Acta Polytechnica doi:10.14311/AP.2016.56.0180 Acta Polytechnica 56(3):180–192, 2016 © Czech Technical University in Prague, 2016 available online at http://ojs.cvut.cz/ojs/index.php/ap ON IMMERSION FORMULAS FOR SOLITON SURFACES Alfred Michel Grundlanda, b, ∗, Decio Levic, Luigi Martinad a Centre de Recherches Mathématiques, Université de Montréal, Montréal CP 6128 (QC) H3C 3J7, Canada b Department of Mathematics and Computer Science, Université du Québec, Trois-Rivières, CP 500 (QC) G9A 5H7, Canada c Dipartimento di Mathematica e Fisica dell’Università Roma Tre, Sezione INFN di Roma Tre, Via della Vasca Navale 84, Roma, 00146 Italy d Dipartimento di Mathematica e Fisica dell’Università del Salento, Sezione INFN di Lecce, Via Arnesano, C.P. 193 Lecce, 73100 Italy ∗ corresponding author: grundlan@crm.umontreal.ca Abstract. This paper is devoted to a study of the connections between three different analytic descriptions for the immersion functions of 2D-surfaces corresponding to the following three types of symmetries: gauge symmetries of the linear spectral problem, conformal transformations in the spectral parameter and generalized symmetries of the associated integrable system. After a brief exposition of the theory of soliton surfaces and of the main tool used to study classical and generalized Lie symmetries, we derive the necessary and sufficient conditions under which the immersion formulas associated with these symmetries are linked by gauge transformations. We illustrate the theoretical results by examples involving the sigma model. Keywords: integrable systems; soliton surfaces; immersion formulas; generalized symmetries. AMS Mathematics Subject Classification: 35Q35, 22E60, 53A05. 1. Introduction Soliton surfaces associated with integrable systems have been shown to play an essential role in many problems with physical applications (see e.g. [2, 5– 9, 11–13, 15, 19, 29, 30, 33–35]). We say that a surface is integrable if the Gauss-Mainardi-Codazzi equations corresponding to it are integrable, i.e. if they can be represented as the compatibility conditions for some “non-fake” Linear Spectral Problem (LSP) [2, 3, 5, 17, 18, 21, 22, 24–26, 31–35]. The possibility of using an LSP to represent a moving frame on the integrable surface has yielded many new results concerning the intrinsic geometric properties of such surfaces (see e.g. [4, 29]). In the present state of development, it has proved most fruitful to extend such characterizations of soliton surfaces via their immersion functions (see e.g. [1, 2, 19, 23, 27, 30] and references therein). The construction of surfaces related to completely integrable models was initiated by A. Sym [33–35]. This construction makes use of the conformal invari- ance of the zero-curvature representation of the system with respect to the spectral parameter. Another ap- proach for finding such surfaces has been formulated by J. Cieslinski and A. Doliwa [5–7] using gauge sym- metries of the LSP. A third approach, using the LSP for integrable systems and their symmetries has been introduced by Fokas and Gel’fand [8, 9], to construct families of soliton surfaces. Most recently, in a series of papers [11–13, 15], a reformulation and extension of the Fokas-Gel’fand immersion formula has been per- formed through the formalism of generalized vector fields and their actions on jet spaces. This extension has provided the necessary and sufficient conditions for the existence of soliton surfaces in terms of the symmetries of the LSP and integrable models. The objective of this paper is to investigate and construct the relation between the three approaches concerning 2D-soliton surfaces associated with integrable systems. This paper is organized as follows. Section 2 con- tains a brief summary of the results concerning the construction of soliton surfaces and symmetries. Us- ing the Sym-Tafel (ST) formula, we get the surface by differentiating the solution of the LSP with respect to the spectral parameter. Through the Cieslinski- Doliwa (CD) formula we apply a gauge transformation and through the Fokas-Gel’fand (FG) formula we con- sider the generalized symmetries of the associated integrable equation. In Section 3 we demonstrate that the immersion problem can be mapped through a gauge to any of the three immersion formulas listed above and we show that these formulas correspond to possibly different parametrizations of the same sur- face. Then, in Section 4, we apply the results on the example of the sigma model. Section 5 contains the concluding remarks. 180 http://dx.doi.org/10.14311/AP.2016.56.0180 http://ojs.cvut.cz/ojs/index.php/ap vol. 56 no. 3/2016 On Immersion Formulas for Soliton Surfaces 2. Summary of results on the construction of soliton surfaces In this section we recall the main tools used to study symmetries suitable for the use of Fokas-Gel’fand formulas for the construction of 2D surfaces. We make use of the formalism of vector fields and their prolongations as presented in [29]. More specifically, we rewrite the formula for the immersion functions of 2D surfaces in terms of the prolongation formalism of the vector fields instead of the Fréchet derivatives. 2.1. Classical and generalized Lie symmetries Let X (with coordinates xα, α = 1, . . . ,p) and U (with coordinates uk, k = 1, . . . ,q) be differential manifolds representing spaces of independent and dependent variables, respectively. Let Jn = Jn(X ×U) denote the n-jet space over X ×U. The coordinates of Jn are given by xα, uk and ukJ = ∂nuk ∂xj1...∂xjn where J = (j1, . . . ,jn) is a symmetric multi-index. We denote these coordinates by x and u(n). On Jn we can define a system of partial differential equations (PDEs) in p independent and q dependent variables given by m equations of the form Ωµ(x,u(n)) = 0, µ = 1, . . . ,m. (2.1) We consider a vector field v tangent to J0 = X ×U v = ξα(x,u)∂α + ϕk(x,u)∂k, (2.2) where ∂α = ∂/∂xα, ∂k = ∂/∂uk and we adopt the summation convention over repeated indices. Such a field defines vector fields, pr(n) v on Jn [29] pr(n) v = ξα∂α + ϕkJ ∂ ∂ukJ . (2.3) The functions ϕkJ are given by ϕkJ = DJR k + ξαukJ,α, R k = ϕk − ξαukα, (2.4) where the operators DJ correspond to multiple total derivatives, each of which is a combination of total derivatives of the form Dα = ∂α + ukJ,α ∂ ∂ukJ , α = 1, . . . ,p (2.5) and Rk are the so-called characteristics of the vector field v. In the following, the representation of v can be written equivalently as v = ξαDα + ωR, ωR = Rk ∂ ∂uk . (2.6) One says that the vector field v is a classical Lie point symmetry of a nondegenerate system of PDEs (2.1) if