Direct method in the calculus of variations

Method for constructing existence proofs and calculating solutions in variational calculus
Part of a series of articles about
Calculus
a b f ( t ) d t = f ( b ) f ( a ) {\displaystyle \int _{a}^{b}f'(t)\,dt=f(b)-f(a)}
  • Rolle's theorem
  • Mean value theorem
  • Inverse function theorem
Differential
Definitions
Concepts
  • Differentiation notation
  • Second derivative
  • Implicit differentiation
  • Logarithmic differentiation
  • Related rates
  • Taylor's theorem
Rules and identities
  • v
  • t
  • e

In mathematics, the direct method in the calculus of variations is a general method for constructing a proof of the existence of a minimizer for a given functional,[1] introduced by Stanisław Zaremba and David Hilbert around 1900. The method relies on methods of functional analysis and topology. As well as being used to prove the existence of a solution, direct methods may be used to compute the solution to desired accuracy.[2]

The method

The calculus of variations deals with functionals J : V R ¯ {\displaystyle J:V\to {\bar {\mathbb {R} }}} , where V {\displaystyle V} is some function space and R ¯ = R { } {\displaystyle {\bar {\mathbb {R} }}=\mathbb {R} \cup \{\infty \}} . The main interest of the subject is to find minimizers for such functionals, that is, functions v V {\displaystyle v\in V} such that J ( v ) J ( u ) {\displaystyle J(v)\leq J(u)} for all u V {\displaystyle u\in V} .

The standard tool for obtaining necessary conditions for a function to be a minimizer is the Euler–Lagrange equation. But seeking a minimizer amongst functions satisfying these may lead to false conclusions if the existence of a minimizer is not established beforehand.

The functional J {\displaystyle J} must be bounded from below to have a minimizer. This means

inf { J ( u ) | u V } > . {\displaystyle \inf\{J(u)|u\in V\}>-\infty .\,}

This condition is not enough to know that a minimizer exists, but it shows the existence of a minimizing sequence, that is, a sequence ( u n ) {\displaystyle (u_{n})} in V {\displaystyle V} such that J ( u n ) inf { J ( u ) | u V } . {\displaystyle J(u_{n})\to \inf\{J(u)|u\in V\}.}

The direct method may be broken into the following steps

  1. Take a minimizing sequence ( u n ) {\displaystyle (u_{n})} for J {\displaystyle J} .
  2. Show that ( u n ) {\displaystyle (u_{n})} admits some subsequence ( u n k ) {\displaystyle (u_{n_{k}})} , that converges to a u 0 V {\displaystyle u_{0}\in V} with respect to a topology τ {\displaystyle \tau } on V {\displaystyle V} .
  3. Show that J {\displaystyle J} is sequentially lower semi-continuous with respect to the topology τ {\displaystyle \tau } .

To see that this shows the existence of a minimizer, consider the following characterization of sequentially lower-semicontinuous functions.

The function J {\displaystyle J} is sequentially lower-semicontinuous if
lim inf n J ( u n ) J ( u 0 ) {\displaystyle \liminf _{n\to \infty }J(u_{n})\geq J(u_{0})} for any convergent sequence u n u 0 {\displaystyle u_{n}\to u_{0}} in V {\displaystyle V} .

The conclusions follows from

inf { J ( u ) | u V } = lim n J ( u n ) = lim k J ( u n k ) J ( u 0 ) inf { J ( u ) | u V } {\displaystyle \inf\{J(u)|u\in V\}=\lim _{n\to \infty }J(u_{n})=\lim _{k\to \infty }J(u_{n_{k}})\geq J(u_{0})\geq \inf\{J(u)|u\in V\}} ,

in other words

J ( u 0 ) = inf { J ( u ) | u V } {\displaystyle J(u_{0})=\inf\{J(u)|u\in V\}} .

