0% found this document useful (0 votes)
287 views79 pages

Wave Equation - Wikipedia PDF

The wave equation describes waves in various physical contexts such as mechanical waves in fluids and solids. It is an important partial differential equation. The one-dimensional wave equation can be derived from properties like Hooke's law and describes waves traveling in opposite directions that can be added through superposition. The three-dimensional wave equation similarly describes wave propagation and behavior in three spatial dimensions.

Uploaded by

elvianadiyah
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
287 views79 pages

Wave Equation - Wikipedia PDF

The wave equation describes waves in various physical contexts such as mechanical waves in fluids and solids. It is an important partial differential equation. The one-dimensional wave equation can be derived from properties like Hooke's law and describes waves traveling in opposite directions that can be added through superposition. The three-dimensional wave equation similarly describes wave propagation and behavior in three spatial dimensions.

Uploaded by

elvianadiyah
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 79

Wave equation

A pulse traveling
through a string
with fixed
endpoints as
modeled by the
wave equation.
Spherical waves
coming from a
point source.

A solution to the 2D
wave equation

The wave equation is an important


second-order linear partial differential
equation for the description of waves—as
they occur in classical physics—such as
mechanical waves (e.g. water waves,
sound waves and seismic waves) or light
waves. It arises in fields like acoustics,
electromagnetics, and fluid dynamics.
Historically, the problem of a vibrating
string such as that of a musical
instrument was studied by Jean le Rond
d'Alembert, Leonhard Euler, Daniel
Bernoulli, and Joseph-Louis
Lagrange.[1][2][3][4] In 1746, d’Alembert
discovered the one-dimensional wave
equation, and within ten years Euler
discovered the three-dimensional wave
equation.[5]

Introduction
The wave equation is a partial differential
equation that may constrain some scalar
function u = u (x1, x2, …, xn; t) of a time
variable t and one or more spatial
variables x1, x2, … xn. The quantity u
may be, for example, the pressure in a
liquid or gas, or the displacement, along
some specific direction, of the particles of
a vibrating solid away from their resting
positions. The equation is

where c is a fixed non-negative real


coefficient.

Using the notations of Newtonian


mechanics and vector calculus, the wave
equation can be written more compactly
as
where denotes double time derivative,
∇ is the nabla operator, and ∇ = ∇ · ∇
2

is the (spatial) Laplacian operator:

A solution of this equation can be quite


complicated, but it can be analyzed as a
linear combination of simple solutions
that are sinusoidal plane waves with
various directions of propagation and
wavelengths but all with the same
propagation speed c. This analysis is
possible because the wave equation is
linear; so that any multiple of a solution is
also a solution, and the sum of any two
solutions is again a solution. This
property is called the superposition
principle in physics.

The wave equation alone does not


specify a physical solution; a unique
solution is usually obtained by setting a
problem with further conditions, such as
initial conditions, which prescribe the
amplitude and phase of the wave.
Another important class of problems
occurs in enclosed spaces specified by
boundary conditions, for which the
solutions represent standing waves, or
harmonics, analogous to the harmonics
of musical instruments.
The wave equation is the simplest
example of a hyperbolic differential
equation. It, and its modifications, play
fundamental roles in continuum
mechanics, quantum mechanics, plasma
physics, general relativity, geophysics,
and many other scientific and technical
disciplines.

Wave equation in one space


dimension
French scientist Jean-Baptiste le Rond d'Alembert
(b. 1717) discovered the wave equation in one
space dimension.[5]

The wave equation in one space


dimension can be written as follows:

This equation is typically described as


having only one space dimension x,
because the only other independent
variable is the time t. Nevertheless, the
dependent variable u may represent a
second space dimension, if, for example,
the displacement u takes place in y-
direction, as in the case of a string that is
located in the x–y plane.

