Number-Theoretic Algorithms - Foundations of Algorithms (2015)

Foundations of Algorithms (2015)

Chapter 11 Number-Theoretic Algorithms

Suppose Bob wants to send Alice a secret love note over the Internet, but he is afraid some of his friends might intercept the message and read it. If he could encode the message so that it appears as gibberish and only Alice could decode the gibberish back to the original message, he would not need to fear his friends intercepting the message. Number-theoretic algorithms can help Bob develop a system for doing this. We discuss such algorithms next.

Number theory is the branch of mathematics concerned with the properties of the integers. Number-theoretic algorithms are algorithms that solve problems involving the integers. For example, a number-theoretic algorithm might find the greatest common divisor of two integers. After reviewing basic number theory in Section 11.1, we present Euclid’s algorithm for finding the greatest common divisor in Section 11.2. Next, Section 11.3 reviews modular arithmetic, Section 11.4 shows an algorithm for solving modular linear equations, and Section 11.5 develops an algorithm for computing modular powers. Section 11.6 concerns algorithms for determining whether a number is prime. An important application of number-theoretic algorithms is in cryptography, which is the discipline concerned with encrypting a message sent from one party to another, so that someone who intercepts the message will not be able to decode it. In Section 11.7 we present the Rivest-Shamir-Adelman (RSA) public-key cryptosystem, which is a system that does this.

Before proceeding, we note that in this chapter we revert back to developing algorithms as we did in Chapters 2 through 6. However, unlike the methods presented in those chapters, number-theoretic algorithms are concerned with solving a certain type of problem (namely ones involving the integers); they are not algorithms that share a common technique.

11.1 Number Theory Review

We review some basic elements of number theory.

• 11.1.1 Composite and Prime Numbers

The set

image

is called the set of integers. For any two integers n, h Z, we say h divides n, denoted h|n, if there is some integer k such that n = kh. If h|n, we also say n is divisible by h, n is a multiple of h, and h is a divisor or factor of n.

Example 11.1

We have that 4|20 since 20 = (5) (4). The integer 3 does not divide 20 since there is no integer k such that 20 = (k) (3).

Example 11.2

The divisors of 12 are

image

An integer n > 1 whose only positive divisors are 1 and n is called a prime number or simply a prime. A prime number has no factors. An integer n > 1 that is not prime is called a composite number. A composite number has at least one factor.

Example 11.3

The first 10 primes are

image

Example 11.4

The integer 12 is composite because it has the divisors 2, 3, 4, and 6. The integer 4 is composite because it has the divisor 2.

• 11.1.2 Greatest Common Divisor

If h|n and h|m, then h is called a common divisor of n and m. If n and m are not both zero, the greatest common divisor of n and m, denoted gcd(n, m), is the largest integer that divides both of them.

Example 11.5

The positive divisors of 24 are

image

and the positive divisors of 30 are

image

So the common positive divisors of 24 and 30 are

image

and the value of gcd(24, 30) is 6.

We have the following theorems concerning common divisors:

image Theorem 11.1

If h|n and h|m, then for any integers i and j

image

Proof: Since h|n and h|m, there exist integers k and l such that n = kh and m = lh. Therefore,

image

which means h| (in + jm).

Before proceeding, we need more definitions. For any two integers n and m, where m ≠ 0, the quotient q of n divided by m is given by

image

and the remainder r of dividing n by m is given by

image

The remainder is denoted n mod m. It is not hard to see that if m > 0 then 0 ≤ r < m, and if m < 0 then m < r ≤ 0. So we have

image

The division algorithm says not only that the integers q and r in Equality 11.1 exist but also that they are unique.

Example 11.6

The following table shows some quotients and remainders:

image

image Theorem 11.2

Let n and m be integers, not both 0, and let

image

That is, d is the smallest positive linear combination of n and m. Then

image

Proof: Let i and j be the integers that yield the minimal value d. That is, d = in + jm. Furthermore, let q and r be the quotient and remainder respectively, of dividing n by d. Then owing to Equality 11.1 and the fact that d is positive,

image

We therefore have

image

which means that r is a linear combination of n and m. Since d is the smallest positive linear combination of n and m and r < d, we conclude r = 0, which means d|n. Similarly, d|m. Therefore, d is a common divisor of m and n, which means

image

Since the gcd(n, m) divides both n and m, and d = in + jm, Theorem 11.1 implies gcd(n, m)|d. We conclude

image

Combining these last two inequalities, we have d = gcd(n, m).

Example 11.7

We have that gcd(12, 8) = 4, and

image

image Corollary 11.1

Suppose n and m are integers, not both 0. Then every common divisor of n and m is a divisor of gcd(n, m). That is, if h|n and h|m, then

image

Proof: By the preceding theorem, gcd(n, m) is a linear combination of n and m. The proof now follows from Theorem 11.1.

Example 11.8

As shown in Example 11.5, the common positive divisors of 24 and 30 are

image

and the value of gcd(24, 30) is 6. As the previous corollary implies, 1, 2, 3, and 6 all divide 6.

image Theorem 11.3

Suppose we have integers n ≥ 0 and m > 0. If we let r = nmodm, then

image

If n = 0, r = m and the equality obviously holds. Otherwise, we will show that gcd(n, m) and gcd(m, r) each divide each other. It is left as an exercise to show that two positive integers divide each other if and only if they are equal, which will complete the proof of the theorem. First we show gcd(n, m)| gcd(m, r). If we let d′ = gcd(n, m), then d′|n and d′|m. Furthermore,

image

where q is the quotient of n divided by m, which means n is a linear combination of m and r. Theorem 11.1 therefore implies d″|n. Since d″|m and d″|n, Corollary 11.1 implies

image

which completes the proof of this direction. Next show gcd(m, r)| gcd(n, m). If we let d″ = gcd(m, r), then d″|m and d″|r. Furthermore,

image

where q is the quotient of n divided by m, which means n is a linear combination of m and r. Theorem 11.1 therefore implies d″|n. Since d″|m and d|n, Corollary 11.1 implies

image

which completes the proof of this direction.

Example 11.9

According to the previous theorem,

image

because 16 = 64 mod 24. We can continue in this manner to determine gcd(64, 24). That is,

image

• 11.1.3 Prime Factorization

Every integer greater than one can be written as a unique product of primes. We next develop theory that proves this assertion.

Two integers n and h, not both zero, are called relatively prime if gcd(n, h) = 1.

Example 11.10

The integers 12 and 25 are relatively prime because gcd(12, 25) = 1. The integers 12 and 15 are not relatively prime because gcd(12, 15) = 3.

image Theorem 11.4

If h and m are relatively prime, and h divides nm, then h divides n. That is, gcd(h, m) = 1 and h|nm implies h|n.

Proof: Theorem 11.2 implies there are integers i and j such that

image

Multiplying this equality by n yields

image

Clearly, h divides the first term on the left-hand side of this equality, and, since h|nm, h also divides the second term. Therefore, h divides the left-hand side, which means h divides n.

Example 11.11

The integers 9 and 4 are relatively prime, 9|72, and 72 = 18 × 4. As the previous theorem implies, 9|18.

image Corollary 11.2

Given integers n, m, and prime integer p, if p|nm, then p|n or p|m (inclusive).

Proof: The proof follows easily from Theorem 11.4.

The theorem that follows is called the unique factorization theorem and the fundamental theorem of arithmetic.

image Theorem 11.5

Every integer n > 1 has a unique factorization as a product of prime numbers. That is,

image

where p1 < p2 < · · · < pj are primes, and this representation of n is unique. The integer ki is called the order of pi in n.

Proof: We use induction to prove such a representation exists.

Induction base: We have 2 = 21.

Induction hypothesis: Suppose all integers m such that 2 ≤ m < n can be written as a product of prime numbers.

Induction step: If n is prime, then n = n1 is our representation. Otherwise, n is composite, which means there exists integers m, h > 1 such that

image

Clearly m, h < n. Therefore, by the induction hypothesis, m and h can each be written as a product of primes. That is,

image

Since n = mh,

image

We obtain our desired representation by grouping primes that are equal and arranging the terms according to increasing values of the primes. This completes the induction proof.

The fact that the product is unique follows from Corollary 11.2 and is left as an exercise.

Example 11.12

We have that

image

The previous theorem says the representation on the right-hand side of this equality is unique.

image Theorem 11.6

The gcd(n, m) is a product of the primes that are common to n and m, where the power of each prime in the product is the smaller of its orders in n and m.

Proof: The proof is left as an exercise.

Example 11.13

We have 300 = 22 × 31 × 52 and 1,125 = 32 × 53. So gcd(300, 1,125) = 20 × 31 × 52 = 75.