and only if its n-th prolongation of v is such that pr(n) vΩµ(x,u(n)) = 0, µ = 1, . . . ,m (2.7) whenever Ωµ(x,u(n)) = 0, µ = 1, . . . ,m are satisfied. By a totally non-degenerate system we mean a system for which all their prolongations have maximal rank and are locally solvable [29]. It follows from the well- known properties of the symmetries of a differential system that the commutator of two symmetries is again a symmetry. Thus, such symmetries form a Lie algebra g, which locally defines an action of a Lie group G on J0. Every solution of (2.1) can be represented by its graph, uk = θk(x), which is a section of J0. The symmetry group G transforms solutions into solutions. This means that a graph corresponding to one solution is transformed into a graph associated with another solution. If the graph is preserved by the group G or equivalently, if the vector fields v from the algebra g are tangent to the graph, then the related solution is said to be G-invariant. Invariant solutions satisfy, in addition to the equations (2.1), the characteristic equations equated to zero ϕka(x,θ) − ξ α a (x,θ)θ k ,α = 0, a = 1, . . . ,r, (2.8) where the index a runs over the generators of g. It may happen that invariant solutions are restricted in number or trivial if the full symmetry group is small. To extend the number of symmetries, and thus of solutions, one looks for generalized symmetries. They exist only if the nonlinear equation (2.1) is integrable [28], i.e. it has been obtained as the compatibility of a Lax pair (see (2.13) and in the following). A generalized vector field is expressed in terms of the characteristics ωR = Rk[u] ∂ ∂uk , (2.9) where [u] = (x,u(n)) ∈ Jn. The prolongation of an evolutionary vector field ωR is given by pr ωR = ωR + DJRk ∂ ∂ukJ . (2.10) A vector field ωR is a generalized symmetry of a non- degenerated system of PDEs (2.1) if and only if [29] pr ωRΩµ(x,u(n)) = 0, (2.11) whenever Ω(x,u(n)) = 0 and its differential conse- quences are satisfied. 2.2. The immersion formulas for soliton surfaces In order to analyse the Fokas-Gel’fand immersion formula for a surface in 2D, we briefly summarize the results obtained in [2, 5–9, 11–13, 15, 33–35]. Let us consider an integrable system of partial differ- ential equations (PDEs) in two independent variables x1,x2 and m dependent variables uk(x1,x2) written as Ω[u] = 0. (2.12) 181 A. M. Grundland, D. Levi, L. Martina Acta Polytechnica Suppose that the system (2.12) is obtained as the compatibility of a matrix LSP written in the form [33] ∂αΦ(x1,x2,λ) −Uα([u],λ)Φ(x1,x2,λ) = 0, α = 1, 2. (2.13) In what follows, the potential matrices Uα can be defined on the extended jet space N = (Jn,λ), where λ is the spectral parameter. The compatibility condi- tion of the LSP (2.13), often called the Zero-Curvature Condition (ZCC) D2U1 −D1U2 + [U1,U2] = 0, (2.14) which is assumed to be valid for all values of λ, implies (2.12). The bracket in (2.14) denotes the Lie algebra commutator. Equation (2.14) provides a representa- tion for the initial system (2.12) under consideration. The m-dimensional matrix functions Uα take values in some semisimple Lie algebra g and the wavefunc- tion Φ takes values in the corresponding Lie group G. We can say that, as long as the potential matri- ces Uα([u],λ) satisfy the ZCC (2.14), there exists a group-valued function Φ which satisfies (2.13). There exists a subclass of Φ which can be defined formally on the extended jet space N . For Φ belonging to this subclass we can write formally Φ = Φ([u],λ) ∈ G, meaning that Φ depends functionally on [u] and mero- morphically in λ. When Φ = Φ([u],λ) the LSP (2.13) can be written as Λα([u],λ) ≡ DαΦ([u],λ) −Uα([u],λ)Φ([u],λ) = 0, α = 1, 2, (2.15) which is a convenient form for the analysis we carry out in the following. In reference [9], the authors looked for a simultane- ous infinitesimal deformation of the associated LSP (2.15) which preserves the ZCC (2.14) Ũ1Ũ2 Φ̃   =  U1U2 Φ   + �  A1A2 Ψ   + O(�2), (2.16) where the matrices Ũ1, Ũ2, A1 and A2 take values in the Lie algebra g, while Φ̃ and Ψ belong to the corresponding Lie group G. The parameter λ is left invariant under the transformation (2.16) and 0 < � � 1. The corresponding infinitesimal generator formally takes the evolutionary form X̂e = A1∂U1 + A 2∂U2 + Ψ∂Φ. (2.17) Equation (2.17) can be written as X̂� = A1j∂Uj1 + A 2j∂ U j 2 + Ψj∂Φj , (2.18) where we decompose the matrix functions A1 and A2 in the basis ej, j = 1, . . . ,s for the Lie algebra g Uα = Ujαej ∈ g, [ei,ej] = c k ijek, (2.19) where [ , ] is the Lie algebra commutator and ckij are the structural constants of g. Since this generator X̂e does not transform λ, these symmetries preserve the singularity structure of the potential matrices Uα in the spectral parameter λ. The infinitesimal deformation of the LSP (2.13) and the ZCC (2.14) under the infinitesimal transformation (2.16) requires that the matrix functions Uα and Ψ satisfy, at first order in �, the equations DαΨ = UαΨ + AαΦ, α = 1, 2 (2.20) and D2A1 −D1A2 + [A1,U2] + [U1,A2] = 0, (2.21) The equation (2.21) coincides with the compatibility condition for (2.20). For the given matrix functions Uα, Aα ∈ g and Φ ∈ G satisfying equations (2.14), (2.15) and (2.21), an infinitesimal symmetry of the matrix system of the integrable PDEs (2.12) allows us to generate a 2D-surface immersed in the Lie algebra g. According to [8] this result is formulated as follows. Theorem 1. If the matrix functions Uα ∈ g, α = 1, 2 and Φ ∈ G of the LSP (2.15) satisfy the ZCC (2.14) and Aα ∈ g are linearly independent matrix func- tions which satisfy (2.21) and Φ ∈ G satisfies the LSP (2.15), then there exists (up to affine transforma- tions) a 2D-surface with a g-valued immersion function F ([u],λ) such that the tangent vectors to this surface are given by DαF([u],λ) = Φ−1Aα([u],λ)Φ, α = 1, 2. (2.22) Proof. The compatibility condition of (2.22) coincides with (2.21). So an immersion function F ([u],λ) exists and can be assumed to take its values in the Lie algebra g. If we define the matrix function Ψ = ΦF, (2.23) then, using (2.22), the function Ψ satisfies (2.20). Hence, since F = Φ−1Ψ, the formula Φ̃ = Φ + εΨ = Φ(I + εF) implies that Φ̃ is in the Lie group G. The immersion function F = ( Φ−1([u],λ) dΦ̃([u],λ,�) d� )∣∣∣∣ �=0 (2.24) is an element of the Lie algebra g. This shows that we have constructed an appropriate infinitesimal de- formation of the wavefunction Φ. In [5, 33] it was shown that the admissible symme- tries of the ZCC (2.14) include a conformal transfor- mation of the spectral parameter λ, a gauge trans- formation of the wavefunction Φ in the LSP (2.13) and generalized symmetries of the integrable system (2.12). All these symmetries can be used to determine explicitly a g-valued immersion function F of a 2D- surface. Thus, a generalization of the FG formula for immersion can be formulated as follows [9, 11]. 