Physics
Physics
Physics
Thomas P. Sotiriou
Center for Fundamental Physics, University of Maryland, College Park, MD 20742-4111, USA
Valerio Faraoni
Physics Department, Bishops University, 2600 College St., Sherbrooke, Qu`ebec, Canada J1M 1Z7
Modified gravity theories have received increased attention lately due to combined motivation
coming from high-energy physics, cosmology and astrophysics. Among numerous alternatives
to Einsteins theory of gravity, theories which include higher order curvature invariants, and
specifically the particular class of f (R) theories, have a long history. In the last five years there has
been a new stimulus for their study, leading to a number of interesting results. We review here f (R)
theories of gravity in an attempt to comprehensively present their most important aspects and
cover the largest possible portion of the relevant literature. All known formalisms are presented
metric, Palatini and metric-affine and the following topics are discussed: motivation; actions,
field equations and theoretical aspects; equivalence with other theories; cosmological aspects and
constraints; viability criteria; astrophysical applications.
Contents
I. Introduction
A. Historical
B. Contemporary Motivation
C. f (R) theories as toy theories
II. Actions and field equations
A. Metric formalism
B. Palatini formalism
C. Metric-affine formalism
1. Preliminaries
2. Field Equations
4. Ghost fields
C. The Cauchy problem
1
1
2
3
4
5
6
8
9
10
11
12
13
13
14
15
15
16
18
18
19
Present address: Department of Applied Mathematics and Theoretical Physics, Centre for Mathematical Sciences, University of
Cambridge, Wilberforce Road, Cambridge, CB3 0WA, UK.
T.Sotiriou@damtp.cam.ac.uk
vfaraoni@ubishops.ca
31
31
34
34
35
36
36
37
40
40
40
41
42
Acknowledgments
43
References
43
I. INTRODUCTION
A. Historical
2
These early attempts were triggered mainly by scientific curiosity and a will to question and, therefore, understand the then newly proposed theory. It is quite
straightforward to realize that complicating the action
and, consequently, the field equations with no apparent
theoretical or experimental motivation is not very appealing. However, the motivation was soon to come.
Beginning in the 1960s, there appeared indications
that complicating the gravitational action might indeed
have its merits. GR is not renormalizable and, therefore,
can not be conventionally quantized. In 1962, Utiyama
and De Witt showed that renormalization at one-loop demands that the EinsteinHilbert action be supplemented
by higher order curvature terms (Utiyama and DeWitt,
1962). Later on, Stelle showed that higher order actions are indeed renormalizable (but not unitary) (Stelle,
1977). More recent results show that when quantum
corrections or string theory are taken into account,
the effective low energy gravitational action admits
higher order curvature invariants (Birrell and Davies,
1982; Buchbinder et al., 1992; Vilkovisky, 1992).
Such considerations stimulated the interest of the
scientific community in higher-order theories of gravity,
i.e., modifications of the EinsteinHilbert action in order
to include higher-order curvature invariants with respect
to the Ricci scalar [see (Schmidt, 2007) for a historical
review and a list of references to early work]. However,
the relevance of such terms in the action was considered
to be restricted to very strong gravity regimes and
they were expected to be strongly suppressed by small
couplings, as one would expect when simple effective
field theory considerations are taken into account.
Therefore, corrections to GR were considered to be
important only at scales close to the Planck scale and,
consequently, in the early universe or near black hole singularities and indeed there are relevant studies, such
as the well-known curvature-driven inflation scenario
(Starobinsky, 1980) and attempts to avoid cosmological and black hole singularities (Brandenberger,
1992,
1993,
1995;
Brandenberger et al.,
1993;
Mukhanov and Brandenberger, 1992; Shahid-Saless,
1990; Trodden et al., 1993).
However, it was not
expected that such corrections could affect the gravitational phenomenology at low energies, and consequently
large scales such as, for instance, the late universe.
B. Contemporary Motivation
Recall that, from GR in the absence of the cosmological constant and under the standard cosmological assumptions (spatial
homogeneity and isotropy etc.), one obtains the second Friedmann equation
a
4 G
=
( + 3P ) ,
(1)
a
3
where a is the scale factor, G is the gravitational constant and
and P are the energy density and the pressure of the cosmological
fluid, respectively. Therefore, if the Strong Energy Condition
+ 3P 0 is satisfied, there can be no acceleration (gravity is
attractive).
3
of the evolution and the matter/energy content of the
universe is at least surprising and definitely calls for an
explanation. The simplest model which adequately fits
the data creating this picture is the so called concordance
model or CDM (-Cold Dark Matter), supplemented by
some inflationary scenario, usually based on some scalar
field called inflaton. Besides not explaining the origin
of the inflaton or the nature of dark matter by itself, the
CDM model is burdened with the well known cosmological constant problems (Carroll, 2001a; Weinberg, 1989):
the magnitude problem, according to which the observed
value of the cosmological constant is extravagantly small
to be attributed to the vacuum energy of matter fields,
and the coincidence problem, which can be summed up
in the question: since there is just an extremely short
period of time in the evolution of the universe in which
the energy density of the cosmological constant is comparable with that of matter, why is this happening today
that we are present to observe it?
These problems make the CDM model more of an
empirical fit to the data whose theoretical motivation
can be regarded as quite poor. Consequently, there
have been several attempts to either directly motivate
the presence of a cosmological constant or to propose
dynamical alternatives to dark energy. Unfortunately,
none of these attempts are problem-free. For instance,
the so-called anthropic reasoning for the magnitude of
(Barrow and Tipler, 1986; Carter, 1974), even when
placed into the firmer grounds through the idea of the
anthropic or string landscape (Susskind, 2003), still
makes many physicists feel uncomfortable due to its
probabilistic nature. On the other hand, simple scenarios for dynamical dark energy, such as quintessence
(Bahcall et al., 1999; Caldwell et al., 1998; Carroll, 1998;
Ostriker and Steinhardt, 1995; Peebles and Ratra, 1988;
Ratra and Peebles, 1988; Wang et al., 2000; Wetterich,
1988) do not seem to be as well motivated theoretically
as one would desire.2
Another perspective towards resolving the issues described above, which might appear as more radical to
some, is the following: gravity is by far the dominant interaction at cosmological scales and, therefore, it is the
force governing the evolution of the universe. Could it be
that our description of the gravitational interaction at the
relevant scales is not sufficiently adequate and stands at
the root of all or some of these problems? Should we consider modifying our theory of gravitation and if so, would
this help in avoiding dark components and answering the
cosmological and astrophysical riddles?
It is rather pointless to argue whether such a perspec-
We are referring here to the fact that, not only the mass of the
scalar turns out to be many orders of magnitude smaller than any
of the masses of the scalar fields usually encountered in particle
physics, but also to the inability to motivate the absence of any
coupling of the scalar field to matter (there is no mechanism or
symmetry preventing this) (Carroll, 2001b).
tive would be better or worse than any of the other solutions already proposed. It is definitely a different way
to address the same problems and, as long as these problems do not find a plausible, well accepted and simple,
solution, it is worth pursuing all alternatives. Additionally, questioning the gravitational theory itself definitely
has its merits: it helps us to obtain a deeper understanding of the relevant issues and of the gravitational interaction, it has high chances to lead to new physics and it has
worked in the past. Recall that the precession of Mercurys orbit was at first attributed to some unobserved
(dark) planet orbiting inside Mercurys orbit, but it
actually took the passage from Newtonian gravity to GR
to be explained.
1
SEH =
d4 x g R,
(2)
2
where 8G, G is the gravitational constant, g is
the determinant of the metric and R is the Ricci scalar
(c = ~ = 1), to become a general function of R, i.e.,
Z
1
(3)
S=
d4 x g f (R).
2
Before going further into the discussion of the details
and the history of such actions this will happen in
the forthcoming section some remarks are in order.
We have already mentioned the motivation coming from
high-energy physics for adding higher order invariants
to the gravitational action, as well as a general motivation coming from cosmology and astrophysics for seeking
generalizations of GR. There are, however, still two questions that might be troubling the reader. The first one
is: Why specifically f (R) actions and not more general
ones, which include other higher order invariants, such
as R R ?
4
The answer to this question is twofold. First of all,
there is simplicity: f (R) actions are sufficiently general to encapsulate some of the basic characteristics of
higher-order gravity, but at the same time they are simple enough to be easy to handle. For instance, viewing f
as a series expansion, i.e.,
f (R) = . . . +
1
R2 R3
2
+
+
+ . . . , (4)
2
+
R
+
R2
R
2
3
probably reasonable.
To conclude, when all of the above is taken into account, f (R) gravity should neither be over- nor underestimated. It is an interesting and relatively simple alternative to GR, from the study of which some useful conclusions have been derived already. However, it is still
a toy-theory, as already mentioned; an easy-to-handle
deviation from Einsteins theory mostly to be used in order to understand the principles and limitations of modified gravity. Similar considerations apply to modifying
gravity in general: we are probably far from concluding
whether it is the answer to our problems at the moment.
However, in some sense, such an approach is bound to be
fruitful since, even if it only leads to the conclusion that
GR is the only correct theory of gravitation, it will still
have helped us to both understand GR better and secure
our faith in it.
As can be found in many textbooks see, for example (Misner et al., 1973; Wald, 1984) there are
actually two variational principles that one can apply
to the EinsteinHilbert action in order to derive Einsteins equations: the standard metric variation and a
less standard variation dubbed Palatini variation [even
though it was Einstein and not Palatini who introduced
it (Ferraris et al., 1982)]. In the latter the metric and the
connection are assumed to be independent variables and
one varies the action with respect to both of them (we
will see how this variation leads to Einsteins equations
shortly), under the important assumption that the matter action does not depend on the connection. The choice
of the variational principle is usually referred to as a formalism, so one can use the terms metric (or second order)
formalism and Palatini (or first order) formalism. However, even though both variational principles lead to the
same field equation for an action whose Lagrangian is linear in R, this is no longer true for a more general action.
Therefore, it is intuitive that there will be two version of
f (R) gravity, according to which variational principle or
formalism is used. Indeed this is the case: f (R) gravity
in the metric formalism is called metric f (R) gravity and
f (R) gravity in the Palatini formalism is called Palatini
f (R) gravity (Buchdahl, 1970).
Finally, there is actually even a third version of f (R) gravity:
metric-affine f (R) gravity
(Sotiriou and Liberati, 2007a,b).
This comes about
if one uses the Palatini variation but abandons the
assumption that the matter action is independent of the
connection. Clearly, metric affine f (R) gravity is the
most general of these theories and reduces to metric or
Palatini f (R) gravity if further assumptions are made.
In this section we will present the actions and field
equations of all three versions of f (R) gravity and point
out their difference. We will also clarify the physical
meaning behind the assumptions that discriminate
5
them.
For an introduction to metric f (R) gravity see
also (Nojiri and Odintsov, 2007a), for a shorter
review of metric and Palatini f (R) gravity see
(Capozziello and Francaviglia, 2008) and for an extensive analysis of all versions of f (R) gravity and other
alternative theories of gravity see (Sotiriou, 2007b).
A. Metric formalism
1
Smet =
(5)
d4 x g f (R) + SM (g , ),
2
where collectively denotes the matter fields. Variation
with respect to the metric gives, after some manipulations and modulo surface terms
1
f (R)R f (R)g [ g ] f (R) = T ,
2
(6)
where, as usual,
2 SM
T =
,
g g
(7)
a prime denotes differentiation with respect to the argument, is the covariant derivative associated with
the Levi-Civita connection of the metric, and
. Metric f (R) gravity was first rigorously studied
in (Buchdahl, 1970).3
It has to be stressed that there is a mathematical jump
in deriving eq. (6) from the action (5) having to do with
the surface terms that appear in the variation: as in the
case of the EinsteinHilbert action, the surface terms do
not vanish just by fixing the metric on the boundary.
For the EinsteinHilbert action, however, these terms
gather into a total variation of a quantity. Therefore,
it is possible to add a total divergence to the action in
order to heal it and arrive to a well-defined variational
principle (this is the well known GibbonsHawkingYork
surface term (Gibbons and Hawking, 1977; York, 1972)).
Unfortunately, the surface terms in the variation of the
action (3) do not consist of a total variation of some
quantity (the reader is urged to calculate the variation in
order to verify this fact) and it is not possible to heal
the action by just subtracting some surface term before
performing the variation.
The way out comes from the fact that the action includes higher order derivatives of the metric and, therefore, it should be possible to fix more degrees of freedom
on the boundary than those of the metric itself. There is
(8)
(9)
6
the left hand side of eq. (6) is divergence-free (generalized
Bianchi identity) implying that T = 0 (Koivisto,
2006a).4
Finally, let us note that it is possible to write the field
equations in the form of Einstein equations with an effective stress-energy tensor composed of curvature terms
moved to the right hand side. This approach is questionable in principle (the theory is not Einsteins theory and
it is artificial to force upon it an interpretation in terms
of Einstein equations) but, in practice, it has been proved
to be useful in scalar-tensor gravity. Specifically, eq. (6)
can be written as
1
g R
2
[f (R) Rf (R)]
T
+ g
=
f (R)
2f (R)
[ f (R) g f (R)]
+
f (R)
G R
(10)
or
(ef f )
T
+
T
,
f (R)
G =
(11)
1 h f (R) Rf (R)
g + f (R)
2
i
g f (R)
(12)
B. Palatini formalism
We have already mentioned that the Einstein equations can be derived using, instead of the standard metric variation of the EinsteinHilbert action, the Palatini
formalism, i.e., an independent variation with respect to
the metric and an independent connection (Palatini variation). The action is formally the same but now the Riemann tensor and the Ricci tensor are constructed with
the independent connection. Note that the metric is not
needed to obtain the latter from the former. For clarity
of notation, we denote the Ricci tensor constructed with
this independent connection as R and the corresponding Ricci scalar5 is R = g R . The action now takes
the form
Z
1
Spal =
(13)
d4 x g f (R) + SM (g , ).
2
GR will come about, as we will see shortly, when f (R) =
R. Note that the matter action SM is assumed to depend only on the metric and the matter fields and not on
the independent connection. This assumption is crucial
for the derivation of Einsteins equations from the linear
version of the action (13) and is the main feature of the
Palatini formalism.
It has already been mentioned that this assumption has
consequences for the physical meaning of the independent
connection (Sotiriou, 2006b,d; Sotiriou and Liberati,
2007b). Let us elaborate on this: recall that an affine
connection usually defines parallel transport and the covariant derivative. On the other hand, the matter action
SM is supposed to be a generally covariant scalar which
includes derivatives of the matter fields. Therefore, these
derivatives ought to be covariant derivatives for a general
matter field. Exceptions exist, such as a scalar field, for
which a covariant and a partial derivative coincide, and
the electromagnetic field, for which one can write a covariant action without the use of the covariant derivative
[it is the exterior derivative that is actually needed, see
next section and (Sotiriou and Liberati, 2007b)]. However, SM should include all possible fields. Therefore,
assuming that SM is independent of the connection can
imply one of two things (Sotiriou, 2006d): either we are
restricting ourselves to specific fields, or we are implicitly assuming that it is the Levi-Civita connection of the
metric that actually defines parallel transport. Since the
first option is implausibly limiting for a gravitational theory, we are left with the conclusion that the independent
connection does not define parallel transport or the
covariant derivative and the geometry is actually pseudoRiemannian. The covariant derivative is actually defined
by the Levi-Civita connection of the metric { }.