• 11.1.4 Least Common Multiple

A concept similar to that of the greatest common divisor is that of the least common multiple. If n and m are both nonzero, the least common multiple of n and m, denoted lcm(n, m), is the smallest positive integer that they both divide.

Example 11.14

The lcm(6, 9) = 18 because 6|18, 9|18, and there is no smaller positive integer that they both divide.

image Theorem 11.7

The lcm(n, m) is a product of the primes that are common to n and m, where the power of each prime in the product is the larger of its orders in n and m.

Proof: The proof is left as an exercise.

Example 11.15

We have 12 = 22 ×31 and 45 = 32 ×51. So lcm(12, 45) = 22 ×32 ×51 = 180.

11.2 Computing the Greatest Common Divisor

Theorem 11.6 gives us a straightforward way to compute the greatest common divisor of two such integers. We simply find the unique factorizations for the two integers, determine which primes they have in common, and determine the greatest common divisor to be a product whose terms are these common primes, where the power of each prime in the product is the smaller of its orders in the two integers. Example 11.13 illustrated this. Another example follows:

Example 11.16

We have

image

So

image

The problem with the previous technique is that it is not easy to find the unique factorization of an integer. You would have had some difficulty factoring the integers in the previous example if we had not given you the factorization. Now imagine the difficulty if the integers had 25 digits instead of 7. Indeed, no one has ever found a polynomial-time algorithm for determining the factorization of an integer. Next we show a more efficient way to compute the greatest common divisor.

• 11.2.1 Euclid’s Algorithm

Theorem 11.1 gives us a straightforward method for determining the greatest common divisor of two integers. Example 11.9 illustrates the method. Namely, to find the gcd(n, m) we recursively apply the equality in the theorem until m = 0, and then we return n. This method is called Euclid’s algorithm because it was developed by Euclid around 300 B.C. The algorithm follows:

Algorithm 11.1

Euclid’s Algorithm

Problem: Compute the greatest common divisor of a positive integer and a nonnegative integer.

Inputs: a positive integer n and a nonnegative integer m.

Outputs: the greatest common divisor of n and m.

image

Example 11.17

We find the greatest common divisor of the numbers in Example 11.16 using Algorithm 11.1.

image

It seemed pretty easy for us to compute the greatest common divisor using Euclid’s algorithm. Let’s analyze Algorithm 11.1 to see just how easy it is. First we need a lemma and a theorem.

image Lemma 11.1

If n > m ≥ 1 and the call gcd(n, m) (in Algorithm 11.1) results in k recursive calls where k ≥ 1, then

image

where fk is the kth number in the Fibonacci sequence.

Proof: The proof is by induction.

Induction base: Suppose the call gcd(n, m) results in 1 recursive call. Since m ≥ 1

image

Since n > m, we have n ≥ 2, which means

image

Induction hypothesis: Assume the lemma is true if k − 1 recursive calls are made.

Induction step: Show the lemma is true if k ≥ 2 recursive calls are made. We need to show

image

The first recursive call is gcd(m, n mod m). Since there are k recursive calls in all, this call must require k − 1 recursive calls. Since k ≥ 2, there is at least one more recursive call, which means n mod m ≥ 1. Therefore, since m > n mod m, the conditions of the induction hypothesis are satisfied, which means

image

So we’ve arrived at one of the inequalities we needed to show. Towards showing the other, we have

image

where q is the quotient of dividing n by m. Since n > m, q ≥ 1. Inequality 11.2 and Equality 11.3 therefore imply

image

This completes the proof.

image Theorem 11.8

(Lamé) For every integer k ≥ 1 if n > m ≥ 1 and m ≤ fk, the kth number in the Fibonacci sequence, then the call gcd(n, m) (in Algorithm 11.1) results in less than k recursive calls.

Proof: The proof follows immediately from the previous lemma.

Next we analyze Algorithm 11.1. Recall from Section 9.2 that in numeric algorithms, the number(s) that is the input is not the input size. Rather, the input size is the number of characters it takes to write the input. If we use binary encoding, the input size is the number of bits it takes to encode the number(s), which is about equal to the lg of the number(s).

image Analysis of Algorithm 11.1

image Worst-Case Time Complexity (Euclid’s Algorithm)

Basic operation: one bit manipulation in the computation of a remainder.

Input size: the number of bits s it takes to encode n and the number of bits t it takes to encode m. That is,

image

Without loss of generality, we will analyze the case where 1 ≤ m < n. That is, if m = n, there will be no recursive calls, while if m > n, the first recursive call will be gcd(m, n), which means the first argument is larger.

We will not determine the exact time complexity. Rather we will first show the number of recursive calls Calls(s, t) is Θ(t), and then we will find a bound for the worst-case time complexity.

First we show Calls(s, t) is in O(t). Assuming m ≥ 2, let fk be the number in the Fibonacci sequence such that

image

In Example B.9 in Section B.2.1 we show that

image

Either k − 1 is odd or k − 2 is odd. Without loss of generality, assume k − 1 is odd. Owing to Equality 11.5 and Inequality 11.4, we have

image

Owing to Equality 11.4, Theorem 11.8 implies

image

Inequalities 11.6 and 11.7 imply that

image

It is left as an exercise to show that the call gcd(fk+1, fk) requires exactly k − 1 recursive calls. We conclude that

image

where W calls(s, t) is the worst-case number of recursive calls for input size s, t.

For each recursive call, we compute one remainder. It is left as an exercise to show that, if we use the standard long division algorithm (as discussed in Section 2.6) to compute the remainder, the worst case number of bit manipulations needed to compute r = n mod m for m < n is bounded above by

image

where q is the quotient of dividing n by m, and c is a constant. We will show if r > 0, then, for the input size sufficiently larger, the worst case number of bit manipulations needed to compute r is bounded above by

image

To that end, since q = (n − r)/m and 1 ≤ r < m,

image

This last inequality along with Relation 11.9 establishes Bound 11.10. Owing to this bound, the total number of bit manipulations needed to compute all remainders in all recursive calls is bounded above by

image

Relation 11.8 implies the number of recursive calls is bounded above by dt, where d is a constant greater than 0. This means the number of terms in Bound 11.11 is bounded above by dt. Therefore, since n > m > r > m mod r > · · · (where the dots denote the remaining terms in Bound 11.11), we conclude from Bound 11.11 that

image

• 11.2.2 An Extension to Euclid’s Algorithm

Theorem 11.1 entails that there are integers i and j such that

image

Knowledge of these integers will be important to our algorithm for solving modular linear equations in the next section. Next we modify Algorithm 11.1 to also produce these integers. In this version we make gcd a variable because so doing makes the proof of the correctness of the algorithm more transparent.

image Algorithm 11.2

Euclid’s Algorithm 2

Problem: Compute the greatest common divisor of a positive integer and a nonnegative integer.

Inputs: a positive integer n and a nonnegative integer m.

Outputs: the greatest common divisor gcd of n and m, and integers i and j such that gcd = in + jm.

image

Algorithm 11.2 clearly has the same time complexity as Algorithm 11.1. So we need only prove it is correct. Before doing this, we show an example of applying it.

Example 11.18

Table 11.1 illustrates the flow of Algorithm 11.2 when the top-level call is

image

The values returned at the top level are gcd = 6, i = −2, and j = 3.

Next we prove Algorithm 11.2 is correct.

Table 11.1 The values determined by Algorithm 11.2 when n = 42 and m = 30. The top-level call is labeled 0; the three recursive calls are labeled 1–3. The arrows show the order in which the values are determined.

image

image Theorem 11.9

The values of i and j returned by Algorithm 11.2 are integers such that

image

Proof: The proof is by induction.

Induction base: In the last recursive call m = 0, which means the gcd(n, m) = n. Since the values of i and j are assigned values 1 and 0, respectively, in that call, we have that

image

Induction hypothesis: Assume in the kth recursive call the values determined for i and j are such that

image

Then the values returned by that call for i′ and j′ [to the (k − 1)st recursive call] are values such that

image

Induction step: We have for the (k − 1)st call that

image

The second to the last equality is due to the induction hypothesis, and the last equality is owing to Theorem 11.3.

Notice that the value Algorithm 11.2 returns for j in the final recursive call can be any integer. We choose 0 for simplicity. Choosing a different integer yields a different (i, j) pair with the gcd(n, m) = in + jm.

11.3 Modular Arithmetic Review

We develop modular arithmetic within the context of group theory. So first we review that theory.

• 11.3.1 Group Theory

A closed binary operation on a set S is a rule for combining two elements of S to yield another element of S. We have the following definition:

Definition

A group G = (S, ∗) is a set S together with a closed binary operation on S satisfying the following:

1. * is associative. That is, for all a, b, cS

image

2. There is an identity element e in S. That is, for each aS

image

3. For each aS there exists an inverse element a′ such that

image

Example 11.19

The integers Z with addition constitute a group. The identity element is 0, and the inverse of a is −a.

