Then $$\eta = \frac{\rho_\mathrm{crit} \Omega_\mathrm{B}}{\langle m \rangle n_\gamma}.$$
And from the Friedmann equation
$$\rho_\mathrm{crit} = \frac{3 H_0^2}{8 \pi
G}.$$
Photon number density today
$$n_\gamma = 2 \zeta(3)
T_0^3 / \pi^2$$
Mean mass per baryon $\langle m \rangle \approx
m_\mathrm{p}$ (but smaller due to Helium binding)
Sakharov conditions
Assume $B=0$ when the universe was created; $B>0$
later.
In 1967 Andrei Sakharov (implicitly) wrote down the
necessary (but not sufficient) conditions for baryogenesis:
Baryon number $B$ violation
$C$ and $CP$ violation
Departure from thermal equilibrium
These specify only what is needed,
not how it works.
More on the Sakharov conditions: $C$
Note that if we had $B$ violation without $C$
violation, then $\overline{B}$ violation
would occur at the same rate:
$$ \Gamma(X \to Y + B) = \Gamma(\overline{X} \to
\overline{Y} + \overline{B}) $$
Thus over time $B=0$ still, unless we have $C$
violation too:
$$ \frac{\mathrm{d} B}{\mathrm{d} t} \propto \Gamma(\overline{X} \to
\overline{Y} + \overline{B}) - \Gamma(X \to Y + B).$$
More on the Sakharov conditions: $CP$
In fact, also need $CP$ violation
Consider $B$-violating $X \to
q_\mathrm{L} q_\mathrm{L}$ process making left handed
baryons
$CP$ symmetry turns this equation into
$ \bar{X} \to \bar{q}_\mathrm{R}
\bar{q}_\mathrm{R}$
The above expression is very similar to that for
the sphaleron rate - the two processes have much in
common
Limiting cases for $S_3$
As discussed above, solve for bounce profile by
shooting.
Identify two limiting cases:
For small supercooling ($T_\mathrm{c} - T_N \ll
T_\mathrm{c}$), bubbles are thin-wall type (with
$\tanh$ walls).
For large supercooling, bubbles are
close to a Gaussian.
Beyond $S_3$
Using $S_3(T)/T$ as the exponential parameter in the
nucleation rate is a high-temperature approximation.
One can also compute the nucleation rate
nonperturbatively, both the prefactor and the
exponential part. Results suggest that (for SM):
True supercooling lies between 1- and
2-loop results
2-loop perturbative surface tension close
to true result
Unfortunately, nucleation rate only studied at one
point in the dimensionally reduced SM theory - so
generally still follow the usual anaysis
Making use of $\Gamma$
The nucleation rate $\Gamma$ gives the probability
of nucleating a bubble per unit volume per unit
time.
More useful for cosmology is to consider the inverse
duration of the phase transition, defined as
$$ \beta \equiv - \left.\frac{d S(t)}{d t}
\right|_{t=t_*} \approx \frac{\dot \Gamma}{\Gamma}
$$
The phase transition completes when the probability
of nucleating one bubble per horizon volume is of order
1
$$ S_3(T_*)/T_* \sim -4 \log \frac{T_*}{m_\mathrm{Pl}}
\approx 100
$$
Making further use of $\Gamma$
Using the adiabaticity of the expansion of the
universe the time-temperature relation is
$$ \frac{d T}{d t} = - T H $$
This gives, for the ratio of the inverse phase
transition duration relative to the Hubble rate,
$$ \frac{\beta}{H_*} = T_* \left. \frac{dS}{dT}
\right|_{T = T_*} = T_* \frac{d}{d T}
\left. \frac{S_3(T)}{T} \right|_{T=T_*}$$
If $\frac{\beta}{H_*} \lesssim 1$ then the phase
transition won't complete...
Nucleation - conclusion
Nucleation rate per unit volume per unit time $\Gamma$
computed from bounce actions $S(T) = \mathrm{min}\{
S_3(T)/T ,S_4(T) \}$
Inverse duration relative to Hubble rate
$\frac{\beta}{H_*}$ computed from $\Gamma$, and controls
GW signal
To get $\beta$:
Find effective potential $V_\text{eff}(\phi,T)$
Compute $S_3(T)/T$ (or $S_4(T)$) for extremal
bubble by solving 'equation of motion'
Determine transition temperature $T_*$
Evaluate $\beta/H$ at $T_*$
Use $\beta/H_*$ as input
to the GW power spectrum.
