74
$\begingroup$

The background of my question comes from an observation that what we teach in schools does not always reflect what we practice. Beauty is part of what drives mathematicians, but we rarely talk about beauty in the teaching of school mathematics.

I'm trying to collect examples of good, accessible proofs that could be used in middle school or high school. Here are two that I have come across thus far:

(1) Pick's Theorem: The area, $A,$ of a lattice polygon, with boundary points $B$ and interior points $I$ is $A = I + B/2 - 1.$

I'm actually not so interested in verifying the theorem (sometimes given as a middle school task) but in actually proving it. There are a few nice proofs floating around, like one given in "Proofs from the Book" which uses a clever application of Euler's formula. A very different, but also clever proof, which Bjorn Poonen was kind enough to show to me, uses a double counting of angle measures, around each vertex and also around the boundary. Both of these proofs involve math that doesn't go much beyond the high school level, and they feel like real mathematics.

(2) Menelaus Theorem: If a line meets the sides $BC,$ $CA,$ and $AB$ of a triangle in the points $D, E,$ and $F$ then $(AE/EC) (CD/DB) (BF/FA) = 1.$ (converse also true) See: http://www.cut-the-knot.org/Generalization/Menelaus.shtml, also for the related Ceva's Theorem.

Again, I'm not interested in the proof for verification purposes, but for a beautiful, enlightening proof. I came across such a proof by Grünbaum and Shepard in Mathematics Magazine. They use what they call the Area Principle, which compares the areas of triangles that share the same base (I would like to insert a figure here, but I do not know how. -- given triangles $ABC$ and $DBC$ and the point $P$ that lies at the intersection of $AD$ and $BC,$ $AP/PD = \operatorname{Area} (ABC) / \operatorname{Area}(DBC).$) This principle is great-- with it, you can knock out Menelaus, Ceva's, and a similar theorem involving pentagons. And it is not hard-- I think that an average high school student could follow it; and a clever student might be able to discover this principle themselves.

Anyway, I'd be grateful for any more examples like these. I'd also be interested in people's judgements about what makes these proofs beautiful (if indeed they are-- is there a difference between a beautiful proof and a clever one?) but I don't know if that kind of discussion is appropriate for this forum.

Edit: I just want to be clear that in my question I'm really asking about proofs you'd consider to be beautiful, not just ones that are neat or accessible at the high school level. (not that the distinction is always so easy to make...)

$\endgroup$
13
  • 5
    $\begingroup$ Surely you know the books ''Math! Encounters with Undergraduates'', ''Math Talks for Undergraduates'' and ''The beauty of doing Mathematics'' by Serge Lang. They were written with your same intention. $\endgroup$
    – agt
    Sep 8, 2011 at 11:15
  • 86
    $\begingroup$ I remember Lang giving a talk at Caltech supposedly aimed at undergraduates, but far more advanced. Lang called on a postdoc in the front row, apparently thinking he was a student, and the postdoc couldn't answer. At one point, he wrote out a complicated formula, and challenged Prof. Ramakrishnan, "Do you teach your students this?" "No." "So you see, Caltech is not better than anywhere [sic] else." After a bit more, he went back to the formula and corrected a sign. Ramakrishnan called out, "That, we teach." $\endgroup$ Sep 14, 2011 at 0:38
  • 6
    $\begingroup$ Even the notion of proof may not be accessible at high school level $\endgroup$ Sep 15, 2011 at 20:33
  • 2
    $\begingroup$ In the US, most high school mathematics classes are not based on proofs. Theorems are introduced without reference to proof. Only in a few limited situations are students asked to prove anything, such as in some geometry classes and some calculus classes. Instead, high school math classes concentrate on introducing objects, their properties, and how to manipulate those objects. Finding neat accessible proofs to show them is reasonable, but this is very different from finding neat accessible math to show them. E.g., I think the Chaos Game is accessible, but few students can handle the proofs. $\endgroup$ Sep 17, 2011 at 0:39
  • 19
    $\begingroup$ As far as I remember, the thing I liked the most in high school maths (age of 14) was the so called Ruffini's rule: $(x-a)$ divides a polynomial $P(x)$ if and only if $P(a)=0$. It looked to me so incredibly easy and so full of consequences. I hope they still learn it with a proof. $\endgroup$ Sep 18, 2011 at 20:12

51 Answers 51

79
$\begingroup$

Extending on Ralph's answer, there is a similar very neat proof for the formula for $Q_n:=1^2+2^2+\dots+n^2$. Write down numbers in an equilateral triangle as follows:

    1
   2 2    
  3 3 3
 4 4 4 4

Now, clearly the sum of the numbers in the triangle is $Q_n$. On the other hand, if you superimpose three such triangles rotated by $120^\circ$ each, then the sum of the numbers in each position equals $2n+1$. Therefore, you can double-count $3Q_n=\frac{n(n+1)}{2}(2n+1)$. $\square$

(I first heard this proof from János Pataki).

How to prove formally that all positions sum to $2n+1$? Easy induction ("moving down-left or down-right from the topmost number does not alter the sum, since one of the three summand increases and one decreases"). This is a discrete analogue of the Euclidean geometry theorem "given a point $P$ in an equilateral triangle $ABC$, the sum of its three distances from the sides is constant" (proof: sum the areas of $APB,BPC,CPA$), which you can mention as well.

How to generalize to sum of cubes? Same trick on a tetrahedron. EDIT: there's some way to generalize it to higher dimensions, but unfortunately it's more complicated than this. See the comments below.

If you wish to tell them something about "what is the fourth dimension (for a mathematician)", this is an excellent start.

