
We consider the damped hyperbolic motion of polygons by a linear semi-discrete analogue of polyharmonic curve diffusion. We show that such flows may transition any polygon to any other polygon, reminiscent of the Yau problem of evolving one curve to another by a curvature flow, before converging exponentially to a point that, under appropriate rescaling, is a planar basis polygon. We also consider a hyperbolic linear semi-discrete flow of the Yau curvature difference flow, where a polygonal curve is able to flow to any other such that we get convergence to the target polygon in infinite time.
Citation: James McCoy, Jahne Meyer. Semi-discrete linear hyperbolic polyharmonic flows of closed polygons[J]. Mathematics in Engineering, 2025, 7(3): 281-315. doi: 10.3934/mine.2025013
[1] | Massimiliano Giona, Luigi Pucci . Hyperbolic heat/mass transport and stochastic modelling - Three simple problems. Mathematics in Engineering, 2019, 1(2): 224-251. doi: 10.3934/mine.2019.2.224 |
[2] | Yu Leng, Lampros Svolos, Dibyendu Adak, Ismael Boureima, Gianmarco Manzini, Hashem Mourad, Jeeyeon Plohr . A guide to the design of the virtual element methods for second- and fourth-order partial differential equations. Mathematics in Engineering, 2023, 5(6): 1-22. doi: 10.3934/mine.2023100 |
[3] | Paola F. Antonietti, Chiara Facciolà, Marco Verani . Unified analysis of discontinuous Galerkin approximations of flows in fractured porous media on polygonal and polyhedral grids. Mathematics in Engineering, 2020, 2(2): 340-385. doi: 10.3934/mine.2020017 |
[4] | Claudio Canuto, Davide Fassino . Higher-order adaptive virtual element methods with contraction properties. Mathematics in Engineering, 2023, 5(6): 1-33. doi: 10.3934/mine.2023101 |
[5] | Gabriele Grillo, Giulia Meglioli, Fabio Punzo . Global existence for reaction-diffusion evolution equations driven by the $ {\text{p}} $-Laplacian on manifolds. Mathematics in Engineering, 2023, 5(3): 1-38. doi: 10.3934/mine.2023070 |
[6] | La-Su Mai, Suriguga . Local well-posedness of 1D degenerate drift diffusion equation. Mathematics in Engineering, 2024, 6(1): 155-172. doi: 10.3934/mine.2024007 |
[7] | Simon Lemaire, Julien Moatti . Structure preservation in high-order hybrid discretisations of potential-driven advection-diffusion: linear and nonlinear approaches. Mathematics in Engineering, 2024, 6(1): 100-136. doi: 10.3934/mine.2024005 |
[8] | Monica Conti, Filippo Dell'Oro, Vittorino Pata . Exponential decay of a first order linear Volterra equation. Mathematics in Engineering, 2020, 2(3): 459-471. doi: 10.3934/mine.2020021 |
[9] | Juan Pablo Borthagaray, Wenbo Li, Ricardo H. Nochetto . Finite element algorithms for nonlocal minimal graphs. Mathematics in Engineering, 2022, 4(2): 1-29. doi: 10.3934/mine.2022016 |
[10] | Luca Formaggia, Alessio Fumagalli, Anna Scotti . A multi-layer reactive transport model for fractured porous media. Mathematics in Engineering, 2022, 4(1): 1-32. doi: 10.3934/mine.2022008 |
We consider the damped hyperbolic motion of polygons by a linear semi-discrete analogue of polyharmonic curve diffusion. We show that such flows may transition any polygon to any other polygon, reminiscent of the Yau problem of evolving one curve to another by a curvature flow, before converging exponentially to a point that, under appropriate rescaling, is a planar basis polygon. We also consider a hyperbolic linear semi-discrete flow of the Yau curvature difference flow, where a polygonal curve is able to flow to any other such that we get convergence to the target polygon in infinite time.
In [2], Chow and Glickenstein considered a semi-discrete analogue of the second order curve shortening flow for smooth, closed curves. Involving only a first time derivative, this may be considered a parabolic-type evolution. Recently we considered the polyharmonic analogue of this work [13]. There we also considered a discrete analogue of the Yau problem of evolving one curve to another by a curvature flow. In this article we replace the first time derivative by the second time derivative plus a first time derivative 'damping term', thus considering a hyperbolic analogue of our previous work. Our ensuing evolution equation now being second order in time requires more conditions in order to have a unique solution. We prescribe not only initial conditions but also conditions at a fixed later time, thus generating flows that transition from one polygonal curve through another and then asymptotically to a limiting shape. These flows thus flow one curve to another in the spirit of the Yau problem and then flow to a limiting polygonal curve.
Our setup is similar as in the earlier works [2,13]. Given an ordered collection of n points in Rp, joined in order to form a piecewise linear, closed immersed 'curve', the flow is given by a second order system of n coupled linear ordinary differential equations (ODEs). The constant coefficient matrix of this system is an mth order difference-type operator.
Linearity permits analysis via finite Fourier series. As in [13] and in the case with nonzero damping term, we find that the higher the order m, the faster the convergence, in the first case under appropriate rescaling, to a regular basis polygon. To our knowledge semi-discrete linear hyperbolic flows have not been considered before, but the hyperbolic analogue, with damping, of the curve shortening flow has been considered (see [5] and the references contained therein).
Fully discrete flows have been considered by C. Rademacher and H. B. Rademacher [16,17]. These are related to discretisation of partial differential equations in space and time using finite differences, while semi-discrete flows are related to the 'method of lines' where there is discretisation in all but one dimension (usually time), reducing a given system of partial differential equations to a system of ordinary differential equations. The equations we consider are closely related to discretisations via finite differences or finite elements; a particular hyperbolic reference is [5]. We would also like to mention that first order semi-discrete flows have been considered in several engineering applications. A linear curve shortening flow scheme is given in [18] to solve the rendezvous problem, where an autonomous robotic system is guided such that each robot converges to a common position. In [7,9,10], curve shortening flows are considered to solve real time path planning for robotic systems and aerial vehicles. In particular, in [9,10], a linearised semi-discrete curve shortening flow is utilised, with an adaption to take into account an ever-changing environment, with the curve shortening flow method overcoming shortcomings of other approaches. We suspect corresponding second order flows to find similar useful applications. In [21], second order systems of linear differential equations with direct reference to engineering applications are discussed, including systems with overdamping. The idea of replacing a first time derivative by a second derivative and including a first derivative term with positive damping coefficient, also appears in the heavy ball with friction system [1,6]. For this system, Attouch et al. [1] explain how the introduction of the second time derivative enables flexibility not found in the steepest descent system where only the first time derivative appears. Additional flexibility is also reflected in our case: rather than only the initial polygonal state determining the evolution of the polygon, we have the ability for an additional intermediate polygonal state to be prescribed at a fixed later time. A reference with related application is [8] which discusses multi-robotic system path planning where robots are required to undertake multiple jobs (where a job is represented by a robot position) within a single trip.
Settings in which curvature flows result in linear ordinary or partial differential equations are quite rare. Apart from the present setting, the main other setting where flow equations are linear is where evolving convex curves and hypersurfaces are expressed via the Gauss map parametrisation and flow speeds have a specific structure. The linear structure was exploited by Smoczyk for curves and hypersurfaces evolving by parabolic equations in [19]; the curve results were extended by the first author, Schrader and Wheeler in [15] to higher order equations and curves with general nonzero rotation number. The first author and Otuf considered linear hyperbolic flows of curves in [14], while the first author considered convex hypersurfaces evolving by analogous second order linear hyperbolic flows in [12].
The structure of this article is as follows: In Section 2 we set up the necessary notation, introduce the semi-discrete linear hyperbolic polyharmonic flows and uncover some fundamental properties of these flows. In Section 3 we provide background for us to be able to state and prove our main results, including key properties of circulant matrices and the differential operators on which our flows are based. In Section 4 we specialise to evolving planar polygons, finding a representation formula for general solutions for the flow that allows us to deduce long-time behaviour. In Section 5, we consider evolving polygonal curves in general codimension. Finally, in Section 6, we consider a nonhomogeneous version of the flow which addresses the discrete analogue of Yau's curvature difference flow.
Definition 2.1. An n sided polygon, or n−gon, →X, is defined as a collection of ordered points →X=(X0,X1,…,Xn−1)T. For j=0,1,…,n−1, each vertex Xj∈Rp for some integer p≥2. We set Xn=X0 so the polygon is closed and the indices of the vertices are considered modulo n.
The vertices of the polygon are the points Xj, j=0,1,,…,n−1. The edges of the polygon are the line segments ¯XjXj+1 joining each vertex Xj to Xj+1. In general, →X is a n×p matrix with real entries. When p=2, the polygon lies in the plane. In this case we may consider each vertex, Xj=(xj,yj)∈R2, to be a point in C, writing Xj=xj+iyj. Thus →X as an n×2 matrix representing a polygon in R2 can also be thought of as a vector in Cn.
Remark 2.1. (1) We can consider the polygonal image of →X where we have each ordered vertex pair Xj and Xj+1 joined by a line segment. If the line segments have crossings, then we say the polygonal curve is a (piecewise smooth) immersion, and if there are no edge crossings the polygon is embedded (i.e. separates the plane into a bounded 'interior' and unbounded 'exterior'). A given polygonal image may be described by different 'polygons' where a different first vertex or a reversed ordering of the vertices may be chosen.
(2) Included in Definition 2.1 are degenerate cases where the collection of points may have repetitions. Our polygonal image of →X may contain duplicated vertices and overlapping edges. In our setting we have basis polygons that are degenerate and also multiply covered, which we discuss further in Section 3.
A normal to each vertex Xj is given by
Nj=(Xj+1−Xj)+(Xj−1−Xj), | (2.1) |
such that the corresponding system of equations for each normal vector can be expressed in matrix form as
→N=M→X, | (2.2) |
where M is the n×n matrix given by
M=[−210⋯011−210⋯001−210⋮⋮0⋱⋱⋱00⋱01−2110⋯01−2]. | (2.3) |
The length of the normal Nj plays the role of the curvature scalar, and the direction is the direction of the 'curvature vector'.
Definition 2.2. Let →X(t) be a family of polygons as given in Definition 2.1. Given m∈N and a constant δ≥0, polygons →X(t) satisfying
d2→Xdt2+δd→Xdt=(−1)m+1Mm→X(SHPFm) |
evolve by the 2mth order semi-discrete linear hyperbolic polyharmonic flow, where M is the matrix given in (2.3).
