You are on page 1of 19

.

3.1 The two-state quantum system


3.1.1 The Pauli matrices: observables of the twostate quantum system
For a quantum system with two-dimensional state
space H=C , observables are self-adjoint linear
operators on C2. With respect to a chosen basis of
C2,
these are 2 by 2 complex matricesM satisfying the
conditionM = My (My is the
conjugate transpose of M). Any such matrix will be a
(real) linear combination
of four matrices:
M = c01 + c1_1 + c2_2 + c3_3
with cj 2 R and the standard choice of basis
elements given by
1=
_
10
01
_
; _1 =
_
01
10
_
; _2 =
_
0 i
i0
_
; _3 =
_
10
0 1

_
The _j are called the \Pauli matrices" and are a
pretty universal choice of basis
in this subject. This choice of basis is a convention,
with one aspect of this
convention that of taking the basis element in the 3direction to be diagonal.
In common physical situations and conventions, the
third direction is the distinguished
\up-down" direction in space, so often chosen when
a distinguished
direction in R3 is needed.
Recall that the basic principle of how measurements
are supposed to work
in quantum theory says that the only states that
have well-de_ned values for
these four observables are the eigenvectors for
these matrices. The _rst matrix
gives a trivial observable (the identity on every
state), whereas the last one, _3,
has the two eigenvectors
_3
_
1
0
_
=
_
1
0
_
and
_3
_
0
1
_

=
_
0
1
_
with eigenvalues +1 and 1. In quantum information
theory, where this is
the qubit system, these two eigenstates are labeled
j0i and j1i because of the
analogy with a classical bit of information. Later on
when we get to the theory of
spin, we will see that 1
2_3 is the observable corresponding to the SO(2) =
U(1)
symmetry group of rotations about the third spatial
axis, and the eigenvalues
24
1
2 ;+1
2 of this operator will be used to label the two
eigenstates
j+
1
2
i=
_
1
0
_
and j
1
2
i=
_
0
1
_

The two eigenstates j+ 1


2 i and j 1
2 i provide a basis for C2, so an arbitrary
vector in H can be written as
j i = _j +
1
2
i + _j
1
2
i
for _; _ 2 C. Only if _ or _ is 0 does the observable _3
correspond to a wellde
_ned number that characterizes the state and can
be measured. This will be
either 1
2 (if _ = 0 so the state is an eigenvector j + 1
2 i), or 1
2 (if _ = 0 so the
state is an eigenvector j 1
2 i).
An easy to check fact is that j+ 1
2 i and j 1
2 i are NOT eigenvectors for the
operators _1 and _2. One can also check that no pair
of the three _j commute,
which implies that one cannot _nd vectors that are
simultaneous eigenvectors for
more than one _j . This non-commutativity of the
operators is responsible for the
characteristic classically paradoxical property of
quantum observables: one can
_nd states with a well de_ned number for the
measured value of one observable
_j , but such states will not have a well-de_ned
number for the measured value

of the other two non-commuting observables. The


physical description of this
phenomenon in the realization of this system as a
spin 1
2 particle is that if one
prepares states with a well-de_ned spin component
in the j-direction, the two
other components of the spin can't be assigned a
numerical value in such a
state. Any attempt to prepare states that
simultaneously have speci_c chosen
numerical
values
for
the
3
observables
corresponding to the _j is doomed. So is
any attempt to simultaneously measure such
values: if one measures the value
for a particular observable _j , then going on to
measure one of the other two
will ensure that the _rst measurement is no longer
valid (repeating it will not
necessarily give the same thing). There are many
subtleties in the theory of
measurement for quantum systems, but this simple
two-state example already
shows some of the main features of how the
behavior of observables is quite
di_erent than in classical physics.
The choice we have made for the _j corresponds to a
choice of basis for H
such that the basis vectors are eigenvectors of _3.
_1 and _2 take these basis
vectors to non-trivial linear combinations of basis
vectors. It turns out that
there are two speci_c linear combinations of _1 and
_2 that do something very
simple to the basis vectors, since
(_1 + i_2) =
_

