Spectroscopic Selection Rules: the Role of Photon States

Total Page:16

File Type:pdf, Size:1020Kb

Spectroscopic Selection Rules: the Role of Photon States Research: Science and Education edited by Advanced Chemistry Classroom and Laboratory Joseph J. BelBruno Dartmouth College Hanover, NH 03755 Spectroscopic Selection Rules: The Role of Photon States Andrew M. Ellis Department of Chemistry, University of Leicester, University Road, Leicester LE1 7RH, UK; [email protected] Selection rules are vital in the interpretation of atomic and In quantum mechanics, all three Cartesian components of molecular spectra. The usual starting point for a derivation this vector cannot be known simultaneously (for a proof see of selection rules is the transition moment which, in intro- ref 2). Instead, there are only two simultaneously observable ductory spectroscopy courses, is normally taken on trust. quantities, the total angular momentum and any one of However, many students find the transition moment to be its Cartesian components. Which component is chosen is a somewhat obscure quantity that does little to reveal the arbitrary in the absence of some external constraint but it is underlying physics. Furthermore, the journey from the tran- conventional to select the z-component. As a result there sition moment to a particular selection rule is not always easy. are two quantum numbers that define the orbital angular This is especially troublesome when all one wishes to achieve momentum of an electron, the total quantum number l and is a justification of certain selection rules, rather than to give its corresponding projection onto the z-axis, quantum a full-blown account. number ml . An easier way of justifying selection rules is to apply con- Figure 1 summarizes the findings in a vector diagram. servation arguments to individual photon–atom (or photon– Similar diagrams are commonplace in textbooks dealing with molecule) interactions. One quantity that must be conserved the quantum theory of angular momentum. The total angular is angular momentum. To invoke this argument, the angular momentum, represented by the length of the vector, has the momentum of the photon must first be known. A transition magnitude that satisfies angular momentum conservation must cause a ll +1 h change in angular momentum state of the absorbing or emit- ting atom (or molecule) that compensates for the loss or gain The quantum number l may have any integer value includ- of photon angular momentum. ing zero. Although only one is indicated in the figure, there It is commonly assumed that a photon possesses an are 2l + 1 possible values of m ranging from {l to + l in unit ⁄ ⁄ l angular momentum of ± h where h = h/2π. In other words, steps. The values of the projection of the orbital angular mo- ⁄ it has a unit quantum of angular momentum and therefore mentum on the z-axis are ml h. unit changes in the angular momentum quantum number Strictly speaking, l and ml are not the only important are expected for an absorbing or emitting atom or molecule. angular momentum quantum numbers. Angular momentum Foss presented this argument in an article in this Journal some states also possess parity. Parity refers to the effect of an in- years ago (1) to rationalize selection rules such as ∆l = ±1 version of all spatial coordinates on the sign of the angular and ∆J = ±1, and it has also appeared in several textbooks momentum wave function. If there is no change in sign the (e.g., see refs 2–5). While it is a useful starting point, this parity is said to be even; a change in sign corresponds to a model fails to account properly for selection rules when sub- state of odd parity. For orbital angular momentum the parity is jected to closer examination. In this article a more compre- ({1)l . This is easily seen by considering the well-known shapes hensive, but nevertheless straightforward, justification of an- of s, p, d, and f orbitals. An inversion of coordinates about the gular momentum selection rules is presented. I begin with a atomic nucleus does not change the sign of s and d orbitals, very brief review of some key results from the quantum and so they possess even parity; but it does change the sign theory of angular momentum. I then show some of the flaws of p and f orbitals, so they must have odd parity. Parity places in the model used by Foss and others. Finally, I explain how restrictions on spectroscopic transitions, as will be seen later. these flaws disappear when a more complete quantum elec- trodynamic model of a photon is employed. Coupling of Two Angular Momenta The coupling of two angular momenta, l1 and l2, yields Review of the Basic Elements of the Quantum Theory a resultant L, which is the vector sum of the two individual of Angular Momentum Most undergraduate courses on quantum mechanics Figure 1. Vector model of quantized include a description of the angular momentum of a single angular momentum. The magnitude rotating or orbiting body. Many also include an elementary of the angular momentum, repre- account of the coupling of angular momenta. Some key facts sented by the length of the vector, is from both topics are summarized below without proof. ll +1 h Single Body The vector precesses (not shown) about the z axis, producing inde- Angular momentum is a vector quantity. An example is the terminate components in the x and orbital angular momentum of an individual electron in an atom y directions but a well-defined com- (the possibility of coupling is dealt with in the next section). ponent along the z axis. JChemEd.chem.wisc.edu • Vol. 76 No. 9 September 1999 • Journal of Chemical Education 1291 Research: Science and Education angular momentum vectors, Photon energy. This is familiar from elementary physics— it depends linearly on the frequency ν; that is, E = hν. L = l + l (1) 1 2 Angular momentum. The quantum mechanics of photon In quantum mechanics, the total orbital angular momentum angular momentum has much in common with any other is quantized and has a quantum number L, which is restricted type of angular momentum. A photon can be ascribed a total to the values angular momentum quantum number, j, and a projection ≥ ≥ |l1 + l2| L |l1 – l2| (2) quantum number mj. The magnitude of this momentum is This important result is central to the arguments about con- given by servation of angular momentum presented later. jj +1 h ⁄ Transitions between Angular Momentum States: The and the projection along the z axis is mj h. Photons may Failure of the Simple Model possess any positive integer j, but j = 0 is impossible; that is, photons cannot have zero angular momentum. Photons possess quantized angular momentum. As dis- Parity. An angular momentum state of a photon must have cussed by Foss (1), experiments have shown that all photons have ± ⁄ parity. For a given j both odd and even parities are possible. an angular momentum of h along the axis of propagation. Proof of the above statements are rather involved and This is known as the helicity of a photon. The two possible can be found in books on quantum electrodynamics (see, for signs refer to the two possible states of circularly polarized example, refs 6, 7). However, the take-home message is very light, left and right circular polarization. Linearly polarized simple. The different possible angular momentum states and light can be thought of as an equal mixture of left and right parities can be identified with different types of photons. Notice circularly polarized light, while any other state of polarization that, in contrast to the electron orbital angular momentum, (e.g., elliptical polarization) will contain unequal proportions both even and odd parities are possible for each value of the of photons with opposite helicities. photon angular momentum quantum number j. Taking j =1 Now consider a transition in which an electron moves and j = 2 as examples, the following types of photon occur: from one atomic orbital to another by photon absorption or j = 1, parity = odd electric dipole photon emission. According to the Laporte selection rule, only tran- sitions for which ∆l = ±1 are allowed. This selection rule can j = 1, parity = even magnetic dipole photon be justified on the grounds of parity conservation, as will be j = 2, parity = even electric quadrupole photon shown later. But it should also be possible to show that it is j = 2, parity = odd magnetic quadrupole photon consistent with angular momentum conservation. The argument In a classical electromagnetic field, the electric and mag- that is normally employed, as mentioned earlier, is that since netic parts contain various multipole components: dipole, a photon possesses unit angular momentum, ± h⁄ , only a quadrupole, and onwards to higher multipoles. In quantum compensating unit change in the electron orbital angular electrodynamics, we can think of these parts as arising from momentum is compatible with the conservation of overall different types of photon, and the names of the photon states angular momentum. given above reflects this. However, a trivial example will show that this explana- It can be shown (6, 7) that all photons, regardless of j tion is flawed. Consider the transition l = 1 ↔ l = 0. The or parity, must possess an angular momentum along their conservation of angular momentum requires that the vector direction of propagation of ± h⁄ , thus concurring with experi- sum of all angular momenta is zero. For l = 0 the orbital ment (1, 8); this is the helicity referred to earlier. However, angular momentum is zero and so in a transition to l = 1 the it is important to recognize that the direction of propagation magnitude of the angular momentum gained by the electron is not the same as the axis (z) along which the projection must exactly equal that of the photon that is being absorbed.
Recommended publications
  • CHAPTER 13 Molecular Spectroscopy 2: Electronic Transitions
    CHAPTER 13 Molecular Spectroscopy 2: Electronic Transitions I. General Features of Electronic spectroscopy. A. Visible and ultraviolet photons excite electronic state transitions. εphoton = 120 to 1200 kJ/mol. Table 13A.1 Colour, frequency, and energy of light B. Electronic transitions are accompanied by vibrational and rotational transitions of any Δν, ΔJ (complex and rich spectra in gas phase). In liquids and solids, this fine structure merges together due to collisional broadening and is washed out. Chlorophyll in solution, vibrational fine structure in gaseous SO2 UV visible spectrum. CHAPTER 13 1 C. Molecules need not have permanent dipole to undergo electronic transitions. All that is required is a redistribution of electronic and nuclear charge between the initial and final electronic state. Transition intensity ~ |µ|2 where µ is the magnitude of the transition dipole moment given by: ˆ d µ = ∫ ψfinalµψ initial τ where µˆ = −e ∑ri + e ∑Ziri € electrons nuclei r vector position of the particle i = D. Franck-Condon principle: Because nuclei are so much more massive and € sluggish than electrons, electronic transitions can happen much faster than the nuclei can respond. Electronic transitions occur vertically on energy diagram at right. (Hence the name vertical transition) Transition probability depends on vibrational wave function overlap (Franck-Condon factor). In this example: ν = 0 ν’ = 0 have small overlap ν = 0 ν’ = 2 have greatest overlap Now because the total molecular wavefunction can be approximately factored into 2 terms, electronic wavefunction and the nuclear position wavefunction (using Born-Oppenheimer approx.) the transition dipole can also be factored into 2 terms also: S µ = µelectronic i,f S d Franck Condon factor i,j i,f = ∫ ψf,vibr ψi,vibr τ = − CHAPTER 13 2 See derivation “Justification 14.2” II.
    [Show full text]
  • Lecture #2: August 25, 2020 Goal Is to Define Electrons in Atoms
    Lecture #2: August 25, 2020 Goal is to define electrons in atoms • Bohr Atom and Principal Energy Levels from “orbits”; Balance of electrostatic attraction and centripetal force: classical mechanics • Inability to account for emission lines => particle/wave description of atom and application of wave mechanics • Solutions of Schrodinger’s equation, Hψ = Eψ Required boundaries => quantum numbers (and the Pauli Exclusion Principle) • Electron configurations. C: 1s2 2s2 2p2 or [He]2s2 2p2 Na: 1s2 2s2 2p6 3s1 or [Ne] 3s1 => Na+: [Ne] Cl: 1s2 2s2 2p6 3s23p5 or [Ne]3s23p5 => Cl-: [Ne]3s23p6 or [Ar] What you already know: Quantum Numbers: n, l, ml , ms n is the principal quantum number, indicates the size of the orbital, has all positive integer values of 1 to ∞(infinity) (Bohr’s discrete orbits) l (angular momentum) orbital 0s l is the angular momentum quantum number, 1p represents the shape of the orbital, has integer values of (n – 1) to 0 2d 3f ml is the magnetic quantum number, represents the spatial direction of the orbital, can have integer values of -l to 0 to l Other terms: electron configuration, noble gas configuration, valence shell ms is the spin quantum number, has little physical meaning, can have values of either +1/2 or -1/2 Pauli Exclusion principle: no two electrons can have all four of the same quantum numbers in the same atom (Every electron has a unique set.) Hund’s Rule: when electrons are placed in a set of degenerate orbitals, the ground state has as many electrons as possible in different orbitals, and with parallel spin.
