Loading [MathJax]/jax/output/SVG/jax.js
Research article Special Issues

Detecting affine equivalences between certain types of parametric curves, in any dimension

  • Two curves are affinely equivalent if there exists an affine mapping transforming one of them onto the other. Thus, detecting affine equivalence comprises, as important particular cases, similarity, congruence and symmetry detection. In this paper we generalized previous results by the authors to provide an algorithm for computing the affine equivalences between two parametric curves of certain types, in any dimension. In more detail, the algorithm is valid for rational curves, and for parametric curves with nonrational but meromorphic components, it admits an also meromorphic, and in fact rational, inverse. Unlike other algorithms already known for rational curves, the algorithm completely avoids polynomial system solving, and instead uses bivariate factoring as a fundamental tool. The algorithm has been implemented in the computer algebra system Maple and can be freely downloaded and used.

    Citation: Juan Gerardo Alcázar, Hüsnü Anıl Çoban, Uğur Gözütok. Detecting affine equivalences between certain types of parametric curves, in any dimension[J]. AIMS Mathematics, 2024, 9(6): 13750-13769. doi: 10.3934/math.2024670

    Related Papers:

    [1] Yan Shi, Qunzhen Zheng, Jingben Yin . Effective outcome space branch-and-bound algorithm for solving the sum of affine ratios problem. AIMS Mathematics, 2024, 9(9): 23837-23858. doi: 10.3934/math.20241158
    [2] Yanxia Hu, Qian Liu . On traveling wave solutions of a class of KdV-Burgers-Kuramoto type equations. AIMS Mathematics, 2019, 4(5): 1450-1465. doi: 10.3934/math.2019.5.1450
    [3] Kareem T. Elgindy, Hareth M. Refat . A direct integral pseudospectral method for solving a class of infinite-horizon optimal control problems using Gegenbauer polynomials and certain parametric maps. AIMS Mathematics, 2023, 8(2): 3561-3605. doi: 10.3934/math.2023181
    [4] Haijun Cao, Fang Xiao . The category of affine algebraic regular monoids. AIMS Mathematics, 2022, 7(2): 2666-2679. doi: 10.3934/math.2022150
    [5] Thierry Dana-Picard, Tomás Recio . Dynamic construction of a family of octic curves as geometric loci. AIMS Mathematics, 2023, 8(8): 19461-19476. doi: 10.3934/math.2023993
    [6] Muhsin Incesu . LS (3)-equivalence conditions of control points and application to spatial Bézier curves and surfaces. AIMS Mathematics, 2020, 5(2): 1216-1246. doi: 10.3934/math.2020084
    [7] Kangqiao Li, Gongxiang Liu . Finite duals of affine prime regular Hopf algebras of GK-dimension one. AIMS Mathematics, 2023, 8(3): 6829-6879. doi: 10.3934/math.2023347
    [8] Tareq M. Al-shami, Abdelwaheb Mhemdi, Radwan Abu-Gdairi, Mohammed E. El-Shafei . Compactness and connectedness via the class of soft somewhat open sets. AIMS Mathematics, 2023, 8(1): 815-840. doi: 10.3934/math.2023040
    [9] Dmitry Sinelshchikov . On an integrability criterion for a family of cubic oscillators. AIMS Mathematics, 2021, 6(11): 12902-12910. doi: 10.3934/math.2021745
    [10] Mudassir Shams, Nasreen Kausar, Serkan Araci, Liang Kong . On the stability analysis of numerical schemes for solving non-linear polynomials arises in engineering problems. AIMS Mathematics, 2024, 9(4): 8885-8903. doi: 10.3934/math.2024433
  • Two curves are affinely equivalent if there exists an affine mapping transforming one of them onto the other. Thus, detecting affine equivalence comprises, as important particular cases, similarity, congruence and symmetry detection. In this paper we generalized previous results by the authors to provide an algorithm for computing the affine equivalences between two parametric curves of certain types, in any dimension. In more detail, the algorithm is valid for rational curves, and for parametric curves with nonrational but meromorphic components, it admits an also meromorphic, and in fact rational, inverse. Unlike other algorithms already known for rational curves, the algorithm completely avoids polynomial system solving, and instead uses bivariate factoring as a fundamental tool. The algorithm has been implemented in the computer algebra system Maple and can be freely downloaded and used.



    We say that two curves are affinely equivalent if one of them is the image of the other curve by means of an affine mapping. If the affine mapping preserves angles, then the two curves are similar, i.e., both correspond to the same shape and differ only in position and/or scaling. If the affine mapping preserves distances, the curves are congruent or isometric, so they differ only in position. Finally, if the two curves coincide, finding the self-congruences of the curve is equivalent to computing its (Euclidean) symmetries.

    Because of the nature of the problem, it has received much attention in applied fields like computer aided geometric design, pattern recognition and computer vision. In the last years, however, the problem has been also addressed in the computer algebra field, and here we follow this trend. Examples of computer algebra papers where this question has been studied are [2,4,7,10]; for other related papers, the interested reader can check the bibliographies of [2,4,7,10]. These papers address the problem for rational curves, i.e., parametric curves whose components are quotients of polynomial functions; while [2] only considers symmetries, and [4,7,10] aim for the more general question of checking projective, and not just affine, equivalence. In [4,10], we can find solutions for this problem with different strategies for curves in any dimension and use it as a fundamental tool for polynomial system solving, whereas the paper [7] addresses the question only for space rational curves, but employing bivariate factoring as an alternative to solving polynomial systems; this leads to better timings and performance. To do this, in [7], two rational invariants, i.e., two functions rationally depending on the parametrizations and their derivatives to be studied, which stay invariant under projective transformations, are found and used.

    In this paper, we generalize the ideas of [7] in three different ways. First, while the development of the invariants in [7] was more of an "art" than a "craft", here we provide a complete algorithm to generate such invariants. Second, the technique is valid for curves in any dimension, and not just space curves, which was the case addressed in [7]: We provide an algorithm that can be downloaded from [8] to generate the corresponding invariants for any dimension, that needs to be executed just once for each dimension. Third, our strategy is valid not only for rational curves, but also for non-algebraic, parametric curves with meromorphic components under certain conditions, so we can apply the algorithm for catenary curves or 3D spirals, for instance, whenever some hypotheses are fulfilled. In order to also include this type of non-algebraic curves, we stick to affine equivalences, and not projective equivalences, although our ideas could be developed in a projective setting for rational curves. Notice that strategies like [4,10], based on polynomial system solving, cannot handle non-algebraic curves because we would need to solve systems involving analytic functions.

