Chapter 11 Wave Mechanics

Total Page:16

File Type:pdf, Size:1020Kb

Chapter 11 Wave Mechanics Chapter 11 Wave Mechanics P. J. Grandinetti Chem. 4300 P. J. Grandinetti Chapter 11: Wave Mechanics Wave Function in Position & Wavenumber Representation In quantum mechanics, we associate wavenumber with particle momentum p⃗ = `k⃗ Remember the Fourier transform relation for waves? ∞ ∞ ; ù1 ; *ikx ; ù1 ; ikx a.k t/ = Ê Ψ(x t/e dx and Ψ(x t/ = Ê a.k t/e dk 2휋 −∞ 2휋 −∞ Let’s replace k with p in the Fourier expansions... P. J. Grandinetti Chapter 11: Wave Mechanics Wave Function in Position and Momentum Representation Wave function in position and momentum representations become ∞ ; ù1 ; *ipx_` : Φ(p t/ = Ê Ψ(x t/ e dx 2휋` −∞ and ∞ ; ù1 ; ipx_` ; Ψ(x t/ = Ê Φ(p t/ e dp 2휋` −∞ We’ll use these expressions to predict a particle’s properties from its wave function. P. J. Grandinetti Chapter 11: Wave Mechanics Operators in Quantum Mechanics P. J. Grandinetti Chapter 11: Wave Mechanics Position Operators, x̂ Mean particle position, or expectation value of x, is weighted average calculated from wave function ∞ ∞ ê ë < ; ; x.t/ = Ê x p.x/ dx = Ê x Ψ .x t)Ψ(x t/ dx −∞ −∞ «­­­­­¯­­­­­¬ p.x/ In QM, we define x̂ operator: x̂ Ψ(x; t/ = x Ψ(x; t/ x̂ operates on Ψ(x; t/ to give back position times wave function. In formalism of quantum mechanics, we write êx.t/ë as ∞ ∞ ∞ ê ë < ; ̂ ; < ; ; < ; ; x.t/ = Ê Ψ .x t/ x Ψ(x t/ dx = Ê Ψ .x t/ x Ψ(x t/ dx = Ê x Ψ .x t)Ψ(x t/ dx −∞ −∞ −∞ «­­­­­¯­­­­­¬ p.x/ P. J. Grandinetti Chapter 11: Wave Mechanics Position Operators, x̂ Other mean or expectation value quantities, such as êx2.t/ë, are written ∞ ∞ ∞ ê 2 ë < ; ̂2 ; < ; 2 ; 2 < ; ; x .t/ = Ê Ψ .x t/ x Ψ(x t/ dx = Ê Ψ .x t/ x Ψ(x t/ dx = Ê x Ψ .x t/ Ψ(x t/ dx −∞ −∞ −∞ «­­­­­­¯­­­­­­¬ p.x/ For any function of x we write ∞ ∞ ∞ ê ; ë < ; ̂ ; ; < ; ; ; ; < ; ; f .x t/ = Ê Ψ .x t/ f .x t/ Ψ(x t/ dx = Ê Ψ .x t/ f .x t/ Ψ(x t/ dx = Ê f .x t/ Ψ .x t/ Ψ(x t/ dx −∞ −∞ −∞ «­­­­­­¯­­­­­­¬ p.x/ As long as the operator f̂.x; t/ depends only on x and t, then f̂.x; t/Ψ(x; t/ = f .x; t)Ψ(x; t/ What’s the point of introducing these extra steps with x̂? P. J. Grandinetti Chapter 11: Wave Mechanics Momentum Operators, p̂ What about mean or expectation value of particle momentum? Using wave function in momentum basis Φ(p; t/ we calculate ∞ ∞ ∞ ê ë < ; ̂ ; < ; ; < ; ; p.t/ = Ê Φ .p t/ p Φ(p t/ dp = Ê Φ .p t/ p Φ(p t/ dp = Ê p Φ .p t/ Φ(p t/ dp −∞ −∞ −∞ «­­­­­­¯­­­­­­¬ p.p/ Here p̂ operates on Φ(p; t/ to give p̂ Φ(p; t/ = pΦ(p; t/ Similarly we have ∞ ê 2 ë < ; ̂ 2 ; p .t/ = Ê Φ .p t/ p Φ(p t/dp −∞ and ∞ ê ; ë < ; ̂ ; ; f .p t/ = Ê Φ .p t/ f .p t/ Φ(p t/dp −∞ Again, what’s the point of introducing these extra steps with p̂? Let’s consider this further. P. J. Grandinetti Chapter 11: Wave Mechanics Momentum