You are on page 1of 23

Symmetric Teleparallel Gauss-Bonnet Gravity and its Extensions

Juan Manuel Armaleoa,b ,∗ Sebastian Bahamondec,d,† Georg Trenklere,f ,‡ and Leonardo G. Trombettae§
a
Universidad de Buenos Aires, Facultad de Ciencias Exactas y Naturales,
Departamento de Fı́sica, 1428 Buenos Aires, Argentina
b
CONICET - Universidad de Buenos Aires, Instituto de Fı́sica de Buenos Aires (IFIBA). Buenos Aires,
Argentina Ciudad Universitaria, Pabellon I, 1428 Buenos Aires, Argentina.
c
Department of Physics, Tokyo Institute of Technology 2-12-1 Ookayama, Meguro-ku, Tokyo 152-8551, Japan.
d
Kavli Institute for the Physics and Mathematics of the Universe (WPI),
The University of Tokyo Institutes for Advanced Study (UTIAS),
The University of Tokyo, Kashiwa, Chiba 277-8583, Japan.
e
CEICO, Institute of Physics of the Czech Academy of Sciences,
Na Slovance 1999/2, 182 21, Prague 8, Czechia.
f
Institute of Theoretical Physics, Faculty of Mathematics and Physics,
Charles University, V Holešovičkách 2, 180 00 Prague 8, Czechia.

General Teleparallel theories assume that curvature is vanishing in which case gravity can be
solely represented by torsion and/or nonmetricity. Using differential form language, we express the
arXiv:2308.07299v2 [gr-qc] 10 Nov 2023

Riemannian Gauss-Bonnet invariant concisely in terms of two General Teleparallel Gauss-Bonnet


invariants, a bulk and a boundary one. Both terms are boundary terms in four dimensions. We also
find that the split is not unique and present two possible alternatives. In the absence of nonmetricity
our expressions coincide with the well-known Metric Teleparallel Gauss-Bonnet invariants for one
of the splits. Next, we focus on the description where only nonmetricity is present and show some
examples in different spacetimes. We finish our discussion by formulating novel modified Symmetric
Teleparallel theories constructed with our new scalars.

I. INTRODUCTION

General Relativity (GR) is the flagship theory of gravity and has been very successful in describing gravitational
physics as well as predicting the existence of gravitational waves, black holes and an expanding universe. It is however
not without its limitations, as many aspects remain unclear, such as the unavoidable appearance of singularities, the
nature of Dark Matter and Dark Energy and the lack of a compelling UV completion. One way to address some
of these is to modify the theory by changing some of its underlying assumptions. Lovelock’s theorem ensures the
Einstein-Hilbert action to uniquely describe (pure) gravity in Riemannian geometry with field equations that are at
most second order in D = 4 dimensions. The next invariant that appears in an expansion in powers of the curvature
is the Gauss-Bonnet invariant [1, 2],
◦ ◦ ◦ ◦ ◦ ◦
G = Rµνρσ Rµνρσ − 4Rµν Rµν + R2 , (1)

which is a topological invariant in D = 4, therefore not affecting the dynamics. However, precisely because of its
topological character, it can lead to interesting features in a theory of gravity. In order to make this quantity relevant
in D = 4 it is necessary instead to let go some of the other assumptions of Lovelock’s theorem. For instance, by
allowing for field equations of higher-than-second order nature, which generically imply the appearance of new degrees
of freedom (d.o.f.), one can then construct a theory containing the Gauss-Bonnet invariant in a nontrivial way, such
◦ ◦
as f (R, G) gravity [3–8].

Another possibility to modify GR is coupling G directly to new degrees of freedom. Indeed, in the context of scalar-
tensor theories of gravity in Riemannian geometry, a lot of attention has been given in recent years to a coupling

between the Gauss-Bonnet invariant and the scalar field ψ of the form f (ψ)G, which can be generated for example at
low energies by string theory [9–11], or in general by the spontaneous breaking of a conformal symmetry, with ψ related
with the corresponding dilaton [12, 13]. This type of operator belongs to the Horndeski class of theories, therefore
leading to second-order field equations for both the scalar and the metric [14], and also inheriting nice properties such

∗ Electronic address: jarmaleo@df.uba.ar


† Electronic address: sbahamondebeltran@gmail.com, bahamonde.s.aa@m.titech.ac.jp
‡ Electronic address: trenkler@fzu.cz
§ Electronic address: trombetta@fzu.cz
2

as nonrenormalization [15]. Phenomenological implications in cosmology include for example the existence of eternal
solutions with zero cosmological constant [16] and inflationary models where the inflaton is nonminimally coupled
to the Gauss-Bonnet invariant [17–19]. On the astrophysical side, the main appeal is the fact that the presence of
such operator generically induces scalarization of black-hole solutions [20, 21], which include interesting scenarios of
spontaneous scalarization [22, 23] (also spin-induced [24, 25]) when the coupling is non-linear in ψ.1

In the special case of a linear coupling ψ G, interestingly the theory enjoys shift-symmetry2 and has been shown
to be the only way of evading known no-hair theorems for shift-symmetric scalar-tensor (Riemannian) theories and
flat asymptotics [27], even when considering the larger Degenerate Higher-Order Scalar-Tensor (DHOST) class of
theories [28]. Black holes in this case are always scalarized, which implies strong constraints from gravitational-wave
observations due to unavoidable scalar-wave emission during the inspiral phase of binary mergers [29–32], although
they could still lead to potentially observable effects in the ringdown phase and provide insight about Dark Energy
scenarios [33]. Strong constraints also follow from theoretical consistency considerations such as the requirement of
causality [34].
Yet another way of modifying GR is by changing the underlying geometrical notion used to describe gravity. It
has been shown that it is possible to formulate GR equivalently either in terms of curvature, torsion or nonmetricity
[35–37]. The scenario in which the general curvature is switched off and still the theory provides the Einstein’s field
equations is known as the General Teleparallel equivalent of GR theory [38–40]. Modifications of that theory are
however different depending on which geometrical formulation is chosen. When not only curvature is zero but also
nonmetricity(torsion) is vanishing, one restricts the theories to be part of the so-called metric(symmetric) Teleparallel
theories or torsional(nonmetricity-based) Teleparallel theories [36, 41–46]. For simplicity, we will use the abbreviation
TG for metric (or torsional) Teleparallel gravity, while Symmetric TG will refer to Symmetric Teleparallel gravity
when only nonmetricity is present. Furthermore, the abbreviation General TG will refer to the General Teleparallel
theories with vanishing curvature and non-zero torsion and nonmetricity.
In particular, any Teleparallel theory could allow for novel ways of including the Gauss-Bonnet invariant in an
action nontrivially. Most of the known results in black hole physics or cosmology assume Riemannian geometry and
the conclusions no longer apply in this case. It is interesting to ask whether they would still hold when considering
instead a Teleparallel formulation. In order to pursue these studies it is necessary to have manageable expressions for
the Riemannian Gauss-Bonnet invariant in terms of teleparallel quantities. For the case of TG this has been already
developed in Ref. [47], where the Riemannian Gauss-Bonnet invariant is split into two terms: a bulk term denoted as
(T ) (T )
TG , and a boundary one as BG . Those terms allow us to construct modified Metric TG theories with non-trivial
dynamics for the Metric TG Gauss-Bonnet invariants, see [47–49] as examples. There are interesting works related
to those torsional theories in the context of cosmology [50–55], and recently, in black hole physics [56]. It has been
found that the Metric TG Gauss-Bonnet invariants provided by torsion generate scalarized black-hole solutions with
spontenous scalarization, and in general, they have different features as in the Riemannian case [57].
In this paper we aim to extend the work of Ref. [47] to the General TG framework, where both torsion and
nonmetricity are present, by formulating a compact and manageable expression for the Riemannian Gauss-Bonnet

invariant G. As a special case, this includes the so-called Symmetric TG where only nonmetricity is different from
zero. In this latter setup, which has been so far the least explored case in the literature, we then use our results to
discuss some novel formulations of Symmetric TG theories that include the Gauss-Bonnet invariant.
The paper is organised as follows: In Sec. II we introduce General TG geometries using both differential form
language and their corresponding tensorial expressions. We also formulate the General TG equivalent of GR. Then,
in Sec. III we split the Riemannian Gauss-Bonnet invariant into two General TG scalars which are both boundary
terms in 4-dimensions. Sec. IV is devoted to studying the particular case of Symmetric TG where we build their
Gauss-Bonnet invariants, construct different new theories, and explore some examples for different spacetimes. We
summarise and discuss our main results in Sec. V.
The notation throughout this paper is as follows: Latin indices a, b, . . . run over the tangent space of the D-
dimensional spacetime while Greek indices µ, ν, . . . run over all coordinates of the D-dimensional spacetime. Quantities
denoted with tildes on top are referring to a general connection, whereas quantities with an overcircle denote the
Riemannian ones (computed using the Levi-Civita connection). Further, quantities without any symbol on top are of
purely Teleparallel nature.

1 The coupling leading to spontanous scalarization of black holes has also been shown to trigger instabilities in the context of inflationary
cosmology [26].

2 Due to the topological character of G in D = 4, a constant shift ψ → ψ + c does not affect the dynamics.
3

II. GENERAL TELEPARALLEL GEOMETRIES

In this section we revise the formalism of differential forms and introduce the relevant geometrical quantities. For
this purpose, we mainly follow the work presented in [47, 58, 59] and encourage the reader to see the aforementioned
references for more details. We list the definitions of the geometrical quantities in the language of differential forms
needed to construct the General TG equivalent of GR.

A. Geometrical quantities in differential form

In the early 1920’s Élie Cartan presented an alternative formulation of GR in terms of two new dynamical quantities:
the (dual) vielbein ea , also known as3 the basis of the cotangent space or tetrads, and the connection 1-form ω̃ a b , also
known as spin connection. As opposed to Einstein’s idea, Cartan’s idea was to distinguish between the metric and the
connection as two different objects (unlike GR in which the Levi-Civita connection is assumed). The connection then
is not written in terms of the metric anymore but they are independent. The information about the geometry (i.e.
gravitation) which in GR is encoded in the Ricci scalar R̃ = g µν R̃µν via the metric, will now be given by the three
quantities (Qab , T a , R̃a b ) known as the nonmetricity 1-form, the torsion 2-form and the curvature 2-form respectively.
They are defined as

Qab := D̃gab := dgab − ω̃ab − ω̃ba , (2)


a a a a c
T := D̃e := de + ω̃ c ∧e , (3)
a a a c
R̃ b := dω̃ b + ω̃ c ∧ ω̃ b , (4)

where d is the exterior derivative and the symbol ∧ denotes the wedge product; gab stands for the metric tensor, with
which we raise and/or lower indices. The above equations are known as Cartan’s structure equations, and they define
the nonmetricity 1-form Qab , the torsion 2-form T a and the curvature 2-form R̃a b in terms of the basis ea and the
connection 1-form ω̃ a b . The general covariant exterior derivative D̃ is defined in the following way: acting on a set of
p−forms Φa b ,

D̃Φa b = dΦa b + ω̃ a c ∧ Φc b − ω̃ c b ∧ Φa c . (5)

Note that the covariant exterior derivative D̃ acts in the same way as the exterior derivative d in the sense that when
applied on a p−form, it gives a (p + 1)−form. The connection terms ensure that D̃ (when applied to a p−form)

transforms as a tensorial object in the same way as the Levi-Civita connection Γa bc ensures that the covariant

derivative ∇ behaves like a tensor when applied to a tensor field in GR. The connection 1-form ω̃ a b defines the
parallel transportation. With the covariant exterior derivative defined, we can write the Bianchi identities as

R̃a b ∧ eb = D̃T a , D̃R̃a b = 0 , (R̃ab + R̃ba ) = −D̃Qab . (6)

The affine connection 1-form can be decomposed into Riemannian (i.e Levi-Civita) and non-Riemannian parts
according to ω̃ a b = ω a b + N a b . Here ω a b stands for the Levi-Civita (or Christoffel) connection 1-form, satisfying
◦ ◦

ωab = −ωba , and N a b is known as the distortion 1-form. As can be seen, the distortion N quantifies the difference
◦ ◦


between the Levi-Civita connection ω and the general connection ω̃. By replacing the decomposition of the affine
connection 1-form into the definition of the curvature 2-form we can get the split of the latter into, again, Riemannian
and non-Riemannian parts as
◦ ◦
R̃a b = Ra b + DN a b + N a c ∧ N c b , (7)

where the Riemannian part of the curvature 2-form is defined, in terms of the Levi-Civita affine connection 1-form,
as

Ra b := dω a b + ω a c ∧ ω c b
◦ ◦ ◦
(8)

3 The formal definition of the vielbein is with the lower index ea while the dual vielbein is defined as ea . The implicit notation is understood
depending whether the index is up or low. Both objects are related via the dual property ea eb = δba where δba is the Kronecker delta.
4

and the covariant exterior derivative w.r.t. the Levi-Civita connection acting on the distortion 1-form is given by

DN a b = dN a b + ω a c ∧ N c b − ω c b ∧ N a c .
◦ ◦
(9)

The distortion 1-form, moreover, can be decomposed as

N a b = K a b + La b . (10)

The first term K a b that satisfies Kab = −Kba is known as the contortion 1-form while the second term is the
disformation 1-form La b , associated to torsion and nonmetricity respectively.

