This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Curves of constant curvature and torsion in the 3-sphere

Debraj Chakrabarti Department of Mathematics, Central Michigan University, Mt. Pleasant, MI 48859, USA chakr2d@cmich.edu http://people.cst.cmich.edu/chakr2d/ Rahul Sahay Central Michigan University, Mt. Pleasant, MI 48859, USA sahay1r@cmich.edu  and  Jared Williams Central Michigan University, Mt. Pleasant, MI 48859, USA willi9jr@cmich.edu
Abstract.

We describe the curves of constant (geodesic) curvature and torsion in the three-dimensional round sphere. These curves are the trajectory of a point whose motion is the superposition of two circular motions in orthogonal planes. The global behavior may be periodic or the curve may be dense in a Clifford torus embedded in the three-sphere. This behavior is very different from that of helices in three-dimensional Euclidean space, which also have constant curvature and torsion.

Key words and phrases:
Curves in the three-sphere, Frenet-Serret equations, Constant curvature and torsion, Geodesic curvature, Helices
2010 Mathematics Subject Classification:
53A35
All three authors were partially supported by a grant from the NSF (#1600371). Debraj Chakrabarti was partially supported by a grant from the Simons Foundation (# 316632) and also by an Early Career internal grant from Central Michigan University.

1. Introduction

Let (M,,)\left(M,\langle,\rangle\right) be a three dimensional Riemannian manifold, let II\subseteq\mathbb{R} be an open interval, and let 𝜸:IM{\boldsymbol{\gamma}}:I\to M be a smooth curve in MM, which we assume to be parametrized by the arc length tt. It is well-known that the local geometry of 𝜸{\boldsymbol{\gamma}} is characterized by the curvature κ\kappa and the torsion τ\tau. These are functions defined along 𝜸{\boldsymbol{\gamma}} and are the coefficients of the well-known Frenet-Serret formulas [2, Vol. IV, pp. 21-23]:

Ddt𝐓(t)\displaystyle\frac{D}{dt}{\mathbf{T}(t)} =\displaystyle= κ(t)𝐍(t)\displaystyle\phantom{-}\kappa(t)\mathbf{N}(t)
Ddt𝐍(t)\displaystyle\frac{D}{dt}{\mathbf{N}(t)} =κ(t)𝐓(t)\displaystyle=-\kappa(t){\mathbf{T}}(t) +τ(t)𝐁(t)\displaystyle+\tau(t)\mathbf{B}(t) (1.1)
Ddt𝐁(t)\displaystyle\frac{D}{dt}{\mathbf{B}(t)} =\displaystyle= τ(t)𝐍(t),\displaystyle-\tau(t){\mathbf{N}}(t),

where the orthogonal unit vector fields 𝐓,𝐍,𝐁\mathbf{T,N,B}, with 𝐓=𝜸\mathbf{T}={\boldsymbol{\gamma}}^{\prime}, along the unit-speed curve 𝜸{\boldsymbol{\gamma}}, constitute its Frenet frame and Ddt\frac{D}{dt} denotes covariant differentiation along 𝜸{\boldsymbol{\gamma}} with respect to the arc length tt. We will assume that each of the functions κ\kappa and τ\tau is either nowhere zero or vanishes identically. Additionally, if κ\kappa is identically zero, then τ\tau is also taken to be identically zero. For completeness, we include a proof of the set of Equations given in (1.1) in Section 3.1 below. We make the following definition:

Definition 1.

Let MM be a Riemannian manifold of dimension 3. A curve
𝜸:IM{\boldsymbol{\gamma}}:I\to M, where II\subseteq\mathbb{R} is an open interval, will be called a helix (plural: helices) if its curvature κ\kappa and torsion τ\tau are non-negative constants. A helix is non-degenerate if κ\kappa and τ\tau are both positive, and degenerate otherwise. We say that the helix 𝜸{\boldsymbol{\gamma}} is periodic if there is a T>0T>0 such that 𝜸(t+T)=𝜸(t){\boldsymbol{\gamma}}(t+T)={\boldsymbol{\gamma}}(t) for each tIt\in I.

We take τ\tau to be non-negative since we use the non-oriented form of the Frenet-Serret Equations (see Section 3). Definition 1 is motivated by the example of the Euclidean space 3\mathbb{R}^{3}, where non-degenerate helices are curves of the form:

𝜸(t)=cos(ωt)𝐀+sin(ωt)𝐁+t𝐂+𝐃\displaystyle{\boldsymbol{\gamma}}(t)=\cos(\omega t)\mathbf{A}+\sin(\omega t)\mathbf{B}+t\mathbf{C}+\mathbf{D} (1.2)

where 𝐀,𝐁,𝐂,𝐃3\mathbf{A},\mathbf{B},\mathbf{C},\mathbf{D}\in\mathbb{R}^{3}, 𝐀,𝐁,𝐂\mathbf{A},\mathbf{B},\mathbf{C} are non-zero and orthogonal with |𝐀|=|𝐁|\left|\mathbf{A}\right|=\left|\mathbf{B}\right|, and ω>0\omega>0. These are elegant curves that are invariant under a one-parameter group of isometries of the ambient space. Note that there are no non-degenerate periodic helices in 3\mathbb{R}^{3}.

The aim of this paper is to study helices in the three dimensional round sphere 𝕊3\mathbb{S}^{3}. Thanks to the fact that 𝕊3\mathbb{S}^{3} is compact, we expect that a non-degenerate helix in 𝕊3\mathbb{S}^{3} should “come back where it started from” provided we wait long enough, and therefore, there is a possibility that, for favorable choice of the curvature and torsion, the helix is actually periodic, though locally it is not much different from a helix in 3\mathbb{R}^{3}. Globally, a non-degenerate helix in 𝕊3\mathbb{S}^{3} has two fundamental angular frequencies, ω1\omega_{1} and ω2\omega_{2}, as opposed to the single fundamental angular frequency, ω\omega, of the helix given by Equation (1.2) in 3\mathbb{R}^{3}. A non-degenerate helix in 𝕊3\mathbb{S}^{3} may be thought of as the trajectory of a particle which performs two superimposed circular motions with frequencies ω1\omega_{1} and ω2\omega_{2}. These fundamental angular frequencies must satisfy

ω2>1>ω1,\omega_{2}>1>\omega_{1},

a constraint which arises because a curve with non-zero curvature and torsion must lie in the positively curved compact space 𝕊3\mathbb{S}^{3}. Unlike in 3\mathbb{R}^{3} where non-degenerate helices are embedded, non-compact submanifolds, depending on the fundamental angular frequencies ω1\omega_{1} and ω2\omega_{2}, a non-degenerate helix in 𝕊3\mathbb{S}^{3} can either be periodic (when it is a compact embedded submanifold) or be dense in a flat 2-torus contained in 𝕊3\mathbb{S}^{3} (when the image of the helix is not an embedded submanifold of 𝕊3\mathbb{S}^{3}). This divergence of global behavior from the flat case is the main topic of this paper.

Of course, the same questions can be asked in any number of spatial dimensions, and for other Riemannian manifolds. Here, for simplicity we consider the special case of 𝕊3\mathbb{S}^{3}, which also allows us to use only elementary calculus-based methods in our investigations. Our methods will likely generalize to round spheres of any number of dimensions.

2. Main results

We consider 𝕊3\mathbb{S}^{3} to be embedded in the Euclidean space 4\mathbb{R}^{4} in the natural way as the hypersurface {x12+x22+x32+x42=1}\{x_{1}^{2}+x_{2}^{2}+x_{3}^{2}+x_{4}^{2}=1\}, and endow 𝕊3\mathbb{S}^{3} with the Riemannian metric induced from 4\mathbb{R}^{4}. This entails no loss of generality because the metric so induced is the same as the standard round metric of 𝕊3\mathbb{S}^{3} with constant sectional curvature 11. To state our results concisely, let us introduce the following definition:

Definition 2.

A smooth curve 𝜸{\boldsymbol{\gamma}} in 4\mathbb{R}^{4} will be called a Lissajous curve if there are numbers ω2>ω10\omega_{2}>\omega_{1}\geq 0 and vectors 𝐀1,𝐁1,𝐀2,𝐁24\mathbf{A}_{1},\mathbf{B}_{1},\mathbf{A}_{2},\mathbf{B}_{2}\in\mathbb{R}^{4} such that, for each tt,

𝜸(t)=cos(ω1t)𝐀1+sin(ω1t)𝐁1+cos(ω2t)𝐀2+sin(ω2t)𝐁2.\displaystyle{\boldsymbol{\gamma}}(t)=\cos(\omega_{1}t)\mathbf{A}_{1}+\sin(\omega_{1}t)\mathbf{B}_{1}+\cos(\omega_{2}t)\mathbf{A}_{2}+\sin(\omega_{2}t)\mathbf{B}_{2}. (2.1)

We will call ω1\omega_{1} and ω2\omega_{2} the fundamental angular frequencies of the curve 𝜸{\boldsymbol{\gamma}} and 𝐀1,𝐁1,𝐀2,𝐁2\mathbf{A}_{1},\mathbf{B}_{1},\mathbf{A}_{2},\mathbf{B}_{2} the coefficient vectors of 𝜸{\boldsymbol{\gamma}}.

Therefore, a Lissajous curve, in our sense, can be thought of as the trajectory of a point in 4\mathbb{R}^{4} which oscillates with frequency ω1\omega_{1} in the 𝐀1𝐁1\mathbf{A}_{1}\mathbf{B}_{1}-plane and with frequency ω2\omega_{2} in the 𝐀2𝐁2\mathbf{A}_{2}\mathbf{B}_{2}-plane. Note also that the projection of 𝜸{\boldsymbol{\gamma}} on any 2-dimensional linear subspace different from the 𝐀1𝐁1\mathbf{A}_{1}\mathbf{B}_{1} and 𝐀2𝐁2\mathbf{A}_{2}\mathbf{B}_{2}-planes is a planar Lissajous curve in the usual sense of the term [1, pp. 114-115]. We are now ready to describe helices in 𝕊3\mathbb{S}^{3}:

Theorem 1.

Let κ,τ0\kappa,\tau\geq 0 be given numbers where, if κ=0\kappa=0, then τ=0\tau=0.

  1. (1)

    There exists a helix 𝜸:𝕊3{\boldsymbol{\gamma}}:\mathbb{R}\to\mathbb{S}^{3} with constant curvature κ\kappa and torsion τ\tau.

  2. (2)

    Such a helix 𝜸{\boldsymbol{\gamma}} is a Lissajous curve in the form of Equation (2.1).

  3. (3)

    The fundamental angular frequencies of 𝜸{\boldsymbol{\gamma}} are distinct and are given by

    ω1\displaystyle\omega_{1} =χ2χ44τ22\displaystyle=\sqrt{\frac{\chi^{2}-\sqrt{\chi^{4}-4\tau^{2}}}{2}} (2.2)
    ω2\displaystyle\omega_{2} =χ2+χ44τ22\displaystyle=\sqrt{\frac{\chi^{2}+\sqrt{\chi^{4}-4\tau^{2}}}{2}} (2.3)

    with

    χ2=κ2+τ2+1.\displaystyle\chi^{2}=\kappa^{2}+\tau^{2}+1. (2.4)
  4. (4)

    If κ>0\kappa>0, then the frequencies ω1\omega_{1} and ω2\omega_{2} satisfy

    ω2>1>ω1.\displaystyle\omega_{2}>1>\omega_{1}. (2.5)
  5. (5)

    If τ0\tau\neq 0 then the four coefficient vectors 𝐀1,𝐁1,𝐀2,𝐁2\mathbf{A}_{1},\mathbf{B}_{1},\mathbf{A}_{2},\mathbf{B}_{2} are orthogonal in 4\mathbb{R}^{4}, and their magnitudes are given by

    |𝐀1|2=|𝐁1|2=1ω22ω12ω22\displaystyle\left|\mathbf{A}_{1}\right|^{2}=\left|\mathbf{B}_{1}\right|^{2}=\frac{1-\omega_{2}^{2}}{\omega_{1}^{2}-\omega_{2}^{2}} (2.6)

    and

    |𝐀2|2=|𝐁2|2=1ω12ω22ω12.\displaystyle\left|\mathbf{A}_{2}\right|^{2}=\left|\mathbf{B}_{2}\right|^{2}=\frac{1-\omega_{1}^{2}}{\omega_{2}^{2}-\omega_{1}^{2}}. (2.7)
  6. (6)

    If τ=0\tau=0, then 𝜸{\boldsymbol{\gamma}} is a circle given by

    𝜸(t)=𝐀1+cos(ωt)𝐀2+sin(ωt)𝐁2{\boldsymbol{\gamma}}(t)=\mathbf{A}_{1}+\cos(\omega t)\mathbf{A}_{2}+\sin(\omega t)\mathbf{B}_{2}

    where ω=κ2+1\omega=\sqrt{\kappa^{2}+1}. Further, the coefficient vectors 𝐀1,𝐀2,𝐁2\mathbf{A}_{1},\mathbf{A}_{2},\mathbf{B}_{2} are mutually orthogonal and we have

    |𝐀2|\displaystyle\left|\mathbf{A}_{2}\right| =|𝐁2|=1ω and |𝐀1|=11ω2.\displaystyle=\left|\mathbf{B}_{2}\right|=\frac{1}{\omega}\quad\text{ and }\quad\left|\mathbf{A}_{1}\right|=\sqrt{1-\frac{1}{\omega^{2}}}.

