We consider several aspects of conjugating symmetry methods, including the method of invariants, with an asymptotic approach. In particular we consider how to extend to the stochastic setting several ideas which are well established in the deterministic one, such as conditional, partial and asymptotic symmetries. A number of explicit examples are presented.
Citation: Giuseppe Gaeta, Roma Kozlov, Francesco Spadaro. Asymptotic symmetry and asymptotic solutions to Ito stochastic differential equations[J]. Mathematics in Engineering, 2022, 4(5): 1-52. doi: 10.3934/mine.2022038
[1] | Juan-Carlos Felipe-Navarro, Tomás Sanz-Perela . Semilinear integro-differential equations, Ⅱ: one-dimensional and saddle-shaped solutions to the Allen-Cahn equation. Mathematics in Engineering, 2021, 3(5): 1-36. doi: 10.3934/mine.2021037 |
[2] | Filippo Gazzola, Gianmarco Sperone . Remarks on radial symmetry and monotonicity for solutions of semilinear higher order elliptic equations. Mathematics in Engineering, 2022, 4(5): 1-24. doi: 10.3934/mine.2022040 |
[3] | Diogo Gomes, Julian Gutierrez, Ricardo Ribeiro . A mean field game price model with noise. Mathematics in Engineering, 2021, 3(4): 1-14. doi: 10.3934/mine.2021028 |
[4] | Mouhamed Moustapha Fall, Veronica Felli, Alberto Ferrero, Alassane Niang . Asymptotic expansions and unique continuation at Dirichlet-Neumann boundary junctions for planar elliptic equations. Mathematics in Engineering, 2019, 1(1): 84-117. doi: 10.3934/Mine.2018.1.84 |
[5] | Dong-Ho Tsai, Xiao-Liu Wang . On an asymptotically log-periodic solution to the graphical curve shortening flow equation. Mathematics in Engineering, 2022, 4(3): 1-14. doi: 10.3934/mine.2022019 |
[6] | Jana Kopfová, Petra Nábělková . Thermoelastoplastic oscillator with Prandtl-Ishlinskii operator. Mathematics in Engineering, 2025, 7(3): 264-280. doi: 10.3934/mine.2025012 |
[7] | Luigi Montoro, Berardino Sciunzi . Qualitative properties of solutions to the Dirichlet problem for a Laplace equation involving the Hardy potential with possibly boundary singularity. Mathematics in Engineering, 2023, 5(1): 1-16. doi: 10.3934/mine.2023017 |
[8] | Giulia Bevilacqua . Symmetry break in the eight bubble compaction. Mathematics in Engineering, 2022, 4(2): 1-24. doi: 10.3934/mine.2022010 |
[9] | Alessandra De Luca, Veronica Felli . Unique continuation from the edge of a crack. Mathematics in Engineering, 2021, 3(3): 1-40. doi: 10.3934/mine.2021023 |
[10] | YanYan Li . Symmetry of hypersurfaces and the Hopf Lemma. Mathematics in Engineering, 2023, 5(5): 1-9. doi: 10.3934/mine.2023084 |
We consider several aspects of conjugating symmetry methods, including the method of invariants, with an asymptotic approach. In particular we consider how to extend to the stochastic setting several ideas which are well established in the deterministic one, such as conditional, partial and asymptotic symmetries. A number of explicit examples are presented.
The theory of symmetry – i.e., what is now known as Lie theory – was developed by Sophus Lie in his attempt to solve (nonlinear) differential equations, generalizing the approach by Evariste Galois for algebraic equations.
The Lie approach, and more generally symmetry methods for differential equations, produced a number of results by Lie and his pupils, but afterwards the theory remained at the same level for quite a long time, until it was taken up by Birkhoff in the USA and by Ovsjannikov in USSR in the sixties [1,2]. One of the reason for this "long sleep" lies in that the applications of the Lie theory do in general require quite extensive computations; these can nowadays be routinely performed on a computer by algebraic manipulation programs, but were quite prohibitive in earlier times.
After the reviving by Birkhoff and Ovsjannikov, the theory developed quickly, and underwent various generalizations; the reader can consult some of the by now standard texts [3,4,5,6,7,8,9] for the basics as well as for the most common extensions and generalizations.
Here we are concerned with two of these extensions and generalizations. That is, on the one side with the generalization to conditional [10], partial [11], and asymptotic [12] symmetries; and on the other side with the extension of the theory to the framework of stochastic differential equations [13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30] (by this we always mean equations of Ito or Stratonovich type [31,32,33,34,35,36,37], see below).
Our goal is to put the two together, i.e., to consider conditional and asymptotic symmetries for stochastic differential equations. Needless to say, our main interest is not in the abstract definition of these, but in showing that they can be of concrete use in determining (conditional or asymptotic) solutions to the concerned equations.
In the first part of the paper we establish some background. In particular, we quickly recall some of the main points of the theory of symmetry for (deterministic) differential equations, mainly to fix notation (Section 2), then recall the (well-known) notion of conditional symmetry (Section 3) and of (less well-known) asymptotic symmetry (Section 4) together with their use. We will also recall the basics of the theory of symmetry for stochastic differential equations (Section 5), and the method of invariants for these (Section 6).
We then proceed to merge the two, i.e., we develop conditional and asymptotic symmetries and invariants for stochastic differential equations; this will be the subject of Sections 7 and 8. We also provide, in Section 9, a number of concrete examples, chosen to be computationally simple in order to keep the conceptual issues in focus.
Our main result can be shortly stated saying that the theory of conditional, partial and asymptotic symmetries can be extended to the framework of stochastic differential equations. The reader will not find "big theorems" in our discussion, rather a number of lemmas; this corresponds to the fact we are not overcoming any special technical difficulty here. We are instead proposing a point of view that has not been considered previously in the literature, i.e., building a method which – we believe – is useful in characterizing solutions to stochastic equations via their asymptotic properties, in the same way as the corresponding deterministic approach proved to be precious in analyzing solutions to deterministic equations having certain asymptotic properties.
As we will discuss in more detail in the following (see Sections 3 and 4), the key observation triggering the present approach is that while in several physical cases one has equations which are invariant – say, for the sake of concreteness, under a rotation or scaling transformation – and correspondingly for these there is a special set of solutions displaying such symmetry, it is of more general interest the case where either (i) the solutions do have such symmetry only asymptotically in time and/or space (think e.g., to the case where the rotationally invariant solutions are stable, so that solutions which are not rotationally invariant at t=0 but which are sufficiently near to those will, in the long run, acquire the rotational symmetry); or (ii) the equations do not have such a symmetry, but there are solutions which, at least asymptotically in time and/or space, display such a symmetry.
The trivial, but not meaningless example, is that of the zero solution, which surely has a very high degree of symmetry, and which can be a valid – and maybe also stable or even attractive – solution for equations with no symmetry.
A less extreme, more concrete and also physically relevant, example is provided, in the deterministic framework, by the Boussinesq equation of Fluid Dynamics: in this case a number of physically relevant – and routinely observed in experiments – solutions do have symmetry properties which are not shared by the equation; and they can be analytically determined though the conditional symmetries approach. Similarly, stable solutions to several types of reaction-diffusion equations (including the prototypical Fisher-Kolmogorov equation) can be analytically determined through the method of asymptotic symmetries.
It must be stressed that in general, in this way one is able to obtain exact analytical expression for these symmetric class of solutions albeit there is no way to obtain a general solution to the equation under study in analytical terms. We will thus extend these approaches, developed in the framework of deterministic equations; and propose a general method to study special classes (characterized indeed by symmetry-like properties) of solutions to stochastic differential equations.
As already mentioned, we will illustrate the method by a number of concrete examples; these will however be chosen to be computationally simple in order to keep the conceptual issues in focus.
Throughout the paper, we will use the Einstein summation convention. We will denote by t the time variable and by xi the spatial ones, while the wk will be Wiener processes; we moreover use the shorthand notation
∂t := ∂∂t , ∂i := ∂∂xi , ˆ∂k := ∂∂wk , |
and denote the time derivative of a function x(t) as either xt or ˙x, depending on typographical convenience.
In this section we recall some basic notions concerning symmetries of deterministic differential equations. In particular we consider first order ODEs, i.e., dynamical systems (DS), as this is the deterministic analogue of Ito or Stratonovich equations to be considered in the following.
For a dynamical system
dxidt = fi(x,t) | (1) |
with t∈R, x∈M0⊆Rn, we define M=M0×R and consider maps Φ:M→M. In particular, we consider one-parameter groups of such maps, with generator
X = θ(x,t)∂t + φi(x,t)∂i . | (2) |
Note that with π0:M→R the projection π0(m0,r)=r, the full phase bundle M=M0×R has a natural structure of fiber bundle over the time real axis, (M,π0,R).
Under the action of X – or more precisely of the maps Φ=exp[εX] in the one-parameter group generated by X – we have, at first order in ε,
t→˜t = t + εθ(x,t) , xi→˜xi = xi + εφi(x,t) . | (3) |
Once we define the transformation of the independent (t) and the dependent (xi) variables, we have also implicitly defined the transformation of the derivatives of the latter with respect to the former. We have indeed (always working at first order in ε)
xit → ˜xit = xit + εψi(x,t,˙x) |
where ψi is given by the prolongation formula [3,4,5,6,7,8,9]
ψi(x,t,˙x) = Dtφi − xit Dtθ , | (4) |
and Dt is the total time derivative (in this case, this is just Dt=∂t+˙xj∂j).
With this, the vector field X on M is extended (prolonged) to a vector field X(1) in J1M (in which we consider local coordinates {t;x1,...,xn;˙x1,...,˙xn}),
X(1) = θ(x,t)∂∂t + φi(x,t)∂∂xi + ψi(x,t;˙x)∂∂˙xi . | (5) |
Here and below J1M is the (first) jet bundle associated to M [3,4,5,6,7,8,9]. Note this can be seen both as a bundle over M (the fiber corresponding to first derivatives ˙x) with projection ˜π1 and as a bundle over R (the fiber corresponding to variables x and first derivatives ˙x) with projection π1,
J1M˜π1⟶Mπ1↘↓π0M0 . |
The projections satisfy π1=π0∘˜π1, i.e. the diagram above is a commutative one.
The DS (1) defines a submanifold in J1M, also called the solution manifold Σf for the Eq (1).
A function x=ξ(t) is identified with its graph
γξ = {(x,t) : x=ξ(t)} ⊂ M , |
i.e., with a section γξ of the bundle (M,π0,R); this is naturally extended to a section γ(1)ξ in J1M. In local coordinates, we have
γ(1)ξ = {t;ξ1(t),...,ξn(t);˙ξ1(t),...,˙ξn(t)} . | (6) |
The function x=ξ(t) is a solution to (1) if and only if
γ(1)ξ ⊂ Σf . | (7) |
The vector field (2) is a Lie-point symmetry (generator) for the Eq (1) if and only if
X(1) : Σf → TΣf ; | (8) |
in this case it also maps (sections of (M,π0,R) representing) solutions to (1) into – generally, but not necessarily, different – (sections of (M,π0,R) representing) solutions to (1) [3,4,5,6,7,8,9].
Note that we will adhere to the usual abuse of notation consisting in calling "symmetries" the vector fields which are actually generators for the symmetry group.
It may be worth noting that the action of the vector field (2) on sections is given by
X : f(t) →˜f(t) = f(t) + ε [φ − θf′(t)] . | (9) |
To avoid any misunderstanding, we specify that this notation means that the section Γf given by Γf={(t,x=f(t))} is mapped into the section Γ˜f given by Γ˜f={(t,x=˜f(t))}.
Actually this is also an equivalent characterization of symmetries: that is, symmetries can be defined as maps M→M which transform solutions into (generally, but not necessarily, different) solutions [3,4,5,6,7,8,9].*
*For a very readable discussion of different – but equivalent – definitions of symmetries, see e.g., Chapter 3 in ref.[8] or Section 2.2. in ref.[6].
If x=ξ(t) is a solution to (1) and if X, satisfying (8), leaves γξ invariant, we say that x=ξ(t) is an X-invariant solution to (1).
Vector fields that are symmetries for a given equations, such as the evolution Eq (1), or of more general form, – or for equations, be these ODEs or PDEs, of more general form – are characterized as solutions to a set of linear PDEs for the coefficients of the vector field, known as determining equations [3,4,5,6,7,8,9]. For equations of the form (1) and vector fields expressed as in (2) with θ=0 (these are the simple vector fields to be considered below), these are written as
∂φi∂t + fj∂φi∂xj − φj∂fi∂xj = 0 . | (10) |
Remark 2.1. It should be mentioned that this sketchy discussion refers to proper Lie-point symmetries. In the case of deterministic Dynamical Systems, it turns out that so-called orbital symmetries are also quite useful [38,39,40]; as far as we know, these have not been investigated in the context of stochastic dynamical systems. ⊙
Remark 2.2. The prolongation formula for ODEs, Eq (4), is a special occurrence of the general prolongation formula, which – in the multi-index notation used in most of the literature, and in particular denoting by xi the independent variables, and by ua the dependent ones – reads
ψaJ,i = DiφaJ − uaJ,k Diξk . |
We refer to [3,4,5,6,7,8,9] for discussion and details. ⊙
The evolution Eq (1) defines a dynamics, and a dynamical vector field Z, in all of M; when f is time-autonomous, we actually also have a vector field Y in M0 and in this case Z=∂t+Y; symmetries X are then characterized by [X,Z]=0 (so called orbital symmetries are characterized by [X,Z]=λZ, for their properties and applications see e.g., [38,39,40]). We are specially interested in this latter case of f time-autonomous, both because the analysis is slightly simpler and due to its relevance in applications.
It may happen that there exist submanifolds of M being invariant under Z. In particular, for f time autonomous, there may be proper submanifolds U⊂M0 which are invariant under Y. The simplest example is that of fixed points for f, i.e., points x0 such that f(x0,t)=0. Other simple but significant examples are provided by periodic or multi-periodic solutions, by heteroclinic or homoclinic orbits, or by the stable and unstable manifolds of hyperbolic fixed points [41,42,43,44].
One would expect that symmetries are constrained by the need to somehow respect the structure of such invariant sets; this is indeed the case, provided one considers time-independent symmetries [4,45]. In particular, it can be shown that Lie-point time independent (LPTI) symmetries maps solutions into solutions whose trajectories in M0 have the same topology.
More specifically, we have the following results (these embody several Lemmas given in [4,45]; see there for proofs; the Corollary makes use of standard results in Topology, see e.g., [46]). We will denote by K(X)⊆M the subset of M which is point-wise invariant under a vector field X (or K(A)⊆M for the set invariant under an algebra A of vector fields), by Z the vector field defining the dynamics of the (autonomous) dynamical system under study, and by G the Lie algebra of symmetries for Z.
Lemma 2.1. The LPTI symmetry algebra of an autonomous ODE transforms stationary solutions into stationary solutions, periodic solutions into periodic solutions of the same period.
Lemma 2.2. Assume there is a compact set K(Z)⊂M invariant under the flow of Z. Then for any subalgebra G0⊆G, the submanifold K0(G0):=(K(G0)∩K(Z))⊆M is either empty or compact and invariant under the flow of Z.
Corollary. If K0(G0) has components isomorphic to an even-dimensional sphere S2n, then on each of these lie stationary solutions to the dynamical system ˙x=Z(x). If K0(G0) has components isomorphic to a disk D2n+1, then on each of these lie stationary solutions to ˙x=Z(x).
In several cases one meets dynamical systems or PDEs which cannot be solved exactly, but whose asymptotic behavior (e.g., for t→∞ and/or for |x|→∞ or for |x|→0) is amenable to exact treatment.
We just mention here two very simple examples of use in the following (these can be readily generalized).
