Next Article in Journal
MTLBORKS-CNN: An Innovative Approach for Automated Convolutional Neural Network Design for Image Classification
Previous Article in Journal
Asymptotic Chow Semistability Implies Ding Polystability for Gorenstein Toric Fano Varieties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

On the Properties of λ-Prolongations and λ-Symmetries

1
School of Applied Mathematics, Jilin University of Finance and Economics, Changchun 130117, China
2
School of Mathematics, Jilin University, Changchun 130012, China
*
Author to whom correspondence should be addressed.
Mathematics 2023, 11(19), 4113; https://doi.org/10.3390/math11194113
Submission received: 7 August 2023 / Revised: 17 September 2023 / Accepted: 25 September 2023 / Published: 28 September 2023

Abstract

:
In this paper, (1) We show that if there are not enough symmetries and λ -symmetries, some first integrals can still be obtained. And we give two examples to illustrate this theorem. (2) We prove that when X is a λ -symmetry of differential equation field Γ , by multiplying Γ a function μ defineded on J n 1 M , the vector fields μ Γ can pass to quotient manifold Q by a group action of λ -symmetry X. (3) If there are some λ -symmetries of equation considered, we show that the vector fields from their linear combination are symmetries of the equation under some conditions. And if we have vector field X defined on J n 1 M with first-order λ -prolongation Y and first-order standard prolongations Z of X defined on J n M , we prove that g Y cannot be first-order λ -prolonged vector field of vector field g X if g is not a constant function. (4) We provide a complete set of functionally independent ( n 1 ) order invariants for V ( n 1 ) which are n 1 th prolongation of λ -symmetry of V and get an explicit n 1 order reduced equation of explicit n order ordinary equation considered. (5) Assume there is a set of vector fields X i , i = 1 , . . . , n that are in involution, We claim that under some conditions, their λ -prolongations also in involution.
MSC:
34C14

1. Introduction

Classical symmetry theory for general systems of differential equations was created by Sophus Lie more than 100 years ago. It plays an important role in analyzing differential equations. The geometric theory of symmetries makes it possible to understand and generalize the standard procedures of the explicit integration of ODE and to obtain new results. They are often used to reduce the order of ordinary differential equations and to reduce the number of independent variables of partial differential equations. The method of symmetry reduction for ODEs can be roughly described as follows: in J n M , which is the jet bundle of order n over the manifold M of independent and dependent variables, if ODEs E of order n > 1 have a vector field X (on M) as a symmetry (generator), then E is invariant under Y = X ( n ) , which is the prolongation of X to J n M . Thus, E can be expressed in terms of the differential invariants of Y , which can be recursively generated starting from those of orders 0 and 1 thanks to the “invariants by differentiation” property (see [1,2,3,4,5,6,7]).
Although the Lie symmetry theory is a good method for analyzing ordinary (and partial) differential equations [4,5], not every technique can be based on symmetry analysis [8,9,10,11], thus people need generalizations of classical Lie methods. In past decades, λ -symmetry was introduced by Muriel and Romero [12] based on a new method of prolonging vector fields known as λ -prolongation, which strictly include symmetries. Generally speaking, λ -symmetries may not be symmetries, as they do not map solutions into solutions. The relationship between λ -symmetries and the first integral of ordinary differential equations has been studied extensively [13,14,15,16,17,18,19]. Moreover, λ -symmetries have been extended to σ -symmetries [20,21] and to μ -symmetries [22,23], which are different from other analytical methods (such as the Kudryashov approach; see [24]) and can be used to derive conservation laws of partial differential equations (PDEs). And the analogous theory of symmetries of stochastic differential equations (SDEs) has also emerged in recent years (see [25,26,27,28,29,30]). In a review of the group analysis methods, symmetries for discrete equations were shown as well. Among the most complete and noteworthy works on this subject at the moment, one can refer to Refs. [31,32]. In [18], the authors considered that if there are not enough symmetries for ODE ( 4 ) , one can use λ -symmetries to obtain some first integrals. Here, we generalize this situation and show that if there are not enough symmetries and λ -symmetries, some first integrals can still be obtained (see Theorem 2). Moreover, we prove that when X is a λ -symmetry of the differential equation field Γ of ( 4 ) , by multiplying Γ , a function μ defined on J n 1 M , the vector fields μ Γ can pass to quotient manifold Q by a group action of λ -symmetry X, i.e., L X ( μ Γ ) = λ ¯ X (see Theorem 3). If there are some λ -symmetries of ( 4 ) , we show that the vector fields from their linear combination are symmetries of ( 4 ) under some conditions (see Theorem 4). And if we have vector field X defined on J n 1 M with first-order λ -prolongation Y and first-order standard prolongations Z of X defined on J n M , we prove that g Y cannot be a first-order λ -prolonged vector field of vector field g X if g is not a constant function (see Theorem 5). We provide a complete set of functionally independent ( n 1 ) -order invariants for V ( n 1 ) , which are ( n 1 )th prolongations of λ -symmetry of V (see Theorem 6) and obtain an explicit n 1 -order-reduced equation of explicit n-order ordinary Equation ( 4 ) . Compared with [12], we deal with an n-order explicit equation system that has a λ -symmetry V , and we just need to consider the invariants of ( n 1 ) -order prolongation V ( n 1 ) of V , but in [12], the authors dealt with implicit equation with a λ -symmetry V , which need to consider the invariants of n-order prolongation V ( n 1 ) of V . At last, we consider that there is a set of vector fields X i , i = 1 , . . . , n that are in involution and claim that under some conditions their λ -prolongations also in involution (see Theorem 7).
This paper is organized as follows. In Section 2, we recall some basic notations we need from the Lie group theory and give some definitions of the λ -symmetries of vector fields, differential forms, distributions and λ -prolongations. In Section 3, we obtain the first integrals of a system of ordinary differential equations by using λ -symmetries. In Section 4, we obtain a quotient manifold and discuss the relationship between λ -symmetry and symmetry. In Section 5, we obtain an explicit n 1 -order-reduced equation of the explicit n-order ordinary equation by using a λ -symmetry. In Section 6, we discuss the involution of λ -prolonged sets of vector fields.

2. Preliminaries and Geometric Properties for λ -Symmetries

Firstly, we give some basic notions that may be used later.

2.1. Equation, Solutions and Symmetries

In this paper, we only consider ordinary differential equations that have the independent variable x R and the dependent one(s) y U = R or y a U R m in the multidimensional case. Let M = X × U be a phase bundle and ( J k M , π k , M ) , or J k M for short, the associated jet bundle of order k .
A differential equation E of order n is a map F : J n M R and is identified with the solution manifold S ( E ) = F 1 ( 0 ) J n M . In the case of l-dimensional systems E , we have l maps F j : J n M R and a solution manifold S ( E ) = F 1 ( 0 ) = ( F 1 ) 1 ( 0 ) . . . ( F l ) 1 ( 0 ) J k M . If differential equation E has m-dimensional dependent variables, then J n M has dimension ( m ( n + 1 ) + 1 ) , thus the dimension of a solution manifold of a system of l independent equations is ( m ( n + 1 ) + 1 l ) .
For a smooth function vector y = f ( x ) , we can obtain an induced function p r ( n ) f ( x ) , which is called the n-th prolongation of f and defined by the equations
d α y a d x α = d α f a d x α , a = 1 , . . . , m , α = 0 , . . . , n .
Thus, p r ( n ) f ( x ) is a function vector from R 1 to the space J n M .
For the differential equation(s) E under study, a function vector f : R 1 U is a solution if and only if its n-th prolongation lies entirely in the solution manifold S ( E ) .
Now, consider a vector field Y on J n M . We say that E is invariant under Y if and only if its solution manifold has the property Y : S ( E ) T S ( E ) . This is also equivalent to condition [ Y ( E ) ] S ( E ) = 0 .
Let Y be the prolongation of a (Lie-point) vector field X on M, i.e., Y = X ( n ) , then X is called a symmetry for E (more precisely, X would be a symmetry generator. We will adopt this standard abuse of notation for ease of language). Thus, X is a symmetry if [ X ( n ) ( E ) ] S ( E ) = 0 .

