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

Quantum interference in the Kerr spacetime

Zhongyou Mo [email protected]    Leonardo Modesto [email protected] Department of Physics, Southern University of Science and Technology, Shenzhen, 518055, China
(today)
Abstract

The gravitational induced interference is here studied in the framework of Teleparallel Gravity. We derive the gravitational phase difference and we apply the result to the case of a Kerr spacetime. Afterwards, we compute the fringe shifts in an interference experiment of particles and discuss how to increase their values by changing the given parameters that include: the area in between the paths, the energy of the particles, the distance from the black hole, the mass and the spin of the black hole. It turns out that it is more difficult to detect the fringe shifts for massless particles than for massive particles. As a further application, we show how the mass of the black hole and its angular momentum can be obtained from the measurement of the fringe shifts. Finally, we compare the phase difference derived in Teleparallel Gravity with a previous work in General Relativity.

1 Introduction

In the year 1959, Aharonov and Bohm proposed an observable effect due to electromagnetic potentials in the quantum domain [1]. They showed that, contrary to the conclusions of classical mechanics, in quantum mechanics there are effects of electromagnetic potentials on charged particles, even in the region where all the fields vanish. In their model, two electron beams go through two cylindrical tubes within two different time-dependent potentials, to finally interfere in a region outside the tubes. In particular, they proved that the interference depends on the time integrals of the potentials. The same two authors proposed also another experiment that we summarize as follows. In the region outside an infinite cylindrical solenoid (in which a magnetic field is confined), an electron beam is split in two, one travels to the right while the other to the left of the solenoid and then they interfere. It turns out that the interference of the two beams depends on the contour integral of the vector potential. These thought experiments prove that even in regions where the fields are absent, the electromagnetic potential can affect the observations.

To clarity, we write the wave function in the presence of the potential as [1]

Ψ=Ψ10ei𝒮1+Ψ20ei𝒮2,\Psi={\Psi_{1}}^{0}e^{-\frac{i}{\hbar}\mathscr{S}_{1}}+{\Psi_{2}}^{0}e^{-\frac{i}{\hbar}\mathscr{S}_{2}}, (1)

where Ψ10{\Psi_{1}}^{0} and Ψ20{\Psi_{2}}^{0} denote the free wave functions. It turns out that the interference depends on the difference between of the two phase factors in (1). In general, the phase difference is given by [2]

1(𝒮1𝒮2)=ecAμdxμ,-\frac{1}{\hbar}(\mathscr{S}_{1}-\mathscr{S}_{2})=\frac{e}{\hbar c}\oint A_{\mu}\mbox{d}x^{\mu}, (2)

where the closed integral is unshrinkable. In the second thought experiment mentioned above, the right hand side of (2) is proportional to the magnetic flux through the cross section of the solenoid. The Aharonov-Bohm effect (AB effect) caused by a magnetic field was experimentally observed by Chambers [3]. Since then, more observations for the AB effect were performed (see Ref. [4] for a review of them).

As a route to connect general relativity with quantum mechanics, it is appealing to image a phase induced by the gravitational field in analogy with the one by the electromagnetic field. The effect of the gravity induced phase, analogous to the AB effect in electromagnetism, is usually referred to as Gravitational Aharonov-Bohm effect [5, 6, 7]. In gravity, the interference of particles moving in a flat spacetime region may be affected by a non vanishing Riemann tensor localized far from the particles. In Ref. [8], Stodolsky argued that such phase is given by

mcds\frac{mc}{\hbar}\int\mbox{d}s (3)

for a massive particle (in the case of a semiclassical limit in which particles travel along the classical path). An interesting feature of this expression is its property under coordinate transformations. As Stodolsky showed, the above phase is gauge invariant under coordinate transformations, as opposite to the gauge variance of the electromagnetic phase under U(1)U(1) transformations of the potential. This discovery reveals the difference between the symmetry properties of the gravitational and the electromagnetic field in the quantum domain.

Concerning our work, we will evaluate the phase in the theory of Teleparallel Gravity (TG). This theory is also known as the Teleparallel Equivalent of General Relativity [9]. In TG, the phase (mc/)ds(mc/\hbar)\int\mbox{d}s can be separated into three parts [9]: the first part represents the free particle, while the second part stands for the inertial effects of the frame, which can be eliminated by choosing an inertial frame, the third part is the one we really have to take care of. Indeed, it represents the gravitational interaction given by the integral of a gauge potential for gravity. Our study is based on this formulation.

Before getting to the heart of our contribution, it deserves to be mentioned the experimental work on the gravitational phase. In 1974, Overhauser and Colella proposed an experiment to detect the gravitational quantum interference [10]. In their proposal, a neutrons’ beam is split into two parts and recombined afterwords. The trajectories of the neutrons approximately form a vertical parallelogram with its base parallel to the surface of the earth. They found that the phase difference between the two beams is related to the gravitational acceleration. In the next year, Colella et al. implemented such idea experimentally [11]. They rotated the interferometer to change the angle between the parallelogram and the surface of the earth, and detected the corresponding counting rates of the interfering beams. With these results they determined the number of the fringes caused by the gravity. Although the influence of the gravitational field of the earth has been found, the gravitational interference caused by small masses is still a difficult task. On this subject, Hohensee et al. proposed an experiment in which matter waves are in a gravitational potential of a pair of masses with vanishing net gravitational force [12]. This thought experiment has not been realized because it requires the optical lattice to be perfect (see the comment in Ref. [13]). Recently the gravitational interference caused by small masses has been detected by Overstreet et al. experimentally [14] 111In Ref. [14] the authors claim they have observed the gravitational Aharonov-Bohm effect. Such result is extremely interesting, but we should notice that the observed effect is not exactly the one in Refs. [5, 6, 7] because the atoms move in a region where the Riemann curvature does not vanish., using laser pulses to split and recombine two atoms vertically at different times. The upper atom goes closer to a ring mass than the lower atom, which leads to a gravity induced phase difference between these atoms.

Now that the gravitational quantum interference has been observed in laboratories, it is essentially to explore more about its theoretical aspects, especially the applications in astronomy. As mentioned above, in TG we can separate the phase (mc/)ds(mc/\hbar)\int\mbox{d}s into three parts with the third term standing for the gravitational interaction. This term called gravitational phase is exactly given by

mcuaBaμdxμ,\frac{mc}{\hbar}\int u_{a}{B^{a}}_{\mu}\mbox{d}x^{\mu}, (4)

where Baμ{B^{a}}_{\mu} is a gauge potential associated to gravity [9]. As Aldrovandi et al. showed [15], in the weak field limit this term gives the same result as the one in the experiment [11] for the interference of neutrons on the earth. This coincidence inspired us to apply this expression and its generalization to other scenes, especially the gravitational quantum interference in the Kerr spacetime, to give a prediction for future observations.

The structure of this paper is arranged as it follows. In Sec. 2, we make a brief introduction to the concept of tetrad in TG. In Sec. 3, we present a method to calculate the gravitational phase. Its integral expression is derived in the inertial frames and applied to the Kerr spacetime. Therefore, we use this expression for an interference experiment in Sec. 4. Finally in Sec. 5, we summarize the results and present the potential extensions.

Throughout this article, we use the units c=G=1c=G=1 and the metric signature (+,,,)(+,-,-,-), unless we explicitly specify.

2 A brief introduction to Teleparallel Gravity

2.1 Tetrad in Teleparallel Gravity

All the formulas in this section are taken from the book [9], which gives a full introduction to TG. We will not show all the details of this theory, but only introduce the core concepts relevant to our study. Let us start with the tetrad, namely

ha=haμμ,ha=haμdxμ,h_{a}={h_{a}}^{\mu}\partial_{\mu},\qquad h^{a}={h^{a}}_{\mu}\mbox{d}x^{\mu}, (5)

a basis which connects the spacetime metric gμνg_{\mu\nu} to the Minkowski’s metric in the tangent space,

ηab=diag(1,1,1,1).\eta_{ab}=\text{diag}(1,-1,-1,-1). (6)

At each point:

gμν=ηabhaμhbν,ηab=gμνhaμhbν,g_{\mu\nu}=\eta_{ab}{h^{a}}_{\mu}{h^{b}}_{\nu},\qquad\eta_{ab}=g_{\mu\nu}{h_{a}}^{\mu}{h_{b}}^{\nu}, (7)

where the Greek letters are used to denote the coordinates in spacetime, while the Latin letters denote the coordinates in the tangent-space. The components of the tetrad satisfy the equations:

haμhaν=δμν,haμhbμ=δba.{h^{a}}_{\mu}{h_{a}}^{\nu}=\delta^{\nu}_{\mu},\qquad{h^{a}}_{\mu}{h_{b}}^{\mu}=\delta^{a}_{b}. (8)

Finally, the tetrad relates the spacetime tensors with the tangent-space tensors:

Vμ=haμVa,Va=haμVμ.V^{\mu}={h_{a}}^{\mu}V^{a},\qquad V_{a}={h^{a}}_{\mu}V^{\mu}. (9)

The components of the tetrad in the presence of gravity are given by:

haμ=μxa+A˙xbabμ+Baμ,{h^{a}}_{\mu}=\partial_{\mu}x^{a}+\dot{A}{{}^{a}}_{b\mu}x^{b}+{B^{a}}_{\mu}, (10)

where A˙=abμΛad(x)μΛbd(x)\dot{A}{{}^{a}}_{b\mu}={\Lambda^{a}}_{d}(x)\partial_{\mu}{\Lambda_{b}}^{d}(x) is the Lorentz connection with Λad(x){\Lambda^{a}}_{d}(x) a local Lorentz transformation from an inertial reference frame to a general frame, and Baμ{B^{a}}_{\mu} is a gauge potential corresponding to a translational transformation δxa(x)=εa(x)\delta x^{a}(x)=\varepsilon^{a}(x) on the tangent space. In TG, gravity is generated from the group of the latter transformations under which the tetrad haμ{h^{a}}_{\mu} is invariant, while the potential Baμ{B^{a}}_{\mu} transforms according to

δB=aμμεaA˙εbabμ.\delta B{{}^{a}}_{\mu}=-\partial_{\mu}\varepsilon^{a}-\dot{A}{{}^{a}}_{b\mu}\varepsilon^{b}. (11)

In Eq. (10) we see that the expression of the tetrad contains three terms. The first one corresponds to a coordinates’ transformation from the spacetime to its tangent-space. As shown in [9], the second one corresponds to the inertia. And the last one corresponds to the gravitational interaction. The expression of the tetrad is obtained by combining (10) with (5), namely

ha=dxa+A˙xbabμdxμ+Bdaμxμ.{h^{a}}=\mbox{d}x^{a}+\dot{A}{{}^{a}}_{b\mu}x^{b}\mbox{d}x^{\mu}+B{{}^{a}}_{\mu}\mbox{d}x^{\mu}. (12)

Opposite to general relativity, in TG, the curvature vanishes while the torsion is non-vanishing, namely

R˙abμν\displaystyle\dot{R}^{a}{{}_{b\mu\nu}} =\displaystyle= μA˙abννA˙a+bμA˙aA˙ccμbνA˙aA˙ccν=bμ0,\displaystyle\partial_{\mu}\dot{A}^{a}{{}_{b\nu}}-\partial_{\nu}\dot{A}^{a}{{}_{b\mu}}+\dot{A}^{a}{{}_{c\mu}}\dot{A}^{c}{{}_{b\nu}}-\dot{A}^{a}{{}_{c\nu}}\dot{A}^{c}{{}_{b\mu}}=0, (13)
T˙aμν\displaystyle\dot{T}^{a}{{}_{\mu\nu}} =\displaystyle= μhaννha+μA˙ahccμνA˙ahccν=μ𝒟˙μBaν𝒟˙νBaμ0,\displaystyle\partial_{\mu}h^{a}{{}_{\nu}}-\partial_{\nu}h^{a}{{}_{\mu}}+\dot{A}^{a}{{}_{c\mu}}h^{c}{{}_{\nu}}-\dot{A}^{a}{{}_{c\nu}}h^{c}{{}_{\mu}}=\dot{\mathscr{D}}_{\mu}{B^{a}}_{\nu}-\dot{\mathscr{D}}_{\nu}{B^{a}}_{\mu}\neq 0, (14)

where the derivative operator 𝒟˙μ\dot{\mathscr{D}}_{\mu} only acts on the indices in the tangent space and it is defined by:

𝒟˙μϕa=μϕa+A˙aϕbbμ.\dot{\mathscr{D}}_{\mu}\phi^{a}=\partial_{\mu}\phi^{a}+\dot{A}^{a}{{}_{b\mu}}\phi^{b}. (15)

The curvature and the torsion can also be expressed in terms of spacetime indices, i.e.

R˙ρλνμ\displaystyle\dot{R}^{\rho}{{}_{\lambda\nu\mu}} =\displaystyle= νΓ˙ρλμμΓ˙ρ+λνΓ˙ρΓ˙ηηνλμΓ˙ρΓ˙ηημ,λν\displaystyle\partial_{\nu}\dot{\Gamma}^{\rho}{{}_{\lambda\mu}}-\partial_{\mu}\dot{\Gamma}^{\rho}{{}_{\lambda\nu}}+\dot{\Gamma}^{\rho}{{}_{\eta\nu}}\dot{\Gamma}^{\eta}{{}_{\lambda\mu}}-\dot{\Gamma}^{\rho}{{}_{\eta\mu}}\dot{\Gamma}^{\eta}{{}_{\lambda\nu}}, (16)
T˙ρνμ\displaystyle\dot{T}^{\rho}{{}_{\nu\mu}} =\displaystyle= Γ˙ρμνΓ˙ρ,νμ\displaystyle\dot{\Gamma}^{\rho}{{}_{\mu\nu}}-\dot{\Gamma}^{\rho}{{}_{\nu\mu}}, (17)

where Γ˙μρν\dot{\Gamma}^{\mu}{{}_{\rho\nu}} is the Weitzenböck connection defined by:

Γ˙μ=ρνhaμ𝒟˙νhaρ.\dot{\Gamma}^{\mu}{{}_{\rho\nu}}={h_{a}}^{\mu}\dot{\mathscr{D}}_{\nu}{h^{a}}_{\rho}. (18)

In TG, the torsion is regarded as a field strength, and from (14) we see that BaμB^{a}{{}_{\mu}} plays a role analogous to the gauge potential in electromagnetism. The torsion is gauge invariant [9] because it can be written in the following form,

T˙a=μν𝒟˙μhaν𝒟˙νhaμ,\dot{T}^{a}{{}_{\mu\nu}}=\dot{\mathscr{D}}_{\mu}{h^{a}}_{\nu}-\dot{\mathscr{D}}_{\nu}{h^{a}}_{\mu}, (19)

while the tetrad is invariant under the gauge transformation (11). The action for gravity is constructed by means of the torsion tensor, which coincides with the Einstein-Hilbert action, and the field equation in TG is equivalent to the Einstein equation (all the details can be found in the book [9]).

2.2 The role of the gauge potential

As mentioned above, the gravitational phase is given by (4) where the gauge potential BaμB^{a}{{}_{\mu}} appears in the integrand. This is reasonable because gravity is represented by the gauge potential, as stated in Ref. [9]. This potential not only appears in the field equation, but also plays an important role in the equation of motion, which is equivalent to the geodesic equation, of a particle in the gravitational field. We now prove the latter claim and finally show that BaμB^{a}{{}_{\mu}} appears in the gravitational phase by an analogy with electromagnetism.

Let us remind the geodesic equation in general relativity, namely

duμds+Γμuρρνuν=0.\frac{\mbox{d}u^{\mu}}{\mbox{d}s}+\Gamma^{\mu}{{}_{\rho\nu}}u^{\rho}u^{\nu}=0. (20)

In TG, the Levi-Civita connection can be written as [9]

Γμ=ρνΓ˙μρνK˙μ,ρν\Gamma^{\mu}{{}_{\rho\nu}}=\dot{\Gamma}^{\mu}{{}_{\rho\nu}}-\dot{K}^{\mu}{{}_{\rho\nu}}, (21)

where Γ˙μρν\dot{\Gamma}^{\mu}{{}_{\rho\nu}} is defined in (18), and K˙μρν\dot{K}^{\mu}{{}_{\rho\nu}} is the contortion

K˙μ=ρν12(T˙ν+μρT˙ρμνT˙μ)ρν,\dot{K}^{\mu}{{}_{\rho\nu}}=\frac{1}{2}(\dot{T}_{\nu}{{}^{\mu}}_{\rho}+\dot{T}_{\rho}{{}^{\mu}}_{\nu}-\dot{T}^{\mu}{{}_{\rho\nu}}), (22)

of the Weitzenböck torsion

T˙μ=ρνhaμT˙a=ρνhaμ(𝒟˙ρBaν𝒟˙νBaρ),\dot{T}^{\mu}{{}_{\rho\nu}}={h_{a}}^{\mu}\dot{T}^{a}{{}_{\rho\nu}}={h_{a}}^{\mu}(\dot{\mathscr{D}}_{\rho}{B^{a}}_{\nu}-\dot{\mathscr{D}}_{\nu}{B^{a}}_{\rho}), (23)

where (14) has been used. Recalling (10), the tetrad depends on Baμ{B^{a}}_{\mu}. Therefore, both Γ˙μρν\dot{\Gamma}^{\mu}{{}_{\rho\nu}} in (18) and K˙μρν\dot{K}^{\mu}{{}_{\rho\nu}} in (22) depend on the gauge potential.

