Introduction

David J. Weir, University of Helsinki

Tentative outline of these lectures:

  1. Today: introduction, key equations and techniques
  2. Tomorrow and Sunday: mostly first order phase transitions

About me

  • David Weir, email david.weir@helsinki.fi
  • Postdoc working on numerical simulations of phase transitions in the early universe
  • Previously:
    • 2014-16 Norway (Stavanger)
    • 2011-14 Helsinki (Finland)
    • 2007-11 Imperial College London (UK) [PhD]

About you ...?

  • Are you PhD students, postdocs, staff?
  • Have you studied GR, gravitational waves, cosmology, particle physics?
  • Are you working in this field?
  • Please feel free to ask questions!
  • And if you have requests, let me know.
Source: NASA
Source: GWplotter

LIGO

Source: (CC-BY) Andrea Nguyen on Flickr

LIGO at the Hanford Site

Source: (CC-BY-NC-ND) Prachatai

About LIGO

  • 2 sites: Livingston, LA; Hanford Site, WA
  • Cost: about a billion USD (most expensive/ambitious project ever funded by NSF)
  • Each site: Michelson interferometers with 4km arms, $1064 \, \mathrm{nm}$ Nd:YAG laser
  • Each arm: Fabry-Pérot cavity (increases path length to equivalent of 280 trips)
  • When a GW passes through: arms detune, photons emitted - signal
  • Sister project in Europe: VIRGO
LIGO design

First direct detection: GW150914

LIGO noise sources
Source: LIGO via Optics and Photonics news

LISA (and LISA pathfinder)

To look at longer wavelengths, need to go into space!


Sources: LISA; BBC

.

LISA Pathfinder

Exceeded design expectations by factor of five!

LISA mission profile

  • LISA: three arms (six laser links), 2.5 M km separation
  • Launch as ESA’s third large-scale mission (L3) in (or before) 2034
  • Proposal officially submitted earlier this year 1702.00786

Source: LISA.

Possible signals

Source: LISA proposal.

Mission requirements

Source: LISA proposal.

From the proposal:

Source: LISA proposal.

Pulsar timing arrays

Source: David Champion via MPIfR

About pulsar timing arrays

  • Millisecond pulsars (MSPs) are very accurate clocks
  • PTA projects time tens of MSPs every week or so
  • Irregularities in the period of one pulsar might be due to a change in the proper distance to that pulsar because of a passing GW (or a similar effect at Earth).
  • Measure correlations between multiple pulsars
    • Disentangle GWs at the pulsar end and at the Earth end
    • Remove effects of 'physics' in the neutron stars
  • Study stochastic background at very low frequencies

More about pulsar timing arrays

  • They also set the some of strongest constraints on the equation of state of dense nuclear matter.
  • Because MSPs are more accurate than atomic clocks, PTAs set the strongest constraints on solar system ephemerides
  • Some of the biggest single users of radio telescope time
Source: NSF via Wikipedia

Cosmological sources of gravitational waves

Electroweak phase transition

  • This is the process by which the Higgs became massive
  • In the minimal Standard Model it is gentle (crossover)
  • It is possible (and theoretically attractive) in extensions that it would experience a first order phase transition


EWPT physics

  • Bubbles produced at a thermal phase transition will be the central focus of these lectures
  • This would be a significant source of gravitational waves

Source: 1504.03291

Where to see it

  • The power spectrum is peaked at around the bubble radius, a fraction of the Hubble radius at the time of the transition.
  • That corresponds to millihertz today, which means it is ideally placed for space-based detectors like LISA
Source: 1512.06239

The end of inflation

  • As the inflaton reheated the universe, it created a lot of particles through violent processes.
  • The inflation oscillating about the bottom of its potential would excite other particles to oscillate with characteristic frequencies given by their masses.
  • These would be an efficient source of gravitational waves, but at high frequencies given by the mass of the field.

Source: 1506.06895.

Where to see it?

  • Some scenarios are potentially observable at earth-based detectors such as advanced LIGO
  • Otherwise, must wait for the more ambitious (post-LISA) space-based detectors (DECIGO, BBO, ...)

