Sturm-Liouville theory: Difference between revisions

From Citizendium
Jump to navigation Jump to search
imported>Dan Nessett
(Initial import of S-L article from Wikipedia)
 
imported>Paul Wormer
(→‎Sturm–Liouville equations as self-adjoint differential operators: Orthogonality explicit; yet another WP error corrected.)
 
(19 intermediate revisions by 4 users not shown)
Line 1: Line 1:
In [[mathematics]] and its applications, a classical '''Sturm–Liouville equation''', named after [[Jacques Charles François Sturm]] (1803–1855) and [[Joseph Liouville]] (1809–1882), is a real second-order linear [[differential equation]] of the form
{{subpages}}


{{NumBlk|:|<math> -\frac{d}{dx}\left[p(x)\frac{dy}{ dx}\right]+q(x)y=\lambda w(x)y.</math>|{{EquationRef|1}}}}
In [[mathematics]] and its applications, a classical '''Sturm–Liouville equation''' is a real second-order linear [[differential equation]] of the form:


where ''y'' is a function of the free variable ''x''. Here the functions ''p''(''x'')&nbsp;>&nbsp;0 has a ''continuous derivative'', ''q''(''x''), and ''w''(''x'')&nbsp;>&nbsp;0 are specified at the outset, and in the simplest of cases are continuous on the finite closed interval [''a'',''b'']. In addition, the function ''y'' is typically required to satisfy some [[boundary condition]]s at ''a'' and ''b''. The function ''w''(''x''), which is sometimes called ''r''(''x''), is called the "weight" or "density" function.
<div style="text-align: right;">
<div style="float: left;  margin-left: 50px;"><math> -\frac{d}{dx}\left[p(x)\frac{dy}{ dx}\right]+q(x)y=\lambda w(x)y,</math></div>
<span id="(1)">(1)</span>
</div><br>


The value of &lambda; is not specified in the equation; finding the values of &lambda; for which there exists a non-trivial solution of '''('''{{EquationNote|1}}''')''' satisfying the boundary conditions is part of the problem called the '''Sturm–Liouville problem''' (S&nbsp;L).  
where ''y'' is a function of the free variable ''x''. Here the functions ''p''(''x'')&nbsp;>&nbsp;0 has a ''continuous derivative'', ''q''(''x''), and ''w''(''x'')&nbsp;>&nbsp;0 are specified at the outset, and in the simplest of cases are continuous on the finite closed interval [''a'',''b'']. In addition, the function ''y'' is typically required to satisfy some [[boundary condition]]s at ''a'' and ''b''. The function ''w''(''x''), which is sometimes called ''r''(''x''), is called the "weight" or "density" function. The equation is named after [[Jacques Charles François Sturm]] (1803&ndash;1855) and [[Joseph Liouville]] (1809&ndash;1882).


Such values of &lambda; when they exist are called the [[eigenvalues]] of the boundary value problem defined by '''('''{{EquationNote|1}}''')''' and the prescribed set of boundary conditions. The corresponding solutions (for such a &lambda;) are the [[eigenfunction]]s of this problem. Under normal assumptions on the coefficient functions ''p''(''x''), ''q''(''x''), and ''w''(''x'') above, they induce a [[hermitian operator|Hermitian]] [[differential operator]] in some [[function space]] defined by [[boundary value problem|boundary conditions]]. The resulting theory of the existence and asymptotic behavior of the eigenvalues, the corresponding qualitative theory of the eigenfunctions and their [[Complete metric space|completeness]] in a suitable [[function space]] became known as '''Sturm–Liouville theory'''. This theory is important in applied mathematics, where S–L problems  occur very commonly, particularly when dealing with linear [[partial differential equation]]s that are [[separation of variables|separable]].
The value of &lambda; is not specified in the equation; finding the values of &lambda;  for which there exists a non-trivial solution of [[#(1) | (1)]] satisfying the boundary conditions is part of the problem called the '''Sturm–Liouville problem''' (S–L).
 
Such values of &lambda; when they exist are called the [[eigenvalues]] of the boundary value problem defined by [[#(1) | (1)]] and the prescribed set of boundary conditions. The corresponding solutions (for such a &lambda;) are the [[eigenfunction]]s of this problem. Under normal assumptions on the coefficient functions ''p''(''x''), ''q''(''x''), and ''w''(''x'') above, they induce a [[hermitian operator|Hermitian]] [[differential operator]] in some [[function space]] defined by [[boundary value problem|boundary conditions]]. The resulting theory of the existence and asymptotic behavior of the eigenvalues, the corresponding qualitative theory of the eigenfunctions and their [[Complete metric space|completeness]] in a suitable [[function space]] became known as '''Sturm–Liouville theory'''. This theory is important in applied mathematics, where S–L problems  occur very commonly, particularly when dealing with linear [[partial differential equation]]s that are [[separation of variables|separable]].


