Skip to content
New issue

Have a question about this project? Sign up for a free GitHub account to open an issue and contact its maintainers and the community.

By clicking “Sign up for GitHub”, you agree to our terms of service and privacy statement. We’ll occasionally send you account related emails.

Already on GitHub? Sign in to your account

bridge derivation: Quantum Harmonic Oscillator and classical Simple Harmonic Oscillator are related by the Correspondence Principle #23

Open
bhpayne opened this issue Jan 26, 2025 · 7 comments
Assignees

Comments

@bhpayne
Copy link
Member

bhpayne commented Jan 26, 2025

as per https://physicsderivationgraph.blogspot.com/2024/03/dichotomy-of-assumptions.html

@bhpayne
Copy link
Member Author

bhpayne commented Jan 26, 2025

prompt to Gemini 2.0 Flash Thinking Experimental 01-21, https://aistudio.google.com/prompts/

The analysis of the harmonic oscillator bridges the assumptions of classical mechanics and quantum mechanics.

For the Classical Harmonic Oscillator, the assumptions of Classical Mechanics include

  • Particles have definite trajectories and properties (position, momentum) at all times.
  • Energy is a continuous quantity.
  • Motion is governed by Newton's laws of motion.

For the Quantum Harmonic Oscillator, the assumptions of Quantum Mechanics include

  • Particles are described by wave functions, and their properties are probabilistic.
  • Energy is often quantized, meaning it can only take on discrete values.
  • Motion is governed by the Schrödinger equation.

Provide a mathematical derivation with explicit steps that shows how the Classical Harmonic Oscillator and Quantum Harmonic Oscillator are related by the Correspondence Principle. Write the output as a single Latex file.

thoughts

Here's a breakdown of the thinking process to construct the LaTeX document explaining the Correspondence Principle for the harmonic oscillator:

  1. Understand the Request: The request asks for a LaTeX document that derives how the Classical Harmonic Oscillator (CHO) and Quantum Harmonic Oscillator (QHO) relate via the Correspondence Principle. This implies showing how, in a certain limit (large quantum numbers), the QHO behavior approaches the CHO behavior. The request also specifically asks for explicit mathematical steps.

  2. Outline the Document Structure: A logical structure for this document would be:

    • Title: Descriptive title.
    • Introduction: Briefly explain the Correspondence Principle and its significance in the context of CHO and QHO.
    • Classical Harmonic Oscillator (CHO) Review: Summarize the key properties of the CHO, including energy, frequency, and probability density. This sets the classical benchmark.
    • Quantum Harmonic Oscillator (QHO) Review: Summarize the key properties of the QHO, including energy levels, wave functions (qualitatively), and probability density. Highlight the quantization and differences from the classical case.
    • Correspondence Principle Application: This is the core section. Demonstrate the principle by focusing on probability density and energy.
      • Probability Density Comparison: Show how the QHO probability density approaches the classical probability density in the high n limit. This is the most visually and conceptually compelling demonstration. This involves:
        • Describing the classical probability density (inverse of velocity, peaked at turning points).
        • Describing the QHO probability density (oscillatory, distributed within classical limits).
        • Arguing (or showing qualitatively) that for large n, the QHO probability density averages to something resembling the classical density within the classical allowed region.
      • Energy Level Spacing: Show how the energy level spacing becomes relatively smaller compared to the energy itself as n increases. This demonstrates the "continuous" energy spectrum approximation in the classical limit.
    • Conclusion: Summarize the findings and reiterate the Correspondence Principle's role in bridging classical and quantum mechanics.
    • References (Optional): Include if referencing specific books or articles.
  3. Mathematical Details and Derivations:

    • CHO Review: Standard formulas for energy (E = 1/2 * m * ω^2 * A^2), frequency (ω = sqrt(k/m)), position as a function of time (x(t) = A cos(ωt + φ)), velocity (v(t) = -Aω sin(ωt + φ)). Probability density needs to be derived: P_classical(x) ∝ 1/|v(x)|. Need to express v(x) in terms of energy and position to get P_classical(x) ∝ 1/sqrt(A^2 - x^2).

    • QHO Review: Energy eigenvalues (E_n = (n + 1/2)ħω), qualitative description of wave functions (Hermite polynomials, increasing oscillations with n), and probability density (|ψ_n(x)|^2). No need to write out full wave function expressions unless specifically requested or necessary for the argument. Focus on the qualitative behavior of the probability densities for low and high n.

    • Correspondence Principle Application (Probability Density):

      • Classical probability density is high at turning points and low in the middle. Graphing or sketching this is very helpful.
      • Quantum probability density for low n is very different. For high n, it becomes more oscillatory and spread out. The key is to argue that averaged over the oscillations, it starts to resemble the classical distribution within the classical allowed region. A rigorous mathematical proof of convergence is complex and likely beyond the scope of a typical request about the correspondence principle at this level. A qualitative argument and visual comparison (if possible in the document) are usually sufficient.
    • Correspondence Principle Application (Energy Levels):

      • Energy levels are quantized: E_n = (n + 1/2)ħω.
      • Energy level spacing is ΔE = E_{n+1} - E_n = ħω, which is constant and independent of n.
      • However, the relative spacing, ΔE / E_n = ħω / ((n + 1/2)ħω) = 1/(n + 1/2), goes to zero as n approaches infinity. This means for very high energies (large n), the energy levels become very close together relative to the total energy, effectively approximating a continuous energy spectrum, which is classical.
  4. LaTeX Implementation: Start writing the LaTeX document section by section based on the outline. Use appropriate LaTeX commands for:

    • Document class (\documentclass{article})
    • Packages (\usepackage{amsmath, amssymb, amsfonts})
    • Title, author, date (\title, \author, \date, \maketitle)
    • Sections and subsections (\section, \subsection)
    • Equations and mathematical notation (\eqref, \begin{equation}, \frac, \hbar, \omega, \int, \infty, \approx, \propto, \text, \mathbf, etc.)
    • Emphasis (\textit, \textbf)
    • Lists (\begin{itemize}, \item)
    • Figures (if adding graphs, which would greatly enhance the probability density comparison, although not explicitly requested for derivation per se). For this request, figures are probably not necessary but would be good for a more complete explanation.
  5. Review and Refine: After drafting the LaTeX document, review it for:

    • Logical flow and clarity: Is the argument easy to follow? Are the explanations clear and concise?
    • Mathematical accuracy: Are the equations correct? Are the derivations sound (even if qualitative for probability density)?
    • Completeness: Does it address all aspects of the request? Does it adequately explain the Correspondence Principle in the context of the harmonic oscillator?
    • LaTeX formatting: Is the LaTeX code correct? Does it compile without errors? Is the formatting visually appealing and readable? Are equations numbered and referenced correctly?

By following these steps, especially the structured outline and focusing on the key points of comparison (probability density and energy level spacing), we can construct a comprehensive and accurate LaTeX document that fulfills the request. The key is to balance mathematical rigor with clarity and focus on the conceptual bridge provided by the Correspondence Principle. For the probability density, a qualitative explanation of convergence in the high n limit is usually sufficient rather than a full rigorous mathematical proof.

response

\documentclass{article}
\usepackage{amsmath}
\usepackage{amssymb}
\usepackage{amsfonts}

\title{The Correspondence Principle and the Harmonic Oscillator}
\author{}
\date{}

\begin{document}
\maketitle

\section{Introduction}

The harmonic oscillator is a fundamental system in physics, serving as a model for a wide range of phenomena from molecular vibrations to oscillations in electromagnetic fields.  It is particularly insightful in the context of the Correspondence Principle, which states that the predictions of quantum mechanics must agree with classical mechanics in the limit of large quantum numbers or high energies. This principle provides a crucial bridge between the seemingly disparate worlds of classical and quantum physics.  In this document, we will mathematically explore how the quantum harmonic oscillator (QHO) approaches the classical harmonic oscillator (CHO) in the classical limit, demonstrating the Correspondence Principle.

\section{Classical Harmonic Oscillator (CHO) Review}

Consider a particle of mass $m$ attached to a spring with spring constant $k$. The potential energy of this system is given by $V(x) = \frac{1}{2}kx^2$.

\subsection{Equations of Motion and Energy}
From Newton's second law, $F = ma = m\ddot{x} = -kx$. This leads to the differential equation for simple harmonic motion:
\begin{equation}
\ddot{x} + \omega^2 x = 0,
\end{equation}
where $\omega = \sqrt{\frac{k}{m}}$ is the angular frequency of oscillation. The general solution is:
\begin{equation}
x(t) = A \cos(\omega t - \phi),
\end{equation}
where $A$ is the amplitude and $\phi$ is the phase constant, determined by initial conditions.

The velocity is given by:
\begin{equation}
v(t) = \dot{x}(t) = -A\omega \sin(\omega t - \phi).
\end{equation}

The total energy of the classical harmonic oscillator is conserved and is the sum of kinetic and potential energies:
\begin{equation}
E_{classical} = T + V = \frac{1}{2}mv^2 + \frac{1}{2}kx^2 = \frac{1}{2}m(A\omega \sin(\omega t - \phi))^2 + \frac{1}{2}k(A \cos(\omega t - \phi))^2.
\end{equation}
Using $k = m\omega^2$ and $\sin^2(\theta) + \cos^2(\theta) = 1$, we get:
\begin{equation}
E_{classical} = \frac{1}{2}mA^2\omega^2 (\sin^2(\omega t - \phi) + \cos^2(\omega t - \phi)) = \frac{1}{2}mA^2\omega^2 = \frac{1}{2}kA^2.
\end{equation}
Classically, energy $E_{classical}$ can take any continuous non-negative value.

\subsection{Classical Probability Density}
The probability of finding the particle in a small interval $dx$ at position $x$ is proportional to the time spent in that interval. Since the particle moves slower near the turning points ($x = \pm A$) and faster near the equilibrium position ($x = 0$), the probability density is inversely proportional to the speed of the particle:
\begin{equation}
P_{classical}(x) \propto \frac{1}{|v(x)|}.
\end{equation}
From energy conservation, $E_{classical} = \frac{1}{2}mv^2 + \frac{1}{2}kx^2$, we can express the speed as:
\begin{equation}
v(x) = \pm \sqrt{\frac{2}{m}(E_{classical} - \frac{1}{2}kx^2)} = \pm \omega \sqrt{A^2 - x^2}.
\end{equation}
Thus, the classical probability density is:
\begin{equation}
P_{classical}(x) = \frac{C}{|v(x)|} = \frac{C}{\omega \sqrt{A^2 - x^2}} = \frac{C'}{\sqrt{A^2 - x^2}}, \quad \text{for } |x| \leq A,
\end{equation}
and $P_{classical}(x) = 0$ for $|x| > A$.  $C'$ is a normalization constant.

\section{Quantum Harmonic Oscillator (QHO) Review}

In quantum mechanics, the harmonic oscillator is described by the time-independent Schrödinger equation:
\begin{equation}
-\frac{\hbar^2}{2m} \frac{d^2\psi_n(x)}{dx^2} + \frac{1}{2}kx^2 \psi_n(x) = E_n \psi_n(x).
\end{equation}
The solutions to this equation are the energy eigenstates $\psi_n(x)$ with corresponding energy eigenvalues $E_n$.

\subsection{Energy Eigenvalues}
The energy eigenvalues for the quantum harmonic oscillator are quantized and given by:
\begin{equation}
E_n = \left(n + \frac{1}{2}\right)\hbar\omega, \quad n = 0, 1, 2, \ldots
\end{equation}
where $n$ is the quantum number and $\hbar$ is the reduced Planck constant.  Energy levels are discrete and equally spaced with spacing $\hbar\omega$.

\subsection{Energy Eigenstates and Probability Density}
The energy eigenstates $\psi_n(x)$ are given by:
\begin{equation}
\psi_n(x) = \frac{1}{\sqrt{2^n n!}} \left(\frac{m\omega}{\pi\hbar}\right)^{1/4} e^{-\frac{m\omega x^2}{2\hbar}} H_n\left(\sqrt{\frac{m\omega}{\hbar}} x\right),
\end{equation}
where $H_n(u)$ are the Hermite polynomials. The probability density of finding the particle at position $x$ in the $n$-th energy state is given by $|\psi_n(x)|^2$.

For low quantum numbers (small $n$), the probability density $|\psi_n(x)|^2$ is significantly different from the classical probability density. For instance, for $n=0$, $|\psi_0(x)|^2$ is a Gaussian centered at $x=0$, with the highest probability at the equilibrium position, unlike the classical case where the probability is lowest at the equilibrium.  Furthermore, there is a non-zero probability of finding the particle outside the classical turning points, which is forbidden in classical mechanics.

\section{Correspondence Principle and the Harmonic Oscillator}

The Correspondence Principle states that in the limit of large quantum numbers ($n \rightarrow \infty$), quantum mechanics should reproduce classical mechanics. Let's examine how this principle applies to the harmonic oscillator.

\subsection{Energy Levels}
The energy levels for the QHO are $E_n = (n + \frac{1}{2})\hbar\omega$. For large $n$, we can approximate $E_n \approx n\hbar\omega$. The energy spacing between adjacent levels is:
\begin{equation}
\Delta E = E_{n+1} - E_n = \left((n+1) + \frac{1}{2}\right)\hbar\omega - \left(n + \frac{1}{2}\right)\hbar\omega = \hbar\omega.
\end{equation}
This spacing $\hbar\omega$ is constant and independent of $n$. However, consider the \textit{relative} spacing compared to the energy itself:
\begin{equation}
\frac{\Delta E}{E_n} = \frac{\hbar\omega}{(n + \frac{1}{2})\hbar\omega} = \frac{1}{n + \frac{1}{2}} \approx \frac{1}{n} \quad \text{for large } n.
\end{equation}
As $n \rightarrow \infty$, the relative spacing $\frac{\Delta E}{E_n} \rightarrow 0$. This means that for very high energy levels, the energy spectrum becomes effectively continuous, resembling the classical case where energy can take continuous values.  In the classical limit of high energy, the quantization of energy becomes less apparent.

\subsection{Probability Density}
Now let's consider the probability density. For large $n$, the wave function $\psi_n(x)$ becomes highly oscillatory, and the Hermite polynomials $H_n(u)$ also exhibit oscillatory behavior within the classical allowed region and decay exponentially outside.  While a rigorous mathematical proof is complex, we can argue qualitatively and consider some properties:

1. \textbf{Classical Turning Points:} The classical turning points are given by $V(x) = E_{classical}$, so $\frac{1}{2}kx^2 = E_{classical}$. For the $n$-th quantum state, we set $E_{classical} = E_n = (n + \frac{1}{2})\hbar\omega$. Thus, the classical turning points are $x = \pm A_n$, where $\frac{1}{2}kA_n^2 = E_n$, so $A_n = \sqrt{\frac{2E_n}{k}} = \sqrt{\frac{2(n + \frac{1}{2})\hbar\omega}{m\omega^2}} = \sqrt{\frac{(2n+1)\hbar}{m\omega}}$.

2. \textbf{Oscillatory Nature of $\psi_n(x)$ for large $n$:} For large $n$, the Hermite polynomials $H_n(u)$ behave approximately as sinusoidal functions within the classically allowed region.  The square of the wave function, $|\psi_n(x)|^2$, will thus exhibit rapid oscillations within $|x| \leq A_n$.

3. \textbf{Averaging over Oscillations:} If we average the quantum probability density $|\psi_n(x)|^2$ over a spatial interval that is large compared to the wavelength of the oscillations but small compared to the overall extent of the wave function, we find that the averaged quantum probability density approaches the classical probability density.  Intuitively, the rapid oscillations in $|\psi_n(x)|^2$ for large $n$ average out to a smoother distribution that resembles the classical distribution, which is peaked at the turning points and minimum at the equilibrium position.

To illustrate this more explicitly, consider the WKB approximation, which is valid for large quantum numbers. In the WKB approximation, the wave function in the classically allowed region oscillates as:
\begin{equation}
\psi_{WKB}(x) \approx \frac{C}{\sqrt{p(x)}} \cos\left(\frac{1}{\hbar} \int^x p(x') dx' - \frac{\pi}{4}\right),
\end{equation}
where $p(x) = \sqrt{2m(E - V(x))} = \sqrt{2m(E - \frac{1}{2}kx^2)}$ is the classical momentum.  The probability density is then approximately:
\begin{equation}
|\psi_{WKB}(x)|^2 \approx \frac{C^2}{p(x)} \cos^2\left(\frac{1}{\hbar} \int^x p(x') dx' - \frac{\pi}{4}\right).
\end{equation}
Averaging the rapidly oscillating $\cos^2$ term over a cycle gives $\langle \cos^2 \rangle = \frac{1}{2}$.  Thus, the average probability density becomes:
\begin{equation}
\langle |\psi_{WKB}(x)|^2 \rangle \approx \frac{C^2}{2p(x)} = \frac{C'^2}{\sqrt{E - \frac{1}{2}kx^2}} = \frac{C''}{\sqrt{A_n^2 - x^2}},
\end{equation}
where we used $E = E_n = \frac{1}{2}kA_n^2$. This form is proportional to the classical probability density $P_{classical}(x) = \frac{C'}{\sqrt{A^2 - x^2}}$.  Hence, in the limit of large quantum numbers, the average quantum probability density approaches the classical probability density within the classically allowed region.

Furthermore, the quantum probability of finding the particle outside the classical turning points (tunneling) becomes relatively less significant for large $n$. The exponentially decaying tails of $\psi_n(x)$ beyond $x = \pm A_n$ become less prominent compared to the oscillatory behavior within the classical region as $n$ increases.

\section{Conclusion}

We have shown how the quantum harmonic oscillator approaches the classical harmonic oscillator in the limit of large quantum numbers, in accordance with the Correspondence Principle.

\begin{itemize}
    \item \textbf{Energy Levels:} The energy levels of the QHO are quantized, but for large quantum numbers $n$, the relative spacing between energy levels becomes vanishingly small compared to the energy itself.  This makes the energy spectrum effectively continuous, as in classical mechanics.
    \item \textbf{Probability Density:} While the quantum probability density for low $n$ is significantly different from the classical probability density, for large $n$, the average quantum probability density (obtained by averaging over oscillations or using approximations like WKB) approaches the classical probability density within the classically allowed region.  The oscillatory nature of the quantum wave function for large $n$ and its localization within the classical boundaries are key to this correspondence.
\end{itemize}

The harmonic oscillator provides a clear and important example of the Correspondence Principle, demonstrating how quantum mechanics smoothly transitions to classical mechanics in the appropriate limit. This principle is not only a historical cornerstone in the development of quantum theory but also a valuable tool for understanding the relationship between quantum and classical descriptions of physical systems.

\end{document}

@bhpayne
Copy link
Member Author

bhpayne commented Jan 26, 2025

prompt

For the following Latex file, write out the file and insert a label for each equation. The label that is added should be descriptive of the equation and contain no spaces, only underscores.

Don't provide any commentary. Just write the revised Latex file content as your answer.

\documentclass{article}
\usepackage{amsmath}

% margins of 1 inch:
\setlength{\topmargin}{-.5in}
\setlength{\textheight}{9in}
\setlength{\oddsidemargin}{0in}
\setlength{\textwidth}{6.5in}

\usepackage[pdftex]{hyperref} % hyperlink equation and bibliographic citations

\title{The Correspondence Principle and the Harmonic Oscillator}
\author{Ben Payne, with Gemini 2.0}
\date{\today}

\begin{document}
\begin{abstract}
    From Gemini 2.0 Flash
\end{abstract}
\maketitle

\section{Introduction}

The harmonic oscillator is a fundamental system in physics, serving as a model for a wide range of phenomena from molecular vibrations to oscillations in electromagnetic fields.  It is particularly insightful in the context of the Correspondence Principle, which states that the predictions of quantum mechanics must agree with classical mechanics in the limit of large quantum numbers or high energies. This principle provides a crucial bridge between the seemingly disparate worlds of classical and quantum physics.  In this document, we will mathematically explore how the quantum harmonic oscillator (QHO) approaches the classical harmonic oscillator (CHO) in the classical limit, demonstrating the Correspondence Principle.

\section{Classical Harmonic Oscillator (CHO) Review}

Consider a particle of mass $m$ attached to a spring with spring constant $k$. The potential energy of this system is given by $V(x) = \frac{1}{2}kx^2$.

\subsection{Equations of Motion and Energy}
From Newton's second law, $F = ma = m\ddot{x} = -kx$. This leads to the differential equation for simple harmonic motion:
\begin{equation}
\ddot{x} + \omega^2 x = 0,
\end{equation}
where $\omega = \sqrt{\frac{k}{m}}$ is the angular frequency of oscillation. The general solution is:
\begin{equation}
x(t) = A \cos(\omega t - \phi),
\end{equation}
where $A$ is the amplitude and $\phi$ is the phase constant, determined by initial conditions.

The velocity is given by:
\begin{equation}
v(t) = \dot{x}(t) = -A\omega \sin(\omega t - \phi).
\end{equation}

The total energy of the classical harmonic oscillator is conserved and is the sum of kinetic and potential energies:
\begin{equation}
E_{classical} = T + V = \frac{1}{2}mv^2 + \frac{1}{2}kx^2 = \frac{1}{2}m(A\omega \sin(\omega t - \phi))^2 + \frac{1}{2}k(A \cos(\omega t - \phi))^2.
\end{equation}
Using $k = m\omega^2$ and $\sin^2(\theta) + \cos^2(\theta) = 1$, we get:
\begin{equation}
E_{classical} = \frac{1}{2}mA^2\omega^2 (\sin^2(\omega t - \phi) + \cos^2(\omega t - \phi)) = \frac{1}{2}mA^2\omega^2 = \frac{1}{2}kA^2.
\end{equation}
Classically, energy $E_{classical}$ can take any continuous non-negative value.

\subsection{Classical Probability Density}
The probability of finding the particle in a small interval $dx$ at position $x$ is proportional to the time spent in that interval. Since the particle moves slower near the turning points ($x = \pm A$) and faster near the equilibrium position ($x = 0$), the probability density is inversely proportional to the speed of the particle:
\begin{equation}
P_{classical}(x) \propto \frac{1}{|v(x)|}.
\end{equation}
From energy conservation, $E_{classical} = \frac{1}{2}mv^2 + \frac{1}{2}kx^2$, we can express the speed as:
\begin{equation}
v(x) = \pm \sqrt{\frac{2}{m}(E_{classical} - \frac{1}{2}kx^2)} = \pm \omega \sqrt{A^2 - x^2}.
\end{equation}
Thus, the classical probability density is:
\begin{equation}
P_{classical}(x) = \frac{C}{|v(x)|} = \frac{C}{\omega \sqrt{A^2 - x^2}} = \frac{C'}{\sqrt{A^2 - x^2}}, \quad \text{for } |x| \leq A,
\end{equation}
and $P_{classical}(x) = 0$ for $|x| > A$.  $C'$ is a normalization constant.

\section{Quantum Harmonic Oscillator (QHO) Review}

In quantum mechanics, the harmonic oscillator is described by the time-independent Schrödinger equation:
\begin{equation}
-\frac{\hbar^2}{2m} \frac{d^2\psi_n(x)}{dx^2} + \frac{1}{2}kx^2 \psi_n(x) = E_n \psi_n(x).
\end{equation}
The solutions to this equation are the energy eigenstates $\psi_n(x)$ with corresponding energy eigenvalues $E_n$.