Remark 2.2. (1) The elements of M in (2.3) are the coefficients in finite difference approximations for the 'second spatial derivatives' associated with →X, as in (2.1). They are related to approximations of the second order derivative using Taylor polynomials where we have
f″(x)(Δx)2≈f(x+Δx)−2f(x)+f(x−Δx), | (2.4) |
for a twice differentiable function f of a real variable. Powers of M then yield higher finite differences, consistent with those obtained for example by Newton's divided difference approach. Elements of Mm correspond to 2mth order derivative approximations by Taylor polynomials. In our case, the Δx of (2.4) relates to the difference between the vertex labels. That is, x±kΔx coincides with vertex index label i±k. The coefficients within the Taylor polynomial approximations relate to the coefficients of polygon vertices in the expression (−1)m+1Mm→X, however we note that depending on the values of n and m and the modulo n nature of the vertex indices, the coefficient of a vertex in (−1)m+1Mm→X may end up being the sum of coefficients as found in the higher order approximations.
(2) We have restricted to δ≥0 as we wish to allow only no damping (δ=0) or damping in the traditional sense of a mechanical or electrical system. One could also consider δ<0 above but this is less relevant for our applications. In the case of negative damping factor, polygons expand under (SHPFm) such that given the same conditions as for the δ>0 case, under appropriate rescaling we have asymptotic convergence to an affine transformation of a regular polygon as t increases. Otherwise the polygon exhibits oscillating behaviour as the polygon expands.
(3) The case m=0 of (SHPFm) is
d2→Xdt2+δd→Xdt=−→X. |
Solution behaviour will depend on the value of δ. If, for example, δ=0 the general solution is
→X(t)=(cost)→Y+(sint)→Z | (2.5) |
for arbitrary polygons →Y and →Z. Polygons →Y and →Z can then be determined for given initial data (or more general data). In particular, for this δ we observe that solutions are breathers, that is, periodic in time. They are also ancient (can be extended back to t→−∞) and eternal (exist for all time). As a specific example, the solution with →X(0)=X0 and →X′(0)=0 (the 'trivial' polygon with all points at the origin) is
→X(t)=(cost)X0. |
On the other hand, if we restrict the general solution (2.5) to the time interval [0,π2] we have a transition from →X(0)=→Y to →X(π2)=→Z, a kind of finite, exact solution to the Yau problem in this setting using a second order flow. Extending the time interval, the solution is breathing between states →Y and →Z.
Alternatively, for 0<δ<2, the general solution has the form
→X(t)=e−δ2t[cos(γt)→Y+sin(γt)→Z] |
where γ=√1−(δ2)2 and we observe solutions again exist for all time but →X(t)→0 as t→∞. The rescaled solutions eδ2→X(t) are exactly
eδ2t→X(t)=cos(γt)→Y+sin(γt)→Z. |
Thus the solutions exhibit asymptotic breathing behaviour as t→∞ and under the same rescaling the ancient solution emanates from an asymptotically breathing solution between the states →Y and →Z.
In the case δ=2 the solution has the form
→X(t)=(1−t)e−t→Y+te(1−t)→Z. |
Again solutions make sense for all time and →X(t)→0 as t→∞. Observe that rescaled solutions satisfy
1tet→X(t)=(1t−1)→Y+e→Z→−→Y+e→Z, |
that is, under this rescaling they approach a precise linear combination of the given data →Y and →Z. One can argue similarly in considering 'where ancient solutions come from'.
Finally for δ>2 the solution has the form
→X(t)=er−t→Y+er+t→Z, |
where r±=−δ2±√(δ2)2−1 are both negative. So in this case solutions again decay to zero but there are no oscillations. Rescaling we observe in this case
e−r+t→X(t)→→Z. |
For the remainder of this article we consider m∈N, the strictly positive integers.
(4) The system of ODEs (SHPFm) is second order in time and a natural approach to working with it would be to set →Y=d→Xdt and consider the associated first order system of 2n equations. This leads to the following first order system to solve,
[d→Xdtd→Ydt]=[0n×nIn(−1)m+1Mm−δIn][→X→Y]. | (2.6) |
Approaches to solving systems of linear second order equations, especially within the context of engineering applications, are given in [21] who consider the following system,
M¨q(t)+C˙q(t)+Kq(t)=f(t), | (2.7) |
for n×n matrices M,K and damping matrix C, and n-vectors q(t) and f(t). When C=0n×n there is no damping, while if C satisfies a certain orthogonality condition it is called proportional damping and (2.7) can be solved as a second order system using what [21] call the modal superposition method. In our case, the damping matrix corresponds to δIn which is proportionally damped. When the system is non-proportionally damped, [21] provide a reformulation into a first order system which as related to our system of equations leads to (2.6). Our second order system of equations can be solved directly which is the approach we take (see Sections 4.2 and 5.2). However the approaches outlined in [21] may be useful in other applications, such as when the damping term varies between each equation of the system.
(5) As (SHPFm) is a homogeneous system of linear differential equations with a constant coefficient matrix, existence of a unique solution in a neighbourhood of any initial data is completely standard as one can write down an explicit formula for the solution. Moreover the formula for the solution reveals that the solution exists for all time, given any initial polygon and indeed ancient solutions also make sense for any initial polygon. These are properties that usually do not hold in full generality for other geometric flows and generally require much work to prove, whether they hold in general or under particular conditions.
In this section we detail the setup of the semi-discrete linear hyperbolic polyharmonic flow and discuss some of its properties. Some of this material is similarly described in [2,13].
The matrix M given in (2.3) is a circulant matrix. A general circulant matrix B has the form
B=[b0b1b2⋯bn−2bn−1bn−1b0b1⋯bn−3bn−2bn−2bn−1b0⋯bn−4bn−3⋮⋮⋮⋯⋮⋮b2b3b4⋯b0b1b1b2b3⋯bn−1b0], |
where each row is produced by shifting each of the elements of the previous row to the right. A matrix of this form is also denoted B=circ(b0,b1,…,bn−1). Many properties of circulant matrices are detailed in [4] with further details about their eigenvalues and eigenvectors in [20]. The product of circulant matrices is circulant, therefore Mm is circulant for any m∈N. For a characterisation of the elements of Mm we refer to [13,Lemmas 3.2 and 3.3].
Our next result is an extension of [13,Lemma 3.5] to include the second time derivative.
Lemma 3.1. Let a vector in Rn with all entries equal to the same constant c be denoted by →c. Also let E denote a p×p matrix. We have the following properties, where →X is a solution to (SHPFm):
(1) Mm→1=→0, and for any ai∈R for i=1,2,…,p,
(2) ddt(→XE+(→a1,→a2,…,→ap))=(−1)m+1Mm(→XE+(→a1,→a2,…,→ap)) and
(3) d2dt2(→XE+(→a1,→a2,…,→ap))=M2m(→XE+(→a1,→a2,…,→ap)).
Proof. The first two parts above are proved in [13]. The third follows by a simple calculation.
Remark 3.1. As remarked in [13], the polygon →X can be considered as a graph G={V,E} where V={X0,…,Xn−1} are the vertices of the polygon, and E is the set of edges between consecutive vertices and the degree of each vertex is 2. Therefore the matrix L=−M is a Laplacian matrix which has a corresponding definition as a Laplacian operator [3,22]. We have −Lm=(−1)m+1Mm and the semi-discrete polyharmonic flow can be associated with a higher order linear hyperbolic flow for graphs.
The eigenvectors of any n×n circulant matrix are in Cn and given by,
Pk=(1,ωk,ω2k,…,ω(n−1)k)T, for k=0,1,…,n−1, | (3.1) |
where ω=e2πi/n. Here powers of ω are the nth roots of unity. Each Pk may be thought of as a polygon by placing the entries of Pk into C as the vertices of the polygon, and joining consecutive entries by arrows. Specifically, P0=(1,1,…,1)T is a point, and the remaining Pk are either regular convex polygons, non-convex regular star polygons, or a line interval in the case of n even. By regular we mean polygons whose symmetry group is the dihedral group. Each pair Pk and Pn−k comprises the same regular polygon, but with the arrows in the opposite orientation. The exception is when n is even then Pn2 in isolation is a line interval where the arrows between points overlap each other n2 times. For each n, the collection {Pk}n−1k=1 contains the regular embedded polygons with n sides, regular immersed star polygons with n sides where all vertices of Pk are distinct, degenerate cases where we have multiply covered regular embedded and immersed polygons whose number of sides is a factor of n, as well as the reversed orientation of these polygons. Also included for n even is the multiply covered line segment Pn2.
Proposition 3.2 ([2]). The set of eigenvectors, {1√nPk}n−1k=0 of M forms an orthonormal basis for Cn where the corresponding eigenvalue, λk, for each Pk is given by
λk=−4sin2(πkn). |
Importantly for us, denoting by λm,k the eigenvalues of matrix (−1)m+1Mm corresponding to eigenvectors Pk, we have
λm,k:=(−1)m+1λmk=−4m[sin2(πkn)]m. | (3.2) |
Observe that all eigenvalues are negative, except for λm,0.
The eigenvalues of a circulant matrix occur in conjugate pairs with some exceptions. That is, λk=¯λn−k with the exception of λ0 and λn2 for n even [4,20]. Furthermore, if the circulant matrix is real and symmetric, the eigenvalues are real and thus each eigenvalue pair λk=λn−k. This holds for (−1)m+1Mm for each m and so we have λm,0=0,λm,k=λm,n−k and λm,n/2 does not have an equal pair for n even. Also λm,k<λm,1<0 for all k=2,…,⌊n2⌋.
Lemma 3.3. Given a vector →x∈Rn, if
(−1)m+1Mm→x=→c, | (3.3) |
where →c=(c,c,…,c)T for a constant c, then c=0.
Proof. We use the diagonalisation of the matrix (−1)m+1Mm which is given by
(−1)m+1Mm=1nFdiag(λm,k)ˉF, | (3.4) |
where F is the Fourier matrix,
F=[P0P1⋯Pn−1]=[111⋯11ω1ω2⋯ωn−11ω2ω4⋯ω2(n−1)⋮⋮⋮⋱⋮1ωn−1ω2(n−1)⋯ω(n−1)(n−1)], | (3.5) |
and diag(λm,k) is the diagonal matrix with diagonal entries given by the eigenvalues λm,0,λm,1,…,λm,n−1. The matrix ˉF in (3.4) is the complex conjugate of the matrix F, where F−1=1nˉF.
