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

Symplectic mechanics of relativistic spinning compact bodies I.:
Covariant foundations and integrability around black holes

Paul Ramond 0000-0001-7123-0039 [email protected] IMCCE, Observatoire de Paris , Université PSL, 77 Avenue Denfert-Rochereau, FR-75014 Paris
Abstract

In general relativity, the motion of an extended body moving in a given spacetime can be described by a particle on a (generally non-geodesic) worldline. In first approximation, this worldline is a geodesic of the underlying spacetime, and the resulting dynamics admit a covariant and 4-dimensional Hamiltonian formulation. In the case of a Kerr background spacetime, the Hamiltonian was shown to be integrable by B. Carter and the now eponymous constant. At the next level of approximation, the particle possesses proper rotation (hereafter spin), which couples the curvature of spacetime and drives the representative worldline away from geodesics. In this article, we lay the theoretical foundations of a series of works aiming at exploiting the Hamiltonian nature of the equations governing the motion of a spinning particle, at linear order in spin. Our formalism is covariant and 10-dimensional. It handles the degeneracies inherent to the local Lorentz invariance of general relativity with tools from Poisson geometry, and accounts for the center-of-mass/spin-supplementary-condition using constrained Hamiltonian system theory. As a first application, we consider the linear-in-spin motion in a Kerr background. We show that the resulting Hamiltonian system admits exactly five functionally independent integrals of motion related to Killing symmetries, thereby proving that the system is integrable. We conclude that linear-in-spin corrections to the geodesic motion do not break integrability, and that the resulting trajectories are not chaotic. We explain how this integrability feature can be used to reduce the computational cost of waveform generation schemes for asymmetric binary systems of compact objects.

Introduction

.1 The relativistic motion of test bodies

The last decades have seen both the advent of gravitational wave astronomy B. P. Abbott et al. (2016) (LIGO Scientific Collaboration and Virgo Collaboration); B. P. Abbott et al. (2019) (LIGO Scientific Collaboration and Virgo Collaboration); B. P. Abbott et al. (2020) (LIGO Scientific Collaboration and Virgo Collaboration); Abbott et al. (2021) and the first direct observations of a black hole’s strong field region Akiyama et al. (2019, 2022). These revolutionary events come as a reward for decades of effort in understanding the family of compact objects, namely black holes, neutron stars, and possibly exotic stars Cardoso et al. (2016); Toubiana et al. (2021). Of particular interest are binary systems of compact objects, which have been at the core of many still ongoing research programs, in view of the future of gravitational wave detectors on Earth Unnikrishnan (2013); Punturo et al. (2010) or in space Amaro-Seoane et al. (2017). Our ability to predict with great accuracy the motion of such systems directly impacts the likeliness of detecting them through their gravitational-wave emission. Thankfully, when it comes to the motion of compact binary systems under their mutual gravitational attraction, not all of their physical properties influence their motion the same way. This “universality” feature, unique to gravitation, is the main reason why multipolar expansions work: one models a real, extended compact object as a point particle endowed with a finite number of multipoles, e.g, mass, linear and angular momentum, deformability coefficients, etc. The higher the number of multipoles, the finer (and the closer to reality) the description of the object. The motion of the body is then obtained by solving a coupled problem: an equation of motion determining the wordline of the particle representing the body, and some evolution equations for the particle’s multipoles along that worldline. This very general result holds for any sufficiently compact object with a conserved stress-energy-momentum tensor Dixon (1974); Harte (2015).

For a small compact object moving in a given, background spacetime, the foundations of the multipolar schemes were laid in 1937 by Miron Mathisson. Key contributions were then made by Papapetrou in the forties, Tulzyjew in the fifties and culminated in the sixties with a series of articles by Dixon, essentially clarifying previous works and giving the “big picture.” We refer to Chap. 2 of Ramond (2021) for a detailed historical account and references for these pioneering works. The most modern and enlightening framework is undoubtedly that based on “generalized Killing fields”, developed by Harte Harte (2012, 2015, 2020). This approach sheds light on many uncovered aspects of previous derivations, makes explicit the difference between kinematical and dynamical effects, and is able to account for the self-fields of the body. The relevant evolution equations, now known as the Mathisson-Papapetrou-Tulczyjew-Dixon (MPTD) equations, are simple manifestations of the attempt to maintain Poincaré invariance as much as possible along the body’s representative worldline Harte (2008). These equations have been thoroughly studied and solved for a plethora of background spacetimes both theoretically and numerically, in particular exact and hairy black holes (see references below) and cosmological metrics Harte (2007); Obukhov and Puetzfeld (2011); Zalaquett et al. (2014). The MPTD equations are also recovered in the context of black hole perturbation theory Pound (2015); Miller and Pound (2021), making them not just “effective” equations replacing the extended body’s true motion, but a real sub-part of Einstein’s field equations.

.2 Relativistic Hamiltonian formulation

When the multipole description of the orbiting body is truncated at dipolar order, the MPTD system possesses two equations: one evolves a linear momentum 1-form pap_{a} and the other an antisymmetric spin tensor SabS^{ab}. Since the MPTD equations describes a purely kinematical evolution Harte (2015) of pa,Sabp_{a},S^{ab}, it is natural to look for a Hamiltonian formulation111By “Hamiltonian formulation”, we mean a system of ordinary differential equations generated by a scalar function and some geometric structure on a phase space. We shall be more precise when necessary. of these equations. In the case of general relativity, two distinct approaches have been proposed in the past. One of them is based on extending, to curved spacetime, the Lagrangian formulation of a free spinning particle in flat spacetime Hanson and Regge (1974). Inspired by this work and similar development in effective field theory Porto (2006), the authors of Barausse et al. (2009) provided a Hamiltonian formulation of the dipolar MPTD system via a singular Legendre transformation of this Lagrangian. Extensions to quadrupolar Vines et al. (2016) and octupolar orders Marsat (2015) have been derived, allowing for the SSC to be understood as a gauge freedom Steinhoff (2011). One drawback of these Lagrangian formulations, however, is their non-covariance, i.e., they require (i) a particular 3+1 split of the background spacetime and (ii) a fine-tuned choice of spin supplementary condition with no clear physical interpretation.

The other approach can be traced back to Souriau’s symplectic approach Souriau (1970) to the problem, cf. the recent note on Souriau’s pioneering ideas Damour (2024). While our treatment does not follow directly Souriau’s, it is solely based on symplectic geometry too, and borrow Souriau’s general philosophy: the Hamiltonian system will be considered the prime ingredient, and not a mere Legendre transform of some Lagrangian. The general symplectic structure that we shall use is already present (but somewhat hidden) in Souriau’s Souriau (1970) and Kunzle’s work Künzle (1972), but it seems that its first clear appearance is in222However, in d’Ambrosi et al. (2015) it is claimed that these brackets are symplectic, although they are not. d’Ambrosi et al. (2015) (see also references Witzany et al. (2019)). Naturally, our formulation will share some similarities with other covariant approaches proposed in d’Ambrosi et al. (2015, 2016); Witzany et al. (2019); Witzany (2019). However, the works d’Ambrosi et al. (2015, 2016) only work at the level of the equations of motion for simple orbits and do not exploit the Hamiltonian nature of the system. Similarly, the Hamiltonians proposed in Witzany et al. (2019); Witzany (2019) have a number of drawbacks that are unsatisfactory for practical applications, in particular ad-hoc mass parameters too many phase space dimensions, both issues related to unresolved constraints.Lastly, the necessary choice of spin supplementary condition for the well-posedness of the MPTD system will lead us, inevitably, to algebraic constraints. In the Lagrangian formalism, this is usually treated using the Dirac-Bergmann algorithm Barausse et al. (2009); Brown (2022). Here, in the context of symplectic geometry, this will be handled within the classical theory of “constrained Hamiltonian systems” following classical textbooks Arnold et al. (2006) (or Deriglazov (2022) for detailed proofs).

.3 The question of integrability around black holes

At monopolar order, the spin tensor SabS^{ab} is not involved in the description of the body. The two MPTD equations reduce to (i) pap^{a} is parallel-transported along the worldline and (ii) pap^{a} is tangent to the worldline. Combining the two implies that the worldline is a geodesic of the background spacetime. The equations of geodesic motion are the most general setup to describe “free systems”, as was first realized by Arnold in Arnold (1966).Owing to their purely free (i.e., purely kinematical) evolution, such geodesic systems are usually simple enough that they can sometimes be integrable, i.e., possess enough integrals of motion that make the solution to their equations of motion solvable by quadrature (i.e., using algebraic manipulations and standard calculus). In general relativity, the most remarkable example of this is the integrability of geodesics in the Kerr spacetime, first demonstrated by Carter Carter (1968), five years after Kerr’s discovery of the eponymous metric describing rotating black holes Kerr (1963). Integrability of Kerr geodesics is far from trivial: there are not enough spacetime symmetries to produce enough constants of motion to ensure integrability. Interestingly, Carter’s proof actually relied on pure symplectic mechanics of the geodesic Hamiltonian, unaware of the existence of the existence of the Killing-Stäckel tensor while which only appeared two years later Walker and Penrose (1970).

While the integrability of the monopolar MPTD system (i.e., geodesics) around black holes is well understood Carter (1968); Frolov et al. (2017), its extension to the dipolar MPTD system (i.e., the motion of spinning particles) is not so clear. Based on existing numerical simulations exhibiting artefacts classically associated to chaos, it is reasonable to believe that the integrability of geodesics is generally broken by the effect of the body’s spin, in the Schwarzschild case Bombelli and Calzetta (1992); Verhaaren and Hirschmann (2010); Zelenka et al. (2020) as well as in Kerr Suzuki and Maeda (1997); Semerák (1999); Kyrian and Semerák (2007); Han (2008); Compère and Druart (2022). This seems to be the case for other types of (non-spin) perturbations, as explored in Witzany et al. (2015); Polcar et al. (2022); Mosna et al. (2022). However, when limiting to spin effect that are linear in SabS^{ab}, no clear answer exists in the literature. Our results fill this gap with an analytical proof of integrability (thus, no chaos arises at linear-in-spin order).

.4 Motivation and main results

In this work, our primary motivation lies in the fact that integrability (or lack thereof) for the motion of a small compact object around a black hole is directly related to the possibility of using gravitational self-force theory to build gravitational waveforms for extreme mass ratio inspiral (EMRI) systems Barack and Pound (2018); Pound and Wardell (2021), one of the primary sources of gravitational waves for Laser Interferometer Space Antenna (LISA) Amaro-Seoane et al. (2017). The Hamiltonian formulation of the equations governing EMRIs is currently of major interest in this particular field because it can account for most of the non-geodesic effects that require implementing to meet LISA’s accuracy requirements Afshordi et al. (2023). This includes the small body’s spin Boyle et al. (2008); Vines et al. (2016); Witzany et al. (2019); Lukes-Gerakopoulos and Witzany (2021), the conservative part of the gravitational self-force Vines and Flanagan (2015a); Fujita et al. (2017); Blanco and Flanagan (2023a) and possibly its dissipative sector too Galley (2013), while also being adapted to efficient and consistent numerical resolution schemes such as symplectic integrators Haier et al. (2006); Wang et al. (2021). In addition, if the Hamiltonian system is integrable, it can be written in terms of action-angle variables, which are Taylor-made for two-timescale expansion methods Hinderer and Flanagan (2008), gauge-invariant analysis of resonances Hinderer and Flanagan (2008), an elementary derivation of the first law of mechanics Isoyama et al. (2014); Le Tiec (2015) as well as flux-balance laws Isoyama et al. (2018); Grant and Moxon (2023a), all of which are invaluable tools used in the modeling and understanding of binary system mechanics. Spinning degrees of freedom of the secondary object are not yet fully implemented in state-of-the art results of EMRI waveform modeling, even though they are under heavy investigation Piovano et al. (2020); Mathews et al. (2022); Rahman and Bhattacharyya (2023). Our work contributes to this ongoing effort, in order to exploit future gravitational wave detectors at their full potential Barausse et al. (2020); Lukes-Gerakopoulos and Witzany (2021).

Our objective, through this series of work (of which this article is the first part), is to extend all the aforementioned results to the linear-in-spin, dipolar case. As they are strongly tied to Kerr geodesics being (i) Hamiltonian and (ii) integrable, our first aim is to (i) give a Hamiltonian formulation of the linear-in-spin dynamics and (ii) apply it the Kerr spacetime and show its integrability.333These two items are this article’s content, while aforementioned applications will be published in subsequent parts.

As argued above, various proposals have already been made for a Hamiltonian formulation of the dipolar MPTD system that only satisfies some of the criteria below. Our first main result is a formulation of the linear-in-spin MPTD system (9) as a Hamiltonian system that is at once:

  • symplectic : it comes with a Poisson structure free of degeneracies

  • a first integral : it is autonomous and not a pure phase space constraint,

  • self-consistent : it does not keep terms of the same order in spin as it omits,

  • covariant : it is independent of a choice of spacetime coordinates or tetrad field,

  • 10-dimensional : as a simple counting of the independent degrees of freedom require,

Besides the formalism, a number of secondary formulae that may be useful in other contexts are given, and clarifications of various statements made in the literature are discussed.

The proof of integrability in the Schwarzschild and Kerr spacetimes relies on exhibiting enough functionally independent integrals of motion. Since our Hamiltonian formulation of the linear-in-spin, dipolar MPTD system is 10-dimensional, integrability follows from the existence of 5 such integrals, denoted (μ,E,Lz,𝔎,𝔔)(\mu,E,L_{z},\mathfrak{K},\mathfrak{Q}). All have a simple physical interpretation and are associated to some kind of symmetry:

  • the mass μ\mu of the body is conserved because the particle evolves freely at dipolar order (no self-induced force or torque from higher-order multipoles), its associated Killing structure is the metric gabg_{ab}, a particular Killing-Stäckel tensor,

  • the particle’s total energy EE, as measured at spatial infinity, is conserved thanks to the stationarity of the Kerr background, with associated timelike Killing vectors (t)a(\partial_{t})^{a},

  • the component LzL_{z} of total angular momentum along the primary black hole’s spin axis, as measured at spatial infinity, is conserved thanks to the axisymmetry of the Kerr background, with associated spacelike Killing vector (ϕ)a(\partial_{\phi})^{a},

  • two additional integrals 𝔎,𝔔\mathfrak{K},\mathfrak{Q}, known as Rüdiger’s invariants, are built from the hidden symmetry associated to a Killing-Yano tensor. While 𝔎\mathfrak{K} is the projection of the particle’s spin 4-vector onto the covariant angular momentum, 𝔔\mathfrak{Q} is a linear-in-spin generalization of the geodesic Carter constant.

We note that we did not discover any of these five integrals of motion: they have been known (at least) since Rüdiger’s work Rüdiger (1981, 1983) on linear-in-spin invariants of motion for the dipolar MPTD system. The novelty rather comes from the rigorous and covariant Hamiltonian setup in which they can be understood as functionally independent, first integrals, ensuring integrability.

.5 Organization of the paper

This article will serve as the basis for many extensions, starting with Ramond and Isoyama (2024), and contains detailed that will be otherwise omitted in further application-oriented publications. Everything is justified or proved as rigorously as possible. It is organized as follows.

  • We start in Sec. I by reviewing the main aspects of the MPTD system, focusing on the dipolar order in Sec. I.1 and the important notion of “spin supplementary condition (SSC) for spinning particles in Sec. I.2. Our choice for the Tulczyjew-Dixon SSC is motivated in Sec. I.3 and the linear-in-spin dynamics are covered in Sec. I.4. The equations of interest for the rest of the article, are Eqs. (9) and (6) (hereafter “the system”).

  • A first Hamiltonian formulation of the system is provided in Sec. II, where it is built on a 14-dimensional phase space, denoted \mathcal{M}, in Sec. II.1. Sections II.2 and II.3 contain the resulting equations of motion and an important change of coordinates, respectively. The phase space \mathcal{M} is then reduced to a 12-dimensional one, denoted 𝒩\mathcal{N}, by lifting the degeneracies on \mathcal{M} associated to the local Lorentz invariance of general relativity, as explained in Sec. II.4.

  • Section III is aimed at covering 𝒩\mathcal{N} with canonical coordinates, and discussing their physical meaning. The derivation, though technical and relegated to App. B, is based on a method that will be relevant for applications, namely transforming non-canonical coordinates into canonical ones. For an explicit example of its use, see Ramond and Isoyama (2024). A summary of the 12D formulation on 𝒩\mathcal{N} is then provided in Sec. III.3,

  • Section IV is divided in two parts: first in Sec. IV.1 we rigorously incorporate the Tulczyjew-Dixon spin supplementary condition into the framework, thereby reducing the phase space once again, from 𝒩\mathcal{N} to a final 10-dimensional phase space 𝒫\mathcal{P} where all physical trajectories lie. The final formulation is summarized in Sec. IV.2 and is the first main result of this paper. Second, in Sec. IV.3, the state-of-the-art results for invariants of the MPTD system are discussed in the context of our Hamiltonian framework.

  • Lastly, in Sec. V with apply the 10-dimensional formulation to the Schwarzschild and Kerr spacetime, in Sec. V.2 and Sec. V.3, respectively. All explicit formulae are given, and the proof of integrability follows from straightforward calculations of Poisson brackets on the 10-dimensional phase space 𝒫\mathcal{P}. This section contains many explicit formulae on which subsequent works in this series will be based.

A concluding section Conclusions and prospects contains a summary of our results in Sec. Concluding summary and a number of possible applications and prospects in Sec. Future applications, most of which are already ongoing. All our notations and conventions for Lorentzian and symplectic geometry are described in App. A, and a Mathematica Companion Notebook, accessible by simple request, contains an explicit verification of all calculations made in Sec. V related to integrability in Kerr.

I Spinning point particles in general relativity

A sometimes overlooked, yet remarkable result of general relativity is that, if a compact object is moving in a given background spacetime (with no back reaction) and modeled as a point particle with (an arbitrarily large number of) multipoles, then its equations of motion can be found in closed form. Moreover, these equations have been derived in a variety of different and independent ways: multipolar expansion Tulczyjew (1959); Dixon (1974); Steinhoff and Puetzfeld (2010), Lagrangian formalism Steinhoff (2015a), gravitational self-force Gralla and Wald (2008); Pound (2012); Mathews et al. (2022), generalized Killing fields Harte (2012, 2015, 2020); see also Chap. 2 in Ramond (2021) for a survey of such methods. Here we only summarize the key results from this branch of relativistic mechanics.

We consider a fixed background spacetime (,gab)(\mathscr{E},g_{ab}) where \mathscr{E} is a 4-dimensional (4D) manifold covered with four coordinates xαx^{\alpha}, and gabg_{ab} is a metric tensor on \mathscr{E}. The actual, physical trajectory of a compact object and all of its continuous, fluid-like properties can be put in correspondence with the worldline \mathscr{L}\subset\mathscr{E} of a representative point particle endowed with multipoles, in a scheme often referred to as a multipolar skeleton formalism Dixon (2015); Harte (2015); Mathews et al. (2022). The higher the number of multipoles, the finer the description of the extended body by the multipolar particle. A whole hierarchy of multipoles enters the description at successive multipolar orders. For example, at monopolar order, the extended body is described by a particle endowed with a single monopole: its four-momentum vector pap^{a}. At dipolar order, an antisymmetric, rank-2 spin tensor SabS^{ab} is added on top of pap^{a}, for a total of 10 degrees of freedom (4 translational ones in pap^{a} and 6 rotational ones in SabS^{ab}). Beyond the monopolar and dipolar order (n=0n=0 and n=1n=1 respectively), each multipolar order n2n\geq 2 adds a tensor of valence n+2n+2 (called a 2n2^{n}-pole) to the description of the body, bringing an additional (n+3)(3n1)(n+3)(3n-1) independent degrees of freedom due to their algebraic symmetries Dixon (1974). These multipoles evolve along the worldline according to a system of ten coupled, ordinary differential equations (ODE) known as the Mathisson-Papapetrou-Tulczyjew-Dixon (MPTD) equations. They take the general form444We note that from the algebric symmetries of RabcdR_{abcd} and the anti-symmetry of SabS^{ab}, one has RabcdSbc=12RbcdaSbcR_{abcd}S^{bc}=\tfrac{1}{2}R_{bcda}S^{bc}.

(vp)a\displaystyle(\nabla_{v}p)_{a} =RabcdSbcvd+Fa,\displaystyle=R_{abcd}S^{bc}v^{d}+F_{a}\,, (1a)
(vS)ab\displaystyle(\nabla_{v}S)^{ab} =2p[avb]+Nab,\displaystyle=2p^{[a}v^{b]}+N^{ab}\,, (1b)

where vav^{a} is an arbitrary tangent vector to \mathscr{L} and v:=vaa\nabla_{v}:=v^{a}\nabla_{a} is the (metric-compatible) covariant derivative along vav^{a}, RabcdR_{abcd} is the Riemann curvature tensor of the background spacetime and Fa,NabF_{a},N^{ab} denote contributions from higher-order multipoles coupled to the geometry, starting with the quadrupole. Note the absence of evolution equations for the quadrupole and higher-order multipoles. Much like in classical mechanics Harte (2015), the MPTD system is not sufficient to describe the system beyond the dipolar order. The quadrupole and other, higher-order multipoles need to be given in terms of pa,Sabp^{a},S^{ab} and the geometry. These expressions are not universal, and can be found by examining the properties of black hole metrics and/or micro-physics for neutron and exotic stars (see for example references in Ramond and Le Tiec (2021)). Even with such models for the quadrupoles and higher-order multipoles, the system is still ill-posed, as there are more unknowns than equations. This comes from two observations. First, the tangent vector vav^{a} is arbitrary in (1) (i.e., not necessarily normalized) as the MPTD equations are actually invariant by re-parametrization of \mathscr{L}. From now, one we will set va=uav^{a}=u^{a}, where uau^{a} is the four-velocity of the particle, normalized as uaua=1u_{a}u^{a}=-1. Second, there is freedom in choosing the worldline that represents the body. At monopolar order, this is not an issue, but starting at dipolar order, this freedom requires a choice to define the angular momentum of the particle properly Steinhoff (2015a); Vines et al. (2016). This is discussed in more details in Sec. I.2.

I.1 The dipolar MPTD system

In the present work, we will truncate the MPTD system (1) at dipolar order, so that all the physical information about the compact object is encoded into a four-momentum form pap_{a} (not tangent to \mathscr{L} in general) and an antisymmetric spin tensor SabS^{ab}. The dipolar MPTD system is obtained by turning off all beyond-dipolar-order multipoles Eqs. (1), which is equivalent to neglecting the force FaF_{a} and torque NabN^{ab} Harte (2015). One is left with the dipolar MPTD equations:

(up)a\displaystyle(\nabla_{u}p)_{a} =RabcdSbcud,\displaystyle=R_{abcd}S^{bc}u^{d}\,, (2a)
(uS)ab\displaystyle(\nabla_{u}S)^{ab} =2p[aub].\displaystyle=2p^{[a}u^{b]}\,. (2b)

The physical interpretation of (2) is clear: the coupling between SabS^{ab} and the background curvature RabcdR_{abcd} drives the evolution of pap^{a}, while the misalignment of pa:=gabpbp^{a}:=g^{ab}p_{b} of uau^{a} drives the evolution of SabS^{ab}. Three scalar fields along \mathscr{L} can be defined from the multipoles of the particle:

μ2:=papa,S2:=12SabSab,andS2:=18εabcdSabScd.\mu^{2}:=-p_{a}p^{a}\,,\qquad S_{\circ}^{2}:=\frac{1}{2}S_{ab}S^{ab}\,,\quad\text{and}\quad S_{\star}^{2}:=\frac{1}{8}\varepsilon_{abcd}S^{ab}S^{cd}\,. (3)

By assumption, pap^{a} is time like while SabS^{ab} is spacelike, resulting in μ0\mu\geq 0, S0S_{\circ}\geq 0 and S0S_{\star}\geq 0. Among these norms, μ\mu and SS_{\circ} will be referred to as the mass, and SS_{\circ} as the spin norm. Note that they are intrinsically defined from the particle’s multipole, and do not depend on any observer. These quantities are, in general, not conserved under the dipolar MPTD system (2).

A remark of importance has to be made about Eqs. (2). It is known that the spin of any compact object will induce a quadratic-in-spin quadrupole Marsat (2015) (see also Chap. 6 in Ramond (2021)), which contributes to both force FaF_{a} and torque NabN^{ab} in the full MPTD system (1). Therefore, when truncating (1) at dipolar order to obtain (9), we have effectively neglected all quadratic-in-SabS^{ab} contributions from the right-hand side of (1). Consequently, to remain self-consistent, Eqs. (9) would also need to be linearized555Technically, one could argue that Eq. (9) are already linear in SabS^{ab}. However, they are not well-posed mathematically, and it will become apparent that they are not linear-in-SabS^{ab} when completed with an additional condition ensuring well-posedness, cf. Sec. I.2. in SabS^{ab}. If not, one would include quadratic-in-spin corrections from the dipolar sector but omits those of the quadrupolar sector, leading to inconsistency. Therefore, it should be kept in mind that the dipolar MPTD equations (2) are only self-consistent at linear order in spin. This linearization, rather ambiguous if the “spin” means the tensor SabS^{ab}, can be done properly with a choice of spin supplementary condition and the introduction of a spin-related, small parameter, as explained in Sec. I.4.

I.2 Spin supplementary conditions

The number of independent unknowns in (vα,pα,Sαβ)(v^{\alpha},p_{\alpha},S^{\alpha\beta}) is 4+4+6=144+4+6=14. Yet, there are only 1010 differential equations in (2). This ambiguity comes from an arbitrariness in the choice of parameter λ\lambda associated to the tangent vector vαv^{\alpha}, and in the definition of the spin tensor, which contains angular momentum to be defined with respect to some reference (as in classical mechanics). A common way of fixing these issues (and thus having a well-posed differential system), is to 1) take va=uav^{a}=u^{a}, the four-velocity along \mathscr{L} so that λ\lambda is the proper time τ\tau and the condition uaua=1u^{a}u_{a}=-1 reduces the number of unknowns to 13, and 2) impose a so-called spin supplementary condition (SSC), reducing this number to 10. An SSC takes the form of 3 algebraic constraints of the form

faSab=0,f_{a}S^{ab}=0\,, (4)

where faf^{a} is some timelike vector defined along \mathscr{L} (only three because this equation projected onto fbf_{b} gives 0=00=0 by antisymmetry of SabS^{ab}). To elucidate the meaning of such SSC, perform a Hodge decomposition of SabS^{ab} with respect to the timelike vector faf^{a}, effectively splitting SabS^{ab} into a spin vector SaS^{a} and mass dipole vector DaD^{a}, according to666Equation (5) assumes that fafa=1f_{a}f^{a}=-1. If not, extra factors of fafaf_{a}f^{a} must be included.

Sab=εabcdfcSd+2D[afb]{Sb=12εabcdfaScd,Db=faSab.S^{ab}=\varepsilon^{abcd}f_{c}S_{d}+2D^{[a}f^{b]}\quad\Leftrightarrow\quad\begin{cases}\,S^{b}=\frac{1}{2}\varepsilon^{abcd}f_{a}S_{cd}\,,\\ D^{b}=f_{a}S^{ab}\,.\end{cases} (5)

The six degrees of freedom contained in the tensor SabS^{ab} are encoded in the spacelike vectors SaS^{a} and DaD^{a}, which have only three degrees of freedom each, as both verify the constraint of being orthogonal to faf^{a}. Clearly, the SSC (4) means, physically, that the mass dipole DaD^{a} measured in the spacelike frame of timelike direction faf^{a} vanishes. In that frame, SabS^{ab} possesses three degrees of freedom and can be entirely described in terms of SaS^{a}. This decomposition is entirely analogous to the decomposition of the Faraday tensor FabF^{ab} of electromagnetism into magnetic BaB^{a} and electric EaE^{a} fields measured by a given observer Gourgoulhon (2013), with the dictionary (Fab,Ba,Ea)(Sab,Sa,Da)(F^{ab},B^{a},E^{a})\leftrightarrow(S^{ab},S^{a},D^{a}).

