ArticlePDF Available

Saturating sets in projective planes and hypergraph covers

Authors:

Abstract

Let Πq\Pi_q be an arbitrary finite projective plane of order q. A subset S of its points is called saturating if any point outside S is collinear with a pair of points from S. Applying probabilistic tools we improve the upper bound on the smallest possible size of the saturating set to 3qlnq\sqrt{3q\ln{q}}. The same result is presented using an algorithmic approach as well, which points out the connection with the transversal number of uniform multiple intersecting hypergraphs.
Saturating sets in projective planes and hypergraph
covers
Zolt´an L´or´ant Nagy
MTA–ELTE Geometric and Algebraic Combinatorics Research Group
H–1117 Budapest, P´azm´any P. s´et´any 1/C, Hungary
nagyzoli@cs.elte.hu
Abstract
Let Πqbe an arbitrary finite projective plane of order q. A subset Sof its points
is called saturating if any point outside Sis collinear with a pair of points from S.
Applying probabilistic tools we improve the upper bound on the smallest possible
size of the saturating set to d3qln qe+d(q+ 1)/2e. The same result is presented
using an algorithmic approach as well, which points out the connection with the
transversal number of uniform multiple intersecting hypergraphs.
Keywords: projective plane, saturating set, dense set, transversal, blocking set,
complete arcs, hypergraph cover
1 Introduction
Let Πqbe an arbitrary finite projective plane of order q, and P,Ldenote the point and
line set of the plane, respectively. A subset Sof its points is called a saturating set if any
point outside Sis collinear with two points in S. In other words, the secants of the point
set cover the whole plane.
The size of saturating (or sometimes also called dense, saturated, determined) sets has
been widely investigated, see [14] for a recent survey. The importance of such sets relies on
connections to covering codes [8], algebraic curves over finite fields [13], sumset theory [15]
and complete arcs [14]. In fact, there is a one-to-one correspondence between saturating
sets in PG(2, q) of cardinality |S|and linear q-ary covering codes of codimension 3 and
covering radius 2, of length |S|. Here we only refer the reader to [7, 8] for the details.
The author is supported by the Hungarian Research Grant (OTKA) No. K 120154 and by the J´anos
Bolyai Research Scholarship of the Hungarian Academy of Sciences
1
Notation 1.1. For any points Pand Q, the set hP, Qidenotes the points of the line
determined by Pand Q. In general, if Sand Tare point sets, then the set hS, T idenotes
the points of the lines determined by a point from Sand a point from T. For convenience,
if one of the sets consists of a single point, we use Pinstead of {P}.
The set hS, Sidenotes the points of the lines determined by a distinct point pair from S.
Clearly, the notation rewrites the concept of saturating sets to the following property on
a set S:hS, Si=Pq).
An arc in the projective plane is a set of points such that no three points of the set are
on a line, and arcs which can not be extended to a larger arc are called complete arcs.
Observe that complete arcs are special separating sets. Due to the pioneer work of Ughi,
Sz˝onyi, Boros, Giulietti, Davydov, Marcugini, Pambianco, Kim and Vu [6, 9, 14, 16, 19]
several bounds are known for particular finite projective planes concerning maximal arcs
and saturating sets.
It is well known (and mentioned first in [19] concerning saturating sets) that the following
proposition must hold:
Proposition 1.2. |S|>2q+ 1 for any saturating set Sin Πq.
This estimate is also known as the Lunelli-Sce bound for complete arcs.
If qis a square and the plane is Desarguesian, then the existence of saturating sets with
the same order of magnitude up to a constant factor is known due to Boros, Sz˝onyi and
Tichler [6].
Proposition 1.3. [6, 19] The union of three non-concurrent Baer sublines in a Baer
subplane from a Desarguesian plane of square order is a saturating set of size 3q.
In fact, this can be even generalized since any 2-blocking set of a Baer-subplane, which
meets every line in at least 2 points, provides a saturating set of the ground plane. Thus
if qis a 4th power, this bound can be even improved to roughly 2qdue to Davydov
et al. [10] and Kiss et al. [17]. Similar order of magnitude can be achieved if qis a 6th
power [10].
Proposition 1.4. [10, 17] The union of the point set of two disjoint Baer subplane of a
Baer subplane from a Desarguesian plane is a saturating set, of size 2q+ 2 4
q+ 2.
(The existence of Baer-subplanes in Baer subplanes of PG(2, q) tacitly implies that qis a
4th power.)
However, in the general case when the plane is not necessarily Desarguesian or the order
is arbitrary, we only have much weaker results. Following the footprints of Boros, Sz˝onyi
and Tichler [6], Bartoli, Davydov, Giulietti, Marcugini and Pambianco [3] obtained an
estimate on the minimal size of a saturating set in Πq.
2
Proposition 1.5 (Bartoli et al.).[3] min |S| ≤ p4(q+ 1) ln (q+ 1) + 2, if Sis a satu-
rating set in Πq.
Our main theorem improves the constant term to 3.
Theorem 1.6. min |S| ≤ (3 + o(1))qln q, if Sis a saturating set in Πq.
In fact, we prove the exact result min |S|≤d3qln qe+d(q+ 1)/2e.