Rearranging (3.3) gives
diag(λm,k)ˉF→x=ˉF→c. |
Considering the first entry of each side in this equation gives λm,0∑n−1j=0xj=nc and, given λm,0=0, we conclude c=0.
We complete this section with a restatement of [13,Lemma 3.7] which will be needed in the next section. The result characterises the solutions →x in Lemma 3.3 as just the vectors with all entries equal.
Lemma 3.4. Given a vector →x∈Rn, if (−1)m+1Mm→x=→0 then →x is a constant vector with all entries equal. That is →x=→c=(c,c,…,c)T for a constant c.
In this section we consider planar solutions of (SHPFm). First, in Section 4.1, we consider self similar solutions, then, in Section 4.2 we consider solutions with general initial polygon →X(0)=→X0. Because (SHPFm) is second order in time, specifying only →X(0)=→X0 does not yield a unique solution in general.
Here we are interested in self-similar solutions to (SHPFm), that is, solutions →X(t) related to the initial polygon →X(0)=→X0 via the formula
→X(t)=g(t)→X0R(f(t))+→h(t), | (4.1) |
where g,f:R→R and →h:R→Mn×2(R) are twice-differentiable functions. Here g(t) represents scaling (g(t)>0 for all t∈[0,T)), R(f(t)) is the 2×2 rotation matrix by angle f(t), and →h(t) is a n×2 matrix corresponding to translation where →h(t)=(→h1(t) →h2(t)) and h1,h2:R→R. We have g(0)=1, f(0)=0, and →h(0)=0n×2, where 0n×2 is the n×2 zero matrix, such that →X(0)=→X0. Note that when considering →X as a polygon in C, Eq (4.1) is instead written
→X(t)=g(t)eif(t)→X0+→h(t), |
where rotation is given by eif(t) and translation by →h(t)∈Cn.
Proposition 4.1. If a family of polygons in the plane →X(t) is a self-similar solution to the flow (SHPFm) by scaling, then →X(t) has the form
→X(t)=g(t)(c1Pk+c2Pn−k) | (4.2) |
for some fixed k∈{0,1,…,⌊n2⌋}, where
g(t)={ct+1, for δ=0 and k=0.cδ+[1−cδ]e−δt, for δ>0 and k=0.e−δ2t[cos(γm,kt)+csin(γm,kt)], for 0≤δ<2√|λm,k| and k∈{1,…,⌊n2⌋} . (1+ct)e−δ2t, for δ=2√|λm,k| and k∈{1,…,⌊n2⌋}.er+t+c(er−t−er+t), for δ>2√|λm,k| and k∈{1,…,⌊n2⌋}. |
Here
r±=−δ2±√(δ2)2+λm,k, γm,k=√|(δ2)2+λm,k| |
and c∈R,c1,c2∈C are constants. In the case n is even and k=n2, we have →X(t)=c1g(t)Pn2.
We observe from above different behaviour corresponding to the zero eigenvalue λm,0. In this case, if δ=0 the point c1P0+c2Pn can be moving (c≠0) or stationary (c=0). Proposition 4.1 shows that each regular polygon Pk is a scaling self-similar solution of (SHPFm). Since the flow (SHPFm) is invariant under affine transformations, scaling self-similar solutions are in the form of affine transformations of the basis polygons. For k>0 we observe that the self-similar polygon solutions decay with oscillations for small damping coefficient, but for large damping there are no oscillations as they decay.
Proof. Since →X(t)=g(t)→X0 is to satisfy Eq (SHPFm) for all t, we use this to find an equivalent expression in terms of →X0. We have
d→Xdt=g′(t)→X0 |
and
d2→Xdt2=g″(t)→X0, |
so (SHPFm) becomes
g″(t)→X0+δg′(t)→X0=(−1)m+1g(t)Mm→X0, |
that is,
[g″(t)g(t)+δg′(t)g(t)]→X0=(−1)m+1Mm→X0. |
The right hand side above is independent of t so the scaling factor g(t) must satisfy
g″(t)+δg′(t)−Cg(t)=0 | (4.3) |
for some constant C.
This being the case, the remaining equation for →X0 is
C→X0=(−1)m+1Mm→X0. |
For this equation to have a nonzero solution →X0, it must be that C is an eigenvalue of (−1)m+1Mm. Thus we have the possibilities C=λm,k for any k with corresponding eigenvectors Pk and Pn−k.
Solving the differential equation (4.3) with C=λm,k for k∈{0,1,…,⌊n2⌋}, δ and the initial condition g(0)=1 we obtain the expressions for g(t) as in the statement of the proposition.
Proposition 4.2. Consider the family of polygons in the plane, →X(t) given by
→X(t)=→X0R(f(t)), | (4.4) |
where R(f(t)) represents the rotation of the polygon →X0 by angle f(t), with f(0)=0.
(1) If →X(t) satisfies (SHPFm) with δ=0 for all t, then →X(t) has the form
→X(t)=(c1Pk+c2Pn−k)R(±√−λm,kt), |
for some k∈{1,2,…,⌊n2⌋} and any constants c1,c2∈C, where λm,k is the corresponding eigenvalue for eigenvectors Pk and Pn−k.
(2) In the case δ>0, there are no nontrivial solutions of (SHPFm) that evolve by pure rotation.
Remark 4.1. (1) We did not include k=0 in Proposition 4.2(1) above because then λm,0=0 and →X(t)=→c for a complex constant c; this is just the trivial solution of (SHPFm).
(2) It is not surprising there are no solutions evolving by purely rotation with δ>0 as this corresponds physically to a damping term.
Proof. To establish an expression for the rotation matrix R(f(t)) in SO(2) we consider the skew symmetric matrix S:R→so(2) such that
S(f(t))=[0−f(t)f(t)0]. |
Note that S(f(t))=f(t)S(1) and so ddtS(f(t))=f′(t)S(1) and d2dt2S(f(t))=f″(t)S(1). Considering the map exp:so(2)→SO(2), we have exp(S(f(t)))=R(f(t)). Furthermore
ddtR(f(t))=f′(t)S(1)R(f(t)) and d2dt2R(f(t))=[f′(t)S(1)]2R(f(t))+f″(t)S(1)R(f(t)). |
We also note that S(1)2=−I2. Therefore for →X(t)=→X0R(f(t)) to satisfy (SHPFm) we have
→X0[(f″(t)+δf′(t))S(1)−(f′(t))2I2]=(−1)m+1Mm→X0. | (4.5) |
The right hand side above is independent of t so for a nonzero solution →X it must be that (f″(t)+δf′(t))S(1)−(f′(t))2I2 is constant for all t. This requires f′(t)≡b for some constant b and so f″(t)≡0. Therefore, (4.5) becomes
→X0(δbS(1)−b2I2)=(−1)m+1Mm→X0. | (4.6) |
If δ=0, then the above reduces to −b2→X0=(−1)m+1Mm→X0, and so
[(−1)m+1Mm+b2In]→X0=0n×2. | (4.7) |
For a nontrivial solution →X0 we require
det[(−1)m+1Mm+b2In]=⌊n2⌋∏k=0(λm,k+b2)=0. |
Hence b=±√−λm,k for some k∈{0,1,…,⌊n2⌋}. Given that −b2=λm,k, the null space of (−1)m+1Mm+b2In is spanned by Pk and Pn−k. Therefore →X0=c1Pk+c2Pn−k for complex constants c1,c2 and the rate of rotation f(t) is given by f(t)=±√−λm,kt in this case. Note that for k=0 we have λm,0=0 and P0=(1,…,1)T and so f(t)≡0 and →X0 is a constant vector.
We now consider the case δ>0. From (4.6) we have
→X0[δbS(1)−b2I2]2=(b4−δb2)→X0−2δb3→X0S(1)=M2m→X0. | (4.8) |
Letting →X0=[→x→y] we therefore have
(b4−δ2b2)[→x→y]−2δb3[→y−→x]=M2m[→x→y], |
and a rearrangement gives
(M2m−(b4−δ2b2)I2)[→x→y]=−2δb3[→y−→x]. | (4.9) |
Multiplying both sides by (M2m−(b4−δb2)I2) we have
(M2m−(b4−δ2b2)I2)2[→x→y]=−2δb3(M2m−(b4−δ2b2)I2)[→y−→x]=−2δb3[2δb3→x2δb3→y] from (4.9), |
which is equivalent to
[(M2m−(b4−δ2b2)In)2+4δ2b6]→X0=0n×2. |
For a non trivial →X0 we require
det[(M2m−(b4−δ2b2)In)2+4δ2b6]=⌊n2⌋∏k=0[(λ2m,k−(b4−δ2b2))2+4δ2b6]=0. |
Each factor in the product above is the sum of two squares. Since δ>0, the only chance of a zero factor is if b=0 which implies f(t)≡0. Substituting b=0 back into (4.8) we find only the trivial solution →X(t)=→c for any complex constant c.
Proposition 4.3. Consider the family of polygons in the plane, →X(t) given by
→X(t)=→X0+→h(t), | (4.10) |
where →h(t) represents the translation of the polygon →X0 and →h(0)=→0. If →X(t) satisfies (SHPFm) for all t then →X0 corresponds to a single point in the plane. That is, there are no non-trivial self-similar solutions by translation under the semi-discrete polyharmonic flow.
Proof. Since →h(t) translates each vertex of →X0 in the same way, it follows that →h(t) must be a vector in Cn with every entry equal to the same function of t. Since (4.10) is to satisfy (SHPFm), it follows that →h(t) must satisfy
→h″(t)+δ→h′(t)=(−1)m+1Mm[→X0+→h(t)]=(−1)m+1Mm→X0. |
Again, the right hand side is independent of t, so each element h(t) of →h(t) must satisfy
h″(t)+δh′(t)=C |
for some complex constant C. In view of Lemmas 3.3 and 3.4, the only solutions of
(−1)m+1Mm→X0=→C |
are the constants →X0=→a0, corresponding to trivial solutions of (SHPFm).
Remark 4.2. One can solve the ODE for h to obtain the expression for the path of translating point in both the δ=0 and δ>0 cases. The results are, for δ=0,
h(t)=dt |
and for δ>0,
h(t)=dδ(1−e−δt) |
for constant d not equal to zero.
