4 M Any-Electron Atoms

Total Page:16

File Type:pdf, Size:1020Kb

4 M Any-Electron Atoms pArt i PART II PART III atomic Structure MOLECULAR MOLECULAR STRUCTURE INTERACTIONS 1 classical and quantum mechanics 2 the Schrödinger equation 3 one-electron atoms 4 Many-Electron Atoms many-electron 4 atoms learning objectiVeS goal Why Are We Here? After this chapter, you will be able to Th e principal goal of this chapter is to introduce the Schrödinger equation do the following: for the many-electron atom, together with some new physics related to ➊ Write the Hamiltonian for spin, and show how we can solve that Schrödinger equation by approxi- any atom. mations to the energies and wavefunctions. We need the energies in order ➋ Find a zero-order wavefunction (the electron to predict the energies of chemical bonds; we need the wavefunctions confi guration) and zero-order because the distribution of the electrons will determine the geometry of energy of any atom. the molecules we build. ➌ Use the Hartree-Fock orbital energies to estimate the context Where Are We Now? effective atomic number of Chapter 3 describes—in considerable detail—a two-body problem: fi nding the orbital. the energies and distributions of one electron bound to an atomic nucleus. ➍ Determine the permutation Chemistry is largely determined by where electrons go, and if we’re to get symmetry of a many-particle function. closer to real chemistry (where most electrons come in pairs), we’ll need ➎ Determine the term symbols more than one electron. In this chapter, we add the other electrons. Th at and energy ordering of the may sound simple enough, but the many -body problem cannot be solved term states resulting from any exactly. Much of our eff ort in this chapter must be directed toward fi nding electron confi guration. approximations that help us approach the right answer, while maintaining some intuitive understanding of the structure of the atom. We can also expect new intricacies, both in the electric fi eld interactions as we add more charges and in the magnetic fi eld interactions because each new electron introduces at least one new magnetic moment. But once we have built a comprehensive model of the many-electron atom, we’ll have the entire periodic table at our disposal as we move on to assembling molecules. Supporting text How Did We Get Here? Th e main qualitative concept we draw on for this chapter is that the electron in an atom is suffi ciently wavelike that quantum mechanics restricts its energy and other parameters to specifi c values. For a more 155 M04_COOK4169_SE_01_C04.indd 155 10/29/12 11:54 AM 156 chapter 4 Many-Electron Atoms detailed background, we will draw on the following equations and sections of text to support the ideas developed in this chapter: • Section 2.1 explained how we can use operators, such as a combination of multiplication and differentiation, to extract information from a function. n In particular, the Hamiltonian H operates on a wavefunction c to return the energy E times the wavefunction. We call this the Schrödinger equation (Eq. 2.1): n Hc = E c. • However, the Schrödinger equation works as written only if the wavefunction is an eigenstate of the Hamiltonian. We will find in this chapter that one useful approximation uses a time-independent wavefunction c that is not an n eigenstate of H, and in that case we can use the average value theorem to find an expectation value of the energy (Eq. 2.10): n A = c*(Ac)dt. all Lspace • Section 2.2 introduced the8 particle9 in a one-dimensional box, a fundamental problem in quantum mechanics in which the Schrödinger equation can be solved analytically to give the wavefunctions (Eq. 2.33) 2 npx c (x) sin n = a a and the energies (Eq. 2.31) A a b n2p2U2 En = . 2ma2 • The one-dimensional box depends only on a single Cartesian coordinate. The atom, however, has a potential energy function that can be more naturally described in spherical coordinates, because the Coulomb attraction between the electron and the nucleus depends only on r, the distance between them. As a result, the solutions to the Schrödinger equation are also easier to express in spherical coordinates. To convert between the two coordinate systems, we use Eqs. A.6: x = r sin u cos f r = x2 + y2 + z2 1/2 z y r sin u sin f u a1rccos 2 = = r a yb z r cos u f arc tan . = = x These relationships also allow us to convert the Cartesiana b volume element dx dy dz to the volume element in spherical coordinates, r2dr sin u dudf, which is an essential component of any