This also implies that Palatini f (R) gravity is a metric theory in the sense that it satisfies the metric postulates (Will, 1981). Let us clarify this: matter is minimally coupled to the metric and not coupled to any other
fields. Once again, as in GR or metric f (R) gravity,
one could use diffeomorphism invariance to show that
5
4
7
the stress energy tensor is conserved by the covariant
derivative defined with the Levi-Civita connection of the
T 6= 0). This can
metric, i.e., T = 0 (but
also be shown by using the field equations, which we will
present shortly, in order to calculate the divergence of T
with respect to the Levi-Civita connection of the metric
and show that it vanishes (Barraco et al., 1999; Koivisto,
2006a).6 Clearly then, Palatini f (R) gravity is a metric
theory according to the definition of (Will, 1981) (not
to be confused with the term metric in metric f (R)
gravity, which simply refers to the fact that one only
varies the action with respect to the metric). Conventionally thinking, as a consequence of the covariant conservation of the matter energy-momentum tensor, test
particles should follow geodesics of the metric in Palatini
f (R) gravity. This can be seen by considering a dust
fluid with T = u u and projecting the conservation equation T = 0 onto the fluid four-velocity u .
Similarly, theories that satisfy the metric postulates are
supposed to satisfy the Einstein Equivalence Principle as
well (Will, 1981). Unfortunately, things are more complicated here and, therefore, we set this issue aside for the
moment. We will return to it and attempt to fully clarify
it in Secs. VI.B and VI.C.2. For now, let us proceed with
our discussion of the field equations.
Varying the action (13) independently with respect to
the metric and the connection, respectively, and using
the formula
.
R =
(14)
(20)
yields
1
f (R)R() f (R)g = T ,
2
gf (R)g
)
+
gf (R)g ( = 0,
(16)
(17)
which implies that we can bring the field equations into
the more economical form
1
f (R)R() f (R)g = G T ,
2
gf (R)g = 0,
f (R)R 2f (R) = 0.
(15)
(18)
(19)
Energy supertensors and pseudotensors in Palatini f (R) gravity were studied in (Barraco et al., 1999; Borowiec et al., 1994,
1998; Ferraris et al., 1992) and alternative energy definitions
were given in (Deser and Tekin, 2002, 2003a,b, 2007).
(21)
h f (R)g .
(22)
8
It can easily be shown that8
h h = g f (R)g .
(23)
Then, eq. (19) becomes the definition of the LeviCivita connection of h and can be solved algebraically
to give
=
1
h ( h + h h ) ,
2
(24)
2 f (R)
i
(f (R)g ) ,
(25)
2 (f (R))2
1
1
g f (R). (26)
f (R)
2
R = R +
R=R +
(27)
f
1
G = T g R
(28)
f
2
f
1
+ ( g ) f
f
3 1
1
2 ( f )( f ) g (f )2 .
2f
2
9
8
This calculation holds in four dimensions. When the number of dimensions D is different from 4 then, instead of using eq. (22), the conformal metric h should be introduced as
h [f (R)]2/(D2) g in order for eq. (23) to still hold.
9
would imply that we would define the covariant derivatives of the matter fields with this connection and, therefore, we would have SM = SM (g , , ). The action of this theory, dubbed metric-affine f (R) gravity
(Sotiriou and Liberati, 2007b), would then be [note the
difference with respect to the action (13)]
Z
1
Sma =
d4 x gf (R) + SM (g , , ). (30)
2
1. Preliminaries
(31)
(32)
which measures the failure of the connection to covariantly conserve the metric, the trace of the non-metricity
tensor with respect to its last two (symmetric) indices,
which is called the Weyl vector,
Q
1
Q ,
4
(33)
(34)
starting with the EinsteinCartan(SciamaKibble) theory (Cartan, 1922, 1923, 1924; Hehl et al., 1976; Kibble,
1961; Sciama, 1964). In this theory, as well as in other
theories with an independent connection, some part of
the connection is still related to the metric (e.g., the
non-metricity is set to zero). In our case, the connection is left completely unconstrained and is to be determined by the field equations. Metric-affine gravity with
the linear version of the action (30) was initially proposed in (Hehl and Kerling, 1978) and the generalization
to f (R) actions was considered in (Sotiriou and Liberati,
2007a,b).
Unfortunately, leaving the connection completely unconstrained comes with a complication. Let us consider
the projective transformation
+ ,
(35)
(36)
(37)
10
for the transformation, i.e., simply four. In practice, this
means that we should start by assuming that the connection is not the most general which one can construct,
but satisfies some constraints.
Since the degrees of freedom that we need to fix
are four and seem to be related to the non-symmetric
part of the connection, the most obvious prescription
is to demand that S = S be equal to zero, which
was first suggested in (Sandberg, 1975) for a linear action and shown to work also for an f (R) action in
(Sotiriou and Liberati, 2007b).10 Note that this does
not mean that should vanish, but merely that
= . Imposing this constraint can easily be done
by adding a Lagrange multiplier B . The additional term
in the action will be
Z
SLM = d4 x g B S .
(38)
The action (30) with the addition of the term in eq. (38)
is, therefore, the action of the most general metric-affine
f (R) theory of gravity.
2. Field Equations
(40)
and for the Ricci tensor of an independent connection
R = R
=
(41)
+ gf (R)g
gf (R)g
g
+2f (R) g S g S + g S
S
10
= 0.
= ( B [ ] ),
(43)
(44)
The proposal of (Hehl and Kerling, 1978) to fix part of the nonmetricity, namely the Weyl vector Q , in order to break projective invariance works only when f (R) = R (Sotiriou, 2007b;
Sotiriou and Liberati, 2007b).
Taking the trace of eq. (43) over the indices and and
using eq. (44) yields
B =
2
.
3
(45)
gf (R)g
+ gf (R)g
g
2
(47)
+2f (R)g S = ( [ ] ),
3
S = 0.
(48)
Next, we examine the role of . By splitting eq. (47)
into a symmetric and an antisymmetric part and performing contractions and manipulations it can be shown
that (Sotiriou and Liberati, 2007b)
[]
= 0 S = 0.
(49)
11
On the contrary, particles with spin, such as Dirac fields,
generically have a non-vanishing hypermomentum and
will, therefore, introduce torsion. A more complicated
case is that of a perfect fluid with vanishing vorticity. If
we set torsion aside, or if we consider a fluid describing
particles that would initially not introduce any torsion
then, as for a usual perfect fluid in GR, the matter action can be written in terms of three scalars: the energy
density, the pressure, and the velocity potential (Schakel,
1996; Stone, 2000). Therefore such a fluid will lead to a
vanishing . However, complications arise when torsion is taken into account: Even though it can be argued
that the spins of the individual particles composing the
fluids will be randomly oriented, and therefore the expectation value for the spin should add up to zero, fluctuations around this value will affect spacetime (Hehl et al.,
1976; Sotiriou and Liberati, 2007b). Of course, such effects will be largely suppressed, especially in situations
in which the energy density is small, such as late time
cosmology.
It should be evident by now that, due to eq. (49), the
field equations of metric f (R) gravity reduce to eqs. (15)
and (16) and, ultimately, to the field equations of Palatini
f (R) gravity (18) and (19), for all cases in which =
0. Consequently, in vacuo, where also T = 0, they
will reduce to the Einstein equations with an effective
cosmological constant given by eq. (29), as discussed at
the end of Sec. II.B for Palatini f (R) gravity.
In conclusion, metric-affine f (R) gravity appears to
be the most general case of f (R) gravity. It includes
enriched phenomenology, such as matter-induced nonmetricity and torsion. It is worth stressing that torsion
comes quite naturally, since it is actually introduced by
particles with spin (excluding gauge fields). Remarkably,
the theory reduces to GR in vacuo or for conformally invariant types of matter, such as the electromagnetic field,
and departs from GR in the same way that Palatini f (R)
gravity does for most matter fields that are usually studied as sources of gravity. However, at the same time, it
exhibits new phenomenology in less studied cases, such
as in the presence of Dirac fields, which include torsion
and non-metricity. Finally, let us repeat once more that
Palatini f (R) gravity, despite appearances, is really a
metric theory according to the definition of (Will, 1981)
(and the geometry is a priori pseudo-Riemannian).11 On
the contrary, metric-affine f (R) gravity is not a metric
theory (hence the name). Consequently, it should also
be clear that T is not divergence-free with respect to
the covariant derivative defined with the Levi-Civita con actually). However, the physical
nection (nor with
meaning of this last statement is questionable and deserves further analysis, since in metric-affine gravity T
11
In the same way that one can make variable redefinitions in classical mechanics in order to bring an equation
describing a system to a more attractive, or easy to handle, form (and in a very similar way to changing coordinate systems), one can also perform field redefinitions in
a field theory, in order to rewrite the action or the field
equations.
There is no unique prescription for redefining the fields
of a theory. One can introduce auxiliary fields, perform
renormalizations or conformal transformations, or even
simply redefine fields to ones convenience.
It is important to mention that, at least within a classical perspective such as the one followed here, two theories are considered to be dynamically equivalent if, under a suitable redefinition of the gravitational and matter
fields, one can make their field equations coincide. The
same statement can be made at the level of the action.
Dynamically equivalent theories give exactly the same
results when describing a dynamical system which falls
within the purview of these theories. There are clear advantages in exploring the dynamical equivalence between
theories: we can use results already derived for one theory in the study of another, equivalent, theory.
The term dynamical equivalence can be considered
misleading in classical gravity. Within a classical perspective, a theory is fully described by a set of field
equations. When we are referring to gravitation theories,
these equations describe the dynamics of gravitating systems. Therefore, two dynamically equivalent theories are
actually just different representations of the same theory
(which also makes it clear that all allowed representations
can be used on an equal footing).
The issue of distinguishing between truly different theories and different representations of the same theory (or
dynamically equivalent theories) is an intricate one. It
has serious implications and has been the cause of many
misconceptions in the past, especially when conformal
transformations are used in order to redefine the fields
(e.g., the Jordan and Einstein frames in scalar-tensor theory). It goes beyond the scope of this review to present
a detailed analysis of this issue. We refer the reader to
the literature, and specifically to (Sotiriou et al., 2007)
and references therein for a detailed discussion. Here, we
simply mention that, given that they are handled carefully, field redefinitions and different representations of
the same theory are perfectly legitimate and constitute
very useful tools for understanding gravitational theories.
In what follows, we review the equivalence between
12
metric and Palatini f (R) gravity with specific theories
within the BransDicke class with a potential. It is shown
that these versions of f (R) gravity are nothing but different representations of BransDicke theory with Brans
Dicke parameter 0 = 0 and 0 = 3/2, respectively. We
comment on this equivalence and on whether preference
to a specific representation should be an issue. Finally,
we use this equivalence to perform a classification of f (R)
gravity.
A. Metric formalism
1
d4 x g f (R) + SM (g , ).
(50)
Smet =
2
One can introduce a new field and write the dynamically equivalent action
Z
1
d4 x g [f () + f ()(R )] +
Smet =
2
+SM (g , ).
(51)
Variation with respect to leads to the equation
f ()(R ) = 0.
(52)
(53)
1
Smet =
d4 x g [R V ()] + SM (g , ). (54)
2
12
1
T
g V ()
2
1
+ ( g ) ,
R = V ().
G =
(55)
(56)
dV
= T.
d
(57)
(58)
d =
,
(59)
2
13
a scalar-tensor theory is mapped into the Einstein frame
in which the new scalar field couples minimally to
the Ricci curvature and has canonical kinetic energy, as
described by the gravitational action
"
#
Z
p
R
1
(g)
4
. (60)
g
S = d x
U ()
2 2
For the 0 = 0 equivalent of metric f (R) gravity we have
f (R) = e
2
3
= Rf (R) f (R) ,
U ()
2
2 (f (R))
(61)
(62)
p
1
R
+
U ()
Smet
=
d4 x
g
2 2
B. Palatini formalism
1
Spal =
d4 x g f (R) + SM (g , ),
(64)
2
and following exactly the same steps as before, i.e., introducing a scalar field which we later redefine in terms
of , yields
Z
1
d4 x g [R V ()] + SM (g , ). (65)
Spal =
2
13
This has been an issue of debate and confusion, see for example
the references in (Faraoni and Nadeau, 2007).
Even though the gravitational part of this action is formally the same as that of the action (54), this action is
not a BransDicke one with 0 = 0: R is not the Ricci
scalar of the metric g . However, we have already seen
that the field equation (18) can be solved algebraically
for the independent connection yielding eq. (25). This
implies that we can replace the connection in the action without affecting the dynamics of the theory (the
independent connection is practically an auxiliary field).
Alternatively, we can directly use eq. (27), which relates
R and R. Therefore, the action (65) can be rewritten,
modulo surface terms, as
Z
1
3
Spal =
V ()
d4 x g R +
2
2
+SM (g , ).
(66)
This is the action of a BransDicke theory with Brans
Dicke parameter 0 = 3/2. The corresponding field
equations obtained from the action (66) through variation with respect to the metric and the scalar are
1
3
T 2 g +
2
2
V
1
g ,
(67)
+ ( g )
1
= (R V ) +
.
(68)
3
2
G =
(69)
Finally, one can also perform the conformal transformation (58) in order to rewrite the action (66) in the
Einstein frame. The result is
#
"
Z
p
R
4
U () + SM (1 g , ),
Spal = d x
g
2
(70)
where U () = V ()/(2 2 ). Note that here we have not
used any redefinition for the scalar.
To conclude, we have established that Palatini f (R)
gravity can be cast into the form of an 0 = 3/2 Brans
Dicke theory with a potential.
C. Classification
14
one extra scalar degree of freedom with respect to GR.
The absence of a kinetic term for the scalar in the action (54) or in eq. (56) should not mislead us to think
that this degree of freedom does not carry dynamics.
As can be seen by eq. (57), is dynamically related to
the matter fields and, therefore, it is a dynamical degree of freedom. Of course, one should also not fail to
mention that eq. (56) does constrain the dynamics of .