182 vol. 56 no. 3/2016 On Immersion Formulas for Soliton Surfaces Theorem 2. Let the set of scalar functions {uk} satisfy a system of integrable PDEs Ω[u] = 0. Let the G-valued function Φ([u],λ) satisfy the LSP (2.15) of g- valued potentials Uα([u],λ). Let us define the linearly independent g-valued matrix functions Aα([u],λ) (α = 1, 2) by the equations Aα([u],λ) = β(λ)DλUα + (DαS + [S,Uα]) + pr ωRUα + (pr ωR(DαΦ −UαΦ)) Φ−1. (2.25) Here β(λ) is an arbitrary scalar function of λ, S = S([u],λ) is an arbitrary g-valued matrix function de- fined on the jet space N , ωR = Rk[u]∂uk is the vector field, written in evolutionary form, of the generalized symmetries of the integrable PDEs Ω[u] = 0 given by the ZCC (2.14). Then there exists a 2D-surface with immersion function F([u],λ) in the Lie algebra g given by the formula (up to an additive g-valued constant) F([u],λ) = Φ−1 (β(λ)DλΦ + SΦ + pr ωRΦ) . (2.26) The integrated form of the surface (2.26) defines a mapping F : N → g and we will refer to it as the ST immersion formula (when S = 0,ωR = 0) [33–35] FST ([u],λ) = β(λ)Φ−1(DλΦ) ∈ g, (2.27) the CD immersion formula (when β = ωR = 0) [5–7] FCD([u],λ) = Φ−1S([u],λ)Φ ∈ g, (2.28) or the FG immersion formula (when β = 0,S = 0) [8, 9] FFG([u],λ) = Φ−1(pr ωRΦ) ∈ g. (2.29) 2.3. Application of the method The construction of soliton surfaces requires three ele- ments for an explicit representation of the immersion function F ∈ g: (1.) An LSP (2.13) for the integrable PDE. (2.) A generalized symmetry ωR of the integrable PDE. (3.) A solution Φ of the LSP associated with the soliton solution of the integrable PDE. Note that item (1.) is always required. In its pres- ence, even without one of the remaining two objects, we can obtain an immersion function F. When a solution Φ of the LSP is unknown, the geometry of the surface F can be obtained using the non-degenerate Killing form on the Lie algebra g. The 2D-surface with the immersion function F can be interpreted as a pseudo-Riemannian manifold. When the generalized symmetries ωR of the inte- grable PDE are unknown but we know a solution Φ of the LSP then we can define the 2D-soliton surface using the gauge transformation and the λ-invariance of the ZCC F = Φ−1(β(λ)DλΦ + SΦ), (2.30) where β(λ) is an arbitrary scalar function of λ and S is an arbitrary g-valued matrix function defined on the extended jet space N . Equation (2.30) is consistent with the tangent vectors DαF = β(λ)Φ−1(DλUα) + Φ−1(DαS + [S,Uα])Φ. (2.31) In all cases, the tangent vectors, given by (2.22), and the unit normal vector to a 2D-surface expressed in terms of matrices are DαF = Φ−1AαΦ ∈ g, N = Φ−1[A1,A2]Φ ( 12 tr[A1,A2] 2)1/2 ∈ g, (2.32) where the g-valued matrices Aα are given by (2.25). The first and second fundamental forms are given by I = gijdxidxj, II = bijdxidxj, i = 1, 2, (2.33) where gij = � 2 tr(AiAj), bij = � 2 tr ( (DjAi + [Ai,Uj])N ) , � = ±1. (2.34) This gives the following expressions for the mean and Gaussian curvatures H = 1 ∆ ( tr(A22) tr ( (D1A1 + [A1,U1])N ) − 8 tr(A1A2) tr ( (D2A1 + [A1,U2])N ) + tr(A21) tr ( (D2A2 + [A2,U2])N )) , K = 1 ∆ ( tr ( (D1A1 + [A1,U1])N ) · tr ( (D2A2 + [A2,U2])N ) − 2 tr2 ( (D2A1 + [A1,U2])N )) , ∆ = tr(A21) tr(A 2 2) − 4 tr(A1A2), (2.35) which are expressible in terms of Uα and Aα only. The study of soliton surfaces defined via the FG formula for immersion provides a unique mechanism for studying the relationship between these various characteristics of integrable systems. For example, do the infinite families of conservation laws have a geometric characterization? What is the geometry behind the Hamiltonian structure? Is there a geo- metric interpretation of the family of surfaces and frames associated with the spectral parameter? These answers will not only serve to construct surfaces with interesting geometric quantities but can also help to clarify some problems in the theory of integrable sys- tem. Following the three terms in (2.26) for the immer- sion of 2D-soliton surfaces in Lie algebras, we now show that there exists a relation between the Sym- Tafel, the Cieslinski-Doliwa and the Fokas-Gel’fand formulas. 183 A. M. Grundland, D. Levi, L. Martina Acta Polytechnica 3. Mapping between the Sym-Tafel, the Cieslinski-Doliwa and the Fokas-Gel’fand immersion formulas 3.1. λ-conformal symmetries and gauge transformations In this subsection we show that the ST immersion formula can always be represented by a gauge transfor- mation through the CD formula for immersion. The converse statement is also true: from a specific gauge it is always possible to determine the ST immersion formula for soliton surfaces. Proposition 1. A symmetry of the ZCC (2.14) of the LSP associated with an integrable system Ω[u] = 0 is a λ-conformal symmetry if and only if there exists a g-valued matrix function S1 = S1([u],λ) which is a solution of the system of differential equations DαS1 + [S1,Uα] = β(λ)DλUα, α = 1, 2. (3.1) Proof. First we show that for any λ-conformal sym- metry of the ZCC of the LSP associated with the integrable system Ω[u] = 0, there exists a g-valued matrix function S1 = S1([u],λ) which is a solution of the system of differential equations (3.1). Indeed, the linearly independent g-valued matrix functions Aα([u],λ) = β(λ)DλUα([u],λ), (3.2) associated with the λ-conformal symmetry of the ZCC (2.14) satisfy the infinitesimal deformation of the ZCC (2.21) and the corresponding ST immersion function is FST ([u],λ) = β(λ)Φ−1DλΦ ∈ g, (3.3) with linearly independent tangent vectors DαF ST = β(λ)Φ−1(DλUα)Φ, α = 1, 2. (3.4) On the other hand any g-valued matrix function can be written as the adjoint group action on its Lie algebra. This implies the existence of a g-valued matrix function S1([u],λ) for which the ST immersion formula (3.3) is the CD formula, i.e. FCD([u],λ) = Φ−1S1([u],λ)Φ ∈ g, (3.5) whose tangent vectors are found to be DαF CD = Φ−1 ( DαS1 +[S1,Uα] ) Φ, α = 1, 2. (3.6) By comparing the tangent vectors (3.4) and (3.6) we obtain (3.1). It remains to show that the system (3.1) is a solvable one. Indeed, from the compatibility condition of (3.1) we get β(λ)D2(DλU1) −β(λ)D1(DλU2) − [ β(λ)DλU2 − [S1,U2],U1 ] − [S1,D2U1] + [ β(λ)DλU1 − [S1,U1],U2 ] + [S1,D1U2] = 0, (3.7) which has to be satisfied whenever (3.1) holds. Using the ZCC (2.14) and the Jacobi identity it is easy to show that (3.7) is identically satisfied. So if we can find a gauge S1([u],λ) which satisfies (3.1), then the ST immersion formula (3.3) can always be represented by a gauge. Conversely, we show that for any g-valued matrix function S1 defined as a solution of the system of PDEs (3.1), there exists a λ-conformal symmetry of the ZCC of the LSP associated with the integrable system Ω[u] = 0. Indeed, comparing the immersion formulas (3.3) with (3.5) we find a linear matrix equa- tion for the wavefunction Φ β(λ)DλΦ = S1([u],λ)Φ. (3.8) If the gauge function S1([u],λ) is known, by solv- ing (3.8) we can determine the wavefunction Φ and consequently obtain the ST immersion formula for 2D-soliton surfaces. Therefore, the ST formula for immersion (2.27) is equivalent to the CD immersion formula (2.28) for the gauge S1, which satisfies differ- ential equation (3.1). 3.2. Generalized symmetries and gauge transformations In this subsection we discuss the links between gauge transformations and generalized symmetries of the ZCC associated with the integrable partial differential system Ω[u] = 0. We show that the immersion formula associated with the generalized symmetries (2.29) can always be obtained by a gauge transformation and the converse statement is also true. Proposition 2. A vector field ωR is a generalized symmetry of the ZCC (2.14) of the LSP associated with an integrable system Ω[u] = 0 if and only if there exists a g-valued matrix function (gauge) S2 = S2([u],λ) which is a solution of the system of differential equations DαS2 + [S2,Uα] = pr ωRUα + ( pr ωR(DαΦ −UαΦ) ) Φ−1. (3.9) Proof. First we demonstrate that for every infinitesi- mal generator ωR which is a generalized symmetry of the ZCC (2.14), there exists a g-valued matrix func- tion (gauge) S2 = S2([u],λ) which is a solution of the system of differential equations (3.9). Indeed an evolutionary vector field ωR is a generalized symmetry of the ZCC (2.14) if and only if pr ωR ( D2U1 −D1U2 + [U1,U2] ) = 0, (3.10) whenever the ZCC (2.14) holds. Equation (3.10) is equivalent to the infinitesimal deformation of the ZCC (2.14) given by (2.21), with linearly independent g- valued matrix functions Aα([u],λ) = pr ωRUα + ( pr ωR(DαΦ −UαΦ) ) Φ−1, α = 1, 2. (3.11) 184 vol. 56 no. 3/2016 On Immersion Formulas for Soliton Surfaces In the derivation of (3.11) we use the fact that the total derivatives Dα commute with the prolongation of a vector field ωR written in the evolutionary form [29] [Dα, pr ωR] = 0, α = 1, 2. (3.12) Using the LSP (2.15), equation (3.11) can be written in the equivalent form Aα([u],λ) = [ −Uα(pr ωRΦ) + pr ωR(DαΦ) ] Φ−1. α = 1, 2 (3.13) Substituting (3.13) into (2.21) we obtain( −D2U1 + D1U2 − [U1,U2] ) (pr ωRΦ)Φ−1 = 0, which is satisfied identically whenever the ZCC (2.14) holds. An integrated form of the immersion function FFG([u],λ) of a 2D-surface associated with a general- ized symmetry ωR of the ZCC (2.14) and the tangent vectors (2.22) is given by the FG formula FFG([u],λ) = Φ−1(pr ωRΦ) ∈ g. (3.14) The fact that any g-valued matrix function can be written under the adjoint group action implies that there exists a g-valued gauge S2, such that (3.5) holds for the CD immersion function and its tangent vectors DαF CD given by (3.6). Comparing equations (2.22) and (3.11) with (3.6) we get (3.9). Let us show that the system (3.9) always possesses a solution. The compatibility condition of (3.9), when- ever (3.9) and (2.21) hold, implies the relation [S2,D2U1 −D1U2] + [ [S2,U1],U2 ] − [ [S2,U2],U1 ] = 0, (3.15) which is identically satisfied in view of the ZCC (2.14) and the Jacobi identity. So, if we can find a gauge function S2([u],λ) which satisfies (3.9), then the FG formula (3.14) can always be represented by a gauge transformation. The converse statement is also true. We show that for any g-valued matrix function S2 defined as a solu- tion of the system of PDEs (3.9), there exists a gener- alized symmetry ωR of the ZCC of the LSP associated with Ω[u] = 0. Indeed, let the LSP of Ω[u] = 0 admit a gauge symmetry. If the gauge S2([u],λ) is given, then the immersion function FCD of a 2D-surface can be integrated explicitly [5–7] FCD([u],λ) = Φ−1S2([u],λ)Φ ∈ g, (3.16) whenever the tangent vectors DαF CD = Φ−1 ( DαS2 + [S2,Uα] ) Φ. (3.17) are linearly independent. It is straightforward to verify that the characteristics of a generalized vector field ωSR, written in evalutionary form, associated with a gauge symmetry S2, can be expressed as Aα = DαS2 + [S2,Uα] ∈ g. (3.18) The matrices Aα identically satisfy the determining equations (2.21) which are required for ωSR to be a generalized symmetry of the ZCC Ω[u] = 0 D2A1 −D1A2 + [A1,U2] + [U1,A2] = pr ωSR ( D2U1 −D1U2 + [U1,U2] ) = 0, (3.19) whenever Ω[u] = 0 holds. Hence, the vector field ωSR associated with a gauge symmetry S2 is given by ωSR = ( DαS2 + [S2,Uα] )j ∂ ∂U j α , (3.20) where we have decomposed the matrix functions Aα and Uα in the basis {ej}n1 for the Lie algebra Uα = Ujαej ∈ g, DαS2 + [S2,Uα] = ( DαS2 + [S2,Uα] )j ej. (3.21) Hence, for any smooth g-valued gauge S2([u],λ) there exists a generalized symmetry ωSR of the ZCC (2.14) and the converse statement holds as well. Comparing the FG formula for immersion (3.14) with the CD immersion formula (3.5) we find the gauge S2 = (pr ωRΦ)Φ−1. (3.22) Hence the FG formula for immersion (2.29) is equiva- lent to the CD immersion formula (2.28) for the gauge S2 satisfying (3.9). 3.3. The Sym-Tafel immersion formula versus the Fokas-Gel’fand immersion formula Under the assumptions of Propositions 1 and 2, we have the following result. Proposition 3. Let S1 and S2 be the two g-valued matrix functions determined in Propositions 1 and 2, respectively in terms of a λ-conformal symmetry and a generalized symmetry of the ZCC (2.14) of the LSP associated with an integrable system Ω[u] = 0. If the gauge S2 is a non-singular matrix then there exists a matrix M = S1S−12 such that β(λ)(DλΦ) = M(pr ωRΦ). (3.23) The matrix M defines a mapping from the FG im- mersion formula (3.14) to the ST immersion formula (3.3). Alternatively, if the gauge S1 is a non-singular ma- trix then there exists a matrix M−1 such that (pr ωRΦ) = M−1β(λ)(DλΦ). (3.24) The matrix M−1 defines a mapping from the ST im- mersion formula (3.3) to the FG immersion formula (3.14). 185 A. M. Grundland, D. Levi, L. Martina Acta Polytechnica FST = β(λ)Φ−1(DλΦ) ∈ g Φ ∈ G FFG = Φ−1(pr ωRΦ) ∈ g S2◦S−11 S1∈g S2∈g S1◦S−12 Figure 1. Representation of the relations between the wavefunction Φ ∈ G and the g-valued ST and FG formulas for immersions of 2D-soliton surfaces. Proof. Equation (3.23) or (3.24) is obtained by elimi- nating the wavefunction Φ from the right-hand side of equations (3.8) and pr ωRΦ = S2Φ, (3.25) respectively. So the link between the immersion func- tions FST and FFG exists, up to a g-valued gauge function. It should be noted that in order to recover soliton surfaces, we have to perform an integration with re- spect to the curvilinear coordinates in the case of the FG formula. Alternatively, by using the ST immer- sion formula, we obtain the same soliton surface by differentiating the wavefunction Φ with respect to the spectral parameter λ. The connection between the FG and ST approaches for determining the immersion functions FST and FFG of 2D-surfaces is obtained through the gauge matrix functions M or M−1 from the equation (3.23) or (3.24), respectively (see Fig. 1). We can also write direct equations relating the generalized symmetries ωR with the Sym-Tafel λ- conformal symmetry for the ZCC (2.14), eliminating the gauge S2([u],λ) in (3.9) by using (3.8). So we get β(λ)(DλΦ)Uα −β(λ)ΦUαΦ−1(DλΦ) + β(λ)(DλUα)Φ + Φ [ −pr ωR(DαΦ) + Uα(pr ωRΦ) ] Φ−1 = 0. (3.26) However, equations (3.26) are nonlinear differential equations for the wavefunction Φ, which in general are not easy to solve. To conclude, in all three cases we give explicit ex- pressions for 2D-soliton surfaces immersed in the Lie algebra g and demonstrate that one such surface can be transformed to another one through a gauge. 4. The sigma model and soliton surfaces For the sake of generality we start by considering the general CPN−1 model. The problem of constructing integrable surfaces associated with the CPN−1 models and their deformations under various types of dynam- ics have generated a great deal of interest over the past decades [1, 23, 37]. The most fruitful approach to the study of general properties of this model has been formulated through descriptions of the model in terms of rank-one Hermitian projectors. A matrix P (z, z̄) is said to be a rank-one Hermitian projector if P 2 = P, P = P†, tr P = 1. (4.1) The target space of the projector P is determined by a complex line in CN, i.e. by a one-dimensional vector function f(z, z̄) given by P = f ⊗f† f†f , (4.2) where f is the mapping C ⊇ Ω 3 z = x + iy 7→ f = (f0,f1, . . . ,fN−1)CN\{0}. Equation (4.2) gives an isomorphism between the equivalence classes of the CPN−1 model and the set of rank-one Hermitian projectors P . The equations of motion Ω(P) = [∂+∂−P,P ] = 0, ∂± = 1 2 (∂1 ± i∂2), ∂1 = ∂x, ∂2 = ∂y (4.3) and other properties of the model take a compact form when the model is written in terms of the projector. Now we present some examples which illustrate the theoretical considerations presented in the previous section. Our first example shows that the integrated form of the surface associated with the CPN−1 model admits conformal symmetries which depend on two arbitrary functions of one complex variable. This model is defined on the Riemann sphere S2 = C∪{∞} and its action functional is finite [37]. An entire class of solutions of (4.3) is obtained by acting on the holomorphic (or anti-holomorphic) solution P [10] with raising and lowering operators. These operators are given by Π±(P) =   (∂±P)P(∂∓P) tr(∂±PP∂∓P) for (∂±P)P(∂∓P) 6= 0, 0 for (∂±P)P(∂∓P) = 0, Π−(Pk) = Pk−1, Π+(Pk) = Pk+1. (4.4) The set of N rank-1 projectors {P0, . . . ,PN−1} acts on orthogonal complements of one-dimensional subspaces in CN and satisfy the orthogonality and completeness relations PjPk = δjkPj, (no summation) and N−1∑ j=0 Pj = IN, (4.5) where IN is the N ×N identity matrix on CN. These projectors provide a basis of commuting elements in the space of the Hermitian matrices on CN and satisfy the Euler-Lagrange equation (written in the form of a conservation law) ∂[∂̄Pk,Pk] + ∂̄[∂Pk,Pk] = 0, k = 0, 1, . . . ,N − 1, (4.6) where ∂ = 12 (∂x − i∂y) and ∂̄ = 1 2 (∂x + i∂y). For a given set of rank-1 projector solutions Pk of (4.6) 186 vol. 56 no. 3/2016 On Immersion Formulas for Soliton Surfaces the su(N)-valued generalized Weierstrass formula for immersion (GWFI) [20] Fk(z, z̄) = i ∫ γ (−[∂Pk,Pk]dz + [∂̄Pk,Pk]dz̄), k = 0, 1, . . . ,N − 1 (4.7) (where γ is a curve locally independent of the trajec- tory in C) can be explicitly integrated [10] Fk(z, z̄) = −i ( Pk + 2 k−1∑ j=0 Pj ) + 1 + 2k N IN. (4.8) The immersion functions Fk satisfy the algebraic con- ditions [Fk − ickIN ][Fk − i(ck−1)IN ][Fk − i(ck−2)IN ] = 0, 0 < k < N − 1, [F0 − ic0IN ][F0 − i(c0 − 1)IN ] = 0, [FN−1 + ic0IN ][FN−1 + i(c0 − 1)IN ] = 0, N−1∑ j=0 (−1)jFj = 0, ck = 1 N (1 + 2k). (4.9) The LSP associated with (4.6) is given by [27, 36] ∂αΦk = UαkΦk, Uαk = 2 1 ±λ [∂αPk,Pk], (U1k)† = −U2k, (4.10) (where α = 1, 2 stands for ±) with soliton solution Φk = Φk([P ],λ) ∈ SU(N) which goes to IN as λ →∞ [36, 37] Φk = IN + 4λ (1 −λ)2 k−1∑ j=0 Pj − 2 1 −λ Pk, Φ−1k = IN − 4λ (1 + λ)2 k−1∑ j=0 Pj − 2 1 + λ Pk, λ = it, t ∈ R. (4.11) The recurrence relation (4.4) is expressed in terms of rank-1 projectors Pk, without any reference to the sequence of functions fk as in (4.2). For the sake of simplicity, in this section, we drop the index k attributed to the N projectors Pk. It is convenient for computational purposes to ex- press the CPN−1 model in terms of the matrix θ ≡ i ( P − 1 N IN ) = θses ∈ su(N), [ej,el] = Csjles, j, l,s = 1, . . . ,N 2 − 1, (4.12) where Csjl are the structural constants of g and es is the basis element for the su(N) algebra. Due to the indempotency of the projector P we get the following algebraic restriction on θ: θ ·θ = −i 2 −N N θ + 1 −N N2 IN ⇐⇒ P 2 = P. (4.13) The equations of motion in terms of the matrix θ are Ωj[θ] = [ (∂21 + ∂ 2 2 )θ,θ ]j = 0, j = 1, . . . ,N2 − 1 (4.14) where [·, ·]j denotes the coefficients of the commutator with respect to the jth basis element ej for the su(N) algebra. The potential matrices Uα in terms of θ are U1 = −2 1 −λ2 ( [∂1θ,θ] − iλ[∂2θ,θ] ) , U2 = −2 1 −λ2 ( iλ[∂1θ,θ] + [∂2θ,θ] ) , λ = it, t ∈ R. (4.15) The wavefunction Φ in terms of θ is Φ([θ],λ) = IN + 4λ (1 −λ)2 k−1∑ j=0 Πj−(θ) − 2 1 −λ ( 1 N IN − iθ ) ∈ SU (N), (4.16) where Π± are the raising and lowering operators acting on the elements θ of the algebra su(N) Π−(θk) = θk−1, Π+(θk) = θk+1. (4.17) In what follows, we use the simplified notation of Π±(θk) by Π±(θ), where the index k is suppressed. The operators (4.17) are written explicitly as Π−(θ) = ∂̄θ(E − iθ)∂θ tr(∂̄θ(E − iθ)∂θ) , Π+(θ) = ∂θ(E − iθ)∂̄θ tr(∂θ(E − iθ)∂̄θ) , E = 1 N IN, (4.18) where the traces in the denominators are different from zero unless the whole matrix is zero. For any functions f and g of one variable, the equations of motion (4.14) and their LSP (2.15) (with the potential matrices (4.15)) admit the conformal symmetries ωCi = [ f(x)∂1θj + g(y)∂2θj ] ∂ ∂θj , i = 1, 2. (4.19) The vector fields ωCi are related to the fields ηci defined on the jet space M = [(Φ,Uα)] ηCi = (∂iΦ j) ∂ ∂Φj + (∂iUjα) ∂ ∂U j α , i = 1, 2, (4.20) which are conformal symmetries of the LSP (2.15). The integrated form of the surface is given by the FG formula [15] FFG = Φ−1 ( f(x)U1 + g(y)U2 ) Φ ∈ su(N). (4.21) 4.1. Soliton surfaces associated with the CP 1 sigma model We give a simple example to illustrate the construc- tion of 2D-soliton surfaces associated with the CP 1 187 A. M. Grundland, D. Levi, L. Martina Acta Polytechnica model (N = 2) introduced in previous section. The only solutions with finite action of the CP 1 model are holomorphic and antiholomorphic projectors [37]. The rank-one Hermitian projectors (i.e. holomorphic P0 and antiholomorphic P1) based on the Veronese sequence f0 = (1,z), take the form P0 = f0 ⊗f † 0 f † 0f0 = 1 1 + |z|2 ( 1 z̄ z |z|2 ) , P1 = f1 ⊗f † 1 f † 1f1 = 1 1 + |z|2 ( |z|2 −z̄ −z 1 ) , (4.22) where f1 = (I2 − P0)∂f0. The corresponding inte- grated forms of the surfaces are given by the GWFI (4.8) F0 = i( 1 2 I2 −P0) = i 1 + |z|2 (1 2 (|z| 2 − 1) −z̄ −z 12 (1 −|z| 2) ) ∈ su(2), F1 = −i(P1 + 2P0) + 3i 2 I2 = F0. (4.23) From equation (4.10) the potential matrices Uαk be- come U10 = U11 = 2 (λ + 1)(1 + |z|2)2 ( −z̄ −z̄2 1 z̄ ) , U20 = U21 = 2 (λ− 1)(1 + |z|2)2 ( −z 1 −z2 z ) , λ = it, t ∈ R. (4.24) The SU(2)-valued soliton wavefunctions Φk in the LSP (4.10) for the CP 1 model have the form Φ0 = 1 1 + |z|2 ( −i+t+(i+t)|z|2 t−i −2iz̄ t−i −2iz t+i i+t+(t−i)|z|2 t+i ) , Φ1 = 1 1 + |z|2 ( 1+t2+(t+i)2|z|2 (t−i)2 2(1−it)z̄ (t−i)2 −2i(t−i)z (t+i)2 1+t2+(t−i)2|z|2 (t+i)2 ) . (4.25) Let us now consider separately four different ana- lytic descriptions for the immersion functions of 2D- soliton surfaces in Lie algebras which are related to four different types of symmetries. I. The ZCC (2.14) of the CP 1 model admits a con- formal symmetry in the spectral parameter λ. The tangent vectors DαFSTk associated with this symme- try are given by DαF ST k = Φ −1 k (DλUαk)Φk, α,k = 1, 2 and are linearly independent. The integrated forms of the 2D-surfaces in su(2) are given by the ST formulas (2.27) FST0 = Φ −1 0 (DλΦ0) = 2i (1 + t2)2(1 + |z|2)2 · ( −|z|2[t2 − 3 + |z|2(1 + t2)] z[(t− i)2 + |z|2(3 − 2it + t2)] z̄[(t + i)2 + |z|2(3 + 2it + t2)] |z|2[t2 − 3 + |z|2(t2 + 1)] ) , FST1 = Φ −1 1 (DλΦ1) = 2i (1 + t2)2(1 + |z|2)2 · ( −[(t2 + 1)(1 + 2|z|4) + 3|z|2(t2 − 3)] z[t2 − 6it− 5 + |z|2(7 + t2 − 6it)] z̄[6it− 5 + t2 + |z|2(7 + 6it + t2)] (t2 + 1)(1 + 2|z|4) + 3|z|2(t2 − 3) ) , (4.26) where, without loss of generality, we can put β(λ) = 1 in the expression (2.27). The surfaces FSTk satisfy (FSTk ) 2 + 1 4 I2 = 0, for k = 0, 1 and have positive constant Gaussian and mean curvatures. Hence, they are spheres (see Fig. 2a) K0 = K1 = 4, H0 = H1 = 4. (4.27) The surfaces FSTk in Cartesian coordinates (x,y) take the form for z = x + iy FSTk = { x 1 + x2 + y2 , y 1 + x2 + y2 , 1 −x2 −y2 2(1 + x2 + y2) } (4.28) The su(2)-valued gauges SSTk associated with the ST immersion functions FSTk (4.26) take the form SST0 = (DλΦ0)Φ −1 0 = 2i 1 + |z|2 ( −|z|2 t2+1 z̄ (t−i)2 z (t+i)2 |z|2 t2+1 ) , SST1 = (DλΦ1)Φ −1 1 = 2i 1 + |z|2 ( −(1+2|z|2) t2+1 z̄(t+i)2 (t−i)4 z(t−i)2 (t+i)4 1+2|z|2 t2+1 ) , det SSTk 6= 0, tr S ST k = 0. (4.29) II. The surfaces Fgk ∈ su(2) associated with the scaling symmetries of the ZCC (2.14) associated with equations of the CP 1 model (4.6) ω g k = ( D1(zU1k) + z̄(D2U1k) ) ∂ ∂θ1 + ( z(D1U2k) + D2(z̄U2k) ) ∂ ∂θ2 , (4.30) have the integrated form [14] F g k = Φ −1 k (zU1k + z̄U2k)Φk, k = 0, 1 (4.31) where θ1 and θ2 are complex-valued functions deter- mined from the su(2) Lie algebra (4.12). The surfaces F g k also have constant positive curvatures K0 = K1 = −4λ2, H0 = H1 = −4iλ, iλ ∈ R. (4.32) 188 vol. 56 no. 3/2016 On Immersion Formulas for Soliton Surfaces The surfaces Fgk are not spheres (as in the previous cases (4.27)) since they have boundaries (see Fig. 2b). The surfaces Fgk can be given in the parametric form F g k = (x3 − 2x2y + x(y2 − 1) − 2y(1 + y2) (1 + x2 + y2)2 , − 2x3 + x2y + y(y2 − 1) + 2x(1 + y2) (1 + x2 + y2)2 , 2(x2 + y2) (1 + x2 + y2)2 ) . (4.33) The su(2)-valued gauges Sgk = (pr ω gΦk)Φ−1k associ- ated with the scaling symmetries ωgk are given by S g 0 = S g 1 = 2 (t2 + 1)(1 + |z|2)2 · ( 2it|z|2 iz̄[i− t + |z|2(t + i)] z[1 − it + |z|2(1 + it)] −2it|z|2 ) , (4.34) where det Sgk 6= 0. III. In the case of surfaces associated with the con- formal symmetries ωck = −gk(z)∂ − ḡk(z̄)∂̄, (4.35) where for simplicity we have assumed that gk(z) = 1 + i, the su(2)-valued immersion functions Fck are given by [12] Fck = Φ −1 k (U1k + U2k)Φk, (4.36) where U10 + U20 = U11 + U21 = 2 (t2 + 1)(1 + |z|2)2 · ( 2z + i(t + i)(z + z̄) −1 − it + iz̄2(t + i) 1 + z2 + it(z2 − 1) −[2z + i(t + i)(z + z̄)] ) . (4.37) By further computation it can be verified that the Gaussian curvature and the mean curvature corre- sponding to the surfaces Fck are not constant for any value of λ. The fact that the surfaces Fck have the Euler-Poincaré characters [10] χk = −1 π ∫∫ S2 ∂∂̄ ln [tr(∂Pk · ∂̄Pk)]dx1dx2 (4.38) equal to 2 and positive Gaussian curvature K > 0 means that the surfaces Fck are homeomorphic to ovaloids (see Fig. 2c). The surfaces Fck associated with the conformal symmetries ωck are cardioid surfaces which can be parametrized as follows Fck = (x2 − 1 − 4xy −y2 (1 + x2 + y2)2 ,− 2(1 + x2 + xy −y2) (1 + x2 + y2)2 , 2(2x−y) (1 + x2 + y2)2 ) (4.39) The su(2)-valued gauges Sck associated with the conformal symmetries ωcktake the form Sc0 = (pr ω cΦ0)Φ−10 = 2 (1 + |z|2)2 · ( −i(1−i)(t−i)z+(1+i)(1−it)z̄ t2+1 i(1+i)(t+i)−(1−i)z2(t−i) (t+i)2 (1−i)(1+it)+(1+i)(1−it)z̄2 (t−i)2 (1−i)(1+it)z+i(1+i)(t+i)z̄ t2+1 ) , Sc1 = (pr ω cΦ1)Φ−11 = 2 (1 + |z|2)2 · ( −i(1−i)(t−i)z+(1+i)(1−it)z̄ t2+1 i(t−i)2[(1+i)(t+i)−(1−i)(t−i)z2] (t+i)4 (t+i)2[(1−i)(1+it)+(1+i)(1−it)z̄2] (t−i)4 (1−i)(1+it)z+i(1+i)(t+i)z̄ t2+1 ) , (4.40) where det Sck 6= 0. IV. If the generalized symmetries of the CP 1 model (4.6) are written in the evolutionary form ωRk = ( D21U1k + D 2 2U1k + [D1U1k,U1k] +[D2U1k,U1k] ) ∂ ∂θ1 + ( D21U2k +D 2 2U2k +[D2U2k,U2k] + [D1U2k,U2k] ) ∂ ∂θ2 , k = 1, 2 (4.41) (where θ1 and θ2 are complex-valued functions ob- tained from (4.12)) then the su(2)-valued integrated form of the immersion becomes [14] FFGk = Φ −1(pr ωRk Φk) = Φ −1 k (D1U1k + D2U2k)Φk. (4.42) The tangent vectors to this surface are given by D1F FG k = Φ −1 k (pr ω R k U1k)Φk = Φ −1 k (D 2 1U1k + D 2 2U1k +[D1U1k,U1k] + [D2U1k,U1k])Φk, D2F FG k = Φ −1 k (pr ω R k U2k)Φk = Φ −1 k (D 2 1U2k + D 2 2U2k +[D2U2k,U2k] + [D1U2k,U2k])Φk. (4.43) The surfaces FFGk also have positive Gaussian curva- tures K > 0 and the Euler-Poincaré characters are equal to 2. In the parametrization x,y, the surfaces FFGk take the form FFGk = ( − x3 − 6x2y −x(1 + 3y2) + 2y(1 + y2) (1 + x2 + y2)3 , 2x3 + y + 3x2y −y3 + x(2 − 6y2) (1 + x2 + y2)3 , − 2(x2 − 4xy −y2) (1 + x2 + y2)3 ) (4.44) and they are homeomorphic to ovaloids. The su(2)-valued gauges SFGk associated with the generalized symmetry ωRk take the form SFGk = (pr ω R k Φk)Φ −1 k = D1U1k + D2U2k. (4.45) 189 A. M. Grundland, D. Levi, L. Martina Acta Polytechnica Under the assumption that Pk are holomorphic or antiholomorphic projectors (4.22) the gauges SFGk take the form SFG0 = S FG 1 = 4 (t2 + 1)(1 + |z|2)3 · ( −z2(1 + it) + z̄2(1 − it) z̄3(1 − it) + z(it + 1) −iz3(t− i) + iz̄(t + i) z2(1 + it) − z̄2(1 − it) ) , (4.46) where det SFGk 6= 0. Hence the mappings Mk = SSTk (S FG k ) −1 from the FG immersion formulas to the ST immersion formulas are given by M0 = SST0 (S FG 0 ) −1 = 1 2(t2 + 1) · ( −2iz3(t−i)+z̄[(t+i)2+|z|2(t2+1)] z(t−i) − [z 3(1+t2)+2z̄(1−it)+z(t−i)2] t+i z[(t+i)2+(t2+1)|z|2+2(1+it)] t−i z(t−i)2+|z|2z(t2+1)+2iz̄3(t+i) z̄(t+i) ) , M1 = SST1 (S FG 1 ) −1 = i 2|z|2 ( −(1+2|z|2) t2+1 (i+t)2z̄ (t−i)4 z(t−i)2 (t+i)4 1+2|z|2 t2+1 ) · ( z2(it + 1) + z̄2(it− 1) −z(it + 1) + z̄3(it− 1) z3(it + 1) + (1 − it)z̄ −z2(1 + it) + (1 − it)z̄2 ) , (4.47) where det Mk 6= 0. Conversely, the gauges M−1k = SFGk (S ST k ) −1 do exist. Hence there exist mappings from the ST immersion formulas to the FG immersion formulas M−10 = S FG 0 (S ST 0 ) −1 = 2 (1 + |z|2)3 ( m11 m12 m21 m22 ) , (4.48) where m11 = (t− i)2z + (1 + t2)|z|2z + 2(it− 1)z̄3 (i + t)z̄ , m12 = 2(1 + it)z2 + (i + t)2z̄2 + (1 + t2)|z|2z̄2 (i− t)z , m21 = (t− i)2z2 + (1 + t2)|z|2z2 + 2(1 − it)z̄2 (i + t)z̄ , m22 = −2(1 + it)z3 + (i + t)2z̄ + (1 + t2)|z|2z̄ (t− i)z , and M−11 =S FG 1 (S ST 1 ) −1 = 2 (t2 + 1)(1 + |z|2)3(1 + 4|z|2) · ( −z2(1 + it) + z̄2(1 − it) z̄3(1 − it) + z(1 + it) −z3(1 + it) + z̄(it− 1) z2(1 + it) − z̄2(1 − it) ) · ( (1 + 2|z|2)(1 + it)(i + t) −iz̄(i+t) 4 (t−i)2 −iz(t−i)4 (i+t)2 (1 + 2|z| 2)(1 − it)(−i + t) ) . (4.49) (a) (b) (c) (d) Figure 2. Surfaces F ST0 in (a), F g 0 in (b), F c 0 in (c) and F F G0 in (d) for k = 0, 1 and λ = i/2, and ξ± = x ± iy with x,y ∈ [−5, 5]. The axes indi- cate the components of the immersion function in the basis for su(2), e1 = ( 0 i i 0 ) , e2 = ( 0 −1 1 0 ) , e3 = ( i 0 0 −i ) . 5. Concluding remarks In this paper we have shown how three different analytic descriptions for the immersion function of 2D-soliton surfaces can be related through different g-valued gauge transformations. The existence of such gauges is demonstrated by reducing the problem to that of mappings between different forms of the immersion formulas for three types of symmetries : conformal transformations in the spectral parameter, gauge symmetries of the LSP and generalized symme- tries of the integrable systems. We have investigated the geometric consequences of these mappings and rephrased them as requirements for the existence of the corresponding vector fields and their prolonga- tions acting on a solution Φ of the associated LSP for an integrable PDE. The explicit expressions for these relations, which we have established, have provided us with a tool for distinguishing between the cases in which soliton surfaces can be or cannot be related among them, see Proposition 3. The task of finding an increasing number of soliton surfaces associated with integrable systems is related to the symmetry properties of these systems. The construction of soliton surfaces started with the con- tribution of Sym [33, 34] and Tafel [35] providing a formula for the immersion of integrable surfaces which are extensively used in the literature (see e.g. 190 vol. 56 no. 3/2016 On Immersion Formulas for Soliton Surfaces [2, 5–16, 30, 33–35]). In this paper we have addressed the question and formulated easily verifiable condi- tions which ensure that the ST formula produces a desired result. This advance can assist future studies of 2D-soliton surfaces with integrable models, which can describe more diverse types of surfaces than the ones discussed in three-dimensional Euclidean space for the CP 1 sigma model. It may be worthwhile to extend the investigation of soliton surfaces to the case of the sigma models defined on other homogeneous spaces via Grassmann models and possibly to models associated with octonion geometry. This case could lead to different classes and types of surfaces than those studied in this paper. This task will be explored in a future work. Acknowledgements AMG has been partially supported by a research grant from NSERC of Canada and would also like to thank the Dipartimento di Mathematica e Fisica of the Universitá Roma Tre and the Dipartimento di Mathematica e Fisica of Universitá del Salento for its warm hospitality. DL has been partly supported by the Italian Ministry of Education and Research, 2010 PRIN Continuous and discrete nonlinear integrable evolutions: from water waves to symplectic maps. LM has been partly supported by the Italian Ministry of Education and Research, 2011 PRIN Teorie geometriche e analitiche dei sistemi Hamiltoniani in dimensioni finite e infinite. DL and LM are also supported by INFN IS-CSN4 Math- ematical Methods of Nonlinear Physics. References [1] Babelon O, Bernard D and Talon M 2006 Introduction to Classical Integrable Systems (Cambridge Monographs on Mathematical Physics) (Cambridge: Cambridge University Press) doi:10.1017/CBO9780511535024 [2] Bobenko AI (1994) Surfaces in terms of 2 by 2 matrices. Old and new integrable cases in Harmonic Maps and Integrable Systems, Eds Fordy A, Wood J (Braunschwieg, Vieweg) doi:10.1007/978-3-663-14092-4_5 [3] Calogero F and Nucci MC, Lax pairs galore, J. Math. Phys., 32 72–74 (1991) doi:10.1063/1.529096 [4] Cartan E 1953 Sur la Structure des Groupes Infinis de Transformation Chapitre I. Les Systèmes Différentiels en Involution (Paris, Gauthier-Villars). [5] Cieśliński J 1997 A generalized formula for integrable classes of surfaces in Lie algebras Journal of Mathematical Physics 38 4255–4272, doi:10.1063/1.532093 [6] Cieslinski J 2007 Pseudospherical surfaces on time scales: a geometric deformation and the spectral approach J. Phys. A: Math. Theor. 40 12525-38, doi:10.1088/1751-8113/40/42/S02 [7] Doliwa A and Sym A 1992 Constant mean curvature surfaces in E3 as an example of soliton surfaces, Nonlinear Evolution Equations and Dynamical Systems (Boiti M, Martina L and Pempinelli F, eds, World Scientific, Singapore, pp 111-7) [8] Fokas A S and Gel’fand I M 1996 Surfaces on Lie groups, on Lie algebras, and their integrability Comm. Math. Phys. 177 203–220 [9] Fokas A S, Gel’fand I M, Finkel F and Liu Q M 2000 A formula for constructing infinitely many surfaces on Lie algebras and integrable equations Sel. Math. 6 347–375 doi:10.1007/PL00001392 [10] Goldstein P P and Grundland A M 2010 Invariant recurrence relations for CP N −1 models J. Phys. A: Math Theor. 43, 265206 (18pp), doi:10.1088/1751-8113/43/26/265206 [11] Grundland AM 2016 Soliton surfaces in generalized symmetry approach J. Theor. Math. Phys. (accepted.) [12] Grundland A M and Post S 2011 Soliton surfaces associated with generalized symmetries of integrable equations J. Phys. A.: Math. Theor.44 165203 (31pp) doi:10.1088/1751-8113/44/16/165203 [13] Grundland A M and Post S 2012 Surfaces immersed in Lie algebras associated with elliptic integrals J Phys A: Math Theor 45, 015204 (20pp), doi:10.1088/1751-8113/45/1/015204 [14] Grundland AM and Post S 2012 Soliton surfaces associated with CP N −1 sigma models J. Phys. Conf. Series 380, 012023 pp 1-14, doi:10.1088/1742-6596/380/1/012023 [15] Grundland AM, Post S and Riglioni D 2014 Soliton surfaces and generalized symmetries of integrable systems J. Phys. A.: Math. Theor.47 015201 (14pp) doi:10.1088/1751-8113/47/1/015201 [16] Grundland AM and Yurdusen I 2009 On analytic descriptions of two-dimensional surfaces associated with the CP N −1 sigma model, J. Phys. A: Math. Theor. 42 172001 doi:10.1088/1751-8113/42/17/172001 [17] Gubbiotti G, Scimiterna C and Levi D 2016 Linearizability and fake Lax pair for a consistent around the cube nonlinear non–autonomous quad–graph equation, Teor. Math. Phys., in press. [18] Hay M and Butler S, Simple identification of fake Lax pair, arXiv:1311.2406v1. [19] Helein F 2001 Constant Mean Curvature Surfaces, Harmonic Maps and Integrable Systems (Lectures in Mathematics) (Boston, MA: Birkhauser), doi:10.1007/978-3-0348-8330-6 [20] Konopelchenko B G, 1996 Induced surfaces and their integrable dynamics. Stud. Appl. Math. 96, 9–51, doi:10.1002/sapm19969619 [21] Levi D, Sym A and Tu GZ, A working algorithm to isolate integrable surfaces in E3, preprint DF INFN 761, Roma Oct. 10th, 1990, doi:10.1016/0375-9601(90)90897-W [22] Li YQ, Li B and Lou SY, Constraints for Evolution Equations with Some Special Forms of Lax Pairs and Distinguishing Lax Pairs by Available Constraints, arXiv:1008.1375v2. [23] Manton N and Sutcliffe P 2004 Topological Solitons (Cambridge Monographs on Mathematical Physics) (Cambridge: Cambridge University Press), doi:10.1017/CBO9780511617034 191 http://dx.doi.org/10.1017/CBO9780511535024 http://dx.doi.org/10.1007/978-3-663-14092-4_5 http://dx.doi.org/10.1063/1.529096 http://dx.doi.org/10.1063/1.532093 http://dx.doi.org/10.1088/1751-8113/40/42/S02 http://dx.doi.org/10.1007/PL00001392 http://dx.doi.org/10.1088/1751-8113/43/26/265206 http://dx.doi.org/10.1088/1751-8113/44/16/165203 http://dx.doi.org/10.1088/1751-8113/45/1/015204 http://dx.doi.org/10.1088/1742-6596/380/1/012023 http://dx.doi.org/10.1088/1751-8113/47/1/015201 http://dx.doi.org/10.1088/1751-8113/42/17/172001 http://arxiv.org/abs/1311.2406v1 http://dx.doi.org/10.1007/978-3-0348-8330-6 http://dx.doi.org/10.1002/sapm19969619 http://dx.doi.org/10.1016/0375-9601(90)90897-W http://arxiv.org/abs/1008.1375v2 http://dx.doi.org/10.1017/CBO9780511617034 A. M. Grundland, D. Levi, L. Martina Acta Polytechnica [24] Marvan M 2002 On the Horizontal Gauge Cohomology and Nonremovability of the Spectral Parameter, Acta Appl. Math. 72 51–65, doi:10.1023/A:1015218422059 [25] Marvan M 2004 Reducibility of zero curvature representations with application to recursion operators, Acta Appl. Math. 83 39–68, doi:10.1023/B:ACAP.0000035588.67805.0b [26] Marvan M 2010 On the spectral parameter problem, Acta Appl. Math. 109 239–255, doi:10.1007/s10440-009-9450-4 [27] Mikhailov AV 1986 Integrable magnetic models Soliton (Modern Problems in Condensed Matter vol 17) ed S E Trullinger et al (Amsterdam: North-Holland) pp 623-90 [28] Mikhailov A V, Shabat A B and Sokolov V V 1991 The Symmetry Approach to Classification of Integrable Equations, in Nonlinear Dynamics, Ed Zakharov V E, Springer, pp 115-184. [29] Olver P J 1993 Applications of Lie Groups to Differential Equations, 2nd edn. (New York, Springer). doi:10.1007/978-1-4612-4350-2 [30] Rogers C and Schief WK 2000 Backlund and Darboux Transformations. Geometry and Modern Applications in Soliton Theory (Cambridge: Cambridge University Press). doi:10.1017/CBO9780511606359 [31] Sakovich S Yu, True and fake Lax pairs: how to distinguish them, arXiv:nlin.SI/0112027. [32] Sakovich S Yu, Cyclic bases of zero-curvature representations: five illustrations to one concept, arXiv:nlin/0212019v1. [33] Sym A, 1982 Soliton surfaces. Lett. Nuovo Cimento 33, 394-400. [34] Sym A, 1995 Soliton surfaces and their applications (Soliton geometry from spectral problems) Geometric Aspect of the Einstein Equation and Integrable Systems (Lectures Notes in Physics vol 239) ed R Martini (Berlin: Springer) pp 154-231. doi:10.1007/3-540-16039-6_6 [35] Tafel J 1995 Surfaces in R3 with prescribed curvature J. Geom. Phys. 17 381-90. doi:10.1016/0393-0440(94)00054-9 [36] Zakharov V E and Mikhailov A V 1979 Relativistically invariant two-dimensional models of field theory which are integrable by means of the inverse scattering problem method Sov. Phys. – JETP 47 1017–49 [37] Zakrzewski W 1989 Low Dimensional Sigma Models (Bristol: Hilger). 192 http://dx.doi.org/10.1023/A:1015218422059 http://dx.doi.org/10.1023/B:ACAP.0000035588.67805.0b http://dx.doi.org/10.1007/s10440-009-9450-4 http://dx.doi.org/10.1007/978-1-4612-4350-2 http://dx.doi.org/10.1017/CBO9780511606359 http://arxiv.org/abs/nlin.SI/0112027 http://arxiv.org/abs/nlin/0212019v1 http://dx.doi.org/10.1007/3-540-16039-6_6 http://dx.doi.org/10.1016/0393-0440(94)00054-9 Acta Polytechnica 56(3):180–192, 2016 1 Introduction 2 Summary of results on the construction of soliton surfaces 2.1 Classical and generalized Lie symmetries 2.2 The immersion formulas for soliton surfaces 2.3 Application of the method 3 Mapping between the Sym-Tafel, the Cieslinski-Doliwa and the Fokas-Gel'fand immersion formulas 3.1 lambda-conformal symmetries and gauge transformations 3.2 Generalized symmetries and gauge transformations 3.3 The Sym-Tafel immersion formula versus the Fokas-Gel'fand immersion formula 4 The sigma model and soliton surfaces 4.1 Soliton surfaces associated with the CP1 sigma model 5 Concluding remarks Acknowledgements References