Abstract 1 Introduction 2 Bottomless Rectangles and Integer Points 3 Disks in the Plane: Separated Setting 4 Disks of Bounded Radii: General Setting 5 Generalization to Positive Homothets of a Convex Body 6 Conclusions and Open Problems References

Online Hitting Sets for Disks of Bounded Radii

Minati De ORCID Dept. of Mathematics, Indian Institute of Technology Delhi, New Delhi, India Satyam Singh ORCID Department of Computer Science, Aalto University, Espoo, Finland Csaba D. Tóth ORCID Department of Mathematics, California State University Northridge, Los Angeles, CA, USA
Department of Computer Science, Tufts University, Medford, MA, USA
Abstract

We present algorithms for the online minimum hitting set problem in geometric range spaces: Given a set P of n points in the plane and a sequence of geometric objects that arrive one-by-one, we need to maintain a hitting set at all times. For disks of radii in the interval [1,M], we present an O(logMlogn)-competitive algorithm. This result generalizes from disks to positive homothets of any convex body in the plane with scaling factors in the interval [1,M]. As a main technical tool, we reduce the problem to the online hitting set problem for a finite subset of integer points and bottomless rectangles. Specifically, for a given N>1, we present an O(logN)-competitive algorithm for the variant where P is a subset of an N×N section of the integer lattice, and the geometric objects are bottomless rectangles.

Keywords and phrases:
Geometric Hitting Set, Online Algorithm, Homothets, Disks
Funding:
Minati De: Research on this paper was supported by SERB MATRICS Grant MTR/2021/ 000584.
Satyam Singh: Research on this paper was supported by the Research Council of Finland, Grant 363444.
Csaba D. Tóth: Research on this paper was supported, in part, by the NSF award DMS-2154347.
Copyright and License:
[Uncaptioned image] © Minati De, Satyam Singh, and Csaba D. Tóth; licensed under Creative Commons License CC-BY 4.0
2012 ACM Subject Classification:
Theory of computation Computational geometry
; Theory of computation Online algorithms
Related Version:
Full Version: https://arxiv.org/abs/2412.04646
Editors:
Anne Benoit, Haim Kaplan, Sebastian Wild, and Grzegorz Herman

1 Introduction

In the general form of the Hitting Set problem, we are given a point set P and a collection of subsets 𝒞={S1,,Sm}, and we need to find a subset HP (hitting set) of minimal size such that every set Si𝒞 contains some point in H. In the Online Hitting Set problem, the set P is known in advance, but the subsets S1,S2, in 𝒞 arrive one at a time (without advance knowledge). We need to maintain a hitting set HiP for the first i sets {S1,,Si} such that HiHi+1 for all i1 (that is, we can add new points to the hitting set, but we cannot delete any point). The study of the Online Hitting Set problem (which is dual to the Online Set Cover problem) was initiated by Alon et al. [2]. They designed a deterministic algorithm with competitive ratio O(log|P|log|𝒞|) and obtained almost matching lower bound of Ω(log|P|log|𝒞|loglog|P|+loglog|𝒞|).

Geometric Hitting Set.

In the geometric Hitting Set problem, we have Pd for some constant dimension d, and the sets in 𝒞 are geometric objects of some type: for example, balls, unit balls, simplices, axis-aligned cubes, or hyper-rectangles. Depending on whether P is finite or infinite, there are different versions of the problem. In this paper, we consider P to be a finite set of points in 2.

Related Previous Work.

When P is finite, Even and Smorodinsky [12] initiated the study of the geometric online Hitting Set problem for various geometric objects. They established an optimal competitive ratio of Θ(log|P|) when the objects are intervals in , or half-planes or congruent disks in the plane. Later, Khan et al. [14] investigated this problem for a finite set of integer points P[0,N)22 and a collection 𝒞 of axis-aligned squares S[0,N)2 with integer coordinates for N>0. They developed an O(logN)-competitive algorithm for this variant. They also established a randomized lower bound of Ω(log|P|), where P2 consists of finitely many points and 𝒞 consists of translates of an axis-aligned square. Recently, De et al. [6] considered the variant when P is set of n points in 2 and 𝒞 consists of homothetic copies of a regular k-gon (for k4) with scaling factors in the interval [1,M], and designed an O(k2logMlogn)-competitive randomized algorithm. Even though a disk can be approximated by a regular k-gon as k, this does not imply any competitive algorithm for disks with radii in the interval [1,M].

Our results and Technical Contribution.

We study the Online Hitting Set problem when P is set of n points in 2. Table 1 summarizes the existing results and the results of the current paper.

Table 1: Summary of known and new results for the geometric Online Hitting Set problem where |P|=n is finite. (#) indicates randomized results. Our results are listed in the last three lines.
Points Objects Lower Bound Upper Bound
P Intervals in Ω(logn) [12] O(logn) [12]
P2 Half-planes and translates of a disk in 2 Ω(logn) [12] O(logn) [12]
P[0,N)22 Axis-aligned squares in [0,N)22 with integer coordinates Ω(logn) [14] (#) O(logN) [14]
P Homothetic copies of a regular k-gon (k4) with scaling factors in the interval [1,M] Ω(logn) [14] (#) O(k2logMlogn) [6] (#)
P[0,N)22 Bottomless rectangles (for definition, see Section 2.1) Ω(logn)[12] O(logN)     [Theorem 1]
P2 Disks having radii in the interval [1,M] Ω(logn) [12] O(logMlogn) [Theorem 12]
P2 Positive homothets of an arbitrary convex body in 2 with scaling factors in the interval [1,M] Ω(logn) [14] (#) O(logMlogn) [Theorem 14]

We now present our contributions and briefly discuss the technical ideas involved.

Bottomless Rectangles in [𝟎,𝑵]𝟐.

We present an O(logN)-competitive deterministic algorithm for the geometric Online Hitting Set problem, where P[0,N)22, and 𝒞 is a sequence of bottomless rectangles of the form [a,b)×[0,c) arriving one by one (Theorem 1 in Section 2). When a bottomless rectangle [a,b)×[0,c) arrives, our algorithm chooses hitting points guided by the canonical partition of the interval [a,b] (see Section 2 for a definition). For each point p in an offline optimum, this structured canonical partition ensures that O(logN) points are sufficient to hit all the incoming rectangles in [0,N]22 that are hit by p. We prove that our algorithm is O(logN)-competitive for a broader class of objects– sets S[a,b)× with lowest-point property (see Section 2.2 for a definition).

