nonautonomous hamiltonian systems and morales–ramis...

19
Copyright © by SIAM. Unauthorized reproduction of this article is prohibited. SIAM J. APPLIED DYNAMICAL SYSTEMS c 2009 Society for Industrial and Applied Mathematics Vol. 8, No. 1, pp. 279–297 Nonautonomous Hamiltonian Systems and Morales–Ramis Theory I. The Case ¨ x = f (x, t) Primitivo B. Acosta-Hum´ anez Abstract. In this paper we present an approach toward the comprehensive analysis of the nonintegrability of differential equations in the form ¨ x = f (x, t) which is analogous to Hamiltonian systems with 1+1/2 degrees of freedom. In particular, we analyze the nonintegrability of some important families of differential equations such as Painlev´ e II, Sitnikov, and the Hill–Schr¨odinger equation. Weemphasize Painlev´ e II, showing its nonintegrability through three different Hamiltonian systems, and also Sitnikov, in which two different versions including numerical results are shown. The main tool for studying the nonintegrability of these kinds of Hamiltonian systems is Morales–Ramis theory. This paper is a very slight improvement to the talk with a similar title delivered by the author at the SIAM Conference on Applications of Dynamical Systems in 2007. Key words. Hill–Schr¨odinger equation, Morales–Ramis theory, nonautonomous Hamiltonian systems, noninte- grability of Hamiltonian systems, Painlev´ e II equation, Sitnikov problem, virtually abelian groups AMS subject classifications. 37J30, 12H05, 70H07 DOI. 10.1137/080730329 1. Introduction. In this section we present the necessary theoretical background for un- derstanding the rest of the paper. 1.1. Differential Galois theory. Our theoretical framework consists of a well-established combination of dynamical systems theory, algebraic geometry, and differential algebra. See [22] or [36] for further information and details. Given a linear differential system with coefficients in C(t), (1.1) ˙ z = A (t) z, a differential field L C(t) exists, unique up to C(t)-isomorphism, which contains all entries of a fundamental matrix Ψ = [ψ 1 ,..., ψ n ] of (1.1). Moreover, the group of differential auto- morphisms of this field extension, called the differential Galois group of (1.1), is an algebraic group G acting over the C-vector space ψ 1 ,..., ψ n of solutions of (1.1) and containing the monodromy group of (1.1). It is worth recalling that the integrability of a linear system (1.1) is equivalent to the solvability of the identity component G 0 of the differential Galois group G of (1.1)—in other words, equivalent to the virtual solvability of G. Received by the editors July 14, 2008; accepted for publication (in revised form) by J. Meiss November 1, 2008; published electronically February 13, 2009. This research was partially supported by grant FPI Spanish Government, project BFM2003-09504-C02-02. http://www.siam.org/journals/siads/8-1/73032.html Departament de Matem` atica Aplicada I, Universitat Polit` ecnica de Catalunya, Diagonal 647, E-08028 Barcelona, Spain ([email protected]). 279

Upload: others

Post on 12-Jul-2020

5 views

Category:

Documents


0 download

TRANSCRIPT

  • Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.

    SIAM J. APPLIED DYNAMICAL SYSTEMS c© 2009 Society for Industrial and Applied MathematicsVol. 8, No. 1, pp. 279–297

    Nonautonomous Hamiltonian Systems and Morales–Ramis Theory I. The Caseẍ = f(x, t)∗

    Primitivo B. Acosta-Humánez†

    Abstract. In this paper we present an approach toward the comprehensive analysis of the nonintegrability ofdifferential equations in the form ẍ = f(x, t) which is analogous to Hamiltonian systems with 1+1/2degrees of freedom. In particular, we analyze the nonintegrability of some important families ofdifferential equations such as Painlevé II, Sitnikov, and the Hill–Schrödinger equation. We emphasizePainlevé II, showing its nonintegrability through three different Hamiltonian systems, and alsoSitnikov, in which two different versions including numerical results are shown. The main tool forstudying the nonintegrability of these kinds of Hamiltonian systems is Morales–Ramis theory. Thispaper is a very slight improvement to the talk with a similar title delivered by the author at theSIAM Conference on Applications of Dynamical Systems in 2007.

    Key words. Hill–Schrödinger equation, Morales–Ramis theory, nonautonomous Hamiltonian systems, noninte-grability of Hamiltonian systems, Painlevé II equation, Sitnikov problem, virtually abelian groups

    AMS subject classifications. 37J30, 12H05, 70H07

    DOI. 10.1137/080730329

    1. Introduction. In this section we present the necessary theoretical background for un-derstanding the rest of the paper.

    1.1. Differential Galois theory. Our theoretical framework consists of a well-establishedcombination of dynamical systems theory, algebraic geometry, and differential algebra. See [22]or [36] for further information and details. Given a linear differential system with coefficientsin C(t),

    (1.1) ż = A (t)z,

    a differential field L ⊃ C(t) exists, unique up to C(t)-isomorphism, which contains all entriesof a fundamental matrix Ψ = [ψ1, . . . ,ψn] of (1.1). Moreover, the group of differential auto-morphisms of this field extension, called the differential Galois group of (1.1), is an algebraicgroup G acting over the C-vector space 〈ψ1, . . . ,ψn〉 of solutions of (1.1) and containing themonodromy group of (1.1).

    It is worth recalling that the integrability of a linear system (1.1) is equivalent to thesolvability of the identity component G0 of the differential Galois group G of (1.1)—in otherwords, equivalent to the virtual solvability of G.

    ∗Received by the editors July 14, 2008; accepted for publication (in revised form) by J. Meiss November 1, 2008;published electronically February 13, 2009. This research was partially supported by grant FPI Spanish Government,project BFM2003-09504-C02-02.

    http://www.siam.org/journals/siads/8-1/73032.html†Departament de Matemàtica Aplicada I, Universitat Politècnica de Catalunya, Diagonal 647, E-08028 Barcelona,

    Spain ([email protected]).

    279

    http://www.siam.org/journals/siads/8-1/73032.htmlmailto:[email protected]

  • Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.

    280 PRIMITIVO B. ACOSTA-HUMÁNEZ

    It is well established (e.g., [23]) that any linear differential equation system with coefficientsin a differential field K,

    (1.2)d

    dt

    (ξ1ξ2

    )=(

    a(t) b(t)c(t) d(t)

    )(ξ1ξ2

    ),

    by means of an elimination process, is equivalent to the second-order equation

    (1.3) ξ̈ −(

    a(t) + d(t) +ḃ(t)b(t)

    )ξ̇ −

    (ȧ(t) + b(t)c(t) − a(t)d(t) − a(t)ḃ(t)

    b(t)

    )ξ = 0,

    where ξ := ξ1. Furthermore, any equation of the form z̈ − 2pż − qz = 0 can be transformed,through the change of variables z = ye

    ∫p, into ÿ = −ry, r, satisfying the Riccati equation

    ṗ = r+q +p2. This change is useful since it restricts the study of the Galois group of ÿ = −ryto that of the algebraic subgroups of SL2 (C).

