• Keine Ergebnisse gefunden

Powers and Polynomials in Z

N/A
N/A
Protected

Academic year: 2022

Aktie "Powers and Polynomials in Z"

Copied!
12
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

Elem. Math. 54 (1999) 118 – 129

0013-6018/99/030118-12 $ 1.50+0.20/0 Elemente der Mathematik

Powers and Polynomials in Z

m

Lorenz Halbeisen, Norbert Hungerbu¨hler, Hans La¨uchli Dedicated to the memory of Prof. Hans La¨uchli

Lorenz Halbeisen, geboren 1964 in Laufen, studierte in Basel und Zu¨rich und pro- movierte an der ETH-Zu¨rich. Nach Forschungsaufenthalten in Caen (Normandie) und Barcelona (Katalonien) ist er gegenwa¨rtig als Research Fellow in Berkeley (Kalifornien) ta¨tig.

Norbert Hungerbu¨hler wurde 1964 geboren. Er studierte an der ETH Zu¨rich, wo er 1994 seine Dissertation bei Michael Struwe abschloss. Anschliessend war er an der Universita¨t Freiburg im Breisgau, an der University of Minnesota in Minneapolis und an der ETH in Zu¨rich ta¨tig. Seit Herbst 1998 arbeitet er am Max-Planck-Institut fu¨r Mathematik in Leipzig.

Hans La¨uchli studierte an der ETH in Zu¨rich und promovierte 1961 bei Ernst Specker mit einer Arbeit u¨ber das Auswahlaxiom. Nach Aufenthalten an der University of California in Berkeley und an der University of Arizona in Tucson wurde er 1966 Professor an der ETH. Seine Interessen galten der ganzen Mathematik, am liebsten aber forschte er im Bereich der Logik, der Mengenlehre und der Kombinatorik. Nach la¨ngerer schwerer Krankheit verstarb er, erst 64ja¨hrig, im Sommer 1997.

.

Beim Rechnen in Zm, dem Restklassenring der ganzen Zahlen modulo m, darf man laufend alle auftretenden Summanden, Faktoren und Zwischenresultate modulo mre- duzieren, so dass man nie mit wirklich grossen Zahlen rechnen muss. Wie steht es aber mit Exponenten? Lorenz Halbeisen, Norbert Hungerbu¨hler und Hans La¨uchli zeigen, dass es nur ganz wenige Moduln m, na¨mlich 1, 2, 6, 42 und 1806, gibt, fu¨r die auch eine Formel vom Typab≡abmodm allgemein zutrifft. Gewisse Reduktionen sind aber auch bei beliebigen Modulnmmo¨glich. So lassen sich Funktionen x7→xb(xZm) mit (grossen) Exponenten b in systematischer Weise durch Polynome x7→ g(x)mit gleichen Werten auf Zm, aber wesentlich niedrigerem Grad, ersetzen.

Hans La¨uchli ist am 13. August 1997 gestorben. Die Elemente der Mathematik rechnen es sich als Ehre an, diese scho¨ne und reizvolle Arbeit, die letzte, an der Hans La¨uchli noch mitgearbeitet hat, publizieren zu du¨rfen. cbl

(2)

1 Introduction and Notations

In this article we consider powers and polynomials in the ring Zm, where m N is arbitrary, and ask for “reduction formulas”. For example, for addition, multiplication and exponentiation, we have the following well known reduction formulas:

a+b≡mod(a,m) +mod(b,m)modm (1) a·b≡mod(a,m)·mod(b,m)modm (2)

abmod(a,m)bmodm (3)

It is much more difficult to find reduction formulas which allow to reduce the exponent.

Of course in general the formula

ab≡amod(b,m)modm (4) is false. In the second section we will investigate for which numbersmsuch a reduction formula holds.

