Ultrarelativistic spinning objects in non-local ghost-free gravity
Abstract
We study the gravitational field of ultrarelativistic spinning objects (gyratons) in a modified gravity theory with higher derivatives. In particular, we focus on a special class of such theories with an infinite number of derivatives known as “ghost-free gravity” that include a non-local form factor such as , where is the scale of non-locality. First, we obtain solutions of the linearized ghost-free equations for stationary spinning objects. To obtain gyraton solutions we boost these metrics and take their Penrose limit. This approach allows us to perform calculations for any number of spacetime dimensions. All solutions are regular at the gyraton axis. In four dimensions, when the scale non-locality tends to zero, the obtained gyraton solutions correctly reproduce the Aichelburg–Sexl metric and its generalization to spinning sources found earlier by Bonnor. We also study the properties of the obtained four-dimensional and higher-dimensional ghost-free gyraton metrics and briefly discuss their possible applications.
I Introduction
The study of the gravitational field of ultrarelativistic particles and beams of light is a very old subject. The first solution describing the gravitational field of beam of light (“pencil”) was found by Tolman, Ehrenfest and Podolski in 1931 Tolman et al. (1931). These authors used a linear approximation of the Einstein equations. One of their main conclusions was that the gravitational force acting on a massless particle moving in the same direction as the beam of light vanishes. Later, Bonnor Bonnor (1969a) presented a solution for the gravitational field produced by a cylindrical beam of a null fluid. This model can be interpreted as a description of a high frequency light beam in the geometric optics approximation when diffraction effects are neglected.111More recent studies of light beams beyond the geometric optics approximation can be found in Schneiter et al. (2018) and references therein. The gravitational field of a spinning pencil of light was obtained by Bonnor in 1970 Bonnor (1970), see also Refs. Mitskievic and Kumaradtya (1989); Lynden-Bell and Bičák (2017). Higher-dimensional solutions describing the gravitational field of spinning ultrarelativistic objects and light beams were obtained in Frolov and Fursaev (2005); Frolov et al. (2005). The latter work introduced the name “gyraton” for such spinning ultrarelativistic objects, which is now used in the literature quite frequently. There exist different generalizations of standard gyraton solutions, such as solutions for charged gyratons Frolov and Zelnikov (2006), gyratons in asymptotically AdS spacetimes Frolov and Zelnikov (2005), in a generalized Melvin universe with cosmological constant Kadlecova and Krtous (2016), and string gyratons in supergravity Frolov and Lin (2006). Gyraton solutions of the Einstein equations belong to the wide class of so-called Kundt metrics Stephani et al. (2003). A comprehensive discussion of gyratons in the Robinson–Trautman and Kundt classes of metrics can be found in Kadlecova et al. (2009); Krtous et al. (2012); Podolsky et al. (2014); Podolsky and Svarc (2019).
There is another problem that has been widely discussed in the literature and which is closely related to gyratons. In 1970, Aichelburg and Sexl Aichelburg and Sexl (1971) constructed a metric of a massive ultrarelativistic particle. In its rest frame, the gravitational field of such a particle of mass is described by the Schwarschild metric. In order to obtain the metric when this particle moves with a very high velocity they applied a boost transformation and considered the limit wherein the velocity of the object tends to the speed of light, and hence the Lorentz factor diverges. They demonstrated that keeping the value of the energy fixed yields a limiting metric which is now called the Aichelburg–Sexl solution. For this solution the gravitational field of a particle is localized at the null plane tangent to the null vector of the particle’s four-velocity. Later, Penrose Penrose (1976) demonstrated that this is a generic property of any metric that is boosted to the speed of light, provided the corresponding energy is kept fixed, and this special limiting case has hence been dubbed “Penrose limit.” Aichelburg–Sexl-type metrics have been widely used for the study of the gravitational interaction of two ultrarelativistic particles as well as black hole production via their collision. The area of the apparent horizon in this process just before the moment of collision was calculated in Eardley and Giddings (2002) and has been widely used for estimating black hole formation cross sections in the collision of ultrarelativistic particles (see e.g. Yoshino and Nambu (2003); Giddings and Rychkov (2004); Yoshino and Rychkov (2005); Yoshino and Mann (2006); Yoshino et al. (2007) and references therein).
Since in the Penrose limit the initial mass of the particle tends to zero, one can obtain the Aichelburg–Sexl metric by starting with a linearized, weak-field gravity solution for a point-like particle. By considering a superposition of such solutions it is easy to construct the gravitational field of extended objects in linearized gravity. In particular, one may consider first a line distribution of mass, and then boost the solution. Due to the Lorentz contraction in the direction of motion the visible size of the body in this direction shrinks. This means that in order to obtain a solution for the ultrarelativistic case featuring a finite energy distribution profile one needs not only to take the Penrose limit keeping constant, but also simultaneously keep the parameter fixed, where is the size of the object in the direction of motion. Such a procedure can be applied to a spinning object provided the rotation takes place within the plane orthogonal to the direction of motion. One can show that in such a procedure one reconstructs the gravitational field of a gyraton. This method is described in details in chapter 5 of the book Frolov and Zelnikov (2011).
The goal of this paper is to construct gyraton-like solutions in so-called “ghost-free” gravity. This is an important special class of modified gravity theories that introduces non-locality by means of non-local form factors of the type . This modification becomes relevant only at small scales comparable to , and hence this type of theories can be considered an ultrviolet (UV) modification of gravity. To that end, the main motivation for this study is that a small scale modification of gravity might become important for the process of mini black hole formation in the collision of ultrarelativistic particles. For example, it was shown that if the Einstein–Hilbert action is modified by the inclusion of higher-derivative as well as infinite-derivative terms, there exists a mass gap for black hole formation Frolov (2015); Frolov et al. (2015); Frolov and Zelnikov (2016a); Giacchini and de Paula Netto (2019a).
