Russian Mathematical Surveys
RUS  ENG    JOURNALS   PEOPLE   ORGANISATIONS   CONFERENCES   SEMINARS   VIDEO LIBRARY   PACKAGE AMSBIB  
General information
Latest issue
Archive
Impact factor
Submit a manuscript

Search papers
Search references

RSS
Latest issue
Current issues
Archive issues
What is RSS



Uspekhi Mat. Nauk:
Year:
Volume:
Issue:
Page:
Find






Personal entry:
Login:
Password:
Save password
Enter
Forgotten password?
Register


Russian Mathematical Surveys, 2023, Volume 78, Issue 1, Pages 65–163
DOI: https://doi.org/10.4213/rm10063e
(Mi rm10063)
 

This article is cited in 3 scientific papers (total in 3 papers)

Left-invariant optimal control problems on Lie groups that are integrable by elliptic functions

Yu. L. Sachkov

Ailamazyan Program Systems Institute of Russian Academy of Sciences
References:
Abstract: Left-invariant optimal control problems on Lie groups are an important class of problems with a large symmetry group. They are theoretically interesting because they can often be investigated in full and general laws can be studied by using these model problems. In particular, problems on nilpotent Lie groups provide a fundamental nilpotent approximation to general problems. Also, left-invariant problems often arise in applications such as classical and quantum mechanics, geometry, robotics, visual perception models, and image processing.
The aim of this paper is to present a survey of the main concepts, methods, and results pertaining to left-invariant optimal control problems on Lie groups that can be integrated by elliptic functions. The focus is on describing extremal trajectories and their optimality, the cut time and cut locus, and optimal synthesis.
Bibliography: 162 titles.
Keywords: optimal control, geometric control theory, left-invariant problems, sub-Riemannian geometry, Lie groups, optimal synthesis, elliptic functions.
Funding agency Grant number
Russian Science Foundation 22-11-00140
The work was supported by the Russian Science foundation under grant no. 22-11-00140, https://rscf.ru/project/22-11-00140/.
Received: 14.06.2022
Russian version:
Uspekhi Matematicheskikh Nauk, 2023, Volume 78, Issue 1(469), Pages 67–166
DOI: https://doi.org/10.4213/rm10063
Bibliographic databases:
Document Type: Article
UDC: 517.977
MSC: Primary 53C17; Secondary 22E25, 49K15
Language: English
Original paper language: Russian

1. Preface

Studying left-invariant control systems on Lie groups and homogeneous spaces is one of the central topics in geometric control theory. This is a natural and important class of systems from the theoretical standpoint, for which an interesting global theory can be developed (such systems arise, for example, when we consider a local nilpotent approximation to smooth systems). On the other hand, such systems can model a number of applied problems (rotation and rolling of rigid bodies, robot motion, quantum mechanics, computer vision).

It is well known that it is very difficult to find an explicit solution of a global non-linear control problem (for example, a problem of controllability or an optimal control problem) when it does not have a large symmetry group. For invariant problems on Lie groups (and their projections onto homogeneous spaces) we can often find an explicit solution by means of the methods of geometric control theory and using the techniques of differential geometry and the theory of Lie groups and algebras. The resulting solution of the invariant problem can be a good approximation to the cor- responding non-linear problem. For instance, the invariant sub-Riemannian geometry of the Heisenberg group is the cornerstone of all sub-Riemannian geometry.

The two main problems considered for left-invariant systems are the controllability problem and the optimal control problem. There are many publications concerning the controllability problem, for instance, [128].

In this survey we consider only problems integrable by elliptic functions. Problems integrable by elementary functions were considered in [143].

2. Problems integrable by elliptic functions and integrals

2.1. Elliptic integrals and functions

Standard references on elliptic integrals and functions are [8], [98], and [162]. Below we provide some minimal information about these that is necessary for the next sections.

Elliptic integrals in the Jacobi form: Legendre elliptic integrals of the first kind

$$ \begin{equation*} F(\varphi,k)=\int_0^{\varphi}\frac{dt}{\sqrt{1-k^2\sin^2 t}}\,, \end{equation*} \notag $$
integrals of the second kind
$$ \begin{equation*} E(\varphi, k)=\int_0^{\varphi} \sqrt{1-k^2 \sin^2 t} \, dt, \end{equation*} \notag $$
integrals of the third kind
$$ \begin{equation*} \Pi(m; \varphi, k)=\int_0^{\varphi} \frac{dt}{(1+m \sin^2 t)\sqrt{1-k^2 \sin^2 t}}\,; \end{equation*} \notag $$
here and in what follows the elliptic modulus $k$ lies in the interval $(0,1)$. The complementary modulus is $k'=\sqrt{1-k^2}$ .

The complete elliptic integrals are

$$ \begin{equation*} K(k)=F\biggl(\frac{\pi}{2}\,, k\biggr)\quad\text{and} \quad E(k)=E\biggl(\frac{\pi}{2}\,, k\biggr). \end{equation*} \notag $$

Jacobi elliptic functions:

$$ \begin{equation*} \begin{gathered} \, \varphi=\operatorname{am}(u,k) \ \ \Longleftrightarrow \ \ u=F(\varphi, k), \\ \operatorname{sn}(u,k)=\sin \operatorname{am}(u,k), \qquad \operatorname{cn}(u,k)=\cos \operatorname{am}(u,k), \\ \operatorname{dn}(u,k)=\sqrt{1-k^2 \operatorname{sn}^2(u,k)}\,, \qquad \operatorname{E}(u,k)=E(\operatorname{am}u,k). \end{gathered} \end{equation*} \notag $$
In the notation for elliptic functions the modulus $K$ is often dropped.

Standard formulae. Derivatives and integrals:

$$ \begin{equation*} \operatorname{am}' u=\operatorname{dn} u, \quad \operatorname{sn}'u=\operatorname{cn} u \operatorname{dn} u, \quad \operatorname{cn}' u=- \operatorname{sn} u \operatorname{dn} u, \quad \operatorname{dn}'u=- k^2 \operatorname{sn} u \operatorname{cn} u \end{equation*} \notag $$
and
$$ \begin{equation*} \int_0^u \operatorname{dn}^2 t \, dt=\operatorname{E}(u). \end{equation*} \notag $$

Degenerate cases:

$$ \begin{equation*} \begin{alignedat}{6} k &\to +0 &\ \ &\Longrightarrow &\ \ \operatorname{sn} u &\to \sin u, &\ \ \operatorname{cn} u &\to \cos u,&\ \ \operatorname{dn} u &\to 1, &\ \ \operatorname{E}(u) &\to u; \\ k &\to 1-0 &\ \ &\Longrightarrow &\ \ \operatorname{sn} u &\to \tanh u, &\ \ \operatorname{cn} u&\to \frac{1}{\operatorname{ch} u}\,, &\ \ \operatorname{dn} u &\to \frac{1}{\cosh u}\,, &\ \ \operatorname{E}(u) &\to \tanh u. \end{alignedat} \end{equation*} \notag $$

2.2. A mathematical pendulum

In all sub-Riemannian problems presented in §§ 2.32.10 the vertical subsystem of the Hamiltonian system of the Pontryagin maximum principle reduces mysteriously to the pendulum equation, so that all of them are integrated by elliptic functions and integrals.

2.2.1. The pendulum equation and its solution

Consider a mathematical pendulum, which is a point mass fixed on a weightless inextensible rod of length $L$, which rotates freely in a vertical plane about the point of suspension. Let $\theta$ denote the angle between the pendulum and its downward vertical position. Then the motion of the pendulum satisfies the equations

$$ \begin{equation} \dot \theta=c, \quad \dot c=-r \sin \theta, \end{equation} \tag{2.1} $$
where $r=g/L > 0$ and $g$ is the gravitational acceleration. The total energy of the pendulum (which is a first integral of equations (2.1)) is
$$ \begin{equation*} E=\frac{c^2}{2}-r\cos \theta \in [-r, +\infty). \end{equation*} \notag $$
The motion pattern of the pendulum is determined by the value of $E$:

We have described above the motion pattern of the pendulum (2.1) for $r=g/L >0$. On the other hand, if $r=0$ (which can be interpreted as the absence of gravitational forces), then

The case $r < 0$ (when the gravitational force is upward) reduces to $r > 0$ after the substitution $(\theta, c, r) \mapsto (\theta+\pi, c, -r)$.

2.2.2. Straightening coordinates

For $r > 0$ the phase cylinder of the pendulum system (2.1),

$$ \begin{equation*} C=\{(\theta, c) \mid \theta\in S^1, \ c \in \mathbb{R}\}, \qquad S^1=\mathbb{R}/{2\pi}\mathbb{Z}, \end{equation*} \notag $$
is stratified depending on the motion pattern:
$$ \begin{equation*} C=\bigsqcup_{i=1}^5 C_i, \end{equation*} \notag $$
where
$$ \begin{equation*} \begin{aligned} \, C_1&=\{(\theta, c) \in C \mid E \in (-r, r)\}, \\ C_2&=\{(\theta, c) \in C \mid E >r\}, \\ C_3&=\{(\theta, c) \in C \mid E=r, \ c \ne 0\}, \\ C_4&=\{(\theta, c) \in C \mid c=0, \ \theta=0\}, \\ C_5&=\{(\theta, c) \in C \mid c=0, \ \theta=\pi\}. \end{aligned} \end{equation*} \notag $$
In the domains $C_1$, $C_2$, and $C_3$ we can introduce coordinates $(\varphi,k)$ straightening the equation of the pendulum.

If $(\theta,c) \in C_1$, then set

$$ \begin{equation*} \begin{gathered} \, k=\sqrt{\frac{E+r}{2r}} \in (0, 1), \qquad \sqrt r\,\varphi \in [0,4K(k)] \ \operatorname{mod}{4K(k)}, \\ \sin \frac{\theta}{2}=k \operatorname{sn}(\sqrt r\,\varphi, k), \qquad \cos \frac{\theta}{2}=\operatorname{dn}(\sqrt r\,\varphi, k), \\ c=2 k \sqrt r\,\operatorname{cn}(\sqrt r\,\varphi, k). \end{gathered} \end{equation*} \notag $$
If $(\theta, c) \in C_2$, then set
$$ \begin{equation*} \begin{gathered} \, k=\sqrt{\frac{2r}{E+r}} \in (0, 1), \qquad \sqrt r\,\varphi\in [0,2k K(k)] \ \operatorname{mod}{2k K(k)}, \\ \sin \frac{\theta}{2}=\pm\operatorname{sn} \biggl(\frac{\sqrt r\,\varphi}{k}\,, k\biggr), \qquad \cos \frac{\theta}{2}=\operatorname{cn} \biggl(\frac{\sqrt r\,\varphi}{k}\,, k\biggr), \\ c=\pm 2 \frac{\sqrt r}{k}\operatorname{dn} \biggl(\frac{\sqrt r \,\varphi}{k}\,,k\biggr), \qquad \pm=\operatorname{sign} c. \end{gathered} \end{equation*} \notag $$
If $(\theta, c) \in C_3$, then set
$$ \begin{equation*} \begin{gathered} \, k=1, \quad \varphi \in \mathbb{R}, \\ \sin \frac{\theta}{2}=\pm \tanh(\sqrt r\,\varphi), \qquad \cos \frac{\theta}{2}=\frac{1}{\cosh(\sqrt r\,\varphi)}\,, \\ c=\pm \frac{2 \sqrt r}{\cosh(\sqrt r\,\varphi)}\,, \qquad \pm=\operatorname{sign} c. \end{gathered} \end{equation*} \notag $$

In the coordinates $(\varphi,k)$ the pendulum equation (2.1) ‘straightens out’:

$$ \begin{equation*} \dot\varphi=1, \quad \dot k=0, \end{equation*} \notag $$
so its solution is
$$ \begin{equation*} \varphi_t=\varphi+t, \quad k \equiv \operatorname{const}. \end{equation*} \notag $$
We use these straightening coordinates and modifications of these to parametrize the extremal trajectories in §§ 2.42.10.

2.2.3. Bibliographic comments

Subsection 2.2.1 is based on [8], while § 2.2.2 is based on [123] (also see [98], [142], and [162]).

2.3. Martinet flat sub-Riemannian problem

2.3.1. The statement of the problem

The Martinet flat sub-Riemannian structure is defined by the metric $ds^2=dx^2+dy^2$ on the Martinet distribution $\Delta=\{dz-(1/2)y^2\,dx= 0\}$ in $M=\mathbb{R}^3_{x, y, z}$. An orthonormal frame can be taken in the form

$$ \begin{equation*} X_1=\frac{\partial}{\partial x}+\frac{y^2}{2}\, \frac{\partial}{\partial z}\,, \qquad X_2=\frac{\partial}{\partial y}\,. \end{equation*} \notag $$
Let $X_3=\partial/\partial z$; then the Lie algebra generated by the fields $X_1$ and $X_2$ has the multiplication table
$$ \begin{equation*} [X_1,X_2]=-yX_3, \quad [X_2,[X_1,X_2]]=-X_3,\quad [X_1,[X_1,X_2]]=0, \quad \operatorname{ad} X_3=0, \end{equation*} \notag $$
so that this is the Engel algebra (see § 2.9).

The Martinet flat sub-Riemannian structure is not left invariant, but we include it in our survey because it plays a special role in sub-Riemannian geometry:

Moreover, the Martinet flat sub-Riemannian structure is a quotient of the left- invariant sub-Riemannian structure on the Engel group (see § 2.9), so the Hamiltonian system for Martinet extremals reduces to the pendulum equation, and the extremals themselves project onto the $(x,y)$-plane as Euler elasticae (see § 2.6).

The optimal control problem for the Martinet flat sub-Riemannian structure has the following form:

$$ \begin{equation*} \begin{gathered} \, \dot q=u_1X_1+u_2X_2, \quad q=(x, y, z) \in\mathbb{R}^3, \quad u=(u_1, u_2)\in\mathbb{R}^2, \\ q(0)=q_0, \quad q(t_1)=q_1, \\ J=\frac{1}{2}\int_0^{t_1}(u_1^2+u_2^2)\,dt \to \min. \end{gathered} \end{equation*} \notag $$

2.3.2. The Pontryagin maximum principle

Proposition 2.1. The abnormal trajectories are $\{y=0,\ z=z_0\}$. They are non-strictly abnormal.

Normal extremals are trajectories of the Hamiltonian field with Hamiltonian

$$ \begin{equation*} H=\frac{1}{2}(h_1^2+h_2^2)= \frac{1}{2}\biggl[\biggl(p_x+\frac{y^2}{2}p_z\biggr)^2+p_y^2\biggr], \end{equation*} \notag $$
where $(p_x,p_y,p_z)$ are the canonical coordinates of a covector $\lambda \in T^*M$ and $h_i(\lambda)=\langle \lambda,X_i(q) \rangle$, $i=1,2,3$. The corresponding Hamiltonian system $\dot \lambda=\vec H(\lambda)$ has the form
$$ \begin{equation*} \begin{alignedat}{2} \dot x&=p_x+\frac{y^2}{2}p_z, &\qquad \dot p_x&=0, \\ \dot y&=p_y, &\qquad \dot p_y&=-\biggl(p_x+\frac{y^2}{2}p_z\biggr)p_z y, \\ \dot z&=\biggl(p_x+\frac{y^2}{2}p_z\biggr)\frac{y^2}{2}\,,&\qquad \dot p_z&=0, \end{alignedat} \end{equation*} \notag $$
that is,
$$ \begin{equation} \begin{alignedat}{2} &\dot x=h_1, &\qquad \dot h_1&=yh_2h_3, \\ &\dot y=h_2, &\qquad \dot h_2&=-yh_1h_3, \\ &\dot z=\frac{y^2}{2}h_1, &\qquad \dot h_3&=0. \end{alignedat} \end{equation} \tag{2.2} $$

We consider extremals on the level surface $\{H=1/2\}$, on which we introduces the coordinates

$$ \begin{equation*} h_1=\cos \theta, \quad h_2=\sin \theta, \quad\text{and}\quad h_3=c. \end{equation*} \notag $$

2.3.3. Symmetries

Reflections. The sub-Riemannian structure $(\Delta,ds^2)$ is invariant under the group of reflections

$$ \begin{equation*} \begin{gathered} \, \operatorname{Sym}=\{ \operatorname{Id}, \varepsilon^1, \varepsilon^2, \varepsilon^3 \} \cong \mathbb{Z}_2 \times \mathbb{Z}_2, \\ \begin{alignedat}{2} \varepsilon^1\colon (x, y, z)&\mapsto (x, -y, z), &\qquad (\theta, c)&\mapsto (\pi-\theta, c), \\ \varepsilon^2\colon (x, y, z)&\mapsto (- x, y, -z), &\qquad (\theta, c)&\mapsto (-\theta,-c), \\ \varepsilon^3\colon (x, y, z)&\mapsto (- x, -y, -z), &\qquad (\theta, c)&\mapsto (\theta-\pi,-c). \end{alignedat} \end{gathered} \end{equation*} \notag $$

Dilations. The Hamiltonian system (2.2) is invariant under the one-parameter group of dilations

$$ \begin{equation*} \begin{aligned} \, (x,y,z) &\mapsto (\delta^{-1}x,\delta^{-1}y,\delta^{-3}z), \\ (h_1,h_2,h_3)&\mapsto (\delta^{-1}h_1,\delta^{-1}h_2,\delta h_3). \end{aligned} \end{equation*} \notag $$

2.3.4. A parametrization of geodesics

We assume below that $q_0=0$.

Proposition 2.2. Geodesics with natural parametrization issuing from the point $q_0=0$ are the curves

$$ \begin{equation*} \begin{aligned} \, x_t&=-t+\frac{2}{\sqrt c}(\operatorname{E}(u)-E(k)), \\ y_t&=- \frac{2 k}{\sqrt c} \operatorname{cn} u, \\ z_t&=\frac{2}{3c^{3/2}}[(2k^2-1)(\operatorname{E}(u) -E(k))+ k'^2t\sqrt c+2k^2\operatorname{sn}u\operatorname{cn}u\operatorname{dn}u], \end{aligned} \end{equation*} \notag $$
where $u=K+t\sqrt c$ , $k=\sin(\pi/4-\theta/2)$, $\theta \in (-\pi/2,\pi/2)$, and $c > 0$, and the curves
$$ \begin{equation*} x_t=t\sin\theta, \quad y_t=t\cos \theta, \quad z_t=\frac{t^3}{6}\sin\theta\cos^2\theta, \end{equation*} \notag $$
where $\theta \in (-\pi/2,\pi/2]$, and also the curves obtained from the above ones by means of the symmetries $\varepsilon^1$ and $\varepsilon^2$.

Let $\operatorname{Exp}$ denote the exponential map

$$ \begin{equation*} \operatorname{Exp}\colon C \times \mathbb{R}_+ \to M, \quad (\lambda, t)\mapsto q_t=\pi \circ e^{t\vec H}(\lambda), \end{equation*} \notag $$
where
$$ \begin{equation*} C=T_{q_0}^* M\cap\biggl\{H=\frac{1}{2} \biggr\}. \end{equation*} \notag $$

2.3.5. Conjugate time

If a geodesic is strictly normal and its projection onto the $(x, y)$-plane is a straight line, then this geodesic is optimal and therefore free from conjugate points. In the abnormal case the geodesic is optimal but consists of conjugate points.

Let $\lambda=(\theta, c) \in C$, and assume that a geodesic $q_t=\operatorname{Exp}(\lambda, t)$ does not project onto the $(x, y)$-plane as a straight line. Because of the symmetries $\varepsilon^1$ and $\varepsilon^2$, we can assume that $c > 0$ and $\theta \in (-\pi/2,\pi/2)$. Then the first conjugate time is

$$ \begin{equation*} t_{\rm conj}^1(\lambda)=\min\{t > 0 \mid v^2 c_1(v)+vc_2(v)+c_3(v)=0\}, \end{equation*} \notag $$
where
$$ \begin{equation*} \begin{aligned} \, c_1(v)&=k'^2\,\frac{\operatorname{cn} v}{\operatorname{dn} v}\,, \\ c_2(v)&=k'^2 \operatorname{sn} v- 2k'^2 \operatorname{E}(v)\frac{\operatorname{cn} v}{\operatorname{dn} v}\,, \\ c_3(v)&=\operatorname{E}^2(v)\frac{\operatorname{cn}v}{\operatorname{dn}v}- \operatorname{E}(v)\operatorname{sn} v \end{aligned} \end{equation*} \notag $$
and $v=t\, \sqrt c$ .

Theorem 2.1. Let $q_t=\operatorname{Exp}(\lambda,t)$, $\lambda \in C$, $t > 0$, be a geodesic whose projection onto the $(x,y)$-plane is not a straight line. Then

$$ \begin{equation*} t_{\rm conj}^1(\lambda )\in \biggl(\frac{2 K}{\sqrt{|c|}}\,, \frac{3 K}{\sqrt{|c|}} \biggr). \end{equation*} \notag $$

As approximate calculations show, the ratio $t_{\rm conj}^1\, \sqrt{|c|}/(3K)$ is roughly a constant, which is equal to $0.97$.

2.3.6. The cut time and cut locus

Theorem 2.2. The geodesics whose projections onto the $(x,y)$-plane are straight lines are length minimizers. Given a geodesic $q_t=\operatorname{Exp}(\lambda,t)$, $\lambda \in C$, $t > 0$, whose projection onto the same plane is not a straight line, the cut time on it is $t_{\rm cut}(\lambda)=2K/\sqrt{|c|}$ and corresponds to its first intersection with the Martinet plane $\{y=0\}$.

The cul locus is

$$ \begin{equation*} \operatorname{Cut}=\{q \in M \mid y=0,\ z \ne 0\}. \end{equation*} \notag $$
This set is disjoint from the first caustic.

2.3.7. A sphere and a wavefront

Dilations take different spheres with centre $q_0=0$ to one another, so we can limit ourselves to considering the unit sphere

$$ \begin{equation*} S=\{q \in M \mid d(q_0,q)=1\}. \end{equation*} \notag $$
It is shown in Fig. 1 in the coordinates $(x,y,v)$, $v=z-xy^2/6$.

Theorem 2.3. The intersection of the sphere $S$ with the cut locus (see Fig. 2) is the curve $k \mapsto \gamma(k)$ in the Martinet plane $\{y=0\}$ given by the parametric equations

$$ \begin{equation} x(k) =-1+2\frac{E(k)}{K(k)}\,, \end{equation} \tag{2.3} $$
$$ \begin{equation} z(k) =\frac{1}{6K^3(k)}[(2 k^2-1)E(k)+k'^2K(k)], \end{equation} \tag{2.4} $$
where $k\in(0,1)$, plus the curve obtained from $\gamma$ by applying the symmetry $\varepsilon^2|_{\{y=0\}}$: $(x,z)\mapsto(-x,-z)$.

As $k \to +0$, the curve $\gamma$ becomes the restriction of the graph of an analytic function

$$ \begin{equation*} z=-\frac{2}{3\pi^2}(x-1)+o(x-1),\qquad x \to 1-0, \end{equation*} \notag $$
to the half-plane $\{z>0\}$.

As $k \to 1-0$, $\gamma$ becomes the graph of a non-analytic smooth function

$$ \begin{equation*} z=\frac{X^3}{6}+F(X), \qquad X=\frac{x+1}{2}\,, \end{equation*} \notag $$
where $F$ is the flat function
$$ \begin{equation*} F(X)=- 4 X^3 e^{-2/X}+o(X^3e^{-2/X}), \qquad X\to +0. \end{equation*} \notag $$

Theorem 2.4. The intersection of the sphere $S$ with the Martinet plane is not subanalytic, so $S$ is not subanalytic either.

Consider the unit-time wavefront from $q_0$:

$$ \begin{equation*} W=\{q \in M \mid q=\operatorname{Exp}(\lambda,1), \ \lambda \in C\}; \end{equation*} \notag $$
other wavefronts from $q_0$ can be obtained from it by dilations.

Theorem 2.5. The intersection of the wavefront $W$ with the Martinet plane $\{y= 0\}$ and the half-space $\{z > 0\}$ is a union of curves $\gamma_n$, $n \in \mathbb{N}$, whose closure has two branch points $x=\pm 1$, $z=0$. Each $\gamma_n$ is defined by the parametric equations

$$ \begin{equation*} \begin{aligned} \, x_n(k)&=-1+2 \frac{E(k)}{K(k)}\,, \\ z_n(k)&=\frac{1}{6n^2K^3(k)}[(2k^2 -1)E(k)+k'^2K(k)]. \end{aligned} \end{equation*} \notag $$
In a neighbourhood of $x=-1$, $z= 0$ this curve is the graph of a function
$$ \begin{equation*} z=\frac{1}{6n^2}X^3+F(X), \end{equation*} \notag $$
where $F(X)=\alpha X^3 e^{-2/X}+o(X^3e^{-2/X})$, $\alpha \ne 0$, and in a neighbourhood of $x=1$, $z=0$ it is the graph of a function
$$ \begin{equation*} z=-\frac{2}{3n^2\pi^2}(x-1)+o(x-1). \end{equation*} \notag $$
The outer curve $\gamma_1$ is the intersection $\gamma$ of the sphere with the Martinet plane $\{y=0\}$ and the half-space $\{z > 0\}$ (see Theorem 2.3).

The intersection of $S$ with the Martinet plane and the half-space $\{z > 0\}$ is the curve given parametrically by $k \mapsto (x(k), z(k))$, $k\! \in \! (0,1)$ (see (2.3) and (2.4)). It extends to $\{z \geqslant 0\}$ by continuity by letting $ k\in [0,1]$. The resulting curve is semianalytic for $k \ne 1$. However, it is not semianalytic at $k=1$, so it is not subanalytic.

Theorem 2.6. In a neighbourhood of the point $X=0$, where $X=(x+1)/2$, the intersection of the sphere $S$ with the Martinet plane $\{y=0\}$ and the half-plane $\{z \geqslant 0\}$ is the graph of a function of the form

$$ \begin{equation*} z=F\biggl(X,\frac{e^{-1/X}}{X^2}\biggr), \end{equation*} \notag $$
where $X \geqslant 0$ and $F$ is an analytic map from a neighbourhood of $(0,0) \in \mathbb{R}^2$ to $\mathbb{R}$.

Hence the intersection of $S$ with the Martinet plane is a curve in the $\exp$-$\log$- category [65], [101].

2.3.8. Bibliographic comments

This section is based on [3].

2.4. Sub-Riemannian problem on the group $\operatorname{SE}(2)$ of Euclidean planar motions

2.4.1. The problem statement

The mechanical setting. Consider the problem of optimal motion for a kinematic model of a mobile robot in the plane. The state of the robot is described by its position on the plane $(x,y) \in \mathbb{R}^2$ and the angle $\theta\in 2\pi \mathbb{Z}$ of its orientation relative to the positive direction of the $x$-axis. The robot can move at an arbitrary linear velocity $u_1 \in \mathbb{R}$, rotating at an arbitrary angular velocity $u_2\in\mathbb{R}$ in the process. The problem is to take the robot from an initial state $g_0=(x_0,y_0,\theta_0)$ to a terminal state $g_1=(x_1, y_1, \theta_1)$ along a shortest path in the state space. The length of a path in the state space $\mathbb{R}^2_{x, y}\times S^1_{\theta}$ is measured by the integral $\displaystyle\int_0^{t_1}(\dot x^2+\dot y^2+\alpha^2\dot\theta^2)^{1/2}\,dt$, where $\alpha>0$ is a prescribed positive number determining a balance between the linear and angular velocity.

The optimal control problem and its normalization. The above problem about a mobile robot can be formalized as an optimal control problem:

$$ \begin{equation*} \begin{gathered} \, \dot x=u_1 \cos\theta, \quad \dot y=u_1 \sin \theta, \quad \dot\theta=u_2, \\ g=(x, y, \theta) \in \mathbb{R}^2_{x,y}\times S^1_{\theta}, \quad u=(u_1, u_2)\in\mathbb{R}^2, \\ g(0)=g_0, \quad g(t_1)=g_1, \\ l=\int_0^{t_1}\sqrt{u_1^2+\alpha^2u_2^2}\,\, dt \to \min. \end{gathered} \end{equation*} \notag $$
After scaling in the $(x,y)$-plane,
$$ \begin{equation*} (x,y,\theta) \mapsto \biggl(\frac{x}{\alpha}\,, \frac{y}{\alpha}\,,\theta\biggr),\quad (u_1, u_2)\mapsto \biggl(\frac{u_1}{\alpha}\,,u_2\biggr), \end{equation*} \notag $$
we can reduce this problem to the case when $\alpha=1$.

Using parallel translations and rotating the $(x,y)$-plane we can achieve that $g_0=(0,0,0)$.

As a result, we obtain the optimal control problem

$$ \begin{equation} \dot x=u_1 \cos \theta, \quad \dot y=u_1 \sin \theta, \quad \dot \theta=u_2, \end{equation} \tag{2.5} $$
$$ \begin{equation} g=(x, y, \theta) \in \mathbb{R}^2_{x, y}\times S^1_{\theta}, \qquad u=(u_1, u_2)\in \mathbb{R}^2, \end{equation} \tag{2.6} $$
$$ \begin{equation} g(0)=g_0=(0, 0, 0), \quad g(t_1)= g_1=(x_1, y_1, \theta_1), \end{equation} \tag{2.7} $$
$$ \begin{equation} l=\int_0^{t_1}\sqrt{u_1^2+u_2^2}\, dt \to \min. \end{equation} \tag{2.8} $$
This is the sub-Riemannian problem specified by the orthonormal frame
$$ \begin{equation} X_1=\cos\theta\,\frac{\partial}{\partial x}+ \sin \theta\,\frac{\partial}{\partial y}\,, \quad X_2=\frac{\partial}{\partial \theta}\,. \end{equation} \tag{2.9} $$

The group of plane motions. The group $G=\operatorname{SE}(2)$ of proper Euclidean motions of the plane is a semidirect product of the group of parallel translations $\mathbb{R}^2$ and the group of rotations $\operatorname{SO}(2)$:

$$ \begin{equation*} \operatorname{SE}(2)=\mathbb{R}^2 \rtimes \operatorname{SO}(2). \end{equation*} \notag $$
It has the linear representation
$$ \begin{equation*} \operatorname{SE}(2)=\left\{\begin{pmatrix} \cos \theta &-\sin \theta & x \\ \sin \theta &\hphantom{-} \cos \theta & y \\ 0 & 0 & 1 \end{pmatrix}\ \bigg|\ \theta \in S^1= \mathbb{R}/2\pi\mathbb{Z}, \ x,y \in \mathbb{R}\right\}. \end{equation*} \notag $$
We can calculate the action of a motion $g=(x, y, \theta)$ on a vector $(a, b) \in \mathbb{R}^2$ using matrix multiplication:
$$ \begin{equation*} \begin{pmatrix} \cos \theta & -\sin \theta & x \\ \sin \theta & \hphantom{-}\cos \theta & y \\ 0 & 0 & 1 \end{pmatrix}\cdot \begin{pmatrix} a \\ b \\ 1 \end{pmatrix}=\begin{pmatrix} a \cos \theta-b \sin\theta+x \\ a\sin\theta+b\cos \theta+y \\ 1 \end{pmatrix}, \end{equation*} \notag $$
so that
$$ \begin{equation*} g\colon (a,b) \mapsto (a\cos\theta-b\sin\theta+x, \ a\sin\theta+b \cos\theta+y). \end{equation*} \notag $$

The Lie algebra of the group $\operatorname{SE}(2)$ is

$$ \begin{equation*} \mathfrak{g}=\mathfrak{se}(2)=\operatorname{span}(E_{21}-E_{12},E_{13},E_{23}), \end{equation*} \notag $$
where $E_{ij}$ is the $ 3\times 3 $ matrix containing a unique non-trivial entry, namely, $1$ in row $i$ and column $j$. Basis left-invariant vector fields on $\operatorname{SE}(2)$ are
$$ \begin{equation*} \begin{gathered} \, X_1=g E_{13}=\cos\theta\,\frac{\partial}{\partial x}+ \sin \theta\,\frac{\partial}{\partial y}\,, \\ X_2=g(E_{21}-E_{12})=\frac{\partial}{\partial \theta}\,, \end{gathered} \end{equation*} \notag $$
and
$$ \begin{equation*} X_3=- g E_{23}=\sin\theta\,\frac{\partial}{\partial x}- \cos \theta\,\frac{\partial}{\partial y} \end{equation*} \notag $$
and their multiplication table is
$$ \begin{equation} [X_1,X_2]=X_3, \quad [X_2,X_3]=X_1, \quad [X_1,X_3]=0. \end{equation} \tag{2.10} $$

The orthonormal frame (2.9) for the sub-Riemannian problem (2.5)(2.8) consists of left-invariant fields, so this is a left-invariant sub-Riemannian problem on the group $G=\operatorname{SE}(2)$.

According to the Agrachev–Barilari classification [1], this is the unique, up to local isometries, totally non-holonomic sub-Riemannian problem on $\operatorname{SE}(2)$; the corresponding invariants are $\chi=\kappa=1$.

That optimal controls exist in (2.5)(2.8) is a consequence of the Rashevskii–Chow and Filippov theorems: the system has full rank because

$$ \begin{equation*} \mathfrak{g}=\operatorname{span}(X_1,X_2,X_3), \qquad X_3=[X_1,X_2]. \end{equation*} \notag $$

2.4.2. The Pontryagin maximum principle

Abnormal trajectories are constant in time.

Normal extremals are trajectories of the Hamiltonian system $\dot\lambda=\vec H(\lambda)$, $\lambda \in T^*G$, where $H=(h_1^2+h_2^2)/2$, $h_i(\lambda)=\langle \lambda,X_i \rangle$, $i=1,2,3$. We can write this system in coordinates as follows:

$$ \begin{equation} \begin{alignedat}{3} \dot h_1&=- h_2 h_3, &\quad \dot h_2&=h_1 h_3, &\quad \dot h_3&=h_1 h_2, \\ \dot x&=h_1 \cos \theta, &\quad \dot y&=h_1 \sin \theta, &\quad \dot \theta&=h_2. \nonumber \end{alignedat} \end{equation} \tag{2.11} $$
On the level surface $\{H=1/2\}$, in the coordinates $(\gamma, c)$, where
$$ \begin{equation*} h_1=\sin \frac{\gamma}{2}\,, \quad h_2=-\cos \frac{\gamma}{2}\,, \quad\text{and}\quad c=2h_3, \end{equation*} \notag $$
the vertical subsystem (2.11) of this Hamiltonian system takes the form of a two- sheeted cover of the pendulum system:
$$ \begin{equation} \dot\gamma=c, \quad \dot c=-\sin\gamma, \quad (\gamma,c) \in C=\mathfrak{g}^*\cap\biggl\{H=\frac{1}{2}\biggr\} \cong (2S^1_{\gamma})\times \mathbb{R}_c, \quad 2S^1=\mathbb{R}/4\pi\mathbb{Z}. \end{equation} \tag{2.12} $$
A first integral of this equation is equal to the energy of the pendulum
$$ \begin{equation} E=\frac{c^2}{2}-\cos \gamma \in [-1,+\infty). \end{equation} \tag{2.13} $$

The symplectic foliation. The Lie coalgebra $\mathfrak{g}^*$ possesses a Casimir function $F=h_1^2+h_3^2$. The symplectic foliation consists of the circular cylinders $\{h_1^2+h_3^2=\operatorname{const} > 0\}$ and the points $\{h_1=h_3=0, \ h_2=\operatorname{const}\}$.

The energy of the pendulum is a linear combination of the Casimir function and the Hamiltonian:

$$ \begin{equation*} E=2F-2H. \end{equation*} \notag $$

Stratification of the cylinder $C$ and straightening coordinates. The cylinder $C$ is subdivided into invariant sets of the pendulum (2.12) by critical level curves of the energy $E$:

$$ \begin{equation} \begin{aligned} \, C&=\bigsqcup_{i=1}^5 C_i, \\ C_1&=\{\lambda \in C \mid E \in (-1, 1) \}, \nonumber \\ C_2&=\{\lambda \in C \mid E \in (1,+\infty) \}, \nonumber \\ C_3&=\{\lambda \in C \mid E =1, \ c \ne 0 \}, \nonumber \\ C_4&=\{\lambda \in C \mid E=- 1 \}= \{ (\gamma, c) \in C \mid \gamma=2 \pi n, \ c=0 \}, \nonumber \\ C_5&=\{\lambda \in C \mid E=1, \ c=0 \}= \{ (\gamma, c) \in C \mid \gamma=\pi+2 \pi n, \ c=0 \}, \nonumber \end{aligned} \end{equation} \tag{2.14} $$
where $n \in \mathbb{Z}$.

For the regular integration of the pendulum equation (2.12), on the strata $C_1$, $C_2$, and $C_3$ we introduce variables $(\varphi, k)$ which straighten this equation.

If $\lambda=(\gamma, c) \in C_1$, then

$$ \begin{equation*} \begin{gathered} \, k=\sqrt{\frac{E+1}{2}}= \sqrt{\sin^2 \frac{\gamma}{2}+\frac{c^2}{4}} \in (0,1), \\ \sin \frac{\gamma}{2}=s_1 k \operatorname{sn}(\varphi,k), \qquad s_1=\operatorname{sign} \cos\frac{\gamma}{2}\,, \\ \cos \frac{\gamma}{2}=s_1 \operatorname{dn}(\varphi,k), \\ \frac{c}{2}=k \operatorname{cn}(\varphi,k), \qquad \varphi \in [0, 4 K(k)]. \end{gathered} \end{equation*} \notag $$

If $\lambda=(\gamma, c) \in C_2$, then

$$ \begin{equation*} \begin{gathered} \, k=\sqrt{\frac{2}{E+1}}=\frac{1}{\sqrt{\sin^2(\gamma/2)+c^2/4}} \in (0,1), \\ \sin \frac{\gamma}{2}= s_2 \operatorname{sn}\biggl(\frac{\varphi}{k}\,,k\biggr), \qquad s_2=\operatorname{sign} c, \\ \cos \frac{\gamma}{2}=\operatorname{cn}\biggl(\frac{\varphi}{k}\,,k\biggr), \\ \frac{c}{2}= \frac{s_2}{k}\operatorname{dn}\biggl(\frac{\varphi}{k}\,,k\biggr), \qquad \varphi \in [0,4 k K(k)]. \end{gathered} \end{equation*} \notag $$

If $\lambda=(\gamma,c) \in C_3$, then

$$ \begin{equation*} \begin{gathered} \, k=1, \\ \sin \frac{\gamma}{2}=s_1 s_2 \tanh \varphi, \qquad s_1=\operatorname{sign} \cos\frac{\gamma}{2}\,, \quad s_2=\operatorname{sign} c, \\ \cos \frac{\gamma}{2}=\frac{s_1}{\cosh \varphi}\,, \\ \frac{c}{2}=\frac{s_2}{\cosh \varphi}\,, \qquad \varphi \in (-\infty, +\infty). \end{gathered} \end{equation*} \notag $$

The pendulum flow (2.12) straightens out in the variables $(\varphi,k)$:

$$ \begin{equation*} \dot \varphi=1, \quad \dot k=0, \qquad \lambda=(\varphi,k) \in \bigcup_{i=1}^3 C_i. \end{equation*} \notag $$

A parametrization on geodesics. If $\lambda=(\varphi,k) \in C_1$, then $\varphi_t=\varphi+t$ and

$$ \begin{equation*} \begin{gathered} \, \begin{aligned} \, \cos \theta_t&=\operatorname{cn} \varphi \operatorname{cn} \varphi_t+ \operatorname{sn} \varphi \operatorname{sn} \varphi_t, \\ \sin \theta_t&=s_1(\operatorname{sn} \varphi \operatorname{cn} \varphi_t- \operatorname{cn} \varphi \operatorname{sn} \varphi_t),\\ \theta_t&=s_1(\operatorname{am} \varphi- \operatorname{am} \varphi_t)\operatorname{mod}{2\pi}, \end{aligned} \\ \begin{aligned} \, x_t&=\frac{s_1}{k}[\operatorname{cn}\varphi(\operatorname{dn}\varphi- \operatorname{dn}\varphi_t)+\operatorname{sn} \varphi(t+\operatorname{E}(\varphi)-\operatorname{E}(\varphi_t))], \\ y_t&=\frac{1}{k}[\operatorname{sn}\varphi(\operatorname{dn}\varphi- \operatorname{dn}\varphi_t)-\operatorname{cn}\varphi(t+ \operatorname{E}(\varphi)-\operatorname{E}(\varphi_t))]. \end{aligned} \end{gathered} \end{equation*} \notag $$

If $\lambda \in C_2$, then

$$ \begin{equation*} \begin{gathered} \, \begin{aligned} \, \cos \theta_t&=k^2 \operatorname{sn} \psi \operatorname{sn} \psi_t+ \operatorname{dn} \psi \operatorname{dn} \psi_t, \\ \sin \theta_t&=k(\operatorname{sn} \psi \operatorname{dn} \psi_t- \operatorname{dn} \psi \operatorname{sn} \psi_t), \end{aligned} \\ \begin{aligned} \, x_t&=s_2 k\biggl[\operatorname{dn} \psi(\operatorname{cn} \psi- \operatorname{cn} \psi_t)+\operatorname{sn} \psi\,\biggl(\frac{t}{k}+ \operatorname{E}(\psi)-\operatorname{E}(\psi_t)\biggr)\biggr], \\ y_t&=s_2\biggl[k^2\operatorname{sn} \psi (\operatorname{cn} \psi- \operatorname{cn} \psi_t)-\operatorname{dn} \psi\,\biggl(\frac{t}{k}+ \operatorname{E}(\psi)-\operatorname{E}(\psi_t)\biggr)\biggr], \end{aligned} \end{gathered} \end{equation*} \notag $$
where $\psi=\varphi/k$ and $\psi_t=\varphi_t/k=\psi+t/k$.