Several interesting features may be noted here:

  1. (1)

    The local existence of helices follows from the existence theorem for solutions of systems of ordinary differential equations on manifolds. However, we prove the existence of helices directly by solving the Frenet-Serret equations and obtain an explicit representation of the solution.

  2. (2)

    When κ,τ>0\kappa,\tau>0, the curve 𝜸{\boldsymbol{\gamma}} may be thought of as the trajectory of a motion consisting of two superimposed circular motions in perpendicular planes: one at a “slow” frequency ω1<1\omega_{1}<1 and the other at a “fast” frequency ω2>1\omega_{2}>1. This global behavior arises from the fact that the curve 𝜸{\boldsymbol{\gamma}} must lie on the compact surface 𝕊3\mathbb{S}^{3}. Observe that there is no such restriction on the angular frequency ω\omega of the Euclidean helix given by Equation (1.2).

  3. (3)

    When κ=0\kappa=0 by definition we have τ=0\tau=0. Such a curve is a geodesic, i.e., its unit tangent field is auto-parallel along the curve. Therefore, geodesics on the sphere 𝕊3\mathbb{S}^{3} are of the form

    𝜸(t)=cos(t)𝐀+sin(t)𝐁{\boldsymbol{\gamma}}(t)=\cos(t)\mathbf{A}+\sin(t)\mathbf{B}

    where |𝐀|=|𝐁|=1\left|\mathbf{A}\right|=\left|\mathbf{B}\right|=1 and 𝐀,𝐁\mathbf{A},\mathbf{B} are mutually orthogonal. Thus, we recapture the well known fact that geodesics in 𝕊3\mathbb{S}^{3} are great circles.

We now turn to the question of uniqueness and periodicity of helices. First, note that if 𝜸{\boldsymbol{\gamma}} is a helix in a Riemannian 3-manifold, MM, and f:MMf:M\to M is a self-isometry of MM, then f𝜸f\circ{\boldsymbol{\gamma}} is also a helix in MM with the same curvature and torsion as that of 𝜸{\boldsymbol{\gamma}}. In 3\mathbb{R}^{3}, the converse holds, i.e., helices with the same curvature and torsion are congruent under an isometry of 3\mathbb{R}^{3}. We show that a similar fact holds in 𝕊3\mathbb{S}^{3} and we also determine when helices are periodic.

Recall that a Clifford torus is a Riemannian 2-manifold which is the metric product of two circles. Clearly, a Clifford torus is flat, i.e., its Gaussian curvature vanishes identically. It is well-known that there are Clifford tori embedded in the sphere 𝕊3\mathbb{S}^{3}, e.g. for 0<λ<10<\lambda<1, the surface in 4\mathbb{R}^{4} given by

𝒞λ={𝐱4:x12+x22=λ,x32+x42=1λ}\mathcal{C}_{\lambda}=\left\{\mathbf{x}\in\mathbb{R}^{4}:x_{1}^{2}+x_{2}^{2}=\lambda,x_{3}^{2}+x_{4}^{2}=1-\lambda\right\}

is clearly contained in 𝕊3\mathbb{S}^{3}, and is therefore a Clifford torus in 𝕊3\mathbb{S}^{3} which is flat in the Riemannian metric induced by the round metric of 𝕊3\mathbb{S}^{3}.

Theorem 2.

Let κ,τ0\kappa,\tau\geq 0 be given numbers where, if κ=0\kappa=0, then τ=0\tau=0.

  1. (1)

    If 𝜶{\boldsymbol{\alpha}} and 𝜷{\boldsymbol{\beta}} are two helices in 𝕊3\mathbb{S}^{3} with the same curvature κ\kappa and torsion τ\tau, then 𝜶{\boldsymbol{\alpha}} and 𝜷{\boldsymbol{\beta}} are congruent, i.e., there is a Riemannian isometry f:𝕊3𝕊3f:\mathbb{S}^{3}\to\mathbb{S}^{3} such that 𝜷=f𝜶{\boldsymbol{\beta}}=f\circ{\boldsymbol{\alpha}}.

  2. (2)

    A helix 𝜸{\boldsymbol{\gamma}} is periodic if and only if the ratio of the angular frequencies

    ω1ω2=χ2χ44τ2χ2+χ44τ2\displaystyle\cfrac{\omega_{1}}{\omega_{2}}=\sqrt{\frac{\chi^{2}-\sqrt{\chi^{4}-4\tau^{2}}}{\chi^{2}+\sqrt{\chi^{4}-4\tau^{2}}}} (2.8)

    is a rational number, where χ2=κ2+τ2+1\chi^{2}=\kappa^{2}+\tau^{2}+1.

  3. (3)

    If κ,τ>0\kappa,\tau>0, there exists a Clifford torus 𝕋𝜸2\mathbb{T}^{2}_{{\boldsymbol{\gamma}}} contained in 𝕊3\mathbb{S}^{3} such that the image of 𝜸{\boldsymbol{\gamma}} lies on 𝕋𝜸2\mathbb{T}^{2}_{{\boldsymbol{\gamma}}}.

  4. (4)

    If κ,τ>0\kappa,\tau>0 and ω1ω2\cfrac{\omega_{1}}{\omega_{2}} is irrational, the image of 𝜸{\boldsymbol{\gamma}} is dense in the torus 𝕋𝜸2\mathbb{T}_{{\boldsymbol{\gamma}}}^{2}.

2.1. Visualization of Helices

One way to visualize 𝕊3\mathbb{S}^{3} is to use the sterographic projection σ:𝕊3{p}3\sigma:\mathbb{S}^{3}\setminus\{p\}\to\mathbb{R}^{3} where pp is a point in 𝕊3\mathbb{S}^{3} which serves as the pole of the projection. It is well-known that σ\sigma is conformal, i.e. it preserves angles but not lengths. Using Wolfram Mathematica™, we produced visualizations of two helices in 𝕊3\mathbb{S}^{3} which are shown in Figures 1 and 2 below. Each of these pictures represents two distinct perspective projections onto 2\mathbb{R}^{2} of the sterographic projection of the helix, where pp is chosen to not be on the helix. The helix in Figure 1 is non-periodic and therefore dense in a Clifford torus. The helix in Figure 2 is periodic and therefore an embedded curve in 𝕊3\mathbb{S}^{3}. The hue and brightness of the following curves are functions of the fourth coordinate of the curve 𝜸{\boldsymbol{\gamma}} in its embedding in 4\mathbb{R}^{4}.

Refer to caption
Refer to caption
Figure 1. Two views of a dense helix in 𝕊3\mathbb{S}^{3} with κ=534\kappa=\frac{5\sqrt{3}}{4} and τ=294\tau=\frac{\sqrt{29}}{4}. The corresponding fundamental angular frequencies are then ω1=12,ω2=292,\omega_{1}=\frac{1}{2},\omega_{2}=\frac{\sqrt{29}}{2}, and thus, their ratio is the irrational number ω2ω1=29\frac{\omega_{2}}{\omega_{1}}=\sqrt{29}.
Refer to caption
Refer to caption
Figure 2. Two views of a periodic helix in 𝕊3\mathbb{S}^{3} with κ=153\kappa=\frac{\sqrt{15}}{3} and τ=512\tau=\frac{5}{12}. The corresponding fundamental angular frequencies are then ω1=14,ω2=53,\omega_{1}=\frac{1}{4},\omega_{2}=\frac{5}{3}, and thus, their ratio is the rational number ω2ω1=203\frac{\omega_{2}}{\omega_{1}}=\frac{20}{3}.

3. The Frenet-Serret Equations

3.1. The Frenet-Serret Equations in a 3-Dimensional Riemannian Manifold

Consider a three dimensional Riemannian manifold (M,,)({M},\langle,\rangle) and an arc length parameterized curve 𝜸:IM\boldsymbol{\gamma}:I\to M where II\subset\mathbb{R} is an open interval. Let Ddt\frac{D}{dt} represent the covariant derivative of a vector field along a curve (parametrized by tt). Let 𝐓=𝜸\mathbf{T}=\boldsymbol{\gamma}^{\prime} denote the unit tangent vector field of 𝜸\boldsymbol{\gamma}. Since 𝐓(t),𝐓(t)=1\left\langle\mathbf{T}(t),\mathbf{T}(t)\right\rangle=1 for each tt we have,

0=ddt𝐓(t),𝐓(t)=2𝐓(t),Ddt𝐓(t).0=\frac{d}{dt}\left\langle\mathbf{T}(t),\mathbf{T}(t)\right\rangle=2\left\langle\mathbf{T}(t),\frac{D}{dt}\mathbf{T}(t)\right\rangle.

Then, the curvature function κ\kappa is defined as

κ(t)=|Ddt𝐓(t)|.\displaystyle\kappa(t)=\left|\frac{D}{dt}\mathbf{T}(t)\right|. (3.1)

We will assume that either κ(t)0\kappa(t)\neq 0 for all tt or that κ0\kappa\equiv 0. In the case where κ\kappa never vanishes, we define the normal vector field to 𝜸\boldsymbol{\gamma} by

𝐍(t)=1κ(t)Ddt𝐓(t).\displaystyle\mathbf{N}(t)=\frac{1}{\kappa(t)}\frac{D}{dt}\mathbf{T}(t).

Then 𝐍\mathbf{N} is a unit vector field along 𝜸\boldsymbol{\gamma} which is always orthogonal to 𝐓\mathbf{T}. The definition of 𝐍\mathbf{N} gives the first Frenet-Serret equation

Ddt𝐓(t)=κ(t)𝐍(t).\displaystyle\frac{D}{dt}\mathbf{T}(t)=\kappa(t)\mathbf{N}(t). (3.2)

Similarly, since 𝐍(t),𝐍(t)=1\left\langle\mathbf{N}(t),\mathbf{N}(t)\right\rangle=1 for each tIt\in I

2𝐍(t),Ddt𝐍(t)\displaystyle 2\left\langle\mathbf{N}(t),\frac{D}{dt}\mathbf{N}(t)\right\rangle =0,\displaystyle=0,

and because 𝐓(t),𝐍(t)=0\left\langle\mathbf{T}(t),\mathbf{N}(t)\right\rangle=0 for each tIt\in I

Ddt𝐓(t),𝐍(t)+𝐓(t),Ddt𝐍(t)\displaystyle\left\langle\frac{D}{dt}\mathbf{T}(t),\mathbf{N}(t)\right\rangle+\left\langle\mathbf{T}(t),\frac{D}{dt}\mathbf{N}(t)\right\rangle =κ(t)+𝐓(t),Ddt𝐍(t)=0\displaystyle=\kappa(t)+\left\langle\mathbf{T}(t),\frac{D}{dt}\mathbf{N}(t)\right\rangle=0

by Equation (3.2). Then,

Ddt𝐍(t)=κ(t)𝐓(t)+vector orthogonal to 𝐓(t) and 𝐍(t).\displaystyle\frac{D}{dt}\mathbf{N}(t)=-\kappa(t)\mathbf{T}(t)+{\text{vector orthogonal to }\mathbf{T}(t)\text{ and }\mathbf{N}(t)}.

