Time-Independent Perturbation Theory (Non-Degenerate) #
This content is primarily based on Section 8, “Approximation Method,” in Chapter 2 of Physics of Atoms and Molecules (2nd Edition) by B. H. Bransden and C. J. Joachain1.
1. Summary #
We aim to solve the time-independent Schrödinger equation (TISE):
$$ \begin{align} (\hat{H}_0+\hat{H}’)\varphi = E \varphi \end{align} $$
to determine the $k$-th eigenenergy $E_k$ and eigenfunction $\varphi_k$.
The effect of $\hat{H}’$ on $\hat{H}_0$ is expressed in terms of the order parameter $\lambda \ll 1$, where $\hat{H}_0 = O(\lambda^0)$ and $\hat{H}’ = O(\lambda^1)$.
The solution $E_k$ and $\varphi_k$ of the TISE \eqref{e20} are expanded as a power series in $\lambda$:
$$ \begin{align} \varphi_k &= \varphi_k^{(0)} + \varphi_k^{(1)} + \varphi_k^{(2)} + \cdots \\ E_k &= E_k^{(0)} + E_k^{(1)} + E_k^{(2)} + \cdots \end{align} $$
Here, $E^{(0)}$ and $\varphi^{(0)}$ are solutions of $\hat{H}_0\varphi^{(0)} = E^{(0)} \varphi^{(0)}$, and $E_k^{(s)} = O(\lambda^s)$, $\varphi_k^{(s)} = O(\lambda^s)$ denote terms of order $\lambda^s$. The solution can then be expressed as:
$$ \begin{align} \varphi_k^{(s)}&=\sum_{m\ne k} a_{mk}^{(s)}\varphi_m^{(0)} \\ E_k^{(s)}&=\langle\varphi_k^{(0)}|\hat{H}’|\varphi_k^{(s-1)}\rangle = \sum_{m\ne k} H’_{km}a_{mk}^{(s-1)} \end{align} $$
where the following substitutions are made:
$$ \begin{align} a_{mk}^{(s)} =\frac{1}{E_m^{(0)}-E_k^{(0)}}\biggl[\sum_{j=1}^{s-1} E_k^{(s-j)}a_{mk}^{(j)} - \sum_{n} H’_{mn} a_{nk}^{(s-1)} \biggr],~~(s\geq 1, m\ne k) \end{align} $$
$$ \begin{align} H’_{lm} = \langle\varphi_l^{(0)}|\hat{H}’|\varphi_m^{(0)}\rangle \end{align} $$
For specific problems, truncating the perturbative solution at an appropriate order results in the wavefunction evolving as follows.
2. Problem Setting #
We consider a system described by a time-independent Hamiltonian:
$$ \begin{align} \label{e1} \hat{H}=\hat{H}_0+\hat{H}’ \end{align} $$
and aim to solve the time-independent Schrödinger equation (TISE) to determine the eigenenergy $E = E_k$ and the corresponding eigenfunction $\varphi = \varphi_k$ of the system.
In other words, the TISE we aim to solve in this study is:
$$ \begin{align} \label{e2} \hat{H}\varphi = E \varphi \end{align} $$
The system Hamiltonian \eqref{e1} is expressed as the sum of two Hamiltonians, distinguished and named as follows:
- $\hat{H}_0$: The unperturbed Hamiltonian, which is of order $\lambda^0$.
- $\hat{H}’$: The perturbation Hamiltonian, which is of order $\lambda^1$.
The difference between $\hat{H}_0$ and $\hat{H}’$ lies in the degree of influence $\lambda$ that they exert on the total system Hamiltonian $\hat{H}$.
Here, $\lambda$ is a small parameter, $\lambda \ll 1$, representing a relative magnitude of $\hat{H}’$ to $\hat{H}_0$. In the strict limit $\lambda = 0$, the system reduces to $\hat{H} = \hat{H}_0$, and the perturbation Hamiltonian $\hat{H}’$ vanishes.
Furthermore, we assume that the TISE for the unperturbed Hamiltonian $\hat{H}_0$ has already been solved. That is:
$$ \begin{align} \label{e3} \hat{H}_0 \varphi_k^{(0)} = E_k^{(0)} \varphi_k^{(0)} \end{align} $$
The eigenenergies $E_k^{(0)}$ and the corresponding eigenfunctions $\varphi_k^{(0)}$ are assumed to be known. The operator $\hat{H}_0$ is assumed to be Hermitian, ensuring that the eigenenergies are real.
The eigenfunctions $\varphi_k^{(0)}$ are normalized and orthogonal, satisfying:
$$ \begin{align} \label{e4} \langle\varphi_n^{(0)}|\varphi_m^{(0)}\rangle &= \delta_{n,m},&&(\text{For discrete state}) \\ &= \delta(n-m),&&(\text{For continuum state}) \end{align} $$
For discrete states, $\delta_{i,j}$ refers to the Kronecker delta, while for continuous states, it represents the Dirac delta function. In subsequent discussions, orthogonality will be expressed using $\delta_{i,j}$, which implicitly includes the continuous case.
Under these conditions, we proceed to solve TISE \eqref{e2}. While the approach differs significantly for degenerate and non-degenerate cases, this study focuses solely on the non-degenerate case.
2.1. Solution Method #
The problem to solve is restated as follows:
$$ \begin{align} \label{e5} \left(\hat{H}_0+\hat{H}’\right)\varphi = E \varphi \end{align} $$
Considering the limit $\lambda \to 0$, the $k$-th eigenstate $\varphi_k$ of TISE \eqref{e5} is assumed to deviate only slightly from $\varphi_k^{(0)}$. This forms the basis for perturbation theory. Conversely, perturbation theory should not be applied to problems where $\varphi_k^{(0)}$ is significantly distorted2.
In the vicinity of $\lambda \to 0$, the solutions of TISE \eqref{e5} can be expanded as a power series in $\lambda$:
$$ \begin{align} E_k &= E_k^{(0)} + E_k^{(1)} + E_k^{(2)} + \cdots \label{e6}\\ \varphi_k &= \varphi_k^{(0)} + \varphi_k^{(1)} + \varphi_k^{(2)} + \cdots \label{e7} \end{align} $$
where $E_k^{(s)}$ and $\varphi_k^{(s)}$ are terms of order $\lambda^s$. Since perturbation theory assumes $\lambda \ll 1$, the contributions of these terms diminish with increasing $s$, as indicated by $\lambda^s$.
Substituting the expansions \eqref{e6} and \eqref{e7} into TISE \eqref{e5} for the $k$-th eigenenergy $E = E_k$ and eigenfunction $\varphi = \varphi_k$, we obtain:
$$ \begin{align} \label{e8} \left(\hat{H}_0+\hat{H}’\right)\left(\varphi_k^{(0)} + \varphi_k^{(1)} +\cdots\right) = \left(E_k^{(0)} + E_k^{(1)} + \cdots\right) \left(\varphi_k^{(0)} + \varphi_k^{(1)} +\cdots\right) \end{align} $$
By noting that $O(\lambda^s)O(\lambda^r) = O(\lambda^{s+r})$, we match terms of the same order in $\lambda$ on both sides to derive identities:
$$ \begin{align} \lambda^{0}&: &&\hat{H}_0\varphi_k^{(0)} = E_k^{(0)}\varphi_k^{(0)} \label{e9_0} \\ \lambda^{1}&: &&\hat{H}_0\varphi_k^{(1)}+\hat{H}’\varphi_k^{(0)} = E_k^{(0)}\varphi_k^{(1)} + E_k^{(1)}\varphi_k^{(0)} \label{e9_1} \\ \lambda^{2}&: &&\hat{H}_0\varphi_k^{(2)}+\hat{H}’\varphi_k^{(1)} = E_k^{(0)}\varphi_k^{(2)} + E_k^{(1)}\varphi_k^{(1)} + E_k^{(2)}\varphi_k^{(0)} \label{e9_2} \\ & &&\vdots \nonumber \\ \lambda^{s}&: &&\hat{H}_0\varphi_k^{(s)}+\hat{H}’\varphi_k^{(s-1)} = \sum_{j=0}^s E_k^{(s-j)}\varphi_k^{(j)} \label{e9_s}\\ & &&\vdots \nonumber \end{align} $$
For a general consideration, we examine \eqref{e9_s}. Multiplying both sides of \eqref{e9_s} from the left by $\psi_l^{(0)*}$ and integrating over the entire space, we derive:
$$ \begin{gather} \langle\varphi_l^{(0)}|\hat{H}_0|\varphi_k^{(s)}\rangle +\langle\varphi_l^{(0)}|\hat{H}’|\varphi_k^{(s-1)}\rangle =\sum_{j=0}^s E_k^{(s-j)}\langle\varphi_l^{(0)}|\varphi_k^{(j)}\rangle \label{e10} \end{gather} $$
and hence:
$$ \begin{align} &\left(E_l^{(0)}-E_k^{(0)}\right)\langle\varphi_l^{(0)}|\varphi_k^{(s)}\rangle -E_k^{(s)}\delta_{l,k} \nonumber\\ &\hspace{10em}=-\langle\varphi_l^{(0)}|\hat{H}’|\varphi_k^{(s-1)}\rangle +\sum_{j=1}^{s-1} E_k^{(s-j)}\langle\varphi_l^{(0)}|\varphi_k^{(j)}\rangle \label{e11} \end{align} $$
For the case $l = k$, the first term on the left-hand side of \eqref{e11} vanishes, requiring a separate treatment for $l = k$ and $l \ne k$.
