Next Article in Journal
Probing Dark Photons Through Gravitational Decoupling of Mass-State Oscillations in Interstellar Media
Previous Article in Journal
Numerical Evaluation of a Soliton Pair with Long-Range Interaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Scaling Behavior and Phases of Nonlinear Sigma Model on Real Stiefel Manifolds Near Two Dimensions

by
Alexandre M. Gavrilik
* and
Andriy V. Nazarenko
Bogolyubov Institute for Theoretical Physics of NAS of Ukraine, 14b, Metrolohichna Street, 03143 Kyiv, Ukraine
*
Author to whom correspondence should be addressed.
Universe 2025, 11(4), 114; https://doi.org/10.3390/universe11040114
Submission received: 30 January 2025 / Revised: 27 March 2025 / Accepted: 28 March 2025 / Published: 31 March 2025
(This article belongs to the Section Field Theory)

Abstract

:
For a quasi-two-dimensional nonlinear sigma model on the real Stiefel manifolds with a generalized (anisotropic) metric, the equations of a two-charge renormalization group (RG) for the homothety and anisotropy of the metric as effective couplings are obtained in a one-loop approximation. Normal coordinates and the curvature tensor are exploited for the renormalization of the metric. The RG trajectories are investigated and the presence of a fixed point common to four critical lines or four phases (tetracritical point) in the general case, or its absence in the case of an Abelian structure group, is established. For the tetracritical point, the critical exponents are evaluated and compared with those known earlier for a simpler particular case.

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 Gr k ( R N ) are their well-known extensions [13], such as the real Stiefel manifolds V k ( R N ) . The latter are nothing but principal fiber bundles with a base being the Grassmannian Gr k ( R N ) and the structure group S O ( k ) . 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 d = 2 + ε , 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 d = 2 and d = 2 + ε . In particular, in contrast to the single-charge RG (scaling) behavior of Grassmannian sigma models, the class of V k ( R N ) sigma models is characterized by a two-charge behavior, and the question of the presence of asymptotic freedom in them at d = 2 becomes quite nontrivial [11,18].
The extension of the Gr k ( R N ) sigma model to the d = ( 2 + ε ) -dimensional nonlinear V k ( R N ) -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 V k ( R N ) 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 V k ( R N ) , the related nonlinear sigma models in d = 2 + ε 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 V k ( R N ) 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 k N k 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 k N k symmetry is an intrinsic feature of the Stiefel V k ( R N ) 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 d = 2 + ε Euclidean dimensions with the action functional A and the Lagrangian L :
A [ U ] = 1 2 T L d d x , L [ U ] = Tr U U + ( λ 1 ) U U U U ,
where we have introduced the temperature T, field U ( x ) valued in the Stiefel manifold V k ( R N ) = S O ( N ) / S O ( N k ) , gradient = ( / x 1 , . . . , / x d ) , and parameter (of the metric anisotropy) λ > 0 .