We define a torsion function τ\tau by

τ(t)=|Ddt𝐍(t)+κ(t)𝐓(t).|\displaystyle\tau(t)=\left|\frac{D}{dt}\mathbf{N}(t)+\kappa(t)\mathbf{T}(t).\right| (3.3)

We will assume that either τ0\tau\neq 0 for all tt, or τ0\tau\equiv 0. If τ(t)0\tau(t)\neq 0 for all tt, then we set

𝐁(t)=1τ(t)(Ddt𝐍(t)+κ(t)𝐓(t))\displaystyle\mathbf{B}(t)=\frac{1}{\tau(t)}\left(\frac{D}{dt}\mathbf{N}(t)+\kappa(t)\mathbf{T}(t)\right)

such that 𝐁\mathbf{B} is a unit vector field along 𝜸{\boldsymbol{\gamma}} which is orthogonal to 𝐓 and 𝐍\mathbf{T}\text{ and }\mathbf{N} for all tt. If τ0\tau\equiv 0, then we choose 𝐁\mathbf{B} to be an auto-parallel vector field along 𝜸{\boldsymbol{\gamma}} such that the vectors 𝐓(t),𝐍(t) and 𝐁(t)\mathbf{T}(t),\mathbf{N}(t)\text{ and }\mathbf{B}(t) form an orthonormal basis of T𝜸(t)𝕊3T_{{\boldsymbol{\gamma}}(t)}\mathbb{S}^{3}. In both cases we have

Ddt𝐍(t)=κ(t)𝐓(t)+τ(t)𝐁(t).\displaystyle\frac{D}{dt}\mathbf{N}(t)=-\kappa(t)\mathbf{T}(t)+\tau(t)\mathbf{B}(t). (3.4)

Finally, since 𝐁(t),𝐁(t)=1\left\langle\mathbf{B}(t),\mathbf{B}(t)\right\rangle=1 for each tIt\in I

𝐁(t),Ddt𝐁(t)=0,\displaystyle\left\langle\mathbf{B}(t),\frac{D}{dt}\mathbf{B}(t)\right\rangle=0,

and because 𝐍(t),𝐁(t)=0\left\langle\mathbf{N}(t),\mathbf{B}(t)\right\rangle=0 for each tIt\in I

Ddt𝐍(t),𝐁(t)+𝐍(t),Ddt𝐁(t)\displaystyle\left\langle\frac{D}{dt}\mathbf{N}(t),\mathbf{B}(t)\right\rangle+\left\langle\mathbf{N}(t),\frac{D}{dt}\mathbf{B}(t)\right\rangle =τ(t)+𝐍(t),Ddt𝐁(t)\displaystyle=\tau(t)+\left\langle\mathbf{N}(t),\frac{D}{dt}\mathbf{B}(t)\right\rangle

by Equation (3.4). Then,

Ddt𝐁(t)\displaystyle\frac{D}{dt}\mathbf{B}(t) =τ(t)𝐍(t)+vector orthogonal to 𝐍(t) and 𝐁(t).\displaystyle=-\tau(t)\mathbf{N}(t)+{\text{vector orthogonal to }\mathbf{N}(t)\text{ and }\mathbf{B}(t)}.
Ddt𝐁(t)=τ(t)𝐍(t)+c𝐓(t)\displaystyle\implies\frac{D}{dt}\mathbf{B}(t)=-\tau(t)\mathbf{N}(t)+c\mathbf{T}(t)

since 𝐓(t)\mathbf{T}(t) is orthogonal to 𝐍(t) and 𝐁(t)\mathbf{N}(t)\text{ and }\mathbf{B}(t) for each tIt\in I by construction. By taking the dot product of both sides with 𝐓(t)\mathbf{T}(t) we have

Ddt𝐁(t),𝐓(t)=τ(t)𝐍(t)+c𝐓(t)=c\displaystyle\left\langle\frac{D}{dt}\mathbf{B}(t),\mathbf{T}(t)\right\rangle=-\tau(t)\mathbf{N}(t)+c\mathbf{T}(t)=c

and by the product rule

Ddt𝐁(t),𝐓(t)\displaystyle\left\langle\frac{D}{dt}\mathbf{B}(t),\mathbf{T}(t)\right\rangle =ddt𝐁(t),𝐓(t)𝐁(t),Ddt𝐓(t)=0\displaystyle=\frac{d}{dt}\left\langle\mathbf{B}(t),\mathbf{T}(t)\right\rangle-\left\langle\mathbf{B}(t),\frac{D}{dt}\mathbf{T}(t)\right\rangle=0
c=0.\displaystyle\implies c=0.

Therefore we have our third and final Frenet-Serret equation

Ddt𝐁(t)=τ(t)𝐍(t)\displaystyle\frac{D}{dt}\mathbf{B}(t)=-\tau(t)\mathbf{N}(t) (3.5)

Equations (3.2), (3.4) and (3.5) constitute the Frenet-Serret equations in a 3-dimensional Riemannian manifold and characterize the local geometry of the curve 𝜸{\boldsymbol{\gamma}}. This concludes the derivation of the Frenet-Serret formulas in the case where κ(t)0\kappa(t)\neq 0 for all tt.

However, in the case where κ0\kappa\equiv 0, we define τ0\tau\equiv 0 and choose 𝐍,𝐁\mathbf{N},\mathbf{B} to be auto-parallel vector fields along 𝜸{\boldsymbol{\gamma}} such that the vectors 𝐓(t),𝐍(t), and 𝐁(t)\mathbf{T}(t),\mathbf{N}(t),\text{ and }\mathbf{B}(t) form an orthonormal basis of T𝜸(t)𝕊3T_{{\boldsymbol{\gamma}}(t)}\mathbb{S}^{3}. Under this choice, the Frenet-Serret Formulas given in (1.1) are again satisfied.

Note that we are not assuming that the manifold MM is orientable. In the case where MM is in fact oriented (i.e., MM is orientable, and one of the two orientations is specified), there is a variant of the Frenet-Serret equations in which one assumes that the frame {𝐓,𝐍,𝐁}\{\mathbf{T,N,B}\} is positively oriented. Then, one must allow the torsion τ\tau to assume negative values. The Equations  1.1 continue to hold with this new interpretation. However, in this paper, we use the non-oriented form of the Frenet-Serret equations, where κ\kappa and τ\tau are always non-negative. Geometrically, this means that while considering helices in 𝕊3\mathbb{S}^{3}, we disregard the chirality.

3.2. The Frenet equations in 𝕊3\mathbb{S}^{3}

We begin by specializing the Frenet-Serret equations given in (1.1) to the case of the embedded sphere 𝕊3\mathbb{S}^{3} in 4\mathbb{R}^{4}. Let

ι:𝕊34\iota:\mathbb{S}^{3}\hookrightarrow\mathbb{R}^{4}

be the natural embedding. Given a curve 𝜸:I𝕊3{\boldsymbol{\gamma}}:I\to\mathbb{S}^{3}, where II\subset\mathbb{R} is an open interval, we may think of 𝜸{\boldsymbol{\gamma}} as a curve in 4\mathbb{R}^{4}, by identifying 𝜸{\boldsymbol{\gamma}} with ι𝜸\iota\circ{\boldsymbol{\gamma}}. Similarly, given a vector field 𝐯\mathbf{v} along the curve 𝜸{\boldsymbol{\gamma}} which assigns to each point tIt\in I a vector 𝐯(t)T𝜸(t)𝕊3\mathbf{v}(t)\in T_{{\boldsymbol{\gamma}}(t)}\mathbb{S}^{3}, we can identify 𝐯\mathbf{v} with the vector field ι𝐯\iota_{*}\mathbf{v} along ι𝜸\iota\circ{\boldsymbol{\gamma}}, which assigns to the point tIt\in I the vector ι𝐯(t)Tι𝜸(t)4\iota_{*}\mathbf{v}(t)\in T_{\iota\circ{\boldsymbol{\gamma}}(t)}\mathbb{R}^{4}. In order to simplify notation, we adopt the following conventions:

  1. (1)

    Consistently identifying 𝕊3\mathbb{S}^{3} with the embedded image, we will omit the map ι\iota and its pushforward ι\iota_{*} from the notation. Thus, we will think of the 𝜸{\boldsymbol{\gamma}} in 𝕊3\mathbb{S}^{3} as a curve in 4\mathbb{R}^{4} whose image lies in 𝕊3\mathbb{S}^{3}. Similarly, we will think of a vector field 𝐯\mathbf{v} in 𝕊3\mathbb{S}^{3} along 𝜸{\boldsymbol{\gamma}} as a vector field in 4\mathbb{R}^{4} along 𝜸{\boldsymbol{\gamma}} such that for each tt, the vector 𝐯(t)T𝜸(t)4\mathbf{v}(t)\in T_{{\boldsymbol{\gamma}}(t)}\mathbb{R}^{4} lies in the subspace T𝜸(t)𝕊3T_{{\boldsymbol{\gamma}}(t)}\mathbb{S}^{3}.

  2. (2)

    We identify the tangent bundle T4T\mathbb{R}^{4} with 4×4\mathbb{R}^{4}\times\mathbb{R}^{4}. Therefore, all vector fields in 4\mathbb{R}^{4} may be identified with 4\mathbb{R}^{4}-valued functions.

  3. (3)

    Given a vector field 𝐯\mathbf{v} along a curve 𝜸{\boldsymbol{\gamma}} in 𝕊3\mathbb{S}^{3}, by the previous two parts, we can identify it with a 4\mathbb{R}^{4}-valued function. We will let 𝐯\mathbf{v}^{\prime} denote its derivative in the Euclidean space 4\mathbb{R}^{4}, i.e., if 𝐯\mathbf{v} is represented using the natural coordinates as

    𝐯(t)=(v1(t),v2(t),v3(t),v4(t)),\mathbf{v}(t)=(v_{1}(t),v_{2}(t),v_{3}(t),v_{4}(t)),

    where vj:I4v_{j}:I\to\mathbb{R}^{4} is smooth, j=1,,4j=1,\dots,4, then

    𝐯(t)=(v1(t),v2(t),v3(t),v4(t)).\mathbf{v}^{\prime}(t)=(v_{1}^{\prime}(t),v_{2}^{\prime}(t),v_{3}^{\prime}(t),v_{4}^{\prime}(t)).

    Of course, this is nothing but a coordinate expression for the covariant derivative of the vector field 𝐯\mathbf{v} along 𝜸{\boldsymbol{\gamma}} with respect to the flat Euclidean metric of 4\mathbb{R}^{4}.

Proposition 3.1.

Let 𝛄:I𝕊3{\boldsymbol{\gamma}}:I\to\mathbb{S}^{3} be a smooth curve in the three-sphere parametrized by arc length, and let 𝐓,𝐍,𝐁\mathbf{T,N,B} be its Frenet frame. Using the notational convention explained above, we think of 𝐓,𝐍,𝐁\mathbf{T,N,B} as functions from II to 4\mathbb{R}^{4}. Then, these vector valued functions satisfy the following differential equations:

𝐓(t)+𝜸(t)\displaystyle\mathbf{T}^{\prime}(t)+{\boldsymbol{\gamma}}(t) =κ(t)𝐍(t)\displaystyle=-\kappa(t)\mathbf{N}(t)
𝐍(t)\displaystyle\mathbf{N}^{\prime}(t) =κ(t)𝐓(t)+τ(t)𝐁(t)\displaystyle=-\kappa(t)\mathbf{T}(t)+\tau(t)\mathbf{B}(t) (3.6)
𝐁(t)\displaystyle\mathbf{B}^{\prime}(t) =τ(t)𝐍(t)\displaystyle=-\tau(t)\mathbf{N}(t)
Proof.

We begin by recalling the following fact from differential geometry [2, Vol. III, p. 2]. Let MM be an embedded submanifold of N\mathbb{R}^{N}, and for each point xMx\in M, let

𝒫x:TxNTxM\mathcal{P}_{x}:T_{x}\mathbb{R}^{N}\to T_{x}M

denote the orthogonal projection (where we identify TxMT_{x}M in the natural way with a subspace of TxN=NT_{x}\mathbb{R}^{N}=\mathbb{R}^{N}). We endow MM with the Riemannian metric induced by the Euclidean metric of N\mathbb{R}^{N}. Let 𝜸:IM{\boldsymbol{\gamma}}:I\to M be a smooth curve in MM, where II\subset\mathbb{R} is an open interval, and assume that 𝜸{\boldsymbol{\gamma}} is parametrized by arc length. If 𝐯\mathbf{v} is a vector field along 𝜸{\boldsymbol{\gamma}}, it is well-known that the covariant derivative of 𝐯\mathbf{v} is given by

Ddt𝐯(t)=𝒫𝜸(t)(𝐯(t))T𝜸(t)M.\frac{D}{dt}\mathbf{v}(t)=\mathcal{P}_{{\boldsymbol{\gamma}}(t)}\left(\mathbf{v}^{\prime}(t)\right)\in T_{{\boldsymbol{\gamma}}(t)}M.