$\endgroup$
8
  • 4
    $\begingroup$ @Frederico: Careful. The sum of cubes does not correspond to a tetrahedron, but rather a pyramid. (Which does not have this nice symmetry) However the slightly modified sum $$Q_N^'=\sum_{n=1}^N n\frac{n(n+1)}{2}=\frac{1}{2}\sum_{n=1}^N n^3+\frac{1}{2}\sum_{n=1}^N n^2$$ will correspond to a tetrahedron. By using this symmetry we find $$Q_N^' = \frac{1}{4} (3N+1)\left(\frac{N(N+1)(N+2)}{6}\right)$$ since each entry will be $3N+1$ and the tetrahedral numbers are $\binom{N+3}{3}=\frac{N(N+1)(N+2)}{6}$. From here we can deduce that $$\sum_{n=1}^N n^3=\frac{(N^2+N)^2}{4}.$$ $\endgroup$ Sep 16, 2011 at 23:31
  • 3
    $\begingroup$ In short, you won't have a nice generalization of the solution for $n=2$ to higher dimensions. Such a generalization is not expected either since Faulhaber's Formula is not so simple. $\endgroup$ Sep 16, 2011 at 23:39
  • 1
    $\begingroup$ @Eric Oh, you're right. :( What a pity, it would've been an even neater generalization. $\endgroup$ Sep 17, 2011 at 9:57
  • 1
    $\begingroup$ Cool. Both the example and the follow up comments. $\endgroup$
    – Manya
    Sep 19, 2011 at 8:30
  • 1
    $\begingroup$ Glad you appreciate it - I like it for the same reason. I don't think there is an easy formula for higher dimensions: since the Bernoulli numbers show up in the final formula, any procedure to prove it should somehow "encode" them. The best proof I know for a generic dimension relies on the identity $ (n+1)^{a+1}=\sum_{k=0}^{n} ((k+1)^{a+1}-k^{a+1}) = \sum_k \sum_{d=0} \binom{a+1}{d}k^d$, which can be used to express $\sum_k k^a$ (unknown) in terms of $(n+1)^{a+1}$ and sums of lower powers (known by induction). $\endgroup$ Jul 24, 2013 at 9:21
58
$\begingroup$

The theorem of "friends and strangers": the Ramsey number $R(3,3)=6$. Not only can the proof be understood by high-school students, a proof can be discovered by students at that level via something akin to the Socratic method. First students can establish the bound $R(3,3) > 5$ by 2-coloring the edges of $K_5$:
           R33
Then they can reason through that a 2-coloring of the edges of $K_6$ must contain a monochromatic triangle, and so $R(3,3)=6$: in every group of six, three must be friends or three must be strangers.

After this exercise, an inductive proof of the 2-color version of Ramsey's theorem is in reach.

An added bonus here is that one quickly reaches the frontiers of mathematics: $R(5,5)$ is unknown! It can be a revelation to students that there is a frontier of mathematics. And then one can tell the Erdős story about $R(6,6)$, as recounted here. :-)

$\endgroup$
9
  • 5
    $\begingroup$ R(3,3)=6 sticks in my memory as the one time I managed to explain mathematics to a non-mathematical friend in the pub $\endgroup$
    – Yemon Choi
    Sep 8, 2011 at 21:57
  • 9
    $\begingroup$ This might be apocryphal, but I've read a story about a sociologist who was surprised to discover such patterns of friendship or non-friendship among his subjects, and pondering the deep psychological origins. If true, a wonderful argument for the need for math literacy. $\endgroup$ Sep 9, 2011 at 1:11
  • 51
    $\begingroup$ From Jacob Fox's lecture notes: "In the 1950’s, a Hungarian sociologist S. Szalai studied friendship relationships between children. He observed that in any group of around 20 children, he was able to find four children who were mutual friends, or four children such that no two of them were friends. Before drawing any sociological conclusions, Szalai consulted three eminent mathematicians in Hungary at that time: Erdős, Turán and Sós. A brief discussion revealed that indeed this is a mathematical phenomenon rather than a sociological one." $\endgroup$ Sep 9, 2011 at 12:07
  • 9
    $\begingroup$ Re the Szalai story: $R(4,4) = 18$. $\endgroup$ Sep 9, 2011 at 13:36
  • 5
    $\begingroup$ I finally tracked down where I must have read about Szalai's story for the first time: N. Alon and M. Krivelevich's article on extremal combinatorics in the Princeton's companion. Also available at: cs.tau.ac.il/~krivelev/papers.html $\endgroup$ Sep 29, 2011 at 17:30
47
$\begingroup$

Euler's Bridges of Konigsberg problem. You can give it to students for five minutes to play with, watch them get annoyed, and then offer them the classical simple and beautiful impossibility proof. I think a lot of high school students, and even bright middle school students, would be totally convinced.

$\endgroup$
1
  • $\begingroup$ Yes! A read about that example in one of Martin Gardner's books at about that age. $\endgroup$ Jun 15, 2012 at 19:16
43
$\begingroup$

The proof, by counting inversions, that you can't interchange the 14 and 15 in the 15 puzzle, just by sliding, is accessible to high-school students, introduces important ideas, and might be found beautiful.

$\endgroup$
32
$\begingroup$

Two of my favorites: (1) The parameterization of all primitive Pythagorean triples; (2) The formula for the $n$th Fibonacci number in terms of the golden ratio $\phi = \frac{1+\sqrt{5}}{2}$, with the corollary that $\displaystyle \lim_{n \longrightarrow \infty} F_n/F_{n-1} = \phi$.

$\endgroup$
1
32
$\begingroup$

The trefoil knot is non-trivial.