\subsection{Energy Eigenvalues}
The energy eigenvalues for the quantum harmonic oscillator are quantized and given by:
\begin{equation}
E_n = \left(n + \frac{1}{2}\right)\hbar\omega, \quad n = 0, 1, 2, \ldots
\end{equation}
where $n$ is the quantum number and $\hbar$ is the reduced Planck constant.  Energy levels are discrete and equally spaced with spacing $\hbar\omega$.

\subsection{Energy Eigenstates and Probability Density}
The energy eigenstates $\psi_n(x)$ are given by:
\begin{equation}
\psi_n(x) = \frac{1}{\sqrt{2^n n!}} \left(\frac{m\omega}{\pi\hbar}\right)^{1/4} e^{-\frac{m\omega x^2}{2\hbar}} H_n\left(\sqrt{\frac{m\omega}{\hbar}} x\right),
\end{equation}
where $H_n(u)$ are the Hermite polynomials. The probability density of finding the particle at position $x$ in the $n$-th energy state is given by $|\psi_n(x)|^2$.

For low quantum numbers (small $n$), the probability density $|\psi_n(x)|^2$ is significantly different from the classical probability density. For instance, for $n=0$, $|\psi_0(x)|^2$ is a Gaussian centered at $x=0$, with the highest probability at the equilibrium position, unlike the classical case where the probability is lowest at the equilibrium.  Furthermore, there is a non-zero probability of finding the particle outside the classical turning points, which is forbidden in classical mechanics.

\section{Correspondence Principle and the Harmonic Oscillator}

The Correspondence Principle states that in the limit of large quantum numbers ($n \rightarrow \infty$), quantum mechanics should reproduce classical mechanics. Let's examine how this principle applies to the harmonic oscillator.

\subsection{Energy Levels}
The energy levels for the QHO are $E_n = (n + \frac{1}{2})\hbar\omega$. For large $n$, we can approximate $E_n \approx n\hbar\omega$. The energy spacing between adjacent levels is:
\begin{equation}
\Delta E = E_{n+1} - E_n = \left((n+1) + \frac{1}{2}\right)\hbar\omega - \left(n + \frac{1}{2}\right)\hbar\omega = \hbar\omega.
\end{equation}
This spacing $\hbar\omega$ is constant and independent of $n$. However, consider the \textit{relative} spacing compared to the energy itself:
\begin{equation}
\frac{\Delta E}{E_n} = \frac{\hbar\omega}{(n + \frac{1}{2})\hbar\omega} = \frac{1}{n + \frac{1}{2}} \approx \frac{1}{n} \quad \text{for large } n.
\end{equation}
As $n \rightarrow \infty$, the relative spacing $\frac{\Delta E}{E_n} \rightarrow 0$. This means that for very high energy levels, the energy spectrum becomes effectively continuous, resembling the classical case where energy can take continuous values.  In the classical limit of high energy, the quantization of energy becomes less apparent.

\subsection{Probability Density}
Now let's consider the probability density. For large $n$, the wave function $\psi_n(x)$ becomes highly oscillatory, and the Hermite polynomials $H_n(u)$ also exhibit oscillatory behavior within the classical allowed region and decay exponentially outside.  While a rigorous mathematical proof is complex, we can argue qualitatively and consider some properties:

1. \textbf{Classical Turning Points:} The classical turning points are given by $V(x) = E_{classical}$, so $\frac{1}{2}kx^2 = E_{classical}$. For the $n$-th quantum state, we set $E_{classical} = E_n = (n + \frac{1}{2})\hbar\omega$. Thus, the classical turning points are $x = \pm A_n$, where $\frac{1}{2}kA_n^2 = E_n$, so $A_n = \sqrt{\frac{2E_n}{k}} = \sqrt{\frac{2(n + \frac{1}{2})\hbar\omega}{m\omega^2}} = \sqrt{\frac{(2n+1)\hbar}{m\omega}}$.

2. \textbf{Oscillatory Nature of $\psi_n(x)$ for large $n$:} For large $n$, the Hermite polynomials $H_n(u)$ behave approximately as sinusoidal functions within the classically allowed region.  The square of the wave function, $|\psi_n(x)|^2$, will thus exhibit rapid oscillations within $|x| \leq A_n$.

3. \textbf{Averaging over Oscillations:} If we average the quantum probability density $|\psi_n(x)|^2$ over a spatial interval that is large compared to the wavelength of the oscillations but small compared to the overall extent of the wave function, we find that the averaged quantum probability density approaches the classical probability density.  Intuitively, the rapid oscillations in $|\psi_n(x)|^2$ for large $n$ average out to a smoother distribution that resembles the classical distribution, which is peaked at the turning points and minimum at the equilibrium position.

To illustrate this more explicitly, consider the WKB approximation, which is valid for large quantum numbers. In the WKB approximation, the wave function in the classically allowed region oscillates as:
\begin{equation}
\psi_{WKB}(x) \approx \frac{C}{\sqrt{p(x)}} \cos\left(\frac{1}{\hbar} \int^x p(x') dx' - \frac{\pi}{4}\right),
\end{equation}
where $p(x) = \sqrt{2m(E - V(x))} = \sqrt{2m(E - \frac{1}{2}kx^2)}$ is the classical momentum.  The probability density is then approximately:
\begin{equation}
|\psi_{WKB}(x)|^2 \approx \frac{C^2}{p(x)} \cos^2\left(\frac{1}{\hbar} \int^x p(x') dx' - \frac{\pi}{4}\right).
\end{equation}
Averaging the rapidly oscillating $\cos^2$ term over a cycle gives $\langle \cos^2 \rangle = \frac{1}{2}$.  Thus, the average probability density becomes:
\begin{equation}
\langle |\psi_{WKB}(x)|^2 \rangle \approx \frac{C^2}{2p(x)} = \frac{C'^2}{\sqrt{E - \frac{1}{2}kx^2}} = \frac{C''}{\sqrt{A_n^2 - x^2}},
\end{equation}
where we used $E = E_n = \frac{1}{2}kA_n^2$. This form is proportional to the classical probability density $P_{classical}(x) = \frac{C'}{\sqrt{A^2 - x^2}}$.  Hence, in the limit of large quantum numbers, the average quantum probability density approaches the classical probability density within the classically allowed region.

Furthermore, the quantum probability of finding the particle outside the classical turning points (tunneling) becomes relatively less significant for large $n$. The exponentially decaying tails of $\psi_n(x)$ beyond $x = \pm A_n$ become less prominent compared to the oscillatory behavior within the classical region as $n$ increases.

\section{Conclusion}

We have shown how the quantum harmonic oscillator approaches the classical harmonic oscillator in the limit of large quantum numbers, in accordance with the Correspondence Principle.

\begin{itemize}
    \item \textbf{Energy Levels:} The energy levels of the QHO are quantized, but for large quantum numbers $n$, the relative spacing between energy levels becomes vanishingly small compared to the energy itself.  This makes the energy spectrum effectively continuous, as in classical mechanics.
    \item \textbf{Probability Density:} While the quantum probability density for low $n$ is significantly different from the classical probability density, for large $n$, the average quantum probability density (obtained by averaging over oscillations or using approximations like WKB) approaches the classical probability density within the classically allowed region.  The oscillatory nature of the quantum wave function for large $n$ and its localization within the classical boundaries are key to this correspondence.
\end{itemize}

The harmonic oscillator provides a clear and important example of the Correspondence Principle, demonstrating how quantum mechanics smoothly transitions to classical mechanics in the appropriate limit. This principle is not only a historical cornerstone in the development of quantum theory but also a valuable tool for understanding the relationship between quantum and classical descriptions of physical systems.

\end{document}

thoughts

The user wants me to add labels to all equations in the provided Latex file.
The labels should be descriptive and use underscores instead of spaces.
I will go through the file, identify each equation environment and add a descriptive label.

Equation 1: Differential equation for simple harmonic motion. Label: DE_SimpleHarmonicMotion
Equation 2: General solution for x(t). Label: GeneralSolution_xt
Equation 3: Velocity v(t). Label: Velocity_vt
Equation 4: Classical energy E_classical (expanded). Label: ClassicalEnergy_Expanded
Equation 5: Classical energy E_classical (simplified). Label: ClassicalEnergy_Simplified
Equation 6: Classical probability density proportionality. Label: ClassicalProbabilityDensity_Proportionality
Equation 7: Velocity v(x) from energy conservation. Label: Velocity_vx_EnergyConservation
Equation 8: Classical probability density P_classical(x). Label: ClassicalProbabilityDensity_Px
Equation 9: Time-independent Schrodinger equation. Label: TimeIndependentSchrodingerEquation
Equation 10: Energy eigenvalues E_n. Label: EnergyEigenvalues_En
Equation 11: Energy eigenstates psi_n(x). Label: EnergyEigenstates_Psix
Equation 12: Energy spacing Delta E. Label: EnergySpacing_DeltaE
Equation 13: Relative energy spacing Delta E / E_n. Label: RelativeEnergySpacing
Equation 14: WKB approximation for wave function. Label: WKB_WaveFunction
Equation 15: WKB probability density (before averaging). Label: WKB_ProbabilityDensity_BeforeAveraging
Equation 16: Averaged WKB probability density. Label: AveragedWKB_ProbabilityDensity

response

\documentclass{article}
\usepackage{amsmath}

% margins of 1 inch:
\setlength{\topmargin}{-.5in}
\setlength{\textheight}{9in}
\setlength{\oddsidemargin}{0in}
\setlength{\textwidth}{6.5in}

\usepackage[pdftex]{hyperref} % hyperlink equation and bibliographic citations

\title{The Correspondence Principle and the Harmonic Oscillator}
\author{Ben Payne, with Gemini 2.0}
\date{\today}

\begin{document}
\begin{abstract}
    From Gemini 2.0 Flash
\end{abstract}
\maketitle

\section{Introduction}

The harmonic oscillator is a fundamental system in physics, serving as a model for a wide range of phenomena from molecular vibrations to oscillations in electromagnetic fields.  It is particularly insightful in the context of the Correspondence Principle, which states that the predictions of quantum mechanics must agree with classical mechanics in the limit of large quantum numbers or high energies. This principle provides a crucial bridge between the seemingly disparate worlds of classical and quantum physics.  In this document, we will mathematically explore how the quantum harmonic oscillator (QHO) approaches the classical harmonic oscillator (CHO) in the classical limit, demonstrating the Correspondence Principle.

\section{Classical Harmonic Oscillator (CHO) Review}

Consider a particle of mass $m$ attached to a spring with spring constant $k$. The potential energy of this system is given by $V(x) = \frac{1}{2}kx^2$.

\subsection{Equations of Motion and Energy}
From Newton's second law, $F = ma = m\ddot{x} = -kx$. This leads to the differential equation for simple harmonic motion:
\begin{equation} \label{DE_SimpleHarmonicMotion}
\ddot{x} + \omega^2 x = 0,
\end{equation}
where $\omega = \sqrt{\frac{k}{m}}$ is the angular frequency of oscillation. The general solution is:
\begin{equation} \label{GeneralSolution_xt}
x(t) = A \cos(\omega t - \phi),
\end{equation}
where $A$ is the amplitude and $\phi$ is the phase constant, determined by initial conditions.

The velocity is given by:
\begin{equation} \label{Velocity_vt}
v(t) = \dot{x}(t) = -A\omega \sin(\omega t - \phi).
\end{equation}

The total energy of the classical harmonic oscillator is conserved and is the sum of kinetic and potential energies:
\begin{equation} \label{ClassicalEnergy_Expanded}
E_{classical} = T + V = \frac{1}{2}mv^2 + \frac{1}{2}kx^2 = \frac{1}{2}m(A\omega \sin(\omega t - \phi))^2 + \frac{1}{2}k(A \cos(\omega t - \phi))^2.
\end{equation}
Using $k = m\omega^2$ and $\sin^2(\theta) + \cos^2(\theta) = 1$, we get:
\begin{equation} \label{ClassicalEnergy_Simplified}
E_{classical} = \frac{1}{2}mA^2\omega^2 (\sin^2(\omega t - \phi) + \cos^2(\omega t - \phi)) = \frac{1}{2}mA^2\omega^2 = \frac{1}{2}kA^2.
\end{equation}
Classically, energy $E_{classical}$ can take any continuous non-negative value.

\subsection{Classical Probability Density}
The probability of finding the particle in a small interval $dx$ at position $x$ is proportional to the time spent in that interval. Since the particle moves slower near the turning points ($x = \pm A$) and faster near the equilibrium position ($x = 0$), the probability density is inversely proportional to the speed of the particle:
\begin{equation} \label{ClassicalProbabilityDensity_Proportionality}
P_{classical}(x) \propto \frac{1}{|v(x)|}.
\end{equation}
From energy conservation, $E_{classical} = \frac{1}{2}mv^2 + \frac{1}{2}kx^2$, we can express the speed as:
\begin{equation} \label{Velocity_vx_EnergyConservation}
v(x) = \pm \sqrt{\frac{2}{m}(E_{classical} - \frac{1}{2}kx^2)} = \pm \omega \sqrt{A^2 - x^2}.
\end{equation}
Thus, the classical probability density is:
\begin{equation} \label{ClassicalProbabilityDensity_Px}
P_{classical}(x) = \frac{C}{|v(x)|} = \frac{C}{\omega \sqrt{A^2 - x^2}} = \frac{C'}{\sqrt{A^2 - x^2}}, \quad \text{for } |x| \leq A,
\end{equation}
and $P_{classical}(x) = 0$ for $|x| > A$.  $C'$ is a normalization constant.

\section{Quantum Harmonic Oscillator (QHO) Review}

In quantum mechanics, the harmonic oscillator is described by the time-independent Schrödinger equation:
\begin{equation} \label{TimeIndependentSchrodingerEquation}
-\frac{\hbar^2}{2m} \frac{d^2\psi_n(x)}{dx^2} + \frac{1}{2}kx^2 \psi_n(x) = E_n \psi_n(x).
\end{equation}
The solutions to this equation are the energy eigenstates $\psi_n(x)$ with corresponding energy eigenvalues $E_n$.

\subsection{Energy Eigenvalues}
The energy eigenvalues for the quantum harmonic oscillator are quantized and given by:
\begin{equation} \label{EnergyEigenvalues_En}
E_n = \left(n + \frac{1}{2}\right)\hbar\omega, \quad n = 0, 1, 2, \ldots
\end{equation}
where $n$ is the quantum number and $\hbar$ is the reduced Planck constant.  Energy levels are discrete and equally spaced with spacing $\hbar\omega$.

\subsection{Energy Eigenstates and Probability Density}
The energy eigenstates $\psi_n(x)$ are given by:
\begin{equation} \label{EnergyEigenstates_Psix}
\psi_n(x) = \frac{1}{\sqrt{2^n n!}} \left(\frac{m\omega}{\pi\hbar}\right)^{1/4} e^{-\frac{m\omega x^2}{2\hbar}} H_n\left(\sqrt{\frac{m\omega}{\hbar}} x\right),
\end{equation}
where $H_n(u)$ are the Hermite polynomials. The probability density of finding the particle at position $x$ in the $n$-th energy state is given by $|\psi_n(x)|^2$.

For low quantum numbers (small $n$), the probability density $|\psi_n(x)|^2$ is significantly different from the classical probability density. For instance, for $n=0$, $|\psi_0(x)|^2$ is a Gaussian centered at $x=0$, with the highest probability at the equilibrium position, unlike the classical case where the probability is lowest at the equilibrium.  Furthermore, there is a non-zero probability of finding the particle outside the classical turning points, which is forbidden in classical mechanics.

\section{Correspondence Principle and the Harmonic Oscillator}

The Correspondence Principle states that in the limit of large quantum numbers ($n \rightarrow \infty$), quantum mechanics should reproduce classical mechanics. Let's examine how this principle applies to the harmonic oscillator.

\subsection{Energy Levels}
The energy levels for the QHO are $E_n = (n + \frac{1}{2})\hbar\omega$. For large $n$, we can approximate $E_n \approx n\hbar\omega$. The energy spacing between adjacent levels is:
\begin{equation} \label{EnergySpacing_DeltaE}
\Delta E = E_{n+1} - E_n = \left((n+1) + \frac{1}{2}\right)\hbar\omega - \left(n + \frac{1}{2}\right)\hbar\omega = \hbar\omega.
\end{equation}
This spacing $\hbar\omega$ is constant and independent of $n$. However, consider the \textit{relative} spacing compared to the energy itself:
\begin{equation} \label{RelativeEnergySpacing}
\frac{\Delta E}{E_n} = \frac{\hbar\omega}{(n + \frac{1}{2})\hbar\omega} = \frac{1}{n + \frac{1}{2}} \approx \frac{1}{n} \quad \text{for large } n.
\end{equation}
As $n \rightarrow \infty$, the relative spacing $\frac{\Delta E}{E_n} \rightarrow 0$. This means that for very high energy levels, the energy spectrum becomes effectively continuous, resembling the classical case where energy can take continuous values.  In the classical limit of high energy, the quantization of energy becomes less apparent.

\subsection{Probability Density}
Now let's consider the probability density. For large $n$, the wave function $\psi_n(x)$ becomes highly oscillatory, and the Hermite polynomials $H_n(u)$ also exhibit oscillatory behavior within the classical allowed region and decay exponentially outside.  While a rigorous mathematical proof is complex, we can argue qualitatively and consider some properties:

1. \textbf{Classical Turning Points:} The classical turning points are given by $V(x) = E_{classical}$, so $\frac{1}{2}kx^2 = E_{classical}$. For the $n$-th quantum state, we set $E_{classical} = E_n = (n + \frac{1}{2})\hbar\omega$. Thus, the classical turning points are $x = \pm A_n$, where $\frac{1}{2}kA_n^2 = E_n$, so $A_n = \sqrt{\frac{2E_n}{k}} = \sqrt{\frac{2(n + \frac{1}{2})\hbar\omega}{m\omega^2}} = \sqrt{\frac{(2n+1)\hbar}{m\omega}}$.

2. \textbf{Oscillatory Nature of $\psi_n(x)$ for large $n$:} For large $n$, the Hermite polynomials $H_n(u)$ behave approximately as sinusoidal functions within the classically allowed region.  The square of the wave function, $|\psi_n(x)|^2$, will thus exhibit rapid oscillations within $|x| \leq A_n$.

3. \textbf{Averaging over Oscillations:} If we average the quantum probability density $|\psi_n(x)|^2$ over a spatial interval that is large compared to the wavelength of the oscillations but small compared to the overall extent of the wave function, we find that the averaged quantum probability density approaches the classical probability density.  Intuitively, the rapid oscillations in $|\psi_n(x)|^2$ for large $n$ average out to a smoother distribution that resembles the classical distribution, which is peaked at the turning points and minimum at the equilibrium position.

To illustrate this more explicitly, consider the WKB approximation, which is valid for large quantum numbers. In the WKB approximation, the wave function in the classically allowed region oscillates as:
\begin{equation} \label{WKB_WaveFunction}
\psi_{WKB}(x) \approx \frac{C}{\sqrt{p(x)}} \cos\left(\frac{1}{\hbar} \int^x p(x') dx' - \frac{\pi}{4}\right),
\end{equation}
where $p(x) = \sqrt{2m(E - V(x))} = \sqrt{2m(E - \frac{1}{2}kx^2)}$ is the classical momentum.  The probability density is then approximately:
\begin{equation} \label{WKB_ProbabilityDensity_BeforeAveraging}
|\psi_{WKB}(x)|^2 \approx \frac{C^2}{p(x)} \cos^2\left(\frac{1}{\hbar} \int^x p(x') dx' - \frac{\pi}{4}\right).
\end{equation}
Averaging the rapidly oscillating $\cos^2$ term over a cycle gives $\langle \cos^2 \rangle = \frac{1}{2}$.  Thus, the average probability density becomes:
\begin{equation} \label{AveragedWKB_ProbabilityDensity}
\langle |\psi_{WKB}(x)|^2 \rangle \approx \frac{C^2}{2p(x)} = \frac{C'^2}{\sqrt{E - \frac{1}{2}kx^2}} = \frac{C''}{\sqrt{A_n^2 - x^2}},
\end{equation}
where we used $E = E_n = \frac{1}{2}kA_n^2$. This form is proportional to the classical probability density $P_{classical}(x) = \frac{C'}{\sqrt{A^2 - x^2}}$.  Hence, in the limit of large quantum numbers, the average quantum probability density approaches the classical probability density within the classically allowed region.

Furthermore, the quantum probability of finding the particle outside the classical turning points (tunneling) becomes relatively less significant for large $n$. The exponentially decaying tails of $\psi_n(x)$ beyond $x = \pm A_n$ become less prominent compared to the oscillatory behavior within the classical region as $n$ increases.

\section{Conclusion}

We have shown how the quantum harmonic oscillator approaches the classical harmonic oscillator in the limit of large quantum numbers, in accordance with the Correspondence Principle.

\begin{itemize}
    \item \textbf{Energy Levels:} The energy levels of the QHO are quantized, but for large quantum numbers $n$, the relative spacing between energy levels becomes vanishingly small compared to the energy itself.  This makes the energy spectrum effectively continuous, as in classical mechanics.
    \item \textbf{Probability Density:} While the quantum probability density for low $n$ is significantly different from the classical probability density, for large $n$, the average quantum probability density (obtained by averaging over oscillations or using approximations like WKB) approaches the classical probability density within the classically allowed region.  The oscillatory nature of the quantum wave function for large $n$ and its localization within the classical boundaries are key to this correspondence.
\end{itemize}

The harmonic oscillator provides a clear and important example of the Correspondence Principle, demonstrating how quantum mechanics smoothly transitions to classical mechanics in the appropriate limit. This principle is not only a historical cornerstone in the development of quantum theory but also a valuable tool for understanding the relationship between quantum and classical descriptions of physical systems.

\end{document}

@bhpayne
Copy link
Member Author

bhpayne commented Jan 26, 2025

prompt

The following latex file contains equations and variables. Provide a list, formatted as Latex, of every unique variable and a description of that variable.

For each entry in the list include a reference to the labeled equations where each variable is used. Write out just the Latex list as your answer.

\documentclass{article}
\usepackage{amsmath}

% margins of 1 inch:
\setlength{\topmargin}{-.5in}
\setlength{\textheight}{9in}
\setlength{\oddsidemargin}{0in}
\setlength{\textwidth}{6.5in}

\usepackage[pdftex]{hyperref} % hyperlink equation and bibliographic citations

\title{The Correspondence Principle and the Harmonic Oscillator}
\author{Ben Payne, with Gemini 2.0}
\date{\today}

\begin{document}

\maketitle
\begin{abstract}
    From Gemini 2.0 Flash
\end{abstract}

\section{Introduction}

The harmonic oscillator is a fundamental system in physics, serving as a model for a wide range of phenomena from molecular vibrations to oscillations in electromagnetic fields.  It is particularly insightful in the context of the Correspondence Principle, which states that the predictions of quantum mechanics must agree with classical mechanics in the limit of large quantum numbers or high energies. This principle provides a crucial bridge between the seemingly disparate worlds of classical and quantum physics.  In this document, we will mathematically explore how the quantum harmonic oscillator (QHO) approaches the classical harmonic oscillator (CHO) in the classical limit, demonstrating the Correspondence Principle.