Computer searches suggest that for Galois planes, this bound is still not sharp and the
correct order of magnitude is probably O(q), see [9]. However, we will see in Section 4
that our estimate might be sharp in general, if we cannot built on any algebraic structure
of the plane. We focus mainly on the order of magnitude of min |S|, partly because it has
been determined for qsmall in P G(2, q), see [4, 5] and the references therein. In particular,
Bartoli et al. showed an upper bound slightly above (3)qln qfor the smallest size of
complete arcs in P G(2, q) under a certain probabilistic conjecture, and their computer
search results suggest that this is indeed the right order of magnitude [4, 5]. These results
also support the assumption that Theorem 1.6 might be sharp for general planes.
The paper is built up as follows. In Section 2, we present the first proof of Theorem
1.6, applying a refined version of the first moment method, where an almost suitable
structure is proved via random argument, which can be fixed. This idea appeared first in
Erd˝os’ work, who proved this way the existence of dense, complete bipartite graph free
and dense, even cycle free graphs, see e.g. in [2] (Chapter 6). In Section 3, we proceed by
showing that an advanced greedy-type algorithm also provides a saturating set of this size.
Finally, in Section 4 we analyze the above approaches and point out their connection to
hypergraph cover problems, which suggests that probably in general (non-desarguesian)
projective planes it would be hard to improve the order of magnitude. Note that similar
connection between geometric problems and transversals in hypergraph appeared before
several times [1]. We finish with a number of open problems.
2 Probabilistic argument - First proof
Proof of Theorem 1.6. The key idea is the following: we put in every point to our future
saturating set Swith a given probability p:= p(q) determined later on, independently
from the other points. Then we complete it to obtain a saturating set via Lemma 2.1.
Lemma 2.1. Consider a set of points Sand the corresponding set Rcontaining those
points which are not determined by the lines of S. One can add at most d|R|/2epoints
to Sso that the resulting set is a saturating set.
Proof. Pair up the points of Rand for each pair {x, x0}, choose the intersection of hx, si
and hx,0s0ifor two different points s, s0in S. Clearly, the addition of the intersection
makes the points x, x0determined.
3
Clearly if Xis random variable counting the number of points in S, we have E(X) =
p(q2+q+ 1). Then if Yis random variable counting the number of points not in hS, Si
we get
E(Y) = (q2+q+ 1)P(a given point is not determined )
= (q2+q+ 1)(1 p)(q2+q+1) p
1p+1 + qp
1pq+1!.(2.1)
Indeed, in order to obtain that a given point is not determined, either a point is chosen,
thus no further points may appear; or it is not chosen and on every line on the point at
most one point is chosen. The expression we get this way simplifies to (2.1).
It is easy to see that (q2+q+ 1)(1 p)(q2+q+1) p
1pdoes not contribute to the main
term of E(Y), so we omit it.
We want to choose the value of p:= p(q) in such a way that the sum E(X) + d1
2E(Y)eis
minimized, which would provide a suitable saturating set via the first moment method,
in view of Lemma 2.1. This yields p1
2exp 2p
p2(q2+q+ 1)exp (q+ 1) ·2pq
1p
pq
1p+2 if we
apply the approximation forms via Taylor’s theorem
(1+1/x)x+0.5=e·1 + 1
12x2O1
x3 (x→ ∞)1+zexp 2z
z+ 2(1+O(z3)) (z0),
for z=pq
1p, as 1 + z2
12 O(z3)2z
z+2 = (1 + O(z3)) if z1, and
(11/x)x0.5=e1·11
12x2O1
x3 (x→ ∞)1pexp 2p
p2(1O(p3)) (p0).
From this, one can derive
1
2E(Y) = 1
2(q2+q+ 1) exp 2p
p2(q2+q+ 1)·q2+q+1
1exp (q+ 1) ·2pq
1p
pq
1p+ 2!·q+1
2,
where ∆1= (1 O(p3)) and ∆2= (1 + O(z3)) denote the error terms with z=pq
1p.
4
By simplifying the main term in E(Y), we get
1
2(q2+q+ 1) ·exp 2p
p2(q2+q+ 1)exp (q+ 1) ·2pq
1p
pq
1p+ 2!=
1
2(q2+q+ 1) ·exp 2p
p2+ 2p(q2+q)·(1
p2+1
pq + 2(1 p))
1
2(q2+q+ 1) ·exp 2p
p2+ 2pq(q+ 1) ·p(q1)
4
1
2(q2+q+ 1) ·exp 1
2q(q21)p2,
(2.2)
after we omit the smaller order terms during the approximation.
The calculation above leads to the choice p=3qlnq
q2+q+1 , which in fact implies that the main
term in E(Y) (2.2) equals to O(q), and it is easy to check that the same holds for the
error terms coming from ∆1and ∆2. This in turn provides the existence of a saturating
set of size
E(X) + d1
2E(Y)e= (1 + o(1))p3qln q.
3 Algorithmic approach - Second proof
Below we will present an algorithm to choose the point set of a saturating set S. We start
with an empty set S0in the beginning, and increase its cardinality by adding one point
in each step.
Notation 3.1. In the ith step, we denote the current set Sithat will be completed to a
saturating set, Didenotes the points of the plane outside Siwhich are determined by Si
and Ridenotes the set of points not in SiDi. For a point set H,σ(H)denotes the set
of lines skew to H.
The benefit of a point b(P)(in step i) is the amount of points from Riwhich would become
determined by the point set Si∪ {P}, that is, b(P) = |hP, Sii ∩ Ri|.
To obtain Di+1 from Di, we would like to add a point to Siwhich has the largest benefit.