02
00
_
and (_1 i_2) =
_
00
20
_
we have
(_1 + i_2)
_
0
1
_
=2
_
1
0
_
(_1 + i_2)
_
1
0
_
=
_
0
0
_
25
and
(_1 i_2)
_
1
0
_
=2

_
0
1
_
(_1 i_2)
_
0
1
_
=
_
0
0
_
(_1 + i_2) is called a \raising operator": on
eigenvectors of _3 it either
increases the eigenvalue by 2, or annihilates the
vector. (_1 i_2) is called
a \lowering operator": on eigenvectors of _3 it either
decreases the eigenvalue
by 2, or annihilates the vector. Note that these
linear combinations are not
self-adjoint, (_1 + i_2) is the adjoint of (_1 i_2) and
vice-versa.
3.1.2 Exponentials of Pauli matrices: unitary
transformations
of the two-state system
We saw in chapter 2 that in the U(1) case, knowing
the observable operator Q on
H determined the representation of U(1), with the
representation matrices found
by exponentiating i_Q. Here we will _nd the
representation corresponding to
the two-state system observables by exponentiating
the observables in a similar
way.

Taking the the identity matrix _rst, multiplication by


i_ and exponentiation
gives the diagonal unitary matrix
ei_1 =
_
ei_ 0
0 ei_
_
This is just exactly the case studied in chapter 2, for
a U(1) group acting on
H = C2, with
Q=
_
10
01
_
This matrix commutes with any other 2 by 2 matrix,
so we can treat its action
on H independently of the action of the _j .
Turning to the other three basis elements of the
space of observables, the
Pauli matrices, it turns out that since all the _j satisfy
_2
j = 1, their exponentials
also take a simple form.
ei__j = 1 + i__j +
1
2
(i_)2_2
j+
1
3!
(i_)3_3
j+___
= 1 + i__j
1
2

_21 i
1
3!
_3_j + _ _ _
= (1
1
2!
_2 + _ _ _ )1 + i(_
1
3!
_3 + _ _ _ )_j
= (cos _)1 + i_j(sin _) (3.1)
As _ goes from _ = 0 to _ = 2_, this exponential
traces out a circle in the
space of unitary 2 by 2 matrices, starting and
ending at the unit matrix. This
circle is a group, isomorphic to U(1). So, we have
found three di_erent U(1)
26
subgroups inside the unitary 2 by 2 matrices, but
only one of them (the case
j = 3) will act diagonally on H, with the U(1)
representation determined by
Q=
_
10
0 1
_
For the other two cases j = 1 and j = 2, by a change
of basis one could put
either one in the same diagonal form, but doing this
for one value of j makes
the other two no longer diagonal. All three values of
j need to be treated
simultaneously, and one needs to consider not just
the U(1)s but the group one

gets by exponentiating general linear combinations


of Pauli matrices.
To compute such exponentials, one can check that
these matrices satisfy the
following relations, useful in general for doing
calculations with them instead of
multiplying out explicitly the 2 by 2 matrices:
[_j ; _k]+ = _j_k + _k_j = 2_jk1
Here [_; _]+ is the anticommutator. This relation
says that all _j satisfy _2
j=1
and distinct _j anticommute (e.g. _j_k = _k_j for j
6= k).
Notice that the anticommutation relations imply
that, if we take a vector
v = (v1; v2; v3) 2 R3 and de_ne a 2 by 2 matrix by
v _ _ = v1_1 + v2_2 + v3_3 =
_
v3 v1 iv2
v1 + iv2 v3
_
then taking powers of this matrix we _nd
(v _ _)2 = (v2
1 + v2
2 + v2
3)1 = jvj21
If v is a unit vector, we have
(v _ _)n =
(
1 n even
(v _ _) n odd
Replacing _j by v _ _, the same calculation as for
equation 3:1 gives (for v
a unit vector)
ei_v__ = (cos _)1 + i(sin _)v _ _
Notice that one can easily compute the inverse of
this matrix:

(ei_v__)1 = (cos _)1 i(sin _)v _ _


since
((cos _)1 + i(sin _)v _ _)((cos _)1 i(sin _)v _ _) =
(cos2 _ + sin2 _)1 = 1
We'll review linear algebra and the notion of a
unitary matrix in chapter 4, but
one form of the condition for a matrix M to be
unitary is
My = M1
27
so the self-adjointness of the _j implies unitarity of
ei_v__ since
(ei_v__)y = ((cos _)1 + i(sin _)v _ _)y
= ((cos _)1 i(sin _)v _ _y)
= ((cos _)1 i(sin _)v _ _)
= (ei_v__)1
One can also easily compute the determinant of
ei_v__, _nding
det(ei_v__) = det((cos _)1 + i(sin _)v _ _)
= det
_
cos _ + i sin _v3 i sin _(v1 iv2)
i sin _(v1 + iv2) cos _ i sin _v3
_
= cos2 _ + sin2 _(v2
1 + v2
2 + v2
3)
=1
So, we see that by exponentiating i times linear
combinations of the selfadjoint
Pauli matrices (which all have trace zero), we get
unitary matrices of
determinant one. These are invertible, and form the
group named SU(2), the
group of unitary 2 by 2 matrices of determinant one.
If we exponentiated not

just i_v _ _, but i(_1 + _v _ _) for some real constant


_ (such matrices will not
have trace zero unless _ = 0), we would get a
unitary matrix with determinant
ei2_. The group of unitary 2 by 2 matrices with
arbitrary determinant is called
U(2). It contains as subgroups SU(2) as well as the
U(1) described at the
beginning of this section. U(2) is slightly di_erent
than the product of these
two subgroups, since the group element
_
1 0
0 1
_
is in both subgroups. In our review of linear algebra
to come we will encounter
the generalization to SU(n) and U(n), groups of
unitary n by n complex matrices.
To get some more insight into the structure of the
group SU(2), consider an
arbitrary 2 by 2 complex matrix
_
__
_
_
Unitarity implies that the rows are orthonormal. One
can see this explicitly
from the condition that the matrix times its
conjugate-transpose is the identity
_
__
_
__
_
__
_

=
_
10
01
_
Orthogonality of the two rows gives the relation
_ + __ = 0 =) _ =
_
_
28
The condition that the _rst row has length one gives
__ + __ = j_j2 + j_j2 = 1
Using these two relations and computing the
determinant (which has to be 1)
gives
__ _ =
__
_
_=
_
(__ + __) =
_
=1
so one must have
= _; _ = _
and an SU(2) matrix will have the form
_
__
_ _
_
where (_; _) 2 C2 and
j_j2 + j_j2 = 1
So, the elements of SU(2) are parametrized by two
complex numbers, with

the sum of their length-squareds equal to one.


Identifying C2 = R4, these are
just vectors of length one in R4. Just as U(1) could
be identi_ed as a space with
the unit circle S1 in C = R2, SU(2) can be identi_ed
with the unit three-sphere
S3 in R4.
3.2 Commutation relations for Pauli matrices
An important set of relations satis_ed by Pauli
matrices are their commutation
relations:
[_j ; _k] = _j_k _k_j = 2i
X3
l=1
_jkl_l
where _jkl satis_es _123 = 1, is antisymmetric under
permutation of two of its
subscripts, and vanishes if two of the subscripts take
the same value. More
explicitly, this says:
[_1; _2] = 2i_3; [_2; _3] = 2i_1; [_3; _1] = 2i_2
One can easily check these relations by explicitly
computing with the matrices.
Putting
together
the
anticommutation
and
commutation relations, one gets a
formula for the product of two Pauli matrices:
_j_k = _jk1 + i
X3
l=1
_jkl_l
While physicists prefer to work with self-adjoint Pauli
matrices and their
real eigenvalues, one can work instead with the
following skew-adjoint matrices
Xj = i
_j
2

29
which satisfy the slightly simpler commutation
relations
[Xj ;Xk] =
X3
l=1
_jklXl
or more explicitly
[X1;X2] = X3; [X2;X3] = X1; [X3;X1] = X2
If these commutators were zero, the SU(2) elements
one gets by exponentiating
linear combinations of the Xj would be commuting
group elements. The
non-triviality of the commutators reects the noncommutativity of the group.
Group elements U 2 SU(2) near the identity satisfy
U ' 1 + _1X1 + _2X2 + _3X2
for _j small and real, just as group elements z 2 U(1)
near the identity satisfy
z ' 1 + i_
One can think of the Xj and their commutation
relations as an in_nitesimal
version of the full group and its group multiplication
law, valid near
the identity. In terms of the geometry of manifolds,
recall that SU(2) is the
space S3. The Xj give a basis of the tangent space
R3 to the identity of
SU(2), just as i gives a basis of the tangent space to
the identity of U(1).
30
3.3 Dynamics of a two-state system
Recall that the time dependence of states in
quantum mechanics is given by the
Schrodinger equation
d
dt

j (t)i = iHj (t)i


where H is a particular self-adjoint linear operator on
H, the Hamiltonian operator.
The most general such operator on C2 will be given
by
H = h01 + h1_1 + h2_2 + h3_3
for four real parameters h0; h1; h2; h3. The solution
to the Schrodinger equation
is just given by exponentiation:
j (t)i = U(t)j (0)i
where
U(t) = eitH
The h01 term in H just contributes an overall phase
factor eih0t, with the
remaining factor of U(t) an element of the group
SU(2) rather than the larger
group U(2) of all 2 by 2 unitaries.
Using our earlier equation
ei_v__ = (cos _)1 + i(sin _)v _ _
valid for a unit vector v, our U(t) is given by taking h
= (h1; h2; h3), v = h
jhj
and _ = tjhj, so we _nd
U(t) =eih0t(cos(tjhj)1 + i sin(tjhj)
h1_1 + h2_2 + h3_3
jhj
)
=eih0t(cos(tjhj)1 i sin(tjhj)
h1_1 + h2_2 + h3_3
jhj
)
=
eih0t(cos(tjhj) i h3
jhj sin(tjhj)) i sin(tjhj) h1ih2
jhj
i sin(tjhj) h1+ih2

jhj eih0t(cos(tjhj) + i h3
jhj sin(tjhj))
!
In the special case h = (0; 0; h3) we have
U(t) =
_
eit(h0+h3) 0
0 eit(h0h3)
_
so if our initial state is
j (0)i = _j +
1
2
i + _j
1
2
i
for _; _ 2 C, at later times the state will be
j (t)i = _eit(h0+h3)j +
1
2
i + _eit(h0h3)j
1
2
i
31
In this special case, one can see that the
eigenvalues of the Hamiltonian are
h0 _ h3.
In the physical realization of this system by a spin
1=2 particle (ignoring its
spatial motion), the Hamiltonian is given by
H=
ge
4mc
(B1_1 + B2_2 + B3_3)

where the Bj are the components of the magnetic


_eld, and the physical constants
are the gyromagnetic ratio (g), the electric charge
(e), the mass (m) and
the speed of light (c), so we have solved the
problem of the time evolution of
such a system, setting hj = ge
4mcBj . For magnetic _elds of size jBj in the 3direction, we see that the two di_erent states with
well-de_ned energy (j + 1
2i
and j 1
2 i) will have an energy di_erence between them of
ge
2mc
jBj
This is known as the Zeeman e_ect and is readily
visible in the spectra of atoms
subjected to a magnetic _eld. We will consider this
example in more detail in
chapter 7, seeing how the group of rotations of R3
appears. Much later, in
chapter 43, we will derive this Hamiltonian term
from general principles of how
electromagnetic _elds couple to such spin 1=2
particles.
3.4 For further reading
Many quantum mechanics textbooks now begin with
the two-state system, giving
a much more detailed treatment than the one given
here, including much
more about the physical interpretation of such
systems (see for example [78]).
Volume III of Feynman's Lectures on Physics [18] is a
quantum mechanics text
with much of the _rst half devoted to two-state
systems. The _eld of \Quantum

Information Theory" gives a perspective on quantum


theory that puts such systems
(in this context called the \qubit") front and center.
One possible reference
for this material is John Preskill's notes on quantum
computation [56].

You might also like