Jump to content
 







Main menu
   


Navigation  



Main page
Contents
Current events
Random article
About Wikipedia
Contact us
Donate
 




Contribute  



Help
Learn to edit
Community portal
Recent changes
Upload file
 








Search  

































Create account

Log in
 









Create account
 Log in
 




Pages for logged out editors learn more  



Contributions
Talk
 



















Contents

   



(Top)
 


1 History  





2 Extrema  





3 EulerLagrange equation  



3.1  Example  







4 Beltrami's identity  





5 EulerPoisson equation  





6 Du Bois-Reymond's theorem  





7 Lavrentiev phenomenon  





8 Functions of several variables  



8.1  Dirichlet's principle  





8.2  Generalization to other boundary value problems  







9 Eigenvalue problems  



9.1  SturmLiouville problems  





9.2  Eigenvalue problems in several dimensions  







10 Applications  



10.1  Optics  



10.1.1  Snell's law  





10.1.2  Fermat's principle in three dimensions  



10.1.2.1  Connection with the wave equation  









10.2  Mechanics  





10.3  Further applications  







11 Variations and sufficient condition for a minimum  





12 See also  





13 Notes  





14 References  





15 Further reading  





16 External links  














Calculus of variations: Difference between revisions






العربية
Asturianu
Башҡортса
Беларуская
Български
Català
Чӑвашла
Čeština
Deutsch
Eesti
Ελληνικά
Español
Esperanto
Euskara
فارسی
Français
Galego

ि
Italiano
עברית
Қазақша
Magyar
Malti
Nederlands

Norsk bokmål
Norsk nynorsk
Piemontèis
Polski
Português
Română
Русский
Shqip
Simple English
Slovenčina
Slovenščina
Српски / srpski
Suomi
Svenska
Українська
Tiếng Vit



 

Edit links
 









Article
Talk
 

















Read
Edit
View history
 








Tools
   


Actions  



Read
Edit
View history
 




General  



What links here
Related changes
Upload file
Special pages
Permanent link
Page information
Cite this page
Get shortened URL
Download QR code
Wikidata item
 




Print/export  



Download as PDF
Printable version
 




Print/export  







In other projects  



Wikimedia Commons
Wikibooks
 
















Appearance
   

 





Help
 

From Wikipedia, the free encyclopedia
 


Browse history interactively
 Previous editNext edit 
Content deleted Content added
No edit summary
Reworded and other copy editing to make this article more compliant with MOS:MATH. Unicode math symbols like ∪, ∩, ≠, ≤, →, ∀, ∃, ∈,∑, ⊕, √, ∞, etc., Unicode letters like σ, ℱ, 𝔉, 𝔽, ℝ, ℂ, ℕ, etc., and other symbols that might not render correctly in some browsers should be avoided (e.g. by replacing them with LaTeX equivalents)
Line 16: Line 16:

[[Adrien-Marie Legendre|Legendre]] (1786) laid down a method, not entirely satisfactory, for the discrimination of maxima and minima. [[Isaac Newton]] and [[Gottfried Leibniz]] also gave some early attention to the subject.<ref name="brunt">{{cite book |last=van Brunt |first=Bruce |title=The Calculus of Variations |publisher=Springer |year=2004 |isbn=978-0-387-40247-5}}</ref> To this discrimination [[Vincenzo Brunacci]] (1810), [[Carl Friedrich Gauss]] (1829), [[Siméon Poisson]] (1831), [[Mikhail Ostrogradsky]] (1834), and [[Carl Gustav Jakob Jacobi|Carl Jacobi]] (1837) have been among the contributors. An important general work is that of [[Pierre Frédéric Sarrus|Sarrus]] (1842) which was condensed and improved by [[Cauchy]] (1844). Other valuable treatises and memoirs have been written by [[Strauch]] (1849), [[John Hewitt Jellett|Jellett]] (1850), [[Otto Hesse]] (1857), [[Alfred Clebsch]] (1858), and [[Carll]] (1885), but perhaps the most important work of the century is that of [[Weierstrass]]. His celebrated course on the theory is epoch-making, and it may be asserted that he was the first to place it on a firm and unquestionable foundation. The [[Hilbert's twentieth problem|20th]] and the [[Hilbert's twenty-third problem|23rd]] [[Hilbert problems|Hilbert problem]] published in 1900 encouraged further development.<ref name="brunt" />

[[Adrien-Marie Legendre|Legendre]] (1786) laid down a method, not entirely satisfactory, for the discrimination of maxima and minima. [[Isaac Newton]] and [[Gottfried Leibniz]] also gave some early attention to the subject.<ref name="brunt">{{cite book |last=van Brunt |first=Bruce |title=The Calculus of Variations |publisher=Springer |year=2004 |isbn=978-0-387-40247-5}}</ref> To this discrimination [[Vincenzo Brunacci]] (1810), [[Carl Friedrich Gauss]] (1829), [[Siméon Poisson]] (1831), [[Mikhail Ostrogradsky]] (1834), and [[Carl Gustav Jakob Jacobi|Carl Jacobi]] (1837) have been among the contributors. An important general work is that of [[Pierre Frédéric Sarrus|Sarrus]] (1842) which was condensed and improved by [[Cauchy]] (1844). Other valuable treatises and memoirs have been written by [[Strauch]] (1849), [[John Hewitt Jellett|Jellett]] (1850), [[Otto Hesse]] (1857), [[Alfred Clebsch]] (1858), and [[Carll]] (1885), but perhaps the most important work of the century is that of [[Weierstrass]]. His celebrated course on the theory is epoch-making, and it may be asserted that he was the first to place it on a firm and unquestionable foundation. The [[Hilbert's twentieth problem|20th]] and the [[Hilbert's twenty-third problem|23rd]] [[Hilbert problems|Hilbert problem]] published in 1900 encouraged further development.<ref name="brunt" />



In the 20th century [[David Hilbert]], [[Emmy Noether]], [[Leonida Tonelli]], [[Henri Lebesgue]] and [[Jacques Hadamard]] among others made significant contributions.<ref name="brunt" /> [[Marston Morse]] applied calculus of variations in what is now called [[Morse theory]].<ref name="ferguson">{{cite arXiv |last=Ferguson |first=James |eprint=math/0402357 |title= Brief Survey of the History of the Calculus of Variations and its Applications |year=2004 }}</ref> [[Lev Pontryagin]], [[R. Tyrrell Rockafellar|Ralph Rockafellar]] and F. H. Clarke developed new mathematical tools for the calculus of variations in [[optimal control theory]].<ref name="ferguson" /> The [[dynamic programming]] of [[Richard Bellman]] is an alternative to the calculus of variations.<ref>[[Dimitri Bertsekas]]. Dynamic programming and optimal control. Athena Scientific, 2005.

In the 20th century [[David Hilbert]], [[Emmy Noether]], [[Leonida Tonelli]], [[Henri Lebesgue]] and [[Jacques Hadamard]] among others made significant contributions.<ref name="brunt" /> [[Marston Morse]] applied calculus of variations in what is now called [[Morse theory]].<ref name="ferguson">{{cite arXiv |last=Ferguson |first=James |eprint=math/0402357 |title= Brief Survey of the History of the Calculus of Variations and its Applications |year=2004 }}</ref> [[Lev Pontryagin]], [[R. Tyrrell Rockafellar|Ralph Rockafellar]] and F. H. Clarke developed new mathematical tools for the calculus of variations in [[optimal control theory]].<ref name="ferguson" /> The [[dynamic programming]] of [[Richard Bellman]] is an alternative to the calculus of variations.<ref>[[Dimitri Bertsekas]]. Dynamic programming and optimal control. Athena Scientific, 2005.

</ref><ref name="bellman">{{cite journal |last=Bellman |first=Richard E. |title= Dynamic Programming and a new formalism in the calculus of variations |year=1954 |journal= Proc. Natl. Acad. Sci. | issue=4 | pages=231–235|pmc=527981 |pmid=16589462 |volume=40 |doi=10.1073/pnas.40.4.231|bibcode=1954PNAS...40..231B }}

</ref><ref name="bellman">{{cite journal |last=Bellman |first=Richard E. |title= Dynamic Programming and a new formalism in the calculus of variations |year=1954 |journal= Proc. Natl. Acad. Sci. | issue=4 | pages=231–235|pmc=527981 |pmid=16589462 |volume=40 |doi=10.1073/pnas.40.4.231|bibcode=1954PNAS...40..231B }}

</ref><ref>{{cite web |title=Richard E. Bellman Control Heritage Award |year=2004 |url=http://a2c2.org/awards/richard-e-bellman-control-heritage-award |work=American Automatic Control Council |access-date=2013-07-28}}</ref>{{efn|See '''[[Harold J. Kushner]] (2004)''': regarding Dynamic Programming, "The calculus of variations had related ideas (e.g., the work of Caratheodory, the Hamilton-Jacobi equation). This led to conflicts with the calculus of variations community."}}

</ref><ref>{{cite web |title=Richard E. Bellman Control Heritage Award |year=2004 |url=http://a2c2.org/awards/richard-e-bellman-control-heritage-award |work=American Automatic Control Council |access-date=2013-07-28}}</ref>{{efn|See '''[[Harold J. Kushner]] (2004)''': regarding Dynamic Programming, "The calculus of variations had related ideas (e.g., the work of Caratheodory, the Hamilton-Jacobi equation). This led to conflicts with the calculus of variations community."}}



== Extrema ==

== Extrema ==



The calculus of variations is concerned with the maxima or minima (collectively called '''extrema''') of functionals. A functional maps [[Function (mathematics)|functions]] to [[scalar (mathematics)|scalars]], so functionals have been described as "functions of functions." Functionals have extrema with respect to the elements {{math|''y''}} of a given [[function space]] defined over a given [[Domain of a function|domain]]. A functional {{math|''J'' [ ''y'' ]}} is said to have an extremum at the function {{math|''f''&nbsp; }}if{{math|''ΔJ'' {{=}} ''J'' [ ''y'' ] ''J'' [ ''f'']}} has the same [[Sign (mathematics)|sign]] for all {{math|''y''}} in an arbitrarily small neighborhood of {{math|''f'' .}}{{efn|The neighborhood of {{math|''f''}} is the part of the given function space where {{math|<nowiki>|</nowiki> ''y'' ''f''<nowiki>|</nowiki> <h}} over the whole domain of the functions, with {{math|h}} a positive number that specifies the size of the neighborhood.<ref name='CourHilb1953P169'>{{cite book |last1=Courant |first1=R |author-link1=Richard Courant |last2=Hilbert |first2=D |author-link2=David Hilbert |title = Methods of Mathematical Physics |volume=I |edition=First English |publisher=Interscience Publishers, Inc. |year=1953 |location=New York |page=169 |isbn=978-0471504474}}</ref>}} The function {{math|''f''}} is called an '''extremal''' function or extremal.{{efn|name=ExtremalVsExtremum| Note the difference between the terms extremal and extremum. An extremal is a function that makes a functional an extremum.}} The extremum {{math|''J'' [ ''f'' ]}} is called a local maximum if {{math|''ΔJ'' 0}} everywhere in an arbitrarily small neighborhood of {{math|''f'' ,}} and a local minimum if {{math|''ΔJ'' 0}} there. For a function space of continuous functions, extrema of corresponding functionals are called '''weak extrema''' or '''strong extrema''', depending on whether the first derivatives of the continuous functions are respectively all continuous or not.<ref name='GelfandFominPP12to13'>{{harvnb|Gelfand|Fomin|2000|pp=12–13}}</ref>

The calculus of variations is concerned with the maxima or minima (collectively called '''extrema''') of functionals. A functional maps [[Function (mathematics)|functions]] to [[scalar (mathematics)|scalars]], so functionals have been described as "functions of functions." Functionals have extrema with respect to the elements <math>y</math> of a given [[function space]] defined over a given [[Domain of a function|domain]]. A functional <math>J[y]</math> is said to have an extremum at the function <math>f</math>if<math>\Delta J = J[y] - J[f]</math> has the same [[Sign (mathematics)|sign]] for all <math>y</math> in an arbitrarily small neighborhood of <math>f.</math>{{efn|The neighborhood of <math>f</math> is the part of the given function space where <math>|y - f| <h</math> over the whole domain of the functions, with <math>h</math> a positive number that specifies the size of the neighborhood.<ref name='CourHilb1953P169'>{{cite book |last1=Courant |first1=R |author-link1=Richard Courant |last2=Hilbert |first2=D |author-link2=David Hilbert |title = Methods of Mathematical Physics |volume=I |edition=First English |publisher=Interscience Publishers, Inc. |year=1953 |location=New York |page=169 |isbn=978-0471504474}}</ref>}} The function <math>f</math> is called an '''extremal''' function or extremal.{{efn|name=ExtremalVsExtremum| Note the difference between the terms extremal and extremum. An extremal is a function that makes a functional an extremum.}} The extremum <math>J[f]</math> is called a local maximum if <math>\Delta J \leq 0</math> everywhere in an arbitrarily small neighborhood of <math>f,</math> and a local minimum if <math>\Delta J \geq 0</math> there. For a function space of continuous functions, extrema of corresponding functionals are called '''weak extrema''' or '''strong extrema''', depending on whether the first derivatives of the continuous functions are respectively all continuous or not.<ref name='GelfandFominPP12to13'>{{harvnb|Gelfand|Fomin|2000|pp=12–13}}</ref>



Both strong and weak extrema of functionals are for a space of continuous functions but strong extrema have the additional requirement that the first derivatives of the functions in the space be continuous. Thus a strong extremum is also a weak extremum, but the [[Converse (logic)|converse]] may not hold. Finding strong extrema is more difficult than finding weak extrema.<ref name='GelfandFominP13'>{{harvnb | Gelfand|Fomin| 2000 | p=13 }}</ref> An example of a [[Necessity and sufficiency|necessary condition]] that is used for finding weak extrema is the [[Euler–Lagrange equation]].<ref name='GelfandFominPP14to15'>{{harvnb | Gelfand|Fomin| 2000 | pp=14–15 }}</ref>{{efn|name=SectionVarSuffCond| For a sufficient condition, see section [[#Variations and sufficient condition for a minimum|Variations and sufficient condition for a minimum]].}}

Both strong and weak extrema of functionals are for a space of continuous functions but strong extrema have the additional requirement that the first derivatives of the functions in the space be continuous. Thus a strong extremum is also a weak extremum, but the [[Converse (logic)|converse]] may not hold. Finding strong extrema is more difficult than finding weak extrema.<ref name='GelfandFominP13'>{{harvnb | Gelfand|Fomin| 2000 | p=13 }}</ref> An example of a [[Necessity and sufficiency|necessary condition]] that is used for finding weak extrema is the [[Euler–Lagrange equation]].<ref name='GelfandFominPP14to15'>{{harvnb | Gelfand|Fomin| 2000 | pp=14–15 }}</ref>{{efn|name=SectionVarSuffCond| For a sufficient condition, see section [[#Variations and sufficient condition for a minimum|Variations and sufficient condition for a minimum]].}}

Line 32: Line 32:

Consider the functional

Consider the functional



:<math> J[y] = \int_{x_1}^{x_2} L(x,y(x),y'(x))\, dx \, .</math>

:<math>J[y] = \int_{x_1}^{x_2} L\left(x,y(x),y^{\prime}(x)\right)\, dx \, .</math>

where

where

:{{math|''x''<sub>1</sub>, ''x''<sub>2</sub>}} are [[Constant (mathematics)|constants]],

:<math>x_1, x_2</math> are [[Constant (mathematics)|constants]],

:{{math|''y'' (''x'')}} is twice continuously differentiable,

:<math>y(x)</math> is twice continuously differentiable,

:{{math|''y'' ′(''x'') {{=}} ''dy / dx'' &nbsp;}},

:<math>y^{\prime}(x) = \frac{dy}{dx},</math>

:{{math|''L''(''x'', ''y'' (''x''), ''y'' ′(''x''))}} is twice continuously differentiable with respect to its arguments {{math|''x'', &nbsp;''y'', &nbsp;''y'' ′}}.