In this sense metric f (R) gravity and 0 = 0 Brans
Dicke theory differs from the general BransDicke theories and constitutes a special case. On the other hand,
in the 0 = 3/2 case which corresponds to Palatini
f (R) gravity, the scalar appears to have dynamics in
the action (66) or in eq. (68). However, once again this is
misleading since, as is clear from eq. (69), is in this case
algebraically related to the matter and, therefore, carries
no dynamics of its own [indeed the field eqs. (67) and (69)
could be combined to give eq. (28), eliminating completely]. As a remark, let us state that the equivalence
between Palatini f (R) gravity and 0 = 3/2 Brans
Dicke theory and the clarifications just made highlight
two issues already mentioned: the fact that Palatini f (R)
gravity is a metric theory according to the definition of
(Will, 1981), and the fact that the independent connection is actually some sort of auxiliary field.
The fact that the dynamics of are not transparent at
the level of the action in both cases should not come as
a big surprise: is coupled to the derivatives of the metric (through the coupling with R) and, therefore, partial
integrations to free or g during the variation are
bound to generate dynamical terms even if they are not
and
i g
tiiii
METRIC-AFFINE f (R)
f (R) GRAVITY
JJ
i
JJ
iiii
JJ
independent
JJ
JJ
= J
J
JJ
JJ
JJ
JJ
$
METRIC f (R)
KS
i
iiii
SM = SM (g , )
PALATINI f (R)
UUUU
KS
U
f (R) = R
UU
f (R) 6= 0
UUUU
UU*
BRANSDICKE, 0 = 23
GR
f (R) = R
iiii
i
i
i
ti
f (R) 6= 0
BRANSDICKE, 0 = 0
FIG. 1 Classification of f (R) theories of gravity and equivalent BransDicke theories. The flowchart shows the list of assumptions that are needed to arrive to the various versions of f (R) gravity and GR beginning from the the general f (R) action. It
also includes the equivalent BransDicke classes. Taken from (Sotiriou, 2006b).
15
work with the BransDicke one, and second, why, since
we know a lot about BransDicke theory, should we regard f (R) gravity as unexplored or interesting?
The answer to the first question is quite straightforward. There is actually no reason to prefer either of the
two representations at least as far as classical gravity is concerned. There can be applications where the
f (R) representation can be more convenient and applications where the BransDicke representation is more convenient. One should probably mention that habit affects
our taste and, therefore, an f (R) representation seems
more appealing to relativists due to its more apparent
geometrical nature, whereas the BransDicke representation seems more appealing to particle physicists. This
issue can have theoretical implications. To give an example: if f (R) gravity is considered as a step towards a more
complicated theory, which generalisation would be more
straightforward will depend on the chosen representation
[see also (Sotiriou et al., 2007) for a discussion].
Whether f (R) theories of gravity are unexplored and
interesting or just an already-studied subcase of Brans
Dicke theory, is a more practical question that certainly
deserves a direct answer. It is indeed true that scalartensor theories and, more precisely, BransDicke theory are well-studied theories which have been extensively
used in many applications, including cosmology. However, the specific choices 0 = 0, 3/2 for the Brans
Dicke parameter are quite exceptional, as already mentioned in the previous section. It is also worthwhile pointing out the following: a) As far as the 0 = 0 case is
concerned, one can probably speculate that it is the apparent absence of the kinetic term for the scalar in the
action which did not seem appealing and prevented the
study of this theory. b) The 0 = 3/2 case leads to a
conformally invariant theory in the absence of the potential [see (Sotiriou, 2006b) and references therein], which
constituted the initial form of BransDicke theory, and
hence it was considered non-viable (a coupling with nonconformally invariant matter is not feasible). However, in
the presence of a potential, the theory no longer has this
feature. Additionally, most calculations which are done
for a general value of 0 in the literature actually exclude
0 = 3/2, mainly because, merely for simplicity purposes, they are done in such a way that the combination
20 + 3 appears in a denominator (see also Sec. V.A).
In any case, the conclusion is that the theories in the
BransDicke class that correspond to metric and Palatini
f (R) gravity had not yet been explored before the recent
re-introduction of f (R) gravity and, as will also become
clear later, several of their special characteristics when
compared with more standard BransDicke theories were
revealed through studies of f (R) gravity.
We now turn our attention to cosmology, which motivated the recent surge of interest in f (R) gravity in order
was
proved
in
(Maartens and Taylor,
1994;
Taylor and Maartens, 1995) and the generalization to
arbitrary metric f (R) gravity was given by (Rippl et al.,
1996). The validity of the EGS theorem can also be seen
through the equivalence between f (R) and BransDicke
theory: the theorem was extended to scalar-tensor
theories in (Clarkson et al., 2001, 2003). Since metric
and Palatini f (R) gravities are equivalent to = 0
and 0 = 3/2 BransDicke theories respectively, it
seems that the results of (Clarkson et al., 2001, 2003)
can be considered as straightforward generalizations
of the EGS theorem in both versions of f (R) gravity
16
as well. However, in the case of Palatini f (R) gravity
there is still some doubt regarding this issue due to
complications in averaging (Flanagan, 2004b).
1. Metric f (R) gravity
(73)
where u denotes the four-velocity of an observer comoving with the fluid and and P are the energy density and
pressure of the fluid, respectively.
Note that the value of k is an external parameter. As
in many other works in the literature, in what follows we
choose k = 0, i.e., we focus on a spatially flat universe.
This choice in made in order to simplify the equations and
should be viewed sceptically. It is sometimes claimed in
the literature that such a choice is favoured by the data.
However, this is not entirely correct. Even though the
data [e.g. (Spergel et al., 2007)] indicate that the current
value of k is very close to zero, it should be stressed that
this does not really reveal the value of k itself. Since
k =
k
,
a2 H 2
Rf f
, (75)
H2 =
3H
Rf
3f
2
h
2
2 f + 2H Rf
2H + 3H = P + (R)
f
i
+ 1 (f Rf ) .
+Rf
(76)
2
With some hindsight, we assume that f > 0 in order to have a positive effective gravitational coupling
and f > 0 to avoid the Dolgov-Kawasaki instability
(Dolgov and Kawasaki, 2003a; Faraoni, 2006a) discussed
in Sec. V.B.
A significant part of the motivation for f (R) gravity
is that it can lead to accelerated expansion without the
need for dark energy (or an inflaton field). An easy way
to see this is to define an effective energy density and
pressure of the geometry as
ef f =
Pef f =
(77)
1
2
(f Rf )
, (78)
(74)
Rf f
3H Rf
2f
f
+ Rf
+
R 2 f + 2H Rf
ef f ,
3
a
= [ef f + 3Pef f ] .
a
6
H2 =
(79)
(80)
14
+ Rf
+ 1 (f Rf )
R 2 f + 2H Rf
Pef f
2
=
.
Rf f
ef f
3H
Rf
2
(81)
17
Since the denominator on the right hand side of eq. (81)
is strictly positive, the sign of wef f is determined by its
numerator. In general, for a metric f (R) model to mimic
the de Sitter equation of state wef f = 1, it must be
RH
f
=
.
f
(R)
(82)
6n2 7n 1
6n2 9n + 3
(83)
2n2 + 3n 1
.
n2
(84)
2(n + 2)
.
3(2n + 1)(n + 1)
(85)
15
ef f + Pef f
"
H
d
ln
=
=
dt
!#
(87)
18
2. Palatini f (R) gravity
As already mentioned, some concerns have been expressed on whether the homogeneity approximation can
justify the use of the FLRW metric as a cosmological solution in Palatini f (R) gravity (Flanagan, 2004b)
[see also (Li et al., 2008)]. Therefore, even though it is
standard practice in the literature to assume a FLRW
background and a perfect fluid description for matter when studying cosmology in Palatini f (R) gravity
[e.g. (Allemandi et al., 2004, 2005a; Amarzguioui et al.,
2006; Meng and Wang, 2003, 2004b, 2005; Sotiriou,
2006a,e; Vollick, 2003)], and we are going to review this
approach here, the reader should approach it with some
reasonable skepticism until this issue is clarified further.
Under the assumptions that the spacetime is indeed
described at cosmological scales by the FLRW metric,
eq. (72), that the stress-energy tensor of matter is that of
eq. (73), and that k = 0, easy manipulations reveal that
the field eqs. (18) and (19) yield the following modified
Friedmann equation [see for instance (Meng and Wang,
2003; Sotiriou, 2006e)]:
1 f
H+
2 f
!2
1 ( + 3P ) 1 f
+
,
6
f
6 f
(88)
(89)
1
2 + Rf f
(Rf 2f ) 2
6f
1 23 ff (Rf
f )
(90)
Considering now that, due to eq. (20), R is just an algebraic function of m , it is easy to realize that eq. (90) is
actually just the usual Friedmann equation with a modified source. The functional form of f will determine how
the dynamics will be affected by this modification.
It seems, therefore, quite intuitive that by tampering
with the function f one can affect the cosmological dynamics in a prescribed way. Indeed it has been shown
that for f (R) = R 2 /(3R) one approaches a de Sitter expansion as the density goes to zero (Vollick, 2003).
In order to match observations of the expansion history,
one needs to choose 1067 (eV)2 1053 m2 . Additionally, in regimes for which , eq. (90) reduces
to high precision to the standard Friedmann equation.
The above can very easily be verified by replacing this
particular choice of f in eq. (90). We refer the reader to
the literature for more details.
One could, of course, consider more general functions
of R. Of particular interest would be having positive powers of R higher than the first power added in
the action (since one could think of the Lagrangian as
a series expansion). Indeed this has been considered
(Meng and Wang, 2003, 2004b, 2005; Sotiriou, 2006a,e).
However, it can be shown that such terms do not really lead to interesting phenomenology as in metric f (R)
gravity: for instance they cannot drive inflation as, unlike in the scenario proposed by (Starobinsky, 1980) in
the metric formalism, here there are no extra dynamics and inflation cannot end gracefully (Meng and Wang,
2004b; Sotiriou, 2006a). As a matter of fact, it is more
likely that positive powers of R will lead to no interesting
cosmological phenomenology unless their coefficients are
large enough to make the models non-viable (Sotiriou,
2006a).
B. Cosmological eras
19
to follow each other, including:
1. early inflation
2. a radiation era during which Big Bang Nucleosynthesis occurs;
3. a matter era;
4. the present accelerated epoch, and
5. a future era.
Big Bang Nucleosynthesis is well constrained
see
(Brookfield et al.,
2006;
Clifton and Barrow,
2005a; Evans et al., 2007; Kneller and Steigman,
2004; Lambiase and Scarpetta, 2006; Nakamura et al.,
2006) for such constraints on f (R) models. The matter
era must last long enough to allow the primordial
density perturbations generated during inflation to grow
and become the structures observed in thd universe
today. The future era is usually found to be a de Sitter
attractor solution, or to be truncated at a finite time by
a Big Rip singularity.
Furthermore, there must be smooth transitions between consecutive eras, which may not happen in all
f (R) models. In particular, the exit from the radiation era has been studied and claimed to originate problems for many forms of f (R) in the metric formalism, including f = R 2(n+1) /Rn , n >
0 (Amendola et al., 2007a,b; Brookfield et al., 2006;
Capozziello et al., 2006c; Nojiri and Odintsov, 2006)
[but not in the Palatini formalism (Carvalho et al., 2008;
Fay et al., 2007b)]. However, the usual model f (R) =
R 4 /R with bad behaviour was studied using singular perturbation methods (Evans et al., 2007), definitely
finding a matter era which is also sufficiently long.
Moreover, one can always find choices of the function f (R) with the correct cosmological dynamics in
the following way: one can prescribe the desired
form of the scale factor a(t) and integrate a differential equation for f (R) that produces the desired scale factor (Capozziello et al., 2005b, 2006c;
de la Cruz-Dombriz and Dobado, 2006; Faulkner et al.,
2007; Fay et al., 2007a,b; Hu and Sawicki, 2007a,b;
Multamaki and Vilja, 2006a; Nojiri and Odintsov, 2006,
2007b,c; Song et al., 2007). In general, this designer
f (R) gravity produces forms of the function f (R) that
are rather contrived. Moreover, the prescribed evolution
of the scale factor a(t) does not determine uniquely the
form of f (R) but, at best, only a class of f (R) models (Multamaki and Vilja, 2006a; Sokolowski, 2007b,c;
Starobinsky, 2007). Therefore, the observational data
providing information on the history of a(t) are not sufficient to reconstruct f (R): one needs additional information, which may come from cosmological density perturbations. There remains a caveat on being careful to
terminate the radiation era and allowing a matter era
that is sufficiently long for scalar perturbations to grow.
20
2006; Song et al., 2007; Stabenau and Jain, 2006;
Tsujikawa,
2007;
White and Kochanek,
2001;
Zhang et al., 2007).
This originated various efforts
to constrain f (R) gravity with cosmic microwave
background data (Amendola and Tsujikawa, 2008;
Appleby and Battye, 2007; Carloni et al., 2008b;
Hu and Sawicki, 2007b; Li and Barrow, 2007; Li et al.,
2007a; Pogosian and Silvestri, 2008; Starobinsky, 2007;
Tsujikawa, 2008; Tsujikawa et al., 2008; Wei and Zhang,
2008).
Most of these works are restricted to specific choices
of the function f (R), but a few general results have
also been obtained. The growth and evolution of local scalar perturbations, which depends on the theory of
gravity employed, was studied in metric f (R) gravity theories which reproduce GR at high curvatures in various
papers (Carroll et al., 2006; de la Cruz-Dombriz et al.,
2008; Song et al., 2007) by assuming a scale factor evolution typical of a CDM model. Vector and tensor
modes are unaffected by f (R) corrections. It is found
that f (R) > 0 is required for the stability of scalar perturbations (Song et al., 2007), which matches the analysis of Sec. (V.B.2) in a locally de Sitter background.
The corrections to the EinsteinHilbert action produce
qualitative differences with respect to Einstein gravity:
they lower the large angle anisotropy of the cosmic microwave background and may help explaining the observed low quadrupole; and they produce different correlations between the cosmic microwave background and
galaxy surveys (Song et al., 2007). Further studies challenge the viability of f (R) gravity in comparison with the
CDM model: in (Bean et al., 2007) it is found that large
scale density fluctuations are suppressed in comparison
to small scales by an amount incompatible with the observational data. This makes it impossible to fit simultaneously large scale data from the cosmic microwave background and small scale data from galaxy surveys. Also,
a quasi-static approximation used in a previous analysis
(Zhang, 2007) is found to be invalid.
In (de la Cruz-Dombriz et al., 2008), the growth of
matter density perturbations is studied in the longitudinal gauge using a fourth order equation for the density contrast /, which reduces to a second order one
for sub-horizon modes. The quasi-static approximation,
which does not hold for general forms of the function
f (R), is however found to be valid for those forms of this
function that describe successfully the present cosmic acceleration and pass the Solar System tests in the weakfield limit. It is interesting that the relation between the
gravitational potentials in the metric which are responsible for gravitational lensing, and the matter overdensities
depends on the theory of gravity; a study of this relation
in f (R) gravity (as well as in other gravitational theories)
is contained in (Zhang et al., 2007).