Derivation of the wave


equation

The wave equation in one space


dimension can be derived in a variety of
different physical settings. Most
famously, it can be derived for the case
of a string that is vibrating in a two-
dimensional plane, with each of its
elements being pulled in opposite
directions by the force of tension.[6]

Another physical setting for derivation of


the wave equation in one space
dimension utilizes Hooke's Law. In the
theory of elasticity, Hooke's Law is an
approximation for certain materials,
stating that the amount by which a
material body is deformed (the strain) is
linearly related to the force causing the
deformation (the stress).

From Hooke's law

The wave equation in the one-


dimensional case can be derived from
Hooke's Law in the following way:
Imagine an array of little weights of mass
m interconnected with massless springs
of length h. The springs have a spring
constant of k:
Here the dependent variable u(x)
measures the distance from the
equilibrium of the mass situated at x, so
that u(x) essentially measures the
magnitude of a disturbance (i.e. strain)
that is traveling in an elastic material. The
forces exerted on the mass m at the
location x + h are:

The equation of motion for the weight at


the location x + h is given by equating
these two forces:
where the time-dependence of u(x) has
been made explicit.

If the array of weights consists of N


weights spaced evenly over the length
L = Nh of total mass M = Nm, and the
total spring constant of the array K = k/N
we can write the above equation as:

Taking the limit N → ∞, h → 0 and


assuming smoothness one gets:
which is from the definition of a second
derivative. KL2/M is the square of the
propagation speed in this particular case.

1-d standing wave as a superposition of two waves


traveling in opposite directions

Stress pulse in a bar

In the case of a stress pulse propagating


through a beam the beam acts much like
an infinite number of springs in series
and can be taken as an extension of the
equation derived for Hooke's law. A
beam of constant cross-section made
from a linear elastic material has a
stiffness K given by

Where A is the cross-sectional area and


E is the Young's modulus of the material.
The wave equation becomes

AL is equal to the volume of the beam


and therefore
where is the density of the material.
The wave equation reduces to

The speed of a stress wave in a bar is

therefore .

General solution

Algebraic approach

The one-dimensional wave equation is


unusual for a partial differential equation
in that a relatively simple general solution
may be found. Defining new variables:[7]
changes the wave equation into

which leads to the general solution

or equivalently:

In other words, solutions of the 1D wave


equation are sums of a right traveling
function F and a left traveling function G.
"Traveling" means that the shape of
these individual arbitrary functions with
respect to x stays constant, however the
functions are translated left and right with
time at the speed c. This was derived by
Jean le Rond d'Alembert.[8]

Another way to arrive at this result is to


note that the wave equation may be
"factored":

As a result, if we define v thus,

then
From this, v must have the form
G(x + ct), and from this the correct form
of the full solution u can be deduced.[9]

For an initial value problem, the arbitrary


functions F and G can be determined to
satisfy initial conditions:

The result is d'Alembert's formula:


In the classical sense if f(x) ∈ Ck and
g(x) ∈ Ck−1 then u(t, x) ∈ Ck. However,
the waveforms F and G may also be
generalized functions, such as the delta-
function. In that case, the solution may
be interpreted as an impulse that travels
to the right or the left.

The basic wave equation is a linear


differential equation and so it will adhere
to the superposition principle. This
means that the net displacement caused
by two or more waves is the sum of the
displacements which would have been
caused by each wave individually. In
addition, the behavior of a wave can be
analyzed by breaking up the wave into
components, e.g. the Fourier transform
breaks up a wave into sinusoidal
components.

Plane wave eigenmodes

Another way to solve for the solutions to


the one-dimensional wave equation is to
first analyze its frequency eigenmodes. A
so-called eigenmode is a solution that
oscillates in time with a well-defined
constant angular frequency , with which
the temporal part of the wave function for
such eigenmode takes a specific form
. The rest of the wave function is
then only dependent on the spatial
variable , hence amounting to
separation of variables. Now writing the
wave function as

we can obtain an ordinary differential


equation for the spatial part

Therefore:

which is precisely an eigenvalue equation


for , hence the name eigenmode. It
has the well-known plane wave solutions
with wave number .

The total wave function for this


eigenmode is then the linear combination

where complex numbers depend in


general on any initial and boundary
conditions of the problem.

