Chapter 7 Relativistic Quantum Mechanics

Total Page:16

File Type:pdf, Size:1020Kb

Chapter 7 Relativistic Quantum Mechanics Chapter 7 Relativistic Quantum Mechanics In the previous chapters we have investigated the Schr¨odinger equation, which is based on the non-relativistic energy-momentum relation. We now want to reconcile the principles of quantum mechanics with special relativity. Schr¨odinger actually first considered a relativistic equation for de Broglie’s matter waves, but was deterred by discovering some apparently unphysical prop- erties like the existence of plane waves with unbounded negative energies E = p2c2 + m2c4: − Classically a particle on the positive branch of the square root will keep E mc2 forever ≥p but quantum mechanically interactions can induce a jump to the negative branch releasing δE 2mc2. Schr¨odinger hence arrived at his famous equation in the non-relativistic context. ≥ Two years later Paul A.M. Dirac found a linearization of the relativistic energy–momentum relation, which explained the gyromagnetic ratio g = 2 of the electron as well as the fine structure of hydrogen. While his equation missed its original task of eliminating negative energy solutions, it was too successful to be wrong so that Dirac went on, inspired by Pauli’s exclusion principle, to solve the problem of unbounded negative energies by inventing the particle-hole duality that is nowadays familiar from semiconductors. In the relativistic context the holes are called anti-particles. While Dirac first tried to identify the anti-particle of the electron with the proton (the only other particle known at that time) this did not work for several reasons and he concluded that there must exist a positively charged particle with the same mass as the electron. The positron was then discovered in cosmic rays in 1932. Dirac thus made the first prediction of a new particle on theoretical grounds and, more generally, showed the existence of antimatter as an implication of the consistency of quantum mechanics with special relativity. This initiated the long development of the modern quantum field theory of elementary particles and interactions. After a brief discussion of the problem with negative energy solutions we now construct the Dirac equation and analyze its nonrelativistic limit. The issue of Lorentz transformations and 127 CHAPTER 7. RELATIVISTIC QUANTUM MECHANICS 128 further symmetries will be taken up in the chapter on symmetries and transformation groups. The Klein-Gordon-equation. If we start with the relativistic energy-momentum relation E2 = m2c4 + c2~p 2. (7.1) and use the correspondence rule E i~∂ and ~p ~ ~ we obtain → t → i ∇ (i~∂ )2 ψ(~x, t)=(m2c4 c2~2∆)ψ(~x, t), (7.2) t − which can be written as m2c2 + ψ(~x, t) = 0 (7.3) ~2 2 in terms of the d’Alembert operator := 1 ∂ ∆, briefly called “box.” While Schr¨odinger c2 ∂t2 − already knew this relativistic wave equation, it was later named after Klein and Gordon who first published it. An immediate problem for the use of (7.3) as an equation for quantum mechanical wave functions is the existence of negative energy solutions i i Et+ ~p~x 2 2 4 2 ψ (~x, t)= e− ~ ~ with E = mc + p c mc (7.4) E − ≤ − p so that the energy is unbounded from below. Once interactions are turned on electrons could thus emit an infinite amount of energy, which is clearly unphysical. If we try to avoid the negative energy solutions of the Klein-Gordon equation (7.3) by using an expansion of the positive square root, 2 2 4 6 2 p 2 p p p 2 E = mc 1+ 2 2 = mc + 3 2 + 5 4 . mc , (7.5) r m c 2m − 8m c 16m c − ≥ the Hamiltonian becomes an infinite series with derivatives of arbitrary order and we loose locality. More concretely, it can be shown that localized wavepackets cannot be constructed without contributions from plane waves with negative energies [Itzykson,Zuber]. 