Skip to main content
\(\newcommand{\trace}{\operatorname{tr}} \newcommand{\real}{\operatorname{Re}} \newcommand{\imaginary}{\operatorname{Im}} \newcommand{\lt}{<} \newcommand{\gt}{>} \newcommand{\amp}{&} \)

Section2.1Homogeneous Linear Equations

Objectives
  • To understand that a second-order linear differential equation with constant coefficients is an equation of the form

    \begin{equation*} a x'' + bx' + cx = 0. \end{equation*}

    and can be solved by examining the roots of the characteristic polynomial \(a \lambda^{2} + b\lambda + c = 0\text{.}\)

  • To understand that simple harmonic oscillator can be modeled by the equation

    \begin{equation*} m \frac{d^2 x}{dt^2} + b \frac{dx}{dt} + k x = 0, \end{equation*}

    where \(m \gt 0\text{,}\) \(k \gt 0\text{,}\) and \(b \geq 0\text{.}\)

A differential equation of the form

\begin{equation*} a(t) x'' + b(t) x' + c(t) x = g(t) \end{equation*}

is called a second-order linear differential equation. We will first consider the case

\begin{equation*} a x'' + b x' + cx = 0, \end{equation*}

where \(a\text{,}\) \(b\text{,}\) and \(c\) are constants and \(a \neq 0\text{.}\) An equation of this form is said to be homogeneous with constant coefficients. In this section, we will develop a technique to solve differential equations of this form.

Subsection2.1.1RLC Circuits

Recall the RC circuits that we studied earlier (see Section1.3). Such circuits contained a voltage source, a capacitor, and a resistor. A battery or generator is an example of a voltage source, and a toaster or an electric stove is an example of something that might provide a resistance in a circuit. Capacitors store an electrical charge and are used in electronic flashes for cameras. We will now add an inductor such as a solenoid, a coil that generates a magnetic field. Inductor applications include transformers, power supplies, televisions, and radios. Our new circuit is called an RLC circuit (Figure2.1.1).

an electrical circuit containing a voltage source, and resistor, a capacitor, and an inductor
Figure2.1.1An RLC Circuit

Current, \(I(t)\text{,}\) is the rate at which a charge flows through this circuit and is measured in amperes or amps. We assign a direction to the current, and a current flowing in the opposite direction will be given negative values. The impressed voltage, \(E(t)\text{,}\) is measured in volts, the resistance \(R\) is measured in ohms, and the capacitance \(C\) is measured in farads. The charge on the capacitor \(Q(t)\) at time \(t\) is measured in coulombs. Inductance on the coil, \(L\text{,}\) is measured in henrys.

The following laws from physics govern how our circuit behaves.

  • \(I = \dfrac{dQ}{dt}\text{.}\)

  • The voltage drop across a resistor is \(IR\) (Ohm's Law).

  • The voltage drop across a capacitor is \(Q/C\text{.}\)

  • The voltage drop across an inductor is \(L (dI/dt)\text{.}\)

  • In a closed circuit the impressed voltage is equal to the sum of the voltage drops in the rest of the circuit (Kirchhoff's Second Law).

Applying Kirchhoff's Second Law to our circuit, we have the differential equation

\begin{equation} L \frac{dI}{dt} + RI + \frac{1}{C} Q = E(t)\label{secondorder01-equation-RLC}\tag{2.1.1} \end{equation}

or

\begin{equation*} LQ'' + RQ' + \frac{1}{C} Q = E(t). \end{equation*}

Differentiating both sides of (2.1.1), we have

\begin{equation*} L I'' + RI' + \frac{1}{C} I = E'(t). \end{equation*}

For example, we might consider an RLC circuit with \(R = 1\text{,}\) \(L = 1\text{,}\) and \(C = 1\text{.}\) At \(t = 0\) when both \(I(0) =0\) and \(I'(0) = Q(0) = 0\text{,}\) the impressed voltage on the circuit is given by \(E(t) = \sin(t)\text{.}\) Our equation becomes

\begin{equation*} I'' + I' + I = E'(t) = \cos t. \end{equation*}

This is an example of a second-order linear differential equation.

Subsection2.1.2Solving Second-Order Linear Equations

Suppose that we have a homogeneous second-order linear differential equation with constant coefficients,

\begin{equation} a x'' + b x' + cx = 0.\label{secondorder01-equation-constant-coef}\tag{2.1.2} \end{equation}

The goal of this section is to be able to solve all such equations. To build some intuition, we start by considering the homogeneous first-order linear differential equation

\begin{equation} ax' + bx = 0.\label{secondorder01-fo-equation-constant-coef}\tag{2.1.3} \end{equation}

Our work in Chapter1 suggests that this equation has a solution of the form \(x(t) = e^{\lambda t}\) for some constant \(\lambda\text{.}\) We can identify \(\lambda\) by substuting our guess for the solution into (2.1.3) and solving algebraically. Since \(\frac{d}{dt}e^{\lambda t} = \lambda e^{\lambda t}\text{,}\) this substitution yields the equation

\begin{equation*} a\lambda e^{\lambda t} + be^{\lambda t} = 0. \end{equation*}

The term \(e^{\lambda t}\) is always non-zero, and therefore we can safely divide both sides of the equation above by \(e^{\lambda t}\) to reach

\begin{equation*} a\lambda + b = 0, \end{equation*}

which has solution \(\lambda = -\frac{b}{a}\text{.}\) Thus (2.1.3) has \(x(t) = e^{-\frac{b}{a}t}\) as a solution.

A similar strategy works for homogeneous second-order linear differential equations. We guess that (2.1.2) has a solution of the form \(x=e^{\lambda t}\) and substitute our guess into the left-hand side of the equation. In detail,

\begin{align*} a\left(\frac{d^2}{dt^2} e^{\lambda t}\right) + b\left(\frac{d}{dt}e^{\lambda t}\right) + c e^{\lambda t} & = a\lambda^{2} e^{\lambda t} + b \lambda e^{\lambda t} + c e^{\lambda t}\\ & = (a\lambda^{2} + b\lambda + c) e^{\lambda t} \end{align*}

Dividing out the non-zero term \(e^{\lambda t}\) as before, we see that \(\lambda\) must satisfy

\begin{equation} a\lambda^{2} + b\lambda + c = 0.\label{secondorder01-equation-characteristic-eq}\tag{2.1.4} \end{equation}

We call (2.1.4) the characteristic equation for \(ax'' + bx' + cx = 0\text{.}\) Its roots are

\begin{equation*} \lambda = \frac{-b \pm \sqrt{b^2 - 4ac}}{2a}. \end{equation*}

If \(b^2 - 4ac \gt 0\text{,}\) we get distinct real roots

\begin{equation*} \lambda_1 = \frac{-b + \sqrt{b^2 - 4ac}}{2a} \quad \text{and} \quad \lambda_2 = \frac{-b - \sqrt{b^2 - 4ac}}{2a}, \end{equation*}

which correspond to a pair of solutions

\begin{equation*} x_{1}(t) = e^{\lambda_{1} t} \quad \text{and} \quad x_{2}(t)=e^{\lambda_{2} t}. \end{equation*}

to (2.1.2). Each solution \(x_{i}(t)\) exhibits exponential growth or decay depending on the sign of \(\lambda_{i}\text{.}\)