When MM is a hypersurface in N\mathbb{R}^{N} and xMx\in M, we may write for 𝐫TxNN\mathbf{r}\in T_{x}\mathbb{R}^{N}\simeq\mathbb{R}^{N},

𝒫x(𝐫)=𝐫(𝐧(x)𝐫)𝐧(x),\mathcal{P}_{x}(\mathbf{r})=\mathbf{r}-\left(\mathbf{n}(x)\cdot\mathbf{r}\right)\mathbf{n}(x),

where 𝐧(x)\mathbf{n}(x) denotes a unit vector normal to the hypersurface MM at the point xx. Consequently we obtain the following formula for differentiating a vector field 𝐯\mathbf{v} along the curve 𝜸{\boldsymbol{\gamma}}:

Ddt𝐯(t)=𝐯(t)(𝐧(𝜸(t))𝐯(t))𝐧(𝜸(t)).\frac{D}{dt}\mathbf{v}(t)=\mathbf{v}^{\prime}(t)-\left(\mathbf{n}({\boldsymbol{\gamma}}(t))\cdot\mathbf{v}^{\prime}(t)\right)\mathbf{n}({\boldsymbol{\gamma}}(t)).

When M=𝕊3M=\mathbb{S}^{3} in 4\mathbb{R}^{4}, we may take for 𝐱𝕊3\mathbf{x}\in\mathbb{S}^{3}

𝐧(𝐱)=𝐱,\mathbf{n}(\mathbf{x})=\mathbf{x},

so that

Ddt𝐯(t)=𝐯(t)(𝜸(t))𝐯(t))𝜸(t)).\displaystyle\frac{D}{dt}\mathbf{v}(t)=\mathbf{v}^{\prime}(t)-\left({\boldsymbol{\gamma}}(t))\cdot\mathbf{v}^{\prime}(t)\right){\boldsymbol{\gamma}}(t)). (3.7)

We now compute 𝜸(t)𝐯(t){\boldsymbol{\gamma}}(t)\cdot\mathbf{v}^{\prime}(t) when 𝐯\mathbf{v} is one of the Frenet frame vector fields 𝐓,𝐁,𝐍\mathbf{T,B,N}. Note that the four vectors 𝜸(t){\boldsymbol{\gamma}}(t), 𝐓(t)\mathbf{T}(t), 𝐍(t)\mathbf{N}(t), and 𝐁(t)\mathbf{B}(t) form an orthonormal set in 4\mathbb{R}^{4}. Observe that for each tt we have the following equations:

𝐓(t)𝜸(t)\displaystyle\mathbf{T}^{\prime}(t)\cdot{\boldsymbol{\gamma}}(t) =(𝐓𝜸)(t)𝐓(t)𝜸(t)=0(𝐓(t)𝐓(t))=1,\displaystyle=(\mathbf{T}\cdot{\boldsymbol{\gamma}})^{\prime}(t)-\mathbf{T}(t)\cdot{\boldsymbol{\gamma}}^{\prime}(t)=0-(\mathbf{T}(t)\cdot\mathbf{T}(t))=-1,
𝐍(t)𝜸(t)\displaystyle\mathbf{N}^{\prime}(t)\cdot{\boldsymbol{\gamma}}(t) =(𝐍𝜸)(t)𝐍(t)𝜸(t)=0(𝐍(t)𝐓(t))=0,\displaystyle=(\mathbf{N}\cdot{\boldsymbol{\gamma}})^{\prime}(t)-\mathbf{N}(t)\cdot{\boldsymbol{\gamma}}^{\prime}(t)=0-(\mathbf{N}(t)\cdot\mathbf{T}(t))=0,
𝐁(t)𝜸(t)\displaystyle\mathbf{B}^{\prime}(t)\cdot{\boldsymbol{\gamma}}(t) =(𝐁𝜸)(t)𝐁(t)𝜸(t)=0(𝐁(t)𝐓(t))=0.\displaystyle=(\mathbf{B}\cdot{\boldsymbol{\gamma}})^{\prime}(t)-\mathbf{B}(t)\cdot{\boldsymbol{\gamma}}^{\prime}(t)=0-(\mathbf{B}(t)\cdot\mathbf{T}(t))=0.

In the first equation, we have used the fact that 𝐓=𝜸\mathbf{T}={\boldsymbol{\gamma}}^{\prime}. Using Equation (3.7) and the above compuations, we obtain the following represenations of the covariant derivatives of the Frenet frame:

Ddt𝐓(t)\displaystyle\frac{D}{dt}\mathbf{T}(t) =𝐓(t)+𝜸(t)\displaystyle=\mathbf{T}^{\prime}(t)+{\boldsymbol{\gamma}}(t)
Ddt𝐍(t)\displaystyle\frac{D}{dt}\mathbf{N}(t) =𝐍(t)\displaystyle=\mathbf{N}^{\prime}(t)
Ddt𝐁(t)\displaystyle\frac{D}{dt}\mathbf{B}(t) =𝐁(t).\displaystyle=\mathbf{B}^{\prime}(t).

Combining these with the Frenet-Serret equations (1.1) in a Riemannian three-manifold, Equations (3.6) follow. ∎

4. Lissajous curves in 𝕊3\mathbb{S}^{3}

In this section, we prove a few results that will be needed to complete the proof of Theorem 1. The following lemma will be required.

Lemma 4.1.

Let α0,α1,,αN\alpha_{0},\alpha_{1},\cdots,\alpha_{N} be distinct non-negative real numbers, and suppose that for each t0t\geq 0, we have

k=0N(aksin(αkt)+bkcos(αkt))=0,\displaystyle\sum_{k=0}^{N}\left(a_{k}\sin{(\alpha_{k}t)}+b_{k}\cos{(\alpha_{k}t)}\right)=0, (4.1)

where the coefficients ak,bka_{k},b_{k} are complex numbers. Then we have ak=bk=0a_{k}=b_{k}=0 for each kk.

Proof.

We can assume without loss of generality that α0=0\alpha_{0}=0 (simply take a0=b0=0a_{0}=b_{0}=0). For k=1,,Nk=1,\dots,N, let us set αk=αk\alpha_{-k}=-\alpha_{k}. Then, Equation (4.1) takes the form

k=NNckeiαkt=0\displaystyle\sum_{k=-N}^{N}c_{k}e^{i\alpha_{k}t}=0 (4.2)

where

ck={a|k|+ib|k|2i,k>0a0,k=0a|k|+ib|k|2i,k<0.\displaystyle c_{k}=\begin{cases}\cfrac{a_{\left|k\right|}+ib_{\left|k\right|}}{2i},&k>0\\ \quad\quad a_{0},&k=0\\ \cfrac{-a_{\left|k\right|}+ib_{\left|k\right|}}{2i},&k<0.\end{cases}

For each k0k\geq 0, it follows that ck=ck=0c_{k}=c_{-k}=0 if and only if ak=bk=0a_{k}=b_{k}=0.

Fix an integer \ell, ||N\left|\ell\right|\leq N, and multiply both sides of Equation (4.2) by eiαte^{-i\alpha_{\ell}t}. Integrating on the interval [0,T][0,T] and dividing by TT, we see that for each T0T\geq 0 we have

k=NkNckT0Tei(αkα)t𝑑t+c=0.\displaystyle\sum_{\begin{subarray}{c}k=-N\\ k\not=\ell\end{subarray}}^{N}\frac{c_{k}}{T}\int_{0}^{T}e^{i(\alpha_{k}-\alpha_{\ell})t}dt+c_{\ell}=0. (4.3)

Note, however, that if kk\not=\ell, we have

|0Tei(αkα)t𝑑t||ei(αkα)T1i(αkα)|2|αkα|.\left|\int_{0}^{T}e^{i(\alpha_{k}-\alpha_{\ell})t}dt\right|\leq\left|\frac{e^{i(\alpha_{k}-\alpha_{\ell})T}-1}{i(\alpha_{k}-\alpha_{\ell})}\right|\leq\frac{2}{\left|\alpha_{k}-\alpha_{\ell}\right|}.

Since for each kk, this is bounded independently of TT, as TT\to\infty, each term in the first sum of Equation (4.3) goes to 0, which shows that c=0c_{\ell}=0. Therefore, a=b=0a_{\ell}=b_{\ell}=0. Since \ell is arbitrary, the lemma is proved. ∎

We will also need the following proposition.

Proposition 4.2.

Suppose that the Lissajous curve given by Equation (2.1) lies in 𝕊3\mathbb{S}^{3}.

  • (a)

    If ω10\omega_{1}\neq 0 and 𝜸{\boldsymbol{\gamma}} has constant speed, then the coefficient vectors of 𝜸{\boldsymbol{\gamma}} satisfy the following relations:

    1. (1)

      𝐀1,𝐁1,𝐀2,𝐁2\mathbf{A}_{1},\mathbf{B}_{1},\mathbf{A}_{2},\mathbf{B}_{2} are orthogonal

    2. (2)

      |𝐀1|=|𝐁1|\left|\mathbf{A}_{1}\right|=\left|\mathbf{B}_{1}\right| and |𝐀2|=|𝐁2|\left|\mathbf{A}_{2}\right|=\left|\mathbf{B}_{2}\right|

    3. (3)

      |𝐀1|2+|𝐀2|2=1|\mathbf{A}_{1}|^{2}+|\mathbf{A}_{2}|^{2}=1.

  • (b)

    If ω1=0\omega_{1}=0, then:

    1. (1)

      𝐀1,𝐀2,𝐁2\mathbf{A}_{1},\mathbf{A}_{2},\mathbf{B}_{2} are orthogonal

    2. (2)

      |𝐀2|=|𝐁2|\left|\mathbf{A}_{2}\right|=\left|\mathbf{B}_{2}\right|

    3. (3)

      |𝐀1|2+|𝐀2|2=1|\mathbf{A}_{1}|^{2}+|\mathbf{A}_{2}|^{2}=1.

Proof.

For use in the later portions of this proof, we will first compute |𝜸(t)|2\left|{\boldsymbol{\gamma}}(t)\right|^{2}. Using Equation (2.1) and the fact that 𝜸{\boldsymbol{\gamma}} lies in 𝕊3\mathbb{S}^{3}, for each tt we have

𝜸(t)𝜸(t)=1=\displaystyle{\boldsymbol{\gamma}}(t)\cdot{\boldsymbol{\gamma}}(t)=1= |𝐀1|2cos2(ω1t)+|𝐁1|2sin2(ω1t)+|𝐀2|2cos2(ω2t)+|𝐁2|2sin2(ω2t)\displaystyle|\mathbf{A}_{1}|^{2}\cos^{2}(\omega_{1}t)+|\mathbf{B}_{1}|^{2}\sin^{2}(\omega_{1}t)+|\mathbf{A}_{2}|^{2}\cos^{2}(\omega_{2}t)+|\mathbf{B}_{2}|^{2}\sin^{2}(\omega_{2}t)
+2(𝐀1𝐁1)cos(ω1t)sin(ω1t)+2(𝐀1𝐀2)cos(ω1t)cos(ω2t)\displaystyle+2(\mathbf{A}_{1}\cdot\mathbf{B}_{1})\cos(\omega_{1}t)\sin(\omega_{1}t)+2(\mathbf{A}_{1}\cdot\mathbf{A}_{2})\cos(\omega_{1}t)\cos(\omega_{2}t)
+2(𝐀1𝐁2)cos(ω1t)sin(ω2t)+2(𝐁1𝐀2)sin(ω1t)cos(ω2t)\displaystyle+2(\mathbf{A}_{1}\cdot\mathbf{B}_{2})\cos(\omega_{1}t)\sin(\omega_{2}t)+2(\mathbf{B}_{1}\cdot\mathbf{A}_{2})\sin(\omega_{1}t)\cos(\omega_{2}t)
+2(𝐁1𝐁2)sin(ω1t)sin(ω2t)+2(𝐀2𝐁2)cos(ω2t)sin(ω2t)\displaystyle+2(\mathbf{B}_{1}\cdot\mathbf{B}_{2})\sin(\omega_{1}t)\sin(\omega_{2}t)+2(\mathbf{A}_{2}\cdot\mathbf{B}_{2})\cos(\omega_{2}t)\sin(\omega_{2}t)
=\displaystyle= (|𝐀1|2+|𝐁1|2+|𝐀2|2+|𝐁2|22)+(|𝐀1|2|𝐁1|22)cos(2ω1t)\displaystyle\left(\frac{|\mathbf{A}_{1}|^{2}+|\mathbf{B}_{1}|^{2}+|\mathbf{A}_{2}|^{2}+|\mathbf{B}_{2}|^{2}}{2}\right)+\left(\frac{|\mathbf{A}_{1}|^{2}-|\mathbf{B}_{1}|^{2}}{2}\right)\cos{(2\omega_{1}t)}
+(𝐀1𝐁1)sin(2ω1t)+(|𝐀2|2|𝐁2|22)cos(2ω2t)+(𝐀2𝐁2)sin(2ω2t)\displaystyle+\left(\mathbf{A}_{1}\cdot\mathbf{B}_{1}\right)\sin{(2\omega_{1}t)}+\left(\frac{|\mathbf{A}_{2}|^{2}-|\mathbf{B}_{2}|^{2}}{2}\right)\cos{(2\omega_{2}t)}+\left(\mathbf{A}_{2}\cdot\mathbf{B}_{2}\right)\sin{(2\omega_{2}t)}
+(𝐀1𝐀2𝐁1𝐁2)cos((ω1+ω2)t)+(𝐀1𝐁2+𝐁1𝐀2)sin((ω1+ω2)t)\displaystyle+\left(\mathbf{A}_{1}\cdot\mathbf{A}_{2}-\mathbf{B}_{1}\cdot\mathbf{B}_{2}\right)\cos{((\omega_{1}+\omega_{2})t)}+\left(\mathbf{A}_{1}\cdot\mathbf{B}_{2}+\mathbf{B}_{1}\cdot\mathbf{A}_{2}\right)\sin{((\omega_{1}+\omega_{2})t)}
+(𝐁1𝐁2+𝐀1𝐀2)cos((ω2ω1)t)+(𝐀1𝐁2𝐁1𝐀2)sin((ω2ω1)t)\displaystyle+\left(\mathbf{B}_{1}\cdot\mathbf{B}_{2}+\mathbf{A}_{1}\cdot\mathbf{A}_{2}\right)\cos{((\omega_{2}-\omega_{1})t)}+\left(\mathbf{A}_{1}\cdot\mathbf{B}_{2}-\mathbf{B}_{1}\cdot\mathbf{A}_{2}\right)\sin{((\omega_{2}-\omega_{1})t)} (4.4)