Example 11.20

The nonzero real numbers with multiplication constitute a group. The identity element is 1, and the inverse of a is 1/a.

Example 11.21

Let S = (a, b, e) and assign

image

It is left as an exercise to show (S, ∗) is a group.

Can an element in a group have more than one inverse? The theorem that follows shows the answer is no.

image Theorem 11.10

The inverse of each element in a group is unique.

Proof: Suppose a′ and a″ are both inverses of a. Then

image

However,

image

We conclude from the previous two equalities that a′ = a″.

image Theorem 11.11

If there exists elements a and b in a group such that a b = a or b a = a, then b is the identity element e in the group.

Proof: If a b = a then

image

where a′ is the inverse of a. However,

image

Combining these two equalities proves this case. The other case is proved in the same way.

The last theorem establishes the result that the identity element in a group is unique.

A group G = (S, ∗) is said to be commutative (or abelian) if for all a, b S,

image

A group G = (S, ∗) is said to be finite if S contains a finite number of elements. The groups in Examples 11.19, 11.20, and 11.21 are all commutative, while only the one in Example 11.21 is finite.

• 11.3.2 Congruency Modulo n

We start with a definition.

Definition

Let m and k be integers and n be a positive integer. If n|(m − k) we say m is congruent to k modulo n, and we write

image

Example 11.22

Since 5|(33 − 18),

image

Recall m mod n gives the remainder when m is divided by n. The following theorem shows we can recognize congruence using the mod function.

image Theorem 11.12

We have that m ≡ k mod n if and only if

image

Proof: The proof is left as an exercise.

Example 11.23

We already know 33 18 mod 5. As the preceding theorem implies, 33 mod 5 = 18 mod 5 = 3.

The following theorem concerning congruency will be needed in Section 11.7 when we develop the RSA cryptosystem.

image Theorem 11.13

Suppose n1, n2, … , nj are pairwise relatively prime and

image

Then for all integer m and k,

image

if and only if, for 1 ≤ i ≤ j,

image

Proof: Suppose, for 1 ≤ i ≤ j, that

image

Then there exist integers h1, h2, … , hj such that

image

It is left as an exercise to use Theorem 11.4 to show that this equality and the fact that n1, n2, … , nj are pairwise relatively prime implies

image

where c is an integer. We then have

image

Since c is an integer, this means m ≡ k mod n.

In the other direction, if m ≡ k mod n, then there is an integer h such that

image

Therefore,

image

which means, for each i, m−k is a multiple of ni. This completes the proof.

Example 11.24

The integers 2, 5, and 9 are pairwise relatively prime, and

image

Since 2 × 5 × 9 = 90, the previous theorem implies

image

Next we show congruency modulo n is an equivalence relation.

image Theorem 11.14

For any positive integer n, congruency modulo n is an equivalence relation on the set of all integers. That is,

1. (reflexivity)

image

2. (symmetry)

image

3. (transitivity)

image

Proof: The proof is left as an exercise.

The set of all integers congruent to m modulo n is called the equivalence class modulo n containing m. Since congruency is an equivalence relation, any integer in a given class determines the class. For a given integer m, it is the set

image

Example 11.25

The equivalence class modulo 5 containing 13 is

image

We denote the equivalence class modulo n containing m by [m]n. So the equivalence class in the preceding example is denoted [13]5. It can also be denoted [3]5, [8]5, and so on. Usually we denote an equivalence class using the smallest nonnegative integer in the class. So the equivalence class in the preceding example is usually denoted [3]5. The set of all equivalence classes modulo n is denoted Zn. That is,

image

Example 11.26

We have

image

For two members of Zn we define addition as follows:

image

For this definition to make sense, the result must not depend on which members of [m]n and [k]n we choose. We show that next. We need to show that if

image

To that end, there exists integers i and j such that

image

So

image

which means s + t [m + k]n.

Example 11.27

We have

image

image Theorem 11.15

For every positive integer n, (Zn, +) is a finite commutative group.

Proof: The associativity and commutivity of + follows readily from the associativity and commutivity of + for the integers. The identity element is [0]5. The additive inverse of [m]n is [n − m]n = [−m]n since

image

Example 11.28

Consider the group (Z5, +). Recall

image

The additive inverse of [1]5 is [5 − 1]5 = [4]5. Note that

image

Similarly, the additive inverse of [2]5 is [3]5. Note that

image

For two members of Zn we define multiplication as follows:

image

For this definition to make sense, the result must not depend on which members of [m]n and [k]n we choose. We need to show that if

image

It is left as an exercise to show this.

Example 11.29

We have

image

Now (Zn, ×) is not always a group because there exists n (namely nonprimes) for which not every element in Zn has a multiplicative inverse. That is, if we let [1]n be the identity element, there exists [m]n such that there is no k for which [m]n × [k]n = [1]n.

Example 11.30

Consider

image

Suppose [6]9 has a multiplicative inverse [k]9. Then

image

which means there exists an integer i such that

image

Therefore, Theorem 11.2 implies 1 is the gcd(6, 9), which clearly is not the case.

The problem in the previous example is that 6 and 9 are not relatively prime. It seems if we include only numbers relatively prime to n, we will have a group. The next theorem proves this is so. First we develop notation for this set. Let

image

Clearly if n is prime, then Zn = Zn − {[0]n}. It is not hard to see that one member of [m]n is relatively prime to n if and only if all members are. So we need only look at the first n − 1 integers to determine the members of Zn.

Example 11.31

We have that

image

image Theorem 11.16

For every positive integer n, (Zn, ×) is a finite commutative group.

Proof: Owing to Corollary 11.2 any divisor of m×k must be either a divisor of m or of k. Therefore, if m and k are each relative prime to n, m × k is relatively prime to n, which means × is a closed binary operation on Zn. The associativity and commutivity of × follows readily from the associativity and commutivity of × for the integers. The identity element is [1]n. If m is relatively prime to n, Theorem 11.2 says there exists integers i and j such that

image

which means

image

So [j]n is the inverse of [m]n.

Example 11.32

For the group (Z9, ×), we have the following multiplicative inverses:

image

image Theorem 11.17

The number of elements in Zn is given by Euler’s totient function, which is

image

where the product is over all primes that divide n, including n if n is prime.

Proof: The proof can be found in Graham et al. (1989).

Note, if p is prime, that

image

Example 11.33

The number of elements in Z*9 is

image

which can be verified by counting in Example 11.31.

Example 11.34

The number of elements in Z60 is

image

Example 11.35

Since 7 is prime, the number of elements in Z7 is

image

• 11.3.3 Subgroups

If G = (S, ∗) is a group, S′ S, and G′ = (S′, ∗) is a group, then G′ is said to be a subgroup of G. It is called a proper subgroup if S′ ≠ S.

Example 11.36

For E, the set of even integers, and Z, the set of integers, (E, +) is a proper subgroup of (Z, +).

Before proceeding, we develop some useful notation. Given a group G = (S, ∗), and a S with inverse a′, for k > 0 we define

image

Suppose we have a group G = (S, ∗) and a subset S′ S such that for every a, b S, a b S. Then S′ is said to be closed relative to .

image Theorem 11.18

Suppose we have a finite group G = (S, ∗), and a nonempty subset S′ S, which is closed relative to . Then (S′, ∗) is a subgroup of G.

Proof: Clearly, associativity holds. Let a S. It is left as an exercise to show that since G is finite, there exist integers k, m ≥ 1 such that

image

Multiplying both sides of this equality by the inverse of ak yields

image

where e is the identity element. Since S′ is closed, this means e is in S′. Next we show the inverse of every a S is in S′. As just shown, for every such a there is an m such that e = am. If m = 1, a = e, which means the inverse of a is e, and we already know e is in S′. Otherwise,

image

which means am−1 is the unique inverse of a. However, since S′ is closed, am−1 is in S′. This completes the proof.

image Theorem 11.19

(Lagrange) Suppose we have a finite group G = (S, ∗), and a subgroup G′ = (S′, ∗) of G. Then

image

where |S| denotes the number of elements in S.

Proof: The proof can be found in Jacobson (1951).

Example 11.37

Consider the group (Z12, +). It is left as an exercise to show that if

image

then (S′, +) is a subgroup of (Z12, +). We have |S′| = 4, |Z12| = 12, and 4|12 as the previous theorem implies.

image Corollary 11.3

If G′ = (S′, ∗) is a proper subgroup of G = (S, ∗), then

image

Proof: The proof follows immediately from the preceding theorem.