The equation

\begin{equation} x'' + 3 x' - 10 x = 0\label{secondorder01-equation-guess}\tag{2.1.5} \end{equation}

has characteristic equation \(\lambda^{2} + 3\lambda - 10 = 0\text{.}\) Rather than using the quadratic formula, we can simply factor the left-hand side as \((\lambda + 5)(\lambda-2)\) to deduce that the roots are \(\lambda_{1} =-5\) and \(\lambda_{2} = 2\text{.}\) The roots correspond to a pair of solutions to (2.1.5):

\begin{equation*} x_1(t) = e^{-5t} \mbox{ and } x_2(t) = e^{2t}. \end{equation*}

The characteristic equation (2.1.4) also has distinct roots when \(b^{2} - 4ac \lt 0\text{,}\) although now the roots are complex numbers. Even so, we can make sense of the corresponding solutions using tools from Section1.8. In this case, it is customary to write \(\lambda = \alpha \pm i\beta\text{,}\) where

\begin{equation*} \alpha = - \frac{b}{2a} \quad \text{and} \quad \beta = \frac{\sqrt{4ac - b^2}}{2a}. \end{equation*}

We then have a pair of complex conjugate roots

\begin{equation*} \lambda_{1} = \alpha + \beta i \quad \text{and} \quad \lambda_{2} = \alpha - \beta i \end{equation*}

corresponding, as in the real case, to a pair of solutions

\begin{equation*} x_{1}(t) = e^{(\alpha + \beta i)t} \quad \text{and} \quad x_{2}(t) = e^{(\alpha - \beta i)t}. \end{equation*}

We can rewrite \(x_{1}(t)\) as follows:

\begin{align*} x_{1}(t) & = e^{(\alpha + \beta i)t}\\ & = e^{\alpha t + \beta i t}\\ & = e^{\alpha t}e^{\beta i t} \end{align*}

From Euler's formula (1.8.1), we have \(e^{i\beta t}=\cos(\beta t) + i\sin(\beta)\text{,}\) and so

\begin{equation*} x_{1}(t) = e^{\alpha t}(\cos(\beta t) + i\sin(\beta t)). \end{equation*}

Using the facts that \(\cos(-\beta t)=\cos(\beta t)\) and \(\sin(-\beta t) = -\sin(\beta t)\text{,}\) a similar computation on \(x_{2}(t)\) results in the expression

\begin{equation*} x_{2}(t) = e^{\alpha t}(\cos(\beta t) - i\sin(\beta t)). \end{equation*}

To get real solutions, we can take the real part of \(x_{1}(t)\) and (negative) the imaginary part of \(x_{2}(t)\text{.}\) Going forward, we will use the resulting real functions as the canonical solutions to (2.1.2) when \(b^{2} - 4ac \lt 0\text{,}\) re-defining \(x_{1}(t)\) and \(x_{2}(t)\) as

\begin{equation*} x_{1}(t) = e^{\alpha t}\cos(\beta t) \quad \text{and} \quad x_{2}(t) = e^{\alpha t}\sin(\beta t). \end{equation*}

Note that complex roots of the characteristic equation induce oscillatory behavior in the solutions of the differential equation due to the presence of the \(\sin\) and \(\cos\) terms.

Consider the equation

\begin{equation*} x'' + 4 x' + 13 x = 0 \end{equation*}

The characteristic equation is \(\lambda^{2} + 4\lambda + 13 = 0\text{.}\) The quadratic formula gives roots

\begin{equation*} \lambda = \frac{-4 \pm \sqrt{4^2-4(1)(13)}}{2} = -2 \pm 3i, \end{equation*}

corresponding to \(\lambda = \alpha \pm \beta i\) with \(\alpha=-2\text{,}\) \(\beta = 3.\) We get the complex solutions:

\begin{equation*} z_{1}(t) = e^{(-2+3i)t} \quad \text{and} \quad z_{2}(t) = e^{(-2-3i)t}. \end{equation*}

From these complex solutions, using the techniques of Subsection1.8.4, from these solutions, we can get the following equivalent set of real solutions:

\begin{equation*} x_{1}(t) = e^{-2t}\cos(3t) \quad \text{and} \quad x_{2}(t) = e^{-2t}\sin(3t). \end{equation*}

There is one more case to consider: when \(b^{2}-4ac=0\text{.}\) In this case, the charactersitic equation has a repeated real root at \(\lambda = -\frac{b}{2a}\text{,}\) which appears to give only one solution \(x_{1}(t) = e^{-\frac{b}{2a}t}.\) However, as demonstrated in the next example, it is possible to generate a second solution using \(x_{1}(t)\text{.}\)

Consider the equation

\begin{equation*} x'' + 2x' + x = 0. \end{equation*}

The characteristic equation is

\begin{equation*} \lambda^{2} + 2\lambda + 1 = (\lambda + 1)^{2} = 0. \end{equation*}

Thus, there is a repeated real root \(\lambda = -1\) and we have a solution \(x_1(t) = e^{-t}\text{.}\)

We seek a second solution that is not a multiple of \(x_1(t) = e^{-t}\text{.}\) We will guess that a solution to \(x'' + 2x' + x = 0\) of the form \(v(t) x_1(t)\) exists for some unknown function \(v(t)\) and attempt to solve for \(v(t)\text{.}\) Indeed, if

\begin{equation*} x(t) = v(t) x_1(t) = v(t) e^{-t}, \end{equation*}

then

\begin{equation*} x'(t) = v(t) x_1'(t) + v'(t) x_1(t) = -v(t) e^{-t} + v'(t) e^{-t} \end{equation*}

and

\begin{align*} x''(t) & = v''(t) x_1(t) + 2v'(t) x_1'(t) + v(t) x_1''(t)\\ & = v''(t) e^{-t} - 2 v'(t) e^{-t} + v(t) e^{-t}. \end{align*}

Consequently,

\begin{align*} x'' + 2 x' + x & = [v'' e^{-t} - 2 v' e^{-t} + v e^{-t}] + 2 [-v e^{-t} + v' e^{-t}] + [v e^{-t}]\\ & = e^{-t} v''\\ & = 0, \end{align*}

and \(v'' = 0\text{.}\) Therefore, \(v = c_1 t + c_2\text{.}\) Letting \(c_1 = 1\) and \(c_2 = 0\) gives \(v(t) = t\text{,}\) and a second solution to our equation is given by

\begin{equation*} x(t) = t e^{-t}\text{.} \end{equation*}

Note that we could just as well have chosen, say, \(c_{1} = 0\) and \(c_{2}=1\text{,}\) or \(c_{1} = 1\) and \(c_{2}=1\text{.}\) The former, however, results in \(x(t)=e^{-t}\text{,}\) which is the solution we already knew. The latter choice gives \(x(t) = te^{-t} + e^{-t}\text{.}\) While this solution is distinct from the one we started with, it still contains information from the original solution in the second term. The assignment \(c_{1}=1\text{,}\) \(c_{2}=0\) is the simplest choice of constants which gives another solution that is, in a sense, independent from the first.

