The fourth sequence does have a limit, but it’s not rational. Write the first few terms of the sequence in decimal notation and see if that helps with your guess.
Exercise 8.8 Parts of the definition are matched with phrases of the sentence: “Regardless of how close [] you want the sequence terms [
] to be to the limiting value [L], you will get your wish [
] so long as you ignore an appropriate number [
] of the initial terms of the sequence.” And order is important here: the value of N depends on the tolerance
. You might wish to write
in place of N, if doing so helps you remember this order.
Exercise 8.9 The same argument that worked for works for
, requiring the solution of
at the end. Now, sketch a (wide!) extension of Figure 1 on page 190 showing that the terms of the sequence eventually stay inside of a thinner band around
.
Exercise 8.10 Since the limit is , we need to show that for every
there is an
such that
for all natural numbers
. Since
for , we can appeal to the first proof given in Theorem 8.5 to conclude that there is an
such that
. Since
, for any natural number
we have
Exercise 8.11 For the first, it’s
For the second, it’s
Exercise 8.12 If , then the sequence of partial sums is simply
, so the series converges to 0. Otherwise, if
and
, the series doesn’t converge. For example, if
, then the series is
and the sequence of partial sums is
, which has no finite limit.
Showing that a geometric series doesn’t converge in other cases where and
requires other arguments, some of which might be assisted by the Comparison Test of Section 8.5.
Exercise 8.13 It is a constant sequence: all of the terms must be equal.
Exercise 8.14
(a) The sequence is monotone and bounded.
(b) The sequence is bounded but not monotone. The sequence presented in Figure 1 (page 190) is another example, and also shows that a sequence need not be monotone to have a limit.
(c) The sequence is monotone but not bounded.
(d) The sequence is neither monotone nor bounded.
Exercise 8.15 Change μ to , “increasing” to “decreasing,” “least upper bound” to “greatest lower bound,” and so on. Also reverse most of the inequalities and change some – signs to
; but
should remain positive!
Exercise 8.16 You can show that the series converges, and find crude upper and lower bounds for the value of the series, by first comparing it with the geometric series , and then with
.
Exercise 8.17 First show that for all
; induction will work, as will a direct argument. This implies that
for all , and the
terms are both 1, so the Comparison Test applies.
Exercise 8.18 Integration by parts yields
where the final term can be replaced via
whose final term can be replaced via
and so on, until a final term is 0 because .
When , the resulting expression is
which evaluates to
as desired.
Exercise 8.19 Among the results you might need are that
when is non-negative and bounded above by M, and that
for any real number c.
Chapter 9
Exercise 9.1 There are eight equally likely outcomes:
Of these eight, three contain exactly two heads: HHT, HTH, and THH. So you might expect this result to occur about three-eighths of the time, if you repeat the activity many, many times.
Exercise 9.2 List the 16 equally likely outcomes and tally the ones you want, as in the previous exercise.
Exercise 9.3 Out of 250 flips, there are 124 heads and 126 tails. Out of all 249 2-flip strings, there are 59 HHs, 64 HTs, 64 THs, and 62 TTs. If you only count 125 2-flip strings starting from flips 1, 3, 5, …, 249, you get 29 HHs, 34 HTs, 32 THs, and 30 TTs.
What does this evidence suggest, if anything?
Exercise 9.4 Only the denominator changes: and
. Since
, the ball is slightly more likely to land on red on a European wheel.
Exercise 9.5 You can determine the numerators of these fractions…
Exercise 9.6 …by counting carefully in Figure 2 (page 212).
Exercise 9.7 Is it close to 1/8?
Exercise 9.8 Symmetry follows from
For unimodality, use the fact that
is true so long as , and so when
.
Exercise 9.9 is a 158-digit integer beginning
and ending with 24 0s.
The real number is approximately
in scientific notation.
Exercise 9.10 The general statement is that, for non-negative integers n,
Mathematical induction seems like an appropriate way to prove this, using the identity in the inductive step along with Theorem 9.8.
Exercise 9.11 Again, count carefully in Figure 2 (page 212).
Exercise 9.12 The values of the binomial coefficients are
The sum of these numbers is 100 146 724, which upon dividing by yields
Now, what’s your conclusion?
Exercise 9.13 If you’re not close, keep rolling!
Exercise 9.14 If X counts the points you earn on a die roll, then
Exercise 9.15 If A is the event that at least one 6 is seen, then is the event that no 6s are seen. It is easier to compute
: by Proposition 2.13,
, so
and thus This number is larger than
.
Exercise 9.16 We know that and
, so we are left to determine
. Since
and
we find that
Thus, .
