Quantum Mechanics

Total Page:16

File Type:pdf, Size:1020Kb

Quantum Mechanics Quantum Mechanics Thomas DeGrand July 18, 2021 Quantum Mechanics 2 Contents 1 Introduction 5 2 Quantum mechanics in the language of Hilbert space 11 3 Time dependence in quantum mechanics 47 4 Propagators and path integrals 67 5 Density matrices 83 6 Wave mechanics 97 7 Angular momentum 147 8 Identical particles 185 9 Time independent perturbation theory 195 10 Variational methods 221 11 Time dependent perturbation theory 231 12 Electromagnetic interactions in semiclassical approximation 249 3 Quantum Mechanics 4 13 Scattering 281 14 Classical waves – quantum mechanical particles 335 15 Atoms and molecules 365 Chapter 1 Introduction 5 Quantum Mechanics 6 The subject of most of this book is the quantum mechanics of systems with a small number of degrees of freedom. The book is a mix of descriptions of quantum mechanics itself, of the general properties of systems described by quantum mechanics, and of techniques for describing their behavior. The examples illustrating quantum mechanical properties are often presented as specific physical systems. They will be drawn from many areas of physics, just to illustrate the fact that many apparently different physical systems naturally behave quite similarly. Ehrenfest famously said that “Physics is simple but subtle,” but he did not mean that it was easy. You cannot learn quantum mechanics out of one book, nor out of these notes. You will have to compare the explanations from many sources and integrate them together into your own whole. You should not be surprised to find that any two books are likely to present completely contradictory explanations for the same topic, and that both explanations are at least partially true. The quantum mechanical description of nature is fundamentally different from a classical description, in that it involves probabilistic statements. The usual causal story of classical mechanics is that in specifying a set of initial conditions, one completely specifies the evolu- tion of the system for all time. That is not possible in quantum mechanics, simply because it is not possible to completely specify all the initial conditions. For example, the uncertainty principle ∆p∆x ~ forbids us from simultaneously knowing the coordinates and momentum ≥ of a particle at any time. However, the evolution of the probability itself is causal, and is encoded in the time-dependent Schr¨odinger equation ∂ψ(t) i~ = Hψˆ (t). (1.1) ∂t Once we specify the wave function (probability amplitude) ψ(t) at some time t0, we know it for all later times. This is of course difficult for us, macroscopic beings that we are, to deal with. Once we have gotten over our classical discomfort with the quantum world, we notice several striking features which recur again and again: The wide applicability of the same ideas, or the same physical systems, to many areas • of physics. The presence of symmetries and their consequences for dynamics. • Quantum Mechanics 7 I like to take advantage of the first of these features and think about applications of quantum mechanics in terms of a few paradigmatic systems, which are approximations to nearly all physical systems you might encounter. It is worthwhile to understand these systems completely and in many different ways. The two most important such systems are The simple harmonic oscillator. Most weakly coupled systems behave like a set of • coupled harmonic oscillators. This is easiest to visualize classically, for a system of particles interacting with a potential. If the system has some equilibrium structure, and if the energy of the system is small, then the particles will be found close to their equilibrium locations. This means that the potential energy is close to its minimum. We can do a Taylor expansion about the minimum, and then 1 2 V (x) V (x )+ V ′′(x )(x x ) + . (1.2) ≃ 0 2 0 − 0 which is the potential of an oscillator. The same situation will happen for a quantum mechanical system. An obvious example would be the motion of nuclear degrees of freedom in a molecule, the origin of the “vibrational spectrum” of molecular excitations, typically seen in the infrared. But there are many other less-obvious examples. Classical systems described by a set of equations of motion d2y i = ω2y (1.3) dt2 − i i can be thought of as collections of harmonic oscillators. Classical wave systems have such equations of motion. Their quantum analogues are also oscillators, and so their quantum descriptions will involve oscillator-like energies and degrees of freedom. Do you recall that the quantum mechanical energy spectrum for an oscillator is a set of equally spaced levels, E = nǫ (up to an overall constant) where ǫ is an energy scale, ǫ = ~ωi in Eq. 1.3, for example, and n is an integer 0, 1, 2,....? Do you also recall the story of the photon, that the quantum electromagnetic field is labeled