Eigenmodes are useful in constructing a


full solution to the wave equation,
because each of them evolves in time
trivially with the phase factor . so
that a full solution can be decomposed
into an eigenmode expansion
or in terms of the plane waves,

which is exactly in the same form as in


the algebraic approach. Functions
are known as the Fourier
component and are determined by initial
and boundary conditions. This is a so-
called frequency-domain method,
alternative to direct time-domain
propagations, such as FDTD method, of
the wave packet , which is
complete for representing waves in
absence of time dilations. Completeness
of the Fourier expansion for representing
waves in the presence of time dilations
has been challenged by chirp wave
solutions allowing for time variation of
.[10] The chirp wave solutions seem
particularly implied by very large but
previously inexplicable radar residuals in
the flyby anomaly, and differ from the
sinusoidal solutions in being receivable at
any distance only at proportionally shifted
frequencies and time dilations,
corresponding to past chirp states of the
source.
Scalar wave equation in
three space dimensions

Swiss mathematician and physicist Leonhard Euler


(b. 1707) discovered the wave equation in three
space dimensions.[5]

A solution of the initial-value problem for


the wave equation in three space
dimensions can be obtained from the
corresponding solution for a spherical
wave. The result can then be also used
to obtain the same solution in two space
dimensions.

Spherical waves

The wave equation can be solved using


the technique of separation of variables.
To obtain a solution with constant
frequencies, let us first Fourier transform
the wave equation in time as

So we get,
This is the Helmholtz equation and can
be solved using separation of variables. If
spherical coordinates are used to
describe a problem, then the solution to
the angular part of the Helmholtz
equation is given by spherical harmonics
and the radial equation now becomes [11]

Here and the complete solution

is now given by
where and are the
spherical Hankel functions. To gain a
better understanding of the nature of
these spherical waves, let us go back
and look at the case when . In this
case, there is no angular dependence
and the amplitude depends only on the
radial distance i.e. . In
this case, the wave equation reduces to

This equation can be rewritten as


where the quantity satisfies the one-
dimensional wave equation. Therefore,
there are solutions in the form

where F and G are general solutions to


the one-dimensional wave equation, and
can be interpreted as respectively an
outgoing or incoming spherical wave.
Such waves are generated by a point
source, and they make possible sharp
signals whose form is altered only by a
decrease in amplitude as r increases
(see an illustration of a spherical wave on
the top right). Such waves exist only in
cases of space with odd dimensions.
For physical examples of non-spherical
wave solutions to the 3D wave equation
that do possess angular dependence,
see dipole radiation.

Monochromatic spherical wave

Cut-away of spherical wavefronts, with a


wavelength of 10 units, propagating from a point
source.

Although the word "monochromatic" is


not exactly accurate since it refers to light
or electromagnetic radiation with well-
defined frequency, the spirit is to
discover the eigenmode of the wave
equation in three dimensions. Following
the derivation in the previous section on
Plane wave eigenmodes, if we again
restrict our solutions to spherical waves
that oscillate in time with well-defined
constant angular frequency , then the
transformed function has simply
plane wave solutions,

or

.
From this we can observe that the peak
intensity of the spherical wave oscillation,
characterized as the squared wave
amplitude

drops at the rate proportional to , an


example of the inverse-square law.

Solution of a general initial-


value problem

The wave equation is linear in u and it is


left unaltered by translations in space
and time. Therefore, we can generate a
great variety of solutions by translating
and summing spherical waves. Let
φ(ξ, η, ζ) be an arbitrary function of three
independent variables, and let the
spherical wave form F be a delta
function: that is, let F be a weak limit of
continuous functions whose integral is
unity, but whose support (the region
where the function is non-zero) shrinks
to the origin. Let a family of spherical
waves have center at (ξ, η, ζ), and let r
be the radial distance from that point.
Thus

If u is a superposition of such waves with


weighting function φ, then
the denominator 4πc is a convenience.