    A natural question which now arises is to determine what happens if the coefficients of thedifferential equation are not all rational. A new method was developed in [3] to transform alinear differential equation of the form ẍ = r(t)x, with transcendental or algebraic nonrationalcoefficients, into its algebraic form—that is, into a differential equation with rational coeffi-cients. This is called the algebrization method and is based on the concept of a Hamiltonianchange of variables [3]. Such a change is derived from the solution of a one-degree-of-freedomclassical Hamiltonian.

    Definition 1.1 (Hamiltonian change of variables). A change of variables τ = τ (t) is calledHamiltonian if (τ(t), τ̇ (t)) is a solution curve of the autonomous Hamiltonian system XHwith Hamiltonian function

    H = H(τ, p) =p2

    2+ V̂ (τ) for some V̂ ∈ C(τ).

    Theorem 1.1 (Acosta–Blázquez algebrization method [3]). Equation ẍ = r(t)x is algebriz-able by means of a Hamiltonian change of variables τ = τ(t) if and only if there exist f, αsuch that ddτ (lnα),

    fα ∈ C(τ), where

    f(τ(t)) = r(t), α(τ) = 2(H − V̂ (τ)) = (τ̇ )2.

    Furthermore, the algebraic form of ẍ = r(t)x is

    (1.4)d2x

    dτ2+(

    12

    d

    dτln α

    )dx

    dτ−(

    f

    α

    )x = 0.

    The next intended step, once a differential equation has been algebrized, is studying itsGalois group and, as a causal consequence, its integrability. Concerning the latter, and byvirtue of the invariance of the identity component of the Galois group by finite branchedcoverings of the independent variable (Morales-Ruiz and Ramis, [25, Theorem 5]), it wasproven in [3, Proposition 1] that the identity component of the Galois group is preserved inthe algebrization mechanism.

  • Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.

    NONAUTONOMOUS HAMILTONIANS AND MORALES–RAMIS I 281

    The final step is analyzing the behavior of t = ∞ (or τ = ∞) by studying the behaviorof η = 0 through the change of variables η = 1/t (or η = 1/τ) in the transformed differentialequation; i.e., t = ∞ (or τ = ∞) is an ordinary point (resp., a regular singular point, anirregular singular point) of the original differential equation if and only if η = 0 is one suchpoint for the transformed differential equation.

    1.2. Morales–Ramis theory. Everything is considered in the complex analytical settingfrom now on. The heuristics of the titular theory rest on the following general principle: ifwe assume system

    (1.5) ż = X (z)

    “integrable” in some reasonable sense, then the corresponding variational equations along anyintegral curve Γ = {ẑ (t) : t ∈ I} of (1.5), defined in the usual manner

    ξ̇ = X ′ (ẑ (t)) ξ,(VEΓ)

    must be also integrable—in the Galoisian sense of the last paragraph in section 1.1. Weassume Γ, a Riemann surface, may be locally parametrized in a disc I of the complex plane;we may now complete Γ to a new Riemann surface Γ, as detailed in [25, sect. 2.1] (see also [22,sect. 2.3]), by adding equilibrium points, singularities of the vector field, and possible pointsat infinity.

    The aforementioned “reasonable” sense in which to define integrability if system (1.5) isHamiltonian is obviously the one given by the Liouville–Arnold theorem, and thus the abovegeneral principle does have an implementation.

    Theorem 1.2 (Morales-Ruiz and Ramis [22, 25]). Let H be an n-degree-of-freedom Hamil-tonian having n independent rational or meromorphic first integrals in pairwise involution,defined on a neighborhood of an integral curve Γ. Then, the identity component Gal

    (VEΓ

    )0is an abelian group (i.e., Gal

    (VEΓ

    )is virtually abelian).

    The disjunctive between meromorphic and rational Hamiltonian integrability in Theo-rem 1.2 is related to the status of t = ∞ as a singularity for the normal variational equations.More specifically, and besides the nonabelian character of the identity component of the Ga-lois group, in order to obtain Galoisian obstructions to the meromorphic integrability of Hthe point at infinity must be a regular singular point of (VEΓ). On the other hand, for thereto be an obstruction to complete sets of rational first integrals, t = ∞ must be an irregularsingular point.

    See [25, Corollary 8] or [22, Theorem 4.1] for a precise statement and a proof.

    1.3. Nonautonomous Hamiltonian systems. Nonautonomous Hamiltonian systems onsymplectic manifolds have long been the subject of study and appear in a most natural wayin classical mechanics and control theory, e.g., [1], [5], [19], [21], [20], [28], [33], [34].

    We consider nonautonomous Hamiltonian systems of the form

    (1.6) H = H(q1, p1, t) =p212

    + V (q1, p1, t).

    H is a nonautonomous Hamiltonian system with 1 + 1/2 degrees of freedom. It is well known(e.g., [28]) that (1.6) can be included as a subsystem of the Hamiltonian system with two

  • Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.

    282 PRIMITIVO B. ACOSTA-HUMÁNEZ

    degrees of freedom given by

    (1.7) Ĥ = Ĥ(q1, q2, p1, p2) =p212

    + V (q1, p1, q2) + p2,

    where q2 and p2 are conjugate variables, i.e., p2 = −H + k, where k is constant, and q2 = t.Furthermore, p2 is easily seen to be the offset or counterbalancing energy of the system [28],[30].

    Also worth mentioning are some recent results on canonical transformations in the ex-tended phase space [30], [31] [32], [35] as well as on definitions and consequences of “integra-bility” under such circumstances or generalizations thereof, even for non-Hamiltonian systems[11], [9], [14], [18] which we will not delve into further at this point.

    2. Main results. Consider the differential equation

    (2.1) ẍ = f(x, t),

    with particular solution x = x(t). We will henceforth order our choice of positions as q1 = xand q2 = t, thus yielding a Hamiltonian system given by

    H =p212

    − F (q1, q2) , Fq1 (q1, q2) =∂F (q1, q2)

    ∂q1= f(q1, q2).

    Equation (2.1) is obviously equivalent to Hamilton’s equations for H,

    q̇1 = p1 = Hp1, ṗ1 = −Hq1 = f(q1, q2);

    this nonautonomous Hamiltonian system is included as a subsystem of XĤ linked to Ĥ :=H + p2, such as in (1.7). Assuming x(t) = q1(t) to be a solution of (2.1) and q2(t) = t, weobtain an integral curve Γ = {z (t)} of Ĥ, where

    z (t) := (q1(t), q2(t), p1(t), p2(t)) = (q1(t), t, q̇1(t),−H(t)) .

