1 zero-sum square matrices paul balister yair caro cecil rousseau raphael yuster

22
1 Zero-sum square matrices Paul Balister Yair Caro Cecil Rousseau Raphael Yuster

Upload: howard-powell

Post on 21-Jan-2016

216 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: 1 Zero-sum square matrices Paul Balister Yair Caro Cecil Rousseau Raphael Yuster

1

Zero-sum square matrices

• Paul Balister

• Yair Caro

• Cecil Rousseau

• Raphael Yuster

Page 2: 1 Zero-sum square matrices Paul Balister Yair Caro Cecil Rousseau Raphael Yuster

2

Definitions• Let A be a matrix over the integers, and let p be a positive

integer. A submatrix B of A is called zero-sum mod p if the sum of each row and each column of B is a multiple of p.

• Let M(p,k) denote the least integer m for which every square matrix of order at least m over the integers, has a square submatrix of order k which is zero-sum mod p. As usual in zero-sum problems, we assume that k is a multiple of p, otherwise M(p,k) may be infinite. If k is a multiple of p, then M(p,k) is finite by a standard Ramsey-type argument: Consider the elements of our matrices as taken from Zp

+ : we have a coloring of an m*m grid with p colors, so, for sufficiently large m, there is a k*k sub-grid which is monochromatic. This translates to a (very specialized) zero-sum mod p submatrix of order k.

1 4 2 1

2 1 0 1

2 1 1 1

Page 3: 1 Zero-sum square matrices Paul Balister Yair Caro Cecil Rousseau Raphael Yuster

3

Graph-theoretic interpretation (p=2)• A zero-one matrix A=Am,n can be interpreted as the adjacency

matrix of a bipartite graph G=(X Y, E), where X=x1,…,xm and Y=y1,…,yn. In this interpretation, A(i,j)=1 if and only if xi is adjacent to yj. Hence, M(2,k) is the least integer m that guarantees that every bipartite graph with equal vertex classes having cardinality m, has an induced subgraph with equal vertex classes having cardinality k, such that all degrees are even.

• By the Ramsey-type argument mentioned above, it is obvious that M(2,k) ≤ B(k) where B(k) is the Bipartite Ramsey Number. The best upper bound and lower bound for B(k) are currently:

√2e-1k2k/2 < B(k) ≤ 2k(k-1)+1In particular, we get M(2,k) ≤ 2k(k-1)+1.

Page 4: 1 Zero-sum square matrices Paul Balister Yair Caro Cecil Rousseau Raphael Yuster

4

Main result• Unlike the exponential lower bound for B(k), we can show that

M(2,k) is linear in k. We can also show that M(3,k) is linear in k for infinitely many values of k.

• Theorem:

– limsup M(2,k)/k ≤ 4.

– liminf M(3,k)/k ≤ 20. In particular, for every positive integer s, M(3,3s) ≤ 20 * 3s(1+o(1))

• Conjecture:

For every positive integer p, there exists a constant cp such that limsup M(p,k)/k ≤ cp.

Page 5: 1 Zero-sum square matrices Paul Balister Yair Caro Cecil Rousseau Raphael Yuster

5

•The theorem implies the conjecture for p=2 with c2 ≤ 4.•It is not difficult to construct examples showing M(2,k) ≥2k+1. Thus, we cannot have c2 < 2.•The construction (example for k=4):

1 1 0 0 0 0 0 00 1 1 0 0 0 0 00 0 1 1 0 0 0 00 0 0 1 1 0 0 00 0 0 0 1 1 0 00 0 0 0 0 1 1 00 0 0 0 0 0 1 11 0 0 0 0 0 0 1

Page 6: 1 Zero-sum square matrices Paul Balister Yair Caro Cecil Rousseau Raphael Yuster

6

• It is extremely difficult to compute M(2,k) (moreover M(p,k) for p >2) even for relatively small values of k. It is an easy exercise to show that M(2,2)=5. A naive approach for computing M(2,4) with a computer program which generates all possible 0-1 matrices and testing them would fail. By using, once again, the fact that M(2,k) ≥ 2k+1, we know that M(2,4) ≥ 9. Even generating all 0-1 matrices of order 9 is not feasible as there are 281 such matrices (and even then, maybe the correct value is larger than 9). We used a sophisticated computer search which enables us to narrow down the checks considerably. With the aid of this search we were able to show:

• Proposition:

– M(2,4)=10. A binary matrix of order 9 that demonstrates M(2,4) > 9 is the following (it is unique up to complement and row/column interchanges):

Page 7: 1 Zero-sum square matrices Paul Balister Yair Caro Cecil Rousseau Raphael Yuster

7

J 0 II J 00 I J

•M(2,6) is already an open problem.