Consider all the lines skew to Si, and choose one of them for which the intersection with
Riis minimal. Adding up the benefits of the points of this line `, we get
X
P`
b(P) = |Ri`|+i· |Ri\`|.(3.1)
Indeed, the double counting counts any point PRi`exactly once as these points
become determined only if we choose Pitself from `. On the other hand, any point Q
5
of Rioutside `will be determined by adding a point from `hQ, Sii. This latter point
set is of cardinality ias Qwas not determined before.
Lemma 3.2.
min
`σ(Si)|Ri`| ≤ |Ri|
q.
Proof. There are at least (q2+q+ 1) iq 1 = q(q+ 1 i) skew lines to any set of i
points. Every point of Riis appearing on exactly (q+ 1 i) lines from σ(Si), hence
X
`σ(Si)|Ri`|=|Ri|(q+ 1 i),
thus the statement follows.
Proposition 3.3.
|Ri+1|≤|Ri|1i
q+ 2
if we add the point having the largest benefit to Si.
Proof. We have b(P)1
q+1 (|Ri`|+i· |Ri\`|) for the point Phaving the largest
benefit on the skew line `of minimal intersection with Ri, in view of Equation 3.1.
Observe that |Ri`| ≤ |Ri|
q, according to Lemma 3.2. By adding this point Pto Si, we
get
|Ri+1|=|Ri|−b(P)≤ |Ri|− |Ri|
q+i(|Ri| − |Ri|
q)
q+ 1 ≤ |Ri|1i(q1)
q(q+ 1)<|Ri|1i
q+ 2.
Lemma 3.4.
k:=d3qln qe
Y
i=1 1i
q+ 2< q3/2.
Proof. Denote Qk
i=1 1i
q+2 =(q+1)!
(q+2)k(qk+1)! by Aq(k). We apply Stirling’s approxima-
tion
n!
t!<2πn n
en
2πt t
et
for t<nwith t=qk+ 1 and n=q+ 1. This implies
Aq(k)<p(q+ 1)
p(qk+ 1) ·(q+ 1)q+1
(q+ 2)k·(qk+ 1)qk+1 ·ek.
One again we use the Taylor-form for the approximation (1+1/x)x+0.5< e·1 + 1
12x2(x > 1),
6
to obtain
1 + k
qk+ 1qk+1
=1 + k
qk+ 1(qk+1
k+1/2)k1 + k
qk+ 1k/2
<
ek·1 + k2
12(qk+ 1)2kqk+ 1
q+ 1 k/2
.
(3.2)
Applying the approximation also for (1 1/x)x0.5< e1·11
12x2(x→ ∞), we can
simplify
qk+ 1
q+ 1 k/2q+ 1
q+ 2k
=1(q+ 2)(k+ 2) + 1
(q+ 2)2k/2
<1(k+ 2)
(q+ 2) k/2
=
1(k+ 2)
(q+ 2) (q+2
k+2 0.5)·k/2·(q+2
k+2 0.5)1
<e1·1(k+ 2)2
12(q+ 2)2k(k+2)
2qk+2
.
(3.3)
In total, if we take into consideration both (3.2) and (3.3), we get that
Aq(k)< exp k(k+ 2)
2qk+ 23(k),
where ∆3(k) is the product of error terms (q+1)
(qk+1) ,1 + k2
12(qk+1)2kand
1(k+2)2
12(q+2)2k(k+2)
2qk+2 .
If one plugs in k=d3qln qe, careful calculations on the error term imply that
Aq(k)< e((1+)3
2ln q),
where =(q)>0.
By applying Proposition 3.3 successively k=d3qln qetimes it follows from Lemma 3.4
that we end up with at most q+ 1 points remaining in Rk. But note that these points
can be covered by adding a further d(q+ 1)/2epoints to the saturating set, via Lemma
2.1, hence this algorithm provides a saturating set of size d3qln qe+d(q+ 1)/2e.
4 Connections with hypergraph coverings, applica-
tions and open problems
To improve further the bound of Theorem 1.6 in the main term, one must ensure at least
one of the followings.
7
(1) Prove the existence of a point that has significantly more benefit then the others,
for several steps.
(2) Prove a lemma, which provides a completion of almost saturating sets to obtain a
saturating set with fewer points than Lemma 2.1. That might enable us to apply
Lemma 3.4 until a smaller summation limit.
Concerning (1), observe that while avoiding lines `which are secants of Siduring the choice
for the newly added point, it is not guaranteed that the new point won’t lie eventually
on a secant of Si, so this method does not enable us to provide arcs. However, as these
tangents does not contain points for Ri, the sum of the benefits on these tangent lines
are certainly less than the sum determined in Equation 3.1. Despite all this, the known
constructions with O(q) points on Galois planes of square order consist of point sets
having large secant sizes — but note that we heavily built on the algebraic (sub)structure.
Concerning (2), notice first that the constant 3 in the bound (and also in the background
of the first probabilistic proof) is explained by the application domain of the approaches,
as the first phase of both methods lasted until the number of not saturated points decreases
from q2to O(q), so if one cannot exploit the structure of the not saturated points hence
cannot obtain a good bound on the variance of the distribution of the benefit values, then
Lemma 3.4 would provide an evidence that O(q) points won’t form a saturating set in
general.
This problem is strongly connected to the theory of bounding the transversal number of
certain hypergraphs. Recall that for an r-uniform hypergraph H, the covering number or
transversal number τ(H) is the the minimum cardinality of a set of vertices that intersects
all edges of H.
