1 Introduction

The Jacobi equation of geodesic deviation, e.g. [1], has an important place in General Relativity. The motion of a small enough particle with no internal structure such as charge or spin is described by a geometrical object, a geodesic, so that by the principle of equivalence an observer freely falling with the particle would not be able to feel the effects of gravity in a small neighbourhood of spacetime. Gravitational effects will become apparent considering objects of a finite size, whose evolution can be thought of as that of a bundle of nearby geodesics: at first order in the separation parameter between the geodesics these effects are described by the Jacobi equation. This happens for example when studying the effects of a passing gravitational wave, as in the gravitational-wave memory effect [2,3,4,5,6]. The Jacobi equation can be generalized to the case of particles with electric charge [7], particles with spin [8], or to the non-linear case where the dependence on relative velocities is not linearized [9].

The geodesic deviation has been used to give a geometrical and physical interpretation of spacetimes in ordinary four dimensions and higher [10,11,12], while using first and higher order it has been used to construct approximations to generic geodesics starting from simple ones [13, 14], that can be used to model extreme mass-ratio systems [15]. In another approach, it has been used to study geodesic (in)stability for dynamical systems, and Lyapunov exponents [16, 17].

In this work we analyze the Jacobi equations from the point of view of Hamiltonian dynamics and symmetries of the dynamics. The Jacobi dynamical system is non-trivial, even though it is linear, because of its time dependence and the complexity of its differential equations. Time dependent Hamiltonian systems have been recently discussed in [18] in the light of the Eisenhart lift technique [19,20,21,22]. On the other hand, due to their complexity, few explicit solutions of the equations are given in the literature [7, 13, 23,24,25,26,27,28].

The equations of geodesic deviation are naturally linked to those of geodesic motion. In [23, 29] a remarkable method is presented allowing to obtain a general solution of the Jacobi equation from a complete integral of the Hamilton–Jacobi equation for geodesics. A Lagrangian formulation of the geodesic deviation equations, including an electromagnetic field and spin, and a treatment of higher order deviation equations can be found in [7, 13, 30], and is obtained by an expansion of that of geodesic motion.

Given this natural connection, it is reasonable to expect that symmetries of dynamics of the geodesic motion, when present, descend to symmetries of the Jacobi equation. This is the focus of the present work, where we show that integrals of geodesic motion give rise, through linearization, to integrals of the Jacobi equation that are expressed in the form of invariant Wronskians.

Of particular interest are the geodesic integrals built from Killing vectors and Killing tensors, associated with the explicit and hidden symmetries of the underlying manifold. In references [26,27,28] it was shown how a given symmetry of geodesic motion, encoded in a Killing vector and/or Killing tensor, gives rise to a solution of the Jacobi equation. In other words, each such symmetry was shown to generate one explicit solution of the Jacobi system.

In our paper we show that in fact more is possible. Namely, given a solution of the Jacobi equation (in particular generated by a Killing vector or a Killing tensor) we obtain a linearized conserved quantity, given by the Wronskian. Since each such conserved quantity “can be used twice”, we can obtain another solution of the Jacobi equation by inverting the (linear) form of the Wronskian. In other words, we show that any symmetry of the geodesic motion in fact yields two solutions of the Jacobi system – twice as many as envisioned in references [26,27,28]. In particular, as we shall see in Sect. 5 this result enables us to integrate the Jacobi equation in the Kerr spacetime.

The obtained linearized integrals of Jacobi equation inherit an algebraic structure via Poisson brackets that is isomorphic to the structure of the geodesic integrals. In particular, if in n dimensions there are n functionally independent, mutually Poisson commuting integrals of the geodesic motion, then these induce via linearization a set of n independent, mutually commuting integrals for the Jacobi equation, thus showing that integrability of the geodesic equations implies integrability of the Jacobi equations – a result already derived in [29] in the language of complete integrals for the corresponding Hamilton–Jacobi theory. Let us stress, however, that our discussion of linearized integrals and their properties remains valid irrespective of whether or not the original geodesic motion is completely integrable.

The structure of the work is as follows. We begin in Sect. 2 by reviewing the concept of geodesic motion and setting some of the notation. In Sect. 3 we present the Jacobi equation and discuss its Lagrangian and Hamiltonian formulation, concentrating on subtle points often glossed over in the literature. We first discuss a coordinate approach where the Hamiltonian is not a scalar with respect to changes of coordinates in the base manifold: in this case there is a coordinate Hamiltonian for each coordinate chart, different such Hamiltonians being related by canonical transformations in overlapping regions, and global motion is obtained sewing solutions from different charts. We then introduce a covariant Hamiltonian approach where the Hamiltonian is a globally defined scalar. Novel results are presented in Sect. 4 where we discuss how geodesic integrals of motion descend to integrals of the Jacobi equation. We present the notion of invariant Wronskians, followed by the conserved quantities generated by Killing tensors and then discuss integrability. In Sect. 5 we apply our results to rotating black holes in four dimensions, and mention how these generalize to higher dimensional Kerr–NUT–(A)dS black holes. The Jacobi equation is integrable in these geometries and from our results it is possible to build a complete set of mutually commuting conserved charges. Section 6 presents concluding remarks and possible future lines of research. For convenience of the reader, and to keep the paper self-contained and with appropriate rigour, we also include two appendices. Appendix 1 presents the Jacobi equation, the linearization procedure and the Wronskians from the point of view of a general phase space and a general set of equations of motion, before the introduction of a cotangent bundle or a specific Hamiltonian, and Appendix 1 details the construction of a covariant Lagrangian and Hamiltonian.

2 Geodesic motion

Let \({\mathcal {M}}\) be a spacetime manifold of dimension n, equipped with metric \(g_{ab}\), and Riemann tensorFootnote 1 \(R_{abcd}\). Having in mind applications to relativity we take the metric to be of almost plus type, though with minor adjustments everything would readily generalize to a generic metric.

There are several approaches to the Lagrangian and Hamiltonian descriptions of the relativistic particle. We follow the approach used by Carter [31] and others , where one identifies the spacetime \({\mathcal {M}}\) with a configuration space of the system. The particle is described by its trajectory \(x^a(\lambda )\) parametrized by an external time \(\lambda \). The phase space is then described by the cotangent space \({\mathbf {T}}^*{\mathcal {M}}\). The Lagrangian for geodesic motion is given by

$$\begin{aligned} L = \frac{1}{2}g_{ab}(x)\frac{d x^a}{d\lambda }\frac{d x^b}{d\lambda }\;. \end{aligned}$$
(2.1)

The momentum and the Hamiltonian read

$$\begin{aligned} p_a= & {} g_{ab} \frac{d x^b}{d\lambda }\;, \end{aligned}$$
(2.2)
$$\begin{aligned} H= & {} \frac{1}{2} g^{ab}(x) p_a p_b\;, \end{aligned}$$
(2.3)

and the Hamilton equations reduce to (2.2) and

$$\begin{aligned} \frac{D p_a}{D\lambda }=0\;. \end{aligned}$$
(2.4)

This gives, of course, the geodesic equation.

This formalism describes a free particle of an arbitrary mass. (In this paper we concentrate on the case of massive particles for which \(dx^a/d\lambda \) is not a null vector.) The mass is fixed by the value of the Hamiltonian, i.e., by the normalization of the momentum

$$\begin{aligned} H = -\frac{1}{2} m^2\;. \end{aligned}$$
(2.5)

Since the Hamiltonian does not depend on the external time explicitly, it is a conserved quantity and the mass is for a given trajectory fixed.