Figure source: Garcia-Bellido, Figueroa and Sastre

Cosmic strings

  • Extended, long lived, very tense pieces of 'string' formed during symmetry breaking in the early universe
  • They produce a scaling network, and lose energy principally by gravitational radiation
  • They can also radiate particles, but exactly how much is an open research question!
  • Loops of string are long-lived, emitting GWs over many wavelengths as they get smaller with time.

Topological stability

In field theory, cosmic strings are topologically stable:


Picture source: Ringeval

Can also be modelled by thin Nambu-Goto strings

Cosmic strings movie 1

Credit: David Daverio

Where to see them?

  • Pulsar timing arrays already place strong constraints on the energy scale on the energy of the strings.
  • Improved PTA measurements will constrain $G\mu$ further, meaning the energy scale is lower

Picture source: Binétruy, Bohé, Caprini and Dufaux

Gravitational waves recap

Further reading:

  • Review articles: Buonnano [follow this to some extent]
  • Textbooks: Hartle; Schutz; Weinberg

Introduction

  • The basic tools of general relativity are central ideas from differential geometry:
    • The metric, $g_{\mu\nu}$
    • The connections (also known as Christoffel symbols) $\Gamma^{\mu}_{\nu\rho}$
    • The Riemann tensor (or curvature tensor) $R_{\mu\nu\rho\sigma}$

The Einstein equations

  • The Einstein equations are written $$ R_{\mu\nu} - \frac{1}{2} g_{\mu\nu} R = 8 \pi G T_{\mu\nu} $$ where
    • The Ricci scalar $R = g^{\mu\nu} R_{\mu\nu}$
    • The Ricci tensor $R_{\mu\nu} = g^{\rho\sigma} R_{\rho\mu\sigma\nu}$
    • The stress-energy tensor is $T_{\mu\nu}$
  • 16 equations as written, 6 independent equations
    • symmetry: 6 constraints
    • Bianchi identity: 4 constraints
  • A gauge theory where the symmetries are diffeomorphisms (carefully constructed coordinate transformations)

Stress-energy tensor

Note that $T_{\mu\nu}$ looks schematically like $$ T_{\mu\nu} = \left( \begin{array}{c|c} \text{[energy density]} & \text{[energy flux]} \\ \hline & \\ \text{[momentum density]} & \text{[pressure and stress]} \\ & \\ \end{array} \right) $$

Linearised gravity

  • As we said, gravity is a gauge theory where the symmetry is diffeomorphism invariance.
  • This means it is invariant under coordinate transformations $$ x^{\mu} \to x'^{\mu} (x) $$ which are invertible, and differentiable.
  • Under a coordinate transformation the metric transforms as $$ g_{\mu\nu}(x) = g'_{\mu\nu}(x') = \frac{\partial x^\rho}{\partial x'^\mu} \frac{\partial x^\sigma}{\partial x'^\nu} g_{\rho\sigma}.$$

Specialising to linear perturbations

  • We now assume that we can write the metric as $$ g_{\mu\nu} = \eta_{\mu\nu} + h_{\mu\nu} \qquad \left| h_{\mu\nu} \right| \ll 1 $$ where $\eta_{\mu\nu}$ is the Minkowski metric $$\mathrm{diag}(-1,+1,+1,+1)$$
  • Even with this assumption, there is an invariance under transformations of the form $$ x^\mu \to x'^\mu = x^\mu + \xi^\mu(x) $$ provided that $|\partial_\mu \xi_\nu | \leq |h_{\mu\nu}|$

Wave equation

  • By substituting $$ g_{\mu\nu} = \eta_{\mu\nu} + h_{\mu\nu}$$ in the Einstein equations $$ R_{\mu\nu} - \frac{1}{2} g_{\mu\nu} R = 8\pi G T_{\mu\nu} $$ and retaining only terms at linear order we will obtain a wave-like equation for $h_{\mu\nu}$.