Consider a point set S0in the projective plane. Let X(S0) = {x1, . . . , xm}be the set
of points not saturated by S0, that is, those points which are not incident to the lines
determined by the point pairs of S0. Assign to each xia set Hiof |S0|(q1) + 1 points
which make xisaturated, i.e. Hi=hxi, S0i \ S0.
Note that the intersection of these sets consists of many points.
Lemma 4.1. |HiHj|=|S0|(|S0| − 1) or |HiHj|= (|S0| − 1)(|S0| − 2) + qfor every
1i<jm.
Proof. Suppose first that hxi, xji ∩ S0=. Then for every z6=z0S0, the points
hxi, zi ∩ hxj, z0iare distinct, hxi, zi ∩ hxj, z0i ∈ HiHj, and they are points outside S0,
hence |HiHj|=|S0|(|S0| − 1).
In the other case when |hxi, xji ∩ S0|= 1, a point zS0is incident to hxi, xji. Now
for every z6=z0S0with z6=z6=z0, the points hxi, zi ∩ hxj, z0i ∈ HiHjare distinct
again and not belonging to S0, while hxi, zi=hxj, zithus hxi, zi \ zalso belongs to
the intersection, hence the claim follows.
8
Lemma 2.1 may thus be altered to a much more general lemma concerning the transversal
number of t-intersecting uniform hypergraphs, applied to the sets {Hi=hxi, S0i\S0}. So
an (approximate) solution for the open problem below would imply stronger results for
the saturation problem as well.
Problem 4.2. Given F={Hi:i= 1 . . . m}an r-uniform t-intersecting set system on
an n-element ground set, prove sharp upper bounds on τ(F)in terms of n, m,tand r.
It is well known that while the trivial bound τ(F)rt+ 1 can be sharp for dense
hypergraphs, the approximation of the transversal number is hard in general. Due to
Lov´asz [18], a connection is made with the fractional transversal number τ(F) (which is
also called fractional covering number) as τ < (1 + ln d)τ, where dis the maximal degree
in the hypergraph.
Note that this result yields instantly the bound τ < (1 + ln d)rm
r+(m1)t.
By exploiting the intersection property, the greedy algorithm yields that there exists a
point in any Hithat is contained in at least 1 + tm
rsets from F. Hence the following
proposition follows from the recursion:
Proposition 4.3. Let F={Hi:i= 1 . . . m}be an r-uniform t-intersecting set system
on an n-element ground set. Then τ(F)≤ d rm
tm+rln me.
For further results, we refer to the survey of F¨uredi [12] and the paper of Alon, Kalai and
Matouˇsek and Meshulam [1]. The investigation of the transversal number of t-intersecting
hypergraphs was proposed also by F¨uredi in [11].
Multiple saturating sets (see e.g. in [3]) and their generalizations in higher dimensional
spaces are also investigated. A point set Sin PG(n, q) is saturating if any point of
PG(n, q)\Sis collinear with two points in S. The two proof techniques presented in
Sections 2 and 3 are applicable in these more general settings as well. Namely, the natural
analogue of the Lunelli-Sce bound provides the following lower bound for a saturating set
Sin PG(n, q):
Proposition 4.4. |S| ≥ 2qn1
2.
The direct analogue of the first (probabilistic) approach shows the existence of a saturating
set in PG(n, q) of size |S| ≤ (1+o(1))p(n+ 1)qn1ln q, which improves previously known
bounds for n= 4.
Acknowledgement
Grateful acknowledgement is due to the anonymous referees for their helpful suggestions
in order to improve the presentation of the paper.
References
[1] Alon, N., Kalai, G., Matouˇ
sek, J., Meshulam, R., Transversal numbers for hyper-
graphs arising in geometry. Advances in Applied Mathematics 29(1), (2002) 79–101.
9
[2] Ball, S., Finite Geometry and combinatorial applications, (Vol. 82), Cambridge University
Press 2015.
[3] Bartoli, D., Davydov, A. A., Giulietti, M., Marcugini, S., Pambianco, F.,
Upper bounds on the smallest size of a saturating set in a projective plane. arXiv preprint
arXiv:1505.01426 (2016).
[4] Bartoli, D., Davydov, A. A., Faina, G., Kreshchuk, A. A., Marcugini, S., Pam-
bianco, F., Upper bounds on the smallest size of a complete arc in PG(2, q) under a certain
probabilistic conjecture. Problems of Information Transmission, 50(4), (2014) 320–339.
[5] Bartoli, D., Davydov, A. A., Faina, G., Kreshchuk, A. A., Marcugini, S., Pam-
bianco, F., Upper bounds on the smallest size of a complete arc in a finite Desarguesian
projective plane based on computer search. J. Geometry, 107(1), (2016) 89–117.
[6] Boros, T., T. Sz˝
onyi, T., Tichler, K., On defining sets for projective planes. Discrete
Mathematics 303(1), (2005) 17–31.
[7] Cohen, G., Honkala, I., Litsyn, S., Lobstein, A., Covering Codes, The Netherlands,
North-Holland:Amsterdam, 1997.
[8] Davydov, A. A., Giulietti, M., Pambianco, F., Marcugini, S., Linear nonbinary
covering codes and saturating sets in projective spaces. Advances in Mathematics of Com-
munications 5(1), (2011) 119–147.
[9] Davydov, A. A., Marcugini, S., Pambianco, F., On saturating sets in projective
spaces. J. Combinatorial Theory A, 103(1), (2003) 1–15.
[10] Davydov, A. A., Giulietti, M., Marcugini, S., Pambianco, F. Linear covering codes
over nonbinary finite fields. In Proc. XI Int. Workshop Algebraic Comb. Coding Theory,
ACCT2008 (2008) 70–75.