B. Tensorial notation

When doing computations in differential form language, it will be useful to be able to recover the known expressions
written in tensorial form. We will focus on the relevant quantities that are going to appear throughout the work. To
be named, we start by stating the tensorial expression for the torsion 2-form,
1 a b
T a := T bc e ∧ ec , (11)
2
In the same way, the expression for the curvature 2-form is given by
1 a
R̃a b := R̃ b cd ec ∧ ed , (12)
2

where R̃a b cd are the components of the general Riemann tensor. It will be useful to keep in mind that the exterior
derivative, when acted on an object, can be written as d(·) = ∂d (·)ed . Finally, the distortion 1-form, when written in
its tensorial form, reads

N a b := N a b c ec . (13)

When moving from the orthonormal to coordinate basis, it is important to remember the “tetrad postulate”
˜ µ ea ν = 0,
∇ (14)

which is valid for any connection.


Finally we write down the well known expressions in the coordinate basis for the objects defined above. We start by
recalling the expressions for the curvature, torsion and nonmetricity, which define the Cartan’s structure equations,

R̃σ ρµν = ∂µ Γ̃σ νρ − ∂ν Γ̃σ µρ + Γ̃σ µλ Γ̃λ νρ − Γ̃σ µλ Γ̃λ νρ , (15)


σ σ σ
T µν = Γ̃ µν − Γ̃ νµ , (16)
Qρµν = ∇˜ ρ gµν = ∂ρ gµν − Γ̃λ ρµ gλν − Γ̃λ ρν gµλ , (17)

In the same way, the distortion tensor N can decomposed into the contortion and disformation tensor as

N λ µν = K λ µν + Lλ µν , (18)

where, expressed in terms of torsion and nonmetricity, they read [60]:


1 λ
K λ µν = T µν − Tµ λ ν − Tν λ µ ,

(19)
2
1
Lλ µν Qλ µν − Qµ λ ν − Qν λ µ .

= (20)
2
We now have all the ingredients both in differential form and in tensorial language that will allow us to formulate GR
in the the Teleparallel framework.
5

C. General teleparallel formulation of GR

In GR, i.e. for the Levi-Civita connection, the action is built from the Riemannian Ricci scalar. In the language of
differential forms this object is obtained from the curvature 2-form in such a way that the Einstein-Hilbert action (in
D dimensions) is written as
1
Z

SEH = 2 dDx LGR , (21)

where κ2 = 8πG and
◦ 1 ◦ ◦
LGR = ǫa1 ...aD Ra1 a2 ∧ ea3 ∧ · · · ∧ eaD = R ∗1 , (22)
(D − 2)!
is the GR Lagrangian, with ∗ standing for the Hodge dual operator that defines the orientation of the manifold
1
∗1 = ǫa ...a ea1 ∧ ea2 ∧ · · · ∧ eaD . (23)
D 1 D
Here, ǫa1 ...aD is the totally antisymmetric Levi-Civita symbol (with ǫ12...D = +1). It is important to note that in

Einstein’s GR, both torsion and nonmetricity are zero, T a = Qa b = 0, but the curvature it is not, Ra b 6= 0. On the
other hand, one can play with different combinations of the three quantities to build different classes of spacetimes.
There are three of such a kind that are going to be of interest for this work: General Teleparallel geometries where
Qab 6= 0, T a 6= 0, R̃ab = 0; Teleparallel geometries (torsional) where Tab = 0, T a 6= 0, R̃ab = 0 and; Symmetric
Teleparallel geometries where Qab 6= 0, T a = 0, R̃ab = 0. Teleparallel geometries are characterise by satisfying the
so-called teleparallel condition:

R̃a b = 0 . (24)

Note that from this expression, and using (7), we can obtain a relationship between Ra b and N a b , which will be useful
later on in this work. Also, for teleparallel quantities, we drop any symbol on top.
The above results will be relevant when computing the Gauss-Bonnet invariant in differential forms language,
defined in the following section. If, instead, we want to recover the tensorial result as in GR, it will be necessary to
keep in mind the expression for the Ricci scalar in the General Teleparalel case [61]:
◦ 1 1 1 1 1
R̃ = R + Tλµν T λµν + Tλµν T µλν − T λ λν T µ µ ν + Tλµν Qνλµ + Qλµν Qλµν − Qλµν Qµλν + Qνλ λ Qµ µν
4 2 4 2 2
1 νλ µ λ νµ λ µν

νµ

µ ν

ν µ
− Q λ Qν µ − T λν Q µ + T λν Q µ − 2∇µ T ν + ∇µ Q ν − ∇µ Q ν , (25)
4
from where we can build an action for the General TG equivalent of GR (zero curvature) as [46, 61, 62]:
1 √
Z
SGeneral TG−GR = − 2 dDx −g G , (26)

with
1 1 1 1 1
G = Tλµν T λµν + Tλµν T µλν − T λ λν T µ µ ν + Tλµν Qνλµ + Qλµν Qλµν − Qλµν Qµλν + Qνλ λ Qµ µν
4 2 4 2 2
1
− Qνλ λ Qν µ µ − T λ λν Qνµ µ + T λ λν Qµν µ , (27)
4

which should not be confused with the Gauss-Bonnet invariant G. In such a way the Einstein-Hilbert action (21)
expressed in terms of the Riemannian Ricci scalar
1 √
Z

SEH = 2 dDx −g R (28)

differs only by a boundary term compared to Eq. (26), meaning that Einstein’s field equations arise from both actions.
For this reason the theory is labelled as equivalent to GR. One can then enforce conditions on the geometry such that
one restricts to the torsional or nonmetricity case. If, for example, we set the torsion to zero, we will get the action
for the Symmetric TG case, which will be the dedicated study in Sec. IV.
6

III. GAUSS-BONNET INVARIANT IN A GENERAL TELEPARALLEL GEOMETRY

We now turn our attention to the computation of the Riemannian Gauss-Bonnet invariant in terms of the distortion
tensor Nλµν in the General TG case. This object is defined as Eq. (1) which in differential-form language in D
dimensions can be obtained as
◦ 1 ◦ ◦
G ∗1 = ǫa1 ...aD Ra1 a2 ∧ Ra3 a4 ∧ ea5 ∧ · · · ∧ eaD , (29)
(D − 4)!

in terms of the Riemannian curvature 2-form Ra b , which in turn can be expressed in terms of the distortion tensor
Nλµν under the teleparallel condition Eq. (24). This would be a rather straight-forward enterprise, however we are
also interested in identifying within the resulting expression bulk and boundary contributions. As we will see, this
split is not unique.

A. Derivation in differential form language

We begin first with an approach to the computation that follows closely Ref. [47], were the Metric TG equivalent

of the Gauss-Bonnet invariant G was first derived. The main difference here is that we will let nonmetricity to also
be present, in addition to torsion. For this we must work with the distortion 1-form N ab which has a priori no
symmetry properties under the exchange of a and b, unlike the purely torsional case where the contortion 1-form K ab
is anti-symmetric. This requires some care when going through the derivation, and expectedly, some new contributions
will arise from the purely symmetric parts of the distortion 1-form, which are uniquely associated with nonmetricity.
Notice that nonmetricity also contributes to the anti-symmetric part of N ab .
Let us start from the Lagrangian containing the Gauss-Bonnet invariant constructed with the general curvature
1
L̃2 = ǫa ...a R̃a1 a2 ∧ R̃a3 a4 ∧ ea5 ∧ · · · ∧ eaD = G̃ ∗1 , (30)
(D − 4)! 1 D

and replace both factors of the general curvature 2-form R̃a b by the Riemmanian curvature form Ra b and the distortion
1-form N a b using (7). One obtains

(D − 4)!L̃2 = (D − 4)!L2 + I1 + 2I2 + 2I3 + 2I4 + I5 , (31)

where

I1 = ǫa1 ...aD N a1 c ∧ N ca2 ∧ N a3 d ∧ N da4 ∧ ea5 ∧ · · · ∧ eaD ,



I2 = ǫa1 ...aD Ra1 a2 ∧ N a3 d ∧ N da4 ∧ ea5 ∧ · · · ∧ eaD ,

I3 = ǫa1 ...aD DN a1 a2 ∧ N a3 d ∧ N da4 ∧ ea5 ∧ · · · ∧ eaD ,
◦ ◦
I4 = ǫa1 ...aD DN a1 a2 ∧ Ra3 a4 ∧ ea5 ∧ · · · ∧ eaD ,
◦ ◦
I5 = ǫa1 ...aD DN a1 a2 ∧ DN a3 a4 ∧ ea5 ∧ · · · ∧ eaD , (32)

and
◦ 1 ◦ ◦ ◦
L2 = ǫa1 ...aD Ra1 a2 ∧ Ra3 a4 ∧ ea5 ∧ · · · ∧ eaD = G ∗1 , (33)
(D − 4)!

which is what we ultimatelly want to compute. Using properties (A1), (A3) and (A4), we readily identify the
contributions to what will be the boundary term. In particular, we can rewrite
 ◦

I4 = d ǫa1 ...aD N a1 a2 ∧ Ra3 a4 ∧ ea5 ∧ · · · ∧ eaD , (34)

and

 ◦
 ◦
I5 = D ǫa1 ...aD N a1 a2 ∧ DN a3 a4 ∧ ea5 ∧ · · · ∧ eaD + ǫa1 ...aD N a1 a2 ∧ D2 N a3 a4 ∧ ea5 ∧ · · · ∧ eaD
 ◦
 ◦
= d ǫa1 ...aD N a1 a2 ∧ DN a3 a4 ∧ ea5 ∧ · · · ∧ eaD + 2ǫa1 ...aD N a1 a2 ∧ Ra3 c ∧ N [ca4 ] ∧ ea5 ∧ · · · ∧ eaD , (35)
7

where we integrated by parts in the first line and used property (A6) in the second line of Eq. (35). Hence, the
Lagrangian in Eq. (31) can be further decomposed as

(D − 4)!L̃2 = (D − 4)!L2 + I1 + 2(I3 + I6 ) + dB , (36)
where
◦ ◦

I6 = ǫa1 ...aD Ra1 a2 ∧ N a3 d ∧ N da4 + N a1 a2 ∧ Ra3 c ∧ N [ca4 ] ∧ ea5 ∧ · · · ∧ eaD ,
 ◦ ◦

B = ǫa1 ...aD 2N a1 a2 ∧ Ra3 a4 + N a1 a2 ∧ DN a3 a4 ea5 ∧ · · · ∧ eaD . (37)

Therefore, I4 and one part of I5 gave rise to the boundary term B, while I2 and the second part of I5 have been
combined into I6 . Next, we realize that
  
◦ ◦ ◦
2(I3 + I6 ) = 2ǫa1 ...aD DN a1 a2 + Ra1 a2 ∧ N a3 d ∧ N da4 + N a1 a2 ∧ Ra3 c ∧ N [ca4 ] ∧ ea5 ∧ · · · ∧ eaD
 ◦

= 2ǫa1 ...aD R̃a1 a2 ∧ N a3 d ∧ N da4 − N a1 c ∧ N ca2 ∧ N a3 d ∧ N da4 + N a1 a2 ∧ Ra3 c ∧ N [ca4 ] ∧ ea5 ∧ · · · ∧ eaD
= 2J0 − 2I1 + 2J1 , (38)
where we used (A5) in the second line of Eq. (38) and defined

J0 = ǫa1 ...aD R̃a1 a2 ∧ N a3 d ∧ N da4 ∧ ea5 ∧ · · · ∧ eaD ,



J1 = ǫa1 ...aD N a1 a2 ∧ Ra3 c ∧ N [ca4 ] ∧ ea5 ∧ · · · ∧ eaD , (39)
such that the Lagrangian in Eq. (36) reads

(D − 4)!L̃2 = (D − 4)!L2 − I1 + 2(J0 + J1 ) + dB . (40)
Finally, we use Eq. (A12) to replace the remaining factor of the Riemmanian curvature in J1 and obtain
 
J1 = ǫa1 ...aD N a1 a2 ∧ R̃a3 c + N a3 d ∧ N d c − DN a3 c ∧ N [ca4 ] ∧ ea5 ∧ · · · ∧ eaD
= Jˆ0 + J2 − J3 , (41)
where we defined
Jˆ0 = ǫa1 ...aD N a1 a2 ∧ R̃a3 c ∧ N [ca4 ] ∧ ea5 ∧ · · · ∧ eaD ,
J2 = ǫa1 ...aD N a1 a2 ∧ N a3 d ∧ N d c ∧ N [ca4 ] ∧ ea5 ∧ · · · ∧ eaD ,
J3 = ǫa1 ...aD N a1 a2 ∧ DN a3 c ∧ N [ca4 ] ∧ ea5 ∧ · · · ∧ eaD , (42)
such that we arrive at

(D − 4)!L̃2 = (D − 4)!L2 − I1 + 2(J0 + Jˆ0 ) + 2J2 − 2J3 + dB . (43)

Enforcing the teleparallel condition R̃ab = 0 ≡ Rab , leads to L̃2 = 0 on the l.h.s. and J0 = 0 = Jˆ0 on the r.h.s. of Eq.
(43), yielding

(D − 4)!L2 = I1 − 2J2 + 2J3 − dB , (44)
or, explicitly
"
◦ 1  ◦ ◦

G ∗1 = ǫa1 ...aD −d 2N a1 a2 ∧ Ra3 a4 ∧ ea5 ∧ · · · ∧ eaD + N a1 a2 ∧ DN a3 a4 ∧ ea5 ∧ · · · ∧ eaD
(D − 4)!

