Category Archives: Quantum Physics

Should the sign of Coulomb potential be positive or negative?

In classical physics, the sign of a potential (say, gravitational potential or electric potential) is merely a convention. For example, since the electric field $\mathbb{E}$ is conservative, it is the gradient of some potential, which we call the electric potential or Coulomb potential $V$, so it can be written as $\mathbb{E}=\nabla V$ or $\mathbb{E}=-\nabla V$. Mathematically, it doesn’t matter whichever you use. If you choose $\mathbb{E}=\nabla V$, Coulomb potential should be $V=-\frac{1}{4\pi\epsilon_0}\frac{Q}{r}$ and if you choose $\mathbb{E}=-\nabla V$, $V=\frac{1}{4\pi\epsilon_0}\frac{Q}{r}$. In physics, the usual convention is $\mathbb{E}=-\nabla V$ and it is interpreted as electric field pointing downhill towards lower voltages.

On the other hand, in quantum mechanics the sign of Coulomb potential matters and it does depend on the problem that you are studying. In a hydrogen atom or a hydrogen-like atom (for example, an ionized helium atom $\mathrm{He}^+$) an electron is trapped in the atom by Coulomb force between proton and electron. In this case, we say the electron is in a bound state. The bound state with the minimum energy is called the ground state and the minimum energy is called the ground state energy. A bound state with energy higher than the ground state energy is said to be unstable and the electron in an unstable bound state always tend to move to the ground state by emitting photons. Modeling bound state of a hydrogen atom or a hydrogen-like atom requires negative Coulomb potential $V(r)=-\frac{Z\alpha}{r}$ as seen in the following figure.

Coulomb potential (in red) with bound state energy (in blue)

In this simple figure, the electron is trapped in the region $0<r<1$. The figure also clearly shows you that bound state energy must be negative. By drawing a picture, one can easily see that bound state cannot be modeled with positive Coulomb potential, but positive Coulomb potential is used to model scattering of a particle.

The Momentum Representation

This note is based on my friend Khin Maung’s short lecture.

Let us begin with Schrödinger equation
$$\left(\frac{\hat p^2}{2m}+\hat V(r)\right)|\psi\rangle=E|\psi\rangle$$
Use the completeness relation
$$1=\int|\vec{p}\rangle\langle\vec{p}|d\vec{p}$$
to get the momentum space representation of the Schrödinger equation
\begin{equation}
\label{eq:schrodingerms}
\frac{p^2}{2m}\psi(\vec{p})+\int\langle\vec{p}|\hat V(r)|\vec{p’}\rangle\psi(\vec{p’})d\vec{p’}=E\psi(\vec{p})
\end{equation}
where $\psi(\vec{p}):=\langle\vec{p}|\psi\rangle$. \eqref{eq:schrodingerms} is called the Schrödinger equation in momentum space. Using the completeness relation
$$1=\int|\vec{r}\rangle\langle\vec{r}|d\vec{r}$$
we obtain
$$\langle\vec{p}|\hat V(r)|\vec{p’}\rangle=\frac{1}{(2\pi\hbar)^3}\int e^{\frac{i}{\hbar}(\vec{p’}-\vec{p})\cdot\vec{r}}V(r)d\vec{r}$$
Here, recall that $\langle\vec{p}|\vec{r}\rangle=\frac{1}{(2\pi\hbar)^{\frac{3}{2}}}e^{-\frac{i}{\hbar}\vec{p}\cdot\vec{r}}$. Let $\vec{q}=\vec{p’}-\vec{p}$ and $V(\vec{q})=\langle\vec{p}|\hat V(r)|\vec{p’}\rangle$. Then we have
$$V(\vec{q})=\frac{1}{(2\pi\hbar)^3}\int e^{\frac{i}{\hbar}\vec{q}\cdot\vec{r}}V(r)d\vec{r}$$
This is just the Fourier transform of $V(r)$. For the Yukawa potential
$$V(r)=V_0\frac{e^{-\mu r}}{r}$$
\begin{align*} V(\vec{q})&=\frac{1}{(2\pi\hbar)^3}\int e^{\frac{i}{\hbar}qr\cos\theta}\int_0^{2\pi}\int_0^\pi\int_0^\infty V(r)r^2dr\sin\theta d\theta d\phi\\ &=\frac{1}{(2\pi)^2\hbar^3}\int_0^\infty V(r)r^2\int_{-1}^1e^{\frac{i}{\hbar}qru}dudr\\ &=\frac{1}{(2\pi\hbar)^2iq}\int_0^\infty V(r)r(e^{\frac{i}{\hbar}qr}-e^{-\frac{i}{\hbar}qr})dr\\ &=\frac{V_0}{(2\pi\hbar)^2iq}\int_0^\infty e^{-\mu r}(e^{\frac{i}{\hbar}qr}-e^{-\frac{i}{\hbar}qr})dr\\ &=\frac{V_0}{2\pi^2\hbar^3}\frac{1}{\mu^2+\frac{q^2}{\hbar^2}} \end{align*}
From here on, we assume that $\hbar=1$ for simplicity.