:<math>L\left(x, y(x), y^{\prime}(x)\right)</math> is twice continuously differentiable with respect to its arguments <math>x, y,</math> and <math>y^{\prime}.</math>



If the functional {{math|''J''[''y'' ]}} attains a [[local minimum]] at {{math|''f'' ,}} and {{math|''η''(''x'')}} is an arbitrary function that has at least one derivative and vanishes at the endpoints {{math|''x''<sub>1</sub>}} and {{math|''x''<sub>2</sub> ,}} then for any number {{math|''ε''}} close to 0,

If the functional <math>J[y]</math> attains a [[local minimum]] at <math>f,</math> and <math>\eta(x)</math> is an arbitrary function that has at least one derivative and vanishes at the endpoints <math>x_1</math> and <math>x_2,</math> then for any number <math>\varepsilon</math> close to 0,

:<math>J[f] \le J[f + \varepsilon \eta] \, .</math>

:<math>J[f] \le J[f + \varepsilon \eta] \, .</math>



The term {{math|''εη''}} is called the '''variation''' of the function {{math|''f''}} and is denoted by {{math|''δf'' .}}<ref name='CourHilb1953P184'/>{{efn|Note that {{math|''η(x)''}} and {{math|''f(x)''}} are evaluated at the ''same'' values of {{math|''x''}}, which is not valid more generally in variational calculus with non-holonomic constraints.}}

The term <math>\varepsilon \eta</math> is called the '''variation''' of the function <math>f</math> and is denoted by <math>\delta f.</math><ref name='CourHilb1953P184'/>{{efn|Note that <math>\eta(x)</math> and <math>f(x)</math> are evaluated at the {{em|same}} values of <math>x,</math> which is not valid more generally in variational calculus with non-holonomic constraints.}}



Substituting &nbsp;{{math|''f'' + ''εη''}} for {{math|''y''}}&nbsp; in the functional {{math|''J''[ ''y'' ] ,}} the result is a function of {{math|''ε''}},

Substituting <math>f + \varepsilon \eta</math> for <math>y</math> in the functional <math>J[y],</math> the result is a function of <math>\varepsilon,</math>



:<math> \Phi(\varepsilon) = J[f+\varepsilon\eta] \, . </math>

:<math>\Phi(\varepsilon) = J[f+\varepsilon\eta] \, .</math>

Since the functional {{math|''J''[ ''y'' ]}} has a minimum for {{math|''y'' {{=}} ''f'' ,}} the function {{math(''ε'')}} has a minimum at {{math|''ε'' {{=}} 0}} and thus,{{efn|The product {{math|''ε''Φ′(0)}} is called the first variation of the functional {{math|''J''}} and is denoted by {{math|''δJ''}}. Some references define the [[first variation]] differently by leaving out the {{math|''ε''}} factor.}}

Since the functional <math>J[y]</math> has a minimum for <math>y = f</math> the function <math>\Phi(\varepsilon)</math> has a minimum at <math>\varepsilon = 0</math> and thus,{{efn|The product <math>\varepsilon \Phi^{\prime}(0)</math> is called the first variation of the functional <math>J</math> and is denoted by <math>\delta J.</math> Some references define the [[first variation]] differently by leaving out the <math>\varepsilon</math> factor.}}

:<math> \Phi'(0) \equiv \left.\frac{d\Phi}{d\varepsilon}\right|_{\varepsilon = 0} = \int_{x_1}^{x_2} \left.\frac{dL}{d\varepsilon}\right|_{\varepsilon = 0} dx = 0 \, . </math>

:<math>\Phi^{\prime}(0) \equiv \left.\frac{d\Phi}{d\varepsilon}\right|_{\varepsilon = 0} = \int_{x_1}^{x_2} \left.\frac{dL}{d\varepsilon}\right|_{\varepsilon = 0} dx = 0 \, .</math>



Taking the [[total derivative]] of {{math|''L''[''x'', ''y'', ''y'' ′] ,}} where {{math|''y'' {{=}} ''f'' + ''ε η''}} and {{math|''y'' {{=}} ''f'' ′ + ''ε η''′}} are considered as functions of {{math|''ε''}} rather than {{math|''x''}}, yields

Taking the [[total derivative]] of <math>L\left[x, y, y^{\prime}\right],</math> where <math>y = f + \varepsilon \eta</math> and <math>y^{\prime} = f^{\prime} + \varepsilon \eta^{\prime}</math> are considered as functions of <math>\varepsilon</math> rather than <math>x,</math> yields

:<math>

:<math>

\frac{dL}{d\varepsilon}=\frac{\partial L}{\partial y}\frac{dy}{d\varepsilon} + \frac{\partial L}{\partial y'}\frac{dy'}{d\varepsilon}

\frac{dL}{d\varepsilon}=\frac{\partial L}{\partial y}\frac{dy}{d\varepsilon} + \frac{\partial L}{\partial y^{\prime}}\frac{dy^{\prime}}{d\varepsilon}

</math>

</math>



and since &nbsp;{{math|''dy'' /''dε'' {{=}} ''η''}} &nbsp;and&nbsp; {{math|''dy'' ′/''dε'' {{=}} ''η' ''}},

and because &nbsp;<math>\frac{dy}{d \varepsilon} = \eta</math> &nbsp;and&nbsp; <math>\frac{d y^{\prime}}{d \varepsilon} = \eta^{\prime},</math>



:<math>

:<math>

\frac{dL}{d\varepsilon}=\frac{\partial L}{\partial y}\eta + \frac{\partial L}{\partial y'}\eta'.

\frac{dL}{d\varepsilon}=\frac{\partial L}{\partial y}\eta + \frac{\partial L}{\partial y^{\prime}}\eta^{\prime}.

</math>

</math>



Line 65: Line 65:

\begin{align}

\begin{align}

\int_{x_1}^{x_2} \left.\frac{dL}{d\varepsilon}\right|_{\varepsilon = 0} dx

\int_{x_1}^{x_2} \left.\frac{dL}{d\varepsilon}\right|_{\varepsilon = 0} dx

& = \int_{x_1}^{x_2} \left(\frac{\partial L}{\partial f} \eta + \frac{\partial L}{\partial f'} \eta'\right)\, dx \\

& = \int_{x_1}^{x_2} \left(\frac{\partial L}{\partial f} \eta + \frac{\partial L}{\partial f^{\prime}} \eta^{\prime}\right)\, dx \\

& = \int_{x_1}^{x_2} \frac{\partial L}{\partial f} \eta \, dx + \left.\frac{\partial L}{\partial f'} \eta \right|_{x_1}^{x_2} - \int_{x_1}^{x_2} \eta \frac{d}{dx}\frac{\partial L}{\partial f'} \, dx \\

& = \int_{x_1}^{x_2} \frac{\partial L}{\partial f} \eta \, dx + \left.\frac{\partial L}{\partial f^{\prime}} \eta \right|_{x_1}^{x_2} - \int_{x_1}^{x_2} \eta \frac{d}{dx}\frac{\partial L}{\partial f^{\prime}} \, dx \\

& = \int_{x_1}^{x_2} \left(\frac{\partial L}{\partial f} \eta - \eta \frac{d}{dx}\frac{\partial L}{\partial f'} \right)\, dx\\

& = \int_{x_1}^{x_2} \left(\frac{\partial L}{\partial f} \eta - \eta \frac{d}{dx}\frac{\partial L}{\partial f^{\prime}} \right)\, dx\\

\end{align}

\end{align}

</math>

</math>

where {{math|''L''[''x'', ''y'', ''y'' ′] ''L''[''x'', ''f'', ''f'' ′]}} when ''ε'' = 0 and we have used [[integration by parts]] on the second term. The second term on the second line vanishes because {{math|''η'' {{=}} 0}}at{{math|''x<sub>1</sub>''}} and {{math|''x<sub>2</sub>''}} by definition. Also, as previously mentioned the left side of the equation is zero so that

where <math>L\left[x, y, y^{\prime}\right] \toL\left[x, f, f^{\prime}\right]</math> when <math>\varepsilon = 0</math> and we have used [[integration by parts]] on the second term. The second term on the second line vanishes because <math>\eta = 0</math>at<math>x_1</math> and <math>x_2</math> by definition. Also, as previously mentioned the left side of the equation is zero so that

:<math> \int_{x_1}^{x_2} \eta (x) \left(\frac{\partial L}{\partial f} - \frac{d}{dx}\frac{\partial L}{\partial f'} \right) \, dx = 0 \, . </math>

:<math>\int_{x_1}^{x_2} \eta (x) \left(\frac{\partial L}{\partial f} - \frac{d}{dx}\frac{\partial L}{\partial f^{\prime}} \right) \, dx = 0 \, .</math>



According to the [[fundamental lemma of calculus of variations]], the part of the integrand in parentheses is zero, i.e.

According to the [[fundamental lemma of calculus of variations]], the part of the integrand in parentheses is zero, i.e.

:<math> \frac{\partial L}{\partial f} -\frac{d}{dx} \frac{\partial L}{\partial f'}=0 </math>

:<math>\frac{\partial L}{\partial f} -\frac{d}{dx} \frac{\partial L}{\partial f^{\prime}}=0</math>

which is called the '''Euler–Lagrange equation'''. The left hand side of this equation is called the [[functional derivative]] of {{math|''J''[''f'']}} and is denoted {{math|''δJ''/''δf''(''x'') .}}

which is called the '''Euler–Lagrange equation'''. The left hand side of this equation is called the [[functional derivative]] of <math>J[f]</math> and is denoted <math>\delta J/\delta f(x).</math>



In general this gives a second-order [[ordinary differential equation]] which can be solved to obtain the extremal function {{math| ''f''(''x'') .}} The Euler–Lagrange equation is a [[necessary condition|necessary]], but not [[sufficient condition|sufficient]], condition for an extremum {{math|''J''[''f'']}}. A sufficient condition for a minimum is given in the section [[#Variations and sufficient condition for a minimum|Variations and sufficient condition for a minimum]].

In general this gives a second-order [[ordinary differential equation]] which can be solved to obtain the extremal function <math>f(x).</math> The Euler–Lagrange equation is a [[Necessary condition|necessary]], but not [[Sufficient condition|sufficient]], condition for an extremum <math>J[f].</math> A sufficient condition for a minimum is given in the section [[#Variations and sufficient condition for a minimum|Variations and sufficient condition for a minimum]].



=== Example ===

=== Example ===

In order to illustrate this process, consider the problem of finding the extremal function {{math|''y'' {{=}} ''f'' (''x'') ,}} which is the shortest curve that connects two points {{math|(''x''<sub>1</sub>, ''y''<sub>1</sub>)}} and {{math|(''x''<sub>2</sub>, ''y''<sub>2</sub>) .}} The [[arc length]] of the curve is given by

In order to illustrate this process, consider the problem of finding the extremal function <math>y = f(x),</math> which is the shortest curve that connects two points <math>\left(x_1, y_1\right)</math> and <math>\left(x_2, y_2\right).</math> The [[arc length]] of the curve is given by



:<math> A[y] = \int_{x_1}^{x_2} \sqrt{1 + [y'(x) ]^2} \, dx \, , </math>

:<math>A[y] = \int_{x_1}^{x_2} \sqrt{1 + [y^{\prime}(x) ]^2} \, dx \, ,</math>



with

with

:<math> y\,'(x) = \frac{dy}{dx} \, , \ \ y_1=f(x_1) \, , \ \ y_2=f(x_2) \, . </math>

:<math>y\,^{\prime}(x) = \frac{dy}{dx} \, , \ \ y_1=f(x_1) \, , \ \ y_2=f(x_2) \, .</math>



{{efn|Note that assuming y is a function of x loses generality; ideally both should be a function of some other parameter. This approach is good solely for instructive purposes.}}

{{efn|Note that assuming y is a function of x loses generality; ideally both should be a function of some other parameter. This approach is good solely for instructive purposes.}}



The Euler–Lagrange equation will now be used to find the extremal function {{math|''f'' (''x'')}} that minimizes the functional {{math|''A''[''y'' ] .}}

The Euler–Lagrange equation will now be used to find the extremal function <math>f(x)</math> that minimizes the functional <math>A[y].</math>



:<math> \frac{\partial L}{\partial f} -\frac{d}{dx} \frac{\partial L}{\partial f'}=0 </math>

:<math>\frac{\partial L}{\partial f} -\frac{d}{dx} \frac{\partial L}{\partial f^{\prime}}=0</math>



with

with

:<math>L = \sqrt{1 + [f'(x) ]^2} \, . </math>

:<math>L = \sqrt{1 + [f^{\prime}(x) ]^2} \, .</math>



Since {{math|''f''}} does not appear explicitly in {{math|''L'' ,}} the first term in the Euler–Lagrange equation vanishes for all {{math|''f'' (''x'') }} and thus,

Since <math>f</math> does not appear explicitly in <math>L,</math> the first term in the Euler–Lagrange equation vanishes for all <math>f(x)</math> and thus,

:<math> \frac{d}{dx} \frac{\partial L}{\partial f'} = 0 \, . </math>

:<math>\frac{d}{dx} \frac{\partial L}{\partial f^{\prime}} = 0 \, .</math>

Substituting for {{math|''L''}} and taking the derivative,

Substituting for <math>L</math> and taking the derivative,

:<math> \frac{d}{dx} \ \frac{ f'(x) } {\sqrt{1 + [ f'(x) ]^2}} \ = 0 \, . </math>

:<math>\frac{d}{dx} \ \frac{f^{\prime}(x)} {\sqrt{1 + [f^{\prime}(x)]^2}} \ = 0 \, .</math>



