C/CS/Phys C191 Particle-in-a-box, Spin 10/02/08 Fall 2008 Lecture 11

Similar documents
1 Position Representation of Quantum State Function

1 Measurement and expectation values

1 Planck-Einstein Relation E = hν

i = cos 2 0i + ei sin 2 1i

1 Recall what is Spin

CHAPTER 8 The Quantum Theory of Motion

Introduction to Quantum Mechanics PVK - Solutions. Nicolas Lanzetti

Brief review of Quantum Mechanics (QM)

Lecture #5: Begin Quantum Mechanics: Free Particle and Particle in a 1D Box

C/CS/Phy191 Problem Set 6 Solutions 3/23/05

Mathematical Tripos Part IB Michaelmas Term Example Sheet 1. Values of some physical constants are given on the supplementary sheet

C/CS/Phys 191 Uncertainty principle, Spin Algebra 10/11/05 Fall 2005 Lecture 13

Physics 342 Lecture 17. Midterm I Recap. Lecture 17. Physics 342 Quantum Mechanics I

The Hydrogen Atom. Dr. Sabry El-Taher 1. e 4. U U r

Chemistry 432 Problem Set 4 Spring 2018 Solutions

The Postulates of Quantum Mechanics Common operators in QM: Potential Energy. Often depends on position operator: Kinetic Energy 1-D case: 3-D case

Angular Momentum Quantization: Physical Manifestations and Chemical Consequences

Lecture-XXVI. Time-Independent Schrodinger Equation

Waves and the Schroedinger Equation

The Schrödinger Equation

Introduction to Quantum Mechanics (Prelude to Nuclear Shell Model) Heisenberg Uncertainty Principle In the microscopic world,

Lecture #1. Review. Postulates of quantum mechanics (1-3) Postulate 1

Problem Set 5: Solutions

Quantum Physics III (8.06) Spring 2007 FINAL EXAMINATION Monday May 21, 9:00 am You have 3 hours.

Lecture 10: Solving the Time-Independent Schrödinger Equation. 1 Stationary States 1. 2 Solving for Energy Eigenstates 3

Chapter 7. Bound Systems are perhaps the most interesting cases for us to consider. We see much of the interesting features of quantum mechanics.

1 Fundamental physical postulates. C/CS/Phys C191 Quantum Mechanics in a Nutshell I 10/04/07 Fall 2007 Lecture 12

Quantum Mechanics: Particles in Potentials

Time part of the equation can be separated by substituting independent equation

If electrons moved in simple orbits, p and x could be determined, but this violates the Heisenberg Uncertainty Principle.

Lecture 2: simple QM problems

BASICS OF QUANTUM MECHANICS. Reading: QM Course packet Ch 5

Lecture 5: Harmonic oscillator, Morse Oscillator, 1D Rigid Rotor

df(x) = h(x) dx Chemistry 4531 Mathematical Preliminaries Spring 2009 I. A Primer on Differential Equations Order of differential equation

Physics 113!!!!! Spring 2009!! Quantum Theory Seminar #9

Chem 467 Supplement to Lecture 19 Hydrogen Atom, Atomic Orbitals

P3317 HW from Lecture 15 and Recitation 8

E = hν light = hc λ = ( J s)( m/s) m = ev J = ev

d 3 r d 3 vf( r, v) = N (2) = CV C = n where n N/V is the total number of molecules per unit volume. Hence e βmv2 /2 d 3 rd 3 v (5)

Probability and Normalization

The Schrodinger Equation and Postulates Common operators in QM: Potential Energy. Often depends on position operator: Kinetic Energy 1-D case:

Quantum Physics II (8.05) Fall 2002 Assignment 11

P3317 HW from Lecture and Recitation 7

The 3 dimensional Schrödinger Equation

(Refer Slide Time: 1:20) (Refer Slide Time: 1:24 min)

where A and α are real constants. 1a) Determine A Solution: We must normalize the solution, which in spherical coordinates means

Applied Nuclear Physics (Fall 2006) Lecture 2 (9/11/06) Schrödinger Wave Equation

Sparks CH301. Quantum Mechanics. Waves? Particles? What and where are the electrons!? UNIT 2 Day 3. LM 14, 15 & 16 + HW due Friday, 8:45 am

PY 351 Modern Physics - Lecture notes, 3

Chem 3502/4502 Physical Chemistry II (Quantum Mechanics) 3 Credits Spring Semester 2006 Christopher J. Cramer. Lecture 7, February 1, 2006

MATH325 - QUANTUM MECHANICS - SOLUTION SHEET 11

Quantum mechanics (QM) deals with systems on atomic scale level, whose behaviours cannot be described by classical mechanics.