by an integer, the number of photons in a particular allowed state? This integer is the n of the harmonic oscillator. We can take the analogy between “counting” and oscillator states still further. Imagine that we had a system which seemed to have nothing to do with an oscillator, like a hydrogen atom. It has an energy spectrum E = ǫi, where i labels the quantum number of the state. Now imagine that we have a whole collection of hydrogen atoms, and imagine also that these atoms do not interact. The energy of the collection is E = niǫi, where again ni, the number of particles in the collection with P Quantum Mechanics 8 energy ǫi, is an integer. If we forget about where the numbers ǫi came from, we will be making an oscillator-like description of the collection. The two state system. In an undergraduate textbook, the paradigm for this system is • the spin-1/2 particle in an external magnetic field. The electron is either spin-up or spin-down, and energy eigenstates are states where the spin aligns along the magnetic field. But this description is often used in situations not involving spin, typically as a way of approximating some complicated dynamics. For example, it often happens that a probe can excite a system from one state to a single different state, or that the probe can excite the state to many different ones, but that there is only one strong transition, and all other ones are small. Then it makes sense to replace the more complicated system by one which contains only two states, and assume that the system is confined only to those states. These states could be different atomic levels (perhaps interacting with a radiation field), different kinds of elementary particles which can exchange roles (neutrino oscillations are a recent example) or could even be systems which are naively very different (cold gases, which can condense as either molecules or as individual atoms, for another recent example). And of course the “qubit” of quantum computing jargon is a two state system. Of course, there are other paradigmatic systems. Often, we study collections of particles which are weakly interacting. A good approximate starting point for these systems is to ignore the interactions, giving a set of free particles as zeroth order states. This is often how we describe scattering. Hydrogen, despite its appearance in every undergraduate quantum mechanics course, is not so paradigmatic. Hydrogen is not even a very typical atom and its “1/n2” Rydberg spectrum is unique, a consequence of a particular special symmetry. It is useful to know about it, though, because it is simple enough that we can solve for its properties completely. The second feature – physics arising from symmetry – appears again and again in these notes. If a physical system has an underlying symmetry, seemingly different processes can be related. An example is the relative intensity of spectral lines of families of states with the same total angular momentum and different z-components: the pattern of spectral lines for an atom decaying from an excited state, placed in an external magnetic field. Often, without knowing anything about a dynamical system other than the symmetries it obeys, one can determine what processes are forbidden. (People speak of “selection rules” as a shorthand for these consequences; “forbidden” means that the rate is zero.) Such predictions are more Quantum Mechanics 9 reliable than ones of the absolute rates for processes actually occur. Knowing how to use symmetry to make predictions is at the bottom of most of the technical parts of this course. There is a deep connection between symmetry and conservation laws. We usually describe this in quantum mechanics by saying that the Hamiltonian possess a symmetry and the presence of a symmetry of the Hamiltonian directly leads to a conservation law. And most of the time we say “Here is a Hamiltonian, what are the consequences?” This is a perfectly good question, except that it may be putting the cart before the horse. Where do Hamiltonians come from, anyway? I think most people would say that it is the symmetries which are at the bottom; a Hamiltonian is a construct which builds the symmetries into the dynamics from the beginning. And not just the Hamiltonian, but one should use dynamical variables that naturally encode the symmetry, or transform in some simple way under the symmetry. Figuring out an appropriate set of variables for dealing with perhaps the most important symmetry transformation – rotational invariance – will be a big part of the course. Then, the job of the physicist is to recognize some symmetry, write down a mathematical system which encodes it, and then look for the consequence of the symmetry in new places. I’ve wandered away from the course, to try to disturb you. Let me say one more disturbing thing, before we get to work: Typically, physics at different energy scales decouples.