2. Some Characteristics of Stiefel Manifold

A real Stiefel manifold V k ( R N ) is the set of all k-frames formed by sets of orthonormal vectors in the space R N . Any element U V k ( R N ) , given by an N × k -matrix, satisfies the orthogonality condition U U = I k , where I k is the identity k × k -matrix and ⊤ denotes the transpose. On such manifolds, the group S O ( N ) acts transitively as the isometry group. The isotropy group corresponding to the “origin” O = { e 1 , e 2 , . . . , e k } consists of matrices of the form diag ( I k , H ) , where H S O ( N k ) . This allows us to identify the manifold V k ( R N ) with the quotient space S O ( N ) / S O ( N k ) with dim V k ( R N ) = N k k ( k + 1 ) / 2 . Thus, V k ( R N ) is defined as a submanifold of the space R N × k . In particular, V 1 ( R ) = { 1 , 1 } ; V 1 ( R N ) = S N 1 is an ( N 1 ) -dimensional sphere; V N ( R N ) = O ( N ) ; V N 1 ( R N ) = S O ( N ) ; and V 2 ( R 4 ) is the first nontrivial case.
Tangent space. Let TM be the tangent space for M = V k ( R N ) . We require that each tangent vector ξ T U M at a point U M belongs to the horizontal subspace, i.e., it has the following structure [21,22,23]:
ξ = U ω ξ + F ξ , ω ξ = 1 2 ( U ξ ξ U ) o ( k ) , F ξ = Π ξ ,
where the projector Π = I N U U , and ω ξ = ω ξ . Since ω ξ consists of k ( k 1 ) / 2 independent matrix components and F ξ contains ( N k ) k components, their total sum restores dim V k ( R N ) . Conversely, the symmetric component ( U ξ + ξ U ) / 2 belongs to the vertical subspace.
Tangent vectors can be defined in the respective Lie algebra so ( N ) , the basis of which is formed by N ( N 1 ) / 2 skew-symmetric N × N -matrices X A :
X A = ( X A ) b , c = δ b , a 1 δ c , a 2 δ b , a 2 δ c , a 1 ,
where the multi-index A = ( a 1 , a 2 ) consists of the two subsets such that 1 a 1 , a 2 N . The matrix X A has only two nonzero elements, 1 and 1 , at positions ( a 1 , a 2 ) and ( a 2 , a 1 ) , so that X A = X A and Tr ( X A X A ) = 2 . The basis elements of the algebra so ( N ) satisfy the commutation relations:
[ X A , X B ] = f A B C X C , f A B C = δ a 1 , b 1 δ a 2 c 1 δ b 2 c 2 + δ a 2 , b 2 δ a 1 c 1 δ b 1 c 2 δ a 2 , b 1 δ a 1 c 1 δ b 2 c 2 δ a 1 , b 2 δ a 2 c 1 δ b 1 c 2 .
In this basis, the orthonormal metric g ( X , Y ) = B ( X , Y ) / [ 2 ( N 2 ) ] based on the bi-invariant Killing form B ( X , Y ) , when B ( X A , X B ) = f A C D f B D C , coincides with the Frobenius metric, which is natural in the description of the σ -model (1):
X A , X B F = 1 2 Tr ( X A X B ) = δ A B .
Then, the norm of the vector ξ TM is written as ξ F 2 = ( 1 / 2 ) Tr ( ξ ξ ) .
To write the ω - and F-components of tangent vectors in the basis introduced, we define X P and X Q with P = ( p , q ) and Q = ( j , p ) , where 1 p , q k and k + 1 j N . Although F R N × k in (2), it has ( N k ) k independent components and is usually included in meaningful expressions as a combination F F .
Metric. The covariant metric g ( U ; λ ) (metric operator) of model (1) and its action on the tangent vector ξ T U M are defined as
g ξ = ξ + ( λ 1 ) U U ξ , g 1 ξ = ξ + ( λ 1 1 ) U U ξ , det g = λ k .
Note that the case λ = 1 / 2 is called canonical in the literature [22,23].
Then the inner (scalar) product of two tangent vectors ξ , η T U M is given as
ξ , η g = Tr ξ η + ( λ 1 ) ξ U U η .
Since ξ , ξ g = Tr ( λ ω ξ ω ξ + F ξ F ξ ) , according to (2), the Lagrangian in (1) becomes
L = Tr ( λ ω u ω u + F u F u ) .
Here we use the following decomposition of the tangent vector U :
U = U ω u + F u , ω u = U U , F u = Π U ,
because U U + U U = 0 due to the constraint U U = I k .
Expression (8) indicates the presence of two ω - and F-subsystems. The condition U U = I k can be easily satisfied by reducing the system to an O ( k ) -model with ω u 0 and F u = 0 . If ω u is expressed through F u 0 , then a model on the Grassmannian Gr k ( R N ) = S O ( N ) [ S O ( N k ) × S O ( k ) ] is obtained. However, our focus here is on incorporating both subsystems, utilizing the full number of degrees of freedom, dim V k ( R N ) .
Linear connection and normal coordinates. Consider the auxiliary problem of one-dimensional evolution, which is generated by the action functional:
S = d t Tr U ˙ U ˙ ( λ 1 ) ( U U ˙ ) 2 + Tr H ( U U I k ) ,
where H = H R k × k is a Lagrange multiplier; U ˙ = d U / d t .
By varying U ( t ) and H and using certain identities, one can obtain the multiplier H = U ˙ U ˙ + 2 ( λ 1 ) ( U U ˙ ) 2 and the equation of the geodesic [21]:
U ¨ + Γ ( U ˙ , U ˙ ) = 0 .
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 U M :
Γ ( ξ , η ) = 1 2 U ξ η + η ξ + ( 1 λ ) Π ξ η + η ξ U .
Now, we present the solution of Equation (11) as a series in t in terms of the initial data U ( 0 ) = U 0 and U ˙ ( 0 ) = V , which are related by U 0 V + V U 0 = 0 . One has
U ( t ) = U 0 + t V t 2 2 ! Γ 0 ( V , V ) t 3 3 ! Γ 0 ( V , V , V ) ,
where Γ 0 ( V , V ) = Γ ( V , V ) U = U 0 , and Γ 0 ( V , V , V ) = D V Γ ( V , V ) 2 Γ ( V , Γ ( V , V ) ) U = U 0 involves the derivative D V at the point U M in the direction V TM , which acts, in particular, as
D ϕ Γ ( ξ , η ) = 1 2 ϕ ξ η + η ξ ( 1 λ ) ϕ U + U ϕ ξ η + η ξ U + ( 1 λ ) Π ξ η + η ξ ϕ .
Note that the expansions (2) and (9) can be applied after calculating such a derivative.
It is easy to verify in the approximation t 2 that I k = U U = U 0 U 0 . Thus, setting t = 1 , the two points U , U 0 M are related by (13). Restoring the dependence of U and V on the coordinates x, formula (13) generalizes the “field shift” U = U 0 + V in Euclidean space when U , U 0 , V R dim M . Thus, fixing U 0 , the variables V are called normal coordinates, which further describe quantum fluctuations in the σ -model.
Curvature. We define the curvature tensor R ( ξ , η ) ϕ in ( 1 , 3 ) -form for ξ , η , ϕ TM as (see [22])
R ( ξ , η ) ϕ = D η Γ ( ξ , ϕ ) D ξ Γ ( η , ϕ ) + Γ ( η , Γ ( ξ , ϕ ) ) Γ ( ξ , Γ ( η , ϕ ) ) .
Let us calculate the components of R ( ξ , η ) ξ = U ( U R ( ξ , η ) ξ ) + Π R ( ξ , η ) ξ :
U R ( ξ , η ) ξ = 1 4 ( ω η ω ξ 2 2 ω ξ ω η ω ξ + ω ξ 2 ω η ) + λ 2 ( ω η F ξ F ξ + F ξ F ξ ω η ) + 3 4 λ 4 ( ω ξ F ξ F η + F η F ξ ω ξ ) 3 2 λ 4 ( ω ξ F η F ξ + F ξ F η ω ξ ) ,
Π R ( ξ , η ) ξ = λ 3 4 λ 2 F ξ ω ξ ω η + λ 3 2 λ 2 F ξ ω η ω ξ λ 2 F η ω ξ 2 + F η F ξ F ξ + 2 3 λ 2 F ξ F ξ F η 4 3 λ 2 F ξ F η F ξ .
The bi-quadratic form K ˜ ( ξ , η ) = R ( ξ , η ) ξ , η g , which characterizes the sectional curvature [24], can be expressed, after substituting expressions (16) and (17), as follows:
K ˜ ( ξ , η ) = λ 2 [ ω ξ , ω η ] F 2 + F η F ξ F ξ F η F 2 + 2 3 λ 2 F η F ξ F ξ F η F 2 + 2 λ 2 F η ω ξ F 2 + F ξ ω η F 2 + 2 λ ( 3 4 λ ) F ξ ω ξ , F η ω η F 2 λ ( 3 2 λ ) F ξ ω η , F η ω ξ F .
A similar formula is derived in the work ([25], p. 405) in terms of rectangular ( N k ) × k -matrices, say M, instead of F R N × k . However, it does not matter, because M and F are entered as combinations M M and F F , which both belong to R k × k .
Using (16) and (17) and contracting the ( 1 , 3 ) -curvature (see Appendix A.1), the diagonal components of the Ricci ( 0 , 2 ) -tensor are obtained [22]:
Ric ( ξ , ξ ) = λ 2 ( N k ) + k 2 4 Tr ( ω ξ ω ξ ) + N 2 λ ( k 1 ) Tr ( F ξ F ξ ) .
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 K ˜ and Ric when renormalizing the σ -model.