Physics 280 Quantum Mechanics Lecture III

4/21/2010. Schrödinger Equation For Hydrogen Atom. Spherical Coordinates CHAPTER 8

Advanced Quantum Mechanics, Notes based on online course given by Leonard Susskind - Lecture 8

Deformed (Nilsson) shell model

I. RADIAL PROBABILITY DISTRIBUTIONS (RPD) FOR S-ORBITALS

Complementi di Fisica Lectures 10-11

Introduction to Electronic Structure Theory

ONE AND MANY ELECTRON ATOMS Chapter 15

L z L L. Think of it as also affecting the angle

Problem 1: Spin 1 2. particles (10 points)

Complementi di Fisica Lectures 5, 6

2. narrative 3. Last year s lecture titles (certain to be modified) 4. Gaussian and FT 5. PS #1 Due 9/13. Read CTDL, pages 9-39, 50-56, 60-85

Let us go back to what you knew in high school, or even earlier...

16.1. PROBLEM SET I 197

2.4. Quantum Mechanical description of hydrogen atom

Lecture 4 (19/10/2012)

5.111 Lecture Summary #6

C/CS/Phys C191 Quantum Mechanics in a Nutshell 10/06/07 Fall 2009 Lecture 12

ECE 487 Lecture 6 : Time-Dependent Quantum Mechanics I Class Outline:

Particle in a 3 Dimensional Box just extending our model from 1D to 3D

3. Quantum Mechanics in 3D

Chemistry 120A 2nd Midterm. 1. (36 pts) For this question, recall the energy levels of the Hydrogenic Hamiltonian (1-electron):

Physics 486 Discussion 5 Piecewise Potentials

Final Exam. Tuesday, May 8, Starting at 8:30 a.m., Hoyt Hall.

P3317 HW from Lecture and Recitation 10

A few principles of classical and quantum mechanics

Relativistic corrections of energy terms

PHY293 Lecture #18. h2 2 [c 1 Ψ 1 + c 2 Ψ 2 ] = 0. + c 2

4 Quantum Mechanical Description of NMR

Approximation Methods in QM

The Particle in a Box

UNIVERSITY OF SURREY FACULTY OF ENGINEERING AND PHYSICAL SCIENCES DEPARTMENT OF PHYSICS. BSc and MPhys Undergraduate Programmes in Physics LEVEL HE2

Quantum Mechanics: Vibration and Rotation of Molecules

Columbia University Department of Physics QUALIFYING EXAMINATION

Lecture 10: The Schrödinger Equation. Lecture 10, p 2

PH 253 Final Exam: Solution

UNIVERSITY OF MARYLAND Department of Physics College Park, Maryland. PHYSICS Ph.D. QUALIFYING EXAMINATION PART II

Harmonic Oscillator I

MITOCW ocw lec8

QUANTUM MECHANICS Intro to Basic Features

Angular Momentum - set 1

We do not derive F = ma; we conclude F = ma by induction from. a large series of observations. We use it as long as its predictions agree

Quantum Mechanics: Postulates

Quantum Orbits. Quantum Theory for the Computer Age Unit 9. Diving orbit. Caustic. for KE/PE =R=-3/8. for KE/PE =R=-3/8. p"...

C/CS/Phys 191 Quantum Mechanics in a Nutshell I 10/04/05 Fall 2005 Lecture 11

Lecture 41 (Hydrogen Atom and Spatial Quantization) Physics Fall 2018 Douglas Fields

Lecture 10: The Schrödinger Equation. Lecture 10, p 2

PHY4604 Introduction to Quantum Mechanics Fall 2004 Final Exam SOLUTIONS December 17, 2004, 7:30 a.m.- 9:30 a.m.

