Next Article in Journal
An Improved Equilibrium Optimizer with a Decreasing Equilibrium Pool
Previous Article in Journal
Model-Free Predictive Power Control for PWM Rectifiers under Asymmetrical Grids
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Noether and Lie Symmetry for Singular Systems Involving Mixed Derivatives

School of Mathematical Sciences, Suzhou University of Science and Technology, Suzhou 215009, China
Symmetry 2022, 14(6), 1225; https://doi.org/10.3390/sym14061225
Submission received: 21 April 2022 / Revised: 8 June 2022 / Accepted: 10 June 2022 / Published: 13 June 2022
(This article belongs to the Topic Fractional Calculus: Theory and Applications)

Abstract

:
Singular systems play an important role in many fields, and some new fractional operators, which are general, have been proposed recently. Therefore, singular systems on the basis of the mixed derivatives including the integer order derivative and the generalized fractional operators are studied. Firstly, Lagrange equations within mixed derivatives are established, and the primary constraints are presented for the singular systems. Then the constrained Hamilton equations are constructed by introducing the Lagrange multipliers. Thirdly, Noether symmetry, Lie symmetry and the corresponding conserved quantities for the constrained Hamiltonian systems are investigated. And finally, an example is given to illustrate the methods and results.

1. Introduction

Fractional calculus is a hot topic recently. Many results have been obtained in fractional calculus and its applications [1,2,3,4,5,6,7]. Since fractional derivatives were used to deal with dissipative forces for nonconservative systems by Riewe [8,9] in 1996, fractional variational problems became popular. For example, Klimek [10] studied Lagrangian and Hamiltonian fractional sequential mechanics; Muslih and Baleanu [11] established the Hamiltonian formulation of the systems with linear velocities within the Riemann–Liouville fractional derivative; Agrawal [12], investigated the fractional variational calculus in terms of the Riesz fractional derivative; Luo [13] studied the fractional Birkhoffian mechanics in terms of the combined Riemann-Liouville fractional derivative and the combined Caputo fractional derivative; Song and Agrawal [14] presented the Euler-Lagrange equations involving the Caputo fractional derivative for singular systems, and so on [15,16]. Especially, in 2010, Agrawal [17] introduced a new kernel κ α t , τ (or κ α τ , t ), on which the generalized fractional derivatives are defined. Only when the parameter set is specified, and the kernel κ α t , τ is equal to t τ α 1 / Γ α , can the Riemann-Liouville fractional derivative, the Caputo fractional derivative, the Riesz-Riemann-Liouville fractional derivative and the Riesz-Caputo fractional derivative be obtained. Besides, the kernel can also be replaced with other kernels, and the entire theories of classical and fractional variational calculus can be redeveloped. Therefore, the generalized fractional derivatives are more general.
Singular system is another keyword of this paper. Singular system plays an important role in field theory, because many important physical systems in field theory are singular ones, such as the Yang-Mills field, the gravitational field, the electromagnetic field, supersymmetry, superstring, supergravity, relativistic moving particles and so on [18]. Singular system has two forms, one is expressed by a Lagrangian, and the other is expressed by a Hamiltonian. When a singular system is expressed in the form of the Hamiltonian, there exist inherent constraints among the canonical variables, and the corresponding system is called a constrained Hamiltonian system [19,20]. The constrained Hamiltonian system also has many applications, such as in quantum field theories of anyons and theories of condensed matter [19,21,22].
In this paper, we intend to study the fractional calculus of variations for singular systems on the basis of generalized fractional derivatives. After the fractional differential equations of motion are established, the next step is to find solutions to them. The symmetry method in mechanics is one of the most effective methods. The symmetry method mainly contains three kinds of methods, namely, the Noether symmetry method, the Lie symmetry method and the Mei symmetry method [23]. This paper focuses on the Noether symmetry method and the Lie symmetry method. The Noether symmetry method was introduced by a German female mathematician Noether [24]. Noether symmetry is an invariance of the Hamilton action under the infinitesimal transformations of time and coordinates, and can lead to a conserved quantity. Lie symmetry is an invariance of the differential equations under the infinitesimal transformations of time and coordinates. Lie symmetry can also lead to a conserved quantity under certain conditions. Many results have been obtained with Noether symmetry and Lie symmetry, including both integer order calculus and fractional order calculus [25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46], we only refer to them briefly here.

2. Preliminaries on the Generalized Operators

Generalized operators were introduced by Agrawal [17] in 2010. He defined the operator K M α as
K M α f t = m t 1 t κ α t , τ f τ d τ + ω t t 2 κ α τ , t f τ d τ ,   α > 0 ,
where t 1 < t < t 2 , M = < t 1 , t , t 2 , m , ω > is a parameter set, m and ω are two real numbers, κ α t , τ is a kernel which may depend on a parameter α . It is easy to verify that the operator K M α is a linear operator and satisfies the following integration by parts formula,
t 1 t 2 g t K M α f t d t = t 1 t 2 f t K M * α g t d t
where M * = < t 1 , t , t 2 , ω , m > .
The operators A M α and B M α were defined by Agrawal as
A M α f t = D n K M n α f t ,   n 1 < α < n ,
B M α f t = K M n α D n f t ,   n 1 < α < n ,
where D means the classical integer order derivative d / d t , n is an integer. Both operators are also linear and they satisfy the following integration by parts formulae,
t 1 t 2 g t A M α f t d t = 1 n t 1 t 2 f t B M * α g t d t + j = 0 n 1 D n 1 j g t A M α + j n f t t = t 1 t = t 2
t 1 t 2 g t B M α f t d t = 1 n t 1 t 2 f t A M * α g t d t + j = 0 n 1 1 j A M * α + j n g t D n 1 j f t t = t 1 t = t 2
where M * = < t 1 , t , t 2 , ω , m > , n is an integer, n 1 < α < n .
Specifically, if we let κ α t , τ = t τ α 1 / Γ α and let M = M 1 = < t 1 , t , t 2 , 1 , 0 > , M = M 2 = < t 1 , t , t 2 , 0 , 1 > and M = M 3 = < t 1 , t , t 2 , 1 / 2 , 1 / 2 > , then the operator A M α reduces to the left Riemann-Liouville, the right Riemann-Liouville and the Riesz-Riemann-Liouville fractional derivative operators, respectively. Similarly, the operator B M α reduces to the left Caputo, the right Caputo and the Riesz-Caputo fractional derivative operators, respectively.
In this text we set 0 < α < 1 . We begin with variational problems and the primary constraints.

