”Postulates of Quantum Mechanics” Are Often Stated in Textbooks. There

Total Page:16

File Type:pdf, Size:1020Kb

”Postulates of Quantum Mechanics” Are Often Stated in Textbooks. There Notes on the Formalism of Quantum Mechanics The ”postulates of quantum mechanics” are often stated in textbooks. There are two main properties of ”physics” upon which these postulates are based: 1)the prob- ability of measurements and 2) the interference properties of the probability amplitudes. These notes are meant to help students understand the ”physics” un- derlying the formalism of quantum mechanics. Throughout the discussion let A and B be two quantities that can be measured (observables). For simplicity, assume that each measurement yields values that are discrete, quantized. We will generalize later to measurements that yield continuous values. Label the possible values of observable A as αi and, and of observable B as βi. The integer i can take on a finite number or an infinite number of values. Probability of Measurements Experiments indicate that one can not determine the exact outcome of a measure- ment before the measurement is carried out. The only information about the outcome of a measurement are the probabilities for the various outcomes. If a measurement of A is guaranteed to yield the result αi, then we will call this state a ”proper state” of the observable A and label it as α >. If a measurement of A is carried out on a | i system in the state αi >, then the outcome is always αi. However, a system is not always in a proper state| that is guaranteed to give a particular result. In general, a system will be in a state that has a certain ”probability distribution” to be in the var- ious αi states. One can represent the state Ψ > of a system as a linear combination of the α > states: | | i Ψ >= a1 α1 > +a2 α2 > +... = ai αi > (1) | | | X | What is the meaning of the ai’? One might be lead to believe that the ai represent the probability that a measurement of A yields a result αi. However, experiments also indicate that interference effects occur with the probabilities. Thus, in order to have the probabilities include interference effects we are lead to interpret the ai as a probability amplitude: The probability for obtaining a value of αi if the observable A is measured is a 2. | i| Similar reasoning and notation can be used with any other observable. For exam- ple, for the observable B, we have Ψ >= b1 β1 > +b2 β2 > +... = bi βi > (2) | | | X | 1 where Ψ > represents the state of the system, and b 2 is the probability that a | | i| measurement of B will yield the value βi. The above expansion is a result of the probabilistic nature of the outcome of a measurement. An important quantity in quantum mechanics is the probability to obtain a value of βi if B is measured when the system is in the state αj >. We will label this 2 | probability as < βi αj > . We will discuss why this method of labeling is valuable in a moment.| The quantity| | < β α > is a probability amplitude, whose square is i| j a probability. In terms of experiments, this would mean that the observable A is measured with a result αj >. Then observable B is immediately measured. The probability of obtaining β| when B is measured is < β α > 2 i | i| j | Interference: Superposition of probability amplitudes An unusual property of systems where quantum effects are important is that probability amplitudes add, which leads to interference effects of probabilities. These interference effects are demonstrated in many experiments. In class for example, we see these effects from single and double slit scattering experiments with electrons and photons. In lecture we let photons pass through two slits in a metal foil and land on a screen behind the slits. The alternating light and dark pattern on the screen can be understood from the superposition of probability amplitudes. The probability amplitude to for the photon to land at a certain point on the screen is the sum of the probability amplitude for the photon to pass through the right slit plus the probability amplitude for the photon to pass through the left slit. The probability itself is the absolute square of the total probability amplitude. For the double slip experiment the probability ampliude for passing though each slit is proportional to eipd/h¯ , where p is the momentum of the photon and d is the total distance traveled from source to screen. The total probability amplitude is: P robamp eipd1/h¯ + eipd2/h¯ (3) ∝ The probability (or probability density) is just the absolute square of the probability amplitude or: P rob eipd1/h¯ + eipd2/h¯ 2 (4) ∝ | | p(d2 d1) 2 + 2 cos( − ) (5) ∝ h¯ p∆d cos2( ) (6) ∝ h¯ 2 π∆d P rob cos2( ) (7) ∝ λ where λ is the DeBroglie wave length of the photon or electron, and ∆d = d2 d1 is the difference in the distance the particle takes going through right versus the− left slit. There is no chance for the particle to land at points on the screen where the path difference ∆d is half a wavelength. Stated in words, the linear superposition of probability amplitudes means (where prob means probability, and probamp means probability amplitude): Prob(for the system to be measured in the state β >) is equal to probamp | i | (system is in state α1 >) < βi α1 > + probamp(system is in state α2 >) < βi α2 > + .... + probamp|(system is in| state α >) < β α > + .... 