The natural question is: which frame should be chosen?, i.e., which vector faf^{a} in (4) ?. This question goes beyond the scope of the present work, and we refer to the thorough review Witzany et al. (2019) and references therein for details on the matter of SSCs (see also Lukes-Gerakopoulos et al. (2014); Costa and Natário (2015); Timogiannis et al. (2021)). While different choices of SSC will lead to more or less simple evolution equations, physical meaning and mathematical well-posedness, one generally expects that physical observables should all remain the same regardless of the chosen SSC, at least at dipolar order Steinhoff (2011, 2015b); Lukes-Gerakopoulos (2017).

I.3 The Tulzyjew-Dixon SSC

In this article and subsequent works, we shall work with the so-called Tulczyjew-Dixon spin supplementary condition (TD SSC) Tulczyjew (1959); Dixon (1964)777Tulczyjew was the first to acknowledge the fact the other SSCs did not necessarily lead to unique center-of-mass trajectories in special relativity Dixon (2015), and adopted it also in GR in Tulczyjew (1959). :

Cb:=paSab=0.C^{b}:=p_{a}S^{ab}=0\,. (6)

Combining the TD SSC (6) with the dipolar MPTD equations (2) leads to a number of well-known results, recalled here for convenience. First, it follows from Eq. (2b) that the spin norm SS_{\circ} is constant along \mathscr{L}. Second, from (5), we find S=0S_{\star}=0 (as enforced by any SSC Witzany et al. (2019)). Third, as shown first in Ehlers and Rudolph (1977) (see also Obukhov and Puetzfeld (2011) for a covariant proof), the TD SSC implies the following exact expression for the tangent vector vav^{a} in terms of pa,Sabp_{a},S^{ab} and the background geometry

ua=mμ2(pa+2SabRbcdepcSde4μ2+RabcdSabScd),u^{a}=\frac{m}{\mu^{2}}\left(p^{a}+\frac{2S^{ab}R_{bcde}p^{c}S^{de}}{4\mu^{2}+R_{abcd}S^{ab}S^{cd}}\right)\,, (7)

where m:=pauam:=-p_{a}u^{a} can be expressed in terms of solely (pa,Sab,Rabcdp_{a},S^{ab},R_{abcd}) by contracting Eq. (7) with vav_{a} and re-inserting the result. Equation (7) readily implies that all the scalars (3) are conserved.

The momentum-velocity relation (7) is a consequence of the dipolar MPTD system (2) and the SSC (4). One could ask if the converse result holds true, i.e., whether the momentum-velocity relation (7) and the dipolar MPTD together imply paSab=0p_{a}S^{ab}=0. This is almost the case, for Eqs. (7) and (2) actually imply that v(paSab)=0\nabla_{v}(p_{a}S^{ab})=0, i.e., that the four-vector paSabp_{a}S^{ab} is parallel-transported along \mathscr{L}. This feature that has no major implication for solving the equations themselves888From the point of view of a Cauchy problem, i.e., a set of ODE’s with some initial conditions, it suffices to require paSab=0p_{a}S^{ab}=0 at some initial time for the TD SSC to hold at all times (by linearity of parallel-transport).. Yet, it must be handled with care in a Hamiltonian formulation of those same equations. This is discussed in Sec. IV.1. Our choice of SSC (6) is not random, but motivated by a number of reasons, detailed below:

  • (i)

    to our knowledge, it is the only one for which a unique center-of-mass-like worldline can be unambiguously defined for the actual extended body, and for which mathematical proofs of unicity exist Beiglböck (1967); Madore (1969); Schattner (1979a, b); Harte (2015). Other SSCs tend to correspond to a family of worldlines rather than a unique one Costa et al. (2012).

  • (ii)

    it leads to a momentum-velocity relation that can be generalized to any multipolar order and can also account for the presence of electromagnetic fields Gralla et al. (2010). To our knowledge, this has never been shown for any other SSC.

  • (iii)

    it naturally emerges when looking at the MPTD equations from the point of view of generalized Killing fields, the formalism developed by A. Harte Harte (2012, 2015) that goes beyond the test body assumption and takes into account self-field effects.

  • (iv)

    several works, including Rüdiger (1981, 1983); Compère and Druart (2022); Compère et al. (2023), aimed at finding invariants of motion for the MPTD system beyond the geodesic limit. These works proved successful when relying on the TD SSC, and methods do not seem to generalize to other SSCs.

  • (v)

    in spacetimes admitting a Killing-Yano tensor, it leads to a covariant notion of angular momentum Dietz and Rüdiger (1981); Dietz and Rudiger (1982); Compère and Druart (2022), with well-behaved geodesic and Newtonian limits for black hole spacetimes Ramond and Isoyama (2024).

  • (vi)

    it is compatible with a covariant Hamiltonian formulation. As pointed out in Witzany et al. (2019), the reason is that the TD SSC is “co-moving” Witzany et al. (2019), i.e., it relies on only the intrinsic properties of the body (pa,Sabp_{a},S^{ab}) and no background fields

  • (vii)

    for such covariant Hamiltonian formulations, the TD SSC uniquely reduces the number of degrees of freedom by the largest amount Witzany et al. (2019). In a canonical setup, this corresponds to two phase space dimensions.

  • (viii)

    as seen in Sec. IV, the TD SSC defines a phase-space sub-manifold 𝒯\mathcal{T} that is invariant under the flow of the free particle Hamiltonian (12). This is a key aspect of our covariant formulation.

  • (ix)

    in a future publication, we shall demonstrate how the linear-in-spin, Hamiltonian formalism presented in this paper can be extended covariantly to the quadratic-in-spin case. We have only been able to do this with the use of the TD SSC.

I.4 Linear-in-spin dynamics under the TD SSC

For convenience, from now on, we take the tangent vector vav^{a} to be the four-velocity uau^{a} of the particle. The associated parameter along \mathscr{L} is the proper time, denoted τ\tau. Inserting (7), into (2) then leads to a closed system of 10 equations for the ten independent components of the momenta (pa,Sab)(p_{a},S^{ab}). This system of equations, however, has a singular limit (observe the denominator in (7)). To make it explicit, we introduce the (only) dimensionless physical parameter at play: If 1/L21/L^{2} is the typical curvature scale of the background spacetime, then Eq. (7) is well-behaved only in the limit where

ϵ:=SμL1.\epsilon:=\frac{S_{\circ}}{\mu L}\lesssim 1\,. (8)

For a generic extended body, the spin angular momentum SS_{\circ} scales as μ\mu\ell where \ell is the typical size of the body, leading to ϵ=/L\epsilon=\ell/L. We recover the small parameter tied to multipolar expansions that lead to the MPTD equations Dixon (1974); Harte (2012); Ramond and Le Tiec (2021). In the particular case of a compact object (defined by μ\ell\simeq\mu) orbiting around black hole of mass MM (for which LML\simeq M) one readily obtains ϵ=μ/M\epsilon=\mu/M. Therefore, the MPTD equations are well-suited to capture the leading effect of the spin in the context of binary systems with asymmetric mass ratio (see the discussions in Witzany et al. (2019); Miller and Pound (2021)).

With the small parameter (8), we can now rigorously expand the dipolar MPTD system to linear order in spin, by (i) setting Sab=SS¯abS^{ab}=S_{\circ}\bar{S}^{ab}, pa=μp¯ap^{a}=\mu\bar{p}^{a} and Rabcd=R¯abcd/L2R_{abcd}=\bar{R}_{abcd}/L^{2}, with barred quantities being dimensionless, (ii) inserting this into Eqs. (2) and (7) and (ii) neglecting any non-linearity in ϵ\epsilon. This leads to

pa\displaystyle p^{a} =μua,\displaystyle=\mu u^{a}\,, (9a)
(up)a\displaystyle(\nabla_{u}p)_{a} =RabcdSbcud,\displaystyle=R_{abcd}S^{bc}u^{d}\,, (9b)
(uS)ab\displaystyle(\nabla_{u}S)^{ab} =0.\displaystyle=0\,. (9c)

It is worth noting at this order, Eq. (9a) implies that pap^{a} is tangent to \mathscr{L}, just as in the non-spinning (geodesic) case. Moreover, Eq. (9c) implies that SabS^{ab} is parallel-transported along \mathscr{L}, just as for a test spin. However, Eq. (9b) implies that \mathscr{L} is not a geodesic of (,gab)(\mathscr{E},g_{ab}) because of the spin-curvature coupling. It is tempting to interpret the right-hand side of Eq. (9b) as a “force”, as it drives the evolution of the momentum. But we stress that there is no dynamical evolution in these equations. In fact, the MPTD system at dipolar order (2) is (i) pure kinematics: body-dependent multipoles only enter at quadrupolar order, with non-zero Fa,NabF^{a},N^{ab} Harte (2015); Harte and Dwyer (2023); and (ii) completely universal: all spinning bodies with the same initial conditions will follow the same trajectories.

System (9) along with the TD SSC (6) is the system that we will to study in this paper. More precisely, what we will investigate in subsequent sections is the Hamiltonian formulation of the differential system given by the components of Eq. (9) onto the natural basis (α)a(\partial_{\alpha})^{a} associated to some coordinate system xαx^{\alpha} on ()(\mathscr{E}). This is the object of Sec. II. We also have to keep in mind that the well-posedness of Eqs. (9) relies on the TD SSC (6), which has to holds along \mathscr{L}999We emphasize again that the system (9) only implies u(paSab)=0\nabla_{u}(p_{a}S^{ab})=0, and not paSab=0p_{a}S^{ab}=0.. We will ensure that the latter is correctly incorporated into the Hamiltonian formulation in Sec. (IV). Lastly, we emphasize that all calculations, from now on, will be performed consistently at linear order in spin, and what we is meant by that, is that in any equation we can restore the dimensionless, small parameter ϵ\epsilon (define in (8)) and neglect any non-linearity in ϵ\epsilon.

II The MPTD equations as a constrained Poisson system

In this section, we examine the linear-in-spin MPTD system (9) and write it as a Poisson system, a particular type of Hamiltonian systems which comes with degeneracies. (See App. (105) for details about different types of Hamiltonian systems.) First, the Poisson system is rigorously defined in Sec. II.1 and the resulting equations of motion are derived in Sec. II.2. Using a change of phase space coordinates in Sec. II.3, the Poisson system is shown to be degenerate (non-symplectic) in Sec. II.4, and the degeneracies are lifted accordingly.

II.1 Definition of the Poisson system

It is well-known that the linearized MPTD + TD SSC system (9) can be written as a Poisson system d’Ambrosi et al. (2015); Witzany et al. (2019). As reviewed in App. D, this type of system is constructed from three ingredients: a phase space \mathcal{M}, a Poisson structure Λ\Lambda on \mathcal{M} (or, equivalently, a set of Poisson brackets {,}\{,\}), and a Hamiltonian HH.

II.1.1 Phase space

In the 4-dimensional spacetime manifold (,gab)(\mathscr{E},g_{ab}), consider a coordinate system xαx^{\alpha} and the natural bases (α)a(\partial_{\alpha})^{a} and (dxα)a(\mathrm{d}x^{\alpha})_{a} associated to it. Let the components of the four-momentum and spin tensor in these bases be pαp_{\alpha} and SαβS^{\alpha\beta}, respectively, and let xα(τ¯x^{\alpha}(\bar{\tau} be a parametrization101010As customary, we use the same notation between the coordinates xαx^{\alpha} covering \mathscr{E} and the parameterization xα(τ¯)x^{\alpha}(\bar{\tau}) of the worldline \mathscr{L}. The context will always be such that this is not ambiguous. of the dipolar particle’s worldline \mathscr{L} with respect to the parameter τ¯:=τ/μ\bar{\tau}:=\tau/\mu, which is affine parameter as μ\mu is constant along \mathscr{L}. We then define a 14-dimensional phase space =14\mathcal{M}=\mathbb{R}^{14}, endowed with 14 phase space coordinates denoted

(xα,pβ,Sγδ)4×4×6,(x^{\alpha},p_{\beta},S^{\gamma\delta})\in\mathbb{R}^{4}\times\mathbb{R}^{4}\times\mathbb{R}^{6}\,, (10)

such that their physical meaning is the one presented above. Note that changing the coordinates on \mathscr{E} corresponds to a different covering of phase space coordinates, not just in xαx^{\alpha}, but also in (pα,Sβγ)(p_{\alpha},S^{\beta\gamma}) because of the natural bases used for their definition.

II.1.2 Poisson structure

The Poisson structure Λ\Lambda can be equivalently defined by a set of Poisson brackets between the coordinates (10), cf. App. D.2 for a summary of Poisson systems. There is a well-known set of such brackets in the context of the MPTD equations, see the references in Witzany et al. (2019). The non-vanishing ones are given explicitly by

{xα,pβ}\displaystyle\{x^{\alpha},p_{\beta}\} =δβα,\displaystyle=\delta^{\alpha}_{\beta}\,, (11a)
{pα,pβ}\displaystyle\{p_{\alpha},p_{\beta}\} =RαγδβSγδ,\displaystyle=R_{\alpha\gamma\delta\beta}S^{\gamma\delta}\,, (11b)
{pα,Sβγ}\displaystyle\{p_{\alpha},S^{\beta\gamma}\} =ΓδαβSδγ+ΓδαγSβδ,\displaystyle=\Gamma^{\beta}_{\delta\alpha}S^{\delta\gamma}+\Gamma^{\gamma}_{\delta\alpha}S^{\beta\delta}\,, (11c)
{Sαβ,Sγδ}\displaystyle\{S^{\alpha\beta},S^{\gamma\delta}\} =gαγSβδgαδSβγ+gβδSαγgβγSαδ,\displaystyle=g^{\alpha\gamma}S^{\beta\delta}-g^{\alpha\delta}S^{\beta\gamma}+g^{\beta\delta}S^{\alpha\gamma}-g^{\beta\gamma}S^{\alpha\delta}\,, (11d)

where Γβγα\Gamma^{\alpha}_{\beta\gamma} are the Christoffel symbols associated to the metric gabg_{ab}, and RαβγδR_{\alpha\beta\gamma\delta} the components of the Riemann tensor in the natural basis. We stress that, in the phase space context, all GR geometric objects gαβ,Γβγα,Rαβγδ,g_{\alpha\beta},\Gamma^{\alpha}_{\beta\gamma},R_{\alpha\beta\gamma\delta},\ldots are viewed as simple functions of the coordinates xαx^{\alpha}. Formulae (11) determine a well-defined set of Poisson brackets, as they satisfy the two requirements to be so: anti-symmetry and the Jacobi identity, as discussed in App. C.2. While the anti-symmetry is easily checked, the Jacobi identity (104b) is verified in App. C.2 using the more practical coordinates defined in Sec. II.3.

II.1.3 Hamiltonian function

The last ingredient is a Hamiltonian function that will generate the same differential system as the MPTD equations. In Witzany et al. (2019), the authors derived a Hamiltonian for each of the commonly used SSC using heuristic arguments, essentially, because no clear strategy exists to do this in general. In that article, one can find (see Eq. (44) there) a Hamiltonian that generates both the (nonlinear) dipolar MPTD equations (2) and the TD momentum-velocity relation (7). However, this Hamiltonian is not well-suited for our purposes for several reasons. First, it is “pure constraint”, i.e., it vanishes identically on the sub-manifold defined by the TD SSC. This poses a problem when applying the Liouville-Arnold theorem Arnold et al. (2006) to discuss integrability, since the Hamiltonian itself cannot be used as an integral of motion. Second, the proposed Hamiltonian is valid under the assumption that the dynamical mass μ2=papa\mu^{2}=-p_{a}p^{a} is a given parameter (independent of the phase space). This is unfortunate as (i) we will need it to find enough invariants of motion, and (ii) we expect the constancy of μ\mu to be a natural consequence of the evolution equations (i.e., Hamilton’s equation, in this context) and not as a ad hoc condition. Third, they were built to generate the nonlinear dipolar MPTD system, and, therefore, contain too much information for our purposes, as already explained in Sec. I.1. Lastly, and perhaps most importantly, this Hamiltonian does not convey the central idea that at dipolar order, the MPTD equations are those of a test particle free of any finite size effect Harte (2012, 2015), or, equivalently, simple conservation laws associated to the local Lorentz invariance of general relativity Harte and Dwyer (2023). The dipolar particle is “free”, in the sense that equations (2) are universal (i.e., any body described at dipolar level satisfies these equations, whether it be a grain of dust, a planet or a compact object), with body-specific multipoles starting only at quadrupolar order Harte (2020). In particular, the spin-curvature coupling term in (2a) is not a general relativistic effect per-se: it already exists in the free particle motion of Newtonian mechanics in non-Euclidean spaces Arnold et al. (2006); Harte (2015). Based on these considerations, we vouch for a more natural choice, that of the free particle Hamiltonian111111In a forthcoming publication, we shall show that this setup is also natural if one is to include the quadrupolar contributions. As expected they will not be part of the brackets, but be a correction to the Hamiltonian (12), to be evolved with exactly the same brackets (11). In fact, we conjecture that this process can be extended to any multipolar order, at least for spin-induced multipolar moments, as in Martelli (2016).

H:(xα,pα,Sαβ)12gαβpαpβ,H:(x^{\alpha},p_{\alpha},S^{\alpha\beta})\mapsto\frac{1}{2}g^{\alpha\beta}p_{\alpha}p_{\beta}\,, (12)

where, in terms of phase space variables, gαβg^{\alpha\beta} are functions of xαx^{\alpha}. Even though it has the same expression, this Hamiltonian is not, strictly speaking, the Hamiltonian that generates geodesics in the metric gabg_{ab} (used for example in Schmidt (2002)). Indeed, even though it does not depend explicitly on SαβS^{\alpha\beta}, it is still considered a function of (xα,pα,Sαβ)(x^{\alpha},p_{\alpha},S^{\alpha\beta}). The equations of motion will, therefore, still contain terms involving the spin because of the Poisson structure (11).

II.2 Equations of motion

The computation of the equations of motion is made with respect to some arbitrary parameter λ\lambda along the worldline (the “time” conjugated to the Hamiltonian), using the (generalized) Hamilton equations:

dFdλ={F,H}={F,xα}Hxα+{F,pα}Hpα+{F,Sαβ}HSαβ,\frac{\mathrm{d}F}{\mathrm{d}\lambda}=\{F,H\}=\{F,x^{\alpha}\}\,\frac{\partial H}{\partial x^{\alpha}}+\{F,p_{\alpha}\}\,\frac{\partial H}{\partial p_{\alpha}}+\{F,S^{\alpha\beta}\}\,\frac{\partial H}{\partial S^{\alpha\beta}}\,, (13)

with FF any function of the phase space variables. To be precise, in Eq. (13) the first equality is a definition (that of the evolution of F:F:\mathcal{M}\rightarrow\mathbb{R} under the flow of HH for the Poisson system (,{,},H)(\mathcal{M},\{,\},H)), and the second equality makes use of the Leibniz rule satisfied by the Poisson bracket. The evolution parameter λ\lambda is that uniquely conjugated to the Hamiltonian (such that {λ,H}=1\{\lambda,H\}=1). It can be physically interpreted only once we compare the Hamilton equations to the system (9).

The partial derivatives of the Hamiltonian (12) are easily found to be

Hxα=Γαβγpγpβ,Hpα=gαβpβandHSαβ=0,\frac{\partial H}{\partial x^{\alpha}}=-\Gamma^{\gamma}_{\alpha\beta}p_{\gamma}p^{\beta}\,,\quad\frac{\partial H}{\partial p_{\alpha}}=g^{\alpha\beta}p_{\beta}\quad\text{and}\quad\frac{\partial H}{\partial S^{\alpha\beta}}=0\,, (14)

where we projected the metric compatibility condition agbc=0\nabla_{a}g_{bc}=0 onto the natural basis to get the first equation. We then replace FF in Eq. (13) by the phase space coordinates xα,pα,Sαβx^{\alpha},p_{\alpha},S^{\alpha\beta} and combine Eqs. (11), (13) and (14) to obtain the evolution for each coordinate:

wα:=dxαdλ\displaystyle w^{\alpha}:=\frac{\mathrm{d}x^{\alpha}}{\mathrm{d}\lambda} =gαβpβ,\displaystyle=g^{\alpha\beta}p_{\beta}\,, (15a)
dpαdλ\displaystyle\frac{\mathrm{d}p_{\alpha}}{\mathrm{d}\lambda} =Γαλνpνwλ12RαβμνSμνwβ,\displaystyle=\Gamma^{\nu}_{\alpha\lambda}p_{\nu}w^{\lambda}-\frac{1}{2}R_{\alpha\beta\mu\nu}S^{\mu\nu}w^{\beta}\,, (15b)
dSαβdλ\displaystyle\frac{\mathrm{d}S^{\alpha\beta}}{\mathrm{d}\lambda} =(ΓλναSλβ+ΓλνβSαλ)wν,\displaystyle=-(\Gamma^{\alpha}_{\lambda\nu}S^{\lambda\beta}+\Gamma^{\beta}_{\lambda\nu}S^{\alpha\lambda})w^{\nu}\,, (15c)

where wαw^{\alpha} is just a notation for dxα/dλ\mathrm{d}x^{\alpha}/\mathrm{d}\lambda. Regarding Eqs. (15b) and (15c), the first term on their right-hand side can be absorbed into their left-hand side by introducing covariant derivatives w=wαα\nabla_{w}=w^{\alpha}\nabla_{\alpha}. We note that Eqs. (15b) and (15c) are exactly the linearized MPTD equations at dipolar order (9), written with respect to tangent vector wα=pαw^{\alpha}=p^{\alpha}. Therefore, the parameter λ\lambda uniquely associated to the Hamiltonian (12) corresponds physically to the worldline parameter associated to pap^{a} as a tangent vector to \mathscr{L}. By virtue of Eq. (9a), we find λ:=τ¯=τ/μ\lambda:=\bar{\tau}=\tau/\mu, and we shall use τ¯\bar{\tau} from now on. We note that this is the same result as in the non-spinning, geodesic case. Since μ\mu is conserved along \mathscr{L}, physically speaking, τ¯\bar{\tau} is an affine parameter.

II.3 Alternative coordinates

The phase space coordinates (10) are useful to prove that the Hamiltonian system (12)-(11) is equivalent to the original system (9). It is also relevant for interpretation purposes. However, when it comes to reducing the system and to doing Hamiltonian mechanics per-se, they are not practical at all. Here, we introduce a new system of coordinates on \mathcal{M} which still retains some physical intelligibility while also admitting drastically simpler Poisson brackets.

First, we require some geometrical construction within spacetime (,gab)(\mathscr{E},g_{ab}). Consider an orthonormal tetrad field {(eA)a}A{0,,3}\{(e_{A})^{a}\}_{A\in\{0,\ldots,3\}}, whose components in the natural basis are (eA)α(e_{A})^{\alpha} (AA labels the four different vectors, and α\alpha the four components of the AA-th vector). Let ωaBC:=gbc(eB)ba(eC)c\omega_{aBC}:=g_{bc}(e_{B})^{b}\nabla_{a}(e_{C})^{c} be the connection 1-forms associated to the tetrad. Since ωaBC=ωaCB\omega_{aBC}=-\omega_{aCB}, there are six independent such 1-forms. Their components in the natural basis, ωαBC\omega_{\alpha BC}, are the connection coefficients and those in the tetrad, ωABC\omega_{ABC}, are the so-called Ricci rotation coefficients (we follow Sec. 3.4b of Wald (1984)). As before, all these objects are defined throughout spacetime, and become simple functions of the phase space coordinates xαx^{\alpha} in the phase space picture. New coordinates on \mathcal{M} are then obtained by (i) keeping the coordinates xαx^{\alpha}, (ii) replacing the linear momentum variables pαp_{\alpha} by πα\pi_{\alpha}, a combination of linear and angular momentum, and (iii) replacing SαβS^{\alpha\beta} by SABS^{AB}, the tetrad components of the spin tensor. Explicitly, we define these new coordinates (xα,πα,SAB)(x^{\alpha},\pi_{\alpha},S^{AB}) by setting121212I thank J. Féjoz for noticing that, at fixed xαx^{\alpha}, the change of coordinates (16) is linear in the momenta pα,Sαβp_{\alpha},S^{\alpha\beta}, thus making connecting it to the notion of fiber bundles.

pα\displaystyle p_{\alpha} =πα+12ωαBCSBC.\displaystyle=\pi_{\alpha}+\frac{1}{2}\omega_{\alpha BC}S^{BC}\,. (16a)
Sαβ\displaystyle S^{\alpha\beta} =SAB(eA)α(eB)β.\displaystyle=S^{AB}(e_{A})^{\alpha}(e_{B})^{\beta}\,. (16b)

Solving these equations for πα\pi_{\alpha} and SABS^{AB} shows that this coordinate change is indeed invertible, and takes the form (xα,pα,Sαβ)(xα,πα(z,p,S),Sαβ(z,S))(x^{\alpha},p_{\alpha},S^{\alpha\beta})\mapsto(x^{\alpha},\pi_{\alpha}(z,p,S),S^{\alpha\beta}(z,S)) where the right-hand side emphasizes the new coordinates’ dependence on the old set (10), denoted schematically by (z,p,S)(z,p,S). The Poisson brackets of the new coordinates can then be computed from the rule (103) and the old brackets (11). This calculation does not appear in the literature, to our knowledge. Therefore, we have detailed it in App. C. It is lengthy but straightforward, using not so well-known formulae involving the connection 1-forms (in particular Eq. (3.4.20) of Wald (1984)). The non-vanishing Poisson brackets are found to be:

{xα,πβ}\displaystyle\{x^{\alpha},\pi_{\beta}\} =δβα,\displaystyle=\delta^{\alpha}_{\beta}\,, (17a)
{SAB,SCD}\displaystyle\{S^{AB},S^{CD}\} =2ηA[DSC]B+2ηB[CSD]A.\displaystyle=2\eta^{A[D}S^{C]B}+2\eta^{B[C}S^{D]A}\,. (17b)

Note that, even though the six SABS^{AB} do not verify the canonical Poisson brackets, the four pairs (xα,πα)(x^{\alpha},\pi_{\alpha}) do, and have vanishing Poisson brackets with SABS^{AB}. The spin variables SABS^{AB} have Poisson brackets akin to the commutators of the so(1,3) algebra (see, e.g., Chap. 7 in Gourgoulhon (2013)). This is reminiscent of the definition of SABS^{AB} as the components in an orthonormal tetrad field, the latter forming a set left invariant by the SO(1,3) group of Lorentz transformations. This property is even more clear when we use, instead of the six independent SABS^{AB}, the 3+33+3 variables (S1,S2,S3,D1,D2,D3)(S^{1},S^{2},S^{3},D^{1},D^{2},D^{3}) defined by

SI=12εJKISJKandDI=S0I.S^{I}=\frac{1}{2}\varepsilon^{I}_{\phantom{I}JK}S^{JK}\quad\text{and}\quad D^{I}=S^{0I}\,. (18)

where εKIJ\varepsilon^{IJ}_{\phantom{IJ}K} is the 3D Levi-Civita symbol, equal to 11 (resp. 1-1) when IJKIJK is an even (resp. odd) permutation of 123123, and 0 otherwise. Although they are but notations in phase space, SIS^{I} and DID^{I} have a clear physical meaning: they are the tetrad components of the Euclidean spin and mass dipole vectors as defined in Eq. (5) (i.e., with (e0)a(e_{0})^{a} replacing faf^{a} there). The brackets (17b), with these notations, take their simplest form

{xα,πβ}\displaystyle\{x^{\alpha},\pi_{\beta}\} =δβα,\displaystyle=\delta^{\alpha}_{\beta}\,, (19a)
{SI,SJ}\displaystyle\{S^{I},S^{J}\} =εKIJSK,\displaystyle=\varepsilon^{IJ}_{\phantom{IJ}K}S^{K}\,, (19b)
{DI,DJ}\displaystyle\{D^{I},D^{J}\} =εKIJSK,\displaystyle=-\varepsilon^{IJ}_{\phantom{IJ}K}S^{K}\,, (19c)
{DI,SJ}\displaystyle\{D^{I},S^{J}\} =εKIJDK.\displaystyle=\varepsilon^{IJ}_{\phantom{IJ}K}D^{K}\,. (19d)

Now the brackets (19) are identical to the commutators of the generators for the Lie algebra so(1,3). We also stress that brackets (19) are equivalent to, it is more a renaming of the spin coordinates (through Eq. (18)) than a coordinate transformation.

