
In this article, we study sums related to the Lehmer problem over short intervals, and give two asymptotic formulae for them. The original Lehmer problem is to count the numbers coprime to a prime such that the number and the its number theoretical inverse are in different parities in some intervals. The numbers which satisfy these conditions are called Lehmer numbers. It prompts a series of investigations, such as the investigation of the error term in the asymptotic formula. Many scholars investigate the generalized Lehmer problems and get a lot of results. We follow the trend of these investigations and generalize the Lehmer problem.
Citation: Yanbo Song. On two sums related to the Lehmer problem over short intervals[J]. AIMS Mathematics, 2021, 6(11): 11723-11732. doi: 10.3934/math.2021681
[1] | Cheng-Hung Huang, Tsung-Yi Lee . Predicting the optimal heating function for uniform exit temperature in a pipe flow. AIMS Mathematics, 2024, 9(1): 1997-2021. doi: 10.3934/math.2024099 |
[2] | Jie Liu, Bo Sang, Lihua Fan, Chun Wang, Xueqing Liu, Ning Wang, Irfan Ahmad . Symmetry, Hopf bifurcation, and offset boosting in a novel chameleon system. AIMS Mathematics, 2025, 10(3): 4915-4937. doi: 10.3934/math.2025225 |
[3] | Rattanasak Hama, Sorin V. Sabau . Randers metrics on two-spheres of revolution with simple cut locus. AIMS Mathematics, 2023, 8(11): 26213-26236. doi: 10.3934/math.20231337 |
[4] | Xiaohang Su, Peng Liu, Haoran Jiang, Xinyu Yu . Neighbor event-triggered adaptive distributed control for multiagent systems with dead-zone inputs. AIMS Mathematics, 2024, 9(4): 10031-10049. doi: 10.3934/math.2024491 |
[5] | Young Joon Ahn . G2/C1 Hermite interpolation of offset curves of parametric regular curves. AIMS Mathematics, 2023, 8(12): 31008-31021. doi: 10.3934/math.20231587 |
[6] | Mauro Costantini, Pierpaolo Soravia . On the optimal second order decrease rate for nonlinear and symmetric control systems. AIMS Mathematics, 2024, 9(10): 28232-28255. doi: 10.3934/math.20241369 |
[7] | Yujun Li . Characteristics of planar sextic indirect-PH curves. AIMS Mathematics, 2024, 9(1): 2215-2231. doi: 10.3934/math.2024110 |
[8] | Maryam T. Aldossary, Rashad A. Abdel-Baky . On the Blaschke approach of Bertrand offsets of spacelike ruled surfaces. AIMS Mathematics, 2022, 7(10): 17843-17858. doi: 10.3934/math.2022983 |
[9] | Li Dong, Jingyong Tang . New convergence analysis of a class of smoothing Newton-type methods for second-order cone complementarity problem. AIMS Mathematics, 2022, 7(9): 17612-17627. doi: 10.3934/math.2022970 |
[10] | Ghulam Mustafa, Syeda Tehmina Ejaz, Dumitru Baleanu, Yu-Ming Chu . The inequalities for the analysis of a class of ternary refinement schemes. AIMS Mathematics, 2020, 5(6): 7582-7604. doi: 10.3934/math.2020485 |
In this article, we study sums related to the Lehmer problem over short intervals, and give two asymptotic formulae for them. The original Lehmer problem is to count the numbers coprime to a prime such that the number and the its number theoretical inverse are in different parities in some intervals. The numbers which satisfy these conditions are called Lehmer numbers. It prompts a series of investigations, such as the investigation of the error term in the asymptotic formula. Many scholars investigate the generalized Lehmer problems and get a lot of results. We follow the trend of these investigations and generalize the Lehmer problem.
Computing the surface at a constant distance from a curve or another surface is a fundamental operation in Geometry Processing [1], which plays a key role in engineering and manufacturing [2,3], or architectural design [4,5]. These surfaces also appear in the description of wavefronts propagating in a homogeneous medium.