3. Background Field Formalism

We assume that the Lagrangian L [ U 0 ] = Tr ( U 0 g 0 U 0 ) is determined by the bare metric g = g ( U ; λ ) , which differs from g 0 = g ( U 0 ; λ ) by the covariant counterterms h :
g = μ ε g 0 + h ,
where ε = d 2 is the renormalization parameter; μ 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 Φ ( x , s ) and quantities similar to those defined earlier:
Φ ( x , 0 ) = U 0 ( x ) , Φ ( x , s ) s s = 0 = V , Φ ( x , 1 ) = U ( x ) .
Let us expand the Lagrangian L [ U ] in a series with respect to the parameter s:
L [ U ] L [ Φ ( x , 1 ) ] = n = 0 1 n ! n L [ Φ ( x , s ) ] s n s = 0 ,
using the covariant derivative ^ s along the geodesic Φ ( x , s ) , as well as ^ :
^ s ξ ξ s + Γ Φ ( x , s ) s , ξ , ^ ξ ξ + Γ Φ ( x , s ) , ξ ,
where the Christoffel function Γ ( ξ , η ) from (12) is taken at U = Φ ( x , s ) .
Applying the decomposition of tangent vectors into ω - and F-components, we obtain
L [ Φ ( x , s ) ] s = 2 Tr Φ g ^ Φ s ,
1 2 2 L [ Φ ( x , s ) ] s 2 = Tr ^ Φ s g ^ Φ s + Tr Φ g ^ s , ^ Φ s ,
taking into account that
^ s Φ ( x , s ) s = 0 ,
^ s Φ ( x , s ) = ^ Φ ( x , s ) s Φ ( x , s ) s + Γ Φ ( x , s ) s , Φ ( x , s ) ,
^ s , ^ ξ = R Φ ( x , s ) , Φ ( x , s ) s ξ ,
where the curvature R ( ξ , η ) ϕ is given by (15).
In the one-loop approximation, we restrict ourselves to the Lagrangian quadratic in V:
L [ U ] = L [ U 0 ] + 2 U 0 , ^ V g 0 + ^ V , ^ V g 0 + U 0 , R ( U 0 , V ) V g 0 ,
where ^ V = V + Γ 0 ( V , U 0 ) , and U 0 , R ( U 0 , V ) V g 0 = K ˜ ( U 0 , V ) by definition. If U 0 ( x ) is a solution of the equation of motion, the term U 0 , ^ V g 0 does not contribute to the action integral. Since expression (29) is general for σ -models within this approximation, we proceed to the next step, which is standard for such models.
We replace V by V to transform Tr ( V g 0 V ) into Tr ( V V ) . To do this, we set V = e ˜ V , where e ˜ = g ( U 0 ; λ 1 / 2 ) . Then e ˜ 2 = g 0 1 and e ˜ g 0 e ˜ = I N , and the relation between the components of the tangent vectors V and V is given by
V = U 0 ω V + F V , V = U 0 ω V + F V , ω V = λ 1 / 2 ω V .
After introducing ω V instead of ω V , the quadratic part of the Lagrangian with respect to V takes on the form
L eff = ^ V , ^ V g 0 K ˜ ( U 0 , e ˜ V ) ,
^ V , ^ V g 0 = Tr ω V ω V + F V F V + L int ,
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 F V Q components of F V R N × k as independent, since the F V P components merely represent the interaction of F V Q with U 0 (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 i is replaced by the real temperature T. We consider U 0 ( x ) as the background field, and ω V and F V as fluctuating quantities. Simultaneously, the components of the metric g , including the positive parameters T and λ (see (1)), play the role of coupling constants. In physical applications of σ -models for describing Anderson localization, the temperature T is often associated with the conductivity coefficient, which undergoes renormalization. On the other hand, the parameter λ (anisotropy) appears in the Lagrangian (8) as a running (scale-dependent) factor.
Quantizing, one should introduce a path integral over the rapidly varying fields ω V and F V (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 U 0 ( x ) 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:
ω V P 1 ( x ) ω V P 2 ( y ) = π P 1 P 2 Δ ( x y ) , F V Q 1 ( x ) F V Q 2 ( y ) = δ Q 1 Q 2 Δ ( x y ) ,
where the brackets . . . denote the quantum average in the vacuum state; and the components of the matrix field F V with multi-indices Q n = ( j n , p n ) , where 1 p n k and k + 1 j n N are chosen as independent. In addition, the symbol δ Q 1 Q 2 = δ j 1 , j 2 δ p 1 , p 2 , and the symbol
π P 1 P 2 = 1 2 δ p 1 , p 2 δ q 1 , q 2 δ p 1 , q 2 δ q 1 , p 2
for multi-indices P n = ( p n , q n ) , 1 p n , q n k , and their contraction gives the number of degrees of freedom. The spatially dependent component is defined by the expression
Δ ( x y ) = T e i k · ( x y ) k 2 d d k ( 2 π ) d ,
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 μ and denoting the area of the unit sphere in d dimensions as Ω d 1 = 2 π d / 2 / Γ ( d / 2 ) , when ε = d 2 0 , we have
Δ ( 0 ) = T Ω d 1 ( 2 π ) d μ k d 3 d k = T 2 π ε T 2 π ln μ + O ( ε ) .
At once, ultraviolet (UV) divergences are eliminated in a standard way for σ -models [28].
Thus, the metric renormalization in the one-loop approximation is induced by the term [26,27,28]
K ˜ ( U 0 , e ˜ V ) d d x = Ric ( U 0 , U 0 ) Δ ( 0 ) d d x .
