Kubiznak 3 Lectures On Variational Principles
Kubiznak 3 Lectures On Variational Principles
Kubiznak 3 Lectures On Variational Principles
principles
David Kubizňák
Institute of Theoretical Physics
Contents i
Bibliography 24
i
Chapter 1: Lecture 1: Variational
principles in GR
where Lm is the scalar Lagrangian density and φ stands for various fields. The
variation gives
√ ∂Lm
Z Z
d δSm ∂Lm
δSm = d x δφ = dd x −g δφ + δ∇µ φ . (1.3)
δφ ∂φ ∂∇µ φ
As always, we now
R interchange the Rcovariant derivative with δ, and, using the
µ
√ √
Stokes theorem, Σ ∇µ V −gd x = ∂Σ γV µ dΣµ , integrate by parts as follows:
d
√ √
Z Z
A (∇µ B) −gd x = − B(∇µ Aµ ) −gdd x + boundary term ,
µ d
(1.4)
to have
√ h ∂Lm
Z ∂L i
m
δSm = dd x −g − ∇µ δφ . (1.5)
∂φ ∂(∇µ φ)
Thus we have derived the generalized Euler–Lagrange field equations:
δSm ∂Lm ∂L
m
=0 ⇔ − ∇µ = 0. (1.6)
δφ ∂φ ∂(∇µ φ)
1
CHAPTER 1. LECTURE 1: VARIATIONAL PRINCIPLES IN GR 2
√
Z
1
δg Sm = − dd x −gTµν δg µν . (1.7)
2
We call the object Tµν the (Rosenfeld’s) energy momentum tensor. It is symmetric
by construction.
For diffeomorphism invariant Lagrangian densities Lm , such an energy momentum
is conserved in the following sense: If the equations of motion for matter are
satisfied, the we have
∇µ T µν = 0 . (1.8)
(In fact, as we shall see, the conservation of energy–momentum is in many cases
equivalent to the equations of motion for the matter.)
The argument for this goes as follows. First, let us remind that an infininitesimal
diffeomorphism, generated by a vector field ξ,
xµ → x µ − ξ µ , (1.9)
induces the following variation of the fields and the components of the metric:
δφ = Lξ φ , δg µν = Lξ g µν = 2∇(µ ξ ν) . (1.10)
δSm
∇µ T µν = 0 ⇔ Lξ φ = 0 , (1.13)
δφ
CHAPTER 1. LECTURE 1: VARIATIONAL PRINCIPLES IN GR 3
√ 1√
δ −g = − −ggµν δg µν , δgµν = −gµρ gνσ δg ρσ . (1.14)
2
Example 1: Scalar field. Using the minimal coupling principle, the Lagrangian
reads
1
L = − g µν (∇µ φ)(∇ν φ) − V (φ) . (1.15)
2
Using (1.6), the corresponding field equation is
dV
∇µ ∇µ φ − = 0. (1.16)
dφ
To calculate the energy momentum tensor, we simply vary the corresponding
action:
Z h √ √ i 1
Z
√ h i
d
δg S = d Lδ −g + −gδL = − dd x −g gµν L + ∇µ φ∇ν φ δg µν (. 1.17)
2 | {z }
Tµν
In the flat space limit, this is the canonical energy momentum tensor for the scalar
field derived from the Noether’s procedure due spacetime translation invariance.
Let us next look at what imposing ∇µ T µν = 0 yields. We have
1 µν 1 µα νβ
L[Aµ , g αβ ] = − F Fµν = − g g Fµν Fαβ , Fµν = 2∇[µ Aν] . (1.20)
16π 16π
It gives rise to the Maxwell equations in curved spacetime:
∇µ F µν = 0 , dF = 0 , (1.21)
CHAPTER 1. LECTURE 1: VARIATIONAL PRINCIPLES IN GR 4
where the latter is automatically satisfied from the definition of the field strength.
