Добавил:
Опубликованный материал нарушает ваши авторские права? Сообщите нам.
Вуз: Предмет: Файл:

Supersymmetry. Theory, Experiment, and Cosmology

.pdf
Скачиваний:
79
Добавлен:
01.05.2014
Размер:
12.6 Mб
Скачать

Cosmological constant 349

12.2Cosmological constant

As is well known, the cosmological constant appears as a constant in Einstein’s equations:

Rµν 21 gµν R = 8πGN Tµν + λgµν ,

(12.54)

where GN is Newton’s constant, Tµν is the energy–momentum tensor and Rµν the Ricci tensor, which is obtained from the Riemann tensor measuring the curvature of spacetime (see Appendix D). The cosmological constant λ is thus of the dimension of an inverse length squared. It was introduced by [127] in order to build a static Universe model, its repulsive e ect compensating the gravitational attraction, but, as we will now see, constraints on the expansion of the Universe impose for it a very small upper limit.

It is more convenient to work in the specific context of a homogeneous and isotropic Friedmann–Lemaˆıtre universe, with a Robertson–Walker metric:

ds2 = dt2 − a2(t)

 

 

 

,

 

 

2

 

 

 

 

 

dr

+ r2

2 + sin2 θdφ2

 

(12.55)

1

kr2

 

where a(t) is the cosmic scale factor, which is time dependent in an expanding or contracting Universe . Implementing energy conservation into the Einstein equations then leads to the Friedmann equation, which gives an expression for the Hubble parameter H measuring the rate of the expansion of the Universe:

H2

a˙ 2

(t)

=

1

(λ + 8πGN ρ)

k

(12.56)

 

 

 

 

.

a2

(t)

3

a2

In this equation, we use standard notation: a˙ is the time derivative of the cosmic scale factor, ρ = T 00 is the energy density and the term proportional to k is a spatial curvature term (see (12.55)). Note that the cosmological constant appears as a constant contribution to the Hubble parameter.

Evaluating each term of the Friedmann equation at present time t0 allows for a rapid estimation of an upper limit on λ (for a more precise determination see Appendix D). Indeed, we have for the Hubble constant H0 i.e. the present value of the Hubble parameter, H0 = h0 × 100 km.s1Mpc1 with h0 of order one, whereas the present energy density ρ0 is certainly within one order of magnitude of the critical energy density ρc = 3H02/(8πGN ) = h20 2.1026 kg m3; moreover the spatial curvature term certainly does not represent presently a dominant contribution to the expansion of the Universe. Thus, (12.56) considered at present time implies the following constraint on λ:

|λ| ≤ H02.

(12.57)

350 The challenges of supersymmetry

In other words, the length scale Λ ≡ |λ|1/2 associated with the cosmological constant must be larger than the Hubble length H0 ≡ cH01 = h0 1 1026 m, and thus be a cosmological distance.

This is not a problem as long as one remains classical: H0 provides a natural cosmological scale for our present Universe. The problem arises when one tries to combine gravity with quantum theory. Indeed, from Newton’s constant and the

Planck constant h¯, we have seen that we can construct the Planck mass scale mP =

"

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

hc/¯ (8πGN ) = 2.4 × 1018

GeV/c2

. The corresponding length scale is the Planck

length

 

 

h¯

 

 

 

 

 

 

 

 

 

 

 

P =

= 8.1 × 1035 m.

 

(12.58)

 

 

 

 

 

 

 

 

 

mP c

 

The above constraint now reads:

 

 

 

 

 

 

 

 

 

 

 

λ

1/2

 

 

=

c

 

1060

.

(12.59)

 

H0

 

 

Λ ≡ | |

 

 

 

H0

 

 

P

 

In other words, there are more than 60 orders of magnitude between the scale associated with the cosmological constant and the scale of quantum gravity.

A rather obvious solution is to take λ = 0. This is as valid a choice as any other in a pure gravity theory. Unfortunately, it is an unnatural one when one introduces any kind of matter. Indeed, set λ to zero but assume that there is a nonvanishing vacuum (i.e. ground state) energy: Tµν = ρvacgµν ; then the Einstein equations (12.54) read

Rµν 21 gµν R = 8πGN Tµν + 8πGN ρvacgµν .