Pipe and offset surfaces can be regarded as the 3D generalization of planar offset curves (Figure 1). Given a distance d and a central curve S (also called spine [6]), a pipe or tube Sd is defined as the locus of points x at a distance d from points p∈S along normal lines N(p) to S. By replacing the spine with a base surface S (also known as progenitor), its offset Sd (dubbed parallel surface in classical differential geometry [7,8]) is defined in the same manner. These two families of surfaces are intimately related, as offsetting a pipe yields another pipe with the same spine. Their extension to a general n-dimensional space are both called offset. Also, throughout this paper, the unifying term progenitor will be employed to denote the spine, base surface, or the base curve when dealing with offsets to planar curves.
This explicit definition of pipes/offsets allows their construction and visualization. However, it requires a regular progenitor S, in the sense that S does not display singularities. Indeed, for pipes, we assume a spine S without cusps or self-intersections (Figure 2), so that we can define a tangent vector at each point of S and its the complementary normal plane. Similarly, the above definition of offset surfaces assumes a progenitor S with no singularities, such as cuspidal edges, cusps, or self-intersections, so that at each point of S a tangent plane is well-defined, and hence it makes sense to speak of normal directions. Observe that, if a progenitor surface S is connected and orientable, its offset consists of two components [8].
As an alternative to this explicit definition, we could characterize an offset implicitly for a general set S of points, not necessarily a regular curve or surface, as the level set at constant distance d from S. This survey employs this implicit form conceptually to determine the offset smoothness and will reveal the precise connection between the two offset versions (implicit and explicit), issues not clearly addressed in the existing literature.
In general, a pipe or offset Sd is topologically more complex than its progenitor S since singularities may appear as d grows. The computation and detection of such singularities is a well-studied topic [2,9]. These singularities alter the original shape of S, so Alcázar and Sendra [10], and Alcázar [11], for plane algebraic curves and rational algebraic surfaces, respectively, put forward the topological concept of good local behavior, meaning that offsetting preserves the local shape. In the case of offsets to plane algebraic curves, this concept extends to good global behavior [12].
In a real design scenario, we need the upper bound of d that guarantees the regularity of Sd. Before the advent of CAGD (Computer Aided Geometric Design), Federer [13] published a seminal work on sets with positive reach that deserves renewed attention. He extended the concept of normal vectors to deal with convex sharp edges and vertices, which enriches the admissible geometries for explicit offset construction. He showed that offsetting any closed set S yields a C1-smooth surface, for distances d<R, where R denotes the so-called reach of the set S. In a CAGD setting, Maekawa et al. [6], and Wallner et al. [14] compute this bound for C2 pipes and offsets, respectively, by carefully analyzing all possible global and self-intersections. This bound plays a fundamental role in the construction of reliable (topologically equivalent) approximations to Sd [6,15] and, in the case of the offset to compact surfaces, it coincides with minimal distance to the cut locus of the surface, i.e., the set of points equally distant from at least two points on the surface. Thus, these works show how to compute the reach (in the closed case), albeit without employing this terminology.
Although a rich literature exists on detecting singularities of pipes and offsets, surprisingly, their general smoothness has received scant attention. Our contribution focuses on clarifying how it depends on that of the progenitor by bringing together existing yet dispersed concepts, making them more accessible, and deriving the relevant results the applied mathematics community needs.
The most remarkable result regarding the smoothness of offset surfaces is due to Hermann [16]. In CAGD, surfaces are usually defined parametrically in a piecewise manner, and their smoothness is characterized by the so-called geometric continuity Gk [17], which means the agreement of i-th derivatives (i=0,…,k) at the joints of adjacent Ck patches, after suitable reparameterization. For a piecewise base surface composed of C∞ patches, joined with geometric continuity Gk along a common boundary, such as non-degenerate NURBS [18,19], Hermann proves that the offset inherits the Gk continuity. His approach relies on checking the existence of the above-mentioned reparameterization, yet without explicitly constructing it.
