Curves of constant curvature and torsion in the 3-sphere
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, Helices2010 Mathematics Subject Classification:
53A351. Introduction
Let be a three dimensional Riemannian manifold, let be an open interval, and let be a smooth curve in , which we assume to be parametrized by the arc length . It is well-known that the local geometry of is characterized by the curvature and the torsion . These are functions defined along and are the coefficients of the well-known Frenet-Serret formulas [2, Vol. IV, pp. 21-23]:
(1.1) | ||||||
where the orthogonal unit vector fields , with , along the unit-speed curve , constitute its Frenet frame and denotes covariant differentiation along with respect to the arc length . We will assume that each of the functions and is either nowhere zero or vanishes identically. Additionally, if is identically zero, then 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 be a Riemannian manifold of dimension 3. A curve
, where is an open interval, will be called a helix (plural: helices) if its curvature and torsion are non-negative constants. A helix is non-degenerate if and are both positive, and degenerate otherwise. We say that the helix is periodic if there is a such that for each .
We take 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 , where non-degenerate helices are curves of the form:
(1.2) |
where , are non-zero and orthogonal with , and . 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 .
The aim of this paper is to study helices in the three dimensional round sphere . Thanks to the fact that is compact, we expect that a non-degenerate helix in 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 . Globally, a non-degenerate helix in has two fundamental angular frequencies, and , as opposed to the single fundamental angular frequency, , of the helix given by Equation (1.2) in . A non-degenerate helix in may be thought of as the trajectory of a particle which performs two superimposed circular motions with frequencies and . These fundamental angular frequencies must satisfy
a constraint which arises because a curve with non-zero curvature and torsion must lie in the positively curved compact space . Unlike in where non-degenerate helices are embedded, non-compact submanifolds, depending on the fundamental angular frequencies and , a non-degenerate helix in can either be periodic (when it is a compact embedded submanifold) or be dense in a flat 2-torus contained in (when the image of the helix is not an embedded submanifold of ). 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 , 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 to be embedded in the Euclidean space in the natural way as the hypersurface , and endow with the Riemannian metric induced from . This entails no loss of generality because the metric so induced is the same as the standard round metric of with constant sectional curvature . To state our results concisely, let us introduce the following definition:
Definition 2.
A smooth curve in will be called a Lissajous curve if there are numbers and vectors such that, for each ,
(2.1) |
We will call and the fundamental angular frequencies of the curve and the coefficient vectors of .
Therefore, a Lissajous curve, in our sense, can be thought of as the trajectory of a point in which oscillates with frequency in the -plane and with frequency in the -plane. Note also that the projection of on any 2-dimensional linear subspace different from the and -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 :
Theorem 1.
Let be given numbers where, if , then .
-
(1)
There exists a helix with constant curvature and torsion .
-
(2)
Such a helix is a Lissajous curve in the form of Equation (2.1).
-
(3)
The fundamental angular frequencies of are distinct and are given by
(2.2) (2.3) with
(2.4) -
(4)
If , then the frequencies and satisfy
(2.5) -
(5)
If then the four coefficient vectors are orthogonal in , and their magnitudes are given by
(2.6) and
(2.7) -
(6)
If , then is a circle given by
where . Further, the coefficient vectors are mutually orthogonal and we have
Several interesting features may be noted here:
-
(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)
When , the curve may be thought of as the trajectory of a motion consisting of two superimposed circular motions in perpendicular planes: one at a “slow” frequency and the other at a “fast” frequency . This global behavior arises from the fact that the curve must lie on the compact surface . Observe that there is no such restriction on the angular frequency of the Euclidean helix given by Equation (1.2).
-
(3)
When by definition we have . Such a curve is a geodesic, i.e., its unit tangent field is auto-parallel along the curve. Therefore, geodesics on the sphere are of the form
where and are mutually orthogonal. Thus, we recapture the well known fact that geodesics in are great circles.
We now turn to the question of uniqueness and periodicity of helices. First, note that if is a helix in a Riemannian 3-manifold, , and is a self-isometry of , then is also a helix in with the same curvature and torsion as that of . In , the converse holds, i.e., helices with the same curvature and torsion are congruent under an isometry of . We show that a similar fact holds in 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 , e.g. for , the surface in given by
is clearly contained in , and is therefore a Clifford torus in which is flat in the Riemannian metric induced by the round metric of .
