Skip to main content
\( \def\negchoose#1#2{\genfrac{[}{]}{0pt}{}{#1}{#2}_{-1}} \newcommand{\mchoose}[2]{\left(\!\binom{#1}{#2}\!\right)} \newcommand{\cycle}[1]{\arraycolsep 5 pt \left(\begin{array}#1\end{array}\right)} \newcommand{\importantarrow}{\Rightarrow} \newcommand{\qchoose}[2]{\left[{#1\atop#2}\right]_q} \newcommand{\bp}{ \begin{enumerate}{\setcounter{enumi}{\value{problemnumber}}}} \newcommand{\ep}{\setcounter{problemnumber}{\value{enumi}} \end{enumerate}} \newcommand{\ignore}[1]{} \renewcommand{\bottomfraction}{.8} \renewcommand{\topfraction}{.8} \newcommand{\apple}{\text{🍎}} \newcommand{\ap}{\apple} \newcommand{\banana}{\text{🍌}} \newcommand{\ba}{\banana} \newcommand{\pear}{\text{🍐}} \newcommand{\pe}{\pear} \DeclareMathOperator{\Fix}{Fix} \DeclareMathOperator{\Orb}{Orb} \newcommand{\F}{\mathcal{F}} \newcommand{\alert}{\fbox} \def\d{\displaystyle} \def\course{Math 228} \newcommand{\f}[1]{\mathfrak #1} \newcommand{\s}[1]{\mathscr #1} \def\N{\mathbb N} \def\B{\mathbf{B}} \def\circleA{(-.5,0) circle (1)} \def\Z{\mathbb Z} \def\circleAlabel{(-1.5,.6) node[above]{$A$}} \def\Q{\mathbb Q} \def\circleB{(.5,0) circle (1)} \def\R{\mathbb R} \def\circleBlabel{(1.5,.6) node[above]{$B$}} \def\C{\mathbb C} \def\circleC{(0,-1) circle (1)} \def\F{\mathbb F} \def\circleClabel{(.5,-2) node[right]{$C$}} \def\A{\mathbb A} \def\twosetbox{(-2,-1.5) rectangle (2,1.5)} \def\X{\mathbb X} \def\threesetbox{(-2,-2.5) rectangle (2,1.5)} \def\E{\mathbb E} \def\O{\mathbb O} \def\U{\mathcal U} \def\pow{\mathcal P} \def\inv{^{-1}} \def\nrml{\triangleleft} \def\st{:} \def\~{\widetilde} \def\rem{\mathcal R} \def\sigalg{$\sigma$-algebra } \def\Gal{\mbox{Gal}} \def\iff{\leftrightarrow} \def\Iff{\Leftrightarrow} \def\land{\wedge} \def\And{\bigwedge} \def\entry{\entry} \def\AAnd{\d\bigwedge\mkern-18mu\bigwedge} \def\Vee{\bigvee} \def\VVee{\d\Vee\mkern-18mu\Vee} \def\imp{\rightarrow} \def\Imp{\Rightarrow} \def\Fi{\Leftarrow} \def\var{\mbox{var}} \def\Th{\mbox{Th}} \def\entry{\entry} \def\sat{\mbox{Sat}} \def\con{\mbox{Con}} \def\iffmodels{\bmodels\models} \def\dbland{\bigwedge \!\!\bigwedge} \def\dom{\mbox{dom}} \def\rng{\mbox{range}} \def\isom{\cong} \DeclareMathOperator{\wgt}{wgt} \newcommand{\vtx}[2]{node[fill,circle,inner sep=0pt, minimum size=4pt,label=#1:#2]{}} \newcommand{\va}[1]{\vtx{above}{#1}} \newcommand{\vb}[1]{\vtx{below}{#1}} \newcommand{\vr}[1]{\vtx{right}{#1}} \newcommand{\vl}[1]{\vtx{left}{#1}} \renewcommand{\v}{\vtx{above}{}} \def\circleA{(-.5,0) circle (1)} \def\circleAlabel{(-1.5,.6) node[above]{$A$}} \def\circleB{(.5,0) circle (1)} \def\circleBlabel{(1.5,.6) node[above]{$B$}} \def\circleC{(0,-1) circle (1)} \def\circleClabel{(.5,-2) node[right]{$C$}} \def\twosetbox{(-2,-1.4) rectangle (2,1.4)} \def\threesetbox{(-2.5,-2.4) rectangle (2.5,1.4)} \def\ansfilename{practice-answers} \def\shadowprops{{fill=black!50,shadow xshift=0.5ex,shadow yshift=0.5ex,path fading={circle with fuzzy edge 10 percent}}} \newcommand{\hexbox}[3]{ \def\x{-cos{30}*\r*#1+cos{30}*#2*\r*2} \def\y{-\r*#1-sin{30}*\r*#1} \draw (\x,\y) +(90:\r) -- +(30:\r) -- +(-30:\r) -- +(-90:\r) -- +(-150:\r) -- +(150:\r) -- cycle; \draw (\x,\y) node{#3}; } \newcommand{\card}[1]{\left| #1 \right|} \newcommand{\twoline}[2]{\begin{pmatrix}#1 \\ #2 \end{pmatrix}} \newcommand{\lt}{<} \newcommand{\gt}{>} \newcommand{\amp}{&} \)