    Although the implementation of our algorithm is relatively simple, and can be downloaded from [8] jointly with the examples worked out in this paper, justifying how the invariants are generated is involved, and extremely technical; because of this, in this paper we focus on the ideas and the algorithm itself, and refer the interested readers to the ArXiv version of the paper [3] for the detailed deduction of the invariants.

    The structure of this paper is the following. In Section 2, we review some previous work related to the problem in order to provide some intuition on our solution. Also in that section, we make precise the kind of curves we can work with and present some necessary tools to understand the algorithm. In Section 3, we develop the method and give the results leading to the algorithm; here, we focus on the ideas and theorems and skip the (very technical) details, for which the interested reader is referred to [3]. Finally, the algorithm itself, together with several examples and an account of the experimentation carried out in Maple, is given in Section 4. We close the paper with our conclusion and some open questions in Section 5.

    The fundamental theorem of space curves [5] states that the curvature κp and the torsion τp of a space curve C defined by a parametrization p,

    κp=|p×p||p|3, τp=p,p,p|p×p|,

    where || denotes the norm and represents the determinant, define the curve up to congruences. As a consequence, if we are given two curves C1,C2 defined by parametrizations p,q, we can check whether the corresponding curves are congruent by checking whether or not their curvature and torsion coincide. Thus, we say that κp,τp are invariants for congruences, in the sense that κp,τp stay invariant when a congruence is applied. Additionally, τp and also κ2p (instead of κp) are said to be rational invariants because they correspond to invariant (under congruences), rational expressions in p and its derivatives.

    Curvature and torsion are used in [2] to compute the Euclidean symmetries of a rational space curve, i.e., a curve defined by a rational parametrization (so that their components are quotients of polynomials). The reason is the following theorem, used in [2] and other papers treating similar questions, like [4,7,10]. Two ingredients are important in this theorem: the first one is that it is required that the parametrizations of the curves are proper, i.e., birational, meaning that they are invertible and have rational inverses. This hypothesis is easy to check [17]; also, if a parametrization is not proper, it can be properly reparametrized [17]. The second ingredient is the fact (see also [17]) that the only birational mappings of the real or complex line are the Möbius transformations,

    φ(z)=az+bcz+d, adbc0. (2.1)

    Theorem 2.1. Let C1,C2Cn be two parametric curves defined by rational proper parametrizations p(z),q(z). If f is a birational mapping such that f(C1)=C2, then there exists a Möbius transformation φ(z) satisfying that

    fp=qφ, (2.2)

    i.e., making commutative the following diagram:

    (2.3)

    The reason why curvature and torsion are useful when applying the above theorem for f a Euclidean symmetry, i.e., an isometry, is not only that they are rational invariants, but the fact that they satisfy the following relationships: If p is a parametization and φ is a Möbius transformation, then [2]

    κpφ=κpφ, τpφ=τpφ.

    We refer to this property by saying that the curvature and torsion commute with Möbius transformations. This allows us to recover the symmetries of a rational curve by using bivariate factoring [2], where the factors we are looking for are the Möbius-like factors H(z,ω)=ω(cz+d)(az+b). This immediately provides us with the Möbius transformations φ in Theorem 2.1; the transformations f themselves are then computed from Eq (2.2).

    This strategy is generalized, also for space curves, in [7] in order to find the projective equivalences, if any, between two space rational curves defined by proper, rational parametrizations p(z),q(z). We say that two curves C1,C2 are projectively equivalent if there exists a projective mapping f such that f(C1)=C2, in which case we call f a projective equivalence between the curves. If we are interested in computing projective equivalences, the curvature and torsion are no longer useful, and we need to produce other rational invariants, namely invariants under the group of projective transformations. Furthermore, since projective mappings are birational, Theorem 2.1 can be applied, but to generalize the strategy of [2] and use also bivariate factoring to solve the problem, we need to produce rational invariants which commute with Möbius transformations. This is exactly the task carried out in [7], for rational space curves.

    In this paper we want to generalize the approach of [7] in two different directions, namely the dimension, so that now we aim at curves living in any dimension n and not only space curves, and the class of curves, which will be not only the class of rational curves but also a more general class of non-algebraic curves parametrized by meromorphic functions, satisfying certain hypotheses. The key idea is that we want to work in a more general setup, but where Theorem 2.1 is still valid.

    Because we want to include nonalgebraic curves, we will limit ourselves to affine equivalences, and not projective equivalences. Given two curves C1,C2Cn we say that C1,C2 are affinely equivalent if there exists a mapping f:CnCn, f(x)=Ax+b with AMn×n(C), i.e., A is an n×n matrix (in general, over the complex), A nonsingular, bCn, such that f(C1)=C2; furthermore, we say that f is an affine equivalence between C1,C2. Although for technical reasons we will consider C1,C2Cn, we will mostly work with real curves, i.e., curves with infinitely many real points, and we will be interested in real affine equivalences; thus, we will be mostly looking at the case when AMn×n(R), bRn. If C1=C2=C is a real curve and A,b are real with A an orthogonal matrix, so ATA=I where I is the identity matrix, then we say that f is a (Euclidean) symmetry of C.

    Remark 2.1. Notice that in order to work with projective equivalences we need to work with homogeneous parametrizations. However, while the homogenization of a rational parametrization is a well-defined notion, the same notion is not well-defined for nonrational, meromorphic parametrizations, since we lack a notion of degree.

    In order to enlarge the class of curves to deal with, which will include as a subset the class of rational curves, we need to take a closer look to the hypotheses in Theorem 2.1. Essentially, we need two things: (1) to guarantee that the parametrizations p(z),q(z) of the curves C1,C2 have global inverses, so the diagram in Eq (2.3) is commutative; (2) to guarantee that the mapping φ at the bottom of Eq (2.3) is a Möbius transformation. An essential observation is that Möbius transformations are not only the birational transformations of the complex line, but the bimeromorphic transformations of the complex line (see Remark 2 in [1]); recall that a transformation g:CC is bimeromorphic iff g is meromorphic, and has an inverse g1 which is also meromorphic. Since the commutativity of the diagram Eq (2.3) implies that φ=q1fp, φ1=p1f1q, if p,q are bimeromorphic parametrizations, Theorem 2.1 works perfectly replacing the hypothesis that p,q are proper rational parametrizations, by the hypothesis that p,q are bimeromorphic parametrizations. We state this as a theorem.