3. Variational Problems and the Primary Constraints

3.1. The Variational Problem and the Primary Constraint with the Operator A M α

Hamilton action with the operator A M α is defined as
I A q A = t 1 t 2 L A t , q A , q ˙ A , A M α q A d t
where q A = q A 1 , q A 2 , , q A n , q A i C 2 t 1 , t 2 ; , i = 1 , 2 , , n , q ˙ A = q ˙ A 1 , q ˙ A 2 , , q ˙ A n , A M α q A = A M α q A 1 , A M α q A 2 , , A M α q A n , 0 < α < 1 and L A , , , C 2 t 1 , t 2 × n × n × n ; .
Then
δ I A = 0
with
q A t 1 = q A 1 ,   q A t 2 = q A 2 ,   δ q ˙ A i = d d t δ q A i ,   δ A M α q A i = A M α δ q A i ,   i = 1 , 2 , , n
is called the Hamilton principle with the operator A M α , where q A 1 = q A 11 , q A 12 , , q A 1 n , q A 2 = q A 21 , q A 22 , , q A 2 n .
From Equations (5), (8) and (9), we obtain
L A q A i d d t L A q ˙ A i B M * α L A A M α q A i + m κ 1 α t 2 , t L A t 2 A M α q A i ω κ 1 α t , t 1 L A t 1 A M α q A i = 0
where L A t 1 = L A t 1 , q A t 1 , q ˙ A t 1 , A M α q A t 1 , L A t 2 = L A t 2 , q A t 2 , q ˙ A t 2 , A M α q A t 2 , i = 1 , 2 , , n . Equation (10) is called the Lagrange equation with the operator A M α .
Define the generalized momenta and the Hamiltonian as
p A i = L A t , q A , q ˙ A , A M α q A q ˙ A i ,   p A i α = L A t , q A , q ˙ A , A M α q A A M α q A i ,
H A = p A i q ˙ A i + p A i α A M α q A i L A t , q A , q ˙ A , A M α q A ,   i = 1 , 2 , , n .
In this paper, we assume that A M α q A i = u A i t , q A , q ˙ A , p A α (or A M α q A = u A t , q A , q ˙ A , p A α ), where p A α = p A 1 α , p A 2 α , , p A n α , u A = u A 1 , u A 2 , , u A n .
Define the elements H A i j , i , j = 1 , 2 , , n , of the Hessian matrix H A i j as
H A i j = 2 L A q ˙ A i q ˙ A j ,   i , j = 1 , 2 , , n ,
then the Lagrangian is called regular if det H A i j 0 , and if det H A i j = 0 , then the Lagrangian L A is called singular. In this text, we assume that det H A i j = 0 and rank H A i j = R , 0 R < n . In the sequel, we will discuss two cases, i.e., 1 R < n and R = 0 .
Firstly, when 1 R < n , which means that only q ˙ A σ , σ = 1 , 2 , , R , can be determined from Equation (11) while q ˙ A ρ , ρ = R + 1 , R + 2 , , n , are random. From Equation (11), we express q ˙ A σ , σ = 1 , 2 , , R , as
q ˙ A σ = f A σ t , q A , p A α , p A E , q ˙ A F ,   ( or   q ˙ A E = f A E t , q A , p A α , p A E , q ˙ A F ) ,  
where p A Ε = p A 1 , p A 2 , , p A R , q ˙ A E = q ˙ A 1 , q ˙ A 2 , , q ˙ A R , q ˙ A F = q ˙ A R + 1 , q ˙ A R + 2 , , q ˙ A n , f A E = f A 1 , f A 2 , , f A R , 1 R < n .
From Equations (11) and (14), we have
p A i = g A i t , q A , f A E t , q A , p A α , p A E , q ˙ A F , q ˙ A F , p A α = g A i t , q A , p A α , p A E , q ˙ A F ,   i = 1 , 2 , , n .
For Equation (15), if i = 1 , 2 , , R , 1 R < n , then Equation (15) always holds. If i = R + 1 , R + 2 , , n , 1 R < n , then from the assumption rank H A i j = R , 1 R < n , we have
p A ρ = g A ρ t , q A , p A E , p A α ,   ( or   p A F = g A F t , q A , p A E , p A α ) ,
where ρ = R + 1 , R + 2 , , n , p A F = p A R + 1 , p A R + 2 , , p A n , g A F = g A R + 1 , g A R + 2 , , g A n , 1 R < n . Equation (16) has another form
ϕ A t , q A , p A , p A α = p A F g A F t , q A , p A E , p A α = 0 ,
where ϕ A = ϕ A 1 , ϕ A 2 , , ϕ A n R , 1 R < n .
Secondly, when R = 0 , which means that no q ˙ A i , i = 1 , 2 , , n , can be determined from Equation (11). Then from Equation (11) and the assumption rank H A i j = R , R = 0 , we have
p A i = g A i t , q A , p A α ,   ( or   p A = g A t , q A , p A α ) ,
where i = 1 , 2 , , n , p A = p A 1 , p A 2 , , p A n , g A = g A 1 , g A 2 , , g A n . Then Equation (18) gives
ϕ A t , q A , p A , p A α = p A g A t , q A , p A α = 0 ,
where ϕ A = ϕ A 1 , ϕ A 2 , , ϕ A n .
Incorporating Equations (17) and (19), we get
ϕ A t , q A , p A , p A α = 0 ,
where ϕ A = ϕ A 1 , ϕ A 2 , , ϕ A n R , 0 R < n . Equation (20) is called primary constraint with the operator A M α .
Remark 1.
Let κ α t , τ = t τ α 1 / Γ α , when M = M 1 , M = M 2 and M = M 3 , Equations (10) and (20) give the Lagrange equations and the primary constraints in terms of the left Riemann-Liouville, the right Riemann-Liouville and the Riesz-Riemann-Liouville fractional derivatives, respectively.

3.2. The Variational Problem and the Primary Constraint with the Operator B M α

Hamilton action with the operator B M α is
I B q B = t 1 t 2 L B t , q B , q ˙ B , B M α q B d t ,
where q B = q B 1 , q B 2 , , q B n , q B i C 2 t 1 , t 2 ; , i = 1 , 2 , , n , q ˙ B = q ˙ B 1 , q ˙ B 2 , , q ˙ B n , B M α q B = B M α q B 1 , B M α q B 2 , , B M α q B n , L B , , , C 2 t 1 , t 2 × n × n × n ; and 0 < α < 1 . Then
δ I B = 0
with
q B t 1 = q B 1 ,   q B t 2 = q B 2 ,   δ q ˙ B i = d d t δ q B i ,   δ B M α q B i = B M α δ q B i ,   i = 1 , 2 , , n
is called the Hamilton principle with the operator B M α , where q B 1 = q B 11 , q B 12 , , q B 1 n , q B 2 = q B 21 , q B 22 , , q B 2 n .
From Equations (6), (22) and (23), we obtain
L B q B i d d t L B q ˙ B i A M * α L B B M α q B i = 0 ,   i = 1 , 2 , , n .
Equation (24) is called the Lagrange equation with the operator B M α .
Define the generalized momenta and the Hamiltonian as
p B i = L B t , q B , q ˙ B , B M α q B q ˙ B i ,   p B i α = L B t , q B , q ˙ B , B M α q B B M α q B i ,
H B = p B i q ˙ B i + p B i α B M α q B i L B t , q B , q ˙ B , B M α q B ,   i = 1 , 2 , , n .
In this paper, we assume that B M α q B i = u B i t , q B , q ˙ B , p B α (or B M α q B = u B t , q B , q ˙ B , p B α ), where p B α = p B 1 α , p B 2 α , , p B n α , u B = u B 1 , u B 2 , , u B n .
Define the elements H B i j , i , j = 1 , 2 , , n , of the Hessian matrix H B i j as
H B i j = 2 L B q ˙ B i q ˙ B j ,   i , j = 1 , 2 , , n ,
then the Lagrangian is called regular if det H B i j 0 , and if det H B i j = 0 , then the Lagrangian L B is called singular. In this text, we assume that det H B i j = 0 and rank H B i j = R , 0 R < n . In the sequel, we will discuss two cases, i.e., 1 R < n and R = 0 .
Firstly, when 1 R < n , which means that only q ˙ B σ , σ = 1 , 2 , , R , can be determined from Equation (25) while q ˙ B ρ , ρ = R + 1 , R + 2 , , n , are random. From Equation (25), we express q ˙ B σ , σ = 1 , 2 , , R , 1 R < n , as
q ˙ B σ = f B σ t , q B , p B α , p B E , q ˙ B F ,   ( or   q ˙ B E = f B E t , q B , p B α , p B E , q ˙ B F ) ,
where p B Ε = p B 1 , p B 2 , , p B R , q ˙ B E = q ˙ B 1 , q ˙ B 2 , , q ˙ B R , q ˙ B F = q ˙ B R + 1 , q ˙ B R + 2 , , q ˙ B n , f B E = f B 1 , f B 2 , , f B R .
From Equations (25) and (28), we have
p B i = g B i t , q B , f B E t , q B , p B α , p B E , q ˙ B F , q ˙ B F , p B α = g B i t , q B , p B α , p B E , q ˙ B F ,   i = 1 , 2 , , n .
For Equation (29), if i = 1 , 2 , , R , then Equation (29) always holds. If i = R + 1 , R + 2 , , n , then from the assumption rank H B i j = R , 1 R < n , we have
p B ρ = g B ρ t , q B , p B E , p B α ,   ( or   p B F = g B F t , q B , p B E , p B α ) ,  
where ρ = R + 1 , R + 2 , , n , p B F = p B R + 1 , p B R + 2 , , p B n , g B F = g B R + 1 , g B R + 2 , , g B n , 1 R < n . Equation (30) has another form
ϕ B t , q B , p B , p B α = p B F g B F t , q B , p B E , p B α = 0 ,
where ϕ B = ϕ B 1 , ϕ B 2 , , ϕ B n R , 1 R < n .
Secondly, when R = 0 , which means that no q ˙ B i , i = 1 , 2 , , n , can be determined from Equation (25). Then from Equation (25) and the assumption rank H B i j = R , R = 0 , we have
p B i = g B i t , q B , p B α ,   ( or   p B = g B t , q B , p B α ) ,
where i = 1 , 2 , , n , p B = p B 1 , p B 2 , , p B n , g B = g B 1 , g B 2 , , g B n . Then Equation (32) gives
ϕ B t , q B , p B , p B α = p B g B t , q B , p B α = 0 ,
where ϕ B = ϕ B 1 , ϕ B 2 , , ϕ B n .
Incorporating Equations (31) and (33), we get
ϕ B t , q B , p B , p B α = 0 ,
where ϕ B = ϕ B 1 , ϕ B 2 , , ϕ B n R , 0 R < n . Equation (34) is called the primary constraint with the operator B M α .
Remark 2.
Let κ α t , τ = t τ α 1 / Γ α , when M = M 1 , M = M 2 and M = M 3 , Equations (24) and (34) give the Lagrange equations and the primary constraints in terms of the left Caputo, the right Caputo and the Riesz-Caputo fractional derivatives, respectively.
We intend to transform the singular Lagrangian systems with the mixed derivatives (Equations (10), (20), (24) and (34)) into the constrained Hamiltonian systems in the following section.