In the third and the two following sections we will consider generalizations of Fermat’s little theorem and Euler’s Theorem which allow to replace (inZm) certain powersab by a polynomial f(a)of degree deg(f)which is strictly less thanb. Such formulas can be useful for various reasons: From an algorithmic point of view, it is cheaper to compute the polynomial f(a)modulomthan the full powerabmodulom. On the other hand one may wish for algebraic reasons to replace an arbitrary polynomialg(a)by a polynomial of fixed (lower) degree (depending only onmbut not on g) which is, as a function in Zm, identical tog(see Section 6).

In the last section, we address the question of the minimal degreee(m)such that every polynomial inZmcan be written as a polynomial of degreeq<e(m). We give a complete answer to this question by determining minimal (normed) null-polynomials modulom.

Throughout this paper, we use the customary shorthand notationa|bfor a,b∈Zwith

b

a Z. We write

a≡bmodm

for numbers a,b Z, m N, if m | a−b, and we adopt the notation (a,b) for the greatest common divisor ofaandb. Furthermore we denote by mod(a,m)the uniquely determined numberr∈ {0,1, . . . ,m−1}such that a= k m+r for some k Z, and Mod(a,m)denotes the numberr∈ {1, . . . ,m}such thata=k m+r for some k∈Z. We had been working on the present article for about two years, when the mournful message of Hans La¨uchli’s death reached us. At that time, only the first part (Section 2), which comprises a theorem resulting from joint work of Hans La¨uchli and Ernst Specker on exponential rings, and the second part (Sections 3–5) had been finished. The third part about minimal polynomials was not yet completed, and we would like to thank Prof. Ernst Specker for inspiring and helpful discussions and for valuable suggestions concerning that last section.

(3)

2 Special values ofm

In this section we investigate for which values of m the reduction formula (4) holds.

The answer is contained in the following theorem.

Theorem 1 LetG:={1,2,6,42,1806}, then the following statements are equivalent:

(a) m∈G.

(b) For all integersa,bone has

ab≡aMod(b,m)modm.

(c) For all integersaone has

am+1≡amodm.

Remark: The equivalence of (b) and (c) is obvious: (c) follows from (b) by choosing b=m+1. The opposite implication follows from (2) by an easy induction argument.

However, notice that in (b) we cannot replace “Mod” by “mod” in the exponent. To make this point more precise we state without proof:

Theorem 2 Letm ∈G, then one has am 1 modm (and hence (b) holds with Mod replaced by mod) if and only if no prime factor of mod(a,m)belongs to the setG+1= {2,3,7,43,1807}.

The proof of the equivalence of (a) and (c) relies on an induction principle, which we prove after the following lemma.

Lemma 1 Let E1 := 2 and En+1 := q+E1E2· · ·En for a fixed, odd q > 0. If A:=E1· · ·Ek such thatEi is prime fori≤ k andx|A, thenx+q∈ {E1, . . . ,Ek+1} orx+qs is not prime for answith 1≤s< k.

Proof. Ifx=A, thenx+q=Ek+1and we are done. Ifx6=A, then letlbe the smallest number such that El-x. If l = 1, then x+q1 is even, therefore x+q = 2 ∈ {E1} or x+q is not prime. Hence, the claim is proved for l = 1 and only the casel >1 remains to be checked: SinceE1, . . . ,El are prime, we haveE1· · ·El−1|x. Notice that E1· · ·El−1 ≡ −qmodEl (for l >1) and that Ej ≡qmodEl for j >l (by definition).

Therefore we conclude x≡ −qsmodEl, wheres is smaller than the number of prime factors ofx, hences< k. ThereforeEl|x+qs and the proof is finished. h We will use the special caseq=1 in the proof of the following

Theorem 3 (Induction Principle) Let H N be a set of natural numbers with the following properties:

(i) 1∈H,

(ii) ifh∈H andh+1 is prime, it follows thath(h+1)∈H, (iii) ifp2|xforp>1, thenx∈/H,

(iv) if h = A p a H, p prime, such that all divisors of a are greater than p, then p−1|A.

ThenH=G.

