Online Hitting Sets for Disks of Bounded Radii
Abstract
We present algorithms for the online minimum hitting set problem in geometric range spaces: Given a set of 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 , we present an -competitive algorithm. This result generalizes from disks to positive homothets of any convex body in the plane with scaling factors in the interval . 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 , we present an -competitive algorithm for the variant where is a subset of an section of the integer lattice, and the geometric objects are bottomless rectangles.
Keywords and phrases:
Geometric Hitting Set, Online Algorithm, Homothets, DisksFunding:
Minati De: Research on this paper was supported by SERB MATRICS Grant MTR/2021/ 000584.Copyright and License:
2012 ACM Subject Classification:
Theory of computation Computational geometry ; Theory of computation Online algorithmsEditors:
Anne Benoit, Haim Kaplan, Sebastian Wild, and Grzegorz HermanSeries and Publisher:
Leibniz International Proceedings in Informatics, Schloss Dagstuhl – Leibniz-Zentrum für Informatik
1 Introduction
In the general form of the Hitting Set problem, we are given a point set and a collection of subsets , and we need to find a subset (hitting set) of minimal size such that every set contains some point in . In the Online Hitting Set problem, the set is known in advance, but the subsets in arrive one at a time (without advance knowledge). We need to maintain a hitting set for the first sets such that for all (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 and obtained almost matching lower bound of .
Geometric Hitting Set.
In the geometric Hitting Set problem, we have for some constant dimension , and the sets in are geometric objects of some type: for example, balls, unit balls, simplices, axis-aligned cubes, or hyper-rectangles. Depending on whether is finite or infinite, there are different versions of the problem. In this paper, we consider to be a finite set of points in .
Related Previous Work.
When 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 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 and a collection of axis-aligned squares with integer coordinates for . They developed an -competitive algorithm for this variant. They also established a randomized lower bound of , where consists of finitely many points and consists of translates of an axis-aligned square. Recently, De et al. [6] considered the variant when is set of points in and consists of homothetic copies of a regular -gon (for ) with scaling factors in the interval , and designed an -competitive randomized algorithm. Even though a disk can be approximated by a regular -gon as , this does not imply any competitive algorithm for disks with radii in the interval .
Our results and Technical Contribution.
We study the Online Hitting Set problem when is set of points in . Table 1 summarizes the existing results and the results of the current paper.
| Points | Objects | Lower Bound | Upper Bound |
| Intervals in | [12] | [12] | |
| Half-planes and translates of a disk in | [12] | [12] | |
| Axis-aligned squares in with integer coordinates | [14] | [14] | |
| Homothetic copies of a regular -gon ( with scaling factors in the interval | [14] | [6] | |
| Bottomless rectangles (for definition, see Section 2.1) | [Theorem 1] | ||
| Disks having radii in the interval | [12] | [Theorem 12] | |
| Positive homothets of an arbitrary convex body in with scaling factors in the interval | [14] | [Theorem 14] |
We now present our contributions and briefly discuss the technical ideas involved.
Bottomless Rectangles in .
We present an -competitive deterministic algorithm for the geometric Online Hitting Set problem, where , and is a sequence of bottomless rectangles of the form arriving one by one (Theorem 1 in Section 2). When a bottomless rectangle arrives, our algorithm chooses hitting points guided by the canonical partition of the interval (see Section 2 for a definition). For each point in an offline optimum, this structured canonical partition ensures that points are sufficient to hit all the incoming rectangles in that are hit by . We prove that our algorithm is -competitive for a broader class of objects– sets with lowest-point property (see Section 2.2 for a definition).
Disks with Radii in .
Our main result is a deterministic -competitive Online Hitting Set algorithm for an arbitrary set of points in the plane, and a sequence of disks of radii in the interval (Theorem 12 in Section 4). Previously, an -competitive algorithm was known only for congruent disks [12]. In particular, our result is the first -competitive algorithm that works for disks of radii in the interval for any constant (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 lie on one side of a line (w.l.o.g., the - or -axis), while 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 -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 -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 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 into 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 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 invocations are relevant. In the competitive analysis, we show that for each point in an offline optimum solution, our algorithm uses hitting points for the disks in each layer that contain . Since there are layers, our algorithm is -competitive.