7.1 The Dirac-equation Dirac tried to avoid the problem with the negative energy solutions by linearization of the equation for the energy. We get an idea for how this could work by recalling the equation σiσj = 2 2 ½ δij ½ + iεijkσk for the Pauli matrices σi, which implies (~σ~v) = σiviσjvj = ~v . For massless particles we thus obtain the relativistic energy-momentum relation from a linear equation 2 2 2 2 2 ½ Eψ = c~σ~pψ E ½ = c (~p~σ) = c p . (7.6) ± ⇒ CHAPTER 7. RELATIVISTIC QUANTUM MECHANICS 129 Upon quantization the suggested Hamiltonian H = c~p~σ becomes a 2 2 matrix of momentum ± × operators because 0 1 0 i 1 0 p3 p1 ip2 ~ ∂3 ∂1 i∂2 ~σ = , − , ~p~σ = piσi = − = i − . (7.7) 1 0 i 0 0 1 ⇒ p1 + ip2 p3 ∂1 + i∂2 ∂3 − − − Equation (7.6) indeed shows up as the massless case of the Dirac equation. It is called Weyl equation and it describes the massless neutrinos of the standard model of particle physics. In 1928 Dirac made the following ansatz for a linear relation between energy and momenta 2 E ½ = cp α + mc β (7.8) · i i with four dimensionless Hermitian matrices αi and β. Taking squares and demanding the relativistic energy-momentum relation (7.1) we find 2 2 4 2 2 3 2 2 4 ½ E ½ =(m c + c pipi) = c αiαjpipj + mc pi(αiβ + βαi)+ β m c , (7.9) which is equivalent to the matrix equations 2 ½ β = ½, α , β = 0, α , α = 2δ , (7.10) { i } { i j} ij where A, B = AB + BA denotes the anticommutator. Assuming the existence of a solution { } we arrive at the free Dirac equation ~ i~∂ ψ = Hψ, H = cα ∂ + βmc2 (7.11) t i i i with H = H†. The coupling to electromagnetic fields is achieved as in the non-relativistic case with the replacement ~p ~ ~ e A~ and E i~∂ V . With V = eφ this leads to the → i ∇ − c → t − interacting Dirac equation ∂ ~ e i~ eφ ψ = cα ∂ A ψ + βmc2ψ, (7.12) ∂t − i i i − c i for charged particles in an electromagnetic field, where we still need to find matrices αi and β representing the Dirac algebra (7.10). While α = σ would satisfy α , α = 2δ ½ it is impossible to find a further 2 2 matrix i i { i j} ij × β solving (7.10). A simple argument shows that the dimension of the Dirac matrices has to be 2 even: Since β = ½ and α β = βα cyclicity of the trace tr(Mβ) = tr(βM) implies i − i tr α = tr α β2 = tr(α β)β = tr β(α β)= tr β(βα )= tr α (7.13) i i i i − i − i and hence tr α = 0 (similarly tr β = tr(βα )α = tr α (βα ) = tr α (α β) = tr β shows i 1 1 1 1 − 1 1 − 2 2 that the trace of β also has to vanish). On the other hand, αi = β = ½ implies that all CHAPTER 7. RELATIVISTIC QUANTUM MECHANICS 130 eigenvalues of α and β must be 1 so that tr α = tr β = 0 entails that half of the eigenvalues i ± i are +1 and the other half are 1. This is only possible for even-dimensional matrices. Dirac − hence tried an ansatz with 4 4 matrices and found the solution × 1 0 σi ½ 0 αi = αi† = αi− αi = , β = , 1 , (7.14) σi 0 0 ½ β = β = β − † − where we used a block notation with 2 2 matrices ½, 0 and σi as matrix entries. In full gory 0001× 0 0 0 i 0 0 1 0 10 0 0 000101 00 0 i −01 00 0 0 11 001 0 0 1 − detail the Dirac matrices read α1 = B0100C, α2 = B0 i 0 0C, α3 = B1 0 0 0 C, β = B0 0 1 0 C, but one B C B − C B C B − C @1000A @i 0 0 0A @0 10 0 A @00 0 1A − − should never use these explicit expressions and rather work with the defining equations (7.10), or possibly with the block notation (7.14) if a splitting of the 4-component spinor wave function T ψ = (ψ1,ψ2,ψ3,ψ4) into two 2-component spinors is unavoidable like in the non-relativistic limit (see below). It can be shown that (7.14) is the unique irreducible unitary representation of 1 1 the Dirac algebra (7.10), up to unitary equivalence α Uα U − , β UβU − with UU † = ½. i → i → Relativistic spin. In order to derive the spin operator for relativistic electrons we observe that the orbital angular momentum L~ = X~ P~ is not conserved for the free Dirac Hamiltonian × [H,L ] = [c~αP~ + βmc2,ε X P ]= ~ cε α P = 0 (7.15) i ijk j k i ijk j k 6 ~ ~ ~ ~ because [Pl, Xj]= i δlj. A conserved operator J = L + S can be constructed by observing that [H,ε α α ]= ε [c~αP~ , α α ]= ε cP ( α , α α α α , α ) (7.16) ijk j k ijk j k ijk l { l j} k − j{ l k} = ε cP (2δ α 2δ α ) = 4cε P α ijk l lj k − lk j ijk j k because [A, BC]= ABC + BAC BAC BCA = A, B C B A, C . The spin operator − − { } − { } ~ ~ σi 0 Si = εijkαjαk = (7.17) 4i 2 0 σi therefore yields a conserved total angular momentum J~ = L~ + S~. 1 Lorentz covariant form of the Dirac equation. Multiplication with the matrix c β from the left recasts the relation E eφ = c~α(~p e A~)+ βmc2 into − − c β( 1 E e φ) β~α(~p e A~) mc ψ = 0. (7.18) c − c − − c − The standard combination of coordinates, energy-momenta and gauge potentials into 4-vectors xµ =(ct, ~x), ∂ =( 1 ∂ , ~ ), pµ =( 1 E, ~p),Aµ =(φ, A~), (7.19) µ c t ∇ c which entails the relativistic version p P = i~∂ of the quantum mechanical correspondence µ → µ µ (2.3), suggests the combination of β and βα into a 4-vector γµ of 4 4 matrices as i × 1 0 0 0 µ µ ν µν µν 00 1 0 01 ½ − γ =(β, β~α) γ ,γ = 2g with g = B0 0 1 0C.
Recommended publications
  • 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]
  • Clifford Algebra and the Interpretation of Quantum
    In: J.S.R. Chisholm/A.K. Commons (Eds.), Cliord Algebras and their Applications in Mathematical Physics. Reidel, Dordrecht/Boston (1986), 321–346. CLIFFORD ALGEBRA AND THE INTERPRETATION OF QUANTUM MECHANICS David Hestenes ABSTRACT. The Dirac theory has a hidden geometric structure. This talk traces the concep- tual steps taken to uncover that structure and points out signicant implications for the interpre- tation of quantum mechanics. The unit imaginary in the Dirac equation is shown to represent the generator of rotations in a spacelike plane related to the spin. This implies a geometric interpreta- tion for the generator of electromagnetic gauge transformations as well as for the entire electroweak gauge group of the Weinberg-Salam model. The geometric structure also helps to reveal closer con- nections to classical theory than hitherto suspected, including exact classical solutions of the Dirac equation. 1. INTRODUCTION The interpretation of quantum mechanics has been vigorously and inconclusively debated since the inception of the theory. My purpose today is to call your attention to some crucial features of quantum mechanics which have been overlooked in the debate. I claim that the Pauli and Dirac algebras have a geometric interpretation which has been implicit in quantum mechanics all along. My aim will be to make that geometric interpretation explicit and show that it has nontrivial implications for the physical interpretation of quantum mechanics. Before getting started, I would like to apologize for what may appear to be excessive self-reference in this talk. I have been pursuing the theme of this talk for 25 years, but the road has been a lonely one where I have not met anyone travelling very far in the same direction.
