Orbits, Integrals, and Chaos

Similar documents
Surfaces of Section I

We have seen how to calculate forces and potentials from the smoothed density ρ.

4 Orbits in stationary Potentials (BT 3 to page 107)

= 0. = q i., q i = E

Lecture 6: The Orbits of Stars

Physics 106b: Lecture 7 25 January, 2018

Lecture XIX: Particle motion exterior to a spherical star

The motions of stars in the Galaxy

Potential/density pairs and Gauss s law

Use conserved quantities to reduce number of variables and the equation of motion (EOM)

Chaos in the Hénon-Heiles system

90 CHAPTER 15. AXISYMMETRIC & TRIAXIAL MODELS the resulting model, or work backwards from a model towards a distribution function. From fèe;j z è to ç

Stellar Dynamics and Structure of Galaxies

Potential-Density pair: Worked Problem In a spherical galaxy, the density of matter varies with radius as. a r 2 (r+a) 2, ρ(r) = M 4π

Physics 113!!!!! Spring 2009!! Quantum Theory Seminar #9

The Two -Body Central Force Problem

M2A2 Problem Sheet 3 - Hamiltonian Mechanics

Graduate Written Examination Spring 2014 Part I Thursday, January 16th, :00am to 1:00pm

arxiv: v1 [astro-ph.ga] 4 Jun 2015

University of Illinois at Chicago Department of Physics

Physics 228 Today: Ch 41: 1-3: 3D quantum mechanics, hydrogen atom

Fig. 1. On a sphere, geodesics are simply great circles (minimum distance). From

Angular momentum. Quantum mechanics. Orbital angular momentum

Hamiltonian Chaos. Niraj Srivastava, Charles Kaufman, and Gerhard Müller. Department of Physics, University of Rhode Island, Kingston, RI

The first order formalism and the transition to the

Chapter 9 Uniform Circular Motion

Chaos in Hamiltonian systems

Projection Effects. Orbits in a static spherical potential B& T sec 3.1

Physics 351, Spring 2015, Homework #6. Due at start of class, Friday, February 27, 2015

Kerr black hole and rotating wormhole

Problems in Magnetostatics

arxiv: v1 [astro-ph.ga] 10 Dec 2015

Two-Body Problem. Central Potential. 1D Motion

A A + B. ra + A + 1. We now want to solve the Einstein equations in the following cases:

arxiv: v1 [gr-qc] 26 Jun 2008

Black Holes. Jan Gutowski. King s College London

A plane autonomous system is a pair of simultaneous first-order differential equations,

Physics 106a, Caltech 4 December, Lecture 18: Examples on Rigid Body Dynamics. Rotating rectangle. Heavy symmetric top

UPDATED: NOVEMBER A particle of mass m and energy E moving in one dimension comes in from - and encounters the repulsive potential:

Goldstein Problem 2.17 (3 rd ed. # 2.18)

CP1 REVISION LECTURE 3 INTRODUCTION TO CLASSICAL MECHANICS. Prof. N. Harnew University of Oxford TT 2017

Phys 622 Problems Chapter 5

PHYS2330 Intermediate Mechanics Fall Final Exam Tuesday, 21 Dec 2010

Ensenada March 2008 Galaxy halos at (very) high resolution

e x3 dx dy. 0 y x 2, 0 x 1.

Epicycles the short form.

Two dimensional oscillator and central forces

arxiv: v4 [nlin.cd] 12 Sep 2011

UPDATED: OCTOBER Problem #1

Lecture 8 Phase Space, Part 2. 1 Surfaces of section. MATH-GA Mechanics

F1.9AB2 1. r 2 θ2 + sin 2 α. and. p θ = mr 2 θ. p2 θ. (d) In light of the information in part (c) above, we can express the Hamiltonian in the form

Overview spherical accretion

Poisson Equation. The potential-energy tensor. Potential energy: work done against gravitational forces to assemble a distribution of mass ρ(x)

Question 1: Spherical Pendulum

Quantum Mechanics in Three Dimensions

AY202a Galaxies & Dynamics Lecture 7: Jeans Law, Virial Theorem Structure of E Galaxies

A645/A445: Exercise #1. The Kepler Problem

HILL S EQUATION WITH RANDOM FORCING TERMS. Fred C. Adams and Anthony M. Bloch Also: Suzie Butler, Jeff Druce and Jake Ketchum

The Schwarzschild Metric

for changing independent variables. Most simply for a function f(x) the Legendre transformation f(x) B(s) takes the form B(s) = xs f(x) with s = df

Homework Two. Abstract: Liu. Solutions for Homework Problems Two: (50 points total). Collected by Junyu

Chaotic Billiards. Part I Introduction to Dynamical Billiards. 1 Review of the Dierent Billiard Systems. Ben Parker and Alex Riina.

Dynamics of Galaxies: Practice. Frontiers in Numerical Gravitational Astrophysics July 3, 2008

MATH H53 : Final exam

( u,v). For simplicity, the density is considered to be a constant, denoted by ρ 0

Central force motion/kepler problem. 1 Reducing 2-body motion to effective 1-body, that too with 2 d.o.f and 1st order differential equations

Kinematics. Chapter Multi-Body Systems

Physics 221A Fall 1996 Notes 12 Orbital Angular Momentum and Spherical Harmonics

Physics 351, Spring 2017, Homework #2. Due at start of class, Friday, January 27, 2017

Lecture 11 : Overview

2 General Relativity. 2.1 Curved 2D and 3D space

is conserved, calculating E both at θ = 0 and θ = π/2 we find that this happens for a value ω = ω given by: 2g

Physics 202 Laboratory 3. Root-Finding 1. Laboratory 3. Physics 202 Laboratory

Collisionless Boltzmann Eq (Vlasov eq)" S+G sec 3.4! Collisionless Boltzmann Eq S&G 3.4!

Curvilinear coordinates

Solutions for the Practice Final - Math 23B, 2016

06. Lagrangian Mechanics II

Pentahedral Volume, Chaos, and Quantum Gravity

Hamiltonian Lecture notes Part 3

Lecture 1: A Preliminary to Nonlinear Dynamics and Chaos

PHZ 6607 Fall 2004 Homework #4, Due Friday, October 22, 2004

Geometry and Physics. Amer Iqbal. March 4, 2010

Lecture 1: Oscillatory motions in the restricted three body problem

A Guide to the Next Few Lectures!

Physics 115/242 The leapfrog method and other symplectic algorithms for integrating Newton s laws of motion

MATH 332: Vector Analysis Summer 2005 Homework

Solution Set Eight. 1 Problem #1: Toroidal Electromagnet with Gap Problem #4: Self-Inductance of a Long Solenoid. 9

ENGI Duffing s Equation Page 4.65

(a) Sections 3.7 through (b) Sections 8.1 through 8.3. and Sections 8.5 through 8.6. Problem Set 4 due Monday, 10/14/02

Hamiltonian chaos and dark matter

Hamiltonian. March 30, 2013

Stability of circular orbits

1. < 0: the eigenvalues are real and have opposite signs; the fixed point is a saddle point

Dynamics of interacting vortices on trapped Bose-Einstein condensates. Pedro J. Torres University of Granada

Noncollision Singularities in the n-body Problem

Chapter 8. Orbits. 8.1 Conics

Solution Set Two. 1 Problem #1: Projectile Motion Cartesian Coordinates Polar Coordinates... 3

arxiv:chao-dyn/ v1 12 Feb 1996

TWO DIMENSIONAL FLOWS. Lecture 5: Limit Cycles and Bifurcations

Hamiltonian Chaos and the standard map

Transcription:

Chapter 7 Orbits, Integrals, and Chaos In n space dimensions, some orbits can be formally decomposed into n independent periodic motions. These are the regular orbits; they may be represented as winding paths on an n-dimensional torus. On the other hand, irregular or stochastic orbits defy any such representation; in theory such orbits may wander anywhere permitted by conservation of energy. 7.1 Constants & Integrals of Motion Constants of motion are functions of phase-space coordinates and time which are constant along orbits: C(r(t);v(t);t)=const:; (7.1) where r(t) and v(t)=dr=dt are a solution to the equations of motion. The functionc(r;v;t) must be constant along every orbit, with a value which depends on the orbit. In a phase-space of 2n dimensions there are always 2n independent constants of motion. For example, the 2n initial conditions (r 0 ;v 0 ) of an orbit are constants of motion; given phase-space coordinates (r;v) at time t, integrate the orbit backwards to t = 0 and read off the initial (r 0 ;v 0 ). Integrals of motion are functions of phase-space coordinates alone which are constant along orbits: I(r(t);v(t)) = const:: (7.2) An integral of motion can t depend on time; thus all integrals are constants of motion, but not all constants are integrals. Integrals come in two varieties: isolating and non-isolating. Isolating integrals are important because they constrain the shapes of orbits; in a phase-space of 2n dimensions, an isolating integral defines a hypersurface of 2n 1 dimensions. Regular orbits are those which have N = n isolating integrals; in such cases each orbit is confined to a hypersurface of 2n N = n dimensions. 7.2 Orbits in Spherical Potentials Consider the motion of a star in a spherically-symmetric potential, Φ = Φ(jrj). The orbit of the star remains in a plane perpendicular to the angular momentum vector, so it s natural to adopt a polar 51

52 CHAPTER 7. ORBITS, INTEGRALS, AND CHAOS coordinate system; call the coordinates R = jrj and φ. The system has n = 2 degrees of freedom, so the phase space has 4 dimensions. All orbits in spherical potentials are regular; they have two isolating integrals. The equations of motion can be derived by starting with the lagrangian, L(R;φ;Ṙ; φ)= 1 2 (Ṙ2 +R 2 φ 2 ) Φ(R); (7.3) where Ṙ = dr=dt and φ = dφ =dt. Differentiating with respect to Ṙ and φ yields the momenta conjugate to R and φ, L Ṙ = Ṙ = v R ; L φ = R2 φ = Rv φ = J ; (7.4) here v R and v φ are velocities in the radial and azimuthal directions. The hamiltonian may now be expressed as a function of the coordinates and conjugate momenta: Then the equations of motion are H(R;φ;v R ;J)= 1 2 (v2 R+J 2 =R 2 )+Φ(R): (7.5) dr dt = H = v R ; v R dv R dt = H R = dφ dr + J2 R 3 ; (7.6) dφ dt = H J = J R 2 ; dj dt = H φ = 0: Here dj=dt = 0 because (7.5) is independent of the conjugate coordinate φ. The two independent integrals of motion are thus the total energy E, numerically equal to the value of H, and the angularmomentumj. These quantities are given by E = 1 2 (v2 R + v 2 φ)+φ(r); J=Rv φ : (7.7) Each of these integrals defines a hypersurface in phase space, and the orbit is confined to the intersection of these hypersurfaces. This can be visualized by ignoring the φ coordinate and drawing surfaces of constant E and J in the three-dimensional space (R;v R ;v φ ), as in Fig. 7.1. Surfaces of constant E are figures of revolution about the R axis, while surfaces of constant J are hyperbolas in the (R;v φ ) plane. The intersection of these surfaces is a closed curve, and an orbit travels around this curve. For an orbit of a given J, the system may be reduced to one degree of freedom by defining the effective potential, Ψ(R)=Φ(R)+ J2 2R 2 (7.8) the corresponding equations of motion are then just dr dt = v R ; dv R dt = dψ dr : (7.9) Because Ψ(R) diverges as R! 0, the star is energetically prohibited from coming too close to the origin, and shuttles back and forth between turning points R min and R max. In addition to its periodic radial motion described by (7.9), a star also executes a periodic azimuthal motion as it orbits the center of the potential. If the radial and azimuthal periods are incommensurate, as is usually the case, the resulting orbit never returns to its starting point in phase space;

7.3. ORBITS IN AXISYMMETRIC POTENTIALS 53 Figure 7.1: An orbit in a Hernquist (1993) potential as the intersection of surfaces of constant E and J. Left: surface of constant E. Middle: surface of constant J. Right: intersection of these surfaces. in coordinate space such an orbit is a rosette (BT87, Fig. 3-1). The Keplerian potential is a very special case in which the radial and azimuthal periods of all bound orbits are equal. The only other potential in which all orbits are closed is the harmonic potential generated by a uniform sphere; here the radial period is half the azimuthal one and all bound orbits are ellipses centered on the bottom of the potential well. Thus in the Keplerian case all stars advance in azimuth by φ = 2π between successive pericenters, while in the harmonic case they advance by φ = π. Galaxies typically have mass distributions intermediate between these extreme cases, so most orbits in spherical galaxies are rosettes advancing by π < φ < 2π between pericenters. This combination of radial and azimuthal motions can be represented as a path on a 2-torus; that is, on a rectangle made periodic by gluing both opposing pairs of edges together. Associate the long direction on the torus with the azimuthal direction of the orbit, and the short direction on the torus with the closed curve (7.9) produces on the (R;v R ) plane; the result is known as an invariant torus. An orbit winds solenoid-like on the surface of the torus, typically making between one and two turns through the hole of the torus for each turn made around its circumference. Points on the surface of the torus may be parameterized by a pair of angles (θ 1 ;θ 2 ); moreover, by stretching the torus appropriately, the motion of a star can be described by a pair of linear relations: θ 1 (t)=θ 1 (0)+ω 1 t; θ 2 (t)=θ 2 (0)+ω 2 t; (7.10) where the ω i are integrals of motion. Together, (θ 1 ;θ 2 ) and their conjugate angular frequencies or actions (ω 1 ;ω 2 ) define a coordinate system for the four-dimensional phase space in which the hamiltonian takes the simplest possible form, these coordinates are known as action-angle variables. H(θ 1 ;θ 2 ;ω 1 ;ω 2 )= 1 2 (ω2 1+ω 2 2); (7.11) 7.3 Orbits in Axisymmetric Potentials In describing axisymmetric galaxy models it s natural to use cylindrical coordinates (R;φ;z), where Rand φ are polar coordinates in the equatorial plane, and z is the coordinate perpendicular to that

54 CHAPTER 7. ORBITS, INTEGRALS, AND CHAOS Figure 7.2: Surface of section for five orbits in the logarithmic potential. plane. In these coordinates, the potential has the form Φ = Φ(R;z). The equations of motion are identical to (7.6), with additional expressions for z and v z : dz dt = v z ; dv z dt Once again, there are two classical integrals of motion: = Φ z : (7.12) E = 1 2 (v2 R + v 2 φ + v 2 z)+φ(r;z); J z =Rv φ : (7.13) Just as for spherical potentials, it s possible to define an effective potential Ψ(R;z)=Φ(R;z)+ J2 z 2R 2 (7.14) Instead of governing motion along a line as in the spherical case, the effective potential now governs the star s motion in the meridional plane, which rotates about the z axis with angular velocity ω = J z =R 2. The radial motion is described by (7.9), while the vertical motion is described by (7.12) with Φ replaced by Ψ. On the meridional plane, which has coordinates (R;z), the effective potential Ψ has a minimum at R > 0andz= 0 and a steep angular momentum barrier as R! 0 (BT87, Fig. 3-2). If only the energy E constrains the motion of a star on this plane, one might expect a star to travel everywhere within some closed contour of constant Ψ. But in many cases this is not observed; instead, stars launched from rest at differentpoints along a contourof constant Ψ follow distinct trajectories. This implies the existence of a third integral besides the classic integrals given by (7.13). No general expression for a third integral exists. The surface of section is a simple and elegant technique for visualizing the non-classical integrals of an orbit. It is normally only applicable to systems with n = 2 dimensions, but in the axisymmetric case, we can use a surface of section to analyze orbits in the meridional plane. To construct a

7.4. ORBITS IN NON-AXISYMMETRIC POTENTIALS 55 surface of section for this case, simply follow the orbit and plot the phase-space coordinates (R;Ṙ) whenever the orbit crosses the z = 0 plane in an upward direction (ż > 0). Fig. 7.2 presents results for five different orbits in the axisymmetric logarithmic potential, Φ(R;z)= 1 2 ln(r2 c + R 2 + z 2 =q 2 ) ; (7.15) where R c is a core radius and q < 0 is the axial ratio. All of these orbits started from (R;φ;z)= (1;0;0)with angular momentum J z = 0:4, and all have the same total energy E. IfEand J z were the only integrals of motion, we would expect each orbit to cross the surface of section anywhere within some energetically permitted region. Instead, each orbit defines a distinct contour on the (R;Ṙ) plane; these contours are level surfaces of the mysterious third integral. The existence of a third integral implies that a star s orbit is a combination of three periodic motions: radial, azimuthal, and vertical. Thus the orbit can be represented as a path on an invariant 3-torus, with action-angle variables (θ i ;ω i ),wherei= 1;2;3. Some axisymmetric potentials do have orbits which wander everywhere energetically permitted on the meridional plane. For such orbits, the description in terms of motion on a invariant 3-torus breaksdown. These are examplesofirregular or stochastic orbits in an axisymmetric potential; they respect only the two classical integrals, E and J z. 7.4 Orbits in Non-Axisymmetric Potentials Non-axisymmetric potentials, with Φ = Φ(x; y) or Φ(x; y; z) in Cartesian coordinates (x;y;z), admit an even richer variety of orbits. The only classical integral of motion in such a potential is the energy, E = 1 2 jvj2 + Φ(r): (7.16) Some potentials nonetheless permit other integrals of motion, and in such potentials regular orbits may be mapped onto invariant tori. But not all regular orbits can be continuously deformed into one another; consequently, orbits can be grouped into topologically distinct orbit families. Each regular orbit family will generally require a different set of invariant tori. Some sense of the variety of possible orbits in non-axisymmetric galaxies is available by examining orbits in separable and scale-free potentials. 7.4.1 Separable Potentials In a separable potential all orbits are regular and the mapping to the invariant tori can be constructed analytically; all integrals of motion are known. Separable potentials are rather special, mathematically speaking, and it s highly unlikely that real galaxies have such potentials. However, numerical experiments show that non-axisymmetric galaxy models with finite cores or shallow cusps usually generate potentials with many key features of separable potentials. The orbits in a separable potential may be classified into distinct families, each associated with a set of closed and stable orbits. In two dimensions, for example, there are two types of closed, stable orbits; one type (i) oscillates back and forth along the major axis, and the other type (ii) loops around the center. Because these orbits are stable, other orbits which start nearby will remain nearby at later times. The families associated with types (i) and (ii) are known as box and loop

56 CHAPTER 7. ORBITS, INTEGRALS, AND CHAOS Figure 7.3: Surface of section for five orbits in the logarithmic potential. orbits, respectively (BT87, Ch. 3.3.1). Fig 7.3 presents a surface of section for orbits in the nonaxisymmetric logarithmic potential Φ(x;y)= 1 2 ln(r2 c + x 2 + y 2 =q 2 ) : (7.17) (Technically speaking, this is not a separable potential, but as noted above it shares many features with separable potentials for R c > 0.) This surface of section was generated by following each orbit and plotting (x;ẋ) whenever the orbit crossed the y = 0 axis with ẏ > 0. All five orbits shown have the same energy; the outer three are box orbits, while the inner two are loops. In three dimensions, a separable potential permits four distinct orbit families: 1. box orbits, 2. short-axis tube orbits, 3. inner long-axis tube orbits, and 4. outer long-axis tube orbits. The short-axis tubes are orbits which loop around the short (minor) axis, while long-axis tubes loop around the long (major) axis. The two families of long-axis tube orbits arise from different closed stable orbits and explore different regions of space (BT87, Fig. 3-20). No intermediate-axis tube orbits exist since closed orbits looping around the intermediate axis are unstable. In general, triaxial potentials with cores have orbit families much like those in separable potentials. 7.4.2 Scale-Free Potentials In scale-free models all properties have either a power-law or a logarithmic dependence on radius. In particular, scale-free models with density profiles proportional to r 2 have logarithmic potentials

7.5. IRREGULAR ORBITS 57 Figure 7.4: Surface of section for five orbits in the perturbed logarithmic potential. and flat rotation curves. While real galaxies are not entirely scale-free, such steep power-law density distributions are reasonable approximations to the central regions of some elliptical galaxies and to the halos of galaxies in general. If the density falls as r 2 or faster, then box orbits are replaced by minor orbital families called boxlets (Gerhard & Binney 1985, Miralda-Escude & Schwarzschild 1989). Each boxlet family is associated with a closed and stable orbit arising from a resonance between the motions in the x and y directions.

58 CHAPTER 7. ORBITS, INTEGRALS, AND CHAOS 7.5 Irregular orbits Fig. 7.4 presents surfaces of section for orbits in the potential Φ(x;y)= 1 2 ln(r2 c + x 2 + y 2 =q 2 p x 2 + y 2 (x 2 y 2 )=R e ) ; (7.18) where R e is a second scale radius (not to be confused with the effective radius of a de Vaucouleurs profile!). In the limit R e! this reduces to (7.17). The top panel shows results for R e = 3:0, while the lower panel shows R e = 1:5; note the larger chaotic zone in the latter case. In principle, an irregular orbit can wander everywhere on the phase-space hypersurface of constant E, but in actuality such orbits show a complicated and often fractal-like structure.