Separation of variables

Last updated

In mathematics, separation of variables (also known as the Fourier method) is any of several methods for solving ordinary and partial differential equations, in which algebra allows one to rewrite an equation so that each of two variables occurs on a different side of the equation.

Contents

Solve proportional first order differential equation Sept proportional differential equation.png
Solve proportional first order differential equation
Solve linear first order differential equation
by separation of variables. Sept linear differential equation.png
Solve linear first order differential equation by separation of variables.

Ordinary differential equations (ODE)

A differential equation for the unknown will be separable if it can be written in the form

where and are given functions. This is perhaps more transparent when written using as:

So now as long as h(y) ≠ 0, we can rearrange terms to obtain:

where the two variables x and y have been separated. Note dx (and dy) can be viewed, at a simple level, as just a convenient notation, which provides a handy mnemonic aid for assisting with manipulations. A formal definition of dx as a differential (infinitesimal) is somewhat advanced.

Alternative notation

Those who dislike Leibniz's notation may prefer to write this as

but that fails to make it quite as obvious why this is called "separation of variables". Integrating both sides of the equation with respect to , we have

 

 

 

 

(A1)

or equivalently,

because of the substitution rule for integrals.

If one can evaluate the two integrals, one can find a solution to the differential equation. Observe that this process effectively allows us to treat the derivative as a fraction which can be separated. This allows us to solve separable differential equations more conveniently, as demonstrated in the example below.

(Note that we do not need to use two constants of integration, in equation ( A1 ) as in

because a single constant is equivalent.)

Example

Population growth is often modeled by the "logistic" differential equation

where is the population with respect to time , is the rate of growth, and is the carrying capacity of the environment. Separation of variables now leads to

which is readily integrated using partial fractions on the left side yielding

where A is the constant of integration. We can find in terms of at t=0. Noting we get

Generalization of separable ODEs to the nth order

Much like one can speak of a separable first-order ODE, one can speak of a separable second-order, third-order or nth-order ODE. Consider the separable first-order ODE:

The derivative can alternatively be written the following way to underscore that it is an operator working on the unknown function, y:

Thus, when one separates variables for first-order equations, one in fact moves the dx denominator of the operator to the side with the x variable, and the d(y) is left on the side with the y variable. The second-derivative operator, by analogy, breaks down as follows:

The third-, fourth- and nth-derivative operators break down in the same way. Thus, much like a first-order separable ODE is reducible to the form

a separable second-order ODE is reducible to the form

and an nth-order separable ODE is reducible to

Example

Consider the simple nonlinear second-order differential equation:

This equation is an equation only of y'' and y', meaning it is reducible to the general form described above and is, therefore, separable. Since it is a second-order separable equation, collect all x variables on one side and all y' variables on the other to get:

Now, integrate the right side with respect to x and the left with respect to y':

This gives

which simplifies to:

This is now a simple integral problem that gives the final answer:

Partial differential equations

The method of separation of variables is also used to solve a wide range of linear partial differential equations with boundary and initial conditions, such as the heat equation, wave equation, Laplace equation, Helmholtz equation and biharmonic equation.

The analytical method of separation of variables for solving partial differential equations has also been generalized into a computational method of decomposition in invariant structures that can be used to solve systems of partial differential equations. [2]

Example: homogeneous case

Consider the one-dimensional heat equation. The equation is

 

 

 

 

(1)

The variable u denotes temperature. The boundary condition is homogeneous, that is

 

 

 

 

(2)

Let us attempt to find a solution which is not identically zero satisfying the boundary conditions but with the following property: u is a product in which the dependence of u on x, t is separated, that is:

 

 

 

 

(3)

Substituting u back into equation ( 1 ) and using the product rule,

 

 

 

 

(4)

Since the right hand side depends only on x and the left hand side only on t, both sides are equal to some constant value −λ. Thus:

 

 

 

 

(5)

and

 

 

 

 

(6)

λ here is the eigenvalue for both differential operators, and T(t) and X(x) are corresponding eigenfunctions.

We will now show that solutions for X(x) for values of λ ≤ 0 cannot occur:

Suppose that λ < 0. Then there exist real numbers B, C such that

From ( 2 ) we get

 

 

 

 

(7)

and therefore B = 0 = C which implies u is identically 0.

Suppose that λ = 0. Then there exist real numbers B, C such that

From ( 7 ) we conclude in the same manner as in 1 that u is identically 0.

Therefore, it must be the case that λ > 0. Then there exist real numbers A, B, C such that

and

From ( 7 ) we get C = 0 and that for some positive integer n,

This solves the heat equation in the special case that the dependence of u has the special form of ( 3 ).

In general, the sum of solutions to ( 1 ) which satisfy the boundary conditions ( 2 ) also satisfies ( 1 ) and ( 3 ). Hence a complete solution can be given as

where Dn are coefficients determined by initial condition.

