Hydrogen and the Central Force Problem

Total Page:16

File Type:pdf, Size:1020Kb

Hydrogen and the Central Force Problem Hydrogen and the Central Force Problem Our treatment of quantum mechanics is getting increasingly more realistic. In the last chapter, we made the leap from 1 to 3 dimensions. In this chapter we solve our first problem with a physically realistic potential– the attractive electrostatic (Coulomb) potential between two oppositely charged particles. In the case of atomic hydrogen these particles are an electron and a proton, but the theory can be easily extended to one electron ions such as Li++, muonic atoms where a muon is captured in orbit about a proton, or positronium (an electron in orbit about an antielectron). In this chapter we will ignore the complications of electron spin. Heretofore we have been discussing the motion of a single particle in an “external” potential of mysterious origin. Because the proton is about 2000 times more massive than the electron and doesn’t move very much, we could think of it as providing an external potential to the “orbiting” electron and still get fairly accurate results. However positronium gives us a great excuse to discuss two particle wave functions which interact in a potential mutually created by both particles. The Two Particle Reduction Lets consider two particles in one dimension. The first particle has mass m1 and is located at x1 ; the second has mass m2 and is located at x2. The wave func- tion becomes a function of both x1 and x2 : ψ(x1,x2). It is natural to generalize 2 the probabilistic interpretation of the wave function PDF(x1,x2)= |ψ(x1,x2)| , where PDF(x1,x2) gives the probability density of finding particle 1 at position x1 and simultaneously finding particle 2 at position x2. To clarify this last point, the probability of finding particle 2 within a distance ∆ of particle 1 would be written as: +∞ x1+∆ ∗ P(|x1 − x2| < ∆) = dx1 dx2 ψ (x1,x2)ψ(x1,x2) −∞ x1−∆ 1 Thewavefunctionψ(x1,x2)obeysaSchr¨odinger equation which is built on 2 2 the total energy of the two particle system: E = p1/(2m1)+p1/(2m1)+V (x1,x2). 2 2 −¯h ∂2 −¯h ∂2 2 ψ(x1,x2)+ 2 ψ(x1,x2)+V (x1,x2) ψ(x1,x2)=Eψ(x1,x2)(1) 2m1 ∂x1 2m2 ∂x2 We thus have a partial differential equation which is traditionally solved using separation of variables. The separation of variables would be trivial if V (x1,x2)=V1(x1)+V2(x2) since we just have a diffiQ in x1 added to a diffiQ in x2. This would be the situation if particle 1 and 2 moved in external potentials but felt no mutual interaction. In the case of positronium, we usually have the opposite problem of zero external potential but the electron and positron (antielectron) attract each other with a potential related to their separation. The one dimensional analogy would be V (x1 −x2). It is very difficult to separate variables for this potential in the original (x1,x2) coordinates but there is a classic change of variables trick which allows for a separation of variables solutions. The clue comes from classical mechanics. In many classical situations motion about center of mass is essentially decoupled from motion of the center of mass. An example would be if one were to throw a system consising of two masses coupled with a spring from the surface of the moon. The masses will oscillate following simple harmonic motion about their center of gravity; while the center of mass follows the same parabolic trajectory as a point like particle in the grav- itational field of the moon. Guided by the classical problem, we tranform the variables of Eq. (1) into the position of the center of mass X and the position of x1 relative to x2: m1 m2 X = x1 + x2 where M = m1 + m2 M M x = x1 − x2 (2) 2 We next transform the derivatives using the chain rule. ∂ ∂X ∂ ∂x ∂ ∂ m1 ∂ ∂ = + , = + ∂x1 ∂x1 ∂X ∂x1 ∂x ∂x1 M ∂X ∂x ∂ ∂X ∂ ∂x ∂ ∂ m2 ∂ ∂ = + , = − (3) ∂x2 ∂x2 ∂X ∂x2 ∂x ∂x2 M ∂X ∂x 2 2 The kinetic energy operator Tˇ = p1/(2m1)+p2/(2m2) can be written in terms of the derivatives of Eq. (3) 2 −¯h m1 ∂ ∂ m1 ∂ ∂ Tˇ = + + 2m1 M ∂X ∂x M ∂X ∂x 2 −¯h m2 ∂ ∂ m2 ∂ ∂ + − − (4) 2m2 M ∂X ∂x M ∂X ∂x We note that the cross terms proportional to ∂2/(∂x∂X) cancel leaving the terms: − 2 2 − 2 2 ˇ ¯h m1 + m2 ∂ ¯h 1 1 ∂ T = 2 2 + + 2 2 M ∂X 2 m1 m2 ∂x − 2 2 − 2 2 ˇ ¯h ∂ ¯h ∂ 1 1 1 T = 2 + 2 where = + (5) 2M ∂X 2µ ∂x µ m1 m2 We see that the kinetic energy has a contribution