If $\lambda=(\varphi,k) \in C_3$ for $k=1$, then $\varphi_t=\varphi+t$, and

$$ \begin{equation*} \begin{gathered} \, \begin{aligned} \, \cos \theta_t&=\frac{1}{\cosh \varphi\cosh \varphi_t}+ \tanh \varphi\tanh \varphi_t, \\ \sin \theta_t&= s_1\biggl(\frac{\tanh \varphi}{\cosh \varphi_t}- \frac{\tanh \varphi_t}{\cosh \varphi}\biggr), \end{aligned} \\ \begin{aligned} \, x_t&=s_1 s_2\biggl[\frac{1}{\cosh \varphi} \biggl(\frac{1}{\cosh \varphi}- \frac{1}{\cosh \varphi_t}\biggr)+\tanh \varphi\, (t+\tanh \varphi-\tanh \varphi_t)\biggr], \\ y_t&=s_2\biggl[\tanh \varphi\, \biggl(\frac{1}{\cosh \varphi}- \frac{1}{\cosh \varphi_t}\biggr)-\frac{1}{\cosh \varphi} (t+\tanh \varphi-\tanh \varphi_t)\biggr]. \end{aligned} \end{gathered} \end{equation*} \notag $$

If $\lambda \in C_4$, then

$$ \begin{equation*} \theta_t=-s_1 t, \qquad x_t=0, \qquad y_t=0. \end{equation*} \notag $$

If $\lambda \in C_5$, then

$$ \begin{equation*} \theta_t=0, \qquad x_t=t\operatorname{sign}\sin\frac{\gamma}{2}\,, \qquad y_t=0. \end{equation*} \notag $$

We show the projections of geodesics onto the $(x,y)$-plane in cases $C_1$, $C_2$, and $C_3$ in Figs. 3, 4, and 5, respectively.

2.4.3. Symmetries and Maxwell strata

The phase portrait of the pendulum system (2.12) is invariant under the symmetry group $\operatorname{Sym}$ generated by the reflections of the cylinder $C$ in the $\gamma$- and $c$-coordinate axes and in the origin $(\gamma,c)=(0,0)$ and the rotation through $2\pi$:

$$ \begin{equation*} \operatorname{Sym}=\{\operatorname{Id},\varepsilon^1,\dots,\varepsilon^7\} \cong \mathbb{Z}_2\times\mathbb{Z}_2\times\mathbb{Z}_2, \end{equation*} \notag $$
where
$$ \begin{equation*} \begin{aligned} \, \varepsilon^1\colon (\gamma,c)&\to (\gamma,-c), \\ \varepsilon^2\colon (\gamma,c)&\to (-\gamma,c), \\ \varepsilon^3\colon (\gamma,c)&\to (-\gamma,-c), \\ \varepsilon^4\colon (\gamma,c)&\to (\gamma+2 \pi,c), \\ \varepsilon^5\colon (\gamma,c)&\to (\gamma+2 \pi,-c), \\ \varepsilon^6\colon (\gamma,c)&\to (-\gamma+2 \pi,c), \\ \varepsilon^7\colon (\gamma,c)&\to (-\gamma+2 \pi,-c). \end{aligned} \end{equation*} \notag $$

These symmetries extend in the natural way to the source space and target space of the exponential map.

If $\nu=(\lambda,t)=(\gamma,c,t) \!\in\! N=C \times \mathbb{R}_+$, then $\varepsilon^i(\nu)=\nu^i=(\lambda^i,t)=(\gamma^i,c^i,t)\! \in\! N$ and

$$ \begin{equation*} \begin{aligned} \, (\gamma^1,c^1)&=(\gamma_t, -c_t), \\ (\gamma^2,c^2)&=(-\gamma_t, c_t), \\ (\gamma^3,c^3)&=(-\gamma, -c), \\ (\gamma^4,c^4)&=(\gamma+2 \pi, c), \\ (\gamma^5,c^5)&=(\gamma_t+2 \pi, -c_t), \\ (\gamma^6,c^6)&=(-\gamma_t+ 2 \pi, c_t), \\ (\gamma^7,c^7)&=(-\gamma, -c). \end{aligned} \end{equation*} \notag $$

If $g=(x,y,\theta) \in G$, then $g^i=\varepsilon^i(g)=(x^i,y^i,\theta^i) \in G$, where

$$ \begin{equation*} \begin{aligned} \, (x^1,y^1,\theta^1)&=(x \cos \theta+y \sin \theta, x \sin \theta- y \cos \theta, \theta), \\ (x^2,y^2,\theta^2)&=(-x \cos \theta- y \sin \theta, -x \sin \theta+y \cos \theta, \theta), \\ (x^3,y^3,\theta^3)&=(-x, -y, \theta), \\ (x^4,y^4,\theta^4)&=(-x, y, -\theta), \\ (x^5,y^5,\theta^5)&=(-x \cos \theta-y \sin \theta, x \sin \theta- y \cos \theta, -\theta), \\ (x^6,y^6,\theta^6)&=(x \cos \theta+y \sin \theta, -x \sin \theta+ y \cos \theta, -\theta), \\ (x^7,y^7,\theta^7)&=(x, -y, -\theta). \end{aligned} \end{equation*} \notag $$

Proposition 2.3. The group $\operatorname{Sym}= \{\operatorname{Id},\varepsilon^1,\dots,\varepsilon^7\}$ is a subgroup of the symmetry group of the exponential map.

Theorem 2.7. For almost all geodesics the first Maxwell time corresponding to the symmetry group $\operatorname{Sym}$ has the following expression:

$$ \begin{equation*} \begin{aligned} \, \lambda \in C_1 &\quad\Longrightarrow\quad t_{\rm Max}^1(\lambda)=2K(k), \\ \lambda \in C_2 &\quad\Longrightarrow\quad t_{\rm Max}^1(\lambda)=2kp_1^1(k), \\ \lambda \in C_3 &\quad\Longrightarrow\quad t_{\rm Max}^1(\lambda)=+\infty, \\ \lambda \in C_4 &\quad\Longrightarrow\quad t_{\rm Max}^1(\lambda)=\pi, \\ \lambda \in C_5 &\quad\Longrightarrow\quad t_{\rm Max}^1(\lambda)=+\infty, \end{aligned} \end{equation*} \notag $$
where $p=p_1^1(k) \in (K(k),2K(k))$ is the first positive zero of the function
$$ \begin{equation*} f_1(p,k)=\operatorname{cn}p(\operatorname{E}(p)-p)- \operatorname{dn} p \operatorname{sn} p. \end{equation*} \notag $$

Remark 2.1. If for a geodesic the first Maxwell time corresponding to the group $\operatorname{Sym}$ is distinct from $t_{\rm Max}^1$, then it is larger than this quantity, while $t_{\rm Max}^1$ is the first conjugate time.

Let $\vec H_v$ denote the vertical component of the Hamiltonian vector field $\vec H$ corresponding to the ordinary differential equation (2.11).

2.4.4. Estimates for the conjugate time

Theorem 2.9. (1) If $\lambda \in C_1\cup C_3\cup C_4\cup C_5$, then $t_{\rm conj}^1(\lambda)=+\infty$.

(2) If $\lambda \in C_2$, then $t_{\rm conj}^1(\lambda)\in [2kp_1^1,4kK]$.

(3) Therefore, $t_{\rm conj}^1(\lambda) \geqslant t_{\rm Max}^1(\lambda)$ for all $\lambda \in C$.

2.4.5. The diffeomorphism structure of the exponential map

Consider a subset of the state space that contains no fixed points of the maps $\varepsilon^i$:

$$ \begin{equation*} \begin{gathered} \, \widetilde G=\{g \in G \mid \varepsilon^i(g) \ne g, \ i=1,\dots,7\}= \{g \in G \mid R_1(g) R_2(g)\sin\theta\ne 0\}, \\ \text{where } R_1=y\cos \frac{\theta}{2}-x \sin\frac{\theta}{2} \quad\text{and}\quad R_2=x \cos \frac{\theta}{2}+y\sin\frac{\theta}{2}\,, \end{gathered} \end{equation*} \notag $$
and consider its partition into connected components:
$$ \begin{equation*} \widetilde G=\bigsqcup_{i=1}^8 G_i, \end{equation*} \notag $$
where each set $G_i$ is characterized by constant signs of the functions $\sin\theta$, $R_1$, and $R_2$, as indicated in Table 1.

Table 1.The definition of the domains $G_i$

$G_i$$G_1$$G_2$$G_3$$G_4$$G_5$$G_6$$G_7$$G_8$
$\operatorname{sign} \sin\theta$$-$$-$$-$$-$$+$$+$$+$$+$
$\operatorname{sign} R_1$$+$$+$$-$$-$$-$$-$$+$$+$
$\operatorname{sign} R_2$$+$$-$$-$$+$$+$$-$$-$$+$

Also consider an open dense subset of the space of potentially optimal geodesics:

$$ \begin{equation*} \widetilde N=\bigl\{(\lambda,t) \in N \mid t < t_{\rm Max}^1(\lambda), \ c_{t/2}\sin\gamma_{t/2} \ne 0\bigr\}, \end{equation*} \notag $$
and consider its connected components
$$ \begin{equation*} \begin{gathered} \, D_1=\bigl\{ (\lambda, t) \in N \mid t < t_{\rm Max}^1(\lambda), \ c_{t/2} > 0, \ \gamma_{t/2}\in(-\pi, 0)\bigr\}, \\ D_2=\bigl\{ (\lambda, t) \in N \mid t < t_{\rm Max}^1(\lambda), \ c_{t/2} > 0, \ \gamma_{t/2}\in(0, \pi)\bigr\}, \\ D_3=\bigl\{ (\lambda, t) \in N \mid t < t_{\rm Max}^1(\lambda), \ c_{t/2} < 0, \ \gamma_{t/2}\in(0, \pi)\bigr\}, \\ D_4=\bigl\{ (\lambda, t) \in N \mid t < t_{\rm Max}^1(\lambda), \ c_{t/2} < 0, \ \gamma_{t/2}\in(-\pi, 0)\bigr\}, \\ D_5=\bigl\{ (\lambda, t) \in N \mid t < t_{\rm Max}^1(\lambda), \ c_{t/2} > 0, \ \gamma_{t/2}\in(\pi, 2\pi)\bigr\}, \\ D_6=\bigl\{ (\lambda, t) \in N \mid t < t_{\rm Max}^1(\lambda), \ c_{t/2} > 0, \ \gamma_{t/2}\in(2\pi, 3\pi)\bigr\}, \\ D_7=\bigl\{ (\lambda, t) \in N \mid t < t_{\rm Max}^1(\lambda), \ c_{t/2} < 0, \ \gamma_{t/2}\in(2\pi, 3\pi)\bigr\}, \end{gathered} \end{equation*} \notag $$
and
$$ \begin{equation*} D_8=\bigl\{ (\lambda, t) \in N \mid t < t_{\rm Max}^1(\lambda), \ c_{t/2} < 0, \ \gamma_{t/2}\in(\pi, 2\pi)\bigr\}; \end{equation*} \notag $$
$\widetilde N=\bigsqcup_{i=1}^8D_i$.

Theorem 2.10. The following maps are diffeomorphisms:

$$ \begin{equation*} \begin{aligned} \, \operatorname{Exp}\colon D_i &\to G_i, \qquad i=1,\dots,8; \\ \operatorname{Exp}\colon \widetilde N &\to \widetilde G. \end{aligned} \end{equation*} \notag $$

2.4.6. The cut time

Theorem 2.11. For each $\lambda \in C$

$$ \begin{equation*} t_{\rm cut}(\lambda)=t_{\rm Max}^1(\lambda). \end{equation*} \notag $$

The cut time is invariant under the vertical component of the Hamiltonian field $\vec H_v$, so the sub-Riemannian structure on $\operatorname{SE}(2)$ is equioptimal.

2.4.7. The cut locus and its stratification

Theorem 2.12. The cut locus is a stratified two-dimensional manifold with the stratification

$$ \begin{equation*} \begin{gathered} \, \operatorname{Cut}=\operatorname{Cut}_{\rm glob} \sqcup \operatorname{Cut}_{\rm loc}^+ \sqcup \operatorname{Cut}_{\rm loc}^-, \\ {\begin{aligned} \, \operatorname{Cut}_{\rm glob}&=\{q \in M \mid \theta=\pi\}, \\ \operatorname{Cut}_{\rm loc}^+&=\{q \in M \mid \theta \in (- \pi, \pi), \ R_2=0, \ R_1 \geqslant R_1^1(|\theta|)\}, \\ \operatorname{Cut}_{\rm loc}^-&=\{q \in M \mid \theta\in (- \pi, \pi), \ R_2=0, \ R_1 \leqslant-R_1^1(|\theta|)\}, \end{aligned}} \end{gathered} \end{equation*} \notag $$
where
$$ \begin{equation*} \begin{alignedat}{2} R_1&=R_1^1(\theta), &\qquad \theta &\in [0,\pi], \\ R_1^1(\theta)&=2(F(v_1^1(k),k)-E(v_1^1(k),k)), &\qquad k&=k_1^1(\theta), \\ v_1^1(k)&=\operatorname{am}(p_1^1(k),k), &\qquad k &\in [0,1), \end{alignedat} \end{equation*} \notag $$
and $k=k_1^1(\theta)$ for $\theta \in [0,\pi]$ is the inverse of the decreasing function
$$ \begin{equation*} \theta(k)=2\arcsin(k\sin v_1^1(k)), \qquad k \in [0,1]. \end{equation*} \notag $$
The reference point $g_0=\operatorname{Id}$ belongs to the closures of both $\operatorname{Cut}_{\rm loc}^+$ and $\operatorname{Cut}_{\rm loc}^-$, but is separated from $\operatorname{Cut}_{\rm glob}$.

We show the cut locus $\operatorname{Cut} \subset \operatorname{SE}(2)$ in Fig. 6 (in the straightening coordinates $R_1=y \cos(\theta/2)-x \sin(\theta/2)$, $R_2=x \cos(\theta/2)+y \sin(\theta/2)$) and Fig. 7 (as embedded in the solid torus modelling the group $\operatorname{SE}(2)$).

2.4.8. Spheres

The sub-Riemannian spheres $S_R$ are homeomorphic (but not diffeomorphic) to

In Figures 8, 9, and 10 we show the sub-Riemannian spheres of radii $\pi/2$, $\pi$, and $3\pi/2$, respectively, as embedded in the solid torus modelling the group $\operatorname{SE}(2)$.

2.4.9. Metric straight lines

The metric straight lines through the identity element $g_0=\operatorname{Id}$ are $g(t)=\operatorname{Exp}(\lambda,t)$, $t \in \mathbb{R}$, where $\lambda \in C_3\cup C_5$. The geodesics $\operatorname{Exp}(\lambda,t)$, $\lambda \in C_3$, are projected onto tractrices on the $(x,y)$-plane, while geodesics $\operatorname{Exp}(\lambda,t)$, $\lambda \in C_5$, are projected onto the lines $(x,y)=(\pm t,0)$.

2.4.10. The bicycle model

We can regard the sub-Riemannian problem on $\operatorname{SE}(2)$ as the problem of the optimal motion of a bicycle model.

Assume that the front and rear wheels of a bicycle touch the ground at points $\mathbf{f}$ and $\mathbf{b}$, and the distance between these points (wheelbase) is fixed and equal to $\ell$. In the process of motion of the bicycle the points $\mathbf{f}$ and $\mathbf{b}$ draw two curves, the front and rear paths. At each moment of time the line segment $\mathbf{f}-\mathbf{b}$ is tangent to the rear path. We say that the motion is optimal if it minimizes the length of the front path. Then the problem of the optimal motion of a bicycle is just the sub-Riemannian problem (2.5)(2.8) on the group $\operatorname{SE}(2)$.

We say that two curves in the plane have the same shape if one of them can be taken to the other by a composition of motions and dilations. The width of a plane curve is the greatest lower bound of the distances between two parallel straight lines bounding a strip which contains this curve.

Theorem 2.13. An optimal trajectory of the front wheel of the bicycle $\mathbf{b}(t)$ is a straight line or an arc of a non-inflectional elastica of width at most $2\ell$. A non- inflectional elastica of any shape can be obtained in this way.

2.4.11. The isometry group and homogeneous geodesics

Theorem 2.15. The isometry group of the sub-Riemannian structure on $\operatorname{SE}(2)$ is

$$ \begin{equation*} \operatorname{Isom}(\operatorname{SE}(2))=\operatorname{SE}(2) \rtimes (\mathbb{Z}_2 \times \mathbb{Z}_2), \end{equation*} \notag $$
where the first factor $\operatorname{SE}(2)$ acts by left shifts on itself, the second factor $\mathbb{Z}_2$ acts on a pair $(\mathbf{b},\mathbf{f})$ as the reflection of the plane in one of the axes, and the third factor $\mathbb{Z}_2$ acts as the reflection $(\mathbf{b},\mathbf{f}) \mapsto (\mathbf{b},2\mathbf{b}-\mathbf{f})$.

A geodesic $\gamma$ on a sub-Riemannian manifold $M$ is said to be homogeneous if it is a homogeneous space of a one-parameter subgroup of $\operatorname{Isom}(M)$, that is, if there exists a one-parameter subgroup $\{\varphi_s \mid s \in \mathbb{R}\} \subset \operatorname{Isom}(M)$ such that

A sub-Riemannian manifold is said to be geodesically orbital if all geodesics on it are homogeneous.

Theorem 2.16. Homogeneous geodesics on $\operatorname{SE}(2)$ have the form $g(t)=\operatorname{Exp}(\lambda,t)$, $\lambda \in C_4\cup C_5$. They are one-parameter subgroups $e^{t X_2}$ and $e^{t X_1}$; their projections onto the $(x,y)$-plane are the point $(0,0)$ and the line $y=0$.

Thus, $\operatorname{SE}(2)$ is not a geodesically orbital space.

2.4.12. Bibliographic comments

Subsections 2.4.12.4.3 are based on [116], §§ 2.4.42.4.6, 2.4.8, and 2.4.9 on [134], § 2.4.7 on [135], and §§ 2.4.10 and 2.4.11 on [12] and [140].

The sub-Riemannian problem on $\operatorname{SE}(2)$ was also considered in [2], [12], [39], [111], and [142].

2.5. The sub-Riemannian problem on the motion group $\operatorname{SH}(2)$ of the pseudo-Euclidean plane

2.5.1. The motion group $\operatorname{SH}(2)$ of the pseudo- Euclidean plane

The pseudo- Euclidean plane is a two-dimensional real linear space endowed with the sign- indefinite bilinear form

$$ \begin{equation*} (\mathbf{x},\mathbf{y})=x_1 y_1-x_2 y_2, \qquad \mathbf{x}=(x_1, x_2), \quad \mathbf{y}=(y_1, y_2). \end{equation*} \notag $$
The distance $r$ between two points $\mathbf{x}=(x_1,x_2)$ and $\mathbf{y}=(y_1,y_2)$ is defined by
$$ \begin{equation*} r^2=(\mathbf{x}-\mathbf{y},\mathbf{x}-\mathbf{y})= (x_1-y_1)^2-(x_2-y_2)^2,\qquad r=\begin{cases} |r| &\text{ for }\ r^2 \geqslant 0, \\ i |r| &\text{ for }\ r^2 < 0. \end{cases} \end{equation*} \notag $$
The locus of points $\mathbf{x}=(x_1,x_2)$ lying at distance zero from the origin $(x_1^2-x_2^2=0$) is called the light cone. The complement of the pseudo-Euclidean plane to the light cone falls into four connected components, quadrants ($\operatorname{sign}(x_1-x_2)=\pm 1$, $\operatorname{sign}(x_1+x_2 )=\pm 1$).

The Lie group $\operatorname{SH}(2)$ and the Lie algebra $\mathfrak{sh}(2)$. A motion of the pseudo- Euclidean plane is a linear transformation that preserves the orientation, the quadrants, and the distances between points in the plane. The group of motions of the pseudo-Euclidean plane is denoted by $\operatorname{SH}(2)$. It has the linear representation

$$ \begin{equation*} \operatorname{SH}(2)=\left\{\begin{pmatrix} \cosh z & \sinh z & x \\ \sinh z & \cosh z & y \\ 0 & 0 & 1 \end{pmatrix}\ \bigg|\ x, y, z \in \mathbb{R}\right\}. \end{equation*} \notag $$
The action of a motion $g=(x,y,z)$ on a point $\mathbf{a}=(a_1,a_2)$ in the pseudo-Euclidean plane can be calculated using matrix multiplication:
$$ \begin{equation*} \begin{pmatrix} \cosh z & \sinh z & x \\ \sinh z & \cosh z & y \\ 0 & 0 & 1 \end{pmatrix}\begin{pmatrix} a_1 \\ a_2 \\ 1 \end{pmatrix}=\begin{pmatrix} a_1 \cosh z+a_2 \sinh z+x \\ a_1 \sinh z+a_2 \cosh z+y \\ 1 \end{pmatrix}, \end{equation*} \notag $$
so that $g \colon (a_1,a_2) \mapsto (a_1 \cosh z+a_2 \sinh z+x, a_1 \sinh z+a_2 \cosh z+y)$.

$G=\operatorname{SH}(2)$ is a Lie group with Lie algebra

$$ \begin{equation*} \mathfrak{g}=\mathfrak{sh}(2)= \operatorname{span}(E_{21}+E_{12},E_{13},E_{23}). \end{equation*} \notag $$
Basis left-invariant vector fields on $\operatorname{SH}(2)$ are
$$ \begin{equation*} \begin{gathered} \, X_1=L_{g*} E_{13}=\cosh z\,\frac{\partial}{\partial x}+ \sinh z\,\frac{\partial}{\partial y}\,, \\ X_2=L_{g*} (E_{21}+E_{12})=\frac{\partial}{\partial z}\,, \end{gathered} \end{equation*} \notag $$
and
$$ \begin{equation*} X_3=L_{g*} E_{23}=\sinh z\,\frac{\partial}{\partial x}+ \cosh z\,\frac{\partial}{\partial y} \end{equation*} \notag $$
with the multiplication table
$$ \begin{equation} [X_1, X_2]=- X_3, \quad [X_2. X_3]=X_1, \quad [X_1, X_3]=0. \end{equation} \tag{2.15} $$

2.5.2. The sub-Riemannian problem on $\operatorname{SH}(2)$

Consider the sub-Riemannian problem on the group $\operatorname{SH}(2)$ with orthonormal frame $(X_1, X_2)$:

$$ \begin{equation} \dot g=u_1 X_1+u_2 X_2, \quad g \in G=\operatorname{SH}(2), \quad u=(u_1, u_2) \in \mathbb{R}^2, \end{equation} \tag{2.16} $$
$$ \begin{equation} g(0)=g_0=\operatorname{Id}, \quad g(t_1)=g_1, \end{equation} \tag{2.17} $$
$$ \begin{equation} l=\int_0^{t_1} \sqrt{u_1^2 +u_2^2}\,dt \to \min. \end{equation} \tag{2.18} $$

By the Agrachev–Barilari classification [1] this is the unique, up to local isometries, non-integrable sub-Riemannian problem of rank 2 on $\operatorname{SH}(2)$; the corresponding invariants are $\chi=-\kappa=1$.

2.5.3. Geodesics

That optimal controls exist in problem (2.16)(2.18) follows from the Rashevskii–Chow and Filippov theorems.

The Pontryagin maximum principle. Abnormal trajectories are constant in time.

Normal extremals are the projections of trajectories of the Hamiltonian system $\dot\lambda=\vec{H}(\lambda)$, $\lambda \in T^*G$, where $H=(h_1^2+h_2^2)/2$, $h_i(\lambda)=\langle \lambda,X_i\rangle$, $i=1,2,3$. In coordinates this system looks as follows:

$$ \begin{equation} \dot{h}_{1} =h_{2}h_{0}, \end{equation} \tag{2.19} $$
$$ \begin{equation} \dot{h}_{2} =-h_{1}h_{0}, \end{equation} \tag{2.20} $$
$$ \begin{equation} \dot{h}_{0} = h_{1}h_{2}, \end{equation} \tag{2.21} $$
$$ \begin{equation} \dot{x} =h_{1}\cosh z, \nonumber \end{equation} \notag $$
$$ \begin{equation} \dot{y} =h_{1}\operatorname{sinh} z, \nonumber \end{equation} \notag $$
$$ \begin{equation} \dot{z} =h_{2}. \nonumber \end{equation} \notag $$
On the level surface $\{H=1/2\}$, in the coordinates $(\gamma,c)$, where
$$ \begin{equation*} h_1=\cos \frac{\gamma}{2}\,, \quad h_2=\sin \frac{\gamma}{2}\,, \quad c=-2 h_3, \end{equation*} \notag $$
the vertical subsystem (2.19)(2.21) assumes the form of a two-sheeted cover of the pendulum system:
$$ \begin{equation} \dot\gamma=c, \quad \dot c=-\sin \gamma, \qquad (\gamma,c) \in \mathfrak{g}^* \cap \biggl\{H=\frac{1}{2}\biggr\} \simeq (2 S^1_{\gamma}) \times \mathbb{R}_c. \end{equation} \tag{2.22} $$
A first integral of this equation is the energy of the pendulum
$$ \begin{equation*} E=\frac{c^2}{2}-\cos \gamma=2 h_3^2-h_1^2+h_2^2 \in [-1,+\infty). \end{equation*} \notag $$

The symplectic foliation. The Lie coalgebra $\mathfrak{g}^*$ possesses a Casimir function $F=h_1^2-h_3^2$. The symplectic foliation consists of

The energy of the pendulum is a linear combination of the Casimir function and the Hamiltonian: $E=2 H-2 F$.

The stratification of the cylinder $C$ and straightening coordinates. Since the vertical subsystem of the Hamiltonian system for the problem on $\operatorname{SH}(2)$, the pendulum system (2.22), coincides with a similar system (2.12) for the problem on $\operatorname{SE}(2)$, the stratification of $C$ and straightening coordinates $(\varphi,k)$ for the problem on $\operatorname{SH}(2)$ also coincide with the ones for the problem on $\operatorname{SE}(2)$ (see § 2.4.2).

A parametrization of geodesics. If $\lambda=(\varphi,k) \in C_1$, then $\varphi_t=\varphi+t$ and

$$ \begin{equation*} \begin{pmatrix} x\\ y\\ z \end{pmatrix}={\begin{pmatrix} \dfrac{s_{1}}{2}\biggl[\biggl(w+\dfrac{1}{w(1-k^{2})}\biggr) [\operatorname{E}(\varphi)-\operatorname{E}(\varphi_{0})]+ \biggl(\dfrac{k}{w(1-k^{2})}-kw\biggr)[\operatorname{sn}\varphi- \operatorname{sn}\varphi_{0}]\biggr] \\ \dfrac{1}{2}\biggl[\biggl(w-\dfrac{1}{w(1-k^{2})}\biggr) [\operatorname{E}(\varphi)-\operatorname{E}(\varphi_{0})]- \biggl(\dfrac{k}{w(1-k^{2})}+kw\biggr)[\operatorname{sn}\varphi- \operatorname{sn}\varphi_{0}]\biggr] \\ s_{1}\log[(\operatorname{dn}\varphi-k\operatorname{cn}\varphi)w] \end{pmatrix}}, \end{equation*} \notag $$
where $w=1/(\operatorname{dn}\varphi_{0}-k\operatorname{cn}\varphi_{0})$.

If $\lambda=(\varphi,k) \in C_2$, then $\psi=\varphi/k$, $\psi_t=\varphi_t/k=\psi+t/k$ and

$$ \begin{equation*} \begin{aligned} \, x&=\frac{1}{2}\biggl(\frac{1}{w(1-k^{2})}-w\biggr)[\operatorname{E}(\psi)- \operatorname{E}(\psi_{0})-k^{\prime2}(\psi-\psi_{0})] \\ &\qquad+\frac{1}{2}(kw+\frac{k}{w(1-k^{2})}) [\operatorname{sn}\psi-\operatorname{sn}\psi_{0}], \\ y&=-\frac{s_{2}}{2}\biggl(\frac{1}{w(1-k^{2})}+w\biggr) [\operatorname{E}(\psi)-\operatorname{E}(\psi_0)-k^{\prime2}(\psi-\psi_{0})] \\ &\qquad+\frac{s_{2}}{2}\biggl(kw-\frac{k}{w(1-k^{2})}\biggr) [\operatorname{sn}\psi-\operatorname{sn}\psi_{0}], \\ z&=s_{2}\log[(\operatorname{dn}\psi-k\operatorname{cn}\psi)w], \end{aligned} \end{equation*} \notag $$
where $w=1/(\operatorname{dn}\psi_{0}-k\operatorname{cn}\psi_{0})$.

If $\lambda=(\varphi,k) \in C_3$, $k=1$, then ${\varphi_t}=\varphi+t$ and

$$ \begin{equation*} \begin{pmatrix} x\\ y\\ z \end{pmatrix}=\begin{pmatrix} \dfrac{s_{1}}{2}\biggl[\dfrac{1}{w}(\varphi-\varphi_{0})+ w(\tanh \varphi-\tanh \varphi_{0})\biggr] \\ \dfrac{s_{2}}{2}\biggl[\dfrac{1}{w}(\varphi-\varphi_{0})- w(\tanh \varphi-\tanh \varphi_{0})\biggr] \\ -s_{1}s_{2}\log(w\operatorname{sech}\varphi) \end{pmatrix}, \end{equation*} \notag $$
where $w=\cos h \varphi_{0}$.

If $\lambda=(\gamma,c) \in C_4$, then

$$ \begin{equation*} \begin{pmatrix} x\\ y\\ z \end{pmatrix}=\begin{pmatrix} \operatorname{sign}\biggl(\cos\dfrac{\gamma}{2}\biggr)t \\ 0 \\ 0 \end{pmatrix}. \end{equation*} \notag $$

If $\lambda=(\gamma,c) \in C_5$, then

$$ \begin{equation*} \begin{pmatrix} x\\ y\\ z \end{pmatrix}=\begin{pmatrix} 0\\ 0\\ \operatorname{sign}\biggl(\sin\dfrac{\gamma}{2}\biggr)t \end{pmatrix}. \end{equation*} \notag $$

The projection of the geodesic onto the $(x,y)$-plane has curvature $\tan(\gamma/2)/\cosh^{3/2}(2z)$. It has inflection points for $\sin(\gamma/2)=0$ (provided that $\lambda \in C_1 \cup C_2 \cup C_3$) and cusp points for $\cos(\gamma/2)=0$ (provided that $\lambda \in C_2$).

2.5.4. Symmetries and Maxwell strata

The phase portrait of the pendulum system (2.22) has the symmetry group $\operatorname{Sym}=\{\operatorname{Id},\varepsilon^1,\dots,\varepsilon^7\}$, described in § 2.4.3. The extension of this group to the source space of the exponential map $N=C \times \mathbb{R}_+$ was also described there. The extension of this symmetry group to the target space of the exponential map has the form

$$ \begin{equation*} \varepsilon^i \colon g=(x,y,z) \mapsto g^i=\varepsilon^i(g)=(x^i,y^i,z^i), \end{equation*} \notag $$
where
$$ \begin{equation} \begin{aligned} \, (x^{1},y^{1},z^{1})&=(x\cosh z-y\operatorname{sinh} z, x\operatorname{sinh} z-y\cosh z,z), \\ (x^{2},y^{2},z^{2})&=(x\cosh z-y\operatorname{sinh} z, -x\operatorname{sinh} z+y\cosh z,-z), \\ (x^{3},y^{3},z^{3})&=(x,-y,-z), \\ (x^{4},y^{4},z^{4})&=(-x,y,-z), \\ (x^{5},y^{5},z^{5})&=(-x\cosh z+y\operatorname{sinh} z, x\operatorname{sinh} z-y\cosh z,-z), \\ (x^{6},y^{6},z^{6})&=(-x\cosh z+y\operatorname{sinh} z, -x\operatorname{sinh} z+y\cosh z,z), \\ (x^{7},y^{7},z^{7})&=(-x,-y,z). \end{aligned} \end{equation} \tag{2.23} $$

The following analogue of Proposition 2.3 holds.

Theorem 2.17. The first Maxwell time corresponding to the symmetry group $\operatorname{Sym}$ has the following expression for almost all geodesics:

$$ \begin{equation*} \begin{aligned} \, \lambda\in C_{1} & \ \ \Longrightarrow\ \ t_{\rm Max}^1(\lambda)=4K(k), \\ \lambda\in C_{2} & \ \ \Longrightarrow\ \ t_{\rm Max}^1(\lambda)=4kK(k), \\ \lambda\in C_{3}\cup C_{4}\cup C_{5} & \ \ \Longrightarrow \ \ t_{\rm Max}^1(\lambda)=+\infty. \end{aligned} \end{equation*} \notag $$

We also have the following result.

Corollary 2.1. For each $\lambda \in C$ the first Maxwell time $t_{\rm Max}^1$ is equal to the period of oscillations of the pendulum (2.22).

An analogue of Theorem 2.8 holds.

2.5.5. Estimates for the cut time

Let $p_1^1(k)\! \in\! (2K,3K)$ denote the first positive root of the equation $\operatorname{cn}p\operatorname{E}(p)- \operatorname{sn}p\operatorname{dn}p=0$.

Theorem 2.18. If $\lambda \in C_1$, then $4K(k) \leqslant t_{\rm conj}^1(\lambda) \leqslant 2 p_1^1(k)$. Moreover,

$$ \begin{equation*} \lim_{k \to+0} t_{\rm conj}^1(\lambda)=2 \pi\quad\textit{and}\quad \lim_{k \to 1-0} t_{\rm conj}^1(\lambda)=+\infty. \end{equation*} \notag $$

Theorem 2.19. If $\lambda \in C_2$, then $4k K(k) \leqslant t_{\rm conj}^1(\lambda) \leqslant 2 k p_1^1(k)$. Moreover,

$$ \begin{equation*} \lim_{k \to+0} t_{\rm conj}^1(\lambda)=0\quad\textit{and}\quad \lim_{k \to 1-0} t_{\rm conj}^1(\lambda)=+\infty. \end{equation*} \notag $$

Theorem 2.20. If $\lambda \in C_4$, then $t_{\rm conj}^1(\lambda)=2\pi$.

If $\lambda \in C_3 \cup C_5$, then $t_{\rm conj}^1(\lambda)=+\infty$.

2.5.6. The cut time

Theorem 2.22. For each $\lambda \in C$

$$ \begin{equation*} t_{\rm cut}(\lambda)=\min\bigl(t_{\rm Max}^1(\lambda),t_{\rm conj}^1(\lambda)\bigr)= \begin{cases} 4K(k), & \lambda \in C_1, \\ 4k K(k), & \lambda \in C_2, \\ 2\pi, & \lambda \in C_4, \\ +\infty, & \lambda \in C_3 \cup C_5. \end{cases} \end{equation*} \notag $$

Theorem 2.23. (1) The function $t_{\rm cut}\colon C\to (0,+\infty]$ depends only on the energy $E$ of the pendulum (2.22).

(2) The function $t_{\rm cut}$ is invariant under the vertical component of the Hamiltonian field $\vec{H}_v$ and the symmetries $\varepsilon^i \in \operatorname{Sym}$.

(3) The function $t_{\rm cut}$ is continuous on $C$ and smooth on $C_1 \cup C_2$.

(4) $\lim_{E \to-1}t_{\rm cut}=2 \pi$, $\lim_{E \to 1}t_{\rm cut}=+ \infty$, and $\lim_{E \to+\infty} t_{\rm cut}=0$.

2.5.7. The diffeomorphism structure of the exponential map

Consider an open dense subset of $G$ that contains no first Maxwell points:

$$ \begin{equation*} \widetilde{G}=\{g \in G \mid z \ne 0\}, \end{equation*} \notag $$
and consider its partition into connected components:
$$ \begin{equation*} \widetilde{G}=G_1 \sqcup G_2, \quad\text{where } G_1=\{g \in G \mid z > 0\}\quad\text{and} \quad G_2=\{g \in G \mid z < 0\}. \end{equation*} \notag $$
Also consider an open dense subset of the space of potentially optimal geodesics:
$$ \begin{equation*} \widetilde N=\biggl\{(\lambda,t) \in \bigcup_{i=1}^3 N_1 \cup N_5 \Bigm| t < t_{\rm cut}(\lambda), \ \sin\frac{\gamma_{t/2}}{2} \ne 0 \biggr\} \end{equation*} \notag $$
and its partition into connected components:
$$ \begin{equation*} \widetilde N=D_1 \sqcup D_2, \end{equation*} \notag $$
where
$$ \begin{equation*} D_1=\biggl\{(\lambda,t) \in \widetilde N \Bigm|\sin\frac{\gamma_{t/2}}{2} > 0 \biggr\} \end{equation*} \notag $$
and
$$ \begin{equation*} D_2=\biggl\{(\lambda,t) \in \widetilde N \Bigm|\sin\frac{\gamma_{t/2}}{2} < 0\biggr\}. \end{equation*} \notag $$

Theorem 2.24. The maps

$$ \begin{equation*} \operatorname{Exp}\colon D_i \to G_i, \qquad i=1, 2, \end{equation*} \notag $$
and
$$ \begin{equation*} \operatorname{Exp}\colon \widetilde N \to \widetilde{G} \end{equation*} \notag $$
are diffeomorphisms.

2.5.8. The cut locus

Theorem 2.25. The cut locus $\operatorname{Cut}$ lies in the plane $\{z=0\}$. It has a partition into connected components

$$ \begin{equation*} \operatorname{Cut}=\operatorname{Cut}_{\rm loc}^+ \sqcup \operatorname{Cut}_{\rm loc}^- \sqcup \operatorname{Cut}_{\rm glob}^+ \sqcup \operatorname{Cut}_{\rm glob}^-, \end{equation*} \notag $$
where The closures of the connected components $\operatorname{Cut}_{\rm loc}^{\pm}$ contain the origin $\operatorname{Id}$, whereas the closure of $\operatorname{Cut}_{\rm glob}^{\pm}$ does not.

We show the cut locus in Fig. 11. In Fig. 12 we show the cut locus and the first caustic $\operatorname{Conj}^1$.

2.5.9. Spheres

The sub-Riemannian spheres $S_R$, $R > 0$, are homeomorphic to the Euclidean 2-sphere (see the sphere $S_{\pi}$ in Fig. 13 and the sphere $S_{2 \pi}$ in Fig. 14).

These spheres have singularities at the points of their intersection with the cut locus (see the intersection of $\operatorname{Cut}$ and $S_{\pi} \cap \{z < 0\}$ in Fig. 15 and the intersection of $\operatorname{Cut}$ and $S_{2\pi} \cap \{z < 0\}$ in Fig. 16).

2.5.10. The structure of optimal synthesis

Theorem 2.26. (1) For each point $g_1 \in \operatorname{Cut} \setminus \operatorname{Conj}^1=\operatorname{int}_{\{z=0\}} \operatorname{Cut}$ there exist precisely two length minimizers connecting $\operatorname{Id}$ with $g_1$. For these curves $g_1$ is a cut point and a Maxwell point, but not a conjugate point.

(2) For each point $g_1 \in \operatorname{Cut} \mathop{\cap} \operatorname{Conj}^1= (\partial_{\{z=0\}} \operatorname{Cut}) \setminus \{\operatorname{Id}\}$ there exists a unique length minimizer connecting $\operatorname{Id}$ with $g_1$. For it $g_1$ is a cut point and a conjugate point, but not a Maxwell point.

(3) For each point $g_1 \in G \setminus(\operatorname{Cut} \cup \operatorname{Id})$ there exists a unique length minimizer connecting $\operatorname{Id}$ with $g_1$. For it $g_1$ is neither a cut point, nor a conjugate or a Maxwell point.

2.5.11. Metric lines

The metric lines through the identity element $\operatorname{Id}$ are

$$ \begin{equation*} g(t)=\operatorname{Exp}(\lambda,t), \qquad t \in \mathbb{R}, \quad \lambda \in C_3 \cup C_5. \end{equation*} \notag $$

2.5.12. Bibliographic comments

Subsection 2.5.1 is based on [159], §§ 2.5.2 and 2.5.3, on [58], §§ 2.5.4 and 2.5.5 on [59], and §§ 2.5.62.5.11 on [60].

2.6. Euler’s elastic problem

2.6.1. The history of the problem

In 1691 J. Bernoulli considered the problem of the shape of an elastic planar rod compressed by an external force. He deduced equations for an elastic rod which is fixed vertically to a horizontal wall and is flexed by a force applied horizontally to its upper end (a rectangular elastic):

$$ \begin{equation*} dy=\frac{x^2\,d x}{\sqrt{1-x^4}}\,, \quad ds=\frac{d x}{\sqrt{1-x^4}}\,, \qquad x \in [0,1), \end{equation*} \notag $$
where $(x,y)$ is the elastic rod and $s$ is the length parameter (the rod is deflected to a distance of 1 in the horizontal direction). Bernoulli integrated these equations by series and obtained two-sided estimates for their solution at the finite point $x=1$ [44].

