• No results found

2 The Number of Subspaces of F

N/A
N/A
Protected

Academic year: 2021

Share "2 The Number of Subspaces of F"

Copied!
46
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

faculteit Wiskunde en Natuurwetenschappen

On the Schur Product of Vector Spaces over Finite Fields

Bachelor Thesis Mathematics

July 2012

Student: Christiaan Koster First Supervisor: Prof. dr. J. Top

Second Supervisor: Prof. dr. H. L. Trentelman

(2)
(3)

Abstract

In this thesis we consider the vector space of all n-tuples over finite fields. We will investigate the Schur product, also known as entry-wise multiplication, of its linear sub- spaces, which is equal to the span of the Schur product over all pairs of vectors of a subspace. The number of d-dimensional linear subspaces will be counted, certain prop- erties of the Schur product will be shown and the Schur product of cyclic codes will also be investigated. The main goal is to find good bounds on the number of subspaces for which the Schur product is equal to the total space. This will then be compared to the total number of subspaces as either the dimension of the vector space or the number of elements of the base field tend to infinity.

(4)

Contents

1 Introduction 5

1.1 Preliminaries . . . 5

1.2 Coding Theory . . . 6

2 The Number of Subspaces of Fnq 9 2.1 Growth Pattern of #G(d, n)(Fq) . . . 11

3 Schur Product 17 4 Schur Product of Subspaces 20 4.1 Bilinear Form . . . 26

4.2 Schur Product of Cyclic Codes . . . 28

4.3 Lower Bounds on #W(d, n, q) . . . 31

4.3.1 First Bound . . . 32

4.3.2 Second Bound . . . 34

5 Conclusion 42 A Magma Code 43 A.1 Code for Schur Product on Subspaces . . . 43

A.2 Code for Schur Product on Cyclic Codes . . . 44

(5)

1 Introduction

In this modern day and age we live in an information-driven society. Private information about people or companies is very valuable, while the people themselves do not wish this information to be known to others. One could think of PIN-codes or credit card numbers, but also the economic situation of a person and very sensitive data such as nuclear missile codes. While the information must not be known to certain others, it also has to be able to be used. The information in such cases can be regarded as a secret. The problem is to find out how a secret from a certain party can be used in combination with other secrets from other parties, without the secret being known to any of the other parties. An important tool used in solving this problem is known as secret-sharing.

In this thesis we consider vector spaces of n-tuples over finite fields and investigate the Schur product, also known as entry-wise multiplication, of linear subspaces of the vector spaces.

The Schur product is a product that is used in secret sharing. The main goal of this thesis is to find the number of subspaces, or upper and lower bounds, for which the Schur product is equal to the whole vector space and compare this to the total number of linear subspaces.

In order to do this we first count the exact number of linear subspaces. Next we investigate some basic properties of the Schur product and then we handle the main question by trying various methods. It will be done by a coding theoretic approach and we also consider how the Schur product acts on cyclic codes.

1.1 Preliminaries

Let K be a finite field, there is the following theorem on the classification of finite fields, found (in Dutch) in [3]:

Theorem 1.

1. Let K be a finite field. Then there is a prime p and an integer m ≥ 1 with #K = pm. 2. Conversely, for each prime p and integer m ≥ 1 there is a finite field with pm elements,

and this field is uniquely determined up to isomorphisms.

The field with q = pm elements will be denoted by Fq. If p is a prime, then Fp ∼= Z/pZ and we write Fp = {0, 1, . . . , p − 1} where a ∈ Fp is identified with a + pZ. For convenience we just write a instead of a + pZ. In this thesis, the vector space Fnq over Fq will be considered, for any integer n, with standard vector addition and scalar multiplication. Define

e1= (1, 0, . . . , 0) e2= (0, 1, . . . , 0)

...

en= (0, . . . , 0, 1), then the set {e1, . . . , en} forms the standard basis of Fnq.

Definition 1. A set V ⊂ Fnq is called a linear subspace of Fnq if for all v, w ∈ V and α ∈ Fq

it holds that:

(6)

1. v + w ∈ V 2. αv ∈ V .

Definition 2. The collection of all d-dimensional linear subspaces of Fnq is denoted by G(d, n)(Fq), 1 ≤ d ≤ n.

The set G(d, n) is also known as the Grassmannian of Fnq. It has a lot of useful properties, but these will not be treated in this thesis. Here it will just be used as easy notation to pick a d-dimensional linear subspace of Fnq.

Definition 3. The general linear group GL(n, Fq) is defined as the group of all n×n invertible matrices over Fq.

Definition 4. For any n ∈ N let S = {1, 2, . . . , n}. The symmetric group of S consisting of all bijections of S onto itself is denoted by Sn. An element σ ∈ Sn is called a permutation.

In this thesis a subspace V ∈ G(d, n)(Fq) will be often be written as the span of d basis vectors v1, . . . , vd. Each vi can be written as vi = ai,1eσ(1)+ · · · + ai,neσ(n) for a permutation σ ∈ Sn and where ai,j ∈ Fq for i = 1, . . . , d, j = 1, . . . , n. Since v1 6= 0 we can assume that a1,1 6= 0 by choosing σ appropriately and since V = Span{v1, . . . , vd} we can take ai,i = 1.

Then subtracting ak,1v1 from vk for each k = 2, . . . , d, a new basis for V is {v1, w2, . . . , wd} where each wj can be written as wj = bj,2eσ(2)+ · · · + bj,neσ(n). Since v1 and w2 are linearly independent, we can assume bj,2 6= 0 and even bj,2 = 1. Now subtract bm,2w2 from wm for each m = 3, . . . , d. This can be continued repeatedly until we find a basis for V of the form

{ui= eη(i)+ ˜ai,i+1eη(i+1)+ · · · + ˜ai,neη(n) : ˜ai,j ∈ Fq for i = 1, . . . d, j = i + 1, . . . , n}

for a η ∈ Sn. For any i > 1 we can now subtract ˜al,iui from ulfor each l = 1, . . . , i − 1. Doing this for each i > 1 and replacing uj with uj− ˜aj,iui whenever j < i we find a basis for V of the form

{eτ (i)+ ci,d+1eτ (d+1)+ · · · + ci,neτ (n): ci,j ∈ Fq for i = 1, . . . d, j = d + 1, . . . , n} (1) for a τ ∈ Sn. For τ fixed, this basis is unique.