Operators If p̂ operates on Ψ(x; t/ we should get same result as p̂ acting on Φ(p; t/ ∞ ê ë < ; ̂ ; ; ̂ ; p.t/ = Ê Ψ .x t/ p Ψ(x t/dx but what is p Ψ(x t/ =?? −∞ Use Fourier transform to expand Ψ(x; t/ in momentum representation ∞ ̂ ; ù1 ̂ ; ipx_` p Ψ(x t/ = Ê p Φ(p t/e dp 2휋` −∞ Take derivative of Fourier expansion for Ψ(x; t/ ∞ ; ù1 ; ipx_` if Ψ(x t/ = Ê Φ(p t/e dp then 2휋` −∞ 0 1 ( ) ( ) ) ∞ ip ∞ ; ù1 ; ipx_` i ù1 ; ipx_` i ̂ ; ) Ψ(x t/ = Ê ` Φ(p t/e dp = ` Ê pΦ(p t/e dp = ` p Ψ(x t/ x 2휋` −∞ 2휋` −∞ ( ) ) i Ψ(x; t/ = p̂ Ψ(x; t/ )x ` P. J. Grandinetti Chapter 11: Wave Mechanics Momentum Operators Rearrange and find that p̂Ψ(x; t/ is given by ) p̂ Ψ(x; t/ = *i` Ψ(x; t/ )x Identify momentum operator when applied to wave function in position basis as ) p̂ = *i` )x Operator notation allows us to work with Ψ(x; t/ basis only and obtain expectation values for both x̂ and p̂. Homework Determine position operator when applied to wave function in momentum basis. P. J. Grandinetti Chapter 11: Wave Mechanics Eigenfunctions and Eigenvalues Whenever operator acts on wave function and gives back numerical value times original wavefunction then that wave function is an eigenfunction for that operator and numerical value returned is called the eigenvalue. Af̂ .§/ = a f .§/ f .§/ is eigenfunction of  and a is eigenvalue. Is Ψ(x; t/ eigenfunction for x̂? x̂Ψ(x; t/ = xΨ(x; t/, Yes Is Φ(p; t/ eigenfunction for p̂? p̂Φ(p; t/ = pΦ(p; t/, Yes dΨ(x; t/ Is Ψ(x; t/ eigenfunction for p̂? p̂Ψ(x; t/ = *i` , No dx Is Φ(p; t/ eigenfunction for x̂? No. Prove this at home. P. J. Grandinetti Chapter 11: Wave Mechanics Operators in Quantum Mechanics must be linear Can only lead to linear differential equations involving Ψ ( ) ̂ ̂ ̂ : A c1Ψ1 + c2Ψ2 = c1AΨ1 + c2AΨ2 Nonlinear operations, such as take square root log cosine sine raise to a power–other than 0 or 1 are not allowed. P. J. Grandinetti Chapter 11: Wave Mechanics Uncertainty and Commutator Relations P. J. Grandinetti Chapter 11: Wave Mechanics Commutation Relations Operators may not commute, i.e., order that operators are applied matters. Example What is result when operator below is applied to a wave function? x̂p̂ * p̂x̂ = ? Hint: not zero ) Recall p̂ = *i` . Apply this difference operator to Ψ(x; t/ we find )x 0 1 ( ) )Ψ(x; t/ ) .x̂p̂ * p̂x̂/Ψ(x; t/ = x̂p̂Ψ(x; t/* p̂x̂Ψ(x; t/ = *i` x * *i` .x Ψ(x; t// )x )x 0 1 )Ψ(x; t/ )Ψ(x; t/ = *i`x * *i`Ψ(x; t/* i`x = i`Ψ(x; t/ )x )x .x̂p̂ * p̂x̂/Ψ(x; t/ = i`Ψ(x; t/ P. J. Grandinetti Chapter 11: Wave Mechanics Commutation Relations .x̂p̂ * p̂x̂/ Ψ(x; t/ = i`Ψ(x; t/ Such operator product differences occur often in quantum mechanics. So often it is called a commutator and given shorthand notation x̂p̂ * p̂x̂ = [x̂;p ̂] = i`; for all Ψ(x; t/ When [ ; B̂ ] =  B̂ * B̂  ≠ 0 we say that  and B̂ do not commute. Why are commutation relations useful? P. J. Grandinetti Chapter 11: Wave Mechanics Uncertainty and Commutator Relations From [x̂;p ̂] = i` we obtain the uncertainty relation: êΔxëêΔpë g `_2 Commutator gives us more general uncertainty relationships. Given [ ; B̂ ] = iC; then êΔAëêΔBë g êCë_2 where ù êΔAë = êA2ë * êAë2 This general relation, êΔAëêΔBë g êCë_2, quantifies ability to specify precisely and simultaneously two observables for  and B̂ . P. J. Grandinetti Chapter 11: Wave Mechanics Kinetic and Potential Energy Operators P. J. Grandinetti Chapter 11: Wave Mechanics Kinetic and Potential Energy Operators Kinetic energy operator in momentum and position bases are p̂ 2 `2 )2 K̂ = and K̂ = * 2m 2m )x2 Potential energy operator, V̂ .x/, depend on system under study. Solving Schrödinger Eq. is easy or impossible depending on V̂ .x/. Examples of some easy ones are... Í Electron trapped in 1D box of length L has potential energy operator V̂ .x/ = 0 if 0 f x f L; V̂ .x/ = ∞ if x < 0 and x < L This is called the infinite well potential. Í 휅 1D harmonic oscillator with force constant f has potential energy operator 1 V̂ .x/ = 휅 x̂2 2 f P. J. Grandinetti Chapter 11: Wave Mechanics Total Energy Operator ≡ Hamiltonian Operator Total energy operator is sum of kinetic and potential energy operators, p̂2 `2 )2 Ĥ = K̂ + V̂ = + V̂ .x/ = * + V̂ .x/ 2m 2m )x2 Ĥ is called Hamiltonian operator. Recall Schrödinger equation 4 5 `2 )2 ) E Ψ(x; t/ = * + V̂ .x/ Ψ(x; t/ = i` Ψ(x; t/ 2m )x2 )t «­­­­­­­­­­¯­­­­­­­­­­¬ Ĥ We can write Schrödinger equation more compactly as ) E Ψ(x; t/ = Ĥ Ψ(x; t/ = i` Ψ(x; t/ )t With this substitution we see that Ĥ is also given by ) Ĥ = i` )t P. J. Grandinetti Chapter 11: Wave Mechanics Solving the Schrödinger Equation P. J. Grandinetti Chapter 11: Wave Mechanics Solving the Schrödinger Equation If potential energy operator, V̂ .x/, depends only on position and not time then separation of variables can be used Ψ(x; t/ = .x/휙.t/ Substituting into Schrödinger equation gives 4 5 `2 )2 `2 )2 .x/ )Ψ(x; t/ )휙.t/ * + V̂ .x/ .x/휙.t/ = * 휙.t/ + 휙.t/V̂ .x/ .x/ = i` = i` .x/ 2m )x2 2m )x2 )t )t Dividing both sides by Ψ(x; t/ = .x/휙.t/ leads to 4 5 1 `2 )2 .x/ 1 )휙.t/ * + V̂ .x/ .x/ = i` = E .x/ 2m )x2 휙.t/ )t «­­­­­­­­­­­­­­­­­­­­­­¯­­­­­­­­­­­­­­­­­­­­­­¬ «­­­­¯­­­­¬ depends only on x depends on on t For equality to remain true for all x and t both sides must equal separation constant, E, which we’ll find is total energy and is time independent. P. J. Grandinetti Chapter 11: Wave Mechanics Solving the Schrödinger Equation This gives 2 uncoupled ODEs: d휙.t/ iE d2 .x/ 2m ( ) + 휙.t/ = 0 and + E * V̂ .x/ .x/ = 0 dt ` dx2 `2 ODE for 휙.t/ has trivial solution: 휙.t/ = e*iEt_` ODE for .x/ is called the time independent Schrödinger equation.