•For general p, it is not difficult to show that M(p,k)/k grows exponentially with p. For fixed p the lower bound is linear in k

M(p,k) ≥ (k/5)ep/2e.

•For p > 3, nothing but the trivial exponential upper bound that follows from bipartite Ramsey numbers with p colors is known. Thus, the conjecture is far from being proved except for p=2 (and p=3 with the liminf version).

Page 8: 1 Zero-sum square matrices Paul Balister Yair Caro Cecil Rousseau Raphael Yuster

8

Proof of the main result• First tool: A theorem of Enomoto, Frankl, Ito and Nomura.• Recall that a linear binary code of length t contains the

distance k if it contains as a codeword a vector with k ones and t-k zeroes. Their theorem:

• Let k be even, and let A be a binary matrix with k-1 rows and 2k columns. Then, A is a parity-check matrix of a linear code which contains the distance k.

• Example: k=4 and some arbitrary 3*8 matrix A

0 1 1 0 1 0 1 11 0 0 0 1 1 0 10 1 0 1 1 1 0 0

The theorem simply says that there exists a binary vector x of length 8 with 4 positions equaling 1 such that Ax=0 (in Z2).

Page 9: 1 Zero-sum square matrices Paul Balister Yair Caro Cecil Rousseau Raphael Yuster

9

.)1(1)(1

k

i

ipGD

• Second tool: A theorem of Olson

• Let G be a finite Abelian group. The Davenport Constant of G, denoted D(G), is the least positive integer t such that in any sequence of t elements of G there is a (nonempty) subsequence whose sum is zero.

• Let p be a prime, and let G be a p-group. Then,

• Olson proved the following:

kpppZZZG 11

Page 10: 1 Zero-sum square matrices Paul Balister Yair Caro Cecil Rousseau Raphael Yuster

10

• An even vector is a binary vector whose number of nonzero coordinates is even. More generally, a p-divisible vector is a vector from (Zp)k whose sum of coordinates is divisible by p. We use the latter two tools to prove the following lemmas:

• The p=2 lemma:Let k be an even positive integer. Then, every sequence of 2k even vectors of length k each, contains a subsequence of exactly k vectors whose sum is the zero vector. (If k is a power of 2 we can improve 2k to 2k-1).

• Proof: Consider the matrix A whose columns are the elements of the sequence of 2k even vectors of length k. Ignoring the last row of A we have a parity check matrix of a linear code which contains the distance k. This means we can choose k columns whose sum is zero. Adding back the last coordinate to these columns still gives a zero sum, since the columns of A are even vectors.

Page 11: 1 Zero-sum square matrices Paul Balister Yair Caro Cecil Rousseau Raphael Yuster

11

• The p=3 lemma:Let k be a power of 3. Every sequence of 5k-2 vectors of length k each that are 3-divisible, contains a subsequence of exactly k vectors whose sum is the zero vector (in (Z3)k).

• Proof: Let a1,…,a5k-2 be a sequence of 3-divisible vectors where ai (Z3)k. Let bi be the same as ai, except that the last coordinate of bi is always one. Note that bi need not be 3-divisible anymore. We consider the bi as elements of G=(Z3)k-1×Z3k. Since k is a power of 3, G is a p-group for p=3. According to Olson’s Theorem, D(G)=1+2(k-1)+(3k-1)=5k-2. Hence, there is a nonempty subsequence of b1,…,b5k-2 whose sum is zero. Since the last coordinate of bi is the element 1 of Z3k, such a subsequence must contain exactly 3k elements. Thus, w.l.o.g., b1+…+b3k=0. Now let us consider these 3k vectors as elements of G’=(Z3)k-1×Zk. Again, G’ is a p-group for p=3, so D(G’)=1+2(k-1)+(k-1)=3k-2. Hence, there is a

Page 12: 1 Zero-sum square matrices Paul Balister Yair Caro Cecil Rousseau Raphael Yuster

12

nonempty subsequence of b1,…,b3k whose sum is zero. Since the last coordinate of bi is the element 1 of Zk, such a subsequence must contain either exactly k or exactly 2k elements. In the latter case, the sum of the remaining k vectors not in the subsequence is also zero. We proved there is always a subsequence of k elements whose sum is zero in G’. Now, since the ai are 3-divisible vectors we immediately get that after setting the last coordinate back to its original value the corresponding subsequence of the ai is also zero.

Page 13: 1 Zero-sum square matrices Paul Balister Yair Caro Cecil Rousseau Raphael Yuster

13

Further definitions• Let x be a vector. A k-subvector of x is taken by selecting k

coordinates of x. Clearly, if x (Zp)t then every k-subvector is a member of (Zp)k, and x has exactly t!/(k!(t-k)!) distinct k-subvectors.