We can always rescale the time variable and introduce the proper time \(\tau \)

$$\begin{aligned} \tau = m\lambda \;, \end{aligned}$$
(2.6)

with the normalization of the velocity \(u^a\)

$$\begin{aligned} u^a = \frac{dx^a}{d\tau }=\frac{p^a}{m}\;,\quad u^a u^b g_{ab} = -1\;. \end{aligned}$$
(2.7)

Of course, free particles of different (non-zero) masses follow the same geometric trajectories – they differ only in time parametrization. We can thus ignore a particular value of the mass and describe the geodesic in proper time parametrization.

In what follows we consider a situation where the geodesic motion admits a nontrivial number of integrals of motion. In particular, the conserved quantities that are homogeneous in particle’s momentum,

$$\begin{aligned} K(x,p) = \frac{1}{r!}\, K^{a_1 \ldots a_{r}}(x)\, p_{a_1} \ldots p_{a_{r}} \,, \end{aligned}$$
(2.8)

are in one to one correspondence with Killing tensors [32]. A Killing tensor of rank r is a symmetric tensor \(K^{a_1 \ldots a_r}=K^{(a_1 \ldots a_r)}\), such that

$$\begin{aligned} \nabla ^{(a_0} K^{a_1 \ldots a_{r})} = 0 \, . \end{aligned}$$
(2.9)

For \(r=1\) it reduces to the Killing vector.

It is the purpose of the present paper to linearize these conserved quantities and show that they give rise to simple integrals of motion for the Jacobi geodesic deviation equation described in the next section. In particular, it implies that when the original geodesic motion is completely integrable, so will be the Jacobi equation.

3 The Jacobi system

3.1 Geodesic deviation equation

Given a geodesic \({\bar{x}}^a(\tau )\), we want to study its nearby trajectories. To that purpose we consider a one-parameter family of curves \(x^a(\sigma , \tau )\), with \(\sigma \) being the parameter labeling different curves and \(\tau \) the time parameter along each curve. We call the geodesic \({\bar{x}}^a(\tau )=x^a(0,\tau )\) a central geodesic. We assume \(\tau \) to be the proper time along the central geodesic, however, it does not have to be the proper time for trajectories with nonvanishing \(\sigma \). Nevertheless, we still denote the velocity with respect to \(\tau \) as \(u^a\),

$$\begin{aligned} u^a(\sigma ,\tau ) = \frac{d x^a}{d\tau }(\sigma ,\tau )\;. \end{aligned}$$
(3.1)

We call by \(n^a\) a vector that links the nearby curves,

$$\begin{aligned} n^a (\sigma ,\tau ) = \frac{dx^a}{d\sigma }(\sigma ,\tau )\;. \end{aligned}$$
(3.2)

For \(\sigma =0\), \({\bar{u}}^a(\tau )\equiv u^a(0,\tau )\) reduces to the normalized velocity of the central geodesic and \(n^a(\tau ) \equiv n^a(0,\tau )\) describes the deviation from the central geodesic. One can think of this vector as a linear approximation for the trajectories close to the central geodesic.

A special degenerate case occurs when the whole family \({x^a(\tau ,\sigma )}\) lies entirely on the central geodesic. Clearly, \(n^a\) is then proportional to \(u^a\).

Specifically, we are interested in the trajectories in the vicinity of the central geodesic which are geodesics as well, i.e., curves that (for all values of the parameter \(\sigma \)) satisfy

$$\begin{aligned} \frac{D^2 x^a}{D \tau ^2} = \frac{d^2 x^a}{d\tau ^2} + \Gamma ^a_{bc} \frac{d x^b}{d\tau } \frac{d x^c}{d\tau } = 0 \, . \end{aligned}$$
(3.3)

In this case it is well known that the deviation \(n^a(\tau )\) connecting nearby geodesics must satisfy the geodesic deviation Jacobi equation,

$$\begin{aligned} \frac{D^2 n^a}{D \tau ^2} + {\bar{R}}^a{}{}_{cbd}\, {\bar{u}}^c {\bar{u}}^d n^b = 0 \, . \end{aligned}$$
(3.4)

Here, the bar above the Riemann tensor (and, similarly, above other quantities) indicates that it is evaluated at the central geodesic, \({\bar{R}}_{abcd} = R_{abcd}({\bar{x}})\).

In order to derive this equation, the key observation is to realize that since \(u^a\) and \(n^a\) are essentially coordinate vectors, \(u\equiv {\partial }_\tau \), \(n\equiv {\partial }_\sigma \), on the 2-surface \(x^a(\sigma ,\tau )\), they Lie-commute, \([u,n]=0\), which when expressed in terms of the metric covariant derivative yields

$$\begin{aligned} \frac{D n^a}{D\tau } = \frac{D u^a}{D\sigma }\,, \end{aligned}$$
(3.5)

see e.g. [1] for more details.

3.2 Lagrangian for the Jacobi equation

The Jacobi equation admits a Lagrangian formulation, with the Lagrangian \({\mathscr {l}}\) given by [7]

$$\begin{aligned} \mathscr {l}= \frac{m}{2} {\bar{g}}_{ab} \frac{D n^a}{D \tau } \frac{D n^b}{D \tau } - \frac{m}{2} {\bar{R}}_{abcd}\, {\bar{u}}^a {\bar{u}}^c n^b n^d \,. \end{aligned}$$
(3.6)

This is the Lagrangian for independent ‘deviation’ variable \(n^a\) which represents a general curve close to the central geodesic. As we will discuss below, it can be understood as a function \(\mathscr {l}(n^a,\dot{n}^a)\) of coordinate velocity \(\dot{n}^a=\frac{dn^a}{d\tau }\), or it can be treated ‘covariantly’, as a function \(\mathscr {l}(n^a, \frac{Dn^a}{d\tau })\) of covariant velocity \(\frac{Dn^a}{d\tau }\). In either case the corresponding Euler–Lagrange equations pick up a nearby geodesic specified by the Jacobi equation.

The Lagrangian (3.6) can be derived by the linearization process, starting from the Lagrangian of the geodesic motion.Footnote 2 We refer to the Appendix 1 for the derivation and further technical details.

The Lagrangian (3.6) is obviously time dependent – it depends on the time-dependent position \({\bar{x}}(\tau )\) and velocity \({\bar{u}}(\tau )\) of the central geodesic. Therefore, there is one such Lagrangian for each central geodesic. While the dependence on the velocity \({\bar{u}}(\tau )\) is explicit, the dependence on \({\bar{x}}(\tau )\) enters through the spacetime dependence of the metric, Christoffel symbols, and the Riemann tensor, and to stress this fact we write these objects with bar. All these should be considered as given functions of the time variable \(\tau \) and provide apriori data for the Jacobi equation. Of course, for a concrete spacetime, it may be difficult to obtain the expression for \({\bar{x}}(\tau )\) and \({\bar{u}}(\tau )\) in an explicit and closed form. We return to the question of time-dependency of the Lagrangian again below, when we discuss it from a covariant perspective.