Let
$$\psi(\vec{p})=\psi_l(p)Y_l^m(\hat p)$$
Here, $\hat p$ stands for the unit vector in the momentum space in spherical coordinates $\hat p=\frac{\vec{p}}{p}=(\theta_l,\phi_p)$ where $\theta_l$ is the polar angle and $\phi_p$ is the azimuth angle corresponding to the momentum vector $\vec{p}$. $V(\vec{q})$ can be written as the series
\begin{align*} V(\vec{q})&=\sum_{l=0}^\infty\sum_{m=-l}^lV_l(p,p’)Y_l^m(\hat p)Y_l^{m}(\hat p’)\\
&=\sum_{l=0}^\infty\frac{2l+1}{4\pi}V_l(p,p’)P_l(x)
\end{align*} The last line is obtained by the addition theorem for spherical harmonics $$\frac{4\pi}{2l+1}\sum_{m=-l}^lY_l^m(\hat p)Y_l^{m}(\hat p’)=P_l(x)$$
where $x=\cos\theta_{pp’}$.
By the orthonormality of spherical harmonics
$$\int_0^{2\pi}\int_0^\pi Y_{l_1}^{m_1}(\hat p)Y_{l_2}^{m_2}(\hat p)\sin\theta d\theta d\phi=\delta_{l_1l_2}\delta_{m_1m_2}$$
the Schrödinger equation in momentum space \eqref{eq:schrodingerms} yields
$$\frac{p^2}{2m}\psi_l(p)+\int_0^\infty V_l(p,p’)\psi_l(p’)p’^2dp’=E\psi_l(p)$$
Using the orthogonality of Legendre polynomials
$$\int_{-1}^1P_{l’}(x)P_l(x)dx=\frac{2}{2l+1}\delta_{ll’}$$
we obtain
$$V_l(p,p’)=2\pi\int_{-1}^1V(\vec{q})P_l(x)dx$$
For the Yukawa potential, we have
$$V(\vec{q})=\frac{V_0}{2\pi^2}\frac{1}{\mu^2+q^2}=\frac{V_0}{2\pi^2}\frac{1}{p^2+p’^2+\mu^2-2pp’x}$$
The Neumann’s formula yields then
$$V_l(p,p’)=\frac{V_0}{pp’\pi}Q_l\left(\frac{p^2+p’^2+\mu^2}{2pp’}\right)$$
Note here that we require $\left|\frac{p^2+p’^2+\mu^2}{2pp’}\right|>1$. Recall that
$$Q_0(z)=\frac{1}{2}\ln\frac{z+1}{z-1}$$
for $|z|>1$. Hence for the Yukawa potential, $V_0(p,p’)$ can be written as
$$V_0(p,p’)=\frac{V_0}{2pp’\pi}\ln\frac{(p+p’)^2+\mu^2}{(p-p’)^2+\mu^2}$$
With $V_0=Z\alpha$ and $\mu=0$, the Yukawa potential reduces to the Coulomb potential $V(r)=\frac{Z\alpha}{r}$. $V_l(p,p’)$ is then given by$$V_l(p,p’)=\frac{Z\alpha}{pp’\pi}Q_l\left(\frac{p^2+p’^2}{2pp’}\right)$$
and accordingly,$$V_0(p,p’)=\frac{Z\alpha}{pp’\pi}\ln\left|\frac{p+p’}{p-p’}\right|$$