Suppose we have a finite group G = (S, ∗), and a S. Let

image

Clearly, ⟨a⟩ is closed under . So, because of Theorem 11.18, (⟨a⟩, ∗) is a subgroup of G. This group is called the subgroup generated by a. If the subgroup generated by a is G, we call a a generator of G.

Example 11.38

Consider the group (Z6, +). We have

image

The previous example illustrates that once we obtain the identity element when generating the subgroup, we can stop because we will only repeat items already generated. The next theorem obtains this result. First we need a definition. Given a group, the order of group element a, denotedord(a), is the least positive integer t such that at = e, where e is the identity element.

image Theorem 11.20

Suppose we have a finite group G = (S, ∗) and a S. Let t = ord(a). Then the subgroup ⟨a⟩ generated by a consists of the following t distinct elements:

image

Proof: First we show these elements are distinct. Suppose for 1 ≤ k < j ≤ t we have

image

Since j > k,

image

where i ≥ 1: Owing to these last two equalities and Theorem 11.11, we have ai = e. Since i < t, we have a contradiction. Next we show there are no other elements in this subgroup. If k > t, then there exist positive integers q and r such that

image

We then have

image

If r = 0, ar = e = at; otherwise 1 ≤ r < t. This completes the proof.

Owing to the previous theorem, we have

image

where t = ord(a), Note that the number of elements in ⟨a⟩ is ord(a).

Example 11.39

Consider the group (Z6, +). We have

image

and

image

Clearly,

image

So [1]6 is a generator of Z6.

Example 11.40

Consider the group (Z9, ×). Recall

image

We have

image

It is left as an exercise to show that

image

So [2]9 is a generator of Z9.,

Example 11.41

Consider the group (Z7, ×). Recall that

image

We have

image

image Corollary 11.4

Suppose we have a finite group G = (S, *) and aS. Let t = ord(a). Consider the sequence

image

Two elements ai and aj in this sequence are equal if and only if

image

This implies the sequence is cyclic with period t, and t is the smallest period for the sequence. That is, for 0 ≤ k < t and all i ≥ 0,

image

and t is the smallest number with this property.

Proof: The proof follows from the previous theorem and is left as an exercise.

Example 11.42

Consider the group (Z5, ×). This table shows the powers of [4]5:

image

whereas this table shows the powers of [3]5:

image

In this group, ord([4]5) = 2 and ord([3]5) = 4. Note the cyclical nature of the powers relative to these orders as implied by Corollary 11.4. Note further that [3]5 is a generator of Z5.

image Corollary 11.5

Suppose we have a finite group G = (S, ∗) with identity element e. Then for all a S,

image

where |S| denotes the number of elements in S.

Proof: From Theorem 11.20, the number of elements in the subgroup ⟨a⟩ generated by a is ord(a). Therefore, from Theorem 11.19, |S| = ord(a) × k for some integer k, and so

image

which completes the proof.

image Theorem 11.21

(Euler) For any integer n > 1, for all [m]nZn,

image

Proof: The proof follows immediately from Theorem 11.17 and Corollary 11.5.

Example 11.43

Consider the group (Z20, ×). We have that

image

and

image

as the previous theorem implies.

image Theorem 11.22

(Fermat) If p is prime, then for all [m]p Zp,

image

Proof: The proof follows from the previous theorem and the fact that ϕ(p) = p − 1 if p is prime.

Example 11.44

Consider the group image. We have that

image

as the previous theorem implies.

image 11.4 Solving Modular Linear Equations

Next we discuss solving the modular equation

image

for x, where x is an equivalence class modulo n, and m, n > 0. We will apply this result in Section 11.7 when we develop the RSA cryptosystem.

Let ⟨[m]n⟩ be the subgroup generated by [m]n relative to the group (Zn+). Since ⟨[m]n⟩= {[0]n, [m]n, [2m]n,}, Equation 11.12 has a solution if and only if

image

Example 11.45

Consider the group (Z8, +). Since

image

the equation

image

has a solution if and only if [k]8 is [0]8, [6]8, [4]8, or [2]8. For example, solutions to

image

are x = [2]8 and x = [6]8.

The following theorem tells us precisely the elements of the set ⟨[m]n

image Theorem 11.23

Consider the group (Zn, +). For any [m]n image Zn, we have that

image

where d = gcd(n, m). This means

image

Proof: First show ⟨[d]nimage ⟨[m]n⟩. Owing to Theorem 11.2, there exists integers i and j such that

image

Owing to the previous equality, for any integer k,

image

which means [kd]n = [kim]n, and therefore [kd]n image ⟨[m]n⟩, We conclude that ⟨[d]nimage ⟨[m]n⟩.

Next show ⟨[m]nimage ⟨[d]n⟩. Since d|m, there is an integer i such that m = id. Therefore, for any integer k,

image

which means [km]n image ⟨[d]n⟩. We conclude ⟨[m]nimage ⟨[d]n⟩.

The final result follows from the results just established and Theorem 11.20.

image Corollary 11.6

The equation

image

has a solution if and only if d|k, where d = gcd(n, m).

Furthermore, if the equation has a solution, it has d solutions.

Proof: As mentioned at the beginning of this section, the equation has a solution if and only if

image

Since Theorem 11.23 implies

image

this means the equation has a solution if and only d|k′ for some kimage [k]n. Since [k]n is an equivalence class modulo n and d|n, clearly d|k′ for one kimage [k]n if and only if it does so for all members of [k]n. This proves the first part of the corollary.

As to the second part, according to Theorem 11.23, ord([m]n) = n/d. Therefore, owing to Corollary 11.4, the sequence

image

has period n/d and the first n/d items are distinct. This means, if [k]n image ⟨[m]n⟩, that [k]n appears exactly d times in the set

image

Clearly, each of these appearances is owing to a different member of Zn. This completes the proof.

image Corollary 11.7

The equation

image

has a solution for every equivalence class [k]n if and only if gcd(n, m) = 1 Furthermore, if this is the case, each [k]n has a unique solution.

Proof: The proof follows immediately from Corollary 11.6.

image Corollary 11.8

The equivalence class [m]n has a multiplicative inverse modulo n if and only if gcd(n, m) = 1. That is, the equation

image

has a solution if and only if gcd(n, m) = 1. Furthermore, if it has an inverse, that inverse is unique.

Proof: The proof follows immediately from Corollary 11.6. The uniqueness actually also follows from Theorem 11.16.

Example 11.46

Consider the group (Z8, +). Since gcd(8, 6) = 2, according to Theorem 11.23,

image

which agrees with our result in Example 11.45. Owing to Corollary 11.6, the equation

image

has exactly two solutions when [k]8 is any member of ⟨[6]8⟩. In Example 11.45, we noted the two solutions when [k]8 = [4]8 are x = [2]8 and x = [6]8.

Example 11.47

Consider again the group (Z8, +). Since gcd(8, 5) = 1, according to Theorem 11.23,

image

According to Corollary 11.7, the equation

image

has exactly one solution when [k]8 is any member of ⟨[5]8⟩. For example, when [k]8 = [3]8, the solution is x = [7]8.

Example 11.48

Consider Z9. Since gcd(9, 6) = 3, Corollary 11.8 implies [6]9 does not have a multiplicative inverse modulo 9. Since gcd(9, 5) = 1, Corollary 11.8 implies [5]9 does have a multiplicative inverse modulo 9. That inverse is [2]9.

image Theorem 11.24

Let d = gcd(n, m) and let i and j be integers such that

image

(We know from Theorem 11.2 that such i and j exist.) Suppose further d|k. Then the equation

image

has solution

image

Proof: Owing to Equality 11.13, we have

image

Since image is an integer, we can multiply both sides of this equality by image, yielding

image

which means

image

This proves the theorem.

Example 11.49

Consider the equation

image

We have gcd(8, 6) = 2,

image

and 2|4. Therefore, the previous theorem implies the equation

image

has solution

image

image Theorem 11.25

Suppose the equation

image

is solvable, x = [j]n is one solution, and d = gcd(n, m). Then the d distinct solutions of this equation are

image

Proof: Owing to Corollary 11.6, there are exactly d solutions. Clearly,

image

So the d modulo classes in Expression 11.14 are all distinct. We complete the theorem by showing each of these classes is a solution to the equation. Since [j]n is a solution of

image

Therefore, for l = 0, 1, · · ·, d − 1,

image

which means image is also a solution to the equation.

Example 11.50

In Example 11.49, we showed that one solution of

image

is [6]8. Since gcd(8, 6) = 2, the previous theorem implies the other solution is

image

Corollary 11.6, Theorem 11.24, and Theorem 11.25 enable us to write a simple algorithm for solving modular linear equations. That is, we first employ Corollary 11.6 to see if a solution exists. If one does, we then use Theorem 11.24 to find one solution and Theorem 11.25 to find the other solutions. The algorithm follows:

image Algorithm 11.3

Solve Modular Linear Equation

Problem: Find all solutions to a modular linear equation

Inputs: positive integers m and n, and integer k.

