Spin Half
Spin Half
Peter Woit Department of Mathematics, Columbia University [email protected] October 15, 2012
The existence of a non-trivial double-cover Spin(3) of the three-dimensional rotation group may seem to be a somewhat obscure mathematical fact but, remarkably, it is Spin(3) rather than SO(3) that is the symmetry group of quantum systems describing elementary particles. Studies of atomic spectra in the early days of quantum mechanics revealed twice as many states as expected, a phenomenon that is now understood to be due to the fact that the state space of an electron at a point is C2 rather than C . This state space is not a representation of the rotational symmetry group SO(3), but it is a representation of the double-cover Spin(3) = SU (2) (the standard representation of SU (2) matrices on C2 ). Particles with this degree of freedom are said to have spin 1/2, with the origin of this 1/2 precisely the same as the 1/2 we saw in the previous class in the discussion of the Lie algebras of the two equivalent forms Sp(1) and SU (2) of the group Spin(3). This same C2 occurs for other matter particles (quarks, neutrinos, etc.) and appears to be a fundamental property of nature. Besides the doubling of the number of states, more complicated physical eects occur when such a particle is subjected to a magnetic eld, and we will examine some of this physics in this section.
In the last section we examined in great detail various ways of looking at the three-dimensional irreducible real representation of the groups SO(3), SU (2) and Sp(1). For SU (2) and Sp(1) however, there is a even simpler non-trivial irreducible representation: the representation of 2 by 2 complex matrices in SU (2) on column vectors C2 by matrix multiplication or the representation of unit quaternions in Sp(1) on H by scalar multiplication. Choosing an identication C2 = H these are isomorphic representations of isomorphic groups, and for convenience we will generally stick to using SU (2) and complex numbers rather than quaternions. This irreducible representation is known as the spinor or
spin representation of Spin(3) The homomorphism spinor dening the representation is just the identity map from SU (2) to itself. The spin representation of Spin(3) is not a representation of SO(3). The double cover map : Spin(3) SO(3) is a homomorphism, so given a representation (, V ) of SO(3) one gets a representation ( , V ) of Spin(3) by composition. But there is no homomorphism SO(3) SU (2) that would allow us to make the standard representation of SU (2) on C2 into an SO(3) representation. What is true is that we can try and dene a representation of SO(3) by : g SO(3) (g ) = spinor ( g ) SU (2) where g is some choice of one of the elements g SU (2) satisfying ( g) = g The problem with this is that we wont quite get a homomorphism. Changing our choice of g will introduce a minus sign, so will only be a homomorphism up to sign (g1 ) (g2 ) = (g1 g2 ) The nontrivial nature of the double-covering ensures that there is no way to completely eliminate all minus signs, no matter how we choose g . Something like this which is not quite a representation, only one up to a sign ambiguity, is known as a projective representation. So, the spinor representation of SU (2) = Spin(3) is only a projective representation of SO(3), not a true representation of SO(3). Quantum mechanics texts often deal with this phenomenon by noting that physically there is an ambiguity in how one species the space of states H, with multiplication by an overall scalar not changing the eigenvalues of operators or the relative probabilities of observing these eigenvalues. As a result, the sign ambiguity has no physical eect. It seems more straightforward though to just work from the beginning with the larger symmetry group Spin(3), accepting that this is the correct symmetry group reecting the action of rotations on three-dimensional quantum systems.
Recall from our earlier discussion of the two-state quantum system, where the state space is H = C2 that there is a four (real)-dimensional space of observables with a general self-adjoint linear operator on H written as M = c0 1 + c1 1 + c2 2 + c3 3 Multiplying by i, one gets skew-Hermitian operators, and thus elements of the Lie algebra u(2). Exponentiating gives elements of the unitary group U (2), with eic0 1 U (1) U (2) and ei(c1 1 +c2 2 +c3 3 ) SU (2) U (2)
For an arbitrary two-state quantum system, neither the operators M nor the group SU (2) have any particular geometric signicance. In some cases though, the group SU (2) does have a geometric interpretation, reecting its role as the double-cover Spin(3) and the fact that the group SO(3) acts on the physical system by rotations of three-dimensional space. In these cases, the quantum system is said to carry spin, in particular spin one-half (we will later on encounter state spaces of higher spin values). We will from now on assume that we are dealing with a spin one-half state space. A standard basis for the observables (besides the unit operator that generates overall U (1) transformations) is taken to be the operators sa , a = 1, 2, 3, where sa = i a 2
with the sa satisfying the commutation relations [s1 , s2 ] = s3 , [s2 , s3 ] = s1 , [s3 , s1 ] = s2 To make contact with the physics formalism, well dene self-adjoint operators a Sa = isa = 2
1 which will have real eigenvalues 2 . Note that the conventional denition of these operators in physics texts includes a factor of a Sa = i sa = 2 This is because rotations of vectors are dened in physics texts using conjugation by the matrix
R(, w) = ei
wS
with the convention of dividing by a factor of appearing here for reasons that have to do with the action of rotations on functions on R3 that we will encounter later on. For now, one can either keep factors of out of the denitions of Sa and the action of rotations, or just assume that we are working in units where = 1. States in H = C2 that have a well-dened value of the observable Sa will be the eigenvectors of Sa , with value for the observable the corresponding eigenvalue, which will be 1 2 . Measurement theory postulates that if we perform the measurement corresponding to Sa on an arbitrary state | , then we will with probability c+ get a value of + 1 2 and leave the state in an 1 eigenvector |a, + 2 of Sa with eigenvalue + 1 2 we will with probability c get a value of 1 2 and leave the state in an 1 eigenvector |a, 1 of S with eigenvalue a 2 2
1 1 + |a, 2 2
After such a measurement, any attempt to measure another orthogonal component of S , say Sb , b = a will give 1 2 with equal probability and put the system in a corresponding eigenvector of Sb . If a quantum system is in an arbitrary state | it may not have a welldened value for some observable A, but one can calculate the expected value of A. This is the sum over a basis of H consisting of eigenvectors (which will all be orthogonal) of the corresponding eigenvalues, weighted by the probability of their occurrence. The calculation of this sum using expansion in eigenvectors of Sa gives
1 1 1 ( a, + 1 |A| 2 | + a, 2 |)A(|a, + 2 + |a, 2 ) = 1 1 1 | ( a, + 1 2 | + a, 2 |)(|a, + 2 + |a, 2 )
1 2 ||2 (+ 1 2 ) + | | ( 2 ) 2 2 || + | | 1 1 =c+ (+ ) + c ( ) 2 2
One often chooses to simplify such calculations by normalizing states so that the denominator | is 1. Note that the same calculation works in general, as long as one has orthogonality and completeness of eigenvectors. Recall that R(, w) = ews = cos( )1 i(w ) sin( ) 2 2 In the case of a spin one-half particle, the group Spin(3) = SU (2) acts on states by the spinor representation with the element R(, w) SU (2) acting as | R(, w)| Taking adjoints and using unitarity, one has (thinking of vectors as column vectors, elements of the dual space as row vectors) the following action on the dual state space | |R(, w)1 The operators Sa transform under this same group according to the vector representation of SU (2), recall that an SO(3) rotation on vectors is given by conjugating by an SU (2) = Spin(3) group element according to Sa R(, w)Sa R(, w)1 4
We see that if our observables transform as vectors, and states as spinors, the expectation values remain invariant: |Sa | |R(, w)1 R(, w)Sa R(, w)1 R(, w)| = |Sa | and all eigenvalues of observables remain invariant. One can also interpret these joint transformations on states and observables as simply a change of coordinates, a rotation from the standard basis to a dierent one. Next semester in this course we may get to the physics electromagnetic elds and how particles interact with them in quantum mechanics, but for now all we need to know is that for a spin one-half particle, the spin degree of freedom that we are describing by H = C2 has a dynamics described by the Hamiltonian H = B Here B is the vector describing the magnetic eld, and =g (e) S 2mc
is an operator called the magnetic moment operator. The constants that appear are: e the electric charge, c the speed of light, m the mass of the particle, and g , a dimensionless number called the gyromagnetic ratio, which is approximately 2 for an electron, about 5.6 for a proton. The Schr odinger equation is d i | (t) = ( )( B)| (t) dt with solution | (t) = U (t)| (0) where U (t) = e
it ge|B| B 2mc s |B|
ge
ge
ge
The time evolution of a state is thus given at time t by a rotation about the B axis w = |B | by an angle ge|B|t 2mc
|B| a rotation taking place with angular velocity ge 2mc . The amount of non-trivial physics that is described by this simple system is impressive, including:
The Zeeman eect: this is the splitting of atomic energy levels that occurs when an atom is put in a constant magnetic eld. With respect to the energy levels for no magnetic eld, where both states in H = C2 have the same energy, the term in the Hamiltonian given above adds 5 ge|B| 4mc
to the two energy levels, giving a splitting between them proportional to the size of the magnetic eld. The Stern-Gerlach experiment: here one passes a beam of spin one-half quantum systems through an inhomogeneous magnetic eld. One can arrange this in such a way as to pick out a specic direction w, and split 1 the beam into two components, of eigenvalue + 2 and 1 2 for the operator w S. Nuclear magnetic resonance spectroscopy: one can subject a spin onehalf system to a time-varying magnetic eld B(t), which will be described by the same Schr odinger equation, although now the solution cannot be found just by exponentiating a matrix. Nuclei of atoms provide spin onehalf systems that can be probed by with time and space-varying magnetic elds, allowing imaging of the material that they make up. Quantum computing: attempts to build a quantum computer involve trying to put together multiple systems of this kind (qubits), keeping them isolated from perturbations by the environment, but still allowing interaction with the system in a way that preserves its quantum behavior. The 2012 Physics Nobel prize was awarded for experimental work making progress in this direction.
So far in this course weve been describing what is known as the Schr odinger picture of quantum mechanics. States in H are functions of time, obeying the Schr odinger equation determined by a Hamiltonian observable H , while observable self-adjoint operators A are time-independent. Time evolution is given by a unitary transformation U (t) = ei
t
One can instead use U (t) to make a unitary transformation that puts the time-dependence in the observables, removing it from the states, as follows: | (t) | (t)
H
where the H subscripts for Heisenberg indicate that we are dealing with Heisenberg picture observables and states. One can easily see that the physically observable quantities given by eigenvalues and expectations values remain the same:
H
(t)|AH | (t)
In the Heisenberg picture the dynamics is given by a dierential equation not for the states but for the operators. Recall from our discussion of the adjoint
representation the formula d tX tX d d (e Y e ) = ( (etX Y ))etX + etX Y ( etX ) dt dt dt = XetX Y etX etX Y etX X Using this with Y = A, X = i we nd H
d H i AH (t) = [i , AH (t)] = [H, AH (t)] dt and this equation determines the time evolution of the observables in the Heisenberg picture. Applying this to the case of the spin one-half system in a magnetic eld, and taking for our observable S we nd d i eg SH (t) = [H, SH (t)] = i [SH (t) B, SH (t)] dt 2mc We know from earlier that the solution will be SH (t) = U (t)SH (0)U (t)1 for U (t) = et
ge|B| B 2mc s |B|
and thus the spin vector observable evolves in the Heisenberg picture by rotating |B| about the magnetic eld vector with angular velocity ge 2mc .
There is a dierent approach one can take to characterizing states of a quantum system with H = C2 . Multiplication of vectors in H by a non-zero complex number does not change eigenvectors, eigenvalues or expectation values, so arguably has no physical eect. Multiplication by a real scalar just corresponds to a change in normalization of the state, and we will often use this freedom to work with normalized states, those satisfying | = 1. With normalized states, one still has the freedom to multiply states by a phase ei without changing eigenvectors, eigenvalues or expectation values. In terms of group theory, the overall U (1) in the unitary group U (2) acts on H acts on H by a representation of U (1), which can be characterized by an integer, the corresponding charge, but this decouples from the rest of the observables and is not of much interest. One is mainly interested in the SU (2) part of the U (2), and the observables that correspond to its Lie algebra.
Working with normalized states in this case corresponds to working with unit-length vectors in C2 , which are given by points on the unit sphere S 3 . If we dont care about the overall U (1) action, we can imagine identifying all states that are related by a phase transformation. Using this equivalence relation we can dene a new set, whose elements are the cosets, elements of S 3 C2 , with elements that dier just by multiplication by ei identied. The set of these elements forms a new geometrical space, called the coset space, often written S 3 /U (1). This structure is called a bering of S 3 by circles, and is known as the Hopf bration. Try an internet search for various visualizations of the geometrical structure involved, a surprising way decomposition of threedimensional space into non-intersecting curves. The same space can be represented in a dierent way, as C2 /C , by taking all elements of C2 and identifying those related by muliplication by a non-zero complex number. If we were just using real numbers, R2 /R can be thought of as the space of all lines in the plane going through the origin.
One sees that each such line hits the unit circle in two opposite points, so this set could be parametrized by a semi-circle, identifying the points at the two ends. This space is given the name RP 1 , the real projective line, and the analog space of lines through the origin in Rn is called RP n1 . What we are interested in is the complex analog CP 1 , which is often called the complex projective line. To better understand CP 1 , one would like to put coordinates on it. A
the complex number z1 /z2 . Overall multiplication by a complex number will drop out in this ratio, so one gets dierent values for each of the of the coordinate z1 /z2 for each dierent coset element, and it appears that elements of CP 1 correspond to points on the complex plane. There is however one problem with this coordinate: the coset of 1 0 does not have a well-dened value: as one approaches this point one moves o to innity in the complex plane. In some sense the space CP 1 is the complex plane, but with a point at innity added. It turns out that CP 1 is best thought of not as plane together with a point, but as a sphere, with the relation to the plane and the point at innity given by stereographic projection. Here one creates a one-to-one mapping by considering the lines that go from a point on the sphere to the north pole of the sphere. Such lines will intersect the plane in a point, and give a one-to-one mapping between points on the plane and points on the sphere, except for the north pole. Now, one can identify the north pole with the point at innity, and thus the space CP 1 can be identied with the space S 2 . The picture looks like this
and the equations relating coordinates (X1 , X2 , X3 ) on the sphere and the complex coordinate z1 /z2 = z = x + iy on the plane are given by x= X1 X2 , y= 1 X3 1 X3
and X1 = x2 2y x2 + y 2 1 2x , X2 = 2 , X3 = 2 2 2 +y +1 x +y +1 x + y2 + 1
z + z1 z2 z +
Such transformations of the complex plane are conformal (angle-preserving) transformations known as M obius transformations. One can check that the corresponding transformation on the sphere is the rotation of the sphere in R3 corresponding to this SU (2) = Spin(3) transformation. In physics language, this sphere CP 1 is known as the Bloch sphere. It provides a useful parametrization of the states of the qubit system, up to scalar multiplication, which is supposed to be physically irrelevant. The North pole is the spin-up state, the South pole is the spin-down state, and along the equator one nds the two states that have denite values for S1 , as well as the two that have denite values for S2 .
10
Notice that the inner product on vectors in H does not correspond at all to the inner product of unit vectors in R3 . The North and South poles of the Bloch sphere correspond to orthogonal vectors in H, but they are not at all orthogonal thinking of the corresponding points on the Bloch sphere as vectors int R3 . Similarly, eigenvectors for S1 and S2 are orthogonal on the Bloch sphere, but not at all orthogonal int H. Many of the properties of the Bloch sphere parametrization of states in H are special to the fact that H = C2 . In the next class we will study systems of spin n n 2 , where H = C . In these cases there is still a two-dimensional Bloch sphere, but only certain states in H are parametrized by it. We will see other examples of systems with coherent states analogous to the states parametrized by the Bloch sphere, but the case H has the special property that all states (up to scalar multiplication) are such coherent states.
Just about every quantum mechanics textbook works out this example of a spin 1/2 particle in a magnetic eld. For one arbitrarily chosen example, see Chapter 14 of [1] or
11
References
[1] Shankar, R., Principles of Quantum Mechanics, 2nd Ed., Springer, 1994.
12