Proof: It has a tricoloring:

alt text (source)

And the existence of a tricoloring is preserved by Reidermeister moves. QED

$\endgroup$
5
  • 4
    $\begingroup$ +1 because it is so obviously technically uncomprehensible to a high school student (at myn hiogh school) yet plausibly beautiful, and beautifully plausible. $\endgroup$
    – roy smith
    Sep 16, 2011 at 5:05
  • 18
    $\begingroup$ @Roy: I disagree with you. I don't think that this is incomprehensible to high school students. The notion of knot is pretty intuitive, and it's very easy to explain what a tricoloring is. Also, the Reidemeister moves are pretty easy things to explain (ok -- I'm not talking about the actual proof that they generate everything). The most difficult aspect of the argument is probably to convince a high school student that there is need for proof. Namely, isn't the trefoil obviously non-trivial!? For that, it might be good to show them some monster unknots first... $\endgroup$ Sep 16, 2011 at 5:17
  • 34
    $\begingroup$ When I was shown the Reidemeister moves in school, several of my classmates and I made the objection, in essence, that it wasn't clear that they generated everything. Worse, since we didn't have topology to work with, we didn't really have a "real" definition to compare it with, so it felt to us that the real issues were being swept under the rug. $\endgroup$ Jun 15, 2012 at 18:49
  • $\begingroup$ I’ve never performed the experiment, but I think it should be possible to explain the completeness of the reidemeister moves to a grade schooler. Ingredients: a knot made of wire and a flashlight to make its shadow. First by shining the light, illustrate that the projection generally has only transverse double points; explain that for this to fail either the light ray must be tangent to the knot or must pass through three points; either can be avoided by jiggling a little. Then move the flashlight around; illustrate that in 1-parameter families one sees no worse than cusps and triple points $\endgroup$ Dec 24, 2020 at 18:12
  • $\begingroup$ ... explain that to see anything worse, the light must be twice tangent or pass through four points or... and again you jiggle to separate the phenomena. (If you’re feeling exceptionally honest, I suppose you could mention that it will be many years before they learn the mathematics of jiggling.) $\endgroup$ Dec 24, 2020 at 18:17
29
$\begingroup$

Going by the parameters of the question, I don't see why the proof would necessarily need to be of a sophisticated theorem. I think Euclid's proof of the infinitude of primes is beautiful and definitely accessible to a high school audience. Having given the proof, one might reflect on some of its features that generalize to many other contexts, like proof by contradiction or the ability to use a clever construction to avoid infinite enumeration.

$\endgroup$
5
  • 38
    $\begingroup$ No! Do not use this proof as an occasion to talk about proof by contradiction. Euclid's proof of this proposition was not by contradiction. The conventional practice of rearranging it into a proof by contradiction not only adds an extra complication that serves no purpose, but also leads to confusions such as the belief that if you multiply the first n primes and add 1, the result is always prime. See the paper by me and Catherine Woodgold on this: "Prime Simplicity", Mathematical Intelligencer, autumn 2009, pages 44--52. $\endgroup$ Sep 15, 2011 at 20:03
  • 2
    $\begingroup$ (I fully agree that this proof is an excellent example for high-school students.) $\endgroup$ Sep 15, 2011 at 20:06
  • 2
    $\begingroup$ It IS by contradiction, just not the kind of contradiction we have in mind. $\endgroup$ Sep 28, 2011 at 1:35
  • 26
    $\begingroup$ To complete Michael Hardy's comment, Euclid's original proof proves the following statement: given any finite list of primes, we can extend the list by finding a prime not in the list. (Proof: multiply the primes in the list and add one; any prime factor of this new number is a prime not in the original list.) So it's a constructive way of taking a list of primes and producing another prime; we don't have to assume that the original list was "all the primes" (as in the contradiction proof) or that they were the first n primes. $\endgroup$
    – shreevatsa
    Dec 23, 2011 at 4:42
  • $\begingroup$ It's certainly a great proof -- certainly immediately convincing. For some reason, it always irritated me. I've thought on and off over the years why this is. The only thing I can think of is that, while it is constructive, it is also outrageously inefficient. Now at the time I learned this, no one that I knew was even talking in those terms, so maybe it's just some instinct that I had. I really wish I liked it more. G.H. Hardy did, right? You can't argue with that. $\endgroup$ Oct 4, 2012 at 0:30
26
$\begingroup$

Cantor's diagonal argument.

The warm-up could be an equally beautiful proof, namely that the rationals are countable.

