You are on page 1of 9

MNRAS 000, 1–9 (2020) Preprint 1 December 2020 Compiled using MNRAS LATEX style file v3.

Noise and waves: a unified kinetic theory for stellar systems


Chris Hamilton1? and Tobias Heinemann2
1 Department of Applied Mathematics and Theoretical Physics, University of Cambridge, Wilberforce Road, Cambridge CB3 0WA, UK
2 Niels Bohr International Academy, Niels Bohr Institute, Blegdamsvej 17, 2100 Copenhagen, Denmark

1 December 2020
arXiv:2011.14812v1 [astro-ph.GA] 30 Nov 2020

ABSTRACT
The traditional Chandrasekhar picture of the slow relaxation of stellar systems assumes that stars’ orbits are only modified by
occasional, uncorrelated, two-body flyby encounters with other stars. However, the long-range nature of gravity means that
in reality large numbers of stars can behave collectively. In stable systems this collective behaviour (i) amplifies the noisy
fluctuations in the system’s gravitational potential, effectively ‘dressing’ the two-body (star-star) encounters, and (ii) allows the
system to support large-scale density waves (a.k.a. normal modes) which decay through resonant wave-star interactions. If the
relaxation of the system is dominated by effect (i) then it is described by the Balescu-Lenard (BL) kinetic theory. Meanwhile
if (ii) dominates, one must describe relaxation using quasilinear (QL) theory, though in the stellar-dynamical context the full
set of QL equations has never been presented. Moreover, in some systems like open clusters and galactic disks, both (i) and (ii)
might be important. Here we present for the first time the equations of a unified kinetic theory of stellar systems in angle-action
variables that accounts for both effects (i) and (ii) simultaneously. We derive the equations in a heuristic, physically-motivated
fashion and work in the simplest possible regime by accounting only for very weakly damped waves. This unified theory is
effectively a superposition of BL and QL theories, both of which are recovered in appropriate limits. The theory is a first step
towards a comprehensive description of those stellar systems for which neither the QL or BL theory will suffice.
Key words: gravitation – stars: kinematics and dynamics – galaxies: star clusters: general – plasmas

1 INTRODUCTION fluctuations δΦ (with characteristic timescale ∼ tcross ) nudge stel-


lar orbits away from their mean field trajectories. The basic premise
When star clusters and galaxies form they rapidly reach stable
of kinetic theory is that while one cannot hope to calculate the effect
quasi-steady states via the process of violent relaxation (Lynden-
of stochastic fluctuations on an arbitrary star’s orbit, the accumu-
Bell 1967; Binney & Tremaine 2008). Thereafter, in the absence
lation of many such weak nudges, when applied to the entire sys-
of strong external perturbations such as mergers, these stellar sys-
tem of stars, produces a deterministic secular evolution of f on the
tems evolve very slowly. More precisely, for a stable N -star sys-
timescale ∼ trelax .
tem the characteristic evolution timescale is the so-called relaxation
time trelax ∼ N tcross , where tcross is a typical orbital period. Since Unfortunately a self-consistent calculation of the collision oper-
N ∼ 106 in globular clusters and N ∼ 1011 in typical galaxies, ator C is difficult. To avoid it, Chandrasekhar (1942, 1943a,b) pi-
there is usually a very large separation between trelax and tcross , and oneered the theory of two-body relaxation. In this theory, one first
the evolution of f is referred to as secular. considers a test star undergoing a series of uncorrelated, independent
Mathematically, secular evolution is described by a kinetic equa- two-body encounters with field stars. Except during these occasional
tion of the form encounters, all stars are assumed to follow straight line trajectories.
∂f One then adds up all the encounters that the test star experiences
= C[f, ...], (1) as if they were statistically independent, and pretends that C may
∂t
be calculated simply by treating the entire system as an ensemble
where f (x, v) is Rthe mean field distribution function (DF), defined of such test stars. Chandrasekhar’s theory has been extremely influ-
such that ρ(x) = dv f is the mean mass density at position x (v is ential in stellar dynamics. However, it has two major shortcomings.
the velocity), and the ‘...’ allows for dependence on other variables. First, it ignores the quasi-periodic nature of stars’ mean field orbits
The quantity C on the right hand side of (1) is referred to as the col- in the potential Φ (replacing them with straight lines). To remedy this
lision operator. Of course, actual physical-contact collisions of stars one must describe phase space not using position and velocity coor-
are extremely rare; instead, C is sourced by correlations between dinates (x, v) but using angle-action variables (θ, J). Second, and
small fluctuations δf of the exact microscopic DF around its mean most crucially for this paper, the long-range nature of gravity means
value f , and small fluctuations δΦ of the exact gravitational poten- that stars can behave collectively, i.e. fluctuations in the gravitational
tial around its mean Φ (Chavanis 2012). Physically, the potential potential δΦ can be sourced by coherent motions of many stars at
once. Many authors have emphasised the importance of these col-
lective effects for determining the evolution of stellar systems (e.g.
? E-mail: ch783@cam.ac.uk Julian & Toomre 1966; Weinberg 1994; Weinberg 1998; Fouvry

© 2020 The Authors


2 C. Hamilton & T. Heinemann
et al. 2015; Hamilton et al. 2018; Tamfal et al. 2020). Collective be- these waves, even if they are present, decay on a short timescale
haviour is manifestly absent in Chandrasekhar’s theory, which deals |γ|−1  trelax ∼ N tcross , where −γ is the Landau damping rate of
only with statistically independent two-body encounters — thus, a the wave (e.g. Fouvry & Bar-Or 2018, Appendix F). Because of this
more sophisticated theory is required. the impact of waves upon secular relaxation is typically neglected.
Roughly speaking, collective behaviour affects δΦ in two ways. However, this assumption is not always valid. Some damped wave
First, it amplifies (or ‘dresses’) the steady-state noise over a contin- modes in star clusters and galaxies can be very long-lasting, decay-
uous range of frequencies, analogous to the phenomenon of Debye ing on a timescale |γ|−1 & 102 tcross (Weinberg 1991; Weinberg
sheilding in an electrostatic plasma. Second, it allows the system to 1994; Sellwood & Pryor 1998; Ideta 2002). Hence in ‘intermediate-
support transient wave modes at a discrete set of modal frequencies1 ; N ’ systems such as open clusters (Danilov & Putkov 2012, 2013)
in an electron plasma these are Langmuir waves, with wavelength or the N = 103 -body simulations of Lau & Binney (2019), the as-
much larger than the Debye length. We now consider each of these sumption |γ|−1  trelax can be violated. Additionally, waves can
contributions to δΦ in turn. be important in systems with very large N . For instance, the spi-
ral structure in galactic disks (N & 1011 ) has long been posited
to consist of a superposition of weakly unstable wave modes (Sell-
1.1 Noise wood & Carlberg 2014, 2019). The interaction of these waves with
stars is thought to heat the stellar velocity distribution (Carlberg &
The exact gravitational potential of a stellar system is always noisy
Sellwood 1985; Binney & Lacey 1988; Griv et al. 2006). Thus there
simply because the system is comprised of a finite number of stars
exist many systems in which wave-star interactions can impact the
N . However, collective effects can significantly amplify this intrin-
secular evolution significantly.
sic Poisson noise (Weinberg 1998; Lau & Binney 2019). The re-
If wave-star interactions dominate the evolution of f then one
sulting amplified or ‘dressed’ noise gradually nudges stars off their
must employ a so-called quasilinear (QL) kinetic theory and take C
mean field orbits, driving secular evolution (Weinberg 1993; Fou-
equal to the QL collision operator CQL [f, {Eg }], where {Eg } is the
vry et al. 2015). Mathematically, this secular evolution is described
set of wave energies. In that case one must also prescribe an equa-
by Balescu-Lenard (BL) theory2 , so that C in equation (1) corre-
tion that determines the time-variation of each Eg . In plasma physics
sponds to the BL collision operator CBL [f ], which is a functional of
there exists a huge literature on the QL kinetic theory of wave-
f alone.
particle interactions (e.g. Tsytovich 1995; Diamond et al. 2010; Ichi-
Moreover, Rostoker’s principle (Rostoker 1964a,b; Gilbert 1968;
maru 2018 and references therein). In the gravitational context, a
Lerche 1971) tells us that BL theory can be thought of as a theory
particular QL theory was developed for application to galactic disks
of ‘dressed’ star-star interactions. In Rostoker’s picture relaxation is
and Saturn’s rings by Griv et al. (1994, 2002, 2003, 2006), but that
driven exclusively by independent two-body encounters just like in
work could not apply to arbitrary geometries since it did not employ
Chandrasekhar’s theory. The only difference is that the ‘bare’ New-
angle-action variables. Meanwhile, QL theory in angle-action vari-
tonian interaction potential ψ = −G/|x − x0 | is replaced by an
ables was developed in the context of cylindrical/toroidal plasmas
effective ‘dressed’ interaction ψ d . Said differently, Chandrasekhar’s
by Kaufman & Nakayama (1970); Kaufman (1971, 1972). Some as-
theory is a theory of star-star interactions in vacuo while the BL
pects of angle-action QL theory, such as the diffusion rate of a test
picture is one of star-star interactions mediated via the ‘dielectric
particle and the Landau damping rate for weakly damped modes,
medium’ of the other N − 2 stars. In Hamilton (2020) we showed
have been derived in stellar dynamics (Lynden-Bell & Kalnajs 1972;
how to derive CBL in a very simple way using Rostoker’s principle.
Tremaine & Weinberg 1984; Binney & Lacey 1988; Tremaine 1998;
If one ignores collective amplification then the BL theory reduces to
Nelson & Tremaine 1999). Despite this, a fully consistent set of QL
Landau theory, which is a theory of ‘bare’ star-star interactions, and
equations in angle-action space has never been written down for stel-
ultimately to Chandrasekhar’s theory (Chavanis 2013a,b).
lar systems.

1.2 Waves 1.3 Plan for this work


Stellar systems also support transient wave modes (Binney & In reality, both noise and waves are always present. Thus star-star
Tremaine 2008, §5). Unlike amplified noise, these are spatially and wave-star interactions are always in effect simultaneously. A
smooth density waves that do not have a statistically stationary am- general kinetic theory of stellar systems must therefore account for
plitude — instead they decay via Landau damping. They tend to both effects at once. In plasma theory several authors have proposed
have characteristic spatial scales comparable to the size of the clus- some form of unification of the BL and QL theories — see for in-
ter/galaxy itself. They are easily excited by external perturbations stance Nishikawa & Osaka (1965); Yamada et al. (1965); Klimon-
such as a massive flyby encounter (Murali 1999; Vesperini & Wein- tovich (1967); Rogister & Oberman (1968, 1969); Hitchcock et al.
berg 2000), or they can be self-excited by the Poisson noise inher- (1983); Baalrud et al. (2008, 2010). However this has never been
ent in the system’s gravitational potential (Weinberg 2001; Heggie done the general stellar-dynamical context.
et al. 2020). When discussing relaxation it is usually assumed that In this paper we present for the first time a unified kinetic theory of
stellar systems in angle-action variables which includes both BL and
1
QL theory as limiting cases. In the spirit of Hamilton (2020) we keep
Collective behaviour can also lead to instabilities — see e.g. Goodman the derivation as short as possible, employing heuristic techniques
(1988); Binney & Tremaine (2008); Rozier et al. (2019) and references
and emphasising physical insight throughout (a rigorous derivation
therein. In this paper we consider only stable systems, so that all wave modes
are damped, but our results also apply to weakly unstable waves (Rogister &
is postponed to future work). When considering wave-star interac-
Oberman 1968). tions we work in the simplest possible limit, accounting only for very
2 First derived by Balescu 1960; Lenard 1960 in plasma physics; the stellar- weakly damped waves (defined precisely in §4.1), ignoring wave-
dynamical equivalent was derived by Luciani & Pellat 1987; Heyvaerts 2010; wave coupling, and making the random phase approximation. In this
Chavanis 2012; Fouvry & Bar-Or 2018). limit the BL and QL parts of the theory effectively decouple so that

MNRAS 000, 1–9 (2020)


Unified kinetic theory for stellar systems 3
the unified theory is a superposition of the two (Yamada et al. 1965; is the dielectric operator which is a functional of f . Here ψ̂ is sim-
Rogister & Oberman 1968). We already derived the BL part of the ply the Newtonian interaction ψ = −G/|x − x0 | written in op-
theory in Hamilton (2020) from Rostoker’s principle; here we derive erator form, while P̂ is the polarisation operator which accounts
the QL part using a similar approach. First we ‘quantise’ each wave for collective amplification. We have chosen the normalisation and
into many small pieces (quasiparticles; in plasma theory these are notation for P̂ to coincide with that of Nelson & Tremaine (1999)
called ‘plasmons’. Then we account for the ‘emission’ and ‘absorp- — see their equation (30) — from whom we will quote various
tion’ of wave quanta by individual stars using Feynman diagrams, general results throughout this paper. An explicit expression for P̂
adding up all possible interactions as if they were independent. The in position-frequency space is given in equation (94) of Nelson &
approach we take resembles that used by Kaufman (1971) for cylin- Tremaine (1999), but will not be required here. If collective ef-
drical plasmas, but in our case the mean field has arbitrary geometry fects are switched off then P̂ = 0. If external perturbations were
(provided it is integrable). To shorten the derivation further we quote present they would also feature on the right hand side of (4) along-
several general results from the linear response theory work of Nel- side δΦball , but we will ignore them here for simplicity.
son & Tremaine (1999).
In §2 we introduce the notation that we will use for the rest of
the paper and write down the fundamental equations of the kinetic
3 DRESSED NOISE, STAR-STAR INTERACTIONS, AND
theory (though we shall not solve them formally). Then we con-
THE BALESCU-LENARD COLLISION OPERATOR
sider separately dressed noise (§3) and waves (§4), and derive their
respective contributions to the kinetic theory. In §5 we collect the In the simplest possible model of a stellar system, all stars are frozen
equations of the unified kinetic theory and discuss how one recovers on to mean field orbits dictated by the smooth potential Φ. This is
its BL and QL limits. We conclude in §6. equivalent to setting the polarization P̂ in equation (4) to zero, so
that δΦ = δΦball . If we draw stars’ initial phase space locations at
random from the mean distribution f , then the only potential fluctu-
ations will be due to Poisson noise owing to the finite number N of
2 FUNDAMENTAL EQUATIONS
stars in the system: δΦball = δΦPoisson . To make this more precise
As in Hamilton (2020), we consider a system of N  1 stars with let us write δΦ as a Fourier series:
equal mass m, and we assume the mean field potential Φ is inte- X
δΦ(x(θ, J), t) = δΦk (J, t)eik·θ . (5)
grable. Then we introduce angle-action coordinates (θ, J) such that
k
a star’s mean field specific Hamiltonian H = v2 /2 + Φ(x) does
not depend on θ, i.e. H = H(J, t), and the angle dependence of all Here k (the ‘wavenumber’) is a dimensionless 3-vector with integer
quantities is periodic under θi → θi + 2π. A star’s mean field equa- entries, and the sum is over all such 3-vectors. Then for the system
tions of motion are then dθ/dt = ∂H/∂J ≡ Ω(J) and dJ/dt = 0. we have just described we would find that the steady-state correla-
Furthermore, by Jeans’ theorem the mean field DF of stars in phase tor of potential fluctuations is (Lifshitz & Pitaevskii 1981; Chavanis
2012; Fouvry & Bar-Or 2018):
R a functionRof actions alone, f = f (J, t). We normalise
space is also
f so that dθdJf = dxdvf = N m, andR so Poisson’s equation
hδΦk (J, t)δΦ∗k0 (J0 , t0 )iPoisson
for the mean quantities reads ∇2 Φ = 4πG dvf . Finally, H and
XZ 0 0 0
f evolve on the timescale trelax  tcross , so from now on we will ∝ dJ0 |ψkk0 (J, J0 )|2 e−ik ·Ω (t−t ) f (J0 ). (6)
suppress their t arguments, simply writing H(J), f (J). k0
The fundamental equations that govern our exact N -body system
are the Klimontovich equation and the Poisson equation (Chavanis Here we used the shorthand Ω0 ≡ Ω(J0 ), and ψkk0 is the Fourier
2012): transform of ψ. Thus the correlator of the potential fluctuations is (i)
proportional to the square of the interaction potential (∼ |ψkk0 |2 ),
d (ii) oscillates at all possible dynamical frequencies available to the
(f + δf ) = 0, (2)
dt system (k0 · Ω0 ), and (iii) is weighted by thePnumber
Z R of stars capable
∇2 (Φ + δΦ) = 4πG dv (f + δf ). (3) of emitting power at those frequencies (∼ k0 dJ0 f (J0 )).
Next, let us start with the same initial conditions but this time
By linearising these equations (i.e. assuming δf  f, δΦ  Φ) switch on collective effects (a.k.a. ‘self-gravity’). That is, we al-
and waiting a few crossing times3 one can straightforwardly relate low each star to move not rigidly in the mean field Φ, but self-
the total potential fluctuation δΦ to the ‘ballistic fluctuation’ δΦball , consistently in the exact gravitational potential Φ + δΦ. Then P̂
i.e. the fluctuation that results if one takes the initial phase space is nonzero and δΦ is given by the sum of the inhomogeneous solu-
density fluctuation δf (t = 0) and assumes that all stars follow their tion and homogeneous solution of (4). These solutions correspond
mean field trajectories indefinitely. They are related via a linear in- to dressed noise and wave modes respectively. We now discuss the
tegral operator equation which schematically we may write as: dressed noise; waves will be discussed in §4.

ˆ · δΦ = δΦball , where ˆ ≡ 1 − 2π ψ̂ P̂ , (4)


3.1 Dressed noise

3 Dressed noise is the name we give to the inhomogeneous solution


One must wait for noise to be synchronised or thermalised (Rogister &
Oberman 1968). For instance, if the initial conditions are drawn randomly to (4), namely δΦ = ˆ−1 · δΦball . This time, if we measure the
from f , so that at t = 0 the only fluctuations are due to bare Poisson noise correlator of the self-consistent potential fluctuations (and ignoring
(§3), then it will take a finite time for that bare noise to be ‘dressed’ by the homogeneous solution of (4) for now) we find it is again given
collective effects (Lau & Binney 2019). In the plasma analogy, it takes a by (6) except with the replacement
finite time for an initially random collection of electrons and ions to group
into a statistically stationary configuration of Debye polarisation clouds. |ψkk0 (J, J0 )|2 → |ψkk
d 0 0 0 2
0 (J, J , k · Ω )| , (7)

MNRAS 000, 1–9 (2020)


4 C. Hamilton & T. Heinemann
0
where ψkk d
0 (J, J , ω) is the Fourier-Laplace transform of the over k and k0 mean that we scan the entire continuum of dynami-
dressed interaction potential ψ d , which in (7) has been evaluated at cal frequencies looking for these resonances. Note also that ψ d gets
the dynamical frequency ω = k0 · Ω0 . We will not write an explicit evaluated at the dynamical frequency k · Ω. The amplification due
expression for ψ d explicitly here, but note that it is a conceptually d
to collective effects will typically be large (|ψkk0|
2
 |ψkk0 |2 ) if
simple exercise in linear response theory (Binney & Tremaine 2008; k · Ω is close to a modal frequency ωg , which we discuss next.
Fouvry & Bar-Or 2018). It is non-zero at all dynamical frequencies
k0 · Ω0 accessible to the system, and can be very large at frequencies
where ˆ−1 is small. With the replacement (7) we say that the Poisson 4 WAVE MODES, WAVE-STAR INTERACTIONS, AND THE
noise has been amplified or dressed by collective effects. Since the QUASILINEAR COLLISION OPERATOR
correlator of potential fluctuations is stationary on timescales shorter
than trelax (i.e. hδΦ(t)δΦ∗ (t0 )i depends only on the time difference Dressed noise is not the only contribution to δΦ. The fact that stars
t − t0 ) we refer to it as representing steady-state dressed noise. In behave collectively also allows the system to support waves or nor-
rigorous BL deviations this is precisely the noise spectrum used to mal modes. These are periodic density oscillations analogous to the
compute the collision operator CBL (Chavanis 2012; Fouvry & Bar- fluid modes in a star or planet. When stellar orbits resonate with a
Or 2018). wave they give energy to the wave or absorb energy from it; in a
stable system the absorption dominates so the wave Landau damps
(Binney & Tremaine 2008).
3.2 Star-star interactions Mathematically, waves correspond to homogeneous solutions to
the linear operator equation (4), i.e. they have potential δΦg such
The interaction of stars with dressed noise drives evolution of f .
that ˆ · δΦg = 0. Thus, the potential of a given wave is an eigen-
Rostoker’s principle tells us that the resulting evolution is mathemat-
function of ˆ corresponding to zero eigenvalue. These eigenfunc-
ically equivalent to that of a superposition of independent two-body
tions exist for a discrete set of frequencies ωg = Ωg + iγg , where
(‘star-star’) encounters, each with effective interaction potential ψ d .
g = 1, 2, ... is an integer that labels the wave. Here Ωg ∈ R is called
In Hamilton (2020) we showed how to use Rostoker’s picture to de-
the pattern frequency (not to be confused with the vector of dynami-
rive the BL collision operator CBL . Here we briefly recap that argu-
cal frequencies Ω), and −γg ∈ R is the Landau damping rate. Now,
ment.
for every wave solution at frequency Ωg + iγg there is another at
First we consider a test star with initial phase space coordinates
−Ωg + iγg , corresponding to a wave with the same damping rate
(θ0 , J) interacting with a field star with initial coordinates (θ00 , J0 ).
but travelling in the opposite ‘direction’. For the full homogeneous
The Hamiltonian that describes this two-body interaction is
solution to (4), which we call δΦw (x, t), to be real the amplitudes of
X i(k·θ−k0 ·θ0 ) d
h = H(J) + H(J0 ) + m e ψkk0 (J, J0 , k0 · Ω0 ), these two waves must be complex conjugates of each other. With this
kk0 information it is easy to see that the general homogeneous solution
(8) is:
X
from which we can deduce the first order change in action after time δΦw (x, t) = δΦg (x, t), (10)
τ , namely δJ(θ0 , J, θ00 , J0 , τ ) — see Hamilton (2020), equation (3). g

Second we consider the entire stellar system to be an ensemble where (c.f. equation (47) of Nelson & Tremaine 1999)
of test stars, and write down a master equation that accounts for the
processes shown in Figure 1a,b. In words, the distribution function δΦg (x, t) ≡ Ag (t)Re[e−iαg φg (x)e−iΩg t ], (11)
f (J) is decremented if the test star is kicked out of state J by a and crucially the sum in (10) is limited to waves g such that Ωg > 0
field star with action J0 (Figure 1a), and incremented if it is instead (Kaufman 1971). The φg in (11) are complex spatial ‘eigenmodes’
kicked into J from some other state J − ∆J (Figure 1b). We must with units of potential (i.e. energy per unit mass); they are not or-
also account for the inverses of both diagrams. In each case the test thogonal in general. The factor Ag (t) is a real number which is pro-
star gets a kick ∆J and the field star gets a kick ∆J0 . We add up all portional to eγg t , but will have other overall slow time dependence
possible interactions by integrating over all field star actions J0 and as well. Finally, αg is an arbitrary phase.
all possible kicks ∆J, ∆J0 . When we expand the resulting master Since the operator ˆ (and hence its inverse ˆ−1 ) is calculated
equation by assuming that the kicks are small, the terms that survive purely from f , waves are a ‘collisionless’ phenomenon in the sense
involve the expectation values τ −1 h∆J∆Jiτ and τ −1 h∆J∆J0 iτ . that ωg and φg (x) are independent of N , and would be the same
Third we put these pieces together: we use the known first order even in a system with N → ∞. The calculation of {ωg , φg (x)}
change δJ(τ ) and integrate over random phases θ0 , θ00 to calculate is not the subject of this paper but is a standard, if often difficult,
h∆J∆Jiτ , and similarly for h∆J∆J0 iτ . Finally we take the limit numerical task (Weinberg 1994; Binney & Tremaine 2008; Heggie
τ → ∞ (since we are interested in the dynamics on a timescale et al. 2020).
 tcross ). The result is the BL collision operator which is given in
Hamilton (2020), equation (11):
∂ X
Z 4.1 Weakly damped modes
CBL [f ] = π(2π)3 m · k dJ0 δ(k · Ω − k0 · Ω0 )
∂J In this paper we will limit our consideration to weakly damped
kk0
  modes, by which we mean that the pattern frequency of the wave
0 0 0 2 ∂ ∂
× |ψkkd
0 (J, J , k · Ω )| k· − k0 · f (J)f (J0 ). is much larger than the Landau damping rate:
∂J ∂J0
(9) |Ωg |  |γg |. (12)
We see that non-zero contributions to CBL arise when there is a res- With this assumption, the factor Ag (t) in (11) is just a slowly vary-
onance k · Ω = k0 · Ω0 between the dynamical frequencies of test ing real amplitude that includes a factor eγg t , while the fast time
(k · Ω) and field (k0 · Ω0 ) stars. The integral over all J0 and sums dependence of any wave ∼ e−iΩg t .

MNRAS 000, 1–9 (2020)


Unified kinetic theory for stellar systems 5

Figure 1. Four possibilities for ‘scattering’ in action coordinates. Diagram (a) shows the interaction of a test star with action J and a field star with action J0 .
As a result the test star is kicked to some new action J + ∆J, decrementing the DF f (J). Similarly, panel (b) shows the test star being kicked in to the state
J, incrementing f (J). Adding up these processes and their inverses leads to the BL collision operator CBL (§3.2). Analogously, diagrams (c) and (d) show
wave-star interactions, whereby a test star is kicked out of (diagram (c)) or in to (diagram (d)) the state J by absorbing a wave quantum (curly line) with pattern
frequency Ωg . Properly weighted, these processes and their inverses add up to give the QL collision operator CQL (§4.2). The proper weighting of processes
(c) and (d) is summarised by the level diagram in Figure 2.

J+ΔJ of the wave amplitude: it does not involve Ag , nor does it depend on
the normalisation of φg (because of the factor Λ−1
g ).
(c)
Ng Ng+1
J 4.2 Wave-star interactions

Ng Ng +1 (d) Waves interact with stars4 , and these interactions modify both the

J _ ΔJ
stellar distribution f and the wave energy. The resulting evolution is
the domain of quasilinear (QL) theory. Our primary aim in the rest
of this section is to derive the QL collision operator CQL .
Figure 2. Diagram showing the processes of absorption and (stimulated To proceed we must address a subtlety behind the wave solution
and spontaneous) emission arising from wave-star interactions, as in Figures (10)-(11). The issue is that (10)-(11) is derived by solving (4) as an
1c,d. A test star’s action increases (decreases) by ∆J when it absorbs (emits) initial value problem starting at t = 0. Suppose that at t = 0 the in-
a wave quantum with frequency Ωg . While absorption is always stimulated, tial conditions are drawn randomly from f , so the only fluctuations
and hence occurs at a rate proportional to the number of wave quanta Ng , in the gravitational potential are from discreteness noise. Then equa-
emission can be either stimulated or spontaneous and so must be weighted
tions (10)-(11) tells us that immediately after t = 0 there are waves
by Ng + 1 (see Thorne & Blandford 2017, §23.3.3).
present in the system. Physically these waves are seeded by the spe-
cific initial noise fluctuation. By analogy with the plasma literature,
We now write down two standard results that apply to a weakly we say that the stellar system ‘emits’ waves by virtue of its inherent
damped wave with potential δΦg (equation (11)). First is the wave discreteness.
energy, which contains both the gravitational potential energy of the However, if we were to take (10)-(11) at face value we would
wave and the kinetic energy of oscillation of the stars that comprise conclude that waves with frequency ωg will have decayed to a neg-
it. Averaged over the period 2π/Ωg , the wave energy is proportional ligible fraction of their initial amplitude by, say, the time t = 3γg−1 .
to the square of the wave amplitude (Nelson & Tremaine 1999, equa- After this time there would be effectively no more wave-star inter-
tion (62)): actions to drive the evolution of f . However, this would be to ignore
h π i the fact that the specific discreteness noise in the system is always
Eg = Ag (t)2 × − Ωg Λg , (13) fluctuating (though it is of course stationary on average). Because
2 of this the system continuously emits new waves at frequency ωg
where which are not in general in phase with each other. So while the ini-
Z
∂PH (x, x0 , ω) tial wave may have ‘died’ after t = 3γg−1 , in the mean time many
Λg ≡ dx dx0 φ∗g (x)φg (x0 ) , (14) other waves of the same frequency will have been born. Said dif-
∂ω ω=Ωg ferently, there ought to be nothing special about choosing the initial
is a real normalisation constant with units of ~ (energy × time) and time to be t = 0 rather than, say, t = 2tcross , or t = 7.34tcross ,
can be positive or negative. Here PH is the Hermitian part of the or whatever, as long as those times are much smaller |γg |−1 , the
polarisation operator P̂ expressed in position-frequency space (an timescale on which the wave distribution changes significantly. For
explicit expression is given in Nelson & Tremaine 1999, equation a more extensive discussion of this point, see §3 of Nishikawa &
(96)). Second is the expression for the Landau growth rate (Nelson Osaka (1965).
& Tremaine 1999, equations (63), (64) and (98)): Thus, the wave ensemble at each frequency ωg really consists
of many waves with different amplitudes and phases. The prob-
(2π)3 X
Z
∂f lem is that we do not know these amplitudes or phases individu-
γg = − dJ |φgk (J)|2 δ(k · Ω − Ωg )k · , (15)
2Λg ∂J ally. However, we can get round this problem by adopting a stan-
k

where we have expanded φg (x) = k φgk (J)eik·θ . As expected,


P
Landau damping arises when a wave’s pattern frequency resonates 4 They also interact with other waves (‘mode coupling’). This nonlinear pro-
with the dynamical frequency of a star. Note that γg is independent cesses lies beyond the domain of QL theory.

MNRAS 000, 1–9 (2020)


6 C. Hamilton & T. Heinemann
dard trick from plasma kinetics (Pines & Schrieffer 1962; Wyld & we assume that ∆J is small compared to J and thereby expand the
Pines 1962; Kaufman 1971; Tsytovich 1995). We replace the com- right hand side. Throwing away terms higher than second order and
plicated set of waves at frequency ωg with a large number Ng  1 using Ng  1 we get
of discrete wave quanta (known in plasma theory as quasiparticles  
∂ X ∂f
or plasmons) of equal amplitude but randomly distributed phase. CQL = · Fg f (J) + Ng Dg · , (18)
∂J g ∂J
Precisely, we label the wave quanta p j = 0, 1, 2, .., Ng and assign
them amplitudes Ajg (t) ≡ Ag (t)/ Ng and phases αgj ≡ 2πj/Ng . where
Obviously, the total energy in these waves then works out to be
h∆Jigτ
Z
−Ag (t)2 πΩg Λg /2 = Eg (t), equivalent to that of a single wave with Fg = g
d∆J w∆J (J) ∆J ≡, (19)
amplitude Ag (t) — see equation (13). τ
h∆J∆Jigτ
Z
As we will see, the total wave energy Eg is a well-defined quantity g
Dg = d∆J w∆J (J) ∆J∆J ≡ , (20)
that obeys a conservation law (§5.1). We will also see that the diffu- τ
sion rate of stars by wave-star interactions is proportional to Eg , so it and hQigτ is the average increment in the quantity Q after time τ
is important that we can calculate it. We choose the integer Ng such (c.f. equation (6) of Hamilton 2020). Note that the 3-vector Fg is in-
that5 Eg = Ng ~Ωg σg , where σg = −sgn(Λg ) = ±1 is chosen dependent of Ng ; this is a dynamical friction term representing the
so that Ng is positive. Then to determine Eg (t) we simply require ‘recoil’ that stars feel when they spontaneously emit wave quanta.
knowledge of Ng (t). We will derive an equation for Ng (t) in §4.3; Meanwhile the 3 × 3 matrix Dg comes multiplied by Ng . This term
for now we simply note that from (13) we have: represents the diffusion of stars induced by the bath of random wave
π A2g Λg π A2g |Λg | quanta. To close the system we must also have an equation that de-
Ng = − σg = ∝ e2γg t . (16) termines Ng (t). This is the so-called wave kinetic equation.
2 ~ 2 ~
Having introduced this quantised picture, we now consider the in-
teraction of stars with wave quanta. Since the number of wave quanta 4.3 The wave kinetic equation
Ng is not conserved, the simplest wave-star interactions possible are
the ‘absorption’ or ‘emission’ of a wave quantum by a star as il- To derive the wave kinetic equation we consider the process shown
lustrated in Figure 1c,d. All other interactions are non-linear and so in6 Figure 1c (wave absorption) and its inverse (wave emission). We
do not enter our theory (Tsytovich 1995). To write down the master note that the average number of stars in the action space volume el-
equation for CQL we sum these processes and their inverses over ement surrounding J is (2π)3 f (J)dJ/m. Then by integrating over
all wave labels g. Crucially, we must take care to weight the sums all possible stellar actions J and all possible kicks ∆J we can easily
properly. The processes shown in (c) and (d) both correspond to the write down a master equation for7 Ng :
absorption of a wave quantum by a star. On the other hand, the in- ∂Ng
verse of these processes — namely emission of a wave by a star ∂t
— can be either stimulated (owing to the fact that the star sits in
(2π)3
Z
g  
a bath of waves) or spontaneous (arising just from the discreteness = dJ d∆J w∆J (J) − Ng f (J) + (Ng + 1)f (J + ∆J)
m
noise produced by the star). While any given microscopic process
(2π)3 Ng (2π)3
Z Z
has the same probability as its inverse, stimulated processes must be ∂f g
= dJ Fg · + dJ d∆Jw∆J (J)f (J),
weighted by the number of wave quanta Ng , while spontaneous pro- m ∂J m
cesses are independent of Ng and so carry no such weighting. The (21)
weighting of each processes is summarised in Figure 2. where to get the second equality we expanded the right hand side for
We are now ready to write down the master equation. We in- small ∆J to first order, and used Ng  1 and equation (19).
g
troduce w∆J (J) as the transition rate density, defined such that Now, we know from equation (16) that Ng ∝ e2γg t ; thus we
g
w∆J (J) d∆J τ is the probability that a given test star with action can identify the first term on the right hand side of (21) as 2γg Ng ,
J is scattered to the volume of phase space d∆J around J + ∆J by and use the fact that we already have an explicit expression for γg ,
a given wave quantum with frequency Ω−1 g after a time interval τ , namely equation8 (15). This term represents the loss of wave energy
where tcross , Ωg  τ  |γg |−1 . Then the master equation for CQL through Landau damping. The second term on the right hand side
reads: of (21), which is independent of the wave amplitude, must there-
Z
fore correspond to the rate of spontaneous emission of wave energy.
CQL = d∆J
Hence by multiplying (22) by ~Ωg σg we arrive at the following
Xn g   equation for the wave energy:
× w∆J (J) − Ng f (J) + (Ng + 1)f (J + ∆J)
g ∂Eg
= 2γg Eg + Sg , (22)
g  o ∂t
+ w∆J (J − ∆J) Ng f (J − ∆J) − (Ng + 1)f (J) . (17)
g
where Sg ≡ [(2π)3 /m]~Ωg σg dJd∆J w∆J
R
(J)f (J) is the spon-
(We again emphasise that the sum over g is restricted only to waves taneous emission rate. There are multiple ways to derive an explicit
with Ωg > 0). In each factor Ng + 1 in this equation, the ‘Ng ’ term expression for Sg . For instance one can calculate the rate at which
corresponds to stimulated emission while the ‘1’ term corresponds
to spontaneous emission. The other terms represent absorption. Next
6 We do not need to consider diagram (d) because we no longer care about
keeping track of the star’s actions (we are about to integrate over all J).
5 Note that we are still pursuing a purely classical calculation. The introduc- 7 We have implicitly chosen the integer N large enough (or equivalently,
g
tion of Planck’s constant ~ is merely to discretise the wave energy into very taken ~ small enough) that it may be approximated as a continuous variable.
small pieces so that Ng  1 can be well-approximated by an integer; any 8 In §4.4 we will demand that (15) be consistent with the first term on the

small constant with the dimensions of ~ will do. right hand side of (21), thereby fixing Fg .

MNRAS 000, 1–9 (2020)


Unified kinetic theory for stellar systems 7
a single star on its mean field orbit, treated as an external perturber, 5 DISCUSSION
does work on the system at frequency Ωg , then sum the contribu-
The unified kinetic theory consists of the kinetic equation for f (J):
tions from all stars (Kaufman & Nakayama 1970; Kaufman 1971).
Alternatively one can demand that the total energy of the system Etot ∂f
= CBL [f ] + CQL [f, {Eg }], (28)
is conserved, and deduce the form of Sg that way. Here we simply ∂t
write down the answer: with CBL given in (9) and CQL given in (27), alongside the wave-
(2π)3 mΩg X kinetic equation for Eg , equation (24). The system is closed by
Z
Sg = − dJδ(k · Ω − Ωg )|φgk (J)|2 f (J). (23) specifying the initial value of f (J) and the initial wave energies
Λg
k
Eg (t = 0). We note that since the eigenmodes φg only appear in
The fact that the work done on the system goes like ∼ mδ(k · Ω − (24), (27) in the form |φgk |2 /Λg , none of our results depend on their
Ωg )|φgk |2 f might have been guessed a priori. To confirm that the normalisation (see equation (14)).
various prefactors are also correct, upon gathering the results of the
unified theory we will show that the system’s total energy Etot is
indeed conserved (§5.1). 5.1 Energy conservation
Putting these pieces together we can now write down the final We can now confirm that our closed system of equations conserves
form of the wave kinetic equation: the total energy in mean field motion and waves:
∂Eg XZ |φgk (J)|2
Z
= −(2π)3 Ωg
X
dJ δ(k · Ω − Ωg ) Etot (t) ≡ (2π)3 dJH(J)f (J, t) + Eg (t). (29)
∂t Λg
k g
 
Eg ∂f
× mf (J) + k· . (24) The rate of change of this energy is
Ωg ∂J Z Z
dEtot
(c.f. equation (33) of Kaufman 1971). = (2π)3 dJH(J) CBL + (2π)3 dJH(J) CQL
dt
X ∂Eg
+ . (30)
4.4 The QL collision operator g
∂t
Our remaining task is to get explicit expressions for the vector Fg It is well known from BL theory that the first term on the right hand
and the matrix Dg that appear in the QL collision operator CQL side is zero (Heyvaerts 2010; Chavanis 2012). Let us define
(equation (18)). First, since we identified
R the first term in (21) as |φgk (J)|2
 
2γg Ng we must have 2γg = (2π)3 m−1 dJ Fg · ∂f /∂J. We can Eg (t) ∂f
βkg (J, t) ≡ − mf (J) + k· , (31)
compare this to (15) and consequently read off a consistent expres- Λg Ωg ∂J
sion for9 Fg : a factor common to both CQL and ∂Eg /∂t. Then what remains of
m X (30) is the QL contribution:
Fg (J) = − k δ(k · Ω − Ωg )|φgk (J)|2 . (25)
Λg XXZ  
k dEtot ∂
=(2π)3 dJ H(J)k · + Ωg
Second, we can compute Dg ≡ τ −1 h∆J∆Jigτ by considering an dt g
∂J
k
encounter between a star, with initial coordinates (θ0 , J), and a wave × δ(k · Ω − Ωg )βkg (J, t). (32)
quantum with frequency Ωg and random phase αg . We calculate the
first order change in the star’s action δJ(τ ), average δJδJ over all Integrating the term involving H(J) by parts and discarding bound-
possible phases θ0 , αg (c.f. equation (7) of Hamilton 2020), divide ary terms, the square bracket can be replaced by [−k · Ω + Ωg ], but
by τ , and take the limit τ → ∞. This calculation was already done this is forced to vanish by the presence of the delta function. Thus,
for example in §3 of Binney & Lacey (1988) and in §V of Kaufman energy is conserved.
(1972), so we relegate the derivation to Appendix A and only present Note that the friction and diffusion parts of the QL theory sepa-
the final result: rately conserve energy. For instance, if we ignored the friction part of
Eg X CQL and the spontaneous emission Sg (as is often done in plasma ki-
Dg (J) = − k k δ(k · Ω − Ωg )|φgk (J)|2 . (26) netics for systems involving weakly unstable waves), then we would
Ωg Λ g Ng
k simply lose the term mf (J) from the right hand side of (31), but the
We may now plug (26) and (25) into (18) to get our final expres- energy conservation argument would work just the same.
sion for the quasilinear collision operator:
∂ X X |φgk (J)|2 5.2 Recovering the QL and BL theories
CQL [f, {Eg }] = − · k δ(k · Ω − Ωg )
∂J Λg
k g The right hand side of the kinetic equation (28) is a superposition
of the BL and QL collision operators CBL , CQL . In the limit where
 
Eg (t) ∂f
× mf (J) + k· . (27) wave-star interactions dominate over the star-star interactions, one
Ωg ∂J
can set CBL = 0 and recover the purely QL theory. Though we
(c.f. equation (39) of Kaufman 1971). As expected, QL evolution is have insisted throughout that our system is stable (γg < 0), our
driven by resonances between a wave’s oscillation frequency Ωg and derivation also applies to weakly unstable systems (γg > 0) as long
a star’s dynamical frequency k · Ω. as γg−1  Ωg ∼ tcross (Rogister & Oberman 1968). The QL limit
is often appropriate for weakly unstable systems like galactic disks
9 This is not the only consistent solution for f but it is the only sensible one wherein the perturbation due to growing waves quickly dwarfs the
given that f (J) must vanish on the boundaries of action space and the result dressed noise (Griv et al. 2002).
must hold for arbitrary m. When considering hot systems like globular clusters one typically

MNRAS 000, 1–9 (2020)


8 C. Hamilton & T. Heinemann
takes the opposite (BL) limit. In this context it is argued that all REFERENCES
waves decay on a timescale |γg |−1 which is typically assumed to
Baalrud S., Callen J., Hegna C., 2008, Physics of Plasmas, 15, 092111
be ∼ tcross , and that after this short time there are effectively no Baalrud S. D., Callen J., Hegna C., 2010, Physics of Plasmas, 17, 055704
more wave-star interactions (e.g. Appendix F of Fouvry & Bar-Or Balescu R., 1960, The Physics of Fluids, 3, 52
2018). However this is not quite correct, for two reasons. First, we Binney J., Lacey C., 1988, Monthly Notices of the Royal Astronomical So-
know that |γg |−1 can be significantly longer than tcross and that in ciety, 230, 597
intermediate-N systems it might even become comparable to trelax Binney J., Tremaine S., 2008, Galactic Dynamics: Second Edition. Princeton
(§1). Second, it does not account for the fact that waves are always University Press
present because they are continuously being spontaneously emit- Carlberg R., Sellwood J., 1985, The Astrophysical Journal, 292, 79
ted by the system’s discreteness noise (§4). Indeed, we see from Chandrasekhar S., 1942, Principles of Stellar Dynamics
the wave kinetic equation (24) that in the absence of secular evolu- Chandrasekhar S., 1943a, The Astrophysical Journal, 97, 255
tion, after a few damping times the wave energy Eg will have almost Chandrasekhar S., 1943b, The Astrophysical Journal, 97, 263
Chavanis P.-H., 2012, Physica A: Statistical Mechanics and its Applications,
reached a steady state value:
391, 3680
|Sg | Chavanis P.-H., 2013a, The European Physical Journal Plus, 128, 126
Egsteady =
2|γg | Chavanis P.-H., 2013b, Astronomy & Astrophysics, 556, A93
Danilov V., Putkov S., 2012, Astronomy reports, 56, 623
dJ δ(k · Ω − Ωg )|φgk (J)|2 f (J)
P R
= mΩg P Rk . Danilov V., Putkov S., 2013, Astrophysical Bulletin, 68, 154
k dJ δ(k · Ω − Ωg )|φgk (J)|2 k · ∂f /∂J Diamond P. H., Itoh S.-I., Itoh K., 2010, Modern Plasma Physics: Volume 1,
(33) Physical Kinetics of Turbulent Plasmas. Cambridge University Press
Hence, the assumption we must actually make to recover the BL Fouvry J.-B., Bar-Or B., 2018, Monthly Notices of the Royal Astronomical
Society, 481, 4566
theory is that all γg are sufficiently large that waves with energy
Fouvry J.-B., Pichon C., Magorrian J., Chavanis P.-H., 2015, Astronomy &
Egsteady have a negligible effect on the diffusion of stars compared
Astrophysics, 584, A129
to that induced by steady-state dressed noise. For a fixed total mass Gilbert I. H., 1968, The Astrophysical Journal, 152, 1043
of the system M = N m we have Egsteady ∝ 1/N , so equivalently Goodman J., 1988, The Astrophysical Journal, 329, 612
this is a requirement that N is sufficiently large that discreteness- Griv E., Chiueh T., Peter W., 1994, Physica A: Statistical Mechanics and its
driven waves are unimportant. Applications, 205, 299
Griv E., Gedalin M., Yuan C., 2002, Astronomy & Astrophysics, 383, 338
Griv E., Gedalin M., Yuan C., 2003, Monthly Notices of the Royal Astro-
6 CONCLUSION nomical Society, 342, 1102
Griv E., Gedalin M., Yuan C., 2006, Advances in Space Research, 38, 47
In this paper we have derived a unified kinetic theory for stellar Hamilton C., 2020, arXiv preprint arXiv:2007.07291
systems that accounts for the effects of both amplified noise and Hamilton C., Fouvry J.-B., Binney J., Pichon C., 2018, Monthly Notices of
transient wave modes on secular evolution. In the approximation we the Royal Astronomical Society, 481, 2041
made (accounting for weakly damped waves only, ignoring mode Heggie D. C., Breen P. G., Varri A. L., 2020, Monthly Notices of the Royal
coupling, and making the random phase approximation) the unified Astronomical Society, 492, 6019
collision operator is essentially a superposition of the BL collision Heyvaerts J., 2010, Monthly Notices of the Royal Astronomical Society, 407,
355
operator (describing dressed star-star interactions) and the QL colli-
Hitchcock D., Hazeltine R., Mahajan S., 1983, The Physics of fluids, 26,
sion operator (describing wave-star interactions). The kinetic equa-
2603
tion for the stars’ phase space distribution f must be complemented Ichimaru S., 2018, Basic principles of plasma physics: a statistical approach.
by a wave-kinetic equation which describes the energy E g in each CRC Press
wave g. Ideta M., 2002, The Astrophysical Journal, 568, 190
We call it ‘a’ unified kinetic theory, rather than ‘the’ unified ki- Julian W. H., Toomre A., 1966, The Astrophysical Journal, 146, 810
netic theory, because throughout this paper we have made the as- Kaufman A. N., 1971, The Physics of Fluids, 14, 387
sumption that waves are weakly damped (12). In practice this may be Kaufman A. N., 1972, The Physics of Fluids, 15, 1063
a poor approximation. For instance Weinberg (1994); Murali (1999) Kaufman A. N., Nakayama T., 1970, The Physics of Fluids, 13, 956
found that dipolar ` = 1 ‘sloshing’ modes in spherical systems like Klimontovich Y. L., 1967, The Statistical Theory of Non-Equilibrium Pro-
globular clusters actually have γdip. ∼ Ωdip.  tcross ; thus they are cesses in a Plasma. International Series of Monographs in Natural Phi-
losophy Vol. 9, Elsevier Science, Kent
certainly long-lived compared to the typical orbital period in the sys-
Lau J. Y., Binney J., 2019, Monthly Notices of the Royal Astronomical So-
tem, but not necessarily weakly damped in the sense of (12). An even
ciety, 490, 478
more general theory might be required to incorporate these wave so- Lenard A., 1960, Annals of Physics, 10, 390
lutions. Lerche I., 1971, The Astrophysical Journal, 166, 207
Lifshitz E. M., Pitaevskii L. P., 1981, Physical kinetics. Butterworth Heine-
mann
ACKNOWLEDGEMENTS Luciani J., Pellat R., 1987, Journal de Physique, 48, 591
Lynden-Bell D., 1967, Monthly Notices of the Royal Astronomical Society,
We thank A. Ellis for several helpful comments on the manuscript. 136, 101
CH is funded by a Science and Technology Facilities Council Lynden-Bell D., Kalnajs A., 1972, Monthly Notices of the Royal Astronom-
(STFC) Studentship. ical Society, 157, 1
Murali C., 1999, The Astrophysical Journal, 519, 580
Nelson R. W., Tremaine S., 1999, Monthly Notices of the Royal Astronomi-
DATA AVAILABILITY cal Society, 306, 1
Nishikawa K., Osaka Y., 1965, Progress of Theoretical Physics, 33, 402
No new data were generated or analysed in support of this research. Pines D., Schrieffer J. R., 1962, Physical Review, 125, 804

MNRAS 000, 1–9 (2020)


Unified kinetic theory for stellar systems 9
Rogister A., Oberman C., 1968, Journal of Plasma Physics, 2, 33 where to get the second equality we substituted the mean field tra-
Rogister A., Oberman C., 1969, Journal of Plasma Physics, 3, 119 jectory J(t) = J(0), θ(t) = θ(0) + Ωt ≡ θ0 + Ωt.
Rostoker N., 1964a, The Physics of Fluids, 7, 479 Next we compute the ensemble average h∆J∆Jigτ , which is the
Rostoker N., 1964b, The Physics of Fluids, 7, 491 value of ∆J∆J arising from the interaction of a star with a wave
Rozier S., Fouvry J.-B., Breen P. G., Varri A. L., Pichon C., Heggie D. C., quantum after time τ (where tcross ∼ Ωg  τ  |γg |−1 ) aver-
2019, Monthly Notices of the Royal Astronomical Society, 487, 711
aged over all wave quanta j. To do so we replace the sum over j
Sellwood J., Carlberg R., 2014, The Astrophysical Journal, 785, 137 PNg R 2π
Sellwood J. A., Carlberg R. G., 2019, Monthly Notices of the Royal Astro- with an integral, Ng−1 j=1 → 0 dαg /2π, and integrate δJδJ
nomical Society, 489, 116 over all possible initial angles θ0 and phases αg (c.f. equation (7) of
Sellwood J., Pryor C., 1998, in , Highlights of Astronomy. Springer, pp 638– Hamilton 2020):
642 Z
dθ0 dαg
Tamfal T., Mayer L., Quinn T. R., Capelo P. R., Kazantzidis S., Babul A., h∆J∆Jigτ =
Potter D., 2020, arXiv e-prints, p. arXiv:2007.13763
(2π)3 2π
Thorne K. S., Blandford R. D., 2017, Modern Classical Physics: Optics, × δJ(θ0 , J, Ωg , αg , τ ) δJ(θ0 , J, Ωg , αg , τ )
Fluids, Plasmas, Elasticity, Relativity, and Statistical Physics. Princeton X Z τ Z τ
0
University Press = kk dt dt0 eik·Ω(t−t ) cgk (J, t − t0 ), (A4)
Tremaine S., 1998, The Astronomical Journal, 116, 2015 k 0 0

Tremaine S., Weinberg M. D., 1984, Monthly Notices of the Royal Astro-
where
nomical Society, 209, 729
Tsytovich V. N., 1995, in , Lectures on Non-linear Plasma Kinetics. Springer, A2g h i
cgk (J, T ) ≡ |φgk (J)|2 e−iΩg T + |φg,−k (J)|2 eiΩg T . (A5)
pp 1–22 4Ng
Vesperini E., Weinberg M. D., 2000, The Astrophysical Journal, 534, 598
Weinberg M. D., 1991, The Astrophysical Journal, 373, 391 Dividing (A4) by τ and taking τ → ∞, we get (c.f. equations (3.8)-
Weinberg M. D., 1993, The Astrophysical Journal, 410, 543 (3.9) of Binney & Lacey 1988):
Weinberg M. D., 1994, The Astrophysical Journal, 421, 481
πA2g X
Weinberg M. D., 1998, Monthly Notices of the Royal Astronomical Society, Dg = k k δ(k · Ω − Ωg )|φgk |2 , (A6)
297, 101 2Ng
k
Weinberg M. D., 2001, Monthly Notices of the Royal Astronomical Society,
which together with equation (13) gives the final expression (26).
328, 311
Wyld H., Pines D., 1962, Physical Review, 127, 1851
Yamada K., Watabe M., Ichikawa Y., 1965, Progress of Theoretical Physics,
34, 383

APPENDIX A: DERIVATION OF THE QL DIFFUSION


TENSOR
Here we will derive the QL diffusion tensor Dg which we quoted in
equation (26). To do so we consider the interaction of a star with a
bath of Ng random wave quanta with frequency ωg . From the dis-
cussion in §4.2 a generic wave quantum j has potential
Ag h j
i
δΦjg (x, t) = p Re φg (x)e−iαg e−iΩg t
Ng
X j
= δΦgk (J, t)eik·θ , (A1)
k

where the Fourier components are


Ag
δΦjgk (J, t) ≡ p
2 Ng
h j j
i
× φgk (J)e−i(αg +Ωg t) + φ∗g,−k (J)ei(αg +Ωg t) . (A2)

Note that δΦjgk = [δΦjg,−k ]∗ , as must be the case since δΦjg is real.
On the contrary φgk 6= φ∗g,−k in general since the eigenfunctions
φg are complex.
By analogy with §3.2 we now consider a test star with Hamilto-
nian h = H(J) + δΦjg . From Hamilton’s equation the change in the
test star’s action after time τ is:
Z τ
∂h
δJ(θ0 , J, Ωg , αgj , τ ) = − dt
0 ∂θ
X Z τ
= −i k dt δΦjgk (J, t)eik·(θ0 +Ωt) ,
k 0

(A3)

MNRAS 000, 1–9 (2020)

You might also like