In 1742 D. Bernoulli, in his letter [43] to Euler, wrote that the elastic energy of the rod is proportional to $J=\displaystyle\int\dfrac{ds}{R^2}$ , where $R$ is the curvature radius of the rod. He proposed to find the shape of elastica on the basis of the variational principle $J \to \min$. At that time Euler was writing his treatise [70] on variational calculus (which was published in 1744). He added a supplement De curvis elasticis to that book, where he applied the methods just developed to the problem of elastic rods. Euler considered a thin homogeneous elastic plate which has the form of a rectangle in its natural (stressless) state. He stated the following problem for the profile of the plate:

… among all curves of the same length which not only pass through the points A and B, but also are tangent to given straight lines at these points, that curve be determined in which the value of $\displaystyle\int_A^B\dfrac{ds}{R^2}$ is a minimum.”

Euler wrote down an equation for the corresponding variational problem, which is now known as the Euler–Lagrange equation, and reduced it to the equations

$$ \begin{equation*} dy=\frac{(\alpha+\beta x+\gamma x^2)\, dx} {\sqrt{a^4 -(\alpha+\beta x+\gamma x^2)^2}}\,,\qquad ds=\frac{a^2\,dx}{\sqrt{a^4-(\alpha+\beta x+\gamma x^2)^2}}\,, \end{equation*} \notag $$
with parameters expressed in terms of the elastic characteristics and length of the rod and the load value. Using the modern language, Euler examined the qualitative behaviour of the elliptic functions parametrizing the elastic curves by means of a qualitative analysis of the equations defining these curves. Subsequently, curves describing the shape of a homogeneous planar rod were called Euler elasticae. Euler describes all types of elasticae and indicated the values of the parameters for which these types appear. He divided the elasticae into nine classes (see Figs. 1925 below): Elasticae of types (ii)–(vi), which have inflection points, are called inflectional elasticae, ones of type (vii) are critical, and elasticae of type (viii), which have no inflection points, are non-inflectional elasticae. We show the family of all elasticae in Fig. 17.

The first explicit parametrization of Euler’s elasticae was due to Saalschütz [122] (1880).

In 1906, the future Nobel prize winner Born defended his thesis Stability of elastic curves in the plane and space [51]. He treated the problem of elasticae by methods of variational calculus and deduced from the Euler–Lagrange equations the equations

$$ \begin{equation*} \begin{gathered} \, \dot x=\cos \theta, \quad \dot y=\sin \theta, \\ A \ddot \theta+R \sin (\theta-\gamma)=0,\quad A,R,\gamma=\operatorname{const}, \end{gathered} \end{equation*} \notag $$
so that the slope angle $\theta$ of an elastica satisfies the equation of a mathematical pendulum. Then Born considered the question of the stability of elasticae with fixed endpoints and prescribed tangents at these endpoints. He proved that an arc without inflection points of an elastica is stable (on it the angle $\theta$ changes monotonically and can be taken as a parameter on the curve; Born showed that the second variation of the elastic energy functional $J=\dfrac{1}{2}\displaystyle\int \dot\theta^2\,dt$ is positive). In the general case Born presented a Jacobian vanishing at conjugate points. Since it involves quite complicated functions, Born limited himself to a numerical examination of conjugate points. He was the first to plot elasticae and verify theoretical results by experiments with elastic rods. In addition, he analyzed the stability of elasticae for various other types of boundary conditions, and obtained some results for elastic three-dimensional curves.

In 1993 Euler elasticae were discovered by Jurdjevic [86] in the problem of ball rolling over a plane without twisting or slipping (see § 2.8) and by Brockett and Dai [56] in the sub-Riemannian problem on the Cartan group (see § 2.10). Euler elasticae also appear miraculously in the Martinet flat sub-Riemannian problem (see § 2.3) and in sub-Riemannian problems on $\operatorname{SE}(2)$ (see § 2.4) and the Engel group (see § 2.9). In could be interesting to understand why Euler elasticae appear in so many optimal control problems.

Subsequently, Euler’s problem of elasticae was considered in [18], [111], [129]–[131], [133], [136], and [137], and our presentation in this section is based on these papers.

2.6.2. The problem statement

The mechanical setting. Assume that a uniform elastic rod on the plane $\mathbb{R}^2$ has length $l > 0$. Fix any points $a_0, a_1 \in \mathbb{R}^2$ and some unit tangent vectors $v_i \in T_{a_i} \mathbb{R}^2$, $|v_i|=1$, $i=0,1$. The problem consists in finding the profile of the rod $\gamma\colon[0,l]\to\mathbb{R}^2$, $|\dot \gamma(s)| \equiv 1$, going out of $a_0$ and coming into $a_1$ with tangent vectors $v_0$ and $v_1$, respectively:

$$ \begin{equation*} \begin{alignedat}{2} \gamma(0)&=a_0, &\qquad \gamma(l)&=a_1, \\ \dot\gamma(0)&=v_0, &\qquad \dot\gamma(l)&=v_1, \end{alignedat} \end{equation*} \notag $$
and with the minimum elastic energy
$$ \begin{equation*} J=\frac{1}{2} \int_0^l k^2(s)\,ds \to \min, \end{equation*} \notag $$
where $k(s)$ is the curvature of the curve $\gamma(s)$.

The optimal control problem. We introduce Cartesian coordinates $(x,y)$ in $\mathbb{R}^2$. We denote the length parameter $s$ on a curve $\gamma$ by $t$; let $t_1=l$. The curve to be determined has a parametrization $\gamma(t)=(x(t),y(t))$, $t \in [0,t_1]$, and its endpoints have coordinates $a_i=(x_i,y_i)$, $i=0,1$. Let $\theta(t)$ denote the angle between the tangent vector $\dot\gamma(t)$ and the positive direction of the $x$-axis. Finally, assume that the tangent vectors at the endpoints of $\gamma$ have coordinates $v_i=(\cos\theta_i,\sin\theta_i)$, $i=0,1$ (see Fig. 18).

Then the required curve $\gamma(t)=(x(t),y(t))$ is the projection of a trajectory of the control system

$$ \begin{equation} \dot x =\cos \theta, \end{equation} \tag{2.24} $$
$$ \begin{equation} \dot y =\sin \theta, \end{equation} \tag{2.25} $$
$$ \begin{equation} \dot \theta =u, \end{equation} \tag{2.26} $$
$$ \begin{equation} g =(x,y,\theta) \in M=\mathbb{R}^2_{x,y} \times S^1_{\theta},\quad u \in \mathbb{R}, \end{equation} \tag{2.27} $$
$$ \begin{equation} g(0) =g_0=(x_0,y_0,\theta_0), \quad g(t_1)=g_1=(x_1,y_1,\theta_1), \quad t_1 \text{ is fixed}. \end{equation} \tag{2.28} $$
For a curve $\gamma$ with natural parametrization the curvature is equal to the angular velocity: $k=\dot \theta=u$, so we obtain the quality functional
$$ \begin{equation} J=\frac{1}{2} \int_0^{t_1} u^2(t) \, dt\to \min\!. \end{equation} \tag{2.29} $$
A natural class of admissible controls in problem (2.24)(2.29) is $u(\,\cdot\,) \in L^2[0, t_1]$, so an admissible trajectory satisfies $g(\,\cdot\,) \in W^{1,2}([0,t_1], M)$.

In the vector notation the problems assumes the following form:

$$ \begin{equation} \begin{gathered} \, \dot g=X_1(g)+u X_2(g),\quad g \in M=\mathbb{R}^2 \times S^1,\quad u \in \mathbb{R}, \\ \nonumber g(0)=g_0, \quad g(t_1)=g_1, \quad t_1 \text{ is fixed}, \\ \nonumber J=\frac{1}{2} \int_0^{t_1} u^2\,dt \to \min, \quad u \in L^2[0,t_1], \end{gathered} \end{equation} \tag{2.30} $$
where the vector fields on the right in (2.30) are
$$ \begin{equation*} X_1=\cos \theta\,\frac{\partial}{\partial x}+ \sin \theta\,\frac{\partial}{\partial y} \quad\text{and}\quad X_2=\frac{\partial}{\partial \theta}\,. \end{equation*} \notag $$
The state space $M=\mathbb{R}^2 \times S^1$ has the natural structure of the planar motion group $G=\mathbb{R}^2 \rtimes \operatorname{SO}(2)$ (see § 2.4). Then $X_1$ and $X_2$ become left-invariant vector fields on the Lie group $G$. We presented the multiplication table for the Lie algebra $\mathfrak{g}=\mathfrak{se}(2)$ in (2.10).

Thus, Euler’s elastic problem (2.24)(2.29) is a left-invariant optimal control problem on the group $\operatorname{SE}(2)$. Hence we can assume that $g_0=\operatorname{Id}=(0,0,0)$.

2.6.3. The attainability set

Theorem 2.27. The attainability set of system (2.30) from the point $\operatorname{Id}=(0,0,0)$ in time $t_1 > 0$ is

$$ \begin{equation*} \mathcal{A}(t_1)=\{(x,y,\theta) \in G \mid x^2+y^2 < t_1^2 \textit{ or } (x,y,\theta)=(t_1,0,0)\}. \end{equation*} \notag $$

Topologically, $\mathcal{A}(t_1)$ is an open solid torus (the interior of a torus) with one boundary point. Below we consider the problem of elasticae under the natural controllability condition $g_1 \in \mathcal{A}(t_1)$.

2.6.4. Existence and boundedness of optimal controls

Theorem 2.28. Let $g_1 \in \mathcal{A}(t_1)$. Then an optimal control $u \in L^2[0, t_1]$ exists; furthermore, $u \in L^{\infty}[0, t_1]$. Hence the optimal control satisfies the Pontryagin maximum principle.

2.6.5. Extremals

Abnormal trajectories. The abnormal trajectory through the point $\operatorname{Id}$ with natural parametrization is $(x,y,\theta)\!=\!(t,0,0)$, $t \in [0,t_1]$. Its projection onto the $(x,y)$-plane is a line segment, an elastic rod in the absence of external effects. In this case the elastic energy attains its absolute minimum $J=0$, so the abnormal trajectory if optimal. This is the trajectory coming into the unique point $(t_1,0,0)$ on the boundary of the attainability set $\mathcal{A}(t_1)$. The abnormal trajectory is incidentally normal.

Normal extremals. Normal extremals satisfy the Hamiltonian system

$$ \begin{equation*} \dot \lambda=\vec{H}(\lambda),\qquad \lambda \in T^*G, \end{equation*} \notag $$
where $H=h_1+h_2^2/2$ and $h_i(\lambda)=\langle \lambda,X_i\rangle$, $i=1,2,3$. In the coordinate form this system looks like
$$ \begin{equation} \dot h_1=- h_2h_3, \quad \dot h_2=h_3, \quad \dot h_3=h_1h_2, \end{equation} \tag{2.31} $$
$$ \begin{equation} \dot g=X_1+h_2 X_2. \end{equation} \tag{2.32} $$
The vertical subsystem (2.31) has a first integral, the Casimir function $F=h_1^2+h_3^2$.

Consider the coordinates

$$ \begin{equation*} c=h_2, \quad h_1=- r \cos \gamma, \quad h_2=- r \sin \gamma, \end{equation*} \notag $$
in which (2.31) takes the form of a mathematical pendulum:
$$ \begin{equation} \dot \gamma=c, \quad \dot c=-r \sin \gamma, \qquad c \in \mathbb{R}, \quad \gamma \in S^1, \quad r \equiv \operatorname{const} \geqslant 0, \end{equation} \tag{2.33} $$
which is known as Kirchhoff’s kinetic analogue for elasticae. The full energy of the pendulum is
$$ \begin{equation*} E=H=\frac{c^2}{2}-r \cos \gamma \in [-r,+\infty). \end{equation*} \notag $$

Stratification of the source space of the exponential map and straightening coordinates. The exponential map for time $t_1 > 0$ in the problem of elasticae is

$$ \begin{equation*} \operatorname{Exp}_{t_1}\colon N=\mathfrak{g}^*\to G, \qquad \lambda \mapsto \pi \circ e^{t_1 \vec{H}}(\lambda), \end{equation*} \notag $$
where $\pi\colon T^*G \to G$ is the canonical projection.

The source space of the exponential map $N=\mathfrak{g}^*$ is partitioned into invariant manifolds of the Hamiltonian field $\vec{H}$ by critical levels of energy $E=H$:

$$ \begin{equation*} N=\bigsqcup_{i=1}^7 N_i, \end{equation*} \notag $$
where
$$ \begin{equation*} \begin{gathered} \, N_1=\{\lambda \in N \mid r \ne 0, \ E \in (-r, r)\}, \\ N_2=\{\lambda \in N \mid r \ne 0, \ E \in (r,+\infty)\}, \\ N_3=\{\lambda \in N \mid r \ne 0, \ E=r, \ \gamma \ne \pi\}, \\ N_4=\{\lambda \in N \mid r \ne 0, \ E =-r \}, \\ N_5=\{\lambda \in N \mid r \ne 0, \ E =r, \ \gamma=\pi\}, \\ N_6=\{\lambda \in N \mid r=0, \ c \ne 0 \}, \end{gathered} \end{equation*} \notag $$
and
$$ \begin{equation*} N_7=\{\lambda \in N \mid r=c=0\}. \end{equation*} \notag $$
On $N_1$, $N_2$, and $N_3$ we introduce a coordinate system $(\varphi,k,r)$ as follows:
$$ \begin{equation*} \begin{aligned} \, \lambda&=(\gamma,c,r) \in N_1 \quad\Longrightarrow\quad \begin{cases} \sin \dfrac{\gamma}{2}=k \operatorname{sn}(\sqrt{r}\,\varphi,k), \\ \dfrac{c}{2}=k \sqrt{r}\,\operatorname{cn}(\sqrt{r}\,\varphi,k), \\ \cos \dfrac{\gamma}{2}=\operatorname{dn}(\sqrt{r}\,\varphi,k), \end{cases} \\ &\qquad k=\sqrt{\dfrac{E+r}{2r}} \in (0,1), \quad \sqrt{r}\,\varphi\ \operatorname{mod}{4K(k)} \in [0,4K(k)]; \\ \lambda&=(\gamma,c,r) \in N_2 \quad\Longrightarrow\quad \begin{cases} \sin \dfrac{\gamma}{2}=\pm \operatorname{sn} \biggl(\dfrac{\sqrt{r}\,\varphi}{k}\,,k\biggr), \\ \dfrac{c}{2}=\pm \dfrac{\sqrt{r}}{k} \operatorname{dn}\biggl(\dfrac{\sqrt{r}\,\varphi}{k}\,,k\biggr), \\ \cos \dfrac{\gamma}{2}=\operatorname{cn} \biggl(\dfrac{\sqrt{r}\,\varphi}{k}\,,k\biggr), \end{cases} \\ &\qquad k=\sqrt{\frac{2 r}{E+r}} \in (0,1), \quad \sqrt{r}\, \varphi\ \operatorname{mod}{2K(k)k} \in [0,2K(k)k], \quad \pm=\operatorname{sign} c; \\ \lambda&=(\gamma,c,r) \in N_3 \quad\Longrightarrow\quad \begin{cases} \sin \dfrac{\gamma}{2}=\pm \tanh (\sqrt{r}\,\varphi), \\ \dfrac{c}{2}=\pm \dfrac{\sqrt{r}}{\cosh (\sqrt{r}\,\varphi)}\,, \\ \cos \dfrac{\gamma}{2}=\dfrac{1}{\cosh (\sqrt{r}\,\varphi)}\,, \end{cases} \\ &\qquad k=1, \quad \varphi \in \mathbb{R}, \quad \pm=\operatorname{sign} c. \end{aligned} \end{equation*} \notag $$

A parametrization of extremals. In the domain $N_1 \cap N_2 \cup N_3$ the pendulum equation ‘straightens out’:

$$ \begin{equation*} \dot \varphi=1, \quad \dot k=\dot r=0, \end{equation*} \notag $$
so it has the solutions
$$ \begin{equation*} \varphi_t=\varphi+t, \quad k,r \equiv \operatorname{const}. \end{equation*} \notag $$
In the original coordinates $(\gamma,c)$ the pendulum equation (2.33) has the following solutions:
$$ \begin{equation*} \begin{aligned} \, \lambda \in N_1 &\quad\Longrightarrow\quad \begin{cases} \sin \dfrac{\gamma_t}{2}=k_1 \operatorname{sn}(\sqrt{r}\,\varphi_t), \\ \cos \dfrac{\gamma_t}{2}=\operatorname{dn}(\sqrt{r}\,\varphi_t), \\ \dfrac{c_t}{2}=k \sqrt{r}\,\operatorname{cn}(\sqrt{r}\,\varphi_t); \end{cases} \\ \lambda \in N_2 &\quad\Longrightarrow\quad \begin{cases} \sin \dfrac{\gamma_t}{2}=\pm \operatorname{sn} \biggl(\dfrac{\sqrt{r}\,\varphi_t}{k}\biggr), \\ \cos \dfrac{\gamma_t}{2}=\operatorname{cn} \biggl(\dfrac{\sqrt{r}\,\varphi_t}{k}\biggr), \\ \dfrac{c_t}{2}=\pm \dfrac{\sqrt{r}}{k} \operatorname{dn} \biggl(\dfrac{\sqrt{r}\,\varphi_t}{k}\biggr), \end{cases} \qquad \pm=\operatorname{sign}c; \\ \lambda \in N_3 &\quad\Longrightarrow\quad \begin{cases} \sin\dfrac{\gamma_t}{2}=\pm\tanh (\sqrt{r}\,\varphi_t), \\ \cos \dfrac{\gamma_t}{2}=\dfrac{1}{\cosh (\sqrt{r}\,\varphi_t)}\,, \\ \dfrac{c_t}{2}=\pm\dfrac{\sqrt{r}}{\cosh (\sqrt{r}\,\varphi_t)}, \end{cases} \qquad \pm=\operatorname{sign}c. \end{aligned} \end{equation*} \notag $$
In the degenerate cases corresponding to $\bigcup_{i=4}^7 N_i$ the pendulum equation (2.33) can be integrated by elementary functions:
$$ \begin{equation*} \begin{alignedat}{2} \lambda \in N_4 &\quad\Longrightarrow\quad \gamma_t \equiv 0, &\qquad c_t &\equiv 0; \\ \lambda \in N_5 &\quad\Longrightarrow\quad \gamma_t \equiv \pi, &\qquad c_t &\equiv 0; \\ \lambda \in N_6 & \quad\Longrightarrow\quad \gamma_t=c t+\gamma, &\qquad c_t &\equiv c; \\ \lambda \in N_7 &\quad\Longrightarrow\quad c_t \equiv 0, &\qquad r &\equiv 0. \end{alignedat} \end{equation*} \notag $$

Solutions of the horizontal subsystem (2.32) have the following parametrization.

If $\lambda \in N_1$, then

$$ \begin{equation*} \begin{aligned} \, \sin \frac{\theta_t}{2} &= k \operatorname{dn}(\sqrt{r}\,\varphi) \operatorname{sn}(\sqrt{r}\,\varphi_t)-k\operatorname{sn}(\sqrt{r}\,\varphi) \operatorname{dn}(\sqrt{r}\,\varphi_t), \\ \cos \frac{\theta_t}{2} &= \operatorname{dn}(\sqrt{r}\,\varphi) \operatorname{dn}(\sqrt{r}\,\varphi_t)+k^2\operatorname{sn}(\sqrt{r}\,\varphi) \operatorname{sn} (\sqrt{r}\,\varphi_t), \\ x_t &= \frac{2}{\sqrt r} \operatorname{dn}^2 (\sqrt r\,\varphi) \bigl(\operatorname{E}(\sqrt r\,\varphi_t)- \operatorname{E}(\sqrt r\,\varphi)\bigr) \\ &\qquad+\frac{4k^2}{\sqrt r}\operatorname{dn}(\sqrt{r}\,\varphi) \operatorname{sn}(\sqrt{r}\,\varphi) \bigl(\operatorname{cn}(\sqrt{r}\,\varphi)- \operatorname{cn} (\sqrt{r}\,\varphi_t)\bigr) \\ &\qquad+\frac{2k^2 }{\sqrt r} \operatorname{sn}^2(\sqrt{r}\,\varphi) \bigl(\sqrt r\,t+\operatorname{E}(\sqrt r\,\varphi)- \operatorname{E}(\sqrt r\,\varphi_t)\bigr)-t, \\ y_t &= \frac{2k}{\sqrt r}(2 \operatorname{dn}^2(\sqrt{r}\,\varphi)-1) \bigl(\operatorname{cn}(\sqrt{r}\,\varphi)- \operatorname{cn}(\sqrt{r}\,\varphi_t)\bigr) \\ &\qquad-\frac{2k}{\sqrt r}\operatorname{sn}(\sqrt{r}\,\varphi) \operatorname{dn}(\sqrt{r}\,\varphi) \bigl[2\bigl(\operatorname{E}(\sqrt r\,\varphi_t)- \operatorname{E}(\sqrt r\,\varphi)\bigr)-\sqrt r\,t\bigr]. \end{aligned} \end{equation*} \notag $$

If $\lambda \in N_2$, then

$$ \begin{equation*} \begin{aligned} \, \sin \frac{\theta_t}{2} &=\pm\bigl(\operatorname{cn}(\sqrt r\,\psi) \operatorname{sn}(\sqrt r\,\psi_t)-\operatorname{sn}(\sqrt r\,\psi) \operatorname{cn}(\sqrt r\,\psi_t)\bigr), \\ \cos \frac{\theta_t}{2} &=\operatorname{cn}(\sqrt r\,\psi) \operatorname{cn}(\sqrt r\,\psi_t)+\operatorname{sn}(\sqrt r\,\psi) \operatorname{sn}(\sqrt r\,\psi_t), \\ x_t &= \frac{1}{\sqrt r}\bigl(1-2 \operatorname{sn}^2(\sqrt r\,\psi)\bigr) \biggl[\frac{2}{k}\bigl(\operatorname{E}(\sqrt r\,\psi_t)- \operatorname{E}(\sqrt r\,\psi)\bigr)-\frac{2-k^2}{k^2}\sqrt r\,t\biggr] \\ &\qquad+ \frac{4}{k \sqrt r} \operatorname{cn}(\sqrt r\,\psi) \operatorname{sn}(\sqrt r\,\psi)\bigl(\operatorname{dn}(\sqrt r\,\psi)- \operatorname{dn}(\sqrt r\,\psi_t)\bigr), \\ y_t &= \pm \biggl(\frac{2}{k\sqrt r} \bigl(2\operatorname{cn}^2(\sqrt r\,\psi)-1\bigr) \bigl(\operatorname{dn}(\sqrt r\,\psi)-\operatorname{dn}(\sqrt r\,\psi_t)\bigr) \\ &\qquad- \frac{2}{\sqrt r} \operatorname{sn}(\sqrt r\,\psi) \operatorname{cn}(\sqrt r\,\psi)\biggl[\frac{2}{k} \bigl(\operatorname{E}(\sqrt r\,\psi_t)-\operatorname{E}(\sqrt r\,\psi)\bigr)- \frac{2-k^2}{k^2}\sqrt r\,t\biggr]\biggr), \end{aligned} \end{equation*} \notag $$
where $\pm=\operatorname{sign} c$ and $\psi_t=\varphi_t/k=(\varphi+t)/k$.

If $\lambda \in N_3$, then

$$ \begin{equation*} \begin{gathered} \, \begin{aligned} \, \sin \frac{\theta_t}{2} &= \pm \biggl(\frac{\tanh (\sqrt r\,\varphi_t)}{\cosh (\sqrt r\,\varphi)}- \frac{\tanh (\sqrt r\,\varphi)} {\cosh (\sqrt r\,\varphi_t)}\biggr), \\ \cos \frac{\theta_t}{2} &= \frac{1}{\cosh (\sqrt r\,\varphi) \cosh (\sqrt r\,\varphi_t)}+\tanh (\sqrt r\,\varphi) \tanh (\sqrt r\,\varphi_t), \end{aligned} \\ \begin{aligned} \, x_t &=(1-2\tanh^2(\sqrt r\,\varphi)) t +\frac{4\tanh (\sqrt r\,\varphi)} {\sqrt r\,\cosh (\sqrt r\,\varphi)} \biggl(\frac{1}{\cosh (\sqrt r\,\varphi)} -\frac{1}{\cosh (\sqrt r\,\varphi_t)}\biggr), \\ y_t &=\pm\biggl[\frac{2}{\sqrt r} \biggl(\frac{2}{\cosh^2(\sqrt r\,\varphi)}-1\biggr) \biggl(\frac{1}{\cosh (\sqrt r\,\varphi)} -\frac{1}{\cosh (\sqrt r\,\varphi_t)}\biggr)\biggr. -2\,\frac{\tanh (\sqrt r\,\varphi)} {\cosh (\sqrt r\,\varphi)}\,t \biggr], \end{aligned} \end{gathered} \end{equation*} \notag $$
where $\pm=\operatorname{sign} c$.

If $\lambda \in N_4 \cup N_5 \cup N_7$, then

$$ \begin{equation*} \theta_t=0, \quad x_t=t, \quad\text{and}\quad y_t=0. \end{equation*} \notag $$

If $\lambda \in N_6$, then

$$ \begin{equation*} \theta_t=ct, \quad x_t=\frac{\sin (ct)}{c}\,, \quad y_t=\frac{1-\cos (ct)}{c}\,. \end{equation*} \notag $$

Euler elasticae. The projections of extremal trajectories onto the $(x,y)$-plane are Euler elasticae. Their equations are

$$ \begin{equation} \begin{gathered} \, \dot x=\cos \theta, \quad \dot y=\sin \theta, \nonumber \\ \ddot \theta=-r \sin(\theta-\gamma), \qquad r,\gamma \equiv \operatorname{const}. \end{gathered} \end{equation} \tag{2.34} $$
Depending on the energy of the pendulum $E=\dot\theta^2/2-r \cos(\theta-\gamma) \in [-r,+\infty)$ and the value of the Casimir function $r \geqslant 0$, elasticae have qualitatively different types, which were discovered by Euler.

If the energy $E$ takes the minimum value $-r < 0$, so that $\lambda \in N_4$, then the elastica $(x_t,y_t)$ is a straight line. The corresponding motion of the pendulum (2.34) (Kirchhoff’s kinetic analogue) is staying in stable equilibrium.

If $E \in (-r,r)$, $r > 0$, so that $\lambda \in N_1$, then the pendulum (2.34) oscillates between the extremal values of the angle and the angular velocity $\dot\theta$ changes sign. The corresponding elasticae have inflection points for $\dot \theta=0$ and vertices for $|\dot\theta|=\max$, because $\dot \theta$ is the curvature of the elastic. Such elasticae are said to be inflectional (see Figs. 1923). Different cases in these figures depend on the value of the modulus of elliptic functions $k=\sqrt{E+r}/(2r) \in (0,1)$:

$$ \begin{equation*} \begin{aligned} \, k \in \biggl(0,\frac{1}{\sqrt 2}\biggr) &\quad\Longrightarrow\quad \text{Fig. 19}, \\ k=\frac{1}{\sqrt 2} &\quad\Longrightarrow\quad \text{Fig. 20}, \\ k \in \biggl(\frac{1}{\sqrt 2}\,,k_0\biggr) &\quad\Longrightarrow\quad \text{Fig. 21}, \\ k=k_0 &\quad\Longrightarrow\quad \text{Fig. 22}, \\ k \in \biggl(k_0,1\biggr) &\quad\Longrightarrow\quad \text{Fig. 23}. \end{aligned} \end{equation*} \notag $$
The value $k=1/\sqrt 2$ corresponds to the rectangular elastica, considered by J. Bernoulli (see § 2.6.1): see Fig. 20. The value $k \approx 0.909$ corresponds to a periodic elastica of the form of a figure-of-eight: see Fig. 22. As Euler noted, for $k \to 0$ inflectional elasticae are similar to sinusoids, which corresponds to a harmonic oscillator $\ddot \theta=- r (\theta-\gamma)$ as Kirchhoff’s kinetic analogue; see Fig. 19.

If $E=r > 0$ and $\theta-\gamma \ne \pi$, so that $\lambda \in N_3$, then the pendulum (2.34) tends to an unstable equilibrium ($\theta-\gamma=\pi$, $\dot \theta=0$) along a separatrix of a saddle, and the corresponding elastica (an ‘Euler soliton’) makes one loop: see Fig. 24.

If $E=r > 0$ and $\theta-\gamma=\pi$, that is, $\lambda \in N_5$, then the pendulum (2.34) is in an unstable equilibrium ($\theta-\gamma=\pi$, $\dot \theta=0$) and the elastica is a straight line.

If $E > r > 0$, so that $\lambda \in N_2$, then Kirchhoff’s kinetic analogue is the pendulum (2.34) rotating anticlockwise (for $\dot \theta > 0$) or clockwise (for $\dot \theta < 0$). The corresponding elasticae have a non-zero curvature $\dot \theta$ and no inflection points; they are said to be non-inflectional: see Fig. 25.

If $r=0$ and $\dot \theta \ne 0$, so that $\lambda \in N_6$, then the pendulum (2.34) rotates uniformly in zero gravity. The corresponding elastica is a circle.

Finally, if $r=0$ and $\dot \theta=0$, so that $\lambda \in N_7$, then the pendulum (2.34) rests in zero gravity (so that the equilibrium is unstable) and the elastica is a straight line.

In the pictures of elasticae in Figs. 1925 the ratio $x/y$ is not always correctly reproduced for reasons of space.

Periodic motions of the pendulum (2.33) (or (2.34)) have the period

$$ \begin{equation*} T=\begin{cases} 4\,\dfrac{K(k)}{\sqrt r}\,, & \lambda \in N_1, \\ 2\,\dfrac{k K(k)}{\sqrt r}\,, & \lambda \in N_2, \\ \dfrac{2\pi}{|c|}\,, & \lambda \in N_6. \end{cases} \end{equation*} \notag $$

2.6.6. Symmetries and Maxwell strata

The phase portrait of the pendulum (2.33) is preserved by the symmetry group $\operatorname{Sym}$, which is generated by the reflection $\varepsilon^1$ in the $\gamma$-axis, the reflection $\varepsilon^2$ in the $c$-axis, and the reflection $\varepsilon^3$ in the origin $(\gamma,c)=(0,0)$:

$$ \begin{equation*} \operatorname{Sym}= \{\operatorname{Id},\varepsilon^1,\varepsilon^2,\varepsilon^3\} \simeq \mathbb{Z}_2 \times \mathbb{Z}_2. \end{equation*} \notag $$
These symmetries extend in a natural way to the source space $N=\mathfrak{g}^*$ and the target space $G$ of the exponential map $\operatorname{Exp}_t$. If $\nu=(\gamma,c,r) \in N$, then
$$ \begin{equation*} \varepsilon^i(\nu)=\nu^i=(\gamma^i,c^i,r) \in N, \end{equation*} \notag $$
where
$$ \begin{equation*} (\gamma^1,c^1)=(\gamma_t,-c_t), \quad (\gamma^2,c^2)=(-\gamma_t, c_t), \quad (\gamma^3,c^3)=(-\gamma,-c). \end{equation*} \notag $$
If $g=(x,y,\theta) \in G$, then
$$ \begin{equation*} \varepsilon^i(g)=(x^i,y^i,\theta^i) \in G, \end{equation*} \notag $$
where
$$ \begin{equation*} \begin{aligned} \, (x^1, y^1, \theta^1)&= (x \cos \theta+y \sin \theta,-x \sin \theta+y \cos \theta,-\theta), \\ (x^2, y^2, \theta^2)&= (x \cos \theta+y \sin \theta, x \sin \theta-y \cos \theta, \theta), \\ (x^3, y^3, \theta^3)&=(x,-y,-\theta). \end{aligned} \end{equation*} \notag $$

Proposition 2.4. The group $\operatorname{Sym}= \{\operatorname{Id},\varepsilon^1,\varepsilon^2,\varepsilon^3\}$ consists of the symmetries of the exponential map.

Theorem 2.29. For almost all extremal trajectories $g_t=\operatorname{Exp}_t(\lambda)$, $\lambda \in N$, the first Maxwell time corresponding to the symmetry group $\operatorname{Sym}$ has the following expression:

$$ \begin{equation*} \begin{gathered} \, \lambda \in N_1 \quad\Longrightarrow\quad t_{\rm Max}^1=\frac{2}{\sqrt r} p_1(k), \\ \lambda \in N_2 \quad\Longrightarrow\quad t_{\rm Max}^1=\frac{2}{\sqrt r} kK(k), \\ \lambda \in N_6 \quad\Longrightarrow\quad t_{\rm Max}^1=\frac{2\pi}{|c|}\,, \\ \lambda \in N_3 \cup N_4 \cup N_5 \cup N_7 \quad\Longrightarrow\quad t_{\rm Max}^1=+\infty, \end{gathered} \end{equation*} \notag $$
where
$$ \begin{equation*} p_1(k)=\min\bigl(2K(k),p_z^1(k)\bigr)=\begin{cases} 2 K(k), & k \in (0,k_0], \\ p_z^1(k), & k \in (k_0,1), \end{cases} \end{equation*} \notag $$
$p=p_z^1(k) \in (K,3K)$ is the first positive root of the equation $\operatorname{sn} p \operatorname{dn} p-(2\operatorname{E}(p)- p)\operatorname{cn}p=0$, and $k_0 \approx 0.909$ is a root of the equation $2E(k)-K(k)=0$.

An observation similar to Remark 2.1 and a result on invariant properties of the function $t_{\rm Max}^1\colon N\to(0,+\infty]$ which is similar to Theorem 2.8 are true.

2.6.7. Estimates for the conjugate time

For Euler elasticae, a very important question from the point of view of applications is their local optimality since it means the stability of the elastica under small perturbations of the profile such that the endpoints and tangents at these endpoints are fixed. From the theoretic point of view answering this question is a step towards the investigation of the global optimality of elasticae.

Theorem 2.30. Let $\lambda=(k,\varphi,r) \in N_1$. Then the first conjugate time $t_{\rm conj}^1(\lambda)$ on the trajectory $\operatorname{Exp}_t(\lambda)$ belongs to the closed interval with endpoints $4 K(k)/\sqrt r$ and $2 p_1(k)/\sqrt r$, namely,

$$ \begin{equation*} \begin{alignedat}{3} &(1)&\quad k &\in (0,k_0) &\quad\Longrightarrow\quad t_{\rm conj}^1 &\in \biggl[\frac{4K(k)}{\sqrt r}\,, \frac{2 p_1^1(k)}{\sqrt r}\biggr], \\ &(2)&\quad k &=k_0 &\quad\Longrightarrow\quad t_{\rm conj}^1&=\frac{4K(k)}{\sqrt r}= \frac{2 p_1^1(k)}{\sqrt r}\,, \\ &(3)&\quad k &\in (k_0,1) &\quad\Longrightarrow\quad t_{\rm conj}^1 &\in \biggl[\frac{2 p_1^1(k)}{\sqrt r}\,, \frac{4K(k)}{\sqrt r}\biggr], \end{alignedat} \end{equation*} \notag $$
where the function $p_1(k)$ was defined in Theorem 2.29.

Corollary 2.2. Let $\lambda=(k,\varphi,r) \in N_1$. Then

$$ \begin{equation*} \begin{alignedat}{3} &(1)&\quad k &\in (0,k_0) &\quad\Longrightarrow\quad t_{\rm conj}^1 &\in [T, t_1^1] \subset \biggl[T,\frac{3T}{2}\biggr), \quad t_1^1=\frac{2 p_1^1}{\sqrt r} \in \biggl(T,\frac{3T}{2}\biggr), \\ &(2)&\quad k &=k_0 &\quad\Longrightarrow\quad t_{\rm conj}^1&=T, \\ &(3)&\quad k &\in (k_0,1) &\quad\Longrightarrow\quad t_{\rm conj}^1 &\in [t_1^1,T] \subset \biggl(\frac{T}{2}\,,T\biggr],\quad t_1^1=\frac{2p_1^1}{\sqrt r} \in \biggl(\frac{T}{2}\,,T\biggr), \end{alignedat} \end{equation*} \notag $$
where $T=4 K(k)/\sqrt r$ is the period of oscillations of the pendulum (2.33) (or (2.34)).

Corollary 2.3. Let $\lambda=(k,\varphi,r) \in N_1$ and $t_1 > 0$, and let

$$ \begin{equation} \Gamma=\{(x_t, y_t) \mid t \in [0, t_1]\},\qquad g(t)=(x_t, y_t, \theta_t)=\operatorname{Exp}_t(\lambda), \end{equation} \tag{2.35} $$
be an arc of the corresponding elastica.

(1) If $\Gamma$ contains no inflection points, then it is locally optimal.

(2) If $k \in (0,k_0]$ and $\Gamma$ contains precisely one inflection point, then it is locally optimal.

(3) If $\Gamma$ contains at least three interior inflection points, then it is not locally optimal.

Consider arcs of inflectional elasticae (2.35) with midpoint at a vertex, that is, assume that the local extremum of the curvature of the elastica is attained at $(x_{t_1/2},y_{t_1/2})$. For examples of such arcs, see Fig. 26.

Set $t_1^1=(2/\sqrt r\,)p_1(k)$, where the function $p_1(k)$ was defined in Theorem 2.29.

Now we consider arcs of inflectional elasticae (2.35) centred at an inflection point, that us, we assume that the elastica has zero curvature at $(x_{t_1/2},y_{t_1/2})$. For examples of such arcs, see Fig. 27.

Theorem 2.33. Let $\lambda \in N_2 \cup N_3 \cup N_6$. Then the extremal trajectory $g(t)=\operatorname{Exp}_t(\lambda)$ contains no conjugate points for $t > 0$.

Thus, if an arc of an elastica contains no inflection points, then it is stable; if it contains at least three interior inflection points, then it is unstable. If there are one or two inflection points, then the elastica can be stable or unstable alike.

2.6.8. The diffeomorphism structure of the exponential map

Let $t_1=1$, $\operatorname{Exp}=\operatorname{Exp}_1$, and

$$ \begin{equation*} \mathcal{A}=\mathcal{A}_1=\{(x,y,\theta) \in G \mid x^2+y^2 < 1 \text{ or } (x,y,\theta)=(1,0,0)\}. \end{equation*} \notag $$
The case of general $t_1 > 0$ reduces to $t_1=0$ by homotheties of the $(x,y)$-plane:
$$ \begin{equation*} (x,y,\theta,t,u,t_1,J) \mapsto (\tilde x,\tilde y,\tilde\theta,\tilde t, \tilde u,\tilde t_1,\tilde J)=(e^sx,e^sy,\theta,e^st,e^{-s}u,e^st_1,e^{-s}J). \end{equation*} \notag $$

Consider a subset of $\mathcal{A}$ containing no fixed points of the reflections $\varepsilon^1$ and $\varepsilon^2$:

$$ \begin{equation*} \begin{gathered} \, \widetilde{G}=\{g \in \mathcal{A} \mid \varepsilon^i (g) \ne g, \ i=1, 2\}= \biggl\{g \in \mathcal{A} \Bigm| \sin\biggl(\frac{\theta}{2}\biggr)P(g) \ne 0 \biggr\}, \\ P(g)=x \sin \frac{\theta}{2}-y \cos \frac{\theta}{2}\,, \end{gathered} \end{equation*} \notag $$
and consider its partition into connected components
$$ \begin{equation*} \widetilde{G}=G_+ \sqcup G_-, \end{equation*} \notag $$
where
$$ \begin{equation*} G_{\pm}=\{ g \in G \mid \theta \in (0, 2\pi), \ x^2+y^2 < 1, \ \operatorname{sign} P(g)=\pm 1 \}. \end{equation*} \notag $$
Also consider the open dense subset of the space of potentially optimal extremal trajectories
$$ \begin{equation*} \widetilde{N}=\biggl\{\lambda \in \bigcup_{i=1}^3 N_i \Bigm| t_1 < t_{\rm Max}^1(\lambda), \ c_{t_1/2} \sin\gamma_{t_1/2} \ne 0\biggr\}, \end{equation*} \notag $$
and its connected components
$$ \begin{equation*} \widetilde{N}=\bigsqcup_{i=1}^4 D_i, \end{equation*} \notag $$
where
$$ \begin{equation*} \begin{gathered} \, D_1=\biggl\{\lambda \in \bigcup_{i=1}^3 N_i \Bigm| t_1 < t_{\rm Max}^1(\lambda), \ c_{t_1/2} > 0, \ \sin \gamma_{t_1/2} > 0\biggr\}, \\ D_2=\biggl\{\lambda \in \bigcup_{i=1}^3 N_i \Bigm| t_1 < t_{\rm Max}^1(\lambda), \ c_{t_1/2} < 0, \ \sin \gamma_{t_1/2} > 0\biggr\}, \\ D_3=\biggl\{\lambda \in \bigcup_{i=1}^3 N_i \Bigm| t_1 < t_{\rm Max}^1(\lambda), \ c_{t_1/2} < 0, \ \sin \gamma_{t_1/2} < 0\biggr\}, \end{gathered} \end{equation*} \notag $$
and
$$ \begin{equation*} D_4=\biggl\{\lambda \in \bigcup_{i=1}^3 N_i \Bigm| t_1 < t_{\rm Max}^1(\lambda), \ c_{t_1/2} > 0, \ \sin \gamma_{t_1/2} < 0\biggr\}. \end{equation*} \notag $$

Theorem 2.34. The following maps are diffeomorphisms:

$$ \begin{equation*} \operatorname{Exp}\colon D_1\to G_+, \quad \operatorname{Exp}\colon D_2\to G_-, \quad \operatorname{Exp}\colon D_3\to G_+, \quad \operatorname{Exp}\colon D_4\to G_-. \end{equation*} \notag $$

Corollary 2.4. The map $\operatorname{Exp}\colon\widetilde{N}\to \widetilde{G}$ is a two-sheeted cover.