Cosmological density perturbations in the Palatini
formalism have been studied in (Amarzguioui et al.,
2006; Carroll et al., 2006; Koivisto, 2006b, 2007;
Koivisto and Kurki-Suonio, 2006; Lee, 2007, 2008;
In addition to having the correct cosmological dynamics and the correct evolution of cosmological perturbations, the following criteria must be satisfied in order for
an f (R) model to be theoretically consistent and compatible with experiment. The model must:
have the correct weak-field limit at both the Newtonian and post-Newtonian levels, i.e., one that is
compatible with the available Solar System experiments;
be stable at the classical and semiclassical level (the
checks performed include the study of a matter instability, of gravitational instabilities for de Sitter
space, and of a semiclassical instability with respect
to black hole nucleation);
not contain ghost fields;
admit a well-posed Cauchy problem;
These independent requirements are discussed separately in the following.
A. Weak-field limit
21
is non-dynamical in the Palatini case. Let us, therefore,
consider the role of the scalar field in the metric formalism as it will turn out to be crucial for the weak-field
limit. Using the notations of Sec. III.A, the action is
given by eq. (54) and the corresponding field equations
by eq. (55).
Equation (52) for has no dynamical content because
it only enforces the equality = R. However, = R is
indeed a dynamical field that satisfies the wave equation,
3f () + 3f ()
+ f () 2f () = T.
(91)
(92)
1
(93)
S=
d4 x g [R V ()] + S (m) ,
2
where V () is given by eq. (53).
Now one may think of studying the dynamics and stability of the model by looking at the shape and extrema
of the effective potential V () but this would be misleading because the dynamics of are not regulated by
V () (indeed, the wave equation (91) does not contain
V ), but are subject the strong constraint = R, and R
(or f (R)) is ruled by the trace equation (8).
The following example shows how the use of the potential V () can be misleading. As is well known, the effective mass of a scalar field (corresponding to the second
derivative of the potential evaluated at the minimum)
controls the range of the force mediated by this field.
Thus, when studying the weak-field limit of the theory it
is important to know the range of the dynamical scalar
field = R present in the metric formalism in addition
to the metric field g , as this field can potentially violate the post-Newtonian constraints obtained from Solar
System experiments if the scalar field gives observable effects at the relevant scales. One way to avoid Solar System constraints, however, is to have have a sufficiently
short range (see Sec. V.A.2 for more details). Consider
the example f (R) = R + aR2 , with a a positive constant.
By naively taking the potential, one obtains
V () = a 2
m21 2
R
T
=
,
6a
6a
(95)
1
6a
(96)
R
0
3 f0
This equation coincides with eq. (6) of (Muller et al.,
1990), with eq. (26) of (Olmo, 2007), and with eq. (17) of
(Navarro and Van Acoleyen, 2007). It also appears in a
calculation of the propagator for f (R) gravity in a locally
flat background (eq. (8) of (Nunez and Solganik, 2004)).
The same expression is recovered in a gauge-invariant
stability analysis of de Sitter space (Faraoni and Nadeau,
2005) reported in Sec. V.B.2 below.
Another possibility is to consider the field f (R)
instead of = R, and to define the effective mass of
by using the Einstein frame scalar-tensor analog of
f (R) gravity instead of its Jordan frame cousin already
discussed (Chiba, 2003). By performing the conformal
transformation
g g = f (R) g g
(98)
(94)
16
17
22
the Ricci curvature and has canonical kinetic energy, as
described by the action
Z
p
1
(g)
4
g R U () +
S
=
d x
2
(100)
+SM (e 2/3 g , ),
(note once more the non-minimal coupling of the matter
in the Einstein frame). For the 0 = 0 equivalent of
metric f (R) gravity we have
2
f (R) = e 3 ,
(101)
Rf
(R)
f
(R)
=
U ()
(102)
2 ,
2 (f (R))
q
3 f
By using d/d
where R = R().
= 2
f , the effective mass of is defined by
m
2ef f
#
"
d2 U
4f
1 1
+
=
2
3 f
f
d2
(f )
(103)
R
.
(104)
0
3f0 f0
f0
In the Einstein frame, it is not the mass m
of a particle or
a field that is measurable, but rather the ratio m/
m
u between m
and the Einstein frame unit of mass m
u , which
1/2
is varying, scaling as m
u = [f (R)]
mu = 1/2 mu ,
where mu is the constant unit of mass in the Jordan frame
(Dicke, 1962; Faraoni et al., 1999; Faraoni and Nadeau,
2007). Therefore,
m
2ef f
m2ef f
=
.
m
2u
m2u
(105)
By using the equivalence between f (R) and scalartensor gravity, Chiba originally suggested that all f (R)
theories are ruled out (Chiba, 2003). This claim was
based on the fact that metric f (R) gravity is equivalent
to an 0 = 0 BransDicke theory, while the observational
constraint is |0 | > 40000 (Bertotti et al., 2003). This
is not quite the case and the weak-field limit is more
subtle than it appears, as the discussion of the previous section might have already revealed: The value of
the parametrized post-Newtonian (PPN) parameter ,
on which the observational bounds are directly applicable, is practically independent of the mass of the scalar
only when the latter is small (Wagoner, 1970). In this
case, the constraints on can indeed be turned into constraints on 0 . However, if the mass of this scalar is
large, it dominates over 0 in the expression of and
drives its value to unity. The physical explanation of
this fact, as mentioned previously, is that the scalar becomes short-ranged and, therefore, has no effect at Solar
System scales. Additionally, there is even the possibility
that the effective mass of the scalar field itself is actually
scale-dependent. In this case, the scalar may acquire a
large effective mass at terrestrial and Solar System scales,
shielding it from experiments performed there while being effectively light at cosmological scales. This is the
chameleon mechanism, well-known in quintessence models (Khoury and Weltman, 2004a,b).
23
ing light deflection and other Solar System experiments18
(Barrow and Clifton, 2006; Clifton and Barrow, 2005a,
2006; Zakharov et al., 2006). Only later was the case
of a general function f (R) discussed (Chiba et al., 2007;
Jin et al., 2006; Olmo, 2007). Chibas result based on the
scalar-tensor equivalence eventually turns out to be valid
subject to certain assumptions which are not always satisfied (Chiba et al., 2007; Jin et al., 2006; Olmo, 2007)
see below. This method, however, does not apply to
the Palatini version of f (R) gravity.
In what follows we adhere to, but streamline, the discussion of (Chiba et al., 2007) with minor modifications,
in order to compute the PPN parameter for metric f (R)
gravity [see also (Olmo, 2007)]. We consider a spherically symmetric, static, non-compact body embedded in
a background de Sitter universe; the latter can exist in
an adiabatic approximation in which the evolution of the
universe is very slow in comparison with local dynamics.
The condition for the existence of a de Sitter space with
R = R0 g /4 and constant curvature R0 = 12H02 is
s
f0
.
(106)
f0 R0 2f0 = 0,
H0 =
6f0
The line element is
ds2 = 1 + 2(r) H02 r2 dt2
+ 1 + 2(r) + H02 r2 dr2 + r2 d2 (107)
r << H01 ,
(109)
18
19
20
(110)
,
3f0
(111)
where
m2 =
f0 f0 R0
.
3f0
(112)
(108)
and
R(r) = R0 + R1 (r),
m2 =
(f0 ) 2f0 f0
.
3f0 f0
(113)
This equation is found in various other treatments of perturbations of de Sitter space (Faraoni and Nadeau, 2005;
Navarro and Van Acoleyen, 2007; Nunez and Solganik,
2004; Olmo, 2007).
Assumption 2 that the scalar R1 is light, which enables
the f (R) theory to produce significant cosmological effects at late times, also allows one to neglect21 the term
m2 R1 in eq. (111). The Green function of the equation
1
2 R1 = 3f
is then G(r) = 4r and the solution is
0
R
)
R1 d3 ~x 3f(r
G(r r ), which yields
R1
21
M
12f0 r
(mr << 1) .
(114)
24
Now, the condition m2 r2 << 1 yields
1 f0
R0 r2 << 1
3 f0
and, using H0 r << 1,
f0 2
r << 1.
f
0
(115)
f
f0 R11 3H02 0 R1 f0 1 1 R1
2
+ f0 R1 R11 + f0 R1 = T11 (125)
(116)
Let us use now the full field equations (6); by expanding f (R) and f (R) and using f0 = 6H02 f0 we get
f
f0 R1 + f0 R 3H02 0 R1
2
f0 R1 + f0 R1 R = T . (117)
f0
R1 f0 2 R1 = .
2
(119)
Recalling that 2 R1 3f
for mr << 1, one obtains
0
f0 2 (r) =
2 f0
R1 .
3
2
(120)
M
2,
=
2
dr
6f0 48f0 r
r
(121)
Rr
where M (r) = 4 0 0 dr (r )2 (r ). The integration constant C1 must be set to zero to guarantee regularity of
the Newtonian potential at r = 0. The potential (r)
then becomes
(r) =
M
M
r.
6f0 r
48f0
(122)
and
(r)
M
.
6f0 r
(124)
R11 3H02
(126)
(127)
f0 2 +
0
+ 0
0. (128)
dr
r dr
2
r dr
Now, using eq. (114) for R1 , one concludes that the third
term in eq. (128) is negligible in comparison with the
fourth term. In fact,
f0 R1 f
(129)
2f 2dR 0 r2 << 1.
0 1
f0
r
dr
Then, using again the expression (114) for dR1 /dr and
eq. (124) for (r), one obtains
d
M
,
=
dr
12f0 r
(130)
M
.
12f0 r
(131)
(r)
1
= .
(r)
2
(132)
25
breaks down. It is important to ascertain whether these
are physically relevant situations. There are three main
cases to consider.
The case of non-analytic f (R): While (Chiba et al.,
2007) consider functions f (R) that are analytic at the
background value R0 of the Ricci curvature, the situation in which this function is not analytical has been
contemplated briefly in (Jin et al., 2006). Assuming that
f (R) has an isolated singularity at R = Rs , it can be
expressed as the sum of a Laurent series,
f (R) =
+
X
n=0
an (R Rs ) .
(133)
65n
2
10 2(1n)
3
H0
n(1 n)
(135)
is obtained (Faulkner et al., 2007). Fifth force experiments give the bounds
1
2(1n)
212n
2
10 1n .
1n
H0
n(1 n)
(136)
Early works on the weak-field limit of Palatini f (R) gravity often led to contradictory results and to several technical problems as well
(Allemandi et al.,
2005b;
Allemandi and Ruggiero,
2007; Barraco and Hamity, 2000; Bustelo and Barraco,
2007; Dominguez and Barraco, 2004; Kainulainen et al.,
2007b; Meng and Wang, 2004a; Olmo, 2005a,b, 2007;
Ruggiero, 2007; Ruggiero and Iorio, 2007; Sotiriou,
2006c) which seem to have been clarified by now.
26
First of all, there seems to have been some confusion
in the literature about the fact that Palatini f (R) gravity reduces to GR with a cosmological constant in vacuum and the consequences that this can have on the
weak-field limit and Solar System tests. It is, of course,
true (see Sec. II.B) that in vacuo Palatini f (R) gravity will have the same solutions of GR plus a cosmological constant and, therefore, the Schwarzschild-(anti)de Sitter solution will be the unique vacuum spherically
symmetric solution (see also Sec. VI.C.1 for a discussion
of the Jebsen-Birkhoff theorem). This was interpreted
in (Allemandi and Ruggiero, 2007; Ruggiero and Iorio,
2007) as an indication that the only parameter that can
be constrained is the effective cosmological constant and,
therefore, models that are cosmologically interesting (for
which this parameter is very small) trivially satisfy Solar
System tests. However, even if one sets aside the fact that
a weak gravity regime is possible inside matter as well,
such claims cannot be correct: they would completely
defeat the purpose of performing a parametrized postNewtonian expansion for any theory for which one can
establish uniqueness of a spherically symmetric solution,
as in this case we would be able to judge Solar System
viability just by considering this vacuum solution (which
would be much simpler).
Indeed, the existence of a spherically symmetric vacuum solution, irrespective of its uniqueness, does not suffice to guarantee a good Newtonian limit. For instance,
the Schwarzschild-de Sitter solution has two free parameters; one of them can be associated with the effective
cosmological constant in a straightforward manner (using
the asymptotics). However, it is not clear how the second parameter, which in GR is identified with the mass
of the object in the Newtonian regime, is related to the
internal structure of the object in Palatini f (R) gravity.
The assumption that it represents the mass defined in
the usual way is not, of course, sufficient. One would
have to actually match the exterior solution to a solution
describing the interior of the Sun within the realm of
the theory in order to express the undetermined parameter in the exterior solution in terms of known physical
quantities, such as Newtons constant and the Newtonian
mass. The essence of the derivation of the Newtonian
limit of the theory consists also in deriving such an explicit relation for this quantity and showing that it agrees
with the Newtonian expression. The parametrized postNewtonian expansion is nothing but an alternative way
to do that without having to solve the full field equations. Therefore, it is clear that more information than
the form of the vacuum solution is needed in order to
check whether the theory can satisfy the Solar System
constraints.
However, some early attempts towards a Newtonian and post-Newtonian expansion were also flawed.
In (Meng and Wang, 2004a) and (Barraco and Hamity,
2000) for instance, a series expansion around a de Sitter
background was performed in order to derive the Newto-
(137)
22
,
R
(138)
(141)
322
T
(143)
27
The situation does not improve even with the unphysical branch of eq. (142) with anegative sign. In fact, in
this case, R1 2 [32 /( T ) + 3] and the correction to
the background curvature is of the order 2 and not much
smaller than that, as it would be required in order to
truncate the expansion (140). In (Barraco and Hamity,
2000), this fact was overlooked and only linear terms in
R1 were kept in the expansion of f (R) and f (R) around
R0 . In (Meng and Wang, 2004a), even though this fact
is noticed in the final stages of the analysis and is actually used, the authors do not take it into account properly
from the outset, keeping again only linear terms [see, e.g.,
eq. (11) of (Meng and Wang, 2004a)].
However, the algebraic dependence of R on the density
does not only signal a problem for the approaches just
mentioned. It actually implies that the outcome of the
post-Newtonian expansion itself depends on the density,
as shown in (Olmo, 2005a,b; Sotiriou, 2006c). Consider
for instance, along the lines of (Sotiriou, 2006a,c), the
conformal metric
h = f (R)g
(144)
2Geff M
V0 2
r + (T ),
(148)
+
r
60
V0 2
2Geff M
(1)
r (T ) ij , (149)
M MV
,
M + MV
(151)
R
where MV 1 0 d3 ~x [V0 /0 V ()/].
As stated in different words in (Olmo,
R 2005b), if the
Newtonian mass is defined as MN d3 ~x (t, ~x ), the
requirement that a theory has a good Newtonian limit
is that Geff M equals GMN , where N denotes Newtonian, and 1 to very high precision. Additionally, the
second term on the right hand side of both eqs. (148)
and (149) should be negligible, since it plays the role of
a cosmological constant term. (T ) should also be small
and have a negligible dependence on T .
