14
Ordinary differential equations

Any equation that expresses the functional dependence of a variable y on its arguments xi(i = 1, 2, …) and the derivatives of y with respect to those arguments, is called a differential equation. Physical systems are almost always described by such equations. For example, a wave f(x, t) travelling with velocity in the x-direction obeys the wave equation

(14.1a) Unnumbered Display Equation

while the motion of a simple pendulum of length l performing small oscillations θ satisfies

(14.1b) Unnumbered Display Equation

where g is a constant. Equation (14.1a) contains partial derivatives because f is a function of more than one variable, and the equation is therefore called a partial differential equation (PDE). These will be discussed in Chapter 16. On the other hand, in Equation (14.1b) the quantity θ depends on the single variable t, and the equation is called an ordinary differential equation (ODE). It is these equations that are the subject of this chapter and the next.

In what follows, we will usually refer to the independent variable as x and the corresponding dependent variable as y(x). Thus, in general, an ordinary differential equation is of the form

(14.2) Unnumbered Display Equation

Examples of ODEs are:

(14.3a) Unnumbered Display Equation

(14.3b) Unnumbered Display Equation

and

(14.3c) Unnumbered Display Equation

It is convenient to classify ordinary differential equations by their order, degree and linearity. The order is defined as the order of the highest derivative in the equation. Thus, Equations (14.3a), (14.3b) and (14.3c) are first, second and third order, respectively. The degree of the equation is defined as the power to which the highest order derivative is raised after the equation is rationalised, that is, only integer powers remain. Thus, (14.3a) and (14.3b) are both first degree equations. Equation (14.3c) is of second degree because when the equation is rationalised, it becomes

numbered Display Equation

so that the highest derivative is d3y/dx3 and it is raised to power 2. Finally, any ODE of order n is said to be linear if it is linear in the dependent variable y and its first n derivatives. If this is not true, the equation is said to be non-linear. Equations (14.3a) and (14.3b) are therefore linear, whereas (14.1b) and (14.3c) are non-linear. The discussion in this chapter will be predominantly about linear ordinary differential equations. This is because, except in a few simple cases, non-linear equations are difficult to solve and one must usually resort to numerical methods. Nonetheless, they are important, especially in describing complex systems, such as the atmosphere, and they play a central role in the so-called chaotic systems.1

The solution of an ODE in the variables x and y is defined as a relation between the two variables, such that when substituted into the ODE gives an identity. It may be of the explicit form y = g(x), or the implicit form φ(x, y) = 0, and is obtained, in principle, by repeated integration. Since each such operation will introduce a constant of integration, we can deduce that a solution of an nth order ordinary differential equation cannot be a general solution unless it contains n arbitrary constants. In physical situations, these constants are fixed by specifying boundary conditions, that is, by requiring that y and/or its derivative has specific values at given points. For example, the linear second-order equation

numbered Display Equation

has the general solution

numbered Display Equation

which may be verified by direct substitution. A and B are two arbitrary constants, as expected for a second-order equation. To determine these, we could specify the values of y at two values of x. Thus, if we were to require that y = 0 when x = 0 and y = 1 when x = 1, then

numbered Display Equation

which gives

numbered Display Equation

and hence the specific solution is

numbered Display Equation

Different boundary conditions lead of course to different solutions.

In the following sections we will discuss a number of methods to solve various types of ordinary differential equations, starting with first-order equations of the first degree. First-order equations of higher than first degree rarely occur in physical science and will not be discussed here.

14.1 First-order equations

The general form for equations of this type is

(14.4) Unnumbered Display Equation

Solutions of (14.4) are found relatively easily if F(x, y) has specific simple forms and we will discuss some of these below.

14.1.1 Direct integration

If F(x, y) = f(x) is a function of x only, then (14.4) takes the form

(14.5a) Unnumbered Display Equation

and may be directly integrated to give

numbered Display Equation

where c is an arbitrary constant. If F(x, y) = f(y) is a function of y only, then we have

(14.5b) Unnumbered Display Equation

so that the solution

numbered Display Equation

is again obtained by direct integration.