2.1.1. Eigenenergy Corrections #
Consider the case $l = k$ in \eqref{e11}. Here, among the variables in the equation, only $E_k^{(s)}$ has the highest order $s$, while the others are corrections of orders lower than $s$. This allows us to solve iteratively from $s = 1$ to determine $E_k^{(s)}$ for any $s$.
Specifically, substituting $l = k$ and solving for $E_k^{(s)}$, we obtain:
$$ \begin{align} E_k^{(s)}=\langle\varphi_k^{(0)}|\hat{H}’|\varphi_k^{(s-1)}\rangle -\sum_{j=1}^{s-1} E_k^{(s-j)}\langle\varphi_k^{(0)}|\varphi_k^{(j)}\rangle \label{e12} \end{align} $$
By choosing the wavefunction to satisfy \eqref{e13a}, the second term in \eqref{e12} vanishes, and we can express:
$$ \begin{align} E_k^{(s)}=\langle\varphi_k^{(0)}|\hat{H}’|\varphi_k^{(s-1)}\rangle \label{e12a} \end{align} $$
2.1.2. Eigenfunction Corrections #
Consider the case $l \ne k$ in \eqref{e11}.
In this case, only $\varphi_k^{(s)}$ has the highest order $s$, while the other variables consist of correction terms of orders lower than $s$. Similar to the eigenenergy corrections, we can solve iteratively from $s = 1$.
Rearranging terms, we obtain:
$$ \begin{align} \langle\varphi_l^{(0)}|\varphi_k^{(s)}\rangle =\frac{1}{E_l^{(0)}-E_k^{(0)}}\biggl[-\langle\varphi_l^{(0)}|\hat{H}’|\varphi_k^{(s-1)}\rangle +\sum_{j=1}^{s-1} E_k^{(s-j)}\langle\varphi_l^{(0)}|\varphi_k^{(j)}\rangle\biggr] \label{e13} \end{align} $$
To derive a more explicit result, we expand $\varphi_k^{(s)}$.
Assuming $\varphi_k^{(0)}$ forms an orthonormal basis, we expand $\varphi_k^{(s)}$ as:
$$ \begin{align} \label{e13a} \varphi_k^{(s)}=\sum_{m\ne k} a_{mk}^{(s)}\varphi_m^{(0)},~~(s\geq 1) \end{align} $$
where the goal is to determine the unknown coefficients $a_{mk}^{(s)}$ for $s \geq 1$ and $m \ne k$. The term $m = k$ is omitted in this expansion, which is justified and explained later.
For the state $k$, we have $a_{mk}^{(0)} = \delta_{mk}$.
In terms of orders, $\varphi_m^{(0)}$ is $O(\lambda^0)$, $\varphi_k^{(s)}$ is $O(\lambda^s)$, and thus $a_{mk}^{(s)}$ is also $O(\lambda^s)$.
Substituting \eqref{e13a} into \eqref{e13}, and noting that $a_{lk}^{(s)} = \langle \varphi_l^{(0)} | \varphi_k^{(s)} \rangle$, we obtain:
$$ \begin{align} a_{lk}^{(s)} =\frac{1}{E_l^{(0)}-E_k^{(0)}}\biggl[\sum_{j=1}^{s-1} E_k^{(s-j)}a_{lk}^{(j)} - \sum_{m\ne k} H’_{lm} a_{mk}^{(s-1)} \biggr],~~(s\geq 1, l\ne k) \label{e13d} \end{align} $$
where we define:
$$ \begin{align} \label{e13e} H’_{lm}\equiv \langle\varphi_l^{(0)}|\hat{H}’|\varphi_m^{(0)}\rangle \end{align} $$
Using \eqref{e13a}, the eigenenergy can be rewritten as:
$$ \begin{align} E_k^{(s)} &= \langle\varphi_k^{(0)}|\hat{H}’|\varphi_k^{(s-1)}\rangle -\sum_{j=1}^{s-1} E_k^{(s-j)}\langle\varphi_k^{(0)}|\varphi_k^{(j)}\rangle \label{e13f} \\ &= \sum_{m\ne k} H’_{km}a_{mk}^{(s-1)}\label{e13g} \end{align} $$
During the transformation to \eqref{e13g}, the second term in \eqref{e13f} is eliminated using \eqref{e13a}.
2.1.3. Reason for Excluding the Term $m = k$ in the Expansion of Perturbative Coefficients #
Here, we explain why the term $m = k$ can be excluded in the expansion \eqref{e13a} of the perturbative terms.