+ 2N a1 a2 ∧ DN a3 c ∧ N [ca4 ] − 2N a1 a2 ∧ N a3 d ∧ N d c ∧ N [ca4 ]
#

a1 f a2 a3 ha4 a5 aD
+N f ∧ N ∧N h∧N ∧ e ∧ ···∧ e . (45)

This is the main result of this section, expressed in differential form language. In the next subsection we provide an
alternative derivation, while in Sec. III C we translate the result back to tensorial notation, given in Eq. (56).
8

B. Alternative split into bulk and boundary terms

We can take a different approach and impose the teleparallel condition (24) from the beginning. It is instructive
to do so because it will show that the split between bulk and boundary terms is not unique, even in D 6= 4 (in D = 4
the Gauss-Bonnet invariant is itself a total derivative).

We start by replacing Eq. (7) with R̃a b = Ra b = 0 in one of the factors of Rab in Eq. (29),
◦ 1 h ◦ ◦ ◦
i
G ∗1 = ǫa1 ...aD −DN a1 a2 ∧ Ra3 a4 − N a1 f ∧ N f a2 ∧ Ra3 a4 ∧ ea5 ∧ · · · ∧ eaD . (46)
(D − 4)!

Here, again by using properties (A1), (A3) and (A4), we see that the first term in Eq. (46) is a total derivative
◦ ◦ ◦
h ◦
i
ǫa1 ...aD DN a1 a2 ∧ Ra3 a4 ∧ ea5 ∧ · · · ∧ eaD = D ǫa1 ...aD N a1 a2 ∧ Ra3 a4 ∧ ea5 ∧ · · · ∧ eaD
h ◦
i
= d ǫa1 ...aD N a1 a2 ∧ Ra3 a4 ∧ ea5 ∧ · · · ∧ eaD
h ◦
i
= ǫa1 ...aD d N a1 a2 ∧ Ra3 a4 ∧ ea5 ∧ · · · ∧ eaD , (47)

and therefore we have


" #
◦ 1  ◦
 ◦
G ∗1 = ǫa ...a −d N a1 a2 ∧ Ra3 a4 ∧ ea5 ∧ · · · ∧ eaD − N a1 f ∧ N f a2 ∧ Ra3 a4 ∧ ea5 ∧ · · · ∧ eaD . (48)
(D − 4)! 1 D

We now replace Ra3 a4 once more time in the second term,
"
◦ 1  ◦

G ∗1 = ǫa1 ...aD −d N a1 a2 ∧ Ra3 a4 ∧ ea5 ∧ · · · ∧ eaD
(D − 4)!
#
 ◦

a1 f a2 a3 a4 a1 f a2 a3 ha4 a5 aD
+ N f ∧N ∧ DN +N f ∧N ∧N h ∧N ∧e ∧··· ∧e . (49)

This very short and simple procedure leads to a different way to split the Gauss-Bonnet invariant into a bulk term
and a boundary term. In Appendix B we show that this expression is equivalent to Eq. (45), while the corresponding
tensorial expressions can be found in Eq. (57).

In general, the presence of mixed terms schematically of the form N ∧N ∧ DN , with at least one derivative acting on
a specific factor, implies there is no absolute split into bulk and boundary pieces, as one can always trade the position
of the derivative at the cost of generating a new boundary contribution. In D = 4 dimensions the Gauss-Bonnet
invariant becomes a topological term, and in particular, a purely boundary term in the form of the divergence of a
non-tensorial quantity [63].

C. Tensorial expressions


We are now interested in reading off the scalar G from our results of Eqs. (45) and (49), which are D-forms. In
order to do so, we express both the contortion 1-form explicitly as in Eq. (13), as well as the curvature 2-form in
terms of the Riemann tensor as given in Eq. (12). From the resulting explicit expression we are able to collect as an
overall factor proportional to the volume element D-form, Eq. (23),

ea ∧ eb ∧ ec ∧ ed ∧ ea5 ∧ · · · ∧ eaD = ǫabcda5 ...aD ∗1 , (50)

and extract the desired scalar expression. For our purposes we will also need to use
1
ǫa ...a ǫabcda5 ...aD = δaa1bac2da3 a4 . (51)
(D − 4)! 1 D

Special care needs to be taken when a term involves the covariant exterior derivative w.r.t. the teleparallel con-
nection; more specifically, DN ab which appears in Eq. (45). When we write the distortion 1-form in terms of the
9

corresponding tensor, Eq. (13), we need to differentiate also the basis ec . Thus - from its definition in Eq. (3) - a term
involving the torsion will emerge:
 
1 a
DN b = ∇d N bc + N be T dc ed ∧ ec ,
a a e
(52)
2

where we also used Eq. (11). This can again be expressed in terms of the anti-symmetric part of the distortion tensor
w.r.t. the last two indices T a bc = −2N a [bc] , where we are using the same convention as Ref. [47]. This is not to be
mistaken with other anti-symmetrizations we have used so far involving the first two indices instead. Then, accounting
also for the raising of an index under the covariant exterior derivative associated to the teleparallel connection D, Eq.
(A9), we obtain

DN ab = − ∇d N ab c + N ab h N h cd ec ∧ ed ,

(53)

where we have dropped the anti-symmetrization square brackets since c and d contract with the anti-symmetric ec ∧ed .
(T,Q)
With this we can immediately obtain the bulk term which we call 1 TG from Eq. (45),
h
1 (T,Q)
TG = δµµ1νµρ2 µσ3 µ4 N µ1 αµ N αµ2 ν N µ3 βρ N βµ4 σ − 2N µ1 µ2 µ N µ3 αν N α βρ N [βµ4 ] σ
+2gαβ N µ1 µ2 µ N [µ3 α] ν N βµ4 γ N γ ρσ + 2gαβ N µ1 µ2 µ N [µ3 α] ν ∇σ N βµ4 ρ
i
+4gαβ gγδ N µ1 µ2 µ N [µ3 α] ν N µ4 γ ρ N (δβ) σ , (54)

where we have also switched from the orthonormal basis back to the coordinate basis, for which it is useful to remember
the “tetrad postulate”, Eq. (14).
Similar manipulations can be applied to the corresponding boundary term of Eq. (45). We first apply Eq. (A5)
once in order to eliminate the explicit appeareance of DN ab , and then by also making use of the properties (A4) and
(A1) one arrives at


  
1 (T,Q) 1 µνρσ µ1 µ2 µ3 λµ4 1 ◦ µ3 µ4
BG = √ ∂µ −g δµ1 µ2 µ3 µ4 N ν N λρ N σ − R ρσ . (55)
−g 2

The Gauss-Bonnet invariant is the sum of Eqs. (54) and (55),



(T,Q) (T,Q)
G = 1 TG + 1BG . (56)

As discussed previously, the split into bulk and boundary terms is not unique, and in fact we provided an alternative
one in the previous subsection, given in Eq. (49). The corresponding tensorial expressions are

(T,Q) (T,Q)
G = 2 TG + 2BG , (57)

with
h ◦
i
2 (T,Q)
TG = δµµ1νµρ2 µσ3 µ4 N µ1 αµ N αµ2 ν N µ3 βρ N βµ4 σ − N µ1 αµ N αµ2 ν ∇σ N µ3 µ4 ρ , (58)
1 1 h√ ◦
i
2 (T,Q)
BG = − √ ∂µ −g δµµ1νµρ2 µσ3 µ4 N µ1 µ2 ν Rµ3 µ4 ρσ . (59)
2 −g
(T,Q) (T,Q)
Theories constructed by breaking down the Gauss-Bonnet invariant and coupling TG and BG independently
will be sensitive to the choice of splitting and will in general not be equivalent. They may coincide accidentally when
(T,Q) i (T,Q)
evaluated in spacetimes with enough symmetries. It is clear that both invariants iTG , BG for each set are
independently boundary terms in D = 4. This can be easily seen from Eq. (56) and Eq. (55) (and Eq. (57) with

Eq. (59)), since the l.h.s. of the split G is a boundary term in D = 4 and both Eq. (55) and Eq. (59) are boundary
terms as well. Or in other words,

i (T,Q) (T,Q)
TG = G − iBG = boundary term . (60)

It is worth saying that then, this property will hold for any version of TG.
10

D. Remarks about the known Metric Teleparallel case

It is worth comparing our general results with the purely Metric TG case originally presented in Ref. [47]. Since
our derivation in Sec. III A follows theirs, our first set of scalars straightforwardly reduces to the known result when
replacing Nµνρ → Kµνρ (vanishing non-metricity) and recalling its anti-symmetry in the first two indices,

(T ) (T )
G = 1 TG + 1BG , (61)

with
1 (T )
= δµµ1νµρ2 µσ3 µ4 K µ1 αµ K αµ2 ν K µ3 βρ K βµ4 σ + 2K µ1 µ2 µ K µ3 αν K αµ4 γ K γ ρσ + 2K µ1 µ2 µ K µ3 αν ∇σ K αµ4 ρ , (62)
 
TG

  
1 (T ) 1 µνρσ µ1 µ2 µ3 λµ4 1 ◦ µ3 µ4
BG = √ ∂µ −g δµ1 µ2 µ3 µ4 K ν K λρ K σ − R ρσ . (63)
−g 2

Notice that some terms of the bulk part have been evaluated to zero due to the anti-symmetry, and that the covariant
derivative is with respect to the Teleparallel (torsional) connection.
Interestingly, we have also derived an alternative split in Sec. III B which was not considered in the literature before,
but it is as legitimate as any other. When evaluated in the Metric TG case it reads

(T ) (T )
G = 2 TG + 2BG , (64)

with
h ◦
i
2 (T )
TG = δµµ1νµρ2 µσ3 µ4 K µ1 αµ K αµ2 ν K µ3 βρ K βµ4 σ − K µ1 αµ K αµ2 ν ∇σ K µ3 µ4 ρ , (65)
1 1 h√ ◦
i
2 (T )
BG = − √ ∂µ −g δµµ1νµρ2 µσ3 µ4 K µ1 µ2 ν Rµ3 µ4 ρσ . (66)
2 −g

This second novel decomposition may lead to significant differences when compared to known results.

IV. SYMMETRIC TELEPARALLEL GAUSS-BONNET GRAVITY

In this section, we focus our study in a particular geometry purely represented by nonmetricity, known as Symmetric
TG. Then, we show how the Symmetric TG Gauss-Bonnet invariants behave, and find examples of different spacetimes
as well as possible theories to consider.