Thus

Thus

:<math>\frac{f'(x)}{\sqrt{1+[f'(x)]^2}} = c \, , </math>

:<math>\frac{f^{\prime}(x)}{\sqrt{1+[f^{\prime}(x)]^2}} = c \, ,</math>

for some constant ''c''. Then

for some constant <math>c.</math> Then

:<math>\frac{[f'(x)]^2}{1+[f'(x)]^2} = c^2 \, , </math>

:<math>\frac{[f^{\prime}(x)]^2}{1+[f^{\prime}(x)]^2} = c^2 \, ,</math>

where

where

:<math>0\le c^2<1.</math>

:<math>0\le c^2<1.</math>

Solving, we get

Solving, we get

:<math>[f'(x)]^2=\frac{c^2}{1-c^2}\, </math>

:<math>[f^{\prime}(x)]^2=\frac{c^2}{1-c^2}\,</math>

which implies that

which implies that

:<math>f'(x)=m </math>

:<math>f^{\prime}(x)=m</math>

is a constant and therefore

is a constant and therefore

that the shortest curve that connects two points {{math|(''x''<sub>1</sub>, ''y''<sub>1</sub>)}} and {{math|(''x''<sub>2</sub>, ''y''<sub>2</sub>) }}is

that the shortest curve that connects two points <math>\left(x_1, y_1\right)</math> and <math>\left(x_2, y_2\right)</math>is

:<math> f(x) = m x + b \qquad \text{with} \ \ m = \frac{y_2 - y_1}{x_2 - x_1} \quad \text{and} \quad b = \frac{x_2 y_1 - x_1 y_2}{x_2 - x_1}</math>

:<math>f(x) = m x + b \qquad \text{with} \ \ m = \frac{y_2 - y_1}{x_2 - x_1} \quad \text{and} \quad b = \frac{x_2 y_1 - x_1 y_2}{x_2 - x_1}</math>

and we have thus found the extremal function {{math|''f''(''x'')}} that minimizes the functional {{math|''A''[''y'']}} so that {{math|''A''[''f'']}} is a minimum. The equation for a straight line is {{math|''y'' {{=}} ''f''(''x'')}}. In other words, the shortest distance between two points is a straight line.{{efn|name=ArchimedesStraight| As a historical note, this is an axiom of [[Archimedes]]. See e.g. Kelland (1843).<ref>{{cite book |last=Kelland |first=Philip |author-link=Philip Kelland|title=Lectures on the principles of demonstrative mathematics |year=1843 |page=58 |url=https://books.google.com/books?id=yQCFAAAAIAAJ&pg=PA58 |via=Google Books}}</ref>}}

and we have thus found the extremal function <math>f(x)</math> that minimizes the functional <math>A[y]</math> so that <math>A[f]</math> is a minimum. The equation for a straight line is <math>y = f(x).</math> In other words, the shortest distance between two points is a straight line.{{efn|name=ArchimedesStraight| As a historical note, this is an axiom of [[Archimedes]]. See e.g. Kelland (1843).<ref>{{cite book |last=Kelland |first=Philip |author-link=Philip Kelland|title=Lectures on the principles of demonstrative mathematics |year=1843 |page=58 |url=https://books.google.com/books?id=yQCFAAAAIAAJ&pg=PA58 |via=Google Books}}</ref>}}



== Beltrami's identity ==

== Beltrami's identity ==

In physics problems it may be the case that <math>\frac{\partial L}{\partial x} = 0</math>, meaning the integrand is a function of <math>f(x)</math> and <math>f'(x)</math> but <math>x</math> does not appear separately. In that case, the Euler–Lagrange equation can be simplified to the [[Beltrami identity]]<ref>{{cite web |author=Weisstein, Eric W. |url=http://mathworld.wolfram.com/Euler-LagrangeDifferentialEquation.html |title=Euler–Lagrange Differential Equation |website=mathworld.wolfram.com |publisher=Wolfram |at=Eq.&nbsp;(5)}}</ref>

In physics problems it may be the case that <math>\frac{\partial L}{\partial x} = 0,</math> meaning the integrand is a function of <math>f(x)</math> and <math>f^{\prime}(x)</math> but <math>x</math> does not appear separately. In that case, the Euler–Lagrange equation can be simplified to the [[Beltrami identity]]<ref>{{cite web |author=Weisstein, Eric W. |url=http://mathworld.wolfram.com/Euler-LagrangeDifferentialEquation.html |title=Euler–Lagrange Differential Equation |website=mathworld.wolfram.com |publisher=Wolfram |at=Eq.&nbsp;(5)}}</ref>



:<math>L-f'\frac{\partial L}{\partial f'}=C \, ,</math>

:<math>L-f^{\prime}\frac{\partial L}{\partial f^{\prime}}=C \, ,</math>



where <math>C</math> is a constant. The left hand side is the [[Legendre transformation]] of <math>L</math> with respect to <math>f'(x)</math>.

where <math>C</math> is a constant. The left hand side is the [[Legendre transformation]] of <math>L</math> with respect to <math>f^{\prime}(x).</math>



The intuition behind this result is that, if the variable {{math|''x''}} is actually time, then the statement <math>\frac{\partial L}{\partial x} = 0</math> implies that the Lagrangian is time-independent. By [[Noether's theorem]], there is an associated conserved quantity. In this case, this quantity is the Hamiltonian, the Legendre transform of the Lagrangian, which (often) coincides with the energy of the system. This is (minus) the constant in Beltrami's identity.

The intuition behind this result is that, if the variable <math>x</math> is actually time, then the statement <math>\frac{\partial L}{\partial x} = 0</math> implies that the Lagrangian is time-independent. By [[Noether's theorem]], there is an associated conserved quantity. In this case, this quantity is the Hamiltonian, the Legendre transform of the Lagrangian, which (often) coincides with the energy of the system. This is (minus) the constant in Beltrami's identity.



== Euler–Poisson equation ==

== Euler–Poisson equation ==

If <math>S</math> depends on higher-derivatives of <math>y(x)</math>, that is, if<blockquote><math>S = \int\limits_{a}^{b} f(x, y(x), y'(x), ..., y^{n}(x)) dx,</math></blockquote>then <math>y</math> must satisfy the Euler–[[Siméon Denis Poisson|Poisson]] equation,<blockquote><math>\frac{\partial f}{\partial y} - \frac{d}{dx} \left( \frac{\partial f}{\partial y'} \right) + ... + (-1)^{n} \frac{d^n}{dx^n} \left[ \frac{\partial f}{\partial y^{(n)}} \right]= 0.</math><ref>{{Cite book|last=Kot|first=Mark|title=A First Course in the Calculus of Variations|publisher=American Mathematical Society|year=2014|isbn=978-1-4704-1495-5|chapter=Chapter 4: Basic Generalizations}}</ref></blockquote>

If <math>S</math> depends on higher-derivatives of <math>y(x),</math> that is, if<blockquote><math>S = \int\limits_{a}^{b} f(x, y(x), y^{\prime}(x), ..., y^{n}(x)) dx,</math></blockquote>then <math>y</math> must satisfy the Euler–[[Siméon Denis Poisson|Poisson]] equation,<blockquote><math>\frac{\partial f}{\partial y} - \frac{d}{dx} \left( \frac{\partial f}{\partial y^{\prime}} \right) + ... + (-1)^{n} \frac{d^n}{dx^n} \left[ \frac{\partial f}{\partial y^{(n)}} \right]= 0.</math><ref>{{Cite book|last=Kot|first=Mark|title=A First Course in the Calculus of Variations|publisher=American Mathematical Society|year=2014|isbn=978-1-4704-1495-5|chapter=Chapter 4: Basic Generalizations}}</ref></blockquote>



== Du Bois-Reymond's theorem ==

== Du Bois-Reymond's theorem ==



The discussion thus far has assumed that extremal functions possess two continuous derivatives, although the existence of the integral ''J'' requires only first derivatives of trial functions. The condition that the first variation vanishes at an extremal may be regarded as a '''weak form''' of the Euler–Lagrange equation. The theorem of Du Bois-Reymond asserts that this weak form implies the strong form. If ''L'' has continuous first and second derivatives with respect to all of its arguments, and if

The discussion thus far has assumed that extremal functions possess two continuous derivatives, although the existence of the integral <math>J</math> requires only first derivatives of trial functions. The condition that the first variation vanishes at an extremal may be regarded as a '''weak form''' of the Euler–Lagrange equation. The theorem of Du Bois-Reymond asserts that this weak form implies the strong form. If <math>L</math> has continuous first and second derivatives with respect to all of its arguments, and if



:<math>\frac{\partial^2 L}{\partial f'^2} \ne 0, </math>

:<math>\frac{\partial^2 L}{\partial f'^2} \ne 0,</math>



then <math>f</math> has two continuous derivatives, and it satisfies the Euler–Lagrange equation.

then <math>f</math> has two continuous derivatives, and it satisfies the Euler–Lagrange equation.

Line 144: Line 144:

:<math>{A} = \{x \in W^{1,1}(0,1) : x(0)=0,\ x(1)=1\}.</math>

:<math>{A} = \{x \in W^{1,1}(0,1) : x(0)=0,\ x(1)=1\}.</math>



Clearly, <math>x(t) = t^{\frac{1}{3}} </math> minimizes the functional, but we find any function <math>x \in W^{1, \infty}</math> gives a value bounded away from the infimum.

Clearly, <math>x(t) = t^{\frac{1}{3}}</math>minimizes the functional, but we find any function <math>x \in W^{1, \infty}</math> gives a value bounded away from the infimum.



Examples (in one-dimension) are traditionally manifested across <math>W^{1,1}</math> and <math>W^{1,\infty}</math>, but Ball and Mizel<ref>{{Cite journal|last=Ball & Mizel|date=1985|title=One-dimensional Variational problems whose Minimizers do not satisfy the Euler-Lagrange equation.|journal=Archive for Rational Mechanics and Analysis|volume=90|issue=4|pages=325–388|doi=10.1007/BF00276295|bibcode=1985ArRMA..90..325B|s2cid=55005550}}</ref> procured the first functional that displayed Lavrentiev's Phenomenon across <math>W^{1,p}</math> and <math>W^{1,q}</math> for <math>1 \leq p < q < \infty. </math> There are several results that gives criteria under which the phenomenon does not occur - for instance 'standard growth', a Lagrangian with no dependence on the second variable, or an approximating sequence satisfying Cesari's Condition (D) - but results are often particular, and applicable to a small class of functionals.

Examples (in one-dimension) are traditionally manifested across <math>W^{1,1}</math> and <math>W^{1,\infty},</math> but Ball and Mizel<ref>{{Cite journal|last=Ball & Mizel|date=1985|title=One-dimensional Variational problems whose Minimizers do not satisfy the Euler-Lagrange equation.|journal=Archive for Rational Mechanics and Analysis|volume=90|issue=4|pages=325–388|doi=10.1007/BF00276295|bibcode=1985ArRMA..90..325B|s2cid=55005550}}</ref> procured the first functional that displayed Lavrentiev's Phenomenon across <math>W^{1,p}</math> and <math>W^{1,q}</math> for <math>1 \leq p < q < \infty.</math> There are several results that gives criteria under which the phenomenon does not occur - for instance 'standard growth', a Lagrangian with no dependence on the second variable, or an approximating sequence satisfying Cesari's Condition (D) - but results are often particular, and applicable to a small class of functionals.



Connected with the Lavrentiev Phenomenon is the repulsion property: any functional displaying Lavrentiev's Phenomenon will display the weak repulsion property.<ref>{{Cite journal|last=Ferriero|first=Alessandro|date=2007|title=The Weak Repulsion property|journal=Journal de Mathématiques Pures et Appliquées|volume=88|issue=4|pages=378–388|doi=10.1016/j.matpur.2007.06.002}}</ref>

Connected with the Lavrentiev Phenomenon is the repulsion property: any functional displaying Lavrentiev's Phenomenon will display the weak repulsion property.<ref>{{Cite journal|last=Ferriero|first=Alessandro|date=2007|title=The Weak Repulsion property|journal=Journal de Mathématiques Pures et Appliquées|volume=88|issue=4|pages=378–388|doi=10.1016/j.matpur.2007.06.002}}</ref>

Line 152: Line 152:

== Functions of several variables ==

== Functions of several variables ==



For example, if φ(''x'',''y'') denotes the displacement of a membrane above the domain ''D'' in the ''x'',''y'' plane, then its potential energy is proportional to its surface area:

For example, if <math>\varphi(x, y)</math> denotes the displacement of a membrane above the domain <math>D</math> in the <math>x,y</math> plane, then its potential energy is proportional to its surface area:

:<math> U[\varphi] = \iint_D \sqrt{1 +\nabla \varphi \cdot \nabla \varphi} \,dx\,dy.\,</math>

:<math>U[\varphi] = \iint_D \sqrt{1 +\nabla \varphi \cdot \nabla \varphi} \,dx\,dy.\,</math>

[[Plateau's problem]] consists of finding a function that minimizes the surface area while assuming prescribed values on the boundary of ''D''; the solutions are called '''minimal surfaces'''. The Euler–Lagrange equation for this problem is nonlinear:

[[Plateau's problem]] consists of finding a function that minimizes the surface area while assuming prescribed values on the boundary of <math>D</math>; the solutions are called '''minimal surfaces'''. The Euler–Lagrange equation for this problem is nonlinear:

:<math> \varphi_{xx}(1 + \varphi_y^2) + \varphi_{yy}(1 + \varphi_x^2) - 2\varphi_x \varphi_y \varphi_{xy} = 0.\,</math>

:<math>\varphi_{xx}(1 + \varphi_y^2) + \varphi_{yy}(1 + \varphi_x^2) - 2\varphi_x \varphi_y \varphi_{xy} = 0.\,</math>

See Courant (1950) for details.

See Courant (1950) for details.



=== Dirichlet's principle ===

=== Dirichlet's principle ===

It is often sufficient to consider only small displacements of the membrane, whose energy difference from no displacement is approximated by

It is often sufficient to consider only small displacements of the membrane, whose energy difference from no displacement is approximated by

:<math> V[\varphi] = \frac{1}{2}\iint_D \nabla \varphi \cdot \nabla \varphi \, dx\, dy.\,</math>

:<math>V[\varphi] = \frac{1}{2}\iint_D \nabla \varphi \cdot \nabla \varphi \, dx\, dy.\,</math>

The functional ''V'' is to be minimized among all trial functions φ that assume prescribed values on the boundary of ''D''. If ''u'' is the minimizing function and ''v'' is an arbitrary smooth function that vanishes on the boundary of ''D'', then the first variation of <math>V[u + \varepsilon v]</math> must vanish:

The functional <math>V</math> is to be minimized among all trial functions <math>\varphi</math> that assume prescribed values on the boundary of <math>D.</math>If<math>u</math> is the minimizing function and <math>v</math> is an arbitrary smooth function that vanishes on the boundary of <math>D,</math> then the first variation of <math>V[u + \varepsilon v]</math> must vanish:

:<math> \frac{d}{d\varepsilon} V[u + \varepsilon v]|_{\varepsilon=0} = \iint_D \nabla u \cdot \nabla v \, dx\,dy = 0.\,</math>

:<math>\frac{d}{d\varepsilon} V[u + \varepsilon v]|_{\varepsilon=0} = \iint_D \nabla u \cdot \nabla v \, dx\,dy = 0.\,</math>

Provided that u has two derivatives, we may apply the divergence theorem to obtain

Provided that u has two derivatives, we may apply the divergence theorem to obtain

:<math> \iint_D \nabla \cdot (v \nabla u) \,dx\,dy =

:<math>\iint_D \nabla \cdot (v \nabla u) \,dx\,dy =

\iint_D \nabla u \cdot \nabla v + v \nabla \cdot \nabla u \,dx\,dy = \int_C v \frac{\partial u}{\partial n} \, ds,</math>

\iint_D \nabla u \cdot \nabla v + v \nabla \cdot \nabla u \,dx\,dy = \int_C v \frac{\partial u}{\partial n} \, ds,</math>

where ''C'' is the boundary of ''D'', ''s'' is arclength along ''C'' and <math> \partial u / \partial n</math> is the normal derivative of ''u''on''C''. Since ''v'' vanishes on ''C'' and the first variation vanishes, the result is

where <math>C</math> is the boundary of <math>D,</math> <math>s</math> is arclength along <math>C</math> and <math>\partial u / \partial n</math> is the normal derivative of <math>u</math>on<math>C.</math> Since <math>v</math> vanishes on <math>C</math> and the first variation vanishes, the result is

:<math>\iint_D v\nabla \cdot \nabla u \,dx\,dy =0 \, </math>

:<math>\iint_D v\nabla \cdot \nabla u \,dx\,dy =0 \,</math>

for all smooth functions v that vanish on the boundary of ''D''. The proof for the case of one dimensional integrals may be adapted to this case to show that

for all smooth functions v that vanish on the boundary of <math>D.</math> The proof for the case of one dimensional integrals may be adapted to this case to show that

:<math> \nabla \cdot \nabla u= 0 \, </math> in''D''.

:<math>\nabla \cdot \nabla u= 0 \,</math>in<math>D.</math>



The difficulty with this reasoning is the assumption that the minimizing function u must have two derivatives. Riemann argued that the existence of a smooth minimizing function was assured by the connection with the physical problem: membranes do indeed assume configurations with minimal potential energy. Riemann named this idea the [[Dirichlet principle]] in honor of his teacher [[Peter Gustav Lejeune Dirichlet]]. However Weierstrass gave an example of a variational problem with no solution: minimize

The difficulty with this reasoning is the assumption that the minimizing function u must have two derivatives. Riemann argued that the existence of a smooth minimizing function was assured by the connection with the physical problem: membranes do indeed assume configurations with minimal potential energy. Riemann named this idea the [[Dirichlet principle]] in honor of his teacher [[Peter Gustav Lejeune Dirichlet]]. However Weierstrass gave an example of a variational problem with no solution: minimize

:<math> W[\varphi] = \int_{-1}^{1} (x\varphi')^2 \, dx\,</math>

:<math>W[\varphi] = \int_{-1}^{1} (x\varphi^{\prime})^2 \, dx\,</math>

among all functions ''φ'' that satisfy <math>\varphi(-1)=-1</math> and

among all functions <math>\varphi</math> that satisfy <math>\varphi(-1)=-1</math> and

<math>\varphi(1)=1.</math>

<math>\varphi(1)=1.</math>

<math>W</math> can be made arbitrarily small by choosing piecewise linear functions that make a transition between −1 and 1 in a small neighborhood of the origin. However, there is no function that makes <math>W=0</math>.{{efn|The resulting controversy over the validity of Dirichlet's principle is explained by Turnbull.<ref>{{cite web |url=http://turnbull.mcs.st-and.ac.uk/~history/Biographies/Riemann.html |title=Riemann biography |publisher=U. St. Andrew |place=UK |author=Turnbull}}</ref>}} Eventually it was shown that Dirichlet's principle is valid, but it requires a sophisticated application of the regularity theory for [[elliptic partial differential equation]]s; see Jost and Li–Jost (1998).

<math>W</math> can be made arbitrarily small by choosing piecewise linear functions that make a transition between −1 and 1 in a small neighborhood of the origin. However, there is no function that makes <math>W=0.</math>{{efn|The resulting controversy over the validity of Dirichlet's principle is explained by Turnbull.<ref>{{cite web |url=http://turnbull.mcs.st-and.ac.uk/~history/Biographies/Riemann.html |title=Riemann biography |publisher=U. St. Andrew |place=UK |author=Turnbull}}</ref>}} Eventually it was shown that Dirichlet's principle is valid, but it requires a sophisticated application of the regularity theory for [[elliptic partial differential equation]]s; see Jost and Li–Jost (1998).



=== Generalization to other boundary value problems ===

=== Generalization to other boundary value problems ===

A more general expression for the potential energy of a membrane is

A more general expression for the potential energy of a membrane is

:<math> V[\varphi] = \iint_D \left[ \frac{1}{2} \nabla \varphi \cdot \nabla \varphi + f(x,y) \varphi \right] \, dx\,dy \, + \int_C \left[ \frac{1}{2} \sigma(s) \varphi^2 + g(s) \varphi \right] \, ds.</math>

:<math>V[\varphi] = \iint_D \left[ \frac{1}{2} \nabla \varphi \cdot \nabla \varphi + f(x,y) \varphi \right] \, dx\,dy \, + \int_C \left[ \frac{1}{2} \sigma(s) \varphi^2 + g(s) \varphi \right] \, ds.</math>

This corresponds to an external force density <math>f(x,y)</math>in''D'', an external force <math>g(s)</math> on the boundary ''C'', and elastic forces with modulus <math>\sigma(s) </math> acting on ''C''. The function that minimizes the potential energy '''with no restriction on its boundary values''' will be denoted by ''u''. Provided that ''f'' and ''g'' are continuous, regularity theory implies that the minimizing function ''u'' will have two derivatives. In taking the first variation, no boundary condition need be imposed on the increment ''v''. The first variation of

This corresponds to an external force density <math>f(x,y)</math>in<math>D,</math> an external force <math>g(s)</math> on the boundary <math>C,</math> and elastic forces with modulus <math>\sigma(s)</math>acting on <math>C.</math> The function that minimizes the potential energy '''with no restriction on its boundary values''' will be denoted by <math>u.</math> Provided that <math>f</math> and <math>g</math> are continuous, regularity theory implies that the minimizing function <math>u</math> will have two derivatives. In taking the first variation, no boundary condition need be imposed on the increment <math>v.</math>The first variation of

<math>V[u + \varepsilon v]</math> is given by

<math>V[u + \varepsilon v]</math> is given by

:<math> \iint_D \left[ \nabla u \cdot \nabla v + f v \right] \, dx\, dy + \int_C \left[ \sigma u v + g v \right] \, ds =0. \,</math>

:<math>\iint_D \left[ \nabla u \cdot \nabla v + f v \right] \, dx\, dy + \int_C \left[ \sigma u v + g v \right] \, ds =0. \,</math>

If we apply the divergence theorem, the result is

If we apply the divergence theorem, the result is

:<math> \iint_D \left[ -v \nabla \cdot \nabla u + v f \right] \, dx \, dy + \int_C v \left[ \frac{\partial u}{\partial n} + \sigma u + g \right] \, ds =0. \,</math>

:<math>\iint_D \left[ -v \nabla \cdot \nabla u + v f \right] \, dx \, dy + \int_C v \left[ \frac{\partial u}{\partial n} + \sigma u + g \right] \, ds =0. \,</math>

If we first set ''v''&nbsp;=&nbsp;0 on ''C'', the boundary integral vanishes, and we conclude as before that

If we first set <math>v = 0</math>on<math>C,</math> the boundary integral vanishes, and we conclude as before that

:<math> - \nabla \cdot \nabla u + f =0 \,</math>

:<math>- \nabla \cdot \nabla u + f =0 \,</math>

in''D''. Then if we allow ''v'' to assume arbitrary boundary values, this implies that ''u'' must satisfy the boundary condition

in<math>D.</math> Then if we allow <math>v</math> to assume arbitrary boundary values, this implies that <math>u</math> must satisfy the boundary condition

:<math> \frac{\partial u}{\partial n} + \sigma u + g =0, \,</math>

:<math>\frac{\partial u}{\partial n} + \sigma u + g =0, \,</math>

on''C''. This boundary condition is a consequence of the minimizing property of ''u'': it is not imposed beforehand. Such conditions are called '''natural boundary conditions'''.

on<math>C.</math> This boundary condition is a consequence of the minimizing property of <math>u</math>: it is not imposed beforehand. Such conditions are called '''natural boundary conditions'''.



The preceding reasoning is not valid if <math> \sigma</math> vanishes identically on ''C''. In such a case, we could allow a trial function <math> \varphi \equiv c</math>, where ''c'' is a constant. For such a trial function,

The preceding reasoning is not valid if <math>\sigma</math> vanishes identically on <math>C.</math> In such a case, we could allow a trial function <math>\varphi \equiv c,</math> where <math>c</math> is a constant. For such a trial function,

:<math> V[c] = c\left[ \iint_D f \, dx\,dy + \int_C g \, ds \right].</math>

:<math>V[c] = c\left[ \iint_D f \, dx\,dy + \int_C g \, ds \right].</math>

By appropriate choice of ''c'', ''V'' can assume any value unless the quantity inside the brackets vanishes. Therefore, the variational problem is meaningless unless

By appropriate choice of <math>c,</math> <math>V</math> can assume any value unless the quantity inside the brackets vanishes. Therefore, the variational problem is meaningless unless

:<math> \iint_D f \, dx\,dy + \int_C g \, ds =0.\,</math>

:<math>\iint_D f \, dx\,dy + \int_C g \, ds =0.\,</math>

This condition implies that net external forces on the system are in equilibrium. If these forces are in equilibrium, then the variational problem has a solution, but it is not unique, since an arbitrary constant may be added. Further details and examples are in Courant and Hilbert (1953).

This condition implies that net external forces on the system are in equilibrium. If these forces are in equilibrium, then the variational problem has a solution, but it is not unique, since an arbitrary constant may be added. Further details and examples are in Courant and Hilbert (1953).



Line 204: Line 204:

{{See also|Sturm–Liouville theory}}

{{See also|Sturm–Liouville theory}}

The Sturm–Liouville eigenvalue problem involves a general quadratic form

The Sturm–Liouville eigenvalue problem involves a general quadratic form

:<math>Q[\varphi] = \int_{x_1}^{x_2} \left[ p(x) \varphi'(x)^2 + q(x) \varphi(x)^2 \right] \, dx, \,</math>

:<math>Q[\varphi] = \int_{x_1}^{x_2} \left[ p(x) \varphi^{\prime}(x)^2 + q(x) \varphi(x)^2 \right] \, dx, \,</math>

where <math> \varphi </math> is restricted to functions that satisfy the boundary conditions

where <math>\varphi</math>is restricted to functions that satisfy the boundary conditions

:<math>\varphi(x_1)=0, \quad \varphi(x_2)=0. \,</math>

:<math>\varphi(x_1)=0, \quad \varphi(x_2)=0. \,</math>

Let ''R'' be a normalization integral

Let <math>R</math> be a normalization integral

:<math> R[\varphi] =\int_{x_1}^{x_2} r(x)\varphi(x)^2 \, dx.\,</math>

:<math>R[\varphi] =\int_{x_1}^{x_2} r(x)\varphi(x)^2 \, dx.\,</math>

The functions <math>p(x)</math> and <math>r(x)</math> are required to be everywhere positive and bounded away from zero. The primary variational problem is to minimize the ratio ''Q''/''R'' among all φ satisfying the endpoint conditions. It is shown below that the Euler–Lagrange equation for the minimizing ''u''is

The functions <math>p(x)</math> and <math>r(x)</math> are required to be everywhere positive and bounded away from zero. The primary variational problem is to minimize the ratio <math>Q/R</math> among all <math>\varphi</math> satisfying the endpoint conditions. It is shown below that the Euler–Lagrange equation for the minimizing <math>u</math>is

:<math> -(pu')' +q u -\lambda r u =0, \,</math>

:<math>-(pu^{\prime})^{\prime} +q u -\lambda r u = 0, \,</math>

where λ is the quotient

where <math>\lambda</math> is the quotient

:<math> \lambda = \frac{Q[u]}{R[u]}. \,</math>

:<math>\lambda = \frac{Q[u]}{R[u]}. \,</math>

It can be shown (see Gelfand and Fomin 1963) that the minimizing ''u'' has two derivatives and satisfies the Euler–Lagrange equation. The associated λ will be denoted by <math>\lambda_1</math>; it is the lowest eigenvalue for this equation and boundary conditions. The associated minimizing function will be denoted by <math>u_1(x)</math>. This variational characterization of eigenvalues leads to the [[Rayleigh–Ritz method]]: choose an approximating ''u'' as a linear combination of basis functions (for example trigonometric functions) and carry out a finite-dimensional minimization among such linear combinations. This method is often surprisingly accurate.

It can be shown (see Gelfand and Fomin 1963) that the minimizing <math>u</math> has two derivatives and satisfies the Euler–Lagrange equation. The associated <math>\lambda</math> will be denoted by <math>\lambda_1</math>; it is the lowest eigenvalue for this equation and boundary conditions. The associated minimizing function will be denoted by <math>u_1(x).</math> This variational characterization of eigenvalues leads to the [[Rayleigh–Ritz method]]: choose an approximating <math>u</math> as a linear combination of basis functions (for example trigonometric functions) and carry out a finite-dimensional minimization among such linear combinations. This method is often surprisingly accurate.



The next smallest eigenvalue and eigenfunction can be obtained by minimizing ''Q'' under the additional constraint

The next smallest eigenvalue and eigenfunction can be obtained by minimizing <math>Q</math> under the additional constraint

:<math> \int_{x_1}^{x_2} r(x) u_1(x) \varphi(x) \, dx=0. \,</math>

:<math>\int_{x_1}^{x_2} r(x) u_1(x) \varphi(x) \, dx=0. \,</math>

This procedure can be extended to obtain the complete sequence of eigenvalues and eigenfunctions for the problem.

This procedure can be extended to obtain the complete sequence of eigenvalues and eigenfunctions for the problem.



The variational problem also applies to more general boundary conditions. Instead of requiring that φ vanish at the endpoints, we may not impose any condition at the endpoints, and set

The variational problem also applies to more general boundary conditions. Instead of requiring that <math>\varphi</math> vanish at the endpoints, we may not impose any condition at the endpoints, and set

:<math>Q[\varphi] = \int_{x_1}^{x_2} \left[ p(x) \varphi'(x)^2 + q(x)\varphi(x)^2 \right] \, dx + a_1 \varphi(x_1)^2 + a_2 \varphi(x_2)^2, \,</math>

:<math>Q[\varphi] = \int_{x_1}^{x_2} \left[ p(x) \varphi^{\prime}(x)^2 + q(x)\varphi(x)^2 \right] \, dx + a_1 \varphi(x_1)^2 + a_2 \varphi(x_2)^2, \,</math>

where <math>a_1</math> and <math>a_2</math> are arbitrary. If we set <math> \varphi = u + \varepsilon v </math> the first variation for the ratio <math> Q/R</math> is

where <math>a_1</math> and <math>a_2</math> are arbitrary. If we set <math>\varphi = u + \varepsilon v</math>the first variation for the ratio <math>Q/R</math>is

:<math> V_1 = \frac{2}{R[u]} \left( \int_{x_1}^{x_2} \left[ p(x) u'(x)v'(x) + q(x)u(x)v(x) -\lambda r(x) u(x) v(x) \right] \, dx + a_1 u(x_1)v(x_1) + a_2 u(x_2)v(x_2) \right) , \,</math>

:<math>V_1 = \frac{2}{R[u]} \left( \int_{x_1}^{x_2} \left[ p(x) u^{\prime}(x)v^{\prime}(x) + q(x)u(x)v(x) -\lambda r(x) u(x) v(x) \right] \, dx + a_1 u(x_1)v(x_1) + a_2 u(x_2)v(x_2) \right), \,</math>

where λ is given by the ratio <math> Q[u]/R[u]</math> as previously.

where λ is given by the ratio <math>Q[u]/R[u]</math> as previously.