The wave equation, simplified

  • We can write it in a fairly clear manner by adopting the trace-reversed metric perturbation $$ \bar{h}^{\mu\nu} = h^{\mu\nu} - \frac{1}{2}\eta^{\mu\nu} h $$
  • It then takes the form $$ \begin{multline} \square \bar{h}_{\nu\sigma} + \eta_{\nu\sigma} \partial^\rho \partial^\lambda \bar{h}_{\rho\lambda} - \partial^\rho \partial_\nu \bar{h}_{\rho \sigma} - \partial^\rho \partial_\sigma \bar{h}_{\rho\nu} + \mathcal{O}(h^2) \\ = - \frac{16 \pi G}{c^4} T_{\nu\sigma} \end{multline} $$

Lorenz gauge

  • Now we impose Lorenz gauge $$ \partial_\nu \bar{h}^{\mu\nu} = 0 $$
  • This is equivalent to Lorenz gauge in electromagnetism: $$ \partial_\mu A^\mu = 0 $$
  • The effect is to cancel every term except $$ \square \bar{h}_{\nu\sigma} = - 16 \pi G T_{\nu\sigma} $$ which indeed looks like a wave equation!

Summary (comparison to electromagnetism)

Source: le Tiec and Novak

Transverse traceless gauge

  • There are still 6 components in $\bar{h}_{\mu\nu}$, only 2 of which are physical - the two polarisations
  • To find these, we consider coordinate transformations that also satisfy Lorenz gauge, meaning $$ x^\mu \to x'^\mu = x^\mu + \xi^\mu(x); \quad |\partial_\mu \xi_\nu | \leq |h_{\mu\nu}| $$ and $\square \xi^\mu = 0 $.
  • $\square \xi^\mu = 0$ gives 4 new constraints. First we choose: $$\bar{h} = 0$$ i.e. make $\bar{h}^{\mu\nu}$ traceless [one constraint]
  • (Note that with this constraint, the trace of $h^{\mu\nu}$ also disappears, so we no longer need the 'trace reversed' tensor!)

More constraints

  • We choose our other 3 constraints to be: $$ h^{i0} = 0 $$ meaning $ \partial_0 h^{00} = 0 $ too, and we can choose $h^{00} = 0$.
  • Put another way, our 4 constraints are: $$h^{00} = h^{0i} = 0$$
  • The remaining spatial entries of $h_{ij}$ must be transverse $$ \partial_{i} h^{ij} = 0 $$ and traceless $$ h^{ii} = 0. $$

Travelling waves

  • A plane wave in the $z$-direction looks like $$ h_{ij}^\text{TT} (t, z) = \left( \begin{array}{ccc} h_+ & h_\times & 0 \\ h_\times & h_+ & 0 \\ 0 & 0 & 0 \\ \end{array} \right)\cos\, \omega \left(t - \frac{z}{c} \right) $$ Note that the polarisation components are perpendicular to the direction of travel.
  • For a more general travelling wave with wave vector $\mathbf{k}$, transverse traceless simplifies to $$ \hat{\mathbf{k}}^i h_{ij}^\text{TT} = 0 .$$

Projection

  • If we have a general metric perturbation, we can use a tensor version of the usual 'transverse' projector: $$ \Lambda_{ij,lm} \equiv P_{il} P_{jm} - \frac{1}{2} P_{ij} P_{lm} $$
  • Here we have $$ P_{ij} = \delta_{ij} - \hat{\mathbf{k}}_i \hat{\mathbf{k}}_j $$
  • And this satisfies $$ P_{ij} = P_{ji}, \quad \hat{\mathbf{k}}^i P_{ij} = 0, \quad P_{ij} P^{jk} = P_i^k, \quad P_{ii} = 2 $$
  • In other words, it projects out the rotational part of a vector field.

Projection of a general $h_{ij}$

  • If we have a general $h_{ij}$ which is in Lorenz gauge but does not otherwise satisfy the transverse traceless (TT) requirements, we can project out the TT parts: $$h_{ij}^\text{TT} = \Lambda_{ij,lm} h^{lm} $$
  • Why do we want to do this?
    • We may want to study the polarisation of the gravitational waves (but for stochastic, cosmological sources, these rarely matter)
    • More importantly, as $h_{ij}^\text{TT}$ contains only the propagating degrees of freedom, measures of energy and power must be done with it.