A. Symmetric Teleparallel Gravity and its Gauss-Bonnet invariant

Symmetric TG is a particular subset of General TG where torsion is absent, hence gravity being purely represented
by nonmetricity. In this case, by taking Eq. (25), we find that the Levi-Civita scalar can be split as

R = Q + BQ , (67)

where Q is the so-called nonmetricity scalar defined as


1 1 1 1
Q=− Qλµν Qλµν + Qλµν Qµνλ + Qµ Qµ − Qµ Q̂µ , Qµ ≡ Qµν ν , Q̂µ ≡ Qνµ ν , (68)
4 2 4 2
and BQ a boundary term given by

BQ = ∇µ (Q̂µ − Qµ ) . (69)

Then, the General Teleparallel equivalent of GR action written in Eq. (26) reduces to the Symmetric Teleparallel
equivalent of GR, whose action reads:

Z
SSTEGR = d4x −g Q . (70)
11

One can also consider modified theories of gravity within this type of geometries by changing the above action. In
this formalism, since both torsion and curvature are vanishing, the connection can be always written as
∂xα
Γα µν = ∂µ ∂ν ξ λ , (71)
∂ξ λ
where ξ α is associated to diffeomorphisms (as a Stückelberg field). Then, the connection only has a maximum of
4 d.o.f.. However, it is always possible to find a gauge, known as the coincident gauge, such that the above vector
trivializes the connection, i.e. Γα µν = 0. In that gauge, ∇µ = ∂µ , and all the d.o.f. of any symmetric teleparallel
theory would be encoded in the metric, at the price of loosing diffeomorphism invariance.
Now, let us focus on the Gauss-Bonnet invariant within Symmetric TG. This means replacing N λ µν → Lλ µν in the
expressions obtained in the previous sections. Explicitly, the first set of Symmetric TG Gauss-Bonnet scalar invariants
using the expressions (54)-(55) become

(Q) (Q)
G = 1 TG + 1BG , (72)

where
h
1 (Q)
TG = δµµ1νµρ2 µσ3 µ4 Lµ1 αµ Lαµ2 ν Lµ3 βρ Lβµ4 σ − 2Lµ1 µ2 µ Lµ3 αν Lα βρ L[βµ4 ] σ
i
+2gαβ Lµ1 µ2 µ L[µ3 α] ν ∇σ Lβµ4 ρ +4gαβ gγδ Lµ1 µ2 µ L[µ3 α] ν Lµ4 γ ρ L(δβ) σ , (73)

  
1 (Q) 1 1 ◦
BG = √ ∂µ −g δµµ1νµρ2 µσ3 µ4 Lµ1 µ2 ν Lµ3 λρ Lλµ4 σ − Rµ3 µ4 ρσ . (74)
−g 2

while using the second split given by (58)-(59) we find



(Q) (Q)
G = 2 TG + 2BG , (75)

h ◦
i
2 (Q)
TG = δµµ1νµρ2 µσ3 µ4 Lµ1 αµ Lαµ2 ν Lµ3 βρ Lβµ4 σ − Lµ1 αµ Lαµ2 ν ∇σ Lµ3 µ4 ρ , (76)
1 h√ ◦
i 1 ◦ h ◦
i
2 (Q)
BG = − √ ∂µ −g δµµ1νµρ2 µσ3 µ4 Lµ1 µ2 ν Rµ3 µ4 ρσ = − ∇µ δµµ1νµρ2 µσ3 µ4 Lµ1 µ2 ν Rµ3 µ4 ρσ . (77)
2 −g 2

Note that the covariant derivative appearing in Eq. (73) is computed with the Symmetric TG connection. Let us also
stress here that the first term on the second line of Eq. (54) is identically zero for the Symmetric Teleparallel case,
(Q)
which is why now 1 TG only contains four terms. In the next sections, we will study these scalars in more detail.

B. Spacetimes examples for the Symmetric Teleparallel Gauss-Bonnet invariant

In this section, we briefly discuss the evaluation of the Symmetric TG Gauss-Bonnet invariants in FLRW cosmology
and spherical symmetry.

1. FLRW cosmology

The line-element for FLRW can be written in spherical coordinates as

a(t)2 dr2
ds2 = −N (t)2 dt2 + + r2 (dθ2 + sin2 θ dφ2 ) , (78)
1 − kr2
where k is the spatial curvature and N (t) and a(t) are the lapse function and scale factor, respectively. In Ref. [64], it
was found that depending on how one solves the curvatureless and torsionless conditions, one can get three different
branches in flat FLRW cosmology having different connections. However, for the non-flat case k 6= 0, there is a unique
branch satisfying those conditions. The first set of Symmetric TG Gauss-Bonnet invariants given by (73)-(74) for
that case become

1 (Q) 6(K − 2H)2 Ḣ 12H(K − H)K̇  k 2 h 6K Ḣ + 12(K − H)K̇


TG = + + 6H 2 −6HK + 4H 2 + 3K 2 + 4
N N a K 3N
12
 
24H H2

k h K − 12 Ḣ + 12 1 − K 2 K̇ 12H −HK + H 2 + K 2 i

6H(H − 2K) i
− + 2 + , (79)
K2 a N K
1 (Q) 6K(4H − K)Ḣ 12H(H − K)K̇ 6k 2   12k 2 H K̇ 6k 2 H(H − 2K)
BG = + + 18H 2 K(2H − K) − 4 2 Ḣ + 2K̇ + 4 3 +
N N ! ! a K N a K N a4 K 2
2
12k 3Ḣ K̇ 12k 2H Ḣ 12kH K̇
+ 2 − HK + 3H 2 − − 2 + H3 + 2 2 , (80)
a N N a K N a K N

where dots are differentiation with respect to t, H = ȧ/(aN ) is the Hubble parameter and K represents an extra
d.o.f. coming from the connection. The above quantities give us the expected value for the Riemannian Gauss-Bonnet
invariant in FLRW cosmology:
! !
1 (Q) 1 (Q)

2 Ḣ 2 24k Ḣ 2
TG + BG = G = 24H +H + 2 +H . (81)
N a N

Similarly, the second set of Symmetric TG Gauss-Bonnet invariants expressed by (76)-(77) for the non-flat FLRW
case gives:
 2 
H
Ḣ + H

12H(K − H)Ḣ 6H 2
K̇ 6k h2 1 − K K 2 + 1 K̇ 
H2
i
2 (Q)
TG = + + 6H 3 (3K − 2H) + 2 +H − + 2H + K
N N a N K
2 h
6k K̇ i
+ 4 H+ , (82)
a K KN  2 
2(H+K)Ḣ H
12H(3H − K)Ḣ 6H 2 K̇ 6k h K − K 2 + 1 K̇  2
H
i
2 (Q) 3
BG = − + 18H (2H − K) + 2 +H + 2H − K
N N a N K
6k 2 h K̇ i
− 4 H+ , (83)
a K KN
which also correctly reproduces the Riemannian Gauss-Bonnet invariant (81). Since the scalars are different for each
set, in principle, one could have different dynamics for a given modified Symmetric TG theory.
Let us now explore the situation in flat FLRW where we have three different branches. Following the notation for
the branches introduced in Ref. [64], we obtain that the first set of Symmetric TG Gauss-Bonnet scalars become
 !
 24H 2 Ḣ 2
+H , First branch





 N
1 (Q) 6(2H + K)2 Ḣ 12H(H + K)K̇
TG = (84)
+ 6H 2 6HK + 4H 2 + 3K 2 , Second branch

+
N N



2
 6(K − 2H) Ḣ + 12H(K − H)K̇ + 6H 2 −6HK + 4H 2 + 3K 2  , Third branch



N N
0 , First branch



 6K(4H + K)Ḣ
 12H(H + K)K̇
1 (Q)
BG = − − − 18H 2 K(2H + K) , Second branch (85)
N N
6K(4H − K) Ḣ 12H(H − K) K̇


2
+ + 18H K(2H − K) , Third branch


N N
and for the second set of Symmetric TG Gauss-Bonnet scalars, we obtain
 !
2 Ḣ 2
−12H +H , First branch






 N
2 (Q) 12H(H + K)Ḣ 6H 2 K̇
TG = (86)
− − − 6H 3 (2H + 3K) , Second branch
N N



2
 12H(K − H)Ḣ + 6H K̇ + 6H 3 (3K − 2H) ,


Third branch

N N
13
 !
2 Ḣ 2
36H +H , First branch






 N
2 (Q) 12H(3H + K)Ḣ 6H 2 K̇
BG = (87)
+ + 18H 3 (2H + K) , Second branch
N N



2
 6H(6H − 2K)Ḣ − 6H K̇ + 18H 3 (2H − K) . Third branch



N N
(Q) (Q)
Clearly, for all the branches the combination i TG + i BG gives the same value as (81) with k = 0, and, as expected,
K which is related to the new d.o.f. related to the connection drops out. It is interesting to notice that the scalars for
the first set, first branch coincide with the Metric TG case [48]. Furthermore, it is also known that in flat FLRW, the
first branch, the scalar Q = 6H 2 also coincides with the torsional scalar case T = 6H 2 . That indicates that the first
set, first branch has similar features in both teleparallel versions. Clearly, for the second set, also the first branch has
(Q) (Q)
a similar feature since 2 BG = −3 2 TG , and then, effectively, the dynamics of any theory constructed from those
terms would be equivalent to their torsional counterpart at the background level. Let us emphasise here again that
our constructed scalars will not have any dynamics in cosmology (or any spacetime) if they appear linearly in an
action since they are boundary terms.

2. Spherical Symmetry

As another important spacetime example for displaying the Symmetric TG Gauss-Bonnet scalars, we will choose
the following spherically symmetric metric
1
ds2 = −A(r)dt2 + dr2 + M (r)2 (dθ2 + sin2 θdφ2 ) , (88)
B(r)
where A, B and M are arbitrary functions depending on the radial coordinate. In Ref. [65], it was found that there are
two branches for spherical symmetry satisfying the torsionless and flat conditions. In Ref. [66], it was found that in
the second branch, there is a particular case in which there are exact scalarized solutions for a particular Symmetric
TG theory. Let us use that particular situation as an example to show the Symmetric TG Gauss-Bonnet scalars
within spherical symmetry. Choosing the same convention as that reference, we find that the first set of Symmetric
TG Gauss-Bonnet invariants for the c = k = 0 case is
BA′2 2BM ′2 + 1

1 (Q) 2BΓr θθ A′2 M ′ 2B 2 A′2 M ′ B 2 A′2 (Γr θθ )2 A′2
TG = − − − − −
A2 M 3 A2 M Γr θθ A2 M 2 2A2 (Γr θθ )2 2A2 M 4
t ′ t ′ t ′ ′t t ′′ 2 ′ ′r
Γ θθ (A (4BΓ θθ M − M (Γ θθ B + 4BΓ θθ )) − 2BM Γ θθ A ) 2B A Γ θθ
+ −
2M 5 A(Γr θθ )3
 
 ′ ′ ′ ′r
 (Γr )2 −2M A′′ + MA′ B ′ + 4A′ M ′
B A (3M B + 4BM (1 − 2Γ θθ )) θθ B
+ 2BA′′ + −
2A(Γr θθ )2 M 2AM 5
2M A B + B 4M (M A + A M (3M B M + 2Γ θθ )) − (Γ θθ ) A + 8B 2 M 2 M ′ (A′′ M ′ + 2A′ M ′′ )
2 ′ ′ ′′ ′ ′ ′ ′ ′r t 2 ′2

+
2AM 4
2Γ θθ A Γ θθ − 2BM + M (2BA M + A (B ′ M ′ + 2BM ′′ ))
r ′ ′r ′2 ′′ ′ ′
 
+
AM 4
2B (2BM A′ M ′′ + M ′ (2BM A′′ + A′ (3M B ′ + 2BM ′ )))
+ , (89)
AM 2 Γr θθ
1 (Q) 2BΓr θθ A′2 M ′ 2B 2 A′2 M ′ 3BA′2 B 2 A′2 (Γr θθ )2 A′2
BG = + + + +
A2 M 3 M A2 Γr θθ A2 M 2 2A2 (Γr θθ )2 2A2 M 4
Γ θθ (2BM Γ θθ A + A (M (Γ θθ B + 4BΓ θθ ) − 4BΓ θθ M ′ )) (Γr θθ )2 (A′ (M B ′ + 4BM ′ ) − 2BM A′′ )
t t ′′ ′ t ′ ′t t
+ +
2M5 2BAM 5
′ t 2 ′ ′ ′r 2 ′

A B (Γ θθ ) A − 8M M Γ θθ − 6M B A (4BM (2Γ θθ − 1) − 3M B ′ )
′ ′ ′r
 
B ′′
+ + − 2BA
2AM 4 2A(Γr θθ )2 M
2Γr θθ A′ Γ′r θθ − 2BM ′2 + M (2BA′′ M ′ + A′ (B ′ M ′ + 2BM ′′ ))
 

AM 4
−6BM A′ B ′ M ′ − 4B 2 M A′′ M ′ + A′ M M ′′ + M ′2

′′ 2 ′ ′r
6BA 2B A Γ θθ
− + + , (90)
AM 2 A(Γr θθ )3 AM 2 Γr θθ
14

where primes are differentiation with respect to r. We can easily see that, as expected

4BA′′ BM ′2 − 1 + A′ B ′ 6BM ′2 − 2 + 8B 2 M ′ M ′′ 2BA′2 BM ′2 − 1


   

1 (Q) 1 (Q)
T G + BG = G = − . (91)
AM 2 A2 M 2
Similar to the cosmological case, the connection components only disappear (in a non-trivial way) when one takes the
combination that creates the Riemannian Gauss-Bonnet invariant. One can also compute similar expressions for the
second set of scalars, finding that they are also different than the above expressions. We will refrain from writing down
their form for simplicity. As a possible application of using those expressions, one can use the minisuperspace approach
to find out the spherically symmetric field equations of a particular Symmetric Teleparallel theory of gravity. This can
be done by introducing the vector ξµ to replace the connection components and then by computing the Euler-Lagrange
equations with respect to {A, B, M } to obtain the metric equations and ξµ , for the connection equations.

