73
$\begingroup$

I've just been asked for a good example of a situation in maths where using infinity can greatly shorten an argument. The person who wants the example wants it as part of a presentation to the general public. The best I could think of was Goodstein sequences: if you take a particular instance of Goodstein's theorem, then the shortest proof in Peano arithmetic will be absurdly long unless the instance is very very small, but using ordinals one has a lovely short proof.

My question is this: does anyone have a more down-to-earth example? It doesn't have to be one where you can rigorously prove that using infinity hugely shortens the shortest proof. Just something where using infinity is very convenient even though the problem itself is finite. (This is related to the question asked earlier about whether finite mathematics needs the axiom of infinity, but it is not quite the same.)

A quick meta-question to add: when I finally got round to registering for this site, I lost the hard-earned reputation I had gained as a non-registered user. I am now disgraced, so to speak. Is that just my tough luck?

$\endgroup$
11
  • 3
    $\begingroup$ I take it by "infinity" you don't mean "one-point compactification," which is also very useful in its own right. $\endgroup$ Nov 16, 2009 at 23:41
  • 5
    $\begingroup$ Alakazam ! $\endgroup$ Nov 17, 2009 at 0:54
  • 5
    $\begingroup$ See also mathoverflow.net/questions/551/… $\endgroup$ Nov 17, 2009 at 0:56
  • 8
    $\begingroup$ If I may make a request, it would be wonderful if our colleague Gowers changed his or her display name to First Last or First M. Last. $\endgroup$ Nov 20, 2009 at 21:35
  • 4
    $\begingroup$ Not the latter as I go by my middle name. Happy to make the change but have not managed to find where I can do it. (Please excuse my utter incompetence.) $\endgroup$
    – gowers
    Nov 21, 2009 at 16:29

31 Answers 31

60
$\begingroup$

A simple instance where infinity makes things simple is when proving that there exist transcendental numbers. Instead of having to come up with ways to tell a transcendental number from an algebraic one, you simply say «well, there are too many numbers for all of them to be algebraic, so there you have it».

$\endgroup$
3
  • $\begingroup$ That's a good example, but I think it may already form part of the presentation (when they talk about infinite sets and the work of Cantor). I'll check though. $\endgroup$
    – gowers
    Nov 16, 2009 at 23:46
  • 3
    $\begingroup$ The technique of proving that A\B is nonempty because |A| > |B| ought to offer many examples. Anyone have good examples where A is countably infinite, and B is finite but hard to get a concrete bound on? $\endgroup$ Nov 17, 2009 at 2:57
  • 6
    $\begingroup$ A related instance would be the existence of normal numbers, just by picking one at random. (It's cheating a little, though, since normality is inherently an infinitary concept.) Another is the existence of non-computable functions from the naturals to the naturals (since there are uncountably many functions, and countably many computable ones), though in this case it's not so hard to finitise the argument. $\endgroup$
    – Terry Tao
    Nov 17, 2009 at 22:59
53
$\begingroup$

One of my colleagues (Silver) gave a general purpose talk last week. You can't untie a knot $A$ by tying another knot $B$ into the rope.

Proof. First show that the connect-sum of long knots is abelian by sliding $A$ through $B$. Then form the decreasing-in-size connect-sum $(AB)(AB)(AB) \dots$. Regroup: $A(BA)(BA) \dots$. If $AB = 0$, i.e. the unknot, then the left-hand grouping is the unknot. The right-hand-grouping is $A$ since $BA=AB=0$. Since $A$ is a knot, then $A=0$, a contradiction. QED

The nice thing about the argument is that it uses an infinitely long piece of rope, an infinite sequence of knots, and the knots shrink to being infinitesimally small.