3: Wall velocities
Motivation
Wall velocity connects the electroweak phase
transition to the two big unknowns:
Baryogenesis (rate of baryon asymmetry
production)
We have so far been using field theory
equations of motion.
Less tricky, but more abstract, are:
Boltzmann equations
Hydrodynamic equations
In particular, the hydrodynamic equations we get are
a valuable motivation for the rest of today's lectures
We will now look at how to arrive at these
higher-level approximations
Boltzmann equations: a reminder
What is a Boltzmann equation?
Phase space is positions $\mathbf{x}$ and momenta
$\mathbf{k}$.
Tells us how our distribution functions
$f_i(\mathbf{x},\mathbf{k})$ evolve.
Consists of four parts:
Time evolution $$\partial_t
f_i(\mathbf{x},\mathbf{k}).$$
Streaming terms in momentum and position space $$\dot{\mathbf{x}} \cdot \nabla_\mathbf{x} f +
\dot{\mathbf{p}} \cdot \nabla_\mathbf{p} f$$
Collision $$C[f]$$
Boltzmann equation for distribution $f$
The Boltzmann equation is
$$ \frac{\mathrm{d} f}{\mathrm{d} t} = \frac{\partial
f}{\partial t} + \dot{\mathbf{x}} \cdot \nabla_\mathbf{x} f +
\dot{\mathbf{p}} \cdot \nabla_\mathbf{p} f = -C[f].$$
This is a semiclassical approximation to the quantum
Liouville equations for all the fields
Only valid when the momenta of the fields is much
higher than the inverse wall thickness:
$$ p \gtrsim g T \gg \frac{1}{L_w}. $$
Very difficult to work with directly, so model the
distribution $f_i$ of each particle with a 'fluid'
ansatz.
Fluid approximation
As mentioned, fluid approximation sets the
scene for the rest of these lectures on the
electroweak phase transition
In short, we have
$$
T_{\mu\nu}^\text{fluid} = \sum_i \int \frac{d^3
k}{(2\pi)^3 E_i} k_\mu k_\nu f_i(k) = w u_\mu u_\nu -
g_{\mu\nu} p $$
but we will try to justify this.
Deriving the fluid approximation
The flow ansatz is
$$ f_i(k,x) = \frac{1}{e^X \pm 1} = \frac{1}{e^{\beta(x)
( u^\mu(x) k_\mu + \mu(x))} \pm 1} $$
with four-velocity $u^\mu(x)$, chemical potential
$\mu(x)$ and inverse temperature $\beta(x)$.
Substituting this ansatz into the Boltzmann
equations for the system yields (after much algebra!) a
(relativistic) Euler momentum equation
$$ u^\mu \partial_\mu u_\nu + \partial_\nu p = C.$$
The field-fluid model
Energy conservation requires that
$$ \partial_\mu T^{\mu\nu} =
\partial_\mu (T^{\mu\nu}_\phi +
T^{\mu\nu}_\text{fluid}) = 0. $$
We are now ready to present the full model:
$$ \begin{align} (\partial_\mu \partial^\mu \phi) \partial^\nu \phi - \frac{\partial
V_\text{eff}(\phi,T)}{\partial \phi} \partial^\nu \phi & = -\eta(\phi, v_\mathrm{w})
u^\mu \partial_\mu \phi \partial^\nu \phi \\
\partial_\mu (w u^\mu u^\nu) - \partial^\nu p + \frac{\partial
V_\text{eff}(\phi,T)}{\partial \phi} \partial^\nu \phi & = + \eta(\phi, v_\mathrm{w})
u^\mu \partial_\mu \phi \partial^\nu \phi
\end{align}
$$
Besides the (dimensionful) definition here, one
choice for $\eta$ that is well motivated is
$\tilde{\eta} \frac{\phi^2}{T}$.
This model is the basis of spherical and 3D
simulations. One can also obtain steady-state equations.
The field-fluid model: observations
Consider the fluid equation:
$$
\partial_\mu (w u^\mu u^\nu) - \partial^\nu p + \frac{\partial
V_\text{eff}(\phi,T)}{\partial \phi} \partial^\nu \phi = \eta(\phi, v_\mathrm{w})
u^\mu \partial_\mu \phi \partial^\nu \phi
$$
Away from the bubble wall, the right hand side goes
to zero. The left hand side has no length scale.
Therefore any fluid solution must be parametrised by
a dimensionless ratio, e.g. radius of the bubble to time
since nucleation - define $\xi = r/t$.
Fluid profiles will scale with the bubble radius:
they are large, extended objects!
Runaway walls?
We have assumed that the wall reaches a terminal
velocity (less than $c$).
But what if it doesn't? Termed a 'runaway wall'.
Consequences would include:
Less interaction with
plasma
Lower amplitude of GWs
Runaway walls are currently a hot topic - with a
recent paper suggesting that they may not exist (due to
subleading corrections arising from the treatment of
gauge bosons)
Wall velocities: conclusion
Detailed studies have been carried out of the wall
velocity, using thermal field theory techniques.
Higher level calculations and simulations use an
effective field-fluid model, with the wall velocity
as an input parameter.
The damping term for field-fluid models (and hence the
wall velocity) is generally obtained by a qualitative
matching to the Boltzmann equations.
4: Thermodynamics
Motivation
In the previous section we described the various
layers of approximation up to the field-fluid model.
Now we will use that field-fluid model (and
steady-state results) to explore the macroscopic behaviour
of the wall.
This is important both for baryogenesis and also for
the GW power spectrum.
Further reading
Energy budget: Espinosa,
Konstandin, No and
Servant arXiv:1004.4187
Bubbles expand with finite velocity
($v_\mathrm{w}$)
Extensive fluid shell around bubble
Latent heat $\mathcal{L}$ turned into fluid
KE???
Object of this section is to quantify how much of
the latent heat ends up as kinetic energy.
Define phase transition strength
$$ \alpha_T = \frac{\mathcal{L(T)}}{g(T) \pi^2 T^4/30}
= \frac{\text{latent heat at $T$}}{\text{radiation energy
at $T$}}
$$
which tell us how much of the energy of the universe
was stored as latent heat in the phase transition.
Computing the efficiency
Larger $\alpha_T$ ⇒ stronger phase transition
But it does not tell us how much of $\mathcal{L}$
ends up as fluid kinetic energy
For that we define the efficiency
$$ \kappa_\mathrm{f} = \frac{w u_i u_i}{\mathcal{L}}= \frac{\text{fluid
KE}}{\text{latent heat}}$$
Then $\kappa_\mathrm{f} \alpha_T$ is the fraction of the energy
density in the universe that ends up as fluid kinetic
energy at the transition.
Very roughly, $\kappa_\mathrm{f} \alpha_T \approx
\overline{U}_f^2$, the Lorentz-boosted mean square fluid velocity as the
transition completes.
Can be computed more accurately either from
spherical simulations or directly solving.
One can also define
$$\kappa_\phi = \frac{\sigma}{\mathcal{L}} \frac{S}{V}
= \frac{\text{scalar field
gradients}}{\text{latent heat}} $$
Note that because this scales as $S/V$, the surface
area over the volume, this is suppressed by the inverse
bubble radius.
Hence for realistic thermal phase transitions,
$\kappa_\phi$ is small.
Thermodynamics: conclusion
Thermal first-order transitions have a reaction
front
Reaction fronts can be deflagrations (generally
subsonic), detonations (supersonic) or hybrids (a
mixture).
The fluid reaches a scaling profile in $\xi = r/t$
based on the available latent heat and wall velocity.
From this, one can compute the efficiency
$\kappa_\mathrm{f}$ and hence how much of the energy in
the universe ends up in the fluid $\kappa_\mathrm{f}
\alpha_T$.
Recap
What parameters have we introduced?
EWPT introduction: latent heat
Nucleation: inverse duration $\beta$
Wall velocities: $v_\mathrm{w}$
Thermodynamics: $\alpha_T$ and $\kappa$
That more or less summarises what we need to know
about the physics of the phase transition, so we can now
talk about the production of GWs.
5: Two approximations
Motivation
In this section we will briefly look at two
widely-used but simple approximations.