First, we prove part (a) of the proposition. Let u begin by assuming that ω23ω1\omega_{2}\neq 3\omega_{1}. Since ω23ω1\omega_{2}\neq 3\omega_{1} and ω10\omega_{1}\neq 0, we see that the five numbers

0, 2ω1, 2ω2,ω2ω1, and ω2+ω10,\ 2\omega_{1},\ 2\omega_{2},\ \omega_{2}-\omega_{1},\ \text{ and }\ \omega_{2}+\omega_{1}

are all distinct. By Lemma 4.1, each of the coefficients in the expression for 𝜸(t)𝜸(t)1{\boldsymbol{\gamma}}(t)\cdot{\boldsymbol{\gamma}}(t)-1 vanishes, as in the left hand side of Equation (4.1). Thus, from the coefficients of Equation (4.4), we obtain:

|𝐀1|2+|𝐁1|2+|𝐀2|2+|𝐁2|22=1|𝐀1|2|𝐁1|22=0𝐀1𝐁1=0|𝐀2|2|𝐁2|22=0𝐀2𝐁2=0𝐀1𝐀2𝐁1𝐁2=0𝐀1𝐁2+𝐁1𝐀2=0𝐁1𝐁2+𝐀1𝐀2=0𝐀1𝐁2𝐁1𝐀2=0}\displaystyle\left.\begin{aligned} &\frac{|\mathbf{A}_{1}|^{2}+|\mathbf{B}_{1}|^{2}+|\mathbf{A}_{2}|^{2}+|\mathbf{B}_{2}|^{2}}{2}=1\\ &\cfrac{|\mathbf{A}_{1}|^{2}-|\mathbf{B}_{1}|^{2}}{2}=0\\ &\mathbf{A}_{1}\cdot\mathbf{B}_{1}=0\\ &\cfrac{|\mathbf{A}_{2}|^{2}-|\mathbf{B}_{2}|^{2}}{2}=0\\ &\mathbf{A}_{2}\cdot\mathbf{B}_{2}=0\\ &\mathbf{A}_{1}\cdot\mathbf{A}_{2}-\mathbf{B}_{1}\cdot\mathbf{B}_{2}=0\\ &\mathbf{A}_{1}\cdot\mathbf{B}_{2}+\mathbf{B}_{1}\cdot\mathbf{A}_{2}=0\\ &\mathbf{B}_{1}\cdot\mathbf{B}_{2}+\mathbf{A}_{1}\cdot\mathbf{A}_{2}=0\\ &\mathbf{A}_{1}\cdot\mathbf{B}_{2}-\mathbf{B}_{1}\cdot\mathbf{A}_{2}=0\end{aligned}\right\}

which yields the following:

|𝐀1|2+|𝐁1|2+|𝐀2|2+|𝐁2|2=2\displaystyle|\mathbf{A}_{1}|^{2}+|\mathbf{B}_{1}|^{2}+|\mathbf{A}_{2}|^{2}+|\mathbf{B}_{2}|^{2}=2 (4.5)
|𝐀1|=|𝐁1|\displaystyle|\mathbf{A}_{1}|=|\mathbf{B}_{1}| (4.6)
|𝐀2|=|𝐁2|\displaystyle|\mathbf{A}_{2}|=|\mathbf{B}_{2}| (4.7)
𝐀1𝐁1=𝐀1𝐀2=𝐀1𝐁2=𝐁1𝐀2=𝐁1𝐁2=𝐀2𝐁2=0\displaystyle\mathbf{A}_{1}\cdot\mathbf{B}_{1}=\mathbf{A}_{1}\cdot\mathbf{A}_{2}=\mathbf{A}_{1}\cdot\mathbf{B}_{2}=\mathbf{B}_{1}\cdot\mathbf{A}_{2}=\mathbf{B}_{1}\cdot\mathbf{B}_{2}=\mathbf{A}_{2}\cdot\mathbf{B}_{2}=0 (4.8)

Equation (4.8) shows that the vectors 𝐀1,𝐁1,𝐀2,and 𝐁2\mathbf{A}_{1},\mathbf{B}_{1},\mathbf{A}_{2},\ \text{and }\mathbf{B}_{2} are orthogonal, which is conclusion 1 of the proposition. Moreover, Equations (4.6) and (4.7) are precisely conclusion 2 of the proposition. Further, recognize that by using Equations (4.5-4.7), we get

|𝐀1|2+|𝐀2|2=1\displaystyle|\mathbf{A}_{1}|^{2}+|\mathbf{A}_{2}|^{2}=1 (4.9)

which is conclusion 3 of the proposition.

To complete the proof of part (a) of the proposition, we now consider the case when ω2=3ω1\omega_{2}=3\omega_{1}. Let us set ω1=ω\omega_{1}=\omega and thus, ω2=3ω\omega_{2}=3\omega. Therefore, we have

2ω1=2ω,2ω2=6ω,ω2ω1=2ω, and ω2+ω1=4ω.2\omega_{1}=2\omega,\quad 2\omega_{2}=6\omega,\quad\omega_{2}-\omega_{1}=2\omega,\text{ and }\omega_{2}+\omega_{1}=4\omega.

Notice, 2ω1=ω2ω1=2ω2\omega_{1}=\omega_{2}-\omega_{1}=2\omega, so in Equation (4.4) there are only 4. Therefore, the relation 𝜸(t)𝜸(t)=1{\boldsymbol{\gamma}}(t)\cdot{\boldsymbol{\gamma}}(t)=1 applied to Equation (4.4) and Lemma 4.1 now give,

|A1|2+|A2|2+|B1|2+|B2|2=2\displaystyle|\textbf{A}_{1}|^{2}+|\textbf{A}_{2}|^{2}+|\textbf{B}_{1}|^{2}+|\textbf{B}_{2}|^{2}=2 (4.10)
|A1|2|B1|2+2(A1A2+B1B2)=0\displaystyle|\textbf{A}_{1}|^{2}-|\textbf{B}_{1}|^{2}+2(\textbf{A}_{1}\cdot\textbf{A}_{2}+\textbf{B}_{1}\cdot\textbf{B}_{2})=0 (4.11)
|A2|=|B2|\displaystyle|\textbf{A}_{2}|=|\textbf{B}_{2}| (4.12)
A1B1+A1B2B1A2=0\displaystyle\textbf{A}_{1}\cdot\textbf{B}_{1}+\textbf{A}_{1}\cdot\textbf{B}_{2}-\textbf{B}_{1}\cdot\textbf{A}_{2}=0 (4.13)
A2B2=0\displaystyle\textbf{A}_{2}\cdot\textbf{B}_{2}=0 (4.14)
A1A2=B1B2\displaystyle\textbf{A}_{1}\cdot\textbf{A}_{2}=\textbf{B}_{1}\cdot\textbf{B}_{2} (4.15)
A1B2=B1A2\displaystyle\textbf{A}_{1}\cdot\textbf{B}_{2}=-\textbf{B}_{1}\cdot\textbf{A}_{2} (4.16)

Since 𝜸{\boldsymbol{\gamma}} has constant speed, there exists a C>0C>0 such that for all tt, we have |𝜸(t)|=C\left|{\boldsymbol{\gamma}}^{\prime}(t)\right|=C. The relation 𝜸(t)𝜸(t)=C2{\boldsymbol{\gamma}}^{\prime}(t)\cdot{\boldsymbol{\gamma}}^{\prime}(t)=C^{2} yields (using Equation (4.4))

0=\displaystyle 0= (ω2|𝐁1|2+9ω2|𝐁2|2+ω2|𝐀1|2+9ω2|𝐀2|22C2)\displaystyle\left(\frac{\omega^{2}|\mathbf{B}_{1}|^{2}+9\omega^{2}|\mathbf{B}_{2}|^{2}+\omega^{2}|\mathbf{A}_{1}|^{2}+9\omega^{2}|\mathbf{A}_{2}|^{2}}{2}-C^{2}\right)
+(ω2|𝐁1|2ω2|𝐀1|22+3ω2𝐁1𝐁2+3ω2𝐀1𝐀2)cos(2ωt)\displaystyle+\left(\frac{\omega^{2}|\mathbf{B}_{1}|^{2}-\omega^{2}|\mathbf{A}_{1}|^{2}}{2}+3\omega^{2}\mathbf{B}_{1}\cdot\mathbf{B}_{2}+3\omega^{2}\mathbf{A}_{1}\cdot\mathbf{A}_{2}\right)\cos(2\omega t)
+9ω2(|𝐁2|2|𝐀2|22)cos(6ωt)(ω2(𝐀1𝐁1)+3ω2(𝐁1𝐀2)3ω2(𝐁2𝐀1))sin(2ωt)\displaystyle+9\omega^{2}\left(\frac{|\mathbf{B}_{2}|^{2}-|\mathbf{A}_{2}|^{2}}{2}\right)\cos(6\omega t)-(\omega^{2}(\mathbf{A}_{1}\cdot\mathbf{B}_{1})+3\omega^{2}(\mathbf{B}_{1}\cdot\mathbf{A}_{2})-3\omega^{2}(\mathbf{B}_{2}\cdot\mathbf{A}_{1}))\sin(2\omega t)
9ω2(𝐀2𝐁2)sin(6ωt)+3ω2(𝐁1𝐁2𝐀1𝐀2)cos(4ωt)3ω2(𝐁1𝐀2+𝐀1𝐁2)sin(4ωt)\displaystyle-9\omega^{2}(\mathbf{A}_{2}\cdot\mathbf{B}_{2})\sin(6\omega t)+3\omega^{2}(\mathbf{B}_{1}\cdot\mathbf{B}_{2}-\mathbf{A}_{1}\cdot\mathbf{A}_{2})\cos(4\omega t)-3\omega^{2}(\mathbf{B}_{1}\cdot\mathbf{A}_{2}+\mathbf{A}_{1}\cdot\mathbf{B}_{2})\sin(4\omega t) (4.17)

