Principles of Quantum Mechanics

Total Page:16

File Type:pdf, Size:1020Kb

Principles of Quantum Mechanics MATH3385/5385. Quantum Mechanics. Handout # 3: The Quantum Mechanics of Structureless Particles Principles of Quantum Mechanics Here we will study the principles of the QM of a structureless particle subject to a (possibly time-dependent) potential V (r,t). Structureless means that there are no other character- istics of the particle other than its position and momentum ( in contrast to e.g. particles with spin, a characteristic that we will introduce at a later stage). Our considerations con- cerning the wave packet description (see Handout # 2) has demonstrated that contrary to the classical perception of a particle (i.e. a point-like object which follows a geometric trajectory characterised at each point in time by its position and velocity) a fundamental uncertainty must be part of the new theory: in QM no longer can we measure both position as well as momentum of a particle simultaneously to arbitrary precision. This imprecision is not a fault of our apparatus or measuring device: it is fundamental to the theory. If we endeavour to measure the position, we lose track of the momentum and vice versa. We have to come to terms with this new and unfamiliar point of view on particles in the quantum theory. To capture both the particle aspect as well as the wave aspect, we need a description that achieves two things: • it allows for superposition of waves, thus incorporating the wave aspect, and • it has a built-in conservation law which accounts for the particle aspect (indeed: the particle is preserved; it is not supposed to vanish). We will now highlight the various ingredients of the theory. Wave function and probability interpretation The main idea behind quantum mechanics is to describe particles by a complex valued wave function ψ(r,t) which represents the spatial state of the particle. The word state is one of the key words in QM: instead of a particle having at each time t a precise value associated with it for the physical observables such as position and momentum, we say that the particle is prepared in a certain state described by the wave function at time t. What does the wave function represent physically? The interpretation (according to the Copenhagen school1) is the following: the absolute value P (r,t)= |ψ(r,t)|2 is a prob- ability density, meaning that 1This interpretation, which was promoted by N. Bohr and M. Born, has not remained undisputed. In fact, A. Einstein objected very strongly against the probability interpretation, saying that “God does not play dice”. Even today the discussions surrounding the probability interpretation of QM is carrying on. However, irrespective of the interpretation, the mathematical consistency of the theory is well-established. 1 |ψ(r,t)|2dr = the probability of finding a particle at time t in an infinitesimally small box of volume dr = dx dy dz located at r. This is quite a new and revolutionary idea: it introduces an amount of subjectiveness in the description of nature. The “observer” (i.e. the person or the device that performs a measurement), is no longer isolated from the object it observes (i.e. the particle), but influences in a fundamental way the state in which the particle is found. In fact, the act of performing a measurement actually prepares the particle in a given state, namely in the state where the measured quantity is relatively well-localised (the bulk of the wave packet lying within the error measurement from the value predicted). Thus, a measurement invokes an irreversible change in the state of the particle. However, moments after the measurement, the wave packet evolves and starts to spread as we lose track of the particle, and we need to perform another measurement to localise the particle once again (in either the momentum or position space). This procedure can be graphically represented as follows ↑ ↑ −→ measurement measurement time The quantum mechanical evolution of a particle: the dashed cloud represents the wave distribution of the particle in space, the fuzziness of which indicates the imprecision before measurement. A measurement localises the wave function, after which the wave packet spreads out, and another measurement is needed to pinch the wave packet again, etc. By the definition of probability, the total probability (i.e. the sum over all values of the relevant, i.e. measured, variables characterising the particle) must be equal to one. In the present case where the wave function is a function of the position we have: ∞ ∞ ∞ P (r,t) dr = 1 , dr = dx dy dz Z Z Z−∞ Z−∞ Z−∞ where the integral represents the integration over three-dimensional space. This expresses the fact that requires that the wave functions are square-integrable i.e. the integral |ψ(r,t)|2 dr < ∞ Z and we will work with wave functions that are normalised to one2. The matter in which function spaces ψ takes values is a rather subtle one and we leave this to later. 