from the center of mass motion X which enters with a mass M = m1 + m2 and a contribution from the relative −1 −1 −1 motion x = x1 − x2 which enters with an “reduced mass” µ = m1 + m2 . Lets now assume a potential of the form V (x1,x2)=Vr(x1 − x2)+Vcg(X) which might be applicable, for example, to two masses connected with a spring 2 which experience a potential Vr(x1 − x2)=κ(x1 − x2) /2, as the mass spring system falls in a gravitational field Vcg(X)=(m1 gx1 +m2 gx2)=MgX.The 3 Schr¨odinger equation reads: 2 2 −¯h ∂2 −¯h ∂2 + V (x) ψ(x, X)+ + V (X) ψ(x, X)=Eψ(x, X) 2µ ∂x2 r 2M ∂X2 cg r cg (6) We now have a classic separation of variables situation. We factor the wave function as ψ(x, X)=ψr(x) × ψcg(X). Inserting this into Eq. (6) and dividing both sides by ψr(x) × ψcg(X) we get two separate Schr¨odinger equations: 2 −¯h d2 ψ (x)+V (x)ψ (x)=E ψ (x) 2µ dx2 r r r r r 2 −¯h d2 ψ (X)+V (X)ψ (X)=E ψ (X) ,E+ E = E (7) 2M dX2 cg g cg cg cg r cg To gain insight into Eq. (7) lets consider the specific example of two masses m1 and m2 undergoing simple harmonic motion, while their center of gravity travels to the right with a total momentumhK ¯ . If the two masses are tied together with a spring with spring constant κ, the oscillation frequency will be −1 −1 −1 ω = κ/µ where the “reduced mass” is µ = m1 + m2 . The ψcg(X) portion of the wave function which describes the center of gravity is just a free particle traveling wave solution of the Schr¨odinger equation of the form: ψcg(X)=exp(iKX) ; while the ψr(x) portion of the wave function which describes the relative coordinate x1−x2 is constructed out of Hermite polynomials √ 2 and given by ψr(x)=Hn( βx)exp(−βx/2). The complete wave function is: 2 µω ψ(x, X)=N H ( βx) e−βx /2 eiKX where β = (8) n ¯h The energy of the system, or E in Eq. (6), is just equal to the sum of Er (the harmonic oscillator energy) and Ecg (the kinetic energy of the center of mass 4 which acts as a particle of mass M traveling with a momentumhK ¯ . 2 1 ¯h K2 E = n + ¯hω + (9) 2 2M A brief summary might be in order. Quite often in reality, we have the problem of describing a situation where two particles interact with each other through a potential which depends on their separation. In such cases, one can affect a separation of variable solution by writing the Schr¨odinger equation in terms of the relative displacement of two particles and location of their center of mass. We’ve shown that the equation describing the center of mass, becomes totally independent of the equation describing the relative coordinates. The Schr¨odinger equation for the relative displacement between the two bodies is exactly the same as the equation for a single particle of reduced mass µ interacting in an external potential V (x)=V (x1 − x2). Hence the fact that we have two particles rather than one creates no additional mathematical complications at all! The Angular Reduction Besides the reduction of a two particle system down to a one particle system with a reduced mass µ, the conservation of angular momentum creates additional simplifications in the central force problem in both classical mechanics and quan- tum mechanics. A few words about the classical orbit problem (illustrated below) wouldn’t hurt. r V(r) r φ r φ Perigee Apogee 2 L = µ r φ r = µ r φ 5 We have described the elliptical orbit in terms of the polar coordinates r and ˙ 1 φ. The tangential velocity is vφ = r φ; while the radial velocity isr ˙. The energy expression is given by: 1 E = µ r˙2 +(rφ˙)2 + V (r) (10) 2 2 ˙ The angular momentum, given by L = µvφ r = µr φ, is a constant of motion. We can express φ˙ in terms of L and insert it in Eq. (10) to obtain an expression in r only: 2 L 1 2 L φ˙ = ,E= µr˙ + + V (r) (11) µr2 2 2 µr2 Eq. (11) is very analogous to the conservation of energy in a one dimensional 1 2 potential well: E = 2 m x˙ + V (x). To complete the analogy we combine the angular momentum term with the potential actual potential energy to form an “effective potential”. L2 Veff = + V (r) (12) 2 µr2 The L2/(2 µr2) term is often called the “centrifugal potential”. We illustrate radial motion of the orbit for a V = −K/r potential in an effective potential well below: V eff 2 -K L V = + 2 eff r 2µ r dominated by centrifugal potential r E Perigee Apogee dominated by Coulomb potential 1 In classical mechanics, one often writesr ˙ to stand for dr/dt.