14.1.2 Separation of variables

One frequently meets applications in which the right-hand side of (14.4) is a product F(x, y) = f(x)g(y), so that it becomes

(14.6a) Unnumbered Display Equation

If either f or g is a constant, then (14.6a) can be evaluated by direct integration, as in Section 14.1.1. Otherwise, rearranging (14.6a) gives

numbered Display Equation

that is, the variables have been separated. Integrating then gives

(14.6b) Unnumbered Display Equation

which expresses y implicitly in terms of x. In doing the integration one must of course remember to include the constant of integration.

14.1.3 Homogeneous equations

In Section 7.2.4, a function f(x1, x2, …, xn) was defined to be homogeneous of degree k if

numbered Display Equation

where λ is an arbitrary parameter. Thus, for example, the function f(x, y) = x3y + x2y2 is a homogeneous function of degree 4, whereas f(x, y) = xy + x/y is inhomogeneous because it does not satisfy this requirement. Homogeneous equations are of the form

(14.7) Unnumbered Display Equation

where g and h are both homogeneous functions of the same degree. The key property is that the right-hand side of (14.7) can be written as a function of the ratio zy/x, i.e.

numbered Display Equation

and

numbered Display Equation

Substituting these equations into (14.7) gives

(14.8a) Unnumbered Display Equation

This is a separable equation, which can be integrated to give

(14.8b) Unnumbered Display Equation

from which z and hence y can be found.

Some equations, although not obviously homogeneous at first sight, may be reduced to this form by suitable transformation of variables. One type that commonly occurs is given in Example 14.3(b).

14.1.4 Exact equations

If

numbered Display Equation

then the first-order equation (14.4) may be written

(14.9) Unnumbered Display Equation

If it is possible to find a function f(x, y) such that

(14.10) Unnumbered Display Equation

then, by Equation (7.17a),

numbered Display Equation

is an exact differential and (14.9) is called an exact equation. We then have

numbered Display Equation

and comparing this with (14.9), we see that the latter has the implicit solution

(14.11) Unnumbered Display Equation

where c is an integration constant. To see whether a function f(x, y) that satisfies (14.10) exists, we note that (14.10) implies

numbered Display Equation

and hence

(14.12) Unnumbered Display Equation

so that (14.12) is a necessary condition for a relation of the form (14.10) to exist. It can also be shown to be a sufficient condition. If it is satisfied, then integrating A with respect to x at fixed y, and B with respect to y at fixed x gives the results

(14.13a) Unnumbered Display Equation

and

(14.13b) Unnumbered Display Equation

where f1(y) and f2(x) are arbitrary functions, which may be identified up to a constant by comparing (14.13a) and (14.13b). Alternatively, (14.13a) may be differentiated with respect to y at fixed x and compared to ∂f/∂y = B to determine c1(y). The solution is then given by (14.11), as we shall illustrate by an example.

14.1.5 First-order linear equations

First-order linear ODEs are equations of the form

(14.14) Unnumbered Display Equation

where p(x) and q(x) are given functions of x, or constants. If the equation is exact, it may be solved by the method of Section 14.1.4. If it is not of this form, that is, it is inexact, it may in principle be solved by multiplying by a function I(x), to be determined below, called an integrating factor. We then obtain

(14.15a) Unnumbered Display Equation

and I(x) is chosen so that the left-hand side of (14.15a) is equal to d[I(x)y]/dx, i.e.,

(14.15b) Unnumbered Display Equation

Equating the linear terms in y in this equation gives

numbered Display Equation

which on integrating gives (providing y ≠ 0) the result

(14.16) Unnumbered Display Equation

for the integrating factor. Finally, from (14.15a) and (14.15b),

numbered Display Equation

and hence the general solution for y is given by

(14.17) Unnumbered Display Equation

with I(x) given by (14.16).

14.2 Linear ODEs with constant coefficients

Linear ODEs are of the form

(14.18) Unnumbered Display Equation

where the ai(x) (i = 0, 1, 2, …, n) and f(x) are given functions of x. For n = 1, this reduces to (14.14), and the method of solution has been discussed in Section 14.1.5. In this section we will consider the case of arbitrary n where the ai(x) are constants, that is, ai(x) = ai so that

(14.19) Unnumbered Display Equation

Other types of linear ODE will be discussed in Section 14.2.4 and Chapter 15.

The solution of equations of the type (14.19) is found in three steps. Firstly, one finds the general solution y0 to the reduced equation obtained by setting f(x) = 0 in (14.19), i.e.