After integration by parts,

After integration by parts,

:<math> \frac{R[u]}{2} V_1 = \int_{x_1}^{x_2} v(x) \left[ -(p u')' + q u -\lambda r u \right] \, dx + v(x_1)[ -p(x_1)u'(x_1) + a_1 u(x_1)] + v(x_2) [p(x_2) u'(x_2) + a_2 u(x_2)]. \,</math>

:<math>\frac{R[u]}{2} V_1 = \int_{x_1}^{x_2} v(x) \left[ -(p u^{\prime})^{\prime} + q u -\lambda r u \right] \, dx + v(x_1)[ -p(x_1)u^{\prime}(x_1) + a_1 u(x_1)] + v(x_2) [p(x_2) u^{\prime}(x_2) + a_2 u(x_2)]. \,</math>

If we first require that ''v'' vanish at the endpoints, the first variation will vanish for all such ''v'' only if

If we first require that <math>v</math> vanish at the endpoints, the first variation will vanish for all such <math>v</math> only if

:<math> -(p u')' + q u -\lambda r u =0 \quad \hbox{for} \quad x_1 < x < x_2.\,</math>

:<math>-(p u^{\prime})^{\prime} + q u -\lambda r u =0 \quad \hbox{for} \quad x_1 < x < x_2.\,</math>

If''u'' satisfies this condition, then the first variation will vanish for arbitrary ''v'' only if

If<math>u</math> satisfies this condition, then the first variation will vanish for arbitrary <math>v</math> only if

:<math> -p(x_1)u'(x_1) + a_1 u(x_1)=0, \quad \hbox{and} \quad p(x_2) u'(x_2) + a_2 u(x_2)=0.\,</math>

:<math>-p(x_1)u^{\prime}(x_1) + a_1 u(x_1)=0, \quad \hbox{and} \quad p(x_2) u^{\prime}(x_2) + a_2 u(x_2)=0.\,</math>

These latter conditions are the '''natural boundary conditions''' for this problem, since they are not imposed on trial functions for the minimization, but are instead a consequence of the minimization.

These latter conditions are the '''natural boundary conditions''' for this problem, since they are not imposed on trial functions for the minimization, but are instead a consequence of the minimization.



=== Eigenvalue problems in several dimensions ===

=== Eigenvalue problems in several dimensions ===

Eigenvalue problems in higher dimensions are defined in analogy with the one-dimensional case. For example, given a domain ''D'' with boundary ''B'' in three dimensions we may define

Eigenvalue problems in higher dimensions are defined in analogy with the one-dimensional case. For example, given a domain <math>D</math> with boundary <math>B</math> in three dimensions we may define



:<math> Q[\varphi] = \iiint_D p(X) \nabla \varphi \cdot \nabla \varphi + q(X) \varphi^2 \, dx \, dy \, dz + \iint_B \sigma(S) \varphi^2 \, dS, \,</math>

:<math>Q[\varphi] = \iiint_D p(X) \nabla \varphi \cdot \nabla \varphi + q(X) \varphi^2 \, dx \, dy \, dz + \iint_B \sigma(S) \varphi^2 \, dS, \,</math>

and

and

:<math> R[\varphi] = \iiint_D r(X) \varphi(X)^2 \, dx \, dy \, dz.\,</math>

:<math>R[\varphi] = \iiint_D r(X) \varphi(X)^2 \, dx \, dy \, dz.\,</math>

Let ''u'' be the function that minimizes the quotient <math> Q[\varphi] / R[\varphi], </math>

Let <math>u</math> be the function that minimizes the quotient <math>Q[\varphi] / R[\varphi],</math>

with no condition prescribed on the boundary ''B''. The Euler–Lagrange equation satisfied by ''u''is

with no condition prescribed on the boundary <math>B.</math> The Euler–Lagrange equation satisfied by <math>u</math>is

:<math> -\nabla \cdot (p(X) \nabla u) + q(x) u - \lambda r(x) u=0,\,</math>

:<math>-\nabla \cdot (p(X) \nabla u) + q(x) u - \lambda r(x) u=0,\,</math>

where

where

:<math> \lambda = \frac{Q[u]}{R[u]}.\,</math>

:<math>\lambda = \frac{Q[u]}{R[u]}.\,</math>

The minimizing ''u'' must also satisfy the natural boundary condition

The minimizing <math>u</math> must also satisfy the natural boundary condition

:<math> p(S) \frac{\partial u}{\partial n} + \sigma(S) u =0,</math>

:<math>p(S) \frac{\partial u}{\partial n} + \sigma(S) u = 0,</math>

on the boundary ''B''. This result depends upon the regularity theory for elliptic partial differential equations; see Jost and Li–Jost (1998) for details. Many extensions, including completeness results, asymptotic properties of the eigenvalues and results concerning the nodes of the eigenfunctions are in Courant and Hilbert (1953).

on the boundary <math>B.</math> This result depends upon the regularity theory for elliptic partial differential equations; see Jost and Li–Jost (1998) for details. Many extensions, including completeness results, asymptotic properties of the eigenvalues and results concerning the nodes of the eigenfunctions are in Courant and Hilbert (1953).



== Applications ==

== Applications ==

=== Optics ===

=== Optics ===

[[Fermat's principle]] states that light takes a path that (locally) minimizes the optical length between its endpoints. If the ''x''-coordinate is chosen as the parameter along the path, and <math>y=f(x)</math> along the path, then the optical length is given by

[[Fermat's principle]] states that light takes a path that (locally) minimizes the optical length between its endpoints. If the <math>x</math>-coordinate is chosen as the parameter along the path, and <math>y=f(x)</math> along the path, then the optical length is given by



:<math> A[f] = \int_{x=x_0}^{x_1} n(x,f(x)) \sqrt{1 + f'(x)^2} dx, \,</math>

:<math>A[f] = \int_{x=x_0}^{x_1} n(x,f(x)) \sqrt{1 + f^{\prime}(x)^2} dx, \,</math>



where the refractive index <math>n(x,y)</math> depends upon the material.

where the refractive index <math>n(x,y)</math> depends upon the material.

If we try <math> f(x) = f_0 (x) + \varepsilon f_1 (x)</math>

If we try <math>f(x) = f_0 (x) + \varepsilon f_1 (x)</math>

then the [[first variation]] of ''A'' (the derivative of ''A'' with respect to ε) is

then the [[first variation]] of <math>A</math> (the derivative of <math>A</math> with respect to ε) is



:<math> \delta A[f_0,f_1] = \int_{x=x_0}^{x_1} \left[ \frac{ n(x,f_0) f_0'(x) f_1'(x)}{\sqrt{1 + f_0'(x)^2}} + n_y (x,f_0) f_1 \sqrt{1 + f_0'(x)^2} \right] dx.</math>

:<math>\delta A[f_0,f_1] = \int_{x=x_0}^{x_1} \left[ \frac{ n(x,f_0) f_0^{\prime}(x) f_1^{\prime}(x)}{\sqrt{1 + f_0^{\prime}(x)^2}} + n_y (x,f_0) f_1 \sqrt{1 + f_0^{\prime}(x)^2} \right] dx.</math>



After integration by parts of the first term within brackets, we obtain the Euler–Lagrange equation

After integration by parts of the first term within brackets, we obtain the Euler–Lagrange equation



:<math> -\frac{d}{dx} \left[\frac{ n(x,f_0) f_0'}{\sqrt{1 + f_0'^2}} \right] + n_y (x,f_0) \sqrt{1 + f_0'(x)^2} =0. \,</math>

:<math>-\frac{d}{dx} \left[\frac{ n(x,f_0) f_0^{\prime}}{\sqrt{1 + f_0'^2}} \right] + n_y (x,f_0) \sqrt{1 + f_0^{\prime}(x)^2} =0. \,</math>



The light rays may be determined by integrating this equation. This formalism is used in the context of [[Lagrangian optics]] and [[Hamiltonian optics]].

The light rays may be determined by integrating this equation. This formalism is used in the context of [[Lagrangian optics]] and [[Hamiltonian optics]].

Line 268: Line 268:

There is a discontinuity of the refractive index when light enters or leaves a lens. Let

There is a discontinuity of the refractive index when light enters or leaves a lens. Let



:<math> n(x,y) = n_{(-)} \quad \hbox{if} \quad x<0, \,</math>

:<math>n(x,y) = n_{(-)} \quad \hbox{if} \quad x<0, \,</math>

:<math> n(x,y) = n_{(+)} \quad \hbox{if} \quad x>0,\,</math>

:<math>n(x,y) = n_{(+)} \quad \hbox{if} \quad x>0,\,</math>



where <math>n_{(-)}</math> and <math>n_{(+)}</math> are constants. Then the Euler–Lagrange equation holds as before in the region where ''x''<0 or ''x''>0, and in fact the path is a straight line there, since the refractive index is constant. At the ''x''=0, ''f'' must be continuous, but ''f' '' may be discontinuous. After integration by parts in the separate regions and using the Euler–Lagrange equations, the first variation takes the form

where <math>n_{(-)}</math> and <math>n_{(+)}</math> are constants. Then the Euler–Lagrange equation holds as before in the region where <math>x < 0</math>or<math>x > 0,</math> and in fact the path is a straight line there, since the refractive index is constant. At the <math>x = 0,</math> <math>f</math> must be continuous, but <math>f^{\prime}</math> may be discontinuous. After integration by parts in the separate regions and using the Euler–Lagrange equations, the first variation takes the form



:<math> \delta A[f_0,f_1] = f_1(0)\left[ n_{(-)}\frac{f_0'(0^-)}{\sqrt{1 + f_0'(0^-)^2}} - n_{(+)}\frac{f_0'(0^+)}{\sqrt{1 + f_0'(0^+)^2}} \right].\,</math>

:<math>\delta A[f_0,f_1] = f_1(0)\left[ n_{(-)}\frac{f_0^{\prime}(0^-)}{\sqrt{1 + f_0^{\prime}(0^-)^2}} - n_{(+)}\frac{f_0'(0^+)}{\sqrt{1 + f_0^{\prime}(0^+)^2}} \right].\,</math>



The factor multiplying <math>n_{(-)}</math> is the sine of angle of the incident ray with the ''x'' axis, and the factor multiplying <math>n_{(+)}</math> is the sine of angle of the refracted ray with the ''x'' axis. [[Snell's law]] for refraction requires that these terms be equal. As this calculation demonstrates, Snell's law is equivalent to vanishing of the first variation of the optical path length.

The factor multiplying <math>n_{(-)}</math> is the sine of angle of the incident ray with the <math>x</math> axis, and the factor multiplying <math>n_{(+)}</math> is the sine of angle of the refracted ray with the <math>x</math> axis. [[Snell's law]] for refraction requires that these terms be equal. As this calculation demonstrates, Snell's law is equivalent to vanishing of the first variation of the optical path length.



==== Fermat's principle in three dimensions ====

==== Fermat's principle in three dimensions ====

It is expedient to use vector notation: let <math>X=(x_1,x_2,x_3),</math> let ''t'' be a parameter, let <math>X(t)</math> be the parametric representation of a curve ''C'', and let <math>\dot X(t)</math> be its tangent vector. The optical length of the curve is given by

It is expedient to use vector notation: let <math>X=(x_1,x_2,x_3),</math> let <math>t</math> be a parameter, let <math>X(t)</math> be the parametric representation of a curve <math>C,</math> and let <math>\dot X(t)</math> be its tangent vector. The optical length of the curve is given by



:<math> A[C] = \int_{t=t_0}^{t_1} n(X) \sqrt{ \dot X \cdot \dot X} \, dt. \,</math>

:<math>A[C] = \int_{t=t_0}^{t_1} n(X) \sqrt{ \dot X \cdot \dot X} \, dt. \,</math>



Note that this integral is invariant with respect to changes in the parametric representation of ''C''. The Euler–Lagrange equations for a minimizing curve have the symmetric form

Note that this integral is invariant with respect to changes in the parametric representation of <math>C.</math> The Euler–Lagrange equations for a minimizing curve have the symmetric form



:<math> \frac{d}{dt} P = \sqrt{ \dot X \cdot \dot X} \, \nabla n, \,</math>

:<math>\frac{d}{dt} P = \sqrt{ \dot X \cdot \dot X} \, \nabla n, \,</math>



where

where

:<math> P = \frac{n(X) \dot X}{\sqrt{\dot X \cdot \dot X} }.\,</math>

:<math>P = \frac{n(X) \dot X}{\sqrt{\dot X \cdot \dot X} }.\,</math>



It follows from the definition that ''P'' satisfies

It follows from the definition that <math>P</math> satisfies



:<math> P \cdot P = n(X)^2. \,</math>

:<math>P \cdot P = n(X)^2. \,</math>



Therefore, the integral may also be written as

Therefore, the integral may also be written as



:<math> A[C] = \int_{t=t_0}^{t_1} P \cdot \dot X \, dt.\,</math>

:<math>A[C] = \int_{t=t_0}^{t_1} P \cdot \dot X \, dt.\,</math>



This form suggests that if we can find a function ψ whose gradient is given by ''P'', then the integral ''A'' is given by the difference of ψ at the endpoints of the interval of integration. Thus the problem of studying the curves that make the integral stationary can be related to the study of the level surfaces of ψ. In order to find such a function, we turn to the wave equation, which governs the propagation of light. This formalism is used in the context of [[Lagrangian optics]] and [[Hamiltonian optics]].

This form suggests that if we can find a function <math\psi</math> whose gradient is given by <math>P,</math> then the integral <math>A</math> is given by the difference of <math\psi</math> at the endpoints of the interval of integration. Thus the problem of studying the curves that make the integral stationary can be related to the study of the level surfaces of <math\psi.</math> In order to find such a function, we turn to the wave equation, which governs the propagation of light. This formalism is used in the context of [[Lagrangian optics]] and [[Hamiltonian optics]].