Example 2.1. For x∈R, consider
˙x = f(x) = −∇Φ(x) |
with Φ(x)=Φ(−x) a convex even function, non-degenerate in x=0. Then for any initial datum x(0)=x0 the dynamics is attracted to x∗=0, and in the region near x∗ (and thus asymptotically in time) the dynamics is well described by the linear equation
˙x = −Ax := [(Df)(0)]x , A=−(Df)(0)>0 ; |
the solutions to this are of course
x(t) = x(t0) exp[−At] .♢ |
Example 2.2. In R2 and with polar coordinates (ρ,ϑ), consider the system
˙ρ=ρ (1 − ρ) ,˙ϑ=1 + (1 − ρ) h(ρ,ϑ) , |
with h(ρ,ϑ) an arbitrary smooth function. The system can not be solved in general, but we know that the dynamics is attracted to the circle ρ=1, and on this attracting manifold we have a trivial dynamics ˙ϑ=1. Thus any solution is asymptotically described by ρ=1, ϑ(t)=k0+t for some phase k0. ♢
As mentioned above, symmetries map (any) solution into a (generally different) solution. It may happen, however, that there are maps such that this property (i.e., mapping solutions into solutions) holds only for a subset of solutions.
The most common case is that of equations admitting solutions which are invariant under maps that are not a symmetry of the equation. For instance, any equation of the form (1) with f linear in x, f(x,t)=A(t)x, will admit the origin as a fixed point, and this in turn will be invariant under any linear (possibly time-dependent) map x→B(t)x. But this will be a symmetry of the equation if and only if
dB(t)dt = [A(t),B(t)] , |
which is definitely a non-generic property.
In this case, and in similar but less trivial ones, one speaks also of conditional symmetries [10]; see also [47,48,49,50,51,52,53,54]. Let us discuss these in some more detail (the reader can see the references given above for a more complete discussion), referring to the case of PDEs for greater generality; we denote by x1,..,xq the independent variables, and by u1,...,up the dependent ones. Our goal is to determine X-invariant solutions for a differential equation (note that by this we will always mean possibly a vector one, i.e., a system of scalar differential equations) Δ=0, where
X = ξi(x,u) ∂∂xi + φa(x,u) ∂∂ua = ξi∂i + φa∂a |
is not necessarily a symmetry of Δ; at the same time, we want to determine the X which admit such solutions.
In order to do this, we complement Δ with an equation ΔX=0 expressing the requirement of X-invariance of the function u=u(x). This is just, see (9) above,
φa(x,u) − ξi(x,u)uai = 0 . | (11) |
We are thus led to consider the system
EX = {Δ = 0ΔX = 0 | (12) |
It is clear that the solutions of this, if they exist, are all and only the solutions to Δ=0 which are also X-invariant (i.e., which also solve ΔX=0). Thus symmetries for the system EX, which transform solutions into solutions, will transform X-invariant solutions to Δ into X-invariant solutions to Δ. In particular this apply for the vector field X itself, which by construction will leave each of such solutions invariant.
We are thus at first sight reduced to a standard problem in symmetry analysis; however we should look at this with some more care. One possibility is that we fix in advance X and look for solutions to Δ which are X-invariant. In this case ΔX reduces to an ansatz on the functional form of u(x) which can then be inserted into Δ, and indeed nothing new is here.
On the other hand, the interesting case is the one where we do not know in advance what X should or could be, and we study the system (12) in order to determine X and the associated invariant solutions, if any.
In this case we get something new: in fact, the usual determining Eq (10) – or more generally their version for PDEs – are linear in the X coefficients {ξi(x,u),φa(x,u)}, while now we will have determining equations which are nonlinear (in particular, quadratic) in these.
We will not enter into the details of how such a system of nonlinear (determining) equations can be solved thanks to the peculiar structure of the problem, referring the reader to the original paper by Levi and Winternitz for this matter [10].
We instead stress a most relevant point: symmetries have to map solutions into solutions, but when we apply symmetry analysis to the system (12), this applies to the common solutions to Δ=0 and to ΔX=0, i.e., only to invariant solutions. Thus these are symmetries of Δ conditional upon the additional invariance condition ΔX.
Remark 3.1. We should mention that one could also investigate the inverse problem for conditional symmetries: that is, for a given number of dependent and independent variables and a given order of the equation, and given a certain vector field X, determine all the differential equations Δ which admit X as a conditional symmetry. This was considered by Levi, Rodriguez and Thomova in [55], and more recently by Pucci and Saccomandi [56,57]. We are not able to provide stochastic extensions of this analysis at the moment. ⊙
A slightly more general case is also possible, i.e., that we have a subset of solutions which are not individually invariant under the map, but which are indeed transformed one into another by the map – which again is not a symmetry of the original equation Δ=0. In this case one speaks of partial symmetries [11] (see also [58]).
A specially relevant case in our dynamical systems context is where the solutions in this set are characterized by living on a manifold which is dynamically invariant (i.e., invariant under the dynamics). The situation is characterized in a way which looks quite similar to the characterization provided above for conditional symmetries. That is, we consider a differential equation (again, in general this can be a system of PDEs) Δ=Δ0, and a vector field X which is not a symmetry for Δ0, thus such that – writing S0:=SΔ0 for ease of writing,
[X(n)Δ0]S0≠ 0 . |
We consider then an auxiliary equation
Δ1 := [X(n)(Δ0)] . |
If we look at solutions to the system
E = {Δ0 = 0Δ1 = 0 |
which are globally invariant under X (that is, the set of solutions is mapped into itself, albeit the single solutions are possibly not invariant), this set is characterized precisely by the property
[X(∗)(Δ1)]S(1) = 0 ; | (13) |
here S(1)=S0∩S1, with of course S1=SΔ1, is the set of solutions to the system E; and we wrote X(∗) to mean the prolongation of X of appropriate order, i.e., the same order as the equation Δ1. In this case, i.e., if (13) is satisfied, we say that X(∗) is a partial symmetry of order one for Δ, as it maps a subset of solutions into solutions (belonging to the same subset).
We stress that (13) may or may not be satisfied, depending on Δ and on our choice of the vector field X. If it is not satisfied, we can iterate our procedure and consider
Δ2 := X(∗)(Δ1) |
and look for S(2)=S0∩S1∩S2, enquiring if – with an obvious notation – the equation
[X(∗)(Δ2)]S(2) = 0 | (14) |
is satisfied. If this is the case, we say that X is a partial symmetry of order two for Δ; otherwise we can still iterate our procedure and so on.
We set this in the form of a Proposition, which is quoted from [11].
Proposition 3.1. Given the general differential problem Δ=Δ0=0 and a vector field X on M, define Δr+1:=X(∗)(Δr). Denote by S(r) the set of simultaneous solutions of the system
Δ0 = Δ1 = ... = Δr = 0 , |
and assume that this is not empty for r≤s. Assume moreover that
[X(∗)(Δr)]S(r) ≠ 0 for r=0,1,...,n−1 , [X(∗)(Δn)]S(n) = 0 . |
Then the set S(n) provides a family of solutions to the initial problem Δ which is mapped into itself by the transformations generated by X.
Remark 3.2. When the situation depicted in Proposition 3.1 is met, we shall say that X is a partial symmetry, or P-symmetry for short, for the problem Δ, and that the globally invariant subset of solutions S(n) obtained in this way is a X-symmetric set. We also refer to the number n appearing in the statement as the order of the P-symmetry. ⊙
The discussion can be specialized to the case of dynamical system; here we just recall a result for this case, quoting again from [11], to which the reader is referred for further detail. Here L(λ)(f) denotes the linearization of f at x(λ) (see below), i.e.,
L(λ)(f) = [∇(f)]x(λ) . |
Proposition 3.2. Assume that the dynamical system ˙x=f(x) admits a LPTI partial symmetry X=φi(x)∂i, and let x(λ)=x(λ)(t) be an orbit of solutions obtained under the action of the group generated by X, with λ the group parameter. Then Φ:=φ(x(λ)) satisfies the equations
dΦdt = L(λ)(f)⋅Φ ; Φ = dx(λ)dλ . |
We give here two Examples (examples 3.1 and 3.2) of conditional symmetries, and three Examples (examples 3.3 through 3.5) dealing with partial symmetries. These are taken respectively from [10] and from [11]. Some Examples deal with PDEs, as the essential of conditional and partial symmetries is better grasped in this context.
Example 3.1. The Boussinesq equation [59] in one spatial dimension, for a function u=u(x,t), reads
utt + uuxx + u2x + uxxxx = 0 . | (15) |
If we consider the vector field
X1 = ∂t + t∂x − 2t∂u , |
this is not a symmetry for the equation, but it turns out to be a conditional symmetry for it.
Under the action of X1 we have a one-parameter flow (group parameter λ) which acts as
{t → ˜t = t+λx → ˜x = x+λt+(1/2)λ2u → ˜u = u−2tλ−λ2 |
Assuming that u=u(x,t) is a solution to (15), we have that ˜u=˜u(˜x,˜t) is a solution if it satisfies the auxiliary condition (we omit the tildes from now on)
Δ1 := (ut + tux)x = 0 . | (16) |
Now we note that the invariance condition ΔX=0, see (11), reads in this case (i.e., for the X we are now considering)
ut + tux + 2t = 0 . |
Thus (16) is just a differential consequence of this, (Δ1)x=0. ♢
Example 3.2. Consider again the Boussinesq Eq (15), and now
X2 = ∂t − (x/t)∂x + [(2/t)u + (6/t3)x2]∂u ; |
again this is not a symmetry for Eq (15). Now we get
˜u(˜x,˜t) = [t+λt]2 u(x,t) + [(t+λ)2t4 − t2(t+λ)4] x2 . |
On the other hand, the invariance condition (11) is now
Δ2 := [2tu + 6t3x2] − ut + xtux = 0 . | (17) |
If u(x,t) is a solution to (15), then the condition for ˜u(˜x,˜t) to be also a solution reads
A1(x,t;λ) F + A2(x,t;λ) Ft + A3(x,t;λ) Fx = 0 , |
where Ai are certain smooth polynomial functions, and
F := ut − (x/t)ux − [(2/t)u + 6(x2/t3)] . |
Thus if u(x,t) is a solution to the Boussinesq equation and F=0, the transformed function ˜u(˜x,˜t) is also a solution. But looking at (17) we see that F=0 is just the invariance condition under X2. ♢
Example 3.3. Consider the modified Laplace equation
Δ := uxx + uyy + g(u)uxxx = 0 , | (18) |
with g(u) an arbitrary smooth function. We also consider the rotation vector field
X = y∂x − x∂y ; | (19) |
this is a symmetry for Δ in the case g(u)=0, but not in general. We will assume g(u)≠0.
Implementing our procedure, we get
Δ1:= uxxy ,Δ2:= 2uxxy − uxxx ,Δ3:= 2uyyy − 7uxxy ,Δ4:= 3uxyy ,Δ5:= uyyy − 2uxxy . |
The last equation is identically satisfied on common solutions to the previous ones, thus we have a partial symmetry of order five.
The set S(5) has the following form, with k1,…,k6 arbitrary constants:
S(5) = {u(x,y) = k1x2 + k2y2 + k3xy + k4x + k5y + k6} ; |
that is, functions which are at most quadratic in the (x,y) variables. Note that for these functions the Eq (18) reduces to the standard Laplace equation, and for this it is obvious that (19) is a symmetry; thus this seemingly complicated computation yields a rather obvious result. ♢
Example 3.4. We look again at the Boussinesq Eq (15), which will be our Δ0, and consider the vector field
X = ∂t + t∂x − 2t∂u . | (20) |
Now we have (disregarding some inessential nonzero multiplicative constants)
Δ1=X(∗)Δ0 = uxt + tuxx ,Δ2=X(∗)Δ1 = 0 . |
We have thus to look for simultaneous solutions to Δ0=0 and to Δ1=0. Solving for Δ1=0 yields
u(x,t) = v(x−t2/2) + w(t) − t2, | (21) |
with v and w arbitrary functions. Note that the functions (21) are X-invariant if and only if w′(t)=0 (this is why the term −t2 has not been absorbed into the arbitrary function w(t) in writing (21)); thus we are considering a setting which is really a generalization of the conditional symmetries approach, even in this case.
Inserting now (21) into Δ0, i.e. into the Boussinesq equation (15), and writing z=x−t2/2, we get that v and w must satisfy
ddz(v‴ + vv′ − v − 2z) + wv″ + d2wdt2 = 0 . | (22) |
We are specially interested, as mentioned above, in the case w(t)≠0 (otherwise we recover the Levi-Winternitz family of X-invariant solutions). E.g., for w(t)=K a nonzero constant, we get solutions u(x,t)=K+v(z)−t2 with v a solution to
v‴ + vv′ − v + Kv′ = c + 2z .♢ |
Example 3.5. We now consider an example involving a Dynamical System in R3 with Cartesian coordinates (x,y,z); we will also write ρ2=x2+y2. Consider
{˙x = x(1−ρ2) − y + zg1(x,y,z)˙y = y(1−ρ2) + x + zg2(x,y,z)˙z = zg3(x,y,z) |
with gi smooth functions. If we also consider the rotation vector field
X = y∂x − x∂y , |
then the partial symmetry condition reads
z Gα(x,y,z) = 0 (α=1,2,3) |
with Gα certain functions which are non-zero for generic gi: e.g.,
G1 = g2 − y(∂g1/∂x) + x(∂g1/∂x) . |
Thus the system exhibits – as rather obvious from its expression – rotational (x,y) symmetry only if restricted to the plane z=0. Moreover, in this plane one can find three different families of solutions (beside the null one, which is rotationally invariant) which are mapped into themselves by the rotations, without being rotationally invariant themselves: the trajectories lying in ρ<1, and respectively in ρ>1, spiralling towards the limit cycle ρ=1; and the solutions running on the single trajectory ρ=1 which is left fixed by the partial symmetry (i.e., the limit cycle). ♢
In many cases, one has equations whose solutions have a complicated behavior, which becomes simple in some asymptotic regime. This may be asymptotic in time (thus for t→±∞, possibly with different behavior in the two limits or with a simple behavior only in one of the limits) or asymptotically in space (i.e., for |x|→∞, or in the one dimensional case for x→±∞, with the same notes as for t); or in a suitable combination of the two.
It is remarkable that in such cases one can adopt a renormalization-group (RG) like approach [60,61]; this in turn can be combined with symmetry considerations [12,62,63]. In particular, we can have asymptotic symmetries, i.e., maps which are symmetries (hence in particular map solutions into solutions) only in this asymptotic regime, being otherwise (that is, in the non-asymptotic regime) only approximate symmetries, at most.
In order to discuss asymptotic symmetry properties (at least in the deterministic case) we consider the general case, i.e., possibly PDEs (of any order n); the independent variables live on a manifold B, the dependent ones take values in a manifold U, and we consider M=U×B. We will follow the discussion given in [4], referring to [12,62,63] for further detail.
It is then convenient to use the geometrical representation of functions as cross sections of the bundle (M,π0,B), and of equations as submanifolds in JnM; in the case of evolution equations, these can also be seen as sections of a suitable fiber bundle (with time derivatives in the fiber, i.e., in the vertical space), and on the other hand some care is needed in order to take into account the special role of the time variable in this case.
Here and in the following we write ∂Ja=(∂/∂uaJ); moreover ψa0=φa, and u[n] will denote (u,u(1),...,u(n)). Recall also that |J| is the order of the multi-index J.
A vector field
Y = ξi(x,u)∂i + φa(x,u)∂a | (23) |
in M, and its prolongation
Y(n) = ξi(x,u)∂i + ψaJ(x,u[|J|])∂Ja | (24) |
to JnM, will then induce an action on the space of sections in (M,π0,B), hence on functions; and on submanifolds in JnM, hence on equations.
In particular, a (vector) function ua=fa(x) is then mapped by the infinitesimal action of the vector field (with parameter ε) into
˜ua(x)=(eεYf)a(x)=ua(x) + ε [φa(x,f(x)) − ξi(x,f(x))∂ifa(x)] + O(ε2) ; | (25) |
this infinitesimal action defines a flow (with parameter λ, not necessarily small) γ(λ)=Φ(λ,γ0) in the set Γ(M) of sections in (M,π0,B), where λ is the flow parameter and γ0 is the initial point for the flow. From this one also computes the transformation rules for derivatives uaJ, i.e., on sections of (JnM,πn,B), and hence for equations; this also defines a flow γ(n)(λ)=Φ(n)(λ,γ(n)0) in the set Γ(n)(M) of sections in (JnM,πn,B), where λ is the flow parameter and γ(n)0 is the initial point for the flow.
We will also write these flow as
fλ=Φ(λ,f0) , Δλ=Φ(n)(λ,Δ0) , | (26) |
for ease of writing.
We say that a function u=f(x) is symmetric under Y if and only if the corresponding section γf∈Γ(M) is invariant under (the flow Φ in Γ induced by) Y; and similarly that the equation Δ in JnM is symmetric under Y if and only if the corresponding submanifold SΔ is invariant under (the flow Φ(n) in JnM induced by) the prolongation Y(n).
We are specially interested in the case where the flow drives functions (or equations) which are not invariant to a fixed point, i.e., to a limit function (or an equation) which is invariant.
In particular, consider the case where the limit
limλ→∞Φ(λ,γf) = γ∞f | (27) |
exists. Then we say that Y is an ω-asymptotic symmetry for (the section γf and hence) the function f:B→U. (In the case where the limit limλ→−∞Φ(λ,γf) = γ−∞f exists, we say that Y is an α-asymptotic symmetry for f. This notation recalls the notions of ω-limit and of α-limit in use in Dynamical Systems theory [41,42,43,44].)
Similarly, in the case where the limit
limλ→∞Φ(n)(λ,SΔ) = S∞Δ | (28) |
exists, we say that Y is an ω-asymptotic symmetry for (the submanifold SΔ and hence) the equation Δ=0; and similarly considering the limit for λ→−∞.
We have the following results, which we quote from [4] (Lemma IX.1 therein):
Lemma 4.1. Let u=f0(x) be a solution to the differential equation Δ0; and let the limits f±∞, Δ±∞ for the flows
fλ = Φ(λ,f0) , Δλ = Φ(n)(λ,Δ0) |
exist. Then f±∞ is a solution to Δ±∞.
Corollary. If Δ0→Δ∗ for λ→±∞, then under the flow Φ(λ,f) all solutions f0 to Δ0 go into solutions (not necessarily invariant) to Δ∗ for λ→±∞.
Remark 4.2. We stress that here "asymptotic" refers to the flow parameter λ, i.e., to the flow Φ and Φ(n) associated to Y. This may or may not correspond to asymptotic properties in time and/or space depending on the form of the vector field Y we are considering. ⊙
Example 4.1. Consider the DS in R2
{˙x = x − Ay − (x2+y2)x,˙y = y + Ax − (x2+y2)y; | (29) |
this is symmetric under both time translations and spatial rotations. Thus if we consider the two-parameters family of vector fields
X = α∂t + β (x∂y−y∂x) , | (30) |
all of these are symmetries for our DS. Note that for α>0 the action of X gives translations forward in time, while for α<0 we have translations backward in time; similarly, the sign of β controls the direction of rotation in the (x,y) plane. We assume α>0 for definiteness and ease of discussion.
The action of X on the space Γ={xλ(t),yλ(t)} of sections of (R2×R,π0,R) – which we denote by ˜X – is described by
dxλdλ=−α [1 − (x2+y2)]xλ − (β − αA)yλ ,dyλdλ=−α [1 − (x2+y2)]yλ + (β − αA)xλ . |
Thus there is no fixed point for the action of ˜X unless we choose β=αA, i.e.,
X = X0 := α[∂t + A (x∂y−y∂x)] . |
With this choice instead, we get
dxλdλ=−α [1 − (x2+y2)]xλ ,dyλdλ=−α [1 − (x2+y2)]yλ ; |
thus all solutions admit the vector field X0 as an α-asymptotic symmetry, and the vector field X∗0=−X0 as an ω-asymptotic one (their role is reversed for α<0 in the definition of X0). Note that the solutions with x20+y20 equal to either zero or one also admit X0 as a full symmetry. ♢
Example 4.2. Consider the FKPP (Fisher-Kolmogorov-Petrovskii-Piskunov) equation [64,65]
ut = A uxx + εu(1−u) | (31) |
for u=u(x,t), with A and ε positive real constants; in biological applications, it is required that u(x,t)≥0 for all x and t. This obviously has some stationary homogeneous (i.e., ux=0) states u(x,t)=0 and u(x,t)=1, with the former being unstable and the latter stable against small perturbations. It is known that if the initial datum is suitably concentrated, in particular if it decays exponentially fast for |x|→∞, then for t→∞ and x→∞ the solution is in the form of a traveling front
u(x,t) ≃ exp[−x − vtλ] ; λ=√A/ε , v = √4Aε . |
It is convenient to rescale coordinates via
t→˜t = εt , x→˜x = √ε/A x . |
Using these coordinates – and omitting the tildes for ease of writing – the equation reads
ut = uxx + u (1−u) , | (32) |
and the asymptotic solution is
u(x,t) ≃ f0(x,t) = A exp[−(x−2t)] . |
As already mentioned, the asymptotic above describes the behavior of the solution for large t and x, which also means for small u. But for small u the equation is approximated by its linearization around u=0, i.e., simply by
ut = uxx + u ; | (33) |
the ansatz u(x,t)=w(z)=w(x−2t) transforms this into
w″ + 2w′ + w = 0 , | (34) |
with solutions
w(z) = c+e−z + c−ez . |
We denote by W the space of such solutions; the constants (c+,c−) provide coordinates in this space. The requirement that w(z)→0 for z→∞ (i.e. for x→∞ at any finite t) yields c−=0 and hence w(x)=c+e−z, which is just the f0(x,t) given above.
The symmetry algebra of the linearized Eq (33) is generated by
X0 = u∂u , X1 = ∂x , X2 = ∂t . |
We also consider
X± := X1 ∓ 12 X2 ; |
note that X+=∂z, X−(z)=0. Moreover, [X0,X±]=0.
The quotient (i.e., symmetry-reduced) Eq (34) admits X0 (now written as X0=w∂w) as a scaling symmetry, and is also invariant under X+. It is easily checked that the propagating front solutions correspond to a subspace of W invariant under the action of the group generated by the vector fields X0 and X+.
As for asymptotic symmetries, the situation is less simple; one can show the following, which we quote from [63]: if we consider a scaling vector field X such that the limit for λ→∞ of exp[λX] extracts the behavior for large x and t, and Δ0 is the FKPP equation, we set Δλ:=eλXΔ0. Then
limλ→∞Δλ = Δ∗ |
is the heat equation ut=uxx.
In fact, consider the most general scaling vector field,
Xs = ax∂x + bt∂t + cu∂u ; |
one of the constants a,b,c can always be set to unity (provided it is nonzero), which amounts to a redefinition of the group parameter. Then choosing Δ0 the FKPP equation in its standardized form (32) we get for Δλ=exp[Xsλ]Δ0 the expression
Δλ = λc−b [ut − λb−2auxx − λbu + λb+cu2] . |
In order to have a meaningful (and nonzero) limit for λ→∞ we set b=c, and we will choose b=c<0; in order to have a nontrivial limit for λ→∞ we also set a=b/2. Moreover, we can reparametrize the flow setting |b|=1, i.e., b=−1. This yields
Δλ = ut − uxx − (u/λ) + (u2/λ2) , |
which reduces to the heat equation for λ→∞. ♢
We have so far discussed symmetries of deterministic differential equations. The symmetry approach has also been applied to stochastic differential equations [13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30].
Here we will be specially concerned with equations (as in the case of deterministic equation, here an equation might be of vector type, i.e., a system of scalar equations) of Ito type, i.e., in the form
dxi = fi(x,t)dt + σi k(x,t)dwk , | (35) |
wherefi and σi k are smooth functions. We will also consider Stratonovich type equations
dxi = bi(x,t)dt + σi k(x,t)∘dwk . | (36) |
As it is well known, to each Ito equation there is an associated Stratonovich equation (and conversely) which carries the same statistical information (the correspondence between the two presents some subtleties, see e.g., [34,37] for a discussion of these), in particular Eqs (35) and (36) are in such a relation if their drifts satisfy
fi(x,t) = bi(x,t) + 12 (∂kσij)σk j . | (37) |
We will consider in particular the approach which parallels the usual treatment of deterministic equations in the stochastic case (to which we gave several contributions in recent years); for a discussion of other approaches, including earlier attempts, see e.g., the review paper [22]. See also [66,67] for an approach relating symmetries to Girsanov theory, and more generally [68,69,70,71,72] for a related concurrent approach.
While in the case of deterministic differential equations there exists a well developed geometric theory, which allows to set the symmetry theory on firm geometrical basis, this does not hold for stochastic equations. Actually, the main characteristic of Ito differentials is that under diffeomorphisms they do not transform by the familiar chain rule, but according to Ito rule. This shows that the lack of a geometrical setting for symmetries of Ito equations is not due to insufficient development of the theory, but to its intrinsic features. The situation is slightly different for Stratonovich equations. In this case the chain rule is preserved, so that a geometrical setting would be possible; but, as it is well known, the Stratonovich formulation presents several delicate points from the point of view of stochastic processes (many of them being actually associated to its more relevant advantage from the point of view of a physicist, i.e., the preservation of the time inversion symmetry), so that a formulation of the basic theory in terms of the Ito theory is preferable. The relation between symmetry theory for Ito and Stratonovich equations is thoroughly discussed e.g., in [27]; see also [19] for earlier results.
Let us consider an Ito Eq (35); this involves the time variable t, the driving Wiener processes wk(t), and the spatial variables xi, which evolve through a stochastic process depending on the realization of the wk(t) and described by the Ito equation itself.
Thus our problem lives in the space E of the variables {t;x;w} (note that x and w are vector variables); a blind application of the Lie approach would consider general vector fields in this space, i.e., vector fields of the form
X = τ(x,t,w)∂∂t + φi(x,t,w)∂∂xi + hk(x,t,w)∂∂wk . | (38) |
It is clear that the action of such a vector field mixes all the variables, discarding the difference between driving and driven stochastic processes, and, even more, introducing a stochastic time. This kind of transformation is well known and widely used by probabilists, and also in the study of stochastic differential equations (including e.g., in the study of normal forms for SDEs, which has close relations with symmetry considerations [73]); but here we are interested in transformations which leave the equation invariant. Transformations which even change the nature of the variables are surely not leaving the equation invariant, and hence will not be considered.
If we want to leave the equation invariant, a necessary (but by no means sufficient) condition is that the nature of the involved variables is not changed. We should also ask the transformed wk variables to be still independent Wiener processes. These requirements (see [30] for how the latter one restricts the form of the hk functions) lead to consider vector fields in E of the form
X = τ(t)∂∂t + φi(x,t,w)∂∂xi + (Rk ℓwℓ)∂∂wk , | (39) |
with R a matrix in the Lie algebra of the conformal linear group CL(n). These vector fields are called admissible in E, and in the following we only consider such fields.
Remark 5.1. We recall CL(n)≈R+×O(n), where O(n) is the (n-dimensional) orthogonal group, and R+ represents here the dilation group. An alternative characterization of matrices A∈CL(n) is simply that they satisfy A+A=λ2I for some λ∈R. We are interested in matrices R in the Lie algebra of CL(n); in practice, such an R is the sum of a diagonal matrix which is a multiple of the identity (corresponding to the linear space spanned by the generator of the dilation group), and a skew-symmetric one (these correspond to the linear space spanned by the generators of the orthogonal part of the group). Thus R has 1+n(n−1)/2 independent elements. ⊙
It is then straightforward to consider the change of variables induced by an infinitesimal action of X, i.e.,
t→t+ετ , xi→xi+εφi , wk→wk+εRk ℓwℓ , |
and hence on the Eq (35). Our restrictions on the form of the vector field guarantees that the transformed equation is still an Ito one, which can hence be written as (recall the variables t,x,w are now the transformed ones)
dxi = ^fi(x,t)dt + ˆσi k(x,t)dwk , | (40) |
where the functions ^fi and ˆσi k – which can be readily computed in explicit terms – are in general different from the original ones, fi and σi k.
Requiring that the equation is left invariant under the action of X means simply requiring that ^fi(x,t)=fi(x,t) and that ˆσi k(x,t)=σi k(x,t). We have thus n+n2 equations involving the unknown functions τ,φi,σi k, which determine the vector fields which leave the Eq (35) invariant, i.e., identify its (Lie-point) symmetries; these are thus called the determining equations.
An important point is that the form (39) for symmetry vector fields is actually still too general: albeit all vector fields of this form are admissible, not all of them are useful when it comes to applying symmetry theory. In particular, as in the deterministic case, the main purpose of computing symmetries is to then apply them for reducing – or even solving – the stochastic equations. But to this purpose, only so called simple symmetries are useful under the present theory [16,17,18,28]. Simple symmetries are those not acting (but generally depending) on the time variable; i.e., those, in the notation of (39), with τ=0). Thus admissible simple symmetries have generators of the form
X = φi(x,t,w)∂∂xi + (Rk ℓwℓ)∂∂wk , | (41) |
with R as discussed above.
We stress that this restriction to simple (admissible) symmetries will be given for granted in the following; it will also simplify a number of matters.
Some further nomenclature within simple symmetries may also be useful: when R=0 and the φi do not depend on the w variables, we have deterministic symmetries; when R=0 and (at least some of) the φi do actually depend on the w variables, we have random symmetries; and finally when R≠0 (and hence the X has a component along the w variables) we have W-symmetries.
Remark 5.2. There is a substantial difference in computing the action of a vector field X on a stochastic equation with respect to the case of a standard (i.e., deterministic) ordinary differential equation. In the latter case the ODE defines a vector field Z in a suitable jet bundle J, and we compute how the vector field X and its prolongations act on the vector field Z; in the case of stochastic Ito equations, we have to compute how X acts on the functions fi(x,t), σi k(x,t) and on the differentials dxi, dt, dwk. Thus, roughly speaking, we are dealing with one-forms rather than with vector fields. This might be compared with recent work by Anco and Wang [74] taking the same point of view for deterministic ODEs. ⊙
When we write down the explicit expressions for ^fi and ˆσi k, we also obtain the explicit expressions for the determining equations. We set this in the form of a Proposition; this is quoted from [30], and we refer to there for a proof.
Here and below the symbol Δ denotes the Ito Laplacian,
ΔF := ∑k,ℓδkℓ∂2F∂wk∂wℓ + ∑i,j(σσT)ij∂2F∂xi∂xj + 2∑j,kσjk∂2F∂xj∂wk . | (42) |
Proposition 5.1. The determining equations for simple symmetries of the Ito Eq (35) are
∂φi∂t + fj∂φi∂xj − φj∂fi∂xj + 12Δφi=0 , | (43) |
∂φi∂wk + σj k∂φi∂xj − φj∂σi k∂xj − σi mRm k=0 . | (44) |
Obviously, several special cases are obtained by considering special forms of the functions τ, φi and R. We will not enter into this discussion, and just refer the reader to [22,30].
We will however stress that solutions with τ=0 (i.e., such that the time variable is not acted upon by the vector field X) are known as simple symmetries, and are specially useful in that the symmetry reduction based on these is specially simple.
Remark 5.3. The determining Eqs (43), (44) are a set of n+n2 equations for the n functions {φ1,...,φn} and the n×n matrix R, having (as noted above) n(n−1)/2+1 independent elements. In order to concretely use the information about the symmetry, it does not suffice to prove that a solution exists, but we need of course to determine it in an explicit way. ⊙
As it is well known – and recalled above – the Ito Eq (35) is associated with the Stratonovich Eq (36), the coefficients of the two being related by (37). With this condition, (36) carries the same statistical information as (35); see e.g., [37] for more detail. From the point of view of symmetry, there is a substantial difference: while the Ito equations transforms, under the action of a diffeomorphism and thus in particular under the flow generated by a vector field, as described by the Ito rule, a Stratonovich equation transforms under the usual chain rule and thus has a geometrical meaning.