3.3 Transverse and tangent splitting

Before we proceed to the Hamiltonian formulation, let us first discuss decoupling of the transverse and parallel degrees of freedom in the deviation variable. The Jacobi equation (3.4) always admits the following two solutions:

$$\begin{aligned} \begin{aligned} {}^1n^a(\tau )&= {\bar{u}}^a \, ,\\ {}^2n^a(\tau )&= \tau {\bar{u}}^a \, , \end{aligned} \end{aligned}$$
(3.7)

which arise from the degenerate case when the entire family \(x^a(\sigma ,\tau )\) lies on the central geodesic \(x^a(\tau )\):

$$\begin{aligned} \begin{aligned} {}^1x^a (\sigma ;\tau )&= x^a (\tau + \sigma )\,, \\ {}^2x^a (\sigma ;\tau )&= x^a (e^\sigma \tau ) \, . \end{aligned} \end{aligned}$$
(3.8)

These solutions of the Jacobi equation correspond to the known freedom of redefining the affine parameter. Since Eq. (3.4) is a linear equation of the second order, in an n-dimensional spacetime the space of solutions is 2n-dimensional. Removing the trivial solutions (3.8) amounts to reducing the problem to the subspace of deviation vectors that are transverse to the geodesics. Namely, writing

$$\begin{aligned} n^a = n_\parallel ^a + n_\perp ^a\,, \quad n_\parallel ^a = \nu (\tau ) {\bar{u}}^a\,, \quad g_{ab}\, n_\perp ^a {\bar{u}}^b =0\,, \end{aligned}$$
(3.9)

we obtain two separate equations

$$\begin{aligned} \frac{D^2 n_\parallel ^a}{D \tau ^2}&= 0 \,, \end{aligned}$$
(3.10)
$$\begin{aligned} \frac{D^2 n_\perp ^a}{D \tau ^2} + {\bar{R}}^a{}_{cbd} {\bar{u}}^c {\bar{u}}^d n_\perp ^b&= 0 \,. \end{aligned}$$
(3.11)

The solutions of (3.10) are precisely those in Eq. (3.7).

Correspondingly, for timelike \({\bar{u}}^a\), the Lagrangian separates as \(\mathscr {l}= \mathscr {l}_\parallel + \mathscr {l}_\perp \), with

$$\begin{aligned} \mathscr {l}_\parallel&= \frac{m}{2} {\bar{g}}_{ab} \frac{D n_\parallel ^a}{D \tau } \frac{D n_\parallel ^b}{D \tau } \, , \end{aligned}$$
(3.12)
$$\begin{aligned} \mathscr {l}_\perp&= \frac{m}{2} {\bar{g}}_{ab} \frac{D n_\perp ^a}{D \tau } \frac{D n_\perp ^b}{D \tau } - \frac{m}{2} {\bar{R}}_{abcd} {\bar{u}}^a {\bar{u}}^c n_\perp ^b n_\perp ^d \,, \end{aligned}$$
(3.13)

as follows from the fact that

$$\begin{aligned} {\bar{g}}_{ab} \frac{D n_\parallel ^a}{D \tau } \frac{D n_\perp ^b}{D \tau }= & {} {\bar{g}}_{ab} \, {\dot{\nu }} {\bar{u}}^a \frac{D n_\perp ^b}{D \tau } \nonumber \\= & {} {\dot{\nu }} \frac{d}{d \tau } \left( {\bar{g}}_{ab} {\bar{u}}^a n_\perp ^b \right) = 0 \, . \end{aligned}$$
(3.14)

When the geodesics are timelike, the metric can be decomposed

$$\begin{aligned} {\bar{g}}_{ab} = - {\bar{u}}_a {\bar{u}}_b + {\bar{g}}_{\perp ab} \, , \end{aligned}$$
(3.15)

where \({\bar{g}}_{\perp ab}\) is the projector to the transverse space, which is positive definite, and (3.13) can be written as

$$\begin{aligned} \mathscr {l}_\perp = \frac{m}{2} {\bar{g}}_{\perp ab} \frac{D n_\perp ^a}{D \tau } \frac{D n_\perp ^b}{D \tau } - \frac{m}{2} {\bar{R}}_{abcd} {\bar{u}}^a {\bar{u}}^c n_\perp ^b n_\perp ^d \, . \end{aligned}$$
(3.16)

3.4 Hamiltonian formalism: coordinate approach

In what follows we shall study the Jacobi equation from the point of the Hamiltonian dynamics using both, coordinate (this subsection) and covariant (next subsection) approaches.

In the coordinate approach, the Lagrangian (3.6) is understood as a function of the ‘positions’ \(n^a\) and the ‘coordinate velocities’ \({\dot{n}}^a\),

$$\begin{aligned} \mathscr {l}= & {} \frac{m}{2} {\bar{g}}_{ab} ({\dot{n}}^a{+}u^k{\bar{\Gamma }}^a_{kc} n^c) ({\dot{n}}^b{+}u^l{\bar{\Gamma }}^b_{ld} n^d) \nonumber \\&- \frac{m}{2} {\bar{R}}_{kalb} {\bar{u}}^k {\bar{u}}^l n^a n^b . \end{aligned}$$
(3.17)

To write down the corresponding Hamiltonian formulation, we define the momentum canonically conjugated to \(n^a\)

$$\begin{aligned} \pi _a = \frac{{\partial }\mathscr {l}}{{\partial }{\dot{n}}^a} = m {\bar{g}}_{ab} \left( {\dot{n}}^b + {\bar{u}}^k{\bar{\Gamma }}^b_{kc} n^c \right) = m {\bar{g}}_{ab} \frac{D n^b}{D \tau } \, , \end{aligned}$$
(3.18)

and introduce what we call the coordinate Hamiltonian, \({\mathscr {h}_{\mathrm {c}} = \pi _a {\dot{n}}^a - \mathscr {l}}\),

$$\begin{aligned} \mathscr {h}_{\mathrm {c}}= & {} \frac{1}{2m} {\bar{g}}^{ab} \pi _a \pi _b - {\bar{u}}^k{\bar{\Gamma }}^a_{kb} \, \pi _a n^b + \frac{m}{2} {\bar{R}}_{abcd} \, {\bar{u}}^a {\bar{u}}^c n^b n^d \, .\nonumber \\ \end{aligned}$$
(3.19)

An explicit calculation shows that the equations of motion obtain from \(\mathscr {h}_{\mathrm {c}}\) imply the Jacobi equation (3.4).

It is obvious that the Hamiltonian (3.19) is not covariant: it depends in a non-tensorial way on a particular choice of coordinates through the Christoffel symbols \(\Gamma ^a_{bc}\), reflecting the fact that we used a non-covariant form of the velocity \({\dot{n}}^a\). Such a velocity does not transform as a vector under a coordinate transformation and therefore the related Hamiltonian is also non-covariant. The Hamiltonian (3.19) generates the coordinate time-evolution of vector quantities.

As we will see in the next subsection, it is possible to proceed in a more covariant way, starting with the covariant velocity \(\frac{D n^a}{D\tau }\), and arrive at a simpler covariant Hamiltonian (3.27) below. This, however, requires an additional care about technical details and before we explore such approach, let us first discuss the behavior of the coordinate Hamiltonian under a change of coordinates.

Evolving the parameter \(\tau \), there can come a moment when the geodesic leaves the given coordinate chart \(\{ x^a \}\). Then it is necessary to use a different set of coordinates

$$\begin{aligned} x^{\prime a} = x^{\prime a} (x^b) \, , \qquad J^a{}_b = \frac{{\partial }x^{\prime a}}{{\partial }x^b} \, . \end{aligned}$$
(3.20)

From the point of view of Hamiltonian dynamics, such a change of coordinates induces a time-dependent canonical transformation of \((n^a, \pi _a)\) variables, and the Hamiltonian in general changes. The non-invariance property is expressed by the non-tensorial term in \(\mathscr {h}_{\mathrm {c}}\). To be explicit, the change of coordinate chart (3.20) induces the following transformation:

$$\begin{aligned} \begin{aligned} n^{\prime a}&= J^a{}_b\, n^b \, , \\ \pi ^\prime _a&= \pi _b\, J^{-1}{}^b {}_a \, . \end{aligned} \end{aligned}$$
(3.21)

Such a transformation is canonical and amounts to the following generating function:

$$\begin{aligned} G(n,\pi ^\prime ) = \pi ^\prime _a \, J^a{}_b (\tau ) \, n^b \, . \end{aligned}$$
(3.22)