Solving a bound state Schrödinger equation in momentum space for a delta potential in position space

In this note, we first obtain Schrödinger equation in momentum space and then solve a bound state problem with delta potential in position space. This is based on what I learned from my friend and a physicist Khin Maung. Let us begin with eigenstate Schrödinger equation
$$\hat H|\psi\rangle=E|\psi\rangle$$
with
$$\hat H=\frac{\hat p^2}{2m}+\hat V$$
in position space. As we will see Schrödinger equation in momentum space becomes an integral equation, while Schrödinger equation in position space is a differential equation.
Before we continue, here is the list of a few things we need for our discussion.

  • The completeness: $1=\int |p\rangle\langle p|dp$, $1=\int |x\rangle\langle x|dx$.
  • Eigenvalue equations: \begin{align*} \hat p|p’\rangle&=p’|p’\rangle\\ \hat x|x’\rangle&=x’|x’\rangle \end{align*}
  • $\langle p|p’\rangle=\delta(p’-p)$, $\langle x|x’\rangle=\delta(x’-x)$.
  • $\langle p|x\rangle=\frac{e^{-\frac{i}{\hbar}px}}{\sqrt{2\pi\hbar}}$, $\langle x|p’\rangle=\frac{e^{\frac{i}{\hbar}p’x}}{\sqrt{2\pi\hbar}}$.

Details about these can be found in any standard quantum mechanics textbook, particularly in the chapter about quantum systems with continuous spectra.

Using the completeness, we have
$$\int\left(\frac{\hat p^2}{2m}+\hat V\right)|p’\rangle\langle p’|\psi\rangle dp’=E|\psi\rangle$$
and
\begin{equation}
\label{eq:seq}
\int\langle p|\left(\frac{\hat p^2}{2m}+\hat V\right)|p’\rangle\tilde\psi(p’)dp’=E\tilde\psi(p)
\end{equation}
where $\tilde\psi(p)=\langle p|\psi\rangle$. Since $\hat p|p’\rangle=p’|p’\rangle$, $\hat p^2|p’\rangle=p’^2|p’\rangle$. So, \eqref{eq:seq} can be written as
\begin{equation}
\label{eq:seq2}
\int\langle p|p’\rangle\frac{p’^2}{2m}\tilde\psi(p’)dp’+\int\langle p|\hat V|p’\rangle\tilde\psi(p’)dp’=E\tilde\psi(p)
\end{equation}
The LHS of \eqref{eq:seq2} is then
$$\int\delta(p’-p)\frac{p’^2}{2m}\tilde\psi(p’)dp’+\int\langle p|\hat V|p’\rangle\tilde\psi(p’)dp’=\frac{p^2}{2m}\tilde\psi(p)+\int\langle p|\hat V|p’\rangle\tilde\psi(p’)dp’$$
Hence, we arrive at the one-dimensional Schrödinger equation in momentum space (or $p$-space)
\begin{equation}
\label{eq:seq3}
\frac{p^2}{2m}\tilde\psi(p)+\int\langle p|\hat V|p’\rangle\tilde\psi(p’)dp’=E\tilde\psi(p)
\end{equation}
\begin{align*} \langle p|\hat V|p’\rangle&=\langle p|I|\hat V|I|p’\rangle\\ &=\langle p|\int|x\rangle\langle x|dx|\hat V|\int|x’\rangle\langle x’|dx’|p’\rangle\\ &=\int\langle p|x\rangle\langle x|\hat V|x’\rangle\langle x’|p’\rangle dxdx’ \end{align*}
Since $\hat V$ is a scalar potential,
\begin{align*} \langle x|\hat V|x’\rangle&=\langle x|x’\rangle\hat V\\ &=\delta(x-x’)V(x) \end{align*}
So, we have
\begin{align*} \langle p|\hat V|p’\rangle&=\int\langle p|x\rangle\delta(x-x’)V(x)\langle x’|p’\rangle dx’dx\\ &=\int\langle p|x\rangle V(x)\langle x|p’\rangle dx\\ &=\int \frac{e^{-ipx}}{\sqrt{2\pi}}V(x)\frac{e^{ip’x}}{\sqrt{2\pi}}dx\ [\mbox{assuming}\ \hbar=1]\\ &=\frac{1}{2\pi}\int V(x)e^{i(p’-p)x}dx \end{align*}
If we want to keep $\hbar$, we have
$$\langle p|\hat V|p’\rangle=\frac{1}{2\pi\hbar}\int V(x) e^{\frac{i}{\hbar}(p’-p)x}dx$$

