Mathematical induction

Mathematical induction is a method of mathematical proof typically used to establish that a given statement is true of all natural numbers (positive integers). It is done by proving that the first statement in the infinite sequence of statements is true, and then proving that if any one statement in the infinite sequence of statements is true, then so is the next one.

The method can be extended to prove statements about more general well-founded structures, such as trees; this generalization, known as structural induction, is used in mathematical logic and computer science. Mathematical induction in this extended sense is closely related to recursion.

Mathematical induction should not be misconstrued as a form of inductive reasoning, which is considered non-rigorous in mathematics (see Problem of induction for more information). In fact, mathematical induction is a form of rigorous deductive reasoning.[1]

Contents

History

In 370 BC, Plato's Parmenides may have contained an early example of an implicit inductive proof.[2] The earliest implicit traces of mathematical induction can be found in Euclid's [3] proof that the number of primes is infinite and in Bhaskara's "cyclic method".[4] An opposite iterated technique, counting down rather than up, is found in the Sorites paradox, where one argued that if 1,000,000 grains of sand formed a heap, and removing one grain from a heap left it a heap, then a single grain of sand (or even no grains) forms a heap.

An implicit proof by mathematical induction for arithmetic sequences was introduced in the al-Fakhri written by al-Karaji around 1000 AD, who used it to prove the binomial theorem and properties of Pascal's triangle.

None of these ancient mathematicians, however, explicitly stated the inductive hypothesis. Another similar case (contrary to what Vacca has written, as Freudenthal carefully showed) was that of Francesco Maurolico in his Arithmeticorum libri duo (1575), who used the technique to prove that the sum of the first n odd integers is n2. The first explicit formulation of the principle of induction was given by Pascal in his Traité du triangle arithmétique (1665). Another Frenchman, Fermat, made ample use of a related principle, indirect proof by infinite descent. The inductive hypothesis was also employed by the Swiss Jakob Bernoulli, and from then on it became more or less well known. The modern rigorous and systematic treatment of the principle came only in the 19th century, with George Boole,[5] Charles Sanders Peirce,[6] Giuseppe Peano, and Richard Dedekind.[4]

Description

The simplest and most common form of mathematical induction proves that a statement involving a natural number n holds for all values of n. The proof consists of two steps:

  1. The basis (base case): showing that the statement holds when n is equal to the lowest value that n is given in the question. Usually, n = 0 or n = 1.
  2. The inductive step: showing that if the statement holds for some n, then the statement also holds when n + 1 is substituted for n.

The assumption in the inductive step that the statement holds for some n is called the induction hypothesis (or inductive hypothesis). To perform the inductive step, one assumes the induction hypothesis and then uses this assumption to prove the statement for n + 1.

The choice between n = 0 and n = 1 in the base case is specific to the context of the proof: If 0 is considered a natural number, as is common in the fields of combinatorics and mathematical logic, then n = 0. If, on the other hand, 1 is taken as the first natural number, then the base case is given by n = 1.

This method works by first proving the statement is true for a starting value, and then proving that the process used to go from one value to the next is valid. If these are both proven, then any value can be obtained by performing the process repeatedly. It may be helpful to think of the domino effect; if one is presented with a long row of dominoes standing on end, one can be sure that:

  1. The first domino will fall
  2. Whenever a domino falls, its next neighbor will also fall,

so it is concluded that all of the dominoes will fall, and that this fact is inevitable.

Axiom of induction

The basic assumption or axiom of induction is, in logical symbols,

(\forall P)[[P(0) \land ( \forall k \in \mathbb{N}) (P(k) \Rightarrow P(k%2B1))] \Rightarrow ( \forall n \in \mathbb{N} ) [ P(n) ]]

where P is any proposition and k and n are both natural numbers.

In other words, the basis P(0) being true along with the inductive case ("P(k) is true implies P(k + 1) is true" for all natural k) being true together imply that P(n) is true for any natural number n. A proof by induction is then a proof that these two conditions hold, thus implying the required conclusion.

This works because k is used to represent an arbitrary natural number. Then, using the inductive hypothesis, i.e. that P(k) is true, show P(k + 1) is also true. This allows us to "carry" the fact that P(0) is true to the fact that P(1) is also true, and carry P(1) to P(2), etc., thus proving P(n) holds for every natural number n.

Note that the first quantifier in the axiom ranges over predicates rather than over individual numbers. This is called a second-order quantifier, which means that the axiom is stated in second-order logic. Axiomatizing arithmetic induction in first-order logic requires an axiom schema containing a separate axiom for each possible predicate. The article Peano axioms contains further discussion of this issue.

Example

Mathematical induction can be used to prove that the following statement, which we will call P(n), holds for all natural numbers n.

0 %2B 1 %2B 2 %2B \cdots %2B n = \frac{n(n %2B 1)}{2}\,.