Recommended publications
  • 1 Notation for States
    The S-Matrix1 D. E. Soper2 University of Oregon Physics 634, Advanced Quantum Mechanics November 2000 1 Notation for states In these notes we discuss scattering nonrelativistic quantum mechanics. We will use states with the nonrelativistic normalization h~p |~ki = (2π)3δ(~p − ~k). (1) Recall that in a relativistic theory there is an extra factor of 2E on the right hand side of this relation, where E = [~k2 + m2]1/2. We will use states in the “Heisenberg picture,” in which states |ψ(t)i do not depend on time. Often in quantum mechanics one uses the Schr¨odinger picture, with time dependent states |ψ(t)iS. The relation between these is −iHt |ψ(t)iS = e |ψi. (2) Thus these are the same at time zero, and the Schrdinger states obey d i |ψ(t)i = H |ψ(t)i (3) dt S S In the Heisenberg picture, the states do not depend on time but the op- erators do depend on time. A Heisenberg operator O(t) is related to the corresponding Schr¨odinger operator OS by iHt −iHt O(t) = e OS e (4) Thus hψ|O(t)|ψi = Shψ(t)|OS|ψ(t)iS. (5) The Heisenberg picture is favored over the Schr¨odinger picture in the case of relativistic quantum mechanics: we don’t have to say which reference 1Copyright, 2000, D. E. Soper [email protected] 1 frame we use to define t in |ψ(t)iS. For operators, we can deal with local operators like, for instance, the electric field F µν(~x, t).
    [Show full text]
  • Math 353 Lecture Notes Intro to Pdes Eigenfunction Expansions for Ibvps
    Math 353 Lecture Notes Intro to PDEs Eigenfunction expansions for IBVPs J. Wong (Fall 2020) Topics covered • Eigenfunction expansions for PDEs ◦ The procedure for time-dependent problems ◦ Projection, independent evolution of modes 1 The eigenfunction method to solve PDEs We are now ready to demonstrate how to use the components derived thus far to solve the heat equation. First, two examples to illustrate the proces... 1.1 Example 1: no source; Dirichlet BCs The simplest case. We solve ut =uxx; x 2 (0; 1); t > 0 (1.1a) u(0; t) = 0; u(1; t) = 0; (1.1b) u(x; 0) = f(x): (1.1c) The eigenvalues/eigenfunctions are (as calculated in previous sections) 2 2 λn = n π ; φn = sin nπx; n ≥ 1: (1.2) Assuming the solution exists, it can be written in the eigenfunction basis as 1 X u(x; t) = cn(t)φn(x): n=0 Definition (modes) The n-th term of this series is sometimes called the n-th mode or Fourier mode. I'll use the word frequently to describe it (rather than, say, `basis function'). 1 00 Substitute into the PDE (1.1a) and use the fact that −φn = λnφ to obtain 1 X 0 (cn(t) + λncn(t))φn(x) = 0: n=1 By the fact the fφng is a basis, it follows that the coefficient for each mode satisfies the ODE 0 cn(t) + λncn(t) = 0: Solving the ODE gives us a `general' solution to the PDE with its BCs, 1 X −λnt u(x; t) = ane φn(x): n=1 The remaining coefficients are determined by the IC, u(x; 0) = f(x): To match to the solution, we need to also write f(x) in the basis: 1 Z 1 X hf; φni f(x) = f φ (x); f = = 2 f(x) sin nπx dx: (1.3) n n n hφ ; φ i n=1 n n 0 Then from the initial condition, we get u(x; 0) = f(x) 1 1 X X =) cn(0)φn(x) = fnφn(x) n=1 n=1 =) cn(0) = fn for all n ≥ 1: Now everything has been solved - we are done! The solution to the IBVP (1.1) is 1 X −n2π2t u(x; t) = ane sin nπx with an given by (1.5): (1.4) n=1 Alternatively, we could state the solution as follows: The solution is 1 X −λnt u(x; t) = fne φn(x) n=1 with eigenfunctions/values φn; λn given by (1.2) and fn by (1.3).
    [Show full text]
  • Math 356 Lecture Notes Intro to Pdes Eigenfunction Expansions for Ibvps
    MATH 356 LECTURE NOTES INTRO TO PDES EIGENFUNCTION EXPANSIONS FOR IBVPS J. WONG (FALL 2019) Topics covered • Eigenfunction expansions for PDEs ◦ The procedure for time-dependent problems ◦ Projection, independent evolution of modes ◦ Separation of variables • The wave equation ◦ Standard solution, standing waves ◦ Example: tuning the vibrating string (resonance) • Interpreting solutions (what does the series mean?) ◦ Relevance of Fourier convergence theorem ◦ Smoothing for the heat equation ◦ Non-smoothing for the wave equation 3 1. The eigenfunction method to solve PDEs We are now ready to demonstrate how to use the components derived thus far to solve the heat equation. The procedure here for the heat equation will extend nicely to a variety of other problems. For now, consider an initial boundary value problem of the form ut = −Lu + h(x; t); x 2 (a; b); t > 0 hom. BCs at a and b (1.1) u(x; 0) = f(x) We seek a solution in terms of the eigenfunction basis X u(x; t) = cn(t)φn(x) n by finding simple ODEs to solve for the coefficients cn(t): This form of the solution is called an eigenfunction expansion for u (or `eigenfunction series') and each term cnφn(x) is a mode (or `Fourier mode' or `eigenmode'). Part 1: find the eigenfunction basis. The first step is to compute the basis. The eigenfunctions we need are the solutions to the eigenvalue problem Lφ = λφ, φ(x) satisfies the BCs for u: (1.2) By the theorem in ??, there is a sequence of eigenfunctions fφng with eigenvalues fλng that form an orthogonal basis for L2[a; b] (i.e.
    [Show full text]
  • Rotation and Spin and Position Operators in Relativistic Gravity and Quantum Electrodynamics
    Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 4 December 2019 doi:10.20944/preprints201912.0044.v1 Peer-reviewed version available at Universe 2020, 6, 24; doi:10.3390/universe6020024 1 Review 2 Rotation and Spin and Position Operators in 3 Relativistic Gravity and Quantum Electrodynamics 4 R.F. O’Connell 5 Department of Physics and Astronomy, Louisiana State University; Baton Rouge, LA 70803-4001, USA 6 Correspondence: [email protected]; 225-578-6848 7 Abstract: First, we examine how spin is treated in special relativity and the necessity of introducing 8 spin supplementary conditions (SSC) and how they are related to the choice of a center-of-mass of a 9 spinning particle. Next, we discuss quantum electrodynamics and the Foldy-Wouthuysen 10 transformation which we note is a position operator identical to the Pryce-Newton-Wigner position 11 operator. The classical version of the operators are shown to be essential for the treatment of 12 classical relativistic particles in general relativity, of special interest being the case of binary 13 systems (black holes/neutron stars) which emit gravitational radiation. 14 Keywords: rotation; spin; position operators 15 16 I.Introduction 17 Rotation effects in relativistic systems involve many new concepts not needed in 18 non-relativistic classical physics. Some of these are quantum mechanical (where the emphasis is on 19 “spin”). Thus, our emphasis will be on quantum electrodynamics (QED) and both special and 20 general relativity. Also, as in [1] , we often use “spin” in the generic sense of meaning “internal” spin 21 in the case of an elementary particle and “rotation” in the case of an elementary particle.
    [Show full text]
  • Motion of the Reduced Density Operator
    Motion of the Reduced Density Operator Nicholas Wheeler, Reed College Physics Department Spring 2009 Introduction. Quantum mechanical decoherence, dissipation and measurements all involve the interaction of the system of interest with an environmental system (reservoir, measurement device) that is typically assumed to possess a great many degrees of freedom (while the system of interest is typically assumed to possess relatively few degrees of freedom). The state of the composite system is described by a density operator ρ which in the absence of system-bath interaction we would denote ρs ρe, though in the cases of primary interest that notation becomes unavailable,⊗ since in those cases the states of the system and its environment are entangled. The observable properties of the system are latent then in the reduced density operator ρs = tre ρ (1) which is produced by “tracing out” the environmental component of ρ. Concerning the specific meaning of (1). Let n) be an orthonormal basis | in the state space H of the (open) system, and N) be an orthonormal basis s !| " in the state space H of the (also open) environment. Then n) N) comprise e ! " an orthonormal basis in the state space H = H H of the |(closed)⊗| composite s e ! " system. We are in position now to write ⊗ tr ρ I (N ρ I N) e ≡ s ⊗ | s ⊗ | # ! " ! " ↓ = ρ tr ρ in separable cases s · e The dynamics of the composite system is generated by Hamiltonian of the form H = H s + H e + H i 2 Motion of the reduced density operator where H = h I s s ⊗ e = m h n m N n N $ | s| % | % ⊗ | % · $ | ⊗ $ | m,n N # # $% & % &' H = I h e s ⊗ e = n M n N M h N | % ⊗ | % · $ | ⊗ $ | $ | e| % n M,N # # $% & % &' H = m M m M H n N n N i | % ⊗ | % $ | ⊗ $ | i | % ⊗ | % $ | ⊗ $ | m,n M,N # # % &$% & % &'% & —all components of which we will assume to be time-independent.
    [Show full text]
  • An S-Matrix for Massless Particles Arxiv:1911.06821V2 [Hep-Th]
    An S-Matrix for Massless Particles Holmfridur Hannesdottir and Matthew D. Schwartz Department of Physics, Harvard University, Cambridge, MA 02138, USA Abstract The traditional S-matrix does not exist for theories with massless particles, such as quantum electrodynamics. The difficulty in isolating asymptotic states manifests itself as infrared divergences at each order in perturbation theory. Building on insights from the literature on coherent states and factorization, we construct an S-matrix that is free of singularities order-by-order in perturbation theory. Factorization guarantees that the asymptotic evolution in gauge theories is universal, i.e. independent of the hard process. Although the hard S-matrix element is computed between well-defined few particle Fock states, dressed/coherent states can be seen to form as intermediate states in the calculation of hard S-matrix elements. We present a framework for the perturbative calculation of hard S-matrix elements combining Lorentz-covariant Feyn- man rules for the dressed-state scattering with time-ordered perturbation theory for the asymptotic evolution. With hard cutoffs on the asymptotic Hamiltonian, the cancella- tion of divergences can be seen explicitly. In dimensional regularization, where the hard cutoffs are replaced by a renormalization scale, the contribution from the asymptotic evolution produces scaleless integrals that vanish. A number of illustrative examples are given in QED, QCD, and N = 4 super Yang-Mills theory. arXiv:1911.06821v2 [hep-th] 24 Aug 2020 Contents 1 Introduction1 2 The hard S-matrix6 2.1 SH and dressed states . .9 2.2 Computing observables using SH ........................... 11 2.3 Soft Wilson lines . 14 3 Computing the hard S-matrix 17 3.1 Asymptotic region Feynman rules .
    [Show full text]
  • 5 the Dirac Equation and Spinors
    5 The Dirac Equation and Spinors In this section we develop the appropriate wavefunctions for fundamental fermions and bosons. 5.1 Notation Review The three dimension differential operator is : ∂ ∂ ∂ = , , (5.1) ∂x ∂y ∂z We can generalise this to four dimensions ∂µ: 1 ∂ ∂ ∂ ∂ ∂ = , , , (5.2) µ c ∂t ∂x ∂y ∂z 5.2 The Schr¨odinger Equation First consider a classical non-relativistic particle of mass m in a potential U. The energy-momentum relationship is: p2 E = + U (5.3) 2m we can substitute the differential operators: ∂ Eˆ i pˆ i (5.4) → ∂t →− to obtain the non-relativistic Schr¨odinger Equation (with = 1): ∂ψ 1 i = 2 + U ψ (5.5) ∂t −2m For U = 0, the free particle solutions are: iEt ψ(x, t) e− ψ(x) (5.6) ∝ and the probability density ρ and current j are given by: 2 i ρ = ψ(x) j = ψ∗ ψ ψ ψ∗ (5.7) | | −2m − with conservation of probability giving the continuity equation: ∂ρ + j =0, (5.8) ∂t · Or in Covariant notation: µ µ ∂µj = 0 with j =(ρ,j) (5.9) The Schr¨odinger equation is 1st order in ∂/∂t but second order in ∂/∂x. However, as we are going to be dealing with relativistic particles, space and time should be treated equally. 25 5.3 The Klein-Gordon Equation For a relativistic particle the energy-momentum relationship is: p p = p pµ = E2 p 2 = m2 (5.10) · µ − | | Substituting the equation (5.4), leads to the relativistic Klein-Gordon equation: ∂2 + 2 ψ = m2ψ (5.11) −∂t2 The free particle solutions are plane waves: ip x i(Et p x) ψ e− · = e− − · (5.12) ∝ The Klein-Gordon equation successfully describes spin 0 particles in relativistic quan- tum field theory.
