1. Introduction
Quantum field theory in anti-de Sitter (AdS) spacetime is an active area of research. This activity is motivated by several reasons. First of all, the corresponding geometry is maximally symmetric and a sufficiently large number of problems are exactly solvable on its background. This helps to give an idea of the influence of the classical gravitational field on quantum phenomena in more complicated geometries. Qualitatively new features in the dynamics of quantum fields on the AdS background are related to the lack of global hyperbolicity and to the existence of both regular and irregular modes. In particular, boundary conditions on propagating fields need to be imposed at timelike infinity to prevent loss of unitarity. The different boundary conditions define different theories in the bulk. Another new feature, which has no analog in Minkowskian field theories, is related to the possibility of regularization for infrared divergences in interacting field theories without reducing the symmetries [
1]. This is closely related to the natural length scale of the AdS spacetime. The high interest to the AdS geometry is also related to its natural appearance as a ground state in supergravity and as the near horizon geometry for extremal black holes, black strings and domain walls.
The AdS spacetime plays a fundamental role in two exciting developments of contemporary theoretical physics. The first one, the AdS/CFT correspondence (for reviews, see [
2,
3,
4]), establishes duality between string theories or supergravity in the AdS bulk and a conformal field theory localized on the AdS boundary. This duality provides an interesting possibility for the investigation of nonperturbative effects in both sides of the correspondence by using the weak coupling regime of the dual theory. The recent developments include applications in various condensed matter systems such as holographic superconductors and topological insulators [
5,
6]. The second development is related to various types of braneworld models with large extra dimensions [
7]. In the corresponding setup, the standard model fields are localized on a brane embedded in a higher dimensional AdS spacetime. Braneworlds naturally appear in string/M-theory context and have been initially proposed for a geometrical resolution of the hierarchy problem between the electroweak and Planck energy scales. They provide an alternative framework to address the problems in particle physics and cosmology from different perspectives.
An inherent feature of field theories in AdS/CFT correspondence and in braneworld models is the need to impose boundary conditions on fields propagating in the AdS bulk. They include the conditions on the AdS boundary and the conditions on the branes in braneworld scenario. In particular, in braneworld models of the Randall–Sundrum type the boundary conditions on the branes are dictated by the
-symmetry. In quantum field theory, the boundary conditions modify the spectrum of the zero-point fluctuations and, as a consequence, the vacuum expectation values (VEVs) are shifted by an amount that depends on the bulk and boundary geometries and also on the boundary conditions. This is the well-known Casimir effect [
8,
9,
10,
11]. In braneworld models the Casimir forces acting on the branes may provide a mechanism for stabilization of the brane location (for mechanisms of moduli stabilization in warped geometries, see, e.g., [
12,
13,
14,
15,
16,
17,
18] and references therein). In particular, this stabilization is required to prevent the variations of physical constants on the branes. In addition, the quantum effects of bulk fields generate a cosmological constant on the brane. Motivated by these points, the Casimir effect in braneworld models on the AdS bulk, with branes parallel to the AdS boundary, has been investigated for scalar [
19,
20,
21,
22,
23,
24,
25,
26,
27,
28,
29,
30,
31,
32,
33,
34,
35,
36,
37,
38], fermionic [
39,
40,
41,
42,
43] and vector fields [
44,
45,
46,
47,
48,
49]. The models with de Sitter branes have been discussed in [
50,
51,
52,
53,
54,
55,
56,
57,
58]. The brane-induced quantum vacuum effects in AdS spacetime with additional compact subspaces were considered in [
59,
60,
61,
62,
63,
64,
65].
The main part of the papers on the Casimir effect in the AdS bulk consider global quantities, such as the Casimir energy and the forces acting on the branes. The local quantities carry more detailed information about the properties of the quantum vacuum. In particular, the expectation value of the energy–momentum tensor is of special interest. It appears as the source in the semiclassical Einstein equations and therefore plays an important role in modeling self-consistent dynamics involving the gravitational field. The VEV of the energy–momentum tensor for a conformally coupled scalar field in conformally-flat geometries has been investigated in [
28]. Massive scalar fields with general curvature coupling in the geometry of two branes on AdS bulk were considered in [
66,
67]. The Casimir densities for a
-symmetric thick brane for the general case of static plane symmetric interior structure have been discussed in [
68]. The VEVs of the energy–momentum tensor for Dirac spinor field and for the electromagnetic field are investigated in [
42,
43,
46,
47,
48,
49]. The geometry of a brane intersecting the AdS boundary is considered in [
69]. For a scalar field with general curvature coupling, the background AdS geometry with an additional compact subspace is discussed in [
61,
62].
