Pauli matrices

Pauli matrices are matrix representations of the spin operators \small \hat{S}_x, \small \hat{S}_y and \small \hat{S}_z.

To derive the Pauli matrices, we let \small \alpha and \small \beta be basis vectors representing the electron spin eigenstates of \small \vert s=1/2,m_s=1/2\rangle and \small \vert s=1/2,m_s=-1/2\rangle respectively, with the following assignments:

\small \alpha=\begin{pmatrix} 1\\0 \end{pmatrix}\; \; \; \; \; \; \beta=\begin{pmatrix} 0\\1 \end{pmatrix}

The general state of an electron can then be written as a linear combination of the two spin states:

\small \chi=c_1\alpha+c_2\beta\; \; \; \; \; \; \; \; 171

where \small \vert c_1\vert^{2} is the probability of finding the electron with the state \small \vert s=1/2,m_s=1/2\rangle, and \small \vert c_2\vert^{2} is the probability of finding the electron with the state \small \vert s=1/2,m_s=-1/2\rangle, with

\small \vert c_1\vert^{2}+\vert c_2\vert^{2}=1\; \; \; \; \; \; \; \; 172

Having defined the two eigenstates, we can work out the corresponding matrix representation of \small \hat{S}^{2} using eq169 by letting \small \hat{S}^{2}=\begin{pmatrix} a &b \\ c &d \end{pmatrix}:

\small \begin{pmatrix} a &b \\ c &d \end{pmatrix}\begin{pmatrix} 1\\0 \end{pmatrix}=\frac{3}{4}\hbar^{2}\begin{pmatrix} 1\\0 \end{pmatrix}\; \; \; \; \; \Rightarrow \; \; \; \; \; \begin{pmatrix} a\\c \end{pmatrix}=\begin{pmatrix} \frac{3}{4}\hbar^{2}\\ 0 \end{pmatrix}

\small \begin{pmatrix} a &b \\ c &d \end{pmatrix}\begin{pmatrix} 0\\1 \end{pmatrix}=\frac{3}{4}\hbar^{2}\begin{pmatrix} 0\\1 \end{pmatrix}\; \; \; \; \; \Rightarrow \; \; \; \; \; \begin{pmatrix} b\\d \end{pmatrix}=\begin{pmatrix} 0\\ \frac{3}{4}\hbar^{2} \end{pmatrix}

So,

\small \hat{S}^{2}=\frac{3}{4}\hbar^{2}\begin{pmatrix} 1 &0 \\ 0 &1 \end{pmatrix}\; \; \; \; \; \; \; \; 173

Similarly, using eq168 for \small \hat{S}_z, we have

\small \hat{S}_z=\frac{\hbar}{2}\begin{pmatrix} 1 &0 \\ 0 &-1 \end{pmatrix}\; \; \; \; \; \; \; \; 174

For \small \hat{S}_x and \small \hat{S}_y, we make use of eq170 where

\small \hat{S}_+\begin{pmatrix} 0\\1 \end{pmatrix}=\hbar\sqrt{\frac{3}{4}+\frac{1}{2}\left ( -\frac{1}{2}+1 \right )}\begin{pmatrix} 1\\0 \end{pmatrix}=\hbar\begin{pmatrix} 1\\0 \end{pmatrix}\; \; \; \; \; \; \Rightarrow \; \; \; \; \; \hat{S}_+=\hbar\begin{pmatrix} 0 &1 \\ 0& 0 \end{pmatrix}\; \; \; \; \; \; \; \; 175

\small \hat{S}_-\begin{pmatrix} 1\\0 \end{pmatrix}=\hbar\sqrt{\frac{3}{4}-\frac{1}{2}\left (\frac{1}{2}-1 \right )}\begin{pmatrix} 0\\1 \end{pmatrix}=\hbar\begin{pmatrix} 0\\1 \end{pmatrix}\; \; \; \; \; \; \Rightarrow \; \; \; \; \; \hat{S}_-=\hbar\begin{pmatrix} 0 &0 \\ 1& 0 \end{pmatrix}\; \; \; \; \; \; \; \; 176

Since \small \hat{S}_\pm=\hat{S}_x\pm i\hat{S}_y, we have \small \hat{S}_x=\hat{S}_+-i\hat{S}_y=\hat{S}_+-(\hat{S}_x-\hat{S}_-) or \small \hat{S}_x=\frac{\hat{S}_++\hat{S}_-}{2}. Similarly, \small \hat{S}_y=\frac{\hat{S}_+-\hat{S}_-}{2i}. Using eq175 and eq176,

\small \hat{S}_x=\frac{\hbar}{2}\begin{pmatrix} 0 &1 \\ 1 & 0 \end{pmatrix}\; \; \; \; \; \; \; \; 177

\small \hat{S}_y=\frac{\hbar}{2}\begin{pmatrix} 0 &-i \\ i & 0 \end{pmatrix}\; \; \; \; \; \; \; \; 178

In an earlier article, when we constructed quantum orbital angular momentum component operators \small \hat{L}_x, \small \hat{L}_y and \small \hat{L}_z, we replaced the position and linear momentum components of the classical angular momentum components \small L_x, \small L_y and \small L_z with their corresponding operators. We also suggested that we could have constructed an angular momentum operator using \small \boldsymbol{\mathit{L}}=\boldsymbol{\mathit{i}}L_x+\boldsymbol{\mathit{j}}L_y+\boldsymbol{\mathit{k}}L_z, such that \small \hat{\boldsymbol{\mathit{L}}}=\boldsymbol{\mathit{i}}\hat{L}_x+\boldsymbol{\mathit{j}}\hat{L}_y+\boldsymbol{\mathit{k}}\hat{L}_z, but for certain reasons chose to construct \small \hat{L}^{2} instead. For electron spin, it is useful to construct the spin operator using that suggestion:

\small \hat{\boldsymbol{\mathit{S}}}=\boldsymbol{\mathit{i}}\hat{S}_x+\boldsymbol{\mathit{j}}\hat{S}_y+\boldsymbol{\mathit{k}}\hat{S}_z=\frac{\hbar}{2}\hat{\boldsymbol{\mathit{\sigma}}}\; \; \; \; \; \; \; \; 179

where \small \hat{\boldsymbol{\mathit{\sigma}}}=\boldsymbol{\mathit{i}}\hat{\sigma}_x+\boldsymbol{\mathit{j}}\hat{\sigma}_y+\boldsymbol{\mathit{k}}\hat{\sigma}_z, with

\small \hat{\sigma}_x=\begin{pmatrix} 0 &1 \\ 1 &0 \end{pmatrix},\; \; \; \; \;\hat{\sigma}_y=\begin{pmatrix} 0 &-i \\ i &0 \end{pmatrix},\; \; \; \; \;\hat{\sigma}_z=\begin{pmatrix} 1 &0 \\ 0 &-1 \end{pmatrix}\; \; \; \; \; \; \; \; 180

\small \hat{\sigma}_x, \small \hat{\sigma}_y and \small \hat{\sigma}_z are called Pauli matrices. They represent observables and are Hermitian (the complex conjugate of each matrix returns the same matrix). Eq179 and eq180 are used to analyse the results of successive Stern-Gerlach experiments.

 

Question

Show that \small [\hat{L}^{2},\hat{\boldsymbol{\mathit{L}}}\cdot\hat{\boldsymbol{\mathit{S}}}]=0 and \small [\hat{S}^{2},\hat{\boldsymbol{\mathit{L}}}\cdot\hat{\boldsymbol{\mathit{S}}}]=0 but \small \left [ \hat{{L}^{2}}^{T},\hat{\boldsymbol{\mathit{L}}_1}\cdot\hat{\boldsymbol{\mathit{S}}_1}+\hat{\boldsymbol{\mathit{L}}_2}\cdot\hat{\boldsymbol{\mathit{S}}_2}\right ]\neq 0.