Given the initial condition

we can get

This is the sine series expansion of f(x) which is amenable to Fourier analysis. Multiplying both sides with and integrating over [0, L] results in

This method requires that the eigenfunctions X, here , are orthogonal and complete. In general this is guaranteed by Sturm–Liouville theory.

Example: nonhomogeneous case

Suppose the equation is nonhomogeneous,

 

 

 

 

(8)

with the boundary condition the same as ( 2 ).

Expand h(x,t), u(x,t) and f(x) into

 

 

 

 

(9)

 

 

 

 

(10)

 

 

 

 

(11)

where hn(t) and bn can be calculated by integration, while un(t) is to be determined.

Substitute ( 9 ) and ( 10 ) back to ( 8 ) and considering the orthogonality of sine functions we get

which are a sequence of linear differential equations that can be readily solved with, for instance, Laplace transform, or Integrating factor. Finally, we can get

If the boundary condition is nonhomogeneous, then the expansion of ( 9 ) and ( 10 ) is no longer valid. One has to find a function v that satisfies the boundary condition only, and subtract it from u. The function u-v then satisfies homogeneous boundary condition, and can be solved with the above method.

Example: mixed derivatives

For some equations involving mixed derivatives, the equation does not separate as easily as the heat equation did in the first example above, but nonetheless separation of variables may still be applied. Consider the two-dimensional biharmonic equation

Proceeding in the usual manner, we look for solutions of the form

and we obtain the equation

Writing this equation in the form

Taking the derivative of this expression with respect to gives which means or and likewise, taking derivative with respect to leads to and thus or , hence either F(x) or G(y) must be a constant, say −λ. This further implies that either or are constant. Returning to the equation for X and Y, we have two cases

and

which can each be solved by considering the separate cases for and noting that .

Curvilinear coordinates

In orthogonal curvilinear coordinates, separation of variables can still be used, but in some details different from that in Cartesian coordinates. For instance, regularity or periodic condition may determine the eigenvalues in place of boundary conditions. See spherical harmonics for example.

Applicability

Partial differential equations

For many PDEs, such as the wave equation, Helmholtz equation and Schrodinger equation, the applicability of separation of variables is a result of the spectral theorem. In some cases, separation of variables may not be possible. Separation of variables may be possible in some coordinate systems but not others, [3] and which coordinate systems allow for separation depends on the symmetry properties of the equation. [4] Below is an outline of an argument demonstrating the applicability of the method to certain linear equations, although the precise method may differ in individual cases (for instance in the biharmonic equation above).

Consider an initial boundary value problem for a function on in two variables:

where is a differential operator with respect to and is a differential operator with respect to with boundary data:

for
for

where is a known function.

We look for solutions of the form . Dividing the PDE through by gives

The right hand side depends only on and the left hand side only on so both must be equal to a constant , which gives two ordinary differential equations

which we can recognize as eigenvalue problems for the operators for and . If is a compact, self-adjoint operator on the space along with the relevant boundary conditions, then by the Spectral theorem there exists a basis for consisting of eigenfunctions for . Let the spectrum of be and let be an eigenfunction with eigenvalue . Then for any function which at each time is square-integrable with respect to , we can write this function as a linear combination of the . In particular, we know the solution can be written as

For some functions . In the separation of variables, these functions are given by solutions to

Hence, the spectral theorem ensures that the separation of variables will (when it is possible) find all the solutions.

For many differential operators, such as , we can show that they are self-adjoint by integration by parts. While these operators may not be compact, their inverses (when they exist) may be, as in the case of the wave equation, and these inverses have the same eigenfunctions and eigenvalues as the original operator (with the possible exception of zero). [5]

Matrices

The matrix form of the separation of variables is the Kronecker sum.

As an example we consider the 2D discrete Laplacian on a regular grid:

where and are 1D discrete Laplacians in the x- and y-directions, correspondingly, and are the identities of appropriate sizes. See the main article Kronecker sum of discrete Laplacians for details.

Software

Some mathematical programs are able to do separation of variables: Xcas [6] among others.

See also

Notes

  1. "Separation of Variables". www.mathsisfun.com. Retrieved 2021-09-18.
  2. Miroshnikov, Victor A. (15 December 2017). Harmonic Wave Systems: Partial Differential Equations of the Helmholtz Decomposition. ISBN   9781618964069.
  3. John Renze, Eric W. Weisstein, Separation of variables
  4. Willard Miller(1984) Symmetry and Separation of Variables, Cambridge University Press
  5. David Benson (2007) Music: A Mathematical Offering, Cambridge University Press, Appendix W
  6. "Symbolic algebra and Mathematics with Xcas" (PDF).

Related Research Articles

In calculus, the chain rule is a formula that expresses the derivative of the composition of two differentiable functions f and g in terms of the derivatives of f and g. More precisely, if is the function such that for every x, then the chain rule is, in Lagrange's notation,

<span class="mw-page-title-main">Catenary</span> Curve formed by a hanging chain

In physics and geometry, a catenary is the curve that an idealized hanging chain or cable assumes under its own weight when supported only at its ends in a uniform gravitational field.

<span class="mw-page-title-main">Laplace's equation</span> Second-order partial differential equation

In mathematics and physics, Laplace's equation is a second-order partial differential equation named after Pierre-Simon Laplace, who first studied its properties. This is often written as

<span class="mw-page-title-main">Heat equation</span> Partial differential equation describing the evolution of temperature in a region

In mathematics and physics, the heat equation is a certain partial differential equation. Solutions of the heat equation are sometimes known as caloric functions. The theory of the heat equation was first developed by Joseph Fourier in 1822 for the purpose of modeling how a quantity such as heat diffuses through a given region.

The calculus of variations is a field of mathematical analysis that uses variations, which are small changes in functions and functionals, to find maxima and minima of functionals: mappings from a set of functions to the real numbers. Functionals are often expressed as definite integrals involving functions and their derivatives. Functions that maximize or minimize functionals may be found using the Euler–Lagrange equation of the calculus of variations.

In mathematics, an implicit equation is a relation of the form where R is a function of several variables. For example, the implicit equation of the unit circle is

In mathematics, the Hodge star operator or Hodge star is a linear map defined on the exterior algebra of a finite-dimensional oriented vector space endowed with a nondegenerate symmetric bilinear form. Applying the operator to an element of the algebra produces the Hodge dual of the element. This map was introduced by W. V. D. Hodge.

<span class="mw-page-title-main">Leibniz's notation</span> Mathematical notation used for calculus

In calculus, Leibniz's notation, named in honor of the 17th-century German philosopher and mathematician Gottfried Wilhelm Leibniz, uses the symbols dx and dy to represent infinitely small increments of x and y, respectively, just as Δx and Δy represent finite increments of x and y, respectively.

In mathematics, integral equations are equations in which an unknown function appears under an integral sign. In mathematical notation, integral equations may thus be expressed as being of the form:

In mathematics and its applications, a Sturm–Liouville problem is a second-order linear ordinary differential equation of the form:

In mathematics, an Euler–Cauchy equation, or Cauchy–Euler equation, or simply Euler's equation is a linear homogeneous ordinary differential equation with variable coefficients. It is sometimes referred to as an equidimensional equation. Because of its particularly simple equidimensional structure, the differential equation can be solved explicitly.

<span class="mw-page-title-main">Differential equation</span> Type of functional equation (mathematics)

In mathematics, a differential equation is an equation that relates one or more unknown functions and their derivatives. In applications, the functions generally represent physical quantities, the derivatives represent their rates of change, and the differential equation defines a relationship between the two. Such relations are common; therefore, differential equations play a prominent role in many disciplines including engineering, physics, economics, and biology.

In mathematics, an exact differential equation or total differential equation is a certain kind of ordinary differential equation which is widely used in physics and engineering.

A differential equation can be homogeneous in either of two respects.

In classical mechanics, holonomic constraints are relations between the position variables that can be expressed in the following form:

In differential calculus, there is no single uniform notation for differentiation. Instead, various notations for the derivative of a function or variable have been proposed by various mathematicians. The usefulness of each notation varies with the context, and it is sometimes advantageous to use more than one notation in a given context. The most common notations for differentiation are listed below.

In mathematics, the spectral theory of ordinary differential equations is the part of spectral theory concerned with the determination of the spectrum and eigenfunction expansion associated with a linear ordinary differential equation. In his dissertation, Hermann Weyl generalized the classical Sturm–Liouville theory on a finite closed interval to second order differential operators with singularities at the endpoints of the interval, possibly semi-infinite or infinite. Unlike the classical case, the spectrum may no longer consist of just a countable set of eigenvalues, but may also contain a continuous part. In this case the eigenfunction expansion involves an integral over the continuous part with respect to a spectral measure, given by the Titchmarsh–Kodaira formula. The theory was put in its final simplified form for singular differential equations of even degree by Kodaira and others, using von Neumann's spectral theorem. It has had important applications in quantum mechanics, operator theory and harmonic analysis on semisimple Lie groups.

The Beltrami identity, named after Eugenio Beltrami, is a special case of the Euler–Lagrange equation in the calculus of variations.

In the study of ordinary differential equations and their associated boundary value problems, Lagrange's identity, named after Joseph Louis Lagrange, gives the boundary terms arising from integration by parts of a self-adjoint linear differential operator. Lagrange's identity is fundamental in Sturm–Liouville theory. In more than one independent variable, Lagrange's identity is generalized by Green's second identity.

The Fokas method, or unified transform, is an algorithmic procedure for analysing boundary value problems for linear partial differential equations and for an important class of nonlinear PDEs belonging to the so-called integrable systems. It is named after Greek mathematician Athanassios S. Fokas.

References