In the present paper, we review the results for the VEV of the energy–momentum tensor in the geometry of a single brane on background of
-dimensional AdS spacetime. The cases of scalar, Dirac and electromagnetic fields are considered. The organization of the paper is as follows. In
Section 2, for a planar brane parallel to the AdS boundary, we consider complete sets of orthonormalized mode functions for both the regions between the AdS boundary and the brane and between the brane and the AdS horizon. The VEVs of the energy–momentum tensors for scalar, Dirac and electromagnetic fields are investigated in
Section 3. The behavior in the asymptotic regions of the parameters is discussed in detail.
Section 4 considers the VEV of the surface energy–momentum tensor for a scalar field on a brane parallel to the AdS boundary.
Section 5 is devoted to the study of the vacuum energy–momentum tensor for a scalar field in the geometry with a brane perpendicular to the AdS boundary. The main results are summarized in
Section 6.
2. Mode Functions in the Geometry with a Brane Parallel to the AdS Boundary
In Poincaré coordinates
, with
,
, the line element for the
-dimensional AdS spacetime is presented as
where
a is the curvature radius of the background geometry sourced by a negative cosmological constant
. For the curvature scalar and the Ricci tensor, one has
and
, with the metric tensor
defined from (
1). Introducing a new coordinate
z,
, in accordance with
, the line element is written in conformally flat form
where
and
is the metric tensor for
-dimensional Minkowski spacetime. The hypersurfaces
and
correspond to the AdS boundary and horizon, respectively. In what follows, we work in the coordinate system defined by (
2).
We are interested in the effects on the local properties of the quantum vacuum induced by a codimension one brane. First, we consider the case where the brane is parallel to the AdS boundary and is located at . It divides the space into two subspaces: the region between the AdS boundary and the brane, (L-region), and the region between the brane and AdS horizon, (R-region). The brane has a nonzero extrinsic curvature tensor and, as a consequence, the properties of the vacuum state in the L- and R-regions are different. The evaluation of the VEVs for local physical observables in those regions requires different procedures and we will discuss them separately. The VEVs are presented in the form of mode-sums over complete set of mode functions for quantum fields and we start by considering the modes for scalar, Dirac and electromagnetic fields.
2.1. Scalar Field
First, we consider a scalar field
with the mass
m. Assuming a general curvature coupling with the parameter
, the field equation reads
where
is the covariant derivative operator. The most important special cases correspond to minimally (
) and conformally (
) coupled fields. Let
be a complete set of positive and negative energy mode functions specified by the set of quantum numbers
. In accordance with the problem symmetry, they can be presented in the form
where
,
, and
. Plugging into the field equation, we get an ordinary differential equation for the function
:
where the energy
is expressed in terms of
and
as
. The general solution of (
5) is expressed in terms of the Bessel and Neumann functions
and
:
where
and
are constants and
To have a stable vacuum we assume the values of the parameters in (
7) for which
(see [
70,
71]). For a conformally coupled massless field, one has
and the mode functions (
4) with (
6) are conformally related to the modes in the Minkowski bulk. The scalar modes are normalized by the condition
where
g is the determinant of the metric tensor. Here,
is the Kronecker delta in the problems with discrete eigenvalues of the quantum number
and
in the problems with continuous spectrum for
.
We are interested in the effects of a codimension one brane, localized at
, on the local properties of the scalar vacuum. The Robin boundary condition is imposed for the field operator on the brane:
where
is the inward pointing normal to the brane and
is a constant. The latter encodes the properties of the brane. In the special cases
and
, the condition (
9) is reduced to the Dirichlet and Neumann boundary conditions, respectively. For the normal in (
9) one has
, where J = L,
in the L-region and J = R,
in the R-region. In general, the values of the constant
could be different for those regions.