(4)

Proof. By (i) and (ii), G H. For the opposite inclusion we claim that 2 ≤h H implies h = E1·El with l 4: In fact, by (iii), we know that h = p1p2· · ·pn with p1 <p2 < . . . <pn being prime numbers. Now we use (iv) with A=1, p=p1 and a=hp. Becausep11|1 (by (iv)), we havep1=2=E1. Now, by induction, we assume that pj =Ej for all j k ≤l. Applying (iv) again, this time with A=E1E2· · ·Ek, p=pk+1anda= Aph , we havepk+11|A. Thus, by Lemma 1,pk+1∈ {E1, . . . ,Ek+1} and sincepk+1>pj for j ≤k, we concludepk+1=Ek+1. h Proof of Theorem 1: Now, we use the induction principle to prove Theorem 1. We have to check properties (i)–(iv) for the setL of numbershwhich satisfy (c):

(i) is trivial.

(ii) follows easily from Fermat’s little theorem (see Section 3).

(iii) Let h= p1· · ·pn L,pk prime. By (c), we know that phk+1 ≡pkmodh. Thus, h | pk(phk 1) and hence we have phk 1 modph

k. For i 6= k it follows that phk 1 modpi and thereforepi 6=pk.

(iv) By (c) we have forh=A p a∈L thatch+1≡cmodhfor allc. Thush|c(ch1) and

ch= (cA a)p1 modp. (5)

Now, letcbe such that(c,p) =1, then (by Fermat’s little theorem)

(cA a)p−1 1 modp. (6)

Combination of (5) and (6) yieldscA a1 modp. Sincepis prime and(c,p) =1, it follows thatp−1|A aand by definition ofawe getp−1|A, which completes

the proof of Theorem 1. h

3 A generalization of Fermat’s little theorem

Let us start with a definition. Letp1, . . . ,pkbe distinct prime numbers andm=pε11· · ·pεkk withεi N be the factorization of a numberm∈N. Then we define the functionϕm for integer numbersn by

ϕm(n) =nmX

i

nmpi +X

i1<i2

n

m

pi1pi2 − · · ·+ (1)knp1···p km

= X

j⊂{1,...,k}

(1)|j|np jm.

Here, the subsetj={j1, . . . ,ji}of{1, . . . ,k}serves as a multi-index with length|j|=i and withpj :=pj1· · ·pji. It is convenient to extend the definition ofϕmbyϕ1(n):=n.

Theorem 4 The functionϕm(n)has the property

ϕm(n)0 modm (7)

for all numbersn∈N.

(5)

Remarks:

(i) Ifnis a prime number, then (7) follows from Gauss’ observation that the number of irreducible polynomials of degree m overZn is given by ϕm(n)/m(see [2]).

Later Serret [8], Lucas [6] and Pellet [7] stated without proof that (7) holds true for arbitrary integern. Later on, several proofs have been given for (7): S. Kantor presented in [3] and [4] geometric proofs and Weyr [9] used an involved inductive method.

(ii) Theorem 4 allows now to determine mod(nm,m)by replacing the full power nm by a polynomial innof degree strictly less thanm, which at least partially answers the question posed in the introduction.

Here, we show that (7) follows easily from a combinatorial fact. To demonstrate the idea we consider the case of a prime numberm=p. Consider the set {(n1, . . . ,np) : ni ∈ {1, . . . ,n}} of points in the discrete p-dimensional cube Q={1, . . . ,n}p. Con- sider the cyclic group Cp whose action on a point (n1, . . . ,np) is generated by σ = σp:(n1, . . . ,np)7→(n2,n3, . . . ,np,n1). According to Burnside’s Lemma the total num- ber of orbits inQgenerated byCp is given by

number of orbits= 1

|Cp| X

g∈Cp

χg (8)

where χg is the number of fix-points of Q under g Cp. Since χσi = n for i = 1, . . . ,p−1 and χσp = χid = np (and of course |Cp| = p) it follows from (8) that np+ (p1)n0 modpand hence

np−n≡0 modp, which is Fermat’s little theorem.

For general mwe proceed similarly, but instead of using Burnside’s Lemma we count directly the orbits of given length.