$\endgroup$
5
  • 5
    $\begingroup$ Another spin on this is that all tame knots in co-dimension larger than two are topologically trivial. So if you apply this argument to Haefliger's non-trivial smooth knots (S^3 in S^6, say -- and these do have inverses in the smooth category) you see a "nice" example of how the topological and smooth category differ. $\endgroup$ Nov 17, 2009 at 3:36
  • 6
    $\begingroup$ This is essentially the Mazur swindle, right? $\endgroup$ Nov 17, 2009 at 4:11
  • 11
    $\begingroup$ The Mazur swindle is so good it gets 1st and 2nd place. :) $\endgroup$ Nov 17, 2009 at 6:08
  • 1
    $\begingroup$ Yes, and according to the wikipedia link above, it is used in Rolfsen to give the proof that I indicated. $\endgroup$ Nov 17, 2009 at 21:07
  • 1
    $\begingroup$ The argument shows that if A has a connected sum-inverse, then it is isotopic to the unknot by some continuous isotopy. I would like to know that if A has a connected sum inverse, then it is isotopic to the unknot by some smooth isotopy. Is there a variation of the argument that shows that? $\endgroup$ Apr 19, 2012 at 20:12
28
$\begingroup$

A down-to-earth example: the Stirling approximation to the factorial.

At least the derivation given on the Wikipedia page compares the explicit sum representation of $\ln(n!)$ with a Riemann sum approximation of the integral $\int \ln(x) \mathrm{d}x$. One only needs finite integers to talk about factorials. On the other hand, one needs at least a couple of kinds of infinities for the real analysis involved in describing the integral and accompanying error estimates. However, the simplicity of this derivation of the Stirling formula is undeniable.

$\endgroup$
26
$\begingroup$

Hindman's theorem is an example where (slightly?) higher infinity simplifies a proof greatly. The theorem asserts that, if you partition the set of positive integers into finitely many subsets, then there is an infinite set $H$ of positive integers such that all sums of finitely many members of $H$ lie in the same piece of your partition. (If, like me, you consider 0 to be such a finite sum, of 0 terms, then explicitly exclude it from the conclusion, so that the statement makes sense.) Hindman's original proof worked in the natural domain of the result, second-order arithmetic, but Hindman himself has suggested that this proof is such as would be useful for torturing students. There is a non-torturous proof, due to Galvin and Glazer, but an essential part of that proof involves considering a semigroup whose points are ultrafilters on the natural numbers - so we're well beyond second-order arithmetic. In fact, a key lemma in the proof depends on applying Zorn's Lemma to a certain poset of subsemigroups of this semigroup. So we're even farther "up" from the natural domain of the theorem. (By the way, the reason for "slightly?" in the first line of this answer is that, although we're well above where most mathematicians like to work, we're nowhere near what set theorists would call "large cardinals".)

$\endgroup$
23
$\begingroup$

Consider the following question of Erdos and Hajnal:

Question (Erdos-Hajnal) Is there a finite $K_4$-free graph which, when the edges colored by $2$ colors, always contains a monocolored triangle.

The question is answered by Shelah, and his argument uses infinity. I think this is the only known proof.

Shelah's proof of the theorem is simply as follows:

He constructs a forcing extension which adds a graph $X$ with the same property but with $\aleph_0$ colors. Then $X$ has the edge-coloring property for $2$ colors, as well. Then using compactness theorem, $X$ must contain a finite subgraph $Y$ with the same property. As forcing cannot create new finite graphs, $Y$ is already present in the ground model. By Godel, $ZFC$ proves that there is such a graph.

$\endgroup$
3
  • 2
    $\begingroup$ very nice, mohammad golshani! $\endgroup$
    – Rahman. M
    Mar 12, 2015 at 16:50
  • 5
    $\begingroup$ There are other proofs of this. Lin Lu has a construction on 9697 vertices see Linyuan Lu. (2008) Explicit Construction of Small Folkman Graphs. SIAM Journal on Discrete Mathematics 21:4, 1053-1060. available at epubs.siam.org/doi/pdf/10.1137/070686743 $\endgroup$ Mar 12, 2015 at 17:42
  • 2
    $\begingroup$ @KristalCantwell As indicated in the title of the paper you cited, this result was first proved (constructively and more generally) by Jon Folkman; it is the result $f(3,3)=3$ in his 1970 paper Graphs with Monochromatic Complete Subgraphs in Every Edge Coloring. $\endgroup$
    – bof
    Mar 27, 2019 at 22:06
18
$\begingroup$

The Mazur Swindle? See the Wikipedia page, for example.