C. New theories constructed with the Symmetric Teleparallel Gauss-Bonnet invariants

One of the simplest modifications of the Symmetric Teleparallel equivalent of GR that has been studied in the last
years relies on promoting Q from appearing linearly as in Eq. (70) to an arbitrary function thereof, explicitly

Z
Sf (Q) = d4x −g f (Q) . (92)

This theory has some similar features to its analogous torsional teleparallel version called f (T ) gravity [46, 67, 68] since
it has second-order field equations of motions which are dynamically quite different from the Riemannian extension

f (R) which is a fourth-order theory in the metric. Moreover, f (Q) theory has been analysed in cosmology finding that
it is compatible with certain cosmological observations [69–71]. However, it has been found that the first cosmological

branch may potentially suffer from strong-coupling problems [72]. Further, since Q is related to R as (67), other more

general theories such as f (Q, BQ ) have been considered in order to include f (R) gravity as a subcase of it [73–75].
One can then generalise the above theory by considering the Symmetric TG Gauss-Bonnet invariant derived in
(Q) (Q)
the previous section. Obviously, a linear combination of either TG or BG would not change the dynamics of the
4
above theory since they are boundary terms in 4-dimensions . Therefore, if one is interested in constructing modified
theories of gravity with them, one could for example consider a non-linear action as with the four mentioned scalars:

Z
(Q) (Q)
Sf (Q,BQ ,T (Q) ,B (Q) ) = d4x −g f (Q, BQ , TG , BG ) , (93)
G G

(Q) (Q)
which as a subcase contains the Riemannian modified Gauss-Bonnet theory by setting f (Q, BQ , TG , BG ) = f (Q +
◦ ◦
(Q) (Q)
BQ , TG + BG ) = f (R, G) that has been widely studied (see for example [3–6]). The theory proposed above, which
(T ) (T )
- in general - is a fourth-order theory in the metric, is also similar to the torsional f (T, B, BG , TG ) introduced
in Refs. [47, 48]. As mentioned in the previous section, the first branch of cosmology in flat FLRW has the same
(T ) (T )
scalars as its torsional counterpart. That means that at least at the background level, the theory f (T, B, TG , BG )
(Q) (Q)
is equivalent to the first branch of f (Q, BQ , TG , BG ) for flat FLRW cosmology (they have the same flat FLRW
(T ) (T )
cosmological equations). The reason for that is the fact that for the first branch of flat FLRW, f (T, B, TG , BG ) =
(T ) (Q) (Q) (Q)
f (T, B, TG ) = f (Q, BQ , TG , BG ) = f (Q, BQ , TG ) (see the scalars and discussion in Sec. IV B 1). However, this
equivalence might be broken at the level of cosmological perturbations. Actually, f (T, B) only has 3 d.o.f. around
flat FLRW [76] while f (Q, BQ ) has at least 4 d.o.f. for the first branch [72].
Another route to construct a modified Symmetric TG with a non-trivial contribution from the Symmetric TG
Gauss-Bonnet invariants is by introducing a scalar field and couple it non-minimally as
1 √ h 1
Z i
(Q) (Q)
SSTsGB = 2 d4x −g Q − β ∂µ ψ∂ µ ψ + α1 G1 (ψ)TG + α2 G2 (ψ)BG . (94)
2κ 2

4 For simplicity in the notation, in this section we removed the numbers in the scalars since the discussion does not change for any set of
scalars considered.
15

The above-proposed theory leads to second-order field equations for all the fields (see Appendix C), and contains
the so-called scalar Gauss-Bonnet gravity theory in the Riemannian sector by taking the limit α1 G1 (ψ) = α2 G2 (ψ)

which would introduce a coupling of the form f (ψ)G. The latter has been widely studied in the context of black hole
physics since it predicts the existence of scalarized black holes with a spontaneous scalarization process [21, 22, 77–80].
Further, two recent studies found the existence of similar solutions in the torsional teleparallel case [56, 57] and then,
the above Symmetric TG theory could be also potentially interesting to study within that direction. Note that the
theory of Eq. (94) is part of the so-called Symmetric TG Horndeski theory proposed in Ref. [81], only when one takes
(Q) (Q)
the second set 2 TG , 2BG given by Eqs. (76)-(77). Although, as mentioned above, both sets satisfy the requirement of
second-order field equations, in that study the authors also made the simplifying assumption that there are only terms
(Q)
with up to quadratic contractions of nonmetricity, which includes G2 (ψ) 2BG as part of the theory5 . Furthermore,

(Q)
since the couplings, Eq. (94) can be always recast as α1 G1 (ϕ)G + αG(ϕ) 2BG and the Symmetric TG Horndeski

theory contains the standard Levi-Civita Horndeski part (which contains α1 G1 (ϕ)G), then, the above theory is part of
(Q) (Q)
it with the second set. On the other hand, since the two scalars in the first set, 1 TG , 1BG given by Eqs. (73)-(74),
always contain cubic or quartic contractions of nonmetricity, they are not included in the Symmetric TG Horndeski
construction of Ref. [81].
It is, of course, trivial to extend such Symmetric TG theories to the General TG case just by replacing the Symmetric
(Q) (Q) (T,Q) (T,Q)
TG Gauss-Bonnet invariants by the general ones from Sec. III, i.e., TG , BG → TG , BG . Those theories will
contain both the Metric Gauss-Bonnet theories introduced in Refs. [47, 48, 56] and also the above proposed Symmetric
TG ones.

V. SUMMARY OF RESULTS

In this paper, we have found how the Riemannian Gauss-Bonnet invariant is related to torsion and nonmetricity
in the General TG framework. We found that it is possible to define newly constructed Teleparallel Gauss-Bonnet
invariants that are always boundary terms in D = 4 dimensions from which the Riemannian Gauss-Bonnet emerges.
Our result coincides with the result found in Ref. [47] when nonmetricity is vanishing, which is related to torsional
Gauss-Bonnet gravity. We then focus on the new terms that are only related to nonmetricity and express all our
quantities in Symmetric TG.
We found that the way of splitting the Riemannian Gauss-Bonnet invariant in terms of teleparallel quantities is
not unique. We present two different ways of splitting it with different bulk and boundary terms. After this, we
focused on the Symmetric TG case and we analysed how these scalars behave in non-flat FLRW cosmology and also
in spherical symmetry. Since those terms are boundary terms in 4-dimensions, if we want them to acquire dynamics,
one would need to consider them in an action non-linearly. We did this by formulating two new Symmetric TG
theories constructed with the Gauss-Bonnet invariants where we allowed more general modifications of the STEGR
action (70). One of those theories contains as a special case, the so-called modified Gauss-Bonnet gravity theory
◦ ◦
constructed in the Riemannian sector and where the action is constructed as f (R, G). The second proposed theory
contains the so-called scalar-Gauss-Bonnet theory where the Gauss-Bonnet acquires dynamics by coupling it with a
scalar field.
In future studies, it would be interesting to analyse further the proposed theories with the Symmetric TG invariants.
They might serve as a starting point to study cosmology in Symmetric TG and also to explore new routes in the process
of understanding the difference between the torsional and nonmetricity versions of Teleparallel gravity. Furthermore,
scalarized black holes with spontaneous scalarization process should exist in (94), but potentially, they might have
different features as their Riemannian or torsional counterparts due to the fact that now, there are more possible
ways of solving the teleparallel condition in spherical symmetry which introduces richer dynamics. Furthermore, the
connection components are still dynamical in this formalism, and one could have different black hole configurations
while having different Symmetric Teleparallel connections. It would be interesting to analyse how the connection
affects the dynamics of such black holes and also their thermodynamics.
Another route that might lead to interesting applications is to follow a similar approach as [82] and construct a
non-trivial D = 4 linear Teleparallel Gauss-Bonnet gravity theory. This can be done by following the construction
of [83, 84] which is to take the Teleparallel GB invariants and use a dimensional reduction method.

5 Doing one integration by parts, it is easy to see that the resulting operator falls into the classification of Ref. [81] as NQ = 1, n = r = 1
and m = l = 0, with arbitrary Nφ .
16

Recently, an interesting and related work appeared [85]. Although the authors study the Gauss-Bonnet invariant
in General TG theories of gravity, which is also this work’s primary goal, here we focus on and exploit the power of
differential form language to obtain a compact expression for our scalars. Moreover, our formalism relies on always
rewriting the invariants as boundary terms (or topological invariants6 ), while in Ref. [85], the authors explore a
different split of the Riemannian Gauss-Bonnet invariant into pieces that are not independently boundary terms. For
example, Eq. (14) in that
√ reference provides an expression for the torsional case in terms of three scalars. Considering
the Lagrangian L = −g(c1 GT + c2 GDT + c3 GDDT ), the only way of having boundary terms without dynamics is
taking the combination c3 = c2 = c1 , which reconstructs the Riemannian Gauss-Bonnet invariant. A similar situation
occurs for the nonmetricity Gauss-Bonnet invariant (given by Eq. (21) in [85]). Thus, none of the scalars are boundary
terms (or topological invariants) on their own, but only the unique combination reconstructing the Riemannian Gauss-
Bonnet term is. In that sense, our approach also differs from the one of those authors’ since our split always keeps
a structure such that the Teleparallel scalars are divided into a bulk and a boundary term; and in 4-dimensions, our
Teleparallel Gauss-Bonnet terms always correspond to boundary terms independently when they appear linearly in
an action.

Acknowledgements

The work of JMA has been supported by CONICET, ANPCyT and UBA. JMA would like to thank the support and
funding of UBA to carry out the current work by means of the grant “Financiamiento de estadı́a en el exterior 2023”,
and to the Institute of Physics of the Czech Academy of Sciences (CEICO) for funding and hospitality during initial
stages of this work. S.B. was supported by JSPS Postdoctoral Fellowships for Research in Japan and KAKENHI
Grant-in-Aid for Scientific Research No. JP21F21789, and by “Agencia Nacional de Investigación y Desarrollo”
(ANID), Grant “Becas Chile postdoctorado al extranjero” No. 74220006. The work of G.T. was supported by the
Grant Agency of the Czech Republic, GACR grant 20-28525S. The work of L.G.T. was supported by European
Union (Grant No. 101063210). S.B., G.T., and L.G.T. would like to thank the support by the Bilateral Czech-
Japanese Mobility Plus Project JSPS-21-12 (JPJSBP120212502). S.B. was also supported by “Agencia Nacional de
Investigación y Desarrollo” (ANID), Grant “Becas Chile postdoctorado al extranjero” No. 74220006.