From the definition of the delta function, u


may also be written as

where α, β, and γ are coordinates on the


unit sphere S, and ω is the area element
on S. This result has the interpretation
that u(t, x) is t times the mean value of φ
on a sphere of radius ct centered at x:

It follows that
The mean value is an even function of t,
and hence if

then

These formulas provide the solution for


the initial-value problem for the wave
equation. They show that the solution at
a given point P, given (t, x, y, z) depends
only on the data on the sphere of radius
ct that is intersected by the light cone
drawn backwards from P. It does not
depend upon data on the interior of this
sphere. Thus the interior of the sphere is
a lacuna for the solution. This
phenomenon is called Huygens'
principle. It is true for odd numbers of
space dimension, where for one
dimension the integration is performed
over the boundary of an interval with
respect to the Dirac measure. It is not
satisfied in even space dimensions. The
phenomenon of lacunas has been
extensively investigated in Atiyah, Bott
and Gårding (1970, 1973).

Scalar wave equation in two


space dimensions
In two space dimensions, the wave
equation is

We can use the three-dimensional


theory to solve this problem if we regard
u as a function in three dimensions that
is independent of the third dimension. If

then the three-dimensional solution


formula becomes

where α and β are the first two


coordinates on the unit sphere, and dω
is the area element on the sphere. This
integral may be rewritten as a double
integral over the disc D with center (x,y)
and radius ct:

It is apparent that the solution at (t, x, y)


depends not only on the data on the light
cone where

but also on data that are interior to that


cone.

Scalar wave equation in


general dimension and
Kirchhoff's formulae
We want to find solutions to utt − Δu = 0
for u : Rn × (0, ∞) → R with u(x, 0) = g(x)
and ut(x, 0) = h(x). See Evans for more
details.

Odd dimensions

Assume n ≥ 3 is an odd integer and


g ∈ Cm+1(Rn), h ∈ Cm(Rn) for
m = (n + 1)/2. Let
and let

then
u ∈ C2(Rn × [0, ∞))
utt − Δu = 0 in Rn × (0, ∞)

Even dimensions

Assume n ≥ 2 is an even integer and


g ∈ Cm+1(Rn), h ∈ Cm(Rn), for
m = (n + 2)/2. Let
and let

then
u ∈ C2(Rn × [0, ∞))
utt − Δu = 0 in Rn × (0, ∞)

Problems with boundaries


One space dimension

The Sturm–Liouville formulation

A flexible string that is stretched between


two points x = 0 and x = L satisfies the
wave equation for t > 0 and 0 < x < L. On
the boundary points, u may satisfy a
variety of boundary conditions. A general
form that is appropriate for applications is
where a and b are non-negative. The
case where u is required to vanish at an
endpoint is the limit of this condition when
the respective a or b approaches infinity.
The method of separation of variables
consists in looking for solutions of this
problem in the special form

A consequence is that

The eigenvalue λ must be determined so


that there is a non-trivial solution of the
boundary-value problem

This is a special case of the general


problem of Sturm–Liouville theory. If a
and b are positive, the eigenvalues are
all positive, and the solutions are
trigonometric functions. A solution that
satisfies square-integrable initial
conditions for u and ut can be obtained
from expansion of these functions in the
appropriate trigonometric series.

Investigation by numerical
methods
Approximating the continuous string with
a finite number of equidistant mass
points one gets the following physical
model:

Figure 1: Three consecutive mass points of the


discrete model for a string

If each mass point has the mass m, the


tension of the string is f, the separation
between the mass points is Δx and ui, i =
1, ..., n are the offset of these n points
from their equilibrium points (i.e. their
position on a straight line between the
two attachment points of the string) the
vertical component of the force towards
point i + 1 is

 
 

 
 
(1)

and the vertical component of the force


towards point i − 1 is

 
 

 
 
(2)

Taking the sum of these two forces and


dividing with the mass m one gets for the
vertical motion:
 
 

 
 
(3)

As the mass density is

this can be written

 
 

 
 

(4)

The wave equation is obtained by letting


Δx → 0 in which case ui(t) takes the form
u(x, t) where u(x, t) is continuous function

of two variables, takes the form

and
But the discrete formulation (3) of the
equation of state with a finite number of
mass point is just the suitable one for a
numerical propagation of the string
motion. The boundary condition

where L is the length of the string takes


in the discrete formulation the form that
for the outermost points u1 and un the
equation of motion are

 
 

 
 
(5)

and
 
 

 
 
(6)

while for 1 < i < n

 
 

 
 
(7)

where

If the string is approximated with 100


discrete mass points one gets the 100
coupled second order differential
equations (5), (6) and (7) or equivalently
200 coupled first order differential
equations.

