∎
e1e-mail: [email protected] \thankstexte2e-mail: [email protected] \thankstexte3e-mail: [email protected]
Effective fluid mixture of tensor-multi-scalar gravity
Abstract
We apply to tensor-multi-scalar gravity the effective fluid analysis based on the representation of the gravitational scalar field as a dissipative effective fluid. This generalization poses new challenges as the effective fluid is now a complicated mixture of individual fluids mutually coupled to each other and many reference frames are possible for its description. They are all legitimate, although not all convenient for specific problems, and they give rise to different physical interpretations. Two of these frames are highlighted.
Keywords:
alternative theories of gravity tensor-multi-scalar gravity1 Introduction
It is well known that the scalar-tensor gravity field equations can be written as effective Einstein equations with an effective dissipative fluid in their right-hand side, built out of the Brans-Dicke-like scalar field present in the theory and of its first and second covariant derivatives Madsen:1988ph ; Pimentel89 ; Faraoni:2018qdr ; Quiros:2019gai . The formalism has been generalized to “viable” Horndeski gravity Giusti:2021sku ; Miranda:2022wkz ; Giusti:2022tgq and applied to Friedmann-Lemaître-Robertson-Walker (FLRW) cosmology Giardino:2022sdv , to theories containing non-propagating scalar degrees of freedom Faraoni:2022doe ; Miranda:2022brj , and to specific scalar-tensor solutions Faraoni:2022jyd ; Faraoni:2022fxo . But what is the analogue of a multi-component fluid? Naturally, the simplest multi-fluid equivalent of a theory of gravity is tensor-multi-scalar gravity. Here we extend the effective fluid formalism to this class of theories. The task is much less obvious than it would appear at first sight because all the gravitational scalar fields couple to gravity, which makes them all couple to each other. In general, there can also be direct mutual couplings through their kinetic and potential terms in the action. In the presence of multiple real fluids decoupled from each other, one can describe this mixture in the frame of an observer with timelike four-velocity . This four-velocity can be that of the comoving frame of one of the fluids, or it can be associated with any other observer. In general, it is difficult to define an average fluid EMMacC . This means that the total stress-energy tensor of the effective fluid mixture, which is a tensor defined unambiguously, can be decomposed in many ways according to the four-velocity selected. Each of these descriptions is legitimate but the description of the total mixture and its physical interpretation will depend on the observer selected to decompose . In particular, the density, pressure, heat flux density, and anisotropic stresses of each fluid as “seen” from a particular observer will differ from those measured in the comoving frame of that fluid. To appreciate the difference between the descriptions of a fluid in different frames, consider a perfect fluid with four-velocity that, in its comoving frame, is described by the stress-energy tensor111We follow the notation and conventions of Ref. Waldbook : the metric signature is , , is Newton’s constant, and units are used in which the speed of light is unity.
(1) |
In the frame of a different observer with timelike four-velocity related to by
(2) | |||||
(3) | |||||
(4) |
this perfect fluid (now “tilted”) will appear dissipative, with the different stress-energy tensor decomposition Maartens:1998xg ; Clarkson:2003ts ; Clarkson:2010uz
(5) |
where Maartens:1998xg ; Clarkson:2003ts ; Clarkson:2010uz
(6) |
(7) |
is the energy density,
(8) |
is the pressure,
(9) |
is the energy flux density, and
(10) |
is the anisotropic stress tensor. It is clear that the (spatial) vector arises solely due to the relative motion between the two frames, i.e., to the (spatial) vector . In this context it is problematic to interpret this purely convective current as a heat flux according to Eckart’s generalization of Fourier’s law Eckart40
(11) |
where is the temperature and is the thermal conductivity. This law expresses the fact that heat conduction is caused not only by spatial temperature gradients but also by an “inertial” contribution due to the fluid acceleration Eckart40 .
The situation becomes more complicated when multiple fluids are coupled to each other and even more when they are effective fluids and they all couple explicitly with the curvature222We do not consider derivative couplings in this work. (more precisely, with the Ricci scalar ) and to each other, which is the situation in tensor-multi-scalar gravity. In this work we discuss two possibilities, but other frames may be more convenient for specific problems.