According to the above expressions, we can prove that the geodesic equation (20) depends on the potential Baμ{B^{a}}_{\mu}. Indeed, we can rewrite the geodesic equation in TG using (21),

duμds+(Γ˙μρνK˙μ)ρνuρuν=0.\frac{\mbox{d}u^{\mu}}{\mbox{d}s}+(\dot{\Gamma}^{\mu}{{}_{\rho\nu}}-\dot{K}^{\mu}{{}_{\rho\nu}})u^{\rho}u^{\nu}=0. (24)

For the last term, according to (22), we get

K˙μuρρνuν=12(T˙ν+μρT˙ρμνT˙μ)ρνuρuν=T˙ρuρμνuν,\dot{K}^{\mu}{{}_{\rho\nu}}u^{\rho}u^{\nu}=\frac{1}{2}(\dot{T}_{\nu}{{}^{\mu}}_{\rho}+\dot{T}_{\rho}{{}^{\mu}}_{\nu}-\dot{T}^{\mu}{{}_{\rho\nu}})u^{\rho}u^{\nu}=\dot{T}_{\rho}{{}^{\mu}}_{\nu}u^{\rho}u^{\nu}, (25)

where the last step follows from the anti-symmetry of T˙μρν\dot{T}^{\mu}{{}_{\rho\nu}} in the last two indices (see (23)), and by re-labeling the indices of the first term. Furthermore, we rewrite (25) as:

K˙μuρρνuν\displaystyle\dot{K}^{\mu}{{}_{\rho\nu}}u^{\rho}u^{\nu} =\displaystyle= gραgμβT˙αuρβνuν\displaystyle g_{\rho\alpha}g^{\mu\beta}\dot{T}^{\alpha}{{}_{\beta\nu}}u^{\rho}u^{\nu} (26)
=\displaystyle= (ηcdhcρhdα)(ηefheμhfβ)haα(𝒟˙βBaν𝒟˙νBaβ)uρuν\displaystyle(\eta_{cd}{h^{c}}_{\rho}{h^{d}}_{\alpha})(\eta^{ef}{h_{e}}^{\mu}{h_{f}}^{\beta}){h_{a}}^{\alpha}(\dot{\mathscr{D}}_{\beta}{B^{a}}_{\nu}-\dot{\mathscr{D}}_{\nu}{B^{a}}_{\beta})u^{\rho}u^{\nu}
=\displaystyle= haμ(ηcehcρ)(ηafhfβ)(𝒟˙βBeν𝒟˙νBeβ)uρuν,\displaystyle{h_{a}}^{\mu}(\eta_{ce}{h^{c}}_{\rho})(\eta^{af}{h_{f}}^{\beta})(\dot{\mathscr{D}}_{\beta}{B^{e}}_{\nu}-\dot{\mathscr{D}}_{\nu}{B^{e}}_{\beta})u^{\rho}u^{\nu},

where (7) and (23) have been used in the second step and (8) has been used in the last step. Then plugging (18) and (26) into (24), we get the equation of motion for a point-like particle:

duμds+haμ[𝒟˙νhaρ(ηcehcρ)(ηafhfβ)(𝒟˙βBeν𝒟˙νBeβ)]uρuν=0,\frac{\mbox{d}u^{\mu}}{\mbox{d}s}+{h_{a}}^{\mu}\Bigl{[}\dot{\mathscr{D}}_{\nu}{h^{a}}_{\rho}-(\eta_{ce}{h^{c}}_{\rho})(\eta^{af}{h_{f}}^{\beta})(\dot{\mathscr{D}}_{\beta}{B^{e}}_{\nu}-\dot{\mathscr{D}}_{\nu}{B^{e}}_{\beta})\Bigr{]}u^{\rho}u^{\nu}=0, (27)

which is equivalent to the geodesic equation (20).

We now show by contradiction that in presence of gravity the gauge potential can not be eliminated from the equation (27). We first replace (10) in (27) and afterwards assume the gauge potential to vanish. Hence, we rewrite (27) in cartesian coordinates of an inertial frame in which the Lorentz connection A˙bμa\dot{A}{{}^{a}}_{b\mu} vanishes and the tetrad components take the form haρ=δaρ{h^{a}}_{\rho}={\delta^{a}}_{\rho} (see Ref. [9]). Therefore, the equation (27) simplifies to:

duμds+haμ(νδaρ)uρuν=0duμds=0,\frac{\mbox{d}u^{\mu}}{\mbox{d}s}+{h_{a}}^{\mu}(\partial_{\nu}{\delta^{a}}_{\rho})u^{\rho}u^{\nu}=0\qquad\Longrightarrow\quad\frac{\mbox{d}u^{\mu}}{\mbox{d}s}=0, (28)

where we used the definition (15) and νδaρ=0\partial_{\nu}{\delta^{a}}_{\rho}=0. Therefore, in cartesian coordinates of an inertial frame and assuming that (27) does not depend on the gauge potential, equation (27) reduces to the equation of a free particle. On the other hand, we know that in the presence of gravity (27) does not reduce to the equation of a free particle because it is equivalent to the geodesic equation (20). Therefore, in the presence of gravity we can not eliminate the gauge potential from the equation (27) and the gauge potential Baμ{B^{a}}_{\mu} represents the effect of gravity on the motion of a point-like particle.

We would also emphasize the role of the Lorentz connection A˙abμ\dot{A}^{a}{{}_{b\mu}}. As stated in Ref. [9], this connection is due to the inertial effects and it appears in the tetrad when a general reference is chosen. Hence, in this case, it also appears in the equation of motion. However, if we take an inertial frame, this connection vanishes. Indeed, such connection is constructed with the local Lorentz transformation Λad(x){\Lambda^{a}}_{d}(x) from an inertial frame to a general frame, namely A˙=abμΛad(x)μΛbd(x)\dot{A}{{}^{a}}_{b\mu}={\Lambda^{a}}_{d}(x)\partial_{\mu}{\Lambda_{b}}^{d}(x). In particular, since the Lorentz transformation from an inertial frame to another inertial frame is a global transformation, A˙bμa\dot{A}{{}^{a}}_{b\mu} vanishes in the inertial frames.

Therefore, based on the above discussions, generally, the motion of the particle is governed by both the Lorentz connection A˙abμ\dot{A}^{a}{{}_{b\mu}} and the gauge potential Baμ{B^{a}}_{\mu}. If an inertial frame is chosen, the motion is only governed by the later. These two quantities together plays a role similar to the Levi-Civita connection in general relativity. Indeed, in general relativity, the motion of the particle is governed by the Levi-Civita connection, as the equation (20) shows.

Finally, let us show that the gauge potential BaμB^{a}{{}_{\mu}} appears in the gravitational phase, though we have proved that it affects the equation of motion of the particle. As shown in the Ref. [9], the equation of motion (24) can be derived directly from the following action principle,

𝒮=mpq(uadxa+uaA˙xbabμdxμ+uaBdaμxμ),\mathscr{S}=-m\int_{p}^{q}(u_{a}\mbox{d}x^{a}+u_{a}\dot{A}{{}^{a}}_{b\mu}x^{b}\mbox{d}x^{\mu}+u_{a}B{{}^{a}}_{\mu}\mbox{d}x^{\mu}), (29)

where the first term stands for the free particle, the second term relates to the inertial effects, and the last term represents the gravitational interaction. Here ua=ηabubu_{a}=\eta_{ab}u^{b} and ubu^{b} is a four-velocity defined in the tangent space (see (33)). In presence of the electromagnetic potential AμA_{\mu}, the action (29), for a charged particle of charge qq, should be modified by adding the term (q/m)Aμdxμ(q/m)A_{\mu}\mbox{d}x^{\mu} under the integral in (29[9]. In special relativity, the action of a particle in presence of the electromagnetic field is just the combination of a free term and the interaction term with the electromagnetic potential. Correspondingly, the electromagnetic phase factor for an Aharonov-Bohm effect [1] is given by eiqAμ𝑑xμe^{\frac{i}{\hbar}\int qA_{\mu}dx^{\mu}}. Thus, for a gravitational field, in strict analogy with the electromagnetism, the last two terms in (29) contribute to the gravitational phase factor [9]. Especially, if we choose an inertial frame, the second term in (29) vanishes and only the last term contributes to the gravitational phase factor. In this frame, the gauge potential BaμB^{a}{{}_{\mu}} dominates the gravitational phase. Indeed, as we see from the definition of the field strength (14), the role of the potential BaμB^{a}{{}_{\mu}} in gravity is similar to the role of the gauge potential in electromagnetism. It deserves to be mentioned that a similar discussion of the gravitational phase can be found in Ref. [15].

3 Gravitational phase

In this section we first provide the general formula for the gravitational phase and afterwards we evaluate it explicitly for the case of the Kerr spacetime.

3.1 Gravitational phase in inertial references

The gravitational phase factor for a massive particle in a generic frame is [9, 15]:

Φg=exp(i𝒮g),\Phi_{g}=\exp\Bigl{(}-\frac{i}{\hbar}\mathscr{S}_{g}\Bigr{)}, (30)

where

𝒮g=mpqua(A˙xbabμdxμ+Bdaμxμ)\mathscr{S}_{g}=-m\int_{p}^{q}u_{a}(\dot{A}{{}^{a}}_{b\mu}x^{b}\mbox{d}x^{\mu}+B{{}^{a}}_{\mu}\mbox{d}x^{\mu}) (31)

is the interaction part of the action

𝒮=mpqds,\mathscr{S}=-m\int_{p}^{q}\mbox{d}s, (32)

and the four-velocities in spacetime and tangent-space are defined respectively as follows,

uμ=dxμds,ua=hads.u^{\mu}=\frac{\mbox{d}x^{\mu}}{\mbox{d}s},\qquad u^{a}=\frac{h^{a}}{\mbox{d}s}. (33)

For simplicity, we choose an inertial coordinate system KK in which A˙=abμ0\dot{A}{{}^{a}}_{b\mu}=0. Hence, according to (31), the interaction action reads:

𝒮g=mpquaBdaβxβ=mpqgμνuμBdνβxβ,\mathscr{S}_{g}=-m\int_{p}^{q}u_{a}B{{}^{a}}_{\beta}\mbox{d}x^{\beta}=-m\int_{p}^{q}g_{\mu\nu}u^{\mu}B{{}^{\nu}}_{\beta}\mbox{d}x^{\beta}, (34)

where the second equation in (9) is used and the function BβνB{{}^{\nu}}_{\beta} is defined as

B=νβhaBν=aβ(hTB).νβB{{}^{\nu}}_{\beta}=h_{a}{{}^{\nu}}B{{}^{a}}_{\beta}=(h^{T}B){{}^{\nu}}_{\beta}. (35)

Here hTh^{T} is the transpose matrix of haνh_{a}{{}^{\nu}}, and BB is the matrix BβaB{{}^{a}}_{\beta}. Therefore, if we have the expressions for gμνg_{\mu\nu}, uμu^{\mu} and Bνβ{B^{\nu}}_{\beta}, we can evaluate 𝒮g\mathscr{S}_{g}. Plugging (34) into (30), we get the gravitational phase factor for a massive particle:

Φg=exp(impqgμνuμBdνβxβ).\Phi_{g}=\exp\Bigl{(}\frac{i}{\hbar}m\int_{p}^{q}g_{\mu\nu}u^{\mu}B{{}^{\nu}}_{\beta}\mbox{d}x^{\beta}\Bigr{)}. (36)

For massless particles, let us consider the light firstly. For a light, its phase factor can be written as:

Φ=exp(iψ)=exp(ipqPμdxμ),\Phi=\exp(i\psi)=\exp\Bigl{(}\frac{i}{\hbar}\int_{p}^{q}P_{\mu}\mbox{d}x^{\mu}\Bigr{)}, (37)

where Pμ=kμP_{\mu}=\hbar k_{\mu} is the four-momentum of the photon, and kμk_{\mu} is the wave vector. In Ref. [8], the optical interferometry is based on (37), but for a weak gravitational field. Unlike in the Ref. [8], we extract the gravitational part from the phase factor in the framework of TG, without need of the weak field approximation. According to (37), we have:

dψ=kμdxμ=kaha=ka(dxa+A˙xbabμdxμ+Bdaμxμ),\mbox{d}\psi=k_{\mu}\mbox{d}x^{\mu}=k_{a}h^{a}=k_{a}(\mbox{d}x^{a}+\dot{A}{{}^{a}}_{b\mu}x^{b}\mbox{d}x^{\mu}+B{{}^{a}}_{\mu}\mbox{d}x^{\mu})\,, (38)

where Eqs. (9), (7), (8), and (12) have been used. Since we only need the interaction part, for the gravitational phase ϕg\phi_{g} we have:

dϕg=ka(A˙xbabμdxμ+Bdaμxμ).\mbox{d}\phi_{g}=k_{a}(\dot{A}{{}^{a}}_{b\mu}x^{b}\mbox{d}x^{\mu}+B{{}^{a}}_{\mu}\mbox{d}x^{\mu}). (39)

Moreover, if we choose an inertial frame for which A˙=abμ0\dot{A}{{}^{a}}_{b\mu}=0, the gravitational phase simplifies to:

ϕg=pqkaBdaμxμ=pqkνBνμdxμ,\phi_{g}=\int_{p}^{q}k_{a}B{{}^{a}}_{\mu}\mbox{d}x^{\mu}=\int_{p}^{q}k_{\nu}{B^{\nu}}_{\mu}\mbox{d}x^{\mu}, (40)

where (9) is used and BβνB{{}^{\nu}}_{\beta} is defined in (35). Finally, the gravitational phase factor for light is:

ΦL=exp(iϕg)=exp(ipqgμνPμBνβdxβ).\Phi_{L}=\exp(i\phi_{g})=\exp\Bigl{(}\frac{i}{\hbar}\int_{p}^{q}g_{\mu\nu}P^{\mu}{B^{\nu}}_{\beta}\mbox{d}x^{\beta}\Bigr{)}. (41)

Although (41) has been derived for photons, we assume it also applicable for other massless particles. Of course, this hypothesis needs a rigorous proof.

In summary, the gravitational phase for a particle (massive or massless) in an inertial frame is given by:

ϕg=1pqSβdxβ,\phi_{g}=\frac{1}{\hbar}\int_{p}^{q}S_{\beta}\mbox{d}x^{\beta}, (42)

where the function SβS_{\beta} is defined as

Sβ=gμνPμBνβ,S_{\beta}=g_{\mu\nu}P^{\mu}{B^{\nu}}_{\beta}, (43)

and PμP^{\mu} is the four-momentum. The gravitational phase factor is given by Φg=exp(iϕg)\Phi_{g}=\exp(i\phi_{g}).

To calculate SβS_{\beta}, we need to know Bνβ{B^{\nu}}_{\beta} firstly. According to (35), the expression of Bνβ{B^{\nu}}_{\beta} is given by haν{h_{a}}^{\nu} and BβaB{{}^{a}}_{\beta}. Thus in addition to haν{h_{a}}^{\nu}, we need to seek the expression for BβaB{{}^{a}}_{\beta}. Before proceeding, let us consider the cartesian coordinate system KK^{\prime} in in which μxa=δμa\partial_{\mu^{\prime}}x^{a}=\delta{{}^{a}}_{\mu^{\prime}} holds [9]. Therefore, according to Eq. (10), in the coordinate KK^{\prime} the gauge potential can be written as:

B=aμhaμδ.aμB{{}^{a}}_{\mu^{\prime}}=h{{}^{a}}_{\mu^{\prime}}-\delta{{}^{a}}_{\mu^{\prime}}. (44)

Moreover, the components of the tetrad in the generic coordinate KK can be expressed as [16]:

h=aρhxνxρaν,h{{}^{a}}_{\rho}=h{{}^{a}}_{\nu^{\prime}}\frac{\partial x^{\nu^{\prime}}}{\partial x^{\rho}}\,, (45)

which can be derived directly by writing the second equation of (5) as:

ha=hdaνxν=hxνxρaνdxρ=hdaρxρ.\displaystyle h^{a}=h{{}^{a}}_{\nu^{\prime}}\mbox{d}x^{\nu^{\prime}}=h{{}^{a}}_{\nu^{\prime}}\frac{\partial x^{\nu^{\prime}}}{\partial x^{\rho}}\mbox{d}x^{\rho}=h{{}^{a}}_{\rho}\mbox{d}x^{\rho}\,. (46)

Now we come back to the expression for BμaB{{}^{a}}_{\mu}. We write the gravitational phase in the coordinate KK:

ϕg=1pqgμνPμBνβdxβ=1pqPaBdaβxβ,\phi_{g}=\frac{1}{\hbar}\int_{p}^{q}g_{\mu\nu}P^{\mu}{B^{\nu}}_{\beta}\mbox{d}x^{\beta}=\frac{1}{\hbar}\int_{p}^{q}P_{a}B{{}^{a}}_{\beta}\mbox{d}x^{\beta}, (47)

where (9) and (8) are used. On the other hand, we write it in the cartesian coordinate KK^{\prime}:

ϕg=1pqPaBdaμxμ=1pqPa(hxσxμaσδ)aμdxμ,\phi_{g}=\frac{1}{\hbar}\int_{p}^{q}P_{a}B{{}^{a}}_{\mu^{\prime}}\mbox{d}x^{\mu^{\prime}}=\frac{1}{\hbar}\int_{p}^{q}P_{a}(h{{}^{a}}_{\sigma}\frac{\partial x^{\sigma}}{\partial x^{\mu^{\prime}}}-\delta{{}^{a}}_{\mu^{\prime}})\mbox{d}x^{\mu^{\prime}}, (48)

where (44) and (45) are used. Furthermore, we write (48) as:

ϕg=1pqPa(hxσxμaσδ)aμxμxβdxβ=1pqPa(haβδxμxβaμ)dxβ.\phi_{g}=\frac{1}{\hbar}\int_{p}^{q}P_{a}(h{{}^{a}}_{\sigma}\frac{\partial x^{\sigma}}{\partial x^{\mu^{\prime}}}-\delta{{}^{a}}_{\mu^{\prime}})\frac{\partial x^{\mu^{\prime}}}{\partial x^{\beta}}\mbox{d}x^{\beta}=\frac{1}{\hbar}\int_{p}^{q}P_{a}(h{{}^{a}}_{\beta}-\delta{{}^{a}}_{\mu^{\prime}}\frac{\partial x^{\mu^{\prime}}}{\partial x^{\beta}})\mbox{d}x^{\beta}. (49)

Comparing (47) and (49), we finally obtain:

B=aβhaβδxμxβaμ.B{{}^{a}}_{\beta}=h{{}^{a}}_{\beta}-\delta{{}^{a}}_{\mu^{\prime}}\frac{\partial x^{\mu^{\prime}}}{\partial x^{\beta}}. (50)

Summarizing. We choose an inertial coordinate system KK. Then we find the expression for the components of the tetrad haβ{h^{a}}_{\beta}, and the transformation between the coordinate KK and the cartesian coordinate KK^{\prime}. Plugging the tetrad into (50), we get the expression of Baβ{B^{a}}_{\beta}. Hence, inserting the latter into (35), we get Bνβ{B^{\nu}}_{\beta}. Pugging the expressions of Bνβ{B^{\nu}}_{\beta} and PμP^{\mu} into (43), we get SβS_{\beta}. According to it, we calculate the integral in (42) to finally get the gravitational phase.

3.2 Gravitational phase in the Kerr spacetime

Using Boyer-Lindquist coordinates K(t,r,θ,φ)K(t,r,\theta,\varphi) in the Kerr spacetime, the matrix form of the metric is [17]:

(gμν)=(g00g03g11g22g30g33)=(1rgr/ρ200argr(sθ)2/ρ20ρ2/Δ0000ρ20argr(sθ)2/ρ200[r2+a2+a2rgr(sθ)2/ρ2](sθ)2),(g_{\mu\nu})=\begin{pmatrix}g_{00}&{}&{}&g_{03}\\ {}&g_{11}&{}&{}\\ {}&{}&g_{22}&{}\\ g_{30}&{}&{}&g_{33}\end{pmatrix}=\begin{pmatrix}1-r_{g}r/\rho^{2}&{0}&{0}&ar_{g}r(\mbox{s}\theta)^{2}/\rho^{2}\\ {0}&-\rho^{2}/\Delta&{0}&{0}\\ {0}&{0}&-\rho^{2}&{0}\\ ar_{g}r(\mbox{s}\theta)^{2}/\rho^{2}&{0}&{0}&-[r^{2}+a^{2}+a^{2}r_{g}r(\mbox{s}\theta)^{2}/\rho^{2}](\mbox{s}\theta)^{2}\end{pmatrix}, (51)

and for the inverse:

(gμν)=(g00g03g11g22g30g33)=(Σ2/(ρ2Δ)00argr/(ρ2Δ)0Δ/ρ200001/ρ20argr/(ρ2Δ)00[Δa2(sθ)2]/[ρ2(sθ)2Δ]),(g^{\mu\nu})=\begin{pmatrix}g^{00}&{}&{}&g^{03}\\ {}&g^{11}&{}&{}\\ {}&{}&g^{22}&{}\\ g^{30}&{}&{}&g^{33}\end{pmatrix}=\begin{pmatrix}\Sigma^{2}/(\rho^{2}\Delta)&{0}&{0}&ar_{g}r/(\rho^{2}\Delta)\\ {0}&-\Delta/\rho^{2}&{0}&{0}\\ {0}&{0}&-1/\rho^{2}&{0}\\ ar_{g}r/(\rho^{2}\Delta)&0&0&-[\Delta-a^{2}(\mbox{s}\theta)^{2}]/[\rho^{2}(\mbox{s}\theta)^{2}\Delta]\end{pmatrix}, (52)

where

rg=2M,ρ2=r2+a2(cθ)2,Δ=r2rgr+a2,Σ2=(r2+a2)2a2(sθ)2Δ.r_{g}=2M,\quad\rho^{2}=r^{2}+a^{2}(\mbox{c}\theta)^{2},\quad\Delta=r^{2}-r_{g}r+a^{2},\quad\Sigma^{2}=(r^{2}+a^{2})^{2}-a^{2}(\mbox{s}\theta)^{2}\Delta. (53)

The parameter MM is the mass of the black hole, and aa is its angular momentum per unit of mass in the units c=1c=1. (In the SI units it is a=J/(Mc)a=J/(Mc), where JJ is the angular momentum of the black hole [18].) Here the symbols sθ\mbox{s}\theta, cθ\mbox{c}\theta, sφ\mbox{s}\varphi and cφ\mbox{c}\varphi denote sin(θ)\sin(\theta), cos(θ)\cos(\theta), sin(φ)\sin(\varphi) and cos(φ)\cos(\varphi) respectively.

We will calculate the gravitational phase in the Kerr spacetime by using the last expression in (34), but before that, we derive the expression of BβaB{{}^{a}}_{\beta} according to Eq. (50). The coordinate transformation from KK^{\prime} to KK is [18]:

{t=t,x=r2+a2sθcφ,y=r2+a2sθsφ,z=rcθ.\begin{cases}t^{\prime}=t,\\ x^{\prime}=\sqrt{r^{2}+a^{2}}\,\mbox{s}\theta\mbox{c}\varphi,\\ y^{\prime}=\sqrt{r^{2}+a^{2}}\,\mbox{s}\theta\mbox{s}\varphi,\\ z^{\prime}=r\,\mbox{c}\theta.\end{cases} (54)

From Eq. (54) we can get the Jacobi matrix:

(xμxβ)=(10000rρ0sθcφρ0cθcφρ0sθsφ0rρ0sθsφρ0cθsφρ0sθcφ0cθrsθ0),\Bigl{(}\frac{\partial x^{\mu^{\prime}}}{\partial x^{\beta}}\Bigr{)}=\begin{pmatrix}1&0&0&0\\ 0&\frac{r}{\rho_{0}}\mbox{s}\theta\mbox{c}\varphi&\rho_{0}\mbox{c}\theta\mbox{c}\varphi&-\rho_{0}\mbox{s}\theta\mbox{s}\varphi\\ 0&\frac{r}{\rho_{0}}\mbox{s}\theta\mbox{s}\varphi&\rho_{0}\mbox{c}\theta\mbox{s}\varphi&\rho_{0}\mbox{s}\theta\mbox{c}\varphi\\ 0&\mbox{c}\theta&-r\mbox{s}\theta&0\end{pmatrix}, (55)

where ρ0=r2+a2\rho_{0}=\sqrt{r^{2}+a^{2}}. The tetrad in the Kerr spacetime is [9, 16]:

(h)aβ=(γ0000η0γ11sθcφγ22cθcφζsφ0γ11sθsφγ22cθsφζcφ0γ11cθγ22sθ0),(h{{}^{a}}_{\beta})=\begin{pmatrix}\gamma_{00}&0&0&\eta\\ 0&\gamma_{11}\mbox{s}\theta\mbox{c}\varphi&\gamma_{22}\mbox{c}\theta\mbox{c}\varphi&-\zeta\mbox{s}\varphi\\ 0&\gamma_{11}\mbox{s}\theta\mbox{s}\varphi&\gamma_{22}\mbox{c}\theta\mbox{s}\varphi&\zeta\mbox{c}\varphi\\ 0&\gamma_{11}\mbox{c}\theta&-\gamma_{22}\mbox{s}\theta&0\end{pmatrix}, (56)

where222In Ref. [16] the authors only give the expression ζ2=η2g33\zeta^{2}=\eta^{2}-g_{33}. We believe ζ=η2g33\zeta=\sqrt{\eta^{2}-g_{33}} also holds, which is in accordance with the tetrad in Schwardschild space time (see (29) in Ref. [16]).

η=g03/γ00,ζ=η2g33,γ00=g00,γjj=gjj.\eta=g_{03}/\gamma_{00},\quad\zeta=\sqrt{\eta^{2}-g_{33}},\quad\gamma_{00}=\sqrt{g_{00}},\quad\gamma_{jj}=\sqrt{-g_{jj}}. (57)

Inserting Eqs. (56) and (55) into (50), we get the gauge potential in the Kerr spacetime:

(B)aβ=(γ00100η0(γ11rρ0)sθcφ(γ22ρ0)cθcφ(ρ0sθζ)sφ0(γ11rρ0)sθsφ(γ22ρ0)cθsφ(ζρ0sθ)cφ0(γ111)cθ(rγ22)sθ0).(B{{}^{a}}_{\beta})=\begin{pmatrix}\gamma_{00}-1&0&0&\eta\\ 0&(\gamma_{11}-\frac{r}{\rho_{0}})\mbox{s}\theta\mbox{c}\varphi&(\gamma_{22}-\rho_{0})\mbox{c}\theta\mbox{c}\varphi&(\rho_{0}\mbox{s}\theta-\zeta)\mbox{s}\varphi\\ 0&(\gamma_{11}-\frac{r}{\rho_{0}})\mbox{s}\theta\mbox{s}\varphi&(\gamma_{22}-\rho_{0})\mbox{c}\theta\mbox{s}\varphi&(\zeta-\rho_{0}\mbox{s}\theta)\mbox{c}\varphi\\ 0&(\gamma_{11}-1)\mbox{c}\theta&(r-\gamma_{22})\mbox{s}\theta&0\end{pmatrix}. (58)

It is easy to check that Baβ=0{B^{a}}_{\beta}=0 in flat spacetime, namely for a=0a=0 and M=0M=0. The matrix form for the inverse of the tetrad is333We think that these are some typos in equation (14.37) in [9]. This equation should be modified as (59).:

(ha)ν=(γ001000ζg03sφγ111sθcφγ221cθcφζ1sφζg03cφγ111sθsφγ221cθsφζ1cφ0γ111cθγ221sθ0).(h_{a}{{}^{\nu}})=\begin{pmatrix}\gamma_{00}^{-1}&0&0&0\\ \zeta g^{03}\mbox{s}\varphi&\gamma_{11}^{-1}\mbox{s}\theta\mbox{c}\varphi&\gamma_{22}^{-1}\mbox{c}\theta\mbox{c}\varphi&-\zeta^{-1}\mbox{s}\varphi\\ -\zeta g^{03}\mbox{c}\varphi&\gamma_{11}^{-1}\mbox{s}\theta\mbox{s}\varphi&\gamma_{22}^{-1}\mbox{c}\theta\mbox{s}\varphi&\zeta^{-1}\mbox{c}\varphi\\ 0&\gamma_{11}^{-1}\mbox{c}\theta&-\gamma_{22}^{-1}\mbox{s}\theta&0\end{pmatrix}. (59)

One can check that (56) and (59) indeed satisfy gμν=ηabhaμhbνg_{\mu\nu}=\eta_{ab}{h^{a}}_{\mu}{h^{b}}_{\nu}, ηab=gμνhaμhbν\eta_{ab}=g_{\mu\nu}{h_{a}}^{\mu}{h_{b}}^{\nu} and haμhbμ=δba{h^{a}}_{\mu}{h_{b}}^{\mu}=\delta^{a}_{b}. Inserting Eqs. (58) and (59) into (35), we obtain

(B)νβ=(1γ00100ηγ001+(ρ0sθζ)ζg0301γ111[rρ01(sθ)2+(cθ)2]γ111sθcθ(rρ0)00γ221sθcθ(1r/ρ0)1γ221[r(sθ)2+ρ0(cθ)2]00001ζ1ρ0sθ).(B{{}^{\nu}}_{\beta})=\begin{pmatrix}1-\gamma_{00}^{-1}&0&0&\eta\gamma_{00}^{-1}+(\rho_{0}\mbox{s}\theta-\zeta)\zeta g^{03}\\ 0&1-\gamma_{11}^{-1}[r\rho_{0}^{-1}(\mbox{s}\theta)^{2}+(\mbox{c}\theta)^{2}]&\gamma_{11}^{-1}\mbox{s}\theta\mbox{c}\theta(r-\rho_{0})&0\\ 0&\gamma_{22}^{-1}\mbox{s}\theta\mbox{c}\theta(1-r/\rho_{0})&1-\gamma_{22}^{-1}[r(\mbox{s}\theta)^{2}+\rho_{0}(\mbox{c}\theta)^{2}]&0\\ 0&0&0&1-\zeta^{-1}\rho_{0}\mbox{s}\theta\end{pmatrix}. (60)

In terms of Eq. (43), the first expression in Eq. (51), and Eq. (60), the matrix SβS_{\beta} can be written as:

(Sβ)=(PμgμνB)νβ=((1γ001)(P0g00+P3g30)P1g11{1γ111[rρ01(sθ)2+(cθ)2]}+P2g22γ221sθcθ(1r/ρ0)P1g11γ111sθcθ(rρ0)+P2g22{1γ221[r(sθ)2+ρ0(cθ)2]}(P0g00+P3g30)[ηγ001+(ρ0sθζ)ζg03]+(P0g03+P3g33)(1ζ1ρ0sθ)).(S_{\beta})=(P^{\mu}g_{\mu\nu}B{{}^{\nu}}_{\beta})=\begin{pmatrix}(1-\gamma_{00}^{-1})(P^{0}g_{00}+P^{3}g_{30})\\ P^{1}g_{11}\{1-\gamma_{11}^{-1}[r\rho_{0}^{-1}(\mbox{s}\theta)^{2}+(\mbox{c}\theta)^{2}]\}+P^{2}g_{22}\gamma_{22}^{-1}\mbox{s}\theta\mbox{c}\theta(1-r/\rho_{0})\\ P^{1}g_{11}\gamma_{11}^{-1}\mbox{s}\theta\mbox{c}\theta(r-\rho_{0})+P^{2}g_{22}\{1-\gamma_{22}^{-1}[r(\mbox{s}\theta)^{2}+\rho_{0}(\mbox{c}\theta)^{2}]\}\\ (P^{0}g_{00}+P^{3}g_{30})[\eta\gamma_{00}^{-1}+(\rho_{0}\mbox{s}\theta-\zeta)\zeta g^{03}]+(P^{0}g_{03}+P^{3}g_{33})(1-\zeta^{-1}\rho_{0}\mbox{s}\theta)\end{pmatrix}. (61)

The latter result can be further simplified by using the following conserved quantities in the Kerr spacetime [17],

E\displaystyle E =\displaystyle= (1rgrρ2)dtdξ+argr(sθ)2ρ2dφdξ=u0g00+u3g30,\displaystyle\Bigl{(}1-\frac{r_{g}r}{\rho^{2}}\Bigr{)}\frac{\mbox{d}t}{\mbox{d}\xi}+\frac{ar_{g}r(s\theta)^{2}}{\rho^{2}}\frac{\mbox{d}\varphi}{\mbox{d}\xi}=u^{0}g_{00}+u^{3}g_{30},
L\displaystyle-L =\displaystyle= argr(sθ)2ρ2dtdξ[r2+a2+rgrρ2a2(sθ)2](sθ)2dφdξ=u0g03+u3g33,\displaystyle\frac{ar_{g}r(s\theta)^{2}}{\rho^{2}}\frac{\mbox{d}t}{\mbox{d}\xi}-\Bigl{[}r^{2}+a^{2}+\frac{r_{g}r}{\rho^{2}}a^{2}(s\theta)^{2}\Bigr{]}(s\theta)^{2}\frac{\mbox{d}\varphi}{\mbox{d}\xi}=u^{0}g_{03}+u^{3}g_{33}, (62)

where ξ\xi is an affine parameter (for massive particles it is the proper time), and EE and LL are defined as

E={m1,for massive particles,,for massless particles,L={m1,for massive particles,,for massless particles.E=\begin{cases}\mathcal{E}m^{-1},&\text{for massive particles,}\\ \mathcal{E},&\text{for massless particles,}\end{cases}\qquad L=\begin{cases}\mathcal{L}m^{-1},&\text{for massive particles,}\\ \mathcal{L},&\text{for massless particles.}\end{cases} (63)

Notice that for massive particles we have Pμ=muμP^{\mu}=mu^{\mu}, while for massless particles we have Pμ=uμP^{\mu}=u^{\mu}. Here the quantity \mathcal{E} has the meaning of energy, while the quantity \mathcal{L} has the meaning of angular momentum along the spin of the black hole. Plugging (62) into (61), we get

(Sβ)=((1γ001)P1g11{1γ111[rρ01(sθ)2+(cθ)2]}+P2g22γ221sθcθ(1r/ρ0)P1g11γ111sθcθ(rρ0)+P2g22{1γ221[r(sθ)2+ρ0(cθ)2]}[ηγ001+(ρ0sθζ)ζg03](1ζ1ρ0sθ)).(S_{\beta})=\begin{pmatrix}(1-\gamma_{00}^{-1})\mathcal{E}\\ P^{1}g_{11}\{1-\gamma_{11}^{-1}[r\rho_{0}^{-1}(\mbox{s}\theta)^{2}+(\mbox{c}\theta)^{2}]\}+P^{2}g_{22}\gamma_{22}^{-1}\mbox{s}\theta\mbox{c}\theta(1-r/\rho_{0})\\ P^{1}g_{11}\gamma_{11}^{-1}\mbox{s}\theta\mbox{c}\theta(r-\rho_{0})+P^{2}g_{22}\{1-\gamma_{22}^{-1}[r(\mbox{s}\theta)^{2}+\rho_{0}(\mbox{c}\theta)^{2}]\}\\ \mathcal{E}[\eta\gamma_{00}^{-1}+(\rho_{0}\mbox{s}\theta-\zeta)\zeta g^{03}]-\mathcal{L}(1-\zeta^{-1}\rho_{0}\mbox{s}\theta)\end{pmatrix}. (64)

Finally, recalling (42), the gravitational phase in the Kerr spacetime is given by:

ϕ=1pqSβdxβ.\phi=\frac{1}{\hbar}\int_{p}^{q}S_{\beta}\mbox{d}x^{\beta}. (65)

4 Particles interference experiment

4.1 Theoretical prediction

We will study an interference experiment in the region rrgr\gg r_{g} with the size of the setup much smaller than its distance from the black hole. Let us start with a review of the Colella-Overhauser-Werner (COW) experiment on the earth [11]. The principle of this experiment is shown in FIG. 1, where the parallelogram is vertical and its base AB is parallel to the surface of the earth. A beam of neutrons is split into two beams along the paths ABC and ADC respectively, and, afterwards they interfere. Since of the presence of gravity, the phase accumulated along the path ABC is different from the phase accumulated along ADC. The theoretical predictions for this experiment were given in Ref. [10], in which the gravitational phase difference between these two paths was found to be:

δϕ=m2glλd2π2s,\delta\phi=\frac{m^{2}gl\lambda_{d}}{2\pi\hbar^{2}}s, (66)

where mm is the mass of the neutron, λd=2π/(mv)\lambda_{d}=2\pi\hbar/(mv) is its de Broglie wavelength, gg is the gravitational acceleration, ll is the height of the parallelogram, and ss is the length of AB.

Refer to caption
Figure 1: Schematic figure for the COW experiment. A beam of neutrons is split into two beams along the sides of a parallelogram which is vertical to the surface of the earth, and, afterwards they interfere.

Let us now to place the parallelogram in the region of the Kerr spacetime for rrgr\gg r_{g}. The particles are not limited to be neutrons and the devise is shown in Fig 2. For simplicity, we assume:
(a) The size of the parallelogram to be much smaller than its distance from the black hole, so that the coordinates rr and θ\theta are approximately constant along the paths AB and DC;
(b) The energy \mathcal{E} of the particle is conserved even when the particle turns direction at the points B and D (The quantity \mathcal{L} changes at these points, but it is conserved on the paths AB, BC, AD and DC), such that the magnitude of its velocity (defined in (153)) does not change at such points.

Refer to caption
Figure 2: The interference experiment in the region rrgr\gg r_{g} of the Kerr spacetime. The particles are split into two beams along the paths ABC and ADC respectively to afterward interfere. In rrgr\gg r_{g}, we can use the Schwarzschild coordinates to approximate the Boyer-Lindquist coordinates. The axis zz of the black hole and the vectors r1\vec{r}_{1} and r2\vec{r}_{2} are in the same plane, which is perpendicular to the base AB. The angle between r1\vec{r}_{1} and the plane of the parallelogram ABCD is γ\gamma. The length of AB is ss, and the height of the parallelogram is ll.

Combining the assumption (a) with (65), we write the accumulated gravitational phase along the path AB as

ϕAB\displaystyle\phi_{AB} \displaystyle\approx 1(S0dt+S3dφ)\displaystyle\frac{1}{\hbar}\Bigl{(}\int S_{0}\mbox{d}t+\int S_{3}\mbox{d}\varphi\Bigr{)} (67)
=\displaystyle= 1(S0tAB+S3φAB)\displaystyle\frac{1}{\hbar}(S_{0}t_{AB}+S_{3}\varphi_{AB})
=\displaystyle= 1(1γ001)tAB+1{[ηγ001+(ρ0sθζ)ζg03]AB(1ζ1ρ0sθ)}φAB,\displaystyle\frac{1}{\hbar}\mathcal{E}(1-\gamma_{00}^{-1})t_{AB}+\frac{1}{\hbar}\Bigl{\{}\mathcal{E}[\eta\gamma_{00}^{-1}+(\rho_{0}\mbox{s}\theta-\zeta)\zeta g^{03}]-\mathcal{L}_{\rm AB}(1-\zeta^{-1}\rho_{0}\mbox{s}\theta)\Bigr{\}}\varphi_{AB},

where tAB=tBtAt_{AB}=t_{B}-t_{A} and φAB=φBφA\varphi_{AB}=\varphi_{B}-\varphi_{A} are defined. We only consider the case a<Ma<M,444We do not consider the case a>Ma>M and a=Ma=M because the former leads to a naked singularity and the latter is unstable [19]. so that a/r1<rg/r1a/r_{1}<r_{g}/r_{1} holds. For convenience, we assume

O(ar1)O(rgr1).O\Bigl{(}\frac{a}{r_{1}}\Bigr{)}\sim O\Bigl{(}\frac{r_{g}}{r_{1}}\Bigr{)}. (68)

Therefore, expanding (67) at the third order in the two quantities (68), we get:

ϕAB1tAB[rg2r13rg28r125rg316r13+a2rg2r13cos2(θ1)]+1φABsin2(θ1)(argr112ABa2rgr13+arg2r12),\phi_{AB}\approx\frac{1}{\hbar}\mathcal{E}t_{AB}\Bigl{[}-\frac{r_{g}}{2r_{1}}-\frac{3r_{g}^{2}}{8r_{1}^{2}}-\frac{5r_{g}^{3}}{16r_{1}^{3}}+\frac{a^{2}r_{g}}{2r_{1}^{3}}\cos^{2}(\theta_{1})\Bigr{]}+\frac{1}{\hbar}\varphi_{AB}\sin^{2}(\theta_{1})\Bigl{(}\mathcal{E}\frac{ar_{g}}{r_{1}}-\frac{1}{2}\mathcal{L}_{\rm AB}\frac{a^{2}r_{g}}{r_{1}^{3}}+\mathcal{E}\frac{ar_{g}^{2}}{r_{1}^{2}}\Bigr{)}, (69)

where we regard (rg/r1)i(a/r1)j(r_{g}/r_{1})^{i}(a/r_{1})^{j} as a term of the order (i+j)(i+j). The quantity \mathcal{L} is given by (see Appendix A)

=g00(g03+vφΓ33g00),\mathcal{L}=\frac{\mathcal{E}}{g_{00}}(-g_{03}+v^{\varphi}\Gamma_{33}\sqrt{g_{00}}), (70)

where vjv^{j} is the three-dimensional velocity and Γij\Gamma_{ij} is the three-dimensional metric tenor defined by [18]555To distinguish the three-velocity (71) from the four-velocity, we emphasize that the velocity vv (given by (153)), which appears in the phase differences, is the ratio between the proper length and the observer’s proper time, namely v=dL/dτv=\mbox{d}L/\mbox{d}\tau. The first equation in (71) is actually equivalent to the definition vk=dxk/dτv^{k}=\mbox{d}x^{k}/\mbox{d}\tau (see Sec. 88 in Ref. [18]). Pay attention that here the metric is not diagonal.

vk=dxkg00(dx0+g0ig00dxi),Γij=gij+g0ig0jg00.v^{k}=\frac{\mbox{d}x^{k}}{\sqrt{g_{00}}(\mbox{d}x^{0}+\frac{g_{0i}}{g_{00}}\mbox{d}x^{i})},\qquad\Gamma_{ij}=-g_{ij}+\frac{g_{0i}g_{0j}}{g_{00}}. (71)

For the gravitational phases ϕBC\phi_{BC}, ϕAD\phi_{AD}, and ϕDC\phi_{DC}, the derivation is similar to ϕAB\phi_{AB}. Combining these phases, we can get the phase difference between the paths ADC and ABC. Hence, with the following relations (see Appendix A):

r2r1+lcos(γ)g11(r1,θ1),θ2|θ1lsin(γ)g22(r1,θ1)|,r_{2}\approx r_{1}+\frac{l\cos(\gamma)}{\sqrt{-g_{11}(r_{1},\theta_{1})}},\qquad\theta_{2}\approx\Bigl{|}\theta_{1}-\frac{l\sin(\gamma)}{\sqrt{-g_{22}(r_{1},\theta_{1})}}\Bigr{|}, (72)

we can expand the phase difference in the neighborhoods of r1r_{1} and θ1\theta_{1}, and for simplicity we only keep the first order terms of l/r1l/r_{1}. Then relate the time, the angle, and the energy with the observations (Appendix A):

tABs(1vg00g03g00Γ33),φABsΓ33,\displaystyle t_{AB}\approx s\Bigl{(}\frac{1}{v\sqrt{g_{00}}}-\frac{g_{03}}{g_{00}\sqrt{\Gamma_{33}}}\Bigr{)},\quad\varphi_{AB}\approx\frac{s}{\sqrt{\Gamma_{33}}}, (73)
={m(1v2)1/2g00,for massive particles,ωg00,for massless particles.\displaystyle\mathcal{E}=\begin{cases}m(1-v^{2})^{-1/2}\sqrt{g_{00}},&\text{for massive particles},\\ \hbar\omega\sqrt{g_{00}},&\text{for massless particles}.\end{cases} (74)

With the above steps, we derive the phase difference between the paths ADC and ABC as follows (see Appendix B for more details)

δϕ\displaystyle\delta\phi \displaystyle\approx 0lsr1{1v[cos(γ)(rg2r1+rg22r12+rg32r13+a2rg4r13(17cos2(θ1)))+sin(2θ1)sin(γ)a2rg2r13]\displaystyle\frac{\mathcal{E}_{0}ls}{\hbar r_{1}}\Bigl{\{}\frac{1}{v}\Bigl{[}\cos(\gamma)\Bigl{(}\frac{r_{g}}{2r_{1}}+\frac{r_{g}^{2}}{2r_{1}^{2}}+\frac{r_{g}^{3}}{2r_{1}^{3}}+\frac{a^{2}r_{g}}{4r_{1}^{3}}\bigl{(}1-7\cos^{2}(\theta_{1})\bigr{)}\Bigr{)}+\sin(2\theta_{1})\sin(\gamma)\frac{a^{2}r_{g}}{2r_{1}^{3}}\Bigr{]} (75)
+a2rgr13sin(θ1)[v(cos(γ)sin(θ1)+32sin(γ)cos(θ1))+12sin(θ1γ)(vv2(vr)2(r1vθ)2)]\displaystyle+\frac{a^{2}r_{g}}{r_{1}^{3}}\sin(\theta_{1})\Bigl{[}v\Bigl{(}\cos(\gamma)\sin(\theta_{1})+\frac{3}{2}\sin(\gamma)\cos(\theta_{1})\Bigr{)}+\frac{1}{2}\sin(\theta_{1}-\gamma)\Bigl{(}v-\sqrt{v^{2}-(v^{r})^{2}-(r_{1}v^{\theta})^{2}}\Bigr{)}\Bigr{]}
argr12(2cos(θ1)sin(γ)+cos(γ)sin(θ1))arg2r13(cos(θ1)sin(γ)+32cos(γ)sin(θ1))},\displaystyle-\frac{ar_{g}}{r_{1}^{2}}\Bigl{(}2\cos(\theta_{1})\sin(\gamma)+\cos(\gamma)\sin(\theta_{1})\Bigr{)}-\frac{ar_{g}^{2}}{r_{1}^{3}}\Bigl{(}\cos(\theta_{1})\sin(\gamma)+\frac{3}{2}\cos(\gamma)\sin(\theta_{1})\Bigr{)}\Bigr{\}},

where vrv^{r} and vθv^{\theta} (defind in (71)) are the velocity components at the point B corresponding to the path BC, and 0\mathcal{E}_{0} is defined by

0={m(1v2)1/2,for massive particles,ω,for massless particles.\mathcal{E}_{0}=\begin{cases}m(1-v^{2})^{-1/2},&\text{for massive particles},\\ \hbar\omega,&\text{for massless particles}.\end{cases} (76)

By the way, the expression v2(vr)2(r1vθ)2\sqrt{v^{2}-(v^{r})^{2}-(r_{1}v^{\theta})^{2}} in (75) can be replaced by |vcos(ζ)||v\cos(\zeta)|, where ζ\zeta is a base angle of the parallelogram666Because this expression only appears in the third order terms in (75), at the point B corresponding to the path BC we have v2(vr)2(r1vθ)2r1sin(θ1)vφ=r1sin(θ1)dφdτ=r1sin(θ1)dldτdLφdldφdLφ=r1sin(θ1)v|cos(ζ)|Γ331|vcos(ζ)|,\displaystyle\sqrt{v^{2}-(v^{r})^{2}-(r_{1}v^{\theta})^{2}}\approx r_{1}\sin(\theta_{1})v^{\varphi}=r_{1}\sin(\theta_{1})\frac{\mbox{d}\varphi}{\mbox{d}\tau}=r_{1}\sin(\theta_{1})\frac{\mbox{d}l}{\mbox{d}\tau}\frac{\mbox{d}L_{\varphi}}{\mbox{d}l}\frac{\mbox{d}\varphi}{\mbox{d}L_{\varphi}}=r_{1}\sin(\theta_{1})v\,|\!\cos(\zeta)|\sqrt{\Gamma_{33}^{-1}}\approx|v\cos(\zeta)|, where (154) and (120) have been used in the first and the last second steps respectively. . In particular, for γ=0\gamma=0 and γ=π/2\gamma=\pi/2, the gravitational phase differences are respectively:

δϕ|γ=0\displaystyle\delta\phi|_{\gamma=0} \displaystyle\approx 0lsr1{1v[rg2r1+rg22r12+rg32r13+a2rg4r13(17cos2(θ1))]\displaystyle\frac{\mathcal{E}_{0}ls}{\hbar r_{1}}\Bigl{\{}\frac{1}{v}\Bigl{[}\frac{r_{g}}{2r_{1}}+\frac{r_{g}^{2}}{2r_{1}^{2}}+\frac{r_{g}^{3}}{2r_{1}^{3}}+\frac{a^{2}r_{g}}{4r_{1}^{3}}\Bigl{(}1-7\cos^{2}(\theta_{1})\Bigr{)}\Bigr{]} (77)
+a2rgr13sin(θ1)[vsin(θ1)+12sin(θ1)(vv2(vr)2(r1vθ)2)]\displaystyle+\frac{a^{2}r_{g}}{r_{1}^{3}}\sin(\theta_{1})\Bigl{[}v\sin(\theta_{1})+\frac{1}{2}\sin(\theta_{1})\Bigl{(}v-\sqrt{v^{2}-(v^{r})^{2}-(r_{1}v^{\theta})^{2}}\Bigr{)}\Bigr{]}
argr12sin(θ1)3arg22r13sin(θ1)},\displaystyle-\frac{ar_{g}}{r_{1}^{2}}\sin(\theta_{1})-\frac{3ar_{g}^{2}}{2r_{1}^{3}}\sin(\theta_{1})\Bigr{\}},
δϕ|γ=π2\displaystyle\delta\phi|_{\gamma=\frac{\pi}{2}} \displaystyle\approx 0lsr1{1vsin(2θ1)a2rg2r13+a2rgr13sin(θ1)[32vcos(θ1)12cos(θ1)(vv2(vr)2(r1vθ)2)]\displaystyle\frac{\mathcal{E}_{0}ls}{\hbar r_{1}}\Bigl{\{}\frac{1}{v}\sin(2\theta_{1})\frac{a^{2}r_{g}}{2r_{1}^{3}}+\frac{a^{2}r_{g}}{r_{1}^{3}}\sin(\theta_{1})\Bigl{[}\frac{3}{2}v\cos(\theta_{1})-\frac{1}{2}\cos(\theta_{1})\Bigl{(}v-\sqrt{v^{2}-(v^{r})^{2}-(r_{1}v^{\theta})^{2}}\Bigr{)}\Bigr{]} (78)
2argr12cos(θ1)arg2r13cos(θ1)}.\displaystyle-\frac{2ar_{g}}{r_{1}^{2}}\cos(\theta_{1})-\frac{ar_{g}^{2}}{r_{1}^{3}}\cos(\theta_{1})\Bigr{\}}.

The prediction (75) can be tested experimentally by measuring the fringe shift as a function of γ\gamma. As for the non-relativistic particles, (75) is reduced to

δϕNR\displaystyle\delta\phi^{\rm NR} \displaystyle\approx mlsr1{(1v+v2)[cos(γ)(rg2r1+rg22r12+rg32r13+a2rg4r13(17cos2(θ1)))+sin(2θ1)sin(γ)a2rg2r13]\displaystyle\frac{mls}{\hbar r_{1}}\Bigl{\{}\Bigl{(}\frac{1}{v}+\frac{v}{2}\Bigr{)}\Bigl{[}\cos(\gamma)\Bigl{(}\frac{r_{g}}{2r_{1}}+\frac{r_{g}^{2}}{2r_{1}^{2}}+\frac{r_{g}^{3}}{2r_{1}^{3}}+\frac{a^{2}r_{g}}{4r_{1}^{3}}\bigl{(}1-7\cos^{2}(\theta_{1})\bigr{)}\Bigr{)}+\sin(2\theta_{1})\sin(\gamma)\frac{a^{2}r_{g}}{2r_{1}^{3}}\Bigr{]} (79)
+a2rgr13sin(θ1)[v(cos(γ)sin(θ1)+32sin(γ)cos(θ1))+12sin(θ1γ)(vv2(vr)2(r1vθ)2)]\displaystyle+\frac{a^{2}r_{g}}{r_{1}^{3}}\sin(\theta_{1})\Bigl{[}v\Bigl{(}\cos(\gamma)\sin(\theta_{1})+\frac{3}{2}\sin(\gamma)\cos(\theta_{1})\Bigr{)}+\frac{1}{2}\sin(\theta_{1}-\gamma)\Bigl{(}v-\sqrt{v^{2}-(v^{r})^{2}-(r_{1}v^{\theta})^{2}}\Bigr{)}\Bigr{]}
argr12(2cos(θ1)sin(γ)+cos(γ)sin(θ1))arg2r13(cos(θ1)sin(γ)+32cos(γ)sin(θ1))},\displaystyle-\frac{ar_{g}}{r_{1}^{2}}\Bigl{(}2\cos(\theta_{1})\sin(\gamma)+\cos(\gamma)\sin(\theta_{1})\Bigr{)}-\frac{ar_{g}^{2}}{r_{1}^{3}}\Bigl{(}\cos(\theta_{1})\sin(\gamma)+\frac{3}{2}\cos(\gamma)\sin(\theta_{1})\Bigr{)}\Bigr{\}},

where we have neglected the terms of O(v2)O(v^{2}) and higher orders.

From (75) we can find that the quantity aa only appears in the second and higher order terms. We can also find that in the Newtonian limit the equation (75) reproduces the result (66) on the earth. Indeed, in such limit we have v1v\ll 1 and the phase difference is dominated by the first term in (75), namely

δϕmassmrgs2r12vlcos(γ)=m2slcos(γ)λd2π2rg2r12.\delta\phi_{\rm mass}\approx\frac{mr_{g}s}{2\hbar r_{1}^{2}v}l\cos(\gamma)=\frac{m^{2}sl\cos(\gamma)\lambda_{d}}{2\pi\hbar^{2}}\frac{r_{g}}{2r_{1}^{2}}. (80)

On the other hand, since rg/(2r12)=gr_{g}/(2r_{1}^{2})=g, (80) is equivalent to (66) by setting γ=0\gamma=0. Equation (80) can also be derived from the gravitational phase directly evaluated in the Newtonian limit:

ϕ1S0dt,\phi\approx\frac{1}{\hbar}\int S_{0}\mbox{d}t, (81)

by expanding the result in the ratios rg/rr_{g}/r and a/ra/r and only keeping the first order term.

Notice that (75) holds only when the condition sΓ33s\ll\sqrt{\Gamma_{33}} is satisfied (recall the second equation in (73)). If the latter condition is violated, the equation (75) should be modified. Take θ1=0\theta_{1}=0 and θ1=π\theta_{1}=\pi as examples, then the second equation in (73) should be replaced by the equation φAB=π\varphi_{AB}=\pi. Correspondingly, (75) should be replaced by the following expression (see the last paragraph in the Appendix B)

δϕ|θ1=0,π=\displaystyle\delta\phi|_{\theta_{1}=0,\pi}= \displaystyle\approx 0lsr1{1v[cos(γ)(rg2r1+rg22r12+rg32r13+a2rg4r13(17cos2(θ1)))+sin(2θ1)sin(γ)a2rg2r13]\displaystyle\frac{\mathcal{E}_{0}ls}{\hbar r_{1}}\Bigl{\{}\frac{1}{v}\Bigl{[}\cos(\gamma)\Bigl{(}\frac{r_{g}}{2r_{1}}+\frac{r_{g}^{2}}{2r_{1}^{2}}+\frac{r_{g}^{3}}{2r_{1}^{3}}+\frac{a^{2}r_{g}}{4r_{1}^{3}}\bigl{(}1-7\cos^{2}(\theta_{1})\bigr{)}\Bigr{)}+\sin(2\theta_{1})\sin(\gamma)\frac{a^{2}r_{g}}{2r_{1}^{3}}\Bigr{]} (82)
arg22r13cos(γ)sin(θ1)}+π0l{a2rgr13vsin2(θ1)(cos(γ)sin(θ1)+32sin(γ)cos(θ1))\displaystyle-\frac{ar_{g}^{2}}{2r_{1}^{3}}\cos(\gamma)\sin(\theta_{1})\Bigr{\}}+\frac{\pi\mathcal{E}_{0}l}{\hbar}\Bigl{\{}\frac{a^{2}r_{g}}{r_{1}^{3}}v\sin^{2}(\theta_{1})\Bigl{(}\cos(\gamma)\sin(\theta_{1})+\frac{3}{2}\sin(\gamma)\cos(\theta_{1})\Bigr{)}
argr12sin(θ1)(cos(θ1)sin(γ)+sin(θ1+γ))arg2r13sin(θ1)sin(θ1+γ)}.\displaystyle-\frac{ar_{g}}{r_{1}^{2}}\sin(\theta_{1})\Bigl{(}\cos(\theta_{1})\sin(\gamma)+\sin(\theta_{1}+\gamma)\Bigr{)}-\frac{ar_{g}^{2}}{r_{1}^{3}}\sin(\theta_{1})\sin(\theta_{1}+\gamma)\Bigr{\}}.

4.2 Impact of the angular momentum of the black hole

We here discuss the contribution coming from the spin of the black hole. According to (69) the quantity aa only appears in the second and higher order terms (we remind that that the order of (rg/r1)i(a/r1)j(r_{g}/r_{1})^{i}(a/r_{1})^{j} is (i+j(i+j).) Furthermore, we claim that in the region rrgr\gg r_{g} the quantity aa does not appear at the first order term in the local gravitational phase (here local means that the path is short enough so that hold SβdxβSβδxβ\int S_{\beta}\mbox{d}x^{\beta}\approx S_{\beta}\delta x^{\beta}, where δxβ\delta x^{\beta} are coordinates’ differences). Indeed, we can expand the function SβS_{\beta} in (64) respect to the parameters κg\kappa_{g} and κa\kappa_{a}, where κg=rg/r\kappa_{g}=r_{g}/r and κa=a/r\kappa_{a}=a/r. Hence, we get:

(Sβ)((12κg38κg2516κg3+12cos2(θ)κa2κg)12P1κg14P2rsin(2θ)κa258P1κg21116P1κg314P1[cos(2θ)3]κa2κg18P1rsin(2θ)(2κa2+κa2κg)sin2(θ)(rκaκg12κa2κg+rκaκg2)),(S_{\beta})\approx\begin{pmatrix}\mathcal{E}\Bigl{(}-\frac{1}{2}\kappa_{g}-\frac{3}{8}\kappa_{g}^{2}-\frac{5}{16}\kappa_{g}^{3}+\frac{1}{2}\cos^{2}(\theta)\kappa_{a}^{2}\kappa_{g}\Bigr{)}\\ -\frac{1}{2}P^{1}\kappa_{g}-\frac{1}{4}P^{2}r\sin(2\theta)\kappa_{a}^{2}-\frac{5}{8}P^{1}\kappa_{g}^{2}-\frac{11}{16}P^{1}\kappa_{g}^{3}-\frac{1}{4}P^{1}[\cos(2\theta)-3]\kappa_{a}^{2}\kappa_{g}\\ \frac{1}{8}P^{1}r\sin(2\theta)(2\kappa_{a}^{2}+\kappa_{a}^{2}\kappa_{g})\\ \sin^{2}(\theta)\Bigl{(}\mathcal{E}r\kappa_{a}\kappa_{g}-\frac{1}{2}\mathcal{L}\kappa_{a}^{2}\kappa_{g}+\mathcal{E}r\kappa_{a}\kappa_{g}^{2}\Bigr{)}\end{pmatrix}, (83)

where P1P^{1} and P2P^{2} are given by the following equations [17],

(u1)2=E2R(r)ρ4,(u2)2=E2Θ(θ)ρ4,(u^{1})^{2}=E^{2}\frac{R(r)}{\rho^{4}},\qquad(u^{2})^{2}=E^{2}\frac{\Theta(\theta)}{\rho^{4}}, (84)

and R(r)R(r) and Θ(θ)\Theta(\theta) are defined as777Note that the forms of R(r)R(r) and Θ(θ)\Theta(\theta) in (84) are different from those in (185) and (186) of Chapter 7 of Ref. [17].

R(r)\displaystyle R(r) =\displaystyle= r4+r2(a2λ2η0)+2Mr[(aλ)2+η0]a2η0δ1r2ΔE2,\displaystyle r^{4}+r^{2}(a^{2}-\lambda^{2}-\eta_{0})+2Mr[(a-\lambda)^{2}+\eta_{0}]-a^{2}\eta_{0}-\delta_{1}\frac{r^{2}\Delta}{E^{2}}, (85)
Θ(θ)\displaystyle\Theta(\theta) =\displaystyle= η0+a2cos2θλ2cot2θδ1a2cos2θE2,\displaystyle\eta_{0}+a^{2}\cos^{2}\theta-\lambda^{2}\cot^{2}\theta-\delta_{1}\frac{a^{2}\cos^{2}\theta}{E^{2}}, (86)

and the parameter δ1\delta_{1} is defined by

δ1={1,for massive particles,0,for massless particles,\delta_{1}=\begin{cases}1,\quad\text{for massive particles,}\\ 0,\quad\text{for massless particles,}\end{cases} (87)

λ=L/E\lambda=L/E, η0=/E2\eta_{0}=\mathscr{L}/E^{2}, and \mathscr{L} is a separation constant in the equations of motion. Then expanding P1P^{1} and P2P^{2}, and plugging them into (83), we can find that κa\kappa_{a} only appears in the second and higher order terms of SβS_{\beta}. On the other hand, the local gravitational phase can be written as ϕSβδxβ\phi\approx S_{\beta}\delta x^{\beta}. Therefore, the quantity aa does not appear in the first order terms of the local gravitational phase.

This conclusion is also true for the gravitational phase difference, as (75) shows. Therefore, the contribution of the quantity aa can be regarded as a small modification to the case of the Schwarzschild spacetime. Theoretically, we can measure the fringe shift between different values of the angle θ1\theta_{1} to detect the contribution of aa. However, this is not an economic way because we need to move the setup significantly. An alternative way is to rotate the parallelogram along the axis ll shown in FIG. 2. For simplicity, we flip it so that the positions of A and B swap. Correspondingly, the second equation in (73) should be changed to

φABsΓ33,\varphi_{AB}\approx-\frac{s}{\sqrt{\Gamma_{33}}}, (88)

while the expression for tABt_{AB} is not changed. Besides, we need to make the change φDCφDC\varphi_{DC}\rightarrow-\varphi_{DC} and reverse the angular momentum such that \mathcal{L}\rightarrow-\mathcal{L}. Plugging these changes into (149), (150) and (151) in the Appendix B, we can derive a new phase difference. Let us denote it as (δφ)2(\delta\varphi)_{2}, then the fringe shift for the rotation is

n\displaystyle n =\displaystyle= |(δϕ)2δϕ2π|\displaystyle\biggl{|}\frac{(\delta\phi)_{2}-\delta\phi}{2\pi}\biggr{|} (89)
=\displaystyle= |0lsπr1[argr12(2sin(γ)cos(θ1)+cos(γ)sin(θ1))+arg2r13sin(γ+θ1)]|,\displaystyle\biggl{|}\frac{\mathcal{E}_{0}ls}{\pi\hbar r_{1}}\Bigl{[}\frac{ar_{g}}{r_{1}^{2}}\Bigl{(}2\sin(\gamma)\cos(\theta_{1})+\cos(\gamma)\sin(\theta_{1})\Bigr{)}+\frac{ar_{g}^{2}}{r_{1}^{3}}\sin(\gamma+\theta_{1})\Bigr{]}\biggr{|},

where δϕ\delta\phi is given in (75) and 0\mathcal{E}_{0} is defined in (76). The quantity nn shows the impact of the angular momentum of the black hole on the interference. In (89) we can find that the fringe shifts vanish for a=0a=0. This is not surprising because flipping the parallelogram along the axis ll does not affect the result of the interference in a Schwarzschild spacetime due to spherical symmetry.

4.3 Numerical results and discussion

For simplicity, we here assume γ=0\gamma=0, we restore the SI units, and use the spin parameter

a=ac2GM=2arg.a_{*}=\frac{ac^{2}}{GM}=\frac{2a}{r_{g}}. (90)

As we showed in (75) and (162), the shape of the parallelogram only makes difference at the third order and higher orders in δϕ\delta\phi. Therefore, in the case γ=0\gamma=0, for simplicity we assume that the path BC is along the radius, such that vr=vv^{r}=v and vθ=0v^{\theta}=0 hold in the phase difference (75).

4.3.1 Massive particles

In the case of non-relativistic massive particles, the phase difference (79) is:

δϕmass(γ=0)δϕ1+δϕ2+δϕ3,\delta\phi_{\rm mass}(\gamma=0)\approx\delta\phi_{1}+\delta\phi_{2}+\delta\phi_{3}, (91)

where δϕ1\delta\phi_{1}, δϕ2\delta\phi_{2}, and δϕ3\delta\phi_{3} are the first, second, and third order terms respectively, namely

δϕ1\displaystyle\delta\phi_{1} =\displaystyle= mcls2r1(mcλd2π+πmcλd)rgr1,δϕ2=mcls2r1[(mcλd2π+πmcλd)asin(θ1)]rg2r12,\displaystyle\frac{mcls}{2\hbar r_{1}}\Bigl{(}\frac{mc\lambda_{d}}{2\pi\hbar}+\frac{\pi\hbar}{mc\lambda_{d}}\Bigr{)}\frac{r_{g}}{r_{1}},\qquad\delta\phi_{2}=\frac{mcls}{2\hbar r_{1}}\Bigl{[}\Bigl{(}\frac{mc\lambda_{d}}{2\pi\hbar}+\frac{\pi\hbar}{mc\lambda_{d}}\Bigr{)}-a_{*}\sin(\theta_{1})\Bigr{]}\frac{r_{g}^{2}}{r_{1}^{2}}, (92)
δϕ3\displaystyle\delta\phi_{3} =\displaystyle= mcls2r1{(mcλd2π+πmcλd)32asin(θ1)+12a2[14(17cos2(θ1))(mcλd2π+πmcλd)\displaystyle\frac{mcls}{2\hbar r_{1}}\Bigl{\{}\Bigl{(}\frac{mc\lambda_{d}}{2\pi\hbar}+\frac{\pi\hbar}{mc\lambda_{d}}\Bigr{)}-\frac{3}{2}a_{*}\sin(\theta_{1})+\frac{1}{2}a_{*}^{2}\Bigl{[}\frac{1}{4}\Bigl{(}1-7\cos^{2}(\theta_{1})\Bigr{)}\Bigl{(}\frac{mc\lambda_{d}}{2\pi\hbar}+\frac{\pi\hbar}{mc\lambda_{d}}\Bigr{)} (93)
+3πmcλdsin2(θ1)]}rg3r13.\displaystyle+\frac{3\pi\hbar}{mc\lambda_{d}}\sin^{2}(\theta_{1})\Bigr{]}\Bigr{\}}\frac{r_{g}^{3}}{r_{1}^{3}}.

We can find that the phase difference is proportional to the area of the parallelogram. The fringe shift (89), corresponding to flipping the parallelogram along the axis ll, is:

nmass|mcls2πr1asin(θ1)(rg2r12+rg3r13)|,n_{\rm mass}\approx\biggl{|}\frac{mcls}{2\pi\hbar r_{1}}a_{*}\sin(\theta_{1})\Bigl{(}\frac{r_{g}^{2}}{r_{1}^{2}}+\frac{r_{g}^{3}}{r_{1}^{3}}\Bigr{)}\biggr{|}, (94)

where we have let γ=0\gamma=0 and neglected the second and higher order terms in v/cv/c, and we remind that the Schwarzschild radius is given by rg=2GM/c2r_{g}=2GM/c^{2}. From (75) we can find δϕ(γ=π)=δϕ(γ=0)\delta\phi(\gamma=\pi)=-\delta\phi(\gamma=0). Therefore, if we change the angle γ\gamma from 0 to π\pi, we get a fringe shift

Nmass=|δϕmass(γ=π)δϕmass(γ=0)2π|=|δϕmass(γ=0)π|.N_{\rm mass}=\Bigl{|}\frac{\delta\phi_{\rm mass}(\gamma=\pi)-\delta\phi_{\rm mass}(\gamma=0)}{2\pi}\Bigr{|}=\Bigl{|}\frac{\delta\phi_{\rm mass}(\gamma=0)}{\pi}\Bigr{|}. (95)

In the following we discuss two examples in which the particles that interfere are neutrons.

(I) The earth as the gravitational source. In this example, we neglect the spin of the earth so that a0a_{*}\approx 0. For the parameter r1r_{1}, we assume the equatorial radius of the earth. For the setup of the experiment, we take the parameters in [10], i.e.

ls=6×104m2,λd=1.42×1010m.ls=6\times 10^{-4}{\rm m}^{2},\qquad\lambda_{d}=1.42\times 10^{-10}{\rm m}. (96)

And for the constants in (92) and (93), we use the values given in [20]. Therefore, we get the results:888Equations (92) and (93) are used here even though (68) is violated, because (69) still holds for the case a=0a=0.

δϕ1=33.502,δϕ2=4.660×108,δϕ3=6.482×1017.\delta\phi_{1}=33.502,\qquad\delta\phi_{2}=4.660\times 10^{-8},\qquad\delta\phi_{3}=6.482\times 10^{-17}. (97)

We can find that the second and the third order terms are much smaller than the first order term. As for the fringe according to (94) we get:

nmass=0n_{\rm mass}=0 (98)

because of a0a_{*}\approx 0. Combining (97), (91), and (95), we get:

Nmass=10.664.N_{\rm mass}=10.664. (99)

The value (99) is nearly the same as the result in [10], which agrees with the claim that the equation (75) produces the result (66) on the earth in the Newtonian limit.

(II) The black hole in Cygnus X-1 as the gravitational source. We take the distance between the black hole and the earth to be the value of r1r_{1}. The parameters are given by [21, 22]999Here we do not consider the uncertainties shown in the references [21, 22]. Moreover, in such papers the authors do not give the angle θ1\theta_{1} directly, but give the binary orbital inclination i=27.51i=27.51^{\circ}. However, as stated in Ref. [22], the spin axis of the black hole is assumed to be aligned with the orbital angular momentum. Therefore, the angle θ1\theta_{1} is equal to the inclination ii.

M=21.2M,a>0.9985,r1=2.22kpc,θ1=27.51,M=21.2M_{\odot},\qquad a_{*}>0.9985,\qquad r_{1}=2.22{\rm kpc},\qquad\theta_{1}=27.51^{\circ}, (100)

where MM_{\odot} is the mass of the sun. For simplicity, we assume the value a=0.9985a_{*}=0.9985. For the area of the parallelogram and the wavelength of the neutron, we still use the parameters (96). Thus we get:

δϕ1=2.049×1018,δϕ2=1.873×1033,δϕ3=7.506×1049.\delta\phi_{1}=2.049\times 10^{-18},\qquad\delta\phi_{2}=1.873\times 10^{-33},\qquad\delta\phi_{3}=7.506\times 10^{-49}. (101)

The gravitational phase difference is totally dominated by the first order term. For the fringe shift we get:

nmass=2.557×1039,n_{\rm mass}=2.557\times 10^{-39}, (102)

when the parallelogram is flipped along the axis ll. Similar to (99), we get the fringe shift

Nmass=6.524×1019,N_{\rm mass}=6.524\times 10^{-19}, (103)

corresponding to changing the angle γ\gamma from 0 to π\pi, which is much smaller than the fringe shift in the example (I).

4.3.2 Massless particles

Similar to (91), according to (75) for massless particles we have:

δϕmassless(γ=0)δϕ1+δϕ2+δϕ3,\delta\phi_{\rm massless}(\gamma=0)\approx\delta\phi_{1}+\delta\phi_{2}+\delta\phi_{3}, (104)

where

δϕ1=πlsλ0r1rgr1,δϕ2=πlsλ0r1(1asin(θ1))rg2r12,δϕ3=πlsλ0r1[132asin(θ1)+a216(113cos(2θ1))]rg3r13.\delta\phi_{1}=\frac{\pi ls}{\lambda_{0}r_{1}}\frac{r_{g}}{r_{1}},\quad\delta\phi_{2}=\frac{\pi ls}{\lambda_{0}r_{1}}\Bigl{(}1-a_{*}\sin(\theta_{1})\Bigr{)}\frac{r_{g}^{2}}{r_{1}^{2}},\quad\delta\phi_{3}=\frac{\pi ls}{\lambda_{0}r_{1}}\Bigl{[}1-\frac{3}{2}a_{*}\sin(\theta_{1})+\frac{a_{*}^{2}}{16}\Bigl{(}1-13\cos(2\theta_{1})\Bigr{)}\Bigr{]}\frac{r_{g}^{3}}{r_{1}^{3}}. (105)

And the equation (89) is simplified to

nmassless=|lsλ0r1sin(θ1)a(rg2r12+rg3r13)|.n_{\rm massless}=\biggl{|}\frac{ls}{\lambda_{0}r_{1}}\sin(\theta_{1})a_{*}\Bigl{(}\frac{r_{g}^{2}}{r_{1}^{2}}+\frac{r_{g}^{3}}{r_{1}^{3}}\Bigr{)}\biggr{|}. (106)

From (105) we can find the phase difference is proportional to the area of the parallelogram and inversely proportional to the wavelength λ0\lambda_{0} of the particles. As an example, we consider gamma rays and adopt the parameters

ls=6×104m2,λ0=1012m.ls=6\times 10^{-4}{\rm m}^{2},\qquad\lambda_{0}=10^{-12}{\rm m}. (107)

Then we repeat the computations in (I) and (II).

For the example (I) we get:

δϕ1=4.111×107,δϕ2=5.719×1016,δϕ3=7.955×1025,\displaystyle\delta\phi_{1}=4.111\times 10^{-7},\qquad\delta\phi_{2}=5.719\times 10^{-16},\qquad\delta\phi_{3}=7.955\times 10^{-25},
nmassless=0,Nmassless=1.309×107.\displaystyle n_{\rm massless}=0,\qquad N_{\rm massless}=1.309\times 10^{-7}. (108)

Note that the definition for the fringe shift NmasslessN_{\rm massless} is similar to (95).

For the example (II) we get:

δϕ1=2.515×1026,δϕ2=1.239×1041,δϕ3=1.973×1057,\displaystyle\delta\phi_{1}=2.515\times 10^{-26},\qquad\delta\phi_{2}=1.239\times 10^{-41},\qquad\delta\phi_{3}=-1.973\times 10^{-57},
nmassless=3.375×1042,Nmassless=8.005×1027.\displaystyle n_{\rm massless}=3.375\times 10^{-42},\qquad N_{\rm massless}=8.005\times 10^{-27}. (109)

4.3.3 Discussion

Comparing the results in the example (I) with those in the example (II), we find that NINIIN_{\rm I}\gg N_{\rm II} holds for both massive and massless particles, where the subscripts denote the two examples. Such inequality is explained by N|δϕ1/π|rg/r12N\approx|\delta\phi_{1}/\pi|\propto r_{g}/r_{1}^{2} and (rg/r12)I(rg/r12)II(r_{g}/r_{1}^{2})_{\rm I}\gg(r_{g}/r_{1}^{2})_{\rm II}. Therefore, if we want to increase the fringe shift NN, we can increase the ratio rg/r12r_{g}/r_{1}^{2}. For example, to let Nmass1N_{\rm mass}\approx 1 in (II), we can decrease the distance to be r15.533×1010m1.793×109kpcr_{1}\approx 5.533\times 10^{10}{\rm m}\approx 1.793\times 10^{-9}{\rm kpc} which is much less than the distance 2.22kpc2.22{\rm kpc} in (100) but still satisfies the condition r1rg6.262×104mr_{1}\gg r_{g}\approx 6.262\times 10^{4}{\rm m}. In order to increase NN we can also increase the area of the parallelogram, according to (92) and (105). Moreover, for this purpose, in the massive case we can use more massive or slower particles. While for the massless case we can use more energetic particles to increase NN. As for nn, to increase its value, we can increase the ratio rg2/r13r_{g}^{2}/r_{1}^{3}, the area of the parallelogram, the spin parameter, or the quantity sin(θ1)\sin(\theta_{1}), according to (94) and (106). For example, in (II) we can decrease the distance to be r19.368×106mr_{1}\approx 9.368\times 10^{6}{\rm m} to let nmass1n_{\rm mass}\approx 1. Furthermore, we can use more massive particles or more energetic massless particles to increase nn.

Now we compare the massive case with the massless case. Comparing the values of NN in (99) and (103) with those in (108) and (109) respectively, we can find that NmassN_{\rm mass} is much greater than NmasslessN_{\rm massless}, although a very small value for λ0\lambda_{0} is chosen. Moreover, comparing the value of nn in (102) with its value in (109), we can find nmassnmasslessn_{\rm mass}\gg n_{\rm massless}. Therefore, we conclude that it is more difficult to detect the fringe shifts for massless particles in comparison with the massive case.

Finally, comparing nmassn_{\rm mass} with NmassN_{\rm mass} and comparing nmasslessn_{\rm massless} with NmasslessN_{\rm massless} in these examples, we find nNn\ll N for both cases. This is because N|δϕ1/π|N\approx|\delta\phi_{1}/\pi| is dominated by the first order terms according to (95), while all the terms of nn have orders higher than one according to (94) and (106). Therefore, it is easier to detect NN than to detect nn. Additionally, according to (92) and (105), the phase difference δϕ1\delta\phi_{1} depends on rgr_{g}, and, according to (94) and (106), the fringe shift nn depends on both aa_{*} and rgr_{g}. Therefore, inversely we can determine the mass of the black hole and its spin parameter according to the measured fringe shifts, following the following steps: First we should measure the fringe shift NN to determine the mass MM. For massive particles it is determined by

M=2π2r122c2λdGls(m2c2λd2+2π22)Nmass,M=\frac{2\pi^{2}r_{1}^{2}\hbar^{2}c^{2}\lambda_{d}}{Gls(m^{2}c^{2}\lambda_{d}^{2}+2\pi^{2}\hbar^{2})}N_{\rm mass}, (110)

while for massless particles it is determined by

M=c2λ0r122GlsNmassless,M=\frac{c^{2}\lambda_{0}r_{1}^{2}}{2Gls}N_{\rm massless}, (111)

then we should measure the fringe shift nn, from which, given the mass MM, one can determine the spin parameter aa_{*} as follows. For simplicity we only keep up to second order terms in (94) and (106). Hence, plugging (110) and (111) into these equations, we can determine the spin parameter aa_{*}. For massive particles it is:

a=ls(m2c2λd2+2π22)28π33sin(θ1)r1mcλd2nmassNmass2,a_{*}=\frac{ls(m^{2}c^{2}\lambda_{d}^{2}+2\pi^{2}\hbar^{2})^{2}}{8\pi^{3}\hbar^{3}\sin(\theta_{1})r_{1}mc\lambda_{d}^{2}}\frac{n_{\rm mass}}{N_{\rm mass}^{2}}, (112)

while for massless particles it is:

a=lssin(θ1)λ0r1nmasslessNmassless2.a_{*}=\frac{ls}{\sin(\theta_{1})\lambda_{0}r_{1}}\frac{n_{\rm massless}}{N_{\rm massless}^{2}}. (113)

5 Conclusion

The gravitational phase difference has been expressed as the integral of a function SβS_{\beta} defined by the product of the four-momentum, the metric, and the gauge gravitational potential, which is expressed by the tetrad. This is the way to calculate the gravitational phase for a general given spacetime. However, as an explicit example, in this paper we considered the case of the Kerr spacetime and we studied a particles’ interference experiment (FIG. 2) analogous to the COW experiment, but in the Kerr spacetime.

We calculated the phase difference for massive and massless particles respectively. We found that the angular momentum of the black hole only appears in the second or higher order terms in the phase difference. As a generalization, we have proved that the angular momentum density aa does not appear in the first order terms of the local gravitational phase at large distance respect to the Schwarzschild radius, namely for rrgr\gg r_{g}. Then we have evaluated the fringe shifts for several examples, compared the results, and discussed how to increase the fringe shifts. Concretely, in order to increase the fringe shifts, we should take a larger black hole’s mass, decrease the distance from it, or increase the area of the parallelogram. For this purpose, we could also choose more massive and slower particles or more energetic massless particles. According to the numerical results, we found that it is more difficult to measure the fringe shifts for massless particles than those for massive particles. In the end, we showed how to determine the mass of the black hole and its spin parameter by the measurement of the fringe shifts.

We here propose some potential extensions of our work. Besides the interference with paths along a parallelogram we could consider other configurations, or, in addition to the asymptotically flat region, we could consider the region closer to the black hole gravitational radius. Additionally, the numerical examples should not be limited to the black hole in Cygnus X-1, but other examples should be discussed in the future. Finally, considering the universality of the gravitational phase (42), we could apply it to other spacetimes. For example, we could consider other compact objects such as binary black holes, neutron stars, and rotating galaxies or black holes beyond Einstein’s theory of gravity [23, 26, 24, 25].

Finally, recalling that the phases derived in this paper are based on Teleparallel Gravity, it is essential to compare our results with those in general relativity. Even though the equation of motion in the former is equivalent to the one in the latter (see Sec. 2.2), the quantum aspects of these theories are not necessary the same. Now let us compare the non-relativistic phase difference (79) with the one obtained in general relativity. In Ref. [27], the authors calculated the phase difference of a quantum interferometer experiments on the earth, with the rotation of the earth taken into account. In the weak field limit and up to the first order in the post-Newtonian approximation, they found

δϕPN\displaystyle\delta\phi^{\rm PN} =\displaystyle= m2gAλ2π2sinμ+2mωA+2m5rgR[ω3(RRω)RR]A\displaystyle\frac{m^{2}gA\lambda}{2\pi\hbar^{2}}\sin\mu+\frac{2m}{\hbar}\vec{\omega}\cdot\vec{A}+\frac{2m}{5\hbar}\frac{r_{g}}{R}\Bigl{[}\vec{\omega}-3\Bigl{(}\frac{\vec{R}}{R}\cdot\vec{\omega}\Bigr{)}\frac{\vec{R}}{R}\Bigr{]}\cdot\vec{A} (114)
12rgR(m2gAλ2π2sinμ)+32(λCλ)2(m2gAλ2π2sinμ),\displaystyle-\frac{1}{2}\frac{r_{g}}{R}\Bigl{(}\frac{m^{2}gA\lambda}{2\pi\hbar^{2}}\sin\mu\Bigr{)}+\frac{3}{2}\Big{(}\frac{\lambda_{C}}{\lambda}\Bigr{)}^{2}\Bigl{(}\frac{m^{2}gA\lambda}{2\pi\hbar^{2}}\sin\mu\Bigr{)},

where gg is the gravitational acceleration, A\vec{A} is the area vector enclosed by the interferometry loop, μ\mu is the angle between A\vec{A} and the position vector R\vec{R} of the interferometer, RR is the radius of the earth, λ\lambda and λC\lambda_{C} are the de Broglie wavelength and the Compton wavelength respectively, and ω\vec{\omega} is the angular velocity vector of the earth with its magnitude related with the Kerr parameter by a=2R2ω/5a=2R^{2}\omega/5. For simplicity, we use δαj\delta\alpha_{j} to denote the jthj^{\rm th} term in (114), where j=1,2,,5j=1,2,...,5. According to Ref. [27], these terms are interpreted as follows: The first term δα1\delta\alpha_{1} is just the result predicted in Ref. [10], verified by the COW experiment [11]; the term δα2\delta\alpha_{2} due to Sagnac effect [28, 29] is caused by the rotation of the interferometer (recall that this experiment is on the earth); the term δα3\delta\alpha_{3} is due to the Lense-Thirring effect [30]; finally, the terms δα4\delta\alpha_{4} and δα5\delta\alpha_{5} correspond to the redshift corrections to the potential energy and the kinetic energy respectively. To compare the result (79) with (114), we need to rewrite the latter according to the parameters in our result. Notice that the second term in (114) is absent here, namely δα2=0\delta\alpha_{2}=0, because in FIG. 2 the interferometer is assumed to be not rotating. Then according to the relation between aa and ω\omega, and the equations:

g=rg2R2,μ=π2γ,Rω=Rωcosθ,RA=RAcos(μ),ωA=ωAcos(θ+μ),g=\frac{r_{g}}{2R^{2}},\qquad\mu=\frac{\pi}{2}-\gamma^{\prime},\qquad\vec{R}\cdot\vec{\omega}=R\omega\cos\theta^{\prime},\qquad\vec{R}\cdot\vec{A}=RA\cos(\mu),\qquad\vec{\omega}\cdot\vec{A}=\omega A\cos(\theta^{\prime}+\mu), (115)

where γ\gamma^{\prime} is the angle between R\vec{R} and the plane of the interferometry, we can rewrite the remaining terms in (114) as follows:

δα1=mAcos(γ)2R1vrgR,δα3=mAR(2cos(θ)sin(γ)+cos(γ)sin(θ))argR2,\displaystyle\delta\alpha_{1}=\frac{mA\cos(\gamma^{\prime})}{2\hbar R}\frac{1}{v}\frac{r_{g}}{R},\qquad\delta\alpha_{3}=-\frac{mA}{\hbar R}\Bigl{(}2\cos(\theta^{\prime})\sin(\gamma^{\prime})+\cos(\gamma^{\prime})\sin(\theta^{\prime})\Bigr{)}\frac{ar_{g}}{R^{2}},
δα4=mAcos(γ)4R1vrg2R2,δα5=3mAcos(γ)4RrgvR,\displaystyle\delta\alpha_{4}=-\frac{mA\cos(\gamma^{\prime})}{4\hbar R}\frac{1}{v}\frac{r_{g}^{2}}{R^{2}},\qquad\delta\alpha_{5}=\frac{3mA\cos(\gamma^{\prime})}{4\hbar R}\frac{r_{g}v}{R}, (116)

where the units c=1c=1 has been used. To compare our result with (116), we neglect the terms of the third order such that the phase difference (79) reduces to

δϕNR\displaystyle\delta\phi^{\rm NR} =\displaystyle= mlscos(γ)2r11vrgr1mlsr1(2cos(θ1)sin(γ)+cos(γ)sin(θ1))argr12\displaystyle\frac{mls\cos(\gamma)}{2\hbar r_{1}}\frac{1}{v}\frac{r_{g}}{r_{1}}-\frac{mls}{\hbar r_{1}}\Bigl{(}2\cos(\theta_{1})\sin(\gamma)+\cos(\gamma)\sin(\theta_{1})\Bigr{)}\frac{ar_{g}}{r_{1}^{2}} (117)
+mlscos(γ)2r11vrg2r12+mlscos(γ)4r1rgvr1+mlscos(γ)4r11vrg2v2r12.\displaystyle+\frac{mls\cos(\gamma)}{2\hbar r_{1}}\frac{1}{v}\frac{r_{g}^{2}}{r_{1}^{2}}+\frac{mls\cos(\gamma)}{4\hbar r_{1}}\frac{r_{g}v}{r_{1}}+\frac{mls\cos(\gamma)}{4\hbar r_{1}}\frac{1}{v}\frac{r_{g}^{2}v^{2}}{r_{1}^{2}}.

The first two terms in (117) coincide with those in (116), while the third and fourth terms are different from those of (116) only in the coefficients, and the last term in (117) can be neglected compared with other terms because (rgv/r1)2(r_{g}v/r_{1})^{2} is very small. We notice that the area AA in Ref. [27] is defined in flat space (see (3.20) in Ref. [27]), while in this paper the area is defined by the length in curved spacetime (see (120)). However, even using the later, the form of the phase differences in (116) are not changed (see the last paragraph in Appendix A), such that the above conclusions do not change. Finally, we mention that the first two terms in (117) also coincide with the result in Ref. [31], where the authors study the same experiment on the earth. They use an approximation to the first order of MM and aa, and neglect the terms of O(v2)O(v^{2}).

Therefore, in the weak field limit, the non-relativistic phase difference (79) based on the theory of Teleparallel Gravity, reproduces partly the result of post-Newtonian approximation in general relativity. In particular, reproduces the result in the COW experiment and the term of the Lense-Thirring effect. It looks a little strange that the predictions from Teleparallel Gravity in the interference experiment are not exactly the same as those from general relativity. Indeed, consider that the equation of motions for a particle in Teleparallel Gravity is identical to the geodesic equation in general relativity (as mentioned in Sec. 2.2). We have to admit that we do not know how to explain such difference, and we simply notice that the derivations for the phase in this paper and in Ref. [27] are different. In Ref. [27], the phase is found by constructing the quantum Hamiltonian of a non-relativistic particle in the weak gravitational field up to the first order of the post-Newtonian approximation, and plugging the Hamiltonian into the Schrödinger equation. While in our paper, following Ref. [9], the phase is constructed by separating a gauge potential BμaB{{}^{a}}_{\mu} analogous to the electromagnetic potential from the Lagrangian of the particle (see the last paragraph in Sec. 2.2). Given that we have not provided a wave equation satisfied by (4), this phase is a conjecture to some extent. In spite of this, we think that it is reasonable in the perspective of the analogy with electromagnetism, and since it successfully reproduces the result of COW experiment. However, the gravitational phase in Teleparallel Gravity deserves more investigations before to be able to help us revealing the differences between Teleparall Gravity and general relativity in such quantum aspects101010As suggested by the referee of this paper, if the Aharonov-Bohm effect is sensitive to the potential Baμ{B^{a}}_{\mu}, it could provide a possible way to distinguish general relativity from Teleparallel Gravity..

Acknowledgements.
This work was supported by the Basic Research Program of the Science, Technology, and Innovation Commission of Shenzhen Municipality (grant no. JCYJ20180302174206969).

Appendix A Derivations for some formulas

In this appendix we derive some equations used in Sec. 4. Let us prove (70) firstly. According to (62) and recalling g30=g03g_{30}=g_{03} in Kerr spacetime, we get

=(g03dtdφ+g33)(g00dtdφ+g03)1.\mathcal{L}=-\mathcal{E}\Bigl{(}g_{03}\frac{\mbox{d}t}{\mbox{d}\varphi}+g_{33}\Bigr{)}\Bigl{(}g_{00}\frac{\mbox{d}t}{\mbox{d}\varphi}+g_{03}\Bigr{)}^{-1}. (118)

The expression for dt/dφ\mbox{d}t/\mbox{d}\varphi is found by letting k=3k=3 in the definition of the velocity (71), namely

dtdφ=1vφg00g03g00.\frac{\mbox{d}t}{\mbox{d}\varphi}=\frac{1}{v^{\varphi}\sqrt{g_{00}}}-\frac{g_{03}}{g_{00}}. (119)

Hence, replacing (119) into (118), we prove (70).

Now we derive (72). As shown in [18], in a spacetime with its metric independent on the time coordinate, a distance LL is defined as the integral of the distance element dL\mbox{d}L given by:

dL2=Γijdxidxj,\mbox{d}L^{2}=\Gamma_{ij}\mbox{d}x^{i}\mbox{d}x^{j}, (120)

where the three-dimensional metric tensor Γij\Gamma_{ij} is defined in (71). Applying (120) to the radial component of ll and the component perpendicular to the radius (see FIG. 2), we find:

lcos(γ)=Γ11drg11(r2r1),\displaystyle l\cos(\gamma)=\sqrt{\Gamma_{11}}\mbox{d}r\approx\sqrt{-g_{11}}(r_{2}-r_{1}), (121)
lsin(γ)=|Γ22dθ|{g22(θ1θ2),when lsin(γ)g221θ1,g22(θ1+θ2)when lsin(γ)g221>θ1,\displaystyle l\sin(\gamma)=\bigl{|}\sqrt{\Gamma_{22}}\mbox{d}\theta\bigr{|}\approx\begin{cases}\sqrt{-g_{22}}(\theta_{1}-\theta_{2}),\quad&\text{when $l\sin(\gamma)\sqrt{-g_{22}^{-1}}\leq\theta_{1}$,}\\ \sqrt{-g_{22}}(\theta_{1}+\theta_{2})\quad&\text{when $l\sin(\gamma)\sqrt{-g_{22}^{-1}}>\theta_{1}$,}\end{cases} (122)

which imply the two equations in (72) respectively.

Now we show how to derive (73). We know that dr0\mbox{d}r\approx 0 and dθ0\mbox{d}\theta\approx 0 on the path AB (assumption (a) in Sec. 4.1), therefore, using (120) to this path, we have the following relation:

dφdLΓ33.\mbox{d}\varphi\approx\frac{\mbox{d}L}{\sqrt{\Gamma_{33}}}. (123)

Hence, combining (123) with (119), we find the relation between time coordinate and length, i.e.,

dt(1vφg00g03g00)dLΓ33.\mbox{d}t\approx\Bigl{(}\frac{1}{v^{\varphi}\sqrt{g_{00}}}-\frac{g_{03}}{g_{00}}\Bigr{)}\frac{\mbox{d}L}{\sqrt{\Gamma_{33}}}. (124)

Integrating (123) and (124), and taking (155) into account, we derive the two equations in (73).

As for the energy of a massive particle in (74), we take directly the result from Ref. [18] (see Sec. 88 in [18] ). While for a massless particle, its energy reads [19]:

=Vω,\mathcal{E}=V\hbar\omega, (125)

where ω\omega is its frequency measured by a static observer, while VV is the redshift factor given by:111111In the metric signature (,+,+,+)(-,+,+,+) the redshift factor is replaced by V=KμKμV=\sqrt{-K_{\mu}K^{\mu}}.

V=KμKμ,V=\sqrt{K_{\mu}K^{\mu}}, (126)

where KμK^{\mu} is the Killing vector related to the time-translation invariance. Here by a static observer we mean that the four-velocity of the observer is proportional to the Killing vector [19]. Inserting the Killing vector Kμ=(1,0,0,0)K^{\mu}=(1,0,0,0) into (126), we obtain V=g00V=\sqrt{g_{00}}. Finally, plugging this result into (125), we obtain the second expression in (74).

Now we prove the statement in Sec. 5 that the forms of the terms in (116) do not change when we use the area defined in the Kerr spacetime to re-express them. Firstly, the metric (2.1) in Ref. [27] can be rewritten as121212We believe that there is a typo in the last term of g00g_{00} in (2.1) of Ref. [27]. Here we have made a modification.

ds2=[1+2Φ+2Φ2(rgar2)2sin2(θ)]dt2+2rgarsin2(θ)dφdt(12Φ)(dr2+r2dθ2+r2sin2(θ)dφ2),\mbox{d}s^{2}=\Bigl{[}1+2\Phi+2\Phi^{2}-\Bigl{(}\frac{r_{g}a}{r^{\prime 2}}\Bigr{)}^{2}\sin^{2}(\theta^{\prime})\Bigr{]}\mbox{d}t^{\prime 2}+\frac{2r_{g}a}{r^{\prime}}\sin^{2}(\theta^{\prime})\mbox{d}\varphi^{\prime}\mbox{d}t^{\prime}-(1-2\Phi)\bigl{(}\mbox{d}r^{\prime 2}+r^{\prime 2}\mbox{d}\theta^{\prime 2}+r^{\prime 2}\sin^{2}(\theta^{\prime})\mbox{d}\varphi^{\prime 2}\bigr{)}, (127)

where Φ=rg/(2r)\Phi=-r_{g}/(2r^{\prime}) is the Newtonian potential, and the coordinates (t,r,θ,φ)(t^{\prime},r^{\prime},\theta^{\prime},\varphi^{\prime}) relate with the asymptotically static coordinates (t,x,y,z)(t,x^{\prime},y^{\prime},z^{\prime}) by

t=t,x=rsin(θ)cos(φ),y=rsin(θ)sin(φ),z=rcos(θ).t^{\prime}=t,\qquad x^{\prime}=r^{\prime}\sin(\theta^{\prime})\cos(\varphi^{\prime}),\qquad y^{\prime}=r^{\prime}\sin(\theta^{\prime})\sin(\varphi^{\prime}),\qquad z^{\prime}=r^{\prime}\cos(\theta^{\prime}). (128)

The terms in (116) are derived by the following integral [27]

δαj=1Hjdt,\delta\alpha_{j}=-\frac{1}{\hbar}\oint H_{j}\mbox{d}t, (129)

where the loop encloses the interferometer, and HjH_{j} are defined by

H1=mΦ,H3=2rgR25r3ωJ,H4=m2Φ2,H5=3Φp22m,H_{1}=m\Phi,\qquad H_{3}=\frac{2r_{g}R^{2}}{5r^{\prime 3}}\vec{\omega}\cdot\vec{J},\qquad H_{4}=\frac{m}{2}\Phi^{2},\qquad H_{5}=\frac{3\Phi\vec{p}\,^{2}}{2m}, (130)

where J=r×p\vec{J}=\vec{r}\times\vec{p} is the angular momentum defined in flat space, and p=mv\vec{p}=m\vec{v}. For simplicity, assume that the loop of the interferometer is a parallelogram. As we mentioned in Sec. 5, the area AA in (116) is defined in flat space [27]. If we take a new area defined by the length in the Kerr spacetime (see (120)), the term δα1\delta\alpha_{1} now reads

δα1=mrg2(1r21r1)tAB=mrg2(1r1+lcos(γ)/g111r1)s(1vg00g03g00Γ33)mlscos(γ)2r11vrgr1,\delta\alpha_{1}=-\frac{mr_{g}}{2\hbar}\Bigl{(}\frac{1}{r^{\prime}_{2}}-\frac{1}{r^{\prime}_{1}}\Bigr{)}t_{AB}=-\frac{mr_{g}}{2\hbar}\Bigl{(}\frac{1}{r^{\prime}_{1}+l\cos(\gamma)/\sqrt{-g_{11}}}-\frac{1}{r^{\prime}_{1}}\Bigr{)}s\Bigl{(}\frac{1}{v\sqrt{g_{00}}}-\frac{g_{03}}{g_{00}\sqrt{\Gamma_{33}}}\Bigr{)}\approx\frac{mls\cos(\gamma)}{2\hbar r^{\prime}_{1}}\frac{1}{v}\frac{r_{g}}{r^{\prime}_{1}}, (131)

where we have used (72) and (73) (they still hold in the coordinates (t,r,θ,φ)(t^{\prime},r^{\prime},\theta^{\prime},\varphi^{\prime})), and have neglected the third order and higher orders terms of rgr_{g} and aa. Here the product lsls is the area of the interferometry loop, and ll and ss are lengths defined in the Kerr spacetime. Similar calculations lead to

δα4mlscos(γ)4r11vrg2r12,δα53mlscos(γ)4r1rgvr1.\delta\alpha_{4}\approx-\frac{mls\cos(\gamma)}{4\hbar r^{\prime}_{1}}\frac{1}{v}\frac{r_{g}^{2}}{{r^{\prime}_{1}}^{2}},\qquad\delta\alpha_{5}\approx\frac{3mls\cos(\gamma)}{4\hbar r^{\prime}_{1}}\frac{r_{g}v}{r^{\prime}_{1}}. (132)

As for δα3\delta\alpha_{3}, it is still given by the result of Ref. [27],

δα3=mAr1(2cos(θ1)sin(γ)+cos(γ)sin(θ1))argr12,\delta\alpha_{3}=-\frac{mA}{\hbar r^{\prime}_{1}}\Bigl{(}2\cos(\theta^{\prime}_{1})\sin(\gamma^{\prime})+\cos(\gamma^{\prime})\sin(\theta^{\prime}_{1})\Bigr{)}\frac{ar_{g}}{{r^{\prime}_{1}}^{2}}, (133)

where A=lsA=l^{\prime}s^{\prime} is the area defined in flat space. We need to re-express δα3\delta\alpha_{3} according to the area lsls. For this purpose, applying (120) to the parallelogram in the radial direction, we obtain

lcos(γ)=LrΓ11lcos(γ),lcos(γ)lcos(γ)Γ11.l\cos(\gamma)=L_{r^{\prime}}\approx\sqrt{\Gamma_{11}}\,l^{\prime}\cos(\gamma^{\prime}),\qquad\Rightarrow\,l^{\prime}\cos(\gamma^{\prime})\approx\frac{l\cos(\gamma)}{\sqrt{\Gamma_{11}}}. (134)

Similarly, in the direction of θ\theta^{\prime} and φ\varphi^{\prime} we get respectively

lsin(γ)r1Γ22lsin(γ),sr1sin(θ1)Γ33s.l^{\prime}\sin(\gamma^{\prime})\approx\frac{r^{\prime}_{1}}{\sqrt{\Gamma_{22}}}l\sin(\gamma),\qquad s^{\prime}\approx\frac{r^{\prime}_{1}\sin(\theta^{\prime}_{1})}{\sqrt{\Gamma_{33}}}s. (135)

Replacing (134) and (135) into (133), expanding the expression, and neglecting the third order and higher order terms, we get

δα3mlsr1(2cos(θ1)sin(γ)+cos(γ)sin(θ1))argr12.\delta\alpha_{3}\approx-\frac{mls}{\hbar r^{\prime}_{1}}\Bigl{(}2\cos(\theta^{\prime}_{1})\sin(\gamma)+\cos(\gamma)\sin(\theta^{\prime}_{1})\Bigr{)}\frac{ar_{g}}{{r^{\prime}_{1}}^{2}}. (136)

Finally, comparing (131), (132) and (136) with (116), we can see that the forms of δαj\delta\alpha_{j} are not changed.

Appendix B Derivations for the phase difference

In this appendix, we show how to derive the phase difference (75) between the paths ADC and ABC in FIG. 2. Lest us start writing the coordinates of the points A, B, C, and D as follows:

A(tA,r1,θ1,φA),B(tA+tAB,r1,θ1,φA+φAB),C(tD+tDC,r2,θ2,φD+φDC),D(tD,r2,θ2,φD),{\rm A}(t_{A},r_{1},\theta_{1},\varphi_{A}),\quad{\rm B}(t_{A}+t_{AB},r_{1},\theta_{1},\varphi_{A}+\varphi_{AB}),\quad{\rm C}(t_{D}+t_{DC},r_{2},\theta_{2},\varphi_{D}+\varphi_{DC}),\quad{\rm D}(t_{D},r_{2},\theta_{2},\varphi_{D}), (137)

where we defined tAB=tBtAt_{AB}=t_{B}-t_{A}, φAB=φBφA\varphi_{AB}=\varphi_{B}-\varphi_{A}, tDC=tCtDt_{DC}=t_{C}-t_{D}, and φDC=φCφD\varphi_{DC}=\varphi_{C}-\varphi_{D}.

Hence, we write down the phases of each path in the following way. As we mentioned in the assumption (a) in Sec. 4.1, we have dr0\mbox{d}r\approx 0 and dθ0\mbox{d}\theta\approx 0 on the paths AB and DC. Moreover, according to (64), we know that all the components SβS_{\beta} are independent on tt and φ\varphi.131313As for P1P^{1} and P2P^{2} which appear in the expressions of S1S_{1} and S2S_{2}, they are also independent of tt and φ\varphi (see (84)). Therefore, (65) simplifies to:

ϕAB\displaystyle\phi_{AB} \displaystyle\approx 1(S0AtAB+S3AφAB)AB,\displaystyle\frac{1}{\hbar}(S_{0}^{A}t_{AB}+S_{3}^{A}\varphi_{AB})_{AB}, (138)
ϕDC\displaystyle\phi_{DC} \displaystyle\approx 1(S0DtDC+S3DφDC)DC,\displaystyle\frac{1}{\hbar}(S_{0}^{D}t_{DC}+S_{3}^{D}\varphi_{DC})_{DC}, (139)

where the superscripts A and D denote the positions, and the subscripts AB and DC denote the paths. As for the path AD, we have

ϕAD\displaystyle\phi_{AD} =\displaystyle= 1(Sβdxβ)AD\displaystyle\frac{1}{\hbar}\Bigl{(}\int S_{\beta}\mbox{d}x^{\beta}\Bigr{)}_{AD} (140)
=\displaystyle= 1[S0(ra)tAD+S1(rb)rAD+S2(rc)θAD+S3(rd)φAD]AD\displaystyle\frac{1}{\hbar}\Bigl{[}S_{0}(\vec{r}_{a})t_{AD}+S_{1}(\vec{r}_{b})r_{AD}+S_{2}(\vec{r}_{c})\theta_{AD}+S_{3}(\vec{r}_{d})\varphi_{AD}\Bigr{]}_{AD}
=\displaystyle= 1{[S0(ra)S0A+S0A]tAD+[S1(rb)S1A+S1A]rAD+[S2(rc)S2A+S2A]θAD+[S3(rd)S3A+S3A]φAD}AD\displaystyle\frac{1}{\hbar}\Bigl{\{}\!\bigl{[}S_{0}(\vec{r}_{a})\!-\!S_{0}^{A}\!+\!S_{0}^{A}\bigr{]}t_{AD}+\bigl{[}S_{1}(\vec{r}_{b})\!-\!S_{1}^{A}\!+\!S_{1}^{A}\bigr{]}r_{AD}+\bigl{[}S_{2}(\vec{r}_{c})\!-\!S_{2}^{A}\!+\!S_{2}^{A}\bigr{]}\theta_{AD}+\bigl{[}S_{3}(\vec{r}_{d})\!-\!S_{3}^{A}\!+\!S_{3}^{A}\bigr{]}\varphi_{AD}\!\Bigr{\}}_{AD}
\displaystyle\approx 1[S0A(tDtA)+S1A(r2r1)+S2A(θ2θ1)+S3A(φDφA)]AD,\displaystyle\frac{1}{\hbar}\bigl{[}S_{0}^{A}(t_{D}-t_{A})+S_{1}^{A}(r_{2}-r_{1})+S_{2}^{A}(\theta_{2}-\theta_{1})+S_{3}^{A}(\varphi_{D}-\varphi_{A})\bigr{]}_{AD},

where we have used the mean value theorem for integrals in the second step, and ra\vec{r}_{a}, rb\vec{r}_{b}, rc\vec{r}_{c} and rd\vec{r}_{d}, are points on AD. The last step in (140) holds because both (Sβ(rp)SβA)AD(S_{\beta}(\vec{r}_{p})-S_{\beta}^{A})_{AD} and xADβx^{\beta}_{AD} are smaller than or equal to O(l/r1)O(l/r_{1}).141414For any point pp on the path AD, we have |rpr1||r2r1||r_{p}-r_{1}|\leq|r_{2}-r_{1}| and |θpθ1||θ2θ1||\theta_{p}-\theta_{1}|\leq|\theta_{2}-\theta_{1}|. Therefore, expanding (Sβ(rp)SβA)AD(S_{\beta}(\vec{r}_{p})-S_{\beta}^{A})_{AD} in the neighborhoods of r1r_{1} and θ1\theta_{1}, we can find it is smaller than or equal to O(l/r1)O(l/r_{1}). As for xADβx^{\beta}_{AD}, both rADr_{AD} and θAD\theta_{AD} are O(l/r1)O(l/r_{1}), according to (72). Finally, for tADt_{AD} and φAD\varphi_{AD}, the geodesic equations in Kerr spacetime imply [17] t=T1(r)dr+T2(θ)dθ,φ=Φ1(r)dr+Φ2(θ)dθ,t=\int T_{1}(r)\mbox{d}r+\int T_{2}(\theta)\mbox{d}\theta,\qquad\varphi=\int\Phi_{1}(r)\mbox{d}r+\int\Phi_{2}(\theta)\mbox{d}\theta, (141) where T1(r)=r2(r2+a2)+2Mar(aλ)ΔR(r),T2(θ)=a2cos2θΘ(θ),Φ1(r)=a(r2+a2aλ)ΔR(r),Φ2(θ)=λcosec2(θ)aΘ(θ),T_{1}(r)=\frac{r^{2}(r^{2}+a^{2})+2Mar(a-\lambda)}{\Delta\sqrt{R(r)}},\qquad T_{2}(\theta)=\frac{a^{2}\cos^{2}\theta}{\sqrt{\Theta(\theta)}},\qquad\Phi_{1}(r)=\frac{a(r^{2}+a^{2}-a\lambda)}{\Delta\sqrt{R(r)}},\qquad\Phi_{2}(\theta)=\frac{\lambda{\rm cosec}^{2}(\theta)-a}{\sqrt{\Theta(\theta)}}, (142) with R(r)R(r) and Θ(θ)\Theta(\theta) defined in (85) and (86). Hence, from the mean value theorem for integrals we have: tAD=T1(re)rAD+T2(θf)θAD,φAD=Φ1(rg)rAD+Φ2(θh)θAD,t_{AD}=T_{1}(r_{e})r_{AD}+T_{2}(\theta_{f})\theta_{AD},\qquad\varphi_{AD}=\Phi_{1}(r_{g})r_{AD}+\Phi_{2}(\theta_{h})\theta_{AD}, (143) where the subscript ee, ff, gg, and hh denote points on AD. Therefore, tADt_{AD} and φAD\varphi_{AD} are also O(l/r1)O(l/r_{1}). (Recall the sentence after (72).) Similar to (140), we can derive

ϕBC\displaystyle\phi_{BC} \displaystyle\approx 1[S0B(tDtA+tDCtAB)+S1B(r2r1)+S2B(θ2θ1)+S3B(φDφA+φDCφAB)]BC,\displaystyle\frac{1}{\hbar}\bigl{[}S_{0}^{B}(t_{D}-t_{A}+t_{DC}-t_{AB})+S_{1}^{B}(r_{2}-r_{1})+S_{2}^{B}(\theta_{2}-\theta_{1})+S_{3}^{B}(\varphi_{D}-\varphi_{A}+\varphi_{DC}-\varphi_{AB})\bigr{]}_{BC}, (144)

where we have used (137). In order to simplify the expression for (ϕADϕBC)(\phi_{AD}-\phi_{BC}) we have to show the equality (SβA)AD=(SβB)BC(S_{\beta}^{A})_{AD}=(S_{\beta}^{B})_{BC}. We know that (gμνA)AD=(gμνB)BCA(g_{\mu\nu}^{A})_{AD}=(g_{\mu\nu}^{B})_{BC}^{A} holds because the metric is independent of tt and φ\varphi. Therefore, taking into account the latter conclusion and the expression of SβS_{\beta} in (64), we have to prove the following equalities,

(A)AD=(B)BC,(A)AD=(B)BC,(P1)ADA=(P1)BCB,(P2)ADA=(P2)BCB.(\mathcal{E}^{A})_{AD}=(\mathcal{E}^{B})_{BC},\qquad(\mathcal{L}^{A})_{AD}=(\mathcal{L}^{B})_{BC},\qquad(P^{1})^{A}_{AD}=(P^{1})^{B}_{BC},\qquad(P^{2})^{A}_{AD}=(P^{2})^{B}_{BC}. (145)

The first equation in (145) holds because of the assumption (b) in Sec. 4.1. This assumption also implies (vA)AD=(vB)BC(\vec{v}_{A})_{AD}=(\vec{v}_{B})_{BC} for a the particle on the parallelogram. Thus combining this conclusion with the expression (70), the second equation in (145) is proved. Finally, according to (84), the last two equations in (145) also hold. Since we have proved (145), the relation (SβA)AD=(SβB)BC(S_{\beta}^{A})_{AD}=(S_{\beta}^{B})_{BC} holds. The latter equality together with (140) and (144) implies:

ϕADϕBC=1[S0B(tDCtAB)+S3B(φDCφAB)]BC.\phi_{AD}-\phi_{BC}=-\frac{1}{\hbar}\bigl{[}S_{0}^{B}(t_{DC}-t_{AB})+S_{3}^{B}(\varphi_{DC}-\varphi_{AB})\bigr{]}_{BC}. (146)

Merging together (146), (138), and (139), the phase difference between the paths ADC and ABC reads:

δϕ\displaystyle\delta\phi =\displaystyle= ϕADCϕABC\displaystyle\phi_{ADC}-\phi_{ABC} (147)
=\displaystyle= ϕDCϕAB+ϕADϕBC\displaystyle\phi_{DC}-\phi_{AB}+\phi_{AD}-\phi_{BC}
=\displaystyle= 1{[(S0D)DC(S0A)AB]tAB+[(S3D)DC(S3A)AB]φAB\displaystyle\frac{1}{\hbar}\biggl{\{}\bigl{[}(S_{0}^{D})_{DC}-(S_{0}^{A})_{AB}\bigr{]}t_{AB}+\bigl{[}(S_{3}^{D})_{DC}-(S_{3}^{A})_{AB}\bigr{]}\varphi_{AB}
+[(S0D)DC(S0B)BC](tDCtAB)+[(S3D)DC(S3B)BC](φDCφAB)}.\displaystyle+\bigl{[}(S_{0}^{D})_{DC}-(S_{0}^{B})_{BC}\bigr{]}(t_{DC}-t_{AB})+\bigl{[}(S_{3}^{D})_{DC}-(S_{3}^{B})_{BC}\bigr{]}(\varphi_{DC}-\varphi_{AB})\biggr{\}}.

Recalling the expression of SβS_{\beta} in (64), we find that S0S_{0} does not depend on \mathcal{L}, such that S0S_{0} keeps its value when the probe particle turns direction at B and D. Hence, we have (S0B)BC=(S0B)AB=(S0A)AB(S_{0}^{B})_{BC}=(S_{0}^{B})_{AB}=(S_{0}^{A})_{AB}. Therefore, taking (138) into account, we can rewrite (147) as:

δϕ=δϕa+δϕb+δϕc,\delta\phi=\delta\phi_{a}+\delta\phi_{b}+\delta\phi_{c}, (148)

where

δϕa\displaystyle\delta\phi_{a} =\displaystyle= 1[(S0D)DCtAB+(S3D)DCφAB]ϕAB,\displaystyle\frac{1}{\hbar}\bigl{[}(S_{0}^{D})_{DC}t_{AB}+(S_{3}^{D})_{DC}\varphi_{AB}\bigr{]}-\phi_{AB}, (149)
δϕb\displaystyle\delta\phi_{b} =\displaystyle= 1[(S0D)DC(S0A)AB](tDCtAB),\displaystyle\frac{1}{\hbar}\bigl{[}(S_{0}^{D})_{DC}-(S_{0}^{A})_{AB}\bigr{]}(t_{DC}-t_{AB}), (150)
δϕc\displaystyle\delta\phi_{c} =\displaystyle= 1[(S3D)DC(S3B)BC](φDCφAB).\displaystyle\frac{1}{\hbar}\bigl{[}(S_{3}^{D})_{DC}-(S_{3}^{B})_{BC}\bigr{]}(\varphi_{DC}-\varphi_{AB}). (151)

In the following, we compute explicitly the above phase differences. We remind that the phase ϕAB\phi_{AB} was obtained in (69). For the first term in (149), comparing it with (138), we only need to replace r1r_{1} with r2r_{2}, θ1\theta_{1} with θ2\theta_{2}, and AB\mathcal{L}_{AB} with DC\mathcal{L}_{DC} in (69). Therefore, (149) is equivalent to:

δϕa=ϕAB(r1r2,θ1θ2,ABDC)ϕAB(r1,θ1,AB).\delta\phi_{a}=\phi_{AB}(r_{1}\rightarrow r_{2},\theta_{1}\rightarrow\theta_{2},\mathcal{L}_{AB}\rightarrow\mathcal{L}_{DC})-\phi_{AB}(r_{1},\theta_{1},\mathcal{L}_{AB}). (152)

Now we focus on the expression of AB\mathcal{L}_{AB}. According to (70), we need to find the expression for vφv^{\varphi}. By means of the equation [18]:

v2=Γijvivj,v^{2}=\Gamma_{ij}v^{i}v^{j}, (153)

in Kerr spacetime we find:

|vφ|=v2+g11(vr)2+g22(vθ)2Γ33.|v^{\varphi}|=\sqrt{\frac{v^{2}+g_{11}(v^{r})^{2}+g_{22}(v^{\theta})^{2}}{\Gamma_{33}}}. (154)

On the path AB we have vφ>0v^{\varphi}>0, vr=0v^{r}=0, and vθ=0v^{\theta}=0, therefore, (154) simplifies to:

vφ=v2Γ33.v^{\varphi}=\sqrt{\frac{v^{2}}{\Gamma_{33}}}. (155)

Replacing (155) into (70), we obtain:

AB=g00(g03+g00Γ33v2)r1sin(θ1)ϵ,\mathcal{L}_{AB}=\frac{\mathcal{E}}{g_{00}}\Bigl{(}-g_{03}+\sqrt{g_{00}\Gamma_{33}v^{2}}\Bigr{)}\approx\mathcal{E}r_{1}\sin(\theta_{1})\epsilon, (156)

where ϵ\epsilon is given by:

ϵ={1m22,for massive particles,1,for massless particles.\epsilon=\begin{cases}\sqrt{1-\frac{m^{2}}{\mathcal{E}^{2}}},&\text{for massive particles},\\ 1,&\text{for massless particles}.\end{cases} (157)

We have neglected the terms of rgr_{g} and aa in the last expression of (156) because, according to (69), AB\mathcal{L}_{AB} only appears in the third order and higher order terms of ϕAB\phi_{AB}. Moreover, we have used (74) and expression (51) in the last step of (156). As for DC\mathcal{L}_{DC}, we only need to replace r1r_{1} with r2r_{2}, and θ1\theta_{1} with θ2\theta_{2} in (156). As for (152), using (72) we expand δϕa\delta\phi_{a} in the neighborhoods of r1r_{1} and θ1\theta_{1} up to the first order. Finally, we get:

δϕa\displaystyle\delta\phi_{a} \displaystyle\approx lr1{tAB[cos(γ)(rg2r1+rg22r12+rg32r13+a2rg4r13(17cos2(θ1)))+sin(2θ1)sin(γ)a2rg2r13]\displaystyle\frac{\mathcal{E}l}{\hbar r_{1}}\Bigl{\{}t_{AB}\Bigl{[}\cos(\gamma)\Bigl{(}\frac{r_{g}}{2r_{1}}+\frac{r_{g}^{2}}{2r_{1}^{2}}+\frac{r_{g}^{3}}{2r_{1}^{3}}+\frac{a^{2}r_{g}}{4r_{1}^{3}}\bigl{(}1-7\cos^{2}(\theta_{1})\bigr{)}\Bigr{)}+\sin(2\theta_{1})\sin(\gamma)\frac{a^{2}r_{g}}{2r_{1}^{3}}\Bigr{]} (158)
+φAB[ϵa2rgr12sin(θ1)(cos(γ)sin2(θ1)+34sin(γ)sin(2θ1))\displaystyle+\varphi_{AB}\Bigl{[}\epsilon\frac{a^{2}r_{g}}{r_{1}^{2}}\sin(\theta_{1})\Bigl{(}\cos(\gamma)\sin^{2}(\theta_{1})+\frac{3}{4}\sin(\gamma)\sin(2\theta_{1})\Bigr{)}
(cos(γ)sin2(θ1)(argr1+3arg22r12)+sin(γ)sin(2θ1)(argr1+arg2r12))]}.\displaystyle-\Bigl{(}\cos(\gamma)\sin^{2}(\theta_{1})\Bigl{(}\frac{ar_{g}}{r_{1}}+\frac{3ar_{g}^{2}}{2r_{1}^{2}}\Bigr{)}+\sin(\gamma)\sin(2\theta_{1})\Bigl{(}\frac{ar_{g}}{r_{1}}+\frac{ar_{g}^{2}}{r_{1}^{2}}\Bigr{)}\Bigr{)}\Bigr{]}\Bigr{\}}.

Plugging (157), (73), and (74) into (158), and expanding the resulting expression, we get:

δϕa\displaystyle\delta\phi_{a} \displaystyle\approx 0lsr1{1v[cos(γ)(rg2r1+rg22r12+rg32r13+a2rg4r13(17cos2(θ1)))+sin(2θ1)sin(γ)a2rg2r13]\displaystyle\frac{\mathcal{E}_{0}ls}{\hbar r_{1}}\Bigl{\{}\frac{1}{v}\Bigl{[}\cos(\gamma)\Bigl{(}\frac{r_{g}}{2r_{1}}+\frac{r_{g}^{2}}{2r_{1}^{2}}+\frac{r_{g}^{3}}{2r_{1}^{3}}+\frac{a^{2}r_{g}}{4r_{1}^{3}}\bigl{(}1-7\cos^{2}(\theta_{1})\bigr{)}\Bigr{)}+\sin(2\theta_{1})\sin(\gamma)\frac{a^{2}r_{g}}{2r_{1}^{3}}\Bigr{]} (159)
+va2rgr13sin(θ1)(cos(γ)sin(θ1)+32sin(γ)cos(θ1))\displaystyle+v\frac{a^{2}r_{g}}{r_{1}^{3}}\sin(\theta_{1})\Bigl{(}\cos(\gamma)\sin(\theta_{1})+\frac{3}{2}\sin(\gamma)\cos(\theta_{1})\Bigr{)}
argr12(2cos(θ1)sin(γ)+cos(γ)sin(θ1))arg2r13(cos(θ1)sin(γ)+32cos(γ)sin(θ1))},\displaystyle-\frac{ar_{g}}{r_{1}^{2}}\Bigl{(}2\cos(\theta_{1})\sin(\gamma)+\cos(\gamma)\sin(\theta_{1})\Bigr{)}-\frac{ar_{g}^{2}}{r_{1}^{3}}\Bigl{(}\cos(\theta_{1})\sin(\gamma)+\frac{3}{2}\cos(\gamma)\sin(\theta_{1})\Bigr{)}\Bigr{\}},

where 0\mathcal{E}_{0} is defined by:

0={m(1v2)1/2,for massive particles,ω,for massless particles.\mathcal{E}_{0}=\begin{cases}m(1-v^{2})^{-1/2},&\text{for massive particles},\\ \hbar\omega,&\text{for massless particles}.\end{cases} (160)

As for δϕb\delta\phi_{b} in (150), it is O(l2/r12)O(l^{2}/r_{1}^{2}) and higher orders because both [(S0D)DC(S0A)AB][(S_{0}^{D})_{DC}-(S_{0}^{A})_{AB}] and (tDCtAB)(t_{DC}-t_{AB}) are of the order of O(l/r1)O(l/r_{1}) and higher orders, according to (72). Therefore, in our approximation (we remind the reader the sentence after (72)), this phase difference is negligible:

δϕb0.\delta\phi_{b}\approx 0. (161)

As for δϕc\delta\phi_{c} in (151), repeating the above calculations, we find:

δϕc0ls2r1a2rgr13sin(θ1)sin(θ1γ)(vv2(vr)2(r1vθ)2),\delta\phi_{c}\approx\frac{\mathcal{E}_{0}ls}{2\hbar r_{1}}\frac{a^{2}r_{g}}{r_{1}^{3}}\sin(\theta_{1})\sin(\theta_{1}-\gamma)\Bigl{(}v-\sqrt{v^{2}-(v^{r})^{2}-(r_{1}v^{\theta})^{2}}\Bigr{)}, (162)

where vrv^{r} and vθv^{\theta} are the components of the velocity at the point B corresponding to the path BC (see the definition (71)). In (162) we can find that the phase difference δϕc\delta\phi_{c} is sensible to the direction of the velocity at B along the path BC, therefore, δϕc\delta\phi_{c} depends on the shape of the parallelogram.

According to the above results, δϕb\delta\phi_{b} and δϕc\delta\phi_{c} are much smaller than δϕa\delta\phi_{a}. This is what we expected because according to (150) and (151), δϕb\delta\phi_{b} and δϕc\delta\phi_{c} are due to the differences between tABt_{AB} and tDCt_{DC}, and φAB\varphi_{AB} and φDC\varphi_{DC} respectively. These differences are very small compared with tABt_{AB} and φAB\varphi_{AB}, thus, δϕb\delta\phi_{b} and δϕc\delta\phi_{c} can be regarded as small modifications to the phase difference. Merging together (159), (161), and (162), we finally obtain the phase difference (75).

Now we show briefly how (82) is derived for θ1=0\theta_{1}=0 and θ1=π\theta_{1}=\pi. As for δϕa\delta\phi_{a}, we only need to insert ϕAB=π\phi_{AB}=\pi into (158), and to repeat the above calculations. Hence, δϕb0\delta\phi_{b}\approx 0 still holds. Finally, δϕc\delta\phi_{c} can be also neglected, because when θ1=0\theta_{1}=0 or θ1=π\theta_{1}=\pi hold, plugging (83) into (151), it results:

δϕc1sin2(θ2)(argr212DCa2rgr23+arg2r22)(φDCφAB).\delta\phi_{c}\approx\frac{1}{\hbar}\sin^{2}(\theta_{2})\Bigl{(}\mathcal{E}\frac{ar_{g}}{r_{2}}-\frac{1}{2}\mathcal{L}_{\rm DC}\frac{a^{2}r_{g}}{r_{2}^{3}}+\mathcal{E}\frac{ar_{g}^{2}}{r_{2}^{2}}\Bigr{)}(\varphi_{DC}-\varphi_{AB}). (163)

Here sin2(θ2)\sin^{2}(\theta_{2}) is of the order of O(l2/r12)O(l^{2}/r_{1}^{2}) and higher orders, hence δϕc\delta\phi_{c} is negligible. Based on above results, we can derive (82).

References

  • [1] Y. Aharonov and D. Bohm, Phys. Rev. 115, 485-491 (1959) doi:10.1103/PhysRev.115.485
  • [2] T. T. Wu and C. N. Yang, Phys. Rev. D 12, 3845-3857 (1975) doi:10.1103/PhysRevD.12.3845
  • [3] R. G. Chambers, Phys. Rev. Lett. 5, no.1, 3-5 (1960) doi:10.1103/physrevlett.5.3
  • [4] A. Tonomura, Lect. Notes Phys. 340, 35-152 (1989) doi:10.1007/BFb0032078
  • [5] J. S. Dowker, Nuovo Cimento B 52, 129 (1967)
  • [6] L. H. Ford and A. Vilenkin, J. Phys. A 14, 2353 (1981)
  • [7] J. Audretsch and C. Lammerzahl, J. Phys. A 16, 2457 (1983) doi:10.1088/0305-4470/16/11/017
  • [8] L. Stodolsky, Gen. Rel. Grav. 11, 391-405 (1979) doi:10.1007/BF00759302
  • [9] R. Aldrovandi and J. G. Pereira, Fundam. Theor. Phys. 173 (2013) doi:10.1007/978-94-007-5143-9
  • [10] A. W. Overhauser and R. Colella, Phys. Rev. Lett. 33, 12377 (1974) doi:10.1103/PhysRevLett.33.1237
  • [11] R. Colella, A. W. Overhauser and S. A. Werner, Phys. Rev. Lett. 34, 1472-1474 (1975) doi:10.1103/PhysRevLett.34.1472
  • [12] M. A. Hohensee, B. Estey, P. Hamilton, A. Zeilinger and H. Muller, Phys. Rev. Lett. 108, 230404 (2012) doi:10.1103/PhysRevLett.108.230404 [arXiv:1109.4887 [quant-ph]].
  • [13] A. Roura, Science 375, no.6577, 142-143 (2021) doi:10.1126/science.abm6854
  • [14] C. Overstreet, P. Asenbaum, J. Curti, M. Kim and M. A. Kasevich, Science 375, no.6577, abl7152 (2021) doi:10.1126/science.abl7152
  • [15] R. Aldrovandi, J. G. Pereira and K. H. Vu, Class. Quant. Grav. 21, 51-62 (2004) doi:10.1088/0264-9381/21/1/004 [arXiv:gr-qc/0310110 [gr-qc]].
  • [16] J. G. Pereira, T. Vargas and C. M. Zhang, Class. Quant. Grav. 18, 833-842 (2001) doi:10.1088/0264-9381/18/5/306 [arXiv:gr-qc/0102070 [gr-qc]].
  • [17] S. Chandrasekhar, Oxford University Press (1998)
  • [18] L. D. Landau and E. M. Lifschits, Pergamon Press, 1975, ISBN 978-0-08-018176-9
  • [19] S. M. Carroll, Cambridge University Press, 2019, ISBN 978-0-8053-8732-2, 978-1-108-48839-6, 978-1-108-77555-7
  • [20] B. W. Carroll, D. A. Ostlie, Cambridge University Press. (2017)
  • [21] J. C. A. Miller-Jones, A. Bahramian, J. A. Orosz, I. Mandel, L. Gou, T. J. Maccarone, C. J. Neijssel, X. Zhao, J. Ziółkowski and M. J. Reid, et al. Science 371, no.6533, 1046-1049 (2021) doi:10.1126/science.abb3363 [arXiv:2102.09091 [astro-ph.HE]].
  • [22] X. Zhao, L. Gou, Y. Dong, X. Zheng, J. F. Steiner, J. C. A. Miller-Jones, A. Bahramian, J. A. Orosz and Y. Feng, Astrophys. J. 908, no.2, 117 (2021) doi:10.3847/1538-4357/abbcd6 [arXiv:2102.09093 [astro-ph.HE]].
  • [23] B. L. Giacchini, T. d. Netto and L. Modesto, Phys. Rev. D 104, no.8, 084072 (2021) doi:10.1103/PhysRevD.104.084072 [arXiv:2105.00300 [gr-qc]].
  • [24] L. Modesto, J. W. Moffat and P. Nicolini, Phys. Lett. B 695, 397-400 (2011) doi:10.1016/j.physletb.2010.11.046 [arXiv:1010.0680 [gr-qc]].
  • [25] L. Modesto and P. Nicolini, Phys. Rev. D 82, 104035 (2010) doi:10.1103/PhysRevD.82.104035 [arXiv:1005.5605 [gr-qc]].
  • [26] P. Nicolini, Int. J. Mod. Phys. A 24, 1229-1308 (2009) doi:10.1142/S0217751X09043353 [arXiv:0807.1939 [hep-th]].
  • [27] S. Wajima, M. Kasai and T. Futamase, Phys. Rev. D 55, 1964-1970 (1997) doi:10.1103/PhysRevD.55.1964
  • [28] M. G. Sagnac, C. R. Acad. Sci. (Paris) 157, 708 (1913).
  • [29] M. G. Sagnac, C. R. Acad. Sci. (Paris) 157, 1410 (1913).
  • [30] J. Lense and H. Thirring, Phys. Z. 19, 156 (1918).
  • [31] J. Kuroiwa, M. Kasai and T. Futamase, Phys. Lett. A 182, 330 (1993).