If we consider each perturbative term being expanded in the basis functions ${\varphi_m^{(0)}}$, the wavefunction can be rewritten as:
$$ \begin{align} \varphi_k & = \varphi_k^{(0)} + \varphi_k^{(1)} + \cdots \nonumber\\ & = \varphi_k^{(0)} + \sum_{m} a_{mk}^{(1)}\varphi_m^{(0)} + \sum_{m} a_{mk}^{(2)}\varphi_m^{(0)} + \cdots \nonumber\\ & = \sum_{m} d_{mk}\varphi_m^{(0)} \nonumber\\ & = d_{kk}\left(\varphi_k^{(0)} + \sum_{m\ne k} c_{mk}\varphi_m^{(0)}\right) \label{e14b} \end{align} $$
where we set:
$$ \begin{gather} d_{mk}\equiv \sum_{s=1} (\delta_{mk} + a_{mk}^{(s)}),~~~~c_{mk}\equiv\frac{d_{mk}}{d_{kk}} \end{gather} $$
Since any complex multiple of the wavefunction also satisfies the TISE, the factor $d_{kk}$ in \eqref{e14b} can be disregarded, leading to the solution:
$$ \begin{align} \label{e14c} \varphi_k = \varphi_k^{(0)} + \sum_{m\ne k} c_{mk}\varphi_m^{(0)} \end{align} $$
Thus, it is valid to always find the perturbative term $\varphi_k^{(s)}$ such that its component along $\varphi_k^{(0)}$ is zero.
However, the probability density of the wavefunction $\varphi_k$ found using \eqref{e14c} is not normalized to 1. Therefore, after computing the perturbative terms to the required order, it is necessary to renormalize the wavefunction to 1 as needed.
2.2. Summary #
We aim to solve the time-independent Schrödinger equation (TISE):
$$ \begin{align} \label{e20} (\hat{H}_0+\hat{H}’)\varphi = E \varphi \end{align} $$
to determine the $k$-th eigenenergy $E_k$ and eigenfunction $\varphi_k$, considering the case where the effect of $\hat{H}’$ is small compared to $\hat{H}_0$.
The degree of influence of $\hat{H}’$ on $\hat{H}_0$ is expressed in terms of the order parameter $\lambda \ll 1$. In terms of order, we have $\hat{H}_0 = O(\lambda^0)$ and $\hat{H}’ = O(\lambda^1)$.
Given that the TISE for $\hat{H}_0$:
$$ \begin{align} \label{e21} \hat{H}_0\varphi^{(0)} = E^{(0)} \varphi^{(0)} \end{align} $$
yields known eigenenergies $E_k^{(0)}$ and eigenfunctions $\varphi_k^{(0)}$, the solutions $E_k$ and $\varphi_k$ of TISE \eqref{e20} can be expanded as a power series in $\lambda$:
$$ \begin{align} \varphi_k &= \varphi_k^{(0)} + \varphi_k^{(1)} + \varphi_k^{(2)} + \cdots \label{e22}\\ E_k &= E_k^{(0)} + E_k^{(1)} + E_k^{(2)} + \cdots \label{e23} \end{align} $$
where $E_k^{(s)} = O(\lambda^s)$ and $\varphi_k^{(s)} = O(\lambda^s)$. The solutions are then obtained as:
$$ \begin{align} \varphi_k^{(s)}&=\sum_{m\ne k} a_{mk}^{(s)}\varphi_m^{(0)} \label{e24}\\ E_k^{(s)}&=\langle\varphi_k^{(0)}|\hat{H}’|\varphi_k^{(s-1)}\rangle = \sum_{m\ne k} H’_{km}a_{mk}^{(s-1)} \label{e25} \end{align} $$
Here, we define:
$$ \begin{align} \label{e27} a_{mk}^{(s)} =\frac{1}{E_m^{(0)}-E_k^{(0)}}\biggl[\sum_{j=1}^{s-1} E_k^{(s-j)}a_{mk}^{(j)} - \sum_{n} H’_{mn} a_{nk}^{(s-1)} \biggr],~~(s\geq 1, m\ne k) \end{align} $$
$$ \begin{align} \label{e28} H’_{lm} = \langle\varphi_l^{(0)}|\hat{H}’|\varphi_m^{(0)}\rangle \end{align} $$
Below, we explicitly write out several perturbative terms.
2.2.1. First-Order Perturbation Term #
$$ \begin{align} E_k^{(1)}&= H’_{kk} \label{e30}\\ \varphi_k^{(1)}&=\sum_{m\ne k} a_{mk}^{(1)}\varphi_m^{(0)} \label{e31}\\ &a_{mk}^{(1)}=\frac{- H’_{mk}}{E_m^{(0)}-E_k^{(0)}},\hspace{1em}(m\ne k) \label{e32} \end{align} $$
Here, $a_{kk}^{(1)} = 0$.