== Sturm–Liouville theory ==
== Sturm–Liouville theory ==


Under the assumptions that the S–L problem is regular, that is, ''p''(''x'')<sup>&minus;1</sup>&nbsp;>&nbsp;0, ''q''(''x''), and ''w''(''x'')&nbsp;>&nbsp;0 are real-valued [[Lebesgue integrable|integrable]] functions over the finite interval [''a'',&nbsp;''b''],
Under the assumptions that the S–L problem is regular, that is, ''p''(''x'')<sup>&minus;1</sup>&nbsp;>&nbsp;0, ''q''(''x''), and ''w''(''x'')&nbsp;>&nbsp;0 are real-valued [[Lebesgue integrable|integrable]] functions over the finite interval [''a'',&nbsp;''b''],
with ''separated boundary conditions'' of the form  
with ''separated boundary conditions'' of the form:


{{NumBlk|:|<math> y(a)\cos \alpha - p(a)y^\prime (a)\sin \alpha = 0,</math>|{{EquationRef|2}}}}
<div style="text-align: right;">
<div style="float: left; margin-left: 50px;"><math> y(a)\cos \alpha - p(a)y^\prime (a)\sin \alpha = 0,</math></div>
<span id="(2)">(2)</span>
</div>


{{NumBlk|:|<math> y(b)\cos \beta - p(b)y^\prime (b)\sin \beta = 0,</math>|{{EquationRef|3}}}}
<div style="text-align: right;">
<div style="float: left; margin-left: 50px;"><math> y(b)\cos \beta - p(b)y^\prime (b)\sin \beta = 0,</math></div>
<span id="(3)">(3)</span>
</div>


:where <math>\alpha, \beta \in [0, \pi),</math>
where <math>\alpha, \beta \in [0, \pi),</math>
the main tenet of '''Sturm–Liouville theory''' states that:
the main tenet of '''Sturm–Liouville theory''' states that:


* The eigenvalues &lambda;<sub>1</sub>, &lambda;<sub>2</sub>, &lambda;<sub>3</sub>, ... of the regular Sturm–Liouville problem '''('''{{EquationNote|1}}''')'''-'''('''{{EquationNote|2}}''')'''-'''('''{{EquationNote|3}}''')''' are real and can be ordered such that
* The eigenvalues &lambda;<sub>1</sub>, &lambda;<sub>2</sub>, &lambda;<sub>3</sub>, ... of the regular Sturm–Liouville problem [[#(1) | (1)]] -[[#(2) | (2)]] -[[#(3) | (3)]] are real and can be ordered such that:


:: <math>\lambda_1 < \lambda_2 < \lambda_3 < \cdots < \lambda_n < \cdots \to \infty; \, </math>
:: <math>\lambda_1 < \lambda_2 < \lambda_3 < \cdots < \lambda_n < \cdots \to \infty; \, </math>


* Corresponding to each eigenvalue &lambda;<sub>''n''</sub> is a unique (up to a normalization constant) eigenfunction ''y''<sub>''n''</sub>(''x'') which has exactly ''n''&nbsp;&minus;&nbsp;1 zeros in (''a'',&nbsp;''b''). The eigenfunction ''y''<sub>''n''</sub>(''x'') is called the ''n''-th ''fundamental solution'' satisfying the regular Sturm–Liouville problem '''('''{{EquationNote|1}}''')'''-'''('''{{EquationNote|2}}''')'''-'''('''{{EquationNote|3}}''')'''.
* Corresponding to each eigenvalue &lambda;<sub>''n''</sub> is a unique (up to a normalization constant) eigenfunction ''y''<sub>''n''</sub>(''x'') which has exactly ''n''&nbsp;&minus;&nbsp;1 zeros in (''a'',&nbsp;''b''). The eigenfunction ''y''<sub>''n''</sub>(''x'') is called the ''n''-th ''fundamental solution'' satisfying the regular Sturm–Liouville problem [[#(1) | (1)]] -[[#(2) | (2)]] -[[#(3) | (3)]].


* The normalized eigenfunctions form an [[orthonormal basis]]
* The normalized eigenfunctions form an [[orthonormal basis]]:


::<math> \int_a^b y_n(x)y_m(x)w(x)\,dx = \delta_{mn},</math>   
::<math> \int_a^b y_n(x)y_m(x)w(x)\,dx = \delta_{mn},</math>   
    
    
:in the [[Hilbert space]] [[Lebesgue space|''L''<sup>2</sup>([''a'',&nbsp;''b''],''w''(''x'')&nbsp;''dx'')]]. Here &delta;<sub>''mn''</sub> is a [[Kronecker delta]].
:in the [[Hilbert space]] [[Lebesgue space|''L''<sup>2</sup>([''a'',&nbsp;''b''],''w''(''x'')&nbsp;''dx'')]]. Here &delta;<sub>''mn''</sub> is a [[Kronecker delta]].
Since by assumption the eigenfunctions are normalized, the result is established by a [[Sturm-Liouville theory/Proofs | proof of their orthogonality]].


Note that, unless  ''p''(''x'') is continuously differentiable and ''q''(''x''), ''w''(''x'') are continuous, the equation has to be understood in a [[weak solution|weak sense]].
Note that, unless  ''p''(''x'') is continuously differentiable and ''q''(''x''), ''w''(''x'') are continuous, the equation has to be understood in a [[weak solution|weak sense]].