    We may now introduce our first main result.Theorem 2.1. Let Γ be an integral curve of X

    Ĥsuch as the one introduced above. If X

    Ĥis integrable by means of rational or meromorphic first integrals, then the Galois group of

    (2.2) ξ̈ = (fq1(q1, q2)|Γ) ξ

    is virtually abelian.Proof. The Hamiltonian field XĤ is given by XĤ = (p1, 1, f(q1, q2), Fq2(q1, q2))

    T . Thevariational equation VEΓ along Γ is

    (2.3)d

    dt

    ⎛⎜⎜⎝ξ1ξ2ξ3ξ4

    ⎞⎟⎟⎠ =⎛⎜⎜⎝

    0 0 1 00 0 0 0

    fq1(q1, q2) fq2(q1, q2) 0 0Fq1q2(q1, q2) Fq2q2(q1, q2) 0 0

    ⎞⎟⎟⎠∣∣∣∣∣∣∣∣Γ

    ⎛⎜⎜⎝ξ1ξ2ξ3ξ4

    ⎞⎟⎟⎠ .

  • Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.

    NONAUTONOMOUS HAMILTONIANS AND MORALES–RAMIS I 283

    More precisely, (2.3) is

    (2.4)

    ⎧⎪⎪⎪⎪⎨⎪⎪⎪⎪⎩ξ̇1 = ξ3,

    ξ̇2 = 0,

    ξ̇3 = (fq1(q1, q2)|Γ) ξ1 + (fq2(q1, q2)|Γ) ξ2,ξ̇4 = (Fq1q2(q1, q2)|Γ) ξ1 + (Fq2q2(q1, q2)|Γ) ξ2.

    Hence ξ2 = k, where k is constant. Assuming k = 0, the normal variational equations((NVEΓ); see [8, sect. 1], [25, sect. 4.3], [22, sect. 4.1.3]) for Ĥ are given by

    (2.5)d

    dt

    (ξ1ξ3

    )=(

    0 1fq1(q1, q2)|Γ 0

    )(ξ1ξ3

    ),

    and solving (2.5) we can obtain ξ4. System (2.5) is equivalent to (2.2), where ξ = ξ1. By virtueof [22, Proposition 4.2], the virtual commutativity of Gal (VEΓ) implies that of Gal (NVEΓ);this, coupled with Theorem 1.2, implies that if XĤ is rationally or meromorphically integrable,then the Galois group of (2.2) is virtually abelian.

    Corollary 2.2. Suppose the Galois group of the differential equation ξ̈ = k(t)ξ is not virtuallyabelian. Define H := p

    212 − k(q2)

    q212 and Ĥ := H + p2. Then XĤ is not integrable by means of

    meromorphic or rational first integrals.Remark 1. If the Galois group of the equation ξ̈ = k(t)ξ is the Borel group G = C∗ � C

    (hence connected, solvable, and nonabelian), then XĤ

    is neither meromorphically nor ratio-nally integrable, although it is still possible to solve the equation as well as, ostensibly, thenonautonomous Hamiltonian system XH .

    In anticipation of the following corollary, consider, for any g(x), a(t), and α = α0 given,the equation

    (2.6) ẍ = gx(x)(g(x) + a(t)) + α, α ∈ C,

    having a certain known particular solution q1 = q1 (t). Let

    H =p212

    − (g(q1) + a(q2))2

    2− αq1, q2 = t,

    be a Hamiltonian linked to (2.6) and Ĥ := H + p2 its autonomous completion.Corollary 2.3. If X

    Ĥis integrable through rational or meromorphic first integrals, then,

    along the integral curve Γ = {z(t) = (q1(t), t, q̇1(t),−H(t))}, the Galois group of the equation

    (2.7) ξ̈ =(g2q1(q1) + gq1q1(q1)(g(q1) + a(q2))

    )∣∣Γ

    ξ

    is virtually abelian.Now, keeping the above hypotheses for g(x), a(t), α = α0, (2.6), and x = q1(t), let us

    define the Hamiltonian system

    (2.8) Ĥ = H + p2, H =p212

    − (g(q1) + a(q2))p1 − (α + aq2(q2))q1, α ∈ C.

  • Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.

    284 PRIMITIVO B. ACOSTA-HUMÁNEZ

    It takes only the following simple calculation to prove that Hamiltonian H is indeed linked to(2.6) in the manner expected, i.e., as introduced in the beginning of section 2. We have

    ẍ = gx(x)(g(x) + a(t)) + α= gx(x)ẋ + gx(x)g(x) + gx(x)a(t) + α + ȧ(t) − gx(x)ẋ − ȧ(t)= gx(x)y + α + ȧ(t) − ddt(g(x) + a(t)),

    where y = ẋ+g(x)+a(t), and requiring x and y to be conjugate variables implies the followingsystem, equivalent to (2.6):

    ẋ = y − (g(x) + a(t)) = Hy,ẏ = gx(x)y + α + ȧ(t) = −Hx.

    A straightforward parallel integration,

    H =y2

    2− (g(x) + a(t)) y + h1(x, t) = h2(y, t) − g(x)y − (α + ȧ(t))x,

    along with the definition of

    h1(x, t) = −(α + ȧ(t))x, h2(y, t) = y2

    2− a(t)y,

    as well as (q1, q2, p1) = (x, t, y), yields the autonomous Hamiltonian system introduced in(2.8).

    Theorem 2.4. If XĤ

    is integrable by means of rational or meromorphic first integrals, then,along Γ = {z (t) = (q1(t), t, p1(t),−H(t))}, the Galois group of the equation

    (2.9) ξ̈ =(g2q1(q1) + p1gq1q1(q1) − gq2q1(q1)

    )∣∣Γ

    ξ

    is virtually abelian.Proof. We may proceed as in the proof of Theorem 2.1. The Hamiltonian field X

    Ĥis

    given by

    XĤ =

    ⎛⎜⎜⎝p1 − (g(q1) + a(q2))

    1gq1(q1)p1 + (α + aq2(q2))aq2(q2)p1 + aq2q2(q2)q1

    ⎞⎟⎟⎠ .The variational equation VEΓ along Γ = {(q1(t), t, p1(t),−H(t))} is

    (2.10) ξ̇ =

    ⎛⎜⎜⎜⎜⎝−∂g(q1)∂q1 −

    ∂a(q2)∂q2

    1 00 0 0 0

    p1∂2g(q1)

    ∂q21

    ∂2a(q2)∂q22

    ∂g(q1)∂q1

    0∂2a(q2)

    ∂q22

    (p1

    ∂2a(q2)∂q22

    + q1∂3a(q2)

    ∂q32

    )∂a(q2)

    ∂q20

    ⎞⎟⎟⎟⎟⎠∣∣∣∣∣∣∣∣∣∣Γ

    ξ,

  • Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.

    NONAUTONOMOUS HAMILTONIANS AND MORALES–RAMIS I 285

    where ξ = (ξ1, ξ2, ξ3, ξ4)T . ξ2 ≡ k ∈ C; in particular, k = 0 renders system (2.10) equal to thefollowing:

    (2.11)

    ⎧⎪⎪⎪⎪⎨⎪⎪⎪⎪⎩ξ̇1 = −gq1(q1)|Γ ξ1 + ξ3,ξ̇2 = 0,

    ξ̇3 = p1gq1q1(q1)|Γ ξ1 + −gq1(q1)|Γ ξ3,ξ̇4 = aq2q2(q2)|Γ ξ1 + aq2(q2)|Γ ξ3,

    NVEΓ corresponding to

    (2.12)d

    dt

    (ξ1ξ3

    )=( −gq1(q1) 1

    p1gq1q1(q1) gq1(q1)

    )∣∣∣∣Γ

    (ξ1ξ3

    ).

    Solving (2.12), we obtain ξ4. Using (1.2) and (1.3) system (2.12) is equivalent to (2.9),where ξ = ξ1. Again, by virtue of [22, Proposition 4.2] as in Theorem 2.1, the integrabilityof Hamiltonian Ĥ by means of meromorphic or rational first integrals implies the virtualcommutativity of the Galois group of (2.9).

    Remark 2. We observe that if ẍ = f(x, t), and if the same particular solution x = x(t)occurs in the statements of Theorem 2.1, Corollary 2.3, and Theorem 2.4, then we obtainthe same NVEΓ ( (2.2), (2.7), and (2.9) are equivalent), despite the fact that their respectivelinked Hamiltonian systems have different expressions.

    3. Examples. In this section, and in the application of Theorems 2.1 and 2.4 as wellas of Corollaries 2.2 and 2.3, we analyze the nonintegrability of the Hamiltonian systemscorresponding to the following differential equations:

    1. Hill–Schrödinger equation: ẍ = k(t)x.2. Sitnikov problem: ẍ = − (1−e cos t)x

    (x2+r2(t))32, r(t) = 1−e cos t2 .

    3. Painlevé II equation: ẍ = 2x3 + tx + α.4. An algebraic toy model: ẍ = − 1

    4x3− t

    x2+ α.

    In order to analyze normal variational equations, a standard procedure is using Maple, andespecially the commands dsolve and kovacicsols. Whenever the command kovacicsolsyields an output “[ ],” it means that the second-order linear differential equation beingconsidered has no Liouvillian solutions, and thus its Galois group is virtually nonsolvable.For equations of the form ÿ = ry with r ∈ C(x), the only virtually nonsolvable group isSL2 (C). In some cases, moreover, dsolve makes it possible to obtain the solutions in termsof special functions such as Airy functions, Bessel functions, and hypergeometric functions,among others [2]. There is a number of second-order linear equations whose coefficients arenot rational and whose solutions Maple cannot find by means of the commands dsolve andkovacicsols alone; this problem, in some cases, can be solved by the stated algebrizationprocedure.

    3.1. Hill–Schrödinger equation ẍ = k(t)x. This example corresponds to Corollary 2.2for k > 0, � � 0, Pn a polynomial of degree n with Pn(0) = 1, and k(t) given by

    k(t) = ke−�t, k(t) = kPn(�t),k(t) = k(1 + sinh(�t)), k(t) = k(1 + sin(�t)),k(t) = k(1 + cosh(�t)), k(t) = k(1 + cos(�t)).

  • Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.

    286 PRIMITIVO B. ACOSTA-HUMÁNEZ

    XĤ is nonintegrable by means of rational first integrals.The integrability of equation ẍ = k(t)x for these examples has been deeply analyzed in [3].

    3.2. The Sitnikov problem. The Sitnikov problem is a symmetrically configured restrictedthree-body problem in which two primaries with equal masses move in ellipses of eccentricitye in a plane π1, and an infinitesimal point mass moves along the line π⊥1 . See [6], [12], [25],[27], [37], [38], [16] for more details. The motion of the infinitesimal point mass is given bythe differential equation

    (3.1) z̈ +z

    (r2 (t) + z2)3/2= 0,

    where z = z (t) is the distance from the infinitesimal mass point to the plane of the primariesand r (t) is half the distance of the primaries,

    r (t) =1 − e cos E (t)

    2,

    where the eccentric anomaly E (t) is the solution of the Kepler equation

    (3.2) E = t + e sin E

    and e is the eccentricity of the ellipses described by the primaries. We will assume 0 ≤ e ≤ 1through subsections 3.2 and 3.3. The Hamiltonian linked to the system is

    (3.3) H =v2

    2− 1

    (z2 (t) + r2 (t))1/2,

    provided v stands for ż in the corresponding equations.Since (3.3) cannot be solved in explicit form, attempts at a Hamiltonian formulation of

    (3.1), whether exact or approximate, require one of at least two options: looking for an exactHamiltonian formulation by means of a change of variables, which we will do in the nextparagraph, and searching an approximate Hamiltonian formulation, which will be done insubsection 3.3.

    Let us now find a Hamiltonian linked to (3.1). We may express r (t) as

    r (t) = (R ◦ ϕ) (t) := 1 − e2

    2 (1 + e cos ϕ (t)),

    where ϕ, the true anomaly, is a solution of

    dt=

    (1 + e cos ϕ)2

    (1 − e2)3/2=

    √1 − e2

    4R2 (ϕ),

    and we may follow the procedure introduced in [38] (see also [16]) by taking ϕ as the newindependent variable and x = z2r(ϕ) as the new dependent variable. Writing t once again todenote ϕ, we have the following differential equation:

    (3.4) ẍ = f (x, t) := −e cos t +(

    14 + x

    2)−3/2

    1 + e cos tx,

  • Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.

    NONAUTONOMOUS HAMILTONIANS AND MORALES–RAMIS I 287

    clearly amenable to the hypotheses in Theorem 2.1 and in the first paragraph of section 2.Defining q1 = x, q2 = t, and p1 = ẋ, the autonomous Hamiltonian system corresponding to(3.4) is given by

    (3.5) Ĥe = He + p2 :=p212

    +eq21 cos q2 − 4

    (1 + 4q21

    )−1/22 (1 + e cos q2)

    + p2,

    always assuming e ∈ [0, 1].The circular Sitnikov problem Ĥ0 is meromorphically integrable in the sense of Liouville–

    Arnold and can be solved using elliptic integrals. The nonintegrability for e = 1 was firststudied by means of straight Morales–Ramis theory in [26, sect. 5] (see also [22, sect. 5.3]); wewill now extend the proof of meromorphic nonintegrability therein to one for every 0 < e ≤ 1by using Theorem 2.1.

    The NVEΓ derived from the Hamiltonian system given by (3.5) is

    ξ̈ =

    (e(4q21 + 1

    )5/2 cos q2 − 64q21 + 8(e cos q2 + 1)

    (4q21 + 1

    )5/2)

    ξ.

    Taking q1 ≡ 0 we have a solution Γ ={z (t) =

    (0, t, 0, 21+e cos t

    )}of X

    Ĥealong which the

    NVEΓ is given by

    ξ̈ =(

    e cos t + 8e cos t + 1

    )ξ,

    which is algebrizable, through the change of variable τ = cos t, into

    (3.6)d2ξ

    dτ2−(

    τ

    1 − τ2)

    dτ−(

    eτ + 8(eτ + 1)(1 − τ2)

    )ξ = 0.

    This equation can be transformed into the differential equation

    (3.7)d2ζ

    dτ2=(

    5eτ3 + 33τ2 − 2eτ − 30(eτ + 1)4(1 − τ2)2

    by means of ξ = ζ4√1−τ2 . Equations (3.6) and (3.7) are integrable in terms of Liouvilliansolutions only if e = 0, whereas for e �= 0 their solutions are given in terms of nonintegrableHeun functions if e �= 1 and nonintegrable hypergeometric functions if e = 1, i.e., their Galoisgroups are virtually nonsolvable and hence virtually nonabelian; furthermore, they have aregular singularity at infinity, which by Theorem 1.2 implies the nonintegrability of XĤ fore ∈ (0, 1] by means of meromorphic first integrals. In particular, the Galois group of (3.7) isexactly SL2 (C).

    3.2.1. Numerical results for the Sitnikov problem. The author is indebted to Sergi Simonin what concerns the following subsection, including the figures. Acknowledgments are alsodue to Carles Simó for further specific suggestions.

    Let Σ = {sin (q2) = 0}. The six figures in this paper show Poincaré sections of the flowwith respect to Σ, projected on the (q1, p1) plane, for the Hamiltonian system XHe obtained

  • Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.

    288 PRIMITIVO B. ACOSTA-HUMÁNEZ

    -2

    -1.5

    -1

    -0.5

    0

    0.5

    1

    1.5

    2

    -4 -3 -2 -1 0 1 2 3 4

    Figure 1. Poincaré section sin q2 = 0 for the Sitnikov problem: e = 0 (figure: Sergi Simon).

    from (3.5). As may be easily deduced from said Hamiltonian, all sections are symmetricalwith respect to the q1 and p1 axes. Different numbers of initial conditions are used for thesake of clarity.

    Figure 1 corresponds to e = 0. In keeping with what was said after (3.5), the wholesubset of Σ transversal to the flow sheds concentric tori (ostensibly, the intersections of theinvariant Liouville–Arnold tori with Σ), a typical sign of integrability; the tori shown are onlya selection of those therein, as the actual area foliated by them is larger.

    A number of these invariant tori break down upon the slightest increase in e, and in theensuing figures the two most interesting features are those invariant sets (usually called KAMtori) whose intersection with Σ prevails in the form of Jordan curves, and the zones of chaoticbehavior between them. Sparse zones of the section will account for chaotic zones as well fore > 0. Figures 2 and 3 show two different close-up views for the Poincaré section correspondingto e = 0.01. The latter figure is actually a detail of the “island” of tori appearing at the rightof the general section. For e = 0.1, Figures 4 and 5 are, respectively, a general view of thesection and a close-up of one of the islands appearing at each side of the central area. As fore = 0.4, Figure 6 is an enlarged view of one of the two islands appearing at each side of acentral area.

    3.3. The approximate Sitnikov problem. As said in subsection 3.2, we now consideran approximation of the Sitnikov problem; see [15] and [12] for more details. As opposedto meromorphic nonintegrability, we will prove nonintegrability by means of rational firstintegrals.

  • Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.

    NONAUTONOMOUS HAMILTONIANS AND MORALES–RAMIS I 289

    -1

    -0.5

    0

    0.5

    1

    1 1.5 2 2.5 3 3.5 4 4.5 5

    Figure 2. Poincaré section sin q2 = 0 for the Sitnikov problem, e = 0.01 (figure: Sergi Simon).

    The fact that ϕ (t) = t + O (e) yields r (t) = 1−e cos t2 + O(e2), and thus the Hamiltonian

    in (3.3) becomes

    H =v2

    2− 1√

    z2 + 14− e cos t(

    z2 + 14)3/2 + O (e2)

    whenever e ≈ 0. In particular, the first-order approximation of this asymptotic expansion ine yields the Hamiltonian

    (3.8) H =v2

    2− 1√

    z2 + 14− e cos t(

    z2 + 14)3/2 .

    Considering q1 = x, q2 = t, and p1 = v, the autonomous Hamiltonian system correspond-ing to this equation is given by

    (3.9) Ĥ = H + p2, H =p212

    − e 2 cos q2(4q21 + 1)3/2

    − 2√4q21 + 1

    ,

    corresponding to the Hamiltonian in Theorem 2.1:

    f(x, t) = − 8x(4x2 + 1)

    32

    − e 24x cos t(4x2 + 1)

    52

    .

  • Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.

    290 PRIMITIVO B. ACOSTA-HUMÁNEZ

    -0.1

    -0.05

    0

    0.05

    0.1

    3.5 3.6 3.7 3.8 3.9 4 4.1 4.2 4.3 4.4

    Figure 3. Poincaré section for e = 0.01 (figure: Sergi Simon).

    The NVEΓ for the Hamiltonian (3.9) is given by

    ξ̈ =

    (e24(16q21 − 1) cos q2

    (4q21 + 1)72

    +8(8q21 − 1)(4q21 + 1)

    52

    )ξ.

    Its general solution may be expressed as

    ξ (t) = K1C(

    32, 48e,t

    2

    )+ K2S

    (32, 48e,

    t

    2

    ),

    where the Mathieu even (resp., odd) function C (a, q, t) (resp., S (a, q, t)) is defined as the even(resp., odd) solution to ÿ + (a − 2q cos (2t))y = 0 [2, Ch. 20].

    Taking q1(t) = 0, we can see that z(t) = (0, t, 0, 2e cos t + 2); hence, defining z(t) =(0, t, 0, 2e cos t + 2) and Γ = {z(t)}, the operator linked to NVEΓ is given by

    ξ̈ = (−24e cos t − 8) ξ,which is algebrizable (see Theorem 1.1) through the change of variables τ = cos t into

    (3.10)d2ξ

    dτ2−(

    τ

    1 − τ2)

    dτ+(

    24eτ + 81 − τ2

    )ξ = 0.

    Now, this equation can be transformed into the differential equation

    (3.11)d2ζ

    dτ2=(

    96eτ3 + 31τ2 − 96eτ − 344(1 − τ2)2

    )ζ, ξ =

    ζ4√

    1 − τ2 .

  • Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.

    NONAUTONOMOUS HAMILTONIANS AND MORALES–RAMIS I 291

    -2

    -1.5

    -1

    -0.5

    0

    0.5

    1

    1.5

    2

    -3 -2 -1 0 1 2 3

    Figure 4. Poincaré section sin q2 = 0 for the Sitnikov problem, e = 0.1 (figure: Sergi Simon).

    Equations (3.10) and (3.11) are integrable in terms of Liouvillian solutions only if e = 0,since for e �= 0 their solutions are given in terms of nonintegrable Mathieu functions; hencetheir Galois groups are virtually nonsolvable and thus virtually nonabelian; furthermore, theyhave an irregular singularity at infinity, implying rational nonintegrability for the Hamiltonianfield XĤ with α = 1 by virtue of Theorem 1.2. In particular, the Galois group of (3.11) isexactly SL2 (C).

    Equations (3.10) and (3.11) have been deeply analyzed in [3] using the Hamiltonian changeof variables τ = eit, obtaining the same result presented here.

    3.4. Painlevé II equation: ẍ = 2x3 + tx + α. The nonintegrability of the secondPainlevé equation for integer α was proved by Morales-Ruiz in [23] and later by Stoyanovaand Christov in [29]. They used only the Hamiltonian (3.14).

    Defining q1 = x, q2 = t, p1 = y, and α ∈ C, the autonomous Hamiltonian systemcorresponding to this equation can be given by any of the following three functions:

    Ĥ = H + p2, H =p212

    − q41

    2− q2 q

    21

    2− αq1,(3.12)

    Ĥ = H + p2, H =p212

    − 12

    (q21 +

    q22

    )2 − αq1,(3.13)Ĥ = H + p2, H =

    p212

    −(q21 +

    q22

    )p1 −

    (α +

    12

    )q1,(3.14)

    where (3.12), (3.13), and (3.14) correspond to the Hamiltonian of Theorem 2.1 (f(x, t) =

  • Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.

    292 PRIMITIVO B. ACOSTA-HUMÁNEZ

    -0.4

    -0.3

    -0.2

    -0.1

    0

    0.1

    0.2

    0.3

    0.4

    1.4 1.6 1.8 2 2.2 2.4 2.6

    Figure 5. Poincaré section for e = 0.1 (figure: Sergi Simon).

    2x3 + tx + α), Corollary 2.3 (g(x) = x2 and a(t) = t/2), and Theorem 2.4 (g(x) = x2

    and a(t) = t/2), respectively. The Hamiltonian system for Painlevé II, studied in [23, 29],corresponds precisely to Hamiltonian (3.14). The NVEΓ for these Hamiltonians is given by

    ξ̈ =(6q21 + q2

    )ξ.

    Taking α = 0 and q1(t) = 0, we have particular solutions z(t) = (0, t, 0, 0), z(t) =(0, t, 0, t2/8

    ), and z(t) =

    (0, t, t/2, t2/8

    ), respectively, for the Hamiltonians (3.12), (3.13), and

    (3.14); hence NVEΓ is given by ξ̈ = tξ, the so-called Airy equation [2, sect. 10.4.1], which hasan irregular singularity at infinity and is not integrable through Liouvillian solutions; i.e., itsGalois group is SL2 (C) and not virtually abelian; thus, by Theorem 1.2, the Hamiltonian fieldXĤ with α = 0 is not integrable through rational first integrals.

    Now, for α = 1 and q1(t) = −1/t, the integral curve z(t) is given by(−1

    t, t,

    1t2

    ,− 12t

    ),

    (−1

    t, t,

    1t2

    ,− 12t

    +t2

    8

    ), and

    (−1

    t, t,

    2t2

    +t

    2,−1

    t+

    t2

    8

    ),

    respectively, for the Hamiltonians (3.12), (3.13), and (3.14), so that NVEΓ is given by

    ξ̈ =(

    6t2

    + t)

    ξ, Γ = {z(t)},

  • Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.

    NONAUTONOMOUS HAMILTONIANS AND MORALES–RAMIS I 293

    -0.05

    0

    0.05

    2.5 3 3.5

    Figure 6. Poincaré section for the Sitnikov problem, e = 0.4 (figure: Sergi Simon).

    whose general solution is

    ξ (t) =√

    t

    [K1I−5/3

    (2t3/2

    3

    )+ K2I5/3

    (2t3/2

    3

    )],

    Iα = 2−αtα(

    1Γ(1+α) +

    t2

    22Γ(2+α)+ O(t4)

    )being, for each α, the modified Bessel function of the

    first kind, i.e., the solution to t2ÿ + tẏ − (t2 + α2) y = 0 [2, sect. 9.6].The normal variational equation has an irregular singularity at infinity and is not inte-

    grable through Liouvillian functions because its solutions are given in term of nonintegrableBessel functions (see [23, 29]); i.e., its Galois group is SL2 (C), which is not virtually abelian;again by the Morales–Ramis theorem, Theorem 1.2, the Hamiltonian field X

    Ĥwith α = 1 is

    not integrable through rational first integrals.

    3.5. An algebraic toy model: ẍ = − 14x3

    − tx2

    + α. Considering q1 = x, q2 = t, p1 = y,and α ∈ C; the autonomous Hamiltonian systems corresponding to this equation are given by

    Ĥ = H + p2, H =p212

    − 18q21

    − q2q1

    − αq1,(3.15)

    Ĥ = H + p2, H =p212

    − 12

    (1

    2q1+ 2q2

    )2− αq1,(3.16)

    Ĥ = H + p2, H =p212

    +(

    12q1

    + 2q2

    )p1 − (α + 2) q1.(3.17)

  • Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.

    294 PRIMITIVO B. ACOSTA-HUMÁNEZ

    Equations (3.15), (3.16), and (3.17) correspond to Theorem 2.1 (f(x, t) = − 14x3

    − tx2

    + α),Corollary 2.3 (g(x) = − 12x and a(t) = −2t), and Theorem 2.4 (g(x) = − 12x and a(t) = −2t),respectively. The NVEΓ for all three is given by

    ξ̈ =(

    34q41

    +2q2q31

    )ξ.

    Now, for α = 1 and q1(t) =√

    t, the integral curve z(t) is given by(√t, t,

    12√

    t, 2√

    t

    ),

    (√t, t,

    12√

    t, 2t2 + 2

    √t

    ), and

    (√t, t,−2t, 2t2

    ),

    respectively, for the Hamiltonians (3.15), (3.16), and (3.17), rendering NVEΓ equal to

    (3.18) ξ̈ =(

    34t2

    +2√t

    )ξ,

    having a solution

    ξ1 = −3t3/2

    2 0F1

    (;73;8t3/2

    9

    )

    = −32t3/2 − 4

    7t3 − 8

    105t9/2 + O

    (t6),

    0F1 (; a; t) = limq→∞ 1F1(q; a; tq

    )=∑∞

    n=0tn

    (a)nn!being the confluent hypergeometric limit func-

    tion [2, Ch. 13], and an independent new solution ξ2 = ξ1∫

    ξ−21 , satisfying

    ξ2 =1

    3√

    t− 8t

    9− 16

    27t5/2 + O

    (t4).

    As is the case for the rest of the normal variational operators appearing in this paper, ourknowledge of the exponents around 0 of a fundamental set of solutions (in this case, ξ1 andξ2), coupled with the basic result on factorization obtained in [7, Th. 8 (Ch. 5)] (see also[8, Criterion 1]), would suffice to prove nonintegrability. Here, however, we will restrict ourattention to Theorems 2.1 and 2.4 and Corollary 2.3.

    Equation (3.18) is algebrizable (Theorem 1.1), through the change of variables τ =√

    tinto

    (3.19)d2ξ

    dτ2−(

    )dξ

    dτ−(

    8τ3 + 3τ2

    )ξ = 0;

    now, this equation can be transformed into the differential equation

    (3.20)d2ζ

    dτ2=(

    32τ3 + 154τ2

    )ζ, ξ = ζ

    √τ .

    Equations (3.19) and (3.20) have an irregular singularity at t = ∞ and are not integrablethrough Liouvillian solutions due to the presence of Bessel functions; i.e., their Galois groupsare virtually nonsolvable and therefore virtually nonabelian, Theorem 1.2 once again settlingrational nonintegrability for α = 1. In particular, the Galois group of (3.20) is exactly SL2 (C).

  • Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.

    NONAUTONOMOUS HAMILTONIANS AND MORALES–RAMIS I 295

    4. Final remarks: Open questions and future work. This paper is the starting point ofa project in which the author is involved. The following questions arose during our work:

    • In [23, 29] it was proven that the autonomous Hamiltonian system related to Painlevé IIis nonintegrable for every α ∈ Z. Is this also true for (2.6)?

    • Does the integrability of (2.6) for arbitrary α ∈ Z depend on the choice of g(x) anda(t)?

    • Assuming the above question has an affirmative answer, in what manner can thechoice and form of g(x) and a(t) ensure nonintegrability for every α ∈ Z? And forevery α ∈ C?

    • Is it possible to find transversal sections of the flow, and thus Poincaré maps, foreither Ĥ or the algebrized equation, even in the absence of nontrivial numerical mon-odromies? Do Stokes multipliers contribute to the answer in a significant manner?

    Among our next goals, the analysis of the following items is due further immediate re-search:

    • the application of Morales–Ramis theory to higher variational equations of nonauton-omous Hamiltonian systems;

    • differential equations in the form ẍ = f(x, ẋ, t);• the rest of the Painlevé equations: Casale in [10] analyzed Painlevé I, and Horozov

    and Stoyanova in [17] analyzed some particular cases for Painlevé VI;• the theoretical aspects of nonautonomous Hamiltonian systems such as their geometry

    and the feasibility of an analogue to Liouville–Arnold theory;• the nonintegrability of nonautonomous Hamiltonian systems with two and a half de-

    grees of freedom;• specific examples of nonautonomous Hamiltonian systems related to control theory, as

    well as others related to celestial mechanics, such as the restricted three- and four-bodyproblems and Hénon–Heiles systems [4], [13].

    • the exact relation, perhaps causal, between separatrix splitting [24] and nonintegra-bility, whether rational or meromorphic.

    Acknowledgments. The author is indebted to Sergi Simon for multiple suggestions con-cerning organization, style, pictures, and numerical remarks, as well as for preliminary correc-tions concerning the Sitnikov problem; for his collaboration in the numerical analysis thereof,acknowledgments are also due to Carles Simó. The author acknowledges Juan J. Morales-Ruizand Jacques-Arthur Weil for their comments and suggestions and for the final proofreadingof this paper.

    REFERENCES

    [1] R. Abraham and J. E. Marsden, Foundations of Mechanics, 2nd ed., revised and enlarged, Benjamin/Cummings, Advanced Book Program, Reading, MA, 1978.

    [2] M. Abramowitz and I. A. Stegun, eds., Handbook of Mathematical Functions with Formulas, Graphs,and Mathematical Tables, John Wiley & Sons, New York, 1984. Reprint of the 1972 edition.

    [3] P. B. Acosta-Humanez and D. Blázquez-Sanz, Non-integrability of some Hamiltonians with rationalpotentials, Discrete Contin. Dyn. Syst. Ser. B, 10 (2008), pp. 265–293.

    [4] M. A. Andreu, Dynamics in the center manifold around L2 in the quasi-bicircular problem, CelestialMech. Dynam. Astronom., 84 (2002), pp. 105–133.

  • Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.

    296 PRIMITIVO B. ACOSTA-HUMÁNEZ

    [5] V. I. Arnold, Mathematical Methods of Classical Mechanics, Graduate Texts in Mathematics 60,Springer-Verlag, New York, 1978.

    [6] G. Benettin, L. Galgani, and J.-M. Strelcyn, Kolmogorov entropy and numerical experiments,Phys. Rev. A, 14 (1976), pp. 2338–2345.

    [7] D. Boucher, Sur les équations différentielles linéaires paramétrées; une application aux systèmes hamil-toniens, Thèse de l’Université de Limoges, Limoges, France, 2000.

    [8] D. Boucher and J.-A. Weil, Application of J.-J. Morales and J.-P. Ramis’ theorem to test the non-complete integrability of the planar three-body problem, in From Combinatorics to Dynamical Systems,IRMA Lect. Math. Theor. Phys. 3, de Gruyter, Berlin, 2003, pp. 163–177.

    [9] A. Dewisme, S. Bouquet, and P. G. L. Leach, Symmetries of time dependent Hamiltonian systems, inModern Group Analysis: Advanced Analytical and Computational Methods in Mathematical Physics(Acireale, 1992), Kluwer, Dordrecht, The Netherlands, 1993, pp. 173–179.

    [10] G. Casale, Irréductibilité de la première équation de Painlevé, C. R. Math. Acad. Sci. Paris, 343 (2006),pp. 95–98.

    [11] A. Dewisme and S. Bouquet, First integrals and symmetries of time-dependent Hamiltonian systems,J. Math. Phys., 34 (1993), pp. 997–1006.

    [12] D. Fernández Hevia and A. F. Rañada, Chaos in the three-body problem: The Sitnikov case, EuropeanJ. Phys., 17 (1996), pp. 295–302.

    [13] F. Gabern, À. Jorba, and P. Robutel, On the accuracy of restricted three-body models for the Trojanmotion, Discrete Contin. Dyn. Syst., 11 (2004), pp. 843–854.

    [14] G. Giachetta, L. Mangiarotti, and G. Sardanashvily, Action-angle coordinates for time-dependentcompletely integrable Hamiltonian systems, J. Phys. A, 35 (2002), pp. L439–L445.

    [15] M. Gidea and F. Deppe, Chaotic orbits in a restricted three-body problem: Numerical experiments andheuristics, Commun. Nonlinear Sci. Numer. Simul., 11 (2006), pp. 161–171.

    [16] J. Hagel and T. Trenkler, A computer-aided analysis of the Sitnikov problem, Celestial Mechanicsand Dynamical Astronomy, 56 (1993), pp. 81–98.

    [17] E. Horozov and T. Stoyanova, Non-integrability of some Painlevé VI-equations and dilogarithms,Regul. Chaotic Dyn., 12 (2007), pp. 622–629.

    [18] R. S. Kaushal, Construction of exact invariants for time dependent classical dynamical systems, Internat.J. Theoret. Phys., 37 (1998), pp. 1793–1856.

    [19] R. Kuwabara, Time-dependent mechanical symmetries and extended Hamiltonian systems, Rep. Math.Phys., 19 (1984), pp. 27–38.

    [20] C. Lanczos, The Variational Principles of Mechanics, Mathematical Expositions 4, University of TorontoPress, Toronto, Ontario, Canada, 1949.

    [21] A. J. Lichtenberg and M. A. Lieberman, Regular and Chaotic Dynamics, 2nd ed., Applied Mathe-matical Sciences 38, Springer-Verlag, New York, 1992.

    [22] J. J. Morales-Ruiz, Differential Galois Theory and Non-Integrability of Hamiltonian Systems, Progressin Mathematics 179, Birkhäuser Verlag, Basel, 1999.

    [23] J. J. Morales-Ruiz, A remark about the Painlevé transcendents, in Théories asymptotiques et équationsde Painlevé, Sémin. Congr. 14, Soc. Math. France, Paris, 2006, pp. 229–235.

    [24] J. J. Morales-Ruiz and J. M. Peris, On a Galoisian approach to the splitting of separatrices, Ann.Fac. Sci. Toulouse Math. (6), 8 (1999), pp. 125–141.

    [25] J. J. Morales-Ruiz and J. P. Ramis, Galoisian obstructions to integrability of Hamiltonian systems I,Methods Appl. Anal., 8 (2001), pp. 33–96.

    [26] J. J. Morales-Ruiz and J. P. Ramis, Galoisian obstructions to integrability of Hamiltonian systemsII, Methods Appl. Anal., 8 (2001), pp. 97–112.

    [27] J. Moser, Stable and Random Motions in Dynamical Systems. With Special Emphasis on Celestial Me-chanics, Princeton Landmarks in Mathematics, Princeton University Press, Princeton, NJ, 2001.Reprint of the 1973 original.

    [28] E. L. Stiefel and G. Scheifele, Linear and Regular Celestial Mechanics. Perturbed Two-Body Motion,Numerical Methods, Canonical Theory, Grundlehren Math. Wiss. 174, Springer-Verlag, New York,1971.

    [29] T. Stoyanova and O. Christov, Non-integrability of the second Painlevé equation as a Hamiltoniansystem, C. R. Acad. Bulgare Sci., 60 (2007), pp. 13–18.

  • Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.

    NONAUTONOMOUS HAMILTONIANS AND MORALES–RAMIS I 297

    [30] J. Struckmeier, Hamiltonian dynamics on the symplectic extended phase space for autonomous andnon-autonomous systems, J. Phys. A, 38 (2005), pp. 1257–1278.

    [31] J. Struckmeier and C. Riedel, Canonical transformations and exact invariants for time-dependentHamiltonian systems, Ann. Phys. (8), 11 (2002), pp. 15–38.

    [32] J. Struckmeier and C. Riedel, Noether’s theorem and Lie symmetries for time-dependent Hamilton-Lagrange systems, Phys. Rev. E (3), 66 (2002), 066605.

    [33] D. R. Stump, Arnol ′d’s transformation—an example of a generalized canonical transformation, J. Math.Phys., 39 (1998), pp. 3661–3669.

    [34] W. Thirring, Lehrbuch der mathematischen Physik. Band 1. Klassische dynamische Systeme, Springer-Verlag, Vienna, 1977.

    [35] A. V. Tsiganov, Canonical transformations of the extended phase space, Toda lattices and the Stäckelfamily of integrable systems, J. Phys. A, 33 (2000), pp. 4169–4182.

    [36] M. van der Put and M. F. Singer, Galois Theory of Linear Differential Equations, Grundlehren Math.Wiss. 328, Springer-Verlag, Berlin, 2003.

    [37] E. T. Whittaker, A Treatise on the Analytical Dynamics of Particles and Rigid Bodies. With an In-troduction to the Problem of Three Bodies, Cambridge Mathematical Library, Cambridge UniversityPress, Cambridge, UK, 1988. Reprint of the 1937 edition.

    [38] K. Wodnar, New formulations of the Sitnikov problem, in Predictability, Stability, and Chaos in N-BodyDynamical Systems (Cortina d’Ampezzo, 1990), NATO Adv. Sci. Inst. Ser. B Phys. 272, Plenum,New York, 1991, pp. 457–466.

    /ColorImageDict > /JPEG2000ColorACSImageDict > /JPEG2000ColorImageDict > /AntiAliasGrayImages false /CropGrayImages true /GrayImageMinResolution 300 /GrayImageMinResolutionPolicy /OK /DownsampleGrayImages true /GrayImageDownsampleType /Bicubic /GrayImageResolution 300 /GrayImageDepth -1 /GrayImageMinDownsampleDepth 2 /GrayImageDownsampleThreshold 1.50000 /EncodeGrayImages true /GrayImageFilter /DCTEncode /AutoFilterGrayImages true /GrayImageAutoFilterStrategy /JPEG /GrayACSImageDict > /GrayImageDict > /JPEG2000GrayACSImageDict > /JPEG2000GrayImageDict > /AntiAliasMonoImages false /CropMonoImages true /MonoImageMinResolution 1200 /MonoImageMinResolutionPolicy /OK /DownsampleMonoImages true /MonoImageDownsampleType /Bicubic /MonoImageResolution 1200 /MonoImageDepth -1 /MonoImageDownsampleThreshold 1.50000 /EncodeMonoImages true /MonoImageFilter /CCITTFaxEncode /MonoImageDict > /AllowPSXObjects false /CheckCompliance [ /None ] /PDFX1aCheck false /PDFX3Check false /PDFXCompliantPDFOnly false /PDFXNoTrimBoxError true /PDFXTrimBoxToMediaBoxOffset [ 0.00000 0.00000 0.00000 0.00000 ] /PDFXSetBleedBoxToMediaBox true /PDFXBleedBoxToTrimBoxOffset [ 0.00000 0.00000 0.00000 0.00000 ] /PDFXOutputIntentProfile () /PDFXOutputConditionIdentifier () /PDFXOutputCondition () /PDFXRegistryName () /PDFXTrapped /False

    /CreateJDFFile false /Description > /Namespace [ (Adobe) (Common) (1.0) ] /OtherNamespaces [ > /FormElements false /GenerateStructure false /IncludeBookmarks false /IncludeHyperlinks false /IncludeInteractive false /IncludeLayers false /IncludeProfiles false /MultimediaHandling /UseObjectSettings /Namespace [ (Adobe) (CreativeSuite) (2.0) ] /PDFXOutputIntentProfileSelector /DocumentCMYK /PreserveEditing true /UntaggedCMYKHandling /LeaveUntagged /UntaggedRGBHandling /UseDocumentProfile /UseDocumentBleed false >> ]>> setdistillerparams> setpagedevice