In this subsection we move from considering those solutions to (SHPFm) that move self-similarly to general solutions given specified initial data. As the governing equation is hyperbolic there are natural analogues of initial position and velocity; the former is clear in our setting while the latter gives rise to several options. Consequently we specify our results in this section for given initial polygon →X0 only and include free parameters that can be determined from an appropriate additional condition. Two possible such conditions are outlined in the remark below.
Theorem 4.4. Given an initial polygon →X0=∑n−1k=0α0kPk with n vertices in R2 and any m∈N, the Eq (SHPFm) with δ=0 and initial data →X(0)=→X0 has a unique solution given by
→X(t)=(α00+a0t)P0+n−1∑k=1[α0kcos(√−λm,kt)+aksin(√−λm,kt)]Pk, | (4.11) |
where the ak, k=0,…,n−1 are arbitrary constants.
Proof. The proof is similar to the parabolic flow cases in [2] for m=1 and in [13] for general m. The set of eigenvectors {Pk}n−1k=0 of (−1)m+1Mm forms a basis for Cn. Therefore by considering →X(t)∈Cn we can write our polygon in the form
→X(t)=n−1∑k=0αk(t)Pk, | (4.12) |
where the coefficients αk(t) are complex. We have
d→Xdt=n−1∑k=0α′k(t)Pk |
and
d2→Xdt2=n−1∑k=0α″k(t)Pk. |
Since →X(t) satisfies (SHPFm) with δ=0, this gives
n−1∑k=0α″k(t)Pk=(−1)m+1Mmn−1∑k=0αk(t)Pk=n−1∑k=0αk(t)λm,kPk. |
Hence
α″k(t)=λm,kαk(t) |
for each t which implies
α0(t)=c0+a0t, |
while for all other k=1,…,n−1,
αk(t)=ckcos(√−λm,kt)+aksin(√−λm,kt), |
for constants ck,ak,k=0,1,…,n−1. The result follows in view of the initial coefficients.
Remark 4.3. The appearance of arbitrary constants a0,…,an−1 in the solution formula (4.11) is not surprising. Prescribing only the initial polygon does not give provide enough information to solve (SHPFm) uniquely. There are of course many ways extra information can be given yielding a unique solution. Two of these are
(1) A 'zero initial velocity' condition. Clearly, in view of (4.11), this will result in a0=a1=…=an−1=0.
(2) A 'prescribed polygon at later time' condition. In other words, not only do we specify the initial polygon, but we also specify another polygon at some later time. In view of the arguments of the sine functions in (4.11), this can be messy in general, however, suppose for a specific example we required
→X(π2√−λm,1)=P0+P1. |
From (4.11), we must therefore have
α00+a0π2√−λm,1=1, a1=1 and α0kcos(π2√λm,kλm,1)+aksin(π2√λm,kλm,1)=0, for k=2,…,n−1. |
This kind of specification of →X at another time is also a kind of approach to the discrete Yau problem where we consider the solution only on a finite time interval.
Next we consider the case of δ>0. The damping ensures convergence to a point that, under certain conditions and under appropriate rescaling, is an affine transformation of a regular polygon. This is fundamentally different behaviour from the undamped case where the solution (4.11) exhibits continued undamped oscillations.
Theorem 4.5. Given an initial polygon →X0=∑n−1k=0α0kPk with n vertices in R2 and any m∈N, the Eq (SHPFm) with δ>0 constant and initial data →X(0)=→X0 has a unique solution given by
→X(t)=[α00+a0δ−a0δe−δt]P0+n−1∑k=1αk(t)Pk, | (4.13) |
where
αk(t)={α0ker+m,kt+ak(er−m,kt−er+m,kt), for |λm,k|<δ24,(α0k+akt)e−δ2t, if λm,k=−δ24,e−δ2t[α0kcos(γm,kt)+aksin(γm,kt)], for |λm,k|>δ24, | (4.14) |
r±m,k=−δ2±√(δ2)2+λm,k, |
and
γm,k=√−λm,k−(δ2)2 . |
The constants ak for k=0,…,n−1 are arbitrary. The solution exists for all time and converges exponentially to a point.
When the dominant eigenvalue λm,d, where λm,d≥λm,k for all k∈{1,…,⌊n2⌋}, satisfies the condition |λm,d|<δ24, or λm,d=−δ24 with at least ad or an−d nonzero, then under appropriate rescaling, the solution is asymptotic as t→∞ to an affine transformation of a regular polygon with n vertices. Otherwise if λm,d=−δ24 and ad=an−d=0, or |λm,d|>δ24, then the solution in (4.13) exhibits continued oscillating behaviour as it shrinks to a point.
Remark: In view of (4.14), all modes of the solution are exponentially decaying except for P0.
Proof. Again writing →X(t)=∑n−1k=0αk(t)Pk, since →X(t) satisfies (SHPFm) with δ>0, this gives
n−1∑k=0[α″k(t)+δα′k(t)]Pk=(−1)m+1Mmn−1∑k=0αk(t)Pk=n−1∑k=0αk(t)λm,kPk. |
Therefore for each k,
α″k(t)+δα′k(t)=λm,kαk(t). |
Solving the above equation for different cases of the eigenvalues λm,k in relation to δ, solves the coefficients αk(t) as given in (4.14).
Since δ>0 and each λm,k<0 for non zero k, then each αk(t) goes to zero as t→∞ for each k=1,…,n−1. Therefore
limt→∞→X(t)=[α00+a0δ]P0, |
that is, each vertex of the polygon converges to the same point, by α00+a0δ where α00 is the complex constant given by the initial polygon, and a0 is an arbitrary complex constant.
To determine the limiting shape of the polygon, we consider an appropriately scaled and translated version of →X(t) given by
→Y(t)=g(t)(→X(t)−[α00+a0δ−a0δe−δt]P0). |
The scaling factor g(t) is determined by the value of the damping term δ and the relationship of the eigenvalues λm,k, to this damping term, such that this therefore determines the coefficient terms as given in (4.14).
Suppose that for the dominant eigenvalue λm,1, we have |λm,1|<δ24. Then we choose the scaling factor to be
g(t)=e−r+m,1t=e(δ2−√δ24+λm,1)t. |
Note that for k=2,…,n−1 we have
r±m,k−r+m,1=±√δ24+λm,k−√δ24+λm,1<0 |
and
−δ2−r+m,1=−√δ24+λm,1<0. |
Therefore for k=2,…,n−1 the expression e−r+m,1tαk(t) will go to zero as t→∞ and we have
limt→∞→Y(t)=(α01−a1)P1+(α0n−1−an−1)Pn−1. |
Therefore →Y(t) converges to an affine transformation of P1.
If we have the condition λm,1=−δ24 with a1 or an−1 nonzero, then we choose scaling factor g(t)=eδ2tt. We therefore obtain
limt→∞→Y(t)=a1P1+an−1Pn−1, |
where the limiting shape is again an affine transformation of P1.
In the case of the original polygon being orthogonal to P1, then we instead consider the next dominant eigenvalue λm,d for some d∈{2,…,⌊n2⌋} where the initial polygon is not orthogonal to Pd. The scaling factor g(t) is therefore chosen in the same way as described above, but instead involving λm,d.
In the case of the dominant eigenvalue with property λm,d=−δ24 and ad=an−d=0, or |λm,d|>δ24, then taking the scaling factor to be g(t)=eδ2t gives
→Y(t)=α0dPd+α0n−dPn−d+n−(d+1)∑k=d+1[α0kcos(γm,kt)+aksin(γm,kt)]Pk |
or
→Y(t)=n−1∑k=1[α0kcos(γm,kt)+aksin(γm,kt)]Pk, |
respectively. Therefore in these cases we have oscillating behaviour of the polygon as it shrinks to a point.
Remark 4.4. As we did earlier for the undamped case, we can supplement the initial condition in Theorem 4.4 with an additional condition to ensure a unique solution. In the case of prescribing a second polygon at a later time, we create a hyperbolic flow that begins with one polygon, transitions through a second polygon at some fixed time and then converges exponentially to a point that, under rescaling, is an affine transformation of a regular polygon for certain values of δ.
Figure 1 depicts the evolution of a pentagon under the semi-discrete hyperbolic polyharmonic flow for m=1,2,3 and with damping term δ=4. All arbitrary constants ak for k=0,1,…,n−1 are chosen to be zero. A selection of nine states of the polygon under the semi-discrete hyperbolic polyharmonic flow are superimposed over the initial polygon. The figures demonstrate the behaviour of the flow as the polygon shrinks and converges to an affine transformation of the regular polygon. In the case of m=1 we have |λm,k|<δ24 for all eigenvalues. In the case of m=2 and m=3, only the dominant eigenvalue λm,1 satisfies this condition, and so the terms in our solution that involve the eigenvectors associated with non-dominant eigenvalues, include coefficients with oscillating expressions as set out in (4.14). Convergence is faster for higher m.
Figure 2 depicts the evolution of the same pentagon as given in Figure 1 under the semi-discrete hyperbolic polyharmonic flow at the same overlayed time steps and for select values of m. In this case however, non zero arbitrary coefficients ak are prescribed such that the polygon flows to a specific polygon at a particular time value before shrinking to a point.
Figure 3 depicts the evolution of a hexagon under (SHPFm) for m=1,2,3 and damping term δ=7, where all arbitrary constants ak for k=0,1,…,n−1 are chosen to be zero. Distinct states of the polygon are overlayed over the initial polygon. In each case the dominant eigenvalue λm,1 has the condition |λ1,k|<δ24. Otherwise for each case of m shown, the coefficients given in the solution vary based on the conditions and expressions given in (4.14). We see that for m=3, where only the dominant eigenvalue satisfies |λ1,k|<δ24, that the convergence to the limiting shape is faster.
We may also examine ancient solutions of our hyperbolic polyharmonic flows of polygons by considering limits as t→−∞ in solution formulae, which always make sense in this setting. In the δ>0 case for planar polygons, we get different outcomes based on the relationships of the least and most dominant eigenvalues with δ, as well as the arbitrary constants ak chosen in the expression (4.13) and (4.14).