Theorem 2.
Let be given numbers where, if , then .
-
(1)
If and are two helices in with the same curvature and torsion , then and are congruent, i.e., there is a Riemannian isometry such that .
-
(2)
A helix is periodic if and only if the ratio of the angular frequencies
(2.8) is a rational number, where .
-
(3)
If , there exists a Clifford torus contained in such that the image of lies on .
-
(4)
If and is irrational, the image of is dense in the torus .
2.1. Visualization of Helices
One way to visualize is to use the sterographic projection where is a point in which serves as the pole of the projection. It is well-known that is conformal, i.e. it preserves angles but not lengths. Using Wolfram Mathematica™, we produced visualizations of two helices in which are shown in Figures 1 and 2 below. Each of these pictures represents two distinct perspective projections onto of the sterographic projection of the helix, where 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 . The hue and brightness of the following curves are functions of the fourth coordinate of the curve in its embedding in .




3. The Frenet-Serret Equations
3.1. The Frenet-Serret Equations in a 3-Dimensional Riemannian Manifold
Consider a three dimensional Riemannian manifold and an arc length parameterized curve where is an open interval. Let represent the covariant derivative of a vector field along a curve (parametrized by ). Let denote the unit tangent vector field of . Since for each we have,
Then, the curvature function is defined as
(3.1) |
We will assume that either for all or that . In the case where never vanishes, we define the normal vector field to by
Then is a unit vector field along which is always orthogonal to . The definition of gives the first Frenet-Serret equation
(3.2) |
Similarly, since for each
and because for each
by Equation (3.2). Then,
We define a torsion function by
(3.3) |
We will assume that either for all , or . If for all , then we set
such that is a unit vector field along which is orthogonal to for all . If , then we choose to be an auto-parallel vector field along such that the vectors form an orthonormal basis of . In both cases we have
(3.4) |
Finally, since for each
and because for each
by Equation (3.4). Then,
since is orthogonal to for each by construction. By taking the dot product of both sides with we have
and by the product rule
Therefore we have our third and final Frenet-Serret equation
(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 . This concludes the derivation of the Frenet-Serret formulas in the case where for all .
However, in the case where , we define and choose to be auto-parallel vector fields along such that the vectors form an orthonormal basis of . Under this choice, the Frenet-Serret Formulas given in (1.1) are again satisfied.
Note that we are not assuming that the manifold is orientable. In the case where is in fact oriented (i.e., 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 is positively oriented. Then, one must allow the torsion 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 and are always non-negative. Geometrically, this means that while considering helices in , we disregard the chirality.
3.2. The Frenet equations in
We begin by specializing the Frenet-Serret equations given in (1.1) to the case of the embedded sphere in . Let
be the natural embedding. Given a curve , where is an open interval, we may think of as a curve in , by identifying with . Similarly, given a vector field along the curve which assigns to each point a vector , we can identify with the vector field along , which assigns to the point the vector . In order to simplify notation, we adopt the following conventions:
-
(1)
Consistently identifying with the embedded image, we will omit the map and its pushforward from the notation. Thus, we will think of the in as a curve in whose image lies in . Similarly, we will think of a vector field in along as a vector field in along such that for each , the vector lies in the subspace .
-
(2)
We identify the tangent bundle with . Therefore, all vector fields in may be identified with -valued functions.
-
(3)
Given a vector field along a curve in , by the previous two parts, we can identify it with a -valued function. We will let denote its derivative in the Euclidean space , i.e., if is represented using the natural coordinates as
where is smooth, , then
Of course, this is nothing but a coordinate expression for the covariant derivative of the vector field along with respect to the flat Euclidean metric of .
Proposition 3.1.
Let be a smooth curve in the three-sphere parametrized by arc length, and let be its Frenet frame. Using the notational convention explained above, we think of as functions from to . Then, these vector valued functions satisfy the following differential equations:
(3.6) | ||||
Proof.