By this we mean that Eqs. (3.21) are obtained via solving

$$\begin{aligned} \begin{aligned} n^{\prime a}&= \frac{\partial G}{\partial \pi ^\prime _a}(n,\pi ^\prime ) \, , \\ \pi _a&= \frac{\partial G}{\partial n^a}(n,\pi ^\prime ) \end{aligned} \end{aligned}$$
(3.23)

for \(n^\prime \) and \(\pi ^\prime \). The standard theory then gives that the Hamiltonian transforms to

$$\begin{aligned} \begin{aligned} \mathscr {h}_{\mathrm {c}}^\prime (n^\prime , \pi ^\prime )&= \mathscr {h}_{\mathrm {c}}(n,\pi ) + \frac{\partial G}{\partial \tau }(n,\pi ^\prime )\\&=\frac{1}{2m} {\bar{g}}^{ab} \pi ^\prime _a \pi ^\prime _b - {\bar{u}}^{\prime k} {\bar{\Gamma }}^{\prime a}_{kb}\, \pi ^\prime _a n^{\prime b} \\&\quad + \frac{m}{2} {\bar{R}}^\prime _{abcd} {\bar{u}}^{\prime a} {\bar{u}}^{\prime c} n^{\prime b} n^{\prime d} \, , \end{aligned} \end{aligned}$$
(3.24)

where the term \(\frac{\partial G}{\partial \tau }\) provides the non-tensorial term in the transformation rule of the connection. The Hamiltonian is thus locally invariant in form, although it is not build from pure tensorial expressions. However, there is no such thing as a ‘global coordinate Hamiltonian’. For each coordinate chart there is a different Hamiltonian which governs time evolution in that chart. These Hamiltonians differ by \(\frac{\partial G}{\partial \tau }\) terms, and so they cannot be understood as just different coordinate expressions for one Hamiltonian function. The global motion has to be sewed from solutions in different charts. Nevertheless, the global covariant Hamiltonian can be defined in the covariant approach as we will show next.

3.5 Hamiltonian formalism: covariant approach

The linearized configuration space of vectors \(n^a\) is time dependent: it is a tangent space at \({\bar{x}}(\tau )\). In the discussion of the coordinate Hamiltonian above, we have implicitly identified such tangent spaces by choosing particular coordinates. Namely, we have naturally identified vectors with the same components with respect to the coordinate frame. More precisely, we regarded such vectors as ‘not changing’. We have used the coordinate-time derivative \({\dot{n}}^a= \frac{d n^a}{d\tau }(\tau )\) to define the velocity and employed it in the construction of the Hamiltonian. Since such an identification of vectors is non-covariant, we obtained a non-covariant Hamiltonian which was dependent explicitly on the Christoffel symbols.

However, we can identify the linearized configuration spaces at different times in a more covariant way: by a parallel transport along the central geodesic. For that, we use the covariant time derivative \(\frac{D n^a}{D\tau }\) to define the velocity.Footnote 3

Thus, in the covariant approach we must interpret the Lagrangian \(\mathscr {l}\) as a function of a \(\underline{\hbox {linearized position}\, n^a}\) and of a \(\underline{\hbox {covariant velocity} \, v^a = \frac{D n^a}{D\tau }}\),

$$\begin{aligned} \mathscr {l}(n,v) = \frac{1}{2}m\, {\bar{g}}_{ab} v^a v^b - \frac{1}{2} m\, {\bar{u}}^c {\bar{u}}^d \,{\bar{R}}_{cadb}\, n^a n^b \;. \end{aligned}$$
(3.25)

A covariant version of canonically conjugate momentum \(\pi _a\) then reads

$$\begin{aligned} \pi _a = \frac{\partial \mathscr {l}}{\partial v^a} = m\, {\bar{g}}_{ab} v^b\,, \end{aligned}$$
(3.26)

and the covariant Hamiltonian, \(\mathscr {h}= \pi _a v^a - \mathscr {l}\), is

$$\begin{aligned} \mathscr {h}(n,\pi ) = \frac{1}{2m}\, {\bar{g}}^{ab} \pi _a \pi _b + \frac{1}{2} m\, {\bar{u}}^c {\bar{u}}^d \,{\bar{R}}_{cadb}\, n^a n^b \;. \end{aligned}$$
(3.27)

The Hamilton equations have to be written again using the covariant time derivative

$$\begin{aligned} \begin{aligned} \frac{D n^a}{D\tau }&= \frac{\partial \mathscr {h}}{\partial \pi _a} = \frac{1}{m}\, {\bar{g}}^{ab}\pi _b\;,\\ \frac{D\pi _a}{D\tau }&= - \frac{\partial \mathscr {h}}{\partial n^a} = -m\, {\bar{u}}^c {\bar{u}}^d \,{\bar{R}}_{cadb}\, n^b \;. \end{aligned} \end{aligned}$$
(3.28)

Combining both equations together yields the Jacobi equation,

$$\begin{aligned} \frac{D^2 n^a}{D\tau ^2} + {\bar{R}}^{a}{}_{cbd}\,{\bar{u}}^c {\bar{u}}^d\, n^b =0\;. \end{aligned}$$
(3.29)

The evolution of a general phase space observable A expressed in new variables n, \(\pi \) is given by

$$\begin{aligned} \frac{d}{d\tau }{A} = \{A,\mathscr {h}\} + \frac{{\partial }A}{{\partial }\tau }\;, \end{aligned}$$
(3.30)

where the Poisson bracket assumes the standard form

$$\begin{aligned} \{A,B\} = \frac{{\partial }A}{{\partial }n^c}\frac{{\partial }B}{{\partial }\pi _c} - \frac{{\partial }A}{{\partial }\pi _c}\frac{{\partial }B}{{\partial }n^c}\;. \end{aligned}$$
(3.31)

Let us stress that even the observables, that were independent of the time parameter in the original variables \((x,\,p)\), typically become explicitly time dependent and the second term in (3.30) is non-trivial. The reason is that a ‘simple’ function of x and p is re-expressed in terms of the central trajectory \({\bar{x}}\), \({\bar{p}}\) and linearized variables n and \(\pi \). The explicit time dependency then enters through the time dependent central trajectory.

In particular, this is true for the covariant Hamiltonian itself, and we have

$$\begin{aligned} \frac{d}{d\tau }\mathscr {h}= \frac{{\partial }\mathscr {h}}{{\partial }\tau }\;, \end{aligned}$$
(3.32)

that is, the new Hamiltonian \(\mathscr {h}\) for the linearized system is not conserved. However, for the Hamiltonian, the time dependency is not caused only by the introduction of the linearized variables with respect to the central trajectory, but it is also related to a time-dependent canonical transformation that relates the original Hamiltonian H and its (linearized) version \(\mathscr {h}\), see Appendix 1 for more details.

4 Integrals of motion for the Jacobi system

In this section we will describe how the integrals of motion for geodesics give rise to the particularly simple integrals of motion for the Jacobi system. We start with a discussion applicable to any linearized dynamical system and only later specify to the case of the geodesic motion.

4.1 Wronskian as a linear integral of motion

Consider a linearized dynamical system, with trajectories near the central trajectory \({\bar{x}}\) described by a linearized trajectory \(n^a(\tau )\). A general feature of linearized systems is that their dynamics is governed by a quadratic Lagrangian \(\mathscr {l}(n,v)\), and, in the Hamiltonian picture, by a quadratic Hamiltonian \(\mathscr {h}(n,\pi )\), see (3.25) and (3.27) for the specific example of linearized geodesic motion.