Recommended publications
  • Further Quantum Physics
    Further Quantum Physics Concepts in quantum physics and the structure of hydrogen and helium atoms Prof Andrew Steane January 18, 2005 2 Contents 1 Introduction 7 1.1 Quantum physics and atoms . 7 1.1.1 The role of classical and quantum mechanics . 9 1.2 Atomic physics—some preliminaries . .... 9 1.2.1 Textbooks...................................... 10 2 The 1-dimensional projectile: an example for revision 11 2.1 Classicaltreatment................................. ..... 11 2.2 Quantum treatment . 13 2.2.1 Mainfeatures..................................... 13 2.2.2 Precise quantum analysis . 13 3 Hydrogen 17 3.1 Some semi-classical estimates . 17 3.2 2-body system: reduced mass . 18 3.2.1 Reduced mass in quantum 2-body problem . 19 3.3 Solution of Schr¨odinger equation for hydrogen . ..... 20 3.3.1 General features of the radial solution . 21 3.3.2 Precisesolution.................................. 21 3.3.3 Meanradius...................................... 25 3.3.4 How to remember hydrogen . 25 3.3.5 Mainpoints.................................... 25 3.3.6 Appendix on series solution of hydrogen equation, off syllabus . 26 3 4 CONTENTS 4 Hydrogen-like systems and spectra 27 4.1 Hydrogen-like systems . 27 4.2 Spectroscopy ........................................ 29 4.2.1 Main points for use of grating spectrograph . ...... 29 4.2.2 Resolution...................................... 30 4.2.3 Usefulness of both emission and absorption methods . 30 4.3 The spectrum for hydrogen . 31 5 Introduction to fine structure and spin 33 5.1 Experimental observation of fine structure . ..... 33 5.2 TheDiracresult ..................................... 34 5.3 Schr¨odinger method to account for fine structure . 35 5.4 Physical nature of orbital and spin angular momenta .
    [Show full text]
  • LYCEN 7112 Février 1972 Notice 92Ft on the Generalized Exchange
    LYCEN 7112 Février 1972 notice 92ft On the Generalized Exchange Operators for SU(n) A. Partensky Institut de Physique Nucléaire, Université Claude Bernard de Lyon Institut National de Physique Nucléaire et de Physique des Particules 43, Bd du 11 novembre 1918, 69-Villeurbanne, France Abstract - By following the work of Biedenharn we have redefined the k-particles generalized exchange operators (g.e.o) and studied their properties. By a straightforward but cur-jersome calculation we have derived the expression, in terms of the SU(n) Casimir operators, of the ?., 3 or 4 particles g.e.o. acting on the A-particles states which span an irreducible representation of the grown SU(n). A striking and interesting result is that the eigenvalues of each of these g.e.o. do not depend on the n in SU(n) but only on the Young pattern associated with the irreducible representation considered. For a given g.e.o. , the eigenvalues corresponding to two conjugate Young patterns are the same except for the sign which depends on the parity of the g.e.o. considered. Three Appendices deal with some related problems and, more specifically, Appendix C 2 contains a method of obtaining the eigenvalues of Gel'fand invariants in a new and simple way. 1 . Introduction The interest attributed by physicists to the Lie groups has 3-4 been continuously growing since the pionner works of Weyl , Wigner and Racah • In particular, unitary groups have been used successfully in various domains of physics, although the physical reasons for their introduction have not always been completely explained.
    [Show full text]
  • 5.1 Two-Particle Systems
    5.1 Two-Particle Systems We encountered a two-particle system in dealing with the addition of angular momentum. Let's treat such systems in a more formal way. The w.f. for a two-particle system must depend on the spatial coordinates of both particles as @Ψ well as t: Ψ(r1; r2; t), satisfying i~ @t = HΨ, ~2 2 ~2 2 where H = + V (r1; r2; t), −2m1r1 − 2m2r2 and d3r d3r Ψ(r ; r ; t) 2 = 1. 1 2 j 1 2 j R Iff V is independent of time, then we can separate the time and spatial variables, obtaining Ψ(r1; r2; t) = (r1; r2) exp( iEt=~), − where E is the total energy of the system. Let us now make a very fundamental assumption: that each particle occupies a one-particle e.s. [Note that this is often a poor approximation for the true many-body w.f.] The joint e.f. can then be written as the product of two one-particle e.f.'s: (r1; r2) = a(r1) b(r2). Suppose furthermore that the two particles are indistinguishable. Then, the above w.f. is not really adequate since you can't actually tell whether it's particle 1 in state a or particle 2. This indeterminacy is correctly reflected if we replace the above w.f. by (r ; r ) = a(r ) (r ) (r ) a(r ). 1 2 1 b 2 b 1 2 The `plus-or-minus' sign reflects that there are two distinct ways to accomplish this. Thus we are naturally led to consider two kinds of identical particles, which we have come to call `bosons' (+) and `fermions' ( ).