(12.60)

The last term is interpreted as an e ective cosmological constant (from now on, we set ¯h = c = 1):

Λ4

 

λe = 8πGN ρvac mP2 .

(12.61)

Generically, ρvac receives a nonzero contribution from symmetry breaking: for instance, the scale Λ would be typically of the order of 100 GeV in the case of the electroweak gauge symmetry breaking or 1 TeV in the case of supersymmetry breaking. But the constraint (12.59) now reads:

Λ 1030 mP 103 eV.

(12.62)

It is this very unnatural fine tuning of parameters (in explicit cases ρvac and thus Λ are functions of the parameters of the theory) that is referred to as the cosmological constant problem, or more accurately the vacuum energy problem.

The most natural reason why vacuum energy would be vanishing is a symmetry argument. We have seen in Section 2.3 of Chapter 2 that global supersymmetry indeed provides such a rationale. The problem is that, at the same time, supersymmetry predicts equal boson and fermion masses and therefore needs to be broken. The amount of breaking necessary to push the supersymmetric partners high enough not to have been observed yet, is incompatible with the limit (12.62).

Cosmological constant 351

Moreover, in the context of cosmology, we should consider supersymmetry in a gravity context and thus work with its local version, supergravity. We have seen in Section 6.3.1 of Chapter 6 that, in this context, the criterion of vanishing vacuum energy is traded for one of vanishing gravitino mass. Local supersymmetry is then absolutely compatible with a nonvanishing vacuum energy, preferably a negative one (although possibly also a positive one). This is both a blessing and a problem: supersymmetry may be broken while the cosmological constant remains small, but we have lost our rationale for a vanishing, or very small, cosmological constant and fine tuning raises again its ugly head.

In some supergravity theories, however, one may recover the vanishing vacuum energy criterion.

Over the last few years, there has been an increasing number of indications that the Universe is presently undergoing accelerated expansion. This appears to be a strong departure from the standard picture of a matter-dominated universe. Indeed, the standard equation for the conservation of energy,

ρ˙ = 3(p + ρ)H,

(12.63)

allows us to derive from the Friedmann equation (12.56), written in the case of a Universe dominated by a component with energy density ρ and pressure p:

a¨

=

4πGN

(ρ + 3p).

(12.64)

a

3

Obviously, a matter-dominated (p 0) Universe is decelerating. One needs instead a component with a negative pressure.

A cosmological constant is associated with a contribution to the energy–momentum tensor as in (12.60), (12.61):

Tνµ = Λ4δνµ = (ρ, −p, −p, −p).

(12.65)

The associated equation of motion is therefore

p = −ρ.

(12.66)

It follows from (12.64) that a cosmological constant tends to accelerate expansion. As explained in Appendix D, recent cosmological observation expressed in the

plane (ΩM , ΩΛ) leads to the constraint given in the plot of Fig. D.2. It is certainly remarkable that a very diverse set of data singles out the same region in this parameter space: ΩM 0.2 to 0.3 and ΩΛ 0.7 to 0.8.

This raises a new problem. Since matter and a cosmological constant evolve very di erently, why should they be of the same order at present times? Indeed, for a component of equation of state p = , we may rewrite (12.63) as

ρ˙

a˙

 

 

 

 

= 3

 

 

(1 + w).

(12.67)

ρ

a

Thus matter (p 0) energy density evolves as a3 whereas a cosmological constant stays constant, as expected. Why should they be presently of similar magnitude (see

352 The challenges of supersymmetry

Fig. 12.4)? This is known as the cosmic coincidence problem. In order to avoid any reference to us (and hence any anthropic interpretation, see below), we may rephrase the problem as follows: why does the vacuum energy starts to dominate at a time tΛ (redshift zΛ 1) which almost coincides with the epoch tG (redshift zG 3 to 5) of galaxy formation?

12.2.1Relaxation mechanisms

From the point of view of high energy physics, it is however di cult to imagine a rationale for a pure cosmological constant, especially if it is nonzero but small compared to the typical fundamental scales (electroweak, strong, grand unified, or Planck scale). There should be some dynamics associated with this form of energy.