    [Show full text]
  • The Quantum Mechanical Model of the Atom
    The Quantum Mechanical Model of the Atom Quantum Numbers In order to describe the probable location of electrons, they are assigned four numbers called quantum numbers. The quantum numbers of an electron are kind of like the electron’s “address”. No two electrons can be described by the exact same four quantum numbers. This is called The Pauli Exclusion Principle. • Principle quantum number: The principle quantum number describes which orbit the electron is in and therefore how much energy the electron has. - it is symbolized by the letter n. - positive whole numbers are assigned (not including 0): n=1, n=2, n=3 , etc - the higher the number, the further the orbit from the nucleus - the higher the number, the more energy the electron has (this is sort of like Bohr’s energy levels) - the orbits (energy levels) are also called shells • Angular momentum (azimuthal) quantum number: The azimuthal quantum number describes the sublevels (subshells) that occur in each of the levels (shells) described above. - it is symbolized by the letter l - positive whole number values including 0 are assigned: l = 0, l = 1, l = 2, etc. - each number represents the shape of a subshell: l = 0, represents an s subshell l = 1, represents a p subshell l = 2, represents a d subshell l = 3, represents an f subshell - the higher the number, the more complex the shape of the subshell. The picture below shows the shape of the s and p subshells: (notice the electron clouds) • Magnetic quantum number: All of the subshells described above (except s) have more than one orientation.
    [Show full text]
  • Lecture B4 Vibrational Spectroscopy, Part 2 Quantum Mechanical SHO
    Lecture B4 Vibrational Spectroscopy, Part 2 Quantum Mechanical SHO. WHEN we solve the Schrödinger equation, we always obtain two things: 1. a set of eigenstates, |ψn>. 2. a set of eigenstate energies, En QM predicts the existence of discrete, evenly spaced, vibrational energy levels for the SHO. n = 0,1,2,3... For the ground state (n=0), E = ½hν. Notes: This is called the zero point energy. Optical selection rule -- SHO can absorb or emit light with a ∆n = ±1 The IR absorption spectrum for a diatomic molecule, such as HCl: diatomic Optical selection rule 1 -- SHO can absorb or emit light with a ∆n = ±1 Optical selection rule 2 -- a change in molecular dipole moment (∆μ/∆x) must occur with the vibrational motion. (Note: μ here means dipole moment). If a more realistic Morse potential is used in the Schrödinger Equation, these energy levels get scrunched together... ΔE=hν ΔE<hν The vibrational spectroscopy of polyatomic molecules gets more interesting... For diatomic or linear molecules: 3N-5 modes For nonlinear molecules: 3N-6 modes N = number of atoms in molecule The vibrational spectroscopy of polyatomic molecules gets more interesting... Optical selection rule 1 -- SHO can absorb or emit light with a ∆n = ±1 Optical selection rule 2 -- a change in molecular dipole moment (∆μ/∆x) must occur with the vibrational motion of a mode. Consider H2O (a nonlinear molecule): 3N-6 = 3(3)-6 = 3 Optical selection rule 2 -- a change in molecular dipole moment (∆μ/∆x) must occur with the vibrational motion of a mode. Consider H2O (a nonlinear molecule): 3N-6 = 3(3)-6 = 3 All bands are observed in the IR spectrum.
    [Show full text]
  • Quantum Spin Systems and Their Local and Long-Time Properties Carolee Wheeler Faculty Advisor: Robert Sims
    URA Project Proposal Quantum Spin Systems and Their Local and Long-Time Properties Carolee Wheeler Faculty Advisor: Robert Sims In quantum mechanics, spin is an important concept having to do with atomic nuclei, hadrons, and elementary particles. Spin may be thought of as a measure of a particle’s rotation about its axis. However, spin differs from orbital angular momentum in the sense that the particles may carry integer or half-integer quantum numbers, i.e. 0, 1/2, 1, 3/2, 2, etc., whereas orbital angular momentum may only take integer quantum numbers. Furthermore, the spin of a charged elementary particle is related to a magnetic dipole moment. All quantum mechanic particles have an inherent spin. This is due to the fact that elementary particles (such as photons, electrons, or quarks) cannot be divided into smaller entities. In other words, they cannot be viewed as particles that are made up of individual, smaller particles that rotate around a common center. Thus, the spin that elementary particles carry is an intrinsic property [1]. An important characteristic of spin in quantum mechanics is that it is quantized. The magnitude of spin takes values S = h s(s + )1 , with h being the reduced Planck’s constant and s being the spin quantum number (a non- negative integer or half-integer). Spin may also be viewed in composite particles, and is calculated by summing the spins of the constituent particles. In the case of atoms and molecules, spin is the sum of the spins of unpaired electrons [1]. Particles with spin can possess a magnetic moment.
    [Show full text]
  • The Principal Quantum Number the Azimuthal Quantum Number The
    To completely describe an electron in an atom, four quantum numbers are needed: energy (n), angular momentum (ℓ), magnetic moment (mℓ), and spin (ms). The Principal Quantum Number This quantum number describes the electron shell or energy level of an atom. The value of n ranges from 1 to the shell containing the outermost electron of that atom. For example, in caesium (Cs), the outermost valence electron is in the shell with energy level 6, so an electron incaesium can have an n value from 1 to 6. For particles in a time-independent potential, as per the Schrödinger equation, it also labels the nth eigen value of Hamiltonian (H). This number has a dependence only on the distance between the electron and the nucleus (i.e. the radial coordinate r). The average distance increases with n, thus quantum states with different principal quantum numbers are said to belong to different shells. The Azimuthal Quantum Number The angular or orbital quantum number, describes the sub-shell and gives the magnitude of the orbital angular momentum through the relation. ℓ = 0 is called an s orbital, ℓ = 1 a p orbital, ℓ = 2 a d orbital, and ℓ = 3 an f orbital. The value of ℓ ranges from 0 to n − 1 because the first p orbital (ℓ = 1) appears in the second electron shell (n = 2), the first d orbital (ℓ = 2) appears in the third shell (n = 3), and so on. This quantum number specifies the shape of an atomic orbital and strongly influences chemical bonds and bond angles.
    [Show full text]
  • Lecture 17 Highlights
    Lecture 23 Highlights The phenomenon of Rabi oscillations is very different from our everyday experience of how macroscopic objects (i.e. those made up of many atoms) absorb electromagnetic radiation. When illuminated with light (like French fries at McDonalds) objects tend to steadily absorb the light and heat up to a steady state temperature, showing no signs of oscillation with time. We did a calculation of the transition probability of an atom illuminated with a broad spectrum of light and found that the probability increased linearly with time, leading to a constant transition rate. This difference comes about because the many absorption processes at different frequencies produce a ‘smeared-out’ response that destroys the coherent Rabi oscillations and results in a classical incoherent absorption. Selection Rules All of the transition probabilities are proportional to “dipole matrix elements,” such as x jn above. These are similar to dipole moment calculations for a charge distribution in classical physics, except that they involve the charge distribution in two different states, connected by the dipole operator for the transition. In many cases these integrals are zero because of symmetries of the associated wavefunctions. This gives rise to selection rules for possible transitions under the dipole approximation (atom size much smaller than the electromagnetic wave wavelength). For example, consider the matrix element z between states of the n,l , m ; n ', l ', m ' Hydrogen atom labeled by the quantum numbers n,,l m and n',','l m : z = ψ()()r z ψ r d3 r n,l , m ; n ', l ', m ' ∫∫∫ n,,l m n',l ', m ' To make progress, first consider just the φ integral.