For two linearized trajectories \(n_1^a(\tau )\) and \(n_2^a(\tau )\), we define the Wronskian as

$$\begin{aligned} W[n_1|n_2] = n_1^a \frac{{\partial }\mathscr {l}}{{\partial }v^a}\Bigl (n_2,\frac{D n_2}{D\tau }\Bigr ) - \frac{{\partial }\mathscr {l}}{{\partial }v^a}\Bigl (n_1,\frac{D n_1}{D\tau }\Bigr ) n_2^a\;.\nonumber \\ \end{aligned}$$
(4.1)

In the phase-space variables the trajectory is characterized by position \(n^a\) and momentum \(\pi _a\) and the Wronskian can be written as

$$\begin{aligned} W[n_1,\pi _1|n_2,\pi _2] = n_1^a \pi _{2a} - \pi _{1a} n_2^a\;. \end{aligned}$$
(4.2)

It is well known (see, e.g., [26, 27] for the case of the Jacobi system) that for any two solutions \(n_1^a(\tau )\) and \(n_2^a(\tau )\) of the equation of motion the Wronskian is conserved in time \(\tau \). To show this, let us use the Hamiltonian picture. By employing the general quadratic Hamiltonian,

$$\begin{aligned} \mathscr {h}(n,\pi ) = \frac{1}{2} \pi _a {\bar{K}}^{ab}\pi _b + \pi _a {\bar{A}}^a{}_b n^b + \frac{1}{2} n^a {\bar{U}}_{ab} n^b\;, \end{aligned}$$
(4.3)

we have the following Hamiltonian equations:

$$\begin{aligned} \begin{aligned} \frac{D n^a}{D\tau }&= \frac{\partial \mathscr {h}}{\partial \pi _a} = {\bar{K}}^{ab}\,\pi _b + {\bar{A}}^a{}_b\, n^b\;,\\ \frac{D\pi _a}{D\tau }&= - \frac{\partial \mathscr {h}}{\partial n^a} = -\, {\bar{U}}_{ab}\, n^b - \pi _b\, {\bar{A}}^b{}_a \;. \end{aligned} \end{aligned}$$
(4.4)

Taking the time derivative of (4.2) and substituting (4.4) for \(\frac{D n_1}{D\tau }\), \(\frac{D\pi _1}{D\tau }\) and \(\frac{D n_2}{D\tau }\), \(\frac{D\pi _2}{D\tau }\), we find

$$\begin{aligned} \frac{d}{d\tau }W[n_1,\pi _1|n_2,\pi _2] =0\,. \end{aligned}$$
(4.5)

This means that any fixed solution \({\tilde{n}}(\tau )\) generates a quantity

$$\begin{aligned} W_{{\tilde{n}}}(n) = W[{\tilde{n}}|n]\,, \end{aligned}$$
(4.6)

which is conserved along any solution \(n(\tau )\). In the phase-space language, any solution \({\tilde{n}}(\tau )\), \({\tilde{\pi }}(\tau )\) defines a conserved quantity

$$\begin{aligned} W_{{\tilde{n}},{\tilde{\pi }}}(n,\pi ) = W[{\tilde{n}},{\tilde{\pi }}|n,\pi ] = {\tilde{n}}^a \pi _{a} - {\tilde{\pi }}_{a} n^a\;. \end{aligned}$$
(4.7)

Clearly, such a conserved quantity is linear in n and \(\pi \).

This observation can be reversed: the most general linear conserved quantity of a linearized system is given by \(W_{{\tilde{n}},{\tilde{\pi }}}\), where \({\tilde{n}}(\tau )\), \({\tilde{\pi }}(\tau )\) are explicit solutions of the Hamilton equations. Indeed, let us consider a general linear observable \({C={\tilde{n}}^a \pi _{a} - {\tilde{\pi }}_{a} n^a}\) with yet unspecified coefficients \({\tilde{\pi }}_{a}(\tau )\) and \({\tilde{n}}^{a}(\tau )\). Its conservation means

$$\begin{aligned} \begin{aligned} 0&=\frac{d}{d\tau }C(n,\pi ) = \{C,\mathscr {h}\} + \frac{{\partial }C}{{\partial }\tau }\\&=-{\tilde{\pi }}_a\frac{{\partial }\mathscr {h}}{{\partial }\pi _a}(n,\pi ) - {\tilde{n}}^a \frac{{\partial }\mathscr {h}}{{\partial }n^a}(n,\pi ) \\&\quad + \frac{D{\tilde{n}}^a}{D\tau } \pi _a - \frac{D{\tilde{\pi }}_a}{D\tau } n^a \;. \end{aligned} \end{aligned}$$
(4.8)

Substituting (4.3), re-arranging terms and using (4.3) again, we get

$$\begin{aligned} 0= & {} \pi _a \Bigl (\frac{D{\tilde{n}}^a}{D\tau } -\frac{{\partial }\mathscr {h}}{{\partial }\pi _a}({\tilde{n}},{\tilde{\pi }})\Bigr ) \nonumber \\&-n^a \Bigl (\frac{D{\tilde{\pi }}_a}{D\tau } +\frac{{\partial }\mathscr {h}}{{\partial }n^a}({\tilde{n}},{\tilde{\pi }})\Bigr )\;. \end{aligned}$$
(4.9)

Since C should be conserved at any phase-space point \((n,\,\pi )\), we obtain that \({\tilde{n}}^a(\tau )\) and \({\tilde{\pi }}_a(\tau )\) must satisfy the Hamilton equations and, thus, the observable C has to be of the form \(C=W_{{\tilde{n}},{\tilde{\pi }}}\).

A similar statement can be obviously formulated in the configuration language: any conserved quantity linear in the trajectory \(n(\tau )\) and its time derivative has to have the form (4.6) for a solution \({\tilde{n}}(\tau )\).

Finally, thanks to the linear structure of the linearized system, the Wronskian of two solutions \(n_1(\tau )\), \(n_2(\tau )\) can be related to the Poisson bracket of the corresponding conserved quantities \(W_{n_1,\pi _1}\) and \(W_{n_2,\pi _2}\),

$$\begin{aligned} \{W_{n_1,\pi _1},W_{n_2,\pi _2}\} = W[n_1,\pi _1|n_2,\pi _2] \;. \end{aligned}$$
(4.10)

Let us finally return back to the geodesic motion and the corresponding linearized Jacobi system. In this case the Wronskian (4.1) takes the particular form

$$\begin{aligned} W[n_1|n_2] = m\Bigl ( n_1^a {\bar{g}}_{ab} \frac{D n_2^b}{D\tau } - \frac{D n_1^a}{D\tau } {\bar{g}}_{ab} n_2^b\Bigr )\;, \end{aligned}$$
(4.11)

and the above formulae directly apply. In what follows we concentrate on this case.