Disks with Radii in [𝟏,𝑴].

Our main result is a deterministic O(logMlogn)-competitive Online Hitting Set algorithm for an arbitrary set P of n points in the plane, and a sequence of disks of radii in the interval [1,M] (Theorem 12 in Section 4). Previously, an O(logn)-competitive algorithm was known only for congruent disks [12]. In particular, our result is the first O(logn)-competitive algorithm that works for disks of radii in the interval [1,1+ε] for any constant ε>0 (Corollary 13).

However, a finite set of disks in the plane do not necessarily have the lowest-point property. We reduce the problem to objects with the lowest-point property in two steps. First, we consider a restricted version, the line-separated setting (Section 3), where the centers of disks in 𝒞 lie on one side of a line (w.l.o.g., the x- or y-axis), while P lies on the other side. We use the concept of disk hull for a point set (introduced by Dumitrescu et al. [10]), which generalizes the notion of convex hulls and α-hulls. Among other important properties, the boundary of the disk hull is monotone w.r.t. the separating line. Using these properties, we reduce the Hitting Set problem in the line-separated setting to objects with the lowest-point property, and obtain an O(logn)-competitive algorithm in the line-separated setting (Theorem 9 in Section 3).

In general, there is no restriction on the location of the points in P and the centers of disks. We reduce the general problem to the line-separated setting as follows: We partition the disks of radii in the interval [1,M] into O(logM) layers, ensuring that the ratio of radii of disks in each layer is bounded by at most 2. For each layer, our algorithm maintains a tiling of the plane into axis-aligned squares such that (a) any disk of a given layer contains the entire tile that contains the disk center, and (b) each disk intersects only O(1) tiles. Our algorithm simultaneously runs several invocations of the line-separating algorithm (one for each directed grid line). When a disk arrives, our algorithm inserts it into all relevant invocations of the line-separating algorithms; we show that only O(1) invocations are relevant. In the competitive analysis, we show that for each point p in an offline optimum solution, our algorithm uses O(logn) hitting points for the disks in each layer that contain p. Since there are O(logM) layers, our algorithm is O(logMlogn)-competitive.

Homothets of a Convex Body with Diameters in [𝟏,𝑴].

We generalize our main result from disks to positive homothets of any convex body in the plane, where the radii in the interval [1,M] are replaced by scaling factors in the interval [1,M] (Theorem 14 in Section 5). Our online algorithm is based on a two-stage approach, similar to the case of disks, and it is O(logMlogn)-competitive. The key technical difficulty arises from the geometric differences between a disk and a general convex body. It is easy to extend the concept of a disk hull to hulls for homothetic convex bodies. However, unlike for disks, the boundary of the hull is not necessarily x- or y-monotone: We show that it is monotone w.r.t. some carefully chosen directions. To generalize a layered decomposition of axis-parallel lines, we need two directions in which the hull is monotone, the two directions must be far apart (in the space of directions), to create a tiling with properties (a) and (b) above. We call a pair of directions satisfying these requirements a good pair of directions. We use a careful geometric argument, which heavily relies on convexity, a suitable affine transformation, and the variational method (i.e., the intermediate value theorem) to prove that every convex body in the plane admits a good pair of directions.

1.1 Further Related Work

When the point set P is infinite, one may further distinguish between the continuous setting where P=d (also known as the piercing problem) and the discrete setting where P is a discrete subset of d (for example, P=d).

Continuous Setting.

In the geometric setting, the duality between the Hitting Set problem and the Set Cover problem only holds when the objects are translates of a convex body [5, Theorem 2]. Hence the results obtained for the Set Cover problem for translates of a convex body also hold for the Hitting Set problem. Charikar et al. [3] studied the Online Set Cover problem for translates of a ball. They proposed an algorithm with a competitive ratio O(2ddlogd). They also proved Ω(logd/logloglogd) as the deterministic lower bound of the competitive ratio for this problem. Dumitrescu et al. [9] improved the bounds on the competitive ratio for translates of a ball, establishing an upper bound of O(1.321d) and a lower bound of Ω(d+1). For translates of a centrally symmetric convex body, they proved that the competitive ratio of every deterministic algorithm is at least I(s), where I(s) is the illumination number of the object s111The illumination number of an object s, denoted by I(s), is the minimum number of smaller homothetic copies of s (λs, where λ(0,1)) whose union contains s.. For translates of an axis-aligned hypercube in d, Dumitrescu and Tóth [11] proved that the competitive ratio of any deterministic algorithm for Online Set Cover is at least 2d. Later, De et al. [5] studied the Online Hitting Set problem for α-fat objects in d with diameters in [1,M] and designed a deterministic algorithm with competitive ratio O((2+2α)dlogM). For hitting axis-aligned homothetic hypercubes with side lengths in [1,M], they gave a deterministic algorithm with competitive ratio at most 3dlog2M+2d. They also proved a Ω(dlogM+2d) lower bound for the problem of hitting homothetic hypercubes in d, with side lengths in the interval [1,M].

Discrete Setting.

De and Singh [7, 8] studied a variant of this problem where P=d and 𝒞 consists of translates of a ball or an axis-aligned hypercube in d. For translates of an axis-aligned hypercube, they showed that there is a randomized algorithm with an expected competitive ratio of O(d2) and also proved that every deterministic algorithm has a competitive ratio of at least d+1. For translates of a ball in d, they proposed a deterministic algorithm having a competitive ratio of O(d4) and proved that every deterministic algorithm has a competitive ratio of at least d+1, for d3. Recently, Alefkhani et al. [1] considered the variant where P=(0,N)dd and 𝒞 is a family of α-fat objects in (0,N)d, for some constant α>0. They proposed a deterministic algorithm with a competitive ratio of at most (4α+1)2dlogN, and proved that the competitive ratio of every deterministic algorithm is Ω(logN1+logα). Very recently, De et al. [6] improved both the upper and lower bounds of Alefkhani et al. [1]. They considered the case where P=d and 𝒞 is a family of α-fat objects having diameters in [1,M], for some constant α>0. They proposed a deterministic algorithm with competitive ratio O((2α)dlogM), and established that the competitive ratio of any randomized algorithm is Ω(dlogM).