Proof of Theorem 4. LetQandCm be as above but now with generalm=pε11· · ·pεkk. We claim that there exist m1ϕm(n)orbits of lengthmand hence the theorem follows. To prove this claim we proceed by induction onm:

1st step:ϕ1(n) =n, hence the assertion is true form=1.

2nd step: “m0 =p1· · ·pk−1→m=p1· · ·pk”: Notice, that the number of orbits gener- ated byCm in{1, . . . ,n}mof length mm0 equals the number of orbits generated byCm/m0

in{1, . . . ,n}m/m0 of length mm0. So, by induction we have that number of orbits of length m

pi

= ϕm

pi(n)

m pi

number of orbits of length m pipj

=

ϕ m

pi pj(n)

m pipj

. . .

(6)

Hence,

number of orbits of lengthm=

= 1

m nmX

i

ϕm

pi(n)X

i<j

ϕ m

pi pj(n)− · · · −ϕi(n)

= 1

m nm X

i⊂{1,...,k}

inot empty

X

j⊂{1,...,k}\{i}

(1)|j|npi pjm

= 1 m(n).

(9)

3rdstep: “m0 =pε11· · ·pεkk1→p0 =pε11· · ·pεkk”: analogous to the second step. h 4 A generalisation of Euler’s Theorem

One disadvantage of (7) is that it reduces in Zm only the power m. Here, we present a formula which reduces yet another power and which is slightly stronger than Euler’s Theorem. Let us recall the definition of Euler’s ϕ function: For any integer n, ϕ(n) denotes the number of integers k∈ {1, . . . ,n−1}which are relatively prime ton, i.e.

ϕ(n):=|{k∈ {1, . . . ,n−1} : (n,k) =1}|. Furthermore, letϑ(n)denote the highest power contained inn, i.e.

ϑ(n):=max{k : mk |n,m∈N, m>1}. Theorem 5 There holds

(a) nϑ(q)(nϕ(q)1)0 modqfor all integersn.

(b) ϑ(q) +ϕ(q)≤qfor allq, with equality if and only ifqis prime.

Proof. (a) Let q= q1ε1· · ·qεkk be the prime factorization of q. If (n,qi) = 1 it follows from Euler’s Theorem (which asserts that nϕ(q) 1 modq provided (n,q) = 1) that qiεi |nϕ(qεii)1. Hence, sinceϕis multiplicative, i.e.ϕ(ab) =ϕ(a)ϕ(b)for(a,b) =1,

qεii |nϕ(q)1 if(n,qi) =1. (10) Furthermore we haveqiεi1 |qand henceqiεi1|q−ϕ(q)>0. On the other hand, it is clear that(n,qi)>1 impliesqi|n. Hence we have

qiεi |nϑ(q) if(n,qi)>1. (11) Now, combining the two cases (10) and (11) the assertion follows.

(b) 1st step: Ifqis prime then obviouslyϑ(q) +ϕ(q) =1+ (q1) =q.

2ndstep: We have to show thatϑ(q) +ϕ(q)<qifqis not prime. Ifq=pnfor a prime numberpandn≥2, the assertion is equivalent ton+ (p1)n <pn, which is easily established by induction onn≥2. Ifq=pnq0 withpprime,q0>1 andn=ϑ(q)≥1,

then ϑ(q) +ϕ(q) =n+ (p1)nϕ(q0)

≤n+ (p1)n(q01)

and hence the assertion follows from the factn+ (p1)n(q01)<pnq0 which is easily

proved by induction onn. h

(7)

Remarks:

(i) Of course, Euler’s Theorem follows from Theorem 5(a).

(ii) It is clear from the proof, that the exponentϑ(q)in (a) is optimal, i.e. it cannot be replaced by a smaller integer.

(iii) Theorem 5 allows to replacenϑ(q)+ϕ(q)inZqby a polynomial innof degree strictly less thanϑ(q) +ϕ(q).