Details

Banach spaces

The direct method may often be applied with success when the space V {\displaystyle V} is a subset of a separable reflexive Banach space W {\displaystyle W} . In this case the sequential Banach–Alaoglu theorem implies that any bounded sequence ( u n ) {\displaystyle (u_{n})} in V {\displaystyle V} has a subsequence that converges to some u 0 {\displaystyle u_{0}} in W {\displaystyle W} with respect to the weak topology. If V {\displaystyle V} is sequentially closed in W {\displaystyle W} , so that u 0 {\displaystyle u_{0}} is in V {\displaystyle V} , the direct method may be applied to a functional J : V R ¯ {\displaystyle J:V\to {\bar {\mathbb {R} }}} by showing

  1. J {\displaystyle J} is bounded from below,
  2. any minimizing sequence for J {\displaystyle J} is bounded, and
  3. J {\displaystyle J} is weakly sequentially lower semi-continuous, i.e., for any weakly convergent sequence u n u 0 {\displaystyle u_{n}\to u_{0}} it holds that lim inf n J ( u n ) J ( u 0 ) {\displaystyle \liminf _{n\to \infty }J(u_{n})\geq J(u_{0})} .

The second part is usually accomplished by showing that J {\displaystyle J} admits some growth condition. An example is

J ( x ) α x q β {\displaystyle J(x)\geq \alpha \lVert x\rVert ^{q}-\beta } for some α > 0 {\displaystyle \alpha >0} , q 1 {\displaystyle q\geq 1} and β 0 {\displaystyle \beta \geq 0} .

A functional with this property is sometimes called coercive. Showing sequential lower semi-continuity is usually the most difficult part when applying the direct method. See below for some theorems for a general class of functionals.

Sobolev spaces

The typical functional in the calculus of variations is an integral of the form

J ( u ) = Ω F ( x , u ( x ) , u ( x ) ) d x {\displaystyle J(u)=\int _{\Omega }F(x,u(x),\nabla u(x))dx}

where Ω {\displaystyle \Omega } is a subset of R n {\displaystyle \mathbb {R} ^{n}} and F {\displaystyle F} is a real-valued function on Ω × R m × R m n {\displaystyle \Omega \times \mathbb {R} ^{m}\times \mathbb {R} ^{mn}} . The argument of J {\displaystyle J} is a differentiable function u : Ω R m {\displaystyle u:\Omega \to \mathbb {R} ^{m}} , and its Jacobian u ( x ) {\displaystyle \nabla u(x)} is identified with a m n {\displaystyle mn} -vector.

When deriving the Euler–Lagrange equation, the common approach is to assume Ω {\displaystyle \Omega } has a C 2 {\displaystyle C^{2}} boundary and let the domain of definition for J {\displaystyle J} be C 2 ( Ω , R m ) {\displaystyle C^{2}(\Omega ,\mathbb {R} ^{m})} . This space is a Banach space when endowed with the supremum norm, but it is not reflexive. When applying the direct method, the functional is usually defined on a Sobolev space W 1 , p ( Ω , R m ) {\displaystyle W^{1,p}(\Omega ,\mathbb {R} ^{m})} with p > 1 {\displaystyle p>1} , which is a reflexive Banach space. The derivatives of u {\displaystyle u} in the formula for J {\displaystyle J} must then be taken as weak derivatives.

Another common function space is W g 1 , p ( Ω , R m ) {\displaystyle W_{g}^{1,p}(\Omega ,\mathbb {R} ^{m})} which is the affine sub space of W 1 , p ( Ω , R m ) {\displaystyle W^{1,p}(\Omega ,\mathbb {R} ^{m})} of functions whose trace is some fixed function g {\displaystyle g} in the image of the trace operator. This restriction allows finding minimizers of the functional J {\displaystyle J} that satisfy some desired boundary conditions. This is similar to solving the Euler–Lagrange equation with Dirichlet boundary conditions. Additionally there are settings in which there are minimizers in W g 1 , p ( Ω , R m ) {\displaystyle W_{g}^{1,p}(\Omega ,\mathbb {R} ^{m})} but not in W 1 , p ( Ω , R m ) {\displaystyle W^{1,p}(\Omega ,\mathbb {R} ^{m})} . The idea of solving minimization problems while restricting the values on the boundary can be further generalized by looking on function spaces where the trace is fixed only on a part of the boundary, and can be arbitrary on the rest.

The next section presents theorems regarding weak sequential lower semi-continuity of functionals of the above type.

Sequential lower semi-continuity of integrals

As many functionals in the calculus of variations are of the form

J ( u ) = Ω F ( x , u ( x ) , u ( x ) ) d x {\displaystyle J(u)=\int _{\Omega }F(x,u(x),\nabla u(x))dx} ,

where Ω R n {\displaystyle \Omega \subseteq \mathbb {R} ^{n}} is open, theorems characterizing functions F {\displaystyle F} for which J {\displaystyle J} is weakly sequentially lower-semicontinuous in W 1 , p ( Ω , R m ) {\displaystyle W^{1,p}(\Omega ,\mathbb {R} ^{m})} with p 1 {\displaystyle p\geq 1} is of great importance.

In general one has the following:[3]

Assume that F {\displaystyle F} is a function that has the following properties:
  1. The function F {\displaystyle F} is a Carathéodory function.
  2. There exist a L q ( Ω , R m n ) {\displaystyle a\in L^{q}(\Omega ,\mathbb {R} ^{mn})} with Hölder conjugate q = p p 1 {\displaystyle q={\tfrac {p}{p-1}}} and b L 1 ( Ω ) {\displaystyle b\in L^{1}(\Omega )} such that the following inequality holds true for almost every x Ω {\displaystyle x\in \Omega } and every ( y , A ) R m × R m n {\displaystyle (y,A)\in \mathbb {R} ^{m}\times \mathbb {R} ^{mn}} : F ( x , y , A ) a ( x ) , A + b ( x ) {\displaystyle F(x,y,A)\geq \langle a(x),A\rangle +b(x)} . Here, a ( x ) , A {\displaystyle \langle a(x),A\rangle } denotes the Frobenius inner product of a ( x ) {\displaystyle a(x)} and A {\displaystyle A} in R m n {\displaystyle \mathbb {R} ^{mn}} ).
If the function A F ( x , y , A ) {\displaystyle A\mapsto F(x,y,A)} is convex for almost every x Ω {\displaystyle x\in \Omega } and every y R m {\displaystyle y\in \mathbb {R} ^{m}} ,
then J {\displaystyle J} is sequentially weakly lower semi-continuous.

When n = 1 {\displaystyle n=1} or m = 1 {\displaystyle m=1} the following converse-like theorem holds[4]

Assume that F {\displaystyle F} is continuous and satisfies
| F ( x , y , A ) | a ( x , | y | , | A | ) {\displaystyle |F(x,y,A)|\leq a(x,|y|,|A|)}
for every ( x , y , A ) {\displaystyle (x,y,A)} , and a fixed function a ( x , | y | , | A | ) {\displaystyle a(x,|y|,|A|)} increasing in | y | {\displaystyle |y|} and | A | {\displaystyle |A|} , and locally integrable in x {\displaystyle x} . If J {\displaystyle J} is sequentially weakly lower semi-continuous, then for any given ( x , y ) Ω × R m {\displaystyle (x,y)\in \Omega \times \mathbb {R} ^{m}} the function A F ( x , y , A ) {\displaystyle A\mapsto F(x,y,A)} is convex.

In conclusion, when m = 1 {\displaystyle m=1} or n = 1 {\displaystyle n=1} , the functional J {\displaystyle J} , assuming reasonable growth and boundedness on F {\displaystyle F} , is weakly sequentially lower semi-continuous if, and only if the function A F ( x , y , A ) {\displaystyle A\mapsto F(x,y,A)} is convex.

However, there are many interesting cases where one cannot assume that F {\displaystyle F} is convex. The following theorem[5] proves sequential lower semi-continuity using a weaker notion of convexity:

Assume that F : Ω × R m × R m n [ 0 , ) {\displaystyle F:\Omega \times \mathbb {R} ^{m}\times \mathbb {R} ^{mn}\to [0,\infty )} is a function that has the following properties:
  1. The function F {\displaystyle F} is a Carathéodory function.
  2. The function F {\displaystyle F} has p {\displaystyle p} -growth for some p > 1 {\displaystyle p>1} : There exists a constant C {\displaystyle C} such that for every y R m {\displaystyle y\in \mathbb {R} ^{m}} and for almost every x Ω {\displaystyle x\in \Omega } | F ( x , y , A ) | C ( 1 + | y | p + | A | p ) {\displaystyle |F(x,y,A)|\leq C(1+|y|^{p}+|A|^{p})} .
  3. For every y R m {\displaystyle y\in \mathbb {R} ^{m}} and for almost every x Ω {\displaystyle x\in \Omega } , the function A F ( x , y , A ) {\displaystyle A\mapsto F(x,y,A)} is quasiconvex: there exists a cube D R n {\displaystyle D\subseteq \mathbb {R} ^{n}} such that for every A R m n , φ W 0 1 , ( Ω , R m ) {\displaystyle A\in \mathbb {R} ^{mn},\varphi \in W_{0}^{1,\infty }(\Omega ,\mathbb {R} ^{m})} it holds:

F ( x , y , A ) | D | 1 D F ( x , y , A + φ ( z ) ) d z {\displaystyle F(x,y,A)\leq |D|^{-1}\int _{D}F(x,y,A+\nabla \varphi (z))dz}

where | D | {\displaystyle |D|} is the volume of D {\displaystyle D} .
Then J {\displaystyle J} is sequentially weakly lower semi-continuous in W 1 , p ( Ω , R m ) {\displaystyle W^{1,p}(\Omega ,\mathbb {R} ^{m})} .

A converse like theorem in this case is the following: [6]

Assume that F {\displaystyle F} is continuous and satisfies
| F ( x , y , A ) | a ( x , | y | , | A | ) {\displaystyle |F(x,y,A)|\leq a(x,|y|,|A|)}
for every ( x , y , A ) {\displaystyle (x,y,A)} , and a fixed function a ( x , | y | , | A | ) {\displaystyle a(x,|y|,|A|)} increasing in | y | {\displaystyle |y|} and | A | {\displaystyle |A|} , and locally integrable in x {\displaystyle x} . If J {\displaystyle J} is sequentially weakly lower semi-continuous, then for any given ( x , y ) Ω × R m {\displaystyle (x,y)\in \Omega \times \mathbb {R} ^{m}} the function A F ( x , y , A ) {\displaystyle A\mapsto F(x,y,A)} is quasiconvex. The claim is true even when both m , n {\displaystyle m,n} are bigger than 1 {\displaystyle 1} and coincides with the previous claim when m = 1 {\displaystyle m=1} or n = 1 {\displaystyle n=1} , since then quasiconvexity is equivalent to convexity.

Notes

  1. ^ Dacorogna, pp. 1–43.
  2. ^ I. M. Gelfand; S. V. Fomin (1991). Calculus of Variations. Dover Publications. ISBN 978-0-486-41448-5.
  3. ^ Dacorogna, pp. 74–79.
  4. ^ Dacorogna, pp. 66–74.
  5. ^ Acerbi-Fusco
  6. ^ Dacorogna, pp. 156.

References and further reading

  • Dacorogna, Bernard (1989). Direct Methods in the Calculus of Variations. Springer-Verlag. ISBN 0-387-50491-5.
  • Fonseca, Irene; Giovanni Leoni (2007). Modern Methods in the Calculus of Variations: L p {\displaystyle L^{p}} Spaces. Springer. ISBN 978-0-387-35784-3.
  • Morrey, C. B., Jr.: Multiple Integrals in the Calculus of Variations. Springer, 1966 (reprinted 2008), Berlin ISBN 978-3-540-69915-6.
  • Jindřich Nečas: Direct Methods in the Theory of Elliptic Equations. (Transl. from French original 1967 by A.Kufner and G.Tronel), Springer, 2012, ISBN 978-3-642-10455-8.
  • T. Roubíček (2000). "Direct method for parabolic problems". Adv. Math. Sci. Appl. Vol. 10. pp. 57–65. MR 1769181.
  • Acerbi Emilio, Fusco Nicola. "Semicontinuity problems in the calculus of variations." Archive for Rational Mechanics and Analysis 86.2 (1984): 125-145