Appendix A: Useful properties of differential forms

This is an incomplete list of properties that are used in the main text and Appendix B to work on the expressions

in the differential form language. First, consider the Levi-Civita convariant exterior derivative D, then it follows

Dea = 0, (A1)

Dηab = 0. (A2)

We also have the Riemannian differential Bianchi identity


◦ ◦
DRa b = 0. (A3)

In an orthonormal frame we also have the property



Dǫa1 ...aD = dǫa1 ...aD = 0, (A4)

where d is the usual exterior derivative. Notice that with Eq. (A2) we can raise the index in Eq. (7)
◦ ◦
Rab = Rab + DN ab + N a c ∧ N cb . (A5)

Acting twice with D gives the usual relation with the Riemannian curvature 2-form
◦ ◦ ◦
D2 N a b = Ra c ∧ N c b − Rc b ∧ N a c . (A6)

6 Note that we are not showing that our scalars are topological invariants separately, but they are always boundary terms in 4-dimensions.
17

Consider now the exterior covariant derivative for the teleparallel connection D,

Dηab = −2ω(ab) = −2N(ab) , (A7)


ab (ab)
Dη = 2N . (A8)

Raising an index under D is now nontrivial when the distortion 1-form has a symmetric part, i.e. nonmetricity,

DN ab = η bc DN a c − 2N a c ∧ N (cb) . (A9)

We can trade D for the Levi-Civita one D plus an extra term

DN a b = dN a b + ω a c ∧ N c b − ω c b ∧ N a c = DN a b + 2N a c ∧ N c b , (A10)

or, equivalently

DN ab = DN ab + 2N a c ∧ N [cb] . (A11)

This allows us to reexpress Eq. (7) in terms of D,



Ra b = Ra b + DN a b − N a c ∧ N c b . (A12)

Appendix B: Equivalence of alternative decompositions

The purpose of this appendix is to prove the equivalence of Eqs. (45) and (49). We work our way starting from
Eq. (49) and integrating by parts the first term of the second line, effectively altering the split between bulk and
boundary,
◦ ◦
ǫa1 ...aD N a1 f ∧ N f a2 ∧ DN a3 a4 ∧ ea5 ∧ · · · ∧ eaD = ǫa1 ...aD DN a1 a2 ∧ N a3 f ∧ N f a4 ∧ ea5 ∧ · · · ∧ eaD
= ǫa1 ...aD d N a1 a2 ∧ N a3 f ∧ N f a4 ∧ ea5 ∧ · · · ∧ eaD


+ǫa1 ...aD N a1 a2 ∧ D(N a3 f ∧ N f a4 ) ∧ ea5 ∧ · · · ∧ eaD , (B1)

which, when plugged back in leads to


"
◦ 1  ◦ ◦

G ∗1 = ǫa1 ...aD −d 2N a1 a2 ∧ Ra3 a4 ∧ ea5 ∧ · · · ∧ eaD + N a1 a2 ∧ DN a3 a4 ∧ ea5 ∧ · · · ∧ eaD
(D − 4)!

+N a1 a2 ∧ D(N a3 f ∧ N f a4 ) ∧ ea5 ∧ · · · ∧ eaD
#
a1 f a2 a3 ha4 a5 aD
+N f ∧N ∧N h ∧N ∧e ∧ ···∧ e , (B2)

◦ ◦
where we also used N a3 f ∧ N f a4 = −Ra3 a4 − DN a3 a4 . The boundary term here already looks the same as in Eq. (45),
but we still need to massage the bulk part, in particular the second line above,

ǫa1 ...aD N a1 a2 ∧ D(N a3 f ∧ N f a4 ) ∧ ea5 ∧ · · · ∧ eaD
◦ ◦

= ǫa1 ...aD ηcd N a1 a2 ∧ DN a3 c ∧ N da4 − N a3 c ∧ DN da4 ∧ ea5 ∧ · · · ∧ eaD
◦ ◦

= ǫa1 ...aD ηcd N a1 a2 ∧ DN a3 c ∧ N da4 + DN ca3 ∧ N a4 d ∧ ea5 ∧ · · · ∧ eaD
◦ ◦

= ǫa1 ...aD ηcd N a1 a2 ∧ 2 DN (a3 c) ∧ N (da4 ) + DN [a3 c] ∧ N [da4 ] ∧ ea5 ∧ · · · ∧ eaD , (B3)

where in the last equality we used N ab = N (ab) + N [ab] . From property (A11), we have

◦ ✘✿ 0  a3
✘ 
DN (a3 c) = ✘✘(a
DN ✘3 c) − N f ∧ N [f c] + N c f ∧ N [f a3 ] , (B4)

 
DN [a3 c] = DN [a3 c] − N a3 f ∧ N [f c] − N c f ∧ N [f a3 ] , (B5)
18

where we used the third Bianchi identity of Eq. (6) in the teleparallel case, which reads DN (ab) = 0 (recalling
N (ab) = Qab ). Then, we have
◦ ◦
DN (a3 c) ∧ N (da4 ) + DN [a3 c] ∧ N [da4 ] = −ηf h N a3 f ∧ N [hc] ∧ N (da4 ) − ηf h N cf ∧ N [ha3 ] ∧ N (da4 ) (B6)
[a3 c] [da4 ] a3 f [hc] [da4 ] cf [ha3 ] [da4 ]
+DN ∧N − ηf h N ∧N ∧N + ηf h N ∧N ∧N ,

and therefore, replacing back and reordering we get



ǫa1 ...aD N a1 a2 ∧ D(N a3 f ∧ N f a4 ) ∧ ea5 ∧ · · · ∧ eaD = 2ǫa1 ...aD ηcd N a1 a2 ∧ DN [a3 c] ∧ N [da4 ] ∧ ea5 ∧ · · · ∧ eaD (B7)
a1 a2 a3 f [hc] (da4 ) a5
− 2ǫa1 ...aD ηcd ηf h N ∧N ∧N ∧N ∧e ∧ · · · ∧ e aD
− 2ǫa1 ...aD ηcd ηf h N a1 a2 ∧ N a3 f ∧ N [hc] ∧ N [da4 ] ∧ ea5 ∧ · · · ∧ eaD
− 2ǫa1 ...aD ηcd ηf h N a1 a2 ∧ N cf ∧ N [ha3 ] ∧ N (da4 ) ∧ ea5 ∧ · · · ∧ eaD
+ 2ǫa1 ...aD ηcd ηf h N a1 a2 ∧ N cf ∧ N [ha3 ] ∧ N [da4 ] ∧ ea5 ∧ · · · ∧ eaD .

Now here notice that the last 4 terms (those without D) contain three factors of N contracted to each other directly,
of the form

ǫ...a3 a4 ... ηcd ηf h N a3 c ∧ N df ∧ N ha4 . (B8)

Due to the presence of the totally antisymmetric tensor ǫ...a3 a4 ... , only the antisymmetric part w.r.t. a3 and a4 of the
above triple N survives. By splitting each N into symmetric and antisymmetric parts we can see which combinations
survive. In particular, we find

ǫa1 ...aD ηcd ηf h N (a3 f ) ∧ N [hc] ∧ N (da4 ) = 0, (B9)


[a3 f ] [hc] [da4 ]
ǫa1 ...aD ηcd ηf h N ∧N ∧N = 0, (B10)
[cf ] [ha3 ] [da4 ]
ǫa1 ...aD ηcd ηf h N ∧N ∧N = 0, (B11)

which leads to

ǫa1 ...aD N a1 a2 ∧ D(N a3 f ∧ N f a4 ) ∧ ea5 ∧ · · · ∧ eaD = 2ǫa1 ...aD ηcd N a1 a2 ∧ DN [a3 c] ∧ N [da4 ] ∧ ea5 ∧ · · · ∧ eaD (B12)
a1 a2 [a3 f ] [hc] (da4 ) a5
− 2ǫa1 ...aD ηcd ηf h N ∧N ∧N ∧N ∧e ∧ · · · ∧ e aD
− 2ǫa1 ...aD ηcd ηf h N a1 a2 ∧ N (a3 f ) ∧ N [hc] ∧ N [da4 ] ∧ ea5 ∧ · · · ∧ eaD
− 2ǫa1 ...aD ηcd ηf h N a1 a2 ∧ N (cf ) ∧ N [ha3 ] ∧ N (da4 ) ∧ ea5 ∧ · · · ∧ eaD
− 2ǫa1 ...aD ηcd ηf h N a1 a2 ∧ N [cf ] ∧ N [ha3 ] ∧ N (da4 ) ∧ ea5 ∧ · · · ∧ eaD
+ 2ǫa1 ...aD ηcd ηf h N a1 a2 ∧ N (cf ) ∧ N [ha3 ] ∧ N [da4 ] ∧ ea5 ∧ · · · ∧ eaD .

Here it is immediate to see that the second and fifth lines cancel each other out. Then, after some reordering and
relabeling we get

ǫa1 ...aD N a1 a2 ∧ D(N a3 f ∧ N f a4 ) ∧ ea5 ∧ · · · ∧ eaD = 2ǫa1 ...aD ηf h N a1 a2 ∧ DN [a3 f ] ∧ N [ha4 ] ∧ ea5 ∧ · · · ∧ eaD (B13)
a1 a2 (a3 f ) [hc] [da4 ] a5
− 2ǫa1 ...aD ηcd ηf h N ∧N ∧N ∧N ∧e ∧ · · · ∧ e aD
+ 2ǫa1 ...aD ηcd ηf h N a1 a2 ∧ N [a3 f ] ∧ N (hc) ∧ N [da4 ] ∧ ea5 ∧ · · · ∧ eaD
+ 2ǫa1 ...aD ηcd ηf h N a1 a2 ∧ N (a3 f ) ∧ N (hc) ∧ N [da4 ] ∧ ea5 ∧ · · · ∧ eaD .

Using Eq. (A9) it simplifies to,



ǫa1 ...aD N a1 a2 ∧ D(N a3 f ∧ N f a4 ) ∧ ea5 ∧ · · · ∧ eaD = 2ǫa1 ...aD N a1 a2 ∧ DN a3 c ∧ N [ca4 ] ∧ ea5 ∧ · · · ∧ eaD (B14)
a1 a2 a3 f (hc) [da4 ] a5
− 2ǫa1 ...aD ηcd ηf h N ∧N ∧N ∧N ∧e ∧ · · · ∧ e aD
− 2ǫa1 ...aD ηcd ηf h N a1 a2 ∧ N (a3 f ) ∧ N [hc] ∧ N [da4 ] ∧ ea5 ∧ · · · ∧ eaD ,

while using Eq. (B10) allows us to combine the last two terms
 
ǫa1 ...aD ηcd ηf h N a3 f ∧ N (hc) ∧ N [da4 ] + N (a3 f ) ∧ N [hc] ∧ N [da4 ]
19
 
= ǫa1 ...aD ηcd ηf h N a3 f ∧ N (hc) ∧ N [da4 ] + N a3 f ∧ N [hc] ∧ N [da4 ]
 
= ǫa1 ...aD ηcd ηf h N a3 f ∧ N hc ∧ N [da4 ] . (B15)

Finally, we arrived at the following identity



ǫa1 ...aD N a1 a2 ∧ D(N a3 f ∧ N f a4 ) ∧ ea5 ∧ · · · ∧ eaD = 2ǫa1 ...aD N a1 a2 ∧ DN a3 c ∧ N [ca4 ] ∧ ea5 ∧ · · · ∧ eaD (B16)
a1 a2 a3 d [ca4 ] a5 aD
− 2ǫa1 ...aD N ∧N d ∧N c ∧N ∧e ∧ ···∧ e ,

which plugged back into Eq. (B2) leads to precisely to Eq. (45).

Appendix C: Second order nature of scalar-ST-Gauss-Bonnet

Here we want to prove that the field equations associated to the theory Eq. (94) are second order in derivatives of all

fields (ψ, gµν , Γα µν ), for the two splits of G proposed in Sec. III and evaluated in the Symmetric TG case in Eqs. (72)
and (75). In the following it is useful to recall that, from its definition in Eqs. (20) and (17), the disformation tensor
contains up to first derivatives of the metric and no derivatives of the Teleparallel connection, schematically

L ∼ ∂g − Γ. (C1)

(Q)
As explained in the main text, we can always recast the last two terms in Eq. (94) as α1 G1 (ψ)G + αG(ψ) BG ,
◦ ◦
(Q) (Q)
where we have absorbed TG into G and redefined the coupling of BG . Since α1 G1 (ψ)G belongs to Riemannian
Horndeski [14], it is already guaranteed to lead to second-order field equations, and therefore it suffices to prove that
(Q)
it is also true for αG(ψ) BG for each split. Moreover, the difference between the two boundary terms associated with
each split trivially leads to second-order field equations

4 √ √
Z h i Z
2 (Q) 1 (Q)
d x −gG(ψ) BG − BG = d4x −gG ′ (ψ)∂µ ψ δµµ1νµρ2 µσ3 µ4 Lµ1 µ2 ν Lµ3 λρ Lλµ4 σ , (C2)

where we integrated by parts once, as it contains at most first derivatives of any given field. Therefore, it only
remains to be shown that any of the two possible boundary terms leads to second-order field equations. We focus on
(Q)
G(ψ) 2BG .

1. Scalar field equation

(Q)
First notice that the bulk terms i TG of both splits, Eqs. (73) and (76) respectively, contain at most second

derivatives of the fields. Then, given that the Riemannian Gauss-Bonnet invariant G itself also has this property, it
(Q)
follows that the corresponding boundary terms iBG , given by Eqs. (74) and (77), also have at most second derivatives.
Therefore, variations w.r.t. the scalar filed ψ are automatically second order in this theory for both splits, and in
(Q)
particular, for G(ψ) 2BG .

