Download Binary evolution in a nutshell

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

History of Solar System formation and evolution hypotheses wikipedia , lookup

Modified Newtonian dynamics wikipedia , lookup

Nebular hypothesis wikipedia , lookup

CoRoT wikipedia , lookup

Dyson sphere wikipedia , lookup

Ursa Minor wikipedia , lookup

Aquarius (constellation) wikipedia , lookup

Lyra wikipedia , lookup

Corvus (constellation) wikipedia , lookup

First observation of gravitational waves wikipedia , lookup

Future of an expanding universe wikipedia , lookup

Stellar evolution wikipedia , lookup

Accretion disk wikipedia , lookup

Cygnus X-1 wikipedia , lookup

Negative mass wikipedia , lookup

Star formation wikipedia , lookup

Transcript
Binary evolution in a nutshell
Marc van der Sluys
August 29, 2013
Contents
1 Stellar timescales
1.1 Dynamical timescale . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.2 Thermal timescale . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.3 Nuclear-evolution timescale . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2
2
2
2
2 Geometry of a binary
2.1 Masses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2 Centre of mass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3 Kepler’s law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2
3
3
3
3 The Roche potential
3.1 Roche-lobe radius . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4
4
4 Orbital energy and angular momentum
4.1 Orbital energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.2 Orbital angular momentum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5
5
5
5 Orbital angular-momentum loss
5.1 Gravitational waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.2 Non-conservative mass transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.3 Magnetic braking . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6
6
7
7
6 Spin
6.1 Black holes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.2 Neutron stars . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7
7
7
7 Mass transfer, mass loss
7.1 Drives of mass transfer . . . . . . . . . . . . .
7.2 Stability of mass transfer . . . . . . . . . . .
7.3 Stable, conservative mass transfer . . . . . . .
7.4 Stellar wind . . . . . . . . . . . . . . . . . . .
7.5 Eddington limit . . . . . . . . . . . . . . . . .
7.6 Classical common envelope . . . . . . . . . .
7.7 Envelope ejection based on AM conservation
7.8 Darwin instability . . . . . . . . . . . . . . .
1
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
8
8
8
9
9
9
10
10
11
1
1.1
Stellar timescales
Dynamical timescale
An estimate of the dynamical timescale of a star can be given by estimating how long it would take a
star of given mass and radius to collapse. This is also known as the free-fall timescale:
s
s
r
R3
(R/R )3
(R/R )3
3
−5
τdyn ∼ τff ∼
≈ 1.6 × 10 s
≈ 5.1 × 10 yr
.
(1)
GM
(M/M )
(M/M )
1.2
Thermal timescale
The thermal timescale is the timescale at which a star with given mass, radius and luminosity could
radiate away its gravitational energy. This is often approximated by the Kelvin-Helmholtz timescale:
τth ∼ τKH ∼
GM 2
(M/M )2
≈ 3.1 × 107 yr
.
RL
(R/R )(L/L )
(2)
For giants (stars with a clear core-envelope structure), one often uses:
τth ∼ τKH ∼
1.3
(M/M )(Menv /M )
GM Menv
≈ 3.1 × 107 yr
.
RL
(R/R )(L/L )
(3)
Nuclear-evolution timescale
For main-sequence stars with masses between roughly 0.2 M and 30 M , the empirical mass-luminosity
relation
3.8
L
M
≈
(4)
L
M
seems to hold reasonably well. Scaling this relation with the nuclear-evolution timescale of the Sun gives