2 Bottomless Rectangles and Integer Points

We present an O(logN)-competitive algorithm for the Online Hitting Set problem where P is a subset of an N×N section of the integer lattice, and the objects are bottomless rectangles (Section 2.1); and then generalize the algorithm for the same point set but with objects that have the lowest-point property (Section 2.2).

2.1 Bottomless Rectangles

In this section we present an O(logN)-competitive algorithm for the Online Hitting Set problem where P is a subset of the integer lattice with nonnegative coordinates less than N, that is, P[0,N)22; and the objects are bottomless rectangles. Bottomless rectangles are of the form ri=[ai,bi)×[0,ci), where 0ai<biN and 0ciN. Note that there are only O(N3) combinatorially different rectangles w.r.t. P, so the general result by Alon et al. [2] gives an algorithm for the online hitting set with competitive ratio O(log2N). In this section, we present an O(logN)-competitive algorithm, which is the best possible (a matching lower bound follows from the lower bound for the Online Hitting Set problem for intervals in one-dimension [12]).

Preliminaries.

We need some preparation before we can present the algorithm. We may assume w.l.o.g. that N is a power of 2, and every bottomless rectangle ri=[ai,bi)×[0,ci) is given with integer parameters ai, bi, and ci. An interval I is canonical if it is of the form I=[q2j,(q+1)2j) for some integers q,j0. For a canonical interval I=[q2j,(q+1)2j), we also define the left neighbor L(I)=[(q1)2j,q2j) and the right neighbor R(I)=[(q+1)2j,(q+2)2j). For every canonical interval I, if (I×[0,N))P, then let p(I) denote a lowest-point in (I×[0,N))P (that is, a point with minimum y-coordinate; ties are broken arbitrarily). If (I×[0,N))P=, then p(I) is undefined.

For every interval [a,b) with nonnegative integer endpoints, we define a canonical partition, i.e., a partition of [a,b) into canonical intervals. This partition is standard – we walk through some of the technical details because we need them for our algorithm and its analysis. Let j0 be the largest integer such that q2j(a,b), for some q. (Note that q is unique. Indeed, suppose that q is not unique, say q2j,(q+1)2j(a,b). Since q or q+1 is even, then q/2 or (q+1)/2 is an integer. Now, we have q22j+1 or q+12 2j+1(a,b), which contradicts the maximality of j.) We call the integer s[a,b):=q2j the splitting point of [a,b). We can partition a given interval [a,b) into canonical intervals as follows. If [a,b) is not canonical, find its splitting point s=s[a,b), partition it into two intervals [a,b)=[a,s)[s,b), and recurse on [a,s) and [s,b). For example, the splitting point of interval [5,11) is 8, and its canonical partition is [5,11)=[5,6)[6,8)[8,10)[10,11); see Figure 1(a)) for an illustration.

Note also that in the canonical partition of [a,s) (resp., [s,b)), there is at most one interval of each size, where the possible sizes are powers of 2 between 1 and sa (resp., bs). Specifically, if I is in the canonical partition of [a,s), then its left neighbor L(I) is not contained in [a,b), consequently aL(I)¯, where L(I)¯ is the closure of L(I). Similarly, if I is in the canonical partition of [s,b), then bR(I).

Online algorithm ALG for bottomless rectangles.

We can now present our online algorithm. We maintain a hitting set HiP, which is initially empty: H0=. When the i-th bottomless rectangle ri=[ai,bi)×[0,ci) arrives, initialize Hi:=Hi1. If riHi, then do not add any new points to Hi. Otherwise, we may assume that riHi=. Compute the splitting point si of [ai,bi), and the canonical partitions 𝒜i and i of [ai,si) and [si,bi), respectively. If ([ai,si)×[0,ci))P, then find the largest canonical interval I𝒜i such that p(I)ri, and put Hi:=Hi{p(I)}. Similarly, if ([si,bi)×[0,ci))P, then find the largest interval Ii such that p(I)ri, and put Hi:=Hi{p(I)}. Overall, we add at most two new points to Hi in step i.

(a)
(b)
Figure 1: (a) When the ith bottomless rectangle ri=[5,11)×[0,ci) arrives, suppose that the hitting set Hi contains the orange points, and riHi=. The splitting point of [5,11) is 8, with canonical partitions 𝒜i=[5,6)[6,8) and i=[8,10)[10,11), respectively. Here, I1=[6,8)𝒜i and I2=[8,10)i are the largest canonical intervals in 𝒜i and i, respectively. The blue (resp., red) point is the lowest-point in I1×[0,N)P (resp., I2×[0,N)P). We add both points to Hi. (b) The black and red colored points denote the set P, while the red colored points denote the set S. The yellow strip denotes I×.

Competitive analysis.

We now prove that ALG is O(logN)-competitive.

Theorem 1.

For the Online Hitting Set problem for a point set P[0,N)22 and a sequence of bottomless rectangles, the online algorithm ALG has competitive ratio O(logN).

Proof.

Let 𝒞 be a sequence of bottomless rectangles. Let H and OPT be the hitting set returned by the online algorithm ALG and an (offline) minimum hitting set of 𝒞, respectively. For a point pOPT, let 𝒞p be the subsequence of bottomless rectangles that contain p. It is enough to show that for every pOPT, our algorithm adds O(logN) points to H in response to the objects in 𝒞p.

Let pOPT, with coordinates p=(px,py); and let r1,,rm be a sequence of bottomless rectangles in 𝒞p for which our algorithm adds new points to the hitting set. We show that m=O(logN). We can distinguish between two types of rectangles ri=[ai,bi)×[0,ci) depending on whether the x-coordinate px of p is on the left or right of the splitting point si of [ai,bi): namely, px<si or sipx. We analyze the two types separately (the two cases are analogous).