Rather surprisingly, in tensor-multi-scalar gravity formulated in the Jordan conformal frame, one can obtain a particular frame as a sort of fictitious “average” frame, which is generally not possible with real fluids EMMacC . It is obtained by identifying the coupling function of the scalars to (which depends on all the scalar fields in the theory) with a new field and amounts to a redefinition of the scalar fields. This procedure is routine in tensor-single-scalar gravity, in which the only Brans-Dicke-like field is redefined for convenience, without much consequence or interpretation. In tensor-multi-scalar gravity, instead, this redefinition takes a new meaning. It identifies a four-velocity and a sort of “average” frame because there is only one Ricci scalar and all the scalar fields in the theory couple to it. This ingredient is missing for real fluids, which do not couple to the curvature and have no “average” frame EMMacC .
In the following we analyze tensor-multi-scalar gravity in its Jordan (conformal) frame formulation. It is possible to discuss it from the point of view of the “average” observer, or from the comoving frame of each fluid, or from that of any other timelike observer . It is important to remember that these descriptions will be different and will provide different physical interpretations of the mechanical and thermal aspects of the fluid mixture, and that these are all legitimate (hence one should not strive to identify the “correct” one). The point is that some of these formulations (originating different decompositions of the total based on different ) will be more convenient, and some others will be less convenient, for specific physical problems. One should adopt the formulation that is most convenient for the particular problem at hand without prejudice. For example, analyses of the quark-gluon plasma created in heavy ion collisions universally employ the Landau (or energy) frame BRAHMS:2004adc ; PHOBOS:2004zne ; STAR:2005gfr ; PHENIX:2004vcz ; Monnai:2019jkc in which there is no heat flux333This frame is found to be non-unique in Ref. Romero-Munoz:2014foa . while in FLRW cosmology, where comoving coordinates are the standard, relativistic fluids are routinely described in their comoving (or Eckart) frame Waldbook ; EMMacC .
Here we are interested in the fluid-mechanical equivalent and in the thermal description of tensor-multi-scalar gravity, where the fluids in the mixture are effective fluids and they all couple explicitly with and with each other. This is a very specific situation and our choices, although convenient in this problem, are not meant to be recipes with universal convenience (although aspects of our discussion may apply to other situations as well). After this discussion, we present an alternative view of the first-order thermodynamics of tensor-multi-scalar gravity in the Einstein conformal frame, while the last section summarizes our conclusions.
2 Tensor-multi-scalar gravity in the Jordan conformal frame
Let us begin with a convenient Jordan frame formulation of tensor-multi-scalar gravity (without derivative couplings). We adopt most of the notations specific to tensor-multi-scalar gravity used in Ref. Hohmann:2016yfd . There are scalar fields of gravitational nature , with , all coupled nonminimally with the Ricci scalar and between themselves, as described by the action
(12) | |||||
where capital indices label the scalar fields in the multiplet , is the determinant of the spacetime metric , is the associated covariant derivative, and is a scalar field potential. The Einstein summation convention is used also on the multiplet indices . The coupling function depends on all the , i.e., , or else some of the scalar fields would not be coupled directly to and would lose their status of gravitational scalar fields.444The nature of these scalar fields (gravitational or not) depends on the conformal frame Sotiriou:2007zu . Here we refer to the Jordan conformal frame. is assumed to be positive to keep the effective gravitational coupling positive.
The matrix acts as a Riemannian metric on the scalar field space of coordinates . can be taken to be symmetric without loss of generality because it multiplies the combination of kinetic terms symmetric in and . The elements of are all positive to avoid introducing unstable phantom fields. In general, also the potential depends on multiple fields (although it is not important that it depends on all these fields, which is instead crucial for the coupling function ).
Since the matrix is real and symmetric, it can be diagonalized at each spacetime point and has positive eigenvalues, turning the sum of kinetic terms appearing in the action (12) into
where a bar denotes fields in the system of principal axes of the matrix in field space, and
(14) |
is the diagonal form of . This diagonalization, however, is not crucial and we will not use it explicitly, retaining the non-diagonal form of in our formulae.
3 Multi-fluid decomposition
The total stress-energy tensor is obtained by varying the action (12) with respect to . Using and , the associated equation of motion reads
(15) |
where is the Einstein tensor, is the matter stress-energy tensor and
(16) |
The equation of motion obtained by variation of the action with respect to reads
(17) |
We can obtain the expression of the Ricci scalar from (15),
where is the trace of the matter stress-energy tensor. With this expression, Eq. (17) turns into
Assuming , we use the matrix to isolate , obtaining
(18) |
The goal of the decomposition given here is to separate so that each part can be decomposed in the frame of a given fluid. Each fluid then receives an individual stress-energy tensor contribution. The number of purely convective terms is minimised by such a decomposition to allow for a clearer description of the intrinsic dissipative properties of each fluid.