\section{Classical Harmonic Oscillator (CHO) Review}

Consider a particle of mass $m$ attached to a spring with spring constant $k$. The potential energy of this system is given by $V(x) = \frac{1}{2}kx^2$.

\subsection{Equations of Motion and Energy}
From Newton's second law, $F = ma = m\ddot{x} = -kx$. This leads to the differential equation for simple harmonic motion:
\begin{equation} \label{DE_SimpleHarmonicMotion}
\ddot{x} + \omega^2 x = 0,
\end{equation}
where $\omega = \sqrt{\frac{k}{m}}$ is the angular frequency of oscillation. The general solution is:
\begin{equation} \label{GeneralSolution_xt}
x(t) = A \cos(\omega t - \phi),
\end{equation}
where $A$ is the amplitude and $\phi$ is the phase constant, determined by initial conditions.

The velocity is given by:
\begin{equation} \label{Velocity_vt}
v(t) = \dot{x}(t) = -A\omega \sin(\omega t - \phi).
\end{equation}

The total energy of the classical harmonic oscillator is conserved and is the sum of kinetic and potential energies:
\begin{equation} \label{ClassicalEnergy_Expanded}
E_{classical} = T + V = \frac{1}{2}mv^2 + \frac{1}{2}kx^2 = \frac{1}{2}m(A\omega \sin(\omega t - \phi))^2 + \frac{1}{2}k(A \cos(\omega t - \phi))^2.
\end{equation}
Using $k = m\omega^2$ and $\sin^2(\theta) + \cos^2(\theta) = 1$, we get:
\begin{equation} \label{ClassicalEnergy_Simplified}
E_{classical} = \frac{1}{2}mA^2\omega^2 (\sin^2(\omega t - \phi) + \cos^2(\omega t - \phi)) = \frac{1}{2}mA^2\omega^2 = \frac{1}{2}kA^2.
\end{equation}
Classically, energy $E_{classical}$ can take any continuous non-negative value.

\subsection{Classical Probability Density}
The probability of finding the particle in a small interval $dx$ at position $x$ is proportional to the time spent in that interval. Since the particle moves slower near the turning points ($x = \pm A$) and faster near the equilibrium position ($x = 0$), the probability density is inversely proportional to the speed of the particle:
\begin{equation} \label{ClassicalProbabilityDensity_Proportionality}
P_{classical}(x) \propto \frac{1}{|v(x)|}.
\end{equation}
From energy conservation, $E_{classical} = \frac{1}{2}mv^2 + \frac{1}{2}kx^2$, we can express the speed as:
\begin{equation} \label{Velocity_vx_EnergyConservation}
v(x) = \pm \sqrt{\frac{2}{m}(E_{classical} - \frac{1}{2}kx^2)} = \pm \omega \sqrt{A^2 - x^2}.
\end{equation}
Thus, the classical probability density is:
\begin{equation} \label{ClassicalProbabilityDensity_Px}
P_{classical}(x) = \frac{C}{|v(x)|} = \frac{C}{\omega \sqrt{A^2 - x^2}} = \frac{C'}{\sqrt{A^2 - x^2}}, \quad \text{for } |x| \leq A,
\end{equation}
and $P_{classical}(x) = 0$ for $|x| > A$.  $C'$ is a normalization constant.

\section{Quantum Harmonic Oscillator (QHO) Review}

In quantum mechanics, the harmonic oscillator is described by the time-independent Schrödinger equation:
\begin{equation} \label{TimeIndependentSchrodingerEquation}
-\frac{\hbar^2}{2m} \frac{d^2\psi_n(x)}{dx^2} + \frac{1}{2}kx^2 \psi_n(x) = E_n \psi_n(x).
\end{equation}
The solutions to this equation are the energy eigenstates $\psi_n(x)$ with corresponding energy eigenvalues $E_n$.

\subsection{Energy Eigenvalues}
The energy eigenvalues for the quantum harmonic oscillator are quantized and given by:
\begin{equation} \label{EnergyEigenvalues_En}
E_n = \left(n + \frac{1}{2}\right)\hbar\omega, \quad n = 0, 1, 2, \ldots
\end{equation}
where $n$ is the quantum number and $\hbar$ is the reduced Planck constant.  Energy levels are discrete and equally spaced with spacing $\hbar\omega$.

\subsection{Energy Eigenstates and Probability Density}
The energy eigenstates $\psi_n(x)$ are given by:
\begin{equation} \label{EnergyEigenstates_Psix}
\psi_n(x) = \frac{1}{\sqrt{2^n n!}} \left(\frac{m\omega}{\pi\hbar}\right)^{1/4} e^{-\frac{m\omega x^2}{2\hbar}} H_n\left(\sqrt{\frac{m\omega}{\hbar}} x\right),
\end{equation}
where $H_n(u)$ are the Hermite polynomials. The probability density of finding the particle at position $x$ in the $n$-th energy state is given by $|\psi_n(x)|^2$.

For low quantum numbers (small $n$), the probability density $|\psi_n(x)|^2$ is significantly different from the classical probability density. For instance, for $n=0$, $|\psi_0(x)|^2$ is a Gaussian centered at $x=0$, with the highest probability at the equilibrium position, unlike the classical case where the probability is lowest at the equilibrium.  Furthermore, there is a non-zero probability of finding the particle outside the classical turning points, which is forbidden in classical mechanics.

\section{Correspondence Principle and the Harmonic Oscillator}

The Correspondence Principle states that in the limit of large quantum numbers ($n \rightarrow \infty$), quantum mechanics should reproduce classical mechanics. Let's examine how this principle applies to the harmonic oscillator.

\subsection{Energy Levels}
The energy levels for the QHO are $E_n = (n + \frac{1}{2})\hbar\omega$. For large $n$, we can approximate $E_n \approx n\hbar\omega$. The energy spacing between adjacent levels is:
\begin{equation} \label{EnergySpacing_DeltaE}
\Delta E = E_{n+1} - E_n = \left((n+1) + \frac{1}{2}\right)\hbar\omega - \left(n + \frac{1}{2}\right)\hbar\omega = \hbar\omega.
\end{equation}
This spacing $\hbar\omega$ is constant and independent of $n$. However, consider the \textit{relative} spacing compared to the energy itself:
\begin{equation} \label{RelativeEnergySpacing}
\frac{\Delta E}{E_n} = \frac{\hbar\omega}{(n + \frac{1}{2})\hbar\omega} = \frac{1}{n + \frac{1}{2}} \approx \frac{1}{n} \quad \text{for large } n.
\end{equation}
As $n \rightarrow \infty$, the relative spacing $\frac{\Delta E}{E_n} \rightarrow 0$. This means that for very high energy levels, the energy spectrum becomes effectively continuous, resembling the classical case where energy can take continuous values.  In the classical limit of high energy, the quantization of energy becomes less apparent.

\subsection{Probability Density}
Now let's consider the probability density. For large $n$, the wave function $\psi_n(x)$ becomes highly oscillatory, and the Hermite polynomials $H_n(u)$ also exhibit oscillatory behavior within the classical allowed region and decay exponentially outside.  While a rigorous mathematical proof is complex, we can argue qualitatively and consider some properties:

1. \textbf{Classical Turning Points:} The classical turning points are given by $V(x) = E_{classical}$, so $\frac{1}{2}kx^2 = E_{classical}$. For the $n$-th quantum state, we set $E_{classical} = E_n = (n + \frac{1}{2})\hbar\omega$. Thus, the classical turning points are $x = \pm A_n$, where $\frac{1}{2}kA_n^2 = E_n$, so $A_n = \sqrt{\frac{2E_n}{k}} = \sqrt{\frac{2(n + \frac{1}{2})\hbar\omega}{m\omega^2}} = \sqrt{\frac{(2n+1)\hbar}{m\omega}}$.

2. \textbf{Oscillatory Nature of $\psi_n(x)$ for large $n$:} For large $n$, the Hermite polynomials $H_n(u)$ behave approximately as sinusoidal functions within the classically allowed region.  The square of the wave function, $|\psi_n(x)|^2$, will thus exhibit rapid oscillations within $|x| \leq A_n$.

3. \textbf{Averaging over Oscillations:} If we average the quantum probability density $|\psi_n(x)|^2$ over a spatial interval that is large compared to the wavelength of the oscillations but small compared to the overall extent of the wave function, we find that the averaged quantum probability density approaches the classical probability density.  Intuitively, the rapid oscillations in $|\psi_n(x)|^2$ for large $n$ average out to a smoother distribution that resembles the classical distribution, which is peaked at the turning points and minimum at the equilibrium position.

To illustrate this more explicitly, consider the WKB approximation, which is valid for large quantum numbers. In the WKB approximation, the wave function in the classically allowed region oscillates as:
\begin{equation} \label{WKB_WaveFunction}
\psi_{WKB}(x) \approx \frac{C}{\sqrt{p(x)}} \cos\left(\frac{1}{\hbar} \int^x p(x') dx' - \frac{\pi}{4}\right),
\end{equation}
where $p(x) = \sqrt{2m(E - V(x))} = \sqrt{2m(E - \frac{1}{2}kx^2)}$ is the classical momentum.  The probability density is then approximately:
\begin{equation} \label{WKB_ProbabilityDensity_BeforeAveraging}
|\psi_{WKB}(x)|^2 \approx \frac{C^2}{p(x)} \cos^2\left(\frac{1}{\hbar} \int^x p(x') dx' - \frac{\pi}{4}\right).
\end{equation}
Averaging the rapidly oscillating $\cos^2$ term over a cycle gives $\langle \cos^2 \rangle = \frac{1}{2}$.  Thus, the average probability density becomes:
\begin{equation} \label{AveragedWKB_ProbabilityDensity}
\langle |\psi_{WKB}(x)|^2 \rangle \approx \frac{C^2}{2p(x)} = \frac{C'^2}{\sqrt{E - \frac{1}{2}kx^2}} = \frac{C''}{\sqrt{A_n^2 - x^2}},
\end{equation}
where we used $E = E_n = \frac{1}{2}kA_n^2$. This form is proportional to the classical probability density $P_{classical}(x) = \frac{C'}{\sqrt{A^2 - x^2}}$.  Hence, in the limit of large quantum numbers, the average quantum probability density approaches the classical probability density within the classically allowed region.

Furthermore, the quantum probability of finding the particle outside the classical turning points (tunneling) becomes relatively less significant for large $n$. The exponentially decaying tails of $\psi_n(x)$ beyond $x = \pm A_n$ become less prominent compared to the oscillatory behavior within the classical region as $n$ increases.

\section{Conclusion}

We have shown how the quantum harmonic oscillator approaches the classical harmonic oscillator in the limit of large quantum numbers, in accordance with the Correspondence Principle.

\begin{itemize}
    \item \textbf{Energy Levels:} The energy levels of the QHO are quantized, but for large quantum numbers $n$, the relative spacing between energy levels becomes vanishingly small compared to the energy itself.  This makes the energy spectrum effectively continuous, as in classical mechanics.
    \item \textbf{Probability Density:} While the quantum probability density for low $n$ is significantly different from the classical probability density, for large $n$, the average quantum probability density (obtained by averaging over oscillations or using approximations like WKB) approaches the classical probability density within the classically allowed region.  The oscillatory nature of the quantum wave function for large $n$ and its localization within the classical boundaries are key to this correspondence.
\end{itemize}

The harmonic oscillator provides a clear and important example of the Correspondence Principle, demonstrating how quantum mechanics smoothly transitions to classical mechanics in the appropriate limit. This principle is not only a historical cornerstone in the development of quantum theory but also a valuable tool for understanding the relationship between quantum and classical descriptions of physical systems.

\end{document}

thoughts

The user wants a Latex formatted list of all unique variables in the provided Latex document.
For each variable, I need to provide a description and a list of equation labels where the variable is used.

Here's the plan:

  1. Read through the Latex document and identify all variables.
  2. For each variable, write a description based on the context in the document.
  3. For each variable, identify all labeled equations where it appears.
  4. Compile the information into a Latex list format.
  5. Output only the Latex list as the answer.

Let's start identifying variables and their descriptions and equation references.

Variables:

  • $m$: mass of the particle. Equations: DE_SimpleHarmonicMotion, GeneralSolution_xt, Velocity_vt, ClassicalEnergy_Expanded, ClassicalEnergy_Simplified, Velocity_vx_EnergyConservation, TimeIndependentSchrodingerEquation, EnergyEigenstates_Psix, A_n equation, WKB_ProbabilityDensity_BeforeAveraging
  • $k$: spring constant. Equations: DE_SimpleHarmonicMotion, ClassicalEnergy_Expanded, ClassicalEnergy_Simplified, Velocity_vx_EnergyConservation, TimeIndependentSchrodingerEquation, ClassicalEnergy_Simplified, A_n equation, AveragedWKB_ProbabilityDensity
  • $x$: position. Equations: DE_SimpleHarmonicMotion, GeneralSolution_xt, Velocity_vt, ClassicalEnergy_Expanded, Velocity_vx_EnergyConservation, ClassicalProbabilityDensity_Px, TimeIndependentSchrodingerEquation, EnergyEigenstates_Psix, WKB_WaveFunction, WKB_ProbabilityDensity_BeforeAveraging, AveragedWKB_ProbabilityDensity
  • $\ddot{x}$: second derivative of position with respect to time (acceleration). Equation: DE_SimpleHarmonicMotion
  • $\omega$: angular frequency. Equations: DE_SimpleHarmonicMotion, GeneralSolution_xt, Velocity_vt, ClassicalEnergy_Expanded, ClassicalEnergy_Simplified, Velocity_vx_EnergyConservation, ClassicalProbabilityDensity_Px, EnergyEigenvalues_En, EnergyEigenstates_Psix, EnergySpacing_DeltaE, RelativeEnergySpacing, A_n equation
  • $t$: time. Equations: GeneralSolution_xt, Velocity_vt, ClassicalEnergy_Expanded
  • $A$: amplitude. Equations: GeneralSolution_xt, Velocity_vt, ClassicalEnergy_Expanded, ClassicalEnergy_Simplified, ClassicalProbabilityDensity_Px, Velocity_vx_EnergyConservation, ClassicalProbabilityDensity_Px
  • $\phi$: phase constant. Equations: GeneralSolution_xt, Velocity_vt, ClassicalEnergy_Expanded, ClassicalEnergy_Simplified
  • $v$: velocity. Equations: Velocity_vt, ClassicalEnergy_Expanded, ClassicalProbabilityDensity_Proportionality, Velocity_vx_EnergyConservation, ClassicalProbabilityDensity_Px
  • $E_{classical}$: classical energy. Equations: ClassicalEnergy_Expanded, ClassicalEnergy_Simplified, Velocity_vx_EnergyConservation, ClassicalProbabilityDensity_Proportionality, A_n equation
  • $T$: kinetic energy. Equation: ClassicalEnergy_Expanded
  • $V$: potential energy. Equation: ClassicalEnergy_Expanded, TimeIndependentSchrodingerEquation, A_n equation, WKB_WaveFunction, AveragedWKB_ProbabilityDensity
  • $P_{classical}(x)$: classical probability density as a function of position. Equations: ClassicalProbabilityDensity_Proportionality, ClassicalProbabilityDensity_Px, AveragedWKB_ProbabilityDensity
  • $C, C', C''$: normalization constants for classical probability density. Equations: ClassicalProbabilityDensity_Px, AveragedWKB_ProbabilityDensity
  • $\psi_n(x)$: energy eigenstate wave function for quantum number n as a function of position. Equations: TimeIndependentSchrodingerEquation, EnergyEigenstates_Psix, WKB_WaveFunction
  • $n$: quantum number. Equations: EnergyEigenvalues_En, EnergyEigenstates_Psix, RelativeEnergySpacing, A_n equation
  • $E_n$: energy eigenvalue for quantum number n. Equations: TimeIndependentSchrodingerEquation, EnergyEigenvalues_En, EnergySpacing_DeltaE, RelativeEnergySpacing, A_n equation, AveragedWKB_ProbabilityDensity
  • $\hbar$: reduced Planck constant. Equations: TimeIndependentSchrodingerEquation, EnergyEigenvalues_En, EnergyEigenstates_Psix, EnergySpacing_DeltaE, RelativeEnergySpacing, A_n equation, WKB_WaveFunction, WKB_ProbabilityDensity_BeforeAveraging
  • $H_n(u)$: Hermite polynomials of order n with argument u. Equation: EnergyEigenstates_Psix
  • $u$: argument of Hermite polynomials, $u = \sqrt{\frac{m\omega}{\hbar}} x$. Equation: EnergyEigenstates_Psix
  • $\Delta E$: energy spacing between adjacent energy levels. Equation: EnergySpacing_DeltaE, RelativeEnergySpacing
  • $A_n$: classical turning points for the n-th quantum state. Equation: A_n equation, AveragedWKB_ProbabilityDensity
  • $\psi_{WKB}(x)$: WKB approximation of the wave function. Equation: WKB_WaveFunction, WKB_ProbabilityDensity_BeforeAveraging
  • $p(x)$: classical momentum as a function of position. Equations: WKB_WaveFunction, WKB_ProbabilityDensity_BeforeAveraging, AveragedWKB_ProbabilityDensity

Now, format as a Latex list. Ensure to only output the Latex list as the answer.

response

\begin{itemize}
    \item $A$: Amplitude of oscillation.
    \begin{itemize}
        \item Used in equations: \eqref{GeneralSolution_xt}, \eqref{Velocity_vt}, \eqref{ClassicalEnergy_Expanded}, \eqref{ClassicalEnergy_Simplified}, \eqref{ClassicalProbabilityDensity_Px}.
    \end{itemize}
    \item $A_n$: Classical turning points for the $n$-th quantum state.
    \begin{itemize}
        \item Used in equations: \eqref{AveragedWKB_ProbabilityDensity}.
    \end{itemize}
    \item $C, C', C''$: Normalization constants for classical probability density.
    \begin{itemize}
        \item Used in equations: \eqref{ClassicalProbabilityDensity_Px}, \eqref{AveragedWKB_ProbabilityDensity}, \eqref{WKB_WaveFunction}, \eqref{WKB_ProbabilityDensity_BeforeAveraging}.
    \end{itemize}
    \item $E_{classical}$: Total energy of the classical harmonic oscillator.
    \begin{itemize}
        \item Used in equations: \eqref{ClassicalEnergy_Expanded}, \eqref{ClassicalEnergy_Simplified}, \eqref{Velocity_vx_EnergyConservation}.
    \end{itemize}
    \item $E_n$: Energy eigenvalue for the $n$-th quantum state of the quantum harmonic oscillator.
    \begin{itemize}
        \item Used in equations: \eqref{TimeIndependentSchrodingerEquation}, \eqref{EnergyEigenvalues_En}, \eqref{EnergySpacing_DeltaE}, \eqref{RelativeEnergySpacing}, \eqref{AveragedWKB_ProbabilityDensity}.
    \end{itemize}
    \item $F$: Force.
    \begin{itemize}
        \item Used in equations: None (mentioned in text before equation \eqref{DE_SimpleHarmonicMotion}).
    \end{itemize}
    \item $H_n(u)$: Hermite polynomial of order $n$ with argument $u$.
    \begin{itemize}
        \item Used in equations: \eqref{EnergyEigenstates_Psix}.
    \end{itemize}
    \item $k$: Spring constant.
    \begin{itemize}
        \item Used in equations: \eqref{DE_SimpleHarmonicMotion}, \eqref{ClassicalEnergy_Expanded}, \eqref{ClassicalEnergy_Simplified}, \eqref{Velocity_vx_EnergyConservation}, \eqref{TimeIndependentSchrodingerEquation}, \eqref{AveragedWKB_ProbabilityDensity}.
    \end{itemize}
    \item $m$: Mass of the particle.
    \begin{itemize}
        \item Used in equations: \eqref{DE_SimpleHarmonicMotion}, \eqref{ClassicalEnergy_Expanded}, \eqref{ClassicalEnergy_Simplified}, \eqref{Velocity_vx_EnergyConservation}, \eqref{TimeIndependentSchrodingerEquation}, \eqref{EnergyEigenstates_Psix}, \eqref{WKB_WaveFunction}, \eqref{WKB_ProbabilityDensity_BeforeAveraging}.
    \end{itemize}
    \item $n$: Quantum number.
    \begin{itemize}
        \item Used in equations: \eqref{EnergyEigenvalues_En}, \eqref{EnergyEigenstates_Psix}, \eqref{RelativeEnergySpacing}, \eqref{AveragedWKB_ProbabilityDensity}.
    \end{itemize}
    \item $P_{classical}(x)$: Classical probability density as a function of position $x$.
    \begin{itemize}
        \item Used in equations: \eqref{ClassicalProbabilityDensity_Proportionality}, \eqref{ClassicalProbabilityDensity_Px}.
    \end{itemize}
    \item $p(x)$: Classical momentum as a function of position $x$.
    \begin{itemize}
        \item Used in equations: \eqref{WKB_WaveFunction}, \eqref{WKB_ProbabilityDensity_BeforeAveraging}, \eqref{AveragedWKB_ProbabilityDensity}.
    \end{itemize}
    \item $T$: Kinetic energy.
    \begin{itemize}
        \item Used in equations: \eqref{ClassicalEnergy_Expanded}.
    \end{itemize}
    \item $t$: Time.
    \begin{itemize}
        \item Used in equations: \eqref{GeneralSolution_xt}, \eqref{Velocity_vt}, \eqref{ClassicalEnergy_Expanded}.
    \end{itemize}
    \item $u$: Argument of Hermite polynomials, $u = \sqrt{\frac{m\omega}{\hbar}} x$.
    \begin{itemize}
        \item Used in equations: \eqref{EnergyEigenstates_Psix}.
    \end{itemize}
    \item $V$: Potential energy.
    \begin{itemize}
        \item Used in equations: \eqref{ClassicalEnergy_Expanded}, \eqref{TimeIndependentSchrodingerEquation}, \eqref{AveragedWKB_ProbabilityDensity}.
    \end{itemize}
    \item $v$: Velocity of the particle.
    \begin{itemize}
        \item Used in equations: \eqref{Velocity_vt}, \eqref{ClassicalEnergy_Expanded}, \eqref{ClassicalProbabilityDensity_Proportionality}, \eqref{ClassicalProbabilityDensity_Px}, \eqref{Velocity_vx_EnergyConservation}.
    \end{itemize}
    \item $x$: Position of the particle.
    \begin{itemize}
        \item Used in equations: \eqref{DE_SimpleHarmonicMotion}, \eqref{GeneralSolution_xt}, \eqref{Velocity_vt}, \eqref{ClassicalEnergy_Expanded}, \eqref{Velocity_vx_EnergyConservation}, \eqref{ClassicalProbabilityDensity_Px}, \eqref{TimeIndependentSchrodingerEquation}, \eqref{EnergyEigenstates_Psix}, \eqref{WKB_WaveFunction}, \eqref{WKB_ProbabilityDensity_BeforeAveraging}.
    \end{itemize}
    \item $\ddot{x}$: Acceleration of the particle.
    \begin{itemize}
        \item Used in equations: \eqref{DE_SimpleHarmonicMotion}.
    \end{itemize}
    \item $\Delta E$: Energy spacing between adjacent energy levels in the quantum harmonic oscillator.
    \begin{itemize}
        \item Used in equations: \eqref{EnergySpacing_DeltaE}, \eqref{RelativeEnergySpacing}.
    \end{itemize}
    \item $\phi$: Phase constant in the general solution for simple harmonic motion.
    \begin{itemize}
        \item Used in equations: \eqref{GeneralSolution_xt}, \eqref{Velocity_vt}, \eqref{ClassicalEnergy_Expanded}, \eqref{ClassicalEnergy_Simplified}.
    \end{itemize}
    \item $\psi_n(x)$: Energy eigenstate wave function for the $n$-th state of the quantum harmonic oscillator as a function of position $x$.
    \begin{itemize}
        \item Used in equations: \eqref{TimeIndependentSchrodingerEquation}, \eqref{EnergyEigenstates_Psix}.
    \end{itemize}
    \item $\psi_{WKB}(x)$: WKB approximation of the wave function.
    \begin{itemize}
        \item Used in equations: \eqref{WKB_WaveFunction}, \eqref{WKB_ProbabilityDensity_BeforeAveraging}.
    \end{itemize}
    \item $\hbar$: Reduced Planck constant.
    \begin{itemize}
        \item Used in equations: \eqref{TimeIndependentSchrodingerEquation}, \eqref{EnergyEigenvalues_En}, \eqref{EnergyEigenstates_Psix}, \eqref{EnergySpacing_DeltaE}, \eqref{RelativeEnergySpacing}, \eqref{WKB_WaveFunction}, \eqref{WKB_ProbabilityDensity_BeforeAveraging}.
    \end{itemize}
    \item $\omega$: Angular frequency of oscillation.
    \begin{itemize}
        \item Used in equations: \eqref{DE_SimpleHarmonicMotion}, \eqref{GeneralSolution_xt}, \eqref{Velocity_vt}, \eqref{ClassicalEnergy_Expanded}, \eqref{ClassicalEnergy_Simplified}, \eqref{Velocity_vx_EnergyConservation}, \eqref{EnergyEigenvalues_En}, \eqref{EnergyEigenstates_Psix}, \eqref{EnergySpacing_DeltaE}, \eqref{RelativeEnergySpacing}.
    \end{itemize}
\end{itemize}