4. Constrained Hamiltonian System and Consistency Condition

4.1. Constrained Hamilton Equation with the Operator A M α

From Equations (11) and (12), we have
δ H A = q ˙ A i δ p A i + δ p A i α A M α q A i L A q A i δ q A i ,   i = 1 , 2 , , n ,
where
L A q A i = p ˙ A i + B M * α p A i α m κ 1 α t 2 , t p A i α t 2 + ω κ 1 α t , t 1 p A i α t 1
From the Hamiltonian H A = H A t , q A , p A , p A α we have,
δ H A = H A q A i δ q A i + H A p A i δ p A i + H A p A i α δ p A i α ,   i = 1 , 2 , , n .
Besides, taking isochronous variation of Equation (20), we have
δ ϕ A a = ϕ A a q A i δ q A i + ϕ A a p A i δ p A i + ϕ A a p A i α δ p A i α = 0 ,   i = 1 , 2 , , n ,   a = 1 , 2 , , n R ,   0 R < n .
Introducing the Lagrange multipliers λ A a t , a = 1 , 2 , , n R , 0 R < n , and from Equations (35)–(38), we have
p ˙ A i = B M * α p A i α H A q A i + m κ 1 α t 2 , t p A i α t 2 ω κ 1 α t , t 1 p A i α t 1 λ A a ϕ A a q A i , q ˙ A i = H A p A i + λ A a ϕ A a p A i ,   A M α q A i = H A p A i α + λ A a ϕ A a p A i α ,   i = 1 , 2 , , n ,   a = 1 , 2 , , n R ,   0 R < n .
Equation (39) is called the constrained Hamilton equation with the operator A M α .
For simplicity, we introduce H A T = H A + λ A a ϕ A a , a = 1 , 2 , , n R , 0 R < n , then Equation (39) can be written as
p ˙ A i = B M * α p A i α H A T q A i + m κ 1 α t 2 , t p A i α t 2 ω κ 1 α t , t 1 p A i α t 1 , q ˙ A i = H A T p A i ,   A M α q A i = H A T p A i α ,   i = 1 , 2 , , n ,   a = 1 , 2 , , n R ,   0 R < n .
Remark 3.
Let κ α t , τ = t τ α 1 / Γ α , when M = M 1 , M = M 2 and M = M 3 , Equation (39) (or Equation (40)) gives the constrained Hamilton equations in terms of the left Riemann-Liouville, the right Riemann-Liouville and the Riesz-Riemann-Liouville fractional derivatives, respectively.

4.2. Constrained Hamilton Equation with the Operator B M α

From Equations (25) and (26), we have
δ H B = q ˙ B i δ p B i + δ p B i α B M α q B i L B q B i δ q B i ,   i = 1 , 2 , , n ,
where
L B q B i = p ˙ B i + A M * α p B i α .
From the Hamiltonian H B = H B t , q B , p B , p B α we have,
δ H B = H B q B i δ q B i + H B p B i δ p B i + H B p B i α δ p B i α ,   i = 1 , 2 , , n .
Besides, taking isochronous variation of Equation (34), we have
δ ϕ B a = ϕ B a q B i δ q B i + ϕ B a p B i δ p B i + ϕ B a p B i α δ p B i α = 0 ,   i = 1 , 2 , , n ,   a = 1 , 2 , , n R ,   0 R < n .
Introducing the Lagrange multipliers λ B a t , a = 1 , 2 , , n R , 0 R < n , and from Equations (41)–(44), we have
p ˙ B i = A M * α p B i α H B q B i λ B a ϕ B a q B i ,   q ˙ B i = H B p B i + λ B a ϕ B a p B i , B M α q B i = H B p B i α + λ B a ϕ B a p B i α ,   i = 1 , 2 , , n ,   a = 1 , 2 , , n R ,   0 R < n .
Equation (45) is called the constrained Hamilton equation with the operator B M α .
For simplicity, we introduce H B T = H B + λ B a ϕ B a , a = 1 , 2 , , n R , 0 R < n , then Equation (45) can be written as
p ˙ B i = A M * α p B i α H B T q B i ,   q ˙ B i = H B T p B i ,   B M α q B i = H B T p B i α , i = 1 , 2 , , n ,   a = 1 , 2 , , n R ,   0 R < n .
Remark 4.
Let κ α t , τ = t τ α 1 / Γ α , when M = M 1 , M = M 2 and M = M 3 , Equation (45) (or Equation (46)) gives the constrained Hamilton equations in terms of the left Caputo, the right Caputo and the Riesz-Caputo fractional derivatives, respectively.

4.3. Consistency Conditions with Generalized Operators

Let F = F t , q , p , p α , G = G t , q , p , p α , we define the Poisson bracket as
F , G = F q i G p i F p i G q i ,   i = 1 , 2 , , n ,
where q = q 1 , q 2 , , q n , p = p 1 , p 2 , , p n , p α = p 1 α , p 2 α , , p n α .
Then using the Poisson bracket and Equation (40), we obtain
ϕ A a , H A + λ A b ϕ A a , ϕ A b + ϕ A a t + ϕ A a p A i α p ˙ A i α ϕ A a p A i B M * α p A i α m κ 1 α t 2 , t p A i α t 2 + ω κ 1 α t , t 1 p A i α t 1 = 0 ,   i = 1 , 2 , , n ,   a , b = 1 , 2 , , n R ,   0 R < n .
Equation (48) is called the consistency condition with the operator A M α .
Similarly, the consistency condition with the operator B M α has the form
ϕ B a , H B + λ B b ϕ B a , ϕ B b + ϕ B a t + ϕ B a p B i α p ˙ B i α ϕ B a p B i A M * α p B i α = 0 , i = 1 , 2 , , n ,   a , b = 1 , 2 , , n R ,   0 R < n .
If det ϕ A a , ϕ A b 0 (resp. det ϕ B a , ϕ B b 0 ), a , b = 1 , 2 , , n R , 0 R < n , then all the Lagrange multipliers λ A a (resp. λ B a ), a = 1 , 2 , , n R , 0 R < n , can be calculated from Equation (48) (resp. Equation (49)).
If det ϕ A a , ϕ A b = 0 (resp. det ϕ B a , ϕ B b = 0 ), then the Lagrange multipliers λ A a (resp. λ B a ), a = 1 , 2 , , n R , 0 R < n , cannot be calculated completely, and then the new constraint, which is called the secondary constraint, will be deduced. Therefore, the secondary constraint arises from the consistency condition of the primary constraint. Similarly, if the consistency condition of the secondary constraint still cannot give all the Lagrange multipliers, then some new secondary constraints will be established. Anyway, no new secondary constraint will be produced after a finite number of steps for a system with finite degrees of freedom.
Remark 5.
Let κ α t , τ = t τ α 1 / Γ α , when M = M 1 , M = M 2 and M = M 3 , Equations (48) and (49) give the consistency conditions in terms of the left Riemann-Liouville, the left Caputo, the right Riemann-Liouville, the right Caputo, the Riesz-Riemann-Liouville and the Riesz-Caputo fractional derivatives, respectively.