Assuming the gradient of each scalar field to be timelike,
(19) |
we define the -fluid four-velocity
(20) |
At this point, in order to avoid ambiguities, all the multiplet summations in this section will be written with an explicit summation symbol. The above identification between a scalar field and an associated effective fluid allows us to rewrite the scalar field derivatives in term of kinematic quantities Miranda:2022wkz . The second derivative in Eq. (16) can be expanded as
(21) |
where is the three-metric of the hypersurface orthogonal to the four-vector , is the expansion tensor associated with the -fluid, and .
With this result, Eq. (16) becomes
(22) |
where
(23) |
Since this equation does not depend on four-velocity gradients, we can interpret it as a purely inviscid contribution to the stress-energy tensor mixture.
If we rewrite the metric as
(24) |
and we define
(25) |
then, writing explicitly the summations, the stress-energy tensor assumes the form {widetext}
(26) |
which is interpreted as a mixture of interacting imperfect fluids.
4 “Average” or “-” description
Let us discuss another possible procedure. In the following we redefine the fields but, before proceeding, it is essential to note (and remember through the rest of this work) that all these fields couple directly with the Ricci scalar through and they all play a role of in determining the properties of the effective fluid equivalent to the tensor-multi-scalar theory and the effective gravitational coupling . (Their role may be different as, in general, is not symmetric in all its arguments.) In particular, the effective temperature of this multi-component fluid is determined by all the fields and the upcoming redefinition of these fields does not change this fact.
We proceed to redefine the scalar field multiplet as in Ref. Hohmann:2016yfd , which is standard practice in single-scalar-tensor gravity. We can rename the coupling function by electing it to be a Brans-Dicke-like scalar,
(27) |
and we then have the scalar fields . This mathematically convenient procedure effectively makes only the field couple explicitly to but the reader should not be fooled into believing that the remaining fields do not couple to gravity. In fact, all the fields are coupled to (and also to each other), which makes them couple also to gravity. Indeed, they were explicitly coupled to gravity before the field redefinition and the physics does not change. The action (12) is recast as Hohmann:2016yfd
(28) | |||||
where
(29) | |||||
(30) | |||||
(31) |
The field equations for obtained by varying the action (28) are Hohmann:2016yfd
(32) | ||||
(33) | ||||
(34) |
where we have used the notation and .
Using the metric field equations we can express the Ricci scalar in terms of the matter and effective stress-energy tensors,
(35) |
where . Then, the equation of motion for turns into
(36) |
Finally, we define the effective stress-energy tensor as
(37) |
We can now move to the effective fluid picture.
5 Comoving (Eckart) frame of -fluid
Assume that the gradient of is timelike; using
(38) |
we define the effective fluid four-velocity
(39) |
which is normalized, (but the sign of the right-hand side of this definition must be adjusted to keep a future-oriented vector, which is crucial in discussions of dissipation which is time-irreversible). In general, the -fluids are tilted with respect to the -fluid, i.e., and have different directions. Using the of the -fluid we perform the usual splitting of spacetime into the time direction and the 3-space “seen” by the observer with four-velocity . This 3-space has Riemannian metric
(40) |
The kinematic quantities (expansion tensor , expansion scalar , shear tensor , shear scalar, and acceleration ) associated with are the same as those calculated for single-scalar-tensor gravity in Faraoni:2018qdr . In fact, their definitions are purely kinematic and theory-independent since they do not use the field equations but only the definition (39) of . These kinematic quantities are straightforward, although lengthy to compute. Since they are used here, we report them in A.