== Sturm–Liouville form ==
== Sturm–Liouville form ==
The differential equation '''('''{{EquationNote|1}}''')''' is said to be in '''Sturm–Liouville form''' or '''self-adjoint form'''. All second-order linear [[ordinary differential equation]]s can be recast in the form on the left-hand side of '''('''{{EquationNote|1}}''')''' by multiplying both sides of the equation by an appropriate [[integrating factor]] (although the same is not true of second-order [[partial differential equation]]s, or if ''y'' is a vector.)
The differential equation [[#(1) | (1)]] is said to be in '''Sturm–Liouville form''' or '''self-adjoint form'''. All second-order linear [[ordinary differential equation]]s can be recast in the form on the left-hand side of [[#(1) | (1)]] by multiplying both sides of the equation by an appropriate [[integrating factor]] (although the same is not true of second-order [[partial differential equation]]s, or if ''y'' is a vector.)


=== Examples ===
=== Examples ===
Line 43: Line 56:
: <math>x^2y''+xy'+(\lambda^2x^2-\nu^2)y=0\,</math>
: <math>x^2y''+xy'+(\lambda^2x^2-\nu^2)y=0\,</math>


can be written in Sturm–Liouville form as
can be written in Sturm–Liouville form as:


: <math>(xy')'+(\lambda^2 x-\nu^2/x)y=0.\,</math>
: <math>(xy')'+(\lambda^2 x-\nu^2/x)y=0.\,</math>


The [[Legendre polynomials|Legendre equation]],
The [[Legendre polynomials|Legendre equation]]:


: <math>(1-x^2)y''-2xy'+\nu(\nu+1)y=0\;\!</math>
: <math>(1-x^2)y''-2xy'+\nu(\nu+1)y=0\;\!</math>


can easily be put into Sturm–Liouville form, since ''D''(1&nbsp;&minus;&nbsp;''x''<sup>2</sup>) = &minus;2''x'', so, the Legendre equation is equivalent to
can easily be put into Sturm–Liouville form, since ''D''(1&nbsp;&minus;&nbsp;''x''<sup>2</sup>) = &minus;2''x'', so, the Legendre equation is equivalent to:


: <math>[(1-x^2)y']'+\nu(\nu+1)y=0\;\!</math>
: <math>[(1-x^2)y']'+\nu(\nu+1)y=0\;\!</math>


Less simple is such a differential equation as
It takes more work to put the following differential equation into Sturm–Liouville form:


: <math>x^3y''-xy'+2y=0.\,</math>
: <math>x^3y''-xy'+2y=0.\,</math>
Line 63: Line 76:
: <math>y''-{x\over x^3}y'+{2\over x^3}y=0</math>
: <math>y''-{x\over x^3}y'+{2\over x^3}y=0</math>


Multiplying throughout by an [[integrating factor]] of  
Multiplying throughout by an [[integrating factor]] of:


: <math>e^{\int -{x / x^3}\,dx}=e^{\int -{1 / x^2}\, dx}=e^{1 / x},</math>  
: <math>e^{\int -{x / x^3}\,dx}=e^{\int -{1 / x^2}\, dx}=e^{1 / x},</math>  


gives
gives:


: <math>e^{1 / x}y''-{e^{1 / x} \over x^2} y'+ {2 e^{1 / x} \over x^3} y = 0</math>
: <math>e^{1 / x}y''-{e^{1 / x} \over x^2} y'+ {2 e^{1 / x} \over x^3} y = 0</math>


which can be easily put into Sturm–Liouville form since
which can be easily put into Sturm–Liouville form since:


: <math>D e^{1 / x} = -{e^{1 / x} \over x^2} </math>
: <math>D e^{1 / x} = -{e^{1 / x} \over x^2} </math>


so the differential equation is equivalent to  
so the differential equation is equivalent to:


: <math>(e^{1 / x}y')'+{2 e^{1 / x} \over x^3} y =0.</math>
: <math>(e^{1 / x}y')'+{2 e^{1 / x} \over x^3} y =0.</math>


In general, given a differential equation
In general, given a differential equation:


: <math>P(x)y''+Q(x)y'+R(x)y=0\,</math>
: <math>P(x)y''+Q(x)y'+R(x)y=0\,</math>


dividing by ''P''(''x''), multiplying through by the integrating factor   
dividing by ''P''(''x''), multiplying through by the integrating factor:  


: <math>e^{\int {Q(x) / P(x)}\,dx},</math>
: <math>e^{\int {Q(x) / P(x)}\,dx},</math>
Line 90: Line 103:


== Sturm–Liouville equations as self-adjoint differential operators ==
== Sturm–Liouville equations as self-adjoint differential operators ==
Let us rewrite equation [[#(1) | (1)]] as
<div style="text-align: right;">
<div style="float: left;  margin-left: 50px;"><math>\Lambda\,y(x) = \lambda\,w(x)\,y(x) </math></div>
<span id="(1a)">(1a)</span>
</div>
with
:<math>
\Lambda \equiv  \left(-{d\over dx}\left[p(x){d\over dx}\right]+q(x) \right).
</math>
The function ''w''(''x'') is positive-definite and hence equation [[#(1a) | (1a)]] has the form of a generalized operator eigenvalue equation. It can be transformed to a regular eigenvalue equation by substitution of
:<math>
u(x) = w(x)^{1/2} y (x)\quad\hbox{and}\quad L = w(x)^{-1/2}\Lambda  w(x)^{-1/2}.
</math>
Equation [[#(1a) | (1a)]] becomes
:<math>
\left[w(x)^{-1/2} \Lambda w(x)^{-1/2}\right] \; w(x)^{1/2} y(x) = \lambda\, w(x)^{1/2}y(x)
</math>
or
<div style="text-align: right;">
<div style="float: left;  margin-left: 50px;"><math>L\, u = \lambda\,u. </math></div>
<span id="(1b)">(1b)</span>
</div>
The map ''L'' can be viewed as a [[linear operator]] mapping a function ''u'' to another function ''Lu''. We may study this linear operator in the context of [[functional analysis]].
Equation [[#(1b)|(1b)]] is precisely the [[eigenvalue]] problem of ''L''; that is, we are trying to find the eigenvalues &lambda;<sub>1</sub>, &lambda;<sub>2</sub>, &lambda;<sub>3</sub>, ... and the corresponding eigenvectors ''u''<sub>1</sub>, ''u''<sub>2</sub>, ''u''<sub>3</sub>, ... of the ''L'' operator. The proper setting for this problem is the [[Hilbert space]] [[Lp space#Weighted Lp spaces|''L''<sup>2</sup>([''a'',&nbsp;''b''],''w''(''x'')&nbsp;''dx'')]] with scalar product:


The map
: <math> \langle u_i, u_j\rangle = \int_{a}^{b} \overline{u_i(x)} u_j(x) \,dx
 
= \int_{a}^{b} \overline{y_i(x)} y_j(x) \, w(x)\, dx, \quad u_{k}(x) \equiv w(x)^{1/2} y_k(x).
: <math>L  u = {1 \over w(x)} \left(-{d\over dx}\left[p(x){du\over dx}\right]+q(x)u \right)</math>
</math>  
 
The functions ''y'' solve the generalized eigenvalue problem (1a) and the functions ''u'' the ordinary eigenvalue problem (1b).
can be viewed as a [[linear operator]] mapping a function ''u'' to another function ''Lu''. We may study this linear operator in the context of [[functional analysis]]. In fact, equation '''('''{{EquationNote|1}}''')''' can be written as
 
: <math>L  u  = \lambda u \,.</math>
 
This is precisely the [[eigenvalue]] problem; that is, we are trying to find the eigenvalues &lambda;<sub>1</sub>, &lambda;<sub>2</sub>, &lambda;<sub>3</sub>, ... and the corresponding eigenvectors ''u''<sub>1</sub>, ''u''<sub>2</sub>, ''u''<sub>3</sub>, ... of the ''L'' operator. The proper setting for this problem is the [[Hilbert space]] [[Lp space#Weighted Lp spaces|''L''<sup>2</sup>([''a'',&nbsp;''b''],''w''(''x'')&nbsp;''dx'')]] with
scalar product


: <math> \langle f, g\rangle = \int_{a}^{b} \overline{f(x)} g(x)w(x)\,dx.</math>   
In this space ''L'' is defined on sufficiently smooth functions which satisfy the above [[boundary value problem|boundary condition]]s. Moreover, ''L'' is a [[self-adjoint operator]].
This can be seen formally by using [[integration by parts]] twice, where the boundary terms vanish by virtue of the boundary conditions.  The functions ''w''(''x''),  ''p''(''x''), and ''q''(''x'') are real. From the vanishing of the boundary terms follows (d/d''x'')<sup>&lowast;</sup> =  &minus; d/d''x'', hence,
:<math>
L^* = (w^{-1/2} \Lambda w^{-1/2})^* = w^{-1/2} \Lambda^* w^{-1/2} \quad\hbox{and}\quad
\Lambda^* = \left(-{d\over dx}\left[p(x){d\over dx}\right]+q(x) \right)^* = \left(-{d\over dx}\left[p(x){d\over dx}\right]+q(x) \right).
</math>
Both ''L'' and &Lambda; are self-adjoint.  It then follows that the eigenvalues &lambda; shared by ''L''  and  &Lambda;  are real and that eigenfunctions of ''L'' corresponding to different eigenvalues are orthogonal. If
:<math>
Lu_k = \lambda_k u_k, \quad Lu_\ell = \lambda_\ell u_\ell \quad\hbox{with}\quad
\lambda_k \ne \lambda_\ell,
</math>
then
:<math>
\int_{a}^{b} \overline{u_k(x)} u_\ell(x) \,dx
= \int_{a}^{b} \overline{y_k(x)} y_\ell(x) \, w(x)\, dx = 0.
</math>   


In this space ''L'' is defined on sufficiently smooth functions which satisfy the above [[boundary value problem|boundary condition]]s. Moreover, ''L'' gives rise to a [[self-adjoint]] operator.
However, the operator ''L'' is unbounded and hence existence of an orthonormal basis of eigenfunctions is not evident. To overcome this problem one looks at the [[resolvent]]:
This can be seen formally by using [[integration by parts]] twice, where the boundary terms vanish by virtue of the boundary conditions. It then follows that the eigenvalues of a Sturm–Liouville operator are real and that eigenfunctions of ''L'' corresponding to different eigenvalues are orthogonal. However, this operator is unbounded and hence existence of an orthonormal basis of eigenfunctions is not evident. To overcome this problem one looks at the [[resolvent]]