$\endgroup$
6
  • 8
    $\begingroup$ I am a bit torn on this one. At least I believe before one would have to also prove, say, that the cardinality of the rationals is equal to that of the naturals (and/or related things). Because if not the 'insight' that there are more reals than naturals might seem too 'obvious' to make a proof enlightening. $\endgroup$
    – user9072
    Sep 8, 2011 at 10:51
  • 2
    $\begingroup$ But the enumeration proving that the rationals are countable is also almost as beautiful as the diagonal argument! Perhaps they could be combined. $\endgroup$
    – J.C. Ottem
    Sep 8, 2011 at 11:11
  • 3
    $\begingroup$ The counting-rationals argument loses some appeal when you have to deal properly with skipping the non-reduced fractions. The visual proof of course establishes that $\mathbb{N} \times \mathbb{N} \cong \mathbb{N}$ as sets, and the countability of $\mathbb{Q}$ follows by applying an unrelated lemma on cardinality of surjective images. Glossing over a lemma like that seems exactly like what a beautiful proof (especially an exemplary proof) should avoid. $\endgroup$
    – Ryan Reich
    Sep 9, 2011 at 21:09
  • 8
    $\begingroup$ There is a simple bijection between the positive rationals and positive integers shown by the fusc function. Let $f(n)$ be the number of ways of representing $n$ as a sum of powers of $2$ with at most $2$ copies of each power. $f(4)=3$ since $4=4,2+2,2+1+1$. The sequence $\{f(n)\}_{n=0} = \{1,1,2,1,3,2,3,1,4,...\}$. Over natural numbers $n$, $g(n) = f(n-1)/f(n)$ hits each positive rational precisely once. See also the Calkin-Wilf tree, how Euclid's algorithm reduces relatively prime ordered pairs. The sequences of numerators and denominators in order are offset copies of $\{f(n)\}$. $\endgroup$ Sep 10, 2011 at 21:33
  • 4
    $\begingroup$ What Douglas Zare said is nicely explained in the 1999 paper Recounting the Rationals (math.upenn.edu/~wilf/website/recounting.pdf) by Calkin and Wilf. Some people have felt (blog.plover.com/math/recounting.html) that it's a good "first math paper" to read. $\endgroup$
    – shreevatsa
    Dec 23, 2011 at 4:48
24
$\begingroup$

Seeing the struggle of many students with standard trigonometry, I especially like the rational parametrization of $x^2+y^2=1$ (which is equivalent to listing all Pythagorean triples) by starting from $\sin^2\phi+\cos^2\phi=1$ and then using $$ \sin\phi=\frac{2t}{1+t^2}, \quad \cos\phi=\frac{1-t^2}{1+t^2}, \qquad t=\tan\frac{\phi}2. $$ Note that the formulas are usually used in the context of integration of rational expressions in sine and cosine.

At the same time, a more general "geometric" argument (applicable to general quadratics), due to Bachet (1620), is still at school level. Namely, fix a single rational point on the curve, $(x _ 0,y _ 0)$ say, and consider the intersection points of the curve and straight lines $y-y_0=t(x-x_0)$ with rational slope $t$ passing through the point.

A beauty here is because of variety of different geometric and analytic methods for solving a classical arithmetic problem.

$\endgroup$
2
  • 2
    $\begingroup$ Since coordinate geometry was invented after 1620 I wonder how Bachet could have done that. $\endgroup$ Sep 8, 2011 at 13:23
  • 10
    $\begingroup$ Franz, I will be definitely happier if you are a little bit more constructive in your critisism: you seem to be the right person to explain why the method is usually attributed to Bachet! $\endgroup$ Sep 10, 2011 at 6:12
19
$\begingroup$

The proof that $\sqrt{2}$ is irrational is a nice example of proof by contradiction.

$\endgroup$
16
  • 1
    $\begingroup$ @Manya My conception of beauty has changed since I first saw that proof. At that time, I did indeed consider it quite beautiful. Now, however, it's so embedded in me that it's hard to appreciate its beauty. $\endgroup$ Sep 8, 2011 at 13:58
  • 12
    $\begingroup$ It is not a proof by contradiction - it is a proof of a negation : $\sqrt{2}=p/q\Rightarrow\bot$. $\endgroup$ May 18, 2013 at 19:23
  • 12
    $\begingroup$ To expand on what Paul Taylor said: to prove a negation $\neg p$, "assume $p$ is true ... derive falsity". A proof by contradiction of $p$: "assume $p$ is false ... derive falsity". The difference is that the second method uses the law of the excluded middle ($\neg \neg p \Rightarrow p$), whereas the first is still valid in intuitionistic logic. See Andrej Bauer's essay: math.andrej.com/2010/03/29/… $\endgroup$
    – Todd Trimble
    Jul 24, 2013 at 10:55
  • 1
    $\begingroup$ @PaulTaylor Some proofs really are proofs by contradiction under that definition. (For example, a student may start with "Assume that $\sqrt{2}$ is not irrational. That means it is rational. So...") Thus, a more correct statement is: "A proof exists that avoids a certain logical non-intuitionistic axiom." But this is about teaching high school students beautiful proofs. Are you really going to tell such a student "it is not a proof by contradiction"? If so, why stop there? One could say: "It is not a proof of a negation, it is a fill in the blank, weaker fragment proof." $\endgroup$ Jan 11 at 3:35
  • 1
    $\begingroup$ @PaceNielsen: when studying logics of various kinds, one might "worry" about $\Rightarrow$, but this discussion was about "proof by contradiction" vs "proof of negation", which are distinguished by what happens with $\lnot$. Someone who "works with classic full logic" thinks that $\lnot\lnot\phi$ is the same as $\phi$ and therefore cannot see the difference. Just like a colour-blind person cannot see the difference between red and green. So which is which? Rational numbers are defined without using negation, irrationality is derived from that by putting $\lnot$ in front. $\endgroup$ Jan 11 at 17:29
19
$\begingroup$

The Gale-Sharpley stable marriage theorem, http://en.wikipedia.org/wiki/Stable_marriage_problem .

The algorithm and its proof are very much accessible to school students. Despite its innocuous look, the algorithm is not easy at all to invent.

On a similar note, Hall's theorem: http://en.wikipedia.org/wiki/Hall%27s_marriage_theorem#Graph_theory . This looks like a recreational puzzle but actually is closer to university mathematics than everything done in high school.

Here is another combinatorial exercise which, properly presented, does not even look like mathematics: the round track puzzle. The thing I don't like about it is that the standard "gotcha" proof (explained in the usual, informal way) requires a bit too much concentration to understand - some students might fail at it and take it as an example that mathematical proofs are something one either believes or not, rather than something one can check. Of course, one can formalize the proof, but this requires quite an amount of time in a high school class. [Update, a decade later: I have written an expanded version of the solution up as the solution to Exercise 5.2.4 in my Fall 2020 Math 235 notes.]