• Let ek(x) denote the number of p-divisible k-subvectors of x. We shall always assume that k is divisible by p. If x=(1,…1) then trivially ek(x)= t!/(k!(t-k)!), but, in general, ek(x) may be a lot smaller.For example, if x=(1,1,1,0,0) (Z2)5 then e2(x)=4<10. Now, put

)(min),,()(

xetkpf kzx t

p

Page 14: 1 Zero-sum square matrices Paul Balister Yair Caro Cecil Rousseau Raphael Yuster

14

• The counting lemma:– Let k be even, and let t satisfy

Then, M(2,k) ≤ t.– Let k be a power of 2, and let t satisfy

Then, M(2,k) ≤ t.– Let k be a power of 3, and let t satisfy

Then, M(3,k) ≤ t.• Proof: We prove the first part. Let A be a square 0-1 matrix of

order t. We must show that A has a square submatrix A’ of order k, such that all rows and all columns of A’ are even vectors. The condition in the lemma implies that there exists a subset of k columns of A, such that the set of t line vectors they induce contains more than 2k-1 even vectors. Thus, there is also a subset of 2k rows such that in the induced 2k * k submatrix each row has zero sum. By the p=2 lemma there is a subset of k rows (out of the 2k rows) such that each column is also zero sum. The proofs of the other cases are identical using the p=2 lemma (Olson version) and the p=3 lemma.

t

k

k

ttkf

22),,2(

t

k

k

ttkf

12),,2(

t

k

k

ttkf

35),,3(

Page 15: 1 Zero-sum square matrices Paul Balister Yair Caro Cecil Rousseau Raphael Yuster

15

• It is not difficult to compute explicit small values of f(p,k,t). Clearly, if x and y are two binary vectors having the same length, and the same number of zeroes then ek(x)=ek(y). More generally, if r denotes the number of zeroes in a binary vector of length t then,

Similarly, for the case p=3, if r denotes the number of zeroes and s denotes the number of ones in a 3-ary vector of length t then,

• Example: f(2,4,13)=343. The minimum is obtained when r=9 (or r=4). We can use the second part of the counting lemma and see that 343 > (6/13)*13!/(4!9!). Thus, we immediately get M(2,4) ≤ 13.

j

rt

jk

rtkf

k

j

t

r 22min),,2(

2/

00

k

j bajba

rt

s

t

r b

srt

a

s

jk

rtkf

0 3mod02,00

minmin),,3(

Page 16: 1 Zero-sum square matrices Paul Balister Yair Caro Cecil Rousseau Raphael Yuster

16

• Computing lower bounds for f(p,k,t) in the general case is a more complicated task. e.g.:what would one guess is the minimum possible fraction of3-divisible k-subvectors of a 20k-vector?

• The next two lemmas give a bound for every p ≥ 2, in case k=(t) and t→∞.

• Before we state the lemmas, we need to define a particular constant: For a sequence of integers a1,…,am let wp(a1,…,am) be the complex number

• Example: let z=e2i/3 w3(4,8)=w3(1,2)=0.25(4+(1+z)(1+z2)+(1+z2)(1+z))=1.5

• This number is always real (in fact, rational).

• Define zp = inf wp(a1,…,am) where the infimum is taken over all finite sequences a1,…,am of all lengths m ≥ 0.

1 1

1 )1(2),,(p

j

m

j

ammp aaw

Page 17: 1 Zero-sum square matrices Paul Balister Yair Caro Cecil Rousseau Raphael Yuster

17

• If all ai=1 and ≠1 then |(1+ aj)/2| ≤ cos(/p)=<1, and sowp ≤ 1+(p-1)m. Letting m→∞ we see that zp ≤ 1. Also note that, trivially, z2=1.

• The zp lemma: zp=p21-p. • A combinatorial proof: The sum of j taken over all roots is p if

p | j and 0 otherwise. Hence wp(a1,…,am) is just p2-m times the number of subsets X of {1,…,m} such that ∑ixai = 0 mod p.If m=p-1 and all ai=1 then the only such subset X is . Hencezp ≤ p21-p and the result will follow if we can show the number of subsets X is always at least 2m+1-p.We prove this by induction on m. Let Sm be the set of residue classes mod p that can be written in the form ∑ixai for someX {1,…,m} and assume that for each element of Sm there are at least 2sm such subsets X. Now Sm=Sm-1 (Sm-1+am). We consider two cases:

Page 18: 1 Zero-sum square matrices Paul Balister Yair Caro Cecil Rousseau Raphael Yuster

18

1. Sm ≠ Sm-1. Then |Sm|>|Sm-1|.

2. Sm=Sm-1. Then Sm-1+am=Sm-1 and so if x Sm then both x and x-am lie in Sm-1. Thus x can be written as a sum of the ai's in at least 2sm-1 ways which do not involve am and 2sm-1

ways that do involve am. Thus sm ≥ sm-1+1.