Transcription:

C/CS/Phys C191 Particle-in-a-box, Spin 10/0/08 Fall 008 Lecture 11 Last time we saw that the time dependent Schr. eqn. can be decomposed into two equations, one in time (t) and one in space (x): space : Ĥψ(x) = Eψ(x) (1) time : i h φ(t) t = Eφ(t) () This time equation is easy: φ(t) = e iet/ h φ(t) = e iωt, ω = Ē h, or E = hω. (Planck-Einstein relation.) The spatial equation is harder. It is called the time-independent Schrodinger equation. However difficult to solve, this equation is a special type of equation known as an eigenvalue problem, where operator ψ = constant ψ. Thus the Schr. eqn. reduces to an eigenvalue problem, which has a storied history in mathematics. To solve the Schr. eqn., one must find the set of wavefunctions {ψ k (x)} that return to themselves (times a constant) when acted on by Ĥ: Ĥψ k (x) = E k ψ k (x). Each solution ψ k has an energy E k associated with it. The states {ψ k } and only they have well-defined energies. The energies {E k } are the set of all physically allowed energies of the system. The full time-dependent solution is ψ(x,t) = ψ(x)φ(t) = ψ k (x)e ie kt/ h. Since these are linear equations we can add solutions, just as above for the free particle case. So, the most general solution to the Schrodinger equation is ψ(x,t) = k A k ψ k (x)e ie kt/ h. This tells us the rule for time evolution in quantum mechanics: if you want to know how a wave function evolves in time then you just break it down into energy eigenfunctions and multiply each one by a timedependent phase factor whose frequency is proportional to its energy. Example: Suppose at time t=0 we have ψ(x) = A 1 ψ 1 (x) + A ψ (x) + A 3 ψ 3 (x) where Ĥψ i (x) = E i ψ i (x). What is ψ(x,t) for t 0? To solve this, we can just add in the phase factors: ψ(x,t 0) = A 1 ψ 1 (x)e ie 1t/ h + A ψ (x)e ie t/ h + A 3 ψ 3 (x)e ie 3t/ h (3) Let s do another example: For a free particle Ĥ = h ψ x k (x) = e ikx,e k = h k. Therefore, ψ(x,t) = e ikx e i hk t as seen before. In general Ĥψ(x) = Eψ(x) can be a very difficult differential equation to solve. Physicists and chemists spend a great deal of time solving this equation! Note that we have already derived this time evolution earlier when analyzing the unitary evolution of a quantum state: ψ(x,t 0) = e iĥt/ h ψ(x,t = 0), with e iĥt/ h the time evolution operator. Recall how we approached it before: C/CS/Phys C191, Fall 008, Lecture 11 1

Any ψ(x,t = 0) can be written ψ(x,t = 0) = k A k ψ k (x), so: ( ) ψ(x,t 0) = e iĥt/ h A k ψ k (x) k = k A k e iĥt/ h ψ k (x) = A k ψ k (x)e ie kt/ h k (4) which is the same as above, since ψ k (x) is an eigenfunction of Ĥ with eigenvalue E k. The time-independent Schr. eqn. sets a condition that determines what the allowed energies of a system are and what the states look like that have well-defined energy ( well-defined means you d get the same answer if you measured the energy of all of the members of an identically prepared ensemble of systems in that state). However, we can also construct states that have well-defined values of other physical quantities, like momentum, position, and angular momentum. This is often called the basis of stationary states. Why? Because if ψ = ψ i (x) where Ĥψ i = E i ψ i then ψ(x,t) = ψ i (x)e ie it/ h. The probability density P(x,t) is then given by P(x,t) = ψ(x,t) = (ψ ) ( ) i (x)e ie it/ h ψ i (x)e ie it/ h = ψ i (x) Therefore the time dependence for the probability density dropped out and does not change in time. But what does well-defined energy mean? It can mean two things: (1) A state ψ has well-defined energy if Ĥψ = Cψ where C = energy of state. () A state ψ has well defined energy if an ensemble (read, many copies) of systems all prepared in the state ψ give the same answer if you measure energy (i.e. E = C if Ĥψ = Eψ). So let s discuss measurement. If ψ >= a 1 ψ E1 > +a ψ E > +a 3 ψ E3 > +, what is the result of a measurement of energy? One of the postulate of QM is that the result of the measurement must be an eigenvalue of Ĥ. ψ will collapse onto one of these eigenstates with some probability. What s the probability of obtaining E 3? P 3 = < ψ e ψ > = a 3 And what is ψ after measurement? ψ is projected to ψ 3 upon an observation of E 3. So, measurement is a random collapse onto one of the eig. states of the observable you are measuring! The same holds for momentum: If we are discussing momentum then it s best to work with momentum eigenstates. ˆpψ p = pψ p {ψ p },{p} Suppose ψ >= b 1 ψ p1 > +b ψ p > +b 3 ψ p3 > + What is a result of a measurement of momentum? We will end up measuring an eigenvalue of momentum with some probability, and then collapse onto that eigenstate (P = b ). The exact same thing happens for the observables ˆx, ˆL, etc. The eigenstates of these observables define bases, and measurement of that observable randomly collapses us onto one of those eigenstates. Question: What if we take an ensemble of identically prepared states and measure the same physical quantity for each? How do we determine (theoretically) the average value of the measurements? This will lead us to the definition of an expectation value. Example: ENERGY. Suppose we know states {ψ E }, {E}. If an ensemble is prepared in ψ >= ψ E > then the situation is simple: < E >= E 0. But what if we prepare an ensemble in a state ψ > in a superposition C/CS/Phys C191, Fall 008, Lecture 11

state which is not an eigenstate of Ĥ, e.g. ψ >= a 1 ψ E1 > +a ψ E > +a 3 ψ E3 > +? What is < E > then? < E >= E 1 Prob[E 1 ]+E Prob[E 1 ]+E 3 Prob[E 3 ]+ where Prob[E i ] is just This yields: Prob[E i ] = < ψ Ei ψ > = a i < E >= a 1 E 1 + a E + a 3 E 3 + Our shorthand for this is given by: < E >=< ψ Ĥ ψ > which is known as the expectation value of the Hamiltonian (or equivalently of the energy). Use completeness to show that < E >=< ψ Ĥ ψ > does indeed give you the same expression as above. We can do this for any observable! Consider arbitrary observable Â. The average value of this quantity for ensemble of systems prepared in ψ > is < A >=< ψ Â ψ >. It should be noted that it is sometimes hard to evaluate the expectation value. Take the continuous basis for example ( x >). Suppose ψ(x) =< x ψ >= Ae x. What is the average value of measured momentum for an ensemble of systems? ( < ˆp >=< ψ ˆp ψ >= ψ (x) ˆpψ(x)dx = A e x)( ) h ( Ae x) dx = 0 i x So, in this instance the expectation value is zero. It is left as an exercise to evaluate < p > (symmetry analysis will immediately tell you if it zero or not). 1 Particle-in-a-box model for atomic qubit Let s do an example now! Let s consider a situation where we want to use the electrons inside atoms as qubits. How do we describe the physical details of these qubits? What are their allowed energies? How do they change in time? What do we do??? We solve the Schr. equation, that s what. As is the case in most QM problems, we must find the Hamiltonian Ĥ. Ĥ in this case is the energy operator for an electron in an atom. To know this then we must make some assumptions about how electrons behave in an atom. Let s assume that atoms are very tiny ( 10 10 meter) 1-D boxes with very hard walls. The walls are located at position x = 0 and x = l. This model works surprisingly well. Inside the box Ĥ is given by the free particle C/CS/Phys C191, Fall 008, Lecture 11 3

Hamiltonian Ĥ = h. Outside the box we model the very hard walls as regions where the potential x energy V. This has the effect of disallowing any ψ to be nonzero outside the box. If it did exist in this region its energy (obtained, as always, by applying the Hamiltonian) would also go to infinity. That s too much energy for our little electrons, so we can say that we will restrict our wavefunctions ψ(x) to functions which vanish at x 0 and x l. ψ(x = 0) = ψ(x = l) = 0 Strictly speaking, we mean that ψ(x 0) = ψ(x l) = 0. We will see that this will allow us to construct wavefunctions which are normalized over our restricted box space x {0, l}. The system as we ve described it can be sketched is sketched in Figure. V(x) = infinity for x < 0 V(x) = infinity for x > L particle never here wavefunction = 0 for x < 0 particle lives in here wavefunction = 0 particle never here wavefunction = 0 for x < 0 x = 0 V(x) = 0 x x = L Figure 1: Particle in a box The first thing to note is that we ve done this problem before! For a free particle we know that we have solutions ψ E (x) = Ae ikx + Be ikx with energies E k = h k. Are we done? No, because we need to impose our boundary condition that ψ(x = 0) = ψ(x = l) = 0 since those walls are hard and do not allow particles to exist outside of the free particle box we ve constructed. Our previous solution ψ E (x) = Ae ikx + Be ikx is fine, but we can also write another general solution as follows: ψ E (x) = C sin(kx)+dcos(kx) As we will see, this is a convenient choice. If we now impose our first boundary conditions: ψ E (x = 0) = 0 = C sin[k(x = 0)]+Dcos[k(x = 0)] = C(0)+D(1) = D So D = 0 and we can forget about the cosine solution. The second boundary condition tells us: ψ E (x = l) = 0 = C sin(kl) = 0 This is satisfied for all kl = nπ, where n is an integer. Therefore, we have k n = nπ l quantized eigenfunction set. The energy eigenvalues are which gives us our C/CS/Phys C191, Fall 008, Lecture 11 4

E n = h kn = h n π with eigenfunctions ( nπ ψ n (x) = Csin l ) x Are we done? No, because we must normalize. l l < ψ n ψ n >= ψ n (x) dx = 1 C sin ( nπ ) x dx = 1 C = 0 0 l l So normalization has given us our proper set of energy eigenfunctions and eigenvalues: ψ n (x) = l sin ( nπ l x ),E n = h n π l Higher energy states have more nodes. Some of the wavefunctions can be sketched as follows: V(x) = infinity for x < 0 V(x) = infinity for x > L n=3 n= n=1 x = 0 x x = L Figure : The first three eigenfunctions of the particle in a box system. What does this have to do with the discrete quantum state picture as described in the context of qubits? To obtain a qubit from this system, we can construct our standard basis 0 > and 1 > by just restricting our state space to the bottom two eigenstates: C/CS/Phys C191, Fall 008, Lecture 11 5

0 >= ψ 1 (x) = ( π ) l sin l x,e n=0 = h π l 1 >= ψ (x) = ( π l sin x ),E n=1 = 4 h π l l Physically this would mean forcing the total energy of the system to be less than E, meaning that the particle could never have any overlap with ψ n for n. What about the energies of qubit states? Suppose I take 10 6 qubits prepared in state 0 = ψ1 and measure their energy and make a histogram. What does the histogram look like? See Figure 1(a). Now suppose that I prepare 10 6 qubits in the superposition state ψ 3 = 5 0 + 5 1, measure their energies, and make a histogram. How does it look? See Figure 1(b) counts counts 10 6 (a) (b) 600,000 400,000 E 1 energy E 1 E energy Figure 3: Histograms of particle energy measurements. Ask yourself, is ψ a state with well-defined energy? NO. Why not? ψ is not an eigenstate of the Hamiltonian operator. Let s check this: Ĥψ = Ĥ ( ) 3 3 5 ψ 1 + 5 ψ = 5 E 1ψ 1 + 5 E ψ Does this equal (constant) (ψ )? No, because as stated E 1 and E are not equal. Therefore ψ is not an eigenstate of the energy operator and has no well-defined energy. However it is a perfectly valid superposition state of the qubit. This extremely simple model of a confined particle is actually very useful physically, and pops up in many real-world applications. In fact, the particle-in-a-box model provides one of the simplest meaningful descriptions of an atom. In an atom the confined particle is an electron and the box is created by the Coulomb C/CS/Phys C191, Fall 008, Lecture 11 6