P(n) gives a formula for the sum of the natural numbers less than or equal to number n. The proof that P(n) is true for each natural number n proceeds as follows.

Basis: Show that the statement holds for n = 0.
P(0) amounts to the statement:

0 = \frac{0\cdot(0 %2B 1)}{2}\,.

In the left-hand side of the equation, the only term is 0, and so the left-hand side is simply equal to 0.
In the right-hand side of the equation, 0·(0 + 1)/2 = 0.
The two sides are equal, so the statement is true for n = 0. Thus it has been shown that P(0) holds.

Inductive step: Show that if P(k) holds, then also P(k + 1) holds. This can be done as follows.

Assume P(k) holds (for some unspecified value of n). It must then be shown that P(k + 1) holds, that is:

(0 %2B 1 %2B 2 %2B \cdots %2B k )%2B (k%2B1) = \frac{(k%2B1)((k%2B1) %2B 1)}{2}

Using the induction hypothesis that P(k) holds, the left-hand side can be rewritten to:

\frac{k(k %2B 1)}{2} %2B (k%2B1)\,.

Algebraically:


\begin{align}
\frac{k(k %2B 1)}{2} %2B (k%2B1) & = \frac {k(k%2B1)%2B2(k%2B1)} 2 \\
& = \frac{(k%2B1)(k%2B2)}{2} \\
& = \frac{(k%2B1)((k%2B1) %2B 1)}{2}.
\end{align}

thereby showing that indeed P(k + 1) holds.

Since both the basis and the inductive step have been proved, it has now been proved by mathematical induction that P(n) holds for all natural n. Q.E.D.

Variants

In practice, proofs by induction are often structured differently, depending on the exact nature of the property to be proved.

Starting at some other number

If we want to prove a statement not for all natural numbers but only for all numbers greater than or equal to a certain number b then:

  1. Showing that the statement holds when n = b.
  2. Showing that if the statement holds for n = mb then the same statement also holds for n = m + 1.

This can be used, for example, to show that n2 ≥ 3n for n ≥ 3. A more substantial example is a proof that

{n^n \over 3^n} < n! < {n^n \over 2^n}\mbox{ for }n\ge 6.

In this way we can prove that P(n) holds for all n ≥1, or even n ≥−5. This form of mathematical induction is actually a special case of the previous form because if the statement that we intend to prove is P(n) then proving it with these two rules is equivalent with proving P(n + b) for all natural numbers n with the first two steps.

Building on n = 2

In mathematics, many standard functions, including operations such as "+" and relations such as "=", are binary, meaning that they take two arguments. Often these functions possess properties that implicitly extend them to more than two arguments. For example, once addition a + b is defined and is known to satisfy the associativity property (a + b) + c = a + (b + c), then the ternary addition a + b + c makes sense, either as (a + b) + c or as a + (b + c). Similarly, many axioms and theorems in mathematics are stated only for the binary versions of mathematical operations and relations, and implicitly extend to higher-arity versions.

Suppose that we wish to prove a statement about an n-ary operation implicitly defined from a binary operation, using mathematical induction on n. Then it should come as no surprise that the n = 2 case carries special weight. Here are some examples.

Example: product rule for the derivative

In this example, the binary operation in question is multiplication (of functions). The usual product rule for the derivative taught in calculus states:

(fg)' = f'g %2B g'f. \!

or in logarithmic derivative form

(fg)'/ (fg) = f'/f %2B g'/g. \!

This can be generalized to a product of n functions. One has

(f_1 f_2 f_3 \cdots f_n)' \!
= (f_1' f_2 f_3 \cdots f_n) %2B (f_1 f_2' f_3 \cdots  f_n) %2B (f_1 f_2 f_3' \cdots  f_n) %2B \cdots %2B(f_1 f_2 \cdots f_{n-1} f_n').

or in logarithmic derivative form

(f_1 f_2 f_3 \cdots f_n)'/(f_1 f_2 f_3 \cdots f_n) \!
= (f_1'/f_1) %2B (f_2'/f_2) %2B (f_3'/f_3) %2B \cdots %2B (f_n'/f_n).

In each of the n terms of the usual form, just one of the factors is a derivative; the others are not.

When this general fact is proved by mathematical induction, the n = 0 case is trivial,(1)' = 0 \! (since the empty product is 1, and the empty sum is 0). The n = 1 case is also trivial, f_1' = f_1' \!. And for each n ≥ 3, the case is easy to prove from the preceding n − 1 case. The real difficulty lies in the n = 2 case, which is why that is the one stated in the standard product rule.

Example: Pólya's proof that there is no "horse of a different color"

In this example, the binary relation in question is an equivalence relation applied to horses, such that two horses are equivalent if they are the same color. The argument is essentially identical to the one above, but the crucial n = 1 case fails, causing the entire argument to be invalid.

In the middle of the 20th century, a commonplace colloquial locution to express the idea that something is unexpectedly different from the usual was "That's a horse of a different color!". George Pólya posed the following exercise: Find the error in the following argument, which purports to prove by mathematical induction that all horses are of the same color:

In general, the basis case is trivial (as any horse is the same color as itself), and the inductive step is correct in all cases n ≥ 2. However, the logic of the inductive step is incorrect going from n = 1 to n+1 = 2, because the statement that "the two sets overlap" is false (there are only two horses). Indeed, the n = 1 case is clearly the crux of the matter; if one could prove the n = 1 case directly, then all higher cases would follow from the inductive hypothesis.

Induction on more than one counter

It is sometimes desirable to prove a statement involving two natural numbers, n and m, by iterating the induction process. That is, one performs a basis step and an inductive step for n, and in each of those performs a basis step and an inductive step for m. See, for example, the proof of commutativity accompanying addition of natural numbers. More complicated arguments involving three or more counters are also possible.

Infinite descent

Another variant of mathematical induction – the method of infinite descent – was one of Pierre de Fermat's favorites. This method of proof works in reverse, and can assume several slightly different forms. For example, it might begin by showing that if a statement is true for a natural number n it must also be true for some smaller natural number m (m < n). Using mathematical induction (implicitly) with the inductive hypothesis being that the statement is false for all natural numbers less than or equal to m, we can conclude that the statement cannot be true for any natural number n.

Complete induction

Another variant, called complete induction (or strong induction or course of values induction), says that in the second step we may assume not only that the statement holds for n = m but also that it is true for all n less than or equal to m.

Complete induction is most useful when several instances of the inductive hypothesis are required for each inductive step. For example, complete induction can be used to show that

 F_n = \frac{\varphi^n - \psi^n}{\varphi - \psi}

where Fn is the nth Fibonacci number, φ = (1 + √5)/2 (the golden ratio) and ψ = (1 − √5)/2 are the roots of the polynomial x2 − x − 1. By using the fact that Fn + 2Fn + 1 + Fn for each n ∈ N, the identity above can be verified by direct calculation for Fn + 2 if we assume that it already holds for both Fn + 1 and Fn. To complete the proof, the identity must be verified in the two base cases n = 0 and n = 1.

Another proof by complete induction uses the hypothesis that the statement holds for all smaller n more thoroughly. Consider the statement that "every natural number greater than 1 is a product of prime numbers", and assume that for a given m > 1 it holds for all smaller n > 1. If m is prime then it is certainly a product of primes, and if not, then by definition it is a product: m = n1 n2, where neither of the factors is equal to 1; hence neither is equal to m, and so both are smaller than m. The induction hypothesis now applies to n1 and n2, so each one is a product of primes. Then m is a product of products of primes; i.e. a product of primes.

This generalization, complete induction, is equivalent to the ordinary mathematical induction described above. Suppose P(n) is the statement that we intend to prove by complete induction. Let Q(n) mean P(m) holds for all m such that 0 ≤ mn. Then Q(n) is true for all n if and only if P(n) is true for all n, and a proof of P(n) by complete induction is just the same thing as a proof of Q(n) by (ordinary) induction.

Transfinite induction

The last two steps can be reformulated as one step:

  1. Showing that if the statement holds for all n < m then the same statement also holds for n = m.

This is in fact the most general form of mathematical induction and it can be shown that it is not only valid for statements about natural numbers, but for statements about elements of any well-founded set, that is, a set with an irreflexive relation < that contains no infinite descending chains.

This form of induction, when applied to ordinals (which form a well-ordered and hence well-founded class), is called transfinite induction. It is an important proof technique in set theory, topology and other fields.

Proofs by transfinite induction typically distinguish three cases:

  1. when m is a minimal element, i.e. there is no element smaller than m
  2. when m has a direct predecessor, i.e. the set of elements which are smaller than m has a largest element
  3. when m has no direct predecessor, i.e. m is a so-called limit-ordinal

Strictly speaking, it is not necessary in transfinite induction to prove the basis, because it is a vacuous special case of the proposition that if P is true of all n < m, then P is true of m. It is vacuously true precisely because there are no values of n < m that could serve as counterexamples.

Proof of mathematical induction

The principle of mathematical induction is usually stated as an axiom of the natural numbers; see Peano axioms. However, it can be proved in some logical systems. For instance, it can be proved if one assumes:

To derive simple induction from these axioms, we must show that if P(n) is some proposition predicated of n, and if:

then P(n) holds for all n.

Proof. Let S be the set of all natural numbers for which P(n) is false. Let us see what happens if we assert that S is nonempty. Well-ordering tells us that S has a least element, say t. Moreover, since P(0) is true, t is not 0. Since every natural number is either zero or some n+1, there is some natural number n such that n+1=t. Now n is less than t, and t is the least element of S. It follows that n is not in S, and so P(n) is true. This means that P(n+1) is true, and so P(t) is true. This is a contradiction, since t was in S. Therefore, S is empty.

It can also be proved that induction, given the other axioms, implies well-ordering.

See also

Notes

  1. ^ Suber, Peter. "Mathematical Induction". Earlham College. http://www.earlham.edu/~peters/courses/logsys/math-ind.htm. Retrieved 26 March 2011. 
  2. ^ Mathematical Induction: The Basis Step of Verification and Validation in a Modeling and Simulation Course
  3. ^ Proof due to Euclid http://primes.utm.edu/notes/proofs/infinite/euclids.html http://www.mathsisgoodforyou.com/conjecturestheorems/euclidsprimes.htm http://www.hermetic.ch/pns/proof.htm
  4. ^ a b Cajori (1918), p. 197

    "The process of reasoning called "Mathematical Induction" has had several independent origins. It has been traced back to the Swiss Jakob (James) Bernoulli, the Frenchman B. Pascal and P. Fermat, and the Italian F. Maurolycus. [...] By reading a little between the lines one can find traces of mathematical induction still earlier, in the writings of the Hindus and the Greeks, as, for instance, in the "cyclic method" of Bhaskara, and in Euclid's proof that the number of primes is infinite."

  5. ^ "It is sometimes required to prove a theorm which shall be true whenever a certain quantity n which it involves shall be an integer or whole number and the method of proof is usually of the following kind. 1st. The theorem is proved to be true when n = 1. 2ndly. It is proved that if the theorem is true when n is a given whole number, it will be true if n is the next greater integer. Hence the theorem is true universally. . .. This species of argument may be termed a continued sorites" (Boole circa 1849 Elementary Treatise on Logic not mathematical pages 40–41 reprinted in Grattan-Guinness, Ivor and Bornet, Gérard (1997), George Boole: Selected Manuscripts on Logic and its Philosophy, Birkhäuser Verlag, Berlin, ISBN 3-7643-5456-9
  6. ^

References

Introduction
History
  • Acerbi, F. (2000). "Plato: Parmenides 149a7-c3. A Proof by Complete Induction?". Archive for History of Exact Sciences 55: 57–76. doi:10.1007/s004070000020. 
  • Bussey, W. H. (1917). "The Origin of Mathematical Induction". The American Mathematical Monthly 24 (5): 199–207. doi:10.2307/2974308. JSTOR 2974308. 
  • Cajori, Florian (1918). "Origin of the Name "Mathematical Induction"". The American Mathematical Monthly 25 (5): 197–201. doi:10.2307/2972638. JSTOR 2972638. 
  • "Could the Greeks Have Used Mathematical Induction? Did They Use It?". Physis XXXI: 253–265. 1994. 
  • Freudenthal, Hans (1953). "Zur Geschichte der vollständigen Induction". Archives Internationales d'Histiore des Sciences 6: 17–37. 
  • Katz, Victor J. (1998). History of Mathematics: An Introduction. Addison-Wesley. ISBN 0321016181.
  • Peirce, C. S. (1881). "On the Logic of Number". American Journal of Mathematics 4 (1–4): pp. 85–95. doi:10.2307/2369151. JSTOR 2369151. MR1507856. http://books.google.com/books?id=LQgPAAAAIAAJ&jtp=85.  Reprinted (CP 3.252-88), (W 4:299-309).
  • Rabinovitch, Nachum L. (1970). "Rabbi Levi Ben Gershon and the origins of mathematical induction". Archive for History of Exact Sciences 6 (3): 237–248. doi:10.1007/BF00327237. 
  • Rashed, Roshdi (1972). "L'induction mathématique: al-Karajī, as-Samaw'al" (in French). Archive for History of Exact Sciences 9 (1): 1–21. doi:10.1007/BF00348537. 
  • Shields, Paul (1997). "Peirce’s Axiomatization of Arithmetic". In Houser et al.. Studies in the Logic of Charles S. Peirce. 
  • Ungure, S. (1991). "Greek Mathematics and Mathematical Induction". Physis XXVIII: 273–289. 
  • Ungure, S. (1994). "Fowling after Induction". Physis XXXI: 267–272. 
  • Vacca, G. (1909). "Maurolycus, the First Discoverer of the Principle of Mathematical Induction". Bulletin of the American Mathematical Society 16 (2): 70–73. doi:10.1090/S0002-9904-1909-01860-9. 
  • Yadegari, Mohammad (1978). "The Use of Mathematical Induction by Abū Kāmil Shujā' Ibn Aslam (850-930)". Isis 69 (2): 259–262. doi:10.1086/352009. JSTOR 230435.