We demonstrate such solutions by following a similar argument as in the proof of Theorem 4.5, rescaling and translating →X(t) as follows
→Y(t)=g(t)(→X(t)−→α0(t)), |
for appropriate rescaling factor g(t). If we have |λm,⌊n2⌋|≤δ24 and ak=0 for all k, then we observe asymptotic convergence to the affine transformation of the regular polygon P⌊n2⌋ as t→−∞. We see this by choosing scaling factor g(t)=e−r+m,⌊n2⌋t if |λm,⌊n2⌋|<δ24, or g(t)=eδ2t if |λm,⌊n2⌋|=δ24, and note
limt→−∞→Y(t)=α0⌊n2⌋P⌊n2⌋+α0n−⌊n2⌋Pn−⌊n2⌋. |
When n is even, this will be the straight line interval of n2 overlapping polygon edges. If the original polygon is orthogonal to P⌊n2⌋ then we consider the next least dominant eigenvalue and get a similar result, provided the arbitrary constants ak are zero.
In the case of nonzero arbitrary constants, if we have at least a1 or an−1 not equal to zero, where λm,1 is the dominant eigenvalue and |λm,1|≤δ24, then choosing a rescaling factor of g(t)=e−r−m,1t if |λm,1|<δ24, or g(t)=eδ2tt if |λm,1|=δ24, demonstrates asymptotic convergence to an affine transformation of regular polygon P1 as t→−∞, given by
limt→−∞→Y(t)=a1P1+an−1Pn−1. |
If a1 and an−1 are both zero, then we consider the next dominant eigenvalue λm,d, d∈{2,…,⌊n2⌋}, where ad or an−d is nonzero, and similarly choose the rescaling factor g(t) as described above but involving eigenvalue λm,d instead. Here we have asymptotic convergence to an affine transformation of the regular polygon Pd as t→−∞.
To consider the flow in higher codimension, we set up similarly as in [2]. Let each vertex Xj∈Rp be denoted as Xj=(x1j,x2j,…,xpj). Consider the ith coordinate for every vertex in the polygon, for i=1,2,…,p, we can define
→xi=(xi0,xi1,…,xi(n−1))T, |
which is a vector in Rn. Therefore
→X=(→x1 ⋯ →xp). | (5.1) |
For k=0,1,…,⌊n2⌋, let us define the following vectors in Rn
→ck=(1,cos(2πkn),cos(4πkn),…,cos(2(n−1)πkn))T, | (5.2) |
→sk=(0,sin(2πkn),sin(2πkn),…,sin(2(n−1)πkn))T. | (5.3) |
The vectors →ck and →sk are the real and imaginary parts respectively of eigenvector Pk. Thinking of each entry of Pk as expressed as a vector in R2, we have Pk=(→ck→sk). Furthermore, nonzero elements from the set {→ck,→sk}k=0,1,…,⌊n2⌋ are mutually orthogonal and form a basis for Rn. Therefore each →xi for i=1,…,p can be expressed as
→xi=⌊n2⌋∑k=0(αik→ck+βik→sk) |
for real coefficients αik and βik. A family of polygons can therefore be expressed as
→X(t)=(→x1(t)⋯→xp(t))=⌊n2⌋∑k=0(→ck→sk)[α1k(t)⋯αpk(t)β1k(t)⋯βpk(t)]. | (5.4) |
We obtain the same behaviour of self-similar solutions in the higher codimension that we see in the plane polygon case. In regards to solutions that are self-similar by scaling, these solution polygons are planar in Rp. There are no non-trivial self similar solutions by translation. In the case of self similar solutions by rotation, when a damping term δ>0 is involved again we have no non-trivial solutions. In the rotation case where δ=0, pure rotations are possible. We provide a simple construction of a planar rotator, but do not give a complete classification of general rotators.
Proposition 5.1. If a family →X(t) of polygons with n vertices in Rp is a self-similar scaling solution to the flow (SHPFm), then →X(t) has the form
→X(t)=g(t)(→ck→sk)[α1⋯αpβ1⋯βp], | (5.5) |
for any k∈{1,2,...,⌊n2⌋} and real constants αj and βj for j=1,…,p, where
g(t)={ct+1, for δ=0 and k=0,cδ+[1−cδ]e−δt, for δ>0 and k=0,e−δ2t[cos(γm,kt)+csin(γm,kt)], for 0≤δ<2√|λm,k| and k∈{1,…,⌊n2⌋},(1+ct)e−δ2t, for δ=2√|λm,k| and k∈{1,…,⌊n2⌋},er+t+c(er−t−er+t), for δ>2√|λm,k| and k∈{1,…,⌊n2⌋}, | (5.6) |
such that
r±=−δ2±√(δ2)2+λm,k, γm,k=√|(δ2)2+λm,k| |
and constant c∈R.
The higher codimension polygon →X(t) is the image of a linear transformation of a regular polygon Pk=(→ck→sk) with n vertices in R2, with linear transformation T:R2→Rp given by
T(x,y)=(xy)[α1⋯αpβ1⋯βp]. | (5.7) |
Proof. Suppose →X(t)=g(t)→X0 for differentiable scaling function g:R→R. Following the same process as in the proof of Proposition 4.1, we find
(g″(t)g(t)+δg′(t)g(t))→X0=(−1)m+1Mm→X0 |
such that since the right hand side of the above equation is independent of t, we have
g″(t)+δg′(t)−Cg(t)=0. |
This results in the following
[(−1)m+1Mm−CIn]→X0=0n×p. | (5.8) |
For nonzero →X0 we therefore require C to be the eigenvalues of the matrix (−1)m+1Mm such that C=λm,k for k∈{0,1,…,⌊n2⌋}. The corresponding eigenvectors are Pk=(→ck→sk) and Pn−k=(→ck−→sk), noting →s0=→0 and →sn2=→0 for n even. Applying a linear transformation as given in (5.7), produces a polygon in Rp given by
→X0=(→ck→sk)[α1⋯αpβ1⋯βp], |
and furthermore this polygon solves (5.8), where αi and βi for all i=1,…,p are any real coefficients.
Similar to the plane polygon case, solving Eq (5.8) for C=λm,k,δ and the required initial condition g(0)=1 gives us the expressions for g(t) as stated in (5.6).
Proposition 5.2. Consider the family →X(t) of polygons in Rp such that
→X(t)=→X0R(t), | (5.9) |
where R:R→SO(p) represents a time-dependent p×p rotation of the polygon →X0 in Rp such that R(0)=Ip.
If →X(t) satisfies (SHPFm) for some δ>0 and all t then R(t)≡Ip and →X0 corresponds to a point in Rp, that is the only self similar solution by rotation in this case is the trivial solution.
If →X(t) satisfies (SHPFm) with δ=0 for all t then pure rotations are possible.
Proof. We first consider the case where δ>0. We also consider the explicit expression for the solutions for (SHPFm) stated and proven in Theorem 5.5. This solution is given in (5.15) where
→X0=⌊n2⌋∑k=0(→ck→sk)[α01k⋯α0pkβ01k⋯β0pk], |
and
→X(t)=(→α10(t)⋯→αp0(t))+⌊n2⌋∑k=1(→ck→sk)[α1k(t)⋯αpk(t)β1k(t)⋯βpk(t)], | (5.10) |
where
αi0(t)=α0i0+ai0δ−ai0δe−δt, |
for each i=1,…,p and ai0 are arbitrary constants. The αik(t) and βik(t) are given in (5.16) and (5.17) also in the statement for Theorem 5.5.
We require →X0R(t), where R(t) is a rotation function, to be of the form given in (5.10) above to satisfy (SHPFm). The expressions for αik(t) and βik(t) all include an exponential decaying term e−δ2t and so we must have ai0=0, αik(t)=0 and βik(t)=0 for all i=1,…,p and k=1,…,⌊n2⌋. This gives
→X0=→c0(α010⋯α0p0), |
and so our initial polygon is a point in Rp and we have only the trivial solution for →X(t) in the δ>0 case.
For the case where δ=0 we give an example of pure rotation self-similar solutions, without classifying all possible solutions of this type.
We consider the skew symmetric matrix S:R→so(p) such that for the exponential map, exp:so(p)→SO(p), from the set of skew symmetric p×p matrices to the set of p×p rotation matrices, we have exp(S(t))=R(t). The second derivative of R is therefore given by
d2Rdt2=d2Sdt2R(t)+(dSdt)2R(t). |
For →X(t)=→X0R(t) to satisfy (SHPFm) where δ=0 we have
→X0d2Rdt2 =(−1)m+1Mm→X0R(t), |
which in terms of the skew symmetric matrix and with rearrangement implies
→X0[d2Sdt2+(dSdt)2]=(−1)m+1Mm→X0. | (5.11) |
For (5.11) to be true for all t then d2Sdt2+(dSdt)2 must be a constant matrix we denote as B, and
→X0B=(−1)m+1Mm→X0. | (5.12) |
We can consider rotations in a plane given by orthogonal axes xi and xj, for i,j∈{1,…,p} and i≠j, where remaining p−2 axes are invariant. We denote this rotation as Rxi,xj(fij(t)) where the differentiable function fij:R→R is the angle of rotation at time t in the xi−xj plane and where fij(0)=0. Furthermore Rxi,xj(fij(t))=exp(Sxi,xj(fij(t))) for skew symmetric matrix Sxi,xj(fij(t)). This matrix Sxi,xj(fij(t))=[sab:a,b=1,…,p] has entries given by sij=−sji=−fij(t), or sij=−sji=fij(t), and all other entries are equal to zero. We also write Sxi,xj(fij(t))=fij(t)Sxi,xj(1) and note that
ddtSxi,xj(fij(t))=f′ij(t)Sxi,xj(1) and d2dt2Sxi,xj(fij(t))=f″ij(t)Sxi,xj(1). |
Denoting by Iij,p the p×p matrix with 1s in the (i,i)th and (j,j)th diagonal entries, and zeroes elsewhere, we have (Sxi,xj(1))2=−Iij,p.
Following from (5.11), for a rotation on the xi−xj sub-plane, the matrix
[f″ij(t)Sxi,xj(1)+(f′ij(t))2(Sxi,xj(1))2] |
has −(f′ij(t))2 in the (i,i)th and (j,j)th diagonal entries, −f″ij(t) in the (i,j)th entry and f″ij(t) in the (j,i)th entry (for i<j), and all other entries are zero.
