Skip to main content
Physics LibreTexts

12.2: Two-State System

  • Page ID
    15800
  • \( \newcommand{\vecs}[1]{\overset { \scriptstyle \rightharpoonup} {\mathbf{#1}} } \) \( \newcommand{\vecd}[1]{\overset{-\!-\!\rightharpoonup}{\vphantom{a}\smash {#1}}} \)\(\newcommand{\id}{\mathrm{id}}\) \( \newcommand{\Span}{\mathrm{span}}\) \( \newcommand{\kernel}{\mathrm{null}\,}\) \( \newcommand{\range}{\mathrm{range}\,}\) \( \newcommand{\RealPart}{\mathrm{Re}}\) \( \newcommand{\ImaginaryPart}{\mathrm{Im}}\) \( \newcommand{\Argument}{\mathrm{Arg}}\) \( \newcommand{\norm}[1]{\| #1 \|}\) \( \newcommand{\inner}[2]{\langle #1, #2 \rangle}\) \( \newcommand{\Span}{\mathrm{span}}\) \(\newcommand{\id}{\mathrm{id}}\) \( \newcommand{\Span}{\mathrm{span}}\) \( \newcommand{\kernel}{\mathrm{null}\,}\) \( \newcommand{\range}{\mathrm{range}\,}\) \( \newcommand{\RealPart}{\mathrm{Re}}\) \( \newcommand{\ImaginaryPart}{\mathrm{Im}}\) \( \newcommand{\Argument}{\mathrm{Arg}}\) \( \newcommand{\norm}[1]{\| #1 \|}\) \( \newcommand{\inner}[2]{\langle #1, #2 \rangle}\) \( \newcommand{\Span}{\mathrm{span}}\)\(\newcommand{\AA}{\unicode[.8,0]{x212B}}\)

    Consider a system in which the time-independent Hamiltonian possesses two eigenstates, denoted \[\begin{aligned} H_0\,\psi_1 &= E_1\,\psi_1,\\[0.5ex] H_0\,\psi_2&=E_2\,\psi_2.\end{aligned}\] Suppose, for the sake of simplicity, that the diagonal elements of the interaction Hamiltonian, \(H_1\), are zero: that is, \[\langle 1|H_1|1\rangle = \langle 2|H_1|2\rangle = 0.\] The off-diagonal elements are assumed to oscillate sinusoidally at some frequency \(\omega\): that is, \[\langle 1|H_1|2\rangle = \langle 2|H_1|1\rangle^\ast = \gamma\,\hbar\,\exp(\,{\rm i}\,\omega\,t),\] where \(\gamma\) and \(\omega\) are real. Note that it is only the off-diagonal matrix elements which give rise to the effect which we are interested in: namely, transitions between states 1 and 2.

    For a two-state system, Equation ([e13.12]) reduces to

    \[\begin{aligned} {\rm i}\,\frac{dc_1}{dt} &= \gamma\,\exp\left[+{\rm i}\,(\omega-\omega_{21})\,t\right]c_2,\\[0.5ex] {\rm i}\,\frac{dc_2}{dt} &=\gamma\,\exp\left[-{\rm i}\,(\omega-\omega_{21})\,t\right]c_1,\label{e13.20}\end{aligned}\] where \(\omega_{21}=(E_2-E_1)/\hbar\). The previous two equations can be combined to give a second-order differential equation for the time-variation of the amplitude \(c_2\): that is,

    \[\label{e13.21} \frac{d^{\,2} c_2}{dt^{\,2}} + {\rm i}\,(\omega-\omega_{21})\,\frac{dc_2}{dt}+\gamma^{\,2}\,c_2=0.\] Once we have solved for \(c_2\), we can use Equation ([e13.20]) to obtain the amplitude \(c_1\). Let us search for a solution in which the system is certain to be in state 1 (and, thus, has no chance of being in state 2) at time \(t=0\). Thus, our initial conditions are \(c_1(0)=1\) and \(c_2(0)=0\). It is easily demonstrated that the appropriate solutions to ([e13.21]) and ([e13.20]) are

    \begin{equation}c_{2}(t)=\left(\frac{-\mathrm{i} \gamma}{\Omega}\right) \exp \left[\frac{-\mathrm{i}\left(\omega-\omega_{21}\right) t}{2}\right] \sin (\Omega t)\end{equation}

    \begin{equation}\begin{aligned}
    c_{1}(t)=& \exp \left[\frac{\mathrm{i}\left(\omega-\omega_{21}\right) t}{2}\right] \cos (\Omega t) \\
    &-\left[\frac{\mathrm{i}\left(\omega-\omega_{21}\right)}{2 \Omega}\right] \exp \left[\frac{\mathrm{i}\left(\omega-\omega_{21}\right) t}{2}\right] \sin (\Omega t)
    \end{aligned}\end{equation}

    where \begin{equation}\Omega=\sqrt{\gamma^{2}+\left(\omega-\omega_{21}\right)^{2} / 4}\end{equation}

    Now, the probability of finding the system in state 1 at time \(t\) is simply \(P_1(t)=|c_1(t)|^{\,2}\). Likewise, the probability of finding the system in state 2 at time \(t\) is \(P_2(t)= |c_2(t)|^{\,2}\). It follows that \[\begin{aligned} P_1(t)&=1-P_2(t),\\[0.5ex] P_2(t)&= \left[\frac{\gamma^{\,2}}{\gamma^{\,2} + (\omega-\omega_{21})^{\,2}/4}\right] \sin^2({\mit\Omega}\,t).\label{e13.25}\end{aligned}\] This result is known as Rabi’s formula .

    Equation ([e13.25]) exhibits all the features of a classic resonance . At resonance, when the oscillation frequency of the perturbation, \(\omega\), matches the frequency \(\omega_{21}\), we find that

    \[\begin{aligned} P_1(t)&=\cos^2(\gamma\,t),\\[0.5ex] P_2(t)&=\sin^2(\gamma\,t).\label{e13.28}\end{aligned}\] According to the previous result, the system starts off in state 1 at \(t=0\). After a time interval \(\pi/(2\,\gamma)\) it is certain to be in state 2. After a further time interval \(\pi/(2\,\gamma)\) it is certain to be in state 1 again, and so on. Thus, the system periodically flip-flops between states 1 and 2 under the influence of the time-dependent perturbation. This implies that the system alternatively absorbs and emits energy from the source of the perturbation.

    The absorption-emission cycle also takes place away from the resonance, when \(\omega\neq \omega_{21}\). However, the amplitude of the oscillation in the coefficient \(c_2\) is reduced. This means that the maximum value of \(P_2(t)\) is no longer unity, nor is the minimum of \(P_1(t)\) zero. In fact, if we plot the maximum value of \(P_2(t)\) as a function of the applied frequency, \(\omega\), then we obtain a resonance curve whose maximum (unity) lies at the resonance, and whose full-width half-maximum (in frequency) is \(4\,\gamma\). Thus, if the applied frequency differs from the resonant frequency by substantially more than \(2\,\gamma\) then the probability of the system jumping from state 1 to state 2 is always very small. In other words, the time-dependent perturbation is only effective at causing transitions between states 1 and 2 if its frequency of oscillation lies in the approximate range \(\omega_{21}\pm 2\,\gamma\). Clearly, the weaker the perturbation (i.e., the smaller \(\gamma\) becomes), the narrower the resonance.

    Contributors and Attributions

    • Richard Fitzpatrick (Professor of Physics, The University of Texas at Austin)

      \( \newcommand {\ltapp} {\stackrel {_{\normalsize<}}{_{\normalsize \sim}}}\) \(\newcommand {\gtapp} {\stackrel {_{\normalsize>}}{_{\normalsize \sim}}}\) \(\newcommand {\btau}{\mbox{\boldmath$\tau$}}\) \(\newcommand {\bmu}{\mbox{\boldmath$\mu$}}\) \(\newcommand {\bsigma}{\mbox{\boldmath$\sigma$}}\) \(\newcommand {\bOmega}{\mbox{\boldmath$\Omega$}}\) \(\newcommand {\bomega}{\mbox{\boldmath$\omega$}}\) \(\newcommand {\bepsilon}{\mbox{\boldmath$\epsilon$}}\)

    This page titled 12.2: Two-State System is shared under a not declared license and was authored, remixed, and/or curated by Richard Fitzpatrick.

    • Was this article helpful?