While non-locality has been explored for quite some time Yukawa (1949a, b, 1950a, 1950b); Efimov (1967, 1972); Efimov et al. (1977); Efimov (1974, 1968), the particular class studied here is motivated from string theory Frampton and Okada (1988); Tseytlin (1995); Tomboulis (1997); Biswas et al. (2006) as well as non-commutative geometry Spallucci et al. (2006). These non-local theories of gravity have appealing UV properties Biswas et al. (2010a); Modesto (2012) and are under active investigation. It has been demonstrated that in the weak-field regime this class of theories regularizes the gravitational field of point-like sources Edholm et al. (2016); Buoninfante et al. (2018a, b); Giacchini and de Paula Netto (2019b) as well as thin brane-like extended objects Boos et al. (2018a); Boos (2020); Kolar and Mazumdar (2020). For results in the strong-field regime in connection with black holes we refer to Conroy et al. (2015); Li et al. (2015); Calcagni and Modesto (2017); Koshelev et al. (2018); Boos et al. (2019a) and references therein; for cosmological applications see Biswas et al. (2010b); Calcagni et al. (2014). Non-local infinite-derivative form factors have also been explored in quantum theory Boos et al. (2018b); Buoninfante et al. (2019a) as well as quantum field theory Shapiro (2015); Frolov and Zelnikov (2016b); Modesto et al. (2018); Asorey et al. (2018); Calcagni et al. (2018); Buoninfante et al. (2019b); Boos et al. (2019b).
This paper is organized as follows: we begin by discussing the solutions of the modified gravity equations in the weak-field approximation. In Sec. II we consider a wide class of theories for which the linearized action is quadratic in curvature and contains an arbitrary number of derivatives. General analysis of such theories shows that their action can be rewritten in a form which contains two scalar functions of the d’Alembert operator, where an additional requirement of the absence of scalar modes establishes a relation between these functions Biswas et al. (2012). We obtain a general solution of the field equations for a stationary distribution of spinning matter in four and higher dimensions, paying special attention to extended pencil-type distribution of spinning matter. In Sec. III we apply the boost transformation to these pencil-like distributions of matter choosing the velocity being directed along the pencil axis. After this, we obtain the Penrose limit for the boosted metrics and find the gravitational field of ultrarelativistic extended spinning objects (gyratons) in four and higher dimensions. The explicit form of these metrics and their properties are discussed in Sec. IV. We summarize our findings and mention possible future applications of these solutions in Sec. V.
II Spinning objects in the weak-field approximation of infinite-derivative gravity
II.1 Linearized equations
Our goal is to obtain the metric for an ultrarelativistic spinning object (gyraton) in infinite-derivative gravity. The method of solving the problem is the following. First one finds a solution of the gravitational field equations in the object’s rest frame. Then one transforms this solution to a new reference frame moving with a constant velocity with respect to the original one. Finally, one takes the limit while keeping the energy fixed. In this Penrose limit the original mass of the object effectively approaches zero, which implies that in order to obtain the corresponding gyraton metric one may start with a solution with very small mass . One can expect that higher-order curvature corrections—which are proportional to second and higher orders in the mass —are therefore small. In this section we will discuss the field equations of linearized infinite-derivative gravity and present their solutions for an extended, slowly spinning object of small mass.
We denote by Cartesian coordinates in -dimensional Minkowski spacetime and use indices from the beginning of the Greek alphabet to label spatial coordinates. The Minkowski metric in the dimensional spacetime is
(1) |
where denotes the flat -dimensional metric. We denote by a small deviation of the metric from the flat background,
(2) |
One can show that the most general linearized action in a Lorentz invariant theory with an arbitrary number of derivatives and quadratic in the perturbation can be written in the form Biswas et al. (2012)
(3) | ||||
where is the d’Alembert operator of Minkowski space, . The functions and can be chosen freely to parametrize different Lorentz-invariant modifications of gravity, subject only to the constraint
(4) |
which guarantees the proper Newtonian limit; see also the related discussions in Refs. Biswas et al. (2012); Buoninfante et al. (2019b). In the case of one recovers the Fierz–Pauli action and linearized General Relativity.
The field equations corresponding to the action (3) are
(5) |
where is the energy-momentum tensor of matter, and denotes the trace of . From now on we shall restrict ourselves to the case of
(6) |
This condition guarantees that no extra scalar modes are present in the theory Biswas et al. (2012). We denote
(7) |
The inverse transformation is
(8) |
We also impose the gauge conditions . Then, Eq. (LABEL:EQN) simplifies greatly and takes the form
(9) |
The conservation law implies that the imposed gauge conditions are consistent.
II.2 Stationary solutions for extended sources
We assume that does not depend on time. For a stationary metric generated by such a stress-energy tensor the -operator reduces to the -dimensional Laplace operator . We denote
(10) |
Then, we can solve the field equations (9) by using the static Green function
(11) |
The solution then takes the form
(12) |
The expression for the perturbation of the metric can be found from (12) by using relation (8). For the stress-energy tensor (87) given in the Appendix one has
(13) |
A solution of the field equations (9) for this source can be written as follows:
(14) | ||||
(15) | ||||
(16) |
Due to the translational symmetry of Eq. (11), the Green function is a function of , while due to the spherical symmetry it depends on the radius variable alone. Thus one has222In order to keep the notation somewhat manageable we shall use the same symbol for the Green function with vectorial arguments and with the scalar radius argument .