The technique that we have used in Example2.1.4 is called reduction of order. In general, given a second-order linear differential equation

\begin{equation*} a x'' + bx' + cx = 0 \end{equation*}

such that \(b^2 - 4ac = 0\text{,}\) we have a pair of solutions

\begin{equation*} x_{1}(t) = e^{\lambda t} \quad\text{and}\quad x_{2}(t)= t e^{\lambda t}, \end{equation*}

where \(\lambda = -b/2a\) is the repeated real root of the characteristic equation.

Subsection2.1.3General Solutions to Second-Order Linear Equations

In Subsection2.1.2 we learned how to construct some solutions to the differential equation (2.1.2) given by

\begin{equation*} ax'' + bx' + cx = 0 \end{equation*}

using the roots of the characteristic equation \(a\lambda^{2} + b \lambda + c = 0.\) In this section, we will show that these solutions can in fact be used to describe all solutions to the differential equation.

A key idea in this direction is the principle of superposition, which states that scalar multiples and sums of solutions to (2.1.2) are again solutions to the same equation.

This fact can be established using basic properties of the derivative. For instance, suppose that \(x(t)\) is a solution to \(ax'' + bx' + cx = 0\text{,}\) and let \(A\) be an arbitrary real constant. Recalling that for any differentiable function \(f\text{,}\) we have \(\frac{d}{dt}Af(t)=A\frac{d}{dt}f(t)\text{,}\) we may compute

\begin{align*} a(Ax)'' + b(Ax)' + c(Ax) & = aA(x'') + bA(x') + cA(x)\\ & = A(ax'' + bx' + cx) \\ & = A \cdot 0 \\ & = 0 \end{align*}

where in the second-to-last line we use the fact that \(x(t)\) is a solution to (2.1.2), that is, \(ax'' + bx' + cx = 0\text{.}\) So, \(Ax(t)\) is also a solution.

Similarly, since \(\frac{d}{dt}(f(t) + g(t)) = \frac{d}{dt}f(t) + \frac{d}{dt}g(t)\) for any differentiable \(f\) and \(g\text{,}\) if \(x_{1}(t)\) and \(x_{2}(t)\) are solutions to (2.1.2), then

\begin{align*} a(x_{1} + x_{2})'' + b(x_{1} + x_{2})' + c(x_{1} + x_{2}) & = ax_{1}'' + ax_{2}'' + bx_{1}' + bx_{2}' + cx_{1} + cx_{2} \\ & = (ax_{1}'' + bx_{1}' + cx_{1}) + (ax_{2}'' + bx_{2}' + cx_{2}) \\ & = 0 + 0 \\ & = 0. \end{align*}

So \(x_{1}(t) + x_{2}(t)\) is also a solution to (2.1.2). In conjunction, the preceding computations yield the following theorem.

Before continuing, it is worth noting that Theorem2.1.5 is still true if the constant coefficients \(a\text{,}\) \(b\text{,}\) and \(c\) are replaced with variable coefficients \(a(t)\text{,}\) \(b(t)\text{,}\) and \(c(t)\) which depend on \(t\text{.}\) For our present purposes, however, we continue to focus solely on the constant coefficient case.

The expression \(c_{1}x_{1}(t) + c_{2}x_{2}(t)\) is called a linear combination of the functions \(x_{1}\) and \(x_{2}\text{.}\) Provided that \(x_{1}\) and \(x_{2}\) solve (2.1.2) and \(x_{2}\) is not simply a constant multiple of \(x_{1}\text{,}\) we can think of the linear combination \(c_{1}x_{1}(t) + c_{2}x_{2}(t)\) as a two-parameter family of solutions to (2.1.2).

Does such a family encompass all solutions? To make this question more tractable, we first state an existence and uniqueness result for second-order linear equations which is analogous to Theorem1.6.1.

In light of Theorem2.1.6, to show that a two-parameter family of solutions \(c_{1}x_{1}(t) + c_{2}x_{2}(t)\) contains all solutions to the differential equation, it is enough to demonstrate that the constants \(c_{1}\) and \(c_{2}\) can be chosen to satisfy arbitrary initial conditions. That is, for any \(t_{0}\text{,}\) \(x_{0}\text{,}\) and \(v_{0}\text{,}\) there exists an assignment of \(c_{1}\) and \(c_{2}\) satisfying the system of equations

\begin{equation} c_{1}x_{1}(t_{0}) + c_{2}x_{2}(t_{0}) = x_{0} \label{secondorder01-IVP-equation1}\tag{2.1.6} \end{equation}
\begin{equation} c_{1}x_{1}'(t_{0}) + c_{2}x_{2}'(t_{0}) = v_{0}. \label{secondorder01-IVP-equation2}\tag{2.1.7} \end{equation}

When this holds, we say that \(x(t) = c_{1}x_{1}(t) + c_{2}x_{2}(t)\) is a general solution to the differential equation.

We begin with an example before proceding with the general theory.

In Example2.1.2, we used the characteristic equation to deduce that

\begin{equation*} x_{1}(t) = e^{-5t}\quad \text{ and } \quad x_{2}(t) = e^{2t} \end{equation*}

both solve the equation \(x'' + 3 x' - 10 x = 0\text{.}\) Taking an arbitrary linear combination of \(x_{1}\) and \(x_{2}\) then gives a two-parameter family of solutions

\begin{equation*} c_{1}e^{-5t} + c_{2}e^{2t}. \end{equation*}

(There is an easy way to verify that \(x_{1}\) and \(x_{2}\) are not constant multiples of one another: just confirm that their ratio \(\frac{e^{-5t}}{e^{2t}} = e^{-7t}\) is not constant!)

Now, let \(t_{0}\text{,}\) \(x_{0}\text{,}\) and \(v_{0}\) be arbitrary. Theorem2.1.6 gives that there is a unique solution \(x(t)\) to the differential equation satisfying \(x(t_{0}) = x_{0}\) and \(x'(t_{0}) = v_{0}\text{.}\) If we are able to choose \(c_{1}\) and \(c_{2}\) so that the system of equations

\begin{align*} c_{1}e^{-5t_{0}} + c_{2}e^{2t_{0}} &= x_{0}\\ -5c_{1}e^{-5t_{0}} + 2c_{2}e^{2t_{0}} &= v_{0}. \end{align*}

is satisfied, then \(x(t) = c_{1}e^{-5t} + c_{2}e^{2t}.\)

Solving the first equation for \(c_{1}\) gives

\begin{equation*} c_{1} = x_{0}e^{5t_{0}} - c_{2}e^{7t_{0}}. \end{equation*}

Substituting this into the second equation and solving for \(c_{2}\) results in

\begin{equation*} c_{2} = \frac{1}{7}\left( v_{0}e^{-2t_{0}} - 5x_{0} \right). \end{equation*}

We have demonstrated assignments of \(c_{1}\) and \(c_{2}\) in terms of arbitary \(t_{0}\text{,}\) \(x_{0}\text{,}\) and \(v_{0}\text{.}\) So

\begin{equation*} x(t) = c_{1}e^{-5t} + c_{2}e^{2t} \end{equation*}

is the general solution to \(x'' + 3 x' - 10 x = 0\text{.}\) Here, we established this by explicitly solving a system of equations for \(c_{1}\) and \(c_{2}\text{.}\) Going forward, we will develop some theory which will allow us to reach the same conclusion with less explicit computation.

As we saw in Example2.1.7, one can demonstrate that \(x(t) = c_{1}x_{1}(t) + c_{2}x_{2}(t)\) is a general solution by explicitly solving the system (2.1.6)-(2.1.7).

Solving the system (2.1.6)-(2.1.7) symbolically yields the following expressions for \(c_{1}\) and \(c_{2}\text{:}\)

\begin{align*} c_{1} =& \frac{x_{2}'(t_{0})x_{0} - x_{2}(t_{0})v_{0}}{x_{1}(t_{0})x_{2}'(t_{0})-x_{2}(t_{0})x_{1}'(t_{0})}\\ c_{2} =& \frac{-x_{1}'(t_{0})x_{0} + x_{1}(t_{0})v_{0}}{x_{1}(t_{0})x_{2}'(t_{0})-x_{2}(t_{0})x_{1}'(t_{0})}. \end{align*}

This represents a valid (and unique) solution to the system if and only if the denominator which appears in the expressions for both \(c_{1}\) and \(c_{2}\) is not equal to \(0\text{.}\) We call this quantity the Wronskian of \(x_{1}\) and \(x_{2}\) and denote it by

\begin{equation*} W(x_{1},x_{2}) = x_{1}(t)x_{2}'(t)-x_{2}(t)x_{1}'(t). \end{equation*}

Notice that \(W(x_{1},x_{2})\) is a function of \(t\text{.}\) Provided that \(W(x_{1},x_{2}) \neq 0\) for all values of \(t\text{,}\) by our computations above we may solve the system (2.1.6)-(2.1.7) for any values of \(t_{0}\text{,}\) \(x_{0}\text{,}\) and \(v_{0}\text{.}\) Thus we have the following criterion.

In the example below, we revisit the differential equation from Example2.1.7 in light of Theorem2.1.8.

Consider again the equation \(x'' + 3 x' - 10 x = 0\) from Example2.1.7. We have a pair of solutions \(x_{1}(t)=e^{-5t}\) and \(x_{2}(t)=e^{2t}\text{.}\) We compute their Wronskian as

\begin{equation*} W(e^{-5t}, e^{2t}) = (e^{-5t})(2e^{2t})-(e^{2t})(-5e^{-5t})=7e^{-3t}. \end{equation*}

Since for all values of \(t\) we have that \(7e^{-3t} \neq 0\text{,}\) we may conclude that

\begin{equation*} x(t)=c_{1}e^{-5t} + c_{2}e^{2t} \end{equation*}

is a general solution to \(x'' + 3 x' - 10 x = 0,\) confirming our observations from Example2.1.7.

We can now give an explicit general solution to the equation \(ax'' + bx' + cx = 0\text{.}\) This task combines our efforts from Subsection2.1.2, where we constructed pairs of solutions using the characteristic equation, with the theory developed in this section.

We prove the theorem only in the case \(b^{2} - 4ac \lt 0\) in detail, since this one is the most computationally taxing. The other two cases can be proved in the same way.

Suppose that \(b^{2} - 4ac \lt 0\text{,}\) and let \(\alpha\) and \(\beta\) be defined as in the theorem statement. We know from Subsection2.1.2 that

\begin{equation*} x_{1}(t) = e^{\alpha t}\cos(\beta t) \quad \text{ and } \quad x_{2}(t)=e^{\alpha t}\sin(\beta t) \end{equation*}

are solutions to \(ax'' + bx' + cx = 0\text{,}\) so to show that \(x(t) = c_{1}x_{1}(t) + c_{2}x_{2}(t)\) is the general solution, by Theorem2.1.8 it suffices to check that their Wronskian \(W(x_{1},x_{2})\) is non-zero for all \(t\text{.}\) We first do some preliminary computations: using the product rule,

\begin{equation*} x_{1}'(t) = \alpha e^{\alpha t}\cos(\beta t) - \beta e^{\alpha t}\sin(\beta t) \end{equation*}

and

\begin{equation*} x_{2}'(t) = \alpha e^{\alpha t}\sin(\beta t) + \beta e^{\alpha t}\cos(\beta t). \end{equation*}

Thus

\begin{equation*} x_{1}(t)x_{2}'(t) = e^{2\alpha t}[\alpha \cos(\beta t)\sin(\beta t) + \beta \cos^{2}(\beta t)] \end{equation*}

and

\begin{equation*} x_{2}(t)x_{1}'(t) = e^{2\alpha t}[\alpha \cos(\beta t)\sin(\beta t) - \beta \sin^{2}(\beta t)]. \end{equation*}

Taking the difference of these expressions results in the cancellation of the \(\alpha \cos(\beta t) \sin(\beta t)\) terms, leaving us with

\begin{equation*} x_{1}(t)x_{2}'(t) - x_{2}(t)x_{1}'(t) = \beta e^{2\alpha t}[\sin^{2}(\beta t) + \cos^{2}(\beta t)]. \end{equation*}

Using the Pythagorean Identity \(\sin^{2}(\theta) + \cos^{2}(\theta) = 1\text{,}\) we concluce

\begin{equation*} W(x_{1}, x_{2}) = \beta e^{2\alpha t}\text{.} \end{equation*}

Since \(e^{2\alpha t} \) is non-zero for all values of \(t\) and \(\beta \neq 0\text{,}\) we conclude that \(W(x_{1},x_{2})\) is non-zero for all values of \(t\text{.}\) Thus \(x(t) = c_{1}x_{1}(t) + c_{2}x_{2}(t)\) is the general solution.

With general solutions in hand, we can solve an initial value problem involving \(ax'' + bx' + cx = 0\) with initial conditions of the form \(x(0) = x_{0}\text{,}\) \(x'(0) = v_{0}\text{.}\) We demonstrate in the next example.

We solve

\begin{equation*} x'' - 8x' + 16x = 0 \end{equation*}

subject to initial conditions

\begin{equation*} x(0) = 10 \quad \text{and} \quad x'(0) = -2. \end{equation*}

We start by writing the general solution to the differential equation. The characteristic equation is

\begin{equation*} \lambda^{2} - 8\lambda + 16 = 0, \end{equation*}

which has a repeated real root at \(\lambda = 4\text{.}\) By Theorem2.1.10, the general soltuion is

\begin{equation*} x(t) = c_{1}e^{4t} +c_{2}t e^{4t}. \end{equation*}

Plugging in \(t = 0\) gives

\begin{equation*} x(0) = c_{1}e^{4\cdot 0} + c_{2} \cdot 0 \cdot e^{4 \cdot 0} = c_{1} \end{equation*}

and so \(x(0) = 10\) imposes \(c_{1} = 10\text{.}\) To solve for \(c_{2}\text{,}\) we must differentiate \(x(t)\) and use \(x'(0) = -2\text{.}\) With some computation, we find

\begin{equation*} x'(t) = (c_{1} + c_{2})e^{4t} + 4c_{2}te^{4t}. \end{equation*}

We already know \(c_{1} = 10\text{.}\) Plugging \(t = 0\) into the above and equating it to \(-2\) yields

\begin{equation*} 10 + c_{2} = -2 \end{equation*}

which gives \(c_{2} = -12.\) So the solution to the IVP is

\begin{equation*} x(t) = 10e^{4t} -12 te^{4t}. \end{equation*}
Activity2.1.1Solving Second-Order Homogeneous Linear Differential Equations

Solve each of the following initial value problems.

(a)
\begin{align*} x'' - 5 x' -14 x & = 0\\ x(0) & = 0\\ x'(0) & = 2. \end{align*}
(b)
\begin{align*} 6x'' - 11 x' + 3 x & = 0\\ x(0) & = 0\\ x'(0) & = 2. \end{align*}
(c)
\begin{align*} x'' - 14 x' + 53 x & = 0\\ x(0) & = 1\\ x'(0) & = 1. \end{align*}
(d)
\begin{align*} 4x'' - 20 x' + 41 x & = 0\\ x(0) & = 0\\ x'(0) & = 1. \end{align*}
(e)
\begin{align*} x'' -1 4 x' +49 x & = 0\\ x(0) & = 1\\ x'(0) & = 0. \end{align*}

Subsection2.1.4Classifying Harmonic Oscillators

Recall from Subsection1.1.3 that we can model harmonic motion using the equation

\begin{equation*} m \frac{d^2 x}{dt^2} + b \frac{dx}{dt} + k x = 0. \end{equation*}

In the case of a spring-mass system, \(m\) is the oscillating mass, \(b\) is the damping coefficient, and \(k\) is the spring constant. It is important to remember that both \(m\) and \(k\) are positive constants and \(b \geq 0\text{.}\)

If \(b = 0\text{,}\) then the oscillator is undamped In this case,

\begin{equation} m \frac{d^2 x}{dt^2} + k x = 0.\label{secondorder01-equation-harmonic}\tag{2.1.8} \end{equation}

The characteristic equation is

\begin{equation*} m \lambda^2 + k = 0, \end{equation*}

and we roots at \(\lambda = \pm i \sqrt{k/m}\text{.}\) Writing \(\omega = \sqrt{k/m}\) for simplicity, Theorem2.1.10 gives the general solution

\begin{equation*} x(t) = c_1 \cos \omega t + c_2 \sin \omega t, \end{equation*}

which represents the position of the mass at time \(t\text{.}\) Now, of course, it is easy to determine the velocity of the mass at time \(t\) to be

\begin{equation*} v(t) = x'(t) = - c_1 \omega \sin \omega t + c_2 \omega \cos \omega t. \end{equation*}

Suppose that an undamped harmonic oscillator is modeled by the initial value problem

\begin{align*} \frac{d^2x}{dt^2} + 9 x \amp = 0\\ x(0) \amp = 2\\ x'(0) \amp = 1. \end{align*}

We can quickly determine the solution of this initial value problem to be

\begin{align*} x(t) \amp = 2 \cos 3t +\frac{1}{3} \sin 3t\\ v(t) \amp = \cos 3t - 6 \sin 3t, \end{align*}

where \(v(t) = x'(t)\) is the velocity of the oscillator (Figure2.1.13). Examining the phase plane of the undamped oscillator, we find that the period of the oscillations is given by \(2 \pi / \omega = 2 \pi / 3 \approx 2.094\) (Figure2.1.14).

position and velocity of an undamped harmonic oscillator plotted agains time
Figure2.1.13Position and velocity of an undamped harmonic oscillator
the velocity of a harmonic oscillator plotted against the position forming an ellipse in the position-velocity plane
Figure2.1.14The phase plane of an undamped harmonic oscillator

If we add damping to the oscillator, the equation becomes

\begin{equation} m \frac{d^2 x}{dt^2} + b \frac{dx}{dt} + k x = 0.\label{secondorder01-equation-damped-oscillator}\tag{2.1.9} \end{equation}

where \(b \gt 0\text{.}\) The charactersitic equation of (2.1.9) is

\begin{equation*} m\lambda^2 + b\lambda + k = 0\text{,} \end{equation*}

which has roots

\begin{equation} \lambda = \frac{-b \pm \sqrt{b^2 - 4mk}}{2m}.\label{secondorder01-equation-characteristic-eq-roots}\tag{2.1.10} \end{equation}

There are three possible types of types of motion for the oscillator depending on the nature of the roots of (2.1.10).

  • If the damping value of \(b\) is small when compared to \(4mk\text{,}\) then \(b^2 - 4mk \lt 0\) and the roots of (2.1.10) will be complex. Furthermore, the real part of each root, \(-b/ 2m\text{,}\) is always negative. In such a situation, we say that the oscillator is under-damped.

  • If the damping value of \(b\) is large , then \(b^2 - 4mk \gt 0\text{,}\) and we obtain distinct real negative roots for (2.1.10). The oscillator is over-damped.

  • Finally, we say that the oscillator is critically-damped if \(b^2 - 4mk = 0\text{.}\)

Suppose that an oscillator is modeled by the initial value problem

\begin{align*} \frac{d^2 x}{dt^2} + 0.4 \frac{dx}{dt} + 1.04 x \amp = 0\\ x(0) \amp = 0\\ x'(0) \amp = 1. \end{align*}

Notice that the damping \(b = 0.4\) is very small compared with the spring constant \(k = 1.04\text{.}\) The characteristic equation of the differential equation is \(\lambda^2 + 0.4 \lambda + 1.04 = 0\text{,}\) which has roots \(\lambda = -0.2 \pm i\text{.}\) Therefore, the general solution is

\begin{equation*} x(t) = c_1 e^{-0.2t} \cos t + c_2 e^{-0.2t} \sin t. \end{equation*}

Applying the initial conditions, our solution becomes

\begin{align*} x(t) \amp = e^{-0.2t} \sin t\\ v(t) \amp = x'(t) = e^{-0.2t} (\cos t - 0.2 \sin t). \end{align*}

Notice that the period of the oscillations, \(2 \pi / \omega = 2 \pi \approx 6.283\text{,}\) does not change; however, the amplitude slowly decreases (Figure2.1.16 and Figure2.1.17).

velocity and position plotted against time with oscillations becoming smaller as time increases
Figure2.1.16Position and velocity of an under-damped harmonic oscillator
the velocity of a harmonic oscillator plotted against the position spiraling towards the origin on the position-velocity plane
Figure2.1.17The phase plane of an under-damped harmonic oscillator

We can expect a different type of behavior in the case of an over-damped oscillator. For example,

\begin{align*} \frac{d^2 x}{dt^2} + 4 \frac{dx}{dt} + 3 x \amp = 0\\ x(0) \amp = 0\\ x'(0) \amp = 1. \end{align*}

The characteristic equation of the differential equation is

\begin{equation*} \lambda^2 + 4 \lambda + 3 = (\lambda + 3)(\lambda + 1) = 0; \end{equation*}

hence, we have the general solution

\begin{equation*} x(t) = c_1 e^{-t} + c_2 e^{-3t}. \end{equation*}

Applying the initial conditions, our solution is

\begin{align*} x(t) \amp = x(t) = \frac{1}{2}e^{-t} - \frac{1}{2} e^{-3t}\\ v(t) \amp = x'(t) = -\frac{1}{2}e^{-t} + \frac{3}{2} e^{-3t}. \end{align*}

Notice that the damping is too strong for any oscillations to occur (Figure2.1.19 and Figure2.1.20).

velocity and position plotted against time with the two curves approaching zero for large times
Figure2.1.19Position and velocity of an over-damped harmonic oscillator
the velocity of a harmonic oscillator plotted against the position with the solution curve approaching the origin on the position-velocity plane
Figure2.1.20The phase plane of an over-damped harmonic oscillator

As we increase the damping, the oscillations will cease to occur for some value of \(b\text{.}\) This will happen when \(b^2 - 4mk = 0\text{.}\) the At this point we have critical damping. Consider the system

\begin{align*} \frac{d^2 x}{dt^2} + 4 \frac{dx}{dt} + 4 x \amp = 0\\ x(0) \amp = 0\\ x'(0) \amp = 1. \end{align*}

The general solution to this initial value problem is

\begin{equation*} x(t) = c_1 e^{-2t} + c_2 t e^{-2t}. \end{equation*}

The solution to the initial value problem is

\begin{align*} x(t) \amp = x(t) = t e^{-2t}\\ v(t) \amp = x'(t) = (1 - 2t)e^{-2t}. \end{align*}

Although we see that no oscillations for this oscillator (Figure2.1.22 and Figure2.1.23), oscillations will commence as soon as we start to reduce the damping constant \(b = 4\text{.}\)

velocity and position plotted against time with the two curves approaching zero for large times
Figure2.1.22Position and velocity of a critically damped harmonic oscillator
the phase plane of a critically damped harmonic oscillator with the solution curve approaching the origin
Figure2.1.23The phase plane of a critically damped harmonic oscillator

Subsection2.1.5Important Lessons

  • A second-order linear differential equation with constant coefficients is an equation of the form

    \begin{equation*} a x'' + bx' + cx = 0. \end{equation*}
  • The characteristic equation of \(a x'' + bx' + cx = 0\) is

    \begin{equation*} a\lambda^{2} + b\lambda + c = 0. \end{equation*}

    The roots of this equation are

    \begin{equation*} \lambda = \frac{-b \pm \sqrt{b^{2}-4ac}}{2a}. \end{equation*}

    These roots govern the behavior of solutions to the differential equation.

  • If \(b^2 - 4ac \gt 0\text{,}\) the characteristic equation has distinct real roots

    \begin{equation*} \lambda_{1} = \frac{-b + \sqrt{b^{2}-4ac}}{2a} \quad \text{and} \quad \lambda_{2} = \frac{-b - \sqrt{b^{2}-4ac}}{2a} \end{equation*}

    and a general solution to \(a x'' + bx' + cx = 0\) is given by

    \begin{equation*} x(t) = c_{1}e^{\lambda_{1}t} + c_{2}e^{\lambda_{2}t}. \end{equation*}
  • If \(b^2 - 4ac = 0\text{,}\) the characteristic equation has a repeated real root at

    \begin{equation*} \lambda = -\frac{b}{2a} \end{equation*}

    and a general solution to \(a x'' + bx' + cx = 0\) is given by

    \begin{equation*} x(t) = c_{1}e^{\lambda t} + c_{2}t e^{\lambda t}. \end{equation*}
  • If \(b^2 - 4ac \lt 0\text{,}\) characteristic equation has complex conjugate roots

    \begin{equation*} \lambda = \alpha \pm \beta i \end{equation*}

    where \(\alpha = -\frac{b}{2a}\) and \(\beta = \frac{\sqrt{4ac - b^{2}}}{2a}\text{.}\) A general solution to \(a x'' + bx' + cx = 0\) is given by

    \begin{equation*} x(t) = c_{1}e^{\alpha t}\cos(\beta t) + c_{2}e^{\alpha t}\sin(\beta t). \end{equation*}
  • A simple harmonic oscillator can be modeled by the equation

    \begin{equation*} m \frac{d^2 x}{dt^2} + b \frac{dx}{dt} + k x = 0, \end{equation*}

    where \(m \gt 0\text{,}\) \(k \gt 0\text{,}\) and \(b \geq 0\text{.}\) There are three possible types of motion for the oscillator depending on the sign of \(b^2 - 4mk\text{.}\)

    • If \(b^2 - 4mk \lt 0\text{,}\) the oscillator is under-damped.

    • If \(b^2 - 4mk \gt 0\text{,}\) the oscillator is over-damped.

    • If \(b^2 - 4mk = 0\text{,}\) the oscillator is critically damped.

Exercise2.1.24

What is the characteristic equation of \(ax'' + bx' + cx = 0\text{?}\)

Exercise2.1.25

Describe the possible types of damping of a harmonic oscillator.

Subsection2.1.6Exercises

Find the general solution for each equation in Exercise Group2.1.6.110.

\(\dfrac{d^2 y}{dx^2} - y = 0\)

\(x'' - 2x' - 8x = 0\)

\(y'' + 5y = 0\)

\(\dfrac{d^2 x}{dt^2} + 6 \dfrac{dx}{dt} + 5x = 0\)

\(x'' - 10x' + 25x = 0\)

\(\dfrac{d^2 y}{dx^2} - 2\dfrac{dy}{dx} + 4y = 0\)

\(y'' - 8y' + 4y = 0\)

\(\dfrac{d^2 x}{dx^2} + 3\dfrac{dx}{dt} - 10x = 0\)

\(\dfrac{d^2 Q}{dt^2} - 4 \dfrac{dQ}{dt} + 9Q = 0\)

\(\dfrac{d^2 y}{dt^2} + 6 \dfrac{dy}{dt} + 9y = 0\)

Solve the initial value problems in Exercise Group2.1.6.1120.

\(\dfrac{d^2 y}{dx^2} - y = 0\text{,}\) \(y(0) = 1\text{,}\) \(y'(0) = 0\)

\(x'' - 2x' - 8x = 0\text{,}\) \(x(0) = 1\text{,}\) \(x'(0) = 2\)

\(y'' + 5y = 0\text{,}\) \(y(0) = 1\text{,}\) \(y'(0) = 1\)

\(\dfrac{d^2 x}{dt^2} + 6 \dfrac{dx}{dt} + 5x = 0\text{,}\) \(x(0) = 2\text{,}\) \(x'(0) = -1\)

\(x'' - 10x' + 25x = 0\text{,}\) \(x(0) = 1\text{,}\) \(x'(0) = 0\)

\(\dfrac{d^2 y}{dx^2} - 2\dfrac{dy}{dx} + 4y = 0\text{,}\) \(x(0) = -1\text{,}\) \(x'(0) = 1\)

\(y'' - 8y' + 4y = 0\text{,}\) \(y(0) = 1\text{,}\) \(y'(0) = -2\)

\(\dfrac{d^2 x}{dx^2} + 3\dfrac{dx}{dt} - 10x = 0\text{,}\) \(x(0) = 1\text{,}\) \(x'(0) = 2\)

\(\dfrac{d^2 Q}{dt^2} - 4 \dfrac{dQ}{dt} + 9Q = 0\text{,}\) \(Q(0) = -1\text{,}\) \(Q'(0) = 2\)

\(\dfrac{d^2 y}{dt^2} + 6 \dfrac{dy}{dt} + 9y = 0\text{,}\) \(y(0) = 0\text{,}\) \(y'(0) = 0\)

Consider the harmonic oscillators with mass \(m\text{,}\) damping coeeficient \(b\text{,}\) and spring constant \(k\) in Exercise Group2.1.6.2128.

  1. Write the second-order initial value problem corresponding for the harmonic oscillator.

  2. Classify the oscillator as undamped, under-damped, over-damped, or critically damped.

  3. Solve the initial value problem.

  4. Sketch the \(x(t)\) and \(v(t)\)-graphs of the solution of the initial value problem.

  5. Sketch the phase portrait of the initial value problem.

\(m = 1\text{,}\) \(b = 1\) \(k = 1\text{,}\) \(x(0) = 1\text{,}\) \(v(0) = 0\)

\(m = 1\text{,}\) \(b = 2\) \(k = 3\text{,}\) \(x(0) = -3\text{,}\) \(v(0) = 4\)

\(m = 1\text{,}\) \(b = 5\) \(k = 3\text{,}\) \(x(0) = 2\text{,}\) \(v(0) = -3\)

\(m = 1\text{,}\) \(b = 0\) \(k = 25\text{,}\) \(x(0) = 2\text{,}\) \(v(0) = 0\)

\(m = 2\text{,}\) \(b = 3\) \(k = 5\text{,}\) \(x(0) = 2\text{,}\) \(v(0) = -1\)

\(m = 4\text{,}\) \(b = 4\) \(k = 1\text{,}\) \(x(0) = 2\text{,}\) \(v(0) = 1\)

\(m = 3\text{,}\) \(b = 4\) \(k = 1\text{,}\) \(x(0) = 2\text{,}\) \(v(0) = 1\)

\(m = 8\text{,}\) \(b = 4\) \(k = 1\text{,}\) \(x(0) = 2\text{,}\) \(v(0) = 1\)

In Exercise Group2.1.6.2931, we will consider the motion of a mass \(m\) hanging at the end of a vertical spring as in Figure2.1.26. The mass stretches the spring in a downward (positive) direction by length \(L\text{.}\) There are two forces acting on the point where the mass is attached to the springthe force exerted by the spring and gravity. The force of gravity, the weight of the mass, acts downward with a magnitude of \(mg\text{,}\) where \(g\) is the acceleration due to gravity. On the other hand, the force of the spring acts upward and is given by \(-kL\text{,}\) where \(L\) is the length of the spring.1This is commonly known as Hooke's law. See Subsection1.1.3. When the mass is hanging in equilibrium the force of gravity and the force of the spring balance each other out; that is,

\begin{equation*} mg - kL = 0. \end{equation*}
three springs hanging from a beam with masses attacheds the right two springs and the rightmost spring extended
Figure2.1.26A spring-mass system

We would like to investigate the motion of the mass if is initially displaced or acted on by an external force. Let \(x(t)\) be the displacement of the mass from its equilibrium position, where a downward displacement is positive. The force acting on the mass are the weight of the mass \(mg\) and the force exerted by the spring, which is the total elongation of the spring, or

\begin{equation*} -k(L + x). \end{equation*}

By Newton's second law of motion,

\begin{equation*} mx'' = mg - k(L + x). \end{equation*}

Furthermore, it is possible to add a damping term, \(bx'\text{,}\) or even an external force \(F(t)\) to obtain

\begin{equation*} mx'' = mg - k(L + x) + bx' + F(t). \end{equation*}

Since \(mg - kL = 0\text{,}\) we obtain the familiar equation

\begin{equation*} mx'' + bx' + kx = F(t). \end{equation*}

Suppose that a mass of \(100\) grams stretches a spring \(2\) centimeters.

  1. Determine the spring constant \(k\text{.}\)

  2. If the mass is displaced an additional \(4\) centimeters and released, write an initial value problem that will model the motion of the oscillating mass.

  3. Solve the initial value problem.

Suppose that a mass of \(1\) kilogram stretches a spring \(5\) centimeters.

  1. Determine the spring constant \(k\text{.}\)

  2. If the mass is displaced an additional \(5\) centimeters and released, write an initial value problem that will model the motion of the oscillating mass.

  3. Suppose the the spring-mass system is suspended in a fluid that exerts a resistance of \(0.25\) kilograms when the mass has a velocity of \(2\) centimeters per second. Modify the intial-value problem that you wrote in (b) to take this fact into account.

  4. Solve the initial value problem.

Suppose that a mass weighing 4 lbs stretches a spring 3 inches.

  1. If \(g = 32 \text{ft/sec}^2\text{,}\) determine \(m\text{.}\)

  2. Determine the spring constant \(k\text{.}\)

  3. If the mass is displaced an additional 6 inches and released, write an initial value problem that will model the motion of the oscillating mass.

  4. Solve the initial value problem.

Hint

Pay careful attention to units.

Let \(a x'' + b x' + cx = 0\text{,}\) where \(a \neq 0\) and \(b^2 - 4ac = 0\text{.}\)

  1. Show that \(x_1(t) = e^{-bt/2a}\) is a solution to \(a x'' + b x' + cx = 0\text{.}\)

  2. Assume that

    \begin{equation*} y = v(t) x_1(t) = v(t) e^{-bt/2a} \end{equation*}

    is a solution to \(a x'' + b x' + cx = 0\) and show that \(v(t) = c_1 + c_2 t\text{.}\) Thus,

    \begin{equation*} x(t) = c_1 e^{-bt/2a} + c_2 t e^{-bt/2a} \end{equation*}

    is a general solution for \(a x'' + b x' + cx = 0\text{.}\)

Hint
  1. Observe that

    \begin{align*} a x_1'' + b_1' + cx_1 & = a \left(\frac{-b}{2a}\right)^2e^{-bt/2a} + b \left( \frac{-b}{2a} \right) e^{-bt/2a} + c e^{-bt/2a}\\ & = e^{-bt/2a} \left( \frac{b^2}{4a} - \frac{b^2}{2a} + c \right)\\ & = e^{-bt/2a} \left( \frac{-b^2 + 4ac}{4a} \right)\\ & = 0. \end{align*}
  2. If \(y = v(t) x_1(t) = v(t) e^{-bt/2a}\) is a solution to our differential equation, then

    \begin{align*} a y'' + b y' + cy & = a (v''x_1 + 2 v'x_1' + vx_1'' ) + b(v' x_1 + v x_1') + cv x_1\\ & = a v''x_1 + 2a v'x_1' + bv' x_1 + v(a x_1'' +b x_1' + c x_1)\\ & = a v'' e^{-bt/2a} + \left[2a \left( \frac{-b}{2a} \right) e^{-bt/2a} + b e^{-bt/2a} \right] v'\\ & = a v'' e^{-bt/2a}\\ & = 0. \end{align*}

    Since \(a \neq 0\text{,}\) we know that \(v'' = 0\text{.}\) Hence, \(v(t) = c_1 + c_2 t\text{.}\)

Reduction of Order. Suppose that \(x_1(t)\) is a solution (not identically zero) to the equation

\begin{equation*} x'' + p(t) x' + q(t) x = 0. \end{equation*}
  1. Assume that \(x(t) = v(t) x_1(t)\) is a solution to \(x'' + p(t) x' + q(t) x = 0\) and derive the equation

    \begin{equation} x_1 v'' +(2x_1' + px_1)v' = 0.\label{secondorder01-equation-exercise-reduction-of-order-1}\tag{2.1.11} \end{equation}
  2. Let \(u = v'\) and show that (2.1.11) is a first-order linear differential equation in \(u\text{.}\)

  3. Show that \(x_1(t) = 1/t\) is a solution to

    \begin{equation} 2 t^2 x'' + 3t x' - x = 0 \label{secondorder01-equation-exercise-reduction-of-order-2}\tag{2.1.12} \end{equation}

    for \(t \gt 0\) and find a second linearly independent solution to (2.1.12).

Hint
  1. \begin{align*} x'' + px' + qx & = (v''x_1 + 2 v' x_1' + vx_1'') + p(v'x_1 + vx_1') + q(vx_1)\\ & = x_1 v'' + 2v' x_1' + p x_1 v' + v (x_1'' + p x_1' + q x_1)\\ & = x_1 v'' +(2x_1' + px_1)v' \\ & = 0. \end{align*}
  2. If \(u = v'\text{,}\) then \(x_1 u' +(2x_1' + px_1)u= 0\text{.}\)

  3. If \(x_1(t) = 1/t\text{,}\) then

    \begin{equation*} 2 t^2 x_1'' + 3t x_1' - x_1 = 2 t^2 \left(\frac{2}{t^3}\right) + 3t \left(\frac{-1}{t^2}\right) - \frac{1}{t} = 0. \end{equation*}

    If we assume that \(x = v/t\) is a second solution, then

    \begin{equation*} 2 t^2 x'' + 3t x' - x = 2tv'' - v' = 0. \end{equation*}

    If we let \(u = v'\text{,}\) then a solution of \(2tu' - u = 0\) is \(u = \sqrt{t}\) and \(v = \int \sqrt{t} \, dt = 2 t^{3/2} / 3\text{.}\) Therefore, the second solution to our equation is

    \begin{equation*} x = \frac{v}{t} = \frac{2}{3} \sqrt{t}. \end{equation*}

Let \(a x'' + b x' + cx = 0\text{,}\) where \(a \neq 0\) and \(b^2 - 4ac = 0\text{.}\)

  1. Show that \(x_1(t) = e^{-bt/2a}\) is a solution to \(a x'' + b x' + cx = 0\text{.}\)

  2. Assume that

    \begin{equation*} y = v(t) x_1(t) = v(t) e^{-bt/2a} \end{equation*}

    is a solution to \(a x'' + b x' + cx = 0\) and show that \(v(t) = c_1 + c_2 t\text{.}\) Thus,

    \begin{equation*} x(t) = c_1 e^{-bt/2a} + c_2 t e^{-bt/2a} \end{equation*}

    is a general solution for \(a x'' + b x' + cx = 0\text{.}\)

Consider the equation

\begin{equation*} y'' - (2 \alpha -1) y' + \alpha (\alpha - 1) y =0. \end{equation*}

Determine all values of \(\alpha\text{,}\) if any, for which all solutions tend toward zero as \(t \to \infty\text{.}\) Also, determine the values of \(\alpha\text{,}\) if any, for which all nonzero solutions become unbounded as \(t \to \infty\text{.}\)

Solve each of the following initial value problems.

  1. \begin{align*} y'' - 2y' + 5y & = 0\\ y(\pi/2) & = 0\\ y'(\pi/2) & = 2 \end{align*}
  2. \begin{align*} 9y'' - 12y' + 4y & = 0\\ y(0) & = 2\\ y'(0) & = -1 \end{align*}
  3. \begin{align*} y'' + 8y' -9y & = 0\\ y(1) & = 1\\ y'(1) & = 0 \end{align*}
  4. \begin{align*} y'' + 2ay' + (a^2 + 1)y & = 0\\ y(0) & = 1\\ y'(0) & = 0 \end{align*}

Reduction of Order. Suppose that \(x_1(t)\) is a solution (not identically zero) to the equation

\begin{equation*} x'' + p(t) x' + q(t) x = 0. \end{equation*}
  1. Assume that \(x(t) = v(t) x_1(t)\) is a solution to \(x'' + p(t) x' + q(t) x = 0\) and derive the equation

    \begin{equation} x_1 v'' +(2x_1' + px_1)v' = 0.\label{secondorder01-equation-exercise-reduction-of-order-1a}\tag{2.1.13} \end{equation}
  2. Let \(u = v'\) and show that ((2.1.13)) is a first-order linear differential equation in \(u\text{.}\)

  3. Show that \(x_1(t) = 1/t\) is a solution to

    \begin{equation} 2 t^2 x'' + 3t x' - x = 0\label{secondorder01-equation-exercise-reduction-of-order-2a}\tag{2.1.14} \end{equation}

    for \(t \gt 0\) and find a second linearly independent solution to ((2.1.14)).

Euler Equations. An important class of second-order linear differential equations are equations of the form

\begin{equation*} t^2 y'' + \alpha t y' + \beta y = 0, \end{equation*}

where \(t \gt 0\) and \(\alpha\) and \(\beta\) are real constants. An equation of this form is called an Euler equation.

  1. Show that the substitution \(x = \ln t\) transforms an Euler equation into an equation of constant coefficients. {\em Hint}: Show that

    \begin{align*} \frac{dy}{dt} & = \frac{dx}{dt} \frac{dy}{dx}\\ \frac{d^2y}{dt^2} & = \left( \frac{dx}{dt} \right)^2 \frac{d^2y}{dx^2} + \frac{d^2x}{dt^2} \frac{dy}{dx}. \end{align*}
  2. Solve the equation

    \begin{equation*} t^2 y'' + 4ty' + 2y = 0. \end{equation*}