5. Noether Symmetry and Conserved Quantity

Noether symmetry is the invariance of the Hamilton action under the infinitesimal transformations of time and coordinates. We begin with Noether symmetry with the operator A M α .

5.1. Noether Symmetry with the Operator A M α

The Hamilton action with the operator A M α is
I A = t 1 t 2 p A i q ˙ A i + p A i α A M α q A i H A t , q A , p A , p A α d t ,   i = 1 , 2 , , n .
The infinitesimal transformations are
t ¯ = t + Δ t ,   q ¯ A i t ¯ = q A i t + Δ q A i ,   p ¯ A i t ¯ = p A i t + Δ p A i , p ¯ A i α t ¯ = p A i α t + Δ p A i α , ( or   t ¯ = t + Δ t ,   q ¯ A t ¯ = q A t + Δ q A ,   p ¯ A t ¯ = p A t + Δ p A , p ¯ A α t ¯ = p A α t + Δ p A α ) ,
and the expanded forms are
t ¯ = t + θ A ξ A 0 t , q A , p A , p A α + ο θ A ,   q ¯ A i t ¯ = q A i t + θ A ξ A i t , q A , p A , p A α + ο θ A , p ¯ A i t ¯ = p A i t + θ A η A i t , q A , p A , p A α + ο θ A , p ¯ A i α t ¯ = p A i α t + θ A η A i α t , q A , p A , p A α + ο θ A ,
where θ A is a small parameter, ξ A 0 , ξ A i , η A i and η A i α , i = 1 , 2 , , n , are the infinitesimal generators of the infinitesimal transformations, ο θ A is the higher order of θ A .
Neglecting the higher order of θ A , we have
Δ I A = I ¯ A I A = t ¯ 1 t ¯ 2 p ¯ A i q ¯ ˙ A i + p ¯ A i α A M ¯ α q ¯ A i H A t ¯ , q ¯ A , p ¯ A , p ¯ A α d t ¯ I A = θ A t 1 t 2 λ A a ϕ A a p A i η A i + λ A a ϕ A a p A i α η A i α H A q A i ξ A i + p A i α d d t A M α q A i H A t ξ A 0 + p A i α A M α ξ A i q ˙ A i ξ A 0 + p A i α A M α q A i H A ξ ˙ A 0 + ω p A i α q A i t 2 ξ A 0 t 2 d d t κ 1 α t 2 , t m p A i α q A i t 1 ξ A 0 t 1 d d t κ 1 α t , t 1 + p A i ξ ˙ A i d t ,  
where ξ A 0 t 1 = ξ A 0 t 1 , q A t 1 , p A t 1 , p A α t 1 , ξ A 0 t 2 = ξ A 0 t 2 , q A t 2 , p A t 2 , p A α t 2 and M ¯ = < t ¯ 1 , t ¯ , t ¯ 2 , m , ω > .
Noether symmetry requires that Δ I A = 0 , that is,
λ A a ϕ A a p A i η A i + λ A a ϕ A a p A i α η A i α H A q A i ξ A i + p A i α d d t A M α q A i H A t ξ A 0 + p A i ξ ˙ A i + p A i α A M α ξ A i q ˙ A i ξ A 0 + p A i α A M α q A i H A ξ ˙ A 0 + ω p A i α q A i t 2 ξ A 0 t 2 d d t κ 1 α t 2 , t m p A i α q A i t 1 ξ A 0 t 1 d d t κ 1 α t , t 1 = 0 .
Equation (54) is called the Noether identity with the operator A M α .
If we let Δ I A = t 1 t 2 d d t Δ G A d t , where Δ G A = θ A G A , G A = G A t , q A , p A , p A α is called a gauge function with the operator A M α , then we obtain
λ A a ϕ A a p A i η A i + λ A a ϕ A a p A i α η A i α H A q A i ξ A i + p A i α d d t A M α q A i H A t ξ A 0 + p A i ξ ˙ A i + p A i α A M α ξ A i q ˙ A i ξ A 0 + p A i α A M α q A i H A ξ ˙ A 0 + ω p A i α q A i t 2 ξ A 0 t 2 d d t κ 1 α t 2 , t m p A i α q A i t 1 ξ A 0 t 1 d d t κ 1 α t , t 1 + G ˙ A = 0 .  
Equation (55) is called the Noether quasi-identity with the operator A M α .
Noether symmetry leads to a conserved quantity. We first present the definition of the conserved quantity.
Definition 1.
A quantity C is called a conserved quantity if and only if d C / d t = 0 holds.
Therefore, we have
Theorem 1.
For the constrained Hamiltonian system with the operator A M α (Equation (39)), if the infinitesimal generators ξ A 0 , ξ A i , η A i and η A i α satisfy Equation (54), then there exists a conserved quantity
C A = p A i ξ A i + p A i α A M α q A i H A ξ A 0 + t 1 t p A i α A M α ξ A i q ˙ A i ξ A 0 + ξ A i q ˙ A i ξ A 0 B M * α p A i α m κ 1 α t 2 , τ p A i α t 2 + ω κ 1 α τ , t 1 p A i α t 1 d τ + ω q A i t 2 ξ A 0 t 2 t 1 t p A i α τ d d τ κ 1 α t 2 , τ d τ m q A i t 1 ξ A 0 t 1 t 1 t p A i α τ d d τ κ 1 α τ , t 1 d τ = const .
Proof of Theorem 1.
From Equations (20), (39) and (54), we have
d C A / d t = p ˙ A i ξ A i + p A i ξ ˙ A i + p A i α A M α q A i H A ξ ˙ A 0 + p A i α A M α ξ A i q ˙ A i ξ A 0 + p ˙ A i α A M α q A i + p A i α d d t A M α q A i H A t H A q A i q ˙ A i H A p A i p ˙ A i H A p A i α p ˙ A i α ξ A 0 + ξ A i q ˙ A i ξ A 0 B M * α p A i α m κ 1 α t 2 , t p A i α t 2 + ω κ 1 α t , t 1 p A i α t 1 + ω q A i t 2 ξ A 0 t 2 p A i α t d d t κ 1 α t 2 , t m q A i t 1 ξ A 0 t 1 p A i α t d d t κ 1 α t , t 1 = H A q A i ξ A i λ A a ϕ A a p A i η A i + p ˙ A i ξ A i + ξ A 0 p ˙ A i α A M α q A i H A q A i q ˙ A i H A p A i p ˙ A i H A p A i α p ˙ A i α + ξ A i q ˙ A i ξ A 0 B M * α p A i α m κ 1 α t 2 , t p A i α t 2 + ω κ 1 α t , t 1 p A i α t 1 λ A a ϕ A a p A i α η A i α = p ˙ A i ξ A 0 λ A a ϕ A a p A i λ A a ϕ A a q A i ξ A i q ˙ A i ξ A 0 λ A a ϕ A a p A i η A i λ A a ϕ A a p A i α η A i α + λ A a ϕ A a p A i α p ˙ A i α ξ A 0 = λ A a ϕ A a q A i δ q A i λ A a ϕ A a p A i δ p A i λ A a ϕ A a p A i α δ p A i α = λ A a δ ϕ A a = 0 .
The proof is completed. □
Theorem 2.
For the constrained Hamiltonian system with the operator A M α (Equation (39)), if there exists a gauge function G A such that the infinitesimal generators ξ A 0 , ξ A i , η A i and η A i α satisfy Equation (55), then there exists a conserved quantity
C A G = p A i ξ A i + p A i α A M α q A i H A ξ A 0 + t 1 t p A i α A M α ξ A i q ˙ A i ξ A 0 + ξ A i q ˙ A i ξ A 0 B M * α p A i α m κ 1 α t 2 , τ p A i α t 2 + ω κ 1 α τ , t 1 p A i α t 1 d τ + ω q A i t 2 ξ A 0 t 2 t 1 t p A i α τ d d τ κ 1 α t 2 , τ d τ m q A i t 1 ξ A 0 t 1 t 1 t p A i α τ d d τ κ 1 α τ , t 1 d τ + G A = const .
Proof of Theorem 2.
From Equations (20), (39) and (55), we have d C A G / d t = 0 . □
Remark 6.
Let κ α t , τ = t τ α 1 / Γ α , when M = M 1 , M = M 2 and M = M 3 , Equation (54), Equation (55), Theorem 1 and Theorem 2 give the Noether identities, Noether quasi-identities and conserved quantities in terms of the left Riemann-Liouville, the right Riemann-Liouville and the Riesz-Riemann-Liouville fractional derivatives, respectively.

5.2. Noether Symmetry with the Operator B M α

The Hamilton action with the operator B M α is
I B = t 1 t 2 p B i q ˙ B i + p B i α B M α q B i H B t , q B , p B , p B α d t ,   i = 1 , 2 , , n .
The infinitesimal transformations are
t ¯ = t + Δ t ,   q ¯ B i t ¯ = q B i t + Δ q B i ,   p ¯ B i t ¯ = p B i t + Δ p B i , p ¯ B i α t ¯ = p B i α t + Δ p B i α , ( or   t ¯ = t + Δ t ,   q ¯ B t ¯ = q B t + Δ q B ,   p ¯ B t ¯ = p B t + Δ p B ,   p ¯ B α t ¯ = p B α t + Δ p B α ) ,
and the expanded forms are
t ¯ = t + θ B ξ B 0 t , q B , p B , p B α + ο θ B ,   q ¯ B i t ¯ = q B i t + θ B ξ B i t , q B , p B , p B α + ο θ B , p ¯ B i t ¯ = p B i t + θ B η B i t , q B , p B , p B α + ο θ B , p ¯ B i α t ¯ = p B i α t + θ B η B i α t , q B , p B , p B α + ο θ B ,
where θ B is a small parameter, ξ B 0 , ξ B i , η B i and η B i α are the infinitesimal generators of the infinitesimal transformations, ο θ B is the higher order of θ B .
Neglecting the higher order of θ B , we have
Δ I B = I ¯ B I B = t ¯ 1 t ¯ 2 p ¯ B i q ¯ ˙ B i + p ¯ B i α B M ¯ α q ¯ B i H B t ¯ , q ¯ B , p ¯ B , p ¯ B α d t ¯ I B = θ B t 1 t 2 λ B a ϕ B a p B i η B i + λ B a ϕ B a p B i α η B i α H B q B i ξ B i + p B i α d d t B M α q B i H B t ξ B 0 + p B i α B M α ξ B i q ˙ B i ξ B 0 + p B i α B M α q B i H B ξ ˙ B 0 + ω p B i α κ 1 α t 2 , t q ˙ B i t 2 ξ B 0 t 2 m p B i α κ 1 α t , t 1 q ˙ B i t 1 ξ B 0 t 1 + p B i ξ ˙ B i d t ,
where ξ B 0 t 1 = ξ B 0 t 1 , q B t 1 , p B t 1 , p B α t 1 , ξ B 0 t 2 = ξ B 0 t 2 , q B t 2 , p B t 2 , p B α t 2 .
Noether symmetry requires that Δ I B = 0 , that is,
λ B a ϕ B a p B i η B i + λ B a ϕ B a p B i α η B i α H B q B i ξ B i + p B i α d d t B M α q B i H B t ξ B 0 + p B i ξ ˙ B i + p B i α B M α ξ B i q ˙ B i ξ B 0 + p B i α B M α q B i H B ξ ˙ B 0 + ω p B i α κ 1 α t 2 , t q ˙ B i t 2 ξ B 0 t 2 m p B i α κ 1 α t , t 1 q ˙ B i t 1 ξ B 0 t 1 = 0 .
Equation (62) is called the Noether identity with the operator B M α .
If we let Δ I B = t 1 t 2 d d t Δ G B d t , where Δ G B = θ B G B , G B = G B t , q B , p B , p B α is called a gauge function with the operator B M α , then we obtain
λ B a ϕ B a p B i η B i + λ B a ϕ B a p B i α η B i α H B q B i ξ B i + p B i α d d t B M α q B i H B t ξ B 0 + p B i ξ ˙ B i + p B i α B M α ξ B i q ˙ B i ξ B 0 + p B i α B M α q B i H B ξ ˙ B 0 + ω p B i α κ 1 α t 2 , t q ˙ B i t 2 ξ B 0 t 2 m p B i α κ 1 α t , t 1 q ˙ B i t 1 ξ B 0 t 1 + G ˙ B = 0 .
Equation (63) is called the Noether quasi-identity with the operator B M α . Therefore, we have
Theorem 3 .
For the constrained Hamiltonian system with the operator B M α (Equation (45)), if the infinitesimal generators ξ B 0 , ξ B i , η B i and η B i α satisfy Equation (62), then there exists a conserved quantity
C B = p B i ξ B i + p B i α B M α q B i H B ξ B 0 + t 1 t p B i α B M α ξ B i q ˙ B i ξ B 0 + ξ B i q ˙ B i ξ B 0 A M * α p B i α d τ + ω q ˙ B i t 2 ξ B 0 t 2 t 1 t p B i α τ κ 1 α t 2 , τ d τ m q ˙ B i t 1 ξ B 0 t 1 t 1 t p B i α τ κ 1 α τ , t 1 d τ = const .
Proof of Theorem 3.
From Equations (34), (45) and (62), we have d C B / d t = 0 . □
Theorem 4.
For the constrained Hamiltonian system with the operator B M α (Equation (45)), if there exists a gauge function G B such that the infinitesimal generators ξ B 0 , ξ B i , η B i and η B i α satisfy Equation (63), then there exists a conserved quantity
C B G = p B i ξ B i + t 1 t p B i α B M α ξ B i q ˙ B i ξ B 0 + ξ B i q ˙ B i ξ B 0 A M * α p B i α d τ + ω q ˙ B i t 2 ξ B 0 t 2 t 1 t p B i α τ κ 1 α t 2 , τ d τ + p B i α B M α q B i H B ξ B 0 m q ˙ B i t 1 ξ B 0 t 1 t 1 t p B i α τ κ 1 α τ , t 1 d τ + G B = const .
Proof of Theorem 4.
From Equations (34), (45) and (63), we have d C B G / d t = 0 . □
Remark 7.
Let κ α t , τ = t τ α 1 / Γ α , when M = M 1 , M = M 2 and M = M 3 , Equation (62), Equation (63), Theorem 3 and Theorem 4 give the Noether identities, the Noether quasi-identities and the conserved quantities in terms of the left Caputo, the right Caputo and the Riesz-Caputo fractional derivatives, respectively.
Remark 8.
When the gauge function G A = 0 (resp. G B = 0 ), Theorem 2 (resp. Theorem 4) reduces to Theorem 1 (resp. Theorem 3). Hence, the Noether-quasi symmetry is more general than the Noether symmetry.

6. Lie Symmetry and Conserved Quantity

Lie symmetry means an invariance of the differential equations under the infinitesimal transformations of time and coordinates. Lie symmetry can also lead to a conserved quantity under certain conditions.