(17) |
As has been shown previously Boos et al. (2018a), the Green function in spatial dimensions is related to the Green function in spatial dimensions via the recursion formulas
(18) | ||||
(19) |
Using this relation one can rewrite (16) in the form
(20) |
For calculations it is convenient to use the following representation of the static Green function given in Frolov and Zelnikov (2016a),
(21) |
Last, let us mention that in the limit the above representation (21) gives
(22) |
In the above we have made use of the constraint (4).333Note that the -regularization is only required for . If one instead calculates the full expression (21) for a given choice of the function , see a detailed description in Appendix C, this regularization is not required, and one may simply take the limit to recover (22). is the static Green function of linearized General Relativity Myers and Perry (1986) in spatial dimensions, which guarantees that for isolated sources in the far field regime one reproduces the standard asymptotics of General Relativity.
II.3 Point particles
The stress-energy of a point-like spinning particle can be written in the form
(23) |
where is the mass of the particle and is a constant antisymmetric matrix parametrizing its angular momentum. A solution for the perturbed metric (14)–(16) for such a source takes the form444In four-dimensional spacetime, this solution can be used to obtain a metric for a spinning ring discussed in Buoninfante et al. (2018c).
(24) |
At large distances one recovers the standard expressions known from linearized General Relativity Myers and Perry (1986):
(25) | ||||
(26) |
We choose the sign of such that in the three-dimensional case one obtains the standard Lense–Thirring expression ( and )
(27) |
II.4 Extended objects: pencils
In order to simplify our presentation further, let us consider a special type of spinning objects. That is, we assume that it has finite extension in one spatial directions, while its transverse size is zero. We call such an object a thin spinning pencil or simply “pencil.”
II.4.1 Coordinates
Let us consider two frames. The first one is frame where the matter creating the gravitational field is at rest. The second frame moves with a constant velocity with respect to . We adapt now the choice of the coordinates which is convenient for this situation. Let be a coordinate along the vector of velocity of and denote by the coordinates orthogonal to the -direction. To distinguish the rest frame coordinates from the coordinates in the boosted frame we use a bar for the rest frame coordinates and write
(28) |
The index enumerates the coordinates transverse to the direction of motion. We omit the bar for the coordinates since the Lorentz transformation for the motion in -direction does not affect their values. The background Minkowski metric is
(29) |
Here, are coordinates in the rest frame and are the corresponding coordinates in the moving frame . In what follows, we denote all quantities defined with respect to the rest frame with a bar. For example, the radial distance from the origin to a point is . Let us specify the coordinates orthogonal to the -direction further:
(30) |
One can say that the -dimensional “transverse space” orthogonal to the -axis is spanned by mutually orthogonal two-planes , and are right-handed coordinates in these planes. We shall refer to these planes as Darboux planes. If the number of spacetime dimensions is odd one has and besides these two-planes there exists an additional one-dimensional -axis which is orthogonal to each of the planes as well as to -axis. In even spacetime dimensions there is no such additional coordinate: for example, in four spacetime dimensions there exists only one two-plane orthogonal to the -direction. We denote by and unit vectors along the -axis and -axis, respectively. The 1-forms dual to these vectors are and such that the volume 2-form for each Darboux plane is given by .
II.4.2 Gravitational field
The stress-energy tensor of a thin spinning pencil is
(31) |
We assume that this object has a finite length in , such that both and vanish when is outside some interval . We call the length of the pencil. The mass and the angular momentum of such a pencil are
(32) | ||||
(33) | ||||
(34) |
see also Appendix A. The quantities and are the mass and angular momentum line densities, respectively. They describe the distribution of the mass and angular momentum along the pencil. In what follows we chose both the total angular momentum and its density to be orthogonal to the -direction. Consequently, they have identical Darboux two-planes such that the antisymmetric matrix is of the form
(35) |
In the above, the are functions of alone. By construction, the total angular momentum has a similar Darboux form.
The gravitational field of a thin spinning pencil is
(36) | |||
(37) | |||
(38) |
where we defined the auxiliary expression
(39) |
For time-independent objects that are extended also in the transverse direction orthogonal to one may use a similar method to construct their gravitational field. Then, however, the energy-momentum (31) no longer factorizes in a -part and a transverse part, but Eq. (12) still applies.
III Ultrarelativistic objects: gyratons
Now that we have found the gravitational field of a thin pencil in the weak-field limit for any number of dimensions in a wide range of infinite-derivative theories, let us address the ultrarelativistic case arising from performing a boost in the -direction.
In particular, we shall be interested in the so-called Penrose limit. This limit consists of (i) boosting a stationary solution to velocity , and then (ii) taking the limit while keeping the product fixed, where
(40) |
is the Lorentz factor, and the mass is given by (32). Moreover, we shall also assume that remains constant during the boost. In this limit the object becomes asymptotically null, and the gravitational fields of these ultrarelativistic objects are called gyraton fields.
III.1 Green function representation
Before performing the boost, and, subsequently, the Penrose limit, let us briefly mention a useful representation of the static Green function given by
(41) | ||||
(42) |
The function is the -dimensional heat kernel in imaginary time and therefore satisfies
(43) | ||||
(44) |
The derivation of the representation (41) for the static Green function is given in Appendix B. The following property makes this representation very useful for the study of the Penrose limit of solutions: Relation (41) expresses the Green function as a double Fourier transform, wherein the radius enters only quadratically via the exponential function . Since , this exponent can be factorized, which in turn allows one to separate the dependence of the integrand on as . Hence, when applying the boost, only this factor is affected. As we shall demonstrate now, this observation allows us to perform the Penrose limit procedure in a very general and convenient form.