2 2 If the wave function is not normalised and we have R |ψ(r,t)| dr = C we can always redefine ψ by dividing by C1/2 such that it becomes normalised as long as ψ is square integrable. 2 The complexity of the wave function ψ allows us to explain the nontrivial interference patterns that were discussed in Handout # 2. Obviously, ψ itself is not a measurable quantity because in physics we can only measure real-valued objects. Thus, it seems that the phase factor exp(iθ) associated with the wave function ψ(r,t)= |ψ(r,t)|eiθ(r,t) is not of physical importance. This, however is an erroneous point of view because when iθ1 iθ2 one superimposes two waves ψ1 = |ψ1|e and ψ2 = |ψ2|e , the superimposed wave function does depend on the relative phase θ1 − θ2 of the waves. In fact, we have 2 2 iθ1 iθ2 |ψ1 + ψ2| = |ψ1|e + |ψ2|e 2 2 ∗ = |ψ1| + |ψ2| + 2Re( ψ1ψ2) 2 2 i(θ1−θ2) = |ψ1| + |ψ2| + 2|ψ1| |ψ2| Re e 2 2 = |ψ1| + |ψ2| + 2|ψ1| |ψ2| cos(θ1 − θ2) Note that the complexity of the wave function is essential in the appearance of the relative phase factor: we do not see it in the individual waves since |eiθ| = 1, but it appears in their interference. Evolution of the system and conservation law The main equation that governs the time evolution of the wave function is the Schr¨odinger equation (SE). We have already derived a simple case of the SE in Handout # 2 in the case of a one-dimenional free particle. In the general case of a particle moving in three- dimensional space subject to a potential V (r,t) (possibly time-dependent) we have the following equation: The Schr¨odinger Equation: ~ ∂ψ ~2 ∇2 ∇2 ∂2 ∂2 ∂2 i ∂t = − 2m ψ + V (r,t)ψ , = ∂x2 + ∂y2 + ∂z2 One important aspect of the SE is that it preserves probability. This is important, because the total probability should be a constant of the motion: once we know that the particle exists it will not disappear and thus we have certainty (i.e. probability equal to one) of finding the particle somewhere. This can easily seen as folows. Using the SE, and its complex conjugate, i.e. ∂ψ∗ ~2 −i~ = − ∇2ψ∗ + V (r,t)ψ∗ ∂t 2m we derive by multiplying the SE by ψ∗ and the above equation by ψ and then subtract: ∂ψ ∂ψ∗ ~2 i~ ψ∗ + ψ = − ψ∗∇2ψ − (∇2ψ∗)ψ ∂t ∂t 2m 3 where the terms containing the potential V cancel eachother. This can be rewritten as: ∂ ~2 i~ |ψ|2 = − ∇ · (ψ∗∇ψ − (∇ψ∗)ψ) ∂t 2m which leads to the conservation law: ∂ i~ P + ∇ · j = 0 , j = − (ψ∗∇ψ − (∇ψ∗)ψ) ∂t 2m and where as before P (r,t)= |ψ(r,t|2. The vector quantity j is called the probability cur- rent, whereas P is the probability density. The conservation law leads to the conservation of probability as follows: let us integrate the formula over a volume V (after which we let V grow infintely large so that we get an integral over the entire three-dimensional space). This leads to d P (r,t) dr = − ∇ · j dr = − j · dS dt ZV ZV ZS where we have used the divergence theorem (see MATH2360: vector calculus) to obtain an integral over the closed surface S that surrounds the volume V. Thus this integral depends only on the values of j (and hence of ψ and ψ∗) on the surface S. Since ψ is assumed to be quadratically integrable, and consequently ψ → 0 as |r|→∞ the integral on the right-hand side vanishes as we let the volume V grow. Thus, we obtain for the integrals over the entire space: d P (r,t) dr = 0 ⇒ P (r,t) dr = constant dt Z Z and consequently the total probability is preserved. This tells us that the Schr¨odinger equation is consistent with the normalisation of the total probability to one. Physical observables and expectation values In QM as in classical mechanics we want to be able to calculate the value of physical ob- servables. In classical mechanics this amounts to trying to predict the values of all relevant variables (position, momentum, etc.) by solving the equations of the motion, and nothing prevents us to do that in principle to arbitrary precision. In QM we can no longer do that: not all relevant variables can be simultaneously measured nor predicted to arbitrary precision: some physical quantities are incommensurable. If two observables (i.e. physical quantites that we can actually measure) are in principle simultaneously measurable then we call them commensurable and we can fully characterize the state of the particle by giving values to a large enough set of commensurable quantities.