$\endgroup$
2
  • 2
    $\begingroup$ I’ll point out that the “combinatorial exercise” is actually useful. Assume that we are given a sequence $s_1,\dots,s_n$ of symbols of arities $a_1,\dots,a_n\ge0$, respectively. It is easy to see that if we can arrange them into a well-formed term, then $\sum_ia_i=n-1$. The exercise says that, conversely, if this identity holds, then there exists a (unique) cyclic permutation of the string $s_1\dots s_n$ which is a valid term in Polish (prefix) notation. Among other things, this allows you to count the well-formed terms consisting of $n$ symbols. $\endgroup$ Sep 8, 2011 at 12:29
  • $\begingroup$ Nice! I think the following proof is another application of this exercise: artofproblemsolving.com/Forum/viewtopic.php?p=2385164#p2385164 . $\endgroup$ Sep 8, 2011 at 12:52
16
$\begingroup$

The Halting Problem. The first time I saw this was my senior year in high school and it completely blew me away. All you need is a notion of what an algorithm is and very basic logic (enough to recognize that assuming $A$ and deriving $\neg A$ is a contradiction)

In a similar vein, Russell's Paradox. The problem here is that you need some basic set theory, so this is more for advanced high school students.

The first beautiful proof I saw in high school (it was beautiful at the time, but now seems too trivial) was the fact that for a geometric series $a, ax, ax^2, \dots$ the sum of the first $n$ terms is $a \cdot \frac{1-x^n}{1-x}$. I thought this was cool because of all the cancellation that seems to come out of nowhere. The treatment here is nice.

$\endgroup$
1
  • $\begingroup$ I haven't taught high school in many years, and I never tried teaching these things there. My impression however, is that the Halting Problem, although the proof is indeed elementary, is conceptually very sophisticated. I think it's one of those things that people have to mull over for years to really get an appreciation for. Maybe some teachers, with some classes, could do this. But I bet that most high school students might be able to follow "every line of the proof", but they'd see it as just a trick. I'd love to be proved wrong. Maybe the same is true for Russell's paradox. $\endgroup$ Dec 3, 2011 at 2:05
13
$\begingroup$

One of the keys to making a proof accessible to high school students (or just non-mathematicians) is to make the answer relevant. This gives a dual responsibility, to ensure that the theorem is motivated and that the proof is accessible. The proof of the infinity of the primes has been mentioned already and is a fantastic example. You can lead students in to it using the (almost trivial) proof that there is no largest integer.

Another example is the classification of the regular polyhedra. With good students and models you can even lead them to the proof there there are at most 6 regular polytopes in 4d (actually showing they all exist is a little harder).

Keeping with polyhedra, the Euler characteristic is also powerful. Start with balloons and get the students to draw lines freely so you get a tiling. Then get them to count faces, vertices and edges. David Eppstein collected 19 proofs to choose from, several of which would be perfect for non-mathematicians: http://www.ics.uci.edu/~eppstein/junkyard/euler/

As a final example (and to show that it does not have to be deep mathematics to motivate) you can consider the question of blocking a square on a chess board and filling the remainder with tromioes. It starts with a puzzle, you can get people to play with, and leads to a lovely induction proof: http://www.cut-the-knot.org/Curriculum/Games/TriggTromino.shtml

Actually polyominoes are a fantastic source of many other fun, non-trivial but accessible proofs.

$\endgroup$
13
$\begingroup$

This topological proof of the fundamental theorem of algebra is accessible to high school students, particularly those at the precalculus level.

    alt text (source)

There are two major problems with this. First, while the winding number is intuitive, it takes effort to define it rigorously. Second, you also want to establish the basic property that the winding number doesn't change as you deform a curve without going over the origin, which again is difficult to establish rigorously without topology. Without these details, you might call this a hand-waving argument instead of a proof. It's good to give references to where these results will be established rigorously, and to give arguments for other results which are more complete.

Nevertheless, I like presenting this proof for several reasons. I think it's beautiful. Geometrically, what $x\mapsto x^n$ does to the complex plane is easy to understand, but many students have little intuition about what this map does, only what polynomials look like on the real line. So, this argument doesn't just say that the statement is true, it is illuminating. The fundamental theorem of algebra is also a result students encountered in algebra, but they usually don't know why it's called a theorem. This is also an opportunity to talk a little about what is studied in more advanced areas of mathematics. It can lead into discussions of topology or the difficulty solving polynomials by radicals.

$\endgroup$
2
  • 7
    $\begingroup$ @Douglas: I ask this question out of ignorance; it's not a criticism. How many high-school students have been introduced to the complex plane sufficiently to grok this? $\endgroup$ Apr 19, 2013 at 23:43
  • 1
    $\begingroup$ @Joseph O'Rourke: I don't think many high school students are really comfortable with the complex plane, but I think you can develop key facts within this argument, particularly for precalculus students who are used to problems like calculating $\lim_{x\to \infty} \frac{x^4+1}{x-3}$. $\endgroup$ Apr 20, 2013 at 2:35
12
$\begingroup$

I've been collecting simple, often one-step, proofs.

http://www.cut-the-knot.org/proofs/index.shtml

Some I judge beautiful - these are listed separately.

$\endgroup$
0
11
$\begingroup$

I recommend Kelly's proof of the Sylvester-Gallai theorem (the original proof of Gallai was about 30 pages long, this one takes a few lines). The theorem and the proof can be read here.