III.2 Penrose limit
We now apply the Penrose limit to our previously described linearized potentials of a spinning “pencil.” We parametrize the boost in the -direction via
(45) |
Let us first make a simple remark concerning the scaling properties of the pencil characteristics under a boost transformation (45). We assume that both mass and angular momentum are uniformly distributed along the pencil and their densities in the rest frame, and are constant. Because of the Lorentz contraction, the length of the same pencil, as measured in the moving frame is , while its energy is . As a result, the linear energy density of the pencil in frame is . In the Penrose limit the energy is taken to be fixed. Thus the energy density grows to infinity as . To keep it finite, one needs to rescale in the boost process, such that the length remains unchanged. It is easy to check that under such rescalings the components of the angular momentum remain the same and finite. Note that this is a result of our assumption that the angular momentum density is orthogonal to the direction of motion, because in that case its components are not affected by the boost.
For fixed , that is, for a fixed point in frame one has . This means that the frame moves in the negative direction of (“left”) with respect to the rest frame . In other words, a pencil which is at rest with respect to moves with a positive velocity in frame.
We introduce the retarded and advanced null coordinates in the frame defined as follows:
(46) |
Then (45) implies
(47) | ||||
(48) |
In the ultrarelativistic limit, , one has
(49) |
This implies that the matter distribution of such an ultrarelativistec pencil is located in the strip between and of spacetime; see Fig. 1.

Because we keep the ratio constant during the Penrose limit, the linear density scales as follows:
(50) |
This guarantees that in the Penrose limit the product and the ratio remain constant,
(51) |
The angular momentum line density “lives” in transverse space and its tensorial structure is unaffected from the boost. Using this property we define the boosted linear density of the angular momentum in the frame as follows:
(52) | ||||
(53) |
The total angular momentum of the boosted pencil therefore remains finite and has the form
(54) |
III.3 Metric
Under this boost and the Penrose limit, as defined above, the resulting metric takes the form
(55) |
where we defined
(56) |
Here, and are given by (37) and (38), respectively. The integrands in their representations contain the Green function . In order to understand their behavior under the Penrose limit we make use of relation (41). In this representation the only quantity which is “sensitive” to the boost is the heat kernel . It factorizes such that the boost-sensitive factor is the exponent of the form , which for large factors takes the form . To take the Penrose limit we use the following relation (see also Shankar (1994)):
(57) |
Denote and apply this relation to (41) to obtain
(58) |
where . Performing the limit in the relations (56) for the potential and the gravitomagnetic potential finally yields
(59) | ||||
(60) |
Introducing polar coordinates in each Darboux plane such that
(61) |
one may use the relation
(62) |
to rewrite the gravitomagnetic potential 1-form as
(63) |
which makes the rotational symmetry in each Darboux plane manifest.
IV Gravitational field of ghost-free gyratons
In this section we present and discuss gyraton-like solutions in General Relativity and in infinite-derivative non-local gravity. In General Relativity, the form factor is simply
(64) |
whereas in infinite-derivative “ghost-free” gravity one may postulate instead
(65) |
The static Green function (21) can be computed for a wide range of theories, but in the context of the present paper we shall consider General Relativity as well as two infinite-derivative theories corresponding to the choices and , which we shall hence refer to as and . It is also possible to extend these studies to arbitrary number of spatial dimensions .
IV.1 Gyratons in
IV.1.1 Gyraton metrics in General Relativity
As a warm-up, let us consider the well-known gyraton solutions of -dimensional General Relativity Bonnor (1969a, b); Frolov and Fursaev (2005); Frolov et al. (2005). The relevant two-dimensional and four-dimensional Green functions are
(66) |
Since in the transverse space is two-dimensional we have and . Therefore we may write , call the polar angle , and denote by the linear density of the angular momentum in the frame. Then, the gravitational potentials and are
(67) | ||||
(68) |
This gravitomagnetic field is locally exact such that
(69) |
Observe, however, that the gravitomagnetic charge does not vanish:
(70) |
Here, denotes a surface in the Darboux plane. For later convenience we may assume to be a circle of radius . However, in a given null plane this charge does not depend on the choice of the contour . As we shall see soon, this property is no longer valid in non-local gravity, and effectively the gravitomagnetic current is spread out of the line in the direction transverse to the motion.
IV.1.2 Gyraton metrics in ghost-free gravity
We consider now a similar gyraton solutions in the non-local theories and . The static Green function for theory can be written as
(71) |
where denotes the complementary exponential integral and is the exponential integral Olver et al. (2010),
(72) | ||||
(73) |
and is the Euler–Mascheroni constant. Then, the gravitational potentials and take the form
(74) | ||||
(75) |
This gravitomagnetic field is no longer exact and hence the gravitomagnetic charge depends on the radius,
(76) |
At large distances, , we recover the gyraton solution obtained in General Relativity. In theory one has
(77) |
where we defined . The gravitomagnetic charge now takes the form
(78) |
See Fig. 2 for a plot of these charges. Interestingly, the charge is monotonic, whereas the charge exhibits an oscillatory behavior.

IV.1.3 Curvature invariants
One may wonder about the geometric properties of the four-dimensional gyraton spacetime
(79) |
This spacetime is a pp-wave because it features a covariantly constant null Killing vector Stephani et al. (2003),
(80) |
This property remains valid for any choice of the functions , and , provided their functional dependence remains the same. Since pp-wave spacetimes have vanishing scalar curvature invariants one finds
(81) |
For this reason they remain unchanged for solutions found in the context of linearized infinite-derivative gravity as compared to linearized General Relativity.