The idea of proving formula (37) is based on the equality of the quantum average of K ˜ over the fields ω V and F V and the average over the basis matrices, when U 0 = U 0 ω u + F u (see Appendix A.1). That is, due to the homogeneity of the Stiefel manifold, the calculations can be shifted to the “origin” O , allowing the use of the basis (3). Then, we perform the substitution ω V P X P , F V Q Π X Q , while the fields themselves are absorbed by Δ . We also express U 0 in the basis (3) and sum over Q and the independent indices of P. When averaging K ˜ ( U 0 , e ˜ V ) , only the terms quadratic in ω V and F V survive, and we obtain
Ric ( U 0 , U 0 ) = λ 2 ( N k ) + k 2 4 Tr ( ω u ω u ) + N 2 λ ( k 1 ) Tr ( F u F u ) .
The divergences of (37) at ε 0 are eliminated by the scaling factors Z g ’s for the metric components:
Z g ω λ T δ P 1 P 2 , Z g F 1 T δ Q 1 Q 2 .
According to (29), (36) and (38), we obtain
Z g ω = 1 τ ε λ 2 ( N k ) + k 2 4 , Z g F = 1 t ε [ N 2 λ ( k 1 ) ] ,
where we have defined the parameters t = T / ( 2 π ) and τ = T / ( 2 π λ ) .
Beta functions. Renormalizing the model metric as z = ln μ changes, we demand
d d z ln μ ε Z g F t 1 = 0 , d d z ln μ ε Z g ω τ 1 = 0 .
These two conditions can be satisfied by assuming that both t and λ (or τ ) depend on z. Otherwise, to renormalize only the temperature t, we would have to require that the Stiefel manifold be Einsteinian when we equate
λ 2 ( N k ) + k 2 4 = λ [ N 2 λ ( k 1 ) ] .
This is the same as P N , k ( λ ) = 0 for P N , k ( λ ) λ 2 ( N 1 ) λ ( N 2 ) + ( k 2 ) / 4 .
Defining the beta functions β t = d t / d z and β λ = d λ / d z that need to be found, we arrive at the set of exact equations:
β t 1 t ln Z g F t β λ ln Z g F λ = ε ,
β t 1 t ln Z g ω t β λ 1 λ + ln Z g F λ = ε .
Determining Z g ’s up to order t, they reduce in the one-loop approximation to
β t t = ε Z g F , β t t β λ λ = ε Z g ω ,
where β t / t β λ / λ = β τ / τ for τ = t / λ .
Note that taking into account all the terms in these equations using approximate Z g ’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
β t = ε t [ N 2 λ ( k 1 ) ] t 2 , β τ = ε τ λ 2 ( N k ) + k 2 4 τ 2 .
Replacing β τ with tantamount β λ = t P N , k ( λ ) , we rewrite it in an equivalent form:
β λ = t ( N 1 ) ( λ λ + ) ( λ λ ) ,
λ ± = N 2 2 ( N 1 ) 1 ± 1 ( k 2 ) ( N 1 ) ( N 2 ) 2 .
The region of admissible values of the model parameters in the plane ( λ ; t ) is the quadrant with λ > 0 and t > 0 . This region is divided into three subregions by two separatrices ( λ ; t ) and ( λ + ; t ) for arbitrary t > 0 . The fixed points are found from the conditions β t = 0 and β λ = 0 . Defining
t ± = ε N 2 λ ± ( k 1 ) ,
we have the Gaussian point ( 0 ; 0 ) , the node ( λ ; t ) , and the saddle point ( λ + ; t + ) (see Figure 1a).
The trajectories of the renormalization group (RG) are given by the equations:
d t d z = ε t [ N 2 λ ( k 1 ) ] t 2 , t ( 0 ) = t 0 ,
d λ d z = t λ 2 ( N 1 ) λ ( N 2 ) + k 2 4 , λ ( 0 ) = λ 0 .
Then, the separatrices in Figure 1, passing through the points ( 0 ; ε ( N 2 ) ) , ( λ ; t ) and ( λ + ; t + ) , are obtained numerically by integrating (50) and (51). It can be seen that in the case of 2 < k < N and d > 2 , the point ( λ ; t ) , at which the four phases meet, seems tetracritical (red circle in Figure 1a). Note that at k = 2 , such tetracriticality disappears, and the point ( λ ; t ) = ( 0 ; ε ( N 2 ) ) becomes bicritical (green circle in Figure 1c). Finally, at d = 2 and fixed λ , phase transitions by varying temperature t are not observed (see Figure 1b,d), although the sink–saddle pair still exists.
By introducing the quantity g = 1 / t , associated with the conductivity in some models, and the corresponding beta function β g d g / d z = g 2 β t ( t = 1 / g ) , 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 ( λ ; t ) . By fixing λ = λ , 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 ν (and η ) in the one-loop approximation for 2 k < N and ε > 0 . Under these conditions, there exists a positive critical temperature t c = t , the determination of which provides [15]:
1 ν = β t ( t c ) = ε + O ( ε 2 ) .
This agrees with the known result for the Grassmannian S O ( N ) [ S O ( N k ) × S O ( k ) ] 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 η similarly to [4]:
η = ζ ( t c ) ε = ε N k + ( k 1 ) / ( 2 λ ) N 2 λ ( k 1 ) 1 + O ( ε 2 ) .
We observe that η 0 + O ( ε 2 ) for large N, and furthermore, when k = 1 , this expression yields the previously established value [4].
Other critical exponents can be found using the scaling and hyperscaling relations:
α = 2 ν d , β = ν d 2 + η 2 , γ = ( 2 η ) ν , δ = d + 2 η d 2 + η .
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 N = 12 in the left neighboring (or k = 2 ) column and then compare the values in both columns. It is also interesting to note that at k = 1 , the critical exponent β does not depend on ε , unlike all other exponents in this column. On the other hand, the critical exponents α and ν do not depend on N and are the same for any k > 0 .

