[Physics] Help trying to understand why the matrix representation of $J^2$ is what it is

angular momentumhilbert-spaceoperatorsquantum mechanicsrepresentation-theory

Say we a basis of kets such as$$
\beta ~ := ~\left\{\left|j=1,\, m=-1\right> , ~ \left|j=1, \, m=0\right>, ~ \left|j=1, \, m=1\right>\right\}
\,.$$

Then it's plain to see why in matrix representation$$J_z~=~h \begin{pmatrix} 1&0&0\\0&0&0\\0&0&-1 \end{pmatrix}$$
as this just represents $m$ being either $0,$$1$ or $-1$.

But why is$$
J^2=2h^2 \begin{pmatrix} 1&0&0\\0&1&0\\0&0&1 \end{pmatrix}
\,?$$

It can't be gotten from just $J_z$ and I don't know what the matrix representation for $J_x, J_y$ nor do I feel i'm expected to know them as they weren't described in class. So how can one compute the matrix representation of $J^2$?

Best Answer

The short answer is that $J^2$ is a scalar under all rotations, and the only matrix that has that property is the identity matrix, so you have to have $J^2\propto I$.

I think that the process and physics are a bit easier to understand if we start from the elemental rotation matrices in $3$-d about each axis \begin{align} R_x & = \left[\begin{array}{ccc} 1 & 0 & 0 \\ 0 & \cos\theta & -\sin\theta \\ 0 & \sin\theta & \cos\theta \end{array}\right] & R_y & = \left[\begin{array}{ccc} \cos\theta & 0 & \sin\theta \\ 0 & 1 & 0 \\ -\sin\theta & 0 & \cos\theta \end{array}\right] & R_z & = \left[\begin{array}{ccc} \cos\theta & -\sin\theta & 0\\ \sin\theta & \cos\theta & 0 \\ 0 & 0 & 1 \end{array}\right]. \end{align} where the pattern of signs is chosen so that a positive rotation gives a rotation around the axis in the sense defined to be positive by the right hand rule. A matrix/operator is said to generate a transformation if, for an infinitesimal transformation, it satisfies $$R(\theta) = I - i \frac{J\theta}{\hbar} + \mathcal{O}(\theta^2),$$ where the transformation is $R$ and the generator is $J$. From that we get that \begin{align} J_x & = \left[\begin{array}{ccc} 0 & 0 & 0 \\ 0 & 0 & -i\hbar \\ 0 & i\hbar & 0 \end{array}\right] & J_y & = \left[\begin{array}{ccc} 0 & 0 & i\hbar \\ 0 & 0 & 0 \\ -i\hbar & 0 & 0 \end{array}\right] & J_z & = \left[\begin{array}{ccc} 0 & -i\hbar & 0 \\ i\hbar & 0 & 0 \\ 0 & 0 & 0 \end{array}\right]. \end{align}

We now look for the eigenvectors and eigenvalues of $J_z$. The defining equation for the eigen-problem is $$J_z \vec{v} = \lambda \vec{v}$$ for some eigenvalue $\lambda$ and its corresponding eigenvector $\vec{v}$. We get that the eigenvalue/eigenvector pairs of $J_z$ are \begin{align} \lambda_{-1} & = -1 \Rightarrow & \vec{v}_{-1} & = \frac{1}{\sqrt{2}}\left[\begin{array}{c} i \\ 1 \\ 0 \end{array}\right], \\ \lambda_0 & = 0 \Rightarrow & \vec{v}_0 & = \left[\begin{array}{c} 0 \\ 0 \\ i \end{array}\right],\ \mathrm{and} \\ \lambda_1 & = 1 \Rightarrow & \vec{v}_1 & = \frac{1}{\sqrt{2}}\left[\begin{array}{c} -i \\ 1 \\ 0 \end{array}\right], \end{align} where I have taken care to normalize the eigenvectors to unit length (i.e. $\vec{v}^* \cdot \vec{v} = 1$), and to set the overall phases of the unit vectors to make the results match certain other conventions.

It is a defining property of the eigenvalue problem for Hermitian matrices that if we define $V \equiv \left[\vec{v}_1,\ \vec{v}_0,\ \vec{v}_{-1}\right]$, the matrix with normalized distinct eigenvectors in the columns, then \begin{align} V & = \left[\begin{array}{ccc} -\frac{i}{\sqrt{2}} & 0 & \frac{i}{\sqrt{2}} \\ \frac{1}{\sqrt{2}} & 0 & \frac{1}{\sqrt{2}} \\ 0 & i & 0 \end{array}\right],\ \mathrm{and} & V^\dagger J_z V &= \hbar \left[\begin{array}{ccc} 1 & 0 & 0 \\ 0 & 0 & 0 \\ 0 & 0 & -1 \end{array}\right]. \end{align} Note that $V^\dagger V=I$ and $\operatorname{det}(V)=1$, meaning $V$ is a special unitary matrix. That fact means that $V$ is a transformation from one orthonormal basis to another. If we say that the above matrices are just representations of a more abstract operator in a particular basis, and use a $\rightarrow$ to signify representation, then we can say that in this basis $$J_z\rightarrow \hbar \left[\begin{array}{ccc} 1 & 0 & 0 \\ 0 & 0 & 0 \\ 0 & 0 & -1 \end{array}\right].$$ If we examine $J_x$ and $J_y$ in this basis, we get \begin{align} J_x &\rightarrow \frac{\hbar}{\sqrt{2}} \left[\begin{array}{ccc} 0 & 1 & 0 \\ 1 & 0 & 1 \\ 0 & 1 & 0 \end{array}\right],\ \mathrm{and} & J_y \rightarrow i\frac{\hbar}{\sqrt{2}} \left[\begin{array}{ccc} 0 & -1 & 0 \\ 1 & 0 & -1 \\ 0 & 1 & 0 \end{array}\right]. \end{align}

Now you can evaluate $J^2 = J_x^2 + J_y^2 + J_z^2$ explicitly.

Notice how our original basis was in terms of the three states that are invariant under rotations about their respective axes, the $m_x=0$, $m_y=0$, and $m_z=0$ states, respectively.

Note also that the choice of phases made won't effect $J^2$. That was done so that the $J_x$ and $J_y$ constructed will match those generated by following the development in Wikipedia's ladder operator article.

You can make a more general construction if you look at the commutation relations among $J_x$, $J_y$ and $J_z$ as defined above, and rename the $x$ through $z$ axes as $1$ through $3$ to get that $$[J_i,\, J_j] = i\hbar \sum_{k=1}^3 \epsilon_{ijk} J_k, \tag1$$ with $\epsilon_{ijk}$ the Levi-Civita symbol. We then say that any collection of three Hermitian matrices that satisfies the commutation relations in (1) are generators of the symmetry transformation we call rotations in physics, in some particular representation/basis. The next step is to prove that $$[J^2,\, J_i] = 0,$$ which shows that the $J^2$ and $J_z$ can be simultaneously digaonalized. I have forgotten the next step in the process, but this more general approach works for all possible values of orbital angular momentum and spin, making it much more powerful.

Related Question