    [Show full text]
  • Chapter 4 MANY PARTICLE SYSTEMS
    Chapter 4 MANY PARTICLE SYSTEMS The postulates of quantum mechanics outlined in previous chapters include no restrictions as to the kind of systems to which they are intended to apply. Thus, although we have considered numerous examples drawn from the quantum mechanics of a single particle, the postulates themselves are intended to apply to all quantum systems, including those containing more than one and possibly very many particles. Thus, the only real obstacle to our immediate application of the postulates to a system of many (possibly interacting) particles is that we have till now avoided the question of what the linear vector space, the state vector, and the operators of a many-particle quantum mechanical system look like. The construction of such a space turns out to be fairly straightforward, but it involves the forming a certain kind of methematical product of di¤erent linear vector spaces, referred to as a direct or tensor product. Indeed, the basic principle underlying the construction of the state spaces of many-particle quantum mechanical systems can be succinctly stated as follows: The state vector à of a system of N particles is an element of the direct product space j i S(N) = S(1) S(2) S(N) ­ ­ ¢¢¢­ formed from the N single-particle spaces associated with each particle. To understand this principle we need to explore the structure of such direct prod- uct spaces. This exploration forms the focus of the next section, after which we will return to the subject of many particle quantum mechanical systems. 4.1 The Direct Product of Linear Vector Spaces Let S1 and S2 be two independent quantum mechanical state spaces, of dimension N1 and N2, respectively (either or both of which may be in…nite).
    [Show full text]
  • Phase Transitions and the Casimir Effect in Neutron Stars
    University of Tennessee, Knoxville TRACE: Tennessee Research and Creative Exchange Masters Theses Graduate School 12-2017 Phase Transitions and the Casimir Effect in Neutron Stars William Patrick Moffitt University of Tennessee, Knoxville, [email protected] Follow this and additional works at: https://trace.tennessee.edu/utk_gradthes Part of the Other Physics Commons Recommended Citation Moffitt, William Patrick, "Phase Transitions and the Casimir Effect in Neutron Stars. " Master's Thesis, University of Tennessee, 2017. https://trace.tennessee.edu/utk_gradthes/4956 This Thesis is brought to you for free and open access by the Graduate School at TRACE: Tennessee Research and Creative Exchange. It has been accepted for inclusion in Masters Theses by an authorized administrator of TRACE: Tennessee Research and Creative Exchange. For more information, please contact [email protected]. To the Graduate Council: I am submitting herewith a thesis written by William Patrick Moffitt entitled "Phaser T ansitions and the Casimir Effect in Neutron Stars." I have examined the final electronic copy of this thesis for form and content and recommend that it be accepted in partial fulfillment of the requirements for the degree of Master of Science, with a major in Physics. Andrew W. Steiner, Major Professor We have read this thesis and recommend its acceptance: Marianne Breinig, Steve Johnston Accepted for the Council: Dixie L. Thompson Vice Provost and Dean of the Graduate School (Original signatures are on file with official studentecor r ds.) Phase Transitions and the Casimir Effect in Neutron Stars A Thesis Presented for the Master of Science Degree The University of Tennessee, Knoxville William Patrick Moffitt December 2017 Abstract What lies at the core of a neutron star is still a highly debated topic, with both the composition and the physical interactions in question.