4.2 Canonical observables

It is easy to see that the Poisson bracket of two observables linear in \((n,\pi )\) is equal to a constant, i.e., an observable independent of \((n,\pi )\). This observation can be used to construct a set of (time dependent) observables for the Jacobi system \((F^j,\,G_j)\), \(j=1,\ldots ,n\), which form canonical coordinates at all times,

$$\begin{aligned} \{F^i,G_j\} = \delta _j^i\;,\quad \{F^i,F^j\} = \{G_i,G_j\} = 0\;. \end{aligned}$$
(4.12)

For example, choosing at time \(\tau _0\) an orthonormal frame of vectors \(e_{(i)}\) and the dual frame of 1-forms \(e^{(i)}\), both at \({\bar{x}}(\tau _0)\), one can define the solutions of the Jacobi system \(f^j(\tau )\) and \(g_j(\tau )\) with initial values at \(\tau _0\) given by

$$\begin{aligned} \begin{aligned} f^j(\tau _0)&= e_{(j)}\;,&\frac{D f^j}{D\tau }(\tau _0)&= 0 \;,\\ g_j(\tau _0)&= 0\;,&\frac{D g_j}{D\tau }(\tau _0)&= e^{(j)} \;. \end{aligned} \end{aligned}$$
(4.13)

Clearly,

$$\begin{aligned} W[f^i,g_j]= \delta _j^i\;,\quad W[f^i,f^j]=W[g_i,g_j]=0 \end{aligned}$$
(4.14)

at time \(\tau _0\), and since the Wronskian is conserved, the relation (4.14) remains true at all times. Using this solution, and thanks to (4.10), we can thus define canonical coordinates \((F^j,G_i)\) satisfying (4.12) by the corresponding Wronskian observables

$$\begin{aligned} F^j = W_{f^j}\;,\quad G_i = W_{g_i}\;. \end{aligned}$$
(4.15)

In particular, the set of coordinates \(F^j\) (as well as the set of \(G_i\)) forms a maximal set of commuting conserved quantities of the linearized system. However, these conserved quantities are not very useful. To find them, one has to find first the solutions \(f^j\) and \(g_j\), i.e., to solve the linearized system.

In the following, we want to discuss more useful conserved quantities – given by the symmetries of the spacetime. To find these observables, in addition to symmetries one only needs to know the central trajectory.

4.3 Conserved quantities generated by Killing tensors

As we reviewed in Sect. 2, generic (homogeneous in momentum) integrals of geodesic motion are generated by Killing tensors, see (2.8). As shown by Caviglia, Zordan and Salmistraro (CZS) [26, 33] these tensors also generate the following linearized solutions \({\tilde{n}}^a(\tau )\), \({\tilde{\pi }}_a(\tau )\) for the Jacobi system:

$$\begin{aligned} {\tilde{n}}^a&= \frac{{\partial }K}{{\partial }p_a}({\bar{x}},{\bar{p}}) = \frac{1}{(r{-}1)!} K^{a b_2 \ldots b_{r}}({\bar{x}})\, {\bar{p}}_{b_2} \ldots {\bar{p}}_{b_{r}} \, ,\nonumber \\ {\tilde{\pi }}_a&= -\frac{\nabla _{a} K}{{\partial }x}({\bar{x}},{\bar{p}}) = - \frac{1}{r!} \nabla _{a}K^{b_1 \ldots b_{r}}({\bar{x}})\, {\bar{p}}_{b_1} \ldots {\bar{p}}_{b_{r}} \, . \end{aligned}$$
(4.16)

Here, K is the conserved quantity of geodesic motion generated by the Killing tensor \(K^{a_1\ldots a_r}\), (2.8), \(\frac{{\partial }K}{{\partial }p}\) denotes the derivative with respect to momentum p with x fixed, and \(n^a \frac{\nabla _{a} K}{{\partial }x}(x,p)\) is the covariant derivative in direction \(n^a\) with p parallelly transported, cf. (A21) in Appendix 1. The latter derivative acts only on x-dependent terms in K and essentially ignores momentum p.

Let us verify the solution (4.16) by checking the Hamilton equations (3.28). Taking advantage of the fact that the central trajectory is geodesic, \(\frac{D{\bar{p}}_a}{D\tau }=0\), and \({\bar{p}}_a = m {\bar{u}}_a\), we obtain

$$\begin{aligned} \frac{D {\tilde{n}}^a}{D\tau } = \frac{1}{m}\frac{1}{(r{-}1)!} {\bar{p}}_{b_1} \nabla ^{b_1} K^{a b_2 \ldots b_{r}}({\bar{x}})\, {\bar{p}}_{b_2} \ldots {\bar{p}}_{b_{r}}\,. \end{aligned}$$
(4.17)

Using the identity

$$\begin{aligned} \nabla ^a K^{b_1 \ldots b_{r}} = - r \nabla ^{(b_1} K^{|a|b_2 \ldots b_{r})}\;, \end{aligned}$$
(4.18)

which follows from the Killing condition (2.9), yields

$$\begin{aligned} \begin{aligned} \frac{D {\tilde{n}}^a}{D\tau }&= - \frac{1}{m}\frac{1}{r!} \nabla ^{a} K^{b_1 \ldots b_{r}}({\bar{x}})\, {\bar{p}}_{b_1} \ldots {\bar{p}}_{b_{r}} \\&= -\frac{1}{m} \frac{\nabla ^{a} K}{{\partial }x} = \frac{1}{m} {\bar{g}}^{ab} {\tilde{\pi }}_b\;. \end{aligned} \end{aligned}$$
(4.19)

For momentum \({\tilde{\pi }}_a\) we get

$$\begin{aligned} \begin{aligned} \frac{D{\tilde{\pi }}_a}{D \tau }&= -\frac{1}{r!} {\bar{u}}^{c}\nabla _{c}\nabla _{a}K^{b_1 \ldots b_{r}}({\bar{x}})\, {\bar{p}}_{b_1} \ldots {\bar{p}}_{b_{r}} \\&= -\frac{1}{m}\frac{1}{r!}\nabla _{a}\nabla ^{(b_0}K^{b_1\ldots b_r)}({\bar{x}})\, {\bar{p}}_{b_0} \ldots {\bar{p}}_{b_{r}} \\&\quad -\frac{1}{(r{-}1)!} {\bar{u}}^c {\bar{R}}_{ca}{}^{b_1}{}_{b}K^{bb_2\ldots b_r}({\bar{x}})\, {\bar{p}}_{b_1} \ldots {\bar{p}}_{b_{r}} \, . \end{aligned} \end{aligned}$$
(4.20)

Here, we have used the Ricci identity and the fact that all r terms with the Riemann tensor give the same contribution. Now, the first term vanishes thanks to (2.9) and in the second term we can identify \({\tilde{n}}^a\),

$$\begin{aligned} \begin{aligned} \frac{D{\tilde{\pi }}_a}{D \tau }&= -m {\bar{u}}^c{\bar{u}}^d {\bar{R}}_{cadb}\frac{1}{(r{-}1)!}K^{bb_2\ldots b_r}({\bar{x}})\, {\bar{p}}_{b_2} \ldots {\bar{p}}_{b_{r}}\\&=-m {\bar{u}}^c{\bar{u}}^d {\bar{R}}_{cadb}{\tilde{n}}^b \, , \end{aligned} \end{aligned}$$
(4.21)

which concludes the proof.

The CZS solution (4.16) is of geometrical nature. It is obtained from the canonical transformation associated with the hidden symmetry of the geodesic equation. Let \(\{\cdot ,\cdot \}_g\) be the Poisson bracket of the full geodesic theory, cf. (A22). Then \({\tilde{n}}^a = \{ x^a , K \}_g\) and \({\tilde{\pi }}_a=\{p_a,K\}_g\). This is the same as the infinitesimal transformation \(\delta x^a\), \(\delta p_a\) of the central geodesic generated by the canonical transformation induced by K.

The solution (4.16) can be related to the linearization \({\mathscr {k}}\) of the conserved quantity K. The expansion of any phase-space observable K(xp) to the first order can be written as

$$\begin{aligned} K = {\bar{K}} + n^a \frac{\nabla _{a}K}{{\partial }x} + \pi _a \frac{{\partial }K}{{\partial }p_a}+\ldots \;. \end{aligned}$$
(4.22)

Obviously, using (4.16) we get

$$\begin{aligned} \begin{aligned} {\mathscr {k}} \equiv K- {\bar{K}} = {\tilde{n}}^a \pi _a - {\tilde{\pi }}_a n^a= W_{{\tilde{n}},{\tilde{\pi }}}\;. \end{aligned} \end{aligned}$$
(4.23)

The linearized observable \({\mathscr {k}}=W_{{\tilde{n}},{\tilde{\pi }}}\) is thus again a conserved quantity that is linear in position and momentum and generated by the CZS solution \(({\tilde{n}},\,{\tilde{\pi }})\).