    [Show full text]
  • Introduction to Differential Equations
    Quantum Dynamics – Quick View Concepts of primary interest: The Time-Dependent Schrödinger Equation Probability Density and Mixed States Selection Rules Transition Rates: The Golden Rule Sample Problem Discussions: Tools of the Trade Appendix: Classical E-M Radiation POSSIBLE ADDITIONS: After qualitative section, do the two state system, and then first and second order transitions (follow Fitzpatrick). Chain together the dipole rules to get l = 2,0,-2 and m = -2, -2, … , 2. ??Where do we get magnetic rules? Look at the canonical momentum and the Asquared term. Schrödinger, Erwin (1887-1961) Austrian physicist who invented wave mechanics in 1926. Wave mechanics was a formulation of quantum mechanics independent of Heisenberg's matrix mechanics. Like matrix mechanics, wave mechanics mathematically described the behavior of electrons and atoms. The central equation of wave mechanics is now known as the Schrödinger equation. Solutions to the equation provide probability densities and energy levels of systems. The time-dependent form of the equation describes the dynamics of quantum systems. http://scienceworld.wolfram.com/biography/Schroedinger.html © 1996-2006 Eric W. Weisstein www-history.mcs.st-andrews.ac.uk/Biographies/Schrodinger.html Quantum Dynamics: A Qualitative Introduction Introductory quantum mechanics focuses on time-independent problems leaving Contact: [email protected] dynamics to be discussed in the second term. Energy eigenstates are characterized by probability density distributions that are time-independent (static). There are examples of time-dependent behavior that are by demonstrated by rather simple introductory problems. In the case of a particle in an infinite well with the range [ 0 < x < a], the mixed state below exhibits time-dependence.
    [Show full text]
  • Chem 1A Chapter 7 Exercises
    Chem 1A Chapter 7 Exercises Exercises #1 Characteristics of Waves 1. Consider a wave with a wavelength = 2.5 cm traveling at a speed of 15 m/s. (a) How long does it take for the wave to travel a distance of one wavelength? (b) How many wavelengths (or cycles) are completed in 1 second? (The number of waves or cycles completed per second is referred to as the wave frequency.) (c) If = 3.0 cm but the speed remains the same, how many wavelengths (or cycles) are completed in 1 second? (d) If = 1.5 cm and the speed remains the same, how many wavelengths (or cycles) are completed in 1 second? (e) What can you conclude regarding the relationship between wavelength and number of cycles traveled in one second, which is also known as frequency ()? Answer: (a) 1.7 x 10–3 s; (b) 600/s; (c) 500/s; (d) 1000/s; (e) ?) 2. Consider a red light with = 750 nm and a green light = 550 nm (a) Which light travels a greater distance in 1 second, red light or green light? Explain. (b) Which light (red or green) has the greater frequency (that is, contains more waves completed in 1 second)? Explain. (c) What are the frequencies of the red and green light, respectively? (d) If light energy is defined by the quantum formula, such that Ep = h, which light carries more energy per photon? Explain. 8 -34 (e) Assuming that the speed of light, c = 2.9979 x 10 m/s, Planck’s constant, h = 6.626 x 10 J.s., and 23 Avogadro’s number, NA = 6.022 x 10 /mol, calculate the energy: (i) per photon; (ii) per mole of photon for green and red light, respectively.