Since this matrix is constant, we must have f′ij(t)=b, for some constant b and so f″ij(t)=0 for all t. Therefore we have
−b2→X0Iij,p=(−1)m+1Mm→X0. | (5.13) |
The matrix →X0Iij,p has zero coordinate vectors at each position, except for the ith and jth coordinate vectors. That is, for →X0Iij,p=(→x1⋯→xp),→xk=→0 for all k≠i,j. We denote this matrix as →X0,ij. For any k∈{1,…,⌊n2⌋}, consider an initial polygon of
→X0=→X0,ij=(→ck→sk)[α1⋯αpβ1⋯βp], |
where αl=βl=0 for all l≠i,j, and is therefore the image of a regular polygon Pk=(→ck→sk) from R2 mapped to Rp by a linear transformation, that is planar in the xi−xj sub-plane of Rp. We note that for this expression of →X0 which corresponds to the eigenvectors Pk, we have
(−1)m+1Mm→X0=λm,k→X0, |
for any k∈{1,…,⌊n2⌋}. From (5.13) we therefore have
−b2→X0=−b2→X0,ij=(−1)m+1Mm→X0, |
and so b=±√−λm,k for k∈{0,1,…,n−1} such that fij(t)=±√−λm,kt. The rotation in the xi−xj plane is therefore given by Rxi,xj(±√−λm,kt). This demonstrates that polygons in Rp given by affine images of regular polygons Pk in R2 such that they are planar polygons in main sub-planes of Rp, are self similar solutions by rotation. We suspect that there are also other rotation self-similar solutions for general rotations.
Remark 5.1. If the rotation R(t) is acting on a subspace that the polygon →X(t) is not in, then →X(t)=→X0R(t)=→X0. In this case for δ≥0, the only solution is the trivial solution →X0 where →X0=(→a1…→ap) for real constants ai, i=1,…,p.
We complete the discussion of higher co-dimension self-similar solutions by stating there are no non-trivial solutions by translation.
Proposition 5.3. Consider the family of polygons with n vertices in Rp, →X(t), such that
→X(t)=→X0+→h(t), | (5.14) |
where →h(t) is a n×p matrix that represents the translation of the polygon →X0 and →h(0)=0n×p. If →X(t) satisfies (SHPFm) for all t then each vertex of →X0 is the same fixed point in Rp. That is, there are no non-trivial self-similar solutions by translation under (SHPFm).
Proof. The proof follows the same process as that for Proposition 4.3. In the higher codimension case we find →h″(t)+δ→h′(t)=→C=(→c1…→cp) for real constants ci,i=1,…,p. Then from Lemma 3.3 and Lemma 3.4, the only solution to
(−1)m+1Mm→X0=→C |
is if →X0=(→a1…→ap) for real constants a1,…,ap.
A similar result to Theorems 4.4 and 4.5 holds for polygons in higher codimensions.
Theorem 5.4. Given an initial polygon →X0 with n vertices in Rp as expressed in (5.4), the Eq (SHPFm), for any m∈N, with δ=0 and initial data →X(0)=→X0, has a unique solution →X(t) for all time given by
→X(t)=→c0(α10(0)+a10t⋯αp0(0)+ap0t)+⌊n2⌋∑k=1(→ck→sk)[α1k(t)⋯αpk(t)β1k(t)⋯βpk(t)], |
where
αik(t)=αik(0)cos(√−λm,kt)+a1ksin(√−λm,kt) |
and
βik(t)=βik(0)cos(√−λm,kt)+b1ksin(√−λm,kt), |
where aik and bik are arbitrary constants for i=1,…,p and k=0,1,…,⌊n2⌋.
Proof. We follow a similar argument to Theorem 4.4 and the process given for the parabolic flow cases in [2] for m=1 and in [13] for general m. As given in the setup above and in (5.4), we consider a polygon given by
→X(t)=(→x1(t)⋯→xp(t))=⌊n2⌋∑k=0(→ck→sk)[α1k(t)⋯αpk(t)β1k(t)⋯βpk(t)]. |
The second derivative is given by
d2→Xdt2=d2dt2(→x1(t)⋯→xp(t))=⌊n2⌋∑k=0(→ck→sk)[α″1k(t)⋯α″pk(t)β″1k(t)⋯β″pk(t)]. |
Since the polygon →X(t) satisfies (SHPFm) with δ=0 then we have
⌊n2⌋∑k=0(→ck→sk)[α″1k(t)⋯α″pk(t)β″1k(t)⋯β″pk(t)]=(−1)m+1Mm⌊n2⌋∑k=0(→ck→sk)[α1k(t)⋯αpk(t)β1k(t)⋯βpk(t)]=⌊n2⌋∑k=0λm,k(→ck→sk)[α1k(t)⋯αpk(t)β1k(t)⋯βpk(t)]. |
Therefore α″ik(t)=λm,kαik and β″ik(t)=λm,kβik for all i=1,…,p and k=0,…,⌊n2⌋.
For k=0 we have λm,0=0 and so
αi0(t)=αi0(0)+ai0t and βi0(t)=βi0(0)+bi0t, |
for i=1,…,p and where ai0,bi0 are constants.
For k=1,…,⌊n2⌋, we have
αik(t)=αik(0)cos(√−λm,kt)+aiksin(√−λm,kt), |
and similarly
βik(t)=βik(0)cos(√−λm,kt)+biksin(√−λm,kt), |
for constants aik,bik, and as such the result follows.
A damping coefficient δ>0 in (SHPFm) causes the polygon to shrink to a point in Rp. Under appropriate rescaling, the solution converges to a polygon in Rp that is the affine image of a regular polygon in the plane.
Theorem 5.5. Given an initial polygon →X0 with n vertices in Rp as given by
→X0=⌊n2⌋∑k=0(→ck→sk)[α01k⋯α0pkβ01k⋯β0pk], |
and any m∈N, the Eq (SHPFm) with δ>0 constant and initial data →X(0)=→X0 has a unique solution given by
→X(t)=(→α10(t)⋯→αp0(t))+⌊n2⌋∑k=1(→ck→sk)[α1k(t)⋯αpk(t)β1k(t)⋯βpk(t)], | (5.15) |
where
αi0(t)=α0i0+ai0δ−ai0δe−δt, |
for each i=1,…,p and where ai0 are arbitrary constants.
Each αik(t) is given by
αik(t)={α0iker+m,kt+aik(er−m,kt−er+m,kt), for |λm,k|<δ24,(α0ik+aikt)e−δ2t, if λm,k=−δ24,e−δ2t[α0ikcos(γm,kt)+aiksin(γm,kt)], for |λm,k|>δ24, | (5.16) |
and similarly for βik(t),
βik(t)={β0iker+m,kt+bik(er−m,kt−er+m,kt), for |λm,k|<δ24,(β0ik+bikt)e−δ2t, if λm,k=−δ24,e−δ2t[β0ikcos(γm,kt)+biksin(γm,kt)], for |λm,k|>δ24, | (5.17) |
where
r±m,k=−δ2±√(δ2)2+λm,k |
and
γm,k=√−λm,k−(δ2)2 . |
The constants aik and bik are arbitrary constants for i=1,…,p and k=1,…,⌊n2⌋.
The solution exists for all time and converges exponentially to a point in Rp. If the dominant eigenvalue λm,d satisfies |λm,d|<δ24 or λm,d=−δ24 where at least aid or bid is nonzero for any i∈{1,…,p}, then under appropriate rescaling, the solution is asymptotic as t→∞ to a planar polygon in Rp with n vertices which is an affine image of a regular polygon in R2. If λm,d=−δ24 with aid=bid=0 for all i=1,…,p, or |λm,d|>δ24, the solution exhibits continued oscillating behaviour as it shrinks to a point.
Proof. Since the polygon →X(t) is to satisfy (SHPFm) with δ>0, we obtain
d2→Xdt2+δd→Xdt=⌊n2⌋∑k=0(→ck→sk)[α″1k(t)+δα′1k(t)⋯α″pk(t)+δα′pk(t)β″1k(t)+δβ′1k(t)⋯β″pk(t)+δβ′pk(t)]=(−1)m+1Mm⌊n2⌋∑k=0(→ck→sk)[α1k(t)⋯αpk(t)β1k(t)⋯βpk(t)]=⌊n2⌋∑k=0λm,k(→ck→sk)[α1k(t)⋯αpk(t)β1k(t)⋯βpk(t)]. |
Therefore we are left to solve
α″ik(t)+δα′ik(t)=λm,kαik(t) |
and
β″ik(t)+δβ′ik(t)=λm,kβik(t), |
for i=1,…,p and k=0,…,⌊n2⌋. This leads to expression for αi0(t) as stated in the theorem, and the cases for αik(t) and βik(t) as given in (5.16) and (5.17).
With the same reasoning as given for the planar case in Theorem 4.5, as t→∞ then each vertex of the polygon will converge to the same point in Rp given by (α010+a10δ,⋯,α0p0+ap0δ).
To determine the limiting shape of the polygon as it shrinks, again we consider an appropriate rescaling and translation of →X(t) given by
→Y(t)=g(t)(→X(t)−(→α10(t)⋯→αp0(t))), | (5.18) |
and where the scaling function is chosen based on the chosen damping term δ and the relationship of eigenvalues λm,k to this damping term, that then determine the expression for the coefficients αik(t) and βik(t).
If the dominant eigenvalue λm,1 satisfies |λm,1|<δ24, then we choose scaling function g(t)=e−r+m,1t. By similar reasoning as in Theorem 4.5, we find that
limt→∞→Y(t)=(→c1→s1)[α011−a11⋯α0p1−ap1β011−b11⋯β0p1−bp1]. |
We have P1=(→c1→s1) which is a regular polygon in plane R2. Therefore the limit of →Y(t) is the polygon P1 where each vertex is mapped to a vertex in Rp by the linear map T1:R2→Rp given by
T1(x,y)=(xy)[α011−a11⋯α0p1−ap1β011−b11⋯β0p1−bp1]. |
T1 is a linear transformation, and since P1 is in the plane, the image of P1 under T1 is two dimensional and therefore planar.
If λm,1=−δ24 and ai1 or bi1 is nonzero for any i∈{1,…,p}, then we choose the scaling factor to be g(t)=eδ2tt in (5.18). Taking the limit gives
limt→∞→Y(t)=(→c1→s1)[a11⋯ap1b11⋯bp1]. |
Therefore considering the linear map T2:R2→Rp given by
T2(x,y)=(xy)[a11⋯ap1b11⋯bp1] |
we have the same result as above where the limiting shape is given by the image of P1 in R2 mapped to Rp by T2.
Given T1 and T2 are linear transformations from R2, the resulting polygon in Rp that is the image of these maps, is planar.