Effective stress-energy tensor

  • Step back and consider a more general background: $$ g_{\mu\nu} = \bar{g}_{\mu\nu} + h_{\mu\nu}, \quad |h_{\mu\nu}| \ll 1 $$
  • What is the background and what is the perturbation?
    • $\bar{g}_{\mu\nu}$ has a scale $L_\mathrm{B}$
    • $h_{\mu\nu}$ have a wavelength $\lambda \ll L_\mathrm{B}$.
  • Or, in frequency:
    • $\bar{g}_{\mu\nu}$ only has frequencies up to $f_\mathrm{B}$
    • $h_{\mu\nu}$ has frequencies $f \gg f_\mathrm{B}$.
  • The overall effect is that, even if $\bar{g}_{\mu\nu}$ has come curvature, it looks slowly varying to the gravitational waves.

The Isaacson argument

  • We can now split the Ricci tensor up as follows: $$ R_{\mu\nu} = \bar{R}_{\mu\nu} + R^{(1)}_{\mu\nu} + R^{(2)}_{\mu\nu} + \ldots $$
    • $\bar{R}_{\mu\nu}$ is the background Ricci tensor
    • $ R^{(1)}_{\mu\nu}$ are high frequency modes at linear order in $h$
    • $ R^{(2)}_{\mu\nu}$ is everything at quadratic order in $h$
  • If we average over a volume bigger than $\lambda$ but smaller than $L_\mathrm{B}$ then we can use the Einstein equations to introduce $ t_{\mu\nu} = -\frac{1}{8\pi G} \left< R_{\mu\nu}^{(2)} - \frac{1}{2} \bar{g}_{\mu\nu} R^{(2)} \right> $
  • And $ \bar{R}_{\mu\nu} - \frac{1}{2}\bar{g}_{\mu\nu} \bar{R} = 8\pi G \left( \bar{T}_{\mu\nu} + t_{\mu\nu} \right).$

The Isaacson expression

  • We end up with (in arbitrary gauge) $$ \begin{multline}t_{\alpha\beta} = \frac{1}{32\pi G} \left< \partial_\alpha \bar{h}_{\mu\nu} \partial_\beta\bar{h}^{\mu\nu} - \frac{1}{2} \partial_\alpha \bar{h} \partial_\beta \bar{h} \right. \\ \left. - \partial_\nu \bar{h}^{\mu\nu} \partial_\beta \bar{h}_{\mu\alpha} - \partial_\nu\bar{h}^{\mu\nu} \partial_\alpha \bar{h}_{\mu\beta} \right> \end{multline}$$ which, for TT, reduces to $$ t_{\alpha\beta} = \frac{1}{32 \pi G} \left< \partial_\alpha h^{\text{TT}}_{\mu\nu} \partial_\beta h_{\text{TT}}^{\mu\nu} \right> $$ which is known as the Isaacson tensor.

Energy density in gravitational waves

  • Now that we have the effective stress-energy tensor, the energy density in gravitational waves is $$ \rho_\text{GW} \equiv t_{00} = \frac{1}{32 \pi G} \langle \dot{h}_{ij}^\text{TT} \dot{h}_{ij}^\text{TT} \rangle $$
  • By Fourier transforming this expression, define the GW power per logarithmic frequency interval $$ \frac{\mathrm{d} \, \rho_\text{GW}(\mathbf{k})}{\mathrm{d} \, \log k } = \frac{1}{32\pi G} \frac{k^3}{2\pi^2} \left< \dot{h}_{ij}^\text{TT}(\mathbf{k}) \dot{h}_{ij}^\text{TT} (-\mathbf{k}) \right> $$
  • These two quantities are what cosmologists most widely quote in papers about gravitational waves, particularly the results of numerical simulations.

Typical assumptions in these lectures

  • Minkowski or FRW spacetime: no, or isotropic expansion
    • Physics on timescales much shorter than expansion
    • All gravitational waves sourced by sub-Horizon physics
  • Homogeneous, stochastic, isotropic source
    • True for most cosmological sources
    • May not be true for, e.g. cosmic string cusps

Results in context

How does this work in a cosmological simulation?