Recommended publications
  • Rotation and Spin and Position Operators in Relativistic Gravity and Quantum Electrodynamics
    Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 4 December 2019 doi:10.20944/preprints201912.0044.v1 Peer-reviewed version available at Universe 2020, 6, 24; doi:10.3390/universe6020024 1 Review 2 Rotation and Spin and Position Operators in 3 Relativistic Gravity and Quantum Electrodynamics 4 R.F. O’Connell 5 Department of Physics and Astronomy, Louisiana State University; Baton Rouge, LA 70803-4001, USA 6 Correspondence: [email protected]; 225-578-6848 7 Abstract: First, we examine how spin is treated in special relativity and the necessity of introducing 8 spin supplementary conditions (SSC) and how they are related to the choice of a center-of-mass of a 9 spinning particle. Next, we discuss quantum electrodynamics and the Foldy-Wouthuysen 10 transformation which we note is a position operator identical to the Pryce-Newton-Wigner position 11 operator. The classical version of the operators are shown to be essential for the treatment of 12 classical relativistic particles in general relativity, of special interest being the case of binary 13 systems (black holes/neutron stars) which emit gravitational radiation. 14 Keywords: rotation; spin; position operators 15 16 I.Introduction 17 Rotation effects in relativistic systems involve many new concepts not needed in 18 non-relativistic classical physics. Some of these are quantum mechanical (where the emphasis is on 19 “spin”). Thus, our emphasis will be on quantum electrodynamics (QED) and both special and 20 general relativity. Also, as in [1] , we often use “spin” in the generic sense of meaning “internal” spin 21 in the case of an elementary particle and “rotation” in the case of an elementary particle.
    [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]
  • Chapter 5 ANGULAR MOMENTUM and ROTATIONS
    Chapter 5 ANGULAR MOMENTUM AND ROTATIONS In classical mechanics the total angular momentum L~ of an isolated system about any …xed point is conserved. The existence of a conserved vector L~ associated with such a system is itself a consequence of the fact that the associated Hamiltonian (or Lagrangian) is invariant under rotations, i.e., if the coordinates and momenta of the entire system are rotated “rigidly” about some point, the energy of the system is unchanged and, more importantly, is the same function of the dynamical variables as it was before the rotation. Such a circumstance would not apply, e.g., to a system lying in an externally imposed gravitational …eld pointing in some speci…c direction. Thus, the invariance of an isolated system under rotations ultimately arises from the fact that, in the absence of external …elds of this sort, space is isotropic; it behaves the same way in all directions. Not surprisingly, therefore, in quantum mechanics the individual Cartesian com- ponents Li of the total angular momentum operator L~ of an isolated system are also constants of the motion. The di¤erent components of L~ are not, however, compatible quantum observables. Indeed, as we will see the operators representing the components of angular momentum along di¤erent directions do not generally commute with one an- other. Thus, the vector operator L~ is not, strictly speaking, an observable, since it does not have a complete basis of eigenstates (which would have to be simultaneous eigenstates of all of its non-commuting components). This lack of commutivity often seems, at …rst encounter, as somewhat of a nuisance but, in fact, it intimately re‡ects the underlying structure of the three dimensional space in which we are immersed, and has its source in the fact that rotations in three dimensions about di¤erent axes do not commute with one another.
    [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]
  • The Position Operator Revisited Annales De L’I
    ANNALES DE L’I. H. P., SECTION A H. BACRY The position operator revisited Annales de l’I. H. P., section A, tome 49, no 2 (1988), p. 245-255 <http://www.numdam.org/item?id=AIHPA_1988__49_2_245_0> © Gauthier-Villars, 1988, tous droits réservés. L’accès aux archives de la revue « Annales de l’I. H. P., section A » implique l’accord avec les conditions générales d’utilisation (http://www.numdam. org/conditions). Toute utilisation commerciale ou impression systématique est constitutive d’une infraction pénale. Toute copie ou impression de ce fichier doit contenir la présente mention de copyright. Article numérisé dans le cadre du programme Numérisation de documents anciens mathématiques http://www.numdam.org/ .