For example, we have seen in Chapter 10 that, in the context of string models, any dimensionful parameter is expressed in terms of the fundamental string scale MS and of vacuum expectation values of scalar fields. The physics of the cosmological constant would then be the physics of the corresponding scalar fields. Indeed, it is di cult to envisage string theory in the context of a true cosmological constant. The corresponding spacetime is known as de Sitter spacetime and has an event horizon (see Appendix D). This is di cult to reconcile with the S-matrix approach of string theory in the context of conformal invariance. More precisely, in the S-matrix approach, states are asymptotically (i.e. at times t → ±∞) free and interact only at finite times: the S-matrix element between an incoming set of free states and an outgoing set yields the probability associated with such a transition. In string theory, the states are strings and a diagram such as the one given in Fig. 10.5 of Chapter 10 gives a contribution to the S-matrix element. But conformal invariance, which we noted in Chapter 10 to be a key element, imposes that the string world-sheet can be deformed at will: this is di cult to reconcile with the presence of a horizon and the requirement of asymptotically free states. Recent advances have been obtained in this domain by the use of fluxes.

 

 

0 6 7

 

 

 

 

 

 

 

T0

 

1

 

 

 

 

 

 

 

 

 

1

0 6 1

 

 

 

 

 

 

 

 

 

1

0 5 5

 

 

 

 

 

 

 

 

 

1

0 4 9

 

 

 

ρradiation

 

 

 

1

0 4 3

 

 

 

 

 

 

1

0 3 7

 

 

 

 

 

 

 

 

]

1 0 3 1

 

 

 

 

 

 

 

 

–3

1

0

2 5

 

 

ρmatter

 

 

 

 

cm

1 0 1 9

 

 

 

 

 

 

[GeV

1

0 1 3

 

 

 

 

 

 

 

 

 

1 0 7

 

 

 

 

 

 

 

 

 

1 0 1

 

 

 

 

ρΛ

 

 

ρ

1 0

5

 

 

 

 

 

 

 

1 0

1 1

 

 

 

 

 

 

 

 

 

1 0

1 7

 

 

 

 

 

 

 

 

 

1 0

2 3

 

 

 

 

 

 

 

 

 

1 0

2 9

 

 

 

 

 

 

 

 

 

1 0 – 3 5

 

 

 

 

 

 

 

 

 

1 0 – 4 1

 

 

 

 

 

 

 

 

 

 

 

106

104

102

100

10–2

10–4

10–6

10–8

10–10 10–12 10–14 10–16 10–18

 

 

 

 

 

 

 

 

T [GeV]

 

 

Fig. 12.4 Evolution of radiation, matter and vacuum energy densities with temperature.

Cosmological constant 353

Steven Weinberg [361] has constrained the possible mechanisms for the relaxation of the cosmological constant by proving the following “no-go” theorem: it is not possible to obtain a vanishing cosmological constant as a consequence of the equations of motion of a finite number of fields.

Indeed, let us consider N such fields ϕn, n = 1, . . . , N . In the equilibrium configuration these fields are constant and their equations of motion simply read

δL

= 0.

(12.68)

δϕn

 

Remembering that λe T µµ where the energy-momentum tensor may be obtained from the metric4, we see that the vanishing of the cosmological constant is a consequence of the equations (12.68) if we can find N functions fn(ϕ) such that

2gµν

δL

=

 

δL

fn(ϕ).

(12.69)

 

 

δgµν

n δϕn

 

This amounts to a symmetry condition, the invariance of the Lagrangian L under

δgµν = 2αgµν , δϕn = −αfn(ϕ).

(12.70)

However, one can redefine the fields ϕn, n = 1, . . . , N , into σa, a = 1, . . . , N − 1, and ϕ in such a way that the invariance reads

δgµν = 2αgµν , δσa = 0, δϕ = −α.

(12.71)

The Lagrangian which satisfies this invariance is written

 

L =

 

L0(σ) = e4ϕ"

 

L0(σ)

 

Det (e2ϕgµν )

(12.72)

|g|

which does not provide a solution to the relaxation of the cosmological constant, as can be seen by redefining the metric: gˆµν = e2ϕgµν (in the new metric, the field ϕ has only derivative couplings).

Obviously, Weinberg’s no-go theorem relies on a series of assumptions: Lorentz invariance, finite number of constant fields, possibility of globally redefining these fields, etc.

