1. Introduction
Various symmetry breaking effects arise at the
propagation of a polarized light in dielectric media. Different transmission
levels for left and right-hand circular polarizations (circular dichroism)
exhibit optically active materials. Conventional optical activity is associated
with intrinsically 3D-chiral molecules, and it is the property of unequal
absorption of right and left hand circular polarized light. In [1,2] directionally asymmetric transmission of
polarized light in planar chiral structures was considered. Optical activity
may also arise from extrinsic chirality. Strong optical activity and circular
dichroism in non-chiral planar microwave and photonic metamaterials was
demonstrated in [3]. It is well known that,
when a light beam is reflected from an interface, the longitudinal shift of the
gravity center of the beam is different for s- and p-polarized
beams [4], while the transverse shift has
reverse signs in the case of right- and left-hand circularly polarized
radiation [5]. Lateral and angular shifts for
strongly focused azimuthally and radially polarized beams at a dielectric
interface were shown in [6].
Polarization-dependent light transmission occurs in a filter with frustrated
total internal reflection (FTIR) due to different resonance conditions for
incident beams with s- and p- polarization [7,8]. Recently, a phenomenon of spin-dependent
splitting of the focal spot of a plasmonic focusing lens was demonstrated
experimentally [9,10]. Polarization-dependent
splitting of the reflected beam from the surface of the subwavelength grating
was also shown in [11,12]. The effect of spin
symmetry breaking via spin-orbit interaction, which occurs even in rotationally
symmetric structures, was observed in plasmonic nanoapertures [13]. Similar effect of polarization-dependent
transmission through subwavelength round and square apertures was demonstrated
in [14].
Polarization-dependent symmetry breaking effects
occur also for a light propagating in optical waveguides. It is known that the
polarization plane in an inhomogeneous medium is rotated on propagation of
light ray on the helical trajectory [15,16].
Such rotation was observed experimentally in a single mode optical fiber wound
on a cylinder [17] and interpreted in terms of
Berry's geometrical phase [18]. In [19] the rotation of the polarization plane was
observed also in a straight multimode fiber with step-index-type profile. It is
of interest also to consider the inverse effect, i.e., the influence of
polarization on the trajectory and the width of a radiation beam. As
demonstrated in [20], these effects can be
also observed in optical fibers, where small shifts are accumulated due to
multiple reflections in the process of radiation propagation through a fiber.
The authors of [20] calculated the rotation of
the speckle pattern produced by circularly polarized light at the output of a
fiber corresponding to the reversal of the sign of circular polarization. In [21] it was experimentally demonstrated that the
rotation angle of the speckle pattern depends on the angle at which a
circularly polarized light beam is coupled into a fiber. It was shown in [22,23] that spin-orbit interaction causes asymmetry
effect for depolarization of the right- and left-handed circularly polarized
light propagating in a graded-index (GRIN) fiber. Depolarization is stronger if
the helicity of the trajectory of rays and photons has the same sign, and less
if they do not coincide. Spin-dependent relative shift between right- and
left-hand circularly polarized light beams propagating along a helical
trajectory in a graded-index fiber was shown in [24,25]. In [26] this
effect was observed experimentally for a laser beam propagating in the glass
cylinder along the smooth helical trajectory. This shift can be regarded as a
manifestation of the optical Magnus effect [27]
and the optical spin-Hall effect [28,29] which
arises due to a spin-orbit coupling. Propagation of light beams in a
graded-index medium is mainly investigated in the paraxial approximation. Both
ray and wave optics is applied for the analysis of light propagation in
graded-index media [30–42]. Effects of the
polarization on the modes in lens-like media were analyzed in [43]. In [44] the
polarization-dependent Goos-Hanchen (GH) beam shift at a graded-index
dielectric interface is examined both theoretically and experimentally. In [45] the beam shifts or corrections with respect to
geometrical optics caused by the nonparaxiality and spin-orbit interaction in a
graded-index optical fiber are investigated.
In this paper, the effect of symmetry breaking for
left- and right-handed circularly polarized light in an isotropic graded-index
fiber due to spin-orbit interaction forces is demonstrated analytically by
solving the full three-component field Maxwell’s equations. It is shown that
the propagation velocities of vortex modes with right- and left- handed
polarizations differ from each other due to spin-orbit interaction.
2. Basic equations
The Maxwell equations for the electric field
in a general inhomogeneous medium with
dielectric constant
ε(x, y)
reduce to:
where is the wavenumber and is the dielectric permittivity of
the medium.
In the paraxial approximation, equation (1) can be reduced to the equivalent time-independent Schrodinger equation
[46]. An analogous approach may be used to obtain a parabolic equation for the two-component vector field wavefunction
[22–24]. Using the same method, the equation for a three-component wave equation can be derived:
is the unperturbed Hamiltonian
corresponding to the first two terms in the equation (1),
and
are the perturbations
corresponding to the third term in the equation (1),
Consider a rotationally symmetric cylindrical
waveguide with a parabolic distribution of the refractive index:
where n0 is the refractive index
on the waveguide axis, ω is the
gradient parameter, .
The Hamiltonian may be rewritten in terms of
annihilation and creation operators [36]:
,
,
,
,
,
,
,
.
These operators satisfy the
commutation relations:
,
.
Here
,
,
,
,
,
,
,
,
,
is the unit matrix and
,
,
,
,
.
Representation of the Hamiltonian
by means of the operators will allow us to calculate the matrix elements
analytically. Note that generalized Stokes vectors consisting of nine real
parameters in terms of vector and tensor operators are considered to completely
describe three-dimensional fields
[47]
.
The solution of the unperturbed
equation is described by radially symmetric Laguerre-Gauss functions
:
where
is the principal quantum number, p
and
are the radial and azimuthal
indices, accordingly, and
or 0,
,
is the radius of the fundamental
mode.
The numbers
and
express the eigenvalues of the
unperturbed Hamiltonian
, and eigenvalues
of the angular momentum operator
.
It was shown in [36,39] that the hybrid wave
functions consisting of transverse and longitudinal components are the
solutions of the equation (2):
where
and
correspond to right-handed and
left-handed circularly polarized beams, accordingly, and
corresponds to the linear
polarization.
The propagation constant, which
takes into account non-paraxial terms of the first
order, is given by the expression [39,48]:
where , is the refractive index on the waveguide axis, is the gradient parameter, is the total angular momentum, σ is the spin angular momentum.
The term in (7) relates to the spin-orbit and spin-spin
interactions.
Consider the incident vector
vortex beams with right-circular and left-circular polarizations, accordingly: and , where is given by (5), and , is the radius of a beam which is
different from the radius of the fundamental mode of the medium .
The arbitrary incident beam may be
expanded into modal solutions, so the evolution of a beam in the medium (3) can be represented as
where
are the coupling coefficients.
If the incident beam is described
by the Laguerre-Gauss function
, the coupling coefficients
can be calculated analytically:
where
are the Jacobi polynomials,
,
,
.
Figure 1.
Beam width change with distance. l = 0, σ = 0, . Dashed line – paraxial approximation. (a) μm; (b) μm.
Figure 1.
Beam width change with distance. l = 0, σ = 0, . Dashed line – paraxial approximation. (a) μm; (b) μm.
Figure 2.
Intensity profiles of the transverse electric field component in axial direction. l = 0, σ = 0, μm. nonparaxial; (b) – paraxial approximation; (c) – intensity profiles at a second focus plane: black line – nonparaxial, red line – paraxial; (d) – intensity profiles at a third focus plane: black line – nonparaxial, red line – paraxial.
Figure 2.
Intensity profiles of the transverse electric field component in axial direction. l = 0, σ = 0, μm. nonparaxial; (b) – paraxial approximation; (c) – intensity profiles at a second focus plane: black line – nonparaxial, red line – paraxial; (d) – intensity profiles at a third focus plane: black line – nonparaxial, red line – paraxial.
Figure 3.
Intensity profiles of the transverse electric field component (left column) and the longitudinal electric field component (right column) for the circularly polarized incident beam with zero radial index in the focal plane : (a, b) ; (c, d) .
Figure 3.
Intensity profiles of the transverse electric field component (left column) and the longitudinal electric field component (right column) for the circularly polarized incident beam with zero radial index in the focal plane : (a, b) ; (c, d) .
Figure 4.
Intensity profiles of the transverse electric field component (left column) and the longitudinal electric field component (right column) for the circularly polarized incident beam with zero radial number in the focal planes : (a, b) ; (c, d) .
Figure 4.
Intensity profiles of the transverse electric field component (left column) and the longitudinal electric field component (right column) for the circularly polarized incident beam with zero radial number in the focal planes : (a, b) ; (c, d) .
Figure 5.
Intensity profiles of the transverse electric field component (left column) and the longitudinal electric field component (right column) for the circularly polarized incident beams with zero radial number in the focal planes : (a, b) ; (c, d) .
Figure 5.
Intensity profiles of the transverse electric field component (left column) and the longitudinal electric field component (right column) for the circularly polarized incident beams with zero radial number in the focal planes : (a, b) ; (c, d) .
Figure 6.
Intensity profiles of the transverse electric field component (left column) and the longitudinal electric field component (right column) for the circularly polarized incident beams with zero radial number in the focal planes : (a, b) ; (c, d) .
Figure 6.
Intensity profiles of the transverse electric field component (left column) and the longitudinal electric field component (right column) for the circularly polarized incident beams with zero radial number in the focal planes : (a, b) ; (c, d) .
Figure 7.
Intensity profiles of the transverse electric field component (left column) and the longitudinal electric field component (right column) for the circularly polarized incident beams with nonzero radial number p = 1 in the focal plane : (a, b) ; (c, d) .
Figure 7.
Intensity profiles of the transverse electric field component (left column) and the longitudinal electric field component (right column) for the circularly polarized incident beams with nonzero radial number p = 1 in the focal plane : (a, b) ; (c, d) .
Figure 8.
Delay times as a function of radial (a) and azimuthal (b) indices, accordingly, z = 1 km, , σ = 0.
Figure 8.
Delay times as a function of radial (a) and azimuthal (b) indices, accordingly, z = 1 km, , σ = 0.
Figure 9.
Relative delay times as a function of topological charge: z = 1 km, ; 1 - delay between beams with σ = -1 and σ = 1 ; 2 - ; 3 - ; 4 - . (a) μm-1, (b) μm-1. Subindex in corresponds to the spin angular momentum σ = 0, 1, −1.
Figure 9.
Relative delay times as a function of topological charge: z = 1 km, ; 1 - delay between beams with σ = -1 and σ = 1 ; 2 - ; 3 - ; 4 - . (a) μm-1, (b) μm-1. Subindex in corresponds to the spin angular momentum σ = 0, 1, −1.