Even though it is not impossible, as mentioned before,
to prescribe f such that all of the above are satisfied for
some range of densities within matter (Sotiriou, 2006c),
this does not seem possible over the wide range of densities relevant for the Solar System tests. As a matter
of fact, is nothing but an algebraic function of T and,
therefore, of the density (since is an algebraic function of R). The presence of the (T ) term in eqs. (148)
and (149) signals an algebraic dependence of the postNewtonian metric on the density. This direct dependence
of the metric on the matter field is not only surprising but
also seriously problematic. Besides the fact that it is evident that the theory cannot have the proper Newtonian
limit for all densities (the range of densities for which it
will fail depends on the functional form of f ), consider
the following: What happens to the post-Newtonian metric if a very weak point source (approximated by a delta
function) is taken into account as a perturbation? And
will the post-Newtonian metric be continuous when going
from the interior of a source to the exterior, as it should?
28
We will refrain from further analysis of these issues
here, since evidence coming from considerations different than the post-Newtonian limit, which we will review
shortly, will be of significant help. We will, therefore,
return to this discussion in Sec. VI.C.2.
B. Stability issues
In the metric formalism, Dolgov and Kawasaki discovered an instability in the prototype model f (R) =
R 4 /R (now called Dolgov-Kawasaki, or Ricci
scalar or matter instability), which manifests itself on
an extremely short time scale and is sufficient to rule out
this model (Dolgov and Kawasaki, 2003a). Their result
was confirmed in (Nojiri and Odintsov, 2003a, 2004b), in
which it was also shown that adding to this specific f (R)
an R2 term removes this instability. The instability was
rediscovered in (Baghram et al., 2007) for a specific form
of the function f (R). The analysis of this instability is
generalized to arbitrary f (R) theories in the metric formalism in the following way (Faraoni, 2006a).
We parametrize the deviations from Einstein gravity
as
f (R) = R + (R),
(152)
(153)
and evaluating f ,
( 1)
T
2
R
+
R=
+
.
3
3
3
(154)
We assume that 6= 0: if = 0 on an interval then the
theory reduces to GR. Isolated zeros of , at which the
theory is instantaneously GR, are in principle possible
but will not be considered here.
Consider a small region of spacetime in the weak-field
regime and approximate locally the metric and the curvature by
R +
g = + h ,
R = T + R1 ,
(155)
1 2 R1 2 T R 1 + 2 T
~ R
~ 1
R
( T + 2)
1
1
,
R1 = T 2 T
+
3
3
(156)
~ and 2 are the gradient and Laplacian in
where
Euclidean three-dimensional space, respectively, and an
overdot denotes differentiation with respect to time. The
function and its derivatives are now evaluated at R =
T . The coefficient of R1 in the fifth term on the left
hand side is the square of an effective mass and is domi1
nated by the term (3 ) due to the extremely small
value of needed for these theories to reproduce the correct cosmological dynamics. Then, the scalar mode R1 of
the f (R) theory is stable if = f > 0, and unstable if
this effective mass is negative, i.e., if = f < 0. The
time scale for this instability to manifest is estimated to
be of the order of the inverse effective mass 1026 s
in the example (R) = 4 /R (Dolgov and Kawasaki,
2003a). The small value of gives a large effective mass
and is responsible for the small time scale over which the
instability develops.
Let us consider now matter with vanishing trace T of
the stress-energy tensor. In this case eq. (156) becomes
2
1 + R1 2 2 R1
~ 1
R
R
2
1
1
+
.
(157)
R1 =
3
29
1
(158)
22
V () ,
2
2
(159)
incorporating both scalar-tensor gravity (if f (, R) =
()R) and modified gravity (if the scalar field is absent and fRR 6= 0). In a spatially flat FLRW universe
the vacuum field equations assume the form
1
2 Rf
f
2
H =
+
+ V 3H f ,(160)
3f 2
2
2
1
(161)
H = 2 + F H f ,
2f
1 d 2 f
dV
+ 3H +
= 0, (162)
+2
2 d
d
where f f /, F f /R, and an overdot denotes
differentiation with respect to t. We choose (H, ) as
dynamical variables; then, the stationary points of the
dynamical system (160)-(162) are de Sitter spaces with
30
constant scalar field (H0 , 0 ). The conditions for these
de Sitter solutions to exist are
6H02 f0 f0 + 2V0 = 0,
dV
df
2
= 0,
d 0
d 0
(163)
(164)
where f0 f (0 , R0 ), f0 f (0 , R0 ), V0 V (0 ),
and R0 = 12H02 . The phase space is a curved twodimensional surface embedded in a three-dimensional
space (de Souza and Faraoni, 2007).
Inhomogeneous perturbations of de Sitter space have
been studied using the covariant and gauge-invariant
formalism of (Bardeen, 1980; Ellis and Bruni, 1989;
Ellis et al., 1990, 1989) in a version provided by (Hwang,
1990a,b, 1997, 1998; Hwang and Noh, 1996) for generalized gravity. The metric perturbations are defined by
g00 = a2 (1 + 2AY ) , g0i = a2 B Yi , (165)
gij = a2 [hij (1 + 2HL Y ) + 2HT Yij ] .
(166)
Here the Y are scalar spherical harmonics, hij is the
i is
three-dimensional metric of the FLRW background,
the covariant derivative of hij , and k is the eigenvalue of
i Y = k 2 Y . The Yi and Yij are vector and tensor
i
harmonics satisfying
1
Yi =
i Y,
k
Yij =
1
1
i j Y + Y hij ,
2
k
3
(167)
a
a
B H T ,
k
k
(170)
H + 3H0 H +
4H0 + H = 0, (171)
a2
3f0
(173)
4
.
R
(174)
23
The generalization of the condition (173) to D spacetime dimensions, derived in (Rador, 2007) for homogeneous perturbations,
is
2
(D 2) f0 Df0 f0
0.
(172)
f0 f0
31
3. Ricci stability in the Palatini formalism
1
S=
d4 x g R + R R + R2 + SM (177)
2
f (R)R 2f (R) = T,
(176)
4. Ghost fields
24
(178)
32
matter and well-formulated otherwise. With the exception of 0 = 3/2, as we will see later, most of Salgados
results can be extended to the more general action
Z
()R ()
4
W () +SM ,
S = d x g
2
2
(180)
which contains the additional coupling function ()
(which is different from the BransDicke parameter 0 )
(Lanahan-Tremblay and Faraoni, 2007).
The field equations, after setting = 1 for this section,
are
1 h
G =
( g )
i
+ ( g )
1h
1
+ g
2
i
(m)
(181)
W ()g + T ,
R W () + = 0, (182)
2
2
where a prime denotes differentiation with respect to .
Eq. (181) can be cast in the form of the effective Einstein
(ef f )
equation G = T , with the effective stress-energy
tensor (Salgado, 2006)
1 ()
(ef f )
()
(m)
T
=
T + T
+ T
,
(183)
()
+
1 ...
Di T 1 ... 1 ... = h1 1 . . . h1 1 . . . h i (3)
1 ...
T
(190)
for any 3-tensor (3) T 1 ... 1 ... , with Di h = 0. The spatial gradient of the scalar is Q D (where D
denotes the covariant derivative of h ), while its momentum is = Ln = n and
1 hij
Kij = i nj =
+ Di Nj + Dj Ni , (191)
2N
t
1
=
(t + N Q ) ,
(192)
N
t Qi + N l l Qi + Ql i N l = Di (N ) . (193)
(ef f )
and
+ () ( g ) ,
(184)
1
= () g
2
W ()g .
(185)
( + 3 ) c .
+W () +
2
2
(186)
The 3 + 1 ArnowittDeserMisner (ADM) formulation of
the theory proceeds by introducing lapse, shift, extrinsic curvature, and gradients of (Reula, 1998; Salgado,
2006; Wald, 1984). Assume that a time function t exists such that the spacetime (M ,g ) admits a foliation
with hypersurfaces t of constant t with unit timelike
normal na . The 3-metric and projection operator on t
are h = g + n n and h , respectively. Moreover,
n n = 1,
n = 0,
h n = h ,
h h = h .
is decomposed
1
(S + J n + J n + En n ) ,
(194)
where
()
T
= () g
()
T
(189)
(187)
1 ()
()
(m)
S + S + S ,(195)
1 ()
(ef f )
J h T n =
J + J() + J(m) ,(196)
1
(ef
f
)
E n n T =
E () + E () + E (m) . (197)
(ef f )
S h h T
R + K 2 Kij K ij = 2E,
(198)
(199)
N
(S + E) , (201)
2
33
where (3) Di Di . Our purpose is to eventually
eliminate all second derivatives. which is present
in eqs. (195)-(197) can actually be eliminated using
eq. (186), provided that 6= 3( )2 /(2).
To be more precise, a direct calculation yields the and -quantities of eqs. (195)-(197)
E () = (D Q + K) + Q2 ,
(202)
J()
()
S
+ D ) Q ,
(203)
= (D Q + K h )
h Q2 2 Q Q ,
(204)
(K Q
where Q2 Q Q , while
S () = (D Q + K 3) + 32 2Q2 ,
(205)
and
E () =
(206)
2 + Q2 + W (),
2
J() = Q ,
(207)
i
h
()
Q2 2 + W () ,(208)
S = Q Q h
2
S () =
(209)
32 Q2 3W ().
2
2h
(3)
(D Q + K) + 2
R + K 2 Kij K ij
2
2
Q2
E (m) + W () , (210)
( + 2 ) =
+
2
1
(Ki c Qc + Di )
Dl K l i Di K +
+ ( + ) Qi =
(m)
Ji
(211)
+ N l K
(3)
j
i
+ K l j N Kj l N
+D Dj N R j N N KK i j
N 2
Q 2 + 2W () + ji
+
2
N
N
+
Dji + K i j
( + ) Qi Qj
i
N h (m)
=
(212)
S
E (m) ji 2S (m) i j ,
2
with trace
t K + N l l K +(3) N N Kij K ij
N
(D Q + K)
N 2
Q (2 + 3 ) 2
+
2
N
=
2W () 3 + S (m) + E (m) (213)
2
2
3( )
2
+ 2
)
+
( + 3 )
(214)
2
2
In vacuo, the initial data (hij , Kij , , Qi , ) on an initial hypersurface 0 obey (210), (211), and Qi = Di ,
Di Qj = Dj Qi . In the presence of matter, the variables
(m)
(m)
E (m) , Ja , Sab must also be assigned on the initial hypersurface 0 . Fixing a gauge corresponds to specifying
the lapse and the shift vector. The system (210)-(213)
contains only first-order derivatives in both space and
time once the dAlembertian is written in terms of
, , , and its derivatives by means of eqs. (186)
or (214). As mentioned earlier, this can be done whenever
6= 3( )2 /(2). As already pointed out in (Salgado,
2006) for the specific case = 1, and can be now generalized for any 6= 3( )2 /(2), the reduction to a
first-order system shows that the Cauchy problem is wellposed in vacuo and well-formulated in the presence of
reasonable matter.
Let us now consider the results specific to f (R) gravity.
Recall that BransDicke theory, which is of interest for
us due to its equivalence with f (R) gravity, corresponds
to () = 0 /, with () = , and W 2V . This
yields the constraints
(3)
2
R + K 2 Kij K ij [D Q + K
i
2 h (m)
0
E
+ V () , (215)
2 + Q2 =
+
2
Dl K l i Di K
(m)
0
1
J
+
K i l Ql + Di +
Qi = i , (216)
t K i j + N l l K i j + K i l j N l Kj l l N i + Di Dj N
N i
(3) Ri j N N KK i j +
(2V () + )
2 j
N 0
N
Di Qj + K i j + 2 Qi Qj
+
i
N h (m)
(m)
(217)
ji 2S (m) i j ,
S
E
=
2
t K + N l l K +(3) N N Kij K ij
0 N
N
(D Q + K) 2 2
i
N h
2V () 3 + S (m) + E (m) ,
=
2
(218)
34
with
T (m)
0
3
=
2 Q2 .
2V ()+V ()+
0 +
2
2
(219)
The condition 6= 3( )2 /(2), which needs to be satisfied in order to for one to be able to use eq. (186) in
order to eliminate can be written in the BransDicke
theory notation as 0 6= 3/2. One could of course had
guessed that by looking at eq. (219). Therefore, metric f (R) gravity, which is equivalent to 0 = 0 Brans
Dicke gravity, has a well-formulated Cauchy problem in
general and is well-posed in vacuo. Further work by
(Salgado et al., 2008) established the well-posedness of
the Cauchy problem for scalar-tensor gravity with = 1
in the presence of matter; this can be translated into the
well-posedness of metric f (R) gravity with matter along
the lines established above.
How about Palatini f (R) gravity, which, corresponding to 0 = 3/2, is exactly the case that the constraint
0 6= 3/2 excludes? Actually, for this value of the
BransDicke parameter, eq. (69), and consequently at
eq. (219), include no derivatives of . Therefore, one can
actually solve algebraically for . [The same could be
done using eq. (186) in the more general case where is
a function of , when = 3( )2 /(2).] We will not
consider cases for which eq. (69) has no roots or when it
is identically satisfied in vacuo. These cases lead to theories for which, in the Palatini f (R) formulation, eq. (21)
has no roots or when it is identically satisfied in vacuo
respectively. As already mentioned in Sec. II.B, the first
case leads to inconsistent field equations, and the second
to a conformally invariant theory (Ferraris et al., 1992),
see also (Sotiriou, 2006b) for a discussion.
Now, in vacuo one can easily show that the solutions
of eq. (69) or (219) will be of the form = constant.
Therefore, all derivatives of vanish and one conclude
that 0 = 3/2 BransDicke theory or Palatini f (R)
gravity have a well-formulated and well-posed Cauchy
problem.25 This could have been expected, as noticed in
(Olmo and Singh, 2009), considering that Palatini f (R)
gravity reduces to GR with a cosmological constant in
vacuo.
In the presence of matter, things are more complicated.
The solutions of eq. (69) or (219) will give as a function
of T , the trace of the stress-energy tensor. This can still
be used to replace in all equations but it will lead to
terms such as T . Therefore, for the Cauchy problem
to be well-formulated in the presence of matter, one does
25
35
(Capozziello et al., 2004, 2005a, 2006a, 2007a,f) computed weak-field limit corrections to the Newtonian
galactic potential and the resulting rotation curves;
when matched to galaxy samples, a best fit yields
n 1.7. (Martins and Salucci, 2007) performed a 2
fit using two broader samples, finding n 2.2 [see
also (Boehmer et al., 2007a) for a variation of this approach focusing on the constant velocity tails of the
rotation curves]. All these values of the parameter n
are in violent contrast with the bounds obtained by
(Barrow and Clifton, 2006; Clifton and Barrow, 2005a,
2006; Zakharov et al., 2006) and have been shown to violate also the current constraints on the precession of perihelia of several Solar System planets (Iorio and Ruggiero,
2007a,b). In addition, the consideration of vacuum metrics used in these works in order to model the gravitational field of galaxies is highly questionable.
