1. Introduction
The nonlinear sigma model (NLSM), first introduced in [
1] for a description of pion properties, has found generalizations which have become a very efficient tool in exploring diverse phenomena in quantum physics, including, e.g., magnetic systems [
2,
3,
4] as well as the conductivity of disordered systems [
5]. Soon after [
1], the importance of using (differential) geometric methods in the construction and study of general phenomenological Lagrangians (NLSMs) was recognized (see [
6,
7,
8]). Moreover, the importance of symmetric homogeneous (or coset) spaces was emphasized in [
8] as well as in [
5,
9]. On the other hand, a few works [
10,
11,
12] paid some attention to the potential role of nonsymmetric coset spaces as target spaces for the respective nonlinear sigma models.
Closely related to the well-known real Grassmann manifolds
are their well-known extensions [
13], such as the real Stiefel manifolds
. The latter are nothing but principal fiber bundles with a base being the Grassmannian
and the structure group
. We are interested in (quasi) two-dimensional sigma models—nonlinear field models with values in the Stiefel (or, in particular, Grassmann) manifolds. We note that the quantum properties of Grassmannian sigma models are quite well studied: in the two-dimensional case, such nonlinear models are characterized by asymptotic freedom [
2,
14], and if we extend to the dimension
, the models have a two-phase structure with a bicritical point [
15]. This behavior, together with the use of the replica method and replica limit, allowed the description (see, e.g., [
5]) of the (de)localization of electrons in disordered systems (or random potential) [
16]. As for the classical two-dimensional Grassmannian models, their integrability was proven for these nonlinear field models [
17].
Nonlinear sigma models on real Stiefel manifolds have radically different properties, both in the classical and especially quantum consideration [
11,
18], both for
and
. In particular, in contrast to the single-charge RG (scaling) behavior of Grassmannian sigma models, the class of
sigma models is characterized by a two-charge behavior, and the question of the presence of asymptotic freedom in them at
becomes quite nontrivial [
11,
18].
The extension of the
sigma model to the
-dimensional nonlinear
-model (of the fields defined on real Stiefel manifolds) is also very desirable in view of the following important reasons.
First, since Grassmannians are Einstein manifolds, the single (effective) charge or coupling in the associated quantum sigma model is of pure geometrical origin—it is the homothety of metrics, i.e., uniform (or isotropic) scaling of a fixed metric. On the contrary, in the extended models on the spaces
the latter in general are not Einstein manifolds, and we come inevitably to the enriched set of couplings: besides homothety, a certain anisotropy of metrics comes into play. Then, the Einsteinian property may occur only very randomly (for one or two special values of the anisotropy).
Second, because of such nontrivial geometry of
, the related nonlinear sigma models in
do manifest more complicated critical behavior: besides/instead of bicritical point, a tetracritical point may also appear.
The third property is very important from the viewpoint of possible physical applications. Namely, we have in mind applying
sigma models to a more complex situation when the system manifesting (de)localization of electrons may also participate in superconductivity. It is clear that the involvement of an additional order parameter could then be related to the presence of a second coupling in the Stiefel model. Finally, the electron-hole symmetry that finds its natural reflection in the
symmetry of Grassmannian sigma models (see, e.g., [
15]) is known to be inevitably broken within the superconductivity context as noted, e.g., in [
19,
20]. However, just the absence of
symmetry is an intrinsic feature of the Stiefel
sigma models.
Because of this, and also in connection with the recent return of interest [
21,
22] to the geometry and other properties of Stiefel manifolds, we will focus on the special properties of the nonlinear sigma model in
Euclidean dimensions with the action functional
and the Lagrangian
:
where we have introduced the temperature
T, field
valued in the Stiefel manifold
, gradient
, and parameter (of the metric anisotropy)
.
2. Some Characteristics of Stiefel Manifold
A real Stiefel manifold is the set of all k-frames formed by sets of orthonormal vectors in the space . Any element , given by an -matrix, satisfies the orthogonality condition , where is the identity -matrix and ⊤ denotes the transpose. On such manifolds, the group acts transitively as the isometry group. The isotropy group corresponding to the “origin” consists of matrices of the form , where . This allows us to identify the manifold with the quotient space with . Thus, is defined as a submanifold of the space . In particular, ; is an -dimensional sphere; ; ; and is the first nontrivial case.
Tangent space. Let
be the tangent space for
. We require that each tangent vector
at a point
belongs to the
horizontal subspace, i.e., it has the following structure [
21,
22,
23]:
where the projector
, and
. Since
consists of
independent matrix components and
contains
components, their total sum restores
. Conversely, the symmetric component
belongs to the
vertical subspace.
Tangent vectors can be defined in the respective Lie algebra
, the basis of which is formed by
skew-symmetric
-matrices
:
where the multi-index
consists of the two subsets such that
. The matrix
has only two nonzero elements, 1 and
, at positions
and
, so that
and
. The basis elements of the algebra
satisfy the commutation relations:
In this basis, the orthonormal metric
based on the bi-invariant Killing form
, when
, coincides with the Frobenius metric, which is natural in the description of the
-model (
1):
Then, the norm of the vector
is written as
.
To write the
- and
F-components of tangent vectors in the basis introduced, we define
and
with
and
, where
and
. Although
in (
2), it has
independent components and is usually included in meaningful expressions as a combination
.
Metric. The covariant metric
(metric operator) of model (
1) and its action on the tangent vector
are defined as
Note that the case
is called
canonical in the literature [
22,
23].
Then the inner (scalar) product of two tangent vectors
is given as
Since
, according to (
2), the Lagrangian in (
1) becomes
Here we use the following decomposition of the tangent vector
:
because
due to the constraint
.
Expression (
8) indicates the presence of two
- and
F-subsystems. The condition
can be easily satisfied by reducing the system to an
-model with
and
. If
is expressed through
, then a model on the Grassmannian
is obtained. However, our focus here is on incorporating both subsystems, utilizing the full number of degrees of freedom,
.
Linear connection and normal coordinates. Consider the auxiliary problem of one-dimensional evolution, which is generated by the action functional:
where
is a Lagrange multiplier;
.
By varying
and
H and using certain identities, one can obtain the multiplier
and the equation of the geodesic [
21]:
To write the latter, we used the expression for the Christoffel function (symbol of the second kind), which defines the linear connection at the point
:
Now, we present the solution of Equation (
11) as a series in
t in terms of the initial data
and
, which are related by
. One has
where
, and
involves the derivative
at the point
in the direction
, which acts, in particular, as
Note that the expansions (
2) and (
9) can be applied
after calculating such a derivative.
It is easy to verify in the approximation
that
. Thus, setting
, the two points
are related by (
13). Restoring the dependence of
U and
V on the coordinates
x, formula (
13) generalizes the “field shift”
in Euclidean space when
. Thus, fixing
, the variables
V are called
normal coordinates, which further describe quantum fluctuations in the
-model.
Curvature. We define the curvature tensor
in
-form for
as (see [
22])
Let us calculate the components of
:
The bi-quadratic form
, which characterizes the sectional curvature [
24], can be expressed, after substituting expressions (
16) and (
17), as follows:
A similar formula is derived in the work ([
25], p. 405) in terms of rectangular
-matrices, say
M, instead of
. However, it does not matter, because
M and
F are entered as combinations
and
, which both belong to
.
Using (
16) and (
17) and contracting the
-curvature (see
Appendix A.1), the diagonal components of the Ricci
-tensor are obtained [
22]:
It is seen that the curvature coefficients are constant due to the homogeneity of the Stiefel manifold (see [
18]). We are discussing the relation between
and
when renormalizing the
-model.
3. Background Field Formalism
We assume that the Lagrangian
is determined by the bare metric
, which differs from
by the covariant counterterms
:
where
is the renormalization parameter;
is the renormalization scale.
This renormalization constitutes our primary concern, though the fields themselves are also subject to renormalization. To address this, we adapt a well-known procedure (see [
26,
27]) to the model defined on the Stiefel manifold. Initially, we present a covariant description utilizing normal coordinates.
Effective Lagrangian. Let us introduce the interpolating field
and quantities similar to those defined earlier:
Let us expand the Lagrangian
in a series with respect to the parameter
s:
using the covariant derivative
along the geodesic
, as well as
:
where the Christoffel function
from (
12) is taken at
.
Applying the decomposition of tangent vectors into
- and
F-components, we obtain
taking into account that
where the curvature
is given by (
15).
In the one-loop approximation, we restrict ourselves to the Lagrangian quadratic in
V:
where
, and
by definition. If
is a solution of the equation of motion, the term
does not contribute to the action integral. Since expression (
29) is general for
-models within this approximation, we proceed to the next step, which is standard for such models.
We replace
V by
to transform
into
. To do this, we set
, where
. Then
and
, and the relation between the components of the tangent vectors
V and
is given by
After introducing
instead of
, the quadratic part of the Lagrangian with respect to
V takes on the form
where the first term in (
32) corresponds to the free evolution of fluctuations and determines the Green’s functions. In this free Lagrangian, we retain only the
components of
as independent, since the
components merely represent the interaction of
with
(see
Appendix B.1).
Quantization. In fact, quantization here means calculating the partition function and average values using the methods of quantum theory, when the imaginary Planck constant
is replaced by the real temperature
T. We consider
as the background field, and
and
as fluctuating quantities. Simultaneously, the components of the metric
, including the positive parameters
T and
(see (
1)), play the role of coupling constants. In physical applications of
-models for describing Anderson localization, the temperature
T is often associated with the conductivity coefficient, which undergoes renormalization. On the other hand, the parameter
(anisotropy) appears in the Lagrangian (
8) as a running (scale-dependent) factor.
Quantizing, one should introduce a path integral over the rapidly varying fields
and
(with a source), which is calculated as a series using Feynman diagrams. In the case under consideration, all expressions correspond to the one-loop approximation. As usual, when averaging expressions, Wick’s theorem for pairing fields of the same sort does work and allows us to limit ourselves to considering only even terms. In principle, the integral over the slowly varying component
also needs to be applied, since it is ambiguous. Some computational aspects of this procedure are outlined in
Appendix B.1.
Let us define the Green’s functions of the fields:
where the brackets
denote the quantum average in the vacuum state; and the components of the matrix field
with multi-indices
, where
and
are chosen as independent. In addition, the symbol
, and the symbol
for multi-indices
,
, and their contraction gives the number of degrees of freedom. The spatially dependent component is defined by the expression
which results from (
32) and (
A27).
4. Beta Functions of Effective Couplings
Renormalization. Due to the form (
35), when averaging local expressions of the model, the infrared (IR) divergences arise, which can be mitigated by introducing a regulator. Softening the expressions at the IR boundary by using the scale
and denoting the area of the unit sphere in
d dimensions as
, when
, we have
At once, ultraviolet (UV) divergences are eliminated in a standard way for
-models [
28].
Thus, the metric renormalization in the one-loop approximation is induced by the term [
26,
27,
28]
The idea of proving formula (
37) is based on the equality of the quantum average of
over the fields
and
and the average over the basis matrices, when
(see
Appendix A.1). That is, due to the homogeneity of the Stiefel manifold, the calculations can be shifted to the “origin”
, allowing the use of the basis (
3). Then, we perform the substitution
,
, while the fields themselves are absorbed by
. We also express
in the basis (
3) and sum over
Q and the independent indices of
P. When averaging
, only the terms quadratic in
and
survive, and we obtain
The divergences of (
37) at
are eliminated by the scaling factors
’s for the metric components:
According to (
29), (
36) and (
38), we obtain
where we have defined the parameters
and
.
Beta functions. Renormalizing the model metric as
changes, we demand
These two conditions can be satisfied by assuming that both
t and
(or
) depend on
z. Otherwise, to renormalize only the temperature
t, we would have to require that the Stiefel manifold be Einsteinian when we equate
This is the same as
for
.
Defining the beta functions
and
that need to be found, we arrive at the set of exact equations:
Determining
’s up to order
t, they reduce in the one-loop approximation to
where
for
.
Note that taking into account all the terms in these equations using approximate ’s may change the picture of the renormalization group (RG) dynamics, but will not correct the parameters of the stable fixed point (sink) described below.
Thus, we arrive at
Replacing
with tantamount
, we rewrite it in an equivalent form:
The region of admissible values of the model parameters in the plane
is the quadrant with
and
. This region is divided into three subregions by two separatrices
and
for arbitrary
. The fixed points are found from the conditions
and
. Defining
we have the Gaussian point
, the node
, and the saddle point
(see
Figure 1a).
The trajectories of the renormalization group (RG) are given by the equations:
Then, the separatrices in
Figure 1, passing through the points
,
and
, are obtained numerically by integrating (
50) and (
51). It can be seen that in the case of
and
, the point
, at which the four phases meet, seems tetracritical (red circle in
Figure 1a). Note that at
, such tetracriticality disappears, and the point
becomes bicritical (green circle in
Figure 1c). Finally, at
and fixed
, phase transitions by varying temperature
t are not observed (see
Figure 1b,d), although the sink–saddle pair still exists.
By introducing the quantity
, associated with the conductivity in some models, and the corresponding beta function
, we complement the case in
Figure 1a with the behavior of
g in
Figure 2.
Critical exponents. As indicated above, the critical point—a stable fixed point (sink)—is given by the parameters
. By fixing
, and thereby reducing the Stiefel manifold to an Einstein manifold (with proportionality between the Ricci tensor and the metric), it is easy to calculate the critical exponent
(and
) in the one-loop approximation for
and
. Under these conditions, there exists a positive critical temperature
, the determination of which provides [
15]:
This agrees with the known result for the Grassmannian
with the Einstein property.
Focusing primarily on the specifics of renormalizing the metric rather than the field, based on the results in
Appendix B.2 we may estimate the exponent
similarly to [
4]:
We observe that
for large
N, and furthermore, when
, this expression yields the previously established value [
4].
Other critical exponents can be found using the scaling and hyperscaling relations:
Their values for the model in the one-loop approximation are given in the
Table 1. It is worth noting that just the rightmost column shows clearly the dependence on
k: that is easily seen if we put
in the left neighboring (or
) column and then compare the values in both columns. It is also interesting to note that at
, the critical exponent
does not depend on
, unlike all other exponents in this column. On the other hand, the critical exponents
and
do not depend on
N and are the same for any
.