Now, we consider a special case of a delta potential
$$V(x)=-\delta(x)V_0$$
\begin{align*} \langle p|\hat V|p’\rangle&=\frac{1}{2\pi}\int_{-\infty}^\infty[-\delta(x)V_0e^{i(p’-p)x}]dx\\ &=-\frac{V_0}{2\pi} \end{align*}
The Schrödinger equation \eqref{eq:seq3} then becomes
\begin{equation}
\label{eq:seq4}
\left(\frac{p^2}{2m}-E\right)\tilde\psi(p)=\frac{V_0}{2\pi}\int_{-\infty}^\infty\tilde\psi(p’)dp’
\end{equation}
Note here that the bound state energy must be negative, so we assume that $E<0$. Let
$$A=\int_{-\infty}^\infty\tilde\psi(p’)dp’$$
Then \eqref{eq:seq4} can be written as
$$\tilde\psi(p)=\frac{V_0}{2\pi}\frac{A}{\frac{p^2}{2m}-E}$$
Integrating this with respect to $p$, we obtain
\begin{align*} 1&=\frac{V_0}{2\pi}\int_{-\infty}^\infty\frac{dp}{\frac{p^2}{2m}-E}\\ &=\frac{V_0}{2}\sqrt{\frac{2m}{-E}} \end{align*}
Hence, we find the bound state energy
\begin{equation}
\label{eq:bde}
E=-\frac{V_0^2m}{2}
\end{equation}
For $\hbar\ne 1$, we have
$$E=-\frac{V_0^2m}{2\hbar^2}$$
Recall that $\tilde\psi(p)$ has $A$ in it. $A$ can be found by normalizing $\tilde\psi(p)$.
$$1=\int_{-\infty}^\infty |\tilde\psi(p)|^2dp$$
Since
$$|\tilde\psi(p)|^2=\frac{V_0^2}{4\pi^2}\frac{|A|^2}{\left(\frac{p^2}{2m}-E\right)^2},$$
we have
\begin{equation}
\label{eq:sqe5}
1=\frac{V_0^2}{4\pi^2}|A|^2\int_{-\infty}^\infty\frac{dp}{\left(\frac{p^2}{2m}-E\right)^2}
\end{equation}
Let
$$I=\int_{-\infty}^\infty\frac{dp}{\left(\frac{p^2}{2m}-E\right)^2}=\sqrt{2m}\pi(-E)^{-\frac{1}{2}}$$
Then
$$\frac{\partial I}{\partial E}=\int_{-\infty}^\infty\frac{dp}{\left(\frac{p^2}{2m}-E\right)^2}=\sqrt{\frac{m}{2}}\pi(-E)^{-\frac{3}{2}}$$
Therefore, from \eqref{eq:sqe5} with \eqref{eq:bde}, we obtain
$$A=\sqrt{2\pi mV_0}e^{i\theta}$$
for some $\theta$.

