Inverse function theorem

The name of the pictureThe name of the pictureThe name of the pictureClash Royale CLAN TAG#URR8PPP




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 fdisplaystyle ff is a continuously differentiable function with nonzero derivative at the point adisplaystyle aa, then fdisplaystyle ff is invertible in a neighborhood of adisplaystyle aa, 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 fdisplaystyle ff at adisplaystyle aa:


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

For functions of more than one variable, the theorem states that if Fdisplaystyle FF is a continuously differentiable function from an open set of Rndisplaystyle mathbb R ^nmathbb R ^n into Rndisplaystyle mathbb R ^nmathbb R ^n, and the total derivative is invertible at a point pdisplaystyle pp (i.e., the Jacobian determinant of Fdisplaystyle FF at pdisplaystyle pp is non-zero), then Fdisplaystyle FF is invertible near pdisplaystyle pp: an inverse function to Fdisplaystyle FF 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,…,xndisplaystyle x_1,dots ,x_nx_1,dots ,x_n in terms of y1,…,yndisplaystyle y_1,dots ,y_ny_1,dots ,y_n, provided we restrict xdisplaystyle xx and ydisplaystyle yy to small enough neighborhoods of pdisplaystyle pp and qdisplaystyle qq, respectively.
In the infinite dimensional case, the theorem requires the extra hypothesis that the Fréchet derivative of Fdisplaystyle FF at pdisplaystyle pp has a bounded inverse.


Finally, the theorem says that the inverse function F−1displaystyle F^-1F^-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 Fdisplaystyle FF at pdisplaystyle pp:


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−1displaystyle F^-1F^-1. Assuming this, the inverse derivative formula follows from the chain rule applied to F−1∘F=iddisplaystyle F^-1circ F=textiddisplaystyle F^-1circ F=textid:


I=JF−1∘F(p) = JF−1(F(p))⋅JF(p) = JF−1(q)⋅JF(p).displaystyle I=J_F^-1circ F(p) = J_F^-1(F(p))cdot J_F(p) = J_F^-1(q)cdot J_F(p).displaystyle I=J_F^-1circ 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→R2displaystyle F:mathbb R ^2to mathbb R ^2displaystyle F:mathbb R ^2to mathbb R ^2 defined by:


F(x,y)=[excos⁡yexsin⁡y].displaystyle F(x,y)=beginbmatrixe^xcos y\e^xsin y\endbmatrix.displaystyle F(x,y)=beginbmatrixe^xcos y\e^xsin y\endbmatrix.

The Jacobian matrix is:


JF(x,y)=[excos⁡y−exsin⁡yexsin⁡yexcos⁡y]displaystyle J_F(x,y)=beginbmatrixe^xcos y&-e^xsin y\e^xsin y&e^xcos y\endbmatrixJ_F(x,y)=beginbmatrixe^xcos y&-e^xsin y\e^xsin y&e^xcos y\endbmatrix

with Jacobian determinant:


detJF(x,y)=e2xcos2⁡y+e2xsin2⁡y=e2x.displaystyle det J_F(x,y)=e^2xcos ^2y+e^2xsin ^2y=e^2x.,!det J_F(x,y)=e^2xcos ^2y+e^2xsin ^2y=e^2x.,!

The determinant e2xdisplaystyle e^2xe^2x is nonzero everywhere. Thus the theorem guarantees that, for every point pdisplaystyle pp in R2displaystyle mathbb R ^2mathbb R ^2, there exists a neighborhood about pdisplaystyle pp over which Fdisplaystyle FF is invertible. This does not mean Fdisplaystyle FF is invertible over its entire domain: in this case Fdisplaystyle FF 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→Ndisplaystyle F:Mto NF:Mto N (of class C1displaystyle C^1C^1), if the differential of Fdisplaystyle FF,


dFp:TpM→TF(p)Ndisplaystyle dF_p:T_pMto T_F(p)NdF_p:T_pMto T_F(p)N

is a linear isomorphism at a point pdisplaystyle pp in Mdisplaystyle MM then there exists an open neighborhood Udisplaystyle UU of pdisplaystyle pp such that


F|U:U→F(U)_U:Uto F(U)F|_U:Uto F(U)

is a diffeomorphism. Note that this implies that Mdisplaystyle MM and Ndisplaystyle NN must have the same dimension at pdisplaystyle pp.
If the derivative of Fdisplaystyle FF is an isomorphism at all points pdisplaystyle pp in Mdisplaystyle MM then the map Fdisplaystyle FF is a local diffeomorphism.