(14.20a) Unnumbered Display Equation

The function y0 contains n free parameters c1, c2, ⋅⋅⋅, cn, since it is the general solution to an equation of order n. It is called the complementary function. The second step is to find a particular solution Y(x) of (14.19), so that

(14.20b) Unnumbered Display Equation

The function Y(x) is called the particular integral. Finally, one adds the complementary function to the particular integral to obtain

(14.21) Unnumbered Display Equation

On substituting (14.21) into (14.19) and using (14.20a) and (14.20b), one easily shows that it is a solution; and since it contains n arbitrary parameters, it is the general solution.

It is relatively easy to find complementary functions for a specific equation, as we shall show in Section 14.2.1, but there is no general method for finding particular integrals for a given f(x). In Sections 14.2.2 and 14.2.3 we shall discuss two methods that work in a wide variety of cases.

14.2.1 Complementary functions

As defined above, complementary functions are the general solutions of reduced equations of the form

(14.22) Unnumbered Display Equation

Equations like (14.22) are often called homogeneous.2 These equations have the important property that if y1(x) and y2(x) are solutions, then any linear combination

(14.23) Unnumbered Display Equation

is also a solution. This result follows directly on substituting (14.23) into the left-hand side of (14.22). In what follows, we shall discuss second-order equations in some detail, because they are by far the most important in physics applications; then we briefly address the extension to higher-orders.

The second-order homogeneous equation is

(14.24) Unnumbered Display Equation

where we have relabelled the constants for later convenience. As a trial solution we will take

(14.25) Unnumbered Display Equation

since with this form, differentiating y just multiplies it by a constant m. Substituting (14.25) in (14.24) gives

numbered Display Equation

and (14.25) is a solution of (14.24) when the bracket vanishes, that is, when m is a root of the auxiliary equation

(14.26) Unnumbered Display Equation

This has roots

(14.27) Unnumbered Display Equation

and three cases must be distinguished.

  1. If b2 > 4ac, there are two real roots m1 and m2, with m1m2. The general solution of (14.24) is then, by the superposition principle (14.23),

    (14.28a) Unnumbered Display Equation

    where A1 and A2 are arbitrary constants.
  2. If b2 = 4ac, then

    (14.29) Unnumbered Display Equation

    so that (14.28a) would contain only one arbitrary constant AA1 + A2 and so cannot be the general solution. In this case, a second solution is obtained by writing

    numbered Display Equation

    where u(x) is a function to be determined. Substituting into (14.24) then gives

    (14.30) Unnumbered Display Equation

    However, using (14.29), we see that

    numbered Display Equation

    so that the second term in (14.30) vanishes and we are left with the equation d2u/dx2 = 0, with solution

    numbered Display Equation

    where again A1 and A2 are arbitrary constants. The general solution of (14.24) when the auxiliary equation has two equal roots is therefore

    (14.28b) Unnumbered Display Equation

  3. Finally, if b2 < 4ac, there are two complex solutions

    (14.31) Unnumbered Display Equation

    where α = −b/2a and . Nonetheless, a real solution of (14.24) is obtained by writing

    numbered Display Equation

    where A is a complex constant. Using (14.31) and ez = cos z + isin z for any z, this can be rewritten in the form

    (14.28c) Unnumbered Display Equation

    where and are two arbitrary real constants.

    This exhausts the types of solution for homogeneous second-order linear equations with constant coefficients.

    In the cases of higher order equations (14.20a), the substitution y = emx leads to the auxiliary equation

    (14.32) Unnumbered Display Equation

    This gives n roots,3 where degenerate and complex roots are treated in the same way as in the second-order case. In particular, if k roots m1, m2, ⋅⋅⋅, mk = m coincide, then the corresponding term in the solution, analogous to (14.28b), becomes

    (14.33) Unnumbered Display Equation

14.2.2 Particular integrals: method of undetermined coefficients

We now turn to the case where f(x) ≠ 0 in (14.19), again restricting the discussion to the case of constant coefficients. We have already stated that the general solution of such an equation is the sum of the complementary function and a particular integral. The standard method for finding the complementary function has already been discussed, so it only remains to find the particular integral. Unfortunately, there is no general method for doing this, but there is a variety of methods, each of which is appropriate for a range of functions f(x). We will discuss two, starting with that known as the method of undetermined coefficients.

This method consists of assuming a trial form for the particular integral Y(x) that resembles f(x), but contains a number of free parameters. The trial function is then substituted into the differential equation and the parameters determined so that the equation is satisfied. It is most useful when f(x) contains only polynomials, exponentials, or sines and cosines. The rules for constructing the appropriate trial functions are as follows.

  1. If f(x) = aebx where a and b are constants, the trial function is Y(x) = Aebx, where A is a constant.
  2. If f(x) = asin px + bcos px, where a, b and p are constants (a or b may be zero), the trial function is Y(x) = Asin px + Bcos px, where A and B are constants.
  3. If , where some of the coefficients may be zero, the trial function is , where the bn are constants.
  4. If any term in these trial functions is contained within the complementary function, then the trial function must be multiplied by the smallest integer power of x such that it then contains no term that is in the complementary function.
  5. Finally, if f(x) is the sum or product of any of these forms, the trial function must be taken to be the sum or product of the appropriate individual trial functions.

*14.2.3 Particular integrals: the D-operator method

There are other methods for finding particular integrals, and by way of contrast we will discuss here the so-called D-operator method. This method has the advantage that it is not necessary to guess a trial function and the numerical coefficients multiplying the functional form of the particular integral are obtained automatically. However, it does require some experience in identifying which manipulations to use to obtain the solution.

The quantity D, defined by D ≡ d/dx, was introduced in Chapter 9, Section 9.3.2. It is a differential operator. That is, it only has a meaning when it acts on a function, which is always written to the right of D. Higher differential operators may be constructed from D. For example,

numbered Display Equation

numbered Display Equation

etc. We see that D satisfies the usual rules of algebra and so may be formally treated as an algebraic quantity, despite the fact that it cannot be evaluated as such to yield a numerical value. From the algebraic rules of differentiation it follows that if f(x) and g(x) are differentiable functions,

(14.36a) Unnumbered Display Equation

and

(14.36b) Unnumbered Display Equation

where c is a constant, and hence D is a linear operator (see Section 9.3.2).

In terms of D we may now write (14.19) as

(14.37a) Unnumbered Display Equation

where

(14.37b) Unnumbered Display Equation

is a polynomial operator in D of order n. Just as D may be formally treated as an algebraic quantity, so may a polynomial function of D, such as F(D), and in particular it may, in suitable cases, be factorised. Moreover the order of the factors is irrelevant provided the coefficients in (14.37) are constants. Thus, for example,

numbered Display Equation

However, the order is relevant if the coefficients are functions of x. For example,

numbered Display Equation

whereas

numbered Display Equation

so

numbered Display Equation

A number of useful results may be derived when the function f(x) = ekx, where k is a constant. For example, using the fact that Dnekx = knekx, it follows that

(14.38a) Unnumbered Display Equation

Similarly, since D2sin kx = −k2sin kx, it follows that

(14.38b) Unnumbered Display Equation

and

(14.38c) Unnumbered Display Equation

If the exponential is multiplied by an arbitrary function V(x), then by considering the terms D[ekxV(x)], D2[ekxV(x)], etc. in succession, the result (14.38a) may be generalised in a straightforward way to

(14.38d) Unnumbered Display Equation

Finally, in Chapter 3 we defined indefinite integration as the inverse operation of differentiation. In an analogous way we now defined the inverse operator D− 1 by

(14.39) Unnumbered Display Equation

with

(14.40) Unnumbered Display Equation

Also by analogy with the notation Dn for successive differentiations, we will use Dn ≡ 1/Dn for the operation of n successive integrations.

We now return to the solution of equations of the type (14.19). If these are rewritten in the form (14.37a) and (14.37b), then a particular integral can be obtained by writing

(14.41) Unnumbered Display Equation

provided we can interpret and evaluate the right-hand side, using the techniques introduced above. To illustrate this, consider the equation

(14.42) Unnumbered Display Equation

In the D-operator formalism, this is written

numbered Display Equation

giving the particular integral

numbered Display Equation

We now have to decide how to handle the polynomial in the denominators, using the relations (14.38). If we use (14.38c) with k = 1, we have

numbered Display Equation

Now we can again use (14.38c) to give

numbered Display Equation

as the desired form for the particular integral.

A different technique is illustrated by considering the equation obtained by replacing cos x in (14.42) by xe2x. The particular integral is then

numbered Display Equation

and using (14.38d) with k = 2 gives

numbered Display Equation

The denominator can be expressed as partial fractions, to give

(14.43) Unnumbered Display Equation

The point of this decomposition is that each fraction can be expanded as a binomial series

numbered Display Equation

where higher derivatives are not required because when acting on x they will be zero. Using these expansions, (14.43) becomes

numbered Display Equation

which is the desired result.

*14.2.4 Laplace Transforms

The D-operator method converts a differential equation to an algebraic equation for D. Another method that does something very similar is based on the use of Laplace transforms. The Laplace transform F(p) of a function f(x) is defined by

(14.44) Unnumbered Display Equation

where the parameter p may in principle be complex, but in the discussion that follows we will take it to be real.

The Laplace transforms of many simple functions may be found by direct evaluation of the defining integral (14.44). Others may then be found by using easily proved general properties of the Laplace transform. These include: linearity,

numbered Display Equation

where a and b are constants; and the shift theorem

If L[f(x)] = F(p), then L[eaxf(x)] = F(pa).

A related version of this, called the translation property, is

numbered Display Equation

where H(x) is the unit step function

numbered Display Equation

For example, the translation property follows by considering the Laplace transform of

numbered Display Equation

which has the same form as f(x) but moved a distance a along the x axis. If we let z = xa, then

numbered Display Equation

Table 14.1 Laplace transforms of some simple functions4

Using a combination of these properties, several useful examples of Laplace transforms may be obtained and are shown in Table 14.1. As an example, we shall derive the result for eaxcos (bx). Firstly,

numbered Display Equation

Then, using the shift theorem,

numbered Display Equation

We can also define an inverse Laplace transform, denoted L− 1, such that

numbered Display Equation

It follows that LL− 1 = 1 and because L is linear, so is L− 1. Inverse transforms for some simple functions follow from the results of Table 14.1. Thus, since L[1] = 1/p, it follows that L− 1[1/p] = 1, but to find inverse transforms in general requires the techniques of complex variable theory and we will not discuss them here.

In order to use Laplace transforms to solve differential equations, we will also need the transforms of the differentials of a function y(x). For example, consider the Laplace transform of dy(x)/dx. From the definition (14.44), this is

numbered Display Equation

Integrating by parts, this is

numbered Display Equation

providing yepx → 0 as x → ∞. So,

(14.45a) Unnumbered Display Equation

Likewise,

numbered Display Equation

so that

numbered Display Equation

where

numbered Display Equation

This procedure may be repeated to obtain expressions for derivatives of any order. These can be used to convert a differential equation into an algebraic equation in F(p), and if the inverse transform L− 1 can be found, the solution of the equation for F(p) can be inverted, and a solution of the original equation ODE results. An advantage of the method is that it can easily incorporate boundary conditions on the function and its first derivative, as can be seen from (14.45). Its use is illustrated in Example 14.10.

Fourier transforms can similarly be used to convert a differential equation for y(x) into an algebraic equation for its Fourier transform. However, in contrast to Laplace transforms, their use is restricted to functions that tend to zero as |x| tends to infinity sufficiently rapidly for (13.45) to be satisfied. In these cases, it can be a useful technique, as illustrated in Problem 14.19, but we will not pursue this here.

It frequently happens that part of the expression for which the inverse transform L− 1 is to be found contains the product of two Laplace transforms. In this the case we can use the method of convolutions that was discussed in Chapter 13, Section 13.3.4 for Fourier transforms. Thus if gi(p) is the Laplace transform of fi(p), that is, if

(14.46) Unnumbered Display Equation

where i = 1,2, then

(14.47a) Unnumbered Display Equation

and hence

(14.47b) Unnumbered Display Equation

where the convolution integral is

numbered Display Equation

To prove (14.47a), we start from the definition (14.46) and form the product

(14.47c) Unnumbered Display Equation

where u and are dummy variables. Now letting , and rewriting in terms of the variables t and u, changes the limits on the integrals, giving

numbered Display Equation

This corresponds to summing the vertical strips on Figure 14.2a. But from the work of Chapter 11, we know that we can equally sum the horizontal strips, as shown in Figure 14.2b, that is, we can reverse the order of integrations. Therefore the double integral may be written

numbered Display Equation

and so (14.47a) follows, which completes the proof.

images

Figure 14.2 Order of integration in (14.47c).

*14.3 Euler's equation

The discussion in Section 14.2 has been exclusively about linear equations with constant coefficients. However, some linear equations with variable coefficients ai(x) can be reduced to linear form with constant coefficients by a suitable transformation. Perhaps the best known of these is Euler's equation

(14.48) Unnumbered Display Equation

On substituting x = et, one obtains

numbered Display Equation

and

numbered Display Equation

so that if f(x) becomes on changing variables, (14.48) becomes

(14.49) Unnumbered Display Equation

This is a linear equation with constant coefficients and so may be solved by the methods of the last section. More generally, the nth order Euler equation

(14.50) Unnumbered Display Equation

is reduced to a linear equation with constant coefficients by the same substitution x = et.

If f(x) = 0, then (14.50) with f(x) = 0 is

numbered Display Equation

where A1, A2, …, An are arbitrary constants. On the other hand, if the roots are not distinct, this simple method fails to give the general solution, which can still be found using the substitution x = et.

Problems 14

  1.   14.1 Find the solution of the equations

    numbered Display Equation

    numbered Display Equation
  2.   14.2 Find the solutions of the equations

    numbered Display Equation

    numbered Display Equation
  3.   14.3 Show that the Fourier transform of exp ( − αx2) with α > 0 is

    by using the symmetry of exp ( − αx2) to derive the equation

    numbered Display Equation

    and then solving it subject the boundary condition

    numbered Display Equation

    which follows from the standard integral derived in Example 11.11. Finally, use ((14.51) to verify (13.71).

  4.   14.4 Find the solution of the equation

    numbered Display Equation
  5.   14.5 Find the solution of the equation

    numbered Display Equation
  6.   14.6 Establish whether each of the following equations is exact and find the solutions of those that are:

    numbered Display Equation
  7.   14.7 Find the solutions of the equations

    numbered Display Equation
  8.   14.8 Find the solutions of the equations

    numbered Display Equation
  9.   14.9 Find the solutions of the equations

    numbered Display Equation

    numbered Display Equation
  10.  14.10 Solve the non-linear equation

    numbered Display Equation

    using the substitution z = y− 2.

  11.  14.11 Find the solution of the equations

    numbered Display Equation

    numbered Display Equation
  12.  14.12 Use the method of undetermined coefficients to find the complete solutions of the equations

    numbered Display Equation
  13.  14.13 Use the method of undetermined coefficients to find the complete solution of the equation

    numbered Display Equation
  14.  14.14 Find the complete solutions of the equations

    numbered Display Equation
  15.  14.15 Find the general solution of the equations

    numbered Display Equation
  16. *14.16 Use the D-operator method to find the particular integrals for the equations

    numbered Display Equation
  17. *14.17 Use the D-operator method to find the particular integral for the equation

    numbered Display Equation
  18.  14.18 Find the complete solution of the equation

    numbered Display Equation
  19.  14.19 A function φ satisfies the equation

    numbered Display Equation

    where C is a constant andf(x) is an arbitrary function such that

    Use the relation (13.53b) to show that

    numbered Display Equation

    is a solution, where h(k) is the Fourier transform of f(x). What is the general solution?

  20. *14.20 Use the Laplace transform method to find the solution of the equation

    numbered Display Equation

    subject to the boundary conditions y(x) = −2 and y′(x) = 4 at x = 0.

  21. *14.21 A one-dimensional system undergoes forced simple harmonic motion, with an equation of motion

    numbered Display Equation

    where ω0 is the natural frequency of vibration and A is a constant. Solve for x(t) with the boundary conditions x(0) = 1 and x′(0) = 0, using the Laplace transform method.

  22.  14.22 Solve Problem 14.21 using the method of undetermined coefficients.

  23. *14.23 Find the solution of the equation

    numbered Display Equation

    subject to the boundary conditions y(0) = 1 and y′(0) = 0, where h(x) is an unknown function of x.

  24. *14.24 Solve the equation

    numbered Display Equation
  25. *14.25 Find the complete solution of the equation

    numbered Display Equation
  26. *14.26 If x = et, show that

    numbered Display Equation

    Hence show that the Euler equation (14.50) is reduced to a linear equation with constant coefficients for any order n by the substitution x = et, as asserted in the text.

Notes

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset