12.07.2015 Views

Updated Lecture Notes (9/23/2013) - University of Illinois at Urbana ...

Updated Lecture Notes (9/23/2013) - University of Illinois at Urbana ...

Updated Lecture Notes (9/23/2013) - University of Illinois at Urbana ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Contentsv15 Quantum Inform<strong>at</strong>ion and Quantum Interpret<strong>at</strong>ion 20315.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20315.2 Quantum Cryptography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20315.2.1 No-cloning Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20315.2.2 Entangled St<strong>at</strong>es . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20415.2.3 A Simple Quantum Encryption Algorithm . . . . . . . . . . . . . . . 20615.3 Quantum Computing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20815.3.1 Quantum Bits (Qubits) . . . . . . . . . . . . . . . . . . . . . . . . . . 20815.3.2 Quantum G<strong>at</strong>es . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20915.3.3 Quantum Computing Algorithms . . . . . . . . . . . . . . . . . . . . . 21015.4 Quantum Teleport<strong>at</strong>ion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21215.5 Interpret<strong>at</strong>ion <strong>of</strong> Quantum Mechanics . . . . . . . . . . . . . . . . . . . . . . 21415.6 EPR Paradox . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21515.7 Bell’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21615.7.1 Prediction by Quantum Mechanics . . . . . . . . . . . . . . . . . . . . 21715.7.2 Prediction by Hidden Variable Theory . . . . . . . . . . . . . . . . . . 21815.8 A Final Word on Quantum Parallelism . . . . . . . . . . . . . . . . . . . . . . 221A Gaussian Wave Packet 225A.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 225A.2 Deriv<strong>at</strong>ion from the Wave Equ<strong>at</strong>ion . . . . . . . . . . . . . . . . . . . . . . . 226A.3 Physical Interpret<strong>at</strong>ion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227A.4 Stability <strong>of</strong> the Plane Wave Solution . . . . . . . . . . . . . . . . . . . . . . . 229B Gener<strong>at</strong>ors <strong>of</strong> Transl<strong>at</strong>or and Rot<strong>at</strong>ion <strong>23</strong>1B.1 Infinitesimal Transl<strong>at</strong>ion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . <strong>23</strong>1B.2 Infinitesimal Rot<strong>at</strong>ion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . <strong>23</strong>2B.3 Deriv<strong>at</strong>ion <strong>of</strong> Commut<strong>at</strong>ion Rel<strong>at</strong>ions . . . . . . . . . . . . . . . . . . . . . . <strong>23</strong>3C Quantum St<strong>at</strong>istical Mechanics <strong>23</strong>5C.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . <strong>23</strong>5C.1.1 Distinguishable Particles . . . . . . . . . . . . . . . . . . . . . . . . . . <strong>23</strong>5C.1.2 Identical Fermions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . <strong>23</strong>7C.1.3 Identical Bosons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . <strong>23</strong>7C.2 Most Probable Configur<strong>at</strong>ion . . . . . . . . . . . . . . . . . . . . . . . . . . . <strong>23</strong>7C.2.1 Distinguishable Particles . . . . . . . . . . . . . . . . . . . . . . . . . . <strong>23</strong>8C.2.2 Identical Fermions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . <strong>23</strong>8C.2.3 Identical Bosons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . <strong>23</strong>9C.3 The Meaning <strong>of</strong> α and β . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 240D Photon Polariz<strong>at</strong>ion 243D.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243D.2 2D Quantum Harmonic Oscill<strong>at</strong>or . . . . . . . . . . . . . . . . . . . . . . . . 243D.3 Single–Photon St<strong>at</strong>e . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 245


viQuantum Mechanics Made Simple


PrefaceThis set <strong>of</strong> supplementary lecture notes is the outgrowth <strong>of</strong> a course I taught, ECE 487,Quantum Electronics, <strong>at</strong> ECE Department, <strong>University</strong> <strong>of</strong> <strong>Illinois</strong> <strong>at</strong> <strong>Urbana</strong>-Champaign. Itwas intended to teach quantum mechanics to undergradu<strong>at</strong>e students as well as gradu<strong>at</strong>estudents. The primary text book for this course is Quantum Mechanics for Scientists andEngineers by D.A.B. Miller. I have learned a gre<strong>at</strong> deal by poring over Miller’s book. Butwhere I feel the book to be incomplete, I supplement them with my lecture notes. I try toreach into first principles as much as I could with these lecture notes. The only backgroundneeded for reading these notes is a background in undergradu<strong>at</strong>e wave physics, and linearalgebra.I would still recommend using Miller’s book as the primary text book for such a course,and use these notes as supplementary to teach this topic to undergradu<strong>at</strong>es.Weng Cho CHEWSeptember <strong>23</strong>, <strong>2013</strong>AcknowledgementsI like to thank Erhan Kudeki who encouraged me to teach this course, and for havingmany interesting discussions during its teaching. I acknowledge interesting discussions withFuchun ZHANG, Guanhua CHEN, Jian WANG, and Hong GUO (McGill U) <strong>at</strong> Hong KongU.I also like to thank many <strong>of</strong> my students and researchers who have helped type the notesand pro<strong>of</strong>read them. They are Phil Atkins, F<strong>at</strong>ih Erden, Tian XIA, Palash Sarker, JunHUANG, Qi DAI, Zuhui MA, Yumao WU, Min TANG, Y<strong>at</strong>-Hei LO, Bo ZHU, and ShengSUN.vii


viiiQuantum Mechanics Made Simple


Chapter 1Introduction1.1 IntroductionQuantum mechanics is an important intellectual achievement <strong>of</strong> the 20th century. It is one<strong>of</strong> the more sophistic<strong>at</strong>ed fields in physics th<strong>at</strong> has affected our understanding <strong>of</strong> nano-meterlength scale systems important for chemistry, m<strong>at</strong>erials, optics, electronics, and quantuminform<strong>at</strong>ion. The existence <strong>of</strong> orbitals and energy levels in <strong>at</strong>oms can only be explained byquantum mechanics. Quantum mechanics can explain the behaviors <strong>of</strong> insul<strong>at</strong>ors, conductors,semi-conductors, and giant magneto-resistance. It can explain the quantiz<strong>at</strong>ion <strong>of</strong> light andits particle n<strong>at</strong>ure in addition to its wave n<strong>at</strong>ure (known as particle-wave duality). Quantummechanics can also explain the radi<strong>at</strong>ion <strong>of</strong> hot body or black body, and its change <strong>of</strong> colorwith respect to temper<strong>at</strong>ure. It explains the presence <strong>of</strong> holes and the transport <strong>of</strong> holes andelectrons in electronic devices.Quantum mechanics has played an important role in photonics, quantum electronics, nanoandmicro-electronics, nano- and quantum optics, quantum computing, quantum communic<strong>at</strong>ionand crytography, solar and thermo-electricity, nano-electromechacnical systems, etc.Many emerging technologies require the understanding <strong>of</strong> quantum mechanics; and hence, itis important th<strong>at</strong> scientists and engineers understand quantum mechanics better.In nano-technologies due to the recent advent <strong>of</strong> nano-fabric<strong>at</strong>ion techniques. Consequently,nano-meter size systems are more common place. In electronics, as transistor devicesbecome smaller, how the electrons behave in the device is quite different from when the devicesare bigger: nano-electronic transport is quite different from micro-electronic transport.The quantiz<strong>at</strong>ion <strong>of</strong> electromagnetic field is important in the area <strong>of</strong> nano-optics andquantum optics. It explains how photons interact with <strong>at</strong>omic systems or m<strong>at</strong>erials. It alsoallows the use <strong>of</strong> electromagnetic or optical field to carry quantum inform<strong>at</strong>ion. Quantummechanics is certainly giving rise to interest in quantum inform<strong>at</strong>ion, quantum communic<strong>at</strong>ion,quantum cryptography, and quantum computing. Moreover, quantum mechanics is alsoneeded to understand the interaction <strong>of</strong> photons with m<strong>at</strong>erials in solar cells, as well as manytopics in m<strong>at</strong>erial science.When two objects are placed close together, they experience a force called the Casimir forceth<strong>at</strong> can only be explained by quantum mechanics. This is important for the understanding1


2 Quantum Mechanics Made Simple<strong>of</strong> micro/nano-electromechanical systems (M/NEMS). Moreover, the understanding <strong>of</strong> spinsis important in spintronics, another emerging technology where giant magneto-resistance,tunneling magneto-resistance, and spin transfer torque are being used. It is obvious th<strong>at</strong> therichness <strong>of</strong> quantum physics will gre<strong>at</strong>ly affect the future gener<strong>at</strong>ion technologies in manyaspects.1.2 Quantum Mechanics is BizarreThe development <strong>of</strong> quantum mechanicsis a gre<strong>at</strong> intellectual achievement, but <strong>at</strong> the sametime, it is bizarre. The reason is th<strong>at</strong> quantum mechanics is quite different from classicalphysics. The development <strong>of</strong> quantum mechanics is likened to w<strong>at</strong>ching two players havinga game <strong>of</strong> chess, but the observers have not a clue as to wh<strong>at</strong> the rules <strong>of</strong> the game are. Byobserv<strong>at</strong>ions, and conjectures, finally the rules <strong>of</strong> the game are outlined. Often, equ<strong>at</strong>ions areconjectured like conjurors pulling tricks out <strong>of</strong> a h<strong>at</strong> to m<strong>at</strong>ch experimental observ<strong>at</strong>ions. Itis the interpret<strong>at</strong>ions <strong>of</strong> these equ<strong>at</strong>ions th<strong>at</strong> can be quite bizarre.Quantum mechanics equ<strong>at</strong>ions were postul<strong>at</strong>ed to explain experimental observ<strong>at</strong>ions, butthe deeper meanings <strong>of</strong> the equ<strong>at</strong>ions <strong>of</strong>ten confused even the most gifted. Even thoughEinstein received the Nobel prize for his work on the photo-electric effect th<strong>at</strong> confirmedth<strong>at</strong> light energy is quantized, he himself was not totally <strong>at</strong> ease with the development <strong>of</strong>quantum mechanicsas charted by the younger physicists. He was never comfortable with theprobabilistic interpret<strong>at</strong>ion <strong>of</strong> quantum mechanics by Born and the Heisenberg uncertaintyprinciple: “God doesn’t play dice,” was his st<strong>at</strong>ement assailing the probabilistic interpret<strong>at</strong>ion.He proposed “hidden variables” to explain the random n<strong>at</strong>ure <strong>of</strong> many experimentalobserv<strong>at</strong>ions. He was thought <strong>of</strong> as the “old fool” by the younger physicists during his time.Schrödinger came up with the bizarre “Schrödinger c<strong>at</strong> paradox” th<strong>at</strong> showed the struggleth<strong>at</strong> physicists had with quantum mechanics’s interpret<strong>at</strong>ion. But with today’s understanding<strong>of</strong> quantum mechanics, the paradox is a thing <strong>of</strong> yesteryear.The l<strong>at</strong>est twist to the interpret<strong>at</strong>ion in quantum mechanics is the parallel universe viewth<strong>at</strong> explains the multitude <strong>of</strong> outcomes <strong>of</strong> the prediction <strong>of</strong> quantum mechanics. All outcomesare possible, but with each outcome occurring in different universes th<strong>at</strong> exist in parallel withrespect to each other ...1.3 The Wave N<strong>at</strong>ure <strong>of</strong> a Particle–Wave Particle DualityThe quantized n<strong>at</strong>ure <strong>of</strong> the energy <strong>of</strong> light was first proposed by Planck in 1900 to successfullyexplain the black body radi<strong>at</strong>ion. Einstein’s explan<strong>at</strong>ion <strong>of</strong> the photoelectric effect furtherasserts the quantized n<strong>at</strong>ure <strong>of</strong> light, or light as a photon. 1 However, it is well known th<strong>at</strong>1 In the photoelectric effect, it was observed th<strong>at</strong> electrons can be knocked <strong>of</strong>f a piece <strong>of</strong> metal only if thelight exceeded a certain frequency. Above th<strong>at</strong> frequency, the electron gained some kinetic energy proportionalto the excess frequency. Einstein then concluded th<strong>at</strong> a packet <strong>of</strong> energy was associ<strong>at</strong>ed with a photon th<strong>at</strong>is proportional to its frequency.


Introduction 3light is a wave since it can be shown to interfere as waves in the Newton ring experiment asfar back as 1717.The wave n<strong>at</strong>ure <strong>of</strong> an electron is revealed by the fact th<strong>at</strong> when electrons pass througha crystal, they produce a diffraction p<strong>at</strong>tern. Th<strong>at</strong> can only be explained by the wave n<strong>at</strong>ure<strong>of</strong> an electron. This experiment was done by Davisson and Germer in 1927. 2 De Brogliehypothesized th<strong>at</strong> the wavelength <strong>of</strong> an electron, when it behaves like a wave, isλ = h p(1.3.1)where h is the Planck’s constant, p is the electron momentum, 3 andWhen an electron manifests as a wave, it is described bywhere k = 2π/λ. Such a wave is a solution to 4A generaliz<strong>at</strong>ion <strong>of</strong> this to three dimensions yieldsWe can definewhere = h/(2π). 5 Consequently, we arrive <strong>at</strong> an equ<strong>at</strong>ionh ≈ 6.626 × 10 −34 Joule · second (1.3.2)ψ(z) ∝ exp(ikz) (1.3.3)∂ 2∂z 2 ψ = −k2 ψ (1.3.4)∇ 2 ψ(r) = −k 2 ψ(r) (1.3.5)p = k (1.3.6)where− 2∇ 2 ψ(r) =p2ψ(r) (1.3.7)2m 0 2m 0m 0 ≈ 9.11 × 10 −31 kg (1.3.8)The expression p 2 /(2m 0 ) is the kinetic energy <strong>of</strong> an electron.considered an energy conserv<strong>at</strong>ion equ<strong>at</strong>ion.Hence, the above can be2 Young’s double slit experiment was conducted in early 1800s to demonstr<strong>at</strong>e the wave n<strong>at</strong>ure <strong>of</strong> photons.Due to the short wavelengths <strong>of</strong> electrons, it was not demonstr<strong>at</strong>ed it until 2002 by Jonsson. But it has beenused as a thought experiment by Feynman in his lectures.3 Typical electron wavelengths are <strong>of</strong> the order <strong>of</strong> nanometers. Compared to 400 nm <strong>of</strong> wavelength <strong>of</strong> bluelight, they are much smaller. Energetic electrons can have even smaller wavelengths. Hence, electron wavescan be used to make electron microscope whose resolution is much higher than optical microscope.4 The wavefunction can be thought <strong>of</strong> as a “halo” th<strong>at</strong> an electron carries th<strong>at</strong> determine its underlyingphysical properties and how it interact with other systems.5 This is also called Dirac constant sometimes.


4 Quantum Mechanics Made SimpleThe Schrödinger equ<strong>at</strong>ion is motiv<strong>at</strong>ed by further energy balance th<strong>at</strong> total energy is equalto the sum <strong>of</strong> potential energy and kinetic energy. Defining the potential energy to be V (r),the energy balance equ<strong>at</strong>ion becomes][− 2∇ 2 + V (r) ψ(r) = Eψ(r) (1.3.9)2m 0where E is the total energy <strong>of</strong> the system. The above is the time-independent Schrödingerequ<strong>at</strong>ion. The ad hoc manner <strong>at</strong> which the above equ<strong>at</strong>ion is arrived <strong>at</strong> usually bothersa beginner in the field. However, it predicts many experimental outcomes. It particular,it predicts the energy levels and orbitals <strong>of</strong> a trapped electron in a hydrogen <strong>at</strong>om withresounding success.One can further modify the above equ<strong>at</strong>ion in an ad hoc manner by noticing th<strong>at</strong> otherexperimental finding shows th<strong>at</strong> the energy <strong>of</strong> a photon is E = ω. Hence, if we lettheni ∂ Ψ(r, t) = EΨ(r, t) (1.3.10)∂tΨ(r, t) = e −iωt ψ(r, t) (1.3.11)Then we arrive <strong>at</strong> the time-dependent Schrödinger equ<strong>at</strong>ion:][− 2∇ 2 + V (r) ψ(r, t) = i ∂ ψ(r, t) (1.3.12)2m 0 ∂tAnother disquieting fact about the above equ<strong>at</strong>ion is th<strong>at</strong> it is written in terms <strong>of</strong> complexfunctions and numbers. In our prior experience with classical laws, they can all be written inreal functions and numbers. We will l<strong>at</strong>er learn the reason for this.Mind you, in the above, the frequency is not unique. We know th<strong>at</strong> in classical physics,the potential V is not unique, and we can add a constant to it, and yet, the physics <strong>of</strong> theproblem does not change. So, we can add a constant to both sides <strong>of</strong> the time-independentSchrödinger equ<strong>at</strong>ion (1.3.10), and yet, the physics should not change. Then the total E onthe right-hand side would change, and th<strong>at</strong> would change the frequency we have arrived <strong>at</strong>in the time-dependent Schrödinger equ<strong>at</strong>ion. We will explain how to resolve this dilemmal<strong>at</strong>er on. Just like potentials, in quantum mechanics, it is the difference <strong>of</strong> frequencies th<strong>at</strong>m<strong>at</strong>ters in the final comparison with experiments, not the absolute frequencies.The setting during which Schrödinger equ<strong>at</strong>ion was postul<strong>at</strong>ed was replete with knowledge<strong>of</strong> classical mechanics. It will be prudent to review some classical mechanics knowledge next.


Chapter 2Classical Mechanics and SomeM<strong>at</strong>hem<strong>at</strong>ical Preliminaries2.1 IntroductionQuantum mechanics cannot be derived from classical mechanics, but classical mechanicscan inspire quantum mechanics. Quantum mechanics is richer and more sophistic<strong>at</strong>ed thanclassical mechanics. Quantum mechanics was developed during the period when physicistshad rich knowledge <strong>of</strong> classical mechanics. In order to better understand how quantummechanics was developed in this environment, it is better to understand some fundamentalconcepts in classical mechanics. Classical mechanics can be considered as a special case <strong>of</strong>quantum mechanics. We will review some classical mechanics concepts here.In classical mechanics, a particle moving in the presence <strong>of</strong> potential 1 V (q) will experiencea force given bydV (q)F (q) = −dq(2.1.1)where q represents the coordin<strong>at</strong>e or the position <strong>of</strong> the particle. Hence, the particle can bedescribed by the equ<strong>at</strong>ions <strong>of</strong> motiondpdt(q)= F (q) = −dV ,dqdqdt= p/m (2.1.2)For example, when a particle is <strong>at</strong>tached to a spring and moves along a frictionless surface,the force the particle experiences is F (q) = −kq where k is the spring constant. Then theequ<strong>at</strong>ions <strong>of</strong> motion <strong>of</strong> this particle aredpdt = ṗ = −kq,dqdt= ˙q = p/m (2.1.3)1 The potential here refers to potential energy.5


6 Quantum Mechanics Made SimpleV(q)qqFigure 2.1: The left side shows a potential well in which a particle can be trapped. The rightside shows a particle <strong>at</strong>tached to a spring. The particle is subject to the force due to thespring, but it can also be described by the force due to a potential well.Given p and q <strong>at</strong> some initial time t 0 , one can integr<strong>at</strong>e (2.1.2) or (2.1.3) to obtain p and q forall l<strong>at</strong>er time. A numerical analyst can think <strong>of</strong> th<strong>at</strong> (2.1.2) or (2.1.3) can be solved by thefinite difference method, where time-stepping can be used to find p and q for all l<strong>at</strong>er times.For instance, we can write the equ<strong>at</strong>ions <strong>of</strong> motion more compactly asdudt= f(u) (2.1.4)where u = [p, q] t , and f is a general vector function <strong>of</strong> u. It can be nonlinear or linear; in theevent if it is linear, then f(u) = A · u.Using finite difference approxim<strong>at</strong>ion, we can rewrite the above asu(t + ∆t) − u(t) . = ∆tf(u(t)),u(t + ∆t) . = ∆tf(u(t)) + u(t) (2.1.5)The above can be used for time marching to derive the future values <strong>of</strong> u from past values.The above equ<strong>at</strong>ions <strong>of</strong> motion are essentially derived using Newton’s law. However, thereexist other methods <strong>of</strong> deriving these equ<strong>at</strong>ions <strong>of</strong> motion. Notice th<strong>at</strong> only two variables pand q are sufficient to describe the st<strong>at</strong>e <strong>of</strong> a particle.2.2 Lagrangian Formul<strong>at</strong>ionAnother way to derive the equ<strong>at</strong>ions <strong>of</strong> motion for classical mechanics is via the use <strong>of</strong> theLagrangian and the principle <strong>of</strong> least action. A Lagrangian is usually defined as the differencebetween the kinetic energy and the potential energy, i.e.,L( ˙q, q) = T − V (2.2.1)where ˙q is the velocity. For a fixed t, q and ˙q are independent variables, since ˙q cannot bederived from q if it is only known <strong>at</strong> one given t. The equ<strong>at</strong>ions <strong>of</strong> motion are derived from theprinciple <strong>of</strong> least action which says th<strong>at</strong> q(t) th<strong>at</strong> s<strong>at</strong>isfies the equ<strong>at</strong>ions <strong>of</strong> motion betweentwo times t 1 and t 2 should minimize the action integralS =∫ t2t 1L( ˙q(t), q(t))dt (2.2.2)


Classical Mechanics and Some M<strong>at</strong>hem<strong>at</strong>ical Preliminaries 7Assuming th<strong>at</strong> q(t 1 ) and q(t 2 ) are fixed, then the function q(t) between t 1 and t 2 shouldminimize S, the action. In other words, a first order perturb<strong>at</strong>ion in q from the optimalanswer th<strong>at</strong> minimizes S should give rise to second order error in S. Hence, taking the firstvari<strong>at</strong>ion <strong>of</strong> (2.2.2), we have∫ t2∫ t2δS = δ L( ˙q, q)dt =t 1=t 1∫ t2∫ t2L( ˙q + δ ˙q, q + δq)dt −t 1δL( ˙q, q)dt =t 1L( ˙q, q)dt∫ t2(δ ˙q ∂L )∂L+ δq dt = 0 (2.2.3)t 1∂ ˙q ∂qIn order to take the vari<strong>at</strong>ion into the integrand, we have to assume th<strong>at</strong> δL( ˙q, q) is takenwith constant time. At constant time, ˙q and q are independent variables; hence, the partialderiv<strong>at</strong>ives in the next equality above follow. Using integr<strong>at</strong>ion by parts on the first term, wehaveδS = δq ∂L∂ ˙q ∣=∫ t2t 1δqt 2∫ t2−t 1 t 1[− d dt( ∂L∂ ˙qδq d dt( ) ∫ ∂Lt2dt +∂ ˙q)dt + ∂L∂qt 1δq ∂L∂q dt]dt = 0 (2.2.4)The first term vanishes because δq(t 1 ) = δq(t 2 ) = 0 because q(t 1 ) and q(t 2 ) are fixed. Sinceδq(t) is arbitrary between t 1 and t 2 , we must haveddt( ) ∂L− ∂L∂ ˙q ∂q = 0 (2.2.5)The above is called the Lagrange equ<strong>at</strong>ion, from which the equ<strong>at</strong>ion <strong>of</strong> motion <strong>of</strong> a particle canbe derived. The deriv<strong>at</strong>ive <strong>of</strong> the Lagrangian with respect to the velocity ˙q is the momentump = ∂L∂ ˙q(2.2.6)The deriv<strong>at</strong>ive <strong>of</strong> the Lagrangian with respect to the coordin<strong>at</strong>e q is the force. HenceF = ∂L∂q(2.2.7)The above equ<strong>at</strong>ion <strong>of</strong> motion is thenṗ = F (2.2.8)Equ<strong>at</strong>ion (2.2.6) can be inverted to express ˙q as a function <strong>of</strong> p and q, namely˙q = f(p, q) (2.2.9)The above two equ<strong>at</strong>ions can be solved in tandem to find the time evolution <strong>of</strong> p and q.


8 Quantum Mechanics Made SimpleFor example, the kinetic energy T <strong>of</strong> a particle is given byThen from (2.2.1), and the fact th<strong>at</strong> V is independent <strong>of</strong> ˙q,T = 1 2 m ˙q2 (2.2.10)orp = ∂L∂ ˙q = ∂T∂ ˙q= m ˙q (2.2.11)˙q = p m(2.2.12)Also, from (2.2.1), (2.2.7), and (2.2.8), we haveṗ = − ∂V∂q(2.2.13)The above pair, (2.2.12) and (2.2.13), form the equ<strong>at</strong>ions <strong>of</strong> motion for this problem.The above can be generalized to multidimensional problems. For example, for a oneparticle system in three dimensions, q i has three degrees <strong>of</strong> freedom, and i = 1, 2, 3. (The q ican represent x, y, z in Cartesian coordin<strong>at</strong>es, but r, θ, φ in spherical coordin<strong>at</strong>es.) But forN particles in three dimensions, there are 3N degrees <strong>of</strong> freedom, and i = 1, . . . , 3N. Theformul<strong>at</strong>ion can also be applied to particles constraint in motion. For instance, for N particlesin three dimensions, q i may run from i = 1, . . . , 3N − k, representing k constraints on themotion <strong>of</strong> the particles. This can happen, for example, if the particles are constraint to movein a manifold (surface), or a line (ring) embedded in a three dimensional space.Going through similar deriv<strong>at</strong>ion, we arrive <strong>at</strong> the equ<strong>at</strong>ion <strong>of</strong> motion( )d ∂L− ∂L = 0 (2.2.14)dt ∂ ˙q i ∂q iIn general, q i may not have a dimension <strong>of</strong> length, and it is called the generalized coordin<strong>at</strong>e(also called conjug<strong>at</strong>e coordin<strong>at</strong>e). Also, ˙q i may not have a dimension <strong>of</strong> velocity, and it iscalled the generalized velocity.The deriv<strong>at</strong>ive <strong>of</strong> the Lagrangian with respect to the generalized velocity is the generalizedmomentum (also called conjug<strong>at</strong>e momentum), namely,p i = ∂L∂ ˙q i(2.2.15)The generalized momentum may not have a dimension <strong>of</strong> momentum. Hence, the equ<strong>at</strong>ion<strong>of</strong> motion (2.2.14) can be written asṗ i = ∂L∂q i(2.2.16)Equ<strong>at</strong>ion (2.2.15) can be inverted to yield an equ<strong>at</strong>ion for ˙q i as a function <strong>of</strong> the othervariables. This equ<strong>at</strong>ion can be used in tandem (2.2.16) as time-marching equ<strong>at</strong>ions <strong>of</strong> motion.


Classical Mechanics and Some M<strong>at</strong>hem<strong>at</strong>ical Preliminaries 92.3 Hamiltonian Formul<strong>at</strong>ionFor a multi-dimensional system, or a many particle system in multi-dimensions, the totaltime deriv<strong>at</strong>ive <strong>of</strong> L isdLdt = ∑ ( ∂L˙q i + ∂L )¨q i (2.3.1)∂qi i ∂ ˙q iSince ∂L/∂q i = d dt (∂L/∂ ˙q i) from the Lagrange equ<strong>at</strong>ion, we havedLdt = ∑ [ ( ) d ∂L˙q i + ∂L ]¨q i = d ∑( ) ∂L˙q i (2.3.2)dt ∂ ˙qii ∂ ˙q i dt ∂ ˙qi ior( )d ∑ ∂L˙q i − L = 0 (2.3.3)dt ∂ ˙qi iThe quantityH = ∑ i∂L∂ ˙q i˙q i − L (2.3.4)is known as the Hamiltonian <strong>of</strong> the system, and is a constant <strong>of</strong> motion, namely, dH/dt = 0.As shall be shown, the Hamiltonian represents the total energy <strong>of</strong> a system. It is a constant<strong>of</strong> motion because <strong>of</strong> the conserv<strong>at</strong>ion <strong>of</strong> energy.The Hamiltonian <strong>of</strong> the system, (2.3.4), can also be written, after using (2.2.15), asH = ∑ i˙q i p i − L (2.3.5)where p i = ∂L/∂ ˙q i is the generalized momentum. The first term has a dimension <strong>of</strong> energy,and in Cartesian coordin<strong>at</strong>es, for a simple particle motion, it is easily seen th<strong>at</strong> it is twice thekinetic energy. Hence, the above indic<strong>at</strong>es th<strong>at</strong> the HamiltonianThe total vari<strong>at</strong>ion <strong>of</strong> the Hamiltonian isH = T + V (2.3.6)( ∑δH = δip i ˙q i)− δL= ∑ ( ˙q i δp i + p i δ ˙q i ) − ∑iiUsing (2.2.15) and (2.2.16), we have( ∂L∂q iδq i + ∂L∂ ˙q iδ ˙q i)(2.3.7)δH = ∑ i( ˙q i δp i + p i δ ˙q i ) − ∑ i(ṗ i δq i + p i δ ˙q i )= ∑ i( ˙q i δp i − ṗ i δq i ) (2.3.8)


10 Quantum Mechanics Made SimpleFrom the above, since the first vari<strong>at</strong>ion <strong>of</strong> the Hamiltonian depends only on δp i and δq i ,we g<strong>at</strong>her th<strong>at</strong> the Hamiltonian is a function <strong>of</strong> p i and q i . Taking the first vari<strong>at</strong>ion <strong>of</strong>the Hamiltonian with respect to these variables, we arrive <strong>at</strong> another expression for its firstvari<strong>at</strong>ion, namely,δH = ∑ ( ∂Hδp i + ∂H )δq i (2.3.9)∂pi i ∂q iComparing the above with (2.3.8), we g<strong>at</strong>her th<strong>at</strong>˙q i = ∂H∂p i(2.3.10)ṗ i = − ∂H∂q i(2.3.11)These are the equ<strong>at</strong>ions <strong>of</strong> motion known as the Hamiltonian equ<strong>at</strong>ions.The (2.3.4) is also known as the Legendre transform<strong>at</strong>ion. The original function L is afunction <strong>of</strong> ˙q i , q i . Hence, δL depends on both δ ˙q i and δq i . After the Legendre transform<strong>at</strong>ion,δH depends on the differential δp i and δq i as indic<strong>at</strong>ed by (2.3.8). This implies th<strong>at</strong> H is afunction <strong>of</strong> p i and q i . The equ<strong>at</strong>ions <strong>of</strong> motion then can be written as in (2.3.10) and (2.3.11).2.4 More on HamiltonianThe Hamiltonian <strong>of</strong> a particle in classical mechanics is given by (2.3.6), and it is a function<strong>of</strong> p i and q i . For a non-rel<strong>at</strong>ivistic particle in three dimensions, the kinetic energyT = p · p2m(2.4.1)and the potential energy V is a function <strong>of</strong> q. Hence, the Hamiltonian can be expressed asH = p · p2m+ V (q) (2.4.2)in three dimensions. When an electromagnetic field is present, the Hamiltonian for an electroncan be derived by letting the generalized momentump i = m ˙q i + eA i (2.4.3)where e = −|e| is the electron charge and A i is component <strong>of</strong> the vector potential A. Consequently,the Hamiltonian <strong>of</strong> an electron in the presence <strong>of</strong> an electromagnetic field isH =(p − eA) · (p − eA)2m+ eφ(q) (2.4.4)The equ<strong>at</strong>ion <strong>of</strong> motion <strong>of</strong> an electron in an electromagnetic field is governed by the Lorentzforce law, which can be derived from the above Hamiltonian using the equ<strong>at</strong>ions <strong>of</strong> motionprovided by (2.3.10) and (2.3.11).


12 Quantum Mechanics Made SimpleA m<strong>at</strong>rix is a m<strong>at</strong>hem<strong>at</strong>ical linear oper<strong>at</strong>or th<strong>at</strong> when oper<strong>at</strong>e (also called ”act”) on avector produces another vector, orb = A · a (2.6.1)where a and b are distinct vectors, and A is a m<strong>at</strong>rix oper<strong>at</strong>or other than the identityoper<strong>at</strong>or. The inner product between two vectors can be <strong>of</strong> the form <strong>of</strong> reaction inner productor energy inner productv t · w (2.6.2)v † · w (2.6.3)where the † implies conjug<strong>at</strong>ion transpose or th<strong>at</strong> v † = (v ∗ ) t . For a finite dimensional m<strong>at</strong>rixand vectors, the above can be written asb j =v t · w =v † · w =N∑A ji a i (2.6.4)i=1N∑v i w i (2.6.5)i=1N∑vi ∗ w i (2.6.6)The above are sometimes written with the summ<strong>at</strong>ion sign removed, namely, asi=1b j = A ji a i (2.6.7)v t · w = v i w i (2.6.8)v † · w = v ∗ i w i (2.6.9)The summ<strong>at</strong>ion is implied whenever repe<strong>at</strong>ed indices occur. This is known varyingly as theindex not<strong>at</strong>ion, indicial not<strong>at</strong>ion, or Einstein not<strong>at</strong>ion.The above concepts can be extended to infinite dimensional system by letting N → ∞. Itis quite clear th<strong>at</strong>N∑N∑v † · v = vi ∗ v i = |v i | 2 > 0 (2.6.10)i=1or v † · v is positive definite.Furthermore, m<strong>at</strong>rix oper<strong>at</strong>ors s<strong>at</strong>isfies associ<strong>at</strong>ivity but not commut<strong>at</strong>ivity, namely,(A · B)· C = A ·(B · C)(2.6.11)i=1A · B ≠ B · A (2.6.12)


Classical Mechanics and Some M<strong>at</strong>hem<strong>at</strong>ical Preliminaries 132.6.1 Identity, Hermitian, Symmetric, Inverse and Unitary M<strong>at</strong>ricesFor discrete, countable systems, the definition <strong>of</strong> the above is quite straightforward.identity oper<strong>at</strong>or I is defined such th<strong>at</strong>TheI · a = a (2.6.13)or the ij element <strong>of</strong> the m<strong>at</strong>rix is[I]ij = δ ij (2.6.14)or it is diagonal m<strong>at</strong>rix with one on the diagonal. A Hermitian m<strong>at</strong>rix A ij has the propertyth<strong>at</strong>orA symmetric m<strong>at</strong>rix A ij is such th<strong>at</strong>orA ij = A ∗ ji (2.6.15)A † = A (2.6.16)A ij = A ji (2.6.17)A t = A (2.6.18)The inverse <strong>of</strong> the m<strong>at</strong>rix A, denoted as A −1 , has the property th<strong>at</strong>A −1 · A = I (2.6.19)So given the equ<strong>at</strong>ionA · x = b (2.6.20)x can be found once A −1 is known. Multiplying the above by A −1 , we havex = A −1 · b (2.6.21)A unitary m<strong>at</strong>rix U has the property th<strong>at</strong>U † · U = I (2.6.22)In other wordsU † = U −1 (2.6.<strong>23</strong>)


14 Quantum Mechanics Made Simple2.6.2 DeterminantAn N × N m<strong>at</strong>rix A can be written as a collection <strong>of</strong> column vectorsA = [a 1 , a 2 , a 3 . . . a N ] (2.6.24)where the column vectors are <strong>of</strong> length N. Determinant <strong>of</strong> A, the det(A), or |A| can bethought <strong>of</strong> as the “volume” subtended by the vectors a n . . . a N . Also, det(A) changes signwhen any two column vectors are swapped. It can be thought <strong>of</strong> as a generalized “crossproduct”.Some useful properties <strong>of</strong> determinants are1. det ( I ) = 12. det(A t) = det ( A )3. det ( A · B ) = det ( A) · det(B )4. det(A −1) = 1/det ( A )5. If Λ is diagonal, then det ( Λ ) = ∏ Ni=1 λ ii6. det ( A ) = ∏ Ni=1 λ i where λ i are the eigenvalues <strong>of</strong> A.7. det ( A ) = 0 implies th<strong>at</strong> A is singular, or there exists a vector u such th<strong>at</strong> A · u =08. det ( cA ) = c N det ( A )2.6.3 Eigenvectors and EigenvaluesAn eigenvector v <strong>of</strong> a m<strong>at</strong>rix A s<strong>at</strong>isfies the property th<strong>at</strong>A · v = λv (2.6.25)where λ is the eigenvalue. The above can be rewritten as(A − λI)· v = 0 (2.6.26)The above implies th<strong>at</strong>det ( A − λI ) = 0 (2.6.27)In general, an N ×N m<strong>at</strong>rix has N eigenvalues with corresponding N eigenvectors. Whentwo or more eigenvalues are the same, they are known as degener<strong>at</strong>e. Some useful propertiesare:1. If A is Hermitian, then λ is real.2. If λ i ≠ λ j , then v † i · v j = 0. In general, v † i · v j = C n δ ij where C n is real.3. If A is positive definite, then λ i > 0 for all i and vice versa for neg<strong>at</strong>ive definiteness.4. P · A · P −1 has the same eigenvalues as A.


Classical Mechanics and Some M<strong>at</strong>hem<strong>at</strong>ical Preliminaries 152.6.4 Trace <strong>of</strong> a M<strong>at</strong>rixThe trace <strong>of</strong> a m<strong>at</strong>rix A is the sum <strong>of</strong> its diagonal elements, orSome properties are:1. tr ( A + B ) =tr ( A ) +tr ( B )2. tr ( cA ) =c tr ( A )3. tr ( A ) =tr(A t)tr ( A ) =4. tr ( A · B ) =tr ( B · A )() (5. tr P −1 (A ) ) −1· A · P =tr · P · P =tr ( A )N∑A ii (2.6.28)i=16. tr ( A ) = ∑ Ni=1 λ i where λ i ’s are the eigenvalues <strong>of</strong> A.2.6.5 Function <strong>of</strong> a M<strong>at</strong>rixAn example <strong>of</strong> a function <strong>of</strong> a m<strong>at</strong>rix isThe above has no meaning unless it oper<strong>at</strong>es on a vector; namelyWe can Taylor series expand to gete A = I + A + A22!e A (2.6.29)e A · x = b (2.6.30)+ A33!+ · · · + Ann!+ · · · (2.6.31)Then if v is an eigenvector <strong>of</strong> A such th<strong>at</strong> A · v = λv, then)e A · v =(I + A + A2+ · · · + An+ · · · · v (2.6.32)2! n!)=(1 + λ + λ22! + · · · + λnn! + · · · · v (2.6.33)= e λ v (2.6.34)In general, we can expandN∑x = a i v i (2.6.35)i=1


Classical Mechanics and Some M<strong>at</strong>hem<strong>at</strong>ical Preliminaries 17Derive th<strong>at</strong>m¨r = q∇A · ṙ − qṙ · ∇A − q∇φ − q ∂A∂tShow th<strong>at</strong> the above is the same as the Lorentz force law th<strong>at</strong>(2.6.48)F = ma = qv × B + qE (2.6.49)where F is the force on the electron, a is the acceler<strong>at</strong>ion, and v is the velocity, B is themagnetic field, and E is the electric field.


18 Quantum Mechanics Made Simple


Chapter 3Quantum Mechanics—SomePreliminaries3.1 IntroductionWith some background in classical mechanics, we may motiv<strong>at</strong>e the Schrödinger equ<strong>at</strong>ionin a more sanguine fashion. Experimental evidence indic<strong>at</strong>ed th<strong>at</strong> small particles such aselectrons behave quite strangely and cannot be described by classical mechanics alone. Inclassical mechanics, once we know p and q and their time deriv<strong>at</strong>ives (or ṗ, ˙q) <strong>of</strong> a particle <strong>at</strong>time t 0 , one can integr<strong>at</strong>e the equ<strong>at</strong>ions <strong>of</strong> motionṗ = F, ˙q = p/m (3.1.1)or use the finite difference method to find p and q <strong>at</strong> t 0 + ∆t, and <strong>at</strong> all subsequent times.In quantum mechanics, the use <strong>of</strong> two variables p and q and their deriv<strong>at</strong>ives is insufficientto describe the st<strong>at</strong>e <strong>of</strong> a particle and derive its future st<strong>at</strong>es. The st<strong>at</strong>e <strong>of</strong> a particle has tobe more richly endowed and described by a wavefunction or st<strong>at</strong>e function ψ(q, t). The st<strong>at</strong>efunction (also known as a st<strong>at</strong>e vector) is a vector in the infinite dimensional space.At this juncture, the st<strong>at</strong>e function or vector is analogous to when we study the controltheory <strong>of</strong> a highly complex system. In the st<strong>at</strong>e variable approach, the st<strong>at</strong>e <strong>of</strong> a control systemis described by the st<strong>at</strong>e vector, whose elements are variables th<strong>at</strong> we think are important tocapture the st<strong>at</strong>e <strong>of</strong> the system. For example, the st<strong>at</strong>e vector v, describing the st<strong>at</strong>e <strong>of</strong> thefactory, can contain variables th<strong>at</strong> represent the number <strong>of</strong> people in a factory, the number <strong>of</strong>machines, the temper<strong>at</strong>ure <strong>of</strong> the rooms, the inventory in each room, etc. The st<strong>at</strong>e equ<strong>at</strong>ion<strong>of</strong> this factory can then be written asdv(t) = A · v(t) (3.1.2)dtIt describes the time evolution <strong>of</strong> the factory. The m<strong>at</strong>rix A causes the coupling betweenst<strong>at</strong>e variables as they evolve. It bears strong similarity to the time-dependent Schrödinger19


20 Quantum Mechanics Made SimpleFigure 3.1: The st<strong>at</strong>e <strong>of</strong> a particle in quantum mechanics is described by a st<strong>at</strong>e function,which has infinitely many degrees <strong>of</strong> freedom.equ<strong>at</strong>ion, which is used to describe the time-evolution <strong>of</strong> the st<strong>at</strong>e function or the wavefunction<strong>of</strong> an electron. In the wavefunction, a complex number is assigned to each loc<strong>at</strong>ion in space.In the Schrödinger equ<strong>at</strong>ion, the wavefunction ψ(q, t) is a continuous function <strong>of</strong> <strong>of</strong> theposition variable q <strong>at</strong> any time instant t; hence, it is described by infinitely many numbers,and has infinite degrees <strong>of</strong> freedom. The time evolution <strong>of</strong> the wavefunction ψ(q, t) is governedby the Schrödinger equ<strong>at</strong>ion. It was motiv<strong>at</strong>ed by experimental evidence and the works <strong>of</strong>many others such as Planck, Einstein, and de Broglie, who were aware <strong>of</strong> the wave n<strong>at</strong>ure <strong>of</strong>a particle and the dual wave-particle n<strong>at</strong>ure <strong>of</strong> light.3.2 Probabilistic Interpret<strong>at</strong>ion <strong>of</strong> the WavefunctionThe wavefunction <strong>of</strong> the Schrödinger equ<strong>at</strong>ion has defied an acceptable interpret<strong>at</strong>ion formany years even though the Schrödinger equ<strong>at</strong>ion was known to predict experimental outcomes.Some thought th<strong>at</strong> it represented an electron cloud, and th<strong>at</strong> perhaps, an electron,<strong>at</strong> the <strong>at</strong>omistic level, behaved like a charge cloud, and hence not a particle. The final, mostaccepted interpret<strong>at</strong>ion <strong>of</strong> this wavefunction (one th<strong>at</strong> also agrees with experiments) is th<strong>at</strong>its magnitude squared corresponds to the probabilistic density function. 1 In other words, theprobability <strong>of</strong> finding an electron in an interval [x, x + ∆x] is equal to|ψ(x, t)| 2 ∆x (3.2.1)For the 3D case, the probability <strong>of</strong> finding an electron in a small volume ∆V in the vicinity<strong>of</strong> the point r is given by|ψ(r, t)| 2 ∆V (3.2.2)Since the magnitude squared <strong>of</strong> the wavefunction represents a probability density function,it must s<strong>at</strong>isfy the normaliz<strong>at</strong>ion condition <strong>of</strong> a probability density function, viz.,∫dV |ψ(r, t)| 2 = 1 (3.2.3)1 This interpret<strong>at</strong>ion is due to Born.


Quantum Mechanics—Some Preliminaries 21with its counterparts in 1D and 2D. The magnitude squared <strong>of</strong> this wavefunction is like somekind <strong>of</strong> “energy” th<strong>at</strong> cannot be destroyed. Electrons cannot be destroyed and hence, chargeconserv<strong>at</strong>ion is upheld by the Schrödinger equ<strong>at</strong>ion.3.3 Time Evolution <strong>of</strong> the Hamiltonian Oper<strong>at</strong>orMotiv<strong>at</strong>ed by the conserv<strong>at</strong>ion <strong>of</strong> the “energy” <strong>of</strong> the wavefunction, we shall consider an“energy” conserving system where the classical Hamiltonian will be a constant <strong>of</strong> motion. Inthis case, there is no “energy” loss from the system. The Schrödinger equ<strong>at</strong>ion th<strong>at</strong> governsthe time evolution <strong>of</strong> the wavefunction ψ isĤψ = i dψdt(3.3.1)where Ĥ is the Hamiltonian oper<strong>at</strong>or.2 One can solve (3.3.1) formally to obtainψ(t) = e −i Ĥ t ψ(t = 0) (3.3.2)Since the above is a function <strong>of</strong> an oper<strong>at</strong>or, it has meaning only if this function acts on theeigenfunctions <strong>of</strong> the oper<strong>at</strong>or Ĥ. As has been discussed in Chapter 2, Subsection 2.6.5, itcan be shown easily th<strong>at</strong> if A · v i = λ i v i ,exp(A) · v i = exp(λ i )v i (3.3.3)To simplify the expression (3.3.2), it is best to express ψ(t = 0) as an eigenfunction(eigenst<strong>at</strong>e or eigenvector) <strong>of</strong> Ĥ, or linear superposition <strong>of</strong> its eigenfunctions. If Ĥ is aHermitian oper<strong>at</strong>or, then there exists eigenfunctions, or special wavefunctions, ψ n , such th<strong>at</strong>Ĥψ n = E n ψ n (3.3.4)Analogous to the Hermitian m<strong>at</strong>rix oper<strong>at</strong>or, it can be shown th<strong>at</strong> E n is purely real and th<strong>at</strong>the ψ n are orthogonal to each other. In this case, the time evolution <strong>of</strong> ψ n from (3.3.2) isEn−iψ n (t) = e t ψ n (t = 0) = e −iωnt ψ n (t = 0) (3.3.5)In the above, E n = ω n , or the energy E n is rel<strong>at</strong>ed to frequency ω n via the reduced Planckconstant . The reduced Planck constant is rel<strong>at</strong>ed to the Planck constant by = h/(2π)and h = 6.626068 × 10 −34 J s. The fact th<strong>at</strong> E n is real means th<strong>at</strong> ω n is real, or th<strong>at</strong> themagnitude squared <strong>of</strong> these functions are time independent or conserved, as is required bytheir probabilistic interpret<strong>at</strong>ion.We can writeψ(r, t = 0) = ∑ nc n ψ n (r, t = 0) (3.3.6)2 Rightfully, one should use the bra and ket not<strong>at</strong>ion to write this equ<strong>at</strong>ion as Ĥ|ψ〉 = i d |ψ〉. In the lessdtrigorous not<strong>at</strong>ion in (3.3.1), we will assume th<strong>at</strong> Ĥ is in the represent<strong>at</strong>ion in which the st<strong>at</strong>e vector ψ is in.Th<strong>at</strong> is if ψ is in coordin<strong>at</strong>e space represent<strong>at</strong>ion, Ĥ is also in coordin<strong>at</strong>es space represent<strong>at</strong>ion.


22 Quantum Mechanics Made Simplewhere c n can be found using the orthonormality rel<strong>at</strong>ion <strong>of</strong> ψ n . Thenψ(r, t) = e − iĤ t ψ(r, t = 0) = ∑ nc n e −iωnt ψ n (r, t = 0) (3.3.7)It can be easily shown th<strong>at</strong> the above, which is derived from (3.3.2) is a solution to (3.3.1).Hence, (3.3.2) is the formal solution to (3.3.1).Scalar variables th<strong>at</strong> are measurable in classical mechanics, such as p and q, are known asobservables in quantum mechanics. They are elev<strong>at</strong>ed from scalar variables to oper<strong>at</strong>ors inquantum mechanics, denoted by a “ˆ” symbol here. In classical mechanics, for a one particlesystem, the Hamiltonian is given byH = T + V = p22m + V (3.3.8)The Hamiltonian contains the inform<strong>at</strong>ion from which the equ<strong>at</strong>ions <strong>of</strong> motion for the particlecan be derived. But in quantum mechanics, this is not sufficient, and H becomes an oper<strong>at</strong>orĤ = ˆp22m + ˆV (3.3.9)This oper<strong>at</strong>or works in tandem with a wavefunction ψ to describe the st<strong>at</strong>e <strong>of</strong> the particle.The oper<strong>at</strong>or acts on a wavefunction ψ(t), where in the coordin<strong>at</strong>e q represent<strong>at</strong>ion, is ψ(q, t).When ψ(q, t) is an eigenfunction with energy E n , it can be expressed asψ n (q, t) = ψ n (q)e −iωnt (3.3.10)where E n = ω n . The Schrödinger equ<strong>at</strong>ion for ψ n (q) then becomes 3( ) ˆp2Ĥψ n (q) =2m + ˆV ψ n (q) = E n ψ n (q) (3.3.11)For simplicity, we consider an electron moving in free space where it has only a constantkinetic energy but not influenced by any potential energy. In other words, there is no forceacting on the electron. In this case, ˆV = 0, and this equ<strong>at</strong>ion becomesˆp 22m ψ n(q) = E n ψ n (q) (3.3.12)It has been observed by de Broglie th<strong>at</strong> the momentum <strong>of</strong> a particle, such as an electronwhich behaves like a wave, has a momentump = k (3.3.13)where k = 2π/λ is the wavenumber <strong>of</strong> the wavefunction. This motiv<strong>at</strong>es th<strong>at</strong> the oper<strong>at</strong>or ˆpcan be expressed byˆp = −i d dq(3.3.14)3 For the Schrödinger equ<strong>at</strong>ion in coordin<strong>at</strong>e space, ˆV turns out to be a scalar oper<strong>at</strong>or or a diagonaloper<strong>at</strong>or.


Quantum Mechanics—Some Preliminaries <strong>23</strong>in the coordin<strong>at</strong>e space represent<strong>at</strong>ion. This is chosen so th<strong>at</strong> if an electron is described by ast<strong>at</strong>e function ψ(q) = c 1 e ikq , then ˆpψ(q) = kψ(q). The above motiv<strong>at</strong>ion for the form <strong>of</strong> theoper<strong>at</strong>or ˆp is highly heuristic. We will see other reasons for the form <strong>of</strong> ˆp when we study thecorrespondence principle and the Heisenberg picture.Equ<strong>at</strong>ion (3.3.12) for a free particle is then− 2d 22m dq 2 ψ n(q) = E n ψ n (q) (3.3.15)Since this is a constant coefficient ordinary differential equ<strong>at</strong>ion, the solution is <strong>of</strong> the formwhich when used in (3.3.15), yieldsNamely, the kinetic energy T <strong>of</strong> the particle is given byψ n (q) = e ±ikq (3.3.16) 2 k 22m = E n (3.3.17)T = 2 k 22m(3.3.18)where p = k is in agreement with de Broglie’s finding.In many problems, the oper<strong>at</strong>or ˆV is a scalar oper<strong>at</strong>or in coordin<strong>at</strong>e space represent<strong>at</strong>ionwhich is a scalar function <strong>of</strong> position V (q). This potential traps the particle within it actingas a potential well. In general, the Schrödinger equ<strong>at</strong>ion for a particle becomes[− 2∂ 2 ]2m ∂q 2 + V (q) ψ(q, t) = i ∂ ψ(q, t) (3.3.19)∂tFor a particular eigenst<strong>at</strong>e with energy E n as indic<strong>at</strong>ed by (3.3.10), it becomes[− 2d 2 ]2m dq 2 + V (q) ψ n (q) = E n ψ n (q) (3.3.20)The above is an eigenvalue problem with eigenvalue E n and eigenfunction ψ n (q). Theseeigenst<strong>at</strong>es are also known as st<strong>at</strong>ionary st<strong>at</strong>es, because they have a time dependence indic<strong>at</strong>edby (3.3.10). Hence, their probability density functions |ψ n (q, t)| 2 are time independent.These eigenfunctions correspond to trapped modes (or bound st<strong>at</strong>es) in the potential welldefined by V (q) very much like trapped guided modes in a dielectric waveguide. These modesare usually countable and they can be indexed by the index n.In the special case <strong>of</strong> a particle in free space, or the absence <strong>of</strong> the potential well, theparticle or electron is not trapped and it is free to assume any energy or momentum indexedby the continuous variable k. In (3.3.17), the index for the energy should rightfully be k andthe eigenfunctions are uncountably infinite. Moreover, the above can be generalized to twoand three dimensional cases.


24 Quantum Mechanics Made Simple3.4 Simple Examples <strong>of</strong> Time Independent SchrödingerEqu<strong>at</strong>ionAt this juncture, we have enough knowledge to study some simple solutions <strong>of</strong> time-independentSchrödinger equ<strong>at</strong>ion such as a particle in a box, a particle impinging on a potential barrier,and a particle in a finite potential well.3.4.1 Particle in a 1D BoxConsider the Schrödinger equ<strong>at</strong>ion for the 1D case where the potential V (x) is defined to bea function with zero value for 0 < x < a (inside the box) and infinite value outside this range.The Schrödinger equ<strong>at</strong>ion is given by[− 2 d 2 ]2m dx 2 + V (x) ψ(x) = Eψ(x) (3.4.1)where we have replaced q with x. Since V (x) is infinite outside the box, ψ(x) has to be zero.Inside the well, V (x) = 0 and the above equ<strong>at</strong>ion has a general solution <strong>of</strong> the formψ(x) = A sin(kx) + B cos(kx) (3.4.2)The boundary conditions are th<strong>at</strong> ψ(x = 0) = 0 and ψ(x = a) = 0. For this reason, a viablesolution for ψ(x) isψ(x) = A sin(kx) (3.4.3)where k = nπ/a, n = 1, . . . , ∞. There are infinitely many eigensolutions for this problem.For each chosen n, the corresponding energy <strong>of</strong> the solution isE n = (nπ/a)22m(3.4.4)These energy values are the eigenvalues <strong>of</strong> the problem, with the corresponding eigenfunctionsgiven by (3.4.3) with the appropri<strong>at</strong>e k. It is seen th<strong>at</strong> the more energetic the electron is(high E n values), the larger the number <strong>of</strong> oscill<strong>at</strong>ions the wavefunction has inside the box.The solutions th<strong>at</strong> are highly oscill<strong>at</strong>ory have higher k values, and hence, higher momentumor higher kinetic energy. The solutions which are even about the center <strong>of</strong> the box are saidto have even parity, while those th<strong>at</strong> are odd have odd parity.One other thing to be noted is th<strong>at</strong> the magnitude squared <strong>of</strong> the wavefunction aboverepresents the probability density function. Hence, it has to be normalized. The normalizedversion <strong>of</strong> the wavefunction isψ(x) = √ 2/a sin(nπx/a) (3.4.5)Moreover, these eigenfunctions are orthonormal to each other, viz.,∫ a0dxψ ∗ n(x)ψ m (x) = δ nm (3.4.6)


Quantum Mechanics—Some Preliminaries 25The orthogonality is the generaliz<strong>at</strong>ion <strong>of</strong> the fact th<strong>at</strong> for a Hermitian m<strong>at</strong>rix system, wherethe eigenvectors are given byH · v i = λ i v i (3.4.7)then it can be proven easily th<strong>at</strong>Moreover, the eigenvalues are real.v † j · v i = C j δ ij (3.4.8)Figure 3.2: The wavefunctions <strong>of</strong> an electron trapped in a 1D box (from DAB Miller).3.4.2 Particle Sc<strong>at</strong>tering by a BarrierIn the previous example, it is manifestly an eigenvalue problem since the solution can befound only <strong>at</strong> discrete values <strong>of</strong> E n . The electron is trapped inside the box. However, in anopen region problem where the electron is free to roam, the energy <strong>of</strong> the electron E can bearbitrary. We can assume th<strong>at</strong> the potential pr<strong>of</strong>ile is such th<strong>at</strong> V (x) = 0 for x < 0 whileV (x) = V 0 for x > 0. The energy <strong>of</strong> the electron is such th<strong>at</strong> 0 < E < V 0 . On the left side,we assume an electron coming in from −∞ with the wavefunction described by A exp(ik 1 x).When this wavefunction hits the potential barrier, a reflected wave will be present, and thegeneral solution on the left side <strong>of</strong> the barrier is given byψ 1 (x) = A 1 e ik1x + B 1 e −ik1x (3.4.9)where (k 1 ) 2 /(2m) = E is the kinetic energy <strong>of</strong> the incident electron. On the right side,however, the Schrödinger equ<strong>at</strong>ion to be s<strong>at</strong>isfied is[− 2 d 2 ]2m dx 2 ψ 2 (x) = (E − V 0 )ψ 2 (x) (3.4.10)The solution <strong>of</strong> the transmitted wave on the right isψ 2 (x) = A 2 e ik2x (3.4.11)wherek 2 = √ 2m(E − V 0 )/ (3.4.12)


26 Quantum Mechanics Made SimpleGiven the known incident wave amplitude A 1 , we can m<strong>at</strong>ch the boundary conditions <strong>at</strong> x = 0to find the reflected wave amplitude B 1 and the transmitted wave amplitude A 2 . By eyeballingthe Schrödinger equ<strong>at</strong>ion (3.4.1), we can arrive <strong>at</strong> the requisite boundary conditionsare th<strong>at</strong> ψ and dd2dxψ(x) are continuous <strong>at</strong> x = 0. This is becausedxψ(x) in (3.4.1) has to be2da finite quantity. Hence,dxψ(x) and ψ(x) cannot have jump discontinuities.Since E < V 0 , k 2 is pure imaginary, and the wave is evanescent and decays when x → ∞.This effect is known as tunneling. The electron as a nonzero probability <strong>of</strong> being found insidethe barrier, albeit with decreasing probability into the barrier. The larger V 0 is compared toE, the more rapidly decaying is the wavefunction into the barrier.However, if the electron is energetic enough so th<strong>at</strong> E > V 0 , k 2 becomes real, and thenthe wavefunction is no more evanescent. It penetr<strong>at</strong>es into the barrier; it can be found evena long way from the boundary.Figure 3.3: Sc<strong>at</strong>tering <strong>of</strong> the electron wavefunction by a 1D barrier (from DAB Miller).It is to be noted th<strong>at</strong> the wavefunction in this case cannot be normalized as the aboverepresents a fictitious situ<strong>at</strong>ion <strong>of</strong> an electron roaming over infinite space. The above exampleillustr<strong>at</strong>es the wave physics <strong>at</strong> the barrier.3.4.3 Particle in a Potential WellIf the potential pr<strong>of</strong>ile is such th<strong>at</strong>⎧⎪⎨ V 1 if x < −a/2, region 1V (x) = V 2 = 0 if |x| < a/2, region 2⎪⎩V 3 if x > a/2, region 3(3.4.13)then there can be trapped modes (or bound st<strong>at</strong>es) inside the well represented by standingwaves, whereas outside the well, the waves are evanescent for eigenmodes for which E < V 1and E < V 3 .The wavefunction for |x| < a/2 can be expressed asψ 2 (x) = A 2 sin(k 2 x) + B 2 cos(k 2 x) (3.4.14)


Quantum Mechanics—Some Preliminaries 27where k 2 = √ 2mE/. In region 1 to the left, the wavefunction isψ 1 (x) = A 1 e α1x (3.4.15)where α 1 = √ 2m(V 1 − E)/. The wave has to decay in the left direction. Similar, in region3 to the right, the wavefunction isψ 3 (x) = B 3 e −α3x (3.4.16)where α 3 = √ 2m(V 3 − E)/. It has to decay in the right direction. Four boundary conditionscan be imposed <strong>at</strong> x = ±a/2 to elimin<strong>at</strong>e the four unknowns A 1 , A 2 , B 2 , and B 3 . These fourboundary conditions are continuity <strong>of</strong> the wavefunction ψ(x) and its deriv<strong>at</strong>ive dψ(x)/dx <strong>at</strong>the two boundaries. However, non-trivial eigensolutions can only exist <strong>at</strong> selected values <strong>of</strong>E which are the eigenvalues <strong>of</strong> the Schrödinger equ<strong>at</strong>ion. The eigenequ<strong>at</strong>ion from which theeigenvalues can be derived is a transcendental equ<strong>at</strong>ion.To illustr<strong>at</strong>e this point, we impose th<strong>at</strong> ψ is continuous <strong>at</strong> x = ±a/2 to arrive <strong>at</strong> thefollowing two equ<strong>at</strong>ions:A 1 e −α1a/2 = −A 2 sin(k 2 a/2) + B 2 cos(k 2 a/2) (3.4.17)A 3 e −α3a/2 = A 2 sin(k 2 a/2) + B 2 cos(k 2 a/2) (3.4.18)We further impose th<strong>at</strong> ∂ψ/∂x is continuous <strong>at</strong> x = ±a to arrive <strong>at</strong> the following two equ<strong>at</strong>ions:α 1 A 1 e −α1a/2 = k 2 A 2 cos(k 2 a/2) + k 2 B 2 sin(k 2 a/2) (3.4.19)−α 3 A 3 e −α3a/2 = k 2 A 2 cos(k 2 a/2) − k 2 B 2 sin(k 2 a/2) (3.4.20)The above four equ<strong>at</strong>ions form a m<strong>at</strong>rix equ<strong>at</strong>ionM · v = 0 (3.4.21)where v = [A 1 , A 2 , B 2 , B 3 ] t , and the elements <strong>of</strong> M, which depend on E, can be gleaned <strong>of</strong>fthe above equ<strong>at</strong>ions. Non-trivial solutions exists for v only ifdet(M(E)) = |M(E)| = 0 (3.4.22)The above is the transcendental eigenequ<strong>at</strong>ion from which the eigenvalues E can be derived.The nontrivial solutions for v are in the null space <strong>of</strong> M. Having known v = [A 1 , A 2 , B 2 , B 3 ] t ,the eigenfunctions <strong>of</strong> the Schrödinger equ<strong>at</strong>ion can be constructed. Notice th<strong>at</strong> v is knownto an arbitrary multiplic<strong>at</strong>ive constant. The normaliz<strong>at</strong>ion <strong>of</strong> the eigenfunction will pin downthe value <strong>of</strong> this constant.When the potential well is symmetric such th<strong>at</strong> V 1 = V 3 = V 0 , then the solutions can bedecomposed into odd and even solutions about x = 0. In this case, either A 2 = 0 for evenmodes, or B 2 = 0 for odd modes. Furthermore, A 1 = ±B 3 for these modes. The problemthen has two unknowns, and two boundary conditions <strong>at</strong> one <strong>of</strong> the interfaces suffice to deducethe eigenequ<strong>at</strong>ion.


28 Quantum Mechanics Made SimpleThe above problem is analogous to the 1D dielectric waveguide problem in classical electromagnetics.In most textbooks, the transcendental eigenequ<strong>at</strong>ion is solved using a graphicalmethod, which can be done likewise here. The plot <strong>of</strong> the eigenmodes and their energy levelsare shown in Figure 3.4. It is an interesting example showing th<strong>at</strong> a trapped electron existswith different energy levels. The more energetic (more kinetic energy) the electron is, thehigher the energy level. For the case when E > V 0 , the wavefunction is not evanescent outsidethe well, and the electron is free to roam outside the well.Modern technology has allowed the engineering <strong>of</strong> nano-structures so small th<strong>at</strong> a quantumwell can be fabric<strong>at</strong>ed. Quantum well technology is one <strong>of</strong> the emerging nano-technologies.It will be shown l<strong>at</strong>er in the course th<strong>at</strong> an electron wave propag<strong>at</strong>ing in a l<strong>at</strong>tice is likean electron wave propag<strong>at</strong>ing in vacuum but with a different effective mass and seeing apotential th<strong>at</strong> is the potential <strong>of</strong> the conduction band. The quantum wells are fabric<strong>at</strong>edwith III-V compound, for example, with alloys <strong>of</strong> the form Al x Ga 1−x As forming a ternarycompound. They have different values <strong>of</strong> valence and conduction band depending on thevalue <strong>of</strong> x. By growing heterostructure layers with different compounds, multiple quantumwells can be made. Hence, the energy levels <strong>of</strong> a quantum well can also be engineered so th<strong>at</strong>laser technology <strong>of</strong> different wavelengths can be fabric<strong>at</strong>ed. 4In the case <strong>of</strong> a hydrogen <strong>at</strong>om, the Coulomb potential around the proton <strong>at</strong> the nucleusyields a potential well described by −q 2 /(ɛ4πr). This well can trap an electron into variouseigenst<strong>at</strong>es, yielding different electronic orbitals. The Schrödinger equ<strong>at</strong>ion has predicted theenergy levels <strong>of</strong> a hydrogen <strong>at</strong>om with gre<strong>at</strong> success.Figure 3.4: Trapped modes representing the electron wavefunctions inside a 1D potential well(from DAB Miller).3.5 The Quantum Harmonic Oscill<strong>at</strong>or–A PreviewThe pendulum, a particle <strong>at</strong>tached to a spring, or many vibr<strong>at</strong>ions in <strong>at</strong>oms and moleculescan be described as a harmonic oscill<strong>at</strong>or. Hence, the harmonic oscill<strong>at</strong>or is one <strong>of</strong> the most1998.4 J. H. Davis, The Physics <strong>of</strong> Low Dimensional Semiconductors: An Introduction, Cambridge U Press,


Quantum Mechanics—Some Preliminaries 29important examples in quantum mechanics. Its quantum mechanical version can be describedby the 1D Schrödinger equ<strong>at</strong>ion.The classical equ<strong>at</strong>ion for a harmonic oscill<strong>at</strong>or is given bym d2 z= −Kz (3.5.1)dt2 The above is Newton’s law, and K is the spring constant, and the force provided by the springis Kz. We can rewrite the above asd 2 zdt 2 = −ω2 z (3.5.2)where ω = √ K/m, and m is the mass <strong>of</strong> the particle. The above has a time harmonic solution<strong>of</strong> the form exp(±iωt) where ω is the oscill<strong>at</strong>ion frequency. Since the force F = −∂V/∂z, thepotential energy <strong>of</strong> a particle <strong>at</strong>tached to a spring is easily shown to be given byV (z) = 1 2 mω2 z 2 (3.5.3)Consequently, the above potential energy can be used in the Schrödinger equ<strong>at</strong>ion to describethe trapping <strong>of</strong> wave modes (or bound st<strong>at</strong>es). The kinetic energy <strong>of</strong> the particle is describedby a term proportional to the square <strong>of</strong> the momentum oper<strong>at</strong>or. Hence, the corresponding1D Schrödinger equ<strong>at</strong>ion is[− 2 d 22m dz 2 + 1 ]2 mω2 z 2 ψ n (z) = E n ψ n (z) (3.5.4)with a parabolic potential well. It turns out th<strong>at</strong> this equ<strong>at</strong>ion has closed-form solutions,yielding the wavefunction for an eigenst<strong>at</strong>e as given by√√ (√ )1 mω mω mωψ n (z) =2 n e− 2 z2 H nn! π z (3.5.5)where H n (x) is a Hermite polynomial, and the wavefunction is Gaussian tapered. The energy<strong>of</strong> the eigenst<strong>at</strong>e is given by(E n = n + 1 )ω (3.5.6)2The energy levels are equally spaced ω apart. Even the lowest energy st<strong>at</strong>e, the groundst<strong>at</strong>e, has a nonzero energy <strong>of</strong> ω/2 known as the zero-point energy. The higher energy st<strong>at</strong>escorrespond to larger amplitudes <strong>of</strong> oscill<strong>at</strong>ion, and vice versa for the lower energy st<strong>at</strong>es. Inorder to kick the quantum harmonic oscill<strong>at</strong>or from the low energy st<strong>at</strong>e to a level above, itneeds a packet <strong>of</strong> energy <strong>of</strong> ω, the quantiz<strong>at</strong>ion energy <strong>of</strong> a photon. The physics <strong>of</strong> quantizedelectromagnetic oscill<strong>at</strong>ions (photons) and quantized mechanical oscill<strong>at</strong>ions (phonons)is intim<strong>at</strong>ely rel<strong>at</strong>ed to the quantum harmonic oscill<strong>at</strong>or.


30 Quantum Mechanics Made SimpleFigure 3.5: Sketch <strong>of</strong> the eigenst<strong>at</strong>es, energy levels, and the potential well <strong>of</strong> a quantumharmonic oscill<strong>at</strong>or (picture from DAB Miller).


