This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Dynamical Friction in Nonlocal Gravity

Mahmood Roshan1,2 [email protected]    Bahram Mashhoon2,3 [email protected] 1Department of Physics, Faculty of Science, Ferdowsi University of Mashhad, P.O. Box 1436, Mashhad, Iran
2School of Astronomy, Institute for Research in Fundamental Sciences (IPM), P. O. Box 19395-5531, Tehran, Iran
3Department of Physics and Astronomy, University of Missouri, Columbia, Missouri 65211, USA
Abstract

We study dynamical friction in the Newtonian regime of nonlocal gravity (NLG), which is a classical nonlocal generalization of Einstein’s theory of gravitation. The nonlocal aspect of NLG simulates dark matter. The attributes of the resulting effective dark matter are described and the main physical predictions of nonlocal gravity, which has a characteristic lengthscale of order 1 kpc, for galactic dynamics are presented. Within the framework of NLG, we derive the analogue of Chandrasekhar’s formula for dynamical friction. The astrophysical implications of the results for the apparent rotation of a central bar subject to dynamical friction in a barred spiral galaxy are briefly discussed.

Gravitation, Dynamical friction, Nonlocal gravity

I Introduction

Imagine the scattering state of a Newtonian two-body system. Under the Newtonian inverse-square law of gravity, the bodies deviate from their original paths during the scattering process, but the mechanical system is conservative and the overall positive energy is conserved. Let us choose one of the bodies as the reference and focus attention on its initial momentum vector. In the scattering process, the component of the final deflected momentum of the reference body along the direction of its initial momentum does not change to first order in the gravitational coupling constant GG, but decreases beyond the linear order. The net loss of momentum of the reference particle along its initial direction of motion is the source of the dynamical friction force. Normally, dynamical friction becomes physically significant when an astronomical body of mass MM moves relatively slowly through a population of stars of average mass mMm\ll M. It was first calculated in a collisionless gravitational system by Chandrasekhar Chandra . For the applications of dynamical friction in galactic dynamics, see Binney . The main purpose of this paper is to study dynamical friction within the framework of nonlocal gravity theory and derive the analogue of Chandrasekhar’s formula in the Newtonian regime of NLG. We now turn to a brief introduction of NLG and its Newtonian regime.

In the special and general theories of relativity, physics is local Einstein . For instance, the standard approach for extending relativity theory to an accelerated system in Minkowski spacetime involves the pointwise application of Lorentz transformations, which makes physical sense if the velocity of the accelerated system is in effect locally uniform during an elementary act of measurement. The locality assumption is certainly valid for pointlike coincidences of classical point particles and rays of radiation. However, wave phenomena are intrinsically nonlocal by the Huygens principle; moreover, such nonlocality extends to the measurement of the electromagnetic field. To determine the frequency content of an incident wave packet, for example, an accelerated observer must in general employ instantaneous Lorentz transformations over an extended period of time, during which the state of the observer varies continuously. At the same time, there are invariant local acceleration scales of length and time associated with the accelerated system. Therefore, accelerated systems are in general nonlocal and the past history of accelerated motion must be taken into account Mashhoon:1993zz , thus leading to nonlocal special relativity Mashhoon:2008vr . The deep connection between inertia and gravitation, as revealed in Einstein’s development of general relativity Einstein , suggests that gravity should be nonlocal as well. The most natural approach to a nonlocal gravity theory would be to introduce history dependence via a nonlocal constitutive relation in close analogy with the nonlocal electrodynamics of media Jackson ; L+L ; HeOb . Thus, nonlocal gravitational field equations would reflect an influence (“memory”) from the past that endures. History dependence could in fact be a natural feature of the universal gravitational interaction.

To implement this idea, the first step would involve expressing Einstein’s general relativity (GR) in a form that resembles Maxwell’s electrodynamics. This can be simply accomplished by the introduction of a preferred frame field in spacetime and employing an extended geometric framework known as teleparallelism. Indeed, there is a well-known teleparallel equivalent of general relativity (TEGR), which is the gauge theory of the group of spacetime translations Cho . Therefore, TEGR, though nonlinear, is formally analogous to electrodynamics and can be rendered nonlocal via history-dependent constitutive relations as in the nonlocal electrodynamics of media Hehl:2008eu ; Hehl:2009es . In the resulting theory of nonlocal gravity (NLG), the gravitational field is locally defined but satisfies partial integro-differential field equations. The only known exact solution of NLG is the trivial solution; that is, we simply recover Minkowski spacetime in the absence of gravity. Thus far, the nonlinearity of NLG has prevented finding exact solutions for strong-field regimes such as those involving black holes or cosmological models Bini:2016phe . However, linearized NLG and its Newtonian limit have been the subject of extensive investigations Blome:2010xn ; Rahvar:2014yta ; Mashhoon:2014jna ; Chicone:2015coa ; MaHe . A comprehensive account of NLG is contained in BMB .

When nonlocal TEGR is expressed as modified GR, the source of gravity turns out to include besides the standard symmetric energy-momentum tensor TμνT_{\mu\nu} of matter, certain purely nonlocal gravity terms which we interpret in terms of nonlocally induced effective dark matter. That is, the nonlocal aspect of gravity appears to simulate dark matter. What is now considered dark matter in astrophysics and cosmology may indeed be the manifestation of the nonlocal component of the gravitational interaction. This circumstance constitutes a significant observational consequence of NLG and has been explored in connection with the gravitational physics of nearby spiral galaxies and clusters of galaxies Rahvar:2014yta . Though much remains to be done, thus far NLG appears to be consistent with observational data regarding the solar system as well as nearby galaxies and clusters of galaxies. However, there are still many issues regarding the astrophysical and cosmological implications of nonlocal gravity that must be resolved Chicone:2015sda ; Chicone:2017oqt ; Roshan:2019xda ; Ghafourian:2020uae ; Roshan:2021mfc .

I.1 Newtonian Regime of NLG

In this paper, we are interested in the Newtonian regime of nonlocal gravity. In the Newtonian limit, the basic equations of nonlocal gravity simply reduce to the nonrelativistic gravitational force

𝐅(𝐱)=mΦ(𝐱)\mathbf{F}(\mathbf{x})=-m\nabla\Phi(\mathbf{x})\, (1)

on a test particle of inertial mass mm in the gravitational potential Φ(𝐱)\Phi(\mathbf{x}), which satisfies the nonlocal Poisson equation BMB

2Φ(𝐱)+χ(𝐱𝐲)2Φ(𝐲)d3y=4πGρ(𝐱),\nabla^{2}\Phi(\mathbf{x})+\int\chi(\mathbf{x}-\mathbf{y})\nabla^{2}\Phi(\mathbf{y})\,d^{3}y=4\pi G\,\rho(\mathbf{x})\,, (2)

where χ\chi is the constitutive kernel in the Newtonian regime. Here, χ\chi is assumed to be a universal function independent of the potential; moreover, χ\chi is a smooth function with certain reasonable mathematical properties which make it possible to write Fredholm integral Eq. (2) in the reciprocal form Chicone:2011me

4πGρ(𝐱)+q(𝐱𝐲)[4πGρ(𝐲)]d3y=2Φ(𝐱),4\pi G\,\rho(\mathbf{x})+\int q(\mathbf{x}-\mathbf{y})[4\pi G\,\rho(\mathbf{y})]\,d^{3}y=\nabla^{2}\Phi(\mathbf{x})\,, (3)

where qq is the reciprocal kernel. Equation (3) can be written as

2Φ=4πG(ρ+ρD),ρD(𝐱)=q(𝐱𝐲)ρ(𝐲)d3y,\nabla^{2}\Phi=4\pi G\,(\rho+\rho_{D})\,,\qquad\rho_{D}(\mathbf{x})=\int q(\mathbf{x}-\mathbf{y})\rho(\mathbf{y})\,d^{3}y\,, (4)

where ρD\rho_{D} is the density of effective dark matter and is given by the convolution of the reciprocal kernel qq with the density of matter ρ\rho. It follows from the convolution theorem for Fourier integral transforms that in Fourier space ρ^D(𝐤)=q^(𝐤)ρ^(𝐤)\hat{\rho}_{D}(\mathbf{k})=\hat{q}(\mathbf{k})\,\hat{\rho}(\mathbf{k}), where we define s^(𝝃)\hat{s}(\bm{\xi}) to be the Fourier integral transform of a suitable function s(𝐱)s(\mathbf{x}) such that

s^(𝝃)=s(𝐱)ei𝝃𝐱d3x,s(𝐱)=1(2π)3s^(𝝃)ei𝝃𝐱d3ξ.\hat{s}(\bm{\xi})=\int s(\mathbf{x})\,e^{-i\,\bm{\xi}\cdot\mathbf{x}}\,d^{3}x\,,\qquad s(\mathbf{x})=\frac{1}{(2\pi)^{3}}\,\int\hat{s}(\bm{\xi})\,e^{i\,\bm{\xi}\cdot\mathbf{x}}\,d^{3}\xi\,. (5)

We note that ρD=0\rho_{D}=0 if ρ=0\rho=0; hence, there is no effective dark matter in the complete absence of matter. Furthermore, for the sake of simplicity, we have suppressed the possible dependence of Φ\Phi, ρ\rho and ρD\rho_{D} upon time tt. There is no retardation in the Newtonian regime; hence, χ\chi and qq have no temporal dependence and gravitational memory is purely spatial in the Newtonian limit.

In NLG, we determine the reciprocal kernel qq on the basis of observational data. Two simple spherically symmetric functions have been considered in detail, namely,

q1(r)=14πλ01+μ0(a0+r)r(a0+r)eμ0rq_{1}(r)=\frac{1}{4\pi\lambda_{0}}\,\frac{1+\mu_{0}\,(a_{0}+r)}{r\,(a_{0}+r)}\,e^{-\mu_{0}\,r}\, (6)

and

q2(r)=14πλ01+μ0(a0+r)(a0+r)2eμ0r,q_{2}(r)=\frac{1}{4\pi\lambda_{0}}\,\frac{1+\mu_{0}\,(a_{0}+r)}{(a_{0}+r)^{2}}\,e^{-\mu_{0}\,r}\,, (7)

where r=|𝐱𝐲|r=|\mathbf{x}-\mathbf{y}|. Here, λ01\lambda_{0}\sim 1 kpc is the basic length scale of NLG, while a0a_{0} and μ0\mu_{0} moderate the short and long distance behaviors of qq, respectively. It turns out that in the Newtonian regime of NLG, we recover the phenomenological Tohline-Kuhn approach to modified gravity Toh ; Kuhn ; Bek . Indeed, kernels (6) and (7) are appropriate generalizations of the Kuhn kernel (4πλ0r2)1(4\pi\lambda_{0}\,r^{2})^{-1} within the framework of NLG BMB .

Let us note that for a0=0a_{0}=0, kernels q1q_{1} and q2q_{2} both reduce to q0q_{0} given by

q0(r)=14πλ01+μ0rr2eμ0r,q_{0}(r)=\frac{1}{4\pi\lambda_{0}}\,\frac{1+\mu_{0}\,r}{r^{2}}\,e^{-\mu_{0}\,r}\,, (8)

where for any finite radial coordinate rr, q0>q1>q2q_{0}>q_{1}>q_{2}, since a0>0a_{0}>0.

For a point mass mm located at the origin of spatial coordinates, ρ(𝐱)=mδ(𝐱)\rho(\mathbf{x})=m\,\delta(\mathbf{x}), the nonlocal Poisson Eq. (4) reduces to

2Φ(𝐱)=4πGm[δ(𝐱)+q(𝐱)],\nabla^{2}\Phi(\mathbf{x})=4\pi G\,m\,[\delta(\mathbf{x})+q(\mathbf{x})]\,, (9)

where mq(𝐱)mq(\mathbf{x}) is the density of dark matter associated with the point mass mm. Therefore, the net amount of effective dark matter associated with the point mass is given by mm times the integral of the reciprocal kernel qq over all space. In this connection, it is useful to define

i(r)=4π0rs2[q0(s)qi(s)]𝑑s,i=1,2,\mathcal{E}_{i}(r)=4\pi\int_{0}^{r}s^{2}[q_{0}(s)-q_{i}(s)]\,ds\,,\qquad i=1,2\,, (10)

where

4π0s2q0(s)𝑑s=α0,α0:=2λ0μ0.4\pi\int_{0}^{\infty}s^{2}\,q_{0}(s)\,ds=\alpha_{0}\,,\qquad\alpha_{0}:=\frac{2}{\lambda_{0}\,\mu_{0}}\,. (11)

It is straightforward to show that

1(r)=12α0ζ0eζ0[E1(ζ0)E1(ζ0+μ0r)],ζ0:=μ0a0,\mathcal{E}_{1}(r)=\frac{1}{2}\alpha_{0}\,\zeta_{0}\,e^{\zeta_{0}}[E_{1}(\zeta_{0})-E_{1}(\zeta_{0}+\mu_{0}r)]\,,\qquad\zeta_{0}:=\mu_{0}a_{0}\,, (12)
2(r)21(r)=12α0ζ0rr+a0eμ0r,\mathcal{E}_{2}(r)-2\,\mathcal{E}_{1}(r)=-\frac{1}{2}\alpha_{0}\,\zeta_{0}\,\frac{r}{r+a_{0}}e^{-\mu_{0}\,r}\,, (13)

where E1E_{1} is the exponential integral function A+S . It follows that 1(r)\mathcal{E}_{1}(r) and 2(r)\mathcal{E}_{2}(r) are positive monotonically increasing functions of rr that start from zero at r=0r=0 and for rr\to\infty approach 1()=12α0ζ0eζ0E1(ζ0)\mathcal{E}_{1}(\infty)=\tfrac{1}{2}\alpha_{0}\,\zeta_{0}\,e^{\zeta_{0}}E_{1}(\zeta_{0}) and 2()=21()\mathcal{E}_{2}(\infty)=2\,\mathcal{E}_{1}(\infty), respectively.

The net amount of effective dark matter associated with a point mass mm is thus

mD=mα0w,m_{D}=m\alpha_{0}\,w\,, (14)

where w=wi,i=1,2w=w_{i},i=1,2, depending upon whether the reciprocal kernel is chosen to be q1q_{1} or q2q_{2}, respectively. Here,

w1=112ζ0eζ0E1(ζ0),w2=1ζ0eζ0E1(ζ0)w_{1}=1-\frac{1}{2}\zeta_{0}\,e^{\zeta_{0}}E_{1}(\zeta_{0})\,,\qquad w_{2}=1-\zeta_{0}\,e^{\zeta_{0}}E_{1}(\zeta_{0})\, (15)

and i()=(1wi)α0\mathcal{E}_{i}(\infty)=(1-w_{i})\alpha_{0}. If we ignore the reciprocal kernel’s short distance parameter a0a_{0}, then ζ0=μ0a0\zeta_{0}=\mu_{0}a_{0} vanishes and w=1w=1. Moreover, w1(ζ0)w_{1}(\zeta_{0}) and w2(ζ0)w_{2}(\zeta_{0}) decrease somewhat from unity with increasing ζ0\zeta_{0}; for instance, for ζ0(0,0.1]\zeta_{0}\in(0,0.1], w1(1,0.9]w_{1}\in(1,0.9] and w2(1,0.8]w_{2}\in(1,0.8]; for the graphs of w1(ζ0)w_{1}(\zeta_{0}) and w2(ζ0)w_{2}(\zeta_{0}), see Fig. 7.2 of BMB . For a reasonable value of a0a_{0}, such as a few parsecs, ζ0104\zeta_{0}\sim 10^{-4} and ww is then very close to unity and we may neglect the contribution of (r)\mathcal{E}(r) to the effective dark matter.

I.2 NLG: Two-Body Force Law

According to the Newtonian regime of NLG, the force of gravity on the point mass mm due to point mass mm^{\prime} at position 𝐫\mathbf{r} is always attractive and is given by

𝐅NLG=Gmm𝐫r3(1+Δ),Δ=(r)+α0[1(1+12μ0r)eμ0r],\mathbf{F}_{\rm NLG}=\frac{Gmm^{\prime}\mathbf{r}}{r^{3}}(1+\Delta)\,,\qquad\Delta=-\mathcal{E}(r)+\alpha_{0}\,\left[1-(1+\tfrac{1}{2}\,\mu_{0}r)\,e^{-\mu_{0}r}\right]\,, (16)

where r=|𝐫|r=|\mathbf{r}| and Δ\Delta is the net contribution of the effective dark matter. This force is conservative and satisfies Newton’s third law of motion. With a0=0a_{0}=0, (r)=0\mathcal{E}(r)=0 and the resulting force law has been employed in the gravitational physics of galaxies and clusters of galaxies to determine parameters α0\alpha_{0} and μ0\mu_{0} Rahvar:2014yta ; indeed, observational data regarding the rotation curves of nearby spiral galaxies imply

α0=10.94±2.56,μ0=0.059±0.028kpc1,λ0=2α0μ03±2kpc.\alpha_{0}=10.94\pm 2.56\,,\quad\mu_{0}=0.059\pm 0.028~{}{\rm kpc}^{-1}\,,\quad\lambda_{0}=\frac{2}{\alpha_{0}\,\mu_{0}}\approx 3\pm 2~{}{\rm kpc}\,. (17)

Furthermore, observational data regarding the solar system indicate that a0>1014a_{0}>10^{14} cm Chicone:2015coa .

It proves useful to present a physical derivation of the formula for the gravitational force (16). In a system of NN point particles, the force on any particle mjm_{j} is given by Eq. (16), namely,

𝐅NLGj=GmjijNmi𝐫r3{1(r)+α0[1(1+12μ0r)eμ0r]},\mathbf{F}_{\rm NLG}^{j}=-Gm_{j}\sum_{i\neq j}^{N}\frac{m_{i}\mathbf{r}}{r^{3}}\left\{1-\mathcal{E}(r)+\alpha_{0}\,[1-(1+\tfrac{1}{2}\,\mu_{0}r)\,e^{-\mu_{0}r}]\right\}\,, (18)

where 𝐫=𝐫i𝐫j\mathbf{r}=\mathbf{r}_{i}-\mathbf{r}_{j}. This result consists of two terms: the Newtonian term due to mim_{i} plus the modification due to the fact that the point particle of mass mim_{i} is surrounded by a spherically symmetric distribution of effective dark matter of density miq(r)m_{i}\,q(r), where qq is the reciprocal kernel and its spherical symmetry is just a convenient simplifying assumption. However, this simplifying assumption has the important result that by Newton’s shell theorem the net attractive force of the dark matter associated with this spherical distribution points in the direction of mim_{i} and involves only that part of the spherical distribution of dark matter that is inside the radius r=|𝐫i𝐫j|r=|\mathbf{r}_{i}-\mathbf{r}_{j}|. To emphasize this point, the contributions of the Newtonian and effective dark matter parts can be written as

mir2+mi0r4πs2q(s)𝑑sr2,\frac{m_{i}}{r^{2}}+\frac{m_{i}\,\int_{0}^{r}4\,\pi\,s^{2}q(s)ds}{r^{2}}\,, (19)

where the rest of the spherical distribution of effective dark matter associated with mim_{i} does not contribute due to Newton’s shell theorem. This is illustrated in Figure 1.

Refer to caption
Figure 1: Schematic diagram illustrates the calculation of the gravitational force on mjm_{j} due to the effective dark mass associated with mim_{i} presented in Eq. (19). Dark matter is indicated here by the spherically symmetric dark region that surrounds mim_{i} and extends to infinity. One can think of the effective dark mass inside the sphere of radius rr as though it were all concentrated at its center at mim_{i} due to Newton’s shell theorem. Moreover, this theorem implies that the contribution of the effective dark mass outside the sphere of radius rr to the gravitational force on mjm_{j} vanishes.

From Eq. (10), we have

0r4πs2q(s)𝑑s=+0r4πs2q0(s)𝑑s,\int_{0}^{r}4\,\pi\,s^{2}q(s)ds=-\mathcal{E}+\int_{0}^{r}4\,\pi\,s^{2}q_{0}(s)ds\,, (20)

where, using Eq. (8), we find

0r4πs2q0(s)𝑑s=α0[1(1+12μ0r)eμ0r].\int_{0}^{r}4\,\pi\,s^{2}q_{0}(s)ds=\alpha_{0}\,\left[1-(1+\tfrac{1}{2}\,\mu_{0}r)\,e^{-\mu_{0}r}\right]\,. (21)

In this way, we recover Eq. (18) for 𝐅NLGj\mathbf{F}_{\rm NLG}^{j} by summing over mim_{i}, iji\neq j.