The potential obtained in the weak-field limit of f (R)
gravity can affect other aspects of galactic dynamics as
well: the scattering probability of an intruder star and
the relaxation time of a stellar system were studied by
(Hadjimichef and Kokubun, 1997), originally motivated
by quadratic corrections to the EinsteinHilbert action.
B. Palatini f (R) gravity and the conflict with the Standard
Model
26
SM =
(221)
d x g g H H 2 H
2~
~
(in units in which G = c = 1). As an example, we choose
f (R) = R 4 /R (Carroll et al., 2004; Vollick, 2003).
For this choice of f , the potential is V () = 22 (
1)1/2 . To go to the local frame, we want to expand the
action to second order around vacuum. The vacuum of
the action (66) with (221) as a matter action is H = 0,
= 4/3 [using eq. (69)] and g (2 acts as
an effective cosmological constant, so its contribution in
the local frame can be safely neglected).
However, when one tries to use a perturbative expansion for , things stop being straightforward:
is algebraically related to the matter fields as is obvious from eq. (69). Therefore, one gets T /2
m2H H 2 /(~3 2 ) at energies lower than the Higgs mass
(mH 100 1000 GeV). Replacing this expression in
the action (66) perturbed to second order, one immediately obtains that the effective action for the Higgs scalar
is
Z
m2H 2
1
effective
4
SM
d x g
g H H 2 H
2~
~
m2H H 2
m2H (H)2
1+
(222)
+
2 ~ 3
4 ~ 3
at energies E mH . Taking into account the fact that
2 H02 where H01 = 4000 Mpc is the Hubble
radius and H mH because E mH , it is not difficult
to estimate the order of magnitude of the corrections:
at an energy E = 103 eV (corresponding to the length
36
scale L = ~/E = 2 104 m), the first correction is
of the order m2H H 2 /(2 ~3 ) (H01 /H )2 (mH /MP )2
1, where H = ~/mH 2 1019 2 1018 m
is the Compton wavelength of the Higgs and MP =
(~c5 /G)1/2 = 1.2 1019 GeV is the Planck mass. The
second correction is of the order m2H (H)2 /(4 ~3 )
(H01 /XH )2 (mH /MP )2 (H01 /L)2 1. Clearly, having
such non-perturbative corrections to the local frame matter action is unacceptable.
An alternative way to see the same problem would be
to replace m2H H 2 /(~3 2 ) directly in (66). Then
the coupling of matter to gravity is described by the interaction Lagrangian
2 g
m2H H 2
g
+
Lint
~3
2
"
1 2 #
H0
m2H H 2
.
(223)
g 1 +
~3
L
This clearly exhibits the fact that gravity becomes nonperturbative at microscopic scales.
It is obvious that the algebraic dependence of on
the matter fields stands at the root of this problem. We
have still not given any explanation for the paradox
of seeing such a behavior in a theory which apparently
satisfies the metric postulates both in the f (R) and the
BransDicke representation. However, this will become
clear in Sec. VI.C.2.
C. Exact solutions and relevant constraints
1. Vacuum and non-vacuum exact solutions
Let us now turn our attention to exact solutions starting from metric f (R) gravity. We have already mentioned
in Sec. II.A that, as it can be seen easily from the form of
the field equations (6), the maximally symmetric vacuum
solution will be either Minkowski spacetime, if R = 0 is a
root of eq. (9), or de Sitter and anti-de Sitter spacetime,
depending on the sign of the root of the same equation.
Things are slightly more complicated for vacuum solutions with less symmetry: by using eq. (6) it is easy to
verify that any vacuum solution (R = g , T = 0)
of Einsteins theory with a (possibly vanishing) cosmological constant, including black hole solutions, is a solution
of metric f (R) gravity (except for pathological cases for
which eq. (9) has no roots). However, the converse is not
true.
For example, when spherical symmetry is imposed, the
Schwarzschild metric is a solution of metric f (R) gravity if R = 0 in vacuum. If R is constant in vacuo,
then Schwarzschild-(anti-)de Sitter spacetime is a solution. As we have already mentioned though, the JebsenBirkhoff theorem (Wald, 1984; Weinberg, 1972) does not
hold in metric f (R) gravity [unless, of course one wishes
to impose further conditions, such as that R is constant
(Capozziello et al., 2007d)]. Therefore, other solutions
can exists as well. An interesting finding is that the cosmic no-hair theorem valid in GR and in pure f (R) gravity
is not valid, in general, in theories of the form
Z
1
S=
d4 x g R + R2 + R R 2 ,
2
(224)
for
which
exact
anisotropic
solutions
that
continue to inflate anisotropically have been
found
(Barrow and Hervik,
2006a,b)
[see
also
(Kluske and Schmidt, 1996; Maeda, 1988)].
However, isotropization during inflation occurs in mixed
f (, R) models (Maeda et al., 1989)
In addition to the exact cosmological solutions
explored for the purpose of explaining the current cosmic acceleration [see, e.g., (Abdalla et al.,
2005; Barrow and Clifton, 2006; Clifton, 2006a,
2007;
Clifton and Barrow,
2005b,
2006)
and
(Capozziello and De Felice, 2008; Capozziello et al.,
2007c; Modak et al., 2005; Vakili, 2008) for an approach
based on Noether symmetries; see (Carloni et al., 2006)
for bouncing solutions and the conditions that they
satisfy], exact spherically symmetric solutions of metric
f (R) gravity have been explored in the literature, with
most recent studies being motivated by the need to understand the weak-field limit of cosmologically-motivated
theories.
Regarding non-vacuum solutions, the most common
matter source is a perfect fluid. Fluid dynamics in metric
f (R) gravity was studied in (Maartens and Taylor,
1994; Mohseni Sadjadi, 2007; Rippl et al., 1996;
Taylor and Maartens, 1995).
Spherically symmetric
solutions were found in (Bronnikov and Chernakova,
2005a,b,c; Bustelo and Barraco, 2007; Capozziello et al.,
2007d; Clifton, 2006a,b; Mignemi and Wiltshire, 1992;
Multamaki and Vilja, 2006b, 2007a,b; Whitt, 1985).
We regret not being able to present these solutions
extensively here due to space limitations and refer the
reader to the literature for more details.
Stability issues for spherically symmetric solutions
were discussed in (Seifert, 2007). In the theory
Z
g
S = d4 x
R R2 R R + , (225)
37
so the classical stability condition for Schwarzschild black
holes can be enunciated as f (R) 0.
Let us now turn our attention to Palatini f (R) gravity. In this case things are simpler in vacuo: as we saw
in Sec. II.B, the theory reduces in this case (or more
precisely even for matter fields with T =const., where
T is the trace of the stress energy tensor) to GR with
a cosmological constant, which might as well be zero
for some models (Barraco et al., 1999; Borowiec et al.,
1998; Ferraris et al., 1992, 1994). Therefore, it is quite
straightforward that Palatini f (R) gravity will have
the same vacuum solutions as GR with a cosmological constant. Also, the Jebsen-Birkhoff theorem (Wald,
1984; Weinberg, 1972) is valid in the Palatini formalism
(Barausse et al., 2008a,b,c; Kainulainen et al., 2007b).
Cosmological solutions in quadratic gravity were obtained in (Shahid-Saless, 1990, 1991). Spherically symmetric interior solutions in the Palatini formalism can
be found by using the generalization of the TolmanOppenheimer-Volkoff equation valid for these theories, which was found in (Barraco and Hamity, 2000;
Bustelo and Barraco, 2007; Kainulainen et al., 2007b).
Indeed, such solutions have been found and matched
with the unique exterior (anti-)de Sitter solution
(Barraco and Hamity, 1998, 2000; Bustelo and Barraco,
2007; Kainulainen et al., 2007a,b).
Nevertheless, a
matching between exterior and interior which can lead
to a sensible solution throughout spacetime is not always
feasible and this seems to have serious consequences for
the viability of f (R) gravity (Barausse et al., 2008a,b,c).
This is discussed extensively in the next section.
Let us close this section with some remarks on black
hole solutions. As is clear from the above discussion, all
black hole solutions of GR (with a cosmological constant)
will also be solutions of f (R) in both the metric and
the Palatini formalism [see also (Barausse and Sotiriou,
2008; Psaltis et al., 2008)]. However, in the Palatini formalism they will constitute the complete set of black
hole solutions of the theory, whereas in the metric formalism other black hole solutions can exist in principle, as the Jebsen-Birkhoff theorem does not hold. For
a discussion on black hole entropy in f (R) gravity see
(Jacobson et al., 1994, 1995; Vollick, 2007).27
2. Surface singularities and the incompleteness of Palatini f (R)
gravity
A =
,
(227)
8GrP +
1+
r
F
r
1 eB
eB
+
1
,(228)
+
8Gr +
B =
1+
r
F
r
"
#
2
F
2F
eB
f
2 3
r
,(229)
+
+
R
4 F
rF
2
F
"
2 #
rF
3 F
2 F
,
, (230)
r
F
2 F
2F
where F f /R.
To determine an interior
solution we need a generalization of the TolmanOppenheimer-Volkoff (TOV) hydrostatic equilibrium
equation. This has been derived for Palatini f (R) gravity in (Barraco and Hamity, 2000; Bustelo and Barraco,
2007; Kainulainen et al., 2007b): Defining mtot (r)
r(1 eB )/2 and using Eulers equation
P =
28
27
See also (Eling et al., 2006) for a derivation of the field equations
of metric f (R) gravity based on thermodynamical arguments applied to local Rindler horizons.
(226)
A
(P + ) ,
2
(231)
38
one can use eqs. (227) and (228) to arrive to the generalised TOV equations:
1
( + P )
(232)
1 + r(r 2mtot )
4r3 P
(r 2mtot ) ,
mtot +
F
2
2
1
4r +
mtot
=
+
(+ ) .(233)
1+
F
2
r
P =
mtot
Assuming that, approaching the surface from the interior, A and mtot indeed take the correct values required
for the matching, it can be shown that continuity of A
r3
R0 .
24
(239)
The sufficient condition for the singularity to occur is that a polytropic EOS with 3/2 < < 2
should adequately describe just the outer layer of
the matter configuration (and not necessarily the
whole configuration).
In practice, there is no dependence of the result
on the functional form of f (R) [a few unrealistic
exceptions can be found in (Barausse et al., 2008a)]
so what is revealed is a generic aspect of Palatini
f (R) gravity as a class of theories.
The singularities discussed are not coordinate, but
true singularities, as can be easily verified by checking that curvature invariants diverge.
(236)
A (rout ) =
.
(237)
3 12r
rout (R0 rout
out + 24m)
dF d
dF
(P + ) =
(P + ) .
dP
d dP
Let us now examine the behavior of mtot at the surface for different values of the polytropic index . For
1 < < 3/2, C = dC/dP P dC/dP (P + ) 0 at
the surface so that the expression (238) is finite and it
gives continuity of mtot across the surface [cf. eq. (236)].
However, for 3/2 < < 2, C as the surface is approached, provided that dF/dR(R0 ) 6= 0 and
dR/dT (T0 ) 6= 0 (note that these conditions are satisfied
by generic forms of f (R), i.e., whenever an R2 term or
a term inversely proportional to R is present). Therefore, even though mtot remains finite (as can be shown
using the fact that P = 0 at the surface), the divergence of mtot drives to infinity the Riemann tensor of
the metric, R , and curvature invariants, such as R
or R R , as can easily be checked.29 Clearly, such
a singular behaviour is bound to give rise to unphysical
phenomena, such as infinite tidal forces at the surface
(cf. the geodesic deviation equation) which would destroy anything present there. We are, therefore, forced
to conclude that no physically relevant solution exists for
any polytropic EOS with 3/2 < < 2.
The following points about the result just presented
should be stressed:
2
3
2F0 R0 rout
+ rout
R0 8mtot C
, (238)
=
16F0
29
39
The only assumptions made regard the EOS and
the symmetries. Thus, the result applies to all
regimes ranging from Newtonian to strong gravity.
Let us now interpret these results. Obviously, one
could object to the use of the polytropic EOS. Even
though it is extensively used for simple stellar models,
it is not a very realistic description for stellar configurations. However, one does not necessarily have to refer
to stars in order to check whether the issue discussed
here leaves an observable signature. Consider two very
well known matter configurations which are exactly describes by a polytropic EOS: a monoatomic isentropic gas
and a degenerate non-relativistic electron gas. For both
of those cases = 5/3, which is well within the range
for which the singularities appear. Additionally, both of
these configuration can be very well described even with
Newtonian gravity. Yet, Palatini f (R) gravity fails to
provide a reasonable description. Therefore, one could
think of such matter configurations as gedanken experiments which reveal that Palatini f (R) gravity is at best
incomplete (Barausse et al., 2008a,b,c).
On the other hand, the use of the polytropic EOS requires a perfect fluid approximation for the description
of matter. One may, therefore, wish to question whether
the length scale on which the tidal forces become important is larger than the length scale for which the fluid
approximation breaks down (Kainulainen et al., 2007a).
However, quantitative results for tidal forces have been
given in (Barausse et al., 2008b), and it has been shown
that the length scales at which the tidal forces become
relevant are indeed larger than it would be required for
the fluid approximation to break down. The observable
consequences on stellar configurations have also been discussed there. To this, one could also add that a theory
which requires a full description of the microscopic structure of the system in order to provide a macroscopic description of the dynamics is not very appealing anyway.
In any case, it should be stated that the problem discussed is not specific to the polytropic EOS. The use of
the latter only simplifies the calculation and allows an
analytic approach. The root of the problem actually lies
with the differential structure of Palatini f (R) gravity.
Consider the field equations in the form (28): it is
not difficult to notice that these are second order partial
differential equations in the metric. However, since f is a
function of R, and R is an algebraic function of T due to
eq. (20), the right hand side of eq. (28) includes second
derivatives of T . Now, T , being the trace of the stress
energy tensor, will include up to first order derivatives of
the matter fields (assuming that the matter action has
to lead to second order field equations when varied with
respect to the matter fields). Consequently, eq. (28) can
be up to third order in the matter fields!
In GR and most of its alternatives, the field equations
are only of first order in the matter fields. This guarantees that gravity is a cumulative effect: the metric is
generated by an integral over the matter sources and,
therefore, any discontinuities (or even singularities) in
40
a general discussion of representation issues in gravitational theories].