$\endgroup$
3
  • $\begingroup$ I don’t understand the proof as given in the Wikipedia article. The picture is very suggestive, but as far as I can see, there is actually nothing in the proof that prevents the intersection of $m$ and $l$ to fall on the other side of the perpendicular (e.g., it may be $A$), in which case it can easily happen that the distance of $B$ from $m$ is larger than the distance of $P$ from $l$. Am I missing something? $\endgroup$ Sep 8, 2011 at 11:58
  • 1
    $\begingroup$ @Emil: There are at least 3 points on $l$, at least two on one side of the perpendicular. Call the closer of these two $B$ and the farther of these two $C$. Now let $m$ denote the line $PC$, then the pair $(B,m)$ has a smaller distance than $(P,l)$. Contradiction. $\endgroup$
    – GH from MO
    Sep 8, 2011 at 12:08
  • 1
    $\begingroup$ This brings up a memory. It was at McGill. Moser had just received an award from the CMS. I asked him about his work and somehow this problem and the related problem of finding the best lower bound for the number of "ordinary lines" came up. The funny thing is that Moser wanted to write on the back of the letter he had just received from CMS! If I remember well, a woman told him not do that (it may have been his wife). $\endgroup$
    – Malkoun
    Jan 11 at 13:41
11
$\begingroup$

Im in high school and i loved the proof of the fermat-toriccelli point of a triangle.

$\endgroup$
1
  • $\begingroup$ Thanks. I don't think I saw that one in high school! $\endgroup$
    – Manya
    Jun 18, 2012 at 19:43
11
$\begingroup$

Minkowski's Theorem (every convex region in the plane of area greater than 4 that's symmetric about the origin contains a lattice point other than (0,0)) is not at all obvious (are you sure you can't squeeze a sufficiently large "blob of irrational slope" in there?) but has a beautiful, simple, and surprising geometric proof.

$\endgroup$
2
  • 3
    $\begingroup$ Do you want to give a hint about the proof? $\endgroup$
    – Manya
    Nov 19, 2011 at 12:22
  • 10
    $\begingroup$ Draw the region $R$ in the plane. Cut the plane into 2x2 squares by cutting along the lines $x=2i$ and $y=2j$ for all integers $i,j$; each square contains some part of $R$ (possible none.) Stack the squares on top of each other. Since $R$ has area greater than 4, there exist two squares whose parts of $R$ overlap. Write down what this means algebraically, apply symmetry and convexity, and construct the nontrivial lattice point. $\endgroup$
    – user6542
    Nov 20, 2011 at 11:14
10
$\begingroup$

For someone in high school, I think it's good to prove that the sum of the interior angles of a triangle is $\pi$ if they don't know why. Personally, I was never shown why this fact is true, and I feel that it's generally a bad idea to not know why something in math is true, especially when the answer is pretty. My favorite proof is to think about how the normal vector changes as you walk around the triangle -- it's nice because it generalizes to other shapes (which may not even be polygons).

$\endgroup$
1
  • 1
    $\begingroup$ Relatedly, the Pythagorean theorem, and its connection to the parallel postulate, is really worthwhile. $\endgroup$
    – Todd Trimble
    Dec 16, 2014 at 21:25
8
$\begingroup$

1) Many elementary binomial identities or identities with Fibonacci numbers have beautiful proofs. Let me only mention the matrix representation of Fibonacci numbers whose determinant gives Cassini's identity.

2) Another elementary problem is the following: Is it possible to cover a checkerboard from which one white and one black square have been removed with dominoes? To show that this is possible run through the board in a cyclical way. Observe that on this path between a white and a black square are an even number of squares. Since I don't know how to make figures I indicate such a path for a 4x4-board: ((1,1),(1,2),(1,3),(1,4),(2,4),(2,3),(2,2),(3,2),(3,3),(3,4),(4,4),(4,3),(4,2),(4,1),(3,1),(2,1)).

$\endgroup$
2
  • 6
    $\begingroup$ I would also add the (checkerboard) proof that if you remove opposite corners from the 8x8 board, the resulting figure can't be tiled with dominoes; if presented without the board coloration it's not immediately obvious, and its proof provides a wonderful a-ha moment that should be easily accessible for high school students (if not earlier). These (along with the various graph-theory problems) also have the advantage of showing students that mathematics is about more than just numbers. $\endgroup$ Sep 8, 2011 at 21:31
  • $\begingroup$ Steven, +1, but even when you present with the board coloration it's not obvious... until they see it! And then it is suddenly obvious. $\endgroup$ Sep 9, 2011 at 0:28
8
$\begingroup$

You should certainly look at the two books by Ross Honsberger, "Mathematical Gems I" and "Mathematical Gems II". A favourite example of mine the proof due to Conway that there are configurations of checkers below the half plane on an infinite board that allow you to move a checker 4 rows into the upper half plane, but not five rows.

The negative result is an ingenious argument using nothing more than the quadratic formula, but provides a great example of to apply mathematics in unexpected contexts.