No $\sqrt{\mathrm{NOT}}$ in PT-Symmetric Quantum Mechanics

In PT-symmetric quantum mechanics, the inner product is given by
\begin{equation}
\label{eq:pt-product}
\langle v_1|P|v_2\rangle
\end{equation}
where $P=\begin{bmatrix}
0 & 1\\
1 & 0
\end{bmatrix}$.
A PT-symmetric operator $H$ is a $2\times 2$ symmetric matrix that satisfies
$$\langle Hv_1|P|v_2\rangle=\langle v_1|P|Hv_2\rangle$$
which is equivalent to
$$\bar HP=PH$$
So, PT-symmetric operators are self-adjoint operators with respect to the inner product \eqref{eq:pt-product}. PT-symmetric operators take the form
$$\begin{bmatrix}
a & b\\
b & \bar a
\end{bmatrix}$$
In PT-symmetric quantum mechanics, $\mathrm{NOT}$ gate is given by
$$\mathrm{NOT}=\begin{bmatrix}
1 & 0\\
0 &-1
\end{bmatrix}$$
The reason for this is that in PT-symmetric quantum mechanics the qubits are
$$|0\rangle=\frac{1}{\sqrt{2}}\begin{bmatrix}
1\\
1
\end{bmatrix},\ |1\rangle=\frac{1}{\sqrt{2}}\begin{bmatrix}
1\\
-1
\end{bmatrix}$$
The eigenvectors of $\mathrm{NOT}$ corresponding to the eigenvalues $\pm 1$ are, respectively, $\begin{bmatrix}
1\\
0
\end{bmatrix}$ and $\begin{bmatrix}
0\\
1
\end{bmatrix}$ and their norms are zero with respect to the inner product \eqref{eq:pt-product}. This means that the spectral decomposition of $\mathrm{NOT}$ in terms of its eigenvectors does not exist in PT-symmetric quantum mechanics. This also hints us that $\sqrt{\mathrm{NOT}}$ does not likely exist. In PT-symmetric quantum mechanics, an operator $U$ is called $P$-unitary if $U^\dagger PU=P$ and $P$-antiunitary if $U^\dagger PU=-P$.

For the same physical reason, in PT-symmetric quantum mechanics, gates are required to be $P$-unitary or $P$-antiunitary. $\mathrm{NOT}$ is $P$-antiunitary. It can be easily seen that $\sqrt{\mathrm{NOT}}$ does not exist. If $\sqrt{\mathrm{NOT}}$ exists, then it is either $P$-unitary or $P$-antiunitary. If $\sqrt{\mathrm{NOT}}$ is $P$-unitary, then
\begin{align*} \mathrm{NOT}^\dagger P\mathrm{NOT}&=\sqrt{\mathrm{NOT}}^\dagger(\sqrt{\mathrm{NOT}}^\dagger P\sqrt{\mathrm{NOT}})\sqrt{\mathrm{NOT}}\\ &=\sqrt{\mathrm{NOT}}^\dagger P\sqrt{\mathrm{NOT}}\\ &=P \end{align*}
This contradicts to the fact that $\mathrm{NOT}$ is $P$-antiunitary. If $\sqrt{\mathrm{NOT}}$ is $P$-antiunitary, then
\begin{align*} \mathrm{NOT}^\dagger P\mathrm{NOT}&=\sqrt{\mathrm{NOT}}^\dagger(\sqrt{\mathrm{NOT}}^\dagger P\sqrt{\mathrm{NOT}})\sqrt{\mathrm{NOT}}\\ &=-\sqrt{\mathrm{NOT}}^\dagger P\sqrt{\mathrm{NOT}}\\ &=P \end{align*}
Again, this is a contradiction. Therefore, $\sqrt{\mathrm{NOT}}$ does not exist. As seen here, the standard hermitian quantum mechanics admits $\sqrt{\mathrm{NOT}}$ gate. This implies that the standard hermitian quantum mechanics and PT-symmetric quantum mechanics cannot be consistent with each other. While no one has yet come up with a physical implementation of $\sqrt{\mathrm{NOT}}$ gate, it is interesting to see that the physical viability of a quantum theory hangs on the physical realizability of a quantum gate.

Square Roots of Operators

Mathematics is full of weird stuff. One of them is the square root of an operator. So far, I have seen only two books that discuss the square root of an operator. They are listed in the references below.

Definition. An operator $R$ is called a square root of an operator $T$ is $R^2=T$.

Example. In quantum computing, $\sqrt{\mathrm{NOT}}$ gate is given by
$$\sqrt{\mathrm{NOT}}=\begin{bmatrix}
\frac{1+i}{2} & \frac{1-i}{2}\\
\frac{1-i}{2} & \frac{1+i}{2}
\end{bmatrix}$$
and
$$\sqrt{\mathrm{NOT}}\cdot\sqrt{\mathrm{NOT}}=\mathrm{NOT}=\begin{bmatrix}
0 & 1\\
1 & 0
\end{bmatrix}$$
As required by quantum mechanics, quantum gates are unitary matrices. There is no counterpart of $\sqrt{\mathrm{NOT}}$ gate in classical computing, so $\sqrt{\mathrm{NOT}}$ gate is a truly quantum gate. As far as I know, no one has come up with a physical implementation of $\sqrt{\mathrm{NOT}}$ gate yet.

Example. An operator does not necessarily have a square root. For example, define $T:\mathbb{C}^3\longrightarrow\mathbb{C}^3$ by
$$T(z_1,z_2,z_3)=(z_2,z_3,0)$$
Then one can easily show that $T$ is linear. Suppose that $T$ has a square root $R$. Let $\begin{bmatrix}
r & s & t\\
u & v & w\\
z & y & z
\end{bmatrix}$ be the matrix associated with $R$. Since the matrix associated with $T$ is $\begin{bmatrix}
0 & 1 & 0\\
0 & 0 & 1\\
0 & 0 & 0
\end{bmatrix}$, we have the equation
$$\begin{bmatrix}
r & s & t\\
u & v & w\\
z & y & z
\end{bmatrix}\cdot\begin{bmatrix}
r & s & t\\
u & v & w\\
z & y & z
\end{bmatrix}=\begin{bmatrix}
0 & 1 & 0\\
0 & 0 & 1\\
0 & 0 & 0
\end{bmatrix}$$
This is a system of 9 scalar equations. One can attempt to solve this equation using a CAS (Computer Algebra System), for example, Maple and see that the system does not have a solution, i.e. $R$ does not exist.

So, what kind of operators have square roots?

Theorem. Identity + Nilpotent has a square root.

Proof. Let $N$ be nilpotent. Then $N^m=0$ for some positive integer $m$. Consider the Taylor series for $\sqrt{1+x}$:
$$\sqrt{1+x}=1+a_1x+a_2x^2+a_3x^3+\cdots$$
Using the nilpotency of $N$, we may guess that $\sqrt{I+N}$ takes the form
$$I+a_1N+a_2N^2+\cdots+a_{m-1}N^{m-1}$$
Now,
\begin{align*} (I+a_1N+a_2N^2+\cdots+a_{m-1}N^{m-1})^2=I&+2a_1N+(2a_2+a_1^2)N^2\\&+(2a_3+2a_1a_2)N^3+\cdots\\ &+(2a_{m-1}+\cdots)N^{m-1}=I+N \end{align*}
and by comparing the coefficients
$a_1=\frac{1}{2}$, $2a_2+a_1^2=0$, so $a_2=-\frac{1}{8}$, $2a_3+2a_1a_2=0$, so $a_3=\frac{1}{16}$, and so and forth.

Theorem. Suppose that $T: V\longrightarrow V$ is invertible. Then $T$ has a square root.

Proof. Let $\lambda_1,\lambda_2,\cdots,\lambda_m$ be the distinct eigenvalues of $T$. Then for each $j$, there exists a nilpotent operator $N_j: G(\lambda_j,T)\longrightarrow G(\lambda_j,T)$ such that $T|{G(\lambda_j,T)}=\lambda_jI+N_j$. (See [1] , p 252, Theorem 8.21 for a proof.) Here, $G(\lambda_j,T)$ is the eigenspace of $T$ corresponding to the eigenvalue $\lambda_j$. Since $T$ is invertible, no $\lambda_j$ is equal to 0, so we can write $$T|{G(\lambda_j,T)}=\lambda_j\left(I+\frac{N}{\lambda_j}\right)$$
Since $\frac{N}{\lambda_j}$ is nipotent, $I+\frac{N}{\lambda_j}$ has a square root. Let $R_j$ be $\sqrt{\lambda_j}$ times the square root of $I+\frac{N}{\lambda_j}$. Any vector $v\in V$ can be written uniquely in the form
$$v=u_1+u_2+\cdots+u_m,$$
where each $u_j$ is in $G(\lambda_j,T)$. Using this decomposition, define an operator $R:V\longrightarrow V$ by
$$Rv=R_1u_1+R_2 u_2+\cdots+R_m u_m$$
Then $R^2u_j=R_j^2 u_j=T|{G(\lambda_j,T)}u_j=\lambda_ju_j$ and so \begin{align*} R^2v&=R_1^2u_1+R_2^2u_2+\cdots+R_m^2u_m\\ &= T|{G(\lambda_1,T)}u_1+T|{G(\lambda_2,T)}u_2+\cdots+T|{G(\lambda_m,T)}u_m=Tv
\end{align*}
Therefore, $R$ is a square root of $T$.

Example. The $\mathrm{NOT}$ gate is invertible. It is an inverse of itself.

The proof of the above theorem suggests that an operator $T$ has a square root if there is a spectral decomposition of $T$ with respect to the standard inner product of $\mathbb{C}^n$. A normal operator is such an operator. Recall that an operator $T: V\longrightarrow V$ is normal if $T^\ast T=TT^\ast$. The following theorem guarantees the existence of such a spectral decomposition of a normal operator.

Theorem. A matrix $T$ is normal if and only if there exists a diagonal matrix $\Lambda$ and an unitary matrix $U$ such that $T=U\Lambda U^\ast$.

Here, the diagonal entries of $\Lambda$ are the eigenvalues of $T$ and the columns of $U$ are the eigenvectors of $T$. The matching eigenvalues in $\Lambda$ come in the same order as the eigenvectors are ordered as columns of $U$.

Let $T: V\longrightarrow V$ be a normal operator and write it as its spectral decomposition, using Dirac’s braket notation,
$$T=\sum_j\lambda_j|v_j\rangle\langle v_j|$$
Define
$$f(T)=\sum_j f(\lambda_j)|v_j\rangle\langle v_j|$$
Using this definition, one can define, for example, the exponential, square root, and logarithmic operator of $T$.

Example. Let $T=\begin{bmatrix}
0 & 1\\
1 & 0
\end{bmatrix}$. Find the square root of $T$.

Solution. $T$ is hermitian (symmetric), so it is normal. The spectral decomposition of $T$ is given by
$$T=|v_1\rangle\langle v_1|-|v_2\rangle\langle v_2|,$$
where
$$|v_1\rangle=\frac{1}{\sqrt{2}}\begin{bmatrix}
1\\
1
\end{bmatrix}\ \mathrm{and}\ |v_2\rangle=\frac{1}{\sqrt{2}}\begin{bmatrix}
1\\
-1
\end{bmatrix}$$
Now,
\begin{align*} \sqrt{T}&=|v_1\rangle\langle v_1|+i|v_2\rangle\langle v_2|\\ &=\frac{1}{2}\begin{bmatrix} 1 & 1\\ 1 & 1 \end{bmatrix}+\frac{i}{2}\begin{bmatrix} 1 & -1\\ -1 & 1 \end{bmatrix}\\ &=\begin{bmatrix} \frac{1+i}{2} & \frac{1-i}{2}\\ \frac{1-i}{2} & \frac{1+i}{2} \end{bmatrix} \end{align*}

References:

  1. Sheldon Axler, Linear Algebra Done Right, Third Edition, Springer, 2015
  2. Michael A. Nielsen and Isaac L. Chuang, Quantum Computation and Quantum Information, Cambridge University Press, 2000