We can now express the original Hamiltonian (12) in terms of the new coordinates (xα,πβ,SI,DJ)(x^{\alpha},\pi_{\beta},S^{I},D^{J}) on \mathcal{M}, by inserting Eqs. (16) into Eq. (12). We obtain

H:(xα,πβ,SI,DJ)12gαβπαπβ+12gαβπαωβCDSCD.H:(x^{\alpha},\pi_{\beta},S^{I},D^{J})\mapsto\frac{1}{2}g^{\alpha\beta}\pi_{\alpha}\pi_{\beta}+\frac{1}{2}g^{\alpha\beta}\pi_{\alpha}\omega_{\beta CD}S^{CD}\,. (20)

Even though (20) calls for a rather natural identification into a “non-spinning” term 12gαβπαπβ\tfrac{1}{2}g^{\alpha\beta}\pi_{\alpha}\pi_{\beta} corrected by a “linear-in-spin term” 12gαβπαωβCDSCD\tfrac{1}{2}g^{\alpha\beta}\pi_{\alpha}\omega_{\beta CD}S^{CD}. This split, however, should be interpreted with care, as there are spin-related quantities hidden in πα\pi_{\alpha}, cf. Eq. (16). We will always avoid any such interpretation, and rather treat the whole Hamiltonian as one describing an object with both linear and angular momentum (orbital and spin) intertwined.

II.4 Degeneracies of the Poisson structure

We now focus on the Poisson brackets (19) between the new coordinates (xα,πα,SI,DI)(x^{\alpha},\pi_{\alpha},S^{I},D^{I}). Following the notations of App. D.2, we are in the case N=14N=14 and we construct the Poisson matrix Λij(y)={yi,yj}\Lambda^{ij}(y)=\{y^{i},y^{j}\}, where (y1,,y14)(y^{1},\ldots,y^{14}) denote the coordinates (xα,πβ,SI,DI)(x^{\alpha},\pi_{\beta},S^{I},D^{I}), in that order.

II.4.1 Rank

Directly reading through Eqs. (19), we find that the matrix Λ\Lambda is block diagonal, and given by

Λ=(𝕁800𝔖),with𝕁8=(0𝕀4𝕀40),𝔖=(𝒮𝒟𝒟𝒮).\Lambda=\left(\begin{array}[]{ll}\mathbb{J}_{8}&0\mathbb{}\\ 0&\mathfrak{S}\end{array}\right)\,,\quad\text{with}\quad\mathbb{J}_{8}=\left(\begin{array}[]{cc}0&\mathbb{I}_{4}\\ -\mathbb{I}_{4}&0\end{array}\right)\,,\quad\mathfrak{S}=\left(\begin{array}[]{cc}\mathcal{S}&\mathcal{D}\\ \mathcal{D}&-\mathcal{S}\end{array}\right)\,. (21)

In the above equations, 𝕀4\mathbb{I}_{4} it the 4×44\times 4 identity matrix, so that 𝕁8\mathbb{J}_{8} is the canonical 8×88\times 8 Poisson matrix (it would be the Poisson matrix of four pairs of canonical coordinates in an 8-dimensional symplectic system), while 𝔖\mathfrak{S} is a 6×66\times 6 antisymmetric matrix constructed from two SO(3) matrices associated to SIS^{I} and DID^{I}, namely

𝒮=(0S3S2S30S1S2S10)and𝒟=(0D3D2D30D1D2D10).\mathcal{S}=\left(\begin{array}[]{ccc}0&S^{3}&-S^{2}\\ -S^{3}&0&S^{1}\\ S^{2}&-S^{1}&0\\ \end{array}\right)\quad\text{and}\quad\mathcal{D}=\left(\begin{array}[]{ccc}0&D^{3}&-D^{2}\\ -D^{3}&0&D^{1}\\ D^{2}&-D^{1}&0\\ \end{array}\right)\,. (22)

The matrix formulation of the Poisson structure Λ\Lambda permits an easy calculation of its rank at each point of \mathcal{M}. Direct inspection reveals that rank(𝕁8)=8\text{rank}(\mathbb{J}_{8})=8, while det(𝔖)=0\text{det}(\mathfrak{S})=0 and det(𝔖)=4\text{det}(\mathfrak{S})=4. Since Λ=diag(𝕁8,𝔖)\Lambda=\text{diag}(\mathbb{J}_{8},\mathfrak{S}), we obtain

rank(Λ)=12<14=dim().\text{rank}(\Lambda)=12<14=\text{dim}(\mathcal{M})\,. (23)

Therefore, as explained in Sec. D.2, the Poisson structure generated by the brackets of the coordinates (19), is degenerate. Consequently, there exists exactly dim()rank(Λ)=1412=2\text{dim}(\mathcal{M})-\text{rank}(\Lambda)=14-12=2 Casimir invariants for this structure. We note that, since the rank (and the vanishing of the determinant) are invariant under diffeomorphisms, these feature (degeneracy and existence of two Casimirs) also hold for the initial brackets (11).

II.4.2 Casimirs

The brackets (19) admit two independent Casimir invariants, which can be explicitly computed.131313For example, using the Killing form of the 𝔰𝔬(p,q)\mathfrak{so}(p,q) algebra, given by B(X,Y)=(q+p2)tr(XY)B(X,Y)=(q+p-2)\,\text{tr}(XY) Baum (2009), or computing of the null-space of Λ\Lambda, of which, by virtue of Eq. (103), the gradients of Casimirs form a basis. Following tradition Witzany et al. (2019), we shall take

𝒞:=SSDDand𝒞:=SD,\mathcal{C}_{\circ}:=\vec{S}\cdot\vec{S}-\vec{D}\cdot\vec{D}\quad\text{and}\quad\mathcal{C}_{\star}:=\vec{S}\cdot\vec{D}\,, (24)

where we have used the Euclidean notations S=(S1,S2,S3)\vec{S}=(S^{1},S^{2},S^{3}) and D=(D1,D2,D3)\vec{D}=(D^{1},D^{2},D^{3}). We can verify directly with the brackets (19) and the Leibniz rule that they are Casimirs, i.e., that for any (z,π,S,D)(z,\pi,S,D)-dependent function FF, {F,𝒞}=0\{F,\mathcal{C}_{\circ}\}=0 and {F,𝒞}=0\{F,\mathcal{C}_{\star}\}=0, or we can also check that their gradient is in the null space of the Poisson matrix Λ\Lambda. Either calculation is easily done and shows, indeed, that (24) are Casimirs. From a physical point of view, these quantities are easily interpreted: re-writing Eqs. (24) in terms of the variables SABS^{AB}, we obtain

𝒞=12ηABηCDSACSBDand𝒞=18εABCDSABSCD.\mathcal{C}_{\circ}=\frac{1}{2}\eta_{AB}\eta_{CD}S^{AC}S^{BD}\quad\text{and}\quad\mathcal{C}_{\star}=\frac{1}{8}\varepsilon_{ABCD}S^{AB}S^{CD}\,. (25)

Most remarkably, Eqs. (25) coincide with the covariant definition Eqs. (3) expanded in an orthonormal tetrad. Therefore, the (phase space) Casimirs are numerically equal to the (squares of the) covariant spin norms:

𝒞=S2and𝒞=S2,\mathcal{C}_{\circ}=S_{\circ}^{2}\quad\text{and}\quad\mathcal{C}_{\star}=S_{\star}^{2}\,, (26)

justifying the notation of the Casimirs. This result should be understood with care. The Poisson structure defined by the brackets (19) (or (11) equivalentmy) automatically implies the conservation of both spin norms S2,S2S_{\circ}^{2},S_{\star}^{2}, as these are Casimirs (𝒞,𝒞)(\mathcal{C}_{\circ},\mathcal{C}_{\star}) of the structure. As pointed out in Witzany et al. (2019), these spin scalars are not constant along the worldline \mathscr{L} for the general MPTD system, but they are conserved the dipolar MPTD system with the TD SSC (regardless of linearization in spin). Indeed, paSab=0p_{a}S^{ab}=0 readily implies u(S)=0\nabla_{u}(S_{\circ})=0 from Eq. (2b) and S=0S_{\star}=0 from Eq. (5).

II.4.3 Symplectic leaves

As reviewed in App. D.2, even though the Poisson brackets (19) define a degenerate structure on the 14-dimensional manifold \mathcal{M}, the latter is foliated by so called “symplectic leaves”, which we denote by 𝒩\mathcal{N}, and defined as the level sets of the Casimirs. Let us, therefore, set both Casimirs invariants (24) to some fixed values (naturally denoted by S2,S2S^{2}_{\circ},S^{2}_{\star}), and consider the 12-dimensional sub-manifold 𝒩\mathcal{N} defined by the two independent algebraic equations

𝒞(z,π,S,D)=S2and𝒞(z,π,S,D)=S2,\mathcal{C}_{\circ}(z,\pi,S,D)=S_{\circ}^{2}\quad\text{and}\quad\mathcal{C}_{\star}(z,\pi,S,D)=S_{\star}^{2}\,, (27)

where we have emphasized that these relations are really algebraic constraints between the coordinates (z,π,S,D)(z,\pi,S,D) on \mathcal{M}. The manifold 𝒩\mathcal{N} is symplectic (non-degenerate), and can be, at least in principle and locally, endowed with canonical coordinates. We shall see in Sec. (III) that such canonical coordinates can be found, and cover 𝒩\mathcal{N} globally. These coordinates hold for a given leaf 𝒩\mathcal{N}. It thus expected that they are parametrized by the Casimirs values S,SS_{\circ},S_{\star}. This is exactly what we shall find. In particular, on 𝒩\mathcal{N}, S,SS_{\circ},S_{\star} are pure parameters, and not functions of the phase space coordinates, as they were on \mathcal{M}. This is important because it makes a distinction between several kinds of invariants of motion: Casimirs (for Poisson system, constant for any Hamiltonian) and first integrals (for symplectic systems, Hamiltonian-dependent). This distinction is important, in particular when counting the “integrals of motion” to check whether a Hamiltonian system is integrable or not.

Lastly, let us discuss the choice of Poisson structure on a given 12D leaf 𝒩\mathcal{N}. Since the latter is defined as the subset of \mathcal{M} where both Casimirs (24) are constant, we are in the presence of a constrained Hamiltonian system, with Eqs. (24) defining 2 constraints. Using the standard theory of constrained systems (reviewed in App. B), it follows that the Poisson structure Λ𝒩\Lambda^{\mathcal{N}} on 𝒩\mathcal{N} is the restriction of Λ\Lambda to 𝒩\mathcal{N}. In generality, Λ\Lambda and Λ𝒩\Lambda^{\mathcal{N}} will be associated to different Poisson brackets {,}{,}𝒩\{,\}\neq\{,\}^{\mathcal{N}}. We will encounter this in section IV, where this difference is explicit in Eq. (43). However, when (and only when) the constraints correspond to Casimir invariants, a classical result of Poisson system theory Deriglazov (2022) shows that Λ𝒩\Lambda^{\mathcal{N}} and Λ\Lambda are essentially identical (this can be traced back to the Casimirs commuting with anything on \mathcal{M}). Therefore, the 7D Hamiltonian system (,Λ,H)(\mathcal{M},\Lambda,H), when restricted to a given leaf 𝒩\mathcal{N}, becomes the 6D Hamiltonian system (𝒩,Λ,H𝒩)\mathcal{N},\Lambda,H_{\mathcal{N}}), where 𝒩\mathcal{N} is endowed with any (invertible) subset of 12 coordinates from (xα,πα,SI,DI)(x^{\alpha},\pi_{\alpha},S^{I},D^{I}), and H𝒩H_{\mathcal{N}} is expressed in terms of these 12 coordinates141414Any occurence of the 2 non-chosen variables in H𝒩H_{\mathcal{N}} (or anywhere else) can be expressed in terms of the 12 chosen ones by inverting the two Casimir relations (24).. However, the original brackets (19) can still be used to evolve the system on 𝒩\mathcal{N}, i.e., {F,G}𝒩={F,G}\{F,G\}^{\mathcal{N}}=\{F,G\} for any functions F,GF,G on 𝒩\mathcal{N}. In practice, it much more practical to derive 12 canonical coordinates on 𝒩\mathcal{N}, instead of picking an arbitrary set of 12 from those on \mathcal{M}. From now on, no distinction will be made between {,}\{,\} and {F,G}𝒩\{F,G\}^{\mathcal{N}}, and a Hamiltonian HH on \mathcal{M} will still be denoted HH on 𝒩\mathcal{N}, with the assumption that the number of variables on which it depends (14 or 12, respectively) leverage any risk of confusion.

III Canonical coordinates on 𝒩\mathcal{N}

In this section, we provide explicit formulae relating the spin tensor tetrad components SABS^{AB} to two symplectic pairs of variables (σ,πσ),(ζ,πζ)(\sigma,\pi_{\sigma}),(\zeta,\pi_{\zeta}), and the two Casimir invariants (25). These variables, which are canonical pairs on any leaf 𝒩\mathcal{N}, are defined in Sec. III.1 and their construction is relegated to App. B. We discuss their physical significance in Sec. III.2 and summarize the 12D canonical formulation on 𝒩\mathcal{N} in Sec. IV, emphasizing on its covariance.

III.1 Statement of the result

Since the 12D leaves 𝒩\mathcal{N} are symplectic (non-degenerate), it is tempting to try and construct 12 canonical coordinates, i.e., 6 pairs (qi,pi)i{1,,6}(q^{i},p_{i})_{i\in\{1,\ldots,6\}} for which the brackets on 𝒩\mathcal{N} satisfy {qi,pj}=δji\{q^{i},p_{j}\}=\delta^{i}_{j}. To derive these coordinates, we note that the Casimirs (24) are only functions of (SI,DI)(S^{I},D^{I}) and not (xα,πα)(x^{\alpha},\pi_{\alpha}). A natural method to construct canonical coordinates on 𝒩\mathcal{N} emerges naturally: (i) keep the 4 pairs (xα,πα)(x^{\alpha},\pi_{\alpha}) of \mathcal{M}, as they will be automatically canonical on 𝒩\mathcal{N}, from Eq. (19a) (recall that {xα,πβ}𝒩={xα,πβ}\{x^{\alpha},\pi_{\beta}\}^{\mathcal{N}}=\{x^{\alpha},\pi_{\beta}\}), and (ii) construct 2 other canonical pairs covering the spin sector (S,D)(\vec{S},\vec{D}), say (σ,πσ),(ζ,πζ)(\sigma,\pi_{\sigma}),(\zeta,\pi_{\zeta}). The latter can be found by solving the nine spin-dependent brackets (17), now seen as canonical brackets on 𝒩\mathcal{N} of the 6 (σ,πσ,ζ,πζ)(\sigma,\pi_{\sigma},\zeta,\pi_{\zeta})-dependent functions (SI,DI)(S^{I},D^{I}). For example, the bracket {S1,S2}=S3\{S^{1},S^{2}\}=S^{3} on \mathcal{M} now becomes a partial differential equation (PDE)

S1σS2πσS2σS1πσ+S1ζS2πζS2ζS1πζ=S3,\frac{\partial S^{1}}{\partial\sigma}\frac{\partial S^{2}}{\partial{\pi_{\sigma}}}-\frac{\partial S^{2}}{\partial\sigma}\frac{\partial S^{1}}{\partial{\pi_{\sigma}}}+\frac{\partial S^{1}}{\partial\zeta}\frac{\partial S^{2}}{\partial{\pi_{\zeta}}}-\frac{\partial S^{2}}{\partial\zeta}\frac{\partial S^{1}}{\partial{\pi_{\zeta}}}=S^{3}\,, (28)

in which S1,S2,S3S^{1},S^{2},S^{3} are viewed as functions on 𝒩\mathcal{N} of the coordinates (σ,πσ),(ζ,πζ)(\sigma,\pi_{\sigma}),(\zeta,\pi_{\zeta}). This calculation is rather non-trivial because the nine Poisson brackets turn into a system of nine nonlinear coupled PDEs, but educated guesses from the so(1,3) symmetries make things manageable.

III.1.1 Formulae

At the end of the aforementioned calculation, detailed in App. B, a remarkably simple solution of the PDE system is found. It reads

S1\displaystyle S^{1} =Xcosσ,\displaystyle=X\cos\sigma\,, (29a)
S2\displaystyle S^{2} =Xsinσ,\displaystyle=X\sin\sigma\,, (29b)
S3\displaystyle S^{3} =πσ,\displaystyle=\pi_{\sigma}\,, (29c)
D1\displaystyle D^{1} =Yπσsinζcosσ+Yπζcosζsinσ+XZcosσ,\displaystyle=Y\pi_{\sigma}\sin\zeta\cos\sigma+Y\pi_{\zeta}\cos\zeta\sin\sigma+XZ\cos\sigma\,, (29d)
D2\displaystyle D^{2} =YπσsinζsinσYπζcosζcosσ+XZsinσ,\displaystyle=Y\pi_{\sigma}\sin\zeta\sin\sigma-Y\pi_{\zeta}\cos\zeta\cos\sigma+XZ\sin\sigma\,, (29e)
D3\displaystyle D^{3} =XYsinζ+Zπσ,\displaystyle=-XY\sin\zeta+Z\pi_{\sigma}\,, (29f)

where X,Y,ZX,Y,Z are functions of the momenta (πσ,πζ)(\pi_{\sigma},\pi_{\zeta}) and the Casimir invariants (S,S)(S_{\circ},S_{\star}), given by

X:=πζ2πσ2,Y:=1S2πζ2S4πζ4,Z:=S2πζ2.X:=\sqrt{\pi_{\zeta}^{2}-\pi_{\sigma}^{2}}\,,\quad Y:=\sqrt{1-\frac{S_{\circ}^{2}}{\pi_{\zeta}^{2}}-\frac{S_{\star}^{4}}{\pi_{\zeta}^{4}}}\,,\quad Z:=\frac{S_{\star}^{2}}{\pi_{\zeta}^{2}}\,. (30)

By construction, the pairs (σ,πσ)(\sigma,\pi_{\sigma}) and (ζ,πζ)(\zeta,\pi_{\zeta}) are symplectic, i.e., their only non-vanishing Poisson brackets are {σ,πσ}=1\{\sigma,\pi_{\sigma}\}=1 and {ζ,πζ}=1\{\zeta,\pi_{\zeta}\}=1.

III.1.2 Domain of definition

From direct inspection of Eqs. (29), we can take both σ\sigma and ζ\zeta in the interval [0,2π][0,2\pi] without loss of generality. Moreover, all expressions are well-defined provided that πζ2πσ20\pi_{\zeta}^{2}-\pi_{\sigma}^{2}\geq 0 and πζ4S2πζ2S40\pi_{\zeta}^{4}-S_{\circ}^{2}\pi_{\zeta}^{2}-S_{\star}^{4}\geq 0. The first condition is straightforward, and the latter implies that πζ{\pi_{\zeta}} must be in the domain ],Π][Π,+[]-\infty,-\Pi]\,\cup\,[\Pi,+\infty[, where

Π=S22+S4+4S420.\Pi=\sqrt{\frac{S_{\circ}^{2}}{2}+\frac{\sqrt{S_{\circ}^{4}+4S_{\star}^{4}}}{2}}\geq 0\,. (31)

Note that one can simply take πζ0{\pi_{\zeta}}\geq 0 without loss of generality, since all expressions in (29) are invariant under πζπζ{\pi_{\zeta}}\mapsto-{\pi_{\zeta}}, except for D1D^{1} and D2D^{2}. But since D1(σ,πσ,ζ,πζ)=D1(σ,πσ,πζ,πζ)D^{1}(\sigma,{\pi_{\sigma}},\zeta,-{\pi_{\zeta}})=D^{1}(\sigma,{\pi_{\sigma}},\pi-\zeta,{\pi_{\zeta}}), the choice πζ0{\pi_{\zeta}}\geq 0 is not restrictive to parametrize the phase space by (σ,ζ)[0,2π]×[0,2π](\sigma,\zeta)\in[0,2\pi]\times[0,2\pi] and (πζ,πσ)Δ({\pi_{\zeta}},{\pi_{\sigma}})\in\Delta, where Δ2\Delta\subset\mathbb{R}^{2} is the open set depicted in Fig. 1.

Refer to caption
Figure 1: The open domain of definition Δ2\Delta\subset\mathbb{R}^{2} for the canonical spin momenta (πσ,πζ)({\pi_{\sigma}},{\pi_{\zeta}}).

III.1.3 Simplification under an SSC

When the TD spin supplementary condition is imposed, we shall see in Sec. IV that the Casimir 𝒞=S2\mathcal{C_{\star}}=S_{\star}^{2} automatically vanishes. From Eq. (30), this implies Z=0Z=0 in Eqs. (29), making these formulae much simpler.151515In the particular case S=0S_{\star}=0, our general formulae (29) should be in a 1-to-1 correspondence with the one derived (by other means) in Witzany et al. (2019), see Eqs. (62) there. The mapping between the two variables is affine (on a given leaf 𝒩\mathcal{N}), and given by (σ,πσ,ζ,πζ)=(ϕ,A+B+S,ψϕ+π/2,B+S)(\sigma,{\pi_{\sigma}},\zeta,{\pi_{\zeta}})=(\phi,A+B+S_{\circ},\psi-\phi+\pi/2,B+S_{\circ}), where our invariant SS_{\circ} is denoted SS in Witzany et al. (2019) and (ϕ,A,ψ,B)(\phi,A,\psi,B) are their canonical coordinates. The vanishing of SS_{\star} is actually independent of the particular choice of SSC, i.e., any timelike vector faf^{a} in (4) will readily enforce S=0S_{\star}=0, as shown in Witzany et al. (2019).

As can be seen on Eqs. (29), an SSC will not enforce that the mass dipole DI0D^{I}\neq 0 will vanish, even though when S=0S_{\star}=0. This is because there is no unique mass dipole, but one for each observer, cf. the discussion in Sec. I.2. However, since S2=SDS_{\star}^{2}=\vec{S}\cdot\vec{D} (cf. Eq. (24)) the Euclidean 3-vectors S\vec{S} and D\vec{D} will be orthogonal (with respect to the Euclidean scalar product) under any SSC, for an observer on a worldline with timelike tangent vector (e0)a(e_{0})^{a}. This gives an general interpretation of whether an SSC has been applied at all, independently of the particular choice (i.e., true for any faf^{a} in Eq. (4)).

III.2 Physical interpretation

The canonical coordinates (29) parametrizing a generic spin tensor SABS^{AB} are, to the best of our knowledge, not discussed in the literature.161616 In the mathematical-oriented literature, one can find a parametrization using complex variables, obtained by combining S\vec{S} and D\vec{D} into two complex-valued 3-vectors V±K=12(ϵIJKSIJ±iDK)V^{K}_{\pm}=\frac{1}{2}(\epsilon_{IJ}^{\phantom{IJ}K}S^{IJ}\pm\mathrm{i}D^{K}). Then one can show that they satisfy the SO(3) brackets, namely {V+I,VJ}=0\{V^{I}_{+},V^{J}_{-}\}=0 and {V±I,V±J}=ϵKIJVK\{V^{I}_{\pm},V^{J}_{\pm}\}=\epsilon^{IJ}_{\phantom{IJ}K}V^{K} from which canonical coordinates can be defined as for the usual SO(3) brackets. This is related to the Weyl representation of the Lorentz group Frankel (2011). This is particularly surprising since this the variables SABS^{AB} satisfy the Lorentz so(1,3) algebra, probably the most important one of modern theoretical physics. In view of applications of the formalism presented in this paper to the case of black hole spacetimes, it is important to know what these coordinate represent physically. In the case of the Schwarzschild spacetime, this interpretation was crucial to construct other adapted coordinates via canonical transformations Ramond and Isoyama (2024).

Let us place ourselves in the reference frame defined by the tetrad (eA)a(e_{A})^{a}, and imagine the three spatial legs as axes of a Cartesian coordinate system. Inspecting the first three equations in (29) shows that πζ{\pi_{\zeta}} is the norm of the Euclidean spin vector S\vec{S}, and πσ{\pi_{\sigma}} is its projection onto the e3\vec{e}_{3}-axis. Moreover, σ\sigma is the angle measured between e1\vec{e}_{1} and the projection of S\vec{S} onto the plane Span(e1,e2)(\vec{e}_{1},\vec{e}_{2}). Notably, in this Cartesian interpretation ((πζ2πσ2)1/2,σ,πσ)((\pi_{\zeta}^{2}-\pi_{\sigma}^{2})^{1/2},\sigma,{\pi_{\sigma}}) are the usual cylindrical coordinates of the vector S\vec{S}. This is, of course, reminiscent of the SO(3)-like Poisson brackets verified by the spin vector; cf. Eq. (19b). Turning to the angle ζ\zeta, we see that it parametrizes only the mass dipole D\vec{D}. To elucidate its physical meaning, let us first note that, as given in Eqs. (29), both 3-vectors S\vec{S} and D\vec{D} clearly undergo a spatial rotation of angle σ\sigma with respect to the e3\vec{e}_{3}-axis. Indeed, we can write

S=σV,D=σ(ZV+YW)whereσ=(cosσsinσ0sinσcosσ0001),\vec{S}=\mathcal{R}_{\sigma}\vec{V}\,,\quad\vec{D}=\mathcal{R}_{\sigma}(Z\vec{V}+Y\vec{W})\quad\text{where}\quad\mathcal{R}_{\sigma}=\begin{pmatrix}\cos\sigma&-\sin\sigma&0\\ \sin\sigma&\cos\sigma&0\\ 0&0&1\end{pmatrix}\,, (32)

and we have introduced two 3-vectors by their coordinates in the (eI)(\vec{e}_{I}) frame as V=((πζ2πσ2)1/2,0,πσ)\vec{V}=((\pi_{\zeta}^{2}-\pi_{\sigma}^{2})^{1/2},0,{\pi_{\sigma}}) and W=(πσsinσ,πζcosσ,(πζ2πσ2)1/2sinσ)\vec{W}=({\pi_{\sigma}}\sin\sigma,{\pi_{\zeta}}\cos\sigma,-(\pi_{\zeta}^{2}-\pi_{\sigma}^{2})^{1/2}\sin\sigma). By construction, V\vec{V} and W\vec{W} are orthogonal for the Euclidean scalar product. Therefore, the expression for D\vec{D} in (32) is the parallel-perpendicular decomposition of D\vec{D} with respect to S\vec{S}, in the frame (e1,e2,e3)(\vec{e}_{1}\phantom{}^{\prime},\vec{e}_{2}\phantom{}^{\prime},\vec{e}_{3}) obtained by rotating (e1,e2,e3)(\vec{e}_{1},\vec{e}_{2},\vec{e}_{3}) by an angle σ\sigma around the third axis. The angle ζ\zeta is then simply that which gives the direction of the perpendicular part of D\vec{D}, as measured from the e2\vec{e}_{2}\phantom{}^{\prime}-axis. This is illustrated in Fig. 2.

Refer to caption
Figure 2: Physical interpretation of the symplectic variables (σ,πσ,ζ,πζ)(\sigma,{\pi_{\sigma}},\zeta,{\pi_{\zeta}}) in the 3-dimensional euclidean subspace (e1,e2,e3)(\vec{e}_{1},\vec{e}_{2},\vec{e}_{3}).

If there is an SSC, then S2=DS=0S_{\star}^{2}=\vec{D}\cdot\vec{S}=0 (as argued in the last section) and, therefore, the parallel part of D\vec{D} vanishes. Hence, with any SSC the mass dipole D\vec{D} is purely orthogonal to S\vec{S} and given by the orange vector in Fig. 2. In addition, under any SSC, formulae (29) can be even more simplified by noting that πσ/πζ[0,1]{\pi_{\sigma}}/{\pi_{\zeta}}\in[0,1]. By introducing an angle χ\chi such that cosχ=πσ/πζ\cos\chi={\pi_{\sigma}}/{\pi_{\zeta}}, Eqs. (29) become

S1\displaystyle S^{1} =πζsinχcosσ,\displaystyle={\pi_{\zeta}}\sin\chi\cos\sigma\,, (33a)
S2\displaystyle S^{2} =πζsinχsinσ,\displaystyle={\pi_{\zeta}}\sin\chi\sin\sigma\,, (33b)
S3\displaystyle S^{3} =πζcosχ,\displaystyle={\pi_{\zeta}}\cos\chi\,, (33c)
D1\displaystyle D^{1} =πζ2S2(cosχsinζcosσ+cosζsinσ),\displaystyle=\sqrt{\pi_{\zeta}^{2}-S_{\circ}^{2}}\bigl{(}\cos\chi\sin\zeta\cos\sigma+\cos\zeta\sin\sigma\bigr{)}\,, (33d)
D2\displaystyle D^{2} =πζ2S2(cosχsinζsinσcosζcosσ),\displaystyle=\sqrt{\pi_{\zeta}^{2}-S_{\circ}^{2}}\bigl{(}\cos\chi\sin\zeta\sin\sigma-\cos\zeta\cos\sigma\bigr{)}\,, (33e)
D3\displaystyle D^{3} =πζ2S2sinχsinζ.\displaystyle=-\sqrt{\pi_{\zeta}^{2}-S_{\circ}^{2}}\sin\chi\sin\zeta\,. (33f)

These formulae clearly show that S\vec{S} and D\vec{D} are two vectors of (Euclidean) norm πζ{\pi_{\zeta}} and (πζ2S2)1/2(\pi_{\zeta}^{2}-S_{\circ}^{2})^{1/2}, respectively, along with some rotation parametrized by the angles σ,χ,ζ\sigma,\chi,\zeta. In fact, if I(α)\mathcal{R}_{I}(\alpha) denotes the 3×33\times 3 matrix of a rotation of angle α\alpha with respect to the axis eI\vec{e}_{I}, then Eqs. (33) can be put into the compact form

S=πζ3(σ)2(χ)e3andD=πζ2S23(σ)2(χ)3(ζ)e2.\vec{S}={\pi_{\zeta}}\mathcal{R}_{3}(\sigma)\mathcal{R}_{2}(\chi)\vec{e}_{3}\quad\text{and}\quad\vec{D}=\sqrt{\pi_{\zeta}^{2}-S_{\circ}^{2}}\mathcal{R}_{3}(\sigma)\mathcal{R}_{2}(\chi)\mathcal{R}_{3}(\zeta)\vec{e}_{2}\,. (34)

Equations (34), along with the relation cosχ=πσ/πζ\cos\chi=\pi_{\sigma}/\pi_{\zeta}, are equivalent to Eqs. (33), with the advantage of making all aforementioned properties particularly clear: norms of S\vec{S} and V\vec{V}, interpretation of the angles (σ,ζ)\sigma,\zeta) and orthogonality properties. In addition, equation (34) that (σ,ζ,χ)(\sigma,\zeta,\chi), are the 3-2-3 Euler-angle parameterization Heard (2008) of D\vec{D} (and of S\vec{S} with ζ=0\zeta=0), in the spatial triad (eI)a(e_{I})^{a}. In particular, the momentum πζ\pi_{\zeta} is the norm of the spin 3-vector S\vec{S} in that frame, therefore, πζ0\pi_{\zeta}\neq 0 and Eqs. (29)–(30) are all well-posed.

III.3 Summary: 12D, covariant, non-degenerate formulation on 𝒩\mathcal{N}

To summarize, at this stage the Hamiltonian system (𝒩,Λ,H)(\mathcal{N},\Lambda,H) that we are considering is made of (i) the 12D phase space 𝒩\mathcal{N}; (ii) the 6 pairs of canonical171717Stating that the coordinates are canonical tells everything there is to know about the Poisson structure Λ\Lambda: it is non-degenerate and the Poisson brackets are the canonical ones (Kronecker symbols), as explained in App. D. coordinates (xα,σ,ζ,πα,πσ,πζ)(x^{\alpha},\sigma,\zeta,\pi_{\alpha},\pi_{\sigma},\pi_{\zeta}) and (iii) the Hamiltonian obtained by inserting Eq. (16) into Eq. (12), namely

H(xα,σ,ζ,πα,πσ,πζ)=12gαβπαπβ+12gαβπαωβCDSCD,H(x^{\alpha},\sigma,\zeta,\pi_{\alpha},\pi_{\sigma},\pi_{\zeta})=\frac{1}{2}g^{\alpha\beta}\pi_{\alpha}\pi_{\beta}+\frac{1}{2}g^{\alpha\beta}\pi_{\alpha}\omega_{\beta CD}S^{CD}\,, (35)

where xαx^{\alpha} are hidden in gαβg^{\alpha\beta} and ωβCD\omega_{\beta CD}, πα\pi_{\alpha} appear explicitly, and the four spin variables (σ,πσ,ζ,πζ)(\sigma,\pi_{\sigma},\zeta,\pi_{\zeta}) are hidden in SCDS^{CD} through formulae (29). We emphasize that the Casimirs/spin norms S,SS_{\circ},S_{\star} are fixed parameters in the formulation (not phase space functions), and that the particle’s mass μ\mu is the conserved value of the Hamiltonian itself along any solution to Hamilton’s equations. trajectory

This formulation is “covariant” in the sense that expression (35) holds for any choice of spacetime coordinates and any choice of background tetrad. To see why, perform a change of coordinates xαx~αx^{\alpha}\mapsto\tilde{x}^{\alpha} on the spacetime (,gab)(\mathscr{E},g_{ab}). It naturally induces a change xα(τ¯)x~(τ¯αx^{\alpha}(\bar{\tau})\mapsto\tilde{x}(\bar{\tau}^{\alpha} in the worldline’s parameterization spacetime. In general, the new coordinates x~α\tilde{x}^{\alpha} will not be canonically conjugated to the old momenta πα\pi_{\alpha}. However, consider the new momenta π~α(z,π)\tilde{\pi}^{\alpha}(z,\pi) such that

π~βπγx~αxγ=δβα.\frac{\partial\tilde{\pi}_{\beta}}{\partial\pi_{\gamma}}\frac{\partial\tilde{x}^{\alpha}}{\partial x^{\gamma}}=\delta^{\alpha}_{\beta}\,. (36)

Then, from the Leibniz rule, {x~α,π~α}=δβα\{\tilde{x}^{\alpha},\tilde{\pi}_{\alpha}\}=\delta^{\alpha}_{\beta}, and the new pairs (x~α,π~β)(\tilde{x}^{\alpha},\tilde{\pi}_{\beta}) are conjugated. In addition, the metric components gαβg^{\alpha\beta} and the connection 1-form components ωαBC\omega_{\alpha BC} change according to the usual chain rule characterizing tensors, and the whole expression in (35) is seen to be invariant under the change of spacetime coordinates.

Similarly, suppose a different tetrad (say (e~A)a(\tilde{e}_{A})^{a} instead of (eA)a(e_{A})^{a}) is chosen for the definition of the spin components SABS^{AB} and the Ricci coefficients. Since, (eA)a(e_{A})^{a} and (e~A)a(\tilde{e}_{A})^{a} are orthonormal tetrads, therefore, (eB)a=LBA(e~A)a(e_{B})^{a}=L^{A}_{\phantom{A}B}(\tilde{e}_{A})^{a}, where LBAL^{A}_{\phantom{A}B} is an arbitrary Lorentz transformation. Then, a direct calculation from their definition shows that the 1-form component ωβCD\omega_{\beta CD} becomes LCALDBω~αABL^{A}_{\phantom{A}C}L^{B}_{\phantom{B}D}\tilde{\omega}_{\alpha AB}, while the spin components in the new tetrad are S¯CD=LACLBDSAB\bar{S}^{CD}=L^{C}_{\phantom{C}A}L^{D}_{\phantom{D}B}S^{AB}. Therefore, since of a Lorentz transformation verifies ηABLCALDB=ηCD\eta_{AB}L^{A}_{\phantom{A}C}L^{B}_{\phantom{B}D}=\eta_{CD} by definition, we obtain ωβCDSCD=ω~βCDS~CD\omega_{\beta CD}S^{CD}=\tilde{\omega}_{\beta CD}\tilde{S}^{CD}, and the expression (35) is left unchanged, as claimed.

IV Constraints and invariants

We have reduced the Hamiltonian formulation on the initial 14D phase space \mathcal{M} to a 12D leaf 𝒩\mathcal{N} with no degeneracies. However, recall that the number of independent unknowns in the linearized MPTD + TD SSC system is 10 and not 12, cf. Sec. 9. In this section, we implement the TD SSC into the formulation on 𝒩\mathcal{N} (cf. the summary in Sec. III.3) by treating it as Hamiltonian constraints. This is done in Sec. IV.1, resulting in a 10D phase space, denoted 𝒫\mathcal{P}, wherein our final result is formulated, as summarized in Sec. IV.2. Lastly, the link between invariants of the MPTD system and integrals of motion of our 10D Hamiltonian formulation is discussed in Sec. IV.3.

IV.1 Algebraic constraints from the TD SSC

The Hamiltonian system we are considering is described by the Hamiltonian (35) and the TD SSC in the form pαSαβ=0p_{\alpha}S^{\alpha\beta}=0. From a phase space point of view, the TD SSC is a set of four algebraic equations relating the 12 canonical phase space variables on 𝒩\mathcal{N} (or the 14 non-canonical variables of \mathcal{M}, equivalently). Therefore, we are in the presence of a constrained Hamiltonian system181818All our results on constrained Hamiltonian systems will follow Sec. I.5 of Arnold et al. (2006) (see also Deriglazov (2022)).. We will now explain how to incorporate these constraints onto 𝒩\mathcal{N}, thus reducing the number of phase space dimensions, and restricting to the only part of 𝒩\mathcal{N} where the physically well-defined trajectories lie.

First, we write the TD SSC in terms of the canonical variables on 𝒩\mathcal{N}. Projecting Eq. (6) onto the tetrad, the TD SSC generates the following four algebraic equations

CB(z,π,S,D):=πASAB=0,C^{B}(z,\pi,S,D):=\pi_{A}S^{AB}=0\,, (37)

where, πA:=(eA)απα\pi_{A}:=(e_{A})^{\alpha}\pi_{\alpha} is a function of (xα,πα)(x^{\alpha},\pi_{\alpha}), and SABS^{AB} is, as always, a function of (σ,πσ,ζ,πζ)(\sigma,\pi_{\sigma},\zeta,\pi_{\zeta}) via Eqs. (18) and (29). Now let 𝒯\mathcal{T} be the sub-manifold of 𝒩\mathcal{N} defined by the four algebraic equations (37), and let us explore its properties.

IV.1.1 Invariance and dimension of 𝒯{\mathcal{T}}

First, we show that 𝒯\mathcal{T} is invariant under the flow of the Hamiltonian (35). Indeed, one can establish using the Leibniz rule and the brackets (17) that191919This formula is best derived using the intermediary brackets {xα,πA}=(eA)α\{x^{\alpha},\pi_{A}\}=(e_{A})^{\alpha} and {πA,πB}=ηCDπD(ωBCAωACB)\{\pi_{A},\pi_{B}\}=\eta^{CD}\pi_{D}(\omega_{BCA}-\omega_{ACB}) which follow from Eqs. (16) and (11), as well as the commutation relation satisfied by the tetrad vectors, see Eq. (3.4.23) in Wald (1984).

dCAdλ={CA,H}=ΩBACB,whereΩBA:=ηADπCωCBD.\frac{\mathrm{d}C^{A}}{\mathrm{d}\lambda}=\{C^{A},H\}=\Omega^{A}_{\phantom{A}B}C^{B}\,,\quad\text{where}\quad\Omega^{A}_{\phantom{A}B}:=\eta^{AD}\pi^{C}\omega_{CBD}\,. (38)

Therefore, any solution to the equations of motion that starts in 𝒯\mathcal{T} (i.e., that satisfies CA=0C^{A}=0 at initial time) will remain in 𝒯\mathcal{T} for all times. That 𝒯\mathcal{T} be invariant is expected: the linearized MPTD system itself is obtained from the TD SSC. From a purely computational viewpoint, we could stop here and evolve the dynamics under the Hamiltonian (35) along with a set of initial conditions such that CA=0C^{A}=0. Then the trajectory in phase space will be confined to the constraint surface202020Even though, in any numerical scheme, the discretization of the equations of motion will necessarily make the solution to venture off 𝒯{\mathcal{T}}, resulting in possibly spurious chaos from the extra dimensions and lack of first integrals., and thus make sense from a physical point of view. However, to show that the system is integrable, in the classical, Liouville-Arnold sense Arnold et al. (2006); Liouville (1855) we need to go further, as these invariants of motion need to be in involution with respect to the symplectic structure on 𝒯{\mathcal{T}}, not that on the unconstrained (bigger) phase-space 𝒩\mathcal{N}.

The Hamiltonian (35) defines a 6D system (12D phase space), so we may be tempted to say that the constrained system is a 4D one (12 minus 4 algebraic equations). However, out of the four algebraic relations obtained by setting B=0,,3B=0,\ldots,3 in Eq. (37), only three are linearly independent. To show this, write the four equations (37) explicitly, with the Euclidean notations introduced earlier:

C0\displaystyle\,\,C^{0} =πD=0,\displaystyle=-\vec{\pi}\cdot\vec{D}=0\,, (39a)
C\displaystyle\vec{C} =π0Dπ×S=0.\displaystyle=\pi_{0}\vec{D}-\vec{\pi}\times\vec{S}=0\,. (39b)

Clearly, Eq. (39a) follows from Eqs. (39b). In addition, Eqs. (39b) readily implies S2:=SD=0S_{\star}^{2}:=\vec{S}\cdot\vec{D}=0. Therefore, the SSC (37) only holds on those symplectic leaves 𝒩\mathcal{N} where the second Casimir 𝒞\mathcal{C}_{\star} vanishes. This result is consistent with the statement made in Sec. III.1.3 about any SSC implying S=0S_{\star}=0. In particular, this means that from now on we have to work on the intersection

𝒫:=𝒯𝒩,\mathcal{P}:=\mathcal{T}\cap\mathcal{N}\,, (40)

where the Poisson structure will be non-degenerate (thanks to 𝒩\mathcal{N}) and the TD SSC will be automatically satisfied (thanks to 𝒯\mathcal{T}). As readily seen from the above considerations, 𝒫\mathcal{P} is 9-dimensional when the leaf 𝒩\mathcal{N} has 𝒞0\mathcal{C}_{\star}\neq 0, and 10-dimensional when 𝒞=0\mathcal{C}_{\star}=0. On the latter, only two out of the four constraints (39) are sufficient to define 𝒫\mathcal{P}, and any two can be chosen. It seems sensible to call 𝒫\mathcal{P} the physical phase space (whence the notation), because all the physically relevant solutions to the MPTD system will lie within 𝒫\mathcal{P}. From now on, we shall denote by 𝒫\mathcal{P} the intersection between the SSC sub-manifold 𝒯{\mathcal{T}} and a leaf 𝒩\mathcal{N} where (𝒞,𝒞)=(S,0)(\mathcal{C}_{\circ},\mathcal{C}_{\star})=(S_{\circ},0).

Refer to caption
Figure 3: Different phase space sub-manifold are necessary to lift the degeneracies associated to the existence of Casimir invariants and correctly implementing the TD SSC.

IV.1.2 Symplectic structure on 𝒫\mathcal{P}

As explained previously, we can consider only two constraints out of the four (39) to describe 𝒫\mathcal{P}. We denote them by C𝒜C^{\mathcal{A}} and CC^{\mathcal{B}}, where 𝒜\mathcal{A} and \mathcal{B} are two fixed numbers in (0,,3)(0,\ldots,3) (as opposed to abstract indices (A,B,C,D,)(A,B,C,D,\ldots) which can be summed over and take any value in (0,,3)(0,\ldots,3)). Coordinates on 𝒫\mathcal{P} can be chosen by taking any subset of 10 from the 12 canonical coordinates on 𝒩\mathcal{N} (just as we did when going from to \mathcal{M} to 𝒩\mathcal{N}).

The Hamiltonian system we consider is now (𝒫,Λ𝒫,H𝒫)(\mathcal{P},\Lambda_{\mathcal{P}},H_{\mathcal{P}}), where dim(𝒫)=10\text{dim}(\mathcal{P})=10 and, as before, the Hamiltonian H𝒫H_{\mathcal{P}} is obtained by writing HH (in Eq. (35))) in terms of the 10 coordinates on 𝒫\mathcal{P} (with the help of the two constraints CA=0,CB=0C^{A}=0,C^{B}=0 to remove the 2 discarded ones, if necessary). According to the classical theory of constrained Hamiltonian systems Deriglazov (2022), the Poisson structure Λ𝒫\Lambda_{\mathcal{P}} on 𝒫\mathcal{P} (inherited from that on 𝒩\mathcal{N}, is well-defined provided that the 𝒩\mathcal{N}-brackets {C𝒜,H}\{C^{\mathcal{A}},H\} and {C,H}\{C^{\mathcal{B}},H\} both vanish on 𝒫\mathcal{P}, i.e., that 𝒫\mathcal{P} is invariant by the flow of HH on 𝒩\mathcal{N}. As follows from Eq. (38), this is the case for us.

However, having a well-defined Λ𝒫\Lambda_{\mathcal{P}} does not mean that it is symplectic (non-degenerate). Again, relying on constrained Hamiltonian system theory Deriglazov (2022), Λ𝒫\Lambda_{\mathcal{P}} will be non-degenerate if and only if {CA,CB}0\{C^{A},C^{B}\}\neq 0 on 𝒫\mathcal{P}.212121In the more general case where there are more than two constraints, the condition is that the matrix of entries {CA,CB}\{C^{A},C^{B}\} be invertible on 𝒫\mathcal{P} Brown (2022). Using the Poisson brackets (11) and the relations (17), we find

{C𝒜,C}=ηCDπCπDS𝒜2π[𝒜C],\{C^{\mathcal{A}},C^{\mathcal{B}}\}=\eta^{CD}\pi_{C}\pi_{D}S^{\mathcal{A}\mathcal{B}}-2\pi^{[\mathcal{A}}C^{\mathcal{B}]}\,, (41)

Evaluated on 𝒫\mathcal{P}, the scalar (41) is numerically to μ2SAB-\mu^{2}S^{AB} up to quadratic-in-spin terms, with μ\mu the mass of the particle, and S𝒜S^{\mathcal{A}\mathcal{B}} is one particular tetrad component of the spin tensor, which is in general non-zero, at least not everywhere on 𝒫\mathcal{P}. Consequently, Λ𝒫\Lambda_{\mathcal{P}} is symplectic (non-degenerate). However, the induced Poisson brackets {,}𝒫\{,\}^{\mathcal{P}} on 𝒫\mathcal{P} will be different to those on 𝒩\mathcal{N} (denoted {,}\{,\} until now). In essence, the 𝒩\mathcal{N}-brackets must be corrected to ensure that the constraints are automatically satisfied for any calculation on 𝒫\mathcal{P}. Thankfully, it is possible to express the 𝒫\mathcal{P}-brackets in terms of those on 𝒩\mathcal{N} and the constraints. The relation is given by222222In the theory of singular Lagrangian systems, this formula is known as the Dirac bracket Brown (2022), introduced by Dirac in the context of quantum mechanics Dirac (1950). In our symplectic geometry approach (no Lagrangian, no Legendre transform), this formula is a consequence of how Poisson structures behave under restrictions/embeddings Deriglazov (2022). Arnold et al. (2006); Brown (2022)

{F,G}𝒫:={F,G}{F,C𝒜}(M1)𝒜{C,G},\{F,G\}^{\mathcal{P}}:=\{F,G\}-\{F,C^{\mathcal{A}}\}(M^{-1})_{\mathcal{A}\mathcal{B}}\{C^{\mathcal{B}},G\}\,, (42)

with M1M^{-1} denoting the matrix inverse of the two-by-two matrix M𝒜={C𝒜,C}M^{\mathcal{A}\mathcal{B}}=\{C^{\mathcal{A}},C^{\mathcal{B}}\}. On the right-hand side of Eq. (42), the 𝒩\mathcal{N}-Poisson brackets must be used first (for example using the 12 canonical coordinates on 𝒩\mathcal{N}), and then the overall result is simplified using the constraints and expressed it in terms of the 10 chosen variables on 𝒫\mathcal{P}. Notice that, by construction, for any function F:𝒫F:\mathcal{P}\rightarrow\mathbb{R}, one has {F,C𝒜}𝒫=0\{F,C^{\mathcal{A}}\}^{\mathcal{P}}=0 and {F,C}𝒫=0\{F,C^{\mathcal{B}}\}^{\mathcal{P}}=0, i.e., the two constraints are Casimirs of the symplectic structure Λ𝒫\Lambda_{\mathcal{P}}. In our case, Eq. (42) can be simplified even more, as we have M𝒜={C𝒜,C}𝕀2M^{\mathcal{A}\mathcal{B}}=\{C^{\mathcal{A}},C^{\mathcal{B}}\}\mathbb{I}_{2}, where 𝕀2\mathbb{I}_{2} is the two-by-two canonical symplectic matrix. Therefore, from (41), we obtain

{F,G}𝒫={F,G}+{F,C𝒜}{G,C}{F,C}{G,C𝒜}μ2S𝒜,\{F,G\}^{\mathcal{P}}=\{F,G\}+\frac{\{F,C^{\mathcal{A}}\}\{G,C^{\mathcal{B}}\}-\{F,C^{\mathcal{B}}\}\{G,C^{\mathcal{A}}\}}{\mu^{2}S^{\mathcal{A}\mathcal{B}}}\,, (43)

where μ2\mu^{2} is a function of the coordinates on 𝒫\mathcal{P} (equal to 2H𝒫-2H_{\mathcal{P}}), just like C𝒜,CC^{\mathcal{A}},C^{\mathcal{B}} and S𝒜S^{\mathcal{A}\mathcal{B}}.

IV.2 Summary of the formulation on 𝒫\mathcal{P}

We can now summarize our final Hamiltonian formulation of the linearized MPTD + TD SSC equations, in the most general way (i.e., for any background metric). A thorough application of this formalism can be found in a joined paper Ramond and Isoyama (2024) for the case of the Schwarzschild metric.

  • Endow 𝒫=10\mathcal{P}=\mathbb{R}^{10} with 10 coordinates chosen from the 12 canonical ones (xα,σ,ζ,πα,πσ,πζ)(x^{\alpha},\sigma,\zeta,\pi_{\alpha},\pi_{\sigma},\pi_{\zeta}) on 𝒩\mathcal{N}. These coordinates all have a natural physical interpretation in spacetime.

  • Build the Hamiltonian H𝒫H_{\mathcal{P}} from expressing Eq. (35) in terms of the 10 coordinates on 𝒫\mathcal{P}. This result in a covariant expression (arbitrary spacetime coordinates and tetrad field).

  • Choose two constraints C𝒜,CC^{\mathcal{A}},C^{\mathcal{B}} directly from TD SSC in (39), and compute the Poisson brackets {,}𝒫\{,\}^{\mathcal{P}} of any two phase space functions.

Doing this will result in a well-defined, covariant and non-degenerate Hamiltonian system, on a 10D phase space 𝒫\mathcal{P}. Any physical solution to the MPTD + TD SSC system (9) will correspond to a trajectory in 𝒫\mathcal{P}, where the TD SSC is automatically enforced at each point. These solutions are described by the law of motion (Hamilton’s equation)

dFdτ¯={F,H𝒫}𝒫,\frac{\mathrm{d}F}{\mathrm{d}\bar{\tau}}=\{F,H_{\mathcal{P}}\}^{\mathcal{P}}\,, (44)

where F:𝒫F:\mathcal{P}\rightarrow\mathbb{R} is any phase space function, H𝒫H_{\mathcal{P}} is the autonomous Hamiltonian and τ¯\bar{\tau} corresponds, physically, to the proper time per unit of the particle’s mass. The formulation is covariant and automatically implies the constancy of the spin norm SS_{\circ} (as a Casimir invariant on 𝒫\mathcal{P}) and of the particle’s mass μ\mu, as the conserved value of the Hamiltonian along any trajectory (μ2=2H𝒫\mu^{2}=-2H_{\mathcal{P}}).

Lastly, it is worth noting that, from Eqs. (38) and (43), we have {F,H𝒫}𝒫={F,H𝒫}\{F,H_{\mathcal{P}}\}^{\mathcal{P}}=\{F,H_{\mathcal{P}}\} for any function F:𝒫F:\mathcal{P}\rightarrow\mathbb{R}, i.e., the equations of motion on 𝒫\mathcal{P} can be computed thanks to Hamilton’s canonical equation on 𝒩\mathcal{N} and simplifying the result with the constraints. However, this is only because 𝒫\mathcal{P} is an invariant sub-manifold of the flow of the quantity H𝒫H_{\mathcal{P}} itself. In general, i.e., for any other function G:𝒫G:\mathcal{P}\rightarrow\mathbb{R}, one will have {F,G}𝒫{F,G}\{F,G\}^{\mathcal{P}}\neq\{F,G\}.

IV.3 Invariants of the MPTD system

With the final Hamiltonian formulation presented, there is only so much we can discuss if we want to remain general, i.e., independent of the metric background. Yet, it is still possible to adapt existing results in the literature regarding invariants of motion for the MPTD system itself. Since our Hamiltonian generates the same equations, we can expect these invariant to have something to do with integrals of motion for our system.

IV.3.1 Killing vectors and invariants

In the context of general relativity, most spacetime metrics of interest possesses Killing vectors, i.e., a vector field ξa\xi^{a} that satisfies Killing’s equation (aξb)=0\nabla_{(a}\xi_{b)}=0. It is well-known for the MPTD equations (2) that, if ξa\xi^{a} is such a Killing vector, then the quantity

Ξ:=pαξα+12Sαβαξβ,\Xi:=p_{\alpha}\xi^{\alpha}+\frac{1}{2}S^{\alpha\beta}\nabla_{\alpha}\xi_{\beta}\,, (45)

remains constant along the worldline of the particle. In the Hamiltonian formulation, this result translates naturally. Let ξα\xi^{\alpha} be the natural components a Killing field ξa\xi^{a}. In our setup, ξα\xi^{\alpha} are four functions of the phase space coordinates xαx^{\alpha}. Then combining the Hamiltonian (12) and the Poisson brackets (11) on \mathcal{M} lead to

{Ξ,H}=pαpβαξβ+12(αβξγRγβαδξδ)pαSβγ=0.\{\Xi,H\}=p^{\alpha}p^{\beta}\nabla_{\alpha}\xi_{\beta}+\frac{1}{2}(\nabla_{\alpha}\nabla_{\beta}\xi_{\gamma}-R_{\gamma\beta\alpha\delta}\xi^{\delta})p^{\alpha}S^{\beta\gamma}=0\,. (46)

where the first and second terms in the right-hand side vanishes independently because of Killing’s equation and the Kostant formula232323Kostant’s formula for Killing fields ξa\xi^{a} reads abξc=Rcbadξd\nabla_{a}\nabla_{b}\xi_{c}=R_{cbad}\xi^{d}, see, e.g., App. E of Ramond and Le Tiec (2021)., respectively. Now, these Killing invariants have an additional interesting property. Let us set Cα:=pβSβαC^{\alpha}:=p_{\beta}S^{\beta\alpha}, so that the TD SSC amounts to the vanishing of the vector CαC^{\alpha}. Then, for any vector field ξα\xi^{\alpha},

{Cα,Ξ}=2Sαβ(βξγ)pγ+Cββξα.\{C^{\alpha},\Xi\}=2S^{\alpha\beta}\nabla_{(\beta}\xi_{\gamma)}p^{\gamma}+C^{\beta}\partial_{\beta}\xi^{\alpha}\,. (47)

The first term on the right-hand side of (47) vanishes by Killing’s equation, and the second one vanishes on 𝒫\mathcal{P}, where Cβ=0C^{\beta}=0. Therefore, Killing invariants Ξ\Xi commute with the constraints induced by the TD SSC. Inserting Eqs. (46) and (47) into the definition of the brackets (43) on 𝒫\mathcal{P}, we find that for any Killing vector field ξα\xi^{\alpha} of the background spacetime,

{Ξ,H}𝒫=0,\{\Xi,H\}^{\mathcal{P}}=0\,, (48)

i.e., Ξ\Xi is also an invariant for the constrained system on 𝒫\mathcal{P}. In addition, we readily see from that if two Killing invariants are in involution in the full phase space, they remain so with respect to the symplectic form on 𝒫\mathcal{P}. We stress that these results hold for Killing invariants (or any scalar function built solely from them), and not, for additional invariants that may not be associated to Killing vector fields.