Exercise 9.17 It was, in fact, faked.
Chapter 10
Exercise 10.1 If and
are two points in
, then the distance between them is
which is equal to
the distance between and
.
Now just see where the two corners go.
Exercise 10.2 It is probably easiest to note that , which can be applied twice for
and three times for
:
Exercise 10.3 and
.
Exercise 10.4 You are interested in , so you first perform
and then
. Since
and
the composition must be .
Exercise 10.5
(a) In order for a subset to be an identity element, it must be the case that
for all
. The empty set can function as an identity element, as
for all . Let A be a non-empty subset of S. The set A cannot serve as the identity element as we would need it to be true that
, but
.
(b) Since S is non-empty, the set S cannot have an inverse with respect to . Such a subset A would have to satisfy
, but
since
.
(c) We’ll let you do these last two parts …
(d) …on your own.
Exercise 10.6 Here are three very strong hints. You can prove that associativity holds by appealing to the fact that it held for G and H; the identity element is the ordered pair of identity elements ; and the inverse of
is
.
Exercise 10.7 Table 1 on the next page is a partially completed Cayley table for . In order to avoid cumbersome notation we express
as
.
Table 1 Part of the Cayley table for .
![]() |
![]() |
![]() |
![]() |
![]() |
![]() |
![]() |
![]() |
|
![]() |
![]() |
![]() |
![]() |
![]() |
![]() |
![]() |
![]() |
![]() |
![]() |
![]() |
![]() |
? | ? | ![]() |
? | ? | ? |
![]() |
![]() |
? | ![]() |
? | ? | ? | ? | ? |
![]() |
![]() |
? | ? | ? | ? | ? | ? | ? |
![]() |
![]() |
![]() |
? | ? | ? | ? | ? | ? |
![]() |
![]() |
? | ? | ? | ? | ? | ![]() |
![]() |
![]() |
![]() |
? | ? | ? | ? | ![]() |
? | ? |
![]() |
![]() |
? | ? | ? | ? | ![]() |
? | ? |
Exercise 10.8 You might do this by first proving that a symmetry of is determined by where it sends any three vertices on a face of
. Then prove that where these three vertices go is determined by where any one of them goes using the fact that the box is irregular.
Exercise 10.9 In Theorem 10.11 you proved that is closed under composition, so you know that
. By tracing out the destinations of the corners you will see that α takes a corner to its opposite corner of the box, which the rotations and reflections cannot do. This map is known as the antipodal map, and it has a simple formula:
.
Exercise 10.10 In doing our work we noticed that for each
.
Exercise 10.11
(a) To show that this is a group, you need to mention that multiplication is associative, 1 is the identity element, and as long as then
. To show it is an Abelian group, you need to highlight that multiplication of rational numbers is a commutative operation.
(b) This would be similar to part (a), except that 2 does not have a multiplicative inverse within , so it’s not even a group!
Exercise 10.12 An examination of Table 2 (page 237) shows that is Abelian. Can you see why a certain form of symmetry in the table implies that the product is commutative? If so, write up your proof by first proving and then using this insight.
Exercise 10.13 The proof of Proposition 10.21 shows that is a cyclic generator in
. The element
is as well, since
Similar computations will show that every non-zero element of is a cyclic generator.
Exercise 10.14 Gauss characterized when is a cyclic group:
For example, is cyclic while
is not.
Exercise 10.15
(a) . (You may want to look up your answer to Exercise 3.23 if you think the answer is different.)
(b) .
(c) .
Exercise 10.16 If you are stuck after a couple of hours …
Exercise 10.17 …then you might want to pause until you get to Section 10.8.
Exercise 10.18 Here’s a proof for part (a). Since we know that for all
, we know that
. Assume by induction that
for an arbitrary
. Then
so this part follows by induction.
Lemma 10.27 establishes (b). An argument like that given above for (a), combined with Lemma 10.28, can be used to prove (c).
Exercise 10.19 There might be some smaller value with
. For example, in
we have
, so
. However it is also the case that
, so
; and in fact
in
.
Exercise 10.20 Let be an isomorphism and assume that G is Abelian. Let
and
be any two elements of H. Since f is a bijection, there is a unique pair
and
in G such that
and
. Thus
. By applying the definition of isomorphism twice and appealing to our assumption that G is Abelian, we find that
Thus H is Abelian.
Similarly, if H is Abelian, we may apply the argument above using the isomorphism to prove that G is Abelian.
Exercise 10.21 Try two adjacent reflections, those whose axes of reflection form an angle of 30. Their product will be a rotation through 60
, but the order in which you multiply them will determine if it is a clockwise or counterclockwise rotation.