Section2.2Proofs in Combinatorics

We have already seen some basic proof techniques when we considered graph theory: direct proofs, proof by contrapositive, proof by contradiction, and proof by induction. In this section, we will consider a few proof techniques particular to combinatorics. Along the way we will establish some basic counting principles and further develop our understanding of binomial coefficients.

Why do we care about proofs? It is not because we doubt the truth of so called theorems or wish to establish new results to further mathematics as an academic discipline (although there is nothing wrong with doing that as well). Rather, it is through creating proofs that we gain a deeper understanding of the big ideas present in this subject.

Subsection2.2.1Two Motivating Examples

To get started, let's consider two typical statements in combinatorics which we might wish to prove.

These are just two of the beautiful identities hidden in Pascal's triangle. But why are they true?

There are a at least four very distinct proofs we could give of each binomial identity. The next activities walk you through these.

Note: throughout this section especially, hints to problems are available. Try the problem first, and if you get stuck, peek at the hint.

First, let's try some proofs of Theorem 2.2.1

Activity72

First, an algebraic proof. We will see later that \(\binom{n}{k} = \frac{n!}{k!(n-k)!}\text{,}\) where \(r! = r \cdot (r-1) \cdot (r-2) \cdot\cdots\cdot 2\cdot 1\) (read “\(r\) factorial”). Using this algebraic formula, prove the identity.

Hint

This should be really easy; don't let that fool you. Note that a correct proof must start with one side of the identity and manipulate it into the other side. The following is NOT a correct proof, because it starts by assuming the statement you wish to prove.

\begin{align*} \binom{n}{k} = \amp \binom{n}{n-k}\\ \frac{n!}{k!(n-k)!} = \amp \frac{n!}{(n-k)!(n-(n-k)!)}\\ \frac{n!}{k!(n-k)!} = \amp \frac{n!}{(n-k)!k!)}\\ \frac{n!}{k!(n-k)!} \cdot (k!(n-k)!)= \amp \frac{n!}{(n-k)!(n-(n-k)!)}\cdot(k!(n-k)!)\\ n! = \amp n! \end{align*}

Note that this proof was not by any means difficult, but it did rely on an algebraic identity we have yet to prove. We will see in Activity 76 that algebraic proofs can be very messy. What's more, nothing about these proofs explain why the identity is true. They lack meaning.

Activity73

We claim the identity is true for all \(n \ge 0\text{,}\) so induction would be a natural proof technique to try. Give a proof by mathematical induction on \(n\text{.}\)

Hint

You will probably want to use the Pascal recurrence \(\binom{n}{k} = \binom{n-1}{k-1} + \binom{n-1}{k}\text{.}\) Note that once you pick your \(n\text{,}\) you should argue that the identity holds for all \(0 \le k \le n\text{.}\) You should not do double induction.

Activity74

Think about what \(\binom{n}{k}\) counts.

(a)

Suppose you own \(n\) bow ties, and want to choose \(k\) of them to take on a trip. How many different collections of bow ties can you take?

Hint

This is supposed to be very easy.

