Inverse function theorem




In mathematics, specifically differential calculus, the inverse function theorem gives a sufficient condition for a function to be invertible in a neighborhood of a point in its domain: namely, that its derivative is continuous and non-zero at the point. The theorem also gives a formula for the derivative of the inverse function.
In multivariable calculus, this theorem can be generalized to any continuously differentiable, vector-valued function whose Jacobian determinant is nonzero at a point in its domain, giving a formula for the Jacobian matrix of the inverse. There are also versions of the inverse function theorem for complex holomorphic functions, for differentiable maps between manifolds, for differentiable functions between Banach spaces, and so forth.




Contents






  • 1 Statement of the theorem


  • 2 Example


  • 3 Methods of proof


  • 4 Generalizations


    • 4.1 Manifolds


    • 4.2 Banach spaces


    • 4.3 Banach manifolds


    • 4.4 Constant rank theorem


    • 4.5 Holomorphic Functions




  • 5 See also


  • 6 Notes


  • 7 References





Statement of the theorem


For functions of a single variable, the theorem states that if f{displaystyle f}f is a continuously differentiable function with nonzero derivative at the point a{displaystyle a}a, then f{displaystyle f}f is invertible in a neighborhood of a{displaystyle a}a, the inverse is continuously differentiable, and the derivative of the inverse function at b=f(a){displaystyle b=f(a)}{displaystyle b=f(a)} is the reciprocal of the derivative of f{displaystyle f}f at a{displaystyle a}a:


(f−1)′(b)=1f′(a).{displaystyle {bigl (}f^{-1}{bigr )}'(b)={frac {1}{f'(a)}}.}{displaystyle {bigl (}f^{-1}{bigr )}'(b)={frac {1}{f'(a)}}.}

For functions of more than one variable, the theorem states that if F{displaystyle F}F is a continuously differentiable function from an open set of Rn{displaystyle mathbb {R} ^{n}}mathbb {R} ^{n} into Rn{displaystyle mathbb {R} ^{n}}mathbb {R} ^{n}, and the total derivative is invertible at a point p{displaystyle p}p (i.e., the Jacobian determinant of F{displaystyle F}F at p{displaystyle p}p is non-zero), then F{displaystyle F}F is invertible near p{displaystyle p}p: an inverse function to F{displaystyle F}F is defined on some neighborhood of q=F(p){displaystyle q=F(p)}{displaystyle q=F(p)}.
Writing F=(F1,…,Fn){displaystyle F=(F_{1},ldots ,F_{n})}{displaystyle F=(F_{1},ldots ,F_{n})}, this means the system of n equations yi=Fi(x1,…,xn){displaystyle y_{i}=F_{i}(x_{1},dots ,x_{n})}y_{i}=F_{i}(x_{1},dots ,x_{n}) has a unique solution for x1,…,xn{displaystyle x_{1},dots ,x_{n}}x_{1},dots ,x_{n} in terms of y1,…,yn{displaystyle y_{1},dots ,y_{n}}y_{1},dots ,y_{n}, provided we restrict x{displaystyle x}x and y{displaystyle y}y to small enough neighborhoods of p{displaystyle p}p and q{displaystyle q}q, respectively.
In the infinite dimensional case, the theorem requires the extra hypothesis that the Fréchet derivative of F{displaystyle F}F at p{displaystyle p}p has a bounded inverse.


Finally, the theorem says that the inverse function F−1{displaystyle F^{-1}}F^{-1} is continuously differentiable, and its Jacobian derivative at q=F(p){displaystyle q=F(p)}{displaystyle q=F(p)} is the matrix inverse of the Jacobian of F{displaystyle F}F at p{displaystyle p}p:


JF−1(q)=[JF(p)]−1.{displaystyle J_{F^{-1}}(q)=[J_{F}(p)]^{-1}.}{displaystyle J_{F^{-1}}(q)=[J_{F}(p)]^{-1}.}