First, let us consider the modes in the R-region. From the boundary condition (
9), it follows that
for the coefficients in (
6). Here and below, for a given function
, we use the notation with the bar defined in accordance with
with the coefficients
The mode functions in the R-region obeying the boundary condition (
9) are presented as
with the function
The spectrum for
is continuous, and, from the normalization condition (
8) with
, we get
The modes are specified by the set
with
,
, and
. The analog of the mode functions (
12) in the region between two parallel branes on the AdS bulk has been considered in [
67].
Note that we could also have modes with purely imaginary
,
. For those modes,
, where
is the MacDonald function (the modes with the modified Bessel function
are not normalizable). From the boundary condition (
9), we get the equation for the allowed values of
:
. The energy corresponding to these modes is given by
and it becomes imaginary for
. This leads to the instability of the vacuum state. To exclude the unstable modes, we restrict the allowed values for the Robin coefficient in the region where the equation
has no roots. It can be seen that, for non-Dirichlet boundary conditions (
), the corresponding condition is expressed as
(for more detailed discussion in models with compact dimensions, see [
72]). For
there is a single root. For the special value
, there exists a mode with
and with the mode functions
. For a minimally coupled massless scalar field, one has
and this special value corresponds to the Neumann boundary condition. The corresponding mode functions do not depend on
z.
In the L-region, the integration over
z in (
8) goes over the interval
. For the solutions (
4) with (
6) and
, the
z-integral diverges at the lower limit
in the range
. Hence, in that range, from the normalizability condition for the mode functions it follows that we should take
. In the range
, the solution (
6) with
is normalizable and in order to uniquely define the mode functions an additional boundary condition at the AdS boundary is required [
70,
73]. The general class of allowed boundary conditions has been discussed in [
74,
75]. In particular, they include the Dirichlet and Neumann boundary conditions, the most frequently used in the literature. Here, for the values of the parameters corresponding to the range
, we choose the Dirichlet condition, which gives
. With this choice, the mode functions in the L-region are specified as
From the boundary condition (
9) on the brane we get the equation for the eigenvalues of the quantum number
:
If we denote by
the positive zeros of the functions
, numerated by
, then the eigenvalues are given by
. Note that the roots
do not depend on the location of the brane. From the normalization condition (
8), with
and with the
z-integration over
, one finds
with the function
.
Similar to the R-region, the stability condition for the vacuum state in the L-region imposes restrictions to the allowed values of the Robin coefficient
. That condition excludes the presence of purely imaginary roots
for the Equation (
16). It can be shown that there are no such modes for
and there is a single mode for
. As seen, the stability condition in the L-region is less restrictive than that for the R-region. In the special case
one has a mode with
with the mode functions
.
2.2. Dirac Field
Now, we turn to a massive Dirac field
. The dynamics of the field is governed by the Dirac equation
where
is the spin connection. The curved spacetime Dirac matrices
are expressed in terms of the flat spacetime matrices
by the relation
, with
being the tetrad fields. In the coordinate system corresponding to (
2), the latter can be taken as
. For the components of the spin connection, this gives
and
for
. In an irreducible representation of the Clifford algebra, the matrices
are
matrices, where
and the square brackets in the exponent mean the integer part. Up to a similarity transformation, the irreducible representation is unique in odd numbers of spatial dimension
D. For even values of
D, one has two inequivalent irreducible representations. We use the flat spacetime gamma matrices in the representation
and
, where
. In odd spatial dimensions, the two values of the parameter
s correspond to two inequivalent representations. The commutation relations for the
matrices
and for their hermitian conjugate matrices
are obtained from those for the Dirac matrices
. They are reduced to the relations
,
,
and
,
, for
. The representation (
19) for the construction of the gamma matrices in AdS spacetime is considered in [
76]. Another representation is taken in [
43].
Assuming the dependence on the coordinates
in the form
and decomposing the spinor
into the upper and lower components, in the representation (
19), the corresponding equations are separated. The dependence of those components on the
z-coordinate is expressed in terms of the function
, where the upper and lower signs correspond to the upper and lower components. The coefficients are determined by the normalization condition and by the boundary condition on the brane at
. The positive and negative energy fermionic modes
, specified by the set of quantum numbers
, are normalized by the condition
As in the case of a scalar field, here
is understood as the Dirac delta function for the continuous components of
and the Kronecker delta for discrete ones. On the brane at
, we impose the bag boundary condition
where
, J = R,L, with
for the L- and R-regions.