    [Show full text]
  • Time-Reversal Symmetry Breaking Abelian Chiral Spin Liquid in Mott Phases of Three-Component Fermions on the Triangular Lattice
    PHYSICAL REVIEW RESEARCH 2, 023098 (2020) Time-reversal symmetry breaking Abelian chiral spin liquid in Mott phases of three-component fermions on the triangular lattice C. Boos,1,2 C. J. Ganahl,3 M. Lajkó ,2 P. Nataf ,4 A. M. Läuchli ,3 K. Penc ,5 K. P. Schmidt,1 and F. Mila2 1Institute for Theoretical Physics, FAU Erlangen-Nürnberg, D-91058 Erlangen, Germany 2Institute of Physics, Ecole Polytechnique Fédérale de Lausanne (EPFL), CH 1015 Lausanne, Switzerland 3Institut für Theoretische Physik, Universität Innsbruck, A-6020 Innsbruck, Austria 4Laboratoire de Physique et Modélisation des Milieux Condensés, Université Grenoble Alpes and CNRS, 25 avenue des Martyrs, 38042 Grenoble, France 5Institute for Solid State Physics and Optics, Wigner Research Centre for Physics, H-1525 Budapest, P.O.B. 49, Hungary (Received 6 December 2019; revised manuscript received 20 March 2020; accepted 20 March 2020; published 29 April 2020) We provide numerical evidence in favor of spontaneous chiral symmetry breaking and the concomitant appearance of an Abelian chiral spin liquid for three-component fermions on the triangular lattice described by an SU(3) symmetric Hubbard model with hopping amplitude −t (t > 0) and on-site interaction U. This chiral phase is stabilized in the Mott phase with one particle per site in the presence of a uniform π flux per plaquette, and in the Mott phase with two particles per site without any flux. Our approach relies on effective spin models derived in the strong-coupling limit in powers of t/U for general SU(N ) and arbitrary uniform charge flux per plaquette, which are subsequently studied using exact diagonalizations and variational Monte Carlo simulations for N = 3, as well as exact diagonalizations of the SU(3) Hubbard model on small clusters.
    [Show full text]
  • 2 Quantum Theory of Spin Waves
    2 Quantum Theory of Spin Waves In Chapter 1, we discussed the angular momenta and magnetic moments of individual atoms and ions. When these atoms or ions are constituents of a solid, it is important to take into consideration the ways in which the angular momenta on different sites interact with one another. For simplicity, we will restrict our attention to the case when the angular momentum on each site is entirely due to spin. The elementary excitations of coupled spin systems in solids are called spin waves. In this chapter, we will introduce the quantum theory of these excita- tions at low temperatures. The two primary interaction mechanisms for spins are magnetic dipole–dipole coupling and a mechanism of quantum mechanical origin referred to as the exchange interaction. The dipolar interactions are of importance when the spin wavelength is very long compared to the spacing between spins, and the exchange interaction dominates when the spacing be- tween spins becomes significant on the scale of a wavelength. In this chapter, we focus on exchange-dominated spin waves, while dipolar spin waves are the primary topic of subsequent chapters. We begin this chapter with a quantum mechanical treatment of a sin- gle electron in a uniform field and follow it with the derivations of Zeeman energy and Larmor precession. We then consider one of the simplest exchange- coupled spin systems, molecular hydrogen. Exchange plays a crucial role in the existence of ordered spin systems. The ground state of H2 is a two-electron exchange-coupled system in an embryonic antiferromagnetic state.
    [Show full text]
  • Total Angular Momentum
    Total Angular momentum Dipanjan Mazumdar Sept 2019 1 Motivation So far, somewhat deliberately, in our prior examples we have worked exclusively with the spin angular momentum and ignored the orbital component. This is acceptable for L = 0 systems such as filled shell atoms, but properties of partially filled atoms will depend both on the orbital and the spin part of the angular momentum. Also, when we include relativistic corrections to atomic Hamiltonian, both spin and orbital angular momentum enter directly through terms like L:~ S~ called the spin-orbit coupling. So instead of focusing only on the spin or orbital angular momentum, we have to develop an understanding as to how they couple together. One of the ways is through the total angular momentum defined as J~ = L~ + S~ (1) We will see that this will be very useful in many cases such as the real hydrogen atom where the eigenstates after including the fine structure effects such as spin-orbit interaction are eigenstates of the total angular momentum operator. 2 Vector model of angular momentum Let us first develop an intuitive understanding of the total angular momentum through the vector model which is a semi-classical approach to \add" angular momenta using vector algebra. We shall first ask, knowing L~ and S~, what are the maxmimum and minimum values of J^. This is simple to answer us- ing the vector model since vectors can be added or subtracted. Therefore, the extreme values are L~ + S~ and L~ − S~ as seen in figure 1.1 Therefore, the total angular momentum will take up values between l+s Figure 1: Maximum and minimum values of angular and jl − sj.