@bhpayne
Copy link
Member Author

bhpayne commented Jan 26, 2025

Resulting Latex

\documentclass{article}
\usepackage{amsmath}

% margins of 1 inch:
\setlength{\topmargin}{-.5in}
\setlength{\textheight}{9in}
\setlength{\oddsidemargin}{0in}
\setlength{\textwidth}{6.5in}

\usepackage[pdftex]{hyperref} % hyperlink equation and bibliographic citations

\title{The Correspondence Principle and the Harmonic Oscillator}
\author{Ben Payne, with Gemini 2.0}
\date{\today}

\begin{document}

\maketitle
\begin{abstract}
    From Gemini 2.0 Flash
\end{abstract}

\section{Variables used in Derivation}

\begin{itemize}
    \item $A$: Amplitude of oscillation.
    \begin{itemize}
        \item Used in equations: \eqref{GeneralSolution_xt}, \eqref{Velocity_vt}, \eqref{ClassicalEnergy_Expanded}, \eqref{ClassicalEnergy_Simplified}, \eqref{ClassicalProbabilityDensity_Px}.
    \end{itemize}
    \item $A_n$: Classical turning points for the $n$-th quantum state.
    \begin{itemize}
        \item Used in equations: \eqref{AveragedWKB_ProbabilityDensity}.
    \end{itemize}
    \item $C, C', C''$: Normalization constants for classical probability density.
    \begin{itemize}
        \item Used in equations: \eqref{ClassicalProbabilityDensity_Px}, \eqref{AveragedWKB_ProbabilityDensity}, \eqref{WKB_WaveFunction}, \eqref{WKB_ProbabilityDensity_BeforeAveraging}.
    \end{itemize}
    \item $E_{classical}$: Total energy of the classical harmonic oscillator.
    \begin{itemize}
        \item Used in equations: \eqref{ClassicalEnergy_Expanded}, \eqref{ClassicalEnergy_Simplified}, \eqref{Velocity_vx_EnergyConservation}.
    \end{itemize}
    \item $E_n$: Energy eigenvalue for the $n$-th quantum state of the quantum harmonic oscillator.
    \begin{itemize}
        \item Used in equations: \eqref{TimeIndependentSchrodingerEquation}, \eqref{EnergyEigenvalues_En}, \eqref{EnergySpacing_DeltaE}, \eqref{RelativeEnergySpacing}, \eqref{AveragedWKB_ProbabilityDensity}.
    \end{itemize}
    \item $F$: Force.
    \begin{itemize}
        \item Used in equations: None (mentioned in text before equation \eqref{DE_SimpleHarmonicMotion}).
    \end{itemize}
    \item $H_n(u)$: Hermite polynomial of order $n$ with argument $u$.
    \begin{itemize}
        \item Used in equations: \eqref{EnergyEigenstates_Psix}.
    \end{itemize}
    \item $k$: Spring constant.
    \begin{itemize}
        \item Used in equations: \eqref{DE_SimpleHarmonicMotion}, \eqref{ClassicalEnergy_Expanded}, \eqref{ClassicalEnergy_Simplified}, \eqref{Velocity_vx_EnergyConservation}, \eqref{TimeIndependentSchrodingerEquation}, \eqref{AveragedWKB_ProbabilityDensity}.
    \end{itemize}
    \item $m$: Mass of the particle.
    \begin{itemize}
        \item Used in equations: \eqref{DE_SimpleHarmonicMotion}, \eqref{ClassicalEnergy_Expanded}, \eqref{ClassicalEnergy_Simplified}, \eqref{Velocity_vx_EnergyConservation}, \eqref{TimeIndependentSchrodingerEquation}, \eqref{EnergyEigenstates_Psix}, \eqref{WKB_WaveFunction}, \eqref{WKB_ProbabilityDensity_BeforeAveraging}.
    \end{itemize}
    \item $n$: Quantum number.
    \begin{itemize}
        \item Used in equations: \eqref{EnergyEigenvalues_En}, \eqref{EnergyEigenstates_Psix}, \eqref{RelativeEnergySpacing}, \eqref{AveragedWKB_ProbabilityDensity}.
    \end{itemize}
    \item $P_{classical}(x)$: Classical probability density as a function of position $x$.
    \begin{itemize}
        \item Used in equations: \eqref{ClassicalProbabilityDensity_Proportionality}, \eqref{ClassicalProbabilityDensity_Px}.
    \end{itemize}
    \item $p(x)$: Classical momentum as a function of position $x$.
    \begin{itemize}
        \item Used in equations: \eqref{WKB_WaveFunction}, \eqref{WKB_ProbabilityDensity_BeforeAveraging}, \eqref{AveragedWKB_ProbabilityDensity}.
    \end{itemize}
    \item $T$: Kinetic energy.
    \begin{itemize}
        \item Used in equations: \eqref{ClassicalEnergy_Expanded}.
    \end{itemize}
    \item $t$: Time.
    \begin{itemize}
        \item Used in equations: \eqref{GeneralSolution_xt}, \eqref{Velocity_vt}, \eqref{ClassicalEnergy_Expanded}.
    \end{itemize}
    \item $u$: Argument of Hermite polynomials, $u = \sqrt{\frac{m\omega}{\hbar}} x$.
    \begin{itemize}
        \item Used in equations: \eqref{EnergyEigenstates_Psix}.
    \end{itemize}
    \item $V$: Potential energy.
    \begin{itemize}
        \item Used in equations: \eqref{ClassicalEnergy_Expanded}, \eqref{TimeIndependentSchrodingerEquation}, \eqref{AveragedWKB_ProbabilityDensity}.
    \end{itemize}
    \item $v$: Velocity of the particle.
    \begin{itemize}
        \item Used in equations: \eqref{Velocity_vt}, \eqref{ClassicalEnergy_Expanded}, \eqref{ClassicalProbabilityDensity_Proportionality}, \eqref{ClassicalProbabilityDensity_Px}, \eqref{Velocity_vx_EnergyConservation}.
    \end{itemize}
    \item $x$: Position of the particle.
    \begin{itemize}
        \item Used in equations: \eqref{DE_SimpleHarmonicMotion}, \eqref{GeneralSolution_xt}, \eqref{Velocity_vt}, \eqref{ClassicalEnergy_Expanded}, \eqref{Velocity_vx_EnergyConservation}, \eqref{ClassicalProbabilityDensity_Px}, \eqref{TimeIndependentSchrodingerEquation}, \eqref{EnergyEigenstates_Psix}, \eqref{WKB_WaveFunction}, \eqref{WKB_ProbabilityDensity_BeforeAveraging}.
    \end{itemize}
    \item $\ddot{x}$: Acceleration of the particle.
    \begin{itemize}
        \item Used in equations: \eqref{DE_SimpleHarmonicMotion}.
    \end{itemize}
    \item $\Delta E$: Energy spacing between adjacent energy levels in the quantum harmonic oscillator.
    \begin{itemize}
        \item Used in equations: \eqref{EnergySpacing_DeltaE}, \eqref{RelativeEnergySpacing}.
    \end{itemize}
    \item $\phi$: Phase constant in the general solution for simple harmonic motion.
    \begin{itemize}
        \item Used in equations: \eqref{GeneralSolution_xt}, \eqref{Velocity_vt}, \eqref{ClassicalEnergy_Expanded}, \eqref{ClassicalEnergy_Simplified}.
    \end{itemize}
    \item $\psi_n(x)$: Energy eigenstate wave function for the $n$-th state of the quantum harmonic oscillator as a function of position $x$.
    \begin{itemize}
        \item Used in equations: \eqref{TimeIndependentSchrodingerEquation}, \eqref{EnergyEigenstates_Psix}.
    \end{itemize}
    \item $\psi_{WKB}(x)$: WKB approximation of the wave function.
    \begin{itemize}
        \item Used in equations: \eqref{WKB_WaveFunction}, \eqref{WKB_ProbabilityDensity_BeforeAveraging}.
    \end{itemize}
    \item $\hbar$: Reduced Planck constant.
    \begin{itemize}
        \item Used in equations: \eqref{TimeIndependentSchrodingerEquation}, \eqref{EnergyEigenvalues_En}, \eqref{EnergyEigenstates_Psix}, \eqref{EnergySpacing_DeltaE}, \eqref{RelativeEnergySpacing}, \eqref{WKB_WaveFunction}, \eqref{WKB_ProbabilityDensity_BeforeAveraging}.
    \end{itemize}
    \item $\omega$: Angular frequency of oscillation.
    \begin{itemize}
        \item Used in equations: \eqref{DE_SimpleHarmonicMotion}, \eqref{GeneralSolution_xt}, \eqref{Velocity_vt}, \eqref{ClassicalEnergy_Expanded}, \eqref{ClassicalEnergy_Simplified}, \eqref{Velocity_vx_EnergyConservation}, \eqref{EnergyEigenvalues_En}, \eqref{EnergyEigenstates_Psix}, \eqref{EnergySpacing_DeltaE}, \eqref{RelativeEnergySpacing}.
    \end{itemize}
\end{itemize}

\section{Introduction}

The harmonic oscillator is a fundamental system in physics, serving as a model for a wide range of phenomena from molecular vibrations to oscillations in electromagnetic fields.  It is particularly insightful in the context of the Correspondence Principle, which states that the predictions of quantum mechanics must agree with classical mechanics in the limit of large quantum numbers or high energies. This principle provides a crucial bridge between the seemingly disparate worlds of classical and quantum physics.  In this document, we will mathematically explore how the quantum harmonic oscillator (QHO) approaches the classical harmonic oscillator (CHO) in the classical limit, demonstrating the Correspondence Principle.

\section{Classical Harmonic Oscillator (CHO) Review}

Consider a particle of mass $m$ attached to a spring with spring constant $k$. The potential energy of this system is given by $V(x) = \frac{1}{2}kx^2$.

\subsection{Equations of Motion and Energy}
From Newton's second law, $F = ma = m\ddot{x} = -kx$. This leads to the differential equation for simple harmonic motion:
\begin{equation} \label{DE_SimpleHarmonicMotion}
\ddot{x} + \omega^2 x = 0,
\end{equation}
where $\omega = \sqrt{\frac{k}{m}}$ is the angular frequency of oscillation. The general solution is:
\begin{equation} \label{GeneralSolution_xt}
x(t) = A \cos(\omega t - \phi),
\end{equation}
where $A$ is the amplitude and $\phi$ is the phase constant, determined by initial conditions.

The velocity is given by:
\begin{equation} \label{Velocity_vt}
v(t) = \dot{x}(t) = -A\omega \sin(\omega t - \phi).
\end{equation}

The total energy of the classical harmonic oscillator is conserved and is the sum of kinetic and potential energies:
\begin{equation} \label{ClassicalEnergy_Expanded}
E_{classical} = T + V = \frac{1}{2}mv^2 + \frac{1}{2}kx^2 = \frac{1}{2}m(A\omega \sin(\omega t - \phi))^2 + \frac{1}{2}k(A \cos(\omega t - \phi))^2.
\end{equation}
Using $k = m\omega^2$ and $\sin^2(\theta) + \cos^2(\theta) = 1$, we get:
\begin{equation} \label{ClassicalEnergy_Simplified}
E_{classical} = \frac{1}{2}mA^2\omega^2 (\sin^2(\omega t - \phi) + \cos^2(\omega t - \phi)) = \frac{1}{2}mA^2\omega^2 = \frac{1}{2}kA^2.
\end{equation}
Classically, energy $E_{classical}$ can take any continuous non-negative value.

\subsection{Classical Probability Density}
The probability of finding the particle in a small interval $dx$ at position $x$ is proportional to the time spent in that interval. Since the particle moves slower near the turning points ($x = \pm A$) and faster near the equilibrium position ($x = 0$), the probability density is inversely proportional to the speed of the particle:
\begin{equation} \label{ClassicalProbabilityDensity_Proportionality}
P_{classical}(x) \propto \frac{1}{|v(x)|}.
\end{equation}
From energy conservation, $E_{classical} = \frac{1}{2}mv^2 + \frac{1}{2}kx^2$, we can express the speed as:
\begin{equation} \label{Velocity_vx_EnergyConservation}
v(x) = \pm \sqrt{\frac{2}{m}(E_{classical} - \frac{1}{2}kx^2)} = \pm \omega \sqrt{A^2 - x^2}.
\end{equation}
Thus, the classical probability density is:
\begin{equation} \label{ClassicalProbabilityDensity_Px}
P_{classical}(x) = \frac{C}{|v(x)|} = \frac{C}{\omega \sqrt{A^2 - x^2}} = \frac{C'}{\sqrt{A^2 - x^2}}, \quad \text{for } |x| \leq A,
\end{equation}
and $P_{classical}(x) = 0$ for $|x| > A$.  $C'$ is a normalization constant.

\section{Quantum Harmonic Oscillator (QHO) Review}

In quantum mechanics, the harmonic oscillator is described by the time-independent Schrödinger equation:
\begin{equation} \label{TimeIndependentSchrodingerEquation}
-\frac{\hbar^2}{2m} \frac{d^2\psi_n(x)}{dx^2} + \frac{1}{2}kx^2 \psi_n(x) = E_n \psi_n(x).
\end{equation}
The solutions to this equation are the energy eigenstates $\psi_n(x)$ with corresponding energy eigenvalues $E_n$.

\subsection{Energy Eigenvalues}
The energy eigenvalues for the quantum harmonic oscillator are quantized and given by:
\begin{equation} \label{EnergyEigenvalues_En}
E_n = \left(n + \frac{1}{2}\right)\hbar\omega, \quad n = 0, 1, 2, \ldots
\end{equation}
where $n$ is the quantum number and $\hbar$ is the reduced Planck constant.  Energy levels are discrete and equally spaced with spacing $\hbar\omega$.

\subsection{Energy Eigenstates and Probability Density}
The energy eigenstates $\psi_n(x)$ are given by:
\begin{equation} \label{EnergyEigenstates_Psix}
\psi_n(x) = \frac{1}{\sqrt{2^n n!}} \left(\frac{m\omega}{\pi\hbar}\right)^{1/4} e^{-\frac{m\omega x^2}{2\hbar}} H_n\left(\sqrt{\frac{m\omega}{\hbar}} x\right),
\end{equation}
where $H_n(u)$ are the Hermite polynomials. The probability density of finding the particle at position $x$ in the $n$-th energy state is given by $|\psi_n(x)|^2$.

For low quantum numbers (small $n$), the probability density $|\psi_n(x)|^2$ is significantly different from the classical probability density. For instance, for $n=0$, $|\psi_0(x)|^2$ is a Gaussian centered at $x=0$, with the highest probability at the equilibrium position, unlike the classical case where the probability is lowest at the equilibrium.  Furthermore, there is a non-zero probability of finding the particle outside the classical turning points, which is forbidden in classical mechanics.

\section{Correspondence Principle and the Harmonic Oscillator}

The Correspondence Principle states that in the limit of large quantum numbers ($n \rightarrow \infty$), quantum mechanics should reproduce classical mechanics. Let's examine how this principle applies to the harmonic oscillator.

\subsection{Energy Levels}
The energy levels for the QHO are $E_n = (n + \frac{1}{2})\hbar\omega$. For large $n$, we can approximate $E_n \approx n\hbar\omega$. The energy spacing between adjacent levels is:
\begin{equation} \label{EnergySpacing_DeltaE}
\Delta E = E_{n+1} - E_n = \left((n+1) + \frac{1}{2}\right)\hbar\omega - \left(n + \frac{1}{2}\right)\hbar\omega = \hbar\omega.
\end{equation}
This spacing $\hbar\omega$ is constant and independent of $n$. However, consider the \textit{relative} spacing compared to the energy itself:
\begin{equation} \label{RelativeEnergySpacing}
\frac{\Delta E}{E_n} = \frac{\hbar\omega}{(n + \frac{1}{2})\hbar\omega} = \frac{1}{n + \frac{1}{2}} \approx \frac{1}{n} \quad \text{for large } n.
\end{equation}
As $n \rightarrow \infty$, the relative spacing $\frac{\Delta E}{E_n} \rightarrow 0$. This means that for very high energy levels, the energy spectrum becomes effectively continuous, resembling the classical case where energy can take continuous values.  In the classical limit of high energy, the quantization of energy becomes less apparent.

\subsection{Probability Density}
Now let's consider the probability density. For large $n$, the wave function $\psi_n(x)$ becomes highly oscillatory, and the Hermite polynomials $H_n(u)$ also exhibit oscillatory behavior within the classical allowed region and decay exponentially outside.  While a rigorous mathematical proof is complex, we can argue qualitatively and consider some properties:

1. \textbf{Classical Turning Points:} The classical turning points are given by $V(x) = E_{classical}$, so $\frac{1}{2}kx^2 = E_{classical}$. For the $n$-th quantum state, we set $E_{classical} = E_n = (n + \frac{1}{2})\hbar\omega$. Thus, the classical turning points are $x = \pm A_n$, where $\frac{1}{2}kA_n^2 = E_n$, so $A_n = \sqrt{\frac{2E_n}{k}} = \sqrt{\frac{2(n + \frac{1}{2})\hbar\omega}{m\omega^2}} = \sqrt{\frac{(2n+1)\hbar}{m\omega}}$.

2. \textbf{Oscillatory Nature of $\psi_n(x)$ for large $n$:} For large $n$, the Hermite polynomials $H_n(u)$ behave approximately as sinusoidal functions within the classically allowed region.  The square of the wave function, $|\psi_n(x)|^2$, will thus exhibit rapid oscillations within $|x| \leq A_n$.

3. \textbf{Averaging over Oscillations:} If we average the quantum probability density $|\psi_n(x)|^2$ over a spatial interval that is large compared to the wavelength of the oscillations but small compared to the overall extent of the wave function, we find that the averaged quantum probability density approaches the classical probability density.  Intuitively, the rapid oscillations in $|\psi_n(x)|^2$ for large $n$ average out to a smoother distribution that resembles the classical distribution, which is peaked at the turning points and minimum at the equilibrium position.

To illustrate this more explicitly, consider the WKB approximation, which is valid for large quantum numbers. In the WKB approximation, the wave function in the classically allowed region oscillates as:
\begin{equation} \label{WKB_WaveFunction}
\psi_{WKB}(x) \approx \frac{C}{\sqrt{p(x)}} \cos\left(\frac{1}{\hbar} \int^x p(x') dx' - \frac{\pi}{4}\right),
\end{equation}
where $p(x) = \sqrt{2m(E - V(x))} = \sqrt{2m(E - \frac{1}{2}kx^2)}$ is the classical momentum.  The probability density is then approximately:
\begin{equation} \label{WKB_ProbabilityDensity_BeforeAveraging}
|\psi_{WKB}(x)|^2 \approx \frac{C^2}{p(x)} \cos^2\left(\frac{1}{\hbar} \int^x p(x') dx' - \frac{\pi}{4}\right).
\end{equation}
Averaging the rapidly oscillating $\cos^2$ term over a cycle gives $\langle \cos^2 \rangle = \frac{1}{2}$.  Thus, the average probability density becomes:
\begin{equation} \label{AveragedWKB_ProbabilityDensity}
\langle |\psi_{WKB}(x)|^2 \rangle \approx \frac{C^2}{2p(x)} = \frac{C'^2}{\sqrt{E - \frac{1}{2}kx^2}} = \frac{C''}{\sqrt{A_n^2 - x^2}},
\end{equation}
where we used $E = E_n = \frac{1}{2}kA_n^2$. This form is proportional to the classical probability density $P_{classical}(x) = \frac{C'}{\sqrt{A^2 - x^2}}$.  Hence, in the limit of large quantum numbers, the average quantum probability density approaches the classical probability density within the classically allowed region.

Furthermore, the quantum probability of finding the particle outside the classical turning points (tunneling) becomes relatively less significant for large $n$. The exponentially decaying tails of $\psi_n(x)$ beyond $x = \pm A_n$ become less prominent compared to the oscillatory behavior within the classical region as $n$ increases.

\section{Conclusion}

We have shown how the quantum harmonic oscillator approaches the classical harmonic oscillator in the limit of large quantum numbers, in accordance with the Correspondence Principle.

\begin{itemize}
    \item \textbf{Energy Levels:} The energy levels of the QHO are quantized, but for large quantum numbers $n$, the relative spacing between energy levels becomes vanishingly small compared to the energy itself.  This makes the energy spectrum effectively continuous, as in classical mechanics.
    \item \textbf{Probability Density:} While the quantum probability density for low $n$ is significantly different from the classical probability density, for large $n$, the average quantum probability density (obtained by averaging over oscillations or using approximations like WKB) approaches the classical probability density within the classically allowed region.  The oscillatory nature of the quantum wave function for large $n$ and its localization within the classical boundaries are key to this correspondence.
\end{itemize}

The harmonic oscillator provides a clear and important example of the Correspondence Principle, demonstrating how quantum mechanics smoothly transitions to classical mechanics in the appropriate limit. This principle is not only a historical cornerstone in the development of quantum theory but also a valuable tool for understanding the relationship between quantum and classical descriptions of physical systems.

\end{document}

@bhpayne
Copy link
Member Author

bhpayne commented Jan 28, 2025

Manually split out expressions with multiple = present. Also manually converted inline expressions to stand-alone expressions.

Resulting Latex:

\documentclass{article}
\usepackage{amsmath}

% margins of 1 inch:
\setlength{\topmargin}{-.5in}
\setlength{\textheight}{9in}
\setlength{\oddsidemargin}{0in}
\setlength{\textwidth}{6.5in}

\usepackage[pdftex]{hyperref} % hyperlink equation and bibliographic citations

\title{The Correspondence Principle and the Harmonic Oscillator}
\author{Ben Payne, with Gemini 2.0}
\date{\today}

\begin{document}

\maketitle
\begin{abstract}
    From Gemini 2.0 Flash Thinking Experimental
\end{abstract}


\section{Introduction}

The harmonic oscillator is a fundamental system in physics, serving as a model for a wide range of phenomena from molecular vibrations to oscillations in electromagnetic fields.  It is particularly insightful in the context of the Correspondence Principle, which states that the predictions of quantum mechanics must agree with classical mechanics in the limit of large quantum numbers or high energies. This principle provides a crucial bridge between the seemingly disparate worlds of classical and quantum physics.  In this document, we will mathematically explore how the quantum harmonic oscillator (QHO) approaches the classical harmonic oscillator (CHO) in the classical limit, demonstrating the Correspondence Principle.

\section{Classical Harmonic Oscillator (CHO) Review}

Consider a particle of mass $m$ attached to a spring with spring constant $k$. The potential energy of this system is given by 
\begin{equation}
    V(x) = \frac{1}{2}kx^2
\end{equation}

\subsection{Equations of Motion and Energy}
From Newton's second law, 
\begin{equation}
    F = ma = m\ddot{x} = -kx
\end{equation}
This leads to the differential equation for simple harmonic motion:
\begin{equation} \label{DE_SimpleHarmonicMotion}
\ddot{x} + \omega^2 x = 0,
\end{equation}
where 
\begin{equation}
    \omega = \sqrt{\frac{k}{m}}
\end{equation}
is the angular frequency of oscillation. The general solution is:
\begin{equation} \label{GeneralSolution_xt}
x(t) = A \cos(\omega t - \phi),
\end{equation}
where $A$ is the amplitude and $\phi$ is the phase constant, determined by initial conditions.

The velocity is given by:
\begin{equation} \label{Velocity_vt}
v(t) = \dot{x}(t) = -A\omega \sin(\omega t - \phi).
\end{equation}

The total energy of the classical harmonic oscillator is conserved and is the sum of kinetic and potential energies:
\begin{equation} \label{ClassicalEnergy_Expanded}
E_{classical} = T + V
\end{equation}

\begin{equation}
    E_{classical} = \frac{1}{2}mv^2 + \frac{1}{2}kx^2
\end{equation}

\begin{equation}
    E_{classical} = \frac{1}{2}m(A\omega \sin(\omega t - \phi))^2 + \frac{1}{2}k(A \cos(\omega t - \phi))^2.
\end{equation}

Using 
\begin{equation}
    k = m\omega^2
\end{equation}
and 
\begin{equation}
    \sin^2(\theta) + \cos^2(\theta) = 1
\end{equation}
we get
\begin{equation}
E_{classical} = \frac{1}{2}mA^2\omega^2 (\sin^2(\omega t - \phi) + \cos^2(\omega t - \phi))
\end{equation}

\begin{equation}
E_{classical} = \frac{1}{2}mA^2\omega^2 
\end{equation}

\begin{equation}
E_{classical} = \frac{1}{2}kA^2
\end{equation}
Classically, energy $E_{classical}$ can take any continuous non-negative value.

\subsection{Classical Probability Density}
The probability of finding the particle in a small interval $dx$ at position $x$ is proportional to the time spent in that interval. Since the particle moves slower near the turning points ($x = \pm A$) and faster near the equilibrium position ($x = 0$), the probability density is inversely proportional to the speed of the particle:
\begin{equation} \label{ClassicalProbabilityDensity_Proportionality}
P_{classical}(x) \propto \frac{1}{|v(x)|}.
\end{equation}
From energy conservation, 
\begin{equation}
    E_{classical} = \frac{1}{2}mv^2 + \frac{1}{2}kx^2
\end{equation}
we can express the speed as:
\begin{equation} \label{Velocity_vx_EnergyConservation}
v(x) = \pm \sqrt{\frac{2}{m}(E_{classical} - \frac{1}{2}kx^2)} 
\end{equation}
\begin{equation} 
v(x) =  \pm \omega \sqrt{A^2 - x^2}
\end{equation}
Thus, the classical probability density is:
\begin{equation}
P_{classical}(x) = \frac{C}{|v(x)|}\quad \text{for } |x| \leq A
\end{equation}
\begin{equation}
P_{classical}(x) = \frac{C}{\omega \sqrt{A^2 - x^2}}\quad \text{for } |x| \leq A
\end{equation}
\begin{equation}
P_{classical}(x) = \frac{C'}{\sqrt{A^2 - x^2}} \quad \text{for } |x| \leq A
\end{equation}
and $P_{classical}(x) = 0$ for $|x| > A$.  $C'$ is a normalization constant.

\section{Quantum Harmonic Oscillator (QHO) Review}

In quantum mechanics, the harmonic oscillator is described by the time-independent Schrödinger equation:
\begin{equation} \label{TimeIndependentSchrodingerEquation}
-\frac{\hbar^2}{2m} \frac{d^2\psi_n(x)}{dx^2} + \frac{1}{2}kx^2 \psi_n(x) = E_n \psi_n(x).
\end{equation}
The solutions to this equation are the energy eigenstates $\psi_n(x)$ with corresponding energy eigenvalues $E_n$.

\subsection{Energy Eigenvalues}
The energy eigenvalues for the quantum harmonic oscillator are quantized and given by:
\begin{equation} \label{EnergyEigenvalues_En}
E_n = \left(n + \frac{1}{2}\right)\hbar\omega, \quad n = 0, 1, 2, \ldots
\end{equation}
where $n$ is the quantum number and $\hbar$ is the reduced Planck constant.  Energy levels are discrete and equally spaced with spacing $\hbar\omega$.

\subsection{Energy Eigenstates and Probability Density}
The energy eigenstates $\psi_n(x)$ are given by:
\begin{equation} \label{EnergyEigenstates_Psix}
\psi_n(x) = \frac{1}{\sqrt{2^n n!}} \left(\frac{m\omega}{\pi\hbar}\right)^{1/4} e^{-\frac{m\omega x^2}{2\hbar}} H_n\left(\sqrt{\frac{m\omega}{\hbar}} x\right),
\end{equation}
where $H_n(u)$ are the Hermite polynomials. The probability density of finding the particle at position $x$ in the $n$-th energy state is given by $|\psi_n(x)|^2$.

For low quantum numbers (small $n$), the probability density $|\psi_n(x)|^2$ is significantly different from the classical probability density. For instance, for $n=0$, $|\psi_0(x)|^2$ is a Gaussian centered at $x=0$, with the highest probability at the equilibrium position, unlike the classical case where the probability is lowest at the equilibrium.  Furthermore, there is a non-zero probability of finding the particle outside the classical turning points, which is forbidden in classical mechanics.

\section{Correspondence Principle and the Harmonic Oscillator}

The Correspondence Principle states that in the limit of large quantum numbers ($n \rightarrow \infty$), quantum mechanics should reproduce classical mechanics. Let's examine how this principle applies to the harmonic oscillator.

\subsection{Energy Levels}
The energy levels for the QHO are 
\begin{equation}
E_n = (n + \frac{1}{2})\hbar\omega    
\end{equation}

For large $n$, we can approximate $E_n \approx n\hbar\omega$. The energy spacing between adjacent levels is:
\begin{equation}
\Delta E = E_{n+1} - E_n
\end{equation}
\begin{equation} 
\Delta E = \left((n+1) + \frac{1}{2}\right)\hbar\omega - \left(n + \frac{1}{2}\right)\hbar\omega
\end{equation}
\begin{equation} \label{EnergySpacing_DeltaE}
\Delta E = \hbar\omega.
\end{equation}
This spacing $\hbar\omega$ is constant and independent of $n$. However, consider the \textit{relative} spacing compared to the energy itself:
\begin{equation} \label{RelativeEnergySpacing}
\frac{\Delta E}{E_n} = \frac{\hbar\omega}{(n + \frac{1}{2})\hbar\omega} 
\end{equation}
\begin{equation} 
\frac{\Delta E}{E_n} = \frac{1}{n + \frac{1}{2}}
\end{equation}
\begin{equation} 
\frac{\Delta E}{E_n}  \approx \frac{1}{n} \quad \text{for large } n.
\end{equation}
As $n \rightarrow \infty$, the relative spacing $\frac{\Delta E}{E_n} \rightarrow 0$. This means that for very high energy levels, the energy spectrum becomes effectively continuous, resembling the classical case where energy can take continuous values.  In the classical limit of high energy, the quantization of energy becomes less apparent.

\subsection{Probability Density}
Now let's consider the probability density. For large $n$, the wave function $\psi_n(x)$ becomes highly oscillatory, and the Hermite polynomials $H_n(u)$ also exhibit oscillatory behavior within the classical allowed region and decay exponentially outside.  While a rigorous mathematical proof is complex, we can argue qualitatively and consider some properties:

1. \textbf{Classical Turning Points:} The classical turning points are given by 
\begin{equation}
    V(x) = E_{classical}
\end{equation}
so 
\begin{equation}
    \frac{1}{2}kx^2 = E_{classical}
\end{equation}
For the $n$-th quantum state, we set 
\begin{equation}
    E_{classical} = E_n
\end{equation}
\begin{equation}
    E_{classical} = (n + \frac{1}{2})\hbar\omega
\end{equation}
 Thus, the classical turning points are $x = \pm A_n$, where 
 \begin{equation}
     \frac{1}{2}kA_n^2 = E_n
 \end{equation}
 so $A_n = \sqrt{\frac{2E_n}{k}} = \sqrt{\frac{2(n + \frac{1}{2})\hbar\omega}{m\omega^2}} = \sqrt{\frac{(2n+1)\hbar}{m\omega}}$.

2. \textbf{Oscillatory Nature of $\psi_n(x)$ for large $n$:} For large $n$, the Hermite polynomials $H_n(u)$ behave approximately as sinusoidal functions within the classically allowed region.  The square of the wave function, $|\psi_n(x)|^2$, will thus exhibit rapid oscillations within $|x| \leq A_n$.

3. \textbf{Averaging over Oscillations:} If we average the quantum probability density $|\psi_n(x)|^2$ over a spatial interval that is large compared to the wavelength of the oscillations but small compared to the overall extent of the wave function, we find that the averaged quantum probability density approaches the classical probability density.  Intuitively, the rapid oscillations in $|\psi_n(x)|^2$ for large $n$ average out to a smoother distribution that resembles the classical distribution, which is peaked at the turning points and minimum at the equilibrium position.

To illustrate this more explicitly, consider the WKB approximation, which is valid for large quantum numbers. In the WKB approximation, the wave function in the classically allowed region oscillates as:
\begin{equation} \label{WKB_WaveFunction}
\psi_{WKB}(x) \approx \frac{C}{\sqrt{p(x)}} \cos\left(\frac{1}{\hbar} \int^x p(x') dx' - \frac{\pi}{4}\right),
\end{equation}
where 
\begin{equation}
p(x) = \sqrt{2m(E - V(x))}
\end{equation}
\begin{equation}
p(x)= \sqrt{2m(E - \frac{1}{2}kx^2)}    
\end{equation}
 is the classical momentum.  The probability density is then approximately:
\begin{equation} \label{WKB_ProbabilityDensity_BeforeAveraging}
|\psi_{WKB}(x)|^2 \approx \frac{C^2}{p(x)} \cos^2\left(\frac{1}{\hbar} \int^x p(x') dx' - \frac{\pi}{4}\right).
\end{equation}
Averaging the rapidly oscillating $\cos^2$ term over a cycle gives $\langle \cos^2 \rangle = \frac{1}{2}$.  Thus, the average probability density becomes:
\begin{equation} \label{AveragedWKB_ProbabilityDensity}
\langle |\psi_{WKB}(x)|^2 \rangle \approx \frac{C^2}{2p(x)} 
\end{equation}
\begin{equation} 
\langle |\psi_{WKB}(x)|^2 \rangle = \frac{C'^2}{\sqrt{E - \frac{1}{2}kx^2}} 
\end{equation}
\begin{equation} 
\langle |\psi_{WKB}(x)|^2 \rangle = \frac{C''}{\sqrt{A_n^2 - x^2}},
\end{equation}
where we used 
\begin{equation}
    E = E_n = \frac{1}{2}kA_n^2
\end{equation}
This form is proportional to the classical probability density 
\begin{equation}
    P_{classical}(x) = \frac{C'}{\sqrt{A^2 - x^2}}
\end{equation}
Hence, in the limit of large quantum numbers, the average quantum probability density approaches the classical probability density within the classically allowed region.

Furthermore, the quantum probability of finding the particle outside the classical turning points (tunneling) becomes relatively less significant for large $n$. The exponentially decaying tails of $\psi_n(x)$ beyond $x = \pm A_n$ become less prominent compared to the oscillatory behavior within the classical region as $n$ increases.

\section{Conclusion}

We have shown how the quantum harmonic oscillator approaches the classical harmonic oscillator in the limit of large quantum numbers, in accordance with the Correspondence Principle.

\begin{itemize}
    \item \textbf{Energy Levels:} The energy levels of the QHO are quantized, but for large quantum numbers $n$, the relative spacing between energy levels becomes vanishingly small compared to the energy itself.  This makes the energy spectrum effectively continuous, as in classical mechanics.
    \item \textbf{Probability Density:} While the quantum probability density for low $n$ is significantly different from the classical probability density, for large $n$, the average quantum probability density (obtained by averaging over oscillations or using approximations like WKB) approaches the classical probability density within the classically allowed region.  The oscillatory nature of the quantum wave function for large $n$ and its localization within the classical boundaries are key to this correspondence.
\end{itemize}

The harmonic oscillator provides a clear and important example of the Correspondence Principle, demonstrating how quantum mechanics smoothly transitions to classical mechanics in the appropriate limit. This principle is not only a historical cornerstone in the development of quantum theory but also a valuable tool for understanding the relationship between quantum and classical descriptions of physical systems.

\end{document}

@bhpayne
Copy link
Member Author

bhpayne commented Jan 28, 2025

prompt

Latex is a software system for typesetting documents. LaTeX is widely used for the communication and publication of scientific documents and technical note-taking in many fields, owing partially to its support for complex mathematical notation.

The Latex file below contains equations, some of which do not have a label. Write out the file and ensure each equation has one label. The label that is added should be descriptive of the equation and contain no spaces, only underscores.

Labels for equations should be unique within the document.

Don't provide any commentary. Just write the revised Latex file content as your answer.

\documentclass{article}
\usepackage{amsmath}

% margins of 1 inch:
\setlength{\topmargin}{-.5in}
\setlength{\textheight}{9in}
\setlength{\oddsidemargin}{0in}
\setlength{\textwidth}{6.5in}

\usepackage[pdftex]{hyperref} % hyperlink equation and bibliographic citations

\title{The Correspondence Principle and the Harmonic Oscillator}
\author{Ben Payne, with Gemini 2.0}
\date{\today}

\begin{document}

\maketitle
\begin{abstract}
    From Gemini 2.0 Flash Thinking Experimental
\end{abstract}


\section{Introduction}

The harmonic oscillator is a fundamental system in physics, serving as a model for a wide range of phenomena from molecular vibrations to oscillations in electromagnetic fields.  It is particularly insightful in the context of the Correspondence Principle, which states that the predictions of quantum mechanics must agree with classical mechanics in the limit of large quantum numbers or high energies. This principle provides a crucial bridge between the seemingly disparate worlds of classical and quantum physics.  In this document, we will mathematically explore how the quantum harmonic oscillator (QHO) approaches the classical harmonic oscillator (CHO) in the classical limit, demonstrating the Correspondence Principle.

\section{Classical Harmonic Oscillator (CHO) Review}

Consider a particle of mass $m$ attached to a spring with spring constant $k$. The potential energy of this system is given by 
\begin{equation}
    V(x) = \frac{1}{2}kx^2
\end{equation}

\subsection{Equations of Motion and Energy}
From Newton's second law, 
\begin{equation}
    F = ma = m\ddot{x} = -kx
\end{equation}
This leads to the differential equation for simple harmonic motion:
\begin{equation} \label{DE_SimpleHarmonicMotion}
\ddot{x} + \omega^2 x = 0,
\end{equation}
where 
\begin{equation}
    \omega = \sqrt{\frac{k}{m}}
\end{equation}
is the angular frequency of oscillation. The general solution is:
\begin{equation} \label{GeneralSolution_xt}
x(t) = A \cos(\omega t - \phi),
\end{equation}
where $A$ is the amplitude and $\phi$ is the phase constant, determined by initial conditions.

The velocity is given by:
\begin{equation} \label{Velocity_vt}
v(t) = \dot{x}(t) = -A\omega \sin(\omega t - \phi).
\end{equation}

The total energy of the classical harmonic oscillator is conserved and is the sum of kinetic and potential energies:
\begin{equation} \label{ClassicalEnergy_Expanded}
E_{classical} = T + V
\end{equation}

\begin{equation}
    E_{classical} = \frac{1}{2}mv^2 + \frac{1}{2}kx^2
\end{equation}

\begin{equation}
    E_{classical} = \frac{1}{2}m(A\omega \sin(\omega t - \phi))^2 + \frac{1}{2}k(A \cos(\omega t - \phi))^2.
\end{equation}

Using 
\begin{equation}
    k = m\omega^2
\end{equation}
and 
\begin{equation}
    \sin^2(\theta) + \cos^2(\theta) = 1
\end{equation}
we get
\begin{equation}
E_{classical} = \frac{1}{2}mA^2\omega^2 (\sin^2(\omega t - \phi) + \cos^2(\omega t - \phi))
\end{equation}

\begin{equation}
E_{classical} = \frac{1}{2}mA^2\omega^2 
\end{equation}

\begin{equation}
E_{classical} = \frac{1}{2}kA^2
\end{equation}
Classically, energy $E_{classical}$ can take any continuous non-negative value.

\subsection{Classical Probability Density}
The probability of finding the particle in a small interval $dx$ at position $x$ is proportional to the time spent in that interval. Since the particle moves slower near the turning points ($x = \pm A$) and faster near the equilibrium position ($x = 0$), the probability density is inversely proportional to the speed of the particle:
\begin{equation} \label{ClassicalProbabilityDensity_Proportionality}
P_{classical}(x) \propto \frac{1}{|v(x)|}.
\end{equation}
From energy conservation, 
\begin{equation}
    E_{classical} = \frac{1}{2}mv^2 + \frac{1}{2}kx^2
\end{equation}
we can express the speed as:
\begin{equation} \label{Velocity_vx_EnergyConservation}
v(x) = \pm \sqrt{\frac{2}{m}(E_{classical} - \frac{1}{2}kx^2)} 
\end{equation}
\begin{equation} 
v(x) =  \pm \omega \sqrt{A^2 - x^2}
\end{equation}
Thus, the classical probability density is:
\begin{equation}
P_{classical}(x) = \frac{C}{|v(x)|}\quad \text{for } |x| \leq A
\end{equation}
\begin{equation}
P_{classical}(x) = \frac{C}{\omega \sqrt{A^2 - x^2}}\quad \text{for } |x| \leq A
\end{equation}
\begin{equation}
P_{classical}(x) = \frac{C'}{\sqrt{A^2 - x^2}} \quad \text{for } |x| \leq A
\end{equation}
and $P_{classical}(x) = 0$ for $|x| > A$.  $C'$ is a normalization constant.

\section{Quantum Harmonic Oscillator (QHO) Review}

In quantum mechanics, the harmonic oscillator is described by the time-independent Schrödinger equation:
\begin{equation} \label{TimeIndependentSchrodingerEquation}
-\frac{\hbar^2}{2m} \frac{d^2\psi_n(x)}{dx^2} + \frac{1}{2}kx^2 \psi_n(x) = E_n \psi_n(x).
\end{equation}
The solutions to this equation are the energy eigenstates $\psi_n(x)$ with corresponding energy eigenvalues $E_n$.

\subsection{Energy Eigenvalues}
The energy eigenvalues for the quantum harmonic oscillator are quantized and given by:
\begin{equation} \label{EnergyEigenvalues_En}
E_n = \left(n + \frac{1}{2}\right)\hbar\omega, \quad n = 0, 1, 2, \ldots
\end{equation}
where $n$ is the quantum number and $\hbar$ is the reduced Planck constant.  Energy levels are discrete and equally spaced with spacing $\hbar\omega$.

\subsection{Energy Eigenstates and Probability Density}
The energy eigenstates $\psi_n(x)$ are given by:
\begin{equation} \label{EnergyEigenstates_Psix}
\psi_n(x) = \frac{1}{\sqrt{2^n n!}} \left(\frac{m\omega}{\pi\hbar}\right)^{1/4} e^{-\frac{m\omega x^2}{2\hbar}} H_n\left(\sqrt{\frac{m\omega}{\hbar}} x\right),
\end{equation}
where $H_n(u)$ are the Hermite polynomials. The probability density of finding the particle at position $x$ in the $n$-th energy state is given by $|\psi_n(x)|^2$.

For low quantum numbers (small $n$), the probability density $|\psi_n(x)|^2$ is significantly different from the classical probability density. For instance, for $n=0$, $|\psi_0(x)|^2$ is a Gaussian centered at $x=0$, with the highest probability at the equilibrium position, unlike the classical case where the probability is lowest at the equilibrium.  Furthermore, there is a non-zero probability of finding the particle outside the classical turning points, which is forbidden in classical mechanics.

\section{Correspondence Principle and the Harmonic Oscillator}

The Correspondence Principle states that in the limit of large quantum numbers ($n \rightarrow \infty$), quantum mechanics should reproduce classical mechanics. Let's examine how this principle applies to the harmonic oscillator.

\subsection{Energy Levels}
The energy levels for the QHO are 
\begin{equation}
E_n = (n + \frac{1}{2})\hbar\omega    
\end{equation}

For large $n$, we can approximate $E_n \approx n\hbar\omega$. The energy spacing between adjacent levels is:
\begin{equation}
\Delta E = E_{n+1} - E_n
\end{equation}
\begin{equation} 
\Delta E = \left((n+1) + \frac{1}{2}\right)\hbar\omega - \left(n + \frac{1}{2}\right)\hbar\omega
\end{equation}
\begin{equation} \label{EnergySpacing_DeltaE}
\Delta E = \hbar\omega.
\end{equation}
This spacing $\hbar\omega$ is constant and independent of $n$. However, consider the \textit{relative} spacing compared to the energy itself:
\begin{equation} \label{RelativeEnergySpacing}
\frac{\Delta E}{E_n} = \frac{\hbar\omega}{(n + \frac{1}{2})\hbar\omega} 
\end{equation}
\begin{equation} 
\frac{\Delta E}{E_n} = \frac{1}{n + \frac{1}{2}}
\end{equation}
\begin{equation} 
\frac{\Delta E}{E_n}  \approx \frac{1}{n} \quad \text{for large } n.
\end{equation}
As $n \rightarrow \infty$, the relative spacing $\frac{\Delta E}{E_n} \rightarrow 0$. This means that for very high energy levels, the energy spectrum becomes effectively continuous, resembling the classical case where energy can take continuous values.  In the classical limit of high energy, the quantization of energy becomes less apparent.

\subsection{Probability Density}
Now let's consider the probability density. For large $n$, the wave function $\psi_n(x)$ becomes highly oscillatory, and the Hermite polynomials $H_n(u)$ also exhibit oscillatory behavior within the classical allowed region and decay exponentially outside.  While a rigorous mathematical proof is complex, we can argue qualitatively and consider some properties:

1. \textbf{Classical Turning Points:} The classical turning points are given by 
\begin{equation}
    V(x) = E_{classical}
\end{equation}
so 
\begin{equation}
    \frac{1}{2}kx^2 = E_{classical}
\end{equation}
For the $n$-th quantum state, we set 
\begin{equation}
    E_{classical} = E_n
\end{equation}
\begin{equation}
    E_{classical} = (n + \frac{1}{2})\hbar\omega
\end{equation}
 Thus, the classical turning points are $x = \pm A_n$, where 
 \begin{equation}
     \frac{1}{2}kA_n^2 = E_n
 \end{equation}
 so $A_n = \sqrt{\frac{2E_n}{k}} = \sqrt{\frac{2(n + \frac{1}{2})\hbar\omega}{m\omega^2}} = \sqrt{\frac{(2n+1)\hbar}{m\omega}}$.

2. \textbf{Oscillatory Nature of $\psi_n(x)$ for large $n$:} For large $n$, the Hermite polynomials $H_n(u)$ behave approximately as sinusoidal functions within the classically allowed region.  The square of the wave function, $|\psi_n(x)|^2$, will thus exhibit rapid oscillations within $|x| \leq A_n$.

3. \textbf{Averaging over Oscillations:} If we average the quantum probability density $|\psi_n(x)|^2$ over a spatial interval that is large compared to the wavelength of the oscillations but small compared to the overall extent of the wave function, we find that the averaged quantum probability density approaches the classical probability density.  Intuitively, the rapid oscillations in $|\psi_n(x)|^2$ for large $n$ average out to a smoother distribution that resembles the classical distribution, which is peaked at the turning points and minimum at the equilibrium position.

To illustrate this more explicitly, consider the WKB approximation, which is valid for large quantum numbers. In the WKB approximation, the wave function in the classically allowed region oscillates as:
\begin{equation} \label{WKB_WaveFunction}
\psi_{WKB}(x) \approx \frac{C}{\sqrt{p(x)}} \cos\left(\frac{1}{\hbar} \int^x p(x') dx' - \frac{\pi}{4}\right),
\end{equation}
where 
\begin{equation}
p(x) = \sqrt{2m(E - V(x))}
\end{equation}
\begin{equation}
p(x)= \sqrt{2m(E - \frac{1}{2}kx^2)}    
\end{equation}
 is the classical momentum.  The probability density is then approximately:
\begin{equation} \label{WKB_ProbabilityDensity_BeforeAveraging}
|\psi_{WKB}(x)|^2 \approx \frac{C^2}{p(x)} \cos^2\left(\frac{1}{\hbar} \int^x p(x') dx' - \frac{\pi}{4}\right).
\end{equation}
Averaging the rapidly oscillating $\cos^2$ term over a cycle gives $\langle \cos^2 \rangle = \frac{1}{2}$.  Thus, the average probability density becomes:
\begin{equation} \label{AveragedWKB_ProbabilityDensity}
\langle |\psi_{WKB}(x)|^2 \rangle \approx \frac{C^2}{2p(x)} 
\end{equation}
\begin{equation} 
\langle |\psi_{WKB}(x)|^2 \rangle = \frac{C'^2}{\sqrt{E - \frac{1}{2}kx^2}} 
\end{equation}
\begin{equation} 
\langle |\psi_{WKB}(x)|^2 \rangle = \frac{C''}{\sqrt{A_n^2 - x^2}},
\end{equation}
where we used 
\begin{equation}
    E = E_n = \frac{1}{2}kA_n^2
\end{equation}
This form is proportional to the classical probability density 
\begin{equation}
    P_{classical}(x) = \frac{C'}{\sqrt{A^2 - x^2}}
\end{equation}
Hence, in the limit of large quantum numbers, the average quantum probability density approaches the classical probability density within the classically allowed region.

Furthermore, the quantum probability of finding the particle outside the classical turning points (tunneling) becomes relatively less significant for large $n$. The exponentially decaying tails of $\psi_n(x)$ beyond $x = \pm A_n$ become less prominent compared to the oscillatory behavior within the classical region as $n$ increases.

\section{Conclusion}

We have shown how the quantum harmonic oscillator approaches the classical harmonic oscillator in the limit of large quantum numbers, in accordance with the Correspondence Principle.

\begin{itemize}
    \item \textbf{Energy Levels:} The energy levels of the QHO are quantized, but for large quantum numbers $n$, the relative spacing between energy levels becomes vanishingly small compared to the energy itself.  This makes the energy spectrum effectively continuous, as in classical mechanics.
    \item \textbf{Probability Density:} While the quantum probability density for low $n$ is significantly different from the classical probability density, for large $n$, the average quantum probability density (obtained by averaging over oscillations or using approximations like WKB) approaches the classical probability density within the classically allowed region.  The oscillatory nature of the quantum wave function for large $n$ and its localization within the classical boundaries are key to this correspondence.
\end{itemize}

The harmonic oscillator provides a clear and important example of the Correspondence Principle, demonstrating how quantum mechanics smoothly transitions to classical mechanics in the appropriate limit. This principle is not only a historical cornerstone in the development of quantum theory but also a valuable tool for understanding the relationship between quantum and classical descriptions of physical systems.

\end{document}

thoughts

The user wants me to add labels to all equations in the provided LaTeX file.
The labels should be descriptive, contain no spaces (only underscores), and be unique.
I will go through the LaTeX file and identify each equation environment.
For each equation environment, I will create a descriptive label and add it using \label{}.
I need to ensure that all labels are unique and descriptive.

Plan:

  1. Read through the LaTeX file.
  2. Identify all equation environments (equation, align, gather, etc. - in this case, only equation is used).
  3. For each equation, create a descriptive label.
    • The label should reflect the content of the equation.
    • Use underscores instead of spaces.
    • Ensure uniqueness.
  4. Insert the \label{} command inside each equation environment, before the equation.
  5. Output the modified LaTeX file content.

Let's start step-by-step.

Equation 1: V(x) = \frac{1}{2}kx^2 - Potential energy of harmonic oscillator - PotentialEnergy_HarmonicOscillator
Equation 2: F = ma = m\ddot{x} = -kx - Newton's second law for harmonic oscillator - NewtonsSecondLaw_HarmonicOscillator
Equation 3: \ddot{x} + \omega^2 x = 0 - Differential equation for simple harmonic motion - Already labeled as DE_SimpleHarmonicMotion - Keep it.
Equation 4: \omega = \sqrt{\frac{k}{m}} - Angular frequency - AngularFrequency_Omega
Equation 5: x(t) = A \cos(\omega t - \phi) - General solution for x(t) - Already labeled as GeneralSolution_xt - Keep it.
Equation 6: v(t) = \dot{x}(t) = -A\omega \sin(\omega t - \phi) - Velocity v(t) - Already labeled as Velocity_vt - Keep it.
Equation 7: E_{classical} = T + V - Classical energy definition - Already labeled as ClassicalEnergy_Expanded - Keep it.
Equation 8: E_{classical} = \frac{1}{2}mv^2 + \frac{1}{2}kx^2 - Classical energy in terms of v and x - ClassicalEnergy_vx
Equation 9: E_{classical} = \frac{1}{2}m(A\omega \sin(\omega t - \phi))^2 + \frac{1}{2}k(A \cos(\omega t - \phi))^2 - Classical energy expanded with x(t) and v(t) - ClassicalEnergy_Expanded_xt_vt
Equation 10: k = m\omega^2 - Relation between k, m, and omega - k_equals_m_omega_squared
Equation 11: \sin^2(\theta) + \cos^2(\theta) = 1 - Trigonometric identity - TrigonometricIdentity_sin_cos_squared
Equation 12: E_{classical} = \frac{1}{2}mA^2\omega^2 (\sin^2(\omega t - \phi) + \cos^2(\omega t - \phi)) - Classical energy with trig identity - ClassicalEnergy_TrigIdentity
Equation 13: E_{classical} = \frac{1}{2}mA^2\omega^2 - Simplified classical energy - ClassicalEnergy_Simplified_mA_omega
Equation 14: E_{classical} = \frac{1}{2}kA^2 - Simplified classical energy - ClassicalEnergy_Simplified_kA
Equation 15: P_{classical}(x) \propto \frac{1}{|v(x)|} - Classical probability density proportionality - Already labeled as ClassicalProbabilityDensity_Proportionality - Keep it.
Equation 16: E_{classical} = \frac{1}{2}mv^2 + \frac{1}{2}kx^2 - Energy conservation again - Already labeled in previous section. No need to label again if it is the same equation and context. But it is used to derive next equation, so label it - EnergyConservation_ClassicalProbability
Equation 17: v(x) = \pm \sqrt{\frac{2}{m}(E_{classical} - \frac{1}{2}kx^2)} - Velocity v(x) from energy conservation - Already labeled as Velocity_vx_EnergyConservation - Keep it.
Equation 18: v(x) = \pm \omega \sqrt{A^2 - x^2} - Velocity v(x) simplified - Velocity_vx_Simplified
Equation 19: P_{classical}(x) = \frac{C}{|v(x)|}\quad \text{for } |x| \leq A - Classical probability density with constant C - ClassicalProbabilityDensity_C_over_v
Equation 20: P_{classical}(x) = \frac{C}{\omega \sqrt{A^2 - x^2}}\quad \text{for } |x| \leq A - Classical probability density with constant C and omega - ClassicalProbabilityDensity_C_omega_sqrt
Equation 21: P_{classical}(x) = \frac{C'}{\sqrt{A^2 - x^2}} \quad \text{for } |x| \leq A - Classical probability density with constant C' - ClassicalProbabilityDensity_Cprime_sqrt
Equation 22: -\frac{\hbar^2}{2m} \frac{d^2\psi_n(x)}{dx^2} + \frac{1}{2}kx^2 \psi_n(x) = E_n \psi_n(x) - Time-independent Schrodinger equation - Already labeled as TimeIndependentSchrodingerEquation - Keep it.
Equation 23: E_n = \left(n + \frac{1}{2}\right)\hbar\omega, \quad n = 0, 1, 2, \ldots - Energy eigenvalues for QHO - Already labeled as EnergyEigenvalues_En - Keep it.
Equation 24: \psi_n(x) = \frac{1}{\sqrt{2^n n!}} \left(\frac{m\omega}{\pi\hbar}\right)^{1/4} e^{-\frac{m\omega x^2}{2\hbar}} H_n\left(\sqrt{\frac{m\omega}{\hbar}} x\right) - Energy eigenstates for QHO - Already labeled as EnergyEigenstates_Psix - Keep it.
Equation 25: E_n = (n + \frac{1}{2})\hbar\omega - Energy levels for QHO - Already labeled in previous section. Label it again for this section - EnergyLevels_QHO
Equation 26: \Delta E = E_{n+1} - E_n - Energy spacing definition - EnergySpacing_Definition
Equation 27: \Delta E = \left((n+1) + \frac{1}{2}\right)\hbar\omega - \left(n + \frac{1}{2}\right)\hbar\omega - Energy spacing expanded - EnergySpacing_Expanded
Equation 28: \Delta E = \hbar\omega - Energy spacing simplified - Already labeled as EnergySpacing_DeltaE - Keep it.
Equation 29: \frac{\Delta E}{E_n} = \frac{\hbar\omega}{(n + \frac{1}{2})\hbar\omega} - Relative energy spacing - Already labeled as RelativeEnergySpacing - Keep it.
Equation 30: \frac{\Delta E}{E_n} = \frac{1}{n + \frac{1}{2}} - Relative energy spacing simplified - RelativeEnergySpacing_Simplified
Equation 31: \frac{\Delta E}{E_n} \approx \frac{1}{n} \quad \text{for large } n - Relative energy spacing approximation - RelativeEnergySpacing_Approximation
Equation 32: V(x) = E_{classical} - Classical turning points condition - ClassicalTurningPoints_Condition
Equation 33: \frac{1}{2}kx^2 = E_{classical} - Classical turning points condition expanded - ClassicalTurningPoints_Condition_Expanded
Equation 34: E_{classical} = E_n - Classical energy equals quantum energy - ClassicalEnergy_equals_QuantumEnergy
Equation 35: E_{classical} = (n + \frac{1}{2})\hbar\omega - Classical energy equals quantum energy expanded - ClassicalEnergy_equals_QuantumEnergy_Expanded
Equation 36: \frac{1}{2}kA_n^2 = E_n - Amplitude in terms of En - Amplitude_An_En
Equation 37: A_n = \sqrt{\frac{2E_n}{k}} = \sqrt{\frac{2(n + \frac{1}{2})\hbar\omega}{m\omega^2}} = \sqrt{\frac{(2n+1)\hbar}{m\omega}} - Amplitude An expanded - Amplitude_An_Expanded
Equation 38: \psi_{WKB}(x) \approx \frac{C}{\sqrt{p(x)}} \cos\left(\frac{1}{\hbar} \int^x p(x') dx' - \frac{\pi}{4}\right) - WKB wave function - Already labeled as WKB_WaveFunction - Keep it.
Equation 39: p(x) = \sqrt{2m(E - V(x))} - Classical momentum definition - ClassicalMomentum_Definition
Equation 40: p(x)= \sqrt{2m(E - \frac{1}{2}kx^2)} - Classical momentum for harmonic oscillator - ClassicalMomentum_HarmonicOscillator
Equation 41: |\psi_{WKB}(x)|^2 \approx \frac{C^2}{p(x)} \cos^2\left(\frac{1}{\hbar} \int^x p(x') dx' - \frac{\pi}{4}\right) - WKB probability density before averaging - Already labeled as WKB_ProbabilityDensity_BeforeAveraging - Keep it.
Equation 42: \langle |\psi_{WKB}(x)|^2 \rangle \approx \frac{C^2}{2p(x)} - Averaged WKB probability density - Already labeled as AveragedWKB_ProbabilityDensity - Keep it.
Equation 43: \langle |\psi_{WKB}(x)|^2 \rangle = \frac{C'^2}{\sqrt{E - \frac{1}{2}kx^2}} - Averaged WKB probability density with C' - AveragedWKB_ProbabilityDensity_Cprime
Equation 44: \langle |\psi_{WKB}(x)|^2 \rangle = \frac{C''}{\sqrt{A_n^2 - x^2}} - Averaged WKB probability density with C'' - AveragedWKB_ProbabilityDensity_Cdoubleprime
Equation 45: E = E_n = \frac{1}{2}kA_n^2 - Energy relation for WKB - EnergyRelation_WKB
Equation 46: P_{classical}(x) = \frac{C'}{\sqrt{A^2 - x^2}} - Classical probability density again - Already labeled in previous section. No need to label again if it is the same equation and context. But it is referred to compare to WKB result, so label it - ClassicalProbabilityDensity_Reiterated

I have now created labels for all equations that were missing labels. I will go through the LaTeX file and insert these labels.

response

\documentclass{article}
\usepackage{amsmath}

% margins of 1 inch:
\setlength{\topmargin}{-.5in}
\setlength{\textheight}{9in}
\setlength{\oddsidemargin}{0in}
\setlength{\textwidth}{6.5in}

\usepackage[pdftex]{hyperref} % hyperlink equation and bibliographic citations

\title{The Correspondence Principle and the Harmonic Oscillator}
\author{Ben Payne, with Gemini 2.0}
\date{\today}

\begin{document}

\maketitle
\begin{abstract}
    From Gemini 2.0 Flash Thinking Experimental
\end{abstract}


\section{Introduction}

The harmonic oscillator is a fundamental system in physics, serving as a model for a wide range of phenomena from molecular vibrations to oscillations in electromagnetic fields.  It is particularly insightful in the context of the Correspondence Principle, which states that the predictions of quantum mechanics must agree with classical mechanics in the limit of large quantum numbers or high energies. This principle provides a crucial bridge between the seemingly disparate worlds of classical and quantum physics.  In this document, we will mathematically explore how the quantum harmonic oscillator (QHO) approaches the classical harmonic oscillator (CHO) in the classical limit, demonstrating the Correspondence Principle.

\section{Classical Harmonic Oscillator (CHO) Review}

Consider a particle of mass $m$ attached to a spring with spring constant $k$. The potential energy of this system is given by
\begin{equation} \label{PotentialEnergy_HarmonicOscillator}
    V(x) = \frac{1}{2}kx^2
\end{equation}

\subsection{Equations of Motion and Energy}
From Newton's second law,
\begin{equation} \label{NewtonsSecondLaw_HarmonicOscillator}
    F = ma = m\ddot{x} = -kx
\end{equation}
This leads to the differential equation for simple harmonic motion:
\begin{equation} \label{DE_SimpleHarmonicMotion}
\ddot{x} + \omega^2 x = 0,
\end{equation}
where
\begin{equation} \label{AngularFrequency_Omega}
    \omega = \sqrt{\frac{k}{m}}
\end{equation}
is the angular frequency of oscillation. The general solution is:
\begin{equation} \label{GeneralSolution_xt}
x(t) = A \cos(\omega t - \phi),
\end{equation}
where $A$ is the amplitude and $\phi$ is the phase constant, determined by initial conditions.

The velocity is given by:
\begin{equation} \label{Velocity_vt}
v(t) = \dot{x}(t) = -A\omega \sin(\omega t - \phi).
\end{equation}

The total energy of the classical harmonic oscillator is conserved and is the sum of kinetic and potential energies:
\begin{equation} \label{ClassicalEnergy_Expanded}
E_{classical} = T + V
\end{equation}

\begin{equation} \label{ClassicalEnergy_vx}
    E_{classical} = \frac{1}{2}mv^2 + \frac{1}{2}kx^2
\end{equation}

\begin{equation} \label{ClassicalEnergy_Expanded_xt_vt}
    E_{classical} = \frac{1}{2}m(A\omega \sin(\omega t - \phi))^2 + \frac{1}{2}k(A \cos(\omega t - \phi))^2.
\end{equation}

Using
\begin{equation} \label{k_equals_m_omega_squared}
    k = m\omega^2
\end{equation}
and
\begin{equation} \label{TrigonometricIdentity_sin_cos_squared}
    \sin^2(\theta) + \cos^2(\theta) = 1
\end{equation}
we get
\begin{equation} \label{ClassicalEnergy_TrigIdentity}
E_{classical} = \frac{1}{2}mA^2\omega^2 (\sin^2(\omega t - \phi) + \cos^2(\omega t - \phi))
\end{equation}

\begin{equation} \label{ClassicalEnergy_Simplified_mA_omega}
E_{classical} = \frac{1}{2}mA^2\omega^2
\end{equation}

\begin{equation} \label{ClassicalEnergy_Simplified_kA}
E_{classical} = \frac{1}{2}kA^2
\end{equation}
Classically, energy $E_{classical}$ can take any continuous non-negative value.

\subsection{Classical Probability Density}
The probability of finding the particle in a small interval $dx$ at position $x$ is proportional to the time spent in that interval. Since the particle moves slower near the turning points ($x = \pm A$) and faster near the equilibrium position ($x = 0$), the probability density is inversely proportional to the speed of the particle:
\begin{equation} \label{ClassicalProbabilityDensity_Proportionality}
P_{classical}(x) \propto \frac{1}{|v(x)|}.
\end{equation}
From energy conservation,
\begin{equation} \label{EnergyConservation_ClassicalProbability}
    E_{classical} = \frac{1}{2}mv^2 + \frac{1}{2}kx^2
\end{equation}
we can express the speed as:
\begin{equation} \label{Velocity_vx_EnergyConservation}
v(x) = \pm \sqrt{\frac{2}{m}(E_{classical} - \frac{1}{2}kx^2)}
\end{equation}
\begin{equation} \label{Velocity_vx_Simplified}
v(x) =  \pm \omega \sqrt{A^2 - x^2}
\end{equation}
Thus, the classical probability density is:
\begin{equation} \label{ClassicalProbabilityDensity_C_over_v}
P_{classical}(x) = \frac{C}{|v(x)|}\quad \text{for } |x| \leq A
\end{equation}
\begin{equation} \label{ClassicalProbabilityDensity_C_omega_sqrt}
P_{classical}(x) = \frac{C}{\omega \sqrt{A^2 - x^2}}\quad \text{for } |x| \leq A
\end{equation}
\begin{equation} \label{ClassicalProbabilityDensity_Cprime_sqrt}
P_{classical}(x) = \frac{C'}{\sqrt{A^2 - x^2}} \quad \text{for } |x| \leq A
\end{equation}
and $P_{classical}(x) = 0$ for $|x| > A$.  $C'$ is a normalization constant.

\section{Quantum Harmonic Oscillator (QHO) Review}

In quantum mechanics, the harmonic oscillator is described by the time-independent Schrödinger equation:
\begin{equation} \label{TimeIndependentSchrodingerEquation}
-\frac{\hbar^2}{2m} \frac{d^2\psi_n(x)}{dx^2} + \frac{1}{2}kx^2 \psi_n(x) = E_n \psi_n(x).
\end{equation}
The solutions to this equation are the energy eigenstates $\psi_n(x)$ with corresponding energy eigenvalues $E_n$.

\subsection{Energy Eigenvalues}
The energy eigenvalues for the quantum harmonic oscillator are quantized and given by:
\begin{equation} \label{EnergyEigenvalues_En}
E_n = \left(n + \frac{1}{2}\right)\hbar\omega, \quad n = 0, 1, 2, \ldots
\end{equation}
where $n$ is the quantum number and $\hbar$ is the reduced Planck constant.  Energy levels are discrete and equally spaced with spacing $\hbar\omega$.

\subsection{Energy Eigenstates and Probability Density}
The energy eigenstates $\psi_n(x)$ are given by:
\begin{equation} \label{EnergyEigenstates_Psix}
\psi_n(x) = \frac{1}{\sqrt{2^n n!}} \left(\frac{m\omega}{\pi\hbar}\right)^{1/4} e^{-\frac{m\omega x^2}{2\hbar}} H_n\left(\sqrt{\frac{m\omega}{\hbar}} x\right),
\end{equation}
where $H_n(u)$ are the Hermite polynomials. The probability density of finding the particle at position $x$ in the $n$-th energy state is given by $|\psi_n(x)|^2$.

For low quantum numbers (small $n$), the probability density $|\psi_n(x)|^2$ is significantly different from the classical probability density. For instance, for $n=0$, $|\psi_0(x)|^2$ is a Gaussian centered at $x=0$, with the highest probability at the equilibrium position, unlike the classical case where the probability is lowest at the equilibrium.  Furthermore, there is a non-zero probability of finding the particle outside the classical turning points, which is forbidden in classical mechanics.

\section{Correspondence Principle and the Harmonic Oscillator}

The Correspondence Principle states that in the limit of large quantum numbers ($n \rightarrow \infty$), quantum mechanics should reproduce classical mechanics. Let's examine how this principle applies to the harmonic oscillator.

\subsection{Energy Levels}
The energy levels for the QHO are
\begin{equation} \label{EnergyLevels_QHO}
E_n = (n + \frac{1}{2})\hbar\omega
\end{equation}

For large $n$, we can approximate $E_n \approx n\hbar\omega$. The energy spacing between adjacent levels is:
\begin{equation} \label{EnergySpacing_Definition}
\Delta E = E_{n+1} - E_n
\end{equation}
\begin{equation} \label{EnergySpacing_Expanded}
\Delta E = \left((n+1) + \frac{1}{2}\right)\hbar\omega - \left(n + \frac{1}{2}\right)\hbar\omega
\end{equation}
\begin{equation} \label{EnergySpacing_DeltaE}
\Delta E = \hbar\omega.
\end{equation}
This spacing $\hbar\omega$ is constant and independent of $n$. However, consider the \textit{relative} spacing compared to the energy itself:
\begin{equation} \label{RelativeEnergySpacing}
\frac{\Delta E}{E_n} = \frac{\hbar\omega}{(n + \frac{1}{2})\hbar\omega}
\end{equation}
\begin{equation} \label{RelativeEnergySpacing_Simplified}
\frac{\Delta E}{E_n} = \frac{1}{n + \frac{1}{2}}
\end{equation}
\begin{equation} \label{RelativeEnergySpacing_Approximation}
\frac{\Delta E}{E_n}  \approx \frac{1}{n} \quad \text{for large } n.
\end{equation}
As $n \rightarrow \infty$, the relative spacing $\frac{\Delta E}{E_n} \rightarrow 0$. This means that for very high energy levels, the energy spectrum becomes effectively continuous, resembling the classical case where energy can take continuous values.  In the classical limit of high energy, the quantization of energy becomes less apparent.

\subsection{Probability Density}
Now let's consider the probability density. For large $n$, the wave function $\psi_n(x)$ becomes highly oscillatory, and the Hermite polynomials $H_n(u)$ also exhibit oscillatory behavior within the classical allowed region and decay exponentially outside.  While a rigorous mathematical proof is complex, we can argue qualitatively and consider some properties:

1. \textbf{Classical Turning Points:} The classical turning points are given by
\begin{equation} \label{ClassicalTurningPoints_Condition}
    V(x) = E_{classical}
\end{equation}
so
\begin{equation} \label{ClassicalTurningPoints_Condition_Expanded}
    \frac{1}{2}kx^2 = E_{classical}
\end{equation}
For the $n$-th quantum state, we set
\begin{equation} \label{ClassicalEnergy_equals_QuantumEnergy}
    E_{classical} = E_n
\end{equation}
\begin{equation} \label{ClassicalEnergy_equals_QuantumEnergy_Expanded}
    E_{classical} = (n + \frac{1}{2})\hbar\omega
\end{equation}
 Thus, the classical turning points are $x = \pm A_n$, where
 \begin{equation} \label{Amplitude_An_En}
     \frac{1}{2}kA_n^2 = E_n
 \end{equation}
 so
 \begin{equation} \label{Amplitude_An_Expanded}
     A_n = \sqrt{\frac{2E_n}{k}} = \sqrt{\frac{2(n + \frac{1}{2})\hbar\omega}{m\omega^2}} = \sqrt{\frac{(2n+1)\hbar}{m\omega}}.
 \end{equation}

2. \textbf{Oscillatory Nature of $\psi_n(x)$ for large $n$:} For large $n$, the Hermite polynomials $H_n(u)$ behave approximately as sinusoidal functions within the classically allowed region.  The square of the wave function, $|\psi_n(x)|^2$, will thus exhibit rapid oscillations within $|x| \leq A_n$.

3. \textbf{Averaging over Oscillations:} If we average the quantum probability density $|\psi_n(x)|^2$ over a spatial interval that is large compared to the wavelength of the oscillations but small compared to the overall extent of the wave function, we find that the averaged quantum probability density approaches the classical probability density.  Intuitively, the rapid oscillations in $|\psi_n(x)|^2$ for large $n$ average out to a smoother distribution that resembles the classical distribution, which is peaked at the turning points and minimum at the equilibrium position.

To illustrate this more explicitly, consider the WKB approximation, which is valid for large quantum numbers. In the WKB approximation, the wave function in the classically allowed region oscillates as:
\begin{equation} \label{WKB_WaveFunction}
\psi_{WKB}(x) \approx \frac{C}{\sqrt{p(x)}} \cos\left(\frac{1}{\hbar} \int^x p(x') dx' - \frac{\pi}{4}\right),
\end{equation}
where
\begin{equation} \label{ClassicalMomentum_Definition}
p(x) = \sqrt{2m(E - V(x))}
\end{equation}
\begin{equation} \label{ClassicalMomentum_HarmonicOscillator}
p(x)= \sqrt{2m(E - \frac{1}{2}kx^2)}
\end{equation}
 is the classical momentum.  The probability density is then approximately:
\begin{equation} \label{WKB_ProbabilityDensity_BeforeAveraging}
|\psi_{WKB}(x)|^2 \approx \frac{C^2}{p(x)} \cos^2\left(\frac{1}{\hbar} \int^x p(x') dx' - \frac{\pi}{4}\right).
\end{equation}
Averaging the rapidly oscillating $\cos^2$ term over a cycle gives $\langle \cos^2 \rangle = \frac{1}{2}$.  Thus, the average probability density becomes:
\begin{equation} \label{AveragedWKB_ProbabilityDensity}
\langle |\psi_{WKB}(x)|^2 \rangle \approx \frac{C^2}{2p(x)}
\end{equation}
\begin{equation} \label{AveragedWKB_ProbabilityDensity_Cprime}
\langle |\psi_{WKB}(x)|^2 \rangle = \frac{C'^2}{\sqrt{E - \frac{1}{2}kx^2}}
\end{equation}
\begin{equation} \label{AveragedWKB_ProbabilityDensity_Cdoubleprime}
\langle |\psi_{WKB}(x)|^2 \rangle = \frac{C''}{\sqrt{A_n^2 - x^2}},
\end{equation}
where we used
\begin{equation} \label{EnergyRelation_WKB}
    E = E_n = \frac{1}{2}kA_n^2
\end{equation}
This form is proportional to the classical probability density
\begin{equation} \label{ClassicalProbabilityDensity_Reiterated}
    P_{classical}(x) = \frac{C'}{\sqrt{A^2 - x^2}}
\end{equation}
Hence, in the limit of large quantum numbers, the average quantum probability density approaches the classical probability density within the classically allowed region.

Furthermore, the quantum probability of finding the particle outside the classical turning points (tunneling) becomes relatively less significant for large $n$. The exponentially decaying tails of $\psi_n(x)$ beyond $x = \pm A_n$ become less prominent compared to the oscillatory behavior within the classical region as $n$ increases.

\section{Conclusion}

We have shown how the quantum harmonic oscillator approaches the classical harmonic oscillator in the limit of large quantum numbers, in accordance with the Correspondence Principle.

\begin{itemize}
    \item \textbf{Energy Levels:} The energy levels of the QHO are quantized, but for large quantum numbers $n$, the relative spacing between energy levels becomes vanishingly small compared to the energy itself.  This makes the energy spectrum effectively continuous, as in classical mechanics.
    \item \textbf{Probability Density:} While the quantum probability density for low $n$ is significantly different from the classical probability density, for large $n$, the average quantum probability density (obtained by averaging over oscillations or using approximations like WKB) approaches the classical probability density within the classically allowed region.  The oscillatory nature of the quantum wave function for large $n$ and its localization within the classical boundaries are key to this correspondence.
\end{itemize}

The harmonic oscillator provides a clear and important example of the Correspondence Principle, demonstrating how quantum mechanics smoothly transitions to classical mechanics in the appropriate limit. This principle is not only a historical cornerstone in the development of quantum theory but also a valuable tool for understanding the relationship between quantum and classical descriptions of physical systems.

\end{document}

@bhpayne
Copy link
Member Author

bhpayne commented Jan 28, 2025

prompt

A derivation in mathematical Physics consists of a sequence of steps. Each step relates mathematical equations.

Modify the Latex file below by adding explanatory text about each step of the derivation. Document the mathematical transformations that relate each equation in the file to other equations in the file.

The following Latex file contains a derivation with many steps. Sometimes a previous expression is referenced implicitly. Edit the following text to manually insert references where appropriate. For example replace text like "substitute the forces" with "substitute Eq.~\ref{} to get". Be sure to reference the relevant equation labels so that Latex can compile the document.

Use the Latex equation labels to reference equations.

Don't provide any commentary. Just write the revised Latex file content as your answer.

\documentclass{article}
\usepackage{amsmath}

% margins of 1 inch:
\setlength{\topmargin}{-.5in}
\setlength{\textheight}{9in}
\setlength{\oddsidemargin}{0in}
\setlength{\textwidth}{6.5in}

\usepackage[pdftex]{hyperref} % hyperlink equation and bibliographic citations

\title{The Correspondence Principle and the Harmonic Oscillator}
\author{Ben Payne, with Gemini 2.0}
\date{\today}

\begin{document}

\maketitle
\begin{abstract}
    From Gemini 2.0 Flash Thinking Experimental
\end{abstract}


\section{Introduction}

The harmonic oscillator is a fundamental system in physics, serving as a model for a wide range of phenomena from molecular vibrations to oscillations in electromagnetic fields.  It is particularly insightful in the context of the Correspondence Principle, which states that the predictions of quantum mechanics must agree with classical mechanics in the limit of large quantum numbers or high energies. This principle provides a crucial bridge between the seemingly disparate worlds of classical and quantum physics.  In this document, we will mathematically explore how the quantum harmonic oscillator (QHO) approaches the classical harmonic oscillator (CHO) in the classical limit, demonstrating the Correspondence Principle.

\section{Classical Harmonic Oscillator (CHO) Review}

Consider a particle of mass $m$ attached to a spring with spring constant $k$. The potential energy of this system is given by
\begin{equation} \label{PotentialEnergy_HarmonicOscillator}
    V(x) = \frac{1}{2}kx^2
\end{equation}

\subsection{Equations of Motion and Energy}
From Newton's second law,
\begin{equation} \label{NewtonsSecondLaw_HarmonicOscillator}
    F = ma = m\ddot{x} = -kx
\end{equation}
This leads to the differential equation for simple harmonic motion:
\begin{equation} \label{DE_SimpleHarmonicMotion}
\ddot{x} + \omega^2 x = 0,
\end{equation}
where
\begin{equation} \label{AngularFrequency_Omega}
    \omega = \sqrt{\frac{k}{m}}
\end{equation}
is the angular frequency of oscillation. The general solution is:
\begin{equation} \label{GeneralSolution_xt}
x(t) = A \cos(\omega t - \phi),
\end{equation}
where $A$ is the amplitude and $\phi$ is the phase constant, determined by initial conditions.

The velocity is given by:
\begin{equation} \label{Velocity_vt}
v(t) = \dot{x}(t) = -A\omega \sin(\omega t - \phi).
\end{equation}

The total energy of the classical harmonic oscillator is conserved and is the sum of kinetic and potential energies:
\begin{equation} \label{ClassicalEnergy_Expanded}
E_{classical} = T + V
\end{equation}

\begin{equation} \label{ClassicalEnergy_vx}
    E_{classical} = \frac{1}{2}mv^2 + \frac{1}{2}kx^2
\end{equation}

\begin{equation} \label{ClassicalEnergy_Expanded_xt_vt}
    E_{classical} = \frac{1}{2}m(A\omega \sin(\omega t - \phi))^2 + \frac{1}{2}k(A \cos(\omega t - \phi))^2.
\end{equation}

Using
\begin{equation} \label{k_equals_m_omega_squared}
    k = m\omega^2
\end{equation}
and
\begin{equation} \label{TrigonometricIdentity_sin_cos_squared}
    \sin^2(\theta) + \cos^2(\theta) = 1
\end{equation}
we get
\begin{equation} \label{ClassicalEnergy_TrigIdentity}
E_{classical} = \frac{1}{2}mA^2\omega^2 (\sin^2(\omega t - \phi) + \cos^2(\omega t - \phi))
\end{equation}

\begin{equation} \label{ClassicalEnergy_Simplified_mA_omega}
E_{classical} = \frac{1}{2}mA^2\omega^2
\end{equation}

\begin{equation} \label{ClassicalEnergy_Simplified_kA}
E_{classical} = \frac{1}{2}kA^2
\end{equation}
Classically, energy $E_{classical}$ can take any continuous non-negative value.

\subsection{Classical Probability Density}
The probability of finding the particle in a small interval $dx$ at position $x$ is proportional to the time spent in that interval. Since the particle moves slower near the turning points ($x = \pm A$) and faster near the equilibrium position ($x = 0$), the probability density is inversely proportional to the speed of the particle:
\begin{equation} \label{ClassicalProbabilityDensity_Proportionality}
P_{classical}(x) \propto \frac{1}{|v(x)|}.
\end{equation}
From energy conservation,
\begin{equation} \label{EnergyConservation_ClassicalProbability}
    E_{classical} = \frac{1}{2}mv^2 + \frac{1}{2}kx^2
\end{equation}
we can express the speed as:
\begin{equation} \label{Velocity_vx_EnergyConservation}
v(x) = \pm \sqrt{\frac{2}{m}(E_{classical} - \frac{1}{2}kx^2)}
\end{equation}
\begin{equation} \label{Velocity_vx_Simplified}
v(x) =  \pm \omega \sqrt{A^2 - x^2}
\end{equation}
Thus, the classical probability density is:
\begin{equation} \label{ClassicalProbabilityDensity_C_over_v}
P_{classical}(x) = \frac{C}{|v(x)|}\quad \text{for } |x| \leq A
\end{equation}
\begin{equation} \label{ClassicalProbabilityDensity_C_omega_sqrt}
P_{classical}(x) = \frac{C}{\omega \sqrt{A^2 - x^2}}\quad \text{for } |x| \leq A
\end{equation}
\begin{equation} \label{ClassicalProbabilityDensity_Cprime_sqrt}
P_{classical}(x) = \frac{C'}{\sqrt{A^2 - x^2}} \quad \text{for } |x| \leq A
\end{equation}
and $P_{classical}(x) = 0$ for $|x| > A$.  $C'$ is a normalization constant.

\section{Quantum Harmonic Oscillator (QHO) Review}

In quantum mechanics, the harmonic oscillator is described by the time-independent Schrödinger equation:
\begin{equation} \label{TimeIndependentSchrodingerEquation}
-\frac{\hbar^2}{2m} \frac{d^2\psi_n(x)}{dx^2} + \frac{1}{2}kx^2 \psi_n(x) = E_n \psi_n(x).
\end{equation}
The solutions to this equation are the energy eigenstates $\psi_n(x)$ with corresponding energy eigenvalues $E_n$.

\subsection{Energy Eigenvalues}
The energy eigenvalues for the quantum harmonic oscillator are quantized and given by:
\begin{equation} \label{EnergyEigenvalues_En}
E_n = \left(n + \frac{1}{2}\right)\hbar\omega, \quad n = 0, 1, 2, \ldots
\end{equation}
where $n$ is the quantum number and $\hbar$ is the reduced Planck constant.  Energy levels are discrete and equally spaced with spacing $\hbar\omega$.

\subsection{Energy Eigenstates and Probability Density}
The energy eigenstates $\psi_n(x)$ are given by:
\begin{equation} \label{EnergyEigenstates_Psix}
\psi_n(x) = \frac{1}{\sqrt{2^n n!}} \left(\frac{m\omega}{\pi\hbar}\right)^{1/4} e^{-\frac{m\omega x^2}{2\hbar}} H_n\left(\sqrt{\frac{m\omega}{\hbar}} x\right),
\end{equation}
where $H_n(u)$ are the Hermite polynomials. The probability density of finding the particle at position $x$ in the $n$-th energy state is given by $|\psi_n(x)|^2$.

For low quantum numbers (small $n$), the probability density $|\psi_n(x)|^2$ is significantly different from the classical probability density. For instance, for $n=0$, $|\psi_0(x)|^2$ is a Gaussian centered at $x=0$, with the highest probability at the equilibrium position, unlike the classical case where the probability is lowest at the equilibrium.  Furthermore, there is a non-zero probability of finding the particle outside the classical turning points, which is forbidden in classical mechanics.

\section{Correspondence Principle and the Harmonic Oscillator}

The Correspondence Principle states that in the limit of large quantum numbers ($n \rightarrow \infty$), quantum mechanics should reproduce classical mechanics. Let's examine how this principle applies to the harmonic oscillator.

\subsection{Energy Levels}
The energy levels for the QHO are
\begin{equation} \label{EnergyLevels_QHO}
E_n = (n + \frac{1}{2})\hbar\omega
\end{equation}

For large $n$, we can approximate $E_n \approx n\hbar\omega$. The energy spacing between adjacent levels is:
\begin{equation} \label{EnergySpacing_Definition}
\Delta E = E_{n+1} - E_n
\end{equation}
\begin{equation} \label{EnergySpacing_Expanded}
\Delta E = \left((n+1) + \frac{1}{2}\right)\hbar\omega - \left(n + \frac{1}{2}\right)\hbar\omega
\end{equation}
\begin{equation} \label{EnergySpacing_DeltaE}
\Delta E = \hbar\omega.
\end{equation}
This spacing $\hbar\omega$ is constant and independent of $n$. However, consider the \textit{relative} spacing compared to the energy itself:
\begin{equation} \label{RelativeEnergySpacing}
\frac{\Delta E}{E_n} = \frac{\hbar\omega}{(n + \frac{1}{2})\hbar\omega}
\end{equation}
\begin{equation} \label{RelativeEnergySpacing_Simplified}
\frac{\Delta E}{E_n} = \frac{1}{n + \frac{1}{2}}
\end{equation}
\begin{equation} \label{RelativeEnergySpacing_Approximation}
\frac{\Delta E}{E_n}  \approx \frac{1}{n} \quad \text{for large } n.
\end{equation}
As $n \rightarrow \infty$, the relative spacing $\frac{\Delta E}{E_n} \rightarrow 0$. This means that for very high energy levels, the energy spectrum becomes effectively continuous, resembling the classical case where energy can take continuous values.  In the classical limit of high energy, the quantization of energy becomes less apparent.

\subsection{Probability Density}
Now let's consider the probability density. For large $n$, the wave function $\psi_n(x)$ becomes highly oscillatory, and the Hermite polynomials $H_n(u)$ also exhibit oscillatory behavior within the classical allowed region and decay exponentially outside.  While a rigorous mathematical proof is complex, we can argue qualitatively and consider some properties:

1. \textbf{Classical Turning Points:} The classical turning points are given by
\begin{equation} \label{ClassicalTurningPoints_Condition}
    V(x) = E_{classical}
\end{equation}
so
\begin{equation} \label{ClassicalTurningPoints_Condition_Expanded}
    \frac{1}{2}kx^2 = E_{classical}
\end{equation}
For the $n$-th quantum state, we set
\begin{equation} \label{ClassicalEnergy_equals_QuantumEnergy}
    E_{classical} = E_n
\end{equation}
\begin{equation} \label{ClassicalEnergy_equals_QuantumEnergy_Expanded}
    E_{classical} = (n + \frac{1}{2})\hbar\omega
\end{equation}
 Thus, the classical turning points are $x = \pm A_n$, where
 \begin{equation} \label{Amplitude_An_En}
     \frac{1}{2}kA_n^2 = E_n
 \end{equation}
 so
 \begin{equation} \label{Amplitude_An_Expanded}
     A_n = \sqrt{\frac{2E_n}{k}} = \sqrt{\frac{2(n + \frac{1}{2})\hbar\omega}{m\omega^2}} = \sqrt{\frac{(2n+1)\hbar}{m\omega}}.
 \end{equation}

2. \textbf{Oscillatory Nature of $\psi_n(x)$ for large $n$:} For large $n$, the Hermite polynomials $H_n(u)$ behave approximately as sinusoidal functions within the classically allowed region.  The square of the wave function, $|\psi_n(x)|^2$, will thus exhibit rapid oscillations within $|x| \leq A_n$.

3. \textbf{Averaging over Oscillations:} If we average the quantum probability density $|\psi_n(x)|^2$ over a spatial interval that is large compared to the wavelength of the oscillations but small compared to the overall extent of the wave function, we find that the averaged quantum probability density approaches the classical probability density.  Intuitively, the rapid oscillations in $|\psi_n(x)|^2$ for large $n$ average out to a smoother distribution that resembles the classical distribution, which is peaked at the turning points and minimum at the equilibrium position.

To illustrate this more explicitly, consider the WKB approximation, which is valid for large quantum numbers. In the WKB approximation, the wave function in the classically allowed region oscillates as:
\begin{equation} \label{WKB_WaveFunction}
\psi_{WKB}(x) \approx \frac{C}{\sqrt{p(x)}} \cos\left(\frac{1}{\hbar} \int^x p(x') dx' - \frac{\pi}{4}\right),
\end{equation}
where
\begin{equation} \label{ClassicalMomentum_Definition}
p(x) = \sqrt{2m(E - V(x))}
\end{equation}
\begin{equation} \label{ClassicalMomentum_HarmonicOscillator}
p(x)= \sqrt{2m(E - \frac{1}{2}kx^2)}
\end{equation}
 is the classical momentum.  The probability density is then approximately:
\begin{equation} \label{WKB_ProbabilityDensity_BeforeAveraging}
|\psi_{WKB}(x)|^2 \approx \frac{C^2}{p(x)} \cos^2\left(\frac{1}{\hbar} \int^x p(x') dx' - \frac{\pi}{4}\right).
\end{equation}
Averaging the rapidly oscillating $\cos^2$ term over a cycle gives $\langle \cos^2 \rangle = \frac{1}{2}$.  Thus, the average probability density becomes:
\begin{equation} \label{AveragedWKB_ProbabilityDensity}
\langle |\psi_{WKB}(x)|^2 \rangle \approx \frac{C^2}{2p(x)}
\end{equation}
\begin{equation} \label{AveragedWKB_ProbabilityDensity_Cprime}
\langle |\psi_{WKB}(x)|^2 \rangle = \frac{C'^2}{\sqrt{E - \frac{1}{2}kx^2}}
\end{equation}
\begin{equation} \label{AveragedWKB_ProbabilityDensity_Cdoubleprime}
\langle |\psi_{WKB}(x)|^2 \rangle = \frac{C''}{\sqrt{A_n^2 - x^2}},
\end{equation}
where we used
\begin{equation} \label{EnergyRelation_WKB}
    E = E_n = \frac{1}{2}kA_n^2
\end{equation}
This form is proportional to the classical probability density
\begin{equation} \label{ClassicalProbabilityDensity_Reiterated}
    P_{classical}(x) = \frac{C'}{\sqrt{A^2 - x^2}}
\end{equation}
Hence, in the limit of large quantum numbers, the average quantum probability density approaches the classical probability density within the classically allowed region.

Furthermore, the quantum probability of finding the particle outside the classical turning points (tunneling) becomes relatively less significant for large $n$. The exponentially decaying tails of $\psi_n(x)$ beyond $x = \pm A_n$ become less prominent compared to the oscillatory behavior within the classical region as $n$ increases.

\section{Conclusion}

We have shown how the quantum harmonic oscillator approaches the classical harmonic oscillator in the limit of large quantum numbers, in accordance with the Correspondence Principle.

\begin{itemize}
    \item \textbf{Energy Levels:} The energy levels of the QHO are quantized, but for large quantum numbers $n$, the relative spacing between energy levels becomes vanishingly small compared to the energy itself.  This makes the energy spectrum effectively continuous, as in classical mechanics.
    \item \textbf{Probability Density:} While the quantum probability density for low $n$ is significantly different from the classical probability density, for large $n$, the average quantum probability density (obtained by averaging over oscillations or using approximations like WKB) approaches the classical probability density within the classically allowed region.  The oscillatory nature of the quantum wave function for large $n$ and its localization within the classical boundaries are key to this correspondence.
\end{itemize}

The harmonic oscillator provides a clear and important example of the Correspondence Principle, demonstrating how quantum mechanics smoothly transitions to classical mechanics in the appropriate limit. This principle is not only a historical cornerstone in the development of quantum theory but also a valuable tool for understanding the relationship between quantum and classical descriptions of physical systems.

\end{document}

thoughts

The user wants me to modify the provided Latex file.
I need to add explanatory text to each step of the derivation.
I also need to add references to equations using Latex equation labels.
I should not add any commentary other than the requested edits.

Plan:

  1. Go through the Latex file section by section.
  2. For each equation, identify the mathematical transformation that leads to the next equation.
  3. Add explanatory text describing the transformation.
  4. Add references to relevant equations using Eq.~\ref{}.
  5. Ensure that all equation labels are correctly used.
  6. Output the revised Latex file content.

response

\documentclass{article}
\usepackage{amsmath}

% margins of 1 inch:
\setlength{\topmargin}{-.5in}
\setlength{\textheight}{9in}
\setlength{\oddsidemargin}{0in}
\setlength{\textwidth}{6.5in}

\usepackage[pdftex]{hyperref} % hyperlink equation and bibliographic citations

\title{The Correspondence Principle and the Harmonic Oscillator}
\author{Ben Payne, with Gemini 2.0}
\date{\today}

\begin{document}

\maketitle
\begin{abstract}
    From Gemini 2.0 Flash Thinking Experimental
\end{abstract}


\section{Introduction}

The harmonic oscillator is a fundamental system in physics, serving as a model for a wide range of phenomena from molecular vibrations to oscillations in electromagnetic fields.  It is particularly insightful in the context of the Correspondence Principle, which states that the predictions of quantum mechanics must agree with classical mechanics in the limit of large quantum numbers or high energies. This principle provides a crucial bridge between the seemingly disparate worlds of classical and quantum physics.  In this document, we will mathematically explore how the quantum harmonic oscillator (QHO) approaches the classical harmonic oscillator (CHO) in the classical limit, demonstrating the Correspondence Principle.

\section{Classical Harmonic Oscillator (CHO) Review}

Consider a particle of mass $m$ attached to a spring with spring constant $k$. The potential energy of this system is given by
\begin{equation} \label{PotentialEnergy_HarmonicOscillator}
    V(x) = \frac{1}{2}kx^2
\end{equation}
This equation describes the potential energy of a harmonic oscillator as a function of displacement $x$.

\subsection{Equations of Motion and Energy}
From Newton's second law,
\begin{equation} \label{NewtonsSecondLaw_HarmonicOscillator}
    F = ma = m\ddot{x} = -kx
\end{equation}
This is Newton's second law applied to the harmonic oscillator, where the restoring force $F$ is proportional to the displacement $x$. This leads to the differential equation for simple harmonic motion:
\begin{equation} \label{DE_SimpleHarmonicMotion}
\ddot{x} + \omega^2 x = 0,
\end{equation}
Eq.~\eqref{DE_SimpleHarmonicMotion} is obtained by rearranging Eq.~\eqref{NewtonsSecondLaw_HarmonicOscillator} and defining $\omega^2 = k/m$, where
\begin{equation} \label{AngularFrequency_Omega}
    \omega = \sqrt{\frac{k}{m}}
\end{equation}
is the angular frequency of oscillation. The general solution to Eq.~\eqref{DE_SimpleHarmonicMotion} is:
\begin{equation} \label{GeneralSolution_xt}
x(t) = A \cos(\omega t - \phi),
\end{equation}
where $A$ is the amplitude and $\phi$ is the phase constant, determined by initial conditions.

The velocity is given by taking the time derivative of Eq.~\eqref{GeneralSolution_xt}:
\begin{equation} \label{Velocity_vt}
v(t) = \dot{x}(t) = -A\omega \sin(\omega t - \phi).
\end{equation}

The total energy of the classical harmonic oscillator is conserved and is the sum of kinetic and potential energies:
\begin{equation} \label{ClassicalEnergy_Expanded}
E_{classical} = T + V
\end{equation}
where $T = \frac{1}{2}mv^2$ and $V = \frac{1}{2}kx^2$. Substituting these expressions into Eq.~\eqref{ClassicalEnergy_Expanded} gives:
\begin{equation} \label{ClassicalEnergy_vx}
    E_{classical} = \frac{1}{2}mv^2 + \frac{1}{2}kx^2
\end{equation}
Substituting the expressions for $x(t)$ from Eq.~\eqref{GeneralSolution_xt} and $v(t)$ from Eq.~\eqref{Velocity_vt} into Eq.~\eqref{ClassicalEnergy_vx} yields:
\begin{equation} \label{ClassicalEnergy_Expanded_xt_vt}
    E_{classical} = \frac{1}{2}m(-A\omega \sin(\omega t - \phi))^2 + \frac{1}{2}k(A \cos(\omega t - \phi))^2.
\end{equation}

Using the relation between spring constant and angular frequency
\begin{equation} \label{k_equals_m_omega_squared}
    k = m\omega^2
\end{equation}
and the trigonometric identity
\begin{equation} \label{TrigonometricIdentity_sin_cos_squared}
    \sin^2(\theta) + \cos^2(\theta) = 1
\end{equation}
we can simplify Eq.~\eqref{ClassicalEnergy_Expanded_xt_vt}. Substituting Eq.~\eqref{k_equals_m_omega_squared} into Eq.~\eqref{ClassicalEnergy_Expanded_xt_vt} and factoring out $\frac{1}{2}mA^2\omega^2$ leads to:
\begin{equation} \label{ClassicalEnergy_TrigIdentity}
E_{classical} = \frac{1}{2}mA^2\omega^2 (\sin^2(\omega t - \phi) + \cos^2(\omega t - \phi))
\end{equation}
Applying the trigonometric identity from Eq.~\eqref{TrigonometricIdentity_sin_cos_squared} to Eq.~\eqref{ClassicalEnergy_TrigIdentity} simplifies to:
\begin{equation} \label{ClassicalEnergy_Simplified_mA_omega}
E_{classical} = \frac{1}{2}mA^2\omega^2
\end{equation}
Substituting $m\omega^2 = k$ from Eq.~\eqref{k_equals_m_omega_squared} into Eq.~\eqref{ClassicalEnergy_Simplified_mA_omega} gives an alternate form for the classical energy:
\begin{equation} \label{ClassicalEnergy_Simplified_kA}
E_{classical} = \frac{1}{2}kA^2
\end{equation}
Classically, energy $E_{classical}$ can take any continuous non-negative value.

\subsection{Classical Probability Density}
The probability of finding the particle in a small interval $dx$ at position $x$ is proportional to the time spent in that interval. Since the particle moves slower near the turning points ($x = \pm A$) and faster near the equilibrium position ($x = 0$), the probability density is inversely proportional to the speed of the particle:
\begin{equation} \label{ClassicalProbabilityDensity_Proportionality}
P_{classical}(x) \propto \frac{1}{|v(x)|}.
\end{equation}
From energy conservation, expressed in Eq.~\eqref{ClassicalEnergy_vx} as:
\begin{equation} \label{EnergyConservation_ClassicalProbability}
    E_{classical} = \frac{1}{2}mv^2 + \frac{1}{2}kx^2
\end{equation}
we can solve for the speed $v(x)$ to get:
\begin{equation} \label{Velocity_vx_EnergyConservation}
v(x) = \pm \sqrt{\frac{2}{m}(E_{classical} - \frac{1}{2}kx^2)}
\end{equation}
Substituting $E_{classical} = \frac{1}{2}kA^2 = \frac{1}{2}m\omega^2A^2$ from Eq.~\eqref{ClassicalEnergy_Simplified_kA} and Eq.~\eqref{k_equals_m_omega_squared} into Eq.~\eqref{Velocity_vx_EnergyConservation} gives:
\begin{equation} \label{Velocity_vx_Simplified}
v(x) =  \pm \omega \sqrt{A^2 - x^2}
\end{equation}
Thus, using the proportionality from Eq.~\eqref{ClassicalProbabilityDensity_Proportionality}, the classical probability density is:
\begin{equation} \label{ClassicalProbabilityDensity_C_over_v}
P_{classical}(x) = \frac{C}{|v(x)|}\quad \text{for } |x| \leq A
\end{equation}
Substituting the expression for $v(x)$ from Eq.~\eqref{Velocity_vx_Simplified} into Eq.~\eqref{ClassicalProbabilityDensity_C_over_v} yields:
\begin{equation} \label{ClassicalProbabilityDensity_C_omega_sqrt}
P_{classical}(x) = \frac{C}{\omega \sqrt{A^2 - x^2}}\quad \text{for } |x| \leq A
\end{equation}
Absorbing the constant $\omega$ into the normalization constant $C$ in Eq.~\eqref{ClassicalProbabilityDensity_C_omega_sqrt} leads to:
\begin{equation} \label{ClassicalProbabilityDensity_Cprime_sqrt}
P_{classical}(x) = \frac{C'}{\sqrt{A^2 - x^2}} \quad \text{for } |x| \leq A
\end{equation}
and $P_{classical}(x) = 0$ for $|x| > A$.  $C'$ is a normalization constant.

\section{Quantum Harmonic Oscillator (QHO) Review}

In quantum mechanics, the harmonic oscillator is described by the time-independent Schrödinger equation:
\begin{equation} \label{TimeIndependentSchrodingerEquation}
-\frac{\hbar^2}{2m} \frac{d^2\psi_n(x)}{dx^2} + \frac{1}{2}kx^2 \psi_n(x) = E_n \psi_n(x).
\end{equation}
This is the time-independent Schrödinger equation for a particle in a harmonic potential. The solutions to this equation are the energy eigenstates $\psi_n(x)$ with corresponding energy eigenvalues $E_n$.

\subsection{Energy Eigenvalues}
The energy eigenvalues for the quantum harmonic oscillator, obtained by solving Eq.~\eqref{TimeIndependentSchrodingerEquation}, are quantized and given by:
\begin{equation} \label{EnergyEigenvalues_En}
E_n = \left(n + \frac{1}{2}\right)\hbar\omega, \quad n = 0, 1, 2, \ldots
\end{equation}
where $n$ is the quantum number and $\hbar$ is the reduced Planck constant.  Energy levels are discrete and equally spaced with spacing $\hbar\omega$.

\subsection{Energy Eigenstates and Probability Density}
The energy eigenstates $\psi_n(x)$ are solutions to Eq.~\eqref{TimeIndependentSchrodingerEquation} and are given by:
\begin{equation} \label{EnergyEigenstates_Psix}
\psi_n(x) = \frac{1}{\sqrt{2^n n!}} \left(\frac{m\omega}{\pi\hbar}\right)^{1/4} e^{-\frac{m\omega x^2}{2\hbar}} H_n\left(\sqrt{\frac{m\omega}{\hbar}} x\right),
\end{equation}
where $H_n(u)$ are the Hermite polynomials. The probability density of finding the particle at position $x$ in the $n$-th energy state is given by $|\psi_n(x)|^2$.

For low quantum numbers (small $n$), the probability density $|\psi_n(x)|^2$ is significantly different from the classical probability density derived in Eq.~\eqref{ClassicalProbabilityDensity_Cprime_sqrt}. For instance, for $n=0$, $|\psi_0(x)|^2$ is a Gaussian centered at $x=0$, with the highest probability at the equilibrium position, unlike the classical case where the probability is lowest at the equilibrium.  Furthermore, there is a non-zero probability of finding the particle outside the classical turning points, which is forbidden in classical mechanics.

\section{Correspondence Principle and the Harmonic Oscillator}

The Correspondence Principle states that in the limit of large quantum numbers ($n \rightarrow \infty$), quantum mechanics should reproduce classical mechanics. Let's examine how this principle applies to the harmonic oscillator.

\subsection{Energy Levels}
The energy levels for the QHO are given by Eq.~\eqref{EnergyEigenvalues_En}:
\begin{equation} \label{EnergyLevels_QHO}
E_n = (n + \frac{1}{2})\hbar\omega
\end{equation}

For large $n$, the term $\frac{1}{2}$ becomes negligible compared to $n$, so we can approximate $E_n \approx n\hbar\omega$. The energy spacing between adjacent levels is defined as:
\begin{equation} \label{EnergySpacing_Definition}
\Delta E = E_{n+1} - E_n
\end{equation}
Substituting the expression for $E_n$ from Eq.~\eqref{EnergyLevels_QHO} into Eq.~\eqref{EnergySpacing_Definition} we get:
\begin{equation} \label{EnergySpacing_Expanded}
\Delta E = \left((n+1) + \frac{1}{2}\right)\hbar\omega - \left(n + \frac{1}{2}\right)\hbar\omega
\end{equation}
Simplifying Eq.~\eqref{EnergySpacing_Expanded} yields:
\begin{equation} \label{EnergySpacing_DeltaE}
\Delta E = \hbar\omega.
\end{equation}
This spacing $\hbar\omega$ is constant and independent of $n$. However, consider the \textit{relative} spacing compared to the energy itself:
\begin{equation} \label{RelativeEnergySpacing}
\frac{\Delta E}{E_n} = \frac{\hbar\omega}{(n + \frac{1}{2})\hbar\omega}
\end{equation}
Substituting the expressions for $\Delta E$ from Eq.~\eqref{EnergySpacing_DeltaE} and $E_n$ from Eq.~\eqref{EnergyLevels_QHO} into Eq.~\eqref{RelativeEnergySpacing}. Simplifying Eq.~\eqref{RelativeEnergySpacing} leads to:
\begin{equation} \label{RelativeEnergySpacing_Simplified}
\frac{\Delta E}{E_n} = \frac{1}{n + \frac{1}{2}}
\end{equation}
For large $n$, we can approximate $n + \frac{1}{2} \approx n$, so Eq.~\eqref{RelativeEnergySpacing_Simplified} becomes:
\begin{equation} \label{RelativeEnergySpacing_Approximation}
\frac{\Delta E}{E_n}  \approx \frac{1}{n} \quad \text{for large } n.
\end{equation}
As $n \rightarrow \infty$, the relative spacing $\frac{\Delta E}{E_n} \rightarrow 0$. This means that for very high energy levels, the energy spectrum becomes effectively continuous, resembling the classical case where energy can take continuous values.  In the classical limit of high energy, the quantization of energy becomes less apparent.

\subsection{Probability Density}
Now let's consider the probability density. For large $n$, the wave function $\psi_n(x)$ becomes highly oscillatory, and the Hermite polynomials $H_n(u)$ also exhibit oscillatory behavior within the classically allowed region and decay exponentially outside.  While a rigorous mathematical proof is complex, we can argue qualitatively and consider some properties:

1. \textbf{Classical Turning Points:} The classical turning points are given by the condition that the potential energy equals the total classical energy:
\begin{equation} \label{ClassicalTurningPoints_Condition}
    V(x) = E_{classical}
\end{equation}
Substituting the potential energy from Eq.~\eqref{PotentialEnergy_HarmonicOscillator} into Eq.~\eqref{ClassicalTurningPoints_Condition} gives:
\begin{equation} \label{ClassicalTurningPoints_Condition_Expanded}
    \frac{1}{2}kx^2 = E_{classical}
\end{equation}
For the $n$-th quantum state, according to the Correspondence Principle, we equate the classical energy with the quantum energy eigenvalue:
\begin{equation} \label{ClassicalEnergy_equals_QuantumEnergy}
    E_{classical} = E_n
\end{equation}
Substituting the expression for $E_n$ from Eq.~\eqref{EnergyLevels_QHO} into Eq.~\eqref{ClassicalEnergy_equals_QuantumEnergy} gives:
\begin{equation} \label{ClassicalEnergy_equals_QuantumEnergy_Expanded}
    E_{classical} = (n + \frac{1}{2})\hbar\omega
\end{equation}
 Thus, the classical turning points are $x = \pm A_n$, where $A_n$ is the amplitude for the $n$-th quantum state, and is determined by equating the potential energy at the turning point to the quantum energy $E_n$:
 \begin{equation} \label{Amplitude_An_En}
     \frac{1}{2}kA_n^2 = E_n
 \end{equation}
 Solving Eq.~\eqref{Amplitude_An_En} for $A_n$ and substituting $E_n$ from Eq.~\eqref{EnergyLevels_QHO} and $k = m\omega^2$ from Eq.~\eqref{k_equals_m_omega_squared} we get:
 \begin{equation} \label{Amplitude_An_Expanded}
     A_n = \sqrt{\frac{2E_n}{k}} = \sqrt{\frac{2(n + \frac{1}{2})\hbar\omega}{m\omega^2}} = \sqrt{\frac{(2n+1)\hbar}{m\omega}}.
 \end{equation}

2. \textbf{Oscillatory Nature of $\psi_n(x)$ for large $n$:} For large $n$, the Hermite polynomials $H_n(u)$ behave approximately as sinusoidal functions within the classically allowed region.  The square of the wave function, $|\psi_n(x)|^2$, will thus exhibit rapid oscillations within $|x| \leq A_n$.

3. \textbf{Averaging over Oscillations:} If we average the quantum probability density $|\psi_n(x)|^2$ over a spatial interval that is large compared to the wavelength of the oscillations but small compared to the overall extent of the wave function, we find that the averaged quantum probability density approaches the classical probability density.  Intuitively, the rapid oscillations in $|\psi_n(x)|^2$ for large $n$ average out to a smoother distribution that resembles the classical distribution, which is peaked at the turning points and minimum at the equilibrium position.

To illustrate this more explicitly, consider the WKB approximation, which is valid for large quantum numbers. In the WKB approximation, the wave function in the classically allowed region oscillates as:
\begin{equation} \label{WKB_WaveFunction}
\psi_{WKB}(x) \approx \frac{C}{\sqrt{p(x)}} \cos\left(\frac{1}{\hbar} \int^x p(x') dx' - \frac{\pi}{4}\right),
\end{equation}
where $p(x)$ is the classical momentum defined as:
\begin{equation} \label{ClassicalMomentum_Definition}
p(x) = \sqrt{2m(E - V(x))}
\end{equation}
Substituting the potential energy for the harmonic oscillator from Eq.~\eqref{PotentialEnergy_HarmonicOscillator} into Eq.~\eqref{ClassicalMomentum_Definition} gives:
\begin{equation} \label{ClassicalMomentum_HarmonicOscillator}
p(x)= \sqrt{2m(E - \frac{1}{2}kx^2)}
\end{equation}
 is the classical momentum.  The probability density in the WKB approximation is then approximately the square of the wave function from Eq.~\eqref{WKB_WaveFunction}:
\begin{equation} \label{WKB_ProbabilityDensity_BeforeAveraging}
|\psi_{WKB}(x)|^2 \approx \frac{C^2}{p(x)} \cos^2\left(\frac{1}{\hbar} \int^x p(x') dx' - \frac{\pi}{4}\right).
\end{equation}
Averaging the rapidly oscillating $\cos^2$ term in Eq.~\eqref{WKB_ProbabilityDensity_BeforeAveraging} over a cycle gives $\langle \cos^2 \rangle = \frac{1}{2}$.  Thus, the average probability density becomes:
\begin{equation} \label{AveragedWKB_ProbabilityDensity}
\langle |\psi_{WKB}(x)|^2 \rangle \approx \frac{C^2}{2p(x)}
\end{equation}
Substituting the expression for $p(x)$ from Eq.~\eqref{ClassicalMomentum_HarmonicOscillator} into Eq.~\eqref{AveragedWKB_ProbabilityDensity} and absorbing constants into $C'$ gives:
\begin{equation} \label{AveragedWKB_ProbabilityDensity_Cprime}
\langle |\psi_{WKB}(x)|^2 \rangle = \frac{C'^2}{\sqrt{E - \frac{1}{2}kx^2}}
\end{equation}
Replacing $E$ with the quantum energy $E_n = \frac{1}{2}kA_n^2$ from Eq.~\eqref{Amplitude_An_En} and absorbing constants into $C''$ in Eq.~\eqref{AveragedWKB_ProbabilityDensity_Cprime} leads to:
\begin{equation} \label{AveragedWKB_ProbabilityDensity_Cdoubleprime}
\langle |\psi_{WKB}(x)|^2 \rangle = \frac{C''}{\sqrt{A_n^2 - x^2}},
\end{equation}
where we used the correspondence between quantum energy $E_n$ and classical energy $E$:
\begin{equation} \label{EnergyRelation_WKB}
    E = E_n = \frac{1}{2}kA_n^2
\end{equation}
Comparing Eq.~\eqref{AveragedWKB_ProbabilityDensity_Cdoubleprime} with the classical probability density from Eq.~\eqref{ClassicalProbabilityDensity_Cprime_sqrt}:
\begin{equation} \label{ClassicalProbabilityDensity_Reiterated}
    P_{classical}(x) = \frac{C'}{\sqrt{A^2 - x^2}}
\end{equation}
We see that the averaged WKB probability density Eq.~\eqref{AveragedWKB_ProbabilityDensity_Cdoubleprime} has the same form as the classical probability density Eq.~\eqref{ClassicalProbabilityDensity_Reiterated}. Hence, in the limit of large quantum numbers, the average quantum probability density approaches the classical probability density within the classically allowed region.

Furthermore, the quantum probability of finding the particle outside the classical turning points (tunneling) becomes relatively less significant for large $n$. The exponentially decaying tails of $\psi_n(x)$ beyond $x = \pm A_n$ become less prominent compared to the oscillatory behavior within the classical region as $n$ increases.

\section{Conclusion}

We have shown how the quantum harmonic oscillator approaches the classical harmonic oscillator in the limit of large quantum numbers, in accordance with the Correspondence Principle.

\begin{itemize}
    \item \textbf{Energy Levels:} The energy levels of the QHO are quantized, but for large quantum numbers $n$, the relative spacing between energy levels becomes vanishingly small compared to the energy itself.  This makes the energy spectrum effectively continuous, as in classical mechanics.
    \item \textbf{Probability Density:} While the quantum probability density for low $n$ is significantly different from the classical probability density, for large $n$, the average quantum probability density (obtained by averaging over oscillations or using approximations like WKB) approaches the classical probability density within the classically allowed region.  The oscillatory nature of the quantum wave function for large $n$ and its localization within the classical boundaries are key to this correspondence.
\end{itemize}

The harmonic oscillator provides a clear and important example of the Correspondence Principle, demonstrating how quantum mechanics smoothly transitions to classical mechanics in the appropriate limit. This principle is not only a historical cornerstone in the development of quantum theory but also a valuable tool for understanding the relationship between quantum and classical descriptions of physical systems.

\end{document}

@bhpayne bhpayne changed the title derivation: Quantum Harmonic Oscillator and classical Simple Harmonic Oscillator are related by the Correspondence Principle bridge derivation: Quantum Harmonic Oscillator and classical Simple Harmonic Oscillator are related by the Correspondence Principle Jan 29, 2025
Sign up for free to join this conversation on GitHub. Already have an account? Sign in to comment
Projects
Status: No status
Development

No branches or pull requests

1 participant