Assume w.l.o.g. that px<si for i=1,,m. This means that p[ai,si)×[0,ci), and so ALG adds the hitting point p(I) for exactly one interval I𝒜i. Suppose that the algorithm adds the hitting point p(I) for I𝒜i. Then I is the largest (hence rightmost) canonical interval in 𝒜i such that (I×[0,ci))P. Recall that aiL(I)¯, where L(I) is the left neighbor of the canonical interval I. This implies that pxL(I)I. That is, either I or its left neighbor L(I) contains px. Note that px is contained in logN canonical intervals (one for each possible size), and each of these canonical intervals has a unique right neighbor. Consequently, I is one of at most 2logN canonical intervals under the assumption that px<si for all i=1,,m. This proves that m4logN.

2.2 Objects with the lowest-point property

In this section, we generalize Theorem 1 to a broader class of objects. Similarly to Section 2.1, let P[0,N)22. For a set SP, the span of S, denoted span(S) is the smallest interval [a,b) with integer endpoints a,b such that S[a,b)×. An object SP has the lowest-point property if for every point s=(sx,sy) in S and every interval Ispan(S) that contains sx, the object S contains all points in P(I×) with the minimum y-coordinates. For an illustration of set S see Figure 1(b). Note, in particular, that every bottomless rectangle ri=[ai,bi)×[0,ci) has the lowest-point property: Indeed, if sxI[ai,bi), then I×[0,sy]ri.

Our online hitting set algorithm and its analysis readily generalize when the objects have the lowest-point property. Let 𝒞=(S1,,Sm) be a sequence of objects with the lowest-point property.

Online algorithm ALG𝟎 for objects with the lowest-point property.

We maintain a hitting set HiP, which is initially empty: H0=. When set Si arrives, initialize Hi:=Hi1. If SiHi, then do not add any new points to Hi. Suppose that SiHi=. Let [ai,bi)=span(Si). Compute the splitting point si of [ai,bi), and the canonical partitions 𝒜i and i of [ai,si) and [si,bi), respectively. If ([ai,si)×)SiP, then find the largest interval I𝒜i such that p(I)Si, and put Hi:=Hi{p(I)}. Similarly, if ([si,bi)×)SiP, then find the largest interval Ii such that p(I)Si, and put Hi:=Hi{p(I)}. Overall, we add at most two new points to Hi in step i.

Correctness and competitive analysis.

When ALG0 adds a points p(I) to Hi in step i, the lowest-point property ensures that p(I)Si. Therefore, ALG0 maintains that Hi is a hitting set for {S1,,Si}, proving the correctness of ALG0. We now show that ALG0 is O(logN)-competitive.

Theorem 2.

For the Online Hitting Set problem for a point set P[0,N]22 and a sequence 𝒞=(S1,,Sm) of objects with the lowest-point property, algorithm ALG0 has competitive ratio O(logN).

Proof.

Let 𝒞 be a sequence of objects with the lowest-point property. Let H and OPT be the hitting set returned by the online algorithm ALG0 and an (offline) minimum hitting set of 𝒞, respectively. For each pOPT, let 𝒞p the subsequence of sets in 𝒞 that contain p. It is enough to show that for every pOPT, our algorithm adds O(logN) points to H in response to the objects in 𝒞p.

Let pOPT, with coordinates p=(px,py); and let S1,,Sm be a sequence of sets in 𝒞p for which our algorithm adds new points to the hitting set. We show that m=O(logN). We can distinguish between two types of sets Si depending on whether the x-coordinate px of p is to the left or right of the splitting point si: namely, px<si or sipx. We analyze the two types separately (the two cases are analogous).

Assume w.l.o.g. that px<si for i=1,,m. This means that p[ai,si)×, and so ALG adds the hitting point p(I) for exactly one interval I𝒜i. Suppose that the algorithm adds the hitting point p(I) for I𝒜i. Then I is the largest (hence rightmost) canonical interval in 𝒜i such that (I×)P. Recall that aiL(I)¯, where L(I) is the left neighbor of the canonical interval I. This implies that pxL(I)I. That is, either I or its left neighbor L(I) contains px. Note that px is contained in logN canonical intervals (one for each possible size), and each of these canonical intervals has a unique right neighbor. Consequently, I is one of at most 2logN canonical intervals under the assumption that px<si for all i=1,,m. This proves that m4logN.

3 Disks in the Plane: Separated Setting

In this section, we consider the Online Hitting Set problem in the plane, where P is a finite set above the x-axis (given in advance); and 𝒞 consists of disks of arbitrary radii with centers located on or below the x-axis (arriving one-by-one). Note that the disks in 𝒞 do not necessarily have the lowest-point property; see Figure 2.

Figure 2: Disk S with center below the x-axis does not have the lowest-point property: We have sS, and the interval Ispan(S) contains sx, but S does not contain the point pP with minimum y-coordinate in the strip I×.

Disk hulls for a point set w.r.t. disks and its properties.

The unit disk hull of a point set was introduced by Dumitrescu et al. [10] as an analogue of the convex hull. Recall that the convex hull conv(P) of a point set P2 is the smallest convex set in the plane that contains P. Equivalently, it is the intersection of all closed half-planes that contain P; it can be computed by the classical “rotating calipers” algorithm, where we continuously rotate a line around P while P remains in one closed half-plane bounded by . Intuitively, we obtain the unit disk hull of P by rolling a unit disk, with center on or below the x-axis, around P. We generalize this notion to disks of any fixed radius t>0.

Figure 3: A point set P (red) and region hull3(P) (pink). The boundary hull3(P) is composed of horizontal lines and circular arcs.
Definition 3.

Let P2 be a finite set of points above the x-axis and let t>0. Let 𝒟t be the set of all disks of radius t with centers on or below the x-axis. Let Mt(P) be the union of all disks D𝒟t such that Pint(D)=. Now, we define the 𝐭-hull of P as hullt(P)=2int(Mt(P)). The boundary of hullt(P) is denoted by hullt(P); for an illustration see Figure 3.