::<math> (L - z)^{-1}, \qquad z \in\mathbb{C},</math>
::<math> (L - z)^{-1}, \qquad z \in\mathbb{C},</math>


where ''z'' is chosen to be some real number which is not an eigenvalue. Then, computing the resolvent amounts to solving the inhomogeneous equation, which can be done using the [[variation of parameters]] formula. This shows that the resolvent is an [[integral operator]] with a continuous symmetric kernel (the [[Green's function]] of the problem). As a consequence of the [[Arzelà–Ascoli theorem]] this integral operator is compact and existence of a sequence of eigenvalues &alpha;<sub>''n''</sub> which converge to 0 and eigenfunctions which form an orthonormal basis follows from the [[compact operator on Hilbert space|spectral theorem for compact operators]]. Finally, note that <math>(L-z)^{-1} u = \alpha u</math> is equivalent to <math>L u = (z+\alpha^{-1}) u</math>.
where ''z'' is chosen to be some complex number which is not an eigenvalue. Then, computing the resolvent amounts to solving the inhomogeneous equation, which can be done using the [[variation of parameters]] formula. This shows that the resolvent is an [[integral operator]] with a continuous symmetric kernel (the [[Green's function]] of the problem). As a consequence of the [[Arzelà–Ascoli theorem]] this integral operator is compact and existence of a sequence of eigenvalues &alpha;<sub>''n''</sub> which converge to 0 and eigenfunctions which form an orthonormal basis follows from the [[compact operator on Hilbert space|spectral theorem for compact operators]].  
Finally, note that <math>(L-z)^{-1} u = \alpha u</math> is equivalent to <math>L u = (z+\alpha^{-1}) u</math>.


If the interval is unbounded, or if the coefficients have singularities at the boundary points, one calls ''L'' singular. In this case the spectrum does no longer consist of eigenvalues alone and can contain a continuous component. There is still an associated eigenfunction expansion (similar to Fourier series versus Fourier transform). This is important in [[quantum mechanics]], since the one-dimensional [[Schrödinger equation]] is a special case of a S–L equation.
If the interval is unbounded, or if the coefficients have singularities at the boundary points, one calls ''L'' singular. In this case the spectrum no longer consists of eigenvalues alone and can contain a continuous component. There is still an associated eigenfunction expansion (similar to Fourier series versus Fourier transform). This is important in [[quantum mechanics]], since the one-dimensional [[Schrödinger equation]] is a special case of a S–L equation.


== Example ==
== Example ==
Line 117: Line 163:
We wish to find a function ''u''(''x'') which solves the following Sturm–Liouville problem:
We wish to find a function ''u''(''x'') which solves the following Sturm–Liouville problem:


{{NumBlk|:|<math> L  u  = \frac{d^2u}{dx^2} = \lambda u</math>|{{EquationRef|4}}}}
<div style="text-align: right;">
<div style="float: left; margin-left: 40px;"><math> L  u  = \frac{d^2u}{dx^2} = \lambda u</math></div>
<span id="(4)">(4)</span>
</div>
 


where the unknowns are ''&lambda;'' and ''u''(''x''). As above, we must add boundary conditions, we take for example
where the unknowns are ''&lambda;'' and ''u''(''x''). As above, we must add boundary conditions, we take for example:


:<math> u(0) = u(\pi) = 0 \, </math>
:<math> u(0) = u(\pi) = 0 \, </math>


Observe that if ''k'' is any integer, then the function
Observe that if ''k'' is any integer, then the function:


:<math> u(x) = \sin kx \, </math>
:<math> u(x) = \sin kx \, </math>


is a solution with eigenvalue &lambda; = &minus;''k''<sup>2</sup>. We know that the solutions of a S–L problem form an orthogonal basis, and we know from Fourier series that this set of sinusoidal functions is an orthogonal basis. Since orthogonal bases are always maximal (by definition) we conclude that the S–L problem in this case has no other eigenvectors.
is a solution with eigenvalue &lambda; = &minus;''k''<sup>2</sup>. We know that the solutions of a S–L problem form an orthogonal basis, and we know from the theory of Fourier series that this set of sinusoidal functions is an orthogonal basis. Since orthogonal bases are always maximal (by definition) we conclude that the S–L problem in this case has no other eigenvectors.