The field equations (32) have the form of effective Einstein equations with an effective stress-energy tensor in their right-hand side, which can be seen as the stress-energy tensor of a dissipative multi-component fluid of the form
(41) |
where
(42) |
is the effective energy density,
(43) |
is the effective heat current density describing heat conduction,
(44) |
is the effective stress tensor,
(45) |
is the effective isotropic pressure, and the trace-free part of the stress tensor
(46) |
is the effective anisotropic stress tensor. , , and are purely spatial with respect to . The fluid description is obtained by expressing the derivatives of in terms of the relative effective fluid four-velocity (39) and kinematic quantities,
(47) | ||||
(48) | ||||
(49) |
Furthermore, we have
(50) |
therefore the -equation of motion reads
(51) |
We need these equations to eliminate the dependence of on and on . Indeed, prior to using the equation of motion for , one obtains
(52) |
Using the decomposition , defining , and taking into account the symmetry , the interacting terms contribute to the density, pressure, heat flux and anisotropic stress,
(53) |
and the stress-energy tensor reads
(54) |
Then, it is straightforward to obtain the effective fluid quantities
(55) | ||||
(56) | ||||
(57) |
where an overdot denotes differentiation along the lines of the -fluid, i.e., .
At this point, we can identify the various contributions to the effective energy tensor as
(58) | ||||
(59) | ||||
(60) | ||||
(61) | ||||
(62) | ||||
(63) |
where
(64) | ||||
(65) | ||||
(66) | ||||
(67) |
while
(68) |
are the bulk and shear viscosity coefficients, respectively.
In this particular case in which the Lagrangian is linear in , the -equation of motion reveal that does not contain derivatives of the -fluid four-velocity, therefore it only contributes to the inviscid pressure. However, it contains -terms related to the interactions.
Finally, the -terms contribute only to the inviscid part of the effective stress-energy tensor (because and depend only on first derivatives of the fields), to the heat flux, and to the shear viscosity. In the general case of the previous section, all the fields contribute to both viscous and inviscid part.
6 Conclusions
The picture of the effective fluid equivalent of tensor-multi-scalar gravity that emerges from the previous sections is the following. Because all the original gravitational scalar fields couple explicitly to the Ricci scalar, they are automatically coupled to each other. In addition, they may have explicit couplings to each other through the functions and , but this is not necessary for them to be mutually coupled. In the multi-fluid interpretation, this property could correspond to these fields being thermalized, but this interpretation is not corroborated in any obvious way by the field equations and remains rather arbitrary.
Acknowledgements.
M. M. is grateful for the support of Istituto Nazionale di Fisica Nucleare (INFN) iniziativa specificha MOONLIGHT2, and for the hospitality at Bishop’s University. This work is supported, in part, by the Natural Sciences & Engineering Research Council of Canada (grant 2016-03803 to V. F.).The authors declare no conflict of interest.
Appendix A Kinematic quantities of the -fluid
The (double) projection of the velocity gradient onto the 3-space orthogonal to
(69) |
is decomposed into its symmetric and antisymmetric parts. The latter is identically zero because the -fluid is derived from a scalar field. The symmetric part is further decomposed into its trace-free and pure trace parts. This results in
(70) |
where the expansion tensor is the symmetric part of , is its trace, the vorticity tensor , and the symmetric, trace-free shear tensor is
(71) |
Expansion, vorticity, and shear are purely spatial,
(72) | |||||
For general fluids, it is Ellis:1971pg ; EMMacC
The kinematic quantities of the -fluid relevant for the present discussion are computed in Faraoni:2018qdr and are as follows:
the acceleration is
(75) | |||||
(76) | |||||
The expansion scalar reads
(77) | |||||
while the shear tensor is
(78) | |||||
References
- (1) C. Eckart, “The thermodynamics of irreversible processes. 3. Relativistic theory of the simple fluid,” Phys. Rev. 58 (1940), 919-924 doi:10.1103/PhysRev.58.919
- (2) V. Faraoni and J. Côté, “Imperfect fluid description of modified gravities,” Phys. Rev. D 98, no.8, 084019 (2018) doi:10.1103/PhysRevD.98.084019 [arXiv:1808.02427 [gr-qc]].
- (3) V. Faraoni and A. Giusti, “Thermodynamics of scalar-tensor gravity,” Phys. Rev. D 103, no.12, L121501 (2021) doi:10.1103/PhysRevD.103.L121501 [arXiv:2103.05389 [gr-qc]].
- (4) V. Faraoni, A. Giusti and A. Mentrelli, “New approach to the thermodynamics of scalar-tensor gravity,” Phys. Rev. D 104, no.12, 124031 (2021) doi:10.1103/PhysRevD.104.124031 [arXiv:2110.02368 [gr-qc]].