(b)

Why is the answer to the above question also \(\binom{n}{n-k}\text{?}\)

Hint

What if you started with all bow ties in the suitcase and got rid of some? How many do you need to remove?

(c)

Why are the two parts above enough to establish the identity?

Finally, here is a subtly different proof.

Activity75

Consider the set \([n] = \{1,2,\ldots,n\}\) and its power set \(\pow([n])\) of all subsets of \([n]\text{.}\) Define the function \(f:\pow([n]) \to \pow([n])\) by \(f(A) = [n]\setminus A\) for any \(A \in \pow([n])\) (that is, \(f(A)\) is the complement of \(A\) in \([n]\)).

(a)

Prove that \(f\) is a bijection. In fact, \(f\) is an involution in that it is its own inverse.

(b)

For any set \(A \in \pow([n])\text{,}\) if \(\card{A} = k\text{,}\) what is the \(\card{f(A)}\text{?}\)

(c)

Is \(f\) still a bijection if we restrict it to just the subsets of \([n]\) that have size \(k\text{?}\) Perhaps we should also restrict our codomain (to what?). Explain.

(d)

Why is the above enough to establish the identity?

Hint

Recall the bijection principle tells us that if \(f:X \to Y\) is a bijection, then \(\card{X} = \card{Y}\text{.}\) What is the cardinality of each of the subsets of \(\pow([n])\) that we are considering?

How are Activity 74 and Activity 75 different? The first consists of answering a single counting question in two different ways, and those ways are the two sides of the identity. Such proofs are sometimes called double counting proofs, or sometimes just combinatorial proofs. On the other hand, Activity 75 proceeds by showing two different sets have the same size, using a bijection, and shows that these two sets are counted by each side of the identity. These proofs are called bijective proofs (and are also sometimes grouped together with double counting proofs as combinatorial proofs).

Both styles of combinatorial proof have the advantage that they do an excellent job of illustrating what is really going on in an identity. They reinforce our understanding of how to solve counting problems and our understanding of basic combinatorial principles.

This is not to say that other proof techniques, especially mathematical induction, are not also helpful. Inductive proofs demonstrate the importance of the recursive nature of combinatorics. Even if we didn't know what Pascal's triangle told us about the real world, we would see that the identity was true entirely based on the recursive definition of its entries.

Now here are four proofs of Theorem 2.2.2.

Activity76

Using the formula \(\binom{n}{k} = \frac{n!}{k!(n-k!)}\text{,}\) you should be able to find a common denominator in the sum \(\sum_{k=0}^n \binom{n}{k}\) and show that this simplifies to \(2^n\text{.}\)

Hint

We say “should”, but please don't waste your time doing this all the way out. Unless you are really bored.

Activity77

We wish to establish this identity for all natural numbers \(n\text{,}\) so it would be natural to give a proof by induction. Do this.

Hint

Again, use the Pascal recurrence \(\binom{n}{k} = \binom{n-1}{k-1} + \binom{n-1}{k}\text{.}\) Doing this for all summands will give you some repeats.

There are lots of ways to give a combinatorial proof. Here is one.

Activity78

Consider the question: how many subsets of \([n]\) are there?

(a)

How many subsets have cardinality 0? How many have cardinality 1? And so on. How many would all of these give us all together?

Hint

Your final answer here should be a sum, since we have counted the number of subsets partitioned into disjoint collections.

(b)

Explain why the answer is also \(2^n\text{.}\)

Hint

Think about how many choices you have for each element. You could put 1 in the subset or not. You could include \(2\) in the subset or not.

Conclude: Since each side of the identity is the answer to the same counting question, we have established the identity.

We will not give a bijective proof, although you might include this as part of the above if you didn't see why \(2^n\) was the number of subsets of \([n]\text{.}\) Instead, you might see directly that \(2^n\) is the number of \(n\)-bit strings (by writing the numbers 0 through \(2^n - 1\) in binary) and then define a bijection from \(n\)-bit strings to subsets of \([n]\text{.}\)

Finally, here is a clever proof of the identity.

Activity79

What does the binomial theorem say again? That is, expand \((x+y)^n\text{.}\) What happens when you substitute \(x = y = 1\text{?}\)

Hint

Just do it.