By now it is clear that the metric tensor of f (R) gravity contains, in addition to the usual massless spin 2
graviton, a massive scalar that shows up in gravitational
waves in the metric version of these theories (in the Palatini version, this scalar is not dynamical and does not
propagate). A scalar gravitational wave mode is familiar
from scalar-tensor gravity (Will, 1981), to which f (R)
gravity is equivalent. Because this scalar is massive,
it propagates at a speed lower than the speed of light
and of massless tensor modes and is, in principle, detectable in the arrival times of signals from an exploding
supernova when gravitational wave detectors are sufficiently sensitive [this possibility has been pointed out as a
discriminator between Tensor-Vector-Scalar theories and
GR (Kahya and Woodard, 2007)]. This massive scalar
mode is longitudinal and is of dipole nature to lowest order (Corda, 2007c; Will, 1981). The study of its generation, propagation, and detection falls within the purview
of scalar-tensor gravity (Will, 1981). The propagation of
gravitational waves in the specific model f (R) = Rn was
studied in (Mendoza and Rosas-Guevara, 2007) where
the massive scalar mode is, however, missed.
The generation of gravitational waves specifically in
f (R) gravity has not received much attention in the
literature. Even though the fact that the black hole
solutions of GR will also be solutions of metric f (R)
gravity (without the converse being true) implies that
determining the geometry around a black hole is unlikely to provide evidence for such modifications of gravity (Psaltis et al., 2008), solutions describing perturbed
black holes do behave differently and could, therefore,
leave a detectable imprint on gravitational wave radiation (Barausse and Sotiriou, 2008). Note the analogy to
the fact that cosmological FLRW solutions are shared by
most gravitational theories, but cosmological perturbations reveal more about the underlying theory of gravity
than the exact solutions themselves. Additionally, gravitational radiation from binary systems would probably
be more revealing than that coming from perturbed black
holes when it comes to modified gravity.
The detection of gravitational waves generated in the
theories f (R) = 1/R [already ruled out by Solar System
data (Barrow and Clifton, 2006; Clifton and Barrow,
2005a, 2006)] and f (R) = R + aR2 were studied in
(Corda, 2007a,c) and (Corda, 2007b), respectively.
The study of cosmological gravitational waves in f (R)
gravity is perhaps more promising than that of astrophysically generated waves. The stochastic gravitational
wave background produced in the early universe was analyzed in (Capozziello et al., 2007b, 2008). The authors of
this last reference consider the model f (R) = R1+ and
derive an evolution equation for the metric perturbations
t
t0
2
n h = 0.
(240)
41
R0 [(1 + R2 /R02 )n 1] with n, > 0 and R0 being of
the order of H02 . Here we have tried to avoid considering
specific models and we have attempted to collect general, model-independent results, with the viewpoint that
these theories are to be seen more as toy theories than
definitive and realistic models.
We are now ready to summarize the main results on
f (R) gravity. On the theoretical side, we have explored
all three versions of f (R) gravity: metric, Palatini and
metric-affine. Several issues concerning dynamics, degrees of freedom, matter couplings, etc. have been extensively discussed. The dynamical equivalence between
both metric/Palatini f (R) gravity and BransDicke theory has been, and continues to be, a very useful tool to
study these theories given some knowledge of the aspects
of interest in scalar-tensor gravity. At the same time, the
study of f (R) gravity itself has provided new insight in
the two previously unexplored cases of BransDicke theory with 0 = 0 and 0 = 3/2. We have also considered
most of the applications of f (R) gravity to both cosmology and astrophysics. Finally, we have explored a large
number of possible ways to constrain f (R) theories and
check their viability. In fact, many avatars of f (R) have
been shown to be subject to potentially fatal troubles,
such as a grossly incorrect post-Newtonian limit, short
time scale instabilities, the absence of a matter era, conflict with particle physics or astrophysics, etc.
To avoid repetition, we will not attempt to summarize here all of the theoretical issues, the applications or
the constraints discussed. This, besides being redundant,
would not be very helpful to the reader, as, in most cases,
the insight gained cannot be summarized in a sentence or
two. Specifically, some of the constraints that have been
derived in the literature are not model or parametrization
independent (and the usefulness of some parametrizations is questionable). This does not allow for them to
be expressed in a straightforward manner through simple
mathematical equations applicable directly to a general
function f (R). Particular examples of such constraints
are those coming from cosmology (background evolution,
perturbations, etc.).
However, we have encountered cases in which clear-cut
viability criteria are indeed easy to derive. We would,
therefore, like to make a specific mention of those. A
brief list of quick-and-easy-to-use results is:
In metric f (R) gravity, the Dolgov-Kawasaki instability is avoided if and only if f (R) 0. The
stability condition of de Sitter space is expressed
by eq. (173).
Metric f (R) gravity might pass the weak-field limit
test and at the same time constitute an alternative
to dark energy only if the chameleon mechanism
is effectivethis restricts the possible forms of the
function f (R) in a way that can not be specified by
a simple formula.
tions. These conclusions are essentially modelindependent. (However, this theory could potentially be fixed by adding extra terms quadratic in
the Ricci and/or Riemann tensors, which would
raise the order of the equations.)
Metric-affine gravity as an extension of the Palatini
formalism is not sufficiently developed yet. At the
moment of writing, it is not clear whether it suffers
or not of the same problems that afflict the Palatini
formalism.
Of course, as already mentioned, the situation is often
more involved and cannot be summarized with a quick
recipe. We invite the reader to consult the previous sections and especially the references that they contain.
B. Extensions and new perspectives on f (R) gravity
30
42
the most crucial viability issues of Palatini f (R) gravity
to the lack of dynamics in the gravity sector, such generalizations could actually help by promoting the connection from the role of an auxiliary field to that of a truly
dynamical field (Barausse et al., 2008b). Such generalizations have been considered in (Li et al., 2007a).
Another extension of metric f (R) gravity that appeared recently is that in which the action includes also
an explicit coupling between R and the matter fields.
In (Bertolami et al., 2007; Bertolami and Paramos, 2007;
Boehmer et al., 2008) the following action was considered:
Z
f1 (R)
4
S = d x g
+ [1 + f2 (R)] Lm , (241)
2
where Lm is the matter Lagrangian and f1,2 are (a priori arbitrary) functions of the Ricci curvature R. Since
the matter is not minimally coupled to R, such theories
will not lead to energy conservation and will generically
exhibit a violation of the Equivalence Principle (which
could potentially be controlled by the parameter ).
The motivation for considering such an action spelled
out in (Bertolami et al., 2007) was that the nonconservation of energy could lead to extra forces, which in
turn might give rise to phenomenology similar to MOdified Newtonian Dynamics (MOND) (Milgrom, 1983)
on galactic scales. Other variants of this action have
also been considered elsewhere: in (Nojiri and Odintsov,
2004a), as an alternative to dark energy by setting
f1 (R) = R and keeping only the nonminimal coupling of
matter to the Ricci curvature; in (Dolgov and Kawasaki,
2003b; Mukohyama and Randall, 2004), where the idea
of making the kinetic term of a (minimally coupled)
scalar field dependent on the curvature, while keeping
f1 (R) = R was exploited in attempts to cure the cosmological constant problem; in (Bertolami and Paramos,
2007) the consequences of such a theory for stellar equilibrium were studied; finally, generalized constraints in
order to avoid the instability discussed in Sec. V.B.1 were
derived in (Faraoni, 2007a). The viability of theories with
such couplings between R and matter is still under investigation. However, the case in which both f1 and f2 are
linear has been shown to be non-viable (Sotiriou, 2008)
and, for the more general case of the action (241), serious doubts have been expressed (Sotiriou and Faraoni,
2008) on whether extra forces are indeed present in galactic environments and, therefore, whether this theory can
really account for the MOND-like phenomenology that
initially motivated its use in (Bertolami et al., 2007) as
a substitute for dark matter.
One could also consider extensions of f (R) gravity in
which extra fields appearing in the action couple to different curvature invariants. A simple example with a scalar
field is the action (159), which is sometimes dubbed extended quintessence (Perrotta et al., 2000), similarly to
the extended inflation realized with BransDicke theory.
However, such generalizations lie beyond the scope of this
review.
C. Concluding remarks
Our goal was to present a comprehensive but still thorough review of f (R) gravity in order to provide a starting
point for the reader less experienced in this field and a
reference guide for the expert. However, even though we
have attempted to cover all angles, no review can replace
an actual study of the literature itself. It seems inevitable
that certain aspects of f (R) might have been omitted, or
analyzed less than rigorously and, therefore, the reader
is urged to resort to the original sources.
Although many shortcomings of f (R) gravity have
been presented which may reduce the initial enthusiasm
43
with which one might have approached this field, the fact
that such theories are mostly considered as toy theories
should not be missed. The fast progress in this field,
especially in the last five years, is probably obvious by
now. And very useful lessons, which have helped significantly in the understanding of (classical) gravity, have
been learned in the study of f (R) gravity. In this sense,
the statement made in the Introduction that f (R) gravity is a very useful toy theory seems to be fully justified.
Remarkably, there are still unexplored aspects of f (R)
theories or their extensions, such as those mentioned in
the previous section, which can turn out to be fruitful.
Acknowledgments
References
Abdalla, M. C. B., S. Nojiri, and S. D. Odintsov, 2005, Class.
Quant. Grav. 22, L35.
Abdelwahab, M., S. Carloni, and P. K. S. Dunsby, 2008, Class.
Quant. Grav. 25, 135002.
Accioly, A., S. Ragusa, E. C. de Rey Neto, and H. Mukai,
1999, N. Cimento B114, 595.
Alcubierre, M., et al., 2002a, Class. Quant. Grav. 19, 5017.
Alcubierre, M., et al., 2002b, eprint astro-ph/0204307.
Allemandi, G., A. Borowiec, and M. Francaviglia, 2004, Phys.
Rev. D70, 043524.
Allemandi, G., A. Borowiec, M. Francaviglia, and S. D.
Odintsov, 2005a, Phys. Rev. D72, 063505.
Allemandi, G., M. Francaviglia, M. L. Ruggiero, and
A. Tartaglia, 2005b, Gen. Rel. Grav. 37, 1891.
Allemandi, G., and M. L. Ruggiero, 2007, Gen. Rel. Grav.
39, 1381.
Amarzguioui, M., O. Elgaroy, D. F. Mota, and T. Multamaki,
2006, Astron. Astrophys. 454, 707.
Amendola, L., S. Capozziello, M. Litterio, and F. Occhionero,
1992, Phys. Rev. D45, 417.
Amendola, L., R. Gannouji, D. Polarski, and S. Tsujikawa,
2007a, Phys. Rev. D75, 083504.
Amendola, L., D. Polarski, and S. Tsujikawa, 2007b, Phys.
Rev. Lett. 98, 131302.
Amendola, L., D. Polarski, and S. Tsujikawa, 2007c, Int. J.
Mod. Phys. D16, 1555.
Amendola, L., and S. Tsujikawa, 2008, Phys. Lett. B660,
125.
44
Boehmer, C. G., L. Hollenstein, and F. S. N. Lobo, 2007b,
Phys. Rev. D76, 084005.
Borowiec, A., M. Ferraris, M. Francaviglia, and I. Volovich,
1994, Gen. Rel. Grav. 26, 637.
Borowiec, A., M. Ferraris, M. Francaviglia, and I. Volovich,
1998, Class. Quant. Grav. 15, 43.
Bosma, A., 1978, The distribution and kinematics of neutral
hydrogen in spiral galaxies of various morphological types,
Ph.D. thesis, Rejksuniversiteit Groningen.
Brandenberger, R. H., 1992, eprint gr-qc/9210014.
Brandenberger, R. H., 1993, eprint gr-qc/9302014.
Brandenberger, R. H., 1995, eprint gr-qc/9509059.
Brandenberger, R. H., V. F. Mukhanov, and A. Sornborger,
1993, Phys. Rev. D48, 1629.
Brans, C., and R. H. Dicke, 1961, Phys. Rev. 124, 925.
Bronnikov, K. A., and M. S. Chernakova, 2005a, Russ. Phys.
J. 48, 940.
Bronnikov, K. A., and M. S. Chernakova, 2005b, eprint grqc/0503025.
Bronnikov, K. A., and M. S. Chernakova, 2005c, Grav. Cosmol. 11, 305.
Brookfield, A. W., C. van de Bruck, and L. M. H. Hall, 2006,
Phys. Rev. D74, 064028.
Buchbinder, I. L., S. D. Odintsov, and I. L. Shapiro, 1992, Effective Actions in Quantum Gravity (IOP Publishing, Bristol).
Buchdahl, H. A., 1970, Mon. Not. Roy. Astron. Soc. 150, 1.
Burles, S., K. M. Nollett, and M. S. Turner, 2001, Astrophys.
J. 552, L1.
Burton, H., and R. B. Mann, 1998a, Phys. Rev. D57, 4754.
Burton, H., and R. B. Mann, 1998b, Class. Quant. Grav. 15,
1375.
Bustelo, A. J., and D. E. Barraco, 2007, Class. Quant. Grav.
24, 2333.
Calcagni, G., B. de Carlos, and A. De Felice, 2006, Nucl.
Phys. B752, 404.
Caldwell, R. R., R. Dave, and P. J. Steinhardt, 1998, Phys.
Rev. Lett. 80, 1582.
Capozziello, S., V. F. Cardone, S. Carloni, and A. Troisi,
2004, Phys. Lett. A326, 292.
Capozziello, S., V. F. Cardone, S. Carloni, and A. Troisi,
2005a, AIP Conf. Proc. 751, 54.
Capozziello, S., V. F. Cardone, and A. Troisi, 2005b, Phys.
Rev. D71, 043503.
Capozziello, S., V. F. Cardone, and A. Troisi, 2006a, JCAP
0608, 001.
Capozziello, S., V. F. Cardone, and A. Troisi, 2007a, Mon.
Not. Roy. Astron. Soc. 375, 1423.
Capozziello, S., S. Carloni, and A. Troisi, 2003, Recent Res.
Dev. Astron. Astrophys. 1, 625.
Capozziello, S., C. Corda, and M. De Laurentis, 2007b, Mod.
Phys. Lett. A22, 1097.
Capozziello, S., and A. De Felice, 2008, eprint 0804.2163.
Capozziello, S., M. De Laurentis, and M. Francaviglia, 2008,
Astropart. Phys. 29, 125.
Capozziello, S., and M. Francaviglia, 2008, Gen. Rel. Grav.
40, 357.
Capozziello, S., S. Nojiri, and S. D. Odintsov, 2006b, Phys.
Lett. B632, 597.
Capozziello, S., S. Nojiri, S. D. Odintsov, and A. Troisi, 2006c,
Phys. Lett. B639, 135.
Capozziello, S., F. Occhionero, and L. Amendola, 1993, Int.
J. Mod. Phys. D1, 615.
Capozziello, S., A. Stabile, and A. Troisi, 2006d, Mod. Phys.
45
Clifton, T., and J. D. Barrow, 2005a, Phys. Rev. D72, 103005.
Clifton, T., and J. D. Barrow, 2005b, Phys. Rev. D72,
123003.
Clifton, T., and J. D. Barrow, 2006, Class. Quant. Grav. 23,
2951.
Cocke, W. J., and J. M. Cohen, 1968, J. Math. Phys 9, 971.
Codello, A., and R. Percacci, 2006, Phys. Rev. Lett. 97,
221301.