IV.2 Gyratons in dimensions
IV.2.1 case
In five spacetime dimensions one has , which—as per Eq. (30)—implies that and . In this case there is only one Darboux plane orthogonal to as well as one additional -axis. Let us write the transverse distance as , where is the radial variable in the Darboux plane. Then, from Eqs. (59) as well as (63), one readily obtains
(82) | ||||
(83) |
where is the polar angle in the Darboux plane, is the angular momentum eigenfunction, and describes the density profile. The explicit expressions for the functions in linearized General Relativity as well as in and theories are given in Appendix C.
IV.2.2 Higher dimensions
In higher dimensions one can proceed analogously to find expressions for the gyraton metrics. Instead of repeating previous steps, we give here an algorithmic procedure of how to construct such solutions in an arbitrary number of higher dimensions.
First, given the number of spatial dimensions , determine the number of Darboux planes using (30). If is even there will be an independent -axis as well. Due to the rotational symmetry around the pre-boost -direction it makes sense to introduce polar coordinates in each Darboux plane called where labels the Darboux planes. This construction is unique, provided one fixes the direction of the polar angles to be right-handed with respect to the original -direction.
Second, one introduces the perpendicular radius variable according to
(84) |
Recall that if is even, and if is odd. Now one can insert this radius variable into (59) and (63). In order to determine the static Green function in higher dimensions one may utilize the recursion formulas (18) and (19) as well as Appendix C.
Last, one may want to start with a known line energy density as well as angular momentum line densities in the original rest frame. In that case, Eqs. (50) and (52) provide prescriptions as to how to retrieve the resulting functions and in retarded time.
Realistic gyratons may also have a finite transverse thickness, but due to the linearity of the problem it is always possible to supplement a transverse density function in (31) and construct the gravitational field of a “thick gyraton” by superposition.
V Discussion
The main goal of this paper is to study the gravitational field of ultrarelativistic spinning objects (gyratons) in the non-local ghost-free theory of gravity. Our starting point is a linearized set of equations for such a theory. In a general case they contain two entire functions of the d’Alembert operator of flat spacetime, and , called form factors, subject to the additional constraint that . We focused on a simple case when , which guarantees the absence of unphysical modes. The set of field equations in Cartesian coordinates takes the form of uncoupled scalar equations for the components of the gravitational field. When the source is time-independent these equations can be solved by using a static Green function defined as a solution of the equation
(85) |
In order to obtain a gyraton solution we first found a stationary solution, and then boosted it to the speed of light by means of the Penrose limit. The key observation which allowed us to obtain such a solution is the following: We demonstrated that the Green function can be expressed as a double Fourier transform of the heat kernel of the Laplace operator . In such a representation all the dependence on the coordinates of the Green function is shifted to the argument of the heat kernel, which has an exponential form , where is the distance between the points which enter as arguments in the Green function, . This exponent can be factorized, such that the dependence on the coordinate is universal and has the standard form . In the Penrose limit this term produces a delta function of the retarded time , . The remaining integral for the Green function in -dimensional space up to a constant coefficient coincides with . Thus in the Penrose limit one has schematically
(86) |
It should be emphasized that this result is quite general. In fact, in its derivation we do not use any special form of the operator , with the exception that and are non-zero on the real line.
To obtain explicit gyraton solutions we made additional assumptions. First of all, we chose a form factor of the form (65) which guarantees that no extra poles are present in the Green function . We also assumed the source of the gravitational field to be in the form of an infinitely thin spinning pencil. Then, we applied the boost transformation in the direction of this pencil and chose its internal rotation such that the resulting Darboux two-planes are orthogonal to the direction of motion. In four spacetime dimensions this just means that the pencil is spinning around the axis of the boost direction. Since the field equations in our approximation are linear, one can easily use the found gyraton solutions for infinitely thin pencils to obtain a similar solutions for “thick” gyratons, which may have non-trivial structure transverse to the direction of motion.
An interesting but expected property of the obtained ghost-free gyraton metrics is that even for infinitely thin -shaped gravitational sources, all solutions are regular at the gyraton axis. This is to be seen in stark contrast to the metrics obtained in linearized General Relativity, wherein the metric functions grow beyond all bounds as , representing a breakdown of the linear approximation scheme. Alternatively, treating the resulting gyraton metrics as geometries beyond the linear approach, the pathological behavior corresponds to a singularity in spacetime. Within linearized ghost-free gravity this pathology disappears entirely, making the linear approximation self-consistent at .
For this reason non-locality effectively spreads the matter and spin distribution of thin gyratons and thereby regularizes them. Another interesting result is that these ghost-free gyraton metrics are vanishing scalar invariant spacetimes: the local curvature invariants vanish. This may be considered as a consequence of a general observation made by Penrose that all metrics after ultrarelativistic boosts take the form of pp-waves Penrose (1976). In four spacetime dimensions it is known that the Aichelburg–Sexl metric Aichelburg and Sexl (1971) and its spinning generalisation Bonnor (1970) obtained by boosting linearized solutions of Einstein equation are in fact exact solutions of these non-linear equations. In higher dimensions this is not true Frolov et al. (2005). However, these higher-dimensional gyraton solutions belong to an important class of so-called Kundt metrics Stephani et al. (2003). It is interesting to check whether this is also true in complete (non-linear) ghost-free gravity.
Let us finally mention that the ghost-free gyraton solutions obtained in this paper may be used to study the collision of ultrarelativistic particles. In particular, they will allow one to understand the role of non-locality and spin in the process of micro-black hole formation Yoshino et al. (2007); Frolov (2015).
Note Added in Proof
It has been brought to our attention that exact pp-wave solutions in ghost-free infinite-derivative gravity have been studied in Ref. Kilicarslan (2019). Recently, these studies have been extended to exact wavelike “impulsive” solutions in anti-de Sitter spacetimes Dengiz et al. (2020), and it would be very interesting to understand the role of the gyraton metrics obtained in this paper in that context.