$\endgroup$
3
  • $\begingroup$ Most of the applications aren't what I would call "finite problems"; probably the closest is the proof of the Cantor-Bernstein-Schroeder theorem given here: terrytao.wordpress.com/2009/10/05/mazurs-swindle $\endgroup$ Nov 16, 2009 at 23:43
  • $\begingroup$ Is there a popular text that describes the Mazur Swindle? $\endgroup$ Feb 23, 2022 at 17:56
  • 1
    $\begingroup$ @AntonPetrunin: suitable for undergrads? I don't know of anything, but perhaps there is. The Wikipedia page looks like a perfect sketch that could be fleshed out into a short article. I wouldn't be surprised if there was a nice Notices article on this topic. If there isn't, there should be. en.wikipedia.org/wiki/Eilenberg%E2%80%93Mazur_swindle $\endgroup$ Feb 23, 2022 at 19:12
17
$\begingroup$

The axiom system PRA of "primitive-recursive arithmetic" is finitistic, but it has been known for a few decades that it has the same set of $\Pi^0_1$ consequences as the infinitistic theory $\text{WKL}_0$ of second-order arithmetic. In particular, there is a primitive recursive function $f$ that turns a formal proof of a $\Pi^0_1$ statement in $\text{WKL}_0$ into a proof in PRA. Roughly put, a $\Pi^0_1$ formula says that all natural numbers have some particular property, depending on the formula, where the property can be stated using a formula in the language of rings with no quantifiers.

The advantage of working in $\text{WKL}_0$ is that the proofs can be much shorter. I think this was always suspected, but Caldon and Ignjatovic recently established (pdf) a formal superexponential lower bound for $f$, at least on an infinite set of formulas. Their result is phrased for a different infinitary system, $I\Sigma^0_1$, that lies between PRA and $\text{WKL}_0$. $I\Sigma^0_1$ is a fragment of Peano arithmetic, unlike $\text{WKL}_0$; its main difference from PRA is that $I\Sigma^0_1$ allows direct universal quantification over the set of natural numbers during the proof, while PRA does not.

In their paper, the set of formulas for the lower bound is explicitly laid out. These may not be particularly concrete, because they relate to consistency statements.

If we expand PRA to allow for existential quantification, we can get a slightly larger theory in which $\Pi^0_2$ statements can be expressed. It is known that $\text{WKL}_0$ is still conservative over this larger theory for $\Pi^0_2$ statements. In 1994, Kikuchi and Tanaka (pdf) gave a nice example of how this could be used to show that the second incompleteness theorem is provable in PRA, by using model-theoretic, infinitary methods in $\text{WKL}_0$ and relying on the conservation result.

$\endgroup$
14
$\begingroup$

Nachum Dershowitz has lots of work on termination proofs and many of these make use of ordinal arithmetic. In particular, if you have a computer program that moves from one state to another, and you can label each state with an ordinal, and you know that the ordinal for each new state is strictly less than the ordinal for the previous state, then the program must eventually terminate. Goodstein's Theorem is a special case, showing termination for a highly contrived sequence. But more realistic examples arise when studying termination of rewrite rules.

A basic example is proving termination of the rewrite rules that define symbolic differentiation. Like in the case of Goodstein's theorem you have the same problem that differentiation might take a step forward in decreasing the powers in a monomial, but then takes what looks like a giant step back as applying the Leibniz rule to a product can lead to many more terms. But Floyd came up with a neat labelling by ordinals resulting in a really simple termination proof. See example 8 here for example.

Also see the conclusion in this paper.

Apparently the idea for this originated with Alan Turing.

$\endgroup$
14
$\begingroup$

I'm not sure whether this is more "down-to-earth" but take 0-1 laws of random graphs. The probability of a first order property to be satisfied by a random graph distributed $G(n,p)$ tends to 0 or 1.