2.6.9. Optimal elasticae for various boundary conditions

Generic boundary conditions. If $g_1 \in G_+$, then there exists a unique pair $(\lambda_1,\lambda_3) \in D_1 \times D_3$ such that $\operatorname{Exp}(\lambda_1)=\operatorname{Exp}(\lambda_3)=g_1$. The optimal trajectory is either $q^1(t)=\operatorname{Exp}_t(\lambda_1)$ or $q^3(t)=\operatorname{Exp}_t(\lambda_3)$, $t \in [0,1]$. As the optimal trajectory we must take the one at which the quality functional $J[q^i(\,\cdot\,)]=\dfrac{1}{2}\displaystyle\int_0^1 (c_t^i)^2\,dt$ takes the smaller value. If $J[q^1(\,\cdot\,)]=J[q^3(\,\cdot\,)]$, then both trajectories are optimal (we show this case in Fig. 28).

If $g_1 \in G_-$, then we select optimal elasticae in a similar way among the ones corresponding to the covectors $\lambda_2 \in D_2$ and $\lambda_4 \in D_4$ for which $\operatorname{Exp}(\lambda_2)=\operatorname{Exp}(\lambda_4)= g_1$.

The case $(x_1,y_1,\theta_1)=(1,0,0)$. The optimal elastica is the line segment $(x,y)=(t,0)$, $t \in [0,1]$.

The case $x_1>0$, $y_1=0$, $\theta_1=\pi$. In this case $g_1 \in G_+$ and the equation $\operatorname{Exp}(\lambda)=g_1$, $\lambda \in \widetilde{G}$, has too roots, $\lambda_1 \in D_1$ and $\lambda_3 \in D_3$. The trajectories $q^1(t)=\operatorname{Exp}_t(\lambda_1)$ and $q^3(t)=\operatorname{Exp}_t(\lambda_3)$ correspond to the same value of $J$, so both of them are optimal. The corresponding optimal inflection elasticae are mirror symmetric with respect to the $x$-axis (see Fig. 29).

The case $x_1<0$, $y_1=0$, $\theta_1=\pi$. This is similar to the previous case (see Fig. 30).

The case $x_1=0$, $y_1=0$, $\theta_1=\pi$. The only optimal elastica, a ‘water drop’, is defined by the parameters $\lambda=(\varphi,k,r) \in N_1$, where $\varphi=\dfrac{\tau}{2p}-\dfrac{1}{2}$ , $r=4 p^2$, $\operatorname{sn}\tau=0$, $1-2k^2\operatorname{sn}^2 p=0$, and $2\operatorname{E}(p)-p=0$ (see Fig. 31).

The case $x_1=0$, $y_1=0$, $\theta_1=0$. There exist two optimal elasticae, which are circles mirror symmetric with respect to the $x$-axis.

The case $x_1>0$, $y_1=0$, $\theta_1=0 $. There are two or four optimal elasticae: there exists $x_* \in (0.4,0.5)$ such that

The case $x_1<0$, $y_1=0$, $\theta_1=0$. There are two optimal non-inflectional elasticae (see Fig. 33).

2.6.10. Bibliographic comments

The first classical study of elasticae was presented by Euler in [70].

Subsection 2.6.1 concerning the history of the problem of elasticae is based on the classical sources [105], [153], and [154]. A remarkable presentation of this history was given in [99].

Note that the problem of elasticae had long been of purely theoretical interest as an example of the use of the theory of elliptic functions (see, for instance, [78] and [105]). In connection with widespread introduction of steel in project design and construction of flexible thin-walled structures, which encouraged the development of the stability theory of deformed systems, the solution of the problem of elasticae became a question of practical importance. In particular, the following questions arose, which are important for applications to engineering: what is the behaviour of a compressed rack under loads higher than the Euler critical value, what is the shape of the rack in this case, is it unique and stable? A number of papers (see [49], [69], [75], [96], [119], [148], [149], [161]) were concerned with these questions, considering various conditions for supporting and loading inextensible elastic rods. During the last decades the interest to elasticae has grown because of applications of the theory of flexible rods to the analysis of micro- and nanostructures in biology and nanotechnology [76], [84], [115], [151]. It has been confirmed that there exists a variety of equilibrium shapes for a fixed load.

Subsections 2.6.22.6.6 are based on [130], § 2.6.7 is based on [131] and [133], and §§ 2.6.8 and 2.6.9 on [137] and [136].

Problems of elasticae were also considered in [2], [9], [12], [79], [87]–[89], [106], [107], [117], and [142].

2.7. A left-invariant sub-Riemannian problem of general form on the group $\operatorname{SO}(3)$

2.7.1. The problem statement

It follows from the classification of contact left- invariant sub-Riemannian structures on three-dimensional Lie groups [1] that, given an arbitrary structure of this type on the group $G=\operatorname{SO}(3)$, we can find an orthonormal frame $(X_1,X_2)$ with multiplication table

$$ \begin{equation} [X_2,X_1]=X_3, \quad [X_1,X_3]=(\kappa+\chi)X_2, \quad [X_2,X_3]=(\chi-\kappa)X_1, \end{equation} \tag{2.36} $$
where $\kappa$ and $\chi$, $\kappa \geqslant \chi \geqslant 0$, are differential invariants of the sub-Riemannian structure. A uniform extension of $X_1$ and $X_2$ changes the distance function and both invariants $\kappa$ and $\chi$ in the same proportion. In [1] the normalization $\kappa^2+\chi^2=1$ was used. In this section it is more convenient to assume that $\kappa+\chi=1$ and use the invariant $a=\sqrt{2\chi} \in [0,1)$. The case $a=0$ corresponds to the axially symmetric sub-Riemannian structure considered in [55].

The following vector fields satisfy the multiplication table (2.36):

$$ \begin{equation*} X_1(g)=L_{g*} A_2, \quad X_2(g)=\sqrt{1-a^2}\,L_{g*} A_1, \quad X_3(g)=\sqrt{1-a^2}\,L_{g*} A_3, \end{equation*} \notag $$
where the basis $A_1$, $A_2$, $A_3$ of the Lie algebra $\mathfrak{g}=\mathfrak{so}(3)$ has the form
$$ \begin{equation} A_1=\begin{pmatrix} 0 & 0 & 0 \\ 0 & 0 & -1 \\ 0 & 1 & 0 \end{pmatrix},\quad A_2=\begin{pmatrix} 0 & 0 & 1 \\ 0 & 0 & 0 \\ -1 & 0 & 0 \end{pmatrix},\quad A_3=\begin{pmatrix} 0 & -1 & 0 \\ 1 & 0 & 0 \\ 0 & 0 & 0 \end{pmatrix}. \end{equation} \tag{2.37} $$

2.7.2. A parametrization of geodesics

Abnormal extremal trajectories are constant in time.

To parametrize normal geodesics consider the Hamiltonians

$$ \begin{equation*} h_i(\lambda)=\langle \lambda,X_i(g)\rangle,\quad i=1,2,3,\qquad H=\frac{1}{2}(h_1^2+h_2^2). \end{equation*} \notag $$
Extremals with normal parametrization are themselves parametrized by points in the cylinder $C=\mathfrak{g}^* \cap \{H=1/2\}$. We introduce the coordinates $(\psi,c)$ on this cylinder:
$$ \begin{equation*} h_1=\cos \psi, \quad h_2=- \sin \psi, \quad h_3=c. \end{equation*} \notag $$

The normal Hamiltonian system of the Pontryagin maximum principle has the form

$$ \begin{equation} \dot h_1=h_2 h_3, \quad \dot h_2=- h_1 h_3, \quad \dot h_3=a^2 h_1 h_2, \end{equation} \tag{2.38} $$
$$ \begin{equation} \dot g=h_1 X_1(g)+h_2 X_2(g). \end{equation} \tag{2.39} $$
The vertical subsystem (2.38) defines the pendulum equation
$$ \begin{equation} \dot \psi=c, \quad \dot c=-\frac{a^2}{2} \sin(2\psi) \end{equation} \tag{2.40} $$
on $C$. The cylinder is stratified by invariant sets of the system (2.40),
$$ \begin{equation*} C=\bigsqcup_{i=1}^5 C_i, \end{equation*} \notag $$
which are defined in terms of the full energy $E=2c^2-a^2\cos(2\psi)$ of the pendulum:
$$ \begin{equation*} \begin{gathered} \, C_1=\{\lambda\in C\mid E\in(-a^2,a^2)\}\quad (\text{the region inside the separatrices}), \\ C_2=\{\lambda\in C\mid E\in(a^2, +\infty)\}\quad (\text{the region outside the separatrices}), \\ C_3=\{\lambda\in C\mid E=a^2, c \ne 0\}\quad (\text{the separatrices}), \\ C_4=\{\lambda\in C\mid E=-a^2\}\quad (\text{the stable equilibrium}), \end{gathered} \end{equation*} \notag $$
and
$$ \begin{equation*} C_5=\{\lambda\in C\mid E=a^2, c=0\}\quad (\text{the unstable equilibrium}). \end{equation*} \notag $$
On $C_1$, $C_2$, and $C_3$ we introduce coordinates $(\theta,k)$ which straighten the equation (2.40). In the domain $C_1$ we have
$$ \begin{equation*} \begin{gathered} \, \sin\psi=s_1k\operatorname{sn}(a\theta,k), \qquad \cos\psi=s_1\operatorname{dn}(a\theta,k), \qquad c=ak\operatorname{cn}(a\theta,k), \\ s_1=\operatorname{sign}\cos\psi,\qquad k=\sqrt{\frac{E+a^2}{2a^2}}\in(0,1),\qquad \theta\in\biggl[0,\frac{4K(k)}{a}\biggr]. \end{gathered} \end{equation*} \notag $$
In the domain $C_2$,
$$ \begin{equation*} \begin{gathered} \, \sin\psi=s_2\operatorname{sn}\biggl(\frac{a\theta}{k}\,,k\biggr),\qquad \cos\psi=\operatorname{cn}\biggl(\frac{a\theta}{k}\,,k\biggr), \qquad c=\frac{s_2a}{k}\operatorname{dn}\biggl(\frac{a\theta}{k}\,,k\biggr), \\ s_2=\operatorname{sign}c,\qquad k=\sqrt{\frac{2a^2}{E+a^2}}\in (0,1),\qquad \theta \in\biggl[0,\frac{4kK(k)}{a}\biggr]. \end{gathered} \end{equation*} \notag $$
On the set $C_3$,
$$ \begin{equation*} \sin\psi=s_1s_2\tanh(a\theta),\ \ \cos\psi=\frac{s_1}{\cosh(a\theta)}\,,\ \ c=\frac{s_2a}{\cosh(a\theta)}\,, \ \ \theta \in(-\infty,+\infty), \ \ k=1. \end{equation*} \notag $$

Then for $(\psi_0,c_0) \in C_1 \cup C_2 \cup C_3$ the pendulum equation has the solution $\theta(t)=t+\theta_0$, $k \equiv \operatorname{const}$. For $(\psi_0,c_0) \in C_4$ we have $\psi \equiv \pi n$, $n \in \mathbb{Z}$, $c=0$, while for $(\psi_0,c_0) \in C_5$ we have $\psi \equiv-\pi/2+\pi n$, $n \in \mathbb{Z}$, $c=0$.

To parametrize the solutions of the horizontal subsystem (2.39) we represent them in terms of the Euler angles:

$$ \begin{equation*} g_t=\exp(-\varphi_1(0) A_3) \exp(-\varphi_2(0) A_1) \exp(\varphi_3(t) A_3) \exp(\varphi_2(t) A_1) \exp(\varphi_1(t) A_3). \end{equation*} \notag $$
Then
$$ \begin{equation} \cos \varphi_2 =\frac{c}{\sqrt M}\,, \qquad \sin \varphi_2 =\sqrt{\frac{M-c^2}{M}}\,, \end{equation} \tag{2.41} $$
$$ \begin{equation} \cos \varphi_1 =\frac{h_1\sqrt{1-a^2}}{\sqrt{M-c^2}}\,, \qquad \sin \varphi_1 =\sqrt{\frac{h_2}{M-c}}\,, \end{equation} \tag{2.42} $$
where $M=h_2^2+(1-a^2) h_1^2+c^2$ is a first integral of the subsystem (2.38).

The angle $\varphi_3$ satisfies the equation

$$ \begin{equation} \dot\varphi_3=\frac{\sqrt{M (1-a^2)}}{M-c^2}= \frac{\sqrt{M(1-a^2)}}{1-a^2 h_1^2}; \end{equation} \tag{2.43} $$
it is a monotone function of time because
$$ \begin{equation*} 0 < \sqrt{M(1-a^2)} \leqslant \dot \varphi_3 \leqslant \sqrt{\frac{M}{1-a^2}}\,. \end{equation*} \notag $$
Solutions of this equation have the following form:
  • $\bullet$ in $C_1$,
    $$ \begin{equation*} \begin{aligned} \, \varphi_3&=\sqrt{\frac{1-a^2(1-k^2)}{a^2(1-a^2)}}\, \biggl[\Pi\biggl(\frac{a^2k^2}{a^2-1}\,; \operatorname{am}(a\theta,k),k\biggr) \\ &\qquad-\Pi\biggl(\frac{a^2k^2}{a^2-1}\,; \operatorname{am}(a\theta_0,k),k\biggr)\biggr]; \end{aligned} \end{equation*} \notag $$
  • $\bullet$ in $C_2$,
    $$ \begin{equation*} \begin{aligned} \, \varphi_3&=\sqrt{\frac{k^2+a^2(1-k^2)}{a^2(1-a^2)}}\, \biggl[\Pi\biggl(\frac{a^2}{a^2-1}\,;\operatorname{am} \biggl(\frac{a\theta}{k}\,,k\biggr),k \biggr) \\ &\qquad-\Pi\biggl(\frac{a^2}{a^2-1}\,;\operatorname{am} \biggl(\frac{a\theta_0}{k}\,,k\biggr),k\biggr)\biggr]; \end{aligned} \end{equation*} \notag $$
  • $\bullet$ in $C_3$,
    $$ \begin{equation*} \begin{aligned} \, \varphi_3&=\sqrt{1-a^2}\,t+\biggl[\arctan\biggl(\frac{a}{\sqrt{1-a^2}}\tanh(a\theta)\biggr) \\ &\qquad-\arctan\biggl(\frac{a}{\sqrt{1-a^2}}\tanh(a\theta_0)\biggr)\biggr]; \end{aligned} \end{equation*} \notag $$
  • $\bullet$ in $C_4$,
    $$ \begin{equation*} \varphi_3=t; \end{equation*} \notag $$
  • $\bullet$ in $C_5$,
    $$ \begin{equation*} \varphi_3=\sqrt{1-a^2}\,t. \end{equation*} \notag $$
Here $\operatorname{am}(\varphi,k)$ is the Jacobi amplitude and $\Pi(n;\varphi,k)$ is an elliptic integral of the third kind. Note that, as we see from the last two expressions, geodesics corresponding to the domains $C_4$ and $C_5$ are rotations about the horizontal basis vectors $e_1=(1,0,0) \in \mathbb{R}^3$ and $e_2=(0,1,0) \in \mathbb{R}^3$.

2.7.3. Periodic geodesics

Proposition 2.5. For each $a \in (0,1)$ there exist infinitely many geodesics in the corresponding sub-Riemannian problem on $\operatorname{SO}(3)$.

For $\lambda \in C_1$ (or $\lambda \in C_2$) the relevant geodesic can only have period $T=4K(k)/a$ ($T=4k K(k)/a$, respectively); such trajectories exist if and only if $\varphi_3(m T)=2 \pi n$ for some $n,m \in \mathbb{N}$. This equality holds along a geodesic if and only if

$$ \begin{equation} \frac{n}{m} > \frac{1}{a} \end{equation} \tag{2.44} $$
in the case of $C_1$ or
$$ \begin{equation} \frac{n}{m} > 1 \end{equation} \tag{2.45} $$
in the case of $C_2$. Different irreducible fractions $n/m \in \mathbb{Q}_+$ correspond to different periodic geodesics.

Proposition 2.6. Every periodic geodesic for $\lambda \in C_1$ (or $\lambda \in C_2$) is uniquely defined by an irreducible fraction $n/m \in \mathbb{Q}_+$ satisfying (2.44) (respectively, (2.45)).

For $\lambda \in C_3$ geodesics are aperiodic.

For $\lambda \in C_4 \cup C_5$ geodesics are periodic.

Since $\pi_1(\operatorname{SO}(3))=\mathbb{Z}_2$, there exist only two homotopy classes of closed paths on $\operatorname{SO}(3)$. The following result shows which periodic geodesics are contractible (null homotopic).

Proposition 2.7. Consider a periodic geodesic $g_t \in \operatorname{SO}(3)$ which is the projection of an extremal $\lambda_t \in C_1$ (or $\lambda_t \in C_2$) and is defined by an irreducible fraction $n/m \in \mathbb{Q}_+$ satisfying (2.44) (respectively, (2.45)). Then $g_t$ is reducible if and only if $n$ is even.

All geodesics corresponding to $\lambda \in C_4 \cup C_5$ are non-contractible.

2.7.4. Conditions for optimality

Consider the unit 3-sphere in the algebra of quaternions

$$ \begin{equation*} S^3=\{q=q^0+i q^1+j q^2+k q^3 \in \mathbb{H} \mid (q^0)^2+(q^1)^2+(q^2)^2+(q^3)^2=1\}. \end{equation*} \notag $$
It is simply connected and covers the group $\operatorname{SO}(3)$ with multiplicity two. The geodesic $g_t \in \operatorname{SO}(3)$ lifts to a curve $q_t \in S^3$, $q_0=1$, of the form
$$ \begin{equation*} q_t=\exp\biggl(-\frac{\varphi_1(0)}{2}k\biggr) \exp\biggl(-\frac{\varphi_2(0)}{2}i\biggr) \exp\biggl(\frac{\varphi_3(t)}{2}k\biggr) \exp\biggl(\frac{\varphi_2(t)}{2}i\biggr) \exp\biggl(\frac{\varphi_1(t)}{2}k\biggr), \end{equation*} \notag $$
where the Euler angles $\varphi_i(t)$ coincide with the analogous angles in (2.41)(2.43).

Theorem 2.35. Let $g_t \in \operatorname{SO}(3)$, $t \in [0,t_1]$, be a geodesic, and let $q_t \in S^3$, $q_0=1$, be its lift to $S^3$. Let $\theta_t$ be the corresponding straightened coordinate variable of the pendulum system (2.40), and let $\tau=a(\theta_0+t/2)$.

Then $g_t$ is not optimal, provided that one of the following conditions holds for some $t \in (0,t_1)$:

(1) $q_t^0=0$;

(2) $q_t^1=0$ and $\operatorname{sn} \tau \ne 0$ for $\lambda_0 \in C_1 \cup C_2$, or $\tau \ne 0$ for $\lambda_0 \in C_3$;

(3) $q_t^2=0$ and $\operatorname{cn} \tau \ne 0$ for $\lambda_0 \in C_1$;

(4) $q_t^3=0$ and $\operatorname{cn} \tau \ne 0$ for $\lambda_0 \in C_2$.

2.7.5. Bibliographic comments

The results presented in this section were obtained in [46].

2.8. The problem of a ball rolling over a plane without twisting or slipping

2.8.1. The history of the problem

In 1983 Hammersley [83] considered the following Oxford ball problem. A unit ball lies on an infinite horizontal plane. The state of the ball is determined by its orientation in space and its position on the plane. We must take the ball from the given initial state to a prescribed terminal state by a sequence of rollings. Each rolling proceeds along a straight line on the plane: we select the distance and direction of each rolling, but there must be no twisting or slipping, that is, the axis of rotation must remain horizontal and the velocity of the ball at the point of contact with the plane must be equal to zero. What minimum number of rollings $N$ do we require to attain any terminal state? With the help of quaternions Hammersley showed that $N \in \{3,4\}$. Next he stated two continuous versions of the ball problem:

In problem (b) Hammersley pointed out that the optimal curve $\Gamma$ is a line segment or an arc of a circle and $0 \leqslant T \leqslant \pi\sqrt 3$ , where the upper bound is only attained when the required change of orientation is the rotation of the sphere through $\pi$ about the vertical axis.

In the final section of [83], entitled “Variants for the twenty-first century” Hammersley stated a number of variants and generalizations of these problems, which are still open.

In 1986, Arthgurs and Walsh [29] examined problem (a). Using quaternions and the Pontryagin maximum principle they showed that the point of contact between the ball and the $(x,y)$-plane satisfies the equations

$$ \begin{equation*} \begin{gathered} \, \dot x=\sin \psi, \quad \dot y=-\cos \psi, \\ \ddot \psi=\lambda \cos(\psi+\varepsilon), \qquad \lambda,\varepsilon \equiv \operatorname{const}. \end{gathered} \end{equation*} \notag $$
They observed that these equations are integrable by elliptic integrals of the first and third kind, and left the problem of optimal control to numerical investigation.

Independently of these papers, in 1993 Brockett and Dai [56] stated a ‘plate-ball problem’. They considered a ball which rolls without twisting or slipping between two horizontal plates which are placed at a distance equal to the diameter of the ball between them. They wrote out a control system of the form (2.46)(2.50) (see below) for the ball and showed that a nilpotent approximation to this system is equivalent to the control system (2.117) on the Cartan group (see § 2.10).

In the same year of 1993 Jurdjevic [86] considered thoroughly the problem of optimal rolling of a ball over a plane without twisting or slipping, where he relied on the statement due to Brockett and Dai [56] and was independent of [29] and [83]. Jurdjevic treated this problem as a left-invariant control problem on the Lie group $G=\mathbb{R}^2\times\operatorname{SO}(3)$:

$$ \begin{equation} \dot x=u_1, \qquad \dot y=u_2, \end{equation} \tag{2.46} $$
$$ \begin{equation} \dot R=R\begin{pmatrix} 0 & 0 &-u_1\\ 0 & 0 & -u_2\\ u_1 & u_2 & 0 \end{pmatrix}, \end{equation} \tag{2.47} $$
$$ \begin{equation} g=(x,y,R) \in G, \quad u=(u_1,u_2) \in \mathbb{R}^2, \end{equation} \tag{2.48} $$
$$ \begin{equation} g(0)=\operatorname{Id}=(0,0,E_{11}+E_{22}+E_{33}), \qquad g(t_1)=g_1 , \end{equation} \tag{2.49} $$
$$ \begin{equation} J=\frac{1}{2} \int_0^{t_1}(u_1^2+u_2^2)\,dt \to \min. \end{equation} \tag{2.50} $$
Next he used the Pontryagin maximum principle in the invariant form for Lie groups (see [6] and [88]) and obtained the following results. The optimal abnormal controls are constant in time, they result in rolling along a straight line; these controls are non-strictly abnormal. Normal extremals are trajectories of the Hamiltonian system with Hamiltonian
$$ \begin{equation*} H=\frac{1}{2}(h_1-H_2)^2+\frac{1}{2}(h_2+H_1)^2, \end{equation*} \notag $$
where the Hamiltonians $h_1$ and $h_2$ correspond to the vector fields $\partial/\partial x$ and $\partial/\partial y$, while $H_1$, $H_2$, and $H_3$ correspond to the left-invariant fields on $\operatorname{SO}(3)$ with generators
$$ \begin{equation*} A_1=E_{32}-E_{23}, \qquad A_2=E_{13}-E_{31}, \quad\text{and}\quad A_3=E_{21}-E_{12}, \end{equation*} \notag $$
defining rotations of three-dimensional space. The vertical subsystem of this Hamiltonian system is as follows:
$$ \begin{equation*} \begin{aligned} \, \dot h_1&=\dot h_2=0, \\ \dot H_1&=(h_1-H_2) H_3, \\ \dot H_2&=(h_2+H_1)H_3, \\ \dot H_3&=-h_1 H_1-h_2 H_2. \end{aligned} \end{equation*} \notag $$
This subsystem has the integrals $h_1$, $h_2$, $H$, and $M=H_1^2+H_2^2+H_3^2$, so it is integrable. In addition, it is reduced to a pendulum equation. To integrate the equations for the orientation $R(t) \in\operatorname{SO}(3)$ of the ball, the Euler angles $\varphi_1$, $\varphi_2$, and $\varphi_3$ are introduced; differential equations for these angles are obtained, qualitatively investigated, and integrated in part. It is shown that the trajectory of the point of contact $(x(t),y(t))$ between the ball and the plane is an Euler elastica (see § 2.6). A connection between the type of the intersection between the cylinder $\{H=\operatorname{const}\}$ and the sphere $\{M=\operatorname{const}\}$, the type of the elastica, and the qualitative behaviour of the Euler angles $\varphi_1$, $\varphi_2$, and $\varphi_3$ is discovered.

The rest of this section is based on [114] and [132].

2.8.2. The statement of the problem

A mechanical setting. Consider the mechanical system of two horizontal planes and a sphere tangent to these planes. The lower plane is fixed, and the sphere rolls over it without twisting or slipping, driven by the horizontal motion of the upper plane. The state of this system is described by the point of contact of the sphere with the lower plane and the orientation of the sphere in three-dimensional space. One must roll the sphere from a prescribed initial state to a prescribed terminal state so that the curve drawn by the point of contact in the plane has the minimum length. The velocity of the upper plane or, equivalently, the velocity of the centre of the sphere is the control.

We consider the kinematics of the system, so can ignore the presence of the upper plane and consider how the sphere rolls over the (lower) plane without slipping or twisting. ‘No slipping’ means that the instantaneous velocity of the point of contact of the sphere to the plane is zero, and ‘no twisting’ means that the vector of angular velocity of the sphere is horizontal. Rolling a surface over other surface without twisting or slipping models the work of a robotic arm, and problems concerning such motions are of great interest in mechanics, robotics, and control theory (see, for instance, [6], [48], [97], [100], and [108]).

A mathematical setting. Let $e_1$, $e_2$, $e_3$ is a fixed right-handed frame in $\mathbb{R}^3$ such that the vectors $e_1$ and $e_2$ lie in the plane $\mathbb{R}^2 \cong (\mathbb{R}^2,0) \subset \mathbb{R}^3$ on which a unit sphere $S^2$ is rolling, and $e_3$ points to the half-space containing this sphere. The frame $e_1$, $e_2$, $e_3$ is fixed at a point $O \in (\mathbb{R}^2,0)$. Let $f_1$, $f_2$, $f_3$ be a movable right-handed frame attached to the rolling sphere $S^2$. We denote the coordinates of a point in $\mathbb{R}^3$ with respect to the basis $e_1$, $e_2$, $e_3$ by $(x,y,z)$, and its coordinates with respect to the basis $f_1$, $f_2$, $f_3$ shifted to $O$ by $(X,Y,Z)$. Thus,

$$ \begin{equation*} x e_1+y e_2+z e_3=X f_1+Y f_2+Z f_3. \end{equation*} \notag $$
Let $R \in \operatorname{SO}(3)$ be the matrix taking the coordinates of a point with respect to the fixed frame $e_1$, $e_2$, $e_3$ to its coordinates with respect to the movable frame $f_1$, $f_2$, $f_3$, that is,
$$ \begin{equation*} \begin{pmatrix} X \\ Y \\ Z \end{pmatrix}= R \begin{pmatrix} x \\ y \\ z \end{pmatrix}. \end{equation*} \notag $$
The state of the system ‘the sphere $S^2$ and the plane $\mathbb{R}^2$’ is described by the coordinates $(x,y)$ of the point of contact between $S^2$ and $\mathbb{R}^2$ and the rotation matrix $R$. As control we take the vector $(u_1,u_2)$ of the velocity of the centre of the sphere. The problem of the optimal rolling of the sphere over the plane is formalized as the optimal control problem
$$ \begin{equation} \dot x=u_1, \end{equation} \tag{2.51} $$
$$ \begin{equation} \dot y=u_2, \end{equation} \tag{2.52} $$
$$ \begin{equation} \dot R=R(u_2 A_1-u_1 A_2), \end{equation} \tag{2.53} $$
$$ \begin{equation} Q=(x,y,R) \in G=\mathbb{R}^2 \times \operatorname{SO}(3), \end{equation} \tag{2.54} $$
$$ \begin{equation} u=(u_1, u_2) \in \mathbb{R}^2, \end{equation} \tag{2.55} $$
$$ \begin{equation} Q(0)=Q_0=(0, 0, \operatorname{Id}), \qquad Q(t_1)=Q_1, \end{equation} \tag{2.56} $$
$$ \begin{equation} l=\int_0^{t_1} \sqrt{u_1^2+u_2^2} \, dt \to \min. \end{equation} \tag{2.57} $$
Here and in what follows we use basis matrices of the Lie algebra $\mathfrak{so}(3)$, namely,
$$ \begin{equation} A_1=\begin{pmatrix} 0 & 0 & 0 \\ 0 & 0 & -1 \\ 0 & 1 & 0 \end{pmatrix},\quad A_2=\begin{pmatrix} 0 & 0 & 1 \\ 0 & 0 & 0 \\ -1 & 0 & 0 \end{pmatrix},\quad\text{and}\quad A_3=\begin{pmatrix} 0 & -1 & 0 \\ 1 & 0 & 0 \\ 0 & 0 & 0 \end{pmatrix}. \end{equation} \tag{2.58} $$

The left-invariant sub-Riemannian problem. Problem (2.51)(2.57) is a left- invariant sub-Riemannian problem on the Lie group $G=\mathbb{R}^2 \times \operatorname{SO}(3)$. We consider the following left-invariant frame on this group:

$$ \begin{equation*} e_1=\frac{\partial}{\partial x}\,, \quad e_2=\frac{\partial}{\partial y}\,, \quad V_i(R)=R A_i, \quad i=1,2,3. \end{equation*} \notag $$
In terms of the left-invariant fields
$$ \begin{equation*} X_1=e_1-V_2\quad\text{and} \quad X_2=e_2+V_1 \end{equation*} \notag $$
the control system (2.51)(2.55) assumes the following form:
$$ \begin{equation} \dot Q=u_1 X_1 (Q)+u_2 X_2(Q), \qquad Q \in G=\mathbb{R}^2 \times \operatorname{SO}(3), \quad (u_1, u_2) \in \mathbb{R}^2. \end{equation} \tag{2.59} $$
The functional (2.57) is the sub-Riemannian length functional for the left-invariant sub-Riemannian structure defined by the fields $X_1$ and $X_2$ as an orthonormal basis:
$$ \begin{equation} \begin{gathered} \, l=\int_0^{t_1} \langle \dot Q,\dot Q\rangle^{1/2} \, dt \to \min, \\ \langle X_i,X_j\rangle=\delta_{ij}, \qquad i,j=1, 2. \nonumber \end{gathered} \end{equation} \tag{2.60} $$

The existence of optimal controls. Recall that the matrix commutators $[A_i,A_j]=A_i A_j-A_j A_i$ are as follows:

$$ \begin{equation*} [A_1, A_2]=A_3, \qquad [A_2, A_3]=A_1, \qquad [A_3, A_1]=A_2. \end{equation*} \notag $$
The multiplication table in the Lie algebra $\mathfrak{g}=\mathbb{R}^2 \oplus \mathfrak{so}(3)= \operatorname{span}(e_1, e_2, V_1, V_2, V_3)$ of $G$ has the following form:
$$ \begin{equation*} \operatorname{ad}e_i=0, \qquad [V_1, V_2]=V_3, \qquad [V_2, V_3]=V_1, \qquad [V_3, V_1]=V_2. \end{equation*} \notag $$

By the equalities

$$ \begin{equation*} [X_1, X_2]=V_3, \qquad [X_1, V_3]=- V_1, \quad\text{and}\quad [X_2, V_3]=- V_2 \end{equation*} \notag $$
the fields $X_1$ and $X_2$ on the right-hand side of (2.59) generate a Lie algebra $\mathfrak{g}$. By the Rashevskii–Chow theorem system (2.59) is completely controllable. By Filippov’s theorem optimal controls in the class of essentially bounded measurable controls exist in problem (2.51)(2.57) for any $Q_0,Q_1 \in G$.

2.8.3. Extremals

Consider the Hamiltonians

$$ \begin{equation*} h_i(\lambda)=\langle \lambda, e_i \rangle, \qquad i=1,2, \end{equation*} \notag $$
and
$$ \begin{equation*} H_i(\lambda)=\langle \lambda, V_i \rangle, \qquad i=1,2,3, \end{equation*} \notag $$
which are linear on fibres of $T^*G$.

Abnormal trajectories. Constant-velocity abnormal trajectories have the form

$$ \begin{equation*} \begin{gathered} \, x_t=u_1 t, \qquad y_t=u_2 t, \\ R_t=\exp(t(u_2 A_1-u_1 A_2)). \end{gathered} \end{equation*} \notag $$
They are non-strictly abnormal. In the abnormal case the sphere rolls uniformly along a straight line.

The normal Hamiltonian system. In the normal case the Hamiltonian system with Hamiltonian

$$ \begin{equation*} H=\frac{1}{2}((h_1-H_2)^2+(h_2+H_1)^2) \end{equation*} \notag $$
is expressed in coordinates as follows:
$$ \begin{equation} \dot h_1 =\dot h_2=0, \end{equation} \tag{2.61} $$
$$ \begin{equation} \dot H_1 =(h_1-H_2)H_3, \end{equation} \tag{2.62} $$
$$ \begin{equation} \dot H_2 =(h_2+H_1) H_3, \end{equation} \tag{2.63} $$
$$ \begin{equation} \dot H_3 =- h_1 H_1-h_2 H_2, \end{equation} \tag{2.64} $$
$$ \begin{equation} \dot Q =(h_1-H_2) X_1+(h_2+H_1) X_2. \end{equation} \tag{2.65} $$
As always in sub-Riemannian problems, we can limit ourselves to geodesics of unit velocity, that is, to extremal trajectories such that $H \equiv 1/2$ on them. Under this condition, in the adjoint space it is convenient to go over from the variables $(h_1,h_2,H_1,H_2,H_3)$ to new variables $(r,\alpha,\theta,c)$:
$$ \begin{equation} h_1=r \cos \alpha, \qquad h_2=r \sin \alpha, \end{equation} \tag{2.66} $$
$$ \begin{equation} h_1-H_2=\cos(\theta+\alpha), \qquad h_2+H_1=\sin (\theta+\alpha), \end{equation} \tag{2.67} $$
$$ \begin{equation} c=H_3. \nonumber \end{equation} \notag $$
Then the Hamiltonian system for normal extremals (2.61)(2.65) takes the following form:
$$ \begin{equation} \dot \theta =c, \end{equation} \tag{2.68} $$
$$ \begin{equation} \dot c =-r \sin \theta, \end{equation} \tag{2.69} $$
$$ \begin{equation} \dot \alpha =\dot r=0, \end{equation} \tag{2.70} $$
$$ \begin{equation} \dot x =\cos(\theta+\alpha), \end{equation} \tag{2.71} $$
$$ \begin{equation} \dot y =\sin(\theta+\alpha), \end{equation} \tag{2.72} $$
$$ \begin{equation} \dot R =R \Omega, \qquad \Omega=\sin(\theta+\alpha) A_1-\cos(\theta+\alpha) A_2. \end{equation} \tag{2.73} $$
The family of normal extremals $\lambda_t$ is parametrized by the cylinder $C$ of initial points $\lambda=\lambda_t\big|_{t=0}$:
$$ \begin{equation*} \begin{aligned} \, C &= \biggl\{\lambda \in\mathfrak{g}^* \mid H(\lambda)=\frac{1}{2}\biggr\} \\ & \cong \{(h_1, h_2, H_1, H_2, H_3) \in \mathbb{R}^5 \mid (h_1-H_2)^2+(h_2+H_1)^2=1\} \\ & \cong \{(\theta, c, \alpha, r) \mid \theta \in S^1, \ c \in \mathbb{R}, \ \alpha \in S^1, \ r \geqslant 0\}. \end{aligned} \end{equation*} \notag $$
The exponential map is defined by
$$ \begin{equation*} \begin{gathered} \, \operatorname{Exp}(\lambda,t)=\pi \circ e^{t \vec H}(\lambda)=Q_t, \\ \operatorname{Exp}\colon N \to M, \qquad N=C \times \mathbb{R}_+=\{(\lambda,t) \mid \lambda \in C, \ t > 0\}. \end{gathered} \end{equation*} \notag $$

If $r=0$, then the elastica $(x_t,y_t)$ is a straight line (for $H_3=c=0$) or a circle (for $H_3=c \ne 0$); we say that such elasticae are degenerate.

If $r \ne 0$, then the elastica $(x_t,y_t)$ belongs to one of the four classes depending on the full energy $E=c^2/2-r\cos \theta$ of the pendulum (2.68), (2.69) (see § 2.6):

We say that elasticae in classes 1)–3) are non-degenerate.

The symplectic foliation. On the Lie coalgebra $\mathfrak{g}^*$ we have the Casimir functions $h_1$, $h_2$, and $M=H_1^2+H_2^2+H_3^2$. The symplectic foliations if formed by

The normal Hamiltonian system has the integrals $h_1$, $h_2$, $M$, and $E=(M+h_1^2+h_2^2)/2-H$; it is integrable by elliptic functions and integrals.

Geodesics of different types projecting onto Euler elasticae $(x_t,y_t)$ correspond to different types of intersection of a level surface of the Hamiltonian $\{H=\operatorname{const}\}$ with symplectic leaves.

Straightening coordinates. The cylinder $C=\{\lambda \in \mathfrak{g}^* \mid H(\lambda)=1/2\}$ is stratified in accordance with types of motion of the pendulum (2.68), (2.69):

$$ \begin{equation*} C=\bigsqcup_{i=1}^7 C_i, \end{equation*} \notag $$
where
$$ \begin{equation*} \begin{gathered} \, C_1=\{\lambda \in C \mid E \in (-r, r), \ r > 0\}, \\ C_2=\{\lambda \in C \mid E \in (r,+\infty), \ r > 0\}, \\ C_3=\{\lambda \in C \mid E=r > 0, \ c \ne 0\}, \\ C_4=\{\lambda \in C \mid E=-r, \ r > 0\}, \\ C_5=\{\lambda \in C \mid E=r > 0, \ c=0\}, \\ C_6=\{\lambda \in C \mid r=0, \ c \ne 0\}, \end{gathered} \end{equation*} \notag $$
and
$$ \begin{equation*} C_7=\{\lambda \in C \mid r=0, \ c=0\}. \end{equation*} \notag $$

In the domain $\bigcup_{i=1}^3 C_i$ we can introduce coordinates $(\varphi,k,\alpha,r)$ straightening the pendulum equations (2.68), (2.69).

If $\lambda=(\theta,c,\alpha,r)\in C_1$, then

$$ \begin{equation*} \sin\frac{\theta}{2}=k\operatorname{sn}(\sqrt{r}\,\varphi,k), \quad \cos\frac{\theta}{2}=\operatorname{dn}(\sqrt{r}\,\varphi,k), \quad\text{and}\quad \frac{c}{2}=k \sqrt{r}\,\operatorname{cn}(\sqrt{r}\,\varphi, k), \end{equation*} \notag $$
here $k=\sqrt{(E+r)/(2r)} \in (0, 1)$ and $\sqrt{r}\,\varphi\,\operatorname{mod}{4K} \in [0,4K]$.

If $\lambda=(\theta,c,\alpha,r)\in C_2$, then

$$ \begin{equation*} \sin\frac{\theta}{2}= \pm\operatorname{sn}\biggl(\frac{\sqrt{r}\,\varphi}{k}\,,k\biggr), \ \ \cos\frac{\theta}{2}= \operatorname{cn}\biggl(\frac{\sqrt{r}\,\varphi}{k}\,,k\biggr), \ \ \text{and}\ \ \frac{c}{2}=\pm \frac{\sqrt{r}}{k} \operatorname{dn}\biggl(\frac{\sqrt{r}\,\varphi}{k}\,,k\biggr), \end{equation*} \notag $$
where $ \pm=\operatorname{sign}c$; here $k=\sqrt{2r/(E+r)} \in (0,1)$ and $\sqrt{r}\,\varphi\,\operatorname{mod}{2kK} \in [0,2kK]$.

If $\lambda \in C_3$, then

$$ \begin{equation*} \sin\frac{\theta}{2}=\pm \tanh(\sqrt{r}\,\varphi), \quad \cos\frac{\theta}{2}=\frac{1}{\cosh(\sqrt{r}\,\varphi)}\,, \quad\text{and}\quad \frac{c}{2}=\pm \frac{\sqrt{r}}{\cosh(\sqrt{r} \varphi)}\,, \end{equation*} \notag $$
where $\pm=\operatorname{sign} c$; here $k=1$ and $\varphi \in (-\infty,+\infty)$.

In the new variables the pendulum equations (2.68), (2.69) take the form

$$ \begin{equation*} \dot \varphi=1, \quad \dot k=0, \quad \dot \alpha=0, \quad \dot r=0, \end{equation*} \notag $$
so that $\varphi_t=\varphi+t$ and $k,\alpha,r=\operatorname{const}$.