2 | | | j i| j | or expressed in summation notation: prob(system to be measured in the state βi >) equals sum over j of probamp(system is in state α >) < β α > 2 | | | j i| j | Linear Vector Space of the States of a System In order for a set of objects to be a vector space, any linear combination of elements must also be an element of the vector space. Consider the set of all possible states of a system. For this ”state space” to be a vector space, any linear combination of states of a system must also be a possible state of the system. If we consider as the ”basis” elements of the system space all the possible outcomes of a measurement, say measurment A, then Ψ >= ai αi > is a possible state of the system for all | 2 P | possible ai as long as ai = 1. The field of the vector space is the type of quantity P | | that the ai are. For the state space of quantum systems, the ai must be complex (at least). Since any linear combination of states is also a state of the system, the set of all states of a system form a vector space over the field complex numbers. For a set of objects to be an inner-product vector space, a scalar product must be defined. The scalar product must be an element of the field (i.e. a complex number). We label the scalar product between state Ψ1 > and Ψ2 > as < Ψ2 Ψ1 >. The scalar product must have the following ”linear” properties:| | | < Ψ ( Ψ1 > + Ψ2 >)=< Ψ Ψ1 > + < Ψ Ψ2 > (8) | | | | | < Ψ (c Ψ1 >)= c< Ψ Ψ1 > (9) | | | 3 and ∗ < Ψ2 Ψ1 >=< Ψ1 Ψ2 > (10) | | The first two properties are true if: the interpretation of the scalar product is a probability amplitude, and probability amplitudes obey the linear super- position principle. The last property is because the field is the complex numbers. Thus, because of the linear superposition property of probability amplitudes, the set of all possible states of a system form a linear vector space with the scalar product interpreted as a probability amplitude. Basis sets of the Vector Space A basis for a vector space is a set of elements such that any state can be expressed as a linear combination of the basis elements. Since the vector space elements are states of a physical system, it is natural to use as a basis the states of an observable corresponding to a measurement outcome. For example, one could use the states of the observable A, αi >, as a basis. This means that every state of the system can be expressed as combinations| of states for all possible values of measurement A: Ψ >= ai αi > (11) | X | where the sum goes over all possible outcomes of A. This is a complete basis when it comes to measurements of A, since very possible outcome is included. However, this basis might be incomplete when it comes to all possible measurements and outcomes. Our physical interpretation of the scalar product allows a nice interpretation of the a ’s. The scalar product < α α > is the probability amplitude that a measurement i i| j of A will result in a value αj when the system is in the state αi >. The probability must be zero unless i = j. If i = j, the probability is 1. Thus, we have: < α α >= δ (12) i| j ij In the terms of linear algebra, the basis αi > is an orthonormal basis. If we take the scalar product of the previous equation| with α >, one obtains | j < αj Ψ >= ai < αj αi >= aj (13) | X | Thus, aj is interpreted as the probability amplitude that the state Ψ > will be found in the state α >, and a 2 is the probability that a measurement will| yield the value j | j| αj. 4 One could use the proper states of any observable as a basis. For example, one could use the proper states of the observable B: β > which correspond to values of | i βi if B is measured. Ψ >= bi βi > (14) | X | For these basis states the following is also true: < βj βi >= δij, bj =< βj Ψ >, with b 2 being the probability that a measurement B on the| system yields β .| | j| j If one knows all the probability amplitudes < βi αj > for all i and j, then all measurement probabilities of the two observables A and| B can be determined.