An example of a relaxation mechanism is provided by the Brown–Teitelboim mechanism [53, 54] where the quantum creation of closed membranes leads to a reduction of the vacuum energy inside. This is easier to understand on a toy model with a single spatial dimension.

Let us thus consider a line and establish along it a constant electric field E0 > 0: the corresponding (vacuum) energy is E02/2. Quantum creation of a pair of ±q-charged particles (q > 0) leads to the formation of a region (between the two charges) where

the electric field is partially

screened to the value E

0

q and thus the vacuum energy

 

2

 

 

is decreased to the value (E0 − q)

 

/2. Quantum creation of pairs in the new region will

4See equation (D.15).

354 The challenges of supersymmetry

subsequently decrease the value of the vacuum energy. The process ends in flat space when the electric field reaches the value E ≤ q/2 because it then becomes insu cient to separate the pairs created.

In a truly three-dimensional universe, the quantum creation of pairs is replaced by the quantum creation of membranes and the one-dimensional electric field is replaced by a tensor field Aµνρ.There are two potential problems with such a relaxation of the cosmological constant.

First, since the region of small cosmological constant originates from regions with large vacuum energies, hence exponential expansion, it is virtually empty: matter has to be produced through some mechanism yet to be specified. The second problem has to do with the multiplicity of regions with di erent vacuum energies: why should we be in the region with the smallest value? Such questions are crying for an anthropic type of answer: some regions of spacetime are preferred because they allow the existence of observers.

More generally, the anthropic principle approach can be sketched as follows. We consider regions of spacetime with di erent values of tG (time of galaxy formation) and tΛ, the time when the cosmological constant starts to dominate, i.e. when the Universe enters a de Sitter phase of exponential expansion. Clearly galaxy formation must precede this phase otherwise no observer (similar to us) would be able to witness it. Thus tG ≤ tΛ. On the other hand, regions with tΛ tG have not yet undergone any de Sitter phase of reacceleration and are thus “phase-space suppressed” compared

>

with regions with tΛ tG. Hence the regions favored have tΛ tG and thus ρΛ ρM .

12.2.2Dark energy

In this section, we will take a slightly di erent route. We assume that some unknown mechanism relaxes the vacuum energy to zero or to a very small value. We then introduce some new dynamical component which accounts for the present observations. Introducing dynamics generally modifies the equation of state (12.66) of this new component of the energy density to the more general form with negative pressure5:

p = wρ, w < 0.

(12.73)

For example, a network of light, nonintercommuting topological defects [343,355] gives w = −n/3 where n is the dimension of the defect, i.e. 1 for a string and 2 for a domain wall. The equation of state for a minimally coupled scalar field necessarily satisfies the condition w ≥ −1.

Experimental data may constrain such a dynamical component, referred to in the literature as dark energy, just as it did with the cosmological constant. For example, in a spatially flat Universe with only matter and an unknown component X with equation of state pX = wX ρX , one obtains from (12.64) with ρ = ρM + ρX , p = wX ρX the following form for the deceleration parameter (see Section D.2.1 of Appendix D)

q0 =

a¨0a0

=

M

+ (1 + 3wX )

X

,

(12.74)

a˙ 02

 

2

2

where ΩX = ρX c. Supernovae results give a constraint on the parameter wX .

5We recall that nonrelativistic matter (dust) has an equation of state p 0 whereas p = ρ/3 corresponds to radiation.

Cosmological constant 355

Another important property of dark energy is that it does not appear to be clustered (just as a cosmological constant). Otherwise, its e ects would have been detected locally, as for the case of dark matter.

A particularly interesting candidate in the context of fundamental theories is a scalar6 field φ slowly evolving in its potential V (φ). Indeed, since the corresponding energy density and pressure are, for a minimally coupled scalar field,

ρφ =

1

˙2

+ V (φ)

(12.75)

2

φ

 

pφ =

1

˙2

− V (φ),

(12.76)

2

φ

 

we have

 

 

˙

 

 

 

 

 

1

2

− V (φ)

 

 

wφ =

2

φ

 

.

(12.77)

1

˙

2

 

 

+ V (φ)

 

 

2

φ

 

 

˙

2

/2 V (φ)), we clearly obtain 1 ≤ wφ 0.