integrals we may need to carry out. • The Schrödinger equation for the one-electron atom can also be solved analytically, yielding the energies (Eq. 3.30) 2 4 2 Z mee Z En = - 2 2 2 K - 2 Eh. 2(4pe0) n U 2n We also obtain the wavefunctions as products of a radial wavefunction Rn,l (r) ml (given in Table 3.2) and an angular wavefunction Yl (u, f) (given in Table 3.1): ml c u f n,l,ml = Rn, l(r)Y l ( , ). M04_COOK4169_SE_01_C04.indd 156 10/29/12 11:54 AM 4.1 Many-Electron Spatial Wavefunctions 157 • One tool we will turn to for a critical step is the Taylor expansion (Eq. A.29), ϱ n n (x - x0) d f (x) f(x) = a n , n=0 n! dx x0 which lets us rewrite any algebraic expressiona asb a` power series. The trick with power series is to set them up so that each higher-order term is smaller than the last term, so that you can ignore all but the first few terms. In this chapter, we use a Taylor series expansion of the Schrödinger equation itself to find approximate solutions to the energies and wavefunctions. 4.1 Many-Electron Spatial Wavefunctions Neglecting some relatively small effects,1 we can solve the Schrödinger equation of the hydrogen atom to get algebraic expressions for the energies and wavefunc- tions. Once we add a second electron to an atom, that becomes impossible. Although many numerical methods work well, the exact solution cannot be written in general form. We will go as far as we can in the space of one chapter, which is pretty far. Checkpoint as in Chapter 3, Electron Configurations we are color-coding the math related to the Schrödinger equation to help distinguish First, let’s write the Hamiltonian for a two-electron atom as an extension of the terms in the Hamiltonian operator Hamiltonian we used for the one-electron atom: (blue) from contributions to the 2 2 2 2 2 scalar energy (red), and to help n U 2 U 2 Ze Ze e illustrate the transformation of , (H ٌ(1) ٌ(2 = - - - - + (4.1) one to the other. 2me 2me 4pe0r1 4pe0r2 4pe0r12 where 2 2 2 2 0 0 0 2 2 2 1 2 ; ( i) = 2 + 2 + 2 , ri = (xi + yi + zi)ٌ 0xi 0yi 0zi > a b xi, yi, and zi are the Cartesian coordinates for electron i; and r12 is the distance between electrons 1 and 2. These coordinates are defined in Fig. 4.1. The potential z energy terms are now written relative to the energy of the completely ionized atom; 1 in other words, the potential energy is zero when r1, r2, and r12 are all infinite. For –e the more general case of the atom with N electrons, we sum up the kinetic energies of all the electrons and their potential energies of Coulomb attraction to the nucleus. Then, foreach pair of electrons we add the repulsion energy: r N 2 2 i- 1 2 1 n U 2 Ze e r12 (H = a - ٌ(i) - + a . (4.2 i=1 2me 4pe0ri j =1 4pe0rij y c d What’s new and disturbing in the many-electron atom Hamiltonians is the +Ze x 2 electron–electron repulsion term, e (4pe0r12), which prevents the electron 1 r2 –e and electron 2 terms of the Hamiltonian in Eq. 4.1 from being separable. 2 > ▲ FigurE 4.1 Coordinate definitions for the two- 1Magnetic interactions, finite mass of the nucleus, special relativity, to name a few. electron atom. M04_COOK4169_SE_01_C04.indd 157 10/29/12 11:54 AM 158 chapter 4 Many-Electron Atoms example 4.1 Many-Electron Atom Hamiltonians context The element beryllium doubles as one of the most useful and frightening metals available to us. It is the lightest of all the structurally useful metals (neglecting lithium, which is soft and highly reactive), and also among the most hazardous. It has a strength comparable to steel, but at a fraction of the mass. Beryllium-based products were widely produced in the United States and Europe in the decades prior to World War II, but then it became gradually clear that the beryllium in factory dust was sickening and killing workers, sometimes years after exposure. Chronic beryllium disease resembles pneumonia, and can be contracted by repeated exposure to 0.5 mg/m3 or more of beryllium in the air. The mechanism for its biological activity is only partly understood, but it apparently combines with a protein and then deposits itself primarily in lung tissue, triggering an over-reaction by the body’s immune system—a response called hypersensitivity that leads to destructive inflammation of the lungs. The high toxicity has led to severe restrictions on its use in manufacturing, but its properties are still tempting. By accurately simulating the electronic properties of beryllium on a computer, we may identify the most valuable applications of this element at minimal risk.
