J = L + S. to this ket and normalize it. In this way we get expressions for all the kets

Similar documents
Lecture 29 Relevant sections in text: 3.9

Singlet State Correlations

The nature of Reality: Einstein-Podolsky-Rosen Argument in QM

The Einstein-Podolsky-Rosen thought-experiment and Bell s theorem

Introduction to Bell s theorem: the theory that solidified quantum mechanics

Bell s Theorem. Ben Dribus. June 8, Louisiana State University

The Einstein-Podolsky-Rosen thought experiment and Bell s theorem

Quantum Entanglement. Chapter Introduction. 8.2 Entangled Two-Particle States

(Again, this quantity is the correlation function of the two spins.) With z chosen along ˆn 1, this quantity is easily computed (exercise):

EPR Paradox and Bell s Inequality

Modern Physics notes Spring 2007 Paul Fendley Lecture 27

1.1.1 Bell Inequality - Spin correlation

2 Quantum Mechanics. 2.1 The Strange Lives of Electrons

If quantum mechanics hasn t profoundly shocked you, you haven t understood it.

Lecture 4. QUANTUM MECHANICS FOR MULTIPLE QUBIT SYSTEMS

EPR Paradox and Bell Inequalities

Einstein-Podolsky-Rosen paradox and Bell s inequalities

EPR Paradox Solved by Special Theory of Relativity

Total Angular Momentum for Hydrogen

Entanglement and information

Bell s inequalities and their uses

Hardy s Paradox. Chapter Introduction

Introduction to Quantum Mechanics

Physics 221A Fall 1996 Notes 14 Coupling of Angular Momenta

Q8 Lecture. State of Quantum Mechanics EPR Paradox Bell s Thm. Physics 201: Lecture 1, Pg 1

Homework 2. Solutions T =

Quantum measurement theory

B. BASIC CONCEPTS FROM QUANTUM THEORY 93

Chem 3502/4502 Physical Chemistry II (Quantum Mechanics) 3 Credits Spring Semester 2006 Christopher J. Cramer. Lecture 20, March 8, 2006

Today s Outline - April 18, C. Segre (IIT) PHYS Spring 2017 April 18, / 23

Photons uncertainty removes Einstein-Podolsky-Rosen paradox. Abstract

Unitary evolution: this axiom governs how the state of the quantum system evolves in time.

Quantum Information Types

Incompatibility Paradoxes

G : Quantum Mechanics II

Honors 225 Physics Study Guide/Chapter Summaries for Final Exam; Roots Chapters 15-18

INTRODUCTORY NOTES ON QUANTUM COMPUTATION

Quantum Mechanics- I Prof. Dr. S. Lakshmi Bala Department of Physics Indian Institute of Technology, Madras

6.896 Quantum Complexity Theory September 9, Lecture 2

Violation of Bell Inequalities

Addition of Angular Momenta

A review on quantum teleportation based on: Teleporting an unknown quantum state via dual classical and Einstein- Podolsky-Rosen channels

Closing the Debates on Quantum Locality and Reality: EPR Theorem, Bell's Theorem, and Quantum Information from the Brown-Twiss Vantage

Wave Mechanics Relevant sections in text: , 2.1

Causation and EPR. But, alas, this hypothesis is mathematically inconsistent with the results of the case b runs, those runs

Bell s Theorem 1964 Local realism is in conflict with quantum mechanics

The controlled-not (CNOT) gate exors the first qubit into the second qubit ( a,b. a,a + b mod 2 ). Thus it permutes the four basis states as follows:

226 My God, He Plays Dice! Entanglement. Chapter This chapter on the web informationphilosopher.com/problems/entanglement

Quantum Mechanics- I Prof. Dr. S. Lakshmi Bala Department of Physics Indian Institute of Technology, Madras

Treatment of Error in Experimental Measurements

Quantum mechanics in one hour

Quantum information and quantum computing

Lecture 19 (Nov. 15, 2017)

Angular momentum operator algebra

Timeline: Bohm (1951) EPR (1935) CHSH (1969) Bell (1964) Theory. Freedman Clauser (1972) Aspect (1982) Weihs (1998) Weinland (2001) Zeilinger (2010)

The Framework of Quantum Mechanics

Chapter 2: Quantum Entanglement

Lecture 2: Introduction to Quantum Mechanics

A proof of Bell s inequality in quantum mechanics using causal interactions

Are these states normalized? A) Yes

ON THE EINSTEIN PODOLSKY ROSEN PARADOX* I. Introduction

The Wave Function. Chapter The Harmonic Wave Function

Cosmology Lecture 2 Mr. Kiledjian

Basic Physical Chemistry Lecture 2. Keisuke Goda Summer Semester 2015

Dependent (Contextual) Events

Quantum Cryptography

Measuring Quantum Teleportation. Team 10: Pranav Rao, Minhui Zhu, Marcus Rosales, Marc Robbins, Shawn Rosofsky

Quantum Openness and the Sovereignty of God. by Don Petcher Department of Physics Covenant College

2 The Density Operator

Ph 219/CS 219. Exercises Due: Friday 3 November 2006

Eigenvectors and Hermitian Operators

Quantum Entanglement, Beyond Quantum Mechanics, and Why Quantum Mechanics

Quantum Mechanics Solutions. λ i λ j v j v j v i v i.

Another Assault on Local Realism

Classical Bell s Inequalities. Vesselin C. Noninski

Quantum Measurement and Bell s Theorem

( ) dσ 1 dσ 2 + α * 2

The quantum state as a vector

arxiv:quant-ph/ v1 14 Sep 1999

3/10/11. Which interpreta/on sounds most reasonable to you? PH300 Modern Physics SP11

Basic Quantum Mechanics Prof. Ajoy Ghatak Department of Physics Indian Institute of Technology, Delhi

Pure Quantum States Are Fundamental, Mixtures (Composite States) Are Mathematical Constructions: An Argument Using Algorithmic Information Theory

Statistics. Lent Term 2015 Prof. Mark Thomson. 2: The Gaussian Limit

NON HILBERTIAN QUANTUM MECHANICS ON THE FINITE GALOIS FIELD

This is the important completeness relation,

Quantum Physics & Reality

Topic 3: Bohr, Einstein, and the EPR experiment

Response to Wiseman, Rieffel, and Cavalcanti on Bell s 1964 Paper

The Wave Function. Chapter The Harmonic Wave Function

Introduction. Introductory Remarks

conventions and notation

A Re-Evaluation of Schrodinger s Cat Paradox

Quantum Physics II (8.05) Fall 2004 Assignment 3

Week 11: April 9, The Enigma of Measurement: Detecting the Quantum World

Lecture 6: Quantum error correction and quantum capacity

Problems with/failures of QM

Lecture 11 Spin, orbital, and total angular momentum Mechanics. 1 Very brief background. 2 General properties of angular momentum operators

A No-Go Result on Common Cause Approaches via Hardy s Paradox

Second quantization: where quantization and particles come from?

Collapse versus correlations, EPR, Bell Inequalities, Cloning

Transcription:

Lecture 3 Relevant sections in text: 3.7, 3.9 Total Angular Momentum Eigenvectors How are the total angular momentum eigenvectors related to the original product eigenvectors (eigenvectors of L z and S z )? We will sketch the construction and give the results. Keep in mind that the general theory of angular momentum tells us that, for a given l, there are only two values for j, namely, j = l ± 1. Begin with the eigenvectors of total angular momentum with the maximum value of angular momentum: j 1 = l, j = 1, j = l + 1, m j = l + 1. Given j 1 and j, there is only one linearly independent eigenvector which has the appropriate eigenvalue for J z, namely so we have j 1 = l, j = 1, m 1 = l, m = 1, j 1 = l, j = 1, j = l + 1, m = l + 1 = j 1 = l, j = 1, m 1 = l, m = 1. To get the eigenvectors with lower eigenvalues of J z we can apply the ladder operator J = L + S to this ket and normalize it. In this way we get expressions for all the kets j 1 = l, j = 1/, j = l + 1, m j, m j = l 1, l + 1,..., l 1, l + 1. The details can be found in the text. Next, we consider j = l 1 for which the maximum m j value is m j = l 1. This eigenvalue is doubly degenerate: there is a -d space with this value for m j. A basis for this space is provided by the product vectors: j 1 = l, j = 1, m 1 = l, m = 1, j 1 = l, j = 1, m 1 = l 1, m = 1 Of course, two of the total angular momentum eigenvectors (which we are trying to compute) will also provide a basis for this space. One of the basis vectors is already known; it has j = l + 1, m j = l 1 and its computation was explained in the previous paragraph. The eigenvector we are after (with j = l 1/, m j = l 1/ ) can now be computed by 1

demanding it be a linear combination of the above product vectors which is orthogonal to the already known eigenvector with j = l + 1, m j = l 1. With the vector characterized by j = l 1, m j = l 1 in hand one can use the ladder operator to get all the other m j values, that is we will obtain j 1 = l, j = 1/, j = l 1, m j, m j = l + 1, l + 3,..., l 3, l 1. In this way we can express all the total angular momentum eigenvectors in terms of the product eigenenvectors. The result of all these considerations is not too complicated. We have, for j = l ± 1, j 1 = l, j = 1, j = l ± 1, m j = 1 { ± l ± m j + 1 l + 1 l, m j 1 + + l m j + 1 l, m j + 1 }. In terms of spinor fields we define the total angular momentum spinors representing the total angular momentum eigenvectors, by taking components in the product basis: Ψ j,m = 1 l + 1 ± l ± m j + 1 a +(r)y l,ml =m j 1 (θ, φ) l m j + 1 a, (r)y l,ml =m j + 1 (θ, φ) The functions a ± (r) are not determined by the angular momentum information. The upper (lower) component of this column vector determines the probability density for finding a particle at the position specified by (r, θ, φ) and with a spin up (down) along z, given that its total angular momentum is known with certainty to have the values specified by j = l ± 1 and m j. Spin correlations and quantum weirdness: Spin 1/ systems The principal difference between quantum mechanics and classical mechanics is the incompatibility of certain observables. This incompatibility can lead to some dramatic results. Recall the results of adding two spin 1/ angular momenta. There the total spin observables are not all compatible with the individual spin observables. This leads to some pretty weird stuff from a classical physics point of view. A key issue was raised by Einstein-Podolsky and Rosen (EPR) in a famous critique of the completeness of quantum mechanics. Much later, Bell showed that the basic classical locality assumption of EPR must be violated, implying in effect that nature is truly as weird as quantum mechanics makes it out to be. Here we give a brief discussion of some of these ideas.

Consider a pair of spin 1/ particles created in a spin singlet state. (Experimentally speaking, this can be done in a variety of ways; see your text.) Thus the spin state of the system is defined by the state vector ψ = 1 ( + + ). Let us suppose that the particles propagate undisturbed and not interacting until they are well-separated. We have two observers. Observer 1 is equipped to measure the spin properties of Particle 1 and Observer is equipped to measure the spin properties of particle. To begin, suppose both observers measure spin component along the z axis. If observer 1 sees spin up, what does observer see? You probably will correctly guess: spin down. But can you prove it? Well, the reason for this result is that the state of the system is an eigenvector of S z with eigenvalue zero. So, the two particles are known with certainty to have opposite values for their z-components of spin. Alternatively, you can see from the expansion of ψ in the product basis that the only states that occur (with equal probability) are states with opposite spins. Let us see how to prove this systematically; it s a good exercise. We can ask the question as the following sequence of simple questions. What is the probability P (S 1z = h ) that observer 1 gets spin up? That s easy:* P (S 1z = h ) = + + ψ + + ψ = 1. Of course, there is nothing special about particle 1 compared to particle ; the same result applies to particle. What s the probability for getting particle 1 with spin up and particle with spin down? We have P (S 1z = h, S z = h ) = + ψ = 1. You can now easily infer that in the singlet state when particle 1 is determined to have spin up along z, then particle will have spin down along z with certainty. We say that * Well, it s easy if you realize that, in a state ψ, the probability for getting an eigenvalue a of an operator A is (exercise) P (a) = d i ψ, where i are a basis for the d-dimensional subspace of vectors with eigenvalue a: i=1 A i = a i, i = 1,,... d. 3

the S z variables for particles 1 and are completely correlated. Another way to state this is the following. If particle 1 is determined to have spin up along z, then we know two things with certainty - the z component of total spin is zero, and particle one has spin up. (Note that [S z, S 1z ] = 0.) The state vector of the system is given by +. A subsequent measurement of the z component of spin for particle is then going to give spin-down with certainty. The foregoing discussion will work for any choice of axis, that is, it holds for any component of the spin (the same component being measured by each observer). We imagine repeating this experiment many times. Each observer measures the single particle spin component and sees a random sequence of spin up or spin down results with a 50-50 probability. Quantum mechanics says that this is the best you can do you cannot predict with certainty the S z value for each particle because the individual particle spin is not compatible with the magnitude of the total spin of the system, which is what is known with certainty. The (singlet) state vector has been specified and there is no more information to be had about the system. On the other hand, if the two observers compare their data, they will see a perfect correlation between their data as described above. How do the particles know that they have to have a perfect correlation in their spin-z components, given that they can be separated by arbitrarily large distances? EPR say: each of the particles does have a definite value of the z component of spin when they are created, and these values are correlated as described above. It s just that quantum mechanics is unable to predict the spin values for each particles in the first place and assigns them a 50-50 probability distribution. Thus EPR ascribe a certain reality to the spin component properties of each particle. EPR go on to argue that if quantum mechanics is the ultimate description (no further information available) then a paradox arises. To see this, reason as follows. If both observers make measurements, then observers 1 and each see a random distribution of spin up and spin down results, with a 50-50 probability. But when they compare their data they see that each observer 1 spin up is matched with an observer spin down and vice versa. Thus, in this particular situation, one could say that particles 1 and have in each experimental run a definite value of S z. (Of course, the choice of the z axis is completely arbitrary.) Now, suppose that observer 1 chooses to measure S x instead of S z. When observer 1 gets spin up/down along x, then the state of the system is (exercise) S 1x = ± h/; S x = h/ and observer s S z measurement will now yield spin up and spin down with a 50-50 probability. To see this, the probability P (S z = h/) for getting S z = h/ in either state S 1x = ± h/; S x, h/ can be computed by P (S z = h/) = S 1x = h/; S z = h/ S 1x = ± h/; S x, h/ + S 1x = h/; S z = h/ S 1x = ± h/; S x, h/ = 1. 4

Evidently, in this case, even knowing that, say, particle one has spin up along x, particle does not have a definite value for S z it has a 50-50 probability distribution. On the other hand, if observer chose to measure S x, with certainty the result will be opposite to the outcomes of observer 1 s measurement. In other words, when observer 1 measures S x, particle can be viewed as having a definite value for S x, but not of S z. Thus it appears that the reality of particle seems to depend upon the causally disconnected choices of Observer 1. A paradox. Put differently we can say: (i) A measurement of one spin component yields random (50-50) values. (ii) When observer 1 measures S z observer s S z value is perfectly correlated with observer 1 s measurement. (iii) When observer 1 measures S x, observer s S z values are random uncorrelated with observer 1 s results. Somehow particle has to know whether to correlate or not with particle 1 instantaneously! To make this last paragraph more precise we can define the correlation function of two observables A and B in a given state to be the normalized covariance function: C(A, B) = AB A B A B. The correlation function takes values in [0, 1]. An observable is always perfectly correlated with itself, C(A, A) = 1, while two statistically independent observables are uncorrelated: C(A, B) = 0. Using the singlet state we see that for any a C(S 1a, S a ) = 1, indicating perfect correlation. On the other hand, Indicating uncorrelated observables. C(S z, S x ) = 0, From this discussion it is clear that according to the description provided by quantum mechanics one cannot assert that the spin observables were a part of the reality of each particle independent of the measurements. That s just how nature works (according to quantum mechanics). This is the resolution of the paradox, but we must accept the fact that quantum mechanics predicts a striking non-locality. EPR argue that this non-locality is unacceptable in a complete description of the system and the only way out is to suppose that all the spin components of each particle are in fact well-defined, determined with certainty by the state preparation process, and quantum mechanics is just an incomplete theory with a concomitant bit of non-locality in its predictions. Remarkably, it is possible to devise experiments (based upon Bell inequalities see your text) which can test whether (1) observables do not have an a priori reality and nature is in 5

fact non-local in the sense provided by quantum mechanics or () nature is fundamentally local, observables have an a priori reality and hence a more complete (non-probabilistic) theory than quantum mechanics is possible. So far all experiments have confirmed the quantum mechanical description nature really is as weird as quantum mechanics says it is! 6