Homothets of a Convex Body with Diameters in .
We generalize our main result from disks to positive homothets of any convex body in the plane, where the radii in the interval are replaced by scaling factors in the interval (Theorem 14 in Section 5). Our online algorithm is based on a two-stage approach, similar to the case of disks, and it is -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 - or -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 is infinite, one may further distinguish between the continuous setting where (also known as the piercing problem) and the discrete setting where is a discrete subset of (for example, ).
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 . They also proved 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 and a lower bound of . For translates of a centrally symmetric convex body, they proved that the competitive ratio of every deterministic algorithm is at least , where is the illumination number of the object 111The illumination number of an object , denoted by , is the minimum number of smaller homothetic copies of (, where ) whose union contains .. For translates of an axis-aligned hypercube in , Dumitrescu and Tóth [11] proved that the competitive ratio of any deterministic algorithm for Online Set Cover is at least . Later, De et al. [5] studied the Online Hitting Set problem for -fat objects in with diameters in and designed a deterministic algorithm with competitive ratio . For hitting axis-aligned homothetic hypercubes with side lengths in , they gave a deterministic algorithm with competitive ratio at most . They also proved a lower bound for the problem of hitting homothetic hypercubes in , with side lengths in the interval .
Discrete Setting.
De and Singh [7, 8] studied a variant of this problem where and consists of translates of a ball or an axis-aligned hypercube in . For translates of an axis-aligned hypercube, they showed that there is a randomized algorithm with an expected competitive ratio of and also proved that every deterministic algorithm has a competitive ratio of at least . For translates of a ball in , they proposed a deterministic algorithm having a competitive ratio of and proved that every deterministic algorithm has a competitive ratio of at least , for . Recently, Alefkhani et al. [1] considered the variant where and is a family of -fat objects in , for some constant . They proposed a deterministic algorithm with a competitive ratio of at most , and proved that the competitive ratio of every deterministic algorithm is . Very recently, De et al. [6] improved both the upper and lower bounds of Alefkhani et al. [1]. They considered the case where and is a family of -fat objects having diameters in , for some constant . They proposed a deterministic algorithm with competitive ratio , and established that the competitive ratio of any randomized algorithm is .
2 Bottomless Rectangles and Integer Points
We present an -competitive algorithm for the Online Hitting Set problem where is a subset of an 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 -competitive algorithm for the Online Hitting Set problem where is a subset of the integer lattice with nonnegative coordinates less than , that is, ; and the objects are bottomless rectangles. Bottomless rectangles are of the form , where and . Note that there are only combinatorially different rectangles w.r.t. , so the general result by Alon et al. [2] gives an algorithm for the online hitting set with competitive ratio . In this section, we present an -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 is a power of 2, and every bottomless rectangle is given with integer parameters , , and . An interval is canonical if it is of the form for some integers . For a canonical interval , we also define the left neighbor and the right neighbor . For every canonical interval , if , then let denote a lowest-point in (that is, a point with minimum -coordinate; ties are broken arbitrarily). If , then is undefined.
For every interval with nonnegative integer endpoints, we define a canonical partition, i.e., a partition of 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 be the largest integer such that , for some . (Note that is unique. Indeed, suppose that is not unique, say . Since or is even, then or is an integer. Now, we have , which contradicts the maximality of .) We call the integer the splitting point of . We can partition a given interval into canonical intervals as follows. If is not canonical, find its splitting point , partition it into two intervals , and recurse on and . For example, the splitting point of interval is 8, and its canonical partition is ; see Figure 1(a)) for an illustration.
Note also that in the canonical partition of (resp., ), there is at most one interval of each size, where the possible sizes are powers of 2 between 1 and (resp., ). Specifically, if is in the canonical partition of , then its left neighbor is not contained in , consequently , where is the closure of . Similarly, if is in the canonical partition of , then .
Online algorithm ALG for bottomless rectangles.
We can now present our online algorithm. We maintain a hitting set , which is initially empty: . When the -th bottomless rectangle arrives, initialize . If , then do not add any new points to . Otherwise, we may assume that . Compute the splitting point of , and the canonical partitions and of and , respectively. If , then find the largest canonical interval such that , and put . Similarly, if , then find the largest interval such that , and put . Overall, we add at most two new points to in step .
