Vous êtes sur la page 1sur 8

Gravitational Quantum Hall Effect

Naoki Yamamoto
Department of Physics, Keio University, Yokohama 223-8522, Japan

In the classical gravitational lensing, a light ray is locally deflected in the direction
of by the gravitational potential . We show that the quantum correction due
to the helicity of photons leads to the displacement of the trajectory of light in the
arXiv:1708.03113v1 [hep-th] 10 Aug 2017

direction perpendicular to both and the classical trajectory, and, in particular, to


the splitting of the trajectories of right- and left-handed circularly polarized light. We
derive the expression for this gravitational quantum Hall effect (GQHE) in terms of
the Berry curvature of photons. We also derive the semiclassical equation of motion
for gravitons taking into account the Berry curvature, and find that the GQHE of
gravitational waves in curved space is twice as large as that of light.

I. INTRODUCTION

Gravitational lensing is the deflection of light rays in a gravitational field around a massive
object. Locally, a light ray is deflected in the direction of by the gravitational potential
. This is one of the important predictions in Einsteins theory of general relativity.
In this paper, we show that the gravitational lensing of light in curved space, described
by the classical general relativity, receives a modification due to the helicity of photons
in the semiclassical regime. We show that the quantum correction leads to the shift of
the ray trajectory in the direction perpendicular to both and the classical trajectory,
and to the splitting of the trajectories of right- and left-handed circularly polarized light.
We find that the correction to the classical gravitational lensing is expressed by the Berry
curvature of photons; see Eq. (9). A closely related effect of light is known in an optically
inhomogeneous medium (not in a gravitational potential) in the context of optics and is
called the optical Magnus effect or Hall effect of light [13]. In our case, the counterpart
of the optical inhomogeneity is produced by the background curved space. This is a
realization of the quantum Hall effect of light by background gravitational fields, and we
call it the gravitational quantum Hall effect (GQHE) of light.1
We also show that the classical gravitational lensing of gravitational waves in curved
space is modified by the helicity of gravitons, and that the quantum correction is expressed
by the Berry curvature of gravitons. In particular, we find that the GQHE of gravitational
waves is twice as large as that of light. This result shows that, although the trajectories of
both light and gravitational wave in the curved space are null geodesic and are degenerate
at the classical level, this degeneracy is resolved by their helical nature in the semiclassical

1
Massless spinning particles in a gravitational field were previously studied, e.g., in Ref. [4], but the GQHE
derived in this paper is new to the best of our knowledge.
2

regime.

II. GRAVITATIONAL QUANTUM HALL EFFECT OF LIGHT

We first illustrate the quantum hall effect of light in a gravitational field. To keep the
relativistic and quantum mechanical nature apparent, we will explicitly write ~ and c in this
section.
Let us first recall the classical gravitational lensing of light in a gravitational potential
(x). Consider the case in a weak gravitational field, where the metric is given by
   
2 2 2 2
ds = 1 + 2 c dt 1 2 (dx)2 .
2
(1)
c c

Here, we assume is static and satisfies ||/c2 1. Because the light propagates along a
null geodesic, ds2 = 0, the velocity of light is given by
s
1 + 2
 
|dx| c2 2
c =c c 1+ 2 . (2)
dt 1 2c2
c

It has been known that the null geodesic equation in the static and weak gravitational
potential is equivalent to the geometric-optical equation of light rays in a medium with re-
fractive index n(x) that varies depending on (x) [5]. Equation (2) shows that the refractive
index is given by
s
1 2
c2 2
n= 1 . (3)
1 + 2
c 2
c 2

For < 0, c is smaller than c in vacuum and n > 1. This feature allows us to describe the
effects of weak gravity using purely the language of the geometric optics. In the following,
we will use the Hamiltonian formulation of the geometric optics.
We now note that the equations of motion for light in curved space are also affected by
the helicity. We consider the regime where the wavelength of light is much smaller than
the radius of curvature of the background gravity. In this semiclassical regime, the helical
nature of photons can be expressed by the Berry connection or Berry curvature [13, 6].
The action for right- and left-handed circularly polarized light in the weak gravity is given
by
Z
I = dt (p x ap p p ) .

(4)

Here, ap is the Berry connection of photons, which is related to the Berry curvature p via

p
p p ap = (5)
p2
3

with p |p| and being the helicity of photons ( = ~ for right- and left-handed photons,
respectively). As we explained above, the effect of the static gravitational potential is
accounted for by the refractive index (3), which modifies the energy dispersion of photons
as
p
p = pc = , (6)
n
where c and n are given by Eqs. (2) and (3).
The semiclassical equations of motion for the wave packet of light are obtained from the
action (4) as

x = p + p p , (7)
 

p = 2p 2 . (8)
c

In the context of the geometric optics, the second term in Eq. (7) was called the optical
Magnus effect [1, 2]. On the other hand, Eq. (8) represents the (classical) gravitational
lensing effect in the gravitational potential . Inserting Eq. (8) into Eq. (7), we have
 
p
x = p 2 2 . (9)
c p

The second term represents the spin quantum Hall effect of light induced by the background
curved geometry. This is the gravitational quantum Hall effect (GQHE) of light. The
trajectory of light is shifted in the direction perpendicular to both and the classical
trajectory p, and in particular, the trajectories of right- and left-handed circularly polarized
light are separated. This effect originates from the interplay between general relativity and
the helical nature of right- or left-handed photons.
As an example, consider the Newtonian potential at a distance r from a point mass M:
GM
(r) = , (10)
r
where G is the universal gravitational constant. In this case, Eq. (9) reduces to

2GM r p
x = p + 2 2
. (11)
c r p

Equation (11) shows that the GQHE becomes larger as M increases and as p decreases (for
fixed r). Thus, the GQHE becomes particularly relevant for electromagnetic waves with
long wavelength around massive astrophysical objects.
It is straightforward to get the generic kinetic theory for photons in the weak gravitational
field. By inserting Eqs. (8) and (9) into the kinetic equation,

f f f
+ x + p = C[f ] , (12)
t x p
4

where f = f (t, x, p) is the distribution function of photons with helicity and C[f ] is
the collision term, we get
     
f p f f
+ p 2 2 2p 2 = C[f ] . (13)
t c p x c p
This equation describes the time evolution of right- and left-handed photons for any given
(weak and static) gravitational field .

III. GRAVITATIONAL QUANTUM HALL EFFECT OF GRAVITATIONAL


WAVES

In this section, we consider the GQHE of gravitational waves in the semiclassical regime.
At the classical level, the gravitational wave travels along a null geodesic, and hence, the
propagation of the gravitational wave in curved space is described by the same geometric-
optical equation as the case of light [7]. We will show that gravitons possess a Berry curvature
in the semiclassical regime, leading to the GQHE of gravitons in curved space, similarly to
that of photons above.

A. Generalized Weyl equation with any helicity

Let us briefly review the generalized Weyl equation for massless fields with any spin
[9, 10], and then we apply it to spin-2 gravitons. In the following, we use the natural units
~ = c = 1 for simplicity, unless stated otherwise.
We first recall the representation of the Poincare algebra for massless fields. The Poincare
symmetry consists of the space-time translations generated by the energy-momentum vector
P and the Lorentz transformations generated by M . In 3+1 space-time dimensions, there
are two Casimir operators that commute with p and M : p2 = p p and W 2 = W W ,
where W is the Pauli-Lubanski vector defined by
1
W = p M . (14)
2
Because the contribution of the orbital angular momentum vanishes due to the antisymmetry
with p , Eq. (14) can also be rewritten as
1 1
W = p S = p S , S S , (15)
2 2
where S is the spin tensor, whose components are S ij = ijk S k and S 0i = iS i with S i
being the spin vector. According to Wigners result [8], we have the relation:

W = p , (16)

which defines the helicity of massless fields.


5

Let us now introduce the (2|| + 1)-component massless field with helicity , which
satisfies the equation (W p ) = 0. Using Eq. (15), this equation can be written as

(S p + p ) = 0. (17)

The temporal ( = 0) and spatial ( = i) components of this equation are given by

(S p p0 ) = 0, (18)
(Sp0 + iS p p) = 0, (19)

respectively. Equation (18) is the generalized Weyl equation for massless field with helicity
, and Eq. (19) is the subsidiary condition [9, 10]. The generalized Weyl Hamiltonian
corresponding to the wave equation (18) is
1
H= S p. (20)

Note that three components of Eq. (19) are not independent. To see this, we first eliminate
0
p in Eq. (19) using Eq. (18) to get
 
1
S(S p) + iS p p = 0 . (21)

It is then easy to check that the inner product of the left-hand side of Eq. (21) with S
vanishes, meaning that only one of three components in Eq. (21) is independent. Without
loss of generality, we take the z-component of Eq. (21),
 
1
Sz (S p) + i (Sx py Sy px ) pz = 0 , (22)

as the subsidiary condition to Eq. (18).
In particular, for gravitons with = 2, a matrix representation of S is

0 1 q0 0 0 0 1 q 0 0 0
3

1 0 0 0 1 0 3 0 0
q 2 q 2
q q
3 3 3 3
Sx = 0

2
0 2
0
, S y = i
0 2
0 2
0 ,
(23)
q q
0 0 3 3
2
0 1

0 0
2
0 1

0 0 0 1 0 0 0 0 1 0

and Sz = diag(2, 1, 0, 1, 2).

B. Path integral formulation of gravitons in flat space

Once the wave equation (18) for gravitons is obtained, the semiclassical equations of
motion for gravitons taking into account the Berry curvature can be derived in a way anal-
ogous to Ref. [11] for chiral fermions. The important difference in our case is, however, the
6

additional constraint (22), which selects out the physical degrees of freedom of the graviton
with helicity 2, similarly to the situation of photons with helicity 1 [6].
Let us consider the path integral quantization for the Hamiltonian (20) for = 2:
Z Z
iI
Z = DxDpPe , I = dt (p x H) , (24)

where P denotes the path-ordered product of the matrices exp(iHt) over the path in
the phase space. The eigenvalues of H are given by |p|, 21 |p|, and 0, and H can be
diagonalized using a unitary matrix Vp as
1
Vp HVp = |p|, Sz . (25)
2
The eigenstates of the eigenvalues |p| and 12 |p| have the negative energies and are not
physical. Also, one can check that the eigenstates of the eigenvalues 0 and 21 |p| are forbidden
due to the subsidiary condition (22). Therefore, we have only one physical eigenstate with
the eigenvalue |p|, which corresponds to helicity = 2.
Following the procedure in Refs. [6, 11], we diagonalize the matrix in the exponential
factor of the path integral (24) at each point of the trajectory as
 i   i 
exp S p2 t exp S p1 t = Vp2 exp(i|p2 |t)Vp2 Vp1 exp(i|p1 |t)Vp1
2 2
= Vp2 exp(i|p2 |t) exp(iaGp pt)

exp(i|p1 |t)Vp1 , (26)

where aG
p iVp p Vp . In deriving the last equation above, we used

Vp2 Vp1 exp(iaG G


p p) = exp(iap pt) (27)

for sufficiently small p p2 p1 .


Taking the semiclassical limit where off-diagonal components of aG
p are ignored, we obtain
the semiclassical action for gravitons in the flat space:
Z
I = dt (p x aG
G
p p p ), (28)

where p = |p| is the energy dispersion and aG


p [ap ]11 is the Berry connection in momentum
space that originates from the helicity of gravitons. From the definition of aG p above, one
finds the Berry curvature of gravitons as
p
G G
p p ap = . (29)
|p|2
This corresponds to the fictitious magnetic field of the magnetic monopole with charge,
1
Z
k= G p dS = . (30)
4
Note that the Berry curvature of gravitons in Eq. (29) is twice as large as that of photons
in Ref. [6] and four times as large as that of chiral fermions in Refs. [11, 12].
7

C. Semiclassical equations of motion for gravitons in curved space

In the weak and static gravitational potential (x), the energy dispersion of gravitons is
modified as Eq. (6) in the same way as photons, where n is the refractive index of space in
Eq. (3). In this case, the action of the graviton is given by Eq. (28) with p being replaced
by Eq. (6). Then, the semiclassical equations of motion for gravitons become

x = p + p Gp, (31)
 

p = 2p 2 . (32)
c

The second term on the right-hand side of Eq. (31) is the Lorentz force in momentum space,
which may be regarded as the gravitational Magnus effect. To the best of our knowledge,
this effect has not been studied so far.
From the two equations above, we obtain Eq. (9) with helicity = 2. The second
term of this equation is the GQHE of the gravitational wave, which is twice as large as that
of light due to the difference of helicity. This means that, although the trajectory of the
gravitational wave in curved space is the same as that of light at the classical level, this
degeneracy of trajectories is lifted by the quantum effects.
Similarly to the case of photons, the kinetic equation for gravitons in the weak gravity
is given by Eq. (13) with f = f (t, x, p) being replaced by the distribution function of
gravitons.

IV. DISCUSSIONS

In this paper, we have derived the quantum corrections to the gravitational lensing of
electromagnetic/gravitational waves in curved space. This gravitational quantum Hall effect
(GQHE) emerges as a consequence of the interplay between general relativity and helicity
of photons/gravitons. In particular, this quantum correction causes the splitting of the
trajectories of right- and left-handed circularly polarized waves.
To quantify the GQHE, consider an electromagnetic wave or a gravitational wave passing
near a Schwarzschild black hole as an example. From Eq. (11), the relative magnitude of
the local shift due to the GQHE, compared with the classical trajectory, is written as
 2  
|| Rs
A= , (33)
2~ r Rs

where Rs = 2GM/c2 is the Schwarzschild radius, r is the distance from the black hole,
and is the wavelength. This relation shows that the GQHE becomes more relevant as
becomes larger, as long as the semiclassical approximation is valid. For example, for a
black hole with solar mass M = M , the relative magnitude is A 103 for r 5Rs and
300 m. It does not seem feasible to observe the total GQHE (given by the integration of
8

the local GQHE over the actual path) both for electromagnetic and gravitational waves by
the current detectors. However, possible future observations of the GQHE of gravitational
waves, in particular, could test the quantum nature of gravitons beyond the classical general
relativity.
From a theoretical point of view, it would be interesting to extend our findings to a
manifestly covariant formulation similar to Ref. [13].

Acknowledgement

The author thanks M. Hongo and X. G. Huang for useful conversations. This work was
supported by JSPS KAKENHI Grant No. 16K17703 and MEXT-Supported Program for
the Strategic Research Foundation at Private Universities, Topological Science (Grant
No. S1511006).

[1] V. S. Liberman and B. Y. Zeldovich, Phys. Rev. A 46, 5199 (1992).


[2] K. Y. Bliokh and Y. P. Bliokh, Phys. Lett. A 333, 181 (2004).
[3] M. Onoda, S. Murakami, and N. Nagaosa, Phys. Rev. Lett. 93, 083901 (2004).
[4] B. Mashhoon, Ann. Phys. 89, 254 (1975).
[5] S. Carroll, Spacetime and Geometry: An introduction to General Relativity (Addison Wesley,
2003).
[6] N. Yamamoto, arXiv:1702.08886 [hep-th].
[7] C. W. Misner, K. S. Thorne, and J. A. Wheeler, Gravitation (Freeman, San Francisco, 1973).
[8] E. P. Wigner, Annals Math. 40, 149 (1939).
[9] H. Bacry, Nuovo Cim. A 32, 448 (1976).
[10] A. Gersten and A. Moalem, J. Phys. Conf. Ser. 330, 012010 (2011).
[11] M. A. Stephanov and Y. Yin, Phys. Rev. Lett. 109, 162001 (2012).
[12] D. T. Son and N. Yamamoto, Phys. Rev. Lett. 109, 181602 (2012); Phys. Rev. D 87, 085016
(2013).
[13] J. Y. Chen, D. T. Son, and M. A. Stephanov, Phys. Rev. Lett. 115, 021601 (2015).

Vous aimerez peut-être aussi