In all cases |Sm|+sm ≥ |Sm-1|+sm-1+1. Now S0={0} and s0=0, so by induction |Sm|+sm ≥ m+1. Since |Sm| ≤ p we havesm ≥ m+1-p. Thus 0 Sm can be written as a sum of ai mod p in at least 2m+1-p ways.

Page 19: 1 Zero-sum square matrices Paul Balister Yair Caro Cecil Rousseau Raphael Yuster

19

• The f(p,k,t) lower bound lemma:Let k ≤ t/2. Then for any fixed prime p:

• Proof: Let T={1,…,t} and let x=(x1,…,xt) (Zp)t be a fixed vector for which ek(x)=f(p,k,t). Assume p | k and k ≤ t/2. We wish to estimate the number f(p,k,t) of sets K T, |K|=k with ∑iKxi = 0 mod p.Fix the k element subset K={1,…,k} of T and instead count the number pk of permutations St such that ∑i=1..kx(i)=0 mod p.It is clear that pk=f(p,k,t) k!(t-k)! Define

Then (1)=1, |()| ≤ 1, pk/t!=p-1 ∑ ().Let be a partition of T into k pairs Si={ai,bi}, i=1,…,k, and a remaining set S0 of t-2k numbers. We also regard as the set of permutations such that {(i), (i+k)}=Si for all i=1,…,k.

11211),,( 1k

pk

p ok

to

k

t

p

ztkpf

k

i

x i

t 1!

1

Page 20: 1 Zero-sum square matrices Paul Balister Yair Caro Cecil Rousseau Raphael Yuster

20

For each there are exactly 2k(t-2k)! choices for .

where and the expectation is over a uniform random choice of (there are exactly t!/(2k(t-2k)!) choices for ). Assume now ≠1. If a ≠ b mod p then | (a+ b)/2| ≤ cos(/p)=<1.Thus |(, )| ≤ n where n is the number of i=1,…,k such that xai ≠ xbi mod p. Hence | ()| ≤ E[n].

Let rs be the number of coordinates with xi= s mod p. ThenE[n]==k ∑i<j (2rirj)/(t(t-1)). It is not hard to show that Var[n]=O(): Indeed, note that n= ∑j=1..k n(j) where

n(j) are indicator random variables that are equal to 1 in casexaj ≠ xbj mod p. Since Cov[n(j1), n(j2)] = O(2/(k2t)) and since ≤ k < t we have: Var[n] ≤ + k2O (2/(k2t)) = O ().

,!

!2

!

1

11

Et

kt

t

k

i

xxk

i

x ibiai

k

i

xxibia

1 2,

Page 21: 1 Zero-sum square matrices Paul Balister Yair Caro Cecil Rousseau Raphael Yuster

21

By Tchebychev's Inequality, if →∞ then n is almost surely large as well and ()=o(1). In this case

Now assume is bounded and k →∞. Let r=t- max ri be the number of coordinates of x not equal to the most common value.Then ≥ 2kr(t/p)/t2, so r=O(t/k) and kr2/t2=ok(1). By adding a constant vector to x and using p | k we can assume the most common coefficient in x is 0. Let m be the number of Si with both xai and xbi non-zero. Then, trivially,Pr[m >0] ≤ E [m]=O(kr2/t2)=o(1).If m=0 then (, ) is a product of n terms of the form(1+ ai)/2. Hence, the sum over all roots of unity of (, ) is at least zp. It follows that

)1()1(111),,(

op

zo

pp

k

ttkpf p

)1(1),,(

op

z

p

k

ttkpf p

Page 22: 1 Zero-sum square matrices Paul Balister Yair Caro Cecil Rousseau Raphael Yuster

22

Completing the proof of the main theorem:Let >0 and let 0< < /(1+ ). Let k be an even integer sufficiently large such that

Such a k exists by the lower bound lemma. Now, Since0.5(1- )>(2k-1)/(4k(1+ )) we get by the counting lemma that M(2,k) ≤ 4k(1+ ). Similarly, by using the second part of the counting lemma we get that for k sufficiently large which is a power of 3,

and so M(3,k) ≤ 20k(1+ ).

)1()1(4

2

1))1(4,,2(

k

kkkf

k

k

k

kkkf

)1(20

)1(20

35))1(20,,3(