Competitive analysis.
We now prove that ALG is -competitive.
Theorem 1.
For the Online Hitting Set problem for a point set and a sequence of bottomless rectangles, the online algorithm ALG has competitive ratio .
Proof.
Let be a sequence of bottomless rectangles. Let and OPT be the hitting set returned by the online algorithm ALG and an (offline) minimum hitting set of , respectively. For a point , let be the subsequence of bottomless rectangles that contain . It is enough to show that for every , our algorithm adds points to in response to the objects in .
Let , with coordinates ; and let be a sequence of bottomless rectangles in for which our algorithm adds new points to the hitting set. We show that . We can distinguish between two types of rectangles depending on whether the -coordinate of is on the left or right of the splitting point of : namely, or . We analyze the two types separately (the two cases are analogous).
Assume w.l.o.g. that for . This means that , and so ALG adds the hitting point for exactly one interval . Suppose that the algorithm adds the hitting point for . Then is the largest (hence rightmost) canonical interval in such that . Recall that , where is the left neighbor of the canonical interval . This implies that . That is, either or its left neighbor contains . Note that is contained in canonical intervals (one for each possible size), and each of these canonical intervals has a unique right neighbor. Consequently, is one of at most canonical intervals under the assumption that for all . This proves that .
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 . For a set , the span of , denoted is the smallest interval with integer endpoints such that . An object has the lowest-point property if for every point in and every interval that contains , the object contains all points in with the minimum -coordinates. For an illustration of set see Figure 1(b). Note, in particular, that every bottomless rectangle has the lowest-point property: Indeed, if , then .
Our online hitting set algorithm and its analysis readily generalize when the objects have the lowest-point property. Let be a sequence of objects with the lowest-point property.
Online algorithm for objects with the lowest-point property.
We maintain a hitting set , which is initially empty: . When set arrives, initialize . If , then do not add any new points to . Suppose that . Let . Compute the splitting point of , and the canonical partitions and of and , respectively. If , then find the largest interval such that , and put . Similarly, if , then find the largest interval such that , and put . Overall, we add at most two new points to in step .
Correctness and competitive analysis.
When adds a points to in step , the lowest-point property ensures that . Therefore, maintains that is a hitting set for , proving the correctness of . We now show that is -competitive.
Theorem 2.
For the Online Hitting Set problem for a point set and a sequence of objects with the lowest-point property, algorithm has competitive ratio .
Proof.
Let be a sequence of objects with the lowest-point property. Let and OPT be the hitting set returned by the online algorithm and an (offline) minimum hitting set of , respectively. For each , let the subsequence of sets in that contain . It is enough to show that for every , our algorithm adds points to in response to the objects in .
Let , with coordinates ; and let be a sequence of sets in for which our algorithm adds new points to the hitting set. We show that . We can distinguish between two types of sets depending on whether the -coordinate of is to the left or right of the splitting point : namely, or . We analyze the two types separately (the two cases are analogous).
Assume w.l.o.g. that for . This means that , and so ALG adds the hitting point for exactly one interval . Suppose that the algorithm adds the hitting point for . Then is the largest (hence rightmost) canonical interval in such that . Recall that , where is the left neighbor of the canonical interval . This implies that . That is, either or its left neighbor contains . Note that is contained in canonical intervals (one for each possible size), and each of these canonical intervals has a unique right neighbor. Consequently, is one of at most canonical intervals under the assumption that for all . This proves that .
3 Disks in the Plane: Separated Setting
In this section, we consider the Online Hitting Set problem in the plane, where is a finite set above the -axis (given in advance); and consists of disks of arbitrary radii with centers located on or below the -axis (arriving one-by-one). Note that the disks in do not necessarily have the lowest-point property; see Figure 2.
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 of a point set is the smallest convex set in the plane that contains . Equivalently, it is the intersection of all closed half-planes that contain ; it can be computed by the classical “rotating calipers” algorithm, where we continuously rotate a line around while remains in one closed half-plane bounded by . Intuitively, we obtain the unit disk hull of by rolling a unit disk, with center on or below the -axis, around . We generalize this notion to disks of any fixed radius .
Definition 3.
Let be a finite set of points above the -axis and let . Let be the set of all disks of radius with centers on or below the -axis. Let be the union of all disks such that . Now, we define the -hull of as . The boundary of is denoted by ; for an illustration see Figure 3.
Dumitrescu et al. [10, Lemma 4] proved that is -monotone222A curve in the plane is -monotone if every vertical line intersects it at most once. for any , and established other properties, which were used by Conroy and Tóth [4], as well.
Lemma 4 (Dumitrescu et al. [10]).
For a finite set above the -axis and , the following holds:
-
1.
lies above the -axis;
-
2.
every vertical line intersects in one point, thus is an -monotone curve;
-
3.
for every disk , the intersection is connected (possibly empty);
-
4.
for every disk , if , then contains a point in .
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 w.r.t. different radii; see Figure 4. We start with an easy observation.
Lemma 5.
Let and be circular arcs lying entirely above the -axis, such that and are arcs of circles and , resp., of radii and , with centers on or below the -axis.
-
1.
Then both and are -monotone and concave curves.
-
2.
Furthermore, if points are contained in both and , and , then lies above (i.e., for every vertical line that separates and , point lies above point ).
Proof.
(1) For every , the center of is below the -axis, and so the leftmost and rightmost points of are also below the -axis. The leftmost and rightmost points partition into two halfcircles, one above the center and one below the center. Both halfcircles are -monotone: The lower halfcircle is convex curve and the upper halfcircle is concave. Since lies entirely above the -axis, it is contained in the upper halfcirlce, which is -monotone and concave.
(2) The locus of centers of circles that contain both and is the orthogonal bisector of the line segment , that we denote by .
Note that is not vertical (or else would be a horizontal line above the -axis, and the centers of and would also be above the -axis). As the center a circle containing and continuously moves from the center of down to , the circular arc between and deforms continuously from to the line segment .
Since concave, it lies above the segment . Since , the arc lies between the arc and the segment and . Consequently,
then lies below , as claimed.
Lemma 6.
For every finite set above the -axis, the following holds:
-
1.
If , then for every disk of radius , the intersection is connected (possibly empty).
-
2.
Suppose that lies on the curve for some . Then there is a radius such that is also on for all , but is below for all .
Proof.
(1) Let . Suppose, to the contrary, that the
intersection has two or more components.
By Lemma 4(2), the -coordinates of the components form
disjoint intervals, and the components have a natural left-to-right ordering.
Let be the rightmost point in the first component, and
let be the leftmost point in the second component. Clearly
. Let be an arbitrary point in between and . Then lies on the boundary
of some disk of radius whose center is below the -axis,
and whose interior is disjoint from . In particular, neither
nor is in the interior of . Since the center of is
below the -axis, contains two interior-disjoint
circular arcs between and the -axis; and both arcs must cross
. We have found two intersection points above the -axis. Furthermore, between and ,
the circular arc lies above the circular arc , contradicting
Lemma 5(2). This completes the proof of Property 1.
(2) Consider a point that lies on the curve for some . Then there exists a disk of radius centered at some point below the -axis such that .
Let be the intersection point of the -axis the line , and the orthogonal projection of to the -axis. We describe two continuous motions, where the disk continuously changes while is in the circle and there is no point in in the interior of : First, a central dilation from center continuously moves to a disk centered at . Second, the center of moves from towards continuously until its center reaches or a point where contains both and another point . Let be radius of at that time.
The continuous motion shows that for all ,
but it is not in for all .
Reduction.
We can reduce the Online Hitting Set problem for a finite set 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:
For a finite point set in the plane above the -axis, let be the set of points such that for some .
Lemma 7.
For a finite point set in the plane above the -axis, has the following property: For every disk centered below the -axis, if , then .
Proof.
Let be a disk of radius centered below the -axis. By Lemma 4(4), contains a point in . By the definition of , this point is in .
We may assume that the points in have distinct -coordinates (if two or more points in have the same -coordinate, w.l.o.g. a minimum hitting set would contain only the point with the smallest -coordinate). Sort by increasing -coordinates such that . For every point , let be the maximum radius such that . Consider the set of radii . Sort the radii in in decreasing order as . We can now define the function . For every , let , that is, the first coordinate of corresponds to the index of (the -order of all points in ), and the second coordinate of corresponds to index of the radius . For every , let ; see Figure 5 for an illustration. Finally, let and . Note that the points in lie above all points in . Since is injective, then it is a bijection between and . Note also that , consequently .
Lemma 8.
For a set of points in the plane above the -axis and for every disk centered below the -axis, the set has the lowest-point property.
Proof.
Let be a disk centered below the -axis. We rephrase the lowest-point property in terms of . Recall that the points in are sorted by -coordinates. Suppose that is in and . Consider the point sets . By Lemma 7, we know that ; let be the largest radius in such that . We need to show that contains all points in .
Let and , resp., be the leftmost and rightmost points in ; and let and be the vertical lines through and . By the definition of , we have and , and by the definition of . Consequently, the intersection point lies at or below , the intersection point lies at or below . Since , then contains both and . We know that is an -monotone curve by Lemma 4(2), and is connected by Lemma 6(2). Since contains both and , then contains the sub-curve of between and . Since all points in are between the vertical lines and , then contains all points in , 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 of points above the -axis and disks centered below the -axis, there is an -competitive algorithm.
Proof.
We are given a set of points above the -axis, and we receive a sequence of disks centered on or below the -axis in an online fashion. Let be a minimum hitting set for .
Initially, we compute the set as defined above Lemma 8. When a disk arrives, we compute the set , which has the lowest-point property by Lemma 8. The bijection maps OPT to a set , where . Here, is a hitting set for the sets .
We run the online algorithm described in Section 2.2 for the point set and the sequence of sets. By Theorem 1, ALG returns a hitting set of size . By Lemma 7, is a hitting set for , and its size is bounded by , as required.
4 Disks of Bounded Radii: General Setting
In this section, we consider the Online Hitting Set problem, where is a finite set (given in advance) in the plane; and the objects are disks with radii in the interval , where is a constant.
Distinguishing layers of disks, according to their radii.
We partition the disks of radii in the interval into layers as follows: for each , let layer be the set of disks of radii in the interval . The index of each layer is denoted by .
Tiling of the plane for each layer index .
For every , let be a two-dimensional lattice spanned by vectors and , where and are the standard basis vectors. Let be a square of side length with lower-left corner at the origin. Translates of (tiles), with translation vectors in the lattice , form the tiling . Let denote the set of axis-parallel lines spanned by the sides of the tiles in .
We observe two key properties of the construction of layers and the tilings.
Observation 10.
For every , if and the center of is in a tile , then ; see Figure 6.
Proof.
Since , the radius of the disk is in at least . The tile is a translate of , and so its diameter is . If the center of is in , then every is within distance from , which implies that .
Observation 11.
For every , every disk of radius at most intersects at most 24 lines in : at most 12 horizontal and 12 vertical lines.
Proof.
Let be a disk of radius ; see Figure 6. The orthogonal projection of to the -axis (resp., -axis) is an interval of length at most . Since the distance between any two consecutive vertical (resp., horizontal) lines in is , then intersects at most horizontal and at most 12 vertical lines in .
Subproblem for a directed line .
For a directed line , we denote by and the closed half-plane on the left and right of , respectively. Given a directed line and the input ) of the Online Hitting Set problem, where is a set of points, and is a sequence of disks in the plane, we define a subproblem as follows: Let , and let be the subsequence of disks such that the center of is in and contains at least one point in . Now for each subproblem , we can run the online algorithm described in Theorem 9, which was developed for the separated setting in Section 3. Let denote the online algorithm, where we run the online algorithm on the subproblem .
Online algorithm.
We can now present our online algorithm ALG. In the current algorithm, we use the online algorithm as a subroutine. For each , let layer be the set of disks of radii in the interval . The algorithm maintains a hitting set for the disks presented so far. Upon the arrival of a new disk with radius , if it is already hit by a point in , then do nothing. Otherwise, proceed as follows.
-
First, find the layer , where , in which belongs.
-
Find the tile that contains the center of .
-
–
If , then choose an arbitrary point and add it to .
-
–
Otherwise, for every line that intersects , direct such that contains the center of , feed the disk to the online algorithm , and add any new hitting point chosen by to .
-
–
Competitive analysis.