    [Show full text]
  • Magnetic Moment of a Spin, Its Equation of Motion, and Precession B1.1.6
    Magnetic Moment of a Spin, Its Equation of UNIT B1.1 Motion, and Precession OVERVIEW The ability to “see” protons using magnetic resonance imaging is predicated on the proton having a mass, a charge, and a nonzero spin. The spin of a particle is analogous to its intrinsic angular momentum. A simple way to explain angular momentum is that when an object rotates (e.g., an ice skater), that action generates an intrinsic angular momentum. If there were no friction in air or of the skates on the ice, the skater would spin forever. This intrinsic angular momentum is, in fact, a vector, not a scalar, and thus spin is also a vector. This intrinsic spin is always present. The direction of a spin vector is usually chosen by the right-hand rule. For example, if the ice skater is spinning from her right to left, then the spin vector is pointing up; the skater is rotating counterclockwise when viewed from the top. A key property determining the motion of a spin in a magnetic field is its magnetic moment. Once this is known, the motion of the magnetic moment and energy of the moment can be calculated. Actually, the spin of a particle with a charge and a mass leads to a magnetic moment. An intuitive way to understand the magnetic moment is to imagine a current loop lying in a plane (see Figure B1.1.1). If the loop has current I and an enclosed area A, then the magnetic moment is simply the product of the current and area (see Equation B1.1.8 in the Technical Discussion), with the direction n^ parallel to the normal direction of the plane.
    [Show full text]
  • Dirac Equation - Wikipedia
    Dirac equation - Wikipedia https://en.wikipedia.org/wiki/Dirac_equation Dirac equation From Wikipedia, the free encyclopedia In particle physics, the Dirac equation is a relativistic wave equation derived by British physicist Paul Dirac in 1928. In its free form, or including electromagnetic interactions, it 1 describes all spin-2 massive particles such as electrons and quarks for which parity is a symmetry. It is consistent with both the principles of quantum mechanics and the theory of special relativity,[1] and was the first theory to account fully for special relativity in the context of quantum mechanics. It was validated by accounting for the fine details of the hydrogen spectrum in a completely rigorous way. The equation also implied the existence of a new form of matter, antimatter, previously unsuspected and unobserved and which was experimentally confirmed several years later. It also provided a theoretical justification for the introduction of several component wave functions in Pauli's phenomenological theory of spin; the wave functions in the Dirac theory are vectors of four complex numbers (known as bispinors), two of which resemble the Pauli wavefunction in the non-relativistic limit, in contrast to the Schrödinger equation which described wave functions of only one complex value. Moreover, in the limit of zero mass, the Dirac equation reduces to the Weyl equation. Although Dirac did not at first fully appreciate the importance of his results, the entailed explanation of spin as a consequence of the union of quantum mechanics and relativity—and the eventual discovery of the positron—represents one of the great triumphs of theoretical physics.
    [Show full text]
  • Quantum Trajectories: Real Or Surreal?
    entropy Article Quantum Trajectories: Real or Surreal? Basil J. Hiley * and Peter Van Reeth * Department of Physics and Astronomy, University College London, Gower Street, London WC1E 6BT, UK * Correspondence: [email protected] (B.J.H.); [email protected] (P.V.R.) Received: 8 April 2018; Accepted: 2 May 2018; Published: 8 May 2018 Abstract: The claim of Kocsis et al. to have experimentally determined “photon trajectories” calls for a re-examination of the meaning of “quantum trajectories”. We will review the arguments that have been assumed to have established that a trajectory has no meaning in the context of quantum mechanics. We show that the conclusion that the Bohm trajectories should be called “surreal” because they are at “variance with the actual observed track” of a particle is wrong as it is based on a false argument. We also present the results of a numerical investigation of a double Stern-Gerlach experiment which shows clearly the role of the spin within the Bohm formalism and discuss situations where the appearance of the quantum potential is open to direct experimental exploration. Keywords: Stern-Gerlach; trajectories; spin 1. Introduction The recent claims to have observed “photon trajectories” [1–3] calls for a re-examination of what we precisely mean by a “particle trajectory” in the quantum domain. Mahler et al. [2] applied the Bohm approach [4] based on the non-relativistic Schrödinger equation to interpret their results, claiming their empirical evidence supported this approach producing “trajectories” remarkably similar to those presented in Philippidis, Dewdney and Hiley [5]. However, the Schrödinger equation cannot be applied to photons because photons have zero rest mass and are relativistic “particles” which must be treated differently.