IV.3.2 Killing tensors and invariants

In addition to vector fields, Killing tensors of higher valence may exist and lead to invariants of motion Frolov et al. (2017). Of particular interest are the (symmetric) Killing-Stäckel tensors KabK_{ab}, such that (aKbc)=0\nabla_{(a}K_{bc)}=0, and the (antisymmetric) Killing-Yano tensors YabY_{ab} such that (aYb)c\nabla_{(a}Y_{b)c}=0. For our purposes, we take inspiration from the pioneering work of Rüdiger on Killing invariants for the MPTD system Rüdiger (1981, 1983) (see also the more recent and complete work on the matter in Compère and Druart (2022)). In these works, the authors look for scalar quantities built from pa,Sabp_{a},S^{ab} and the geometry that are conserved under the dipolar MPTD system under the TD SSC. As a result, they find that in spacetimes with a Killing-Yano tensor YabY_{ab}, two such invariants exist. The first one is

𝔎:=14εabcdYabScd.\mathfrak{K}:=\frac{1}{4}\varepsilon_{abcd}Y^{ab}S^{cd}\,. (49)

In contrast to Killing vector invariants (45) which are exactly conserved for the general dipolar MPTD system (without any SSC), the Rüdiger invariant 𝔎\mathfrak{K} is referred to as a “quasi-invariant”, since its derivative along \mathscr{L} is not zero, but of quadratic order in spin: u𝔎=O(ϵ2)\nabla_{u}\mathfrak{K}=O(\epsilon^{2}), in the sense of (8). The other Rüdiger invariant, which we denote by 𝔔\mathfrak{Q}, is

𝔔:=Kabpapb+4εade[bYc]e𝔶aSdbpc,\mathfrak{Q}:=K^{ab}p_{a}p_{b}+4\varepsilon_{ade[b}Y^{e}_{~{}c]}\mathfrak{y}^{a}S^{db}p^{c}\,, (50)

where KabK^{ab} is the Killing-Stäckel tensor Carter (2009) associated to YabY_{ab} and 𝔶a\mathfrak{y}^{a} is the Hodge dual of aYbc\nabla_{a}Y_{bc}, both uniquely defined from YabY_{ab} through Kab:=YcaYbcK^{ab}:=Y^{a}_{~{}c}Y^{bc} and aYbc:=εabcd𝔶d\nabla_{a}Y_{bc}:=\varepsilon_{abcd}\mathfrak{y}^{d}. Again, 𝔔\mathfrak{Q} is only conserved at linear order in spin.

Interestingly, in the non-spinning (geodesic) case Sab=0S^{ab}=0, one has 𝔎=0\mathfrak{K}=0 and 𝔔=Kabpapb\mathfrak{Q}=K^{ab}p_{a}p_{b}, which only requires the existence of a Killing-Stäckel tensor KabK_{ab}, not a Killing-Yano tensor Dietz and Rüdiger (1981); Dietz and Rudiger (1982). In the Kerr spacetime, KabpapbK^{ab}p_{a}p_{b} is the well-known Carter constant Carter (1968) that ensures the integrability of geodesics.

IV.3.3 Covariant notion of angular momentum

Lastly, let us mention that the two Rüdiger invariants 𝔎,𝔔\mathfrak{K},\mathfrak{Q} can be given a physical interpretation related to a covariant notion of angular momentum tied to the Killing-Yano tensor. This was discussed in Frolov et al. (2017); Witzany (2019); Compère and Druart (2022), but first pointed out by Dietz and Rüdiger in Dietz and Rüdiger (1981); Dietz and Rudiger (1982). Let us define the following quantities

STDb:=12εabcdp¯aScdandLb:=paYab,\quad S^{b}_{\text{\tiny{TD}}}:=\frac{1}{2}\varepsilon^{abcd}\bar{p}_{a}S_{cd}\quad\text{and}\quad L^{b}:=p_{a}Y^{ab}\,, (51)

where STDaS^{a}_{\text{\tiny{TD}}} is the spin 4-vector measured by an observer of four-velocity p¯a\bar{p}_{a} (recall the general decomposition in IV.1). Then LaL^{a} can be interpreted as a covariant notion of angular momentum in the sense that, in the Schwarzschild spacetime, its spatial components admit the correct Newtonian limit L=r×p\vec{L}=\vec{r}\times\vec{p} at spatial infinity, like in a spherically symmetric Newtonian potential. This analogy is explored in our follow-up paper within Schwarzschild Ramond and Isoyama (2024), where explicit formulae are given.

Combining the definitions (51) with the decomposition (5) (with fa=p¯af^{a}=\bar{p}^{a}), we can re-write the two Rüdiger invariants (49),(50) as

𝔎\displaystyle\mathfrak{K} =1μLaSTDa+O(Ca),\displaystyle=-\frac{1}{\mu}L_{a}S^{a}_{\text{\tiny{TD}}}+O(C^{a})\,, (52a)
𝔔\displaystyle\mathfrak{Q} =LaLa2μYabSTDa𝔶b+𝔎pa𝔶a+O(Ca),\displaystyle=L^{a}L_{a}-2\mu Y_{ab}S_{\text{\tiny{TD}}}^{a}\mathfrak{y}^{b}+\mathfrak{K}\,p_{a}\mathfrak{y}^{a}+O(C^{a})\,, (52b)

where the fact that LaL^{a} and STDaS^{a}_{\text{\tiny{TD}}} are orthogonal to pap_{a} was used, and O(Ca)O(C^{a}) describes a term proportional to Cb:=paSabC^{b}:=p_{a}S^{ab}, which vanishes under the TD SSC (6). From Eq. (52), we thus can interpret the first invariant 𝔎\mathfrak{K} as the projection of the spin 4- vector onto the angular momentum 4-vector LaL^{a}, and the second invariant 𝔔\mathfrak{Q} as the (spin-corrected) squared norm of the LaL^{a}. Interestingly Compère and Druart (2022), in the Kerr spacetime one has 𝔶a=(t)a\mathfrak{y}^{a}=(\partial_{t})^{a} so the third term in (52b) is actually independently conserved from 𝔔\mathfrak{Q}, as the product 𝔎(E)\mathfrak{K}\cdot(-E), where E:=πtE:=-\pi_{t} is the conserved total energy of the particle, as discussed in Sec. V.3 below.

V Application: integrability around black holes

In this section, we apply our Hamiltonian formulation to the case of a Schwarzschild and Kerr background. In both cases, the resulting system is found to be integrable, in the usual sense of Liouville, i.e, as defined in the classical reference Arnold et al. (2006) (or Hinderer and Flanagan (2008) for the general relativistic case including non-compact dimensions). Even though the integrability in Kerr evidently implies that in Schwarzschild as a particular case, the latter admits a number of unique features worth discussing first (in Sec. V.2) before moving on to the general, Kerr case (in Sec. V.3).

V.1 Preliminaries

The matter of integrability of the motion of spinning particles in black hole spacetimes has been studied in the past, but its current status is still, mostly, in a state of conjecture, and based on numerical studies. To be as clear as possible, we start by recalling the classical definition of integrable Hamiltonian systems, following classical references, like Arnold et al. (2006), but see also Hinderer and Flanagan (2008) and references therein for subtleties in the relativistic case.

Let (𝒫,{,}𝒫,H𝒫)(\mathcal{P},\{,\}^{\mathcal{P}},H_{\mathcal{P}}) be a Hamiltonian system, where dim(𝒫)=2N\text{dim}(\mathcal{P})=2N\in\mathbb{N} and where {,}𝒫\{,\}^{\mathcal{P}} is a non-degenerate Poisson bracket on 𝒫\mathcal{P}. A first integral of the system (𝒫,{,}𝒫,H𝒫)(\mathcal{P},\{,\}^{\mathcal{P}},H_{\mathcal{P}}) is a function :𝒫\mathcal{I}:\mathcal{P}\rightarrow\mathbb{R} such that {,H𝒫}𝒫=0\{\mathcal{I},H_{\mathcal{P}}\}^{\mathcal{P}}=0. Two first integrals (1,2)(\mathcal{I}_{1},\mathcal{I}_{2}) are in involution if they satisfy {1,2}𝒫=0\{\mathcal{I}_{1},\mathcal{I}_{2}\}^{\mathcal{P}}=0. A set of nn\in\mathbb{N} first integrals (1,,n)(\mathcal{I}_{1},\ldots,\mathcal{I}_{n}) is said to be linearly independent if the matrix constructed from their gradients has maximal rank (equal to nn) on 𝒫\mathcal{P}. The system (𝒫,{,}𝒫,H𝒫)(\mathcal{P},\{,\}^{\mathcal{P}},H_{\mathcal{P}}) is integrable if it admits (i) N=dim(𝒫)/2N=\text{dim}(\mathcal{P})/2 first integrals that are (ii) linearly independent and (iii) in pairwise involution. It is crucial that all three items above be verified. In particular, we stress that a Casimir invariant can be used as a first integral, as its mere existence violates the requirement of non-degeneracy242424The proof of integrability in Schwarzschild in Kunst et al. (2016) is thus misguided, as the list of first integrals includes a Casimir of SO(3). of {,}𝒫\{,\}^{\mathcal{P}}. Similarly, the particle mass μ2=\mu^{2}=, as the on-shell value of our autonomous Hamiltonian, is readily promoted to a first integral.252525This explains why the Hamiltonian formulation of Witzany et al. (2019), where μ\mu is an ad hoc external parameter, is unadapted for our goals. Lastly, we emphasize that everything that follows (like the rest of the article) is done consistently at linear order in spin, i.e., in the small parameter ϵ=S/μM\epsilon=S_{\circ}/\mu M, with MM the black hole mass (cf. the discussion in Sec. I.4).

With all necessary precautions now taken, we can now apply the above result with N=5N=5 to our Hamiltonian formalism, summarized in Sec. IV.2, in the Schwarzschild and Kerr spacetimes.

V.2 Integrability in the Schwarzschild spacetime

V.2.1 Geometry and Hamiltonian

In this subsection, we let (,gab)(\mathscr{E},g_{ab}) be the Schwarzschild spacetime. To specify the canonical variables on 𝒩\mathcal{N}, by making a choice of coordinates covering \mathscr{E} and a choice of orthonormal tetrad. We choose the Schwarzschild-Droste (SD) coordinates262626We chose the SD coordinates for simplicity and because they are adapted to the spherical symmetry. By covariance of the Hamiltonian, any other coordinate system can be used. xα=(t,r,θ,ϕ)x^{\alpha}=(t,r,\theta,\phi), in which the contravariant components gαβg^{\alpha\beta} of the metric read Gourgoulhon (2021)

gtt=1f(r),grr=f(r),gθθ=1r2,gϕϕ=csc2θr2,g^{tt}=-\frac{1}{f(r)}\,,\quad g^{rr}=f(r)\,,\quad g^{\theta\theta}=\frac{1}{r^{2}}\,,\quad g^{\phi\phi}=\frac{\csc^{2}\theta}{r^{2}}\,, (53)

where f(r)=12M/rf(r)=1-2M/r, MM being the Schwarzschild mass parameter. Because the metric is diagonal, an orthonormal tetrad can be obtained by normalizing the natural basis (α)a(\partial_{\alpha})^{a} associated to the coordinates xαx^{\alpha}. We find, in terms of this basis, the tetrad

(e0)a=1f(t)a,(e1)a=f(r)a,(e2)a=1r(θ)a,(e3)a=cscθr(ϕ)a.(e_{0})^{a}=\frac{1}{\sqrt{f}}(\partial_{t})^{a}\,,\quad(e_{1})^{a}=\sqrt{f}(\partial_{r})^{a}\,,\quad(e_{2})^{a}=\frac{1}{r}(\partial_{\theta})^{a}\,,\quad(e_{3})^{a}=\frac{\csc\theta}{r}(\partial_{\phi})^{a}\,. (54)

For the Hamiltonian (35) we also need the connection 1-forms ωaBC=(eB)ba(eC)b\omega_{aBC}=(e_{B})_{b}\nabla_{a}(e_{C})^{b}, and, more specifically, their components in the natural basis, ωαBC\omega_{\alpha BC}, which are explicitly needed for the transformation (16). A direct calculation shows that, out of the 24 Ricci coefficients, only eight are non vanishing, and found to be (see also Müller and Grave (2009) but with different conventions):

ωt01=Mr2,ωθ12=f,ωϕ31=fsinθ,ωϕ23=cosθ,\omega_{t01}=-\frac{M}{r^{2}}\,,\quad\omega_{\theta 12}=-\sqrt{f}\,,\quad\omega_{\phi 31}=\sqrt{f}\sin\theta\,,\quad\omega_{\phi 23}=-\cos\theta\,, (55)

and the four other obtained by the algebraic antisymmetry ωαBC=ωαCB\omega_{\alpha BC}=-\omega_{\alpha CB}. Inserting formulae (53) and (54) into the generic Hamiltonian (35), we obtain

H𝒩=πt22f+fπr22+12r2(πθ2+πϕ2sin2θ)πϕcosθr2sin2θS1+πϕfr2sinθS2fπθr2S3+Mπtr2fD1,H_{\mathcal{N}}=-\frac{\pi_{t}^{2}}{2f}+\frac{f\pi_{r}^{2}}{2}+\frac{1}{2r^{2}}\left(\pi_{\theta}^{2}+\frac{\pi_{\phi}^{2}}{\sin^{2}\theta}\right)-\frac{\pi_{\phi}\cos\theta}{r^{2}\sin^{2}\theta}S^{1}+\frac{\pi_{\phi}\sqrt{f}}{r^{2}\sin\theta}S^{2}-\frac{\sqrt{f}\pi_{\theta}}{r^{2}}S^{3}+\frac{M\pi_{t}}{r^{2}f}D^{1}\,, (56)

where it is understood that one replaces the spin components S,D\vec{S},\vec{D} using Eqs. (29). By construction, the Hamiltonian (56) generates the linear-in-spin dynamics of a spinning particle in a Schwarzschild spacetime. It depends on 3 parameters M,S,SM,S_{\circ},S_{\star} and 12 phase space variables arranged as symplectic pairs: the six degrees of freedom (t,r,θ,ϕ,σ,ζ)(t,r,\theta,\phi,\sigma,\zeta) and their conjugated momenta (πt,πr,πθ,πϕ,πσ,πζ)(\pi_{t},\pi_{r},\pi_{\theta},\pi_{\phi},\pi_{\sigma},\pi_{\zeta}). Lastly, the two constraints coming from the TD SSC (37), as prescribed in Sec. IV.2, are chosen to be setting 𝒜=0\mathcal{A}=0 and =1\mathcal{B}=1. In terms of the phase space coordinates, they read, respectively,

C0\displaystyle C^{0} =fπrD1πθrD2πϕrsinθD3=0,\displaystyle=-\sqrt{f}\pi_{r}D^{1}-\frac{\pi_{\theta}}{r}D^{2}-\frac{\pi_{\phi}}{r\sin\theta}D^{3}=0\,, (57a)
C1\displaystyle C^{1} =πtfD1+πϕrsinθS2πθrS3=0,\displaystyle=\frac{\pi_{t}}{\sqrt{f}}D^{1}+\frac{\pi_{\phi}}{r\sin\theta}S^{2}-\frac{\pi_{\theta}}{r}S^{3}=0\,, (57b)

where, again, S,D\vec{S},\vec{D} are functions of (σ,πσ,ζ,πζ)(\sigma,\pi_{\sigma},\zeta,\pi_{\zeta}) via Eqs. (29).

V.2.2 First integrals

The symmetries of the Schwarzschild spacetime are in correspondence with four Killing vectors. Given in the natural basis (α)a(\partial_{\alpha})^{a} associated to the SD coordinates xα=(t,r,θ,ϕ)x^{\alpha}=(t,r,\theta,\phi), they read

ξ(t)a\displaystyle\xi_{(t)}^{a} =(t)a,\displaystyle=(\partial_{t})^{a}\,, (58a)
ξ(x)a\displaystyle\xi_{(x)}^{a} =sinϕ(θ)acosϕcotθ(ϕ)a,\displaystyle=-\sin\phi(\partial_{\theta})^{a}-\cos\phi\cot\theta\,(\partial_{\phi})^{a}\,, (58b)
ξ(y)a\displaystyle\xi_{(y)}^{a} =cosϕ(θ)asinϕcotθ(ϕ)a,\displaystyle=\cos\phi(\partial_{\theta})^{a}-\sin\phi\cot\theta\,(\partial_{\phi})^{a}\,, (58c)
ξ(z)a\displaystyle\xi_{(z)}^{a} =(ϕ)a.\displaystyle=(\partial_{\phi})^{a}\,. (58d)

The four Killing vectors (58) imply the existence of four Killing invariants of the form (45). In terms of the canonical coordinates on 𝒩\mathcal{N}, these read

E\displaystyle E :=πt,\displaystyle:=-\pi_{t}\,, (59a)
Jx\displaystyle J_{x} :=sinϕπθcosϕcotθπϕ+cosϕcscθS1,\displaystyle:=-\sin\phi\,\pi_{\theta}-\cos\phi\cot\theta\,\pi_{\phi}+\cos\phi\csc\theta\,S^{1}\,, (59b)
Jy\displaystyle J_{y} :=cosϕπθsinϕcotθπϕ+sinϕcscθS1,\displaystyle:=\cos\phi\,\pi_{\theta}-\sin\phi\cot\theta\,\pi_{\phi}+\sin\phi\csc\theta\,S^{1}\,, (59c)
Jz\displaystyle J_{z} :=πϕ,\displaystyle:=\pi_{\phi}\,, (59d)

where the first invariant, associated to the Killing vector ξ(t)a-\xi_{(t)}^{a}, is denoted EE according to its interpretation as the total energy of the particle as measured by an observer at rest at infinity Gourgoulhon (2021), and the last three are denoted according to their classical interpretation as three Cartesian components of a conserved angular-momentum 3-vector J\vec{J}. This vector is orthogonal to an invariant Euclidean plane to which the motion of the particle stays close, but in which it is not confined.272727Unlike in the geodesic case where the motion is always planar. We refer to the details of Ramond and Isoyama (2024) where this feature is fully exploited to derive a canonical system based on our Hamiltonian formulation. We emphasize that, in spite of their notation, these invariants include both orbital and spin contributions hidden in the πα,πσ\pi_{\alpha},\pi_{\sigma} variables (cf. Eq. (16)).

V.2.3 Invariants from the Killing-Yano tensor

The Schwarzschild spacetime admits a Killing-Yano tensor Tanaka et al. (1996); Açık et al. (2010); Compère and Druart (2022), whose components in the natural basis are given by

Yab=2rsinθ(θ)[a(ϕ)b],Y^{ab}=\frac{2}{r\sin\theta}(\partial_{\theta})^{[a}(\partial_{\phi})^{b]}\,, (60)

and it is manifestly antisymmetric. Squaring the Killing-Yano tensor generates a Killing-Stäckel tensor KabK^{ab} though the combination Kab:=YacYcbK^{ab}:=Y^{ac}Y^{b}_{\phantom{c}c}. Using (60), we find

Kab=(θ)a(θ)b+1sin2θ(ϕ)a(ϕ)b.K^{ab}=(\partial_{\theta})^{a}(\partial_{\theta})^{b}+\frac{1}{\sin^{2}\!\theta}\,(\partial_{\phi})^{a}(\partial_{\phi})^{b}\,. (61)

When we substitute Eqs. (16), (60) and (61) within Eqs. (49) and (50), we obtain the explicit expressions for the Rüdiger quasi-invariants 𝔎\mathfrak{K} and 𝔔\mathfrak{Q} in terms of the canonical coordinate on 𝒩\mathcal{N}:

𝔎\displaystyle\mathfrak{K} =rD1,\displaystyle=rD^{1}\,, (62a)
𝔔\displaystyle\mathfrak{Q} =πθ2+πϕ2sin2θ2πϕcosθsin2θS1,\displaystyle=\pi_{\theta}^{2}+\frac{\pi_{\phi}^{2}}{\sin^{2}\theta}-\frac{2\pi_{\phi}\cos\theta}{\sin^{2}\theta}S^{1}\,, (62b)

where D1D^{1} is a function of (σ,πσ,ζ,πζ)(\sigma,\pi_{\sigma},\zeta,\pi_{\zeta}) via Eq. (29).

V.2.4 Integrability

To prove the integrability, we need to (i) exhibit five first integrals, which are (ii) linearly independent on 𝒫\mathcal{P} and (iii) in pairwise involution.

The first item is easy to verify. There are seven first integral “candidates” obtain from Killing symmetries: four in Eqs. (59), two in Eqs. (62) and the Hamiltonian (56). The fact that they are (indeed) first integrals for H𝒫H_{\mathcal{P}} is also easy to check, as follows. First, H𝒫H_{\mathcal{P}} is trivially so as an autonomous Hamiltonian. Second, all four (E,Jx,Jy,Jz)(E,J_{x},J_{y},J_{z}) are first integrals by the general result (48) for Killing vectors. Third, a direct calculation reveals that

{𝔎,H𝒩}=fC0,and{𝔔,H𝒩}=0.\{\mathfrak{K},H_{\mathcal{N}}\}=-\sqrt{f}\,{C}^{0}\,,\quad\text{and}\quad\{\mathfrak{Q},H_{\mathcal{N}}\}=0\,. (63)

This equation reveals that the invariant 𝔎\mathfrak{K} is not a first integral on 𝒩\mathcal{N} (i.e., when the TD SSC is not applied), as should be expected Rüdiger (1981, 1983); Compère and Druart (2022). However, inserting Eqs. (63) into the 𝒫\mathcal{P}-brackets (43) and applying the constraints to the right-hand side gives {𝔎,H𝒫}𝒫=0\{\mathfrak{K},H_{\mathcal{P}}\}^{\mathcal{P}}=0 and {𝔔,H𝒫}𝒫=0\{\mathfrak{Q},H_{\mathcal{P}}\}^{\mathcal{P}}=0. Now, to their linear independence. It is already compromised from the fact that

𝔔=Jx2+Jy2+Jz2\mathfrak{Q}={J_{x}}^{2}+{J_{y}}^{2}+{J_{z}}^{2} (64)

as follows from Eqs. (59) and (62b). This is, again, somewhat expected from the spherical symmetry of the Schwarzschild spacetime, as discussed in Ramond and Isoyama (2024). Removing, say, JxJ_{x} from the list leaves us with 6 first integrals (H𝒫,E,Jy,Jz,𝔎,𝔔)(H_{\mathcal{P}},E,J_{y},J_{z},\mathfrak{K},\mathfrak{Q}), which are easily checked to be linearly independent. Lastly, their involution. This is compromised too, since, as is well-known, the angular-momentum components verify {Jy,Jz}=Jx\{J_{y},J_{z}\}=J_{x}, again this is associated to the spherical symmetry (see, e.g., Zelenka et al. (2020)). We can thus drop one of them, say JyJ_{y}. However, all five first integrals in the final set (H𝒫,E,Jz,𝔎,𝔔)(H_{\mathcal{P}},E,J_{z},\mathfrak{K},\mathfrak{Q}) are in involution, since the only non-trivial brackets to verify is {𝔎,𝔔}𝒫\{\mathfrak{K},\mathfrak{Q}\}^{\mathcal{P}}. However, from combining Eqs. (57) and (62), we easily show the following identities

{𝔎,𝔔}=0,{𝔎,C0}=πt𝔎f,{𝔎,C1}=fπr𝔎,{𝔔,C0}=0,{𝔔,C1}=0.\{\mathfrak{K},\mathfrak{Q}\}=0\,,\quad\{\mathfrak{K},C^{0}\}=\frac{\pi_{t}\mathfrak{K}}{\sqrt{f}}\,,\quad\{\mathfrak{K},C^{1}\}=-\sqrt{f}\pi_{r}\mathfrak{K}\,,\quad\{\mathfrak{Q},C^{0}\}=0\,,\quad\{\mathfrak{Q},C^{1}\}=0\,. (65)

Inserting all these identities into Eq. (43) readily leads to {𝔎,𝔔}𝒫=0\{\mathfrak{K},\mathfrak{Q}\}^{\mathcal{P}}=0. This confirms that (H𝒫,E,Jz,𝔎,𝔔)(H_{\mathcal{P}},E,J_{z},\mathfrak{K},\mathfrak{Q}) are in involution on 𝒫\mathcal{P}, thereby concluding the proof of integrability.

It is clear that the spherical symmetry of the Schwarzschild spacetime plays a leading role into the integrability of the system. However, this symmetry is also the reason why, out of the seven initial first integrals, two were discarded. In addition, the Killing vectors are not enough: while 𝔔\mathfrak{Q} actually comes from the spherical invariance (because it equals JJ\vec{J}\cdot\vec{J}), 𝔎\mathfrak{K} exists only thanks to the Killing-Yano tensor, and is independent of the Killing vectors. Therefore, it seems that the Killing-Yano structure, even on the Schwarzschild spacetime, is key to the integrability.

V.3 Integrability in the Kerr spacetime

Most remarkably, all five first integrals (H𝒫,E,Jz,𝔎,𝔔)(H_{\mathcal{P}},E,J_{z},\mathfrak{K},\mathfrak{Q}) that ensure integrability in Schwarzschild have an equivalent in the Kerr spacetime, thus suggesting that integrability can be naturally extended to the latter. We will show that this is precisely the case, following the exact same steps. We use the following shorthands in this section (and only in this section): Σ:=r2+a2cosθ\Sigma:=r^{2}+a^{2}\cos\theta, Δ:=r22Mr+a2\Delta:=r^{2}-2Mr+a^{2}, δ2=a2+r2\delta^{2}=a^{2}+r^{2}, c=cos2θc=\cos^{2}\theta and s=sinθs=\sin\theta.