5. Conclusions

In this work, starting from the field Lagrangian for the ( 2 + ε ) -dimensional nonlinear sigma model on the real Stiefel manifold, we applied the background-field method and normal coordinates for the quantum treatment and obtained a renormalization group description of the renormalized model in terms of two effective or running (depending on the scaling parameter) charges. This result fully confirms the corresponding formulas for beta functions and the properties of the renormalization group behavior of the studied model, previously obtained [11,18] by applying a pure geometric approach to the general case of the matrix beta function of a nonlinear model on an arbitrary Riemannian manifold. The special tetracritical behavior—the presence of a tetracritical point common to four phases—requires further detailed study (of the symmetry properties, etc.) of each of these phases and the corresponding critical exponents. No less interesting is the subsequent application of the considered model to the description of the properties of real tetracritical physical systems, in particular to the problem of the possible coexistence and mutual influence of the phenomenon of (de)localization in a quasi-two-dimensional electronic system and the phenomenon of superconductivity, as well as other systems with two independent order parameters (see, e.g., Refs. [29,30] for two such examples of symmetry-based systems that manifest tetracritical behavior). More generally speaking, the obtained results may also be useful from the viewpoint of studying new classes of universality and their systematics in the physics of (multi)critical phenomena.

Author Contributions

The contributions of both authors to the article preparation were equal. Conceptualization and draft preparation, A.M.G. and A.V.N.; numerical analysis, A.V.N. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Academy of Sciences of Ukraine by its priority project No. 0122U000888 and was supported by the Simons Foundation grant 1290587.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to authors.

Acknowledgments

The authors acknowledge anonymous reviewers for valuable comments and remarks which helped to improve the manuscript.

Conflicts of Interest

The authors declare no conflicts of interest.

Appendix A. Calculating the Curvature Tensors

Appendix A.1. Ricci Curvature Tensor

Let us calculate the diagonal components of the Ricci ( 0 , 2 ) -tensor in the basis (3), using the contraction
Ric ( ξ , ξ ) = η Tr η R ( ξ , η ) ξ = ω η , F η Tr ω η U R ( ξ , η ) ξ + F η Π R ( ξ , η ) ξ .
Due to the homogeneity of Stiefel manifolds, we perform calculations at the “origin” O , that is, using the matrix U O and the corresponding projector Π O :
U O = I k 0 ( N k ) × k , Π O = I N U O U O .
Define the components of the tangent vector η in the basis (3) and at the “origin” as
ω η ( P ) = 2 1 / 2 X P , F η ( Q ) = Π O X Q ,
where multi-indices P = ( p , q ) and Q = ( j , r ) consist of p , q , r = 1 , k ¯ and j = k + 1 , N ¯ . This means that (1) we embed the components ω η and F η in R N × N , setting all remaining components to zero; (2) we normalize η such that Tr [ ω η ( P ) ω η ( P ) ] = 1 and Tr [ F ξ ( Q ) F ξ ( Q ) ] = 1 .
Arbitrarily fixing the multi-indices P 0 and Q 0 of the components ω ξ and F ξ , we write
ω ξ ( P 0 ) = 2 1 / 2 ω X P 0 , F ξ ( Q 0 ) = F Π O X Q 0 ,
where ω and F are the values such that Tr [ ω ξ ( P 0 ) ω ξ ( P 0 ) + F ξ ( Q 0 ) F ξ ( Q 0 ) ] = ω 2 + F 2 .
Substituting the defined vector components into (A1) and numerically summing over P = ( p , q ) , p < q , and Q = ( j , r ) in the corresponding terms, one obtains
r ( P 0 , Q 0 ) = λ 2 ( N k ) + k 2 4 ω 2 + N 2 λ ( k 1 ) F 2 .
Next, summing all terms r ( P 0 , Q 0 ) with admissible P 0 and Q 0 , we obtain an expression for the diagonal components of the Ricci tensor:
Ric ( ξ , ξ ) = λ 2 ( N k ) + k 2 4 Tr ( ω ξ ω ξ ) + N 2 λ ( k 1 ) Tr ( F ξ F ξ ) .
Due to the homogeneity of Stiefel manifolds, the curvature coefficients are constant at all points of space.
Decompose (A6) into orthogonal components for admissible P and Q to obtain
Ric ( X P , X P ) = 2 λ 2 ( N k ) + k 2 2 , Ric ( X Q , X Q ) = N 2 λ ( k 1 ) ,
where we replace Tr ( ω ξ ω ξ ) Tr ( X P X P ) = 2 and Tr ( F ξ F ξ ) Tr ( X Q Π O X Q ) = 1 .
We would also like to note another derivation of the tensors for λ = 1 / 2 in Appendix A.2.