Recommended publications
  • Path Integrals in Quantum Mechanics
    Path Integrals in Quantum Mechanics Dennis V. Perepelitsa MIT Department of Physics 70 Amherst Ave. Cambridge, MA 02142 Abstract We present the path integral formulation of quantum mechanics and demon- strate its equivalence to the Schr¨odinger picture. We apply the method to the free particle and quantum harmonic oscillator, investigate the Euclidean path integral, and discuss other applications. 1 Introduction A fundamental question in quantum mechanics is how does the state of a particle evolve with time? That is, the determination the time-evolution ψ(t) of some initial | i state ψ(t ) . Quantum mechanics is fully predictive [3] in the sense that initial | 0 i conditions and knowledge of the potential occupied by the particle is enough to fully specify the state of the particle for all future times.1 In the early twentieth century, Erwin Schr¨odinger derived an equation specifies how the instantaneous change in the wavefunction d ψ(t) depends on the system dt | i inhabited by the state in the form of the Hamiltonian. In this formulation, the eigenstates of the Hamiltonian play an important role, since their time-evolution is easy to calculate (i.e. they are stationary). A well-established method of solution, after the entire eigenspectrum of Hˆ is known, is to decompose the initial state into this eigenbasis, apply time evolution to each and then reassemble the eigenstates. That is, 1In the analysis below, we consider only the position of a particle, and not any other quantum property such as spin. 2 D.V. Perepelitsa n=∞ ψ(t) = exp [ iE t/~] n ψ(t ) n (1) | i − n h | 0 i| i n=0 X This (Hamiltonian) formulation works in many cases.
    [Show full text]
  • 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]
  • Path Probabilities for Consecutive Measurements, and Certain "Quantum Paradoxes"
    Path probabilities for consecutive measurements, and certain "quantum paradoxes" D. Sokolovski1;2 1 Departmento de Química-Física, Universidad del País Vasco, UPV/EHU, Leioa, Spain and 2 IKERBASQUE, Basque Foundation for Science, Maria Diaz de Haro 3, 48013, Bilbao, Spain (Dated: June 20, 2018) Abstract ABSTRACT: We consider a finite-dimensional quantum system, making a transition between known initial and final states. The outcomes of several accurate measurements, which could be made in the interim, define virtual paths, each endowed with a probability amplitude. If the measurements are actually made, the paths, which may now be called "real", acquire also the probabilities, related to the frequencies, with which a path is seen to be travelled in a series of identical trials. Different sets of measurements, made on the same system, can produce different, or incompatible, statistical ensembles, whose conflicting attributes may, although by no means should, appear "paradoxical". We describe in detail the ensembles, resulting from intermediate measurements of mutually commuting, or non-commuting, operators, in terms of the real paths produced. In the same manner, we analyse the Hardy’s and the "three box" paradoxes, the photon’s past in an interferometer, the "quantum Cheshire cat" experiment, as well as the closely related subject of "interaction-free measurements". It is shown that, in all these cases, inaccurate "weak measurements" produce no real paths, and yield only limited information about the virtual paths’ probability amplitudes. arXiv:1803.02303v3 [quant-ph] 19 Jun 2018 PACS numbers: Keywords: Quantum measurements, Feynman paths, quantum "paradoxes" 1 I. INTRODUCTION Recently, there has been significant interest in the properties of a pre-and post-selected quan- tum systems, and, in particular, in the description of such systems during the time between the preparation, and the arrival in the pre-determined final state (see, for example [1] and the Refs.
    [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]
  • Physical Quantum States and the Meaning of Probability Michel Paty
    Physical quantum states and the meaning of probability Michel Paty To cite this version: Michel Paty. Physical quantum states and the meaning of probability. Galavotti, Maria Carla, Suppes, Patrick and Costantini, Domenico. Stochastic Causality, CSLI Publications (Center for Studies on Language and Information), Stanford (Ca, USA), p. 235-255, 2001. halshs-00187887 HAL Id: halshs-00187887 https://halshs.archives-ouvertes.fr/halshs-00187887 Submitted on 15 Nov 2007 HAL is a multi-disciplinary open access L’archive ouverte pluridisciplinaire HAL, est archive for the deposit and dissemination of sci- destinée au dépôt et à la diffusion de documents entific research documents, whether they are pub- scientifiques de niveau recherche, publiés ou non, lished or not. The documents may come from émanant des établissements d’enseignement et de teaching and research institutions in France or recherche français ou étrangers, des laboratoires abroad, or from public or private research centers. publics ou privés. as Chapter 14, in Galavotti, Maria Carla, Suppes, Patrick and Costantini, Domenico, (eds.), Stochastic Causality, CSLI Publications (Center for Studies on Language and Information), Stanford (Ca, USA), 2001, p. 235-255. Physical quantum states and the meaning of probability* Michel Paty Ëquipe REHSEIS (UMR 7596), CNRS & Université Paris 7-Denis Diderot, 37 rue Jacob, F-75006 Paris, France. E-mail : [email protected] Abstract. We investigate epistemologically the meaning of probability as implied in quantum physics in connection with a proposed direct interpretation of the state function and of the related quantum theoretical quantities in terms of physical systems having physical properties, through an extension of meaning of the notion of physical quantity to complex mathematical expressions not reductible to simple numerical values.
    [Show full text]
  • Quantum Computing Joseph C
    Quantum Computing Joseph C. Bardin, Daniel Sank, Ofer Naaman, and Evan Jeffrey ©ISTOCKPHOTO.COM/SOLARSEVEN uring the past decade, quantum com- underway at many companies, including IBM [2], Mi- puting has grown from a field known crosoft [3], Google [4], [5], Alibaba [6], and Intel [7], mostly for generating scientific papers to name a few. The European Union [8], Australia [9], to one that is poised to reshape comput- China [10], Japan [11], Canada [12], Russia [13], and the ing as we know it [1]. Major industrial United States [14] are each funding large national re- Dresearch efforts in quantum computing are currently search initiatives focused on the quantum information Joseph C. Bardin ([email protected]) is with the University of Massachusetts Amherst and Google, Goleta, California. Daniel Sank ([email protected]), Ofer Naaman ([email protected]), and Evan Jeffrey ([email protected]) are with Google, Goleta, California. Digital Object Identifier 10.1109/MMM.2020.2993475 Date of current version: 8 July 2020 24 1527-3342/20©2020IEEE August 2020 Authorized licensed use limited to: University of Massachusetts Amherst. Downloaded on October 01,2020 at 19:47:20 UTC from IEEE Xplore. Restrictions apply. sciences. And, recently, tens of start-up companies have Quantum computing has grown from emerged with goals ranging from the development of software for use on quantum computers [15] to the im- a field known mostly for generating plementation of full-fledged quantum computers (e.g., scientific papers to one that is Rigetti [16], ION-Q [17], Psi-Quantum [18], and so on). poised to reshape computing as However, despite this rapid growth, because quantum computing as a field brings together many different we know it.
    [Show full text]
  • Uniting the Wave and the Particle in Quantum Mechanics
    Uniting the wave and the particle in quantum mechanics Peter Holland1 (final version published in Quantum Stud.: Math. Found., 5th October 2019) Abstract We present a unified field theory of wave and particle in quantum mechanics. This emerges from an investigation of three weaknesses in the de Broglie-Bohm theory: its reliance on the quantum probability formula to justify the particle guidance equation; its insouciance regarding the absence of reciprocal action of the particle on the guiding wavefunction; and its lack of a unified model to represent its inseparable components. Following the author’s previous work, these problems are examined within an analytical framework by requiring that the wave-particle composite exhibits no observable differences with a quantum system. This scheme is implemented by appealing to symmetries (global gauge and spacetime translations) and imposing equality of the corresponding conserved Noether densities (matter, energy and momentum) with their Schrödinger counterparts. In conjunction with the condition of time reversal covariance this implies the de Broglie-Bohm law for the particle where the quantum potential mediates the wave-particle interaction (we also show how the time reversal assumption may be replaced by a statistical condition). The method clarifies the nature of the composite’s mass, and its energy and momentum conservation laws. Our principal result is the unification of the Schrödinger equation and the de Broglie-Bohm law in a single inhomogeneous equation whose solution amalgamates the wavefunction and a singular soliton model of the particle in a unified spacetime field. The wavefunction suffers no reaction from the particle since it is the homogeneous part of the unified field to whose source the particle contributes via the quantum potential.