  1. Evolve Lorenz-gauge wave equation in position space $$ \nabla^2 h_{ij} (\mathbf{x},t) - \frac{\partial}{\partial t^2} h_{ij}(\mathbf{x},t) = 8 \pi G T_{ij}^\text{source}(\mathbf{x},t)$$ during simulation, using relevant $T^\text{source}_{ij}$ of 'source system'.
  2. Projection to TT-gauge requires expensive Fourier transform, so only project when measurement desired: $$ h^{\text{TT}}_{ij}(\mathbf{k},t_\text{meas}) = \Lambda_{ij,lm}(\hat{\mathbf{k}}) h^{lm}(\mathbf{k},t) $$
  3. Measure energy density (or power) in gravitational waves $$ \rho_\text{GW}(t_\text{meas}) = \frac{1}{32 \pi G} \left< \dot{h}_{ij}^\text{TT} \dot{h}_{ij}^\text{TT} \right> $$
  4. Redshift to present day.

Other techniques

  • Two other approaches are sometimes seen in numerical cosmology:
    • Quadrupole approximation - as we will see, this is a poor approximation for bubble collisions we will be studying, but it still provides insight
    • "Weinberg formula" - this gives a clean time-domain formula where the stress-energy tensor takes simple forms
  • We will look at these, and the properties of general sources, next.

Compact, distant sources

  • This is also relevant, e.g. for colliding pairs of bubbles.
  • Start from the Lorentz-gauge wave equation $$ \square \bar{h}_{\mu\nu} = - 16 \pi G T_{\mu\nu}. $$
  • Solve in position space with retarded Green's functions $$ \bar{h}_{\mu\nu}(x) = - 16 \pi G \int \mathrm{d}^4 x' \, G(x-x') T_{\mu\nu}(x'). $$
  • If we now specialise to TT gauge and write in terms of the retarded time $t - |\mathbf{x} - \mathbf{x}'|$, $$ \begin{multline} h_{ij}^\text{TT} (t, \mathbf{x}) = \Lambda_{ij,lm} (\hat{\mathbf{n}})\, 4 G \int d^3 x' \frac{1}{|\mathbf{x} - \mathbf{x}'|} \\ \times T_{lm} \left( t- |\mathbf{x} - \mathbf{x}|; \mathbf{x}'\right) \end{multline} $$

Far field approximation

  • And if we are also far from the source (where $T_{lm}(t, \mathbf{x}) \neq 0$), $$ |\mathbf{x} - \mathbf{x}'| \approx r - \mathbf{x}' \cdot \hat{\mathbf{n}} $$ and $$ \begin{multline} h_{ij}^\text{TT} (t, \mathbf{x}) \approx \frac{1}{r} \, 4G \, \Lambda_{ij,lm} (\hat{\mathbf{n}}) \int_{\text{source}} d^3 x' \frac{1}{|\mathbf{x} - \mathbf{x}'|} \\ \times T_{lm} \left( t- r + \mathbf{x}'\cdot \hat{\mathbf{n}}; \mathbf{x}'\right) \end{multline}$$
  • We will assume(!) that velocities inside the source are non-relativistic
  • In other words gravitational waves have lower frequencies $\omega$ than the source diameter: $$ \omega \mathbf{x}' \cdot \hat{\mathbf{n}} \ll 1 $$

Multipole expansion

  • We can write the source using a Fourier transform $$ \begin{multline} T_{lm} \left(t - r + \mathbf{x}'\cdot \hat{\mathbf{n}}; \mathbf{x}'\right) \\ = \int \frac{\mathrm{d}^4 k}{(2\pi)^4} T_{lm}(\omega, \mathbf{k}) e^{-i \omega \left( t- r + \mathbf{x}' \cdot \hat{\mathbf{n}} \right) + i \mathbf{k} \cdot \mathbf{x}' } \end{multline} $$ and the leading order term expanding in $\omega \mathbf{x}' \cdot \hat{\mathbf{n}}$ is $$ T_{lm} \left(t - r + \mathbf{x}'\cdot \hat{\mathbf{n}}; \mathbf{x}'\right) \approx \int \frac{\mathrm{d}^4 k}{(2\pi)^4} T_{lm}(\omega,\mathbf{k}) $$ This is the quadrupole term.