Recommended publications
  • Forced Mechanical Oscillations
    169 Carl von Ossietzky Universität Oldenburg – Faculty V - Institute of Physics Module Introductory laboratory course physics – Part I Forced mechanical oscillations Keywords: HOOKE's law, harmonic oscillation, harmonic oscillator, eigenfrequency, damped harmonic oscillator, resonance, amplitude resonance, energy resonance, resonance curves References: /1/ DEMTRÖDER, W.: „Experimentalphysik 1 – Mechanik und Wärme“, Springer-Verlag, Berlin among others. /2/ TIPLER, P.A.: „Physik“, Spektrum Akademischer Verlag, Heidelberg among others. /3/ ALONSO, M., FINN, E. J.: „Fundamental University Physics, Vol. 1: Mechanics“, Addison-Wesley Publishing Company, Reading (Mass.) among others. 1 Introduction It is the object of this experiment to study the properties of a „harmonic oscillator“ in a simple mechanical model. Such harmonic oscillators will be encountered in different fields of physics again and again, for example in electrodynamics (see experiment on electromagnetic resonant circuit) and atomic physics. Therefore it is very important to understand this experiment, especially the importance of the amplitude resonance and phase curves. 2 Theory 2.1 Undamped harmonic oscillator Let us observe a set-up according to Fig. 1, where a sphere of mass mK is vertically suspended (x-direc- tion) on a spring. Let us neglect the effects of friction for the moment. When the sphere is at rest, there is an equilibrium between the force of gravity, which points downwards, and the dragging resilience which points upwards; the centre of the sphere is then in the position x = 0. A deflection of the sphere from its equilibrium position by x causes a proportional dragging force FR opposite to x: (1) FxR ∝− The proportionality constant (elastic or spring constant or directional quantity) is denoted D, and Eq.
    [Show full text]
  • 1 the Basic Set-Up 2 Poisson Brackets
    MATHEMATICS 7302 (Analytical Dynamics) YEAR 2016–2017, TERM 2 HANDOUT #12: THE HAMILTONIAN APPROACH TO MECHANICS These notes are intended to be read as a supplement to the handout from Gregory, Classical Mechanics, Chapter 14. 1 The basic set-up I assume that you have already studied Gregory, Sections 14.1–14.4. The following is intended only as a succinct summary. We are considering a system whose equations of motion are written in Hamiltonian form. This means that: 1. The phase space of the system is parametrized by canonical coordinates q =(q1,...,qn) and p =(p1,...,pn). 2. We are given a Hamiltonian function H(q, p, t). 3. The dynamics of the system is given by Hamilton’s equations of motion ∂H q˙i = (1a) ∂pi ∂H p˙i = − (1b) ∂qi for i =1,...,n. In these notes we will consider some deeper aspects of Hamiltonian dynamics. 2 Poisson brackets Let us start by considering an arbitrary function f(q, p, t). Then its time evolution is given by n df ∂f ∂f ∂f = q˙ + p˙ + (2a) dt ∂q i ∂p i ∂t i=1 i i X n ∂f ∂H ∂f ∂H ∂f = − + (2b) ∂q ∂p ∂p ∂q ∂t i=1 i i i i X 1 where the first equality used the definition of total time derivative together with the chain rule, and the second equality used Hamilton’s equations of motion. The formula (2b) suggests that we make a more general definition. Let f(q, p, t) and g(q, p, t) be any two functions; we then define their Poisson bracket {f,g} to be n def ∂f ∂g ∂f ∂g {f,g} = − .