If the kinetic energy is subdominant (φ

 

A consequence is that the corresponding speed of sound cs = δp/δρ is of the order of the speed of light: the scalar field pressure resists gravitational clustering.

We will see below that the scalar field must be extremely light. Two situations have been considered:

A scalar potential slowly decreasing to zero as φ goes to infinity ( [63, 322, 365]). This is often referred to as quintessence or runaway quintessence.

A very light field (pseudo-Goldstone boson) which is presently relaxing to its vacuum state [166].

In both cases one is relaxing to a position where the vacuum energy is zero. This is associated with our assumption that some unknown mechanism wipes the cosmological constant out. We discuss the two cases in turn.

Runaway quintessence

A runaway potential is frequently present in models where supersymmetry is dynamically broken. We have seen that supersymmetric theories are characterized by a scalar potential with many flat directions, i.e. directions φ in field space for which the potential vanishes. The corresponding degeneracy is lifted through dynamical supersymmetry breaking. In some instances (dilaton or compactification radius), the field expectation value φ actually provides the value of the strong interaction coupling. Then at infinite φ value, the coupling e ectively goes to zero together with the supersymmetry breaking e ects and the flat direction is restored: the potential decreases monotonically to zero as φ goes to infinity.

Let us take the example of supersymmetry breaking by gaugino condensation in e ective superstring theories. We recall briefly the discussion of Section 7.4.2. The value g0 of the gauge coupling at the string scale MS is provided by the vacuum expectation value of the dilaton field s (taken to be dimensionless by dividing by mP )

6A vector field or any field which is not a Lorentz scalar must have settled down to a vanishing value, otherwise, Lorentz invariance would be spontaneously broken.

356 The challenges of supersymmetry

present among the massless string modes: g02 = s 1. If the gauge group has a oneloop beta function coe cient b > 0, then the running gauge coupling becomes strong at the scale

Λ MS e8π2/(bg02) = MS e8π2s/b.

(12.78)

At this scale, the gaugino fields are expected to condense. Through dimensional analy-

¯

3

. Terms quadratic in the

sis, the gaugino condensate λλ is expected to be of order Λ

 

gaugino fields thus yield in the e ective theory below condensation scale a potential for the dilaton:

 

 

(12.79)

V λλ¯

2 e48π2s/b.

The s-dependence of the potential is of course more complicated and one usually looks for stable minima with vanishing cosmological constant. But the behavior (12.78) is characteristic of the large s region and provides a potential slopping down to zero at infinity as required in the quintessence approach. Similar behavior is observed for moduli fields whose vev describes the radius of the compact manifolds which appear from the compactification from 10 or 11 dimensions to four in superstring theories.

Let us take therefore the example of an exponentially decreasing potential. More explicitly, we consider the following action

S =

d4x

 

m2

1

µφ∂µφ − V (φ) ,

 

g

P

R +

 

(12.80)

2

2

which describes a real scalar field φ minimally coupled with gravity and the selfinteractions of which are described by the potential:

V (φ) = V0e−λφ/mP ,

(12.81)

where V0 is a positive constant.

The energy density and pressure stored in the scalar field are given by (12.75) and (12.76). We will assume that the background (matter and radiation) energy density

ρB and pressure pB obey a standard equation of state

 

pB = wB ρB .

(12.82)

If one neglects the spatial curvature (k 0), the equation of motion for φ simply reads

¨

˙

 

dV

 

(12.83)

φ + 3=

,

with

 

 

 

 

 

H2 =

1

(ρB + ρφ).

(12.84)

3m2

 

 

 

 

 

 

P

 

 

 

 

This can be rewritten as

 

 

 

 

 

 

 

˙2

 

(12.85)

ρ˙φ = 3Hφ .

 

We are looking for scaling solutions, i.e. solutions where the φ energy density scales as a power of the cosmic scale factor: ρφ a−nφ or ρ˙φφ = −nφH (nφ being constant).

Cosmological constant 357

In this case, one easily obtains from (12.75), (12.76), and (12.85) that the φ field obeys a standard equation of state

pφ = wφρφ,

(12.86)

with

nφ

 

 

 

wφ =

 

1.

(12.87)

3

Hence

 

 

 

ρφ a3(1+wφ).