Varying the action w.r.t. the metric, we have
√ √
Z
1 h i
δg S = − dd x F 2 δ −g + 2 −gFµδ Fν δ δg µν . (1.22)
16π
This immediately yields the (automatically symmetric and gauge invariant) energy–
momentum tensor:
1 α 1 2
Tµν = Fµα Fν − gµν F . (1.23)
4π 4
It is this electromagnetic energy momentum tensor that couples to gravity (not
the canonical one, which is not symmetric and not gauge invariant). Moreover,
we find
1
4π∇µ T µ ν = ∇µ F µα Fνα + F µα ∇µ Fνα − ∇ν Fαβ F αβ = ∇µ F µα Fνα , (1.24)
| {z } 2
1 βα
2
F (∇β Fνα +∇α Fβν )
where the last three terms vanished upon using the second set of Maxwell’s equa-
tions.1 Obviously, if Fνα has an inverse, then conservation of the energy momen-
tum tensor implies the first set of Maxwell’s equations.
Variation. Varying this action, and using the dirty trick that R = Rµν g µν , we get
√ √ √ √
Z Z
1 αβ 1
δSEH = δ( −gRαβ g ) = (Rδ −g+ −gRµν δg µν + −gg µν δRµν ) .
16πG 16πG
(1.27)
1
We could have derived this directly using the result (1.13). Namely, using the Cartan’s lemma
Lξ ω = ξ · (dω) + d(ξ · ω) , valid for any p-form ω, we have
δSm √ √ √
Lξ Aµ ∝ −g∇ν F νµ Lξ Aµ = −g∇ν F νµ (dA)µα ξ α + −g∇ν F νµ ∇µ (ξ · A) . (1.25)
δAµ
However, by integrating the second term by parts (we are under an integral really), we turn it into
∇µ ∇ν F µν (ξ · A) = 0. Thus we arrive at the same conclusion as above.
CHAPTER 1. LECTURE 1: VARIATIONAL PRINCIPLES IN GR 5
To prove this lemma, we shall use ‘local inertial frame’ equipped with Riemann normal coordinates,
that is Γµ νλ = 0 at a point, but ∂Γ non-vanishing (gµν,λ = 0 but gµν,λδ is not) together with the fact
that if the tensor identity is proved in one frame, it is valid in any frame. So,
.
δRαβ = δRγ αγβ = δ[Γγ αβ,γ − Γγ αγ,β + ΓΓ − ΓΓ] = δ[Γγ αβ,γ − Γγ αγ,β ]
= ∇γ (δΓ)γ αβ − ∇β (δΓ)γ αγ = δΓγ αβ;γ − δΓγ αγ;β , (1.30)
where we have used the fact that in normal coordinates ∂ and ∇ are the same and that δΓ is a tensor.
We also have
1 µσ 1 µσ
δΓµ νλ = g (δgσν,λ + δgσλ,ν − δgνλ,σ ) + δg (gσν,λ + gσλ,ν − gνλ,σ )
2 2
1 µσ 1 µσ
= g (δgσν;λ + δgσλ;ν − δgνλ;σ ) + δg (gσν;λ + gσλ;ν − gνλ;σ )
2 2
1 µσ
= g (δgσν;λ + δgσλ;ν − δgνλ;σ ) , (1.31)
2
which also yields
1 ασ 1
δΓα αβ = g (∇β δgσα + ∇α δgσβ − ∇σ δgαβ ) = g ασ ∇β δgασ . (1.32)
2 2
Hence we found
1
∇µ ∇λ δgλν + ∇ν ∇λ δgλµ − ∇2 δgµν − ∇µ ∇ν (g λσ δgλσ ) .
δRµν = (1.33)
2
Everywhere above “δg” means variation of gµν . We also need variation w.r.t. g µν , which is given by
δg µν = −g σµ g λν δgλσ . (1.34)
So we have
where
δg −1 = gµν δg µν , (1.36)
Note also a useful identity
1
δΓα αβ = − ∇β δg −1 . (1.37)
2
CHAPTER 1. LECTURE 1: VARIATIONAL PRINCIPLES IN GR 6
which does not contribute to the equations of motion. Thus we have the following
Einstein equations in the absence of matter:
Gµν = 0 . (1.39)
∇µ Gµν = 0 . (1.40)
The same conclusion will be true for ‘generalized Einstein tensor’ of any
diffeomorphism invariant (higher-curvature) theory of gravity.