II Effective Dark Matter

It follows from Eq. (4) that in the Newtonian regime of NLG, Newtonian gravitation theory can be employed provided we take due account of the corresponding effective dark matter as well. Indeed, every point particle of mass mm is surrounded by a spherically symmetric distribution of effective dark matter of density mq(r)m\,q(r), where rr is the radial coordinate in a spherical polar coordinate system centered on the point particle mm. The net amount of effective dark matter is mD=mα0wm_{D}=m\,\alpha_{0}\,w, where α011\alpha_{0}\approx 11. If we choose q=q0q=q_{0}, then w=1w=1. The result is slightly less if we choose q1q_{1} or q2q_{2}, see Fig. 7.2 of BMB . In any case, because of spherical symmetry, the center of mass of the total effective dark matter associated with mm is exactly at the location of mm.

If we now have a Newtonian system of NN baryons, say, each of inertial mass mim_{i}, i=1,2,,Ni=1,2,\cdots,N, then each baryon brings with it the accompanying spherically symmetric effective dark matter of mass miα0wm_{i}\,\alpha_{0}\,w centered on the baryon. Since α0w\alpha_{0}\,w is a constant and the same for each baryon, the Newtonian center of mass coincides with the center of mass of the total effective dark matter. This will not be true, however, if we ask for the effective dark matter inside a galaxy, for example; then, the range of integration of the effective dark matter will be restricted by the boundary of the galaxy and instead of α0w\alpha_{0}\,w for each baryon, we will get a fraction of this quantity depending on the location of the baryon mim_{i} in the galactic system. Therefore, for a galaxy the baryonic center of mass is in general different from the center of mass of the corresponding effective dark matter.

For any isolated system of NN baryons, we have the total baryonic mass MBM_{B} and the corresponding total effective dark matter α0wMB\alpha_{0}\,wM_{B} over all space, where w<1w<1. However, the amount of dark matter, MDM_{D}, confined within the system depends on the shape and volume of the system such that

MD<α0wMB<α0MB,MB:=i=1Nmi.M_{D}<\alpha_{0}\,wM_{B}<\alpha_{0}\,M_{B}\,,\qquad M_{B}:=\sum_{i=1}^{N}\,m_{i}\,. (22)

The dynamic mass of the system is thus given by MB+MD=MB(1+ϕ)M_{B}+M_{D}=M_{B}(1+\phi), where ϕ\phi for the system is defined by

ϕ:=MDMB<α0w<α0.\phi:=\frac{M_{D}}{M_{B}}<\alpha_{0}\,w<\alpha_{0}\,. (23)

In NLG, we have the basic result that for any finite system the fraction ϕ\phi must be smaller than α011\alpha_{0}\approx 11. If the system is so huge that most of the baryonic matter is at distances larger than μ0117\mu_{0}^{-1}\approx 17 kpc, then we expect that the effective dark matter fraction approaches the limiting value of α0w\alpha_{0}\,w. Moreover, very far away from a baryonic system of mass MBM_{B}, it exhibits an effective gravitational mass that approaches MB(1+α0w)M_{B}\,(1+\alpha_{0}\,w). Therefore, for a cluster of galaxies, NLG predicts that we should have ϕα0w\phi\approx\alpha_{0}\,w. This appears to be consistent with observation of nearby clusters of galaxies Rahvar:2014yta . The standard definition of dark matter fraction in astrophysics is

fDM:=MDMB+MD=ϕ1+ϕ<ϕ.f_{DM}:=\frac{M_{D}}{M_{B}+M_{D}}=\frac{\phi}{1+\phi}<\phi\,. (24)

Let us note that fDMf_{DM} as a function of ϕ\phi increases monotonically from zero and approaches unity as ϕ\phi\to\infty. In NLG, ϕ<α0\phi<\alpha_{0}; hence, fDMf_{DM} is less than α0/(1+α0)\alpha_{0}/(1+\alpha_{0}). Therefore, for α011\alpha_{0}\approx 11, the effective dark matter ratio in NLG is such that fDM<0.917f_{DM}<0.917.

These considerations are naturally reflected in the gravitational force (16) on a point mass mm due to a point mass mm^{\prime} located at 𝐫\mathbf{r}. For r1/μ0r\gg 1/\mu_{0}, Eq. (16) reduces to

𝐅NLGGmm𝐫r3(1+α0w).\mathbf{F}_{\rm NLG}\approx\frac{Gmm^{\prime}\mathbf{r}}{r^{3}}(1+\alpha_{0}\,w)\,. (25)

That is, as the distance rr between mm and mm^{\prime} increases, the magnitude of 𝐅NLG\mathbf{F}_{\rm NLG} in Eq. (16) slowly increases from the Newtonian 1/r21/r^{2} and for r1μ017r\gg\frac{1}{\mu_{0}}\approx 17 kpc becomes 1/r21/r^{2} again, but with mm(1+α0w)m^{\prime}\to m^{\prime}\,(1+\alpha_{0}\,w). The physical interpretation of this result is that mm is now subject to the attractive Newtonian gravitational force of mm as well as the total spherically symmetric effective dark matter distribution associated with mm^{\prime}. For rr\to\infty, this net amount of dark matter becomes mα0wm^{\prime}\,\alpha_{0}\,w, as expected, and hence the effective gravitational mass of mm^{\prime} becomes m(1+α0w)m^{\prime}\,(1+\alpha_{0}w). As is evident in the force formula above, this situation is symmetric between mm and mm^{\prime} and hence our discussion could be repeated for the gravitational force experienced by mm^{\prime}.

It is interesting to mention an alternative explanation of the behavior of the two-body gravitational force that is important for the consideration of dynamical friction that is the focus of the present work. We can imagine that the strength of the gravitational coupling increases with distance in NLG; that is,

GG(1+Δ),Δ=1+0r4πs2q(s)𝑑s,G\to G(1+\Delta)\,,\qquad\Delta=1+\int_{0}^{r}4\,\pi\,s^{2}q(s)ds\,, (26)

so that the gravitational coupling “constant” ranges from GG to G(1+α0w)G(1+\alpha_{0}\,w) as rr goes from 0 to \infty.

Let us briefly digress here and mention the connection of these results with Fourier space. From

ρ^(𝐤)=ρ(𝐱)ei𝐤𝐱d3x,\hat{\rho}(\mathbf{k})=\int\rho(\mathbf{x})\,e^{-i\,\mathbf{k}\cdot\mathbf{x}}\,d^{3}x\,, (27)

we note that ρ^(0)=MB\hat{\rho}(0)=M_{B} and similarly ρ^D(0)\hat{\rho}_{D}(0) is the total mass of dark matter over all space, namely, α0wMB\alpha_{0}\,w\,M_{B}. The convolution theorem implies ρ^D(0)=q^(0)ρ^(0)\hat{\rho}_{D}(0)=\hat{q}(0)\,\hat{\rho}(0), which is consistent with the fact that q^(0)=α0w\hat{q}(0)=\alpha_{0}\,w. Moreover, q^0(0)>q^1(0)>q^2(0)\hat{q}_{0}(0)>\hat{q}_{1}(0)>\hat{q}_{2}(0), a relation that actually holds for any 𝐤\mathbf{k} BMB .

Finally, let us imagine an isolated astrophysical system (“galaxy”) consisting of NN particles and let DD be the baryonic diameter of the system, namely, the diameter of the smallest sphere that completely surrounds the baryonic system at the present time. For each baryon mim_{i} in the system, i=1,2,,Ni=1,2,\cdots,N, the corresponding effective dark matter that is within the confines of the system and contributes to the dynamic mass of the whole system is evidently less than

mi0D4πs2q(s)𝑑s.m_{i}\,\int_{0}^{D}4\,\pi\,s^{2}q(s)ds\,. (28)

This means that the total effective dark matter in the galaxy, MDM_{D}, is such that

MD<MB0D4πs2q(s)𝑑s.M_{D}<M_{B}\,\int_{0}^{D}4\,\pi\,s^{2}q(s)ds\,. (29)

On the other hand, q<q0q<q_{0} implies

0D4πs2q(s)𝑑s0D4πs2q0(s)𝑑s.\int_{0}^{D}4\,\pi\,s^{2}q(s)ds\leq\int_{0}^{D}4\,\pi\,s^{2}q_{0}(s)ds\,. (30)

Moreover, Eq. (21) can be written as

0D4πs2q0(s)𝑑s=α0H(D),\int_{0}^{D}4\,\pi\,s^{2}q_{0}(s)ds=\alpha_{0}\,H(D)\,, (31)

where

H(D):=1(1+12μ0D)eμ0DH(D):=1-(1+\frac{1}{2}\mu_{0}\,D)\,e^{-\mu_{0}\,D}\, (32)

is a function that starts from H(0)=0H(0)=0 at D=0D=0, monotonically increases with increasing DD and asymptotically approaches unity as DD\to\infty. It follows that in the Newtonian regime of NLG, we have

ϕ:=MDMB<α0H(D)=α0[1(1+12μ0D)eμ0D].\phi:=\frac{M_{D}}{M_{B}}<\alpha_{0}\,H(D)=\alpha_{0}\,\left[1-(1+\tfrac{1}{2}\mu_{0}\,D)\,e^{-\mu_{0}\,D}\right]\,. (33)

For systems with D1/μ017D\gg 1/\mu_{0}\approx 17\,kpc, such as for nearby clusters of galaxies, H(D)1H(D)\approx 1 and we find

ϕα0,\phi\lessapprox\alpha_{0}\,, (34)

in agreement with the relation ϕα0w\phi\approx\alpha_{0}\,w discussed earlier for such systems. On the other hand, for systems in the intermediate regime where DD is less than or comparable to 1/μ0171/\mu_{0}\approx 17\,kpc, we note that H(D)H(D) always stays below the line μ0D/2\mu_{0}\,D/2 for D>0D>0. Hence,

ϕ<Dλ0,\phi<\frac{D}{\lambda_{0}}\,, (35)

where λ03\lambda_{0}\approx 3 kpc. Thus for a globular star cluster, the effective amount of dark matter is at most a few percent of the baryonic content of the globular cluster. For a dwarf galaxy, the effective amount of dark matter could be at most comparable to its baryonic content. For giant galaxies and clusters of galaxies, ϕ\phi can at most approach α0w\alpha_{0}\,w. Further discussion of these issues is contained in BMB . For recent observational advances in connection with dwarf galaxies, see vanDokkum:2018vup ; Guo:2019wgb ; Pina:2019rer ; Hammer:2020qcd ; Shen:2021zka and the references cited therein; the prediction of NLG that there is less dark matter in dwarf galaxies seems to be consistent with these studies.

III Gravitational Drag

The purpose of this section is to discuss the gravitational two-body system in accordance with the Newtonian limit of NLG in order to derive the analogue of Chandrasekhar’s formula for dynamical friction. For the sake of simplicity and convenience, we follow closely the standard treatment presented in Binney . An astronomical body of mass MM moves without direct collision through a population of stars, each with a typical mass mm. The result of two-body interaction of MM with a star of mass mm will then be extended to include all of the stars.