Outputs: if the equation [m]n x = [k]n is solvable, all solutions to it.

image

The input size in Algorithm 11.3 is the number of bits it takes to encode the input, which is given by

image

The time complexity includes the time complexity of Algorithm 11.2 (Euclid’s Algorithm 2), which we already know is O(st), plus the time complexity of the for-l loop. Since d can be as large as m or n, this time complexity is worst-case exponential in terms of the input size. However, we can do nothing about this since the problem requires that we compute and display an exponential number of values in the worst case.

image 11.5 Computing Modular Powers

For an element [m]n image Zn and nonnegative integer k, the problem of computing

image

is called the problem of computing modular powers. Examples 11.43 and 11.44 showed the computation of some modular powers. Another example follows.

Example 11.51

Consider Z20. We have

image

In Example 11.51 we determined the power simply by taking 7 to the 11th power. Clearly, this method has exponential time complexity in terms of the input size (approximately the logarithms of the numbers). Next we develop a more efficient algorithm.

The algorithm requires we represent k as a binary number. Let

image

be the ordered set of binary digits in that representation. For example, since the binary representation of 13 is 1101, if k = 13,

image

The algorithm now follows. We say that the algorithm uses the method of repeated squaring. The reason is obvious.

image Algorithm 11.4

Compute Modular Power

Problem: For an element [m]n image Zn and nonnegative integer k, compute ([m]n)k.

Inputs: positive integer n, and nonnegative integers m and k.

Outputs: ([m]n)k.

image

Next we show an example of applying Algorithm 11.4.

Example 11.52

Suppose

image

Since the binary representation of 45 is 101101,

image

Table 11.2 shows the state of a after each iteration of the for-i loop in Algorithm 11.4 given this input. The final value of a, namely [147]257, is the value of ([5]257)45.

In Table 11.2, ki is the value determined by the high-order j − i + 1 bits in the binary representation of k. That is,

image

Table 11.2 The state of a after each iteration of the for-i loop in Algorithm 11.4 when n = 257, m = 5, and k = 45. The third row shows the value determined by the high-order j − i + 1 bits in the binary representation of k.

image

For example,

image

Clearly, k0 = k. We will use these variables to prove the correctness of the algorithm, which we do next.

image Theorem 11.26

After each iteration of the for-i loop in Algorithm 11.4,

image

Since k0 = k, this means the final value of a is ([m]n)k.

Proof: The proof is by induction.

Induction base: We assume a minimal binary representation; so the high-order bit bj equals 1. Therefore,

image

Before entering the for-i loop, a = [1]n. Since bj equals 1, the if condition is true in the first iteration, which means the instruction a = a × [m]n executes. Therefore, we have

image

after the first iteration.

Induction hypothesis: Suppose after the iteration with index value i that

image

Induction step: We need to show that, after the iteration with index value (i − 1),

image

There are two cases: If bi−1 = 0, clearly

image

Since the condition in the if statement is false, only the instruction a = a2 changes the value of a. Since by the induction hypothesis the previous value of a is ([m]n)ki, we have

image

after this iteration.

If bi−1 = 1, clearly

image

Since the condition in the if statement is true, the instruction a = [m]n also executes. So in this case,

image

after this iteration.

It is straightforward that, if we let s be the number of bits it takes to encode the input, the time complexity of Algorithm 11.4 is in O(s3).

11.6 Finding Large Prime Numbers

Finding large prime numbers is necessary to the success of the RSA public-key cryptosystem, which is discussed in Section 11.7. After discussing searching for a large prime number, we show an algorithm for testing whether a number is prime.

• 11.6.1 Searching for a Large Prime

To find a large prime number, we randomly select integers of the appropriate size and test whether each selected integer is prime until one is found to be prime. When we use this method, an important consideration is the likelihood of finding a prime when an integer is chosen at random.

The prime distribution theorem, which we give next, enables us to approximate this likelihood.

The prime distribution function π(n) is the number of primes that are less than or equal to n. For example, π(12) = 5 since there are five primes, namely 2, 3, 5, 7, and 11, that are less than or equal to 12. The prime number theorem shown in Theorem 11.27, gives an approximation of π(n).

image Theorem 11.27

We have that

image

Proof: The proof can be found in Hardy and Wright (1960).

Owing to the previous theorem, for large values of n we can use n/ ln n as an estimate of the number of primes less than or equal to n. Therefore, if we randomly choose an integer between 1 and n according to the uniform distribution, the probability of it being prime is about

image

Example 11.53

If we randomly choose an integer between 1 and n = 1016 according to the uniform distribution, the probability of it being prime is about

image

Suppose we choose 200 such numbers at random. The probability of them all not being prime is then about

image

Example 11.54

If we randomly choose an integer between 1 and n = 10100 according to the uniform distribution, the probability of it being prime is about

image

Suppose we choose 200 such numbers at random. The probability of them all not being prime is then about

image

As noted at the beginning of this section, to find a large prime number we randomly select numbers in the appropriate range, and we then check them for being prime until one is found to be prime. We see from the previous examples that it should not take very long (relative to the size of the numbers in the range) to find a prime. Next we discuss how we check whether a number is prime.

image • 11.6.2 Checking if a Number Is Prime

At the beginning of Section 9.2 we showed a straightforward algorithm for determining if a number is prime. However, as discussed in that section, the algorithm has exponential-time complexity in the worst case. So it cannot be used to check large numbers. Note that this inefficient algorithm finds a factor of the number if it is composite. So the algorithm can be used (repeatedly) to factor a number.

Until recently no one had ever found a polynomial-time algorithm for the prime-checking problem, and many thought it to be NP-complete. The standard efficient algorithm for prime-checking was a Monte Carlo algorithm called the Miller-Rabin Randomized Primality Test, which appeared in Rabin (1980). If a number is prime, then this algorithm says the number is always prime. However, very rarely, the algorithm will declare a composite number to be prime. This happens so rarely, that the algorithm could essentially be counted upon for accuracy. The Miller-Rabin algorithm does not find a factor of the number if it determines the number is composite. So it cannot be used for factoring.

Finally, in 2002, Agrawal et al. succeeded in developing a polynomial-time algorithm for the prime-checking problem. That algorithm is described here.

A Polynomial-Time Algorithm

First we need further results from number theory.

Definition

Let f(x) and g(x) be polynomials with integral coefficients. If the coefficients of each power of x are congruent modulo n, we say that f(x) and g(x) are congruent modulo n, and we write

image

Example 11.55

We have

image

because

image

Example 11.56

We have

image

because

image

Missing powers are assumed to have coefficient 0.

Example 11.57

We have

image

because

image

Example 11.58

We have

image

because

image

This last example illustrates something interesting. Owing to Theorem 11.22, for all integers x,

image

but yet the polynomials x5 and x are not congruent modulo 5.

Next we show a theorem that gives a straightforward way to determine whether a number is prime. But first we need a lemma.

image Lemma 11.2

If n is prime then for all integers m,

image

Proof: We have that

image

Therefore,

image

Since n is prime, clearly for 1 < i ≤ n − 1,

image

Therefore, for each i the coefficient of xi in Equality 11.15 is congruent to 0 modulo n. So we only need to show

image

If n = 2

image

The last step is owing to the fact that either m or m + 1 is even. If n > 2, then n is odd and therefore

image

Now if m is a multiple of n, then m ≡ 0 mod n. Otherwise, owing to Theorem 11.22, (−mn−1 + 1) 0 mod n. This completes the proof.

image Theorem 11.28

Suppose m and n are relatively prime. Then n is prime if and only if

image

Proof: Owing to Lemma 11.2, the congruence is satisfied if n is prime.

In the other direction, suppose n is composite. Let q be a factor of n and let k be the order of q in n. It is left as an exercise to show image Since n is relatively prime to m, clearly qk is relatively prime to (−m)n−q. Therefore, image does not contain qk as a factor, which means it is not congruent to 0 modulo n. However, this expression is one of the coefficients in Equality 11.15. This completes the proof.

Example 11.59

The numbers 9 and 2 are relatively prime. As Theorem 11.28 implies,

image

since

image

Example 11.60

The numbers 9 and 4 are relatively prime and 4 is not prime. Theorem 11.28 implies

image

It is left as an exercise to expand the exponent and illustrate this result.