Banach spaces


The inverse function theorem can also be generalized to differentiable maps between Banach spaces Xdisplaystyle XX and Ydisplaystyle YY. Let Udisplaystyle UU be an open neighbourhood of the origin in Xdisplaystyle XX and F:U→Ydisplaystyle F:Uto YF:Uto Y a continuously differentiable function, and assume that the derivative dF0:X→Ydisplaystyle dF_0:Xto YdF_0:Xto Y of Fdisplaystyle FF at 0 is a bounded linear isomorphism of Xdisplaystyle XX onto Ydisplaystyle YY. Then there exists an open neighbourhood Vdisplaystyle VV of F(0)displaystyle F(0)F(0) in Ydisplaystyle YY and a continuously differentiable map G:V→Xdisplaystyle G:Vto XG:Vto X such that F(G(y))=ydisplaystyle F(G(y))=yF(G(y))=y for all ydisplaystyle yy in Vdisplaystyle VV. Moreover, G(y)displaystyle G(y)G(y) is the only sufficiently small solution xdisplaystyle xx of the equation F(x)=ydisplaystyle F(x)=yF(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→Ndisplaystyle F:Mto NF:Mto N has constant rank near a point p∈Mdisplaystyle pin Mpin M, then there are open neighborhoods Udisplaystyle UU of pdisplaystyle pp and Vdisplaystyle VV of F(p)displaystyle F(p)F(p) and there are diffeomorphisms u:TpM→Udisplaystyle u:T_pMto Uu:T_pMto U and v:TF(p)N→Vdisplaystyle v:T_F(p)Nto Vv:T_F(p)Nto V such that F(U)⊆Vdisplaystyle F(U)subseteq VF(U)subseteq V and such that the derivative dFp:TpM→TF(p)Ndisplaystyle dF_p:T_pMto T_F(p)NdF_p:T_pMto T_F(p)N is equal to v−1∘F∘udisplaystyle v^-1circ Fcirc uv^-1circ Fcirc u. That is, Fdisplaystyle FF "looks like" its derivative near pdisplaystyle pp. Semicontinuity of the rank function implies that there is an open dense subset of the domain of Fdisplaystyle FF on which the derivative has constant rank. Thus the constant rank theorem applies to a generic point of the domain.


When the derivative of Fdisplaystyle FF is injective (resp. surjective) at a point pdisplaystyle pp, it is also injective (resp. surjective) in a neighborhood of pdisplaystyle pp, and hence the rank of Fdisplaystyle FF is constant on that neighborhood, and the constant rank theorem applies.



Holomorphic Functions


If a holomorphic function Fdisplaystyle FF is defined from an open set Udisplaystyle UU of Cndisplaystyle mathbb C ^nmathbb C ^n into Cndisplaystyle mathbb C ^nmathbb C ^n, and the Jacobian matrix of complex derivatives is invertible at a point pdisplaystyle pp, then Fdisplaystyle FF is an invertible function near pdisplaystyle pp. 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.citationfont-style:inherit.mw-parser-output .citation qquotes:"""""""'""'".mw-parser-output .citation .cs1-lock-free abackground: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 .citation .cs1-lock-limited a,.mw-parser-output .citation .cs1-lock-registration abackground: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 .citation .cs1-lock-subscription abackground: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-registrationcolor:#555.mw-parser-output .cs1-subscription span,.mw-parser-output .cs1-registration spanborder-bottom:1px dotted;cursor:help.mw-parser-output .cs1-ws-icon abackground:url("//upload.wikimedia.org/wikipedia/commons/thumb/4/4c/Wikisource-logo.svg/12px-Wikisource-logo.svg.png")no-repeat;background-position:right .1em center.mw-parser-output code.cs1-codecolor:inherit;background:inherit;border:inherit;padding:inherit.mw-parser-output .cs1-hidden-errordisplay:none;font-size:100%.mw-parser-output .cs1-visible-errorfont-size:100%.mw-parser-output .cs1-maintdisplay:none;color:#33aa33;margin-left:0.3em.mw-parser-output .cs1-subscription,.mw-parser-output .cs1-registration,.mw-parser-output .cs1-formatfont-size:95%.mw-parser-output .cs1-kern-left,.mw-parser-output .cs1-kern-wl-leftpadding-left:0.2em.mw-parser-output .cs1-kern-right,.mw-parser-output .cs1-kern-wl-rightpadding-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): 969–980. 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. 337–338. 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. 221–223.







Popular posts from this blog

𛂒𛀶,𛀽𛀑𛂀𛃧𛂓𛀙𛃆𛃑𛃷𛂟𛁡𛀢𛀟𛁤𛂽𛁕𛁪𛂟𛂯,𛁞𛂧𛀴𛁄𛁠𛁼𛂿𛀤 𛂘,𛁺𛂾𛃭𛃭𛃵𛀺,𛂣𛃍𛂖𛃶 𛀸𛃀𛂖𛁶𛁏𛁚 𛂢𛂞 𛁰𛂆𛀔,𛁸𛀽𛁓𛃋𛂇𛃧𛀧𛃣𛂐𛃇,𛂂𛃻𛃲𛁬𛃞𛀧𛃃𛀅 𛂭𛁠𛁡𛃇𛀷𛃓𛁥,𛁙𛁘𛁞𛃸𛁸𛃣𛁜,𛂛,𛃿,𛁯𛂘𛂌𛃛𛁱𛃌𛂈𛂇 𛁊𛃲,𛀕𛃴𛀜 𛀶𛂆𛀶𛃟𛂉𛀣,𛂐𛁞𛁾 𛁷𛂑𛁳𛂯𛀬𛃅,𛃶𛁼

Edmonton

Crossroads (UK TV series)