    [Show full text]
  • The Damped Harmonic Oscillator
    THE DAMPED HARMONIC OSCILLATOR Reading: Main 3.1, 3.2, 3.3 Taylor 5.4 Giancoli 14.7, 14.8 Free, undamped oscillators – other examples k m L No friction I C k m q 1 x m!x! = !kx q!! = ! q LC ! ! r; r L = θ Common notation for all g !! 2 T ! " # ! !!! + " ! = 0 m L 0 mg k friction m 1 LI! + q + RI = 0 x C 1 Lq!!+ q + Rq! = 0 C m!x! = !kx ! bx! ! r L = cm θ Common notation for all g !! ! 2 T ! " # ! # b'! !!! + 2"!! +# ! = 0 m L 0 mg Natural motion of damped harmonic oscillator Force = mx˙˙ restoring force + resistive force = mx˙˙ ! !kx ! k Need a model for this. m Try restoring force proportional to velocity k m x !bx! How do we choose a model? Physically reasonable, mathematically tractable … Validation comes IF it describes the experimental system accurately Natural motion of damped harmonic oscillator Force = mx˙˙ restoring force + resistive force = mx˙˙ !kx ! bx! = m!x! ! Divide by coefficient of d2x/dt2 ! and rearrange: x 2 x 2 x 0 !!+ ! ! + " 0 = inverse time β and ω0 (rate or frequency) are generic to any oscillating system This is the notation of TM; Main uses γ = 2β. Natural motion of damped harmonic oscillator 2 x˙˙ + 2"x˙ +#0 x = 0 Try x(t) = Ce pt C, p are unknown constants ! x˙ (t) = px(t), x˙˙ (t) = p2 x(t) p2 2 p 2 x(t) 0 Substitute: ( + ! + " 0 ) = ! 2 2 Now p is known (and p = !" ± " ! # 0 there are 2 p values) p t p t x(t) = Ce + + C'e " Must be sure to make x real! ! Natural motion of damped HO Can identify 3 cases " < #0 underdamped ! " > #0 overdamped ! " = #0 critically damped time ---> ! underdamped " < #0 # 2 !1 = ! 0 1" 2 ! 0 ! time ---> 2 2 p = !" ± " ! # 0 = !" ± i#1 x(t) = Ce"#t+i$1t +C*e"#t"i$1t Keep x(t) real "#t x(t) = Ae [cos($1t +%)] complex <-> amp/phase System oscillates at "frequency" ω1 (very close to ω0) ! - but in fact there is not only one single frequency associated with the motion as we will see.
    [Show full text]
  • VIBRATIONAL SPECTROSCOPY • the Vibrational Energy V(R) Can Be Calculated Using the (Classical) Model of the Harmonic Oscillator
    VIBRATIONAL SPECTROSCOPY • The vibrational energy V(r) can be calculated using the (classical) model of the harmonic oscillator: • Using this potential energy function in the Schrödinger equation, the vibrational frequency can be calculated: The vibrational frequency is increasing with: • increasing force constant f = increasing bond strength • decreasing atomic mass • Example: f cc > f c=c > f c-c Vibrational spectra (I): Harmonic oscillator model • Infrared radiation in the range from 10,000 – 100 cm –1 is absorbed and converted by an organic molecule into energy of molecular vibration –> this absorption is quantized: A simple harmonic oscillator is a mechanical system consisting of a point mass connected to a massless spring. The mass is under action of a restoring force proportional to the displacement of particle from its equilibrium position and the force constant f (also k in followings) of the spring. MOLECULES I: Vibrational We model the vibrational motion as a harmonic oscillator, two masses attached by a spring. nu and vee! Solving the Schrödinger equation for the 1 v h(v 2 ) harmonic oscillator you find the following quantized energy levels: v 0,1,2,... The energy levels The level are non-degenerate, that is gv=1 for all values of v. The energy levels are equally spaced by hn. The energy of the lowest state is NOT zero. This is called the zero-point energy. 1 R h Re 0 2 Vibrational spectra (III): Rotation-vibration transitions The vibrational spectra appear as bands rather than lines. When vibrational spectra of gaseous diatomic molecules are observed under high-resolution conditions, each band can be found to contain a large number of closely spaced components— band spectra.