Theorem 11.28 yields a straightforward algorithm for determining if a number is prime. That is, given an integer n, we choose an integer m relatively prime to n and determine whether Congruence 11.16 is satisfied. However, since we need to evaluate n + 1 coefficients in the left-hand side of Congruence 11.16, this algorithm has exponential time complexity. We do a ‘trick’ to improve on this. However, first we need more definitions.

As is the case for integers, the mod function returns the remainder when one polynomial is divided by the other. That is,

image

is the polynomial that is the remainder when f(x) is divided by g(x). The following example illustrates this idea.

Example 11.61

Since

image

then

image

Now suppose f(x), g(x), and h(x) are polynomials with integral coefficients. If

image

we write

image

Example 11.62

We have that

image

and

image

So

image

Example 11.63

We have that

image

and

image

So

image

Example 11.64

We have that

image

and

image

So

image

The previous examples illustrate the truth of the following theorem:

image Theorem 11.29

Suppose n and r are prime. Then for all integers m,

image

Proof: The proof follows from Lemma 11.2 and is left as an exercise.

If the previous theorem said that n is prime if and only if Congruence 11.17 were satisfied, we would immediately be able to improve on using Congruence 11.16 to test whether a number was prime. That is, the computation of (xn − m) mod(xr 1) is trivial, and Knuth (1998) develops an algorithm for computing (x − m)n mod (xr 1), which uses fast Fourier multiplication and whose time complexity is in image. So, if Theorem 11.29 were ‘if and only if’ and r were small, we could fairly quickly replace the determination of Congruence 11.16 to that of a congruence involving far fewer coefficients. The following example illustrates this.

Example 11.65

Suppose we want to determine whether 11 is prime and Theorem 11.29 is ‘if and only if.’ Choose m = 4 and r = 3. We have

image

and

image

We now check whether

image

instead of checking whether

image

So we have reduced the problem of testing for 12 congruencies to that of testing for three.

The problem in this procedure is that Theorem 11.29 is not ‘if and only if’ even if we choose m relatively prime to n. That is, some composites will satisfy Congruence 11.17 for certain values of m and r, where m is relatively prime to n. So if we simply check whether the congruence is satisfied for particular values, a composite may pass the test. We solve this problem by choosing an appropriate r and checking for several values of m. In this way, no composites will be found to be prime, and we still have a polynomial-time algorithm. After presenting the algorithm, we obtain these results. Before we do this, we need some additional notation.

Recall that the order of group element a, denoted ord(a), is the least positive integer t such that at = e, where e is the identity element. Given the group image is then the least positive integer t such that ([n]r)t = [1]r. That is, it is the least positive integer such that

image

We call this the order of n modulo r and we denote it as ordr(n).

image Algorithm 11.5

Polynomial Determine Prime

Problem: Determine whether an integer is prime.

Inputs: an integer n > 1.

Outputs: if n is prime, true; if n is composite, false.

image

Correctness of the Algorithm

Next we prove the algorithm is correct. First we prove the algorithm always determines that a prime is prime.

image Theorem 11.30

If a prime number is the input to Algorithm 11.5, the algorithm returns true.

Proof: If n is prime, then for all r < n,

image

which means false cannot be returned in the first while loop. Since n and r are prime, owing to Theorem 11.29, false cannot be returned in the second

while loop. We conclude the algorithm exits with the last instruction, and that instruction returns true.

It is more difficult to prove the algorithm always determines a composite number to be composite. We do that next. First we need the following two lemmas, which generalize Lemma 11.2 and Theorem 11.29.

image Lemma 11.3

Suppose g(x) is a polynomial with integer coefficients and n is prime. Then

image

Proof: Let

image

Then

image

Furthermore, the coefficient of xi in [g(x)]n is

image

Case 1: i ≠ jn for any j. Then, as can be seen from Equality 11.18, the coefficient of xi in g(xn) is 0. Furthermore, as can be seen from Equality 11.19, ij ≠ n for all j in all of the terms that comprise bi. However, then n divides each term, which means n|bi, and bi 0 mod n. This proves this case.

Case 2: i = jn for some j where 0 ≤ j ≤ d. Then, as can be seen from Equality 11.18, the coefficient of xi in g(xn) is aj. Furthermore, if we take ij = n, we have jij = jn = i. Therefore, as can be seen from Equality 11.19, one of the terms in bi is

image

For each of the other terms in bi, clearly we will have ik ≠ n for every k, which means n divides the term, and the term is congruent to 0 modulo n. This case now follows from the fact that Theorem 11.22 says

image

This completes the proof.

image Lemma 11.4

Suppose g(x) is a polynomial with integer coefficients, and n and r are prime. Then

image

Proof: The proof follows from the previous lemma and is left as an exercise.

We also need the following lemmas:

image Lemma 11.5

If r and q are prime, q divides r − 1, and q ≥ 4√rlg n, then q|ordr(n) if and only if

image

Proof: If we let t = ordr(n), then, owing to the fact that r is prime, Theorem 11.17, and Theorem 11.19, there exists an integer k such that r − 1 = tk. Since q divides r − 1 and q is prime, Corollary 11.2 implies q|t or q|k. Now we prove the ‘if and only if’:

Suppose image and by way of contradiction, suppose q. Since q|t or q|k, this means q|k. Therefore, since r − 1 = tk, there exists an integer j such that r − 1 = tjq, which means (r − 1) /q = tj. However, then we have image which means

image

This contradiction proves q|t.

In the other direction, suppose q divides t. Since q ≥ 4√rlgn,

image

Since t = ordr(n), this means

image

which completes the proof.

image Lemma 11.6

If n is composite, q is prime, and q divides q|ordr(n), then there is a prime factor p of n such that

image

Proof: Let p1, p2,…, pk be the prime factors of n. It is left as an exercise to show

image

where lcm is the least common multiple. Since q|ordr(n) and q is prime, Theorem 11.7 implies there is some pi such that q|ordr(pi). This completes the proof.

Our proof of the next theorem relies heavily on the following lemma. The proof of this lemma requires algebra that is beyond the scope of this text. For the reader familiar with abstract algebra, we present the results that prove this lemma at the end of this section. Now we merely state it.

image Lemma 11.7

Suppose the second while loop in Algorithm 11.5 is exited because switch is true. If p is as in Lemma 11.6, and we let l = images2√rlg nimages, then there is a polynomial

image

with the following property: If we let

image

then

1. Jg(x) is closed under multiplication.

2. There is an integer such that for

image

then

image

Proof: See the end of this section.

We can now prove our main result.

image Theorem 11.31

If a composite number is the input to Algorithm 11.5, the algorithm returns false.

Proof: If the number is composite and the first while loop is abruptly ended owing to the return statement in it, false is returned and we are done. If the first while loop is exited because r = n − 1, then n must be prime. So we will assume a composite is entered and the first while loop is exited because switch is true. Then, owing to Lemma 11.5, q divides ordr(n). Suppose by way of contradiction that the algorithm returns true. Then, owing to the second while loop, for 1 ≤ m ≤ l = images2√rlg nimages,

image

which means

image

where p is as in Lemma 11.6. Therefore, each term (x−m) in g(x) in Lemma 11.7 individually satisfies Congruence 11.20, which means

image

Therefore, n image Jg(x) where Jg(x) is as defined in Lemma 11.7. Furthermore, p image Jg(x) owing to Lemma 11.4, and trivially 1 image Jg(x).

Now consider the set

image

Owing to Part 1 of Lemma 11.7, EJg(x). Furthermore,

image

Therefore, by the pigeonhole principle, there are two elements nip j and nhpk in E with i ≠ h or j ≠ k such that

image

So, owing to Part 2 of Lemma 11.7, we have

image

where a is as in that lemma. Since p|n, n is composite, and i, j ≤ images√rimages,

image

Similarly, since h, k ≤ images√rimages,

image

However, since a > n2r/2, Congruence 11.21 therefore implies

image

Since p|n and either i ≠ h or j ≠ k, this last equality implies for some integer s ≥ 1 that

image

However, in the first step of the algorithm we checked whether n is of the form ps for s ≥ 2. Therefore, s = 1 and n is prime. This contradiction proves the theorem.

Time Complexity of the Algorithm

Next we discuss the time complexity of Algorithm 11.5. First we need some lemmas and a theorem.

image Lemma 11.8

Let qm be the largest prime factor of m. Then there exists a positive constant c and integer N such that for n > N

image

Proof: The proof can be found in Baker and Harman (1996).

image Lemma 11.9

Let π(m) be the number of primes less than or equal to m. Then for m ≥ 1,

image

Proof: The proof can be found in Apostol (1997).

image Lemma 11.10

Given positive integers m and n, the product

image

has at most m2 lg n prime factors.