This use of the binomial theorem is an example of one of the many uses for generating functions which we will return to later. For now, you might enjoy plugging in other values to the binomial theorem to uncover new binomial identities.

Subsection2.2.2Interlude: The Sum and Product Principles

Any proof you write in mathematics must assume some foundational principles. In Euclidean geometry, you have axioms, common notions, and often other propositions, used as justification in your proof. Let's pause to consider two foundational principles in combinatorics that, while intuitively obvious, are worth pointing out as foundational.

Activity80
(a)

A restaurant offers 8 appetizers and 14 entrées. How many choices do you have if:

  1. you will eat one dish, either an appetizer or an entrée?
  2. you are extra hungry and want to eat both an appetizer and an entrée?
(b)

Think about the methods you used to solve question (a). Write down the rules for these methods.

(c)

Do your rules work? A standard deck of playing cards has 26 red cards and 12 face cards.

  1. How many ways can you select a card which is either red or a face card?

  2. How many ways can you select a card which is both red and a face card?

  3. How many ways can you select two cards so that the first one is red and the second one is a face card?

When solving a counting question, it is often advantageous to break the problem into parts. If you think of the counting question as asking how many ways there are to complete a task, then you might break that task up in various ways, count the number of ways to complete each sub-task, and them combine those somehow to get the number of ways to complete the overall task. How do you combine them?

Sum Principle

The sum principle states that if task \(A\) can be completed in \(m\) ways, and task \(B\) can be completed in \(n\) disjoint ways, then completing “\(A\) or \(B\)” is a task that can be completed in \(m + n\) ways.

Product Principle

The product principle states that if task \(A\) can be completed in \(m\) ways, and each possibility for \(A\) allows for exactly \(n\) ways for task \(B\) to be completed, then the task “\(A\) and \(B\)” (or “\(A\) followed by \(B\)”) can be completed in \(m \cdot n\) ways.

Note, while both the sum and product principle are true for more than two subtasks, which you could prove easily by induction from the versions given above. But how do we know these principles are true?

To make these principles more mathematically rigorous, consider operations on sets.

This suggests that the union of (disjoint) sets somehow corresponds to the disjunction (or) of completing tasks. You might guess that the set operation that corresponds to conjunction (and) of tasks is set intersection. But this cannot be correct: the intersection of two sets is no larger than the smallest of the two sets, while combining tasks with conjunction apparently corresponds to multiplication.

A better way to think of the conjunction is as concatenation. You are performing one task, followed by the other. Recall that \(A \times B = \{(a,b) \st a \in A, b \in B\}\) is the Cartesian product of \(A\) and \(B\text{;}\) the set of all ordered pairs where the first element comes from \(A\) and the second from \(B\text{.}\) This is exactly what we need.

We could prove these two propositions without too much trouble. Then we must explain what they have to do with the sum and product principle.

Activity81
(b)

An incomplete deck of cards only has four face cards (the Jack, King and Queen of diamonds and the King of hearts) and five black cards (the 2 through 6 of spades). A magician asks you to pick a card. How many choices do you have?

(c)

How exactly does the counting question above relate to Proposition 2.2.3? How exactly does it relate to the Sum Principle? Illustrate both by listing specific outcomes.

(d)

Make explicit the connection between the Sum Principle and Proposition 2.2.3. That is, justify the Sum Principle in terms of cardinality of sets.

Activity82
(b)

Using the same incomplete deck of cards (containing the Jack, King and Queen of diamonds, King of hearts, and 2 through 6 of spades), the magician asks you to put any face card face down under any face down black card. How many choices do you have for this two card stack?

(c)

How exactly does the counting question above relate to Proposition 2.2.4? How exactly does it relate to the Product Principle? Illustrate both by listing specific outcomes.

(d)

Make explicit the connection between the Product Principle and Proposition 2.2.4. That is, justify the Product Principle in terms of cardinality of sets.

The previous activity might obscure an important subtlety. Perhaps you said that the set \(A\) corresponds to the face cards (or to the set of outcomes of the task of selecting a face card) and the set \(B\) corresponded to the black cards. The set of outcomes of picking a two card stack is exactly then \(A \times B\text{,}\) so there are \(4\cdot 5 = 20\) outcomes. This is completely correct. But does it generalize?