III.1 Two-Body Problem in NLG

Imagine the “Newtonian” gravitational interaction between masses mm and MM according to NLG. The main equations are

md2𝐱mdt2=GmM𝐫r3[1+Δ(r)],Md2𝐱Mdt2=GmM𝐫r3[1+Δ(r)],m\,\frac{d^{2}\mathbf{x}_{m}}{dt^{2}}=\frac{GmM\mathbf{r}}{r^{3}}\,[1+\Delta(r)]\,,\qquad M\,\frac{d^{2}\mathbf{x}_{M}}{dt^{2}}=-\frac{GmM\mathbf{r}}{r^{3}}\,[1+\Delta(r)]\,, (36)

where Δ(r)\Delta(r) is given by Eq. (16) and

𝐫:=𝐱M𝐱m,r:=|𝐫|.\mathbf{r}:=\mathbf{x}_{M}-\mathbf{x}_{m}\,,\qquad r:=|\mathbf{r}|\,. (37)

As in the Newtonian Kepler problem, the center of mass moves uniformly and it is possible to separate the center of mass motion from the relative motion, so that

𝐫¨=G(m+M)𝐫r3[1+Δ(r)].\ddot{\mathbf{r}}=-\frac{G(m+M)\mathbf{r}}{r^{3}}\,[1+\Delta(r)]\,. (38)

In NLG, Δ\Delta is a universal function of three new gravitational constants: a0a_{0}, α0\alpha_{0} and μ0\mu_{0}. Moreover, Δ(r)\Delta(r) is a monotonically increasing function of rr; indeed, Δ(r)\Delta(r) starts from zero at r=0r=0 and eventually approaches α0w\alpha_{0}w as rr\to\infty. Equation (38) is rather complicated but possible to solve exactly; however, such a solution does not appear to be expressible in terms of the standard functions of mathematical physics and would hence be impractical for the problem under consideration in this paper. We therefore follow a different approach. Let us note that in the classical two-body problem within the framework of NLG, the Newtonian gravitational constant GG is in effect replaced by G(1+Δ)G(1+\Delta), indicating that the strength of the gravitational interaction increases by an order of magnitude in NLG as the relative distance goes to infinity. That is, for rμ01r\ll\mu_{0}^{-1} and rμ01r\gg\mu_{0}^{-1}, we have the inverse-square law of gravity, while the gravitational coupling “constant” ranges from GG to G(1+α0w)G(1+\alpha_{0}w). It seems natural to employ an approximation scheme involving a constant η\eta such that

G(1+Δ)G(1+η):=Gη,0<η<α0w.G(1+\Delta)\to G(1+\eta):=G_{\eta}\,,\qquad 0<\eta<\alpha_{0}w\,. (39)

The problem in NLG can thus be reduced in this approximation to the corresponding Newtonian problem; therefore, we can simply use the results of the standard treatment Binney , except that GGη=G(1+η)G\to G_{\eta}=G(1+\eta). We defer the determination of η\eta to the end of this section.

III.2 Dynamical Friction

Next, we follow the standard treatment of this problem given in Binney . In the background inertial reference frame, mass MM with state (𝐱M,𝐯M)(\mathbf{x}_{M},\mathbf{v}_{M}) interacts gravitationally with a star of mass mm with state (𝐱m,𝐯m)(\mathbf{x}_{m},\mathbf{v}_{m}). Let 𝐕:=𝐫˙\mathbf{V}:=\dot{\mathbf{r}} be the relative velocity. The center of mass of the two-body system moves uniformly; therefore, the net change in the scattering process in the velocity of MM is given by δ𝐯M=mδ𝐕/(m+M)\delta\mathbf{v}_{M}=m\delta\mathbf{V}/(m+M) and, similarly, δ𝐯m=Mδ𝐕/(m+M)\delta\mathbf{v}_{m}=-M\delta\mathbf{V}/(m+M). In the standard scattering picture in terms of relative Cartesian coordinates, 𝐕0\mathbf{V}_{0} is the initial relative velocity at t=t=-\infty. Hence, the magnitude of the relative specific orbital angular momentum is given by L=bV0L=b\,V_{0}, where bb is the impact parameter. As a consequence of its distant encounter with a star of mass mm, mass MM suffers a loss in its initial velocity along the original direction of motion due to the attractive drag of mm. We are interested in the net change in the velocity of MM, (δ𝐯M)||(\delta\mathbf{v}_{M})_{||}, along its initial direction of motion. The end result is

(δ𝐯M)||=2m𝐕0Gη2(m+M)Gη2(m+M)2+b2V04,(\delta\mathbf{v}_{M})_{||}=-2\,m\,\mathbf{V}_{0}\,\frac{G_{\eta}^{2}(m+M)}{G_{\eta}^{2}(m+M)^{2}+b^{2}V_{0}^{4}}\,, (40)

which is the same as given by Binney , except that we have replaced GG by Gη=G(1+η)G_{\eta}=G(1+\eta), where η\eta is a constant parameter such 0<η<α0w0<\eta<\alpha_{0}w. The magnitude of η\eta for a given system is estimated at the end of this section. The presence of η\eta means that the net effect is naturally stronger in NLG due to the effective dark matter of mm.

The rest of the analysis is exactly as in Ref. Binney and we find the Chandrasekhar dynamical friction formula

d𝐯Mdt=16π2Gη2m(m+M)lnΛ𝐯MvM30vMf(v)v2𝑑v,\frac{d\mathbf{v}_{M}}{dt}=-16\pi^{2}G_{\eta}^{2}\,m\,(m+M)\ln\Lambda\,\frac{\mathbf{v}_{M}}{v_{M}^{3}}\,\int_{0}^{v_{M}}f(v)v^{2}\,dv\,, (41)

where GG has again been replaced by GηG_{\eta} and f(v)f(v) is the isotropic velocity distribution function of the background stars. Here, lnΛ\ln\Lambda is the Coulomb logarithm given by

Λ=bmaxV02Gη(m+M)Dvtyp2Gη(m+M),\Lambda=\frac{b_{\rm max}V_{0}^{2}}{G_{\eta}(m+M)}\approx\frac{D\,v_{\text{typ}}^{2}}{G_{\eta}(m+M)}\,, (42)

where vtypv_{\text{typ}} is the typical velocity of the background particles. For MmM\gg m, the magnitude of dynamical friction force \mathcal{F} can be expressed as

NLG=4πGη2M2vM2ρ(<vM)lnΛ,\mathcal{F}_{\text{NLG}}=-\frac{4\pi\,G_{\eta}^{2}\,M^{2}}{v_{M}^{2}}\rho(<v_{M})\ln\Lambda\,, (43)

where ρ\rho is the mass density of the background stars and

ρ(<vM)=4πm0vMf(v)v2𝑑v.\rho(<v_{M})=4\pi\,m\int_{0}^{v_{M}}f(v)v^{2}\,dv\,. (44)

It is possible to write Λ=bmax/bmin\Lambda=b_{\rm max}/b_{\rm min}, where in the present approach bmin=Gη(m+M)/vtyp2b_{\rm min}=G_{\eta}(m+M)/v_{\text{typ}}^{2}. As revealed by simulations, Chandrasekhar’s formula is quite useful in astrophysical applications Binney ; nevertheless, it is an approximate result that takes into account only the two-body gravitational scatterings of MM with the stars of the background infinite homogeneous medium.

It is important to compare our result, namely, NLG\mathcal{F}_{\text{NLG}} with the standard Λ\LambdaCDM picture involving particles of dark matter. Let us first note that one can recover the Newtonian dynamical friction force, N\mathcal{F}_{\text{N}}, by simply setting η\eta equal to zero. On the other hand, it would then be necessary to take into account the contribution of the hypothetical dark matter particles. We recall that such particles are postulated to be nonexistent in NLG. Assuming that dark matter particles exist with mass density ρd\rho_{d}, N\mathcal{F}_{\text{N}} can be written as

N=4πG2M2vM2[ρ(<vM)lnΛ~+ρd(<vM)lnΛ~d].\mathcal{F}_{\text{N}}=-\frac{4\pi\,G^{2}\,M^{2}\,}{v_{M}^{2}}\Big{[}\rho(<v_{M})\ln\tilde{\Lambda}+\rho_{d}(<v_{M})\ln\tilde{\Lambda}_{d}\Big{]}\,. (45)

Here, lnΛ~\ln\tilde{\Lambda} and lnΛ~d\ln\tilde{\Lambda}_{d} with

Λ~=Dvtyp2G(m+M),Λ~d=Ddvtyp2G(md+M)\tilde{\Lambda}=\frac{D\,v_{\text{typ}}^{2}}{G(m+M)},\qquad\tilde{\Lambda}_{d}=\frac{D_{d}\,v_{\text{typ}}^{2}}{G(m_{d}+M)}\, (46)

are the Coulomb logarithms associated with the baryonic and dark matter particles, respectively, and DdD_{d} is the diameter of the smallest sphere that completely surrounds the corresponding dark matter particles of the system at the present epoch. Comparing Eq. (43) with Eq. (45), it turns out that depending on the physical properties of the system, dynamical friction can be stronger, equivalent or weaker in NLG as compared to the Newtonian theory that includes dark matter particles. It appears that |NLG/N||\mathcal{F}_{\text{NLG}}/\mathcal{F}_{\text{N}}| depends sensitively upon the ratio ρd/ρ\rho_{d}/\rho. For instance, when ρd<ρ\rho_{d}<\rho, it is possible that |NLG||\mathcal{F}_{\text{NLG}}| is greater than, or equal to, |N||\mathcal{F}_{\text{N}}|. In the latter case, dynamical friction cannot be employed to distinguish between nonlocal gravity and Newtonian theory with particle dark matter. In Section V, we consider a certain astrophysical system explicitly and compare the role of dynamical friction in NLG with the standard Lambda cold dark matter (Lambda-CDM) case.