Dumitrescu et al. [10, Lemma 4] proved that hullt(P) is x-monotone222A curve in the plane is x-monotone if every vertical line intersects it at most once. for any t>0, and established other properties, which were used by Conroy and Tóth [4], as well.

Lemma 4 (Dumitrescu et al. [10]).

For a finite set P2 above the x-axis and t>0, the following holds:

  1. 1.

    hullt(P) lies above the x-axis;

  2. 2.

    every vertical line intersects hullt(P) in one point, thus hullt(P) is an x-monotone curve;

  3. 3.

    for every disk D𝒟t, the intersection D(hullt(P)) is connected (possibly empty);

  4. 4.

    for every disk D𝒟t, if PD, then PD contains a point in hullt(P).

Figure 4: A point set P (red), hull3(P) (pink), and hull1(P) (light blue or pink). A disk D𝒟3 of radius 3 (dotted blue), where the intersection D(hull1(P)) has two components.

Since we consider the case of disks with bounded radii, for our purposes, we need to compare two disk hulls for the same point set P w.r.t. different radii; see Figure 4. We start with an easy observation.

Lemma 5.

Let γ1 and γ2 be circular arcs lying entirely above the x-axis, such that γ1 and γ2 are arcs of circles C1 and C2, resp., of radii r1 and r2, with centers on or below the x-axis.

  1. 1.

    Then both γ1 and γ2 are x-monotone and concave curves.

  2. 2.

    Furthermore, if points p1,p22 are contained in both γ1 and γ2, and r1<r2, then γ1 lies above γ2 (i.e., for every vertical line L that separates p1 and p2, point γ1L lies above point γ2L).

Proof.

(1) For every i{1,2}, the center of Ci is below the x-axis, and so the leftmost and rightmost points of C1 are also below the x-axis. The leftmost and rightmost points partition Ci into two halfcircles, one above the center and one below the center. Both halfcircles are x-monotone: The lower halfcircle is convex curve and the upper halfcircle is concave. Since γi lies entirely above the x-axis, it is contained in the upper halfcirlce, which is x-monotone and concave.
(2) The locus of centers of circles that contain both p1 and p2 is the orthogonal bisector of the line segment p1p2, that we denote by (p1p2). Note that p1p2 is not vertical (or else (p1p2) would be a horizontal line above the x-axis, and the centers of C1 and C2 would also be above the x-axis). As the center a circle containing p1 and p2 continuously moves from the center of C1 down to y=, the circular arc between p1 and p2 deforms continuously from γ1 to the line segment p1p2. Since γ1 concave, it lies above the segment p1p2. Since r1<r2, the arc γ2 lies between the arc γ1 and the segment p1 and p2. Consequently, then γ2 lies below γ1, as claimed.

Lemma 6.

For every finite set P2 above the x-axis, the following holds:

  1. 1.

    If 0<s<t, then for every disk D𝒟s of radius s, the intersection D(hullt(P)) is connected (possibly empty).

  2. 2.

    Suppose that pP lies on the curve hullt(P) for some t>0. Then there is a radius rp(0,t) such that p is also on hulls(P) for all s[rp,t], but p is below hulls(P) for all s[0,rp).

Proof.

(1) Let D𝒟s. Suppose, to the contrary, that the intersection D(hullt(P)) has two or more components. By Lemma 4(2), the x-coordinates of the components form disjoint intervals, and the components have a natural left-to-right ordering. Let q1 be the rightmost point in the first component, and let q2 be the leftmost point in the second component. Clearly q1,q2D. Let q be an arbitrary point in hull(A) between q1 and q2. Then q lies on the boundary of some disk D of radius t whose center is below the x-axis, and whose interior is disjoint from P. In particular, neither q1 nor q2 is in the interior of D. Since the center of D is below the x-axis, D contains two interior-disjoint circular arcs between q and the x-axis; and both arcs must cross D. We have found two intersection points p1,p2DD above the x-axis. Furthermore, between p1 and p2, the circular arc D lies above the circular arc D, contradicting Lemma 5(2). This completes the proof of Property 1.
(2) Consider a point pP that lies on the curve hullt(P) for some t>0. Then there exists a disk D𝒟t of radius t centered at some point c below the x-axis such that pD. Let c1 be the intersection point of the x-axis the line cp, and c2 the orthogonal projection of p to the x-axis. We describe two continuous motions, where the disk D continuously changes while p is in the circle D and there is no point in P in the interior of D: First, a central dilation from center p continuously moves D to a disk D1 centered at c1. Second, the center of D moves from c1 towards c2 continuously until its center reaches c2 or a point c3 where D contains both p and another point pP. Let rp be radius of D at that time. The continuous motion shows that phulls(P) for all s[rp,t], but it is not in hulls(P) for all s<rp.

Note that Lemma 6(1) is not symmetric for s<t: For a disk D𝒟t of radius t, the intersection D(hulls(P)) is not necessarily connected; see Figure 4 for an example.

Reduction.

We can reduce the Online Hitting Set problem for a finite set P2 and disks of bounded radii in the separated setting, to the Online Hitting Set problem for a finite subset of integer points and objects with the lowest-point property. We achieve the reduction in two steps:

  1. (1)

    We choose a subset QP of points that are relevant for a hitting set (Lemma 7); and

  2. (2)

    we map the points in P into a set of integer points P[0,n]22 (Lemma 8).

For a finite point set P in the plane above the x-axis, let Q=Q(P) be the set of points pP such that phullt(P) for some t>0.

Lemma 7.

For a finite point set P in the plane above the x-axis, Q=Q(P) has the following property: For every disk D centered below the x-axis, if DP, then DQ.

Proof.

Let D be a disk of radius t>0 centered below the x-axis. By Lemma 4(4), DP contains a point in hullt(P). By the definition of Q, this point is in Q.

We may assume that the points in P have distinct x-coordinates (if two or more points in P have the same x-coordinate, w.l.o.g. a minimum hitting set would contain only the point with the smallest y-coordinate). Sort P by increasing x-coordinates such that P={p0,,pn1}. For every point qQ, let t(q)>0 be the maximum radius such that qhullt(q)(P). Consider the set of radii T={t(q):qQ}. Sort the radii in T in decreasing order as t0>t1>>t|T|1. We can now define the function π:P[0,n)22. For every piQ, let π(pi)=(i,j) if and only if t(pi)=tj, that is, the first coordinate of π(pi) corresponds to the index i of pi (the x-order of all points in Q), and the second coordinate of π(pi) corresponds to index j of the radius tj=t(pi). For every piPQ, let π(pi)=(i,|T|); see Figure 5 for an illustration. Finally, let P=π(P)={π(pi):piP} and Q=π(Q)={π(pi):piQ}. Note that the points in PQ lie above all points in Q. Since π is injective, then it is a bijection between P and P. Note also that |T||Q||P|=n, consequently P[0,n]22.

Figure 5: Description of the function π. Left: hull1(P) is orange, hull2(P) is green, and hull3(P) is blue. Right: The grid points π(p0),,π(p9) corresponding to p0,,p9.
Lemma 8.

For a set P of n points in the plane above the x-axis and for every disk D centered below the x-axis, the set π(DP) has the lowest-point property.

Proof.

Let D be a disk centered below the x-axis. We rephrase the lowest-point property in terms of DP. Recall that the points in P are sorted by x-coordinates. Suppose that s=(sx,sy) is in DP and sxIspan(DP). Consider the point sets P(I):={p=(px,py)DP:pxI}. By Lemma 7, we know that DQ; let t be the largest radius in T such that Q(I)hullt(P). We need to show that D contains all points in P(I)hullt(P).

Let qleft and qrightP(I), resp., be the leftmost and rightmost points in P(I)Q; and let Lleft and Lright be the vertical lines through qleft and qright. By the definition of Q, we have qlefthullt(qleft)(P) and qrighthullt(qright)(P), and tmax{t(qleft),t(qright)} by the definition of t. Consequently, the intersection point :=Llefthullt(P) lies at or below qleft, the intersection point r:=Lrighthullt(P) lies at or below qright. Since qleft,qrightD, then D contains both and r. We know that hullt(P) is an x-monotone curve by Lemma 4(2), and Dhullt(P) is connected by Lemma 6(2). Since D contains both and r, then D contains the sub-curve of hullt(P) between r and . Since all points in P(I) are between the vertical lines Lleft and Lright, then D contains all points in P(I)hullt(P), as required.

Online algorithm for disks in the separated setting.

We can now complete the reduction.

Theorem 9.

For the Online Hitting Set problem for a set P2 of n points above the x-axis and disks centered below the x-axis, there is an O(logn)-competitive algorithm.

Proof.

We are given a set P2 of n points above the x-axis, and we receive a sequence 𝒞=(D1,,Dm) of disks centered on or below the x-axis in an online fashion. Let OPTP be a minimum hitting set for 𝒞.

Initially, we compute the set P[0,n]22 as defined above Lemma 8. When a disk Di arrives, we compute the set Si=π(DiP), which has the lowest-point property by Lemma 8. The bijection π maps OPT to a set OPT=π(OPT)P, where |OPT|=|OPT|. Here, OPT is a hitting set for the sets 𝒞=(S1,,Sm).

We run the online algorithm ALG0 described in Section 2.2 for the point set P and the sequence 𝒞 of sets. By Theorem 1, ALG returns a hitting set HP of size |OPT|O(logn). By Lemma 7, H=π1(H)P is a hitting set for 𝒞, and its size is bounded by |H|=|H||OPT|O(logn)=|OPT|O(logn), as required.

4 Disks of Bounded Radii: General Setting

In this section, we consider the Online Hitting Set problem, where P is a finite set (given in advance) in the plane; and the objects are disks with radii in the interval [1,M), where M>1 is a constant.

Distinguishing layers of disks, according to their radii.

We partition the disks of radii in the interval [1,M) into logM+1 layers as follows: for each j{0,1,,logM}, let layer Lj be the set of disks of radii in the interval [2j,2j+1). The index of each layer Lj is denoted by j.

Tiling of the plane for each layer index 𝐣.

For every j{0,1,,logM}, let Λj={α1𝐯1+α2𝐯2:(α1,α2) 2} be a two-dimensional lattice spanned by vectors 𝐯1=2j1/2𝐞1 and 𝐯2=2j1/2𝐞2, where 𝐞1=(1,0) and 𝐞2=(0,1) are the standard basis vectors. Let τj=[0,2j1/2]2 be a square of side length 2j1/2 with lower-left corner at the origin. Translates of τj (tiles), with translation vectors in the lattice Λj, form the tiling 𝒯j. Let j denote the set of axis-parallel lines spanned by the sides of the tiles in 𝒯j.

Figure 6: A section of the tiling 𝒯j, the tile τj of side length 2j1/2, and a disk D of radius 2j+2.

We observe two key properties of the construction of layers and the tilings.

Observation 10.

For every j{0,1,,logM}, if σLj and the center of σ is in a tile τ𝒯j, then τσ; see Figure 6.

Proof.

Since σLj, the radius of the disk σ is in at least 2j. The tile τ is a translate of τj=[0,2j1/2]2, and so its diameter is 22j1/2=2j. If the center c of σ is in S, then every pτ is within distance 2j from c, which implies that τσ.

Observation 11.

For every j{0,1,,logM}, every disk D of radius at most 2j+2 intersects at most 24 lines in j: at most 12 horizontal and 12 vertical lines.

Proof.

Let D be a disk of radius 2j+2; see Figure 6. The orthogonal projection of D to the x-axis (resp., y-axis) is an interval of length at most 2j+3. Since the distance between any two consecutive vertical (resp., horizontal) lines in j is 2j1/2, then D intersects at most 2j+3/2j1/2=27/2=12 horizontal and at most 12 vertical lines in j.

Subproblem for a directed line 𝐋.

For a directed line L, we denote by L and L+ the closed half-plane on the left and right of L, respectively. Given a directed line L and the input (P,𝒞) of the Online Hitting Set problem, where P is a set of points, and 𝒞 is a sequence of disks in the plane, we define a subproblem (PL,𝒞L) as follows: Let PL=PL, and let 𝒞L be the subsequence of disks σi𝒞 such that the center of σi is in L+ and σi contains at least one point in PL. Now for each subproblem (PL,𝒞L), we can run the online algorithm ALG0 described in Theorem 9, which was developed for the separated setting in Section 3. Let ALG0(L) denote the online algorithm, where we run the online algorithm ALG0 on the subproblem (PL,𝒞L).