Quadrupole source

  • To leading order, then, the metric perturbation is $$ h_{ij}^\text{TT}(t,\mathbf{x}) \approx \frac{1}{r} \, 4G \, \Lambda_{ij,lm}(\hat{\mathbf{n}}) \int d^3 x \, T^{lm} \left(t - r, \mathbf{x} \right) + \ldots $$
  • This is zero for a spherically symmetric source (or linear superposition of spherically symmetric sources)
    • Vacuum fluctuations at the end of inflation
    • Freshly nucleated bubbles in the early universe
    • Isolated massive objects
  • Need some non-spherical dynamics:
    • Particle resonances
    • Bubbles colliding
    • Binary compact massive objects

Why we must go beyond the quadrupole approximation

  • In the early universe, velocities within the source are not small, and the gravitational waves are typically the same scale as the bubbles.

Weinberg formula

  • So far, we have seen two methods of computing $h_{ij}$
    • Numerically solving the equation of motion $$ \nabla^2 h_{ij} - \frac{\partial}{\partial t^2} h_{ij}= 8\pi G T_{ij} $$ e.g. during a simulation
    • Using the quadrupole approximation if the wavelength of the gravitational waves is long compared to the size of the source(s)
  • However, sometimes can simplify the source so that it is simple in Fourier space, and we do not need to do the quadrupole approximation
  • Then we can use the Weinberg formula

Using the Weinberg formula

  • For radiation in a direction $\hat{\mathbf{k}}$ and frequency $\omega$, the power spectrum per logarithmic frequency interval, per unit solid angle, $$ \frac{\mathrm{d}\rho_\text{GW}}{\mathrm{d} \log \omega \, \mathrm{d} \Omega} = 2 G \omega^3 \Lambda_{ij,lm}(\hat{\mathbf{k}}) T_{ij}^* (\hat{\mathbf{k}},\omega) T_{lm} (\hat{\mathbf{k}},\omega) $$
  • One application of this is the 'envelope approximation', which we shall revisit later.

Useful insight 1

Source: Dufaux, Felder, Kofman, Navros

  • Begin with the momentum-space Green's function expression (assume source off at $t' < 0$) $$ h_{ij}^\text{TT}(\mathbf{k},t) = 16\pi G \, \Lambda_{ij,lm} \int_0^t \, \mathrm{d} t' \frac{\sin[k(t-t')]}{k} T_{lm}(\mathbf{k},t') $$
  • If the source is slowly varying in space at low $\mathbf{k}$: $$T_{lm}(\mathbf{k}) \to \text{const.}; \qquad k \ll k_\text{max} $$ (equivalent to the quadrupole approximation) we get $$ h_{ij}^\text{TT}(\mathbf{k},t) \approx 16\pi G \, \Lambda_{ij,lm} \int_0^t \, \mathrm{d} t' \frac{\sin[k(t-t')]}{k} T_{lm}(0,t') $$

Useful insight 1

  • If the source $T_{lm}(0,t)$ varies faster than the $\sin[k(t-t')]$, the equation reduces to $$ h_{ij}^\text{TT}(\mathbf{k},t) \approx 16\pi G \, \Lambda_{ij,lm} \int_0^t \, \mathrm{d} t' T_{lm}(0,t') $$ This gives $$ \frac{\mathrm{d} \rho_\text{GW}(k)}{\mathrm{d} \, \log \, k} \propto k^3 $$
  • In other words, at sufficiently long sub-horizon scales, the quadrupole approximation always works and all the matters is how long the source is on for.
  • The power law is $k^3$.
  • This holds for e.g. first order phase transitions and the end of inflation.

Useful insight 2

  • If the source has an intermediate regime where $T_{lm}(0,t)$ varies slower than $\sin[k(t-t')]$, then the source stays in the integral, and there is an additional $1/k$ factor
  • Therefore, in some cases we can expect a $k^1$ power law at higher wavenumbers than the $k^3$ is valid
  • This is less generally true than the previous $k^3$ regime, so it might not be observed at all.
  • In general, though, where a power law is seen in simulation results, it is worth seeing if the underlying physics is amenable to simpliciation!
  • We will encounter more power laws later in these lectures...

Useful insight: graphical summary

Conclusions

  • With pulsar timing arrays, space- and earth-based detectors, we now (or very soon) will view the gravitational wave sky from nanohertz through to kilohertz
  • The basic equations of gravitational radiation share a lot of features with electromagnetism (or other gauge theories)
  • There are some useful regimes that one can explore with only very limited knowledge of the form of a source of gravitational waves.