<<

arXiv:1503.06789v1 [physics.class-ph] 23 Mar 2015 uewt h aitna eg,frashrclysmercHmloinw Hamiltonian symmetric commut spherically mutually a of for of Schr¨o set (e.g., eigenfunctions the complete Hamiltonian solve a the to of with employed mute eigenfunctions methods the common of the one for to analogous is rem . the on tha explicitly fact depend the that emphasizing examples, illustrative of some give we 3 Section oso rce aihs hn epoeta,i odtos(1)– conditions if that, prove we of Then, vanishes. Poisson hwta w functions, two that show digto adding sacmo iefnto ftefunctions the of eigenfunction common a is where lsrt h tnadfraimo unu ehnc.Specifically mechanics. quantum of formalism standard the indices). to repeated closer over summation is there follows, what in and and stedffrnilo oefunction, some of differential the is ota,lclya es,w a xrs the express can we least, at locally that, so 1 ,3 ,5 6]. 5, eq 4, (HJ) 3, Hamilton–Jacobi 2, the of [1, solution complete a then bracket), ht o ehnclsse with system mechanical involution, a mechan classical for of formulation that, Hamiltonian the of framework the In Introduction 1 h iuil hoe sapolmo omneigenfunctions common of problem a as theorem Liouville The F h ttmn fteLovletermpeetdhr losu os to us allows here presented theorem Liouville the of statement The nScin2w rsn h ento fteegnucin fafun a of eigenfunctions the of definition the present we 2 Section In h i fti ae st hwta h iuil hoe a eformu be can theorem Liouville the that show to is paper this of aim The oepeiey if precisely, More 1 F , efruae stepolmo nigcmo eigenfunction common finding of problem of the solutions as on theorem formulated Liouville be the , phase extended nouin where involution, { 2 , F , . . . , ti hw ht yaporaeydfiigteeigenfunctio the defining appropriately by that, shown is It } S F eoe h oso rce wt h convention the (with bracket Poisson the denotes naporaefnto of function appropriate an nvria u´nm ePel,750Pel,Pe,M´exico Pue., Puebla, 72570 Puebla, Aut´onoma de Universidad 1 F , eatmnod ´sc ae´tc,Isiuod Ciencias de F´ısica Matem´atica, Instituto de Departamento n 2 s pt nadtv ucinof function additive an to up is, L F , . . . , 2 , F L n 1 z ( n stenme fdgeso reo ftesystem. the of freedom of degrees of number the is q and f i p , ta is, (that ( q i i H t , p , p ∂F i ) ). ∂t i ( F , . . . , { t , q F j i and ) F , i + ..Tre e Castillo del Torres G.F. F , { S F j { ( n t , j i q n F } F , i t ( d ) uut2,2021 20, August i ere ffedm fw have we if freedom, of degrees t , g q H , 0 = ny n ban opeeslto fteH equation. HJ the of solution complete a obtains one only, ( det j i ,wihi opeeslto fteH qain(here equation HJ the of solution complete a is which ), q q p , } i i F p } p , ,j i, , i − for 0 = Abstract i  1 t , 0 = F , i sfntosof functions as H ∂F t , ∂p are ) 2 1 ,hv omnegnucin fadol ftheir if only and if eigenfunctions common have ), i , t j F , . . . , i q  ny opeeslto fteH qain In equation. HJ the of solution complete a only, i p , n 0 6= ,j i, 1 = i ( osat fmto nivlto,ta is, that involution, in motion of constants q 1 = n j , , 1 = F , 2 acnett edfie eo)te,by then, below) defined be to concept (a n, , . . . , , j 2 t , , h aitnJcb qaincan equation Hamilton–Jacobi the q { n , . . . , 2 j ) q n , . . . , F , t , i so ucindfie nthe on defined function a of ns p , aincnb on yquadrature by found be can uation 3 od omneigenfunction common a hold, (3) of s  j j c,teLovletermasserts theorem Liouville the ics, d and , } igreuto,weew look we where equation, dinger t ecnmk s fconstants of use make can we t esalso htif that show shall we , = n where , n prtr htas com- also that operators ing osat fmto in motion of constants δ eta h iuil theo- Liouville the that ee t ij n h ieeta form differential the , ,te,asmn that assuming then, ), ction osat fmto in motion of constants osdrtecommon the consider e ae naohrform, another in lated { , f ( } q i stePoisson the is p , i t , n we and ) S ( q i (2) (1) (4) (3) t , ) 2 Eigenfunctions of a function and complete solutions of the Hamilton–Jacobi equation

We start by giving the definition of the eigenfunctions of a real-valued function f(qi,pi,t): We shall say that S(qi,t) is an eigenfunction of f(qi,pi,t), with eigenvalue λ, if S is a solution of the first-order partial differential equation ∂S f(qi, ,t)= λ. (5) ∂qi (It may be noticed that if f is a time-independent Hamiltonian, then (5) is the corresponding time- independent HJ equation.) We note that if S(qi,t) is an eigenfunction of f(qi,pi,t) with eigenvalue λ, then so it is S(qi,t)+φ(t), for any function φ(t) of t only, and that the solutions of (5) will depend parametrically on λ. Of course, in order for (5) to be a differential equation, f must depend on one of the pi, at least. For instance, according to this definition, the eigenfunctions of the function

F (q,p,t)= mωq sin ωt + p cos ωt, (6) where m and ω are constants, are the solutions of the differential equation ∂S mωq sin ωt + cos ωt = λ, ∂q which can be readily integrated giving mω S = λq sec ωt − q2 tan ωt + φ(t), (7) 2 where φ(t) is an arbitrary function of t only. Note that λ may be a function of t. If S(qi,t) is a common eigenfunction of f(qi,pi,t) and g(qi,pi,t), with eigenvalues λ and µ, respectively, that is, S satisfies (5) and

∂S g(qi, ,t)= µ, ∂qi then, differentiating with respect to qi, making use of the , we obtain

∂f ∂f ∂2S ∂g ∂g ∂2S + = 0 and + =0, ∂qi ∂pj ∂qj ∂qi ∂qi ∂pj ∂qj∂qi hence,

∂f ∂g ∂g ∂f ∂f ∂2S ∂g ∂g ∂2S ∂f {f,g} = − = − + ∂qi ∂pi ∂qi ∂pi ∂pj ∂qj ∂qi ∂pi ∂pj ∂qj ∂qi ∂pi ∂2S ∂2S ∂f ∂g = − =0. ∂q ∂q ∂q ∂q ∂p ∂p  j i i j  i j

Thus, if f(qi,pi,t) and g(qi,pi,t) possess common eigenfunctions, then {f,g} = 0. In order to see that the converse is also true, we now assume that {f,g} = 0. If f and g are functionally independent, then there exists, locally at least, a set of , Qi, Pi, such that, P1 = f and P2 = g. Then, the eigenvalue equations for f and g are ∂S/∂Q1 = λ and ∂S/∂Q2 = µ, which have the simultaneous solutions S = λQ1 + µQ2 + φ(Q3,...,Qn,t), where φ is an arbitrary function of n − 1 variables, thus showing that f and g have common eigenfunctions. (Note that this does not mean that every eigenfunction of f is an eigenfunction of g (cf. [7], Sec. 2.9).) The expression for S in terms of the original coordinates, (qi,t), is not given by the simple substitution of the Qi as functions of (qi,pi,t) [8]; what is relevant here is the existence of common eigenfunctions for f and g. In the case where f and g are functionally dependent, the eigenvalue equations for f and g are equivalent to each other and, trivially, possess common solutions.

2 2.1 Alternative formulation of the Liouville theorem

We now assume that F1,...,Fn are n functions satisfying (1)–(3), and we consider a common eigenfunction S(qi,t) of F1,...,Fn, with eigenvalues λ1,...,λn, respectively, then, assuming that the eigenvalues are constant, differentiating with respect to t both sides of the equation ∂S Fi(qj , ,t)= λi, (8) ∂qj making use of the chain rule, the Hamilton equations and (1), we have

∂Fi ∂Fi d ∂S ∂Fi 0 = q˙j + + ∂qj ∂pj dt ∂qj ∂t ∂F ∂H ∂F ∂2S ∂2S ∂F ∂H ∂F ∂H = i + i + q˙ − i + i ∂q ∂p ∂p ∂t∂q ∂q ∂q k ∂q ∂p ∂p ∂q j j j  j k j  j j j j ∂F ∂2S ∂2S ∂H ∂H = i + + , i =1, . . . , n. ∂p ∂t∂q ∂q ∂q ∂p ∂q j  j k j k j  By virtue of (3), the last equations are equivalent to

∂2S ∂2S ∂H ∂H 0 = + + ∂t∂qj ∂qk∂qj ∂pk ∂qj ∂ ∂S ∂S = + H(q , ,t) , j =1, . . . , n, ∂q ∂t k ∂q j  k  which implies that the expression inside the brackets is a function of t only, ∂S ∂S + H(qk, ,t)= χ(t). ∂t ∂qk Thus, t S˜ = S − χ(u) du Z is a solution of the HJ equation. We can verify that this solution is complete by differentiating (8) with respect to λj , which gives 2 ∂Fi ∂ S = δij . ∂pk ∂λj ∂qk 2 Taking into account (3), this last equation shows that det(∂ S/∂λj∂qk) 6= 0.

3 Examples

In this section we give some examples of the method presented above.

3.1 One-dimensional The function F (q,p,t)= mωq sin ωt + p cos ωt already considered above [see (6)], is a if the Hamiltonian is given by

2 2 p mω 2 H = + q , (9) 2m 2 where ω is a constant. According to the results of the preceding section, if λ is a constant, mω S = λq sec ωt − q2 tan ωt + φ(t) (10) 2

3 must be a solution of the HJ equation for the Hamiltonian (9), if the function φ is appropriately chosen. A direct computation yields

2 2 2 1 ∂S mω ∂S λ ′ + q2 + = sec2 ωt + φ (t), 2m ∂q 2 ∂t 2m   and, therefore, choosing φ(t)= −λ2 tan ωt/2mω, we obtain the complete solution of the HJ equation

λ2 mω2 tan ωt S = λq sec ωt − + q2 . 2m 2 ω   Note that, in this case, H is also a constant of motion and, as pointed out above, the equation that determines the eigenfunctions of H is just the time-independent HJ equation. However, the constant of motion (6) leads to simpler expressions.

3.2 Particle in a time-dependent field As a second example we consider the time-dependent Hamiltonian p2 H = − ktq, 2m where k is a constant. One can readily verify that kt2 F = p − 2 is a constant of motion and that the eigenfunctions of F , i.e., the solutions of ∂S kt2 − = λ, ∂q 2 are given by kt2 S = λq + q + φ(t), (11) 2 where φ(t) is an arbitrary function of t only. Then

2 2 2 2 2 4 1 ∂S kt ∂S λ λkt k t ′ − + = + + + φ (t), 2m ∂q 2 ∂t 2m 2m 8m   hence, choosing λ2t λkt3 k2t5 φ(t)= − − − , 2m 6m 40m (11) is a complete solution of the HJ equation (which is not separable).

3.3 Particle in two dimensions As a final example we consider the Hamiltonian

2 2 1 eB eB H = p + y + p − x , (12) 2m x 2c y 2c "    # which corresponds to a charged particle of m and electric charge e in a uniform magnetic field B. The functions 1 1 mω mω F1 = (1 + cos ωt)p − p sin ωt + x sin ωt − (1 − cos ωt)y, (13) 2 x 2 y 4 4 1 1 mω mω F2 = (1 + cos ωt)p + p sin ωt + y sin ωt + (1 − cos ωt)x, (14) 2 y 2 x 4 4 where ω ≡ eB/mc, are constants of motion in involution, which correspond to the values of the canonical momenta px and py, respectively, at t = 0.

4 From (13) and (14) one finds that the common eigenfunctions of F1 and F2, with eigenvalues λ1 and λ2, respectively, are

1 mω 2 2 S = λ1x + λ2y + tan ωt λ2x − λ1y − (x + y ) + φ(t), 2 4 h i where φ(t) is an arbitrary function of t only. Substituting this expression into the HJ equation one finds that S is a solution of this equation if and only if

2 2 λ1 + λ2 ′ sec2 1 ωt + φ (t)=0, 2m 2 hence, 2 2 1 mω mω tan 2 ωt S = λ1x + λ2y − λ1 + y + λ2 − x 2 2 mω      is a complete solution of the HJ equation.

4 Concluding remarks

As pointed out above, the formulation of the Liouville theorem given here makes use of terms anal- ogous to those employed in the standard formalism of quantum mechanics, thus providing another example of the parallelism between both theories. Another advantage of the version of the Liouville theorem given above is that its proof is shorter than those usually presented in the textbooks.

References

[1] Whittaker, E.T.: A Treatise on the of Particles and Rigid Bodies, 4th ed. Cambridge University Press, Cambridge (1993). [2] Vilasi, G.: Hamiltonian Dynamics. World Scientific, Singapore (2001). [3] Babelon, O., Bernard, D., Talon, M.: Introduction to Classical Integrable Systems. Cambridge University Press, Cambridge (2003). [4] Fasano, A., Marmi, S.: . Oxford University Press, Oxford (2006). [5] DiBenedetto, E.: . Birkh¨auser, New York (2011). [6] Torres del Castillo, G.F.: Applications and extensions of the Liouville theorem on constants of motion, Rev. Mex. F´ıs. 57, 245-249 (2011). [7] Sneddon, I.N.: Elements of Partial Differential Equations. Dover, New York (2006). [8] Torres del Castillo, G.F., Cruz Dom´ınguez, H.H., de Yta Hern´andez, A., Herrera Flores, J.E., Sierra Mart´ınez, A.: Mapping of solutions of the Hamilton–Jacobi equation by an arbitrary , Rev. Mex. F´ıs. 60, 301-304 (2014).

5