Recommended publications
  • Applications of the Variational Monte Carlo Method to the Ground States of the Lithium Atom and Its Ions up to Z = 10 in the Presence of Magnetic Field
    Applications of the Variational Monte Carlo Method to the Ground States of the Lithium Atom and its Ions up to Z = 10 in the Presence of Magnetic Field S. B. Doma1), M. O. Shaker2), A. M. Farag2) and F. N. El-Gammal3) 1)Mathematics Department, Faculty of Science, Alexandria University, Egypt E-mail address: [email protected] 2) Mathematics Department, Faculty of Science, Tanta University, Tanta, Egypt 3)Mathematics Department, Faculty of Science, Menofia University, Shebin El-Kom, Egypt Abstract The variational Monte Carlo method is applied to investigate the ground state energy of the lithium atom and its ions up to = 10 in the presence of an external magnetic field regime with = 0 ~ 100 a.u. Our calculations are based on using three forms of compact and accurate trial wave functions, which were put forward in calculating energies in the absence of magnetic field. The obtained results are in good agreement with the most recent accurate values and also with the exact values. Key Words: Atoms in external magnetic field, Variational Monte Carlo method, Lithium atom, Lithium like ions, Total energy. Introduction Over the last decade continuing effort has gone into calculating, with ever increasing accuracy and with various methods, the energies of atoms and ions in neutron star magnetic fields. The motivation comes largely from the fact that features discovered [1–3] in the thermal emission spectra of isolated neutron stars may be due to absorption of photons by heavy atoms in the hot, thin atmospheres of these strongly magnetized cosmic objects [4]. Also, features of heavier elements may be present in the spectra of magnetic white dwarf stars [5, 6].
    [Show full text]
  • Coordination Chemistry III: Electronic Spectra
    160 Chapter 11 Coordination Chemistry III: Electronic Spectra CHAPTER 11: COORDINATION CHEMISTRY III: ELECTRONIC SPECTRA 11.1 a. p3 There are (6!)/(3!3!) = 20 microstates: MS –3/2 –1/2 1/2 3/2 + – – + – + +2 1 1 0 1 1 0 + – – + – + 1 1 –1 1 1 –1 +1 – – + + – + 1 0 0 1 0 0 + – – + + – 1 0 –1 1 0 –1 – – – – + – + – + + + + ML 0 1 0 –1 1 0 –1 1 0 –1 1 0 –1 1– 0– –1+ 1– 0+ –1+ + – – + – + –1 –1 1 –1 –1 1 –1 – – + + – + –1 0 0 –1 0 0 + – – + – + –2 –1 –1 0 –1 –1 0 Terms: L = 0, S = 3/2: 4S (ground state) L = 2, S = 1/2: 2D 2 L = 1, S = 1/2: P 6! 10! b. p1d1 There are 60 microstates: 1!5! 1!9! M S –1 0 1 – – – + + – + + 3 1 2 1 2 , 1 2 1 2 – – + – – + + + 1 1 1 1 , 1 1 1 1 2 + + 0– 2– 0– 2+, 0+ 2– 0 2 1– 0– 1+ 0–, 0+ 1– 1+ 0+ + + 1 0– 1– 1– 0+, 0– 1+ 0 1 + + –1– 2– –1– 2+, –1+ 2– –1 2 – – + – – + + + 1 –1 1 –1 , 1 –1 1 –1 – – – + + – + + ML 0 0 0 0 0 , 0 0 0 0 + + –1– 1– –1+ 1–, –1– 1+ –1 1 – – + – + – + + –1 0 –1 0 , 0 –1 –1 0 – – – + – + + + –1 0 –1 –1 0 , 0 –1 0 –1 – – – + + – 1+ –2+ 1 –2 1 –2 , 1 –2 –1– –1– –1+ –1–, –1– –1+ –1– –1+ –2 + – – + + + 0– –2– 0 –2 , 0 –2 0 –2 –3 –1– –2– –1– –2+, –1+ –2– –1+ –2+ Terms: L = 3, S = 1 3F (ground state) L = 2, S = 0 1D L = 3, S = 0 1F L = 1, S = 1 3P 3 1 L = 2, S = 1 D L = 1, S = 0 P The two electrons have quantum numbers that are independent of each other, because the electrons are in different orbitals.