Activity83

Suppose the magician hands you 4 cards (the Ace, King, Queen, and Jack of clubs) and asks you to select any two cards (placing them face down in a stack). How many choices do you have? Assume the order the cards are placed down matters.

(a)

If you divide the task into two sub-tasks, what are they? How many ways can you complete the first task? How many ways can you complete the second? What is the final answer?

(b)

If you think of this problem as counting a Cartesian product, what is set \(A\text{?}\) What is set \(B\text{?}\) What is wrong?

The Product Principle is not exactly analogous to finding the cardinality of a Cartesian product after all, since tasks \(A\) and \(B\) do not need to be disjoint. How do we make sense of the Product Principle then?

Activity84
(a)

Write out all stacks of cards that are possible outcomes in Activity 83. Arrange these into a rectangle of appropriate dimensions.

(b)

How might you group the outcomes in a meaningful way? How many groups and how many outcomes in each group?

(c)

Find some sets \(A\) and \(B\) and a bijection between the outcomes (stacks of 2 cards) and the set \(A \times B\text{.}\) Describe how this would work in general.

To illustrate just how fundamental these two principles are, here are a few counting problems that can be solved with them. For each of these, say exactly where and how the sum and product principles are used.

Activity85

We have seen that the number of lattice paths from \((0,0)\) to \((x,y)\) is given by \(\binom{x+y}{x}\text{.}\) Now with the sum and product principles, we can answer lots of questions about lattice paths.

(a)

How many lattice paths from \((0,0)\) to \((10,10)\) pass through the point \((4,7)\text{?}\)

Hint

Many students want this sort of question to be an example of the sum principle (presumably because to get an acceptable path you travel to \((4,7)\) and then add on a path that gets you the rest of the way. But notice that you are really concatenating two paths, and there are lots of choices for the first path and lots for the second. This sounds more like an ordered pair.

(b)

How many lattice paths from \((0,0)\) to \((10,10)\) do NOT pass through the point \((4,7)\text{?}\) In what way does this problem use the sum principle?

Hint

If you knew the answer to this part and the previous part, what is another way you could compute the number of paths from \((0,0)\) to \((10,10)\text{?}\)

(c)

How many lattice paths from \((0,0)\) to \((10,10)\) pass through \((4,7)\) or \((7,4)\text{?}\)

(d)

Explain why can you NOT use the sum principle to answer this question: How many lattice paths from \((0,0)\) to \((10,10)\) pass through \((4,7)\) or \((7,8)\text{?}\) What is the answer?

Hint

Compare to the question about how many playing cards are either red or a face card, and why the answer is not \(26 + 12\text{.}\)

The last part of the previous activity illustrates what can go wrong if the sets \(A\) and \(B\) are not disjoint. In general, we have,

\begin{equation*} \card{A \cup B} = \card{A} + \card{B} - \card{A \cap B}. \end{equation*}

This is an example of the Principle of Inclusion/Exclusion, which we will investigate further later.

This next set of questions you can answer entirely using the product principle. While all the activities ask questions with a general \(n\) and \(k\text{,}\) it would be a good idea to first try to answer the corresponding question with specific numbers and then generalize.

Activity86

A tennis club has \(2n\) members. We want to pair up the members by twos for singles matches.

(a)

In how many ways may we pair up all the members of the club? (Hint: consider the cases of 2, 4, and 6 members.)

Hint

Suppose you have a list in alphabetical order of names of the members of the club. In how many ways can you pair up the first person on the list? In how many ways can you pair up the next person who isn't already paired up?

(b)

Suppose that in addition to specifying who plays whom, for each pairing we say who serves first. Now in how many ways may we specify our pairs?

Activity87

A roller coaster car has \(n\) rows of seats, each of which has room for two people. If \(n\) men and \(n\) women get into the car with a man and a woman in each row, in how many ways may they choose their seats?

Hint

In how many ways may you assign the men to their rows? The women? Once a woman and a man have a row to share, in how many ways may they choose their seats?