Chapter 4Time-Dependent SchrödingerEqu<strong>at</strong>ion4.1 IntroductionEach eigenst<strong>at</strong>e <strong>of</strong> Schrödinger equ<strong>at</strong>ion has its own time dependence <strong>of</strong> exp(−iω n t). Whenwe consider one eigenst<strong>at</strong>e alone, its time dependence is unimportant, as the time dependencedisappears when we convert the wavefunction into a probability density function by takingits magnitude squared. Moreover, the absolute frequency <strong>of</strong> an eigenst<strong>at</strong>e is arbitrary. Hence,the probability density function is quite uninteresting. This is an antithesis to the classicalharmonic oscill<strong>at</strong>or where the position <strong>of</strong> the particle moves with respect to time.However, when a quantum st<strong>at</strong>e is described by a wavefunction which is a linear superposition<strong>of</strong> two eigenst<strong>at</strong>es, it is important th<strong>at</strong> we take into account their individual frequencyvalue and time dependence. The two eigenst<strong>at</strong>es will “be<strong>at</strong>” with each other to produce adifference in frequencies when we take the magnitude squared <strong>of</strong> the wavefunction.4.2 Quantum St<strong>at</strong>es in the Time DomainConsider a quantum st<strong>at</strong>e which is a linear superposition <strong>of</strong> two eigenst<strong>at</strong>esψ(r, t) = c a e −iω<strong>at</strong> ψ a (r) + c b e −iω bt ψ b (r) (4.2.1)where c a and c b are properly chosen to normalize the corresponding probability density function.Then the probability function is[]|ψ(r, t)| 2 = |c a | 2 |ψ a (r)| 2 + |c b | 2 |ψ b (r)| 2 + 2Re c a c ∗ bψ a (r)ψb ∗ (r)e −i(ωa−ω b)t(4.2.2)It is clearly time varying. We are seeing the motion <strong>of</strong> the particle through the potential well.31


32 Quantum Mechanics Made SimpleFigure 4.1: The time evolution <strong>of</strong> the linear superposition <strong>of</strong> the first two eigenst<strong>at</strong>es <strong>of</strong> thesimple quantum harmonic oscill<strong>at</strong>or. (i) Beginning <strong>of</strong> a period (solid line). (ii) One quarterand three quarters way through the period (dotted line). (iii) Half way through the period(picture from DAB Miller).4.3 Coherent St<strong>at</strong>eThe eigenst<strong>at</strong>es <strong>of</strong> a quantum harmonic oscill<strong>at</strong>or do not resemble its classical st<strong>at</strong>e. First,its magnitude squared is time independent, whereas a classical harmonic oscill<strong>at</strong>or is timevarying. In order to see the classical st<strong>at</strong>e emerging from the quantum harmonic oscill<strong>at</strong>oreigenst<strong>at</strong>es, we need to take a judicious linear superposition <strong>of</strong> them. Such a st<strong>at</strong>e is calledthe coherent st<strong>at</strong>e. 1 The coherent st<strong>at</strong>e is given byψ N (ξ, t) =∞∑c Nn e −i(n+1/2)ωt ψ n (ξ) (4.3.1)n=0where ξ = √ mω/z, and c Nn = √ N n e −N /n!. 2 It forms a time-varying wave packet th<strong>at</strong>emul<strong>at</strong>es the motion <strong>of</strong> the classical harmonic oscill<strong>at</strong>ors such as a pendulum. The totalenergy <strong>of</strong> this wave packet is given by (N + 1/2)ω as shall be shown l<strong>at</strong>er. The larger N is,the more energy the particle has, and the closer is the coherent st<strong>at</strong>e to a classical st<strong>at</strong>e. Aplot <strong>of</strong> the coefficient |c Nn | 2 = P n as a function <strong>of</strong> n is shown in Fig. 4.2. Here, P n is theprobability <strong>of</strong> finding the st<strong>at</strong>e in the st<strong>at</strong>ionary st<strong>at</strong>e ψ n . Notice the the probability <strong>of</strong> beingin the eigenst<strong>at</strong>es peak around n = N, the dominant eigenst<strong>at</strong>e <strong>of</strong> the packet.1 The st<strong>at</strong>e was derived by R. Glauber who won the Nobel Prize in 2005 for his contribution.2 The coefficient is also written as e −|α|2 /2 α n√n!where N = |α| 2 .


Time-Dependent Schrödinger Equ<strong>at</strong>ion 330.400.350.300.25Eigenfunction Coefficents |c Nn | 2 N=1N=2N=5N=10N=50N=100|c Nn | 20.200.150.100.050.000 20 40 60 80 100 120 140Eigenfunction Index nFigure 4.2: The plot <strong>of</strong> |c Nn | 2 = P n as a function <strong>of</strong> n for different values <strong>of</strong> N.As shall be seen, as N increases, the coherent st<strong>at</strong>e looks more like a localized wavepacket describing a particle oscill<strong>at</strong>ing back and forth time harmonically. This is the case fora particle traveling in s parabolic potential well. Even when a particle is traveling in vacuum,it should be described by a localized wavefunction, such as a Gaussian wave packet. Thestudy <strong>of</strong> the Gaussian wave packet is given in Appendix A. It can be seen th<strong>at</strong> the classicallimit emerges when the momentum <strong>of</strong> the particle becomes very large.4.4 Measurement Hypothesis and Expect<strong>at</strong>ion ValueThe wavefunction <strong>of</strong> a quantum system can be written as a linear superposition <strong>of</strong> the st<strong>at</strong>ionaryst<strong>at</strong>esψ(r, t) = ∑ nc n (t)ψ n (r) (4.4.1)The magnitude squared <strong>of</strong> this wavefunction should integr<strong>at</strong>e to one due to its probabilisticinterpret<strong>at</strong>ion. In other words,∫∫∑|ψ(r, t)| 2 dr = c n (t)ψ n (r) ∑ c ∗ n ′(t)ψ∗ n ′(r)dr = 1 (4.4.2)n ′VVnExchanging the order <strong>of</strong> integr<strong>at</strong>ion and summ<strong>at</strong>ions, we arrive <strong>at</strong>∫|ψ(r, t)| 2 dr = ∑ ∑c n (t)c ∗ n ′(t) ψ n (r)ψn Vn n∫V∗ ′(r)dr (4.4.3)′


34 Quantum Mechanics Made SimpleFigure 4.3: Coherent st<strong>at</strong>e for different values <strong>of</strong> N (picture from DAB Miller).Using the orthonormality <strong>of</strong> the eigenst<strong>at</strong>es, viz., ∫ V ψ n(r)ψn ∗ ′(r)dr = δ nn ′, we arrive <strong>at</strong> thefact th<strong>at</strong>∑|c n (t)| 2 = 1 (4.4.4)nSince the squares <strong>of</strong> the magnitudes add up to one, they can be assigned probabilistic interpret<strong>at</strong>ionas well. Hence, P n = |c n (t)| 2 represents the probability <strong>of</strong> finding the electron ineigenst<strong>at</strong>e n.The quantum measurement hypothesis st<strong>at</strong>es th<strong>at</strong> before the measurement, the electronlives as a linear superposition <strong>of</strong> different eigenst<strong>at</strong>es. After the measurement, the electroncollapses into one <strong>of</strong> the eigenst<strong>at</strong>es, say the n eigenst<strong>at</strong>e, with probability proportional to|c n (t)| 2 .The above can go further by saying th<strong>at</strong> an electron is described by a wavefunction whereits position is indetermin<strong>at</strong>e: it can be found any place where the wavefunction is nonzero.Its probability <strong>of</strong> being found is proportional to the magnitude squared <strong>of</strong> the wavefunction.However, once the measurement is made to determine the loc<strong>at</strong>ion <strong>of</strong> the electron, the electron’sposition collapses to one loc<strong>at</strong>ion as is discovered by the experiment. Its position isdetermin<strong>at</strong>e after the measurement.One can think <strong>of</strong> the electron to be like “ghost” or “angel” which could exist as a linearsuperposition <strong>of</strong> many st<strong>at</strong>es before a measurement. After the measurement, it collapses tothe st<strong>at</strong>e “discovered” by the measurement.Due to this probabilistic interpret<strong>at</strong>ion, the expected energy <strong>of</strong> the quantum system is


36 Quantum Mechanics Made SimpleIt is seen th<strong>at</strong>〈Ĥ〉 ∫=drψ ∗ (r)Ĥψ(r) = 〈E〉 = ∑ nE n P n (4.4.12)in accordance with (4.4.5) and (4.4.9). Hence, the expect<strong>at</strong>ion <strong>of</strong> an oper<strong>at</strong>or is the average<strong>of</strong> its eigenvalues with respect to the st<strong>at</strong>e th<strong>at</strong> it is in.4.4.1 Uncertainty Principle–A Simple VersionThe uncertainty principle is asserted by the fact th<strong>at</strong> Fourier transform <strong>of</strong> a Gaussian functionis another Gaussian function. For instance, if we have a wave packet th<strong>at</strong> is formed bysuperposing waves with different k or momentum, we can express it aswhereψ(z) =∫ ∞−∞˜ψ(k)e ikz dk (4.4.13)˜ψ(k) = Ce −( k−k 02∆k ) 2 (4.4.14)The above represents a Gaussian-tapered function with a spread <strong>of</strong> wavenumbers centeredaround k = k 0 . It can be Fourier transformed in closed form. First, we find the Fouriertransform <strong>of</strong> a Gaussian, namely,∫ ∞∫ ∞∫ ∞ψ g (z) = e −k2 e ikz dk = e −(k−iz/2)2 e −z2 /4 dk = e −z2 /4e −η2 dη (4.4.15)−∞−∞The second equality follows by completing the square <strong>of</strong> the exponent. The last equalityfollows by a change <strong>of</strong> variable by letting k − iz/2 = η. The integr<strong>at</strong>ion <strong>of</strong> a Gaussianfunction can be obtained in closed form. 3 Subsequently,−∞ψ g (z) = e −z2 /4 √ π (4.4.16)proving the fact th<strong>at</strong> the Fourier transform <strong>of</strong> a Gaussian is a Gaussian. The scaling property<strong>of</strong> Fourier transform says th<strong>at</strong>ψ s (z) =∫ ∞−∞˜ψ(k/a)e ikz dk = a∫ ∞−∞˜ψ(k/a)e i(k/a)az d(k/a) = aψ(az) (4.4.17)The above implies th<strong>at</strong> if ψ(z) is the Fourier transform <strong>of</strong> ˜ψ(k), then aψ(az) is the Fouriertransform <strong>of</strong> ˜ψ(k/a). This together with the shifting property <strong>of</strong> Fourier transform allows usto deduce th<strong>at</strong> the Fourier transform <strong>of</strong> (4.4.14) isψ(z) = 2∆kCe −(∆kz)2 +ik 0z(4.4.18)3 The integral can be evalu<strong>at</strong>ed in closed form by using the identity th<strong>at</strong> I = ∫ ∞−∞ dxe−x2 = √ π. This canbe proved by noticing th<strong>at</strong> I 2 = ∫ ∞ ∫ ∞−∞ dxe−x2 −∞ dye−y2 = ∫∫ ∞−∞ dxdye−(x2 +y 2) = 2π ∫ ∞0 ρdρe−ρ2 = π.The last integral can be integr<strong>at</strong>ed in closed form.


Time-Dependent Schrödinger Equ<strong>at</strong>ion 37The probability density functions are proportional to the magnitude squared <strong>of</strong> the above. Thefirst pulse, whose probability density function is given by | ˜ψ(k)| 2 = C 2 e −(k−k0)2 /[2(∆k) 2] , hasstandard devi<strong>at</strong>ion <strong>of</strong> ∆k. 4 The standard devi<strong>at</strong>ion <strong>of</strong> |ψ(z)| 2 ∝ e −(2∆kz)2 /2 is ∆z = 1/(2∆k).The product <strong>of</strong> these two standard devi<strong>at</strong>ions yields∆k∆z = 1 2(4.4.19)or∆p∆z = 2(4.4.20)4.4.2 Particle CurrentIt turns out th<strong>at</strong> even though the st<strong>at</strong>e <strong>of</strong> an electron is defined by a wavefunction, otherequ<strong>at</strong>ions do not see this wavefunction. For instance, Maxwell’s equ<strong>at</strong>ions, will produceelectric field from sources, but they will only produce the electric field from the charge cloudand the current produced by the charge cloud. We see th<strong>at</strong> for st<strong>at</strong>ionary st<strong>at</strong>es <strong>of</strong> a trappedelectron in a potential well, the charge cloud is st<strong>at</strong>ic. Hence, it will not radi<strong>at</strong>e accordingto electromagnetic theory. This resolves the conflict in early days as to why the electron,supposedly “orbiting” around the nucleus <strong>of</strong> an <strong>at</strong>om, does not radi<strong>at</strong>e.However, when an electron is in more than one st<strong>at</strong>ionary st<strong>at</strong>e, the charge cloud is timevarying, and can potentially couple to an external electric field and radi<strong>at</strong>e. 5 For conserv<strong>at</strong>ion<strong>of</strong> charge, we have the rel<strong>at</strong>ion th<strong>at</strong>∂ρ p∂t = −∇ · J p (4.4.21)where ρ p is the particle density, and J p is the particle current density. The particle densityρ p (r, t) = |ψ(r, t)| 2 . We can take the time deriv<strong>at</strong>ive <strong>of</strong> ρ p to yield∂∂t [ψ∗ (r, t)ψ(r, t)] = ∂ψ∗ (r, t)ψ(r, t) + ψ ∗ ∂ψ(r, t)(r, t)∂t∂t(4.4.22)We can use Schrödinger equ<strong>at</strong>ion to replace the time deriv<strong>at</strong>ives on the right-hand side to get∂∂t [ψ∗ (r, t)ψ(r, t)] = − i (ψ ∗ Ĥψ −ψĤ∗ ψ ∗) (4.4.<strong>23</strong>)Substituting in for the definition <strong>of</strong> the Hamiltonian oper<strong>at</strong>or, we have further th<strong>at</strong>∂∂t [ψ∗ (r, t)ψ(r, t)] = i (ψ ∗ ∇ 2 ψ − ψ∇ 2 ψ ∗) (4.4.24)2m∂∂t [ψ∗ (r, t)ψ(r, t)] = − i2m ∇ · (ψ∇ψ∗ − ψ ∗ ∇ψ) (4.4.25)4 A Gaussian PDF <strong>of</strong> the form Ce −x2 /(2σ 2) has a standard devi<strong>at</strong>ion <strong>of</strong> σ.5 We will learn how to tre<strong>at</strong> such coupling l<strong>at</strong>er on.


38 Quantum Mechanics Made SimpleHence, we can define the particle current asJ p = i2m (ψ∇ψ∗ − ψ ∗ ∇ψ) (4.4.26)When a st<strong>at</strong>ionary eigenst<strong>at</strong>e is substituted into the above, due to the product <strong>of</strong> the functionand its complex conjug<strong>at</strong>e above, the particle current becomes time independent. Hence,st<strong>at</strong>ionary st<strong>at</strong>es can only give rise to non-time-varying current (st<strong>at</strong>ic or DC current), andsuch current does not radi<strong>at</strong>e according to electromagnetic theory.


40 Quantum Mechanics Made SimpleMoreover, an inner product space which is complete is a Hilbert space. Hilbert spaces can beinfinite dimensional. The above inner product facilit<strong>at</strong>es the definition <strong>of</strong> a norm since 〈f|f〉is a positive definite number. Hence, the norm <strong>of</strong> a vector can be defined to be‖f‖ = (〈f|f〉) 1/2 (5.1.3)It is the measure <strong>of</strong> the length <strong>of</strong> the vector. We can use this norm to define the distancebetween two vectors to bed(f, g) = ‖f − g‖ (5.1.4)In general, if we have a set <strong>of</strong> orthonormal eigenfunctions, {ψ n (x), n = 1, . . . ∞}, th<strong>at</strong>spans a linear vector space, we can expand an arbitrary function in the same space asg(x) = ∑ nd n ψ n (x) (5.1.5)The set {ψ n (x), n = 1, . . . ∞} also forms the orthonormal basis or the orthonormal basis setfor spanning the vector space. A member <strong>of</strong> the set is known as a basis function or a basisvector. 3 Eigenfunctions <strong>of</strong> an oper<strong>at</strong>or can be used as basis functions.The above can be written using Dirac’s not<strong>at</strong>ion as|g〉 = ∑ nd n |ψ n 〉 (5.1.6)Inner product the above with 〈ψ m | from the left, we arrive <strong>at</strong> th<strong>at</strong>〈ψ m |g〉 = ∑ nd n 〈ψ m |ψ n 〉 (5.1.7)Using the orthonormality <strong>of</strong> the eigenfunction such th<strong>at</strong> 〈ψ m |ψ n 〉 = δ nm , the above yieldsth<strong>at</strong>Consequently, we haved m = 〈ψ m |g〉 (5.1.8)|g〉 = ∑ n|ψ n 〉〈ψ n |g〉 (5.1.9)We can identify the oper<strong>at</strong>orÎ = ∑ n|ψ n 〉〈ψ n | (5.1.10)as an identity oper<strong>at</strong>or since when it oper<strong>at</strong>es on a vector, it returns the same vector. Wecan <strong>of</strong>ten construct an identity oper<strong>at</strong>or <strong>of</strong> a vector space once we have identify the set<strong>of</strong> orthonormal vectors th<strong>at</strong> span the space. The product |ψ n 〉〈φ n | in Dirac’s not<strong>at</strong>ion, is3 We will use “function” and “vector” interchangeably since they are the same.


More M<strong>at</strong>hem<strong>at</strong>ical Preliminaries 41analogous to outer product between two vectors u · v † in linear algebraic not<strong>at</strong>ion. Thisouter product produces a m<strong>at</strong>rix, and in the case <strong>of</strong> infinite dimensional linear vector space,produces an oper<strong>at</strong>or.For instance, if we are in a 3-space (the short for 3D vector space), the unit vectors th<strong>at</strong>span the 3-space are {â 1 , â 2 , â 3 }. They are complete and orthonormal. Hence, the identityoper<strong>at</strong>or in a 3-space is 4 3∑I = â i â i (5.1.11)i=1To be strictly correct, the above is an outer product between two vectors, and a transpose signshould be <strong>at</strong>tached to one <strong>of</strong> the unit vectors. But it is customary in the physics liter<strong>at</strong>ure toignore this transpose sign. Written in terms <strong>of</strong> x, y, z not<strong>at</strong>ion, the identity oper<strong>at</strong>or becomesI = ˆxˆx + ŷŷ + ẑẑ (5.1.12)One can easily be convinced th<strong>at</strong> I · a = a, confirming the above. 5In an N dimensional vector space spanned by a set <strong>of</strong> orthonormal vectors {U n , n =1, . . . , N}, the identity oper<strong>at</strong>or is formed by their outer products; namely,I =N∑u n · u † n (5.1.13)n=1One can easily show th<strong>at</strong> when I oper<strong>at</strong>es on a member <strong>of</strong> the vector space, it returns thesame member. Namely,I · v =N∑u n · u † n · v =n=1N∑ (u n u†n · v ) = v (5.1.14)The last equality follows because the second last term is just an orthonormal eigenvectorexpansion <strong>of</strong> the vector v.A vector space is also defined by a set <strong>of</strong> axioms. For u, v, w th<strong>at</strong> belong to a vectorspace, the following axioms hold:Associ<strong>at</strong>ivity <strong>of</strong> addition u + (v + w) = (u + v) + w (5.1.15)Commut<strong>at</strong>ivity <strong>of</strong> addition u + v = v + u (5.1.16)Identity element <strong>of</strong> addition v + 0 = v (5.1.17)Inverse elements <strong>of</strong> addition v + (−v) = 0 (5.1.18)Distributivity <strong>of</strong> scalar multiplic<strong>at</strong>ion over vectors a(u + v) = au + av (5.1.19)Distributivity <strong>of</strong> scalar multiplic<strong>at</strong>ion by a vector (a + b)v = av + bv (5.1.20)Comp<strong>at</strong>ibility <strong>of</strong> scalar multiplic<strong>at</strong>ion a(bv) = (ab)v (5.1.21)Identity element <strong>of</strong> scalar multiplic<strong>at</strong>ion 1v = v (5.1.22)4 Throughout these lecture notes, the h<strong>at</strong> (“ˆ”) symbol is used to denote an oper<strong>at</strong>or, but here, it denotesa unit vector.5 It is to be noted th<strong>at</strong> an outer product in m<strong>at</strong>rix not<strong>at</strong>ion is u · v † , while in physics not<strong>at</strong>ion for 3 space,it is <strong>of</strong>ten just written as EB. In Dirac not<strong>at</strong>ion, an outer product is |ψ〉〈φ|.n=1


42 Quantum Mechanics Made Simple5.2 Oper<strong>at</strong>orsAn oper<strong>at</strong>or maps vectors from one space to vectors in another space. It is denoted m<strong>at</strong>hem<strong>at</strong>icallyas  : V → W where  is the oper<strong>at</strong>or, while V and W are two different vectorspaces: V is the domain space, while W is the range space. In linear algebra, the oper<strong>at</strong>or isa m<strong>at</strong>rix oper<strong>at</strong>or. In Hilbert spaces, it can be a differential oper<strong>at</strong>or such asg(x) = d f(x) (5.2.1)dxIt can be an integral oper<strong>at</strong>or such as a Fourier transform oper<strong>at</strong>orIn Dirac not<strong>at</strong>ion,g(k) =Linear oper<strong>at</strong>ors are defined such th<strong>at</strong>∫ ∞−∞dxe ikx f(x) (5.2.2)|g〉 = Â|f〉 (5.2.3)ˆL(c 1 |g 1 〉 + c 2 |g 2 〉) = c 1 ˆL|g1 〉 + c 2 ˆL|g2 〉 (5.2.4)It is quite clear th<strong>at</strong> m<strong>at</strong>rix oper<strong>at</strong>ors s<strong>at</strong>isfy the above, and hence, they are linear. Theyare also known as linear maps. In general, like m<strong>at</strong>rix oper<strong>at</strong>ors, linear oper<strong>at</strong>ors are notcommut<strong>at</strong>ive; namely ˆB ≠ˆB (5.2.5)5.2.1 M<strong>at</strong>rix Represent<strong>at</strong>ion <strong>of</strong> an Oper<strong>at</strong>orAn oper<strong>at</strong>or equ<strong>at</strong>ion can be written as 6 |g〉 = Â|f〉 (5.2.6)We can convert the above into a m<strong>at</strong>rix equ<strong>at</strong>ion by inserting an identity oper<strong>at</strong>or on theright-hand side to give|g〉 = ∑ nÂ|ψ n 〉〈ψ n |f〉 (5.2.7)Furthermore, we can multiply the above from the left by the basis vector 〈ψ m |, m = 1, . . . , ∞to yield 7〈ψ m |g〉 = ∑ n〈ψ m |Â|ψ n〉〈ψ n |f〉, m = 1, . . . , ∞ (5.2.8)6 In the m<strong>at</strong>hem<strong>at</strong>ics liter<strong>at</strong>ure, this is <strong>of</strong>ten just denoted as Âf = g.7 This process is called testing or weighting, and ψ m is called the testing or weighting function.


More M<strong>at</strong>hem<strong>at</strong>ical Preliminaries 43The above is an infinite dimensional m<strong>at</strong>rix equ<strong>at</strong>ion which can be written asg = A · f (5.2.9)where[A]mn = 〈ψm|Â|ψ n〉 (5.2.10)[g] m= 〈ψ m |g〉 (5.2.11)[f] n= 〈ψ n |f〉 (5.2.12)The m<strong>at</strong>rix equ<strong>at</strong>ion can be solved approxim<strong>at</strong>ely by trunc<strong>at</strong>ing its size to N × N, or withouttrunc<strong>at</strong>ion, it can be solved iter<strong>at</strong>ively.As a simple example <strong>of</strong> an iter<strong>at</strong>ive solution, we write A = D+T where D is the diagonalpart <strong>of</strong> A while T is its <strong>of</strong>f-diagonal part. Then (5.2.9) can be rewritten asThe above can be solved iter<strong>at</strong>ively asD · f = g − T · f (5.2.13)D · f n = g − T · f n−1 , n = 0, 1, 2, . . . , (5.2.14)until convergence is reached, or th<strong>at</strong> f n does not change with n.The m<strong>at</strong>rix denoted byA mn = 〈ψ m |Â|ψ n〉 (5.2.15)is the m<strong>at</strong>rix represent<strong>at</strong>ion <strong>of</strong> the oper<strong>at</strong>or Â. By the same token, 〈ψ m|g〉 and 〈ψ n |f〉 arethe vector represent<strong>at</strong>ions <strong>of</strong> the functions g and f respectively.In the above, we have assumed th<strong>at</strong> the range space and the domain space <strong>of</strong> the oper<strong>at</strong>orare the same, and hence, they can be spanned by the same basis set. For a Hermitian oper<strong>at</strong>or,this is usually the case. However, for some oper<strong>at</strong>ors where the range space and the domainspace are different, we may choose to test (5.2.7) with a different set <strong>of</strong> basis functions.5.2.2 Bilinear Expansion <strong>of</strong> an Oper<strong>at</strong>orWe have seen how the use <strong>of</strong> the identity oper<strong>at</strong>or allows us to expand a function in terms<strong>of</strong> a set <strong>of</strong> basis functions as in (5.1.9). The same can be done with an oper<strong>at</strong>or. Pre- andpost-multiply an oper<strong>at</strong>or with the identity oper<strong>at</strong>or as given by (5.1.10), we have = ∑ ∑|ψ n 〉〈ψ n |Â|ψ m〉〈ψ m | (5.2.16)n mThe above can be rewritten as = ∑ ∑|ψ n 〉A nm 〈ψ m | = ∑ ∑A nm |ψ n 〉〈ψ m | (5.2.17)n mn mwhere A nm is the m<strong>at</strong>rix represent<strong>at</strong>ion <strong>of</strong> the oper<strong>at</strong>or Â. The above is the bilinear expansion<strong>of</strong> the oper<strong>at</strong>or in terms <strong>of</strong> orthonormal functions. Notice th<strong>at</strong> the expansion <strong>of</strong> an identityoper<strong>at</strong>or given by (5.1.10) is a bilinear expansion.


44 Quantum Mechanics Made Simple5.2.3 Trace <strong>of</strong> an Oper<strong>at</strong>orThe trace <strong>of</strong> a m<strong>at</strong>rix oper<strong>at</strong>or is defined to be the sum <strong>of</strong> its diagonal elements; namely,tr ( M ) =N∑M ii (5.2.18)If an oper<strong>at</strong>or  has m<strong>at</strong>rix represent<strong>at</strong>ion given by 〈ψi|Â|ψ j〉, the trace <strong>of</strong> the oper<strong>at</strong>or  isdefined to be)tr( = ∑ 〈ψ i |Â|ψ i〉 (5.2.19)ii=1It can be shown th<strong>at</strong> the trace <strong>of</strong> an oper<strong>at</strong>or is independent <strong>of</strong> the basis used for its represent<strong>at</strong>ion.To this end, we insert the identity oper<strong>at</strong>orÎ = ∑ m|φ m 〉〈φ m | (5.2.20)into (5.2.19) to get)tr( = ∑ i∑〈ψ i |φ m 〉〈φ m |Â|ψ i〉 (5.2.21)mExchanging the order <strong>of</strong> summ<strong>at</strong>ion above, and the order <strong>of</strong> the two scalar numbers in thesummand, we have)tr( = ∑ ∑〈φ m |Â|ψ i〉〈ψ i |φ m 〉 (5.2.22)miThe inner summ<strong>at</strong>ion reduces to an identity oper<strong>at</strong>or which can be removed, and the abovebecomes)tr( = ∑ 〈φ m |Â|φ m〉 (5.2.<strong>23</strong>)mThis is the trace <strong>of</strong>  using another basis set th<strong>at</strong> is complete and orthonormal. Hence, thetrace <strong>of</strong> an oper<strong>at</strong>or is invariant with respect to the choice <strong>of</strong> basis.If we choose a basis function th<strong>at</strong> is the eigenfunction <strong>of</strong>  such th<strong>at</strong> Â|ψ i〉 = λ i |ψ i 〉, then(5.2.19) becomes)tr( = ∑ λ i 〈ψ i |ψ i 〉 = ∑ λ i (5.2.24)iiHence, the trace <strong>of</strong> an oper<strong>at</strong>or is also the sum <strong>of</strong> its eigenvalues.It can also be shown th<strong>at</strong>)tr( ˆB = tr( ˆBÂ) (5.2.25)


More M<strong>at</strong>hem<strong>at</strong>ical Preliminaries 45This is quite easy to show for m<strong>at</strong>rix oper<strong>at</strong>ors sincetr ( A · B ) = ∑ ∑A ij B ji = ∑i jj∑B ji A ij (5.2.26)iTrace is usually used in quantum mechanics as an altern<strong>at</strong>ive way to write the expect<strong>at</strong>ionvalue <strong>of</strong> an oper<strong>at</strong>or. As mentioned before, for a quantum system in a st<strong>at</strong>e defined by thest<strong>at</strong>e function |ψ〉, the expect<strong>at</strong>ion value <strong>of</strong> a quantum oper<strong>at</strong>or in such a st<strong>at</strong>e is〈ψ|Â|ψ〉 (5.2.27)For denumerable indices, the above is analogous tou † · A · u = ∑ ∑u ∗ nA nm u m = ∑ ∑A nm u m u ∗ n = tr ( A · u · u †) (5.2.28)n mn mwhere u m u ∗ n is the outer product <strong>of</strong> two vectors u m and u n . Converting the above to Diracnot<strong>at</strong>ion, we haveThe oper<strong>at</strong>or〈ψ|Â|ψ〉 = tr{Â|ψ〉〈ψ|} (5.2.29)ˆρ = |ψ〉〈ψ| (5.2.30)is known as the density oper<strong>at</strong>or. It is used in quantum mechanics to denote the st<strong>at</strong>e <strong>of</strong> aquantum system as an altern<strong>at</strong>ive to the st<strong>at</strong>e vector.5.2.4 Unitary Oper<strong>at</strong>orsUnitary oper<strong>at</strong>ors s<strong>at</strong>isfy the property th<strong>at</strong>U † · U = I (5.2.31)Unitary oper<strong>at</strong>ors are oper<strong>at</strong>ors th<strong>at</strong> when acted on a vector does not change its length. Inm<strong>at</strong>rix not<strong>at</strong>ion, a unitary transform isThe length squared <strong>of</strong> the new vector v ′ is defined to beMaking use <strong>of</strong> (5.2.31), the above implies th<strong>at</strong>U · v = v ′ (5.2.32)v † · U † · U · v = v ′† · v ′ (5.2.33)v † · v = v ′† · v ′ (5.2.34)or th<strong>at</strong> the length <strong>of</strong> the vector has not changed. Furthermore, it can be easily shown th<strong>at</strong>unitary transform<strong>at</strong>ions preserve the inner product between two vectors, namely,v † · u = v ′† · u ′ (5.2.35)


46 Quantum Mechanics Made SimpleFor unitary m<strong>at</strong>rix, it is clear th<strong>at</strong> U † = U −1 . The above can be rewritten using Diracnot<strong>at</strong>ion. Since in quantum mechanics, 〈ψ|ψ〉 = 1, the length <strong>of</strong> the vector is 1. The timeevolution oper<strong>at</strong>or by integr<strong>at</strong>ing Schrödinger equ<strong>at</strong>ion given belowˆτ = e − i Ĥt (5.2.36)where Ĥ is the Hamiltonian oper<strong>at</strong>or, is a unitary oper<strong>at</strong>or, since the st<strong>at</strong>e vector it acts oncannot change its length.Since many basis set comprises orthonormal basis vectors, a unitary m<strong>at</strong>rix is needed forthe change <strong>of</strong> basis from one set to another. For example, we can define a Fourier transformto bef(x) = √ 1 ∫ ∞dk e ikx ˜f(k).2π−∞It is used to transform the momentum represent<strong>at</strong>ion <strong>of</strong> a wavefunction to the coordin<strong>at</strong>erepresent<strong>at</strong>ion. Using the above, we can show th<strong>at</strong>∫∫dx f ∗ (x) g(x) = dk ˜f ∗ (k) ˜g(k).The above can be proved using the fact th<strong>at</strong>12π∫ ∞−∞e i(k′ −k)x dx = δ(k ′ − k)Hence, a Fourier transform oper<strong>at</strong>or is a unitary oper<strong>at</strong>or th<strong>at</strong> preserves inner productsbetween two vectors and their lengths.5.2.5 Hermitian Oper<strong>at</strong>orsHermitian oper<strong>at</strong>ors appear frequently in quantum mechanics since they have real eigenvalues.As such, the expect<strong>at</strong>ion value <strong>of</strong> such oper<strong>at</strong>ors are real valued, so th<strong>at</strong> it can be connectedto measurable or observable quantities in the real world.A Hermitian m<strong>at</strong>rix is one where its conjug<strong>at</strong>e transpose (also called Hermitian transposeor Hermitian conjug<strong>at</strong>e) is itself. For oper<strong>at</strong>ors, the term “adjoint” is <strong>of</strong>ten used. A Hermitianoper<strong>at</strong>or is also called a self-adjoint oper<strong>at</strong>or. The not<strong>at</strong>ion for this isˆM † = ˆM (5.2.37)For m<strong>at</strong>rix oper<strong>at</strong>or with denumerable indices, the above is the same asM ∗ ij = M ji (5.2.38)For oper<strong>at</strong>ors with nondenumerable indices, it is better to define the adjoint <strong>of</strong> the oper<strong>at</strong>orvia inner products. Using the rule for conjug<strong>at</strong>e transpose, we have(a† · M · b ) ∗= b† · M † · a (5.2.39)


More M<strong>at</strong>hem<strong>at</strong>ical Preliminaries 47Generalizing the above to infinite dimensional space, we have〈f| ˆM|g〉 ∗ = 〈f| ˆM|g〉 † = 〈g| ˆM † |f〉 (5.2.40)The first equality follows because the above are scalar quantities: hence, the conjug<strong>at</strong>e transpose<strong>of</strong> a scalar is the same as its conjug<strong>at</strong>ion. To obtain the second equality, we have usedthe rule th<strong>at</strong>(A · B · C) t= Ct· Bt· At(5.2.41)in linear algebra, where the m<strong>at</strong>rices need not be square. Hence, A and C can be 1 × N andN × 1 respectively representing vectors. This identity immedi<strong>at</strong>ely implies th<strong>at</strong>(A · B · C) †= C†· B†· A†(5.2.42)These rules can easily be shown to apply to inner product <strong>of</strong> oper<strong>at</strong>ors and vectors underthe Dirac not<strong>at</strong>ion. Hence, (5.2.40) defines the adjoint <strong>of</strong> the oper<strong>at</strong>or ˆM † when f and g arearbitrary. It does so by using just inner products.A Hermitian oper<strong>at</strong>or has real eigenvalues and their eigenvectors are orthogonal. Thepro<strong>of</strong> is analogous to the pro<strong>of</strong> for m<strong>at</strong>rix oper<strong>at</strong>ors, for which we will provide next. GivenM · v i = λ i v i (5.2.43)M · v j = λ j v j (5.2.44)Dot-multiply the first equ<strong>at</strong>ion by v † jfrom the left, and likewise for the second equ<strong>at</strong>ion withv † i , we have v † j · M · v i = λ i v † j · v i (5.2.45)v † i · M · v j = λ j v † i · v j (5.2.46)We can take the conjug<strong>at</strong>e transpose <strong>of</strong> the first equ<strong>at</strong>ion, using the rule in (5.2.42), the lefthandside becomes the same as th<strong>at</strong> for the second equ<strong>at</strong>ion after making use <strong>of</strong> the Hermitianproperty <strong>of</strong> the m<strong>at</strong>rix M. On subtracting the two equ<strong>at</strong>ions, the following equ<strong>at</strong>ion ensues:0 = (λ ∗ i − λ j )v † i · v j (5.2.47)If i = j, on the right-hand side, we have v † i · v i = |v i | 2 which is a positive definite number.The above can be zero only if λ ∗ i = λ i or th<strong>at</strong> λ i is real. On the other hand, if λ i ≠ λ j , it isnecessary th<strong>at</strong> v † i · v j = 0. In other words,v † i · v j = C n δ ij (5.2.48)The above pro<strong>of</strong> can be repe<strong>at</strong>ed using Dirac not<strong>at</strong>ion, and the conclusion will be the same.The eigenvectors <strong>of</strong> a Hermitian oper<strong>at</strong>or are also complete in the space th<strong>at</strong> the oper<strong>at</strong>or actson. It is obvious in the finite dimensional case, but not so obvious in the infinite dimensionalspace.


48 Quantum Mechanics Made SimpleWe can use the rule expressed in (5.2.40) to see if an oper<strong>at</strong>or is Hermitian. For instance,the momentum oper<strong>at</strong>or is ˆp. Using (5.2.40), we have−∞〈f|ˆp|g〉 ∗ = 〈g|ˆp † |f〉 (5.2.49)The above defines the adjoint <strong>of</strong> the oper<strong>at</strong>or ˆp = −id/dx. Writing the above explicitly in1D space using coordin<strong>at</strong>e space represent<strong>at</strong>ion, 8 we have on the left-hand side∫ ∞dxf(x)i d ∫ ∞(dx g∗ (x) = dxg ∗ (x) −i d )f(x) (5.2.50)dxWe have arrived <strong>at</strong> the form <strong>of</strong> the right-hand side by using integr<strong>at</strong>ion by parts, and assumingth<strong>at</strong> the functions are vanishing <strong>at</strong> infinity. By comparing the above, we identify th<strong>at</strong>−∞ˆp † = −i ddx(5.2.51)Therefore, we note th<strong>at</strong> (ˆp) † = ˆp implying th<strong>at</strong> it is Hermitian or self-adjoint. The eigenfunction<strong>of</strong> the momentum oper<strong>at</strong>or is f k (x) ∝ e ikx . It is quite clear th<strong>at</strong> ˆpf k (x) = kf k (x), andhence, its eigenvalues are also real, as is required <strong>of</strong> a Hermitian oper<strong>at</strong>or.The above can be generalized to 3D. It can also be shown th<strong>at</strong> the kinetic energy oper<strong>at</strong>orˆT = ˆp 2 /2m = − ˆ∇ 22m(5.2.52)is Hermitian. In 3D, ˆp 2 is proportional to the Laplacian oper<strong>at</strong>or ∇ 2 in coordin<strong>at</strong>e represent<strong>at</strong>ion.We can use integr<strong>at</strong>ion by parts in 3D to show th<strong>at</strong> the above oper<strong>at</strong>or is alsoHermitian. Using the fact th<strong>at</strong>∇ · [f(r)∇g ∗ (r)] = f(r)∇ 2 g ∗ (r) + ∇f(r)∇g ∗ (r) (5.2.53)we have in coordin<strong>at</strong>e represent<strong>at</strong>ion∫∫∫drf(r)∇ 2 g ∗ (r) = dr [∇ · (f(r)∇g ∗ (r)) − ∇f(r)∇g ∗ (r)] = −VVVdr∇f(r)∇g ∗ (r)(5.2.54)The last form is symmetric between f and g, and hence, we can easily show th<strong>at</strong>∫∫drg ∗ (r)∇ 2 f(r) = − dr∇g ∗ (r)∇f(r) (5.2.55)VConsequently, we can show th<strong>at</strong>[∫] ∗〈f| ˆ∇ 2 |g〉 ∗ = drf ∗ (r)∇ 2 g(r)V∫= drg ∗ (r)∇ 2 f(r)VV= 〈g| ˆ∇ 2 |f〉 (5.2.56)8 A note is in order here regarding the term “coordin<strong>at</strong>e space”, since there is only one Hilbert space. Itis understood th<strong>at</strong> when the term “coordin<strong>at</strong>e space represent<strong>at</strong>ion” is used, it means th<strong>at</strong> the Hilbert spaceis represented in the space where the basis is the coordin<strong>at</strong>e space basis. It is <strong>of</strong>ten just called “coordin<strong>at</strong>erepresent<strong>at</strong>ion”.


More M<strong>at</strong>hem<strong>at</strong>ical Preliminaries 49indic<strong>at</strong>ing the Hermitian property <strong>of</strong> the Laplacian oper<strong>at</strong>or. Consequently, the ˆT oper<strong>at</strong>oris also Hermitian or self-adjoint. If we have chosen f = g in the above, (5.2.55) is alwaysneg<strong>at</strong>ive, implying th<strong>at</strong> the Laplacian oper<strong>at</strong>or is a neg<strong>at</strong>ive definite oper<strong>at</strong>or.The definition <strong>of</strong> (5.2.40) can be extended to defining the transpose <strong>of</strong> an oper<strong>at</strong>or. Form<strong>at</strong>rix algebra,(a† · M · b ) t= bt · M t · a ∗ (5.2.57)Generalizing the above to infinite dimensional space, we havewhich is the defining equ<strong>at</strong>ion for the transpose <strong>of</strong> an oper<strong>at</strong>or.〈f| ˆM|g〉 t = 〈g ∗ | ˆM t |f ∗ 〉 (5.2.58)5.3 *Identity Oper<strong>at</strong>or in a Continuum Space 9We have seen the deriv<strong>at</strong>ion <strong>of</strong> the identity oper<strong>at</strong>or when the basis functions are indexedby integers, such as the eigenfunctions <strong>of</strong> an infinite potential well, or those <strong>of</strong> a harmonicoscill<strong>at</strong>or in a parabolic potential well. In many situ<strong>at</strong>ions, we are required to work withindices th<strong>at</strong> are a continuum (nondenumerable) like the coordin<strong>at</strong>e space indices x, y, z. Whenan electron is freely roaming, its energy values are also a continuum.We have seen how we have used denumerable eigenfunctions, such as ψ n to project afunction into its vector represent<strong>at</strong>ion. Namely,f n = 〈ψ n |f〉 (5.3.1)Very <strong>of</strong>ten, a denumerable basis function is just written as 〈n|, and the above becomesf n = 〈n|f〉 (5.3.2)Similar, we can think <strong>of</strong> a coordin<strong>at</strong>e (or position) basis function 〈p x | whose property is th<strong>at</strong>its inner product with |f〉 yields the value <strong>of</strong> the function f <strong>at</strong> position x; namely,The above is <strong>of</strong>ten abbrevi<strong>at</strong>ed asf(x) = 〈p x |f〉 (5.3.3)f(x) = 〈x|f〉 (5.3.4)Assuming th<strong>at</strong> this basis |x〉 is complete and orthogonal in 0 < x < a, then we can definean identity oper<strong>at</strong>or such th<strong>at</strong>I =∫ a9 Sections with * may be skipped on the first reading.0dx ′ |x ′ 〉〈x ′ | (5.3.5)


50 Quantum Mechanics Made Simpleso th<strong>at</strong>|f〉 =∫ a0dx ′ |x ′ 〉〈x ′ |f〉 =∫ a0dx ′ |x ′ 〉f(x ′ ) (5.3.6)Notice th<strong>at</strong> the above is quite different from the identity oper<strong>at</strong>or when the basis functionsare denumerable, which isI =∞∑|f n 〉〈f n | (5.3.7)n=1Taking the product <strong>of</strong> (5.3.6) with 〈x|, we have〈x|f〉 = f(x) =∫ a0dx ′ 〈x|x ′ 〉f(x ′ ) (5.3.8)Hence, in order for (5.3.5) to be an identity oper<strong>at</strong>or, the basis function must s<strong>at</strong>isfy〈x|x ′ 〉 = δ(x − x ′ ) (5.3.9)Notice th<strong>at</strong> in the denumerable case, the orthonormal rel<strong>at</strong>ion <strong>of</strong> the basis function is expressedin terms <strong>of</strong> Kronecker delta function δ ij , but in the continuum case, the equivalentrel<strong>at</strong>ionship is expressed in terms <strong>of</strong> the Dirac delta function δ(x − x ′ ).We note th<strong>at</strong> |x〉 is not orthonormal in the strict sense th<strong>at</strong> (5.3.9) is infinite when x = x ′ .The identity oper<strong>at</strong>or (5.3.5) is different from (5.3.7) because <strong>of</strong> the extra weight dx in theintegral summ<strong>at</strong>ion. One may think th<strong>at</strong> |x〉 √ dx ′ as analogous to the orthonormal vector <strong>of</strong>the countable case. So (5.3.5) can be thought <strong>of</strong> as the limiting case <strong>of</strong> a countable summ<strong>at</strong>ionI =N∑ √∆x|xi 〉〈x i | √ ∆x (5.3.10)i=1as ∆x → 0, so the orthonormal basis is actually √ ∆x|x i 〉. Similar, (5.3.9) as the limitingcase <strong>of</strong>The inner product between two vectors is written as〈f|g〉 =∫ a〈x i |x j 〉 = δ ij /∆x (5.3.11)dx〈f|x〉〈x|g〉 =∫ adxf ∗ (x)g(x) (5.3.12)00where we have inserted the identity oper<strong>at</strong>or (5.3.5) in the first expression above to get thesecond expression, and making use <strong>of</strong> (5.3.4) to get the third expression. Furthermore, wenote th<strong>at</strong> 〈f|x〉 is the complex conjug<strong>at</strong>e <strong>of</strong> 〈x|f〉, since 〈f| and |x〉 are conjug<strong>at</strong>e transpose<strong>of</strong> |f〉 and 〈x|, respectively.


More M<strong>at</strong>hem<strong>at</strong>ical Preliminaries 51A vector |f〉 can have other represent<strong>at</strong>ions. For example, we can define a set <strong>of</strong> “orthonormal”vectors 〈k| such th<strong>at</strong>〈k|f〉 = f(k), (5.3.13)where f(k) is the Fourier transform <strong>of</strong> f(x) via the rel<strong>at</strong>ionshipThe above can be written asf(x) = 1 √2πf(x) = 〈x|f〉 =where we have defined the identity oper<strong>at</strong>orI =∫ ∞−∞∫∞−∞∫ ∞−∞Comparing (5.3.13), (5.3.14) and (5.3.15), we deduce th<strong>at</strong>dke ikx f(k) (5.3.14)dk〈x|k〉〈k|f〉 (5.3.15)dk|k〉〈k| (5.3.16)〈x|k〉 = 1 √2πe ikx (5.3.17)In other words, the coordin<strong>at</strong>e represent<strong>at</strong>ion <strong>of</strong> the vector |k〉 is given by (5.3.17). Sincek in quantum mechanics is rel<strong>at</strong>ed to the momentum <strong>of</strong> a particle, f(k) = 〈k|f〉 is alsocalled the momentum represent<strong>at</strong>ion <strong>of</strong> the vector |f〉, while f(x) = 〈x|f〉 is the coordin<strong>at</strong>erepresent<strong>at</strong>ion <strong>of</strong> the vector |f〉.From the above, we can derive Parseval’s theorema∫a∫〈f|g〉 = dx〈f|x〉〈x|g〉 = dxf ∗ (x)g(x)00∞∫∞∫(5.3.18)= dk〈f|k〉〈k|g〉 = dkf ∗ (k)g(k)−∞Parseval’s theorem is the st<strong>at</strong>ement th<strong>at</strong> the inner product between two vectors is invariantwith respect to the basis th<strong>at</strong> represent it.The above can be generalized to 3-space where a position basis function is 〈r|, with theproperty th<strong>at</strong>−∞f(r) = 〈r|f〉 (5.3.19)The identity oper<strong>at</strong>or can then be expressed as∫I = dr ′ |r ′ 〉〈r ′ | (5.3.20)And the “orthonormality” rel<strong>at</strong>ionship isSimilar basis can be defined for the 3D k space.V〈r|r ′ 〉 = δ(r − r ′ ) (5.3.21)


52 Quantum Mechanics Made Simple5.4 *Changing Between Represent<strong>at</strong>ionsIf we have an oper<strong>at</strong>or equ<strong>at</strong>ion such asˆp|ψ〉 = |g〉 (5.4.1)The above equ<strong>at</strong>ion is analogous to the equ<strong>at</strong>ion P · f = g in m<strong>at</strong>rix algebra. The explicitforms and the represent<strong>at</strong>ions <strong>of</strong> the m<strong>at</strong>rix oper<strong>at</strong>or P and the vector f are not given. Inorder to express the above explicitly, we first insert the identity oper<strong>at</strong>or to transform (5.4.1)to∫dx ˆp|x〉〈x|ψ〉 = |g〉 (5.4.2)Then multiply the above from the left by 〈x ′ |, we have∫dx〈x ′ |ˆp|x〉〈x|ψ〉 = 〈x ′ |g〉 (5.4.3)∫dxp(x ′ , x)ψ(x) = g(x ′ ) (5.4.4)The above is the coordin<strong>at</strong>e represent<strong>at</strong>ion <strong>of</strong> (5.4.1). It is valid for any oper<strong>at</strong>or in a 1Dspace.5.4.1 Momentum Oper<strong>at</strong>orBut if ˆp is the momentum oper<strong>at</strong>or, we know th<strong>at</strong> the above equ<strong>at</strong>ion in coordin<strong>at</strong>e spacerepresent<strong>at</strong>ion should be−i ddx ψ(x) = g(x′ ) (5.4.5)Hence we conclude th<strong>at</strong> the coordin<strong>at</strong>e represent<strong>at</strong>ion <strong>of</strong> ˆp, which is 〈x ′ |ˆp|x〉 isp(x ′ , x) = −iδ(x ′ − x) ddx(5.4.6)Therefore, a differential oper<strong>at</strong>or is a highly localized oper<strong>at</strong>or, or a quasi-diagonal oper<strong>at</strong>or.Next, we will study the action <strong>of</strong> the momentum oper<strong>at</strong>or ˆp on a momentum eigenst<strong>at</strong>e|k〉; namely,ˆp|k〉 (5.4.7)Transforming the above to coordin<strong>at</strong>e represent<strong>at</strong>ion similar to wh<strong>at</strong> have been done above,we have∫∫(〈x ′ |ˆp|k〉 = dx〈x ′ |ˆp|x〉〈x|k〉 = dxδ(x ′ − x) −i d ) (〈x|k〉 = −i d )dxdx ′ 〈x ′ |k〉 (5.4.8)


More M<strong>at</strong>hem<strong>at</strong>ical Preliminaries 53The expression for 〈x|k〉 is given in equ<strong>at</strong>ion (5.3.17). Hence, the above becomes〈x ′ |ˆp|k〉 = k〈x ′ |k〉 (5.4.9)Therefore, we see th<strong>at</strong> |k〉 is an eigenvector <strong>of</strong> the momentum oper<strong>at</strong>or ˆp. In other words,ˆp|k〉 = k|k〉 (5.4.10)In fact, any function <strong>of</strong> ˆp, when oper<strong>at</strong>ing on a momentum eigenst<strong>at</strong>e givesf(ˆp)|k〉 = f(k)|k〉 (5.4.11)We can repe<strong>at</strong> (5.4.2) and (5.4.3) using the basis |k〉 to arrive <strong>at</strong>∫dk〈k ′ |ˆp|k〉〈k|ψ〉 = 〈k ′ |g〉 (5.4.12)Since |k〉 is an eigenvector <strong>of</strong> ˆp, the above becomes∫∫dkk〈k ′ |k〉〈k|ψ〉 = dkkδ(k ′ − k)〈k|ψ〉 = k ′ 〈k ′ |ψ〉 = 〈k ′ |g〉 (5.4.13)which is just a scalar equ<strong>at</strong>ion. In other words, the momentum represent<strong>at</strong>ion <strong>of</strong> (5.4.1) is5.4.2 Position Oper<strong>at</strong>or〈k ′ |ˆp|ψ〉 = k ′ 〈k ′ |ψ〉 = k ′ ψ(k ′ ) = g(k ′ ) (5.4.14)Similarly, we can show th<strong>at</strong> the position oper<strong>at</strong>or ˆx, when oper<strong>at</strong>ing on a vector, can beexpressed as∫∫ ∫∫ ∫ˆx|ψ〉 = dkˆx|k〉〈k|ψ〉 = dk dxˆx|x〉〈x|k〉〈k|ψ〉 = dk dxx|x〉〈x|k〉〈k|ψ〉 (5.4.15)where we have used ˆx|x〉 = x|x〉 to arrive <strong>at</strong> the last equality. Furthermore, we notice th<strong>at</strong>x|x〉〈x|k〉 = − i dk dk(|x〉〈x|k〉) after making use <strong>of</strong> (5.3.17). Consequently,∫ (ˆx|ψ〉 = dk − i ∫) ∫ (ddx|x〉〈x|k〉 〈k|ψ〉 = dk − i )dk dkk dk |k〉 〈k|ψ〉 (5.4.16)Taking the inner product <strong>of</strong> the above with 〈k ′ |, we have∫ (〈k ′ |ˆx|ψ〉 = dk − i ) ∫ (dk dk 〈k′ |k〉 〈k|ψ〉 = dk − i )dk dk δ(k′ − k) ψ(k)∫( ) i= dkδ(k ′ d− k)k dk ψ(k)(5.4.17)The last equality follows from integr<strong>at</strong>ion by parts. Therefore, the momentum represent<strong>at</strong>ion<strong>of</strong> ˆx|ψ〉, which is just 〈k ′ |ˆx|ψ〉, is〈k ′ |ˆx|ψ〉 = i kddk ′ ψ(k′ ) (5.4.18)


54 Quantum Mechanics Made Simple5.4.3 The Coordin<strong>at</strong>e Basis FunctionUsing the coordin<strong>at</strong>e basis function |r〉, we can write many <strong>of</strong> our previously derived identitiesin coordin<strong>at</strong>e represent<strong>at</strong>ion. Say if the functions defined in (5.1.9) and (5.1.10) are functions<strong>of</strong> 3-space, indexed by r, then multiplying (5.1.9) from the left by 〈r|,〈r|g〉 = g(r) = ∑ n〈r|ψ n 〉〈ψ n |g〉 = ∑ nψ n (r)〈ψ n |g〉 (5.4.19)We can pre- and post-multiply the identity oper<strong>at</strong>or defined in (5.1.10) by 〈r| and |r ′ 〉 respectively,we can identify the oper<strong>at</strong>or〈∣r ∣Î ∣ r ′〉 = 〈r|r ′ 〉 = δ(r − r ′ ) = ∑ 〈r|ψ n 〉〈ψ n |r ′ 〉 = ∑ ψ n (r)ψn(r ∗ ′ ) (5.4.20)nnwhere we have noted th<strong>at</strong> 〈ψ n |r ′ 〉 = 〈r ′ |ψ n 〉 ∗ since they are conjug<strong>at</strong>e transpose <strong>of</strong> each other.The above is the bilinear eigenfunction expansion <strong>of</strong> the Dirac delta function. We can alsoapply the above to the bilinear expansion <strong>of</strong> an oper<strong>at</strong>or given by (5.2.17). Going throughsimilar oper<strong>at</strong>ions, we have〈∣r ∣ ∣ r ′〉 = A(r, r ′ ) = ∑ ∑A nm ψ n (r)ψm(r ∗ ′ ) (5.4.21)n mWe have let A(r, r ′ ) be the coordin<strong>at</strong>e represent<strong>at</strong>ion <strong>of</strong> the oper<strong>at</strong>or Â. The above is justthe bilinear expansion in coordin<strong>at</strong>e represent<strong>at</strong>ion.We can also choose the momentum basis set |k〉 (or any other basis sets) and project thepreviously obtained identities in momentum represent<strong>at</strong>ion (or any other represent<strong>at</strong>ions).5.5 Commut<strong>at</strong>ion <strong>of</strong> Oper<strong>at</strong>orsOper<strong>at</strong>ors, like m<strong>at</strong>rices, are non-commuting for most parts. In other words, ˆB ≠ˆB (5.5.1)The commut<strong>at</strong>or is defined to be[Â, ˆB]=  ˆB −ˆB (5.5.2)If  and ˆB are commuting, thenBut for most parts,[Â, ˆB]= 0 (5.5.3)[Â, ˆB]= iĈ (5.5.4)It can be shown easily th<strong>at</strong> Ĉ is a Hermitian oper<strong>at</strong>or if  and ˆB are Hermitian.


More M<strong>at</strong>hem<strong>at</strong>ical Preliminaries 55If  and ˆB are commuting, they share the same set <strong>of</strong> eigenfunctions, but not the sameeigenvalues. Say if |ψ n 〉 is an eigenfunction <strong>of</strong> ˆB, with eigenvalue B n , thenˆBÂ|ψ n〉 =  ˆB|ψ n 〉 = ÂB n|ψ n 〉 = B n Â|ψ n 〉 (5.5.5)From the above, it is clear th<strong>at</strong> Â|ψ n〉 is also eigenfunction <strong>of</strong> ˆB with the same eigenvalueB n . Therefore, Â|ψ n 〉 is proportional to |ψ n 〉, or th<strong>at</strong> Â|ψ n〉 = F n |ψ n 〉. In other words, |ψ n 〉is also an eigenfunction <strong>of</strong> Â, but with a different eigenvalue.As an example, the position oper<strong>at</strong>or ˆx = x and the momentum oper<strong>at</strong>or ˆp = −id/dxdo not commute with each other. It can be easily shown th<strong>at</strong>[x, ˆp] = iÎ = iTherefore, they cannot share the same set <strong>of</strong> eigenfunctions.5.6 Expect<strong>at</strong>ion Value and Eigenvalue <strong>of</strong> Oper<strong>at</strong>orsThe expect<strong>at</strong>ion value <strong>of</strong> a random variable is a concept from probability. In quantum mechanics,a measurable quantity or an observable in the real world is a random variable. Foreach observable in the real world, there is a corresponding oper<strong>at</strong>or in the quantum world.The expect<strong>at</strong>ion value <strong>of</strong> an oper<strong>at</strong>or for a quantum system in st<strong>at</strong>e |f〉 is defined to be〈Â〉=〈f〉∣∣ ∣ f(5.6.1)It is prudent to understand this expect<strong>at</strong>ion value in terms <strong>of</strong> the eigenst<strong>at</strong>es and eigenvalues<strong>of</strong> Â. Letting|f〉= ∑ n|ψ n 〉f n (5.6.2)then (5.6.1) can be rewritten as〈Â〉= ∑ n∑ 〈 ∣ 〉 ∣∣ ψ n Â∣ ψ m fnf ∗ m (5.6.3)mWhen ψ m is the eigenst<strong>at</strong>e <strong>of</strong>  with eigenvalue A m, the above can be written as〈Â〉= ∑ ∑A m 〈ψ n |ψ m 〉fnf ∗ m = ∑ A n |f n | 2 = ∑ A n P n (5.6.4)n mnnwhere |f n | 2 = P n is the probability <strong>of</strong> finding the st<strong>at</strong>e in eigenst<strong>at</strong>e n. 10 Hence, the above isthe st<strong>at</strong>istical average or the expect<strong>at</strong>ion value <strong>of</strong> the eigenvalue A n . The eigenvalue A n canbe thought <strong>of</strong> as a random variable. Hence, the expect<strong>at</strong>ion value <strong>of</strong> an oper<strong>at</strong>or is the average10 Because in (5.6.2), 〈f|f〉 = 1, it implies th<strong>at</strong> ∑ n |fn|2 = 1. Therefore, |f n| 2 = P n, the probability <strong>of</strong>finding the particle in st<strong>at</strong>e |ψ n〉.


56 Quantum Mechanics Made Simplevalue <strong>of</strong> its eigenvalue when the quantum st<strong>at</strong>e is in st<strong>at</strong>e f. Therefore, it is customary isdenote〈Â〉 〈〉∣= f ∣Â ∣ f = 〈A〉 = Ā (5.6.5)where the scalar variable A denotes the eigenvalue <strong>of</strong> Â, which is a random variable, while theangular brackets over A, or 〈A〉, indic<strong>at</strong>e st<strong>at</strong>istical average. An overbar in the last equalityabove is <strong>of</strong>ten used as a short-hand for angular brackets to denote an average <strong>of</strong> a randomvariable.The above indic<strong>at</strong>es th<strong>at</strong> if we prepare a quantum st<strong>at</strong>e th<strong>at</strong> is exactly the eigenst<strong>at</strong>e <strong>of</strong>the quantum oper<strong>at</strong>or Â, then its expect<strong>at</strong>ion value in this quantum eigenst<strong>at</strong>e is just theeigenvalue <strong>of</strong> the eigenst<strong>at</strong>e. In a sense, we can “measure” or “observe” the eigenvalue <strong>of</strong> thequantum oper<strong>at</strong>or through such an experiment. Namely,〈Â〉 〈 ∣ 〉 ∣∣ = ψ n Â∣ ψ n = A n (5.6.6)where A n is the eigenvalue <strong>of</strong> the eigenst<strong>at</strong>e.We see in the previous section th<strong>at</strong> when two oper<strong>at</strong>ors commute, they share the sameeigenfunctions but with different eigenvalues. Therefore, if we prepare an eigenst<strong>at</strong>e shared bythese two oper<strong>at</strong>ors, their respective eigenvalues can be “measured” exactly. In other words,〈Â〉 〈 ∣ 〉 ∣∣ = ψ n Â∣ ψ n = A n (5.6.7)〈 〈 ∣ 〉ˆB〉∣∣ = ψ n ˆB ∣ ψ n = B n (5.6.8)On the other hand, if the two oper<strong>at</strong>ors do not commute, they do not share the same set<strong>of</strong> eigenst<strong>at</strong>es. If we prepare an eigenst<strong>at</strong>e ψ n th<strong>at</strong> is the eigenst<strong>at</strong>e <strong>of</strong> Â, but it is not theeigenst<strong>at</strong>e <strong>of</strong> ˆB. However, we can expand the eigenst<strong>at</strong>e ψ n in terms <strong>of</strong> the eigenst<strong>at</strong>es <strong>of</strong> theoper<strong>at</strong>or ˆB; namely,|ψ n 〉 = ∑ ia i |φ i 〉 (5.6.9)where |φ i 〉 are the eigenst<strong>at</strong>es <strong>of</strong> ˆB. We now see th<strong>at</strong> the expect<strong>at</strong>ion value <strong>of</strong> ˆB due to theeigenst<strong>at</strong>e |ψ n 〉 is〈 〈 ∣ 〉ˆB〉∣∣ = ψ n ˆB ∣ ψ n = ∑ a ∗ i a j 〈φ i | ˆB|φ j 〉 = ∑ |a i | 2 B i (5.6.10)ijiThe expect<strong>at</strong>ion value <strong>of</strong> ˆB is due to a range <strong>of</strong> eigenvalues <strong>of</strong> ˆB, and not just due to onesingle eigenvalue like before. This is the gist <strong>of</strong> the uncertainty principle. When two oper<strong>at</strong>orsdo not commute, the precise measurement <strong>of</strong> the eigenvalue <strong>of</strong> one oper<strong>at</strong>or implies th<strong>at</strong> themeasurement <strong>of</strong> the second oper<strong>at</strong>or will involve a range <strong>of</strong> eigenvalues.For example, the eigenfunction <strong>of</strong> the momentum oper<strong>at</strong>or ˆp = −id/dx is proportionalto e ikx with eigenvalue k, since ˆpe ikx = ke ikx . The eigenfunction <strong>of</strong> the position oper<strong>at</strong>or


More M<strong>at</strong>hem<strong>at</strong>ical Preliminaries 57is δ(x − x 0 ) since xδ(x − x 0 ) = x 0 δ(x − x 0 ) where x 0 is a constant. Sincee ikx =∫ ∞−∞e ikx′ δ(x − x ′ )dx ′It is seen th<strong>at</strong> the pure eigenst<strong>at</strong>e <strong>of</strong> the momentum oper<strong>at</strong>or has to be expanded in terms <strong>of</strong>infinitely many eigenst<strong>at</strong>es <strong>of</strong> the position oper<strong>at</strong>or. Conversely,δ(x − x ′ ) = 12π∫ ∞−∞e i(x−x′ )k dkindic<strong>at</strong>es th<strong>at</strong> the eigenst<strong>at</strong>e <strong>of</strong> the position oper<strong>at</strong>or can be expanded in terms <strong>of</strong> infinitelymany eigenst<strong>at</strong>es <strong>of</strong> the momentum oper<strong>at</strong>or. 115.7 *Generalized Uncertainty PrincipleWe have seen the Heisenberg uncertainty principle expressed for uncertainty in momentum∆p and position ∆x as ∆p∆x ≥ /2. In this case, p and x are both observables in the realworld. In quantum mechanics, all observables are replaced by oper<strong>at</strong>ors. To connect theoper<strong>at</strong>ors to real world observables, we take the expect<strong>at</strong>ion values <strong>of</strong> the oper<strong>at</strong>ors. Th<strong>at</strong> isĀ = 〈A〉 = 〈f|Â|f〉 (5.7.1)where A is a scalar variable representing the eigenvalue <strong>of</strong>  and f is a st<strong>at</strong>e vector th<strong>at</strong>defines th<strong>at</strong> st<strong>at</strong>e the quantum system is in. The expect<strong>at</strong>ion value <strong>of</strong> an oper<strong>at</strong>or alsogives the st<strong>at</strong>istical mean <strong>of</strong> the observable expressed as the mean <strong>of</strong> the eigenvalue <strong>of</strong> thecorresponding oper<strong>at</strong>or. In general,〈f|(Â)n |f〉 = 〈(A) n 〉 (5.7.2)〈f|F (Â)|f〉 = 〈F (A)〉 (5.7.3)where the left-hand sides are the expect<strong>at</strong>ion values <strong>of</strong> the oper<strong>at</strong>ors, while the right-handsides are the expect<strong>at</strong>ion values <strong>of</strong> their eigenvalues.We can define the devi<strong>at</strong>ion from its mean by the oper<strong>at</strong>or∆ =  − Ā (5.7.4)On the right-hand side, the first term is an oper<strong>at</strong>or while the second term is the mean. 12The above oper<strong>at</strong>or has zero expect<strong>at</strong>ion value or zero mean.11 It is to be noted th<strong>at</strong> the eigenst<strong>at</strong>es e ikx and δ(x − x ′ ) are both not square integrable, and hence, shouldbe thought <strong>of</strong> as limiting cases <strong>of</strong> some square integrable functions.12 To be strictly correct, we should multiply the second term by the identity oper<strong>at</strong>or, but this is usuallyunderstood.


58 Quantum Mechanics Made Simple( 2We can study the variance <strong>of</strong> the oper<strong>at</strong>or ∆Â)which has nonzero expect<strong>at</strong>ion value.Namely,〈 〉 〈〉f|(∆Â)2 |f = f|( − Ā)2 |f= 〈 (A − Ā)2〉= 〈 A 2 − 2ĀA + Ā2〉= 〈 A 2〉 − 2Ā〈A〉 + Ā2= 〈 A 2〉 − Ā2 = 〈 (∆A) 2〉 = (∆A) 2 = σ 2 A (5.7.5)The above is just the definition <strong>of</strong> variance <strong>of</strong> random variable A as in st<strong>at</strong>istics. The standarddevi<strong>at</strong>ion is obtained by taking the square root <strong>of</strong> the variance to get 13σ A =√(∆A) 2 (5.7.6)We can derive a similar expression for σ B , the standard devi<strong>at</strong>ion for B.The generalized uncertainty principle is obtained by using the Schwartz inequality:∫∣ f ∗ gdx∣2(∫≤) (∫|f| 2 dxWe can rewrite the above using Dirac not<strong>at</strong>ion to get)|g| 2 dx(5.7.7)|〈f|g〉| 2 ≤ 〈f|f〉〈g|g〉 (5.7.8)The above is the generaliz<strong>at</strong>ion <strong>of</strong> the cosine inequality we have for 3-vectors 14|A · B| 2 = |A| 2 |B| 2 | cos θ| 2 ≤ |A| 2 |B| 2 (5.7.9)It can be generalized to N-vectors or vectors in N dimensional space, and then to vectorsin infinite dimensional space if the integrals converge. If we define, for a quantum system inst<strong>at</strong>e ψ,)|f〉 =( − Ā |ψ〉 = â|ψ〉 (5.7.10)( )|g〉 = ˆB − ¯B |ψ〉 = ˆb|ψ〉 (5.7.11)where â =  − Ā and ˆb = ˆB − ¯B. Then〈f|f〉 = 〈 ψ ∣ ∣â 2∣ ∣ ψ 〉 〈∣ 〉2∣∣∣ = ψ ∣∣( − Ā)ψ = (∆A) 2 (5.7.12)〈∣ 〉 〈∣ 〉∣〈g|g〉 = ψ ∣ˆb 2 ∣∣ 2∣∣∣ ψ = ψ ∣∣(ˆB − ¯B)ψ = (∆B) 2 (5.7.13)13 It is tempting to denote the standard devi<strong>at</strong>ion as ∆A but this could be confusing in view <strong>of</strong> (5.7.4).14 The integral for the inner products above can be generalized to 3D space.