    [Show full text]
  • Chapter 5 ANGULAR MOMENTUM and ROTATIONS
    Chapter 5 ANGULAR MOMENTUM AND ROTATIONS In classical mechanics the total angular momentum L~ of an isolated system about any …xed point is conserved. The existence of a conserved vector L~ associated with such a system is itself a consequence of the fact that the associated Hamiltonian (or Lagrangian) is invariant under rotations, i.e., if the coordinates and momenta of the entire system are rotated “rigidly” about some point, the energy of the system is unchanged and, more importantly, is the same function of the dynamical variables as it was before the rotation. Such a circumstance would not apply, e.g., to a system lying in an externally imposed gravitational …eld pointing in some speci…c direction. Thus, the invariance of an isolated system under rotations ultimately arises from the fact that, in the absence of external …elds of this sort, space is isotropic; it behaves the same way in all directions. Not surprisingly, therefore, in quantum mechanics the individual Cartesian com- ponents Li of the total angular momentum operator L~ of an isolated system are also constants of the motion. The di¤erent components of L~ are not, however, compatible quantum observables. Indeed, as we will see the operators representing the components of angular momentum along di¤erent directions do not generally commute with one an- other. Thus, the vector operator L~ is not, strictly speaking, an observable, since it does not have a complete basis of eigenstates (which would have to be simultaneous eigenstates of all of its non-commuting components). This lack of commutivity often seems, at …rst encounter, as somewhat of a nuisance but, in fact, it intimately re‡ects the underlying structure of the three dimensional space in which we are immersed, and has its source in the fact that rotations in three dimensions about di¤erent axes do not commute with one another.
    [Show full text]
  • Sturm-Liouville Expansions of the Delta
    Gauge Institute Journal H. Vic Dannon Sturm-Liouville Expansions of the Delta Function H. Vic Dannon [email protected] May, 2014 Abstract We expand the Delta Function in Series, and Integrals of Sturm-Liouville Eigen-functions. Keywords: Sturm-Liouville Expansions, Infinitesimal, Infinite- Hyper-Real, Hyper-Real, infinite Hyper-real, Infinitesimal Calculus, Delta Function, Fourier Series, Laguerre Polynomials, Legendre Functions, Bessel Functions, Delta Function, 2000 Mathematics Subject Classification 26E35; 26E30; 26E15; 26E20; 26A06; 26A12; 03E10; 03E55; 03E17; 03H15; 46S20; 97I40; 97I30. 1 Gauge Institute Journal H. Vic Dannon Contents 0. Eigen-functions Expansion of the Delta Function 1. Hyper-real line. 2. Hyper-real Function 3. Integral of a Hyper-real Function 4. Delta Function 5. Convergent Series 6. Hyper-real Sturm-Liouville Problem 7. Delta Expansion in Non-Normalized Eigen-Functions 8. Fourier-Sine Expansion of Delta Associated with yx"( )+=l yx ( ) 0 & ab==0 9. Fourier-Cosine Expansion of Delta Associated with 1 yx"( )+=l yx ( ) 0 & ab==2 p 10. Fourier-Sine Expansion of Delta Associated with yx"( )+=l yx ( ) 0 & a = 0 11. Fourier-Bessel Expansion of Delta Associated with n2 -1 ux"( )+- (l 4 ) ux ( ) = 0 & ab==0 x 2 12. Delta Expansion in Orthonormal Eigen-functions 13. Fourier-Hermit Expansion of Delta Associated with ux"( )+- (l xux2 ) ( ) = 0 for any real x 14. Fourier-Legendre Expansion of Delta Associated with 2 Gauge Institute Journal H. Vic Dannon 11 11 uu"(ql )++ [ ] ( q ) = 0 -<<pq p 4 cos2 q , 22 15. Fourier-Legendre Expansion of Delta Associated with 112 11 ymy"(ql )++- [ ( ) ] (q ) = 0 -<<pq p 4 cos2 q , 22 16.
    [Show full text]
  • Lecture 15.Pdf
    Lecture #15 Eigenfunctions of hermitian operators If the spectrum is discrete , then the eigenfunctions are in Hilbert space and correspond to physically realizable states. It the spectrum is continuous , the eigenfunctions are not normalizable, and they do not correspond to physically realizable states (but their linear combinations may be normalizable). Discrete spectra Normalizable eigenfunctions of a hermitian operator have the following properties: (1) Their eigenvalues are real. (2) Eigenfunctions that belong to different eigenvalues are orthogonal. Finite-dimension vector space: The eigenfunctions of an observable operator are complete , i.e. any function in Hilbert space can be expressed as their linear combination. Continuous spectra Example: Find the eigenvalues and eigenfunctions of the momentum operator . This solution is not square -inferable and operator p has no eigenfunctions in Hilbert space. However, if we only consider real eigenvalues, we can define sort of orthonormality: Lecture 15 Page 1 L15.P2 since the Fourier transform of Dirac delta function is Proof: Plancherel's theorem Now, Then, which looks very similar to orthonormality. We can call such equation Dirac orthonormality. These functions are complete in a sense that any square integrable function can be written in a form and c(p) is obtained by Fourier's trick. Summary for continuous spectra: eigenfunctions with real eigenvalues are Dirac orthonormalizable and complete. Lecture 15 Page 2 L15.P3 Generalized statistical interpretation: If your measure observable Q on a particle in a state you will get one of the eigenvalues of the hermitian operator Q. If the spectrum of Q is discrete, the probability of getting the eigenvalue associated with orthonormalized eigenfunction is It the spectrum is continuous, with real eigenvalues q(z) and associated Dirac-orthonormalized eigenfunctions , the probability of getting a result in the range dz is The wave function "collapses" to the corresponding eigenstate upon measurement.