    [Show full text]
  • 1. the Hamiltonian, H = Α Ipi + Βm, Must Be Her- Mitian to Give Real
    1. The hamiltonian, H = αipi + βm, must be her- yields mitian to give real eigenvalues. Thus, H = H† = † † † pi αi +β m. From non-relativistic quantum mechanics † † † αi pi + β m = aipi + βm (2) we know that pi = pi . In addition, [αi,pj ] = 0 because we impose that the α, β operators act on the spinor in- † † αi = αi, β = β. Thus α, β are hermitian operators. dices while pi act on the coordinates of the wave function itself. With these conditions we may proceed: To verify the other properties of the α, and β operators † † piαi + β m = αipi + βm (1) we compute 2 H = (αipi + βm) (αj pj + βm) 2 2 1 2 2 = αi pi + (αiαj + αj αi) pipj + (αiβ + βαi) m + β m , (3) 2 2 where for the second term i = j. Then imposing the Finally, making use of bi = 1 we can find the 26 2 2 2 relativistic energy relation, E = p + m , we obtain eigenvalues: biu = λu and bibiu = λbiu, hence u = λ u the anti-commutation relation | | and λ = 1. ± b ,b =2δ 1, (4) i j ij 2. We need to find the λ = +1/2 helicity eigenspinor { } ′ for an electron with momentum ~p = (p sin θ, 0, p cos θ). where b0 = β, bi = αi, and 1 n n idenitity matrix. ≡ × 1 Using the anti-commutation relation and the fact that The helicity operator is given by 2 σ pˆ, and the posi- 2 1 · bi = 1 we are able to find the trace of any bi tive eigenvalue 2 corresponds to u1 Dirac solution [see (1.5.98) in http://arXiv.org/abs/0906.1271].
    [Show full text]
  • Physics 461 / Quantum Mechanics I
    Physics 461 / Quantum Mechanics I P.E. Parris Department of Physics University of Missouri-Rolla Rolla, Missouri 65409 January 10, 2005 CONTENTS 1 Introduction 7 1.1WhatisQuantumMechanics?......................... 7 1.1.1 WhatisMechanics?.......................... 7 1.1.2 PostulatesofClassicalMechanics:.................. 8 1.2TheDevelopmentofWaveMechanics..................... 10 1.3TheWaveMechanicsofSchrödinger..................... 13 1.3.1 Postulates of Wave Mechanics for a Single Spinless Particle . 13 1.3.2 Schrödinger’sMechanicsforConservativeSystems......... 16 1.3.3 The Principle of Superposition and Spectral Decomposition . 17 1.3.4 TheFreeParticle............................ 19 1.3.5 Superpositions of Plane Waves and the Fourier Transform . 23 1.4Appendix:TheDeltaFunction........................ 26 2 The Formalism of Quantum Mechanics 31 2.1 Postulate I: SpecificationoftheDynamicalState.............. 31 2.1.1 PropertiesofLinearVectorSpaces.................. 31 2.1.2 Additional Definitions......................... 33 2.1.3 ContinuousBasesandContinuousSets................ 34 2.1.4 InnerProducts............................. 35 2.1.5 ExpansionofaVectoronanOrthonormalBasis.......... 38 2.1.6 Calculation of Inner Products Using an Orthonormal Basis . 39 2.1.7 ThePositionRepresentation..................... 40 2.1.8 TheWavevectorRepresentation.................... 41 2.2PostulateII:ObservablesofQuantumMechanicalSystems......... 43 2.2.1 OperatorsandTheirProperties.................... 43 2.2.2 MultiplicativeOperators.......................