    Theorem 2.2. Let C1,C2Cn be two parametric curves defined by bimeromorphic parametrizations p(z),q(z). If f is an affine mapping such that f(C1)=C2, then there exists a Möbius transformation φ(z) satisfying Eq (2.2), i.e., making commutative the diagram in Eq (2.3).

    Notice that proper rational parametrizations are bimeromorphic, so we are definitely enlarging the class of curves we work with. However, while checking whether or not a rational parametrization is proper is easy and fast, checking whether a non-rational parametrization is bimeromorphic is extremely difficult. This is understandable, since it requires to verify whether a non-rational mapping admits a global inverse, which is a very hard problem. For this reason, we will present now a scheme, which includes rational parametrizations and a wider class of non-rational, meromorphic parametrizations, where the requirement of being bimeromorphic is guaranteed, and can be algorithmically checked. In order to do that, we start with a meromorphic function ξ:CC. Defining

    Π:CC2, Π(z)=(z,ξ(z)),

    we observe that Π is an invertible function over its image, which is the graph Gξ of the function ξ,

    Gξ={(z,ξ(z))|zC}C2. (2.4)

    Indeed, for (z,ω)C2, ω=ξ(z), we have Π1(z,ω)=z. Next, consider a rational mapping Φ:C2Cn. If we compose these two mappings, we get a new mapping

    p=ΦΠ, p:CCn, (2.5)

    which provides a parametrization p(z)=Φ(z,ξ(z)) of a curve CCn, which is the image of Gξ under Φ. Notice that p is a vector function with meromorphic components. Of course if ξ is a rational function, p is just a rational parametrization. We will also assume that the curve defined by p is not contained in a hyperplane.

    Now we want to impose sufficient conditions on Φ to ensure that p1 exists and is meromorphic, in which case p is bimeromorphic. Of course, this holds when ξ is rational and p=ΦΠ is proper, since p is a proper rational parametrization. However let us see that this is also the case whenever Φ is a birational mapping, so that Φ1 exists and is rational, and ξ is a non-algebraic meromorphic function; we will see that this is also a condition that we can algorithmically verify. Thus, let us assume that Φ is a birational mapping, and let p=ΦΠ. Recall that the cardinality of the fiber of Φ, which we denote by #(Φ), is the number of points in the pre-image of Φ(q) with qC2 a generic point. The birationality of Φ is equivalent to #(Φ)=1 (see for instance Proposition 7.16 in [9]), so we can check this condition by just picking a random point q, and computing the number of points in the pre-image of Φ(q). Furthermore, we have the following lemma, inspired by [15,16].

    Lemma 2.1. Let Φ:C2Cn be a birational mapping, then the set of points qC2 such that #(Φ(q))>1 is included in an algebraic variety VC2 of dimension at most 1.

    Proof. Let

    Φ(x1,x2)=(Φ1(x1,x2),,Φn(x1,x2))=(y1,,yn).

    If Φ is birational, then Φ1 exists and is rational, i.e.,

    Φ1(y1,,yn)=(Ψ1(y1,,yn),Ψ2(y1,,yn)),

    where for i=1,2,

    Ψi(y1,,yn)=Ai(y1,,yn)Bi(y1,,yn)

    with Ai,Bi polynomials. The set of points qC2 such that #(Φ(q))>1 is included in the set VC2 where Φ1 is not defined, V being the union of the sets defined by (BiΦ)(x1,x2)=0 with i=1,2, and Ni(x1,x2)=0, i=1,2, where Ni is the denominator of (AiΦ)(x1,x2). Notice that V is an algebraic planar curve.

    Proposition 2.1. Assume that Φ is a birational mapping, ξ is a non-algebraic, meromorphic function, Gξ corresponds to Eq (2.4), and p(z)=(ΦΠ)(z) as in Eq (2.5), then p(z) is invertible over C=Φ(Gξ), and the inverse p1 has rational components. As a consequence, p(z) is bimeromorphic.

    Proof. Since by assumption Φ is a birational mapping, i.e., #(Φ)=1, Φ1 exists and is rational. Next from Lemma 2.1 we have that #(Φ) is constant except perhaps for the points of an algebraic variety VC2 of dimension at most one. Since ξ is not an algebraic function, Gξ is not an algebraic curve. Therefore, by the identity theorem (see Theorem 3.1.9 in [11]) GξV is either finite, or infinite but without any accumulation point. Thus, for a generic point qGξ we get that the cardinality of Φ(q) is 1, so Φ1 is well-defined for almost all points in Φ(Gξ), and p1=Π1Φ1. Since Φ1,Π1 are rational, p1 is rational as well. Since every rational function is meromorphic, p1 is meromorphic; and since p is meromorphic by construction, p is bimeromorphic.

    Therefore, we have found a new class of parametric, non-rational curves, for which Theorem 2.2 holds. We formulate this as a corollary.

    Corollary 2.1. Let C1,C2Cn be two parametric curves defined by p(z),q(z), where p(z),q(z) are either proper rational parametrizations, or can be written as

    p=Φ1Π1,q=Φ2Π2,

    where for i=1,2, Πi(z)=(z,ξi(z)), ξi(z) is a non-algebraic, meromorphic function, and Φi:C2Cn is a birational mapping. If f is an affine mapping such that f(C1)=C2, then there exists a Möbius transformation φ(z) satisfying Eq (2.2), i.e., making commutative the diagram in Eq (2.3).

    Thus, the curves meeting the hypotheses in Corollary 2.1 will be the curves that we will consider as the input to our problem. The next example provides some non-algebraic, planar and space, parametric curves satisfying the hypotheses in Corollary 2.1; these curves are plotted in Figure 1.

    Figure 1.  Some parametric curves fitting our scheme. Left: catenary (Eq (2.6)); middle: image of the exponential curve under an inversion (Eq (2.7)); Right; 3D spiral (Eq (2.9)).

    Example 2.1. The following curves satisfy the requirements in Corollary 2.1.

    (1) Catenary. Consider the planar curve C parametrized by

    p(z)=(z,cosh(z)), (2.6)

    where cosh denotes the hyperbolic cosine (see Figure 1, left). Here p(z)=(ΦΠ)(z), with Π(z)=(z,ξ(z)), where ξ(z)=cosh(z), which is meromorphic, and Φ(x,y)=(x,y), which is clearly birational. In fact, any other planar curve p(z)=(z,ξ(z)) with ξ(z) a non-algebraic, meromorphic function also satisfies the hypotheses in Corollary 2.1.