Propagating these up to the times


using an 8th order multistep method the
6 states displayed in figure 2 are found:

Figure 2: The string at 6 consecutive epochs, the


first (red) corresponding to the initial time with the
string in rest
Figure 3: The string at 6 consecutive epochs

Figure 4: The string at 6 consecutive epochs


Figure 5: The string at 6 consecutive epochs

Figure 6: The string at 6 consecutive epochs


Figure 7: The string at 6 consecutive epochs

The red curve is the initial state at time


zero at which the string is "let free" in a
predefined shape[12] with all . The
blue curve is the state at time
i.e. after a time that corresponds to the
time a wave that is moving with the

nominal wave velocity would

need for one fourth of the length of the


string.
Figure 3 displays the shape of the string
at the times .
The wave travels in direction right with

the speed without being

actively constraint by the boundary


conditions at the two extremes of the
string. The shape of the wave is
constant, i.e. the curve is indeed of the
form f(x − ct).

Figure 4 displays the shape of the string


at the times
. The constraint on the right extreme
starts to interfere with the motion
preventing the wave to raise the end of
the string.
Figure 5 displays the shape of the string
at the times
when the direction of motion is reversed.
The red, green and blue curves are the
states at the times
while the 3
black curves correspond to the states at
times with
the wave starting to move back towards
left.

Figure 6 and figure 7 finally display the


shape of the string at the times
and
. The wave
now travels towards left and the
constraints at the end points are not
active any more. When finally the other
extreme of the string the direction will
again be reversed in a way similar to
what is displayed in figure 6.

Several space dimensions

A solution of the wave equation in two dimensions


with a zero-displacement boundary condition along
the entire outer edge.

The one-dimensional initial-boundary


value theory may be extended to an
arbitrary number of space dimensions.
Consider a domain D in m-dimensional x
space, with boundary B. Then the wave
equation is to be satisfied if x is in D and t
> 0. On the boundary of D, the solution u
shall satisfy

where n is the unit outward normal to B,


and a is a non-negative function defined
on B. The case where u vanishes on B is
a limiting case for a approaching infinity.
The initial conditions are

where f and g are defined in D. This


problem may be solved by expanding f
and g in the eigenfunctions of the
Laplacian in D, which satisfy the
boundary conditions. Thus the
eigenfunction v satisfies

in D, and

on B.

In the case of two space dimensions, the


eigenfunctions may be interpreted as the
modes of vibration of a drumhead
stretched over the boundary B. If B is a
circle, then these eigenfunctions have an
angular component that is a
trigonometric function of the polar angle
θ, multiplied by a Bessel function (of
integer order) of the radial component.
Further details are in Helmholtz equation.

If the boundary is a sphere in three


space dimensions, the angular
components of the eigenfunctions are
spherical harmonics, and the radial
components are Bessel functions of
half-integer order.

Inhomogeneous wave
equation in one dimension
The inhomogeneous wave equation in
one dimension is the following:
with initial conditions given by

The function s(x, t) is often called the


source function because in practice it
describes the effects of the sources of
waves on the medium carrying them.
Physical examples of source functions
include the force driving a wave on a
string, or the charge or current density in
the Lorenz gauge of electromagnetism.