    [Show full text]
  • Bound-State Solutions of Dirac Equation Plan of the Lecture
    Lecture 3 Bound-state solutions of Dirac equation Plan of the lecture • Few comments about Dirac equation • Free- and bound-state solutions • Dirac’s spectroscopic notations – Integrals of motion – Parity of states • Energy levels of the bound-state Dirac’s particle • Structure of Dirac’s wavefunction • Radial components of the Dirac’s wavefunction Four-vectors In the relativistic world it is more convenient to work with four-vectors: Contravariant vectors Covariant vectors 휇 푥 = 푡, 푥, 푦, 푧 푥휇 = 푡, −푥, −푦, −푧 휕 휕 휕 휕 휕 휕 휕 휕 휕휇 = , − , − , − 휕 = , , , 휕푡 휕푥 휕푦 휕푧 휇 휕푡 휕푥 휕푦 휕푧 휇 푝 = 퐸, 푝푥, 푝푦, 푝푧 푝휇 = 퐸, −푝푥, −푝푦, −푝푧 Lorentz transformation 푥′휇 = ෍ 푎휈 푥휇 ′ 휈 휇 푥휇 = ෍ 푎휇 푥휇 휈 휈 훾 −훾훽 0 0 훾 훾훽 0 0 −훾훽 훾 훾훽 훾 0 0 푎휈 = 0 0 푎휈 = 휇 0 0 1 0 휇 0 0 1 0 0 0 0 1 0 0 0 1 Klein-Gordon equation Based on the relativistic energy-mass equation: 퐸2 = 푝ҧ2 + 푚2 One can derive Klein-Gordon equation for scalar (zero-spin) relativistic particles: Oscar Klein 휇 2 휕 휕휇 + 푚 휑 푥 = 0 By introducing d'Alembert operator: 휕2 휕휇휕 =⊡= − 휵2 휇 휕푡2 We can re-write Klein-Gordon equation as: ⊡ + 푚2 휑 푥 = 0 Klein-Gordon equation We can derive Klein-Gordon equation for scalar (zero-spin) relativistic particles: 휇 2 휕 휕휇 + 푚 휑 푥 = 0 Free-particle solutions of this equation: Oscar Klein 휑 푥 = 푁 푒−푖 푝푥 = 푁 푒−푖퐸푡+푖풑풓 Allow particles with both positive and negative energy: 퐸 = ± 풑2 + 푚2 And with positive and negative probability density: 푗0 = 2 푁 2 퐸 How do we understand negative-energy solutions? And what is much worse, the negative probability density? Dirac equation We can re-write
    [Show full text]
  • 5.80 Small-Molecule Spectroscopy and Dynamics Fall 2008
    MIT OpenCourseWare http://ocw.mit.edu 5.80 Small-Molecule Spectroscopy and Dynamics Fall 2008 For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms. Lecture #1 Supplement Contents A. Spectroscopic Notation . 1 1. H. N. Russell, A. G. Shenstone, and L. A. Turner, \Report on Notation for Atomic Spectra," 1 2. W. F. Meggers and C. E. Moore, \Report of Subcommittee f (Notation for the Spectra of Diatomic Molecules)" . 2 3. F. A. Jenkins, \Report of Subcommittee f (Notation for the Spectra of Diatomic Molecules)" 2 4. No author, \Report on Notation for the Spectra of Polyatomic Molecules" . 2 B. Good Quantum Numbers . 2 C. Perturbation Theory and Secular Equations . 3 D. Non-Orthonormal Basis Sets . 6 E. Transformation of Matrix Elements of any Operator into Perturbed Basis Set . 7 A. Spectroscopic Notation The language of spectroscopy is very explicit and elegant, capable of describing a wide range of unanticipated situations concisely and unambiguously. The coherence of this language is diligently preserved by a succession of august committees, whose agreements about notation are codified. These agreements are often published as authorless articles in major journals. The following list of citations include the best of these notation-codifying articles. 1. H. N. Russell, A. G. Shenstone, and L. A. Turner, \Report on Notation for Atomic Spectra," Phys. Rev. 33, 900-906 (1929). At an informal meeting of spectroscopists at Washington in April, 1928, the writers of this report were requested to draw up a scheme for the clarification of spectroscopic notation. After much discussion and correspondence with spectroscopists both in this country and abroad we are able to present the following recommendations.
    [Show full text]