To derive a more general result, we assume no particular representation (implicit or parametric) of the progenitor, vindicate Federer's results, and extend them to Ck progenitors, so the paper is arranged as follows. In Section 2, we introduce the different geometries (sets of points) required for a unified handling of progenitors and resulting pipes and offsets. We begin with the classical concept of smooth submanifolds and extend them to sets of positive reach in Federer's framework involving the cone of normal vectors. Next, in Section 3, we define the untrimmed offset as the set obtained by displacing points of S a constant distance d along normal lines, and the trimmed offsets as the level set at distance d. In Section 4, we review the concept of reach and associated results noting that, if d<R, trimmed and untrimmed offset coincides. In Section 5, we employ existing results on the smoothness of the distance function to a closed Ck submanifold to show that, if d<R, then Sd inherits the class Ck of S. Finally, conclusions are drawn in Section 6.
Rather than in the constructional way of geometric continuity, we will characterize the smoothness of pipe/offset surfaces by defining them as a subset of points fulfilling certain conditions. For this purpose, we could employ the classical concept [20] of surface of class Ck in R3, or, for a more comprehensive scope, that of submanifold of class Ck [21], in Euclidean space Rn:
Definition 1. Let m,n be integers such that 1≤m<n. A subset S⊂Rn is a m-dimensional submanifold of class Ck, k≥1, if either of the following equivalent conditions [21] is fulfilled:
(i) For each point x∈S there is an open neighborhood U⊂Rn and a homeomorphism X:V⊂Rm→S∩U of class Ck such that the differential dX has rank=m for all points in U.
(ii) For each point x∈S there is an open neighborhood U and a local function F:U⊂Rn→Rn−m of class Ck such that F−1(0)=S∩U and the differential dF has rank=n−m for all points over S∩U.
Characterization (i) requires the existence of a local parametrization X of class Ck around every point x∈S, whereas (ii) requires that S be defined locally by an implicit equation F(x)=0, with F of class Ck.
Remark 1. For k=1, the first characterization reduces to that of a regular surface. Therefore, this concept of class extends the notion of regularity.
Remark 2. The concept of Ck class, stronger than that of Ck curves or surfaces employed in CAGD, encompasses the case of regular Gk piecewise surfaces: according to characterization (i), they are of class Ck.
Remark 3. For hypersurfaces (m=n−1), in characterization (ii) dF simplifies to dF=∇F, with ∇F≠0. Later, we will come across the so called hypersurfaces of class C1,1 [23], weaker than C2, but stronger than C1:
Definition 2. A C1,1 hypersurface is a (n−1)-dimensional submanifold characterized by a C1,1 implicit function F, i.e, such that ∇F is 1-Lipschitz.
CAGD applications such as solid modeling [22] deal with a region of Rn, rather than a hypersurface. The closed set bounded by a hypersurface (without boundary) corresponds to the concept of n-dimensional manifold with boundary. A general and formal definition for submanifolds of arbitrary dimension with boundary is found in the classical book [21]. The particular case of a n-dimensional manifold with boundary is, in essence, a set locally diffeomorphic to the half-space Hn=[0,∞)×Rn−1.
The concept of submanifold allows a unified handling of smooth progenitor curves and surfaces, and hence of the resulting pipes and offsets. However, it excludes progenitors with singularities, a severe restriction in solid modeling. For instance, a simple polyhedral solid does not qualify as a submanifold. Federer's work provides a more general setting by extending the concept of tangent and normal lines [13,24]:
Definition 3. The tangent cone of vectors Tan(p,S) at each point p∈S of a given set S⊂Rn consists of the limits of all (secant) vectors originating from p and passing through a sequence of points pi∈S∖ p that converges to p. The corresponding dual cone of normal vectors Nor(p,S) consists of all vectors v such that ⟨u,v⟩≤0 whenever u∈Tan(p,S).
This elegant framework extends beyond the submanifold realm to the so-called sets with positive reach. Although the concept of reach will be formally in Section 4, we advance that these sets are characterized [25,26] by admitting a value R>0 such that we can roll up a ball of a radius d<R all over the boundary ∂S. Thus, they encompass the following closed sets S∈Rn (Figure 3):
a) Traditional submanifolds S (without boundary): The cone Nor(p,S) coincides with the customary normal direction that, in Definition 1, the conditions on the rank guarantee.
b) n-dimensional submanifolds S with boundary ∂S: The cone consists of the outward normal vectors to ∂S.
c) Closed sets whose boundary ∂S may contain convex sharp edges and vertices: At these singularities, the cone (highlighted in orange in Figure 3c) fills the gaps between the normal directions of adjacent faces. In [27], the term normal pyramid is employed instead to denote this cone.
In the general framework of Section 2.2, from Definition 3 we can construct the normal bundle of a given set S:
ν(S):={(p,v)∈Rn×Rn: p∈S and v∈Nor(p,S)}. | (3.1) |
and the map σ:
σ:ν(S)→Rn,σ(p,v)=p+v. | (3.2) |
We assume that ∀p∈S the cone Nor(p,S) does not degenerate to a null vector v=0, so that the above concepts make sense. Consequently, we exclude sets whose boundary contains sharp concave edges or vertices.
Now, we are ready to give a formal definition of the explicit offset, dubbed untrimmed in contrast to its trimmed (level set) sibling:
Definition 4. The untrimmed offset Sd at distance d to S is the set constructed by displacing all points p∈S a constant distance d along normal lines N(p):
Sd:={σ(p,v)∈Rn : (p,v)∈ν(S) and ‖v‖=d}. | (3.3) |
Remark 4. This definition encompasses pipes Sd, generated from a submanifold S of dimension m=1 (a curve).
Remark 5. When S is a solid with convex sharp edges and vertices, such as the cone of Figure 4, Sd (3.3) includes the resulting vertex and edge offsets. These components (highlighted in orange) fill the gaps between face offsets (blue), making up the topology Farouki [28] identified.
Since the map σ (3.2) is Ck−1 [29], we could jump to the conclusion that a Ck submanifold S generates another Ck−1 submanifold Sd (3.3). However, on the one hand we must constrain values d to ensure that Sd actually meets the definition of submanifold, because Sd may have singularities, such as the self-intersections of the inner component of the planar offset in Figure 5a. These singularities substantially alter the shape of the original progenitor S. On the other hand, this class can be improved to Ck.
To characterize the smoothness of Sd (3.3), the key idea is using an alternative offset construction, what Farouki [30,31] calls trimmed or true offset:
Definition 5. Given a non-empty set S, its trimmed offset at distance d is the level set δ−1(d,S) of the distance function δ to S:
δ:Rn→R,δ(x,S)=infp∈S‖x−p‖, | (3.4) |
When no ambiguity is possible, we drop the symbol S in (3.4) and upcoming definitions. The catch is that, once again, δ−1(d) may display singularities (Figure 5b). Furthermore, δ−1(d) does not always coincide with its untrimmed counterpart Sd (3.3), since Sd may contain points x at distance δ(x,S)<d, points trimmed away [12,32,33,34,35] in δ−1(d). In the example of Figure 5b, the inner component of δ−1(d) would even vanish for a large d. The concept of reach of a set, reviewed in the upcoming Section, will elucidate the connection between the offsets Sd and δ−1(d), i.e., Definitions 4 and 5.
Definition 6. Given a set S, we denote U(S) the set of all x∈Rn with the unique nearest point property, i.e., ∀x∈U(S), there is a unique point π(x,S)∈S such that δ(x)=‖x−π(x,S)‖. The map π is called the nearest point function, or projection onto S:
π:U(S)→S. | (4.1) |
Definition 7. For points p∈S, if Br(p) denotes the open n−ball of radius r centered at p, then the local reach(p,S) around p and global reach(S) are defined as:
reach(p,S):=sup{r:Br(p)⊂U(S)},R=reach(S):=infp∈Sreach(p,S). | (4.2) |
We say that a set S has positive reach if R>0. Geometrically [25,26], as already noted, this property means that we can roll up a ball of radius at most R all over the boundary ∂S.
All sets with positive reach are necessarily closed [13], so, from now on, we restrict ourselves to closed sets S. Thus, R admits a more familiar interpretation [41,42], using the complement Rn∖U(S), i.e., the set of points having more than one nearest point. This complement set corresponds to the so-called cut locus L=L(S) (without its limits points) [14,25,36,37]. In terms of L, the local and global reach (4.2) are given by (Figure 6a):
reach(p,S)=δ(p,L),R=infp∈S,q∈L‖p−q‖. | (4.3) |
Consequently, positive reach is tantamount to saying that L(S) does not touch S [25].
Equipped with the above definitions, we show that the reach R determines when Sd and the level set δ−1(d) coincide. For closed sets, Theorem 4.8 in [13] furnishes two fundamental results:
(i) The projection (4.1) occurs always along a normal line N(p) to S, as Figure 6b sketches:
x−π(x)=v∈Nor(π(x),S),∀x∈U(S). | (4.4) |
(ii) Moving from p along N(p) up to a distance reach(p,S), all points x project onto p, and hence the distance δ(x,S) is precisely along N(p). In terms of the map σ (3.2):
π(σ(p,v))=pδ(σ(p,v))=‖v‖,for‖v‖<reach(p,S). | (4.5) |
Result (i) means that δ−1(d)⊂Sd, whereas (ii) implies that, for d<reach(S), then Sd⊂δ−1(d) and Sd⊂U(S). Let us summarize these relationships:
Theorem 1. Let S be a closed set with reach(S)=R>0. Then, for d∈(0,R), its untrimmed offset Sd (3.3) at distance d coincides with the level set δ−1(d), and Sd⊂U(S).
Remark 6. This Theorem can be regarded as the generalized version, to sets with positive reach, of Corollary 8 in [15], formulated for the offset to compact C2 surfaces. Sakkalis et al. [15] do not employ the terminology reach to denote the upper bound of the size of the neighborhood enjoying the unique nearest points property. However, and more importantly, they give a practical recipe for its numerical computation.
In general, both the local and global reach improve when we consider a set S instead of its boundary ∂S (Figure 7a). Indeed ([13], Remark 4.2), for points p∈∂S:
reach(p,S)≥reach(p,∂S)⇒reach(S)≥reach(∂S). | (4.6) |
A sharp edge or vertex in ∂S (Figure 7b) implies that L(∂S) touches ∂S, so reach(∂S)=0, whereas S may still have positive reach. If S is the interior region bounded by ∂S, then, in terms of cut loci, L(S) contains only the exterior of L(∂S), i.e., excluding its interior component, which coincides with the medial axis of S [25,36,37]. The medial axis can hence be defined as the closure of the locus of centers of maximal balls inscribed in S [2,31,36,37]. The exterior component of L(∂S) is involved in trimming the outer offsets to the solid S, whereas offsets inside the pocket S are trimmed against the medial axis [31,38,39,40]. A warning: in [41,42], and also in [43] in a submanifold context, the term medial axis is used instead to denote the cut locus, but the terminology we follow is more widespread [2,25,31,36,37,38,39,40,44,45].
A remarkable case is the generalized tubular neighborhood
S≤r:={x:δ(x,S)≤r},r<reach(S), | (4.7) |
called r-offset in [41,42]. Corollary 4.9 in [13] relates the properties of S≤r and S:
δ(x,S≤r)=δ(x,S)−rwheneverδ(x,S≤r)≥r,reach(S≤r)≥reach(S)−r. | (4.8) |
In particular, we are interested in the case where S is a spine and hence S≤r is bounded by a pipe Sr, as Figure 8 illustrates at a normal section through a point p∈S. Combining relationships (4.8) and Theorem 1, we formalize the relationship between pipes and their offsets mentioned in the Introduction:
Lemma 1. Let S≤r be the r-offset (4.7) bounded by the pipe Sr. Then, the offset to S≤r at a distance (d−r), with r<d<reach(S), coincides with the level set δ−1(d−r,S≤r) and the pipe Sd=δ−1(d,S).
We recall the well-known result on the reach of general m-dimensional submanifolds [26,29,46]:
Lemma 2. Compact C2 submanifolds have positive reach R>0.
Violating either of the conditions required in this Lemma may lead to a zero reach:
a) The planar curve S of Figure 9a, adapted from [21], illustrates the unbounded case. As S tapers off to infinity along an asymptote contained in L, the points p∈S get closer and closer. Consequently, reach(p,S)→0 and R=0.
b) Figure 9b shows that a C1 curve may have points p with zero reach, contrary to intuition. Consider the graph of the C1 function y(x)=|x|3/2, an example taken from [46,47] and clearly a C1 curve. If we draw the normal lines to S of constant length d>0 (upper side), no matter how small d is, if we zoom in near the origin O, there exists an interval x∈[−ϵ,ϵ] such that the normals at f(x) intersect. Therefore, O does not admit any neighborhood with the unique nearest point property, i.e., reach(O,S)=0. Equivalently by (4.3), L (the non-negative vertical semi-axis) touches O.
Remark 7. By inequality (4.6), Lemma 2 carries over to n-dimensional submanifolds S with compact C2 boundary.
Remark 8. For compact hypersurfaces S, Lucas [23] proved that the C2 condition can be relaxed, since S is of positive reach if and only if it enjoys C1,1 continuity.
As a consequence of Theorem 1, given a closed set S the function
F(x)=δ(x)−d | (5.1) |
characterizes Sd as a submanifold according to Definition 1(ii). Since ∇F=∇δ, the smoothness of the distance function δ carries over to Sd. Whereas Federer [13] already analyzed the differentiability of the distance function δ to general closed sets, the smoothness of higher derivatives is well-studied in the setting of submanifolds of class Ck. Thus, we proceed with a separate analysis for general closed sets, which include closed C1 submanifolds, and smoother Ck submanifolds, k≥2.
For general closed sets, it is well-known that ∇δ(x) enjoys Lipschitzian character everywhere but S∪L(S), as Federer [13] (Theorem 4.8) proved:
Theorem 2. Let S⊂Rn be a nonempty closed set. Then, on Int(U(S)∖S) the distance (3.4) δ(x) to S is a C1,1 function with gradient:
∇δ(x)=x−π(x)δ(x)≠0,x∈Int(U(S)∖S). | (5.2) |
In a CAGD setting, Wolter [36] rederived this result, without reference to Federer's work. For d<R, Theorem 2 implies a C1,1 function F (5.1) and, by invoking Definition 2, the C1,1 class of the trimmed offset δ−1(d). Since trimmed and untrimmed offset Sd coincide (Theorem 1), Sd inherits this class [13,42]:
Corollary 1. Let S be a closed set with positive reach R>0. Then, its offset Sd at distance d∈(0,R) is a C1,1 hypersurface.
Remark 9. In the case of a closed set S with sharp edges and vertices, hence not eligible as a smooth submanifold, Corollary 1 implies that offsetting with d<R improves the smoothness. Figure 4 illustrates this property with a solid cone S, which enjoys R=∞ for being a convex set [13], thereby admitting a C1,1 offset Sd for any d.
Regarding the smoothness of the distance to Ck submanifolds, k≥2, at first glance, δ seems Ck−1, since δ is always continuous and expressible in terms of the direction normals to S. However, Gilbarg and Trudinger [48] first noted that, for hypersurfaces, δ is actually Ck in a certain neighborhood of S when k≥2. By incorporating Federer's work, Krantz and Parks [47] added the case k=1, whereas Foote [29], with an elegant coordinate-free proof, extended the result to general submanifolds:
Lemma 3. Let S be a compact Ck submanifold, with k≥2. Then, it admits a generalized tubular neighborhood S≤r⊂U(S) such that the distance δ(x,S) is Ck for x∈Int(S≤r∖S).
Krantz and Parks provide another proof in [49] and, in Section 4.4 of their book [46], summarize the most relevant results on the smoothness of δ. Also, the survey by Jones et al. [25] on distance fields briefly reviews the differentiability of the distance function.
The main shortcoming of Lemma 3, limiting their applicability to characterize the smoothness of pipes/offsets, is that it does not specify the size r of the tubular neighborhood. However, when S is a Ck hypersurface, it can locally be expressed as the graph of a Ck function [21], which leads to the more specific result about r. Giusti [50] (Appendix B), adapting the original result by Gilbarg and Trudinger [48] and without citing Federer's paper [13], proves that δ(x,S) to a compact set S bounded by a Ck submanifold is Ck, for distances δ(x,S)<R such that points x have a unique nearest point and property (4.4) applies. Extending the argumentation to the boundary by considering its normal space in (4.4), since this bound is precisely R=reach(S), we can rewrite Giusti's result in a more precise way by quantifying r:
Lemma 4. Let S be a compact Ck hypersurface, with k≥2. Then, it has reach R>0 and the distance δ(x) is Ck for points x such that δ(x)∈(0,R).
Lemma 4 is still incomplete for our purposes, as it only applies to hypersurfaces, not to general m-dimensional submanifolds. To find a more powerful result, we make a short foray into Riemannian geometry [51] and extend our setting, replacing Rn and the Euclidean distance d with a general n-dimensional Riemannian space M and its induced inner metric. The concepts of submanifold, unique nearest point property [52], set of positive reach [53] and hence cut locus extend to this new framework. Moreover, Mantegazza and Mennucci [54] (Proposition 4.6, point 7) derive the equivalent of Theorem 2 for Ck submanifolds. For our purposes, where M=Rn, their remarkable result particularizes as follows:
Theorem 3. Let S be a closed Ck submanifold, with k≥2. Then, the distance δ(x) is Ck for points x everywhere but L(S)∪S.
Remark 10. With respect to Lemmas 3, 4, Theorem 3 relaxes the condition on S, from compact to closed. However, it no longer guarantees R>0, as the unbounded curve in Figure 8a illustrates.
With a similar argumentation to that leading from Theorem 2 to Corollary 1, from Theorem 3 we derive the sibling of Corollary 1 for closed Ck progenitors:
Corollary 2. Let S be a closed Ck progenitor with k≥2 and of positive reach R>0. Then, its offset Sd at distance d∈(0,R) is a Ck hypersurface.
Remark 11. Corollary 2 applies, with improved reach (4.6), to the offset to an n−dimensional submanifold with Ck boundary. It also implies that offsetting with d<R keeps the smoothness of S. Figure 10 illustrates this property in the case k=2, with the C2 tube Sd generated from a C2 spine S, defined as a periodic cubic B-spline [3] from its control polygon. Figure 11 shows another example, namely offsetting a solid (i.e., a 3-dimensional submanifold) with C2 boundary, using a commercial solid-modeling system. This operation yields a C2 offset surface if d<R, whereas trying a value d≥R triggers a system error, warning that the resulting geometry is unsupported.
To analyze the smoothness of a pipe or offset Sd, at distance d from a progenitor S, we have brought together and adapted existing concepts and results, especially those from Federer's pioneering paper [13] on sets with positive reach. This work is usually overlooked in the literature on CAGD, with few exceptions [41,42,43], as well as in classical books on differential geometry [7,8] when dealing with tubes and offsets. On the other hand, the mathematical research on the smoothness of the distance function to submanifolds does not focus on the construction of pipes/offsets and seems unaware of the CAGD applied results on their regularity.
The reach R of S, i.e., the minimal distance from S to its cut locus, plays the key role. For closed sets S with positive reach, the condition d<R guarantees:
● The coincidence between two offset versions: the explicit untrimmed offset Sd (obtained by displacing points of S a constant distance d along normal lines) and the trimmed offset (level set δ−1(d) of the distance δ to S).
● The containment of Sd in a neighborhood with the unique nearest point property, i.e., that any point has a unique point on the progenitor at minimal distance.
● That Sd is a C1,1 hypersurface. If S is a submanifold of class Ck (or the region bounded by a closed Ck hypersurface), this class is inherited by the distance function to S and hence by Sd. Hence, offsetting with d<R improves or keeps the smoothness of S.
Sets with R>0 are hence those admitting a singularity-free offset for d<R. Along with compactness, a C2 class of a submanifold implies R>0, but, contrary to intuition, C1 smoothness does not.
Unlike the most relevant previous work [16] on higher-order smoothness of offset surfaces, our results hold for general closed progenitors of class Ck, not necessarily piecewise made up of C∞ patches with Gk joints along common boundaries. Also, they hold in Euclidean space Rn of arbitrary dimension n, for pipe or offset hypersurfaces Sd, at constant distance d from a progenitor spine or hypersurface.
Our discussion applies only to the classical offset to a hypersurface, that is, along its normal. We do not consider the so-called generalized offsets [55,56,57,58], i.e., with translation along a vector different from the normal. Admittedly, a generalized offset to S can be rewritten as a standard offset Sd (at constant distance d) to a new progenitor ˆS [59,60]. However, ˆS involves tangents and normals to Sd. Consequently, a progenitor S of class Ck generates a new ˆS of class only Ck−1.
This research was supported by grant PID2019-104586RB-I00 funded by MCIN/AEI/10.13039/501100011033; grant SBPLY/19/180501/000247 by Consejería de Educación Cultura y Deportes (Junta de Comunidades de Castilla-La Mancha); and grant 2021-GRIN-31214 by Univ. de Castilla-La Mancha; co-financed by the ERDF (European Regional Development Fund). We are grateful to a referee for their valuable suggestions, which substantially improved this article.
The authors declare that there is no conflict of interest.
[1] | R. K. Guy, Unsolved problems in number theory, New York: Springer-Verlag, 2004. |
[2] | W. P. Zhang, On D. H. Lehmer problem, Chin. Sci. Bull., 21 (1992), 1765–1769. |
[3] | W. P. Zhang, On a problem of D. H. Lehmer and its generalization, Compos. Math., 86 (1993), 307–316. |
[4] | W. P. Zhang, A problem of D. H. Lehmer and its generalization (II), Compos. Math., 91 (1994), 47–56. |
[5] |
Y. Yi, W. P. Zhang, On the generalization of a problem of D. H. Lehmer, Kyushu J. Math., 56 (2002), 235–241. doi: 10.2206/kyushujm.56.235
![]() |
[6] | I. E. Shparlinski, On a generalised Lehmer problem for arbitrary powers, East-West J. Math., 2008. |
[7] |
I. E. Shparlinski, On a generalisation of a Lehmer problem, Math. Z., 263 (2009), 619–631. doi: 10.1007/s00209-008-0434-2
![]() |
[8] |
Y. Lu, Y. Yi, On the generalization of the D. H. Lehmer problem, Acta Math. Sinica, Engl. Ser., 25 (2009), 1269–1274. doi: 10.1007/s10114-009-7652-3
![]() |
[9] |
Z. F. Xu, W. P. Zhang, On a problem of D. H. Lehmer over short intervals, J. Math. Anal. Appl., 320 (2006), 756–770. doi: 10.1016/j.jmaa.2005.07.054
![]() |
[10] | H. N. Liu, W. P. Zhang, On a problem of D. H. Lehmer, Acta Math. Sinica, Engl. Ser., 22 (2006), 61–68. |
[11] |
Z. Y. Zheng, On the ordinary problem of D. H. Lehmer, Chin. Sci. Bull., 38 (1993), 1060–1065. doi: 10.1360/csb1993-38-12-1060
![]() |
[12] |
P. Xi, Y. Yi, Kloosterman sums and a problem of D. H. Lehmer, Chin. Ann. Math. Ser. B, 41 (2020), 361–370. doi: 10.1007/s11401-020-0203-z
![]() |
[13] |
H. Zhang, W. P. Zhang, Some new sums related to D. H. Lehmer problem, Czech. Math. J., 65 (2015), 915–922. doi: 10.1007/s10587-015-0217-y
![]() |
[14] | S. Chowla, On Kloosterman's Sum, Norske Vid. Selsk. Forh., 40 (1967), 70–72. |
[15] |
P. Xi, Y. Yi, Generalized D. H. Lehmer problem over short intervals, Glasgow Math. J., 53 (2011), 293–299. doi: 10.1017/S0017089510000704
![]() |
[16] | C. D. Pan, C. B. Pan, Goldbach conjecture, Beijing: Science Press, 1981. |
1. | Feifei Qu, Xin Wei, Limit behaviour of constant distance boundaries of Jordan curves, 2022, 7, 2473-6988, 11311, 10.3934/math.2022631 | |
2. | Javier Sánchez-Reyes, Self-Intersections of Cubic Bézier Curves Revisited, 2024, 12, 2227-7390, 2463, 10.3390/math12162463 |