attraction between the negatively charged electron and the positively charged nucleus. Most atoms have a lot of electrons, but all atoms behave (at some level) in a manner that is very similar to the simplest atom which is hydrogen. Hydrogen has just one electron that circles around one proton. Solving the exact Schroedinger equation for the motion of an electron around a proton involves some complexities that we dont want to worry about right now, but suffice to say that the energy is determined most strongly by the radial motion of electron. (i.e. the electrons radial distance from the proton). The radial Schroedinger Equation of the electron then looks very much like the simple particle-in-a-box model that we have just solved. The ground state and quantized excited states of hydrogen look like standing waves in much the same way as the particle-in-a-box wave-functions. Higher energy states of hydrogen have more nodes just like the particle-in-a-box states. Spin Qubit systems can always be mapped onto an effective spin 1/ system so it is important to understand what this is and where it comes from. Elementary particles and composite particles carry an intrinsic angular momentum called spin. For our purposes, the most important particles are electrons and protons. To each of these is associated an angular momentum vector that can point up > or down >. The quantum mechanical spin state of an electron or proton is thus ψ >= α > +β >. Therefore, spins can be used as qubits with 0 >= >, 1 >= >. The spin angular momentum is intrinsic and signals the presence of an intrinsic magnetic moment. Uhlenbeck and Goudsmit introduced the concept of spin in 195 to explain the behavior of hydrogen atoms in a magnetic field: The extra transitions can be explained if an electron has an intrinsic magnetic moment µ, since a magnetic moment in a magnetic field B has an energy E = µ B. In the context of QM, new energy levels can derive from µ being oriented parallel or anti-parallel to B. Where does µ come from? The simplest explanation is classical : classically, a magnetic moment µ comes from a loop of current. The energy E = µ B comes from I B force of current in a B-field (Lorentz force). The lowest energy, and thereby the place where the system wants to go, is obtained when the magnetic moment and B-field C/CS/Phys C191, Fall 008, Lecture 11 7

line up. If an isolated electron has intrinsic µ then the simplest explanation for this is that electron spins about some axis. This is independent of its orbital motion in an atom, just like the Earth s spin about the north pole is independent of its orbit around the sun. Since µ is associated with a spinning charge, then we can write µ in terms of angular momentum. Anything that spins has angular momentum! The simplest way to see this is classically for a spinning charge. For an electron the charge q is equal to e. Angular momentum is given by L = r p = r m v. L = mvr for a charge of mass = m moving in a a circle with velocity = v. The magnetic moment can be obtained as follows: µ = (current)(area) = q τ πr But the revolution period τ = πr v. Substituting for τ and v in terms of L, we obtain µ = q L Now comes the tricky part. The electron is not actually spinning about some axis! It only acts as though it is. Electrons are point particles which, as far as we know, have no size in the traditional sense. Therefore the r in the previous discussion of spinning charge is not meaningful. The intrinsic angular momentum of an electron has nothing to do with orbital motion, but it does lead to an intrinsic µ. This is a relativistic effect that can be derived from the Dirac Equation (Relativistic Schrodinger equation for spin- 1 particles), but it holds for electrons that are not moving fast. This intrinsic angular momentum is called spin = S. For an electron, classically: µ = e L, while quantum mechanically: µ = ge S. What is g? g is called the g-factor and it is a unitless correction factor due to QM. For electrons, g. For protons, g 5.6. You should also note that m proton m electron 000, so we conclude that µ proton µ electron. So, to understand behavior of the electron s intrinsic magnetic moment µ (which is an observable we can measure) then we must understand the behavior of its intrinsic angular momentum = S. This is why spin is important. Since the electron is small, S must be described by QM. See also notes electrons in atoms (look for spin section) on the web page Science Trek at http://www.colorado.edu/phy 1.1 The Bloch Sphere A very nice way to think of the quantum states of qubits is via the Bloch Sphere. This is a convenient mapping for all possible single-qubit states. See Figure 1 below. C/CS/Phys C191, Fall 008, Lecture 11 8

Figure 4: The Bloch sphere. 0 is at the North pole, 1 at the South pole. θ and φ are the usual spherical coordinates, with 0 θ π,0 < φ π. Every point on the sphere represents a possible qubit. All possible qubits (within an overall multiplicative phase factor) can be thought of as vectors on this unit sphere. A vector on the Bloch Sphere represents this qubit: ψ >= cos θ 0 > +eiφ sin θ 1 > C/CS/Phys C191, Fall 008, Lecture 11 9