One method to solve the initial value


problem (with the initial values as posed
above) is to take advantage of a special
property of the wave equation in an odd
number of space dimensions, namely
that its solutions respect causality. That
is, for any point (xi, ti), the value of u(xi, ti)
depends only on the values of f(xi + cti)
and f(xi − cti) and the values of the
function g(x) between (xi − cti) and
(xi + cti). This can be seen in d'Alembert's
formula, stated above, where these
quantities are the only ones that show up
in it. Physically, if the maximum
propagation speed is c, then no part of
the wave that can't propagate to a given
point by a given time can affect the
amplitude at the same point and time.
In terms of finding a solution, this
causality property means that for any
given point on the line being considered,
the only area that needs to be
considered is the area encompassing all
the points that could causally affect the
point being considered. Denote the area
that casually affects point (xi, ti) as RC.
Suppose we integrate the
inhomogeneous wave equation over this
region.

To simplify this greatly, we can use


Green's theorem to simplify the left side
to get the following:

The left side is now the sum of three line


integrals along the bounds of the
causality region. These turn out to be
fairly easy to compute

In the above, the term to be integrated


with respect to time disappears because
the time interval involved is zero, thus dt
= 0.
For the other two sides of the region, it is
worth noting that x ± ct is a constant,
namely xi ± cti, where the sign is chosen
appropriately. Using this, we can get the
relation dx ± cdt = 0, again choosing the
right sign:

And similarly for the final boundary


segment:
Adding the three results together and
putting them back in the original integral:

Solving for u(xi, ti) we arrive at


In the last equation of the sequence, the
bounds of the integral over the source
function have been made explicit.
Looking at this solution, which is valid for
all choices (xi, ti) compatible with the
wave equation, it is clear that the first two
terms are simply d'Alembert's formula, as
stated above as the solution of the
homogeneous wave equation in one
dimension. The difference is in the third
term, the integral over the source.

Other coordinate systems


In three dimensions, the wave equation,
when written in elliptic cylindrical
coordinates, may be solved by
separation of variables, leading to the
Mathieu differential equation.

Further generalizations
Elastic waves

The elastic wave equation in three


dimensions describes the propagation of
waves in an isotropic homogeneous
elastic medium. Most solid materials are
elastic, so this equation describes such
phenomena as seismic waves in the
Earth and ultrasonic waves used to
detect flaws in materials. While linear, this
equation has a more complex form than
the equations given above, as it must
account for both longitudinal and
transverse motion:

where:

λ and μ are the so-called Lamé


parameters describing the elastic
properties of the medium,
ρ is the density,
f is the source function (driving force),
and u is the displacement vector.

Note that in this equation, both force and


displacement are vector quantities. Thus,
this equation is sometimes known as the
vector wave equation. As an aid to
understanding, the reader will observe
that if f and ∇ ⋅ u are set to zero, this
becomes (effectively) Maxwell's equation
for the propagation of the electric field E,
which has only transverse waves.

Dispersion relation

In dispersive wave phenomena, the


speed of wave propagation varies with
the wavelength of the wave, which is
reflected by a dispersion relation

where ω is the angular frequency and k


is the wavevector describing plane wave
solutions. For light waves, the dispersion
relation is ω = ±c |k|, but in general, the
constant speed c gets replaced by a
variable phase velocity:

See also
Acoustic attenuation
Acoustic wave equation
Bateman transform
Electromagnetic wave equation
Helmholtz equation
Inhomogeneous electromagnetic wave
equation
Laplace operator
Mathematics of oscillation
Maxwell's equations
Schrödinger equation
Standing wave
Vibrations of a circular membrane
Wheeler–Feynman absorber theory