Integrating the vertical subsystem of the Pontryagin maximum principle. If $\lambda \in C_1$, then

$$ \begin{equation*} \sin\frac{\theta_t}{2}=k\operatorname{sn}(\sqrt{r}\,\varphi_t,k), \quad \cos\frac{\theta_t}{2}=\operatorname{dn}(\sqrt{r}\,\varphi_t,k), \quad\text{and}\quad \frac{c_t}{2}=k\sqrt{r}\,\operatorname{cn}(\sqrt{r}\,\varphi_t,k). \end{equation*} \notag $$
If $\lambda \in C_2$, then
$$ \begin{equation*} \sin\frac{\theta_t}{2}= \pm \operatorname{sn}\biggl(\frac{\sqrt{r}\,\varphi_t}{k}\,,k\biggr), \quad \cos\frac{\theta_t}{2}= \operatorname{cn}\biggl(\frac{\sqrt{r}\,\varphi_t}{k}\,,k\biggr), \quad \frac{c_t}{2}=\pm \frac{\sqrt{r}}{k}\operatorname{dn} \biggl(\frac{\sqrt{r}\,\varphi_t}{k}\,,k\biggr), \end{equation*} \notag $$
where $\pm=\operatorname{sign}c$.

If $\lambda \in C_3$, then

$$ \begin{equation*} \sin\frac{\theta_t}{2}= \pm \tanh(\sqrt{r}\,\varphi_t), \quad \cos\frac{\theta_t}{2}=\frac{1}{\cosh(\sqrt{r}\,\varphi_t)}\,, \quad \frac{c_t}{2}=\pm \frac{\sqrt{r}}{\cosh(\sqrt{r}\,\varphi_t)}\,, \end{equation*} \notag $$
where $\pm=\operatorname{sign}c$.

In the case when $\lambda \in \bigcup_{i=4}^7 C_i$ system (2.68)(2.70) is integrated directly:

$$ \begin{equation*} \begin{gathered} \, \theta_t \equiv 0,\quad c_t \equiv 0 \quad\text{for } \lambda \in C_4;\qquad \theta_t \equiv \pi,\quad c_t \equiv 0\quad\text{for } \lambda \in C_5; \\ \theta_t=c t+\theta,\quad c_t \equiv c \ne 0\quad\text{for } \lambda \in C_6;\qquad \theta_t \equiv \theta,\quad c_t \equiv 0\quad\text{for }\lambda \in C_7. \end{gathered} \end{equation*} \notag $$

Integrating the equations for $x$ and $y$. To integrate equations (2.71), (2.72) with initial condition $x_0=y_0=0$ we use the symmetry of the problem and perform the rotation

$$ \begin{equation*} x=\bar x\cos \alpha-\bar y \sin \alpha, \quad y=\bar x \sin \alpha+\bar y \cos \alpha. \end{equation*} \notag $$
In the new variables we obtain the Cauchy problem
$$ \begin{equation} \dot{\bar x}_t=\cos \theta_t, \quad \dot{\bar y}_t=\sin \theta_t, \qquad \bar x_0= \bar y_0=0, \end{equation} \tag{2.74} $$
whose solutions can be parametrized as follows.

If $\lambda \in C_1$, then

$$ \begin{equation*} \bar{x}_t=\frac{2(\operatorname{E}(\sqrt{r}\,\varphi_t)- \operatorname{E}(\sqrt{r}\,\varphi))-\sqrt{r}\,t}{\sqrt{r}} \end{equation*} \notag $$
and
$$ \begin{equation*} \bar{y}_t=\frac{2k(\operatorname{cn}(\sqrt{r}\,\varphi)- \operatorname{cn}(\sqrt{r}\,\varphi_t))}{\sqrt{r}}\,. \end{equation*} \notag $$

If $\lambda \in C_2$, then

$$ \begin{equation*} \bar{x}_t=\frac{2(\operatorname{E}(\sqrt{r}\,\varphi_t/k)- \operatorname{E}(\sqrt{r}\,\varphi/k)-(2-k^2)\sqrt{r}\,t/(2k))}{k\sqrt{r}} \end{equation*} \notag $$
and
$$ \begin{equation*} \bar{y}_t=\pm\frac{2(\operatorname{dn}(\sqrt{r}\,\varphi/k)- \operatorname{dn}(\sqrt{r}\,\varphi_t/k))}{k\sqrt{r}}\,, \qquad \pm=\operatorname{sign}c. \end{equation*} \notag $$

If $\lambda \in C_3$, then

$$ \begin{equation*} \bar{x}_t=\frac{2(\tanh(\sqrt{r}\,\varphi_t)-\tanh(\sqrt{r}\,\varphi))- \sqrt{r}\,t}{\sqrt{r}} \end{equation*} \notag $$
and
$$ \begin{equation*} \bar{y}_t=\pm \frac{2(1/\cosh(\sqrt{r}\,\varphi)- 1/\cosh(\sqrt{r}\,\varphi_t))}{\sqrt{r}}\,, \qquad \pm=\operatorname{sign}c. \end{equation*} \notag $$

For $\lambda \in \bigcup_{i=4}^7 C_i$ equations (2.74) are integrated directly:

$$ \begin{equation*} \begin{gathered} \, \bar{x}_t=t,\quad \bar{y}_t=0\quad\text{for } \lambda \in C_4;\qquad \bar{x}_t=-t,\quad \bar{y}_t=0\quad\text{for } \lambda \in C_5; \\ \text{and}\quad \bar{x}_t=\frac{\sin(c t+\theta)-\sin \theta}{c}\,,\quad \bar{y}_t=\frac{\cos \theta-\cos(c t+\theta)}{c}\quad\text{for } \lambda \in C_6; \\ \bar{x}_t=t \cos \theta,\quad \bar{y}_t=t \sin \theta\quad\text{for } \lambda \in C_7. \end{gathered} \end{equation*} \notag $$

Integrating the equations for $R$. Assume that $M=H_1^2+H_2^2+H_3^2 > 0$. Then

$$ \begin{equation} \begin{aligned} \, \nonumber R(t)&=\exp\bigl((\alpha-\varphi_3^0) A_3\bigr) \exp\bigl(-\varphi_2^0 A_2\bigr)\exp\bigl(\varphi_1(t) A_3\bigr) \\ &\qquad\times\exp\bigl(\varphi_2(t) A_2\bigr) \exp\bigl((\varphi_3(t)-\alpha)A_3\bigr), \end{aligned} \end{equation} \tag{2.75} $$
where the angles $\varphi_i$ are defined from relations (2.76)(2.82) for $r \ne 1$ and from (2.79)(2.83) for $r=1$ (see below), and the angle $\varphi_1$ satisfies the initial condition $\varphi_1^0=0$.

The exponentials of matrices containing $\varphi_2$ and $\varphi_3$ and involved in (2.75) are expressed in terms of $\cos \varphi_2$, $\sin \varphi_2$, $\cos \varphi_3$, and $\sin \varphi_3$. These latter are expressed in terms of the variables $c$, $\cos(\theta/2)$, and $\sin(\theta/2)$ with the help of (2.76), (2.77), (2.79), (2.80). These variables were represented above as functions of elliptic coordinates or can be used directly. For $r=1$ we have $\varphi_1(t)=\sqrt M\,t/2$. We defer the integration of equation (2.78) for $r \ne 1$ to the next subsection.

In the case $M=0$ we have $r=1$, $c=0$, and $\theta=0$, so that $u_1=\cos\alpha$ and $u_2=\sin \alpha$. Therefore, $\Omega=u_2 A_1-u_1 A_2 \equiv \operatorname{const}$ and $R(t)=e^{t\Omega}$.

Integrating the equations for $\varphi_1$. Along normal geodesics, for $r \ne 1$ the angles $\varphi_i$ satisfy the equalities

$$ \begin{equation} \cos \varphi_2 =\frac{c}{\sqrt{M}}\,, \qquad \sin \varphi_2 =\pm \frac{\sqrt{M-c^2}}{\sqrt{M}}\,, \end{equation} \tag{2.76} $$
$$ \begin{equation} \cos \varphi_3 =\mp\frac{\sin \theta}{\sqrt{M-c^2}}\,, \qquad \sin \varphi_3 =\pm\frac{r-\cos \theta}{\sqrt{M-c^2}}\,, \end{equation} \tag{2.77} $$
$$ \begin{equation} \dot\varphi_1=\frac{\sqrt{M}(1-r \cos \theta)}{M-c^2}\,, \end{equation} \tag{2.78} $$
while for $r=1$ they satisfy
$$ \begin{equation} \cos \varphi_2 =\frac{c}{\sqrt{M}}\,, \qquad \sin \varphi_2 =\pm \frac{2 \sin(\theta/2)}{\sqrt{M}}\,, \end{equation} \tag{2.79} $$
$$ \begin{equation} \cos \varphi_3 =\mp \cos\frac{\theta}{2}\,, \qquad \sin \varphi_3 =\pm \sin\frac{\theta}{2}\,, \end{equation} \tag{2.80} $$
$$ \begin{equation} \dot\varphi_1=\frac{\sqrt{M}}{2}\,. \end{equation} \tag{2.81} $$

We introduce an elliptic integral of the third kind in the following form:

$$ \begin{equation} \Pi(n, u, k)=\int_0^u \frac{dt}{(1-n \sin^2 t)\sqrt{1-k^2 \sin^2 t}}=\int_0^{F(u,k)} \frac{dv}{1-n \operatorname{sn}^2 v}\,. \end{equation} \tag{2.82} $$

Let $r \ne 1$. If $\lambda_1 \in C_1$, then

$$ \begin{equation} \varphi_1(t)=\frac{\sqrt{M}}{2}\,t+\frac{\sqrt{M}\,(1+r)}{2\,\sqrt{r}\,(1-r)} \bigl[\Pi(l,\operatorname{am}(\sqrt{r}\,(\varphi+t)),k)- \Pi(l,\operatorname{am}(\sqrt{r}\,\varphi),k)\bigr], \end{equation} \tag{2.83} $$
where $l=-4 k^2 r/(1-r)^2$.

If $\lambda_1 \in C_2$, then

$$ \begin{equation*} \varphi_1(t)=\frac{\sqrt{M}}{2}\,t+\frac{\sqrt{M}\,k(1+r)}{2\,\sqrt{r}\,(1-r)}\, \biggl[\Pi\biggl(l,\operatorname{am} \biggl(\frac{\sqrt{r}\,(\varphi+t)}{k}\biggr),k\biggr)- \Pi\biggl(l,\operatorname{am} \biggl(\frac{\sqrt{r}\,\varphi}{k}\biggr),k\biggr)\biggr], \end{equation*} \notag $$
where $l=-4r/(1-r)^2$.

If $\lambda_1 \in C_3$, then

$$ \begin{equation*} \varphi_1(t)=\frac{\sqrt{M}}{2}\,t+\frac{\sqrt{M}\,k(1-r^2)}{8 r^{3/2}} \bigl[I(\sqrt{r}\,(\varphi+t),a)-I(\sqrt{r}\,\varphi,a)\bigr], \end{equation*} \notag $$
and
$$ \begin{equation*} I(v,a)=\int_0^v\frac{dt}{a^2+\tanh^2t}= \frac{a t-\arctan a+\arctan (e^t(a^2\cosh t+\sinh t)/a)}{a+a^3}\,, \end{equation*} \notag $$
where $ a=(1-r)/(2\,\sqrt{r}\,)$.

If $\lambda_1 \in C_6$, then $\varphi_1(t)=\sqrt{1+c^2}\, t$.

If $\lambda \in C_4 \cup C_5 \cup C_7$, then

$$ \begin{equation*} \theta_t \equiv \operatorname{const}=\theta,\quad \Omega=\sin(\alpha+\theta)A_1-\cos(\alpha+\theta) A_2 \equiv \operatorname{const},\quad\text{and}\quad R(t)=e^{t\Omega}. \end{equation*} \notag $$

The control system in terms of quaternions. To describe the orientation of the rolling sphere it is convenient to use quaternions along with the rotation matrix $R$.

Let

$$ \begin{equation*} \mathbb{H}=\{q=q_0+iq_1+j q_2+k q_3 \mid q_0,\dots,q_3 \in \mathbb{R}\} \end{equation*} \notag $$
be the algebra of quaternions, let $S^3=\{q \in \mathbb{H} \mid |q|^2=q_0^2+q_1^2+q_2^2+q_3^2=1\}$ be the unit sphere, and
$$ \begin{equation*} I=\{q \in \mathbb{H} \mid \operatorname{Re}q=q_0=0\} \end{equation*} \notag $$
be the subspace of purely imaginary quaternions. A quaternion $q \in S^3$ defines a rotation of the Euclidean space $I$:
$$ \begin{equation*} q \in S^3 \quad\Longrightarrow \quad R_q(a)=q a q^{-1}, \quad a \in I, \quad R_q \in \operatorname{SO}(3)\cong \operatorname{SO}(I). \end{equation*} \notag $$

The correspondence between the quaternions $q=q_0+i q_1+j q_2+k q_3 \in S^3$ and the matrices $R \in \operatorname{SO}(3)$ looks as follows:

$$ \begin{equation} R=\begin{pmatrix} q_0^2+q_1^2-q_2^2-q_3^2 & 2 q_1 q_2-2 q_0 q_3& 2 q_0 q_2 + 2 q_1 q_3 \\ 2 q_1 q_2+2 q_0 q_3 &q_0^2-q_1^2+q_2^2-q_3^2 & -2 q_0q_1+2 q_2 q_3 \\ -2 q_0 q_2+2 q_1 q_3 & 2 q_0 q_1+2 q_2 q_3 & q_0^2-q_1^2 - q_2^2+q_3^2 \end{pmatrix}. \end{equation} \tag{2.84} $$

In terms of quaternions the control system (2.47) takes the following form:

$$ \begin{equation} \begin{cases} \dot{q}_0=(q_2 u_1-q_1 u_2)/2, \\ \dot{q}_1=(q_3 u_1+q_0 u_2)/2, \\ \dot{q}_2=(-q_0 u_1+q_3 u_2)/2, \\ \dot{q}_3=(-q_1 u_1-q_2 u_2)/2, \end{cases}\qquad q=q_0+i q_1+j q_2+k q_3 \in S^3, \quad (u_1,u_2) \in \mathbb{R}^2. \end{equation} \tag{2.85} $$

The control system on $\mathbb{R}^2\times\operatorname{SO}(3)$ (2.46), (2.47) with initial condition $g(0)=\operatorname{Id}$ lifts to $\mathbb{R}^2\times S^3$ as a system (2.46), (2.85) with initial conditions $(x,y)(0)=(0,0)$ and $q(0)=1$.

2.8.4. Symmetries

Symmetries of the family of extremal trajectories. The rotations of elasticae $(x_s,y_s)$ around the origin in the $(x,y)$-plane generate a one- parameter group of symmetries of trajectories of the Hamiltonian system (2.68)(2.73)

$$ \begin{equation*} \{\Phi^{\beta} \mid \beta \in S^1 \}, \end{equation*} \notag $$
where the rotation $\Phi^{\beta}$ is defined as follows:
$$ \begin{equation} \Phi^{\beta}\colon\{\lambda_s \mid s \in [0,t]\}\to \{\lambda_s^{\beta} \mid s \in [0,t]\}, \end{equation} \tag{2.86} $$
where
$$ \begin{equation} \lambda_s=(\theta_s, c_s, \alpha, r, Q_s), \qquad Q_s=(x_s, y_s, R_s), \end{equation} \tag{2.87} $$
$$ \begin{equation} \lambda_s^{\beta}= (\theta_s^{\beta},c_s^{\beta},\alpha^{\beta},r,Q_s^{\beta}),\qquad Q_s^{\beta}=(x_s^{\beta},y_s^{\beta},R_s^{\beta}), \end{equation} \tag{2.88} $$
$$ \begin{equation} \theta_s^{\beta}=\theta_s, \qquad c_s^{\beta}=c_s, \qquad \alpha^{\beta}=\alpha+\beta, \end{equation} \tag{2.89} $$
$$ \begin{equation} \begin{pmatrix} x_s^{\beta} \\ y_s^{\beta}\end{pmatrix}= \begin{pmatrix} \cos \beta &-\sin \beta \\ \sin \beta & \cos \beta \end{pmatrix} \begin{pmatrix} x_s \\ y_s\end{pmatrix}, \end{equation} \tag{2.90} $$
$$ \begin{equation} R_s^{\beta}=e^{\beta A_3} R_s e^{-\beta A_3}. \end{equation} \tag{2.91} $$

Proposition 2.8. If $\{\lambda_s \mid s \in [0,t]\}$ is a trajectory of system (2.68)(2.73), then for each $\beta \in S^1$ the curve $\{\lambda_s^{\beta} \mid s \in [0,t]\}$ is a trajectory of this system too.

The reflections of trajectories $(\theta_s,c_s)$ of the pendulum (2.68), (2.69) in the $\theta$- and $c$-coordinate axes and in the origin extend to discrete symmetries $\varepsilon^1$, $\varepsilon^2$, and $\varepsilon^3$ of the family of trajectories of system (2.68)(2.73):

$$ \begin{equation*} \begin{gathered} \, \varepsilon^i\colon\{\lambda_s \mid s \in [0,t]\}\to \{\lambda_s^{i} \mid s \in [0,t]\}, \qquad i=1,2,3, \\ \begin{alignedat}{2} \lambda_s&=(\theta_s, c_s, \alpha, r, Q_s), &\qquad Q_s&=(x_s, y_s, R_s), \\ \lambda_s^{i}&=(\theta_s^{i}, c_s^{i}, \alpha^{i}, r, Q_s^{i}), &\qquad Q_s^{i}&=(x_s^{i}, y_s^{i}, R_s^{i}). \end{alignedat} \end{gathered} \end{equation*} \notag $$

The reflection of trajectories $(\theta_s,c_s)$ of the pendulum (2.68), (2.69) in the $\theta$-axis corresponds to a discrete symmetry $\varepsilon^1$ of the family of extremal trajectories:

$$ \begin{equation*} \begin{gathered} \, \theta_s^{1}=\theta_{t-s}, \qquad c_s^{1}=-c_{t-s}, \qquad \alpha^{1}=\alpha+\pi, \\ x_s^1=x_{t-s}-x_t, \qquad y_s^1=y_{t-s}-y_t, \\ R_s^{1}=(R_t)^{-1} R_{t-s}, \qquad \Omega_s^{1}=-\Omega_{t-s}. \end{gathered} \end{equation*} \notag $$

The reflection of trajectories $\theta_s,c_s)$ of the pendulum in the $c$-axis generates a symmetry $\varepsilon^2$ of extremal trajectories:

$$ \begin{equation*} \begin{gathered} \, \theta_s^{2}=-\theta_{t-s}, \qquad c_s^{2}=c_{t-s}, \qquad \alpha^{2}=\pi-\alpha, \\ x_s^2=x_{t-s}-x_t, \qquad y_s^2=y_t-y_{t-s}, \\ R_s^{2}=I_2 (R_t)^{-1} R_{t-s} I_2, \qquad \Omega_s^{2}=- I_2 \Omega_{t-s} I_2, \\ I_2=I_2^{-1}=e^{\pi A_2}=\begin{pmatrix} - 1 & 0 & \hphantom{-}0 \\ \hphantom{-}0 & 1 & \hphantom{-}0 \\ \hphantom{-}0 & 0 & -1 \end{pmatrix}. \end{gathered} \end{equation*} \notag $$

The reflection of trajectories $(\theta_s,c_s)$ of the pendulum in the origin $(\theta,c)=(0,0)$ extends to a symmetry $\varepsilon^3$ of extremal trajectories:

$$ \begin{equation*} \begin{gathered} \, \theta_s^{3}=-\theta_{s}, \qquad c_s^{3}=-c_{s}, \qquad \alpha^{3}=- \alpha, \\ x_s^3=x_{s}, \qquad y_s^3=- y_{s}, \\ R_s^{3}=I_2 R_{s} I_2, \qquad \Omega_s^{3}=I_2 \Omega_{s} I_2. \end{gathered} \end{equation*} \notag $$

Proposition 2.9. If $\{\lambda_s \mid s \in [0,t]\}$ is a trajectory of system (2.68)(2.73), then the curves $\{\lambda_s^{i} \mid s \in [0,t]\}$, $i=1,2,3$, are also trajectories of these system.

Symmetries of the exponential map. The actions of rotations $\Phi^\beta$ and reflections $\varepsilon^i$ on the source space and target space of the exponential map are defined so as to commute with the action of the exponential map.

Rotations $\Phi^\beta\colon\lambda\to\lambda^{\beta}$ (2.86)(2.91) are symmetries of the Hamiltonian system, so their actions on $T^*G$ decompose in a natural way into a direct sum of actions on $N=\mathfrak{g}^* \times \mathbb{R}_+$ (on points $(\lambda,t)$, where $\lambda$ is the initial point of the extremal) and on $G$ (on the terminal points $Q_t$ of the corresponding extremal trajectories):

$$ \begin{equation*} \begin{gathered} \, \Phi^\beta\colon N \to N, \qquad (\lambda,t) \mapsto (\lambda^{\beta},t), \\ \lambda=(\theta,c,\alpha,r), \qquad \lambda^{\beta}=(\theta,c,\alpha^{\beta},r),\qquad \alpha^{\beta}=\alpha+\beta, \end{gathered} \end{equation*} \notag $$
and
$$ \begin{equation*} \begin{gathered} \, \Phi^\beta\colon G\to G, \qquad Q \mapsto Q^{\beta}, \\ Q=(x, y, R), \qquad Q^{\beta}=(x^{\beta}, y^{\beta}, R^{\beta}), \\ \begin{pmatrix} x^{\beta} \\ y^{\beta}\end{pmatrix}=\begin{pmatrix} \cos \beta &-\sin \beta \\ \sin \beta & \hphantom{-}\cos \beta \end{pmatrix} \begin{pmatrix} x \\ y\end{pmatrix},\qquad R^{\beta}=e^{\beta A_3} R e^{-\beta A_3}. \end{gathered} \end{equation*} \notag $$

The action of the reflections $\varepsilon^i$ on $N$ is defined by their restrictions to the vertical components of extremal trajectories at the initial moment of time $s=0$:

$$ \begin{equation*} \begin{gathered} \, \varepsilon^i\colon N\to N, \quad (\lambda, t) \mapsto (\lambda^{i}, t), \qquad i=1, 2, 3, \\ \lambda=(\theta, c, \alpha, r), \qquad \lambda^{i}=(\theta^i, c^i, \alpha^{i}, r), \end{gathered} \end{equation*} \notag $$
where $\lambda=\lambda_s\big|_{s=0}$ and $\lambda^i=\lambda_s^i\big|_{s=0}$. The explicit expressions for the actions of the $\varepsilon^i$ on $N$ are as follows:
$$ \begin{equation*} \begin{aligned} \, \varepsilon^1\colon(\theta,c,\alpha,r,t)\to(\theta^1,c^1,\alpha^1,r,t)&= (\theta_t,-c_t,\alpha+\pi,r,t), \\ \varepsilon^2\colon(\theta,c,\alpha,r,t)\to(\theta^2,c^2,\alpha^2,r,t)&= (-\theta_t, c_t, \pi-\alpha, r, t), \\ \varepsilon^3\colon(\theta,c,\alpha,r,t)\to(\theta^3,c^3,\alpha^3,r,t)&= (-\theta, -c, -\alpha, r, t). \end{aligned} \end{equation*} \notag $$

The action of reflections on $G$ is determined by their action on the extremal trajectories at the terminal moment of time $s=t$:

$$ \begin{equation*} \begin{gathered} \, \varepsilon^i\colon G\to G, \quad Q \mapsto Q^{i}, \qquad i=1,2,3, \\ Q=(x,y,R), \qquad Q^i=(x^i,y^i,R^i), \end{gathered} \end{equation*} \notag $$
where $Q=Q_s\big|_{s=t}$ and $Q^i=Q_s^i\big|_{s=t}$. The explicit formulae are
$$ \begin{equation*} \begin{aligned} \, \varepsilon^1\colon(x, y, R)\to(x^1, y^1, R^1)&=(-x, -y, (R)^{-1}), \\ \varepsilon^2\colon(x, y, R)\to(x^2, y^2, R^2)&=(-x, y, I_2 (R)^{-1}I_2), \\ \varepsilon^3\colon(x, y, R)\to(x^3, y^3, R^3)&=(x, -y, I_2 R I_2). \end{aligned} \end{equation*} \notag $$

Thus we have defined the actions of rotations and reflections on the target space of the exponential map:

$$ \begin{equation} \Phi^\beta,\varepsilon^i \colon N \to N, \qquad (\lambda,t) \mapsto (\lambda^{\beta},t),(\lambda^i,t), \end{equation} \tag{2.92} $$
$$ \begin{equation} \Phi^\beta,\varepsilon^i \colon G \to G, \qquad Q \mapsto Q^{\beta},Q^i. \end{equation} \tag{2.93} $$
It is essential that the image $Q^i=\varepsilon^i(Q)$ depends only on the preimage $Q$, rather than on the moment of time $t$.

Proposition 2.10. The maps $\Phi^\beta$ and $\varepsilon^i$ are symmetries of the exponential map.

Consider the group of symmetries of the exponential map generated by the rotations and reflections:

$$ \begin{equation*} \operatorname{Sym}= \langle \Phi^\beta,\varepsilon^1,\varepsilon^2,\varepsilon^3 \rangle. \end{equation*} \notag $$
The multiplication table in this group is as follows:

$\cdot \, \circ \, \cdot$$\varepsilon^1$$\varepsilon^2$$\varepsilon^3$$\Phi^\beta$
$\varepsilon^1$$\operatorname{Id}$$\varepsilon^3$$\varepsilon^2$$\Phi^\beta \circ \varepsilon^1$
$\varepsilon^2$$\varepsilon^3$$\operatorname{Id}$$\varepsilon^1$$\Phi^{-\beta} \circ \varepsilon^2$
$\varepsilon^3$$\varepsilon^2$$\varepsilon^1$$\operatorname{Id}$$\Phi^{-\beta} \circ \varepsilon^3$
$\Phi^{\gamma}$$\varepsilon^1 \circ \Phi^{\gamma}$$\varepsilon^2 \circ \Phi^{-\gamma}$$\varepsilon^3 \circ \Phi^{-\gamma}$$\Phi^{\beta+\gamma}$

Hence we obtain an explicit description of the symmetry group of the exponential map:

$$ \begin{equation*} \operatorname{Sym}=\{\Phi^\beta,\Phi^\beta \circ \varepsilon^i \mid \beta \in S^1, \ i=1,2,3\} \cong \operatorname{SO}(2)\times (\mathbb{Z}_2\times\mathbb{Z}_2). \end{equation*} \notag $$

The Maxwell set corresponding to the group $\langle \varepsilon^i,\Phi^\beta\rangle$, $i=1,2,3$, is defined by

$$ \begin{equation*} \begin{aligned} \, \operatorname{MAX}^i=\bigl\{(\lambda,t)\in N\mid \exists\beta \in S^1&\colon (\tilde\lambda,t)=\varepsilon^i \circ \Phi^\beta(\lambda,t), \\ &\quad\operatorname{Exp}(\lambda, s) \not\equiv \operatorname{Exp}(\tilde\lambda, s), \ \operatorname{Exp}(\lambda,t)= \operatorname{Exp}(\tilde \lambda,t)\bigr\}. \end{aligned} \end{equation*} \notag $$

2.8.5. Conditions for optimality

Theorem 2.36. Let $t > 0$, and let $Q_s=(x_s, y_s, R_s)=\operatorname{Exp}(\lambda,s)$ be an extremal trajectory such that

Then $(\lambda,t) \in \operatorname{MAX}^1$, so that for no $t_1 > t$ is the trajectory $Q_s$, $s \in [0,t_1]$, optimal.

Theorem 2.37. Let $t > 0$, and let $Q_s=(x_s,y_s,R_s)=\operatorname{Exp}(\lambda,s)$ be an extremal trajectory such that

Then $(\lambda,t) \in \operatorname{MAX}^2$, so that for no $t_1 > t$ is the trajectory $Q_s$, $s \in [0, t_1]$, optimal.

Theorem 2.38. Let $t > 0$, and let $Q_s=(x_s, y_s, R_s)=\operatorname{Exp}(\lambda,s)$ be an optimal trajectory such that

Then $(\lambda,t) \in \operatorname{MAX}^3$, so that for no $t_1 > t$ is the trajectory $Q_s$, $s \in [0,t_1]$, optimal.

Remark 2.2. Bearing in mind that for each quaternion $q=q_0+iq_1+jq_2+kq_3 \in S^3$ the corresponding motion $R_q\colon \mathbb{R}^3 \to \mathbb{R}^3$ is a rotation about the vector $(q_1,q_2,q_3) \in\mathbb{R}^3$, we can give the following geometric interpretation of conditions (1) in Theorems 2.362.38.

1. Condition (1) in Theorem 2.36 means that the rotation $R_t$ of the sphere is a rotation about some horizontal axis.

2. Condition (1) in Theorem 2.37 means that $R_t$ is a rotation about some axis orthogonal to the displacement vector of the point of contact of the sphere and the plane $(x_t,y_t,0)$.

3. Condition (1) in Theorem 2.38 means that $R_t$ is a rotation about the horizontal axis orthogonal to the vector $(x_t,y_t,0)$, or the rotation through $\pi$ about an axis lying in the vertical plane containing $(x_t,y_t,0)$.

2.8.6. Bibliographic comments

Subsection 2.8.1 is based on [29], [56], [83], and [86], § 2.8.2 is based on [132], § 2.8.3 on [114] and [132], and § 2.8.4 on [132].

The problem of rolling a ball over a plane without twisting or slipping was also considered in [2] and [88].

2.9. The sub-Riemannian problem on the Engel group

2.9.1. The problem statement

The geometric setting. Fix two points $a_0,a_1 \in \mathbb{R}^2$ on the plane which are connected by a curve $\gamma_0 \subset \mathbb{R}^2$. Also fix $S \in \mathbb{R}$ and a line $L \subset \mathbb{R}^2$. The problem consists in joining $a_0$ with $a_1$ by a shortest curve $\gamma \subset \mathbb{R}^2$ such that $\gamma_0$ and $\gamma$ bound a plane domain with algebraic area $S$ whose centre of mass lies on $L$. Thus this is a generalization (a sophistication) of the Dido problem (see [143] and [157]).

The optimal control problem. The above geometric problem can also be formulated as an optimal control problem:

$$ \begin{equation} \dot{g}=u_1 X_1(g)+u_2 X_2(g), \quad g=(x, y, z, v) \in \mathbb{R}^4, \end{equation} \tag{2.94} $$
$$ \begin{equation} g(0)=g_0, \quad g(t_1)=g_1, \end{equation} \tag{2.95} $$
$$ \begin{equation} l=\int_0^{t_1}\sqrt{u_1^2+u_2^2}\,\,dt \to \min, \end{equation} \tag{2.96} $$
$$ \begin{equation} X_1=\frac{\partial}{\partial x}- \frac{y}{2}\,\frac{\partial}{\partial z}\,, \quad X_2=\frac{\partial}{\partial y}+ \frac{x}{2}\,\frac{\partial}{\partial z}+ \frac{x^2+y^2}{2}\,\frac{\partial}{\partial v}\,. \end{equation} \tag{2.97} $$
This is a sub-Riemannian problem for the sub-Riemannian structure on $\mathbb{R}^4$ defined by the vector fields $X_1$ and $X_2$ treated as an orthonormal frame.

The Engel algebra and Engel group. The Engel algebra is the Lie algebra $\mathfrak{g}$, with basis $(X_1,\dots,X_4)$ such that the non-trivial commutators of its elements are

$$ \begin{equation*} [X_1, X_2]=X_3\quad\text{and} \quad [X_1, X_3]=X_4 \end{equation*} \notag $$
(see Fig. 34).

The Engel algebra is a nilpotent Lie algebra with grading:

$$ \begin{equation*} \mathfrak{g}=\mathfrak{g}^{(1)} \oplus\mathfrak{g}^{(2)} \oplus \mathfrak{g}^{(3)}, \end{equation*} \notag $$
where
$$ \begin{equation*} \mathfrak{g}^{(1)}=\operatorname{span}(X_1,X_2),\quad \mathfrak{g}^{(2)}=\mathbb{R}X_3,\quad \mathfrak{g}^{(3)}=\mathbb{R} X_4,\qquad [\mathfrak{g}^{(1)},\mathfrak{g}^{(i)}]=\mathfrak{g}^{(i+1)}, \end{equation*} \notag $$
and $\mathfrak{g}^{(4)}=\{0\}$, so it is a Carnot algebra. The corresponding connected and simply connected Lie group $G$ is called the Engel group.

The Engel group has a linear representation as

$$ \begin{equation*} \left\{\begin{pmatrix} 1 & b & c & d \\ 0 & 1 & a & a^2/2\\ 0 & 0 & 1 & a \\ 0 & 0 & 0 & 1 \end{pmatrix}\,\Bigg|\, a,b,c,d \in \mathbb{R}\right\}. \end{equation*} \notag $$
In $\mathbb{R}^4_{x,y,z,v}$ we introduce multiplication by
$$ \begin{equation*} \begin{aligned} \, \begin{pmatrix} x_1\\ y_1\\ z_1\\ v_1 \end{pmatrix} \cdot \begin{pmatrix} x_2\\ y_2\\ z_2\\ v_2 \end{pmatrix}=\begin{pmatrix} x_1+x_2\\ y_1+y_2\\ z_1+z_2+\dfrac{x_1 y_2-x_2 y_1}{2}\\ v_1+v_2+\dfrac{y_1y_2(y_1+y_2)}{2}+x_1z_2+\dfrac{x_1y_2(x_1+x_2)}{2} \end{pmatrix}, \end{aligned} \end{equation*} \notag $$
which makes of it the Engel group: $G \cong \mathbb{R}^4_{x,y,z,v}$ so that the fields (2.97) become left-invariant fields on this group. Thus, (2.94)(2.96) is a left-invariant sub- Riemannian problem on the Engel group. Hence we can assume that the reference point in (2.95) is the identity element of the Engel group: $g_0=\operatorname{Id}=(0,0,0,0)$.

All completely non-holonomic left-invariant sub-Riemannian problems of rank 2 on the Engel group are transformed into one another by isomorphisms of this group [124].

Special features of the problem. The sub-Riemannian problem on the Engel group is a simplest left-invariant sub-Riemannian problem, which has the following properties:

This problem presents a nilpotent approximation to each sub-Riemannian problem of Engel type (that is, with growth vector $(2,3,4)$: see [45] and [143]) and, in particular, to a mobile robot with trailer.

2.9.2. Symmetries of the distribution and sub-Riemannian structure

Theorem 2.39. The Lie algebra of infinitesimal symmetries of the distribution $\operatorname{span}(X_1,X_2)$ on the Engel group is parametrized by the smooth functions on this group that are constant along the field $X_2$.

Theorem 2.40. The Lie algebra of infinitesimal symmetries of the nilpotent sub- Riemannian structure on the Engel group is isomorphic to the Engel algebra of right-invariant vector field on this group.

2.9.3. Geodesics

Optimal trajectories exist in problem (2.94)(2.96) as follows from the Rashevskii–Chow and Filippov theorems.

The Pontryagin maximum principle. From the problem of the minimization of length (2.96) we go over to the equivalent problem of the minimization of energy

$$ \begin{equation} J=\frac{1}{2}\int_0^{t_1} (u_1^2+u_2^2)\,dt \to \min\!. \end{equation} \tag{2.98} $$

Consider the Hamiltonians $h_i(\lambda)=\langle\lambda,X_i\rangle$, $i=1,\dots,4$, which are constant on fibres of $T^*G$. Then the Pontryagin maximum principle for problem (2.94), (2.95), (2.98) assumes the following form:

$$ \begin{equation*} \begin{gathered} \, \dot h_1=-u_2 h_3,\quad \dot h_2=u_1 h_3,\quad \dot h_3=u_1 h_4,\quad \dot h_4=0, \\ \dot{g}=u_1 X_1+u_2 X_2, \\ u_1 h_1+u_2 h_2+\frac{\nu}{2}(u_1^2+u_2^2) \to \max_{(u_1,u_2) \in \mathbb{R}^2}, \\ \nu \leqslant 0,\quad (h_1,\dots,h_4,\nu) \ne 0. \end{gathered} \end{equation*} \notag $$

Abnormal extremals. Abnormal extremals of constant velocity can be parametrized as follows:

$$ \begin{equation} \begin{gathered} \, h_1=h_2=h_3=0, \quad h_4 \equiv \operatorname{const} \ne 0, \nonumber \\ u_1\equiv 0, \quad u_2\equiv \pm 1, \nonumber \\ x=z \equiv 0, \quad y=\pm t, \quad v=\pm \frac{t^3}{6}\,. \end{gathered} \end{equation} \tag{2.99} $$
Abnormal trajectories (2.99) are one-parameter subgroups $g(t)=e^{\pm tX_2}$. They project onto the $(x,y)$-plane as straight lines, so they are sub-Riemannian length minimizers. The abnormal set is a smooth one-dimensional manifold diffeomorphic to a line:
$$ \begin{equation*} \operatorname{Abn}=\biggl\{g \in G \Bigm| x=z=v-\frac{y^3}{6}=0\biggr\}. \end{equation*} \notag $$

Normal extremals. Normal extremals are trajectories of the normal Hamiltonian system

$$ \begin{equation} \begin{aligned} \, \dot{\lambda}=\vec{H}(\lambda), \quad \lambda \in T^*G, \end{aligned} \end{equation} \tag{2.100} $$
with Hamiltonian $H=(h_1^2+h_2^2)/2$. On the level surface $\{H=1/2\}$ we introduce variables $(\theta,c,\alpha)$:
$$ \begin{equation*} h_1=-\sin\theta, \quad h_2=\cos\theta, \quad h_3=c, \quad h_4=\alpha. \end{equation*} \notag $$
Then the Hamiltonian system (2.100) takes the form
$$ \begin{equation} \dot{\theta}=c, \quad \dot c=-\alpha\sin\theta, \quad \dot{\alpha}=0, \end{equation} \tag{2.101} $$
$$ \begin{equation} \dot{g}=-\sin \theta\cdot X_1+\cos\theta\cdot X_2. \end{equation} \tag{2.102} $$
The vertical subsystem (2.101) is the equation of a pendulum in a gravity field with gravitational acceleration $g=\alpha l$, where $l$ is the length of the pendulum. Thus, for $\alpha> 0$ (for $\alpha < 0$) the gravity force is directed upwards (downwards, respectively) relative to the axis from which we measure $\theta$, while for $\alpha=0$ the pendulum moves in zero gravity.

The projections of normal extremals onto the $(x,y)$-plane are Euler elasticae (see § 2.6).

Abnormal length minimizers satisfy the normal Hamiltonian system (2.101), (2.102) for $\theta=\pi+2\pi n$ and $c=0$, so they are non-strictly abnormal.

The symplectic foliation and Casimir functions There exist two independent Casimir functions on the Lie coalgebra $\mathfrak{g}^*$:

$$ \begin{equation*} \begin{aligned} \, h_4 \quad\text{and}\quad E=\frac{h_3^2}{2}-h_2 h_4, \end{aligned} \end{equation*} \notag $$
where $E$ is the full energy of the pendulum (2.101).

The symplectic foliation $\mathfrak{g}^*$ consists of

  • $\bullet$ parabolic cylinders
    $$ \begin{equation*} \{E=\operatorname{const}, \ h_4=\operatorname{const} \ne 0, \ h_3^2+h_4^2 \ne 0\}, \end{equation*} \notag $$
  • $\bullet$ pairs of planes
    $$ \begin{equation*} \{E=\operatorname{const}, \ h_4=0, \ h_3 \ne 0\}, \end{equation*} \notag $$
  • $\bullet$ points
    $$ \begin{equation*} \{h_i=\operatorname{const}, \ i=1,\dots,4, \ h_3^2+h_4^2=0\}. \end{equation*} \notag $$
Symplectic leaves are two- and zero-dimensional, so the vertical subsystem (2.101) is Liouville integrable. The phase portrait of the Hamiltonian system (2.100) on the cylinder $C\cap \{h_4=\operatorname{const}\}$, where $C=\mathfrak{g}^* \cap \{H=1/2\}$, is obtained as the intersections of this cylinder with level surfaces of the energy $E$.

A parametrization of normal geodesics. The family of normal extremals on the level surface $\{H=1/2\}$ is parametrized by their initial points on the cylinder $C$.