Let us finally mention that the CZS solution for \({\tilde{n}}^a\), (4.16), need not be generated from a Killing tensor. As noted in [26, 33] (see also [34, 35]), its existence is in one-to-one correspondence with a new object, called the affine tensor. An affine tensor of rank r, \(K^{a_1\ldots a_r}=K^{(a_1\ldots a_r)}\), is an object that satisfies

$$\begin{aligned} \nabla _{(a}K_{a_1\ldots a_r)}=h_{aa_1\ldots a_r}\,,\quad \nabla _{b} h_{aa_1\ldots a_r}=0\,. \end{aligned}$$
(4.24)

That is, the definition of an affine tensor ‘generalizes’ that of a Killing tensor by requiring that its symmetrized derivative need not vanish but can be a covariantly constant tensor. Of course, the above presented construction of conserved quantities through Wronskians immediately generalizes to the CZS solutions generated by affine tensors. Let us stress, however, that although the Killing tensors are formally a subfamily of affine tensors, the requirement on the existence of non-trivial \(h_{aa_1\ldots a_r}\) is very strong and at the moment there are no known physical spacetimes admitting affine tensors that are not at the same time Killing tensors. For this reason we shall not probe this possibility in this paper any further.

4.4 Integrability of the linearized system

Let us now assume that we have at least two Killing tensors \(K_1^{a\ldots }\) and \(K_2^{a\ldots }\) corresponding to the conserved quantities \(K_1\) and \(K_2\) of the full geodesic motion. In general, such integrals of motion do not Poisson-commute. Their Poisson bracket generates a new conserved quantity K,

$$\begin{aligned} K = \{K_1, K_2\}_g\;, \end{aligned}$$
(4.25)

which corresponds also to a Killing tensor \(K^{a\ldots }\), given by the (symmetric) Schouten–Nijenhuis bracket [36,37,38]

$$\begin{aligned} K = [ K_1, K_2 ]_{\scriptscriptstyle \mathrm {SN}}\;, \end{aligned}$$
(4.26)

cf., e.g., [39].

For the linearized quantities \({\mathscr {k}}_1\), \({\mathscr {k}}_2\) we have

$$\begin{aligned} \{{\mathscr {k}}_1,{\mathscr {k}}_2\} = \{W_{{\tilde{n}}_1,{\tilde{\pi }}_1},W_{{\tilde{n}}_2,{\tilde{\pi }}_2}\} = W[{\tilde{n}}_1,{\tilde{\pi }}_1|{\tilde{n}}_2,{\tilde{\pi }}_2]\;. \end{aligned}$$
(4.27)

The Wronskian can be expressed using the quantities related to the central trajectory. Substituting (4.2) and (4.16), we get

$$\begin{aligned} \begin{aligned} \{{\mathscr {k}}_1,{\mathscr {k}}_2\}&=\Bigl (-\frac{{\partial }K_1}{{\partial }p_a}\frac{\nabla _{a}K_2}{{\partial }x} + \frac{\nabla _{a}K_1}{{\partial }x} \frac{{\partial }K_2}{{\partial }p_a}\Bigr )\bigg |_{{\bar{x}},{\bar{p}}}\\&=\{K_1,K_2\}_g\big |_{{\bar{x}},{\bar{p}}}={\bar{K}}\;, \end{aligned} \end{aligned}$$
(4.28)

the result already shown in [33].

It follows that if the original integrals of motion \(K_1\), \(K_2\) Poisson-commute, \(K=0\), the linearized conserved quantities \({\mathscr {k}}_1\), \({\mathscr {k}}_2\) also Poisson-commute. In particular, if the spacetime geometry possesses a full set of commuting integrals of motion \(K_j\) generated by Killing tensors, the linearized system has also a full set of mutually commuting integrals of motion \({\mathscr {k}}_j\equiv W_{{\tilde{n}}_j,{\tilde{\pi }}_j}\). The complete integrability of the geodesic motion thus naturally implies the complete integrability of the Jacobi system.

5 Integrability of Jacobi equation in rotating black hole spacetimes

Let us now apply the above developed formalism to explicitly demonstrate the integrability of the Jacobi equation in the Kerr black hole spacetime [40] and its higher-dimensional generalizations. Such an integrability stems from the existence of hidden symmetries in these spacetimes and derives from the integrability of the full geodesic motion. The same results remain also true for charged black holes and will be discussed elsewhere.

The Kerr metric represents a unique rotating black hole solution of vacuum Einstein equations. In the Boyer–Lindquist coordinates it reads

$$\begin{aligned} ds^2= & {} - \frac{\Delta }{\Sigma } \left( dt - a \sin ^2\theta d\varphi \right) ^2 + \frac{\sin ^2 \theta }{\Sigma } \left[ (r^2 + a^2) d\varphi - a dt \right] ^2 \nonumber \\&+ \frac{\Sigma }{\Delta } dr^2 + \Sigma \, d\theta ^2 \, , \end{aligned}$$
(5.1)

where \(\Sigma = r^2 + a^2 \cos ^2\theta \), and \(\Delta = r^2 - 2Mr + a^2\).

The metric admits two Killing vectors \({\partial }_t\) and \({\partial }_\varphi \). For a geodesic motion, these imply the following conserved charges:

$$\begin{aligned} K_E&= - p_t \, , \end{aligned}$$
(5.2)
$$\begin{aligned} K_L&= p_\varphi \, . \end{aligned}$$
(5.3)

In addition, there is a hidden symmetry encoded in the Killing tensor \(K_{ab}\) that gives rise to Carter’s constant [31, 32, 41]

$$\begin{aligned} K_{C}= & {} \frac{1}{2} K^{ab} p_a p_b = \frac{1}{2\Sigma } \left[ - \Delta a^2 \cos ^2\theta \, p_r^2 + r^2 p_\theta ^2 \right. \nonumber \\&\left. + \frac{a^2 \cos ^2\theta }{\Delta } \left( (r^2+a^2)p_t + a p_\varphi \right) ^2 \right. \nonumber \\&\left. + \frac{r^2}{\sin ^2\theta } \left( a \sin ^2\theta \, p_t + p_\varphi \right) ^2 \right] \,. \end{aligned}$$
(5.4)

Lastly there is a conserved quantity generated by the metric \(g_{ab}\), seen as a (covariantly constant) Killing tensor,

$$\begin{aligned} K_{m^2}= & {} - g^{ab} p_a p_b = - \frac{1}{\Sigma } \left[ \Delta p_r^2 + p_\theta ^2 + \frac{\Delta - a^2 \sin ^2\theta }{\Delta \sin ^2\theta } p_\varphi ^2 \right. \nonumber \\&\left. - \frac{4 M a r}{\Delta } p_t p_\varphi - \frac{(r^2+a^2)^2 - a^2 \Delta \sin ^2\theta }{\Delta } p_t^2 \right] \,.\quad \nonumber \\ \end{aligned}$$
(5.5)

The four integrals of motion, \(\{K_E, K_L, K_C, K_{m^2}\}\), are functionally independent and mutually Poisson commute, yielding the geodesic motion completely integrable [41]. The explicit solution in terms of special functions can be for example found in [42, 43].