    [Show full text]
  • Harmonic Oscillator with Time-Dependent Effective-Mass And
    Harmonic oscillator with time-dependent effective-mass and frequency with a possible application to 'chirped tidal' gravitational waves forces affecting interferometric detectors Yacob Ben-Aryeh Physics Department, Technion-Israel Institute of Technology, Haifa,32000,Israel e-mail: [email protected] ; Fax: 972-4-8295755 Abstract The general theory of time-dependent frequency and time-dependent mass ('effective mass') is described. The general theory for time-dependent harmonic-oscillator is applied in the present research for studying certain quantum effects in the interferometers for detecting gravitational waves. When an astronomical binary system approaches its point of coalescence the gravitational wave intensity and frequency are increasing and this can lead to strong deviations from the simple description of harmonic oscillations for the interferometric masses on which the mirrors are placed. It is shown that under such conditions the harmonic oscillations of these masses can be described by mechanical harmonic-oscillators with time- dependent frequency and effective-mass. In the present theoretical model the effective- mass is decreasing with time describing pumping phenomena in which the oscillator amplitude is increasing with time. The quantization of this system is analyzed by the use of the adiabatic approximation. It is found that the increase of the gravitational wave intensity, within the adiabatic approximation, leads to squeezing phenomena where the quantum noise in one quadrature is increased and in the other quadrature it is decreased. PACS numbers: 04.80.Nn, 03.65.Bz, 42.50.Dv. Keywords: Gravitational waves, harmonic-oscillator with time-dependent effective- mass 1 1.Introduction The problem of harmonic-oscillator with time-dependent mass has been related to a quantum damped oscillator [1-7].
    [Show full text]
  • Integrals of Motion Key Point #15 from Bachelor Course
    KeyKey12-10-07 see PointPoint http://www.strw.leidenuniv.nl/˜ #15#10 fromfrom franx/college/ BachelorBachelor mf-sts-07-c4-5 Course:Course:12-10-07 see http://www.strw.leidenuniv.nl/˜ IntegralsIntegrals ofof franx/college/ MotionMotion mf-sts-07-c4-6 4.2 Constants and Integrals of motion Integrals (BT 3.1 p 110-113) Much harder to define. E.g.: 1 2 Energy (all static potentials): E(⃗x,⃗v)= 2 v + Φ First, we define the 6 dimensional “phase space” coor- dinates (⃗x,⃗v). They are conveniently used to describe Lz (axisymmetric potentials) the motions of stars. Now we introduce: L⃗ (spherical potentials) • Integrals constrain geometry of orbits. • Constant of motion: a function C(⃗x,⃗v, t) which is constant along any orbit: examples: C(⃗x(t1),⃗v(t1),t1)=C(⃗x(t2),⃗v(t2),t2) • 1. Spherical potentials: E,L ,L ,L are integrals of motion, but also E,|L| C is a function of ⃗x, ⃗v, and time t. x y z and the direction of L⃗ (given by the unit vector ⃗n, Integral of motion which is defined by two independent numbers). ⃗n de- • : a function I(x, v) which is fines the plane in which ⃗x and ⃗v must lie. Define coor- constant along any orbit: dinate system with z axis along ⃗n I[⃗x(t1),⃗v(t1)] = I[⃗x(t2),⃗v(t2)] ⃗x =(x1,x2, 0) I is not a function of time ! Thus: integrals of motion are constants of motion, ⃗v =(v1,v2, 0) but constants of motion are not always integrals of motion! → ⃗x and ⃗v constrained to 4D region of the 6D phase space.
    [Show full text]
  • STUDENT HIGHLIGHT – Hassan Najafi Alishah - “Symplectic Geometry, Poisson Geometry and KAM Theory” (Mathematics)
    SEPTEMBER OCTOBER // 2010 NUMBER// 28 STUDENT HIGHLIGHT – Hassan Najafi Alishah - “Symplectic Geometry, Poisson Geometry and KAM theory” (Mathematics) Hassan Najafi Alishah, a CoLab program student specializing in math- The Hamiltonian function (or energy function) is constant along the ematics, is doing advanced work in the Mathematics Departments of flow. In other words, the Hamiltonian function is a constant of motion, Instituto Superior Técnico, Lisbon, and the University of Texas at Austin. a principle that is sometimes called the energy conservation law. A This article provides an overview of his research topics. harmonic spring, a pendulum, and a one-dimensional compressible fluid provide examples of Hamiltonian Symplectic and Poisson geometries underlie all me- systems. In the former, the phase spaces are symplec- chanical systems including the solar system, the motion tic manifolds, but for the later phase space one needs of a rigid body, or the interaction of small molecules. the more general concept of a Poisson manifold. The origins of symplectic and Poisson structures go back to the works of the French mathematicians Joseph-Louis A Hamiltonian system with enough suitable constants Lagrange (1736-1813) and Simeon Denis Poisson (1781- of motion can be integrated explicitly and its orbits can 1840). Hermann Weyl (1885-1955) first used “symplectic” be described in detail. These are called completely in- with its modern mathematical meaning in his famous tegrable Hamiltonian systems. Under appropriate as- monograph “The Classical groups.” (The word “sym- sumptions (e.g., in a compact symplectic manifold) plectic” itself is derived from a Greek word that means the motions of a completely integrable Hamiltonian complex.) Lagrange introduced the concept of a system are quasi-periodic.