    [Show full text]
  • 4 Nuclear Magnetic Resonance
    Chapter 4, page 1 4 Nuclear Magnetic Resonance Pieter Zeeman observed in 1896 the splitting of optical spectral lines in the field of an electromagnet. Since then, the splitting of energy levels proportional to an external magnetic field has been called the "Zeeman effect". The "Zeeman resonance effect" causes magnetic resonances which are classified under radio frequency spectroscopy (rf spectroscopy). In these resonances, the transitions between two branches of a single energy level split in an external magnetic field are measured in the megahertz and gigahertz range. In 1944, Jevgeni Konstantinovitch Savoiski discovered electron paramagnetic resonance. Shortly thereafter in 1945, nuclear magnetic resonance was demonstrated almost simultaneously in Boston by Edward Mills Purcell and in Stanford by Felix Bloch. Nuclear magnetic resonance was sometimes called nuclear induction or paramagnetic nuclear resonance. It is generally abbreviated to NMR. So as not to scare prospective patients in medicine, reference to the "nuclear" character of NMR is dropped and the magnetic resonance based imaging systems (scanner) found in hospitals are simply referred to as "magnetic resonance imaging" (MRI). 4.1 The Nuclear Resonance Effect Many atomic nuclei have spin, characterized by the nuclear spin quantum number I. The absolute value of the spin angular momentum is L =+h II(1). (4.01) The component in the direction of an applied field is Lz = Iz h ≡ m h. (4.02) The external field is usually defined along the z-direction. The magnetic quantum number is symbolized by Iz or m and can have 2I +1 values: Iz ≡ m = −I, −I+1, ..., I−1, I.
    [Show full text]
  • Perturbation Theory and Exact Solutions
    PERTURBATION THEORY AND EXACT SOLUTIONS by J J. LODDER R|nhtdnn Report 76~96 DISSIPATIVE MOTION PERTURBATION THEORY AND EXACT SOLUTIONS J J. LODOER ASSOCIATIE EURATOM-FOM Jun»»76 FOM-INST1TUUT VOOR PLASMAFYSICA RUNHUIZEN - JUTPHAAS - NEDERLAND DISSIPATIVE MOTION PERTURBATION THEORY AND EXACT SOLUTIONS by JJ LODDER R^nhuizen Report 76-95 Thisworkwat performed at part of th«r«Mvchprogmmncof thcHMCiattofiafrccmentof EnratoniOTd th« Stichting voor FundtmenteelOiutereoek der Matctk" (FOM) wtihnnmcWMppoft from the Nederhmdie Organiutic voor Zuiver Wetemchap- pcigk Onderzoek (ZWO) and Evntom It it abo pabHtfMd w a the* of Ac Univenrty of Utrecht CONTENTS page SUMMARY iii I. INTRODUCTION 1 II. GENERALIZED FUNCTIONS DEFINED ON DISCONTINUOUS TEST FUNC­ TIONS AND THEIR FOURIER, LAPLACE, AND HILBERT TRANSFORMS 1. Introduction 4 2. Discontinuous test functions 5 3. Differentiation 7 4. Powers of x. The partie finie 10 5. Fourier transforms 16 6. Laplace transforms 20 7. Hubert transforms 20 8. Dispersion relations 21 III. PERTURBATION THEORY 1. Introduction 24 2. Arbitrary potential, momentum coupling 24 3. Dissipative equation of motion 31 4. Expectation values 32 5. Matrix elements, transition probabilities 33 6. Harmonic oscillator 36 7. Classical mechanics and quantum corrections 36 8. Discussion of the Pu strength function 38 IV. EXACTLY SOLVABLE MODELS FOR DISSIPATIVE MOTION 1. Introduction 40 2. General quadratic Kami1tonians 41 3. Differential equations 46 4. Classical mechanics and quantum corrections 49 5. Equation of motion for observables 51 V. SPECIAL QUADRATIC HAMILTONIANS 1. Introduction 53 2. Hamiltcnians with coordinate coupling 53 3. Double coordinate coupled Hamiltonians 62 4. Symmetric Hamiltonians 63 i page VI. DISCUSSION 1. Introduction 66 ?.