Given the preceding, let us now solve the inhomogeneous problem
Given the preceding, let us now solve the inhomogeneous problem:


:<math>L  u  =x, \qquad x\in(0,\pi)</math>
:<math>L  u  =x, \qquad x\in(0,\pi)</math>


with the same boundary conditions. In this case, we must write ''f''(''x'') = ''x'' in a Fourier series. The reader may check, either by integrating &int;exp(''ikx'')''x''&nbsp;d''x'' or by consulting a table of Fourier transforms, that we thus obtain
with the same boundary conditions. In this case, we must write ''f''(''x'') = ''x'' in a Fourier series. The reader may check, either by integrating &int;exp(''ikx'')''x''&nbsp;d''x'' or by consulting a table of Fourier transforms, that we thus obtain:


:<math>L  u  =\sum_{k=1}^{\infty}-2\frac{(-1)^k}{k}\sin kx.</math>
:<math>L  u  =\sum_{k=1}^{\infty}-2\frac{(-1)^k}{k}\sin kx.</math>
Line 139: Line 189:
This particular Fourier series is troublesome because of its poor convergence properties. It is not clear ''a priori'' whether the series converges pointwise. Because of Fourier analysis, since the Fourier coefficients are "square-summable", the Fourier series converges in ''L''<sup>2</sup> which is all we need for this particular theory to function. We mention for the interested reader that in this case we may rely on a result which says that Fourier's series converges at every point of differentiability, and at jump points (the function ''x'', considered as a periodic function, has a jump at &pi;) converges to the average of the left and right limits (see [[convergence of Fourier series]]).
This particular Fourier series is troublesome because of its poor convergence properties. It is not clear ''a priori'' whether the series converges pointwise. Because of Fourier analysis, since the Fourier coefficients are "square-summable", the Fourier series converges in ''L''<sup>2</sup> which is all we need for this particular theory to function. We mention for the interested reader that in this case we may rely on a result which says that Fourier's series converges at every point of differentiability, and at jump points (the function ''x'', considered as a periodic function, has a jump at &pi;) converges to the average of the left and right limits (see [[convergence of Fourier series]]).


Therefore, by using formula '''('''{{EquationNote|4}}''')''', we obtain that the solution is
Therefore, by using formula [[#(4) | (4)]], we obtain the solution:


:<math>u=\sum_{k=1}^{\infty}2\frac{(-1)^k}{k^3}\sin kx.</math>
:<math>u=\sum_{k=1}^{\infty}2\frac{(-1)^k}{k^3}\sin kx.</math>


In this case, we could have found the answer using antidifferentiation. This technique yields ''u'' =&nbsp;(''x''<sup>3</sup>&nbsp;&minus;&nbsp;''&pi;''<sup>2</sup>''x'')/6, whose Fourier series agrees with the solution we found. The antidifferentiation technique is no longer useful in most cases when the differential equation is in many variables.
In this case, we could have found the answer using antidifferentiation. This technique yields ''u'' =&nbsp;(''x''<sup>3</sup>&nbsp;&minus;&nbsp;''&pi;''<sup>2</sup>''x'')/6, whose Fourier series agrees with the solution we found. The antidifferentiation technique is not generally useful when the differential equation has many variables.


== Application to normal modes ==
== Application to normal modes ==
Line 151: Line 201:
:<math>\frac{\partial^2W}{\partial x^2}+\frac{\partial^2W}{\partial y^2} = \frac{1}{c^2}\frac{\partial^2W}{\partial t^2}.</math>
:<math>\frac{\partial^2W}{\partial x^2}+\frac{\partial^2W}{\partial y^2} = \frac{1}{c^2}\frac{\partial^2W}{\partial t^2}.</math>


The equation is [[separation of variables|separable]] (substituting ''W'' = ''X''(''x'') &times; ''Y''(''y'') &times; ''T''(''t'')), and the normal mode solutions that have [[harmonic]] time dependence and satisfy the boundary conditions ''W'' = 0 at ''x'' = 0, ''L''<sub>1</sub> and ''y'' = 0, ''L''<sub>2</sub> are given by
The equation is [[separation of variables|separable]] (substituting ''W'' = ''X''(''x'') &times; ''Y''(''y'') &times; ''T''(''t'')), and the normal mode solutions that have [[harmonic]] time dependence and satisfy the boundary conditions ''W'' = 0 at ''x'' = 0, ''L''<sub>1</sub> and ''y'' = 0, ''L''<sub>2</sub> are given by:


:<math>W_{mn}(x,y,t) = A_{mn}\sin\left(\frac{m\pi x}{L_1}\right)\sin\left(\frac{n\pi y}{L_2}\right)\cos\left(\omega_{mn}t\right)</math>
:<math>W_{mn}(x,y,t) = A_{mn}\sin\left(\frac{m\pi x}{L_1}\right)\sin\left(\frac{n\pi y}{L_2}\right)\cos\left(\omega_{mn}t\right)</math>


where ''m'' and ''n'' are non-zero [[integer]]s, ''A<sub>mn</sub>'' is an arbitrary constant and
where ''m'' and ''n'' are non-zero [[integer]]s, ''A<sub>mn</sub>'' is an arbitrary constant and:


: <math>\omega^2_{mn} = c^2 \left(\frac{m^2\pi^2}{L_1^2}+\frac{n^2\pi^2}{L_2^2}\right).</math>
: <math>\omega^2_{mn} = c^2 \left(\frac{m^2\pi^2}{L_1^2}+\frac{n^2\pi^2}{L_2^2}\right).</math>


Since the eigenfunctions ''W<sub>mn</sub>'' form a basis, an arbitrary initial displacement can be decomposed into a sum of these modes, which each vibrate at their individual frequencies <math>\omega_{mn}</math>. Infinite sums are also valid, as long as they [[convergence|converge]].
Since the eigenfunctions ''W<sub>mn</sub>'' form a basis, an arbitrary initial displacement can be decomposed into a sum of these modes, which each vibrate at their individual frequencies <math>\omega_{mn}</math>. Infinite sums are also valid, as long as they [[convergence|converge]].
==See also==
* [[Normal mode]]
* [[Self-adjoint]]
== References ==
* P. Hartman, ''Ordinary Differential Equations'', SIAM, Philadelphia, 2002 (2nd edition). ISBN 978-0-898715-10-1
* A. D. Polyanin and V. F. Zaitsev, ''Handbook of Exact Solutions for Ordinary Differential Equations'', Chapman & Hall/CRC Press, Boca Raton, 2003 (2nd edition). ISBN 1-58488-297-2
* G. Teschl, ''Ordinary Differential Equations and Dynamical Systems'', http://www.mat.univie.ac.at/~gerald/ftp/book-ode/ (Chapter 5)
* G. Teschl, ''Mathematical Methods in Quantum Mechanics and Applications to Schrödinger Operators'', http://www.mat.univie.ac.at/~gerald/ftp/book-schroe/ (see Chapter 9 for singular S-L operators and connections with quantum mechanics)
* A. Zettl, ''Sturm–Liouville Theory'', American Mathematical Society, 2005. ISBN 0-8218-3905-5.
[[Category:Ordinary differential equations]]
[[Category:Operator theory]]
[[Category:Spectral theory]]
[[de:Sturm-Liouville-Problem]]
[[es:Teoría de Sturm-Liouville]]
[[fr:Théorie de Sturm-Liouville]]
[[ko:스텀-리우빌 이론]]
[[it:Teoria di Sturm-Liouville]]
[[he:תורת שטורם-ליוביל]]
[[nl:Sturm-Liouvillevraagstuk]]
[[pt:Teoria de Sturm-Liouville]]
[[ru:Задача Штурма — Лиувилля]]
[[sv:Sturm-Liouvilles problem]]

Latest revision as of 04:52, 18 October 2009

This article is basically copied from an external source and has not been approved.
Main Article
Discussion
Related Articles  [?]
Bibliography  [?]
External Links  [?]
Citable Version  [?]
Proofs [?]
 
This editable Main Article is under development and subject to a disclaimer.
The content on this page originated on Wikipedia and is yet to be significantly improved. Contributors are invited to replace and add material to make this an original article.

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

(1)


where y is a function of the free variable x. Here the functions p(x) > 0 has a continuous derivative, q(x), and w(x) > 0 are specified at the outset, and in the simplest of cases are continuous on the finite closed interval [a,b]. In addition, the function y is typically required to satisfy some boundary conditions at a and b. The function w(x), which is sometimes called r(x), is called the "weight" or "density" function. The equation is named after Jacques Charles François Sturm (1803–1855) and Joseph Liouville (1809–1882).

The value of λ is not specified in the equation; finding the values of λ for which there exists a non-trivial solution of (1) satisfying the boundary conditions is part of the problem called the Sturm–Liouville problem (S–L).

Such values of λ when they exist are called the eigenvalues of the boundary value problem defined by (1) and the prescribed set of boundary conditions. The corresponding solutions (for such a λ) are the eigenfunctions of this problem. Under normal assumptions on the coefficient functions p(x), q(x), and w(x) above, they induce a Hermitian differential operator in some function space defined by boundary conditions. The resulting theory of the existence and asymptotic behavior of the eigenvalues, the corresponding qualitative theory of the eigenfunctions and their completeness in a suitable function space became known as Sturm–Liouville theory. This theory is important in applied mathematics, where S–L problems occur very commonly, particularly when dealing with linear partial differential equations that are separable.

Sturm–Liouville theory

Under the assumptions that the S–L problem is regular, that is, p(x)−1 > 0, q(x), and w(x) > 0 are real-valued integrable functions over the finite interval [ab], with separated boundary conditions of the form:

(2)

(3)

where the main tenet of Sturm–Liouville theory states that:

  • The eigenvalues λ1, λ2, λ3, ... of the regular Sturm–Liouville problem (1) - (2) - (3) are real and can be ordered such that:
  • Corresponding to each eigenvalue λn is a unique (up to a normalization constant) eigenfunction yn(x) which has exactly n − 1 zeros in (ab). The eigenfunction yn(x) is called the n-th fundamental solution satisfying the regular Sturm–Liouville problem (1) - (2) - (3).
in the Hilbert space L2([ab],w(xdx). Here δmn is a Kronecker delta.

Since by assumption the eigenfunctions are normalized, the result is established by a proof of their orthogonality.

Note that, unless p(x) is continuously differentiable and q(x), w(x) are continuous, the equation has to be understood in a weak sense.

Sturm–Liouville form

The differential equation (1) is said to be in Sturm–Liouville form or self-adjoint form. All second-order linear ordinary differential equations can be recast in the form on the left-hand side of (1) by multiplying both sides of the equation by an appropriate integrating factor (although the same is not true of second-order partial differential equations, or if y is a vector.)