III.3 Estimation of η\eta

To estimate η\eta, the simplest possibility would be to average the function Δ(r)\Delta(r) that appears in Eqs. (16) and (36) and represents the contribution of effective dark matter to the “Newtonian” gravitational force. Let us recall here that Δ\Delta as defined by Eq. (16) consists of two parts. The term involving \mathcal{E} vanishes in the absence of the short-distance parameter a0a_{0}. Indeed, for a reasonable value of a0a_{0} — for instance, for a0a_{0} equal to a few parsecs — the contribution of (r)\mathcal{E}(r) to Δ(r)\Delta(r) can be neglected. Then,

Δ(r)α0H(r),\Delta(r)\approx\alpha_{0}\,H(r)\,, (47)

where HH is defined by Eq. (32). From

H(r)𝑑r=α0r+12α0μ0(3+μ0r)eμ0r,\int H(r)\,dr=\alpha_{0}\,r+\frac{1}{2}\,\frac{\alpha_{0}}{\mu_{0}}\,(3+\mu_{0}\,r)\,e^{-\mu_{0}\,r}\,, (48)

we find that averaging Δ(r)\Delta(r) from r1r_{1} to r2r_{2} results in

ηα0(r2r1)1r1r2H(r)𝑑r=α0+α02μ0(r2r1)[(3+μ0r2)eμ0r2(3+μ0r1)eμ0r1].\eta\approx\alpha_{0}\,(r_{2}-r_{1})^{-1}\int_{r_{1}}^{r_{2}}H(r)\,dr=\alpha_{0}+\frac{\alpha_{0}}{2\mu_{0}(r_{2}-r_{1})}\Big{[}(3+\mu_{0}r_{2})e^{-\mu_{0}\,r_{2}}-(3+\mu_{0}r_{1})e^{-\mu_{0}\,r_{1}}\Big{]}\,. (49)

Here, r1bminr_{1}\approx b_{\rm min} could be considered to be the maximum of Gη(m+M)/vtyp2G_{\eta}(m+M)/v_{\text{typ}}^{2} and the size of the reference body MM, while r2bmaxr_{2}\approx b_{\rm max} could be taken to be the diameter of the baryonic system DD. That is, as before, DD is the diameter of the smallest sphere that completely surrounds the baryonic system (galaxy) at the present time. Let us recall that Λr2/r1\Lambda\approx r_{2}/r_{1} and assume for the sake of definiteness that

r2=D,r1=DΛ.r_{2}=D\,,\qquad r_{1}=\frac{D}{\Lambda}\,. (50)

Then, η\eta can be expressed as

ηα0+α0Λ2(Λ1)μ0D[(3+μ0D)eμ0D(3+μ0DΛ)eμ0DΛ].\eta\approx\alpha_{0}+\frac{\alpha_{0}\Lambda}{2(\Lambda-1)\mu_{0}D}\left[(3+\mu_{0}D)\,e^{-\mu_{0}D}-(3+\tfrac{\mu_{0}D}{\Lambda})\,e^{-\frac{\mu_{0}D}{\Lambda}}\right]\,. (51)

Normally, we have r2r1r_{2}\gg r_{1} (or equivalently Λ1\Lambda\gg 1) and μ0r11\mu_{0}\,r_{1}\ll 1. In fact, the limiting case of Λ\Lambda\to\infty exists and we find

η(Λ)α0(1321eμ0Dμ0D+12eμ0D).\eta(\Lambda\to\infty)\to\alpha_{0}\,\left(1-\frac{3}{2}\,\frac{1-e^{-\mu_{0}\,D}}{\mu_{0}\,D}+\frac{1}{2}\,e^{-\mu_{0}\,D}\right)\,. (52)

For μ0D(0,)\mu_{0}\,D\in(0,\infty), the function in the parentheses starts from zero with slope 1/41/4 and slowly but monotonically approaches unity as μ0D\mu_{0}\,D\to\infty.

To make a crude estimate for η\eta in a globular star cluster, a dwarf galaxy and a galaxy, let us take the following typical sizes D10D\approx 10\,pc, 1\approx 1\,kpc and 10\approx 10\,kpc, respectively. Using equation (51) with Λ100\Lambda\approx 100, η\eta for a globular star cluster is small, i.e. η103\eta\approx 10^{-3}, implying that effects of NLG do not appear to be significant in globular star clusters. On the other hand, for a dwarf galaxy and a normal galaxy we find η0.163\eta\approx 0.163 and η1.596\eta\approx 1.596, respectively.

IV Gravitational Wake

There is an alternative way to view Chandrasekhar’s dynamical friction formula. As the reference body of mass MM moves through an infinite homogeneous medium consisting of stars of average mass mm, the ensuing disturbance leads to a density enhancement in the reference body’s wake, thereby slowing it down via gravitational attraction. This approach has been developed by a number of authors; see TW and the references cited therein. To determine the role of NLG in this process, we follow the treatment presented in TW .

In the background medium, we ignore the gravitational interaction between the stars. Therefore, the stars in their unperturbed states move according to 𝐱=𝐱0+𝐯mt\mathbf{x}=\mathbf{x}_{0}+\mathbf{v}_{m}t with constant momenta 𝐩=m𝐯m\mathbf{p}=m\mathbf{v}_{m} starting from t=t=-\infty with 𝐱0\mathbf{x}_{0} ranging over all space. The motion of the reference body generates a density perturbation

ρM(t,𝐱)=Meβtδ(𝐱𝐯Mt),\rho_{M}(t,\mathbf{x})=Me^{\beta t}\delta(\mathbf{x}-\mathbf{v}_{M}t)\,, (53)

where β>0\beta>0 is a constant auxiliary parameter that we need in the course of our calculations; however, we will eventually assume that β0\beta\to 0. In the Newtonian regime of NLG, Poisson’s Eq. (4) takes the form

2Φ=4πGMeβt[δ(𝐱𝐯Mt)+q(𝐱𝐯Mt)],\nabla^{2}\Phi=4\pi GMe^{\beta t}\,[\delta(\mathbf{x}-\mathbf{v}_{M}t)+q(\mathbf{x}-\mathbf{v}_{M}t)]\,, (54)

where qq is the reciprocal kernel. Working in Fourier space, we find

Φ(t,𝐱)=GM2π21+q^(k)k2ei𝐤(𝐱𝐯Mt)+βtd3k,\Phi(t,\mathbf{x})=-\frac{GM}{2\pi^{2}}\int\frac{1+\hat{q}(k)}{k^{2}}\,e^{i\mathbf{k}\cdot(\mathbf{x}-\mathbf{v}_{M}t)+\beta t}\,d^{3}k\,, (55)

where we have used the fact that the reciprocal kernel is spherically symmetric by assumption and its Fourier integral transform q^(k)\hat{q}(k) is thus only a function of k=|𝐤|k=|\mathbf{k}|. Neglecting any interaction between the stars, the motion of a star is perturbed by the gravitational influence of the reference body alone. In general,

d𝐩dt=mΦ(𝐱).\frac{d\mathbf{p}}{dt}=-m\nabla\Phi(\mathbf{x})\,. (56)

We can express the resulting perturbation in 𝐱\mathbf{x} and 𝐩\mathbf{p} of a star in powers series in terms of the gravitational coupling constant GG. That is,

𝐱=𝐱0+𝐯mt+Δ1𝐱+Δ2𝐱+,𝐩=m𝐯m+Δ1𝐩+Δ2𝐩+,\mathbf{x}=\mathbf{x}_{0}+\mathbf{v}_{m}t+\Delta_{1}\mathbf{x}+\Delta_{2}\mathbf{x}+\cdots\,,\qquad\mathbf{p}=m\mathbf{v}_{m}+\Delta_{1}\mathbf{p}+\Delta_{2}\mathbf{p}+\cdots\,, (57)

where, for instance, Δn𝐩\Delta_{n}\mathbf{p} is the momentum perturbation of order nn. Substituting Eq. (57) in Eq. (56), the first-order perturbation in 𝐩\mathbf{p} can be obtained from Eq. (56) with unperturbed 𝐱=𝐱0+𝐯mt\mathbf{x}=\mathbf{x}_{0}+\mathbf{v}_{m}t. Therefore,

d(Δ1pj)dt=iGMm2π21+q^(k)k2kjei(ωt+𝐤𝐱0)d3k,\frac{d(\Delta_{1}p^{j})}{dt}=i\frac{GMm}{2\pi^{2}}\int\frac{1+\hat{q}(k)}{k^{2}}\,k^{j}\,e^{i(-\omega t+\mathbf{k}\cdot\mathbf{x}_{0})}\,d^{3}k\,, (58)

where

ω:=𝐤(𝐯M𝐯m)+iβ.\omega:=\mathbf{k}\cdot(\mathbf{v}_{M}-\mathbf{v}_{m})+i\beta\,. (59)

It is important to note that the net force of gravity on the stars vanishes to first order in GG. For instance, the jj-component of the net force can be calculated by integrating Eq. (58) over all the stars. To this end, let us multiply the right-hand side of Eq. (58) by Nd3x0N\,d^{3}x_{0}, where NN is the constant density of stars. The integration over the jj-component of 𝐱0\mathbf{x}_{0} results in ξδ(ξ)=0\xi\,\delta(\xi)=0 in the integrand, where ξ=kj\xi=k^{j}. Thus, the net rate of momentum transfer is zero to linear order in GG. We therefore proceed to the evaluation of this quantity to second order in GG.

The second-order perturbation in the momentum Δ2𝐩\Delta_{2}\mathbf{p} can be determined from

d(Δ2pj)dt=m2ΦxjxlΔ1xl,\frac{d(\Delta_{2}p_{j})}{dt}=-m\frac{\partial^{2}\Phi}{\partial x^{j}\partial x^{l}}\,\Delta_{1}x^{l}\,, (60)

where Kjl=2Φ/xjxlK_{jl}=\partial^{2}\Phi/\partial x^{j}\partial x^{l} is the tidal matrix evaluated along the unperturbed path of a star. Moreover, the first-order change in the position of a star can be calculated by integrating the force Eq. (58) twice over time starting from t=t=-\infty. The result is

Δ1xj=iGM2π21+q^(k)k2ω2kjei(ωt+𝐤𝐱0)d3k.\Delta_{1}x^{j}=-i\frac{GM}{2\pi^{2}}\int\frac{1+\hat{q}(k)}{k^{2}\omega^{2}}\,k^{j}\,e^{i(-\omega t+\mathbf{k}\cdot\mathbf{x}_{0})}\,d^{3}k\,. (61)

To evaluate this second-order rate of momentum transfer, let us first note that

2Φxjxl=GM2π21+q^(k)k2kjklei(ωt+𝐤𝐱0)d3k,\frac{\partial^{2}\Phi}{\partial x^{j}\partial x^{l}}=\frac{GM}{2\pi^{2}}\int\frac{1+\hat{q}(k^{\prime})}{k^{\prime 2}}\,k^{\prime}_{j}k^{\prime}_{l}\,e^{i(-\omega^{\prime}t+\mathbf{k^{\prime}}\cdot\mathbf{x}_{0})}\,d^{3}k^{\prime}\,, (62)

where ω:=𝐤(𝐯M𝐯m)+iβ\omega^{\prime}:=\mathbf{k^{\prime}}\cdot(\mathbf{v}_{M}-\mathbf{v}_{m})+i\beta. It proves convenient at this point to replace 𝐤\mathbf{k^{\prime}} by 𝐤-\mathbf{k^{\prime}} and hence ω\omega^{\prime} by ω-\omega^{{}^{\prime}*} in Eq. (62); then, combining Eqs. (61) and (62) we find

d(Δ2pj)dt=im(GM2π2)2[1+q^(k)][1+q^(k)]k2k2ω2kjklklei(ωω)t+i(𝐤𝐤)𝐱0d3kd3k.\frac{d(\Delta_{2}p^{j})}{dt}=i\,m\left(\frac{GM}{2\pi^{2}}\right)^{2}\int\frac{[1+\hat{q}(k)][1+\hat{q}(k^{\prime})]}{k^{2}k^{\prime 2}\omega^{2}}\,k^{\prime j}k^{\prime}_{l}k^{l}\,e^{i(\omega^{{}^{\prime}*}-\omega)t+i(\mathbf{k}-\mathbf{k^{\prime}})\cdot\mathbf{x}_{0}}\,d^{3}k\,d^{3}k^{\prime}\,. (63)

To find the net rate of momentum transfer to second order, we must sum over all the stars; therefore, let us integrate Eq. (63) over Nd3x0N\,d^{3}x_{0}. The result is

Σstarsd(Δ2pj)dt=2iπG2M2mNe2βt[1+q^(k)]2k2ω2kjd3k,\Sigma_{\rm stars}\frac{d(\Delta_{2}p^{j})}{dt}=\frac{2i}{\pi}G^{2}M^{2}mNe^{2\beta t}\int\frac{[1+\hat{q}(k)]^{2}}{k^{2}\omega^{2}}\,k^{j}\,d^{3}k\,, (64)

where we have used the fact that i(ωω)=2βi(\omega^{{}^{\prime}*}-\omega)=2\beta when 𝐤=𝐤\mathbf{k^{\prime}}=\mathbf{k}. In Appendix A, we prove a useful identity, namely,

i[1+q^(k)]2k2ω2kjd3k=βvmj[1+q^(k)]2k2|ω|2d3k.i\int\frac{[1+\hat{q}(k)]^{2}}{k^{2}\omega^{2}}\,k_{j}\,d^{3}k=\beta\frac{\partial}{\partial v_{m}^{j}}\int\frac{[1+\hat{q}(k)]^{2}}{k^{2}|\omega|^{2}}\,d^{3}k\,. (65)

Substituting this relation in Eq. (64) and taking advantage of the following representation of Dirac’s delta function:

δ(x)=limβ0+1πβ|x+iβ|2,\delta(x)=\lim_{\beta\to 0^{+}}\frac{1}{\pi}\frac{\beta}{|x+i\beta|^{2}}\,, (66)

we find by letting β0+\beta\to 0^{+} that exp(2βt)1\exp(2\beta t)\to 1 and

Σstarsd(Δ2pj)dt=2G2M2mNvmj[1+q^(k)]2k2δ[𝐤(𝐯M𝐯m)]d3k.\Sigma_{\rm stars}\frac{d(\Delta_{2}p_{j})}{dt}=2G^{2}M^{2}mN\frac{\partial}{\partial v_{m}^{j}}\int\frac{[1+\hat{q}(k)]^{2}}{k^{2}}\,\delta[\mathbf{k}\cdot(\mathbf{v}_{M}-\mathbf{v}_{m})]\,d^{3}k\,. (67)

Thus far, we have been working with Cartesian coordinates in Fourier space. Let us now introduce spherical polar coordinates and write

k1=ksinϑcosφ,k2=ksinϑsinφ,k3=kcosϑ.k_{1}=k\sin\vartheta\cos\varphi\,,\qquad k_{2}=k\sin\vartheta\sin\varphi\,,\qquad k_{3}=k\cos\vartheta\,. (68)

Without any loss in generality, we can choose the Cartesian axes in Fourier space such that 𝐯M𝐯m\mathbf{v}_{M}-\mathbf{v}_{m} is a vector in the polar direction; then, integrating over angles in Eq. (67), we get

Σstarsd(Δ2pj)dt=4πG2M2mNvmj1|𝐯M𝐯m|[1+q^(k)]2dkk.\Sigma_{\rm stars}\frac{d(\Delta_{2}p_{j})}{dt}=4\pi G^{2}M^{2}mN\frac{\partial}{\partial v_{m}^{j}}\frac{1}{|\mathbf{v}_{M}-\mathbf{v}_{m}|}\int[1+\hat{q}(k)]^{2}\frac{dk}{k}\,. (69)

It follows from the results of Appendix A and Eq. (72) below that this net force is in the direction of 𝐯M\mathbf{v}_{M}, once we take due account of the isotropic velocity distribution of the background stars.

We note that the mass density of stars is given by

ρ=mN=mf(vm)d3vm,\rho=mN=m\int f(v_{m})d^{3}v_{m}\,, (70)

where f(vm)f(v_{m}) is the isotropic velocity distribution function of the stars. Therefore, in Eq. (69), the part that depends on the velocity of the stars should be replaced by

mNvmj1|𝐯M𝐯m|mf(vm)(𝐯M𝐯m)j|𝐯M𝐯m|3d3vm,mN\frac{\partial}{\partial v_{m}^{j}}\frac{1}{|\mathbf{v}_{M}-\mathbf{v}_{m}|}\mapsto m\int f(v_{m})\frac{(\mathbf{v}_{M}-\mathbf{v}_{m})_{j}}{|\mathbf{v}_{M}-\mathbf{v}_{m}|^{3}}\,d^{3}v_{m}\,, (71)

since Eq. (69) involves the sum over all stars and that includes their velocity distribution. Using Newton’s shell theorem in the present context, we find

f(vm)(𝐯M𝐯m)|𝐯M𝐯m|3d3vm=4π𝐯MvM30vMf(vm)vm2𝑑vm.\int f(v_{m})\frac{(\mathbf{v}_{M}-\mathbf{v}_{m})}{|\mathbf{v}_{M}-\mathbf{v}_{m}|^{3}}\,d^{3}v_{m}=4\pi\frac{\mathbf{v}_{M}}{v_{M}^{3}}\int_{0}^{v_{M}}f(v_{m})v_{m}^{2}\,dv_{m}\,. (72)

Finally, Eq. (69) can be written as

Σstarsd(Δ2pj)dt=16π2G2M2mvMjvM30vMf(vm)vm2𝑑vm[1+q^(k)]2dkk,\Sigma_{\rm stars}\frac{d(\Delta_{2}p^{j})}{dt}=16\pi^{2}G^{2}M^{2}m\frac{v_{M}^{j}}{v_{M}^{3}}\int_{0}^{v_{M}}f(v_{m})v_{m}^{2}\,dv_{m}\int[1+\hat{q}(k)]^{2}\frac{dk}{k}\,, (73)

where

ρ(<vM)=4πm0vMf(vm)vm2𝑑vm.\rho(<v_{M})=4\pi m\int_{0}^{v_{M}}f(v_{m})v_{m}^{2}\,dv_{m}\,. (74)

The total rate of momentum transfer to the stars (73) should be equal and opposite to the dynamical friction force \mathcal{F} experienced by the reference body; hence,

NLG=Md𝐯Mdt=4πG2M2𝐯MvM3ρ(<vM)[1+q^(k)]2dkk,\mathcal{F}_{\rm NLG}=M\frac{d\mathbf{v}_{M}}{dt}=-4\pi G^{2}M^{2}\frac{\mathbf{v}_{M}}{v_{M}^{3}}\rho(<v_{M})\int[1+\hat{q}(k)]^{2}\frac{dk}{k}\,, (75)

which reduces to Chandrasekhar’s result in the absence of the reciprocal kernel. That is, the Coulomb logarithm is given by

kminkmaxdkk=ln(kmaxkmin),\int_{k_{\rm min}}^{k_{\rm max}}\frac{dk}{k}=\ln\left(\frac{k_{\rm max}}{k_{\rm min}}\right)\,, (76)

where 1/kmax1/k_{\rm max} has to do with the distance of closest approach (``bmin")(``b_{\rm min}"), while 1/kmin1/k_{\rm min} has to do with the extent of the background medium (``bmax")(``b_{\rm max}") which we assumed to be equal to DD.

To connect with our approach in the previous section, we can define a constant parameter η^\hat{\eta} via

kminkmax[1+q^(k)]2dkk=(1+η^)2kminkmaxdkk.\int_{k_{\rm min}}^{k_{\rm max}}[1+\hat{q}(k)]^{2}\frac{dk}{k}=(1+\hat{\eta})^{2}\int_{k_{\rm min}}^{k_{\rm max}}\frac{dk}{k}\,. (77)

Then, Eq. (75) takes the form of Chandrasekhar’s formula with GG replaced by Gη^=G(1+η^)G_{\hat{\eta}}=G(1+\hat{\eta}), where η^\hat{\eta} depends upon the boundaries of the domain of integration kmink_{\rm min} and kmaxk_{\rm max}; that is,

NLG=Md𝐯Mdt=4πGη^2M2𝐯MvM3ρ(<vM)lnΛ,\mathcal{F}_{\rm NLG}=M\frac{d\mathbf{v}_{M}}{dt}=-4\pi G_{\hat{\eta}}^{2}M^{2}\frac{\mathbf{v}_{M}}{v_{M}^{3}}\rho(<v_{M})\ln\Lambda\,, (78)

where Λ=bmax/bmin\Lambda=b_{\rm max}/b_{\rm min}.

IV.1 Estimation of η^\hat{\eta}

The reciprocal kernel depends upon three parameters, namely, a0a_{0}, λ0\lambda_{0} and μ0\mu_{0}. To estimate η^\hat{\eta}, we neglect the short-distance parameter a0a_{0} for the sake of simplicity. In this case, qq reduces to q0q_{0} given by Eq. (8). Therefore,

kminkmax[1+q^0(k)]2dkk(1+η^)2kminkmaxdkk,\int_{k_{\rm min}}^{k_{\rm max}}[1+\hat{q}_{0}(k)]^{2}\frac{dk}{k}\approx(1+\hat{\eta})^{2}\int_{k_{\rm min}}^{k_{\rm max}}\frac{dk}{k}\,, (79)

where q^0(k)\hat{q}_{0}(k) is the Fourier integral transform of q0q_{0} and is given by

q^0(k)=α02(11+k2/μ02+μ0karctankμ0)>0.\hat{q}_{0}(k)=\frac{\alpha_{0}}{2}\Big{(}\frac{1}{1+k^{2}/\mu_{0}^{2}}+\frac{\mu_{0}}{k}\arctan\frac{k}{\mu_{0}}\Big{)}>0\,. (80)

Notice that q^0\hat{q}_{0} goes from α0\alpha_{0} to 0 when kk goes from 0 to \infty; therefore, we expect from Eq. (79) that 0<η^<α00<\hat{\eta}<\alpha_{0}. It is straightforward to evaluate the integral containing q^0\hat{q}_{0} in Eq. (79) in order to find an approximate analytic expression for η^\hat{\eta}. Let us define

:=[1+q^0(k)]2dkk=[1+α02(11+u2+arctanuu)]2duu,\mathcal{I}:=\int[1+\hat{q}_{0}(k)]^{2}\frac{dk}{k}=\int\Big{[}1+\frac{\alpha_{0}}{2}\Big{(}\frac{1}{1+u^{2}}+\frac{\arctan u}{u}\Big{)}\Big{]}^{2}\frac{du}{u}\,, (81)

where u=k/μ0u=k/\mu_{0}. This integral can be evaluated explicitly and the result is

(u)=\displaystyle\mathcal{I}(u)={} α0281u2+1+(α0+1)2lnu12α0(α0+2)ln(u2+1)\displaystyle\frac{\alpha_{0}^{2}}{8}\frac{1}{u^{2}+1}+(\alpha_{0}+1)^{2}\ln u-\frac{1}{2}\alpha_{0}(\alpha_{0}+2)\ln(u^{2}+1) (82)
α04(3α0+4)arctanuuα028(3u2+1)arctan2uu2.\displaystyle-\frac{\alpha_{0}}{4}(3\alpha_{0}+4)\frac{\arctan u}{u}-\frac{\alpha_{0}^{2}}{8}(3u^{2}+1)\frac{\arctan^{2}u}{u^{2}}\,.

We recall that kmin=1/bmax1/Dk_{\rm min}=1/b_{\rm max}\approx 1/D and kmax=1/bminΛ/Dk_{\rm max}=1/b_{\rm min}\approx\Lambda/D. With u=k/μ0u=k/\mu_{0},

umin1μ0D,umaxΛμ0D,u_{\rm min}\approx\frac{1}{\mu_{0}D}\,,\qquad u_{\rm max}\approx\frac{\Lambda}{\mu_{0}D}\,, (83)

as before, we can evaluate η^\hat{\eta} from

(umax)(umin)(1+η^)2ln(umaxumin).\mathcal{I}(u_{\rm max})-\mathcal{I}(u_{\rm min})\approx(1+\hat{\eta})^{2}\ln\left(\frac{u_{\rm max}}{u_{\rm min}}\right)\,. (84)

It is interesting to note that η^(Λ)0\hat{\eta}(\Lambda\to\infty)\to 0, which should be compared and contrasted with the corresponding Eq. (52) for η\eta.

To estimate η^\hat{\eta}, which is in some sense the Fourier analogue of η\eta discussed at the end of Section III, we use Λ100\Lambda\approx 100, as before, and the same typical values for DD used to estimate η\eta, namely, 10\approx 10 pc, 1\approx 1\,kpc and 10\approx 10\,kpc for a globular cluster, a dwarf galaxy and a galaxy, and the corresponding estimates for η^\hat{\eta} turn out to be 103\approx 10^{-3}, 0.116\approx 0.116 and 1.409\approx 1.409, respectively. It is clear that our estimates for η^\hat{\eta} are in fairly good agreement with those for η\eta.

V Dynamical Friction in Barred Spiral Galaxies

In the standard Λ\LambdaCDM model, dynamical friction plays a central role in the secular evolution of spiral galaxies. In particular, a stellar bar in a spiral galaxy transfers angular momentum between the inner and outer parts of the galactic disk. More importantly, the stellar bar exchanges angular momentum with the dark matter halo due to dynamical friction on the stellar bar in its interaction with the particles of dark matter. In comparison, the corresponding dynamical friction on the stellar bar due to its interaction with the baryonic content in the galactic disk is negligibly small. One may simply regard the first term in Eq. (45) to be negligible in comparison with the second term. That is, the wake induced by the motion of the bar in the normal disk particles is much weaker than the corresponding wake induced in the particles of dark matter.

The angular momentum exchange via dynamical friction means that the pattern speed of the bar decreases with time Debattista:2000ey . This circumstance implies that barred spiral galaxies in Λ\LambdaCDM model should host relatively slow bars, as indeed occurs in almost all of the state-of-the-art cosmological hydrodynamic simulations (Roshan:2021liy, ). On the other hand, most of the observed bars are relatively fast, which has posed a challenge for the standard Λ\LambdaCDM model. One possibility involves postulating ultralight axions as the particles of dark matter Hui:2016ltb . One of the advantages of such fuzzy dark matter particles is that dynamical friction can be substantially suppressed in spiral galaxies. However, a recent study of such models claims that these axions are in gross disagreement with Lyman-alpha forest observations at 99.7% confidence level Rogers:2020ltq ; Rogers:2020cup .

In the context of NLG, there is no dark matter halo. Depending on the value of η^\hat{\eta}, we can quantify the impact of dynamical friction on the stellar bar due to the baryonic content of the disk. As already mentioned, this makes a negligible contribution to the total dynamical friction in the standard Λ\LambdaCDM picture. It is necessary to mention that our result for dynamical friction, i.e. Eqs. (43) and (78), have been obtained for a moving point mass and not a stellar bar. However, modeling a bar with a dumbbell consisting of two point masses captures some important features of a more detailed description MDW ; Binney . To estimate η^\hat{\eta} for the apparent rotation of a stellar bar parallel to the disk of a spiral galaxy, it is appropriate to take bmaxhb_{\rm max}\approx h, where hh is the thickness of the disk. The typical value for hh is 2\approx 2\,kpc. In this case, we find η^0.244\hat{\eta}\approx 0.244 using Eq. (84). Therefore, the contribution of the baryonic matter to the dynamical friction should still be essentially negligible even though it is stronger in NLG by a factor of (1+η^)21.55(1+\hat{\eta})^{2}\approx 1.55. That is, it is rather unlikely that this 55% increase in dynamical friction by baryonic matter in NLG can mimic in magnitude the strong contribution of the dark matter particles in the standard Λ\LambdaCDM picture. The numerical galactic simulations in NLG confirm this description; that is, for all of the maximal disks simulated in Roshan:2019xda ; Roshan:2021mfc , dynamical friction is essentially negligible in the NLG disks, while a substantial reduction in the bar speed appears in the standard Λ\LambdaCDM dark matter models.

VI Discussion

Chandrasekhar’s dynamical friction appears as an important aspect of astrophysical systems helping to distinguish between particle dark matter and modified gravity. We have therefore studied dynamical friction in the Newtonian regime of nonlocal gravity (NLG). Within the context of NLG, we have followed the standard approaches to dynamical friction and demonstrated that nonlocal gravity effects appear effectively as an enhancement in the strength of Newton’s gravitational constant GG. Specifically, we have shown that Chandrasekhar’s expression holds approximately in NLG as well and the only essential difference is that GG is replaced by G(1+η^)G\,(1+\hat{\eta}), where η^\hat{\eta} is a constant such that 0<η^<α0110<\hat{\eta}<\alpha_{0}\approx 11. This means that in systems where there is no actual dark matter component, NLG leads to stronger dynamical friction. On the other hand, in systems hosting a substantial amount of hypothetical dark matter, depending on the magnitude of η^\hat{\eta} and the other physical properties, dynamical friction can be stronger or weaker than in NLG. As an instance of possible astrophysical implications of our results, we have argued that, in general agreement with numerical simulations Roshan:2019xda ; Roshan:2021mfc , dynamical friction on stellar bars in barred galaxies is much weaker in NLG compared to the standard Λ\LambdaCDM picture.

Acknowledgments

The work of M.R. has been supported by the Ferdowsi University of Mashhad.

Appendix A Proof of Identity (65)

Employing spherical polar coordinates in Fourier space introduced in Eq. (68), we can write the left-hand side of identity (65) as

𝕀j=i[1+q^(k)]2k2ω2kjd3k=i[1+q^(k)]2(kUcosϑ+iβ)2kjsinϑdkdϑdφ,\mathbb{I}_{j}=i\int\frac{[1+\hat{q}(k)]^{2}}{k^{2}\omega^{2}}\,k_{j}\,d^{3}k=i\int\frac{[1+\hat{q}(k)]^{2}}{(kU\cos\vartheta+i\beta)^{2}}\,k_{j}\sin\vartheta\,dkd\vartheta d\varphi\,, (85)

where U:=|𝐯M𝐯m|U:=|\mathbf{v}_{M}-\mathbf{v}_{m}| and 𝐯M𝐯m\mathbf{v}_{M}-\mathbf{v}_{m} has been assumed to be a vector in the polar direction with no loss in generality. Integrating over the azimuthal angle φ\varphi, we find 𝕀1=𝕀2=0\mathbb{I}_{1}=\mathbb{I}_{2}=0 for j=1j=1 and j=2j=2. For j=3j=3, we have

𝕀3=2πi[1+q^(k)]2(k2U2ζ2+β2)2(kUζiβ)2kζ𝑑k𝑑ζ,\mathbb{I}_{3}=2\pi i\int\frac{[1+\hat{q}(k)]^{2}}{(k^{2}U^{2}\zeta^{2}+\beta^{2})^{2}}(kU\zeta-i\beta)^{2}k\zeta\,dkd\zeta\,, (86)

where ζ=cosϑ(1,1)\zeta=\cos\vartheta\in(-1,1). Thus, only even terms in ζ\zeta contribute to the integrand for 𝕀3\mathbb{I}_{3} and we have

𝕀3=4πβU[1+q^(k)]2(k2U2ζ2+β2)2k2ζ2𝑑k𝑑ζ.\mathbb{I}_{3}=4\pi\beta U\int\frac{[1+\hat{q}(k)]^{2}}{(k^{2}U^{2}\zeta^{2}+\beta^{2})^{2}}k^{2}\zeta^{2}\,dkd\zeta\,. (87)

Let us consider next the right-hand side of identity (65),

βvmj[1+q^(k)]2k2|ω|2d3k\beta\frac{\partial}{\partial v_{m}^{j}}\int\frac{[1+\hat{q}(k)]^{2}}{k^{2}|\omega|^{2}}\,d^{3}k\, (88)

and note that

vmj1|ω|2=vmj1[𝐤(𝐯M𝐯m)]2+β2=2kUζkj(k2U2ζ2+β2)2.\frac{\partial}{\partial v_{m}^{j}}\,\frac{1}{|\omega|^{2}}=\frac{\partial}{\partial v_{m}^{j}}\,\frac{1}{[\mathbf{k}\cdot(\mathbf{v}_{M}-\mathbf{v}_{m})]^{2}+\beta^{2}}=2kU\zeta\frac{k_{j}}{(k^{2}U^{2}\zeta^{2}+\beta^{2})^{2}}\,. (89)

Substituting Eq. (89) in Eq. (88), we find

2βU[1+q^(k)]2(k2U2ζ2+β2)2kjkζ𝑑k𝑑ζ𝑑φ,2\beta U\int\frac{[1+\hat{q}(k)]^{2}}{(k^{2}U^{2}\zeta^{2}+\beta^{2})^{2}}k_{j}k\zeta\,dkd\zeta d\varphi\,, (90)

which vanishes for j=1j=1 and j=2j=2, while for j=3j=3 integrates over the azimuthal angle φ\varphi to the same result as 𝕀3\mathbb{I}_{3}. This completes the proof of identity (65).

References

  • (1) S. Chandrasekhar, Astrophys. J. 97, 255-262 (1943).
  • (2) J. Binney and S. Tremaine, Galactic Dynamics (Princeton University Press, Princeton, NJ, USA, 1987).
  • (3) A. Einstein, The Meaning of Relativity (Princeton University Press, Princeton, NJ, USA, 1955).
  • (4) B. Mashhoon, “Nonlocal theory of accelerated observers”, Phys. Rev. A 47, 4498-4501 (1993).
  • (5) B. Mashhoon, “Nonlocal Special Relativity”, Ann. Phys. (Berlin) 17, 705-727 (2008). [arXiv:0805.2926 [gr-qc]]
  • (6) J. D. Jackson, Classical Electrodynamics, 3rd edn (Wiley, Hoboken, NJ, 1999).
  • (7) L. D. Landau and E. M. Lifshitz, Electrodynamics of Continuous Media (Pergamon, Oxford, UK, 1960).
  • (8) F. W. Hehl, and Y. N. Obukhov, Foundations of Classical Electrodynamics: Charge, Flux, and Metric (Birkhäuser, Boston, MA, USA, 2003).
  • (9) Y. M. Cho, “Einstein Lagrangian as the translational Yang-Mills Lagrangian”, Phys. Rev. D 14, 2521-2525 (1976).
  • (10) F. W. Hehl and B. Mashhoon, “Nonlocal Gravity Simulates Dark Matter”, Phys. Lett. B 673, 279-282 (2009). [arXiv:0812.1059 [gr-qc]]
  • (11) F. W. Hehl and B. Mashhoon, “A Formal framework for a nonlocal generalization of Einstein’s theory of gravitation”, Phys. Rev. D 79, 064028 (2009). [arXiv:0902.0560 [gr-qc]]
  • (12) D. Bini and B. Mashhoon, “Nonlocal gravity: Conformally flat spacetimes”, Int. J. Geom. Meth. Mod. Phys. 13, no.06, 1650081 (2016). [arXiv:1603.09477 [gr-qc]]
  • (13) H. J. Blome, C. Chicone, F. W. Hehl and B. Mashhoon, “Nonlocal Modification of Newtonian Gravity”, Phys. Rev. D 81, 065020 (2010). [arXiv:1002.1425 [gr-qc]]
  • (14) S. Rahvar and B. Mashhoon, “Observational Tests of Nonlocal Gravity: Galaxy Rotation Curves and Clusters of Galaxies”, Phys. Rev. D 89, 104011 (2014). [arXiv:1401.4819 [astro-ph.GA]]
  • (15) B. Mashhoon, “Nonlocal Gravity: The General Linear Approximation”, Phys. Rev. D 90, 124031 (2014). [arXiv:1409.4472 [gr-qc]]
  • (16) C. Chicone and B. Mashhoon, “Nonlocal Gravity in the Solar System”, Classical Quantum Gravity 33, no.7, 075005 (2016). [arXiv:1508.01508 [gr-qc]]
  • (17) B. Mashhoon and F. W. Hehl, “Nonlocal Gravitomagnetism”, Universe 5, 195 (2019). [arXiv:1908.05431 [gr-qc]]
  • (18) B. Mashhoon, Nonlocal Gravity (Oxford University Press, Oxford, UK, 2017).
  • (19) C. Chicone and B. Mashhoon, “Nonlocal Newtonian Cosmology”, J. Math. Phys. 57, no.7, 072501 (2016). [arXiv:1510.07316 [gr-qc]]
  • (20) C. Chicone and B. Mashhoon, “Cosmological Implications of Nonlocal Gravity”, [arXiv:1703.04205 [gr-qc]].
  • (21) M. Roshan and S. Rahvar, “Evolution of spiral galaxies in nonlocal gravity”, Astrophys. J. 872, no.1, 6 (2019). [arXiv:1901.01087 [gr-qc]]
  • (22) N. Ghafourian, M. Roshan and S. Abbassi, “Does modified gravity predict fast stellar bars in spiral galaxies?”, Astrophys. J. 895, no.1, 13 (2020). [arXiv:2004.10104 [astro-ph.GA]]
  • (23) M. Roshan, I. Banik, N. Ghafourian, I. Thies, B. Famaey, E. Asencio and P. Kroupa, “Barred spiral galaxies in modified gravity theories”, Mon. Not. Roy. Astron. Soc. 503, 2833 (2021). [arXiv:2103.01794 [astro-ph.GA]]
  • (24) C. Chicone and B. Mashhoon, “Nonlocal Gravity: Modified Poisson’s Equation”, J. Math. Phys. 53, 042501 (2012). [arXiv:1111.4702 [gr-qc]]
  • (25) J. E. Tohline, “Does gravity exhibit a 1/r1/r force on the scale of galaxies?”, Ann. N.Y. Acad. Sci. 422, 390 (1984).
  • (26) J. R. Kuhn and L. Kruglyak, “Non-Newtonian forces and the invisible mass problem”, Astrophys. J. 313, 1-12 (1987).
  • (27) J. D. Bekenstein, “The missing light puzzle: a hint about gravity?”, in Second Canadian Conference on General Relativity and Relativistic Astrophysics, A. Coley, C. Dyer, and T. Tupper, eds. (World Scientific, Singapore, 1988), pp. 68-104.
  • (28) M. Abramowitz and I. A. Stegun, Handbook of Mathematical Functions (National Bureau of Standards, Washington, D.C., 1964).
  • (29) P. van Dokkum, S. Danieli, Y. Cohen, A. Merritt, A. J. Romanowsky, R. Abraham, J. Brodie, C. Conroy, D. Lokhorst and L. Mowla, et al. “A galaxy lacking dark matter”, Nature 555, no.7698, 629-632 (2018). [arXiv:1803.10237 [astro-ph.GA]]
  • (30) Q. Guo, H. Hu, Z. Zheng, S. Liao, W. Du, S. Mao, L. Jiang, J. Wang, Y. Peng and L. Gao, et al. “Further evidence for a population of dark-matter-deficient dwarf galaxies”, Nature Astron. 4, no.3, 246-251 (2019). [arXiv:1908.00046 [astro-ph.GA]]
  • (31) P. E. Mancera Piña, F. Fraternali, E. A. K. Adams, A. Marasco, T. Oosterloo, K. A. Oman, L. Leisman, E. M. di Teodoro, L. Posti and M. Battipaglia, et al. “Off the Baryonic Tully–Fisher Relation: A Population of Baryon-dominated Ultra-diffuse Galaxies”, Astrophys. J. Lett. 883, no.2, L33 (2019). [arXiv:1909.01363 [astro-ph.GA]]
  • (32) F. Hammer, Y. Yang, F. Arenou, J. Wang, H. Li, P. Bonifacio and C. Babusiaux, “Orbital evidences for dark-matter-free Milky Way dwarf spheroidal galaxies”, Astrophys. J. 892, no.1, 3 (2020). [arXiv:2002.09493 [astro-ph.GA]]
  • (33) Z. Shen, S. Danieli, P. van Dokkum, R. Abraham, J. P. Brodie, C. Conroy, A. E. Dolphin, A. J. Romanowsky, J. M. Diederik Kruijssen and D. Dutta Chowdhury, “A Tip of the Red Giant Branch Distance of 22.1 ±\pm 1.2 Mpc to the Dark Matter Deficient Galaxy NGC 1052–DF2 from 40 Orbits of Hubble Space Telescope Imaging”, Astrophys. J. Lett. 914, no.1, L12 (2021). [arXiv:2104.03319 [astro-ph.GA]]
  • (34) S. Tremaine and M. D. Weinberg, “Dynamical friction in spherical systems”, Mon. Not. Roy. Astron. Soc. 209, 729-757 (1984).
  • (35) V. P. Debattista and J. A. Sellwood, “Constraints from dynamical friction on the dark matter content of barred galaxies”, Astrophys. J. 543, 704-721 (2000). [arXiv:astro-ph/0006275 [astro-ph]]
  • (36) M. Roshan, N. Ghafourian, T. Kashfi, I. Banik, M. Haslbauer, V. Cuomo, B. Famaey and P. Kroupa, “Fast galaxy bars continue to challenge standard cosmology”, Mon. Not. Roy. Astron. Soc. 508, 926-939 (2021). [arXiv:2106.10304 [astro-ph.GA]]
  • (37) L. Hui, J. P. Ostriker, S. Tremaine and E. Witten, “Ultralight scalars as cosmological dark matter”, Phys. Rev. D 95, no.4, 043541 (2017). [arXiv:1610.08297 [astro-ph.CO]]
  • (38) K. K. Rogers and H. V. Peiris, “Strong Bound on Canonical Ultralight Axion Dark Matter from the Lyman-Alpha Forest”, Phys. Rev. Lett. 126, no.7, 071302 (2021). [arXiv:2007.12705 [astro-ph.CO]]
  • (39) K. K. Rogers and H. V. Peiris, “General framework for cosmological dark matter bounds using NN-body simulations”, Phys. Rev. D 103, no.4, 043526 (2021). [arXiv:2007.13751 [astro-ph.CO]]
  • (40) M. D. Weinberg, Mon. Not. Roy. Astron. Soc. 213, 451-471 (1985).