Notes
1. Cannon, John T.; Dostrovsky, Sigalia
(1981). The evolution of dynamics,
vibration theory from 1687 to 1742.
Studies in the History of Mathematics
and Physical Sciences. 6. New York:
Springer-Verlag. pp. ix + 184 pp.
ISBN 978-0-3879-0626-3. GRAY, JW
(July 1983). "BOOK REVIEWS". Bulletin
(New Series) of the American
Mathematical Society. 9 (1). (retrieved 13
Nov 2012).
2. Gerard F Wheeler. The Vibrating
String Controversy, (retrieved 13 Nov
2012). Am. J. Phys., 1987, v55, n1, p33–
37.
3. For a special collection of the 9
groundbreaking papers by the three
authors, see First Appearance of the
wave equation: D'Alembert, Leonhard
Euler, Daniel Bernoulli. – the controversy
about vibrating strings (retrieved 13 Nov
2012). Herman HJ Lynge and Son.
4. For de Lagrange's contributions to the
acoustic wave equation, one can consult
Acoustics: An Introduction to Its Physical
Principles and Applications Allan D.
Pierce, Acoustical Soc of America, 1989;
page 18 (retrieved 9 Dec 2012).
5. Speiser, David. Discovering the
Principles of Mechanics 1600–1800 , p.
191 (Basel: Birkhäuser, 2008).
6. Tipler, Paul and Mosca, Gene.
Physics for Scientists and Engineers,
Volume 1: Mechanics, Oscillations and
Waves; Thermodynamics , pp. 470–471
(Macmillan, 2004).
7. Eric W. Weisstein. "d'Alembert's
Solution" . MathWorld. Retrieved
2009-01-21.
8. D'Alembert (1747) "Recherches sur la
courbe que forme une corde tenduë
mise en vibration" (Researches on the
curve that a tense cord forms [when] set
into vibration), Histoire de l'académie
royale des sciences et belles lettres de
Berlin, vol. 3, pages 214–219.
See also: D'Alembert (1747) "Suite
des recherches sur la courbe que
forme une corde tenduë mise en
vibration" (Further researches on
the curve that a tense cord forms
[when] set into vibration), Histoire de
l'académie royale des sciences et
belles lettres de Berlin, vol. 3, pages
220–249.
See also: D'Alembert (1750)
"Addition au mémoire sur la courbe
que forme une corde tenduë mise
en vibration," Histoire de l'académie
royale des sciences et belles lettres
de Berlin, vol. 6, pages 355–360.
9.
http://math.arizona.edu/~kglasner/math4
56/linearwave.pdf .
10. V Guruprasad (2015), "Observational
evidence for travelling wave modes
bearing distance proportional shifts" ,
EPL, 110 (5): 54001, arXiv:1507.08222 ,
Bibcode:2015EL....11054001G ,
doi:10.1209/0295-5075/110/54001
11. John David Jackson, Classical
Electrodynamics, 3rd Edition, Wiley,
page 425. ISBN 978-0-471-30932-1
12. The initial state for "Investigation by
numerical methods" is set with quadratic
splines as follows:

for

for

for
with
References
M. F. Atiyah, R. Bott, L. Garding,
"Lacunas for hyperbolic differential
operators with constant coefficients I ",
Acta Math., 124 (1970), 109–189.
M.F. Atiyah, R. Bott, and L. Garding,
"Lacunas for hyperbolic differential
operators with constant coefficients
II ", Acta Math., 131 (1973), 145–206.
R. Courant, D. Hilbert, Methods of
Mathematical Physics, vol II.
Interscience (Wiley) New York, 1962.
L. Evans, "Partial Differential
Equations". American Mathematical
Society Providence, 1998.
"Linear Wave Equations ", EqWorld:
The World of Mathematical Equations.
"Nonlinear Wave Equations ",
EqWorld: The World of Mathematical
Equations.
William C. Lane, "MISN-0-201 The
Wave Equation and Its Solutions ",
Project PHYSNET .

External links
Nonlinear Wave Equations by
Stephen Wolfram and Rob Knapp,
Nonlinear Wave Equation Explorer by
Wolfram Demonstrations Project.
Mathematical aspects of wave
equations are discussed on the
Dispersive PDE Wiki .
Graham W Griffiths and William E.
Schiesser (2009). Linear and nonlinear
waves . Scholarpedia , 4(7):4308.
doi:10.4249/scholarpedia.4308

Wikimedia Commons has media


related to Wave equation.

Retrieved from
"https://en.wikipedia.org/w/index.php?
title=Wave_equation&oldid=891912306"

Last edited 15 days ago by Jorge …

Content is available under CC BY-SA 3.0


unless otherwise noted.

You might also like