    [Show full text]
  • 22.51 Course Notes, Chapter 9: Harmonic Oscillator
    9. Harmonic Oscillator 9.1 Harmonic Oscillator 9.1.1 Classical harmonic oscillator and h.o. model 9.1.2 Oscillator Hamiltonian: Position and momentum operators 9.1.3 Position representation 9.1.4 Heisenberg picture 9.1.5 Schr¨odinger picture 9.2 Uncertainty relationships 9.3 Coherent States 9.3.1 Expansion in terms of number states 9.3.2 Non-Orthogonality 9.3.3 Uncertainty relationships 9.3.4 X-representation 9.4 Phonons 9.4.1 Harmonic oscillator model for a crystal 9.4.2 Phonons as normal modes of the lattice vibration 9.4.3 Thermal energy density and Specific Heat 9.1 Harmonic Oscillator We have considered up to this moment only systems with a finite number of energy levels; we are now going to consider a system with an infinite number of energy levels: the quantum harmonic oscillator (h.o.). The quantum h.o. is a model that describes systems with a characteristic energy spectrum, given by a ladder of evenly spaced energy levels. The energy difference between two consecutive levels is ∆E. The number of levels is infinite, but there must exist a minimum energy, since the energy must always be positive. Given this spectrum, we expect the Hamiltonian will have the form 1 n = n + ~ω n , H | i 2 | i where each level in the ladder is identified by a number n. The name of the model is due to the analogy with characteristics of classical h.o., which we will review first. 9.1.1 Classical harmonic oscillator and h.o.
    [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]
  • Assignment 7: Solution
    Quantum Mechanics I / Quantum Theory I. 14 December, 2015 Assignment 7: Solution 1. In class, we derived that the ground state of the harmonic oscillator j0 is given in the position space as 1=4 m! 2 −m!x =2~ 0(x) = xj0 = e : π~ (a) This question deals with finding the state in the momentum basis. Let's start off by finding the position operatorx ^ in the momentum basis. Show that @ pjx^j = i pj : ~@p This is analogous to the expression of the momentum in the position basis, @ xjp^j = −i~ @x xj . You will use [Beck Eq. 10.56] expressing the momentum eigenstate in the position basis as well as the definition of the Fourier transform. @ (b) Sincea ^j0 = 0 and the position operator pjx^ = i~ @p , construct a differential equation to find the wavefunction in momentum space pj0 = f0(p). Solve this equation to find f0(p). Plot f0(p) in the momentum space. (c) Verify that your f0(p) = pj0 derived above is the Fourier transform of xj0 = 0(x). Answer (a) Lets find the position operator in momentum space pjx^j = pjx^1^j Z = dx pjx^jx xj Z = dx pjx x xj * x^jx = xjx : Now the normalized wavefunction of the momentum eigenstate in the position representation is 1 xjp = p eipx=~ Eq 10.56 in Beck 2π~ Z 1 −ipx= ) pjx^j = p dxe ~x xj : (1) 2π~ Due Date: December. 14, 2015, 05:00 pm 1 Quantum Mechanics I / Quantum Theory I. 14 December, 2015 Now @ −ix e−ipx=~ = e−ipx=~ @p ~ @ i e−ipx=~ = xe−ipx=~: ~ @p Inserting into Eq (1): i Z @ pjx^j = p ~ dx e−ipx=~ xj 2π~ @p Z @ 1 −ipx= = i~ p dxe ~ xj @p 2π~ @ = i pj ; ~@p where e(p) is the Fourier transform of (x).