===== Connection with the wave equation =====

===== Connection with the wave equation =====

The [[wave equation]] for an inhomogeneous medium is

The [[wave equation]] for an inhomogeneous medium is



:<math> u_{tt} = c^2 \nabla \cdot \nabla u, \,</math>

:<math>u_{tt} = c^2 \nabla \cdot \nabla u, \,</math>



where ''c'' is the velocity, which generally depends upon ''X''. Wave fronts for light are characteristic surfaces for this partial differential equation: they satisfy

where <math>c</math> is the velocity, which generally depends upon <math>X.</math> Wave fronts for light are characteristic surfaces for this partial differential equation: they satisfy



:<math> \varphi_t^2 = c(X)^2 \, \nabla \varphi \cdot \nabla \varphi. \,</math>

:<math>\varphi_t^2 = c(X)^2 \, \nabla \varphi \cdot \nabla \varphi. \,</math>



We may look for solutions in the form

We may look for solutions in the form



:<math> \varphi(t,X) = t - \psi(X). \,</math>

:<math>\varphi(t,X) = t - \psi(X). \,</math>



In that case, ψ satisfies

In that case, <math\psi</math> satisfies



:<math> \nabla \psi \cdot \nabla \psi = n^2, \,</math>

:<math>\nabla \psi \cdot \nabla \psi = n^2, \,</math>



where <math>n=1/c.</math> According to the theory of [[first-order partial differential equation]]s, if <math>P = \nabla \psi, </math> then ''P'' satisfies

where <math>n=1/c.</math> According to the theory of [[first-order partial differential equation]]s, if <math>P = \nabla \psi,</math> then <math>P</math> satisfies



:<math> \frac{dP}{ds} = n \, \nabla n,</math>

:<math>\frac{dP}{ds} = n \, \nabla n,</math>



along a system of curves ('''the light rays''') that are given by

along a system of curves ('''the light rays''') that are given by



:<math> \frac{dX}{ds} = P. \,</math>

:<math>\frac{dX}{ds} = P. \,</math>



These equations for solution of a first-order partial differential equation are identical to the Euler–Lagrange equations if we make the identification

These equations for solution of a first-order partial differential equation are identical to the Euler–Lagrange equations if we make the identification



:<math> \frac{ds}{dt} = \frac{\sqrt{ \dot X \cdot \dot X} }{n}. \,</math>

:<math>\frac{ds}{dt} = \frac{\sqrt{ \dot X \cdot \dot X} }{n}. \,</math>



We conclude that the function ψ is the value of the minimizing integral ''A'' as a function of the upper end point. That is, when a family of minimizing curves is constructed, the values of the optical length satisfy the characteristic equation corresponding the wave equation. Hence, solving the associated partial differential equation of first order is equivalent to finding families of solutions of the variational problem. This is the essential content of the [[Hamilton–Jacobi theory]], which applies to more general variational problems.

We conclude that the function <math\psi</math> is the value of the minimizing integral <math>A</math> as a function of the upper end point. That is, when a family of minimizing curves is constructed, the values of the optical length satisfy the characteristic equation corresponding the wave equation. Hence, solving the associated partial differential equation of first order is equivalent to finding families of solutions of the variational problem. This is the essential content of the [[Hamilton–Jacobi theory]], which applies to more general variational problems.



=== Mechanics ===

=== Mechanics ===

{{main|Action (physics)}}

{{main|Action (physics)}}

In classical mechanics, the action, ''S'', is defined as the time integral of the Lagrangian, ''L''. The Lagrangian is the difference of energies,

In classical mechanics, the action, <math>S,</math> is defined as the time integral of the Lagrangian, <math>L.</math> The Lagrangian is the difference of energies,

:<math> L = T - U, \,</math>

:<math>L = T - U, \,</math>

where ''T'' is the [[kinetic energy]] of a mechanical system and ''U'' its [[potential energy]]. [[Hamilton's principle]] (or the action principle) states that the motion of a conservative holonomic (integrable constraints) mechanical system is such that the action integral

where <math>T</math> is the [[kinetic energy]] of a mechanical system and <math>U</math> its [[potential energy]]. [[Hamilton's principle]] (or the action principle) states that the motion of a conservative holonomic (integrable constraints) mechanical system is such that the action integral

:<math> S = \int_{t=t_0}^{t_1} L(x, \dot x, t) \, dt </math>

:<math>S = \int_{t=t_0}^{t_1} L(x, \dot x, t) \, dt</math>

is stationary with respect to variations in the path ''x''(''t'').

is stationary with respect to variations in the path <math>x(t).</math>

The Euler–Lagrange equations for this system are known as Lagrange's equations:

The Euler–Lagrange equations for this system are known as Lagrange's equations:

:<math> \frac{d}{dt} \frac{\partial L}{\partial \dot x} = \frac{\partial L}{\partial x}, \,</math>

:<math>\frac{d}{dt} \frac{\partial L}{\partial \dot x} = \frac{\partial L}{\partial x}, \,</math>

and they are equivalent to Newton's equations of motion (for such systems).

and they are equivalent to Newton's equations of motion (for such systems).



The conjugate momenta ''P'' are defined by

The conjugate momenta <math>P</math> are defined by

:<math> p = \frac{\partial L}{\partial \dot x}. \,</math>

:<math>p = \frac{\partial L}{\partial \dot x}. \,</math>

For example, if

For example, if

:<math> T = \frac{1}{2} m \dot x^2, \,</math>

:<math>T = \frac{1}{2} m \dot x^2, \,</math>

then

then

:<math> p = m \dot x. \,</math>

:<math>p = m \dot x. \,</math>

[[Hamiltonian mechanics]] results if the conjugate momenta are introduced in place of <math>\dot x</math> by a Legendre transformation of the Lagrangian ''L'' into the Hamiltonian ''H'' defined by

[[Hamiltonian mechanics]] results if the conjugate momenta are introduced in place of <math>\dot x</math> by a Legendre transformation of the Lagrangian <math>L</math> into the Hamiltonian <math>H</math> defined by

:<math> H(x, p, t) = p \,\dot x - L(x,\dot x, t).\,</math>

:<math>H(x, p, t) = p \,\dot x - L(x,\dot x, t).\,</math>

The Hamiltonian is the total energy of the system: ''H'' = ''T'' + ''U''.

The Hamiltonian is the total energy of the system: <math>H = T + U.</math>

Analogy with Fermat's principle suggests that solutions of Lagrange's equations (the particle trajectories) may be described in terms of level surfaces of some function of ''X''. This function is a solution of the [[Hamilton–Jacobi equation]]:

Analogy with Fermat's principle suggests that solutions of Lagrange's equations (the particle trajectories) may be described in terms of level surfaces of some function of <math>X.</math> This function is a solution of the [[Hamilton–Jacobi equation]]:

:<math> \frac{\partial \psi}{\partial t} + H\left(x,\frac{\partial \psi}{\partial x},t\right) =0.\,</math>

:<math>\frac{\partial \psi}{\partial t} + H\left(x,\frac{\partial \psi}{\partial x},t\right) = 0.\,</math>



=== Further applications ===

=== Further applications ===

Line 370: Line 370:

Calculus of variations is concerned with variations of functionals, which are small changes in the functional's value due to small changes in the function that is its argument. The '''first variation'''{{efn|name=AltFirst| The first variation is also called the variation, differential, or first differential.}} is defined as the linear part of the change in the functional, and the '''second variation'''{{efn|name=AltSecond| The second variation is also called the second differential.}} is defined as the quadratic part.<ref name='GelfandFominP11–12,99'>{{harvnb|Gelfand|Fomin|2000|pp=11–12, 99}}</ref>

Calculus of variations is concerned with variations of functionals, which are small changes in the functional's value due to small changes in the function that is its argument. The '''first variation'''{{efn|name=AltFirst| The first variation is also called the variation, differential, or first differential.}} is defined as the linear part of the change in the functional, and the '''second variation'''{{efn|name=AltSecond| The second variation is also called the second differential.}} is defined as the quadratic part.<ref name='GelfandFominP11–12,99'>{{harvnb|Gelfand|Fomin|2000|pp=11–12, 99}}</ref>



For example, if {{math|''J''[''y'']}} is a functional with the function {{math|''y'' {{=}} ''y''(''x'')}} as its argument, and there is a small change in its argument from {{math|''y''}}to{{math|''y'' + ''h''}}, where {{math|''h'' {{=}} ''h''(''x'')}} is a function in the same function space as {{math|''y''}}, then the corresponding change in the functional is

For example, if <math>J[y]</math> is a functional with the function <math>y = y(x)</math> as its argument, and there is a small change in its argument from <math>y</math>to<math>y + h,</math> where <math>h = h(x)</math> is a function in the same function space as <math>y,</math> then the corresponding change in the functional is



:<math> \Delta J[h] = J[y+h] - J[y]. </math>&nbsp;&nbsp;{{efn|name=SimplifyNotation|Note that {{math ''J''[''h'']}} and the variations below, depend on both {{math|''y''}} and {{math|''h''}}. The argument {{math|''y''}} has been left out to simplify the notation. For example, {{math ''J''[''h'']}} could have been written {{math ''J''[''y'' ; ''h'']}}.<ref name='GelfandFominP12FN6'>{{harvnb | Gelfand|Fomin|2000 | p=12, footnote 6}}</ref>}}

:<math>\Delta J[h] = J[y+h] - J[y].</math>&nbsp;&nbsp;{{efn|name=SimplifyNotation|Note that <math>\Delta J[h]</math> and the variations below, depend on both <math>y</math> and <math>h.</math> The argument <math>y</math> has been left out to simplify the notation. For example, <math>\Delta J[h]</math> could have been written <math>\Delta J[y; h].</math><ref name='GelfandFominP12FN6'>{{harvnb | Gelfand|Fomin|2000 | p=12, footnote 6}}</ref>}}



The functional {{math|''J''[''y'']}} is said to be '''differentiable''' if

The functional <math>J[y]</math> is said to be '''differentiable''' if



:<math> \Delta J[h] = \varphi [h] + \varepsilon \|h\|, </math>

:<math>\Delta J[h] = \varphi [h] + \varepsilon \|h\|,</math>



where {{math|''φ''[''h'']}} is a linear functional,{{efn|name=Linear| A functional {{math|''φ''[''h'']}} is said to be '''linear''' if {{math|''φ''[''αh''] {{=}} ''α φ''[''h'']}} &nbsp; and &nbsp; {{math|''φ''[''h''<sub>1</sub> +''h''<sub>2</sub>] {{=}} ''φ''[''h''<sub>1</sub>] + ''φ''[''h''<sub>2</sub>]}} , where {{math|''h'', ''h''<sub>1</sub>, ''h''<sub>2</sub>}} are functions and {{math|''α''}} is a real number.<ref name='GelfandFominP8'>{{harvnb | Gelfand|Fomin| 2000 | p=8 }}</ref>}} {{math|''<nowiki>||</nowiki>h<nowiki>||</nowiki>''}} is the norm of {{math|''h''}},{{efn|name=Norm| For a function {{math|''h'' {{=}} ''h''(''x'')}} that is defined for {{math|''a'' ''x'' ''b''}}, where {{math|''a''}} and {{math|''b''}} are real numbers, the norm of {{math|''h''}} is its maximum absolute value, i.e. {{math|''<nowiki>||</nowiki>h<nowiki>||</nowiki>''}} {{=}} max {{math|<nowiki>|</nowiki>''h''(''x'')<nowiki>|</nowiki>}} for {{math|''a'' ''x'' ''b''}}.<ref name='GelfandFominP6'>{{harvnb | Gelfand|Fomin| 2000 | p=6 }}</ref>}} and {{math|''ε 0''}} as{{math|''<nowiki>||</nowiki>h<nowiki>||</nowiki> 0''}}. The linear functional {{math|''φ''[''h'']}} is the first variation of {{math|''J''[''y'']}} and is denoted by,<ref name='GelfandFominP11–12'>{{harvnb | Gelfand|Fomin| 2000 | pp=11–12}}</ref>

where <math>\varphi[h]</math> is a linear functional,{{efn|name=Linear|A functional <math>\varphi[h]</math> is said to be '''linear''' if <math>\varphi[\alpha h] = \alpha \varphi[h]</math> &nbsp; and &nbsp; <math>\varphi\left[h + h_2\right] = \varphi[h] + \varphi\left[h_2\right],</math> where <math>h, h_2</math> are functions and <math>\alpha</math> is a real number.<ref name='GelfandFominP8'>{{harvnb | Gelfand|Fomin| 2000 | p=8 }}</ref>}} <math>\|h\|</math> is the norm of <math>h,</math>{{efn|name=Norm| For a function <math>h = h(x)</math> that is defined for <math>a\leqx\leq b,</math> where <math>a</math> and <math>b</math> are real numbers, the norm of <math>h</math> is its maximum absolute value, i.e. <math>\|h\| = \displaystyle\max_{a \leqx\leq b} |h(x)|.</math><ref name='GelfandFominP6'>{{harvnb | Gelfand|Fomin| 2000 | p=6 }}</ref>}} and <math>\varepsilon \to 0</math>as<math>\|h\| \to 0.</math> The linear functional <math>\varphi[h]</math> is the first variation of <math>J[y]</math> and is denoted by,<ref name='GelfandFominP11–12'>{{harvnb | Gelfand|Fomin| 2000 | pp=11–12}}</ref>



:<math>\delta J[h] = \varphi[h]. </math>

:<math>\delta J[h] = \varphi[h].</math>



The functional {{math|''J''[''y'']}} is said to be '''twice differentiable''' if

The functional <math>J[y]</math> is said to be '''twice differentiable''' if



:<math> \Delta J[h] = \varphi_1 [h] + \varphi_2 [h] + \varepsilon \|h\|^2, </math>

:<math>\Delta J[h] = \varphi_1 [h] + \varphi_2 [h] + \varepsilon \|h\|^2,</math>



where {{math|''φ''<sub>1</sub>[''h'']}} is a linear functional (the first variation), {{math|''φ''<sub>2</sub>[''h'']}} is a quadratic functional,{{efn|name=Quadratic| A functional is said to be '''quadratic''' if it is a bilinear functional with two argument functions that are equal. A '''bilinear functional''' is a functional that depends on two argument functions and is linear when each argument function in turn is fixed while the other argument function is variable.<ref name='GelfandFominP97–98'>{{harvnb | Gelfand|Fomin| 2000 | pp=97–98 }}</ref>}} and {{math|''ε 0''}} as{{math|''<nowiki>||</nowiki>h<nowiki>||</nowiki> 0''}}. The quadratic functional {{math|''φ''<sub>2</sub>[''h'']}} is the second variation of {{math|''J''[''y'']}} and is denoted by,<ref name='GelfandFominP99'>{{harvnb | Gelfand|Fomin| 2000 | p=99 }}</ref>

where <math>\varphi_1[h]</math> is a linear functional (the first variation), <math>\varphi_2[h]</math> is a quadratic functional,{{efn|name=Quadratic| A functional is said to be '''quadratic''' if it is a bilinear functional with two argument functions that are equal. A '''bilinear functional''' is a functional that depends on two argument functions and is linear when each argument function in turn is fixed while the other argument function is variable.<ref name='GelfandFominP97–98'>{{harvnb | Gelfand|Fomin| 2000 | pp=97–98 }}</ref>}} and <math>\varepsilon \to 0</math>as<math>\|h\| \to 0.</math> The quadratic functional <math>\varphi_2[h]</math> is the second variation of <math>J[y]</math> and is denoted by,<ref name='GelfandFominP99'>{{harvnb | Gelfand|Fomin| 2000 | p=99 }}</ref>



:<math>\delta^2 J[h] = \varphi_2[h]. </math>

:<math>\delta^2 J[h] = \varphi_2[h].</math>



The second variation {{math|''δ''<sup>2</sup>''J''[''h'']}} is said to be '''strongly positive''' if

The second variation <math>\delta^2 J[h]</math> is said to be '''strongly positive''' if

:<math>\delta^2J[h] \ge k \|h\|^2, </math>

:<math>\delta^2J[h] \ge k \|h\|^2,</math>

for all {{math|''h''}} and for some constant {{math|''k'' > ''0''}} .<ref name='GelfandFominP100'>{{harvnb | Gelfand|Fomin| 2000 | p=100 }}</ref>

for all <math>h</math> and for some constant <math>k > 0</math> .<ref name='GelfandFominP100'>{{harvnb | Gelfand|Fomin| 2000 | p=100 }}</ref>



Using the above definitions, especially the definitions of first variation, second variation, and strongly positive, the following sufficient condition for a minimum of a functional can be stated.