Proof: Each term has at most m lg n prime factors, and there are m terms.

image Theorem 11.32

There exists positive constants c1 and c2 and integer N such that for every n > N there is a prime r in the interval

image

such that the largest prime factor q of r − 1 satisfies

image

Proof: For the moment, let c1 and c2 be arbitrary positive integers. Owing to Lemma 11.8. there exists a positive constant c and integer N such that for c2 (lg n)6 > N,

image

It is left as an exercise to show that Equality 11.22 implies

image

If we let m = c1 (lg n)6 in Lemma 11.9, we obtain, after some manipulations, that the number of primes less than or equal to c1 (lg n)6 is no greater than

image

Combining Inequalities 11.22 and 11.23, we obtain that the number of primes p in the interval

image

satisfying

image

is greater than or equal to

image

Call such primes special primes. Recall that c1 and c2 were arbitrary positive integers. Now choose c1 46. We then have, for any special prime p,

image

The first inequality is owing to Inequalities 11.24 and 11.25, and the second is owing to Inequality 11.25.

Now choose c2 so that image. We then have for n > N that the number of special primes is greater than or equal to

image

Now let x = c2 (lg n)6. Then

image

Owing to Lemma 11.10, the product

image

has at most imagesx2/3imageslg n prime factors. Therefore, due to Inequalities 11.27 and 11.28, there is at least one special prime r that does not divide image.

We will show that r satisfies the conditions of the theorem. Clearly, r satisfies the first two conditions owing to Inequalities 11.24 and 11.26. So if we let q = qr−1, we need only show

image

To that end, suppose

image

Then image Therefore, since r does not divide image, we only need to show

image

to obtain a contradiction. Since Inequality 11.25 says image, we only need to show

image

However, this last inequality does hold, due to Inequality 11.24. This contradiction completes the proof.

Analysis of Algorithm 11.5

image Worst-Case Time Complexity (Polynomial Determine Prime)

Basic operation: one-bit manipulation.

Input size: the number of bits s it takes to encode n, which is given by

image

To determine whether n is of the form kj, the number of roots checked is in O(s). That is, we need to check n1/2, n1/3,…, n1/m, where m = imageslg nimages. Using the brute-force technique discussed in Section 2.6, the time complexity of checking each root is in O(s2). Therefore, the total time complexity of this determination is in

image

Owing to Theorem 11.32, the number of passes through the first while loop is in O(s6). Let’s discuss the work done in each pass. Since r < n, if we use Algorithm 11.1, the time complexity of the computation of the gcd(n, r) is in O(s2). Suppose we use the algorithm presented at the beginning of Section 9.2 to determine if r is prime and to find the largest prime factor of r − 1. There are at most r1/2 passes through the while loop in that algorithm, and, using the brute-force technique discussed in Section 2.6, the time complexity of the computation of the remainder in each pass is in O(s2). Therefore, the total time complexity is in O(r1/2s2), which means, owing to Theorem 11.32, that time complexity is in O(s3s2) = O(s5). Again using the techniques in Section 2.6, the time complexity of incrementing r and of checking the exit condition in the first while loop is in O(s2). Therefore, the total time complexity of the first while loop is in

image

Since r < c(lg n)6, it is possible to obtain a tighter bound than the one just developed. However, as we shall see, the second while loop dominates our final bound anyway.

The number of passes through the second while loop is in O(r1/2s), which means, owing to Theorem 11.32, that the number of passes is in O(s3s) = O(s4). As we mentioned following Theorem 11.29, the time complexity of the congruence determination in that loop is in O(rs2) if we use Fast Fourier multiplication, which means, again owing to Theorem 11.32, that time complexity is in O(s6s2) = O(s8). Therefore, the total time complexity of the second while loop is in

image

We conclude that

image

Agrawal et al. (2002) note that in practice Algorithm 11.5 might be much faster than the bound just obtained. Indeed, they prove that, if a conjecture they make is true, a heuristic time complexity for the algorithm is in O(s6).

image Results that Prove Lemma 11.7

Next we state the lemmas that prove Lemma 11.7. The section requires knowledge of abstract algebra. First we explain our notation. Given a ring R, we denote the corresponding polynomial ring by R[x]. For example, the polynomial ring corresponding to Zn is denoted by Zn[x]. Given a ring Rand an ideal L in R, the factor ring consisting of the forms {r + l: l image L} is denoted R/L. Furthermore, if r image R, and L is the ideal consisting of all multiples of r, we denote R/L by R/r. For example, Zn could be denoted by Z/n. As another example, given a polynomial ring R(x) and polynomial f(x) imageR(x), the factor ring determined by the ideal consisting of all multiples of f(x) is denoted R(x)/f(x). Given a prime p, we will be interested in the ring (actually field) Zp, the polynomial ring Zp[x], and the factor ring Zp[x]/h(x).

We have the following lemmas, which we state without proof. The proofs use standard algebraic manipulations, can be developed in sequence, and can be found in Agrawal et al. (2002).

image Lemma 11.11

Let p and r be prime with p ≠ r.

1. If h(x) is a factor of xr 1, and m ≡ k mod r, then

image

2. If we let ordr(p) be the order of p modulo r, then (xr 1)/(x − 1) factors into polynomials, which are irreducible over Zp and which have degree ordr(p).

Suppose the second while loop in Algorithm 11.5 is exited because switch is true, p is as in Lemma 11.6, and we let l = images2√r lg nimages. Owing to Part 2 of the preceding lemma, there is a polynomial h(x), which is a factor of xr 1, is irreducible over Zp, and has degree ordr(p). The following lemmas all assume the second while loop in Algorithm 11.5 is exited because switch is true, p is as in Lemma 11.6, h(x) is the polynomial just described, and l = images2√r lg nimages.

image Lemma 11.12

In Zp[x]/h(x), let G be the group generated by the l binomials

image

That is,

image

Then G is cyclic and

image

where |G| is the number of elements in G.

image Lemma 11.13

Let G be the group in the previous lemma, g(x) be a generator of G, and ag be the order of g(x) in Zp[x]/h(x). Then

image

Proof: Since l = images2√r lg nimages and q ≥ 4√r lgn, we have q ≥ 2l. Since Lemma 11.6 says q|ordr(p), this means ordr(p) 2l. Owing to the previous lemma, we then have

image

The proof now follows from the fact that g(x) is a generator of G.

image Lemma 11.14

Let g(x) be as in the previous lemma. Then the set

image

is closed under multiplication.

image Lemma 11.15

Let g(x), ag, and Jg(x) be as in the previous lemmas. Then for all m, k image Jg(x), if

image

then

image

It is straightforward that Lemma 11.7 follows from the preceding lemmas if we let a = ag.

11.7 The RSA Public-Key Cryptosystem

Recall the situation discussed at the beginning of this chapter concerning Bob wanting to send Alice a secret love note over the Internet. We noted that if he could encode the message so that it appears as gibberish and only Alice could decode the gibberish back to the original message, he would not need to fear his friends intercepting the message. Public-key cryptosystems enable Bob to do this. After describing such systems, we present the RSA public-key cryptosystem.

• 11.7.1 Public-Key Cryptosystems

A public-key cryptosystem consists of a set of permissible messages, a set of participants such that each participant has a public key and a secret key, and a network for sending messages among the participants. The set of permissible messages might include all character sequences of some given length or less. If we let

image

then each participant x’s public key pkeyx and secret key skeyx determine functions pubx and secx, respectively, from M to M, which are inverses of each other. That is, for each b image M

image

Now the public keys of all participants are known to all the participants, but the secret key of x is known only to x. So, for example, if Bob wants to send Alice a secret love note b, he and Alice do the following:

1. Bob computes c = pubAlice(b) using Alice’s public key, pkeyAlice. The message c is called ciphertext. It is unreadable.

2. Bob sends ciphertext c to Alice.

3. Alice computes b = secAlice(c) using her secret key skeyAlice.

Example 11.66

Suppose Bob wants to send Alice the message ‘I love you.’ The steps are as follows:

1. Bob computes

image

Suppose the result is ‘@!##% * (!’.

2. Bob sends this message to Alice. Bob’s friends see ‘@!##% * (!’.

3. Alice computes

image

The application of pubAlice in Step 1 is called encryption, while the application of secAlice in Step 3 is called decryption. These steps are illustrated in Figure 11.1. Note that since only Alice knows secAlice, only she can decode the ciphertext c back to the original message b. So Bob’s friends see only the ciphertext.

This method will work as long as it is not possible (or at least it is very difficult) to determine skeyx from pkeyx. Next we show a system that makes it very difficult to do so.

Figure 11.1 Bob encrypts his message b using Alice’s public key, and Alice decrypts it using her secret key. Bob’s friends see only the ciphertext c.

image

• 11.7.2 The RSA Cryptosystem

The RSA cryptosystem relies on the facts that we can find large primes fairly readily, but we have no efficient method for factoring a large number. After presenting the method, we discuss these facts further.