The proof is by showing that the limit theory satisfied by these graphs has only one countable model, and is therefore complete.

$\endgroup$
11
$\begingroup$

If you want a situation where there is only an infinite proof then there is the strengthened finite Ramsey theorem which is not provable in PA but can be deduced from the infinite Ramsey theorem. see:

http://en.wikipedia.org/wiki/Paris%E2%80%93Harrington_theorem

$\endgroup$
11
$\begingroup$

The Poincare conjecture might qualify as an example. One can phrase this conjecture in a purely combinatorial fashion, involving finite simplicial complexes, but the only known proof of this conjecture uses Ricci flow, and in particular the various monotonicity properties of this flow which are proven by integration by parts, as well as the repeated use of limits. In principle, this proof can be finitised by working with finite element versions of Ricci flow, but it would definitely lengthen the proof substantially. (On the other hand, Perelman's proof is not exactly short to begin with, so this may not be the best example...)

$\endgroup$
0
10
$\begingroup$

Every finite partial order has a maximal element.

One can prove that by induction and looking at cases, but I think the most natural way is to argue that if there is no maximal element, one must have a strictly increasing infinite sequence, which is impossible.

$\endgroup$
2
  • 1
    $\begingroup$ The statement that "nonempty + no maximal element" implies the existence of an infinite chain actually uses the dependent choice axiom, so it's perhaps not so natural. On the other hand, the proof by induction is very simple. $\endgroup$ Dec 4, 2012 at 10:17
  • 1
    $\begingroup$ But since the poset is finite, it is well-orderable, and so DC is not needed here. I think Michael's argument is sound, and simple: there must be a maximal element, since otherwise you could keep going up forever. $\endgroup$ Aug 13, 2013 at 4:03
10
$\begingroup$

There is a theorem that if a set of edge colored square tiles can be used to tile the positive orthant then it can also be used to tile the entire plane. (The edges are colored and in the edge-to edge tiling you need to match the colors).

The easy argument (using infinity) is: draw a rooted tree where the level $k$ vertices are all possible tilings of a $(2k-3) \times (2k-3)$ squares (and there is a root at level 1). This is a tree with each vertex has a finite number of children. The fact about the positive quadrant asserts that there are vertices of arbitrary large level, and Konig's theorem then asserts that there is an infinite path which then gives the tiling of the plane. (Proving Konig's theorem is based on moving from a vertex with infinitely many descendants to one of his children with infinitely many descendants.)

I am not sure that this qualifies as an answer since I am not aware of any "finite" proof and there is something infinite in the statement itself. But there may be special cases where a tiling of the positive orthant be explicitly presented, and it will allow a finite proof which can be very complicated.

$\endgroup$
8
$\begingroup$

Is there anything wrong with the simplest case - mathematical induction? If I want to show that the sum of the first million cubes is whatever it is, much easier to prove the general formula then specialise to n = 1000000.

Now I see from above Qiaochu Yuan has discounted induction as an answer. But why? Induction relies on the natural numbers being a certain initial algebra, i.e., being taken as an entity, which happens to have infinite cardinality.

$\endgroup$
3
  • 3
    $\begingroup$ I'm not sure I count induction either. It seems to me that one is not relying on infinity in a serious way to sum the first million cubes: one is saying, "If it's valid for 1 then it's valid for 2, and if it's valid for 2 then it's valid for 3, and ... and if it's valid for 999999 then it's valid for 1000000." That is a ridiculously long proof, but induction allows one to shorten it by giving a rule for when it's OK to put in the "..." (A more formal way of making the point is that the axiom of infinity isn't part of first-order Peano arithmetic.) $\endgroup$
    – gowers
    Nov 18, 2009 at 10:01
  • $\begingroup$ Induction does not reach actual infinty, but could be viewed as an example of reasoning with potential infinity. Particularly note that from $IsFinite(0)=True$ and $IsFinite(x) −> isFinite(x+1)$ we prove by induction that all natural numbers are finite. $\endgroup$ Feb 2, 2012 at 3:48
  • 1
    $\begingroup$ Technically, ZFC doesn't have induction until the axiom of infinity is added. Point against this line of argument: most mathematicians don't really work in ZFC, but in a pidgin higher-order logic with sets, in which induction is probably axiomatic. Point for this line of argument: it probably matches the audience better, who (I would guess) don't distinguish between the potential infinity of induction and the completed infinity of the naturals. $\endgroup$ Apr 19, 2012 at 17:31
8
$\begingroup$

A good example is a solution to Hilbert's third problem: it is not possible to cut the unit cube into finitely many polyhedral pieces and reassemble then as the regular tetrahedron of unit volume. The shortest proof requires construction of a Dehn's invariant which, in turn, requires the existence of a Hamel basis of the set of real numbers regarded as a linear space over $\mathbb Q$. The existence of such Hamel basis follows from the axiom of choice.

There are variants of the proof that reqire working with finitely dimensional linear spaces only, but the argument mentioned above is more elegant and leads to more general results.

$\endgroup$
7
$\begingroup$

An example that I came to think of even though it doesn't have to involve infinity is the existence of a finite field with a specified number $q=p^n$ of elements. One can show by an enumerative argument (Möbius inversion) that there must be an irreducible polynomial of degree $n$ over $\mathbb{Z}_p$ (there simple aren't enough lower degree polynomials to factor all of them). If $f$ is such a polynomial, then $\mathbb{Z}_p[x]/f(x)$ provides a field of $q$ elements.

An alternative is to use the fact that every field has an algebraic closure, and to verify that the set of zeros of $x^q-x$ in the algebraic closure of $\mathbb{Z}_p$ is closed under the field operations (and that they are distinct).

It's true that we really only need the splitting field of $x^q-x$ which, as it turns out, consists of only those $q$ zeros, but that's something that emerges from the proof, and it might conceivably have been much bigger.

$\endgroup$
7
$\begingroup$

I once had a teaching assistant in calculus who admitted that he was unable to give an epsilon, delta proof that a the Heaviside stepfunction was not continuous. To take a slightly less trivial example, the proof that $y=x^2$ is not uniformly continuous using $\epsilon, \delta$ is somewhat involved for a freshman. Using an infinite number $H$, this becomes transparent. Consider the infinitesimal $e=1/H$. Then the points $x=H$ and $x=H+e$ are infinitely close, but the corresponding values of $y$ are not by a trivial calculation. This implies a failure of uniform continuity by general theory.

$\endgroup$
2
  • 2
    $\begingroup$ yet what's so involved in $x=n$ and $x'=n+1/n$ and $x'^2-x^2>2$ ? $\endgroup$ Mar 11, 2018 at 21:42
  • $\begingroup$ @PietroMajer, what's involved is getting the quantifier order right and that sort of thing. $\endgroup$ Mar 20, 2018 at 11:30
5
$\begingroup$

The precise asymptotics of the partition function $p(n)$ are, as far as I know, totally inaccessible by finitary tools. More generally the methods of complex analysis are a great way to obtain incredibly precise asymptotics on many finitary sequences of concrete interest, i.e. trees satisfying certain properties or sequences related to the running times of algorithms, etc.

A related example is comparing the elementary and non-elementary proofs of the prime number theorem; my recollection is that the elementary proof is quite a bit more involved.

$\endgroup$
5
  • $\begingroup$ The elementary proof is far tougher to slog through, although there's a bunch of background hidden in the complex-analytic proof which makes it probably about as hard to learn "from scratch." $\endgroup$ Nov 17, 2009 at 0:14
  • 1
    $\begingroup$ You can get $n! \geq n^n/e^n$ just by looking at the Taylor series of $e^x$, but that is perhaps still not hands on. $\endgroup$ Nov 17, 2009 at 20:53
  • 1
    $\begingroup$ You can get the inequality quite easily I think. The log of n! is the sum of log m from 1 to n, which is at least the integral from 1 to n of log x, which is nlogn-n. Done. $\endgroup$
    – gowers
    Nov 17, 2009 at 21:22
  • $\begingroup$ I guess what I meant to say is I don't know of a finitary proof, but now that I think about it, that might not be true. Hmm. $\endgroup$ Nov 17, 2009 at 22:41
  • $\begingroup$ Never mind; you can prove the statement by induction, so that was a bad example. $\endgroup$ Nov 17, 2009 at 22:51
5
$\begingroup$

