To be used in future
 Section 3.2: Examples of Symmetries Up Chapter 3: Symmetries Section 3.4: Unitary operators 

3.3 Consequences for Quantum Mechanics

\textbookShankar, chapter 11
Note: In this section we shall only look at non-relativistic problems; relativity will reappear later!

3.3.1 Role of symmetry

We first need to address the question “what is a symmetry in quantum mechanics?”. If we work in the standard Schrödinger picture with time-dependent wave functions, and time-independent operators [E]  [E] This allows for time-dependent external forces. --there are alternatives-- we must be able to link the symmetry directly to the wave function, which is of course the most fundamental quantity in QM. At the same time we can also link it to the behaviour of general matrix elements if we change the coordinates in the both wave functions under, e.g., one of the transformations discussed above. The question we now need to ask ourselves, is “how would we see a symmetry if we transform the argument of both wave functions in a matrix element under a transformation in a symmetry group?”. The answer is simple: since the matrix element describes physics, the matrix element itself should be invariant (unchanged) for any pair of wave functions if our system (particles and operator) have a symmetry. In most cases the operator that we shall use is the Hamiltonian.
Let us thus start by calculating the matrix elements of the Hamiltonian If the quantum Hamiltonian is invariant under a symmetry, where the transformation gives the effect of the symmetry on the coordinates (i.e., is the active transformation), we must require that For instance, is translationally invariant and invariant under space inversion (also under time reversal, but that will be investigated later). A single-particle Hamiltonian with an external potential depending on the magnitude of , is rotationally invariant, but not invariant under translations.

3.3.2 Active and passive view

What we have shown above is again the active view of such transformations in quantum mechanics, where we transform the wave functions. We shall analyse below the detailed behaviour of such transformations, but we shall assume that we can write [F]  [F] In mathematical physics we often use the term “a representation” of the transformation for the calculation of the operator on the wave function giving the action of a symmetry operator on the coordinates. We are looking at a linear representation here. where is a linear operator giving the effect of the change of coordinates, but by acting on the wave function, rather than acting on its arguments. This is a new intermediate stage--and it may not even be clear that such an operator actually exists; we shall construct these explicitly. We now derive the passive view from Eq. (): In the quantum passive view we transform operators (rather than the wave function), and we find that the operators transform according to Using the passive view, if we have a symmetry, we must require that the transformed operator is invariant, i.e., is the same as the original one,

3.3.3 Symmetry generators

In the case of a continuous symmetry, one with a set of parameters that vary continuously (which are all transformations but the discrete ones above), we can take the continuous parameter arbitrarily small, where the transformation has almost no effect, and analyse its behaviour when acting on the coordinates inside a wave function using a Taylor expansion. Note that since we transform coordinates, all our transformations carry the minus sign common to active transformations.

3.3.3.1 Translations: momentum

Consider a translation over a very small distance in passive form, [G]  [G] We ignore the effect of a potential phase acquired by the wave function when we make such a transformation. See [1] and [3]. We use this expression and say that the operator is the generator of translations.
There is no great surprise that we find this generator. If were a function of time, and represents a wave packet, we clearly add momentum to our system (since it moves). Conversely, if we give a state a small momentum we expect it to translate as time increases. No
If we expand the expectation value () above (suppressing time dependence) to first order in , we find that or, expanding the left-hand side to first order in and equating the coefficient of to zero, Since and are arbitrary functions, we must have and thus the generator of the symmetry commutes with the Hamiltonian if it is compatible with the symmetry.
This works exactly the same way for the other transformations:
Time translations
This one is easy! Look at ,
where we used the time-dependent Schrödinger equation The generator is thus .
Rotations
This one is a little harder; we use the following three rotation matrices If we make the angles small, we find that We now calculate the effect on of a general transformation  [H]  [H] The dot product of and denotes the matrix . If we analyse each of the three ’s separately, we find for instance for a rotation about that
Equivalently this can be done using the Levi-Civitta symbol, gives
We find that the three generators of rotational symmetry are . Each commutes with the Hamiltonian, but they do not commute with each other,  [I]  [I] See also example sheet

3.3.4Discrete symmetries

Discrete symmetries don’t have generators, since there is no small parameter to expand in. If they are realised by an action on the arguments of the wave function (again, excluding an external phase that could occur in general) they are rather simple. We then define the linear symmetry operator by its action on the wave function.

3.3.4.1 Space inversion

For space inversion we define the “parity operator (sometimes denoted as ) that inverts the coordinates The fact that , as we see from () shows that we have eigenvalues and : Thus in a problem with parity invariance, the wave function is either even or odd under space inversion. We say it has positive or negative parity. We leave the fact that momentum changes sign as an exercise.

3.3.4.2 Time reversal

For time reversal we define the time reversal operator The fact that , as we see from () shows again that we have eigenvalues and .
Since time place a special role in the Schrödinger equation, we need to interpret this result carefully. Using the fact that the time evolution can be found from the Schrödinger equation, we find that for a Hermitian Hamiltonian. (We thus conclude that .) We conclude Since time reversal gives we conclude that for a time reversal invariant problem we can choose basis functions that satisfy and thus eigenstates of are either purely real or purely imaginary. We leave the fact that the momentum operator changes sign under this transformation as an exercise.

3.3.5 Simultaneous diagonalisation of operators

You should be aware from your previous work in quantum mechanics that two operators that commute can be simultaneously diagonalised; i.e., a state can be an energy eigenstate and a symmetry eigenstate at the same time.
This fails if the symmetry operators are not mutually commuting, as for the three components of angular momentum. We then need to choose a maximal commuting set of operators to label the states (in the example, usually , and ).

3.3.6 Lorentz Boosts

Even though we will not linger on Lorentz symmetry for too long before having discussed relativistic wave equations, it is easy to show that given the wave function and a small velocity boost along the axis, we find We shall leave this expression at this stage: we have carelessly defined the operator ; since we know there is a link between time derivatives and the action of the Hamiltonian, this needs to be done much more carefully, as will be done later on.
 Section 3.2: Examples of Symmetries Up Chapter 3: Symmetries Section 3.4: Unitary operators