The hard part of the theorem is the existence and differentiability of F−1{displaystyle F^{-1}}F^{-1}. Assuming this, the inverse derivative formula follows from the chain rule applied to F−1∘F=id{displaystyle F^{-1}circ F={text{id}}}{displaystyle F^{-1}circ F={text{id}}}:


I=JF−1∘F(p) = JF−1(F(p))⋅JF(p) = JF−1(q)⋅JF(p).{displaystyle I=J_{F^{-1}circ F}(p) = J_{F^{-1}}(F(p))cdot J_{F}(p) = J_{F^{-1}}(q)cdot J_{F}(p).}{displaystyle I=J_{F^{-1}circ F}(p) = J_{F^{-1}}(F(p))cdot J_{F}(p) = J_{F^{-1}}(q)cdot J_{F}(p).}


Example


Consider the vector-valued function F:R2→R2{displaystyle F:mathbb {R} ^{2}to mathbb {R} ^{2}}{displaystyle F:mathbb {R} ^{2}to mathbb {R} ^{2}} defined by:


F(x,y)=[excos⁡yexsin⁡y].{displaystyle F(x,y)={begin{bmatrix}{e^{x}cos y}\{e^{x}sin y}\end{bmatrix}}.}{displaystyle F(x,y)={begin{bmatrix}{e^{x}cos y}\{e^{x}sin y}\end{bmatrix}}.}

The Jacobian matrix is:


JF(x,y)=[excos⁡y−exsin⁡yexsin⁡yexcos⁡y]{displaystyle J_{F}(x,y)={begin{bmatrix}{e^{x}cos y}&{-e^{x}sin y}\{e^{x}sin y}&{e^{x}cos y}\end{bmatrix}}}J_{F}(x,y)={begin{bmatrix}{e^{x}cos y}&{-e^{x}sin y}\{e^{x}sin y}&{e^{x}cos y}\end{bmatrix}}

with Jacobian determinant:


detJF(x,y)=e2xcos2⁡y+e2xsin2⁡y=e2x.{displaystyle det J_{F}(x,y)=e^{2x}cos ^{2}y+e^{2x}sin ^{2}y=e^{2x}.,!}det J_{F}(x,y)=e^{2x}cos ^{2}y+e^{2x}sin ^{2}y=e^{2x}.,!

The determinant e2x{displaystyle e^{2x}}e^{{2x}} is nonzero everywhere. Thus the theorem guarantees that, for every point p{displaystyle p}p in R2{displaystyle mathbb {R} ^{2}}mathbb {R} ^{2}, there exists a neighborhood about p{displaystyle p}p over which F{displaystyle F}F is invertible. This does not mean F{displaystyle F}F is invertible over its entire domain: in this case F{displaystyle F}F is not even injective since it is periodic: F(x,y)=F(x,y+2π){displaystyle F(x,y)=F(x,y+2pi )}{displaystyle F(x,y)=F(x,y+2pi )}.



Methods of proof


As an important result, the inverse function theorem has been given numerous proofs. The proof most commonly seen in textbooks relies on the contraction mapping principle, also known as the Banach fixed point theorem (which can also be used as the key step in the proof of existence and uniqueness of solutions to ordinary differential equations).
Since the fixed point theorem applies in infinite-dimensional (Banach space) settings, this proof generalizes immediately to the infinite-dimensional version of the inverse function theorem (see "Generalizations", below).
An alternate proof in finite dimensions hinges on the extreme value theorem for functions on a compact set.[1]
Yet another proof uses Newton's method, which has the advantage of providing an effective version of the theorem: bounds on the derivative of the function imply an estimate of the size of the neighborhood on which the function is invertible.[2]