    (2) Image of the graph of the exponential curve under an inversion. Let the planar curve C (see Figure 1, middle) be parametrized by

    p(z)=(zz2+e2z,ezz2+e2z). (2.7)

    Here p(z)=(ΦΠ)(z), with Π(z)=(z,ξ(z)), where ξ(z)=ez, which is clearly meromorphic, and

    Φ(x,y)=(xx2+y2,yx2+y2),

    which is an inversion from the origin, and therefore a birational mapping. In fact, substituting the above Φ by any other birational planar mapping (e.g., a projective or affine mapping), and the function ez by any other meromorphic function, we also get a curve satisfying our hypotheses.

    (3) 3D spiral. Consider the space curve C parametrized by (zcos(z),zsin(z),z), which is a 3D spiral (see Figure 1, right). Writing

    cos(z)=e2iz+12eiz, sin(z)=e2iz12ieiz, i2=1, (2.8)

    the parametrization of C can be expressed as

    p(z)=(ze2iz+12eiz,ze2iz12ieiz,z). (2.9)

    Thus, p(z)=(ΦΠ)(z), with Π(z)=(z,ξ(z)), where ξ(z)=eiz, which is meromorphic, and

    Φ(x,y)=(yx2+12x,yx212ix,y),

    which is a birational mapping.

    Certainly, not all parametric curves meet the hypotheses of Corollary 2.1. In order to clarify our input curves, we present two examples of such curves.

    Example 2.2. The following two curves do not satisfy the requirements in Corollary 2.1.

    (1) Cycloid. Consider the planar curve C parametrized by (zsin(z),1cos(z)). Using Eq (2.8), we can rewrite this parametrization as

    p(z)=(ze2iz12ieiz,1e2iz+12eiz).

    We can see p(z) as p(z)=(ΦΠ)(z), with Π(z)=(z,eiz), where eiz is a meromorphic function, and

    Φ(x,y)=(Φ1(x,y),Φ2(x,y))=(xy212iy,1y2+12y).

    However, Φ(x,y) is a rational mapping, but it is not birational, since a generic point (u,v)=(Φ1(x,y),Φ2(x,y)) has two pre-images (x,y); indeed, notice that imposing v=Φ2(x,y) we get two different values for y.

    (2) Spherical spiral. Consider the space curve C parametrized by

    p(z)=(cos(z)1+z2,sin(z)1+z2,z1+z2).

    Using again Eq (2.8), we can write p(z)=(ΦΠ)(z) with Π(z)=(z,eiz). However, Φ is no longer a rational function in this case, because of the presence of a square-root in the denominators.

    Therefore, we are finally ready to state the problem that we want to solve.

    Problem: Given two curves C1,C2Cn, not contained in hyperplanes, parametrized by mappings p(z),q(z) satisfying the hypotheses in Corollary 2.1, compute the affine equivalences, if any, between C1,C2.

    Notice that for these curves the diagram Eq (2.3) is commutative, and the function φ(z) at the bottom of Eq (2.3) is a Möbius function. So in order to solve the problem generalizing the approach in [7], we need to find rational invariants, in any dimension. This will be addressed in Section 3.

    In this subsection we recall two notions that we will be using later in the paper. The first one is the Schwartzian derivative: Given a holomorphic function f:CC, the Schwartzian derivative [14] S(f) of f is

    S(f)(z)=f(z)f(z)32(f(z)f(z))2.

    The Schwartzian derivative of any Möbius transformation is identically zero. The following lemma is a consequence of this.

    Lemma 2.2. Let ω:=φ(z) a Möbius transformation, and let ω(k) denote the k-th derivative of ω with respect to k. For k3,

    ω(k)=k!2k1(ω)k1(ω)k2. (2.10)

    Proof. Since the Schwartzian derivative of a Möbius transformation is identically zero, we get that

    ω=32(ω)2ω,

    which corresponds to Eq (2.10) for k=3, then the result follows by induction on k.

    The second tool is the Lah number L(k,m) (see for instance [12]),

    L(k,m)=(k1m1)k!m!,

    which allows us to define the following function, which we spell here for future reference:

    ˜Bk,m={1nkm(n+1)kmL(k,m)k>m,1k=m,0k<m. (2.11)

    We want to exploit Eq (2.2) to first find the Möbius transformation φ, if any, and then derive f from φ. If we expand Eq (2.2), we get

    Ap(z)+b=(qφ)(z). (3.1)

    Our overall strategy will consist of three steps that we will refer to as steps (ⅰ)–(ⅲ), which somehow mimic the strategy in [7], although for a completely general dimension:

    (ⅰ) Find initial invariants: We start by constructing certain functions I1,,In satisfying that Ii(p)=Ii(qφ), which are rational in the sense that they are rational functions of p and its derivatives. Since by Eq (3.1) we observe that qφ is the image of p under an affine mapping f(x)=Ax+b, we say that I1,,In are affine invariants, i.e., functions depending on a parametrization (and its derivatives) that stay the same when an affine transformation is applied.

    (ⅱ) Find Möbius-commuting invariants: Recall, from Subsection 2.1 that we say that a function F depending on a parametrization u=u(z) and its derivatives is Möbius-commuting if for any Möbius function we have

    F(uφ)=F(u)φ.

    The functions Ii found in step (ⅰ) are not, in general, Möbius-commuting. Thus, in a second step we will compute Möbius-commuting functions F1,,Fn1 from the Ii. The Fj not only satisfies that Fj(p)=Fj(qφ) for j=1,,n1, but they also satisfy that Fj(qφ)=Fj(q)φ. In turn, for j=1,,n1 we have

    Fj(p)=Fj(q)φ.

    Notice that while we have n initial invariants Ii, we have n1 Möbius-commuting invariants. Furthermore, the Fj will also be rational invariants.

    (ⅲ) Compute φ using bivariate factoring, and derive f from φ: Setting ω:=φ(z), the equalities Fj(p)=Fj(q)φ, after clearing denominators, are translated into n1 conditions Mj(z,ω)=0, with j=1,,n1. The Möbius function φ corresponds to a common factor of all the Mj, and the affine equivalence itself, f(x)=Ax+b, follows from Eq (3.1).