    [Show full text]
  • The Five Common Particles
    The Five Common Particles The world around you consists of only three particles: protons, neutrons, and electrons. Protons and neutrons form the nuclei of atoms, and electrons glue everything together and create chemicals and materials. Along with the photon and the neutrino, these particles are essentially the only ones that exist in our solar system, because all the other subatomic particles have half-lives of typically 10-9 second or less, and vanish almost the instant they are created by nuclear reactions in the Sun, etc. Particles interact via the four fundamental forces of nature. Some basic properties of these forces are summarized below. (Other aspects of the fundamental forces are also discussed in the Summary of Particle Physics document on this web site.) Force Range Common Particles It Affects Conserved Quantity gravity infinite neutron, proton, electron, neutrino, photon mass-energy electromagnetic infinite proton, electron, photon charge -14 strong nuclear force ≈ 10 m neutron, proton baryon number -15 weak nuclear force ≈ 10 m neutron, proton, electron, neutrino lepton number Every particle in nature has specific values of all four of the conserved quantities associated with each force. The values for the five common particles are: Particle Rest Mass1 Charge2 Baryon # Lepton # proton 938.3 MeV/c2 +1 e +1 0 neutron 939.6 MeV/c2 0 +1 0 electron 0.511 MeV/c2 -1 e 0 +1 neutrino ≈ 1 eV/c2 0 0 +1 photon 0 eV/c2 0 0 0 1) MeV = mega-electron-volt = 106 eV. It is customary in particle physics to measure the mass of a particle in terms of how much energy it would represent if it were converted via E = mc2.
    [Show full text]
  • Review Sheet on Determining Term Symbols
    Review Sheet on Determining Term Symbols Term symbols for electronic configurations are useful not only to the spectroscopist but also to the inorganic chemist interested in understanding electronic and magnetic properties of molecules. We will concentrate on the method of Douglas and McDaniel (p. 26ff); however, you may find other treatments equal or superior to the D & M method. Term symbols are a shorthand method used to describe the energy, angular momentum, and spin multiplicity of an atom in any particular state. The general form is a given as Tj where T is a capital letter corresponding to the value of L (the angular momentum quantum number) and may be assigned as S, P, D, F, G, … for |L| = 0, 1, 2, 3, 4, … respectively. The superscript “a” is called the spin multiplicity and can be evaluated as a = 2S +1 where S is the spin quantum number. The subscript “j” is the numerical value of J, a new quantum number defined as: J = l +S, which corresponds to the total orbital and spin angular momentum of the system. The term symbol 3P is read as triplet – Pee state and indicates that there are two unpaired electrons in a state with maximum orbital angular momentum, L=1. The number of microstates (N) of a system corresponds to the total number of distinct arrangements for “e” number of electrons to be placed in “n” number of possible orbital positions. N = # of microstates = n!/(e!(n-e)!) For a set of p orbitals n = 6 since there are 2 positions in each orbital.