    [Show full text]
  • INTRODUCTION to CARBON-13 NMR the Following Table Shows The
    INTRODUCTION TO CARBON-13 NMR The following table shows the spin quantum numbers of several nuclei. Only elements that have a nuclear spin quantum number I other than zero are NMR active. Such elements have odd atomic number, or odd atomic mass, or both. Elements whose atomic number and atomic mass are both even have I = 0 and are not NMR active. SPIN QUANTUM NUMBERS OF COMMON NUCLEI 1 2 12 13 14 16 17 19 31 35 1H 1H 6C 6C 7N 8O 8O 9F 15P 17Cl Nuclear spin quantum 1/2 1 0 1/2 1 0 5/2 1/2 1/2 3/2 number (I) Number of spin states 2 3 0 2 3 0 6 2 2 4 n = 2I + 1 Because both proton and carbon have the same nuclear spin value, they behave similarly in NMR and simi- lar guidelines apply to the analysis of their spectra. However, there are also some important differences. Examination of the following table reveals the basis for some of those differences. FIELD STRENGTHS AND RESONANCE FREQUENCIES FOR SELECTED NUCLEI Natural Abundance Field Strength (B ) Absorption Frequency (n) Isotope o (%) in Tesla* in MHz 1.00 42.6 2.35 100 1H 99.98 4.70 200 7.05 300 2H 0.02 1.00 6.5 1.00 10.7 2.35 25 13C 1.11 4.70 50 7.05 75 19F 100 1.00 40 31P 100 1.00 17 *1 Tesla = 10,000 Gauss Natural abundance effect. The natural abundance of proton means that virtually all the hydrogen atoms in a molecule are the 1H isotope, and therefore are in high abundance.
    [Show full text]
  • HJ Lipkin the Weizmann Institute of Science Rehovot, Israel
    WHO UNDERSTANDS THE ZWEIG-IIZUKA RULE? H.J. Lipkin The Weizmann Institute of Science Rehovot, Israel Abstract: The theoretical basis of the Zweig-Iizuka Ru le is discussed with the emphasis on puzzles and paradoxes which remain open problems . Detailed analysis of ZI-violating transitions arising from successive ZI-allowed transitions shows that ideal mixing and additional degeneracies in the particle spectrum are required to enable cancellation of these higher order violations. A quantitative estimate of the observed ZI-violating f' + 2rr decay be second order transitions via the intermediate KK state agrees with experiment when only on-shell contributions are included . This suggests that ZI-violations for old particles are determined completely by on-shell contributions and cannot be used as input to estimate ZI-violation for new particle decays where no on-shell intermediate states exist . 327 INTRODUCTION - SOME BASIC QUESTIONS 1 The Zweig-Iizuka rule has entered the folklore of particle physics without any clear theoretical understanding or justification. At the present time nobody really understands it, and anyone who claims to should not be believed. Investigating the ZI rule for the old particles raises many interesting questions which may lead to a better understanding of strong interactions as well as giving additional insight into the experimentally observed suppression of new particle decays attributed to the ZI rule. This talk follows an iconoclastic approach emphasizing embarrassing questions with no simple answers which might lead to fruitful lines of investigation. Both theoretical and experimental questions were presented in the talk. However, the experimental side is covered in a recent paper2 and is not duplicated here.
    [Show full text]
  • Free Electron Fermi Gas
    Phys463.nb 33 6 Free Electron Fermi Gas 6.1. Electrons in a metal 6.1.1. Electrons in one atom One electron in an atom (a hydrogen-like atom): the nucleon has charge +Z e, where Z is the atomic number, and there is one electron moving around this nucleon Four quantum number: n, l and lz, sz. Energy levels En with n = 1, 2, 3 ... Z2 m e4 1 En = - (6.1) 2 2 2 2 32 p e0 Ñ n where m = me mN me + mN » me where me is the mass of an electron and mN is the mass of the nucleon. 13.6 eV Z2 En = - H L (6.2) n2 H L For each n, the angular momentum quantum number [L2 y = l l + 1 y] can take the values of l = 0, 1, 2, 3, 4 … , n - 1. These states are known as the s, p, d, f , g,… states For each l, the quantum number for Lz can be any integer betweenH -Ll and +l lz = -l, -l + 1,…, l - 1, l . For fixed n, l and lz, the spin quantum number sz can be +1/2 or -1/2 (up or down). H L At each n, there are 2 n2 quantum states. In a real atom In a hydrogen-like atom, for the same n, all the 2 n2 quantum number has the same energy. In a real atom, the energy level splits according to the total angular momentum quantum number l. Typically, the state with lower l has lower energy (not always true).
    [Show full text]