Activity88

In how many ways can we pass out \(k\) distinct pieces of fruit to \(n\) children (with no restriction on how many pieces of fruit a child may get)?

Activity89

How many subsets does a set \(S\) with \(n\) elements have?

Hint

Try applying the product principle in the case \(n = 2\) and \(n = 3\text{.}\) How might you apply it in general?

Activity90

Assuming \(k\le n\text{,}\) in how many ways can we pass out \(k\) distinct pieces of fruit to \(n\) children if each child may get at most one? What is the number if \(k>n\text{?}\) Assume for both questions that we pass out all the fruit.

Hint1

Ask yourself if either the sum principle or product principle applies.

Hint2

Remember that zero is a number.

The last activity is an example of a permutation. We are asking how to arrange \(k\) of the \(n\) kids (arranged by which fruit they get). There are \(n!\) ways to arrange \(n\) elements, so there are \(n!\) permutations of \(n\) elements. Often we only want to arrange some of these, so we would consider a \(k\)-permutation. We will use this often enough that it deserves its own notation.

\(k\)-permutations of \(n\) elements

\(P(n,k)\) is the number of \(k\)-permutations of \(n\) elements, the number of ways to arrange \(k\) objects chosen from \(n\) distinct objects.

\begin{equation*} P(n,k) = n\cdot (n-1) \cdot (n-2) \cdot \cdots \cdot (n-k+1) = \frac{n!}{(n-k)!}. \end{equation*}

Some other examples of questions where the answer is a permutation:

  • How many 5-letter “words” (i.e., strings of symbols) contain no repeated letters? There are 26 choices for the first letter, 25 for the second, 24 for the third, 23 for the fourth and 22 for the fifth, so there are \(26\cdot 25 \cdot 24 \cdot 23 \cdot 22 = P(26,5)\) such words.

  • How many different shuffled decks of cards are possible? There are 52 choices for the bottom card, 51 for the next, and so on. So the answer is \(P(52,52) = 52!\)

  • How many injective functions from \([k]\) to \([n]\) are there? There are \(n\) choices for where 1 is sent, \(n-1\) choices for where \(2\) is sent (since the function is injective, 2 cannot be sent to the same element as 1), and so on. So the answer is \(P(n,k)\text{.}\)

While we are looking at permutations, let's also consider their counterpart, combinations. The key distinction is that while a permutation counts different arrangements as different outcomes, a combination does not.

Activity91

Suppose you have 17 books strewn around your room. You resolve to get organized.

(a)

You decide to throw 10 books in the trash. How many choices do you have for which books to get rid of?

Hint

We are not necessarily looking for a numerical answer here, but you should be able to express the answer using notation we have developed already. Or by using some sort of triangle.

(b)

You think twice about tossing your books, and you find a bookshelf that can hold 10 books. How many ways can you fill the bookshelf?

(c)

How many ways could you fill the 10 book bookshelf if the books were arranged alphabetically?

Hint

Should your answer be larger or smaller than if the books could be in any order? How is this any different from throwing the books away?

Hopefully you realized that the number of \(k\)-combinations of \(n\) elements is just \(\binom{n}{k}\text{.}\) This makes sense: you just need to choose which elements to take, you don't need to arrange them.

Notice though that combinations and permutations are related in some way. What is that way? Well, once you have chosen \(k\) elements from an \(n\)-element set, you could then arrange those \(k\) elements to get a permutation.

Activity92

How many 5-letter words can be made with distinct letters?

(a)

Explain why the answer is \(P(26,5)\text{.}\)

(b)

Explain why the answer is also \(\binom{26}{5}\cdot 5!\text{.}\)

(c)

Express \(P(26,5)\) as the quotient of two factorials. Use this to find a numerical value for \(\binom{26}{5}\text{.}\)

Hint

You have \(P(26,5) = \binom{26}{5}\cdot 5!\text{.}\) You know that \(P(26,5) = \frac{26!}{21!}\text{,}\) so you should be able to easily “solve” for \(\binom{26}{5}\text{.}\)

There is nothing special about 26 and 5 in the above activity.