    [Show full text]
  • Lecture 8 Symmetries, Conserved Quantities, and the Labeling of States Angular Momentum
    Lecture 8 Symmetries, conserved quantities, and the labeling of states Angular Momentum Today’s Program: 1. Symmetries and conserved quantities – labeling of states 2. Ehrenfest Theorem – the greatest theorem of all times (in Prof. Anikeeva’s opinion) 3. Angular momentum in QM ˆ2 ˆ 4. Finding the eigenfunctions of L and Lz - the spherical harmonics. Questions you will by able to answer by the end of today’s lecture 1. How are constants of motion used to label states? 2. How to use Ehrenfest theorem to determine whether the physical quantity is a constant of motion (i.e. does not change in time)? 3. What is the connection between symmetries and constant of motion? 4. What are the properties of “conservative systems”? 5. What is the dispersion relation for a free particle, why are E and k used as labels? 6. How to derive the orbital angular momentum observable? 7. How to check if a given vector observable is in fact an angular momentum? References: 1. Introductory Quantum and Statistical Mechanics, Hagelstein, Senturia, Orlando. 2. Principles of Quantum Mechanics, Shankar. 3. http://mathworld.wolfram.com/SphericalHarmonic.html 1 Symmetries, conserved quantities and constants of motion – how do we identify and label states (good quantum numbers) The connection between symmetries and conserved quantities: In the previous section we showed that the Hamiltonian function plays a major role in our understanding of quantum mechanics using it we could find both the eigenfunctions of the Hamiltonian and the time evolution of the system. What do we mean by when we say an object is symmetric?�What we mean is that if we take the object perform a particular operation on it and then compare the result to the initial situation they are indistinguishable.�When one speaks of a symmetry it is critical to state symmetric with respect to which operation.
    [Show full text]
  • Conservation of Total Momentum TM1
    L3:1 Conservation of total momentum TM1 We will here prove that conservation of total momentum is a consequence Taylor: 268-269 of translational invariance. Consider N particles with positions given by . If we translate the whole system with some distance we get The potential energy must be una ected by the translation. (We move the "whole world", no relative position change.) Also, clearly all velocities are independent of the translation Let's consider to be in the x-direction, then Using Lagrange's equation, we can rewrite each derivative as I.e., total momentum is conserved! (Clearly there is nothing special about the x-direction.) L3:2 Generalized coordinates, velocities, momenta and forces Gen:1 Taylor: 241-242 With we have i.e., di erentiating w.r.t. gives us the momentum in -direction. generalized velocity In analogy we de✁ ne the generalized momentum and we say that and are conjugated variables. Note: Def. of gen. force We also de✁ ne the generalized force from Taylor 7.15 di ers from def in HUB I.9.17. such that generalized rate of change of force generalized momentum Note: (generalized coordinate) need not have dimension length (generalized velocity) velocity momentum (generalized momentum) (generalized force) force Example: For a system which rotates around the z-axis we have for V=0 s.t. The angular momentum is thus conjugated variable to length dimensionless length/time 1/time mass*length/time mass*(length)^2/time (torque) energy Cyclic coordinates = ignorable coordinates and Noether's theorem L3:3 Cyclic:1 We have de ned the generalized momentum Taylor: From EL we have 266-267 i.e.
    [Show full text]
  • 22.51 Course Notes, Chapter 9: Harmonic Oscillator
    9. Harmonic Oscillator 9.1 Harmonic Oscillator 9.1.1 Classical harmonic oscillator and h.o. model 9.1.2 Oscillator Hamiltonian: Position and momentum operators 9.1.3 Position representation 9.1.4 Heisenberg picture 9.1.5 Schr¨odinger picture 9.2 Uncertainty relationships 9.3 Coherent States 9.3.1 Expansion in terms of number states 9.3.2 Non-Orthogonality 9.3.3 Uncertainty relationships 9.3.4 X-representation 9.4 Phonons 9.4.1 Harmonic oscillator model for a crystal 9.4.2 Phonons as normal modes of the lattice vibration 9.4.3 Thermal energy density and Specific Heat 9.1 Harmonic Oscillator We have considered up to this moment only systems with a finite number of energy levels; we are now going to consider a system with an infinite number of energy levels: the quantum harmonic oscillator (h.o.). The quantum h.o. is a model that describes systems with a characteristic energy spectrum, given by a ladder of evenly spaced energy levels. The energy difference between two consecutive levels is ∆E. The number of levels is infinite, but there must exist a minimum energy, since the energy must always be positive. Given this spectrum, we expect the Hamiltonian will have the form 1 n = n + ~ω n , H | i 2 | i where each level in the ladder is identified by a number n. The name of the model is due to the analogy with characteristics of classical h.o., which we will review first. 9.1.1 Classical harmonic oscillator and h.o.
    [Show full text]
  • Exact Solution for the Nonlinear Pendulum (Solu¸C˜Aoexata Do Pˆendulon˜Aolinear)
    Revista Brasileira de Ensino de F¶³sica, v. 29, n. 4, p. 645-648, (2007) www.sb¯sica.org.br Notas e Discuss~oes Exact solution for the nonlinear pendulum (Solu»c~aoexata do p^endulon~aolinear) A. Bel¶endez1, C. Pascual, D.I. M¶endez,T. Bel¶endezand C. Neipp Departamento de F¶³sica, Ingenier¶³ade Sistemas y Teor¶³ade la Se~nal,Universidad de Alicante, Alicante, Spain Recebido em 30/7/2007; Aceito em 28/8/2007 This paper deals with the nonlinear oscillation of a simple pendulum and presents not only the exact formula for the period but also the exact expression of the angular displacement as a function of the time, the amplitude of oscillations and the angular frequency for small oscillations. This angular displacement is written in terms of the Jacobi elliptic function sn(u;m) using the following initial conditions: the initial angular displacement is di®erent from zero while the initial angular velocity is zero. The angular displacements are plotted using Mathematica, an available symbolic computer program that allows us to plot easily the function obtained. As we will see, even for amplitudes as high as 0.75¼ (135±) it is possible to use the expression for the angular displacement, but considering the exact expression for the angular frequency ! in terms of the complete elliptic integral of the ¯rst kind. We can conclude that for amplitudes lower than 135o the periodic motion exhibited by a simple pendulum is practically harmonic but its oscillations are not isochronous (the period is a function of the initial amplitude).
    [Show full text]
  • Solving the Harmonic Oscillator Equation
    Solving the Harmonic Oscillator Equation Morgan Root NCSU Department of Math Spring-Mass System Consider a mass attached to a wall by means of a spring. Define y=0 to be the equilibrium position of the block. y(t) will be a measure of the displacement from this equilibrium at a given time. Take dy(0) y(0) = y0 and dt = v0. Basic Physical Laws Newton’s Second Law of motion states tells us that the acceleration of an object due to an applied force is in the direction of the force and inversely proportional to the mass being moved. This can be stated in the familiar form: Fnet = ma In the one dimensional case this can be written as: Fnet = m&y& Relevant Forces Hooke’s Law (k is FH = −ky called Hooke’s constant) Friction is a force that FF = −cy& opposes motion. We assume a friction proportional to velocity. Harmonic Oscillator Assuming there are no other forces acting on the system we have what is known as a Harmonic Oscillator or also known as the Spring-Mass- Dashpot. Fnet = FH + FF or m&y&(t) = −ky(t) − cy&(t) Solving the Simple Harmonic System m&y&(t) + cy&(t) + ky(t) = 0 If there is no friction, c=0, then we have an “Undamped System”, or a Simple Harmonic Oscillator. We will solve this first. m&y&(t) + ky(t) = 0 Simple Harmonic Oscillator k Notice that we can take K = m and look at the system: &y&(t) = −Ky(t) We know at least two functions that will solve this equation.
    [Show full text]