[11] F¨
uredi, Z., Intersecting designs from linear programming and graphs of diameter two.
Discrete Mathematics, 127(1-3), (1994) 187–207.
[12] F¨
uredi, Z., Matchings and covers in hypergraphs. Graphs and Combinatorics, 4(1), (1988)
115–206.
[13] Giulietti, M., Torres, F., On dense sets related to plane algebraic curves. Ars Combi-
natoria, 72, (2004) 33–40.
[14] Giulietti, M., The geometry of covering codes: small complete caps and saturating sets
in Galois spaces, Surveys in combinatorics, 2013.
[15] Grynkiewicz, D. J., Lev, V. F. 1-saturating sets, caps, and doubling-critical sets in
binary spaces. SIAM J. Discrete Mathematics, 24(1), (2010) 169–190.
[16] Kim, J. H., Vu, V. H. Small complete arcs in projective planes. Combinatorica, 23(2),
(2003) 311–363.
[17] Kiss, Gy., Kov´
acs, I., Kutnar, K., Ruff, J., Sparl, P. A Note on a geometric
construction of large Cayley graphs of given degree and diameter. Studia Universitatis
Babes-Bolyai, Mathematica, 3, (2009) 77–84.
[18] Lov´
asz, L., On the ratio of optimal integral and fractional covers. Discrete Mathematics,
13(4), (1975) 383–390.
[19] Ughi, E., Saturated configurations of points in projective Galois spaces. European J. Com-
bin. 8, (1987) 325–334.
10

Supplementary resource (1)

... For r = t R + 1, R ≥ 2, t ≥ 1, in [2,[11][12][13]32], see also the references therein, upper bounds of the following forms are obtained: ...
Article
Full-text available
The length function ℓq(r,R)q(r,R)\ell _q(r,R) is the smallest possible length n of a q-ary linear [n,n-r]qR[n,nr]qR[n,n-r]_qR code with codimension (redundancy) r and covering radius R. Let sq(N,ρ)sq(N,ρ)s_q(N,\rho ) be the smallest size of a ρρ\rho -saturating set in the projective space PG(N,q)PG(N,q)\textrm{PG}(N,q). There is a one-to-one correspondence between [n,n-r]qR[n,nr]qR[n,n-r]_qR codes and (R-1)(R1)(R-1)-saturating n-sets in PG(r-1,q)PG(r1,q)\textrm{PG}(r-1,q) that implies ℓq(r,R)=sq(r-1,R-1)q(r,R)=sq(r1,R1)\ell _q(r,R)=s_q(r-1,R-1). In this work, for R≥3R3R\ge 3, new asymptotic upper bounds on ℓq(tR+1,R)q(tR+1,R)\ell _q(tR+1,R) are obtained in the following form: ∙ℓq(tR+1,R)=sq(tR,R-1)≤R!RR-2R·q(r-R)/R·lnqR+o(q(r-R)/R),r=tR+1,t≥1,qis an arbitrary prime power,qis large enough;∙if additionallyRis large enough, thenR!RR-2R∼1e≈0.3679.\begin{aligned}&\bullet ~\ell _q(tR+1,R) =s_q(tR,R-1)\\&\hspace{0.4cm} \le \root R \of {\frac{R!}{R^{R-2}}}\cdot q^{(r-R)/R}\cdot \root R \of {\ln q}+o(q^{(r-R)/R}), \hspace{0.3cm} r=tR+1,~t\ge 1,\\&\hspace{0.4cm}~ q\text { is an arbitrary prime power},~q\text { is large enough};\\&\bullet ~\text { if additionally }R\text { is large enough, then }\root R \of {\frac{R!}{R^{R-2}}}\thicksim \frac{1}{e}\thickapprox 0.3679. \end{aligned}The new bounds are essentially better than the known ones. For t=1t=1, a new construction of (R-1)(R1)(R-1)-saturating sets in the projective space PG(R,q)PG(R,q)\textrm{PG}(R,q), providing sets of small sizes, is proposed. The [n,n-(R+1)]qR[n,n(R+1)]qR[n,n-(R+1)]_qR codes, obtained by the construction, have minimum distance R+1R+1R + 1, i.e. they are almost MDS (AMDS) codes. These codes are taken as the starting ones in the lift-constructions (so-called “qmqmq^m-concatenating constructions”) for covering codes to obtain infinite families of codes with growing codimension r=tR+1r=tR+1, t≥1t1t\ge 1.
... To determine this value is a very difficult problem in general and indeed a classical problem in coding theory is the covering problem, that is the determination of the smallest the minimal size of a ρ-covering code of length n. This problem has been deeply investigated during the last years, see [12,16,19,29] for some references. One of the main tool used to study the covering problem has been the geometric interpretation of the covering radius. ...
Article
Bonini, Borello and Byrne started the study of saturating linear sets in Desarguesian projective spaces, in connection with the covering problem in the rank metric. In this paper we study 1-saturating linear sets in PG(2, q 4), that is F q-linear sets in PG(2, q 4) with the property that their secant lines cover the entire plane. By making use of a characterization of generalized Gabidulin codes, we prove that the rank of such a linear set is at least 5. This answers to a recent question posed by Bartoli, Borello and Marino.
... Proof. The proof follows the probabilistic argument of [35] of the second author, on the size of saturating sets of projective planes. Let H 0 be a random subset of G consisting of elements g ∈ G where each element is chosen independently, uniformly at random with probability p. ...