Activity93

Give a combinatorial proof for the identity \(P(n,k) = \binom{n}{k}\cdot k!\text{.}\)

Conclude that

\begin{equation*} \binom{n}{k} = \frac{n!}{(n-k)!\cdot k!}\text{.} \end{equation*}

Notice that the only thing we needed to find the algebraic formula for binomial coefficients was the product principle and a willingness to solve a counting problem in two ways. In Section 2.3 we will consider this formula again from the other direction.

Subsection2.2.3More Proofs

We will use the proof techniques of double counting and bijections throughout the rest of the book, but for now, let's practice a bit. The activities below ask you to give a combinatorial proof, and often also an alternative proof that has some nice features.

Activity94

Prove the identity \(\binom{2n}{2} = 2 \binom{n}{2} + n^{2}\text{.}\)

Hint

Think of selecting two balls from a set of \(n\) distinct red balls and \(n\) distinct green balls.

The reader should attempt an algebraic proof of the above result using the factorial formula for \(\binom{n}{k}\text{.}\)

Activity95

Prove the identity \(\binom{m + n}{2} - \binom{m}{2} - \binom{n}{2} = mn\text{.}\)

Hint

The \(mn\) looks like an application of the product principle. Say you had \(m\) bow ties and \(n\) fezzes. What question should you ask, and why does the other side also answer this?

You could also attempt an algebraic proof or perhaps a geometric proof making use of figures consisting of triangular numbers. Also, as a challenge, the reader could formulate a similar result involving \(\binom{a + b + c}{3}\) and a corresponding proof.

Activity96

We have already proved this identity: \(\binom{n}{0} + \binom{n}{1} + \binom{n}{2} + \ldots + \binom{n}{n} = 2^{n}\text{.}\) Now try giving a combinatorial proof that uses lattice paths.

Hint

You will want to consider a lot of lattice paths, not all ending at the same point. In fact, what do all the lattice paths have in common?

Becoming an expert in combinatorial proof will help you solve actual counting problems. For example, you might wonder how many positive integers have their digits in strictly increasing order.

Activity97

Prove that the number of positive integers that have their digits in strictly increasing order is \(2^{9} - 1\text{.}\) Include single digit numbers.

Hint

The answer looks like it might be almost the sum of a row in Pascal's triangle. Why does this make sense? You could think of including, or excluding numbers.

Now back to some combinatorial proofs.

Activity98

Prove \(\binom{n}{1} + 2 \binom{n}{2} + 3 \binom{n}{3} + \ldots + n \binom{n}{n} = n2^{n - 1}\text{.}\)

Hint

What if some number of a group of \(n\) people wanted to go to an escape room, and among those going, one needed to be the team captain?

Some other ways to approach the previous activity might be one or more of the following: \(n2^{n - 1}\) looks like a derivative, so try differentiating \(\left( 1 + x \right)^{n}\text{;}\) a reverse and add approach also works; or, first prove \(k \binom{n}{k} = n \binom{n - 1}{k - 1}\) and then use it.

Activity99

Prove the Pascal recurrence: \(\binom{n}{k} = \binom{n - 1}{k-1} + \binom{n - 1}{k}\text{.}\) Do this in two ways, first with a double counting style proof, and then a bijective proof.

Hint

Bit strings are nice here. What could the last bit of the bit string be?

The above activity illustrates another important reason to use combinatorial proofs: to establish recurrence relations. The next identity is also a recurrence, this time about derangements.

Activity100

A derangement of a set of elements is a permutation of those elements so that no element appears in its original position. For example, a derangement of \([4]\) might be \(3,1,4,2\text{,}\) but the permutation \(3,2,4,1\) is not a derangement because \(2\) is still in the second position.

Let \(d_{n}\) denote the number of derangements of \([n]\text{.}\) We will take \(d_{0} = 1\) and \(d_{1} = 0\text{.}\) Prove that \(d_{n} = (n - 1)(d_{n - 1}+ d_{n - 2})\) for \(n \geq 2\text{.}\)

Hint

Where can \(n\) go?

Later we will use the Principle of Inclusion/Exclusion to derive a formula for \(d_{n}\) from which the new recursion \(d_{n} = nd_{n - 1} + \left( - 1 \right)^{n}\text{,}\) and the above recursion, can be derived.