In the R-region,
, from the boundary condition (
21), for the ratio of the coefficients in the linear combination of the Bessel and Neumann functions, one finds
. The positive and negative energy mode functions, obeying the boundary condition (
21), are expressed as
where
and
In (
22), the one-column matrices
,
are introduced with the elements
and having
rows. The normalization constant is determined from the condition (
20):
The set of quantum numbers
is specified as
.
In the L-region and for
, from the normalizability condition, we get
. In the range of the mass corresponding to
, the modes with
are normalizable as well, and an additional boundary condition is required to uniquely define the mode functions. Here, we consider a special case that corresponds to the choice
for all values of the mass. The fermionic modes are given by the expressions
The allowed values of
are determined from the boundary condition (
21) on the brane. They are roots of the equation
and are expressed as
. The normalization coefficient is obtained from (
20) and is given by
Note that the eigenvalues for
are the same for both the representations of the Clifford algebra.
2.3. Electromagnetic Field
For the electromagnetic field with the vector potential
,
, the field equation reads
where
is the field strength tensor. To find a complete set of mode functions
for the vector potential, we impose the Lorenz condition
and an additional gauge condition
. For the positive energy modes, presenting the dependence on the coordinates
in the form
, from the field equation, we can see that
Here,
correspond to different polarizations specified by the polarization vector
. For the latter, one has the normalization condition
and the constraints
and
. The latter two relations follow from the gauge conditions. The modes (
29) are normalized by the condition
where the set of quantum numbers is given by
.
The coefficients
and
in (
29) are determined from (
30) and from the boundary condition on the brane. We consider two types of gauge invariant constraints. The first condition is the analog of the perfect conductor boundary condition in
Maxwell electrodynamics and is given by
with
being the dual of the field tensor and
is the normal vector to the boundary. The second boundary conditions is expressed as
This type of condition has been used in quantum chromodynamics to confine the gluons. In the gauge under consideration and for the mode functions (
29), the boundary condition (
31) yields
, whereas the condition (
32) gives
.
In the R-region, from the boundary condition on the brane, we get
, where
The corresponding mode functions for the vector potential are presented as
where the function
is given by (
23). The normalization constant is found from (
30):
with
.
In the L-region, from the normalizability of the mode functions, it follows that in (
29)
for
. For
, an additional condition on the AdS boundary is required in order to uniquely define the modes. Here, we consider a special case with
for
as well. For the mode functions, we obtain
The allowed values for
are determined by the boundary condition on the brane and they are the roots of the equation
with
from (
33). Hence, we have
. For the normalization coefficient, one gets
The mode functions in the region between two branes have recently been considered [
49].
2.4. Boundary Conditions in -Symmetric Braneworlds
An example of the
-symmetric braneworld is provided by the Randall–Sundrum model with a single brane (RSII model) formulated in background of (4+1)-dimensional AdS spacetime (see [
77,
78] and the review [
7] for the RSI and RSII models). For an arbitrary number of spatial dimensions, the line element is given by (
1) with
replaced by
. The regions
and
are identified by the
-symmetry. The brane is located at
. Hence, in the corresponding setup, two copies of the R-region are employed with
. The boundary conditions on the bulk fields at the location of the brane are obtained by integrating the field equations about
(see, e.g., the discussions in [
26,
40,
67,
76,
79,
80]).
For scalar fields, even under the reflection with respect to the brane, the Robin boundary condition is obtained with the coefficient
, where
is the brane mass term. The latter appears in the part of the action located on the brane,
. For odd scalar fields, the Dirichlet boundary condition is obtained. For fermionic fields two types of boundary conditions are obtained. The first one is reduced to the bag boundary condition (
21) and the second one is obtained from (
21) by the change of the sign in the term containing the normal to the brane. For vector fields, even under the reflection with respect to the brane, the boundary condition is reduced to (
32) and for odd fields the condition (
31) is obtained. For fermionic fields with the boundary condition obtained from (
21) by the change of the sign in the second term, the VEV of the energy–momentum tensor is evaluated in a way similar to that we have demonstrated for the bag boundary condition. The corresponding mode functions are obtained from (
22) by the replacement
in the first index of the functions
,
, and in the expression (
24) for the normalization coefficient.
Hence, we conclude that the VEVs of the energy–momentum tensors for scalar, Dirac and electromagnetic fields in -symmetric braneworlds with a single brane are obtained from the results given below for the R-region by an appropriate choice of the boundary conditions on the brane. The only difference is that an additional factor should be added. The latter is related to the presence of two copies of the R-region.
4. VEV of the Surface Energy—Momentum Tensor for a Scalar Field
In the discussion above, we consider the VEV of the bulk energy–momentum tensor. In manifolds with boundaries, in addition to the latter, a surface energy–momentum tensor may present, which is localized on the boundaries. In the general case of bulk and boundary geometries, the expression of the surface energy–momentum tensor for a scalar field with general curvature coupling has been obtained in [
96] by using the standard variational procedure. The VEV of the surface energy–momentum tensor for branes parallel to the AdS boundary is investigated in [
97] by using the generalized zeta function technique.
For a given field, the expression for the surface energy–momentum tensor
, in addition to the bulk action, depends on the surface action. In [
96], for a spacetime region
M with boundary
, the surface action for a scalar field is taken in the form
where
for spacelike and
for timelike elements of the boundary, and
h is the determinant of the induced metric
, with
being the inward pointing unit normal to
,
. In (
82),
is the trace of the extrinsic curvature tensor
of the boundary and
is a parameter.
consists of the initial and final spacelike hypersurfaces and a timelike smooth boundary
. The variation of the total action with respect to the field
leads to the standard field Equation (
3) in the bulk and to the boundary condition
The variation of the action with respect to the metric tensor gives the metric energy–momentum tensor. In addition to the bulk part, the latter contains a contribution
located on the boundary
:
, where
is the ’one-sided’
-function. By using the boundary condition (
83), the expression for
is presented in the form [
96]
Note that the boundary condition (
83) is of the Robin type. By using the boundary condition, one can exclude the derivative term for the field in (
84).
In the geometry (
2) with a single brane at
, the extrinsic curvature tensor for the R- and L-regions (J = R,L) has the form
for
, and
. The boundary condition (
83) is reduced to (
9) with
. The VEV of the surface energy–momentum tensor,
, is evaluated by using the mode-sum formula
with the mode functions given by (
12) and (
15) for the R- and L-regions, respectively. The VEV has the form
,
,
and, from the point of view of an observer living on the brane it corresponds to a gravitational source of the cosmological constant type. An essential difference compared with the bulk energy–momentum tensor is that the subtraction of the part corresponding to the geometry without a brane is not sufficient and an additional renormalization is required. The latter is reduced to the renormalization of the VEV for the field squared on the brane. In [
97], the generalized zeta function technique is used.
The VEV of the surface energy–momentum tensor for the region J = R,L,
, is expressed in terms of the VEV of the field squared on the brane,
, as
The VEV
is obtained by the analytic continuation of the function (the details can be found in [
97])
to the physical point
. Here, the parameter
is the renormalization scale,
is defined by (
7) and
The representation (
86) is valid in the slice
of the complex
s-plane. In the first step of the analytic continuation, the integral in (
86) is presented in the form of the sum of the integrals over the regions
and
. In the first integral the substitution
can be made directly. In the second integral, we subtract and add to the function
in the integrand the
N leading terms of the corresponding asymptotic expansion for large values of
x and integrate the asymptotic part. For
the asymptotic expansion has the form
, where the coefficients
are found from those for the expansions of the modified Bessel functions. The function
has a simple pole at
and the leading term in the Laurent expansion is given by
In this way, the VEVs
for J = R and J = L are decomposed into the pole and finite contributions. The pole terms can be absorbed by adding to the brane action the respective counterterms. The expressions for the finite parts in separate regions will not be given here and can be found in [
97]. We will consider the total energy density.
Combining the results for the R- and L-regions, one obtains the total surface energy density
. Comparing the pole parts (
88) for
in those regions, we can see that in odd spatial dimensions the pole parts in the energy density cancel out and the finite part does not depend on the renormalization scale
. Taking
for the number of the terms taken in the asymptotic expansions of the function
, for the total surface energy density in odd dimensions
D, one gets the formula
Note that this quantity does not depend on the location of the brane. Depending on the value of the Robin coefficient
, the surface energy density (
89) can be either positive or negative (see the graphs in [
97] for minimally and conformally coupled scalar fields).
In the geometry of two branes, the VEV of the surface energy–momentum tensor on a given brane is decomposed into two parts. The first one corresponds to the VEV in the problem where the second brane is absent and the corresponding evaluation procedure has been described in this section. The second part is induced by the presence of the second brane and it requires no additional renormalization. As it has been discussed in [
97], in the Randall–Sundrum model the surface energy density induced on the visible brane by the presence of the hidden brane gives rise to naturally suppressed cosmological constant. The surface energy density in models with additional compact dimensions is discussed in [
98,
99] for neutral and charged scalar fields. In the latter case, the value of the induced cosmological constant on the brane is additionally controlled by tuning the magnetic flux enclosed by compact dimensions.
5. Geometry with a Brane Perpendicular to the AdS Boundary
In a number of recent developments of the AdS/CFT correspondence, branes intersecting the AdS boundary are considered. They include the extensions of the correspondence for conformal field theories with boundaries (AdS/BCFT correspondence) [
100,
101] and the geometric procedure for the evaluation of the entanglement entropy for a bounded region in CFT [
102,
103]. In this section, based on [
69], we consider the effects on the scalar vacuum induced by a brane perpendicular to the AdS boundary.
The background geometry is described by the line element (
2) and the brane is located at
. The problem is symmetric with respect to the brane and we will consider the region
. The scalar field
obeys the field Equation (
3) and the boundary condition
on the brane. Introducing the notations
and
, the normalized mode functions obeying the boundary condition have the form
where
, the energy is given by
with
and
is defined by (
7). The function
is determined from the relation
Note that in the case
we impose the same boundary condition on the AdS boundary as that in
Section 2 for the L-region.
For positive values of the Robin parameter
in (
90), in addition to (
91), there is a mode for which the part depending on the coordinate
is expressed in terms of the exponential function
. The corresponding energy is given by
. In the subspace
of the quantum numbers, the energy is imaginary and the vacuum state is unstable. Note that for a Robin plate in the Minkowski bulk the energy for the corresponding bound states is positive in the range
. To have a stable vacuum in the AdS bulk, we assume that
.
With the mode functions (
91), the VEV of the energy–momentum tensor is evaluated by using the mode-sum formula (
40) (for the procedure based on the point-splitting regularization technique see [
69]). The mode-sum contains the integration over the region
. In the integrand, we write the parts containing the products of the trigonometric functions with the arguments
in the form of the sum of three terms. The first one does not depend on
and its contribution corresponds to the VEV in the AdS spacetime when the brane is absent. The second and third terms depend on the
-coordinate in the form of the exponents
and
. By taking into account that those terms exponentially decay in the upper and lower half-planes of the complex variable
, we rotate the integration contour over
by the angle
for the part with the exponent
and by the angle
for the part with
. The VEV of the energy–momentum tensor is presented in the decomposed form (
41), where the diagonal components of the brane-induced part are given by the expression (no summation over
)
with
,
,
for
. The operators
are defined as
with
and
is given by (
56). In (
93), we introduce the function
where
is the hypergeometric function. The diagonal components in the region
are given by the expression (
93) with
replaced by
.
An important difference from the geometry with a brane parallel to the AdS boundary is the presence of the off-diagonal component of the vacuum energy–momentum tensor:
This expression for the off-diagonal component in the region
is obtained from (
96) changing the sign and replacing
. Note that
. Due to the nonzero off-diagonal component
, the Casimir force acting on the brane has two components. The first one is determined by the stress
and corresponds to the component normal to the brane. The second one is obtained from
and is directed along the
z-axis. It corresponds to the shear force. Of course, because of the surface divergences in the local VEVs, both these components require an additional renormalization. Note that the Casimir forces acting tangential to the boundaries (lateral Casimir forces) may arise also in condensed matter systems if the properties of the corresponding surfaces are anisotropic or inhomogeneous (see, for example, [
104,
105] and references therein). In the problem under consideration, the tangential force is a consequence of the
z-dependence of the background geometry.
From the expressions (
93) and (
96), it follows that for the Dirichlet and Neumann boundary conditions the brane-induced contributions to the VEV of the energy–momentum tensor differ by the signs. In these special cases the corresponding expressions are further simplified as (no summation over
)
where the upper and lower signs correspond to the Dirichlet and Neumann boundary conditions, respectively, and
. The operators
in the expressions for the diagonal components are given by
for
, and
The function
is expressed in terms of the hypergeometric function
:
For a conformally coupled massless field, one has
and the problem under consideration is conformally related to the problem in the Minkowski spacetime with the line element
,
, and with planar codimension one boundaries located at
(the conformal image of the AdS boundary) and
(the conformal image of the brane). The Minkowskian field obeys the Dirichlet boundary condition at
and the Robin boundary condition (
90) at
. The Dirichlet boundary condition at
is a consequence of the condition we have imposed on the AdS boundary. Taking
, from the results given above, we can obtain the VEV of the energy–momentum tensor for a conformally coupled massless field in the geometry of perpendicular planar boundaries in the Minkowski bulk by using the relation
(see (
45)). In the Minkowskian limit we get the result
57 (with the replacement
) for the components
.
Near the brane and for non-Dirichlet boundary conditions,
, the leading term in the asymptotic expansions for diagonal components with
is given by (no summation over
)
This coincides with the Minkowskian result where the distance from the boundary is replaced by the ratio
(compare with (
58)). Note that, in accordance with (
2), the latter is the proper distance from the brane measured by an observer at rest with respect to the brane. Note that the proper distance
is different from the geodesic distance
. The latter between the spacetime points
and
is given as
. For the normal stress and for the off-diagonal component near the brane one gets
The leading terms for the Dirichlet boundary condition differ from the ones given above by the sign. The expressions (
101) and (
102) also describe the asymptotic behavior of the brane-induced VEV near the AdS horizon (large values of
z for fixed
).
Now, let us consider the asymptotics at large distances from the brane,
. For non-Neumann boundary conditions, additionally assuming that
, for the components
, the leading term is given by (no summation over
)
The asymptotics for the remaining components are expressed as
As seen, the leading order terms for
coincide with those for the Dirichlet boundary condition. For a scalar field with the Neumann boundary condition (
), the leading terms differ from those for the Dirichlet condition by the signs. Hence, the Dirichlet boundary condition is the attractor in a class of Robin boundary conditions with
. At large distances, the brane-induced contributions, considered as functions of the proper distance from the brane, exhibit a power-law fall-off for both massless and massive fields. This behavior is in clear contrast with the case of the Minkowski bulk, where the boundary-induced VEVs decay exponentially for massive fields, as
(see (
57)). Note that for large
one has the relation
, with
being the geodesic distance. For fixed
, the asymptotic formulas (
103) and (
104) describe the behavior of the VEVs near the AdS boundary. The diagonal components decay as
, whereas the off-diagonal component tends to zero as
. The qualitative behavior of the brane-induced energy density, as a function of the distance from the brane, is similar to what we describe in the previous section for a scalar field with the Robin boundary condition on the brane parallel to the AdS boundary.
6. Summary
We consider the influence of a brane in AdS bulk on the properties of quantum vacuum. Two geometries are discussed: (i) a brane parallel to the AdS boundary; and (ii) a brane perpendicular to the AdS boundary. In the first geometry, as a local characteristic of the vacuum state, the VEV of the energy–momentum tensor is investigated for scalar, Dirac and electromagnetic fields. For calar field, a general Robin boundary condition is considered and the Dirac field is constrained by the bag boundary condition. In the case of the electromagnetic field, two types of boundary conditions are discussed. The first one corresponds to the perfect conductor boundary conditions in 3D electrodynamics and the second one is the analog of the boundary condition used in bag models of hadrons to confine the gluons. The VEV of the energy–momentum tensor is expressed as a mode-sum over complete set of mode functions and for all these cases the corresponding sets are given. The brane divides the background geometry into two regions: the region between the brane and AdS horizon (R-region) and the region between the brane and AdS boundary (L-region). Although the AdS spacetime is homogeneous, the brane has a nonzero extrinsic curvature tensor and the properties of the quantum vacuum in those regions are different. In particular, the spectrum of the quantum number , corresponding to the momentum along the direction normal to the AdS boundary, is continuous in the R-region and discrete in the L-region. In the latter region, the eigenvalues are zeros of cylinder functions. The mode-sum for the VEV of the energy–momentum tensor contains series over those zeros and for the summation we have employed the generalized Abel–Plana formula. That allows extracting from the VEV the part corresponding to the geometry without a brane and to present the brane-induced contribution in terms of integral, exponentially convergent for points away form the brane. A similar decomposition is provided for the R-region.
Near the brane, the leading terms in the asymptotic expansions for the energy density and parallel stresses coincide with the corresponding expressions for a single boundary in the Minkowski bulk, where the distance from the boundary is replaced by the proper distance from the brane on the AdS bulk. For those VEVs, the effect of gravity is weak. This is related to the fact that, near the brane, the main contribution to the corresponding VEVs come from the vacuum fluctuations with the wavelengths smaller than the curvature radius of the background geometry and influence of the gravitational field on those modes is weak. For a boundary in the Minkowski bulk, the normal stress is zero. The nonzero normal stress in the geometry of a brane on the AdS bulk is a purely gravitational effect. The effect of gravity on the brane-induced VEVs is essential at distances from the brane larger than the curvature radius. In particular, for the R-region, at large distances, the decay of the brane-induced contribution in the vacuum energy–momentum tensor, as a function of the proper distance, is exponential for both massless and massive fields. For the Minkowski bulk and for massless fields, the fall-off of the boundary-induced contribution is as power-law. On the AdS boundary, the brane-induced contributions tend to zero as
, where
for scalar field (with
given by (
7)),
for the Dirac field and
for the electromagnetic field. Near the AdS boundary, one has a simple relation between the energy density and the normal stress, given by
. This correspond to the barotropic equation of state for the vacuum pressure
along the
z-direction and vacuum energy density. Note that, for the pressures along the directions parallel to the brane, the equation of state is of the cosmological constant type. By using the generalized zeta function technique, we also investigate the VEV of the surface energy–momentum tensor. From the viewpoint of the observer living on the brane, the latter corresponds to a gravitational source of cosmological constant type. Depending on the value of the coefficient in the boundary condition, the induced cosmological constant can be either positive and negative.
The brane-induced effects on the quantum vacuum for Geometry (ii), we consider the example of a scalar field with general curvature coupling parameter. For the Robin boundary condition the mode functions have the form (
91). The diagonal components of the brane-induced energy–momentum tensor are given by the expressions (
93). An important difference from the problem with a brane parallel to the AdS boundary is the presence of nonzero off-diagonal component (
96) of the vacuum energy–momentum tensor. As a consequence, the Casimir force acting on the brane, in addition to the normal component, contains a component directed parallel to the brane (shear force). At large distances from the brane, the decay of the brane-induced contribution to the energy–momentum tensor, as a function of the proper distance from the brane, is as power-law for both massive and massless field. As mentioned above, in the Minkowski bulk, the decay for massive fields is exponential.
For charged fields, another important local characteristic of the vacuum state is the expectation value of the current density. This VEV in models with local AdS geometry and with toroidally compact spatial dimensions, in the presence of single and two branes has been investigated in [
72,
76,
106,
107] for charged scalar and fermionic fields. The vacuum currents have nonzero components along the compact dimensions only. They are periodic functions of the magnetic flux with the period equal to the flux quantum. Depending on the boundary conditions imposed on the fields at the locations of the branes, the brane-induced effects lead to increase or decrease of the current density. Applications are discussed to Randall–Sundrum type braneworld models as well as curved graphene tubes.
In the discussion above, we assume that the background geometry is fixed. Among the interesting directions for the further research is the investigation of the back-reaction of quantum effects on the geometry by using the semiclassical Einstein equations with the VEV of the energy–momentum tensor in the right-hand side. The vacuum energy–momentum tensor may violate the energy conditions in the singularity theorems, and this leads to interesting cosmological dynamics of the bulk and on the brane. In this regard, the next step in the study of local quantum effects in braneworlds could be the investigation of the vacuum energy–momentum tensor in models with dS branes.