6.1. Lie Symmetry with the Operator A M α

We rewrite the constrained Hamilton equation with the operator A M α (Equation (39)) as
p ˙ A i = B M * α p A i α + f A i t , q A , p A , p A α ,   q ˙ A i = S A i t , q A , p A , p A α , A M α q A i = h A i t , q A , p A , p A α ,   i = 1 , 2 , , n .  
Then under the condition κ 1 α t , t = 0 , we have
q ¯ ˙ A i S A i t ¯ , q ¯ A , p ¯ A , p ¯ A α = d q ¯ A i d t ¯ S A i t + Δ t , q A + Δ q A , p A + Δ p A , p A α + Δ p A α = d q A i + d Δ q A i d t + d Δ t S A i t + Δ t , q A + Δ q A , p A + Δ p A , p A α + Δ p A α = d q A i d t + d Δ q A i d t 1 d Δ t d t 1 + d Δ t d t 1 d Δ t d t S A i t + Δ t , q A + Δ q A , p A + Δ p A , p A α + Δ p A α = q ˙ A i + d Δ q A i d t q ˙ A i d Δ t d t S A i t , q A , p A , p A α S A i t Δ t S A i q A k Δ q A k S A i p A k Δ p A k S A i p A k α Δ p A k α = q ˙ A i S A i t , q A , p A , p A α + θ A ξ ˙ A i q ˙ A i ξ ˙ A 0 X A 0 S A i ,
where X A 0 = ξ A 0 t + ξ A k q A k + η A k p A k + η A k α p A k α , k = 1 , 2 , , n .
A M α q ¯ A i t ¯ h A i t ¯ , q ¯ A , p ¯ A , p ¯ A α = A M α q ¯ A i t ¯ h A i t + Δ t , q A + Δ q A , p A + Δ p A , p A α + Δ p A α = A M α q A i + A M α δ q A i + Δ t d d t A M α q A i + d d t m κ 1 α t , t 1 Δ t 1 q A i t 1 + ω κ 1 α t 2 , t Δ t 2 q A i t 2 h A i t , q A , p A , p A α h A i t Δ t h A i q A k Δ q A k h A i p A k Δ p A k h A i p A k α Δ p A k α = A M α q A i h A i t , q A , p A , p A α + θ A A M α ξ A i q ˙ A i ξ A 0 + ξ A 0 d d t A M α q A i + d d t m κ 1 α t , t 1 ξ A 0 t 1 q A i t 1 + ω κ 1 α t 2 , t ξ A 0 t 2 q A i t 2 X A 0 h A i ,  
where
ξ A 0 t 1 = ξ A 0 t 1 , q A t 1 , p A t 1 , p A α t 1 ,   ξ A 0 t 2 = ξ A 0 t 2 , q A t 2 , p A t 2 , p A α t 2 , A M α q ¯ A i t ¯ = A M α q A i + Δ t d d t A M α q A i + d d t m κ 1 α t , t 1 Δ t 1 q A i t 1 + ω κ 1 α t 2 , t Δ t 2 q A i t 2 + A M α δ q A i . p ¯ ˙ A i + B M * α p ¯ A i α t ¯ f A i t ¯ , q ¯ A , p ¯ A , p ¯ A α = d p A i + d Δ p A i d t + d Δ t + B M * α p ¯ A i α t ¯ f A i t ¯ , q ¯ A , p ¯ A , p ¯ A α = d p A i d t + d Δ p A i d t 1 d Δ t d t 1 + d Δ t d t 1 d Δ t d t + B M * α p ¯ A i α f A i t + Δ t , q A + Δ q A , p A + Δ p A , p A α + Δ p A α = p ˙ A i + B M * α p A i α f A i t , q A , p A , p A α + θ A η ˙ A i p ˙ A i ξ ˙ A 0 + B M * α η A i α p ˙ A i α ξ A 0 + ξ A 0 d d t B M * α p A i α ω κ 1 α t , t 1 p ˙ A i α t 1 ξ A 0 t 1 + m κ 1 α t 2 , t p ˙ A i α t 2 ξ A 0 t 2 X A 0 f A i ,  
where
B M * α p ¯ A i α = B M * α p A i α + B M * α δ p A i α + Δ t d d t B M * α p A i α ω κ 1 α t , t 1 p ˙ A i α t 1 Δ t 1 + m κ 1 α t 2 , t p ˙ A i α t 2 Δ t 2 .
Lie symmetry requires that
ξ ˙ A i q ˙ A i ξ ˙ A 0 X A 0 S A i = 0 , A M α ξ A i q ˙ A i ξ A 0 + d d t m κ 1 α t , t 1 ξ A 0 t 1 q A i t 1 + ω κ 1 α t 2 , t ξ A 0 t 2 q A i t 2 + ξ A 0 d d t A M α q A i X A 0 h A i = 0 , η ˙ A i p ˙ A i ξ ˙ A 0 + B M * α η A i α p ˙ A i α ξ A 0 + ξ A 0 d d t B M * α p A i α ω κ 1 α t , t 1 p ˙ A i α t 1 ξ A 0 t 1 + m κ 1 α t 2 , t p ˙ A i α t 2 ξ A 0 t 2 X A 0 f A i = 0
Equation (70) is called the determined equation for the constrained Hamiltonian system with the operator A M α (Equation (39)). Lie symmetry also leads to a conserved quantity under certain conditions. Therefore, we have
Theorem 5.
For the constrained Hamiltonian system with the operator A M α (Equation (39)), if the infinitesimal generators ξ A 0 , ξ A i , η A i and η A i α satisfy Equation (54) and the determined equation (Equation (70)), then there exists a conserved quantity Equation (56).
Proof of Theorem 5.
From Equations (20), (39) and (54), we can get the intended result. □

6.2. Lie Symmetry with the Operator B M α

We rewrite the constrained Hamilton equation with the operator B M α (Equation (45)) as
p ˙ B i = A M * α p B i α + f B i t , q B , p B , p B α ,   q ˙ B i = S B i t , q B , p B , p B α , B M α q B i = h B i t , q B , p B , p B α ,   i = 1 , 2 , , n
Then under the condition κ 1 α t , t = 0 , we have
q ¯ ˙ B i S B i t ¯ , q ¯ B , p ¯ B , p ¯ B α = d q ¯ B i d t ¯ S B i t + Δ t , q B + Δ q B , p B + Δ p B , p B α + Δ p B α = d q B i + d Δ q B i d t + d Δ t S B i t + Δ t , q B + Δ q B , p B + Δ p B , p B α + Δ p B α = q ˙ B i S B i t , q B , p B , p B α + θ B ξ ˙ B i q ˙ B i ξ ˙ B 0 X B 0 S B i ,
where
X B 0 = ξ B 0 t + ξ B k q B k + η B k p B k + η B k α p B k α ,   k = 1 , 2 , , n . B M α q ¯ B i t ¯ h B i t ¯ , q ¯ B , p ¯ B , p ¯ B α = B M α q ¯ B i t ¯ h B i t + Δ t , q B + Δ q B , p B + Δ p B , p B α + Δ p B α = B M α q B i h B i t , q B , p B , p B α + θ B B M α ξ B i q ˙ B i ξ B 0 + ξ B 0 d d t B M α q B i m κ 1 α t , t 1 ξ B 0 t 1 q ˙ B i t 1 + ω κ 1 α t 2 , t ξ B 0 t 2 q ˙ B i t 2 X B 0 h B i ,
where
ξ B 0 t 1 = ξ B 0 t 1 , q B t 1 , p B t 1 , p B α t 1 ,   ξ B 0 t 2 = ξ B 0 t 2 , q B t 2 , p B t 2 , p B α t 2 . p ¯ ˙ B i + A M * α p ¯ B i α t ¯ f B i t ¯ , q ¯ B , p ¯ B , p ¯ B α = d p B i + d Δ p B i d t + d Δ t + A M * α p ¯ B i α t ¯ f B i t ¯ , q ¯ B , p ¯ B , p ¯ B α = p ˙ B i + A M * α p B i α f B i t , q B , p B , p B α + θ B η ˙ B i p ˙ B i ξ ˙ B 0 + A M * α η B i α p ˙ B i α ξ B 0 + ξ B 0 d d t A M * α p B i α + d d t ω κ 1 α t , t 1 p B i α t 1 ξ B 0 t 1 + m κ 1 α t 2 , t p B i α t 2 ξ B 0 t 2 X 0 f B i .
Lie symmetry requires that
ξ ˙ B i q ˙ B i ξ ˙ B 0 X B 0 S B i = 0 , B M α ξ B i q ˙ B i ξ B 0 + ξ B 0 d d t B M α q B i m κ 1 α t , t 1 ξ B 0 t 1 q ˙ B i t 1 + ω κ 1 α t 2 , t ξ B 0 t 2 q ˙ B i t 2 X B 0 h B i = 0 , η ˙ B i p ˙ B i ξ ˙ B 0 + A M * α η B i α p ˙ B i α ξ B 0 + ξ B 0 d d t A M * α p B i α X 0 f B i + d d t ω κ 1 α t , t 1 p B i α t 1 ξ B 0 t 1 + m κ 1 α t 2 , t p B i α t 2 ξ B 0 t 2 = 0 .
Equation (75) is called the determined equation for the constrained Hamiltonian system with the operator B M α (Equation (45)). Lie symmetry also leads to a conserved quantity under certain conditions. Therefore, we have
Theorem 6.
For the constrained Hamiltonian system with the operator B M α (Equation (45)), if the infinitesimal generators ξ B 0 , ξ B i , η B i and η B i α satisfy Equation (62) and the determined equation (Equation (75), then there exists a conserved quantity Equation (64).
Proof of Theorem 6.
From Equations (34), (45) and (62), we can get the intended result. □