Acknowledgments
J.B. is grateful for a Vanier Canada Graduate Scholarship administered by the Natural Sciences and Engineering Research Council of Canada as well as for the Golden Bell Jar Graduate Scholarship in Physics by the University of Alberta. V.F. thanks the Natural Sciences and Engineering Research Council of Canada and the Killam Trust for their financial support.
Appendix A Mass and angular momentum of extended objects in higher dimensions
We denote by Cartesian coordinates in dimensional Minkowski spacetime and use indices from the beginning of the Greek alphabet to label spatial coordinates. Let us consider distribution of matter described by the stress-energy of the form
(87) |
where is an anti-symmetric tensor function. It is easy to check that this stress-energy tensor satisfies the required conservation law . Denote by a generator of the space-time translations, and by the generators of the rigid spatial rotations, then one has
(88) | ||||
(89) |
The conserved quantities related to these symmetries are
(90) | |||||
(91) |
or in an explicit form
(92) | |||||
(93) | |||||
(94) |
We assume that the stress-energy tensor (87) either vanishes outside some compact region, or it is sufficiently fast decreasing at far spatial distance, so that the surface terms arising as a result of integration by parts in (94) vanish. Simple calculations give
(95) |
The relation implies that the stress-energy tensor (87) is written in the center of mass frame.
Appendix B Heat kernel representation of ghost-free static Green functions
The static Green function considered in this paper satisfies the relation
(96) |
Here is a Laplace operator in -dimensional space. We denote by the -dimensional heat kernel of . It is defined as a solution of the equation
(97) |
obeying the boundary conditions
(98) |
It has the following explicit form:
(99) |
Let us define the object as a solution of the equation
(100) |
Then it is easy to check the required Green function can be written in the form
(101) |
We introduce now the Fourier transform of and its inverse by means of the relations
(102) |
Then we may write
(103) | ||||
(104) | ||||
(105) | ||||
(106) |
In the first equality we have used (100), then used the properties of the heat kernel via Eq. (99), and finally integrated by parts where the boundary terms vanish due to (98). The integral over can be easily calculated assuming that one takes care about its asymptotic behavior and uses the standard regularization. By using the relation
(107) |
one obtains
(108) |
which is the double Fourier representation for the Green function used in the main body of the paper.
Appendix C Static infinite-derivative ghost-free Green functions
Let us consider theories with the form factor of the form , where is a positive integer number. We refer to such a theory as ghost-free gravity and use the abbreviation for such a theory. For , and the corresponding theory is nothing but linearized General Relativity. Let us write and denote by a static Green function for theory in a space with dimensions. Such a Green function obeys the equation
(109) |
For , that is, in General Relativity, we also use the notation . The static Green functions can be found by using Eqs. (19)–(21). In this appendix we collect exact expressions for these Green functions for General Relativity as well as and theory for the number of spatial dimensions . Using the recursive relations (19) one can obtain their expression for . In what follows we will use the abbreviation .
(110) | ||||
(111) | ||||
(112) | ||||
(113) | ||||
(114) | ||||
(115) | ||||
(116) | ||||
(117) | ||||
(118) | ||||
(119) | ||||
(120) | ||||
(121) |
Here we use the standard notation for the hypergeometric function Olver et al. (2010).
References
- Tolman et al. (1931) R. C. Tolman, P. Ehrenfest, and B. Podolsky, Phys. Rev. 37, 602 (1931).
- Bonnor (1969a) W. B. Bonnor, Communications in Mathematical Physics 13, 163 (1969a).
- Schneiter et al. (2018) F. Schneiter, D. Ratzel, and D. Braun, Classical and Quantum Gravity 35, 195007 (2018).
- Bonnor (1970) W. B. Bonnor, Int. J. Theor. Phys. 3, 257 (1970).
- Mitskievic and Kumaradtya (1989) N. V. Mitskievic and K. K. Kumaradtya, Journal of Mathematical Physics 30, 1095 (1989), https://doi.org/10.1063/1.528380 .
- Lynden-Bell and Bičák (2017) D. Lynden-Bell and J. Bičák, Phys. Rev. D 96, 104053 (2017).
- Frolov and Fursaev (2005) V. Frolov and D. Fursaev, Phys.Rev. D71, 104034 (2005), arXiv:hep-th/0504027 [hep-th] .
- Frolov et al. (2005) V. P. Frolov, W. Israel, and A. Zelnikov, Phys.Rev. D72, 084031 (2005), arXiv:hep-th/0506001 [hep-th] .
- Frolov and Zelnikov (2006) V. P. Frolov and A. Zelnikov, Class. Quant. Grav. 23, 2119 (2006), arXiv:gr-qc/0512124 [gr-qc] .
- Frolov and Zelnikov (2005) V. P. Frolov and A. Zelnikov, Phys. Rev. D72, 104005 (2005), arXiv:hep-th/0509044 [hep-th] .
- Kadlecova and Krtous (2016) H. Kadlecova and P. Krtous, in Proceedings, Workshop on Black Holes and Neutron Stars (RAGtime 10-13): Opava, Czech Republic, September 15-17, 2008; September 20-22, 2009; September 15-17, 2010; September 14-16, 2011 (2016) arXiv:1610.08225 [gr-qc] .
- Frolov and Lin (2006) V. P. Frolov and F.-L. Lin, Phys. Rev. D73, 104028 (2006), arXiv:hep-th/0603018 [hep-th] .
- Stephani et al. (2003) H. Stephani, D. Kramer, M. A. MacCallum, C. Hoenselaers, and E. Herlt, Exact solutions of Einstein’s field equations, Cambridge Monographs on Mathematical Physics (Cambridge Univ. Press, Cambridge, 2003).
- Kadlecova et al. (2009) H. Kadlecova, A. Zelnikov, P. Krtous, and J. Podolsky, Phys. Rev. D80, 024004 (2009), arXiv:0905.2476 [gr-qc] .
- Krtous et al. (2012) P. Krtous, J. Podolsky, A. Zelnikov, and H. Kadlecova, Phys. Rev. D86, 044039 (2012), arXiv:1201.2813 [gr-qc] .
- Podolsky et al. (2014) J. Podolsky, R. Steinbauer, and R. Svarc, Phys. Rev. D90, 044050 (2014), arXiv:1406.3227 [gr-qc] .
- Podolsky and Svarc (2019) J. Podolsky and R. Svarc, Phys. Rev. D99, 044004 (2019), arXiv:1812.02635 [gr-qc] .
- Aichelburg and Sexl (1971) P. Aichelburg and R. Sexl, Gen.Rel.Grav. 2, 303 (1971).
- Penrose (1976) R. Penrose, “Any space-time has a plane wave as a limit,” in Differential Geometry and Relativity: A Volume in Honour of André Lichnerowicz on His 60th Birthday, edited by M. Cahen and M. Flato (Springer Netherlands, Dordrecht, 1976) pp. 271–275.
- Eardley and Giddings (2002) D. M. Eardley and S. B. Giddings, Phys. Rev. D66, 044011 (2002), arXiv:gr-qc/0201034 [gr-qc] .
- Yoshino and Nambu (2003) H. Yoshino and Y. Nambu, Phys. Rev. D67, 024009 (2003), arXiv:gr-qc/0209003 [gr-qc] .
- Giddings and Rychkov (2004) S. B. Giddings and V. S. Rychkov, Phys. Rev. D70, 104026 (2004), arXiv:hep-th/0409131 [hep-th] .
- Yoshino and Rychkov (2005) H. Yoshino and V. S. Rychkov, Phys. Rev. D71, 104028 (2005), [Erratum: Phys. Rev.D77,089905(E)(2008)], arXiv:hep-th/0503171 [hep-th] .
- Yoshino and Mann (2006) H. Yoshino and R. B. Mann, Phys. Rev. D74, 044003 (2006), arXiv:gr-qc/0605131 [gr-qc] .
- Yoshino et al. (2007) H. Yoshino, A. Zelnikov, and V. P. Frolov, Phys. Rev. D75, 124005 (2007), arXiv:gr-qc/0703127 [gr-qc] .
- Frolov and Zelnikov (2011) V. P. Frolov and A. Zelnikov, Introduction to black hole physics (Oxford Univ. Press, Oxford, 2011).
- Frolov (2015) V. P. Frolov, Phys. Rev. Lett. 115, 051102 (2015), arXiv:1505.00492 [hep-th] .
- Frolov et al. (2015) V. P. Frolov, A. Zelnikov, and T. de Paula Netto, JHEP 06, 107 (2015), arXiv:1504.00412 [hep-th] .
- Frolov and Zelnikov (2016a) V. P. Frolov and A. Zelnikov, Phys. Rev. D93, 064048 (2016a), arXiv:1509.03336 [hep-th] .
- Giacchini and de Paula Netto (2019a) B. L. Giacchini and T. de Paula Netto, Eur. Phys. J. C 79, 217 (2019a), arXiv:1806.05664 [gr-qc] .
- Yukawa (1949a) H. Yukawa, Phys. Rev. 76, 300 (1949a).
- Yukawa (1949b) H. Yukawa, Phys. Rev. 76, 1731 (1949b).
- Yukawa (1950a) H. Yukawa, Phys. Rev. 77, 219 (1950a).
- Yukawa (1950b) H. Yukawa, Phys. Rev. 80, 1047 (1950b).
- Efimov (1967) G. V. Efimov, Comm. Math. Phys. 5, 42 (1967).
- Efimov (1972) G. V. Efimov, Annals Phys. (NY) 71, 466 (1972).
- Efimov et al. (1977) G. V. Efimov, M. A. Ivanov, and O. Mogilevsky, Annals Phys. (NY) 103, 169 (1977).
- Efimov (1974) G. V. Efimov, Int. J. Theor. Phys. 10, 19 (1974).
- Efimov (1968) G. V. Efimov, Comm. Math. Phys. 7, 138 (1968).
- Frampton and Okada (1988) P. H. Frampton and Y. Okada, Phys. Rev. D37, 3077 (1988).
- Tseytlin (1995) A. A. Tseytlin, Phys. Lett. B363, 223 (1995), arXiv:hep-th/9509050 [hep-th] .
- Tomboulis (1997) E. Tomboulis, (1997), arXiv:hep-th/9702146 [hep-th] .
- Biswas et al. (2006) T. Biswas, A. Mazumdar, and W. Siegel, JCAP 0603, 009 (2006), arXiv:hep-th/0508194 [hep-th] .
- Spallucci et al. (2006) E. Spallucci, A. Smailagic, and P. Nicolini, Phys. Rev. D73, 084004 (2006), arXiv:hep-th/0604094 [hep-th] .
- Biswas et al. (2010a) T. Biswas, J. A. R. Cembranos, and J. I. Kapusta, Phys. Rev. D82, 085028 (2010a), arXiv:1006.4098 [hep-th] .
- Modesto (2012) L. Modesto, Phys. Rev. D86, 044005 (2012), arXiv:1107.2403 [hep-th] .
- Edholm et al. (2016) J. Edholm, A. S. Koshelev, and A. Mazumdar, Phys. Rev. D94, 104033 (2016), arXiv:1604.01989 [gr-qc] .
- Buoninfante et al. (2018a) L. Buoninfante, A. S. Koshelev, G. Lambiase, J. Marto, and A. Mazumdar, (2018a), arXiv:1804.08195 [gr-qc] .
- Buoninfante et al. (2018b) L. Buoninfante, G. Harmsen, S. Maheshwari, and A. Mazumdar, (2018b), arXiv:1804.09624 [gr-qc] .
- Giacchini and de Paula Netto (2019b) B. L. Giacchini and T. de Paula Netto, JCAP 1907, 013 (2019b), arXiv:1809.05907 [gr-qc] .
- Boos et al. (2018a) J. Boos, V. P. Frolov, and A. Zelnikov, Phys. Rev. D97, 084021 (2018a), arXiv:1802.09573 [gr-qc] .
- Boos (2020) J. Boos, (2020), arXiv:2003.13847 [gr-qc] .
- Kolar and Mazumdar (2020) I. Kolar and A. Mazumdar, (2020), arXiv:2004.07613 [gr-qc] .
- Conroy et al. (2015) A. Conroy, A. Mazumdar, and A. Teimouri, Phys. Rev. Lett. 114, 201101 (2015), arXiv:1503.05568 [hep-th] .
- Li et al. (2015) Y.-D. Li, L. Modesto, and L. Rachwał, JHEP 12, 173 (2015), arXiv:1506.08619 [hep-th] .
- Calcagni and Modesto (2017) G. Calcagni and L. Modesto, Phys. Lett. B773, 596 (2017), arXiv:1707.01119 [gr-qc] .
- Koshelev et al. (2018) A. S. Koshelev, J. Marto, and A. Mazumdar, Phys. Rev. D98, 064023 (2018), arXiv:1803.00309 [gr-qc] .
- Boos et al. (2019a) J. Boos, V. P. Frolov, and A. Zelnikov, Phys. Rev. D 100, 104008 (2019a), arXiv:1909.01494 [hep-th] .
- Biswas et al. (2010b) T. Biswas, T. Koivisto, and A. Mazumdar, JCAP 1011, 008 (2010b), arXiv:1005.0590 [hep-th] .
- Calcagni et al. (2014) G. Calcagni, L. Modesto, and P. Nicolini, Eur. Phys. J. C74, 2999 (2014), arXiv:1306.5332 [gr-qc] .
- Boos et al. (2018b) J. Boos, V. P. Frolov, and A. Zelnikov, Phys. Lett. B782, 688 (2018b), arXiv:1805.01875 [hep-th] .
- Buoninfante et al. (2019a) L. Buoninfante, A. Mazumdar, and J. Peng, Phys. Rev. D 100, 104059 (2019a), arXiv:1906.03624 [gr-qc] .
- Shapiro (2015) I. L. Shapiro, Phys. Lett. B744, 67 (2015), arXiv:1502.00106 [hep-th] .
- Frolov and Zelnikov (2016b) V. P. Frolov and A. Zelnikov, Phys. Rev. D93, 105048 (2016b), arXiv:1603.00826 [hep-th] .
- Modesto et al. (2018) L. Modesto, L. a. Rachwał, and I. L. Shapiro, Eur. Phys. J. C 78, 555 (2018), arXiv:1704.03988 [hep-th] .
- Asorey et al. (2018) M. Asorey, L. Rachwal, and I. L. Shapiro, Proceedings, 4th Workshop on Cosmology and the Quantum Vacuum: Segovia, Spain, September 4-8, 2017, Galaxies 6, 23 (2018), arXiv:1802.01036 [hep-th] .
- Calcagni et al. (2018) G. Calcagni, L. Modesto, and G. Nardelli, (2018), arXiv:1803.07848 [hep-th] .
- Buoninfante et al. (2019b) L. Buoninfante, G. Lambiase, and A. Mazumdar, Nucl. Phys. B 944, 114646 (2019b), arXiv:1805.03559 [hep-th] .
- Boos et al. (2019b) J. Boos, V. P. Frolov, and A. Zelnikov, Phys. Rev. D99, 076014 (2019b), arXiv:1901.07096 [hep-th] .
- Biswas et al. (2012) T. Biswas, E. Gerwick, T. Koivisto, and A. Mazumdar, Phys. Rev. Lett. 108, 031101 (2012), arXiv:1110.5249 [gr-qc] .
- Myers and Perry (1986) R. C. Myers and M. J. Perry, Annals Phys. 172, 304 (1986).
- Buoninfante et al. (2018c) L. Buoninfante, A. S. Cornell, G. Harmsen, A. S. Koshelev, G. Lambiase, J. Marto, and A. Mazumdar, Phys. Rev. D98, 084041 (2018c), arXiv:1807.08896 [gr-qc] .
- Shankar (1994) R. Shankar, Principles of Quantum Mechanics, 2nd ed. (Plenum Press, New York, USA, 1994).
- Bonnor (1969b) W. B. Bonnor, Int. J. Theor. Phys. 2, 373 (1969b).
- Olver et al. (2010) F. W. Olver, D. W. Lozier, R. F. Boisvert, and C. W. Clark, NIST Handbook of Mathematical Functions, 1st ed. (Cambridge University Press, New York, NY, USA, 2010).
- Kilicarslan (2019) E. Kilicarslan, Phys. Rev. D99, 124048 (2019), arXiv:1903.04283 [gr-qc] .
- Dengiz et al. (2020) S. Dengiz, E. Kilicarslan, I. Kolar, and A. Mazumdar, (2020), arXiv:2006.07650 [gr-qc] .