– Remark 2. The obtained equation is second order PDE for the metric, Gµν =
Gµν (g, ∂g, ∂ 2 g). How is this possible? We started from the action, given by
R, which already is of second order in derivatives. We should thus have
received 4th order equations of motion. However, this is not the case and
the reason is simple. It can be shown that the Lagrangian density can be
split to a piece that depends only on the first derivatives, plus a piece that
is a total derivative (none of which is a tensor):
√ √
−gR(g, ∂g, ∂ 2 g) = −g R̃(g, ∂g) + ∂µ R̂µ (g, ∂g) , (1.41)
√
where δ R̂µ = −gV µ .3 The latter term does not contribute to the equa-
tions of motion, provided we impose the corresponding boundary conditions.
These, however, do not give rise to the standard Dirichlet problem. To
achieve that, one needs to add the so called York–Gibbons–Hawking term,
as we shall see in the next lecture.
– Remark 3. In the above variation we used the so called second-order formalism:
the action was treated as a function of the metric gµν and contained its first
and second derivatives.
3
In fact, one has
√
R̃ = g ικ (Γλ µκ Γµ λι − Γλ ικ Γµ λµ ) , R̂µ = −g(g ικ Γµ ικ − g ιµ Γκ ικ ) . (1.42)
Moreover, there is a ‘holographic relation’ between the bulk and surface part of the Lagrangian, with
the surface terms obtainable from the bulk on by differentiation, [2].
CHAPTER 1. LECTURE 1: VARIATIONAL PRINCIPLES IN GR 7
Perhaps more useful is the first-order (Palatini) formalism where the action
is treated as action for two fields: the metric gµν and the connection Γα βγ :
√
Z
1
SPalatini [g, Γ] = −gg µν Rµν (Γ) , (1.43)
16πG
where we used the fact that Rµν can be entirely written in terms of the
connection and its derivatives, Rµν = Rµν (Γ). Thus the variation w.r.t.
the metric yields immediately the Einstein equations (1.39). It can then be
shown that the variation w.r.t. the connection establishes that the connec-
tion is given by Christoffel symbols (this was an input in the second order
formalism.)
Note that in this spirit, one can also write down the purely connection
dependent action for gravity with cosmological constant Λ, known as the
Eddington’s action:
Z
1
q
4
SEddington [Γ] = d x − det Rµν (Γ) . (1.44)
8πGΛ
If you want to, please show the equivalence with the Einstein-Hilbert action
in the presence of Λ.
– Remark 4: Einstein equations. Recovering now the full Einstein equations is
easy. We simply add the corresponding matter Lagrangian density:
√
Z
S = SEH [g] + d4 x −gLm . (1.45)
The variation w.r.t. the metric and throwing away the boundary terms then
yields
√ 1
Z
4 µν 1 µν
δS = d x −g Gµν δg − Tµν δg (1.46)
16πG 2
Thus we recover the famous 1915 Einstein’s field equations:4
However, one can show that this is not the case, as we have the following [3]:
Lovelock theorem. In four dimensions, the Einstein–Hilbert action is the only local
action (apart from the cosmological constant and topological terms) that leads to
the second order differential equations for the metric.
In other words, in four dimensions Einstein’s theory is the unique theory we
can obtain from the action principle that yields 2nd-order EOM for the metric.
In higher dimensions this is no longer true – we have a possibility of the so
called Lovelock gravities. The corresponding Lagrangian is given in terms of the
Euler densities, a certain combination of curvature invariants constructed from
the powers of the Riemann tensor. They lead to the second order equations of
motion for the metric, naturally generalizing the Einstein equations to higher
dimensions.
Gauss–Bonnet gravity. To give and example, let us consider the so called 2nd-
order Lovelock gravity, also know as the Gauss–Bonnet gravity. The correspond-
ing Lagrangian reads:
√
Z
1
S=− dd x −g (R + αG) , G = R2 − 4Rαβ Rαβ + Rαβγδ Rαβγδ , (1.49)
16πGN
where α is a coupling constant with the dimension [α] = L2 . This extension of
General Relativity naturally appears in the low energy effective action of heterotic
string theory. It yields the following equations of motion:
ni (σ A ) = 0 i = 1, . . . , n , (2.2)
9
CHAPTER 2. LECTURE 2: SUBMANIFOLDS & YORK–GIBBONS–HAWKING
TERM 10
for all coordinate functions σ A on Σ:
We will take
ni · nj = ±δij , (2.3)
where − is for timelike n and + for spacelike n.