We can now pick our favourite geodesic and turn to the corresponding linearized Jacobi system. The CZS solution, (4.16), yields the following independent solutions:

$$\begin{aligned} {\tilde{n}}_E= & {} - {\partial }_t \, , \end{aligned}$$
(5.6)
$$\begin{aligned} {\tilde{n}}_L= & {} {\partial }_\varphi \, , \end{aligned}$$
(5.7)
$$\begin{aligned} \Sigma {\tilde{n}}_{C}= & {} - \Delta a^2 \cos ^2{\bar{\theta }} {\bar{p}}_r {\partial }_r + {\bar{r}}^2 {\bar{p}}_\theta {\partial }_\theta \nonumber \\&+ \left( \frac{{\bar{p}}_\varphi }{\sin ^2{\bar{\theta }}} + a {\bar{p}}_t \right) {\bar{r}}^2 {\partial }_\varphi \nonumber \\&+ \frac{a^3 \cos ^2{\bar{\theta }} \left( a {\bar{p}}_\varphi + {\bar{p}}_t ({\bar{r}}^2 + a^2)\right) }{\Delta } {\partial }_\varphi \nonumber \\&+ \frac{a^2 \cos ^2{\bar{\theta }} ({\bar{r}}^2 + a^2)\left( a {\bar{p}}_\varphi + ({\bar{r}}^2+a^2){\bar{p}}_t \right) }{\Delta } {\partial }_t \nonumber \\&+ a {\bar{r}}^2 ( {\bar{p}}_\varphi + a \sin ^2{\bar{\theta }} {\bar{p}}_t) {\partial }_t \, , \end{aligned}$$
(5.8)
$$\begin{aligned} - \frac{\Sigma }{2} {\tilde{n}}_{m^2}= & {} \Delta {\bar{p}}_r {\partial }_r + {\bar{p}}_\theta {\partial }_\theta \nonumber \\&+ \left( \frac{\Delta - a^2 \sin ^2{\bar{\theta }}}{\Delta \sin ^2{\bar{\theta }}} {\bar{p}}_\varphi - \frac{2 M a r}{\Delta } {\bar{p}}_t \right) {\partial }_\varphi \nonumber \\&- \left( \frac{({\bar{r}}^2+a^2)^2 - a^2 \Delta \sin ^2{\bar{\theta }}}{\Delta } {\bar{p}}_t + \frac{2 M a r}{\Delta } {\bar{p}}_\varphi \right) {\partial }_t\,.\nonumber \\ \end{aligned}$$
(5.9)

From these we construct the independent conserved quantities for the Jacobi equation in Kerr

$$\begin{aligned} {\mathscr {k}}_i = {\tilde{n}}_{i}^a \pi _a - n^a \frac{D{\tilde{n}}_{i a}}{D\tau } \, , \qquad i = 1, \ldots , 4 \, . \end{aligned}$$
(5.10)

We set the constant values of the Wronskians to \(w_i\), \(i=1, \ldots , 4\), and introduce the abbreviated notation \(g_i = n^a \frac{D{\tilde{n}}_{i a}}{D\tau }\), where the \(g_i\) do not depend on the momenta \(\pi \). From (5.10) it is easy to extract the value of the momenta \(\pi _t\) and \(\pi _\varphi \), which are given by

$$\begin{aligned} \pi _t= & {} \pi _t ({\bar{x}}^a, n^a) = - g_1 - w_1 \, , \end{aligned}$$
(5.11)
$$\begin{aligned} \pi _\varphi= & {} \pi _\varphi ({\bar{t}},{\bar{r}}, {\bar{\theta }}, {\bar{\varphi }}) = g_2 + w_2 \, . \end{aligned}$$
(5.12)

For geodesics with \(p_r \ne 0\), \(p_\theta \ne 0\) it is possible to invert (5.10) with respect to \(\pi _r\), \(\pi _\theta \). The result is

$$\begin{aligned} \pi _r= & {} \alpha _r + \beta _r \pi _\varphi + \gamma _r \pi _t \, , \end{aligned}$$
(5.13)
$$\begin{aligned} \pi _\theta= & {} \alpha _\theta + \beta _\theta \pi _\varphi + \gamma _\theta \pi _t \, , \end{aligned}$$
(5.14)

with

$$\begin{aligned} \alpha _r= & {} - \frac{{\bar{r}}^2 f_3 + 2 f_4}{2 \Delta p_r} \, , \end{aligned}$$
(5.15)
$$\begin{aligned} \beta _r= & {} \frac{a}{\Delta ^2 p_r} \left[ (a^2+{\bar{r}}^2) p_t + a p_\varphi \right] \, , \end{aligned}$$
(5.16)
$$\begin{aligned} \gamma _r= & {} \frac{a^2+{\bar{r}}^2}{a} \beta _r \, , \end{aligned}$$
(5.17)
$$\begin{aligned} \alpha _\theta= & {} \frac{1}{2 p_\theta } \left( - a^2 \cos ^2{\bar{\theta }} f_3 + 2 f_4 \right) \, , \end{aligned}$$
(5.18)
$$\begin{aligned} \beta _\theta= & {} - \frac{1}{p_\theta } \left[ a p_t + \frac{1}{\sin ^2{\bar{\theta }}} p_\varphi \right] \, , \end{aligned}$$
(5.19)
$$\begin{aligned} \gamma _\theta= & {} a \sin ^2{\bar{\theta }} \beta _\theta \,. \end{aligned}$$
(5.20)

Here we have set \(f_3 = w_3 + g_3\), \(f_4 = w_4 + g_4\), these are functions of \(n^a\) and not of the momenta. These expressions are involved, although in a closed form: it is a good example of the fact that the Jacobi equation is complicated even if it is linear.

Let us finally mention that the procedure described in this section directly generalizes to higher-dimensional Kerr-NUT-(A)dS black hole spacetimes [44]. Such spacetimes are known to admit a number of hidden symmetries that yield the geodesic motion completely integrable [45, 46]. It follows that the corresponding Jacobi system is also integrable and in principle can be solved by the same steps described in this section. Let us, however, stress that in higher dimensions, the generic geodesic is given only in terms of complicated integrals [39], see also [47,48,49] for special cases, and the solution of the Jacobi system thus becomes far from explicit.

6 Conclusions

In this paper we have analyzed the Jacobi geodesic deviation equation from a point of view of Hamiltonian dynamics. It represents a dynamical system that is (although linear) explicitly time dependent. Consequently, the coordinate Hamiltonian is not covariant and varies from chart to chart. Nevertheless, we have shown that a covariant Hamiltonian can be constructed and shown (see Appendix 1) how it can be obtained by the canonical transformation (accompanied by a due linearization) of the geodesic Hamiltonian. Although the geodesic Hamiltonian is a constant of motion, the linearized Hamiltonian for the geodesic deviation depends explicitly on time.

The main result of our paper regards the observation that the integrals of geodesic motion give rise to the corresponding integrals for the Jacobi system that are linear and given by the invariant Wronskians. In particular, this is true for the integrals generated by hidden symmetries of the spacetime. We have shown that if the geodesic motion is completely integrable, so will be the corresponding linearized motion described by the Jacobi equation, see [29] for an alternative demonstration of this result. This has been further illustrated on an example of rotating black hole spacetimes in four and higher dimensions.

There is a number of topics that we have not discussed and that we point out as suitable for future research. One of these is the inclusion of spin. For example, a Lagrangian for the Jacobi deviation in the presence of Grassmannian spin variables can be found in [8], and a discussion of conserved quantities for geodesics in the presence of spin in Kerr–NUT–(A)dS spaces in [50]. Another one is the possibility of extending our results to higher order geodesic perturbations. These have been used to build analytic approximations of generic geodesics from simple exact solutions [13, 14], and have been used to model extreme mass-ratio systems [15]. It would be interesting to find out if for example the standard conserved charges of Kerr can be used to build conserved charges for higher order geodesic perturbations. Lastly our results can be used to discuss the issue of (in)stability of dynamical systems [16, 17].