Preprint
Full-text available
Let G denote an abelian group, written additively. A 3-term arithmetic progression of G, 3-AP for short, is a set of three distinct elements of G of the form g, g + d, g + 2d, for some g, d ∈ G. a 3-AP of G consisting of x and two elements of S, we call S saturating w.r.t. 3-APs. A classical problem in additive combinatorics is to determine the maximal size of a 3-AP free set. A classical problem in finite geometry and coding theory is to find the minimal size of a complete cap in affine spaces, where a cap is a point set without a collinear triplet and in F n 3 caps and 3-AP free sets are the same objects. In this paper we determine the minimum size of complete 3-AP free sets and 3-AP saturating sets for several families of vector spaces and cyclic groups, up to a small multiplicative constant factor.
... This is equivalent, in the coding-theoretical language, to the so-called covering problem (see [12,Chapter 1]). Many bounds and construction of small saturating sets has been given over the last two decades (see for example [15,17,28] and references therein). ...
Preprint
Full-text available
Saturating sets are combinatorial objects in projective spaces over finite fields that have been intensively investigated in the last three decades. They are related to the so-called covering problem of codes in the Hamming metric. In this paper, we consider the recently introduced linear version of such sets, which is, in turn, related to the covering problem in the rank metric. The main questions in this context are how small the rank of a saturating linear set can be and how to construct saturating linear sets of small rank. Recently, Bonini, Borello, and Byrne provided a lower bound on the rank of saturating linear sets in a given projective space, which is shown to be tight in some cases. In this paper, we provide construction of saturating linear sets meeting the lower bound and we develop a link between the saturating property and the scatteredness of linear sets. The last part of the paper is devoted to show some parameters for which the bound is not tight.
... Here we only mention the paper of Gács and Szőnyi [23] on various applications, the celebrated paper on complete arcs of Kim and Vu [31] which applies Rödl's nibble, and the paper of second author [35] on the topic of saturating sets of projective planes which we revisit later on. ...
Article
Full-text available
Minimal linear codes are in one-to-one correspondence with special types of blocking sets of projective spaces over a finite field, which are called strong or cutting blocking sets. Minimal linear codes have been studied since decades but their tight connection with cutting blocking sets of finite projective spaces was unfolded only in the past few years, and it has not been fully exploited yet. In this paper we apply finite geometric and probabilistic arguments to contribute to the field of minimal codes. We prove an upper bound on the minimal length of minimal codes of dimension k over the q-element Galois field which is linear in both q and k, hence improve the previous superlinear bounds. This result determines the minimal length up to a small constant factor. We also improve the lower and upper bounds on the size of so called higgledy-piggledy line sets in projective spaces and apply these results to present improved bounds on the size of covering codes and saturating sets in projective spaces as well.
... For an introduction to the projective geometry over finite fields and its connection with coding theory, see [17, 25, 30, 34-37, 41, 45] and the references therein. Note also that in the papers [3][4][5][6][14][15][16][17][18][19][20][21][22][23][24]30,31,33,49,52], distinct aspects of covering codes and saturating sets, including their joint investigations, are considered. ...
Preprint
Full-text available
The length function q(r,R)\ell_q(r,R) is the smallest length of a q-ary linear code with codimension (redundancy) r and covering radius R. In this work, new upper bounds on q(tR+1,R)\ell_q(tR+1,R) are obtained in the following forms: \begin{equation*} \begin{split} &(a)~\ell_q(r,R)\le cq^{(r-R)/R}\cdot\sqrt[R]{\ln q},~ R\ge3,~r=tR+1,~t\ge1,\\ &\phantom{(a)~} q\text{ is an arbitrary prime power},~c\text{ is independent of }q. \end{split} \end{equation*} \begin{equation*} \begin{split} &(b)~\ell_q(r,R)< 3.43Rq^{(r-R)/R}\cdot\sqrt[R]{\ln q},~ R\ge3,~r=tR+1,~t\ge1,\\ &\phantom{(b)~} q\text{ is an arbitrary prime power},~q\text{ is large enough}. \end{split} \end{equation*} In the literature, for q=(q)Rq=(q')^R with qq' a prime power, smaller upper bounds are known; however, when q is an arbitrary prime power, the bounds of this paper are better than the known ones.
Article
The smallest possible length of a q-ary linear code of covering radius R and codimension (redundancy) r is called the length function and is denoted by q(r,R)\ell_q(r,R). In this work, for q \emph{an arbitrary prime power}, we obtain the following new constructive upper bounds on q(3t+1,3)\ell_q(3t+1,3): \begin{equation*} \begin{split} &\bullet~\ell_q(r,3)\lessapprox \sqrt[3]{k}\cdot q^{(r-3)/3}\cdot\sqrt[3]{\ln q},~r=3t+1, ~t\ge1,~ q\ge\lceil\W(k)\rceil,\\ &\phantom{\bullet~}18 <k\le20.339,~\W(k)\text{ is a decreasing function of }k ;\\ &\bullet~\ell_q(r,3)\lessapprox \sqrt[3]{18}\cdot q^{(r-3)/3}\cdot\sqrt[3]{\ln q},~r=3t+1,~t\ge1,~ q\text{ large enough}. \end{split} \end{equation*} For t=1t = 1, we use a one-to-one correspondence between codes of covering radius 3 and codimension 4, and 2-saturating sets in the projective space PG(3,q)\mathrm{PG}(3,q). A new construction providing sets of small size is proposed. The codes, obtained by geometrical methods, are taken as the starting ones in the lift-constructions (so-called ``qmq^m-concatenating constructions'') to obtain infinite families of codes with radius 3 and growing codimension r=3t+1r = 3t + 1, t1t\ge1. The new bounds are essentially better than the known ones.
Preprint
The smallest possible length of a q-ary linear code of covering radius R and codimension (redundancy) r is called the length function and is denoted by q(r,R)\ell_q(r,R). In this work, for q \emph{an arbitrary prime power}, we obtain the following new constructive upper bounds on q(3t+1,3)\ell_q(3t+1,3): q(r,3)k3q(r3)/3lnq3, r=3t+1, t1, qW(k),18<k20.339, W(k) is a decreasing function of k;\ell_q(r,3)\lessapprox \sqrt[3]{k}\cdot q^{(r-3)/3}\cdot\sqrt[3]{\ln q},~r=3t+1, ~t\ge1, ~ q\ge\lceil\mathcal{W}(k)\rceil, 18 <k\le20.339,~\mathcal{W}(k)\text{ is a decreasing function of }k ; q(r,3)183q(r3)/3lnq3, r=3t+1, t1, q large enough.\ell_q(r,3)\lessapprox \sqrt[3]{18}\cdot q^{(r-3)/3}\cdot\sqrt[3]{\ln q},~r=3t+1,~t\ge1,~ q\text{ large enough}. For t=1t = 1, we use a one-to-one correspondence between codes of covering radius 3 and codimension 4, and 2-saturating sets in the projective space PG(3,q)\mathrm{PG}(3,q). A new construction providing sets of small size is proposed. The codes, obtained by geometrical methods, are taken as the starting ones in the lift-constructions (so-called ``qmq^m-concatenating constructions'') to obtain infinite families of codes with radius 3 and growing codimension r=3t+1r = 3t + 1, t1t\ge1. The new bounds are essentially better than the known ones.
Article
Full-text available
The length function q(r,R)\ell_q(r,R) is the smallest length of a q -ary linear code with codimension (redundancy) r and covering radius R. In this work, new upper bounds on q(tR+1,R)\ell_q(tR+1,R) are obtained in the following forms: \begin{equation*} \begin{split} &(a)~\ell_q(r,R)\le cq^{(r-R)/R}\cdot\sqrt[R]{\ln q},~ R\ge3,~r=tR+1,~t\ge1,\\ &\phantom{(a)~} q\text{ is an arbitrary prime power},~c\text{ is independent of }q. \end{split} \end{equation*} \begin{equation*} \begin{split} &(b)~\ell_q(r,R)< 3.43Rq^{(r-R)/R}\cdot\sqrt[R]{\ln q},~ R\ge3,~r=tR+1,~t\ge1,\\ &\phantom{(b)~} q\text{ is an arbitrary prime power},~q\text{ is large enough}. \end{split} \end{equation*} In the literature, for q=(q)Rq=(q')^R with qq' a prime power, smaller upper bounds are known; however, when q is an arbitrary prime power, the bounds of this paper are better than the known ones. For t=1, we use a one-to-one correspondence between [n,n(R+1)]qR[n,n-(R+1)]_qR codes and (R1)(R-1)-saturating n-sets in the projective space PG(R,q)\mathrm{PG}(R,q). A new construction of such saturating sets providing sets of small size is proposed. Then the [n,n(R+1)]qR[n, n - (R+1)]_qR codes, obtained by geometrical methods, are taken as the starting ones in the lift-constructions (so-called ``qmq^m-concatenating constructions'') for covering codes to obtain infinite families of codes with growing codimension r=tR+1, t1t\ge1.
Conference Paper
Full-text available
In a projective plane Π q (not necessarily Desar-guesian) of order q, a point subset S is saturating (or dense) if any point of Π q \S is collinear with two points in S. Using probabilistic methods, more general than those previously used for saturating sets, the following upper bound on the smallest size s(2, q) of a saturating set in Π q is proved: s(2, q) 1 a random point set of size k in Π q with 2c√(q + 1) ln(q + 1) + 2 2 −1/q+2 ∼ q is a saturating set with probability greater than 1 − 1/(q + 1) 2c2 −2 . Our probabilistic approach is also applied to multiple saturating sets. A point set S ⊂ Π q is (1,μ)-saturating if for every point Q of Π q \S the number of secants of S through Q is at least μ, counted with multiplicity. The multiplicity of a secant l is computed as ( #(l∩S) 2 ). The following upper bound on the smallest size s μ (2, q) of a (1,μ)-saturating set in Π q is proved: s μ (2, q) < 2(μ + 1)√(q + 1)ln(q + 1) + 2 for 2 < μ < √q. By using inductive constructions, upper bounds on the smallest size of a saturating set (as well as on a (1, μ)-saturating set) in the projective space PG(N, q) are obtained. All the results are also stated in terms of linear covering codes.
Article
Full-text available
In the projective plane PG(2, q), upper bounds on the smallest size t 2(2, q) of a complete arc are considered. For a wide region of values of q, the results of computer search obtained and collected in the previous works of the authors and in the present paper are investigated. For q ≤ 301813, the search is complete in the sense that all prime powers are considered. This proves new upper bounds on t 2(2, q) valid in this region, in particular t2(2,q)    <0.9983qlnq    for        7q160001;t2(2,q)    <1.053qlnqfor        7q301813;t2(2,q)    <qln0.7295qfor    109q160001;t2(2,q)    <qln0.7404qfor    160001<q301813.\begin{array}{ll}t_{2}(2, q) \;\; < 0.998 \sqrt{3q {\rm ln}\,q} \quad\;\;\,\,{\rm for} \;\; \qquad \quad \;\;7 \leq q \leq 160001;\\ t_{2}(2, q) \;\; < 1.05 \sqrt{3q {\rm ln}\, q}\qquad\,\,{\rm for}\;\; \qquad \quad \;\;7 \leq q \leq 301813;\\ t_{2}(2,q)\;\; < \sqrt{q}{\rm ln}^{0.7295}\,q \qquad \,\,\,\,\,{\rm for} \;\; \quad \quad \,\,\,109 \leq q \leq 160001;\\ t_{2}(2,q) \;\; < \sqrt{q}{\rm ln}^{0.7404}\,q \qquad \,\,\,\,\,{\rm for }\;\;\, \quad 160001 < q \leq 301813.\end{array}The new upper bounds are obtained by finding new small complete arcs in PG(2,q) with the help of a computer search using randomized greedy algorithms and algorithms with fixed (lexicographical) order of points (FOP). Also, a number of sporadic q’s with q ≤ 430007 is considered. Our investigations and results allow to conjecture that the 2-nd and 3-rd bounds above hold for all q ≥ 109. Finally, random complete arcs in PG(2, q), q ≤ 46337, q prime, are considered. The random complete arcs and complete arcs obtained by the algorithm FOP have the same region of sizes; this says on the common nature of these arcs.
Article
Full-text available
The following upper bound on the smallest size s(2,q) of a saturating set in a projective plane Πq\Pi _{q} (not necessary Desarguesian) of order q is proven. \begin{equation*} s(2,q)\leq 2\sqrt{(q+1)\ln q}+2\thicksim 2\sqrt{q\ln q}. \end{equation*} Our approach is probabilistic rather than geometric. The proof uses some known results on the classical Birthday problem, and also provides the following result. In Πq\Pi _{q} any point k-set, with k2c(q+1)lnq+2 k\leq 2c\sqrt{(q+1)\ln q}+2 and an universal constant c1,c\geq 1, is a saturating set with probability \begin{equation*} p>1-\frac{1}{q^{2c^{2}-2}}. \end{equation*}
Article
Full-text available
In the projective plane PG(2,q), an iterative construction of complete arcs adding a new point on every step is considered. It is proven that uncovered points are uniformly distributed on the plane. For more than half of the steps of the iterative process, an estimation of the number of new covered points on every step is proven. A natural (and well-founded) conjecture that the estimation holds for the rest of steps is done. As a result, upper bounds on the smallest size t_{2}(2,q) of a complete arc in PG(2,q) are obtained, in particular, \begin{align*} & t_{2}(2,q)<\sqrt{q}\sqrt{3\ln q+\ln \ln q+\ln 3}+\sqrt{\frac{q}{3\ln q}}+3,\\ & t_{2}(2,q)<1.87\sqrt{q\ln q}. \end{align*} Non-standard types of upper bounds on t_{2}(2,q), including a new one, are considered. The effectiveness of the new bounds is illustrated by comparison with the smallest known sizes of complete arcs obtained in the recent works of the authors and in the present paper via computer search for a wide region of q. A connection of the problems considered with the Birthday problem (or the Birthday paradox) is noted.
Article
Full-text available
An infinite series and some sporadic examples of large Cayley graphs with given degree and diameter are constructed. The graphs arise from arcs, caps and other objects of finite projective spaces.
Article
Full-text available
In the late 1950's, B. Segre introduced the fundamental notion of arcs and complete arcs [?, ?]. An arc in a finite projective plane is a set of points with no three on a line and it is complete if cannot be extended without violating this property. Given a projective plane P, determining n(P), the size of its smallest complete arc, has been a major open question in finite geometry for several decades. Assume that P has order q, it was shown by Lunelli and Sce [?], more than 40 years ago, that n(P) ≥ √ 2q. Apart from this bound, practically nothing was known about n(P), except for the case P is the Galois plane. For this case, the best upper bound, prior to this paper, was O(q 3/4) obtained by Sz˝ onyi using the properties of the Galois field GF (q). In this paper, we prove that n(P) ≤ √ q log c q for any projective plane P of order q, where c is a universal constant. Together with Lunelli-Sce's lower bound, our result determines n(P) up to a polylogarithmic factor. Our proof uses a probabilistic method known as the dynamic random contruction or Rödl's nibble. The proof also gives a quick randomized algorithm which produces a small complete arc with high probability. The key ingredient of our proof is a new concentration result, which applies for non-Lipschitz functions and is of independent interest.
Book
The projective and polar geometries that arise from a vector space over a finite field are particularly useful in the construction of combinatorial objects, such as latin squares, designs, codes and graphs. This book provides an introduction to these geometries and their many applications to other areas of combinatorics. Coverage includes a detailed treatment of the forbidden subgraph problem from a geometrical point of view, and a chapter on maximum distance separable codes, which includes a proof that such codes over prime fields are short. The author also provides more than 100 exercises (complete with detailed solutions), which show the diversity of applications of finite fields and their geometries. Finite Geometry and Combinatorial Applications is ideal for anyone, from a third-year undergraduate to a researcher, who wishes to familiarise themselves with and gain an appreciation of finite geometry.
Article
Almost all combinatorial question can be reformulated as either a matching or a covering problem of a hypergraph. In this paper we survey some of the important results.