First, the quadrupole approximation makes a
reappearance.
We will see why (a version of) the quadrupole formula
is a bad approximation for bubbles
The next approximation is the envelope
approximation
This was widely used until recently for studying
bubble collisions.
It is still important for vacuum transitions where the
scalar field walls are all that matters (and $\kappa_\phi$
can dominate)
Starting point is the Weinberg formula $$ \frac{d
E_\text{GW}}{d\omega \, d\Omega} = 2 G \omega^2
\Lambda_{ij,lm} (\hat{\bf{k}}) T_{ij}^* (\hat{\bf{k}},
\omega) T_{lm}(\hat{\bf{k}},\omega)$$ with $$ T_{ij}
(\hat{\bf{k}}, \omega) = \frac{1}{2\pi} \int \mathrm{d}t
\, e^{i\omega t} \int \mathrm{d}^3 x \, e^{-i\omega
\hat{\bf{k}}\cdot \bf{x}} T_{ij}(\bf{x},t)$$ and
$$ \Lambda_{ij,lm} \equiv P_{ij}(\hat{\bf{k}})
P_{lm}(\hat{\bf{k}}) - \frac{1}{2} P_{ij} (\hat{\bf{k}})
P_{lm}(\hat{\bf{k}}) $$
where
$$ P_{ij}(\hat{\bf{k}}) = \delta_{ij} - \hat{\mathbf{k}}_i
\hat{\mathbf{k}}_j $$
Quadrupole approximation
Consider a pair of vacuum scalar bubbles along the $z$-axis
In integral for $T_{ij}$ take $\hat{\mathbf{k}}\cdot
\mathbf{x} \to 0$, such that
$$ T_{ij}(\hat{\mathbf{k}},\omega) \to T_{ij}^Q(\omega)
\equiv \frac{1}{2\pi} \mathrm{d}t \, e^{i\omega t} \int
\mathrm{d}^3 x \, T_{ij} (\mathbf{x},t) $$
Using cylindrical symmetry...
$$T_{ij}^Q (\omega) = T_{xx}^Q (\omega) + T_{yy}^Q
(\omega) + T_{zz}^Q (\omega) = D(\omega) \delta_{ij} + \Delta
(\omega) \delta_{iz} \delta_{jz}$$
where only $\Delta(\omega)$ sources gravitational waves.
So, in the quadrupole approximation
$$ \frac{\mathrm{d}E}{\mathrm{d}\omega \,
\mathrm{d}\Omega} = G \omega^2 \left| \Delta(\omega)
\right|^2 \sin^4 \theta $$
Here $\Delta(\omega)$ can encode details of the bubble
walls interacting, and can be found numerically.
The stress energy tensor of the system
$T_{ij}(\mathbf{x},t)$ can be turned into a sum of
uncollided areas $S_n$ of each of the $n$ bubbles:
$$ T_{ij}(\mathbf{k},\omega) = \frac{1}{2\pi} \int
\mathrm{d} t \, e^{i\omega t} \sum_n \int_{S_n}
\mathrm{d}\Omega \int \mathrm{d}r \, r^2 e^{-i\omega
\hat{\mathbf{k}} \cdot (\mathbf{x}_n + r\hat{\mathbf{x}})}
T_{ij,n}(r,t)$$
and then if we assume the walls are thin
$$
4 \pi \int \mathrm{d}r \, r^2 e^{-i\omega
\hat{\mathbf{k}} \cdot (\mathbf{x}_n + r\hat{\mathbf{x}})}
T_{ij,n}(r,t) \\ \approx \frac{4\pi}{3} e^{-i\omega
\hat{\mathbf{k}} \cdot (\mathbf{x}_n + R_n(t)
\hat{\mathbf{x}})} \hat{\mathbf{x}}_i \hat{\mathbf{x}}_j
R_n(t)^3 \underbrace{\kappa \rho_\text{vac}}_{\text{i.e.}\, \sigma}.$$
Envelope approximation: implementation
With the approximation listed above, we get a double
oscillatory integral:
$$ \begin{align}
T_{ij} (\hat{\mathbf{k}}, \omega) & = \kappa
\rho_\text{vac} v_\mathrm{w}^3
C_{ij}(\hat{\mathbf{k}},\omega) \\
C_{ij} (\hat{\mathbf{k}},\omega) & = \frac{1}{6\pi}
\sum_n \int \mathrm{d} t \, e^{i \omega(t
-\hat{\mathbf{k}}\cdot \mathbf{x}_n)} (t-t_n)^3
A_{n,ij}(\hat{\mathbf{k}},\omega) \\
A_{n,ij}(\hat{\mathbf{k}},\omega) & = \int_{S_n}
\mathrm{d} \Omega e^{- i \omega v_\mathrm{w} (t -
t_n)\hat{\mathbf{k}}\cdot \hat{\mathbf{x}}}
\hat{\mathbf{x}}_i \hat{\mathbf{x}}_j
\end{align}
$$
Then evaluate these time-domain Fourier transforms
numerically
Integrate over uncollided areas $S_n$ at each
timestep.
Note that all $\hat{\mathbf{k}} \cdot \mathbf{x} \neq
0$, i.e. full result
Wall velocities top to bottom $v_\mathrm{w} =
\{1,0.1,0.01\}$.
Total power scales as $v_\mathrm{w}^3$.
Peak at $\omega/\beta \approx 1$.
Power laws on both sides of peak.
Envelope approximation: results
Simple power spectrum:
One length scale (average radius $R_*$)
Two power laws ($\omega^3$, $ \sim
\omega^{-1}$)
Amplitude
⇒ 4 numbers define spectral form
NB: Used to be applied to shock waves (fluid KE),
now only use for bubble wall (field gradient energy)
Envelope approximation
4-5 numbers parametrise the transition:
$\alpha_{T_*}$, vacuum energy fraction
$v_\mathrm{w}$, bubble wall speed
$\kappa_\phi$, conversion 'efficiency' into gradient energy
$(\nabla \phi)^2$
Transition rate:
$H_*$, Hubble rate at transition
$\beta$, bubble nucleation rate
[only matters for vacuum/runaway transitions]
Envelope approximation: comparison with full scalar
field simulations
Envelope approximation: comparison with fluid source
Envelope approx.: recent developments
The envelope approximation is a semi-numerical method
which depends on multidimensional oscillatory integrals.
It is difficult to implement accurately at high $f$,
so the high-frequency power laws are not fully understood.
In a recent paper, Jinno and Takimoto reproduced the
results of the envelope approximation in a novel way
The calculation of Jinno and Takimoto
Working in the same framework as the envelope
approximation, further analytical progress
Express the total power spectrum in terms of the
unequal time correlator $\langle T_{ij}(\mathbf{x},t_x)T_{lm}(\mathbf{y},t_y) \rangle $. The authors split it into two parts:
A 'single-bubble' part, where the two points
$\mathbf{x}$ and $\mathbf{y}$ lie on the surface of
the same bubble.
A 'double-bubble' parts, where they lie on two
intersecting bubble walls.
These contributions are summed over.
This allows the $k^{-1}$ high-power behaviour to be
seen analytically by Taylor expanding the resulting
correlator.
Equation of motion is
$$ - \ddot{h}_{ij}(x,t) +\nabla^2 h_{ij}(x,t) = 16 \pi G
T^\text{source}_{ij}(x,t). $$
where the sources are
$$ T^{\text{source},\,\phi}_{ij} = \partial_i \phi
\partial_j \phi; \qquad T^{\text{source},\,\text{fluid}}
= w u_i u _j $$
Power law behaviour above peak is between $k^{-2}$ and
$k^{-1}$
“Ringing” due to simultaneous nucleation, unimportant
From $\phi$ and $u_\mu$ to $h_{ij}$ and $\Omega_\text{GW}$
As discussed, simply evolve:
$$ \square h_{ij}(x,t) = 16 \pi G
T_{ij}^\text{source}(x,t). $$
Note that when $T_{ij}^\text{source}(x,t) =
w(x)u_i(x)u_j(x)$ this is basically a convolution of the
fluid velocity power (assuming $w(x) \approx \bar w$)
Caprini, Durrer and Servant
When we want to measure the energy in gravitational
waves, we do the projection to TT and measure:
$$ t_{\mu\nu}^\text{GW}= \frac{1}{32 \pi G}\langle
\partial_\mu h^\text{TT}_{ij} \partial_\nu
h^\text{TT}_{ij} \rangle; \quad
\rho_\text{GW} = \frac{1}{32 \pi G} \langle
\dot{h}^\text{TT}_{ij} \dot{h}^\text{TT}_{ij}
\rangle. $$
We can then redshift this to present day to get
$\Omega_\text{GW} h^2$.
Without solving the field theory equations of motion
for everything (e.g. with hard thermal loops) or
doing the Boltzmann equations, simulating the field-fluid
model is the best we can do.
Current cutting-edge simulations are still
frustratingly small in size, need to extrapolate.
Simulations too short to study turbulence.
Therefore, use simulation results to derive ansätze
and models, and combine with theoretical results where
required to make predictions.
Models and predictions
Motivation
For a given model - Higgs singlet, 2HDM, ... - compute
the GW power spectrum.
Approximately 4 inputs $\alpha$, $\beta$,
$v_\mathrm{w}$, $T_*$, all derivable from the
phenomenological model
Perturbation theory
(effective potential, etc.)
Nonperturbative simulations
Output: $\Omega_\text{gw} h^2$
Then compare to LISA sensitivity curve (and others) and see if we
could detect it
The peak frequency in the spectral shape is given by
$$
f_\text{env} = 16.5 \, \mu\mathrm{Hz} \,
\left(\frac{f_*}{\beta} \right) \left( \frac{\beta}{H_*} \right)
\left( \frac{T_*}{100 \, \mathrm{GeV}} \right) \left( \frac{g_*}{100}
\right)^{\frac{1}{6}}
$$
The wall velocity dependence of $f_\text{env}$ is
$$\frac{f_*}{\beta} = \frac{0.35}{1 + 0.069
v_\mathrm{w} + 0.69 v_\mathrm{w}^4}. $$
The amplitude is given by
$$h^2 \Omega_\text{sw}(f) =
8.5 \times 10^{-6} \left(\frac{100}{g_*} \right)^{\frac{1}{3}}
\Gamma^2 \overline{U}_\mathrm{f}^4 \left(\frac{H}{\beta}\right)
v_\mathrm{w} \, S_\text{sw}(f)$$
where $\Gamma =
\overline{w}/\overline{\epsilon} \approx 4/3$; $\overline{w}$ and
$\overline{\epsilon}$ are the volume-averaged enthalpy and energy
density
$\overline{U}_\mathrm{f}$ is a
measure of the rms fluid velocity
$$\overline{U}_\mathrm{f}^2 = \frac{1}{\overline w} \frac{1}{\mathcal{V}} \int_\mathcal{V}
d^3 x \, \tau_{ii}^\mathrm{f} \approx \frac{3}{4} \kappa_\mathrm{f} \alpha_{T_*}$$
Here $\kappa_\text{turb}$ is the efficiency of conversion
of latent heat into turbulent flows. On short
timescales it is very small (a few percent at
most).
Shocks and
turbulence develop on timescale:
$ \tau_\text{sh} \sim \mathrm{L}_f /
\overline{U}_\mathrm{f}. $
Turbulence: spectral shape
Although the amplitude is uncertain and will have to
wait for future simulations, the peak frequency is known
exactly,
$$ S_\text{turb}(f) =
\frac{(f/f_\text{turb})^3}{[1+(f/f_\text{turb})]^{\frac{11}{3}}} ( 1 + 8 \pi f/h_*).
$$
Here $h_*$ is the Hubble rate at $T_*$:
$$ h_* = 16.5 \, \mu\mathrm{Hz} \left( \frac{T_*}{100 \, \mathrm{GeV}
}\right) \left(\frac{g_*}{100} \right)^{\frac{1}{6}}$$
Turbulence: peak frequency
The peak frequency $f_\text{turb}$ is slightly higher than for the
sound wave contribution,
$$
f_\text{turb} = 27 \, \mu\mathrm{Hz} \,
\frac{1}{v_\mathrm{w}} \left( \frac{\beta}{H_*} \right)
\left(\frac{T_*}{100 \, \mathrm{GeV}} \right)
\left(\frac{g_*}{100}\right)^{\frac{1}{6}}.
$$