We begin by recalling the following fact from differential geometry [2, Vol. III, p. 2]. Let be an embedded submanifold of , and for each point , let
denote the orthogonal projection (where we identify in the natural way with a subspace of ). We endow with the Riemannian metric induced by the Euclidean metric of . Let be a smooth curve in , where is an open interval, and assume that is parametrized by arc length. If is a vector field along , it is well-known that the covariant derivative of is given by
When is a hypersurface in and , we may write for ,
where denotes a unit vector normal to the hypersurface at the point . Consequently we obtain the following formula for differentiating a vector field along the curve :
When in , we may take for
so that
(3.7) |
We now compute when is one of the Frenet frame vector fields . Note that the four vectors , , , and form an orthonormal set in . Observe that for each we have the following equations:
In the first equation, we have used the fact that . Using Equation (3.7) and the above compuations, we obtain the following represenations of the covariant derivatives of the Frenet frame:
4. Lissajous curves in
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 be distinct non-negative real numbers, and suppose that for each , we have
(4.1) |
where the coefficients are complex numbers. Then we have for each .
Proof.
We can assume without loss of generality that (simply take ). For , let us set . Then, Equation (4.1) takes the form
(4.2) |
where
For each , it follows that if and only if .
Fix an integer , , and multiply both sides of Equation (4.2) by . Integrating on the interval and dividing by , we see that for each we have
(4.3) |
Note, however, that if , we have
Since for each , this is bounded independently of , as , each term in the first sum of Equation (4.3) goes to 0, which shows that . Therefore, . Since 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 .
-
(a)
If and has constant speed, then the coefficient vectors of satisfy the following relations:
-
(1)
are orthogonal
-
(2)
and
-
(3)
.
-
(1)
-
(b)
If , then:
-
(1)
are orthogonal
-
(2)
-
(3)
.
-
(1)
Proof.
For use in the later portions of this proof, we will first compute . Using Equation (2.1) and the fact that lies in , for each we have
(4.4) |
First, we prove part (a) of the proposition. Let u begin by assuming that . Since and , we see that the five numbers
are all distinct. By Lemma 4.1, each of the coefficients in the expression for vanishes, as in the left hand side of Equation (4.1). Thus, from the coefficients of Equation (4.4), we obtain:
which yields the following:
(4.5) | |||
(4.6) | |||
(4.7) | |||
(4.8) |
Equation (4.8) shows that the vectors 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
(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 . Let us set and thus, . Therefore, we have
Notice, , so in Equation (4.4) there are only 4. Therefore, the relation applied to Equation (4.4) and Lemma 4.1 now give,
(4.10) | ||||
(4.11) | ||||
(4.12) | ||||
(4.13) | ||||
(4.14) | ||||
(4.15) | ||||
(4.16) |
Since has constant speed, there exists a such that for all , we have . The relation yields (using Equation (4.4))
(4.17) |
Using Lemma 4.1, this gives the relations
(4.18) | ||||
(4.19) | ||||
(4.20) | ||||
(4.21) | ||||
(4.22) | ||||
(4.23) | ||||
(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
Similarly, Equations (4.13) and (4.21) imply that
Combining these last two relations with Equations (4.10 - 4.16), we get that the coefficient vectors are mutually orthogonal, , and . Conclusions (1), (2) and (3) follow again.
Now we will prove part (b) of the proposition. We set and in Equation (4.4). Then in Equation (4.4) we have the following three distinct frequencies:
Therefore, by Lemma 4.1 we have
(4.25) | |||
(4.26) | |||
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 , one can construct Lissajous curves in (of non-constant speed) for which the coefficient vectors are not orthogonal.
5. Proof of Theorem 1
5.1. Part 1
Adjoining the relation to the Frenet-Serret equations (Equation (3.6)) we obtain the system of equations
(5.1) | ||||||||
where and are the given constants. We now rewrite these equations in matrix form. Note that the four vectors form an orthonormal basis of . Let denote the matrix whose rows are these four vectors. Then for each , the matrix is orthogonal. When the curvature and the torsion 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
(5.2) |
where denotes the skew-symmetric matrix
(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 and is given by
(5.4) |
This proves part 1 of the theorem.
5.2. Part 2
In order to calculate the matrix exponential , we first recognize that since is skew-symetric, it can be diagonalized and thus written as
where is a diagonal matrix with real entries of the form
(5.5) |
where . This is because the eigenvalues of the real skew-symmetric matrix are purely imaginary and occur in complex conjugate pairs. Therefore,
(5.6) |
From Equation (5.5) it follows that
Since is the first row of the matrix it follows that
where the coefficient vectors are constant vectors in . This proves part 2 of the theorem.