$\endgroup$
5
  • $\begingroup$ The way I remember the negative result was that you associate a weight to the checkers so that jumping and removing keeps the weight the same, and the infinite sum $\sum (2n+1) \Phi^{-n}$ for the weight of the lower half plane converges and can be calculated. That last calculation seems too complicated for middle school students who aren't comfortable with $\sum 1/2^n$ or $\Phi$. The positive result is a simple exercise requiring no background. Since the whole question is very abstract I don't think this is an unexpected application of mathematics. $\endgroup$ Sep 8, 2011 at 20:40
  • $\begingroup$ I was under the impression that 5 rows up is attainable if you use all of the spaces in the lower half plane (assuming a suitable notion of doing infinitely many hopping moves). $\endgroup$
    – S. Carnahan
    Sep 9, 2011 at 13:37
  • $\begingroup$ Well, $5$ is not attainable if you allow infinite well-ordered sequences of moves and take position-wise limits at the limit points. $\endgroup$ Sep 9, 2011 at 20:55
  • $\begingroup$ I don't understand how infinitely many moves can help. If a checker arrives at a certain spot then it did so in finitely many moves. Of course one could use an ultrafilter and say a checker arrives at a spot if it lands there ultrafilter often, but I don't see how this would be much use either The infinite dimensions of the board simply allow all finite configurations. $\endgroup$ Sep 13, 2011 at 20:44
  • $\begingroup$ In answer to Douglas Zare, the argument I remember did involve summing a geometric series of number that are the roots of a quadratic polynomial. You are right that this is probably too advanced for most high school students, bug I ythink it could be taught to bright students of the type that this hypothetical course would be geared towards. $\endgroup$ Sep 13, 2011 at 20:47
8
$\begingroup$

I suggest Euler's polyeder formula with application to the Platonic solids. One first observes that instead of polyeders one can consider graphs in the plane, counting the unbounded region as a face. One observes next that one can just as well allow the edges to be pieces of curves. Then one observes that the formula $F - E + V = 2$ is preserved if one edge or one vertex is added, thus the formula is proved by induction. Then use the formula to prove that there are no regular polyeders except the five Platonic solids as follows: faces can only be triangles, squares or pentagons, edges are always common to exactly two faces, and for the number of faces that meet in a vertex there are a number of possibilities that one checks. For each case, $E$ and $V$ can be expressed in terms of $F$, and the formula gives the possible values of $F$.

$\endgroup$
2
7
$\begingroup$

i suggest the proof archimedes wanted on his tombstone and its relatives. i.e. since two solids with the same horizontal slice area at every height have the same volume, hence by pythagoras, the volume of a cylinder equals the sum of the volumes of an inscribed cone and an inscribed solid hemisphere.

to generalize this, the volume generated by revolving a solid hemisphere around a planar axis in 4 space equal that generated by revolving a cylinder minus that generated by revolving a cone. Using the fact that the center of gravity of a cone is 1/4 the way up from the base, one obtains the volume of a 4-sphere, as pi^2/2 R^4.

a generalization of the first computation is that of the volume of a bicylinder (intersection of two perpendicular cylinders), since it is the difference of the volumes of a cube containing the bicylinder and a square based double pyramid also inscribed in the cube.

I find these beautiful, but of course that is subjective.

I also like euclid's argument for pythagoras, and for constructing a regular pentagon, but they are hard to reproduce here briefly.

$\endgroup$
7
$\begingroup$

Here is one that I like and used it for different purposes, e.g. introduction to proofs, algebraic thinking, beauty, and so one. Shuffle a deck of cards. Divide it into two halves. Magic: The number of the red cards in one of the halves is exactly equal to the number of black cards in the other half. It has a simple algebraic proof. However, The bigger magic is yet to come.

Theorem: Vertically opposite angles are equal.

Bigger Magic: The two theorems/proofs are essentially the same!

$\endgroup$
7
$\begingroup$

The formula $1 + 2 + \cdots + n = n(n+1)/2$ can be proved on middle school level: Assume first $n$ is even. Then there are $n/2$ pairs $(1,n), (2,n-1), \ldots, (n/2,n/2+1)$ those sum is always $n+1.$ Thus the overall sum is $n/2\cdot(n+1).$ The case when $n$ is odd can be treated in the same manner.

$\endgroup$
6
  • $\begingroup$ And do you think this proof is really beautiful (as opposed to neat, cool, or something like that)? $\endgroup$
    – Manya
    Sep 8, 2011 at 11:30
  • 15
    $\begingroup$ Actually this proof would be more approriate for primary school (age 10 or so). Smarter kids (like Gauss) figure it out themselves. $\endgroup$
    – GH from MO
    Sep 8, 2011 at 12:04
  • 3
    $\begingroup$ I prefer the "visual proof" because you don't have to reason, you just "see" it: put 1 little square in the first raw, two in the second,... n in the last raw, and you've obtained a figure of which you can easily compute the area as: (AreaOfBigSquare-AreaOfSquaresOnTheDiagonal)/2+ AreaOfSquaresOnTheDiagonal=$(n^2-n)/2+n=n(n+1)/2$. Of course to pass from $(n^2-n)/2+n$ to $n(n+1)/2$ the kid must have learned some "algebra". $\endgroup$
    – Qfwfq
    Sep 8, 2011 at 13:48
  • 3
    $\begingroup$ No, the best proof is the visual one where you draw a triangle with 1, 2, ..., n circles in each row, and another row below the last with n + 1 circles. Anything in the smaller triangle can be specified by two "coordinates" in the last row, obtained by dropping down parallel to the sides. Thus, $\binom{n + 1}{2}$ using the definition of "n + 1 choose 2"; to get a formula for that number you could use the argument in Dan's comment. This one is nice because it is both visual and a bijective proof, rather than computational. $\endgroup$
    – Ryan Reich
    Sep 8, 2011 at 15:31
  • 6
    $\begingroup$ How about this: write the numbers from $1$ to $n$ in row, then write them again in a row below, but backwards. Add up the columns, to get $n+1$ each time, so $n(n+1)$ in total, because there are $n$ columns. Divide by $2$. $\endgroup$
    – euklid345
    Sep 9, 2011 at 4:22
6
$\begingroup$

There is a very elegant proof that there exists no continuous injection from the plane into the real line. The proof can basically be given by drawing a picture on the blackboard.