Generalizations



Manifolds


The inverse function theorem can be rephrased in terms of differentiable maps between differentiable manifolds. In this context the theorem states that for a differentiable map F:M→N{displaystyle F:Mto N}F:Mto N (of class C1{displaystyle C^{1}}C^{1}), if the differential of F{displaystyle F}F,


dFp:TpM→TF(p)N{displaystyle dF_{p}:T_{p}Mto T_{F(p)}N}dF_{p}:T_{p}Mto T_{F(p)}N

is a linear isomorphism at a point p{displaystyle p}p in M{displaystyle M}M then there exists an open neighborhood U{displaystyle U}U of p{displaystyle p}p such that


F|U:U→F(U){displaystyle F|_{U}:Uto F(U)}F|_{U}:Uto F(U)

is a diffeomorphism. Note that this implies that M{displaystyle M}M and N{displaystyle N}N must have the same dimension at p{displaystyle p}p.
If the derivative of F{displaystyle F}F is an isomorphism at all points p{displaystyle p}p in M{displaystyle M}M then the map F{displaystyle F}F is a local diffeomorphism.



Banach spaces


The inverse function theorem can also be generalized to differentiable maps between Banach spaces X{displaystyle X}X and Y{displaystyle Y}Y. Let U{displaystyle U}U be an open neighbourhood of the origin in X{displaystyle X}X and F:U→Y{displaystyle F:Uto Y}F:Uto Y a continuously differentiable function, and assume that the derivative dF0:X→Y{displaystyle dF_{0}:Xto Y}dF_{0}:Xto Y of F{displaystyle F}F at 0 is a bounded linear isomorphism of X{displaystyle X}X onto Y{displaystyle Y}Y. Then there exists an open neighbourhood V{displaystyle V}V of F(0){displaystyle F(0)}F(0) in Y{displaystyle Y}Y and a continuously differentiable map G:V→X{displaystyle G:Vto X}G:Vto X such that F(G(y))=y{displaystyle F(G(y))=y}F(G(y))=y for all y{displaystyle y}y in V{displaystyle V}V. Moreover, G(y){displaystyle G(y)}G(y) is the only sufficiently small solution x{displaystyle x}x of the equation F(x)=y{displaystyle F(x)=y}F(x)=y.



Banach manifolds


These two directions of generalization can be combined in the inverse function theorem for Banach manifolds.[3]



Constant rank theorem


The inverse function theorem (and the implicit function theorem) can be seen as a special case of the constant rank theorem, which states that a smooth map with constant rank near a point can be put in a particular normal form near that point.[4] Specifically, if F:M→N{displaystyle F:Mto N}F:Mto N has constant rank near a point p∈M{displaystyle pin M}pin M, then there are open neighborhoods U{displaystyle U}U of p{displaystyle p}p and V{displaystyle V}V of F(p){displaystyle F(p)}F(p) and there are diffeomorphisms u:TpM→U{displaystyle u:T_{p}Mto U}u:T_{p}Mto U and v:TF(p)N→V{displaystyle v:T_{F(p)}Nto V}v:T_{{F(p)}}Nto V such that F(U)⊆V{displaystyle F(U)subseteq V}F(U)subseteq V and such that the derivative dFp:TpM→TF(p)N{displaystyle dF_{p}:T_{p}Mto T_{F(p)}N}dF_{p}:T_{p}Mto T_{F(p)}N is equal to v−1∘F∘u{displaystyle v^{-1}circ Fcirc u}v^{{-1}}circ Fcirc u. That is, F{displaystyle F}F "looks like" its derivative near p{displaystyle p}p. Semicontinuity of the rank function implies that there is an open dense subset of the domain of F{displaystyle F}F on which the derivative has constant rank. Thus the constant rank theorem applies to a generic point of the domain.


When the derivative of F{displaystyle F}F is injective (resp. surjective) at a point p{displaystyle p}p, it is also injective (resp. surjective) in a neighborhood of p{displaystyle p}p, and hence the rank of F{displaystyle F}F is constant on that neighborhood, and the constant rank theorem applies.