Using Lemma 4.1, this gives the relations

ω2(|𝐁1|2+9|𝐁2|2+|𝐀1|2+9|𝐀2|2)=2C2\displaystyle\omega^{2}(|\mathbf{B}_{1}|^{2}+9|\mathbf{B}_{2}|^{2}+|\mathbf{A}_{1}|^{2}+9|\mathbf{A}_{2}|^{2})=2C^{2} (4.18)
|𝐁1|2|𝐀1|2+6(𝐁1𝐁2+𝐀1𝐀2)=0\displaystyle|\mathbf{B}_{1}|^{2}-|\mathbf{A}_{1}|^{2}+6(\mathbf{B}_{1}\cdot\mathbf{B}_{2}+\mathbf{A}_{1}\cdot\mathbf{A}_{2})=0 (4.19)
|𝐁2|=|𝐀2|\displaystyle|\mathbf{B}_{2}|=|\mathbf{A}_{2}| (4.20)
𝐀1𝐁1+3(𝐁1𝐀2𝐁2𝐀1)=0\displaystyle\mathbf{A}_{1}\cdot\mathbf{B}_{1}+3(\mathbf{B}_{1}\cdot\mathbf{A}_{2}-\mathbf{B}_{2}\cdot\mathbf{A}_{1})=0 (4.21)
𝐀2𝐁2=0\displaystyle\mathbf{A}_{2}\cdot\mathbf{B}_{2}=0 (4.22)
𝐁1𝐁2=𝐀1𝐀2\displaystyle\mathbf{B}_{1}\cdot\mathbf{B}_{2}=\mathbf{A}_{1}\cdot\mathbf{A}_{2} (4.23)
𝐁1𝐀2=𝐀1𝐁2\displaystyle\mathbf{B}_{1}\cdot\mathbf{A}_{2}=-\mathbf{A}_{1}\cdot\mathbf{B}_{2} (4.24)

Comparing Equations (4.10 - 4.16) and Equations (4.18 - 4.24), we see that we have obtained three new relations, which are Equations (4.18), (4.19) and (4.21). Combining Equations (4.11) with (4.19) we see that

|𝐀1|=|𝐁1| and 𝐀1𝐀2=𝐁1𝐁2.\left|\mathbf{A}_{1}\right|=\left|\mathbf{B}_{1}\right|\text{ and }\mathbf{A}_{1}\cdot\mathbf{A}_{2}=-\mathbf{B}_{1}\cdot\mathbf{B}_{2}.

Similarly, Equations (4.13) and (4.21) imply that

𝐀1𝐁1=0 and 𝐀1𝐁2=𝐁1𝐀2.\mathbf{A}_{1}\cdot\mathbf{B}_{1}=0\text{ and }\mathbf{A}_{1}\cdot\mathbf{B}_{2}=\mathbf{B}_{1}\cdot\mathbf{A}_{2}.

Combining these last two relations with Equations (4.10 - 4.16), we get that the coefficient vectors 𝐀1,𝐁1,𝐀2,𝐁2\mathbf{A}_{1},\mathbf{B}_{1},\mathbf{A}_{2},\mathbf{B}_{2} are mutually orthogonal, |𝐀1|=|𝐁1|\left|\mathbf{A}_{1}\right|=\left|\mathbf{B}_{1}\right|, and |𝐀2|=|𝐁2|\left|\mathbf{A}_{2}\right|=\left|\mathbf{B}_{2}\right|. Conclusions (1), (2) and (3) follow again.

Now we will prove part (b) of the proposition. We set ω1=0\omega_{1}=0 and ω2=ω\omega_{2}=\omega in Equation (4.4). Then in Equation (4.4) we have the following three distinct frequencies:

0=2ω1,2ω2=2ω,ω2+ω1=ω2ω1=ω0=2\omega_{1},\quad 2\omega_{2}=2\omega,\quad\omega_{2}+\omega_{1}=\omega_{2}-\omega_{1}=\omega

Therefore, by Lemma 4.1 we have

2|𝐀1|2+|𝐀2|2+|𝐁2|2=2\displaystyle 2\left|\mathbf{A}_{1}\right|^{2}+\left|\mathbf{A}_{2}\right|^{2}+\left|\mathbf{B}_{2}\right|^{2}=2 (4.25)
|𝐀2|=|𝐁2|\displaystyle\left|\mathbf{A}_{2}\right|=\left|\mathbf{B}_{2}\right| (4.26)
𝐀1𝐀2=0,𝐀1𝐁2=0, and 𝐀2𝐁2=0\displaystyle\mathbf{A}_{1}\cdot\mathbf{A}_{2}=0,\ \mathbf{A}_{1}\cdot\mathbf{B}_{2}=0,\text{ and }\mathbf{A}_{2}\cdot\mathbf{B}_{2}=0

Equations (4.25) and (4.26) are precisely conclusions 1 and 2 of part (b) of the proposition. Additionally, combining Equations (4.25), and (4.26) we have conclusion 3 of part (b) of the proposition.

In connection with the proof of part (a) of the above proposition, we note that if ω2=3ω1\omega_{2}=3\omega_{1}, one can construct Lissajous curves in 𝕊3\mathbb{S}^{3} (of non-constant speed) for which the coefficient vectors are not orthogonal.

5. Proof of Theorem 1

5.1. Part 1

Adjoining the relation 𝜸(t)=𝐓(t){\boldsymbol{\gamma}}^{\prime}(t)=\mathbf{T}(t) to the Frenet-Serret equations (Equation (3.6)) we obtain the system of equations

𝜸(t)\displaystyle{\boldsymbol{\gamma}}^{\prime}(t) =\displaystyle= 𝐓(t)\displaystyle\phantom{-}\phantom{\kappa}\mathbf{T}(t)
𝐓(t)\displaystyle\mathbf{T}^{\prime}(t) =\displaystyle= 𝜸(t)\displaystyle-{\boldsymbol{\gamma}}(t) κ𝐍(t)\displaystyle-\kappa\mathbf{N}(t)
𝐍(t)\displaystyle\mathbf{N}^{\prime}(t) =\displaystyle= κ𝐓(t)\displaystyle-\kappa\mathbf{T}(t) +τ𝐁(t)\displaystyle+\tau\mathbf{B}(t) (5.1)
𝐁(t)\displaystyle\mathbf{B}^{\prime}(t) =\displaystyle= τ𝐍(t)\displaystyle-\tau\mathbf{N}(t)

where κ\kappa and τ\tau are the given constants. We now rewrite these equations in matrix form. Note that the four vectors 𝜸(t),𝐓(t),𝐍(t),𝐁(t){\boldsymbol{\gamma}}(t),\mathbf{T}(t),\mathbf{N}(t),\mathbf{B}(t) form an orthonormal basis of 4\mathbb{R}^{4}. Let 𝐗(t)\mathbf{X}(t) denote the 4×44\times 4 matrix whose rows are these four vectors. Then for each tt, the matrix 𝐗(t)\mathbf{X}(t) is orthogonal. When the curvature κ\kappa and the torsion τ\tau are constants, the augmented Frenet-Serret equations given by the set of Equations in (5.1) in the sphere Equation may be written in matrix form as

𝐗(t)=𝐂𝐗(t)\displaystyle\mathbf{X}^{\prime}(t)=\mathbf{C}\cdot\mathbf{X}(t) (5.2)

where 𝐂\mathbf{C} denotes the skew-symmetric matrix

𝐂=[010010κ00κ0τ00τ0]\displaystyle\mathbf{C}=\begin{bmatrix}0&1&0&0\\ -1&0&\kappa&0\\ 0&-\kappa&0&\tau\\ 0&0&-\tau&0\end{bmatrix} (5.3)

From the theory of ordinary differential equations we know that the solution to the constant coefficient system presented in Equation (5.2) exists for all tt and is given by

𝐗(t)=et𝐂𝐗(0).\displaystyle\mathbf{X}(t)=e^{t\mathbf{C}}\cdot\mathbf{X}(0). (5.4)

This proves part 1 of the theorem.

5.2. Part 2

In order to calculate the matrix exponential et𝐂e^{t\mathbf{C}}, we first recognize that since 𝐂\mathbf{C} is skew-symetric, it can be diagonalized and thus written as

𝐂=𝐏(i𝐃)𝐏1\displaystyle\mathbf{C}=\mathbf{P}\left(i\mathbf{D}\right)\mathbf{P}^{-1}

where 𝐃\mathbf{D} is a diagonal matrix with real entries of the form

𝐃=diag(ω1,ω1,ω2,ω2),\displaystyle\mathbf{D}={\rm diag}(\omega_{1},-\omega_{1},\omega_{2},-\omega_{2}), (5.5)

where ω1,ω20\omega_{1},\omega_{2}\geq 0. This is because the eigenvalues of the real skew-symmetric matrix 𝐂\mathbf{C} are purely imaginary and occur in complex conjugate pairs. Therefore,

et𝐂=𝐏eit𝐃𝐏1.\displaystyle e^{t\mathbf{C}}=\mathbf{P}e^{it\mathbf{D}}\mathbf{P}^{-1}. (5.6)

From Equation (5.5) it follows that

eit𝐃=diag(eiω1t,eiω1t,eiω2t,eiω2t).e^{it\mathbf{D}}={\rm diag}(e^{i\omega_{1}t},e^{-i\omega_{1}t},e^{i\omega_{2}t},e^{-i\omega_{2}t}).

Since 𝜸(t){\boldsymbol{\gamma}}(t) is the first row of the matrix 𝐗(t)=𝐏eit𝐃𝐏1𝐗(0)\mathbf{X}(t)=\mathbf{P}e^{it\mathbf{D}}\mathbf{P}^{-1}\cdot\mathbf{X}(0) it follows that

𝜸(t)=cos(ω1t)𝐀1+sin(ω1t)𝐁1+cos(ω2t)𝐀2+sin(ω2t)𝐁2{\boldsymbol{\gamma}}(t)=\cos{(\omega_{1}t)}\mathbf{A}_{1}+\sin{(\omega_{1}t)}\mathbf{B}_{1}+\cos{(\omega_{2}t)}\mathbf{A}_{2}+\sin{(\omega_{2}t)}\mathbf{B}_{2}

where the coefficient vectors 𝐀1,𝐁1,𝐀2, and 𝐁2\mathbf{A}_{1},\mathbf{B}_{1},\mathbf{A}_{2},\text{ and }\mathbf{B}_{2} are constant vectors in 4\mathbb{R}^{4}. This proves part 2 of the theorem.

5.3. Part 3

The diagonal entries of i𝐃i\mathbf{D}, where 𝐃\mathbf{D} is as in Equation (5.5), are the eigenvalues of the matrix 𝐂\mathbf{C} of Equation (5.3). We find them by solving the characteristic equation

det(𝐂x𝐈)=x4+(κ2+τ2+1)x2+τ2=x4+χ2x2+τ2=0\displaystyle\det\left(\mathbf{C}-x\mathbf{I}\right)=x^{4}+(\kappa^{2}+\tau^{2}+1)x^{2}+\tau^{2}=x^{4}+\chi^{2}x^{2}+\tau^{2}=0 (5.7)

with χ\chi as in Equation (2.4). The solutions of the characteristic equation are

x=±iω1 or x=±iω2,x=\pm i\omega_{1}\text{ or }x=\pm i\omega_{2},

where ω1,ω2\omega_{1},\omega_{2} are as in Equations (2.2) and (2.3). This proves part 3 of the theorem.

5.4. Part 4

Since κ>0\kappa>0, we have

χ44τ2\displaystyle\chi^{4}-4\tau^{2} =(κ2+τ2+1)24τ2\displaystyle=(\kappa^{2}+\tau^{2}+1)^{2}-4\tau^{2}
=κ4+(τ21)2+2κ2τ2+2κ2\displaystyle=\kappa^{4}+(\tau^{2}-1)^{2}+2\kappa^{2}\tau^{2}+2\kappa^{2}
>(τ21)2.\displaystyle>(\tau^{2}-1)^{2}. (5.8)

Therefore, from Equations (2.2) and (2.3) we see that ω2>ω1\omega_{2}>\omega_{1}. by definition of χ\chi in (2.4) we have,

χ2=κ2+τ2+1>1+τ2.\displaystyle\chi^{2}=\kappa^{2}+\tau^{2}+1>1+\tau^{2}. (5.9)

Combining Equations (5.8) and (5.9) we have

χ2+χ44τ2\displaystyle\chi^{2}+\sqrt{\chi^{4}-4\tau^{2}} >(1+τ2)+(τ21)2\displaystyle>(1+\tau^{2})+\sqrt{(\tau^{2}-1)^{2}} (5.10)
=1+τ2+|τ21|\displaystyle=1+\tau^{2}+|\tau^{2}-1|
={2τ2,τ12,τ<1\displaystyle=\begin{cases}2\tau^{2},&\tau\geq 1\\ 2,&\tau<1\end{cases} (5.11)

Therefore, χ2+χ44τ2>2\chi^{2}+\sqrt{\chi^{4}-4\tau^{2}}>2, and we have

ω22=χ2+χ44τ22>1.\displaystyle\omega_{2}^{2}=\frac{\chi^{2}+\sqrt{\chi^{4}-4\tau^{2}}}{2}>1. (5.12)

Then, by making use of Equation (5.11),

ω12=χ2χ44τ22=2τ2χ2+χ24τ2<{1,τ1τ2,τ<1\displaystyle\omega_{1}^{2}=\frac{\chi^{2}-\sqrt{\chi^{4}-4\tau^{2}}}{2}=\frac{2\tau^{2}}{\chi^{2}+\sqrt{\chi^{2}-4\tau^{2}}}<\begin{cases}1,&\tau\geq 1\\ \tau^{2},&\tau<1\end{cases} (5.13)

Thus,

ω1<1\displaystyle\omega_{1}<1 (5.14)

This proves part 4 of the theorem.

5.5. Part 5

Note that |𝜸(t)|=1\left|{\boldsymbol{\gamma}}(t)\right|=1 for all tt and |𝜸(t)|=1\left|{\boldsymbol{\gamma}}^{\prime}(t)\right|=1 for all tt as well, since 𝜸{\boldsymbol{\gamma}} lies in 𝕊3\mathbb{S}^{3} and has unit speed. Since τ0\tau\neq 0 by Equation (2.2), we know that ω1>0\omega_{1}>0. Therefore, by part (a) of Proposition 4.2,

  1. (1)

    The coefficient vectors 𝐀1,𝐁1,𝐀2,𝐁2\mathbf{A}_{1},\mathbf{B}_{1},\mathbf{A}_{2},\mathbf{B}_{2} are mutually orthogonal, which is one of the conclusions of part 5 of Theorem 1,

  2. (2)

    |𝐀1|=|𝐁1| and |𝐀2|=|𝐁2|\left|\mathbf{A}_{1}\right|=\left|\mathbf{B}_{1}\right|\text{ and }\left|\mathbf{A}_{2}\right|=\left|\mathbf{B}_{2}\right| which is part of the content of Equations (2.6) and (2.7). We will however need to work further to obtain the remaining content of these equations and to therefore complete the proof of part 5,

  3. (3)

    We have that

    |𝐀1|2+|𝐀2|2=1.\displaystyle\left|\mathbf{A}_{1}\right|^{2}+\left|\mathbf{A}_{2}\right|^{2}=1. (5.15)

Now, let 𝜶(t)=𝜸(t){\boldsymbol{\alpha}}(t)={\boldsymbol{\gamma}}^{\prime}(t). Then, |𝜶(t)|=1\left|{\boldsymbol{\alpha}}(t)\right|=1 (since 𝜸(t){\boldsymbol{\gamma}}(t) is parameterized by arc length) and differentiating Equation (2.1), we see that 𝜶(t){\boldsymbol{\alpha}}(t) may be represented as

𝜶(t)\displaystyle{\boldsymbol{\alpha}}(t) =ω1cos(ω1t)𝐁1ω1sin(ω1t)𝐀1+ω2cos(ω2t)𝐁2ω2sin(ω2t)𝐀2\displaystyle=\omega_{1}\cos(\omega_{1}t)\mathbf{B}_{1}-\omega_{1}\sin(\omega_{1}t)\mathbf{A}_{1}+\omega_{2}\cos(\omega_{2}t)\mathbf{B}_{2}-\omega_{2}\sin(\omega_{2}t)\mathbf{A}_{2}
=cos(ω1t)𝐏1+sin(ω1t)𝐐1+cos(ω2t)𝐏2+sin(ω2t)𝐐2,\displaystyle=\cos(\omega_{1}t)\mathbf{P}_{1}+\sin(\omega_{1}t)\mathbf{Q}_{1}+\cos(\omega_{2}t)\mathbf{P}_{2}+\sin(\omega_{2}t)\mathbf{Q}_{2},

where 𝐏1=ω1𝐁1\mathbf{P}_{1}=\omega_{1}\mathbf{B}_{1}, 𝐐1=ω1𝐀1\mathbf{Q}_{1}=-\omega_{1}\mathbf{A}_{1}, 𝐏2=ω2𝐁2\mathbf{P}_{2}=\omega_{2}\mathbf{B}_{2}, and 𝐐2=ω2𝐀2\mathbf{Q}_{2}=-\omega_{2}\mathbf{A}_{2}. This shows that 𝜶{\boldsymbol{\alpha}} is a Lissajous curve in 𝕊3\mathbb{S}^{3}.

Now, we claim that |𝜶(t)|\left|{\boldsymbol{\alpha}}^{\prime}(t)\right| is constant independently of tt. Recall, that

𝜶(t)=𝜸′′(t)=𝐓(t){\boldsymbol{\alpha}}^{\prime}(t)={\boldsymbol{\gamma}}^{\prime\prime}(t)=\mathbf{T}^{\prime}(t)

where 𝐓\mathbf{T} is the tangent vector field in the Frenet Frame (𝐓,𝐍,𝐁)(\mathbf{T,N,B}). Therefore, by the first equation in (3.6), we have

𝜸′′(t)=κ𝐍(t)𝜸(t),{\boldsymbol{\gamma}}^{\prime\prime}(t)=-\kappa\mathbf{N}(t)-{\boldsymbol{\gamma}}(t),

which yields,

𝜶(t)𝜶(t)\displaystyle{\boldsymbol{\alpha}}^{\prime}(t)\cdot{\boldsymbol{\alpha}}^{\prime}(t) =κ2+1+2κ(𝐍(t)𝜸(t))\displaystyle=\kappa^{2}+1+2\kappa(\mathbf{N}(t)\cdot{\boldsymbol{\gamma}}(t))
=κ2+1\displaystyle=\kappa^{2}+1 (5.16)

where the term 𝐍(t)𝜸(t)=0\mathbf{N}(t)\cdot{\boldsymbol{\gamma}}(t)=0 since 𝐍(t)T𝜸(t)𝕊3\mathbf{N}(t)\in T_{{\boldsymbol{\gamma}}(t)}\mathbb{S}^{3}. We may now apply conclusion 3 of part (a) of Proposition 4.2 to obtain that |𝐏1|2+|𝐏2|2=1\left|\mathbf{P}_{1}\right|^{2}+\left|\mathbf{P}_{2}\right|^{2}=1, which is equivalent to the statement that

ω12|𝐀1|2+ω22|𝐀2|2=1.\displaystyle\omega_{1}^{2}\left|\mathbf{A}_{1}\right|^{2}+\omega_{2}^{2}\left|\mathbf{A}_{2}\right|^{2}=1. (5.17)

Combining Equations (5.15) and (5.17), we get

ω12|𝐀1|2+ω22(1|𝐀1|2)=1\omega_{1}^{2}\left|\mathbf{A}_{1}\right|^{2}+\omega_{2}^{2}(1-\left|\mathbf{A}_{1}\right|^{2})=1

and

ω12(1|𝐀2|2)+ω22|𝐀2|2=1.\omega_{1}^{2}(1-\left|\mathbf{A}_{2}\right|^{2})+\omega_{2}^{2}\left|\mathbf{A}_{2}\right|^{2}=1.

Solving these equations for |𝐀1|2\left|\mathbf{A}_{1}\right|^{2} and |𝐀2|2\left|\mathbf{A}_{2}\right|^{2}, we obtain Equations (2.6) and (2.7).

5.6. Part 6

If τ=0\tau=0, by Equation (2.2), we have ω1=0\omega_{1}=0. We set ω2=ω=κ2+1\omega_{2}=\omega=\sqrt{\kappa^{2}+1} by Equation (2.3) and then by parts 1,2, and 3 of this theorem, proved above, 𝜸{\boldsymbol{\gamma}} is given by

𝜸(t)=𝐀1+cos(ωt)𝐀2+sin(ωt)𝐁2.{\boldsymbol{\gamma}}(t)=\mathbf{A}_{1}+\cos(\omega t)\mathbf{A}_{2}+\sin(\omega t)\mathbf{B}_{2}.

Furthermore, by part (b) of Proposition 4.2 we know that |𝐀2|=|𝐁2|\left|\mathbf{A}_{2}\right|=\left|\mathbf{B}_{2}\right|, that the coefficient vectors 𝐀1,𝐀2, and 𝐁2\mathbf{A}_{1},\mathbf{A}_{2},\text{ and }\mathbf{B}_{2} are mutually orthogonal. Note that,

𝜸(t)=ωsin(ωt)𝐀2+ωcos(ωt)𝐁2.{\boldsymbol{\gamma}}^{\prime}(t)=-\omega\sin(\omega t)\mathbf{A}_{2}+\omega\cos(\omega t)\mathbf{B}_{2}.

Since 𝜸(t){\boldsymbol{\gamma}}^{\prime}(t) has unit speed, we have

𝜸(t)𝜸(t)=1=ω2|𝐀2|2sin2(ωt)+ω2|𝐁2|2cos2(ωt)=ω2|𝐀2|2,{\boldsymbol{\gamma}}^{\prime}(t)\cdot{\boldsymbol{\gamma}}^{\prime}(t)=1=\omega^{2}\left|\mathbf{A}_{2}\right|^{2}\sin^{2}(\omega t)+\omega^{2}\left|\mathbf{B}_{2}\right|^{2}\cos^{2}(\omega t)=\omega^{2}\left|\mathbf{A}_{2}\right|^{2},

which implies,

|𝐀2|=1ω.\displaystyle\left|\mathbf{A}_{2}\right|=\frac{1}{\omega}. (5.18)

Combining Equation (5.18) with conclusion 3 of part (b) of Proposition 4.2, we get that

|𝐀1|=11ω2.\displaystyle\left|\mathbf{A}_{1}\right|=\sqrt{1-\frac{1}{\omega^{2}}}. (5.19)

6. Proof of Theorem  2

6.1. Part 1

Suppose that we have two helices 𝜶{\boldsymbol{\alpha}} and 𝜷{\boldsymbol{\beta}} in 𝕊3\mathbb{S}^{3} with the same curvature κ0\kappa\geq 0 and torsion τ0\tau\geq 0. Then, by part 3 of Theorem 1, the fundamental angular frequencies ω1 and ω2\omega_{1}\text{ and }\omega_{2} of these two curves are the same. Thus, the curves are represented as

𝜶(t)\displaystyle{\boldsymbol{\alpha}}(t) =cos(ω1t)𝐀1+sin(ω1t)𝐁1+cos(ω2t)𝐀2+sin(ω2t)𝐁2\displaystyle=\cos(\omega_{1}t)\mathbf{A}_{1}+\sin(\omega_{1}t)\mathbf{B}_{1}+\cos(\omega_{2}t)\mathbf{A}_{2}+\sin(\omega_{2}t)\mathbf{B}_{2}
𝜷(t)\displaystyle{\boldsymbol{\beta}}(t) =cos(ω1t)𝐂1+sin(ω1t)𝐃1+cos(ω2t)𝐂2+sin(ω2t)𝐃2\displaystyle=\cos(\omega_{1}t)\mathbf{C}_{1}+\sin(\omega_{1}t)\mathbf{D}_{1}+\cos(\omega_{2}t)\mathbf{C}_{2}+\sin(\omega_{2}t)\mathbf{D}_{2}

If τ0\tau\neq 0, by part 5 of Theorem 1, we also know that

|𝐀1|\displaystyle\left|\mathbf{A}_{1}\right| =|𝐁1|=|𝐂1|=|𝐃1|=1ω22ω12ω22,\displaystyle=\left|\mathbf{B}_{1}\right|=\left|\mathbf{C}_{1}\right|=\left|\mathbf{D}_{1}\right|=\sqrt{\cfrac{1-\omega_{2}^{2}}{\omega_{1}^{2}-\omega_{2}^{2}}},
|𝐀2|\displaystyle\left|\mathbf{A}_{2}\right| =|𝐁2|=|𝐂2|=|𝐃2|=1ω12ω22ω12,\displaystyle=\left|\mathbf{B}_{2}\right|=\left|\mathbf{C}_{2}\right|=\left|\mathbf{D}_{2}\right|=\sqrt{\cfrac{1-\omega_{1}^{2}}{\omega_{2}^{2}-\omega_{1}^{2}}},

and that the sets of vectors {𝐀1,𝐁1,𝐀2,𝐁2}\{\mathbf{A}_{1},\mathbf{B}_{1},\mathbf{A}_{2},\mathbf{B}_{2}\} are mutually orthogonal and {𝐂1,𝐃1,𝐂2,𝐃2}\{\mathbf{C}_{1},\mathbf{D}_{1},\mathbf{C}_{2},\mathbf{D}_{2}\} are also mutually orthogonal. Therefore, there exists an orthogonal map, G:44G:\mathbb{R}^{4}\to\mathbb{R}^{4} in O(4)O(4) such that G(𝐀1)=𝐂1G(\mathbf{A}_{1})=\mathbf{C}_{1}, G(𝐁1)=𝐃1G(\mathbf{B}_{1})=\mathbf{D}_{1}, G(𝐀2)=𝐂2G(\mathbf{A}_{2})=\mathbf{C}_{2}, and G(𝐁2)=𝐃2G(\mathbf{B}_{2})=\mathbf{D}_{2}. Then f=G|𝕊3f=G|_{\mathbb{S}^{3}} is an isometry of 𝕊3\mathbb{S}^{3} and it is clear that 𝜷=f𝜶{\boldsymbol{\beta}}=f\circ{\boldsymbol{\alpha}}.

If τ=0\tau=0, then by part 6 of Theorem 1, we know that 𝜶{\boldsymbol{\alpha}} and 𝜷{\boldsymbol{\beta}} take the form

𝜶(t)\displaystyle{\boldsymbol{\alpha}}(t) =𝐀1+cos(ωt)𝐀2+sin(ωt)𝐁2\displaystyle=\mathbf{A}_{1}+\cos(\omega t)\mathbf{A}_{2}+\sin(\omega t)\mathbf{B}_{2}
𝜷(t)\displaystyle{\boldsymbol{\beta}}(t) =𝐂1+cos(ωt)𝐂2+sin(ωt)𝐃2\displaystyle=\mathbf{C}_{1}+\cos(\omega t)\mathbf{C}_{2}+\sin(\omega t)\mathbf{D}_{2}

where ω1=0\omega_{1}=0 and ω=ω2=κ2+1\omega=\omega_{2}=\sqrt{\kappa^{2}+1} by Equations (2.2) and (2.3). Furthermore, by part 6 of Theorem 1, we know that

|𝐀2|=|𝐁2|=|𝐂2|=|𝐃2|=1ω,|𝐀1|=|𝐂1|=11ω2,\displaystyle\left|\mathbf{A}_{2}\right|=\left|\mathbf{B}_{2}\right|=\left|\mathbf{C}_{2}\right|=\left|\mathbf{D}_{2}\right|=\frac{1}{\omega},\quad\left|\mathbf{A}_{1}\right|=\left|\mathbf{C}_{1}\right|=\sqrt{1-\cfrac{1}{\omega^{2}}},

and that the sets of vectors {𝐀1,𝐀2,𝐁2}\{\mathbf{A}_{1},\mathbf{A}_{2},\mathbf{B}_{2}\} are mutually orthogonal and {𝐂1,𝐂2,𝐃2}\{\mathbf{C}_{1},\mathbf{C}_{2},\mathbf{D}_{2}\} are also mutually orthogonal. Therefore, there again exists an orthogonal map, G:44G:\mathbb{R}^{4}\to\mathbb{R}^{4} in O(4)O(4) such that G(𝐀1)=𝐂1G(\mathbf{A}_{1})=\mathbf{C}_{1}, G(𝐀2)=𝐂𝟐G(\mathbf{A}_{2})=\mathbf{C_{2}}, G(𝐁2)=𝐃2G(\mathbf{B}_{2})=\mathbf{D}_{2}. Then f=G|𝕊3f=G|_{\mathbb{S}^{3}} is again an isometry of 𝕊3\mathbb{S}^{3} and it is clear that 𝜷=f𝜶{\boldsymbol{\beta}}=f\circ{\boldsymbol{\alpha}}.

6.2. Part 2

Let 𝜸{\boldsymbol{\gamma}} be a helix in 𝕊3\mathbb{S}^{3} which can be written, thanks to Theorem 1, in the form of Equation (2.1):

𝜸(t)=cos(ω1t)𝐀1+sin(ω1t)𝐁1+cos(ω2t)𝐀2+sin(ω2t)𝐁2{\boldsymbol{\gamma}}(t)=\cos(\omega_{1}t)\mathbf{A}_{1}+\sin(\omega_{1}t)\mathbf{B}_{1}+\cos(\omega_{2}t)\mathbf{A}_{2}+\sin(\omega_{2}t)\mathbf{B}_{2}

Now suppose that 𝜸{\boldsymbol{\gamma}} is periodic with period TT. Then, for each tt\in\mathbb{R}, we have

𝜸(t)=𝜸(t+T).{\boldsymbol{\gamma}}(t)={\boldsymbol{\gamma}}(t+T).

First, let us assume that τ0\tau\neq 0 and consequently, because of Equation (2.2), ω10\omega_{1}\neq 0. Then, comparing the coefficients of 𝐀1 and 𝐁1\mathbf{A}_{1}\text{ and }\mathbf{B}_{1}, we obtain,

cos(ω1(t+T))=cos(ω1t) and sin(ω1(t+T))=sin(ω1t)\displaystyle\cos(\omega_{1}(t+T))=\cos(\omega_{1}t)\quad\text{ and }\quad\sin(\omega_{1}(t+T))=\sin(\omega_{1}t) (6.1)

for each tt\in\mathbb{R}. This shows that there exists a non-zero mm\in\mathbb{Z} such that,

T=2πmω1.T=\frac{2\pi m}{\omega_{1}}.

Similarly, we compare the coefficients of 𝐀2 and 𝐁2\mathbf{A}_{2}\text{ and }\mathbf{B}_{2}, to get

cos(ω2(t+T))=cos(ω2t) and sin(ω2(t+T))=sin(ω2t)\displaystyle\cos(\omega_{2}(t+T))=\cos(\omega_{2}t)\quad\text{ and }\quad\sin(\omega_{2}(t+T))=\sin(\omega_{2}t) (6.2)

for each tt\in\mathbb{R}. This shows that there exists a non-zero nn\in\mathbb{Z} such that,

T=2πnω2.T=\frac{2\pi n}{\omega_{2}}.

It follows that

ω1ω2=mn\frac{\omega_{1}}{\omega_{2}}=\frac{m}{n}\in\mathbb{Q}

Now we prove the converse. Suppose that ω1ω2=mn\cfrac{\omega_{1}}{\omega_{2}}=\cfrac{m}{n}\in\mathbb{Q}. Then, let

T=2πmω1=2πnω2.T=\frac{2\pi m}{\omega_{1}}=\frac{2\pi n}{\omega_{2}}.

Then, Equations (6.1) and (6.2) hold. Therefore,

𝜸(t+T)=𝜸(t){\boldsymbol{\gamma}}(t+T)={\boldsymbol{\gamma}}(t)

Which proves part 2 of Theorem 2 in the case where τ0\tau\neq 0. If, on the other hand, τ=0\tau=0 and consequently ω1=0\omega_{1}=0, then 𝜸{\boldsymbol{\gamma}} is given by,

𝜸(t)=𝐀1+cos(ω2t)𝐀2+sin(ω2t)𝐁2.{\boldsymbol{\gamma}}(t)=\mathbf{A}_{1}+\cos(\omega_{2}t)\mathbf{A}_{2}+\sin(\omega_{2}t)\mathbf{B}_{2}.

Which is always periodic with a period of

T=2πω2.T=\frac{2\pi}{\omega_{2}}.

This proves part 2 of the theorem.

6.3. Part 3

Let κ,τ>0\kappa,\tau>0 and let 𝐚1,𝐛1,𝐚2, and 𝐛2\mathbf{a}_{1},\mathbf{b}_{1},\mathbf{a}_{2},\text{ and }\mathbf{b}_{2} be the orthonormal basis of 4\mathbb{R}^{4} consisting of the unit vectors along the coefficient vectors 𝐀1,𝐁1,𝐀2, and 𝐁2\mathbf{A}_{1},\mathbf{B}_{1},\mathbf{A}_{2},\text{ and }\mathbf{B}_{2} of 𝜸{\boldsymbol{\gamma}} as given in Equation (2.1). We denote the coordinates of a point 𝐱4\mathbf{x}\in\mathbb{R}^{4} by

𝐱=x1𝐚1+x2𝐛1+x3𝐚2+x4𝐛2.\mathbf{x}=x_{1}\mathbf{a}_{1}+x_{2}\mathbf{b}_{1}+x_{3}\mathbf{a}_{2}+x_{4}\mathbf{b}_{2}.

By parts 2 and 5 of Theorem 1, we have that 𝜸{\boldsymbol{\gamma}} is represented in these coordinates by x1=|𝐀1|cos(ω1t),x2=|𝐀1|sin(ω1t),x3=|𝐀2|cos(ω2t), and x4=|𝐀2|sin(ω2t)x_{1}=\left|\mathbf{A}_{1}\right|\cos(\omega_{1}t),x_{2}=\left|\mathbf{A}_{1}\right|\sin(\omega_{1}t),x_{3}=\left|\mathbf{A}_{2}\right|\cos(\omega_{2}t),\text{ and }x_{4}=\left|\mathbf{A}_{2}\right|\sin(\omega_{2}t). Consider the torus in 4\mathbb{R}^{4} given by

𝕋𝜸2={𝐱4:x12+x22=|𝐀1|2,x32+x42=|𝐀2|2}.\displaystyle\mathbb{T}^{2}_{{\boldsymbol{\gamma}}}=\left\{\mathbf{x}\in\mathbb{R}^{4}:x_{1}^{2}+x_{2}^{2}=\left|\mathbf{A}_{1}\right|^{2},x_{3}^{2}+x_{4}^{2}=\left|\mathbf{A}_{2}\right|^{2}\right\}. (6.3)

Clearly 𝜸{\boldsymbol{\gamma}} lies on 𝕋𝜸2\mathbb{T}^{2}_{{\boldsymbol{\gamma}}}. It is clear that 𝕋𝜸2\mathbb{T}^{2}_{{\boldsymbol{\gamma}}} is a flat Clifford torus in 4\mathbb{R}^{4}, and is contained in 𝕊3\mathbb{S}^{3}, since if 𝐱=(x1,x2,x3,x4)𝕋𝜸2\mathbf{x}=(x_{1},x_{2},x_{3},x_{4})\in\mathbb{T}^{2}_{{\boldsymbol{\gamma}}}, then

|𝐱|2=x12+x22+x32+x42\displaystyle\left|\mathbf{x}\right|^{2}=x_{1}^{2}+x_{2}^{2}+x_{3}^{2}+x_{4}^{2} =|𝐀1|2+|𝐀2|2\displaystyle=\left|\mathbf{A}_{1}\right|^{2}+\left|\mathbf{A}_{2}\right|^{2}
=1ω22ω12ω22+1ω12ω22ω12\displaystyle=\frac{1-\omega_{2}^{2}}{\omega_{1}^{2}-\omega_{2}^{2}}+\frac{1-\omega_{1}^{2}}{\omega_{2}^{2}-\omega_{1}^{2}}
=1,\displaystyle=1,

where the last equality follows by Equations (2.6) and (2.7).

6.4. Part 4

Note that the helix 𝜸{\boldsymbol{\gamma}} is a solution of the differential equations on the torus 𝕋𝜸2\mathbb{T}^{2}_{{\boldsymbol{\gamma}}}

dθ1dt=ω1 and dθ2dt=ω2\frac{d\theta_{1}}{dt}=\omega_{1}\text{ and }\frac{d\theta_{2}}{dt}=\omega_{2}

where θ1\theta_{1} (resp. θ2\theta_{2}) is an angular coordinate on the circle x12+x22=|𝐀1|2x_{1}^{2}+x_{2}^{2}=\left|\mathbf{A}_{1}\right|^{2} (resp. x32+x42=|𝐀2|2x_{3}^{2}+x_{4}^{2}=\left|\mathbf{A}_{2}\right|^{2}). If ω1ω2\cfrac{\omega_{1}}{\omega_{2}}\notin\mathbb{Q}, then a classical result in the theory of dynamical systems [1, Proposition 4.2.8, p. 113] shows that the image of 𝜸{\boldsymbol{\gamma}} is dense in 𝕋𝜸2\mathbb{T}^{2}_{{\boldsymbol{\gamma}}}. Therefore, part 3 of the theorem is proven.

References

  • [1] Hasselblatt, Boris; Katok, Anatole. “A first course in dynamics. With a panorama of recent developments.” Cambridge University Press, New York, 2003.
  • [2] Spivak, Michael. “A comprehensive introduction to differential geometry. Vol. I - V. Second edition. ”  Publish or Perish, Inc., Houston, Texas., 1999.