Appendix A.2. Curvature in the Canonical Case λ=1/2

Focusing on the historically earlier and instructive case λ = 1 / 2 for X , Y , Z m , when so ( N ) m h at h = so ( N k ) , let us analyze the curvature R at the “origin” O (see Proposition 3.4 in [24], II, p. 202):
( R ( X , Y ) Z ) O = 1 4 [ X , [ Y , Z ] m ] m 1 4 [ Y , [ X , Z ] m ] m 1 2 [ [ X , Y ] m , Z ] m [ [ X , Y ] h , Z ] ,
where the commutator markings denote projections onto m and h , and there is an obvious decomposition [ X , Y ] = [ X , Y ] m + [ X , Y ] h .
Due to Proposition 3.4 from ([24], II, p. 202), define the bi-quadratic form:
K ˜ ( X , Y ) g ( R ( X , Y ) Y , X ) = 1 4 g ( [ X , Y ] m , [ X , Y ] m ) + g ( [ X , Y ] h , [ X , Y ] h ) ,
which has the properties: K ˜ ( Y , X ) = K ˜ ( X , Y ) , K ˜ ( X , X ) = 0 , K ˜ ( X + Y , Y ) = K ˜ ( X , Y ) , and K ˜ ( α X , Y ) = K ˜ ( X , α Y ) = α 2 K ˜ ( X , Y ) , where α is a scalar.
Taking into account the commutation relation (4) and the fact that given f A B C is not anti-symmetrized over indices ( c 1 , c 2 ) = C , we can write that [ X A , X B ] m = f A B ( p , q ) X ( p , q ) + f A B ( p , j ) X ( p , j ) + f A B ( j , p ) X ( j , p ) m and [ X A , X B ] h = f A B ( j 1 , j 2 ) X ( j 1 , j 2 ) h . Then, we have for matrices X A , X B m that
K ˜ ( X A , X B ) = 1 4 p , q = 1 k ( 1 δ p , q ) f A B ( p , q ) 2 + 1 4 p = 1 k j = k + 1 N f A B ( p , j ) 2 + f A B ( j , p ) 2 + j 1 , j 2 = k + 1 N ( 1 δ j 1 , j 2 ) f A B ( j 1 , j 2 ) 2 .
Using K ˜ ( X , Y ) , the sectional curvature of V k ( R N ) is determined by
K ( X , Y ) = K ˜ ( X , Y ) X 2 Y 2 g ( X , Y ) 2 .
where the divisor equals unit in the case of orthonormal metric g ( X A , X B ) = δ A B , see (5).
The Riemann curvature tensor for V k ( R N ) is given by ([24], I, p. 201):
R ( X 1 , X 2 , X 3 , X 4 ) = g ( R ( X 3 , X 4 ) X 2 , X 1 ) .
For instance, it can be presented for distinct X 1 , , X 4 as
R ( X 1 , X 2 , X 3 , X 4 ) = 1 6 K ˜ ( X 1 + X 3 , X 2 + X 4 ) K ˜ ( X 2 + X 3 , X 1 + X 4 ) + K ˜ ( X 3 , X 1 ) + K ˜ ( X 4 , X 2 ) K ˜ ( X 3 , X 2 ) K ˜ ( X 4 , X 1 ) .
Therefore, contracting the Riemann tensor, the Ricci curvature tensor ([24], I, p. 249) is determined by
R ( X , Y ) = Z m g ( R ( Z , X ) Y , Z ) ,
Its diagonal components for X m are written as
Ric ( X , X ) = Z m K ˜ ( Z , X )
= p < q = 1 k K ˜ ( X ( p , q ) , X ) + p = 1 k j = k + 1 N K ˜ ( X ( j , p ) , X ) ,
Using (A10), one can numerically verify for arbitrary indices P = ( p , q ) and Q = ( j , p ) that
Ric ( X P , X P ) = N 2 2 , Ric ( X Q , X Q ) = N 2 k 1 2 .
These numerical coefficients coincide with (A7) at λ = 1 / 2 .

Appendix B. Certain Aspects of Quantization

Appendix B.1. Path Integral

Using the action integral (1), the partition function Z is given by the path integral with an invariant measure defined up to a constant factor involving det g ( U ( x ) ; λ ) = λ k / 2 :
Z = exp ( A [ U ] ) δ ( U U I k ) D U ,
where the formal δ -function fixes k ( k + 1 ) / 2 constraints at each point x.
Note that, unlike quantum theory, where the imaginary Planck constant i is used, we employ low-temperature T as a measure of positive definite action.
Introducing the on-shell background field U 0 ( x ) , U 0 U 0 = I k , which justifies the equation 2 U 0 + Γ 0 ( U 0 , U 0 ) = 0 , we set U = U 0 + υ , where the shift field υ is expressed in terms of the normal coordinates V T M according to (13):
υ = V 1 2 ! Γ 0 ( V , V ) 1 3 ! Γ 0 ( V , V , V ) .
The orthogonality condition U U = I k in terms of new variables is satisfied due to the vanishing of s V U 0 V + V U 0 . Then, decomposing V = U 0 ( ω V + s V ) + F V , where dim ω V = k ( k 1 ) / 2 and dim F V = ( N k ) k , the k ( k + 1 ) / 2 degrees of freedom of s V are eliminated by the δ -functions from (A18). Thus, we are left with V = U 0 ω V + F V = λ 1 / 2 U 0 ω V + F V , as it must be for a tangent vector.
However, we still need to parameterize F V R N × k , U 0 F V = 0 , by independent variables. By choosing for this F V Q R ( N k ) × k , where the multi-index Q = ( j , p ) with 1 p , q k and k + 1 j N , we can express the dependent components F V P with P = ( p , q ) as
F V ( p , q ) = j = k + 1 N E 0 ( p , j ) F V ( j , q ) , E 0 E 0 ( p , j ) = U 0 ( p , q ) 1 U 0 ( j , q ) ,
where one would expect U 0 ( p , q ) = ± ( I k U ˜ 0 U ˜ 0 ) 1 / 2 with U ˜ 0 = U 0 ( j , p ) to obtain U 0 U 0 = I k . Obviously, E 0 = U ˜ 0 = 0 , when we take U 0 = U O at the “origin”.
Let us emphasize that below by F V we mean independent components F V Q , for example, in the integration measure, unless otherwise specified.
Under these conditions, we come to the integral over fluctuations:
Ω [ U 0 ] = exp ( A [ U 0 ] A [ U 0 + υ ] ) Det δ υ ( x ) δ ω V P ( y ) , δ υ ( x ) δ F V Q ( y ) D ω V D F V .
Further, we are using a perturbative approach based on (vacuum) averaging:
( ) = Z 0 1 ( ) exp ( S 0 [ ω V , F V ] ) D ω V D F V ,
Z 0 = exp ( S 0 [ ω V , F V ] ) D ω V D F V ,
S 0 [ ω V , F V ] 1 2 T Tr ω V ω V + F V F V d d x .
It leads to Ω [ U 0 ] = Z 0 exp ( A int [ U 0 , ω V , F V ] ) with
A int [ U 0 , ω V , F V ] A [ U 0 + υ ] A [ U 0 ] S 0 [ ω V , F V ] ln D ,
where D denotes the Jacobi determinant from (A21).
According to the Dyson–Wick algorithm, let us introduce the generating functional with a skew-symmetric source J:
Z 0 [ J ] Z 0 1 exp S 0 + Tr J ω ω V + J F F V d d x D ω V D F V = exp T 2 Tr J ω 2 J ω + J F 2 J F d d x .
The Green’s function (35) results from
2 δ d ( x ) = e i k · x k 2 d d k ( 2 π ) d .
Thus, by expanding the interaction functional, the multi-point Green’s function is
ω V P 1 ( x 1 ) ω V P n ( x n ) F V Q 1 ( y 1 ) F V Q m ( y m ) = δ n + m Z 0 [ J ] δ J F Q m ( y m ) δ J F Q 1 ( y 1 ) δ J ω P n ( x n ) δ J ω P 1 ( x 1 ) J ω = 0 , J F = 0 .
Furthermore, the connected Green’s functions are generated by the functional T ln Ω [ U 0 ] .