This is probably neither the sort of infinity nor the sort of simplification you're looking for, but Bézout's theorem is a challenge to prove, let alone state, without a line at infinity. I don't know if it would be suitable for a general audience either, but in its favor, it has a very nice statement using both the infinite and the infinitesimal (i.e., contact multiplicity).

$\endgroup$
5
$\begingroup$

Perhaps fitting in here is Lindenstrauss's proof of Lyapunov's convexity theorem.

http://atlas-conferences.com/c/a/e/y/06.htm

"the slickest proof to end all proofs" (Halmos)

although the result is in $\mathbb{R}^n$, for the proof we go to an infinite-dimensional space

$\endgroup$
1
  • $\begingroup$ The link is broken. $\endgroup$
    – user76284
    Jan 9, 2021 at 20:19
5
$\begingroup$

Using Eilenberg-Maclane spaces (namely $K(\mathbb{Z},2)=\mathbb{P}^{\infty}(\mathbb{C})$) and cellular approximation,one can show that any 3-fold with a positive Betti number has a map to the sphere $S^2$ which is not nullhomotopic. I am not sure how "down to earth" this example is, but it shows that even when studying finite dimensional objects it is natural to consider infinite-dimensional things.

$\endgroup$
5
$\begingroup$

Short version of "there exist functions not computable by Turing machine" using infinities:

There are an uncountable number of predicates, but only a countable number of Turing machines, so there are some predicates that are not computable by a Turing machine.

(That's weaker than the usual proof of the uncomputability of the Halting Problem of course, because it doesn't tell you that "does this Turing machine halt?" is one of those predicates.)

$\endgroup$
5
$\begingroup$

This isn't particularly down-to-earth, but of course before February or March you had the situation where the density Hales-Jewett theorem (which can be phrased in completely finitary language) was only provable via ergodic-theory methods (which are infinitary.)

$\endgroup$
4
  • $\begingroup$ What does DHJ stand for? $\endgroup$
    – Todd Trimble
    Dec 3, 2012 at 18:26
  • $\begingroup$ @Todd: Density Hales-Jewett, I suppose: arxiv.org/abs/0910.3926 $\endgroup$
    – Martin
    Dec 3, 2012 at 18:52
  • $\begingroup$ Much obliged, user49437. I now recognize the phrase from peripheral awareness of a polymath project, but I wouldn't have been able to guess it. $\endgroup$
    – Todd Trimble
    Dec 3, 2012 at 20:26
  • $\begingroup$ @Todd: If it is any consolation: it took me a moment, too :-) $\endgroup$
    – Martin
    Dec 3, 2012 at 22:08
5
$\begingroup$

I am surprised no one has mentioned any examples from physics. Fluid mechanics for example obtains some beautiful results by assuming that a fluid is continuous, while at the molecular level it is a discrete object. An example from my physics highschool textbook:

Suppose you have a circular grain storage, with a diameter of 2 meters and 10 meters high. The grain weights about $X \ kg/m^3$. What is the pressure at two meters depth? at four meters depth?

The problem is easier to solve if we think of the grain as a fluid and apply, say, Pascal's principle.

$\endgroup$
4
$\begingroup$

I can't think of any good specific examples off the top of my head, but I'm sure that there must be lots of examples where you can much more easily prove something about an affine variety by adding "points at infinity", i.e. by looking at the projectivization of the variety.