We now have two ways to describe things. i) spacetime point of view (in terms
of spacetime objects) and ii) submanifold point of view (in terms of objects on a
submanifold). Let us start with the first:
Spacetime point of view. In this description we define the first fundamental form,
or induced metric of Σ as
n
X
hµν = gµν + j njµ njν , (2.4)
j=1
where j = + for timelike and − for spacelike nj . hµν is the metric Σ inherits
from M (but lies in T ∗ M ⊗ T ∗ M ) At the same time
hµ ν (2.5)
K = g µν Kµν = ∇µ nµ . (2.7)
Indeed, we have
Manifold point of view. Since Σ is a manifold in its own right, we can also consider
quantities intrinsic to Σ. To this purpose, can think of Σ as a map Σ → M , given
by:
Σ : xµ (σ A ) . (2.9)
∂xµ
eµ A = : Tp∗ (M ) → Tp∗ (Σ) . (2.10)
∂σ A
For example,
ωµ → eµ A ωµ = ωA . (2.11)
In particular, we define the intrinsic (pull-back) metric (element of T ∗ Σ ⊗ T ∗ Σ):
∂xµ
eA · ni = niµ = 0, i = 1, . . . , n . (2.15)
∂σ A
Can also define the corresponding extrinsic curvature (from the intrinsic point of
view)
KAB = eµ A eν B ∇µ nν = −nµ DA eµ B , (2.16)
where D is the intrinsic covariant derivative (again Levi-Civita) inherited from
the covariant derivative ∇:
Dµ Vν = hµ λ hν σ ∇λ Vσ , (2.17)
(d) 0 0 0 0
Rα βγδ = (D) Rα β 0 γ 0 δ0 hα α0 hβ β hγ γ hδ δ ∓ (K α γ Kβδ − K α δ Kβγ ) , (2.25)
where
V = V µ nµ = ∇n δg −1 − nα ∇β δg αβ , δg −1 = gαβ δg αβ . (2.27)
The second term does not yield the standard Dirichlet boundary conditions. In
what follows we would like to eliminate it, by adding a proper boundary term to
the Einstein–Hilbert action, to get the Dirichlet variational principle for gravity.
CHAPTER 2. LECTURE 2: SUBMANIFOLDS & YORK–GIBBONS–HAWKING
TERM 14
To this purpose we use the Gauss–Codazzi formalism introduced above, thinking
of ∂M as Σ. That is, we consider co-dimension one Σ – a “wall”. To simplify
the calculations, we take ∂M to be a timelike boundary, that is nµ spacelike, and
(w.l.o.g.) extend nµ geodesically into the bulk. Thus we have
Consider next
1 1 1
= ∇α δnα − ∇n δg −1 = ∇α ( nβ δg αβ ) − ∇n δg −1
2 2 2
1 1
nβ ∇α δg αβ − ∇n δg −1 + ∇α nβ δg αβ .
= (2.30)
2 | {z } 2
−V
and so
V = −2δK + Kαβ δhαβ . (2.32)
1
Z √
d3 x h 2δK − Kµν δhµν
δS∂M =
16πG
1
Z
3
√ 1
Z √
= d xδ( hK) − d3 x h(Kµν − Khµν )δhµν . (2.33)
8πG 16πG
The first term is a (minus) variation of the famous York–Gibbons–Hawking bound-
ary term:
1
Z √
SGH = − d3 x hK , (2.34)
8πG
whose variation cancels the unwanted boundary term in (2.26).
1
Under construction – or up to you :)
CHAPTER 2. LECTURE 2: SUBMANIFOLDS & YORK–GIBBONS–HAWKING
TERM 15
To summarize, we should consider the following gravitational action:
1
Z
4 √ 1
Z √
Sg = SEH + SYGH =− d x gR − d3 x hK . (2.35)
16πG 8πG
The first terms gives the E-L equations plus the boundary derivative term. The
latter is cancelled by the second term. However, we still get the “Kµν − Khµν ”
term:
1
Z
4 √ µν 1
Z √
δSg = − d x gGµν δg − d3 x h(Kµν − Khµν )δhµν . (2.36)
16πG M 16πG
For fixed boundary and Dirichlet boundary conditions we have δhµν = 0, which
yields a well defined variational principle.