    [Show full text]
  • Chapter 3 Mathematical Formalism of Quantum Mechanics
    Chapter 3 Mathematical Formalism of Quantum Mechanics 3.1 Hilbert Space To gain a deeper understanding of quantum mechanics, we will need a more solid math- ematical basis for our discussion. This we achieve by studying more thoroughly the structure of the space that underlies our physical objects, which as so often, is a vector space, the Hilbert space . Definition 3.1 A Hilbert space is a finite- or infinite-dimensional, linear vector space with scalar product in C. The Hilbert space provides, so to speak, the playground for our analysis. In analogy to a classical phase space, the elements of the vector space, the vectors, are our possible physical states. But the physical quantities we want to measure, the observables, are now operators acting on the vectors. As mentioned in Definition 3.1, Hilbert spaces can be finite- or infinite-dimensional, as opposed to classical phase spaces (which are always 6n-dimensional, where n is the particle number). We shall therefore investigate those two cases separately. 3.1.1 Finite–Dimensional Hilbert Spaces In finite dimensions the vectors of a Hilbert space, denoted by , and the corresponding scalar products differ from the standard Euklidean case onlyH by the choice of complex quantities C instead of real ones R. It means that for vectors x, y ∈ H x1 y1 . i i x = . , y = . , x , y , C , (3.1) ∈ xn yn 55 56 CHAPTER 3. MATHEMATICAL FORMALISM OF QUANTUM MECHANICS where represents the n-dimensional Hilbert space under consideration, the scalar prod- uct canH be written as 1 y n i .
    [Show full text]
  • Photon Wave Mechanics: a De Broglie - Bohm Approach
    PHOTON WAVE MECHANICS: A DE BROGLIE - BOHM APPROACH S. ESPOSITO Dipartimento di Scienze Fisiche, Universit`adi Napoli “Federico II” and Istituto Nazionale di Fisica Nucleare, Sezione di Napoli Mostra d’Oltremare Pad. 20, I-80125 Napoli Italy e-mail: [email protected] Abstract. We compare the de Broglie - Bohm theory for non-relativistic, scalar matter particles with the Majorana-R¨omer theory of electrodynamics, pointing out the impressive common pecu- liarities and the role of the spin in both theories. A novel insight into photon wave mechanics is envisaged. 1. Introduction Modern Quantum Mechanics was born with the observation of Heisenberg [1] that in atomic (and subatomic) systems there are directly observable quantities, such as emission frequencies, intensities and so on, as well as non directly observable quantities such as, for example, the position coordinates of an electron in an atom at a given time instant. The later fruitful developments of the quantum formalism was then devoted to connect observable quantities between them without the use of a model, differently to what happened in the framework of old quantum mechanics where specific geometrical and mechanical models were investigated to deduce the values of the observable quantities from a substantially non observable underlying structure. We now know that quantum phenomena are completely described by a complex- valued state function ψ satisfying the Schr¨odinger equation. The probabilistic in- terpretation of it was first suggested by Born [2] and, in the light of Heisenberg uncertainty principle, is a pillar of quantum mechanics itself. All the known experiments show that the probabilistic interpretation of the wave function is indeed the correct one (see any textbook on quantum mechanics, for 2 S.
    [Show full text]
  • An Approach to Quantum Mechanics
    An Approach to Quantum Mechanics Frank Rioux Department of Chemistry St. John=s University | College of St. Benedict The purpose of this tutorial is to introduce the basics of quantum mechanics using Dirac bracket notation while working in one dimension. Dirac notation is a succinct and powerful language for expressing quantum mechanical principles; restricting attention to one-dimensional examples reduces the possibility that mathematical complexity will stand in the way of understanding. A number of texts make extensive use of Dirac notation (1-5). In addition a summary of its main elements is provided at: http://www.users.csbsju.edu/~frioux/dirac/dirac.htm. Wave-particle duality is the essential concept of quantum mechanics. DeBroglie expressed this idea mathematically as λ = h/mv = h/p. On the left is the wave property λ, and on the right the particle property mv, momentum. The most general coordinate space wave function for a free particle with wavelength λ is the complex Euler function shown below. ⎛⎞⎛⎞⎛⎞x xx xiλπ==+exp⎜⎟⎜⎟⎜⎟ 2 cos 2 π i sin 2 π ⎝⎠⎝⎠⎝⎠λ λλ Equation 1 Feynman called this equation “the most remarkable formula in mathematics.” He referred to it as “our jewel.” And indeed it is, because when it is enriched with de Broglie’s relation it serves as the foundation of quantum mechanics. According to de Broglie=s hypothesis, a particle with a well-defined wavelength also has a well-defined momentum. Therefore, we can obtain the momentum wave function (unnormalized) of the particle in coordinate space by substituting the deBroglie relation into equation (1).
    [Show full text]