    [Show full text]
  • Chapter 5 Angular Momentum and Spin
    Chapter 5 Angular Momentum and Spin I think you and Uhlenbeck have been very lucky to get your spinning electron published and talked about before Pauli heard of it. It appears that more than a year ago Kronig believed in the spinning electron and worked out something; the first person he showed it to was Pauli. Pauli rediculed the whole thing so much that the first person became also the last ... – Thompson (in a letter to Goudsmit) The first experiment that is often mentioned in the context of the electron’s spin and magnetic moment is the Einstein–de Haas experiment. It was designed to test Amp`ere’s idea that magnetism is caused by “molecular currents”. Such circular currents, while generating a magnetic field, would also contribute to the angular momentum of a ferromagnet. Therefore a change in the direction of the magnetization induced by an external field has to lead to a small rotation of the material in order to preserve the total angular momentum. For a quantitative understanding of the effect we consider a charged particle of mass m and charge q rotating with velocity v on a circle of radius r. Since the particle passes through its orbit v/(2πr) times per second the resulting current I = qv/(2πr), which encircles an area A = r2π, generates a magnetic dipole moment µ = IA/c, qv IA qv r2π qvr q q I = µ = = = = L = γL, γ = , (5.1) 2πr ⇒ c 2πr c 2c 2mc 2mc where L~ = m~r ~v is the angular momentum. Now the essential observation is that the × gyromagnetic ratio γ = µ/L is independent of the radius of the motion.
    [Show full text]
  • Type of Presentation: Poster IT-16-P-3287 Electron Vortex Beam
    Type of presentation: Poster IT-16-P-3287 Electron vortex beam diffraction via multislice solutions of the Pauli equation Edström A.1, Rusz J.1 1Department of Physics and Astronomy, Uppsala University Email of the presenting author: [email protected] Electron magnetic circular dichroism (EMCD) has gained plenty of attention as a possible route to high resolution measurements of, for example, magnetic properties of matter via electron microscopy. However, certain issues, such as low signal-to-noise ratio, have been problematic to the applicability. In recent years, electron vortex beams\cite{Uchida2010,Verbeeck2010}, i.e. electron beams which carry orbital angular momentum and are described by wavefunctions with a phase winding, have attracted interest as potential alternative way of measuring EMCD signals. Recent work has shown that vortex beams can be produced with a large orbital moment in the order of l = 100 [6, 7]. Huge orbital moments might introduce new effects from magnetic interactions such as spin-orbit coupling. The multislice method[2] provides a powerful computational tool for theoretical studies of electron microscopy. However, the method traditionally relies on the conventional Schrödinger equation which neglects relativistic effects such as spin-orbit coupling. Traditional multislice methods could therefore be inadequate in studying the diffraction of vortex beams with large orbital angular momentum. Relativistic multislice simulations have previously been done with a negligible difference to non-relativistic simulations[4], but vortex beams have not been considered in such work. In this work, we derive a new multislice approach based on the Pauli equation, Eq. 1, where q = −e is the electron charge, m = γm0 is the relativistically corrected mass, p = −i ∇ is the momentum operator, B = ∇ × A is the magnetic flux density while A is the vector potential and σ = (σx , σy , σz ) contains the Pauli matrices.
    [Show full text]
  • Measuring the Gyromagnetic Ratio Through Continuous Nuclear Magnetic Resonance Amy Catalano and Yash Aggarwal, Adlab, Boston University, Boston MA 02215
    1 Measuring the gyromagnetic ratio through continuous nuclear magnetic resonance Amy Catalano and Yash Aggarwal, Adlab, Boston University, Boston MA 02215 Abstract—In this experiment we use continuous wave nuclear magnetic resonance to measure the gyromagnetic ratios of Hydro- gen, Fluorine, the H nuclei in Water, and Water with Gadolinium. Our best result for Hydrogen was 25.4 ± 0.19 rad/Tm, compared to the previously reported value of 23 rad/Tm [1]. Our results for the gyromagnetic ratio of water and water with gadolinium contained a greater difference than expected. I. INTRODUCTION Nuclear magnetic resonance has to do with the absorption of radiation by a nucleus in a magnetic field. The energy Fig. 1. Schematic diagram for continuous nmr of a nucleus in a B-field can be written in terms of the gyromagnetic ratio, γ the magnetic field, B, and the frequency, f: III. DATA AND ERROR ANALYSIS ∆E = γB = 2πf (1) For Hydrogen, we plotted 16 points on a graph of frequency The gyromagnetic ratio is a fundamental nuclear constant (Mhz) and magnetic field (Gs). Using the numpy library in and is a different value for every nucleus. Every nucleus Python we calculated a linear best fit line of has a magnetic moment. The principle of nuclear magnetic resonance is to put the nucleus with an intrinsic nuclear spin y = mx + b (2) in a magnetic field and measure at which B-field and frequency where the spin of the nucleus flips. As we vary the magnetic field and the nucleus absorbs the signal from the spectrometer, the Gs m = 217:89 ± 1:95 (3) nucleus will go from a lower energy state to a higher energy MHz state.