2. Teleparallel connection equation

(Q)
We continue by integrating G(ψ) 2BG by parts

4 √ 1 √
Z Z

2 (Q)
d x −gG(ψ) BG = d4x −gG ′ (ψ)∂µ ψ δµµ1νµρ2 µσ3 µ4 Lµ1 µ2 ν Rµ3 µ4 ρσ , (C3)
2
which is a second order expression. Given that the Teleparallel connection Γα µν appears here only linearly and without
any derivatives acting on it, as per Eq. (C1), it follows then that variations w.r.t. to it are automatically second order
too.
20

3. Metric field equations

Finally, let us consider variations w.r.t. the metric, which are the less trivial ones. Indeed, by counting derivatives
in Eq. (C3), we get schematically ∂ψ ∂g ∂ 2 g, which could potentially lead to higher-than-second order field equations.
Let us work on the variation explicitly in order to show that those higher-order terms actually cancel,

4 √ 1 √
Z Z h i

2 (Q)
δg d x −gG(ψ) BG ⊃ d4x −g ∂µ ψ δµµ1νµρ2 µσ3 µ4 δg Lµ1 µ2 ν Rµ3 µ4 ρσ ≡ Eg , (C4)
2
where we are not tracking terms which lead automatically to second-order field equations. It is useful to recall that
under the Teleparallel condition, Eq. (24), from Eq. (25) it is possible to write
◦ ◦ ◦ ◦
Rσ ρµν = ∇ν Lσ µρ − ∇µ Lσ νρ + · · · = 2∇[ν Lσ µ]ρ + . . . , (C5)

where we also dropped terms with less than two derivatives acting on the metric, which are safe. The above implies
◦ ◦
δg Rσ ρµν = 2∇[ν δLσ µ]ρ , (C6)

and therefore

Z h i
◦ ◦
Eg = d4x −g ∂µ ψ δµµ1νµρ2 µσ3 µ4 δg Lµ1 µ2 ν ∇σ Lµ3 µ4 ρ + Lµ1 µ2 ν ∇σ δg Lµ3 µ4 ρ , (C7)

where we used the antisymmetry of ρ and σ to drop the antisymmetrization brackets, and we also used the symmetry
of L in the last two indices. Let us work on the second term to show that contributes the same as the first. We first
integrate by parts the second term

Z h i
◦ ◦
Eg = d4x −g ∂µ ψ δµµ1νµρ2 µσ3 µ4 δg Lµ1 µ2 ν ∇σ Lµ3 µ4 ρ − δg Lµ3 µ4 ρ ∇σ Lµ1 µ2 ν

Z
◦ ◦
− d4x −g ∇σ ∇µ ψ δµµ1νµρ2 µσ3 µ4 Lµ1 µ2 ν δg Lµ3 µ4 ρ . (C8)

Notice that the second term with two derivatives acting on ψ is zero due to antisymmetry of µ and σ. We then do an
exchange and relabel of ρ and ν (one minus sign) and a cyclic permutation of µ1 µ2 µ3 µ4 (no sign change),

Z h i
◦ ◦
Eg = d4x −g δµµ1νµρ2 µσ3 µ4 δg Lµ1 µ2 ν ∇σ Lµ3 µ4 ρ + δg Lµ1 µ2 ν ∇σ Lµ3 µ4 ρ

Z

= 2 d4x −g δµµ1νµρ2 µσ3 µ4 δg Lµ1 µ2 ν ∇σ Lµ3 µ4 ρ

Z

= 2 d4x −g δµµ1νµρ2 µσ3 µ4 δg Lµ1 µ2 ν Rµ3 µ4 ρσ , (C9)

where in the last step we again have the Levi-Civita Riemann tensor, up to terms that do not have more that second
deriviatives of the metric.
We finally now express δg Lµ1 µ2 ν in terms of δgµν ,

δg L[µ1 µ2 ] ν = 2g α[µ1 g µ2 ]β ∇α δgβν , (C10)

which allows us to pull out another derivative and actually see the dangerous term:

4 √ √
Z Z
◦ ◦ ◦ ◦
µνρσ
Eg = 4 d x −g δµ1 µ2 µ3 µ4 R µ3 µ4
ρσ g g ∇α δgβν = −4 d4x −g δµµ1νµρ2 µσ3 µ4 g µ2 β δgβν ∇µ1 Rµ3 µ4 ρσ . (C11)
αµ1 µ2 β

The last factor contains the potentially dangerous third derivatives of g. However, because of total antisymmetrization
of µ1 µ3 µ4 , it is proportional to
◦ ◦
∇[µ1 Rµ3 µ4 ] ρσ = 0, (C12)

which vanishes due to the second Bianchi identity, proving the absence of higher-than-second order terms in the field
equations.
21

[1] D. Lovelock, “The four-dimensionality of space and the einstein tensor,” J. Math. Phys. 13 (1972) 874–876.
[2] D. Lovelock, “The Einstein tensor and its generalizations,” J. Math. Phys. 12 (1971) 498–501.
[3] S. Nojiri, S. D. Odintsov, and M. Sasaki, “Gauss-Bonnet dark energy,” Phys. Rev. D 71 (2005) 123509,
arXiv:hep-th/0504052.
[4] S. Nojiri and S. D. Odintsov, “Modified Gauss-Bonnet theory as gravitational alternative for dark energy,”
Phys. Lett. B 631 (2005) 1–6, arXiv:hep-th/0508049.
[5] S. Nojiri, S. D. Odintsov, and V. K. Oikonomou, “Modified Gravity Theories on a Nutshell: Inflation, Bounce and
Late-time Evolution,” Phys. Rept. 692 (2017) 1–104, arXiv:1705.11098 [gr-qc].
[6] G. Cognola, E. Elizalde, S. Nojiri, S. D. Odintsov, and S. Zerbini, “Dark energy in modified Gauss-Bonnet gravity:
Late-time acceleration and the hierarchy problem,” Phys. Rev. D 73 (2006) 084007, arXiv:hep-th/0601008.
[7] B. Li, J. D. Barrow, and D. F. Mota, “The Cosmology of Modified Gauss-Bonnet Gravity,”
Phys. Rev. D 76 (2007) 044027, arXiv:0705.3795 [gr-qc].
[8] S. Nojiri, S. D. Odintsov, and V. K. Oikonomou, “Ghost-free Gauss-Bonnet Theories of Gravity,”
Phys. Rev. D 99 (2019) no. 4, 044050, arXiv:1811.07790 [gr-qc].
[9] E. S. Fradkin and A. A. Tseytlin, “Effective Field Theory from Quantized Strings,” Phys. Lett. B 158 (1985) 316–322.
[10] D. J. Gross and J. H. Sloan, “The Quartic Effective Action for the Heterotic String,” Nucl. Phys. B 291 (1987) 41–89.
[11] C. G. Callan, Jr., E. J. Martinec, M. J. Perry, and D. Friedan, “Strings in Background Fields,”
Nucl. Phys. B 262 (1985) 593–609.
[12] Z. Komargodski and A. Schwimmer, “On Renormalization Group Flows in Four Dimensions,” JHEP 12 (2011) 099,
arXiv:1107.3987 [hep-th].
[13] M. A. Luty, J. Polchinski, and R. Rattazzi, “The a-theorem and the Asymptotics of 4D Quantum Field Theory,”
JHEP 01 (2013) 152, arXiv:1204.5221 [hep-th].
[14] T. Kobayashi, M. Yamaguchi, and J. Yokoyama, “Generalized G-inflation: Inflation with the most general second-order
field equations,” Prog. Theor. Phys. 126 (2011) 511–529, arXiv:1105.5723 [hep-th].
[15] D. Pirtskhalava, L. Santoni, E. Trincherini, and F. Vernizzi, “Weakly Broken Galileon Symmetry,” JCAP 09 (2015) 007,
arXiv:1505.00007 [hep-th].
[16] J. M. Armaleo, J. Osorio Morales, and O. Santillan, “Gauss–Bonnet models with cosmological constant and non zero
spatial curvature in D = 4,” Eur. Phys. J. C 78 (2018) no. 2, 85, arXiv:1711.09484 [gr-qc].
[17] B. M. N. Carter and I. P. Neupane, “Towards inflation and dark energy cosmologies from modified Gauss-Bonnet
theory,” JCAP 06 (2006) 004, arXiv:hep-th/0512262.
[18] E. O. Pozdeeva, M. R. Gangopadhyay, M. Sami, A. V. Toporensky, and S. Y. Vernov, “Inflation with a quartic potential
in the framework of Einstein-Gauss-Bonnet gravity,” Phys. Rev. D 102 (2020) no. 4, 043525,
arXiv:2006.08027 [gr-qc].
[19] S. D. Odintsov, V. K. Oikonomou, I. Giannakoudi, F. P. Fronimos, and E. C. Lymperiadou, “Recent Advances on
Inflation,” Symmetry 15 (2023) 9, arXiv:2307.16308 [gr-qc].
[20] T. P. Sotiriou and S.-Y. Zhou, “Black hole hair in generalized scalar-tensor gravity,” Phys. Rev. Lett. 112 (2014) 251102,
arXiv:1312.3622 [gr-qc].
[21] T. P. Sotiriou and S.-Y. Zhou, “Black hole hair in generalized scalar-tensor gravity: An explicit example,”
Phys. Rev. D 90 (2014) 124063, arXiv:1408.1698 [gr-qc].
[22] H. O. Silva, J. Sakstein, L. Gualtieri, T. P. Sotiriou, and E. Berti, “Spontaneous scalarization of black holes and compact
stars from a Gauss-Bonnet coupling,” Phys. Rev. Lett. 120 (2018) no. 13, 131104, arXiv:1711.02080 [gr-qc].
[23] C. F. B. Macedo, J. Sakstein, E. Berti, L. Gualtieri, H. O. Silva, and T. P. Sotiriou, “Self-interactions and Spontaneous
Black Hole Scalarization,” Phys. Rev. D 99 (2019) no. 10, 104041, arXiv:1903.06784 [gr-qc].
[24] A. Dima, E. Barausse, N. Franchini, and T. P. Sotiriou, “Spin-induced black hole spontaneous scalarization,”
Phys. Rev. Lett. 125 (2020) no. 23, 231101, arXiv:2006.03095 [gr-qc].
[25] C. A. R. Herdeiro, E. Radu, H. O. Silva, T. P. Sotiriou, and N. Yunes, “Spin-induced scalarized black holes,”
Phys. Rev. Lett. 126 (2021) no. 1, 011103, arXiv:2009.03904 [gr-qc].
[26] T. Anson, E. Babichev, C. Charmousis, and S. Ramazanov, “Cosmological instability of scalar-Gauss-Bonnet theories
exhibiting scalarization,” JCAP 06 (2019) 023, arXiv:1903.02399 [gr-qc].
[27] L. Hui and A. Nicolis, “No-Hair Theorem for the Galileon,” Phys. Rev. Lett. 110 (2013) 241104,
arXiv:1202.1296 [hep-th].
[28] P. Creminelli, N. Loayza, F. Serra, E. Trincherini, and L. G. Trombetta, “Hairy Black-holes in Shift-symmetric
Theories,” JHEP 08 (2020) 045, arXiv:2004.02893 [hep-th].
[29] K. Yagi, “A New constraint on scalar Gauss-Bonnet gravity and a possible explanation for the excess of the orbital decay
rate in a low-mass X-ray binary,” Phys. Rev. D 86 (2012) 081504, arXiv:1204.4524 [gr-qc].
[30] N. Yunes, K. Yagi, and F. Pretorius, “Theoretical Physics Implications of the Binary Black-Hole Mergers GW150914 and
GW151226,” Phys. Rev. D 94 (2016) no. 8, 084002, arXiv:1603.08955 [gr-qc].
[31] H. Witek, L. Gualtieri, P. Pani, and T. P. Sotiriou, “Black holes and binary mergers in scalar Gauss-Bonnet gravity:
scalar field dynamics,” Phys. Rev. D 99 (2019) no. 6, 064035, arXiv:1810.05177 [gr-qc].
[32] Z. Lyu, N. Jiang, and K. Yagi, “Constraints on Einstein-dilation-Gauss-Bonnet gravity from black hole-neutron star
gravitational wave events,” Phys. Rev. D 105 (2022) no. 6, 064001, arXiv:2201.02543 [gr-qc]. [Erratum: Phys.Rev.D
22

106, 069901 (2022), Erratum: Phys.Rev.D 106, 069901 (2022)].