Suppose there is such an injection $f$. Let $x$ and $y$ be distinct points in the plane and let $g_1$ and $g_2$ be paths from $x$ to $y$ such that $g_1(r_1)\neq g_2(r_2)$ for $r_1,r_2\in (0,1)$. Now this implies that $f\circ g_1((0,1))\cap f\circ g_2((0,1))=\emptyset$, contradicting the intermediate value theorem.

$\endgroup$
3
  • 12
    $\begingroup$ But what is the point of proving a theorem as intuitively obvious as this in the eyes of a high-schooler? $\endgroup$ Sep 8, 2011 at 10:18
  • 15
    $\begingroup$ Also, continuous functions in more than 1 dimension are usually not defined in high school... $\endgroup$ Sep 8, 2011 at 10:19
  • $\begingroup$ In fact, one might not even explicitly define 'continuous' at all in high school, as everything they're going to work with (polynomials, exp, log) is continuous anyway. $\endgroup$ Sep 9, 2011 at 12:00
6
$\begingroup$

Morley's Theorem.

Wikipedia's proof is completely elementary and only involves trigonometric identities and euclidean plane geometry.

There is also a proof by Alain Connes, based on affine geometry techniques. Of course it is a bit more technical, but again it involves math that doesn't go much beyond the high school level, and could be appreciated by the most gifted students

$\endgroup$
7
  • 1
    $\begingroup$ Thanks. Actually this example was given by Rota in a famous paper about the phenomenology of mathematical beauty as an example of a theorem that is surprising but not beautiful (in response to Hardy's claim that beauty arises from a feeling of surprise.) Of course, it is open to debate... Thanks for the Connes reference, I didn't know about that. $\endgroup$
    – Manya
    Sep 8, 2011 at 11:40
  • 1
    $\begingroup$ See also Conway's proof (linked at the Wikipedia article). $\endgroup$
    – Todd Trimble
    Sep 8, 2011 at 12:05
  • 2
    $\begingroup$ @Manya: Well, I guess it's a matter of taste. Personally, I do find this theorem beautiful. Thank you for the remark $\endgroup$ Sep 8, 2011 at 12:40
  • 1
    $\begingroup$ But the question was asking for a beautiful proof, not a beautiful theorem. $\endgroup$
    – euklid345
    Sep 9, 2011 at 4:51
  • 1
    $\begingroup$ P.S. (@ eucklid) Good point. Rota actually claimed that neither the theorem nor any of the proofs of it (thus far?) are beautiful. $\endgroup$
    – Manya
    Sep 19, 2011 at 8:16
6
$\begingroup$

Those are pretty nice, and can be done at a fairly low level (say, from 12 years old onwards) :

  • The proof of the formula "half base times height" for the area of a triangle, by first considering a right triangle and completing a rectangle, then considering an arbitrary triangle and breaking it in two along an height (two cases : inside(+) or outside(-)) : it is a nice example of how mathematicians treat the general case by reduction to particular cases ;
  • Euclid's proof of the pythagorean theorem using the previous formula, as in this animation.
$\endgroup$
1
  • $\begingroup$ Good example:I remember being excited by this when I understood why this worked at school. $\endgroup$ Mar 12, 2012 at 9:21
6
$\begingroup$

Using Euler's formula ($F-E+V=2$, as mentioned earlier), it can be proved that the graphs $K_5$ and $K_{3,3}$ are not planar. I think these proof and also the proof of Euler's formula are simple enough to be understood by an interested high school student.

$\endgroup$
0
5
$\begingroup$

In the general game "Poset Chomp" the first player always has a winning strategy. The proof is a one-line strategy stealing argument, hence nonconstructive. In fact, a winning strategy is unknown in most cases, which makes the result interesting and mysterious. For a good quick account see here.

$\endgroup$
6
  • $\begingroup$ I don't expect middle school students to be familiar with posets or abstract games. $\endgroup$ Sep 8, 2011 at 20:33
  • $\begingroup$ @Douglas: Read carefully the link. Special cases of "Poset Chomp", e.g. the one where the players choose divisors of a given integer do not need any knowledge of posets or abstract games: it can be played by a 10-year old. $\endgroup$
    – GH from MO
    Sep 8, 2011 at 22:21
  • $\begingroup$ @Douglas: Quote from the link: "Schuh published in 1952 his "game of divisors" ([1]). Here the partially ordered set is that of all divisors of a fixed number N, with x below y when y|x." $\endgroup$
    – GH from MO
    Sep 8, 2011 at 22:23
  • 1
    $\begingroup$ @Douglas: Another quote from the link: "David Gale reinvented this game ([2,2a]). His version is that of the m-by-n chocolate bar, where square (0,0) (say, the lower left-hand corner) is poisoned, and players take turns eating a "chomp": a square (a,b) together with all squares to the left and/or above it." $\endgroup$
    – GH from MO
    Sep 8, 2011 at 22:25
  • 5
    $\begingroup$ @GH: I'm aware of the history of chomp and of simple posets and the strategy-stealing argument. Have you worked with middle school students who were not selected to be competitors in a math competition? What percentage do you think will understand and be impressed by a nonconstructive existence result about an abstract game they have not seen before? When I tell members of the general public about mathematics, I try to relate it to concrete objects and situations I think they know beforehand. $\endgroup$ Sep 9, 2011 at 0:08

Your Answer

By clicking “Post Your Answer”, you agree to our terms of service and acknowledge you have read our privacy policy.

Not the answer you're looking for? Browse other questions tagged or ask your own question.