5 Another application of Burnside’s Lemma

In this section, we consider a variant of the arguments of Section 3. There, we considered the cyclic groupCm, i.e. the group with one generating element of orderm. Notice that the set of points of the cube Q = {1, . . . ,n}m (on which Cm acts) may as well be considered as the set of colorings with n colors of the Cayley graph of Cm generated by the generating element. (The Cayley graphG[A]of a groupG generated by a subset A ={a1, . . . ,ak} ⊂G has the elements {g1, . . . ,gl} of G as its vertex set and edges between gi andgj iff there exists an A with gi ◦an = gj.) By applying Burnside’s Lemma to this situation, we obtained (7).

A natural variant of this idea would be to look at the group G = Cp1 × · · · ×Cpk of k generating elements a1, . . . ,ak of orders p1, . . . ,pk, acting on the Cayley graph G[a1, . . . ,ak] over the generating elements and colored with n colors. In fact, if the pi are chosen to be prime (but not necessarily different), we recover (7) by applying Burnside’s Lemma. But we do in fact obtain a new congruence if we look at a “reduced Cayley graph” instead. More precisely we consider the graph Cp1[p1]× · · · ×Cpk[pk] colored with n colors, and g1ε1· · ·gεkk ∈G acting on it by application of gi on Cpi[pi].

Counting orbits in a similar way as in Section 3 we find

Theorem 6 Ifm=p1· · ·pk (pi prime, but not necessarily distinct), then there holds

for all integersn X

j⊂{1,...,k}

(−n)|j|ns(m)−s(pj)0 modm

where we used the multi-index notation of Section 3 ands(m):=p1+· · ·+pk denotes the sum of the primes inm(with multiplicity).

Remarks:

(i) Theorem 6 now allows to reducens(m) by a polynomial of lower degree inZm. (ii) If one does not insist onpi being prime, one ends up with a polynomial of degree

p1+· · ·+pk ≥s(m)which vanishes inZm. 6 Minimal null-polynomials

6.1 Normed null-polynomials. Usually one defines two polynomials f and g to be congruent modulom, written f ≡gmodm, if corresponding coefficients are congruent integers modulom. This equivalence relation provides a nice structure in particular ifmis chosen to be prime. On the other hand we will say that two polynomials (or, more general, two functions) f andgare graph-congruent modulom, written f ≡g graph modm, if

(8)

they have the same graph as functions fromZm toZm, i.e. if f(n)≡g(n)modmfor all integersn. Of course, two congruent polynomials are graph-congruent, but the converse implication does not hold in general, e.g. x2 ≡x graph mod 2, but x2 andx are not congruent modulo 2. We say f is a normed null-polynomial modulom, if f is graph- congruent to the polynomial 0 and if f is normed (i.e. the leading coefficient equals 1). Of course, for allmthere exist normed null-polynomials, e.g. f(x) = (x1)(x 2)· · ·(x−m), and hence it makes sense to look for minimal normed null-polynomials modulom, i.e. normed null-polynomials of minimal degree e(m). It is easy to see, that ifm=pis prime, the polynomial

xp−x≡(x1)· · ·(x−p)modp

is (up to congruence) the unique minimal normed null-polynomial, and hence e(p) =p for pprime. Minimal normed null-polynomials are useful since they allow to replace arbitrary polynomials by graph-congruent polynomials of degree less than or equal to e(m)−1 modulom. To find a minimal normed null-polynomial on a computer by just checking polynomial after polynomial, would be extremely time consuming. On the other hand from Theorem 5 and 6 we infer, that

e(q)≤min{q,s(q), ϑ(q) +ϕ(q)}. Example: Letm=35 andf(n) =P35

i=0ni. Find a polynomialgof lower degree which is as a function inZm identical to f.

Theorem 4 provides a normed null-polynomial of degree 35, which would allow to find a polynomial g of degree 34. Theorem 5 gives a normed null-polynomial of degree

ϑ(m) +ϕ(m) =25 which is better, but Theorem 6 gives a polynomial of even lower

degree, namelys(m) =12, in fact

n12≡n(n5+n7−n) graph mod 35.

Replacing in f successively all powersn12byn(n5+n7−n)one finds X35

i=0

ni1+n−15(n2+n3)13(n4+n5)+

+5(n6+n7) +21(n8+n9) +19(n10+n11) graph mod 35.

We include the following list, which decides for which m Theorem 5 or Theorem 6 yields a normed null-polynomial of lower degree:

(1) ϑ(q) +ϕ(q) =s(q)if and only ifqis prime orq∈ {4,18} (2) ϑ(q) +ϕ(q)<s(q)ifq=2p,pprime, orq∈ {12,30} (3) for all otherqthere holdsϑ(q) +ϕ(q)>s(q)

Since for m = 18 both theorems give a polynomial of degree 8, we can look at the difference which is the (normed) null-polynomialn7+2n62n5−n4+n3−n2. But still, it is not minimal. In factn6+n42n2is a minimal normed null-polynomial modulo 18, i.e.e(18) =6. The following theorem gives the general answer to the problem:

(9)

Theorem 7 The polynomialg(x) =Qs(m)

i=1 (x+i)is a minimal normed null-polynomial inZm and hencee(m) =s(m). Here,s(m)denotes the Smarandache functions(m):=

min{k N : m| k!}.

The functions(m)is named after the Romanian Mathematician Florentin Smarandache, but it has been introduced already in 1918 by Kempner in [5]. It has many interest- ing properties and applications in number theory (see e.g. the Smarandache Function Journal).

Proof. 1st step:g(x)is a normed null-polynomial inZm: This follows immediately from the fact that for allx∈Z

g(x) =

x+s(m) s(m)

s(m)!

Now, the first factor is an integer, ands(m)!0 modm.

2nd step: e(m)≥s(m): Let us consider the normed polynomial f(x):=a1x+a2x2+ . . .+ar−1xr−1+xr withai Zandr>1. We define

M=





1 1 · · · 1

2 22 · · · 2r−1 3 32 · · · 3r−1 ... ... . .. ... r−1 (r1)2 · · · (r1)r−1





and the vectors

a=



a1

a2

... ar−1



, h=



f(1) f(2) ... f(r1)



, ρ=



 1r 2r ... (r1)r



.

In this notation, we have

Ma=h−ρ.

Now, suppose that

f(x)0 modmfor allx=1,2, . . . ,r−1,

i.e. h= m q for some q∈ Zr−1. Notice that M is a Vandermonde matrix and that in particular det(M) 6= 0. Hence, the equation Ma = m q−ρ determines for any given right hand side a unique solutiona. From Lemma 2 below we infer

f(r) =rr+

r−1

X

i=1

airi=rr+

r−1

X

i=1

(1)i+r r

i

(ir−mqi) Xr

i=1

(1)i+r r

i

irmodm. Lemma 3 below now gives that f(r)≡r!≡0 modmimpliesr≥s(m). This completes

the proof. h

(10)

Lemma 2 LetMbe the Vandermonde matrix(ij)i,j=1,...,r−1as above. Then, fora∈Rr−1 andb=Ma there holds

r−1

X

i=1

airi=

r−1

X

i=1

(1)i+r r

i

bi. (12)

Proof. By linearity, it suffices to show (12) forai=δi,j, j=1,2, . . . ,r−1. That is, we have to show that for 1≤j≤r−1

rj=

r−1

X

i=1

(1)i+r r

i

ij.

This follows also from Lemma 3. h

Lemma 3 Forr∈N0 and j∈N0, there holds Xr

i=0

(1)r−i r

i

ij=r!S2(j,r),

whereS2 is the Stirling number of the second kind.

Proof. A proof of this well-known lemma can be found e.g. in [1]. But for the sake of completeness, we like to give a proof by combinatorial arguments which are similar to those in the proof of Burnside’s Lemma. Moreover, we shall give a special proof for the case j=rand will consider the general case afterwards in a slightly different way.

First notice, that from the binomial expansion of(1+x)r withx=1, we get Xr

i=0

(1)r−i r

i

=0r, ()

which is (forr>0) obviously equivalent to

r−1

X

i=0

(1)i+1 r

r−i

= (1)r. ()

Let A := {a0, . . . ,ar−1} be an alphabet of r > 0 symbols and let wr(k) denote the set of words of length r, such that every word inwr(k)consists of exactly k different letters. Further, letWr(k)denote the cardinality ofwr(k). Obviously we haveWr(0) = Wr(r+i) =0 (fori≥1) andWr(r) =r!. And in general we have

Wr(k) = r

k

kr

r−k+1 1

Wr(k1)

r−k+2 2

Wr(k2)−. . . . . .−

r−1 k−1

Wr(1) r

k

Wr(0).

(∗∗)

(11)

To see this, remember that with k different letters we can form kr words for length r, but of course, not all of them contain kdifferent letters. So, to computeWr(k), we have to exclude the words which contain less than k different letters.

Combining()and(∗∗)we get Wr(r) =

r r

rr r

r−1

(r1)r+ r

r−2

(r2)r−. . .(1)r r

0

0r =r!. BecauseS2(n,n) =1 (for alln∈N0), this proves the Lemma forr= jeven in the case whenr=j=0 (becauseW0(0) = 00

0! =0!).

Now we consider the general case. Again,wj(k)denotes the set of all words of length j, such that every word inwj(k)consists of exactly k different letters. For an arbitrary wordu(of length j) let ¯ube the set of all letters occurring inuand|u|¯ be its cardinality.

So, ifu∈wj(k), then|u|¯ = k. For a set of lettersI ⊆Aletvj(I)be the set of all indexed I-wordsuI of length j, such that ¯uI ⊆I. To each indexed worduI there corresponds in a natural way the (non-indexed) wordu. For two differentI andI0 such that|I|=|I0| we call two indexedI-wordsuI anduI0 equivalent (uI ∼uI0) if the (non-indexed) words are equal. Let[u]i :={vI :vI ∼uI0∧ |I|=|I0|=i}. Finally let

Vj(i):=X

I⊆A|I|=i

|vj(I)|.

Evidentially we haveVj(i) = ri

ij. For an arbitrary worduof length j with ¯u⊆I ⊆A where|I|=i we get

|[u]i|=

r− |u¯| r−i

. For a worduwith ¯u<r, we have by ()that

Xr i=|u|¯

(1)r−i|[u]i| =0.

Therefore,Pr−1

i=0(1)r−i ri

ij=0=r!S2(j,r). Now, with the alphabetAwe can form r!S2(j,r)wordsuof length j, such that ¯u=A, which completes the proof. h Remark: As a corollary of the previous lemma, we obtain Wilson’s Theorem:(p−1)!

1 modpif and only ifpis prime. To see this, notice first that ifp=ab, witha,bboth bigger than 1 and(a,b) =1, thena|(p−1)! andb|(p−1)!, therefore(p−1)!0 modp.

Forpprime, setr= j=p−1 and use Fermat’s little theorem in the Lemma 3 (for the only even prime numberp=2, notice that11 mod 2).

6.2 General null-polynomials. Except in the case whenmis prime, the minimal normed null-polynomials are far from unique. For example, given a normed null-polynomial, one can add a general (not normed) null-polynomial of lower degree. So, let us look now for

(12)

non-trivial minimal null-polynomials (which need not be normed). Let ˜e(m)denote the degree of a general non-trivial minimal null-polynomial modulom. Then there holds:

Theorem 8 ˜e(m)equals the smallest prime factor inm.

Proof. Letm=pε11· · ·pεkk withpi prime andp1 ≤pj for all j>1.

1st step: Ifp1>2, then the polynomial f(x) = m

p1

x

p1−1

Y2

i=1

(x2−i2)

is a null-polynomial. Forp1=2 the polynomial f(x) = m2x(1+x)is a null-polynomial.

Thus we have ˜e(m)≤p1.

2ndstep: Let f(x)be a non-trivial null-polynomial inZm. Without loss of generality, we may assume that the coefficients of f do not contain a common divisorpδii withδi> εi (otherwise, one can divide f bypδiiεi which would still be a non-trivial null-polynomial inZm, but with the desired property). Let Qk

i=1pγii be the largest common divisor (of this form) of the coefficients of f. In particular, we have that 0 γi εi for all i.

Thus, we have f(x) =Qk

i=1pγiig(x)for a polynomialg(x)with integer coefficients and for all x Z there exists an integer hx such that f(x) = m hx. Hence, we conclude for g(x)that g(x) =hxQk

i=1pεiiγi. This means thatgis a null-polynomial inZm0 with m0 =Qk

i=1pεiiγi >1. Furthermore, gis non-trivial in Zm0 since the greatest common divisor of the coefficients ofgdoes not contain a factorpi. Now, let jdenote the smallest index with the property thatεj−γj >0. Then,gis a non-trivial null-polynomial in the field Zpj. Since a non-trivial polynomial has in a field at most as many zeros as the degree indicates, we conclude deg(f) =deg(g)≥pj ≥p1. h References

[1] M. Aigner: Kombinatorik. Springer, 1975

[2] C.F. Gauss. Posthumous papers, Werke 2 (1863), p. 222

[3] S. Kantor: Wie viele cyclische Gruppen gibt es in einer quadratischen Transformation der Ebene? Annali di matematica, Serie II, 10 (1880), 64–70

[4] S. Kantor: Beantwortung derselben Frage fu¨r Cremona’sche Transformation, Annali di matematica, Serie II, 10 (1880), 71–73

[5] A.J. Kempner: Concerning the smallest integerm! divisible by a given integern. Amer. Math. Monthly 25 (1918), 204–210.

[6] E. Lucas: Sur la ge´ne´ralisation du the´ore`me de Fermat. Comptes rendus 96 (1883), 1300–1301 [7] M. Pellet: Sur une ge´ne´ralisation du the´ore`me de Fermat. Comptes rendus 96 (1883), 1301–1302 [8] M. Serret: The´ore`me de Fermat ge´ne´ralise´. Nouvelles annales de mathe´matiques 14 (1855), 261–262 [9] E. Weyr: O jiste´ veˇteˇ cˇı´selne´. Cˇ asopis mathematiky a fysiky 11 (1882), 39–47

Lorenz Halbeisen Mathematics U.C. Berkeley Evans Hall 938 Berkeley, CA 94720

Norbert Hungerbu¨hler

Max-Planck-Institut fu¨r Mathematik Inselstr. 22-26

D - 04103 Leipzig

Referenzen

ÄHNLICHE DOKUMENTE

If the minimum distance d ⊥ of the dual code is at least 3 (this is the same as saying that the generator matroid is simple), we also show how one can identify important coefficients

In the introductory Section 1.3, we summarized statements known for univariate and multivariate stable polynomials. Some of them can be transferred to conic stability. For example,

Zum Beispiel fu¨hrt die Frage nach der Irreduzibilita¨t quadratischer Polynome u¨ber den ganzen Zahlen auf das Studium quadratischer Irra- tionalita¨ten, welches bereits in der

In this paper, we applied the variational iteration method using He’s polynomials (MVIM) for finding the solutions of Burgers’ and coupled Burgers’ equa- tions.. The use of

It seems that Theorem 3 together with [MZ2], [MZ4] makes it possible to prove the claim for curves over C when the functions are elements of the function field of an elliptic curve..

These include special orthogonal polynomials, such as associated Laguerre polynomials and Chebyshev polynomials of the second kind.. Also included are polynomials attached

This method can be used in other cases to prove polynomial identities combinatorially: when x is an integer, the two sides of the identity could count the same thing in two

As in the case of symmetric polynomials, the algebra of multisymmetric polynomials admits an obvious module basis: the one made of the symmetrizations of the monomials, called