Online algorithm.

We can now present our online algorithm ALG. In the current algorithm, we use the online algorithm ALG0(L) as a subroutine. For each j{0}, let layer Lj be the set of disks of radii in the interval [2j,2j+1). The algorithm maintains a hitting set HP for the disks presented so far. Upon the arrival of a new disk σ with radius r, if it is already hit by a point in H, then do nothing. Otherwise, proceed as follows.

  • First, find the layer Lj, where j=logr, in which σ belongs.

  • Find the tile τ𝒯j that contains the center of σ.

    • If Pτ, then choose an arbitrary point pPτ and add it to H.

    • Otherwise, for every line Lj that intersects σ, direct L such that L+ contains the center of σ, feed the disk σ to the online algorithm ALG0(L), and add any new hitting point chosen by ALG0(L) to H.

Competitive analysis.

We now prove that ALG is O(logMlogn)-competitive.

Theorem 12.

For the Online Hitting Set problem for a set P of n points in the plane and a sequence 𝒞=(σ1,,σm) of disks of radii in the interval [1,M], the online algorithm ALG has competitive ratio O(logMlogn).

Proof.

Let 𝒞 be a sequence of disks. For each j{0,1,,logM}, let 𝒞j be the collection of disks in 𝒞 with radii in the interval [2j,2j+1). Let H and OPT, resp., be the hitting set returned by the online algorithm ALG and an (offline) minimum hitting set for 𝒞. For every point pOPT, let 𝒞p be the set of disks in 𝒞 containing p. For each j{0,1,,logM}, let 𝒞pj be the set of disks in 𝒞j containing p, i.e., 𝒞pj=𝒞j𝒞p. Let HpjH be the set of points that ALG adds to H in response to hit objects in 𝒞pj. It is enough to show that for every j{0,1,,logM} and pOPT, we have |Hpj|O(logn).

Let τ be the tile in 𝒯j that contains p, and let 𝒞pj𝒞pj be the subset of disks whose centers are located in τ. To hit the first disk σ𝒞pj, our algorithm adds a point from Pτ to H. By Observation 10, any point in Pτ hits σ, as well as any subsequent disks in 𝒞pj. Our algorithm adds at most 1 point to H to hit all the disks in 𝒞pj.

It remains to bound the number of points our algorithm adds for disks in 𝒞pj𝒞pj. Notice that a disk D0 centered at p of radius 2j+1 contains all the centers of the disks in 𝒞pj𝒞pj. By the triangle inequality, a disk D centered at p of radius 2j+2 contains all disks in 𝒞pj𝒞pj. For any disk σ𝒞pj𝒞pj, our algorithm uses algorithm ALG0(L) for a line Lj, directed such that L+ contains the center of σ. According to Observation 11, the disk D intersects at most 24 lines in j. However, depending on the location of the center of σ, each line may be used in either direction for ALG0(L). As a result, for all disks in 𝒞pj𝒞pj, algorithm ALG0(L) is called with at most 48 directed lines L.

For each directed line L, the online algorithm ALG0(L) maintains a hitting set H(L) for the disks fed into this algorithm. For the point p, let Hpj(L) denote the set of points that algorithm ALG0(L) adds to H(L) in response to a disk in 𝒞pj𝒞pj that it receives as input. By Theorem 9, we have |Hpj(L)|O(log|𝒞pj𝒞pj|)O(logn) for every directed line L. This yields |Hpj|1+48O(logn)=O(logn), as required.

By construction, we have H=j=0logMpOPTHpj. We have shown that |Hpj|=O(logn), for all j{0,1,,logM} and pOPT. Consequently, we obtain |H|j=0logMpOPTO(logn)=(logM+1)|OPT|O(logn)=O(logMlogn)|OPT|.

For disks of radii in [1,1+ε] where ε>0 is constant, Theorem 12 implies the following.

Corollary 13.

For the Online Hitting Set problem for a set P of n points in the plane and a sequence 𝒞=(σ1,,σm) of disks of radii in the interval [1,1+ε], where ε>0 is a constant, the online algorithm ALG is O(logn)-competitive.

5 Generalization to Positive Homothets of a Convex Body

In this section, we generalize Theorem 12 for positive homothets of an arbitrary convex body C in the plane. A set C2 is a convex body if it is convex and has a nonempty interior; and it is centrally symmetric (w.r.t. the origin) if C=C, where C={p:pC}.

The key components of our O(logn)-competitive algorithm for disks of comparable sizes were an O(logn)-competitive online algorithm in the line-separated setting and a grid tiling that allowed a reduction to the line-separated setting. Specifically, Observation 10 and Observation 11 formulate the two essential properties of a tiling: If a center of disk σ lies in a tile τ, then τσ (Observation 10); and every disk intersects O(1) grid lines (Observation 11).

We state the main result of this section here. For a detailed explanation and the complete proof, please refer to the full version of the paper.

Theorem 14 ().

Given any convex body σ2 and a parameter M1, there is an online algorithm with a competitive ratio of O(logMlogn) for the Online Hitting Set problem for a set P of n points in the plane and a sequence 𝒞=(σ1,,σm) of positive homothets σi=aiσ+bi, where ai[1,M].

For positive homothets of a convex object with scaling factor in [1,1+ε], where ε>0 is a constant, Theorem 14, implies the following.

Corollary 15.

Given any convex body σ2 and constant ε>0, there is an online algorithm of competitive ratio O(logn) for the Online Hitting Set problem for a set P of n points in the plane and a sequence 𝒞=(σ1,,σm) of positive homothets σi=aiσ+bi, where ai[1,1+ε].

A good pair of lines.