Appendix B.2. Renormalizing the Field

Renormalizing, due to introducing the field scaling factor Z U , let us require finiteness of the anomalous quantum mean Z U 1 / 2 U of the field U V k ( R N ) at the point x R d . Given the background field U 0 V k ( R N ) , and using the normal coordinates V (see, for instance, (13)), in the one-loop approximation we can write that
U = U 0 + V 1 2 ! Γ 0 ( V , V ) = U 0 I k 1 2 λ 1 ω V ω V + F V F V ,
where U 0 = U 0 , and the Christoffel function, averaged as Γ 0 ( V , V ) = U 0 V V , is expressed in terms of the ω - and F-components of the field V with V = 0 .
According to the rules (33) and (36), we extract the divergent part
U 1 + t 2 ε k 1 2 λ + N k U 0 ,
that immediately leads to
Z U = 1 t ε k 1 2 λ + N k .
In particular, the substitution k = 1 reproduces the result for the vector model [4].
Then, the scaling properties in our approximation can be given by the “beta” function:
ζ = d ln Z U d z = k 1 2 λ + N k t .

References

  1. Gell-Mann, M.; Levy, M. The axial vector current in beta decay. Nuovo Cim. 1960, 16, 705. [Google Scholar] [CrossRef]
  2. Polyakov, A.M. Interaction of goldstone particles in two dimensions. Applications to ferromagnets and massive Yang-Mills fields. Phys. Lett. B 1975, 59, 79. [Google Scholar] [CrossRef]
  3. Volkov, D.V.; Zheltukhin, A.A.; Bliokh, Y.P. Phenomenological lagrangian for spin waves. Fiz. Tverd. Tela 1971, 13, 1668. [Google Scholar]
  4. Brezin, E.; Zinn-Justin, J. Renormalization of the nonlinear σ model in 2+ε dimensions—Application to the Heisenberg ferromagnets. Phys. Rev. Let. 1976, 36, 691. [Google Scholar] [CrossRef]
  5. Houghton, A.; Jevicki, A.; Kenway, R.D.; Pruisken, A.M.M. Noncompact σ models and the existence of a mobility edge in disordered electronic systems near two dimensions. Phys. Rev. Lett. 1980, 45, 394. [Google Scholar] [CrossRef]
  6. Coleman, S.; Wess, J.; Zumino, B. Structure of phenomenological Lagrangians. I. Phys. Rev. 1969, 177, 2230. [Google Scholar] [CrossRef]
  7. Callan, C.G.; Coleman, S.; Wess, J.; Zumino, B. Structure of phenomenological Lagrangians. II. Phys. Rev. 1969, 177, 2247. [Google Scholar] [CrossRef]
  8. Volkov, D.V. Phenomenological Lagrangians. Sov. J. Part. Nucl. 1973, 4, 1. [Google Scholar]
  9. Hikami, S.; Larkin, A.I.; Nagaoka, Y. Spin-orbit interaction and magnetoresistance in the two dimensional random system. Prog. Theor. Phys. 1980, 63, 707. [Google Scholar] [CrossRef]
  10. Gavrilik, A.M. Gell-Mann-Low function of the chiral model on the homogeneous Berger space. Theor. Math. Phys. 1986, 65, 1075. [Google Scholar] [CrossRef]
  11. Gavrilik, A.M. Self-interaction anisotropy of Stiefel systems of Goldstonions. Mod. Phys. Lett. A 1989, 4, 1783. [Google Scholar]
  12. Gavrilik, A.M. Coset-space string compactification leading to 14 subcritical dimensions. Acta Phys. Hung. A 2000, 11, 35. [Google Scholar]
  13. Husemoller, D. Fibre Bundles, 3rd ed.; Springer: New York, NY, USA, 1993. [Google Scholar]
  14. McKane, A.; Stone, M. Non-linear σ models: A perturbative approach to symmetry restoration. Nucl. Phys. B 1980, 163, 169. [Google Scholar]
  15. Hikami, S. Renormalization group functions of orthogonal and symplectic non-linear σ model. Prog. Theor. Phys. 1980, 64, 1466. [Google Scholar]
  16. Abrahams, E.; Anderson, P.W.; Licciardello, D.C.; Ramakishnan, T.V. Scaling theory of localization: Absence of quantum diffusion in two dimensions. Phys. Rev. Lett. 1979, 42, 673. [Google Scholar]
  17. Zakharov, V.E.; Mikhailov, A.M. Relativistically invariant two-dimensional models in field theory integrable by the inverse problem technique. Zh. Eksp. Teor. Fiz. 1978, 74, 1953. [Google Scholar]
  18. Gavrilik, A.M. Geometry in nonlinear quantum-field models on Stiefel manifolds and bifurcations of autonomous systems. Ukr. J. Math. 1991, 43, 1527. (In Russian) [Google Scholar]
  19. Hirsch, J.E. Electron-hole asymmetry and superconductivity. Phys. Rev. B 2003, 68, 012510. [Google Scholar]
  20. Ghosh, S.S.; Quan, Y.-D.; Pickett, W.E. Strong particle-hole asymmetry in a 200 Kelvin superconductor. Phys. Rev. B 2019, 100, 094521. [Google Scholar]