Let's get back to binomial identities.

Activity101

Prove, \(\binom{n}{0}^{2} + \binom{n}{1}^{2} + \binom{n}{2}^{2} + \ldots + \binom{n}{n}^{2} = \binom{2n}{n}\text{.}\)

Hint

First note that \(\binom{n}{k}^2 = \binom{n}{k}\binom{n}{n-k}\text{,}\) by the symmetry of Pascal's triangle. Then, there is a nice proof using lattice paths here, or you select \(n\) things from a collection that has \(n\) things of one type and \(n\) of another.

An alternate algebraic proof is less interesting: Extract the coefficient of \(x^{n}\) from both sides of \(\left\lbrack \left( x + 1 \right)^{n} \right\rbrack^{2} = \left(x + 1 \right)^{2n}\)

Activity102

\(\binom{2}{2} + \binom{3}{2} + \binom{4}{2} + \ldots + \binom{n}{2} = \binom{n + 1}{3}\text{.}\)

Hint

Try bit strings here. What length and weight are we looking at? Where might the first (left-most) one be?

For a non-combinatorial proof of the above, attempting a proof by mathematical induction is an easy option. An algebraic approach is not!

If you spend enough time working on these proofs, you might find yourself discovering new identities that use the same basic setup.

Activity103

Prove \(1\cdot n + 2 \cdot (n-1) + 3 \cdot (n-2) + \cdots + n \cdot 1 = \binom{n+2}{3}\)

Hint

You could use bit strings again, but this time think about where the second 1 could go. Or for variety, why not try asking about subsets of \([n+2]\text{?}\)

Activity104

\(\binom{3n}{3} = 3 \binom{n}{3} + 6n \binom{n}{2} + n^{3}\text{.}\)

Hint

Compare to Activity 94. What if you also have yellow balls?

Activity105

\(1 \cdot 1! + 2 \cdot 2! + 3 \cdot 3! + \ldots + n \cdot n! = \left( n + 1 \right)! - 1\text{.}\)

Hint

In how many ways can you arrange the \(n+1\) numbers \(0, 1, 2, \ldots, n\) so that they are not in ascending order?

The previous identity can also be established using a collapsing sum or induction proof.

Activity106

\(k \binom{n}{k} = n \binom{n - 1}{k - 1}\text{.}\)

Hint

Look back at Activity 98. This one is easier.

Activity107

\(\binom{n}{0} d_{0} + \binom{n}{1} d_{1} + \binom{n}{2} d_{2} + \ldots + \binom{n}{n} d_{n} = n!\) where \(d_{n}\) denotes the \(n\)th derangement number, \(d_{0} = 1,d_{1} = 0\text{.}\)

To close this section, let's return to a binomial identity we considered before, and consider a few additional proofs.

Proof

We give three proofs:

  1. How many lines that are not vertical or horizontal can you form by connecting the points that are on the positive \(x\) and \(y\) axes? Since we really want only lines that are formed by connecting the \(m\) points with the \(n\) points, we could say we have \(mn\) lines. Or we could consider all \(m + n\) points and pick 2 in \(\binom{m+n}{2}\) ways and then delete those choices where you took 2 from \(m\) or 2 from \(n\text{,}\) since these formed vertical and horizontal lines, respectively. Thus \(\binom{m + n}{2} - \binom{m}{2} - \binom{n}{2} = mn\text{.}\)

  2. There are \(mn\) one by one squares in the subdivided \(m\) by \(n\) rectangle. Each choice of arrows (one horizontal, one vertical) specifies one of these squares. Pick two arrows but don't take two from the top or two from the side. Then \(\binom{m + n}{2} - \binom{m}{2} - \binom{n}{2} = mn\text{.}\)

  3. Take \(m\) people in one group and \(n\) in another. How many handshakes can be accomplished? Among the \(m\) people there are \(\binom{m}{2}\) handshakes; among the \(n\) people there are \(\binom{n}{2}\) handshakes and between the two groups, \(mn\text{.}\) But, \(\binom{m + n}{2}\) also represents the total number of handshakes among the people. Then we get: \(\binom{m}{2} + \binom{n}{2} + mn = \binom{m + n}{2}\text{.}\)