Examples

The Bessel equation:

can be written in Sturm–Liouville form as:

The Legendre equation:

can easily be put into Sturm–Liouville form, since D(1 − x2) = −2x, so, the Legendre equation is equivalent to:

It takes more work to put the following differential equation into Sturm–Liouville form:

Divide throughout by x3:

Multiplying throughout by an integrating factor of:

gives:

which can be easily put into Sturm–Liouville form since:

so the differential equation is equivalent to:

In general, given a differential equation:

dividing by P(x), multiplying through by the integrating factor:

and then collecting gives the Sturm–Liouville form.

Sturm–Liouville equations as self-adjoint differential operators

Let us rewrite equation (1) as

(1a)

with

The function w(x) is positive-definite and hence equation (1a) has the form of a generalized operator eigenvalue equation. It can be transformed to a regular eigenvalue equation by substitution of

Equation (1a) becomes

or

(1b)

The map L can be viewed as a linear operator mapping a function u to another function Lu. We may study this linear operator in the context of functional analysis. Equation (1b) is precisely the eigenvalue problem of L; that is, we are trying to find the eigenvalues λ1, λ2, λ3, ... and the corresponding eigenvectors u1, u2, u3, ... of the L operator. The proper setting for this problem is the Hilbert space L2([ab],w(xdx) with scalar product:

The functions y solve the generalized eigenvalue problem (1a) and the functions u the ordinary eigenvalue problem (1b).

In this space L is defined on sufficiently smooth functions which satisfy the above boundary conditions. Moreover, L is a self-adjoint operator. This can be seen formally by using integration by parts twice, where the boundary terms vanish by virtue of the boundary conditions. The functions w(x), p(x), and q(x) are real. From the vanishing of the boundary terms follows (d/dx) = − d/dx, hence,

Both L and Λ are self-adjoint. It then follows that the eigenvalues λ shared by L and Λ are real and that eigenfunctions of L corresponding to different eigenvalues are orthogonal. If

then

However, the operator L is unbounded and hence existence of an orthonormal basis of eigenfunctions is not evident. To overcome this problem one looks at the resolvent:

where z is chosen to be some complex number which is not an eigenvalue. Then, computing the resolvent amounts to solving the inhomogeneous equation, which can be done using the variation of parameters formula. This shows that the resolvent is an integral operator with a continuous symmetric kernel (the Green's function of the problem). As a consequence of the Arzelà–Ascoli theorem this integral operator is compact and existence of a sequence of eigenvalues αn which converge to 0 and eigenfunctions which form an orthonormal basis follows from the spectral theorem for compact operators. Finally, note that is equivalent to .

If the interval is unbounded, or if the coefficients have singularities at the boundary points, one calls L singular. In this case the spectrum no longer consists of eigenvalues alone and can contain a continuous component. There is still an associated eigenfunction expansion (similar to Fourier series versus Fourier transform). This is important in quantum mechanics, since the one-dimensional Schrödinger equation is a special case of a S–L equation.

Example

We wish to find a function u(x) which solves the following Sturm–Liouville problem:

(4)


where the unknowns are λ and u(x). As above, we must add boundary conditions, we take for example:

Observe that if k is any integer, then the function:

is a solution with eigenvalue λ = −k2. We know that the solutions of a S–L problem form an orthogonal basis, and we know from the theory of Fourier series that this set of sinusoidal functions is an orthogonal basis. Since orthogonal bases are always maximal (by definition) we conclude that the S–L problem in this case has no other eigenvectors.

Given the preceding, let us now solve the inhomogeneous problem:

with the same boundary conditions. In this case, we must write f(x) = x in a Fourier series. The reader may check, either by integrating ∫exp(ikx)x dx or by consulting a table of Fourier transforms, that we thus obtain:

This particular Fourier series is troublesome because of its poor convergence properties. It is not clear a priori whether the series converges pointwise. Because of Fourier analysis, since the Fourier coefficients are "square-summable", the Fourier series converges in L2 which is all we need for this particular theory to function. We mention for the interested reader that in this case we may rely on a result which says that Fourier's series converges at every point of differentiability, and at jump points (the function x, considered as a periodic function, has a jump at π) converges to the average of the left and right limits (see convergence of Fourier series).

Therefore, by using formula (4), we obtain the solution:

In this case, we could have found the answer using antidifferentiation. This technique yields u = (x3 − π2x)/6, whose Fourier series agrees with the solution we found. The antidifferentiation technique is not generally useful when the differential equation has many variables.

Application to normal modes

Suppose we are interested in the modes of vibration of a thin membrane, held in a rectangular frame, 0 < x < L1, 0 < y < L2. We know the equation of motion for the vertical membrane's displacement, W(x, y, t) is given by the wave equation:

The equation is separable (substituting W = X(x) × Y(y) × T(t)), and the normal mode solutions that have harmonic time dependence and satisfy the boundary conditions W = 0 at x = 0, L1 and y = 0, L2 are given by:

where m and n are non-zero integers, Amn is an arbitrary constant and:

Since the eigenfunctions Wmn form a basis, an arbitrary initial displacement can be decomposed into a sum of these modes, which each vibrate at their individual frequencies . Infinite sums are also valid, as long as they converge.