  21. Hüper, K.; Markina, I.; Leite, F.S. A Lagrangian approach to extremal curves on Stiefel manifolds. J. Geom. Mech. 2021, 13, 55. [Google Scholar]
  22. Nguyen, D. Curvatures of Stiefel manifolds with deformation metrics. J. Lie Theory 2022, 32, 563. [Google Scholar]
  23. Edelman, A.; Arias, T.A.; Smith, S.T. The geometry of algorithms with orthogonality constraints. SIAM J. Matrix Anal. Appl. 1998, 20, 303. [Google Scholar]
  24. Kobayashi, S.; Nomizu, K. Foundations of Differential Geometry; John Wiley & Sons: New York, NY, USA, 1969; Volume I–II. [Google Scholar]
  25. Jensen, G.R. Imbeddings of Stiefel manifolds into Grassmannians. Duke Math. J. 1975, 42, 397. [Google Scholar]
  26. Alvarez-Gaumé, L.; Freedman, D.; Mukhi, S. The background field method and the ultraviolet structure of the supersymmetric nonlinear σ-model. Ann. Phys. 1981, 134, 85. [Google Scholar]
  27. Howe, P.S.; Papadopoulos, G.; Stelle, K.S. The background field method and the non-linear σ-model. Nucl. Phys. B 1988, 296, 26. [Google Scholar]
  28. Braaten, E.; Curtright, T.L.; Zachos, C.K. Torsion and geometrostasis in nonlinear sigma models. Nucl. Phys. B 1985, 260, 630. [Google Scholar]
  29. Birman, J.L.; Wege, M. Theory of coexistence of superconductivity and ferroelectricity: A dynamical symmetry. Phys. Rev. B 2001, 64, 174503. [Google Scholar]
  30. Sannino, F.; Tuominen, K. Tetracritical behavior in strongly interacting theories. Phys. Rev. D 2004, 70, 034019. [Google Scholar]
Figure 1. Phase regions of the model. (Top row) N = 12 , k = 9 , when d = 2.2 in (a) and d = 2 in (b). (Bottom row) N = 5 , k = 2 , when d = 2.1 in (c) and d = 2 in (d). Gray arrows denote velocity field. Colored circles are sinks, while the white ones are saddle points. Black lines are phase separatrices, while blue curves are RG trajectories.
Figure 1. Phase regions of the model. (Top row) N = 12 , k = 9 , when d = 2.2 in (a) and d = 2 in (b). (Bottom row) N = 5 , k = 2 , when d = 2.1 in (c) and d = 2 in (d). Gray arrows denote velocity field. Colored circles are sinks, while the white ones are saddle points. Black lines are phase separatrices, while blue curves are RG trajectories.
Universe 11 00114 g001
Figure 2. Phase regions of the model in the ( g , λ ) -plane for N = 12 , k = 9 , d = 2.2 . The red and white circles, as before, indicate the stable fixed point (sink) and the saddle point, respectively.
Figure 2. Phase regions of the model in the ( g , λ ) -plane for N = 12 , k = 9 , d = 2.2 . The red and white circles, as before, indicate the stable fixed point (sink) and the saddle point, respectively.
Universe 11 00114 g002
Table 1. Critical exponents for the temperature transition at λ = λ ( N , k ) .
Table 1. Critical exponents for the temperature transition at λ = λ ( N , k ) .
Parameters N > 2 , k = 1 N > 2 , k = 2 N = 12 , k = 9
d 2 + ε 2 + ε 2.2
t c ε / ( N 2 ) ε / ( N 2 ) 0.0247
α 1 2 / ε 1 2 / ε 9
β ( N 1 ) / [ 2 ( N 2 ) ] + 1.2278
γ 2 / ε 1 / ( N 2 ) 8.5443
δ ( 4 + ε ) N 8 3 ε ε ( N 1 ) 1 7.9589
ν 1 / ε 1 / ε 5
η ε / ( N 2 ) + 0.2911
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Gavrilik, A.M.; Nazarenko, A.V. Scaling Behavior and Phases of Nonlinear Sigma Model on Real Stiefel Manifolds Near Two Dimensions. Universe 2025, 11, 114. https://doi.org/10.3390/universe11040114

AMA Style

Gavrilik AM, Nazarenko AV. Scaling Behavior and Phases of Nonlinear Sigma Model on Real Stiefel Manifolds Near Two Dimensions. Universe. 2025; 11(4):114. https://doi.org/10.3390/universe11040114

Chicago/Turabian Style

Gavrilik, Alexandre M., and Andriy V. Nazarenko. 2025. "Scaling Behavior and Phases of Nonlinear Sigma Model on Real Stiefel Manifolds Near Two Dimensions" Universe 11, no. 4: 114. https://doi.org/10.3390/universe11040114

APA Style

Gavrilik, A. M., & Nazarenko, A. V. (2025). Scaling Behavior and Phases of Nonlinear Sigma Model on Real Stiefel Manifolds Near Two Dimensions. Universe, 11(4), 114. https://doi.org/10.3390/universe11040114

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop