The Dirac Equation C4: Particle Physics

Total Page:16

File Type:pdf, Size:1020Kb

The Dirac Equation C4: Particle Physics Malcolm John The Dirac Equation C4: particle physics These notes consider the properties of the Dirac equation introducing gamma-matrices, the Dirac current, helicity and chirality. 1 The Klein-Gordon problem Thus for a free particle with a plane wavefunction, = e i(p·x−Et) ∗ = e−i(p·x−Et) We remember the nonrelativistic energy-momentum relation ˙ − i(p·x−Et) for a free particle of mass, m, is = iEe ˙∗ = iEe−i(p·x−Et) E = p2=2m : (1) the probability density and probability current are deduced, With quantum mechanical operators, ρ = 2E ; j = 2p (9) @ E ! i~ ; p ! −i~r ; (2) @t and together in the covariant from: Eq. 1 becomes an operator equation acting on a complex jµ = 2pµ ; where @ jµ = 0 : (10) wavefunction, (x; t), µ @ ~ The problem is revealed when the plane wavefunction is sub- i + r2 = 0 ; (3) @t 2m stituted into Eq. 5 to give the energy eigenvalues, which is the time-dependent Schrödinger equation in the ab- q E = ± jpj2 − m2 : (11) sence of an external potential. We understand ρ = j j2 as the probability density and j j2d3 x as the probability of finding the In addition to the expected positive energy solution are an particle in elemental volume d3 x. equal number of negative energy solutions; and by Eq. 9 these negative energy solutions have negative probability. For the relativistic particle case, one might consider repeat- ing the quantum mechanical operator substitution for the rela- tivistic energy-momentum relation, 2 The Dirac revelation (1928) E2 = jpj2c2 + m2c4 : (4) This gives the Klein Gordon equation, 2 Paul Dirac sought to overcome the problem of negative proba- 2 @ 2 2 2 4 −~ = −~ c r + mc bility in relativistic quantum mechanics by defining an equation @t2 (5) linear in both time and spatial derivatives, @2 h i or = r2 − m ; @t2 @ E = i = (α · p + βm) p = −ir : @t in natural units, ~ = c = 1. It is of interest to form the con- | {z } (12) tinuity equation by multiplying Eq. 5 by −i ∗ and its complex Hamiltonian conjugate by −i , To show this Dirac equation satisfies Klein Gordon (which is 2 2 2 ∗ ∗ ∗ j j −i ¨ = −i r2 + im just E = p + m ), the Hamiltonian is applied twice, (6) ¨∗ 2 ∗ ∗ 2 −i = −i r + im ; @ 2 − = −iαx@x − iαy@y − iαz@z + βm (13) @t2 and subtracting one from another, @2 h α2@2 α2@2 α2@2 β2m2 (14) 2 = x x + y y + z z + i ∗ ¨ − i ¨∗ = i ∗r2 − i r2 ∗ : (7) @t + (αxαy + αyαx)@x@y − m(αxβ + βαx)@x @ h i Using the product rule a∗a˙ − aa˙∗ = a∗a¨ − aa¨∗ (in reverse), + (α α + α α )@ @ − m(α β + βα )@ @x y z z y y z y y y on both time and space parts independently gives the continu- i +(αzαx + αxαz)@z@x − m(αzβ + βαz)@z ity equation for the wavefunction probability, 2 2 requiring of the terms: αi = β = 1; (αiα j + α jαi) = 0; and @ h ∗ ∗i ∗ ∗ i ˙ − i ˙ + r · −i r + i r = 0 : (α β + βα ) = 0. Or more succinctly with β ≡ α , @t i i 0 | {z } | {z } (8) ρ j αµαν + αναµ = 2δµν : (15) Malcolm John - 2020 This condition cannot be satisfied by numbers, αµ must be ma- current and probability density defined as, trices. The matrices are shown to be traceless by, j = yα ρ = y : (19) y Tr(αµ) = Tr(αναµαν) The probability density, y = j j2 + j j2 + j j2 + j j2 is = − Tr(αyα α ) using Eq. 15 1 2 3 4 ν ν µ always positive, as required. Each four-vector component is = − Tr(αµ) = 0 ; generated from the multiplication of a spinor with its transpose conjugate, sometimes through a matrix. In this sense, a spinor where the first equality is proved by, is sometimes described as a “square-root" of a four-vector. y y Tr(Ai j) = Tr(B jk Bki Ai j) = Tr(BkiAi jB jk) |{z} 3 Intrinsic spin = 1 if B is hermitian. This is appropriate as we require energy to be observable (real eigenvalues) so the Hamiltonian in Eq. 12, and thus αi and β, A satisfying aspect of the Dirac equation is that it requires the are hermitian. From the eigenvalue formula, particles it describes to have an intrinsic property that behaves like angular momentum. The proof start by recalling standard αµv = λv quantum mechanics commutations, vy αyα v = λ∗λ vyv µ µ |{z} |{z} [pi; p j] = 0; [xi; x j] = 0; [pi; x j] = −i~δi j: (20) = 1 jλj2 = 1 ) λ = ±1 : Also the commutation relation between Pauli spin matrices, PN But the traceless condition, Tr(αµ) = i λi = 0, requires the [σ ; σ ] = 2i σ : (21) dimension of αµ to be even. For N=2 Pauli spin matrices, i j i jk k ! ! ! 0 1 0 −i 1 0 We require angular momentum to commute with the Hamil- σ = σ = σ = (16) x 1 0 y i 0 z 0 −1 tonian because angular momentum quantum states are well established to have define states of definite energies. So, satisfy Eq. 15 but we need four such matrices (αx; αy; αz; β) so N = 4 is required. Specifically the set of four hermitian, trace- [H; L] = [α · p + βm; r × p]: (22) less 4 × 4 matrices define the Dirac representation, The βm term commutes trivially, and for the rest, we exemplify ! ! I 0 0 σi the x component, β = αi = (17) 0 −1 σi 0 h i [α · p; Lx] = αx px + αy py + αz pz; ypz − zpy : (23) where σi are the Pauli spin matrices, satisfies Eq. 15. This is not unique, the Weyl representation is also valid, see Sec. 8. The only non-zero parts are terms of the form [xi; pi] = −i~, [H; Lx] = αy pz[py; y] − αz py[pz; z] = −i~(αy pz − αz py) The 4 × 4 structure of the Dirac equation give four solu- = −i~(α ^ p) tions, . To check if this first-order equation indeed resolves x the problems of Klein Gordon, the probability density and cur- ) [H; L] = −i~ α ^ p , 0 ! (24) rent is calculated in a similar manner as in Eq. 6. However, The Dirac equation appears not to commute with angular mo- as this equation has matrices, the Hermitian conjugate is used mentum, which cannot be a correct description of fermion rather than just the complex conjugate. states with definite angular momentum having definite energy. @ @ y i y − (−i) = @t @t y This is solved by connecting to SU(2), which described fun- −iαx@x − iαy@y − iαz@z + mβψ 1 damental spin- 2 objects (spinors), the rotations of which are y y y y − i@x αx + i@y αy + i@z αz + m β : generated by the Pauli spin matrices. Here we have 4 × 4 so try, ! then using the product rule on each term, σ 0 S = 1 ~ ; (25) 2 0 σ @ @ y X y + + yα @ + @ yα = 0 @t @t i i i i (σ are Pauli spin matrices). Setting ~ = 1, S x; S y; S z are: | {z } i | {z } 00 1 1 00 −i 1 01 0 1 B C B C B C @ y y 1 B1 0 C 1 Bi 0 C 1 B0 −1 C + r · α = 0 (18) = B C ; B C ; B C : @t 2 B 0 1C 2 B 0 −iC 2 B 1 0 C B C B C B C @ 1 0A @ i 0 A @ 0 −1A ∂ρ which is of the form = −r · j with the Dirac probability @t The commutation with the Hamiltonian proceeds as above: Malcolm John - 2020 1 identify the non-trivial part and exemplify the x component. mechanical angular momentum with S = 2 ~, [H; S] = [α · p + βm; S] S 2 = S (S + 1) ; |{z} (29) 1 ! ! ! ! S = ~ ; I 0 σ 0 σ I z 2 1 − 1 = 0 : 2 −I σ 0 2 σ 0 −I and it is intrinsic to the Dirac Hamiltonian. = [α · p; S] e.g. ) [H; S x] = [αx px + αy py + αz pz; S x] 4 Gamma matrices = px[αx; S x] + py[αy; S x] + pz[αz; S x] : Dealing with each of the three terms as well as employing the We continue by rearranging the Dirac equation of Eq. 12 and commutation relation for Pauli spin matrices, multiplying by the matrix β, ! ! ! ! 1 0 σx σx 0 1 σx 0 0 σx [αx; S x] = 2 − 2 " # σx 0 0 σx 0 σx σx 0 @ 0 1 0 1 β i + iα · r − βm = 0 0 σ2 0 σ2 1 B xC − 1 B xC @t = 2 @B 2 AC 2 @B 2 AC = 0 : σx 0 σx 0 " # 2 iβ@t + i βαx @x + i βαy @y + i βαz @z − β m = ! ! ! ! # |{z} |{z} |{z} # 1 0 σy σx 0 1 σx 0 0 σy [αy; S x] = 2 − 2 0 1 2 3 σy 0 0 σx 0 σx σy 0 γ γ γ γ I ! ! 1 0 σyσx 1 0 σxσy = 2 − 2 σyσx 0 σxσy 0 to arrive at a manifestly covariant form of the Dirac equation, ! 1 0 [σyσx] = 2 h i [σyσx] 0 µ iγ @µ − mI = 0 (30) ! 1 0 −2iσz = 2 = −iαz : −2iσz 0 where the identity matrix I is shown to make it explicit this is a ! matrix equation, acting on the four-component Dirac spinor. 1 0 [σzσx] [αz; S x] = 2 (by the same reasoning) [σzσx] 0 ! 1 0 +2iσy = 2 = iαy : +2iσy 0 It is important to understand that, though written with a µ super- script, γµ are not fourvectors.
Recommended publications
  • Symmetry Conditions on Dirac Observables
    Proceedings of Institute of Mathematics of NAS of Ukraine 2004, Vol. 50, Part 2, 671–676 Symmetry Conditions on Dirac Observables Heinz Otto CORDES Department of Mathematics, University of California, Berkeley, CA 94720 USA E-mail: [email protected] Using our Dirac invariant algebra we attempt a mathematically and philosophically clean theory of Dirac observables, within the environment of the 4 books [10,9,11,4]. All classical objections to one-particle Dirac theory seem removed, while also some principal objections to von Neumann’s observable theory may be cured. Heisenberg’s uncertainty principle appears improved. There is a clean and mathematically precise pseudodifferential Foldy–Wouthuysen transform, not only for the supersymmeytric case, but also for general (C∞-) potentials. 1 Introduction The free Dirac Hamiltonian H0 = αD+β is a self-adjoint square root of 1−∆ , with the Laplace HS − 1 operator ∆. The free Schr¨odinger√ Hamiltonian 0 =1 2 ∆ (where “1” is the “rest energy” 2 1 mc ) seems related to H0 like 1+x its approximation 1+ 2 x – second partial sum of its Taylor expansion. From that aspect the early discoverers of Quantum Mechanics were lucky that the energies of the bound states of hydrogen (a few eV) are small, compared to the mass energy of an electron (approx. 500 000 eV), because this should make “Schr¨odinger” a good approximation of “Dirac”. But what about the continuous spectrum of both operators – governing scattering theory. Note that today big machines scatter with energies of about 10,000 – compared to the above 1 = mc2. So, from that aspect the precise scattering theory of both Hamiltonians (with potentials added) should give completely different results, and one may have to put ones trust into one of them, because at most one of them can be applicable.
    [Show full text]
  • Relational Quantum Mechanics and Probability M
    Relational Quantum Mechanics and Probability M. Trassinelli To cite this version: M. Trassinelli. Relational Quantum Mechanics and Probability. Foundations of Physics, Springer Verlag, 2018, 10.1007/s10701-018-0207-7. hal-01723999v3 HAL Id: hal-01723999 https://hal.archives-ouvertes.fr/hal-01723999v3 Submitted on 29 Aug 2018 HAL is a multi-disciplinary open access L’archive ouverte pluridisciplinaire HAL, est archive for the deposit and dissemination of sci- destinée au dépôt et à la diffusion de documents entific research documents, whether they are pub- scientifiques de niveau recherche, publiés ou non, lished or not. The documents may come from émanant des établissements d’enseignement et de teaching and research institutions in France or recherche français ou étrangers, des laboratoires abroad, or from public or private research centers. publics ou privés. Noname manuscript No. (will be inserted by the editor) Relational Quantum Mechanics and Probability M. Trassinelli the date of receipt and acceptance should be inserted later Abstract We present a derivation of the third postulate of Relational Quan- tum Mechanics (RQM) from the properties of conditional probabilities. The first two RQM postulates are based on the information that can be extracted from interaction of different systems, and the third postulate defines the prop- erties of the probability function. Here we demonstrate that from a rigorous definition of the conditional probability for the possible outcomes of different measurements, the third postulate is unnecessary and the Born's rule naturally emerges from the first two postulates by applying the Gleason's theorem. We demonstrate in addition that the probability function is uniquely defined for classical and quantum phenomena.
    [Show full text]
  • A Study of Fractional Schrödinger Equation-Composed Via Jumarie Fractional Derivative
    A Study of Fractional Schrödinger Equation-composed via Jumarie fractional derivative Joydip Banerjee1, Uttam Ghosh2a , Susmita Sarkar2b and Shantanu Das3 Uttar Buincha Kajal Hari Primary school, Fulia, Nadia, West Bengal, India email- [email protected] 2Department of Applied Mathematics, University of Calcutta, Kolkata, India; 2aemail : [email protected] 2b email : [email protected] 3 Reactor Control Division BARC Mumbai India email : [email protected] Abstract One of the motivations for using fractional calculus in physical systems is due to fact that many times, in the space and time variables we are dealing which exhibit coarse-grained phenomena, meaning that infinitesimal quantities cannot be placed arbitrarily to zero-rather they are non-zero with a minimum length. Especially when we are dealing in microscopic to mesoscopic level of systems. Meaning if we denote x the point in space andt as point in time; then the differentials dx (and dt ) cannot be taken to limit zero, rather it has spread. A way to take this into account is to use infinitesimal quantities as ()Δx α (and ()Δt α ) with 01<α <, which for very-very small Δx (and Δt ); that is trending towards zero, these ‘fractional’ differentials are greater that Δx (and Δt ). That is()Δx α >Δx . This way defining the differentials-or rather fractional differentials makes us to use fractional derivatives in the study of dynamic systems. In fractional calculus the fractional order trigonometric functions play important role. The Mittag-Leffler function which plays important role in the field of fractional calculus; and the fractional order trigonometric functions are defined using this Mittag-Leffler function.
    [Show full text]
  • Introductory Lectures on Quantum Field Theory
    Introductory Lectures on Quantum Field Theory a b L. Álvarez-Gaumé ∗ and M.A. Vázquez-Mozo † a CERN, Geneva, Switzerland b Universidad de Salamanca, Salamanca, Spain Abstract In these lectures we present a few topics in quantum field theory in detail. Some of them are conceptual and some more practical. They have been se- lected because they appear frequently in current applications to particle physics and string theory. 1 Introduction These notes are based on lectures delivered by L.A.-G. at the 3rd CERN–Latin-American School of High- Energy Physics, Malargüe, Argentina, 27 February–12 March 2005, at the 5th CERN–Latin-American School of High-Energy Physics, Medellín, Colombia, 15–28 March 2009, and at the 6th CERN–Latin- American School of High-Energy Physics, Natal, Brazil, 23 March–5 April 2011. The audience on all three occasions was composed to a large extent of students in experimental high-energy physics with an important minority of theorists. In nearly ten hours it is quite difficult to give a reasonable introduction to a subject as vast as quantum field theory. For this reason the lectures were intended to provide a review of those parts of the subject to be used later by other lecturers. Although a cursory acquaintance with the subject of quantum field theory is helpful, the only requirement to follow the lectures is a working knowledge of quantum mechanics and special relativity. The guiding principle in choosing the topics presented (apart from serving as introductions to later courses) was to present some basic aspects of the theory that present conceptual subtleties.
    [Show full text]
  • Identical Particles
    8.06 Spring 2016 Lecture Notes 4. Identical particles Aram Harrow Last updated: May 19, 2016 Contents 1 Fermions and Bosons 1 1.1 Introduction and two-particle systems .......................... 1 1.2 N particles ......................................... 3 1.3 Non-interacting particles .................................. 5 1.4 Non-zero temperature ................................... 7 1.5 Composite particles .................................... 7 1.6 Emergence of distinguishability .............................. 9 2 Degenerate Fermi gas 10 2.1 Electrons in a box ..................................... 10 2.2 White dwarves ....................................... 12 2.3 Electrons in a periodic potential ............................. 16 3 Charged particles in a magnetic field 21 3.1 The Pauli Hamiltonian ................................... 21 3.2 Landau levels ........................................ 23 3.3 The de Haas-van Alphen effect .............................. 24 3.4 Integer Quantum Hall Effect ............................... 27 3.5 Aharonov-Bohm Effect ................................... 33 1 Fermions and Bosons 1.1 Introduction and two-particle systems Previously we have discussed multiple-particle systems using the tensor-product formalism (cf. Section 1.2 of Chapter 3 of these notes). But this applies only to distinguishable particles. In reality, all known particles are indistinguishable. In the coming lectures, we will explore the mathematical and physical consequences of this. First, consider classical many-particle systems. If a single particle has state described by position and momentum (~r; p~), then the state of N distinguishable particles can be written as (~r1; p~1; ~r2; p~2;:::; ~rN ; p~N ). The notation (·; ·;:::; ·) denotes an ordered list, in which different posi­ tions have different meanings; e.g. in general (~r1; p~1; ~r2; p~2)6 = (~r2; p~2; ~r1; p~1). 1 To describe indistinguishable particles, we can use set notation.
    [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]
  • WEYL SPINORS and DIRAC's ELECTRON EQUATION C William O
    WEYL SPINORS AND DIRAC’SELECTRON EQUATION c William O. Straub, PhD Pasadena, California March 17, 2005 I’ve been planning for some time now to provide a simplified write-up of Weyl’s seminal 1929 paper on gauge invariance. I’m still planning to do it, but since Weyl’s paper covers so much ground I thought I would first address a discovery that he made kind of in passing that (as far as I know) has nothing to do with gauge-invariant gravitation. It involves the mathematical objects known as spinors. Although Weyl did not invent spinors, I believe he was the first to explore them in the context of Dirac’srelativistic electron equation. However, without a doubt Weyl was the first person to investigate the consequences of zero mass in the Dirac equation and the implications this has on parity conservation. In doing so, Weyl unwittingly anticipated the existence of a particle that does not respect the preservation of parity, an unheard-of idea back in 1929 when parity conservation was a sacred cow. Following that, I will use this opportunity to derive the Dirac equation itself and talk a little about its role in particle spin. Those of you who have studied Dirac’s relativistic electron equation may know that the 4-component Dirac spinor is actually composed of two 2-component spinors that Weyl introduced to physics back in 1929. The Weyl spinors have unusual parity properties, and because of this Pauli was initially very critical of Weyl’sanalysis because it postulated massless fermions (neutrinos) that violated the then-cherished notion of parity conservation.
    [Show full text]
  • Two-State Systems
    1 TWO-STATE SYSTEMS Introduction. Relative to some/any discretely indexed orthonormal basis |n) | ∂ | the abstract Schr¨odinger equation H ψ)=i ∂t ψ) can be represented | | | ∂ | (m H n)(n ψ)=i ∂t(m ψ) n ∂ which can be notated Hmnψn = i ∂tψm n H | ∂ | or again ψ = i ∂t ψ We found it to be the fundamental commutation relation [x, p]=i I which forced the matrices/vectors thus encountered to be ∞-dimensional. If we are willing • to live without continuous spectra (therefore without x) • to live without analogs/implications of the fundamental commutator then it becomes possible to contemplate “toy quantum theories” in which all matrices/vectors are finite-dimensional. One loses some physics, it need hardly be said, but surprisingly much of genuine physical interest does survive. And one gains the advantage of sharpened analytical power: “finite-dimensional quantum mechanics” provides a methodological laboratory in which, not infrequently, the essentials of complicated computational procedures can be exposed with closed-form transparency. Finally, the toy theory serves to identify some unanticipated formal links—permitting ideas to flow back and forth— between quantum mechanics and other branches of physics. Here we will carry the technique to the limit: we will look to “2-dimensional quantum mechanics.” The theory preserves the linearity that dominates the full-blown theory, and is of the least-possible size in which it is possible for the effects of non-commutivity to become manifest. 2 Quantum theory of 2-state systems We have seen that quantum mechanics can be portrayed as a theory in which • states are represented by self-adjoint linear operators ρ ; • motion is generated by self-adjoint linear operators H; • measurement devices are represented by self-adjoint linear operators A.
    [Show full text]
  • Chapter 6. Time Evolution in Quantum Mechanics
    6. Time Evolution in Quantum Mechanics 6.1 Time-dependent Schrodinger¨ equation 6.1.1 Solutions to the Schr¨odinger equation 6.1.2 Unitary Evolution 6.2 Evolution of wave-packets 6.3 Evolution of operators and expectation values 6.3.1 Heisenberg Equation 6.3.2 Ehrenfest’s theorem 6.4 Fermi’s Golden Rule Until now we used quantum mechanics to predict properties of atoms and nuclei. Since we were interested mostly in the equilibrium states of nuclei and in their energies, we only needed to look at a time-independent description of quantum-mechanical systems. To describe dynamical processes, such as radiation decays, scattering and nuclear reactions, we need to study how quantum mechanical systems evolve in time. 6.1 Time-dependent Schro¨dinger equation When we first introduced quantum mechanics, we saw that the fourth postulate of QM states that: The evolution of a closed system is unitary (reversible). The evolution is given by the time-dependent Schrodinger¨ equation ∂ ψ iI | ) = ψ ∂t H| ) where is the Hamiltonian of the system (the energy operator) and I is the reduced Planck constant (I = h/H2π with h the Planck constant, allowing conversion from energy to frequency units). We will focus mainly on the Schr¨odinger equation to describe the evolution of a quantum-mechanical system. The statement that the evolution of a closed quantum system is unitary is however more general. It means that the state of a system at a later time t is given by ψ(t) = U(t) ψ(0) , where U(t) is a unitary operator.