- (5) C. Eling, R. Guedens, and T. Jacobson, “Non-equilibrium thermodynamics of spacetime,” Phys. Rev. Lett. 96 (2006) 121301, doi:10.1103/PhysRevLett.96.121301 [arXiv:gr-qc/0602001 [gr-qc]].
- (6) T. Jacobson, “Thermodynamics of space-time: The Einstein equation of state,” Phys. Rev. Lett. 75 (1995) 1260, doi:10.1103/PhysRevLett.75.1260 [arXiv:gr-qc/9504004 [gr-qc]].
- (7) M. S. Madsen, “Scalar Fields in Curved Space-times,” Class. Quant. Grav. 5, 627-639 (1988) doi:10.1088/0264-9381/5/4/010
- (8) L. O. Pimentel, “Energy Momentum Tensor in the General Scalar-Tensor Theory,” Class. Quant. Grav. 6 (1989), L263-L265 doi:10.1088/0264-9381/6/12/005
- (9) U. Nucamendi, R. De Arcia, T. Gonzalez, F. A. Horta-Rangel and I. Quiros, “Equivalence between Horndeski and beyond Horndeski theories and imperfect fluids,” Phys. Rev. D 102 (2020) no.8, 084054, doi:10.1103/PhysRevD.102.084054 [arXiv:1910.13026 [gr-qc]].
- (10) A. Giusti, S. Zentarra, L. Heisenberg and V. Faraoni, “First-order thermodynamics of Horndeski gravity,” [arXiv:2108.10706 [gr-qc]].
- (11) M. Miranda, D. Vernieri, S. Capozziello and V. Faraoni, “Fluid nature constrains Horndeski gravity,” [arXiv:2209.02727 [gr-qc]].
- (12) A. Giusti, S. Giardino and V. Faraoni, “Past-directed scalar field gradients and scalar-tensor thermodynamics,” [arXiv:2210.15348 [gr-qc]].
- (13) S. Giardino, V. Faraoni and A. Giusti, “First-order thermodynamics of scalar-tensor cosmology,” JCAP 04, no.04, 053 (2022) doi:10.1088/1475-7516/2022/04/053 [arXiv:2202.07393 [gr-qc]].
- (14) V. Faraoni, A. Giusti, S. Jose and S. Giardino, “Peculiar thermal states in the first-order thermodynamics of gravity,” Phys. Rev. D 106, no.2, 024049 (2022) doi:10.1103/PhysRevD.106.024049 [arXiv:2206.02046 [gr-qc]].
- (15) M. Miranda, D. Vernieri, S. Capozziello and V. Faraoni, Phys. Rev. D 105, no.12, 124024 (2022) doi:10.1103/PhysRevD.105.124024 [arXiv:2204.09693 [gr-qc]].
- (16) V. Faraoni and T. B. Françonnet, “Stealth metastable state of scalar-tensor thermodynamics,” Phys. Rev. D 105, no.10, 104006 (2022) doi:10.1103/PhysRevD.105.104006 [arXiv:2203.14934 [gr-qc]].
- (17) V. Faraoni, P. A. Graham and A. Leblanc, “Critical solutions of nonminimally coupled scalar field theory and first-order thermodynamics of gravity,” [arXiv:2207.03841 [gr-qc]].
- (18) R. Maartens, “Causal thermodynamics in relativity,” [arXiv:astro-ph/9609119 [astro-ph]].
- (19) N. Andersson and G. L. Comer, “Relativistic fluid dynamics: Physics for many different scales,” Living Rev. Rel. 10, 1 (2007), doi:10.12942/lrr-2007-1 [arXiv:gr-qc/0605010 [gr-qc]].
- (20) G.F.R. Ellis, R. Maartens, and M.A.H. MacCallum, Relativistic Cosmology (Cambridge University Press, Cambridge, UK, 2012).
- (21) R. M. Wald, General Relativity (Chicago University Press, Chicago, 1984).
- (22) R. Maartens, T. Gebbie and G. F. R. Ellis, “Covariant cosmic microwave background anisotropies. 2. Nonlinear dynamics,” Phys. Rev. D 59, 083506 (1999) doi:10.1103/PhysRevD.59.083506 [arXiv:astro-ph/9808163 [astro-ph]].