Using the above definitions, especially the definitions of first variation, second variation, and strongly positive, the following sufficient condition for a minimum of a functional can be stated.

Line 398: Line 398:

{{quote box|align=left |fontsize=100% |border=2px |quote='''Sufficient condition for a minimum:'''

{{quote box|align=left |fontsize=100% |border=2px |quote='''Sufficient condition for a minimum:'''



:The functional {{math|''J''[''y'']}} has a minimum at {{math|''y'' {{=}} ''ŷ''}} if its first variation {{math|''δJ''[''h''] {{=}} 0}}at{{math|''y'' {{=}} ''ŷ''}} and its second variation {{math|''δ''<sup>2</sup>''J''[''h'']}} is strongly positive at {{math|''y'' {{=}} ''ŷ''}} .<ref name='GelfandFominP100Theorem2'>{{harvnb | Gelfand|Fomin| 2000 | p=100 |loc=Theorem 2}}</ref> {{efn|name=sufficient| For other sufficient conditions, see in {{harvnb|Gelfand|Fomin|2000}},

:The functional <math>J[y]</math> has a minimum at <math>y = \hat{y}</math> if its first variation <math>\delta J[h] = 0</math>at<math>y = \hat{y}</math> and its second variation <math>\delta^2 J[h]</math> is strongly positive at <math>y = \hat{y}.</math><ref name='GelfandFominP100Theorem2'>{{harvnb | Gelfand|Fomin| 2000 | p=100 |loc=Theorem 2}}</ref> {{efn|name=sufficient| For other sufficient conditions, see in {{harvnb|Gelfand|Fomin|2000}},

* '''Chapter{{nbsp}}5: "The Second Variation. Sufficient Conditions for a Weak Extremum" – ''' Sufficient conditions for a weak minimum are given by the theorem on p.{{nbsp}}116.

* '''Chapter{{nbsp}}5: "The Second Variation. Sufficient Conditions for a Weak Extremum" – ''' Sufficient conditions for a weak minimum are given by the theorem on p.{{nbsp}}116.

* '''Chapter{{nbsp}}6: "Fields. Sufficient Conditions for a Strong Extremum" – ''' Sufficient conditions for a strong minimum are given by the theorem on p.{{nbsp}}148.}}{{efn|name=FuncMin| One may note the similarity to the sufficient condition for a minimum of a function, where the first derivative is zero and the second derivative is positive.}} }}

* '''Chapter{{nbsp}}6: "Fields. Sufficient Conditions for a Strong Extremum" – ''' Sufficient conditions for a strong minimum are given by the theorem on p.{{nbsp}}148.}}{{efn|name=FuncMin| One may note the similarity to the sufficient condition for a minimum of a function, where the first derivative is zero and the second derivative is positive.}} }}


Revision as of 21:22, 14 May 2021

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.[a] 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.

A simple example of such a problem is to find the curve of shortest length connecting two points. If there are no constraints, the solution is a straight line between the points. However, if the curve is constrained to lie on a surface in space, then the solution is less obvious, and possibly many solutions may exist. Such solutions are known as geodesics. A related problem is posed by Fermat's principle: light follows the path of shortest optical length connecting two points, where the optical length depends upon the material of the medium. One corresponding concept in mechanics is the principle of least/stationary action.

Many important problems involve functions of several variables. Solutions of boundary value problems for the Laplace equation satisfy the Dirichlet principle. Plateau's problem requires finding a surface of minimal area that spans a given contour in space: a solution can often be found by dipping a frame in a solution of soap suds. Although such experiments are relatively easy to perform, their mathematical interpretation is far from simple: there may be more than one locally minimizing surface, and they may have non-trivial topology.

History

The calculus of variations may be said to begin with Newton's minimal resistance problem in 1687, followed by the brachistochrone curve problem raised by Johann Bernoulli (1696).[2] It immediately occupied the attention of Jakob Bernoulli and the Marquis de l'Hôpital, but Leonhard Euler first elaborated the subject, beginning in 1733. Lagrange was influenced by Euler's work to contribute significantly to the theory. After Euler saw the 1755 work of the 19-year-old Lagrange, Euler dropped his own partly geometric approach in favor of Lagrange's purely analytic approach and renamed the subject the calculus of variations in his 1756 lecture Elementa Calculi Variationum.[3][4][1]

Legendre (1786) laid down a method, not entirely satisfactory, for the discrimination of maxima and minima. Isaac Newton and Gottfried Leibniz also gave some early attention to the subject.[5] To this discrimination Vincenzo Brunacci (1810), Carl Friedrich Gauss (1829), Siméon Poisson (1831), Mikhail Ostrogradsky (1834), and Carl Jacobi (1837) have been among the contributors. An important general work is that of Sarrus (1842) which was condensed and improved by Cauchy (1844). Other valuable treatises and memoirs have been written by Strauch (1849), Jellett (1850), Otto Hesse (1857), Alfred Clebsch (1858), and Carll (1885), but perhaps the most important work of the century is that of Weierstrass. His celebrated course on the theory is epoch-making, and it may be asserted that he was the first to place it on a firm and unquestionable foundation. The 20th and the 23rd Hilbert problem published in 1900 encouraged further development.[5]

In the 20th century David Hilbert, Emmy Noether, Leonida Tonelli, Henri Lebesgue and Jacques Hadamard among others made significant contributions.[5] Marston Morse applied calculus of variations in what is now called Morse theory.[6] Lev Pontryagin, Ralph Rockafellar and F. H. Clarke developed new mathematical tools for the calculus of variations in optimal control theory.[6] The dynamic programmingofRichard Bellman is an alternative to the calculus of variations.[7][8][9][b]

Extrema

The calculus of variations is concerned with the maxima or minima (collectively called extrema) of functionals. A functional maps functionstoscalars, so functionals have been described as "functions of functions." Functionals have extrema with respect to the elements of a given function space defined over a given domain. A functional is said to have an extremum at the function if has the same sign for all in an arbitrarily small neighborhood of [c] The function is called an extremal function or extremal.[d] The extremum is called a local maximum if everywhere in an arbitrarily small neighborhood of and a local minimum if there. For a function space of continuous functions, extrema of corresponding functionals are called weak extremaorstrong extrema, depending on whether the first derivatives of the continuous functions are respectively all continuous or not.[11]

Both strong and weak extrema of functionals are for a space of continuous functions but strong extrema have the additional requirement that the first derivatives of the functions in the space be continuous. Thus a strong extremum is also a weak extremum, but the converse may not hold. Finding strong extrema is more difficult than finding weak extrema.[12] An example of a necessary condition that is used for finding weak extrema is the Euler–Lagrange equation.[13][e]

Euler–Lagrange equation

Finding the extrema of functionals is similar to finding the maxima and minima of functions. The maxima and minima of a function may be located by finding the points where its derivative vanishes (i.e., is equal to zero). The extrema of functionals may be obtained by finding functions where the functional derivative is equal to zero. This leads to solving the associated Euler–Lagrange equation.[f]

Consider the functional

where

are constants,
is twice continuously differentiable,
is twice continuously differentiable with respect to its arguments and

If the functional attains a local minimumat and is an arbitrary function that has at least one derivative and vanishes at the endpoints and then for any number close to 0,

The term is called the variation of the function and is denoted by [1][g]

Substituting for in the functional the result is a function of

Since the functional has a minimum for the function has a minimum at and thus,[h]

Taking the total derivativeof where and are considered as functions of rather than yields

and because    and 

Therefore,

where when and we have used integration by parts on the second term. The second term on the second line vanishes because at and by definition. Also, as previously mentioned the left side of the equation is zero so that

According to the fundamental lemma of calculus of variations, the part of the integrand in parentheses is zero, i.e.

which is called the Euler–Lagrange equation. The left hand side of this equation is called the functional derivativeof and is denoted

In general this gives a second-order ordinary differential equation which can be solved to obtain the extremal function The Euler–Lagrange equation is a necessary, but not sufficient, condition for an extremum A sufficient condition for a minimum is given in the section Variations and sufficient condition for a minimum.

Example

In order to illustrate this process, consider the problem of finding the extremal function which is the shortest curve that connects two points and The arc length of the curve is given by

with

[i]

The Euler–Lagrange equation will now be used to find the extremal function that minimizes the functional

with

Since does not appear explicitly in the first term in the Euler–Lagrange equation vanishes for all and thus,

Substituting for and taking the derivative,

Thus

for some constant Then

where

Solving, we get

which implies that

is a constant and therefore that the shortest curve that connects two points and is

and we have thus found the extremal function that minimizes the functional so that is a minimum. The equation for a straight line is In other words, the shortest distance between two points is a straight line.[j]

Beltrami's identity

In physics problems it may be the case that meaning the integrand is a function of and but does not appear separately. In that case, the Euler–Lagrange equation can be simplified to the Beltrami identity[16]

where is a constant. The left hand side is the Legendre transformationof with respect to

The intuition behind this result is that, if the variable is actually time, then the statement implies that the Lagrangian is time-independent. By Noether's theorem, there is an associated conserved quantity. In this case, this quantity is the Hamiltonian, the Legendre transform of the Lagrangian, which (often) coincides with the energy of the system. This is (minus) the constant in Beltrami's identity.

Euler–Poisson equation

If depends on higher-derivatives of that is, if

then must satisfy the Euler–Poisson equation,

[17]

Du Bois-Reymond's theorem

The discussion thus far has assumed that extremal functions possess two continuous derivatives, although the existence of the integral requires only first derivatives of trial functions. The condition that the first variation vanishes at an extremal may be regarded as a weak form of the Euler–Lagrange equation. The theorem of Du Bois-Reymond asserts that this weak form implies the strong form. If has continuous first and second derivatives with respect to all of its arguments, and if

then has two continuous derivatives, and it satisfies the Euler–Lagrange equation.

Lavrentiev phenomenon

Hilbert was the first to give good conditions for the Euler–Lagrange equations to give a stationary solution. Within a convex area and a positive thrice differentiable Lagrangian the solutions are composed of a countable collection of sections that either go along the boundary or satisfy the Euler–Lagrange equations in the interior.

However Lavrentiev in 1926 showed that there are circumstances where there is no optimum solution but one can be approached arbitrarily closely by increasing numbers of sections. The Lavrentiev Phenomenon identifies a difference in the infimum of a minimization problem across different classes of admissible functions. For instance the following problem, presented by Manià in 1934:[18]

Clearly, minimizes the functional, but we find any function gives a value bounded away from the infimum.

Examples (in one-dimension) are traditionally manifested across and but Ball and Mizel[19] procured the first functional that displayed Lavrentiev's Phenomenon across and for There are several results that gives criteria under which the phenomenon does not occur - for instance 'standard growth', a Lagrangian with no dependence on the second variable, or an approximating sequence satisfying Cesari's Condition (D) - but results are often particular, and applicable to a small class of functionals.

Connected with the Lavrentiev Phenomenon is the repulsion property: any functional displaying Lavrentiev's Phenomenon will display the weak repulsion property.[20]

Functions of several variables

For example, if denotes the displacement of a membrane above the domain in the plane, then its potential energy is proportional to its surface area:

Plateau's problem consists of finding a function that minimizes the surface area while assuming prescribed values on the boundary of ; the solutions are called minimal surfaces. The Euler–Lagrange equation for this problem is nonlinear:

See Courant (1950) for details.

Dirichlet's principle

It is often sufficient to consider only small displacements of the membrane, whose energy difference from no displacement is approximated by

The functional is to be minimized among all trial functions that assume prescribed values on the boundary of If is the minimizing function and is an arbitrary smooth function that vanishes on the boundary of then the first variation of must vanish:

Provided that u has two derivatives, we may apply the divergence theorem to obtain

where is the boundary of is arclength along and is the normal derivative of on Since vanishes on and the first variation vanishes, the result is

for all smooth functions v that vanish on the boundary of The proof for the case of one dimensional integrals may be adapted to this case to show that

in

The difficulty with this reasoning is the assumption that the minimizing function u must have two derivatives. Riemann argued that the existence of a smooth minimizing function was assured by the connection with the physical problem: membranes do indeed assume configurations with minimal potential energy. Riemann named this idea the Dirichlet principle in honor of his teacher Peter Gustav Lejeune Dirichlet. However Weierstrass gave an example of a variational problem with no solution: minimize

among all functions that satisfy and can be made arbitrarily small by choosing piecewise linear functions that make a transition between −1 and 1 in a small neighborhood of the origin. However, there is no function that makes [k] Eventually it was shown that Dirichlet's principle is valid, but it requires a sophisticated application of the regularity theory for elliptic partial differential equations; see Jost and Li–Jost (1998).

Generalization to other boundary value problems

A more general expression for the potential energy of a membrane is

This corresponds to an external force density in an external force on the boundary and elastic forces with modulus acting on The function that minimizes the potential energy with no restriction on its boundary values will be denoted by Provided that and are continuous, regularity theory implies that the minimizing function will have two derivatives. In taking the first variation, no boundary condition need be imposed on the increment The first variation of is given by

If we apply the divergence theorem, the result is

If we first set on the boundary integral vanishes, and we conclude as before that

in Then if we allow to assume arbitrary boundary values, this implies that must satisfy the boundary condition

on This boundary condition is a consequence of the minimizing property of : it is not imposed beforehand. Such conditions are called natural boundary conditions.

The preceding reasoning is not valid if vanishes identically on In such a case, we could allow a trial function where is a constant. For such a trial function,

By appropriate choice of can assume any value unless the quantity inside the brackets vanishes. Therefore, the variational problem is meaningless unless

This condition implies that net external forces on the system are in equilibrium. If these forces are in equilibrium, then the variational problem has a solution, but it is not unique, since an arbitrary constant may be added. Further details and examples are in Courant and Hilbert (1953).

Eigenvalue problems

Both one-dimensional and multi-dimensional eigenvalue problems can be formulated as variational problems.

