$\newcommand{\vect}[1]{\mathrm{ \vec{\boldsymbol{#1}}}} $ $\newcommand{\vecta}[3]{ \begin{bmatrix} #1\\ #2\\ #3 \end{bmatrix} } $ $\newcommand{\hata}[1]{\mathrm{ \hat{\boldsymbol{#1}}}} $ $ \newcommand{\blbox}[1]{\bbox[8px,border:1px solid black]{#1}}$ $ \newcommand{\ddp}[2]{ \frac{\partial ^2 {#1}}{\partial {#2}^2}} $ $ \newcommand{\ddf}[2]{ \frac{d ^2 {#1}}{d {#2}^2}} $
Home Page
Next Page

Damped Harmonic Oscillator


Constructing and Justifying our Differential Equation

Hooke's law describes a restorative force: $F=-kx$. What makes this force restorative?

Each contribution to the net force $\vect{F}$ may be written:


Newton's Equation for our spring system follows:

$$ F_{total} - F_{spring} - F_{damp} = 0$$ $$m \ddot{x} + \beta \dot{x} + kx = 0 $$ $$\blbox{m \frac{d^2 x}{dt^2} + \beta \frac{dx}{dt} + kx = 0} \tag{Eq. 1} $$

Eq. 1 - Linear, second-order, ordinary differential equation. Force assocaited with a mass on a spring, with friction accounted for in our second term.

Undoubtably, constructing (Eq. 1) was far easier than jumping straight to its solution, x(t). In solving (Eq. 1), we will demonstrate that x(t) is not trivial, and would not have been easily-obtainable with algebra or empirical methods alone.


Complex-Valued Solution to the Differential Equation for Real Spring Motion

For completeness, it is necessary to consider x(t) to be complex-valued when solving our differential equation. However, our physical (observable) system is described only by its real part (not its imaginary). We will write Re[x(t)] after working out our complex solution.

First observation: (Eq. 1) is actually homogonous in addition to being linear and ordinary. Given these properties, our complete (most general) solution must satisfy the following:

Examine the most genearlized oscillatory function we can come up with, but label it as being a part of some index j to remind ourselves that we expect to find exactly two solutions of x(t). In other words, j={1,2}: $$x_n(t)=C_n e^{i \omega_n t} \tag {Eq. 2.1}$$

Eq. 2.1 - $C_n$ are complex-valued constants, and may have imaginary parts!

Our most generalized solution is some linear combination of these two solutions:

$$\blbox{x(t)=C_1 e^{i \omega_1 t} + C_2 e^{i \omega_2 t}} \tag{Eq. 2.2} $$

Reduction to an Eigenvalue Problem

Substitute this full solution into our equation of motion to start determining our solution parameters ($C_n, \omega_n$):

$$ \frac{d^2}{dt^2} [C_n e^{i \omega_n t}] + \left(\frac{\beta}{m}\right) \frac{d}{dt} [C_n e^{i \omega_n t}] + \left(\frac{k}{m}\right) [C_n e^{i \omega_n t} ]= 0 $$ Perform our time-derivatives in terms one and two: $$ i^2 \omega_n^2 [C_n e^{i \omega_n t}] + \left(\frac{\beta}{m}\right) i \omega_n[C_n e^{i \omega_n t}] + \left(\frac{k}{m}\right) [C_n e^{i \omega_n t} ]= 0 $$ Pull out common factors to reveal a quadratic equation, and our eigenvalue problem. Recall $i^2=-1$: $$\blbox{C_n e^{i \omega_n t} \left( - \omega_n^2 + i\frac{\beta}{m}\omega_n + \frac{k}{m} \right) = 0} \tag {Eq. 3.1} $$

We need to eliminate this exponential term in order to solve the auxillary (or secular) equation (3.2).

$$\blbox{ - \omega_n^2 + i\frac{\beta}{m}\omega_n + \frac{k}{m} = 0 } \tag{Eq. 3.2}$$

Eq. 3.2 - Polynomial which is degree-two in $\omega_n$.

The upcoming math may be simplified by defining the following, algebraic substitutions: $$\begin{align} \beta /m \rightarrow 2\gamma \\ k/m \rightarrow \kappa^2 \end{align}$$

Solve the auxillary equation to obtain two complex-valued eigenvalues:

$$ - \omega_n^2 + 2i\gamma \omega_n + \kappa^2 = 0 $$ $$\blbox{\omega_{1,2} = i\gamma \pm \sqrt{\kappa^2 - \gamma^2}} \tag {Eq. 3.3}$$ Therefore, our general solution for damped spring motion is given by substituting (Eq. 3.3) in (Eq. 2.2): $$ x(t)= C_1 e^{i^2 \gamma t} e^{+i \left(\sqrt{\kappa^2 - \gamma^2}\right)t} + C_2 e^{i^2 \gamma t} e^{-i \left(\sqrt{\kappa^2 - \gamma^2}\right)t} $$ $$ \blbox{x(t)= e^{-\gamma t} \left(C_1 e^{+i \left(\sqrt{\kappa^2 - \gamma^2}\right)t} + C_2 e^{-i \left(\sqrt{\kappa^2 - \gamma^2}\right)t} \right) } \tag{Eq. 4}$$ We already see an interesting quality of our solution to damped motion: our particle is oscillating with some new (relative to undamped motion) frequency that is related by its natural frequency ($\kappa$) and the systen damping (encoded in $\beta$).

Interestingly, x(t) is a product of two distinctly-shaped functions: a factor that decays purely-exponentially in time, and a purely-oscillatory factor:

damping1

Figure 1 - Plot of (Eq. 4) for arbitrary parameters ($\gamma \lt \kappa$). Predicts oscillating motion that decays in amplitude over time. Matches our expectation of how a damped harmonic oscillator will behave. Changes shape if $\gamma \ge \kappa$ (critical/overdamping).


The Real-Valued Part of x(t)

Our complex-valued solution is the most complete description for damped oscillations in 1-D. However, the imaginary part ( Im[x(t)] ) does not contribute to any observable behavior, so we may disregard it and pursue a trigonometric form for (Eq. 4). To drop the imaginary part and clean notation, make the following substitutions:

$$ x(t) \rightarrow \text{Re}[x(t)] $$ $$ C_n \rightarrow A_n+iB_n $$ $$ \sqrt{\kappa^2 - \gamma^2} \rightarrow \Omega$$

Note that $A_n, B_n$ are real-valued coefficients. The damped frequency, $\Omega$, varies between the natural frequency and 0 for positive damping (i.e., $\gamma>0$). Damping results in energy loss, meaning the damped frequency ($\Omega$) is always less than the natural (undamped) frequency ($\kappa$). This mathemtical result makes intuitive sense (a luxury of solving classical systems).

Given these substitutions, the real-valued part of (Eq. 4) may be expressed in a very manageable form:

$$x(t) \rightarrow e^{-\gamma t}\ \text{Re}\left[ (A_1 + iB_1) e^{+i \Omega t} + (A_2 + iB_2) e^{-i \Omega t} \right] $$ $$\begin{align} x(t) = e^{\gamma t}\ ( &\text{Re}[ A_1 \cos{\Omega t} + A_2 \cos{\Omega t} + B_1 \sin{\Omega t} - B_2 \sin{\Omega t}] \\ - &\text{Re}[ i A_1 \sin{ \Omega t} - i A_2 \sin{\Omega t} - iB_1 \cos{\Omega t} - iB_2 \cos{ \Omega t} ] ) \end{align} \tag{Eq. 5}$$

I have carefully grouped terms in (Eq. 5) such that it is clear which of our trigonometric terms are purely real versus purely imaginary -- our real solution for x(t) is clearly visible:

$$x(t) = e^{-\gamma t} \left[ (A_1 + A_2) \cos{\Omega t} + (B_1 - B_2) \sin{\Omega t} \right] $$ Compressing our solution constants yet again, we obtain an elegant solution to damped harmonic motion: $$\blbox{x(t) = e^{-\gamma t} \left[ A\cos{\Omega t} + B{\sin{\Omega t}} \right]} \tag {Eq. 6} $$

Eq. 6 - The real-valued solution to Newton's equation for damped, harmonic motion (Eq. 1). The damped oscillation-frequency ($\Omega$) may be imaginary, necessitating the use of complex analysis (or Eq. 4) to evaluate behavior when $\gamma > \kappa$.

re-underdamped.png re-critical.bmp re-underdamped.bmp

Figure 2 - Plots of x(t) for different damping strengths. The complex nature of x(t) encodes our lack of real roots in the overdamped case, and we lose our oscillatory behavior, slowing the system before equillibrium is reached (hence no real roots).

Our eigenfrequency $\Omega$ is influenced by the damping strength, implying that the mechanism for energy loss causes decay in amplitude in addition to down-shifting the oscillation periodicity from its undamped value (when $\gamma=0$). Indeed, the overdamped case is an exaggerated form of why amplitude decay must be occompanied by a down-shift in oscillation frequency.


Brief Verification of our Solution

Prior to a discussion, we should demonstrate explicitly that x(t) really does solve (Eq. 1). Use the complex-valued form (Eq. 4) over the real part (Eq. 6) in order to keep terms compact and avoid errors.

$$ \ddf{x}{t} + 2\gamma \frac{dx}{dt}+\kappa^2 x = 0$$ $$ \frac{d^2}{dt^2} \left[ e^{-\gamma t} \left( C_1 e^{i \Omega t} + C_2 e^{-i \Omega t} \right) \right] + 2\gamma \frac{d}{dt} \left[ e^{-\gamma t} \left( C_1 e^{i \Omega t} + C_2 e^{-i \Omega t} \right) \right] + \kappa^2 \left[ e^{-\gamma t} \left( C_1 e^{i \Omega t} + C_2 e^{-i \Omega t} \right) \right] \stackrel{?}{=} 0 $$ $$ -\left( \Omega^2 -\kappa^2 + \gamma^2) e^{-\gamma t} (C_1 e^{i \Omega t} + C_2 e^{-i \Omega t} \right) \stackrel{?}{=} 0 $$ $$ \cancelto{0}{([ \kappa^2 - \gamma^2 ] -\kappa^2 + \gamma^2)} e^{-\gamma t} \left( C_1 e^{i \Omega t} + C_2 e^{-i \Omega t} \right) \stackrel{\unicode{10004}}{=} 0 $$ Therefore, x(t) is a solution to (Eq. 1) for any choice for $\beta, m, k$, and t.

Understanding Initial Conditions -- A, B

For a brief moment, consider how certain configurations of A and B influence x(t) at initial time $t=t_0$:

A0.png B0.bmp AB.bmp

Figure 3 - Plots of motion for different values of A,B, illustrating that these parameters vary according to initial system configuration.