2.2.2. Second-Order Perturbation Term #
$$ \begin{align} E_k^{(2)}&=\langle\varphi_k^{(0)}|\hat{H}’|\varphi_k^{(1)}\rangle = \sum_{m\ne k} \frac{- |H’_{mk}|^2}{E_m^{(0)}-E_k^{(0)}} \label{e33}\\ \varphi_k^{(2)}&=\sum_{m\ne k} a_{mk}^{(2)}\varphi_m^{(0)} \label{e34} \end{align} $$
$$ \begin{align} a_{mk}^{(2)}&=\frac{1}{E_m^{(0)}-E_k^{(0)}}\biggl[E_k^{(1)}a_{mk}^{(1)} - \sum_{n\ne k} H’_{mn} a_{nk}^{(1)} \biggr] \label{e35}\\ &=\frac{1}{E_m^{(0)}-E_k^{(0)}}\biggl[E_k^{(1)}\frac{- H’_{mk}}{E_m^{(0)}-E_k^{(0)}} + \sum_{n\ne k} \frac{ H’_{mn}H’_{nk}}{E_n^{(0)}-E_k^{(0)}} \biggr],\hspace{1em}(m\ne k) \label{e36} \end{align} $$
Here, $a_{kk}^{(1)} = a_{kk}^{(2)} = 0$ (for all $k$).
2.2.3. s-th Order Perturbation Term #
$$ \begin{align} E_k^{(s)}&=\langle\varphi_k^{(0)}|\hat{H}’|\varphi_k^{(s-1)}\rangle = \sum_{m\ne k} H’_{km}a_{mk}^{(s-1)} \label{e37}\\ \varphi_k^{(s)}&=\sum_{m\ne k} a_{mk}^{(s)}\varphi_m^{(0)} \label{e38} \end{align} $$
$$ \begin{align} \label{e39} a_{mk}^{(s)} =\frac{1}{E_m^{(0)}-E_k^{(0)}}\biggl[\sum_{j=1}^{s-1} E_k^{(s-j)}a_{mk}^{(j)} - \sum_{n\ne k} H’_{mn} a_{nk}^{(s-1)} \biggr],~~(s\geq 1, m\ne k) \end{align} $$
Here, $a_{kk}^{(s)} = 0$ (for $s \geq 1$ and all $k$).
3. Examples #
3.1. Harmonic Oscillator with Perturbation #
To verify the validity of perturbation theory, we examine a harmonic oscillator with an added perturbation.
3.1.1. Analytical Solution #
$$ \begin{align} \label{e40} (\hat{H}_0+\hat{H}’)\varphi = E \varphi \end{align} $$
Here, the unperturbed Hamiltonian is written as:
$$ \begin{align} \label{e41} \hat{H}_0=-\frac{1}{2}\frac{d^2}{dx^2}+\frac{1}{2}x^2 \end{align} $$
and the perturbation Hamiltonian is:
$$ \begin{align} \hat{H}’&=-ax+\frac{1}{2}a^2+b \label{e42}\\ &\equiv Ax+B \label{e43} \end{align} $$
where $A = -a$ and $B = \frac{1}{2}a^2 + b$ are introduced for computational convenience.
Without considering perturbations, the Schrödinger equation is transformed as:
$$ \begin{gather} \label{e44} \left(-\frac{1}{2}\frac{d^2}{dx^2}+\frac{1}{2}(x-a)^2\right)\varphi(x) = (E-b) \varphi(x) \end{gather} $$
This implies that $a$ shifts the eigenfunctions by $a$, and $b$ shifts the eigenenergies by $b$. The solution to the TISE can be expressed using the unperturbed eigenenergies $E^{(0)}$ and eigenfunctions $\varphi^{(0)}(x)$ as:
$$ \begin{align} \varphi(x)&=\varphi^{(0)}(x-a) \label{e45} \\ E&=E^{(0)}+b \label{e46}\\ &=E^{(0)}+B-\frac{A^2}{2} \end{align} $$
We will now verify whether the results derived using perturbation theory match \eqref{e45} and \eqref{e46}.
The eigenenergies and eigenstates of the TISE $\hat{H}_0\varphi^{(0)} = E^{(0)}\varphi^{(0)}$ are given explicitly as:
$$ \begin{gather} E^{(0)}_n=n+\frac{1}{2}\\ \varphi^{(0)}_n(x)=\frac{\pi^{-1/4}}{\sqrt{2^n n!}}H_n(x)e^{-x^2/2} \end{gather} $$
3.1.2. Results from Perturbation Theory #
We now calculate the results following perturbation theory. The first-order perturbation results, obtained using \eqref{e30}, \eqref{e31}, and \eqref{e32}, are calculated as follows.
The calculation of $H’_{lm}$ yields below.Details of the calculations are provided in the appendix, specifically in equation \eqref{x61}:
$$ \begin{align} H’_{lm} &= \langle\varphi_l^{(0)}|(Ax+B)|\varphi_m^{(0)}\rangle \nonumber\\ &= A\sqrt{\frac{m}{2}}\delta_{l,m-1} + B\delta_{l,m} + A\sqrt{\frac{m+1}{2}}\delta_{l,m+1} \label{e61} \end{align} $$
Additionally, $H’{l,m} = H’{m,l}$ holds.