2.2. Local Coordinates

We will consider local coordinates ( x , y a ) , a = 1 , . . . , m in M = X × U , and correspondingly local coordinates ( x , y a ( k ) ) (with k = 0 , . . . , n ), where y a ( k ) : = ( k y a / x k ) , in J n M .
A general vector field defined on J n M will be written in local coordinates (in this paper, we will use the Einstein summation convention in some places. The notation y ( k ) denotes the k-order derivative of y) as
Y = ξ ( x , y 1 , . . . , y m ) x + ψ k a ( x , y 1 , . . . , y m , y 1 ( 1 ) , . . . , y m ( 1 ) , . . . , y m ( n ) ) y a ( k ) ,
which is the prolongation of vector field
X = ξ ( x , y 1 , . . . , y m ) x + ϕ a ( x , y 1 , . . . , y m ) y a
if and only if the ψ k a satisfies the (standard) prolongation formula
ψ k + 1 a = D x ψ k a y a ( k + 1 ) D x ξ , ψ 0 a = ϕ a .
We use D x to denote the total derivative with respect to x, i.e.,
D x = x + y a ( 1 ) y a + y a ( 2 ) y a ( 1 ) .

2.3. λ -Prolongations

After the work of Muriel and Romero [12,33], people also considered λ -symmetries of ODEs. A vector field Y written in local coordinates in the form ( 1 ) is a λ -prolonged vector field if its coefficients satisfy
ψ k + 1 a = D x ψ k a y a ( k + 1 ) D x ξ + λ ( ψ k a y a ( k + 1 ) ξ )
with λ being a smooth function λ : J 1 M R . We also say that Y is the nth λ -prolongation of X (see ( 2 ) ) if ψ 0 a = ϕ a , i.e., if X is the restriction of Y to M .
We say that X is a λ -symmetry for E if the λ -prolongation Y of X leaves an equation (or system) E invariant.

2.4. Explicit n-Order Equation

Consider the m-dimension system of nth-order differential equations
d n y a d x n = f a ( x , y a , . . . , y a ( n 1 ) )
where ( x , y 1 , . . . , y m ) M R m + 1 denote ( y a ) = ( y 1 , . . . , y m ) , a = 1 , . . . , m .   x , y a , y a ( 1 ) , . . . , y a ( n 1 ) for a = 1 , . . . , m as natural coordinates on the ( n 1 ) th jet bundle J n 1 M . Put σ 1 a = d y a y a ( 1 ) d x , σ j a = d y a ( j 1 ) y a ( j ) d x . Then, { σ 1 a , . . . , σ n 1 a } a = 1 m is the natural basis for the contact forms on J n 1 M , which are the forms on J n 1 M that are zero on curves or surfaces lifted from J 0 M . Let f a 0 J n 1 M , then put σ n a = d y a ( n 1 ) f a d x . Then, the σ n a s are the natural force forms for ( 4 ) . Put
Ω = σ 1 1 σ 1 m σ n 1 σ n m ,
meaning the m n -fold wedge product, then Ω is the natural characteristic form for ( 4 ) . Let
Γ = x + y a ( 1 ) y a + y a ( 2 ) y a ( 1 ) + + f a y a ( n 1 ) ,
then Γ is an ( n 1 ) th-order differential equation field. It is the characteristic vector field of ( 4 ) , tangent to the lifted solution curves, and represents differentiation along the solution curves. For any p-form ω and vector field X, we denote i X ω as the interior product of X and ω [34] (the interior product of the vector field with a differential form reduces the degree of the latter by one), so it is a vector field in the kernel of Ω normalized by i Γ d x = 1 .
Proposition 1.
We can obtain that i Γ σ j a = ( d y a ( j 1 ) y a ( j ) d x ) ( x + y a ( 1 ) y a + y a ( 2 ) y a ( 1 ) + + f a y a ( n 1 ) ) = 0 , a = 1 , . . . , m , j = 1 , . . . , n , so i Γ Ω = 0 and i Γ Ω V = Ω , where Ω V = d x d y 1 d y 2 . . . d y m d y 1 ( 1 ) d y 2 ( 1 ) . . . d y m ( n 1 ) is the volume form of J n 1 M , here J n 1 M is the ( n 1 ) th-order jet bundle.
Through the whole paper, all calculations involving differential forms such as Cartan’s identity, one can see [34].
Let C ( Δ ) : = { ( x , y a , y a ( 1 ) , . . . , y a ( n ) ) | y a ( n ) = f ( x , y a , y a ( 1 ) , . . . , y a ( n 1 ) ) } , We know that a λ -symmetry, or more precisely a Lie-point λ -symmetry generator of ( 4 ) is a vector field X on M such that its n-order λ -prolongation Y satisfies that
Y d n y i d x n f i ( x , y a , . . . , y a ( n 1 ) ) C ( Δ ) = 0 , i = 1 , . . . , m .
Lemma 1.
(1)
We suppose that, for some λ C ( J 1 M ) , the vector field X = ξ ( x , y a ) x + η ( x , y a ) y a is a λ-symmetry of Equation ( ) . Then,
[ Y , Γ ] = λ · Y + μ · Γ
for some μ C ( J 1 M ) , where Y is the n 1 -order λ-prolongation of X, and [ · , · ] denotes the Lie bracket.
(2)
Conversely, if
X = ξ ( x , y a ) x + η a ( x , y a ) y a + η 1 a ( x , y a , y a ( 1 ) ) y a ( 1 ) + + η n 1 a ( x , y a , y a ( 1 ) , . . . , y a ( n 1 ) ) y a ( n 1 )
is a vector field defined on J n 1 M such that
[ X , Γ ] = λ X + μ Γ
for some λ , μ C ( J 1 M ) , then the vector field
v = ξ ( x , y a ) x + η a ( x , y a ) y a ,
defined on M , is a λ-symmetry of the Equation (4), and X is the n 1 -order λ-prolongation of v.
Proof. 
See [18]. □
Definition 1.
A differential p-form is simple or decomposable if it is the wedge product of p 1-forms.
Definition 2.
We call a set of vector fields D defined on manifold M a distribution if it is a vector subspace of T p M , p M , and for any vector fields X , Y D , [ X , Y ] D .
Definition 3.
We say a distribution D is a Frobenius integral if it is rank-invariant, which means that for any point p M , dim D is a constant.
Definition 4.
We say a p-form ω is a Frobenius integrable if its kernel is a Frobenius integrable and of maximal dimension everywhere.
Based on the properties of λ -symmetries for a system, we give three more general definitions of λ -symmetries for vector fields, distributions and differential forms.
Definition 5.
Let X and Y be vector fields, D be a distribution, and Θ be a simple differential form defined on manifold J n 1 M .
1. X is called a λ-symmetry of Y if there exist functions λ C ( J 1 M ) , μ C ( J 1 M ) such that L X Y = [ X , Y ] = λ X + μ Y .
2. X is called a λ-symmetry of D if L X Y λ X D , Y D for some function λ C ( J 1 M ) .
3. X is called a λ-symmetry of Θ if X is a λ-symmetry of kerΘ for some function λ C ( J 1 M ) , where kerΘ is the kernel or characteristic space of Θ defined by ker Θ = { Y : i Y Θ = 0 } .
Remark 1.
A vector field defined on J n 1 M may not be an ( n 1 ) -order prolongation of any vector field defined on M .

3. First Integral of System of Ordinary Differential Equation

Lemma 2.
Given a natural number N , 3 N m n , let span { X N , . . . , X m n , Γ } be a Frobenius integral and X j be a λ j -symmetry of span { X j + 1 , . . . , X m n , Γ } ,   j = 2 , . . . , N 1 such that X 1 , . . . , X n and Γ are linearly independent everywhere. Then, i X j i X m n Ω is a Frobenius integral j = 2 , . . . , N .
Proof. 
Since X j is a λ j -symmetry of span { X j + 1 , . . . , X m n , Γ } , the Lie bracket on span { X j , . . . , X m n , Γ } = ker ( i X j i X m n Ω ) is closed. Then, i X j i X m n Ω is Frobenius integral j = 2 , . . . , N 1 .
Lemma 3.
If X j , j = 1 , . . . , m n , are vector fields and X 1 , . . . , X m n , Γ are linearly independent, then i X 1 i X m n Ω 0 .
Proof. 
Since X 1 , . . . , X m n , Γ are linearly independent, we have i X j 0 . If i X 1 i X m n Ω = 0 , then ker ( i X 2 i X m n Ω ) = span { X 1 , . . . , X m n , Γ } . On each point, p of M ( n 1 ) , T p M ( n 1 ) = span { X 1 , . . . , X m n , Γ } | p , thus i X 2 i X m n Ω = 0 , so we know ker ( i X 3 i X m n Ω ) span { X 2 , . . . , X m n , Γ } . On the other hand, i X 1 i X m n Ω = 0 = i X 2 i X 1 i X 3 i X m n , ker ( i X 1 i X 3 i X m n Ω ) = span { X 1 , . . . , X m n , Γ } , thus i X 2 i X m n Ω = 0 , so we know ker ( i X 3 i X m n Ω ) span { X 1 , X 3 , . . . , X m n , Γ } . Hence, ker ( i X 3 i X m n Ω ) = span { X 1 , . . . , X m n , Γ } , and we obtain i X 3 i X m n Ω = 0 . Clearly the process is inductive, and finally, we can obtain that i X m n Ω = 0 , which contradicts the assumption, and the proof is complete. □
In [18], Zhang showed that λ -symmetries can be used to obtain some first integrals if there are not enough symmetries, i.e., the following Theorem 1.
Theorem 1
([18]). Given natural numbers N and k , 3 N m n , 1 k N 1 , suppose that
(1) span { X N , . . . , X m n , Γ } is Frobenius integrable;
(2) X j are λ-symmetries of span { X j + 1 , . . . , X N 1 , X N , . . . , X m n , Γ } , j = k + 1 , . . . , N 1 ;
(3) X l is a symmetry of span { X 1 , . . . , X l 1 , X l + 1 , . . . , X m n , Γ } , l = 1 , . . . , k ;
(4) X 1 , . . . , X m n and Γ are linearly independent everywhere.
Put
σ l = i X 1 i X l 1 i X l + 1 i X m n Ω , l = 1 , . . . , k .
Then
ω l = σ l i X l σ l , l = 1 , . . . , k ,
are closed, and locally provide k functionally independent first integrals of ( 4 ) .
Proof. 
From Lemma 2, we know that σ l is a Frobenius integrable 1-form, because X l is a symmetry of span { X 1 , . . . , X ^ l , . . . , X k + 1 , . . . , X m n , Γ } = ker σ i , l = 1 , . . . , k . This implies that X i is a symmetry of σ i . By [35], we know that for a Frobenius integrable 1-form θ , which is nowhere zero, if it has a symmetry Z ker θ , then locally, there is an integrating factor 1 i Z Ω such that θ i Z Ω is a closed 1-form; therefore, ω l . So, by the converse of the Poincaré Lemma, locally, ω l = d I l , where I 1 , . . . , I k are functions. Next, we prove that I l , l = 1 , . . . , k are functionally independent first integrals of ( 4 ) . Due to ker ω l ker ω j as l j , where l , j = 1 , . . . , k , we obtain that ω l , l = 1 , . . . , k are linearly independent, which implies that I l , l = 1 , . . . , k are linearly independent. Moreover, by i Γ Ω = 0 , we know that Γ ( I l ) = i Γ d I l = i Γ i X 1 i X l 1 i X l + 1 i X m n Ω i X l i X 1 i X l 1 i X l + 1 i X m n Ω = 0 , so I l , l = 1 , . . . , k are functionally independent first integrals of ( 4 ) .
Here, we generalize the theorem above in [18] and show that if there are not enough symmetries and λ -symmetries, some first integrals can still be obtained.
Theorem 2.
Given natural numbers N 1 , N 2 and k , 3 N 2 < N 1 m n , 1 k N 1 , suppose that
(1) Span { X N 1 , . . . , X m n , Γ } is Frobenius integrable;
(2) Vector fields X j , N 2 j N 1 1 satisfy
[ X j , X k ] s p a n { X N 2 , . . . , X m n , Γ } , N 2 k m n ;
(3) X i are λ-symmetries of span { X i + 1 , . . . , X m n , Γ } , i = k + 1 , . . . , N 2 1 ;
(3) X l is a symmetry of span { X 1 , . . . , X l 1 , X l + 1 , . . . , X m n , Γ } , l = 1 , . . . , k ;
(4) X 1 , . . . , X m n and Γ are linearly independent everywhere.
Put
σ l = i X 1 i X l 1 i X l + 1 i X m n Ω , l = 1 , . . . , k .
Then
ω l = σ l i X l σ l , l = 1 , . . . , k ,
are closed, and locally provide k functionally independent first integrals of ( 4 ) .
Proof. 
Similar to Theorem 1. □
Here, we give two examples to illustrate the utility of Theorem 2.
Example 1.
Consider the following second-order equation
y = x 2 4 y 3 + y + 1 2 y .
In [12], the authors pointed out this equation has no nontrivial symmetries, and they give a λ-symmetry X = y y , λ = x / y 2 . Now we know that Γ = x + y y x 2 4 y 3 + y + 1 2 y y , Ω = d y d y y d x d p x 2 4 y 3 + y + 1 2 y d x d y . We can calculate W = c y y + c ( y + x y ) y is a λ-symmetry of ( 7 ) , for λ = x y 2 , c is some constant. It is easy to verify that Y = 2 x 1 2 y y is a symmetry of span { W , Γ } . Then, by Theorem 1 or 2,
ω = i W Ω i Y i W Ω = d x + a r c t a n y y + x 2 y 2
is closed, and I = x + a r c t a n y y + x 2 y 2 is a first integral.
Remark 2.
Although [18] showed that Theorem 2 can be used to obtain the first integral for Example 1, here, we point out we can also use Theorem 2 to obtain the first integral for this equation.
Example 2.
Consider system
y 1 = x y 1 + x 2 y 2 + x 3 y 3 y 2 = 2 x y 1 + 2 x 2 y 2 + 2 x 3 y 3 y 3 = x y 1 + x 2 y 2 + x 3 y 3
so we can obtain Γ = x + ( x y 1 + x 2 y 2 + x 3 y 3 ) y 1 + ( 2 x y 1 + 2 x 2 y 2 + 2 x 3 y 3 ) y 2 + ( x y 1 + x 2 y 2 + x 3 y 3 ) y 3 , and Ω = i Γ d x d y 1 d y 2 d y 3 . We can verify that the vector field set { X 1 = y 1 + y 2 , X 2 = y 2 + y 3 } satisfies
[ X 1 , Γ ] = X 1 Γ Γ X 1 = ( x + x 2 ) X 1 + ( x + x 2 ) X 2 , [ X 2 , Γ ] = X 2 Γ Γ X 2 = ( x 2 + x 3 ) X 1 + ( x 2 + x 3 ) X 2 , [ X 1 , X 2 ] = 0 .
Thus, X 1 and X 2 are not the λ-symmetries of ( 8 ) . We can also verify that X 3 = y 1 + y 3 is a symmetry of { X 1 , X 2 , Γ } ; in fact, [ X 3 , X 2 ] = [ X 3 , X 1 ] = 0 , [ X 3 , Γ ] = X 3 Γ Γ X 3 = ( x + x 3 ) X 1 + ( x + x 3 ) X 2 . Then, by using Theorem 2, we can obtain a first integral of ( 8 ) . Calculate
i X 2 i X 1 i Γ ( d x d y 1 d y 2 d y 3 ) = d y 3 d y 2 + d y 1 ,
moreover, i X 3 i X 2 i X 1 i Γ ( d x d y 1 d y 2 d y 3 ) = 2 , so i X 2 i X 1 i Γ ( d x d y 1 d y 2 d y 3 ) i X 3 i X 2 i X 1 i Γ ( d x d y 1 d y 2 d y 3 ) = d y 1 y 2 + y 3 2 , and it is obvious d y 1 y 2 + y 3 2 d x = 1 2 ( x y 1 + x 2 y 2 + x 3 y 3 ( 2 x y 1 + 2 x 2 y 2 + 2 x 3 y 3 ) + x y 1 + x 2 y 2 + x 3 y 3 ) = 0 . Thus, y 1 y 2 + y 3 2 is a first integral of ( 8 ) .

4. Quotient Manifold, Relationship between λ -Symmetry and Symmetry

In [4], Olver stated that if X as a vector field defined on J n 1 M has the property [ X , Γ ] = σ X , then X is a symmetry of System ( 4 ) , where Γ is the equation field defined as ( 5 ) , and σ is defined on J n 1 M . Thus, a quotient manifold Q by a group action of symmetry X can be obtained, and we call the equation field Γ to pass to it. Here, we prove that when X is a λ -symmetry of Γ , by multiplying Γ , a function μ defined on J n 1 M , we can obtain a vector field that passes to quotient manifold Q , i.e., we obtain a vector field μ Γ that passes to Q. Moreover, we also discuss the relationship between λ -symmetries and symmetries.
Theorem 3.
Let Γ be the ( n 1 ) th-order equation field of ( 4 ) . Let X be a λ-symmetry of ( 4 ) , and let Γ , X be two linearly independent vector fields. Assume Q is the quotient of J n 1 M by the one-parameter group generated by X . Then, there exists a function μ defined on J n 1 M such that μ Γ passes to Q .
Proof. 
A vector field W passes to Q if and only if L X W = ρ X for some function ρ . On the other hand, we have L X Γ = λ Γ + λ ¯ X , so we look for a multiple μ Γ of Γ such that L X ( μ Γ ) = λ ¯ X . Because L X ( μ Γ ) = ( L X μ ) Γ + μ L X Γ = ( L X μ + λ μ ) Γ + λ ¯ X = ( X ( μ ) + λ μ ) Γ + λ ¯ X and Γ , X are two linearly independent vector fields, μ must satisfy X ( μ ) + μ λ = 0 . If we attempt μ = ( i Γ ω ) 1 in which ω satisfies i Γ ω 0 for some appropriate 1-form ω , we find that if i Γ ( L X ω ) = 0 , i Γ ω 0 and ω ( X ) = 0 then μ = ( i Γ ω ) 1 satisfies the condition X ( μ ) + μ λ = 0 , because
0 = i Γ ( L X ω ) = i Γ ( i X d ω + d i X ω ) = i Γ i X d ω + i Γ d i X ω = d ω ( X , Γ ) + Γ ( ω ( X ) ) = X ( ω ( Γ ) ) ω ( [ X , Y ] ) = X ( ω ( Γ ) ) λ ω ( Γ ) λ ¯ ω ( X ) = X ( ω ( Γ ) ) λ ω ( Γ ) = ( i Γ ω ) 2 X ( ω ( Γ ) ) + λ ω ( Γ ) ( i Γ ω ) 2 = ( i Γ ω ) 2 ( X ( ( i Γ ω ) 1 ) + λ ( i Γ ω ) 1 ) .
For example, if g is a differential invariant of X, so that X ( g ) = 0 , with Γ ( g ) 0 , then we take ω = d g . Clearly, i Γ ( L X ω ) = i Γ ( i X d ω + d i X ω ) = i Γ ( i X ( d ( d g ) ) + d ( i X d g ) ) = i Γ ( d ( i X d g ) ) = i Γ ( d ( X ( g ) ) ) = 0 , so we know Γ ^ / Γ ( g ) passes to the quotient manifold Q. □
Example 3.
As for ordinary equation
d u d x = e v d v d x = 1
we have characteristic vector fields Γ = x + e v u + v . Given vector fields X = x + v and Z = ( 1 + u 2 ) u , let Q be the quotient of R 3 by the one-parameter group generated by X . We can compute that
[ Γ , X ] = Γ X X Γ = Γ + X , [ X , Z ] = 0 ,
thus, Γ does not pass to Q. That is to say X is a λ-symmetry of Γ, and Z passes to Q. According to Theorem 3 , let μ = e v , then [ e v Γ , X ] = e v X , so e v Γ passes to Q . Recall that for vector fields X , Y defined on a manifold M , we say X is a symmetry of Y if we have
[ X , Y ] = g Y ,
where g is a function defined on M .
Lemma 4.
Let X , Γ be defined as Theorem 3 and g be some function defined on J n 1 M . Then, g X is a symmetry of ( 4 ) if and only if
λ g = Γ g .
Proof. 
Because
[ g X , Γ ] = g X Γ Γ ( g X ) = g X Γ Γ ( g ) X g Γ X = g [ X , Γ ] Γ ( g ) X = g ( λ X + λ ¯ Γ ) Γ ( g ) X = ( λ g Γ ( g ) ) X + g λ ¯ Γ ,
we know that if λ g = Γ g , then g X is a symmetry of ( 4 ) and vice versa. □
If there are l λ -symmetries, similar to Lemma 4, we show that under some conditions, the vector fields from their linear combination are symmetries.
Theorem 4.
Let X i be the λ-symmetries of ( 4 ) , i = 1 , . . . , l , Y i : = A i j X j , i , j = 1 , . . . , l , A i j be functions defined on J n 1 M , Γ be the equation field as Lemma 3. Then, Y i are the symmetries of ( 4 ) if and only if
A i j λ j = Γ ( A i j ) .
where λ j are functions defined on J 1 M . Moreover, if g is the common first integral of X i , i = 1 , . . . l , then Y i is the symmetry of g Γ .
Proof. 
These follow from explicit computation. In fact,
[ Y i , Γ ] = [ A i j X j , Γ ] = A i j X j Γ Γ ( A i j ) X j A i j Γ X j = A i j [ X j , Γ ] Γ ( A i j ) X j = A i j ( λ j X j + λ ¯ j Γ ) Γ ( A i j ) X j = ( A i j λ j Γ ( A i j ) ) X j + A i j λ ¯ j Γ ,
which implies (9). Also,
[ Y i , g Γ ] = [ A i j X j , g Γ ] = A i j g X j Γ g Γ ( A i j ) X j g A i j Γ X j = g A i j [ X j , Γ ] g Γ ( A i j ) X j = g A i j ( λ j X j + λ ¯ j Γ ) g Γ ( A i j ) X j = g ( A i j λ j Γ ( A i j ) ) X j + A i j λ ¯ j ( g Γ ) ,
which completes the proof. □
If we have vector field X defined on J n 1 M with first-order λ -prolongation Y defined on J n M , then we want to find the connection between first-order standard prolongations Z of X and their first-order λ -prolongation Y. It clearly suffices to discuss the situation n = 1 .
Lemma 5.
Let X be a vector field in M and Y be its first-order λ-prolongation. Then, g Y coincides with the standard prolongation Z of vector field W : = g X with g smooth function on M satisfying
D x g = g λ .
Proof. 
We will work in local coordinates and write
X = ξ x + ϕ a y a , Y = X + ψ a y a ( 1 ) ,
with
ψ a = ( D x ϕ a y a ( 1 ) D x ξ ) + λ ( ϕ a y a ( 1 ) ξ ) ,
and
W = g X , Z = W + Θ a y a ( 1 ) ,
where (standard prolongation formula)
Θ a = D x ( g ϕ a ) y a ( 1 ) D x ( g ξ ) = D x ( g ) ϕ a + g D x ( ϕ a ) y a ( 1 ) ξ D x ( g ) y a ( 1 ) g D x ( ξ ) .
Now, requiring that Z = g Y amounts to requiring that Θ a = g ψ a , this is written as
g ( D x ϕ a y a ( 1 ) D x ξ ) + g λ ( ϕ a y a ( 1 ) ξ ) = D x ( g ) ϕ a + g D x ( ϕ a ) y a ( 1 ) ξ D x ( g ) y a ( 1 ) g D x ( ξ ) .
Eliminating the equal terms on both sides yields
( g λ D x g ) ( ϕ a y a ( 1 ) ξ ) = 0 ,
which implies ( 10 ) .
Example 4.
Let us consider M = R 2 with coordinates ( x , y ) . We now take the vector fields X 1 = y u and λ = y x , g = e y , then we can compute that
D x g = g λ = y x e y .
The 1-order λ-prolongation Y of X 1 and standard prolongation Z of g X 1 are equal. And Y = ( y x + y x y ) , Z = y x ( e y y + e y ) y , clearly, g Y = Z .
Inspired by Lemma 5, we prove that g Y cannot be a first-order λ -prolonged vector field of the vector field g X if g is not a constant function.
Theorem 5.
Let X = ξ ( x , y a ) x + ϕ a y a be vector field defined on M and Y = X + ψ 1 a y a ( 1 ) be a first-order λ-prolonged vector field of vector field X. There is no nontrivial function g defined on M that makes g Y be the first-order λ-prolonged vector field of vector field g X .
Proof. 
In fact,
( D x + λ ) ( g ϕ a ) y a ( 1 ) ( D x + λ ) ( g ξ ) = ( D x g ) ϕ a + g D x ϕ a + λ g ϕ a y a ( 1 ) ( ( D x g ) ξ + g ( D x ξ ) + λ g ξ ) = ( D x g ) ( ϕ a y a ( 1 ) ξ ) + g ( ( D x + λ ) ϕ a y a ( 1 ) ( D x + λ ) ξ ) = ( D x g ) ( ϕ a y a ( 1 ) ξ ) + g ψ 1 a ,
which implies that g ψ 1 a = ( D x + λ ) ( g ϕ a ) y a ( 1 ) ( D x + λ ) ( g ξ ) if and only if D x g = 0 . So, g Y is the 1-order λ -prolonged vector field of vector field g X if and only if D x g = 0 , i.e., g = C , C is a constant. □

5. Reduction of System of Explicit Ordinary Differential Equations

In this section, we consider the reduction of the system of explicit ordinary differential Equations ( 4 ) . Let V be a vector field defined on M .
Definition 6.
A function ζ defined on J j M is a j-th-order invariant of V if V ( j ) ( ζ ) = 0 , where V ( j ) is the j-order λ-prolongation of V .
Clearly, a j-th-order invariant is a ( j + 1 ) -th-order invariant.
Lemma 6.
If V ( k ) is a k-order λ-prolonged vector field, then it satisfies
[ V ( k ) , D x ] = λ V ( k ) [ ( D x ξ ) + λ ξ ] D x .
Proof. 
This follows from explicit computation. In fact,
D x = x + y a ( 1 ) y a + y a ( 2 ) y a ( 1 ) + + y a ( n + 1 ) y a ( n ) + ,
by ( 1 ) and ( 3 ) , we have immediately
[ V ( k ) , D x ] = V ( k ) D x D x V ( k ) = ( ψ k + 1 a D x ψ k a ) y a ( k ) ( D x ξ ) x = ( y a ( k + 1 ) D x ξ + λ ( ψ k a y a ( k + 1 ) ξ ) ) y a ( k ) ( D x ξ ) x = ( D x ξ ) D x + λ ( V ( k ) ξ x ) λ ξ ( D x x ) = [ ( D x ξ ) + λ ξ ] D x + λ V ( k ) .
Lemma 7.
Assume that if ξ and ζ are j-th-order invariants of a ( k ) -order λ-prolonged vector field V ( k ) , then Θ : = D x ( ζ ) / D x ( η ) is a ( j + 1 ) -th-order invariant of ( k + 1 ) -order λ-prolonged vector field V ( k + 1 ) .
Proof. 
It is obvious that Θ is a function defined on J k + 1 M . To show that it is invariant under the ( k + 1 ) -order λ -prolonged vector field V ( k + 1 ) , we just proceed by straightforward computation. First of all, by
V ( k + 1 ) ( Θ ) = ( V ( k + 1 ) ( D x ζ ) ) · ( D x η ) ( D x ζ ) · ( V ( k + 1 ) ( D x η ) ) ( D x η ) 2 : = χ ( D x η ) 2 ,
it is clear that we just have to show that the numerator χ vanishes. On the other hand, we have (recalling that by assumption V ( k ) ( ζ ) = V ( k ) ( η ) = 0 , and using Lemma 6)
χ = ( V ( k + 1 ) ( D x ζ ) ) · ( D x η ) ( D x ζ ) · ( V ( k + 1 ) ( D x η ) ) = ( D x ( V ( k + 1 ) ζ ) + [ V ( k + 1 ) , D x ] ζ ) · ( D x η ) ( D x ζ ) · ( ( D x ( V ( k + 1 ) η ) ) + [ V ( k + 1 ) , D x ] η ) = ( [ V ( k + 1 ) , D x ] ζ ) · ( D x η ) ( D x ζ ) · ( [ V ( k + 1 ) , D x ] η ) = ( ( λ V ( k + 1 ) [ ( D x ξ ) + λ ξ ] D x ) ζ ) · ( D x η ) ( D x ζ ) · ( ( λ V ( k + 1 ) [ ( D x ξ ) + λ ξ ] D x ) η ) = 0 .
This shows indeed χ = 0 and hence the lemma. □
Remark 3.
Let V i k be k-th λ i -prolonged vector fields i = 1 , . . . , l , and let η , ζ be independent common differential invariants of order k for all of them. Then, by using the same method as Lemma 7, we can know that
Θ : = D x ( ζ ) / D x ( η )
is a common differential invariant of order k + 1 for all of them.
Suppose we have find a zero-th-order invariant u and m independent first-order invariants z 1 , . . . , z m for λ -symmetry V . Define z a ( 0 ) = z a , a = 1 , . . . , m and z a ( j ) = D x ( z a ( j 1 ) ) / D x u for j = 1 , . . . , n 2 so that z a ( j ) are ( j + 1 ) th invariants of V.
Remark 4.
Because Γ = x + y a ( 1 ) y a + + y a ( n 1 ) y a ( n 2 ) + f a y a ( n 1 ) = x + y a ( 1 ) y a + + y a ( n 1 ) y a ( n 2 ) + y a ( n ) y a ( n 1 ) and z a ( j ) is defined on J j + 1 M , j = 1 , . . . , n 2 , z a ( j ) = D x ( z a ( j 1 ) ) / D x u = Γ ( z a ( j 1 ) ) / Γ ( u ) , j = 1 , . . . , n 2 .
Theorem 6.
Suppose u and z a , a = 1 , . . . , m are functionally independent zero-order and first-order invariants, respectively, of the λ-symmetry V, that is V ( u ) = V ( 1 ) ( z a ) = 0 , a = 1 , . . . , m , and assume that Γ ( z a ) and Γ ( u ) , a = 1 , . . . , m are nowhere zero. Put z a ( 0 ) = z linebreak and z a ( j ) = Γ ( z a ( j 1 ) ) / Γ ( u ) for j = 1 , . . . , n 2 . Then, u , z a , z a ( 1 ) , . . . , z a ( n 2 ) , a = 1 , . . . , m provide a complete set of functionally independent ( n 1 ) th-order invariant for V n 1 .
Proof. 
We use induction by assumption, [ V ( n 1 ) , Γ ] = λ Γ + λ ¯ V ( n 1 ) , and suppose that z a ( j ) is an invariant of V ( n 1 ) . Because V ( u ) = V ( 1 ) ( z a ) = 0 , we can obtain V ( n 1 ) ( u ) = 0 , V ( n 1 ) ( z a ( j ) ) = V ( j + 1 ) ( z a ( j ) ) = 0 , j = 0 , . . . , n 2 . Then, we prove that z a ( j + 1 ) is an invariant of V ( n 1 ) ,
( Γ ( u ) ) 2 V ( n 1 ) ( z a ( j + 1 ) ) = ( Γ ( u ) ) 2 V ( n 1 ) Γ ( z a ( j ) ) Γ ( u ) = ( Γ ( u ) ) 2 Γ ( u ) V ( n 1 ) ( Γ ( z a ( j ) ) ) Γ ( z a ( j ) ) V ( n 1 ) ( Γ ( u ) ) ( Γ ( u ) ) 2 = Γ ( u ) V ( n 1 ) ( Γ ( z a ( j ) ) ) Γ ( z a ( j ) ) V ( n 1 ) ( Γ ( u ) ) = Γ ( u ) ( V ( n 1 ) ( Γ ( z a ( j ) ) ) Γ ( V ( n 1 ) ( z a ( j ) ) ) ) Γ ( z a ( j ) ) ( V ( n 1 ) ( Γ ( u ) ) Γ ( V ( n 1 ) ( u ) ) ) = Γ ( u ) [ V ( n 1 ) , Γ ] ( z a ( j ) ) Γ ( z a ( j ) ) [ V ( n 1 ) , Γ ] ( u ) = Γ ( u ) ( λ Γ ( z a ( j ) ) + λ ¯ V ( n 1 ) ( z a ( j ) ) ) Γ ( z a ( j ) ) ( λ Γ ( u ) + λ ¯ V ( n 1 ) ( u ) ) = Γ ( u ) λ Γ ( z a ( j ) ) Γ ( z a ( j ) ) λ Γ ( u ) = 0 ,
hence V ( n 1 ) ( z a ( j ) ) = 0 for j = 0 , . . . , n 2 , a = 1 , . . . , m .
To see the functional independence, we note how Γ acting on a function effectively increases the order of derivatives, which it depends on by one. Because u is a zero-order invariant with no y a ( 1 ) , a = 1 , . . . , m dependence and z a , a = 1 , . . . , m must have nowhere zero y a ( 1 ) dependence, u and z a , a = 1 , . . . , m are functionally independent. Similarly, we know Γ ( z a ) will have nowhere zero y a ( 2 ) and only z a ( 1 ) depends on y a ( 2 ) , a = 1 , . . . , m , for instance, only z 1 ( 1 ) depends on y 1 ( 2 ) . From here, the result follows by induction. □
Proposition 2.
Denote f ^ a = Γ ( z a ( n 2 ) ) / Γ ( u ) , a = 1 , . . . , m , then f ^ a is also a first integral of V ( n 1 ) .
Proof. 
( Γ ( u ) ) 2 V ( n 1 ) ( f ^ a ) = ( Γ ( u ) ) 2 V ( n 1 ) Γ ( z a ( n 2 ) ) Γ ( u ) = ( Γ ( u ) ) 2 Γ ( u ) V ( n 1 ) ( Γ ( z a ( n 2 ) ) ) Γ ( z a ( n 2 ) ) V ( n 1 ) ( Γ ( u ) ) ( Γ ( u ) ) 2 = Γ ( u ) V ( n 1 ) ( Γ ( z a ( n 2 ) ) ) Γ ( z a ( n 2 ) ) V ( n 1 ) ( Γ ( u ) ) = Γ ( u ) ( V ( n 1 ) ( Γ ( z ( n 2 ) ) ) Γ ( V ( n 1 ) ( z a ( n 2 ) ) ) ) Γ ( z a ( n 2 ) ) ( V ( n 1 ) ( Γ ( u ) ) Γ ( V ( n 1 ) ( u ) ) ) = Γ ( u ) [ V ( n 1 ) , Γ ] ( z a ( n 2 ) ) Γ ( z a ( n 2 ) ) [ V ( n 1 ) , Γ ] ( u ) = Γ ( u ) ( λ Γ ( z a ( n 2 ) ) + λ ¯ V ( n 1 ) ( z a ( n 2 ) ) ) Γ ( z a ( n 2 ) ) ( λ Γ ( u ) + λ ¯ V ( n 1 ) ( u ) ) = Γ ( u ) λ Γ ( z a ( n 2 ) ) Γ ( z a ( n 2 ) ) λ Γ ( u ) = 0 .
By Theorem 6, V ( n 1 ) has a complete set of functionally independent invariants { u , z a , z a ( 1 ) , . . . , z a ( n 2 ) } . And on the set S ( Δ ) : = { ( x , y a , y a ( 1 ) , . . . , y a ( n ) ) | y a ( n ) = f a ( x , y a , y a ( 1 ) , . . . , y a ( n 1 ) ) , a = 1 , . . . , m } , z a ( n 1 ) = D x ( z a ( n 2 ) ) / D x u = d n 1 z a / d u n 1 = Γ ( z a ( n 2 ) ) / Γ ( u ) = f ^ a , which is also a first integral of V ( n 1 ) , so it depends on functions u , z a , z a ( 1 ) , . . . , z a ( n 2 ) , i.e., f ^ a = f ^ a ( u , z a , z a ( 1 ) , . . . , z a ( n 2 ) ) . Thus, we obtain the ( n 1 ) th-order equation
d n 1 z a / d u n 1 = f ^ ( u , z a , z a ( 1 ) , . . . , z a ( n 2 ) ) .
We can recover the general solution of the original equation from the general solution of ( 12 ) and the corresponding system of the first-order auxiliary equation
z a = z a ( x , y a , y a ( 1 ) ) , a = 1 , . . . , m .
In fact, as for Equation ( 12 ) , if we obtain the solution z a ( u ) , together with ( 13 ) , we only need to solve system of the first-order equation z a ( u ( x , y a ) ) = z a ( x , y a , y a ( 1 ) ) to obtain the solution y a ( x ) of ( 4 ) .
Remark 5.
In [12], the authors studied ordinary differential Equations ( 4 ) with m = 1 . Here, we discuss the system of ordinary differential Equations ( 4 ) with m 1 , and our method states that if we deal with an n-order explicit equation that has a λ-symmetry V , then we just need to consider the invariants of ( n 1 ) -order prolongation V ( n 1 ) of V . In [12], the authors only dealt with implicit equation with a λ-symmetry V , which needs to consider the invariants of n-order prolongation V ( n 1 ) of V .
On the other hand, putting θ a j = d z a ( j ) z a ( j + 1 ) d u for j = 0 , . . . , n 3 , a = 1 , . . . , m and θ a n 2 = d z a ( n 2 ) f ^ a d u , we have m ( n 1 ) independent 1-forms.
Proposition 3.
On S ( Δ ) , i Γ θ a j = 0 , i V ( n 1 ) θ a j = 0 , j = 0 , . . . , n 2 , a = 1 , . . . , m and i V ( n 1 ) Ω = α θ a 0 θ a n 2 , α is a some function defined on J n 1 M .
Proof. 
Because z a ( j + 1 ) = Γ ( z a ( j ) ) Γ ( u ) = d z a ( j ) ( Γ ) d u ( Γ ) , j = 0 , . . . , n 3 , f ^ a = Γ ( z a ( n 2 ) ) Γ ( u ) = d z a ( n 2 ) ( Γ ) d u ( Γ ) , V ( n 1 ) ( z a ( j ) ) = 0 , V ( n 1 ) ( u ) = 0 . We have ( d z a ( j ) z a ( j + 1 ) d u ) ( Γ ) = i Γ θ a j = 0 , j = 0 , . . . , n 3 , ( d z a ( n 2 ) f ^ a d u ) ( Γ ) = i Γ θ a n 2 = 0 , and i V ( n 1 ) θ a j = V n 1 ( z a ( j ) ) z a ( j + 1 ) V ( n 1 ) ( u ) = 0 , j = 0 , . . . , n 3 , i V ( n 1 ) θ a n 2 = V n 1 ( z a ( n 2 ) ) f ^ a V ( n 1 ) ( u ) = 0 , thus i V ( n 1 ) ( θ a 0 θ a n 2 ) = i Γ ( θ a 0 θ a n 2 ) = 0 , moreover, we know that ker i V ( n 1 ) Ω = span { V ( n 1 ) , Γ } , hence i V ( n 1 ) Ω = α θ a 0 θ a n 2 . □
Example 5.
Let M = R 2 with coordinate ( x , y ) , and consider the equation
y = y + x y .
Then, Γ = x + y y + ( y + x y ) y , and we can also compute that X = y is a λ-symmetry of ( 14 ) , λ = x . In fact, using the λ-prolongation formula ( 3 ) , we obtain the 1-order λ-prolongation Y of X, i.e., Y = X + ( D x + x ) 1 y = y + x y . Then,
[ Y , Γ ] = Y Γ Γ Y = Y Γ Γ Y = y + x y + x 2 y y = x ( y + x y ) = x Y .
X ( x ) = 0 , Y ( y + x 2 x y ) = 0 , u : = x , z : = y + x 2 x y are 0-order and 1-order invariants of X, respectively. So, we obtain an explicit order equation
z = Γ ( z ) Γ ( u ) = 2 x = 2 u ,
thus, we can obtain a solution z = u 2 + C , C as a constant. Together with z = y + x 2 x y , we have y + x 2 x y = x 2 + C , i.e.,
y = x y + C .
Hence, we have reduced Equation ( 14 ) to the first-order Equation ( 15 ) .