[33] J. Noller, L. Santoni, E. Trincherini, and L. G. Trombetta, “Black Hole Ringdown as a Probe for Dark Energy,”
Phys. Rev. D 101 (2020) 084049, arXiv:1911.11671 [gr-qc].
[34] F. Serra, J. Serra, E. Trincherini, and L. G. Trombetta, “Causality constraints on black holes beyond GR,”
JHEP 08 (2022) 157, arXiv:2205.08551 [hep-th].
[35] J. M. Nester and H.-J. Yo, “Symmetric teleparallel general relativity,” Chin. J. Phys. 37 (1999) 113,
arXiv:gr-qc/9809049.
[36] J. Beltrán Jiménez, L. Heisenberg, and T. S. Koivisto, “The Geometrical Trinity of Gravity,”
Universe 5 (2019) no. 7, 173, arXiv:1903.06830 [hep-th].
[37] L. Heisenberg, “A systematic approach to generalisations of General Relativity and their cosmological implications,”
Phys. Rept. 796 (2019) 1–113, arXiv:1807.01725 [gr-qc].
[38] Y. Obukhov and J. Pereira, “Metric affine approach to teleparallel gravity,” Phys. Rev. D 67 (2003) 044016,
arXiv:gr-qc/0212080.
[39] J. Beltrán Jiménez, L. Heisenberg, and T. S. Koivisto, “Teleparallel Palatini theories,” JCAP 08 (2018) 039,
arXiv:1803.10185 [gr-qc].
[40] J. Beltrán Jiménez, L. Heisenberg, and T. Koivisto, “Coincident General Relativity,”
Phys. Rev. D 98 (2018) no. 4, 044048, arXiv:1710.03116 [gr-qc].
[41] R. Aldrovandi and J. G. Pereira, Teleparallel Gravity, vol. 173. Springer, Dordrecht, 2013.
[42] J. W. Maluf, “The teleparallel equivalent of general relativity,” Annalen Phys. 525 (2013) 339–357,
arXiv:1303.3897 [gr-qc].
[43] M. Krššák and E. N. Saridakis, “The covariant formulation of f(T) gravity,”
Class. Quant. Grav. 33 (2016) no. 11, 115009, arXiv:1510.08432 [gr-qc].
[44] Y.-F. Cai, S. Capozziello, M. De Laurentis, and E. N. Saridakis, “f(T) teleparallel gravity and cosmology,”
Rept. Prog. Phys. 79 (2016) no. 10, 106901, arXiv:1511.07586 [gr-qc].
[45] M. Krssak, R. J. van den Hoogen, J. G. Pereira, C. G. Böhmer, and A. A. Coley, “Teleparallel theories of gravity:
illuminating a fully invariant approach,” Class. Quant. Grav. 36 (2019) no. 18, 183001, arXiv:1810.12932 [gr-qc].
[46] S. Bahamonde, K. F. Dialektopoulos, C. Escamilla-Rivera, G. Farrugia, V. Gakis, M. Hendry, M. Hohmann, J. L. Said,
J. Mifsud, and E. Di Valentino, “Teleparallel Gravity: From Theory to Cosmology,” arXiv:2106.13793 [gr-qc].
[47] G. Kofinas and E. N. Saridakis, “Teleparallel equivalent of Gauss-Bonnet gravity and its modifications,”
Phys. Rev. D90 (2014) 084044, arXiv:1404.2249 [gr-qc].
[48] S. Bahamonde and C. G. Böhmer, “Modified teleparallel theories of gravity: Gauss–Bonnet and trace extensions,”
Eur. Phys. J. C76 (2016) no. 10, 578, arXiv:1606.05557 [gr-qc].
[49] P. A. Gonzalez and Y. Vasquez, “Teleparallel Equivalent of Lovelock Gravity,” Phys. Rev. D 92 (2015) no. 12, 124023,
arXiv:1508.01174 [hep-th].
[50] G. Kofinas, G. Leon, and E. N. Saridakis, “Dynamical behavior in f (T, TG ) cosmology,”
Class. Quant. Grav. 31 (2014) 175011, arXiv:1404.7100 [gr-qc].
[51] G. Kofinas and E. N. Saridakis, “Cosmological applications of F (T, TG ) gravity,” Phys. Rev. D90 (2014) 084045,
arXiv:1408.0107 [gr-qc].
[52] S. Capozziello, M. De Laurentis, and K. F. Dialektopoulos, “Noether symmetries in Gauss–Bonnet-teleparallel
cosmology,” Eur. Phys. J. C76 (2016) no. 11, 629, arXiv:1609.09289 [gr-qc].
[53] S. Bahamonde, U. Camci, and S. Capozziello, “Noether symmetries and boundary terms in extended Teleparallel gravity
cosmology,” Class. Quant. Grav. 36 (2019) no. 6, 065013, arXiv:1807.02891 [gr-qc].
[54] A. de la Cruz-Dombriz, G. Farrugia, J. L. Said, and D. Saez-Gomez, “Cosmological reconstructed solutions in extended
teleparallel gravity theories with a teleparallel Gauss–Bonnet term,” Class. Quant. Grav. 34 (2017) no. 23, 235011,
arXiv:1705.03867 [gr-qc].
[55] S. Bahamonde, M. Marciu, S. D. Odintsov, and P. Rudra, “String-inspired Teleparallel cosmology,”
Nucl. Phys. B 962 (2021) 115238, arXiv:2003.13434 [gr-qc].
[56] S. Bahamonde, D. D. Doneva, L. Ducobu, C. Pfeifer, and S. S. Yazadjiev, “Spontaneous scalarization of black holes in
Gauss-Bonnet teleparallel gravity,” Phys. Rev. D 107 (2023) no. 10, 104013, arXiv:2212.07653 [gr-qc].
[57] S. Bahamonde, D. D. Doneva, L. Ducobu, C. Pfeifer, and S. S. Yazadjiev, “Distinctive Features of Hairy Black Holes in
Teleparallel Gauss-Bonnet Gravity,” arXiv:2307.14720 [gr-qc].
[58] M. Adak, T. Dereli, T. S. Koivisto, and C. Pala, “General teleparallel metrical geometries,”
Int. J. Geom. Meth. Mod. Phys. 20 (2023) 2350215, arXiv:2303.17812 [gr-qc].
[59] M. Gasperini, Theory of Gravitational Interactions. UNITEXT for Physics. Springer International Publishing, Cham,
2017.
[60] F. W. Hehl, P. Von Der Heyde, G. D. Kerlick, and J. M. Nester, “General Relativity with Spin and Torsion: Foundations
and Prospects,” Rev. Mod. Phys. 48 (1976) 393–416.
[61] J. Beltrán Jiménez, L. Heisenberg, D. Iosifidis, A. Jiménez-Cano, and T. S. Koivisto, “General teleparallel quadratic
gravity,” Phys. Lett. B 805 (2020) 135422, arXiv:1909.09045 [gr-qc].
[62] D. Iosifidis, Metric-Affine Gravity and Cosmology/Aspects of Torsion and non-Metricity in Gravity Theories. PhD thesis,
2019. arXiv:1902.09643 [gr-qc].
[63] A. Yale and T. Padmanabhan, “Structure of Lanczos-Lovelock Lagrangians in Critical Dimensions,”
Gen. Rel. Grav. 43 (2011) 1549–1570, arXiv:1008.5154 [gr-qc].
[64] M. Hohmann, “General covariant symmetric teleparallel cosmology,” Phys. Rev. D 104 (2021) no. 12, 124077,
23

arXiv:2109.01525 [gr-qc].
[65] F. D’Ambrosio, S. D. B. Fell, L. Heisenberg, and S. Kuhn, “Black holes in f(Q) gravity,”
Phys. Rev. D 105 (2022) no. 2, 024042, arXiv:2109.03174 [gr-qc].
[66] S. Bahamonde, J. Gigante Valcarcel, L. Järv, and J. Lember, “Black hole solutions in scalar-tensor symmetric teleparallel
gravity,” JCAP 08 (2022) 082, arXiv:2206.02725 [gr-qc].
[67] R. Ferraro and F. Fiorini, “Modified teleparallel gravity: Inflation without inflaton,” Phys. Rev. D75 (2007) 084031,
arXiv:gr-qc/0610067 [gr-qc].
[68] R. Ferraro and F. Fiorini, “On Born-Infeld Gravity in Weitzenbock spacetime,” Phys. Rev. D78 (2008) 124019,
arXiv:0812.1981 [gr-qc].
[69] F. D’Ambrosio, L. Heisenberg, and S. Kuhn, “Revisiting cosmologies in teleparallelism,”
Class. Quant. Grav. 39 (2022) no. 2, 025013, arXiv:2109.04209 [gr-qc].
[70] F. K. Anagnostopoulos, S. Basilakos, and E. N. Saridakis, “First evidence that non-metricity f(Q) gravity could challenge
ΛCDM,” Phys. Lett. B 822 (2021) 136634, arXiv:2104.15123 [gr-qc].
[71] N. Frusciante, “Signatures of f (Q)-gravity in cosmology,” Phys. Rev. D 103 (2021) no. 4, 044021,
arXiv:2101.09242 [astro-ph.CO].
[72] J. Beltrán Jiménez, L. Heisenberg, T. S. Koivisto, and S. Pekar, “Cosmology in f (Q) geometry,”
Phys. Rev. D 101 (2020) no. 10, 103507, arXiv:1906.10027 [gr-qc].
[73] V. Gakis, M. Krššák, J. Levi Said, and E. N. Saridakis, “Conformal gravity and transformations in the symmetric
teleparallel framework,” Phys. Rev. D 101 (2020) no. 6, 064024, arXiv:1908.05741 [gr-qc].
[74] S. Capozziello, V. De Falco, and C. Ferrara, “The role of the boundary term in f (Q, B) symmetric teleparallel gravity,”
arXiv:2307.13280 [gr-qc].
[75] T.-H. Loo, A. De, and E. N. Saridakis, “Non-metricity with bounday terms: f (Q, C) gravity and cosmology,”
arXiv:2308.00652 [gr-qc].
[76] S. Bahamonde, V. Gakis, S. Kiorpelidi, T. Koivisto, J. Levi Said, and E. N. Saridakis, “Cosmological perturbations in
modified teleparallel gravity models: Boundary term extension,” Eur. Phys. J. C 81 (2021) no. 1, 53,
arXiv:2009.02168 [gr-qc].
[77] P. Kanti, N. E. Mavromatos, J. Rizos, K. Tamvakis, and E. Winstanley, “Dilatonic black holes in higher curvature string
gravity,” Phys. Rev. D 54 (1996) 5049–5058, arXiv:hep-th/9511071.
[78] T. Torii, H. Yajima, and K.-i. Maeda, “Dilatonic black holes with Gauss-Bonnet term,” Phys. Rev. D 55 (1997) 739–753,
arXiv:gr-qc/9606034.
[79] P. Pani and V. Cardoso, “Are black holes in alternative theories serious astrophysical candidates? The Case for
Einstein-Dilaton-Gauss-Bonnet black holes,” Phys. Rev. D 79 (2009) 084031, arXiv:0902.1569 [gr-qc].
[80] D. D. Doneva and S. S. Yazadjiev, “New Gauss-Bonnet Black Holes with Curvature-Induced Scalarization in Extended
Scalar-Tensor Theories,” Phys. Rev. Lett. 120 (2018) no. 13, 131103, arXiv:1711.01187 [gr-qc].
[81] S. Bahamonde, G. Trenkler, L. G. Trombetta, and M. Yamaguchi, “Symmetric teleparallel Horndeski gravity,”
Phys. Rev. D 107 (2023) no. 10, 104024, arXiv:2212.08005 [gr-qc].
[82] D. Glavan and C. Lin, “Einstein-Gauss-Bonnet Gravity in Four-Dimensional Spacetime,”
Phys. Rev. Lett. 124 (2020) no. 8, 081301, arXiv:1905.03601 [gr-qc].
[83] H. Lu and Y. Pang, “Horndeski gravity as D → 4 limit of Gauss-Bonnet,” Phys. Lett. B 809 (2020) 135717,
arXiv:2003.11552 [gr-qc].
[84] R. A. Hennigar, D. Kubizňák, R. B. Mann, and C. Pollack, “On taking the D → 4 limit of Gauss-Bonnet gravity: theory
and solutions,” JHEP 07 (2020) 027, arXiv:2004.09472 [gr-qc].
[85] F. Bajardi, D. Blixt, and S. Capozziello, “The Gauss-Bonnet topological scalar in the Geometric Trinity of Gravity,”
arXiv:2308.03632 [gr-qc].

You might also like