Holomorphic Functions


If a holomorphic function F{displaystyle F}F is defined from an open set U{displaystyle U}U of Cn{displaystyle mathbb {C} ^{n}}mathbb {C} ^{n} into Cn{displaystyle mathbb {C} ^{n}}mathbb {C} ^{n}, and the Jacobian matrix of complex derivatives is invertible at a point p{displaystyle p}p, then F{displaystyle F}F is an invertible function near p{displaystyle p}p. This follows immediately from the real multivariable version of the theorem. One can also show that the inverse function is again holomorphic.[5]



See also


  • Implicit function theorem


Notes





  1. ^ Michael Spivak, Calculus on Manifolds.


  2. ^ John H. Hubbard and Barbara Burke Hubbard, Vector Analysis, Linear Algebra, and Differential Forms: a unified approach, Matrix Editions, 2001.


  3. ^ Lang 1995, Lang 1999, pp. 15–19, 25–29.


  4. ^ William M. Boothby, An Introduction to Differentiable Manifolds and Riemannian Geometry, Revised Second Edition, Academic Press, 2002, .mw-parser-output cite.citation{font-style:inherit}.mw-parser-output q{quotes:"""""""'""'"}.mw-parser-output code.cs1-code{color:inherit;background:inherit;border:inherit;padding:inherit}.mw-parser-output .cs1-lock-free a{background:url("//upload.wikimedia.org/wikipedia/commons/thumb/6/65/Lock-green.svg/9px-Lock-green.svg.png")no-repeat;background-position:right .1em center}.mw-parser-output .cs1-lock-limited a,.mw-parser-output .cs1-lock-registration a{background:url("//upload.wikimedia.org/wikipedia/commons/thumb/d/d6/Lock-gray-alt-2.svg/9px-Lock-gray-alt-2.svg.png")no-repeat;background-position:right .1em center}.mw-parser-output .cs1-lock-subscription a{background:url("//upload.wikimedia.org/wikipedia/commons/thumb/a/aa/Lock-red-alt-2.svg/9px-Lock-red-alt-2.svg.png")no-repeat;background-position:right .1em center}.mw-parser-output .cs1-subscription,.mw-parser-output .cs1-registration{color:#555}.mw-parser-output .cs1-subscription span,.mw-parser-output .cs1-registration span{border-bottom:1px dotted;cursor:help}.mw-parser-output .cs1-hidden-error{display:none;font-size:100%}.mw-parser-output .cs1-visible-error{font-size:100%}.mw-parser-output .cs1-subscription,.mw-parser-output .cs1-registration,.mw-parser-output .cs1-format{font-size:95%}.mw-parser-output .cs1-kern-left,.mw-parser-output .cs1-kern-wl-left{padding-left:0.2em}.mw-parser-output .cs1-kern-right,.mw-parser-output .cs1-kern-wl-right{padding-right:0.2em}
    ISBN 0-12-116051-3.



  5. ^ K. Fritzsche, H. Grauert, "From Holomorphic Functions to Complex Manifolds", Springer-Verlag, (2002). Page 33.




References




  • Lang, Serge (1995). Differential and Riemannian Manifolds. Springer. ISBN 0-387-94338-2.


  • Lang, Serge (1999). Fundamentals of Differential Geometry. Graduate Texts in Mathematics. New York: Springer. ISBN 978-0-387-98593-0.


  • Nijenhuis, Albert (1974). "Strong derivatives and inverse mappings". Amer. Math. Monthly. 81 (9): 969&ndash, 980. doi:10.2307/2319298.


  • Renardy, Michael; Rogers, Robert C. (2004). An introduction to partial differential equations. Texts in Applied Mathematics 13 (Second ed.). New York: Springer-Verlag. pp. 337&ndash, 338. ISBN 0-387-00444-0.


  • Rudin, Walter (1976). Principles of mathematical analysis. International Series in Pure and Applied Mathematics (Third ed.). New York: McGraw-Hill Book Co. pp. 221&ndash, 223.









這個網誌中的熱門文章

Tangent Lines Diagram Along Smooth Curve

Yusuf al-Mu'taman ibn Hud

Zucchini