Sturm–Liouville problems

The Sturm–Liouville eigenvalue problem involves a general quadratic form

where is restricted to functions that satisfy the boundary conditions

Let be a normalization integral

The functions and are required to be everywhere positive and bounded away from zero. The primary variational problem is to minimize the ratio among all satisfying the endpoint conditions. It is shown below that the Euler–Lagrange equation for the minimizing is

where is the quotient

It can be shown (see Gelfand and Fomin 1963) that the minimizing has two derivatives and satisfies the Euler–Lagrange equation. The associated will be denoted by ; it is the lowest eigenvalue for this equation and boundary conditions. The associated minimizing function will be denoted by This variational characterization of eigenvalues leads to the Rayleigh–Ritz method: choose an approximating as a linear combination of basis functions (for example trigonometric functions) and carry out a finite-dimensional minimization among such linear combinations. This method is often surprisingly accurate.

The next smallest eigenvalue and eigenfunction can be obtained by minimizing under the additional constraint

This procedure can be extended to obtain the complete sequence of eigenvalues and eigenfunctions for the problem.

The variational problem also applies to more general boundary conditions. Instead of requiring that vanish at the endpoints, we may not impose any condition at the endpoints, and set

where and are arbitrary. If we set the first variation for the ratio is

where λ is given by the ratio as previously. After integration by parts,

If we first require that vanish at the endpoints, the first variation will vanish for all such only if

If satisfies this condition, then the first variation will vanish for arbitrary only if

These latter conditions are the natural boundary conditions for this problem, since they are not imposed on trial functions for the minimization, but are instead a consequence of the minimization.

Eigenvalue problems in several dimensions

Eigenvalue problems in higher dimensions are defined in analogy with the one-dimensional case. For example, given a domain with boundary in three dimensions we may define

and

Let be the function that minimizes the quotient with no condition prescribed on the boundary The Euler–Lagrange equation satisfied by is

where

The minimizing must also satisfy the natural boundary condition

on the boundary This result depends upon the regularity theory for elliptic partial differential equations; see Jost and Li–Jost (1998) for details. Many extensions, including completeness results, asymptotic properties of the eigenvalues and results concerning the nodes of the eigenfunctions are in Courant and Hilbert (1953).

Applications

Optics

Fermat's principle states that light takes a path that (locally) minimizes the optical length between its endpoints. If the -coordinate is chosen as the parameter along the path, and along the path, then the optical length is given by

where the refractive index depends upon the material. If we try then the first variationof (the derivative of with respect to ε) is

After integration by parts of the first term within brackets, we obtain the Euler–Lagrange equation

The light rays may be determined by integrating this equation. This formalism is used in the context of Lagrangian optics and Hamiltonian optics.

Snell's law

There is a discontinuity of the refractive index when light enters or leaves a lens. Let

where and are constants. Then the Euler–Lagrange equation holds as before in the region where or and in fact the path is a straight line there, since the refractive index is constant. At the must be continuous, but may be discontinuous. After integration by parts in the separate regions and using the Euler–Lagrange equations, the first variation takes the form

The factor multiplying is the sine of angle of the incident ray with the axis, and the factor multiplying is the sine of angle of the refracted ray with the axis. Snell's law for refraction requires that these terms be equal. As this calculation demonstrates, Snell's law is equivalent to vanishing of the first variation of the optical path length.

Fermat's principle in three dimensions

It is expedient to use vector notation: let let be a parameter, let be the parametric representation of a curve and let be its tangent vector. The optical length of the curve is given by

Note that this integral is invariant with respect to changes in the parametric representation of The Euler–Lagrange equations for a minimizing curve have the symmetric form

where

It follows from the definition that satisfies

Therefore, the integral may also be written as

This form suggests that if we can find a function <math\psi</math> whose gradient is given by then the integral is given by the difference of <math\psi</math> at the endpoints of the interval of integration. Thus the problem of studying the curves that make the integral stationary can be related to the study of the level surfaces of <math\psi.</math> In order to find such a function, we turn to the wave equation, which governs the propagation of light. This formalism is used in the context of Lagrangian optics and Hamiltonian optics.

Connection with the wave equation

The wave equation for an inhomogeneous medium is

where is the velocity, which generally depends upon Wave fronts for light are characteristic surfaces for this partial differential equation: they satisfy

We may look for solutions in the form

In that case, <math\psi</math> satisfies

where According to the theory of first-order partial differential equations, if then satisfies

along a system of curves (the light rays) that are given by

These equations for solution of a first-order partial differential equation are identical to the Euler–Lagrange equations if we make the identification

We conclude that the function <math\psi</math> is the value of the minimizing integral as a function of the upper end point. That is, when a family of minimizing curves is constructed, the values of the optical length satisfy the characteristic equation corresponding the wave equation. Hence, solving the associated partial differential equation of first order is equivalent to finding families of solutions of the variational problem. This is the essential content of the Hamilton–Jacobi theory, which applies to more general variational problems.

Mechanics

In classical mechanics, the action, is defined as the time integral of the Lagrangian, The Lagrangian is the difference of energies,

where is the kinetic energy of a mechanical system and its potential energy. Hamilton's principle (or the action principle) states that the motion of a conservative holonomic (integrable constraints) mechanical system is such that the action integral

is stationary with respect to variations in the path The Euler–Lagrange equations for this system are known as Lagrange's equations:

and they are equivalent to Newton's equations of motion (for such systems).

The conjugate momenta are defined by

For example, if

then

Hamiltonian mechanics results if the conjugate momenta are introduced in place of by a Legendre transformation of the Lagrangian into the Hamiltonian defined by

The Hamiltonian is the total energy of the system: Analogy with Fermat's principle suggests that solutions of Lagrange's equations (the particle trajectories) may be described in terms of level surfaces of some function of This function is a solution of the Hamilton–Jacobi equation:

Further applications

Further applications of the calculus of variations include the following:

Variations and sufficient condition for a minimum

Calculus of variations is concerned with variations of functionals, which are small changes in the functional's value due to small changes in the function that is its argument. The first variation[l] is defined as the linear part of the change in the functional, and the second variation[m] is defined as the quadratic part.[22]

For example, if is a functional with the function as its argument, and there is a small change in its argument from to where is a function in the same function space as then the corresponding change in the functional is

  [n]

The functional is said to be differentiableif

where is a linear functional,[o] is the norm of [p] and as The linear functional is the first variation of and is denoted by,[26]

The functional is said to be twice differentiableif

where is a linear functional (the first variation), is a quadratic functional,[q] and as The quadratic functional is the second variation of and is denoted by,[28]

The second variation is said to be strongly positiveif

for all and for some constant .[29]

Using the above definitions, especially the definitions of first variation, second variation, and strongly positive, the following sufficient condition for a minimum of a functional can be stated.

Sufficient condition for a minimum:

The functional has a minimum at if its first variation at and its second variation is strongly positive at [30] [r][s]

See also

  • Isoperimetric inequality
  • Variational principle
  • Variational bicomplex
  • Fermat's principle
  • Principle of least action
  • Infinite-dimensional optimization
  • Functional analysis
  • Ekeland's variational principle
  • Inverse problem for Lagrangian mechanics
  • Obstacle problem
  • Perturbation methods
  • Young measure
  • Optimal control
  • Direct method in calculus of variations
  • Noether's theorem
  • De Donder–Weyl theory
  • Variational Bayesian methods
  • Chaplygin problem
  • Nehari manifold
  • Hu–Washizu principle
  • Luke's variational principle
  • Mountain pass theorem
  • ‹The template Cat is being considered for merging.› Category:Variational analysts
  • Measures of central tendency as solutions to variational problems
  • Stampacchia Medal
  • Fermat Prize
  • Convenient vector space
  • Notes

    1. ^ Whereas elementary calculus is about infinitesimally small changes in the values of functions without changes in the function itself, calculus of variations is about infinitesimally small changes in the function itself, which are called variations.[1]
  • ^ See Harold J. Kushner (2004): regarding Dynamic Programming, "The calculus of variations had related ideas (e.g., the work of Caratheodory, the Hamilton-Jacobi equation). This led to conflicts with the calculus of variations community."
  • ^ The neighborhood of is the part of the given function space where over the whole domain of the functions, with a positive number that specifies the size of the neighborhood.[10]
  • ^ Note the difference between the terms extremal and extremum. An extremal is a function that makes a functional an extremum.
  • ^ For a sufficient condition, see section Variations and sufficient condition for a minimum.
  • ^ The following derivation of the Euler–Lagrange equation corresponds to the derivation on pp. 184–185 of Courant & Hilbert (1953).[14]
  • ^ Note that and are evaluated at the same values of which is not valid more generally in variational calculus with non-holonomic constraints.
  • ^ The product is called the first variation of the functional and is denoted by Some references define the first variation differently by leaving out the factor.
  • ^ Note that assuming y is a function of x loses generality; ideally both should be a function of some other parameter. This approach is good solely for instructive purposes.
  • ^ As a historical note, this is an axiom of Archimedes. See e.g. Kelland (1843).[15]
  • ^ The resulting controversy over the validity of Dirichlet's principle is explained by Turnbull.[21]
  • ^ The first variation is also called the variation, differential, or first differential.
  • ^ The second variation is also called the second differential.
  • ^ Note that and the variations below, depend on both and The argument has been left out to simplify the notation. For example, could have been written [23]
  • ^ A functional is said to be linearif   and   where are functions and is a real number.[24]
  • ^ For a function that is defined for where and are real numbers, the norm of is its maximum absolute value, i.e. [25]
  • ^ A functional is said to be quadratic if it is a bilinear functional with two argument functions that are equal. A bilinear functional is a functional that depends on two argument functions and is linear when each argument function in turn is fixed while the other argument function is variable.[27]
  • ^ For other sufficient conditions, see in Gelfand & Fomin 2000,
    • Chapter 5: "The Second Variation. Sufficient Conditions for a Weak Extremum" – Sufficient conditions for a weak minimum are given by the theorem on p. 116.
    • Chapter 6: "Fields. Sufficient Conditions for a Strong Extremum" – Sufficient conditions for a strong minimum are given by the theorem on p. 148.
  • ^ One may note the similarity to the sufficient condition for a minimum of a function, where the first derivative is zero and the second derivative is positive.
  • References

  • ^ Gelfand, I. M.; Fomin, S. V. (2000). Silverman, Richard A. (ed.). Calculus of variations (Unabridged repr. ed.). Mineola, New York: Dover Publications. p. 3. ISBN 978-0486414485.
  • ^ a b Thiele, Rüdiger (2007). "Euler and the Calculus of Variations". In Bradley, Robert E.; Sandifer, C. Edward (eds.). Leonhard Euler: Life, Work and Legacy. Elsevier. p. 249. ISBN 9780080471297.
  • ^ Goldstine, Herman H. (2012). A History of the Calculus of Variations from the 17th through the 19th Century. Springer Science & Business Media. p. 110. ISBN 9781461381068.
  • ^ a b c van Brunt, Bruce (2004). The Calculus of Variations. Springer. ISBN 978-0-387-40247-5.
  • ^ a b Ferguson, James (2004). "Brief Survey of the History of the Calculus of Variations and its Applications". arXiv:math/0402357.
  • ^ Dimitri Bertsekas. Dynamic programming and optimal control. Athena Scientific, 2005.
  • ^ Bellman, Richard E. (1954). "Dynamic Programming and a new formalism in the calculus of variations". Proc. Natl. Acad. Sci. 40 (4): 231–235. Bibcode:1954PNAS...40..231B. doi:10.1073/pnas.40.4.231. PMC 527981. PMID 16589462.
  • ^ "Richard E. Bellman Control Heritage Award". American Automatic Control Council. 2004. Retrieved 2013-07-28.
  • ^ Courant, R; Hilbert, D (1953). Methods of Mathematical Physics. Vol. I (First English ed.). New York: Interscience Publishers, Inc. p. 169. ISBN 978-0471504474.
  • ^ Gelfand & Fomin 2000, pp. 12–13
  • ^ Gelfand & Fomin 2000, p. 13
  • ^ Gelfand & Fomin 2000, pp. 14–15
  • ^ Courant, R.; Hilbert, D. (1953). Methods of Mathematical Physics. Vol. I (First English ed.). New York: Interscience Publishers, Inc. ISBN 978-0471504474.
  • ^ Kelland, Philip (1843). Lectures on the principles of demonstrative mathematics. p. 58 – via Google Books.
  • ^ Weisstein, Eric W. "Euler–Lagrange Differential Equation". mathworld.wolfram.com. Wolfram. Eq. (5).
  • ^ Kot, Mark (2014). "Chapter 4: Basic Generalizations". A First Course in the Calculus of Variations. American Mathematical Society. ISBN 978-1-4704-1495-5.
  • ^ Manià, Bernard (1934). "Sopra un esempio di Lavrentieff". Bollenttino dell'Unione Matematica Italiana. 13: 147–153.
  • ^ Ball & Mizel (1985). "One-dimensional Variational problems whose Minimizers do not satisfy the Euler-Lagrange equation". Archive for Rational Mechanics and Analysis. 90 (4): 325–388. Bibcode:1985ArRMA..90..325B. doi:10.1007/BF00276295. S2CID 55005550.
  • ^ Ferriero, Alessandro (2007). "The Weak Repulsion property". Journal de Mathématiques Pures et Appliquées. 88 (4): 378–388. doi:10.1016/j.matpur.2007.06.002.
  • ^ Turnbull. "Riemann biography". UK: U. St. Andrew.
  • ^ Gelfand & Fomin 2000, pp. 11–12, 99
  • ^ Gelfand & Fomin 2000, p. 12, footnote 6
  • ^ Gelfand & Fomin 2000, p. 8
  • ^ Gelfand & Fomin 2000, p. 6
  • ^ Gelfand & Fomin 2000, pp. 11–12
  • ^ Gelfand & Fomin 2000, pp. 97–98
  • ^ Gelfand & Fomin 2000, p. 99
  • ^ Gelfand & Fomin 2000, p. 100
  • ^ Gelfand & Fomin 2000, p. 100, Theorem 2
  • Further reading

    External links


    Retrieved from "https://en.wikipedia.org/w/index.php?title=Calculus_of_variations&oldid=1023181484"

    Categories: 
    Calculus of variations
    Optimization in vector spaces
    Hidden categories: 
    Articles with short description
    Short description is different from Wikidata
    Pages using sidebar with the child parameter
    Articles with J9U identifiers
    Articles with LCCN identifiers
    Articles with NDL identifiers
    Articles with NKC identifiers
     



    This page was last edited on 14 May 2021, at 21:22 (UTC).

    This version of the page has been revised. Besides normal editing, the reason for revision may have been that this version contains factual inaccuracies, vandalism, or material not compatible with the Creative Commons Attribution-ShareAlike License.



    Privacy policy

    About Wikipedia

    Disclaimers

    Contact Wikipedia

    Code of Conduct

    Developers

    Statistics

    Cookie statement

    Mobile view



    Wikimedia Foundation
    Powered by MediaWiki