    [Show full text]
  • From the Atomic to the Femtoscale Linda Young Argonne National
    From the Atomic to the Femtoscale Linda Young Argonne National Laboratory, Argonne, Illinois 60439 Abstract: Atom traps of lithium can be used to provide a new window on few-body atomic and nuclear systems. The trapped atoms form an excellent sample, dense and motionless, for precision measurements. This talk describes experiments using ultracold lithium atoms to study ionization dynamics (a persistent few-body dynamical problem) and outline proposed precision measurements of isotope shifts to determine charge radii of short-lived lithium isotopes (a challenging few-body nuclear physics problem). Introduction The lithium atom, with three electrons and six through eleven nucleons, is a hotbed of activity for few-body theorists in both atomic and nuclear physics. Atomic theorists have the advantage that the forces in the problem, Coulomb interactions, are well known. This advantage simplifies development of many-body techniques for both structure and dynamics [1]. On the atomic physics side, there has been recent effort devoted both to precision calculations of atomic structure [2-5] and to understanding the dynamical correlation between the outgoing electrons in photo triple-ionization [6-9]. On the nuclear physics side, a long-standing program to calculate properties of few-body nuclei has now reached A ≤ 10 [10,11]. In these calculations, the nucleon-nucleon potentials are adjusted to fit the large collection of pp and np scattering data. Impressively, the calculations have fit the binding energies of all A ≤ 10 nuclei to an rms deviation of ≈400 keV, predicted the absence of stable A=5, 8 nuclei [12], and simultaneously predicted rms proton radii, rms neutron radii, quadrupole moments and magnetic moments.
    [Show full text]
  • A Discussion on Characteristics of the Quantum Vacuum
    A Discussion on Characteristics of the Quantum Vacuum Harold \Sonny" White∗ NASA/Johnson Space Center, 2101 NASA Pkwy M/C EP411, Houston, TX (Dated: September 17, 2015) This paper will begin by considering the quantum vacuum at the cosmological scale to show that the gravitational coupling constant may be viewed as an emergent phenomenon, or rather a long wavelength consequence of the quantum vacuum. This cosmological viewpoint will be reconsidered on a microscopic scale in the presence of concentrations of \ordinary" matter to determine the impact on the energy state of the quantum vacuum. The derived relationship will be used to predict a radius of the hydrogen atom which will be compared to the Bohr radius for validation. The ramifications of this equation will be explored in the context of the predicted electron mass, the electrostatic force, and the energy density of the electric field around the hydrogen nucleus. It will finally be shown that this perturbed energy state of the quan- tum vacuum can be successfully modeled as a virtual electron-positron plasma, or the Dirac vacuum. PACS numbers: 95.30.Sf, 04.60.Bc, 95.30.Qd, 95.30.Cq, 95.36.+x I. BACKGROUND ON STANDARD MODEL OF COSMOLOGY Prior to developing the central theme of the paper, it will be useful to present the reader with an executive summary of the characteristics and mathematical relationships central to what is now commonly referred to as the standard model of Big Bang cosmology, the Friedmann-Lema^ıtre-Robertson-Walker metric. The Friedmann equations are analytic solutions of the Einstein field equations using the FLRW metric, and Equation(s) (1) show some commonly used forms that include the cosmological constant[1], Λ.