(12.88)

If one can neglect the background energy ρB , then (12.84) yields a simple di erential equation for a(t) which is solved as:

a t2/[3(1+wφ)].

(12.89)

˙2 = (1 + wφ)ρφ t2, one deduces that φ varies logarithmically with time. Since φ

One then easily obtains from (12.83,12.84) that

φ = φ0 +

2

mP

ln(t/t0).

(12.90)

λ

 

 

 

 

 

 

and7

 

λ2

 

 

wφ =

1,

(12.91)

 

3

 

It is clear from (12.91) that, for λ su ciently small, the field φ can play the rˆole of quintessence. We note that, even if we started with a small value φ0, φ reaches a value of order mP .

But the successes of the standard Big Bang scenario indicate that clearly ρφ cannot have always dominated: it must have emerged from the background energy density ρB . Let us thus now consider the case where ρB dominates. It turns out that the solution just discussed with ρφ ρB and (12.91) is a late time attractor [218] only if λ2 < 3(1 + wB ). If λ2 > 3(1 + wB ), the global attractor turns out to be a scaling solution [87, 160, 365] with the following properties:

φ

ρφ

=

3

(1 + wB )

(12.92)

ρφ + ρB

λ2

 

wφ = wB .

(12.93)

The second equation (12.93) clearly indicates that this does not correspond to a dark energy solution (12.73).

The semirealistic models discussed earlier tend to give large values of λ and thus the latter scaling solution as an attractor. For example, in the case (12.79) where the scalar field is the dilaton, λ = 48π2/b with b = 90 for a E8 gauge symmetry down to b = 9 for SU (3). Moreover [161], on the observational side, the condition that

7Under the condition λ2 6 (wφ 1 since V (φ) 0).

358 The challenges of supersymmetry

ρφ should be subdominant during nucleosynthesis (in the radiation-dominated era) imposes to take rather large values of λ. Typically requiring ρφ/(ρφ + ρB ) to be then smaller than 0.2 imposes λ2 > 20.

Ways to obtain a quintessence component have been proposed however. Let us sketch some of them in turn.

One is the notion of tracker field8 [387]. This idea also rests on the existence of scaling solutions of the equations of motion which play the rˆole of late time attractors, as illustrated above. An example is provided by a scalar field described by the action (12.80) with a potential

V (φ) = λ

Λ4+α

(12.94)

φα

 

 

with α > 0. In the case where the background density dominates, one finds an attractor scaling solution [304, 322] φ a3(1+wB )/(2+α), ρφ a3α(1+wB )/(2+α) (see Exercise 1). Thus ρφ decreases at a slower rate than the background density (ρB a3(1+wB )) and tracks it until it becomes of the same order at a given value aQ. We thus have:

 

 

φ

 

a

3(1+wB )/(2+α)

 

 

 

,

(12.95)

 

 

 

 

 

mP

aQ

 

 

ρφ

 

a

 

6(1+wB )/(2+α)

 

 

 

.

(12.96)

 

ρB

aQ

 

One finds

 

 

 

α(1 + wB )

 

 

 

wφ = 1 +

(12.97)

 

 

 

 

.

 

 

 

2 + α

Shortly after φ has reached for a = aQ a value of order mP , it satisfies the standard slow-roll conditions (equations (D.109) and (D.110) of Appendix D) and therefore (12.97) provides a good approximation to the present value of wφ. Thus, at the end of the matter-dominated era, this field may provide the quintessence component that we are looking for.

Two features are interesting in this respect. One is that this scaling solution is reached for rather general initial conditions, i.e. whether ρφ starts of the same order or much smaller than the background energy density [387]. Regarding the cosmic coincidence problem, it can be rephrased here as follows (since φ is of order mP in this scenario): why is V (mP ) of the order of the critical energy density ρc? It is thus the scale Λ which determines the time when the scalar field starts to emerge and the Universe expansion reaccelerates. Indeed, using (12.94), the constraint reads:

Λ H02mP2+α 1/(4+α) .

(12.98)

We may note that this gives for α = 2, Λ 10 MeV, not such an atypical scale for high energy physics.

8Somewhat of a misnomer since in this solution, as we see below, the field φ energy density tracks the radiation-matter energy density before overcoming it, in contradistinction with (12.92). One should rather describe it as a transient tracker field.