V.3.1 Geometry and Hamiltonian

In terms of the Boyer-Lindquist (BL) coordinates xα=(t,r,θ,ϕ)x^{\alpha}=(t,r,\theta,\phi), the contravariant components of the Kerr metric read Gourgoulhon (2021)

gtt=2a2Mrs2+δ2ΣΣΔ,gθθ=1Σ,grr=ΔΣ,gϕϕ=Σ2MrΣΔs2,gϕt=gtϕ=2aMrΔΣ,g^{tt}=-\frac{2a^{2}Mrs^{2}+\delta^{2}\Sigma}{\Sigma\Delta}\,,\,g^{\theta\theta}=\frac{1}{\Sigma}\,,\,g^{rr}=\frac{\Delta}{\Sigma}\,,\,g^{\phi\phi}=\frac{\Sigma-2Mr}{\Sigma\Delta s^{2}}\,,\,g^{\phi t}=g^{t\phi}=-\frac{2aMr}{\Delta\Sigma}\,, (66)

where M0M\geq 0 is interpreted as black hole’s mass and the spin parameter a[0;M]a\in[0;M] is such the angular momentum of the black hole is aMaM. For a=0a=0, one recovers the Schwarzschild metric (53). To construct the canonical coordinates for the spin on 𝒩\mathcal{N}, we shall use the Carter tetrad Gourgoulhon (2021). In terms of the natural basis (α)a)(\partial_{\alpha})^{a}) associated to the BL coordinates, the Carter tetrad is defined by

(e0)a:=δ2(t)a+a(ϕ)aΔΣ,(e1)a:=ΔΣ(r)a,(e2)a:=(θ)aΣ,(e3)a:=as(t)a+(ϕ)a/sΣ.(e_{0})^{a}:=\frac{\delta^{2}\,(\partial_{t})^{a}+a\,(\partial_{\phi})^{a}}{\sqrt{\Delta\Sigma}}\,,\,(e_{1})^{a}:=\sqrt{\frac{\Delta}{\Sigma}}\,(\partial_{r})^{a}\,,\,(e_{2})^{a}:=\frac{(\partial_{\theta})^{a}}{\sqrt{\Sigma}}\,,\,(e_{3})^{a}:=\frac{as\,(\partial_{t})^{a}+(\partial_{\phi})^{a}/s}{\sqrt{\Sigma}}\,. (67)

The non-vanishing components of the connection 1-forms ωaBC\omega_{aBC} associated to the Carter tetrad are easily computed and read:

ωt01\displaystyle\omega_{t01} =MΣ2(a2c2r2),ωt23=2acMrΣ2,ωr03=arsΔΣ,ωr12=a2csΔΣ,\displaystyle=\frac{M}{\Sigma^{2}}(a^{2}c^{2}-r^{2})\,,\,\,\omega_{t23}=\frac{2acMr}{\Sigma^{2}}\,,\,\,\omega_{r03}=\frac{ars}{\sqrt{\Delta}\,\Sigma}\,,\,\,\omega_{r12}=-\frac{a^{2}cs}{\sqrt{\Delta}\,\Sigma}\,, (68a)
ωθ03\displaystyle\omega_{\theta 03} =acΔΣ,ωθ12=ΔrΣ,ωϕ01=as2Σ2(a2c2(rM)+r2(M+r)),\displaystyle=-\frac{ac\sqrt{\Delta}}{\Sigma}\,,\,\,\omega_{\theta 12}=-\frac{\sqrt{\Delta}r}{\Sigma}\,,\,\,\omega_{\phi 01}=\frac{as^{2}}{\Sigma^{2}}\left(a^{2}c^{2}(r-M)+r^{2}(M+r)\right)\,, (68b)
ωϕ02\displaystyle\omega_{\phi 02} =acΔsΣ,ωφ13=ΔrsΣ,ωϕ23=cΣ2(a2Δs2δ2).\displaystyle=\frac{ac\sqrt{\Delta}s}{\Sigma}\,,\,\omega_{\varphi 13}=-\frac{\sqrt{\Delta}rs}{\Sigma}\,,\,\omega_{\phi 23}=\frac{c}{\Sigma^{2}}(a^{2}\Delta s^{2}-\delta^{2})\,. (68c)

Again, setting a=0a=0 in the coefficients leads to their Schwarzschild expressions (55). The Hamiltonian on 𝒩\mathcal{N}, which we re-write from (12) using Eqs. (18), reads

H𝒩=12gαβπαπβ+12gαβπα(ωβIJεKIJSK+2ωβ0IDI),H_{\mathcal{N}}=\frac{1}{2}g^{\alpha\beta}\pi_{\alpha}\pi_{\beta}+\frac{1}{2}g^{\alpha\beta}\pi_{\alpha}\left(\omega_{\beta IJ}\varepsilon^{IJ}_{\phantom{IJ}K}S^{K}+2\omega_{\beta 0I}D^{I}\right)\,, (69)

in which all formulae (66) and (68) can be inserted to get an explicit expression. Similarly, all the above ingredients can be used to compute the two constraints (C0,C1)=(0,0)(C^{0},C^{1})=(0,0) given by Eq. (39), which read

ΣC0\displaystyle\sqrt{\Sigma}\,C^{0} :=ΔπrD1(aπtsinθ+πϕcscθ)D3πθD2,\displaystyle:=-\sqrt{\Delta}\pi_{r}D^{1}-(a\pi_{t}\sin\theta+\pi_{\phi}\csc\theta)D^{3}-\pi_{\theta}D^{2}\,, (70a)
ΔΣC1\displaystyle\sqrt{\Delta\Sigma}\,C^{1} :=(aπϕ+δ2πt)D1+Δ((aπtsinθ+πφcscθ)S2πθS3).\displaystyle:=(a\pi_{\phi}+\delta^{2}\pi_{t})\,D^{1}+\sqrt{\Delta}\left((a\pi_{t}\sin\theta+\pi_{\varphi}\csc\theta)\,S^{2}-\pi_{\theta}S^{3}\right)\,. (70b)

V.3.2 First integrals

The Kerr spacetime is stationary just like the Schwarzschild one, but spherical symmetry is lost and replaced by asymmetry. Consequently, only two Killing vectors of (58) remain, namely:

ξ(t)a=(t)aandξ(z)a=(ϕ)a.\xi_{(t)}^{a}=(\partial_{t})^{a}\quad\text{and}\quad\xi_{(z)}^{a}=(\partial_{\phi})^{a}\,. (71)

These Killing vectors lead to the existence of two Killing invariants from Eq. (45), expressed in terms of the coordinates on 𝒩\mathcal{N} as

E:=πtandLz:=πϕ,E:=-\pi_{t}\quad\text{and}\quad L_{z}:=\pi_{\phi}\,, (72)

and have the same physical interpretation as in the Schwarzschild case (total energy, and angular-momentum along the symmetry axis, measured at infinity). They are automatically promoted as first integrals thanks to the general result (48). In the Kerr spacetime, there alse exists a Killing-Yano tensor, which admits a simple expression in terms of the Carter tetrad (67), namely Floyd (1973); Compère and Druart (2022)

Yab:=2r(e2)[a(e3)b]2acosθ(e0)[a(e1)b],Y^{ab}:=2r(e_{2})^{[a}(e_{3})^{b]}-2a\cos\theta(e_{0})^{[a}(e_{1})^{b]}\,, (73)

Once again, this Killing-Yano tensor generates the Killing-Stäckel Kab:=YacYcbK^{ab}:=Y^{ac}Y^{b}_{\phantom{b}c}, whose expression is readily obtained as

Kab=r2((e2)a(e2)b+(e3)a(e3)b)+a2cos2θ((e0)a(e0)b(e1)a(e1)b).K^{ab}=r^{2}\,\bigl{(}(e_{2})^{a}(e_{2})^{b}+(e_{3})^{a}(e_{3})^{b}\bigr{)}+a^{2}\cos^{2}\!\theta\,\bigl{(}(e_{0})^{a}(e_{0})^{b}-(e_{1})^{a}(e_{1})^{b}\bigr{)}\,. (74)

Inserting Eqs. (73) and (74) within the definition of the two Rüdiger invariants (49) and (50) leads to their expression in terms of the canonical coordinates on 𝒩\mathcal{N}

𝔎\displaystyle\mathfrak{K} =rD1+acosθS1,\displaystyle=r\,D^{1}+a\cos\theta\,S^{1}\,, (75a)
𝔔\displaystyle\mathfrak{Q} =𝔔0+𝔔SS+𝔔DD,\displaystyle=\mathfrak{Q}_{0}+\vec{\mathfrak{Q}}_{S}\cdot\vec{S}+\vec{\mathfrak{Q}}_{D}\cdot\vec{D}\,, (75b)

where 𝔔0,𝔔S,𝔔S\mathfrak{Q}_{0},\vec{\mathfrak{Q}}_{S},\vec{\mathfrak{Q}}_{S} are spin-independent quantities given by

ΣΔ𝔔0\displaystyle\Sigma\,\Delta\,\mathfrak{Q}_{0} :=πϕ2(a4c2s2+Δr2)/s2+2a2c2Mπt2r3+Δ(πθ2r2a2c2Δπr2)\displaystyle:=\pi_{\phi}^{2}(a^{4}c^{2}s^{2}+\Delta r^{2})/s^{2}+2a^{2}c^{2}M\pi_{t}^{2}r^{3}+\Delta(\pi_{\theta}^{2}r^{2}-a^{2}c^{2}\Delta\,\pi_{r}^{2}) (76a)
+aπt(aπt+2πϕ)(a2c2δ2+Δr2),\displaystyle\,\,\,+a\pi_{t}(a\pi_{t}+2\pi_{\phi})(a^{2}c^{2}\delta^{2}+\Delta\,r^{2})\,, (76b)
Σ𝔔S1\displaystyle\Sigma\,\mathfrak{Q}_{S}^{1} :=2(aπϕ+δ2πt)ac2r2πϕc/s2,\displaystyle:=2(a\pi_{\phi}+\delta^{2}\pi_{t})\,ac-2r^{2}\pi_{\phi}c/s^{2}\,, (76c)
ΔΣ2𝔔S2\displaystyle\sqrt{\Delta}\,\Sigma^{2}\,\mathfrak{Q}_{S}^{2} :=4ar(ac2((2Mra2)πϕcscθ+aΔsπt)r2(sπtδ2+2aπϕcscθ)),\displaystyle:=4ar\left(ac^{2}((2Mr-a^{2})\pi_{\phi}\csc{\theta}+a\Delta s\pi_{t})-r^{2}(s\pi_{t}\delta^{2}+2a\pi_{\phi}\csc{\theta})\right)\,, (76d)
Σ2𝔔S3\displaystyle\Sigma^{2}\,\mathfrak{Q}_{S}^{3} :=2a2Δ(πθcrπrs)rc,\displaystyle:=2a^{2}\sqrt{\Delta}\left(\pi_{\theta}c-r\pi_{r}s\right)rc\,, (76e)
ΔΣ𝔔D1\displaystyle\Delta\,\Sigma\,\mathfrak{Q}_{D}^{1} :=2a2c2((a23r2)Mπt+aπϕ(Mr))2ar(Δ(πϕ+aπt)c2δ2aπt),\displaystyle:=2a^{2}c^{2}\left((a^{2}-3r^{2})M\pi_{t}+a\pi_{\phi}(M-r)\right)-2ar\left(\Delta(\pi_{\phi}+a\pi_{t})-c^{2}\delta^{2}a\pi_{t}\right)\,, (76f)
ΔΣ2𝔔D2\displaystyle\sqrt{\Delta}\,\Sigma^{2}\,\mathfrak{Q}_{D}^{2} :=2a2(a2c3Δsπt+a(c2(a22Mr)+r2)πϕcotθ+cδ2r2sπt),\displaystyle:=-2a^{2}\left(a^{2}c^{3}\Delta s\pi_{t}+a(c^{2}(a^{2}-2Mr)+r^{2})\pi_{\phi}\cot\theta+c\delta^{2}r^{2}s\pi_{t}\right)\,, (76g)
Σ2𝔔D3\displaystyle\Sigma^{2}\,\mathfrak{Q}_{D}^{3} :=2aΔ(a2c3πθ+r3πrs).\displaystyle:=2a\sqrt{\Delta}(a^{2}c^{3}\pi_{\theta}+r^{3}\pi_{r}s)\,. (76h)

Unlike in the Schwarzschild case, we have at our disposal five first integral “candidates”, namely (H𝒫,E,Lz,𝔎,𝔔)(H_{\mathcal{P}},E,L_{z},\mathfrak{K},\mathfrak{Q}). This is just enough to ensure integrability, as shown below.

V.3.3 Integrability

We do exactly the same checks as we did in Schwarzschild to ensure that all five invariants defined are (i) first integrals for H𝒫H_{\mathcal{P}} and (ii) linearly independent. The calculations become quickly complicated by hand, and can be found in the companion Mathematica Notebook MMA (2024). The last ingredient to check is the pairwise involution of all these first integrals. Again, this present no difficulty and only the bracket {𝔎,𝔔}𝒫\{\mathfrak{K},\mathfrak{Q}\}^{\mathcal{P}} is non-trivial. The latter is zero, ensuring that (H𝒫,E,Lz,𝔎,𝔔)(H_{\mathcal{P}},E,L_{z},\mathfrak{K},\mathfrak{Q}) satisfy all properties for the 10D Hamiltonian system (𝒫,{,}𝒫,H𝒫)(\mathcal{P},\{,\}^{\mathcal{P}},H_{\mathcal{P}}) to be integrable in Kerr, thus concluding the proof.

Before we end this section, we want to emphasize on an interesting feature that unexpectedly stumbled upon. Quite remarkably, we found a much stronger result than {𝔎,𝔔}𝒫=0\{\mathfrak{K},\mathfrak{Q}\}^{\mathcal{P}}=0, namely

{𝔔,H}=0,and{CA,𝔔}=αBACB,\{\mathfrak{Q},H\}=0\,,\quad\text{and}\quad\{C^{A},\mathfrak{Q}\}=\alpha^{A}_{\phantom{A}B}C^{B}\,, (77)

for some coefficients282828We do not display them here because their explicit form is irrelevant for the result, but they can be found in the Companion Notebook MMA (2024). αBA\alpha^{A}_{\phantom{A}B} that only depend on (πt,r,πr,θ,πθ,πϕ)(\pi_{t},r,\pi_{r},\theta,\pi_{\theta},\pi_{\phi}). The result (77) readily implies {𝔎,𝔔}𝒫=0\{\mathfrak{K},\mathfrak{Q}\}^{\mathcal{P}}=0 (owing to Eq. (43)), but also that 𝔔\mathfrak{Q}, as given in (50), is actually a first integral of the 12D system on 𝒩\mathcal{N}, just like a Killing vector invariant would be. This is a very special result: it does not hold for 𝔎\mathfrak{K}, and it is unclear if it holds for other Ricci-flat spacetimes.

Conclusions and prospects

Concluding summary

This article contains two main results. The first one is a rigorous and covariant Hamiltonian framework for the linear-in-spin dynamics in any given metric, under the TD SSC. The framework is summarized in Sec. IV.2. The final formulation on the 10D phase space 𝒫\mathcal{P} is the only one that is well-posed theoretically, in the sense that the SSC is automatically enforced along any trajectory, and practically, because the structure is symplectic and one can use the classical tool from Hamiltonian mechanics. The reduction to 𝒫\mathcal{P} went in two steps. First, we went from a 14D, degenerate formulation on the phase space (xα,pα,Sαβ)(x^{\alpha},p_{\alpha},S^{\alpha\beta})\in\mathcal{M}, to a 12D, symplectic (non-degenerate) formulation on (xα,πα,σ,πσ,ζ,πζ)𝒩(x^{\alpha},\pi_{\alpha},\sigma,\pi_{\sigma},\zeta,\pi_{\zeta})\in\mathcal{N} (the symplectic leaves of \mathcal{M}). In particular, these coordinates on 𝒩\mathcal{N} are canonical, and we provided a simple physical interpretation for them. Lastly, we enforced the Tulczyjew-Dixon spin supplementary condition, treating it a phase space constraint, thereby restricting to a 10D, symplectic (non-degenerate) formulation on 𝒫\mathcal{P}. All physical solutions to the algebraic-differental system (9)-(6) fill the 10D phase space 𝒫\mathcal{P}.

The general covariance of the framework, its 10-dimensional nature and the rigorous treatment of different kinds of “constants” of motion, are the key aspects of the formulation, compared to other existing Hamiltonian constructions available in the literature, in particular Barausse et al. (2009) and Witzany et al. (2019). These features are made possible by the use of the Tulczyjew-Dixon spin supplementary condition (SSC). As we argued in Sec. I.3, this SSC plays a special role in our framework, and it is possible that it is the only SSC for which such rigorous, covariant and 10D treatment is possible in the first place.

The second main result is the application of our formalism to the case of the Schwarzschild and Kerr spacetimes. By exhibiting enough independent integrals of motion in Sec. V, we have shown that the system is integrable, at linear order in spin. This result disproves the conjecture that exists chaotic motion for the linear-in-spin MPTD + TD SSC system. The five functionally independent integrals of motion admit a very simple interpretation, summarized in Table. 1. They were all known already in 1981, following Rüdiger’s work on the MPTD + TD SSC invariants. These invariants simply lacked a well-defined Hamiltonian framework in which to be interpreted as first integrals. Our work clearly identifies the role of the Killing-Yano tensor (or hidden symmetry) of both the Kerr and Schwarzschild spacetime as the source of this integrability at linear-in-spin order. In particular, Kerr is once again very special, since all the first integrals E,Lz,𝔎,𝔔E,L_{z},\mathfrak{K},\mathfrak{Q} can be related to existence of the sole Killing-Yano tensor of the Kerr spacetime, as the latter uniquely seeds the two Killing vectors Hughston and Sommers (1973), a property unique to Kerr.

Symmetry First integral Killing structure
free particle μ:\mu: mass  Metric gabg_{ab}
stationarity E:E: total energy  Killing vector (t)a(\partial_{t})^{a}
axisymmetry Lz:L_{z}: zz-comp. of ang. mom.  Killing vector (ϕ)a(\partial_{\phi})^{a}
hidden symmetry (𝔎,𝔔):(\mathfrak{K},\mathfrak{Q}): Rüdiger’s spin invariants  Killing-Yano tensor YabY_{ab}
Table 1: Symmetries at play behind integrability of spinning particles around Kerr black holes.

Future applications

This article is the first in a series aiming at exploiting the powerful machinery of Hamiltonian mechanics to study the MPTD equations. Now that the general framework is established and that integrability for black hole backgrounds is proven, our next step is to perform practical implementations and solve specific problems. We now make a (non-exhaustive) list of potential prospects, some of which are already ongoing.

Canonical formulation and solutions by quadrature

The final, 10-dimensional formulation on the phase space 𝒫\mathcal{P} (summarized in Sec. IV.2) is not endowed with canonical coordinates. We do not think this is possible to do in the general case, i.e., without first making a choice for the background metric. However, once such choice is made, nothing prevents us from constructing such coordinates from the non-canonical ones. A method for doing just this is, in fact, is already present in this paper, in App. B, where we construct the map SAB6(σ,πσ,ζ,πζ)S^{AB}\in\mathbb{R}^{6}\mapsto(\sigma,\pi_{\sigma},\zeta,\pi_{\zeta}) by explicitly solving the SABS^{AB}-brackets by viewing them as canonical brackets between (σ,πσ,ζ,πζ)(\sigma,\pi_{\sigma},\zeta,\pi_{\zeta})-dependent functions. This is an example of non-canonical to canonical coordinate transformation.

In a follow-up article Ramond and Isoyama (2024), we have managed to do this reduction in a Schwarzschild background, i.e., we have explicitly constructed 10 canonical coordinates on 𝒫\mathcal{P}, thus giving a canonical formulation of the linear-in-spin dynamics. These coordinates turn the 5D Hamiltonian into an effective 1D one for the radial motion, with the 8 other phase space coordinates taking the form of (cyclic angle,first integral) pairs; with the first integrals being precisely (E,Jz,𝔔,𝔎)(E,J_{z},\mathfrak{Q},\mathfrak{K}). This formulation permits an analytic solution in terms of elliptic functions. Generalization to the Kerr case is ongoing.

Action-angle variables, frequencies and resonances

Since the linear-in-spin dynamics in the Kerr spacetime form an integrable system, the Liouville-Arnold theorem Liouville (1855); Arnold et al. (2006) ensures that there exists a particular set of canonical variables (𝒥i,ϑi){i=1,,5}(\mathcal{J}_{i},\vartheta_{i})_{\{i=1,\ldots,5\}}, called action-angle variables.292929In practice, we expect that there are different patches of action-angle coordinates for each type of bound, plunging and scattering orbits, all separated by separatrices. The actions 𝒥i\mathcal{J}_{i} are in a 1-to-1 correspondence with the five integrals of motion (μ,E,Lz,𝔎,𝔔)(\mu,E,L_{z},\mathfrak{K},\mathfrak{Q}). To our knowledge, these have never been derived for generic Schwarzschild or Kerr orbits in linear-in-spin motion. These variables are probably the best (if not the only) tool to study resonances in the system linear-in-system, with the well-defined, gauge-invariant notion of Hamiltonian frequencies Ωi:=H𝒫/𝒥i\Omega_{i}:=\partial H_{\mathcal{P}}/\partial\mathcal{J}_{i}. These frequencies are presented for the Schwarzschild case in the follow-up article Ramond and Isoyama (2024). They will allow us to study the impact of linear-in-spin terms on orbital resonances, as well as spin-orbit resonances. In essence, this will extend the geodesic analyses of Brink et al. (2015) for conservative resonances, and Flanagan and Hinderer (2012) for transient resonances (when dissipation is turned on).

First law of mechanics and flux-balance laws

By adapting the methods of Le Tiec (2014), it should be possible to derive a first law of mechanics for a two-body system made of a dipolar particle orbiting a Kerr black hole, at linear order in spin. The first law, at this order, has been derived in the past, in different contexts (e.g.g, using the ADM Hamiltonian formalism Blanchet et al. (2013), or for circular orbits using the helical symmetry in Ramond and Le Tiec (2022)). This new derivation would give some insight as to the status of these laws, extend their domain of validity, but most of all, it would allow for a very convenient way to extract the conservative part of the second-order gravitational self-force, as linear-in-spin effects enter precisely at this order Mathews et al. (2022).

In the context of adiabatic evolution for asymmetric binaries Pound and Wardell (2021), it is possible to derive so-called “flux-balance laws” for the constants of motion that are slowly evolving under the emission of gravitational waves. These laws encode crucial information about the dissipative sector of the gravitational self-force. When these constants of motion are associated to spacetime isometries (like EE or LzL_{z} in Kerr), the derivation of such laws is well-understood for geodesics Isoyama et al. (2019). When linear-in-spin effects are included, the study Akcay et al. (2020) extended these results for energy and angular-momentum. Recently, the authors of Grant and Moxon (2023b) proposed a method based on symplectic currents to construct these laws for constants of motion that are not associated to spacetime isometries, for example the Carter constant in the non-spinning case. This method could be generalized for the Rüdiger constants 𝔎,𝔔\mathfrak{K},\mathfrak{Q} that fit in this category.

Adding the self-force to the linear-in-spin formalism

It has been shown that for a non-spinning particle orbiting a Kerr black hole, the conservative part of the self-force does not break the integrability of the system Vines and Flanagan (2015b); Fujita et al. (2017). It is still unclear if this holds true at linear order in spin, or whether couplings between the self-force effects and spin breaks integrability. Nevertheless, it is now well-established that the conservative self-force can be absorbed into a re-definition of the background metric to an effective one Shah et al. (2011); Barack (2014); Pound and Wardell (2021); Harte (2015), such that there is hope to write a (possibly) covariant Hamiltonian framework that does just this (see also Blanco and Flanagan (2023a, b) on this matter). Ref Isoyama et al. (2019) was able to do this for a Kerr background and non-spinning case, and we hope to extend this analysis to the linear-in-spin case.

Regarding the dissipative sector of the self-force, one must leave the realm of classical conservative Hamiltonian systems. Indeed, the prime effect of dissipation is to make the constants of motion evolve adiabatically. However, there is hope that Galley’s Hamiltonian formulation of non-conservative mechanics Galley (2013) may apply to this context. It was already successfully applied in the context of effective field theory Galley and Leibovich (2012), and we plan to explore its application in ours in the future.

Hairy black holes, electromagnetic fields

The extension of our integrability result to other spacetimes (in particular other black hole metrics) should be straightforward, even though integrability will most probably only follow if there are enough Killing symmetries. It would be interesting to investigate integrability for axisymmetric, stationary spacetimes admitting Killing-Yano tensors, relying on classification to explore all possibilities. In addition, if one accounts for electromagnetic fields, for example in the case of charged black holes, then an additional linear-in-spin coupling appears in the Hamiltonian (cf. the generalized momentum-velocity relation in Gralla et al. (2010) or the added terms in the Poisson structure d’Ambrosi et al. (2016)). One cannot say if this additional term will break integrability or not, because of what becomes of the Killing tensors under these additional fields. We leave all these extensions for future investigations.

Quadratic-in-spin dynamics

In this article, all calculations are done consistently at linear order in the spin SabS^{ab}, cf. Sec. I.4. This linearization is not so much a simplifying constraint than it is a necessary assumption for the dipolar MPTD equations to be physically self-consistent. A natural extension to our work is, therefore, to include quadratic-in-spin effects in our Hamiltonian. While this is easily done at dipolar order by adapting the construction of Witzany et al. (2019), self-consistency requires that we go to quadrupolar order, where a spin-induced quadrupole arises on the left-hand-side of Eq. (2), see Harte (2020); Compère et al. (2023). We have managed to extend our covariant Hamiltonian formulation to this quadratic-in-spin order, which will be the object a forthcoming publication Ramond et al. (2024). To our knowledge, there exists no quadratic-in-spin and covariant Hamiltonian formulation of these quadratic-in-spin equations, the closest to this being the self-consistent (but non-covariant) quadratic-in-spin Hamiltonian of Vines et al. (2016). One interesting aspect of having a quadratic-in-spin Hamiltonian is that it is a perturbation of the linear-in-spin one, presented here. Therefore, it should be possible to analytically study the emergence of chaos there, using the Melnikov-Poincaré theory Wiggins (2013); Guckenheimer and Holmes (2013), and the impact it can have on the motion, especially around resonances, as described by Kolmogorov-Arnold-Moser theory, and around fixed points, as described by Poincaré-Birkhoff theory Arnold et al. (2006). We aim to explore this effect in the future.

We note that the non-integrability of quadratic-in-spin dynamics, in both Schwarzschild and Kerr, have currently no analytical proofs but only numerical evidence. If the secondary is a neutron star, the system is probably non-integrable, because of the arbitrarily complex matter fields at play. However, when the secondary is a black hole, both Rüdiger constants 𝔎,𝔔\mathfrak{K},\mathfrak{Q} that we have used in this work admit a generalization to quadratic-in-spin order Compère et al. (2023). Whether these invariants are real first integrals of the corresponding (quadratic-in-spin) Hamiltonian system and whether they are functionally independent to ensure integrability will be the subject of subsequent work.

Acknowledgement

I am grateful to Soichiro Isoyama for countless helpful discussions during this project. I thank Karim Van Aelst, Adrien Druart, Jacques Fejoz, Josh Mathews, Sashwat Tanay and Alexandre Le Tiec for useful comments and feedback. I am also indebted to David Brown for his guidance through the Dirac-Bergmann algorithm at earlier stages of this work.