7. An Example

Try to find the conserved quantity for the following singular system with the operator A M α , whose Lagrangian is
L A = q ˙ A 1 q A 2 q A 1 q ˙ A 2 + q A 1 2 + q A 2 2 + 1 2 A M α q A 1 2 + A M α q A 2 2 .
From Equations (11) and (12), we have
p A 1 = L A q ˙ A 1 = q A 2 ,   p A 2 = L A q ˙ A 2 = q A 1 ,   p A 1 α = L A A M α q A 1 = A M α q A 1 ,   p A 2 α = L A A M α q A 2 = A M α q A 2 , H A = p A 1 q ˙ A 1 + p A 2 q ˙ A 2 + p A 1 α A M α q A 1 + p A 2 α A M α q A 2 L A = 1 2 A M α q A 1 2 + A M α q A 2 2 q A 1 2 q A 2 2 = 1 2 p A 1 α 2 + p A 2 α 2 q A 1 2 q A 2 2 .  
Therefore, the Hamiltonian and the two primary constraints have the form
H A = 1 2 p A 1 α 2 + p A 2 α 2 q A 1 2 q A 2 2 ,
ϕ A 1 = p A 1 q A 2 = 0 ,   ϕ A 2 = p A 2 + q A 1 = 0 .
Then from Equation (48), the two Lagrange multipliers λ A 1 and λ A 2 can be calculated as
λ A 1 = q A 2 + 1 2 B M * α p A 2 α m p A 2 α t 2 κ 1 α t 2 , t + ω p A 2 α t 1 κ 1 α t , t 1 , λ A 2 = q A 1 1 2 B M * α p A 1 α m p A 1 α t 2 κ 1 α t 2 , t + ω p A 1 α t 1 κ 1 α t , t 1 .
Therefore, the constrained Hamilton equation with the operator A M α can be obtained. And we can also verify that under the condition d d t κ α t , τ = d d τ κ α t , τ ,
ξ A 0 = 1 ,   ξ A 1 = ξ A 2 = 0 ,   η A 1 = η A 2 = 0 ,   η A 1 α = η A 2 α = 0 ,   G A = 0
is a solution to the Noether quasi-identity (Equation (55)). Then Theorem 2 gives
C A G = 1 2 p A 1 α 2 + p A 2 α 2 + q A 1 2 + q A 2 2 t 1 t p A 1 α d d t A M α q A 1 + p A 2 α d d t A M α q A 2 + q ˙ A 1 B M * α p A 1 α m κ 1 α t 2 , τ p A 1 α t 2 + ω κ 1 α τ , t 1 p A 1 α t 1 + q ˙ A 2 B M * α p A 2 α m κ 1 α t 2 , τ p A 2 α t 2 + ω κ 1 α τ , t 1 p A 2 α t 1 d τ = const .
Specially, let κ α t , τ = t τ α 1 / Γ α , when M = M 1 (or M = M 2 or M = M 3 ) and α 1 , we have C A G = H A = const .

8. Results and Discussion

Based on the mixed integer order derivative and generalized operators, the singular Lagrange equations, primary constraints, constrained Hamilton equations, consistency conditions and conserved quantities were investigated. All are new works. In fact, Lie symmetry can lead to the Noether type conserved quantity as well as the Hojman conserved quantity. Here, we only presented the Noether type conserved quantity simply. Next, Lie symmetry and the Hojman conserved quantity, Mei symmetry and the Mei type conserved quantity and the relationships among the three symmetry methods will be studied. Singular systems on time scales is also a hot topic that needs to be investigated further.