The System

In the RSA public-key cryptosystem, each participant creates his public key and secret key according to the following steps:

1. Select two very large prime numbers, p and q. The number of bits needed to represent p and q might be 1024.

2. Compute

image

The formula for ϕ(n) is owing to Theorem 11.17.

3. Select a small prime number g that is relatively prime to ϕ(n).

4. Using Algorithm 11.3, compute the multiplicative inverse [h]ϕ(n) of [g]ϕ(n). That is,

image

Owing to Corollary 11.8, this inverse exists and is unique.

5. Let pkey = (n, g) by the public key, and skey = (n, h) be the secret key.

The set of permissible messages is Zn. The function corresponding to the public key pkey = (n, g) is

image

where b image Zn, and the function corresponding to the secret key skey = (n, h) is

image

The values of these functions can be computed using Algorithm 11.5.

For this system to be correct, the functions corresponding to the public and secret keys must be inverses of each other. Next we prove this is the case.

image image Theorem 11.33

The functions in Equalities 11.29 and 11.30 are inverses of each other.

Proof: Owing to Equalities 11.29 and 11.30, for any b image Zn

image

So we need to show only that

image

To that end, let m image b. (Recall b image Zn.) Then mgh image bgh. We will show

image

Since g and h are multiplicative inverses modulo ϕ(n) = (p − 1)(q − 1),

image

which means there is an integer k such that

image

There are two cases.

Case 1: Assume [m]p [0]p. We then have

image

The third equality above is due to Theorem 11.22.

Case 2: If [m]p = [0]p,

image

So we’ve established Equality 11.31. Similarly,

image

Owing to Equalities 11.31 and 11.32,

image

and

image

Due to Theorem 11.13, we therefore have

image

which means

image

This completes the proof.

image Discussion

As mentioned previously, the success of the RSA cryptosystem relies on the facts that we can find large primes fairly readily, but we have no efficient method for factoring a large number. That is, we can find a large prime as follows. First, we randomly choose integers of the appropriate size. As discussed at the beginning of Section 11.6.1, it should not take very long even to find a fairly large prime. For each integer chosen, we can then use Algorithm 11.5, which is polynomial-time, to check whether the number is prime. We do this until we find two large primes. As discussed in Section 11.6.2, even before Algorithm 11.5 was developed, the Miller-Rabin Randomized Primality Test was used to efficiently check with a very low error rate whether a number was prime. Indeed, the Miller-Rabin Randomized Primality Test may still be the algorithm of choice for prime checking since its time complexity is in O(es3), where s is the number of bits it takes to encode the input, and e is an integer chosen such that the probability the algorithm makes an error is no greater than 2−e. So if we choose e to be only 40, the probability of an error is no greater than 240 = 9. 094 9 × 1013.

On the other hand, no one has ever found a polynomial-time algorithm for factoring a number. The possibility exists that one could find the value of h in the secret key skey = (n, h) without factoring n. However, no one has found an efficient way to do this either. Currently we can achieve security with the RSA cryptosystem if integers containing around 1024 or more bits are used.

EXERCISES

Section 11.1

1. Find the positive divisors of the following integers.

(a) 72

(b) 31

(c) 123

2. Prove that if h|m and m|n, then h|n.

3. Show that two integers divide each other if and only if they are equal.

4. Let p and q be two prime numbers. If p = q + 2, then p and q are called “twin prime numbers.” Find two pairs of twin prime numbers.

5. Prove that gcd(n, m) = gcd(m, n).

6. Prove that if m and n are both even, then gcd (m, n) = 2 gcd (m/2, n/2).

7. Prove that if n ≥ m > 0, then gcd (m, n) = gcd (m, n − m).

8. Prove that if p is a prime number and 0 < h < p, then gcd(p, h) = 1.

9. Use Corollary 11.2 to show that the prime factorization of an integer, as discussed in Theorem 11.5, is unique.

10. Write each of the following integers as a product of prime numbers.

(a) 123

(b) 375

(c) 927

11. Prove Theorem 11.6.

12. Prove Theorem 11.7.

13. Prove that for positive integers m and n, gcd(m, n) = lcm(m, n) iff m = n.

Section 11.2

14. Illustrate the flow of Algorithm 11.1 when the top-level call is gcd (68, 40).

15. Write an iterative version of Algorithm 11.1. Your algorithm should only use a constant amount of memory [i.e., the space complexity function is in θ(1)].

16. Write an algorithm that uses Algorithm 11.1 to express a rational number in its lowest terms. You may assume that this rational number is given in the form of a fraction m/n where m and n are integers.

17. Illustrate the flow of Algorithm 11.2 when the top-level call is Euclid (64,40,gcd,i,j).

18. Write an algorithm that uses subtraction to compute the greatest common divisor. (See Exercise 7.) Analyze your algorithm.

Section 11.3

19. Show that (S, *) of Example 11.21 is a group.

20. Prove Theorem 11.12.

21. The following was left as an exercise in the proof of Theorem 11.13. Show that there exists an integer c such that h1 = cn2n3 · · · nj.

22. Prove Theorem 11.14.

23. Show that if s image [m]n and t image [k]n, then s × t image [m × k]n.

24. Show that if G=(S, *) is a finite group and a image S, then there exists integers k,m ≥ 1 such that ak = akam.

25. Show that if S={[0]12,[3]12,[6]12,[9]12}, then (S,+) is a subgroup of (Z12,+).

26. Use Theorem 11.19 to prove Corollary 11.3.

27. Consider the group image. Show that image.

Section 11.4

28. Solve the following modular equations.

(a) [8]10 x = [4]10 (b) [4]17 x ≡ [5]17

29. Implement Algorithm 11.3 and run it on various problem instances.

30. Find all solutions to the equations [1]7 x = [3]7 and [12]9 x = [6]9.

Section 11.5

31. Compute ([3]73)12 by raising 3 to the 12th power.

32. Compute ([7]73)15 by raising 7 to the 15th power.

33. Use Algorithm 11.4 to compute ([3]73)12.

34. Use Algorithm 11.4 to compute ([7]73)15.

35. Implement Algorithm 11.4, and run it on various problem instances.

Section 11.6

36. Find the number of prime numbers that are less than or equal to 100.

37. If an integer between 1 and 10,000 is randomly chosen according to the uniform distribution, approximately what is the probability of it being prime?

38. Suppose we randomly choose 100 numbers between 1 and 10,000 according to the uniform distribution. Approximately, what is the probability of all of them not being prime?

39. Show that if n is prime and 1 < k ≤ n − 1, then B(n,k) 0 mod n, where B(n,k) denotes the binomial coefficient.

40. Show that if q is a factor of n and k is the order of q in n, then qk where B(n,q) denotes the binomial coefficient.

41. Are 9x3 + 2x and x2− 4 congruent modulo 2?

42. Show that (x − 9)4 is not congruent to (x4 − 9) modulo 4.

43. Show that (x − 5)3 is congruent to (x3 − 5) modulo 3.

44. Prove Theorem 11.29.

45. Implement Algorithm 11.5 and run it on different problem instances.

46. Use Lemma 11.3 to prove Lemma 11.4.

47. The following was left as an exercise in the proof of Lemma 11.6. Show ordr(n)|lcm(ordr(p1), ordr(p2),…,ordr(pk)).

48. Use Inequality 11.22 to obtain the inequality that follows it.

Section 11.7

49. What is the difference between a public key and a secret key?

50. Consider an RSA cryptosystem using p = 7, q = 11 and g = 13.

(a) Compute n.

(b) Compute ϕ.

(c) Find h.

51. Consider an RSA cryptosystem using p = 23, q = 41 and g = 3. Encipher the message [847]943.

52. Use the RSA cryptosystem of Exercise 51 to decrypt the encrypted message found in that exercise.

53. In an RSA cryptosystem, show that if ϕ(n) can be discovered, then the cryptosystem may be compromised.

Additional Exercises

54. Prove that there are infinitely many prime numbers.

55. Show that the gcd operator is associative. That is, for all integers m, n, and h, we have gcd (m, gcd(n, h)) = gcd (gcd(m, n),h).

56. Prove that if m is odd and n is even, then gcd (m, n) = gcd (m, n/2).

57. Prove that if m and n are both odd, then gcd (m, n) = gcd ((m − n)/2, n). 58. Find the necessary condition to have equation mx ≡ my mod n imply x ≡ y mod n.

59. Assuming that p is a prime number, find the solutions of the equation x2 = [1]p.

60. In an RSA cryptosystem, let p and q be the large primes, let n = pq, and let pub be the public key. Show that pub(a)pub(b) is congruent to pub(ab) modulo n.