Massive string

We have considered systems of infinite masses connected by massless string. We will now consider what happens to a massive string. We start with the familiar system of masses:

We write the equation of motion in terms of xx instead of jj.

y¨(x)=Tma(y(xa)2y(x)+y(x+a)). \ddot y(x) = \frac{T}{ma} (y(x-a) -2y(x) + y(x+a)).

As each mass decreases and the distance between them shrinks this system begins to approximate a massive string. We substitute in ρL=masslength=ma\rho_L = \frac{\text{mass}}{\text{length}} = \frac ma.

y¨(x)=TρLa2(y(xa)2y(x)+y(x+a)). \ddot y(x) = \frac{T}{\rho_L a^2} (y(x-a) - 2y(x) +y(x+a)).

Now let m,a0m,a \to 0 such that ρL\rho_L remains constant. We Taylor expand y(xa)y(x-a) and y(x+a)y(x+a). Since a0a \to 0 the first few terms of the Taylor series are a very good approximation.

y(xa)y(x)ay(x)1!+a2y(x)2y(x+a)y(x)+ay(x)1!+a2y(x)2. \begin{align*} y(x-a) &\approx y(x) - \frac{ay'(x)}{1!} + \frac{a^2y''(x)}{2} \\ y(x+a) &\approx y(x) + \frac{ay'(x)}{1!} + \frac{a^2y''(x)}{2}. \end{align*}

We substitute our approximations in and see that things cancel nicely.

y¨(x)=TρLa2(yay+12a2y2y+y+ay+12a2y)=TρLy(x)2yt2=TρL2yx2. \begin{align*} \ddot y(x) &= \frac{T}{\rho_La^2} (y-ay' + \frac12a^2y''-2y+y+ay'+\frac12a^2y'') \\ &= \frac{T}{\rho_L} y''(x) \\ \frac{\partial^2 y}{\partial t^2} &= \frac{T}{\rho_L} \frac{\partial^2y}{\partial x^2}. \tag{1} \end{align*}

Let’s consider the dispersion relation ω(K)\omega(\mathcal K) as a0a \to 0.

ω2=2ω02(1cosKa)cosKa1(Ka)22ω2ω02K2a2=TK2ρLωTρLK. \begin{align*} \omega^2 &= 2\omega_0^2(1-\cos\mathcal Ka) \\ \cos\mathcal Ka &\approx 1-\frac{(\mathcal Ka)^2}{2} \\ \omega^2 &\approx \omega_0^2 \mathcal K^2 a^2 = \frac{T\mathcal K^2}{\rho_L} \\ \omega &\approx \sqrt{\frac{T}{\rho_L}} \mathcal K. \end{align*}

We now generalize (1) by substituting vp2=TρLv_p^2 = \frac{T}{\rho_L} and replacing yy with ψ\psi. This gives us the general wave equation:

2ψt2=vp22ψx2. \frac{\partial^2\psi}{\partial t^2} = v_p^2 \frac{\partial^2\psi}{\partial x^2}.

We guess a solution ψ(x,t)=A(x)B(t)\psi(x,t) = A(x)B(t).

2(A(x)B(t))t2=vp22(A(x)B(t))x2A(x)2B(t)t2=vp2B(t)2A(x)x21vpB(t)2B(t)t2=1A(x)2A(x)x2. \begin{align*} \frac{\partial^2 (A(x) B(t))}{\partial t^2} &= v_p^2 \frac{\partial^2 (A(x)B(t))}{\partial x^2} \\ A(x) \frac{\partial^2 B(t)}{\partial t^2} &= v_p^2 B(t) \frac{\partial^2 A(x)}{\partial x^2} \\ \frac{1}{v_p B(t)} \frac{\partial^2 B(t)}{\partial t^2} &= \frac{1}{A(x)} \frac{\partial^2 A(x)}{\partial x^2}. \tag{2} \end{align*}

Notice that the left side of (2) is a function only of tt and the right side is a function only of xx. The only way this equality can hold for all tt and xx is if it is constant. We call this constant value Km2-\mathcal K_m^2 since it will make our result useful later. We can now separate equation (2) and set each side equal to Km2-\mathcal K_m^2.

1vp2B(t)2Bt2=Km2B¨=Km2vp2BB=Bmsin(vpKmt+βm). \begin{align*} \frac{1}{v_p^2 B(t)} \frac{\partial^2B}{\partial t^2} &= -\mathcal K_m^2 \\ \ddot B &= -\mathcal K_m^2 v_p^2 B \\ B &= B_m \sin(v_p \mathcal K_m t + \beta_m). \end{align*}

Similarly

1A(x)2A(x)x2=Km2A=Km2AA=Amsin(Kmx+αm). \begin{align*} \frac{1}{A(x)} \frac{\partial^2 A(x)}{\partial x^2} &= -\mathcal K_m^2 \\ A'' &= -\mathcal K_m^2 A \\ A &= A_m \sin(\mathcal K_m x + \alpha_m). \end{align*}

We plug AA and BB into our guess to find the general solution:

ψm(x,t)=Amsin(Kmx+αm)sin(ωmt+βm)ωm=vpKm \begin{align*} \psi_m (x,t) &= A_m \sin(\mathcal K_m x + \alpha_m) \sin(\omega_m t + \beta_m) \\ \omega_m &= v_p \mathcal K_m \end{align*}

where mm stands for the mthm^\text{th} normal mode.

Boundary conditions

We have found the general solution for infinite massive string, now we will consider how this solution applies to finite systems. Consider a string fixed between two walls as pictured below.

From the boundary conditions

ym(x,t)=Amsin(Kmx+αm)sin(ωmt+βm)ym(0,t)=0=Amsin(αm)sin(ωmt+βm)αm=0ym(L,t)=0=Amsin(KmL)sin(ωmt+βm)Km=mπL. \begin{align*} y_m(x,t) &= A_m \sin (\mathcal K_m x + \alpha_m) \sin(\omega_m t + \beta_m) \\ y_m(0,t) &= 0 = A_m \sin(\alpha_m) \sin(\omega_m t + \beta_m) \\ \alpha_m &= 0 \\ y_m(L,t) &= 0 = A_m \sin (\mathcal K_m L) \sin(\omega_m t + \beta_m) \\ \mathcal K_m &= \frac{m \pi}{L}. \end{align*}

Which means our solution for a string of length LL fixed between two walls is

ym(x,t)=Amsin(mπxL)sin(ωmt+βm). y_m(x,t) = A_m \sin\left(\frac{m \pi x}{L}\right) \sin(\omega_m t + \beta_m).

Where AmA_m and βm\beta_m are defined by initial conditions.

Initial conditions

Consider a string released from rest in the initial configuration shown below. f(x)f(x) models the initial displacement of the string.

f(x)={hif aLxbL0otherwise f(x) = \begin{cases} h & \text{if } aL \le x \le bL \\ 0 & \text{otherwise} \end{cases}

Since the string is released from rest, the initial velocity is zero.

ytt=0=0=ωmAmsin(mπxL)cos(βm) \left. \frac{\partial y}{\partial t} \right|_{t=0} = 0 = \omega_m A_m \sin\left(\frac{m \pi x}{L}\right) \cos(\beta_m)

So βm=π2+lπ\beta_m = \frac\pi2 + l \pi. However, all ll values will give the same solution, so we can just say βm=π2\beta_m = \frac\pi2. We find our solution is the sum of the normal modes

y(x,t)=m=1Amsin(mπxL)cos(ωmt). y(x,t) = \sum_{m=1}^\infty A_m \sin \left(\frac{m \pi x}{L}\right) \cos(\omega_m t).

Now to plug in our initial condition we want y(x,0)=f(x)=m=1Amsin(mπxL)y(x,0) = f(x) = \sum_{m=1}^\infty A_m \sin\left(\frac{m\pi x}{L}\right). Note that this is a Fourier series. Following the derivation on the Fourier series page, we find

An=2L0Lsin(nπxL)f(x)dx. A_n = \frac{2}{L} \int_0^L \sin\left(\frac{n\pi x}{L}\right) f(x) dx.

Using the definition of f(x)f(x) above, we evaluate the integral. When x[aL,bL]x \notin [aL, bL], f(x)=0f(x) = 0 so only the part of the integral from aLaL to bLbL contributes.

An=2hLaLbLsin(nπxL)dx=2hnπ(cos(nπa)cos(nπb)). \begin{align*} A_n &= \frac{2h}{L} \int_{aL}^{bL} \sin\left(\frac{n\pi x}{L}\right) dx \\ &= \frac{2h}{n\pi}\left( \cos(n \pi a) - \cos(n \pi b) \right). \end{align*}

Which gives us the solution

y=m=12hmπ(cos(nπa)cos(nπb))sin(mπxL)cos(ωmt). y = \sum_{m=1}^\infty \frac{2h}{m\pi} \left( \cos(n \pi a) - \cos(n \pi b) \right) \sin \left(\frac{m\pi x}{L}\right) \cos(\omega_m t).

We visualize this solution (with the sum up to 100 terms) below.

Traveling wave

As shown here, a traveling wave f(x+vt)f(x+vt) solves the wave equation. A traveling wave makes it much easier to incorporate initial conditions than a series of infinite standing waves as we used above. Let’s consider how we can use this.

We saw from the visualization above that our initial wave split into two waves of half the amplitude traveling in opposite directions. Let f(x)f(x) be the initial displacement of the string. Then we can model the traveling wave as

y=f(x+vt)+f(xvt)2. y = \frac{f(x+vt) + f(x-vt)}{2}.

Energy in oscillating massive string

Since we have not considered any damping so far, energy must be conserved in our system. Let us consider how energy is stored along an oscillating string.

We first consider the kinetic energy along the string. Kinetic energy K=12mv2K = \frac12 mv^2, where dm=ρdxdm = \rho \, dx and v=ytv = \frac{\partial y}{\partial t}. This gives us dK=12ρdx(yt)2dK = \frac12 \rho \, dx \left(\frac{\partial y}{\partial t}\right)^2.

We integrate to find

K=0L12ρ(yt)2dx. K = \int_0^L \frac12 \rho \left(\frac{\partial y}{\partial t}\right)^2 dx.

Now consider potential energy U=W=FdsU = W = Fds.

From the figure above we see that the displacement in string length ds=dx2+dy2dxds = \sqrt{dx^2 + dy^2} - dx.

dU=T(dx2+dy2dx)=T(1+(dydx)21)dx12T(dydx)2dxU=T20L(dydx)2dx. \begin{align*} dU &= T(\sqrt{dx^2 + dy^2} - dx) \\ &= T\left(\sqrt{1 + \left(\frac{dy}{dx}\right)^2} - 1\right)dx \\ &\approx \frac12 T \left(\frac{dy}{dx}\right)^2 dx \\ U &= \frac{T}{2}\int_0^L \left(\frac{dy}{dx}\right)^2 dx. \end{align*}

Power flowing across an oscillating massive string

We want to know how much power is flowing across a point on the wave. Power has units of energy over time, or force times velocity. Consider the force from the left acting on an infinitesimal mass along the string. We only care about the force from the left and not the net force because we want to find the power flowing from left to right.

Because θ\theta is very small (since we are considering an infinitesimal slice of the string), we can approximate sinθtanθdydx\sin \theta \approx \tan \theta \approx \frac{dy}{dx}. This gives us the force in the vertical direction FTdydxF \approx -T \frac{dy}{dx}. Multiplying FvF \cdot v we get

P=Tyxyt. P = -T \frac{\partial y}{\partial x} \frac{\partial y}{\partial t}.

For a more detailed explanation of energy and power in a string see this page by Mathew Schwartz at Harvard.