From these results, the corrections to the eigenenergies and eigenfunctions derived via perturbation theory are given as follows:
-
First-Order Perturbation
-
Correction to Eigenenergy
$$ \begin{align} E_k^{(1)}= H’_{kk} = B \label{e62} \end{align} $$
-
Correction to Eigenfunction
$$ \begin{align} \varphi_k^{(1)}(x)&=\frac{A}{\sqrt{2}}\left[\sqrt{k}\varphi_{k-1}^{(0)}(x) - \sqrt{k+1} \varphi_{k+1}^{(0)}(x)\right] \label{e63} \\ &=-a \frac{d\varphi_k^{(0)}}{dx} \end{align} $$
*Note: Utilizes $A = -a$ and the recurrence relation for $\varphi_{k}^{(0)}(x)$ (Appendix \eqref{x80}).*
-
Coefficients Related to the Correction of Eigenfunctions
$$ \begin{align} a_{mk}^{(1)} = A\sqrt{\frac{k}{2}}\delta_{m,k-1} - A\sqrt{\frac{k+1}{2}}\delta_{m,k+1} \label{e64} \end{align} $$
*Note: The transformation steps for $a_{mk}^{(1)}$ are detailed in Appendix \eqref{x64}.*
-
Wavefunction Including First-Order Perturbation
Based on the above, the wavefunction up to first-order perturbation is expressed as:
$$ \begin{align} \varphi_k(x)&=\varphi_k^{(0)}(x)+\varphi_k^{(1)}(x) + O(\lambda^2)\\ &= \varphi_k^{(0)}(x) + \frac{A}{\sqrt{2}}\left[\sqrt{k}\varphi_{k-1}^{(0)}(x) - \sqrt{k+1} \varphi_{k+1}^{(0)}(x)\right] \\ &=\varphi_k^{(0)}(x)- a \frac{d\varphi_k^{(0)}}{dx} + O(\lambda^2) \label{e80} \\ \end{align} $$
This result matches the Taylor expansion of the analytical solution \eqref{e45}:
$$ \begin{align} \varphi^{(0)}(x-a) &= \varphi^{(0)}(x) - a\frac{d\varphi^{(0)}}{dx} + a^2\frac{1}{2!}\frac{d^2\varphi^{(0)}}{dx^2} + O(\lambda^3) \end{align} $$
up to the $O(\lambda^1)$ term.
If the probability density is normalized to $1$, the wavefunction becomes:
$$ \begin{align} \small \varphi_k(x)=\frac{1}{\sqrt{1+\frac{A^2}{2}(2k+1)}}\left(\varphi_k^{(0)}(x) + \frac{A}{\sqrt{2}}\left[\sqrt{k}\varphi_{k-1}^{(0)}(x) - \sqrt{k+1} \varphi_{k+1}^{(0)}(x)\right]\right) + O(\lambda^2) \end{align} $$
-
-
Second-Order Perturbation
-
Correction to Eigenenergy
$$ \begin{align} E_k^{(2)}&=\langle\varphi_k^{(0)}|\hat{H}’|\varphi_k^{(1)}\rangle =\frac{A}{\sqrt{2}}\left[\sqrt{k}H’_{k,k-1} -\sqrt{k+1} H’_{k,k+1}\right] \nonumber \\ &=-\frac{A^2}{2} \label{e65} \end{align} $$
-
Correction to Eigenfunction
$$ \begin{align} \label{e66} \varphi_k^{(2)}=\frac{A^2}{4}\biggl[\sqrt{k(k-1)}\varphi_{k-2}(x) + \sqrt{(k+1)(k+2)}\varphi_{k+2}(x) \biggr] \end{align} $$
-
Coefficients Related to the Correction of Eigenfunctions
$$ \begin{align} \label{e67} a_{mk}^{(2)} = \frac{1}{2}\frac{A^2}{2}\biggl[\sqrt{k(k-1)}\delta_{m,k-2} + \sqrt{(k+1)(k+2)}\delta_{m,k+2} \biggr] \end{align} $$
*Note: The transformation steps for $a_{mk}^{(2)}$ are detailed in Appendix \eqref{x67a} to \eqref{x67b}.*
-
Wavefunction Including Second-Order Perturbation
$$ \begin{align} \varphi_k(x)&=\varphi_k^{(0)}(x)+\varphi_k^{(1)}(x)+\varphi_k^{(2)}(x) + O(\lambda^3)\\ &=\left[1+\frac{a^2}{2}\left(k+\frac{1}{2}\right)\right]\varphi_k^{(0)} - a\frac{d\varphi_k^{(0)}}{dx} + \frac{a^2}{2}\frac{d^2\varphi_k^{(0)}}{dx^2} + O(\lambda^3) \\ &\approx \varphi_k^{(0)} - a\frac{d\varphi_k^{(0)}}{dx} + \frac{a^2}{2}\frac{d^2\varphi_k^{(0)}}{dx^2} + O(\lambda^3) \label{e90}\\ \end{align} $$
The result in \eqref{e90} matches the Taylor expansion of $\varphi_k^{(0)}(x-a)$ up to the $O(\lambda^2)$ term.
Regarding the coefficients of $\varphi_k^{(0)}$ in \eqref{e90}, the first term is $O(\lambda^0)$, and the second term is $O(\lambda^2)$, which is considered negligible3.
If the probability density is normalized to $1$, the wavefunction becomes:
$$ \begin{align} \varphi_k(x)&=\frac{1}{\sqrt{1+\frac{A^2}{2}(2k+1)+\frac{A^4}{8}(k^2+k+1)}} \nonumber\\ &\times \biggl(\varphi_k^{(0)}(x) + \frac{A}{\sqrt{2}}\left[\sqrt{k}\varphi_{k-1}^{(0)}(x) - \sqrt{k+1} \varphi_{k+1}^{(0)}(x)\right] \nonumber \\ &+ \frac{A^2}{4}\biggl[\sqrt{k(k-1)}\varphi_{k-2}(x) + \sqrt{(k+1)(k+2)}\varphi_{k+2}(x)\biggr]\biggr) + O(\lambda^3) \end{align} $$
-
-
Third-Order Perturbation
-
Correction to Eigenenergy
$$ \begin{align} E_k^{(3)}&=\langle\varphi_k^{(0)}|\hat{H}’|\varphi_k^{(2)}\rangle = 0 \label{e68} \end{align} $$
From \eqref{e61}, $H’{l,m}$ has non-zero values only when $l = m$ or $l = m \pm 1$. However, as shown by the results for the second-order correction term \eqref{e66}, only $H’{l,l \pm 2}$ terms can arise, which implies that the value is zero.
-
Correction to Eigenfunction
$$ \begin{align} \varphi_k^{(3)}&=\frac{A^3}{12\sqrt{2}} \biggl[\sqrt{k(k-1)(k-2)}\varphi_{k-3}(x) + 3(k+1)\sqrt{k}\varphi_{k-1}(x) \nonumber\\ & \hspace{3em} - 3k\sqrt{k+1}\varphi_{k+1}(x) - \sqrt{(k+1)(k+2)(k+3)}\varphi_{k+3}(x)\biggr] \label{e68a} \end{align} $$
The third-order term of the Taylor expansion of the exact solution is given as:
$$ \begin{align} -\frac{a^3}{3!}\frac{d^3\varphi_k^{(3)}}{dx^3}&=\frac{A^3}{12\sqrt{2}} \biggl[\sqrt{k(k-1)(k-2)}\varphi_{k-3}(x) - 3k^{3/2}\varphi_{k-1}(x) \nonumber\\ &\hspace{1em} + 3(k+1)^{3/2}\varphi_{k+1}(x) - \sqrt{(k+1)(k+2)(k+3)}\varphi_{k+3}(x)\biggr] \label{e68b} \end{align} $$
The result \eqref{e68b} was expected to match \eqref{e68a}, but as with the second-order case, there is no agreement. There may be a calculation error, but none has been identified so far.
-
3.1.3. Visualization #
The wavefunctions truncated at appropriate orders in perturbation theory are visualized as follows:
- For $a=0.5$
- For $a=1.0$
- For $a=2.0$
When incorporating up to the second-order perturbation, the wavefunction for the smaller perturbation magnitude $a = 0.5$ closely matches the analytical solution. However, for the larger perturbation magnitude $a = 2.0$, the approximation significantly deviates from the analytical solution.
4. Appendix #
4.1. Transformation of Equation \eqref{e61} #
$$ \begin{align} \langle\varphi_l^{(0)}|x|\varphi_m^{(0)}\rangle &=\frac{\pi^{-1/2}}{\sqrt{2^{l+m} l!m!}} \int_{-\infty}^{\infty} x H_l(x) H_m(x)e^{-x^2} dx \nonumber\\ &=\frac{\pi^{-1/2}}{\sqrt{2^{l+m} l!m!}} \left[\frac{1}{2}\sqrt{\pi}2^m m!\delta_{l+1,m}+l\sqrt{\pi}2^m m!\delta_{l-1,m}\right] \nonumber\\ &=\sqrt{2^{m-l}}\sqrt{\frac{m!}{l!}}\left(\frac{1}{2}\delta_{l+1,m}+l\delta_{l-1,m}\right) \nonumber\\ &=\sqrt{\frac{m}{2}}\delta_{l,m-1}+\sqrt{\frac{m+1}{2}}\delta_{l,m+1} \label{x61} \end{align} $$
Utilizing the three-term recurrence relation of Hermite polynomials4:
$$ \begin{gather} x H_n(x) = \frac{1}{2}H_{n+1}(x) + n H_{n-1}(x),\hspace{1em}(n\geq 0)\label{x62} \\ H_{-1}(x) = 0 \end{gather} $$
and the orthogonality of Hermite polynomials5:
$$ \begin{gather} \int H_m(x)H_n(x)e^{-x^2} = \sqrt{\pi}2^n n! \delta_{m,n} \end{gather} $$
4.2. Transformation of Equation \eqref{e64} #
$$ \begin{align} a_{mk}^{(1)} &= \frac{- H’_{mk}}{E_m^{(0)}-E_k^{(0)}}, \hspace{1em} (m\ne k) \nonumber\\ &= \frac{-1}{m-k}\left[A\sqrt{\frac{m}{2}}\delta_{m-1,k} + B\delta_{m,k} + A\sqrt{\frac{m+1}{2}}\delta_{m+1,k}\right] \nonumber\\ &= A\sqrt{\frac{k}{2}}\delta_{m,k-1} - A\sqrt{\frac{k+1}{2}}\delta_{m,k+1} \label{x64} \end{align} $$
4.3. Transformation of Equation \eqref{e80} #
For the orthogonal polynomial $\varphi_k^{(0)}$, the following recurrence relation holds:
$$ \begin{align} \label{x80} \frac{d\varphi_k^{(0)}}{dx} = \sqrt{\frac{k}{2}}\varphi_{k-1}^{(0)}(x) - \sqrt{\frac{k+1}{2}}\varphi_{k+1}^{(0)}(x) \end{align} $$
4.4. Transformation of Equation \eqref{e67} #
To calculate $a_{mk}^{(2)}$, the following sum required for the calculation is evaluated:
$$ \begin{align} \sum_{n\ne k} H’_{mn} a_{nk}^{(1)} &= \sum_{n\ne k}H’_{m,n}\left(A\sqrt{\frac{k}{2}}\delta_{n,k-1}-A\sqrt{\frac{k+1}{2}}\delta_{n,k+1}\right) \nonumber\\ &=H’_{m,k-1}A\sqrt{\frac{k}{2}}-H’_{m,k+1}A\frac{k+1}{2} \nonumber\\ &=\frac{A^2}{2}\sqrt{k(k-1)}\delta_{m,k-2} + AB\sqrt{\frac{k}{2}}\delta_{m,k-1} - \frac{A^2}{2}\delta_{m,k} \nonumber\\ &\hspace{5em} - AB\sqrt{\frac{k+1}{2}}\delta_{m,k+1} - \frac{A^2}{2}\sqrt{(k+1)(k+2)}\delta_{m,k+2} \label{x67a} \end{align} $$
This results in:
$$ \begin{align} a_{mk}^{(2)} &= \frac{1}{E_m^{(0)}-E_k^{(0)}}\biggl[E_k^{(1)}a_{mk}^{(1)} - \sum_{n\ne k} H’_{mn} a_{nk}^{(1)} \biggr] \nonumber\\ &= \frac{1}{m-k}\biggl[B a_{mk}^{(1)} - \sum_{n\ne k} H’_{mn} a_{nk}^{(1)} \biggr] \nonumber\\ &= \frac{-1}{m-k}\frac{A^2}{2}\biggl[\sqrt{k(k-1)}\delta_{m,k-2}+\delta_{m,k}+\sqrt{(k+1)(k+2)}\delta_{m,k+2}\biggr] \end{align} $$
Specifically, since only the values for $m \ne k$ are needed:
$$ \begin{align} a_{mk}^{(2)} &= \frac{-1}{m-k}\frac{A^2}{2}\biggl[\sqrt{k(k-1)}\delta_{m,k-2}+\sqrt{(k+1)(k+2)}\delta_{m,k+2}\biggr],\hspace{1em}(m\ne k) \nonumber\\ &= \frac{1}{2}\frac{A^2}{2}\biggl[\sqrt{k(k-1)}\delta_{m,k-2} + \sqrt{(k+1)(k+2)}\delta_{m,k+2} \biggr] \label{x67b} \end{align} $$
5. References #
-
B. H. Bransden and C. J. Joachain, Physics of Atoms and Molecules (2nd Edition), Addison-Wesley(2003). ↩︎
-
It might seem impossible to judge whether the solution will be significantly distorted before solving the problem, and indeed, this is true. In practice, if the perturbative results converge, the approximation may be valid; otherwise, it is not applicable. ↩︎
-
While exploring how to handle the second-order coefficients of $\varphi_k^{(0)}$, the only plausible approach seems to be ignoring them for now. The second-order correction term \eqref{e66} leads to the appearance of $\varphi_k^{(0)}$ components when rewritten in the form of a second derivative. Since this arises from a separate cause than the normalization technique setting the coefficient of $\varphi_k^{(0)}$ to 1, arbitrarily discarding it may not be justified. ↩︎
-
Abramowitz and Stegun, Handbook of Mathematical Functions. 22.7. Recurrence Relations ↩︎
-
Abramowitz and Stegun, Handbook of Mathematical Functions. 22.2. Orthogonality Relations ↩︎