Cognola, G., E. Elizalde, S. Nojiri, S. D. Odintsov, and
S. Zerbini, 2005, JCAP 0502, 010.
Cognola, G., E. Elizalde, S. Nojiri, S. D. Odintsov, and
S. Zerbini, 2006, Phys. Rev. D73, 084007.
Cognola, G., M. Gastaldi, and S. Zerbini, 2008, Int. J. Theor.
Phys. 47, 898.
Cognola, G., and S. Zerbini, 2006, J. Phys. A39, 6245.
Cognola, G., and S. Zerbini, 2008, eprint arXiv:0802.3967
[hep-th].
Coley, A. A., 2003, Dynamical systems and cosmology
(Kluwer, Dordrecht).
Comelli, D., 2005, Phys. Rev. D72, 064018.
Corda, C., 2007a, eprint 0710.2605.
Corda, C., 2007b, eprint 0711.4917.
Corda, C., 2007c, JCAP 0704, 009.
de la Cruz-Dombriz, A., and A. Dobado, 2006, Phys. Rev.
D74, 087501.
de la Cruz-Dombriz, A., A. Dobado, and A. L. Maroto, 2008,
eprint arXiv:0802.2999 [astro-ph].
Dabrowski, M. P., T. Denkiewicz, and D. Blaschke, 2007, Annalen Phys. 16, 237.
Davidson, A., 2005, Class. Quant. Grav. 22, 1119.
De Felice, A., 2007, J. Phys. A40, 7061.
De Felice, A., and M. Hindmarsh, 2007, JCAP 0706, 028.
De Felice, A., M. Hindmarsh, and M. Trodden, 2006, JCAP
0608, 005.
DeDeo, S., and D. Psaltis, 2007, eprint 0712.3939.
Deser, S., 1970, Ann. Phys. (NY) 59, 248.
Deser, S., and B. Tekin, 2002, Phys. Rev. Lett. 89, 101101.
Deser, S., and B. Tekin, 2003a, Phys. Rev. D67, 084009.
Deser, S., and B. Tekin, 2003b, Class. Quant. Grav. 20, 4877.
Deser, S., and B. Tekin, 2007, Phys. Rev. D75, 084032.
Dick, R., 2004, Gen. Rel. Grav. 36, 217.
Dicke, R. H., 1962, Phys. Rev. 125, 2163.
Dolgov, A., and D. N. Pelliccia, 2006, Nucl. Phys. B734, 208.
Dolgov, A. D., and M. Kawasaki, 2003a, Phys. Lett. B573,
1.
Dolgov, A. D., and M. Kawasaki, 2003b, eprint astroph/0307442.
Dominguez, A. E., and D. E. Barraco, 2004, Phys. Rev. D70,
043505.
Dvali, G. R., G. Gabadadze, and M. Porrati, 2000, Phys. Lett.
B485, 208.
Easson, D. A., 2004, Int. J. Mod. Phys. A19, 5343.
Eddington, A. S., 1923, The Mathematical Theory of Relativity (Cambridge University Press, Cambridge).
Ehlers, J., P. Geren, and R. K. Sachs, 1968, J. Math. Phys.
9, 1344.
Eisenstein, D. J., et al. (SDSS), 2005, Astrophys. J. 633, 560.
Eling, C., R. Guedens, and T. Jacobson, 2006, Phys. Rev.
Lett. 96, 121301.
Ellis, G. F. R., and M. Bruni, 1989, Phys. Rev. D40, 1804.
Ellis, G. F. R., M. Bruni, and J. Hwang, 1990, Phys. Rev.
D42, 1035.
Ellis, G. F. R., J. Hwang, and M. Bruni, 1989, Phys. Rev.
D40, 1819.
46
733.
Hehl, F. W., and G. D. Kerling, 1978, Gen. Rel. Grav. 9, 691.
Hehl, F. W., J. D. McCrea, E. W. Mielke, and Y. Neeman,
1995, Phys. Rept. 258, 1.
Hehl, F. W., P. Von Der Heyde, G. D. Kerlick, and J. M.
Nester, 1976, Rev. Mod. Phys. 48, 393.
Hindawi, A., B. A. Ovrut, and D. Waldram, 1996, Phys. Rev.
D53, 5597.
Hoyle, C. D., et al., 2001, Phys. Rev. Lett. 86, 1418.
Hu, W., and I. Sawicki, 2007a, Phys. Rev. D76, 104043.
Hu, W., and I. Sawicki, 2007b, Phys. Rev. D76, 064004.
Hwang, J. C., 1990a, Class. Quant. Grav. 7, 1613.
Hwang, J. C., 1990b, Phys. Rev. D42, 2601.
Hwang, J.-C., 1997, Class. Quant. Grav. 14, 3327.
Hwang, J.-C., 1998, Class. Quant. Grav. 15, 1401.
Hwang, J.-C., and H. Noh, 1996, Phys. Rev. D54, 1460.
Hwang, J.-C., and H. Noh, 2002, Phys. Rev. D65, 023512.
Iglesias, A., N. Kaloper, A. Padilla, and M. Park, 2007, Phys.
Rev. D76, 104001.
Iorio, L., 2007, eprint arXiv:0710.0022 [gr-qc].
Iorio, L., and M. L. Ruggiero, 2007a, Int. J. Mod. Phys. A22,
5379.
Iorio, L., and M. L. Ruggiero, 2007b, eprint arXiv:0711.0256
[gr-qc].
Israel, W., 1966, Nuovo Cimento 44B, 1, see erratum (Israel,
1967).
Israel, W., 1967, Nuovo Cimento 49B, 463.
Jacobson, T., G. Kang, and R. C. Myers, 1994, Phys. Rev.
D49, 6587.
Jacobson, T., G. Kang, and R. C. Myers, 1995, Phys. Rev.
D52, 3518.
Jacobson, T., and D. Mattingly, 2001, Phys. Rev. D64,
024028.
Jhingan, S., et al., 2008, eprint 0803.2613.
Jin, X.-H., D.-J. Liu, and X.-Z. Li, 2006, eprint astroph/0610854.
Kahn, F. D., and L. Woltjer, 1959, Astrophys. J. 130, 705.
Kahya, E. O., and R. P. Woodard, 2007, Phys. Lett. B652,
213.
Kainulainen, K., J. Piilonen, V. Reijonen, and D. Sunhede,
2007a, Phys. Rev. D76, 024020.
Kainulainen, K., V. Reijonen, and D. Sunhede, 2007b, Phys.
Rev. D76, 043503.
Khoury, J., and A. Weltman, 2004a, Phys. Rev. D69, 044026.
Khoury, J., and A. Weltman, 2004b, Phys. Rev. Lett. 93,
171104.
Kibble, T. W. B., 1961, J. Math. Phys. 2, 212.
Kluske, S., and H.-J. Schmidt, 1996, Astron. Nachr. 317, 337.
Kneller, J. P., and G. Steigman, 2004, New J. Phys. 6, 117.
Knop, R. A., et al. (Supernova Cosmology Project), 2003,
Astrophys. J. 598, 102.
Knox, L., Y.-S. Song, and J. A. Tyson, 2006, Phys. Rev. D74,
023512.
Koivisto, T., 2006a, Class. Quant. Grav. 23, 4289.
Koivisto, T., 2006b, Phys. Rev. D73, 083517.
Koivisto, T., 2007, Phys. Rev. D76, 043527.
Koivisto, T., and H. Kurki-Suonio, 2006, Class. Quant. Grav.
23, 2355.
Kolb, E. W., and M. S. Turner, 1992, The Early Universe
(Addison-Wesley, California).
Koyama, K., and R. Maartens, 2006, JCAP 0601, 016.
Lambiase, G., and G. Scarpetta, 2006, Phys. Rev. D74,
087504.
Lanahan-Tremblay, N., and V. Faraoni, 2007, Class. Quant.
47
[astro-ph].
Multamaki, T., and I. Vilja, 2007b, Phys. Rev. D76, 064021.
Myers, R. C., and J. Z. Simon, 1988, Phys. Rev. D38, 2434.
Myers, R. C., and J. Z. Simon, 1989, Gen. Rel. Grav. 21, 761.
Nakamura, R., M. Hashimoto, S. Gamow, and K. Arai, 2006,
Astron. Astrophys. 448, 23.
Navarro, I., and K. Van Acoleyen, 2005, Phys. Lett. B622, 1.
Navarro, I., and K. Van Acoleyen, 2006, JCAP 0603, 008.
Navarro, I., and K. Van Acoleyen, 2007, JCAP 0702, 022.
Noakes, D. R., 1983, J. Math. Phys. 24, 1846.
Nojiri, S., 2004, TSPU Vestnik 44N7, 49.
Nojiri, S., and S. D. Odintsov, 2003a, Phys. Rev. D68,
123512.
Nojiri, S., and S. D. Odintsov, 2003b, Phys. Lett. B576, 5.
Nojiri, S., and S. D. Odintsov, 2004a, Phys. Lett. B599, 137.
Nojiri, S., and S. D. Odintsov, 2004b, Gen. Rel. Grav. 36,
1765.
Nojiri, S., and S. D. Odintsov, 2005, Phys. Lett. B631, 1.
Nojiri, S., and S. D. Odintsov, 2006, Phys. Rev. D74, 086005.
Nojiri, S., and S. D. Odintsov, 2007a, Int. J. Geom. Meth.
Mod. Phys. 4, 115.
Nojiri, S., and S. D. Odintsov, 2007b, J. Phys. Conf. Ser. 66,
012005.
Nojiri, S., and S. D. Odintsov, 2007c, J. Phys. A40, 6725.
Nojiri, S., and S. D. Odintsov, 2007d, Phys. Lett. B657, 238.
Nojiri, S., and S. D. Odintsov, 2008a, eprint 0801.4843.
Nojiri, S., and S. D. Odintsov, 2008b, Phys. Rev. D77,
026007.
Nojiri, S., and S. D. Odintsov, 2008c, Phys. Lett. B659, 821.
Nojiri, S., and S. D. Odintsov, 2008d, eprint 0804.3519.
Nordtvedt, J., Kenneth, 1970, Astrophys. J. 161, 1059.
Novello, M., and S. E. P. Bergliaffa, 2008, eprint 0802.1634.
Nunez, A., and S. Solganik, 2004, eprint hep-th/0403159.
OHanlon, J., 1972a, J. Phys. A 5, 803.
OHanlon, J., 1972b, Phys. Rev. Lett. 29, 137.
OHanlon, J., and B. Tupper, 1972, Nuovo Cimento B 7, 305.
Olmo, G. J., 2005a, Phys. Rev. Lett. 95, 261102.
Olmo, G. J., 2005b, Phys. Rev. D72, 083505.
Olmo, G. J., 2007, Phys. Rev. D75, 023511.
Olmo, G. J., 2008, Phys. Rev. D77, 084021.
Olmo, G. J., and P. Singh, 2009, JCAP 0901, 030.
Ostriker, J. P., and P. J. Steinhardt, 1995, Nature 377, 600.
Ostrogradski, M., 1850, Mem. Ac. St. Petersbourg VI 4, 385.
Paul, B. C., and D. Paul, 2006, Phys. Rev. D74, 084015.
Paul, D., and B. C. Paul, 2005, Phys. Rev. D72, 064012.
Peebles, P. J. E., and B. Ratra, 1988, Astrophys. J. 325, L17.
Perez Bergliaffa, S. E., 2006, Phys. Lett. B642, 311.
Perlmutter, S., et al. (Supernova Cosmology Project), 1998,
Nature 391, 51.
Perrotta, F., C. Baccigalupi, and S. Matarrese, 2000, Phys.
Rev. D61, 023507.
Persic, M., P. Salucci, and F. Stel, 1996, Mon. Not. Roy.
Astron. Soc. 281, 27.
Pogosian, L., and A. Silvestri, 2008, Phys. Rev. D77, 023503.
Psaltis, D., D. Perrodin, K. R. Dienes, and I. Mocioiu, 2008,
Phys. Rev. Lett. 100, 091101.
Querella, L., 1998, eprint gr-qc/9902044.
Rador, T., 2007, Phys. Rev. D75, 064033.
Rajaraman, A., 2003, eprint astro-ph/0311160.
Ratra, B., and P. J. E. Peebles, 1988, Phys. Rev. D37, 3406.
Reula, O. A., 1998, Living Rev. Rel. 1, 3.
Riegert, R. J., 1984, Phys. Lett. A105, 110.
Riess, A. G., et al. (Supernova Search Team), 1998, Astron.
J. 116, 1009.
48
Sotiriou, T. P., 2007b, Modified Actions for Gravity: Theory
and Phenomenology, Ph.D. thesis, SISSA, eprint 0710.4438.
Sotiriou, T. P., 2008, eprint 0805.1160.
Sotiriou, T. P., and V. Faraoni, 2008, eprint 0805.1249.
Sotiriou, T. P., V. Faraoni, and S. Liberati, 2007, eprint
0707.2748.
Sotiriou, T. P., and S. Liberati, 2007a, J. Phys. Conf. Ser. 68,
012022.
Sotiriou, T. P., and S. Liberati, 2007b, Annals Phys. 322,
935.
Soussa, M. E., and R. P. Woodard, 2004, Gen. Rel. Grav. 36,
855.
de Souza, J. C. C., and V. Faraoni, 2007, Class. Quant. Grav.
24, 3637.
Spergel, D. N., et al. (WMAP), 2007, Astrophys. J. Suppl.
170, 377.
Stabenau, H. F., and B. Jain, 2006, Phys. Rev. D74, 084007.
Starobinsky, A. A., 1980, Phys. Lett. B91, 99.
Starobinsky, A. A., 2007, JETP Lett. 86, 157.
Stelle, K., 1978, Gen. Rel. Grav. 9, 353.
Stelle, K. S., 1977, Phys. Rev. D16, 953.
Stoeger, S. J., William R., R. Maartens, and G. F. R. Ellis,
1995, Astrophys. J. 443, 1.
Stone, M., 2000, Phys. Rev. E62, 1341.
Strominger, A., 1984, Phys. Rev. D30, 2257.
Susskind, L., 2003, eprint hep-th/0302219.
Taylor, D. R., and R. Maartens, 1995, Gen. Rel. Grav. 27,
1309.
Teyssandier, P., and P. Tourrenc, 1983, J. Math. Phys. 24,
2793.
Tonry, J. L., et al. (Supernova Search Team), 2003, Astrophys. J. 594, 1.
Trodden, M., V. F. Mukhanov, and R. H. Brandenberger,
1993, Phys. Lett. B316, 483.
Tsujikawa, S., 2007, Phys. Rev. D76, 023514.
Tsujikawa, S., 2008, Phys. Rev. D77, 023507.
Tsujikawa, S., K. Uddin, and R. Tavakol, 2008, Phys. Rev.
D77, 043007.
Uddin, K., J. E. Lidsey, and R. Tavakol, 2007, Class. Quant.