- (23) C. A. Clarkson, A. A. Coley, E. S. D. O’Neill, R. A. Sussman and R. K. Barrett, “Inhomogeneous cosmologies, the Copernican principle and the cosmic microwave background: More on the EGS theorem,” Gen. Rel. Grav. 35, 969-990 (2003) doi:10.1023/A:1024094215852 [arXiv:gr-qc/0302068 [gr-qc]].
- (24) C. Clarkson and R. Maartens, “Inhomogeneity and the foundations of concordance cosmology,” Class. Quant. Grav. 27, 124008 (2010) doi:10.1088/0264-9381/27/12/124008 [arXiv:1005.2165 [astro-ph.CO]].
- (25) V. Faraoni, S. Giardino, A. Giusti and R. Vanderwee, “Scalar field as a perfect fluid: thermodynamics of minimally coupled scalars and Einstein frame scalar-tensor gravity,” [arXiv:2208.04051 [gr-qc]].
- (26) I. Arsene et al. [BRAHMS], “Quark gluon plasma and color glass condensate at RHIC? The Perspective from the BRAHMS experiment,” Nucl. Phys. A 757, 1-27 (2005) doi:10.1016/j.nuclphysa.2005.02.130 [arXiv:nucl-ex/0410020 [nucl-ex]].
- (27) B. B. Back et al. [PHOBOS], “The PHOBOS perspective on discoveries at RHIC,” Nucl. Phys. A 757, 28-101 (2005) doi:10.1016/j.nuclphysa.2005.03.084 [arXiv:nucl-ex/0410022 [nucl-ex]]
- (28) J. Adams et al. [STAR], “Experimental and theoretical challenges in the search for the quark gluon plasma: The STAR Collaboration’s critical assessment of the evidence from RHIC collisions,” Nucl. Phys. A 757, 102-183 (2005) doi:10.1016/j.nuclphysa.2005.03.085 [arXiv:nucl-ex/0501009 [nucl-ex]].
- (29) K. Adcox et al. [PHENIX], “Formation of dense partonic matter in relativistic nucleus-nucleus collisions at RHIC: Experimental evaluation by the PHENIX collaboration,” Nucl. Phys. A 757, 184-283 (2005) doi:10.1016/j.nuclphysa.2005.03.086 [arXiv:nucl-ex/0410003 [nucl-ex]].
- (30) A. Monnai, “Landau and Eckart frames for relativistic fluids in nuclear collisions,” Phys. Rev. C 100, no.1, 014901 (2019) doi:10.1103/PhysRevC.100.014901 [arXiv:1904.11940 [nucl-th]].
- (31) M. Romero-Munoz, L. Dagdug and G. Chacon-Acosta, “Vanishing condition for the heat flux of a relativistic fluid in a moving frame,” J. Phys. Conf. Ser. 545, no.1, 012012 (2014) doi:10.1088/1742-6596/545/1/012012
- (32) M. Hohmann, L. Jarv, P. Kuusk, E. Randla and O. Vilson, “Post-Newtonian parameter for multiscalar-tensor gravity with a general potential,” Phys. Rev. D 94, no.12, 124015 (2016) doi:10.1103/PhysRevD.94.124015 [arXiv:1607.02356 [gr-qc]].
- (33) T. P. Sotiriou, V. Faraoni and S. Liberati, “Theory of gravitation theories: A No-progress report,” Int. J. Mod. Phys. D 17, 399-423 (2008) doi:10.1142/S0218271808012097 [arXiv:0707.2748 [gr-qc]].
- (34) L. D. Landau and E. M. Lifshitz, Fluid Mechanics (Pergamon Press, New York, 1959).
- (35) G. F. R. Ellis, “Relativistic cosmology,” Proc. Int. Sch. Phys. Fermi 47, 104-182 (1971). Reprinted in Gen. Relativ. Gravit. 41, 581–660 (2009), doi:10.1007/s10714-009-0760-7
- (36) T. Damour and G. Esposito-Farese, “Testing gravity to second postNewtonian order: A field theory approach,” Phys. Rev. D 53, 5541-5578 (1996) doi:10.1103/PhysRevD.53.5541 [arXiv:gr-qc/9506063 [gr-qc]].
- (37) D. I. Kaiser, “Conformal Transformations with Multiple Scalar Fields,” Phys. Rev. D 81, 084044 (2010) doi:10.1103/PhysRevD.81.084044 [arXiv:1003.1159 [gr-qc]].