References

  • B. P. Abbott et al. (2016) (LIGO Scientific Collaboration and Virgo Collaboration) B. P. Abbott et al. (LIGO Scientific Collaboration and Virgo Collaboration), Phys. Rev. Lett. 116, 061102 (2016), arXiv:1602.03837 [gr-qc] .
  • B. P. Abbott et al. (2019) (LIGO Scientific Collaboration and Virgo Collaboration) B. P. Abbott et al. (LIGO Scientific Collaboration and Virgo Collaboration), Phys. Rev. X 9, 031040 (2019), arXiv:1811.12907 [astro-ph.HE] .
  • B. P. Abbott et al. (2020) (LIGO Scientific Collaboration and Virgo Collaboration) B. P. Abbott et al. (LIGO Scientific Collaboration and Virgo Collaboration), arXiv preprint arXiv:2010.14527  (2020).
  • Abbott et al. (2021) R. Abbott, T. Abbott, F. Acernese, K. Ackley, C. Adams, N. Adhikari, R. Adhikari, V. Adya, C. Affeldt, D. Agarwal, et al., arXiv preprint arXiv:2111.03606  (2021).
  • Akiyama et al. (2019) K. Akiyama, A. Alberdi, W. Alef, K. Asada, R. Azulay, A.-K. Baczko, D. Ball, M. Baloković, J. Barrett, D. Bintley, et al., The Astrophysical Journal Letters 875, L4 (2019).
  • Akiyama et al. (2022) K. Akiyama, A. Alberdi, W. Alef, J. C. Algaba, R. Anantua, K. Asada, R. Azulay, U. Bach, A.-K. Baczko, D. Ball, et al., The Astrophysical Journal Letters 930, L14 (2022).
  • Cardoso et al. (2016) V. Cardoso, S. Hopper, C. F. Macedo, C. Palenzuela,  and P. Pani, Physical review D 94, 084031 (2016).
  • Toubiana et al. (2021) A. Toubiana, S. Babak, E. Barausse,  and L. Lehner, Physical Review D 103, 064042 (2021).
  • Unnikrishnan (2013) C. Unnikrishnan, International Journal of Modern Physics D 22, 1341010 (2013).
  • Punturo et al. (2010) M. Punturo, M. Abernathy, F. Acernese, B. Allen, N. Andersson, K. Arun, F. Barone, B. Barr, M. Barsuglia, M. Beker, et al., Classical and Quantum Gravity 27, 194002 (2010).
  • Amaro-Seoane et al. (2017) P. Amaro-Seoane, H. Audley, S. Babak, J. Baker, E. Barausse, P. Bender, E. Berti, P. Binetruy, M. Born, D. Bortoluzzi, et al., arXiv preprint arXiv:1702.00786  (2017).
  • Dixon (1974) W. G. Dixon, Phil. Trans. R. Soc. Lond. A 277, 59 (1974).
  • Harte (2015) A. I. Harte, Fund. Theor. Phys. 179, 327 (2015), arXiv:1405.5077 [gr-qc] .
  • Ramond (2021) P. Ramond, The First Law of Mechanics in General Relativity & Isochrone Orbits in Newtonian GravityPhD Thesis, Paris Cité Univ. (2021) (2021).
  • Harte (2012) A. I. Harte, Class. Quant. Grav. 29, 055012 (2012), arXiv:1103.0543 [gr-qc] .
  • Harte (2020) A. I. Harte, Physical Review D 102, 124075 (2020)arXiv:2011.00110 [gr-qc] .
  • Harte (2008) A. I. Harte, Classical and Quantum Gravity 25, 205008 (2008), arXiv:0805.4259 [gr-qc] .
  • Harte (2007) A. I. Harte, Classical and Quantum Gravity 24, 5161 (2007), arXiv:0706.2909 [gr-qc] .
  • Obukhov and Puetzfeld (2011) Y. N. Obukhov and D. Puetzfeld, Physical Review D 83, 044024 (2011), arXiv:1010.1451 [gr-qc] .
  • Zalaquett et al. (2014) N. Zalaquett, S. A. Hojman,  and F. A. Asenjo, Classical and Quantum Gravity 31, 085011 (2014).
  • Pound (2015) A. Pound, Fund. Theor. Phys. 179, 399 (2015), arXiv:1506.06245 [gr-qc] .
  • Miller and Pound (2021) J. Miller and A. Pound, Physical Review D 103, 064048 (2021), arXiv:2006.11263 [gr-qc] .
  • Hanson and Regge (1974) A. J. Hanson and T. Regge, Ann. Phys. 87, 498 (1974).
  • Porto (2006) R. A. Porto, Phys. Rev. D 73, 104031 (2006), arXiv:gr-qc/0511061 .
  • Barausse et al. (2009) E. Barausse, E. Racine,  and A. Buonanno, Phys. Rev. D 80, 104025 (2009), Erratum: Phys. Rev. D 85, 069904(E) (2012), arXiv:0907.4745 [gr-qc] .
  • Vines et al. (2016) J. Vines, D. Kunst, J. Steinhoff,  and T. Hinderer, Physical Review D 93, 103008 (2016), arXiv:1601.07529 [gr-qc] .
  • Marsat (2015) S. Marsat, Class. Quant. Grav. 32, 085008 (2015), arXiv:1411.4118 [gr-qc] .
  • Steinhoff (2011) J. Steinhoff, Annalen der Physik 523, 296 (2011), arXiv:1106.4203 [gr-qc] .
  • Souriau (1970) J.-M. Souriau, CR Acad. Sci. Paris 271, 751 (1970).
  • Damour (2024) T. Damour,   (2024), arXiv:2401.10013 [gr-qc] .
  • Künzle (1972) H. Künzle, Journal of Mathematical Physics 13, 739 (1972).
  • d’Ambrosi et al. (2015) G. d’Ambrosi, S. S. Kumar,  and J. van Holten, Physics Letters B 743, 478 (2015), arXiv:1501.04879 [gr-qc] .
  • Witzany et al. (2019) V. Witzany, J. Steinhoff,  and G. Lukes-Gerakopoulos, Classical and Quantum Gravity 36, 075003 (2019), arXiv:1808.06582 [gr-qc] .
  • d’Ambrosi et al. (2016) G. d’Ambrosi, S. S. Kumar, J. Van de Vis,  and J. Van Holten, Physical Review D 93, 044051 (2016).
  • Witzany (2019) V. Witzany, Physical Review D 100, 104030 (2019), arXiv:1903.03651 [gr-qc] .
  • Brown (2022) J. D. Brown, Universe 8, 171 (2022), arXiv:2201.06558 .
  • Arnold et al. (2006) V. Arnold, V. Kozlov,  and A. Neishtadt, Mathematical aspects of classical and celestial mechanics, Vol. 3 (Springer, 2006).
  • Deriglazov (2022) A. A. Deriglazov, Universe 8, 536 (2022), arXiv:2210.09131 .
  • Arnold (1966) V. Arnold, in Annales de l’institut Fourier, Vol. 16 (1966) pp. 319–361.
  • Carter (1968) B. Carter, Physical Review 174, 1559 (1968).
  • Kerr (1963) R. P. Kerr, Phys. Rev. Lett. 11, 237 (1963).
  • Walker and Penrose (1970) M. Walker and R. Penrose, Communications in Mathematical Physics 18, 265 (1970).
  • Frolov et al. (2017) V. P. Frolov, P. Krtouš,  and D. Kubizňák, Living reviews in relativity 20, 1 (2017).
  • Bombelli and Calzetta (1992) L. Bombelli and E. Calzetta, Classical and Quantum Gravity 9, 2573 (1992).
  • Verhaaren and Hirschmann (2010) C. Verhaaren and E. W. Hirschmann, Physical Review D 81, 124034 (2010).
  • Zelenka et al. (2020) O. Zelenka, G. Lukes-Gerakopoulos, V. Witzany,  and O. Kopáček, Physical Review D 101, 024037 (2020), arXiv:1911.00414 [gr-qc] .
  • Suzuki and Maeda (1997) S. Suzuki and K.-i. Maeda, Physical Review D 55, 4848 (1997), arXiv:9604020 [gr-qc] .
  • Semerák (1999) O. Semerák, Mon. Not. R. Astron. Soc. 308, 863 (1999).
  • Kyrian and Semerák (2007) K. Kyrian and O. Semerák, Mon. Not. R. Astron. Soc. 382, 1922 (2007).
  • Han (2008) W. Han, General Relativity and Gravitation 40, 1831 (2008).
  • Compère and Druart (2022) G. Compère and A. Druart, SciPost Physics 12, 012 (2022), arXiv:2105.12454 [gr-qc] .
  • Witzany et al. (2015) V. Witzany, O. Semerák,  and P. Suková, Monthly Notices of the Royal Astronomical Society 451, 1770 (2015), arXiv:1503.09077 [gr-qc] .
  • Polcar et al. (2022) L. Polcar, G. Lukes-Gerakopoulos,  and V. Witzany, arXiv preprint arXiv:2205.08516  (2022).
  • Mosna et al. (2022) R. A. Mosna, F. F. Rodrigues,  and R. S. Vieira, Physical Review D 106, 024016 (2022), arXiv:2207.04341 [gr-qc] .
  • Barack and Pound (2018) L. Barack and A. Pound, Rep. Prog. Phys. 82, 016904 (2018), arXiv:1805.10385 [gr-qc] .
  • Pound and Wardell (2021) A. Pound and B. Wardell, “Black hole perturbation theory and gravitational self-force,”  (2021), invited chapter for ”Handbook of Gravitational Wave Astronomy” (Eds. C. Bambi, S. Katsanevas, and K. Kokkotas; Springer, Singapore, 2021), arXiv:2101.04592 [gr-qc] .
  • Afshordi et al. (2023) N. Afshordi, S. Akçay, P. A. Seoane, A. Antonelli, J. C. Aurrekoetxea, L. Barack, E. Barausse, R. Benkel, L. Bernard, S. Bernuzzi, et al., arXiv preprint arXiv:2311.01300  (2023).
  • Boyle et al. (2008) M. Boyle et al., Phys. Rev. D 78, 104020 (2008), arXiv:0804.4184 [gr-qc] .
  • Lukes-Gerakopoulos and Witzany (2021) G. Lukes-Gerakopoulos and V. Witzany, in Handbook of Gravitational Wave Astronomy (Springer, 2021) pp. 1–44, arXiv:2103.06724v1 [gr-qc] .
  • Vines and Flanagan (2015a) J. Vines and É. É. Flanagan, Phys. Rev. D 92, 064039 (2015a), arXiv:1503.04727 [gr-qc] .
  • Fujita et al. (2017) R. Fujita, S. Isoyama,  and others., Class. Quant. Grav. 34, 134001 (2017), arXiv:1612.02504 .
  • Blanco and Flanagan (2023a) F. M. Blanco and É. É. Flanagan, Physical Review Letters 130, 051201 (2023a).
  • Galley (2013) C. R. Galley, Phys. Rev. Lett. 110, 174301 (2013), arXiv:1210.2745 .
  • Haier et al. (2006) E. Haier, C. Lubich,  and G. Wanner, Geometric Numerical integration: structure-preserving algorithms for ordinary differential equations (Springer, 2006).
  • Wang et al. (2021) Y. Wang, W. Sun, F. Liu,  and X. Wu, The Astrophysical Journal 907, 66 (2021).
  • Hinderer and Flanagan (2008) T. Hinderer and E. E. Flanagan, Physical Review D 78, 064028 (2008), arXiv:0805.3337 [gr-qc] .
  • Isoyama et al. (2014) S. Isoyama et al., Phys. Rev. Lett. 113, 161101 (2014), arXiv:1404.6133 [gr-qc] .
  • Le Tiec (2015) A. Le Tiec, Phys. Rev. D 92, 084021 (2015), arXiv:1506.05648 [gr-qc] .
  • Isoyama et al. (2018) S. Isoyama, H. Nakano,  and T. Nakamura, Progress of Theoretical and Experimental Physics 2018, 073E01 (2018).
  • Grant and Moxon (2023a) A. M. Grant and J. Moxon, Physical Review D 108, 104029 (2023a).
  • Piovano et al. (2020) G. A. Piovano, A. Maselli,  and P. Pani, Physical Review D 102, 024041 (2020).
  • Mathews et al. (2022) J. Mathews, A. Pound,  and B. Wardell, Physical Review D 105, 084031 (2022), arXiv:2112.13069 .
  • Rahman and Bhattacharyya (2023) M. Rahman and A. Bhattacharyya, Physical Review D 107, 024006 (2023).
  • Barausse et al. (2020) E. Barausse et al., Gen. Rel. Grav. 52, 81 (2020), arXiv:2001.09793 [gr-qc] .
  • Rüdiger (1981) R. Rüdiger, Proceedings of the Royal Society of London. A. Mathematical and Physical Sciences 375, 185 (1981).
  • Rüdiger (1983) R. Rüdiger, Proceedings of the Royal Society of London. A. Mathematical and Physical Sciences 385, 229 (1983).
  • Ramond and Isoyama (2024) P. Ramond and S. Isoyama,   (2024).
  • Tulczyjew (1959) W. Tulczyjew, Acta Phys. Polon. 18, 393 (1959).
  • Steinhoff and Puetzfeld (2010) J. Steinhoff and D. Puetzfeld, Phys. Rev. D 81, 044019 (2010), arXiv:0909.3756 [gr-qc] .
  • Steinhoff (2015a) J. Steinhoff, Fund. Theor. Phys. 179, 615 (2015a), arXiv:1412.3251 [gr-qc] .
  • Gralla and Wald (2008) S. E. Gralla and R. M. Wald, Class. Quant. Grav. 25, 205009 (2008), arXiv:0806.3293 [gr-qc] .
  • Pound (2012) A. Pound, Phys. Rev. Lett. 109, 051101 (2012), arXiv:1201.5089 [gr-qc] .
  • Dixon (2015) W. G. Dixon, in Equations of Motion in Relativistic Gravity (Springer, 2015) pp. 1–66.
  • Ramond and Le Tiec (2021) P. Ramond and A. Le Tiec, Class. Quant. Grav 38 (2021), arXiv:2005.00602 [gr-qc] .
  • Gourgoulhon (2013) E. Gourgoulhon, Special relativity in general frames, Graduate Texts in Physics (Springer, New York, 2013).
  • Lukes-Gerakopoulos et al. (2014) G. Lukes-Gerakopoulos, J. Seyrich,  and D. Kunst, Physical Review D 90, 104019 (2014), arXiv:1409.4314 [gr-qc] .
  • Costa and Natário (2015) L. F. O. Costa and J. Natário, Fund. Theor. Phys. 179, 215 (2015), arXiv:1410.6443 [gr-qc] .
  • Timogiannis et al. (2021) I. Timogiannis, G. Lukes-Gerakopoulos,  and T. A. Apostolatos, Physical Review D 104, 024042 (2021), arXiv:2104.11183 [gr-qc] .
  • Steinhoff (2015b) J. Steinhoff, arXiv preprint arXiv:1501.04951  (2015b).
  • Lukes-Gerakopoulos (2017) G. Lukes-Gerakopoulos, Physical Review D 96, 104023 (2017), arXiv:1709.08942 [gr-qc] .
  • Dixon (1964) W. G. Dixon, Il Nuovo Cimento 34, 317 (1964).
  • Ehlers and Rudolph (1977) J. Ehlers and E. Rudolph, Gen. Rel. Grav. 8, 197 (1977).
  • Beiglböck (1967) W. Beiglböck, Commun. Math. Phys. 5, 106 (1967).
  • Madore (1969) J. Madore, in Annales de l’IHP Physique théorique, Vol. 11 (1969) pp. 221–237.
  • Schattner (1979a) R. Schattner, General Relativity and Gravitation 10, 377 (1979a).
  • Schattner (1979b) R. Schattner, General Relativity and Gravitation 10, 395 (1979b).
  • Costa et al. (2012) L. F. Costa, C. Herdeiro, J. Natário,  and M. Zilhão, Phys. Rev. D 85, 024001 (2012), arXiv:1109.1019 [gr-qc] .
  • Gralla et al. (2010) S. E. Gralla, A. I. Harte,  and R. M. Wald, Physical Review D 81, 104012 (2010), arXiv:1004.0679 [gr-qc] .
  • Compère et al. (2023) G. Compère, A. Druart,  and J. Vines, SciPost Phys. 15, 226 (2023)arXiv:2302.14549 [gr-qc] .
  • Dietz and Rüdiger (1981) W. Dietz and R. Rüdiger, Proceedings of the Royal Society of London. A. Mathematical and Physical Sciences 375, 361 (1981).
  • Dietz and Rudiger (1982) W. Dietz and R. Rudiger, Proceedings of the Royal Society of London. A. Mathematical and Physical Sciences 381, 315 (1982).
  • Harte and Dwyer (2023) A. I. Harte and D. Dwyer, Physical Review D 108, 124005 (2023).
  • Martelli (2016) B. Martelli, An introduction to geometric topology (CreateSpace Independent Publishing Platform, 2016) arXiv:1610.02592 .
  • Schmidt (2002) W. Schmidt, Classical and Quantum Gravity 19, 2743 (2002), arXiv:gr-qc/0202090 [gr-qc] .
  • Wald (1984) R. M. Wald, General relativity (University of Chicago Press, Chicago, 1984).
  • Baum (2009) H. Baum, “Gauge theory. an introduction into differential geometry on fibre bundles.(eichfeldtheorie. eine einführung in die differentialgeometrie auf faserbündeln.),”  (2009).
  • Frankel (2011) T. Frankel, The geometry of physics: an introduction (Cambridge university press, 2011).
  • Heard (2008) W. B. Heard, Rigid body mechanics: mathematics, physics and applications (John Wiley & Sons, 2008).
  • Liouville (1855) J. Liouville, Journal de Mathématiques pures et appliquées 20, 137 (1855).
  • Dirac (1950) P. A. M. Dirac, Canadian journal of mathematics 2, 129 (1950).
  • Carter (2009) B. Carter, General Relativity and Gravitation 41, 2873 (2009).
  • Kunst et al. (2016) D. Kunst, T. Ledvinka, G. Lukes-Gerakopoulos,  and J. Seyrich, Physical Review D 93, 044004 (2016), arXiv:1506.01473 [gr-qc] .
  • Gourgoulhon (2021) É. Gourgoulhon, “Geometry and physics of black holes,” (2021).
  • Müller and Grave (2009) T. Müller and F. Grave, arXiv preprint arXiv:0904.4184  (2009).
  • Tanaka et al. (1996) T. Tanaka, Y. Mino, M. Sasaki,  and M. Shibata, Phys. Rev. D 54, 3762 (1996)arXiv:gr-qc/9602038 .
  • Açık et al. (2010) Ö. Açık, Ü. Ertem, M. Önder,  and A. Vercin, Journal of mathematical physics 51, 022502 (2010), arXiv:0803.3328 .
  • Floyd (1973) R. Floyd, PhD Thesis, University of London  (1973).
  • MMA (2024) “Mathematica companion notebook,”  (2024).
  • Hughston and Sommers (1973) L. P. Hughston and P. Sommers, Communications in Mathematical Physics 33, 129 (1973).
  • Brink et al. (2015) J. Brink, M. Geyer,  and T. Hinderer, Physical review letters 114, 081102 (2015).
  • Flanagan and Hinderer (2012) É. É. Flanagan and T. Hinderer, Phys. Rev. Lett. 109, 071102 (2012), arXiv:1009.4923 [gr-qc] .
  • Le Tiec (2014) A. Le Tiec, Classical and Quantum Gravity 31, 097001 (2014), arXiv:1311.3836 [gr-qc] .
  • Blanchet et al. (2013) L. Blanchet, A. Buonanno,  and A. Le Tiec, Phys. Rev. D 87, 024030 (2013), arXiv:1211.1060 [gr-qc] .
  • Ramond and Le Tiec (2022) P. Ramond and A. Le Tiec, Physical Review D 106, 044057 (2022), arXiv:2202.09345 [gr-qc] .
  • Isoyama et al. (2019) S. Isoyama, R. Fujita, H. Nakano, N. Sago,  and T. Tanaka, Progress of Theoretical and Experimental Physics 2019, 013E01 (2019).
  • Akcay et al. (2020) S. Akcay, S. R. Dolan, C. Kavanagh, J. Moxon, N. Warburton,  and B. Wardell, Physical Review D 102, 064013 (2020), arXiv:1912.09461 [gr-qc] .
  • Grant and Moxon (2023b) A. M. Grant and J. Moxon, Physical Review D 108, 104029 (2023b).
  • Vines and Flanagan (2015b) J. Vines and É. É. Flanagan, Phys. Rev. D 92, 064039 (2015b), arXiv:1503.04727 .
  • Shah et al. (2011) A. G. Shah, T. S. Keidl, J. L. Friedman, D.-H. Kim,  and L. R. Price, Phys. Rev. D 83, 064018 (2011), arXiv:1009.4876 [gr-qc] .
  • Barack (2014) L. Barack, Fund. Theor. Phys. 177, 147 (2014).
  • Blanco and Flanagan (2023b) F. M. Blanco and É. É. Flanagan, Physical Review D 107, 124017 (2023b).
  • Galley and Leibovich (2012) C. R. Galley and A. K. Leibovich, Phys. Rev. D 86, 044029 (2012).
  • Ramond et al. (2024) P. Ramond, A. Druart,  and S. Isoyama,   (2024).
  • Wiggins (2013) S. Wiggins, Global bifurcations and chaos: analytical methods, Vol. 73 (Springer Science & Business Media, 2013).
  • Guckenheimer and Holmes (2013) J. Guckenheimer and P. Holmes, Nonlinear oscillations, dynamical systems, and bifurcations of vector fields, Vol. 42 (Springer Science & Business Media, 2013).
  • Vaisman (2012) I. Vaisman, Lectures on the geometry of Poisson manifolds, Vol. 118 (Birkhäuser Basel, 2012).

Appendix A Conventions and notations

Throughout the paper, we use geometric units in which G=c=1G=c=1, and our conventions for the Riemann curvature tensor and other geometric tensors are as in Wald (1984), in particular abωc=Rabcdωd\nabla_{a}\nabla_{b}\omega_{c}=R_{abc}^{\phantom{abc}d}\omega_{d} for any 1-form ωa\omega_{a}. We use a,b,c,d,a,b,c,d,\ldots to denote abstract indices. For the components, we use two kinds of indices varying in {0,1,2,3}\{0,1,2,3\} : α,β,γ,δ,\alpha,\beta,\gamma,\delta,\ldots are used for components in the natural basis (α)a(\partial_{\alpha})^{a} associated to a spacetime coordinate system xαx^{\alpha}, and A,B,C,D,A,B,C,D,\ldots are used for components in an orthonormal tetrad basis (eA)a(e_{A})^{a}. Calligraphic indices 𝒜,,𝒞,\mathcal{A},\mathcal{B},\mathcal{C},\ldots are used to denote a fixed index (never summed over). An arrow is sometimes used to denote Euclidean 3-vectors, using interchangeably v=(v1,v2,v3)=vI\vec{v}=(v^{1},v^{2},v^{3})=v^{I}, as well as usual notations \cdot for the scalar product and ×\times for the cross product. In the phase space, pairs of canonical coordinates are always of the form (q,πq)(q,\pi_{q}), with πq\pi_{q} being the conjugated momentum of some degree of freedom qq. Poisson brackets are denoted with the brackets {,}\{\cdot,\cdot\}. A table for the frequently used symbols, description and references is given in Table. 2.

Table 2: List of frequently used symbols.
Notation Description Relevant reference
Spacetime
\mathscr{E} Lorentzian manifold
xαx^{\alpha} spacetime coordinates
(α)a(\partial_{\alpha})^{a} natural vector basis
gabg_{ab} metric tensor on \mathscr{E}
RabcdR_{abcd} Riemann curvature tensor
a\nabla_{a} covariant derivative
εabcd\varepsilon_{abcd} Levi-Civita tensor
ξa\xi^{a} Killing vector (58), (71)
YabY^{ab} Killing-Yano tensor (60), (73)
KabK^{ab} Killing-Stäckel tensor (61), (74)
(eA)a(e_{A})^{a} orthonormal tetrad (54), (67)
ωaBC\omega_{aBC} connection 1-forms (55), (68)
Particle
\mathscr{L} particle’s worldline
pap_{a} momentum 1-form
SabS^{ab} antisymmetric spin tensor
μ,S,S\mu,S_{\circ},S_{\star} dynamical mass, spin norms (3)
τ,ua\tau,u^{a} proper time, four-velocity (7)
STDa,CaS^{a}_{\text{TD}},C^{a} spin and mass dipole wrt pap^{a} (5)   [fa=p¯a][f^{a}=\bar{p}^{a}]
Sa,DaS^{a},D^{a} spin and mass dipole wrt (e0)a(e_{0})^{a} (5)   [fa=(e0)a][f^{a}=(e_{0})^{a}]
Ξ\Xi Killing vector invariant (59), (72)
𝔎\mathfrak{K} First Rüdiger invariant (62a), (75a)
𝔔\mathfrak{Q} Second Rüdiger invariant (62b), (75b)
Phase space
\mathcal{M} 14D phase-space II.1.1
𝒩\mathcal{N} 12D symplectic leaves II.4.3
𝒯\mathcal{T} 9D sub-manifold (TD SSC) IV.1
𝒫\mathcal{P} 10D physical phase space (40)
𝒞,𝒞\mathcal{C}_{\circ},\mathcal{C}_{\star} Casimir invariants (24)
H,H𝒫H,H_{\mathcal{P}} Hamiltonian on 𝒩\mathcal{N}, on 𝒫\mathcal{P} (12), (35)
Λ,{,}\Lambda,\{\,,\} Poisson structure, brackets on \mathcal{M} or 𝒩\mathcal{N} (11), (19)
Λ𝒫,{,}𝒫\Lambda_{\mathcal{P}},\{\,,\}^{\mathcal{P}} Poisson structure, brackets on 𝒫\mathcal{P} (43)
𝒞𝒜,𝒞\mathcal{C}^{\mathcal{A}},\mathcal{C}^{\mathcal{B}} Constraints from TD SSC (37)

Appendix B Construction of canonical coordinates for (SI,DI)(S^{I},D^{I})

We now turn to the derivation of the formulae (29). This is a technical section, but with ideas that can be transposed to other constructions of symplectic variables from non-symplectic ones. Motivated by the discussion in Sec. III, we shall refer to S=(S1,S2,S3)\vec{S}=(S^{1},S^{2},S^{3}) as the spin variables, and to D=(D1,D2,D3)\vec{D}=(D^{1},D^{2},D^{3}) as the dipole variables. We shall perform the derivation in four successive steps detailed below. The strategy is to use brute force and to solve for the system of nine coupled partial differential equations (PDE) obtained by writing explicitly the Poisson brackets (19) for (S,D)(\vec{S},\vec{D}) in terms of (σ,πσ,ζ,πζ)(\sigma,{\pi_{\sigma}},\zeta,{\pi_{\zeta}}), i.e., using the formula

{F,G}=FσGπσGσFπσ+FζGπζGζFπζ,\{F,G\}=\frac{\partial F}{\partial\sigma}\frac{\partial G}{\partial{\pi_{\sigma}}}-\frac{\partial G}{\partial\sigma}\frac{\partial F}{\partial{\pi_{\sigma}}}+\frac{\partial F}{\partial\zeta}\frac{\partial G}{\partial{\pi_{\zeta}}}-\frac{\partial G}{\partial\zeta}\frac{\partial F}{\partial{\pi_{\zeta}}}\,, (78)

which holds for any functions F,GF,G of the (looked-for) symplectic variables (σ,πσ,ζ,πζ)(\sigma,{\pi_{\sigma}},\zeta,{\pi_{\zeta}}).

B.1 Spin sector SIS^{I}

Our starting point is the assumption that we can parameterize the vector S\vec{S} using only one (say σ\sigma) out of the two angles (σ,ζ)(\sigma,\zeta) available. The motivation is simple: in the limit where the dipole vanishes D=0\vec{D}=0, the variables should resemble the usual symplectic variables of SO(3). Under our assumption, in the vanishing dipole limit, the coordinate conjugated to ζ\zeta will be absent, and its conjugate πζ{\pi_{\zeta}} a constant of motion (thanks to Hamilton’s equations). This invariant should correspond to the Casimir invariant of the SO(3) algebra, namely the norm of S\vec{S}. Guided by this SO(3) analogy, our initial Ansatz for S\vec{S} reads

S1=F1(σ,πσ,πζ),S2=F2(σ,πσ,πζ)andS3=F3(σ,πσ,πζ),S^{1}=F_{1}(\sigma,{\pi_{\sigma}},{\pi_{\zeta}})\,,\quad S^{2}=F_{2}(\sigma,{\pi_{\sigma}},{\pi_{\zeta}})\quad\text{and}\quad S^{3}=F_{3}(\sigma,{\pi_{\sigma}},{\pi_{\zeta}})\,, (79)

where the functions F1,F2,F3F_{1},F_{2},F_{3} are to be determined. Next, we make a symplectic transformation that makes F3(σ,πσ,πζ)F_{3}(\sigma,{\pi_{\sigma}},{\pi_{\zeta}}) the new variable conjugated to σ\sigma. The other variables are then changed accordingly, but since they are unconstrained at this point, we may as well rename them and consider, without loss of generality, that our Ansatz is

S1=F(πσ,πζ,σ),S2=G(πσ,πζ,σ)andS3=πσ,S^{1}=F({\pi_{\sigma}},{\pi_{\zeta}},\sigma)\,,\quad S^{2}=G({\pi_{\sigma}},{\pi_{\zeta}},\sigma)\quad\text{and}\quad S^{3}={\pi_{\sigma}}\,, (80)

where the functions F,GF,G are to be determined. With the form (80), the Poisson brackets (19b), written in terms of the symplectic coordinates as in (78), take the simple form

FσGπσGσFπσ=πσ,Gσ=FandFσ=G.\frac{\partial F}{\partial\sigma}\frac{\partial G}{\partial{\pi_{\sigma}}}-\frac{\partial G}{\partial\sigma}\frac{\partial F}{\partial{\pi_{\sigma}}}={\pi_{\sigma}}\,,\quad\frac{\partial G}{\partial\sigma}=F\quad\text{and}\quad-\frac{\partial F}{\partial\sigma}=G\,. (81)

The last two brackets readily imply that S1=f1cosσ+f2sinσS^{1}=f_{1}\cos\sigma+f_{2}\sin\sigma and S2=f2cosσ+f1sinσS^{2}=-f_{2}\cos\sigma+f_{1}\sin\sigma, for some functions f1(πσ,πζ)f_{1}({\pi_{\sigma}},{\pi_{\zeta}}) and f2(πσ,πζ)f_{2}({\pi_{\sigma}},{\pi_{\zeta}}) to be determined. Inserting the latter in the first equation of (81) and solving leads to f12+f22=f(πζ)2πσ2f_{1}^{2}+f_{2}^{2}=f({\pi_{\zeta}})^{2}-\pi_{\sigma}^{2} for some arbitrary positive function denoted f(πζ)2f({\pi_{\zeta}})^{2}. Without loss of generality, one can take f=idf=\text{id} and f2=0f_{2}=0, the former thanks to the simple symplectic change of coordinates (ζ,πζ)(ζ/f(πζ),f(πζ))(\zeta,{\pi_{\zeta}})\mapsto(\zeta/f^{\prime}({\pi_{\zeta}}),f({\pi_{\zeta}})), and the latter because of the inherent SO(3) invariance of the brackets (19b) for S\vec{S}. In the end, this leads to our final formulae for S\vec{S}, namely

S1=πζ2πσ2cosσ,S2=πζ2πσ2sinσandS3=πσ,S^{1}=\sqrt{\pi_{\zeta}^{2}-\pi_{\sigma}^{2}}\cos\sigma\,,\quad S^{2}=\sqrt{\pi_{\zeta}^{2}-\pi_{\sigma}^{2}}\sin\sigma\quad\text{and}\quad S^{3}={\pi_{\sigma}}\,, (82)

as given in our main result Eqs. (29).