More M<strong>at</strong>hem<strong>at</strong>ical Preliminaries 59Using the inequality in (5.7.8), we have((∆A) 2 ) ((∆B) 2 )≥ |〈ψ|âˆb|ψ〉| 2 (5.7.14)It can be shown easily th<strong>at</strong> if [Â, ˆB] = iĈwhere C is Hermitian. Furthermore,âˆb = âˆb + ˆbâ2[â, ˆb] = iĈ (5.7.15)+ âˆb − ˆbâ2= âˆb + ˆbâ2+ iĈ2(5.7.16)Taking the expect<strong>at</strong>ion value <strong>of</strong> the above, we have〈âˆb +〈âˆb〉 =ˆbâ〉+ i 〈Ĉ〉 (5.7.17)2 2Since â and ˆb are Hermitian, the oper<strong>at</strong>or in first term on the right-hand side is also Hermitian.Hence, its expect<strong>at</strong>ion value is purely real, while the second term is purely imaginary.Therefore, the amplitude squared <strong>of</strong> the above isUsing the above in (5.7.14), we have〈|〈âˆb〉| 2 âˆb +=ˆbâ〉∣ ∣∣∣∣2+ 1 ∣ 2 4 |〈Ĉ〉|2 ≥ 1 4 |〈Ĉ〉|2 (5.7.18)) )((∆A)((∆B) 2 2 ≥ 1 4 |〈Ĉ〉|2 (5.7.19)The above is the generalized uncertainty principle for two observables A and B. We can takethe square root <strong>of</strong> the above to getσ A σ B ≥ 1 |〈Ĉ〉| (5.7.20)25.8 *Time Evolution <strong>of</strong> the Expect<strong>at</strong>ion Value <strong>of</strong> an Oper<strong>at</strong>orThe expect<strong>at</strong>ion value <strong>of</strong> an oper<strong>at</strong>or is given byTaking the time deriv<strong>at</strong>ive <strong>of</strong> the above yields〈Â〉 = 〈ψ(t)|Â|ψ(t)〉 (5.8.1)∂ t 〈Â〉 = 〈∂ tψ(t)|Â|ψ(t)〉 + 〈ψ(t)|Â|∂ tψ(t)〉 (5.8.2)


60 Quantum Mechanics Made SimpleAssuming th<strong>at</strong>  is time independent, and using the fact th<strong>at</strong>i∂ t |ψ(t)〉 = i|∂ t ψ(t)〉 = Ĥ|ψ(t)〉 (5.8.3)we have∂ t 〈Â〉 = i 〈[Ĥψ(t)]|Â|ψ(t) 〉− i 〈ψ(t)|Â|Ĥψ(t)〉= i 〈ψ(t)|ĤÂ|ψ(t)〉 − i 〈ψ(t)|ÂĤ|ψ(t)〉= i 〈ψ(t)|Ĥ − ÂĤ|ψ(t)〉 (5.8.4)or]〉i∂ t〈[Â, 〈Â〉 = 〈ÂĤ − ĤÂ〉 = Ĥ(5.8.5)In other words, if  commutes with Ĥ, its expect<strong>at</strong>ion value is time independent, or it isa constant <strong>of</strong> motion. The oper<strong>at</strong>or  in this case represents an observable th<strong>at</strong> is conserved,such as linear momentum or angular momentum. If such an oper<strong>at</strong>or commutes withthe Hamiltonian <strong>of</strong> the quantum system, the observable th<strong>at</strong> corresponds to the oper<strong>at</strong>or isconserved.Furthermore, it can be shown th<strong>at</strong> if  is time dependent, we need only to augment theabove equ<strong>at</strong>ion as]〉i∂ t〈[Â, 〈Â〉 = Ĥ + i〈∂ t Â〉 (5.8.6)5.9 Periodic Boundary ConditionOne technique <strong>of</strong>ten used to cre<strong>at</strong>e countably infinite modes is to use periodic boundarycondition. In this method, we require th<strong>at</strong> the wavefunction s<strong>at</strong>isfies the boundary conditionth<strong>at</strong> ψ(x = 0) = ψ(x = L). This can be s<strong>at</strong>isfied easily by a traveling waveThe periodic boundary condition implies th<strong>at</strong>or th<strong>at</strong>ψ(x) = Ce ikx (5.9.1)e ikL = 1 (5.9.2)k = k n = 2nπL(5.9.3)where we have indexed k with subscript n, and n is all integer values on the real line from−∞ to ∞. Furthermore, we can pick C = 1/ √ L to normalize the above function. It is quiteeasy to show th<strong>at</strong> withψ n (x) =√1L ei2nπx/L (5.9.4)


More M<strong>at</strong>hem<strong>at</strong>ical Preliminaries 61then〈ψ n |ψ m 〉 = δ nm (5.9.5)Hence, given any function f(x) defined between 0 < x < L, we can expandf(x) =∞∑n=−∞f n ψ n (x) =∞∑n=−∞f n√1L ei2nπx/L (5.9.6)The above is just a Fourier series expansion <strong>of</strong> the function f(x). The Fourier coefficient f ncan be found by using the orthonormality <strong>of</strong> ψ n , to get√ ∫ 1 Lf n = 〈ψ n |f〉 = dxe −i2nπx/L f(x) (5.9.7)LIn the above, since ψ n (x) is a periodic function with period L, f(x) is also a periodic functionwith period L.Fourier transform can be derived from Fourier series expansion. To this end, we look<strong>at</strong> (5.9.3), and note th<strong>at</strong> the k n values are equally spaced by 2π/L on the real k line. Thespacing becomes increasingly small as L → ∞. We can define the spacing to be ∆k = 2π/L.Consequently, in view <strong>of</strong> changing a summ<strong>at</strong>ion to an integral, we can rewrite (5.9.6) moresuggestively as√ ∞∑ L 1f(x) = ∆kf n2π L ei2nπx/L = 1 ∞∑∆k2π˜f(k n )e iknx (5.9.8)n=−∞0n=−∞where we have defined ˜f(k n ) = f n√L. From (5.9.7), we have˜f(k n ) = √ Lf n =∫ L0dxe −iknx f(x) (5.9.9)In the limit when L tends to infinity, ∆k → 0, the sum in (5.9.8) becomes an integral; namelyf(x) = 12π∫ ∞−∞dk ˜f(k)e ikx (5.9.10)The above is just a Fourier inverse transform integral with ˜f(k) from (5.9.9)˜f(k) =∫ ∞0dxe −ikx f(x) (5.9.11)which is a Fourier transform. Here, f(x) is originally a periodic function, but now the periodL is infinite. Also, if f(x) → 0 when x → ∞, then by its infinite periodicity, f(x) → 0 whenx → −∞. We can replace the semi-infinite integral above with an infinite integral:˜f(k) =∫ ∞−∞The above, (5.9.10) and (5.9.12), form a Fourier transform pair.dxe −ikx f(x) (5.9.12)


62 Quantum Mechanics Made Simple


Chapter 6Approxim<strong>at</strong>e Methods inQuantum Mechanics6.1 IntroductionThere are many problems in quantum mechanics where closed form or simple solutions cannotbe found. Wh<strong>at</strong> we have shown as examples are usually problems th<strong>at</strong> have simple solutions,such a infinite potential well, harmonic oscill<strong>at</strong>or, or a finite potential well. Even the case <strong>of</strong>finite potential well requires the solution <strong>of</strong> a transcendental equ<strong>at</strong>ion. Nevertheless, theseare called textbook problems, because they are wonderful examples in textbooks to teachbeginning students on the subject m<strong>at</strong>ter. But in general, most quantum mechanics problemsdo not have closed form solution. We have to resort to approxim<strong>at</strong>e or numerical methods tosolve such problems.6.2 Use <strong>of</strong> an Approxim<strong>at</strong>e SubspaceWe have seen th<strong>at</strong> in general, a Schrödinger equ<strong>at</strong>ion problem can be cast into a m<strong>at</strong>rix equ<strong>at</strong>ionby projecting it into a space spanned by countably infinite orthonormal basis functions.With such a basis, we define an identity oper<strong>at</strong>or∞∑Î = |ψ n 〉〈ψ n | (6.2.1)n=1Given a general time independent Schrödinger equ<strong>at</strong>ion, which isĤ|ψ〉 = E|ψ〉 (6.2.2)we can first insert an identity oper<strong>at</strong>or between Ĥ and |ψ〉. Then the above becomes∞∑Ĥ|ψ n 〉〈ψ n |ψ〉 = E|ψ〉 (6.2.3)n=163


64 Quantum Mechanics Made SimpleTesting the above by multiplying it from the left with 〈ψ m |, m = 1, . . . , ∞, we have 1∞∑〈ψ m |Ĥ|ψ n〉〈ψ n |ψ〉 = E〈ψ m |ψ〉, m = 1, . . . , ∞ (6.2.4)n=1The above is a m<strong>at</strong>rix equ<strong>at</strong>ion <strong>of</strong> the form∞∑H mn f n = Ef m , m = 1, . . . , ∞ (6.2.5)n=1It corresponds to an infinite dimension m<strong>at</strong>rix, but it can be trunc<strong>at</strong>ed to form a finite m<strong>at</strong>rixsystemH · f = Ef (6.2.6)The above is a m<strong>at</strong>rix eigenvalue problem th<strong>at</strong> can be solved numerically. If H is N ×N, it willhave N eigenvalues and N eigenvectors. If we choose our basis functions appropri<strong>at</strong>ely, andare only interested in certain eigenvalues, which represent the energy levels <strong>of</strong> the st<strong>at</strong>ionaryst<strong>at</strong>es, we need only to pick a small set <strong>of</strong> basis functions th<strong>at</strong> approxim<strong>at</strong>e these st<strong>at</strong>ionaryst<strong>at</strong>es well. Therefore, N can be a small number. This is important because the number <strong>of</strong>computer oper<strong>at</strong>ions needed to find the eigenvectors and eigenvalues <strong>of</strong> the above problem isproportional to N 3 .In solving the above, we will have to form the m<strong>at</strong>rix elements H mn . For an infinitepotential well, whose well bottom has been distorted, we can use the eigenfunctions <strong>of</strong> thefl<strong>at</strong>-bottom well as our basis functions. Then, the m<strong>at</strong>rix elements H mn are explicitly givenas∫ L(H mn = dxψm(x)∗ − d 2 )2m o dx 2 + V (x) ψ n (x) (6.2.7)0We can use integr<strong>at</strong>ion by parts to convert the term th<strong>at</strong> involves second deriv<strong>at</strong>ive to a formth<strong>at</strong> involves only first deriv<strong>at</strong>ives. Th<strong>at</strong> is we letψm(x) ∗ d2dx 2 ψ n(x) = d [ψ ∗dxm(x) d ]dx ψ n(x) − ddx ψ∗ m(x) ddx ψ n(x)in the above to arrive <strong>at</strong>H mn = − ψ2mm(x) ∗ do dx ψ n(x)∣L0+∫ L0( ddx2m o dx ψ∗ m(x) d)dx ψ n(x) + ψm(x)V ∗ (x)ψ n (x)(6.2.8)The first term on the right-hand side can be made to vanish by the virtue <strong>of</strong> the boundarycondition, leaving the second term, which is a more symmetric form.1 This procedure is called testing or weighting in the m<strong>at</strong>hem<strong>at</strong>ics liter<strong>at</strong>ure, 〈ψ m| is called the testing orweighting function.


Approxim<strong>at</strong>e Methods in Quantum Mechanics 65Modern numerical methods have allowed us to use flexible basis set to expand our unknownfunctions. For example, we can let|ψ〉 =N∑a n |f n 〉 (6.2.9)n=1where f n (x) need not be orthogonal or complete in Hilbert space, but they can approxim<strong>at</strong>ethe unknown function ψ(x) well. Also, f n (x), n = 1, . . . , N spans an N dimensional vectorspace th<strong>at</strong> is the subspace <strong>of</strong> the Hilbert space. We assume th<strong>at</strong> the solution in the subspaceis a good approxim<strong>at</strong>ion <strong>of</strong> the solution in the original infinite dimensional Hilbert space.Examples <strong>of</strong> such finite basis set are triangle functions shown in Figure 6.1.Using (6.2.9) in (6.2.2), we haveN∑N∑Ĥ|f n 〉a n = E a n |f n 〉 (6.2.10)n=1n=1Multiplying the above from 〈f m | from the left, we haveN∑N∑〈f m |Ĥ|f n〉a n = E 〈f m |f n 〉a n (6.2.11)n=1The above is a m<strong>at</strong>rix system <strong>of</strong> the formn=1˜H · a = EB · a (6.2.12)In the above,[ ˜H][ ]Bmnmn= 〈f m |Ĥ|f n〉= 〈f m |f n 〉[a] n= a n (6.2.13)The equ<strong>at</strong>ion (6.2.12) above is a generalized eigenvalue problem with eigenvalue E and eigenvectora. The difference <strong>of</strong> it from (6.2.6) is th<strong>at</strong> a m<strong>at</strong>rix B is on the right-hand side. 2 Butnumerical s<strong>of</strong>tware to seek the eigenvalues and eigenvectors for such generalized eigenvalueproblems are widely available, e.g., in MATLAB.The approxim<strong>at</strong>e subspace method is varyingly known as the subspace projection method,Galerkin’s method, Petrov-Galerkin method, weighted-residual method, and the method <strong>of</strong>moments. The finite element method falls under the same c<strong>at</strong>egory.6.3 *Time Independent Perturb<strong>at</strong>ion TheoryLet us assume th<strong>at</strong> we can solve a problem when it is unperturbed, such as the infinitepotential well problem. We like to solve the problem when, say there is a small electric field2 The m<strong>at</strong>rix B is also known as the Gram m<strong>at</strong>rix.


66 Quantum Mechanics Made SimpleFigure 6.1: The triangle functions for a piecewise linear approxim<strong>at</strong>ion <strong>of</strong> a function. This isa basis th<strong>at</strong> is not orthogonal but yet can be used to seek approxim<strong>at</strong>e solutions to (6.2.12).Figure 6.2: The infinite potential well on the left represents the unperturbed problem. Themiddle figure represents a perturb<strong>at</strong>ion due to a tiny electric field. The right figure representsa perturb<strong>at</strong>ion due to imperfection in fabric<strong>at</strong>ion or impurities.in the well indic<strong>at</strong>ed by a potential with a gradient, or th<strong>at</strong> the well is imperfect due to someimpurities indic<strong>at</strong>ed by a small bump (see Figure 6.2). The Schrödinger equ<strong>at</strong>ion for theperturbed system can be written as(Ĥ0 + γĤp)|φ〉 = E|φ〉 (6.3.1)where Ĥ0 is the Hamiltonian <strong>of</strong> the unperturbed system whose solution is known, and γĤpis due to the small perturb<strong>at</strong>ion where γ is a small parameter. Here, Ĥ 0 can be the Hamiltonian<strong>of</strong> the infinite potential well, for instance. In the above equ<strong>at</strong>ion, |φ〉 and E are bothunknowns, but we can write them in a perturb<strong>at</strong>ion series or expansion, namely|φ〉 = |φ (0) 〉 + γ|φ (1) 〉 + γ 2 |φ (2) 〉 + . . . (6.3.2)E = E (0) + γE (1) + γ 2 E (2) + . . . (6.3.3)


Approxim<strong>at</strong>e Methods in Quantum Mechanics 67Upon substituting the above series into (6.3.1), we obtain) (Ĥ0 ( )+ γĤp |φ (0) 〉 + γ|φ (1) 〉 + γ 2 |φ (2) 〉 + . . .() ()= E (0) + γE (1) + γ 2 E (2) + . . . |φ (0) 〉 + γ|φ (1) 〉 + γ 2 |φ (2) 〉 + . . .(6.3.4)The left-hand side <strong>of</strong> (6.3.4) can be expanded and rewritten on a power series in γwhile the right-hand side is similarly written asa 0 + a 1 γ + a 2 γ 2 + ... (6.3.5)b 0 + b 1 γ + b 2 γ 2 + ... (6.3.6)These two power series in γ are equal only if a i = b i , i = 0, 1, ..., ∞. 3Equ<strong>at</strong>ing the coefficients <strong>of</strong> the power series on both sides <strong>of</strong> (6.3.4) we have the followingequ<strong>at</strong>ions:Zeroth Order:Ĥ 0 |φ (0) 〉 = E (0) |φ (0) 〉 (6.3.7)First Order:Ĥ 0 |φ (1) 〉 + Ĥp|φ (0) 〉 = E (0) |φ (1) 〉 + E (1) |φ (0) 〉 (6.3.8)Second Order:Ĥ 0 |φ (2) 〉 + Ĥp|φ (1) 〉 = E (0) |φ (2) 〉 + E (1) |φ (1) 〉 + E (2) |φ (0) 〉 (6.3.9)We assume th<strong>at</strong> the zeroth order equ<strong>at</strong>ion is known in terms <strong>of</strong> an eigenst<strong>at</strong>e |ψ m 〉 with energyE m . In other words|φ (0) 〉 = |ψ m 〉, E (0) = E m (6.3.10)We will use this knowledge to solve the first order equ<strong>at</strong>ion (6.3.8) above.Before we proceed further, a note is in order regarding the uniqueness <strong>of</strong> the eigenvalueproblem (6.3.1). An eigenvector is known only within a multiplic<strong>at</strong>ive factor. Hence, itslength is indetermin<strong>at</strong>e. This non-uniqueness in its length manifests in the non-uniqueness<strong>of</strong> the value <strong>of</strong> the perturb<strong>at</strong>ion series (6.3.2) as we shall see l<strong>at</strong>er. To achieve uniqueness, itis best to pin down the length <strong>of</strong> the total eigenvector given by (6.3.2). We fix the length <strong>of</strong>the eigenvector |φ〉 by requiring th<strong>at</strong>〈ψ m |φ〉 = 1 (6.3.11)With this requirement, we substitute (6.3.2) into the above. Since 〈ψ m |φ (0) 〉 = 1, because|φ (0) 〉 = |ψ m 〉, it is easy to show th<strong>at</strong> 〈ψ m |φ (i) 〉 = 0, i > 0. As a consequence, |φ (i) 〉 isorthogonal to |ψ m 〉. The perturb<strong>at</strong>ion series is not necessarily normalized, but it can benormalized l<strong>at</strong>er after the series has been calcul<strong>at</strong>ed.3 This can be easily proved by setting γ = 0, which immedi<strong>at</strong>ely implies th<strong>at</strong> a 0 = b 0 . By differenti<strong>at</strong>ingthe power series again and let γ = 0, we prove th<strong>at</strong> a i = b i etc.


68 Quantum Mechanics Made Simple6.3.1 First Order Perturb<strong>at</strong>ionNext, to find the first order corrections to the eigenvalue and the eigenvector, we move theunknowns |φ (1) 〉 to the left <strong>of</strong> (6.3.8). We then have(Ĥ0 − E m)|φ (1) 〉 = E (1) |ψ m 〉 − Ĥp|ψ m 〉 (6.3.12)where we have made)use <strong>of</strong> (6.3.10). Notice th<strong>at</strong> the above equ<strong>at</strong>ion is non-unique since theoper<strong>at</strong>or(Ĥ0 − E m has a null space with a null space vector |ψ m 〉.Testing the above equ<strong>at</strong>ion with 〈ψ m |, we have〈ψ m |Ĥ0 − E m |φ (1) 〉 = E (1) − 〈ψ m |Ĥp|ψ m 〉 (6.3.13))But 〈ψ m |(Ĥ0 − E m = 0 because Ĥ0|ψ m 〉 = E m |ψ m 〉. Hence, the above givesE (1) = 〈ψ m |Ĥp|ψ m 〉 (6.3.14)the first order correction to the energy <strong>of</strong> the perturbed system.Therefore, one <strong>of</strong> the two unknowns in (6.3.12) is found. The remaining unknown |φ (1) 〉can be expanded in terms <strong>of</strong> the eigenfunctions <strong>of</strong> the unperturbed system, th<strong>at</strong> is|φ (1) 〉 = ∑ na (1)n |ψ n 〉 (6.3.15)First, testing the equ<strong>at</strong>ion (6.3.12) with 〈ψ i |, we have〈ψ i |Ĥ0 − E m |φ (1) 〉 = E (1) 〈ψ i |ψ m 〉 − 〈ψ i |Ĥp|ψ m 〉 (6.3.16)Upon substituting (6.3.15) into the above, the left-hand side evalu<strong>at</strong>es to〈ψ i |E i − E m |φ (1) 〉 = (E i − E m ) ∑ n〈ψ i |a (1)n |ψ n 〉 = (E i − E m )a (1)i (6.3.17)The right-hand side <strong>of</strong> (6.3.16), for i ≠ m, is just −〈ψ i |Ĥp|ψ m 〉. Hencea (1)i = 〈ψ i|Ĥp|ψ m 〉E m − E i, i ≠ m (6.3.18)When i = m, (6.3.17) evalu<strong>at</strong>es to zero, while the right hand side <strong>of</strong> (6.3.16) also evalu<strong>at</strong>esto zero because <strong>of</strong> (6.3.14). Hence a (1)m is undefined. We choose a (1)m = 0 for a number <strong>of</strong>reasons: It makes the correction term unique since |ψ (1) 〉 is orthogonal to |ψ (0) 〉. It makesthe normaliz<strong>at</strong>ion <strong>of</strong> the eigenvector |φ〉 accur<strong>at</strong>e to second order even though the correctionis first order. It will also make the second order corrections much simpler to find.


Approxim<strong>at</strong>e Methods in Quantum Mechanics 696.3.2 Second Order Perturb<strong>at</strong>ionTo find the second order corrections, we rewrite (6.3.9) with the unknown |φ (2) 〉 on the lefthand side. Then (6.3.9) becomes(Ĥ0 − E m)|φ (2) 〉 = E (1) |φ (1) 〉 + E (2) |ψ m 〉 − Ĥp|φ (1) 〉 (6.3.19)Testing the above with 〈ψ m |, the left hand side becomes zero as before. 4 Since we havemade |φ (1) 〉 orthogonal to |ψ m 〉, on the right-hand side, only the last two terms remain.Consequently,orLetting0 = E (2) − 〈ψ m |Ĥp|φ (1) 〉 (6.3.20)E (2) = 〈ψ m |Ĥp|φ (1) 〉 (6.3.21)|φ (2) 〉 = ∑ na (2)n |ψ n 〉 (6.3.22)and substituting it into the left side <strong>of</strong> (6.3.19), testing with 〈ψ i |, we haveTherefore, for i ≠ m,(E i − E m )a (2)i = E (1) 〈ψ i |φ (1) 〉 + E (2) δ im − 〈ψ i |Ĥp|φ (1) 〉 (6.3.<strong>23</strong>)a (2)i = 〈ψ i|Ĥp|φ (1) 〉− E(1) a (1)i(6.3.24)E m − E i E m − E iWhen i = m, both sides <strong>of</strong> (6.3.<strong>23</strong>) vanish, and a (2)m is undefined. Again, we pick a (2)m = 0 toobtain a unique solution.6.3.3 Higher Order CorrectionsThe above procedure can be generalized to arbitrary order. By induction, we notice th<strong>at</strong> theequivalence <strong>of</strong> (6.3.9) to p-th order isĤ 0 |φ (p) 〉 + Ĥp|φ (p−1) 〉 = E (0) |φ (p) 〉 + E (1) |φ (p−1) 〉 + · · · + E (p) |φ (0) 〉 (6.3.25)The above can be rewritten as(Ĥ0 − E (0)) |φ (p) 〉 = E (1) |φ (p−1) 〉 + · · · + E (p) |φ (0) 〉 − Ĥp|φ (p−1) 〉 (6.3.26)Testing the above with 〈ψ m | gives4 Again, the above is non-unique for the same reason cited for (6.3.12).E (p) = 〈ψ m |Ĥp|φ (p−1) 〉 (6.3.27)


70 Quantum Mechanics Made SimpleLetting|φ (p) 〉 = ∑ na (p)n |ψ n 〉 (6.3.28)in (6.3.25), and testing with 〈ψ i |, one can easily show, for i ≠ m, and after noting th<strong>at</strong>E (0) = E m , |φ (0) 〉 = |ψ m 〉, th<strong>at</strong>a (p) 1()i =〈ψ i |Ĥp|φ (p−1) 〉 − E (1) a (p−1)i − E (2) a (p−2)i + · · · − E (p−1) a (1)iE m − E i(6.3.29)It is to be noted th<strong>at</strong> with modern advent <strong>of</strong> computer technology, and given the availability<strong>of</strong> numerical methods, the calcul<strong>at</strong>ion <strong>of</strong> perturb<strong>at</strong>ion theory to very high order islaborious and not necessary. However, a perturb<strong>at</strong>ion correction can give us insight on howa small change in the Hamiltonian can change the solution.6.4 Tight Binding ModelThe tight binding model is a way to find the solution to two weakly coupled quantum systemswhen the solution to each isol<strong>at</strong>ed quantum system is known (see Figure 6.3).Figure 6.3: The tight binding model can be used to find the approxim<strong>at</strong>e eigenst<strong>at</strong>es <strong>of</strong> twoquantum wells th<strong>at</strong> are weakly coupled to each other. The eigenst<strong>at</strong>es <strong>of</strong> the isol<strong>at</strong>ed quantumwells are known (Figure is from DAB Miller).


Approxim<strong>at</strong>e Methods in Quantum Mechanics 71To solve this system, we can use the subspace projection method <strong>of</strong> nonorthogonal basisto derive the relevant m<strong>at</strong>rix equ<strong>at</strong>ion. Given the Schrödinger equ<strong>at</strong>ionwe letĤ |ψ〉 = E |ψ〉 (6.4.1)|ψ〉 =2∑a i |ψ i 〉 (6.4.2)i=1where |ψ 1 〉 is the eigenfunction <strong>of</strong> the ground st<strong>at</strong>e <strong>of</strong> well 1 in isol<strong>at</strong>ion and |ψ 2 〉 is theeigenfunction <strong>of</strong> the ground st<strong>at</strong>e <strong>of</strong> well 2. Going through the same procedure we have hadbefore, we arrive <strong>at</strong>2∑a i 〈ψ j | Ĥ |ψ i〉 = Ei=1The above is the same as the m<strong>at</strong>rix equ<strong>at</strong>ion2∑a i 〈ψ j | ψ i 〉 (6.4.3)i=1H · a = EB · a (6.4.4)where[ ]H[ ]Bijij= 〈ψ i | Ĥ |ψ j〉 (6.4.5)= 〈ψ i | ψ j 〉 (6.4.6)It is easy to show th<strong>at</strong> H and B are Hermitian m<strong>at</strong>rices. 5 In the above, we can assume th<strong>at</strong>|ψ i 〉 is quasi-orthonormal, so th<strong>at</strong>〈ψ i | ψ j 〉 ≃ δ ij (6.4.7)In other words, the B m<strong>at</strong>rix is unity on the diagonal, and small on the <strong>of</strong>f diagonal terms.Then we can write the system (6.4.4) as[E1 ∆E∆E ∗E 1] [a1a 2]= E[ ] [ ]1 ∆B a1∆B ∗ 1 a 2(6.4.8)The <strong>of</strong>f diagonal elements are conjug<strong>at</strong>e <strong>of</strong> each other because the m<strong>at</strong>rix is Hermitian. Moreover,∆E ≪ E 1 and ∆B ≪ 1. These terms are small because the overlap between the eigenfunctionsψ 1 and ψ 2 is small. The further apart the wells are, the smaller these terms wouldbe, as the overlap is smaller. When the wells are infinite far apart, the <strong>of</strong>f diagonal termsvanish and the two quantum systems are uncoupled from each other. The two eigenvaluesare then degener<strong>at</strong>e. To begin, we can assume weak coupling between the modes in the wells.5 It can be easily shown th<strong>at</strong> the m<strong>at</strong>rix represent<strong>at</strong>ion <strong>of</strong> a Hermitian oper<strong>at</strong>or remains Hermitian.


72 Quantum Mechanics Made SimpleorThe eigenvalue <strong>of</strong> (6.4.8) above is obtained by solvingUpon solving the above, we have[Edet 1 − E∆E ∗ − E∆B ∗]∆E − E∆B= 0 (6.4.9)E 1 − E(E 1 − E) 2 − |∆E − E∆B| 2 = 0 (6.4.10)E = E 1 ± |∆E 1 − E∆B| ∼ = E 1 ± |∆E 1 − E 1 ∆B| (6.4.11)Since the E∆B on the right-hand side <strong>of</strong> the first equality above is small, we do not incurmuch error by replacing it with E 1 ∆B. Also, it can be shown th<strong>at</strong> st<strong>at</strong>ionary st<strong>at</strong>es can bemade completely real valued. Hence, ∆E and ∆B can be made to be real valued. Therefore,without the lost <strong>of</strong> generality, the above can be replaced withE ∼ = E 1 ± (∆E 1 − E 1 ∆B) (6.4.12)Note th<strong>at</strong> the degener<strong>at</strong>e eigenvalues are now split into two non-degener<strong>at</strong>e eigenvalues.The equ<strong>at</strong>ion to be s<strong>at</strong>isfied by the eigenvectors is[ ]±(∆E − E1 ∆B) ∆E − E 1 ∆B∆E − E 1 ∆B ±(∆E − E 1 ∆B)[ ]a1· = 0 (6.4.13)a 2On solving, and normalizing, 6 we obtain th<strong>at</strong>[a1a 2]= √ 1 [ ]12 1(6.4.14)or[a1a 2]= √ 1 [ ] 12 −1(6.4.15)The corresponding eigenvectors can be found, and the eigenfunctions are <strong>of</strong> the formψ + = 1 √2(ψ 1 + ψ 2 ) , ψ − = 1 √2(ψ 1 − ψ 2 ) (6.4.16)The eigenfunctions are due to even and odd coupling between the isol<strong>at</strong>ed quantum st<strong>at</strong>es <strong>of</strong>the two wells as they are brought closer together.6 The normaliz<strong>at</strong>ion <strong>of</strong> the eigenfunction in (6.4.2) implies th<strong>at</strong> |a 1 | 2 + |a 2 | 2 = 1.


Approxim<strong>at</strong>e Methods in Quantum Mechanics 736.4.1 Vari<strong>at</strong>ional MethodGiven an eigenvalue problem denoted byĤ |φ〉 = E |φ〉 (6.4.17)we can convert the above into a m<strong>at</strong>rix generalized eigenvalue problemH · a = E B · a (6.4.18)We next convert the above into a functional by multiplying it with ā † to obtaina † · H · a = E a † · B · a (6.4.19)By setting a = a 0 + δa, E = E 0 + δE, where a 0 and E 0 are the exact solution to (6.4.18),and then taking the first vari<strong>at</strong>ion <strong>of</strong> (6.4.19) we haveδa † · H · a 0 + a † 0 · H · δa = δE a† 0 · B · a 0 + E 0 · δa † · B · a 0+E 0 a † 0 · B · δa (6.4.20)If a 0 and E 0 are exact solution to (6.4.20), they s<strong>at</strong>isfyH · a 0 = E 0 B · a 0 (6.4.21)a † 0 · H = E 0a † 0 · B 0 (6.4.22)Then the term multiplied by δa † and δa cancel each other on both sides and we have, to firstorder,0 = δE a † 0 · B · a 0 (6.4.<strong>23</strong>)The above implies th<strong>at</strong> δE is zero. It means th<strong>at</strong> if we were to use (6.4.19) to find E bywriting it asE = a† · H · a(6.4.24)a † · B · aand if we substitute inexact or approxim<strong>at</strong>e value for a = a 0 + δa, the value we obtain forE is second order accur<strong>at</strong>e since δE = 0. The above is known as the Rayleigh quotient. Tobegin with, we need an estim<strong>at</strong>e <strong>of</strong> the solution th<strong>at</strong> is quite close to the exact solution soth<strong>at</strong> δa is small. In other words, it allows one to substitute in approxim<strong>at</strong>e value <strong>of</strong> a th<strong>at</strong>is just first-order accur<strong>at</strong>e, and yet get an estim<strong>at</strong>e <strong>of</strong> the eigenvalue or energy E th<strong>at</strong> issecond-order accur<strong>at</strong>e.A more general form <strong>of</strong> the Rayleigh quotient is to test (6.4.17) with 〈φ| and convert itinto a functional. The corresponding Rayleigh quotient isE = 〈φ|Ĥ|φ〉〈φ|φ〉(6.4.25)It assumes th<strong>at</strong> |φ〉 is not normalized. In the above, it can be proved by the same token th<strong>at</strong>an approxim<strong>at</strong>e, first-order accur<strong>at</strong>e value <strong>of</strong> |φ〉 can yield a second-order accur<strong>at</strong>e estim<strong>at</strong>e<strong>of</strong> E. Rayleigh quotient is gre<strong>at</strong> for estim<strong>at</strong>ing the eigenvalues when only approxim<strong>at</strong>e eigenfunctionsare known. These eigenfunctions are also called trial functions. A system<strong>at</strong>ic wayto estim<strong>at</strong>e these eigenfunctions is known as the Rayleigh-Ritz method.


74 Quantum Mechanics Made Simple6.5 Time Dependent Perturb<strong>at</strong>ion TheoryNow, we assume th<strong>at</strong> the perturbing Hamiltonian is a function <strong>of</strong> time. In other words, wehave a Schrödinger equ<strong>at</strong>ion whereandi ∂ |Ψ〉 = Ĥ |Ψ〉 (6.5.1)∂tĤ = Ĥ0 + γĤp (t) (6.5.2)where Ĥp (t) is time varying. 7 It can come from an external electric field th<strong>at</strong> is time-harmonic,for instance. We also assume th<strong>at</strong> we can solve the unperturbed system in terms <strong>of</strong> st<strong>at</strong>ionaryst<strong>at</strong>es, namelyĤ 0 |Ψ n 〉 = E n |Ψ n 〉 (6.5.3)The total time dependent wavefunction can be expanded in terms <strong>of</strong> the st<strong>at</strong>ionary st<strong>at</strong>es,namely|Ψ〉 = ∑ a n (t) e −iωnt |Ψ n 〉 (6.5.4)nwhere ω n = E n , and a n (t) is a function <strong>of</strong> time.Substituting the above into (6.5.1), we have∂∂t |Ψ〉 = ∑ n(ȧ n (t) − iω n ) e −iωnt |Ψ n 〉 = 1 ∑ a n (t) e −iωnt |Ψ n 〉 (6.5.5)iĤ n= 1 ∑( )a n (t) e −iωnt E n |Ψ n 〉 + γĤp(t) |Ψ n 〉in(6.5.6)where ȧ n (t) = da n (t) /dt.Since E n = ω n , the E n term on the right-hand side cancels the ω n term on the left-handside. Finally we have∑ȧ n (t) e −iωnt |Ψ n 〉 = γ ∑a n (t) e −iωnt Ĥ p (t) |Ψ n 〉 (6.5.7)inTo simplify the equ<strong>at</strong>ion further, we test it with 〈Ψ q | to getȧ q (t) e −iωqt = γ ∑a n (t) e −iωnt 〈Ψ q |iĤp(t) |Ψ n 〉 (6.5.8)nThe above equ<strong>at</strong>ion is exact as <strong>of</strong> this point. To solve it further, we expandna n = a (0)n + γa (1)n + γ 2 a (2)n · · · (6.5.9)7 To be precise, the Hamiltonian <strong>of</strong> a system has to be a constant <strong>of</strong> motion, since it represents theconserv<strong>at</strong>ion <strong>of</strong> energy. But when a quantum system is coupled to another quantum system to which energyis transferred, the first Hamiltonian may not be a constant anymore. But we can assume th<strong>at</strong> the perturbingHamiltonian is small.


Approxim<strong>at</strong>e Methods in Quantum Mechanics 75Using the above in (6.5.8), and m<strong>at</strong>ching terms <strong>of</strong> like orders, we haveor th<strong>at</strong> a (0)qȧ (0)q (t) = 0 (6.5.10)is a constant. By m<strong>at</strong>ching terms <strong>of</strong> the first order, we haveȧ (1)q (t) = 1i∑na (0)n e iωqnt 〈ψ q |Ĥp(t)|ψ n 〉 (6.5.11)where ω qn = ω q −ω n . To make things simpler, we can assume th<strong>at</strong> the electron is in the m-thst<strong>at</strong>ionary st<strong>at</strong>e before we turn on the perturb<strong>at</strong>ion Ĥp(t). Namely,In the above, since if only one st<strong>at</strong>e exists, a (0)m|ψ〉 = a (0)m e −iωmt |ψ m 〉 = e −iωmt |ψ m 〉 (6.5.12)= 1 by normaliz<strong>at</strong>ion. Then, (6.5.11) becomesȧ (1)q (t) = 1i a(0) m e iωqmt 〈ψ q |Ĥp(t)|ψ m 〉 (6.5.13)We assume the perturbing Hamiltonian to be <strong>of</strong> the form:⎧⎪⎨ 0, t < 0Ĥ p (t) = Ĥ p0 (e⎪⎩−iωt + e iωt ), 0 < t < t 0(6.5.14)0, t > t 0so th<strong>at</strong> it is turned on only for a time window 0 < t < t 0 . Then for t > t 0 , by integr<strong>at</strong>ing(6.5.13), we havea (1)q = 1i∫ t00〈ψ q |Ĥp(t ′ )|ψ m 〉e iωqmt′ dt ′ (6.5.15)∫ t0(e i(ωqm−ω)t′ + e i(ωqm+ω)t′) dt ′ (6.5.16)= 1i 〈ψ q|Ĥp0)|ψ m 〉0= − 1( ei(ω qm−ω)ti 〈ψ 0)− 1− 1q|Ĥp0)|ψ m 〉+ ei(ωqm+ω)t0ω qm − ω ω qm + ω= − t [(0i 〈ψ q|Ĥp0)|ψ m 〉 e i(ωqm−ω)t 0 (ωqm − ω)t 02 sinc2The probability <strong>of</strong> finding the st<strong>at</strong>e in q-st<strong>at</strong>e is(6.5.17))])(+ e i(ωqm+ω)t 0 (ωqm + ω)t 02 sinc2(6.5.18)P (q) = |a (1)q | 2 (6.5.19)or[ ( ) ( )]P (q) ≈ t2 0 2 |〈ψ q|Ĥp0)|ψ m 〉| 2 sinc 2 (ωqm − ω)t 0+ sinc 2 (ωqm + ω)t 022(6.5.20)


76 Quantum Mechanics Made Simplewhere we have ignored the cross terms since the sinc functions in (6.5.18) are highly peaked<strong>at</strong> ω = ±ω qm when t 0 → ∞; and hence their cross terms are small.The above equ<strong>at</strong>ion means th<strong>at</strong> the probability <strong>of</strong> transition from the m to the q eigenst<strong>at</strong>eis high only ifor th<strong>at</strong>ω = ±ω qm = ±(ω q − ω m ) (6.5.21)ω = ±(ω q − ω m ) = ±(E q − E m ) (6.5.22)The “+” sign corresponds to when the electron jumps from a low energy E m to high energyE q requiring the absorption <strong>of</strong> a photon with energy ω = (E q − E m ). The “−” sign in(6.5.22) corresponds to the electron dropping from a high energy st<strong>at</strong>e E m to a low energyst<strong>at</strong>e E q emitting a photon with energy ω = (E m −E q ). This is known as the Fermi’s goldenrule.


Chapter 7Quantum Mechanics in Crystals7.1 IntroductionIn a crystal, the <strong>at</strong>oms are loc<strong>at</strong>ed in a periodic l<strong>at</strong>tice. Hence, when an electron wavepropag<strong>at</strong>es on a l<strong>at</strong>tice, it is propag<strong>at</strong>ing in a periodic structure. However, the real world isnot th<strong>at</strong> simple. Usually, there are more than one electron traveling in a l<strong>at</strong>tice. The electronssee each other’s electric field or potentials. There will be electron-electron interaction th<strong>at</strong> hasto be accounted for. Moreover, electrons are fermions meaning th<strong>at</strong> they obey Pauli exclusionprinciple. Two electrons cannot be in the same st<strong>at</strong>e simultaneously. There are many-bodyeffects, but we will ignore them here. The many-body effect can be lumped approxim<strong>at</strong>elyinto an effective potential. We will assume th<strong>at</strong> the effective potential is still periodic, eventhough this may not be true.Another effect is th<strong>at</strong> as an electron moves through a l<strong>at</strong>tice, the <strong>at</strong>tractive force betweenthe electron and a nucleus <strong>of</strong> the <strong>at</strong>oms distorts the l<strong>at</strong>tice loc<strong>at</strong>ions. This gives rise to l<strong>at</strong>ticevibr<strong>at</strong>ions called phonons. We will ignore electron-phonon coupling here.In a periodic l<strong>at</strong>tice, the <strong>at</strong>oms are loc<strong>at</strong>ed <strong>at</strong> the l<strong>at</strong>tice points given byR L = n 1 a 1 + n 2 a 2 + n 3 a 3 (7.1.1)where a 1 , a 2 , a 3 are l<strong>at</strong>tice vectors, and n 1 , n 2 , n 3 are integers. Every l<strong>at</strong>tice point can berel<strong>at</strong>ed to every other l<strong>at</strong>tice point by a proper choice <strong>of</strong> n 1 , n 2 , and n 3 .Assuming th<strong>at</strong> an electron sees a periodic potential in a l<strong>at</strong>tice, thenV P (r + R L ) = V P (r) (7.1.2)The single electron sees this periodic potential and its wavefunction s<strong>at</strong>isfies][− 2∇ 2 + V P (r) ψ(r) = Eψ(r) (7.1.3)2m eWe will study next the kind <strong>of</strong> wave th<strong>at</strong> can propag<strong>at</strong>e in this periodic potential, known asBloch-Floquet waves. 11 In the subsequent discussions, we will replace m e with m, with implicit understanding th<strong>at</strong> this is themass <strong>of</strong> the electron.77


78 Quantum Mechanics Made Simple7.2 Bloch-Floquet WavesLet us study the case <strong>of</strong> a wave propag<strong>at</strong>ing in a 1-D periodic structure, where V (x) maylook like th<strong>at</strong> <strong>of</strong> Figure 7.1.Figure 7.1: A 1D periodic structure <strong>of</strong> a potential pr<strong>of</strong>ile for V (x) where a Bloch-Floquetwave can travel on it.The periodicity <strong>of</strong> the potential is a. Let us assume th<strong>at</strong> the wave th<strong>at</strong> can exist on sucha structure is <strong>of</strong> the formψ(k, x) = e ikx u p (x) (7.2.1)This wavefunction must s<strong>at</strong>isfy the 1-D Schrödinger equ<strong>at</strong>ion th<strong>at</strong>[− 2d 2 ]2m dx 2 + V P (x) ψ(k, x) = Eψ(k, x) (7.2.2)In order for ψ(k, x) to s<strong>at</strong>isfy the above, u p (x) has to s<strong>at</strong>isfy a certain condition. Since theabove is a periodic structure, if one transl<strong>at</strong>e the coordin<strong>at</strong>e such th<strong>at</strong> x ⇒ x + na, thenV P (x) ⇒ V P (x + na) = V P (x) which remains unchanged. Letting x ⇒ x + na in (7.2.1), wehaveψ(k, x) ⇒ ψ(k, x + na) = e ikx e ikna u p (x + na) (7.2.3)This new wavefunction must be a solution <strong>of</strong> Schrödinger equ<strong>at</strong>ion as well. In order for it tos<strong>at</strong>isfy (7.2.2), we require th<strong>at</strong>u p (x + na) = u p (x), for all n. (7.2.4)In other words, u p (x) is a periodic function <strong>of</strong> x. Since e ikna is just a constant, it is clearth<strong>at</strong> ψ(x + na) s<strong>at</strong>isfies the same equ<strong>at</strong>ion as (7.2.2). Hence, the form <strong>of</strong> (7.2.1) th<strong>at</strong> s<strong>at</strong>isfies(7.2.2) has to be such th<strong>at</strong> u p (x) is a periodic function. Equ<strong>at</strong>ion (7.2.1) represents a travelingwave modul<strong>at</strong>ed by a periodic function u p (x). However, u p (x) is more rapidly varying thane ikx . Such a wave is a Bloch wave or a Bloch-Floquet wave. It is to be noted th<strong>at</strong> u p (x)is a function <strong>of</strong> k as well, and we should have written u p (k, x), but we assume th<strong>at</strong> the kdependence is understood.


Quantum Mechanics in Crystals 79Since u p (x) is periodic function with period a, it can be expanded as a Fourier series,namely∞∑u p (x) = a n e inπx/(2a) (7.2.5)n=−∞Noting th<strong>at</strong> the second deriv<strong>at</strong>ive <strong>of</strong> (7.2.1) is given byd 2[e ikxdx 2 u p (x) ] = −k 2 e ikx u p (x) + 2ike ikx u ′ p(x) + e ikx u ′′p(x), (7.2.6)substituting (7.2.1) into (7.2.2), and using the above, we have− 2[−k 2 u p (x) + 2iku ′ p(x) + u ′′p(x) ] + V p (x)u p (x) = Eu p (x) (7.2.7)2mIn the above, if V p (x) = 0, a simple solution to the above is u p (x) a constant with E =(k) 2 /(2m). Then the Bloch-Floquet wave becomes a simple traveling wave. Otherwise, wecan rewrite (7.2.7) as[ {− 2 d22m dx 2 + 2ik d ] }dx − k2 + V p (x) u p (x) = Eu p (x) (7.2.8)with u p (x) s<strong>at</strong>isfying the periodic boundary condition within period a. The above is a simpleeigenvalue problem with eigenvector u p (x) and eigenvalue E. We can use m<strong>at</strong>rix method toconvert the above into a m<strong>at</strong>rix equ<strong>at</strong>ion∑H mn a n = E ∑ B mn a n (7.2.9)nn∑by letting u p (x) = N a n φ n (x), and testing with φ m (x). One also ensures th<strong>at</strong> the choice <strong>of</strong>n=1basis is such th<strong>at</strong> the periodic boundary condition is s<strong>at</strong>isfied. Here, φ n (x) are not necessarilyorthogonal. In the above,whereH mn = 〈φ m |Ĥp|φ n 〉, B mn = 〈φ m |φ n 〉 (7.2.10)[Ĥ p = − 2 d22m dx 2 + 2ik d ]dx − k2 + V p (x) (7.2.11)Notice th<strong>at</strong> Ĥp is a function <strong>of</strong> k, and so H mn is a function <strong>of</strong> k. It is to be noted th<strong>at</strong> Ĥ isa Hermitian oper<strong>at</strong>or since we have shown th<strong>at</strong> −id/dx is a Hermitian oper<strong>at</strong>or in (5.2.51)in Chapter 2.Therefore, (7.2.9) can be written as a generalized eigenvalue problemH(k) · a = E(k)B · a (7.2.12)with real eigenvalues. In the above, we have explicitly denoted th<strong>at</strong> H(k) is a function <strong>of</strong> k.Hence, we have to fix k in order to solve the eigenvalue problem (7.2.8), or the generalizedm<strong>at</strong>rix eigenvalue problem (7.2.12). Therefore, the eigenvalue E obtained is also a function<strong>of</strong> k, the wavenumber in the Bloch-Floquet wave (7.2.1). In general, a E(k) plot will look liketh<strong>at</strong> in Figure 7.2.


80 Quantum Mechanics Made SimpleFigure 7.2: The E-k diagram or band structure diagram <strong>of</strong> a Bloch-Floquet wave propag<strong>at</strong>ingin a 1D periodic structure.7.2.1 Periodicity <strong>of</strong> E(k)For each k value, there are many possible values <strong>of</strong> E. Also, in (7.2.1), if we let k ⇒ k + 2pπa ,where p is an integer value, then, (7.2.1) becomesψ (k, x) ⇒ ψ (k + 2pπ/a, x) = e ikx u p (x) e 2ipπa x = e ikx ũ p (x) (7.2.13)Notice th<strong>at</strong> e 2ipπa x is a periodic function; hence, it can be lumped with u p (x) to form anew ũ p (x) periodic function. They s<strong>at</strong>isfy the same equ<strong>at</strong>ion (7.2.8) with the same k value.Therefore, u p (x) and ũ p (x) share the same set <strong>of</strong> eigenvalues. Consequently, E (k) is aperiodic function in k with period 2π a. Each <strong>of</strong> this period is called the Brillouin zone. Thezone th<strong>at</strong> is centered about the origin is called the first Brillouin zone.7.2.2 Symmetry <strong>of</strong> E(k) with respect to kFurthermore, since every term in the oper<strong>at</strong>or <strong>of</strong> (7.2.2) is real, ψ ∗ (k, x) is also a solution <strong>of</strong>it whenever ψ(k, x) is a solution. Butψ ∗ (k, x) = e −ikx u ∗ p(x) (7.2.14)is also a Bloch-Floquet wave with −k wavenumber. In fact, u ∗ p(x) remains a periodic functionth<strong>at</strong> s<strong>at</strong>isfies an equ<strong>at</strong>ion th<strong>at</strong> is the conjug<strong>at</strong>e <strong>of</strong> (7.2.8) but with the same eigenvalue E.Therefore, E(−k) = E(k).7.3 Bloch-Floquet Theorem for 3DIn three dimensions, the Bloch-Floquet wave looks likeψ(k, r) = e ik·r u p (r) (7.3.1)


Quantum Mechanics in Crystals 81with the property th<strong>at</strong>u p (r + R L ) = u p (r) (7.3.2)It is to be noted th<strong>at</strong> u p (r) is a function <strong>of</strong> k as well. We can expand u p (r) as a generalizedFourier seriesu p (r) = ∑ Ga G e iG·r (7.3.3)where G represents points on the reciprocal l<strong>at</strong>tice denoted byG = l 1 b 1 + l 2 b 2 + l 3 b 3 (7.3.4)with the property th<strong>at</strong>wheree iG·R L= 1 (7.3.5)R L = n 1 a 1 + n 2 a 2 + n 3 a 3 (7.3.6)In the above, l i and n i , i = 1, 2, 3 are integers. The above property indic<strong>at</strong>es th<strong>at</strong> u p (r) in(7.3.3) is periodic as indic<strong>at</strong>ed by (7.3.2). The summ<strong>at</strong>ion in (7.3.2) is over all possible values<strong>of</strong> G or l 1 , l 2 , and l 3 .One can solve for b 1 ,b 2 and b 3 to yieldb 1 =a 2 × a 3a 1 · (a 2 × a 3 ) 2π, b 2 =a 3 × a 1a 1 · (a 2 × a 3 ) 2π, b 3 =for this G. Then by back substitution, (7.3.5) is s<strong>at</strong>isfied. Futhermore,wherea 1 × a 22π (7.3.7)a 1 · (a 2 × a 3 )ψ(k + G, r) = e ik·r e iG·r u p (r) (7.3.8)e iG·(r+R L) = e iG·r (7.3.9)Since it is a periodic function on the l<strong>at</strong>tice, the e iG·r term can be lumped with the unit cellfunction u p (r); therefore, the new Bloch-Floquet wave isψ(k + G, r) = e ik·r ũ p (r) (7.3.10)It is <strong>of</strong> the same form as (7.3.1); and hence will yield the same set <strong>of</strong> eigenvalues as ψ(k, r).Consequently, the l<strong>at</strong>tice points are defined by the l<strong>at</strong>tice vector R L , and the reciprocal l<strong>at</strong>ticepoints are defined by the reciprocal l<strong>at</strong>tice vector G. The Brillouin zone is defined by thereciprocal l<strong>at</strong>tice points. The first Brillouin zone is the first unique zone centered around theorigin. These zones in 3D can be quite complic<strong>at</strong>ed as shown in Figure 7.3.The band structure diagram is r<strong>at</strong>her complic<strong>at</strong>ed and is usually plotted with E versus kvector along prescribed lines in the first Brillouin zone as shown in Figure 7.4.We can take yet another viewpoint <strong>of</strong> the existence <strong>of</strong> band structure in a crystal. Wehave studied the tight-binding model <strong>of</strong> two identical quantum wells. This can be extended totwo identical <strong>at</strong>oms as shown in Figure 7.5. The figure shows how the energy <strong>of</strong> the trapped


82 Quantum Mechanics Made SimpleFigure 7.3: The first Brillouin zone <strong>of</strong> the FCC (face centered cubic) l<strong>at</strong>tice (from Wikipedia).Figure 7.4: The band structure diagram <strong>of</strong> Silicon along prescribed lines in the first Brillouinzone (from Warwich Physics Dept.).


Quantum Mechanics in Crystals 83Figure 7.5: The energy levels <strong>of</strong> two <strong>at</strong>oms versus the <strong>at</strong>omic spacing. (from Seeger, SemiconductorPhysics).modes <strong>of</strong> the <strong>at</strong>oms are split versus the spacing <strong>of</strong> the <strong>at</strong>oms (as we have found on our tightbinding model in the previous chapter). The closer the spacing, the stronger is the couplingbetween the modes, and the larger is the split in the levels <strong>of</strong> the degener<strong>at</strong>e modes.If we have a system with billions <strong>of</strong> <strong>at</strong>oms, each <strong>at</strong>om has its energy levels due to trappedmodes in the <strong>at</strong>om. If the inter-<strong>at</strong>omic distance is large, there is no interaction betweenthe <strong>at</strong>oms, and energy levels will be many billion times degener<strong>at</strong>e. However, when theinter<strong>at</strong>omic spacing becomes smaller, the wavefunctions from one <strong>at</strong>om will overlap with theneighboring <strong>at</strong>oms through tunneling. The inter<strong>at</strong>omic coupling, like the example <strong>of</strong> thetight-binding model, will split the energy levels and make them into a continuous band <strong>of</strong>energy. This give rise to energy band where energy st<strong>at</strong>es exist and energy band where nost<strong>at</strong>es exist, yielding the band structures we have seen.As we have seen in the simple case <strong>of</strong> the tight-binding model study <strong>of</strong> two quantum wells,the degener<strong>at</strong>e modes split into even and odd mode coupling between them. In one case,the modes are in phase, and the other case, the modes are out <strong>of</strong> phase with respect to eachother. When multitude <strong>of</strong> mode coupling prevails, there could be infinitely many possiblephase rel<strong>at</strong>ions between the <strong>at</strong>omic modes, giving rise to Bloch-Floquet wave with differentexp(ik · r) across the l<strong>at</strong>tice. The Bloch-Floquet wave establishes the phase rel<strong>at</strong>ionshipbetween different trapped modes <strong>of</strong> the <strong>at</strong>oms as they couple to each other.7.4 Fermi-Dirac Distribution FunctionNow th<strong>at</strong> we know how band structures come about, it will be prudent to introduce theFermi-Dirac distribution function. We will first present this function and derive it l<strong>at</strong>er inthe course. The Fermi-Dirac distribution function, also called the Fermi-Dirac function, isderived based on Boltzmann law and the Pauli exclusion principle for fermions since electronsare fermions. The Pauli exclusion principle st<strong>at</strong>es th<strong>at</strong> no two identical electrons can be inthe same st<strong>at</strong>e simultaneously. But they are non identical if they have different spins. Hence,each st<strong>at</strong>e can admit a spin-up electron, and a spin-down electron.


84 Quantum Mechanics Made SimpleFigure 7.6: The genesis <strong>of</strong> band structure as the inter<strong>at</strong>omic spacing <strong>of</strong> the <strong>at</strong>oms <strong>of</strong> thel<strong>at</strong>tice becomes smaller and smaller. (From M. Fox, Quantum Optics.)The Fermi-Dirac distribution function is given byf(E) =11 + e (E−E f )/(k B T )(7.4.1)where E f is the Fermi level also called the chemical potential. E f is actually a function <strong>of</strong>temper<strong>at</strong>ure, but if we assume it to be a constant, f(E) for different T ’s looks like thoseshown in Figure 7.7.The Fermi-Dirac function gives the probability th<strong>at</strong> a st<strong>at</strong>e with energy E is being occupied.Hence, 0 < f(E) < 1. For T = 0, it has a binary value <strong>of</strong> 1 for E < E f and 0for E > E f . So all energy st<strong>at</strong>es below E f are occupied and unoccupied for E > E f . ForT > 0, due to thermal agit<strong>at</strong>ion, some electrons are removed from E < E f and moved toE > E f st<strong>at</strong>es, giving rise to the distribution as shown. Therefore, as electrons are addedto the crystalline m<strong>at</strong>erial, the low energy st<strong>at</strong>es will be filled first, followed by high energyst<strong>at</strong>es. By the Pauli exclusion principle, each energy st<strong>at</strong>e can admit <strong>at</strong> most two electrons,one with spin up, and the other with spin down. Their probability <strong>of</strong> occupying the st<strong>at</strong>e isin accordance with Fermi-Dirac distribution function.7.4.1 Semiconductor, Metal, and Insul<strong>at</strong>orIn a semiconductor, E f is sandwiched between the conduction band and the valence band. AtT = 0, the valence band is completely occupied, and there are no electrons in the conductionband. Therefore, no conduction is possible. For T > 0, or <strong>at</strong> room temper<strong>at</strong>ure, the band


Quantum Mechanics in Crystals 85f(E)T=0T>0E fEFigure 7.7: The Fermi-Dirac distribution function <strong>at</strong> T = 0 and <strong>at</strong> T > 0.EEEConduction BandConduction BandConduction BandE fValence BandE fValence BandValence BandE fSemiconductorMetalInsul<strong>at</strong>orFigure 7.8: The rel<strong>at</strong>ionship <strong>of</strong> the Fermi level to the conduction and valence bands forsemiconductor, metal, and insul<strong>at</strong>or.gap is not large compared to k B T so th<strong>at</strong> some electrons can jump from the valence bandto the conduction band according to Fermi-Dirac distribution function. The electrons in theconduction band start to conduct. Moreover, empty st<strong>at</strong>es are left behind in the valence bandknown as holes, and they can also conduct electricity. For a metal, the Fermi level is in theconduction band. So even if T = 0, there are electrons in the conduction band allowing theflow <strong>of</strong> electrons. For an insul<strong>at</strong>or, the band gap is rel<strong>at</strong>ively large compared to k B T <strong>at</strong> roomtemper<strong>at</strong>ure, and the Fermi level is far from the conduction band, giving zero probabilityth<strong>at</strong> an electron is in the conduction band. Hence, it cannot conduct electricity.In semiconductors, the Fermi level moves toward the conduction band when the m<strong>at</strong>erialis doped with n type impurity since more electron carriers are gener<strong>at</strong>ed. Conversely, theFermi level moves toward the valence band when the m<strong>at</strong>erial is doped with p type impuritysince more holes are gener<strong>at</strong>ed.


86 Quantum Mechanics Made SimpleEE-∆EE+∆EkFigure 7.9: When the quantum system <strong>of</strong> a crystalline m<strong>at</strong>erial is in equilibrium, the Bloch-Floquet modes correspond to +k are equally likely compared to −k. Therefore, there is nonet momentum, and hence, no net movement <strong>of</strong> electrons. However, in the partially filledconduction band, the electrons can readjust themselves so th<strong>at</strong> some have larger +k valuescompared to −k values, giving rise to a net momentum, and hence movement <strong>of</strong> the electrons.7.4.2 Why Do Electrons and Holes Conduct Electricity?In some courses, we are told th<strong>at</strong> electrons conduct electricity because there are empty sites soth<strong>at</strong> electrons can move about and hence gives rise to conduction. But the wave picture <strong>of</strong> anelectron implies th<strong>at</strong> it can be simultaneously everywhere. From a wave picture viewpoint,this story can be quite different. In the valence band, for every electron th<strong>at</strong> has a kmomentum, there is a st<strong>at</strong>e with a −k momentum. The net momentum is always zero,and a fully filled valence band cannot conduct electricity. But in the conduction band, whenan electric field is applied, some electrons can move to a higher energy st<strong>at</strong>e E + ∆E withk + ∆k, while others can move to E − ∆E with −k + ∆k. Hence, there is a net gain <strong>of</strong>2∆k in the total momentum <strong>of</strong> the electrons giving rise to electron flow. Similarly, if emptyst<strong>at</strong>es are cre<strong>at</strong>ed in the valence band, the st<strong>at</strong>es can rearrange themselves so th<strong>at</strong> there is anet momentum for the electrons, and hence electricity flows. This is the wave picture <strong>of</strong> howelectrons and holes conduct electricty.7.5 Effective Mass Schrödinger Equ<strong>at</strong>ionThe bottom <strong>of</strong> many conduction band is parabolic. In the vicinity <strong>of</strong> k = 0, we can write theE-k rel<strong>at</strong>ion asE k = 2 k 22m e+ V (7.5.1)where m e is the effective mass chosen to fit the curv<strong>at</strong>ure <strong>of</strong> the parabola, and V is the energy<strong>at</strong> the bottom <strong>of</strong> the conduction band. It is to be emphasized th<strong>at</strong> this V is very differentfrom the V p th<strong>at</strong> has been discussed earlier.


Quantum Mechanics in Crystals 87We can write the Bloch-Floquet wave packet as a superposition <strong>of</strong> waves with differentenergies or k values, namelyψ(r, t) = ∑ kc k u k (r)e ik·r e −iω kt , k ≃ 0, ω k = E k (7.5.2)When k ≃ 0, we can approxim<strong>at</strong>eu k (r) ≃ u 0 (r) (7.5.3)so th<strong>at</strong>ψ(r, t) = u 0 (r) ∑ kc k e ik·r e −iω ktwhere ψ e (r, t) is the slowly varying envelope function, namely= u 0 (r)ψ e (r, t) (7.5.4)ψ e (r, t) = ∑ kc k e ik·r e −iω kt(7.5.5)We can show th<strong>at</strong>i ∂ψ e(r, t)∂t= ∑ c k E k e ik·r−iω ktk= ∑ ( 2 k 2 )c k + V e ik·r−iω kt2m ek=(− 2 ∇ 2 ) ∑+ V c k e ik·r−iω kt2m ek(7.5.6)ori ∂ψ e(r, t)∂t=(− 2 ∇ 22m e)+ V ψ e (r, t) (7.5.7)The above is the effective mass Schrödinger equ<strong>at</strong>ion. It is valid when k is small, or e ik·r isslowly varying compared to the l<strong>at</strong>tice spacing a. In the above, we can assume th<strong>at</strong> V (r) isslowly varying, and th<strong>at</strong> the effective mass m e (r) is also a slowly varying function <strong>of</strong> r. Thenthe above can be correctly written as− 22 ∇ · 1m e (r) ∇ψ e(r, t) + V (r)ψ e (r, t) = i ∂ ∂t ψ e(r, t) (7.5.8)The essence <strong>of</strong> the above approxim<strong>at</strong>ion is th<strong>at</strong> if the m<strong>at</strong>erial properties are slowly varyingwith respect to the l<strong>at</strong>tice spacing, an electron “sees” locally an environment <strong>of</strong> an infinite


88 Quantum Mechanics Made Simplel<strong>at</strong>tice. Consequently, if there are two domains with different V and m e , the above equ<strong>at</strong>ioninduces the boundary conditions <strong>at</strong> the interface <strong>of</strong> the domains as1m e1ψ e1 (r) = ψ e2 (r), r ∈ S (7.5.9)∂∂n ψ e1(r) = 1 ∂m e2 ∂n ψ e2(r), r ∈ S (7.5.10)where S is the interface separ<strong>at</strong>ing the two domains and ∂/∂n refers to normal deriv<strong>at</strong>ive.Even though V (r) and m e (r) have jump discontinuities <strong>at</strong> such an interface, we assume th<strong>at</strong>the discontinuities are still slowly varying compared to the l<strong>at</strong>tice spacings.It is to be noted th<strong>at</strong> the particle current th<strong>at</strong> we have defined earlier in (4.4.26), Chapter4, is proportional to1m e (r) ψ∗ e(r, t)∇ψ e (r, t)Hence, the boundary conditions (7.5.10) and (7.5.10) correspond to the continuity <strong>of</strong> particlecurrent across an interface.7.6 Heterojunctions and Quantum WellsVacuum Level1<strong>23</strong>Conduction BandValance BandFigure 7.10: Heterojunctions where the energy levels can be aligned using Anderson’s rule.Multiple heterojunctions can be used to make quantum wells.When two identical semiconductor crystalline m<strong>at</strong>erials are joint together, but dopeddifferently, the junction is known as a homojunction. However, when the two crystalline m<strong>at</strong>erialsare different, we have a heterojunction. We now know th<strong>at</strong> an electron propag<strong>at</strong>ingin a semiconductor crystalline l<strong>at</strong>tice resembles an electron in vacuum, except th<strong>at</strong> it has adifferent effective mass, and it sees a potential th<strong>at</strong> is the bottom <strong>of</strong> the conduction band.Hence, when two different m<strong>at</strong>erials are joint together to form a junction, we can use Anderson’srule to line up the energy level. Anderson’s rule requires th<strong>at</strong> the vacuum levels <strong>of</strong> thetwo m<strong>at</strong>erials be aligned. But the conduction band <strong>of</strong> the two m<strong>at</strong>erials need not be aligned,since the two m<strong>at</strong>erials can have different electron affinity χ. The electron affinity χ is theenergy needed to lift an electron from the bottom <strong>of</strong> the conduction band to the vacuum level.Because the electron follows the bottom <strong>of</strong> the conduction band, it will see a quantumwell when more than one heterojunctions is used. However, Anderson’s rule is not perfect,as the effective mass model is only good for infinite l<strong>at</strong>tice and imperfections can exist <strong>at</strong> theheterojunctions. Therefore, the rule is used as a guide.