    [Show full text]
  • XSAMS: XML Schema for Atomic, Molecular and Solid Data
    XSAMS: XML Schema for Atomic, Molecular and Solid Data Version 0.1.1 Draft Document, January 14, 2011 This version: http://www-amdis.iaea.org/xsams/docu/v0.1.1.pdf Latest version: http://www-amdis.iaea.org/xsams/docu/ Previous versions: http://www-amdis.iaea.org/xsams/docu/v0.1.pdf Editors: M.L. Dubernet, D. Humbert, Yu. Ralchenko Authors: M.L. Dubernet, D. Humbert, Yu. Ralchenko, E. Roueff, D.R. Schultz, H.-K. Chung, B.J. Braams Contributors: N. Moreau, P. Loboda, S. Gagarin, R.E.H. Clark, M. Doronin, T. Marquart Abstract This document presents a proposal for an XML schema aimed at describing Atomic, Molec- ular and Particle Surface Interaction Data in distributed databases around the world. This general XML schema is a collaborative project between International Atomic Energy Agency (Austria), National Institute of Standards and Technology (USA), Universit´ePierre et Marie Curie (France), Oak Ridge National Laboratory (USA) and Paris Observatory (France). Status of this document It is a draft document and it may be updated, replaced, or obsoleted by other documents at any time. It is inappropriate to use this document as reference materials or to cite them as other than “work in progress”. Acknowledgements The authors wish to acknowledge all colleagues and institutes, atomic and molecular database experts and physicists who have collaborated through different discussions to the building up of the concepts described in this document. Change Log Version 0.1: June 2009 Version 0.1.1: November 2010 Contents 1 Introduction 11 1.1 Motivation..................................... 11 1.2 Limitations..................................... 12 2 XSAMS structure, types and attributes 13 2.1 Atomic, Molecular and Particle Surface Interaction Data XML Schema Structure 13 2.2 Attributes ....................................
    [Show full text]
  • Atomic Spectroscopy Summary
    ATOMIC SPECTROSCOPY SUMMARY TERM SYMBOLS Because the total angular momentum and energy commute, we can use the angular momentum (L) to label the energy states. Note that electron configurations are ambiguous in that we don’t specify which orbital or which spin each electron has. This means that there are a number of different sets of 푚ℓ and 푚푠 that are consistent with a given configuration. Why are there different energies? Because of electron/electron repulsion and spin-orbit coupling, the ways we can fill the orbitals DO NOT NESCESSARILLY have the same energy. (Remember that we talked about how Hund’s rules come from this. How spin up elelctrons interact less with other spin up electrons in the same subshell.) Components of a Term Symbol: 2푆+1 퐿퐽 RULES FOR TERM SYMBOLS Term Symbols are assigned based on two main rules: 1. Unsöld’s Rule: All filled shells/subshells are ignored when calculating L and S. Basically, all of the electrons will have their contributions to spin and orbital angular momentum canceled by other electrons in the same subshell. 2. “holes”: The state of the “hole” (an empty spot in an orbital where an electron could be but isn’t) is the same as the state of an electron. This means that when we look at a p5 we can pretend it is a p1. This is incredibly useful as it means that instead of 5 electrons to tinker with, you can just act like it is 1. CALCULATING L: L is the Total Angular Momentum L= (sum of ℓ values) ….(smallest positive difference of ℓ values) o 2 s electrons (different shells) L=0+0,…,0-0 L=0 o 1 d electron and 1 p electron L=2+1,…,2-1 L=3, 2, 1 S is the Total Spin S is the total spin quantum number of the electrons in an atom.
    [Show full text]
  • Lesson 1: the Single Electron Atom: Hydrogen
    Lesson 1: The Single Electron Atom: Hydrogen Irene K. Metz, Joseph W. Bennett, and Sara E. Mason (Dated: July 24, 2018) Learning Objectives: 1. Utilize quantum numbers and atomic radii information to create input files and run a single-electron calculation. 2. Learn how to read the log and report files to obtain atomic orbital information. 3. Plot the all-electron wavefunction to determine where the electron is likely to be posi- tioned relative to the nucleus. Before doing this exercise, be sure to read through the Ins and Outs of Operation document. So, what do we need to build an atom? Protons, neutrons, and electrons of course! But the mass of a proton is 1800 times greater than that of an electron. Therefore, based on de Broglie’s wave equation, the wavelength of an electron is larger when compared to that of a proton. In other words, the wave-like properties of an electron are important whereas we think of protons and neutrons as particle-like. The separation of the electron from the nucleus is called the Born-Oppenheimer approximation. So now we need the wave-like description of the Hydrogen electron. Hydrogen is the simplest atom on the periodic table and the most abundant element in the universe, and therefore the perfect starting point for atomic orbitals and energies. The compu- tational tool we are going to use is called OPIUM (silly name, right?). Before we get started, we should know what’s needed to create an input file, which OPIUM calls a parameter files. Each parameter file consist of a sequence of ”keyblocks”, containing sets of related parameters.