Tµν ∼ Sµν δ xµ − xµ (σ A ) .
(2.38)
where h · i stands for averaging of “ +00 +“−00 quantities, e.g. K(n+ ) + K(n− ).
These are the Israel junction conditions.
CHAPTER 2. LECTURE 2: SUBMANIFOLDS & YORK–GIBBONS–HAWKING
TERM 16
To derive these, Israel [8] took the limit δ → 0 of the Einstein equations, Gµν =
8πGTµν , in the following setup (note that in this case we have a continuous nor-
mal across Σ):
He then used these to study the gravitational collapse of thin matter shells.
Alternatively, one could consider a certain kind of the Neumann boundary con-
ditions, see e.g. [9].
Chapter 3: Lecture 3: Black hole
thermodynamics
dr2 2M Q2 r2 Q
ds2 = −f dt2 + + r2 dΩ2 , f =1− + 2 + 2, A = − dt , (3.1)
f r r l r
where dΩ2 = dθ2 +sin2 θdϕ2 . This is a solution of Einstein Maxwell-AdS equations
(EM )
Gµν = −Λgµν + 8πTµν , ∇µ F µν = 0 , F = dA , (3.2)
Basic properties. The above solution is characterized by its mass M and charge
Q. For given M, Q, l, its horizon is located at the horizon radius r+ , given by the
largest root of
f (r+ , M, Q, l) = 0 . (3.5)
It is a Killing horizon: a null surface generated by Killing field k = ∂t . Associated
with it is the concept of surface gravity κ:
17
CHAPTER 3. LECTURE 3: BLACK HOLE THERMODYNAMICS 18
We may also calculate the horizon area. Taking dt = 0 = dr, the induced spatial
metric ‘on the horizon’ is dγ 2 = r+
2
dΩ2 . The area then reads
Z p Z
2 2
A= det γdθdϕ = r+ sin θdθdϕ = 4πr+ . (3.7)
First law of black hole mechanics. Let’s now consider a physical process under
which the black hole spacetime is perturbed and eventually settles to a new
charged AdS black hole spacetimes with modified spacetime parameters, {M +
δM, Q + δQ, l + δl}. Consequently, the horizon radius also modifies to r+ + δr+ ,
determined from f (r+ + δr+ , M + δM, Q + δQ, l + δl) = 0. Using the Taylor
expansion to linear order in perturbation, together with (3.5), we thus have
∂f ∂f ∂f ∂f
0= δM + δr+ + δQ + δl . (3.8)
∂M ∂r ∂Q ∂l r=r+
where in the second line we used (3.6) and the specific form of f for the charged
AdS black hole. We thus find
κ δA Q 4 3
δM = + δQ + πr+ δP , (3.10)
2π 4 r+ 3
treating M = M (A, Q, P ). Let’s now make the following definitions of electrostatic potential
and black hole volume:
∂M Q ∂M 4 3
φ= = , V = = πr+ . (3.11)
∂Q A,P r+ ∂P A,Q 3
However, this looks a lot like a 1st law of (black hole) thermodynamics (especially
because of the work terms), provided we identify M with enthalpy and
~κ A
T = , S= . (3.13)
2πkB 4~GN
While this seems strange for classical black holes, as shown by Hawking in 1974,
[11, 12], when quantum effects are taken into account, black holes radiate away
as black body with these characteristics. Derivation used QFT in curved space.