Answer

For \small \left [ \hat{L}^{2},\hat{\boldsymbol{\mathit{L}}}\cdot\hat{\boldsymbol{\mathit{S}}}\right ], substitute in \small \hat{\boldsymbol{\mathit{L}}}=\boldsymbol{\mathit{i}}\hat{L}_x+\boldsymbol{\mathit{j}}\hat{L}_y+\boldsymbol{\mathit{k}}\hat{L}_z and \small \hat{\boldsymbol{\mathit{S}}}=\boldsymbol{\mathit{i}}\hat{S}_x+\boldsymbol{\mathit{j}}\hat{S}_y+\boldsymbol{\mathit{k}}\hat{S}_z and expand the expression.  From eq102, \small \hat{L}^{2} commutes with all components of \small \hat{\boldsymbol{\mathit{L}}}; and also all components of \small \hat{\boldsymbol{\mathit{S}}} because orbital angular momentum operators and spin angular momentum operators act on different vector spaces. Thus, \small \left [ \hat{L}^{2},\hat{\boldsymbol{\mathit{L}}}\cdot\hat{\boldsymbol{\mathit{S}}}\right ]=0. Similarly, from eq165, eq166 and eq167, \small \hat{S}^{2} commutes with all components of \small \hat{\boldsymbol{\mathit{L}}} and \small \hat{\boldsymbol{\mathit{S}}} and so \small \left [ \hat{S}^{2},\hat{\boldsymbol{\mathit{L}}}\cdot\hat{\boldsymbol{\mathit{S}}}\right ]=0.

Now,

\small \hat{{L}^{2}}^{T}=\hat{\boldsymbol{\mathit{L}}}^{T}\cdot\hat{\boldsymbol{\mathit{L}}}^{T}=\left (\hat{\boldsymbol{\mathit{L}}_1}+\hat{\boldsymbol{\mathit{L}}_2}\right )\cdot\left (\hat{\boldsymbol{\mathit{L}}_1}+\hat{\boldsymbol{\mathit{L}}_2}\right )=\hat{L}_{1}^{\: 2} +\hat{L}_{2}^{\: 2}+2\left (\hat{\boldsymbol{\mathit{L}}_1}\cdot\hat{\boldsymbol{\mathit{L}}_2}\right ) \; \; \; \; \; \; \; \; 181

Substituting eq181 in \small \left [{\hat{L^{2}}^{T}},\hat{\boldsymbol{\mathit{L}}_1}\cdot\hat{\boldsymbol{\mathit{S}}_1}+\hat{\boldsymbol{\mathit{L}}_2}\cdot\hat{\boldsymbol{\mathit{S}}_2}\right ]\neq 0 and expanding the expression, we find, after some algebra, that

\small \left [{\hat{L^{2}}^{T}},\hat{\boldsymbol{\mathit{L}}_1}\cdot\hat{\boldsymbol{\mathit{S}}_1}+\hat{\boldsymbol{\mathit{L}}_2}\cdot\hat{\boldsymbol{\mathit{S}}_2}\right ]\neq 0\; \; \; \; \; \; \; \; 182

Similarly,

\left [ \hat{{S}^2}^T,\hat{\boldsymbol{\mathit{L}}}_1\cdot\hat{\boldsymbol{\mathit{S}}}_1+\hat{\boldsymbol{\mathit{L}}}_2\cdot\hat{\boldsymbol{\mathit{S}}}_2\right ]\neq0\; \; \; \; \; \; \; \; 182a

 

 

Next article: Sequential Stern-Gerlach experiments
Previous article: spin angular momentum
Content page of quantum mechanics
Content page of advanced chemistry
Main content page

Leave a Reply

Your email address will not be published. Required fields are marked *