    [Show full text]
  • 1.1. Introduction the Phenomenon of Positron Annihilation Spectroscopy
    PRINCIPLES OF POSITRON ANNIHILATION Chapter-1 __________________________________________________________________________________________ 1.1. Introduction The phenomenon of positron annihilation spectroscopy (PAS) has been utilized as nuclear method to probe a variety of material properties as well as to research problems in solid state physics. The field of solid state investigation with positrons started in the early fifties, when it was recognized that information could be obtained about the properties of solids by studying the annihilation of a positron and an electron as given by Dumond et al. [1] and Bendetti and Roichings [2]. In particular, the discovery of the interaction of positrons with defects in crystal solids by Mckenize et al. [3] has given a strong impetus to a further elaboration of the PAS. Currently, PAS is amongst the best nuclear methods, and its most recent developments are documented in the proceedings of the latest positron annihilation conferences [4-8]. PAS is successfully applied for the investigation of electron characteristics and defect structures present in materials, magnetic structures of solids, plastic deformation at low and high temperature, and phase transformations in alloys, semiconductors, polymers, porous material, etc. Its applications extend from advanced problems of solid state physics and materials science to industrial use. It is also widely used in chemistry, biology, and medicine (e.g. locating tumors). As the process of measurement does not mostly influence the properties of the investigated sample, PAS is a non-destructive testing approach that allows the subsequent study of a sample by other methods. As experimental equipment for many applications, PAS is commercially produced and is relatively cheap, thus, increasingly more research laboratories are using PAS for basic research, diagnostics of machine parts working in hard conditions, and for characterization of high-tech materials.
    [Show full text]
  • Bouncing Oil Droplets, De Broglie's Quantum Thermostat And
    Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 28 August 2018 doi:10.20944/preprints201808.0475.v1 Peer-reviewed version available at Entropy 2018, 20, 780; doi:10.3390/e20100780 Article Bouncing oil droplets, de Broglie’s quantum thermostat and convergence to equilibrium Mohamed Hatifi 1, Ralph Willox 2, Samuel Colin 3 and Thomas Durt 4 1 Aix Marseille Université, CNRS, Centrale Marseille, Institut Fresnel UMR 7249,13013 Marseille, France; hatifi[email protected] 2 Graduate School of Mathematical Sciences, the University of Tokyo, 3-8-1 Komaba, Meguro-ku, 153-8914 Tokyo, Japan; [email protected] 3 Centro Brasileiro de Pesquisas Físicas, Rua Dr. Xavier Sigaud 150,22290-180, Rio de Janeiro – RJ, Brasil; [email protected] 4 Aix Marseille Université, CNRS, Centrale Marseille, Institut Fresnel UMR 7249,13013 Marseille, France; [email protected] Abstract: Recently, the properties of bouncing oil droplets, also known as ‘walkers’, have attracted much attention because they are thought to offer a gateway to a better understanding of quantum behaviour. They indeed constitute a macroscopic realization of wave-particle duality, in the sense that their trajectories are guided by a self-generated surrounding wave. The aim of this paper is to try to describe walker phenomenology in terms of de Broglie-Bohm dynamics and of a stochastic version thereof. In particular, we first study how a stochastic modification of the de Broglie pilot-wave theory, à la Nelson, affects the process of relaxation to quantum equilibrium, and we prove an H-theorem for the relaxation to quantum equilibrium under Nelson-type dynamics.
    [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]