    [Show full text]
  • Positive Eigenfunctions of Markovian Pricing Operators
    Positive Eigenfunctions of Markovian Pricing Operators: Hansen-Scheinkman Factorization, Ross Recovery and Long-Term Pricing Likuan Qin† and Vadim Linetsky‡ Department of Industrial Engineering and Management Sciences McCormick School of Engineering and Applied Sciences Northwestern University Abstract This paper develops a spectral theory of Markovian asset pricing models where the under- lying economic uncertainty follows a continuous-time Markov process X with a general state space (Borel right process (BRP)) and the stochastic discount factor (SDF) is a positive semi- martingale multiplicative functional of X. A key result is the uniqueness theorem for a positive eigenfunction of the pricing operator such that X is recurrent under a new probability mea- sure associated with this eigenfunction (recurrent eigenfunction). As economic applications, we prove uniqueness of the Hansen and Scheinkman (2009) factorization of the Markovian SDF corresponding to the recurrent eigenfunction, extend the Recovery Theorem of Ross (2015) from discrete time, finite state irreducible Markov chains to recurrent BRPs, and obtain the long- maturity asymptotics of the pricing operator. When an asset pricing model is specified by given risk-neutral probabilities together with a short rate function of the Markovian state, we give sufficient conditions for existence of a recurrent eigenfunction and provide explicit examples in a number of important financial models, including affine and quadratic diffusion models and an affine model with jumps. These examples show that the recurrence assumption, in addition to fixing uniqueness, rules out unstable economic dynamics, such as the short rate asymptotically arXiv:1411.3075v3 [q-fin.MF] 9 Sep 2015 going to infinity or to a zero lower bound trap without possibility of escaping.
    [Show full text]
  • Hamilton Equations, Commutator, and Energy Conservation †
    quantum reports Article Hamilton Equations, Commutator, and Energy Conservation † Weng Cho Chew 1,* , Aiyin Y. Liu 2 , Carlos Salazar-Lazaro 3 , Dong-Yeop Na 1 and Wei E. I. Sha 4 1 College of Engineering, Purdue University, West Lafayette, IN 47907, USA; [email protected] 2 College of Engineering, University of Illinois at Urbana-Champaign, Urbana, IL 61820, USA; [email protected] 3 Physics Department, University of Illinois at Urbana-Champaign, Urbana, IL 61820, USA; [email protected] 4 College of Information Science and Electronic Engineering, Zhejiang University, Hangzhou 310058, China; [email protected] * Correspondence: [email protected] † Based on the talk presented at the 40th Progress In Electromagnetics Research Symposium (PIERS, Toyama, Japan, 1–4 August 2018). Received: 12 September 2019; Accepted: 3 December 2019; Published: 9 December 2019 Abstract: We show that the classical Hamilton equations of motion can be derived from the energy conservation condition. A similar argument is shown to carry to the quantum formulation of Hamiltonian dynamics. Hence, showing a striking similarity between the quantum formulation and the classical formulation. Furthermore, it is shown that the fundamental commutator can be derived from the Heisenberg equations of motion and the quantum Hamilton equations of motion. Also, that the Heisenberg equations of motion can be derived from the Schrödinger equation for the quantum state, which is the fundamental postulate. These results are shown to have important bearing for deriving the quantum Maxwell’s equations. Keywords: quantum mechanics; commutator relations; Heisenberg picture 1. Introduction In quantum theory, a classical observable, which is modeled by a real scalar variable, is replaced by a quantum operator, which is analogous to an infinite-dimensional matrix operator.
    [Show full text]