If λm,1=−δ24 and ai1=bi1=0 for all i=1,…,p, then we choose the scaling factor g(t)=eδ2t for the expression in (5.18) which gives
→Y(t)=(→c1→s1)[α011⋯α0p1β011⋯β0p1]+⌊n2⌋∑k=2(→ck→sk)[α01kcos(γm,kt)+a1ksin(γm,kt)⋯α0pkcos(γm,kt)+apksin(γm,kt)β01kcos(γm,kt)+b1ksin(γm,kt)⋯β0pkcos(γm,kt)+bpksin(γm,kt)]. |
Similarly, for |λm,1|>δ24 we also choose the scaling factor g(t)=eδ2t which gives
→Y(t)=⌊n2⌋∑k=1(→ck→sk)[α01kcos(γm,kt)+a1ksin(γm,kt)⋯α0pkcos(γm,kt)+apksin(γm,kt)β01kcos(γm,kt)+b1ksin(γm,kt)⋯β0pkcos(γm,kt)+bpksin(γm,kt)]. |
As such, in both these cases the polygon exhibits continued oscillating behaviour.
In the case of αi1(0)=0 and βi1(0)=0 for all i=1,2,…,p, we instead scale the polygon by an expression involving λm,d such that λm,d is the next dominant eigenvalue where we have nonzero αid,βid, and carry out the same process as described above depending on the relationship λm,d has with the damping term δ.
Remark 5.2. Results regarding ancient solutions also hold for solutions in higher codimension using the scaled polygon
→Y(t)=g(t)(→X(t)−(→α10(t)⋯→αp0(t))), |
with g(t) chosen similarly as described in Section 4.3. Here we need to consider the values of the arbitrary constants aik and bik in the general solution given in (5.15). Whether these constants are equal to zero or not will determine the limiting behaviour of the polygon as t→−∞ and whether we have asymptotic convergence to an affine image of P1 or P⌊n2⌋ in Rp. The outcome is based on the same reasoning as in the planar polygon case.
In [11], Lin and Tsai described Yau's curvature difference flow whose objective is to evolve one curve to another, either in finite time or in infinite time, possibly up to an isometry, using a parabolic flow. In [13], we considered a first order in time semi-discrete polyharmonic flow analogue of Yau's curvature difference flow. Here we consider a second order in time analogue that includes a linear damping term.
Theorem 6.1. Given a target polygon →Y with n vertices in Rp, for p∈N,p≥2, and any fixed δ>0, all solutions →X(t) to the second order semi-discrete Yau difference flow,
d2→Xdt2+δd→Xdt=(−1)m+1Mm[→X(t)−→Y](SYDFm) |
with any initial polygon →X(0)=→X0 with n vertices, exist for all time and converges as t→∞ to a translate of →Y.
Remark 6.1. (1) Recalling the operator (−1)m+1Mm is a higher order linear curvature-type operator, the right hand side of (SYDFm) is precisely the difference in this type of curvature of →X as compared with that of →Y. When →X≡→Y, the right hand side is identically equal to zero and →X≡→Y is a stationary solution.
(2) Of particular interest is that convergence to →Y is obtained for any initial polygon.
(3) Above we say 'all solutions' because the problem as written is underdetermined. As earlier, to obtain a unique solution we must provide some additional data, for example, an initial 'velocity' or a specific polygon through which the solution transitions at a specific time. Our result says whatever the initial polygon and additional data, we will always have long time existence, and convergence as t→∞ to →Y.
(4) If, instead of having the same number of vertices, one of →X0 or →Y has fewer vertices, we can simply duplicate vertices or add points on the line segments joining vertices to create initial and target polygons with the same number of 'vertices'. This process is described in more detail in [13].
Proof. As in [13] we set up a difference polygon, →Z(t)=→X(t)−→Y.
Write
→Y=⌊n2⌋∑k=0(→ck→sk)[y(1)1k⋯y(1)pky(2)1k⋯y(2)pk], |
and
→X0=⌊n2⌋∑k=0(→ck→sk)[α01k⋯α0pkβ01k⋯β0pk]. |
Then
→Z0=⌊n2⌋∑k=0(→ck→sk)[α01k−y(1)1k⋯α0pk−y(1)pkβ01k−y(2)1k⋯β0pk−y(2)pk]. |
Since →Y is constant, it follows that →Z(t) satisfies Eq (SHPFm) and therefore, by Theorem 5.5 has solution
→Z(t)=⌊n2⌋∑k=0(→ck→sk)[z(1)1k(t)⋯z(1)pk(t)z(2)1k(t)⋯z(2)pk(t)], |
where for each i=1,…,p,
z(1)i0(t)=(α0i0−y(1)i0)+ai0δ−ai0δe−δt, | (6.1) |
where ai0 are arbitrary constants, and for k=1,2,…,n−1,
z(1)ik(t)={(α0ik−y(1)ik)er−m,kt+aik(er+m,kt−er−m,kt), for |λm,k|<δ24,(α0ik−y(1)ik+aikt)erm,kt, if λm,k=−δ24,e−δ2t[(α0ik−y(1)ik)cos(γm,kt)+aiksin(γm,kt)], for |λm,k|>δ24, | (6.2) |
and
z(2)ik(t)={(β0ik−y(2)ik)er−m,kt+bik(er+m,kt−er−m,kt), for |λm,k|<δ24,(β0ik−y(2)ik+bikt)erm,kt, if λm,k=−δ24,e−δ2t[(β0ik−y(2)ik)cos(γm,kt)+biksin(γm,kt)], for |λm,k|>δ24, | (6.3) |
where
r±m,k=−δ2±√(δ2)2+λm,k, γm,k=√−λm,k−(δ2)2. |
The constants aik and bik are completely free. Hence
→X(t)=→c0[α010+a10δ−a10δe−δt⋯α0p0+ap0δ−ap0δe−δt]+⌊n2⌋∑k=1(→ck→sk)[z(1)1k(t)+y(1)1k⋯z(1)pk(t)+y(1)pkz(2)1k(t)+y(2)1k⋯z(2)pk(t)+y(2)pk]. |
In view of (6.2) and (6.3), provided ai0 is chosen such that
α0i0+ai0δ=y(1)i0 |
for all i=1,…,p, we will have →X(t)→→Y as t→∞. Otherwise →X will converge to a translate of →Y as t→∞.
The case of a moving target polygon →Y(t) may be handled using Green's functions as follows. Considering polygons in the plane, our equation can now be written as
d2→Xdt2+δd→Xdt=(−1)m+1Mm[→X(t)−→Y(t)], | (6.4) |
where
→Y(t)=n−1∑k=0yk(t)Pk. |
Specifically, in view of Proposition 3.2 we may seek a solution to (6.4) of the form
→X(t)=n−1∑k=0αk(t)Pk. | (6.5) |
Then, by linearity of (6.4), the coefficients of the evolving polygon →X(t) satisfy
α″k(t)+δα′k(t)=λm,k[αk(t)−yk(t)] | (6.6) |
with solution for k=0
α0(t)=c0e−δt+d0 |
with arbitrary constants c0,d0. The given data will put at least one condition on these constants, so only in special cases will the translation term approach that of →Y.
For k=1,2,…,n−1, writing as usual
r±m,k=−δ2±√(δ2)2+λm,k, |
we have the independent real solutions to the homogeneous equation
α(1)k(t)=er−m,kt and α(2)k(t)=er+m,kt for |λm,k|<δ24, |
α(1)k(t)=erm,kt and α(2)k(t)=term,kt if λm,k=−δ24, |
and
α(1)k(t)=e−δ2tcos(γm,kt) and α(2)k(t)=e−δ2tsin(γm,kt) for |λm,k|>δ24, |
where
γm,k=√−λm,k−(δ2)2. |
These allow us to write down a Green's function corresponding to each coefficient function:
Gk(x,t)=1detWk(t)[α(1)k(x)α(2)k(t)−α(1)k(t)α(2)k(x)], |
where
Wk(t)=[α(1)k(t)α(2)k(t)ddtα(1)k(t)ddtα(2)k(t)]. |
The solution to the nonhomogeneous Eq (6.6) is then
αk(t)=c−kα(1)k(t)+c+kα(2)k(t)+∫t0Gk(x,t)yk(x)dx, |
where c±k are arbitrary constants. The solution polygon is then
→X(t)=[c0e−δt+d0]P0+n−1∑k=1[c−kα(1)k(t)+c+kα(2)k(t)+∫t0Gk(x,t)yk(x)dx]Pk. |
The limiting solution will not translate as per →Y unless d0=y0 is constant. In view of the form of the coefficients of the other Pk, whether the solution will approach →Y(t) as t→∞ depends precisely on the behaviour of the terms involving the Green's functions.
Figure 4 depicts examples of the semi-discrete Yau difference flow. In Figure 4c, nonzero values of ak are chosen such that the polygon passes through an intermediate polygon at a fixed time before converging to the target polygon. Figure 5 shows alternative cases for the Yau difference flow including where a polygon can be flowed to a polygon of different number of vertices by setting excess vertices along the line segments or by duplicating excess vertices.
Remark 6.2. (1) Hyperbolic flows that evolve one smooth curve to another are discussed in [14]. Parabolic flows that achieve this are described in [11,15]. In general, some conditions on the smooth initial and target curves are needed. For example, they might need to be strictly locally convex. Alternatively, for curves given as radial graphs one may flow the radial graph function by the heat equation.
(2) In the case of δ=0, there would be no exponential decay factors in the solution formula and instead of convergence to the target →Y we would have oscillating about the target polygon for all time.
(3) By considering a sequence t→−∞ of initial polygons, we can construct a flow with 'initial' (limiting t→−∞) polygon →X−∞ that passes through two states, say →X0 and →X1 at two distinct times before converging to the target polygon →Y. The hyperbolic flow with damping (SYDFm) allows two intermediate states →X0 and →X1 in such a process, as compared with a parabolic flow that would allow one intermediate state. More intermediate states could be accommodated by using higher order flows. Such considerations could be relevant in practical applications, for example, where a collection of robots or drones need to pass through several specific states before approach a long-term target state.
(4) The evolution Eq (SYDFm) can flow any initial polygon →X0 with n sides to any target polygon →Y with n sides. As in the smooth case and as discussed in [11] for example, there are also other ways of deforming one polygon to another that do not involve a curvature flow. In our setting we could simply take, for example, →X:[0,1]→Rp given by
→X(t)=t→Y+(1−t)→X0. |
(5) As with our earlier flows, one may consider ancient solutions to (SYDFm). Again in comparison with the smooth case any solution at all may be extended back in time.