It is quite natural to wonder if symmetries of an Ito equation and of the associated Stratonovich equation are related or coincide. The question was first tackled by Unal [15], and then fully solved by Kozlov [27], who showed that deterministic and random symmetries of an Ito and of the associated Stratonovich stochastic equations are just the same – as indeed natural given the equivalence of the two formulations.
The situation is slightly different for W-symmetries; the relations of these for an Ito and the associated Stratonovich equation are discussed in [30], see Section VII (and in particular Section VII.C) therein.
One could of course also analyze symmetries of a Stratonovich Eq (36) along the lines of Section 5.1, with the same limitation on the form if the admitted vector fields; in this way one would obtain the determining equations for (Lie-point) symmetries of a Stratonovich equation. See again [30] for a proof to the following statement.
Proposition 5.2. The determining equations for simple symmetries of the Stratonovich Eq (36) are
∂φi∂t + bj∂φi∂xj − φj∂bi∂xj=0 , | (45) |
∂φi∂wk + σj k∂φi∂xj − φj∂σi k∂xj=σi mRm k . | (46) |
Note that (46) coincides with (44); as for (45), this coincides with (43) when we impose the correspondence (37). See also [24,25,26,27] for further detail.
In the view of Sophus Lie, the symmetry analysis of differential equations was not a mathematical curiosity, but a concrete method to obtain order/dimensional reduction, and when possible exact solutions, of the differential equations.
The same applies for symmetry analysis of stochastic equations: knowledge of their symmetries allows to obtain a reduction, and when we have enough symmetries (e.g., one symmetry for a scalar equation) there is a concrete method to obtain a complete solution of the equation.
Remark 5.4. It is maybe worth specifying what we mean by this, i.e., by a complete solution of an Ito equation. We mean an explicit correspondence between a concrete realization of the driving (n-dimensional) Wiener process and the corresponding realization of the (n-dimensional) stochastic process x(t). Example of this correspondence are provided below in this subsection. ⊙
As for deterministic equations, the key to integration of a symmetric equation lies in passing to symmetry-adapted variables (an alternative method uses invariants, and will be discussed in the next Section). In these, the equation takes a specially simple form and is integrated in an elementary way; that is, the difficulty of integrating the equation is moved to the difficulty of determining the symmetry and the symmetry-adapted variables.
In fact, once we have determined (simple) symmetries of a given Ito equation, these can be used pretty much in the same way as in the case of deterministic equations to integrate or reduce the equation at hand.
Proposition 5.3. Let the Eq (35) for x∈M≈Rn admit a r-dimensional solvable algebra G of simple deterministic Lie symmetries. Then, passing to G-adapted variables, the equation can be reduced to a stochastic equation for y∈M/G≈Rn−r plus a set of r reconstruction equations; the latter amount to simple Ito integrations. In the case r=n, the equation is explicitly integrated.
Proof. This is given and proved e.g., in [25]; it is however worth sketching the proof. The first step, i.e., the determination of (simple) Lie symmetries, has already been mapped in the previous part of this section to the problem of solving the determining equations. Assume that a symmetry vector field X – written in the form (41) – has been found. We want then to pass from (x,t;w) to new variables (y,θ;z), where the zk are still independent Wiener processes, such that in the new variables X reads simply as X=∂/∂yn. In fact, if this is the case the Ito equation will be written, in the new variables, as
dyi = Fi(y,θ)dθ + Si k(y,θ)dzk | (47) |
with both the Fi and the Si k being independent of yn. The equation decouples then into an (n−1)-dimensional Ito equation plus a "reconstruction equation", which is written in terms of an Ito integral as
yn(t) = yn(0) + ∫t0Fn[y(θ),θ]dθ + ∫t0Sn k[y(θ),θ]dzk . | (48) |
In the case of a scalar equation, this gives directly a full solution of the equation.
When we have a higher dimensional equations and several symmetries spanning a solvable Lie algebra, we can proceed step by step by reducing the dimension of the system provided we operate according to the structure of the Lie algebra, i.e., according to the derived series for G, as in the case of deterministic equations [3,4,5,6,7,8,9]. ♢
We stress that for simple Lie-point symmetries, i.e., vector fields of the form
X = φi(x,t) ∂i , | (49) |
there is a simple way to relate the sought-for change of variables to the symmetry X [16,17,18]. First of all we note that in this case the time variable can be left unaffected by the change of variables. It suffices then to require that
X = φi(x,t)∂∂xi = φi[x(y),t]∂yℓ∂xi∂∂yℓ = ∂∂yn ; |
this requires to solve the linear equations
{φi(x,t)(∂yℓ/∂xi) = 0(ℓ=1,...,n−1)φi(x,t)(∂yn/∂xi) = 1, |
which are solved by the classical method of characteristics.
In particular, for n=1 we have to solve only the last equation, and its solution is promptly written as
y = ∫1φ(x,t) dx . |
Remark 5.5. The reader can wonder why this Proposition considers only deterministic symmetries. The reason is that in the case of random or W-symmetries, one is not guaranteed that the transformed equation is still of Ito type (this depends on a certain compatibility condition being satisfied, see [25,30]). Note that the problem is specially present when attempting multiple reduction. In fact, the first reduced equation could be not of Ito type, and in this case our theory can not say anything about successive reductions.
We also stress that in the scalar case the reduced equation, albeit not of Ito type, could be integrable; in this case its solution provides – inverting the transformation – the solution to the original equation.
Finally, we note that in [30], see in particular Sect. V therein, condition ensuring the symmetry reduced equations are still of Ito type are discussed in some detail (in any case, one can perform the reduction and verify if the reduced equation is Ito or otherwise). When all the reduced equations are of Ito type, Proposition 3 extends also to random or W-symmetries. ⊙
It is well known that symmetries of a deterministic differential equations form a Lie algebra [3,4,5,6,7,8,9] (for an explicit proof, see e.g., Sect. 2.3 (and Sect. 5.1 for the case of generalized symmetries) in ref.[6]). The same holds for stochastic equations, at least for deterministic or random symmetries; note this includes the class of symmetries (deterministic ones) which are guaranteed to allow multiple reductions (see Remark 5.5 above):
Proposition 5.4. The set of simple deterministic or random symmetries of a given Ito equation has the standard Lie algebra structure, i.e., if X,Y are symmetries of a given equation, their commutator Z=[X,Y] is also a symmetry for the same equation.
Proof. This is Theorem 4.3 in [26]; see there for the proof. Alternatively, note that symmetries of an Ito equation correspond to symmetries of the associated Stratonovich equation; for the latter the standard chain rule holds, so it is obvious that symmetries have the standard Lie algebra structure, proceeding just as in the case of deterministic equations. ♢
Remark 5.6. It should be noted that [20] contains a statement in contrast with Proposition 5.4 (see Sect. IX.A therein). That statement concerns transformations of a rather more general form than the one considered here† and it is thus wrong if seen in the present context, while Proposition 5.4 above is correct. ⊙
†In fact, far too general transformations, in particular functions h(x,t,w), are allowed at that point of the discussion in [20]; restricting to h(x,t,w)=Rw, as we do here, produces the right result.
Remark 5.7. Our Proposition 5.4 does not say anything concerning W-symmetries. In fact, the following Examples will show that in some cases these also form a Lie algebra (or possibly even a Lie module), while in other cases this is not the case. In other words, no general statement is possible about the Lie algebraic structure of W-symmetries of a given equation. ⊙
Finally, we recall that for any Ito Eq (35) there is an associated diffusion (Fokker-Planck) equation
ut = Aijuij − Biui − Cu , | (50) |
describing the evolution of the density u(x,t) in time; here ui:=(∂u/∂xi) and similarly uij:=(∂2u/∂xi∂xj), and the coefficients in the Fokker-Planck Eq (50) are related to those in the Ito Eq (35) by
Aij = 12(σσT)ij , Bi = fi − ∂j(σσT)ij , C = (∂ifi) − 12∂2ij(σσT)ij . | (51) |
Remark 5.7. The Fokker-Planck equation is often also written, more synthetically, as
ut = 12∂2∂xi∂xj[σσT u] − ∂∂xi[fiu]. | (52) |
Combining this equation with those for the transition from an Ito equation to the associated Stratonovich one, see Eq (37), it is immediate to get the Fokker-Planck equation associated to a Stratonovich equation. ⊙
Remark 5.8. We also recall that, as discussed in [13,19,22], any symmetry of the Ito Eq (35) leads (adding a standard term proportional to ∂u) to a symmetry of the associated Fokker-Planck Eq (50), but the converse is not true. This is rather natural, given that different Ito equations can have the same associated Fokker-Planck equation; in fact, symmetries of a Fokker Planck equation will in general map these different Ito equations one into the other. We refer to [13,14,19,22] for further detail and examples. ⊙
We consider here some examples illustrating the results discussed in this Section. We recall that we are considering simple admissible symmetries or W-symmetries; that is, vector fields of the form (41) with R a constant matrix of the form discussed in Sect. 5.1.
Example 5.1. Consider the equation
dx = t dw . |
The determining equations are then
φt + 12Δ(φ) = 0 , φw + tφx − Rt = 0 ; |
their general solution is
φ = Rx + k , |
with k a constant. Thus we have two symmetries (a deterministic one and a W-symmetry), i.e.,
X = ∂x , Y = Rx∂x + Rw∂w . |
It is immediate to check that
[X,Y] = R∂x |
is still a symmetry: thus in this case the set of (simple) symmetries, involving both "ordinary" and W-symmetries, is indeed a Lie algebra. ♢
Example 5.2. Consider the scalar Ito equation
dx = αxdt + βxdw , | (53) |
where α and β are real constants. This describes the geometric Brownian motion in one dimension, and it is well known that this equation can be integrated (see e.g., [36], Chapter 5). Note that if we start with x(0)>0, we are guaranteed to have x(t)≥0 for all times. The determining equations (43), (44) for simple symmetries‡ of (53) are
‡Recall this means τ=0 in terms of (39), i.e., that X should be of the form (41).
φt + αxφx − αφ + 12Δφ=0 , | (54) |
φw + βxφx − βφ − βxR=0 . | (55) |
Solving (55) yields
φ(x,t;w) = x [Rlog(x) + ψ(z,t)] ; z := w − β−1 log(x) . |
Inserting this into (54), the latter reads
(x2β) [β(β2 + 2α)R + 2βψt + (β2 − 2α)ψz] = 0 ; |
the solution to this equation is
ψ(z,t) = β(2α+β2)Rz + (2α−β2)Q(ζ)(2α−β2) , |
having defined
ζ = (2α − β2)t + 2βw − 2log(x)2α − β2 . |
Going back to evaluate φ, we thus we have symmetries depending on the arbitrary function Q and on the arbitrary constant R
X = [x (β2α−β2((2α+β2)w − 2βlog(x)) R + Q(ζ))] ∂x + (Rw)∂w . | (56) |
The simplest choice is of course R=0, Q(ζ)=1, which gives φ=x, i.e. the scaling symmetry
X0 = x ∂x . |
The associated integrating change of variables is simply
y = ∫1φ dx = ∫1x dx = log(x) ; |
In fact, Ito formula for the change of variables gives immediately (using (53) for the time evolution)
dy = (α − β2/2) dt + β dw , |
which yields at once (setting for ease of writing t0=0 and w(0)=0)
y(t) = y0 + (α−β2/2) t + βw(t) , |
and hence
x(t) = exp[y(t)] = x0 exp[(α−β2/2) t + βw(t)] . |
We can use this Example also to illustrate the Lie algebraic structure of the set of symmetries. Let us first set R=0, i.e., consider random symmetries, and two symmetries of the general form (56) with R=0, say
X = xP(ζ) ∂x , Y = xQ(ζ) ∂x . |
An explicit computation shows that
Z := [X,Y] = x H(ζ) ∂x |
where the function H is given explicitly by
H(ζ) := 1α−β2/2 [P′(ζ)Q(ζ) − P(ζ)Q′(ζ)] . |
This is still of the functional form (56) with R=0, and hence is immediately known to be still a (random) symmetry for our Eq (53).
On the other hand, let us consider the general case, i.e., let us not enforce R=0. We set a simplified notation writing
μ(x,w) = β2α−β2((2α+β2)w−2βlog(x)) , |
and consider vector fields
X=x (μ(x,w)R1 + P(ζ))∂x + (R1w)∂w ,Y=x (μ(x,w)R2 + Q(ζ))∂x + (R2w)∂w ; |
by our discussion above, they are symmetries of (53); we could actually also take R1=R2=R≠0. Then an explicit computation yields
Z = [X,Y] = x h(x,w;ζ) ∂x |
where the function h is given by
h(x,w;ζ):=1(α−β2/2)2[β2(βw−log(x))(R2P′(ζ)−R1Q′(ζ))+(α−β2/2)[(R1Q−R2P)β2+P′(ζ)Q(ζ)−P(ζ)Q′(ζ)]]. |
The vector field Z thus defined is not of the form (56) and hence is not a symmetry of the Eq (53).
This shows that W-symmetries of a given equation, contrary to simple deterministic and random symmetries, in general do not form a Lie algebra. ♢
Example 5.3. Consider the stochastic logistic equation with environmental noise
dx = (αx − βx2) dt + γxdw . | (57) |
Here α,β,γ are nonzero real constants; in the biological applications they have positive sign, but for our present purposes this is not relevant.
We will again look at the case with R=0, i.e., search for symmetries of the form X=φ(x,t;w)∂x. In this case the second determining equation (44) reads
φw + γ (xφx − φ) = 0 , |
which yields immediately
φ(x,t,w) = x ψ(z,t) ; z = w − γ−1 log(x) . |
Then the first determining Eq (43) splits into two equations (for the coefficients of x and x2 in the resulting expression):
ψt=(αγ − γ2) ψzψz=−γ ψ . |
These are easily solved, and we get in the end
φ(x,t,w) = x2 exp[−(At + γw)] , A := α − γ2/2 . |
Note this yields a random symmetry
X = x2 exp[−(At + γw)] ∂x . |
The associated change of variables and its inverse are given by
y=∫1φ(x,t,w) dx = −1x exp[At + γw] ; | (58) |
x=−1y exp[At + γw] . | (59) |
When we express the evolution of y using Ito's rule and (57), we get
dy = −β exp[At + γw] dt ; |
note this is not an Ito equation. It can however be directly integrated to give
y(t) = y0 − β ∫t0exp[Aτ+γw(τ)] dτ . |
Transforming back to the original variable x via (59) we obtain the solution to (57) in closed – albeit involved – form. The reader can see [75] for numerical experiments confirming the analytical procedure. ♢
Example 5.4. Consider the equation
dx = xe−tdt + x dw . | (60) |
(note that x(0)>0 entails x(t)≥0 for all t≥0.) This admits as symmetry
X = x F(ζ) ∂x , | (61) |
where F is an arbitrary function of
ζ := e−t + t2 − w + log(x) . |
The simplest choice is of course F(ζ)=1, i.e.,
X = x ∂x ; |
note this is a deterministic symmetry. The associated change of variables is just y=log(x), and with this our Eq (60) is mapped into
dy = (e−t−1/2) dt + dw , |
which is directly integrated to
y(t) = y(t0) − [(t−t0)2 + (e−t−e−t0)] + (w(t)−w(t0)) . |
This yields the solution to our Eq (60) upon inversion of the change of variables, i.e., setting x=ey.
Suppose we do not want to make this simplest choice, and go for the next simpler one; this will produce a random symmetry. So let us choose F(ζ)=ζ, and hence
X = x [e−t + t2 − w + log(x)] ∂x . |
Then the associated change of variables is
y = log[e−t+t2−w+log(x)] + β(t,w) . |
the equation for y is then
dy = (βt + 12βww)dt + βwdw . |
Thus we should choose β(t,w)=b(t)+cw in order to get an Ito equation. With this choice, the initial equation is mapped into
dy = [b′(t)] dt + c dw , |
which is readily integrated to yield
y(t) = y(t0) + [b(t) − b(t0)] + c[w(t) − w(t0)] . |
Inverting the change of coordinates – which is now a more complex operation than in the previous case – we obtain the solution for x(t).
We can again check that the symmetries, i.e., vector fields of the form (61) (which are random symmetries), form a Lie algebra. In this case, it is in fact immediate to check that if
X = x F(ζ) ∂x , Y = x G(ζ) ∂x , |
then we have
Z := [X,Y] = x H(ζ) ∂x |
where we have defined
H(ζ) = F′(ζ) G(ζ) − F(ζ) G(ζ) . |
Thus Z is of the form (61), and hence a symmetry for Eq (60), as claimed. ♢
Example 5.5. Consider the two-dimensional equation (here and elsewhere in concrete examples, we write vector indices as lower ones, to avoid any possible confusion with squares)
dx1=[ex1 − 12e−2x1]dt + e−x1dw1dx2=12ex2[2ex1 + ex2 + e2x1+x2]dt − ex1+x2dw1 − ex2dw2 . | (62) |
It is easily checked that this admits the symmetry vector field
X = −ex2∂2 ; | (63) |
symmetry-adapted variables are
y1 = exp[x1] , y2 = exp[−x2] ; | (64) |
in these variables, the vector field is
X = ∂/∂y2 |
and the Eq (62) reads simply
dy1=y21dt + dw1dy2=−y1dt + y1dw1 + dw2 . | (65) |
That is, we have an autonomous Ito equation for y1; if this is solved, say with y1=Φ[t,w1(t)], then the equation for y2 is immediately integrated to give
y2(t) = y2(0) − ∫t0Φ[t,w1(t)]dt + ∫t0Φ[t,w1(t)]dw1(t) + ∫t0dw2(t) . |
The solution to the initial problem (62) is then obtained by inverting the change of variables (64). ♢
Example 5.6. Consider, as in [30], the two-dimensional isotropic linear stochastic equation
dxi = λxidt + μdwi (i=1,2) . | (66) |
This was considered, for what concerns W-symmetries in [30] (this was Example 7 therein), but only W-symmetries were considered in that work; we will thus have to derive the full symmetry algebra.
We write, for ease of notation, R in the form
R = (r0r1−r1r0) . |
Solving the σ-determining equations for φi(x,t;w) yields
φ1(x1,x2,t;w1,w2)=r0x1 + r1x2 + ψ1(z1,z2,t) ,φ2(x1,x2,t;w1,w2)=−r1x1 + r0x2 + ψ2(z1,z2,t) ;zi:=wi − xi/μ . |
With this, the f-determining equations read
∂ψ1∂t=λμ(x1∂ψ1∂z1 + x2∂ψ1∂z2) + λψ1 ,∂ψ2∂t=λμ(x1∂ψ2∂z1 + x2∂ψ2∂z2) + λψ2 . |
Note that the ψ are functions of the zi variables alone; thus coefficients of the xi must vanish separately, and we get
ψi(z1,z2,t) = eλt ci |
where ci are arbitrary constants.
In conclusion, the symmetry algebra G for (66) is spanned by the vector fields
X1=x1∂1 + x2∂2 + w1ˆ∂1 + w2ˆ∂2 ,X2=x2∂1 − x1∂2 + w2ˆ∂1 − w1ˆ∂2 ;Y1=exp[λt]∂1 ,Y2=exp[λt]∂2 . |
Note that the Xi are W-symmetries, while the Yi are deterministic symmetries. A trivial computation shows that the commutation relations are given by
[X1,X2]=0 , [Y1,Y2]=0 ;[Y1,X1]=Y1 , [Y1,X2]=−Y2 ,[Y2,X1]=Y2 , [Y2,X2]=Y1 . |
This shows that G is indeed a Lie algebra; moreover GX={X1,X2} and GY={Y1,Y2} are abelian subalgebras, and GY an abelian ideal in G. ♢
Consider a general function
F = F(x,t,w) ; | (67) |
its Ito differential is
dF = ∂F∂tdt + ∂F∂xidxi + ∂F∂wadwa + 12ΔFdt . |
When we evaluate this on the dynamics described by the Ito Eq (35), we get
dF = [∂F∂t + fi(x,t)∂F∂xi + 12ΔF]dt + [∂F∂wa + σi a(x,t)∂F∂xi]dwa . | (68) |
It follows immediately that if (and only if) F satisfies the equations
∂F∂t + fi(x,t)∂F∂xi + 12ΔF=0 , | (69) |
∂F∂wa + σi a(x,t)∂F∂xi=0 (a=1,...,m) , | (70) |
then F is an invariant for the SDEs (35).
Note that in general the invariants will be arbitrary functions of some "basic invariants" functions.
Example 6.1. Consider the equation
dxi = fi(t)dt + σi k(t)dwk (i=1,...,n) ; | (71) |
then the Eq (70) yield immediately that
F(x,t,w) = F(z,t) ; zi := xi − ∫σi k(t)dwk(t) . |
A straightforward computation shows that in this case ΔF=0 (note in particular Δzi=0). The Eq (69) reads then
∂F∂t − fi∂F∂zi = 0 , |
and it follows immediately that
F = F(ζ) , |
having defined
ζi := zi − Φi(t) = xi − Φi(t) − ∫σi k(t)dwk , | (72) |
where Φ is the primitive of f,
Φi(t) = ∫ fi(τ)dτ . | (73) |
Obviously, the ζi are in this case the basic invariants mentioned above. ♢
The existence of invariants allows to express the solutions to a SDE system in terms of a reduced system of Ito equations; if there are as many independent invariants as degrees of freedom, then the solution can be expressed purely in terms of invariants.
In other words, the presence of a sufficient number of invariants guarantees the integrability of the equation, pretty much as in the deterministic case the presence of a sufficient number of conserved quantities guarantees integrability.
Example 6.2. In the case of Eq (71), we have the basic invariants defined in (72); at the initial time t=0 these will satisfy ζi=ζi(0). If wi(0)=0, and choosing the constant in the primitives of (73) so that Φi(0)=0, at t=0 we have
ζi(0) = ζi0 = xi0 = xi(0) . |
Thus we have at all times and for each realization of the Wiener processes wk(t), and with Φi as in (73) above,
xi(t) = xi0 + Φi(t) + ∫t0σi k(τ) dwk(τ) . |
This yields directly the solution to (71). ♢
It is natural to wonder if and how invariants are related to symmetries of the stochastic equation.
Comparing Eqs (43), (44) one the one hand, and (69), (70) on the other, we immediately observe a rather trivial (but useful) relation: if both fi and σi k do not actually depend on the x variables (but possibly depend on t) and set R=0, the symmetry coefficients φi are also invariants for the Ito equation. (Note that such relation does not exist for R≠0, i.e., for proper W-symmetries [30].)
We would be interested in knowing if other relations exist. This matter has been studied in recent papers by one of us [26,27,28] (to which we refer for details and examples), and we briefly report the relevant results here.
Proposition 6.1. The set of invariants for a given Ito equation is a ring over R, i.e., if F,G are invariants for the given equations and α,β are real constants, so are
H = αF + βG , K = F⋅G . |
Proposition 6.2. If a given Ito Eq (35) admits the symmetry X and the invariant F, then G=X(F) is also an invariant for the same equation.
Proposition 6.3. The set of symmetries for a given Ito equation has, beside the structure of Lie algebra, the structure of a Lie module over the ring of invariants; i.e., if X,Y are symmetries for the given equation, and F,G invariants for the same equation, then
Z = FX + GY |
is still a symmetry for the equation.
Proof. Proposition 6.1 is trivial; Propositions 6.2 and 6.3 are respectively Theorems 4.4 and 4.6 in [26]; see there for proofs. ♢
We stress that, obviously, a generic stochastic equation has no invariants, even in the case it has symmetries.
Example 6.3. Consider again the two-dimensional isotropic linear stochastic Eq (66) considered in Example 5.6 above (where it was found it has a nontrivial symmetry algebra). Now a direct computation shows that there are no invariants. In fact, invariants should satisfy the equations
Jw1+μJx1 = 0 ,Jw2+μJx2 = 0 ;Jt+λ(x1Jx1+x2Jx2) = 0 . | (74) |
Note in the last equation there should also be a term (1/2)ΔJ with Δ the Ito Laplacian (42), but in this case it results ΔJ=0.) Solving the first two equations yields
J(x1,x2,t,w1,w2) = H(z1,z2,t) ; zi := wi − xi/μ . |
Plugging this into the last Eq (74) yields H=const, thus showing that only trivial invariants are present. ♢
We now want to argue that – pretty much in the same way as a great extent of the standard theory of symmetry of deterministic differential equations can be extended to the framework of stochastic differential equations – the concepts of conditional, partial and asymptotic symmetry also apply for stochastic differential equations. What is more, these are also useful to determine solutions to stochastic equations.
It will happen that these are intimately related – much more than in the deterministic setting – to similar concepts for invariants, i.e., to conditional, partial or asymptotic invariants. These, and the relation with symmetries, will be discussed in the next Sect. 8. We will first discuss the matter in abstract terms – actually with a more formal writing than in previous Sections, as now we develop new material and results; and then provide, in Sect. 9 below, a number of concrete Examples.
Our first observation is that conditional and partial symmetries can be, in principle, defined for stochastic equations exactly as in the case of deterministic ones.
The difference – and the reason to write "in principle" in the lines above – is that while a solution of a deterministic equation is determined by the initial condition, in the case of stochastic equation it is determined by the initial condition and by the realization of the driving Wiener processes. But, of course, we do not know a priori what the realization of the Wiener process will be, so we do not want to discuss properties which depend on such realization. (A partial exception to this is that often we are satisfied with properties which hold for generic realizations of the Wiener processes, i.e., for a full measure set of the driving Wiener paths.)
Thus, in practice, it is quite difficult – and makes little sense in view of applications – to study symmetries of a given solution, or of a given set of solutions.
The exception is provided by the case where the solutions in the set which is studied are identified by their living on an invariant manifold for the equation. By this we mean that the Eq (35) describes a variable (x,t)∈M (M is the phase manifold), and there is a proper submanifold M0⊂M such that if (x(0),t0)∈M0, then for any realization of the driving Wiener processes, and for any t≥t0, (x(t),t)∈M0.
We will actually confine ourselves to the case of autonomous equations, i.e., the case where the coefficients fi(x,t) and σi k(x,t) in (35) are actually independent of t. In this case we can work with the reduced phase manifold§, and in the following M should be interpreted as the reduced phase manifold. Extension of our discussion and results to the non-autonomous is rather simple, and in any case can be obtained – remaining within the autonomous formalism – by adding a new variable x0 with evolution governed by dx0=dt.
§This was denoted as M0 in Sect. 2; but here we reserve this notation to the dynamically invariant submanifold. We trust no confusion will arise.
We will see in the following that the existence of these invariant manifolds is naturally related to "conditional" (or "partial") invariants for the equation.¶
¶The simplest – but rather trivial – case in which there is an invariant manifold is that of a fixed point; for example the stochastic Eq (35) with coefficients satisfying fi(0,t)=0, σi j(0,t)=0, automatically admit x=0 as a fixed point, and hence a (highly symmetric) solution x(t)≡0. This is of little interest.
The simplest nontrivial case is that in which the dynamics admits a nontrivial invariant submanifold M0⊂M (this is of course more significant if M0 is stable or attracting, as will be discussed later on). In this case we can consider the restriction of our system to M0, and this may have symmetries which are not present for the full system. These in turn may allow for the integration of the restricted equation; solutions obtained in this way are also special solutions for the full system.
We start by stating some very simple – but useful – facts.
Lemma 7.1. Let the SDE Δ (35) in the n-dimensional manifold M admit an invariant submanifold M0⊂M of dimension m<n. Consider the restriction Δ0 of Δ to M0. Any solution to Δ0 is also a solution to Δ, and the most general solution to Δ with x(0)∈M0 is obtained as a solution to Δ0.
Proof. Obvious.
Lemma 7.2. In the setting of Lemma 7.1, let Δ0 admit m simple symmetries spanning a solvable algebra. Then it can be integrated, and hence all the solutions to the initial equation Δ with initial conditions on M0 can be obtained.
Proof. Integrability of Δ0 follows from Proposition 5.3 above. Once Δ0 is integrable, the rest of the statement follows immediately from Lemma 7.1. ♢
This construction is also generalized to the case where the symmetry algebra has smaller dimension, as stated in the following Lemma.
Lemma 7.3. Let the Ito SDE Δ in the n-dimensional manifold M admit an invariant submanifold M0⊂M of dimension m<n. Let Δ0 be the restriction of Δ to M0. Then
(a) If Δ0 admits s<m simple symmetries spanning a solvable algebra G which generates a local Lie group G, then Δ0 can be reduced to an equation ˆΔ0 on M0/G.
(b) Let ˆx(t) be any solution to ˆΔ0; then ˆx(t) extends to a solution x(t) of Δ0, and this is turn describes the most general solution to Δ with x(0)∈M0.
Proof. Point (a) follows by Proposition 5.3 above. Solutions to ˆΔ0 are extended to solutions to Δ0 again by means of Proposition 5.3, i.e., via the reconstruction equations. In view of Lemma 7.1, the exact solutions obtained in this way are also all the solutions to the initial equation Δ with initial conditions on M0. ♢
Remark 7.1. The statement in these Lemmas are rather trivial, as shown by their very simple proofs. Actually these Lemma are more the description of a strategy to obtain a set of solutions (identified by living on M0) to Δ. ⊙
Remark 7.2. If the equation admits a standard (global) invariant, then all the level manifolds for this invariant are invariant manifolds of the differential equation; see Example 9.1 below. Thus the Lemmas 7.1, 7.2 and 7.3 given above apply in general, in the sense that any initial datum lies on such an invariant manifold. But they apply also to the case where no global invariant exists (we will see in a moment a weaker type of invariant is involved here), i.e., when we have isolated invariant manifolds. ⊙
Remark 7.3. The presence of an isolated invariant manifold can be described in terms of a conditional invariant for the stochastic equation. That is, if M0 is the submanifold described by F(x)=m0 (here m0∈R) and M0 is an invariant manifold while there are no nearby invariant manifolds, then F(x) is a conditional invariant for our equation. Note that this means that on the dynamics we have
dF = [F(x)−m0] [α(x,t)dt + βk(x,t)dwk] , |
with nonzero α,βk. In this case only the m0 level set is invariant, while the other ones are in general not invariant, as in our Examples 9.2 and 9.3 below (in both cases the conditional invariant is just ρ and the invariant level set is ρ=1). We will discuss this point in more detail in Sect. 8 below. ⊙
Remark 7.4. This situation is similar to the setting considered by Misawa and by Albeverio & Fei [76,77,78,79] in their early study of symmetry of stochastic equations (see the discussion in [22]). In their case, however, we had a global (proper) invariant, hence a global constant of motion – which allowed for a dimensional reduction of the stochastic system – as in Example 9.1 below, while in the setting discussed here we only have a conditional one, hence reduction can take place only on the specific level set of the conditional constant of motion, viz. the manifold M0. ⊙
Remark 7.5. The situation discussed in this Section is related to conditional constants of motion [80,81,82,83,84] and conditional invariants, so that it would be natural to speak of conditional symmetries. On the other hand, in this case the solutions (to the restricted problem) are not invariant under the conditional symmetry: they are actually mapped one into another by the action of the latter. In this sense, the situation is more similar to that encountered when dealing, in the deterministic framework, with partial symmetries; from this point of view it might be more appropriate to speak of partial invariants; but we will prefer to use the term conditional invariant in view of the first mentioned parallel with the deterministic case, and the established use in the latter. ⊙
In the previous discussion, in particular Lemmas 7.1 through 7.3, we have assumed the existence of an invariant manifold M0⊂M for the dynamics described by our Ito Eq (35). Needless to say, this is a highly non-generic situation. In fact, as anticipated in Remark 7.3, the existence of invariant manifolds is related to the existence of (conditional) invariants for the dynamics.
In order to identify invariant manifolds in M, or in M×R, these invariants should depend only on the x, or the (x,t), variables. We have thus the following:
Definition 7.1. When the invariant J=J(x,t;w) for an Ito equation does not depend on the w variables, i.e., J(x,t;w)=Ψ(x,t), we say that J is a phase invariant; if moreover Ψ is independent of t, we say J is a configurational invariant.
Remark 7.6. We stress that while configurational invariants can identify invariant manifolds, and phase invariant can identify time-varying invariant manifolds, in order to identify solutions in terms of invariants we need invariants which also depend on the wi variables, and hence on the realization of the Wiener processes. See also Example 9.1 in this respect. ⊙
We recall that if J:M→R is a smooth function defined on the phase manifold M, or more generally J:M×W→R where W≈Rn is the space in which the driving Wiener processes wi=wi(t) take values, then its evolution on the dynamics described by the Ito equation Δ – written as (35) – is given by
dJ=(∂J∂t)dt + (∂J∂xi)dxi + 12 Δ(J)dt=[(∂J∂t) + (∂J∂xi) fi(x,t) + 12 Δ(J)] dt + [(∂J∂wk) + (∂J∂xi)σi k(x,t)] dwk . | (75) |
We are specially interested in the case where J is a function of x and t alone, J(x,t;w)=Ψ(x,t); we say then that J is a phase space function, and if Ψ does not depend on t we say it is a configurational function. In both these cases the Ito Laplacian reduces to
ΔJ = ΔΨ = σikσj k ∂2Ψ∂xi∂xj . | (76) |
Lemma 7.4. The evolution of a phase space function under the dynamics described by an Ito Eq (35) is itself described by an Ito equation.
Proof. Using (75) and specializing it to the case J=Ψ(x,t), we have
dJ=[(∂Ψ∂t) + (∂Ψ∂xi) fi(x,t) + 12 Δ(Ψ)] dt + [(∂Ψ∂xi)σi k(x,t)] dwk . | (77) |
Recalling also (76), we note that now all terms within the square brackets depend on (x,t) alone, i.e., we have again an equation of Ito type. ♢
It is now natural to introduce the following Definition, which leads immediately to the foregoing Lemma 7.5.
Definition 7.2. (a) If the function J is such that dJ=0 on solutions to the Ito equation (35), i.e., (75) is identically zero, we say that it is an invariant for the Ito equation (a phase space invariant if J=Ψ(x,t), and (78) is identically satisfied, a configurational invariant if J=Ψ(x)).
(b) If this relation is valid only on the level set J=c, we say that J is a conditional invariant (a conditional phase space invariant if J=Ψ(x,t), a conditional configurational invariant if J=Ψ(x)) for the Ito equation.
Lemma 7.5. Phase space invariants are identified by the n+1 equations
(∂Ψ∂t) + (∂Ψ∂xi) fi(x,t) + 12 Δ(Ψ) = 0 ,(∂Ψ∂xi)σi k(x,t) = 0 . | (78) |
Proof. This follows at once from (77). ♢
Remark 7.7. If the solutions to the (78) identify one or more level sets of J=Ψ (in practice, if the l.h.s. of the above equations can be factorized with a factor being just a function of Ψ), then this or these level sets correspond to invariant manifolds for the dynamics. We will see in concrete Examples that this may happen. ⊙
Remark 7.8. As configurational invariants (and functions) are special cases of phase space invariants (and functions), the Lemmas 7.4 and 7.5 above also apply to them. Note that the Eq (78) read in this case
(∂Ψ∂xi) fi(x) + 12 Δ(Ψ) = 0 ,(∂Ψ∂xi)σi k(x) = 0 , | (79) |
where we have taken into account the fact that configurational invariants naturally arise in autonomous systems. ⊙
We have so far discussed the case where an invariant submanifold M0⊂M exists, and considered dynamics on it. We have not discussed neither what happens nearby the invariant manifold, nor the stability of such manifold.
As we are considering stochastic systems, we should consider at least two types of (local or global) stability [85,86].
(A) On the one hand, it is possible that for any initial datum x0 in a neighborhood B0⊃M0 and any realization of the Wiener processes wi(t), the dynamics is attracted towards M0.
(B) On the other hand, it is possible that for any x0∈B0, considering the Fokker-Planck equation with initial condition u(x,0)=δ(x−x0) the probability density function u(x,t) remains concentrated around M0.
Definition 7.3. When the situation described in case (A) above occur, we say that M0 is locally strongly attractive; if this holds for any x0∈M, we say it is globally strongly attractive.
Definition 7.4. When the situation described in case (B) above occur, we say that M0 is locally weakly attractive; if this holds for any x0∈M, we say it is globally weakly attractive.
Remark 7.9. Note that in case (B) above typically there will be a parameter σ0≥0 – related in physical terms to the intensity of the noise described by the Wiener processes – which controls the spreading of the asymptotic measure around M0, and such that in the limit σ0→0 the asymptotic measure u∗(x)=limt→∞u(x,t) collapses to a measure on M0. ⊙
Lemma 7.6. In the case (A), the most general asymptotic solution to Δ will be described by the most general solution to Δ0.
Proof. In fact, under these hypotheses any dynamics starting in B will be attracted to the invariant manifold M0, and on this the dynamics is described by Δ0. The conclusion follows then by Lemma 7.1 above. ♢
Remark 7.10. Needless to say, even in the case Δ0 is fully integrable we will in general not be able to describe which ones of the solutions to Δ0 will describe the asymptotic behavior for a given solution to Δ, i.e., for given initial datum x0∈B and given realization of the Wiener processes, as we will not be able to integrate the transient dynamics. ⊙
Remark 7.11. The Remark above refers to case (A). In case (B), on the other hand, the dynamics can not be reduced, even asymptotically, exactly to the dynamics on M0 and hence to Δ0. Thus even in the case where Δ0 is exactly integrable – and even for x0∈B0 – this does not give a description of the asymptotic dynamics for Δ. In this case we have a weaker result. ⊙
Lemma 7.7. In the case (B), for small enough σ0, the asymptotic dynamics will be described for most of the time, by the linearization ˜Δ0 of Δ around M0.
Proof. Under the hypotheses holding in case (B) and for small enough σ0, the measure will be asymptotically concentrated around M0, so that the linearization ˜Δ0 (which, in turn, will be a linearization around Δ0) will describe "most" of the asymptotic solutions; by this we mean that there could be from time to time large fluctuations driving the dynamics for a short time outside of the immediate neighborhood of M0, i.e., outside the region where Δ reduces to ˜Δ0. These large fluctuations will occur with a frequency exponentially small in σ0 and will decay rapidly so that the dynamics is quickly driven back to the region around M0. ♢
Remark 7.12. This Lemma does not give a result of the same strength as the previous one; on the other hand, it says that one will have a reliable description in terms of ˜Δ0, except for rare and short fluctuations. ⊙
As far as symmetries are concerned, our discussion suggests naturally the following definitions.
Definition 7.5. Let M0⊂M be a strongly attractive submanifold for the equation Δ, and let Δ0 be the restriction of Δ to M0; let X be a symmetry for Δ0, hence a conditional symmetry for Δ. We say that X is also a strong asymptotic symmetry for Δ.
Definition 7.6. Let M0⊂M be a weakly attractive submanifold for the equation Δ, let Δ0 be the restriction of Δ to M0, and let ˜Δ0 be the linearization of Δ around M0; let X be a symmetry for Δ0, hence a conditional symmetry for Δ. If X is also a symmetry for ˜Δ0, then we say that X is also a weak asymptotic symmetry for Δ.
Both these cases are related to the presence of asymptotic conditional invariants, as we are now going to discuss in the next Section 8.
We stress that we are again interested in the case of phase space or configurational invariants, i.e., where the would-be invariants depend on (x,t) or on the x variables alone; thus we are in the setting discussed above, and the evolution is described by (77).
The discussion of Section 7.2, referring to conditional symmetries, can be extended to the case of asymptotic symmetries.
The key observation is that the evolution of a phase space function under the dynamics of an Ito Eq (35) is described by (77), i.e., by an equation of Ito type; this will be denoted, for ease of reference, as ΔJ. So our discussion of (stochastic) stability for submanifolds M0 of M under Δ can be repeated for the stability under ΔJ.
Definition 8.1. If J=Ψ(x) is a conditional configurational invariant for the Ito equation Δ on the invariant manifold M0, and M0 is a strongly attractive manifold for Δ, we say that J is a strong asymptotic invariant for Δ.
Definition 8.2. If J=Ψ(x) is a conditional configurational invariant for the Ito equation Δ on the invariant manifold M0, and M0 is a weakly attractive manifold for Δ, se say that J is a weak asymptotic invariant for Δ.
Lemma 8.1. Let J be a strongly asymptotic configurational invariant for the Ito equation Δ, associated to the invariant manifold M0, and let Δ0 be the reduction of Δ to M0. Then the asymptotic solutions to Δ are described by the solutions to Δ0.
Proof. By definition, under the hypotheses of the Lemma M0 is a strongly attractive invariant manifold for Δ, hence the statement follows immediately from Lemma 7.6, which itself uses Lemma 7.1. ♢
Lemma 8.2. Let J be a weakly asymptotic configurational invariant for the Ito equation Δ, associated to the invariant manifold M0, and let ˜Δ0 be the linearization of Δ around M0. Then the asymptotic solutions to Δ are described by the solutions to ˜Δ0, up to large fluctuations.
Proof. By definition, under the hypotheses of the Lemma M0 is a weakly attractive manifold for Δ. Thus, according to Lemma 7.7, the dynamics of Δ lives, for most of the time and up to rare and short-lived large fluctuations, in a neighborhood of M0; thus it is described by ˜Δ0, and the Lemma is just a restatement of these facts in terms of (asymptotic) invariants. ♢
Remark 8.1. The approach discussed in this subsection is strongly related to stochastic Lyapounov functions, see e.g., [85,86]. Some of our ideas are also related to ideas developed in the context of bifurcation for stochastic systems [87]. ⊙
Remark 8.2. Our discussion used configurational invariants to relate invariants to attractive invariant submanifolds. As stressed above, see Remark 7.6, configurational or phase space (asymptotic) invariants can only do this, i.e., they cannot be used to obtain solutions to Ito equations, at difference with general (asymptotic) invariants, i.e., invariants depending also on the w variables. See also the discussion in Sect. 6 for the case of full invariants. ⊙
We will now consider concrete Examples related to the matters discussed in Sections 7 and 8 above. These matters are of course strongly interconnected, but we have separated the Examples in different types – and correspondingly separated the present Section in different subsections – for ease of reading.
In most cases, our Examples are built so that the compact invariant manifold is just a circle (usually the unit one), as we want to focus on conceptual issues rather than on computational difficulties. Thus it is not surprising that things will be simpler in polar or spherical coordinates, and indeed we will mostly work in polar coordinates, albeit in some cases we will also start from Cartesian ones.
We start by considering an Example due to Misawa [79]; this is Example 1 in there, where it is considered in the equivalent Stratonovich form, and only up to checking J is an invariant.
Example 9.1. Consider the Ito equation
dx1=(x3−x2−12(2x1−x2−x3))dt + (x3−x2)dwdx2=(x1−x3−12(2x2−x3−x1))dt + (x1−x3)dwdx3=(x2−x1−12(2x3−x1−x2))dt + (x2−x1)dw | (80) |
(note that only one Wiener process is involved). In this case the function
J = x21+x22+x23 = ρ2 | (81) |
is an invariant. In fact one has
ΔJ = 4 [J − (x1x2+x2x3+x3x1)] , |
and then dJ=0 on the Eq (80) follows immediately from a direct computation. Thus any sphere of radius ρ is invariant under the evolution described by (80).
It is also easily checked, by the same procedure or by just looking at (80), that
H = x1 + x2 + x3 | (82) |
is another invariant.
By considering the Jacobian matrix for the change of coordinates (x1,x2,x3)→(J,H,z) with z=x3, we observe that its determinant is 2(x1−x2), i.e., this is singular on the plane x1=x2; we will thus have to consider two regions.
Implementing the change of variables (x1,x2,x3)→(J,H,z) we have that the inverse change is given (in the two regions x1<x2 and x1>x2) by
x1=12 [(H − z) ∓ √2(J−z2) − (H−z)2] ,x2=12 [(H − z) ± √2(J−z2) − (H−z)2] . |
Note that by elementary algebra, and in view of the definitions of J,H, and z, we have
2(J−z2) − (H−z)2 = (x1 − x2)2 ; |
thus the argument of the square root in the formula above is always positive. We introduce, for ease of writing, the new real function
χ(J,H;z) = √2(J−z2) − (H−z)2 , |
so that the above inverse change of variables reads
x1=12 [(H − z) ∓ χ(J,H;z)] ,x2=12 [(H − z) ± χ(J,H;z)] . | (83) |
On the one-dimensional manifolds (corresponding to the ± sign in the formulas above) identified by given values for J=ρ2 and H we have the reduced equation
dz = [12 (H−3z) ± χ(J,H;z)] dt ± χ(J,H;z) dw . | (84) |
Note that here z∈[−ρ,ρ] (recall J=ρ2), and that for z=±ρ we have J=z2, H=z and x1=x2=0.
We can further look for symmetries of the reduced equation. We disregard W-symmetries, i.e., set R=0 in (44)||; with straightforward computations we get
||The case with R≠0 can also be analyzed completely, yielding of course more complex formulas.
φ±(z,t;w) = χ(J,H;z) ψ±(u±) , |
where the ± refers to the two determinations given above for x1 and x2 in terms of z, ψ± are arbitrary function of their argument, and
u± = w + t ± 1√3 arctan[H−3z√3χ(J,H;z)] . |
The simplest case does of course correspond to ψ±(u±)=c±, say
ψ±(u±)=±1 |
for ease of discussion (note the two ± signs are independent), i.e., to**
**This notation might be confusing: a more precise notation would be φ±=s±χ, with s± a sign.
φ±(z,t;w) = ± χ(J,H;z) . |
With this, the integrating change of variables is given by
ξ = ∫1φ±(z,t;w) dz = ∓ 1√3 arctan[H−3z√3 χ(J,H;z)] . |
With a direct application of Ito rule, the evolution equation for the random variable ξ turns out to be
dξ = dt + dw . | (85) |
This is immediately integrated, yielding
ξ(t) = ξ(t0) + (t−t0) + [w(t)−w(t0)] . |
Going back to the z variable we obtain (note that 3J−H2≥0)
z = H3 ± √23√3J−H2 sin(√3ξ) . |
Having obtained z(t), i.e., x3(t), now x1(t) and x2(t) are also obtained via (83). ♢
We now consider several examples with conditional invariants; at this stage we will not yet look at asymptotic symmetries or invariants.
Example 9.2. Let us work in R2 with polar coordinates (ρ,ϑ), and consider the Ito equations
dρ=a(1−ρ2)ρdt + σ(1−ρ2)dw1dϑ=[b(1−ρ2) + ωρ2] dt + σdw2 | (86) |
with a,b,ω,σ nonzero real constants.
It is clear that the unit circle ρ=1 is an invariant manifold for these equations; the dynamics on the unit circle is described by
dϑ = ωdt + σdw2 |
i.e., by a uniform rotation with a superimposed Wiener process. This yields immediately
ϑ(t) = ϑ(0) + ωt + σ[w2(t)−w2(0)] ; ρ(t)=1 . |
Thus albeit we have no information about the dynamics for generic initial conditions, we have a full description of the dynamics with initial condition on the unit circle.
The discussion is immediately generalized – in its entirety – to equations of the form
dρ=α(ρ,ϑ)dt + σ1(ρ,ϑ)dw1dϑ=β(ρ,ϑ) dt + σ1(ρ,ϑ)dw2 |
with α,β,σi smooth functions and
α(ρ0,ϑ) = 0 = σ1(ρ0,ϑ) ,∂ϑβ(ρ0,ϑ) = ∂ϑσ1(ρ0,ϑ) = 0 |
for some ρ0.
If we drop the last condition, i.e., require only α(ρ0,ϑ)=0=σ1(ρ0,ϑ), we still have reduction to a one dimensional equation on the circle of radius ρ0, albeit we have no information on the solutions living on this circle.
Note that in this case the full equations are rotationally invariant, as immediately seen from (86); but the associated function J=ρ is a conditional invariant. ♢
Example 9.3. Consider R3 with cylindrical coordinates (ρ,ϑ,z) and the Ito equations
dρ=a(1−ρ2)dt + κ(1−ρ2)dw1dϑ=[ω+b(1−ρ2)+ccos(z)]dt + σ1(ρ,ϑ,z)dw2dz=γ(r,ϑ,z)dt + σ2(ρ,ϑ,z)dw3 , | (87) |
with {a,b,c,κ,ω} nonzero real constants and {γ,σ1,σ2} smooth functions. By construction, the cylinder M0 of equation ρ=1 is an invariant submanifold (the function J1=ρ is a conditional invariant), and on it the system reduces to
dϑ=[ω+ccos(z)]dt + σ1(1,ϑ,z)dw2dz=γ(1,ϑ,z)dt + σ2(1,ϑ,z)dw3 , |
Suppose now that in general {γ,σ1,σ2} depend effectively on ϑ, but that
(∂ϑγ(ρ,ϑ,z))ρ=1 = (∂ϑσ1(ρ,ϑ,z))ρ=1 = (∂ϑσ2(ρ,ϑ,z))ρ=1 = 0 . |
Then the restricted system admits rotations in ϑ as symmetries, and hence the system can be further reduced to an equation in z. Albeit we are in general not able to solve this (one-dimensional) equation, we know that given a solution we can reconstruct a solution of the two-dimensional restricted system, and this is a special solution to the full three-dimensional problem.
Similarly it is easy to produce examples in which not only the cylinder ρ=1 is invariant, but within this we have the circle S10={ρ=1 z=0} which is also invariant. In this case the equation is fully integrated for initial data on S10, partially integrated for initial data on the cylinder ρ=1, and in general nothing can be said for general initial data (but Eq (87) show that an asymptotic analysis will be possible). ♢
Example 9.4. In the previous examples, everything was rather obvious by construction, thanks to the use of polar coordinates and the fact the invariant manifold was just a circle of radius one. Things are slightly less evident if we use Cartesian coordinates, and we want to show now that one can indeed work also with a less "ready to use" setting.
Consider, with r2=x21+x22 for short, the equation
dx1=[(α(1−r2)−(σ2/2))x1 − (β(1−r2)+ωr2)x2] dt + σ(1−r2)x1dw1 − σx2dw2dx2=[(β(1−r2)+ωr2)x1 + (α(1−r2)−(σ2/2))x2] dt + σ(1−r2)x2dw1 + σx1dw2 . | (88) |
We consider now the function J=r2, and look at its variation on the dynamics (88). This yields
dJ=∂J∂tdt + ∂J∂x1dx1 + ∂J∂x2dx2 + 12Δ(J)dt=[(2a + (1−J)σ2)J(1−J)] dt + 2σJ(1−J)dw1 . | (89) |
Thus we obtain that J is not an invariant; but it is a conditional invariant: on the level sets J=0 (corresponding to the origin in the Cartesian plane) and J=1 (the unit circle in the Cartesian plane) we have dJ=0, and hence these are invariant manifolds for the dynamics (88).
Actually, passing to polar coordinates (ρ,θ) (on ρ=0 the change of coordinates is singular, and θ ill-defined) the Eq (88) reads simply
dρ=αρ(1−ρ2)dt + σρ(1−ρ2)dw1dθ=[β(1−ρ2)+ωρ2]dt + σdw2 . | (90) |
It is thus evident that on J=ρ2=1 the dynamics reduces to
dρ=0dθ=ωdt + σdw2 . |
It should be noted that, as immediately apparent from (90), in this case rotations in the (x1,x2) plane are a symmetry of the full equation, and not only of the reduced one. We thus see again that one can have a conditional invariant associated to a full symmetry. ♢
Example 9.5. We generalize the previous Example; we use Cartesian coordinates (x,y) and write again r=√x2+y2. Consider the equations
dx1=[((1−r)x1f1−x2f2)−(1−r)22((s221+s222)x1+2(s11s21+s12s22)x2)] dt + (1−r)(s11x1−s21x2)dw1 + (1−r)(s12x1−s22x2)dw2dy=[(x1f2+(1−r)x2f1)+(1−r)22(2(s11s21+s12s22)x1−(s221+s222)x2)] dt + (1−r)(s21x1+s11x2) dw1 + (1−r)(s22x1+s12x2) dw2 . | (91) |
Here fi=fi(x,y) and sij=sij(x,y) are arbitrary smooth functions, and of course the system is written in this rather involved way so that it produces a simple result.
We consider again the function J=r2; it results that on solutions to (91) we have
dJ=[(1−r)J(2f1(x,y)+s211+s212−r(s211+s212))]dt + 2(1−r)Js11 dw1 + 2(1−r)Js12dw2 . |
It is thus immediately apparent that J is in general not an invariant for (91), but that it is always a conditional invariant on the circle r=1. Thus the latter is an invariant manifold for all the equations of the form (91).
In fact, passing to polar coordinates (r,θ), and writing
Fi(r,θ)=fi[x1(r,θ),x2(r,θ)] , Sij(r,θ)=sij[x1(r,θ),x2(r,θ)] , |
the Eq (91) reads simply
dr=r (1−r) [F1(r,θ) dt + S11(r,θ) dw1 + S12(r,θ) dw2]dθ=F2(r,θ) dt + (1−r) [S21(r,θ) dw1 + S22(r,θ) dw2] . | (92) |
Now, unless all the functions Fi and Sij are actually independent of θ, rotations are not a symmetry of the full Eq (92), and hence of the original Eq (91).
On the other hand, the equation reduced to the invariant manifold r=1 reads just
dr = 0 , dθ = F2(1,θ) dt | (93) |
(note this is a deterministic equation), and if F2(1,θ) is independent of θ, it has a rotational symmetry. On the other hand, if F2(1,θ) does actually depend on θ, we do not have a rotational conditional symmetry, despite the presence of the conditional invariant J on r=1. ♢
The reader has surely noted that, albeit we only discussed conditional symmetries and/or invariants in our previous Examples, these are actually tailored so to have asymptotic symmetries and/or invariants. We will look again at them under this point of view, but we will first consider an even simpler one dimensional Example.
Example 9.6. Consider the one-dimensional equation
dx = −(x + x3) dt + σxdw ; | (94) |
in this case the invariant manifold is just the single point x=0; this is of course of little interest in itself.
This equation does not have continuous symmetries (an obvious reflection symmetry (x,t,w)→(−x,t,w) is present), but as the dynamics is attracted towards x=0 we know that asymptotically the system dynamics will be described by its linearization
dx = −x dt + σx dw . | (95) |
This has a scaling symmetry,
X = x ∂∂x := φ(x) ∂∂x ; | (96) |
thus we can just use the prescription discussed in Sect. 5 and pass to the new variable
y = ∫1φ(x) dx = log(x) . |
Note here that we should actually write y=log(|x|) and distinguish the cases x>0 and x<0. This is just fine as x=0 is a barrier for this dynamics. We will then use x>0 (so that in particular ∂y/∂x=1/x), and x=exp[y]. A similar discussion would apply for x<0.
If we apply this transformation to the full equation, we get
dy=1x dx + 12Δ(y)dt=1x [−(x+x3) dt + σxdw] − 12σ2dt=−(1+x2+σ2/2)dt + σdw=−[(1+σ2/2)+e2y]dt + σdw . |
Now we should observe that we are interested in the region x≈0, i.e., y≈−∞; in this limit the term e2y can be discarded. Actually, the y dynamics will drive the system towards this limit. Thus in this asymptotic regime the dynamics is just governed by
dy = −(1+σ2/2)dt + σdw ; |
it is clear that this equation is readily integrated to
y(t) = y(t0) + (1+σ2/2)(t−t0) + [w(t)−w(t0)] , |
and also that it admits ∂y as a symmetry. ♢
Example 9.7. Consider again the system of Example 9.2, i.e., (86), which we rewrite here for ease of reference:
dρ=a(1−ρ2)ρdt + σ(1−ρ2)dw1dϑ=[b(1−ρ2) + ωρ2] dt + σdw2 . | (97) |
The equation for ρ is autonomous, and we readily note that ρ=1 is an invariant manifold, and actually for a>0 a weakly attractive one.
It appears from (97) that – at least for |σ|≪1, i.e. when the drift overcomes the diffusion term – the dynamics converges to the circle ρ=1. Note this will be a weakly attracting manifold. We will thus linearize (97) around ρ=1; we do this by writing
ρ=1+η |
and taking only first order terms in η. This yields
dη=−2aηdt − 2σηdw1dϑ=[ω + 2η(ω−b)]dt + σdw2 . | (98) |
The dynamics is attracted to η=0 (corresponding to ρ=1) and on this it is described by dϑ=ωdt+σdw2, while in the region around the attracting manifold it is described by (98).
The original Eq (97) is rotationally invariant; this symmetry is also present in the reduced Eq (98).
Note also that the first equation in (98) admits X=η∂η as a (scaling) symmetry; thus it can be integrated (passing to the variable y=log(ρ)=log(1+η)), which yields η=η(t) as an explicit function, depending on the realization of the Wiener process w1(t). With this, the second equation in (98) is also readily integrated.
It may be interesting to see what happens if we proceed in a slightly different way. We can perform the change of variables y=log(ρ) directly on the original equation (97); this yields in mixed notation (obviously we have a singularity at ρ=0) for the first equation
dy = [(a−σ22ρ2+σ22)(1−ρ2)] dt + σ1−ρ2ρ dw1 |
Passing now to write ρ=1+η as above, the full equation reads
dy=− [η(2+η)(a−η(2+η)σ22(1+η)2)] dt − ση2+η1+η dw1 ; |
expanding in series at first order in η, we get y=log(1+η)≃η, hence dy≃dη, and
dη = −2aηdt − 2σηdw1 , |
where |σ|≪1 has been taken into account, exactly as before. ♢
Example 9.8. Consider the equation
dx1=[(1−r)(ax1 − bx2) − ω x2 − (α2+β2)(x1/2)] dt − αx2dw1 − βx2dw2dx2=[(1−r)(bx1 + ax2) + ωx1 − (α2+β2)(x2/2)] dt + αx1dw1 + βx1dw2 . | (99) |
Here a,b are real constants, α,β,ω are arbitrary smooth functions of x1,x2. We assume a>0 for definiteness.
In this case the evolution of J=r2 is given by
dJ = 2a r2(1−r) dt ; |
note this is a deterministic equation, so that any initial datum with r≠0 is attracted by the circle r=1. In this case, we have an invariant and strongly attractive submanifold. Correspondingly, J is a strongly asymptotic invariant.
Passing to polar coordinates, we have
dr=a r (1−r) dtdθ=[b(1−r) + ω] dt + α dw1 + β dw2 . |
Needless to say, the functions α,β,ω should now be thought as functions of (r,θ); we will denote by ˆα,ˆβ,ˆω the restrictions of these to the circle r=1. On the invariant strongly attractive submanifold r=1 the evolution is described by
dθ = ˆω(θ) dt + ˆα(θ) dw1 + ˆβ(θ) dw2 . | (100) |
It is rather clear that choosing suitably the functions α,β,ω we can have different situations:
● If these do not depend on θ, the full Eq (99) is rotationally invariant, i.e., X=−x2∂1+x1∂2 is a full symmetry of (99);
● If these do depend on θ, but their restrictions ˆα,ˆβ,ˆω do not depend on θ, then (99) does not admit X as a full symmetry, but admits it as a conditional symmetry on the invariant (and attractive) manifold M0 identified by r=1;
● If (at least one of) the ˆα,ˆβ,ˆω depend on θ, then (99) does not even admit X as a conditional symmetry.
It is maybe worth stressing again that in all these cases J is a strongly asymptotic invariant. ♢
In many physically interesting situations, one observes that equations not enjoying a given symmetry can have special solutions which are symmetric. Even more frequently, the asymptotic behavior of the system under study is characterized in terms of symmetry, independently of the symmetry of initial conditions; typically this involves a simple, i.e., a rotational or a scaling symmetry. A theory of conditional and asymptotic symmetries for deterministic equations has been developed in the literature.
In the present paper, we have extended this theory to the case of stochastic differential equations, considering also the interplay between (asymptotic) symmetry and (asymptotic) integrability for stochastic equations. A special role in this frame is played by (full or conditional or asymptotic) invariants for stochastic equations. We have also shown by concrete (simple) examples how the theory developed here can be applied in practice.
Our examples show that the interrelation between the presence of (conditional and/or asymptotic) symmetry and invariants is not unique: the presence of conditional of asymptotic configurational invariants is needed to have a conditional or asymptotic symmetries, but the converse is not true.
Our discussion shows that the pre-existing theory for deterministic equations can be fully extended to stochastic ones, and the examples show that it can be applied with a well definite algorithm and standard computations; these were specially simple in the considered examples – which were built precisely in order to show the conceptual issues without blurring them by computational details – but our discussion shows that even in more involved cases (apart from computational complications) the procedure is rather straightforward.
Needless to say, its success depends on the equations under study possessing the required (full, conditional or asymptotic) symmetry properties. Symmetry is a non-generic property; but conditional or even more asymptotic symmetry is more common than full symmetry, so we are confident that our method can be widely applicable, as it already happens in the context of deterministic equations.
Our approach was focusing – again, as in the deterministic case – on continuous (Lie) symmetries. At difference with the deterministic case, we have dealt only with ordinary stochastic differential equations, and not discussed partial stochastic differential equations. The reason for this is that a symmetry theory for these equations is not as well developed as for ordinary ones; we expect that once this (substantial) obstacle is removed, the theory can be extended from full to conditional and asymptotic symmetries as in this work. In view of the relevance of conditional and asymptotic symmetries for the study of deterministic PDEs, we expect this extension to stochastic PDEs to be specially relevant for applications, and we hope it may soon be obtained.
Finally, we would like to note that in a way the approach pursued in the literature dealing with conditional or asymptotic symmetries is inverse to that of the well known symmetry breaking approach. In that case, attention is focused on the fact that equations possessing a given symmetry G can have solutions which are invariant only under a subgroup G0⊂G (possibly with G0={I}); here instead one is focusing on the fact that an equation possessing a symmetry G0 (possibly with G0={I}) can have solutions with symmetry G⊃G0. Similarly, for an equation with symmetry G0 it may happen that the asymptotic dynamics is described by an effective equation with symmetry G⊃G0, and this is turn may have solutions with symmetry G or even ˜G⊃G⊃G0.
The work of GG is also supported by SMRI, which he was visiting while the first draft of this paper was prepared and again at the time of final writing. The work of FS is supported by ERC SG CONSTAMIS grant. GG and FS are also members of GNFM-INdAM. We thank these Institutions for their support.
The authors declare no conflict of interest.
[1] | G. Birkhoff, Hydrodynamics, Princeton UP, 1960. |
[2] | L. V. Ovsjiannikov, Group properties of differential equations, Novosibirsk, 1962. |
[3] | D. V. Alexseevsky, A. M. Vinogradov, V. V. Lychagin, Basic ideas and concepts of differential geometry, Springer, 1991. |
[4] | G. Cicogna, G. Gaeta, Symmetry and perturbation theory in nonlinear dynamics, Springer, 1999. |
[5] | I. S. Krasil'schik, A. M. Vinogradov, Symmetries and conservation laws for differential equations of mathematical physics, A.M.S., 1999. |
[6] | P. J. Olver, Application of Lie groups to differential equations, Springer, 1986. |
[7] | P. J. Olver, Equivalence, invariants and symmetry, Cambridge University Press, 1995. |
[8] | H. Stephani, Differential equations. Their solution using symmetries, Cambridge University Press, 1989. |
[9] | P. Winternitz, What is new in the study of differential equations by symmetry methods?, In: Group theoretical methods in physics (Proceedings XV ICGTMP), World Scientific, 1987. |
[10] |
D. Levi, P. Winternitz, Non-classical symmetry reduction: example of the Boussinesq equation, J. Phys. A, 22 (1989), 2915–2924. doi: 10.1088/0305-4470/22/15/010
![]() |
[11] |
G. Cicogna, G. Gaeta, Partial Lie-point symmetries of differential equations, J. Phys. A, 34 (2001), 491–512. doi: 10.1088/0305-4470/34/3/312
![]() |
[12] |
G. Gaeta, Asymptotic symmetries and asymptotically symmetric solutions of partial differential equations, J. Phys. A, 27 (1994), 437–451. doi: 10.1088/0305-4470/27/2/027
![]() |
[13] |
G. Gaeta, N. Rodríguez-Quintero, Lie-point symmetries and stochastic differential equations, J. Phys. A, 32 (1999), 8485–8505. doi: 10.1088/0305-4470/32/48/310
![]() |
[14] |
G. Gaeta, Lie-point symmetries and stochastic differential equations II, J. Phys. A, 33 (2000), 4883–4902. doi: 10.1088/0305-4470/33/27/306
![]() |
[15] |
G. Unal, Symmetries of Ito and Stratonovich dynamical systems and their conserved quantities, Nonlin. Dyn., 32 (2003), 417–426. doi: 10.1023/A:1025669920594
![]() |
[16] |
R. Kozlov, Symmetry of systems of stochastic differential equations with diffusion matrices of full rank, J. Phys. A, 43 (2010), 245201. doi: 10.1088/1751-8113/43/24/245201
![]() |
[17] |
R. Kozlov, The group classification of a scalar stochastic differential equations, J. Phys. A, 43 (2010), 055202. doi: 10.1088/1751-8113/43/5/055202
![]() |
[18] |
R. Kozlov, On maximal Lie point symmetry groups admitted by scalar stochastic differential equations, J. Phys. A, 44 (2011), 205202. doi: 10.1088/1751-8113/44/20/205202
![]() |
[19] |
R. Kozlov, On symmetries of stochastic differential equations, Commun. Nonlinear. Sci., 17 (2012), 4947–4951. doi: 10.1016/j.cnsns.2012.05.019
![]() |
[20] |
G. Gaeta, F. Spadaro, Random Lie-point symmetries of stochastic differential equations, J. Math. Phys., 58 (2017), 053503. doi: 10.1063/1.4982639
![]() |
[21] |
G. Gaeta, F. Spadaro, Erratum: "Random Lie-point symmetries of stochastic differential equations" [J. Math. Phys. 58, 053503 (2017)], J. Math. Phys., 58 (2017), 129901. doi: 10.1063/1.5012089
![]() |
[22] |
G. Gaeta, Symmetry of stochastic non-variational differential equations, Phys. Rep., 686 (2017), 1–62. doi: 10.1016/j.physrep.2017.05.005
![]() |
[23] |
G. Gaeta, Erratum for "Symmetry of stochastic non-variational differential equations" [Phys. Rep. 686 (2017) 1–62], Phys. Rep., 713 (2017), 18. doi: 10.1016/j.physrep.2017.11.002
![]() |
[24] | G. Gaeta, C. Lunini, On Lie-point symmetries for Ito stochastic differential equations, J. Nonlinear Math. Phys., 24 (2017), 90–102. |
[25] | G. Gaeta, C. Lunini, Symmetry and integrability for stochastic differential equations, J. Nonlinear Math. Phys., 25 (2018), 262–289. |
[26] |
R. Kozlov, Random Lie symmetries of Ito stochastic differential equations, J. Phys. A, 51 (2018), 305203. doi: 10.1088/1751-8121/aac95a
![]() |
[27] |
R. Kozlov, Lie-point symmetries of Stratonovich stochastic differential equations, J. Phys. A, 51 (2018), 505201. doi: 10.1088/1751-8121/aae753
![]() |
[28] | R. Kozlov, Symmetries of Ito stochastic differential equations and their applications, In: Nonlinear systems and their remarkable mathematical structures, CRC Press, 2018. |
[29] | G. Gaeta, C. Lunini, F. Spadaro, Recent advances in symmetry of stochastic differential equations, Rend. Mat. Appl., 39 (2018), 293–306. |
[30] |
G. Gaeta, W-Symmetries of Ito stochastic differential equations, J. Math. Phys., 60 (2019), 053501. doi: 10.1063/1.5080434
![]() |
[31] | L. Arnold, Random dynamical systems, Springer, 1988. |
[32] | L. C. Evans, An introduction to stochastic differential equations, A.M.S., 2013. |
[33] | D. Freedman, Brownian motion and diffusion, Springer, 1983. |
[34] | N. Ikeda, S. Watanabe, Stochastic differential equations and diffusion processes, North Holland, 1981. |
[35] | N. G. van Kampen, Stochastic processes in physics and chemistry, North Holland: Elsevier, 1992. |
[36] | B. Oksendal, Stochastic differential equations, 4 Eds., Springer, 1985. |
[37] | D. W. Stroock, Markov processes from K.Ito's perspective, Princeton UP, 2003. |
[38] | S. Walcher, Orbital symmetries of first order ODEs, In: Symmetry and Perturbation Theory (SPT98), World Scientific, 1999, 96–113. |
[39] |
M. K. Kinyon, S. Walcher, On ordinary differential equations admitting a finite linear group of symmetries, J. Math. Anal. Appl., 216 (1997), 180–196. doi: 10.1006/jmaa.1997.5668
![]() |
[40] | S. Walcher, Multi-parameter symmetries of first order ordinary differential equations, J. Lie Theory, 9 (1999), 249–269. |
[41] | J. Guckenheimer, Ph. Holmes, Nonlinear oscillations, dynamical systems, and bifurcation of vector fields, Springer, 1983. |
[42] | P. Glendinning, Stability, instability and chaos: an introduction to the theory of nonlinear differential equations, Cambridge UP, 1994. |
[43] | G. Iooss, M. Adelmeyer, Topics in bifurcation theory and applications, World Scientific, 1992. |
[44] | F. Verhulst, Nonlinear differential equations and dynamical systems, Springer, 1989. |
[45] | G. Gaeta, Autonomous systems, dynamical systems, LPTI symmetries, topology of trajectories, and periodic solutions, Int. J. Theor. Phys., 32 (1993), 191–199. |
[46] | J. Milnor, Topology from the differentiable viewpoint, Princeton UP, 1997. |
[47] | G. W. Bluman, J. D. Cole, The general similarity solution of the heat equation, J. Math. Mech., 18 (1969), 1025–1042. |
[48] | G. W. Bluman, J. D. Cole, Similarity methods for differential equations, Springer, 1974. |
[49] |
P. A. Clarkson, M. D. Kruskal, New similarity reductions of the Boussinesq equation, J. Math. Phys., 30 (1989), 2201–2213. doi: 10.1063/1.528613
![]() |
[50] |
P. A. Clarkson, E. L. Mansfield, Symmetry reductions and exact solutions of a class of nonlinear heat equations, Physica D, 70 (1994), 250–288. doi: 10.1016/0167-2789(94)90017-5
![]() |
[51] |
G. Gaeta, On the conditional symmetries of Levi and Winternitz, J. Phys. A, 23 (1990), 3643–3645. doi: 10.1088/0305-4470/23/15/033
![]() |
[52] |
E. Pucci, G. Saccomandi, On the weak symmetry group of partial differential equations, J. Math. Anal. Appl., 163 (1992), 588–598. doi: 10.1016/0022-247X(92)90269-J
![]() |
[53] |
E. Pucci, G. Saccomandi, On the reduction methods for ordinary differential equations, J. Phys. A, 35 (2002), 6145–6155. doi: 10.1088/0305-4470/35/29/314
![]() |
[54] |
R. Z. Zhdanov, Conditional Lie-Backlund symmetry and reduction of evolution equations, J. Phys. A, 28 (1995), 3841–3850. doi: 10.1088/0305-4470/28/13/027
![]() |
[55] | D. Levi, M. A. Rodríguez, Z. Thomova, Differential equations invariant under conditional symmetries, J. Nonlinear Math. Phys., 26 (2019), 281–293. |
[56] |
E. Pucci, G. Saccomandi, Partial differential equations admitting a given non-classical point symmetry, Stud. Appl. Math., 145 (2020), 81–96. doi: 10.1111/sapm.12312
![]() |
[57] |
E. Pucci, G. Saccomandi, Using symmetries à rebours, Stud. Appl. Math., 146 (2021), 99–117. doi: 10.1111/sapm.12342
![]() |
[58] |
R. B. Gardner, N. Kamran, Characteristics and the geometry of hyperbolic equations in the plane, J. Differ. Equations, 104 (1993), 60–116. doi: 10.1006/jdeq.1993.1064
![]() |
[59] | G. B. Whitham, Linear and nonlinear waves, Wiley, 2011. |
[60] | N. Golfenfeld, O. Martin, Y. Oono, F. Liu, Anomalous dimensions and the renormalization group in a non-linear diffusion process, Phys. Rev. Lett., 65 (1990), 1361–1364. |
[61] |
J. Bricmont, A. Kupiainen, Renormalizaion group and the Ginzburg-Landau equation, Commun. Math. Phys., 150 (1992), 193–208. doi: 10.1007/BF02096573
![]() |
[62] |
G. Gaeta, R. Mancinelli, Asymptotic symmetries of anomalous reaction-diffusion equations, J. Nonlinear. Math. Phys., 12 (2005), 550–566. doi: 10.2991/jnmp.2005.12.4.9
![]() |
[63] | G. Gaeta, R. Mancinelli, Asymptotic scaling symmetries for nonlinear PDEs, Int. J. Geom. Methods. M., 2 (2005), 1081–1114. |
[64] | J. D. Murray, Mathematical biology I: an introduction, Springer, 2002. |
[65] | J. D. Murray, Mathematical biology II: spatial models and biomedical applications, Springer, 2003. |
[66] |
F. C. De Vecchi, P. Morando, S. Ugolini, Symmetries of stochastic differential equations using Girsanov transformations, J. Phys. A: Math. Theor., 53 (2020), 135204. doi: 10.1088/1751-8121/ab757d
![]() |
[67] |
F. C. De Vecchi, P. Morando, S. Ugolini, Reduction and reconstruction of SDEs via Girsanov and quasi Doob symmetries, J. Phys. A, 54 (2021), 185203. doi: 10.1088/1751-8121/abef7f
![]() |
[68] |
F. C. De Vecchi, P. Morando, S. Ugolini, Symmetries of stochastic differential equations: A geometric approach, J. Math. Phys., 57 (2016), 063504. doi: 10.1063/1.4953374
![]() |
[69] |
F. C. De Vecchi, P. Morando, S. Ugolini, Reduction and reconstruction of stochastic differential equations via symmetries, J. Math. Phys., 57 (2016), 123508. doi: 10.1063/1.4973197
![]() |
[70] |
F. C. De Vecchi, P. Morando, S. Ugolini, A note on symmetries of diffusions within a martingale problem approach, Stoch. Dynam., 19 (2019), 1950011. doi: 10.1142/S0219493719500114
![]() |
[71] | S. Albeverio, F. C. De Vecchi, P. Morando, S. Ugolini, Weak symmetries of stochastic differential equations driven by semimartingales with jumps, Electron. J. Probab., 25 (2020), 1–34. |
[72] | F. C. De Vecchi, E. Mastrogiacomo, M. Turra, S. Ugolini, Noether theorem in stochastic optimal control problems via contact symmetries, 2021, arXiv: 2102.03172. |
[73] |
L. Arnold, P. Imkeller, Normal forms for stochastic differential equations, Probab. Theory Rel., 110 (1998), 559–588. doi: 10.1007/s004400050159
![]() |
[74] |
S. C. Anco, B. Wang, Geometrical formulation for adjoint-symmetries of Partial differential equations, Symmetry, 12 (2020), 1547. doi: 10.3390/sym12111840
![]() |
[75] | G. Gaeta, Integration of the stochastic logistic equation via symmetry analysis, J. Nonlinear Math. Phys., 26 (2019), 454–467. |
[76] |
S. Albeverio, S. M. Fei, A remark on symmetry of stochastic dynamical systems and their conserved quantities, J. Phys. A, 28 (1995), 6363–6371. doi: 10.1088/0305-4470/28/22/012
![]() |
[77] |
T. Misawa, Noether's theorem in symmetric stochastic calculus of variations, J. Math. Phys., 29 (1988), 2178–2180. doi: 10.1063/1.528145
![]() |
[78] |
T. Misawa, Conserved quantities and symmetry for stochastic dynamical systems, Phys. Lett. A, 195 (1994), 185–189. doi: 10.1016/0375-9601(94)90150-3
![]() |
[79] |
T. Misawa, New conserved quantities derived from symmetry for stochastic dynamical systems, J. Phys. A, 27 (1994), L777–L782. doi: 10.1088/0305-4470/27/20/004
![]() |
[80] |
W. Sarlet, P. G. L. Leach, F. Cantrijn, First integrals versus configurational invariants and a weak form of complete integrability, Physica D, 17 (1985), 87–98. doi: 10.1016/0167-2789(85)90136-8
![]() |
[81] |
K. Rosquist, G. Pucacco, Invariants at fixed and arbitrary energy. A unified geometric approach, J. Phys. A, 28 (1995), 3235–3252. doi: 10.1088/0305-4470/28/11/021
![]() |
[82] |
G. Pucacco, K. Rosquist, Configurational invariants of Hamiltonian systems, J. Math. Phys., 46 (2005), 052902. doi: 10.1063/1.1888565
![]() |
[83] |
L. S. Hall, Invariants polynomial in momenta for integrable Hamiltonians, Phys. Rev. Lett., 54 (1985), 614–615. doi: 10.1103/PhysRevLett.54.614
![]() |
[84] |
L. S. Hall, A theory of exact and approximate configurational invariants, Physica D, 8 (1983), 90–116. doi: 10.1016/0167-2789(83)90312-3
![]() |
[85] | H. J. Kushner, Stochastic stability and control, Academic Press, 1967. |
[86] | R. Z. Khas'minskii, Stochastic stability of differential equations, Noordhoff, 1980. |
[87] |
T. S. Doan, M. Engel, J. S. Lamb, M. Rasmussen, Hopf bifurcation with additive noise, Nonlinearity, 31 (2018), 4567–4601. doi: 10.1088/1361-6544/aad208
![]() |
1. | G. Gaeta, On the integration of Ito equations with a random or a W-symmetry, 2023, 64, 0022-2488, 10.1063/5.0141333 | |
2. | Moh Shahid Khan, Ravi Kumar Mandava, Design of dynamically balanced gait for the biped robot while crossing the obstacle, 2024, 238, 0954-4062, 9125, 10.1177/09544062241246878 | |
3. | Giuseppe Gaeta, Miguel Angel Rodriguez, Integrable Ito equations with multiple noises, 2022, Volume 2, 2802-9356, 10.46298/ocnmp.10275 | |
4. | Giuseppe Gaeta, Miguel Angel Rodriguez, Symmetry classification of scalar autonomous Ito stochastic differential equations with simple noise, 2022, Volume 2, 2802-9356, 10.46298/ocnmp.9770 | |
5. | Juan Arturo Alvarez-Valdez, Guillermo Fernandez-Anaya, Roadmap of the Multiplier Method for Partial Differential Equations, 2023, 11, 2227-7390, 4572, 10.3390/math11224572 |