1.2 Coding Theory

In this thesis we will also look at linear subspaces of Fnq from a coding theory perspective. All definitions and theorems in this part can be found (in Dutch) in [2]. Since we only consider linear subspaces of Fnq, the following definition of codes will be used.

Definition 5. Let K be a field. A code of length n is a linear subspace C of Kn. An element c ∈ C is called a word of the code.

Definition 6. A cyclic code is a code C ⊂ Fnq such that if c = (c1, . . . , cn) ∈ C then (cn, c1, . . . , cn−1) ∈ C.

Definition 7. The dual C of a code C is defined as C = {(a1, . . . , an) ∈ Fnq :

n

X

i=1

aici = 0 ∀(c1, . . . , cn) ∈ C}

(7)

Note that C is always linear as well.

Proposition 1. Let C ⊂ Fnq be a code, then Dim(C) + Dim(C) = n.

Example 1. Let C ⊂ F32 be linear and cyclic. We can construct cyclic codes C of dimension 0, 1, 2 and 3. If ei ∈ C for any i ∈ {1, 2, 3}, then, since C is cyclic, ei ∈ C for i = 1, 2, 3.

Hence, in this case Dim(C) = 3 and C = F32. If C = Span{(1, 1, 1)}, then C is a cyclic code with Dim(C) = 1. Finally, let C = {(0, 0, 0), (1, 1, 0), (0, 1, 1), (1, 0, 1)}. Then C is a cyclic code with Dim(C) = 2. These codes, along with C = {(0, 0, 0)}, are the only cyclic codes in F32.

Let Fq[X] denote the polynomial ring over Fq, then for any n ≥ 1 define

R := Fq[X]/(Xn− 1) = {a0+ a1X + · · · + an−1Xn−1 mod (Xn− 1) : a0, . . . , an−1 ∈ Fq} Next define a mapping g : Fnq → R as

g((b0, . . . , bn−1)) = b0+ b1X + · · · + bn−1Xn−1 mod (Xn− 1)

for (b0, . . . , bn−1) ∈ Fnq. Then g is an isomorphism between the vector spaces Fnq and R, that is g is a linear bijective map. From now on, elements of R will be denoted by ¯f ∈ R where f ∈ Fq[X]. The point of the identification of Fnq with R will become clear when cyclic codes are considered.

Let C ⊂ Fnq be a cyclic code, then C gets mapped by g onto ˜C ⊂ R. Now we will determine what properties ˜C has, caused by the fact that C is a cyclic code.

That C is linear implies the following:

1. 0 ∈ ˜C, since (0, . . . , 0) ∈ C.

2. if f1, f2 ∈ ˜C and a ∈ Fq, then fi= g(ci) for a ci∈ C. Since C is linear, also c1−ac2 ∈ C and hence f1− af2 = g(c1− ac2) ∈ ˜C.

So ˜C is linear as well.

That C is cyclic means that if (b0, . . . , bn−1) ∈ C, then (bn−1, b0, . . . , bn−2) ∈ C. In ˜C this means that if ¯f = b0 + b1X + · · · + bn−1Xn−1 mod (xn− 1) ∈ ˜C then also bn−1+ b0X +

· · · + bn−2Xn−1 mod (Xn− 1) ∈ ˜C. However in ˜C it holds that 1 mod (Xn − 1) = Xn mod (xn−1). So bn−1+b0X +· · ·+bn−2Xn−1 mod (Xn−1) = b0X +· · ·+bn−2Xn−1+bn−1Xn mod (Xn− 1) = Xf . Hence ˜C being cyclic means that if f ∈ ˜C then Xf ∈ ˜C as well.

Combining this with the linearity of ˜C the third property becomes 3. for any f ∈ ˜C, g ∈ R it holds that f g, gf ∈ ˜C.

All 3 properties together are exactly the properties that make ˜C an ideal in R. Hence, we can conclude that

Proposition 2. ˜C ⊂ R is linear and cyclic ⇔ ˜C is an ideal in R.

So to find cyclic codes in Fnq is equivalent to finding ideals in R. The following proposition states how to do that.

Proposition 3. I is an ideal in R ⇔ I = f R with f |(Xn− 1).

(8)

For an ideal I generated by a polynomial f , that is I = f R, we will use the alternative notation I = (f ).

Example 2. We will again look at F32 like in example 1. Now R = F2[X]/(X3 − 1) and X3 − 1 = (X + 1)(X2 + X + 1). All cyclic codes can be found by looking at the ideals (1), (X +1), (X2+X +1), (X3−1) (for convenience we won’t denote the bar on each generator).

Now (1) = R and thus has dimension 3. Next, (X + 1) = {0, X + 1, X2+ X, 1 + X2} which is a 2-dimensional cyclic code and it corresponds with C = {(0, 0, 0), (1, 1, 0), (0, 1, 1, ), (1, 0, 1)}

like in example 1. Furthermore, (X2+ X + 1) = {0, X2+ X + 1} and (X3− 1) = {0}.

For a cyclic code (f ) ⊂ R with f |(Xn− 1) we can also determine the dimension.

Proposition 4. Let f |(Xn− 1) ∈ Fq[X], then the cyclic code (f ) has dimension n − deg(f ).

Using proposition 4 there is an easy way of constructing a basis for a cyclic code C = (f ) ⊂ R with f |(Xn− 1). Let k =deg(f ) and write f = Pk

i=0aiXi where each ai ∈ Fq. Since C is cyclic, each of the elements Xjf ∈ C. For 0 ≤ j ≤ n − k − 1 these elements correspond with

(a0, . . . , ak, 0, . . . , 0), (0, a0, . . . , ak, 0, . . . , 0), . . . , (0, . . . , 0, a0, . . . , ak) ∈ Fnq

These vectors are all linearly independent and there are n − k of these vectors which is exactly the dimension of C, so they are a basis of C.

(9)

2 The Number of Subspaces of F

nq

In this section we will count the number of d-dimensional subspaces of Fnq. We start with the most simple case, namely d = 1.

Example 3. To count all the 1-dimensional subspaces of Fnq, it is needed to count how many vectors there are that have a different span. So we have to count all the non-zero vectors up to scalar multiples. There are qn− 1 non-zero vectors in Fnq, so the number of 1-dimensional subspaces is equal to (qn− 1)/(q − 1) =Pn−1

i=1 qi.

To count how many subspaces there are of arbitrary dimension, a counting argument using group theory will be used. Let Vd ⊂ Fnq be the linear subspace of dimension d, 0 < d ≤ n, defined as Vd= Span{e1, . . . , ed}. Let V be another d-dimensional linear subspace of Fnq with a chosen basis {f1, . . . , fd}. We extend this basis to a basis of Fnq: {f1, . . . , fd, fd+1, . . . , fn} is a basis of Fnq for certain fd+1, . . . , fn∈ Fnq. Now define the matrix A : Fnq → Fnq by ei 7→ fi for i = 1, . . . , n. Observe that A ∈ GL(n, Fq) and that AVd = V . But V is an arbitrary d-dimensional linear subspace of Fnq and for each V we can find the matrix as defined above.

So

G(d, n)(Fq) : = {V ⊂ Fnq : V a linear subspace of Fnq, Dim(V ) = d}

= {AVd: A ∈ GL(n, Fq)}.

Note that it is possible that A1Vd= A2Vd for A1, A2∈ GL(n, Fq), A1 6= A2, so the same sub- space could be obtained by different invertible matrices. For example, let {f1, . . . , fd, fd+1, . . . , fn} and {f1, . . . , fd, ˜fd+1, . . . , ˜fn} be two different bases of Fnq and let A1 : ei 7→ fi, i = 1, . . . , n and A2 : ej 7→ fj, for j = 1, . . . d and ej 7→ ˜fi for j = d + 1, . . . , n. Then A1Vd = A2Vd, but A16= A2.

We now look at what happens when AVd= Vd. We get the following lemma.

Lemma 1. The set H := {A ∈ GL(n, Fq) : AVd= Vd} is a subgroup of GL(n, Fq).

Proof. For the identity matrix I we have IVd = Vd, so I ∈ H. If A, B ∈ H, so AVd = Vd and BVd = Vd, then ABVd = AVd = Vd and hence AB ∈ H. Also, since AVd = Vd we get Vd= A−1Vd, so A−1∈ H. Hence H is a subgroup of GL(n, Fq).

We use the following result from group theory found in [1].

Theorem 2. [Lagrange’s Theorem] If H is a subgroup of a finite group G, then #H|#G.

The proof of this theorem shows that #G = [G : H] · #H and G = Sm

i=1giH for certain gi0s ∈ G where m = [G : H] is the index of H in G and giH ∩ gjH = ∅ if i 6= j.

Using this result for the groups used in this section, we get

GL(n, Fq) =

[GL(n,Fq):H]

[

i=1

giH

for certain gi0s ∈ GL(n, Fq) with giH ∩ gjH = ∅ if i 6= j. We get the following lemma.

(10)

Lemma 2. Let G1 = {g1, . . . .gm}, where m = [GL(n, Fq) : H], such that GL(n, Fq) = S

g∈G1gH and Vd= Span{e1, . . . , ed}. Then

G(d, n)(Fq) = {gVd: g ∈ G1}.

Furthermore #G(d, n)(Fq) = [GL(n, Fq) : H].

Proof. We have already seen that for any g ∈ GL(n, Fq) we get gVd ∈ G(d, n)(Fq). So {gVd: g ∈ G1} ⊆ G(d, n)(Fq). Let V ∈ G(d, n)(Fq), then we have seen there is a g ∈ GL(n, Fq) such that V = gVd. Since GL(n, Fq) = S

gi∈G1giH, we get that g = g1h for unique g1 ∈ G1 and h ∈ H. Then V = gVd = g1hVd = g1Vd. So V ∈ {gVd : g ∈ G1}. Hence G(d, n)(Fq) = {gVd : g ∈ G1}. Now the following holds for gi, gj ∈ G1: giVd = gjVd ⇔ gj−1giVd = Vd ⇔ gj−1gi ∈ H ⇔ gj−1giH = H ⇔ giH = gjH ⇔ i = j. So each gi ∈ G1 gives a different giVd∈ G(d, n)(Fq)). Hence #G(d, n)(Fq) = #{gVd: g ∈ G1} = #G1 = [GL(n, Fq) : H].

Now we want to determine [GL(n, Fq) : H] = #GL(n, Fq)/#H.

Lemma 3. The number of elements of GL(n, Fq) is equal to Qn

i=1(qn− qi−1).

Proof. A matrix A ∈ GL(n, Fq) ⇔ A has n linearly independent columns. Let A = [v1, . . . , vn], with vi ∈ Fnq, i = 1, . . . , n. Then v1 needs to have at least one non-zero element for A to be invertible. So there are qn− 1 possibilities. For v2 we require that it is not a multiple of the first column, that is v2 6= a1v1 with a1 ∈ Fq. So for v2 there are qn− q possibilities.

Continuing in this way, we find that vi cannot be of the form vi = a1v1 + · · · + ai−1vi−1

with aj ∈ Fq, j = 1, . . . , i − 1. So, for vi there are qn− qi−1 possibilities. Then we get

#GL(n, Fq) =Qn

i=1(qn− qi−1).

Now that #GL(n, Fq) is known, we need to know how many elements H has.

Lemma 4. The number of elements of H = {A ∈ GL(n, Fq) : AVd= Vd} is equal to

#H = #GL(d, Fq) · #GL(n − d, Fq) · qd(n−d)=

d

Y

i=1

(qd− qi−1) ·

n−d

Y

i=1

(qn−d− qi−1) · qd(n−d) Proof. To compute the number of elements of H, we first determine what the matrices of H look like as a block matrix. For A ∈ H we have that AVd= Vd, where Vd= Span{e1, . . . , ed}.

So A maps each ei, i = 1, . . . d, to a linear combination of e1, . . . , ed. Taking into account that A has to be invertible, we get that A is of the form

A =A11 A12 0 A22



where A11∈ GL(d, Fq), A22∈ GL(n − d, Fq) and A12a d × (n − d) matrix. Since each aij ∈ Fq, there are qd(n−d)possibilities for A12. So we get

#H = #GL(d, Fq) · #GL(n − d, Fq) · qd(n−d)=

d

Y

i=1

(qd− qi−1) ·

n−d

Y

i=1

(qn−d− qi−1) · qd(n−d)

Now we can count #G(d, n)(Fq).

(11)

Corollary 1. The number of elements of G(d, n)(Fq) is equal to

#G(d, n)(Fq) =

Qn

i=1(qn− qi−1) qd(n−d)·Qd

i=1(qd− qi−1) ·Qn−d

i=1(qn−d− qi−1) (2) Proof. From lemma 2 we have #G(d, n)(Fq) = [G : H] = #GL(n, Fq)/#H. The result then follows from lemma 3 and 4.

Observe that #G(d, n)(Fq) is symmetric in the variable d around n/2, that is #G(d, n)(Fq) =

#G(n − d, n)(Fq). Another way to see this, is by using proposition 1. That is, if V ∈ G(d, n)(Fq), then V∈ G(n − d, n)(Fq). So #G(d, n)(Fq) ≤ #G(n − d, n)(Fq). Interchanging d and n − d, it shows that #G(n − d, n)(Fq) ≤ #G(d, n)(Fq) and hence #G(d, n)(Fq) =

#G(n − d, n)(Fq).

2.1 Growth Pattern of #G(d, n)(Fq)

Now that we know the number of d-dimensional subspaces of Fnq, we wish to determine how that number grows as q increases. If q is large enough, then qn  qn−1 for any n ∈ N, so approximately #GL(n, Fq) =Qn

i=1(qn− qi−1) ≈ qn2. To be exact, it holds that

#GL(n, Fq) qn2 =

Qn

i=1(qn− qi−1)

qn2 =

n

Y

i=1

qn− qi−1 qn

=

n

Y

i=1

(1 − qi−1−n) =

n

Y

i=1

(1 − q−i) (3)

Hence

q→∞lim

#GL(n, Fq)

qn2 = lim

q→∞

n

Y

i=1

(1 − q−i) = 1

Using the same argument for #G(d, n)(Fq), if q is large enough, we get approximately, using (2),

#G(d, n)(Fq) ≈ qn2

qd(n−d)· qd2 · q(n−d)2 = qn2−nd−n2+2nd−d2 = qd(n−d) So for fixed d and n this gives

q→∞lim

#G(d, n)(Fq)

qd(n−d) = 1 (4)

Note that #G(d, n)(Fq) has a maximum in the variable d when d = n2 if n is even and when d = n±12 if n is odd. This allows us to determine the growth rate of the total number of subspaces. If n is even we write n = 2m for m ∈ N and use (4) to obtain

q→∞lim Pn

d=0#G(d, n)(Fq)

#G(m, n, q) = lim

q→∞

P2m

d=0qd(2m−d)

qm2 = lim

q→∞

2m

X

d=0

q−(m−d)2 = 1

Hence, as q → ∞ and n is even, it is enough to look at the number of subspaces of dimension

n

2, instead of the total number of subspaces. This will be useful later on. If n is odd instead,

(12)

we write n = 2m + 1. When d = n±12 we know #G(d, n)(Fq) has a maximum. Here that is when d = m or d = m + 1. Using (4) again we obtain

q→∞lim Pn

d=0#G(d, n)(Fq) 2#G(m, n, q) = lim

q→∞

P2m+1

d=0 qd(2m+1−d) 2qm(m+1) = 1

2 lim

q→∞

2m+1

X

d=0

q−(d−m)(d−m−1)

= 1

Thus for n odd and q → ∞ only the subspaces of dimension n±12 are important for counting the total number of subspaces.

Now that the growth patterns of #G(d, n)(Fq) are known as q → ∞, we are interested in the same growth pattern as n tends to infinity, both for fixed dimensions and the total number of subspaces. We start again with (3), that is what can be said about the convergence of

n→∞lim

n

Y

i=1

(1 − q−i)

First, note that for any i ∈ N it holds that 1 − q−i < 1. Using this bound for each i > 1 it gives the upper bound

n→∞lim

n

Y

i=1

(1 − q−i) < 1 −1 q. Second, for the convergence it is required that P

i=1log(1 − q−i) converges where log is the natural logarithm. Now, the logarithm satisfies log(1 − x) ≥ −2x for any x ∈ [0,12]. To show this, define g : [0,12] → R by g(x) = log(1 − x) + 2x. Then g(0) = 0 and

g0(x) = −1

1 − x+ 2 = 2x − 1

x − 1 ≥ 0 for x ∈ [0,1 2].

The claim now follows. Since 0 < q−i12 we find log(1 − q−i) ≥ −2q−i and

−2

X

i=1

q−i= −2( 1

1 −1q − 1) = −2( q

q − 1− 1) = −2 q − 1 so P

i=1log(1 − q−i) converges by the comparison test and P

i=1log(1 − q−i) ≥q−1−2. HenceQ

i=1(1 − q−i) converges, so define φ as φ(q) :=Q

i=1(1 − q−i) and then φ(q) = lim

n→∞

n

Y

i=1

(1 − q−i) = exp( lim

n→∞log(1 − q−i)) ≥ exp( −2 q − 1) See figure 1 for a plot of φ.

We can conclude that

n→∞lim

#GL(n, Fq) φ(q)qn2 = 1

φ(q) lim

n→∞

#GL(n, Fq) qn2 = 1

φ(q) lim

n→∞

n

Y

i=1

(1 − q−i) = 1

(13)

Figure 1: Plot of φ

For the growth rate #G(d, n)(Fq) for d and q fixed while n tends to infinity we will find the growth rate of each term separately and then combine them. We have the following

n→∞lim Qn

i=1(1 − q−i) φ(q)qn2 = 1

n→∞lim

qd(n−d)Qd

i=1(qd− qi−1) qd(n−d)Qd

i=1(qd− qi−1) = 1

n→∞lim

φ(q)q(n−d)2 Qn−d

i=1(qn−d− qi−1) = 1 Now

n→∞lim #G(d, n)(Fq)/ φ(q)qn2 qd(n−d)Qd

i=1(qd− qi−1)φ(q)q(n−d)2

!

= lim

n→∞

Qn

i=1(1 − q−i)

φ(q)qn2 ·qd(n−d)Qd

i=1(qd− qi−1) qd(n−d)Qd

i=1(qd− qi−1) · φ(q)q(n−d)2 Qn−d

i=1(qn−d− qi−1) = 1 Combining the powers of q in the first denominator gives

qn2

qd(n−d)+(n−d)2 = qn2

qnd−d2+n2−2nd+d2 = qnd Hence the growth rate of #G(d, n)(Fq) for d and q fixed is

n→∞lim

#G(d, n)(Fq) qnd/Qd

i=1(qd− qi−1) = 1 (5)

If now d → ∞ as well, this gives us

n→∞lim

d→∞

#G(d, n)(Fq) qd(n−d)

φ(q)

= 1 (6)

(14)

For the growth rate of the total number of subspaces as n tends to infinity, we again have to distinguish the cases n even and n odd. Since for n even only subspaces of dimension n2 need to be considered, but for n odd the subspaces of dimensions 12(n ± 1) are the ones that are important. We first take n = 2m for m ∈ N and use both (5) and (6) and the symmetry of G(d, n)(Fq) to obtain

m→∞lim P2m

d=0#G(d, 2m)(Fq)

#G(m, 2m)(Fq) = lim

m→∞

φ(q)P2m

d=0q2md/Qd

i=1(qd− qi−1) qm2

= lim

m→∞φ(q)

2m

X

d=0

qm(2d−m) Qd

i=1(qd− qi−1)

= lim

m→∞φ(q) qm2 Qm

i=1(qm− qi−1)+ lim

m→∞2φ(q)

2m

X

d=m+1

qm(2d−m) Qd

i=1(qd− qi−1)

= 1 + 2 lim

m→∞

2m

X

d=m+1

q−(m−d)2 = 1 + 2 lim

m→∞

m

X

d=1

q−d2

The series limm→∞Pm

d=1q−d2 = P

d=1q−d2 converges since q−d2 ≤ q−d for any d ≥ 1 and P

d=1q−d = q−11 , so converges. Now define the function φeven by φeven(q) = 1 + 2P d=1q−d2, for a plot of φeven see figure 2, then we find

m→∞lim P2m

d=0#G(d, 2m)(Fq) φeven(q)qm2

φ(q)

= 1

φeven(q) lim

m→∞

φ(q)P2m

d=0#G(d, 2m)(Fq)

qm2 = 1

Figure 2: Plot of φeven

Now we take n = 2m + 1, then the maximum number of subspaces occur at d = m and d = m + 1. Since G(d, n)(Fq) is symmetric, we take 2#G(m, 2m + 1)(Fq) in the limit. Using (5) and (6) again we obtain

(15)

m→∞lim

P2m+1

d=0 #G(d, 2m + 1)(Fq)

2#G(m, 2m + 1)(Fq) = lim

m→∞

φ(q)P2m+1

d=0 q(2m+1)d/Qd

i=1(qd− qi−1) 2qm(m+1)

= lim

m→∞

1 2φ(q)

2m+1

X

d=0

q(d−m)(m+1)+md

Qd

i=1(qd− qi−1)

= lim

m→∞φ(q) q(m+1)2 Qm+1

i=1 (qm− qi−1) + lim

m→∞φ(q)

2m+1

X

d=m+2

q(d−m)(m+1)+md

Qd

i=1(qd− qi−1)

= 1 + lim

m→∞

2m+1

X

d=m+2

q−(d−m)(d−m−1) = 1 + lim

m→∞

m

X

d=1

q−d(d+1)

The series limm→∞Pm

d=1q−d(d+1)=P

d=1q−d(d+1) converges as well since q−d(d+1)≤ q−d for any d ≥ 1. Now define the function φodd by φodd(q) = 1 +P

d=1q−d(d+1), for a plot of φodd see figure 3, then we find

m→∞lim

P2m+1

d=0 #G(d, 2m + 1)(Fq) 2φodd(q)qm(m+1)

φ(q)

= 1

φodd(q) lim

m→∞

φ(q)P2m+1

d=0 #G(d, 2m + 1)(Fq)

2qm(m+1) = 1

Figure 3: Plot of φodd

From figure 2 and figure 3 we can see that φodd(q) → 1 much faster than φeven(q) → 1. This is not surprising, since for n even we took the subspaces of dimension n2, but for n odd we took all the subspaces of dimensions 12(n ± 1). We present the conclusions in the following proposition.

(16)

Proposition 5. Define the functions φ and φ2 as φ(q) = limn→∞Qn

i=1(1 − q−i), φeven(q) = 1 + 2P

d=1q−d2 and φodd(q) = 1 +P

d=1q−d(d+1), then the following limits for the number of subspaces hold.

q→∞lim

#G(d, n)(Fq)

qd(n−d) = 1

q→∞lim P2m

d=0#G(d, 2m)(Fq)

qm2 = 1

q→∞lim

P2m+1

d=0 #G(d, 2m + 1)(Fq)

2qm(m+1) = 1

n→∞lim

#G(d, n)(Fq) qnd/Qd

i=1(qd− qi−1) = 1

n→∞lim

d→∞

#G(d, n)(Fq) qd(n−d)

φ(q)

= 1

m→∞lim P2m

d=0#G(d, 2m)(Fq) φeven(q)qm2

φ(q)

= 1

m→∞lim

P2m+1

d=0 #G(d, 2m + 1)(Fq) 2φodd(q)qm(m+1)

φ(q)

= 1

(17)

3 Schur Product

In this section the Schur product of vectors and subspaces of Fnq will be defined and some basic properties will be shown. We start with the definition.

Definition 8. Let v = (v1, . . . , vn), w = (w1, . . . , wn) ∈ Fnq. The Schur product between v and w, denoted by v ∗ w, is defined as

v ∗ w := (v1w1, . . . , vnwn) For V, W linear subspaces of Fnq the Schur product is defined as

V ∗ W := Span{v ∗ w : v ∈ V, w ∈ W }

In this thesis we only consider the Schur product V ∗ V for a V ∈ G(d, n)(Fq).

Lemma 5. For any u, v, w ∈ Fnq and α ∈ Fq the following properties of the Schur product hold.

1. v ∗ w = w ∗ v 2. (αv) ∗ w = α(v ∗ w)

3. (u + v) ∗ w = u ∗ w + v ∗ w

4. v ∗ w = 0 if and only if for any i ∈ {1, . . . , n} it holds that vi= 0 or wi = 0.

5. For V, W subspaces of Fnq with W ⊂ V , then W ∗ W ⊂ V ∗ V . Proof. 1. Since vi, wi ∈ Fq, we have that viwi = wivi for i = 1, . . . n. So

v ∗ w = (v1w1, . . . , vnwn) = (w1v1, . . . , wnvn) = w ∗ v 2. (αv) ∗ w = (αv1w1, . . . , αvnwn) = α(v1w1, . . . , vnwn) = α(v ∗ w)

3. (u+v)∗w = ((u1+v1)w1, . . . , (un+vn)wn) = (u1w1+v1w1, . . . , unwn+vnwn) = u∗w +v ∗w 4. v ∗ w = 0 if and only if viwi = 0 for i = 1, . . . , n. Since Fq is a domain, viwi = 0 if and only if vi = 0 or wi= 0.

5. Let v, w ∈ W , then v, w ∈ V . So v ∗ w ∈ V ∗ V . Hence W ∗ W ⊂ V ∗ V .

If V ⊂ Fnq is given as V = Span{v1, . . . , vm} with vi ∈ Fnq, i = 1, . . . , m, then any u, w ∈ V can be written as u = Pm

j=1ajvj and w = Pm

j=1bjvj for certain aj, bj ∈ Fq, j = 1, . . . , m.

Then

v ∗ w = (

m

X

j=1

ajvj) ∗ (

m

X

k=1

bkvk) =

m

X

j=1 m

X

k=1

ajbkvj ∗ vk

So V ∗ V is spanned by {vj∗ vk : j = 1, . . . , m, k = 1, . . . , m}. Note, that if all vectors in {v1, . . . , vm} are linearly independent, it is in most cases not true that all vectors in {vj∗vk : j = 1, . . . , m, k = 1, . . . , m} are linearly independent. As a consequence of this, we have the following result.

Corollary 2. If V ⊂ Fnq has dimension d, then Dim(V ∗ V ) ≤ d2 + d.

(18)

Proof. Let {v1. . . . , vd} be a basis of V . Then V ∗ V = Span{vj ∗ vk : j = 1, . . . , m, k = 1, . . . , m} and since the Schur product is symmetric the set {vj ∗ vk |j = 1, . . . , m, k = 1, . . . , m} consists of at most d2 + d elements. Hence Dim(V ∗ V ) ≤ d2 + d.

Some examples to show what the Schur product of subspaces can be will now be given.

Example 4. 1. Let V ⊂ F3q be given as V = Span{v1 = (1, 1, 0), v2 = (0, 1, 1)}. Then v1∗ v1 = (1, 1, 0) = v1

v1∗ v2 = (0, 1, 0) v2∗ v2 = (0, 1, 1) = v2 These are all linearly independent, and so V ∗ V = F3q.

2. Let W ⊂ F3q be given as W = Span{w1= (1, 1, 0), w2 = (0, 0, 1)}. Then w1∗ w1 = (1, 1, 0) = w1

w1∗ w2 = (0, 0, 0) w2∗ w2 = (0, 0, 1) = w2

We see that here W ∗ W = W .

3. Let U ⊂ F33 be given as U = Span{u1= (1, 0, 0), u2 = (0, 2, 1)}. Then u1∗ u1= (1, 0, 0) = u1

u1∗ u2= (0, 0, 0) u2∗ u2= (0, 1, 1)

From this we get U ∗ U = Span{u1, (0, 1, 1)} and we notice that now U ∗ U 6= U . From these examples it is clear that the Schur product of subspaces can vary greatly. The Schur product of a subspace can be the whole space Fnq, the subspace itself, and it can be different from the subspace but with the same dimension. In particular, if V ⊂ Fnq, then it does not hold in general that V ⊂ V ∗ V . However, over F2 we have that v ∗ v = v for any v ∈ Fn2. So then it is always true that V ⊂ V ∗ V .

The Schur product of vectors was defined with respect to the standard basis of Fnq. The next example shows what can happen if a different basis is chosen.

Example 5. Let (e1, e2, e3) be the standard basis of F33 and take v1 = (2, 1, 0), v2 = (0, 1, 2) with respect to this basis. If the subspace V is defined as V = Span{v1, v2}, we have

v1∗ v1 = (1, 1, 0) v1∗ v2 = (0, 1, 0) v2∗ v2 = (0, 1, 1) and thus V ∗ V = F33.

(19)

However, if we take B = {v1, v2, e2} as a basis of F33, then, with respect to the new basis, v1 = (1, 0, 0)B, v2 = (0, 1, 0)B. For the Schur product of V we now get

v1∗ v1= (1, 0, 0)B v1∗ v2= (0, 0, 0)B v2∗ v2= (0, 1, 0)B and so, with respect to the basis B, V ∗ V = V .

From example 5 it is clear that the Schur product depends on the basis chosen for Fnq. This won’t pose much of a problem though, since if B1, B2 are two different basis of Fnq and V, W ⊂ Fnq such that if v = (v1, . . . , vn)B1 ∈ V then there is a w ∈ W with w = (v1, . . . , vn)B2 and vice versa. Let (V ∗ V )Bi denote the Schur product of V with respect to the basis Bi. Then for any u = (u1, . . . , un)B1 ∈ (V ∗ V )B1) there is a x ∈ (W ∗ W )B2 such that x = (u1, . . . , un)B2. So if for a V ⊂ Fnq the Schur product V ∗ V has a certain property with respect to a basis B of Fnq, then there is a W ⊂ Fnq such that W ∗ W has the same property with respect to the standard basis of Fnq.

In example 4 there were examples of subspaces for which the Schur product either increased in dimension or for which the dimension stayed the same. This leads us to wondering whether the Schur product can also decrease in dimension. The following lemma deals with this.

Lemma 6. Let V be a d-dimensional subspace of Fnq, then Dim(V ∗ V ) ≥ d.

Proof. Let V ∈ G(d, n)(Fq), then from (1) there is a basis of V consisting of vectors of the form vi = eσ(i)+Pn

j=d+1ai,jeσ(j) where σ ∈ Sn, i = 1, . . . , d and ai,j ∈ Fq ∀i, j. Then vi ∗ vi = eσ(i)+Pn

j=d+1a2i,jeσ(j). Since eσ(i) ∈ Span{e/ σ(1), . . . , eσ(i−1), eσ(i+1), . . . , eσ(d)} for any i ∈ {1, . . . , d}, it follows that vi∗ vi ∈ Span{v/ 1∗ v1, . . . , vi−1∗ vi−1, vi+1∗ vi+1, . . . , vd∗ vd} for any i ∈ {1, . . . , d}. Hence the vectors v1∗ v1, . . . , vd∗ vd are all linearly independent and thus Dim(V ∗ V ) ≥ d.

(20)

4 Schur Product of Subspaces

Now that some of the basic properties of Schur products are known, we wish to answer the main question: How many linear subspaces of Fnq are there for which the Schur product is not equal to Fnq. An exact answer might not be possible, but upper or lower bounds might be able to be found. In this section various methods will be used to try and find these upper or lower bounds.

Definition 9. The set of all d-dimensional linear subspaces of Fnq for which the Schur prod- uct is, respectively is not, equal to Fnq is denoted by W(d, n, q), respectively V(d, n, q), i.e.

W(d, n, q) = {W ⊂ Fnq : Dim(W ) = d, W ∗ W = Fnq} and V(d, n, q) = G(d, n)(Fq)\W(d, n, q).

A simple case is given by Corollary 2, since if d2 + d < n it holds automatically that for a d-dimensional subspace the Schur product is not the whole space. Determining when

d 2



+ d − n = d(d + 1)

2 − n = 0 we find that d = 12 ± 12

1 + 8n. Since d ≥ 0, the condition d2 + d < n holds when 0 ≤ d <

12+12

1 + 8n. Since√

1 + 8n >√

8n = 2√

2n, a more intuitive bound is 0 ≤ d < −12+√ 2n.

Lemma 7. Let V ∈ G(d, n)(Fq) with 0 ≤ d < −12 +12

1 + 8n then V ∈ V(d, n, q).

Next, from lemma 5 (5), if W ⊂ V ⊂ Fnq, then W ∗ W ⊂ V ∗ V . Hence, if V ∗ V 6= Fnq, then W ∗ W 6= Fnq as well. Thus, if we can count #V(n − 1, n, q), then we can find a lower bound for all the lower dimensional subspaces for which the property holds. This can be done by counting how many different subspaces there are in all the (n − 1)-dimensional subspaces for which the Schur product is not the whole space. However, counting this amount may be quite difficult as one needs to take into account all the subspaces which are counted multiple times.

Using Magma Calculator to count the number of (n − 1)-dimensional subspaces for which the Schur product is not the whole space for various n and q, the following results were obtained.

n \ q 2 3 4 5

1 1 1 1 1

2 3 4 5 6

3 6 9 12 15

4 10 16 22 28

5 15 25 35 45

6 21 49 51 -

From the table, there seems to be a pattern in each column. For q = 2 it seems to be

#V(n − 1, n, 2) =Pn

k=1k. We will prove this is true to gain insight in the general case.

Example 6. Let V ∈ G(n − 1, n)(F2), then V ⊂ V ∗ V . Thus, V ∈ V(n − 1, n, 2) is equivalent to V ∗ V = V in this case. There is a basis of V of the form

{vi = eσ(i)+ aieσ(n): ai ∈ F2, i = 1, . . . , n − 1}

for a certain permutation σ ∈ Sn. Then vi∗ vj = aiajeσ(n) if i 6= j and vi∗ vi = vi. So, if V ∗ V = V , this implies that vi∗ vj = 0 for all i 6= j. Hence, there is at most one ai 6= 0.

(21)

There are two possibilities now: either ai = 0 for all i, or there is exactly one ai 6= 0. If all ai = 0, then V is just the span of n−1 standard basis vectors. So there are n−1n  = n different subspaces for which this is true. Now assume there is exactly one ai 6= 0. Without loss of generality we can take a1 6= 0, so a1 = 1. Then V = Span{eσ(1)+ eσ(n), eσ(2), . . . , eσ(n−1)}.

If σ(1) and σ(n) are chosen, V is fixed. For σ(1) and σ(n) there are n2 different choices.

Hence

#V(n − 1, n, 2) = n +n 2



= n + n(n − 1)

2 = n(n + 1)

2 =

n

X

k=1

k.

From the Magma Calculator data, the number of elements of V(n − 1, n, q) seems to increase as follows: for q = 3 the data suggest #V(n − 1, n, 3) =Pn

k=1(2k − 1), for q = 4 it is #V(n − 1, n, 4) =Pn

k=1(3k − 2). In general this seems to be #V(n − 1, n, q) =Pn

k=1((q − 1)k − (q − 2).

This can be rewritten as

n

X

k=1

((q − 1)k − (q − 2) = (q − 1)n(n + 1)

2 − n(q − 2)

= q(n(n + 1)

2 − n) −n(n + 1) 2 + 2n

= qn 2



−n 2

 + n

= (q − 1)n 2

 + n

We will prove this is true, using mostly the same arguments used in example 6 but adjusted where needed.

Lemma 8. The number of elements of V(n − 1, n, q) is equal to

#V(n − 1, n, q) = qn(n − 1)

2 − n(n − 3)

2 .

Proof. Let V ∈ V(n − 1, n, q). We choose a basis of V of the form {vi = eσ(i)+ aieσ(n): ai ∈ Fq, i = 1, . . . , n − 1}

for a certain σ ∈ Sn. Then vi∗ vi = eσ(i)+ a2ieσ(n) and vi∗ vj = aiajeσ(n) when i 6= j. The vectors v1∗ v1, . . . , vn−1∗ vn−1 are linearly independent and vi∗ vj ∈ Span{v/ 1∗ v1, . . . , vn−1∗ vn−1}. Since V ∈ V(n − 1, n, q), this implies vi ∗ vj = 0 for all i 6= j. Hence there is at most one ai 6= 0. If all ai = 0, then V is just the span of n − 1 standard basis vectors.

So there are n−1n 

= n different subspaces for which this is true. Now suppose there is exactly one ai 6= 0. Without loss of generality we can take a1 6= 0. Then V = Span{eσ(1)+ a1eσ(n), eσ(2), . . . , eσ(n−1)}. For a1 there are q − 1 possibilities. If σ(1) and σ(n) are chosen, V is fixed. For σ(1) and σ(n) there are n2

different choices. Hence #V(n − 1, n, q) = (q − 1) n2 + n.

From the proof it is seen that

V(n − 1, n, q) = {V = Span{eσ(1)+ a1eσ(n), eσ(2), . . . , eσ(n−1)} : σ ∈ Sn, a1∈ Fq}

(22)

For (n − 1)-dimensional subspaces we could determine #V(n − 1, n, q) exact. One could wonder if, using similar arguments, #V(d, n, q) can be determined for arbitrary d. This might be possible for some specific d’s, however the combinatorics involved quickly become complicated as d approaches n2. For d = n − 2 we will compute #V(n − 2, n, q) analytically, which will also make clear how the computations are more difficult from the n − 1-dimensional case.

Any element V of G(n − 2, n)(Fq) has a basis of the form

{vi= eσ(i)+ aieσ(n−1)+ bieσ(n) : ai, bi∈ Fq, i = 1, . . . , n − 2, σ ∈ Sn}

If V ∈ V(n − 2, n, q), then either Dim(V ∗ V ) = n − 2 or Dim(V ∗ V ) = n − 1. When n = 2 or n = 3 it is trivial to compute #V(n − 2, n, q), since then #V(n − 2, n, q) = #G(n − 2, n)(Fq).

So let n ≥ 4, we will look at both cases separately. Let Dim(V ∗ V ) = n − 2. Since the vectors v1∗ v1, . . . , vn−2∗ vn−2 are linearly independent, this means that vi∗ vj = 0 whenever i 6= j.

Hence there are at most one ai 6= 0 and one bi 6= 0. If there is one ai6= 0, then we can choose i arbitrarily since permutations are used to count all the different possibilities. There are the following cases:

1. ai = bi = 0 for i = 1, . . . , n − 2, so V = Span{eσ(1), . . . , eσ(n−2)} for which there are n2 possibilities.

2. an−2 6= 0 and bj = 0 for j = 1, . . . , n − 2. So V = Span{eσ(1), . . . , eσ(n−3), eσ(n−2)+ an−2eσ(n−1)}. There are q − 1 choices for an−2, n2 different choices for σ(n − 2) and σ(n − 1). There are then n − 2 choices left for σ(n), which each give rise to a different subspace. Hence there are n2(q − 1)(n − 2) different subspaces of this form.

3. There is exactly one ai 6= 0 and bj 6= 0 with i 6= j. Take i = n − 3 and j = n − 2, so V = Span{eσ(1), . . . , eσ(n−4), eσ(n−3)+ an−3eσ(n−1), eσ(n−2)+ bn−2eσ(n)}

There are (q − 1)2 different choices for an−3 and bn−2 together and n4 different possi- bilities for σ(n − 3), . . . , σ(n). To determine how many different subspaces there are for σ(n − 3), . . . , σ(n) fixed, it is equal to counting how many different two 2cycles there are on S4, which is 3. So in this case there are 3 n4(q − 1)2 different possibilities.

4. an−26= 0 and bn−2 6= 0 and ai = bi = 0 when i 6= n−2. Here V = Span{eσ(1), . . . , eσ(n−2)+ an−2eσ(n−1)+ bn−2eσ(n)}. For the an−2 and bn−2 there are (q − 1)2 choices. Switching any of the σ(n − 2), σ(n − 1) and σ(n) doesn’t give any new subspaces and when they are chosen the rest is fixed. Hence there are n3(q − 1)2 different subspaces of this form.

These were all the cases of Dim(V ∗ V ) = n − 2. So now we let Dim(V ∗ V ) = n − 1. Since the vectors v1 ∗ v1, . . . , vn−2∗ vn−2 are linearly independent, this means that there are k, l, k 6= l such that vk∗ vl∈ Span{v/ 1∗ v1, . . . , vn−1∗ vn−1} and vi∗ vj = αvk∗ vl for all i 6= j and α ∈ Fq. This happens in the following cases.

5. There are at least two ai6= 0 and bj = 0 for all j. So V is of the form V = Span{eσ(1)+ a1eσ(n−1), . . . , eσ(n−2)+ an−2eσ(n−1)}. First note that each choice of σ(n) gives rise to the same number of subspaces with all the choices for the ai’s and σ(1), . . . , σ(n − 1) and all these subspaces are different. Consider now the case with σ the identity, then the matrix of V with respect to the basis used here is:

Referenties

GERELATEERDE DOCUMENTEN

Let V be the 3-dimensional vector space of polynomials of degree at most 2 with coefficients in R.. Give the signature of φ and a diagonalizing basis

In this thesis, we aim to design and compare different biomaterials in terms of growth factor (BMP-2) release profile and bone regenerative capacity with a final goal to develop

We estimate the relative importance of tidal stripping for the two dSphs we study here, and we introduce the key concepts used in this paper to derive our lower and upper bounds on

In this thesis we show how to construct convexity preserving upper and lower bounds in the case that no derivative information is available, and construct new Sandwich algorithms