(6) The flow (SYDFm) can evolve an n1-gon →X0 to an n2-gon →Y where n1≠n2 are not necessarily the same. We can either duplicate vertices or add additional vertices on the line segments of the polygon of fewer vertices. If for example n1<n2, then →X0 is adjusted to a n2-gon where some of its original vertices are duplicated (Figure 5a). Similarly if n1>n2,→Y can be adjusted to a n1-gon with the same approach (Figure 5b). An example of adding additional vertices along connecting line segments is given in Figure 5c.
In this paper we considered damped hyperbolic motion of closed polygons in Rp, p≥2, by a linear semi-discrete hyperbolic analogue of polyharmonic curve diffusion flows. Our flows correspond to second order linear ODEs, and solutions in both the plane and in higher codimension are given explicitly and their behaviour explored. For positive damping, an initial polygon may evolve to any other prescribed polygon at a fixed later time, before converging exponentially to a point in Rp. Under certain conditions and appropriate rescaling, a plane polygon asymptotically converges to an affine transformation of a regular polygon. In the higher codimension case, given similar conditions and again with appropriate rescaling, the polygon asymptotically converges to a planar polygon in Rp that is an affine image of are regular plane polygon. For systems with zero damping factor, the solution undergoes continued undamped oscillations. Self-similar solutions under such semi-discrete hyperbolic flows are also examined. We also introduced the semi-discrete linear hyperbolic analogue of Yau's curvature difference flow, where we are able to evolve any initial closed polygonal curve to any other polygon. The introduction of the second time derivative for the hyperbolic flows in question allows for an additional polygonal state to be prescribed, such that an initial polygon flows to an intermediate state at a fixed time, before converging exponentially to a given target polygon in infinite time.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
This work was completed while the first author was supported by Discovery Project DP180100431 of the Australian Research Council and the second author was supported by an UNRS Central Scholarship. Parts of this work was completed while the first author was visiting Nanjing University of Information Science and Technology and the MATRIX Institute. The authors are grateful for this support. The authors are grateful to Professor Ben Andrews for a discussion on self-similar solutions in higher codimension.
The authors declare no conflicts of interest.
Calculations for example in Figures 1a and 2a
We provide further detail for the calculation of the examples given in Figures 1a and 2a. In both these cases we have an initial plane polygon with five vertices given as a vector in C5,
→X0=(−1+10i,0,9−i,3+9i,10+2i)T. | (A.1) |
We consider damping term δ=4 and m=1.
The basis eigenvectors {Pk}4k=0 are given by P0=(1,1,1,1,1)T, P1=(1,ω,ω2,ω3,ω4)T, P2=(1,ω2,ω4,ω,ω3)T, P3=(1,ω3,ω,ω4,ω2)T, and P4=(1,ω4,ω3,ω2,ω)T, where ω=e2πi5. Furthermore we have eigenvalues
λ1,k=−4sin2(πk5) |
for k=0,1,2,3,4. We note that |λ1,k|<δ24=4 for all k=0,…,4. The coefficients α0k in the expression →X0=∑4k=0α0kPk are calculated as follows,
α0k=15⟨→X0,Pk⟩ |
for →X0 as in (A.1), and we note that α00=215+4i. The solution to (SHPFm) for this example is given by
→X(t)=[215+4i+a04−a04e−4t]P0+4∑k=1αk(t)Pk, | (A.2) |
where from (4.14) we have
αk(t)=(α0k−ak)e(−2+√4+λ1,k)t+ake(−2−√4+λ1,k)t |
for k=1,2,3,4, and the ak are arbitrary constants. All five vertices converge to 215+a04+4i as t increases. We consider appropriate rescaling by scaling factor
g(t)=e(2−√4+λ1,1)t=e(3−√5)t2, |
and translation of →X(t) such that
→Y(t)=g(t)(→X(t)−[215+4i+a04−a04e−4t]P0). |
Therefore
limt→∞→Y(t)=(α01−a1)P1+(α04−a4)P4, | (A.3) |
which is an affine transformation of the regular pentagon P1.
For the example given in Figure 1a we have ak=0 for all k=0,…,4, and so the coefficients αk(t) in (A.2) for k=1,2,3,4 are given by
αk(t)=α0ke(−2+√4+λ1,k)t, |
and the polygon converges to (215+4i)P0 as t increases.
For the example given in Figure 2a, we choose the constants ak such that
→X(1.2)=((215+4i)P0+3P1). |
Therefore, we find
a0=0, a1=3−α01e1.2r+1,1e1.2r−1,1−e1.2r+1,1=3−α01e35(√5−3)e−35(√5+5)−e35(√5−3) |
and
ak=−α0ke1.2r+1,ke1.2r−1,k−e1.2r+1,k, for k=2,3,4, |
such that
ak=−α0ke35(√5−1)e−35(√5−1)−e35(√5−1), for k=2,3, and a4=−α04e35(√5+1)e−35(√5+1)−e35(√5+1). |
[1] |
H. Attouch, X. Goudou, P. Redont, The heavy ball with friction method, I. The continuous dynamical system: global exploration of the local minima of a real-valued function by asymptotic analysis of a dissipative dynamical system, Commun. Contemp. Math., 2 (2000), 1–34. https://doi.org/10.1142/S0219199700000025 doi: 10.1142/S0219199700000025
![]() |
[2] |
B. Chow, D. Glickenstein, Semidiscrete geometric flows of polygons, Am. Math. Mon., 114 (2007), 316–328. https://doi.org/10.1080/00029890.2007.11920419 doi: 10.1080/00029890.2007.11920419
![]() |
[3] | F. R. K. Chung, Spectral graph theory, CBMS Regional Conference Series in Mathematics, American Mathematical Society, 1997. |
[4] | P. Davis, Circulant matrices, John Wiley & Sons, 1979. |
[5] |
K. Deckelnick, R. Nürnberg, Discrete hyperbolic curvature flow in the plane, SIAM J. Numer. Anal., 61 (2023), 1835–1857. https://doi.org/10.1137/22M1493112 doi: 10.1137/22M1493112
![]() |
[6] |
X. Goudou, J. Munier, The gradient and heavy ball with friction dynamical systems: the quasiconvex case, Math. Program., 116 (2009), 173–191. https://doi.org/10.1007/s10107-007-0109-5 doi: 10.1007/s10107-007-0109-5
![]() |
[7] |
K. Groh, S. Röck, A contribution to collision-free trajectory planning for handling systems in varying environments, Prod. Eng. Res. Devel., 4 (2010), 101–106. https://doi.org/10.1007/s11740-009-0202-0 doi: 10.1007/s11740-009-0202-0
![]() |
[8] | B. Hu, S. Xu, Z. Cao, Multi-robot path planning for each robot with several jobs in a single trip, IFAC-PapersOnLine, 53, (2020), 279–284. https://doi.org/10.1016/j.ifacol.2021.04.106 |
[9] |
M. Huptych, S. Röck, Online path planning in dynamic environments using the curve shortening flow method, Prod. Eng. Res. Devel., 9 (2015), 613–621. https://doi.org/10.1007/s11740-015-0641-8 doi: 10.1007/s11740-015-0641-8
![]() |
[10] |
M. Huptych, S. Röck, Real-time path planning in dynamic environments for unmanned aerial vehicles using the curve-shortening flow method, Int. J. Adv. Robotic Syst., 18 (2021), 1–16. https://doi.org/10.1177/1729881420968687 doi: 10.1177/1729881420968687
![]() |
[11] |
Y. C. Lin, D. H. Tsai, Evolving a convex closed curve to another one via a length-preserving linear flow, J. Differ. Equations, 247 (2009), 2620–2636. https://doi.org/10.1016/j.jde.2009.07.024 doi: 10.1016/j.jde.2009.07.024
![]() |
[12] |
J. McCoy, Representation formulae for higher-order linear curvature flows, J. Evol. Equ., 25 (2025), 40. https://doi.org/10.1007/s00028-025-01067-9 doi: 10.1007/s00028-025-01067-9
![]() |
[13] | J. McCoy, J. Meyer, Linear semi-discrete polyharmonic flows of closed polygons, arXiv, 2025. https://doi.org/10.48550/arXiv.2502.05243 |
[14] |
J. McCoy, I. Otuf, Representation formulae for linear hyperbolic curvature flows, J. Differ. Equations, 397 (2024), 166–198. https://doi.org/10.1016/j.jde.2024.03.007 doi: 10.1016/j.jde.2024.03.007
![]() |
[15] |
J. McCoy, P. Schrader, G. Wheeler, Representation formulae for higher order curvature flows, J. Differ. Equations, 344 (2023), 1–43. https://doi.org/10.1016/j.jde.2022.10.011 doi: 10.1016/j.jde.2022.10.011
![]() |
[16] |
C. Rademacher, H. B. Rademacher, Solitons of discrete curve shortening, Results Math., 71 (2017), 455–482. https://doi.org/10.1007/s00025-016-0572-5 doi: 10.1007/s00025-016-0572-5
![]() |
[17] |
C. Rademacher, H. B. Rademacher, Solitons of the midpoint mapping and affine curvature, J. Geom., 112 (2021), 7. https://doi.org/10.1007/s00022-020-00567-y doi: 10.1007/s00022-020-00567-y
![]() |
[18] |
S. Smith, M. Broucke, B. Francis, Curve shortening and the rendezvous problem for mobile autonomous robots, IEEE Trans. Autom. Control, 52 (2007), 1154–1159. https://doi.org/10.1109/TAC.2007.899024 doi: 10.1109/TAC.2007.899024
![]() |
[19] | K. Smoczyk, A representation formula for the inverse harmonic mean curvature flow, Elem. Math., 60 (2005), 57–65. |
[20] | G. J. Tee, Eigenvectors of block circulant and alternating circulant matrices, N. Z. J. Math., 36 (2007), 195–211. |
[21] |
F. Tisseur, K. Meerbergen, The quadratic eigenvalue problem, SIAM Rev., 43 (2001), 235–286. https://doi.org/10.1137/S0036144500381988 doi: 10.1137/S0036144500381988
![]() |
[22] | M. Wardetzky, S. Mathur, F. Kälberer, E. Grinspun, Discrete Laplace operators: no free lunch, Eurographics Symposium on Geometry Processing, 7 (2007), 33–37. |