Quantum Mechanics in Crystals 89In order to grow one m<strong>at</strong>erial on top <strong>of</strong> another m<strong>at</strong>erial without defects, the l<strong>at</strong>ticeconstants <strong>of</strong> the two m<strong>at</strong>erials have to be m<strong>at</strong>ched. A popular alloy compound for suchgrowth is alluminum gallium arsenide, Al x Ga 1−x As, where x denotes the percentage <strong>of</strong> Alin the alloy. The l<strong>at</strong>tice constant remains invariant as x is varied, but the bandgap <strong>of</strong> thecompound changes.The use <strong>of</strong> heterojunction quantum well confines the electrons to the well, allowing forthe making <strong>of</strong> the first room temper<strong>at</strong>ure semiconductor laser. A Nobel Prize was awardedto Herbert Kroemer and Zhores Alferov in 2000 for their contributions in this area.7.7 Density <strong>of</strong> St<strong>at</strong>es (DOS)We can use periodic boundary condition to discretize the k vector in the Bloch-Floquet waveso th<strong>at</strong> it is countable, or th<strong>at</strong>In one dimension, it isk L = l 12πL 1â 1 + l 22πL 2â 2 + l 32πL 3â 3 (7.7.1)k l = l 2π = l∆k (7.7.2)Lwhere ∆k = 2π L. When we count the number <strong>of</strong> discrete modes, we have 1 mode per ∆k onthe k l real line. Or the density <strong>of</strong> st<strong>at</strong>es isg 1D (k) = 1/∆k = L/2π (7.7.3)The number <strong>of</strong> st<strong>at</strong>es is proportional to the length <strong>of</strong> L. In 2D, the density <strong>of</strong> st<strong>at</strong>es can beeasily derived to beg 2D (k) = L 1 L 2 / (2π) 2 = A/ (2π) 2 (7.7.4)In 3D, it becomesg 3D (k) = L 1 L 2 L 3 / (2π) 3 = V/ (2π) 3 (7.7.5)We can normalize the density <strong>of</strong> st<strong>at</strong>es with respect to length L, area A, and volume V for1D, 2D, and 3D, respectively. Theng 1D (k) = 1/2π, g 2D (k) = 1/ (2π) 2 , g 3D (k) = 1/ (2π) 3 (7.7.6)We may want to count the density <strong>of</strong> st<strong>at</strong>es per unit energy instead as shown by the aboveE-k diagram. Even though the st<strong>at</strong>es are evenly spaced on the k line, they are not evenlyspread on the E line, thereby, altering the density <strong>of</strong> st<strong>at</strong>es per ∆E. In general, we can write,in 3D,g 3D (E) dE = 2g 3D (k) d 3 k = 2g (k) 4πk 2 dk= 8πg (k) k 2 dkdE dE (7.7.7)where we put a factor <strong>of</strong> 2 for two electron st<strong>at</strong>es per energy st<strong>at</strong>e.Therefore, the rel<strong>at</strong>ion between DOS on the E line and the k line isg 3D (E) = 8πg 3D (k)k 2 dkdE(7.7.8)


90 Quantum Mechanics Made SimpleE0 k 1 k 2 k 3 k 4kFigure 7.11: If every st<strong>at</strong>e on the k line corresponds to a st<strong>at</strong>e on the E line, then the density<strong>of</strong> st<strong>at</strong>es per unit length on the E line is different from th<strong>at</strong> on the k line.Sincefor a free electron, we haveE = 2 k 22m e+ V (7.7.9)k = √ √ 2meE − V(7.7.10)dkdE = √me2 2 1√E − V(7.7.11)andg 3D (E) = 1 ( ) 32me21(E − V )2π 2 2 2 (7.7.12)The above is the DOS in the conduction band <strong>of</strong> a bulk m<strong>at</strong>erial.7.7.1 Fermi Level and Fermi EnergyThe electron density in a bulk m<strong>at</strong>erial is given byN =∫ ∞0g 3D (E)f(E)dE (7.7.13)where f(E) is the Fermi-Dirac distribution function given in (7.4.1). When T changes, theelectron density remains constant, but f(E changes in shape. Therefore, E f in (7.4.1) has to


Quantum Mechanics in Crystals 91change to maintain the constancy <strong>of</strong> N: E f is a function <strong>of</strong> temper<strong>at</strong>ure T . The value <strong>of</strong> E f<strong>at</strong> T = 0 is known as the Fermi energy E F . In this case, the integral above can be trunc<strong>at</strong>edabruptly <strong>at</strong> E = E F , we can rewrite it asN =∫ EF0g 3D (E)f(E)dE (7.7.14)From knowing N which can be ascertained by counting the number <strong>of</strong> conduction electrons,we can obtain the value <strong>of</strong> E F , the Fermi energy, from the above equ<strong>at</strong>ion.7.7.2 DOS in a Quantum WellIn a 1D quantum well, the potential varies as a function <strong>of</strong> z, but is independent <strong>of</strong> x and y.Then the effective mass Schrödinger equ<strong>at</strong>ion becomes− 22 ∇ 1 ∇ψ(r) + V (z)ψ(r) = Eψ(r) (7.7.15)m e (z)We can further separ<strong>at</strong>e the z vari<strong>at</strong>ion from the x and y vari<strong>at</strong>ions to get− 2∇ 22msψ(r) − 2 ∂e 2 ∂zwhere ∇ 2 s = ∂2∂x+ ∂22 ∂y. 2By separ<strong>at</strong>ion <strong>of</strong> variables, we let1m ewhere r s = ˆxx + ŷy. Then, we can show th<strong>at</strong>andIn the above− 22ddz∂ψ(r) + V (z)ψ(r) = Eψ(r) (7.7.16)∂zψ(r) = ψ n (z)ψ s (r s ) (7.7.17)− 22m e∇ 2 sψ s (r s ) = E s ψ s (r s ) (7.7.18)1 dm e (z) dz ψ n(z) + V (z)ψ n (z) = E n ψ n (z) (7.7.19)E = E s + E n (7.7.20)Equ<strong>at</strong>ion (7.7.19) can be solved for the eigenvalues E n . For example, it can be the 1-D finitepotential well problem, or the infinite potential well problem. Altern<strong>at</strong>ively, it can have acomplic<strong>at</strong>ed V (z) and an m e (z) th<strong>at</strong> are amenable to numerical methods only.In the x-y direction, we letψ(r s ) ∝ e iks·rs (7.7.21)where k s = ˆxk x + ŷk y . Then,E s = 2 k 2 s2m e(7.7.22)where k 2 s = k 2 x + k 2 y. To make the solution simple, it is better to assume th<strong>at</strong> m e is constant. 2Therefore, the total E-k s diagram looks like th<strong>at</strong> shown in Figure 7.122 When m e is not a constant, the problem is still separable, but the solution is more complic<strong>at</strong>ed.


92 Quantum Mechanics Made SimpleFigure 7.12: The subband E-k s rel<strong>at</strong>ion <strong>of</strong> a quantum well.In general, g 2D (k s ) = 1(2π) 2 , and we haveg 2D (E s )dE s = 2g 2D (k s )2πk sdk sdE sdE s (7.7.<strong>23</strong>)Finally, we getg 2D (E s ) = m eπ 2 (7.7.24)which is a constant independent <strong>of</strong> E s .Hence, if we march along the E-line from 0 upward, the DOS as a function <strong>of</strong> total E isshown in Figure 7.13. We pick up density <strong>of</strong> st<strong>at</strong>es contribution from subband n as E > E n .Hence, the total DOS can be written asg 2D (E s ) = m ∑eπ 2 θ(E − E n ) (7.7.25)nFigure 7.13: The density <strong>of</strong> st<strong>at</strong>es <strong>of</strong> a 2D quantum well compared to th<strong>at</strong> <strong>of</strong> a 3D bulkm<strong>at</strong>erial.If the 1D quantum well is an infinite potential well, the energy is given by(E n =2 nπ) 2(7.7.26)2m e L


Quantum Mechanics in Crystals 93But density <strong>of</strong> st<strong>at</strong>es for a 3D bulk m<strong>at</strong>erial isg 3D (E) = 1 ( ) 32me21E2π 2 2 2 (7.7.27)If we evalu<strong>at</strong>e this <strong>at</strong> E = E n , we haveg 3D (E n ) = m (e n)nπ 2 = g 2DL L(7.7.28)Therefore, g 2DLtouches the g 3D <strong>at</strong> E n values as shown in Figure 7.13.When L becomes large, so th<strong>at</strong> the quantiz<strong>at</strong>ion level in the quantum well becomes finer,the DOS plot versus total E is shown in Figure 7.14. The DOS for a quantum well resemblesth<strong>at</strong> <strong>of</strong> a bulk m<strong>at</strong>erial.Figure 7.14: The density <strong>of</strong> st<strong>at</strong>es <strong>of</strong> a 2D quantum well compared to th<strong>at</strong> <strong>of</strong> a 3D bulkm<strong>at</strong>erial when the width <strong>of</strong> the quantum well L is large.7.7.3 Quantum WiresQuantum wires can be made by etching a quantum well until the electron is confined in arod-like region, like a dielectric rod waveguide. In this case, the wavefunction s<strong>at</strong>isfies thefollowing equ<strong>at</strong>ion− 22 ▽ s · 1m e▽ s ψ(r) −2 ∂ 22m e ∂zSimilar to separ<strong>at</strong>ion <strong>of</strong> variables, we assume th<strong>at</strong>Substituting (7.7.30) into (7.7.29), we have2ψ(r) + V (x, y)ψ(r) = Eψ(r) (7.7.29)ψ(r) = e ikzz ψ mn (x, y) (7.7.30)− 22 ▽ s · 1m e▽ s ψ(r s ) + V (r s )ψ mn (r s ) = E mn ψ mn (r) (7.7.31)


94 Quantum Mechanics Made Simpleg 1D vs Energy21.81.61.4g1D(E)/ √ 2me¯h1.210.80.60.40.200 1 2 3 4 5 6 7 8 9 10E/ ¯h22me10g 1D vs Energy987g1D(E)/ √ 2me¯h65432100 10 20 30 40 50 60 70E/ ¯h22meFigure 7.15: The density <strong>of</strong> st<strong>at</strong>es <strong>of</strong> a quantum wire showing the effect <strong>of</strong> the subbandsplotted on different scales.


Quantum Mechanics in Crystals 95E mn = E − 2 k 2 z2m e(7.7.32)We can assume m e to be constant to make the solution simple. The above form gives us√ 2me √k z = E − Emn (7.7.33)For infinite rectangular potential well case, an E mn can be found in closed form yielding[ (mπ ) 2 ( ) ] 2E mn =2nπ+(7.7.34)2m e L 1 L 2where L 1 and L 2 are the two sides <strong>of</strong> the rectangular rod. We can get the density <strong>of</strong> st<strong>at</strong>es asg 1D (E)dE = 4g 1D (k z ) dk zdE dE = 1 √ 2me 1√ dE (7.7.35)π E − EmnThe factor <strong>of</strong> 4 after the first equality comes about because for every energy st<strong>at</strong>e, twoelectrons can occupy it, and th<strong>at</strong> there are two k z st<strong>at</strong>es, corresponding to ±k z st<strong>at</strong>es, forevery energy st<strong>at</strong>e. Consequently,g 1D (E) = 1 π√ 2me1√ E − Emn(7.7.36)which is singular <strong>at</strong> E = E mn . When many subbands are included the DOS isg 1D (E) = 1 √ 2me ∑ 1√ θ(E − E mn ) (7.7.37)π E − Emnm,nwhere θ is a unit step function. The plot <strong>of</strong> density <strong>of</strong> st<strong>at</strong>es versus E is shown in Figure7.15. When the size <strong>of</strong> the quantum wire becomes large, the DOS resembles th<strong>at</strong> <strong>of</strong> the bulkm<strong>at</strong>erial.The density <strong>of</strong> st<strong>at</strong>es is important for determining the absorption and conduction properties<strong>of</strong> semiconductor m<strong>at</strong>erials. The availability <strong>of</strong> electron st<strong>at</strong>es indic<strong>at</strong>e the number <strong>of</strong>electrons th<strong>at</strong> can particip<strong>at</strong>e in electricity conduction. The availability <strong>of</strong> st<strong>at</strong>es in differentsubbands affects the optical absorption property <strong>of</strong> quantum wells as shown in Figure 7.16.The first peak corresponds to the absorption from the first subband, and the second peakcorresponds to the second subband absorption.In this chapter, when we discuss the propag<strong>at</strong>ion <strong>of</strong> an electron in a crystalline l<strong>at</strong>tice, wehave ignored the many-body effect, or the inter-electron effect. Each electron has a Coulombpotential around itself, and it will affect the nearby electrons. But this many body problem isa very difficult to solve, and is usually handled with the density function theory, introducedby Kohn and Sham. Walter Kohn was awarded the Nobel Prize in 1998 with John Pople onthis work.


96 Quantum Mechanics Made SimpleFigure 7.16: The density <strong>of</strong> st<strong>at</strong>es <strong>of</strong> a quantum well affects the intersubband optical absorption<strong>of</strong> a quantum well.


Chapter 8Angular Momentum8.1 IntroductionIn quantum mechanics, we learn th<strong>at</strong> travelling wave e ikx carries a linear momentum proportionalto k. Hence, the more oscill<strong>at</strong>ory the wave is, the higher the momentum it carries.However, when an electron is trapped inside the potential <strong>of</strong> an <strong>at</strong>om, the wave is not a linearlytravelling wave, but a wave th<strong>at</strong> swirls around the nucleus <strong>of</strong> the <strong>at</strong>om. Hence, the wavecarries mostly angular momentum. The magnitude <strong>of</strong> this angular momentum, intuitively, isproportional to the angular vari<strong>at</strong>ion <strong>of</strong> the wavefunction.If we take Schrödinger’s Equ<strong>at</strong>ion in free space, it is given byWe can rewrite this as− 22m 0∇ 2 ψ(r) = Eψ(r) (8.1.1)(∇ 2 + k02 )ψ(r) = 0 (8.1.2)where k 2 0 = 2m 0 E/ 2 . The above is the Helmholtz wave equ<strong>at</strong>ion whose solution is wellknown. In cylindrical coordin<strong>at</strong>es, it is( 1ρ∂∂ρ ρ ∂ ∂ρ + 1 ∂ 2)ρ 2 ∂φ 2 + ∂2∂z 2 + k2 0 ψ(ρ, φ, z) = 0 (8.1.3)where ρ = √ x 2 + y 2 , φ = tan −1 (y/x). In spherical coordin<strong>at</strong>es, it is( 1r 2 ∂∂r r2 ∂ ∂r + 1r 2 sin θ∂∂θ sin θ ∂ ∂θ + 1r 2 sin 2 θwhere r = √ x 2 + y 2 + z 2 , θ = cos −1 (z/r), φ = tan −1 (y/x).97∂ 2 )∂φ 2 + k2 0 ψ(r, θ, φ) = 0 (8.1.4)


98 Quantum Mechanics Made Simple8.1.1 Electron Trapped in a Pill BoxWe will first solve the above equ<strong>at</strong>ion in cylindrical coordin<strong>at</strong>es to gain physical insight intothe wave-angular momentum rel<strong>at</strong>ionship. Hence, we study the case <strong>of</strong> an electron trappedinside a pill box. The solutions to Helmholtz equ<strong>at</strong>ion can be found by the separ<strong>at</strong>ion <strong>of</strong>variables. For example, to solve (8.1.3), we letUsing (8.1.5) in (8.1.3) yieldsψ(ρ, φ, z) = R(ρ)Φ(φ)Z(z) (8.1.5)Φ(φ)Z(z) 1 ∂ρ ∂ρ ρ ∂ ∂ρ R(ρ) + R(ρ)Z(z) 1 ∂ 2ρ 2 ∂φ 2 Φ(φ)To solve the above, we propose two eigenvalue problems,The solutions to the above are easily found:By so doing, Equ<strong>at</strong>ion (8.1.6) becomes+ R(ρ)Φ(φ) ∂2∂z 2 Z(z) + k2 0R(ρ)Φ(φ)Z(z) = 0 (8.1.6)∂ 2∂z 2 Z(z) = −k2 zZ(z) (8.1.7)∂ 2∂φ 2 Φ(φ) = −n2 Φ(φ) (8.1.8)Z(z) = e ±ikzz (8.1.9)Φ(φ) = e ±inφ (8.1.10)where1 dρ dρ ρ d n2R(ρ) −dρ ρ 2 R(ρ) + k2 ρR(ρ) = 0 (8.1.11)k 2 ρ = k 2 0 − k 2 z (8.1.12)Equ<strong>at</strong>ion (8.1.11) is the Bessel equ<strong>at</strong>ion where solutions are Bessel function J n (k ρ ρ) andNeumann function Y n (k ρ ρ). Here, n is the order <strong>of</strong> these functions. So the general solution(8.1.5) becomesψ(ρ, φ, z) = [AJ n (k ρ ρ) + BY n (k ρ ρ)] e ±inφ e ±ikzz (8.1.13)We can now put an electron in a pill box whose wall is an infinite potential as shownin Figure 8.1. Since Y n (k ρ ρ) → ∞, when ρ → 0, we set B = 0 to elimin<strong>at</strong>e the Neumannfunction. Also, ψ(ρ, φ, z) has to be regular and finite inside the pill box. M<strong>at</strong>ching boundaryconditions <strong>at</strong> the two ends <strong>of</strong> the pill box, we have( pπ)ψ (ρ, φ, z) = AJ n (k ρ ρ) e ±inφ sinL z (8.1.14)


Angular Momentum 99zz=L2ayxφρFigure 8.1: The geometry <strong>of</strong> a cylindrical pill box where an electron is trapped.Table 8.1: The first few zeros <strong>of</strong> Bessel functions:m ζ 0m ζ 1m ζ 2m ζ 3m ζ 4m ζ 5m1 2.4048 3.8317 5.1356 6.3802 7.5883 8.77152 5.5201 7.0156 8.4172 9.7610 11.0647 12.33863 8.6537 10.1735 11.6198 13.0152 14.3725 15.70024 11.7915 13.3<strong>23</strong>7 14.7960 16.2<strong>23</strong>5 17.6160 18.98015 14.9309 16.4706 17.9598 19.4094 20.8269 22.2178where k z = pπ/L, for all integer p.The boundary condition <strong>at</strong> ρ = a is th<strong>at</strong> ψ (a, φ, z) = 0, or th<strong>at</strong>The zeros <strong>of</strong> Bessel functions are tabul<strong>at</strong>ed and given byThe first few zeros are given in Table 8.1. Hence, we obtain th<strong>at</strong>So (8.1.14) becomesJ n (k ρ a) = 0 (8.1.15)J n (ζ nm ) = 0 (8.1.16)k ρ = ζ nm /a (8.1.17)(ρ) ( pπ)ψ (ρ, φ, z) = AJ n ζ nm e ±inφ sinaL z(8.1.18)In the above, we see a standing wave sin (pπz/L) in the z direction, a standing wave representedby J n(ζnmρa)in the ρ direction, and a travelling wave e ±inφ in the φ direction. The


100 Quantum Mechanics Made Simple1J 0(x)J 1(x)J 2(x)0.5J n(x)0-0.50 5 10 15xFigure 8.2: The Bessel function <strong>of</strong> different order as a function <strong>of</strong> its argument. Notice th<strong>at</strong>the higher the order, the smaller the Bessel function is <strong>at</strong> the origin.larger n is, the more rapidly varying is the travelling wave as a function <strong>of</strong> φ, and the moreangular momentum it carries.In fact, plots <strong>of</strong> J n (x) versus x for different n is shown in Figure 8.2. 1 We notice th<strong>at</strong>the larger n is, the more rapidly varying e ±inφ is, and the more the Bessel function J n (x) ispulled away from the origin. This is characteristic <strong>of</strong> a particle swirling around in a circle.The centrifugal force is keeping the particle away from the origin.From the fact th<strong>at</strong>( ) 2k0 2 = kρ 2 + kz 2 ζnm( pπ) 2 2m 0 E= + =a L 2 (8.1.19)we can derive th<strong>at</strong> the energy levels <strong>of</strong> the trapped electron are given by[ (ζnm ) ]2 (E nmp =2pπ) 2+2m 0 a L(8.1.20)Hence, the energy levels <strong>of</strong> the trapped electron assume quantized values. This is the characteristic<strong>of</strong> the wave n<strong>at</strong>ure <strong>of</strong> the electron.8.1.2 Electron Trapped in a Spherical BoxTo obtain different physical insight with different wavefunctions, next, we place the electronin a spherical box with infinite potential <strong>at</strong> the wall. We also solve this problem by separ<strong>at</strong>ion<strong>of</strong> variables by lettingψ (r, θ, φ) = R (r) Θ (θ) Φ (φ) (8.1.21)1 Ways to compute J n(x) is well documented and is available in programming toolbox like M<strong>at</strong>lab.


Angular Momentum 101Figure 8.3: Legendre Polynomial with argument cos θ plotted for l = 5 and different m values.Notice th<strong>at</strong> as m increases, the z component <strong>of</strong> the angular increases, and the wavefunctionspirals away from the z axis, due to the centra-fugal force <strong>of</strong> the electron. Also, the θ vari<strong>at</strong>ion<strong>of</strong> the wavefunction slows down, indic<strong>at</strong>ing a lesser component <strong>of</strong> the angular momentum inthe transverse direction.Using (8.1.21) in (8.1.4), we haveΘ (θ) Φ (φ) 1r 2∂∂r r2 ∂ ∂r R (r) + R (r) Φ (φ) 1+R (r) Θ (θ)1r 2 sin 2 θ(∂r 2 sin θ ∂θ sin θ∂∂θ Θ (θ))∂ 2∂φΦ (φ) + k0R 2 (r) Θ (θ) Φ (φ) = 02(8.1.22)We propose another eigenvalue problem such th<strong>at</strong>∂ 2∂φ 2 Φ (φ) = −m2 Φ (φ) (8.1.<strong>23</strong>)Using the above in (8.1.22), Φ (φ) can be canceled in the equ<strong>at</strong>ion. Then we propose anothereigenvalue problem such th<strong>at</strong>1 ∂sin θ ∂θ sin θ ∂ ∂θ Θ (θ) −m2sin 2 Θ (θ) = −l (l + 1) Θ (θ) (8.1.24)θ


102 Quantum Mechanics Made SimpleThe solution to (8.1.<strong>23</strong>) is e ±imφ while the solution to (8.1.24) is in terms <strong>of</strong> associ<strong>at</strong>edLegendre polynomials, Pl m (x), or in terms <strong>of</strong> θ, 2Eventually, the equ<strong>at</strong>ion for R (r) isΘ (θ) = P ml(cos θ) , −l m l (8.1.25)1 d d []r 2 dr r2 dr R (r) + k0 2 l (l + 1)−r 2 R (r) = 0 (8.1.26)The last equ<strong>at</strong>ion is the spherical Bessel equ<strong>at</strong>ion whose solution isR (r) = Aj l (k 0 r) + Bh (1)l(k 0 r) (8.1.27)In the above, j l (x) represents a spherical Bessel function <strong>of</strong> order l, which is regular <strong>at</strong> x = 0,while h (1)l(x) represents a spherical Hankel function <strong>of</strong> the first kind <strong>of</strong> order l, which issingular <strong>at</strong> x = 0.When we put this solution inside a spherical box, we set B = 0 since h (1)l(k 0 r) is singular<strong>at</strong> r = 0. Then the general solution to (8.1.4) inside a box isψ(r, θ, φ) = Aj l (k 0 r)P ml (cos θ)e ±imφ , −l m l (8.1.28)Assuming the radius <strong>of</strong> the sphere is a, to s<strong>at</strong>isfy the boundary condition, we require th<strong>at</strong>j l (k 0 a) = 0, (8.1.29)Since j l (ζ lp ) = 0, we deduce th<strong>at</strong> k 0 = ζ lp /a.functions are given in Table 8.2.The first few zeros <strong>of</strong> the spherical BesselTable 8.2: The first few zeros <strong>of</strong> the spherical Bessel functions:p ζ 0p ζ 1p ζ 2p ζ 3p ζ 4p1 3.142 4.493 5.763 6.988 8.1832 6.283 7.725 9.095 10.417 11.7053 9.425 10.904 12.3<strong>23</strong> 13.698 15.0404 12.566 14.066 15.515 16.924 18.301Therefore, the general solution becomes(r)ψ(r, θ, φ) = Aj l ζ lp Pl m (cos θ)e ±imφ ,a−l m l (8.1.30)In the above, e ±imφ represents a traveling wave in the φ direction, Plm (cos θ) represents astanding wave in the θ direction. When l becomes large, Plm (x) is a higher order polynomial2 The other solution to (8.1.24) is the associ<strong>at</strong>e Legendre function <strong>of</strong> the second kind, Q n l (cos θ) but thisfunction is singular for 0 θ π, and hence, is not admissible as a solution.


Angular Momentum 103and Plm (cos θ) is a rapidly varying function <strong>of</strong> θ. In this case, e ±imφ , −l m l could alsobe a rapidly varying function <strong>of</strong> φ implying the presence <strong>of</strong> high angular momentum. FromFigure 8.3, it is seen th<strong>at</strong> the larger m is, the more slowly varying is the associ<strong>at</strong>ed Legendrepolynomial. The vari<strong>at</strong>ions from Plm (cos θ) and e ±imφ compens<strong>at</strong>e each other for a fixed l soas to keep the “sum” <strong>of</strong> the angular momentum a constant when l is fixed.The function j l (x) is well known and is plotted in Figure 8.4, showing th<strong>at</strong> when n islarge, the centrifugal force <strong>of</strong> the particle pulls it away from the origin as in the pill-box casein cylindrical coordin<strong>at</strong>es. From the fact th<strong>at</strong>k 2 0 =we deduce th<strong>at</strong> the energy levels <strong>of</strong> the trapped electron areE lp =( ) 2 ζlp= 2m 0Ea 2 (8.1.31)( ) 2 2 ζlp(8.1.32)2m 0 aThis is again quantized due to the wave n<strong>at</strong>ure <strong>of</strong> the electron.The above cases illustr<strong>at</strong>e the trapping <strong>of</strong> an electron by a pill box and a spherical box.In n<strong>at</strong>ure, the positive charge <strong>of</strong> a nucleus forms a potential th<strong>at</strong> can trap an electron. Inthe case <strong>of</strong> the hydrogen <strong>at</strong>om, closed form solutions can be obtained for the trapping <strong>of</strong>an electron by the Coulomb potential <strong>of</strong> the positive nucleus. This has been documented inmany quantum mechanics books.8.2 M<strong>at</strong>hem<strong>at</strong>ics <strong>of</strong> Angular Momentum10.80.6j 0(x)j 1(x)j 2(x)j n(x)0.40.20-0.20 5 10 15xFigure 8.4: Spherical Bessel function <strong>of</strong> different order versus its argument. The higher orderfunctions are smaller around the origin.


104 Quantum Mechanics Made SimpleWe have shown previously th<strong>at</strong> an electron trapped inside a pill box or a spherical boxdoes display angular momentum. Next, we derive the oper<strong>at</strong>or th<strong>at</strong> is rel<strong>at</strong>ed to angularmomentum. In classical mechanics, the angular momentum is given byL = r × p (8.2.1)where p is the linear momentum. Notice th<strong>at</strong> by taking the cross product <strong>of</strong> p with r, onlythe circumferential component <strong>of</strong> the linear momentum is extracted. Written explicitly, it isL x = yp z − zp y , L y = zp x − xp z , L z = xp y − yp x (8.2.2)We raise these to oper<strong>at</strong>ors, and write them as, in coordin<strong>at</strong>e space represent<strong>at</strong>ion,(ˆL x = −i y ∂ ∂z − z ∂ ) (,∂yˆL y = −i z ∂∂x − x ∂ ) (,∂zˆL z = −i x ∂ ∂y − y ∂ ), (8.2.3)∂xx, y, z can be raised to be oper<strong>at</strong>ors as well, but in coordin<strong>at</strong>e space represent<strong>at</strong>ion, they arejust scalars. Actually, we can write (8.2.1) as the oper<strong>at</strong>orˆL = ˆr × ˆp (8.2.4)The h<strong>at</strong> signs above indic<strong>at</strong>e th<strong>at</strong> they are oper<strong>at</strong>ors and not unit vectors. The coordin<strong>at</strong>espace represent<strong>at</strong>ion <strong>of</strong> the ˆr oper<strong>at</strong>or is just r. The coordin<strong>at</strong>e space represent<strong>at</strong>ion represent<strong>at</strong>ion<strong>of</strong> the momentum oper<strong>at</strong>or ˆp has been derived before. Consequently, the aboveoper<strong>at</strong>ors can be shown to s<strong>at</strong>isfy the following commut<strong>at</strong>ion rel<strong>at</strong>ions[ˆLx , ˆL]y = iˆL z ,[ˆLy , ˆL]z = iˆL x ,[ˆLz , ˆL]x = iˆL y (8.2.5)The ˆL 2 oper<strong>at</strong>or is defined asˆL 2 = ˆL · ˆL = ˆL 2 x + ˆL 2 y + ˆL 2 z (8.2.6)8.2.1 Transforming to Spherical Coordin<strong>at</strong>esThe above can be transformed to spherical coordin<strong>at</strong>e system. To show this, we write (8.2.4)in coordin<strong>at</strong>e represent<strong>at</strong>ion yielding 3(ˆL = −ir × ∇ = −i ˆφ ∂ ∂θ − ˆθ 1 )∂(8.2.7)sin θ ∂φThe last equality can be obtained by writing gradient oper<strong>at</strong>or ∇ in spherical coordin<strong>at</strong>es.Then(ˆL x = ˆx · ˆL = −i ˆx · ˆφ ∂ )1 ∂− ˆx · ˆθ (8.2.8)∂θ sin θ ∂φBut ˆx · ˆφ = − sin φ, ˆx · ˆθ = ˆx · ˆρ cos θ = cos φ cos θ. Then (8.2.8) becomes(ˆL x = i sin φ ∂ )∂+ cot θ cos φ∂θ ∂φ3 We will use the ˆx, ŷ, ẑ, ˆr, ˆθ, ˆφ to denote unit vectors in the respective directions.(8.2.9)


Angular Momentum 105Similarly,(ˆL y = ŷ · ˆL = −i ŷ · ˆφ ∂ ∂θ − ŷ · ˆθ 1 )∂sin θ ∂φ(8.2.10)with ŷ · ˆφ = cos φ, ŷ · ˆθ = sin φ cos θ, we have(ˆL y = i − cos φ ∂ )∂+ cot θ cos φ∂θ ∂φSimilarly,ˆL z = iẑ · ˆθ 1 ∂sin θ ∂φ = −i ∂∂φThe eigenfunction <strong>of</strong> the ˆL z oper<strong>at</strong>or is Φ = e ±imφ so th<strong>at</strong>(8.2.11)(8.2.12)ˆL z Φ(φ) = mΦ(φ) (8.2.13)For ˆL 2 oper<strong>at</strong>or, we haveˆL 2 = − 2 (r × ∇) · (r × ∇) (8.2.14)Using a vector identity th<strong>at</strong> (a × b) · c = a · b × c, we rewrite the above asˆL 2 = − 2 r · ∇ × (r × ∇) (8.2.15)By expressing the ∇× oper<strong>at</strong>or in spherical coordin<strong>at</strong>es, and th<strong>at</strong> r × ∇ can be expressed asin (8.2.7), the above becomes(ˆL 2 = − 2 1 ∂sin θ ∂θ sin θ ∂ ∂θ + ∂∂φ1sin θ)∂∂φ(8.2.16)The above is just the circumferential part <strong>of</strong> the Laplacian oper<strong>at</strong>or in (8.1.4). It is quiteeasy to show th<strong>at</strong> the above commutes with ˆL z . Hence, ˆL 2 commutes with ˆL z . By symmetry,ˆL 2 also commutes with ˆL x and ˆL y .We can find the eigenfunctions <strong>of</strong> the ˆL 2 oper<strong>at</strong>or. Traditionally, this eigenfunction andeigenvalue problem is expressed asˆL 2 Y lm (θ, φ) = 2 l(l + 1)Y lm (θ, φ), −l m l (8.2.17)Its solution is similar to the method <strong>of</strong> solving for the solution <strong>of</strong> Helmholtz wave equ<strong>at</strong>ionin spherical coordin<strong>at</strong>es. By the separ<strong>at</strong>ion <strong>of</strong> variables, the solution isY lm (θ, φ) = CP ml (cos θ)e imφ , −l m l (8.2.18)which can be normalized to yield√Y lm (θ, φ) = (−1) m (2l + 1)(l − m)!Pl m (cos θ)e imφ , −l m l (8.2.19)4π(l + m)!


106 Quantum Mechanics Made SimpleIn Hilbert space, this eigenfunction is denoted as |l, m〉. We can define another oper<strong>at</strong>orˆL 2 s = ˆL 2 − ˆL 2 z = ˆL 2 x + ˆL 2 y (8.2.20)It is quite clear th<strong>at</strong> ˆL z commutes with ˆL 2 s, and ˆL 2 commutes with ˆL 2 s. Hence, ˆL z , ˆL 2 s andˆL 2 are mutually commut<strong>at</strong>ive, and share the same eigenfunctions |l, m〉. Therefore,ˆL 2 s|l, m〉 = 2 [l(l + 1) − m 2 ]|l, m〉 (8.2.21)A pictorial illustr<strong>at</strong>ion <strong>of</strong> this concept is given in Figure 8.5. It is to be noted th<strong>at</strong> thispicture can be drawn because ˆL z , ˆL 2 s, and ˆL 2 are mutually commut<strong>at</strong>ive. For the commoneigenfunction |l, m〉, their eigenvalues can be “observed” uniquely and simultaneously. 4 Thecorresponding increase <strong>of</strong> L z and decrease <strong>of</strong> L s also correl<strong>at</strong>e with the wavefunctions shownin Figure 8.3.However, ˆL z does not commute with ˆL x or ˆL y . Therefore, for an eigenfunction |l, m〉 th<strong>at</strong>gives the value <strong>of</strong> 〈ˆL z 〉 precisely, the value <strong>of</strong> 〈ˆL x 〉 〈ˆL y 〉 are indetermin<strong>at</strong>e. In fact, from thepicture <strong>of</strong> |Y lm | 2 , the function is completely indetermin<strong>at</strong>e in the φ direction.4 We are actually drawing the eigenvalues <strong>of</strong> ˆL s =√ˆL2 x + ˆL 2 y and ˆL =√ˆL2 x + ˆL 2 y + ˆL 2 z , but they are stillmutually commut<strong>at</strong>ive.


Angular Momentum 107Figure 8.5: A picture showing the angular momentum components, L z and L s , for differenttilt angles <strong>of</strong> the total angular momentum L.


108 Quantum Mechanics Made Simple


Chapter 9Spin9.1 IntroductionSpin is a special property <strong>of</strong> <strong>at</strong>omic or sub<strong>at</strong>omic particles th<strong>at</strong> has no classical analogue.Electron has spin. We can think <strong>of</strong> it as being due to the self spinning <strong>of</strong> the electron, butwe should not let our imagin<strong>at</strong>ion run further than th<strong>at</strong>. Spin <strong>of</strong> an electron gives it a spinangular momentum in addition to the orbital angular momentum it possesses. The spin alsoendows an electron with a magnetic dipole moment th<strong>at</strong> causes it to interact with a magneticfield.The spin <strong>of</strong> some particles is found to have binary values <strong>of</strong> “spin up” and “spin down”experimentally by the famous Stern-Gerlach experiment. This binary n<strong>at</strong>ure, as we shall see,fits nicely in the m<strong>at</strong>hem<strong>at</strong>ical structure <strong>of</strong> angular momentum in quantum mechanics, butit cannot be described by a wavefunction or wave mechanics. Instead, it can be representedby m<strong>at</strong>rix mechanics.9.2 Spin Oper<strong>at</strong>orsWe have seen th<strong>at</strong> the z component <strong>of</strong> the orbital angular momentum, represented by theoper<strong>at</strong>or ˆL z , is quantized to be m where −l m l, l being an integer rel<strong>at</strong>ed to the totalangular momentum square oper<strong>at</strong>or ˆL 2 with eigenvalue l (l + 1) 2 .It can be shown th<strong>at</strong> the rel<strong>at</strong>ionship between the total angular momentum number l andthe z-component <strong>of</strong> the angular number m is not restricted to orbital angular momenta. Itcan be established for all quantum mechanical angular momenta, as is shown in AppendixA. A generalized angular momentum can be J = L + S, J = L 1 + L 2 , or J = S 1 + S 2 . Likelinear momentum, angular momentum can be added like vectors. Also, they are conservedquantities. In the quantum world, these observables will be elev<strong>at</strong>ed to become oper<strong>at</strong>ors.A more general framework for angular momentum is th<strong>at</strong> for Ĵ 2 = Ĵ 2 x+Ĵ 2 y +Ĵ 2 z , an oper<strong>at</strong>orth<strong>at</strong> represents the square <strong>of</strong> the total angular momentum, and Ĵx, Ĵ y , Ĵ z , oper<strong>at</strong>ors th<strong>at</strong>represent the x, y, and z components <strong>of</strong> angular momenta, thenĴ 2 |L, M〉 = L (L + 1) 2 |L, M〉 (9.2.1)109


110 Quantum Mechanics Made SimpleĴ z |L, M〉 = M 2 |L, M〉 , −L M L (9.2.2)We have proved the above results for orbital angular momentum by using wave mechanicsand wavefunctions, but they can be proven for general angular momentum by using rot<strong>at</strong>ionalsymmetry <strong>of</strong> 3D coordin<strong>at</strong>e space, and m<strong>at</strong>hem<strong>at</strong>ics <strong>of</strong> raising and lowering oper<strong>at</strong>ors (seeExercise <strong>at</strong> the end <strong>of</strong> the Chapter).Spin angular momentum oper<strong>at</strong>ors also fit under the framework <strong>of</strong> general angular momentumoper<strong>at</strong>or, and can be thought <strong>of</strong> as a special case <strong>of</strong> the above framework. Hence,the above can also be applied to spin 1 2 angular momentum where L = 1 2 , − 1 2 M 1 2 . Thevalues <strong>of</strong> M has to be spaced by value one apart, hence, M = ± 1 2 .For spins, we let Ŝ represent the total angular momentum oper<strong>at</strong>or, while Ŝz representsthe z component <strong>of</strong> the spin angular momentum. Applying (9.2.1) and (9.2.2) to spin 1 2angular momentum, we haveŜ ∣ 2 12 , ± 〉 ∣ 12 =34 2 12 , ± 〉 12 , L =12 , M = ± 1 (9.2.3)2∣Ŝ z 12 , ± 〉 12 = ±12 ∣ 12 , ± 〉12(9.2.4)As a result, the corresponding z component <strong>of</strong> the spin angular momentum, representedby the oper<strong>at</strong>or Ŝz, has only two eigenvalues and two eigenst<strong>at</strong>es: an up st<strong>at</strong>e with angularmomentum <strong>of</strong> 1 2 and a down st<strong>at</strong>e with angular momentum <strong>of</strong> − 1 2 .1The corresponding x and y components <strong>of</strong> the spin angular momentum can be representedby oper<strong>at</strong>ors Ŝx and Ŝy. Together with Ŝz, they s<strong>at</strong>isfy the following commut<strong>at</strong>ion rel<strong>at</strong>ions,[Ŝx Ŝy], = iŜz,[Ŝy Ŝz], = iŜx,[Ŝz Ŝx], = iŜy (9.2.5)The above is similar to the commut<strong>at</strong>ion rel<strong>at</strong>ions s<strong>at</strong>isfied by ˆL x , ˆL y , and ˆL z , where theyhave been motiv<strong>at</strong>ed by wave mechanics. However, as has been shown in Appendix A, th<strong>at</strong> ifan oper<strong>at</strong>or is to represent an angular momentum, then their x, y, and z components have tos<strong>at</strong>isfy the above commut<strong>at</strong>ion rel<strong>at</strong>ions by rot<strong>at</strong>ional symmetry <strong>of</strong> the 3D coordin<strong>at</strong>e space.It is expedient <strong>at</strong> this point to define Pauli spin m<strong>at</strong>rices given byˆσ x = 2 Ŝx, ˆσ y = 2 Ŝy, ˆσ z = 2 Ŝz (9.2.6)with the commut<strong>at</strong>ion rel<strong>at</strong>ion[ˆσ x , ˆσ y ] = 2iˆσ z , [ˆσ y , ˆσ z ] = 2iˆσ x , [ˆσ z , ˆσ x ] = 2iˆσ y (9.2.7)Since there are only two spin st<strong>at</strong>es, they can be represented by a two dimensional columnvector. Hence, we have|↑〉 = [ 1 ] [ 0, |↓〉 =(9.2.8)1]1 This was discovered in silver <strong>at</strong>oms by the famous Stern-Gerlach experiment.


Spin 111The spin oper<strong>at</strong>ors can in turn be represented by 2 × 2 m<strong>at</strong>rices. Then, the m<strong>at</strong>rix represent<strong>at</strong>ion<strong>of</strong> ˆσ z is[ ] 1 0ˆσ z =(9.2.9)0 −1such th<strong>at</strong>ˆσ z |↑〉 = |↑〉 , ˆσ z |↓〉 = − |↓〉 (9.2.10)The other Pauli m<strong>at</strong>rices can be found to be[ ]0 1ˆσ x = , ˆσ1 0y =[ ]0 −ii 0(9.2.11)It can be shown th<strong>at</strong> any 2 × 2 m<strong>at</strong>rix can be expanded in terms <strong>of</strong> the Pauli m<strong>at</strong>rices andthe identity m<strong>at</strong>rix; namely,A = a 1ˆσ x + a 2ˆσ y + a 3ˆσ z + a 4 IFurthermore, it can be shown th<strong>at</strong>ˆσ 2 x = ˆσ 2 y = ˆσ 2 z = I =[ ] 1 00 1(9.2.12)Hence, we can define a ˆσ 2 oper<strong>at</strong>or such th<strong>at</strong>ˆσ 2 = ˆσ 2 x + ˆσ 2 y + ˆσ 2 z = 3I (9.2.13)By the same token, the Ŝ2 oper<strong>at</strong>or, from (9.2.6) and the above, can be evalu<strong>at</strong>ed to beŜ 2 = Ŝ2 x + Ŝ2 y + Ŝ2 z = 24 ˆσ2 = 3 I (9.2.14)4It can be shown th<strong>at</strong> the eigenvector <strong>of</strong> the ˆσ x oper<strong>at</strong>or is|x〉 = √ 1 [ ] 1= √ 1 [| ↑〉 + | ↓〉]2 1 2while th<strong>at</strong> for the ˆσ y oper<strong>at</strong>or is|y〉 = 1 √2[ 1i]= 1 √2[| ↑〉 + i| ↓〉]It indic<strong>at</strong>es th<strong>at</strong> the st<strong>at</strong>e vector th<strong>at</strong> represents a spin pointing in the x or y direction isa linear supposition <strong>of</strong> an up spin st<strong>at</strong>e vector and a down spin st<strong>at</strong>e vector with differentphases.In the above, the quantum mechanics <strong>of</strong> spin half particle such as an electron is describedby m<strong>at</strong>rix mechanics.


112 Quantum Mechanics Made Simple9.3 The Bloch SphereWe have seen th<strong>at</strong> the eigenst<strong>at</strong>es <strong>of</strong> ˆσ z are[ ] [ 1 0|↑〉 = , |↓〉 =01](9.3.1)In the two dimensional space in which spin st<strong>at</strong>es are represented, these two st<strong>at</strong>es are orthonormaland complete. Hence, the eigenst<strong>at</strong>es <strong>of</strong> the ˆσ x and ˆσ y oper<strong>at</strong>ors can be expandedas well in terms <strong>of</strong> |↑〉 and |↓〉 st<strong>at</strong>es. Consequently, a general spin st<strong>at</strong>e can be written as|s〉 = a ↑ |↑〉 + a ↓ |↓〉 = cos( θ2)|↑〉 + e iφ sin( θ2)|↓〉 (9.3.2)We have chosen a ↑ and a ↓ judiciously so th<strong>at</strong> the wavefunction is clearly normalized. Therel<strong>at</strong>ive phase between a ↑ and a ↓ is in e iφ as absolute phase is unimportant. Letσ = ıˆσ x + jˆσ y + kˆσ z (9.3.3)be an oper<strong>at</strong>or th<strong>at</strong> represents a vector <strong>of</strong> the spin angular momentum, where ı, j, and kare unit vectors in the x, y, and z directions. Then the expect<strong>at</strong>ion value <strong>of</strong> this oper<strong>at</strong>orfor a given quantum st<strong>at</strong>e |s〉 in (9.3.2) should point in a vectorial direction. This is a basictenet <strong>of</strong> quantum mechanics since σ represents angular momentum, which is an observable.Consequently, we find the vectorP s = 〈s| σ |s〉 = ı 〈s| ˆσ x |s〉 + j 〈s| ˆσ y |s〉 + k 〈s| ˆσ z |s〉 (9.3.4)After substituting |s〉 from (9.3.2) into the above, we can show th<strong>at</strong>P s = ı sin θ cos φ + j sin θ sin φ + k cos θ (9.3.5)The vector P s maps out a sphere called the Bloch sphere with corresponding θ and φ valuesas shown in Figure 9.1.The above analysis shows th<strong>at</strong> by superposing the up and down spin st<strong>at</strong>es judiciously,the observed spin vector, which is the expect<strong>at</strong>ion <strong>of</strong> the vector spin oper<strong>at</strong>or, can be made topoint in any directions in the 3D space. This fe<strong>at</strong>ure has been <strong>of</strong> gre<strong>at</strong> interest in constructinga quantum computer and storing quantum inform<strong>at</strong>ion.9.4 SpinorIn general, we need a two component vector to describe both the wavefunction and the spinst<strong>at</strong>e <strong>of</strong> an electron, namely,|ψ〉 =[|ψup 〉|ψ down 〉][ 1= |ψ up 〉0] [ 0+ |ψ down 〉1](9.4.1)These wave fucntions with two components are known as spinors. The vector space for spinorsconsists <strong>of</strong> the vector space for the electron wavefunction multiplying with the two-dimensionalvector space for the spin. This is known as a direct product space.


Spin 1139.5 Pauli Equ<strong>at</strong>ionFigure 9.1: The Bloch sphere showing the vector P s .The classical angular momentum <strong>of</strong> a particle moving in a circle iswhich can also be written as a vectorL = m 0 vr (9.5.1)L = r × p = r × m 0 v (9.5.2)If this particle is an electron, it will complete an orbit in 2πr/v time, and we can denote thecurrent due to this orbiting electron asI = − ev2πrA circul<strong>at</strong>ing current loop produces a magnetic dipole moment(9.5.3)µ d = Ia (9.5.4)where |a| is the area <strong>of</strong> the current loop, the vector a points in the direction normal to thesurface |a|. For circul<strong>at</strong>ing electron, this moment isWe can then write|µ d | = − ev2πr πr2 = − evr2µ d = − er × v2(9.5.5)= − eL2m 0(9.5.6)


114 Quantum Mechanics Made SimpleThe interaction energy <strong>of</strong> a magnetic dipole and a background magnetic field B is given byAssume th<strong>at</strong> B = ẑB z , thenE u =E u = −µ d · B (9.5.7)e B z L z =e B z m (9.5.8)2m 0 2m 0since the eigenvalue <strong>of</strong> ˆL z is m and the expect<strong>at</strong>ion value <strong>of</strong> ˆL z can be precisely m. Equ<strong>at</strong>ion(9.5.8) can also be written aswhere µ B is a Bohr magneton defined to beE u = mµ B B z (9.5.9)µ B = e2m 0(9.5.10)The Hamiltonian <strong>of</strong> quantum mechanics is inspired by the classical Hamiltonian mechanics.The Hamiltonian represents the total energy <strong>of</strong> the system. If an applied external magneticfield contribution to a small change in the total energy, we can add a perturb<strong>at</strong>ion Hamiltonianin accordance to (9.5.8) and letĤ p = eB z2m 0ˆLz (9.5.11)Originally, for angular momentum with number l, the z component <strong>of</strong> the angular momentumis m with (2l + 1) values for m th<strong>at</strong> are degener<strong>at</strong>e. The applied B field will split thisdegeneracy into 2l + 1 distinct energy values. This is known as the Zeeman effect.Equ<strong>at</strong>ion (9.5.11) can be written asĤ p =e ˆL · B (9.5.12)2m 0When the spin angular momentum is added, the perturbing Hamiltonian becomesĤ p =e (ˆL + gŜ) · B (9.5.13)2m 0where g is the gyromagnetic factor found to be approxim<strong>at</strong>ely 2 experimentally.In general, the Hamiltonian for an electron in the presence <strong>of</strong> an electromagnetic field isĤ p = 1 (ˆp − eA) 2 + V +e (ˆL + gŜ) · B (9.5.14)2m 0 2m 0The above Hamiltonian is again motiv<strong>at</strong>ed by the Hamiltonian for classical mechanics for anelectron in the presence <strong>of</strong> an electromagnetic field, which isĤ p = 12m 0[p − eA(r)] 2 + V (r) (9.5.15)One can show, using the classical equ<strong>at</strong>ion <strong>of</strong> motion for Hamiltonian mechanics, th<strong>at</strong> Lorentzforce law emerges from the above. Equ<strong>at</strong>ion (9.5.14) is motiv<strong>at</strong>ed by (9.5.15) where themomentum p in (9.5.15) is elev<strong>at</strong>ed to be an oper<strong>at</strong>or ˆp in (9.5.14). The vector potential A(r),however remains as a classical variable. As a result, (9.5.14) is a semi-classical Hamiltonian.


Spin 1159.5.1 Splitting <strong>of</strong> Degener<strong>at</strong>e Energy LevelThe interaction <strong>of</strong> spins with magnetic field can cause the degener<strong>at</strong>e energy levels <strong>of</strong> aquantum system to split. If no magnetic field is present, the spin up st<strong>at</strong>e and the spindown st<strong>at</strong>e can share the same wave function with the same energy level, and hence, they aredegener<strong>at</strong>e. In the presence <strong>of</strong> a magnetic field, the st<strong>at</strong>e where the spin is parallel with themagnetic field is <strong>of</strong> lower energy compared to the case where the spin is antiparallel to themagnetic field. This causes the splitting <strong>of</strong> a degener<strong>at</strong> energy level as shown in Figure 9.2.This fact can be used to design interesting spin devices.EBE downE upFigure 9.2: The splitting <strong>of</strong> degener<strong>at</strong>e energy level when the spin <strong>of</strong> the electron interactswith an ambient magnetic field.9.6 SpintronicsSome m<strong>at</strong>erials are highly magnetic because there are many unpaired electrons in these m<strong>at</strong>erials.These unpaired electrons give these m<strong>at</strong>erials magnetic dipole moments, such asferromagnetic m<strong>at</strong>erials. In ferromagnets, the magnetic dipoles <strong>of</strong> the same orient<strong>at</strong>ion clustertogether to form microscopic domains th<strong>at</strong> are random on a macroscopic scale. Thesedomains can be aligned macroscopically by magnetiz<strong>at</strong>ion, making these m<strong>at</strong>erials into magnets.Fe, Co, Ni and their alloys have this property.The conduction property <strong>of</strong> these m<strong>at</strong>erials can also be affected by remnant magneticfield in the domain, or externally applied magnetic field. This can be explained by the DOSdiagram. The electrons in a magnetic m<strong>at</strong>erial see a self field th<strong>at</strong> affects their energy levels:electrons th<strong>at</strong> are aligned with the magnetic field are in the lower energy st<strong>at</strong>es comparedto the anti-parallel electrons. Hence the DOS <strong>of</strong> a ferromagnets is different for up and downspins compared to normal metals.Now, the conduction electrons available for up spins do not exist, but occur in abundancefor the down spins. (The skewness in DOS can be further affected by an externally appliedmagnetic field.) This gives rise to conductivity th<strong>at</strong> is spin dependent, or spin polarizedtransport. Also, electrons th<strong>at</strong> pass through a ferromagnets become spin polarized. Devices


116 Quantum Mechanics Made SimpleEEEE fDOSE fDOSE fDOSDOSNormalDOSFerromagneticDOSFerromagneticFigure 9.3: The shifting <strong>of</strong> the energy levels <strong>of</strong> the up and down spins st<strong>at</strong>es in the DOS plotdue to different ambient or external magnetic field. The left half is for DOS <strong>of</strong> down spinsand conversely for the right half.<strong>of</strong> high resistance and low resistance can be made by sandwiching three layers together <strong>of</strong> anormal metal between two ferromagnets layers as shown in Figure 9.4. Such phenomenon isknown as giant magneto-resistance (GMR).Moreover, the external magnetic field can alter the resistance <strong>of</strong> the device using it tosense magnetic field or store inform<strong>at</strong>ion in magnetic field. Hence, such devices can be usedas magnetic read head or non-vol<strong>at</strong>ile memory as shown in Figure 9.5.Exercise 2In the Appendix, we have derived the commut<strong>at</strong>ion rel<strong>at</strong>ion for general angular momentumin equ<strong>at</strong>ion (B.3.9). Use the same technique, derive the other commut<strong>at</strong>ion rel<strong>at</strong>ions forangular momentum in (B.3.10).1. The raising and lowering oper<strong>at</strong>ors in angular momentum are defined asJ ± = J x ±iJ yUsing the commut<strong>at</strong>ion rel<strong>at</strong>ions in (B.3.10), show th<strong>at</strong>J ∓ J ± = J 2 x + J 2 y ±i[J x , J y ]J 2 = J 2 x + J 2 y + J 2 z = J ∓ J ± + J 2 z ±J z[J ± , J z ] = ∓J ±[J 2 , J z ] = 0[J 2 , J ± ] = 0The above means th<strong>at</strong> J ± shares the same eigenfunctions with J 2 but not with J z .However, J 2 shares the same set <strong>of</strong> eigenfunctions with J z .


Spin 1172. Denoting the eigenfunctions <strong>of</strong> J 2 and J z aswhereand J and M are integers. Show th<strong>at</strong>3. By using the fact th<strong>at</strong>show th<strong>at</strong>|J, M〉J 2 |J, M〉 = J (J + 1) 2 |J, M〉J z |J, M〉 = M|J, M〉J z J ± |J, M〉 = (M ± 1) J ± |J, M〉〈J, M|J ∓ J ± |J, M〉 ≥ 0−J≤M≤JThe wonder <strong>of</strong> the above is th<strong>at</strong> it could be derived from commut<strong>at</strong>ion rel<strong>at</strong>ions alone,and the explicit form <strong>of</strong> |J, M〉 need not be known.


118 Quantum Mechanics Made SimpleFigure 9.4: The sandwiching <strong>of</strong> a normal metal between two ferromagnets can give rise tohigh or low resistance depending on their rel<strong>at</strong>ive orient<strong>at</strong>ions (from G.A. Prinz, Science, v.282, 1660, 1998).


Spin 119Figure 9.5: When a GMR read head passes over magnetic domains <strong>of</strong> different polariz<strong>at</strong>ion ina storage media, the leakage field in between the domains can be detected, as the resistance<strong>of</strong> the GMR is affected by such leakage field (Gary A. Prinz, Science v. 282, 1660, 1998).


120 Quantum Mechanics Made Simple


Chapter 10Identical Particles10.1 IntroductionThe physics <strong>of</strong> identical particles is another intellectual triumph in quantum mechanics wherethe phenomenon is not observed in the classical world. When particles are identical or indistinguishable,their wavefunctions have to assume a certain form to s<strong>at</strong>isfy a certain symmetrywhen the positions <strong>of</strong> the two particles are swapped. In quantum mechanics, it is themagnitude squared <strong>of</strong> a wavefunction th<strong>at</strong> has physical meaning. If the particles are indistinguishable,the magnitude squared <strong>of</strong> a wavefunction does not change when the two particlesare swapped.The many-particle problem cannot be solved in closed form, but we can use a vari<strong>at</strong>ionalmethod such as the Rayleigh-Ritz method to find an approxim<strong>at</strong>e eigenvalue <strong>of</strong> the systemas described in Section 6.4.1. To this end, we have to construct basis functions for identicalparticles. These basis functions can be constructed from the single-particle solutions whenthe particles are non-interacting with and isol<strong>at</strong>ed from each other.First, we can consider the particles to be non-interacting and isol<strong>at</strong>ed. If there are twonon-interacting particles, we can write down two independent equ<strong>at</strong>ions for them, namely]Ĥ 1 ψ a (r 1 ) =[− 22m ∇2 1 + V (r 1 ) ψ a (r 1 ) = E a ψ a (r 1 ) (10.1.1)Ĥ 2 ψ b (r 2 ) =][− 22m ∇2 2 + V (r 2 ) ψ b (r 2 ) = E b ψ b (r 2 ) (10.1.2)In the above, the two particles share the same potential function V , but are in differentworlds. Altern<strong>at</strong>ively, we can write down the equ<strong>at</strong>ion for the entire quantum system whichis the sum <strong>of</strong> the two quantum systems, namely(Ĥ1 + Ĥ2)ψ a (r 1 )ψ b (r 2 ) = (E a + E b )ψ a (r 1 )ψ b (r 2 ) (10.1.3)We use a direct product space consisting <strong>of</strong> ψ a (r 1 )ψ b (r 2 ) to expand the eigenfunction <strong>of</strong> thesystem. This is alright if the particles are non-interacting, but if they are interacting, the121


122 Quantum Mechanics Made Simpleissue <strong>of</strong> identical particles and their indistinguishability emerges. The wave function has tobe constructed with caution.In general, we can write the general wavefunction for two particles asψ tp (r 1 , r 2 ) = c 12 ψ a (r 1 )ψ b (r 2 ) + c 21 ψ a (r 2 )ψ b (r 1 ) (10.1.4)where ψ η (r i ) is a one-particle wavefunction for a particle in st<strong>at</strong>e η described by coordin<strong>at</strong>er i .But when r 1 ⇋ r 2 , the assumption is th<strong>at</strong> the magnitude squared <strong>of</strong> the wavefunctionremains unchanged, since these particles are indistinguishable. Namely,which means th<strong>at</strong>where γ = e iθ . Using (10.1.4) in (10.1.6), we have|ψ tp (r 1 , r 2 )| 2 = |ψ tp (r 2 , r 1 )| 2 (10.1.5)ψ tp (r 1 , r 2 ) = γψ tp (r 2 , r 1 ) (10.1.6)c 12 ψ a (r 1 )ψ b (r 2 ) + c 21 ψ a (r 2 )ψ b (r 1 ) = γ [c 12 ψ a (r 2 )ψ b (r 1 ) + c 21 ψ a (r 1 )ψ b (r 2 )] (10.1.7)From the above, we getConsequently,or γ 2 = 1, γ = ±1, c 21 = ±c 12 , andc 21 = γc 12 , c 12 = γc 21 (10.1.8)c 12 = γc 21 = γ 2 c 12 (10.1.9)ψ tp (r 1 , r 2 ) = c 12 [ψ a (r 1 )ψ b (r 2 ) ± ψ a (r 2 )ψ b (r 1 )] (10.1.10)When the “+” sign is chosen in (10.1.10), it corresponds to the wavefunction <strong>of</strong> two identicalbosons. When the “−” sign is chosen, it corresponds to the wavefunction <strong>of</strong> two identicalfermions.10.2 Pauli Exclusion PrincipleFor fermions, the two-particle wavefunction is given byψ tp (r 1 , r 2 ) = c [ψ a (r 1 )ψ b (r 2 ) − ψ a (r 2 )ψ b (r 1 )] (10.2.1)when a = b, the above vanishes, meaning th<strong>at</strong> two fermions cannot be in the same st<strong>at</strong>e.The above vanishes when r 1 = r 2 , meaning th<strong>at</strong> two fermions cannot be in the same positionsimultaneously. This known as the Pauli exclusion principle. No two identical electrons canbe in the same st<strong>at</strong>e nor the same position simultaneously. However, two electrons withdifferent spins, up spin and down spin, are considered different, and they can be in the samest<strong>at</strong>e, like in the same orbital in an <strong>at</strong>om. Also, the sign <strong>of</strong> the total wavefunction, ψ tp (r 1 , r 2 ),changes sign when the positions <strong>of</strong> the two particles are swapped.


Identical Particles 1<strong>23</strong>10.3 Exchange EnergyWhen we have two electrons, each will see the Coulomb potential <strong>of</strong> the other electron. Hence,the Hamiltonian is given byĤ = − 2 (∇22m 1 + ∇ 2 ) e 22 + V 1 + V 2 +e 4πɛ 0 |r 1 − r 2 | = e 2Ĥ1 + Ĥ2 +4πɛ 0 |r 1 − r 2 |(10.3.1)where V 1 = V (r 1 ), V 2 = V (r 2 ) and ∇ 2 i is the Laplacian oper<strong>at</strong>or expressed in the i-thcoordin<strong>at</strong>e. The two particles are in the same environment and their wavefunctions canoverlap. The two-particle fermion wavefunction can be written as|ψ tp 〉 = 1 √2[|1, a〉 |2, b〉 − |2, a〉 |1, b〉] (10.3.2)where1 √2 ensures th<strong>at</strong> |ψ tp 〉 is also normalized. In the above, |i, η〉 is the Hilbert spacerepresent<strong>at</strong>ion <strong>of</strong> the wavefunction ψ η (r i ). The above does not solve the Schrödinger equ<strong>at</strong>ioncorresponding to the above Hamiltonian, but it can be used to estim<strong>at</strong>e the eigenvalue or theenergy <strong>of</strong> the Hamiltonian in the spirit <strong>of</strong> the vari<strong>at</strong>ional method described previously, usingthe Rayleigh quotient.To find the corresponding expect<strong>at</strong>ion value <strong>of</strong> the energy, we evalu<strong>at</strong>e〈E〉 = 〈ψ tp | Ĥ |ψ tp〉 (10.3.3)Since the function is normalized, the above is also the Rayleigh quotient for the energy. Whenexpanded to four terms, gives〈E〉 = 1 [〈1, a| 〈2, b| Ĥ |1, a〉 |2, b〉 + 〈2, a| 〈1, b| Ĥ |2, a〉 |1, b〉2]− 〈1, a| 〈2, b| Ĥ |2, a〉 |1, b〉 − 〈2, a| 〈1, b| Ĥ |1, a〉 |2, b〉 (10.3.4)The first two terms above are equal to each other. Specifically,〈1, a| 〈2, b| Ĥ |1, a〉 |2, b〉= 〈1, a| 〈2, b| Ĥ1 |1, a〉 |2, b〉 + 〈1, a| 〈2, b| Ĥ2 |1, a〉 |2, b〉+ 〈1, a| 〈2, b|e 2|1, a〉 |2, b〉4πɛ 0 |r 1 − r 2 |= 〈1, a| Ĥ1 |1, a〉 + 〈2, b| Ĥ2 |2, b〉+ 〈1, a| 〈2, b|e 2|1, a〉 |2, b〉4πɛ 0 |r 1 − r 2 |= E a + E b + E IE,ab (10.3.5)E a is the energy <strong>of</strong> particle 1 in st<strong>at</strong>e a, while E b is the energy <strong>of</strong> particle 2 in st<strong>at</strong>e b. TheE IE,ab term is the interaction energy th<strong>at</strong> can be derived explicitly to yieldE IE,ab = e 2 ∫drdr ′′ |ψ a (r)| 2 |ψ b (r ′ )| 24πɛ 0 |r − r ′ |(10.3.6)


124 Quantum Mechanics Made SimpleThis is exactly the same as the potential energy from the second term on the right-hand side<strong>of</strong> (10.3.4) if we were to write it explicitly. Consequently, we haveE a + E b + E IE,ab = 1 ][〈1, a| 〈2, b|2Ĥ |1, a〉 |2, b〉 + 〈2, a| 〈1, b| Ĥ |2, a〉 |1, b〉 (10.3.7)However, the third and fourth terms on the right-hand side <strong>of</strong> (10.3.4) have no classical analog.They are there because <strong>of</strong> the presence <strong>of</strong> the second term in (10.1.10) be it for bosons orfermions. The second term is needed because the particles are indistinguishable or identicalif we exchange their positions.Due to the Hermiticity <strong>of</strong> the Hamiltonian oper<strong>at</strong>or, the fourth term on the right-handside <strong>of</strong> (10.3.4) is the complex conjug<strong>at</strong>e <strong>of</strong> the third term. Consequently, these two termscan be written as) ∗ ][〈1, a| 〈2, b|(〈1, Ĥ |2, a〉 |1, b〉 + a| 〈2, b| Ĥ |2, a〉 |1, b〉E EX,ab = − 1 2[∫= −Re]drdr ′ ψa ∗ (r) ψb ∗ (r ′ ) Ĥψ a (r ′ ) ψ b (r)(10.3.8)(10.3.9)The exchange energy is proportional to the overlap between the wavefunctions ψ a (r) andψ b (r). This term is non-zero even if the Coulomb interaction is absent, but the wavefunctionsoverlap and are non-orthogonal. When their wavefunctions overlap, the issue <strong>of</strong>indistinguishability arises, and the exchange term can be non-zero even for non-interactingparticles.10.4 Extension to More Than Two ParticlesIn the following, we will discuss methods to construct basis functions for different kinds <strong>of</strong>particles. These basis functions does not solve the Schrödinger equ<strong>at</strong>ion yet, but they haveto s<strong>at</strong>isfy certain symmetry conditions depending on the kind <strong>of</strong> particles they represent.The basis functions can be used in the Rayleigh-Ritz procedure to estim<strong>at</strong>e the energy <strong>of</strong> thesystem. We can use the eigenst<strong>at</strong>es <strong>of</strong> the isol<strong>at</strong>ed particles to construct basis functions.1. Non-identical Particle Case:Let us assume th<strong>at</strong> we have N particles, and M modes to fit this N particles. We canconstruct a st<strong>at</strong>e for non-identical particles th<strong>at</strong> looks like|ψ diff 〉 = |1, a〉 |2, b〉 |3, c〉 ... |N, n〉 (10.4.1)In terms <strong>of</strong> basis function, we may express the above as|ψ ab···n 〉 = |1, a〉|2, b〉|3, c〉 · · · |N, n〉 (10.4.2)orψ ab···n (r 1 , r 2 , · · · , r N ) = ψ a (r 1 )ψ b (r 2 ) · · · ψ n (r N ) (10.4.3)We can fit the N particles in M modes, and these M modes can be repe<strong>at</strong>ing or nonrepe<strong>at</strong>ing.In other words, a, b, c, . . . , n are chosen from M eigenst<strong>at</strong>es with energy


Identical Particles 125values E 1 , E 2 , E 3 , . . . , E M with possibility <strong>of</strong> repetition. For non-repe<strong>at</strong>ing case, it isnecessary for M N.However, the above wavefunction cannot be used for bosons and fermions, as we willget a new wavefunction when we swap the positions <strong>of</strong> two particles. But bosons andfermions are indistinguishable particles. We will consider them separ<strong>at</strong>ely.2. Boson Case:For the N boson particle case, we can write the legitim<strong>at</strong>e wavefunction, which can beused as a basis function, as|ψ identical-bosons 〉 ∝ ∑ˆPˆP |1, a〉|2, b〉|3, c〉 · · · |N, n〉 (10.4.4)where ˆP is a permut<strong>at</strong>ion oper<strong>at</strong>or, and the above summ<strong>at</strong>ion is over all possible permut<strong>at</strong>ions<strong>of</strong> the coordin<strong>at</strong>e r i over the one-particle eigenst<strong>at</strong>es a, b, c, · · · , n. Repetition<strong>of</strong> these energy eigenst<strong>at</strong>es is allowed since more than one particle can be placed inone energy st<strong>at</strong>e. The above wavefunction remains unchange when we permute thepositions <strong>of</strong> two particles, because for every |1, a〉 · · · |i, l〉 · · · |j, p〉 · · · |N, n〉, there is a|1, a〉 · · · |j, l〉 · · · |i, p〉 · · · |N, n〉 in the above summ<strong>at</strong>ion. Hence, swapping <strong>of</strong> i and j willnot change the sign <strong>of</strong> the above wavefunction. The above can also be written as a basisfunction as|ψ ab···n 〉 ∝ ∑ˆPˆP |1, a〉|2, b〉|3, c〉 · · · |N, n〉 (10.4.5)3. Fermion Case:For the N fermion case, we can write the wavefunction, which can be used as a basisfunction, as|ψ identical-fermion 〉 = √ 1 ∑N!N!ˆP =1± ˆP |1, a〉|2, b〉|3, c〉 · · · |N, n〉 (10.4.6)where the “+” sign is chosen for even permut<strong>at</strong>ion while the “−” sign is chosen for oddpermut<strong>at</strong>ion. A permut<strong>at</strong>ion involves a unique pairwise exchange <strong>of</strong> two particles . Thepermut<strong>at</strong>ion is even or odd depending on the number <strong>of</strong> pairwise exchanges th<strong>at</strong> havetaken place.Therefore, given a term |1, a〉 · · · |i, l〉 · · · |j, p〉 · · · |N, n〉, there always exists another term:−|1, a〉 · · · |j, l〉 · · · |i, p〉 · · · |N, n〉 in the above summ<strong>at</strong>ion since they differ by one permut<strong>at</strong>ion.If i = j, implying th<strong>at</strong> r i = r j , the two terms cancel each other implyingth<strong>at</strong> they cannot be in the same position. Likewise all the terms in the sum cancel eachother since every term th<strong>at</strong> contains i and j can be paired up with every other terms inthe sum. Moreover, If l = p, all terms in the summ<strong>at</strong>ion above cancel as well implyingth<strong>at</strong> they cannot be in the same mode or st<strong>at</strong>e. Therefore, the above is a legitim<strong>at</strong>ebasis function th<strong>at</strong> represents the fermions as it obeys Pauli exclusion principle. Also,there is a sign change when the position <strong>of</strong> two particles are swapped.


126 Quantum Mechanics Made SimpleAnother way <strong>of</strong> writing the fermion case is∣ ∣∣∣∣∣∣∣∣ |1, a〉 |2, a〉 · · · |N, a〉|ψ identical-fermion 〉 = √ 1 |1, b〉 |2, b〉 · · · |N, b〉). N! . . .. .|1, n〉 |2, n〉 · · · |N, n〉 ∣(10.4.7)The above is known as the Sl<strong>at</strong>er determinant. When two particles are in the sameposition, two columns are the same, and the above determinant is zero. When two st<strong>at</strong>esare the same, two rows are the same, and the determinant <strong>of</strong> the above is again zero,implying th<strong>at</strong> these two cases are not allowed. Also when two columns are swapped,the sign <strong>of</strong> the determinant changes, because it corresponds to two particles exchangingpositions. In terms <strong>of</strong> basis function, we can express the above as|ψ ab···n 〉 = √ 1 ∑N!N!ˆP =1± ˆP |1, a〉|2, b〉|3, c〉 · · · |N, n〉∣ ∣∣∣∣∣∣∣∣ |1, a〉 |2, a〉 · · · |N, a〉= √ 1 |1, b〉 |2, b〉 · · · |N, b〉). N! . . .. .|1, n〉 |2, n〉 · · · |N, n〉 ∣(10.4.8)10.5 Counting the Number <strong>of</strong> Basis St<strong>at</strong>esUsually, the one particle eigenfucntion has infinitely many possible st<strong>at</strong>es. In the finite basismethod, we may want to choose a subset <strong>of</strong> this eigenst<strong>at</strong>es, say M st<strong>at</strong>es. At this point, ifwe have N particles and M st<strong>at</strong>es to put this N particles in, it may be prudent to count howmany possible basis st<strong>at</strong>es there are. Such counting scheme will also be used in the appendixto derive the thermal distribution functions.1. Non-identical Particles:The first particle has M st<strong>at</strong>e to choose from in (10.4.2), the second particle also has Mst<strong>at</strong>es to choose from, and eventually, there are M N st<strong>at</strong>es possible for (10.4.2), sincerepetition <strong>of</strong> the st<strong>at</strong>es is allowed.2. Identical Bosons:N basis = M N (10.5.1)In this case, the number <strong>of</strong> particles th<strong>at</strong> can be put into one st<strong>at</strong>e is not limited. Sincewe need to permute the st<strong>at</strong>e (10.4.2) to arrive <strong>at</strong> new st<strong>at</strong>es to be linear superposed inthe manner <strong>of</strong> (10.4.5), the ordering <strong>of</strong> the particles within the st<strong>at</strong>es is not important.We will first derive the number <strong>of</strong> permut<strong>at</strong>ions when ordering is important. One canimaging the M st<strong>at</strong>es to be equivalent to M bins. The first particle has M possiblebins to fit into. However, once the particle is put into one <strong>of</strong> the bins, since ordering


Identical Particles 127is important, repetition is allowed, there are two ways to put in the second particle inthe same bin, before or after the first particle. In other words, the first particle splitsthe bin into two compartments, allowing two ways to put two particles in the same bin.Together with the other M − 1 bins, there are M + 1 ways to insert the second particleinto the bins. In other words, adding a new particle always adds a new compartmentfor the next particle to fit in. By the same token, the third particle has M + 2 ways t<strong>of</strong>it into the bins etc. Finally the number <strong>of</strong> possible ways isM(M + 1)(M + 2)...(M + N − 1) (10.5.2)One can verify th<strong>at</strong> the above gives the correct answer <strong>of</strong> N! if only one bin is availablefor N particles. Since ordering is unimportant, we have3. Identical Fermions:N basis =(M + N − 1)!N!(M − 1)!(10.5.3)For fermions, the first particle has M st<strong>at</strong>es to choose from. Since each st<strong>at</strong>e can onlyadmit one fermion particle, it is necessary th<strong>at</strong> M ≥ N. The second particle has M − 1st<strong>at</strong>es to choose from, since repetition is not allowed. Also, ordering is not importantsince all permut<strong>at</strong>ions are used in the summ<strong>at</strong>ion (10.4.8). Consequently,10.6 Examples1. Non-identical Particles:N basis =M!N!(M − N)!Say if we have two electrons with different spins, the distinct st<strong>at</strong>es are(10.5.4)|1, a〉|2, a〉, |1, b〉|2, b〉, |1, a〉|2, b〉, |1, b〉|2, a〉 (10.6.1)The above is in agreement with M = 2, N = 2, M N = 2 2 = 4.2. Identical Bosons:Consider the 4 He (helium four) <strong>at</strong>oms which are bosons. Then the possible boson st<strong>at</strong>esare1|1, a〉|2, a〉, |1, b〉|2, b〉, √ (|1, a〉|2, b〉 + |1, b〉|2, a〉) (10.6.2)2Again this is in agreement with M = 2, N = 2, (M+N−1)!N!(M−1)!= 3!2!1! = 3.3. Identical Fermions:If we have two identical fermions, the only st<strong>at</strong>e is1√2(|1, a〉|2, b〉 − |1, b〉|2, a〉) (10.6.3)Again this is in agreement with M = 2, N = 2,M!N!(M−N)! = 2!2!0! = 1.


128 Quantum Mechanics Made SimpleIn the above simple examples, we see th<strong>at</strong> for the boson case, there are more st<strong>at</strong>es withparticles in the same st<strong>at</strong>e compared to the other cases. In the boson case, there are two out<strong>of</strong> three st<strong>at</strong>es where the particles are in the same st<strong>at</strong>e (see (10.6.2)). In the non-identicalparticle case, there are two out <strong>of</strong> four such st<strong>at</strong>es. In the fermion case, there is none. Thisis indic<strong>at</strong>ive <strong>of</strong> the fact th<strong>at</strong> bosons like to cluster together.10.7 Thermal Distribution FunctionsThere are some important thermal distribution functions in understanding semiconductorphysics. These are the Maxwell-Boltzmann distribution, the Fermi-Dirac distribution, and theBose-Einstein distribution. These distribution functions are derived via st<strong>at</strong>istical mechanics,and the results will just quoted here. The first-principles deriv<strong>at</strong>ions are given in AppendixC.1. Maxwell-Boltzmann Distribution:For this distribution, the number <strong>of</strong> particles N(E) in a given energy st<strong>at</strong>e is given byN(E) = e −(E−µ)k B T(10.7.1)where µ is the chemical potential, k B is the Boltzmann constant, and T is temper<strong>at</strong>urein Kelvin.2. Fermi-Dirac Distribution:This is given byN(E) =11 + e E−µk B T(10.7.2)This is the most important distribution function for semiconductors. The chemicalpotential µ is loosely called the Fermi level. To be precise, µ is the same as the Fermilevel <strong>at</strong> absolute zero temper<strong>at</strong>ure. When T is zero, the distribution function looks likeas shown in Figure 10.1. The electrons are frozen into the ground st<strong>at</strong>e for E < µ.When T > 0, some <strong>of</strong> the electrons from the st<strong>at</strong>es where E < µ are dislodged bythermal agit<strong>at</strong>ion into st<strong>at</strong>es where E > µ. This phenomenon is more pronounced as Tincreases. This distribution also explains the physical character <strong>of</strong> semiconductors.In semiconductors, the Fermi level is midway in between the valence band and theconduction band. When T = 0, all the electrons are frozen in the valence band, and thesemiconductor cannot conduct electricity, as there are no free electrons in the conductionband. When T > 0, some electrons in the valence band are dislodged into the conductionband. This gives rise to electrons in the conduction band, and holes in the valence band.They contribute to the flow <strong>of</strong> electric current and the semiconductor starts to conduct.The conductivity <strong>of</strong> the semi-conductor m<strong>at</strong>erial increases with increasing temper<strong>at</strong>ure.3. Bose-Einstein Distribution:


Identical Particles 129Figure 10.1: Fermi-Dirac Distribution for changing temper<strong>at</strong>ure T .The Bose-Einstein distribution is for bosons, and it is given byN(E) =1e (E−µ)/(k BT )− 1(10.7.3)This distribution has a divergence when E = µ. It also reflects the fact th<strong>at</strong> bosonslike to cluster together. When the temper<strong>at</strong>ure is low, they condense to around thechemical potential µ.


130 Quantum Mechanics Made SimpleFigure 10.2: Comparison <strong>of</strong> different thermal distribution functions. For high energy st<strong>at</strong>es,they are similar to each other (from DAB Miller).


Chapter 11Density M<strong>at</strong>rix11.1 Pure and Mixed St<strong>at</strong>esGiven a quantum st<strong>at</strong>e described by|ψ〉 = a 1 |ψ 1 〉 + a 2 |ψ 2 〉 (11.1.1)which is a linear superposition <strong>of</strong> two st<strong>at</strong>es, the interpret<strong>at</strong>ion <strong>of</strong> quantum mechanics is th<strong>at</strong>the particle is in a linear superposition <strong>of</strong> the two st<strong>at</strong>es, and one does not know wh<strong>at</strong> st<strong>at</strong>ethe particle is in until the measurement is done. The measurement collapses the particle intoone <strong>of</strong> the two quantum st<strong>at</strong>es. For instance, the two st<strong>at</strong>es could be the spin st<strong>at</strong>es <strong>of</strong> an<strong>at</strong>om. The famous Stern-Gerlach experiment separ<strong>at</strong>es the spin st<strong>at</strong>es into the up st<strong>at</strong>e andthe down st<strong>at</strong>e. Such a st<strong>at</strong>e indic<strong>at</strong>ed in (11.1.1) is known as a pure quantum st<strong>at</strong>e. Thephase rel<strong>at</strong>ionship between a 1 and a 2 is maintained precisely. When the quantum st<strong>at</strong>e is insuch a st<strong>at</strong>e, we say th<strong>at</strong> the two st<strong>at</strong>es are coherent.The above interpret<strong>at</strong>ion <strong>of</strong> quantum mechanics is known as the Copenhagen school <strong>of</strong>thought, lead by Niels Bohr. The probabilistic interpret<strong>at</strong>ion <strong>of</strong> quantum mechanics wasfraught with controversies in its early days. Many physicists, including Einstein, cannot bringthemselves to term with it. “God does not play dice!” was one famous saying <strong>of</strong> Einstein. Heposited th<strong>at</strong> “a quantum system is already in a known st<strong>at</strong>e before the measurement” withthe hidden variable description. However, experimental findings are with the Copenhagenschool for coherent quantum st<strong>at</strong>es. For incoherent quantum st<strong>at</strong>es, they are with Einstein’sposition.Nevertheless, it is difficult for a particle to be in a pure quantum st<strong>at</strong>e if it involves a linearsuperposition <strong>of</strong> many quantum st<strong>at</strong>es unless the quantum system is completely isol<strong>at</strong>ed.Coupling to the thermal b<strong>at</strong>h can destroy the coherence between these st<strong>at</strong>es. Hence, thereare st<strong>at</strong>es for which partial coherence still exists, and the linear superposition <strong>of</strong> quantumst<strong>at</strong>es de-coheres with respect to each other, before the linear superposition <strong>of</strong> st<strong>at</strong>es becomesentirely a mixed st<strong>at</strong>e. Coherence lifetime is used to measure when a linear superposition <strong>of</strong>quantum st<strong>at</strong>es, which starts out being a coherent sum, becomes an incoherent superposition.However, there are st<strong>at</strong>es where the particle has already collapsed into st<strong>at</strong>e 1 or st<strong>at</strong>e 2even before the measurement. Or these two st<strong>at</strong>es are entirely uncorrel<strong>at</strong>ed or incoherent.131


132 Quantum Mechanics Made SimpleThe particle is either in st<strong>at</strong>e 1 or st<strong>at</strong>e 2. The probability <strong>of</strong> finding the particle in st<strong>at</strong>e 1 is|a 1 | 2 while th<strong>at</strong> for st<strong>at</strong>e 2 is |a 2 | 2 . The phase rel<strong>at</strong>ionship between a 1 and a 2 is completelyrandom or incoherent. St<strong>at</strong>es for which there is no coherence between a 1 and a 2 are knownas mixed st<strong>at</strong>es. If we take an ensemble average <strong>of</strong> the measurement outcomes <strong>of</strong> the mixedst<strong>at</strong>es, they s<strong>at</strong>isfy the aforementioned probability property.The probabilistic interpret<strong>at</strong>ion <strong>of</strong> quantum mechanics cannot be verified by a singleexperiment, but an ensemble <strong>of</strong> measurements whose setups are identical to each other. Anensemble average 1 <strong>of</strong> an experimental outcome is the average over these outcomes over theidentical ensemble <strong>of</strong> measurements. It will be prudent to find a way to describe a quantumsystem th<strong>at</strong> is rel<strong>at</strong>ed to its ensemble average and its experimental outcome.11.2 Density Oper<strong>at</strong>orAn elegant way to represent a quantum st<strong>at</strong>e th<strong>at</strong> is either in the pure st<strong>at</strong>e, mixed st<strong>at</strong>e, orpartial coherent st<strong>at</strong>e, is via the density oper<strong>at</strong>orˆρ = |ψ〉〈ψ| (11.2.1)We can easily show th<strong>at</strong> the expect<strong>at</strong>ion <strong>of</strong> the oper<strong>at</strong>or  in this st<strong>at</strong>e is) (ˆρ 〈Â〉 = 〈ψ|Â|ψ〉 = tr(11.2.2)This has been shown in Section 5.2.3. Therefore, knowing the density oper<strong>at</strong>or is equivalentto knowing the quantum st<strong>at</strong>e <strong>of</strong> a system which is denoted by st<strong>at</strong>e |ψ〉. We can calcul<strong>at</strong>e thetrace <strong>of</strong> an oper<strong>at</strong>or in terms <strong>of</strong> the sum <strong>of</strong> the diagonal elements <strong>of</strong> its m<strong>at</strong>rix represent<strong>at</strong>ion.Hence,)tr(ˆρ = ∑ 〈φ n |ˆρ|φ n〉 (11.2.3)nWe show th<strong>at</strong> the above is basis independent by inserting the identity oper<strong>at</strong>or Î = ∑ m |φ m〉〈φ m |twice in (11.2.2), to yield〈ψ|Â|ψ〉 = ∑ n,m〈ψ|φ n 〉〈φ n |Â|φ m〉〈φ m |ψ〉 (11.2.4)Since the factors in the summand are scalars, we can rearrange them to give〈ψ|Â|ψ〉 =∑〈φ m |ψ〉〈ψ|φ n 〉〈φ n |Â|φ m〉 = ∑ ρ mn A nm = tr ( ρ · A ) (11.2.5)n,mn,mwhere ρ mn is the m<strong>at</strong>rix represent<strong>at</strong>ion <strong>of</strong> ˆρ while A nm is the m<strong>at</strong>rix represent<strong>at</strong>ion <strong>of</strong> Â.The above is clearly independent <strong>of</strong> the orthonormal basis we choose.1 An ensemble average <strong>of</strong> a random variable is the average <strong>of</strong> the variable over its different outcomes forthe same system or identical systems.


Density M<strong>at</strong>rix 133Say if we start with (11.1.1) for a quantum st<strong>at</strong>e, using it in (11.2.1) will yieldˆρ = |a 1 | 2 |ψ 1 〉〈ψ 1 | + |a 2 | 2 |ψ 2 〉〈ψ 2 | + a 1 a ∗ 2|ψ 1 〉〈ψ 2 | + a 2 a ∗ 1|ψ 2 〉〈ψ 1 | (11.2.6)Notice th<strong>at</strong> only rel<strong>at</strong>ive phases between a 1 and a 2 are needed to form the above. Absolutephase has no meaning in quantum mechanics.For a mixed st<strong>at</strong>e, we assume th<strong>at</strong> the <strong>of</strong>f diagonal terms involving a ∗ 1a 2 and a ∗ 2a 1 willeither time average or ensemble average to zero since they are incoherent. 2 Hence, we supposeth<strong>at</strong>ˆρ = |a 1 | 2 |ψ 1 〉〈ψ 1 | + |a 2 | 2 |ψ 2 〉〈ψ 2 |= p 1 |ψ 1 〉〈ψ 1 | + p 2 |ψ 2 〉〈ψ 2 | (11.2.7)The form (11.2.7) can be thought <strong>of</strong> as having been ensemble averaged or time averaged. Fora general mixed st<strong>at</strong>e we can further write the density oper<strong>at</strong>or asˆρ = ∑ jp j |ψ j 〉〈ψ j | (11.2.8)where |ψ j 〉 are pure quantum st<strong>at</strong>es but not necessarily st<strong>at</strong>ionary st<strong>at</strong>es nor orthogonal, andp j is the probability <strong>of</strong> finding the quantum system in st<strong>at</strong>e |ψ j 〉. Hence, ∑ j p j = 1.When we find the expect<strong>at</strong>ion value <strong>of</strong> an oper<strong>at</strong>or  with a mixed st<strong>at</strong>e, it can be thought<strong>of</strong> as having been ensemble averaged. For example,〈Â〉 = tr(ˆρ) =∑jp j tr(Â|ψ j〉〈ψ j |)= ∑ jp j 〈ψ j |Â|ψ j〉 (11.2.9)The trace <strong>of</strong> the m<strong>at</strong>rix represent<strong>at</strong>ion <strong>of</strong> an oper<strong>at</strong>or is independent <strong>of</strong> the basis st<strong>at</strong>eschosen to represent the oper<strong>at</strong>or as explained in Chapter 5. It will be interesting to see howthe density oper<strong>at</strong>or changes under the change <strong>of</strong> basis st<strong>at</strong>e. To this end, we let|ψ j 〉 = ∑ nc (j)n (t)|φ n 〉 (11.2.10)Using this in (11.2.8), we haveˆρ = ∑ jp j∑n,m= ∑ |φ n 〉〈φ m | ∑n,mjc j n|φ n 〉〈φ m |(c (j)m ) ∗p j c (j)n (c (j)m ) ∗ (11.2.11)2 St<strong>at</strong>istical processes whose time average and ensemble average are equivalent to each other are known asergodic processes.


134 Quantum Mechanics Made SimpleWe can find the uv element <strong>of</strong> the m<strong>at</strong>rix represent<strong>at</strong>ion <strong>of</strong> ˆρ asρ uv = 〈φ u |ˆρ|φ v 〉 = ∑ jp j c (j)u [c (j)v ] ∗ = c u c ∗ v (11.2.12)where the overbar stands for ensemble average. From the above, it is quite clear th<strong>at</strong>ρ uv = ρ ∗ vu (11.2.13)or th<strong>at</strong> the density m<strong>at</strong>rix is Hermitian. The Hermiticity property <strong>of</strong> the density oper<strong>at</strong>or isobvious from (11.1.1) and (11.2.8). Furthermore,tr (ˆρ) = ∑ ρ uu = ∑ ∑p j |c (j)u | 2 = ∑ p j = 1 (11.2.14)uj ujThe second last equality follows from the normality <strong>of</strong> |ψ j 〉 in (11.2.10). Since a trace <strong>of</strong> anoper<strong>at</strong>or is independent <strong>of</strong> the basis st<strong>at</strong>e in which it is represented, the trace <strong>of</strong> the densityoper<strong>at</strong>or is defined astr(ˆρ) = ∑ u〈φ u |ˆρ|φ u 〉 (11.2.15)where φ u constitutes an orthonormal basis.For a pure quantum st<strong>at</strong>e,Therefore,ˆρ 2 = |ψ〉〈ψ|ψ〉〈ψ| = |ψ〉〈ψ| = ˆρ (11.2.16)tr(ˆρ 2 ) = tr(ˆρ) = 1 (11.2.17)The trace <strong>of</strong> the above is 1 is obvi<strong>at</strong>ed by setting  to 1 in Equ<strong>at</strong>ion (11.2.2). The aboveshows th<strong>at</strong> the trace <strong>of</strong> a density oper<strong>at</strong>or is always one irrespective <strong>of</strong> it is for a pure-st<strong>at</strong>eor a mixed-st<strong>at</strong>e quantum system. But this is not true for the trace <strong>of</strong> the ˆρ 2 oper<strong>at</strong>or.For a mixed st<strong>at</strong>eˆρ 2 = ∑ ∑p i p j |ψ i 〉〈ψ i |ψ j 〉〈ψ j | (11.2.18)i jTaking the trace <strong>of</strong> the above, we havetr(ˆρ 2 ) = ∑ n= ∑ i〈φ n |ˆρ 2 |φ n 〉 = ∑ n∑p i p j |〈ψ i |ψ j 〉| 2 ≤ ∑ji∑ ∑p i p j 〈φ n |ψ i 〉〈ψ i |ψ j 〉〈ψ j |φ n 〉ij(∑ ∑p i p j =jip i) 2= 1 (11.2.19)In the above, we did not assume th<strong>at</strong> |ψ j 〉 are orthogonal, but they are normalized. Theinequality follows from th<strong>at</strong> |〈ψ i |ψ j 〉| 2 ≤ 1. Hence, the trace <strong>of</strong> the density oper<strong>at</strong>or has theproperty th<strong>at</strong>tr(ˆρ 2 ) < 1, mixed st<strong>at</strong>e (11.2.20)tr(ˆρ 2 ) = 1, pure st<strong>at</strong>e (11.2.21)


Density M<strong>at</strong>rix 13511.3 Time Evolution <strong>of</strong> the M<strong>at</strong>rix Element <strong>of</strong> an Oper<strong>at</strong>orThe m<strong>at</strong>rix represent<strong>at</strong>ion <strong>of</strong> an oper<strong>at</strong>or is given byA mn (t) = 〈ψ m (t)|Â|ψ n(t)〉 (11.3.1)Taking the time deriv<strong>at</strong>ive <strong>of</strong> the above,we have∂ t A mn (t) = 〈∂ t ψ m (t)|Â|ψ n(t)〉 + 〈ψ m (t)|Â|∂ tψ n (t)〉 (11.3.2)assuming th<strong>at</strong>  is time independent, using the fact th<strong>at</strong>i∂ t |ψ m (t)〉 = i|∂ t ψ m (t)〉 = Ĥ|ψ m(t)〉 (11.3.3)we have∂ t A mn (t) = i ] 〉〈[Ĥψm (t) |Â|ψ n(t) − i 〈ψm(t)|Â|Ĥψ n(t)〉= i 〈ψ m(t)|ĤÂ|ψ n(t)〉 − i 〈ψ m(t)|ÂĤ|ψ n(t)〉= i 〈ψ m(t)|Ĥ − ÂĤ|ψ n(t)〉 (11.3.4)ori∂ t A mn (t) = (ÂĤ − ĤÂ) mn =[Â, Ĥ]mn(11.3.5)The above applies to expect<strong>at</strong>ion value <strong>of</strong>  as well, as shown in Chapter 5, where〈Â〉 = 〈ψ(t)|Â|ψ(t)〉 (11.3.6)and]〉i∂ t〈[Â, 〈Â〉 = 〈ÂĤ − ĤÂ〉 = Ĥ(11.3.7)In other words, if  commutes with Ĥ, its expect<strong>at</strong>ion value is time independent. In theabove deriv<strong>at</strong>ion, we have assumed th<strong>at</strong>  is independent <strong>of</strong> time.The time evolution <strong>of</strong> the density oper<strong>at</strong>or can be derived, similar to the time evolution<strong>of</strong> the expect<strong>at</strong>ion value <strong>of</strong> an oper<strong>at</strong>or. However, the density oper<strong>at</strong>or is in general timedependent, whereas oper<strong>at</strong>ors considered previously are time independent; and hence, thederiv<strong>at</strong>ion is slightly different. Giventhenˆρ = |ψ〉〈ψ| (11.3.8)∂ t ˆρ = |∂ t ψ〉〈ψ| + |ψ〉〈∂ t ψ| (11.3.9)


136 Quantum Mechanics Made SimpleUsing (11.3.3), we have∂ t ˆρ = − i Ĥ|ψ〉〈ψ| + i |ψ〉〈ψ|Ĥ = i ][ˆρĤ − Ĥ ˆρ = i [ ]ˆρ, Ĥ(11.3.10)There is a sign difference between (11.3.7) and (11.3.10). The above holds if ˆρ is in a mixedst<strong>at</strong>e as wellˆρ = ∑ jp j |ψ j 〉〈ψ j | (11.3.11)This follows from the linearity <strong>of</strong> the above equ<strong>at</strong>ions and th<strong>at</strong> p j is time-independent.Equ<strong>at</strong>ion (11.3.10) can be easily generalized to m<strong>at</strong>rix elements when the basis functionsused for seeking the m<strong>at</strong>rix represent<strong>at</strong>ion are from time-independent basis. Then,∂ t ρ mn = i ][ˆρĤ − Ĥ ˆρ = i [ ]ˆρ,mn Ĥ(11.3.12)mnwhere A mn = 〈ψ m |Â|ψ n〉, and ψ i is from a time-independent basis. By inserting identityoper<strong>at</strong>or defined using the same basis set, we can convert the above into wholly a m<strong>at</strong>rixform:∂ t ρ = i [ρ, H](11.3.13)where ρ and H are the m<strong>at</strong>rix represent<strong>at</strong>ions <strong>of</strong> ˆρ and Ĥ, respectively. We <strong>of</strong>ten do notdistinguished the m<strong>at</strong>rix represent<strong>at</strong>ions ρ and H, and oper<strong>at</strong>or represent<strong>at</strong>ions ˆρ and Ĥ, asm<strong>at</strong>hem<strong>at</strong>ically, they are the same. The above equ<strong>at</strong>ion replaces the Schrödinger equ<strong>at</strong>ion asthe equ<strong>at</strong>ion <strong>of</strong> motion for the quantum system.11.4 Two-Level Quantum SystemsTwo-level systems are encountered in nuclear magnetic resonance when the spin st<strong>at</strong>es <strong>of</strong>a particle interact with a magnetic field. They can also be used to study the Josephsonjunction effect in superconductive devices. Many more complex systems, for simplicity, canbe approxim<strong>at</strong>ed by a two-level system. When an electric field (optical field) interacts withan <strong>at</strong>om and causes an <strong>at</strong>omic transition between two energy levels, while the other energylevels are far away, the system can be approxim<strong>at</strong>ed by a simpler two-level system. Also,approxim<strong>at</strong>e two-level systems are <strong>of</strong>ten used to represent the value “0” and “1” in quantumcomputing as we shall see l<strong>at</strong>er. We can describe a simple two-level system with a coupledmodeequ<strong>at</strong>ion;i∂ t ψ 1 = E 1 ψ 1 + Kψ 2 (11.4.1)i∂ t ψ 2 = E 2 ψ 2 + Kψ 1 (11.4.2)where if K = 0, the two st<strong>at</strong>es will be st<strong>at</strong>ionary st<strong>at</strong>es evolving independently <strong>of</strong> each other.When K ≠ 0, their time evolution will be different evolving according to the coupled-modeequ<strong>at</strong>ion. We will discuss how to find K.


Density M<strong>at</strong>rix 13711.4.1 Interaction <strong>of</strong> Light with Two-Level SystemsFor an optical system, one assumes th<strong>at</strong> before it is perturbed, the <strong>at</strong>om (or quantum well)is described by a simple two-level system with two st<strong>at</strong>ionary st<strong>at</strong>es bearing energies E 1 andE 2 . The unperturbed Hamiltonian can be described by[ ]E1 0Ĥ 0 =(11.4.3)0 E 2In the above, we assume th<strong>at</strong> the Hamiltonian oper<strong>at</strong>or and its m<strong>at</strong>rix represent<strong>at</strong>ion are thesame. In this case, the perturbing Hamiltonian due to an electric field is assumed to beĤ p = ezE = −E ˆµ (11.4.4)where ˆµ = −eẑ. 3 The m<strong>at</strong>rix represent<strong>at</strong>ion <strong>of</strong> the dipole moment in terms <strong>of</strong> two st<strong>at</strong>ionaryst<strong>at</strong>es before perturb<strong>at</strong>ion is given by 4where ẑ is the position oper<strong>at</strong>or, and(Ĥp)µ mn = −e〈ψ m |ẑ|ψ n 〉 (11.4.5)mn= H p,mn = −Eµ mn (11.4.6)In the above µ 11 = µ 22 = 0, so are H p11 = H p22 = 0, because ψ m and ψ n have differentparities. In other words, if one <strong>of</strong> them has odd parity, the other one has even parity. Thenin the coordin<strong>at</strong>e represent<strong>at</strong>ion, the function |ψ i | 2 z will have odd parity, and its integralevalu<strong>at</strong>es to zero. But for the <strong>of</strong>f diagonal elements, the integrands will have even parity,and their integrals evalu<strong>at</strong>e to nonzero values. The m<strong>at</strong>rix elements above are formed withst<strong>at</strong>ionary st<strong>at</strong>es, and µ 12 can be made pure real so th<strong>at</strong> µ 12 = µ 21 = µ d . Hence, theperturbing Hamiltonian becomes[Ĥ p =0 −Eµ d−Eµ d 0](11.4.7)The above can be simplified to, and approxim<strong>at</strong>ed by a two-level system with the use<strong>of</strong> simple m<strong>at</strong>rix mechanics. This is also very much in the spirit <strong>of</strong> a finite-basis method,where a small number <strong>of</strong> basis can form a good approxim<strong>at</strong>ion <strong>of</strong> the quantum system. Theunperturbed Hamiltonian, Ĥ0 is assumed to have eigenvalues E 1 and E 2 with respect to thesetwo st<strong>at</strong>ionary st<strong>at</strong>es. The total Hamiltonian is then[ ]E1 −EµĤ = Ĥ0 + Ĥp =d(11.4.8)−Eµ d E 2The equ<strong>at</strong>ion <strong>of</strong> motion according to Schrödinger equ<strong>at</strong>ion isi∂ t |ψ〉 = Ĥ|ψ〉 (11.4.9)3 The coordin<strong>at</strong>e represent<strong>at</strong>ion <strong>of</strong> ẑ is just z.4 This is an approxim<strong>at</strong>ion since after perturb<strong>at</strong>ion, the quantum st<strong>at</strong>e is not necessarily describable by justtwo st<strong>at</strong>ionary st<strong>at</strong>es. We saw th<strong>at</strong> in the time-dependent perturb<strong>at</strong>ion theory, all eigenst<strong>at</strong>es <strong>of</strong> the systemwas necessary to approxim<strong>at</strong>e the perturbed system.


138 Quantum Mechanics Made SimpleIn the above, the perturbing Hamiltonian introduces the <strong>of</strong>f diagonal terms which are thecross-coupling terms in coupled-mode theory. They are responsible for causing the transition<strong>of</strong> eigenst<strong>at</strong>es between st<strong>at</strong>e 1 and st<strong>at</strong>e 2. If the <strong>of</strong>f diagonal terms are absent, thest<strong>at</strong>ionary st<strong>at</strong>es will remain in their respective st<strong>at</strong>es without transition. In the above finitebasis approxim<strong>at</strong>ion to the Hamiltonian, the basis used consists <strong>of</strong> the two eigenst<strong>at</strong>es <strong>of</strong> theoriginal unperturbed problem. This is a subspace approxim<strong>at</strong>ion method, whereby an infinitedimensional Hilbert space has been replaced by a two-dimensional subspace.The density m<strong>at</strong>rix which is the m<strong>at</strong>rix represent<strong>at</strong>ion <strong>of</strong> the density oper<strong>at</strong>or is 5[ ]ρ11 ρˆρ =12(11.4.10)ρ 21 ρ 22Equ<strong>at</strong>ion (11.2.6) expresses the density oper<strong>at</strong>or as oper<strong>at</strong>or. But by finding its m<strong>at</strong>rixrepresent<strong>at</strong>ion in terms <strong>of</strong> the two st<strong>at</strong>es assuming th<strong>at</strong> they are orthonormal, the densitym<strong>at</strong>rix is[ ]|a1 (t)|ˆρ =2 a 1 (t)a ∗ 2(t)a 2 (t)a ∗ 1(t) |a 2 (t)| 2 (11.4.11)In (11.2.6), we assume here th<strong>at</strong> the two st<strong>at</strong>es are st<strong>at</strong>ionary st<strong>at</strong>es (they need not be),and the time vari<strong>at</strong>ions <strong>of</strong> the st<strong>at</strong>es have been put into a 1 (t) and a 2 (t). Hence, the <strong>of</strong>fdiagonal terms can be rapidly varying functions <strong>of</strong> time due to the interference between thetwo st<strong>at</strong>ionary st<strong>at</strong>es.Next, we derive the equ<strong>at</strong>ion <strong>of</strong> motion for the density m<strong>at</strong>rix. Sincedˆρdt = i ][ˆρĤ − Ĥ ˆρ= i = i [From the above([ρ11 ρ 12dρ 21dt] [ ]E1 −Eµ d−ρ 21 ρ 22 −Eµ d E 2[ ] [ ])E1 −Eµ d ρ11 ρ 12−Eµ d E 2 ρ 21 ρ 22−Eµ d (ρ 12 − ρ 21 ) −Eµ d (ρ 11 − ρ 22 ) + (E 2 − E 1 )ρ 12−Eµ d (ρ 22 − ρ 11 ) + (E 1 − E 2 )ρ 21 −Eµ d (ρ 21 − ρ 12 )](11.4.12)= i [(ρ 11 − ρ 22 )Eµ d − (E 2 − E 1 )ρ 21 ] = −iω 21 ρ 21 + i µ dE∆ρ (11.4.13)where ω 21 = E 2 − E 1 and ∆ρ = ρ 11 − ρ 22 . We need not write down the equ<strong>at</strong>ion for ρ 12since ρ 12 = ρ ∗ 21.From (11.4.12)dρ 11= − i dt Eµ d(ρ 12 − ρ 21 ),From them, we getd∆ρdtdρ 22dt= − i Eµ d(ρ 21 − ρ 12 ) (11.4.14)= d dt (ρ 11 − ρ 22 ) = − i Eµ d(ρ 12 − ρ 21 − ρ 21 + ρ 12 ) = 2i Eµ d(ρ 21 − ρ ∗ 21)= −4 Eµ dIm(ρ 21 ) (11.4.15)5 Again, we ignore the difference between oper<strong>at</strong>or and its m<strong>at</strong>rix represent<strong>at</strong>ion and use the same not<strong>at</strong>ion.


Density M<strong>at</strong>rix 139since ρ 12 = ρ ∗ 21. If this system models an ensemble <strong>of</strong> identical <strong>at</strong>oms, ∆ρ is proportionalto the popul<strong>at</strong>ion difference <strong>of</strong> <strong>at</strong>oms in st<strong>at</strong>es 1 and 2. The above equ<strong>at</strong>ions, (11.4.13) and(11.4.15), form two coupled equ<strong>at</strong>ions from which the unknowns, ∆ρ = ρ 11 − ρ 22 and ρ 12 canbe solved for.It is noted th<strong>at</strong> if no external field is applied, there is no change <strong>of</strong> the popul<strong>at</strong>ion densitiesin the two st<strong>at</strong>es. Moreover, if the external field is switched <strong>of</strong>f, then (11.4.13) evolves in timeaccording to exp(−iω 21 t). Since these are st<strong>at</strong>ionary st<strong>at</strong>es, their phase rel<strong>at</strong>ionship foran ideal quantum system remains locked in this manner. The driving term, the last termin (11.4.13), due to the electric field being on will cause it to veer away from this phaserel<strong>at</strong>ionship.At his point, we are going to add phenomenological terms to the above equ<strong>at</strong>ions to makethem agree with experimental observ<strong>at</strong>ions. The above two level system, so far, has beenisol<strong>at</strong>ed from a thermal b<strong>at</strong>h or interaction with other <strong>at</strong>oms. It more aptly describes the twolevel system <strong>of</strong> a single <strong>at</strong>om in isol<strong>at</strong>ion. In actuality, there will be a collection <strong>of</strong> <strong>at</strong>oms th<strong>at</strong>will be similarly excited by the electric field. These <strong>at</strong>oms will be coupled to other thermalsources such as by collision with other <strong>at</strong>oms or with the wall <strong>of</strong> the container. The densitym<strong>at</strong>rix can be used to represent the ensemble averaged st<strong>at</strong>e <strong>of</strong> the collection <strong>of</strong> these <strong>at</strong>oms.Here, for a collection <strong>of</strong> <strong>at</strong>oms, ρ 11 − ρ 22 is the fractional popul<strong>at</strong>ion difference in electrons<strong>of</strong> the <strong>at</strong>oms between st<strong>at</strong>e 1 and st<strong>at</strong>e 2. But when coupled to a thermal b<strong>at</strong>h, there will besome electrons excited to level 2 even in the absence <strong>of</strong> an external electric field. Hence, wemodify (11.4.15) to account for the coupling to a thermal b<strong>at</strong>h to becomed 2i∆ρ =dt Eµ d (ρ 21 − ρ ∗ 21) − ∆ρ − ∆ρ o(11.4.16)T 1where ∆ρ o = (ρ 11 − ρ 22 ) ois the quiescent steady st<strong>at</strong>e value <strong>of</strong> ∆ρ in the absence <strong>of</strong> externaldriving field. The transient value ∆ρ is assumed to relax to the quiescent value ∆ρ o in timeT 1 .Also, ρ 21 = C e −iω21t in the absence <strong>of</strong> external driving field. But if the two level systemis coupled to an external he<strong>at</strong> b<strong>at</strong>h the phase coherence between the two energy st<strong>at</strong>es willbe lost and average to zero. We can describe this by adding a term to (11.4.13) to accountfor dephasing, ordρ 21dt= −iω 21 ρ 21 + i µ d E ∆ρ − ρ 21T 2(11.4.17)T 2 is the dephasing time which is usually shorter than T 1 . The first term on the right-handside <strong>of</strong> the above gives rise to a resonant solution if the other terms are absent. The secondterm is a driving term due to the presence <strong>of</strong> the electric field and the popul<strong>at</strong>ion differencebetween st<strong>at</strong>es 1 and 2. The last term is an <strong>at</strong>tenu<strong>at</strong>ion due to the decoherence between thetwo quantum st<strong>at</strong>es. If the two st<strong>at</strong>es are entirely incoherent or uncorrel<strong>at</strong>ed, ρ 21 = 0.At this point, it is prudent to discuss the role <strong>of</strong> the exciting field, which is <strong>of</strong> the formE (t) = E o cos ωt = E o2[e iωt + e −iωt] (11.4.18)When ∆ρ is slowly varying in (11.4.17), the driving term, which is the second term on theright-hand side in (11.4.17) has two rot<strong>at</strong>ing signals e iωt and e −iωt . The response, ρ 21 will


140 Quantum Mechanics Made Simplehave two rot<strong>at</strong>ing signals as well. Hence, ρ 21 = C 1 e −iωt +C 2 e iωt when driven by a sinusoidalsource (11.4.18). If ω ≈ ω 21 , then C 1 ≫ C 2 . Therefore, we defineβ 21 (t) = ρ 21 (t) e iωt = C 1 + C 2 e 2iωt ≈ C 1 (11.4.19)where C 1 is slowly varying so th<strong>at</strong> β 21 (t) is slowly varying when we ignore the rapid term.Using this in (11.4.16) and making us <strong>of</strong> (11.4.18), and keeping only the slowly varying terms,we haveddt ∆ρ = i E oµ d (β 21 − β12) ∗ − ∆ρ − ∆ρ o(11.4.20)T 1In the above, we have kept only the slowest varying term <strong>of</strong> the first term on the right-handside. The rapid term is assumed to only result in a small response in ∆ρ. This is known asthe rot<strong>at</strong>ing wave approxim<strong>at</strong>ion.From (11.4.18), we deduce th<strong>at</strong>dρ 21 (t)dtConsequently, (11.4.17) becomes=[ ddt β 21 (t)]e −iωt − iωβ 21 (t) e −iωt (11.4.21)ddt β 21 = i ∆ω β 21 + i µ dE o2 ∆ρ − β 21T 2(11.4.22)where ∆ω = ω − ω 21 , and we have kept only the slowly varying terms.In this calcul<strong>at</strong>ion, we would eventually like to find out the effect <strong>of</strong> the electric field on theensemble <strong>of</strong> <strong>at</strong>oms. The result is to polarize each <strong>at</strong>om yielding a dipole moment producing apolariz<strong>at</strong>ion current. To this end, we will find the dipole moment produced. The expect<strong>at</strong>ionvalue <strong>of</strong> the dipole moment is〈ˆµ〉 = tr(ˆρˆµ) (11.4.<strong>23</strong>)whereHenceˆρˆµ =[ ] [ ] [ ]ρ11 ρ 12 0 µd ρ12 µ= d ρ 11 µ dρ 21 ρ 22 µ d 0 ρ 22 µ d ρ 21 µ d(11.4.24)〈ˆµ〉 = µ d (ρ 12 + ρ 21 ) = 2µ d Re(ρ 12 ) (11.4.25)because ρ 12 = ρ ∗ 21.After making the rot<strong>at</strong>ing wave approxim<strong>at</strong>ion th<strong>at</strong> ρ 12 = β 12 e iωt , we arrive <strong>at</strong>after using (11.4.19). Therefore〈ˆµ〉 = µ d (ρ 12 + ρ 21 ) = µ d(β12 e iωt + β 21 e −iωt) (11.4.26)〈ˆµ〉 = 2µ d [Re (β 21 ) cos ωt + Im (β 21 ) sin ωt] (11.4.27)


Density M<strong>at</strong>rix 141since β 21 = β ∗ 12. Also, 〈ˆµ〉 has to be a measurable, real-valued quantity, and th<strong>at</strong> is wh<strong>at</strong> theabove shows. In general, it is not in phase with the exciting electric field, having in-phaseand quadr<strong>at</strong>ure components.In steady st<strong>at</strong>e, d∆ρ/dt = 0 in (11.4.20), or ∆ρ in (11.4.22) is a constant. And hence, β 21must tend to a constant in steady st<strong>at</strong>e with dβ 21 /dt = 0. Defining Ω = µ d E o / (2), whereE is the electric field amplitude, we have from (11.4.22) and (11.4.20) th<strong>at</strong>0 = i∆ωβ 21 + iΩ∆ρ − β 21 /T 2 (11.4.28)0 = −∆ΩIm (β 21 ) − (∆ρ − ∆ρ o ) /T 1 (11.4.29)Taking the real and imaginary parts <strong>of</strong> (11.4.28), we arrive <strong>at</strong>0 = −∆ωIm (β 21 ) − Re (β 21 ) /T 2 (11.4.30)0 = ∆ωRe (β 21 ) + Ω∆ρ − Im (β 21 ) /T 2 (11.4.31)Equ<strong>at</strong>ions (11.4.29)-(11.4.31) constitute three equ<strong>at</strong>ions with three unknowns ∆ρ, Im (β 21 ),Re (β 21 ), from which they can be solved for. ThereforeIf there are N <strong>at</strong>oms, then we defineandIn electromagnetics, we have∆ρ = ∆ρ o1 + ∆ω 2 T 2 21 + ∆ω 2 T 2 2 + 4Ω2 T 1 T 2(11.4.32)∆ωΩT2 2 ∆ρ oRe (β 21 ) = −1 + ∆ω 2 T2 2 + (11.4.33)4Ω2 T 1 T 2ΩT 2 ∆ρ oIm (β 21 ) = −1 + ∆ω 2 T2 2 + (11.4.34)4Ω2 T 1 T 2∆N = N∆ρ, ∆N o = N∆ρ o (11.4.35)∆N = ∆N o1 + ∆ω 2 T 2 21 + ∆ω 2 T 2 2 + 4Ω2 T 1 T 2(11.4.36)P = ɛ o ˜χẼ (11.4.37)in phasor represent<strong>at</strong>ion where ˜χ and Ẽ are in frequency domain represent<strong>at</strong>ions. In timedomainP = Re [ ɛ o (χ ′ + iχ ′′ ) E o e −iωt]= ɛ o (χ ′ cos ωt + χ ′′ sin ωt) E o (11.4.38)ButFrom (11.4.27), we haveP = N 〈ˆµ〉 (11.4.39)P = 2Nµ d [Re (β 21 ) cos ωt + Im (β 21 ) sin ωt] (11.4.40)


142 Quantum Mechanics Made SimpleComparing (11.4.38) and (11.4.40), we haveχ ′ = 2Nµ dɛ o E oRe (β 21 )= − µ2 d T 2∆N oɛ o ∆ωT 21 + (∆ω) 2 T 2 2 + 4Ω2 T 1 T 2(11.4.41)χ ′′ = 2Nµ dɛ o E oIm (β 21 )= µ2 d T 2∆N oɛ o 11 + (∆ω) 2 T 2 2 + 4Ω2 T 1 T 2(11.4.42)When the applied electric field is small, we can ignore the 4Ω 2 T 1 T 2 term to getχ ′ (ω) = µ2 d T 2∆N oɛ o χ ′′ (ω) = µ2 d T 2∆N oɛ o (ω 21 − ω) T 21 + (ω − ω 21 ) 2 T 2 211 + (ω − ω 21 ) 2 T 2 2(11.4.43)(11.4.44)Figure 11.1: Plots <strong>of</strong> χ ′ and χ ′′ according to (11.4.43) and (11.4.44) (from DAB Miller).Note th<strong>at</strong> in this analysis, the perturbing field need not be small. The approxim<strong>at</strong>ion wehave made is the two-st<strong>at</strong>e approxim<strong>at</strong>ion. This is a decent approxim<strong>at</strong>ion if the other eigenst<strong>at</strong>esare far away from these two st<strong>at</strong>es in terms <strong>of</strong> energy levels. The other approxim<strong>at</strong>ionis the rot<strong>at</strong>ing wave approxim<strong>at</strong>ion which is good if the exciting frequency <strong>of</strong> the electric fieldis high and is much faster than the relax<strong>at</strong>ion times T 1 and T 2 , and th<strong>at</strong> it is close to ω 21 .Also, this analysis is not a small perturb<strong>at</strong>ion analysis since the exciting field need notbe weak. In fact, Ω above, which is proportional to the exciting electric field, can be large.


Density M<strong>at</strong>rix 143When it is large, we note from (11.4.32) and (11.4.36) th<strong>at</strong> the fractional popul<strong>at</strong>ion differenceapproaches zero, or th<strong>at</strong> ρ 11 = ρ 22 . In this case, the absorption transition is equalto the stimul<strong>at</strong>ed emission. From (11.4.42), when ∆ω = 0, or ω = ω 21 we can express thedenomin<strong>at</strong>or as1 + 4Ω 2 T 1 T 2 = 1 + I I s(11.4.45)where I is the field intensity, and I s is the s<strong>at</strong>ur<strong>at</strong>ion intensity, since I is proportional to thesquare <strong>of</strong> the electric field. The s<strong>at</strong>ur<strong>at</strong>ion intensity is defined to be the intensity <strong>at</strong> whichthe absorption peak <strong>of</strong> χ ′′ will drop to half <strong>of</strong> its value compared to the weak excit<strong>at</strong>ion casewhere I is very small.Equ<strong>at</strong>ions (11.4.20) and (11.4.22) are also known as the optical Bloch equ<strong>at</strong>ions. Theseequ<strong>at</strong>ions were first used to analyze another two-st<strong>at</strong>e system, the nuclear magnetic resonancesystem. With a magnetic field pointing in the z direction, when the nuclear spins are pointingupward and aligned with the magnetic field, they are in the lower energy st<strong>at</strong>e. However,when they are pointing anti-parallel to the magnetic field, they are in a higher energy st<strong>at</strong>e.RF (radio frequency) field can be used to flip the spin st<strong>at</strong>es, and similar equ<strong>at</strong>ions as abovecan be derived to describe the spin dynamics.Exercise 3A Josephson junction can also be described by a two-level quantum system where a Cooperpair (quasi-particle) is trapped on either side <strong>of</strong> a thin insul<strong>at</strong>ing junction. The Hamiltonoper<strong>at</strong>or <strong>of</strong> the system can be described by[ ]E1 KĤ =(11.4.46)K E 2where E 1 and E 2 are the potential energy <strong>at</strong> the bottom <strong>of</strong> the conduction band, accordingto the effective mass theory for such quasi-particle. By connecting the two regions <strong>of</strong> thesuperconductor to a voltage source, a difference in the potential energy between the regionscan be cre<strong>at</strong>ed, yielding E 2 −E 1 = −qV where q = −2e for Cooper pairs, and V is the voltage<strong>of</strong> the source.Use the two-level quantum system theory, whereShow th<strong>at</strong>ρ 11 = |a 1 | 2 , ρ 22 = |a 2 | 2 , ρ 12 = a 1 a ∗ 2 = |a 1 ||a 2 |e i(θ1−θ2) = ρ ∗ 21 (11.4.47)ρ 12 = √ ρ 11 ρ 22 e −iδ (11.4.48)where δ = θ 2 − θ 1 . Next show th<strong>at</strong> the current flow through the Josephson junction is equaltoJ = dρ 11= 2K √ρ11 ρ 22 sin δ (11.4.49)dt Hence, the current flow between the two regions is proportional to the sine <strong>of</strong> the phasedifference between the wave functions <strong>of</strong> the two regions. Furthermore, show th<strong>at</strong>d √ρ11 ρ 22 = 0 (11.4.50)dt


144 Quantum Mechanics Made Simpleand th<strong>at</strong>d(ρ 12 + ρ 21 )= iω 21 (ρ 12 − ρ 21 ) (11.4.51)dtFrom the above, show th<strong>at</strong> the phase difference can be affected by the applied voltage, andth<strong>at</strong> the r<strong>at</strong>e <strong>of</strong> change <strong>of</strong> the phase difference is given byDeduce th<strong>at</strong>If˙δ = −(E 2 − E 1 )/ = qV/ (11.4.52)δ(t) = δ 0 + q ∫ t0V (t ′ )dt ′ (11.4.53)V = V 0 + v cos(ωt) (11.4.54)thenδ(t) = δ 0 + q V 0t + q vsin(ωt) (11.4.55) ωSince is small, the above gives rise to rapidly varying phases, and the associ<strong>at</strong>ed currentcannot be measured, or averages to zero. Usingsin(x + ∆x) ≈ sin(x) + ∆x cos(x) (11.4.56)Show th<strong>at</strong> the current is nowJ≈J 0 [sin(δ 0 + q V 0t) + q v ω sin(ωt) cos(δ 0 + q V 0t)] (11.4.57)Explain how you would pick the frequency ω so th<strong>at</strong> a measurable DC current ensues.Because <strong>of</strong> the sensitivity <strong>of</strong> the Josephson junction to an applied EMF, it can be usedto sense magnetic field in a SQUID (superconducting quantum interference detector). Themagnetic field through the ring in the above figure gener<strong>at</strong>es a difference in the EMF andtherefore, a phase difference on the two arms <strong>of</strong> the ring superconductor with Josephsonjunctions in them. The interference p<strong>at</strong>tern <strong>of</strong> the output current is very sensitive to themagnetic field, and hence, can be used to detect small field differences.


Density M<strong>at</strong>rix 145


146 Quantum Mechanics Made Simple


Chapter 12Quantiz<strong>at</strong>ion <strong>of</strong> Classical Fields12.1 IntroductionThe quantum harmonic oscill<strong>at</strong>or is extremely important for the quantiz<strong>at</strong>ion <strong>of</strong> classicalfields, be it electromagnetic, acoustic, or elastic in n<strong>at</strong>ure. Classical fields can be thought <strong>of</strong>as due to a set <strong>of</strong> coupled classical harmonic oscill<strong>at</strong>ors. In the case <strong>of</strong> electromagnetics field,these classical harmonic oscill<strong>at</strong>ors are dipoles in space polarized by the electric field. In thecase <strong>of</strong> acoustic or elastic field, they are due to coupled <strong>at</strong>omic or molecular vibr<strong>at</strong>ions in amedium. Many experimental evidence suggest th<strong>at</strong> there are photon particles associ<strong>at</strong>ed withelectromagnetic field, and phonon particles associ<strong>at</strong>ed with vibr<strong>at</strong>ions <strong>of</strong> <strong>at</strong>oms or moleculesin crystalline solids.When the classical harmonic oscill<strong>at</strong>ors are replaced by quantum harmonic oscill<strong>at</strong>ors,the classical field becomes quantized as shall be seen. The energy th<strong>at</strong> can propag<strong>at</strong>e andassoci<strong>at</strong>ed with the field is quantized, as in the case <strong>of</strong> the quantum harmonic oscill<strong>at</strong>or. Inclassical electromagnetics when we calcul<strong>at</strong>e energy, e.g. as stored energiesW e = 1 2 ɛ |E|2 , W h = 1 2 µ |H|2 (12.1.1)we think <strong>of</strong> W e and W h as a continuum, capable <strong>of</strong> assuming all values from zero to a largevalue. Also, the power flow by a plane waveS = 12η 0|E| 2 (12.1.2)assumes continuous values in classical electromagnetics. As shall be shown, this cannot bethe case if the electromagnetic field is due to a set <strong>of</strong> coupled quantum harmonic oscill<strong>at</strong>ors.The energy carried by a plane wave is also quantized corresponding to packets <strong>of</strong> energy.The quantized n<strong>at</strong>ure <strong>of</strong> electromagnetic radi<strong>at</strong>ion has been observed historically. Thiswas first suggested for explaining black-body radi<strong>at</strong>ion and deriving the Planck radi<strong>at</strong>ionlaw. L<strong>at</strong>er, it was used by Einstein to explain experimental findings in the photoelectriceffect. From the experimental d<strong>at</strong>a, it was deduced th<strong>at</strong> the packet <strong>of</strong> energy associ<strong>at</strong>ed with147


148 Quantum Mechanics Made Simpleelectromagnetic field isE photon = ω (12.1.3)This quantized energy is very small when ω is small but sizeable in the optical regime. Thegrainy n<strong>at</strong>ure <strong>of</strong> electromagnetic field is unimportant <strong>at</strong> radio frequency (RF), and microwavefrequencies, but is important <strong>at</strong> optical frequencies. We shall show th<strong>at</strong> this is the case whenthe electromagnetic field is quantized when the classical harmonic oscill<strong>at</strong>ors are replaced byquantum harmonic oscill<strong>at</strong>ors.To introduce the quantized n<strong>at</strong>ure <strong>of</strong> electromagnetic field, one way is to regard the fieldin a resonant cavity or a box with periodic boundary condition. The electromagnetic fieldin such a system oscill<strong>at</strong>es harmonically. We can replace the classical harmonic oscill<strong>at</strong>orsassoci<strong>at</strong>ed with such a system with the quantum harmonic oscill<strong>at</strong>or; and hence, quantize theamplitude and the associ<strong>at</strong>ed energy <strong>of</strong> the electromagnetic field.A prevailing view is to think <strong>of</strong> a photon as a particle, but is associ<strong>at</strong>ed with a wave fieldth<strong>at</strong> s<strong>at</strong>isfies Maxwell’s equ<strong>at</strong>ions. This is analogous to an electron being a particle and isassoci<strong>at</strong>ed with a wavefunction as its “halo”. As photons are bosons, we can have a collection<strong>of</strong> photons described by the same wave field, with different amplitudes, since this wave fieldis now endowed with more energy. As shall be seen, the wave field associ<strong>at</strong>ed with photons isquite different from the wave field associ<strong>at</strong>ed with electrons, or other boson particles obeyingSchrödinger equ<strong>at</strong>ion.We will first revisit the quantum harmonic oscill<strong>at</strong>or because <strong>of</strong> its importance here. Nextwe study the wave on a linear <strong>at</strong>omic chain, first as a classical wave, and then as a set <strong>of</strong>coupled quantum harmonic oscill<strong>at</strong>ors. Then we study the quantiz<strong>at</strong>ion <strong>of</strong> electromagneticwave field along similar spirit.12.2 The Quantum Harmonic Oscill<strong>at</strong>or RevisitedThe governing equ<strong>at</strong>ion for the quantum harmonic oscill<strong>at</strong>or isĤψ =(− 2d 22m dz 2 + 1 )2 mω2 z 2 ψ = Eψ (12.2.1)We divide the above by ω to make the oper<strong>at</strong>or and the eigenvalue dimensionless. Lettingξ = √ mω z, the above becomes )1(− d22 dξ 2 + ξ2 ψ = E ω ψ (12.2.2)The above looks almost like A 2 − B 2 with the exception th<strong>at</strong> oper<strong>at</strong>ors are involved instead<strong>of</strong> scalars. One can show th<strong>at</strong>(1√ − d ) ( )1 d2 dξ + ξ √2dξ + ξ = 1 ) (− d22 dξ 2 + ξ2 − 1 ( d2 dξ ξ − ξ d )(12.2.3)dξ


Quantiz<strong>at</strong>ion <strong>of</strong> Classical Fields 149Furthermore, it can be easily shown th<strong>at</strong>( ddξ ξ − ξ d dξ)= Î (12.2.4)Hence,)1(− d22 dξ 2 + ξ2 = √ 1 (− d ) ( )1 d2 dξ + ξ √2dξ + ξ + 1 2(12.2.5)We defineâ † = √ 1 (− d )2 dξ + ξ , cre<strong>at</strong>ion or raising oper<strong>at</strong>or (12.2.6)â = √ 1 ( ) d2 dξ + ξ , annihil<strong>at</strong>ion or lowering oper<strong>at</strong>or (12.2.7)It can easily be shown by integr<strong>at</strong>ion by parts th<strong>at</strong> â † is in fact the conjug<strong>at</strong>e transposeoper<strong>at</strong>or <strong>of</strong> â.Consequently, from (12.2.1) and (12.2.2), we have(â † â + 1 )ψ = E 2 ω ψ (12.2.8)and(Ĥ = ω â † â + 1 )2(12.2.9)SinceE n =(n + 1 )ω (12.2.10)2from the Section 3.5, we conclude th<strong>at</strong>â † â |ψ n 〉 = n |ψ n 〉 (12.2.11)In other words, if |ψ n 〉 is an eigenst<strong>at</strong>e <strong>of</strong> (12.2.1) or (12.2.8), the above must be true. Wedefine a number oper<strong>at</strong>orˆn = â † â (12.2.12)such th<strong>at</strong>ˆn |ψ n 〉 = n |ψ n 〉 (12.2.13)One can also show by direct substitution using (12.2.6) and (12.2.7) th<strong>at</strong>[â,↠] = ââ † − â † â = Î (12.2.14)One can easily show, using the commut<strong>at</strong>ion rel<strong>at</strong>ion, th<strong>at</strong>â † ââ |ψ n 〉 = ( ââ † − 1 ) â |ψ n 〉 = â ( â † â − 1 ) |ψ n 〉 = (n − 1) â |ψ n 〉 (12.2.15)


150 Quantum Mechanics Made SimpleThereforeâ |ψ n 〉 = A n |ψ n−1 〉 (12.2.16)Hence, â is a lowering oper<strong>at</strong>or. Similarly, using the commut<strong>at</strong>ion rel<strong>at</strong>ion,Thereforeâ † ââ † |ψ n 〉 = â † ( 1 + â † â ) |ψ n 〉 = (n + 1) â † |ψ n 〉 (12.2.17)â † |ψ n 〉 = B n+1 |ψ n+1 〉 (12.2.18)Hence, â † is a raising oper<strong>at</strong>or. Testing (12.2.16) with 〈ψ n−1 |, and (12.2.18) with 〈ψ n+1 |, wehave〈ψ n−1 | â |ψ n 〉 = A n , 〈ψ n+1 | â † |ψ n 〉 = B n+1 (12.2.19)By comparing the two equ<strong>at</strong>ions in (12.2.19), taking the transpose <strong>of</strong> the second equ<strong>at</strong>ion,and let n + 1 → n, we deduce th<strong>at</strong>〈ψ n−1 | â |ψ n 〉 = B ∗ n (12.2.20)It is clear th<strong>at</strong>SinceA n = B ∗ n (12.2.21)We conclude th<strong>at</strong>â † â |ψ n 〉 = A n â † |ψ n−1 〉 = A n B n |ψ n 〉 (12.2.22)= |A n | 2 |ψ n 〉 = n |ψ n 〉 (12.2.<strong>23</strong>)A n = √ n (12.2.24)to within an arbitrary phase factor. But the phase factor can be absorbed into the definition<strong>of</strong> â and â † . Therefore, for simplicity, we assume th<strong>at</strong> A n and B n are both real. Thereforeâ |ψ n 〉 = √ n |ψ n−1 〉 (12.2.25)â † |ψ n 〉 = √ n + 1 |ψ n+1 〉 (12.2.26)12.2.1 Eigenfunction by the Ladder ApproachWith the raising oper<strong>at</strong>ors, we can construct the requisite eigenfunctions if we only know theground st<strong>at</strong>e eigenfunction. The ground st<strong>at</strong>e eigenfunction s<strong>at</strong>isfiesWritten explicitly in coordin<strong>at</strong>e space represent<strong>at</strong>ion we haveâ |ψ 0 〉 = 0 (12.2.27)1√2( ddξ + ξ )ψ 0 (ξ) = 0 (12.2.28)


Quantiz<strong>at</strong>ion <strong>of</strong> Classical Fields 151We can solve the above ordinary differential equ<strong>at</strong>ion to getψ 0 (ξ) = 1π 1/4 e−ξ2 /2(12.2.29)We can also verify th<strong>at</strong> ψ 0 (ξ) s<strong>at</strong>isfies (12.2.28) by direct substitution. Consequently, onecan show th<strong>at</strong> (↠) n|ψ0 〉 = √ n! |ψ n 〉 (12.2.30)or th<strong>at</strong> the n-th eigenst<strong>at</strong>e is|ψ n 〉 = 1 √n!(↠) n|ψ0 〉 (12.2.31)12.3 Quantiz<strong>at</strong>ion <strong>of</strong> Waves on a Linear Atomic Chain–PhononsWhen a particle is in simple harmonic motion, quantum mechanics allows us to quantize itsmotion using Schrödinger equ<strong>at</strong>ion for the harmonic oscill<strong>at</strong>or. When we have an <strong>at</strong>omicchain, there will be coupling forces between the <strong>at</strong>oms. When one <strong>of</strong> the <strong>at</strong>oms is set intosimple harmonic oscill<strong>at</strong>ion, the whole chain will be in simple harmonic oscill<strong>at</strong>ion due tocoupling between them. Since the energy <strong>of</strong> one single particle is quantized in the quantumrealm, we expect th<strong>at</strong> harmonic oscill<strong>at</strong>ion <strong>of</strong> the whole chain to be quantized in energy aswell. This is a much easier notion to accept if we accept th<strong>at</strong> the energy <strong>of</strong> one particle isquantized.Figure 12.1: The linear <strong>at</strong>omic chain (from Haken).For <strong>at</strong>oms on a linear <strong>at</strong>omic chain, assuming only nearest neighbor forces between theparticles, the equ<strong>at</strong>ion <strong>of</strong> motion ism 0¨q l (t) = f [q l+1 (t) − q l (t)] − f [q l (t) − q l−1 (t)]= f [q l+1 (t) − 2q l (t) + q l−1 (t)] (12.3.1)The right-hand side <strong>of</strong> the above can be simplified by assuming th<strong>at</strong> q l ∼ e ikla . Furthermore,we impose the periodic boundary condition so th<strong>at</strong>q l = q l+N (12.3.2)


152 Quantum Mechanics Made SimpleConsequently,q l (t) = 1 √Ne ikνla B kν (t) (12.3.3)wherek ν = 2νπNa(12.3.4)where ν is an integer ranging from −∞ to ∞. Since k ν is countable, we will drop the subscriptν and use k as an index. Using (12.3.3) in (12.3.1), we havewithWe let¨B k (t) = g ( e ika + e −ika − 2 ) B k (t) (12.3.5)g = f m 0(12.3.6)B k (t) = B k (0)e ±iω kt(12.3.7)Using (12.3.7) in (12.3.5) yields−ω 2 k = 2g [cos (ka) − 1] (12.3.8)orThe kinetic energy <strong>of</strong> the system isω k = ±2 √ g sin( ) ka2(12.3.9)while the potential energyT =N∑l=1m 02 ˙q2 l (12.3.10)such th<strong>at</strong> the force on the l-th <strong>at</strong>om isV = 1 N 2 f ∑(q l − q l+1 ) 2 (12.3.11)l=1The classical Hamiltonian <strong>of</strong> the system is thenF l = − ∂∂q lV (12.3.12)∑H = N p 2 l+ 1 2m 0 2 f ∑ N (q l − q l+1 ) 2 (12.3.13)l=1l=1


Quantiz<strong>at</strong>ion <strong>of</strong> Classical Fields 153The equ<strong>at</strong>ion <strong>of</strong> motion (12.3.1) can be derived from the above using Hamiltonian equ<strong>at</strong>ions<strong>of</strong> motion derive in Section 2.3, reproduced here as q l = ∂H/∂p l , and p l = −∂H/∂q l . Theywill yieldso th<strong>at</strong>˙q l (t) = 1 m 0p l (t) , ṗ l (t) = f [q l+1 (t) − 2q l (t) + q l−1 (t)]q l (t) = 1 √Ne ikla B k (t) + c.c. (12.3.14)p l (t) = 1 √Ne ikla A k (t) + c.c. (12.3.15)Ḃ k (t) = 1 m 0A k (t) (12.3.16)A˙k (t) = f [ e ika + e −ika − 2 ] ( ) kaB k (t) = −4f sin 2 B k (t) (12.3.17)2In the above, “c.c.” stands for “complex conjug<strong>at</strong>e”. Equ<strong>at</strong>ion (12.3.17) can be written as˙ A k (t) = −ω 2 km 0 B k (t) (12.3.18)whereCombining (12.3.16) and (12.3.18), we haveωk 2 = 4f ( ) kasin 2m 0 2(12.3.19)Ä k (t) = −ω 2 kA k (t) (12.3.20)¨B k (t) = −ω 2 kB k (t) (12.3.21)To have a travelling wave, we letA k (t) = A k (0) e ±iω kt(12.3.22)From (12.3.18), we see th<strong>at</strong>We also find th<strong>at</strong>B k (t) = B k (0) e ±iω kt(12.3.<strong>23</strong>)A k (t) = ±iω k m 0 B k (t) (12.3.24)p 2 l(t) = 2 [ ]1N |A k(t)| 2 +N e2ikla A 2 k(t) + c.c.where the second term is an oscill<strong>at</strong>ory function <strong>of</strong> l. Also(12.3.25)(q l − q l+1 ) = √ 1 e ikla (1 − e ika )B k (t) + c.c.N= −√ 1( ) kae ikla 2i sin e ik a 2 Bk (t) + c.c. (12.3.26)N 2


154 Quantum Mechanics Made SimpleHence(q l − q l+1 ) 2 = 8 ( ) kaN sin2 |B k (t)| 2 + OSC (12.3.27)2where “OSC” stands for “oscill<strong>at</strong>ory functions <strong>of</strong> l”. The oscill<strong>at</strong>ory terms sum to zero overl when we plug (12.3.25) and (12.3.27) into (12.3.13). Therefore,H =1 ( ) ka|A k (t)| 2 + 4f sin 2 |B k (t)| 2m 0 2= 1 m 0|A k (t)| 2 + m 0 ω 2 k|B k (t)| 2= 2m 0 ω 2 k|B k (t)| 2 (12.3.28)after using (12.3.24). 1 If we let√4m0 ω k B k (t) = p k (t) ± iq k (t)to within a phase factor, orthen4m 0 ω k |B k (t)| 2 = p 2 k(t) + q 2 k(t) (12.3.29)H = ω k2[p2k (t) + q 2 k(t) ] (12.3.30)which is identical to the Hamiltonian <strong>of</strong> a classical harmonic oscill<strong>at</strong>or for normalized coordin<strong>at</strong>esand momenta. The reason why the above simple form ensues is:1. By imposing periodic boundary condition on the linear <strong>at</strong>omic chain, only countable,discrete modes with wavenumber k are allowed to propag<strong>at</strong>e on the linear <strong>at</strong>omic chain;2. Only a single mode is allowed to propag<strong>at</strong>e on the linear <strong>at</strong>omic chain so as to makethe solution and the corresponding Hamiltonian very simple.The physical picture we shall have <strong>of</strong> the above is th<strong>at</strong> because only one mode exists, allthe <strong>at</strong>oms on the chain are oscill<strong>at</strong>ing <strong>at</strong> the same frequency and in unison except for a phaselag. So each <strong>of</strong> them is in a simple harmonic motion. Since we have learned how to converta classical harmonic oscill<strong>at</strong>or to a quantum harmonic oscill<strong>at</strong>or, the same principles can beapplied to turn the above into a chain <strong>of</strong> coupled quantum harmonic oscill<strong>at</strong>ors.To convert the above into a quantum harmonic oscill<strong>at</strong>or, we elev<strong>at</strong>e the above amplitudesto oper<strong>at</strong>ors, namelyĤ = ω k2[ˆp2k + ˆq 2 k](12.3.31)1 It is to be noted th<strong>at</strong> because <strong>of</strong> the assumption <strong>of</strong> a right traveling wave in (12.3.22) and (12.3.<strong>23</strong>),|A k (t)| 2 and |B k (t)| 2 in the above are constants. Wh<strong>at</strong> this means is th<strong>at</strong> both the kinetic energy and thepotential energy parts <strong>of</strong> the Hamiltonian are constants <strong>of</strong> motion. However, had we assumed th<strong>at</strong> both rightand left traveling waves exist in (12.3.22) and (12.3.<strong>23</strong>), this would not have been true: both |A k (t)| 2 and|B k (t)| 2 are time varying. However, it can be proved th<strong>at</strong> their sum in the Hamiltonian is a constant.


Quantiz<strong>at</strong>ion <strong>of</strong> Classical Fields 155whereˆp k = −i ddq k, ˆq 2 k = q 2 k (12.3.32)in coordin<strong>at</strong>e space represent<strong>at</strong>ion. We can further define dimensionless entitiesso th<strong>at</strong>ξ k = q k / √ ,Ĥ = ω k2[− d2dξ 2 kˆΠk = ˆp k / √ = −i ddξ k(12.3.33)+ ξ 2 k]= ω k2 (ˆΠ 2 k + ξ 2 k) (12.3.34)(12.3.35)The above is the quantum analogue <strong>of</strong> (12.3.30). Furthermore, we can show th<strong>at</strong>â † k = 1 √2(−iˆΠ k + ξ k)â k = 1 √2(iˆΠ k + ξ k)Hence, by direct use <strong>of</strong> the above, we can show th<strong>at</strong>Ĥ = ω k2[]â † kâk + â k â † k(12.3.36)(12.3.37)We note in the above th<strong>at</strong> â † k and â k are the quantum analogue <strong>of</strong> B k and Bk ∗ . Hence, wecan rewrite (12.3.28) more suggestively asH = m 0 ω 2 k (B ∗ k(t)B k (t) + B k (t)B ∗ k(t)) (12.3.38)The quantum analogue <strong>of</strong> the above is elev<strong>at</strong>e the B k and Bk ∗(Ĥ = m 0 ωk2 ˆB† ˆB k k + ˆB)k ˆB†kto oper<strong>at</strong>ors or(12.3.39)We can identify th<strong>at</strong>√ˆB k = â k2m 0 ω√ kˆB † k = â † k(12.3.40)2m 0 ω kby comparing (12.3.39) with (12.3.37). Also, the above analogue is in the spirit <strong>of</strong> theSchrödinger picture where the time dependence is releg<strong>at</strong>ed to the st<strong>at</strong>e vector and the oper<strong>at</strong>orsare time independent. An arbitrary phase factor can be added to the above but this


156 Quantum Mechanics Made Simpleresult in a time shift for the field. With the above analogue, we can elev<strong>at</strong>e the field to becomeoper<strong>at</strong>ors, namely,√ˆq l =2m 0 ω k N eikla â k + c.c. (12.3.41)√m0 ω kˆp l = −i2N eikla â k + c.c. (12.3.42)These field oper<strong>at</strong>ors are time independent because they are in the Schrödinger picture. Noticeth<strong>at</strong> it is the Hamiltonian <strong>of</strong> the total system involving N <strong>at</strong>oms, or the total energy <strong>of</strong> thetotal system, th<strong>at</strong> is quantized with with energy proportional to ω k . This quantized energyper <strong>at</strong>om is then proportional to ω k /N. This spread <strong>of</strong> the quantized energy per <strong>at</strong>ombecomes smaller as N becomes larger, which becomes unphysical, unless we keep N finite.Hence, to be more realistic, one will have to consider a localized wave packet. We will returnto this point l<strong>at</strong>er.If we had assumed th<strong>at</strong> the solutions are summ<strong>at</strong>ion <strong>of</strong> different k modes, then (12.3.14)and (12.3.15) becomeq l (t) = √ 1 ∑e ikla B k (t) + c.c. (12.3.43)Nkp l (t) = √ 1 ∑e ikla A k (t) + c.c. (12.3.44)NConsequently, (12.3.38), after using Parseval’s theorem, becomesk∑H = m 0 ωk 2 (Bk(t)B ∗ k (t) + B k (t)Bk(t)) ∗ (12.3.45)kand (12.3.41) and (12.3.42) becomeˆq l = ∑ k√2m 0 ω k N eikla â k + c.c. (12.3.46)ˆp l = ∑ k√m0 ω k−i2N eikla â k + c.c. (12.3.47)The above represents a multitude <strong>of</strong> propag<strong>at</strong>ing modes with different k values. They cancome together to form a wave packet th<strong>at</strong> more realistically represent a wave th<strong>at</strong> propag<strong>at</strong>eson an infinite linear <strong>at</strong>omic chain.12.4 Schrödinger Picture versus Heisenberg PictureAt this juncture, it is prudent to discuss the difference between the Schrödinger pictureversus the Heisenberg picture <strong>of</strong> quantum mechanics. In the Schrödinger picture, the timedependence is in the wavefunction or the st<strong>at</strong>e vector; whereas in the Heisenberg picture, thetime dependence is in the oper<strong>at</strong>or th<strong>at</strong> represents an observable.


Quantiz<strong>at</strong>ion <strong>of</strong> Classical Fields 157The Schrödinger equ<strong>at</strong>ion is given byThe formal solution to the above can be written asĤ|ψ〉 = i∂ t |ψ〉 (12.4.1)|ψ〉 = e − i Ĥt |ψ 0 〉 (12.4.2)where |ψ 0 〉 is the st<strong>at</strong>e vector <strong>at</strong> t = 0 and is independent <strong>of</strong> time. The expect<strong>at</strong>ion value <strong>of</strong>an oper<strong>at</strong>or Ô is 〈Ô〉 = 〈ψ|Ô|ψ〉 = 〈ψ 0|e i Ĥt Ôe − i Ĥt |ψ 0 〉 (12.4.3)Hence, in general, the rel<strong>at</strong>ionship between an oper<strong>at</strong>or in the Schrödinger picture and in theHeisenberg picture is given byÔ H = e i Ĥt Ô S e − i Ĥt (12.4.4)where the subscripts “H” and “S” mean the Heisenberg picture and the Schrödinger picturerespectively. It can be shown easily th<strong>at</strong>dÔH= i ] [Ĥ, Ô H (12.4.5)dt For the simple quantum harmonic oscill<strong>at</strong>or,(Ĥ = ω â † â + 1 )(12.4.6)2It can be shown easily th<strong>at</strong>] [Ĥ, â = −ωâ (12.4.7)or th<strong>at</strong>Consequently, solving giveSimilarly, one can show th<strong>at</strong>dâdt= −iωâ (12.4.8)â(t) = â(0)e −iωt (12.4.9)â † (t) = â † (0)e iωt (12.4.10)The above are the annihil<strong>at</strong>ion and cre<strong>at</strong>ion oper<strong>at</strong>ors in the Heisenberg picture. Using theabove in (12.3.46) and (12.3.47), we haveˆq l (t) = ∑ √2m 0 ω k N eikla−iωt â k + c.c. (12.4.11)kˆp l (t) = ∑ √m0 ω k−i2N eikla−iωt â k + c.c. (12.4.12)kThe above are the field oper<strong>at</strong>ors in the Heisenberg picture. They are time dependent.


158 Quantum Mechanics Made Simple12.5 The Continuum LimitIn the continuum limit, we letm 0 = ρa (12.5.1)Then (12.3.1) becomesorρa¨q(x, t) = fa 2 ∂2q(x, t) (12.5.2)∂x2 where g = fa. Assume a traveling wavewith periodic boundary condition; thenρ¨q(t) = g ∂2q(x, t) (12.5.3)∂x2 q(x, t) = 1 √Le ikx B k (t) + c.c. (12.5.4)We define the momentum density to bek ν = 2πνL(12.5.5)Then a Hamiltonian th<strong>at</strong> will lead to (12.5.3) isπ(x) = ρ ˙q(x) (12.5.6)We letH = 12ρ∫ L0π 2 (x)dx + g 2∫ L0( ) 2 ∂q(x)dx (12.5.7)∂xq(x, t) = 1 √Le ikx B k (t) + c.c. (12.5.8)π(x, t) = 1 √Le ikx A k (t) + c.c. (12.5.9)ThenḂ k (t) = 1 ρ A k(t) (12.5.10)˙ A k (t) = −gk 2 B k (t) (12.5.11)


Quantiz<strong>at</strong>ion <strong>of</strong> Classical Fields 159From the above, we havewhere ω 2 k = gk2 /ρ, or¨B k (t) = − gk2ρ B k(t) = −ω 2 kB k (t) (12.5.12)B k (t) = B k (0)e −iω kt(12.5.13)A k (t) = −iω k ρB k (t) (12.5.14)From (12.5.9), we obtain th<strong>at</strong>π 2 (x, t) = 2 L |A k(t)| 2 + OSC (12.5.15)( ∂q(x, t)∂x) 2= 2k2L |B k(t)| 2 + OSC (12.5.16)where “OSC” stands for oscill<strong>at</strong>ory term.terms integr<strong>at</strong>e to zero, andWhen substituted into (12.5.7), the oscill<strong>at</strong>oryH = 1 ρ |A k(t)| 2 + k 2 g|B k (t)| 2 = 2ρω 2 k|B k (t)| 2= ρω 2 k [B ∗ k(t)B k (t) + B k (t)B ∗ k(t)] (12.5.17)Defining4ρω k |B k (t)| 2 = p 2 k(t) + q 2 k(t) (12.5.18)the above becomesH = ω k2[p2k (t) + q 2 k(t) ] (12.5.19)which is the classical Hamiltonian <strong>of</strong> a simple harmonic oscill<strong>at</strong>or. The above can be quantizedto yieldĤ = ω k2()â † kâk + â k â † kand (12.5.17) can be elev<strong>at</strong>ed to quantized form giving[Ĥ = ρωk2 ˆB† ˆB k k + ˆB]k ˆB†k(12.5.20)(12.5.21)Hence, we conclude th<strong>at</strong>√ˆB k = â k (12.5.22)2ρω k


160 Quantum Mechanics Made Simpleup to an arbitrary phase. With this, we can elev<strong>at</strong>e the fields to oper<strong>at</strong>ors, namely,√ˆq(x, t) =ktâ2ρω k L eikx−iω k + c.c. (12.5.<strong>23</strong>)√ωk ρˆπ(x, t) = −iktâ2L eikx−iω k + c.c. (12.5.24)The above are in Heisenberg picture where oper<strong>at</strong>ors are functions <strong>of</strong> time. With multi-modefield, the above generalize toˆq(x, t) = ∑ √ktâ2ρω k L eikx−iω k + c.c. (12.5.25)kˆπ(x, t) = −i ∑ k√ωk ρ2L eikx−iω ktâk + c.c. (12.5.26)12.6 Quantiz<strong>at</strong>ion <strong>of</strong> Electromagnetic FieldFor electromagnetic field, the harmonic oscill<strong>at</strong>ors are dipoles in a medium th<strong>at</strong> are polarizedby an electric field. In general, the electric flux in a medium is given byD = ɛ 0 E + P (12.6.1)The first term is the contribution to the electric flux D due to vacuum, while the second termis the polariz<strong>at</strong>ion density contributed from dipoles polarized in a m<strong>at</strong>erial medium by anelectric field. It is customary to write P = ɛ 0 χE to indic<strong>at</strong>e th<strong>at</strong> the polariz<strong>at</strong>ion density isproportional to the electric field.The time vari<strong>at</strong>ion <strong>of</strong> the electric flux, ∂ t D, is instrumental in giving rise to displacementcurrent, and hence, yielding a propag<strong>at</strong>ing wave. This electric flux exists even in vacuum.Hence, we can imagine th<strong>at</strong> even vacuum is polarized by an electric field to produce dipoledensity. 2 In other words, the oscill<strong>at</strong>ing electric dipole will produce a magnetic field cre<strong>at</strong>ingan inductive effect. Together, they can form a reson<strong>at</strong>ing system behaving like a harmonicoscill<strong>at</strong>or. They resemble a tiny LC tank circuit.As seen previously, the quantiz<strong>at</strong>ion <strong>of</strong> classical field is most easily described by thequantiz<strong>at</strong>ion <strong>of</strong> a modal field. A mode can be formed by a periodic boundary conditionor a perfect electric/magnetic conductor boundary condition. In this manner, the number <strong>of</strong>modes is countably infinite. In view <strong>of</strong> this, we can writeE z = α(t)De ikx + c.c. (12.6.2)2 It is deb<strong>at</strong>able as to wh<strong>at</strong> gives rise to these dipoles in vacuum. In my opinion, it could be electronpositronpairs th<strong>at</strong> are embedded in vacuum. When vacuum is bombarded with an energetic photon, it isknown to produce an electron-position pair.


Quantiz<strong>at</strong>ion <strong>of</strong> Classical Fields 161By using periodic boundary condition, we getk ν = 2νπL(12.6.3)making it countably infinite. (We will drop the subscript ν subsequently.) The associ<strong>at</strong>ed Bfield isB y = β(t) D c eikx + c.c. (12.6.4)where c is the velocity <strong>of</strong> light.Maxwell’s equ<strong>at</strong>ions. FromThe rel<strong>at</strong>ion between α(t) and β(t) can be derived from∇ × E = −∂ t B (12.6.5)we haveUsing (12.6.2) and (12.6.4) in the above, we have∂ x E z = ∂ t B y (12.6.6)From the above, we g<strong>at</strong>her th<strong>at</strong>Usingor more specificallyFinally, we find th<strong>at</strong>or th<strong>at</strong>ikα(t)De ikx + c.c. = ˙β D c eikx + c.c. (12.6.7)˙β = iωα (12.6.8)∇ × B = µ 0 ɛ 0 ∂ t E (12.6.9)∂ x B y = µ 0 ɛ 0 ∂ t E z (12.6.10)˙α = iωβ (12.6.11)¨α = −ω 2 α (12.6.12)¨β = −ω 2 β (12.6.13)A traveling wave solution is th<strong>at</strong> α = α 0 e ±iωt , β = β 0 e ±iωt . From (12.6.8) and (12.6.11), wededuce th<strong>at</strong> α 0 = ±β 0


162 Quantum Mechanics Made Simple12.6.1 HamiltonianThe Hamiltonian <strong>of</strong> the electromagnetic system is given by the total energy <strong>of</strong> the systemwhich iswhereH =W (x) = 1 2∫ L0W (x)dx (12.6.14)(ɛ 0 |E| 2 + 1 µ 0|B| 2 )From (12.6.2), we derive th<strong>at</strong> for a propag<strong>at</strong>ing mode,(12.6.15)E 2 z = 2|α(t)| 2 |D| 2 + OSC (12.6.16)B 2 y = 2|β(t)| 2 |D|2c 2 + OSC (12.6.17)where “OSC” stands for “oscill<strong>at</strong>ory terms”. When substituted into (12.6.14) and integr<strong>at</strong>ed,the oscill<strong>at</strong>ory terms disappear; we haveH = L|D| 2 ɛ 0 |α(t)| 2 + L|D|2µ 0 c 2 |β(t)|2= L|D| 2 ɛ 0 [|α(t)| 2 + |β(t)| 2 ] (12.6.18)Definingthe above can be written as√ ωD =(12.6.19)4Lɛ 0By letting α = p ± iq to within a phase factor, it becomesH = ω 4 [|α|2 + |β| 2 ] = ω 2 |α|2 (12.6.20)H = ω 2 (p2 + q 2 ) (12.6.21)The above resembles the Hamiltonian for a classical harmonic oscill<strong>at</strong>or. Again, the reasonfor its simplicity is th<strong>at</strong> we have assumed a single propag<strong>at</strong>ing mode. This single propag<strong>at</strong>ingmode is causing all the dipoles in the system to be oscill<strong>at</strong>ing time-harmonically in unison,but with a phase lag between them. 3But p and q are observables with Hermitian oper<strong>at</strong>ors. Therefore, the quantum analogueisĤ = ω 2 (ˆp2 + ˆq 2 ) (12.6.22)3 Notice th<strong>at</strong> there is no radi<strong>at</strong>ion damping in these dipoles as the periodic boundary condition wraps thewave around, and energy is never lost from the system.


Quantiz<strong>at</strong>ion <strong>of</strong> Classical Fields 163In coordin<strong>at</strong>e space represent<strong>at</strong>ion,ˆp = −i d , ˆq = q (12.6.<strong>23</strong>)dqAs in the linear <strong>at</strong>omic chain case, we can show th<strong>at</strong>Ĥ = ω 2(â†â + ââ †) = ω(â † â + 1 )2(12.6.24)12.6.2 Field Oper<strong>at</strong>orsAltern<strong>at</strong>ively, we can rewrite (12.6.20) asH = ω 4 [α(t)α∗ (t) + α(t)α ∗ (t)] (12.6.25)The above can be elev<strong>at</strong>ed to be quantum mechanical oper<strong>at</strong>ors so th<strong>at</strong>Ĥ = ω [ˆαˆα † + ˆαˆα †] (12.6.26)4Comparing the above with (12.6.24), we infer th<strong>at</strong>ˆα = √ 2â (12.6.27)to within an arbitrary phase. Hence, α in (12.6.2) can be elev<strong>at</strong>ed to be a field oper<strong>at</strong>or toyield, in the Heisenberg picture,√ωkÊ z = â k e ikx−iωkt + c.c. (12.6.28)2Lɛ 0ˆB y =√ωk2Lɛ 01c âke ikx−iω kt + c.c. (12.6.29)The above is the field oper<strong>at</strong>or for a monochrom<strong>at</strong>ic plane wave th<strong>at</strong> is not localized. Alocalized wave can be obtained by superposing plane waves with different frequencies or wavenumbers. To this end, we haveÊ z = ∑ k√ωk2Lɛ 0â k e ikx−iω kt + c.c. (12.6.30)ˆB y = ∑ k√ωk2Lɛ 01c âke ikx−iω kt + c.c. (12.6.31)The above can be generalized to a plane wave mode propag<strong>at</strong>ing in arbitrary direction,namely,Ê = ∑ √ωke s â k,s e ik·r−iωkt + c.c. (12.6.32)2V ɛ 0k,s


164 Quantum Mechanics Made SimpleˆB = ∑ k,s√ωk2V ɛ 01c ˆk × e s â k,s e ik·r−iω kt + c.c. (12.6.33)where e s is a unit vector denoting the polariz<strong>at</strong>ion <strong>of</strong> the electric field. It is orthogonal to thek vector. It can be either linearly polarized or circularly polarized. In the linearly polarizedcase, e s is either vertical or horizontal. In the circularly polarized case, e s is either right-handor left-hand circularly polarized. In the above, we have defined the Hamiltonian as an integralover a volume V , and hence, the normaliz<strong>at</strong>ion above is with respect to V .A photon is a packet <strong>of</strong> energy th<strong>at</strong> propag<strong>at</strong>es due to coupled quantum harmonic oscill<strong>at</strong>ors.This is similar to the phonon case which is a packet <strong>of</strong> energy due to coupled quantumharmonic oscill<strong>at</strong>ors <strong>of</strong> mechanical origin. A photon can also propag<strong>at</strong>e in a dielectric medium,and it is quite clear then th<strong>at</strong> in this case, the photon is due to coupled quantum harmonicoscill<strong>at</strong>ors <strong>of</strong> dipolar origin (or electronic origin).The quantiz<strong>at</strong>ion <strong>of</strong> energy is with respect to the entire Hamiltonian which describes a set<strong>of</strong> coupled quantum harmonic oscill<strong>at</strong>ors. For the purely monochrom<strong>at</strong>ic photon, as is seenbefore, if the quantiz<strong>at</strong>ion energy is ω k , it will be shared among all the coupled harmonic oscill<strong>at</strong>ors.This is untenable physically as the number <strong>of</strong> coupled harmonic oscill<strong>at</strong>ors becomeslarge. Hence, it is more appropri<strong>at</strong>e to think <strong>of</strong> a photon as quasi-monochrom<strong>at</strong>ic as a linearsuperposition <strong>of</strong> wavenumbers k, allowing a localized wave packet to be formed. A photoncan be thought <strong>of</strong> as a “particle” possessing an expected or average packet <strong>of</strong> energy ω e , butit exists as a linear superposition <strong>of</strong> different modes with different k’s. Each mode can have aenergy ω k associ<strong>at</strong>ed with it, but as the particle is a linear superposition <strong>of</strong> different modes,ω e is the expect<strong>at</strong>ion value <strong>of</strong> the photon energy when the photon is in a superposition <strong>of</strong>different quantum st<strong>at</strong>es. 412.6.3 Multimode Case and Fock St<strong>at</strong>eWhen polariz<strong>at</strong>ion is considered, we may write the Hamiltonian for a single mode photon asĤ k = ∑ sĤ k,s = ∑ s(ω k â † k,sâk,s + 1 )2(12.6.34)where s stands for either horizontal or vertical polariz<strong>at</strong>ion. The above can be easily derivedby considering the electric field with both polariz<strong>at</strong>ions present.In the above, we can denote eigenst<strong>at</strong>e <strong>of</strong> the Hamiltonian as 5|ψ〉 = |n v 〉|n h 〉 (12.6.35)where v stands for vertical polariz<strong>at</strong>ion and h stands for horizontal polariz<strong>at</strong>ion. In theabove, there are n v photons in the vertical polariz<strong>at</strong>ion st<strong>at</strong>e and there are n h photons in thehorizontal polariz<strong>at</strong>ion st<strong>at</strong>e.4 This is my personal opinion.5 When the Hamiltonian consists <strong>of</strong> sum <strong>of</strong> the Hamiltonians <strong>of</strong> two quantum systems, the st<strong>at</strong>es <strong>of</strong> theHamiltonian oper<strong>at</strong>or can be represented by the product space <strong>of</strong> the st<strong>at</strong>es <strong>of</strong> the individual quantum system.


Quantiz<strong>at</strong>ion <strong>of</strong> Classical Fields 165When many modes are considered simultaneously, it can be shown th<strong>at</strong> the correspondingHamiltonian isĤ = ∑ Ĥ k,s = ∑ (ω k â † k,sâk,s + 1 )(12.6.36)2k,sk,sThe eigenst<strong>at</strong>e <strong>of</strong> the above Hamiltonian can be written as|ψ〉 = ∏ k,s|n k,s 〉 (12.6.37)The above can be written altern<strong>at</strong>ively as|ψ〉 = ∏ s|n k1,s〉|n k2,s〉|n k3,s〉 · · · = ∏ s|n k1,s, n k2,s, n k3,sn k4,s · · ·〉 = ∏ s|{n k,s }〉 (12.6.38)The above st<strong>at</strong>e is known as a Fock st<strong>at</strong>e or the occup<strong>at</strong>ional number st<strong>at</strong>e. It is customaryto leave out the st<strong>at</strong>es where the occup<strong>at</strong>ional number is zero, listing only the st<strong>at</strong>es withnon-zero number <strong>of</strong> photons. For example|ψ〉 = |3 k1,v, 2 k2,h, 1 k3,v〉 (12.6.39)indic<strong>at</strong>es a quantum st<strong>at</strong>e with three V (vertical) polarized photons in k 1 mode, two H(horizontal) polarized photons in k 2 mode, and one V polarized photon in k 3 mode.12.6.4 One-Photon St<strong>at</strong>eThe understanding <strong>of</strong> the one-photon st<strong>at</strong>e is important because <strong>of</strong> its usefulness in quantuminform<strong>at</strong>ion, quantum communic<strong>at</strong>ion, and quantum cryptography. It is prudent to elabor<strong>at</strong>eon it further. For example, when only one k mode is present, to denote a photon in the verticalpolariz<strong>at</strong>ion st<strong>at</strong>e, and no photons in the horizontal polariz<strong>at</strong>ion st<strong>at</strong>e, the st<strong>at</strong>e vector shouldbeOften, this is just written as|ψ〉 = |1 v 〉|0 h 〉 (12.6.40)|ψ〉 = |1 v 〉 (12.6.41)where the second part is understood. For a photon in arbitrary polariz<strong>at</strong>ion, it is written as|ψ〉 = a v |1 v 〉 + a h |1 h 〉 (12.6.42)The above denotes a one-photon st<strong>at</strong>e where the photon is in the linear superposition <strong>of</strong> twoorthonormal one-photon st<strong>at</strong>es. Hence,|a v | 2 + |a h | 2 = 1 (12.6.43)


166 Quantum Mechanics Made SimpleOften, photons are gener<strong>at</strong>ed by <strong>at</strong>omic transitions whereby the electron changes from ahigh energy st<strong>at</strong>e to a low energy st<strong>at</strong>e giving rise to one or more photons. Since this is acausal event, the wave field associ<strong>at</strong>ed with a photon is a localized wave in space time. 6 Thelocalized wave is formed by linear superposing wave field with different k values. For instance,a localized one-photon, vertically polarized st<strong>at</strong>e is a linear superposition <strong>of</strong> a number <strong>of</strong> onephotonst<strong>at</strong>es with different k values as follows:|ψ〉 = ∑ ka k,v |1 k,v 〉 (12.6.44)The wave field associ<strong>at</strong>ed with the above st<strong>at</strong>e can be localized in space time, and the wavefield s<strong>at</strong>isfies causality.It is to be noted th<strong>at</strong> (12.6.36) is a quantum mechanical Hamiltonian involving the sumover individual Hamiltonians <strong>of</strong> different quantum systems. The quantum system arises from aset <strong>of</strong> coupled quantum harmonic oscill<strong>at</strong>ors. The choice <strong>of</strong> appropri<strong>at</strong>e modes “diagonalizes”the system, giving rise to apparently uncoupled systems. When a photon is in a quantumst<strong>at</strong>e, it can be in a linear superposition <strong>of</strong> different quantum st<strong>at</strong>es <strong>of</strong> these different quantumsystems. Therefore, one-photon st<strong>at</strong>es (12.6.42) and (12.6.44) should be thought <strong>of</strong> as a linearsuperposition <strong>of</strong> different quantum st<strong>at</strong>es, subject to the quantum interpret<strong>at</strong>ion <strong>of</strong> quantummechanics. The particle is in a linear superposition <strong>of</strong> st<strong>at</strong>es before the measurement, and itcollapses to one <strong>of</strong> the st<strong>at</strong>es after the measurement.Even though a photon is associ<strong>at</strong>ed with a packet <strong>of</strong> energy, when it is detected with apolarizer (th<strong>at</strong> detects it either in the vertical or horizontal polarized st<strong>at</strong>e), it is found eitherin one st<strong>at</strong>e or the other. The packet <strong>of</strong> energy is never split between the two st<strong>at</strong>es indic<strong>at</strong>edby (12.6.42). Hence, experiment evidence suggests th<strong>at</strong> a photon is a quantum particle in thesense th<strong>at</strong> an electron is a quantum particle <strong>at</strong> the quantum level. The same interpret<strong>at</strong>ionapplies to (12.6.44). The subject <strong>of</strong> quantum interpret<strong>at</strong>ion will be discussed l<strong>at</strong>er.12.6.5 Coherent St<strong>at</strong>e RevisitedThe expect<strong>at</strong>ion value <strong>of</strong> the annihil<strong>at</strong>ion and cre<strong>at</strong>ion oper<strong>at</strong>ors with respect to the photonnumber st<strong>at</strong>es is always zero. Therefore, the expect<strong>at</strong>ion value <strong>of</strong> the field oper<strong>at</strong>ors, whichare proportional to the annihil<strong>at</strong>ion and cre<strong>at</strong>ion oper<strong>at</strong>ors, is always zero. Hence, the numberst<strong>at</strong>es are non-classical. To arrive <strong>at</strong> an expect<strong>at</strong>ion value <strong>of</strong> field oper<strong>at</strong>ors th<strong>at</strong> are non-zero,and resemble a classical field, one has to work with the coherent st<strong>at</strong>es.We have studied the coherent st<strong>at</strong>e previously. It is defined to be the eigenst<strong>at</strong>e <strong>of</strong> theannihil<strong>at</strong>ion oper<strong>at</strong>or. Namely, if |α〉 represents a coherent st<strong>at</strong>e, thenâ|α〉 = α|α〉 (12.6.45)where we have used the eigenvalue α to index the coherent st<strong>at</strong>e. Since the number st<strong>at</strong>e |n〉is complete, we can expand the coherent st<strong>at</strong>e in terms <strong>of</strong> the number st<strong>at</strong>e, or|α〉 =∞∑C n |n〉 (12.6.46)n=06 See the appendix for a discussion on localized wave packets.


Quantiz<strong>at</strong>ion <strong>of</strong> Classical Fields 167When the annihil<strong>at</strong>ion oper<strong>at</strong>or is applied to both sides, we obtain∞∑∞∑∞∑ √ ∑∞ √â|α〉 = C n â|n〉 = C n â|n〉 = C n n|n − 1〉 = C n+1 n + 1|n〉 (12.6.47)n=0n=1n=1where we have used â|0〉 = 0, â|n〉 = √ n|n − 1〉. Equ<strong>at</strong>ing the above with α|α〉, we haven=0∞∑ √ ∑∞C n+1 n + 1|n〉 = α C n |n〉 (12.6.48)n=0By the orthogonality <strong>of</strong> the number st<strong>at</strong>es, m<strong>at</strong>ching the coefficients, we have C n+1 =αC n / √ n + 1, orConsequently,n=0C n = C n−1 α/ √ n = C n−2 α 2 / √ (n(n − 1)) = C 0 α n / √ n! (12.6.49)|α〉 = C 0∞ ∑n=0α n√n!|n〉 (12.6.50)The above can be normalized to show th<strong>at</strong> C 0 = exp(−|α| 2 /2).The photon number oper<strong>at</strong>or is ˆn = â † â. It can be shown th<strong>at</strong> the average number <strong>of</strong>photons associ<strong>at</strong>ed with a coherent st<strong>at</strong>e is given bywhere 〈α|â † = 〈α|α ∗ is used. Moreover〈ˆn〉 = ¯n = 〈α|ˆn|α〉 = 〈α|â † â|α〉 = |α| 2 (12.6.51)〈ˆn 2 〉 = 〈α|ˆn 2 |α〉 = 〈α|â † ââ † â|α〉 = 〈α|â † â † ââ + â † â|α〉 = |α| 4 + |α| 2 = ¯n 2 + ¯n (12.6.52)where we have used ââ † − â † â = 1. Then∆n = √ 〈ˆn 2 〉 − 〈ˆn〉 2 = ¯n 1/2 (12.6.53)The above is characteristic <strong>of</strong> a Poisson process. The probability <strong>of</strong> detecting n photons in acoherent st<strong>at</strong>e isP n = |〈n|α〉| 2 |α|2n−|α|2 −¯n ¯nn= e = en! n!(12.6.54)typical <strong>of</strong> a Poisson distribution. 7 The fractional uncertainty in the photon number is∆n¯n = 1 √¯n(12.6.55)7 An example <strong>of</strong> a Poisson distribution is a shot noise source, producing ¯n photons in interval τ. Theprobability <strong>of</strong> find n photons in this time interval follows the Poisson distribution.


168 Quantum Mechanics Made SimpleFigure 12.2: Two typical Poisson distributions.¯n = 200 (from Haus).Case (a) has ¯n = 50 while case (b) hasFor a photon number st<strong>at</strong>e |n ′ 〉, it can be shown th<strong>at</strong>P n = |〈n|n ′ 〉| 2 = δ nn ′The above distribution has zero standard devi<strong>at</strong>ion. A light with photon distribution narrowerthan a Poisson distribution is known as sub-Poissonian. It illustr<strong>at</strong>es the quantum n<strong>at</strong>ure <strong>of</strong>light. On the other hand, light with photon distribution broader than a Poisson distributionis known as super-Poissonian. It illustr<strong>at</strong>es the incoherent n<strong>at</strong>ure <strong>of</strong> light since coherent lighthas Poissonian distribution.Time Evolution <strong>of</strong> the Coherent St<strong>at</strong>eIn the Schrödinger picture, the time dependence is with the eigenfunctions; hence, the timedependence <strong>of</strong> the coherent st<strong>at</strong>e isThe above can be written as|α, t〉 = e −iωt/2 e − 1 2 |α|2|α, t〉 = e − 1 2 |α|2∞ ∑n=0∞ ∑n=0α n e −i(n+1/2)ωt√n!|n〉 (12.6.56)(αe −iωt ) n√n!|n〉 = e −iωt/2 |αe −iωt 〉 (12.6.57)In the Heisenberg picture, the coherent st<strong>at</strong>e is time independent, and time-dependent


Quantiz<strong>at</strong>ion <strong>of</strong> Classical Fields 169eigenst<strong>at</strong>e can be obtained by applying the oper<strong>at</strong>or|α, t〉 = e −iĤt/ |α〉 = e − 1 2 |α|2∞ ∑n=0= e − 1 2 |α|2 ∞ ∑n=0α n√n!e −iĤt/ |n〉α n√n!e −i(n+1/2)ωt |n〉 = e −iωt/2 e − 1 2 |α|2∞ ∑n=0(αe −iωt ) n√n!|n〉 = e −iωt/2 |αe −iωt 〉(12.6.58)It can be further shown th<strong>at</strong> the above can be summed in terms <strong>of</strong> a “Gaussian pulse” 8indic<strong>at</strong>ing th<strong>at</strong> this pulse evolves in time without shape distortion.Figure 12.3: The time evolution <strong>of</strong> the coherent st<strong>at</strong>e. It follows the motion <strong>of</strong> a classicalpendulum or harmonic oscill<strong>at</strong>or (From Gerry and Knight).The expect<strong>at</strong>ion value <strong>of</strong> the field oper<strong>at</strong>or in the Heisenberg picture with respect to thecoherent st<strong>at</strong>e is thenE = 〈α|Ê|α〉 = ∑ k,sSimilarly,√ωk2V ɛ 0e s 〈α|â k,s |α〉e ik·r−iω kt + c.c. = ∑ k,sB = 〈α| ˆB|α〉 = ∑ k,s√ωk2V ɛ 0e s α k,s e ik·r−iω kt + c.c.(12.6.59)√ωk2V ɛ 01c ˆk × e s α k,s e ik·r−iω kt + c.c. (12.6.60)The above certainly look like a classical field where α k,s are complex numbers.8 It is to be noted th<strong>at</strong> the Gaussian pulse we refer to is with respect to the displacement <strong>of</strong> the quantumharmonic oscill<strong>at</strong>or, not with respect to the direction <strong>of</strong> propag<strong>at</strong>ion <strong>of</strong> the photon.


170 Quantum Mechanics Made Simple12.7 Thermal LightCoherent st<strong>at</strong>e is a st<strong>at</strong>e where all the photon number st<strong>at</strong>es are in phase and in quantumcoherence. Hence the density represent<strong>at</strong>ion <strong>of</strong> the coherent st<strong>at</strong>e isˆρ = |α〉〈α| = e −|α|2∞ ∑∞∑n=0 n ′ =0α n√n!α ∗n′√n′ ! |n〉〈n′ | (12.7.1)This is indic<strong>at</strong>ed by th<strong>at</strong> the <strong>of</strong>f-diagonal elements <strong>of</strong> the density oper<strong>at</strong>or multiply |n〉〈n ′ |where n ≠ n ′ are non-zero due to incoherence.Thermal light is a superposition <strong>of</strong> different photon number st<strong>at</strong>es and the st<strong>at</strong>e vectorcan be expressed as∞∑|ψ〉 = a n |n〉 (12.7.2)n=0However, these photon number st<strong>at</strong>es are incoherent with respect to each other since thelights are excited randomly by a thermal b<strong>at</strong>h. The density oper<strong>at</strong>or for such a st<strong>at</strong>e can bewritten as∞∑ˆρ T = |ψ〉〈ψ| = |a n | 2 |n〉〈n| (12.7.3)where the <strong>of</strong>f diagonal terms ensemble average to zero.According to Boltzmann’s law, <strong>at</strong> thermal equilibrium, the probability <strong>of</strong> finding a systemto be energy st<strong>at</strong>e E n isP (n) = e−En/k BT(12.7.4)Zwhere∞∑Z = e −En/k BT(12.7.5)n=0n=0where Z is the partition function. It is clear th<strong>at</strong> ∑ ∞For photon, E n = ( n + 2) 1 ω. Hence,Z = e − 1 2 ω/k BT∞∑e −nω/kBT = e − 1 2 ω/k BTn=0where x = e −ω/k BT and from (12.7.4)n=0n=0n=0P (n) = 1.∞∑x n =e− 1 2 ω/k BT1 − e −ω/k BT(12.7.6)P (n) = e −nω/k BT (1 − e −ω/k BT ) (12.7.7)The mean photon number is∞∑∞∑¯n = nP (n) = nx n (1 − x) =x1 − x = 1e ω/k BT− 1n=0(12.7.8)The probability distribution given by (12.7.7) definitely has a broader spread compared to thecase for coherent st<strong>at</strong>e, which has a Poissonian probability distribution. Light <strong>of</strong> this n<strong>at</strong>ureis termed super-Poissonian.


Quantiz<strong>at</strong>ion <strong>of</strong> Classical Fields 171Exercise 4We have shown in class th<strong>at</strong> for a linear <strong>at</strong>omic chain, if we start out with a single modetravelling wave solution with periodic boundary condition such th<strong>at</strong>q l (t) = 1 √Ne iknla B n (t) + c.c.p l (t) = 1 √Ne iknla A n (t) + c.c.the Hamiltonian for a linear <strong>at</strong>omic chain is given byH = 1 m |A n (t) | 2 + 4f sin 2 (k n a/2) |B n (t) | 21. Show rigorously th<strong>at</strong> if we start out with a superposition <strong>of</strong> travelling wavesq l (t) =p l (t) =∞∑n=−∞∞∑n=−∞1√Ne iknla B n (t) + c.c.1√Ne iknla A n (t) + c.c.the Hamiltonian for a linear <strong>at</strong>omic chain is given byH =∞∑n=−∞1m |A n (t) | 2 + 4f sin 2 (k n a/2) |B n (t) | 22. First write down the above Hamiltonian so th<strong>at</strong> it involves only |B n (t) | 2 . Then writedown the quantum mechanical analog <strong>of</strong> this Hamiltonian.Hint: It may be simpler to define A −n = A ∗ n and B −n = B ∗ n to make the Fourier expansionreal-valued without the “complex conjug<strong>at</strong>ion” term.


172 Quantum Mechanics Made Simple


Chapter 13Schrödinger Wave Fields13.1 IntroductionWe notice th<strong>at</strong> electromagnetic field can be viewed as a collection <strong>of</strong> photons each with anfield <strong>at</strong>tached. These photons collectively form the coherent st<strong>at</strong>e th<strong>at</strong> produces a wave fieldth<strong>at</strong> is analogous to the classical electromagnetic wave field. Photons are bosons whose wavefield s<strong>at</strong>isfies Maxwell’s equ<strong>at</strong>ions.Similarly, if a boson wavefunction s<strong>at</strong>isfies Schrödinger equ<strong>at</strong>ion, a collection <strong>of</strong> bosonsgives rise to a wave field th<strong>at</strong> s<strong>at</strong>isfies Schrödinger equ<strong>at</strong>ion. Similarly, a collection <strong>of</strong> fermionsgives rise to a wave field th<strong>at</strong> s<strong>at</strong>isfies Schrödinger equ<strong>at</strong>ion if each individual particle wavefunctions<strong>at</strong>isfies the same equ<strong>at</strong>ion. We will call such fields Schrödinger wave fields.We have seen th<strong>at</strong> the number <strong>of</strong> bosons in an electromagnetic wave field can be tracked bythe annihil<strong>at</strong>ion and cre<strong>at</strong>ion oper<strong>at</strong>ors; similar oper<strong>at</strong>ors are needed for tracking the particle<strong>of</strong> a Schrödinger wave field. The wave field can be a collection <strong>of</strong> bosons or a collection <strong>of</strong>fermions. Eventually, the bookkeeping <strong>of</strong> the many-particle system becomes simpler.For Schrödinger equ<strong>at</strong>ion, this is the second quantiz<strong>at</strong>ion. The first quantiz<strong>at</strong>ion was thediscovery th<strong>at</strong> electrons can be associ<strong>at</strong>ed with a wavefunction, and the trapped modes inpotential well are quantized. The second quantiz<strong>at</strong>ion refers to the fact th<strong>at</strong> the Schrödingerwave field, which can be tre<strong>at</strong>ed as a continuum, can be made granular or quantized. This isthe advanced way <strong>of</strong> expressing the wave-particle duality concept inherent in a wave field. 113.2 Fock Space for FermionsFor fermions, each eigenmode or eigenst<strong>at</strong>e is either occupied or unoccupied. In general, therewill be infinitely many modes th<strong>at</strong> the fermions can occupy. We could denote a two-particlest<strong>at</strong>e as|ψ 2p 〉 = | · · · , 0, 1 m , 0, · · · , 0, 1 v , 0, · · · 〉 (13.2.1)1 The quantiz<strong>at</strong>ion <strong>of</strong> classical fields is also known as first quantiz<strong>at</strong>ion for them.173


174 Quantum Mechanics Made Simplewhere only the m and v modes are occupied and the rest <strong>of</strong> the modes are unoccupied. Thenot<strong>at</strong>ion above is cumbersome, and it is common to denote the two-particle fermion Fockst<strong>at</strong>e aswhere for fermions, it is necessary th<strong>at</strong>We define a fermion cre<strong>at</strong>ion oper<strong>at</strong>or such th<strong>at</strong>|ψ 2p 〉 = |m, v〉 (13.2.2)|m, v〉 = −|v, m〉 (13.2.3)ˆb†k|m, n〉 = |k, m, n〉 (13.2.4)The newly cre<strong>at</strong>ed particle always occupies the first space in the Fock st<strong>at</strong>e vector.definition thenˆb† ˆb † l k| · · · 〉 = |l, k, · · · 〉 (13.2.5)ByThereforeˆb† kˆb † l | · · · 〉 = |k, l, · · · 〉 = −|l, k, · · · 〉 = −ˆb † l ˆb † k | · · · 〉 (13.2.6)Similarly, we can show th<strong>at</strong>ˆb† kˆb † l + ˆb † l ˆb † k = 0 (13.2.7)ˆbkˆbl + ˆb lˆbk = 0 (13.2.8)or th<strong>at</strong> the above is the Hermitian conjug<strong>at</strong>e <strong>of</strong> (13.2.7). To show (13.2.8) more rigorously,we start with a fermion Fock st<strong>at</strong>e|ψ〉 = | · · · , l, k, · · · 〉 (13.2.9)This st<strong>at</strong>e acquires a minus sign when we permute two <strong>of</strong> the particles. Finally, we can havea st<strong>at</strong>e such th<strong>at</strong>An annihil<strong>at</strong>ion oper<strong>at</strong>or can be defined such th<strong>at</strong>It can be shown th<strong>at</strong>Similarly,|ψ F 〉 = |l, k, · · · 〉 = −|k, l, · · · 〉 (13.2.10)ˆbl |l, k, · · · 〉 = |k, · · · 〉 (13.2.11)ˆbkˆbl |ψ F 〉 = ˆb kˆbl |l, k, · · · 〉 = ˆb k |k, · · · 〉 = | · · · 〉 (13.2.12)−ˆb lˆbk |ψ F 〉 = ˆb lˆbk |k, l, · · · 〉 = ˆb l |l, · · · 〉 = | · · · 〉 (13.2.13)


Schrödinger Wave Fields 175Therefore,The above equ<strong>at</strong>ions (13.2.7) and (13.2.14) imply th<strong>at</strong> if k = l,ˆbkˆbl + ˆb lˆbk = 0 (13.2.14)ˆb† kˆb † k = 0 (13.2.15)ˆbkˆbk = 0 (13.2.16)It means th<strong>at</strong> no two particles can be cre<strong>at</strong>ed in the same mode k or no two particles can beannihil<strong>at</strong>ed from the same mode k.By the same token, we can show th<strong>at</strong> for k ≠ l,When k = l, we have a st<strong>at</strong>e |k, · · · 〉 and thenIn conclusion,In summary,ˆb† kˆb l + ˆb lˆb†k = 0 (13.2.17)(ˆb † kˆb k + ˆb kˆb†k )|k, · · · 〉 = ˆb † kˆb k |k, · · · 〉 + ˆb kˆb†k |k, · · · 〉= ˆb † k | · · · 〉 + 0|k, · · · 〉 = ˆb † k | · · · 〉 = |k, · · · 〉= Î|k, · · · 〉 (13.2.18)[ˆb†k , ˆb l]+][ˆb†k , ˆb † l[ˆbk , ˆb]lˆb† kˆb l + ˆb lˆb†k = δ klÎ (13.2.19)++= ˆb † kˆb l + ˆb lˆb†k = δ klÎ (13.2.20)= ˆb † kˆb † l + ˆb † l ˆb † k = 0 (13.2.21)= ˆb kˆbl + ˆb lˆbk = 0 (13.2.22)where [A, B] += AB + BA. These commut<strong>at</strong>ion rel<strong>at</strong>ions, as we shall see, do wonders insimplifying the book-keeping <strong>of</strong> many particle problems.13.3 Field Oper<strong>at</strong>orsJust as in the case <strong>of</strong> electromagnetic field oper<strong>at</strong>ors th<strong>at</strong> represent many photons whichare bosons, we can define field oper<strong>at</strong>ors for fermions and bosons in general. A one-particleSchrödinger field oper<strong>at</strong>or is defined such th<strong>at</strong>ˆψ(r) = ∑ jˆbj φ j (r) (13.3.1)


176 Quantum Mechanics Made Simplewhere φ j is an eigenmode from a complete set <strong>of</strong> orthonormal functions. It may or may not bethe j-th eigenst<strong>at</strong>e <strong>of</strong> the quantum system governed by the one-particle Schrödinger equ<strong>at</strong>ion.The above field oper<strong>at</strong>or acts on a st<strong>at</strong>e vector in the Fock space. For instance,ˆψ(r)|m〉 = ˆψ(r)ˆb † m|0〉= ∑ j= ∑ jφ j (r)ˆb jˆb† m |0〉φ j (r)(δ jm − ˆb † mˆb j )|0〉= φ m (r)|0〉 (13.3.2)where the commut<strong>at</strong>ion rel<strong>at</strong>ion (13.2.20) has been used. Notice th<strong>at</strong> the above field oper<strong>at</strong>or,when oper<strong>at</strong>ing on a Fock vector with a m-th mode occupied, produces a vector tagged withthe sp<strong>at</strong>ial dependence <strong>of</strong> the m-th eigenmode.After using the completeness property <strong>of</strong> the orthonormal basis, plus the use <strong>of</strong> commut<strong>at</strong>ionrel<strong>at</strong>ions for the annihil<strong>at</strong>ion and cre<strong>at</strong>ion oper<strong>at</strong>ors, the field oper<strong>at</strong>or can be shownto s<strong>at</strong>isfy the following commut<strong>at</strong>ion rel<strong>at</strong>ions:[ ˆψ† (r), ˆψ(r )]′ = ˆψ † (r) ˆψ(r ′ ) + ˆψ(r ′ ) ˆψ † (r) = δ(r − r ′ ) (13.3.3)+[ ˆψ† (r), ˆψ † (r )]′ = ˆψ † (r) ˆψ † (r ′ ) + ˆψ † (r ′ ) ˆψ † (r) = 0 (13.3.4)+[ˆψ(r), ˆψ(r )]′ = ˆψ(r) ˆψ(r ′ ) + ˆψ(r ′ ) ˆψ(r) = 0 (13.3.5)As an extension, one can define two-particle field oper<strong>at</strong>or+ˆψ (r 1 , r 2 ) = 1 √2∑n 1,n 2ˆbn2ˆbn1 φ n1 (r 1 ) φ n2 (r 2 ) (13.3.6)where the factor 1/ √ 2 is needed for normaliz<strong>at</strong>ion. Here, φ n1 (r 1 ) and φ n2 (r 2 ) may or maynot be eigenst<strong>at</strong>es <strong>of</strong> the one-particle Schrödinger equ<strong>at</strong>ion. The requirements on them areth<strong>at</strong> they are complete and orthonormal. In the above, the n 1 = n 2 terms vanish becauseˆbnˆbn = 0. It can be shown easily th<strong>at</strong>ˆψ (r 1 , r 2 ) |l, m〉 = ˆψ (r 1 , r 2 ) ˆb † l ˆb † m |0〉 = 1 √2[φ l (r 1 ) φ m (r 2 ) − φ l (r 2 ) φ m (r 1 )] |0〉= φ lm (r 1 , r 2 )|0〉 (13.3.7)where φ lm (r 1 , r 2 ) is a two-particle wave function s<strong>at</strong>isfying the symmetry requirements forfermion particles. It is to be noted th<strong>at</strong> the above is zero when r 1 = r 2 . It is a consequence<strong>of</strong> the field oper<strong>at</strong>or oper<strong>at</strong>ing on the two-particle Fock st<strong>at</strong>e, and the commut<strong>at</strong>ion rel<strong>at</strong>ionsfor fermions. A three-particle field oper<strong>at</strong>or is defined to beˆψ (r 1 , r 2 , r 3 ) = √ 1 ∑ˆbn3ˆbn2ˆbn1 φ n1 (r 1 ) φ n2 (r 2 ) φ n3 (r 3 ) (13.3.8)3!n 1,n 2,n 3


Schrödinger Wave Fields 177In the above, none <strong>of</strong> the n 1 , n 2 , n 3 can be repe<strong>at</strong>ed for the reasons <strong>of</strong> (13.2.15) and (13.2.16).Then, using the commut<strong>at</strong>ion rel<strong>at</strong>ions derived in the previous section, (13.2.20) to (13.2.22),it can be shown th<strong>at</strong>ˆψ (r 1 , r 2 , r 3 ) |l, m, n〉 = ˆψ (r 1 , r 2 , r 3 ) ˆb † l ˆb † mˆb † n |0〉= 1 √3![φ l (r 1 ) φ m (r 2 ) φ n (r 3 ) + φ l (r 2 ) φ m (r 3 ) φ n (r 1 )+ φ l (r 3 ) φ m (r 1 ) φ n (r 2 ) − φ l (r 1 ) φ m (r 3 ) φ n (r 2 )−φ l (r 2 ) φ m (r 1 ) φ n (r 3 ) − φ l (r 3 ) φ m (r 2 ) φ n (r 1 )] |0〉= φ lmn (r 1 , r 2 , r 3 )|0〉 (13.3.9)where φ lmn (r 1 , r 2 , r 3 ) is the three-particle wave function with the requisite symmetry requirements.In general, an N-particle field oper<strong>at</strong>or isˆψ (r 1 , r 2 , · · · , r N ) = 1 √N!∑n 1,n 2,··· ,n NˆbnN · · · ˆb n2ˆbn1 φ n1 (r 1 ) φ n2 (r 2 ) · · · φ nN (r N ) (13.3.10)In the above sum, only terms where n 1 , n 2 , · · · , n N are distinct are contributing for the samereason given in (13.3.8). It is seen th<strong>at</strong>ˆψ NP (r 1 , r 2 , · · · , r N ) = 1 √N!ˆψ1P (r N ) · · · ˆψ 1P (r 2 ) ˆψ 1P (r 1 ) (13.3.11)where ˆψ 1P is a one-particle field oper<strong>at</strong>or while ˆψ NP is the N-particle field oper<strong>at</strong>or. However,the used <strong>of</strong> such factorized form has to be handled with caution, as each one-particle oper<strong>at</strong>orhas to oper<strong>at</strong>e on distinct modes in order for the product <strong>of</strong> factors to be non-zero.A general Fock st<strong>at</strong>e for N fermion particles can be written as|Φ N 〉 = ˆb † n 1ˆb† n2ˆb† n3 · · · ˆb † n N|0〉 = |n 1 , n 2 , n 3 , · · · , n N 〉 (13.3.12)In the above, the st<strong>at</strong>e vector changes sign when any two <strong>of</strong> the annihil<strong>at</strong>ion oper<strong>at</strong>ors swapposition, and th<strong>at</strong> the st<strong>at</strong>e vector is zero if any two <strong>of</strong> the modes are identical.When the Fock st<strong>at</strong>e is oper<strong>at</strong>ed on by the N-particle field oper<strong>at</strong>or, the coordin<strong>at</strong>e spacerepresent<strong>at</strong>ion <strong>of</strong> the N-particle fermion st<strong>at</strong>e is obtained as shown by (13.3.7) and (13.3.9).This approach avoids the cumbersome use <strong>of</strong> Sl<strong>at</strong>er determinant in the book-keeping <strong>of</strong> theN-particle fermion st<strong>at</strong>es. In generalˆψ NP (r 1 , r 2 , · · · , r N ) |Φ N 〉 = φ N (r 1 , r 2 , · · · , r N ) |0〉 (13.3.13)The above is an elegant and compact way to express an N-particle fermion wave function.13.4 Similarity TransformIt is prudent to note from the above th<strong>at</strong> one can change between coordin<strong>at</strong>e space basisand Fock space basis via the algebra shown. This allows us to effect a change <strong>of</strong> basis for a


178 Quantum Mechanics Made Simplecomplex quantum system. Some quantum systems, when expressed in the Fock space basis, isa lot simpler than the coordin<strong>at</strong>e space basis. This was first shown for electromagnetic field.It can be shown for other quantum fields. Hence, it is worthwhile to review the m<strong>at</strong>hem<strong>at</strong>ics<strong>of</strong> similarity transform next.We can view the change <strong>of</strong> basis as a similarity transform when we change from eigenfunctionspace or coordin<strong>at</strong>e space represent<strong>at</strong>ion to Fock space represent<strong>at</strong>ion. Given a m<strong>at</strong>rixequ<strong>at</strong>ionwe can define a new represent<strong>at</strong>ion for the unknowns asA · x = λx (13.4.1)x = S · y (13.4.2)In the above, S is unitary since the length <strong>of</strong> the vectors does not change. The above equ<strong>at</strong>ionbecomesMultiplying the above by S † , we haveA · S · y = λS · y (13.4.3)S † · A · S · y = λS † · S · y = λy (13.4.4)where the unitary property <strong>of</strong> the S oper<strong>at</strong>or has been used. Hence,whereEqu<strong>at</strong>ion (13.4.6) is in the form <strong>of</strong> a similarity transform.A s · y = λy (13.4.5)A s = S † · A · S (13.4.6)13.5 Additive One-Particle Oper<strong>at</strong>orWe can consider a simple identical many particle Hamiltonian where the particles do notinteract with each other, say via the Coulomb potential. For instance, for fermions, the onlyway they interact is via Pauli’s exclusion principle. An example <strong>of</strong> such a Hamiltonian for Nparticles, in coordin<strong>at</strong>e space represent<strong>at</strong>ion, isĤ r =N∑Ĥ ri (13.5.1)i=1whereĤ ri = − 22m ∇2 i + V (r i ) (13.5.2)


Schrödinger Wave Fields 179As an example, consider the three particle case. The three-particle wavefunction in coordin<strong>at</strong>erepresent<strong>at</strong>ion isφ lmn (r 1 , r 2 , r 3 ) = 1 √3![φ l (r 1 ) φ m (r 2 ) φ n (r 3 ) + φ l (r 2 ) φ m (r 3 ) φ n (r 1 )+ φ l (r 3 ) φ m (r 1 ) φ n (r 2 ) − φ l (r 1 ) φ m (r 3 ) φ n (r 2 )−φ l (r 2 ) φ m (r 1 ) φ n (r 3 ) − φ l (r 3 ) φ m (r 2 ) φ n (r 1 )] (13.5.3)If the eigenmodes above are the eigenmodes <strong>of</strong> the Ĥri oper<strong>at</strong>or, thenĤ r φ lmn (r 1 , r 2 , r 3 ) =N∑Ĥ ri φ lmn (r 1 , r 2 , r 3 ) = (E l + E m + E n )φ lmn (r 1 , r 2 , r 3 ) (13.5.4)i=1The above approach gets unwieldy as the number <strong>of</strong> particles increases. However, if we wereto test the above equ<strong>at</strong>ion with φ ∗ l ′ m ′ n ′(r 1, r 2 , r 3 ) and integr<strong>at</strong>e over space, we have∫dr 1 dr 2 dr 3 φ ∗ lmn(r 1 , r 2 , r 3 )Ĥrφ lmn (r 1 , r 2 , r 3 ) = (E l + E m + E n )δ l′ lδ m′ mδ n′ n (13.5.5)The above indic<strong>at</strong>es th<strong>at</strong> the m<strong>at</strong>rix represent<strong>at</strong>ion <strong>of</strong> Ĥr is diagonal and very simple despitethe complic<strong>at</strong>ed n<strong>at</strong>ure <strong>of</strong> the wave functions. It signals an altern<strong>at</strong>ive simpler represent<strong>at</strong>ion<strong>of</strong> the N-particle Hamilton. This simplic<strong>at</strong>ion can be achieved by the change <strong>of</strong> basis usingthe identity in (13.3.9) to obtain th<strong>at</strong>〈n ′ , m ′ , l ′ |Ĥ|l, m, n〉 = (E l + E m + E n )δ l′ lδ m′ mδ n′ n = 〈n ′ , m ′ , l ′ |(E l + E m + E n )|l, m, n〉(13.5.6)where∫Ĥ =dr 1 dr 2 dr 3 ˆψ† (r 1 , r 2 , r 3 )Ĥr ˆψ(r 1 , r 2 , r 3 ) (13.5.7)By inspection, if we letĤ = ∑ pE pˆb† pˆbp (13.5.8)it will have the same m<strong>at</strong>rix represent<strong>at</strong>ion as (13.5.6) above. But we can also derive theabove by a lengthier exercise starting with (13.5.7).In general, for a simpler approach, we can perform a change <strong>of</strong> basis or similarity transformon the original Hamiltonian. The new Hamiltonian is∫Ĥ =N∑i=1ˆψ † NP (r 1, r 2 , · · · , r N ) Ĥri ˆψ NP (r 1 , r 2 , · · · , r N ) dr 1 dr 2 · · · dr N (13.5.9)The oper<strong>at</strong>or above now acts on a vector in the Fock space, and transforms it to anothervector in the same space.


180 Quantum Mechanics Made SimpleEqu<strong>at</strong>ion (13.5.9) can be written, after using (13.3.11), asĤ = 1 N!N∑∫i=1[∫ˆψ † 1P (r N ) · · · ∼ · · · ˆψ † 1P (r 1)]dr i ˆψ†1P (r i) Ĥri ˆψ 1P (r i )ˆψ 1P (r 1 ) · · · ∼ · · · ˆψ 1P (r N ) dr 1 · · · ∼ · · · dr N (13.5.10)where the ∼ sign implies th<strong>at</strong> r i is excluded from the sequence. We can move ˆψ † 1P (r i) andˆψ 1P (r i ) as there are an even number <strong>of</strong> sign change as we anti-commute the field oper<strong>at</strong>orsinward.13.5.1 Three-Particle CaseIt is quite complex to sort out the algebra <strong>of</strong> the above system. Much insight, however,can be gotten by studying a simpler three-particle case. In this case, a typical term <strong>of</strong> thetransformed Hamiltonian isĤ = 1 ∫[∫]ˆψ † 1P3!(r 3) ˆψ † 1P (r 2) dr 1 ˆψ†1P (r 1) Ĥr1 ˆψ 1P (r 1 )ˆψ 1P (r 2 ) ˆψ 1P (r 3 ) dr 2 dr 3 + · · · (13.5.11)where the + · · · above refers to other two terms where r i = r 2 and r i = r 3 , and r i here refersto the coordin<strong>at</strong>es for the term inside the square brackets.The space-dependent parts involving φ ni (r i ) in the field oper<strong>at</strong>ors can be grouped togetherand the sp<strong>at</strong>ial integr<strong>at</strong>ions can be performed first. These modes are orthonormal giving riseto δ n2n ′ and δ 2 n 3n ′ which can be used to reduce double summ<strong>at</strong>ions into single summ<strong>at</strong>ions.3Finally, one obtainsĤ = 1 6⎡∑ˆb† n3ˆb† n2⎣ ∑n 2,n 3ˆb†n ′n ′ 11 ,n1H (1)n ′ 1 ,n1ˆbn1 ⎤⎦ ˆbn2ˆbn3 + · · · (13.5.12)The general case where φ n1 (r 1 ) is not an eigenst<strong>at</strong>e <strong>of</strong> the Hamiltonian Ĥr1 is assumed here.Hence,H (1)n ′ 1 ,n1 = 〈φ n ′ 1 |Ĥr1|φ n1 〉 (13.5.13)In the event th<strong>at</strong> φ n1 (r 1 ) is an eigenst<strong>at</strong>e <strong>of</strong> Ĥr1, the above becomesH (1)n ′ 1 ,n1 = E n 1δ n ′1 n 1(13.5.14)To evalu<strong>at</strong>e the action <strong>of</strong> the above oper<strong>at</strong>or in (13.5.12) on the three-fermion-particlest<strong>at</strong>e |l, m, n〉 = ˆb † l ˆb † mˆb † n|0〉, one needs to first show th<strong>at</strong>ˆbn3 |l, m, n〉 = ˆb ˆb † n3ˆb†l mˆb † n|0〉(= δ ˆb)†nn3ˆb†l m + δ ln3ˆb† mˆb† n + δ mn3ˆb† nˆb†l|0〉 (13.5.15)


Schrödinger Wave Fields 181lb n1 b n2 b n3m|l, m, n >nFigure 13.1: The three annihil<strong>at</strong>ion oper<strong>at</strong>ors ˆb n1ˆbn2ˆbn3 acting on the three-particle st<strong>at</strong>e|l, m, n〉 produces six threads th<strong>at</strong> eventually gives rise to six terms.Next, one can show th<strong>at</strong>Finally,ˆbn2ˆbn3 |l, m, n〉 =(δ ln3 δ mn2ˆb† n − δ ln3ˆb† m δ nn2 − δ ln2 δ mn3ˆb† n+ˆb † l δ mn 3δ nn2 + δ ln2ˆb† m δ nn3 − ˆb † l δ mn 2δ nn3)|0〉 (13.5.16)ˆbn1ˆbn2ˆbn3 |l, m, n〉 = (δ ln3 δ mn2 δ nn1 − δ ln3 δ mn1 δ nn2 − δ ln2 δ mn3 δ nn1+δ ln1 δ mn3 δ nn2 + δ ln2 δ mn1 δ nn3 − δ ln1 δ mn2 δ nn3 ) |0〉 (13.5.17)A total <strong>of</strong> six terms is found. In the above ˆb n3 , oper<strong>at</strong>ing on the three-particle st<strong>at</strong>e, producesthree terms each <strong>of</strong> which is a two-particle st<strong>at</strong>e as shown in Figure 13.1. Next, ˆb n2 oper<strong>at</strong>ingon them will produce two terms per each term <strong>of</strong> the two-particle st<strong>at</strong>e, producing a net <strong>of</strong>six terms <strong>of</strong> one-particle st<strong>at</strong>es. The final oper<strong>at</strong>ion ˆb n1 produces one term for each particle.In general, if we start with an N-particle st<strong>at</strong>e, the oper<strong>at</strong>ion <strong>of</strong> ˆb n1 · · · ˆb nN on it will produceN! terms.The ˆb † oper<strong>at</strong>ors in (13.5.12) anti-commute with each other. Hence, ˆb † n 3ˆb† n2can be movedto the right <strong>of</strong> the H (1)nelement, together with the summ<strong>at</strong>ions. Therefore, we finally need′1 ,n1to evalu<strong>at</strong>e∑ˆb† n3ˆb† n2ˆbn1ˆbn2ˆbn3 |l, m, n〉 (13.5.18)n 2,n 3The above can be shown to evalu<strong>at</strong>e to∑(ˆb† n3ˆb† n2ˆbn1ˆbn2ˆbn3 |l, m, n〉 = 2 δ ˆb)†nn1ˆb†l m + δ ln1ˆb† mˆb† n + δ mn1ˆb† nˆb†l|0〉n 2,n 3= 2ˆb n1ˆb†l ˆb † mˆb † n|0〉 = 2ˆb n1 |l, m, n〉 (13.5.19)The first equality above can be obtained by using (13.5.17). The second equality follows from


182 Quantum Mechanics Made Simple(13.5.15). The above can be generalized to the N-particle case yielding 2∑n 2,··· ,n Nˆb† nN · · · ˆb † n 2ˆbn1ˆbn2 · · · ˆb nN |l 1 , · · · , l N 〉 = (N − 1)!ˆb n1 |l 1 , · · · , l N 〉 (13.5.20)Consequently,Ĥ|l, m, n〉 = 1 3∑ˆb†nH (1)′n ′ 1 n ′ ,n1ˆbn1 |l, m, n〉 + · · · (13.5.21)11 ,n1The other two terms indic<strong>at</strong>ed by the + · · · would contribute to exactly the same expressionas they are from indistinguishable particles. Finally, we haveĤ|l, m, n〉 = ∑H (1)n|l, m, n〉 (13.5.22)′ ,n1ˆbn11ˆb†n ′n ′ 11 ,n1In general, the additive one-particle oper<strong>at</strong>or in Fock space isĤ = ∑ n ′ ,nˆb†n ′H(1) n ′ ,nˆb n (13.5.<strong>23</strong>)The above can be used for the N-particle case as long as they are indistinguishable. It canbe shown to yield the same m<strong>at</strong>rix represent<strong>at</strong>ion using Fock st<strong>at</strong>e vectors, compared to thecase in coordin<strong>at</strong>e space represent<strong>at</strong>ion. It can be easily valid<strong>at</strong>ed using the two-particle orthree-particle st<strong>at</strong>e vectors.For the case when the one-particle eigenst<strong>at</strong>es are also the eigenst<strong>at</strong>es <strong>of</strong> the Ĥr ioper<strong>at</strong>or,the above m<strong>at</strong>rix becomes diagonal yielding,Ĥ = ∑ nE nˆb† nˆbn (13.5.24)The above means th<strong>at</strong> if we have an N-particle fermion field, it can be represented by thephysics <strong>of</strong> one-particle Hamiltonian if the eigenst<strong>at</strong>es chosen for the similarity transform arealso the eigenst<strong>at</strong>es <strong>of</strong> the one-particle Hamiltonian. Also, the above Hamiltonian is verysimilar to the Hamiltonian for photons derived in the previous chapter, except th<strong>at</strong> photonsare bosons and the above deriv<strong>at</strong>ion is for fermions. A similar deriv<strong>at</strong>ion for bosons showsth<strong>at</strong> the Hamiltonian is similar to the above.The above Hamiltonian (13.5.24) does not distinguish between one particle or N particles.This difference <strong>of</strong> the quantum systems is expressed by the Fock st<strong>at</strong>es <strong>of</strong> the particles.13.6 Additive Two-Particle Oper<strong>at</strong>orIn general, the N particles in a Schrödinger wave field will interact with each other. They mayinteract pair-wise, for instance, via Coulomb potential other than just the Pauli’s exclusion2 For the fermion case, since n 1 , · · · , n N are distinct, we can anti-commute the oper<strong>at</strong>ors to arrange themin pairs <strong>of</strong> ˆb †ˆb and arrive <strong>at</strong> (13.5.19) and (13.5.20) more succinctly.


Schrödinger Wave Fields 183principle. An example <strong>of</strong> an additive two particle oper<strong>at</strong>or in coordin<strong>at</strong>e space isV (r 1 , r 2 ) =e 24πɛ 0 |r 1 − r 2 |(13.6.1)This will be part <strong>of</strong> a two-particle HamiltonianĤ r (r 1 , r 2 ) = − 2 (∇22m 1 + ∇ 2 )2 + V (r1 , r 2 ) (13.6.2)When N-particle Hamiltonian is considered, the Coulomb interaction appears asV = ∑ i


184 Quantum Mechanics Made SimpleThe above needs to oper<strong>at</strong>e on an N-particle Fock st<strong>at</strong>e, namely,ˆV |l 1 , · · · , l N 〉 = 12N!∑ˆb† nN · · · ˆb † n 3n 3,··· ,n N⎡∑⎣nˆb† ′2 n ′ 1n ′ 1 ,n1,n′ 2 ,n2 ˆb†The right-most term th<strong>at</strong> acts on the Fock st<strong>at</strong>e is <strong>of</strong> the formV (1,2)n ′ 1 ,n′ 2 ,n1,n2ˆbn1ˆbn2 ⎤⎦ ˆbn3 · · · ˆb nN |l 1 , · · · , l N 〉 + · · ·(13.6.8)ˆbn1ˆbn2 · · · ˆb nN |l 1 , · · · , l N 〉 (13.6.9)The above yields N! terms similar to (13.5.17). But ˆb † n N · · · ˆb † n 3commutes with ˆb † n 2ˆb† n1in(13.6.6) above. We can then move ˆb † n N · · · ˆb † n 3to the right <strong>of</strong> V (1,2)n ′ 1 ,n′ 2 ,n1,n2.Then we need toevalu<strong>at</strong>e∑n 3,··· ,n Nˆb† nN · · · ˆb † n 3ˆbn1ˆbn2 · · · ˆb nN |l 1 , · · · , l N 〉 = (N − 2)!ˆb n1ˆbn2 |l 1 , · · · , l N 〉 (13.6.10)The above equality can be proved by induction from the three-particle case or from equ<strong>at</strong>ions(13.5.19) and (13.5.20). As a result,ˆV |l 1 , · · · , l N 〉 =12N(N − 1)∑ˆb†n ′n ′ 21 ,n1,n′ 2 ,n2ˆb†nV (1,2)′1 n|l ′1 ,n′ ,n1,n2ˆbn1ˆbn2 1 , · · · , l N 〉 + · · · (13.6.11)2There are N(N − 1) terms <strong>of</strong> the similar kind in the + · · · above. Summing them up, we haveˆV |l 1 , · · · , l N 〉 = 1 2∑ˆb†nˆb† ′n ′ 2 nV (1,2)′1 n|l ′1 ,n′ ,n1,n2ˆbn1ˆbn2 1 , · · · , l N 〉 (13.6.12)21 ,n1,n′ 2 ,n2In general, the additive two-particle oper<strong>at</strong>or in an N-particle fermion field is given byˆV = 1 2∑ˆb†nˆb† ′n ′ 2 nV (1,2)′1 n ′ 1 ,n′ ,n1,n2ˆbn1ˆbn2 (13.6.13)21 ,n1,n′ 2 ,n2The above can be easily valid<strong>at</strong>ed to produce the same m<strong>at</strong>rix represent<strong>at</strong>ion compared tothe coordin<strong>at</strong>e represent<strong>at</strong>ion using the two-particle or three-particle st<strong>at</strong>e vectors.The above deriv<strong>at</strong>ions for fermions can be repe<strong>at</strong>ed for the boson case. In this case, wewill have a field <strong>of</strong> bosons. It is pleasing th<strong>at</strong> by starting out with the N-particle coordin<strong>at</strong>espace represent<strong>at</strong>ion <strong>of</strong> the Schrödinger equ<strong>at</strong>ion, one arrives <strong>at</strong> a much simpler Fockspace represent<strong>at</strong>ion <strong>of</strong> the same equ<strong>at</strong>ion. It allows one to tre<strong>at</strong> N-particle problems moresuccinctly using such a represent<strong>at</strong>ion.


Schrödinger Wave Fields 18513.7 More on Field Oper<strong>at</strong>orsThe above shows th<strong>at</strong> the one-particle additive oper<strong>at</strong>or is independent <strong>of</strong> the number <strong>of</strong>particles. So wh<strong>at</strong> distinguishes one-particle quantum system from a two-particle one is inthe Fock st<strong>at</strong>e: the first one will have a Fock st<strong>at</strong>e for one particle while the l<strong>at</strong>ter will havea Fock st<strong>at</strong>e for two particles.A general one-particle Fock st<strong>at</strong>e is|ψ〉 = ∑ µc µˆb† µ |0〉 (13.7.1)When the one-particle field oper<strong>at</strong>or acts on the above, it projects the coordin<strong>at</strong>e spacerepresent<strong>at</strong>ion <strong>of</strong> the one-particle wavefunction. Namely,ˆψ(r)|ψ〉 = ∑ µc µ φ m (r)|0〉 = ψ(r)|0〉 (13.7.2)The above one-particle st<strong>at</strong>e is in a linear superposition <strong>of</strong> different eigenst<strong>at</strong>es. Hence, it canconstitute a wave packet. 3Also, from (13.3.1), we can easily show th<strong>at</strong>∫ˆb† µ = φ µ (r) ˆψ † (r)dr (13.7.3)On combining with (13.7.3), we have|ψ〉 = ∑ ∫c µ φ µ (r) ˆψ † (r)dr|0〉µ∫ [ ]∑= c µ φ µ (r) ˆψ † (r)dr|0〉µ∫= f(r) ˆψ † (r)dr|0〉 (13.7.4)Therefore, a general one-particle st<strong>at</strong>e can be written as a linear superposition <strong>of</strong> the oneparticlefield oper<strong>at</strong>or ˆψ † (r) weighted by f(r). It is clear th<strong>at</strong> f(r) s<strong>at</strong>isfies the one-particleSchrödinger equ<strong>at</strong>ion.For a general two-particle st<strong>at</strong>e in Fock space,|ψ〉 = ∑µ 1,µ 2c µ1,µ 2ˆb† µ1ˆb† µ2|0〉 (13.7.5)upon substituting (13.7.3) into the above, we have∫ ( )∑|ψ〉 = φ µ1 (r 1 )φ µ2 (r 2 )c µ1,µ 2ˆψ † (r 1 ) ˆψ † (r 2 )dr 1 dr 2 |0〉µ 1µ 2∫= f(r 1 , r 2 ) ˆψ † (r 1 ) ˆψ † (r 2 )dr 1 dr 2 |0〉 (13.7.6)3 See the appendix for a discussion <strong>of</strong> wave packets.


186 Quantum Mechanics Made Simplewheref(r 1 , r 2 ) = ∑µ 1,µ 2φ µ1 (r 1 )φ µ2 (r 2 )c µ1,µ 2. (13.7.7)For fermions, it can be easily shown, using the commut<strong>at</strong>ion rel<strong>at</strong>ions for the field oper<strong>at</strong>ors,th<strong>at</strong>f (r 1 , r 2 ) = −f (r 2 , r 1 ) (13.7.8)The above implies th<strong>at</strong> c µ1,µ 2= −c µ2,µ 1. The above s<strong>at</strong>isfies the two-particle Schrödingerequ<strong>at</strong>ion. It can be generalized to N particles, yielding∫|ψ〉 = f (r 1 , r 2 , · · · , r N ) ˆψ † (r 1 ) ˆψ † (r 2 ) · · · ˆψ † (r N ) dr 1 dr 2 · · · dr N |0〉 (13.7.9)13.8 Boson Wave FieldThe results for fermion wave field in the previous section can also be derived for boson wavefield. The difference is th<strong>at</strong> more than one boson can occupy a given st<strong>at</strong>e: The Pauli exclusionprinciple does not apply to bosons. A boson st<strong>at</strong>e involving two modes can be denoted as:|ψ〉 = |n m , n v 〉 (13.8.1)where there are n m particles in mode m, and n v particles in mode v. Since ordering isunimportant, it is necessary th<strong>at</strong>A boson cre<strong>at</strong>ion oper<strong>at</strong>or can be defined such th<strong>at</strong>|n m , n v 〉 = |n v , n m 〉 (13.8.2)â † k |n k, n m , n v 〉 = C nk +1|n k + 1, n m , n v 〉 (13.8.3)It raises the number <strong>of</strong> bosons in st<strong>at</strong>e k by one. By definition,â † l ↠k |n k, n l , · · · 〉 = C nk +1C nl +1|n k + 1, n l + 1, · · · 〉 (13.8.4)â † k↠l |n l, n k , · · · 〉 = C nl +1C nk +1|n l + 1, n k + 1, · · · 〉 = â † l ↠k |n k, n l , · · · 〉 (13.8.5)The last equality follows since ordering is unimportant. Consequently,â † k↠l − ↠l ↠k = 0 (13.8.6)Similarly, one can define annihil<strong>at</strong>ion oper<strong>at</strong>ors for bosons th<strong>at</strong> have the opposite effect asthe cre<strong>at</strong>ion oper<strong>at</strong>ors. Hence, going through a similar process, one also hasFrom definition,â k â l − â l â k = 0 (13.8.7)â † |n〉 = C n+1 |n + 1〉 (13.8.8)â|n〉 = B n |n − 1〉 (13.8.9)


Schrödinger Wave Fields 187Furthermore, from (13.8.9), we have〈n + 1|â † |n〉 = C n+1 , Cn+1 ∗ = 〈n|â|n + 1〉 = B n+1 (13.8.10)Then, we can show th<strong>at</strong>â † l âk|n k , n l , · · · 〉 = Cn ∗ kC nl +1|n k − 1, n l + 1, · · · 〉 (13.8.11)â k â † l |n l, n k , · · · 〉 = Cn ∗ l +1C nk |n l + 1, n k − 1, · · · 〉 (13.8.12)If C n is a real number, then when l ≠ k,â k â † l − ↠l âk = 0 (13.8.13)For the case when l = k, we can drop the subscripts k and l and look <strong>at</strong>â † â|n〉 = Cn|n〉 2 (13.8.14)If C 2 n+1 − C 2 n = 1, thenââ † |n〉 = C 2 n+1|n〉 (13.8.15)ââ † − â † â = Î (13.8.16)Then above implies th<strong>at</strong> C 2 n = n, since C 0 =0. Also, C n = √ n. In summary, for bosons,â † k↠l − ↠l ↠k = 0 (13.8.17)â k â l − â l â k = 0 (13.8.18)â k â † l − ↠l âk = Îδ kl (13.8.19)The above is derived without resorting to the use <strong>of</strong> the quantum harmonic oscill<strong>at</strong>or, butthe assumption th<strong>at</strong> C n is real and th<strong>at</strong> C 2 n = n. One can further conclude th<strong>at</strong>â † â|n〉 = n|n〉 (13.8.20)which is the number oper<strong>at</strong>or, which has been derived using a different approach.13.9 Boson Field Oper<strong>at</strong>orsSimilar to fermions, we can define field oper<strong>at</strong>ors for bosons, so th<strong>at</strong>ˆψ(r) = ∑ jâ j φ j (r) (13.9.1)In gereral, for N particles,ˆψ(r 1 , r 2 , · · · , r N ) = √ 1 ∑â nN · · · â n1 φ n1 (r 1 )φ n2 (r 2 ) · · · φ nN (r N ) (13.9.2)N!n 1,n 2,··· ,n N


188 Quantum Mechanics Made SimpleWe can illustr<strong>at</strong>e with a three-particle field oper<strong>at</strong>or:ˆψ(r 1 , r 2 , r 3 ) = 1 √3!∑n 1,n 2,n 3â n3 â n3 â n1ˆφ(r1 )φ(r 1 )φ(r 2 )φ(r 3 ) (13.9.3)When this oper<strong>at</strong>es on |1 l , 1 m , 1 n 〉 where the three particles are in different modes, we get(13.3.9) except th<strong>at</strong> − signs are now replaced by + signs. If we have a st<strong>at</strong>e denoted bythenˆψ(r 1 , r 2 , r 3 )|2 l , 1 m 〉|2 l , 1 m 〉 = 1 √2!(â † l) 2am |0〉 (13.9.4)= 1 √2!3![2φ l (r 1 )φ m (r 2 )φ l (r 3 ) + 2φ l (r 2 )φ m (r 3 )φ l (r 1 ) + 2φ l (r 3 )φ m (r 1 )φ l (r 2 )] |0〉 (13.9.5)It can be shown th<strong>at</strong> the above is the correctly normalized wavefunction. If the st<strong>at</strong>e isthen|3 l 〉 = 1 √3!(â † l) 3|0〉 (13.9.6)ˆψ(r 1 , r 2 , r 3 )|3 l 〉 = 6 3! φ l(r 1 )φ l (r 2 )φ l (r 3 )|0〉 (13.9.7)The above is in fact normalized.13.10 Additive One-Particle Oper<strong>at</strong>orIn this case, we can illustr<strong>at</strong>e with the three-particle case for bosons, arriving <strong>at</strong> an expressionsimilar to (13.5.12)whereĤ = 1 6⎡∑â † n 3 â † ⎣ ∑n 2n 2,n 3n ′ 1 ,n1 â † n ′ 1H (1)n ′ 1 ,n1ân 1⎤⎦ â n2 â n3 + · · · (13.10.1)H (1)n ′ 1 ,n1 = 〈φ n ′ 1 |Ĥr 1|φ n1 〉 (13.10.2)When the three particles are in the |1 l , 1 m , 1 n 〉 st<strong>at</strong>e, (13.5.17) follows for bosons except th<strong>at</strong>we replace − signs with + signs. Equ<strong>at</strong>ion (13.5.19) follows similarly for bosons, so does(13.5.21) and (13.5.22). When two <strong>of</strong> the particles are in the same st<strong>at</strong>e as (13.9.4), we haveâ n1 â n2 â n3 |2 l , 1 m 〉 = √ 2!(δ ln3 δ mn2 δ ln1 + δ ln3 δ mn1 δ ln2 + δ ln2 δ mn3 δ ln1 )|0〉 (13.10.3)


Schrödinger Wave Fields 189Similar to (13.5.19), we need to evalu<strong>at</strong>e∑â † n 3 â † n 2 â n1 â n2 â n3 |2 l , 1 m 〉n 2,n 3= √ ()2! 2δ ln1 â † l ↠m + δ mn1 â † l ↠l|0〉= √ 2!â n1(â † l) 2↠m |0〉 = (2!)â n1 |2 l , 1 m 〉 (13.10.4)The first equality is established using (13.10.3), while the second equality is obtained byworking backward using commut<strong>at</strong>ion rel<strong>at</strong>ions to expand â n1(â † l) 2↠m |0〉. When the threeparticles are in the same st<strong>at</strong>e as indic<strong>at</strong>ed by (13.9.6), we haveandIn general,∑n 2n 3â † n 3 â † n 2 â n1 â n2 â n3 |3 l 〉= (3!) 3/2 δ ln1(â † l∑â n1 â n2 â n3 |3 l 〉 = (3!) 3/2 δ ln3 δ ln2 δ ln1 |0〉 (13.10.5)) 2 1) 3|0〉 =3 (3!)3/2 â n1(â † l |0〉 = (2!)ân1 |3 l 〉 (13.10.6)n 2,n 3â † n 3 â † n 2 â n1 â n2 â n3 |ψ 3p 〉 = (2!) â n1 |ψ 3p 〉 (13.10.7)where |ψ 3p 〉 is the three-particle Fock st<strong>at</strong>e <strong>of</strong> either|1 l , 1 m , 1 n 〉 , |2 l , 1 m 〉 , |3 l 〉 (13.10.8)or other combin<strong>at</strong>ions. The pro<strong>of</strong> for the rest follows th<strong>at</strong> <strong>of</strong> fermion particles <strong>of</strong> (13.5.21) to(13.5.24). In general, the Hamiltonian for N bosons is given byĤ = ∑ n ′ ,nH n′ ,nâ † n ′â n (13.10.9)For the diagonal case, it becomesĤ = ∑ nE n â † nâ n (13.10.10)13.11 The Difference between Boson Field and PhotonFieldThe above is the Hamiltonian for a boson field or gas in Fock space. It looks strikingly similarto the Hamiltonian for photon field or electromagnetic field. However, it is to be noted th<strong>at</strong>a boson field is quite different from the field <strong>of</strong> N photons.


190 Quantum Mechanics Made Simple1. A coordin<strong>at</strong>e r i can be associ<strong>at</strong>ed with every boson particle, but not with every photon;2. All photons share the same electromagnetics field function in its field oper<strong>at</strong>or represent<strong>at</strong>ion,but N bosons will have product <strong>of</strong> N one-particle field oper<strong>at</strong>or;3. The Hamiltonian <strong>of</strong> a boson gas has no zero-point energy, but for photons, there existsthe zero-point energy <strong>of</strong> 1 2 ω.A photon should be thought <strong>of</strong> as a quantized energy packet due to coupled quantumharmonic oscill<strong>at</strong>ion <strong>of</strong> “space” or “vacuum”, similar to the case <strong>of</strong> a phonon, which is coupledquantum harmonic oscill<strong>at</strong>ion <strong>of</strong> crystalline l<strong>at</strong>tice.


Chapter 14Interaction <strong>of</strong> Different Particles14.1 IntroductionIn the previous chapter, it was shown th<strong>at</strong> the Hamiltonian <strong>of</strong> a many particle system canbe written succinctly using Fock space represent<strong>at</strong>ion. This opens up the study <strong>of</strong> complexsystems, such as excitons th<strong>at</strong> involve the interaction <strong>of</strong> electron-hole pair, as well assuperconductivity th<strong>at</strong> involves the interaction <strong>of</strong> electrons with phonons. In this chapter,we will study the interaction <strong>of</strong> the electron <strong>of</strong> an <strong>at</strong>om with photons <strong>of</strong> a cavity using themany-particle formalism.14.2 Interaction <strong>of</strong> ParticlesSay if we have two quantum systems th<strong>at</strong> are initially non-interacting, incoherent, and independent,their individual quantum systems can be described by eigenst<strong>at</strong>es th<strong>at</strong> s<strong>at</strong>isfyĤ 1 |ψ 1 〉 = E 1 |ψ 1 〉 (14.2.1)Ĥ 2 |ψ 2 〉 = E 2 |ψ 2 〉 (14.2.2)Even though the two systems are entirely non-interacting with each other, nevertheless, wecan combine the two systems together and write)Ĥ 0 |ψ〉 =(Ĥ1 + Ĥ2 |ψ 1 〉 |ψ 2 〉 = (E 1 + E 2 ) |ψ 1 〉 |ψ 2 〉 = E |ψ〉 (14.2.3)where |ψ 1 〉 |ψ 2 〉 represents a st<strong>at</strong>e in the direct product space. It can be used to representthe eigenst<strong>at</strong>es <strong>of</strong> the combined system. Equ<strong>at</strong>ion (14.2.3) is entirely equivalent to (14.2.1)and (14.2.2). In the above, Ĥ 1 acts only on |ψ 1 〉 and Ĥ2 acts on |ψ 2 〉. The above allowsus to add an interaction term between system 1 and system 2, and the same direct productspace can be used to span the solution space. By using Fock space represent<strong>at</strong>ion, the aboveHamiltonians can be written in terms <strong>of</strong> annihil<strong>at</strong>ion and cre<strong>at</strong>ion oper<strong>at</strong>ors for a system191


192 Quantum Mechanics Made Simpleconsisting <strong>of</strong> fermions and bosons,Ĥ 1 = ∑ jE jˆb† jˆb j ,Ĥ 2 = ∑ λω λ â † λâλ (14.2.4)where Ĥ1 can be the Hamiltonian for electrons, and Ĥ2 can be the Hamiltonian for photonswhere the 1 2 ω λ term or the zero-point energy has been ignored. This term will just introducea phase shift in the solution.When electric dipole interaction exists between the electron <strong>of</strong> an <strong>at</strong>om and the electromagneticfield, the interaction Hamiltonian may be written asĤ ed,r = eE · r (14.2.5)The above has been added as a perturb<strong>at</strong>ion to the unperturbed equ<strong>at</strong>ion (14.2.3) where theelectric field is tre<strong>at</strong>ed classically in a previous chapter. We have enough knowledge now totre<strong>at</strong> both E and r quantum mechanically. Here, r represents the position <strong>of</strong> the electron.For the electron in the i-th <strong>at</strong>om, the above becomesĤ ed,ri= ie ∑ λ( ) √â λ − â † ωλu λ (r i ) · r i (14.2.6)2ɛ 0We have used a quantized cavity mode electric field in (14.2.5) to arrive <strong>at</strong> the above. Also,r in coordin<strong>at</strong>e space represent<strong>at</strong>ion remains unchanged. When N <strong>at</strong>oms are present,Ĥ ed,r =N∑ie ∑ λi=1( ) √â λ − â † ωλu λ (r i ) · r i (14.2.7)2ɛ 0A similarity transform <strong>of</strong> the above Hamiltonian can be performed with the N particle fieldoper<strong>at</strong>or to yield∫Ĥ ed = ˆψ † NP (r 1, · · · , r N )Ĥed,r ˆψ NP (r 1 , · · · , r N )dr 1 · · · dr N (14.2.8)As mentioned in the previous chapter, since only one-particle interaction is involved, theabove can be transformed with the one-particle field oper<strong>at</strong>or yieldingĤ ed = ∑ )H ed,λ,j,kˆb† jˆb k(â λ − â † λ(14.2.9)j,k,λwhere√ ∫ ωH ed,λ,j,k = ie2ɛ 0drφ ∗ j (r)u λ (r) · rφ k (r) (14.2.10)In the above, j, k represents the orbital wave functions <strong>of</strong> the <strong>at</strong>omic energy st<strong>at</strong>es with energyE j and E k , respectively, and λ denotes the electromagnetic mode.


Interaction <strong>of</strong> Different Particles 19314.3 Time-Dependent Perturb<strong>at</strong>ion TheoryIn the time-dependent perturb<strong>at</strong>ion theory, we first seek the solution <strong>of</strong> the unperturbedsystem. The eigenst<strong>at</strong>e <strong>of</strong> the unperturbed system is defined asĤ 0 |N fm ; N bm 〉 = E m |N fm ; N bm 〉 (14.3.1)where |N fm ; N bm 〉 stands for a vector in the direct product space <strong>of</strong> fermions and bosons:N fm stands for the m-th st<strong>at</strong>e <strong>of</strong> the fermions, while N bm stands for the m-th st<strong>at</strong>e <strong>of</strong> thebosons.When the system is perturbed, we denote the solution |ψ〉 as a linear superposition <strong>of</strong> theeigenst<strong>at</strong>es <strong>of</strong> the unperturbed solution. Th<strong>at</strong> is|ψ〉 = ∑ mc m (t) e −iωmt |N fm ; N bm 〉 (14.3.2)where ω m = E m . The st<strong>at</strong>e |ψ〉 evolves in time according to the equ<strong>at</strong>ioni ∂ ) (Ĥ0∂t |ψ〉 = + Ĥp |ψ〉 (14.3.3)Applying the time-dependent perturb<strong>at</strong>ion method as we have done previously, and testingthe equ<strong>at</strong>ion with 〈N fq ; N bq | givesċ (1)q (t) = 1i∑mc (0)m e −i(ωm−ωq)t 〈N fq ; N bq | Ĥp |N fm ; N bm 〉 (14.3.4)If the starting st<strong>at</strong>e is assumed such th<strong>at</strong> c (0)s = 1, and all other st<strong>at</strong>es are zero, then theabove becomesċ (1)q (t) = 1i ei(ωq−ωs)t 〈N fq ; N bq | Ĥp |N fs ; N bs 〉 (14.3.5)withĤ p = Ĥed = ∑ j,k,λ)H ed,λ,j,kˆb† jˆb k(â λ − â † λ(14.3.6)14.3.1 AbsorptionThe electron in an <strong>at</strong>om is assumed to have two st<strong>at</strong>es: a lower energy st<strong>at</strong>e E 1 , and an upperenergy st<strong>at</strong>e E 2 . We assume a photon in mode λ 1 in the cavity. Then the starting st<strong>at</strong>e canbe written as|N fs ; N bs 〉 = ˆb † 1↠λ 1|0〉 (14.3.7)with energyConsequently,Ĥ p |N fs ; N bs 〉 = ∑ j,k,λE s = E 1 + ω λ1 (14.3.8))H ed,λ,j,kˆb† jˆb k(â λ − â † λˆb† 1↠λ 1|0〉 (14.3.9)


194 Quantum Mechanics Made SimpleIt can be shown th<strong>at</strong>)ˆb† jˆb k(â λ − â † λˆb† 1↠λ 1|0〉 = δ k1 δ λλ1ˆb† j |0〉 − δ k1ˆb † j↠λ↠λ 1|0〉 (14.3.10)In order for 〈N fq ; N bq | Ĥp |N fs ; N bs 〉 to be nonzero, it is necessary th<strong>at</strong> |N fq ; N bq 〉 containsˆb† j |0〉 or ˆb † j↠λ↠λ 1|0〉. If|N fq ; N bq 〉 = ˆb † j |0〉 (14.3.11)then E q = E j , andċ (1)q (t) = 1i ei(ωj−ω1−ω λ 1 )tH ed,λ,j,k δ k1 δ λλ1 〈0| ˆb jˆb† j |0〉j,k,λ(14.3.12)= 1 ∑e i(ωj−ω1−ω λ )t 1 H ed,λ1,j,1i(14.3.13)jwhere ω j = E j . The above is an oscill<strong>at</strong>ory function <strong>of</strong> t unless ω j − ω 1 − ω λ1 = 0.When this happens, ċ (1)q (t) will integr<strong>at</strong>e to a large value, indic<strong>at</strong>ing the high likelihood<strong>of</strong> transition to his eigenst<strong>at</strong>e. Therefore, for this to happen, we needE 2 − E 1 = ω λ1 (14.3.14)In this process, we start with the electron in E 1 and a photon in the cavity, and end up withthe electron in E 2 and no photon as indic<strong>at</strong>ed by (14.3.11).If we integr<strong>at</strong>e the above equ<strong>at</strong>ion, a sinc function results, which can be approxim<strong>at</strong>ed bya delta function for long integr<strong>at</strong>ion time, as was done previously. One can show th<strong>at</strong>|c (1)q | 2 ≈ 2π t 0 |H edλ1,2,1| 2 δ(E 2 − E 1 − ω λ1 ), t o → ∞ (14.3.15)One can derive the transition r<strong>at</strong>e as before to arrive <strong>at</strong>w q = 2π | H edλ 1,2,1 | 2 δ(E 2 − E 1 − ω λ1 ) (14.3.16)In the above, the other possibility is for the final st<strong>at</strong>e to bewith energyHence|N fq ; N bq 〉 = ˆb † j↠λ↠λ 1|0〉 (14.3.17)E q = E j + ω λ + ω λ1 (14.3.18)E q − E s = E j − E 1 + ω λ (14.3.19)and E q −E s = 0 in order for transition to occur or c (1)q to be large. However, (14.3.19) cannotbe zero. If E j = E 2 , E q − E s > 0 always. If E j = E 1 , then ω λ > 0. Wh<strong>at</strong> this means is th<strong>at</strong>it is not possible to start with an electron in the ground st<strong>at</strong>e and a photon in the cavity toend up with the electron in the excited st<strong>at</strong>e with the emission <strong>of</strong> a photon.


Interaction <strong>of</strong> Different Particles 19514.4 Spontaneous EmissionIn the case <strong>of</strong> spontaneous emission, we assume th<strong>at</strong> the starting st<strong>at</strong>e <strong>of</strong> the electron is inthe excited st<strong>at</strong>e with energy E 2 . Then, E s = E 2 initially andThen|N fs ; N bs 〉 = ˆb † 2 |0〉 (14.4.1)Ĥ p |N fs ; N bs 〉 = ∑ )H ed,λ,j,k ˆb† jˆb k(â λ − â † λˆb† 2 |0〉 (14.4.2)j,k,λThe above can be reduced toĤ p |N fs ; N bs 〉 = − ∑ j,k,λH ed,λ,j,k δ 2kˆb† j↠λ|0〉 (14.4.3)Therefore, for 〈N fq ; N bq |Ĥp|N fs ; N bs 〉 to be nonzero, we needwithThereforeIn order for a sizeable c (1)q , we need E j = E 1 , and then|N fq ; N bq 〉 = ˆb † j↠λ|0〉 (14.4.4)E q = E j + ω λ (14.4.5)E q − E s = E j − E 2 + ω λ (14.4.6)E 2 − E 1 = ω λ (14.4.7)Then electron starts with the excited st<strong>at</strong>e E 2 , spontaneously emits a photon, and drops tolower st<strong>at</strong>e E 1 . Then energy <strong>of</strong> the emitted photon s<strong>at</strong>isfies (14.4.7) by energy conserv<strong>at</strong>ion.14.5 Stimul<strong>at</strong>ed EmissionIn this case, the electron in an <strong>at</strong>om is in the excited st<strong>at</strong>e. The presence <strong>of</strong> a photon in thecavity stimul<strong>at</strong>es the emission <strong>of</strong> another photon from the electron. The initial st<strong>at</strong>e iswith|N fs ; N bs 〉 = ˆb † 2↠λ 1|0〉 (14.5.1)E s = E 2 + ω λ1 (14.5.2)


196 Quantum Mechanics Made SimpleWe can show th<strong>at</strong>Ĥ p |N fs ; N bs 〉 = ∑ j,k,λ= ∑ j,k,λ)H ed,λ,j,k ˆb† jˆb k(â λ − â † λˆb† 2↠λ 1|0〉 (14.5.3)(H ed,λ,j,k δ k2 δ λλ1ˆb† j − ˆb)†j↠λ↠λ 1|0〉 (14.5.4)In order for the above to transition to the final st<strong>at</strong>e, or one requires a non-zero result forit is necessary th<strong>at</strong>with energyor〈N fq ; N bq |Ĥp|N fs ; N bs 〉In other words, the only possibility is for E j = E 1 , yielding|N fq ; N bq 〉 = ˆb † j↠λ↠λ 1|0〉 (14.5.5)E q = E j + ω λ + ω λ1 (14.5.6)E q − E s = E j − E 2 + ω λ = 0 (14.5.7)E 2 − E 1 = ω λ (14.5.8)The above is just the spontaneous emission <strong>of</strong> a photon with the above energy, regardless ifwe already have a photon with energy ω λ1 in the cavity. So it is not the most interestingcase.Next we consider the case when λ = λ 1 . Then|N fq ; N bq 〉 = √ 1 ( ) 2â † λ 1|0〉 (14.5.9)2!ˆb†withConsequently,1〈N fq ; N bq |Ĥp|N fs ; N bs 〉 = H ed,λ1,j,2〈0| √ (â λ1 ) 2ˆbjˆb†2!jE q = E j + 2ω λ1 (14.5.10)j(â † λ 1) 2|0〉= √ 12!H ed,λ1,j,2〈0| √ (â λ1 ) 2ˆbj 1√2! 2!ˆb†j(â † λ 1) 2|0〉= √ 2H ed,λ1,j,2 (14.5.11)The √ 2 factor is important implying th<strong>at</strong> the transition is two times more likely to occurcompared to the previous case. This is peculiar to stimul<strong>at</strong>ed emission where the emission <strong>of</strong>a photon is enhanced by the presence <strong>of</strong> a photon <strong>of</strong> the same frequency.


Interaction <strong>of</strong> Different Particles 19714.6 Multi-photon CaseIn the multi-photon case, the transition r<strong>at</strong>e for stimul<strong>at</strong>ed emission can be shown to bew q = 2π (n λ 1+ 1)|H ed,λ1,1,2| 2 δ(E 1 − E 2 + ω λ1 ) (14.6.1)implying th<strong>at</strong> the presence <strong>of</strong> n λ1 photons in the cavity enhances the emission by (n λ1 + 1)times.The spontaneous emission, however, is not affected by the presence <strong>of</strong> photons <strong>of</strong> otherfrequencies in the cavity. For the absorption case, it can be shown th<strong>at</strong> the formula isw q = 2π n λ 1|H ed,λ1,1,2| 2 δ(E 2 − E 1 − ω λ1 ) (14.6.2)14.7 Total Spontaneous Emission R<strong>at</strong>eWhen an electron emits a photon into the cavity, there are many modes with the samefrequency th<strong>at</strong> the emission can occur. In general, the total spontaneous emission r<strong>at</strong>e isW spon = ∑ w q = 2π ∑|H ed,λ,1,2 | 2 δ(E 1 − E 2 + ω λ ) (14.7.1)qλwhere in the above summ<strong>at</strong>ion, we have replaced q with λ since λ is the index for the photonmode th<strong>at</strong> the quantum st<strong>at</strong>e transition to in the spontaneous emission <strong>of</strong> one photon.Furthermore,√ ∫ ωλH ed,λ,1,2 = ie φ ∗ j (r)[u λ (r) · r]φ k (r)dr2ɛ 0√ωλ≃ ie u λ (r 0 ) · r jk (14.7.2)2ɛ 0where∫r jk = φ ∗ j (r)rφ k (r)dr (14.7.3)We have assumed here th<strong>at</strong> u λ (r 0 ) is slowly varying compared to φ l (r), l = j, k, and th<strong>at</strong>φ l (r) is highly localized around an <strong>at</strong>om.The modes in the cavity can be made countable by imposing a periodic boundary conditionarriving <strong>at</strong>u λ (r) = e 1 √Vbe ik λ·r(14.7.4)where we have denoted the index <strong>of</strong> k by λ where ordinarily, it is omitted or implied.The summ<strong>at</strong>ion over different electromagnetic modes then becomes∑→ ∑ ∑→ ∑ ∑ V b(2π) 3 ∆k = ∑ ∫V bdk (14.7.5)(2π)3λ pol k pol kpol


198 Quantum Mechanics Made SimpleFigure 14.1: The vectors in the polariz<strong>at</strong>ion <strong>of</strong> the emitted photon are aligned to simplify thecalcul<strong>at</strong>ion (from DAB Miller).In the k space, the spacings between the modes are given by (2π/L x , 2π/L y , 2π/L z ). Hence,each mode occupies a volume <strong>of</strong> (2π) 3 /V b where V b = L x L y L z , and the number <strong>of</strong> modes indk is dV b /(2π) 3 dk.We can also pick the polariz<strong>at</strong>ion such th<strong>at</strong> one <strong>of</strong> them is always orthogonal to r 12 , soth<strong>at</strong>u λ (r 0 ) · r 12 = u λ (r 0 )r 12 sin θ (14.7.6)The above is also equivalent to making r 12 to coincide with the z-axis <strong>of</strong> a spherical coordin<strong>at</strong>esystem, so th<strong>at</strong> the radi<strong>at</strong>ion problem is axi-symmetric.Consequently,W spon = 2π ∫ ∣V ∣∣∣∣ √ 2b ωk 1 (2π) 3 ie √ e ik·r0 r 12 sin θδ(E2ɛ 0 Vb ∣ 1 − E 2 + ω k )dk (14.7.7)orW spon = e2 |r 12 | 28π 2 ɛ 0= e2 |r 12 | 28π 2 ɛ 0∫ω k sin 2 θδ(E 1 − E 2 + ω k )dk∫ ∞ ∫ πk=0θ=0ω k δ(E 1 − E 2 + ω k )2π sin 3 θk 2 dθdk (14.7.8)In the above, ω k = ck, or ck = ω k . ThenSinceW spon = e2 |r 12 | 24πɛ 0 c 3 4 ∫ ∞k=0∫ πθ=0(ω k )δ(E 1 − E 2 + ω k ) sin 3 θdθ(ω k ) 2 d(ω k ) (14.7.9)∫ πθ=0∫ 1sin 3 θdθ = − − cos−1(1 2 θ)d cos θ = 4 3(14.7.10)


Interaction <strong>of</strong> Different Particles 199then,W spon = e2 |r 12 | 2 ω 3 1<strong>23</strong>πɛ 0 c 3 (14.7.11)where ω 12 = E 2 − E 1 . The life time <strong>of</strong> a st<strong>at</strong>e is then τ = W −1spon.14.8 More on Atom-Field InteractionIt is shown th<strong>at</strong> when an electron is in the presence <strong>of</strong> a field, the classical Hamiltonian foran <strong>at</strong>om isH A = 12m [p + eA(r, t)]2 − eΦ(r, t) + V (r) (14.8.1)Lorentz force law can be derived from the above. For quantum mechanics, we elev<strong>at</strong>e p toˆp = −i∇ to become an oper<strong>at</strong>or arriving <strong>at</strong>Ĥ A = 12m [ˆp + eA]2 − eΦ + V (r) (14.8.2)The scalar potential Φ and vector potential A describe an electromagnetic field, whereE(r, t) = −∇Φ − ∂ t A(r, t) (14.8.3)B(r, t) = ∇ × A(r, t) (14.8.4)The definitions <strong>of</strong> A and Φ are not unique. It is common to apply the radi<strong>at</strong>ion gauge where∇ · A = 0, Φ = 0, where E = −∂ t A(r, t). 1 In this gauge, E and A are linearly rel<strong>at</strong>ed,∇ · E = 0, and it works for a source-free medium. Consequently,Ĥ A = 12m [ˆp + eA]2 + V (r) (14.8.5)In the long wavelength approxim<strong>at</strong>ion, we let A(r, t) = A(t) be independent <strong>of</strong> r. In thiscase we can show th<strong>at</strong>(ˆp + eA) e − ie A(t)·r φ(r, t) = e − ie A(t)·rˆpφ(r, t) (14.8.6)But(ˆp + eA) 2 e − ie A(t)·r φ(r, t) = e − ie A(t)·rˆp 2 φ(r, t) (14.8.7)∂∂t e− ie A(t)·r φ(r, t) = − ie ∂A(t)· rφ(r, t) + e − ie A(t)·r ∂ t φ(r, t) (14.8.8) ∂t1 This is Coulomb gauge with Φ = 0.


200 Quantum Mechanics Made SimpleGiven the original Schrödinger equ<strong>at</strong>ion <strong>of</strong> the form( )1Ĥ A φ ′ (r, t) =2m [ˆp + eA]2 + V (r) φ ′ (r, t) = i ∂ ∂t φ′ (r, t) (14.8.9)with the following transform<strong>at</strong>ionφ ′ (r, t) = e − ie A(t)·r φ(r, t) (14.8.10)and making use <strong>of</strong> (14.8.6) to (14.8.8), it becomes[ ]12m ˆp2 + V (r) − e∂ t A(t) · r φ(r, t) = i∂ t φ(r, t) (14.8.11)Using the fact th<strong>at</strong> E = −∂ t A, the above becomes[ ]12m ˆp2 + V (r) + eE · r φ(r, t) = i∂ t φ(r, t) (14.8.12)The above is equivalent to performing a similarity transform to the Hamiltonian ĤA asor a change <strong>of</strong> basis fromHowever, the total Hamiltonian <strong>of</strong> the <strong>at</strong>om-field system isĤ ′ A = e ie A(t)·r Ĥ A e − ie A(t)·r (14.8.13)φ ′ (r, t) = e − ie A(t)·r φ(r, t) (14.8.14)Ĥ = ĤA + ĤF (14.8.15)whereĤ F = ∑ λ12 ω λ[â † λâλ + â λ â † λ ] (14.8.16)In a fully quantum system A → Â or fields are elev<strong>at</strong>ed to oper<strong>at</strong>ors in ĤA or in (14.8.5).It can be shown th<strong>at</strong> the above similarity transform works ever if we include ĤF part in theHamiltonian. After the similarity transform, the above becomeswhereandĤ ′ = ĤA0 + Ĥ′ F + Ĥ′ I (14.8.17)Ĥ A0 = ˆp2 + V (r) (14.8.18)2mĤ ′ I = eÊ · r (14.8.19)H ′ F retains the form similar to H F . 2 The above manipul<strong>at</strong>ion is also known as the electricfield gauge or E-gauge.2 See C. Cohen-Tannoudji, J. Dupont-Roc and G. Grynberg, Photons and Atoms: Introduction to QuantumElectrodynamics, Wiley Pr<strong>of</strong>essional, Feb 1997.


Interaction <strong>of</strong> Different Particles 201Another way to derive the dipole interaction term is to rewrite (14.8.5) asĤ A = ˆp2 eA · ˆp+2m m+ e2 A · A+ V (r) (14.8.20)2mwhere we have assumed th<strong>at</strong> ˆp · A = 0 because ∇ · A = 0. Since ˆp acts on the <strong>at</strong>omic orbitalwave functions, which are rapidly varying, we can assume th<strong>at</strong> ˆp ≫ eA. Then (14.8.20)becomesĤ A∼ ˆp 2 eA · ˆp= + V (r) + (14.8.21)2m mThe above requires no similarity transform. When the field Hamiltonian is added, and allfields are elev<strong>at</strong>ed to oper<strong>at</strong>ors, we can write the total Hamiltonian aswhere ĤA0 is as defined before in (14.8.18), andĤ = ĤA0 + ĤF + ĤI (14.8.22)Ĥ I =e · ˆpm(14.8.<strong>23</strong>)14.8.1 Interaction PictureIf the quantum system is initially described by a Hamiltonian Ĥ0, and with the addition <strong>of</strong>the interaction Hamiltonian ĤI, we can write the exact solution <strong>of</strong> the quantum system interms <strong>of</strong> the interaction picture. For Schrödinger equ<strong>at</strong>ioni ∂ ∂t |ψ〉 = (Ĥ0 + ĤI)|ψ〉 (14.8.24)we first letThen|ψ〉 = e − i Ĥ0t |φ〉 (14.8.25)i ∂ ∂t |ψ〉 = Ĥ0|ψ + e − i Ĥ0t ∂ |φ〉 (14.8.26)∂tUsing the above in (14.8.24), we haveThe above equ<strong>at</strong>ion can be integr<strong>at</strong>ed formally to yield(|φ(t)〉 = exp − i i ∂ ∂t |φ〉 = e i Ĥ0t Ĥ I e − i Ĥ0t |φ〉 (14.8.27)∫ t0dt ′ e i Ĥ0t′ Ĥ I e − i Ĥ0t′ )|φ (0)〉 (14.8.28)If instead we use a perturb<strong>at</strong>ion concept and let|φ〉 ≈ ∑ []c (0)m + c (1)m (t) |φ m 〉 (14.8.29)m


202 Quantum Mechanics Made SimpleUsing the fact th<strong>at</strong> ∂ t c (0)m = 0, ∂ t |φ m 〉 = 0, because |φ m 〉 is a st<strong>at</strong>ionary st<strong>at</strong>e with no timedependence, from (14.8.27), we havei ∑ mċ (1)m |φ m 〉 = ∑ mc (0)m e i Ĥ0t Ĥ I e − i Ĥ0t |φ m 〉 (14.8.30)Assuming th<strong>at</strong> c (0)m = δ sm , i.e., the quantum system is in only one initial st<strong>at</strong>e, then testing(14.8.30) with |φ q 〉 where |φ q 〉 is a final st<strong>at</strong>ionary st<strong>at</strong>e, we haveiċ (1)q = e i(ωq−ωs)t 〈φ q |ĤI|φ s 〉 (14.8.31)The above is the same as the one derived from time-dependent perturb<strong>at</strong>ion theory.14.8.2 E · r or A · p InteractionTo find the time evolution <strong>of</strong> the quantum system due to interaction between the <strong>at</strong>om andthe field, we need to evalu<strong>at</strong>e〈φ q |ĤI|φ s 〉 (14.8.32)However, the interaction Hamiltonian can be written asin one case, andĤ I = eÊ · r (14.8.33)Ĥ I ′ = e · p/m (14.8.34)in another case. In order to reconcile the difference, we note th<strong>at</strong> in (14.8.14), φ ′ and φ arerel<strong>at</strong>ed by a transform. When this transform is accounted for, the difference disappears. 33 For details, see M.O. Scully and M. Suhail Zubairy, Quantum Optics,CUP, 1997.


Chapter 15Quantum Inform<strong>at</strong>ion andQuantum Interpret<strong>at</strong>ion15.1 IntroductionOne important tenet <strong>of</strong> quantum mechanics is th<strong>at</strong> one does not know wh<strong>at</strong> the st<strong>at</strong>e <strong>of</strong>the system is until one performs a measurement. Before the measurement, the quantumsystem is described by a st<strong>at</strong>e th<strong>at</strong> is in a linear superposition <strong>of</strong> different st<strong>at</strong>es. After themeasurement, the system collapses to the st<strong>at</strong>e th<strong>at</strong> is “discovered” by the measurement.It is the existence as a linear superposition <strong>of</strong> st<strong>at</strong>es th<strong>at</strong> gre<strong>at</strong>ly enriches the inform<strong>at</strong>ioncontent <strong>of</strong> a quantum system. The possibility <strong>of</strong> a system to be simultaneously in differentst<strong>at</strong>es is peculiar to quantum mechanics. Objects in the classical world cannot be in such ast<strong>at</strong>e. It puts quantum systems in the realm <strong>of</strong> “ghosts” and “angels” in fairy tales and ghoststories <strong>of</strong> different cultures; these ghosts and angels can be simultaneously in many places orin different st<strong>at</strong>es. This “spookiness” <strong>of</strong> quantum mechanics is only recently confirmed byexperiments in the l<strong>at</strong>e 70’s and early 80’s. It is because <strong>of</strong> these “ghost-angel” st<strong>at</strong>es, th<strong>at</strong> wecan have quantum cryptography, quantum communic<strong>at</strong>ion, quantum circuits, and quantumcomputing.15.2 Quantum CryptographyQuantum cryptography can be used for secure quantum communic<strong>at</strong>ion. It is secure becausea quantum st<strong>at</strong>e cannot be replic<strong>at</strong>ed without disturbing the quantum st<strong>at</strong>e. Hence, whoeverwants to replic<strong>at</strong>e a quantum st<strong>at</strong>e to steal the d<strong>at</strong>a will be easily detected. The property <strong>of</strong>non-replic<strong>at</strong>ion follows from the no-cloning theorem.15.2.1 No-cloning TheoremFirst we assume th<strong>at</strong> a quantum oper<strong>at</strong>or can be designed such th<strong>at</strong> it can clone a quantumst<strong>at</strong>e without measuring it. But such a capability will viol<strong>at</strong>e the principle <strong>of</strong> linear superpo-203


204 Quantum Mechanics Made Simplesition, as we shall show. Therefore, such an oper<strong>at</strong>or cannot be designed. This is a pro<strong>of</strong> bycontradiction.A quantum system is described by a quantum st<strong>at</strong>e th<strong>at</strong> evolves from an initial st<strong>at</strong>e toa final st<strong>at</strong>e following the laws <strong>of</strong> quantum mechanics. An example <strong>of</strong> such an evolutionaryoper<strong>at</strong>or isˆT = e −i Ĥ t (15.2.1)The above quantum oper<strong>at</strong>or is a unitary oper<strong>at</strong>or as well as a linear oper<strong>at</strong>or. First, assumeth<strong>at</strong> it has the capability <strong>of</strong> replic<strong>at</strong>ing a quantum st<strong>at</strong>e in system 2 to be identical to thest<strong>at</strong>e in system 1 after acting on such a quantum system. It does so without altering thequantum st<strong>at</strong>e <strong>of</strong> system 1, e.g., by a measurement. We denote this byBy the same token, it should replic<strong>at</strong>eNow if the st<strong>at</strong>e to be replic<strong>at</strong>ed isˆT |Ψ s 〉 2 |Ψ a 〉 1 = |Ψ a 〉 2 |Ψ a 〉 1 (15.2.2)ˆT |Ψ s 〉 2 |Ψ b 〉 1 = |Ψ b 〉 2 |Ψ b 〉 1 (15.2.3)|Ψ c 〉 1 = 1 √2[|Ψ a 〉 1 + |Ψ b 〉 1 ] (15.2.4)ThenˆT |Ψ s 〉 2 |Ψ c 〉 1 = 1 √2[ˆT |Ψs 〉 2 |Ψ a 〉 1 + ˆT |Ψ s 〉 2 |Ψ b 〉 1]= 1 √2[|Ψ a 〉 2 |Ψ a 〉 1 + |Ψ b 〉 2 |Ψ b 〉 1 ] ≠ |Ψ c 〉 2 |Ψ c 〉 1 (15.2.5)Clearly, the above viol<strong>at</strong>es the principle <strong>of</strong> linear superposition if the last equality is true.Hence, such a cloning oper<strong>at</strong>or is impossible due to the viol<strong>at</strong>ion <strong>of</strong> the principle <strong>of</strong> linearsuperpostion. The above proves the no-cloning theorem.15.2.2 Entangled St<strong>at</strong>esA multi-mode photon can be described by the following field oper<strong>at</strong>orÊ(r) = ∑ √ωke s â k,s e ik·r−iωt + c.c (15.2.6)2V ɛ 0k,sThe single photon st<strong>at</strong>e can be denoted by|ψ〉 = |1 k,v 〉 (15.2.7)The above denotes a photon in a pure k st<strong>at</strong>e with polariz<strong>at</strong>ion v. It has a packet <strong>of</strong> energyE = ω k . However, a photon with a pure k st<strong>at</strong>e is not localized. But a photon gener<strong>at</strong>ed by a


Quantum Inform<strong>at</strong>ion and Quantum Interpret<strong>at</strong>ion 205source like an <strong>at</strong>omic transition must be causal, and hence, localized. A localized wave packetdescribing this photon field can be constructed by using a linear superposition <strong>of</strong> wavenumberk or frequencies. For high frequency photons, this localized st<strong>at</strong>e can have a center frequencywith a small spread <strong>of</strong> frequencies around the center frequency. The single-photon Fock st<strong>at</strong>ecan hence be written as|ψ〉 = ∑ kc k |1 k,v 〉 (15.2.8)For quasi-monochrom<strong>at</strong>ic photons, the above will be domin<strong>at</strong>ed by one term and we candenote this photon approxim<strong>at</strong>ely with the st<strong>at</strong>e vector (15.2.7).Figure 15.1: Two photons traveling in different directions.With the above picture in mind, we can think <strong>of</strong> two localized photons traveling in differentdirections. A direct product space can be used to represent the st<strong>at</strong>e <strong>of</strong> these two photons:|ψ〉 ab = |1 ka,v〉 a |1 kb ,v〉 b (15.2.9)For simplicity, we denote a two-where we have assumed quasi-mono-chrom<strong>at</strong>ic photons.photon st<strong>at</strong>e as|ψ〉 12 = |V 〉 1 |V 〉 2 (15.2.10)If the two photons are gener<strong>at</strong>ed from the same source, entangled photon st<strong>at</strong>es mayresult. Entangled two-particle st<strong>at</strong>es are those th<strong>at</strong> cannot be written as a product (outerproduct or tensor product) <strong>of</strong> simpler st<strong>at</strong>es. An example is the EPR (Einstein, Podolsky,and Rosen) pair|Φ + 〉 12 = 1 √2(|H〉1 |H〉 2 + |V 〉 1 |V 〉 2)(15.2.11)Other entangled st<strong>at</strong>es are|Φ − 〉 12 = 1 √2(|H〉1 |H〉 2 − |V 〉 1 |V 〉 2)|Ψ + 〉 12 = 1 √2(|H〉1 |V 〉 2 + |V 〉 1 |H〉 2)|Ψ − 〉 12 = 1 √2(|H〉1 |V 〉 2 − |V 〉 1 |H〉 2)(15.2.12)(15.2.13)(15.2.14)


206 Quantum Mechanics Made SimpleThe above four st<strong>at</strong>es are also called the Bell st<strong>at</strong>es. They are usually gener<strong>at</strong>ed due to theparticle pair needing to s<strong>at</strong>isfy conserv<strong>at</strong>ion <strong>of</strong> angular momentum. For instance, two photonsare gener<strong>at</strong>ed by <strong>at</strong>omic transitions where the initial angular momentum <strong>of</strong> the system is zero.An angular momentum conserving st<strong>at</strong>e with two counter-propag<strong>at</strong>ing photon is|Φ〉 = 1 √2(|R〉1 |R〉 2 + |L〉 1 |L〉 2)(15.2.15)where R and L stand for right-handed and left-handed circular polariz<strong>at</strong>ions, respectively.By letting|R〉 1 = (|H〉 1 + i|V 〉 1 ) , |R〉 2 = (|H〉 2 + i|V 〉 2 ) (15.2.16)|L〉 1 = (|H〉 1 − i|V 〉 1 ) , |L〉 2 = (|H〉 2 − i|V 〉 2 ) (15.2.17)which is a change <strong>of</strong> basis, the above becomes|Φ〉 = − 1 √2(|H〉1 |H〉 2 − |V 〉 1 |V 〉 2)(15.2.18)one <strong>of</strong> the Bell st<strong>at</strong>es.The entangled st<strong>at</strong>es are bewildering because it means th<strong>at</strong> for two counter-propag<strong>at</strong>ingphotons, if one measures one photon is in |H〉 st<strong>at</strong>e, the other photon immedi<strong>at</strong>ely collapsesto an |H〉 st<strong>at</strong>e as well, regardless <strong>of</strong> how far apart the two photons are. Whereas before themeasurement, the photons are in a linear superposition <strong>of</strong> a |H〉 and |V 〉 st<strong>at</strong>es.15.2.3 A Simple Quantum Encryption AlgorithmLet us assume th<strong>at</strong> Alice and Bob communic<strong>at</strong>e by the use <strong>of</strong> simple photon polarizers. WhenAlice sends out a |V 〉 st<strong>at</strong>e photon, it represents a “1” and similarly, an |H〉 st<strong>at</strong>e photonrepresents a “0”. If Bob receives with a similarly aligned polarizer, he receives the inform<strong>at</strong>ioncorrectly. (We call this the VH mode.) If Alice aligns her polarizer <strong>at</strong> 45 ◦ and Bob followssuit, he continues to receive the inform<strong>at</strong>ion correctly.However, if Alice aligns her polarizer vertically, and Bob aligns his <strong>at</strong> 45 ◦ , the bit inform<strong>at</strong>ionreceived by him is correct only 50% <strong>of</strong> the time. This is because the |+45〉 st<strong>at</strong>e and|−45〉 st<strong>at</strong>e are expressible as a linear superposition <strong>of</strong> the |H〉 and |V 〉 st<strong>at</strong>es. Normally,|+45〉 = 1 √2(|H〉 + |V 〉) (15.2.19)|−45〉 = 1 √2(|H〉 − |V 〉) (15.2.20)By the tenet <strong>of</strong> quantum measurement, these st<strong>at</strong>es are measured with equal likelihood<strong>of</strong> being |H〉 or |V 〉. Similarly, if Alice transmits with her polarizer aligned in the 45 ◦ angle,and Bob receives with polarizers with H and V polariz<strong>at</strong>ion, he receives the polariz<strong>at</strong>ion <strong>of</strong>|H〉 and |V 〉 with equal likelihood (50% chance) regardless <strong>of</strong> wh<strong>at</strong> polariz<strong>at</strong>ion Alice sends.This is because|H〉 = √ 1 (|+45〉 + |−45〉) (15.2.21)2


Quantum Inform<strong>at</strong>ion and Quantum Interpret<strong>at</strong>ion 207Figure 15.2: Communic<strong>at</strong>ion between Alice and Bob using single-photon source and simplifiedpolarizer measurement schemes (from DAB Miller).|V 〉 = 1 √2(|+45〉 − |−45〉) (15.2.22)Now if Alice decides to use her polarizer randomly, and Bob receives with the polarizerrandomly so th<strong>at</strong> they are equally likely to use VH mode or 45 ◦ mode. The probability th<strong>at</strong>their polarizers are aligned is correctly 50%. During this time, they communic<strong>at</strong>e with noerror. The other 50% time, when their polarizers are misaligned, they communic<strong>at</strong>e with 50%error. Hence, 25% <strong>of</strong> the d<strong>at</strong>a are wrong.After a preliminary quantum communic<strong>at</strong>ion, Alice and Bob can communic<strong>at</strong>e the inform<strong>at</strong>ionabout the alignments <strong>of</strong> their polarizers, say, by a phone call. Bob will retain onlythe d<strong>at</strong>a when their polarizers are correctly aligned and discard the rest. For the preliminarycommunic<strong>at</strong>ion, they can compare their d<strong>at</strong>a over the aligned case, and there should be errorfree in principle.If an eavesdropper, Eve, <strong>at</strong>tempts to steal the inform<strong>at</strong>ion, she does so by using a polarizer


208 Quantum Mechanics Made Simpleto intercept the photon. If Eve knows th<strong>at</strong> Alice is using the VH mode, Eve aligns her polarizerin the VH mode. After Eve has received the d<strong>at</strong>a, she can retransmit the d<strong>at</strong>a to Bob, thusstealing the d<strong>at</strong>a. However, if the polariz<strong>at</strong>ion used by Alice is random, Eve’s polarizer isnot aligned with Alice’s half the time. Eve would have corrupted her d<strong>at</strong>a making the wrongtransmission 50% <strong>of</strong> the time. This would increase the error in transmission <strong>of</strong> the d<strong>at</strong>a fromAlice to Bob making it wrong 25% <strong>of</strong> the time. If Alice and Bob communic<strong>at</strong>e by a phonecall to check the security <strong>of</strong> their d<strong>at</strong>a transmission and found th<strong>at</strong> it is wrong approxim<strong>at</strong>ely25% <strong>of</strong> the time, they would have suspected an eavesdropper.Notice th<strong>at</strong> Eve has to collapse the st<strong>at</strong>e <strong>of</strong> the photon sent out by Alice into one <strong>of</strong> thetwo st<strong>at</strong>es <strong>of</strong> Eve’s polarizer before she can duplic<strong>at</strong>e the photon and send it to Bob. Because<strong>of</strong> the no-cloning theorem, Eve cannot duplic<strong>at</strong>e the st<strong>at</strong>e <strong>of</strong> the photon th<strong>at</strong> Alice has sentwithout measuring it.In a secure communic<strong>at</strong>ion system, Alice and Bob do not send the real message in preliminarytesting <strong>of</strong> the security <strong>of</strong> time channel. Alice will first send Bob the secret key andtest if the channel is secure. If it is a secure channel, then she would send the rest <strong>of</strong> theinform<strong>at</strong>ion. The above is known as the BB84 protocol, <strong>at</strong>tributed to Bennett and Brassard’swork in 1984.Also, notice th<strong>at</strong> the above secure communic<strong>at</strong>ion system does not work in a classicaloptical communic<strong>at</strong>ion channel where a bunch <strong>of</strong> photons is sent. If a bunch <strong>of</strong> photon is sentby Alice to denote a V or an H polariz<strong>at</strong>ion to send “1” and “0”, when Eve eavesdrops withher misaligned polarizer by 45 ◦ , she would have noticed th<strong>at</strong> equal number <strong>of</strong> photons areemerging from her two orthogonal polariz<strong>at</strong>ions. By checking the phase <strong>of</strong> the two streams<strong>of</strong> photons, she can easily duplic<strong>at</strong>e a classical photon bunch and send it to Bob, meanwhilestealing the d<strong>at</strong>a <strong>of</strong>f the communic<strong>at</strong>ion channel. Hence, the security <strong>of</strong> the quantum communic<strong>at</strong>ionchannel comes from the interpret<strong>at</strong>ion <strong>of</strong> quantum mechanics: a particle is in thelinear superposition <strong>of</strong> quantum st<strong>at</strong>es before the measurement. A measurement collapsesthe quantum st<strong>at</strong>e into one <strong>of</strong> the st<strong>at</strong>es “discovered” by the measurement.The above discussion <strong>of</strong> a secure channel is based on ideal single-photon sources. In practice,non-ideality will give rise to communic<strong>at</strong>ion errors. Quantum error correction schemeshave been devised to minimize the errors in a quantum communic<strong>at</strong>ion channel.15.3 Quantum ComputingThe distinguishing fe<strong>at</strong>ure <strong>of</strong> quantum computing is quantum parallelism. Again, this followsfrom the tenet <strong>of</strong> quantum measurement. A quantum st<strong>at</strong>e can be in a linear superposition <strong>of</strong>many st<strong>at</strong>es before the measurement. After the measurement, the quantum st<strong>at</strong>e collapses toone <strong>of</strong> the quantum st<strong>at</strong>es. The prowess <strong>of</strong> quantum computing, as mentioned before, comesfrom the “ghost-angel” st<strong>at</strong>e <strong>of</strong> a quantum system.15.3.1 Quantum Bits (Qubits)A quantum bit or a qubit is a bit in a quantum st<strong>at</strong>e th<strong>at</strong> is in the linear superposition <strong>of</strong>two st<strong>at</strong>es representing the |0〉 bit and the |1〉 bit. Namely,|ψ〉 = C 0 |0〉 + C 1 |1〉 (15.3.1)


Quantum Inform<strong>at</strong>ion and Quantum Interpret<strong>at</strong>ion 209where |C 0 | 2 + |C 1 | 2 = 1. The two st<strong>at</strong>es |0〉 and |1〉 can be the vertical and horizontalpolariz<strong>at</strong>ion <strong>of</strong> a photon. Altern<strong>at</strong>ively, it can be the up and down st<strong>at</strong>e <strong>of</strong> a spin, or any twoenergy levels <strong>of</strong> a multi-level system.The richness <strong>of</strong> quantum inform<strong>at</strong>ion is already manifested in this very simple example.Unlike classical bits in classical computers, which can only have binary values, a qubit canhave multitudes <strong>of</strong> possible values depending on the values <strong>of</strong> C 0 and C 1 . A two-st<strong>at</strong>e quantumsystem, as has been shown in the spin case, can be represented by a Bloch sphere. Everypoint on the Bloch sphere represents a possible quantum st<strong>at</strong>e, depending on C 1 and C 2 , andthere could be infinitely many st<strong>at</strong>es.15.3.2 Quantum G<strong>at</strong>esAnalogous to classical logic g<strong>at</strong>es, there are quantum g<strong>at</strong>es th<strong>at</strong> manipul<strong>at</strong>e the |0〉 and |1〉st<strong>at</strong>es <strong>of</strong> a qubit. A qubit as indic<strong>at</strong>ed by (15.3.1) can be represented by a column vector <strong>of</strong>length two. For example, the qubit in (15.3.1) can be represented by [C 0 , C 1 ] t . A quantumg<strong>at</strong>e transforms the quantum st<strong>at</strong>e [C 0 , C 1 ] t to another st<strong>at</strong>e [C 0, ′ C 1] ′ t . Such a m<strong>at</strong>rix[ ] [ ] [ ]C′0 M11 M=12 C0(15.3.2)M 22 C 1C ′ 1M 21has to be unitary since all quantum g<strong>at</strong>es must oper<strong>at</strong>e by the time evolution according toˆM = e −i Ĥ t (15.3.3)Figure 15.3: Some single qubit g<strong>at</strong>es showing their input and output st<strong>at</strong>es.Examples <strong>of</strong> quantum g<strong>at</strong>es, expressed in their m<strong>at</strong>rix represent<strong>at</strong>ions with a slight abuse<strong>of</strong> not<strong>at</strong>ion, are[ ]0 1 ˆX =(15.3.4)1 0[ ] 1 0Ẑ =(15.3.5)0 −1Ĥ = √ 1 [ ] 1 1(15.3.6)2 1 −1


210 Quantum Mechanics Made SimpleˆX represents the NOT g<strong>at</strong>e while Ẑ represents one th<strong>at</strong> flips the sign <strong>of</strong> |1〉 bit. Ĥ is calledthe Hadamard g<strong>at</strong>e th<strong>at</strong> is almost like a “square root” g<strong>at</strong>e.When these quantum g<strong>at</strong>es oper<strong>at</strong>e on the qubit denoted by (15.3.1), the results are asfollows:Moreover, one can show th<strong>at</strong>ˆX|ψ〉 = C 1 |0〉 + C 0 |1〉 (15.3.7)Ẑ|ψ〉 = C 0 |0〉 − C 1 |1〉 (15.3.8)Ĥ|ψ〉 = C 0√2(|0〉 + |1〉) + C 1√2(|0〉 − |1〉) (15.3.9)Ĥ 2 = Î (15.3.10)When expressed in terms <strong>of</strong> m<strong>at</strong>rix algebra,[ ] [ ]C0 C1ˆX =C 1 C 0[ ] [ ]C0 C0Ẑ =C 1 −C 1[ ]C0Ĥ =C 1[ C0+C 1 √2]C 0−C √2 1(15.3.11)(15.3.12)(15.3.13)In addition to one qubit g<strong>at</strong>e, there are also two qubit g<strong>at</strong>es. In this case, there are twoinput bits for these g<strong>at</strong>es. They can be arranged as |0, 0〉, |0, 1〉, |1, 0〉, and |1, 1〉 as the fourinput and output st<strong>at</strong>es. A very important one is the CNOT g<strong>at</strong>e shown in Figure 15.4. Theupper line represents the first bit which is also the control bit, while the lower line representsthe second bit, whose output is the target bit. The ⊕ symbol represents addition modulus 2.Hence, its transform<strong>at</strong>ion m<strong>at</strong>rix is given by⎡ ⎤1 0 0 0Û CNOT = ⎢0 1 0 0⎥⎣0 0 0 1⎦ (15.3.14)0 0 1 0The above can also be implemented with a unitary transform.15.3.3 Quantum Computing AlgorithmsAs mentioned before, the most important aspect <strong>of</strong> quantum computing algorithm is quantumparallelism. We will illustr<strong>at</strong>e this with the Deutsch algorithm. It can be implemented withthe quantum circuit shown below:We start with a st<strong>at</strong>e|ψ 0 〉 = |0, 1〉 (15.3.15)


Quantum Inform<strong>at</strong>ion and Quantum Interpret<strong>at</strong>ion 211Figure 15.4: A quantum circuit <strong>of</strong> a two-qubit g<strong>at</strong>e representing the CNOT g<strong>at</strong>e. It takestwo input streams, and has two output streams. The ⊕ symbol represents an exclusive oroper<strong>at</strong>ion, or addition modulus 2.Figure 15.5: The implement<strong>at</strong>ion <strong>of</strong> the Deutsch algorithm with quantum g<strong>at</strong>es and circuits.At stage |ψ 1 〉, we haveHence,x = Ĥ · q 1 = 1 √2(|0〉 + |1〉) (15.3.16)y = Ĥ · q 2 = 1 √2(|0〉 − |1〉) (15.3.17)|ψ 1 〉 = 1 2 (|0〉 + |1〉) (|0〉 − |1〉) = 1 (|0, 0〉 − |0, 1〉 + |1, 0〉 − |1, 1〉) (15.3.18)2The unitary oper<strong>at</strong>or û f has no effect on the x qubit, but performs the oper<strong>at</strong>ion y ⊕ f(x) onthe y qubit. The function f(x) takes input x which is either 0 or 1. It produces an outputwhich is either constant or balanced, but the output is either 0 or 1.Hence, according to the rules above,|ψ 2 〉 = 1 (|0, f(0)〉 − |0, 1 ⊕ f(0)〉 + |1, f(1)〉 − |1, 1 ⊕ f(1)〉) (15.3.19)2If f(x) is a constant function, then f(0) = f(1), and the above becomes|ψ 2 〉 const = 1 (|0, f(0)〉 − |0, 1 ⊕ f(0)〉 + |1, f(0)〉 − |1, 1 ⊕ f(0)〉)2= 1 (|0〉 + |1〉) (|f(0)〉 − |1 ⊕ f(0)〉) (15.3.20)2


212 Quantum Mechanics Made SimpleWith the Hadamard oper<strong>at</strong>ion on the upper qubit, we have|ψ 3 〉 const = |0〉 1 √2(|f(0)〉 − |1 ⊕ f(0)〉) (15.3.21)If for the balanced case, f(0) ≠ f(1), hence f(1) = 1 ⊕ f(0). Then|ψ 2 〉 bal = 1 (|0, f(0)〉 − |0, 1 ⊕ f(0)〉 + |1, 1 ⊕ f(0)〉 − |1, f(0)〉)2After the H g<strong>at</strong>e oper<strong>at</strong>ion, we have= 1 (|0〉 − |1〉) (|f(0)〉 − |1 ⊕ f(0)〉) (15.3.22)2|ψ 3 〉 bal = |1〉 1 √2(|f(0)〉 − |1 ⊕ f(0)〉) (15.3.<strong>23</strong>)The above shows the prowess <strong>of</strong> quantum parallelism with just one oper<strong>at</strong>ion, one candetermine if a function f(x) is balanced or constant. More sophistic<strong>at</strong>ed algorithms exploitingquantum parallelism, such as the Shor’s algorithm and the Grover’s algorithm, have beendevised. The Shor’s algorithm can perform a Fourier transform in (log N) 2 oper<strong>at</strong>ions r<strong>at</strong>herthan the classical N log N oper<strong>at</strong>ions. The Grover’s algorithm can search a d<strong>at</strong>a base withN d<strong>at</strong>a in √ N oper<strong>at</strong>ions r<strong>at</strong>her than the classical N oper<strong>at</strong>ions. Because <strong>of</strong> quantumparallelism, quantum computer can also perform quantum simul<strong>at</strong>ion <strong>of</strong> quantum systemwhich is not possible on classical computers.15.4 Quantum Teleport<strong>at</strong>ionQuantum teleport<strong>at</strong>ion is the idea <strong>of</strong> Alice being able to send a photon <strong>of</strong> unknown st<strong>at</strong>e toBob, without having to perform a measurement on this photon, nor disturb its st<strong>at</strong>e. Wedenote the photon, called photon 1, in the unknown st<strong>at</strong>e by|ψ〉 1= C 0 |0〉 1+ C 1 |1〉 1(15.4.1)In the beginning, this photon is only accessible to Alice. Alice, however, is accessible toanother photon, called photon 2, <strong>of</strong> an entangled Bell st<strong>at</strong>e. The other photon <strong>of</strong> the Bellst<strong>at</strong>e, photon 3, is not accessible to Alice but is accessible to Bob.We can define the st<strong>at</strong>e <strong>of</strong> the three photons by the direct product st<strong>at</strong>e|Ψ〉 1<strong>23</strong>= 1 √2(C 0 |0〉 1+ C 1 |1〉 1) (|0〉 2|1〉 3− |1〉 2|0〉 3) (15.4.2)where the Bell st<strong>at</strong>e is assumed to be∣ Ψ− 〉 <strong>23</strong> = 1 √2(|0〉 2|1〉 3− |1〉 2|0〉 3) (15.4.3)


Quantum Inform<strong>at</strong>ion and Quantum Interpret<strong>at</strong>ion 213Figure 15.6: A teleport<strong>at</strong>ion experiment setup with photons. Alice makes measurement onthe two photons accessible to her using the Bell st<strong>at</strong>e measurement device, and communic<strong>at</strong>ethe outcome to Bob via a classical channel. Bob then performs a unitary transform<strong>at</strong>ion onhis photon to obtain the input photon st<strong>at</strong>e. (From M. Fox, Quantum Optics.)Expanding (15.4.2) gives rise to|Ψ〉 1<strong>23</strong>= 1 √2(C 0 |0〉 1|0〉 2|1〉 3− C 0 |0〉 1|1〉 2|0〉 3+C 1 |1〉 1|0〉 2|1〉 3− C 1 |1〉 1|1〉 2|0〉 3) (15.4.4)The four Bell st<strong>at</strong>es are complete and orthogonal, and the st<strong>at</strong>es <strong>of</strong> photon 1 and photon 2can be expanded in the four Bell st<strong>at</strong>es; namely∣∣Φ +〉 12 = 1 √2(|0〉 1|0〉 2+ |1〉 1|1〉 2) (15.4.5)∣ Φ− 〉 12 = 1 √2(|0〉 1|0〉 2− |1〉 1|1〉 2) (15.4.6)∣∣Ψ +〉 12 = 1 √2(|0〉 1|1〉 2+ |1〉 1|0〉 2) (15.4.7)∣ Ψ− 〉 12 = 1 √2(|0〉 1|1〉 2− |1〉 1|0〉 2) (15.4.8)Projecting (15.4.4) onto the four Bell st<strong>at</strong>es, and subsequently expanding (15.4.4) in terms<strong>of</strong> them, we have|Ψ〉 1<strong>23</strong>= 1 [∣ ∣ Φ +〉 212 (C 0 |1〉 3− C 1 |0〉 3)+ ∣ Φ− 〉 12 (C 0 |1〉 3+ C 1 |0〉 3)+ ∣ ∣ Ψ+ 〉 12 (−C 0 |0〉 3+ C 1 |1〉 3)− ∣ ∣Ψ −〉 12 (C 0 |0〉 3+ C 1 |1〉 3) ] (15.4.9)


214 Quantum Mechanics Made SimpleIn the above, the first two photons, photon 1 and photon 2, are grouped into different Bellst<strong>at</strong>es. Moreover, the st<strong>at</strong>e <strong>of</strong> photon 3 resembles the st<strong>at</strong>e <strong>of</strong> the original photon 1. Thequantum system now is in a linear superposition <strong>of</strong> different Bell st<strong>at</strong>es.The Bell st<strong>at</strong>e measurement device projects the first two photons onto a Bell st<strong>at</strong>e. TheBell st<strong>at</strong>e measurement device collapses the quantum system into one <strong>of</strong> the four Bell st<strong>at</strong>es.For example, when Alice finds th<strong>at</strong> the first two photons are in the Bell st<strong>at</strong>e |Φ + 〉 12, thenthe third photon must be in the st<strong>at</strong>e|ψ〉 3= C 0 |1〉 3− C 1 |0〉 3(15.4.10)Bob can apply a unitary oper<strong>at</strong>or, similar to qubit g<strong>at</strong>e oper<strong>at</strong>ors described in the previoussection, to obtain the original st<strong>at</strong>e <strong>of</strong> photon 1.The above does not viol<strong>at</strong>e the no-cloning theorem, because the original st<strong>at</strong>e <strong>of</strong> thephoton 1 is destroyed, and its semblance is reproduced in photon 3.15.5 Interpret<strong>at</strong>ion <strong>of</strong> Quantum MechanicsQuantum mechanics has the basic tenet th<strong>at</strong> a quantum st<strong>at</strong>e is in a linear superposition <strong>of</strong>st<strong>at</strong>es before a measurement. A measurement projects a quantum st<strong>at</strong>e into one <strong>of</strong> the st<strong>at</strong>es.The ability <strong>of</strong> a quantum st<strong>at</strong>e to be in a linear superposition <strong>of</strong> st<strong>at</strong>es is surreal, and it hasbothered a gre<strong>at</strong> many physicists. In the classical world, a system can only be in one st<strong>at</strong>eor the other, but not in a linear superposition <strong>of</strong> st<strong>at</strong>es. Only the world <strong>of</strong> ghosts and angelscan we imagine th<strong>at</strong> an object is in a linear superposition <strong>of</strong> st<strong>at</strong>es.In the coordin<strong>at</strong>e space, an electron, represented by its wavefunction, can be simultaneously<strong>at</strong> all loc<strong>at</strong>ions where the wavefunction is non-zero. In the Young’s double slit experiment,the electron, represented by its wavefunction, can go through both slits simultaneouslylike a wave. 1When the ghost-angel st<strong>at</strong>e concept is extended to classical objects, such as a c<strong>at</strong>, it givesrise to the ludicrous result: the story <strong>of</strong> the Schrödinger c<strong>at</strong>. The Schrödinger c<strong>at</strong> is a linearsuperposition <strong>of</strong> a dead c<strong>at</strong> and a live c<strong>at</strong>. To understand why the Schrödinger c<strong>at</strong> does notexist, we need to understand the concept <strong>of</strong> quantum coherence.Two st<strong>at</strong>es are in quantum coherence if the phase rel<strong>at</strong>ionships between them are deterministicand not random. When this coherence is lost, the phase rel<strong>at</strong>ionship between themis lost. The quantum system has already collapsed into one <strong>of</strong> the two st<strong>at</strong>es. Hence, inpractice, a measurement is not always necessary before the quantum system collapses intoone or more <strong>of</strong> the quantum st<strong>at</strong>es. The interaction <strong>of</strong> quantum system with its environmentcan cause such a collapse.From a st<strong>at</strong>istical physics viewpoint, it is impossible for a quantum system to be completelyisol<strong>at</strong>ed. Almost all systems are in a thermal b<strong>at</strong>h <strong>of</strong> the universe with which they areseeking equilibrium. Macroscopic objects cannot be in a pure quantum st<strong>at</strong>e which has thecharacteristics <strong>of</strong> the ghost-angel st<strong>at</strong>e. It is impossible for the huge number <strong>of</strong> <strong>at</strong>oms in theSchrödinger c<strong>at</strong> to be coherent with respect to each other.1 Or the apparition <strong>of</strong> the ghost-angel st<strong>at</strong>e.


Quantum Inform<strong>at</strong>ion and Quantum Interpret<strong>at</strong>ion 215The density m<strong>at</strong>rix is a nice way <strong>of</strong> representing a st<strong>at</strong>e <strong>of</strong> a quantum system wherethe physics <strong>of</strong> quantum coherence surfaces explicitly. This concept is expressed in the <strong>of</strong>fdiagonalterms <strong>of</strong> the density m<strong>at</strong>rix. If one allows time average or ensemble average 2 to thedensity m<strong>at</strong>rix, when the system is expressed by quantum st<strong>at</strong>es th<strong>at</strong> are not coherent, the<strong>of</strong>f-diagonal elements will average to zero. The system is in a mixed st<strong>at</strong>e r<strong>at</strong>her than a purequantum st<strong>at</strong>e. The system is similar to the local hidden variable theory: the st<strong>at</strong>e <strong>of</strong> thequantum system is already predetermined before the measurement.Another uneasiness about the philosophical interpret<strong>at</strong>ion <strong>of</strong> quantum mechanics is th<strong>at</strong>one does not know wh<strong>at</strong> st<strong>at</strong>e the quantum system is in before the measurement. This hasprompted Einstein to ask,“Is the moon there if you don’t look <strong>at</strong> it?” The uncertainty <strong>of</strong>the st<strong>at</strong>e applied to quantum mechanics is only true for a linear superposition <strong>of</strong> coherentquantum st<strong>at</strong>es, which I term the ghost-angel st<strong>at</strong>e. This st<strong>at</strong>e has not been found to existfor macroscopic objects.However, if one insists th<strong>at</strong>, “One does not know if the moon is there before one looks <strong>at</strong>it.” as a true st<strong>at</strong>ement, it cannot be refuted nor confirmed by experiments. The mere act <strong>of</strong> anexperiment already means th<strong>at</strong> we have “looked” <strong>at</strong> the moon. The same claim goes th<strong>at</strong> “If Isaw a fallen tree in the forest, it did not necessary follow from the act <strong>of</strong> falling before I arrivedthere.” Altern<strong>at</strong>ively, “If we found dinosaur bones, it did not necessary mean th<strong>at</strong> dinosaursroamed the earth over 200 million years ago.” We believe th<strong>at</strong> the moon is there even if we donot look <strong>at</strong> it, the tree fell because it went through the act <strong>of</strong> falling, and th<strong>at</strong> dinosaur boneswere found because they roamed the earth 200 million years ago, because we believe in therealism <strong>of</strong> the world we live in. This realism cannot be proved but is generally accepted bythose who live in this world. Hence, it is this surreal interpret<strong>at</strong>ion <strong>of</strong> quantum mechanics th<strong>at</strong>causes the uneasiness among many physicists. But the interpret<strong>at</strong>ion <strong>of</strong> quantum mechanicsis slightly better than the above: a quantum st<strong>at</strong>e is in a linear superposition <strong>of</strong> st<strong>at</strong>es, theprecise one <strong>of</strong> which we are not sure <strong>of</strong> until a measurement is performed. However, thissurrealism <strong>of</strong> this ghost-angel st<strong>at</strong>e exists in our minds in fairy tales and ghost stories <strong>of</strong>many cultures. Experimental effort has agreed with the surreal interpret<strong>at</strong>ion <strong>of</strong> quantummechanics in terms <strong>of</strong> the Bell’s theorem, th<strong>at</strong> will be discussed.The ghost-angel st<strong>at</strong>e <strong>of</strong> a quantum system is wh<strong>at</strong> enriches the inform<strong>at</strong>ion in it. However,for a quantum system to be in such a st<strong>at</strong>e, the linear superposition <strong>of</strong> st<strong>at</strong>es must be coherentwith each other. Quantum coherence is the largest stumbling block to the construction <strong>of</strong>quantum computers; however, rapid advances are being made, and one day, it can be a reality.15.6 EPR ParadoxThe interpret<strong>at</strong>ion <strong>of</strong> quantum mechanics went through difficult times. The fact th<strong>at</strong> a particlecan be in a superposition <strong>of</strong> many st<strong>at</strong>es before a quantum measurement, and the probabilisticinterpret<strong>at</strong>ion <strong>of</strong> a quantum measurement behooves the challenge by many gre<strong>at</strong> physicists,especially Einstein. The most severe challenge <strong>of</strong> quantum mechanics and its interpret<strong>at</strong>ioncomes from the EPR (Einstein, Podolsky and Rosen) paradox. To describe it simply, we2 Processes for which time average is the same as ensemble average are known as ergodic processes.


216 Quantum Mechanics Made Simpleimagine a π meson th<strong>at</strong> decays into an electron-positron pair:π o → e − + e + (15.6.1)The π meson originally has spin zero. So for conserv<strong>at</strong>ion <strong>of</strong> angular momentum, the electronpositronpair will have opposite spins: spin up and spin down. Since the total angularmomentum is zero, they are in the singlet st<strong>at</strong>e which has total spin <strong>of</strong> zero, or|Ψ〉 = 1 √2(|↑ − ↓ + 〉 − |↓ − ↑ + 〉) (15.6.2)The electron-positron pair is in the linear superposition <strong>of</strong> two st<strong>at</strong>es, but the electron andpositron are flying in opposite directions. According to the interpret<strong>at</strong>ion <strong>of</strong> quantum mechanics,one does not know the spin st<strong>at</strong>e <strong>of</strong> the electron nor the positron before the measurement.After one measures the spin st<strong>at</strong>e <strong>of</strong>, say the electron, irrespective <strong>of</strong> how far the positronis away from the electron, we immedi<strong>at</strong>ely know the spin st<strong>at</strong>e <strong>of</strong> the positron according tothe above equ<strong>at</strong>ion. The spins <strong>of</strong> the two particles are always opposite to each other. Thisnotion is unpal<strong>at</strong>able to many physicists, and hence, is called “spooky action <strong>at</strong> a distance”by Einstein. Inform<strong>at</strong>ion cannot travel faster than the speed <strong>of</strong> light. How could the st<strong>at</strong>e <strong>of</strong>one particle be immedi<strong>at</strong>ely determined after a measurement is made <strong>at</strong> another particle faraway? This paradox <strong>at</strong>tempts to prove th<strong>at</strong> quantum mechanics is incomplete by reductio adabsurdum.15.7 Bell’s Theorem 3Quantum measurements are known to be random, and the d<strong>at</strong>a can only be interpretedprobabilistically. If one were to measure the spin <strong>of</strong> one <strong>of</strong> the particle, it is equally likely tobe in the spin up or spin down st<strong>at</strong>e randomly according to (15.6.2). In the hidden variabletheory, it is suggested th<strong>at</strong> the outcome <strong>of</strong> the experiment is already predetermined evenbefore the measurement. The outcome is determined by a hidden random variable λ. It isthe randomness <strong>of</strong> this variable th<strong>at</strong> gives rise to the randomness <strong>of</strong> the outcome in quantummeasurements. This is contrary to the now accepted quantum interpret<strong>at</strong>ion 4 where onedoes not know wh<strong>at</strong> st<strong>at</strong>e a quantum system is until after a measurement is performed.Before the measurement, the quantum system can be in a linear superposition <strong>of</strong> st<strong>at</strong>es. Themeasurement collapses the quantum system into the st<strong>at</strong>e “discovered” by the measurement.Many hidden variable theories were proposed shortly after the EPR paradox was published.In 1964, J. S. Bell, in the spirit <strong>of</strong> proving the correctness <strong>of</strong> the hidden variable theory, cameup with an inequality th<strong>at</strong> showed the incomp<strong>at</strong>ibility <strong>of</strong> quantum mechanics and the hiddenvariable theory. If hidden variable theory is correct, the inequality will be s<strong>at</strong>isfied, but ifquantum mechanics is correct, the inequality is viol<strong>at</strong>ed. This is known as Bell’s theorem.We can discuss the deriv<strong>at</strong>ion <strong>of</strong> the Bell’s theorem in the context <strong>of</strong> the two-photonexperiment, since the experiment th<strong>at</strong> verifies the theorem has been done using photons. Theactual experiment is quite complex, but we will reduce it to a simplified case. The simplified3 This section is written with important input from Y. H. Lo and Q. Dai.4 This was espoused by the Copenhagen school.


Quantum Inform<strong>at</strong>ion and Quantum Interpret<strong>at</strong>ion 217experiment involves a photon source th<strong>at</strong> produces an entangled photon pair, each <strong>of</strong> whichis traveling in opposite directions. The photon pair is in one <strong>of</strong> the Bell st<strong>at</strong>e, say the EPRpair:|Ψ〉 = 1 √2(|V 1 V 2 〉 + |H 1 H 2 〉) (15.7.1)In the above st<strong>at</strong>e, if one <strong>of</strong> the photons is measured to be V (vertical) polarized, the otherphoton must be V polarized. However, if one photon is measured to be H (horizontal)polarized, the other photon must be H polarized. We will detect the photon st<strong>at</strong>e with asimple polarizer. In the above st<strong>at</strong>e, which is a linear superposition <strong>of</strong> two st<strong>at</strong>es, the photonsare equally likely to be found in the first st<strong>at</strong>e, |V 1 V 2 〉, or the second st<strong>at</strong>e, |H 1 H 2 〉. Thepolarizer will detect an H or a V polariz<strong>at</strong>ion with equal likelihood, but the moment th<strong>at</strong>one photon is determined to be H polarized, the other photon is immedi<strong>at</strong>ely known to be Hpolarized, and vice versa. This is the “spookiness” <strong>of</strong> quantum interpret<strong>at</strong>ion.Imagine an <strong>at</strong>omic source th<strong>at</strong> gener<strong>at</strong>es two photons propag<strong>at</strong>ing in opposite directions.The <strong>at</strong>om initially has zero angular momentum, so th<strong>at</strong> the two photons are either bothhorizontally polarized or both vertically polarized. Hence, the photon can be described inone <strong>of</strong> the Bell st<strong>at</strong>es or an entangled st<strong>at</strong>e as shown in (15.7.1).Figure 15.7: Experimental verific<strong>at</strong>ion <strong>of</strong> Bell’s theorem15.7.1 Prediction by Quantum MechanicsFirst, we will predict the experimental outcome by invoking the quantum measurement hypothesisand quantum interpret<strong>at</strong>ion. Photon 1 is measured with a polarizer P 1 with verticalpolariz<strong>at</strong>ion pointing in the a direction. If P 1 measures a V polariz<strong>at</strong>ion, we set A(a) = 1, andif P 1 measures an H polariz<strong>at</strong>ion, we set A(a) = −1. Here, A(a) denotes the measurementoutcome, and it is completely random according to quantum mechanics.Similarly, polarizer P 2 has its vertical polariz<strong>at</strong>ion oriented in the b direction. When itmeasures a V polariz<strong>at</strong>ion, we set B(b) = 1, and when it measures an H polariz<strong>at</strong>ion, it setsB(b) = −1. Again, B(b) is completely random. In the above, we set A and B to be functions<strong>of</strong> a and b respectively, as the experimental outcomes are expected to be functions <strong>of</strong> theorient<strong>at</strong>ion <strong>of</strong> the polarizers.If a = b, we expect th<strong>at</strong>〈A(a)B(b)〉 = E (a, b) = AB = 1 (15.7.2)


218 Quantum Mechanics Made SimpleIf a⊥b, if P 1 measures a V polariz<strong>at</strong>ion, P 2 will measure a H polariz<strong>at</strong>ion, we expect th<strong>at</strong>〈A(a)B(b)〉 = E (a, b) = AB = −1 (15.7.3)Even though A and B are random, their products are deterministic in the above two cases.An interesting case ensues if a and b are <strong>at</strong> an incline with respect to each other. In thiscase, we know by quantum mechanics th<strong>at</strong>|V 〉 a = cos θ|V 〉 b − sin θ|H〉 b (15.7.4)|H〉 a = sin θ|V 〉 b + cos θ|H〉 b (15.7.5)Figure 15.8: The case when the two polarizers are <strong>at</strong> an incline with respect to each otherfor proving the Bell’s theorem.If P 1 measures an outcome with A(a) = 1, then the photon th<strong>at</strong> propag<strong>at</strong>es to P 2 mustbe polarized in the a direction. However, according to (15.7.4), for P 2 , such a photon has theprobability <strong>of</strong> cos 2 θ being detected in the V polariz<strong>at</strong>ion, and the probability <strong>of</strong> sin 2 θ beingdetected in the H polariz<strong>at</strong>ion. Hence, the expect<strong>at</strong>ion value <strong>of</strong> B, or 〈B〉 A=1 = cos 2 θ−sin 2 θ.If P 1 finds th<strong>at</strong> A = −1, by similar arguments, the expect<strong>at</strong>ion value <strong>of</strong> B or 〈B〉 A=−1 =sin 2 θ − cos 2 θ. Then 5 E(a, b) = 〈A(a)B(b)〉 = cos 2 θ − sin 2 θ = cos(2θ) (15.7.6)Notice th<strong>at</strong> the above reduces to the special cases <strong>of</strong>: (i) when the polarizers P 1 and P 2 arealigned, namely when θ = 0 ◦ , as in (15.7.2), and (ii) when the polarizers are perpendicularto each other, with θ = 90 ◦ , as in (15.7.3).15.7.2 Prediction by Hidden Variable TheoryIn the hidden variable theory deriv<strong>at</strong>ion, a particle is assumed to be already in a polariz<strong>at</strong>ionst<strong>at</strong>e even before a measurement. A and B are random, but they are predetermined by ahidden random variable λ. We letA(a, λ) = ±1 (15.7.7)5 The following can be obtained using 〈A, B〉 = ∑ A,B P (A, B)AB = ∑ A,B P (B|A)P (A)AB.


Quantum Inform<strong>at</strong>ion and Quantum Interpret<strong>at</strong>ion 219B(b, λ) = ±1 (15.7.8)Here, A and B hence flip-flop between ±1 entirely due to randomness <strong>of</strong> the variable λ. Thistheory predicts the randomness <strong>of</strong> quantum mechanics experiments nicely. The expect<strong>at</strong>ionvalue <strong>of</strong> AB then is∫E(a, b) = 〈AB〉 = ρ(λ)A(a, λ)B(b, λ)dλ (15.7.9)where ρ(λ) is the probability distribution <strong>of</strong> the random variable λ. The above is a verygeneral represent<strong>at</strong>ion <strong>of</strong> the hidden variable theory where we have not explicitly st<strong>at</strong>ed thefunctions ρ(λ), A(a, λ) nor B(b, λ).If a = b, thenA(a, λ) = B(a, λ) (15.7.10)and the above becomes∫∫E(a, a) = ρ(λ)A 2 (a, λ)dλ = ρ(λ)dλ = 1 (15.7.11)Same as we would have found in (15.7.2).If a⊥b, thenA(a, λ) = −B(b, λ) (15.7.12)and (15.7.9) becomes∫E(a, b) = −∫ρ(λ)A 2 (a, λ)dλ = −ρ(λ)dλ = −1 (15.7.13)as would have been found in (15.7.3). Hence, hidden variable theory is in good agreementwith quantum mechanics interpret<strong>at</strong>ion <strong>of</strong> (15.7.2) and (15.7.3).So far, everything is fine and dainty until when a and b do not belong to any <strong>of</strong> the abovec<strong>at</strong>egory, but in general are inclined with respect to each other. If a ≠ b in general, then thehidden variable gener<strong>at</strong>es random A and B in such a manner th<strong>at</strong>|E(a, b)| ≤ 1 (15.7.14)The above is less than one because now a and b are not exactly correl<strong>at</strong>ed or anti-correl<strong>at</strong>ed.This is quite different from why (15.7.6) is less than one when θ ≠ 0. As we shall see, thisgives rise to quite a different n<strong>at</strong>ure for E(a, b) for hidden variable theory.To this end, we can show th<strong>at</strong>∫E(a, b) − E(a, b ′ ) = [A(a, λ)B(b, λ) − A(a, λ)B(b ′ , λ)]ρ(λ)dλ (15.7.15)The above can be rewritten as∫E(a, b) − E(a, b ′ ) =∫−A(a, λ)B(b, λ)[1 ± A(a ′ , λ)B(b ′ , λ)]ρ(λ)dλA(a, λ)B(b ′ , λ)[1 ± A(a ′ , λ)B(b, λ)]ρ(λ)dλ (15.7.16)


220 Quantum Mechanics Made SimpleWe have just added and subtracted identical terms in the above. After using the triangularinequality∫|E(a, b) − E(a, b ′ )| ≤ |A(a, λ)B(b, λ)||1 ± A(a ′ , λ)B(b ′ , λ)|ρ(λ)dλ∫+ |A(a, λ)B(b ′ , λ)||1 ± A(a ′ , λ)B(b, λ)|ρ(λ)dλ (15.7.17)Using the fact th<strong>at</strong>|AB| = 1, 1 ± AB ≥ 0 (15.7.18)we have∫|E(a, b) − E(a, b ′ )| ≤ [1 ± A(a ′ , λ)B(b ′ , λ)]ρ(λ)dλ∫+ [1 ± A(a ′ , λ)B(b, λ)]ρ(λ)dλ (15.7.19)The above is the same asIt is <strong>of</strong> the formwhich implies th<strong>at</strong>orConsequently, we have|E(a, b) − E(a, b ′ )| ≤ 2 ± [E(a ′ , b ′ ) + E(a ′ , b)] (15.7.20)|X| ≤ 2 ± Y (15.7.21)|X| ± Y ≤ 2 (15.7.22)|X + Y | ≤ |X| + |Y | ≤ 2 (15.7.<strong>23</strong>)|E(a, b) − E(a, b ′ ) + E(a ′ , b ′ ) + E(a ′ , b)| ≤ 2 (15.7.24)which is the Clauser-Horne-Shimony-Holt (CHSH) inequality.In (15.7.20), if a ′ = b ′ = c, then (15.7.20) becomes|E(a, b) − E(a, c)| ≤ 2 ± [E(c, c) + E(b, c)] (15.7.25)From (15.7.11), E(c, c) = 1. We pick the smaller <strong>of</strong> the right-hand side <strong>of</strong> (15.7.25) and arrive<strong>at</strong>|E(a, b) − E(a, c)| ≤ 1 − E(b, c) (15.7.26)


Quantum Inform<strong>at</strong>ion and Quantum Interpret<strong>at</strong>ion 221Figure 15.9: The case for a, b, and c where the quantum mechanics prediction viol<strong>at</strong>es theBell’s inequality.The above is the Bell’s inequality. It can be easily shown th<strong>at</strong> it cannot be s<strong>at</strong>isfied byquantum mechanics th<strong>at</strong> predicts (15.7.6). Say if we pick a, b, and c as shown in the Figure15.9, then according to quantum mechanics or (15.7.6),E(a, b) = cos(90 ◦ ) = 0 (15.7.27)E(a, c) = cos(45 ◦ ) = 1 √2(15.7.28)E(b, c) = cos(45 ◦ ) = 1 √2(15.7.29)Then in (15.7.26), we have∣ 0 − √ 1 ∣ ∣∣∣= 0.707 ≰ 1 − √ 1 = 0.293 (15.7.30)2 2In experimental tests <strong>of</strong> Bell’s theorem, quantum mechanics triumphs over hidden variabletheory so far. 6 Hence, the spookiness <strong>of</strong> ghost-angel st<strong>at</strong>es will prevail in quantum mechanics.15.8 A Final Word on Quantum ParallelismQuantum interpret<strong>at</strong>ion gives rise to the spookiness <strong>of</strong> quantum mechanics, in a way givingit the capability th<strong>at</strong> empowers ghosts and angels. Quantum parallelism is such an empowerment.It reminds me <strong>of</strong> a novel th<strong>at</strong> I have read when I was young on the Monkey King. ThisMonkey King is <strong>of</strong> Indian origin, but has perme<strong>at</strong>ed Chinese culture to take on a differentpersona. He is known as Hanuman in India. According to the story in The Journey to theWest, this Monkey King had unusual capabilities. One <strong>of</strong> them was th<strong>at</strong> he could pull astrand <strong>of</strong> hair from his body, and with a puff <strong>of</strong> air, he could turn it into many duplic<strong>at</strong>es <strong>of</strong>6 see a paper by A. Aspect, 1982.


222 Quantum Mechanics Made Simplehimself. He could then fight his enemies from all angles and all sides. He thus made himselfinvincible.He was arrogant, mischievous, and wreaked havoc in the Heavenly Palace where the othergods lived. Finally, he could only be tamed and subdued by Lord Buddha, summoning himto accompany and protect the monk Xuan Zang in his treacherous journey to collect theBuddhist Sutra from the West (in this case India).Oh Lord, if we were to empower ourselves with the capabilities <strong>of</strong> ghosts and angels, whois there to curb our power!


Quantum Inform<strong>at</strong>ion and Quantum Interpret<strong>at</strong>ion 2<strong>23</strong>Figure 15.10: The Monkey King sending replicas <strong>of</strong> himself against his enemy, defe<strong>at</strong>ingeveryone in his p<strong>at</strong>h.


224 Quantum Mechanics Made Simple


Appendix AGaussian Wave PacketA.1 IntroductionIn order to rel<strong>at</strong>e the wave n<strong>at</strong>ure to the particle n<strong>at</strong>ure <strong>of</strong> an object, it is necessary to obtaina wave packet picture <strong>of</strong> the object. Experimental observ<strong>at</strong>ion has indic<strong>at</strong>ed th<strong>at</strong> electrons arehighly localized objects, so are photons. A wavefunction with a pure or single wavenumberk has equal amplitude everywhere. According to the probabilistic interpret<strong>at</strong>ion <strong>of</strong> quantummechanics, it is untenable physically to think <strong>of</strong> a particle as all pervasive and can be foundequally likely everywhere. A photon wave can have many k wavenumbers, and they can belinearly superposed to form a localized photon wave packet. Since the E-k rel<strong>at</strong>ionship <strong>of</strong>the photon wave is linear (or almost linear in the m<strong>at</strong>erial-media case), the form<strong>at</strong>ion <strong>of</strong> sucha wave packet is quite straightforward. It is less clear for an electron wave, since the E-krel<strong>at</strong>ionship is nonlinear.In this appendix, we will explain the particle n<strong>at</strong>ure <strong>of</strong> an electron in the classical limitwhen the momentum <strong>of</strong> the electron becomes large. One expects th<strong>at</strong> the wavefunctionresembles th<strong>at</strong> <strong>of</strong> a localized particle in this limit so th<strong>at</strong> the electron is found with highprobability only in a confined loc<strong>at</strong>ion. This understanding can be achieved by studying theGaussian wave packet solution <strong>of</strong> Schrödinger equ<strong>at</strong>ion.When the Schrödinger equ<strong>at</strong>ion is solved in vacuum for electron, the solution ise ik·rThis unlocalized solution means th<strong>at</strong> the electron can be everywhere. A more realistic wavefunctionfor an electron is a wave packet which is localized. This is more akin to the motionth<strong>at</strong> an electron is a localized particle. A wave packet can be constructed by linear superposingwaves <strong>of</strong> different momenta k or different energies ω. We will derive the Gaussian wavepacket solution to Schrödinger equ<strong>at</strong>ion. This can be constructed by studying the solution tothe wave equ<strong>at</strong>ion.225


226 Quantum Mechanics Made SimpleA.2 Deriv<strong>at</strong>ion from the Wave Equ<strong>at</strong>ionIt is well known th<strong>at</strong> the wave equ<strong>at</strong>ion( ∂2∂x 2 + ∂2∂y 2 + ∂2∂z 2 + k2 )φ (x, y, z) = 0(A.2.1)admits solution <strong>of</strong> the formφ (x, y, z) = C e±ik √x 2 +y 2 +(z−ib) 2√x 2 + y 2 + (z − ib) 2 (A.2.2)for√x 2 + y 2 + (z − ib) 2 ≠ 0(A.2.3)The above corresponds to a spherical wave where the source point is loc<strong>at</strong>ed in the complexcoordin<strong>at</strong>es x = 0, y = 0, and z = ib. It was first suggested by G. Deschamps. From theabove, it is quite clear th<strong>at</strong>, after letting z = vt,√φ (x, y, vt) = A e±ik x 2 +y 2 +(vt−ib) 2(A.2.4)√x 2 + y 2 + (vt − ib) 2is a solution to ( ∂2∂x 2 + ∂2∂y 2 + 1 ∂ 2 )v 2 ∂t 2 + k2 φ (x, y, vt) = 0The above equ<strong>at</strong>ion can be factored( √) ( √)v −1 ∂ t − i k 2 + ∂x 2 + ∂y2 v −1 ∂ t + i k 2 + ∂x 2 + ∂y2 φ (x, y, vt) = 0(A.2.5)(A.2.6)where ∂ t = ∂/∂t, ∂x 2 = ∂ 2 /∂x 2 and so on. In the above, function <strong>of</strong> an oper<strong>at</strong>or has meaningonly when the function is Taylor expanded into an algebraic series. One can assume th<strong>at</strong>k 2 → ∞, while ∂x 2 and ∂y 2 are small. 1 Taylor expanding the above and keeping leading orderterms only, we have(v −1 ∂ t − ik − i (∂22k x + ∂y) ) ( 2 v −1 ∂ t + ik + i (∂22k x + ∂y2 ) ) φ (x, y, vt) ∼ = 0 (A.2.7)In (A.2.4), we can let 2 |vt − ib| 2 ≫ x 2 + y 2 to arrive <strong>at</strong> the approxim<strong>at</strong>ionφ (x, y, vt) ≈ A e± (ik(vt−ib)+ik x2 +y 2 )vt−ibvt − ib(A.2.8)1 This is known as the paraxial wave approxim<strong>at</strong>ion. It implies th<strong>at</strong> the vari<strong>at</strong>ion <strong>of</strong> the solution is mainlyin the vt direction with slow vari<strong>at</strong>ion in the x and y directions.2 It can be shown th<strong>at</strong> when this is valid, the paraxial wave approxim<strong>at</strong>ion used in (A.2.7) is good.


Quantum Inform<strong>at</strong>ion and Quantum Interpret<strong>at</strong>ion 227It can be shown th<strong>at</strong> when we pick the plus sign above in the ± sign, the above is the exactsolution to (v −1 ∂ t − ik − i (∂22k x + ∂y2 ) ) φ + (x, y, vt) = 0 (A.2.9)whereandφ + (x, y, vt) = Ae ik(vt−ib) eik x2 +y 22(vt−ib)2 (vt − ib) = eik(vt−ib) ψ (x, y, t) (A.2.10)ψ (x, y, t) = −A 0ibe ik x2 +y 22(vt−ib)(vt − ib)(A.2.11)Furthermore, ψ (x, y, t) is an exact solution to(v −1 ∂ t − i (∂22k x + ∂y2 ) ) ψ (x, y, t) = 0 (A.2.12)If we multiply the above by iv, the above becomes(i∂ t + v (∂22k x + ∂y2 ) ) ψ (x, y, t) = 0 (A.2.13)By letting v = k/m, the above becomes(i∂ t + 2 (∂22m x + ∂y2 ) ) ψ (x, y, t) = 0 (A.2.14)which is just Schrödinger equ<strong>at</strong>ion. Equ<strong>at</strong>ion (A.2.11) represents the Gaussian wave packetsolution <strong>of</strong> Schrödinger equ<strong>at</strong>ion. It can be studied further to elucid<strong>at</strong>e its physical contents.A.3 Physical Interpret<strong>at</strong>ionIn (A.2.11), one can writewherek ( x 2 + y 2)2 (vt − ib) = k ( x 2 + y 2) (vt + ib)2 (v 2 t 2 + b 2 = k ( x 2 + y 2)+ i x2 + y 2)2R W 2R = v2 t 2 + b 2, W 2 = 2b (1 + v2 t 2 )vtk b 2Then Gaussian wave packet can be more suggestively written asψ (x, y, t) =A 0√1 + v2 t 2 /b 2 e− x2 +y 2W 2 e ik x2 +y 22R e−iϕ(t)(A.3.1)(A.3.2)(A.3.3)whereϕ(t) = tan −1 (vt/b)(A.3.4)


228 Quantum Mechanics Made SimpleThe above reveals a wave packet which is Gaussian tapered with width W and modul<strong>at</strong>edby oscill<strong>at</strong>ory function <strong>of</strong> space and time. However, the width <strong>of</strong> this packet is a function <strong>of</strong>time as indic<strong>at</strong>ed by (A.3.2).At vt = 0, the width <strong>of</strong> the packet is given byW 0 =√2bk , or b = 1 2 kW 2 0 (A.3.5)This width can be made independent <strong>of</strong> k if b is made proportional to k. Nevertheless, astime progresses with vt > 0, the width <strong>of</strong> the packet grows according to (A.3.2). However, tomaintain a fixed-width W 0 , it is necessary th<strong>at</strong> b → ∞ as k → ∞. Subsequently, the effect <strong>of</strong>v 2 t 2 /b 2 becomes small in (A.3.2) as b → ∞. This means th<strong>at</strong> the width <strong>of</strong> the Gaussian wavepacket remains almost constant for the time when vt ≤ b, but b is a large number proportionalto k. The dur<strong>at</strong>ion over which the Gaussian wave packet’s width remains unchange becomesincreasingly longer as k becomes larger.In the above, k represents the momentum <strong>of</strong> a particle. When the particle carries highmomentum, it can be represented by a Gaussian wave packet th<strong>at</strong> hardly changes in shapewith respect to time. This is wh<strong>at</strong> is expected <strong>of</strong> a classical picture <strong>of</strong> a moving particle: theGaussian wave packet does reproduce the classical picture <strong>of</strong> a high momentum particle.It is quite easy to design a wave packet for a photon th<strong>at</strong> does not spread with time bylinear superposing waves with different frequencies. This is because the ω-k diagram (or E-kdiagram, since E = ω) for photons is a straight line. It implies th<strong>at</strong> all waves with differentk’s travel with the same phase velocity. These Fourier modes are locked in phase, and thepulse shape does not change as the wave packet travels.ωFigure A.1: ω − k diagram for a photon wave which is a straight line.kBut for electron wave in a vacuum, the ω-k diagram is quadr<strong>at</strong>ic. It implies th<strong>at</strong> waveswith different k numbers travel with different phase velocity, giving rise to distortion <strong>of</strong> thepulse shape. But if we have a narrow band pulse oper<strong>at</strong>ing in the high k regime, the ω-kdiagram is quasi-linear locally, and there should be little pulse spreading. Hence, one canconstruct a quasi-distortionless pulse in the high k regime.


Quantum Inform<strong>at</strong>ion and Quantum Interpret<strong>at</strong>ion 229ωΔωΔkkFigure A.2: ω − k diagram <strong>of</strong> an electron wave in vacuum.A.4 Stability <strong>of</strong> the Plane Wave SolutionIt is quite obvious from our understanding <strong>of</strong> quantum mechanics, wave packets, and coherence,th<strong>at</strong> the plane wave solution e ikx is not stable as k → ∞. One can always expresse ikx =∫ ∞−∞dx ′ e ikx′ δ(x − x ′ )(A.4.1)If one can think <strong>of</strong> δ(x − x ′ ) as the limiting case <strong>of</strong> a wave packet, the above implies th<strong>at</strong>a plane wave can be expanded as a linear superposition <strong>of</strong> wave packets <strong>at</strong> each loc<strong>at</strong>ion x,but bearing a phase exp(ikx). As k → ∞, this phase is rapidly varying among the differentwave packets. Hence, their coherence is extremely difficult to maintain, and upset easilyby coupling to other quantum systems th<strong>at</strong> exists in the environment. In other words, theparticle cannot be stably maintained in the plane-wave st<strong>at</strong>e classically: it has to collapse toa wave-packet st<strong>at</strong>e. Consequently, classically, particles are localized when its momentum kbecomes large.


<strong>23</strong>0 Quantum Mechanics Made Simple


Appendix BGener<strong>at</strong>ors <strong>of</strong> Transl<strong>at</strong>or andRot<strong>at</strong>ionB.1 Infinitesimal Transl<strong>at</strong>ionFrom Taylor series expansion, we havef(x + a) = f(x) + af(x) + a22! f 2 (x) + · · ·[= 1 + a ddx + a2 d 2 ]2! dx 2 + · · · f(x)= e a ddx f(x)(B.1.1)The exponential to an oper<strong>at</strong>or is interpreted as a Taylor series when it needs to be evalu<strong>at</strong>ed.The above can be written with a momentum oper<strong>at</strong>or (assuming th<strong>at</strong> = 1)or in Dirac not<strong>at</strong>ionf(x + a) = e iaˆp f(x)(B.1.2)|f a 〉 = e iaˆp |f〉 (B.1.3)where |f a 〉 is the st<strong>at</strong>e vector represent<strong>at</strong>ion <strong>of</strong> the function f(x + a). If a Hamiltonian istransl<strong>at</strong>ional invariant, it will commute with the transl<strong>at</strong>ion oper<strong>at</strong>or, namelyĤe iaˆp |f〉 = e iaˆp Ĥ|f〉(B.1.4)or [ ]e iaˆp , Ĥ = 0 (B.1.5)For example, if the Hamiltonian is such th<strong>at</strong> − 1 2 d2 /dx 2 , then it is quite clear th<strong>at</strong> the left-handside <strong>of</strong> (B.1.4) isd 2d 2− 1 eiaˆp 2 dx 2 f(x) = − 1 2 dx 2 f(x + a) = −1 2 f ′′ (x + a)(B.1.6)<strong>23</strong>1


<strong>23</strong>2 Quantum Mechanics Made SimpleThe right-hand side <strong>of</strong> (B.1.4) is[e iaˆp − 1 d 2 ]2 dx 2 f(x)= e[− iaˆp 1 ]2 f ′′ (x) = − 1 2 f ′′ (x + a) (B.1.7)which is the same as the left-hand side.By assuming th<strong>at</strong> e iaˆp ≈ 1 + iaˆp + · · · , the above also means th<strong>at</strong>Ĥ ˆp = ˆpĤ, [Ĥ, ˆp] = 0(B.1.8)This means th<strong>at</strong> ˆp is a constant <strong>of</strong> motion as shown in Section 5.8, ord〈ˆp〉dt= 0 (B.1.9)In other words, momentum is conserved in a system where the Hamiltonian is transl<strong>at</strong>ionalinvariant.B.2 Infinitesimal Rot<strong>at</strong>ionSimilarly, the L z oper<strong>at</strong>or is −i ∂∂φwhich is a gener<strong>at</strong>or <strong>of</strong> rot<strong>at</strong>ion,1f(φ + α) = e α ddφ f(φ) = eiα ˆL zf(φ)(B.2.1)Going through the deriv<strong>at</strong>ion as we have before, for a Hamiltonian th<strong>at</strong> is rot<strong>at</strong>ionally symmetricabout the z axis, then it commutes with e iα ˆL z. In this caseor Taylor series expanding the above, we derive th<strong>at</strong>[e iα ˆL z, Ĥ] = 0 (B.2.2)[ˆL z , Ĥ] = 0(B.2.3)Hence, the ẑ component <strong>of</strong> the angular momentum is conserved for a system th<strong>at</strong> is rot<strong>at</strong>ionallysymmetric about the z axis.The above has been motiv<strong>at</strong>ed by ˆL z th<strong>at</strong> follows from the orbital angular momentum,whose form has been motiv<strong>at</strong>ed by the classical angular momentum. Wh<strong>at</strong> if the angularmomentum has no classical analogue like the spin angular momentum? Or the st<strong>at</strong>e vectormay not be written using wavefunctions <strong>at</strong> all. In this case, we can postul<strong>at</strong>e a generalizedgener<strong>at</strong>or <strong>of</strong> rot<strong>at</strong>ion <strong>of</strong> the forme iαĴz |j〉(B.2.4)The above will take the angular momentum st<strong>at</strong>e |j〉 and gener<strong>at</strong>e rot<strong>at</strong>ed st<strong>at</strong>e about the zaxis. We postul<strong>at</strong>e the above form for three reasons:1 Assume again th<strong>at</strong> = 1.


Quantum Inform<strong>at</strong>ion and Quantum Interpret<strong>at</strong>ion <strong>23</strong>31. The gener<strong>at</strong>or <strong>of</strong> rot<strong>at</strong>ion for orbital angular momentum is already <strong>of</strong> this form. It mustbe the special case <strong>of</strong> the above form.2. If the above gener<strong>at</strong>or commutes with the Hamiltonian with similar rot<strong>at</strong>ional symmetry,this component <strong>of</strong> angular momentum will be conserved.3. The rot<strong>at</strong>ion will cause the expect<strong>at</strong>ion value <strong>of</strong> the angular momentum vector, anobservable, to rot<strong>at</strong>e according to a coordin<strong>at</strong>e rot<strong>at</strong>ion about the z axis.It is easy to see th<strong>at</strong> the expect<strong>at</strong>ion value <strong>of</strong> the Ĵz oper<strong>at</strong>or remains unchanged underthis rot<strong>at</strong>ion. We can easily show th<strong>at</strong>〈j|e −iαĴzĴ ze iαĴz |j〉 = 〈j|Ĵz|j〉(B.2.5)In the above, functions <strong>of</strong> oper<strong>at</strong>or Ĵz commute with each other, a fact th<strong>at</strong> can be easilyproved by expanding the functions as a power series. The conjug<strong>at</strong>e transpose <strong>of</strong> e iαĴz ise −iαĴz since Ĵz is Hermitian because it represents an observable. Hence, the last equality in(B.2.5) follows.B.3 Deriv<strong>at</strong>ion <strong>of</strong> Commut<strong>at</strong>ion Rel<strong>at</strong>ionsIn general, we can define the angular momentum oper<strong>at</strong>or to beĴ = iĴx + jĴy + kĴz(B.3.1)The above is an observable so are the components <strong>of</strong> the oper<strong>at</strong>or in the x, y, z directions.Hence, the expect<strong>at</strong>ion value <strong>of</strong> the above with respect to the st<strong>at</strong>e |j〉 gives rise to〈Ĵ〉 = i〈Ĵx〉 + j〈Ĵy〉 + k〈Ĵz〉 = iJ x + jJ y + kJ z = J(B.3.2)where we denote the expect<strong>at</strong>ion values <strong>of</strong> Ĵi with scalar number J i , and th<strong>at</strong> <strong>of</strong> Ĵ with J. Wewill test point 3 above with respect to this vector J. This vector will have to rot<strong>at</strong>e accordingto coordin<strong>at</strong>e rot<strong>at</strong>ion as the st<strong>at</strong>e vector is rot<strong>at</strong>ed according to (B.2.4).If e iαĴz is a gener<strong>at</strong>or <strong>of</strong> rot<strong>at</strong>ion about the z axis, it will leave J z unchanged as shownabove. But it will not leave J x and J y components unchanged. Now, we can find the expect<strong>at</strong>ionvalue <strong>of</strong> Ĵx under rot<strong>at</strong>ion about z axis, or th<strong>at</strong>J x = 〈j|Ĵx|j〉, J ′ x = 〈j|e −iαĴzĴ xe iαĴz |j〉 (B.3.3)Before rot<strong>at</strong>ion, this would have represented the expect<strong>at</strong>ion values <strong>of</strong> the x component <strong>of</strong> J,the angular momentum. After rot<strong>at</strong>ion,J = i ′ J ′ x + j ′ J ′ y + kJ z(B.3.4)and J ′ x will contain both the x and y component <strong>of</strong> J. When we rot<strong>at</strong>e the st<strong>at</strong>e vector byangle α, the expect<strong>at</strong>ion value <strong>of</strong> the vector J will rot<strong>at</strong>e by α in the same direction. Butfrom the Figure B.1, it is clear th<strong>at</strong>J ′ x = J x cos α + J y sin α(B.3.5)


<strong>23</strong>4 Quantum Mechanics Made SimpleThe equivalence <strong>of</strong> rot<strong>at</strong>ions in (B.3.3) and (B.3.5) can be proved by lengthy algebra. Tosimplify the algebra, we can show the equivalence for infinitesimal rot<strong>at</strong>ion. To this end, weassume th<strong>at</strong> α is small so th<strong>at</strong> we keep only the first two terms <strong>of</strong> the Taylor expansion fore iαĴz or e iαĴz ≈ 1 + iαĴz. Then,J ′ x = 〈j|Ĵx − iαĴzĴx + iαĴxĴz + · · · |j〉(B.3.6)For small α, (B.3.5) becomesBut (B.3.6) can be written asJ ′ x ∼ = J x + αJ y(B.3.7)J ′ x = 〈j|Ĵx|j〉 − iα〈j|ĴzĴx − ĴxĴz|j〉(B.3.8)Figure B.1: Coordin<strong>at</strong>e rot<strong>at</strong>ion <strong>of</strong> the xy plane about the z axis by angle α.Comparing (B.3.7) and (B.3.8), we haveĴ z Ĵ x − ĴxĴz = iĴy(B.3.9)We can let ≠ 1 to arrive back <strong>at</strong> the previously postul<strong>at</strong>ed commut<strong>at</strong>or rel<strong>at</strong>ions. Theabove is arrived <strong>at</strong> using only rot<strong>at</strong>ional symmetry argument th<strong>at</strong> the angular momentumoper<strong>at</strong>or is a gener<strong>at</strong>or <strong>of</strong> rot<strong>at</strong>ion. The other commut<strong>at</strong>ion rel<strong>at</strong>ions for angular oper<strong>at</strong>orscan be derived by similar arguments. In general, for all oper<strong>at</strong>ors th<strong>at</strong> represent angularmomentum, we have[Ĵx Ĵy], = iĴz,[Ĵy Ĵz], = iĴx,[Ĵz Ĵx], = iĴy (B.3.10)From the above, we can define raising and lowering oper<strong>at</strong>ors and use the ladder approach toderive the properties <strong>of</strong> the eigenst<strong>at</strong>es <strong>of</strong> angular momentum oper<strong>at</strong>ors.


Appendix CQuantum St<strong>at</strong>istical MechanicsC.1 IntroductionWe would like to derive Maxwell-Boltzmann, Fermi-Dirac, and Bose-Einstein distributions.It is based on st<strong>at</strong>istical mechanics th<strong>at</strong> assumes th<strong>at</strong> all identical energy levels have equallikelihood <strong>of</strong> being occupied. Only the counting is based on quantum mechanics.We consider a quantum system with energy levels E 1 , E 2 , E 3 , · · · . Each level E i couldhave degeneracy <strong>of</strong> d i , i = 1, 2, 3, · · · , and is filled by N i , i = 1, 2, 3, · · · particles. Whenthe system is coupled to its environment, energy will be exchanged with its environment viacollision, vibr<strong>at</strong>ion, and radi<strong>at</strong>ion. At thermal equilibrium, there is no net energy flowing intoand out <strong>of</strong> the system. We expect th<strong>at</strong>E =∞∑E n N nn=1(C.1.1)to be a constant due to energy conserv<strong>at</strong>ion. Also we expect the number <strong>of</strong> particles to remainconstant, or th<strong>at</strong>N =∞∑n=1N n(C.1.2)due to particle conserv<strong>at</strong>ion.The particles will acquire energy from their environment due to energy exchange, and distributethemselves across the energy levels subject to the above constraints, but also accordingto the availability <strong>of</strong> energy levels. Energy levels with higher degeneracies are more likely tobe filled. We will consider three different cases: (i) when the particles are distinguishable, (ii)when they are identical fermions, and (iii) when they are identical bosons.C.1.1Distinguishable ParticlesWe will find the configur<strong>at</strong>ion function Q(N 1 , N 2 , N 3 , · · · ) whose value equals the number <strong>of</strong>ways th<strong>at</strong> there are N 1 particles in E 1 , N 2 particles in E 2 , N 3 particles in E 3 and so on. In<strong>23</strong>5


<strong>23</strong>6 Quantum Mechanics Made SimpleE nd n, N nE 3d 3, N 3E 2d 2, N 2E 1d 1, N 1Figure C.1: Particles are distributed among different energy levels according to quantumst<strong>at</strong>istics. An energy level is denoted with E i with degeneracy d i and N i particles occupyingit.order to fill energy level E 1 with N 1 particles, the number <strong>of</strong> ways is( ) N N!=N 1 N 1 ! (N − N 1 )!(C.1.3)If E 1 has degeneracy d 1 , all the degener<strong>at</strong>e levels are equally likely to be filled. Then thereare d N11 ways th<strong>at</strong> the N 1 particles can go into E 1 level (see Section 10.5). Hence, the number<strong>of</strong> ways th<strong>at</strong> E 1 can be filled isQ E1 =The number <strong>of</strong> ways th<strong>at</strong> E 2 can be filled isN!d N11N 1 ! (N − N 1 )!(C.1.4)Then the total number <strong>of</strong> waysQ E2 =Q(N 1 , N 2 , N 3 , · · · ) = Q E1 Q E2 Q E3 · · ·(N − N 1 )!d N22N 2 ! (N − N 1 − N 2 )!(C.1.5)N!d N11 (N − N 1 )!d N22 (N − N 1 − N 2 )!d N33=N 1 ! (N − N 1 )! N 2 ! (N − N 1 − N 2 )! N 3 ! (N − N 1 − N 2 − N 3 )! · · ·= N! dN1 1 dN2 2 dN3 3 · · ·∞N 1 !N 2 !N 3 ! · · · = N! ∏ d Nnn(C.1.6)N n !n=1


Quantum Inform<strong>at</strong>ion and Quantum Interpret<strong>at</strong>ion <strong>23</strong>7Notice th<strong>at</strong> the above is independent <strong>of</strong> the order in which the levels are filled, as expected.The above can also be interpreted in a different light: There are d Nnn ways th<strong>at</strong> N n particlescan fit into the E n energy level. But order is unimportant and a division by N n ! is necessary.But there are N! ways th<strong>at</strong> these distinguishable particles can be selected in an ordered way,and hence, a prefactor <strong>of</strong> N! is needed in the above.C.1.2Identical FermionsDue to Pauli exclusion principle, each energy level can admit only one particle. In this case,the number <strong>of</strong> ways energy E n can be filled is (see Section 10.5)Q En =d n !N n ! (d n − N n )! , d n ≥ N n(C.1.7)(The above has value 1 when d n = 1.) The total number <strong>of</strong> ways is then∏Q (N 1 , N 2 , N 3 , ...) = ∞ d n !N n ! (d n − N n )!n=1(C.1.8)C.1.3Identical BosonsFor bosons, repetition for filling a given st<strong>at</strong>e is allowed, but the particles are indistinguishablefrom each other. Then when N n particles are picked to fill the d n degener<strong>at</strong>e E n levels, thefirst particle from N n particles has d n slots to fill in, while the second particle has d n + 1 slotsto fill in, and the third particle has d n + 2 slots to fill in and so on, since repetition is allowed.Th<strong>at</strong> is the new particle can take the position <strong>of</strong> the old particle as well. So the number <strong>of</strong>ways th<strong>at</strong> E n with d n degeneracy can be filled is (see Section 10.5)Q En = d n (d n + 1) (d n + 2) ... (d n + N n − 1)N n != (N n + d n − 1)!N n ! (d n − 1)!(C.1.9)(C.1.10)ThereforeQ (N 1 , N 2 , N 3 , ...) = ∞ ∏n=1(N n + d n − 1)!N n ! (d n − 1)!(C.1.11)Unlike the distinguishable particle case, no prefactor <strong>of</strong> N! is needed for the identicalparticle case, since the order with which they are selected is unimportant.C.2 Most Probable Configur<strong>at</strong>ionA way <strong>of</strong> filling the energy levels {E 1 , E 2 , E 3 , ...} with {N 1 , N 2 , N 3 , ...} is called a configur<strong>at</strong>ion.The most likely configur<strong>at</strong>ion to be filled is the one with the largest Q (N 1 , N 2 , N 3 , ...).At st<strong>at</strong>istical equilibrium, the system will gravit<strong>at</strong>e toward this configur<strong>at</strong>ion. Hence, to findthis configur<strong>at</strong>ion, we need to maximize Q with respect to different {N 1 , N 2 , N 3 , ...} subjectto the energy conserv<strong>at</strong>ion constraint (C.1.1) and particle conserv<strong>at</strong>ion constraint (C.1.2).


<strong>23</strong>8 Quantum Mechanics Made SimpleWe use the Lagrange multiplier technique to find the optimal Q subject to constraints(C.1.1) and (C.1.2). We define a function[ ] []∑G = ln Q + α N − ∞ ∑N n + β E − ∞ N n E n (C.2.1)n=1The optimal Q value is obtained by solving[ ] []∂G∂G∂N n= 0, n = 1, 2, 3, · · · ,∂α = ∑N − ∞ ∂GN n = 0,∂β = ∑E − ∞ N n E n = 0n=1(C.2.2)Notice th<strong>at</strong> the constraint conditions for particle and energy conserv<strong>at</strong>ions are imposed in thesecond and third equ<strong>at</strong>ions above.C.2.1In this case,Distinguishable ParticlesWe use Stirling’s formula th<strong>at</strong>Consequently,From the aboveAs a result, we have∑G = ln N! + ∞ [N n ln d n − ln N n !] (C.2.3)n=1[ ] []∑+ α N − ∞ ∑N n + β E − ∞ N n E n (C.2.4)n=1ln (z!) ≈ z ln z − zn=1n=1n=1(C.2.5)∑G = ∞ [N n ln d n − N n ln N n + N n − αN n − βE n N n ] (C.2.6)n=1+ ln N! + αN + βE (C.2.7)∂G∂N n= ln d n − ln N n − α − βE n = 0N n = d n e −(α+βEn)The above is the precursor to the Maxwell-Boltzmann distribution from first principles.(C.2.8)(C.2.9)C.2.2 Identical FermionsIn this case∞∑G = [ln(d n !) − ln(N n !) − ln((d n − N n )!)]n=1[] []∞∑∞∑+ α N − N n + β E − N n E nn=1n=1(C.2.10)(C.2.11)


Quantum Inform<strong>at</strong>ion and Quantum Interpret<strong>at</strong>ion <strong>23</strong>9After applying Stirling’s formula to N n dependent terms, we have∞∑G = [ln d n ! − N n ln N n + N n − (d n − N n ) ln(d n − N n ) + (d n − N n ) − αN n − βN n E n ]n=1From the above,(C.2.12)+ αN + βE (C.2.13)or∂G∂N n= − ln N n + ln(d n − N n ) − α − βE n = 0(C.2.14)N n =d n1 + e α+βEn (C.2.15)In the above deriv<strong>at</strong>ion, we have assumed th<strong>at</strong> d n is large so th<strong>at</strong> N n is large since N n ≤ d nin (C.1.7). The above is the precursor to the Fermi-Dirac distribution.C.2.3 Identical BosonsIn this case∞∑G = [ln((N n + d n − 1)!) − ln(N n !) − ln((d n − 1)!)]n=1[] []∞∑∞∑+ α N − N n + β E − N n E nn=1n=1(C.2.16)(C.2.17)With Stirling’s approxim<strong>at</strong>ion to the N n dependent terms,∞∑G = [(N n + d n − 1) ln(N n + d n − 1) − (N n + d n − 1) − N n ln N n + N nn=1− ln((d n − 1)!) − αN n − βN n E n ] + αN + βE(C.2.18)(C.2.19)Then∂G∂N n= ln(N n + d n − 1) − ln(N n ) − α − βE n = 0(C.2.20)yieldingN n = d n − 1e α+βEn − 1 ≃ d ne α+βEn − 1(C.2.21)where we assume th<strong>at</strong> d n ≫ 1. The above is the precursor to the Bose-Einstein distribution.


240 Quantum Mechanics Made SimpleC.3 The Meaning <strong>of</strong> α and βWe will apply the above to a simple quantum system in order to infer wh<strong>at</strong> α and β shouldbe. To find them, we need to solve the constraint equ<strong>at</strong>ions (C.1.1) and (C.1.2) th<strong>at</strong> followfrom the optimiz<strong>at</strong>ion <strong>of</strong> (C.2.1). To this end, we need to find the degeneracy per energyst<strong>at</strong>e, d k <strong>of</strong> a quantum system.We consider N electrons inside a bulk m<strong>at</strong>erial describable by a single electron in aneffective mass approxim<strong>at</strong>ion. In such a case, the kinetic energy <strong>of</strong> the electron is describedbyE k = 2 k 22m(C.3.1)If we assume periodic boundary conditions on a box <strong>of</strong> lengths L x , L y , and L z , then( ) 2 ( ) 2 ( ) 2k 2 2nx π 2ny π 2nz π= + +(C.3.2)L x L y L zIn the k space, there is a st<strong>at</strong>e associ<strong>at</strong>ed with a unit box <strong>of</strong> volume∆V k =(2π)3 = 8π3L x L y L z V(C.3.3)where V is the volume <strong>of</strong> the box <strong>of</strong> bulk m<strong>at</strong>erial. In k space, in a spherical shell <strong>of</strong> thickness∆k, the number <strong>of</strong> st<strong>at</strong>es in the neighborhood <strong>of</strong> k isd k = 4πk2 ∆k∆V k= V2π 2 k2 ∆kThen the total number <strong>of</strong> particles is, using (C.2.9) for Maxwell-Boltzmann,By using the fact th<strong>at</strong>I 1 =∫ ∞the above integr<strong>at</strong>es to0N = ∑ d k e −(α+βEk) =V ∑ 2π 2 e−α k 2 ∆ke −βE kkk= V ∫ ∞2π 2 e−α e −β2 k 2 /(2m) k 2 dke −Ak2 k 2 dk = − ddASimilarly, the total E is given byE = ∑ kE k N k = ∑ k0∫ ∞0(C.3.4)(C.3.5)e −Ak2 dk = − d √ √1 π πdA 2 A = 4 A− 3 2 (C.3.6)( ) 3mN = V e −α 22πβ 2E k d k e −(α+βE k) = − ddβ∑kd k e −(α+βE k) = − ddβ N(β)(C.3.7)(C.3.8)


Quantum Inform<strong>at</strong>ion and Quantum Interpret<strong>at</strong>ion 241It can be shown th<strong>at</strong>E = 3N2βFrom the equipartition theorem <strong>of</strong> energy from st<strong>at</strong>istical mechanics we know th<strong>at</strong>Hence, we conclude th<strong>at</strong>It is customary to writeEN = 3 2 k BTβ = 1k B T(C.3.9)(C.3.10)α = − µ(T )k B T(C.3.11)so th<strong>at</strong> for Maxwell-Boltzmann distribution,N n = d n e −(En−µ)/(k BT )(C.3.12)where µ(T ) is the chemical potential. When normalized with respect to degeneracy, we haven mb = e −(En−µ)/(k BT ) , Maxwell-Boltzmann (C.3.13)Fermi-Dirac and Bose-Einstein distribution become Maxwell-Boltzmann when (βE n +α) ≫ 1. So they have the same α and β. Therefore,n fd =n be =1e (En−µ)/(k BT ), Fermi-Dirac (C.3.14)+ 11e (En−µ)/(k BT ), Bose-Einstein (C.3.15)− 1When (E n − µ) ≫ k B T the above distributions resemble the Maxwell-Boltzmann distribution.This is because n becomes small per energy level, and the indistinguishability <strong>of</strong>the particles plays a less important role. When (E n − µ) < 0, the Fermi-Dirac distribution“freezes” to 1 for all energy levels below µ, since the Pauli exclusion principle only allowsone fermion per level. When (E n − µ) ≪ k B T , the Bose-Einstein distribution diverges, implyingth<strong>at</strong> the particles condense to an energy level close to µ. This condens<strong>at</strong>ion is morepronounced if k B T is small.For photons in a box, the number <strong>of</strong> photons need not be conserved as photons can freelythe box than entering the box. Hence, we can set α or µ, the chemical potential, to zero. Inthis case, the Bose-Einstein distribution becomes the Planck distribution.n pl =1e En/(k BT ), Planck (C.3.16)− 1


242 Quantum Mechanics Made Simple


Appendix DPhoton Polariz<strong>at</strong>ionD.1 IntroductionPhoton polariz<strong>at</strong>ion is used in many quantum inform<strong>at</strong>ion experiments. Hence, it is importantto understand the quantum n<strong>at</strong>ure <strong>of</strong> photon polariz<strong>at</strong>ion. In quantum electrodynamics, theelectric field is an observable. Its oper<strong>at</strong>or represent<strong>at</strong>ion can be written as (see (12.6.32))Ê = C [e x â x + e y â y ] + c.c.(D.1.1)for a mode th<strong>at</strong> is propag<strong>at</strong>ing in the z direction. We can ignore the constant C and the c.c.part for the time being, and consider the field oper<strong>at</strong>orÊ = e x â x + e y â y(D.1.2)To understand the st<strong>at</strong>es th<strong>at</strong> this oper<strong>at</strong>or can acts on, it is prudent to study the 2D quantumharmonic oscill<strong>at</strong>or first.D.2 2D Quantum Harmonic Oscill<strong>at</strong>orThe field oper<strong>at</strong>or expressed by (D.1.2) above is a consequence <strong>of</strong> a 2D harmonic oscill<strong>at</strong>or.It has a Hamiltonian given byĤ 2D = ˆp2 x2m + ˆp2 y2m + mω2 02 (x2 + y 2 ) = Ĥx + Ĥy (D.2.1)where ˆp x and ˆp y are respectively the momentum oper<strong>at</strong>ors, and hence, Ĥ x and Ĥy are respectivelythe harmonic oscill<strong>at</strong>or in 1D in the x and y directions. Similar to the definitionin the 1D case, we can rewrite the above as(Ĥ 2D = ω 0 â † xâ x + 1 ) (+ ω 0 â †2yâ y + 1 )(D.2.2)2243


244 Quantum Mechanics Made Simplewhere similar to (12.2.7)â † x = 1 √2(− ddξ x+ ξ x), â x = 1 √2( ddξ x+ ξ x)â † y = √ 1 (− d )+ ξ y , â y = 1 ( ) d√ + ξ y2 dξ y 2 dξ y(D.2.3)(D.2.4)ξ x =√ mω0 x, ξ y =The Hamiltonian for photons is given by (12.6.36) asĤ = ∑ (ω k â † k,sâk,s + 1 )2k,s√ mω0 y (D.2.5)(D.2.6)The above represents a chain <strong>of</strong> coupled quantum harmonic oscill<strong>at</strong>ors in space. Hence, we canstudy the case <strong>of</strong> (D.2.1) and (D.2.2) to better understand the more complex case describedby (D.2.6).The eigenst<strong>at</strong>es <strong>of</strong> Ĥx and Ĥy, which correspond to 1D harmonic oscill<strong>at</strong>ors, are given byĤ x |ψ mx 〉 = E mx |ψ mx 〉(D.2.7)Ĥ y |ψ ny 〉 = E ny |ψ ny 〉(D.2.8)where E mx = ( m + 2) 1 ω0 , E ny = ( n + 2) 1 ω0 . The eigenst<strong>at</strong>es <strong>of</strong> Ĥ2D <strong>of</strong> (D.2.2) are givenbyĤ 2D |ψ i 〉 = E i |ψ i 〉(D.2.9)By letting|ψ i 〉 = |ψ mx 〉|ψ ny 〉, (D.2.10)then E i = E mx + E ny . In general, the eigenst<strong>at</strong>es <strong>of</strong> Ĥ2D can be constructed from the directproduct <strong>of</strong> spaces spanned by the eigenst<strong>at</strong>es <strong>of</strong> Ĥx and Ĥy, or[(E i = E m,n = m + 1 ) (+ n + 1 )]ω 0(D.2.11)2 2The above eigenvalues are degener<strong>at</strong>e when m + n = M, a constant. For a given M,there are M + 1 degener<strong>at</strong>e st<strong>at</strong>es with energy E M = (M + 1)ω 0 . In general, a quantizedelectromagnetic field has two components, and it corresponds to a 2D quantum harmonicoscill<strong>at</strong>or. For a st<strong>at</strong>ionary st<strong>at</strong>e with energy E M , it corresponds to an M-photon st<strong>at</strong>e withM + 1 degeneracies. In general, an M-photon st<strong>at</strong>e needs to be expanded in terms <strong>of</strong> thesedegener<strong>at</strong>e st<strong>at</strong>esM∑|ψ M 〉 = c n |ψ M−n,x 〉|ψ ny 〉(D.2.12)n=0with the requirement th<strong>at</strong> ∑ n |c n| 2 = 1 and th<strong>at</strong> the above s<strong>at</strong>isfies rot<strong>at</strong>ional symmetryabout the z axis.


Quantum Inform<strong>at</strong>ion and Quantum Interpret<strong>at</strong>ion 245D.3 Single–Photon St<strong>at</strong>eWe consider the simpler case <strong>of</strong> a single photon, or th<strong>at</strong> M = m + n = 1. Then, there aretwo degener<strong>at</strong>e eigenst<strong>at</strong>es|1 x 〉|0 y 〉, |0 x 〉|1 y 〉 (D.3.1)For simplicity, we write these st<strong>at</strong>es as |1 x 〉, |1 y 〉 st<strong>at</strong>es suppressing the associ<strong>at</strong>ed groundst<strong>at</strong>es. Then, a single photon st<strong>at</strong>e polarized in the α direction can be expressed as|1 α 〉 = c x |1 x 〉 + c y |1 y 〉 (D.3.2)where |c x | 2 + |c y | 2 = 1. Altern<strong>at</strong>ively, we can write the above as|1 x ′〉 = cos φ|1 x 〉 + sin φ|1 y 〉 (D.3.3)The above implies th<strong>at</strong> a single photon polarized in the x ′ direction can be written as asuperposition <strong>of</strong> two single photon st<strong>at</strong>es, |1 x 〉 and |1 y 〉 representing a photon polarized inthe x and y directions respectively. The above is identical to the decomposition <strong>of</strong> the unitvectore x ′ = e x cos φ + e y sin φ (D.3.4)To check the consistency <strong>of</strong> this rule, we use it to write|1 x 〉 = cos φ|1 x ′〉 − sin φ|1 y ′〉 (D.3.5)|1 y 〉 = sin φ|1 x ′〉 + cos φ|1 y ′〉 (D.3.6)Upon substituting (D.3.5) and (D.3.6) into (D.3.3), consistency is found.Looking <strong>at</strong> the vector oper<strong>at</strong>or Ê, in (D.1.2), lettingthen, (D.1.2), becomese x = e x ′ cos φ − e y ′ sin φe y = e x ′ sin φ + e y ′ cos φ(D.3.7)(D.3.8)Ê = e x ′ (cos φâ x + sin φâ y ) + e y ′ (− sin φâ x + cos φâ y )(D.3.9)Letting (D.3.9) oper<strong>at</strong>e on (D.3.3), we have(Ê|1 x ′〉 = e x ′ cos 2 φ|0 x , 0 y 〉 + sin 2 φ|0 x , 0 y 〉 ) + e y ′ (− sin φ cos φ|0 x , 0 y 〉 + cos φ sin φ|0 x , 0 y 〉)(D.3.10)Note th<strong>at</strong> the |0 x 〉 and |0 y 〉 ground st<strong>at</strong>es have been suppressed in our not<strong>at</strong>ion in (D.3.2),but in (D.3.10), we write them explicitly. Hence (D.3.10) becomesÊ|1 x ′〉 = e x ′|0 x , 0 y 〉 = e x ′|0 x ′, 0 y ′〉 (D.3.11)Since there is only one ground st<strong>at</strong>e, it can be shown th<strong>at</strong> this ground st<strong>at</strong>e is rot<strong>at</strong>ionallysymmetric, and hence, the above equality.


246 Quantum Mechanics Made SimplewhereHad we written (D.3.9) asÊ = e x ′â x ′ + e y ′â y ′(D.3.12)â x ′ = cos φâ x + sin φâ y (D.3.13)â y ′ = − sin φâ x + cos φâ y (D.3.14)and apply (D.3.12) directly to |1 x ′〉, we would have gotten (D.3.11). The above shows theconsistency <strong>of</strong> (D.3.3) to (D.3.6). It is interesting to note th<strong>at</strong> |1 x 〉 , |1 y 〉, â x , and â y transformas unit vectors e x and e y under coordin<strong>at</strong>e rot<strong>at</strong>ion.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!