The key technical tool for the proof of Theorem 14 is Definition 16. Given a convex body C, we first consider an inscribed triangle of the maximum area (see [13]). We then apply an area-preserving (unary) affine transformation to transform C so that this inscribed triangle of the maximum area becomes an equilateral triangle. (This is similar to mapping the minimum enclosing ellipse of C into a circle, or assuming that C is “fat” after a suitable affine transformation.) We may further assume, by scaling, that the inscribed circle of this triangle has a unit diameter.

Definition 16.

Let C be a convex body in the plane such that an inscribed triangle of the maximum area is an equilateral triangle Tin, and the circle inscribed in Tin is a circle of a unit diameter. A pair of lines {1,2} is a good pair for 𝐂 if they satisfy the following properties:

  1. 1.

    The angle between the two lines is bounded from below by (1,2)π/15.

  2. 2.

    For i{1,2}, there exist points pi,qiC such that the two lines tangent to C parallel to i contain pi and qi, respectively; furthermore, C contains the disk B(x,150) of diameter 125 centered at the intersection point x=p1q1p2q2.

In the full version, we prove that every convex body C specified in Definition 16 admits a good pair of lines, which can be computed in polynomial time if C is a convex polygon. No attempts were made to optimize the constants π/15 and 150 in Definition 16.

6 Conclusions and Open Problems

We revisited the Online Hitting Set problem for a set of n points in the plane and geometric objects that arrive in an online fashion, such as disks or homothets of a convex body of comparable sizes, or bottomless rectangles in the plane. In all these cases, we designed O(logn)-competitive online algorithms, which is the best possible. It remains an open problem whether our results generalize to 3- or higher dimensions. In fact, no O(logn)-competitive algorithm is currently known for simple geometric objects in 3-space, for example, a set of n points and a sequence of unit balls in 3; or a set of n points P[0,n)33 and a sequence of axis-aligned cubes in 3.

Our results provide further evidence that there may exist O(logn)-competitive algorithms for the Online Hitting Set problem for n points in d and any sequence of objects 𝒞 of bounded VC-dimension – an open problem raised by Even and Smorodinsky [12]; see also [14]. This problem remains open: The current best lower and upper bounds are Ω(logn) and O(log2n) [2]. No better bounds are known even in some of the most common geometric range spaces, for example, when P is a subset of the grid [0,n)22 and 𝒞 is a sequence of axis-aligned rectangles in the plane; or when P is a set of n points in the plane, and 𝒞 is a sequence of disks of arbitrary radii.

References

  • [1] Shanli Alefkhani, Nima Khodaveisi, and Mathieu Mari. Online hitting set of d-dimensional fat objects. In Proc. 21st International Workshop on Approximation and Online Algorithms (WAOA), volume 14297 of LNCS, pages 134–144. Springer, 2023. doi:10.1007/978-3-031-49815-2_10.
  • [2] Noga Alon, Baruch Awerbuch, Yossi Azar, Niv Buchbinder, and Joseph Naor. The online set cover problem. SIAM J. Comput., 39(2):361–370, 2009. doi:10.1137/060661946.
  • [3] Moses Charikar, Chandra Chekuri, Tomás Feder, and Rajeev Motwani. Incremental clustering and dynamic information retrieval. SIAM J. Comput., 33(6):1417–1440, 2004. doi:10.1137/S0097539702418498.
  • [4] Jonathan B. Conroy and Csaba D. Tóth. Hop-spanners for geometric intersection graphs. Journal of Computational Geometry, 14(2):26–64, 2023. doi:10.20382/jocg.v14i2a3.
  • [5] Minati De, Saksham Jain, Sarat Varma Kallepalli, and Satyam Singh. Online geometric covering and piercing. Algorithmica, 86:2739–2765, 2024. doi:10.1007/s00453-024-01244-1.
  • [6] Minati De, Ratnadip Mandal, and Satyam Singh. New lower bound and algorithm for online geometric hitting set problem. In Computing and Combinatorics - 31st International Computing and Combinatorics Conference, COCOON 2025, Proceedings, Part I, volume 15983 of Lecture Notes in Computer Science, pages 234–247. Springer, 2025. doi:10.1007/978-981-95-0215-8\_18.
  • [7] Minati De and Satyam Singh. Hitting geometric objects online via points in d. In Proc. 28th Conf. Computing and Combinatorics (COCOON), volume 13595 of LNCS, pages 537–548. Springer, 2022. doi:10.1007/978-3-031-22105-7_48.
  • [8] Minati De and Satyam Singh. Online hitting of unit balls and hypercubes in d using points from d. Theor. Comput. Sci., 992:114452, 2024. doi:10.1016/J.TCS.2024.114452.
  • [9] Adrian Dumitrescu, Anirban Ghosh, and Csaba D. Tóth. Online unit covering in Euclidean space. Theor. Comput. Sci., 809:218–230, 2020. doi:10.1016/J.TCS.2019.12.010.
  • [10] Adrian Dumitrescu, Anirban Ghosh, and Csaba D. Tóth. Sparse hop spanners for unit disk graphs. Comput. Geom., 100:101808, 2022. doi:10.1016/J.COMGEO.2021.101808.
  • [11] Adrian Dumitrescu and Csaba D. Tóth. Online unit clustering and unit covering in higher dimensions. Algorithmica, 84(5):1213–1231, 2022. doi:10.1007/S00453-021-00916-6.
  • [12] Guy Even and Shakhar Smorodinsky. Hitting sets online and unique-max coloring. Discret. Appl. Math., 178:71–82, 2014. doi:10.1016/J.DAM.2014.06.019.
  • [13] Ivor van der Hoog, Vahideh Keikha, Maarten Löffler, Ali Mohades, and Jérôme Urhausen. Maximum-area triangle in a convex polygon, revisited. Inf. Process. Lett., 161:105943, 2020. doi:10.1016/J.IPL.2020.105943.
  • [14] Arindam Khan, Aditya Lonkar, Saladi Rahul, Aditya Subramanian, and Andreas Wiese. Online and dynamic algorithms for geometric set cover and hitting set. In Proc. 39th Symposium on Computational Geometry (SoCG), volume 258 of LIPIcs, pages 46:1–46:17. Schloss Dagstuhl, 2023. doi:10.4230/LIPICS.SOCG.2023.46.