    [Show full text]
  • Relativistic Quantum Mechanics 1
    Relativistic Quantum Mechanics 1 The aim of this chapter is to introduce a relativistic formalism which can be used to describe particles and their interactions. The emphasis 1.1 SpecialRelativity 1 is given to those elements of the formalism which can be carried on 1.2 One-particle states 7 to Relativistic Quantum Fields (RQF), which underpins the theoretical 1.3 The Klein–Gordon equation 9 framework of high energy particle physics. We begin with a brief summary of special relativity, concentrating on 1.4 The Diracequation 14 4-vectors and spinors. One-particle states and their Lorentz transforma- 1.5 Gaugesymmetry 30 tions follow, leading to the Klein–Gordon and the Dirac equations for Chaptersummary 36 probability amplitudes; i.e. Relativistic Quantum Mechanics (RQM). Readers who want to get to RQM quickly, without studying its foun- dation in special relativity can skip the first sections and start reading from the section 1.3. Intrinsic problems of RQM are discussed and a region of applicability of RQM is defined. Free particle wave functions are constructed and particle interactions are described using their probability currents. A gauge symmetry is introduced to derive a particle interaction with a classical gauge field. 1.1 Special Relativity Einstein’s special relativity is a necessary and fundamental part of any Albert Einstein 1879 - 1955 formalism of particle physics. We begin with its brief summary. For a full account, refer to specialized books, for example (1) or (2). The- ory oriented students with good mathematical background might want to consult books on groups and their representations, for example (3), followed by introductory books on RQM/RQF, for example (4).
    [Show full text]
  • Relativistic Inversion, Invariance and Inter-Action
    S S symmetry Article Relativistic Inversion, Invariance and Inter-Action Martin B. van der Mark †,‡ and John G. Williamson *,‡ The Quantum Bicycle Society, 12 Crossburn Terrace, Troon KA1 07HB, Scotland, UK; [email protected] * Correspondence: [email protected] † Formerly of Philips Research, 5656 AE Eindhoven, The Netherlands. ‡ These authors contributed equally to this work. Abstract: A general formula for inversion in a relativistic Clifford–Dirac algebra has been derived. Identifying the base elements of the algebra as those of space and time, the first order differential equations over all quantities proves to encompass the Maxwell equations, leads to a natural extension incorporating rest mass and spin, and allows an integration with relativistic quantum mechanics. Although the algebra is not a division algebra, it parallels reality well: where division is undefined turns out to correspond to physical limits, such as that of the light cone. The divisor corresponds to invariants of dynamical significance, such as the invariant interval, the general invariant quantities in electromagnetism, and the basis set of quantities in the Dirac equation. It is speculated that the apparent 3-dimensionality of nature arises from a beautiful symmetry between the three-vector algebra and each of four sets of three derived spaces in the full 4-dimensional algebra. It is conjectured that elements of inversion may play a role in the interaction of fields and matter. Keywords: invariants; inversion; division; non-division algebra; Dirac algebra; Clifford algebra; geometric algebra; special relativity; photon interaction Citation: van der Mark, M.B.; 1. Introduction Williamson, J.G. Relativistic Inversion, Invariance and Inter-Action.
    [Show full text]