    [Show full text]
  • Assignment 2 Solutions 1. the General State of a Spin Half Particle
    PHYSICS 301 QUANTUM PHYSICS I (2007) Assignment 2 Solutions 1 1. The general state of a spin half particle with spin component S n = S · nˆ = 2 ~ can be shown to be given by 1 1 1 iφ 1 1 |S n = 2 ~i = cos( 2 θ)|S z = 2 ~i + e sin( 2 θ)|S z = − 2 ~i where nˆ is a unit vector nˆ = sin θ cos φ ˆi + sin θ sin φ jˆ + cos θ kˆ, with θ and φ the usual angles for spherical polar coordinates. 1 1 (a) Determine the expression for the the states |S x = 2 ~i and |S y = 2 ~i. 1 (b) Suppose that a measurement of S z is carried out on a particle in the state |S n = 2 ~i. 1 What is the probability that the measurement yields each of ± 2 ~? 1 (c) Determine the expression for the state for which S n = − 2 ~. 1 (d) Show that the pair of states |S n = ± 2 ~i are orthonormal. SOLUTION 1 (a) For the state |S x = 2 ~i, the unit vector nˆ must be pointing in the direction of the X axis, i.e. θ = π/2, φ = 0, so that 1 1 1 1 |S x = ~i = √ |S z = ~i + |S z = − ~i 2 2 2 2 1 For the state |S y = 2 ~i, the unit vector nˆ must be pointed in the direction of the Y axis, i.e. θ = π/2 and φ = π/2. Thus 1 1 1 1 |S y = ~i = √ |S z = ~i + i|S z = − ~i 2 2 2 2 1 1 2 (b) The probabilities will be given by |hS z = ± 2 ~|S n = 2 ~i| .
    [Show full text]
  • Lecture 2: Basics of Quantum Mechanics
    Lecture 2: Basics of Quantum Mechanics Scribed by: Vitaly Feldman Department of Mathematics, MIT September 9, 2003 In this lecture we will cover the basics of Quantum Mechanics which are required to under­ stand the process of quantum computation. To simplify the discussion we assume that all the Hilbert spaces mentioned below are finite­dimensional. The process of quantum computation can be abstracted via the following four postulates. Postulate 1. Associated to any isolated physical system is a complex vector space with inner product (that is, a Hilbert space) known as the state space of the system. The state of the system is completely described by a unit vector in this space. Qubits were the example of such system that we saw in the previous lecture. In its physical realization as a polarization of a photon we have two basis vectors: |�� and |↔� representing the vertical and the horizontal polarizations respectively. In this basis vector polarized at angle θ can be expressed as cos θ |↔� − sin θ |��. An important property of a quantum system is that multiplying a quantum state by a unit complex factor (eiθ ) yields the same complex state. Therefore eiθ |�� and |�� represent essentially the same state. Notation 1. State χ is denoted by |χ� (often called a ket) is a column vector, e.g., ⎛ ⎞ √1/2 ⎝ i 3/2 ⎠ 0 † |χ� = �χ| (often √called a bra) denotes a conjugate transpose of |χ�. In the previous example we would get (1/2, −i 3/2, 0). It is easy to verify that �χ| χ� = 1 and �x|y� ≤ 1.
    [Show full text]
  • Path Integrals in Quantum Mechanics
    Path Integrals in Quantum Mechanics Emma Wikberg Project work, 4p Department of Physics Stockholm University 23rd March 2006 Abstract The method of Path Integrals (PI’s) was developed by Richard Feynman in the 1940’s. It offers an alternate way to look at quantum mechanics (QM), which is equivalent to the Schrödinger formulation. As will be seen in this project work, many "elementary" problems are much more difficult to solve using path integrals than ordinary quantum mechanics. The benefits of path integrals tend to appear more clearly while using quantum field theory (QFT) and perturbation theory. However, one big advantage of Feynman’s formulation is a more intuitive way to interpret the basic equations than in ordinary quantum mechanics. Here we give a basic introduction to the path integral formulation, start- ing from the well known quantum mechanics as formulated by Schrödinger. We show that the two formulations are equivalent and discuss the quantum mechanical interpretations of the theory, as well as the classical limit. We also perform some explicit calculations by solving the free particle and the harmonic oscillator problems using path integrals. The energy eigenvalues of the harmonic oscillator is found by exploiting the connection between path integrals, statistical mechanics and imaginary time. Contents 1 Introduction and Outline 2 1.1 Introduction . 2 1.2 Outline . 2 2 Path Integrals from ordinary Quantum Mechanics 4 2.1 The Schrödinger equation and time evolution . 4 2.2 The propagator . 6 3 Equivalence to the Schrödinger Equation 8 3.1 From the Schrödinger equation to PI’s . 8 3.2 From PI’s to the Schrödinger equation .
    [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]