5.3. Part 3
The diagonal entries of , where is as in Equation (5.5), are the eigenvalues of the matrix of Equation (5.3). We find them by solving the characteristic equation
(5.7) |
with as in Equation (2.4). The solutions of the characteristic equation are
where are as in Equations (2.2) and (2.3). This proves part 3 of the theorem.
5.4. Part 4
Since , we have
(5.8) |
Therefore, from Equations (2.2) and (2.3) we see that . by definition of in (2.4) we have,
(5.9) |
Combining Equations (5.8) and (5.9) we have
(5.10) | ||||
(5.11) |
Therefore, , and we have
(5.12) |
Then, by making use of Equation (5.11),
(5.13) |
Thus,
(5.14) |
This proves part 4 of the theorem.
5.5. Part 5
Note that for all and for all as well, since lies in and has unit speed. Since by Equation (2.2), we know that . Therefore, by part (a) of Proposition 4.2,
-
(1)
The coefficient vectors are mutually orthogonal, which is one of the conclusions of part 5 of Theorem 1,
- (2)
-
(3)
We have that
(5.15)
Now, let . Then, (since is parameterized by arc length) and differentiating Equation (2.1), we see that may be represented as
where , , , and . This shows that is a Lissajous curve in .
Now, we claim that is constant independently of . Recall, that
where is the tangent vector field in the Frenet Frame . Therefore, by the first equation in (3.6), we have
which yields,
(5.16) |
where the term since . We may now apply conclusion 3 of part (a) of Proposition 4.2 to obtain that , which is equivalent to the statement that
(5.17) |
Combining Equations (5.15) and (5.17), we get
and
Solving these equations for and , we obtain Equations (2.6) and (2.7).
5.6. Part 6
If , by Equation (2.2), we have . We set by Equation (2.3) and then by parts 1,2, and 3 of this theorem, proved above, is given by
Furthermore, by part (b) of Proposition 4.2 we know that , that the coefficient vectors are mutually orthogonal. Note that,
Since has unit speed, we have
which implies,
(5.18) |
Combining Equation (5.18) with conclusion 3 of part (b) of Proposition 4.2, we get that
(5.19) |
6. Proof of Theorem 2
6.1. Part 1
Suppose that we have two helices and in with the same curvature and torsion . Then, by part 3 of Theorem 1, the fundamental angular frequencies of these two curves are the same. Thus, the curves are represented as
If , by part 5 of Theorem 1, we also know that
and that the sets of vectors are mutually orthogonal and are also mutually orthogonal. Therefore, there exists an orthogonal map, in such that , , , and . Then is an isometry of and it is clear that .
If , then by part 6 of Theorem 1, we know that and take the form
where and by Equations (2.2) and (2.3). Furthermore, by part 6 of Theorem 1, we know that
and that the sets of vectors are mutually orthogonal and are also mutually orthogonal. Therefore, there again exists an orthogonal map, in such that , , . Then is again an isometry of and it is clear that .
6.2. Part 2
Let be a helix in which can be written, thanks to Theorem 1, in the form of Equation (2.1):
Now suppose that is periodic with period . Then, for each , we have
First, let us assume that and consequently, because of Equation (2.2), . Then, comparing the coefficients of , we obtain,
(6.1) |
for each . This shows that there exists a non-zero such that,
Similarly, we compare the coefficients of , to get
(6.2) |
for each . This shows that there exists a non-zero such that,
It follows that
Now we prove the converse. Suppose that . Then, let
Then, Equations (6.1) and (6.2) hold. Therefore,
Which proves part 2 of Theorem 2 in the case where . If, on the other hand, and consequently , then is given by,
Which is always periodic with a period of
This proves part 2 of the theorem.
6.3. Part 3
Let and let be the orthonormal basis of consisting of the unit vectors along the coefficient vectors of as given in Equation (2.1). We denote the coordinates of a point by
By parts 2 and 5 of Theorem 1, we have that is represented in these coordinates by . Consider the torus in given by
(6.3) |
Clearly lies on . It is clear that is a flat Clifford torus in , and is contained in , since if , then
where the last equality follows by Equations (2.6) and (2.7).
6.4. Part 4
Note that the helix is a solution of the differential equations on the torus
where (resp. ) is an angular coordinate on the circle (resp. ). If , then a classical result in the theory of dynamical systems [1, Proposition 4.2.8, p. 113] shows that the image of is dense in . 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.