Funding

This research was funded by the National Natural Science Foundation of China, grant numbers 12172241, 11802193, 11972241, the Natural Science Foundation of Jiangsu Province, grant number BK20191454 and the “Qinglan Project” of Jiangsu Province.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kusnezov, D.; Bulgac, A.; Dang, G.D. Quantum Lévy processes and fractional kinetics. Phys. Rev. Lett. 1999, 82, 1136–1139. [Google Scholar] [CrossRef] [Green Version]
  2. Naber, M. Time fractional Schrdinger equation. J. Math. Phys. 2004, 45, 3339–3352. [Google Scholar] [CrossRef]
  3. Muslih, S.I.; Agrawal, O.P.; Baleanu, D. A fractional Dirac equation and its solution. J. Phys. A Math. Theor. 2010, 43, 055203. [Google Scholar] [CrossRef]
  4. Metzler, R.; Klafter, J. The random walk’s guide to anomalous diffusion: A fractional dynamics approach. Phys. Rep. 2000, 339, 1–77. [Google Scholar] [CrossRef]
  5. Herrmann, R. Gauge invariance in fractional field theories. Phys. Lett. A 2008, 372, 5515–5522. [Google Scholar] [CrossRef] [Green Version]
  6. Lazo, M.J. Gauge invariant fractional electromagnetic fields. Phys. Lett. A 2011, 375, 3541–3546. [Google Scholar] [CrossRef] [Green Version]
  7. Wu, Q.; Huang, J.H. Fractional Calculus; Tsinghua University Press: Beijing, China, 2016. [Google Scholar]
  8. Riewe, F. Nonconservative Lagrangian and Hamiltonian mechanics. Phys. Rev. E 1996, 53, 1890–1899. [Google Scholar] [CrossRef] [PubMed]
  9. Riewe, F. Mechanics with fractional derivatives. Phys. Rev. E 1997, 55, 3581–3592. [Google Scholar] [CrossRef]
  10. Klimek, M. Lagrangian and Hamiltonian fractional sequential mechanics. Czechoslov. J. Phys. 2002, 52, 1247–1253. [Google Scholar] [CrossRef]
  11. Muslih, S.I.; Baleanu, D. Hamiltonian formulation of systems with linear velocities within Riemann–Liouville fractional derivatives. J. Math. Anal. Appl. 2005, 304, 599–606. [Google Scholar] [CrossRef] [Green Version]
  12. Agrawal, O.P. Fractional variational calculus in terms of Riesz fractional derivatives. J. Phys. A Math. Theor. 2007, 40, 6287–6303. [Google Scholar] [CrossRef]
  13. Luo, S.K.; Xu, Y.L. Fractional Birkhoffian mechanics. Acta Mech. 2015, 226, 829–844. [Google Scholar] [CrossRef]
  14. Song, C.J.; Agrawal, O.P. Hamiltonian formulation of systems described using fractional singular Lagrangian. Acta Appl. Math. 2021, 172, 9. [Google Scholar] [CrossRef]
  15. Rabei, E.M.; Nawafleh, K.I.; Hijjawi, R.S.; Muslih, S.I.; Baleanu, D. The Hamilton formalism with fractional derivatives. J. Math. Anal. Appl. 2007, 327, 891–897. [Google Scholar] [CrossRef]
  16. Malinowska, A.B.; Torres, D.F.M. Introduction to the Fractional Calculus of Variations; Imp. Coll. Press: London, UK, 2012. [Google Scholar]
  17. Agrawal, O.P. Generalized variational problems and Euler-Lagrange equations. Comput. Math. Appl. 2010, 59, 1852–1864. [Google Scholar] [CrossRef] [Green Version]
  18. Li, Z.P. Canonical symmetry and Dirac conjecture for constrained system. In Thirty Years of Nonholonomic Mechanics in China; Chen, B., Mei, F.X., Eds.; Henan University Press: Kaifeng, China, 1994. [Google Scholar]
  19. Li, Z.P. Classical and Quantal Dynamics of Contrained Systems and Their Symmetrical Properties; Beijing Polytechnic University Press: Beijing, China, 1993. [Google Scholar]
  20. Cao, S. Canonicalization and Symmetry Theories of the Constrained Hamiltonian System. Master’s Thesis, Zhejiang Sci-Tech University, Hangzhou, China, 2017. [Google Scholar]
  21. Li, Z.P. Contrained Hamiltonian Systems and Their Symmetrical Properties; Beijing Polytechnic University Press: Beijing, China, 1999. [Google Scholar]
  22. Li, Z.P.; Jiang, J.H. Symmetries in Constrained Canonical Systems; Science Press: Beijing, China, 2002. [Google Scholar]
  23. Mei, F.X.; Wu, H.B.; Zhang, Y.F. Symmetries and conserved quantities of constrained mechanical systems. Int. J. Dynam. Control 2014, 2, 285–303. [Google Scholar] [CrossRef] [Green Version]
  24. Noether, A.E. Invariante Variationsprobleme. Nachr. Akad. Wiss. Gött. Math-Phys. 1918, KI, 235–258. [Google Scholar]
  25. Song, C.J.; Zhang, Y. Conserved quantities for Hamiltonian systems on time scales. Appl. Math. Comput. 2017, 313, 24–36. [Google Scholar] [CrossRef]
  26. Song, C.J.; Zhang, Y. Noether symmetry and conserved quantity for fractional Birkhoffian mechanics and its applications. Fract. Calc. Appl. Anal. 2018, 21, 509–526. [Google Scholar] [CrossRef]
  27. Zhang, Y.; Tian, X. Conservation laws of nonholonomic nonconservative system based on Herglotz variational problems. Phys. Lett. A 2019, 383, 691–696. [Google Scholar] [CrossRef]
  28. Zhang, Y. Lie symmetry and invariants for a generalized Birkhoffian system on time scales. Chaos Solitons Fractals 2019, 128, 306–312. [Google Scholar] [CrossRef]
  29. Ding, J.J.; Zhang, Y. Conserved quantities and adiabatic invariants of fractional Birkhoffian system of Herglotz type. Chin. Phys. B 2020, 29, 044501. [Google Scholar] [CrossRef]
  30. Zhou, Y.; Zhang, Y. Noether symmetries for fractional generalized Birkhoffian systems in terms of classical and combined Caputo derivatives. Acta Mech. 2020, 231, 3017–3029. [Google Scholar] [CrossRef]
  31. Mei, F.X. Analytical Mechanics (II); Beijing Institute of Technology Press: Beijing, China, 2013. [Google Scholar]
  32. Mei, F.X. Symmetry and Conserved Quantity of Constrained Mechanical Systems; Beijing Institute of Technology Press: Beijing, China, 2004. [Google Scholar]
  33. Mei, F.X.; Wu, H.B. Dynamics of Constrained Mechanical Systems; Beijing Institute of Technology Press: Beijing, China, 2009. [Google Scholar]
  34. Cai, P.P.; Fu, J.L.; Guo, Y.X. Lie symmetries and conserved quantities of the constraint mechanical systems on time scales. Rep. Math. Phys. 2017, 79, 279–298. [Google Scholar] [CrossRef]
  35. Han, Y.L.; Wang, X.X.; Zhang, M.L.; Jia, L.Q. Lie symmetry and approximate Hojman conserved quantity of Appell equations for a weakly nonholonomic system. Nonlinear Dyn. 2013, 71, 401–408. [Google Scholar] [CrossRef] [Green Version]
  36. Chen, X.W.; Li, Y.M.; Zhao, Y.H. Lie symmetries, perturbation to symmetries and adiabatic invariants of Lagrange system. Phys. Lett. A 2005, 337, 274–278. [Google Scholar] [CrossRef]
  37. Ding, N.; Fang, J.H. Lie symmetry and conserved quantities for nonholonomic Vacco dynamical systems. Commun. Theor. Phys. 2006, 46, 265–268. [Google Scholar]
  38. Zhang, Y.; Xun, Y. Lie symmetries of constrained Hamiltonian system with the second type of constraints. Acta Phys. Sin. 2001, 50, 816–819. [Google Scholar] [CrossRef]
  39. Song, C.J.; Shen, S.L. Noether symmetry method for Birkhoffian systems in terms of generalized fractional operators. Theor. Appl. Mech. Lett. 2021, 11, 100298. [Google Scholar] [CrossRef]
  40. Jia, Q.L.; Wu, H.B.; Mei, F.X. Noether symmetries and conserved quantities for fractional forced Birkhoffian systems. J. Math. Anal. Appl. 2016, 442, 782–795. [Google Scholar] [CrossRef]
  41. Zhou, S.; Fu, H.; Fu, J.L. Symmetry theories of Hamiltonian systems with fractional derivatives. Sci. China Phys. Mech. Astron. 2011, 54, 1847–1853. [Google Scholar]
  42. Song, C.J.; Cheng, Y. Noether symmetry method for Hamiltonian mechanics involving generalized operators. Adv. Math. Phys. 2021, 2021, 1959643. [Google Scholar] [CrossRef]
  43. Song, C.J.; Cheng, Y. Noether’s theorems for nonshifted dynamic systems on time scales. Appl. Math. Comput. 2020, 374, 125086. [Google Scholar] [CrossRef]
  44. Zhang, Y.; Zhai, X.H. Perturbation to Lie symmetry and adiabatic invariants for Birkhoffian systems on time scales. Commun. Nonlinear Sci. Numer. Simul. 2019, 75, 251–261. [Google Scholar] [CrossRef]
  45. Song, C.J.; Cheng, Y. Conserved quantity and adiabatic invariant for Hamiltonian system with variable order. Symmetry 2019, 11, 1270. [Google Scholar] [CrossRef] [Green Version]
  46. Zhang, Y. Herglotz’s variational problem for non-conservative system with delayed arguments under Lagrangian framework and its Noether’s Theorem. Symmetry 2020, 12, 845. [Google Scholar] [CrossRef]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Song, C.-J. Noether and Lie Symmetry for Singular Systems Involving Mixed Derivatives. Symmetry 2022, 14, 1225. https://doi.org/10.3390/sym14061225

AMA Style

Song C-J. Noether and Lie Symmetry for Singular Systems Involving Mixed Derivatives. Symmetry. 2022; 14(6):1225. https://doi.org/10.3390/sym14061225

Chicago/Turabian Style

Song, Chuan-Jing. 2022. "Noether and Lie Symmetry for Singular Systems Involving Mixed Derivatives" Symmetry 14, no. 6: 1225. https://doi.org/10.3390/sym14061225

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