    [Show full text]
  • L the Probability Current Or Flux
    Atomic and Molecular Quantum Theory Course Number: C561 L The Probability Current or flux 1. Consider the time-dependent Schr¨odinger Equation and its complex conjugate: 2 ∂ h¯ ∂2 ıh¯ ψ(x, t)= Hψ(x, t)= − 2 + V ψ(x, t) (L.26) ∂t " 2m ∂x # 2 2 ∂ ∗ ∗ h¯ ∂ ∗ −ıh¯ ψ (x, t)= Hψ (x, t)= − 2 + V ψ (x, t) (L.27) ∂t " 2m ∂x # 2. Multiply Eq. (L.26) by ψ∗(x, t), andEq. (L.27) by ψ(x, t): ∂ ψ∗(x, t)ıh¯ ψ(x, t) = ψ∗(x, t)Hψ(x, t) ∂t 2 2 ∗ h¯ ∂ = ψ (x, t) − 2 + V ψ(x, t) (L.28) " 2m ∂x # ∂ −ψ(x, t)ıh¯ ψ∗(x, t) = ψ(x, t)Hψ∗(x, t) ∂t 2 2 h¯ ∂ ∗ = ψ(x, t) − 2 + V ψ (x, t) (L.29) " 2m ∂x # 3. Subtract the two equations to obtain (see that the terms involving V cancels out) 2 2 ∂ ∗ h¯ ∗ ∂ ıh¯ [ψ(x, t)ψ (x, t)] = − ψ (x, t) 2 ψ(x, t) ∂t 2m " ∂x 2 ∂ ∗ − ψ(x, t) 2 ψ (x, t) (L.30) ∂x # or ∂ h¯ ∂ ∂ ∂ ρ(x, t)= − ψ∗(x, t) ψ(x, t) − ψ(x, t) ψ∗(x, t) (L.31) ∂t 2mı ∂x " ∂x ∂x # 4. If we make the variable substitution: h¯ ∂ ∂ J = ψ∗(x, t) ψ(x, t) − ψ(x, t) ψ∗(x, t) 2mı " ∂x ∂x # h¯ ∂ = I ψ∗(x, t) ψ(x, t) (L.32) m " ∂x # (where I [···] represents the imaginary part of the quantity in brackets) we obtain ∂ ∂ ρ(x, t)= − J (L.33) ∂t ∂x or in three-dimensions ∂ ρ(x, t)+ ∇·J =0 (L.34) ∂t Chemistry, Indiana University L-37 c 2003, Srinivasan S.
    [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]
  • On Relational Quantum Mechanics Oscar Acosta University of Texas at El Paso, [email protected]
    University of Texas at El Paso DigitalCommons@UTEP Open Access Theses & Dissertations 2010-01-01 On Relational Quantum Mechanics Oscar Acosta University of Texas at El Paso, [email protected] Follow this and additional works at: https://digitalcommons.utep.edu/open_etd Part of the Philosophy of Science Commons, and the Quantum Physics Commons Recommended Citation Acosta, Oscar, "On Relational Quantum Mechanics" (2010). Open Access Theses & Dissertations. 2621. https://digitalcommons.utep.edu/open_etd/2621 This is brought to you for free and open access by DigitalCommons@UTEP. It has been accepted for inclusion in Open Access Theses & Dissertations by an authorized administrator of DigitalCommons@UTEP. For more information, please contact [email protected]. ON RELATIONAL QUANTUM MECHANICS OSCAR ACOSTA Department of Philosophy Approved: ____________________ Juan Ferret, Ph.D., Chair ____________________ Vladik Kreinovich, Ph.D. ___________________ John McClure, Ph.D. _________________________ Patricia D. Witherspoon Ph. D Dean of the Graduate School Copyright © by Oscar Acosta 2010 ON RELATIONAL QUANTUM MECHANICS by Oscar Acosta THESIS Presented to the Faculty of the Graduate School of The University of Texas at El Paso in Partial Fulfillment of the Requirements for the Degree of MASTER OF ARTS Department of Philosophy THE UNIVERSITY OF TEXAS AT EL PASO MAY 2010 Acknowledgments I would like to express my deep felt gratitude to my advisor and mentor Dr. Ferret for his never-ending patience, his constant motivation and for not giving up on me. I would also like to thank him for introducing me to the subject of philosophy of science and hiring me as his teaching assistant.
    [Show full text]