Consider the stratification of $C$ by the submanifolds corresponding to different types of trajectories of the pendulum (2.101):

$$ \begin{equation*} C=\bigsqcup_{i=1}^7 C_i, \end{equation*} \notag $$
where
$$ \begin{equation*} \begin{gathered} \, C_1=\{\lambda \in C \mid \alpha \ne 0, \ E\in(-|\alpha|,|\alpha|)\}, \\ C_2=\{\lambda \in C \mid \alpha \ne 0, \ E\in(|\alpha|,+\infty)\}, \\ C_3=\{\lambda \in C \mid \alpha \ne 0, \ E=|\alpha|, c \ne 0 \}, \\ C_4=\{\lambda \in C \mid \alpha \ne 0, \ E=-|\alpha|\}, \\ C_5=\{\lambda \in C \mid \alpha \ne 0, \ E=|\alpha|, c=0\}, \\ C_6=\{\lambda \in C \mid \alpha=0, \ c \ne 0\}, \end{gathered} \end{equation*} \notag $$
and
$$ \begin{equation*} C_7=\{\lambda \in C \mid \alpha=c=0\}. \end{equation*} \notag $$
Next, the sets $C_i$, $i=1,\dots,5$, are partitioned into subsets depending on the sign of $\alpha$:
$$ \begin{equation*} C_i^+=C_i \cap \{\alpha>0\}, \quad C_i^-=C_i \cap \{\alpha<0\}, \quad i\in\{1,\dots,5\}. \end{equation*} \notag $$

Moreover, the subsets $C_6$, $C_2^{\pm}$, and $C_3^{\pm}$ fall into connected components depending on the sign of $c$:

$$ \begin{equation*} \begin{alignedat}{2} C_{6+}&=C_6 \cap \{c>0\}, &\quad C_{6-}&=C_6 \cap \{c<0\}, \\ C_{i+}^{\pm}&=C_i^{\pm} \cap \{c>0\}, &\quad C_{i-}^{\pm}&=C_i^{\pm}\cap \{c<0\}, \qquad i\in\{2,3\}. \end{alignedat} \end{equation*} \notag $$

For the normalization of normal geodesics, on the strata $C_1$, $C_2$, and $C_3$ we introduce elliptic coordinates $(\varphi,k,\alpha)$ such that the pendulum equation (2.101) straightens out in these coordinates.

In the domain $C_1^+$ we set

$$ \begin{equation*} \begin{gathered} \, k=\sqrt{\frac{E+\alpha}{2 \alpha}}= \sqrt{\frac{c^2}{4\alpha}+\sin^2 \frac{\theta}{2}}\in (0,1), \\ \sin\frac{\theta}{2}=k\operatorname{sn}(\sqrt{\alpha}\,\varphi),\qquad \cos\frac{\theta}{2}=\operatorname{dn}(\sqrt{\alpha}\,\varphi), \\ \frac{c}{2}=k\sqrt{\alpha}\,\operatorname{cn}(\sqrt{\alpha}\,\varphi),\qquad \varphi \in [0,4K]. \end{gathered} \end{equation*} \notag $$

In the domain $C_2^+$ we set

$$ \begin{equation*} \begin{gathered} \, k=\sqrt{\frac{2\alpha}{E+\alpha}}=\frac{1}{\sqrt{c^2/(4\alpha)+ \sin^2(\theta/2)}}\in (0,1), \\ \sin\frac{\theta}{2}=\operatorname{sign} c\cdot \operatorname{sn}\biggl(\frac{\sqrt{\alpha}\,\varphi}{k}\biggr),\qquad \cos\frac{\theta}{2}= \operatorname{cn}\biggl(\frac{\sqrt{\alpha}\,\varphi}{k}\biggr), \\ \frac{c}{2}=\operatorname{sign}c \cdot \frac{\sqrt{\alpha}}{k} \operatorname{dn}\biggl(\frac{\sqrt{\alpha}\,\varphi}{k}\biggr), \qquad \varphi \in [0,2kK], \\ \psi=\frac{\varphi}{k}\,. \end{gathered} \end{equation*} \notag $$

On the set $C_3^+$,

$$ \begin{equation*} \begin{gathered} \, k=1,\\ \sin\frac{\theta}{2}=\operatorname{sign}c\cdot \tanh(\sqrt{\alpha}\,\varphi),\qquad \cos\frac{\theta}{2}=\frac{1}{\cosh(\sqrt{\alpha}\,\varphi)}\,, \\ \frac{c}{2}=\operatorname{sign} c \cdot \frac{\sqrt{\alpha}}{\cosh(\sqrt{\alpha}\,\varphi)}\,, \qquad \varphi \in (-\infty,+\infty). \end{gathered} \end{equation*} \notag $$

On $C_1^-$, $C_2^-$, and $C_3^-$ we define the new coordinates as follows:

$$ \begin{equation} \varphi(\theta,c,\alpha)=\varphi(\theta-\pi,c,-\alpha), \end{equation} \tag{2.103} $$
$$ \begin{equation} k(\theta,c,\alpha)=k(\theta-\pi,c,-\alpha). \end{equation} \tag{2.104} $$

In the new variables the vertical subsystem (2.101) assumes the following form:

$$ \begin{equation*} \dot{\varphi}=1, \qquad \dot{k}=0, \qquad \dot{\alpha}=0, \end{equation*} \notag $$
so it has the solutions
$$ \begin{equation} \varphi_t=\varphi+t, \qquad k=\operatorname{const}, \qquad \alpha=\operatorname{const}. \end{equation} \tag{2.105} $$
The problem is invariant under left shifts on the Engel group, under the dilations
$$ \begin{equation} \delta_s\colon (t,x,y,z,v) \mapsto (e^s t,e^s x,e^s y,e^{2s}z,e^{3s}v), \end{equation} \tag{2.106} $$
$$ \begin{equation} (\theta,c,\alpha) \mapsto (\theta,e^{-s}c,e^{-2s}\alpha), \quad (\varphi,k,\alpha) \mapsto (e^s\varphi,k,e^{-2s}\alpha) \end{equation} \tag{2.107} $$
and under the reflections
$$ \begin{equation*} \begin{gathered} \, (t,x,y,z,v) \mapsto (t,-x,-y,z,-v), \\ (\theta,c,\alpha) \mapsto (\theta-\pi,c,-\alpha), \quad (\varphi,k,\alpha) \mapsto (\varphi,k,-\alpha). \end{gathered} \end{equation*} \notag $$
Dilations define the flow of the vector field
$$ \begin{equation*} Y=x\,\frac{\partial}{\partial x}+y\,\frac{\partial}{\partial y}+ 2z\,\frac{\partial}{\partial z}+3v\,\frac{\partial}{\partial v} \end{equation*} \notag $$
on the Engel group.

For $\lambda=(\varphi,k,\alpha) \in \bigcup_{i=1}^3 C_i$, and $\alpha=1$, geodesics can be parametrized as follows.

If $\lambda \in C_1$, then

$$ \begin{equation} \begin{aligned} \, x_t&=2 k (\operatorname{cn} \varphi_t-\operatorname{cn} \varphi), \nonumber \\ y_t&=2 \bigl(\operatorname{E}(\varphi_t)-\operatorname{E}(\varphi)\bigr)-t, \nonumber \\ z_t&=2k\biggl(\operatorname{sn}\varphi_t\operatorname{dn}\varphi_t- \operatorname{sn}\varphi\operatorname{dn}\varphi- \frac{y_t}{2}(\operatorname{cn}\varphi_t+\operatorname{cn}\varphi)\biggr), \nonumber \\ v_t&=\frac{y_t^3}{6}+2 k^2 \operatorname{cn}^2 \varphi\cdot y_t- 4k^2\operatorname{cn}\varphi\cdot(\operatorname{sn}\varphi_t \operatorname{dn}\varphi_t-\operatorname{sn}\varphi\operatorname{dn}\varphi) \nonumber \\ &\qquad+2k^2\biggl(\frac{2}{3}\operatorname{cn}\varphi_t\operatorname{dn} \varphi_t \operatorname{sn}\varphi_t- \frac{2}{3} \operatorname{cn} \varphi \operatorname{dn} \varphi \operatorname{sn} \varphi+\frac{1-k^2}{3 k^2}\,t \nonumber \\ &\qquad+\frac{2 k^2 -1}{3 k^2}\bigl(\operatorname{E}(\varphi_t)- \operatorname{E}(\varphi)\bigr)\biggr). \end{aligned} \end{equation} \tag{2.108} $$

If $\lambda \in C_2$, then

$$ \begin{equation} \begin{gathered} \, \begin{aligned} \, x_t&=\frac{2\operatorname{sign}c}{k}(\operatorname{dn}\psi_t- \operatorname{dn}\psi), \nonumber \\ y_t&=\frac{k^2-2}{k^2}\,t+\frac{2}{k}\bigl(\operatorname{E}(\psi_t)- \operatorname{E}(\psi)\bigr), \nonumber \\ z_t&=-\frac{x_t y_t}{2}-\frac{2\operatorname{sign}c\, \operatorname{dn}\psi}{k}\,y_t+2\operatorname{sign}c\, (\operatorname{cn}\psi_t\operatorname{sn}\psi_t- \operatorname{cn} \psi \operatorname{sn} \psi), \nonumber \\ v_t&=\frac{4}{k}\biggl(\!\frac{1}{3}\operatorname{cn}\psi_t \operatorname{dn} \psi_t \operatorname{sn} \psi_t -\frac{1}{3} \operatorname{cn} \psi \operatorname{dn} \psi \operatorname{sn} \psi-\frac{1-k^2}{3 k^3}\,t- \frac{k^2-2}{6 k^2}\bigl(\operatorname{E}(\psi_t)- \operatorname{E}(\psi)\bigr)\!\biggr) \nonumber \\ &\qquad+\frac{y_t^3}{6}+\frac{2y_t}{k^2}\operatorname{dn}^2\psi- \frac{4}{k}\operatorname{dn}\psi\bigl(\operatorname{cn}\psi_t \operatorname{sn}\psi_t-\operatorname{cn}\psi\operatorname{sn}\psi\bigr), \nonumber \end{aligned} \\ \psi=\frac{\varphi}{k}\,, \quad \psi_t=\psi+\frac{t}{k}\,. \end{gathered} \end{equation} \tag{2.109} $$

If $\lambda \in C_3$, then

$$ \begin{equation} \begin{aligned} \, x_t&=2 \operatorname{sign}c\biggl(\frac{1}{\cosh\varphi_t}- \frac{1}{\cosh\varphi}\biggr), \nonumber \\ y_t&=2(\tanh \varphi_t-\tanh \varphi)-t , \nonumber \\ z_t&=-\frac{x_t y_t}{2}- \frac{2\operatorname{sign} c}{\cosh \varphi}\, y_t+ 2\operatorname{sign}c\biggl(\frac{\tanh\varphi_t}{\cosh\varphi_t}- \frac{\tanh\varphi}{\cosh\varphi}\biggr), \nonumber \\ v_t&=\frac{2}{3}\biggl(\tanh\varphi_t-\tanh\varphi+ 2\,\frac{\tanh\varphi_t}{\cosh^2\varphi_t}- 2\,\frac{\tanh\varphi}{\cosh^2\varphi}\biggr) \nonumber \\ &\qquad+\frac{y_t^3}{6}+\frac{2y_t}{\cosh^2\varphi}- \frac{4}{\cosh\varphi}\biggl(\frac{\tanh\varphi_t}{\cosh\varphi_t}- \frac{\tanh\varphi}{\cosh\varphi}\biggr). \end{aligned} \end{equation} \tag{2.110} $$
For arbitrary $\lambda=(\varphi,k,\alpha) \in \bigcup_{i=1}^3 C_i$ a parametrization of geodesics can be obtained from the case $\alpha=1$ using dilations and reflections:
  • $\bullet$ if $\alpha > 0$, then
    $$ \begin{equation*} (x_t,y_t,z_t,v_t)(\varphi,k,\alpha)=\biggl(\frac{x_{t'}}{\alpha^{1/2}}\,, \frac{y_{t'}}{\alpha^{1/2}}\,,\frac{z_{t'}}{\alpha}\,, \frac{v_{t'}}{\alpha^{3/2}}\biggr)(\sqrt\alpha\,\varphi,k,1), \quad t'=t\sqrt\alpha\,; \end{equation*} \notag $$
  • $\bullet$ if $\alpha < 0$, then
    $$ \begin{equation*} (x_t,y_t,z_t,v_t)(\varphi,k,\alpha)=(-x_t,-y_t,z_t,-v_t)(\varphi,k,-\alpha). \end{equation*} \notag $$
In the remaining cases, when $\lambda \in \bigcup_{i=4}^7C_i$, geodesics are parametrized by elementary functions.

If $\lambda \in C_4$, then

$$ \begin{equation*} x_t=0, \qquad y_t=t \operatorname{sign}\alpha, \qquad z_t=0, \qquad v_t=\frac{t^3}{6} \operatorname{sign} \alpha. \end{equation*} \notag $$

If $\lambda \in C_5$, then

$$ \begin{equation*} x_t=0, \qquad y_t=- t \operatorname{sign} \alpha, \qquad z_t=0, \qquad v_t=-\frac{t^3}{6} \operatorname{sign} \alpha. \end{equation*} \notag $$

If $\lambda \in C_6$, then

$$ \begin{equation*} \begin{gathered} \, x_t=\frac{\cos (c t+\theta)-\cos \theta}{c}\,,\qquad y_t=\frac{\sin(c t+\theta)-\sin \theta}{c}\,, \qquad z_t=\frac{ct-\sin(ct)}{2c^2}\,, \\ v_t=\frac{3\cos\theta-2ct\sin\theta- 4\cos(ct+\theta)+\cos(2ct+\theta)}{4c^3}\,. \end{gathered} \end{equation*} \notag $$

If $\lambda \in C_7$, then

$$ \begin{equation*} x_t=-t\sin\theta, \qquad y_t=t\cos\theta, \qquad z_t=0, \qquad v_t=\frac{t^3}{6}\cos\theta. \end{equation*} \notag $$

The projections of geodesics onto the $(x,y)$-plane are Euler elasticae (see § 2.6), namely, inflectional ones for $\lambda \in C_1$, non-inflectional ones for $\lambda \in C_2$, critical ones for $\lambda \in C_3$, straight lines for $\lambda \in C_4 \cup C_5 \cup C_7$, and circles for $\lambda \in C_6$.

The family of all geodesics is parametrized by means of the exponential map

$$ \begin{equation*} \operatorname{Exp} \colon N=C \times \mathbb{R}_+ \to M,\quad \operatorname{Exp}(\lambda,t)=g_t=(x_t,y_t,z_t,v_t). \end{equation*} \notag $$

2.9.4. Symmetries of the exponential map and the Maxwell time

The dilations (2.106), (2.107) form a one-parameter group of symmetries of the exponential map. It also has the discrete group of symmetries formed by reflections,

$$ \begin{equation*} \operatorname{Sym}=\{\operatorname{Id},\varepsilon^1,\dots,\varepsilon^7\} \cong \mathbb{Z}_2\times\mathbb{Z}_2\times\mathbb{Z}_2. \end{equation*} \notag $$

Let $\vec H_v=c\,\dfrac{\partial}{\partial\theta}- \alpha\sin\theta\,\dfrac{\partial}{\partial c}\in\operatorname{Vec}(C)$ be the vertical component of the normal Hamiltonian field $\vec H$. The maps $\varepsilon^i\colon C \to C$ below preserve the direction field of the vector field $\vec H_v$:

$$ \begin{equation*} \begin{aligned} \, \varepsilon^1\colon (\theta,c,\alpha) &\mapsto (\theta,-c,\alpha), \\ \varepsilon^2\colon (\theta,c,\alpha) &\mapsto (-\theta,c,\alpha), \\ \varepsilon^3\colon (\theta,c,\alpha) &\mapsto (-\theta,-c,\alpha), \\ \varepsilon^4\colon (\theta,c,\alpha) &\mapsto (\theta+\pi,c,-\alpha), \\ \varepsilon^5\colon (\theta,c,\alpha) &\mapsto (\theta+\pi,-c,-\alpha), \\ \varepsilon^6\colon (\theta,c,\alpha) &\mapsto (-\theta+\pi,c,-\alpha), \\ \varepsilon^7\colon (\theta,c,\alpha) &\mapsto (-\theta+\pi,-c,-\alpha). \end{aligned} \end{equation*} \notag $$
Namely, $\varepsilon^i_*\vec H_v=\vec H_v$ for $i=3,4,7$ and $\varepsilon^i_*\vec H_v=-\vec H_v$ for $i=1,2,5,6$. The actions of the reflections $\varepsilon^i\colon C \to C$ extend to symmetries of the exponential map as follows.

The action $\varepsilon^i\colon N \to N$, $N=C \times \mathbb{R}_{+}$, is defined by

$$ \begin{equation*} {\varepsilon}^i(\lambda,t)=\begin{cases} \bigl({\varepsilon}^i(\lambda),t\bigr) & \text{if}\ {\varepsilon}^i_* \vec{H}_v=\vec{H}_v, \\ \bigl({\varepsilon}^i \circ e^{t \vec{H}_v}(\lambda),t\bigr) & \text{if } \ {\varepsilon}^i_* \vec{H}_v=-\vec{H}_v. \end{cases} \end{equation*} \notag $$

The action $\varepsilon^i\colon G \to G$ is defined by

$$ \begin{equation*} \varepsilon^i(q)=\varepsilon^i(x,y,z,v)=g^i=(x^i,y^i,z^i,v^i), \end{equation*} \notag $$
where
$$ \begin{equation*} \begin{aligned} \, (x^1, y^1, z^1, v^1)&=(x, y, -z, v-x z), \\ (x^2, y^2, z^2, v^2)&=(-x, y, z, v-x z), \\ (x^3, y^3, z^3, v^3)&=(-x, y, -z, v), \\ (x^4, y^4, z^4, v^4)&=(-x, -y, z, -v), \\ (x^5, y^5, z^5, v^5)&=(-x, -y, -z, -v+x z), \\ (x^6, y^6, z^6, v^6)&=(x, -y, z, -v+x z), \\ (x^7, y^7, z^7, v^7)&=(x, -y, -z, -v). \end{aligned} \end{equation*} \notag $$

Proposition 2.11. The group $\operatorname{Sym}$ is a subgroup of the symmetry group of the exponential map.

Theorem 2.41. On almost all geodesics the first Maxwell time corresponding to the symmetry group $\operatorname{Sym}$ can be expressed as follows:

$$ \begin{equation} \lambda \in C_1 \quad\Longrightarrow\quad t_{\rm Max}^1= \frac{\min\bigl(2 p_z^1(k),4 K(k)\bigr)}{\sigma}\,, \end{equation} \tag{2.111} $$
$$ \begin{equation} \lambda \in C_2 \quad\Longrightarrow\quad t_{\rm Max}^1= \frac{2kK(k)}{\sigma}\,, \end{equation} \tag{2.112} $$
$$ \begin{equation} \lambda \in C_6 \quad\Longrightarrow\quad t_{\rm Max}^1=\frac{2\pi}{|c|}\,, \end{equation} \tag{2.113} $$
$$ \begin{equation} \lambda \in C_3 \cup C_4 \cup C_5 \cup C_7 \quad\Longrightarrow\quad t_{\rm Max}^1=+\infty, \end{equation} \tag{2.114} $$
where $\sigma=\sqrt{|\alpha|}$, and $p^1_z(k)\in \bigl(K(k),3K(k)\bigr)$ is the first positive zero of the function $f_z(p,k)=\operatorname{dn} p \, \operatorname{sn}p+(p-2\operatorname{E}(p))\operatorname{cn}p$.

Remark 2.3. On geodesics with first Maxwell time distinct from $t_{\rm Max}^1$, it is lager than this quantity, while $t_{\rm Max}^1$ is the first conjugate time.

Theorem 2.42. The function $t_{\rm Max}^1\colon C \to (0,+\infty]$ has the following invariance properties:

(1) $t_{\rm Max}^1(\lambda)$ depends only on $E$ and $|\alpha|$;

(2) $t_{\rm Max}^1(\lambda)$ in a first integral of the field $\vec H_v$;

(3) $t_{\rm Max}^1(\lambda)$ is reflection invariant: if $(\lambda,t) \in C\times \mathbb{R}_+$, $(\lambda^i, t)=\varepsilon^i(\lambda, t)$, then

$$ \begin{equation*} t_{\rm Max}^1(\lambda^i)=t_{\rm Max}^1(\lambda); \end{equation*} \notag $$

(4) $t_{\rm Max}^1(\lambda)$ is dilation homogeneous: if $\lambda \in C$ and $\lambda_s=\delta_s(\lambda) \in C$, then

$$ \begin{equation*} t_{\rm Max}^1(\lambda_s)=e^st_{\rm Max}^1(\lambda),\qquad s \in \mathbb{R}. \end{equation*} \notag $$

2.9.5. A lower bound for the conjugate time

Theorem 2.43. For each $\lambda \in C$

$$ \begin{equation*} t_{\rm conj}^1(\lambda) \geqslant t_{\rm Max}^1(\lambda). \end{equation*} \notag $$

2.9.6. The diffeomorphism structure of the exponential map

Consider the following subset of the state space, which does not contain fixed points of the symmetries $\varepsilon^1$ and $\varepsilon^2$:

$$ \begin{equation*} \widetilde G=\{g \in G \mid \varepsilon^1(g) \ne g \ne \varepsilon^2(g)\}= \{g \in G \mid xz \ne 0\}, \end{equation*} \notag $$
and consider its connected components
$$ \begin{equation*} \begin{gathered} \, G_1=\{g \in G \mid x < 0, z > 0\}, \\ G_2=\{g \in G \mid x < 0, z < 0\}, \\ G_3=\{g \in G \mid x > 0, z < 0\}, \end{gathered} \end{equation*} \notag $$
and
$$ \begin{equation*} G_4=\{g \in G \mid x > 0, z > 0\}. \end{equation*} \notag $$
Also consider the open dense subset of the space of potentially optimal geodesics
$$ \begin{equation*} \widetilde{N}=\{(\lambda, t) \in N \mid t < t_{\rm Max}^1(\lambda), \ c_{t/2} \sin \theta_{t/2} \ne 0\} \end{equation*} \notag $$
and its connected components
$$ \begin{equation*} \begin{gathered} \, D_1=\bigl\{(\lambda, t) \in N \mid t \in \bigl(0,t_{\rm Max}^1(\lambda)\bigr), \ \sin\theta_{t/2} > 0, \ c_{t/2} > 0\bigr\}, \\ D_2=\bigl\{(\lambda, t) \in N \mid t \in \bigl(0,t_{\rm Max}^1(\lambda)\bigr), \ \sin\theta_{t/2} > 0, \ c_{t/2} < 0\bigr\}, \\ D_3=\bigl\{(\lambda, t) \in N \mid t \in \bigl(0, t_{\rm Max}^1(\lambda)\bigr), \ \sin\theta_{t/2} < 0, \ c_{t/2} < 0\bigr\}, \end{gathered} \end{equation*} \notag $$
and
$$ \begin{equation*} D_4=\bigl\{(\lambda, t) \in N \mid t \in \bigl(0, t_{\rm Max}^1(\lambda)\bigr), \ \sin\theta_{t/2} < 0, \ c_{t/2} > 0\bigr\}. \end{equation*} \notag $$

Theorem 2.44. The following maps are difeomorphisms:

$$ \begin{equation*} \begin{aligned} \, \operatorname{Exp}\colon D_i &\to G_i, \quad i=1,\dots,4, \\ \operatorname{Exp}\colon \widetilde N &\to \widetilde G. \end{aligned} \end{equation*} \notag $$

2.9.7. The cut time

Theorem 2.45. For each $\lambda \in C$

$$ \begin{equation*} t_{\rm cut}(\lambda)=t_{\rm Max}^1(\lambda). \end{equation*} \notag $$

2.9.8. The cut locus and its stratification

Theorem 2.46. The cut locus $\operatorname{Cut}$ lies in the union of coordinate subspaces $\{x=0\}$ and $\{z=0\}$. It is invariant under dilations and discrete symmetries:

$$ \begin{equation*} e^{tY}(\operatorname{Cut})=\operatorname{Cut}, \quad t \in \mathbb{R};\qquad \varepsilon^i(\operatorname{Cut})=\operatorname{Cut}, \quad i=1,\dots,7. \end{equation*} \notag $$

Theorem 2.47. The cut locus has the stratification

$$ \begin{equation*} \begin{aligned} \, \operatorname{Cut}&= (\mathcal{I}_{x+} \sqcup \mathcal{I}_{x-}) \sqcup (\mathcal{N}{x+} \sqcup \mathcal{N}{x-}) \sqcup (\mathcal{I}_{z+} \sqcup \mathcal{I}_{z-}) \\ &\qquad\sqcup (\mathcal{CI}_{x+}^{+} \sqcup \mathcal{CI}_{x+}^{-} \sqcup \mathcal{CI}_{x-}^{+}\sqcup \mathcal{CI}_{x-}^{-})\sqcup (\mathcal{CN}_{x+}^{+}\sqcup \mathcal{CN}_{x+}^{-} \sqcup \mathcal{CN}_{x-}^{+} \sqcup \mathcal{CN}_{x-}^{-}) \\ &\qquad\sqcup (\mathcal{CI}_{z+}^{+} \sqcup \mathcal{CI}_{z+}^{-} \sqcup \mathcal{CI}_{z-}^{+} \sqcup \mathcal{CI}_{z-}^{-}) \\ &\qquad\sqcup (\mathcal{E}_{+} \sqcup \mathcal{E}_{-}). \end{aligned} \end{equation*} \notag $$
The intersections of the cut locus with coordinate subspaces are stratified as follows:
$$ \begin{equation*} \begin{aligned} \, \operatorname{Cut} \cap \{z=0\} &= (\mathcal{I}_{z+} \mathbin{\sqcup} \mathcal{I}_{z-}) \mathbin{\sqcup} (\mathcal{CI}_{z+}^{+} \mathbin{\sqcup} \mathcal{CI}_{z+}^{-} \mathbin{\sqcup} \mathcal{CI}_{z-}^{+} \mathbin{\sqcup} \mathcal{CI}_{z-}^{-}) \\ &\qquad\mathbin{\sqcup} (\mathcal{I}_{x+}^0 \mathbin{\sqcup} \mathcal{I}_{x-}^0)\mathbin{\sqcup} (\mathcal{E}_{+} \mathbin{\sqcup} \mathcal{E}_{-}), \\ \operatorname{Cut} \cap \{x=0\} &=(\mathcal{I}_{x+} \mathbin{\sqcup} \mathcal{I}_{x-}) \mathbin{\sqcup} (\mathcal{CI}_{x+}^{+} \mathbin{\sqcup} \mathcal{CI}_{x+}^{-} \mathbin{\sqcup} \mathcal{CI}_{x-}^{+} \mathbin{\sqcup} \mathcal{CI}_{x-}^{-}) \\ &\qquad\mathbin{\sqcup} (\mathcal{N}_{x+} \mathbin{\sqcup} \mathcal{N}_{x-})\mathbin{\sqcup} (\mathcal{CN}_{x+}^{+} \mathbin{\sqcup} \mathcal{CN}_{x+}^{-} \mathbin{\sqcup} \mathcal{CN}_{x-}^{+} \mathbin{\sqcup} \mathcal{CN}_{x-}^{-}) \\ &\qquad\mathbin{\sqcup} (\mathcal{I}_{z+}^0 \mathbin{\sqcup} \mathcal{I}_{z-}^0)\mathbin{\sqcup} (\mathcal{E}_{+} \mathbin{\sqcup} \mathcal{E}_{-}), \\ \operatorname{Cut} \cap \{x=z=0\} &= (\mathcal{I}_{z+}^0 \mathbin{\sqcup} \mathcal{I}_{z-}^0) \mathbin{\sqcup} (\mathcal{I}_{x+}^0 \mathbin{\sqcup} \mathcal{I}_{x-}^0) \mathbin{\sqcup} (\mathcal{E}_{+} \mathbin{\sqcup} \mathcal{E}_{-}). \end{aligned} \end{equation*} \notag $$
Here $\mathcal{I}_{x\pm}^0 \subset \mathcal{I}_{x\pm}$, $\mathcal{I}_{z\pm}^0 \subset \mathcal{I}_{z\pm}$, and, in addition,
$$ \begin{equation*} \begin{gathered} \, \begin{aligned} \, \mathcal{I}_{z+} &= \bigl\{g \in G \mid z=0, \ y > Y_0^1 |x|, \ w < G_1(x,y)\bigr\} \simeq \mathbb{R}^3, \\ \mathcal{I}_{x+} &= \bigl\{g \in G \mid x=0, \ y > 0, \ w > G_2(z,y)\bigr\} \simeq \mathbb{R}^3, \\ \mathcal{N}{x\pm} &= \bigl\{g \in G \mid x=0, \ \operatorname{sign} z=\pm 1, \ -G_3(z,-y) < w < G_3(z,y)\bigr\} \simeq \mathbb{R}^3, \\ \mathcal{CI}_{z+}^{\pm} &= \bigl\{g \in G \mid z=0, \ y > Y_0^1 |x|, \ w=G_1(x,y), \operatorname{sign} x=\pm 1\bigr\} \simeq \mathbb{R}^2, \\ \mathcal{CI}_{x+}^{\pm} &= \bigl\{g \in G \mid z=0, \ y > 0, \ w=G_2(x,y), \operatorname{sign} z=\pm 1\bigr\} \simeq \mathbb{R}^2, \\ \mathcal{CN}_{x\pm}^{+} &= \bigl\{g \in G \mid x=0, \ \operatorname{sign} z =\pm 1, \ w=G_3(z,y)\bigr\} \simeq \mathbb{R}^2, \\ \mathcal{I}_{z\pm}^0 &= \bigl\{g \in G \mid x=z=0, \ y w < 0, \ \operatorname{sign} y=\pm 1 \bigr\} \simeq \mathbb{R}^2, \\ \mathcal{I}_{x\pm}^0 &= \bigl\{g \in G \mid x=z=0, \ y w > 0, \ \operatorname{sign} y=\pm 1 \bigr\} \simeq \mathbb{R}^2, \\ \mathcal{E}_{\pm} &= \bigl\{g \in G \mid x=y=z=0, \ \operatorname{sign} w=\pm 1 \bigr\} \simeq \mathbb{R}^1, \\ \end{aligned} \\ \begin{gathered} \, \mathcal{I}_{z-} = \varepsilon^4(\mathcal{I}_{z+}), \quad \mathcal{I}_{x-}=\varepsilon^4(\mathcal{I}_{z+}), \\ \mathcal{CI}_{z-}^{\pm}=\varepsilon^4(\mathcal{CI}_{z+}^{\pm}), \quad \mathcal{CI}_{x-}^{\pm}=\varepsilon^4(\mathcal{CI}_{x+}^{\pm}), \quad \mathcal{CN}_{x\pm}^{-}=\varepsilon^4(\mathcal{CN}_{x\pm}^{+}), \end{gathered} \end{gathered} \end{equation*} \notag $$
where $Y_0^1 < 0$ and the $G_i$, $i=1,2,3,$ are certain smooth functions with the following properties:
$$ \begin{equation*} \begin{alignedat}{4} G_1(0,y)&=0, &\qquad G_1(-x,y)&=G_1(x,y), &\qquad G_1(\rho x,\rho y)&=\rho^3 G_1(x,y),&\quad \rho&>0; \\ G_2(0,y)&=0, &\qquad G_2(-z,y)&=G_2(z,y),&\qquad G_2(\rho^2 z,\rho y)&=\rho^3 G_2(z,y),&\quad \rho&>0; \\ &&G_3(-z,y)&=G_3(z,y), &\qquad G_3(\rho^2 z,\rho y)&=\rho^3 G_3(z,y),&\quad \rho&>0. \end{alignedat} \end{equation*} \notag $$

The three-dimensional strata $\mathcal{I}_{x\pm}$ and $\mathcal{I}_{z\pm}$ (or $\mathcal{N}_{x\pm}$) consist of points such that the projections of length minimizers onto the $(x,y)$-plane are elasticae, which are inflectional, that is, with inflection points (or non-inflectional, that is, without inflection points, respectively); see § 2.6. For the one-dimensional strata $\mathcal{E}_{\pm}$ the corresponding elastica are closed (have the shape of a ‘figure-of-eight’).

In Figs 35 and 36 we show the stratifications of the cut locus and its intersection with a coordinate subspace. In Fig. 35 the reader can see the adjacency topology of the quotients of strata of the cut locus by the dilations $Y$. In Fig. 36 we show the intersection $\operatorname{Cut} \mathrel{\cap} \{x=z=0\}$.

In Fig. 37 we show the quotient of the set $\operatorname{Cut}\cap \{z=0\}$ by the dilations $Y$; the quotient $\{z=0\}/ e^{\mathbb{R} Y}$ is shown there as the topological sphere $\{g \in G \mid x^6+y^6+w^2=1\}$. In a similar way, in Fig. 38 the quotient $(\operatorname{Cut}\mathrel{\cap} \{x=0\})/ e^{\mathbb{R}Y}$ is shown on the topological sphere $\{g \in G \mid y^6+|z|^3+w^2=1\}$.

It is obvious that a unique sub-Riemannian length minimizer comes into each point $g_1 \in G \setminus \operatorname{Cut}$. Below we describe a similar property of points $g_1 \in \operatorname{Cut}$.

Theorem 2.48. (1) Precisely two length minimizers come into each point in the three- dimensional strata of the cut locus (these strata consist of Maxwell points that are not conjugate points).

(2) A unique length minimizer comes into each point in the two-dimensional strata (these strata consist of conjugate points that are not Maxwell points).

(3) A one-parameter family of length minimizers come into each point in the one- dimensional strata (these strata consist of Maxwell points that are incidentally conjugate points).

The cut locus is not closed because it contains points lying arbitrarily close to the reference point $q_0$, but not elasticae itself (this is a general fact of sub-Riemannian geometry). The closure of the cut locus in the sub-Riemannian problem on the Engel group has the following simple description.

Theorem 2.49. The equality $\operatorname{cl}(\operatorname{Cut})=\operatorname{Cut} \mathrel{\sqcup} \mathcal{A}_+ \sqcup \mathcal{A}_- \sqcup \{g_0\}$ holds.

The abnormal trajectories $\mathcal{A}_{\pm}$ adjoin strata of the cut locus as shown in Fig. 35.

Theorem 2.50. The following stratifications hold:

$$ \begin{equation*} \operatorname{Cut} \cap \operatorname{Conj} = \bigsqcup_{i \in \{+,-\}\, j\in\{+,-\}} (\mathcal{CI}_{zi}^j \sqcup \mathcal{CI}_{xi}^j \sqcup \mathcal{CN}_{xi}^j) \sqcup \mathcal{E}_+\sqcup \mathcal{E}_- \end{equation*} \notag $$
and
$$ \begin{equation*} \operatorname{Cut} \cap \operatorname{Max} = \bigsqcup_{i \in \{+,-\}} (\mathcal{I}_{zi} \sqcup \mathcal{I}_{xi} \sqcup \mathcal{N}_x^{i} \sqcup \mathcal{E}_i). \end{equation*} \notag $$

2.9.9. Spheres

Sub-Riemannian spheres are taken to one another by left shifts:

$$ \begin{equation*} L_g(S_R(g_0))=S_R(g g_0), \end{equation*} \notag $$
and by dilations:
$$ \begin{equation*} \delta_s(S_R(\operatorname{Id}))=S_{R'}(\operatorname{Id}), \qquad R'=e^s R, \end{equation*} \notag $$
so it is sufficient to examine the unit sphere $S=S_1(\operatorname{Id})$.

The unit sphere is invariant under reflections:

$$ \begin{equation*} \varepsilon^i(S)=S, \qquad i=1,\dots,7. \end{equation*} \notag $$
Consider the cross-section of the unit sphere by the two-dimensional fixed-point manifold of the basic symmetries $\varepsilon^1$ and $\varepsilon^2$:
$$ \begin{equation*} \widetilde{S}=\{g \in S \mid \varepsilon^1(g)= \varepsilon^2(g)=g\}=S \cap \{x=z=0\} \end{equation*} \notag $$
(see Fig. 39).

The cross-section $\widetilde{S}$ is centrally symmetric because of the reflection $\varepsilon^4$:

$$ \begin{equation*} \begin{gathered} \, \varepsilon^4(\gamma_i)=\gamma_{i+2}, \qquad i=1, 2, \\ \varepsilon^4(A_+)=A_-, \qquad \varepsilon^4(C_+)=C_-. \end{gathered} \end{equation*} \notag $$
Different points in $\widetilde{S}$ can be characterized as follows: Points in $\widetilde{S}$ have the following multiplicities $\mu$ (the number of length minimizers from $\operatorname{Id}$ to the point in question):

Theorem 2.51. The cross-section $\widetilde{S}$ has the following regularity at different points:

(1) the curves $\gamma_i$ are analytic and regular;

(2) $A_{\pm}$ and $C_{\pm}$ are singular points; $\widetilde{S}$ is not smooth but Lipschitz at these points;

(3) $\overline{\gamma}_2=\gamma_2 \cup \{C_+, A_+\}$ is smooth of class $C^{\infty}$;

(4) $\gamma_1 \cup \{C_+\}$ is smooth of class $C^{\infty}$;

(5) $\gamma_1 \cup \{A_-\}$ is smooth of class $C^{1}$.

Theorem 2.52. (1) The set $\widetilde{S} \setminus \{A_+, A_-\}$ is semianalytic and therefore subanalytic.

(2) In a neighbourhood of $A_-$ the curve $\gamma_1$ is the graph of a non-analytic function

$$ \begin{equation*} w=\frac 16 Y^3-4 Y^3 \exp\biggl(-\frac{2}{Y}\biggr)(1+ o(1)), \qquad Y=\frac{y+1}{2} \to 0. \end{equation*} \notag $$

(3) The set $\widetilde{S}$ is not semianalytic, so it is not subanalytic.

(4) The sphere $S$ is not subanalytic.

Remark 2.4. That the Engel sphere $S$ is not subanalytic also follows from its projection onto the (non-subanalytic) Martinet sphere (see § 2.3).

Theorem 2.53. In a neighbourhood of $A_-$ the curve $\gamma_1$ is the graph of a function

$$ \begin{equation*} w=F\biggl(Y,\frac{e^{-1/Y}}{Y}\biggr), \qquad Y=\frac{y+1}{2} \to 0, \end{equation*} \notag $$
in the $\exp$-$\log$-category, where $F(\xi,\eta)$ is an analytic function in a neighbourhood of the point $(\xi,\eta )=(0,0)$.

Hence $\widetilde{S}$ is a set in the $\exp$-$\log$-category.

Theorem 2.54. The partition

$$ \begin{equation*} \widetilde{S}=\bigcup_{i=1}^4 \gamma_i \cup \{A_+,A_-,C_+,C_-\} \end{equation*} \notag $$
is a Whitney stratification.

2.9.10. Explicit expressions for the sub-Riemannian distance

For some points in the Engel group we know their distances to the identity element:

2.9.11. Metric lines

Theorem 2.55. The following geodesics and only they are metric straight lines with natural parametrization on the Engel group:

(1) one-parameter subgroups tangent to the distribution:

$$ \begin{equation} \begin{aligned} \, &e^{(u_1X_1+u_2X_2) t}=\operatorname{Exp}(\lambda, t), \qquad t \in \mathbb{R}, \\ &u_1=-\sin \theta, \quad u_2=\cos \theta, \quad \lambda=(\theta, c=0, \alpha) \in C_4\cup C_5, \nonumber \end{aligned} \end{equation} \tag{2.115} $$

(2) critical geodesics:

$$ \begin{equation} \operatorname{Exp}(\lambda, t), \qquad \lambda \in C_3, \quad t\in \mathbb{R}. \end{equation} \tag{2.116} $$

Remark 2.5. The geodesics (2.115) project onto the $(x,y)$-plane as Euclidean straight lines. Among them only the curves

$$ \begin{equation*} e^{X_2 t}=\operatorname{Exp}(\lambda,t), \qquad \lambda=(\theta=0, c=0, \alpha)\in C_4 \cup C_5, \end{equation*} \notag $$
are abnormal. The geodesics (2.116) project onto the $(x,y)$-plane as critical Euler elasticae (see Fig. 24), called Euler solitons.

2.9.12. Bibliographic comments

Subsections 2.9.1, 2.9.3, and 2.9.4 are based on [19], § 2.9.2 is based on [124], § 2.9.5 on [20], §§ 2.9.6 and 2.9.11 on [21], § 2.9.8 on [22], and § 2.9.9 on [145].

A parametization of the Riemannian geodesics on the Engel group was originally obtained in [158].

The sub-Riemannian problem on the Engel group was also considered in [23] and [142].

2.10. The sub-Riemannian problem on the Cartan group

2.10.1. The statement of the problem

The geometric setting. Consider the following generalization (sophistication) of problems on the Heisenberg group [143], [157] and Engel group (§ 2.9): the generalized Dido problem. Let $a_0,a_1 \in \mathbb{R}^2$ be points in the plane which are joined by a curve $\gamma_0 \subset \mathbb{R}^2$. Also fix $S \in \mathbb{R}$ and $c\in \mathbb{R}^2$. The problem consists in connecting $a_0$ with $a_1$ by a shortest curve $\gamma \subset \mathbb{R}^2$ such that the two curves $\gamma_0$ and $\gamma$ bound a plane domain of algebraic area $S$ with centre of mass $c$.

The optimal control problem. We can also state this geometric problem as an optimal control problem:

$$ \begin{equation} \dot{g}=u_1 X_1(g)+u_2 X_2(g), \qquad g=(x, y, z, v, w) \in \mathbb{R}^5, \end{equation} \tag{2.117} $$
$$ \begin{equation} g(0)=g_0, \quad g(t_1)=g_1, \end{equation} \tag{2.118} $$
$$ \begin{equation} l=\int_0^{t_1}\sqrt{u_1^2+u_2^2}\,\,dt \to \min, \end{equation} \tag{2.119} $$
$$ \begin{equation} X_1=\frac{\partial}{\partial x}-\frac{y}{2}\,\frac{\partial}{\partial z}- \frac{x^2 +y^2}{2}\,\frac{\partial}{\partial w}\,, \quad X_2=\frac{\partial}{\partial y}+\frac{x}{2}\,\frac{\partial}{\partial z}+ \frac{x^2 +y^2}{2}\,\frac{\partial}{\partial v}\,. \end{equation} \tag{2.120} $$
This is a sub-Riemannian problem for the sub-Riemannian structure on $\mathbb{R}^5$ defined by the vector fields $X_1$ and $X_2$ as an orthonormal frame.

The Cartan algebra and Cartan group. The Cartan algebra if the free five- dimensional nilpotent Lie algebra $\mathfrak{g}$ with two generators which has depth 3. There exists a basis $\mathfrak{g}=\operatorname{span}(X_1,\dots,X_5)$ such that the nontrivial Lie brackets in this basis are

$$ \begin{equation*} [X_1,X_2]=X_3, \quad [X_1,X_3]=X_4, \quad\text{and}\quad [X_2,X_3]=X_5 \end{equation*} \notag $$
(see Fig. 40).

The Cartan algebra has the grading

$$ \begin{equation*} \mathfrak{g}=\mathfrak{g}^{(1)} \oplus \mathfrak{g}^{(2)} \oplus \mathfrak{g}^{(3)}, \end{equation*} \notag $$
where
$$ \begin{equation*} \begin{gathered} \, \mathfrak{g}^{(1)}=\operatorname{span}(X_1,X_2),\quad \mathfrak{g}^{(2)}=\mathbb{R} X_3,\quad \mathfrak{g}^{(3)}=\operatorname{span}(X_4,X_5),\qquad [\mathfrak{g}^{(1)}, \mathfrak{g}^{(i)}]=\mathfrak{g}^{(i+1)}, \\ \text{and}\quad \mathfrak{g}^{(4)}=\mathfrak{g}^{(5)}=\{0\}, \end{gathered} \end{equation*} \notag $$
so it is a Carnot algebra. The corresponding connected and simply connected Lie group $G$ is called the Cartan group.

On the space $\mathbb{R}^5_{x,y,z,v,w}$ we can introduce multiplication by

$$ \begin{equation*} \begin{pmatrix} x_1 \\ y_1 \\ z_1 \\ v_1 \\ w_1\end{pmatrix}\cdot \begin{pmatrix} x_2 \\ y_2 \\ z_2 \\ v_2 \\ w_2\end{pmatrix}= \begin{pmatrix} x_1+x_2 \\ y_1+y_2 \\ z_1+z_2+(x_1 y_2-y_1 x_2)/2 \\ v_1+v_2+(x_1^2+y_1^2+x_1 x_2+y_1 y_2)y_2/2+x_1 z_2 \\ w_1+w_2-(x_1^2+y_1^2+x_1x_2+y_1 y_2)x_2/2+y_1z_2 \end{pmatrix}, \end{equation*} \notag $$
which makes of it the Cartan group, $G \cong \mathbb{R}^5_{x,y,z,v,w}$, so that the fields (2.120) are left invariant on this group. Thus, problem (2.117)(2.119) is a left-invariant sub-Riemannian problem on the Cartan group. Hence we can assume that $g_0=\operatorname{Id}=(0,\dots,0)$.

Apart from the model (2.120), there are also other models of the sub-Riemannian problem on the Cartan group [10], [56], [124].

The left-invariant sub-Riemannian problem with growth vector $(2,3,5)$ on the Cartan group is unique up to an automorphism of this group [124].

Special features of the problem. The sub-Riemannian problem on the Cartan group is the simplest left-invariant problem with the following properties:

This is the unique free nilpotent sub-Riemannian problem of depth 3 with Liouville-integrable normal Hamiltonian field of the Pontryagin maximum principle (Liouville non-integrable problems are free nilpotent problems of depth 3 and rank greater than 2 [50] and ones of depth greater than 3 and rank at least 2 [104]).

The distribution $\Delta=\operatorname{span}(X_1,X_2)$ has a 14-dimensional algebra of infinitesimal symmetries, which is the special Lie algebra $\mathfrak{g}_2$: this fact goes back to È, Cartan’s famous ‘five-dimensional’ paper [62] (also see § 2.10.2 below).

Finally, the sub-Riemannian problem on the Cartan group provides a nilpotent approximation to any problem with growth vector $(2,3,5)$, for instance,

Any of these reasons would suffice to investigate the sub-Riemannian problem on the Cartan group thoroughly.

2.10.2. Symmetries of the distribution and sub-Riemannian structure

Theorem 2.56. The Lie algebra of infinitesimal symmetries of the distribution $\Delta$ on the Cartan group is the $14$-dimensions algebra $\mathfrak{g}_2$ that is the non-compact real form of the special complex Lie algebra $\mathfrak{g}_2^{\mathbb{C}}$.

Theorem 2.57. The Lie algebra of infinitesimal symmetries of the nilpotent sub- Riemannian structure on the Cartan group is the $6$-dimensional Lie algebra in which there exists a basis $X_0,Y_1,\dots,Y_5$ with non-trivial brackets

$$ \begin{equation*} \begin{gathered} \, [X_0,Y_1]=-Y_2, \qquad [X_0,Y_2]=Y_1, \\ [X_0,Y_4]=-Y_5, \qquad [X_0,Y_5]=Y_4, \\ [Y_1, Y_2]=Y_3,\quad [Y_1, Y_3]=Y_4, \quad [Y_2, Y_3]=Y_5. \end{gathered} \end{equation*} \notag $$
The vector fields $Y_1,\dots,Y_5$ are right-invariant fields on $G$, while $X_0$ vanishes at the identity element. The commutators of these symmetries with the basis fields of the sub-Riemannian structure have the following form:
$$ \begin{equation*} \begin{gathered} \, [Y_i, X_j]=0, \qquad i,j=1,\dots,5, \\ [X_0, X_1]= -X_2, \quad [X_0, X_2]=X_1, \quad [X_0, X_3]=0, \\ [X_0, X_4]=-X_5, \quad [X_0, X_5]=X_4. \end{gathered} \end{equation*} \notag $$
In the model (2.120)
$$ \begin{equation*} X_0=-y\,\frac{\partial}{\partial x}+x\,\frac{\partial}{\partial y}- w\,\frac{\partial}{\partial v}+v\,\frac{\partial}{\partial w}\,. \end{equation*} \notag $$

A representation of the Lie algebra of symmetries of the distribution and sub- Riemannian structure by vector fields in $\mathbb{R}^5$ was presented in [124].

2.10.3. Geodesics

The existence of optimal controls in problem (2.117)(2.119) is a consequence of the Rashevskii–Chow and Filippov theorems.

The Pontryagin maximum principle. Going over from the minimization of length (2.119) to the minimization of the energy $J=\dfrac{1}{2}\displaystyle\int_0^{t_1} (u_1^2+u_2^2)\,dt$ and using the Hamiltonians $h_i(\lambda)=\langle\lambda, X_i\rangle$, $i=1,\dots,5$, which are linear on fibres of $T^*G$ we obtain the assumptions of the Pontryagin maximum principle

$$ \begin{equation*} \begin{gathered} \, \dot h_1=-u_2 h_3,\quad \dot h_2=u_1 h_3,\quad \dot h_3=u_1 h_4+u_2 h_5,\quad \dot h_4=0,\quad \dot h_5=0, \\ \dot{g}=u_1 X_1+u_2 X_2, \\ u_1 h_1+u_2 h_2+\frac{\nu}{2}(u_1^2+u_2^2) \to \max_{(u_1,u_2) \in \mathbb{R}^2}, \\ \nu \leqslant 0,\qquad (h_1,\dots,h_5,\nu) \ne 0. \end{gathered} \end{equation*} \notag $$

Abnormal extremals. Abnormal extremals of constant velocity can be parametrized as follows:

$$ \begin{equation} h_1= h_2=h_3=0, \quad (h_4, h_5) \equiv \operatorname{const} \ne 0, \nonumber \end{equation} \notag $$
$$ \begin{equation} (u_1, u_2) \equiv \operatorname{const}, \nonumber \end{equation} \notag $$
$$ \begin{equation} x=u_1 t, \end{equation} \tag{2.121} $$
$$ \begin{equation} y=u_2 t, \end{equation} \tag{2.122} $$
$$ \begin{equation} z=0, \end{equation} \tag{2.123} $$
$$ \begin{equation} v=\frac{1}{6}(u_1^2+u_2^2)u_1 t^3, \end{equation} \tag{2.124} $$
$$ \begin{equation} w=-\frac{1}{6}(u_1^2+u_2^2)u_2 t^3. \end{equation} \tag{2.125} $$

Abnormal trajectories (2.121)(2.125) are one-parameter subgroups $g_t=e^{t(u_1X_1+u_2X_2)}$ tangent to the distribution $\Delta$. They project onto the $(x,y)$-planes as straight lines, so they are length minimizers.

The abnormal set is a smooth two-dimensional manifold diffeomorphic to $\mathbb{R}^2$:

$$ \begin{equation*} \operatorname{Abn}=\biggl\{g \in G \Bigm| z=v-\frac{1}{6}(x^2+y^2)x= w+\frac{1}{6}(x^2+y^2)y=0\biggr\}. \end{equation*} \notag $$

Normal extremals. Normal extremals satisfy the Hamiltonian system

$$ \begin{equation} \dot{\lambda}=\vec{H}(\lambda), \quad \lambda \in T^*G, \end{equation} \tag{2.126} $$
with Hamiltonian $H=(h_1^2+h_2^2)/2$. On the level surface $\{H=1/2\}$ we introduce the coordinates $(\theta,c,\alpha,\beta) \in S^1\times\mathbb{R}\times\mathbb{R}_+\times S^1$ such that
$$ \begin{equation} h_1=\cos \theta, \quad h_2=\sin \theta, \quad h_3=c, \quad h_4=\alpha \sin \beta, \quad h_5=-\alpha \cos \beta. \end{equation} \tag{2.127} $$
Then the Hamiltonian system (2.126) becomes
$$ \begin{equation} \dot{\theta}=c, \quad \dot c=-\alpha \sin(\theta-\beta), \quad \dot{\alpha}=\dot{\beta}=0, \end{equation} \tag{2.128} $$
$$ \begin{equation} \dot{g}=\cos \theta\, X_1+\sin \theta\,X_2. \end{equation} \tag{2.129} $$
The vertical subsystem (2.128) is the pendulum equations.

The projections of normal geodesics onto the $(x,y)$-plane are Euler elasticae (see § 2.6).

Abnormal length minimizers (2.121)(2.125) satisfy the normal Hamiltonian system (2.128), (2.129) for $\theta=\beta$ and $c=0$, so they are non-strictly abnormal.

The symplectic foliation and Casimir functions. There exist three independent Casimir functions on the Lie coalgebra $\mathfrak{g}^*$:

$$ \begin{equation*} h_4, \quad h_5, \quad E=\frac{h_3^2}{2}+h_1 h_5-h_2 h_4. \end{equation*} \notag $$
The symplectic foliation on $\mathfrak{g}^*$ consists of Symplectic leaves have dimension at most 2, so the vertical subsystem (2.128) is Liouville integrable.

Parametrization of normal geodesics. The family of normal extremals on the level surface $\{H=1/2\}$ is parametrized by their initial points on the cylinder

$$ \begin{equation*} C=\mathfrak{g}^* \cap \biggl\{H=\frac{1}{2}\biggr\}. \end{equation*} \notag $$
This cylinder is stratified depending on different types of trajectories of the pendulum (2.128):
$$ \begin{equation*} C=\bigsqcup_{i=1}^7 C_i, \end{equation*} \notag $$
where
$$ \begin{equation*} \begin{gathered} \, C_1=\{\lambda \in C \mid \alpha > 0, \ E \in (-\alpha,\alpha)\}, \\ C_2=\{\lambda \in C \mid \alpha > 0, \ E \in (\alpha,+\infty)\}, \\ C_3=\{\lambda \in C \mid \alpha > 0, \ E=\alpha, \ \theta-\beta \ne \pi\}, \\ C_4=\{\lambda \in C \mid \alpha > 0, \ E=-\alpha\}, \\ C_5=\{\lambda \in C \mid \alpha > 0, \ E=\alpha, \ \theta-\beta=\pi\}, \\ C_6=\{\lambda \in C \mid \alpha=0, \ c \ne 0\}, \end{gathered} \end{equation*} \notag $$
and
$$ \begin{equation*} C_7=\{\lambda \in C \mid \alpha=c=0\}. \end{equation*} \notag $$
To parametrize normal geodesics, on the strata $C_1$, $C_2$, and $C_3$ we introduce elliptic coordinates $(\varphi,k,\alpha,\beta)$ in which the pendulum equation (2.128) straightens out: if $\lambda \in C_1$, then
$$ \begin{equation*} \begin{gathered} \, k=\sqrt{\frac{E+\alpha}{2\alpha}}= \sqrt{\sin^2\frac{\theta-\beta}{2}+\frac{c^2}{4\alpha}} \in (0,1), \qquad \varphi \in [0,4K], \\ \begin{cases} \sin\dfrac{\theta-\beta}{2}=k\operatorname{sn}(\sqrt{\alpha}\,\varphi), \\ \dfrac{c}{2}=k\sqrt{\alpha}\,\operatorname{cn}(\sqrt{\alpha}\,\varphi); \end{cases} \end{gathered} \end{equation*} \notag $$

if $\lambda \in C_2$, then

$$ \begin{equation*} \begin{gathered} \, k=\sqrt{\frac{2\alpha}{E+\alpha}}=\biggl(\sqrt{\sin^2\frac{\theta-\beta}{2}+ \frac{c^2}{4\alpha}}\biggr)^{-1}\in (0, 1),\qquad \varphi \in [0,2kK], \\ \begin{cases} \sin\dfrac{\theta-\beta}{2}=\pm \operatorname{sn} \dfrac{\sqrt{\alpha}\,\varphi}{k}\,, \\ \dfrac{c}{2}=\pm \dfrac{\sqrt{\alpha}}{k} \operatorname{dn} \dfrac{\sqrt{\alpha}\,\varphi}{k}\,, \end{cases} \qquad \pm=\operatorname{sign} c, \\ \psi=\frac{\varphi}{k}\,; \end{gathered} \end{equation*} \notag $$
if $\lambda \in C_3$, then
$$ \begin{equation*} \begin{gathered} \, k=1, \qquad \varphi \in (-\infty,+\infty), \\ \begin{cases} \sin\dfrac{\theta-\beta}{2}=\pm\tanh(\sqrt{\alpha}\,\varphi), \\ \dfrac{c}{2}=\pm\dfrac{\sqrt{\alpha}}{\cosh(\sqrt{\alpha}\,\varphi)}\,, \end{cases} \qquad \pm=\operatorname{sign} c. \end{gathered} \end{equation*} \notag $$
Then
$$ \begin{equation*} \dot{\varphi}=1, \quad \dot{k}=\dot{\alpha}=\dot{\beta}=0. \end{equation*} \notag $$
The problem is invariant under left shifts on the Cartan group, the dilations
$$ \begin{equation*} \begin{gathered} \, e^{sY}\!\colon (t,x,y,z,v,w)\mapsto (e^st,e^sx,e^sy,e^{2s}z,e^{3s}v,e^{3s}w), \\ (\theta,c,\alpha,\beta)\mapsto (\theta,e^{-s}c,e^{-2s}\alpha,\beta), \\ (\varphi,k,\alpha,\beta)\mapsto (e^s\varphi,k,e^{-2s}\alpha,\beta), \\ Y=x\,\frac{\partial}{\partial x}+y\,\frac{\partial}{\partial y}+ 2z\,\frac{\partial}{\partial z}+3v\,\frac{\partial}{\partial v}+ 3w\,\frac{\partial}{\partial w}\,, \end{gathered} \end{equation*} \notag $$
and the rotations
$$ \begin{equation} \begin{aligned} \, e^{rX_0}\colon (x,y,z,v,w)&\mapsto (x\cos r-y\sin r,x\sin r+y\cos r,z, \nonumber \\ &\qquad v\cos r-w\sin r,v\sin r+w\cos r). \end{aligned} \end{equation} \tag{2.130} $$
Using dilations and rotations we can take any covector $\lambda=(\varphi,k,\alpha,\beta) \in\bigcup_{i=1}^3C_i$ to the fundamental set $\{\alpha=1, \ \beta=0\}$. For $\alpha=1$, $\beta=0$, and $\lambda \in\bigcup_{i=1}^3C_i$ geodesics $g_t=(x_t,y_t,z_t,v_t,w_t)$ are parametrized as follows.

If $\lambda \in C_1$, then

$$ \begin{equation*} \begin{gathered} \, x_t=2(\operatorname{E}(\varphi_t)-\operatorname{E}(\varphi))- (\varphi_t-\varphi), \\ y_t=2k(\operatorname{cn} \varphi-\operatorname{cn} \varphi_t), \\ z_t=2k(\operatorname{sn} \varphi_t \operatorname{dn} \varphi_t- \operatorname{sn}\varphi\operatorname{dn}\varphi)- k(\operatorname{cn}\varphi+\operatorname{cn}\varphi_t)x_t, \\ v_t=2k \operatorname{sn} \varphi_t \operatorname{dn} \varphi_t\cdot x_t- k \operatorname{cn} \varphi_t\cdot x_t^2-(1-2k^2+ 2k^2 \operatorname{cn} \varphi \operatorname{cn}\varphi_t) y_t, \end{gathered} \end{equation*} \notag $$
and
$$ \begin{equation*} \begin{aligned} \, w_t&=-\frac{1}{6}\bigl[x_t^3+2(2k^2-1+6k^2\operatorname{cn}^2\varphi)x_t+ 2(\varphi_t-\varphi) \\ &\qquad+8k^2(\operatorname{sn} \varphi_t \operatorname{cn} \varphi_t \operatorname{dn} \varphi_t-\operatorname{sn} \varphi\operatorname{cn} \varphi \operatorname{dn} \varphi) \\ &\qquad-24 k^2 \operatorname{cn} \varphi (\operatorname{sn} \varphi_t \operatorname{dn} \varphi_t- \operatorname{sn} \varphi \operatorname{dn} \varphi)\bigr], \end{aligned} \end{equation*} \notag $$
where $\varphi_t=\varphi+t$.

If $\lambda \in C_2$, then

$$ \begin{equation*} \begin{gathered} \, x_t=\frac{2}{k}\biggl(\operatorname{E}(\psi_t)- \operatorname{E}(\psi)-\frac{2-k^2}{2}(\psi_t-\psi)\biggr), \\ y_t=\pm \frac{2}{k}(\operatorname{dn} \psi-\operatorname{dn} \psi_t), \\ z_t=\pm \biggl(2(\operatorname{sn} \psi_t \operatorname{cn}\psi_t- \operatorname{sn} \psi \operatorname{cn} \psi)- \frac{1}{k}(\operatorname{dn}\psi+\operatorname{dn}\psi_t)x_t\biggr), \\ v_t=\pm \biggl(2\operatorname{sn} \psi_t \operatorname{cn} \psi_t\cdot x_t- \frac{1}{k} \operatorname{dn} \psi_t\cdot x_t^2 \biggr)+ \frac{1}{k^2}(2-k^2-2\operatorname{dn} \psi \operatorname{dn} \psi_t) y_t, \end{gathered} \end{equation*} \notag $$
and
$$ \begin{equation*} \begin{aligned} \, w_t&=-\frac{1}{6}\biggl(x_t^3+\frac{2}{k^2}(2-k^2+ 6 \operatorname{dn}^2 \psi) x_t+2 k(\psi_t-\psi) \\ &\qquad+\frac{8}{k}(\operatorname{sn}\psi_t\operatorname{cn} \psi_t \operatorname{dn}\psi_t-\operatorname{sn} \psi \operatorname{cn} \psi \operatorname{dn}\psi) \\ &\qquad-\frac{24}{k}\operatorname{dn}\psi\,(\operatorname{sn}\psi_t \operatorname{cn}\psi_t-\operatorname{sn}\psi\operatorname{cn}\psi)\biggr), \end{aligned} \end{equation*} \notag $$
where $\pm=\operatorname{sign}c$ and $\psi_t=\psi+t/k$.

If $\lambda \in C_3$, then

$$ \begin{equation*} \begin{gathered} \, x_t=2(\tanh \varphi_t-\tanh\varphi)-(\varphi_t-\varphi), \\ y_t=\pm 2 \biggl(\frac{1}{\cosh\varphi}-\frac{1}{\cosh\varphi_t}\biggr), \\ z_t=\pm \biggl( 2 \biggl( \frac{\operatorname{sinh} \varphi_t} {\cosh^2 \varphi_t}-\frac{\operatorname{sinh}\varphi}{\cosh^2 \varphi} \biggr)- \biggl( \frac{1}{\cosh \varphi}+\frac{1}{\cosh \varphi_t} \biggr) x_t \biggr), \\ v_t=\pm \biggl(\frac{2}{\operatorname{sinh} \varphi_t}\,x_t- \frac{1}{\cosh \varphi_t}\,x_t^2\biggr)+ \biggl(1-\frac{2}{\cosh \varphi \cosh \varphi_t}\biggr) y_t, \end{gathered} \end{equation*} \notag $$
and
$$ \begin{equation*} \begin{aligned} \, w_t&=-\frac{1}{6}\biggl(x_t^3+ 6\,\frac{2+\cosh^2\varphi}{\cosh^2\varphi}\,x_t+6(\varphi_t-\varphi) \\ &\qquad-\frac{24}{\cosh\varphi}\biggl(\frac{\operatorname{sinh} \varphi_t} {\cosh^2 \varphi_t}-\frac{\operatorname{sinh} \varphi}{\cosh^2 \varphi}\biggr)- 8(\tanh^3 \varphi_t-\tanh^3 \varphi) \biggr), \end{aligned} \end{equation*} \notag $$
where $\pm=\operatorname{sign} c$ and $\varphi_t=\varphi+t$.

We obtain a parametrization of geodesics for arbitrary $\lambda=(\varphi,k,\alpha,\beta) \in\bigcup_{i=1}^3C_i$ from the case where $\alpha=1$ and $\beta=0$ by use of rotations and dilations:

$$ \begin{equation*} \begin{gathered} \, g_t(\varphi,k,\alpha,\beta)=e^{-rX_0}\circ e^{-sY}(g_{t'}(\varphi',k,\alpha'=1,\beta'=0)), \\ t'=t\sqrt{\alpha}\,, \quad \varphi'=\varphi\sqrt{\alpha}\,, \quad r=-\beta, \quad s=\frac{1}{2} \log \alpha. \end{gathered} \end{equation*} \notag $$
In the remaining cases, where $\lambda=(\theta,c,\alpha,\beta) \in \bigcup_{i=4}^7C_i$, geodesics are parametrized by elementary functions:

if $\lambda=(\theta,c,\alpha,\beta) \in C_4 \cup C_5 \cup C_7$ and $\beta=0$, then

$$ \begin{equation*} (x_t,y_t,z_t,v_t,w_t)=\biggl(t,0,0,0,-\frac{t^3}{6}\biggr); \end{equation*} \notag $$
while in the general case, where $\lambda \in C_4 \cup C_5 \cup C_7$,
$$ \begin{equation*} g_t(\theta,c,\alpha,\beta)= e^{-r X_0}\bigl(g_t(\theta',c,\alpha,\beta'=0)\bigr), \quad \theta'=\theta-\beta, \quad r=-\beta; \end{equation*} \notag $$

if $\lambda=(\theta=0,c,\alpha=0) \in C_6$, then

$$ \begin{equation*} \begin{gathered} \, x_t=\frac{\sin \tau}{c}\,, \qquad y_t=\frac{1-\cos \tau}{c}\,, \qquad z_t=\frac{\tau-\sin \tau}{2c^2}\,, \\ v_t=\frac{\cos(2\tau)-4 \cos \tau+3}{4c^3}\,, \qquad w_t=\frac{\sin(2\tau)-4 \sin \tau+2\tau}{4c^3}\,, \qquad \tau=ct; \end{gathered} \end{equation*} \notag $$
while in the general case where $\lambda \in C_6$,
$$ \begin{equation*} g_t(\theta,c,\alpha=0,t)=e^{\theta X_0}(g_t(\theta' =0,c,\alpha=0,t)). \end{equation*} \notag $$

The family of all geodesics is parametrized by means of the exponential map:

$$ \begin{equation*} \operatorname{Exp}\colon(\lambda,t) \mapsto g_t=\pi \circ e^{t \vec H}(\lambda), \quad C \times \mathbb{R}_+ \to G. \end{equation*} \notag $$

2.10.4. Symmetries and Maxwell strata

Continuous symmetries. Dilations and rotations form a two-parameter group of continuous symmetries of the exponential map.

Consider the Haimltonians

$$ \begin{equation*} h_0(\lambda)=\langle \lambda,X_0(g) \rangle \quad\text{and}\quad h_Y(\lambda)=\langle \lambda,Y(g) \rangle,\qquad \lambda \in T^*G, \end{equation*} \notag $$
which are linear on fibres of $T^*G$, and consider the corresponding Hamiltonian vector fields
$$ \begin{equation*} \vec h_0,\vec h_Y \in \operatorname{Vec}(T^* G). \end{equation*} \notag $$
Then
$$ \begin{equation*} \begin{alignedat}{2} [\vec h_0, \vec H]&=0, &\qquad \vec h_0 H&=0, \\ [\vec h_Y, \vec H]&=-2 \vec H, &\qquad \vec h_Y H&=-2 H. \end{alignedat} \end{equation*} \notag $$
Let $e=\displaystyle\sum_{i=1}^5 h_i\,\dfrac{\partial}{\partial h_i}$ be a vertical Euler field on $T^*G$. As $H$ is quadratic on fibres of $T^*G$ , the Hamiltonian field $\vec H$ is linear on fibres, so that
$$ \begin{equation*} [e,\vec H]=\vec H \quad\text{and}\quad e H=2H. \end{equation*} \notag $$
Hence the vector field $Z=\vec h_Y+e$ satisfies the equalities
$$ \begin{equation} \nonumber [Z,\vec H]=-\vec H \quad\text{and}\quad Z H=0. \end{equation} \notag $$
Moreover,
$$ \begin{equation*} [\vec h_0,Z]=0. \end{equation*} \notag $$

Proposition 2.12. For all $t,s,r \in \mathbb{R}$ and $\lambda \in T^*G$

$$ \begin{equation*} e^{rZ} \circ e^{s \vec h_0} \circ e^{t \vec H}(\lambda)= e^{t' \vec H} \circ e^{r Z} \circ e^{s \vec h_0}(\lambda),\quad \textit{where } t'=t e^r. \end{equation*} \notag $$

Discrete symmetries. The quotient of the vertical system (2.128) by the rotations $X_0$ and dilations $Y$ is the standard pendulum equation

$$ \begin{equation*} \dot \theta=c, \quad \dot c=-\sin \theta, \qquad (\theta,c) \in S^1 \times \mathbb{R}. \end{equation*} \notag $$
Its direction field has obvious discrete symmetries, namely, the reflections in the coordinate axes and in the origin
$$ \begin{equation*} \begin{gathered} \, \varepsilon^1\colon(\theta, c) \to (\theta,-c), \\ \varepsilon^2\colon(\theta, c) \to (-\theta, c), \end{gathered} \end{equation*} \notag $$
and
$$ \begin{equation*} \varepsilon^3\colon (\theta, c) \to (-\theta,-c). \end{equation*} \notag $$
These reflections generate the dihedral group
$$ \begin{equation*} D_2=\{\operatorname{Id},\varepsilon^1,\varepsilon^2,\varepsilon^3\}= \mathbb{Z}_2 \times \mathbb{Z}_2. \end{equation*} \notag $$
The actions of reflections extend in a natural way to Euler elasticae $(x_t,y_t)$, so modulo the rotations of the $(x,y)$-plane, The actions of reflections also extend in a natural way to the source space of the exponential map:
$$ \begin{equation*} \varepsilon^i\colon C \times\mathbb{R}_+\to C \times \mathbb{R}_+,\qquad i=1, 2, 3, \end{equation*} \notag $$
and the image of this map:
$$ \begin{equation*} \varepsilon^i\colon G \to G, \qquad i=1, 2, 3, \end{equation*} \notag $$
so that
$$ \begin{equation*} \varepsilon^i \circ \operatorname{Exp}(\lambda, t)= \operatorname{Exp} \circ \varepsilon^i(\lambda,t), \qquad (\lambda, t) \in C \times \mathbb{R}_+, \quad i=1, 2, 3. \end{equation*} \notag $$
More explicitly,
$$ \begin{equation*} \begin{aligned} \, \varepsilon^1 \colon (\theta, c, \alpha, \beta, t) &\mapsto (\theta^1, c^1, \alpha, \beta^1, t)=(\theta_t, -c_t, \alpha, \beta, t), \\ \varepsilon^2 \colon (\theta, c, \alpha, \beta, t) &\mapsto (\theta^2, c^2, \alpha, \beta^2, t)=(-\theta_t, c_t, \alpha, -\beta, t), \\ \varepsilon^3 \colon (\theta, c, \alpha, \beta, t) &\mapsto (\theta^3, c^3, \alpha, \beta^3, t)=(-\theta, -c, \alpha, -\beta, t) \end{aligned} \end{equation*} \notag $$
and
$$ \begin{equation*} \begin{aligned} \, \varepsilon^1\colon (x, y, z, v, w) &\mapsto (x, y, -z, v-xz, w-yz), \\ \varepsilon^2\colon (x, y, z, v, w) &\mapsto (x, -y, z, -v+xz, w-yz), \\ \varepsilon^3\colon (x, y, z, v, w) &\mapsto (x, -y, -z, -v, w). \end{aligned} \end{equation*} \notag $$
The symmetry group $\operatorname{Sym}$ of the exponential map consists of rotations, reflections, and their compositions:
$$ \begin{equation*} \begin{aligned} \, e^{s \vec h_0},e^{s\vec h_0} \circ \varepsilon^i&\colon C \times \mathbb{R}_+\to C \times \mathbb{R}_+, \\ e^{s X_0},e^{sX_0} \circ \varepsilon^i&\colon G\to G. \end{aligned} \end{equation*} \notag $$

Theorem 2.58. Let $\lambda \in C$. Then for almost all geodesics $\operatorname{Exp}(\lambda,t)$ the first Maxwell time corresponding to the group $\operatorname{Sym}$ of symmetries of the exponential map has the following expression:

$$ \begin{equation*} \begin{aligned} \, \lambda \in C_1 &\quad\Longrightarrow\quad t_{\rm Max}^1(\lambda)= \min \biggl(\frac{2}{\sqrt{\alpha}}\,p_1^z(k), \frac{2}{\sqrt{\alpha}}\,p_1^V(k)\biggr), \\ \lambda \in C_2 &\quad\Longrightarrow\quad t_{\rm Max}^1(\lambda)= \frac{2k}{\sqrt{\alpha}}\,p_1^V(k), \\ \lambda \in C_6 &\quad\Longrightarrow\quad t_{\rm Max}^1(\lambda)= \frac{4}{|c|}\,p_1^V(0) \end{aligned} \end{equation*} \notag $$
and
$$ \begin{equation*} \lambda \in C_i, \quad i=3, 4, 5, 7 \quad\Longrightarrow\quad t_{\rm Max}^1(\lambda)=+ \infty. \end{equation*} \notag $$
Here $p=p_1^z(k) \in (K,3K)$ is the first positive zero of the function
$$ \begin{equation*} f_z(p,k)=\operatorname{sn} p \operatorname{dn} p- (2 \operatorname{E}(p)-p) \operatorname{cn} p, \end{equation*} \notag $$
$p=p_1^V(k)$ is the first positive zero of
$$ \begin{equation*} \begin{aligned} \, f_V(p)&=\frac{4}{3}\operatorname{sn}p \, \operatorname{dn}p \bigl[-p-2(1-2 k^2+6 k^2 \operatorname{cn}p^2)(2\operatorname{E}(p)-p)+ (2\operatorname{E}(p)-p)^3 \\ &\qquad+8 k^2\operatorname{cn}p \, \operatorname{sn}p \, \operatorname{dn}p\bigr]+4\operatorname{cn}p \, (1-2 k^2 \operatorname{sn}p^2)(2 \operatorname{E}(p)-p)^2 \end{aligned} \end{equation*} \notag $$
for $\lambda \in C_1$ (in this case $p_1^V(k) \in [2K,4K)$) and of
$$ \begin{equation*} \begin{aligned} \, f_V(p)&=\frac{4}{3}\bigl\{ 3\,\operatorname{dn}p\, (2\operatorname{E}(p)-(2-k^2)p)^2+\operatorname{cn}p\, \bigl[8\operatorname{E}^3(p)-4\operatorname{E}(p)(4+k^2) \\ &\qquad-12\operatorname{E}^2(p)(2-k^2)p +6\operatorname{E}(p)(2-k^2)^2 p^2 \\ &\qquad+p(16-4k^2-3k^4-(2-k^2)^3p^2)\bigr]\,\operatorname{sn}p \\ &\qquad-2\,\operatorname{dn}p\,(-4k^2+3(2\operatorname{E}(p)- (2-k^2)p)^2)\,\operatorname{sn}p^2 \\ &\qquad+12k^2\operatorname{cn}p\,(2\operatorname{E}(p)- (2-k^2)p)\operatorname{sn}p^3- 8k^2\,\operatorname{sn}p^4\operatorname{dn}p\bigr\} \end{aligned} \end{equation*} \notag $$
for $\lambda \in C_2$ (in this case $p_1^V(k) \in (K,2K)$), and $p=p_1^V(0) \in (\pi/2,\pi)$ is the first positive zero of the function
$$ \begin{equation*} f_V^0(p)=\frac{1}{512}[(32p^2-1)\cos(2p)-8p\sin(2p)+\cos(6p)]. \end{equation*} \notag $$

Remark 2.6. On those geodesics where the first Maxwell time corresponding to the group $\operatorname{Sym}$ is distinct from $t_{\rm Max}^1$, it is greater than this quantity, while $t_{\rm Max}^1$ is the first conjugate time.

Theorem 2.59. The function $t_{\rm Max}^1\colon C \to (0,+\infty]$ has the following invariance properties:

(1) $t_{\rm Max}^1(\lambda)$ depends only on $E$ and $|\alpha|$;

(2) $t_{\rm Max}^1(\lambda)$ is a first integral of the field $\vec H_v$;

(3) $t_{\rm Max}^1(\lambda)$ is rotation invariant: if $(\lambda,t) \in C \times \mathbb{R}_+$ and $(\lambda^i, t)=\varepsilon^i(\lambda,t)$, then

$$ \begin{equation*} t_{\rm Max}^1(\lambda^i)=t_{\rm Max}^1(\lambda); \end{equation*} \notag $$

(4) $t_{\rm Max}^1(\lambda)$ is dilation homogeneous: if $\lambda \in C$ and $\lambda_s=\delta_s(\lambda) \in C$, then

$$ \begin{equation*} t_{\rm Max}^1(\lambda_s)=e^s\, t_{\rm Max}^1(\lambda),\qquad s \in \mathbb{R}. \end{equation*} \notag $$

2.10.5. A lower bound for the conjugate time

Theorem 2.60. For each $\lambda \in C$

$$ \begin{equation*} t_{\rm conj}^1(\lambda) \geqslant t_{\rm Max}^1(\lambda). \end{equation*} \notag $$

2.10.6. The cut time and length minimizers

Theorem 2.61. For each $\lambda \in C$

$$ \begin{equation*} t_{\rm cut}(\lambda)=t_{\rm Max}^1(\lambda). \end{equation*} \notag $$

Theorem 2.62. Let $g_1=(x_1,y_1,z_1,v_1,w_1) \in G$. If $z_1\ne 0$ and $x_1 v_1+y_1 w_1-{(x_1^2+y_1^2)z_1}/{2}\ne 0$, then there exists a unique length minimizer connecting $g_0=\operatorname{Id}$ with $g_1$.

2.10.7. Metric straight lines

Theorem 2.63. The following geodesics (and only they) are metric straight lines with natural parametrization on the Cartan group:

(1) the one-parameter groups tangent to the distribution:

$$ \begin{equation} \begin{gathered} \, e^{t(u_1 X_1+u_2 X_2) }=\operatorname{Exp}(\lambda,t), \\ u_1=\cos \theta, \quad u_2=\sin \theta, \quad \lambda=(\theta, c=0, \alpha, \beta)\in C_4\cup C_5\cup C_7, \nonumber \end{gathered} \end{equation} \tag{2.131} $$

(2) the critical geodesics

$$ \begin{equation} \operatorname{Exp}(\lambda, t), \quad \lambda \in C_3. \end{equation} \tag{2.132} $$

Remark 2.7. Geodesics (2.131) project onto the $(x,y)$-plane as Euclidean straight lines, and the geodesics (2.132) project as critical Euler elasticae (see Fig. 24), called Euler solitons.

2.10.8. Bibliographic comments

The sub-Riemannian problem on the Cartan group was considered for the first time by Brockett and Dai [56], who showed that geodesics are integrated by elliptic functions.

Subsections 2.10.1 and 2.10.3 are based on [123]; § 2.10.2 is based on [124]; § 2.10.4 on [125]–[127]; § 2.10.5 on [141]; §§ 2.10.6 and 2.10.7 on [14]. Also see the recent paper [144].

3. Instead of a conclusion: several questions left out

For reasons of space we have not addressed several questions close to those considered above. We list them here.

1. Works on vaconomic mechanics [28]. In particular, authors have considered the ‘Chaplygin sledge’ problem in the variational form, and its hydrodynamic analogue, a planar inertial motion of a rigid body in a fluid; this problem has been integrated by elliptic functions [95]. If one of the added masses is let go to infinity, then we obtain a sub-Riemannian problem in the limit [92]. The problem of ball rolling over a plane was integrated by elliptic functions in [90]. In mechanics the sub-Riemannian problem on the motion group of three-dimensional space corresponds to problems in the dynamics of a rigid body with fixed point; it was considered in [93] together with Poinsot’s geometric picture.

2. Left-invariant sub-Finsler problems [16], [17], [24], [25], [30], [37], [38], [41], [42], [52], [57], [102], [138], [139].

3. Left-invariant sub-Lorentzian problems [26], [61], [80], [82], [91], [147].

4. Left-invariant sub-Riemannian problems with non-integrable geodesic flow [50], [77], [103], [104], [146].

5. Applications of left-invariant problems to nilpotent approximation and to a constructive solution of a two-point control problem [4], [7], [31]–[33], [47], [63], [71], [81], [85], [94], [97], [109], [118], [150], [152], [155], [156], [160].

6. Applications of left-invariant problems to image processing and vision models [34]–[36], [53], [54], [64], [66]–[68], [72]–[74], [110], [112], [113], [120], [121].

7. Applications of left-invariant problems to robotics [11], [13], [15], [27].

The author is thankful to A. A. Agrachev, V. V. Kozlov, A. V. Podobryaev, A. P. Mashtakov, A. A. Ardentov, and I. Yu. Beschastnyi for their helpful advice on the contents and presentation of this paper.

The author is also grateful to E. F. Sachkova for her help in typesetting the text survey and her permanent support during the work on this survey.