Hawking basically showed “stimulated emission”. The problem with his deriva-
tion is that due to the bluehift near the horizon, the test field approximation
breaks down and we cannot really trust the result. However, since then the same
result has been reproduced by many other approaches, e.g: Euclidean path in-
tegral, tunneling, string theory, LQG. Let’s exploit our knowledge of variational
principles to ‘derive’ these results.
dr2
ds2 = f dτ 2 + + r2 dΩ2 . (3.15)
f
Near the horizon we may expand
Therefore, the near horizon limit of the Euclidean solution takes the following
form:
dr2
ds2 = 2κ∆rdτ 2 + 2
+ r+ dΩ2 . (3.17)
2κ∆r
We can now introduce a new coordinate ρ by
dr2 dr κ 2
dρ2 = ⇔ dρ = √ ⇔ ∆r = ρ , (3.18)
2κ∆r 2κ∆r 2
CHAPTER 3. LECTURE 3: BLACK HOLE THERMODYNAMICS 20
getting
ds2 = κ2 ρ2 dτ 2 + dρ2 + r+
2
dΩ2 = ρ2 dϕ2 + dρ2 + r+
2
dΩ2 , (3.19)
upon introducing a new angle coordinate, ϕ = κτ . This looks like a flat space
written in polar coordinates, provided the angle ϕ has a period 2π, otherwise
there is a conical singularity at ρ = 0, which corresponds to the original black
hole horizon. The reasoning now goes as follows: since the original black hole was
originally non-singular (there is no matter there), we expect it to be non-singular
again. This is achieved by setting (we want to avoid conical singularity)
κ
ϕ ∼ ϕ + 2π ⇔ τ ∼ τ + 2π/κ ⇔ T = , (3.20)
| {z } 2π
β
IE 1
F = = − log Z . (3.29)
β β
Our goal is thus to calculate the Euclidean action for the given black hole solution.
We then find the corresponding entropy by the standard thermodynamic relation:
∂F
S=− = β 2 ∂β F . (3.30)
∂T
Calculation for Schwarzschild. Recall that we need to calculate
1
Z
4 √ 1
Z √
IE = − d x gR − d3 x hK (3.31)
16πG M 8πG ∂M
for the Euclidean Schwarzschild solution
dr2 2M
ds2 = f dτ 2 + + r2 dΩ2 , f =1− . (3.32)
f r
The Euclidean Schwarzschild geometry corresponds to a “cigar”:
Even after the Wick rotation, this is a vacuum solution of Einstein equations,
Rµν = 0 and thence
SEH = 0 . (3.33)
CHAPTER 3. LECTURE 3: BLACK HOLE THERMODYNAMICS 22
Extrinsic curvature is
1 √ 1 2p 1 f 0 (R)
K = ∇µ nµ = √ ∂µ ( gnµ ) = 2 (r2 nr )0 = f (R) + p . (3.36)
g r r=R R 2 f (R)
To deal with this divergence we consider the appropriate background that has the
same time periodicity β and matches our spacetime exactly at the boundary:
This ‘cylinder’ corresponds to ‘thermal flat space’ filled with radiation of T = 1/β.
Its metric and induced metric are
We now use the ‘renormalization’ subtraction procedure and write the total
gravitational action as:
1
Z
4 √ 1
Z √
IE = − d x gR − d3 x h(K − K0 ) , (3.42)
16πG M 8πG ∂M
where the first boundary term, with K, is called the York term, and its purpose
is to yield a well posed Dirichlet variational principle, and the second term, with
K0 , is called the Gibbons-Hawking term, and its purpose is to ‘tune’ the value of
the action.
Thus we have
βM
ISch = . (3.43)
2
The corresponding free energy is thus
ISch M β
F = = = , (3.44)
β 2 16π
and the corresponding entropy reads:
β2 A
S = β 2 ∂β F = = 4πM 2 = πr+
2
= , (3.45)
16π 4
deriving the Bekenstein formula. At the same time, we can check that
F = M − TS , (3.46)
[3] D. Lovelock, The einstein tensor and its generalizations, Journal of Mathematical
Physics 12 (1971) 498–501.
[8] W. Israel, Singular hypersurfaces and thin shells in general relativity, Il Nuovo
Cimento B (1965-1970) 44 (1966) 1–14.
[9] C. Krishnan and A. Raju, A Neumann Boundary Term for Gravity, Mod. Phys.
Lett. A 32 (2017) 1750077, [1605.01603].
[10] D. Kastor, S. Ray and J. Traschen, Enthalpy and the Mechanics of AdS Black
Holes, Class. Quant. Grav. 26 (2009) 195011, [0904.2765].
24