We now prove that ALG is -competitive.
Theorem 12.
For the Online Hitting Set problem for a set of points in the plane and a sequence of disks of radii in the interval , the online algorithm ALG has competitive ratio .
Proof.
Let be a sequence of disks. For each , let be the collection of disks in with radii in the interval . Let and OPT, resp., be the hitting set returned by the online algorithm ALG and an (offline) minimum hitting set for . For every point , let be the set of disks in containing . For each , let be the set of disks in containing , i.e., . Let be the set of points that ALG adds to in response to hit objects in . It is enough to show that for every and , we have .
Let be the tile in that contains , and let be the subset of disks whose centers are located in . To hit the first disk , our algorithm adds a point from to . By Observation 10, any point in hits , as well as any subsequent disks in . Our algorithm adds at most 1 point to to hit all the disks in .
It remains to bound the number of points our algorithm adds for disks in . Notice that a disk centered at of radius contains all the centers of the disks in . By the triangle inequality, a disk centered at of radius contains all disks in . For any disk , our algorithm uses algorithm for a line , directed such that contains the center of . According to Observation 11, the disk intersects at most 24 lines in . However, depending on the location of the center of , each line may be used in either direction for . As a result, for all disks in , algorithm is called with at most 48 directed lines .
For each directed line , the online algorithm maintains a hitting set for the disks fed into this algorithm. For the point , let denote the set of points that algorithm adds to in response to a disk in that it receives as input. By Theorem 9, we have for every directed line . This yields , as required.
By construction, we have . We have shown that , for all and . Consequently, we obtain .
For disks of radii in where is constant, Theorem 12 implies the following.
Corollary 13.
For the Online Hitting Set problem for a set of points in the plane and a sequence of disks of radii in the interval , where is a constant, the online algorithm ALG is -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 in the plane. A set is a convex body if it is convex and has a nonempty interior; and it is centrally symmetric (w.r.t. the origin) if , where .
The key components of our -competitive algorithm for disks of comparable sizes were an -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 lies in a tile , then (Observation 10); and every disk intersects 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 and a parameter , there is an online algorithm with a competitive ratio of for the Online Hitting Set problem for a set of points in the plane and a sequence of positive homothets , where .
For positive homothets of a convex object with scaling factor in , where is a constant, Theorem 14, implies the following.
Corollary 15.
Given any convex body and constant , there is an online algorithm of competitive ratio for the Online Hitting Set problem for a set of points in the plane and a sequence of positive homothets , where .
A good pair of lines.
The key technical tool for the proof of Theorem 14 is Definition 16. Given a convex body , we first consider an inscribed triangle of the maximum area (see [13]). We then apply an area-preserving (unary) affine transformation to transform so that this inscribed triangle of the maximum area becomes an equilateral triangle. (This is similar to mapping the minimum enclosing ellipse of into a circle, or assuming that 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 be a convex body in the plane such that an inscribed triangle of the maximum area is an equilateral triangle , and the circle inscribed in is a circle of a unit diameter. A pair of lines is a good pair for if they satisfy the following properties:
-
1.
The angle between the two lines is bounded from below by .
-
2.
For , there exist points such that the two lines tangent to parallel to contain and , respectively; furthermore, contains the disk of diameter centered at the intersection point .
In the full version, we prove that every convex body specified in Definition 16 admits a good pair of lines, which can be computed in polynomial time if is a convex polygon. No attempts were made to optimize the constants and in Definition 16.
6 Conclusions and Open Problems
We revisited the Online Hitting Set problem for a set of 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 -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 -competitive algorithm is currently known for simple geometric objects in 3-space, for example, a set of points and a sequence of unit balls in ; or a set of points and a sequence of axis-aligned cubes in .
Our results provide further evidence that there may exist -competitive algorithms for the Online Hitting Set problem for points in and any sequence of objects 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 and [2]. No better bounds are known even in some of the most common geometric range spaces, for example, when is a subset of the grid and is a sequence of axis-aligned rectangles in the plane; or when is a set of points in the plane, and is a sequence of disks of arbitrary radii.
References
- [1] Shanli Alefkhani, Nima Khodaveisi, and Mathieu Mari. Online hitting set of -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 . 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 using points from . 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.