Having now fully exploited the three brackets (19b) satisfied by S\vec{S} to simplify the calculations, we now turn to finding coordinates for the dipole D\vec{D} with the formulae (82) at hand.

B.2 Dipole sector DID^{I}

B.2.1 Preliminaries

The fact that S3(σ,πσ,ζ,πζ)=πσS^{3}(\sigma,{\pi_{\sigma}},\zeta,{\pi_{\zeta}})={\pi_{\sigma}} leads to remarkably simple brackets in the dipole sector. Starting with the three brackets {DI,S3}\{D^{I},S^{3}\} with I=1,2,3I=1,2,3 from (19d), we find

D3σ=0,D1σ=D2andD2σ=D1.\frac{\partial D^{3}}{\partial\sigma}=0\,,\quad\frac{\partial D^{1}}{\partial\sigma}=-D^{2}\quad\text{and}\quad\frac{\partial D^{2}}{\partial\sigma}=D^{1}\,. (83)

The first equation simply means that D3=D3(πσ,ζ,πζ)D^{3}=D^{3}({\pi_{\sigma}},\zeta,{\pi_{\zeta}}), and the last two readily imply that D1,D2D^{1},D^{2} are of the form

D1=Acosσ+BsinσandD2=AsinσBcosσ,D^{1}=A\cos\sigma+B\sin\sigma\quad\text{and}\quad D^{2}=A\sin\sigma-B\cos\sigma\,, (84)

where A,BA,B are functions of (πσ,ζ,πζ)({\pi_{\sigma}},\zeta,{\pi_{\zeta}}) to be determined. The Poisson bracket {D2,S2}=0\{D^{2},S^{2}\}=0 from (19d) then implies that BB is actually independent of πσ{\pi_{\sigma}}. Therefore, so far

D1\displaystyle D^{1} =A(πσ,ζ,πζ)cosσ+B(ζ,πζ)sinσ,\displaystyle=A({\pi_{\sigma}},\zeta,{\pi_{\zeta}})\cos\sigma+B(\zeta,{\pi_{\zeta}})\sin\sigma\,, (85a)
D2\displaystyle D^{2} =A(πσ,ζ,πζ)sinσB(ζ,πζ)cosσ,\displaystyle=A({\pi_{\sigma}},\zeta,{\pi_{\zeta}})\sin\sigma-B(\zeta,{\pi_{\zeta}})\cos\sigma\,, (85b)
D3\displaystyle D^{3} =D3(πσ,ζ,πζ).\displaystyle=D^{3}({\pi_{\sigma}},\zeta,{\pi_{\zeta}})\,. (85c)

B.2.2 Finding AA and BB

From the brackets {S1,D2}\{S^{1},D^{2}\} and {D1,S2}\{D^{1},S^{2}\} in Eq. (19d) we obtain after some lengthy but straightforward algebra two PDEs for AA and BB, namely

πσB=πζAζandπζBζ+πσA=(πσ2πζ2)Aπσ,{\pi_{\sigma}}B={\pi_{\zeta}}\frac{\partial A}{\partial\zeta}\quad\text{and}\quad{\pi_{\zeta}}\frac{\partial B}{\partial\zeta}+{\pi_{\sigma}}A=(\pi_{\sigma}^{2}-\pi_{\zeta}^{2})\frac{\partial A}{\partial{\pi_{\sigma}}}\,, (86)

where we used B/πσ=0\partial B/\partial{\pi_{\sigma}}=0 along the way. Differentiating the second equation with respect to ζ\zeta and using the first one to eliminate AA with the help of Schwartz’ theorem to swap the second derivatives, we obtain the simple equation 2B/ζ2+B=0\partial^{2}B/\partial\zeta^{2}+B=0, resulting in

B(ζ,πζ)=b1(πζ)cosζ+b2(πζ)sinζ,B(\zeta,{\pi_{\zeta}})=b_{1}({\pi_{\zeta}})\cos\zeta+b_{2}({\pi_{\zeta}})\sin\zeta\,, (87)

for some functions b1b_{1} and b2b_{2}. Inserting this result back into the first equation of (86) and integrating with respect to ζ\zeta then leads to the following expression for AA,

A(πσ,ζ,πζ)=πσπζ(b1(πζ)sinζb2(πζ)cosζ)+a(πζ,πσ),A({\pi_{\sigma}},\zeta,{\pi_{\zeta}})=\frac{{\pi_{\sigma}}}{{\pi_{\zeta}}}\bigl{(}b_{1}({\pi_{\zeta}})\sin\zeta-b_{2}({\pi_{\zeta}})\cos\zeta\bigr{)}+a({\pi_{\zeta}},{\pi_{\sigma}})\,, (88)

for some function a(πζ,πσ)a({\pi_{\zeta}},{\pi_{\sigma}}). The latter can be found by inserting both Eqs. (87) and (88) into the second equation of (86). One obtains a PDE303030After many simplifications this PDE reads ((πσ2πζ2)πσπσ)a(πζ,πσ)=0\bigl{(}(\pi_{\sigma}^{2}-\pi_{\zeta}^{2})\frac{\partial}{\partial{\pi_{\sigma}}}-{\pi_{\sigma}}\bigr{)}a({\pi_{\zeta}},{\pi_{\sigma}})=0. for a(πσ,πζ)a({\pi_{\sigma}},{\pi_{\zeta}}) that integrates into

a(πσ,πζ)=a0(πζ)πζ2πσ2,a({\pi_{\sigma}},{\pi_{\zeta}})=a_{0}({\pi_{\zeta}})\sqrt{\pi_{\zeta}^{2}-\pi_{\sigma}^{2}}\,, (89)

for some unknown function a0(πζ)a_{0}({\pi_{\zeta}}). Lastly, from the expressions (87)–(89) for AA, BB and aa we obtain that of D3D^{3} by using either of the two brackets {S1,D2}=D3\{S^{1},D^{2}\}=D^{3} and {D1,S2}=D3\{D^{1},S^{2}\}=D^{3}. They lead to the same result, namely

D3(πσ,ζ,πζ)=1πσ2πζ2(b2(πζ)cosζb1(πζ)sinζ)+a0(πζ)πσπζ2πσ2.D^{3}({\pi_{\sigma}},\zeta,{\pi_{\zeta}})=\sqrt{1-\frac{\pi_{\sigma}^{2}}{\pi_{\zeta}^{2}}}\left(b_{2}({\pi_{\zeta}})\cos\zeta-b_{1}({\pi_{\zeta}})\sin\zeta\right)+a_{0}({\pi_{\zeta}}){\pi_{\sigma}}\sqrt{\pi_{\zeta}^{2}-\pi_{\sigma}^{2}}\,. (90)

It turns out that the remaining brackets from (19d) are of no additional help because of the Jacobi identity, which gives a relation between all Poisson brackets, making them not all independent from one another. Hence, we turn to the brackets (19c) to finish the calculation.

B.2.3 Integration “constants”

At this point the expressions for D\vec{D} depend on three unknown functions of the variable πζ{\pi_{\zeta}} alone, namely a0a_{0}, b1b_{1} and b2b_{2}. To find them we start from either of the two brackets {D2,D3}=S1\{D^{2},D^{3}\}=-S^{1} or {D3,D1}=S2\{D^{3},D^{1}\}=-S^{2} from Eq. (19c). They lead to the same PDE, of the form

(AD3πσ+{D3,B}(ζ,πζ)πζ2πσ2)cosσ+(BD3πσ{D3,A}(ζ,πζ))sinσ=0,\left(A\frac{\partial D^{3}}{\partial{\pi_{\sigma}}}+\{D^{3},B\}_{(\zeta,{\pi_{\zeta}})}-\sqrt{\pi_{\zeta}^{2}-\pi_{\sigma}^{2}}\right)\cos\sigma+\left(B\frac{\partial D^{3}}{\partial{\pi_{\sigma}}}-\{D^{3},A\}_{(\zeta,{\pi_{\zeta}})}\right)\sin\sigma=0\,, (91)

where {}(ζ,πζ)\{\cdot\}_{(\zeta,\pi_{\zeta})} is the bracket with respect to the sole pair (ζ,πζ)(\zeta,\pi_{\zeta}). Since both terms in parentheses are independent of σ\sigma, we obtain, thanks to the linear independence of σcosσ\sigma\mapsto\cos\sigma and σsinσ\sigma\mapsto\sin\sigma, two independent PDEs. Now we can substitute for the current expressions of A,B,D3A,B,D^{3} given in Eqs. (88), (87) and (90) into each, and after many simplifications, the first one turns into an elementary ODE for the unknown a0(πζ)a_{0}({\pi_{\zeta}}):

da0dπζ+2a0πζ=0a0(πζ)=C1πζ2,\frac{\mathrm{d}a_{0}}{\mathrm{d}{\pi_{\zeta}}}+\frac{2a_{0}}{{\pi_{\zeta}}}=0\quad\Rightarrow\quad a_{0}({\pi_{\zeta}})=\frac{C_{1}}{\pi_{\zeta}^{2}}\,, (92)

for some constant C1C_{1}\in\mathbb{R}; while the second PDE implies an algebraic relation between b1b_{1} and b2b_{2}:

b1db1dπζ+b2db2dπζ=πζ+C12πζ3b1(πζ)2+b2(πζ)2=πζ2C12πζ2C2,b_{1}\frac{\mathrm{d}b_{1}}{\mathrm{d}{\pi_{\zeta}}}+b_{2}\frac{\mathrm{d}b_{2}}{\mathrm{d}{\pi_{\zeta}}}={\pi_{\zeta}}+\frac{C_{1}^{2}}{{\pi_{\zeta}}^{3}}\quad\Rightarrow\quad b_{1}({\pi_{\zeta}})^{2}+b_{2}({\pi_{\zeta}})^{2}=\pi_{\zeta}^{2}-\frac{C_{1}^{2}}{\pi_{\zeta}^{2}}-C_{2}\,, (93)

for some constant C2C_{2}\in\mathbb{R}. At this point, the only remaining unknown function is either b1b_{1} or b2b_{2}. It can be checked that all remaining Poisson brackets are consistent with any choice of b1,b2b_{1},b_{2} provided that the constraint (93) is verified. This is the same degeneracy that occurs in the SO(3) case, as already mentioned in Sec. B.1 for the Ansatz for S\vec{S}. In order to obtain the simplest end result, we set b2(πζ)=0b_{2}({\pi_{\zeta}})=0, such that our last remaining unknown function is found to be (we chose the positive square root with no loss of generality)

b1(πζ)=πζ2C12πζ2C2,b_{1}({\pi_{\zeta}})=\sqrt{\pi_{\zeta}^{2}-\frac{C_{1}^{2}}{\pi_{\zeta}^{2}}-C_{2}}\,, (94)

and the two constants C1,C2C_{1},C_{2} remain arbitrary.

At this point our results are (82) for S\vec{S}, and a combination of (94), (92) and (85) through (90) for D\vec{D}. We can use these expressions to compute the two Casimir invariants S,SS_{\circ},S_{\star} given by (24). Most conveniently, we obtain the very simple result

C1=S2andC2=S2.C_{1}=S_{\star}^{2}\quad\text{and}\quad C_{2}=S_{\circ}^{2}\,. (95)

This concludes the proof of the expressions of (S,D)(\vec{S},\vec{D}) in terms of the symplectic coordinates (σ,πσ,ζ,πζ)(\sigma,{\pi_{\sigma}},\zeta,{\pi_{\zeta}}) and the invariants (S,S)(S_{\circ},S_{\star}), as provided in our main result (29).

Appendix C Identities for the coordinates (xα,πα,SI,DI)(x^{\alpha},\pi_{\alpha},S^{I},D^{I})

C.1 Derivation of Poisson brackets

C.2 Jacobi identity

Showing that the Poisson brackets (11) satisfy the Jacobi identity is straightforward but lengthy. However, the calculation can be shortened as follows. First, we use the coordinates (y1,,y14)=(xα,πα,SI,DI)(y^{1},\ldots,y^{1}4)=(x^{\alpha},\pi_{\alpha},S^{I},D^{I}), which are in a 1-to-1 correspondence with (xα,πα,Sαβ)(x^{\alpha},\pi_{\alpha},S^{\alpha\beta}), and, second, the equivalence between the Jacobi identity. Whether the Jacobi identity holds is, indeed, invariant under isomorphisms. Second, we can use the Jacobi identity in the equivalent form cycle{yi,{yj,yk}}=0\sum_{\text{cycle}}\{y^{i},\{y^{j},y^{k}\}\}=0, cf. Deriglazov (2022). Since any bracket involving one xαx^{\alpha} or one πα\pi_{\alpha} is either 0 or 1, the Jacobi identity will be verified for any triplet that includes at least one instance of xαx^{\alpha} or πα\pi_{\alpha}. For the other triplets, only four combinations are possible, thanks to the anti-symmetry of Poisson brackets. For example,

{SI,{SJ,SK}}+{SK,{SI,SJ}}+{SJ,{SK,SI}}=(εIJLεKLM+εJKLεILM+εKILεJLM)SM,\{S^{I}\!,\{S^{J}\!,S^{K}\}\}+\{S^{K}\!,\{S^{I}\!,S^{J}\}\}+\{S^{J}\!,\{S^{K}\!,S^{I}\}\}=(\varepsilon^{IJL}\varepsilon^{KLM}\!+\varepsilon^{JKL}\varepsilon^{ILM}\!+\varepsilon^{KIL}\varepsilon^{JLM})S^{M}\,,

where we used the brackets (19)\eqref{PBsSD} and the fact that εIJK{1,0,1}\varepsilon^{IJK}\in\{-1,0,1\}. Using the identity εIJKεLMK=δILδJMδIMδJL\varepsilon^{IJK}\varepsilon^{LMK}=\delta^{IL}\delta^{JM}-\delta^{IM}\delta^{JL}, the right-hand side of this equation is readily seen to vanish. All three other possibilities (schematically {S,{S,D}}\{S,\{S,D\}\}, {S,{D,D}}\{S,\{D,D\}\} and {D,{D,D}}\{D,\{D,D\}\}), follow exactly the same pattern and vanish for the same reason.

Appendix D Hamiltonian systems and generalizations

In this Appendix, we provide a summary of several non-equivalent notions related to Hamiltonian systems. To be clear, by Hamiltonian system, we mean a general framework in which a scalar function HH defined on some phase space manifold \mathcal{M} gives rise, through some geometrical structure Λ\Lambda, to a set of first order differential equations (ODES). Our aim is to be concise and self-contained, and to give an explicit definition of every notion that is, usually, defined implicitly. There exists many different formulations of Hamiltonian systems in the literature, with different levels of abstraction and generality. These differences are mainly concerned with the dimensionality of the phase space and the geometric structure thereon. We shall cover the ones necessary for our purposes. They include (i) canonical symplectic systems, (ii) non-canonical symplectic systems (iii) Poisson (non-symplectic) systems Our Hamiltonian formulation of the linearized MPTD + TD SSC system requires each of these four notions.

Even though all these notions can be given a coordinate-free exposition (with tensor algebra and differential geometry on the phase space manifold), we will instead choose to always work with coordinates since 1) in applications one will practically never use the coordinate-free application and 2) we do not want to mix things up with the covariant formulation of objects in GR. We refer to the recent introductory exposition Deriglazov (2022) which is at once detailed and pedagogical. Otherwise, classical textbooks such as Arnold et al. (2006) and Vaisman (2012) are always recomended.

D.1 Symplectic systems

The main distinction between different notions of Hamiltonian systems has to do with the geometrical structure on the phase space, which generates a vector field (trajectories) from the Hamiltonian function. This structure is a closed bilinear form, which, when non-degenerate, is called “symplectic”. Therefore, in what follows, symplectic is a place-holder for “non-degenerate”. The degenerate case will be dealt with later in Sec .D.2.

D.1.1 Canonical

Perhaps without knowing it, one is always introduced to symplectic systems during one’s first course on analytical mechanics. There, one learns that a 2n2n-dimensional Hamiltonian system is a set of ordinary differential equations (ODEs) for some unknowns quantities (q,p)=(qi,pi)i{1,,n}(q,p)=(q^{i},p_{i})_{i\in\{1,\ldots,n\}}. There are always an even number of such unknowns (here 2n2n for some nn\in\mathbb{N}), and the ODEs that they satisfy are called Hamilton’s equations:

dqidλ=Hpianddpidλ=Hqi,\frac{\mathrm{d}q^{i}}{\mathrm{d}\lambda}=\frac{\partial H}{\partial p_{i}}\quad\text{and}\quad\frac{\mathrm{d}p_{i}}{\mathrm{d}\lambda}=-\frac{\partial H}{\partial q^{i}}\,, (96)

where H:2nH:\mathbb{R}^{2n}\rightarrow\mathbb{R} is a scalar function of the (q,p)(q,p) called the Hamiltonian, and λ\lambda is the parameter associated to HH, usually (but not necessarily) related to some kind of “time”.313131We only cover the case where the Hamiltonian HH is autonomous, i.e., that HH does not depend explicitly on the time parameter λ\lambda. We will always work with autonomous systems in this series of work. What makes this formulation canonical is that and the 2n2n variables (qi,pi)(q_{i},p_{i}) always come in nn pairs, referred to as conjugated pairs. These variables span a 2n2n-dimensional phase space 𝒫2n\mathcal{P}\subset\mathbb{R}^{2n}, on which each point corresponds to a (possible) configuration of the underlying physical system. On top of this classical picture, one will occasionally be introduced to the notion of Poisson bracket. This is defined as an operation between two smooth functions F:𝒫F:\mathcal{P}\rightarrow\mathbb{R} and G:𝒫G:\mathcal{P}\rightarrow\mathbb{R}, defined by

{F,G}=j=1n(FqjGpjGqjFpj).\{F,G\}=\sum_{j=1}^{n}\,\left(\,\frac{\partial F}{\partial q_{j}}\frac{\partial G}{\partial p^{j}}-\frac{\partial G}{\partial q_{j}}\frac{\partial F}{\partial p^{j}}\,\right)\,. (97)

The Poisson bracket is useful because it can be used to re-write the evolution of any function F:𝒫F:\mathcal{P}\rightarrow\mathbb{R} along a solution to Hamilton’s equations (96) as follows. Using the chain rule, as well as Eqs. (96) and (97), one readily finds

dFdλ={F,H}.\frac{\mathrm{d}F}{\mathrm{d}\lambda}=\{F,H\}\,. (98)

Therefore, the evolution of any function F:𝒫F:\mathcal{P}\rightarrow\mathbb{R} along a phase space trajectory is given by its Poisson bracket with HH. From the definition (97), the Poisson brackets of two coordinates, say qi,pjq^{i},p^{j}, are also easy to find. One obtains the well-known formula

{qi,pj}=δji,\{q^{i},p_{j}\}=\delta^{i}_{j}\,, (99)

where δji\delta^{i}_{j} is the Kronecker symbol. Equation (99) is called the canonical Poisson brackets, and is a characterization of canonical coordinates. In other words, Eq. (99) is equivalent to the word “canonical” in the phrase “canonical Hamiltonian system”, or to the word “conjugated” in “conjugated pairs of variables”, all these notions being tautological. There are infinitely many distinct systems of canonical coordinates, all linked to one another through canonical transformations. The reason canonical variables on 𝒫\mathcal{P} are so useful in Hamiltonian mechanics is because 1) the general form of Hamilton’s equation (96), 2) the definition (98) of the Poisson bracket of two arbitrary functions, and 3) the fundamental brackets between canonical coordinates (99), are all invariant under canonical transformations. In other words, if (q,p)2n(Q,P)2n(q,p)\in\mathbb{R}^{2n}\mapsto(Q,P)\in\mathbb{R}^{2n} is a canonical transformation, then one can simply substitute (q,p)(q,p) by (Q,P)(Q,P) in Eqs. (96), (98) and (99), where it is understood that the Hamiltonian HH and the functions F,GF,G are expressed in terms of (Q,P)(Q,P).

D.1.2 Non-canonical

In spite of their usefulness, a canonical formulation of is but a very special way of doing Hamiltonian mechanics. Sometimes, as is the case in the present article, such a formulation is not available in the first place. What happens if we use some coordinates that do not satisfy the canonical brackets (99)? This, in essence, is answered by the broader field of symplectic geometry: a more abstract generalization of Hamiltonian mechanics, of which we give a simplified overview below.

Let us start from some canonical coordinates (q,p)2n(q,p)\in\mathbb{R}^{2n} covering a phase space 𝒫\mathcal{P}, and a non-canonical change of coordinates, i.e., a diffeomorphism Φ:(q,p)2ny2n\Phi:(q,p)\in\mathbb{R}^{2n}\mapsto y\in\mathbb{R}^{2n}. Note that, since the new coordinates y=(y1,,y2n)y=(y^{1},\ldots,y^{2n}) are not canonical, there is no reason to identify them as pairs, like the (q,p)(q,p). Let us pick two, say (yi,yj)(y^{i},y^{j}). Seing these as two functions of the (q,p)(q,p), we can compute their Poisson bracket using the canonical formula Eq. (97), and obtain the result in terms of (q,p)(q,p). According to the characterization (99) of canonical coordinates, we expect {yi,yj}δij\{y^{i},y^{j}\}\neq\delta^{ij} since yy are not canonical. Nevertheless, using the map Φ\Phi, we can express any occurrence of (q,p)(q,p) in the resulting formula in terms of the yy. We thus obtain the Poisson bracket {yi,yj}\{y^{i},y^{j}\} expressed solely in terms of yy. Doing this for each pair determines a set of Poisson brackets for the yy coordinates, given by 4n24n^{2} formulae of the form:

{yi,yj}=Λij(y),\{y^{i},y^{j}\}=\Lambda^{ij}(y)\,, (100)

where the explicit form of the right-hand side will strongly depend on the chosen coordinates. The yy-dependent matrix Λ\Lambda whose entries are Λij(y)={yi,yj}\Lambda^{ij}(y)=\{y^{i},y^{j}\} is called the Poisson matrix for the coordinates yy. Equation (100) generalizes the canonical formula (99) to non-canonical coordinates.

The Poisson matrix (100) of the coordinates yy on 𝒫\mathcal{P} is necessary and sufficient to do anything one would have done with canonical coordinates. For example, the Poisson bracket of any two functions F(y),G(y)F(y),G(y) on is simply given by

{F,G}=i,jΛij(y)FyiGyj.\{F,G\}=\sum_{i,j}\,\Lambda^{ij}(y)\,\frac{\partial F}{\partial y^{i}}\,\frac{\partial G}{\partial y^{j}}\,. (101)

where only the Leibniz rule and (100) was used. Equation (101) generalizes the canonical formula (97) to non-canonical coordinates yy. Similarly, the evolution of one coordinate yiy^{i} under some Hamiltonian H(y)H(y) is obtained directly from (101) with F=yiF=y^{i} and G=HG=H, leading to

dyidλ=jΛij(y)Hyj,\frac{\mathrm{d}y^{i}}{\mathrm{d}\lambda}=\sum_{j}\,\Lambda^{ij}(y)\,\frac{\partial H}{\partial y^{j}}\,, (102)

again, generalizing Hamilton’s canonical equations (96) to the non-canonical case. Note that Eqs. (101) and (102) can be simply written as the product of gradients of scalar functions and the matrix Λ\Lambda.

The non-canonical formulation shows that, given some phase space \mathcal{M} and a Hamiltonian HH, the key quantity that encode all information about the geometry (and thus the dynamics of the system) are the Poisson brackets between some coordinates yy on \mathcal{M} (non-necessarily canonical). Naturally, one could ask whether, given some manifold \mathcal{M} endowed with some coordinates zz on it, a set of brackets is actually sufficient in order to define a Hamiltonian system. The answer lies in a yet more general formulation, that of Poisson systems, which, in general, are degenerate.

D.2 Non-symplectic (Poisson) systems

The Poisson system formulation of a Hamiltonian system relies on the following ingredients. Let it be given (i) a manifold \mathcal{M} of dimension NN\in\mathbb{N} endowed with some coordinates y=(y1,,yN)y=(y^{1},\ldots,y^{N}), (ii) a scalar function H(y)H(y) on \mathcal{M} and (iii) a set of Poisson brackets between the yy. In what we have presented so far, items (i) and (ii) are well-defined, but (iii) is not: the definition (101) of Poisson bracket relied on the assumption that the system was non-degenerate. In the present context, the Poisson bracket is defined as an operation between two functions F(y),G(y)F(y),G(y) on \mathcal{M}, defined by

{F,G}=i,jΛij(y)FyiGyj.\{F,G\}=\sum_{i,j}\,\Lambda^{ij}(y)\,\frac{\partial F}{\partial y^{i}}\,\frac{\partial G}{\partial y^{j}}\,. (103)

where the coefficients Λij(y)\Lambda_{ij}(y) satisfy the following two properties:323232These two properties are equivalent to the other definition involving bi-vectors or Poisson brackets, Deriglazov (2022).

Skew-symmetry: Λij=Λji,\displaystyle\Lambda^{ij}=-\Lambda^{ji}\,, (104a)
Jacobi identity: ΛiΛjk+ΛjΛki+ΛkΛij=0,\displaystyle\Lambda^{i\ell}\partial_{\ell}\Lambda^{jk}+\Lambda^{j\ell}\partial_{\ell}\Lambda^{ki}+\Lambda^{k\ell}\partial_{\ell}\Lambda^{ij}=0\,, (104b)

where a sum from =1\ell=1 to =N\ell=N is implied and =/y\partial_{\ell}=\partial/\partial y^{\ell}. In other words, defining a Poisson system is a very general process: it is done by the choice of a N×NN\times N antisymmetric matrix whose coefficients Λij(y)\Lambda^{ij}(y) verify the condition (104b). Again, one refers to Λ(y)\Lambda(y) as the Poisson matrix of the system. Replacing (F,G)(F,G) in Eq. (104) by two coordinates (yi,yj)(y^{i},y^{j}), we find the meaning of its coefficients:

Λij(y)={yi,yj}.\Lambda^{ij}(y)=\{y^{i},y^{j}\}\,. (105)

What then makes a Poisson system different from a symplectic system in non-canonical coordinates? This question is natural since, after all, Eqs. (103) and (105) for Poisson systems look very much like Eqs. (101) and  (100) for non-canonical systems.

There is essentially one difference, hidden in the Poisson matrix Λ\Lambda. In a symplectic system, Λ\Lambda is necessarily non-singular, i.e., it has maximal rank at every point in phase space. This is because the yy coordinates are always a diffeomorphism away from canonical coordinates (q,p)(q,p), in which Λ=𝕁2n\Lambda=\mathbb{J}_{2n}. Since the determinant is invariant under changes of coordinates, we have det(Λ)=det(𝕁2n)=2n0\text{det}(\Lambda)=\text{det}(\mathbb{J}_{2n})=2n\neq 0, and thus rank(Λ)=2n\text{rank}(\Lambda)=2n. In a Poisson system, however, the Poisson matrix Λ\Lambda can be degenerate. For example, this is automatically the case if \mathcal{M} is odd-dimensional, because rank(Λ)\text{rank}(\Lambda) is always even by the skew-symmetry of Λ\Lambda. Yet, even when \mathcal{M} is even-dimensional can the rank of Λ\Lambda be non-maximal rank, i.e., non-equal to NN. This is the case for the brackets (19), for example.

If Λ\Lambda is degenerate, then rank(Λ)=p\text{rank}(\Lambda)=p for some 0<pN0<p\leq N. From the rank theorem, this means that their exists k=Npk=N-p independent vectors spanning the null space of Λ\Lambda. It can be shown that these vectors are the gradients of some scalar field on \mathcal{M}, called Casimirs.333333They also bear the name of Casimir element, quadratic invariant or Casimir operator in the field theory of Lie algebra and/or quantum mechanics. If C(y)C(y) is such a Casimir, then since its gradient C\nabla C belongs to the null space of Λ\Lambda, if follows from Eq. (103) that for any F:F:\mathcal{M}\rightarrow\mathbb{R}, {F,C}=0\{F,C\}=0. In other words, Casimirs are special functions of the yiy^{i} whose Poisson brackets with anything vanish on \mathcal{M}. There are exactly p=rank(Λ)p=\text{rank}(\Lambda) independent such Casimirs. A remarkable result of Poisson geometry is that \mathcal{M} is foliated by sub-manifolds of constant Casimirs. These sub-manifolds, of co-dimension pp, are called symplectic leaves and they are symplectic, i.e., non-degenerate.