$\endgroup$
5
  • 2
    $\begingroup$ Given that this is for "a presentation to the general public" - if this is in the UK then I'm pessimistic about even saying the words "affine variety" without losing half the attention of half the audience $\endgroup$
    – Yemon Choi
    Nov 17, 2009 at 18:57
  • 1
    $\begingroup$ There's somewhere where a general audience's eyes wouldn't start to glaze over at that point? In the immortal words of Liz Lemon, "I want to go to there." $\endgroup$ Nov 17, 2009 at 22:45
  • 1
    $\begingroup$ Of course you don't have to say the words "affine variety." The example that comes to mind is that projectivizing the circle turns the almost-isomorphism between the line and the circle into an isomorphism, but if you only care about finitary things (like Pythagorean triples) this doesn't really matter... $\endgroup$ Nov 17, 2009 at 22:50
  • $\begingroup$ @Harrison I was being slightly tongue in cheek there. Didn't say anything about what happened to the other half of the audience... $\endgroup$
    – Yemon Choi
    Nov 17, 2009 at 23:55
  • $\begingroup$ Yes, of course you don't say the words "affine variety". But it is not unreasonable to talk about A^1 and P^1, and then A^2 and P^2, at the very least. $\endgroup$ Nov 18, 2009 at 12:24
4
$\begingroup$

One way of generating very simple examples of this kind is to define a family of predicates $F(n)$ by recursion over the natural numbers. Then, there can usually be no proof of $\forall n. F(n)$ in Peano arithmetic, since $F(n)$ won't usually be a schema in the sense the first-order Peano induction axiom requires -- it can be a different formula at each numeral.

For example, one very simple example is to define

$F(0) = (0 = 0)$

$F(n+1) = (n+1 = n+1) \wedge F(n)$

Now, it's clearly the case that $\forall n. F(n)$. However, since it's a different tautology for each argument, Peano arithmetic can't prove this proposition correct (indeed, it's not even a formula of first-order logic). We can only prove each instance correct, which becomes an infinite number of (incredibly trivial) propositions to prove....

$\endgroup$
2
  • $\begingroup$ This seems to be a statement that you have to go beyond first order reasoning to prove the theorem, but that doesn't mean you are using infinity. $\endgroup$ Nov 18, 2009 at 20:35
  • 3
    $\begingroup$ This idea of a real use of infinity has appeared several times. Can you give an example of such a use? It seems like there is a more subtle idea in play than simple cardinality. (E.g., a topos model can be made of extremely "big" objects, but an argument in its internal language can be completely constructive and hence "small".) $\endgroup$ Nov 19, 2009 at 10:46
4
$\begingroup$

While browsing Wikipedia, I came across this statement:

"The introduction of projective techniques made many theorems in algebraic geometry simpler and sharper: For example, Bézout's theorem on the number of intersection points between two varieties can be stated in its sharpest form only in projective space".

I think this fact can be stated in simple terms without any explicit reference to algebraic geometry and projective space, just talking about intersections of graphics of polynomials, the product of their degrees, and "combing space to get the point of infinity" (or some better and longer explanation).

$\endgroup$
4
$\begingroup$

Maybe you want something more proof-theoretic, like Gödel's speed-up theorem. The example from the Wikipedia article says: the statement "This statement cannot be proved in Peano arithmetic in fewer than a googolplex symbols" is provable in Peano arithmetic, but the shortest proof has at least a googolplex symbols (explanation given there). On the other hand, the theory PA+CON(PA) gives a short proof.

Similarly, second-order arithmetic (since it proves CON(PA)) gives a short proof, and I guess that would count as "adding infinity". Or you could look at hereditarily finite set theory (ZFC with the axiom of infinity replaced by its negation). I'm pretty sure regular ZFC proves this theory consistent, so again the speed-up theorem works out about the same way as it does with PA.

$\endgroup$
2
$\begingroup$

Maybe "analytic number theory" belongs here as an answer. Certain number theory results are most simply proved using complex analysis.

$\endgroup$
1
$\begingroup$

Maybe you could use some complex transformation related to Möebius functions (of the kind "circles and lines are the same") to prove several 2D-geometric statements at once.

$\endgroup$

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.