an approximation of the nuclear-evolution timescale of a star with given mass:
τnuc ∼
M
M
L
L
−1
τnuc, ≈ 1010 yr
M
M
−2.8
.
(5)
Here we made the assumption that the fraction of the stellar mass available for hydrogen fusion is the
same for all stars. In reality, this factor will vary; it is ∼ 0.15 for the Sun, but may be much larger for
more massive stars, for example because they have convective cores and can mix material from outside
into the ‘burning’ region to be fused.
For evolutionary phases other than the main sequence, one could scale Eq.5 with the energy efficiency
of the nuclear reaction involved. However, this is only true for other core-burning phases, in which case
a burning shell often provides a significant fraction (if not the majority) of the luminosity.
If the timescale is needed in the context of mass transfer, the relevant timescale is often that of the
rate Ṙ of change of the stellar radius due to the nuclear evolution of the star. If a stellar-evolution code
is available, so that R and hence Ṙ can be computed, one can use
τnuc ≈
R
.
Ṙ
(6)
Of course, this number is only meaningful if stellar evolution is indeed the drive for the change in radius.
It is useless if for instance the star fills its Roche lobe.
2
Geometry of a binary
See Figure 1 for a sketch of the geometry. For most expressions a circular orbit is assumed. The subscript
i denotes a binary component (i = 1, 2); the subscript (3 − i) indicates the other binary component.
2
Figure 1: Schematic geometry of a binary with masses M1 > M2 , and orbital separation a. The distances
a1 , a2 are from the binary members to the centre of mass (cm).
2.1
Masses
The total mass MT , mass ratio q and reduced mass µ of a binary are given by:
M T = M1 + M2 ,
(7)
qi =
Mi
,
M(3−i)
(8)
µ=
M1 M2
.
MT
(9)
and
When working with gravitational waves, the chirp mass M and symmetric mass ratio 0 < η ≤ 1 are
often used:
q
η = M1 M2 /MT2 =
;
(10)
(1 + q)2
M = MT η 3/5 .
To compute 0 < q ≤ 1 from η, use
q=
2.2
√
1 − 1 − 4η
√
.
1 + 1 − 4η
(11)
(12)
Centre of mass
The distance to the centre of mass (cm) for each of the binary members (ai , see Fig. 1) is given by:
ai =
2.3
M(3−i)
a
MT
(13)
a2
a
a1
=
=
M2
M1
MT
(14)
M1 a1 = M2 a2 = µ a
(15)
Kepler’s law
Kepler’s law can be written as:
2π
P
2
= ω2 =
GMT
a3
1/2
4π 2
a3/2
GMT
1/3
GMT
a=
P 2/3
4π 2
(16)
P =
3
(17)
(18)
Figure 2: Three-dimensional representation of the Roche potential for a binary with q = 2. The
droplet-shaped areas in the equipotential plot on the bottom of the figure are the Roche lobes of the
two stars (thick lines — the more massive companion has the larger lobe). The points L1 , L2 en L3 are
the Lagrangian points where forces cancel (L2 lies behind the lower-mass companion). Matter can be
transferred from a Roche-lobe-filling star to its companion through L1 , and matter can be lost from the
binary into a circumbinary disc through L3 and to infinity through L2 if both stars fill their Roche lobes.
Note that L1 is not the centre of mass. From van der Sluys (2006).
3
The Roche potential
The gravitational and rotational potential in a frame corotating with the binary is called the Roche
potential (see Fig. 2), and defined as:
ΦRoche (~r) = −
GM1
GM2
1
2
−
− (~
ω × ~r) ,
|~r − ~r1 | |~r − ~r2 | 2
F~ = −∇ΦRoche .
(19)
(20)
Note that, because we’re in a corotating frame:
1
2
ω × ~r) = −∞.
lim − (~
r→∞
2
This is fine, because this applies to the corotating frame, which does not extend far beyond the binary.
Note also that in a co-rotating frame, moving particles will experience the fictitious Coriolis force:
F~c = −2m (~
ω × ~v ) .
3.1
(21)
Roche-lobe radius
An approximation for the Roche-lobe radius accurate within 2% for 0 < qi < 0.8 is given by Paczyński
(1971) (note that there is a typo in this equation in Paczyński, 1967):
1/3
2
Mi
RRl,i
≈ 4/3
.
(22)
a
MT
3
An approximation accurate (the first part) within 1% for 0 < qi < ∞ was derived by Eggleton (1983)
(see Fig. 3):
2/3
RRl,i
0.49 qi
0.44 qi0.33
≈
≈
.
(23)
1/3
2/3
a
(1 + qi )0.2
0.6 q
+ ln 1 + q
i
i
4
Figure 3: Fit of the Roche-lobe radius by Eggleton (1983) and its approximation in Eq. 23.
For a binary in which star 1 fills its Roche lobe:
Porb ≈ 0.35
R13
M1
1/2 2
1 + q1
0.2
days.
(24)
In general,
−1/2
Porb ∝ (ρ̄1 )
4
4.1
.
(25)
Orbital energy and angular momentum
Orbital energy
The orbital energy of a circular binary is given by:
Eorb = Epot + Ekin = −
GM1 M2
GM1 M2
GM1 M2
+
=−
.
a
2a
2a
(26)
The “stellar” unit for energy is
2
GM
≈ 3.79 × 1048 erg.
R
4.2
(27)
Orbital angular momentum
The angular momentum (AM) of binary component i in a circular orbit:
Ji
J = J1 + J2
=
|J~i | = Mi |~vi × ~ai | = Mi vi ai
(28)
=
Mi a2i ω
(29)
=
2
(30)
µa ω
The total orbital angular momentum J can be written in terms of the orbital period P , orbital separation
a, or both:
2π
;
P
1/3
P
M 1 M2
;
J(P ) = G2/3 1/3
2π
MT
1/2
p
Ga
J(a) = M1 M2
= µ G MT a.
MT
J(a, P )
= µ a2
5
(31)
(32)
(33)
The “stellar” unit for angular momentum is
3/2
1/2
G1/2 M R ≈ 6.05 × 1051 erg s.
(34)
Inversely, the orbital period and separation can be computed from the angular momentum:
P =
3
3
J
J
MT
2π
=
2π
;
µ
G2 MT2
G2 M1 M2
J2
a= 2
=
µ GMT
J
M1 M 2
2
MT
.
G
(35)
(36)
The orbital angular momentum for binary component i is given by:
Ji = J
M(3−i)
J
=µ
.
MT
Mi
(37)
The specific orbital angular momentum for binary component i can be written as:
hi =
5
J
Ji
=µ 2
Mi
Mi
(38)
Orbital angular-momentum loss
A general equation for angular-momentum loss from a binary can be obtained by taking the logarithmic
derivatives of Eqs. 32 and 33:
J˙
J
=
=
5.1
Ṁ1
Ṁ2
1 ṀT
1 Ṗ
+
−
+
;
M1
M2
3 MT
3P
Ṁ1
Ṁ2
1 ṀT
1 ȧ
+
−
+
.
M1
M2
2 MT
2a
(39)
(40)
Gravitational waves
Angular momentum lost due the emission of gravitational waves is given by:
dJ
dt
1/2
=
GW
=
32 G7/2 M12 M22 MT
5 c5
a7/2
!2 7/3
32 G7/3 M1 M2
2π
−
1/3
5 c5
P
M
−
(41)
(42)
T
(Peters, 1964). This can be expressed as a logarithmic derivative as:
!
8/3
J˙
32 G5/3 2π
M1 M2
= −
5
1/3
J
5 c
P
M
(43)
T
GW
= −
32 G3 M1 M2 MT
.
5 c5
a4
(44)
The timescale for angular-momentum loss due to GWs is then:
τGW
−5/3 8/3
2 J (1 + q)
MT
P
= ≈ 380 Gyr
,
q
M
day
J˙ GW
(45)
and the time until contact or merger for a given detached binary is:
tmerge =
6
τGW
.
8
(46)
5.2
Non-conservative mass transfer
If, during mass transfer in a binary, a fraction β of the transferred mass is accreted (Ṁ(3−i) = −β Ṁi )
and a fraction (1 − β) is lost from the binary with a fraction α of the specific angular momentum of the
accretor, the angular momentum lost from the system is
dJ
= −α(1 − β) a2(3−i) ω Ṁi
(47)
dt MT
1/3
P
Mi2
2/3
= −α(1 − β) G
Ṁi .
(48)
4/3
2π
MT
Written as a logarithmic derivative, this becomes
!
Mi Ṁi
J˙
= −α(1 − β)
.
J
M(3−i) MT
(49)
MT
5.3
Magnetic braking
Magnetic braking removes angular momentum from a rotating star with a magnetic field. If such a
star is in a binary, and (close to) Roche-lobe filling, tidal effects may force the star to corotate with
the orbit, essentially removing this angular momentum from the binary orbit. A prescription for the
angular-momentum loss due to magnetic braking is given by e.g. Verbunt & Zwaan (1981):
dJ
= −3.8 × 10−30 M R4 ω 3 dyn cm.
(50)
dt MB
For rapidly-rotating stars, it is assumed that the magnetic field strength may no longer increase with the
rotational period of the star, “saturating” the magnetic braking. A prescription for saturated magnetic
braking is given by e.g. Sills et al. (2000) (see van der Sluys et al., 2005, for a complete set of equations):
dJ
dt
= −K
MB
= −K
6
R
R
1/2 R
R
1/2 M
M
−1/2
M
M
−1/2
ω3 ,
ω ≤ ωcrit
2
ωωcrit
,
(51)
ω > ωcrit
(52)
Spin
6.1
Black holes
In black holes and neutron stars, the spin is often indicated by a dimensionless quantity between zero
and unity
cJ
,
(53)
aspin ≡
GM 2
where J is the spin angular momentum, M the mass and c and G are the speed of light and Newton’s
constant, respectively.
6.2
Neutron stars
For a theoretical spherical neutron star with constant density (I = 52 M R2 ) we find
aspin
4πc R2
=
≈ 0.58
5G M P
R
12 km
2 M
1.4 M
−1 P
ms
−1
.
Realistic neutron stars rotate at breakup speed around aspin ∼ 0.7 (Miller et al., 2011).
7
(54)
7
Mass transfer, mass loss
7.1
Drives of mass transfer
Mass transfer between the donor star and the accretor can be driven by a number of processes. The
mass-transfer rate can be estimated roughly from the timescale of the appropriate process τ :
Ṁ ∼
Mdonor
.
τ
(55)
Mass transfer can be driven by intrinsic changes, i.e. changes in the donor star (see Sect. 7.2):
• dynamical instability of the donor. Ṁdyn ∼
• thermal evolution of the donor: Ṁth ∼
• nuclear evolution of the donor: Ṁnuc ∼
Mdonor
τdyn
Mdonor
τth
(see Sect.1.1);
(see Sect.1.2);
Mdonor
τnuc
(see Sect.1.3);
Mass transfer may also be driven by extrinsic changes, i.e. loss of angular momentum, e.g.:
• gravitational waves;
• magnetic braking;
• mass loss from the system;
• tidal dissipation.
7.2
Stability of mass transfer
Once mass transfer (MT) commences, its stability depends on the response of the radius of the donor
star and the response of the Roche-lobe radius to the mass loss of the donor. These responses are usually
expressed using the logarithmic derivative and the Greek letter ζ:
d log R
ζ≡
(56)
d log M
(see e.g. Hjellming & Webbink, 1987; Soberman et al., 1997). In order for the mass transfer to be stable,
the donor star must remain within its Roche lobe; hence the radius of the donor star must shrink more
rapidly or expand less rapidly than the radius of the donor’s Roche lobe during the MT:
ζd ≥ ζRl 1 .
(57)
In the range of stability of mass transfer, there are three regimes:
1. MT is thermally stable, and proceeds on the nuclear-evolution timescale;
2. MT is thermally unstable, but dynamically stable, and proceeds on the thermal timescale;
3. MT is dynamically unstable, and proceeds on the dynamical timescale (as a common envelope).
In order to determine on which timescale the mass transfer occurs, one will have test whether the MT will
be thermally stable by comparing the Kelvin-Helmholtz mass-radius exponent ζd,KH to the Roche-lobe
exponent ζRl . If this test indicates thermally unstable MT, one will have to do a second test, comparing
the adiabatic exponent ζd,ad to ζRl .
The adiabatic response of a giant star’s radius to mass loss, based on a polytrope with n = 3/2, is
given to 1% accuracy by
ζad =
2 mc
1 1 − mc
mc
−
− 0.03mc + 0.2
,
3 1 − mc
3 1 + 2mc
1 + (1 − mc )−6
(58)
where mc = MHe,d /Md is the mass fraction of the donor’s helium core (Soberman et al., 1997).
1 In fact, violation of this inequality indicates that the mass-transfer rate increases. If this is only temporary, the mass
transfer may still stabilise after an initial increase in Ṁ .
8
The response of the donor’s Roche-lobe radius to the mass transfer depends on the mass ratio of the
binary q ≡ Md /Ma and the accretion efficiency 0 ≤ β ≤ 1, defined as β = 1 for conservative (all mass
lost from the donor is accreted by the secondary) and β = 0 as completely non-conservative MT:
ζRl (q, β)
=
∂ ln(RRl /a) ∂ ln q
∂ ln a
+
∂ ln Md
∂ ln q
∂ ln Md
(59)
#
M
2Md2 − 2Ma2 − Md Ma (1 − β)
2 q 1/3 1.2 q 1/3 + 1/ 1 + q 1/3
1 + β d . (60)
+
−
Ma (Md + Ma )
3
3 0.6 q 2/3 + ln 1 + q 1/3
Ma
"
=
The second term in Eq. 60 comes from the approximation for the Roche-lobe radius by Eggleton (1983,
see our Eq. 23). For a derivation, see Sect. 2.3 in Woods et al. (2012).
7.3
Stable, conservative mass transfer
If mass transfer (MT) is conservative, and ignoring stellar winds and other sources of angular-momentum
loss, Ṁ2 = −Ṁ1 , ṀT = 0 and J˙ = 0. We then find from Eq. 40:
M1 − M2
Ṁ1
ȧ
= 2Ṁ1
=2
(q − 1)
(61)
a
M1 M2
M1
Note that
• if M1 > M2 and MT happens from star 1 to star 2, Ṁ1 < 0, hence ȧ < 0 and the orbit shrinks;
• ȧ = 0 when M1 = M2 , hence the orbit starts expanding when the mass ratio flips.
Then the orbital evolution is described by:
a
amin
=
MT2
4M1 M2
where
amin =
7.4
2
,
16J 2
.
GMT3
(62)
(63)
Stellar wind
Assume a fast (vwind vorb ), isotropic stellar wind from star 1: Ṁ1 = ṀT and Ṁ2 = 0. Then:
!
J˙
M2 ṀT
1
=
= ṀT MT ,
J
MT M1
q1
(64)
wind
and:
ȧ
=2
a
J˙
J
!
−
wind
M1 + 2M2 ṀT
ṀT
=−
.
M1
MT
MT
(65)
Note that the result is independent of which star loses the mass and is generally applicable to wind mass
loss. Hence:
ȧ
Ṁ
=− ;
a
M
Ṗ
Ṁ
= −2 ;
P
M
7.5
a ∝ MT−1 ;
(66)
Porb ∝ MT−2 .
(67)
Eddington limit
When a compact object (e.g. NS, BH) accretes matter, it radiates away energy produced from the infalling matter. If this luminosity becomes sufficiently high, the radiation pressure may prevent further
accretion. Hence, there is a maximum accretion rate for an accretor, known as the Eddington limit.
9
We assume that the accreted matter is a plasma and each “particle” has the mass of a proton mp
and the Thomson cross section of an electron σT . If the luminosity force cancels out gravity on such a
particle:
GM∗ mp
L σT
=
,
(68)
FL = Fg →
2
c 4πr
r2
so that the Eddington luminosity (Ledd ) is defined as:
4πc Gmp
M∗
Ledd =
M∗ ≈ 3.3 × 104
L ,
(69)
σT
M
and the Eddington accretion limit (Ṁedd ) as:
Ṁedd =
7.6
4πc mp
R∗ ≈ 1.5 × 10−8
σT
R∗
10 km
M yr−1 .
(70)
Classical common envelope
A common envelope (CE) is used to explain many observed compact binaries (Paczynski, 1976). A CE
is initiated when mass transfer is dynamically unstable and hence (the initial phase of) a CE takes place
on the short dynamical timescale. The orbital change due to a classical common envelope is described
by energy balance between the orbit and the binding energy of the donor’s envelope:
GM1,i M2
GM1,f M2
−
Ebind,env = αCE ∆Eorb = αCE
(71)
2af
2ai
(Webbink, 1984).
In many papers and especially population-synthesis codes, the envelope binding energy is approximated by
GMenv M∗
Ebind,env ≈
,
(72)
λenv R∗
where λenv is the envelope structure parameter, often assumed to be constant. However, see Dewi &
Tauris (2000); Tauris & Dewi (2001); van der Sluys et al. (2006, 2010) on values for λenv , and Loveridge
et al. (2010) on how to approximate Ebind,env (so that you no longer need λenv ).
Still, much of the uncertainty in Ebind,env comes from the discussion of which energy sources should
be included (in particular, can the recombination energy actually be used to expel the envelope), and,
for massive stars (M & 5 − 8 M ), from the uncertainty of which fraction of the star is left after the CE
(i.e., the definition of the core mass).
7.7
Envelope ejection based on AM conservation
Nelemans et al. (2000) showed that a combination of stable mass transfer and a classical CE cannot
explain observed double white dwarfs (DWDs). They suggest a mode of envelope ejection (EE) where
angular-momentum, rather than energy, balance is important:
Jf − Ji
M1f − M1i
=γ
,
Ji
Mtot,i
(73)
where γ describes how many times the specific angular momentum of the system is carried away by the
mass loss. Nelemans et al. (2000, 2001); Nelemans & Tout (2005) show that this works well for observed
(populations of) DWDs for γ ≈ 1.5 − 1.75. However, a clear physical picture of why this works and why
γ should have these or any other values is currently lacking.
van der Sluys et al. (2006) used this idea and designed similar prescriptions for an envelope ejection
with the specific angular momentum of the donor (γd , e.g. a strong stellar wind) and accretor (γa ,
isotropic wind from the accretor) respectively:
M1f − M1i M2i
Jf − Ji
= γd
;
Ji
Mtot,f M1i
Jf − Ji
Mtot,i
M1f − M1i
= γa
exp
−1 .
Ji
Mtot,f
M2
10
(74)
(75)
They show that a γd -EE for the formation of the first WD followed by a γa -EE for the formation of
the second WD explains observed values of P , q and the difference in cooling age for DWDs well, for
γd , γa ≈ 1.0.
The γ-EE scenarios do not have to occur on the dynamical timescale, but must happen on a timescale
that is short compared to the nuclear-evolution timescale (since the companion does not accrete, and the
donor’s core mass does not grow).
7.8
Darwin instability
Apart from Roche-lobe overflow, an CE may be initiated by the Darwin instability (Darwin, 1879). If
the spin angular momentum of one of the binary components exceeds 13 Jorb , the star is tidally locked to
the orbit and the star expands (due to evolution), a stable binary orbit is no longer possible, and the two
stars are bound to plunge into each other. When the star expands, it needs to spin up by extracting AM
from the orbit, which then spins up itself, so that the star needs more AM to keep pace, et cetera. This
will result in an “AM catastrophe”, an orbital spiral-in and a CE.
References
Darwin, G. H. 1879, Proc. R. Soc., 29, 168 [LINK]
Dewi, J. D. M. & Tauris, T. M. 2000, A&A, 360, 1043 [ADS]
Eggleton, P. P. 1983, ApJ, 268, 368 [ADS]
Hjellming, M. S. & Webbink, R. F. 1987, ApJ, 318, 794 [ADS]
Loveridge, A. J., van der Sluys, M., & Kalogera, V. 2010, ArXiv e-prints [ADS]
Miller, J. M., Miller, M. C., & Reynolds, C. S. 2011, ApJL, 731, L5 [ADS]
Nelemans, G. & Tout, C. A. 2005, MNRAS, 356, 753 [ADS]
Nelemans, G., Verbunt, F., Yungelson, L. R., & Portegies Zwart, S. F. 2000, A&A, 360, 1011 [ADS]
Nelemans, G., Yungelson, L. R., Portegies Zwart, S. F., & Verbunt, F. 2001, A&A, 365, 491 [ADS]
Paczyński, B. 1967, Acta Astronomica, 17, 287 [ADS]
—. 1971, ARA&A, 9, 183 [ADS]
Paczynski, B. 1976, in IAU Symposium, Vol. 73, Structure and Evolution of Close Binary Systems, ed.
P. Eggleton, S. Mitton, & J. Whelan, 75–+ [ADS]
Peters, P. C. 1964, Physical Review, 136, 1224 [ADS]
Sills, A., Pinsonneault, M. H., & Terndrup, D. M. 2000, ApJ, 534, 335 [ADS]
Soberman, G. E., Phinney, E. S., & van den Heuvel, E. P. J. 1997, A&A, 327, 620 [ADS]
Tauris, T. M. & Dewi, J. D. M. 2001, A&A, 369, 170 [ADS]
van der Sluys, M., Politano, M., & Taam, R. E. 2010, in American Institute of Physics Conference Series,
Vol. 1314, American Institute of Physics Conference Series, ed. V. Kologera & M. van der Sluys, 13–18
[ADS]
van der Sluys, M. V. 2006, Formation and evolution of compact binaries (PhD thesis) [LINK]
van der Sluys, M. V., Verbunt, F., & Pols, O. R. 2005, A&A, 440, 973 [ADS]
—. 2006, A&A, 460, 209 [ADS]
Verbunt, F. & Zwaan, C. 1981, A&A, 100, L7 [ADS]
Webbink, R. F. 1984, ApJ, 277, 355 [ADS]
Woods, T. E., Ivanova, N., van der Sluys, M. V., & Chaichenets, S. 2012, ApJ, 744, 12 [ADS]
11