Bibliography

1. A. Agrachev and D. Barilari, “Sub-Riemannian structures on 3D Lie groups”, J. Dyn. Control Syst., 18:1 (2012), 21–44  crossref  mathscinet  zmath
2. A. Agrachev, D. Barilari, and U. Boscain, A comprehensive introduction to sub-Riemannian geometry. From the Hamiltonian viewpoint, Cambridge Stud. Adv. Math., 181, Cambridge Univ. Press, Cambridge, 2020, xviii+745 pp.  crossref  mathscinet  zmath
3. A. Agrachev, B. Bonnard, M. Chyba, and I. Kupka, “Sub-Riemannian sphere in Martinet flat case”, ESAIM Control Optim. Calc. Var., 2 (1997), 377–448  crossref  mathscinet  zmath
4. A. Agrachev and A. Marigo, “Nonholonomic tangent spaces: intrinsic construction and rigid dimensions”, Electron. Res. Announc. Amer. Math. Soc., 9 (2003), 111–120  crossref  mathscinet  zmath
5. A. A. Agrachev and Yu. L. Sachkov, “An intrinsic approach to the control of rolling bodies”, Proceedings of the 38th IEEE conference on decision and control (Phoenix, AZ 1999), v. 1, IEEE, 1999, 431–435  crossref
6. A. A. Agrachev and Yu. L. Sachkov, Control theory from the geometric viewpoint, Encyclopaedia Math. Sci., 87, Control theory and optimization II, Springer-Verlag, Berlin, 2004, xiv+412 pp.  crossref  mathscinet  zmath
7. A. A. Agrachev and A. V. Sarychev, “Filtrations of a Lie algebra of vector fields and nilpoint approximation of controlled systems”, Dokl. Akad. Nauk SSSR, 295:4 (1987), 777–781  mathnet  mathscinet  zmath; English transl. in Dokl. Math., 36:1 (1988), 104–108
8. N. I. Akhiezer, Elements of the theory of elliptic functions, 2nd ed., Nauka, Moscow, 1970, 304 pp.  mathscinet  zmath; English transl. Transl. Math. Monogr., 79, Amer. Math. Soc., Providence, RI, 1990, viii+237 pp.  crossref  mathscinet  zmath
9. S. S. Antman, “The influence of elasticity on analysis: modern developments”, Bull. Amer. Math. Soc. (N. S.), 9:3 (1983), 267–291  crossref  mathscinet  zmath
10. A. Anzaldo-Menezes and F. Monroy-Pérez, “Charges in magnetic fields and sub-Riemannian geodesics”, Contemporary trends in nonlinear geometric control theory and its applications (México City 2000), World Sci. Publ., River Edge, NJ, 2002, 183–202  crossref  mathscinet  zmath
11. A. A. Ardentov, “Controlling of a mobile robot with a trailer and its nilpotent approximation”, Regul. Chaotic Dyn., 21:7-8 (2016), 775–791  mathnet  crossref  mathscinet  zmath
12. A. Ardentov, G. Bor, E. Le Donne, R. Montgomery, and Yu. Sachkov, “Bicycle paths, elasticae and sub-Riemannian geometry”, Nonlinearity, 34:7 (2021), 4661–4683  crossref  mathscinet  zmath  adsnasa
13. A. A. Ardentov and I. S. Gubanov, “Modelling trailer car parking along Markov–Dubins and Reeds–Shepp paths”, Program. Sist. Theor. Prilozhen., 10:4 (2019), 97–110 (Russian)  mathnet  crossref
14. A. Ardentov and E. Hakavuori, “Cut time in the sub-Riemannian problem on the Cartan group”, ESAIM Control Optim. Calc. Var., 28 (2022), 12, 19 pp.  crossref  mathscinet  zmath
15. A. A. Ardentov, Y. L. Karavaev, and K. S. Yefremov, “Euler elasticas for optimal control of the motion of mobile wheeled robots: the problem of experimental realization”, Regul. Chaotic Dyn., 24:3 (2019), 312–328  mathnet  crossref  mathscinet  zmath  adsnasa
16. A. A. Ardentov, E. Le Donne, and Yu. L. Sachkov, “Sub-Finsler geodesics on the Cartan group”, Regul. Chaotic Dyn., 24:1 (2019), 36–60  mathnet  crossref  mathscinet  zmath  adsnasa
17. A. A. Ardentov, E. Le Donne, and Yu. L. Sachkov, “A sub-Finsler problem on the Cartan group”, Optimal control and differential equations, Tr. Mat. Inst. Steklova, 304, Seklov Mathematical Institute, Moscow, 2019, 49–67  mathnet  crossref  mathscinet  zmath; English transl. in Proc. Steklov Inst. Math., 304 (2019), 42–59  crossref
18. A. A. Ardentov and Yu. L. Sachkov, “Solution to Euler's elastic problem”, Avtomat. Tekemekh., 2009, no. 4, 78–88  mathnet  mathscinet  zmath; English transl. in Autom. Remote Control, 70:4 (2009), 633–643  crossref
19. A. A. Ardentov and Yu. L. Sachkov, “Extremal trajectories in a nilpotent sub-Riemannian problem on the Engel group”, Mat. Sb., 202:11 (2011), 31–54  mathnet  crossref  mathscinet  zmath; English transl. in Sb. Math., 202:11 (2011), 1593–1615  crossref  adsnasa
20. A. A. Ardentov and Yu. L. Sachkov, “Conjugate points in nilpotent sub-Riemannian problem on the Engel group”, J. Math. Sci. (N. Y.), 195:3 (2013), 369–390  crossref  mathscinet  zmath
21. A. A. Ardentov and Yu. L. Sachkov, “Cut time in sub-Riemannian problem on Engel group”, ESAIM Control Optim. Calc. Var., 21:4 (2015), 958–988  crossref  mathscinet  zmath
22. A. A. Ardentov and Yu. L. Sachkov, “Maxwell strata and cut locus in the sub-Riemannian problem on the Engel group”, Regul. Chaotic Dyn., 22:8 (2017), 909–936  mathnet  crossref  mathscinet  zmath  adsnasa
23. A. A. Ardentov and Yu. L. Sachkov, “Cut locus in the sub-Riemannian problem on Engel group”, Dokl. Ross. Akad. Nauk, 478:6 (2018), 623–626  crossref  mathscinet  zmath; English transl. in Dokl. Math., 97:1 (2018), 82–85  crossref
24. A. A. Ardentov and Yu. L. Sachkov, “Sub-Finsler problem on the Cartan group”, Dokl. Ross. Akad. Nauk, 484:2 (2019), 138–141  crossref  zmath; English transl. in Dokl. Math., 99:1 (2019), 20–22  crossref  mathscinet
25. A. A. Ardentov and Yu. L. Sachkov, “Sub-Finsler structures on the Engel group”, Dokl. Ros. Akad. Nauk, 485:4 (2019), 395–398  crossref  zmath; English transl. in Dokl. Math., 99:2 (2019), 171–174  crossref  mathscinet
26. A. A. Ardentov, Yu. L. Sachkov, T. Huang, and X. Yang, “Extremal trajectories in the sub-Lorentzian problem on the Engel group”, Mat. Sb., 209:11 (2018), 3–31  mathnet  crossref  mathscinet  zmath; English transl. in Sb. Math., 209:11 (2018), 1547–1574  crossref  adsnasa
27. A. A. Ardentov and A. V. Smirnov, “Control of a mobile robot along Euler elasticae”, Program. Sistemy Teor. Prilozhen., 8:4 (2017), 163–178 (Russian)  mathnet  crossref
28. V. I. Arnold, V. V. Kozlov, and A. I. Neishtadt, Mathematical aspects of classical and celestial mechanics, 2nd revised and augmented ed., Editorial URSS, 2002, 416 pp.; English transl. Encyclopaedia Math. Sci., 3, Dynamical systems. III, 3rd rev. ed., Springer-Verlag, Berlin, 2006, xiv+518 pp.  crossref  mathscinet  zmath
29. A. M. Arthurs and G. R. Walsh, “On Hammersley's minimum problem for a rolling sphere”, Math. Proc. Cambridge Philos. Soc., 99:3 (1986), 529–534  crossref  mathscinet  zmath
30. D. Barilari, U. Boscain, E. Le Donne, and M. Sigalotti, “Sub-Finsler structures from the time-optimal control viewpoint for some nilpotent distributions”, J. Dyn. Control Syst., 23:3 (2017), 547–575  crossref  mathscinet  zmath
31. A. Bellaïche, “The tangent space in sub-Riemannian geometry”, Sub-Riemannian geometry, Progr. Math., 144, Birkhäuser, Basel, 1996, 1–78  crossref  mathscinet  zmath
32. A. Bellaïche, J.-P. Laumond, and M. Chyba, “Canonical nilpotent approximation of control systems: application to nonholonomic motion planning”, Proceedings of 32nd IEEE conference on decision and control (San Antonio, TX), v. 3, IEEE, 1993, 2694–2699  crossref
33. A. Bellaïche, J.-P. Laumond, and J.-J. Risler, “Nilpotent infinitesimal approximations to a control Lie algebra”, Nonlinear control systems design 1992, Selected papers from the 2nd IFAC symposium (Bordeaux 1992), IFAC Symposia Series, Pergamon Press, Oxford, 1993, 101–108  crossref
34. E. J. Bekkers, R. Duits, A. Mashtakov, and G. R. Sanguinetti, “A PDE approach to data-driven sub-Riemannian geodesics in $\mathrm{SE}(2)$”, SIAM J. Imaging Sci., 8:4 (2015), 2740–2770  crossref  mathscinet  zmath
35. E. J. Bekkers, R. Duits, A. Mashtakov, and Yu. Sachkov, “Vessel tracking via sub-Riemannian geodesics on the projective line bundle”, Geometric science of information, Lecture Notes in Comput. Sci., 10589, Springer, Cham, 2017, 773–781  crossref  mathscinet  zmath
36. G. Ben-Yosef and O. Ben-Shahar, “A tangent bundle theory for visual curve completion”, IEEE Trans. Pattern Anal. Mach. Intell., 34:7 (2012), 1263–1280  crossref
37. V. N. Berestovskii, “Homogeneous manifolds with intrinsic metric. II”, Sibirsk. Mat. Zh., 30:2 (1989), 14–28  mathnet  mathscinet  zmath; English transl. in Siberian Math. J., 30:2 (1989), 180–191  crossref
38. V. N. Berestovskii, “Structure of homogeneous locally compact spaces with intrinsic metric”, Sibirsk. Mat. Zh., 30:1 (1989), 23–34  mathnet  mathscinet  zmath; English transl. in Siberian Math. J., 30:1 (1989), 16–25  crossref
39. V. N. Berestovskii, “Geodesics of a left-invariant nonholonomic Riemannian metric on the group of motions of the Euclidean plane”, Sibirsk. Mat. Zh., 35:6 (1994), 1223–1229  mathnet  mathscinet  zmath; English transl. in Siberian Math. J., 35:6 (1994), 1083–1088  crossref
40. V. N. Berestovskii, “Geodesics of nonholonomic left-invariant intrinsic metrics on the Heisenberg group and isoperimetric curves on the Minkowski plane”, Sibirsk. Mat. Zh., 35:1 (1994), 3–11  mathnet  mathscinet  zmath; English transl. in Siberian Math. J., 35:1 (1994), 1–8  crossref
41. V. N. Berestovskii and I. A. Zubareva, “Extremals of a left-invariant sub-Finsler metric on the Engel group”, Sibirsk. Mat. Zh., 61:4 (2020), 735–751  mathnet  crossref  mathscinet  zmath; English transl. in Siberian Math. J., 61:4 (2020), 575–588  crossref
42. V. N. Berestovskii and I. A. Zubareva, “PMP, (co)adjoint representation, and normal geodesics, of left-invariant (sub-)Finsler metric on Lie groups”, Chebyshevskii Sb., 21:2 (2020), 43–64  mathnet  crossref  mathscinet  zmath
43. D. Bernoulli, “26th letter to L. Euler (October, 1742)”: P. H. Fuss, Correspondance mathématique et physique, v. 2, St. Petersburg, 1843, 499–507
44. J. Bernoulli, “Véritable hypothèse de la résistance des solides, avec la demonstration de la corbure des corps qui font ressort”, Collected works, v. 2, Geneva, 1744
45. I. Beschastnyi and A. Medvedev, “Left-invariant Sub-Riemannian Engel structures: abnormal geodesics and integrability”, SIAM J. Control Optim., 56:5 (2018), 3524–3537  crossref  mathscinet  zmath
46. I. Yu. Beschastnyi and Yu. L. Sachkov, “Geodesics in the sub-Riemannian problem on the group $\operatorname{SO}(3)$”, Mat. Sb., 207:7 (2016), 29–56  mathnet  crossref  mathscinet  zmath; English transl, in Sb. Math., 207:7 (2016), 915–941  crossref  adsnasa
47. R. M. Bianchini and G. Stefani, “Graded approximations and controllability along a trajectory”, SIAM J. Control Optim., 28:4 (1990), 903–924  crossref  mathscinet  zmath
48. A. Bicchi, D. Prattichizzo, and S. Sastry, “Planning motions of rolling surfaces”, Proceedings of 1995 34th IEEE conference on decision and control (New Orleans, LA 1995), v. 3, IEEE, 1995, 2812–2817  crossref
49. K. E. Bisshopp and D. C. Drucker, “Large deflection of cantilever beams”, Quart. Appl. Math., 3 (1945), 272–275  crossref  mathscinet  zmath
50. I. A. Bizyaev, A. V. Borisov, A. A. Kilin, and I. S. Mamaev, “Integrability and nonintegrability of sub-Riemannian geodesic flows on Carnot groups”, Regul. Chaotic Dyn., 21:6 (2016), 759–774  mathnet  crossref  mathscinet  zmath  adsnasa
51. M. Born, Stabilität der elastischen Linie in Ebene und Raum, Preisschrift und Dissertation, Dieterichsche Universitäts-Buchdruckerei Göttingen, Göttingen, 1906, 101 pp.  zmath
52. U. Boscain, Th. Chambrion, and G. Charlot, “Nonisotropic 3-level quantum systems: complete solutions for minimum time and minimum energy”, Discrete Contin. Dyn. Syst. Ser. B, 5:4 (2005), 957–990  crossref  mathscinet  zmath
53. U. Boscain, R. Duits, F. Rossi, and Yu. Sachkov, “Curve cuspless reconstruction via sub-Riemannian geometry”, ESAIM Control Optim. Calc. Var., 20:3 (2014), 748–770  crossref  mathscinet  zmath
54. U. Boscain, R. A. Chertovskih, J.-P. Gauthier, and A. O. Remizov, “Hypoelliptic diffusion and human vision: a semidiscrete new twist”, SIAM J. Imaging Sci., 7:2 (2014), 669–695  crossref  mathscinet  zmath
55. U. Boscain and F. Rossi, “Invariant Carnot–Caratheodory metrics on $S^3$, $\operatorname{SO}(3)$, $\operatorname{SL}(2)$ and lens spaces”, SIAM J. Control Optim., 47:4 (2008), 1851–1878  crossref  mathscinet  zmath
56. R. W. Brockett and L. Dai, “Non-holonomic kinematics and the role of elliptic functions in constructive controllability”, Nonholonomic motion planning, Kluwer Int. Ser. Eng. Comput. Sci., 192, Kluwer Acad. Publ., Boston, MA, 1993, 1–21  crossref  zmath
57. H. Busemann, “The isoperimetric problem in the Minkowski plane”, Amer. J. Math., 69:4 (1947), 863–871  crossref  mathscinet  zmath
58. Y. A. Butt, Yu. L. Sachkov, and A. I. Bhatti, “Extremal trajectories and Maxwell strata in sub-Riemannian problem on group of motions of pseudo-Euclidean plane”, J. Dyn. Control Syst., 20:3 (2014), 341–364  crossref  mathscinet  zmath
59. Y. A. Butt, Yu. L. Sachkov, and A. I. Bhatti, “Maxwell strata and conjugate points in the sub-Riemannian problem on the Lie group $\operatorname{SH}(2)$”, J. Dyn. Control Syst., 22:4 (2016), 747–770  crossref  mathscinet  zmath
60. Y. A. Butt, Yu. L. Sachkov, and A. I. Bhatti, “Cut locus and optimal synthesis in sub-Riemannian problem on the Lie group $\operatorname{SH}(2)$”, J. Dyn. Control Syst., 23:1 (2017), 155–195  crossref  mathscinet  zmath
61. Q. Cai, T. Huang, Yu. L. Sachkov, and X. Yang, “Geodesics in the Engel group with a sub-Lorentzian metric”, J. Dyn. Control Syst., 22:3 (2016), 465–484  crossref  mathscinet  zmath
62. E. Cartan, “Les systèmes de Pfaff, à cinq variables et les équations aux dérivées partielles du second ordre”, Ann. Sci. École Norm. Sup. (3), 27 (1910), 109–192  crossref  mathscinet  zmath
63. Y. Chitour, F. Jean, and R. Long, “A global steering method for nonholonomic systems”, J. Differential Equations, 254:4 (2013), 1903–1956  crossref  mathscinet  zmath
64. G. Citti and A. Sarti, “A cortical based model of perceptual completion in the roto-translation space”, J. Math. Imaging Vision, 24:3 (2006), 307–326  crossref  mathscinet  zmath
65. L. van den Dries, A. Macintyre, and D. Marker, “The elementary theory of restricted analytic fields with exponentiation”, Ann. of Math. (2), 140:1 (1994), 183–205  crossref  mathscinet  zmath
66. R. Duits, U. Boscain, F. Rossi, and Y. L. Sachkov, “Association fields via cuspless sub-Riemannian geodesics in $\operatorname{SE}(2)$”, J. Math. Imaging Vision, 49:2 (2014), 384–417  crossref  mathscinet  zmath
67. R. Duits, M. Felsberg, G. Granlund, and B. H. Romeny, “Image analysis and reconstruction using a wavelet transform constructed from a reducible representation of the Euclidean motion group”, Int. J. Comput. Vis., 72:1 (2007), 79–102  crossref
68. R. Duits, A. Ghosh, T. C. J. Dela Haije, and A. Mashtakov, “On sub-Riemannian geodesics in $\operatorname{SE}(3)$ whose spatial projections do not have cusps”, J. Dyn. Control Syst., 22:4 (2016), 771–805  crossref  mathscinet  zmath
69. M. S. El Naschie, “Thermal initial post buckling of the extensional elastica”, Int. J. Mech. Sci., 18:6 (1976), 321–324  crossref
70. L. Euler, “De curvis elasticis”, Additamentum I, Methodus inveniendi lineas curvas maximi minimive proprietate gaudentes, sive solutio problematis isoperimetrici latissimo sensu accepti, M.-M. Bousquet, Lausanne–Geneva, 1744, 245–310; English transl. “Leonhard Euler's elastic curves”, Isis, 20:1 (1933), 72–160
71. C. Fernandes, L. Gurvits, and Z. X. Li, “A variational approach to optimal nonholonomic motion planning”, Proceedings. 1991 IEEE international conference on robotics and automation (Sacramento, CA 1991), v. 1, IEEE, 1991, 680–685  crossref
72. B. Franceschiello, A. Mashtakov, G. Citti, and A. Sarti, “Modelling of the Poggendorff illusion via sub-Riemannian geodesics in the roto-translation group”, New trends in image analysis and processing–ICIAP 2017, Lecture Notes in Comput. Sci., 10590, Springer, Cham, 2017, 37–47  crossref  mathscinet
73. B. Franceschiello, A. Mashtakov, G. Citti, and A. Sarti, “Geometrical optical illusion via sub-Riemannian geodesics in the roto-translation group”, Differential Geom. Appl., 65 (2019), 55–77  crossref  mathscinet  zmath
74. E. Franken and R. Duits, “Crossing-preserving coherence-enhancing diffusion on invertible orientation scores”, Int. J. Comput. Vis., 85:3 (2009), 253–278  crossref
75. R. Frisch-Fay, Flexible bars, Butterworths and Co., London, 1962, vii+220 pp.  zmath
76. N. J. Glassmaker and C. Y. Hui, “Elastica solution for a nanotube formed by self-adhesion of a folded thin film”, J. Appl. Phys., 96:6 (2004), 3429–3434  crossref  adsnasa
77. C. Golé and R. Karidi, “A note on Carnot geodesics in nilpotent Lie groups”, J. Dyn. Control Syst., 1:4 (1995), 535–549  crossref  mathscinet  zmath
78. A. G. Greenhill, The applications of elliptic functions, Macmillan, London–New York, 1892, xi+357 pp.  mathscinet  zmath
79. P. A. Griffiths, Exterior differential systems and the calculus of variations, Progr. Math., 25, Birkhäuser, Boston, MA, 1983, ix+335 pp.  crossref  mathscinet  zmath
80. M. Grochowski, “Reachable sets for the Heisenberg sub-Lorentzian metric on $\mathbb{R}^3$. An estimate for the distance function”, J. Dyn. Control Syst., 12:2 (2006), 145–160  crossref  mathscinet  zmath
81. M. Gromov, “Carnot–Carathéodory spaces seen from within”, Sub-Riemannian geometry, Progr. Math., 144, Birkhäuser, Basel, 1996, 79–323  crossref  mathscinet  zmath
82. E. Grong and A. Vasil'ev, “Sub-Riemannian and sub-Lorentzian geometry on $\operatorname{SU}(1,1)$ and on its universal cover”, J. Geom. Mech., 3:2 (2011), 225–260  crossref  mathscinet  zmath
83. J. M. Hammersley, “Oxford commemoration ball”, Probability, statistics and analysis, London Math. Soc. Lecture Note Ser., 79, Cambridge Univ. Press, Cambridge–New York, 1983, 112–142  crossref  mathscinet  zmath
84. G. H. M. van der Heijden, S. Neukirch, V. G. A. Goss, and J. M. T. Thompson, “Instability and self-contact phenomena in the writhing of clamped rods”, Int. J. Mech. Sci., 45 (2003), 161–196  crossref  zmath
85. H. Hermes, “Nilpotent and high-order approximations of vector field systems”, SIAM Rev., 33:2 (1991), 238–264  crossref  mathscinet  zmath
86. V. Jurdjevic, “The geometry of the plate-ball problem”, Arch. Rational Mech. Anal., 124:4 (1993), 305–328  crossref  mathscinet  zmath  adsnasa
87. V. Jurdjevic, “Non-Euclidean elastica”, Amer. J. Math., 117:1 (1995), 93–124  crossref  mathscinet  zmath
88. V. Jurdjevic, Geometric control theory, Cambridge Stud. Adv. Math., 52, Cambridge Univ. Press, Cambridge, 1997, xviii+492 pp.  crossref  mathscinet  zmath
89. V. Jurdjevic, Optimal control and geometry: integrable systems, Cambridge Stud. Adv. Math., 154, Cambridge Univ. Press, Cambridge, 2016, xx+415 pp.  crossref  mathscinet  zmath
90. P. V. Kharlamov, “A critique of some mathematical models of mechanical systems with differential constraints”, Prikl. Mat. Mech., 56:4 (1992), 683–692  mathscinet  zmath; English transl. in J. Appl. Math. Mech., 56:4 (1992), 584–594  crossref  adsnasa
91. A. Korolko and I. Markina, “Nonholonomic Lorentzian geometry on some $\mathbb H$-type groups”, J. Geom. Anal., 19:4 (2009), 864–889  crossref  mathscinet  zmath
92. V. V. Kozlov, “Dynamics of systems with nonintegrable constraints. I”, Vestn. Moskov. Univ. Ser. 1 Mat. Mekh., 1982, no. 3, 92–100  mathnet  mathscinet  zmath; II, 1982, № 4, 70–76  mathnet  mathscinet  zmath; III, 1983, № 3, 102–111  mathnet  mathscinet  zmath; English transl. in I, Moscow Univ. Mech. Bull., 37:3-4 (1982), 27–34; II, 37:3-4 (1982), 74–80; III, 38:3 (1983), 40–51
93. V. V. Kozlov, “Realization of nonintegrable constraints in classical mechanics”, Dokl. Akad. Nauk SSSR, 272:3 (1983), 550–554  mathnet  mathscinet  zmath; English transl. in Soviet Phys. Dokl., 28:9 (1983), 735–737
94. G. Lafferriere and H. J. Sussmann, “A differential geometric approach to motion planning”, Nonholonomic motion planning, Kluwer Int. Ser. Eng. Comput. Sci., 192, Kluwer Acad. Publ., Boston, MA, 1993, 235–270  crossref  zmath
95. H. Lamb, Hydrodynamics, 6th ed., Cambridge Univ. Press, Cambridge, 1932, xv+738 pp.  mathscinet  zmath
96. T. J. Lardner, “A note on the elastica with large loads”, Internat. J. Solids Structures, 21:1 (1985), 21–26  crossref  mathscinet
97. J. P. Laumond, Nonholonomic motion planning for mobile robots, Preprint no. 98211, LAAS-CNRS, Toulouse, 1998
98. D. F. Lawden, Elliptic functions and applications, Appl. Math. Sci., 80, Springer-Verlag, New York, 1989, xiv+334 pp.  crossref  mathscinet  zmath
99. R. Levien, The elastica: a mathematical history, Tech. rep. no. UCB/EECS-2008-103, Univ. of California, Berkeley, 2008, 25 pp. https://www2.eecs.berkeley.edu/Pubs/TechRpts/2008/EECS-2008-103.html
100. Z. Li and J. Canny, “Motion of two rigid bodies with rolling constraint”, IEEE Trans. Robot. Automat., 6:1 (1990), 62–72  crossref
101. J.-M. Lion and J.-P. Rolin, “Théorème de préparation pur les fonctions logarithmico-exponentielles”, Ann. Inst. Fourier (Grenoble), 47:3 (1997), 859–884  crossref  mathscinet  zmath
102. L. V. Lokutsievskiy, “Convex trigonometry with applications to sub-Finsler geometry”, Mat. Sb., 210:8 (2019), 120–148  mathnet  crossref  mathscinet  zmath; English transl. in Sb. Math., 210:8 (2019), 1179–1205  crossref  adsnasa
103. L. V. Lokutsievskii and Yu. L. Sachkov, “Liouville nonintegrability of sub-Riemannian problems on free Carnot groups of step 4”, Dokl. Ross. Akad. Nauk, 474:1 (2017), 19–21  mathnet  crossref  mathscinet  zmath; English transl. in Dokl. Math., 95:3 (2017), 211–213  crossref
104. L. V. Lokutsievskiy and Yu. L. Sachkov, “Liouville integrability of sub-Riemannian problems on Carnot groups of step 4 or greater”, Mat. Sb., 209:5 (2018), 74–119  mathnet  crossref  mathscinet  zmath; English transl. in Sb. Math., 209:5 (2018), 672–713  crossref  adsnasa
105. A. E. H. Love, A treatise on the mathematical theory of elasticity, 4th ed., Cambridge Univ. Press, Cambridge, 1927, xviii+643 pp.  mathscinet  zmath
106. R. S. Manning, J. H. Maddocks, and J. D. Kahn, “A continuum rod model of sequence-dependent DNA structure”, J. Chem. Phys., 105:13 (1996), 5626–5646  crossref  adsnasa
107. R. S. Manning, K. A. Rogers, and J. H. Maddocks, “Isoperimetric conjugate points with application to the stability of DNA minicircles”, R. Soc. Lond. Proc. Ser. A Math. Phys. Eng. Sci., 454:1980 (1998), 3047–3074  crossref  mathscinet  zmath  adsnasa
108. A. Marigo and A. Bicchi, “Rolling bodies with regular surface: the holonomic case”, Differential geometry and control (Boulder, CO 1997), Proc. Sympos. Pure Math., 64, Amer. Math. Soc., Providence, RI, 1999, 241–256  crossref  mathscinet  zmath
109. A. P. Mashtakov, “Algorithms and software for the solution of a constructive control problem using non-holonomic five-dimensional systems”, Program. Sistemy Teor. Prilozhen., 3:1 (2012), 3–29 (Russian)  mathnet
110. A. P. Mashtakov, A. A. Ardentov, and Yu. L. Sachkov, “Parallel algorithm and software for image inpainting via sub-Riemannian minimizers on the group of rototranslations”, Numer. Math. Theory Methods Appl., 6:1 (2013), 95–115  crossref  mathscinet  zmath
111. A. Mashtakov, A. Ardentov, and Yu. L. Sachkov, “Relation between Euler's elasticae and sub-Riemannian geodesics on $\operatorname{SE}(2)$”, Regul. Chaotic Dyn., 21:7-8 (2016), 832–839  mathnet  crossref  mathscinet  zmath  adsnasa
112. A. P. Mashtakov, R. Duits, Yu. L. Sachkov, E. Bekkers, and I. Yu. Beschastnyi, “Sub-Riemannian geodesics in $\operatorname{SO}(3)$ with application to vessel tracking in spherical images of retina”, Dokl. Ross. Akad. Nauk, 473:5 (2017), 521–524  crossref  mathscinet  zmath; English transl. in Dokl. Math., 95:2 (2017), 168–171  crossref
113. A. Mashtakov, R. Duits, Yu. Sachkov, E. J. Bekkers, and I. Beschastnyi, “Tracking of lines in spherical images via sub-Riemannian geodesics in $\operatorname{SO}(3)$”, J. Math. Imaging Vision, 58:2 (2017), 239–264  crossref  mathscinet  zmath
114. A. P. Mashtakov and Yu. L. Sachkov, “Extremal trajectories and the asymptotics of the Maxwell time in the problem of the optimal rolling of a sphere on a plane”, Mat. Sb., 202:9 (2011), 97–120  mathnet  crossref  mathscinet  zmath; English transl. in Sb. Math., 202:9 (2011), 1347–1371  crossref  adsnasa
115. Y. Mikata, “Complete solution of elastica for a clamped-hinged beam, and its applications to a carbon nanotube”, Acta Mech., 190:4 (2007), 133–150  crossref  zmath
116. I. Moiseev and Yu. L. Sachkov, “Maxwell strata in sub-Riemannian problem on the group of motions of a plane”, ESAIM Control Optim. Calc. Var., 16:2 (2010), 380–399  crossref  mathscinet  zmath
117. D. Mumford, “Elastica and computer vision”, Algebraic geometry and its applications (West Lafayette, IN 1990), Springer, New York, 1994, 491–506  crossref  mathscinet  zmath
118. R. M. Murray, “Nilpotent bases for a class of nonintegrable distributions with applications to trajectory generation for nonholonomic systems”, Math. Control Signals Systems, 7:1 (1994), 58–75  crossref  mathscinet  zmath; CDS Tech. Rep. 92-002, Univ. of California, Berkeley, 1992, 19 pp. http://www.cds.caltech.edu/~murray/papers/1992i_mur92-cds.html
119. D. E. Panayotounakos and P. S. Theocaris, “Analytic solutions for nonlinear differential equations describing the elastica of straight bars: theory”, J. Franklin Inst., 325:5 (1988), 621–633  crossref  mathscinet  zmath
120. J. Petitot, “The neurogeometry of pinwheels as a sub-Riemannian contact structure”, J. Physiol. Paris, 97:2-3 (2003), 265–309  crossref
121. J. Petitot, Neurogéométrie de la vision. Modèles mathématiques et physiques des architectures fonctionnelles, Éditions de l'École Polytechnique, Palaiseau, 2008, 419 pp.  mathscinet
122. L. Saalschütz, Der belastete Stab unter Einwirkung einer seitlichen Kraft, B. G. Teubner, Leipzig, 1880, 280 pp.  zmath
123. Yu. L. Sachkov, “Exponential map in the generalized Dido problem”, Mat. Sb., 194:9 (2003), 63–90  mathnet  crossref  mathscinet  zmath; English transl. in Sb. Math., 194:9 (2003), 1331–1359  crossref  adsnasa
124. Yu. L. Sachkov, “Symmetries of flat rank two distributions and sub-Riemannian structures”, Trans. Amer. Math. Soc., 356:2 (2004), 457–494  crossref  mathscinet  zmath
125. Yu. L. Sachkov, “Discrete symmetries in the generalized Dido problem”, Mat. Sb., 197:2 (2006), 95–116  mathnet  crossref  mathscinet  zmath; English transl. in Sb. Math., 197:2 (2006), 235–257  crossref  adsnasa
126. Yu. L. Sachkov, “The Maxwell set in the generalized Dido problem”, Mat. Sb., 197:4 (2006), 123–150  mathnet  crossref  mathscinet  zmath; English transl. in Sb. Math., 197:4 (2006), 595–621  crossref  adsnasa
127. Yu. L. Sachkov, “Complete description of the Maxwell strata in the generalized Dido problem”, Mat. Sb., 197:6 (2006), 111–160  mathnet  crossref  mathscinet  zmath; English transl. in Sb. Math., 197:6 (2006), 901–950  crossref  adsnasa
128. Yu. L. Sachkov, “Control theory on Lie groups”, Optimal control, Sovr. Mat. Fund. Napravl., 27, RUDN University, Moscow, 2008, 5–59  mathnet  mathscinet  zmath; English transl. in J. Math. Sci. (N. Y.), 156:3 (2009), 381–439  crossref
129. Yu. L. Sachkov, “Optimality of Euler's elasticae”, Dokl. Ross. Akad. Nauk, 417:1 (2007), 23–25  mathnet  mathscinet  zmath; English transl. in Dokl. Math., 76:3 (2007), 817–819  crossref
130. Yu. L. Sachkov, “Maxwell strata in Euler elastic problem”, J. Dyn. Control Syst., 14:2 (2008), 169–234  crossref  mathscinet  zmath
131. Yu. L. Sachkov, “Conjugate points in Euler elastic problem”, J. Dyn. Control Syst., 14:3 (2008), 409–439  crossref  mathscinet  zmath
132. Yu. L. Sachkov, “Maxwell strata and symmetries in the problem of optimal rolling of a sphere over a plane”, Mat. Sb., 201:7 (2010), 99–120  mathnet  crossref  zmath; English transl. in Sb. Math., 201:7 (2010), 1029–1051  crossref  mathscinet  adsnasa
133. Yu. L. Sachkov and S. V. Levyakov, “Stability of inflectional elasticae centered at vertices or inflection points”, Differential equations and topology. II, Tr. Mat. Inst. Steklova, 271, MAIK Nauka/Interperiodika, Moscow, 2010, 187–203  mathnet  mathscinet  zmath; English transl. in Proc. Steklov Inst. Math., 271 (2010), 177–192  crossref
134. Yu. L. Sachkov, “Conjugate and cut time in the sub-Riemannian problem on the group of motions of a plane”, ESAIM Control Optim. Calc. Var., 16:4 (2010), 1018–1039  crossref  mathscinet  zmath
135. Yu. L. Sachkov, “Cut locus and optimal synthesis in the sub-Riemannian problem on the group of motions of a plane”, ESAIM Control Optim. Calc. Var., 17:2 (2011), 293–321  crossref  mathscinet  zmath
136. Yu. L. Sachkov, “Closed Euler elasticae”, Differential equations and dynamical systems, Tr. mat. Inst. Steklova, 278, MAIK Nauka/Interperiodika, Moscow, 2012, 227–241  mathnet  mathscinet  zmath; Proc. Steklov Inst. Math., 278 (2012), 218–232  crossref
137. Yu. L. Sachkov and E. F. Sachkova, “Exponential mapping in Euler's elastic problem”, J. Dyn. Control Syst., 20:4 (2014), 443–464  crossref  mathscinet  zmath
138. Yu. L. Sachkov, “Optimal bang-bang trajectories in sub-Finsler problem on the Cartan group”, Nelineinaya Dinamika, 14:4 (2018), 583–593  mathnet  crossref  mathscinet  zmath
139. Yu. L. Sachkov, “Periodic controls in step 2 strictly convex sub-Finsler problems”, Regul. Chaotic Dyn., 25:1 (2020), 33–39  mathnet  crossref  mathscinet  zmath  adsnasa
140. Yu. L. Sachkov, “Homogeneous sub-Riemannian geodesics on a group of motions of the plane”, Differ. Uravn., 57:11 (2021), 1568–1572  crossref  mathscinet  zmath; English transl. in Differ. Equ., 57:11 (2021), 1550–1554  crossref
141. Yu. L. Sachkov, “Conjugate time in the sub-Riemannian problem on the Cartan group”, J. Dyn. Control Syst., 27:4 (2021), я709–751  crossref  mathscinet  zmath
142. Yu. L. Sachkov, Introduction to geometric control, URSS, Moscow, 2021, 160 pp.; augmented English transl. Springer Optim. Appl., 192, Springer, Cham, 2022, xvii+162 pp.  crossref  mathscinet  zmath
143. Yu. L. Sachkov, “Left-invariant optimal control problems on Lie groups: classification and problems integrable by elementary functions”, Uspekhi Mat. Nauk, 77:1(463) (2022), 109–176  mathnet  crossref  mathscinet  zmath; English transl. in Russian Math. Surveys, 77:1 (2022), 99–163  crossref  adsnasa
144. Yu. L. Sachkov, “Sub-Riemannian Cartan sphere”, Dokl. Ross. Akad. Nauk Mat. Informat. Prots. Upravl., 507 (2022), 66–70 (Russian)  crossref
145. Yu. L. Sachkov, “Sub-Riemannian Siegel sphere”, Dokl. Ross. Akad. Nauk Mat. Informat. Prots. Upravl., 500 (2021), 97–101  mathnet  crossref  zmath; English transl. in Dokl. Math., 104:2 (2021), 301–305  crossref  mathscinet
146. Yu. L. Sachkov and E. F. Sachkova, “Degenerate abnormal trajectories in a sub-Riemannian problem with growth vector $(2, 3, 5, 8)$”, Differ. Uravn., 53:3 (2017), 362–374  mathnet  crossref  zmath; English transl. in Differ. Equ., 53:3 (2017), 352–365  crossref  mathscinet
147. Yu. L. Sachkov and E. F. Sachkova, “Sub-Lorentzian problem on the Heisenberg group”, Mat. Zametki, 113:1 (2023), 154–157  mathnet  crossref; English transl. in Math. Notes, 113:1 (2023), 160–163  crossref
148. P. Seide, “Large deflections of a simply supported beam subjected to moment at one end”, J. Appl. Mech., 51:3 (1984), 519–525  crossref  adsnasa
149. I. H. Stampouloglou, E. E. Theotokoglou, and P. N. Andriotaki, “Asymptotic solutions to the non-linear cantilever elastica”, Internat. J. Non-Linear Mech., 40:10 (2005), 1252–1262  crossref  mathscinet  zmath  adsnasa
150. G. Stefani, “Polynomial approximations to control systems and local controllability”, 1985 24th IEEE conference on decision and control (Fort Lauderdale, FL 1985), IEEE, 1985, 33–38  crossref
151. T. Tang and N. J. Glassmaker, “On the inextensible elastica model for the collapse of nanotubes”, Math. Mech. Solids, 15:5 (2010), 591–606  crossref  zmath
152. D. Tilbury, R. M. Murray, and S. S. Sastry, “Trajectory generation for the $N$-trailer problem using Goursat normal form”, IEEE Trans. Automat. Control, 40:5 (1995), 802–819  crossref  mathscinet  zmath
153. S. P. Timoshenko, History of strength of materials, McGraw-Hill Book Co., Inc., New York–Toronto–London, 1953, x+452 pp.  mathscinet  zmath
154. C. Truesdell, “The influence of elasticity on analysis: the classic heritage”, Bull. Amer. Math. Soc. (N. S.), 9:3 (1983), 293–310  crossref  mathscinet  zmath
155. M. Vendittelli, G. Oriolo, and J.-P. Laumond, “Steering nonholonomic systems via nilpotent approximations: the general two-trailer system”, Proceedings 1999 IEEE international conference on robotics and automation (Detroit, MI 1999), v. 1, IEEE, 1999, 823–829  crossref
156. M. Venditelli, G. Oriolo, F. Jean, and J.-P. Laumond, “Nonhomogeneous nilpotent approximations for nonholonomic systems with singularities”, IEEE Trans. Automat. Control, 49:2 (2004), 261–266  crossref  mathscinet  zmath
157. A. M. Vershik and V. Ya. Gershkovich, “Nonholonomic dynamical systems, geometry of distributions and variational problems”, Dynamical systems, 3, Sovr. Probl. mat. Fund. Napavl., 16, VINITI, Moscow, 1987, 5–85  mathnet  mathscinet  zmath; English transl. in Dynamical systems VII, Encyclopaedia Math. Sci., 16, Springer, Berlin, 1994, 1–81  crossref  mathscinet
158. A. M. Vershik and O. A. Granichina, “Reduction of nonholonomic variation problems to isoperimetric ones and connections in principal bundles”, Mat. Zametki, 49:5 (1991), 37–44  mathnet  mathscinet  zmath; English transl. in Math. Notes, 49:5 (1991), 467–472  crossref
159. N. Ja. Vilenkin, Special functions and the theory of group representations, Nauka, Moscow, 1965, 588 pp.  mathscinet  zmath; English transl. Transl. Math. Monogr., 22, Amer. Math. Soc., Providence, RI, 1968, x+613 pp.  mathscinet  zmath
160. G. C. Walsh, R. Montgomery, and S. S. Sastry, “Optimal path planning on matrix Lie groups”, Proceedings of 1994 33rd IEEE conference on decision and control (Lake Buena Vista, FL 1994), v. 2, IEEE, 1994, 1258–1263  crossref
161. C. Y. Wang, “Post-buckling of a clamped-simply supported elastica”, Internat. J. Non-Linear Mech., 32:6 (1997), 1115–1122  crossref  zmath
162. E. T. Whittaker and G. N. Watson, A course of modern analysis. An introduction to the general theory of infinite processes and of analytic functions; with an account of the principal transcendental functions, 4th ed., Cambridge Univ. Press, Cambridge, 1927, vi+608 pp.  mathscinet  zmath

Citation: Yu. L. Sachkov, “Left-invariant optimal control problems on Lie groups that are integrable by elliptic functions”, Uspekhi Mat. Nauk, 78:1(469) (2023), 67–166; Russian Math. Surveys, 78:1 (2023), 65–163
Citation in format AMSBIB
\Bibitem{Sac23}
\by Yu.~L.~Sachkov
\paper Left-invariant optimal control problems on Lie groups that are integrable by elliptic functions
\jour Uspekhi Mat. Nauk
\yr 2023
\vol 78
\issue 1(469)
\pages 67--166
\mathnet{http://mi.mathnet.ru/rm10063}
\crossref{https://doi.org/10.4213/rm10063}
\mathscinet{http://mathscinet.ams.org/mathscinet-getitem?mr=4634796}
\adsnasa{https://adsabs.harvard.edu/cgi-bin/bib_query?2023RuMaS..78...65S}
\transl
\jour Russian Math. Surveys
\yr 2023
\vol 78
\issue 1
\pages 65--163
\crossref{https://doi.org/10.4213/rm10063e}
\isi{https://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcApp=Publons&SrcAuth=Publons_CEL&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=001057003200002}
\scopus{https://www.scopus.com/record/display.url?origin=inward&eid=2-s2.0-85163991145}
Linking options:
  • https://www.mathnet.ru/eng/rm10063
  • https://doi.org/10.4213/rm10063e
  • https://www.mathnet.ru/eng/rm/v78/i1/p67
  • This publication is cited in the following 3 articles:
    Citing articles in Google Scholar: Russian citations, English citations
    Related articles in Google Scholar: Russian articles, English articles
    Успехи математических наук Russian Mathematical Surveys
    Statistics & downloads:
    Abstract page:531
    Russian version PDF:76
    English version PDF:87
    Russian version HTML:315
    English version HTML:167
    References:42
    First page:11
     
      Contact us:
     Terms of Use  Registration to the website  Logotypes © Steklov Mathematical Institute RAS, 2024