6. Involution of λ -Prolonged Sets of Vector Fields

Assume there is a set of vector fields X i , i = 1 , . . . , n that are in involution, and all of them have the form X i = ϕ i a y a , i = 1 , . . . , n . We will give a theorem that claims under what conditions are their λ -prolongations also in involution.
Theorem 7.
Assume vector fields X i = ϕ i a y a , i = 1 , . . . , n defined on M are involution, with [ X i , X j ] = μ i j k X k for μ i j k smooth functions on M . Then, their λ-prolongations Y i , i = 1 , . . . , n satisfy the same involution relations, i.e., [ Y i , Y j ] = μ i j k Y k , if
(1) Y i ( λ j ) = 0 for i , j = 1 , . . . , n , i j ;
(2) D x ( μ i j k ) + λ i μ i j k + λ j μ i j k + λ k μ i j k = 0 , for i , j , k = 1 , . . . , n .
Proof. 
This follows from an explicit computation. We proceed by induction on the order of the prolongation. Denoting by Z i the ( q 1 ) th λ i -prolongation of X i , we have Y i = k = 0 q ψ i , k a y a ( k ) + Z i + ψ i , q a y a ( q ) , thus
[ Y i , Y j ] = [ Y i , Z j + ψ j , q a y a ( q ) ] = [ Y i , Z j ] + [ Y i , ψ j , q a y a ( q ) ] = [ Z i + ψ i , q a y a ( q ) , Z j ] + [ Y i , ψ j , q a y a ( q ) ] = [ Z i , Z j ] Z j ( ψ i , q a ) y a ( q ) + Y i ( ψ j , q a ) y a ( q ) ψ j , q a ψ i , q a y a ( q ) y a ( q ) = [ Z i , Z j ] + ( Y i ( ψ j , q a ) Y j ( ψ i , q a ) ) y a ( q ) .
Denote F i j , q a = Y i ( ψ j , q a ) Y j ( ψ i , q a ) , because by induction [ Z i , Z j ] = μ i j k Z k (i.e., the involution relations are satisfied for ( q 1 ) th prolongations), the requirement that [ Y i , Y j ] = μ i j k Y k = μ i j k Z k + μ i j k ψ k , q a y a ( q ) is equivalent to the requirement that
F i j , q q = μ i j k ψ k , q a .
F i j , k a needs to be rewritten for easier comparison with the rhs of ( 16 ) , so by X i = ϕ i a y a , we can write ( 16 ) as
F i j , q q = μ i j k ( D x ψ k , q 1 a + λ k ψ k , q 1 a ) .
Using Lemma 6, Condition (1) and Y i ψ j , q 1 a = Z i ψ j , q 1 a , with standard computation, we obtain
F i j , q a = Y i ( ψ j , q a ) Y j ( ψ i , q a ) = Y i ( ( D x + λ j ) ψ j , q 1 a ) Y j ( ( D x + λ i ) ψ i , q 1 a ) = Y i D x ψ j , q 1 a Y j D x ψ i , q 1 a + ψ j , q 1 a Y i ( λ j ) ψ i , q 1 a Y j ( λ i ) + λ j Y i ( ψ j , q 1 a ) λ i Y j ( ψ i , q 1 a ) = D x Y i ψ j , q 1 a + λ i Y i ψ j , q 1 a D x Y j ψ i , q 1 a λ j Y j ψ i , q 1 a + λ j Y i ( ψ j , q 1 a ) λ i Y j ( ψ i , q 1 a ) = D x ( Y i ψ j , q 1 a Y j ψ i , q 1 a ) + λ i ( Y i ψ j , q 1 a Y j ψ i , q 1 a ) + λ j ( Y i ψ j , q 1 a Y j ψ i , q 1 a ) = ( D x + λ i + λ j ) ( Y i ψ j , q 1 a Y j ψ i , q 1 a ) = ( D x + λ i + λ j ) ( Z i ψ j , q 1 a Z j ψ i , q 1 a ) = ( D x + λ i + λ j ) ( μ i j k ψ k , q 1 a ) = ( λ i + λ j ) ( μ i j k ψ k , q 1 a ) + D x ( μ i j k ) ψ k , q 1 a + μ i j k D x ψ k , q 1 a = ( D x ( μ i j k ) + λ i μ i j k + λ j μ i j k ) ψ k , q 1 a + μ i j k D x ψ k , q 1 a .
Comparing this with ( 17 ) , we must require
F i j , q a = ( D x ( μ i j k ) + λ i μ i j k + λ j μ i j k ) ψ k , q 1 a + μ i j k D x ψ k , q 1 a = μ i j k ( D x ψ k , q 1 a + λ k ψ k , q 1 a ) .
Eliminating equal terms on both sides, we obtain
( D x ( μ i j k ) + λ i μ i j k + λ j μ i j k + λ k μ i j k ) ψ k , q 1 a = 0 ,
which holds under Condition ( 2 ) , so the involution properties holding up to order q 1 are also holding at order q .
Remark 6.
Noting that in Condition (2) of Theorem 7, λ k μ i j k is not the Einstein summation.
Remark 7.
If vector fields X i = ϕ i a y a , i = 1 , . . . , n defined on M commute, i.e., [ X i , X j ] = 0 , μ i j k = 0 , and their λ i -prolongations Y i , i = 1 , . . . , n satisfy Condition (1) of Theorem 7, then Y i commute i = 1 , . . . , n .
Remark 8.
If λ i is a common first integral of X j , j i , j = 1 , . . . , n . It is natural Y i ( λ j ) = 0 for i , j = 1 , . . . , n , i j .
Example 6.
Let us consider X = R , with coordinate x and U = R 2 with coordinates ( u , v ) . We now take the vector fields
X 1 = u u , X 2 = u v ,
these are in involution, in fact,
[ X 1 , X 2 ] = X 1 X 2 X 2 X 1 = u v = X 2 ,
thus μ 12 1 = μ 21 1 = 0 , μ 12 2 = 1 , μ 21 2 = 1 , μ i i k = 0 , i , k = 1 , 2 . We take λ 1 = 2 x , λ 2 = x , so
D x ( μ i j k ) + λ i μ i j k + λ j μ i j k + λ k μ i j k = 0 ,
and
Y 1 = u u + ( u x + 2 x u ) u x , Y 1 ( λ 2 ) = Y 1 ( x ) = 0 ,
Y 2 = u v ( u x x u ) v x , Y 2 ( λ 1 ) = 2 Y 2 ( x ) = 0 .
Hence, by Theorem 7, we know [ Y 1 , Y 2 ] span { Y 1 , Y 2 } . In fact,
[ Y 1 , Y 2 ] = [ u u + ( u x + 2 x u ) u x , u v ( u x x u ) v x ] = u v + x u v x u x v x = Y 2 .
Example 7.
Let us consider X = R , with coordinate x and U = R 2 with coordinates ( u , v ) . We now take the vector fields
X 1 = u u , X 2 = v v ,
these are in involution, i.e., [ X 1 , X 2 ] = 0 , by Remark 8 . Just take λ 1 = v , λ 2 = u . We know that X 1 ( v ) = X 2 ( u ) = 0 , so for n-order λ 1 -prolongations Y 1 and λ 2 -prolongations Y 2 of X 1 , X 2 , we can obtain [ Y 1 , Y 2 ] = 0 . For instance, Y 1 , Y 2 are 1-order λ 1 -prolongation and 1-order λ 2 -prolongation of X 1 , X 2 , i.e.,
Y 1 = u u + ( u x + v u ) u x ,
Y 2 = v v + ( v x + v u ) v x .
It is clear that [ Y 1 , Y 2 ] = 0 . Moreover, we can also compute the 2-order λ 1 -prolongations and 2-order λ 2 -prolongation of X 1 , X 2 , i.e.,
Y ˜ 1 = u u + ( u x + v u ) u x + ( 2 u x v + u x x + v 2 u ) u x x , Y ˜ 2 = v v + ( v x + v u ) v x + ( 2 v x u + v x x + v u 2 ) v x x ,
it straightly follows that [ Y ˜ 1 , Y ˜ 2 ] = 0 .

7. Conclusions

In this paper, we generalize a theorem in [18] and show that if there are not enough symmetries and λ -symmetries, some first integrals can still be obtained. Moreover, we give two examples to illustrate this theorem. Secondly, we prove that when X is a λ -symmetry of differential equation field Γ of ( 4 ) , by multiplying Γ , a function μ defined on J n 1 M , and the vector fields μ Γ can pass to quotient manifold Q by a group action of λ -symmetry X. Thirdly, if there are some λ -symmetries of ( 4 ) , we show that the vector fields from their linear combination are symmetries of ( 4 ) under some conditions. And if we have vector field X defined on J n 1 M with first-order λ -prolongation Y and first-order standard prolongations Z of X defined on J n M , we prove that g Y cannot be first-order λ -prolonged vector field of vector field g X if g is not a constant function. Fifth, we provide a complete set of functionally independent ( n 1 ) -order invariants for V ( n 1 ) , which are n 1 th prolongations of λ -symmetry of V, and obtain an explicit n 1 -order-reduced equation of explicit n-order ordinary Equation ( 4 ) . Finally, assume there is a set of vector fields X i , i = 1 , . . . , n that are in involution and all of them have the form X i = ϕ i a y a , i = 1 , . . . , n . We give Theorem 7 to claim under what conditions are their λ -prolongations also in involution.

Author Contributions

Conceptualization, W.L. and X.L.; methodology, W.L. and Y.P.; writing—original draft preparation, W.L. and X.L.; writing—review and editing, W.L. and Y.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research is supported by the NNSF of China (Grant No. 41301182), Natural Science Foundation of Jilin Province (Grant No. 20210101153JC), and the 13th Five-Year Science and Technology Project of Jilin Provincial Department of Education (Grant No. JJKH20200136KJ).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Alekseevsky, A.D.V.; Vinogradov, A.M.; Lychagin, V.V. Basic Ideas and Concepts of Differential Geometry; Springer: Berlin/Heidelberg, Germany, 1991. [Google Scholar]
  2. Bhuvaneswari, A.; Chandrasekar, V.K.; Senthilvelan, M.; Lakshmanan, M. On the complete integrability of a nonlinear oscillator from group theoretical perspective. J. Math. Phys. 2012, 53, 073504. [Google Scholar] [CrossRef]
  3. Cicogna, G.; Gaeta, G. Symmetry and Perturbation Theory in Nonlinear Dynamics; Springer: Berlin/Heidelberg, Germany, 1999. [Google Scholar]
  4. Olver, P.J. Application of Lie Groups to Differential Equations; Springer: Berlin/Heidelberg, Germany, 1986. [Google Scholar]
  5. Olver, P.J. Equivalence, Invariants and Symmetry; Cambridge University Press: Cambridge, UK, 1995. [Google Scholar]
  6. Ovsiannikov, L.V. Group Analysis of Differential Equations; Academic Press: New York, NY, USA; London, UK, 1982. [Google Scholar]
  7. Stephani, H. Differential Equations: Their Solution Using Symmetries; Cambridge University Press: Cambridge, UK, 1989. [Google Scholar]
  8. González-López, A. Symmetry and integrability by quadratures of ordinary differential equations. Phys. Lett. A 1988, 133, 190–194. [Google Scholar]
  9. Gónzalez-Gascón, F.; González-López, A. Newtonian systems of differential equations, integrable via quadratures, with trivial group of point symmetries. Phys. Lett. A 1988, 129, 153–156. [Google Scholar] [CrossRef]
  10. Meleshko, S.V.; Moyo, S.; Muriel, C.; Romero, J.L.; Guha, P.; Choudhury, A.G. On first integrals of second-order ordinary differential equation. J. Eng. Math. 2013, 82, 17–30. [Google Scholar]
  11. Zhao, X.F.; Li, Y. The Reduction, First Integral and Kam Tori for n-Dimensional Volume-Preserving Systems. J. Dyn. Diff. Equation 2023, 1–20. [Google Scholar] [CrossRef]
  12. Muriel, C.; Romero, J.L. New methods of reduction for ordinary differential equations. IMA J. Appl. Math. 2001, 66, 111–125. [Google Scholar]
  13. Muriel, C.; Romero, J.L. Integrating factors and λ-symmetries. J. Nonlinear Math. Phys. 2008, 15, 300–309. [Google Scholar] [CrossRef]
  14. Muriel, C.; Romero, J.L. First integrals, integrating factors and λ-symmetries of second-order differential equations. J. Phys. A Math. Theor. 2009, 42, 365207. [Google Scholar] [CrossRef]
  15. Pan-Collantes, A.J.; Ruiz, A.; Muriel, C.; Romero, J.L. C-symmetries of distributions and integrability. J. Differ. Equation 2023, 348, 126–153. [Google Scholar]
  16. Pucci, E.; Saccomandi, G. On the reduction methods for ordinary differential equations. J. Phys. A Math. Gen. 2002, 35, 6145–6155. [Google Scholar]
  17. Yasar, E. λ-symmetries, nonlocal transformations and first integrals to a class of Painlevé-Gambier equations. Math. Methods Appl. Sci. 2012, 35, 684–692. [Google Scholar] [CrossRef]
  18. Zhang, J. A relationship between λ-symmetries and first integrals for ordinary differential equations. Math. Methods Appl. Sci. 2019, 42, 6139–6154. [Google Scholar] [CrossRef]
  19. Zhang, J.; Li, Y. Symmetries and first integrals of differential equations. Acta Appl. Math. 2008, 103, 147–159. [Google Scholar] [CrossRef]
  20. Cicogna, G.; Gaeta, G.; Walcher, S. A generalization of λ-symmetry reduction for systems of ODEs: σ-symmetries. J. Phys. A 2012, 45, 355205. [Google Scholar] [CrossRef]
  21. Zhao, X.F.; Li, Y. σ-Symmetries and First Integral of Differential Equations. J. Lie Theory 2024, in press. 34. [Google Scholar]
  22. Cicogna, G.; Gaeta, G. Noether theorem for μ-symmetries. J. Phys. A Math. Theor. 2007, 40, 11899–11921. [Google Scholar] [CrossRef]
  23. Morando, P. Deformation of Lie derivative and μ-symmetries. J. Phy. A Math. Theor. 2007, 40, 11547–11559. [Google Scholar] [CrossRef]
  24. Malik, S.; Hashemi, M.S.; Kumar, S.; Rezazadeh, H.; Mahmoud, W.; Osman, M.S. Application of new Kudryashov method to various nonlinear partial differential equations. Opt. Quant. Electron. 2023, 55, 8. [Google Scholar] [CrossRef]
  25. De Lara, M.C. Geometric and symmetry properties of a nondegenerate diffusion process. Ann. Probab. 1995, 23, 1557–1604. [Google Scholar] [CrossRef]
  26. Gaeta, G. W-symmetries of Ito stochastic differential equations. J. Math. Phys. 2019, 60, 053501. [Google Scholar] [CrossRef]
  27. Glover, J. Symmetry groups and translation invariant representations of Markov processes. Ann. Probab. 1991, 19, 562–586. [Google Scholar] [CrossRef]
  28. Glover, J.; Mitro, J. Symmetries and functions of Markov process. Ann. Probab. 1990, 18, 655–668. [Google Scholar] [CrossRef]
  29. De Vecchi, F.C.; Morando, P.; Ugolini, S. Symmetries of stochastic differential equation: A geometric approach. J. Math. Phy. 2016, 57, 123508. [Google Scholar] [CrossRef]
  30. De Vecchi, F.C.; Morando, P.; Ugolini, S. A note on symmetries of diffusions with a martingale problem approach. Stoch. Dyn. 2019, 19, 1950011. [Google Scholar] [CrossRef]
  31. Dorodnitsyn, V. Applications of Lie Groups to Difference Equations; Differential and Integral Equations and Their Applications; CRC Press: Boca Raton, FL, USA, 2011. [Google Scholar]
  32. Levi, D.; Winternitz, P.; Yamilov, R.I. Continuous Symmetries and Integrability of Discrete Equations; American Mathematical Society: Providence, RI, USA, 2022. [Google Scholar]
  33. Muriel, C.; Romero, J.L. C symmetries and nonsolvable symmetry algebras. IMA J. Appl. Math. 2001, 66, 477–498. [Google Scholar] [CrossRef]
  34. Von Westenholz, C. Differential Forms in Mathematical Physics; Elsevier: Amsterdam, The Netherlands, 1981. [Google Scholar]
  35. Sherring, J.; Prince, G. Geometric aspects of reduction of order. Trans. Am. Math. Soc. 1992, 334, 433–453. [Google Scholar] [CrossRef]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Li, W.; Li, X.; Pang, Y. On the Properties of λ-Prolongations and λ-Symmetries. Mathematics 2023, 11, 4113. https://doi.org/10.3390/math11194113

AMA Style

Li W, Li X, Pang Y. On the Properties of λ-Prolongations and λ-Symmetries. Mathematics. 2023; 11(19):4113. https://doi.org/10.3390/math11194113

Chicago/Turabian Style

Li, Wenjin, Xiuling Li, and Yanni Pang. 2023. "On the Properties of λ-Prolongations and λ-Symmetries" Mathematics 11, no. 19: 4113. https://doi.org/10.3390/math11194113

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

Article Metrics

Back to TopTop