    In this subsection we will present step (ⅰ); the remaining steps will be described in the next subsections. Also, in the rest of the paper we will use the notation [w1,,wn] for an n×n matrix whose columns are w1,,wnCn, and w1,,wn for the determinant of the matrix [w1,,wn].

    The description of step (ⅰ) is analogous to Section 3.2 in [7]. Thus, here we focus on the main ideas, and refer the interested reader to [7] for details and proofs. Going back to Eq (2.2), let us write u:=p(z), v:=(qφ)(z), so that Eq (3.1) becomes simply Au+b=v. Repeatedly differentiating this equation with respect to z yields AD(u)=D(v) where

    D(u)=[u,u,,u(n)], D(v)=[v,v,,v(n)],

    i.e., D(u),D(v) are matrices whose columns consist of the first n derivatives of u,v with respect to z. Whenever p,q and u,v are not contained in hyperplanes, D(u),D(v) are invertible [18]. Thus, we can write A=D(v)(D(u))1. Differentiating this equality with respect to z and taking into account that A is a constant matrix, we get that

    d(D(v)(D(u))1)dz=0.

    Expanding the derivative in the left-hand side of the above equation, we arrive at

    (D(u))1dD(u)dz=(D(v))1dD(v)dz. (3.2)

    Denoting

    U=(D(u))1dD(u)dz, V=(D(v))1dD(v)dz,

    one can check that

    U=[000u(n+1),u,,u(n)u,u,,u(n)010u,u(n+1),,u(n)u,u,,u(n)001u,u,,u(n+1)u,u,,u(n)],V=[000v(n+1),v,,v(n)v,v,,v(n)010v,v(n+1),,v(n)v,v,,v(n)001v,v,,v(n+1)v,v,,v(n)]. (3.3)

    Next, let us define

    Ai(u):=u,,u(i1),u(n+1),ui+1,,u(n), Δ(u):=u,u,,un. (3.4)

    Thus, Ai(u) is the result of replacing u(i) in uuu(n) by u(n+1). Finally, for i=1,,n, let

    Ii(u):=Ai(u)Δ(u), (3.5)

    which corresponds to the entries of the last column of U; notice that whenever u,v are not contained in hyperplanes Δ(u) is not identically zero [18], so the Ii are well-defined.

    By Eq (3.2), U,V are equal and therefore their last columns coincide. Thus, Ii(u)=Ii(v) for i=1,,n, i.e., Ii(p)=Ii(qφ), which, by Theorem 2.1, is a necessary condition for affine equivalence. The following result, analogous to Theorem 7 in [7] and which can be proved in a similar way using Corollary 2.1 shows that this condition is also sufficient.

    Theorem 3.1. Let C1,C2Cn be two curves, not contained in a hyperplane, parametrized by mappings p,q satisfying the hypotheses in Corollary 2.1. If C1,C2 are affinely equivalent, then there exists a Möbius transformation φ such that

    Ii(p)=Ii(qφ) (3.6)

    for i=1,,n.

    In order to carry out step (ⅱ), which will be addressed in the next subsection, we need an auxiliary invariant, I0, defined as

    I0(u):=u,,u(n1),u(n+2)u,u,,u(n). (3.7)

    The following lemma proves that I0 lies in the differential field spanned by I1,,In.

    Lemma 3.1. I0=dIndz+In1+I2n.

    Proof. Differentiating In we get

    dIn(u)dz=(u,,u(n2),u(n),u(n+1)+u,,u(n1),u(n+2))u,,u(n)u,,u(n)2u,,u(n1),u(n+1)u,,u(n1),u(n+1)u,,u(n)2=u,,u(n2),u(n+1)u(n)u,,u(n)+u,,u(n1),u(n+2)u,,u(n)u,,u(n1),u(n+1)2u,,u(n)2=In1+I0I2n. (3.8)

    Isolating I0 from the above equality, we get I0=dIndz+In1+I2n.

    Since, according to Lemma 3.1, I0 is generated by I1,,In, the result in Theorem 3.1 also holds when we add I0 to the list of the Iis.

    Corollary 3.1. Let C1,C2Cn be two curves, not contained in a hyperplane, parametrized by mappings p,q satisfying the hypotheses in Corollary 2.1. If C1,C2 are affinely equivalent then there exists a Möbius transformation φ such that

    Ii(p)=Ii(qφ) (3.9)

    for i{0,1,n}.

    The Ii developed in the previous section are not Möbius-commuting, i.e., Ii(qφ)Ii(q)φ; in other words, calling ω:=φ(z), Ii(q(ω))Ii(q)(ω). For instance, in the case n=3, expanding Ii(q(ω)) for i=1,2,3 we get that

    ω3I1(q(ω))=3ω3+32ω2ω2I3(q)(ω)ω4ωI2(q)(ω)+ω6I1(q)(ω)ω2I2(q(ω))=9ω2+ω4I2(q)(ω)3ω2ωI3(q)(ω)ωI3(q(ω))=6ω+ω2I3(q)(ω), (3.10)

    where ω,ω are the first and second derivatives of ω=φ(z) with respect to z; to produce these equalities, we have taken into account the definition of I1,I2,I3 as quotients of determinants, the chain rule, and the fact that, because of Eq (2.10) in Lemma 2.2, the derivatives of ω of order higher than 2 can be written in terms of ω,ω. However, by eliminating ω,ω in Eq (3.10), one can show that

    [36I1(q(ω))+6I2(q(ω))I3(q(ω))+I3(q(ω))3]2[4I2(q(ω))+I3(q(ω))2]3=[36I1(q)(ω)+6I2(q)(ω)I3(q)(ω)+I3(q)(ω)3]2[4I2(q)(ω)+I3(q)(ω)2]3, (3.11)

    so that

    F=[36I1(q)+6I2(q)I3(q)+I33(q)]2[4I2(q)+I23(q)]3 (3.12)

    is Möbius-commuting, i.e., F(qφ)=F(q)φ.

    One can certainly manipulate Eq (3.10) by hand to get rid of ω,ω, reach Eq (3.11), and therefore find the invariant in Eq (3.12). However, we want to produce invariants like the one in Eq (3.12) in an algorithmic fashion, and for any dimension: that is the task in step (ⅱ). The rough idea, as in Eq (3.10), is to get rid of the derivatives ω(k), k=1,2,,n+2, in the system consisting of the expressions

    Ii(q(ω))=ξi(I0(q)(ω),,In(q)(ω),ω,ω,,ω(n+2)), (3.13)

    where ξi is the result of expanding Ii(q(ω)), with i=0,1,,n. In fact, because of Eq (2.10) in Lemma 2.2, the left-hand side only depends on ω,ω.

    The most difficult part is to provide an explicit expression for the righthand side of Eq (3.13). This is a long, technical process involving far from trivial combinatorial questions, so we will skip the details here, and refer the interested reader to the ArXiv version of this paper [3] for a complete deduction. We will just point out that one can write ω at the righthand side of Eq (3.13) in terms of the powers of ω, In(q(ω)) and In(q)(ω) (see Lemma 11 of [3]), and that it is the final elimination of the powers of ω in the resulting equations that yield the Möbius-commuting invariants (see Section 4.3 of [3]). In order to introduce these invariants, we recall the function ˜Bk,m defined in Eq (2.11) (see Subsection 2.3), and denote by Mn+1,i the (n+2)×(n+2) determinant satisfying that:

    ● If j<i, the j-th column of Mn+1,i is ˜Bn+1,n+1j,˜Bn,n+1j,,˜B,n+1j.

    ● If ji, the j-th column of Mn+1,i is ˜Bn+1,nj,˜Bn,nj,,˜B,nj.

    See Section 4.3 of [3] for the motivation for introducing this determinant. Additionally, let

    F1(q):=I0(q)12n+2nI2n(q)In1(q)+Mn+1,n2I2n(q). (3.14)

    We have the following theorem (see Section 4.3 of [3] for a proof).

    Theorem 3.2. Let F1(q) be the expression in Eq (3.14), and for k=3,,n, let

    Fk1(q):=(ki=0Mn+1,nk+2iInk+1+i(q)Iin(q))ek/k(In1(q)+Mn+1,n2I2n(q))ek/2,3kn, (3.15)

    where ek is, for 3kn, the least common multiple of 2,k, i.e., ek=lcm(2,k). The F, for =1,,n1, are Möbius-commuting.

    The generation of the Möbius-commuting invariants, for any dimension n, is implemented in [8], which can be freely downloaded, and can be done just once for each dimension n. In Table 1, we spell the invariants for low dimension, 2n4.

    Table 1.  Möbius-commuting invariants for low dimension.
    n Möbius-commuting invariants
    2 F1=I0I226I1+I22
    3 F1=6I05I234I2+I23 F2=(36I1+6I2I3+I33)2(4I2+I23)3
    4 F1=4I03I2410I3+3I24 F2=(50I2+15I3I4+3I34)2(10I3+3I24)3 F3=4000I1+400I2I4+60I3I24+9I44(10I3+3I24)2

     | Show Table
    DownLoad: CSV

    Next, let us address step (ⅲ). Let Fj be a Möbius-commuting invariant, j{1,,n1}. Since Fj is a rational function of the Ii, Fj is also an affine invariant, i.e., from Theorem 3.1 we get that Fj(p)=Fj(qφ). Therefore, in terms of the variables z and ω:=φ(z), and taking into account that Fj(qφ)=Fj(q)φ, we deduce that Fj(p)(z)=Fj(q)(ω), thus we have the following result.

    Proposition 3.1. Let C1,C2Cn be two curves, not contained in a hyperplane, parametrized by mappings p,q satisfying the hypotheses in Corollary 2.1. C1,C2 are affinely equivalent if and only if there exists a Möbius transformation φ such that

    Fj(p)(z)Fj(q)(ω)=0, (3.16)

    for j{1,2,,n1} with ω=φ(z), such that D(qφ)(D(p))1(z) is a constant matrix A and b=(qφAp)(z) is a constant vector. Furthermore, f(x)=Ax+b is an affine equivalence between C1,C2.

    Proof. () Let f be an affine equivalence between C1,C2. By Theorem 2.1, there exists a Möbius function φ such that fp=qφ. By Corollary 3.1 we have that Ii(p)(z)=Ii(q(ω)) for all i{0,,n}. Since the Fj are rational functions of the Ii, Ii(p)(z)=Ii(q(ω)) yields Fj(p)(z)=Fi(q)(ω) for j{1,2,,n1}. Finally, writing f(x)=Ax+b, the condition fp=qφ implies that Ap(z)+b=q(φ(z)), so b=(qφAp)(z), which is a constant vector. Furthermore, by differentiating the condition Ap(z)+b=q(φ(z)) (see Subsection 3.1) we deduce that A=D(qφ)(D(p))1(z). () Let φ be a Möbius transformation satisfying Fi(p(z))Fi(q)(ω)=0 for ω=φ(z). If A=D(qφ)(D(p))1(z) is a constant matrix, then D(Ap)(z)=D(qφ)(z), so Ap(z)(qφ)(z) is a constant equal to b. Therefore, Ap(z)+b=q(φ(z)). However this equality implies that Ap(z)+b, which is the image of C1 under the affine mapping f(x)=Ax+b, and q(z) parametrize the same curve, namely, C2. Thus, f(x)=Ax+b is an affine equivalence between C1 and C2.

    To finally turn Proposition 3.1 into an algorithm, let Mj(z,ω) be obtained by clearing denominators in Fj(p)(z)Fj(q)(ω). We need to request that Mj(z,ω) is not identically zero, which amounts to requiring that not all the Fj are constant; although this is rare, it can happen: two examples are conic planar curves, and helices, i.e., space curves where the quotient between curvature and torsion is constant (including circular helices). If Mj(z,ω) is not zero, then Mj(z,ω)=0 defines an analytic curve in the plane z,ω. Now if φ(z), as in Eq (2.1), is a Möbius function satisfying Proposition 3.1, calling ω=φ(z) we get that all the points (z,ω) of the curve

    ω(cz+d)(az+d)=0,

    which is an irreducible analytic curve, are also points of the curve Mj(z,ω). As a consequence of Study's Lemma (see Section 6.13 of[6]), H(z,ω)=ω(cz+d)(az+d) must be a factor of Mj(z,ω); we say that H(z,ω)=ω(cz+d)(az+d) is a Möbius-like factor of Mj(z,ω), and that the Möbius function φ in Eq (2.1) is associated with H(z,ω). So we have the following theorem, which follows from Proposition 3.1.

    Theorem 4.1. Let C1,C2Cn be two curves, not contained in a hyperplane, parametrized by mappings p,q satisfying the hypotheses in Corollary 2.1 and where not all the Fj are constant. C1,C2 are affinely equivalent if and only if there exists a Möbius-like factor H(z,ω) common to Mj(z,ω), j=1,,n1 such that the corresponding associated Möbius function φ satisfies that: (1) D(qφ)(D(p))1(z) is a constant matrix A, (2) b=(qφAp)(z) is a constant vector. Furthermore, in that case, f(x)=Ax+b is an affine equivalence between C1,C2.

    Thus, we get the following procedure AffineEquivalences to find the affine equivalences between the curves C1,C2 defined by p,q.

    AffineEquivalences
    Input: Two parametrizations p and q satisfying the hypotheses in Corollary 2.1.
    Output: Either the list of affine equivalences between the curves, or the warning The curves are not affinely equivalent
       1: procedure AffEq(p,q)
       2:       Compute Mj(x,z), j=1,,n1, by clearing denominators in Fj(p)(z)Fj(q)(ω).
       3:       if all the Mj are identically zero then
       4:             return Failure: allvthe Möbiuscommuting invariants are constant
       5:       else
       6:             Compute the common factor L(x,z) of the Mj(x,z).
       7:             Let L be the list of Möbius-like factors of L(x,z)
       8:             if L= then
       9:                   return The curves are not affinely equivalent
       10:             else
       11:                   for φL do
       12:                         Check whether or not A=D(q(φ))D(p)1, b=qφAp are constant
       13:                         In the affirmative case, return f(x)=Ax+b.

    If p,q are rational, the Mj(x,z) are rational and H(z,ω) is a factor of gcd(M1(z,ω),,Mn1(z,ω)). However, the computer algebra system Maple [13] where we implemented the procedure (see [8]) can compute H(z,ω) also in the case when p,q are not rational, but satisfies the hypotheses of the procedure. In this last case, we ask Maple to solve H(z,ω) for ω to find the Möbius functions.

    Remark 4.1. Although Maple HelpSystem is not too specific about this, in the case when the Mj(x,z) are not rational the idea seems to be that Maple renames repeated non-rational expressions found in the Mj(x,z) (e.g., cos(z),ez, etc.) to form rational functions, and then proceeds by applying the algorithm for the rational case. Furthermore, in the case of non-rational parametrizations we have considered examples where the adjoined function ξ(z) (see Section 2.2) is the same for both p,q, since it is not guaranteed that Maple can solve H(z,ω) for ω otherwise.

    In order to illustrate the performance of the procedure AffineEquivalences, we consider now two examples where we compute the affine equivalences between curves taken from Example 2.1, and the images of these curves under an affine mapping. These examples were computed with Maple and executed in a PC with a 3.60 GHz Intel Core i7 processor and 32 GB RAM, and are accessible in [8] as well.

    Example 4.1. (2D catenary curves) Consider the curves C1 and C2 parametrized by

    p(z)=(2zcosh(2z)+14z+cosh(2z)),q(z)=(zcosh(z)).

    The curve q(z) corresponds to the first curve in Example 2.1, which is a catenary curve. After appying our algorithm, we find two factors ˆHi,(z,ω), i=1,2, common to the Mj, namely

    ˆH1(z,ω)=cosh(ω)sinh(2z)cosh(2z)sinh(ω), ˆH2(z,ω)=cosh(ω)sinh(2z)+cosh(2z)sinh(ω).

    When solving for ω, we get infinitely many (complex) Möbius functions leading to infinitely many (complex) affine equivalences, which reveals that the ˆHi(ω,z) contain Möbius-like factors. The affine equivalences can be classified in three classes fj(x)=Ajx+bj, j{1,2,3}, with associated Möbius functions φj(z):

    A1=(1313(1)k1+123(1)k113),b1=(13+ik1π(1)k123),φ1(z)=2z+ik1π,k1Z,
    A2=(13132313),b2=(13+2ik2π23),φ2(z)=2z+ik2π,k2Z,

    and

    A3=(13132313),b3=(13+(2k2+1)iπ23),φ3(z)=2z+(2k2+1)iπ,k2Z,

    where i2=1. If we just consider real affine equivalences, we have three of them, which correspond to fixing k1=0 for f1(x), k2=0 for f2(x), and k2=1/2 for f3(x). The whole computation took 0.172 seconds.

    Example 4.2. (3D spirals) Consider the curves C1 and C2 parametrized by

    p(z)=(ze4iz+1e2izize4iz1e2iz+12ze4iz+1e2izize4iz1e2iz2z2z1),q(z)=(ze2iz+12eizize2iz12eizz).

    The curve q(z) corresponds to the third curve in Example 2.1, which is a 3D spiral. After applying our algorithm, we find two Möbius-like factors Hi(z,ω), i=1,2, common to the Mj(z,ω), namely

    H1(z,ω)=ω2z, H2(z,ω)=ω+2z.

    When solving for ω, we get two Möbius transformations φ1(z)=2z and φ2(z)=2z corresponding to the affine equivalences f1(x)=A1x+b1 and f2(x)=A2x+b2 with

    A1=(111211001),b1=(011),

    and

    A2=(111211001),b1=(011).

    The whole computation took 0.032 seconds.

    Example 4.3. (Rational curves in n-th dimension) Finally, in Table 2, we present the results of performance tests to compute affine equivalences between rational curves of various degrees in different dimensions. The rational curves in the experiments were randomly generated [8] with coefficients between 10 and 10. After generating the first curve, the second curve was obtained by applying an affine mapping f(x)=Ax+b to the first curve, where the matrix and the translation vector, for each dimension, are shown in Table 3; additionally, the resulting curve was reparametrized using a Möbius transformation φ(z)=2z1. The timings to recover the affine equivalences are shown in Table 3: the rows of Table 3 correspond to dimensions from n=2 to n=6, and the columns, to degrees from d=6 to d=12. For degrees up to 10, we can compute the affine equivalences between the curves in less than a minute, for all the dimensions tested.

    Table 2.  CPU time in seconds for affine equivalences of random rational curves with various degrees in various dimensions.
    Degree
    n 6 7 8 9 10 11 12
    2 0.109 0.188 0.125 0.203 0.453 0.750 0.532
    3 0.969 1.969 3.750 6.406 8.579 12.281 15.703
    4 1.343 2.063 4.359 7.453 12.531 17.688 30.234
    5 2.813 6.047 14.000 28.609 48.406 89.203 138.546
    6 0.922 6.281 12.203 26.609 51.328 90.344 153.719

     | Show Table
    DownLoad: CSV
    Table 3.  Affine mappings used in the examples.
    n A b
    2 (1120) (01)
    3 (112203004) (010)
    4 (1121203000410102) (0100)
    5 (1121320301004130041001021) (01000)
    6 (112130203012004131004102010211012013) (010000)

     | Show Table
    DownLoad: CSV

    We have presented an algorithm, generalizing the algorithm in [7], to compute the affine equivalences, if any, between two parametric curves in any dimension. Our strategy relies on bivariate factoring, and avoids polynomial system solving. The algorithm works for rational curves and also certain types of non-algebraic parametric curves with bimeromorphic parametrizations, where we are adjoining a non-algebraic, meromorphic function ξ(z). We have implemented the algorithm in Maple, and evidence of its performance has been presented.

    The algorithm works whenever not all the Möbius-commuting invariants are constant. This happens generically, but identifying the curves where this does not occur, as well as providing a solution to the problem for this special case, are questions that we pose here as open problems.

    Additionally, in the case of non-algebraic curves, right now we need some hypotheses that are not always satisfied: for instance, planar curves like the cycloid, or the tractrix, or classical planar spirals, do not satisfy our hypotheses. However, we have observed that the algorithm seems to work also for many of those curves, which makes us think that our hypotheses could be relaxed. This requires more theoretical work regarding analytic curves.

    It would be desirable to extend our ideas to the case of rational surfaces/hypersurfaces. This probably requires some extra hypotheses, e.g., nonexistence of base points or special types of surfaces/hypersurfaces, that allow us to guess the type of transformation that we have in the parameter space: such transformation would play a role similar to the role played by Möbius transformations here. These are questions that we would like to address in the future.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    Juan Gerardo Alcázar is supported by the grant PID2020-113192GB-I00 (Mathematical Visualization: Foundations, Algorithms and Applications) from the Spanish MICINN. Juan G. Alcázar is also a member of the Research Group asynacs (Ref. ccee2011/r34). Uğur Gözütok is supported by the grant 121C421, in the scope of 2218-National Postdoctoral Research Fellowship Program, from TUBITAK (The Scientific and Technological Research Council of Türkiye). We thank the reviewer of the paper for his/her comments, which helped to improve the first version of the paper.

    Juan Gerardo Alcázar is the Guest Editor of special issue "Computer Algebra, Geometry and Applications" for AIMS Mathematics. Juan Gerardo Alcázar was not involved in the editorial review and the decision to publish this article.

    The authors declare no conflict of interest in this paper.



    [1] J. G. Alcázar, G. Muntingh, Affine equivalences of surfaces of translation and minimal surfaces, and applications to symmetry detection and design, J. Comput. Appl. Math., 411 (2022), 114206. https://doi.org/10.1016/j.cam.2022.114206 doi: 10.1016/j.cam.2022.114206
    [2] J. G. Alcázar, C. Hermoso, G. Muntingh, Symmetry detection of rational space curves from their curvature and torsion, Comput. Aided Geom. D., 33 (2015), 51–65. https://doi.org/10.1016/j.cagd.2015.01.003 doi: 10.1016/j.cagd.2015.01.003
    [3] J. G. Alcázar, H. A. Çoban, U. Gözütok, Detecting affine equivalences between certain types of parametric curves, in any dimension, 2024. https://arXiv.org/abs/2403.16636
    [4] M. Bizzarri, M. Làvi˘cka, J. Vr˘sek, Computing projective equivalences of special algebraic varieties, J. Comput. Appl. Math., 367 (2020), 112438. https://doi.org/10.1016/j.cam.2019.112438 doi: 10.1016/j.cam.2019.112438
    [5] M. Do Carmo, Differential geometry of curves and surfaces, 1976.
    [6] G. Fischer, Plane algebraic curves, American Mathematical Society, 2001.
    [7] U. Gözütok, H. A. Çoban, Y. Sağiroğlu, J. G. Alcázar, A new method to detect projective equivalences and symmetries of 3D rational curves, J. Comput. Appl. Math., 419 (2023), 114782. https://doi.org/10.1016/j.cam.2022.114782 doi: 10.1016/j.cam.2022.114782
    [8] U. Gözütok, https://www.ugurgozutok.com/academics/software.
    [9] J. Harris, Algebraic geometry, New York: Springer Science & Business Media, 1992. https://doi.org/10.1007/978-1-4757-2189-8
    [10] M. Hauer, B. Jüttler, Projective and affine symmetries and equivalences of rational curves in arbitrary dimension, J. Symb. Comput., 87 (2018), 68–86. https://doi.org/10.1016/j.jsc.2017.05.009 doi: 10.1016/j.jsc.2017.05.009
    [11] T. Jong, G. Pfister, Local analytic geometry, In: Advanced lectures in mathematics, 2000. https://doi.org/10.1007/978-3-322-90159-0
    [12] T. Kim, D. S. Kim, L. C. Jang, H. Lee, H. Y. Kim, Complete and incomplete Bell polynomials associated with Lah-Bell numbers and polynomials, Adv. Differ. Equ., 2021 (2021), 101. https://doi.org/10.1186/s13662-021-03258-3 doi: 10.1186/s13662-021-03258-3
    [13] MapleTM, 2021. Maplesoft, a division of Waterloo Maple Inc. Waterloo, Ontario.
    [14] V. Ovsienko, S. Tabachnikov, What is the Schwarzian derivative? Not. Ame. Math. Soc., 56 (2009), 34–36. https://www.ams.org/notices/200901/tx090100034p.pdf
    [15] S. Pérez-Díaz, J. R. Sendra, Computation of the degree of rational surface parametrizations, J. Pure Appl. Algebra, 193 (2004), 99–121. https://doi.org/10.1016/j.jpaa.2004.02.011 doi: 10.1016/j.jpaa.2004.02.011
    [16] S. Pérez-Díaz, J. R. Sendra, C. Villarino, Computing the singularities of rational surfaces, Math. Comp., 84 (2015), 1991–2021. https://www.jstor.org/stable/24489185
    [17] J. R. Sendra, F. Winkler, S. Pérez-Díaz, Rational algebraic curves, Springer, 2008. https://doi.org/10.1007/978-3-540-73725-4
    [18] R. Sulanke, The fundamental theorem for curves in the n-dimensional Euclidean space, 2020.
  • Reader Comments
  • © 2024 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(1066) PDF downloads(63) Cited by(0)

Figures and Tables

Figures(1)  /  Tables(3)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog