fbpx
Wikipedia

Intermediate value theorem

In mathematical analysis, the intermediate value theorem states that if is a continuous function whose domain contains the interval [a, b], then it takes on any given value between and at some point within the interval.

Intermediate value theorem: Let be a continuous function defined on and let be a number with . Then there exists some between and such that .

This has two important corollaries:

  1. If a continuous function has values of opposite sign inside an interval, then it has a root in that interval (Bolzano's theorem).[1] [2]
  2. The image of a continuous function over an interval is itself an interval.

Motivation edit

 
The intermediate value theorem

This captures an intuitive property of continuous functions over the real numbers: given   continuous on   with the known values   and  , then the graph of   must pass through the horizontal line   while   moves from   to  . It represents the idea that the graph of a continuous function on a closed interval can be drawn without lifting a pencil from the paper.

Theorem edit

The intermediate value theorem states the following:

Consider an interval   of real numbers   and a continuous function  . Then

  • Version I. if   is a number between   and  , that is,
     
    then there is a   such that  .
  • Version II. the image set   is also an interval (closed), and it contains  .

Remark: Version II states that the set of function values has no gap. For any two function values   with  , even if they are outside the interval between   and  , all points in the interval   are also function values,

 
A subset of the real numbers with no internal gap is an interval. Version I is naturally contained in Version II.

Relation to completeness edit

The theorem depends on, and is equivalent to, the completeness of the real numbers. The intermediate value theorem does not apply to the rational numbers Q because gaps exist between rational numbers; irrational numbers fill those gaps. For example, the function   for   satisfies   and  . However, there is no rational number   such that  , because   is an irrational number.

Proof edit

Proof version A edit

The theorem may be proven as a consequence of the completeness property of the real numbers as follows:[3]

We shall prove the first case,  . The second case is similar.

Let   be the set of all   such that  . Then   is non-empty since   is an element of  . Since   is non-empty and bounded above by  , by completeness, the supremum   exists. That is,   is the smallest number that is greater than or equal to every member of  .

Note that, due to the continuity of   at  , we can keep   within any   of   by keeping   sufficiently close to  . Since   is a strict inequality, consider the implication when   is the distance between   and  . No   sufficiently close to   can then make   greater than or equal to  , which means there are values greater than   in  . A more detailed proof goes like this:

Choose  . Then   such that  ,

 
Consider the interval  . Notice that   and every   satisfies the condition  . Therefore for every   we have  . Hence   cannot be  .

Likewise, due to the continuity of   at  , we can keep   within any   of   by keeping   sufficiently close to  . Since   is a strict inequality, consider the similar implication when   is the distance between   and  . Every   sufficiently close to   must then make   greater than  , which means there are values smaller than   that are upper bounds of  . A more detailed proof goes like this:

Choose  . Then   such that  ,

 
Consider the interval  . Notice that   and every   satisfies the condition  . Therefore for every   we have  . Hence   cannot be  .

With   and  , it must be the case  . Now we claim that  .

Fix some  . Since   is continuous at  ,   such that  ,  .

Since   and   is open,   such that  . Set  . Then we have

 
for all  . By the properties of the supremum, there exists some   that is contained in  , and so
 
Picking  , we know that   because   is the supremum of  . This means that
 
Both inequalities
 
are valid for all  , from which we deduce   as the only possible value, as stated.

Proof version B edit

We will only prove the case of  , as the   case is similar.[4]

Define   which is equivalent to   and lets us rewrite   as  , and we have to prove, that   for some  , which is more intuitive. We further define the set  . Because   we know, that   so, that   is not empty. Moreover, as  , we know that   is bounded and non-empty, so by Completeness, the supremum   exists.

There are 3 cases for the value of  , those being   and  . For contradiction, let us assume, that  . Then, by the definition of continuity, for  , there exists a   such that   implies, that  , which is equivalent to  . If we just chose  , where  , then   and  , so  . It follows that   is an upper bound for  . However,  , contradicting the upper bound property of the least upper bound  , so  . Assume then, that  . We similarly chose   and know, that there exists a   such that   implies  . We can rewrite this as   which implies, that  . If we now chose  , then   and  . It follows that   is an upper bound for  . However,  , which contradict the least property of the least upper bound  , which means, that   is impossible. If we combine both results, we get that   or   is the only remaining possibility.

Remark: The intermediate value theorem can also be proved using the methods of non-standard analysis, which places "intuitive" arguments involving infinitesimals on a rigorous[clarification needed] footing.[5]

History edit

A form of the theorem was postulated as early as the 5th century BCE, in the work of Bryson of Heraclea on squaring the circle. Bryson argued that, as circles larger than and smaller than a given square both exist, there must exist a circle of equal area.[6] The theorem was first proved by Bernard Bolzano in 1817. Bolzano used the following formulation of the theorem:[7]

Let   be continuous functions on the interval between   and   such that   and  . Then there is an   between   and   such that  .

The equivalence between this formulation and the modern one can be shown by setting   to the appropriate constant function. Augustin-Louis Cauchy provided the modern formulation and a proof in 1821.[8] Both were inspired by the goal of formalizing the analysis of functions and the work of Joseph-Louis Lagrange. The idea that continuous functions possess the intermediate value property has an earlier origin. Simon Stevin proved the intermediate value theorem for polynomials (using a cubic as an example) by providing an algorithm for constructing the decimal expansion of the solution. The algorithm iteratively subdivides the interval into 10 parts, producing an additional decimal digit at each step of the iteration.[9] Before the formal definition of continuity was given, the intermediate value property was given as part of the definition of a continuous function. Proponents include Louis Arbogast, who assumed the functions to have no jumps, satisfy the intermediate value property and have increments whose sizes corresponded to the sizes of the increments of the variable.[10] Earlier authors held the result to be intuitively obvious and requiring no proof. The insight of Bolzano and Cauchy was to define a general notion of continuity (in terms of infinitesimals in Cauchy's case and using real inequalities in Bolzano's case), and to provide a proof based on such definitions.

Converse is false edit

A Darboux function is a real-valued function f that has the "intermediate value property," i.e., that satisfies the conclusion of the intermediate value theorem: for any two values a and b in the domain of f, and any y between f(a) and f(b), there is some c between a and b with f(c) = y. The intermediate value theorem says that every continuous function is a Darboux function. However, not every Darboux function is continuous; i.e., the converse of the intermediate value theorem is false.

As an example, take the function f : [0, ∞) → [−1, 1] defined by f(x) = sin(1/x) for x > 0 and f(0) = 0. This function is not continuous at x = 0 because the limit of f(x) as x tends to 0 does not exist; yet the function has the intermediate value property. Another, more complicated example is given by the Conway base 13 function.

In fact, Darboux's theorem states that all functions that result from the differentiation of some other function on some interval have the intermediate value property (even though they need not be continuous).

Historically, this intermediate value property has been suggested as a definition for continuity of real-valued functions;[11] this definition was not adopted.

Generalizations edit

Multi-dimensional spaces edit

The Poincaré-Miranda theorem is a generalization of the Intermediate value theorem from a (one-dimensional) interval to a (two-dimensional) rectangle, or more generally, to an n-dimensional cube.

Vrahatis[12] presents a similar generalization to triangles, or more generally, n-dimensional simplices. Let Dn be an n-dimensional simplex with n+1 vertices denoted by v0,...,vn. Let F=(f1,...,fn) be a continuous function from Dn to Rn, that never equals 0 on the boundary of Dn. Suppose F satisfies the following conditions:

  • For all i in 1,...,n, the sign of fi(vi) is opposite to the sign of fi(x) for all points x on the face opposite to vi;
  • The sign-vector of f1,...,fn on v0 is not equal to the sign-vector of f1,...,fn on all points on the face opposite to v0.

Then there is a point z in the interior of Dn on which F(z)=(0,...,0).

It is possible to normalize the fi such that fi(vi)>0 for all i; then the conditions become simpler:

  • For all i in 1,...,n, fi(vi)>0, and fi(x)<0 for all points x on the face opposite to vi. In particular, fi(v0)<0.
  • For all points x on the face opposite to v0, fi(x)>0 for at least one i in 1,...,n.

The theorem can be proved based on the Knaster–Kuratowski–Mazurkiewicz lemma. In can be used for approximations of fixed points and zeros.[13]

General metric and topological spaces edit

The intermediate value theorem is closely linked to the topological notion of connectedness and follows from the basic properties of connected sets in metric spaces and connected subsets of R in particular:

  • If   and   are metric spaces,   is a continuous map, and   is a connected subset, then   is connected. (*)
  • A subset   is connected if and only if it satisfies the following property:  . (**)

In fact, connectedness is a topological property and (*) generalizes to topological spaces: If   and   are topological spaces,   is a continuous map, and   is a connected space, then   is connected. The preservation of connectedness under continuous maps can be thought of as a generalization of the intermediate value theorem, a property of real valued functions of a real variable, to continuous functions in general spaces.

Recall the first version of the intermediate value theorem, stated previously:

Intermediate value theorem (Version I) — Consider a closed interval   in the real numbers   and a continuous function  . Then, if   is a real number such that  , there exists   such that  .

The intermediate value theorem is an immediate consequence of these two properties of connectedness:[14]

Proof

By (**),   is a connected set. It follows from (*) that the image,  , is also connected. For convenience, assume that  . Then once more invoking (**),   implies that  , or   for some  . Since  ,   must actually hold, and the desired conclusion follows. The same argument applies if  , so we are done. Q.E.D.

The intermediate value theorem generalizes in a natural way: Suppose that X is a connected topological space and (Y, <) is a totally ordered set equipped with the order topology, and let f : XY be a continuous map. If a and b are two points in X and u is a point in Y lying between f(a) and f(b) with respect to <, then there exists c in X such that f(c) = u. The original theorem is recovered by noting that R is connected and that its natural topology is the order topology.

The Brouwer fixed-point theorem is a related theorem that, in one dimension, gives a special case of the intermediate value theorem.

In constructive mathematics edit

In constructive mathematics, the intermediate value theorem is not true. Instead, one has to weaken the conclusion:

  • Let   and   be real numbers and   be a pointwise continuous function from the closed interval   to the real line, and suppose that   and  . Then for every positive number   there exists a point   in the unit interval such that  .[15]

Practical applications edit

A similar result is the Borsuk–Ulam theorem, which says that a continuous map from the  -sphere to Euclidean  -space will always map some pair of antipodal points to the same place.

Proof for 1-dimensional case

Take   to be any continuous function on a circle. Draw a line through the center of the circle, intersecting it at two opposite points   and  . Define   to be  . If the line is rotated 180 degrees, the value d will be obtained instead. Due to the intermediate value theorem there must be some intermediate rotation angle for which d = 0, and as a consequence f(A) = f(B) at this angle.

In general, for any continuous function whose domain is some closed convex  -dimensional shape and any point inside the shape (not necessarily its center), there exist two antipodal points with respect to the given point whose functional value is the same.

The theorem also underpins the explanation of why rotating a wobbly table will bring it to stability (subject to certain easily met constraints).[16]

See also edit

  • Mean value theorem – On the existence of a tangent to an arc parallel to the line through its endpoints
  • Non-atomic measure – A measurable set with positive measure that contains no subset of smaller positive measure
  • Hairy ball theorem – Theorem in differential topology
  • Sperner's lemma – Theorem on triangulation graph colorings

References edit

  1. ^ Weisstein, Eric W. "Bolzano's Theorem". MathWorld.
  2. ^ Cates, Dennis M. (2019). Cauchy's Calcul Infinitésimal. p. 249. doi:10.1007/978-3-030-11036-9. ISBN 978-3-030-11035-2. S2CID 132587955.
  3. ^ Essentially follows Clarke, Douglas A. (1971). Foundations of Analysis. Appleton-Century-Crofts. p. 284.
  4. ^ Slightly modified version of Abbot, Stephen (2015). Understanding Analysis. Springer. p. 123.
  5. ^ Sanders, Sam (2017). "Nonstandard Analysis and Constructivism!". arXiv:1704.00281 [math.LO].
  6. ^ Bos, Henk J. M. (2001). "The legitimation of geometrical procedures before 1590". Redefining Geometrical Exactness: Descartes' Transformation of the Early Modern Concept of Construction. Sources and Studies in the History of Mathematics and Physical Sciences. New York: Springer. pp. 23–36. doi:10.1007/978-1-4613-0087-8_2. MR 1800805.
  7. ^ Russ, S.B. (1980). "A translation of Bolzano's paper on the intermediate value theorem". Historia Mathematica. 7 (2): 156–185. doi:10.1016/0315-0860(80)90036-1.
  8. ^ Grabiner, Judith V. (March 1983). "Who Gave You the Epsilon? Cauchy and the Origins of Rigorous Calculus" (PDF). The American Mathematical Monthly. 90 (3): 185–194. doi:10.2307/2975545. JSTOR 2975545.
  9. ^ Karin Usadi Katz and Mikhail G. Katz (2011) A Burgessian Critique of Nominalistic Tendencies in Contemporary Mathematics and its Historiography. Foundations of Science. doi:10.1007/s10699-011-9223-1 See link
  10. ^ O'Connor, John J.; Robertson, Edmund F., "Intermediate value theorem", MacTutor History of Mathematics Archive, University of St Andrews
  11. ^ Smorynski, Craig (2017-04-07). MVT: A Most Valuable Theorem. Springer. ISBN 9783319529561.
  12. ^ Vrahatis, Michael N. (2016-04-01). "Generalization of the Bolzano theorem for simplices". Topology and Its Applications. 202: 40–46. doi:10.1016/j.topol.2015.12.066. ISSN 0166-8641.
  13. ^ Vrahatis, Michael N. (2020-04-15). "Intermediate value theorem for simplices for simplicial approximation of fixed points and zeros". Topology and Its Applications. 275: 107036. doi:10.1016/j.topol.2019.107036. ISSN 0166-8641.
  14. ^ Rudin, Walter (1976). Principles of Mathematical Analysis. New York: McGraw-Hill. pp. 42, 93. ISBN 978-0-07-054235-8.
  15. ^ Matthew Frank (July 14, 2020). "Interpolating Between Choices for the Approximate Intermediate Value Theorem". Logical Methods in Computer Science. 16 (3). arXiv:1701.02227. doi:10.23638/LMCS-16(3:5)2020.
  16. ^ Keith Devlin (2007)

External links edit

intermediate, value, theorem, mathematical, analysis, intermediate, value, theorem, states, that, displaystyle, continuous, function, whose, domain, contains, interval, then, takes, given, value, between, displaystyle, displaystyle, some, point, within, interv. In mathematical analysis the intermediate value theorem states that if f displaystyle f is a continuous function whose domain contains the interval a b then it takes on any given value between f a displaystyle f a and f b displaystyle f b at some point within the interval Intermediate value theorem Let f displaystyle f be a continuous function defined on a b displaystyle a b and let s displaystyle s be a number with f a lt s lt f b displaystyle f a lt s lt f b Then there exists some x displaystyle x between a displaystyle a and b displaystyle b such that f x s displaystyle f x s This has two important corollaries If a continuous function has values of opposite sign inside an interval then it has a root in that interval Bolzano s theorem 1 2 The image of a continuous function over an interval is itself an interval Contents 1 Motivation 2 Theorem 3 Relation to completeness 4 Proof 4 1 Proof version A 4 2 Proof version B 5 History 6 Converse is false 7 Generalizations 7 1 Multi dimensional spaces 7 2 General metric and topological spaces 8 In constructive mathematics 9 Practical applications 10 See also 11 References 12 External linksMotivation edit nbsp The intermediate value theoremThis captures an intuitive property of continuous functions over the real numbers given f displaystyle f nbsp continuous on 1 2 displaystyle 1 2 nbsp with the known values f 1 3 displaystyle f 1 3 nbsp and f 2 5 displaystyle f 2 5 nbsp then the graph of y f x displaystyle y f x nbsp must pass through the horizontal line y 4 displaystyle y 4 nbsp while x displaystyle x nbsp moves from 1 displaystyle 1 nbsp to 2 displaystyle 2 nbsp It represents the idea that the graph of a continuous function on a closed interval can be drawn without lifting a pencil from the paper Theorem editThe intermediate value theorem states the following Consider an interval I a b displaystyle I a b nbsp of real numbers R displaystyle mathbb R nbsp and a continuous function f I R displaystyle f colon I to mathbb R nbsp Then Version I if u displaystyle u nbsp is a number between f a displaystyle f a nbsp and f b displaystyle f b nbsp that is min f a f b lt u lt max f a f b displaystyle min f a f b lt u lt max f a f b nbsp then there is a c a b displaystyle c in a b nbsp such that f c u displaystyle f c u nbsp Version II the image set f I displaystyle f I nbsp is also an interval closed and it contains min f a f b max f a f b displaystyle bigl min f a f b max f a f b bigr nbsp Remark Version II states that the set of function values has no gap For any two function values c d f I displaystyle c d in f I nbsp with c lt d displaystyle c lt d nbsp even if they are outside the interval between f a displaystyle f a nbsp and f b displaystyle f b nbsp all points in the interval c d displaystyle bigl c d bigr nbsp are also function values c d f I displaystyle bigl c d bigr subseteq f I nbsp A subset of the real numbers with no internal gap is an interval Version I is naturally contained in Version II Relation to completeness editThe theorem depends on and is equivalent to the completeness of the real numbers The intermediate value theorem does not apply to the rational numbers Q because gaps exist between rational numbers irrational numbers fill those gaps For example the function f x x2 displaystyle f x x 2 nbsp for x Q displaystyle x in mathbb Q nbsp satisfies f 0 0 displaystyle f 0 0 nbsp and f 2 4 displaystyle f 2 4 nbsp However there is no rational number x displaystyle x nbsp such that f x 2 displaystyle f x 2 nbsp because 2 displaystyle sqrt 2 nbsp is an irrational number Proof editProof version A edit The theorem may be proven as a consequence of the completeness property of the real numbers as follows 3 We shall prove the first case f a lt u lt f b displaystyle f a lt u lt f b nbsp The second case is similar Let S displaystyle S nbsp be the set of all x a b displaystyle x in a b nbsp such that f x lt u displaystyle f x lt u nbsp Then S displaystyle S nbsp is non empty since a displaystyle a nbsp is an element of S displaystyle S nbsp Since S displaystyle S nbsp is non empty and bounded above by b displaystyle b nbsp by completeness the supremum c supS displaystyle c sup S nbsp exists That is c displaystyle c nbsp is the smallest number that is greater than or equal to every member of S displaystyle S nbsp Note that due to the continuity of f displaystyle f nbsp at a displaystyle a nbsp we can keep f x displaystyle f x nbsp within any e gt 0 displaystyle varepsilon gt 0 nbsp of f a displaystyle f a nbsp by keeping x displaystyle x nbsp sufficiently close to a displaystyle a nbsp Since f a lt u displaystyle f a lt u nbsp is a strict inequality consider the implication when e displaystyle varepsilon nbsp is the distance between u displaystyle u nbsp and f a displaystyle f a nbsp No x displaystyle x nbsp sufficiently close to a displaystyle a nbsp can then make f x displaystyle f x nbsp greater than or equal to u displaystyle u nbsp which means there are values greater than a displaystyle a nbsp in S displaystyle S nbsp A more detailed proof goes like this Choose e u f a gt 0 displaystyle varepsilon u f a gt 0 nbsp Then d gt 0 displaystyle exists delta gt 0 nbsp such that x a b displaystyle forall x in a b nbsp x a lt d f x f a lt u f a f x lt u displaystyle x a lt delta implies f x f a lt u f a implies f x lt u nbsp Consider the interval a min a d b I1 displaystyle a min a delta b I 1 nbsp Notice that I1 a b displaystyle I 1 subseteq a b nbsp and every x I1 displaystyle x in I 1 nbsp satisfies the condition x a lt d displaystyle x a lt delta nbsp Therefore for every x I1 displaystyle x in I 1 nbsp we have f x lt u displaystyle f x lt u nbsp Hence c displaystyle c nbsp cannot be a displaystyle a nbsp Likewise due to the continuity of f displaystyle f nbsp at b displaystyle b nbsp we can keep f x displaystyle f x nbsp within any e gt 0 displaystyle varepsilon gt 0 nbsp of f b displaystyle f b nbsp by keeping x displaystyle x nbsp sufficiently close to b displaystyle b nbsp Since u lt f b displaystyle u lt f b nbsp is a strict inequality consider the similar implication when e displaystyle varepsilon nbsp is the distance between u displaystyle u nbsp and f b displaystyle f b nbsp Every x displaystyle x nbsp sufficiently close to b displaystyle b nbsp must then make f x displaystyle f x nbsp greater than u displaystyle u nbsp which means there are values smaller than b displaystyle b nbsp that are upper bounds of S displaystyle S nbsp A more detailed proof goes like this Choose e f b u gt 0 displaystyle varepsilon f b u gt 0 nbsp Then d gt 0 displaystyle exists delta gt 0 nbsp such that x a b displaystyle forall x in a b nbsp x b lt d f x f b lt f b u f x gt u displaystyle x b lt delta implies f x f b lt f b u implies f x gt u nbsp Consider the interval max a b d b I2 displaystyle max a b delta b I 2 nbsp Notice that I2 a b displaystyle I 2 subseteq a b nbsp and every x I2 displaystyle x in I 2 nbsp satisfies the condition x b lt d displaystyle x b lt delta nbsp Therefore for every x I2 displaystyle x in I 2 nbsp we have f x gt u displaystyle f x gt u nbsp Hence c displaystyle c nbsp cannot be b displaystyle b nbsp With c a displaystyle c neq a nbsp and c b displaystyle c neq b nbsp it must be the case c a b displaystyle c in a b nbsp Now we claim that f c u displaystyle f c u nbsp Fix some e gt 0 displaystyle varepsilon gt 0 nbsp Since f displaystyle f nbsp is continuous at c displaystyle c nbsp d1 gt 0 displaystyle exists delta 1 gt 0 nbsp such that x a b displaystyle forall x in a b nbsp x c lt d1 f x f c lt e displaystyle x c lt delta 1 implies f x f c lt varepsilon nbsp Since c a b displaystyle c in a b nbsp and a b displaystyle a b nbsp is open d2 gt 0 displaystyle exists delta 2 gt 0 nbsp such that c d2 c d2 a b displaystyle c delta 2 c delta 2 subseteq a b nbsp Set d min d1 d2 displaystyle delta min delta 1 delta 2 nbsp Then we havef x e lt f c lt f x e displaystyle f x varepsilon lt f c lt f x varepsilon nbsp for all x c d c d displaystyle x in c delta c delta nbsp By the properties of the supremum there exists some a c d c displaystyle a in c delta c nbsp that is contained in S displaystyle S nbsp and so f c lt f a e lt u e displaystyle f c lt f a varepsilon lt u varepsilon nbsp Picking a c c d displaystyle a in c c delta nbsp we know that a S displaystyle a not in S nbsp because c displaystyle c nbsp is the supremum of S displaystyle S nbsp This means that f c gt f a e u e displaystyle f c gt f a varepsilon geq u varepsilon nbsp Both inequalities u e lt f c lt u e displaystyle u varepsilon lt f c lt u varepsilon nbsp are valid for all e gt 0 displaystyle varepsilon gt 0 nbsp from which we deduce f c u displaystyle f c u nbsp as the only possible value as stated Proof version B edit We will only prove the case of f a lt u lt f b displaystyle f a lt u lt f b nbsp as the f a gt u gt f b displaystyle f a gt u gt f b nbsp case is similar 4 Define g x f x u displaystyle g x f x u nbsp which is equivalent to f x g x u displaystyle f x g x u nbsp and lets us rewrite f a lt u lt f b displaystyle f a lt u lt f b nbsp as g a lt 0 lt g b displaystyle g a lt 0 lt g b nbsp and we have to prove that g c 0 displaystyle g c 0 nbsp for some c a b displaystyle c in a b nbsp which is more intuitive We further define the set S x a b g x 0 displaystyle S x in a b g x leq 0 nbsp Because g a lt 0 displaystyle g a lt 0 nbsp we know that a S displaystyle a in S nbsp so that S displaystyle S nbsp is not empty Moreover as S a b displaystyle S subseteq a b nbsp we know that S displaystyle S nbsp is bounded and non empty so by Completeness the supremum c sup S displaystyle c sup S nbsp exists There are 3 cases for the value of g c displaystyle g c nbsp those being g c lt 0 g c gt 0 displaystyle g c lt 0 g c gt 0 nbsp and g c 0 displaystyle g c 0 nbsp For contradiction let us assume that g c lt 0 displaystyle g c lt 0 nbsp Then by the definition of continuity for ϵ 0 g c displaystyle epsilon 0 g c nbsp there exists a d gt 0 displaystyle delta gt 0 nbsp such that x c d c d displaystyle x in c delta c delta nbsp implies that g x g c lt g c displaystyle g x g c lt g c nbsp which is equivalent to g x lt 0 displaystyle g x lt 0 nbsp If we just chose x c dN displaystyle x c frac delta N nbsp where N gt db c displaystyle N gt frac delta b c nbsp then g x lt 0 displaystyle g x lt 0 nbsp and c lt x lt b displaystyle c lt x lt b nbsp so x S displaystyle x in S nbsp It follows that x displaystyle x nbsp is an upper bound for S displaystyle S nbsp However x gt c displaystyle x gt c nbsp contradicting the upper bound property of the least upper bound c displaystyle c nbsp so g c 0 displaystyle g c geq 0 nbsp Assume then that g c gt 0 displaystyle g c gt 0 nbsp We similarly chose ϵ g c 0 displaystyle epsilon g c 0 nbsp and know that there exists a d gt 0 displaystyle delta gt 0 nbsp such that x c d c d displaystyle x in c delta c delta nbsp implies g x g c lt g c displaystyle g x g c lt g c nbsp We can rewrite this as g c lt g x g c lt g c displaystyle g c lt g x g c lt g c nbsp which implies that g x gt 0 displaystyle g x gt 0 nbsp If we now chose x c d2 displaystyle x c frac delta 2 nbsp then g x gt 0 displaystyle g x gt 0 nbsp and a lt x lt c displaystyle a lt x lt c nbsp It follows that x displaystyle x nbsp is an upper bound for S displaystyle S nbsp However x lt c displaystyle x lt c nbsp which contradict the least property of the least upper bound c displaystyle c nbsp which means that g c gt 0 displaystyle g c gt 0 nbsp is impossible If we combine both results we get that g c 0 displaystyle g c 0 nbsp or f c u displaystyle f c u nbsp is the only remaining possibility Remark The intermediate value theorem can also be proved using the methods of non standard analysis which places intuitive arguments involving infinitesimals on a rigorous clarification needed footing 5 History editA form of the theorem was postulated as early as the 5th century BCE in the work of Bryson of Heraclea on squaring the circle Bryson argued that as circles larger than and smaller than a given square both exist there must exist a circle of equal area 6 The theorem was first proved by Bernard Bolzano in 1817 Bolzano used the following formulation of the theorem 7 Let f f displaystyle f varphi nbsp be continuous functions on the interval between a displaystyle alpha nbsp and b displaystyle beta nbsp such that f a lt f a displaystyle f alpha lt varphi alpha nbsp and f b gt f b displaystyle f beta gt varphi beta nbsp Then there is an x displaystyle x nbsp between a displaystyle alpha nbsp and b displaystyle beta nbsp such that f x f x displaystyle f x varphi x nbsp The equivalence between this formulation and the modern one can be shown by setting f displaystyle varphi nbsp to the appropriate constant function Augustin Louis Cauchy provided the modern formulation and a proof in 1821 8 Both were inspired by the goal of formalizing the analysis of functions and the work of Joseph Louis Lagrange The idea that continuous functions possess the intermediate value property has an earlier origin Simon Stevin proved the intermediate value theorem for polynomials using a cubic as an example by providing an algorithm for constructing the decimal expansion of the solution The algorithm iteratively subdivides the interval into 10 parts producing an additional decimal digit at each step of the iteration 9 Before the formal definition of continuity was given the intermediate value property was given as part of the definition of a continuous function Proponents include Louis Arbogast who assumed the functions to have no jumps satisfy the intermediate value property and have increments whose sizes corresponded to the sizes of the increments of the variable 10 Earlier authors held the result to be intuitively obvious and requiring no proof The insight of Bolzano and Cauchy was to define a general notion of continuity in terms of infinitesimals in Cauchy s case and using real inequalities in Bolzano s case and to provide a proof based on such definitions Converse is false editA Darboux function is a real valued function f that has the intermediate value property i e that satisfies the conclusion of the intermediate value theorem for any two values a and b in the domain of f and any y between f a and f b there is some c between a and b with f c y The intermediate value theorem says that every continuous function is a Darboux function However not every Darboux function is continuous i e the converse of the intermediate value theorem is false As an example take the function f 0 1 1 defined by f x sin 1 x for x gt 0 and f 0 0 This function is not continuous at x 0 because the limit of f x as x tends to 0 does not exist yet the function has the intermediate value property Another more complicated example is given by the Conway base 13 function In fact Darboux s theorem states that all functions that result from the differentiation of some other function on some interval have the intermediate value property even though they need not be continuous Historically this intermediate value property has been suggested as a definition for continuity of real valued functions 11 this definition was not adopted Generalizations editMulti dimensional spaces edit The Poincare Miranda theorem is a generalization of the Intermediate value theorem from a one dimensional interval to a two dimensional rectangle or more generally to an n dimensional cube Vrahatis 12 presents a similar generalization to triangles or more generally n dimensional simplices Let Dn be an n dimensional simplex with n 1 vertices denoted by v0 vn Let F f1 fn be a continuous function from Dn to Rn that never equals 0 on the boundary of Dn Suppose F satisfies the following conditions For all i in 1 n the sign of fi vi is opposite to the sign of fi x for all points x on the face opposite to vi The sign vector of f1 fn on v0 is not equal to the sign vector of f1 fn on all points on the face opposite to v0 Then there is a point z in the interior of Dn on which F z 0 0 It is possible to normalize the fi such that fi vi gt 0 for all i then the conditions become simpler For all i in 1 n fi vi gt 0 and fi x lt 0 for all points x on the face opposite to vi In particular fi v0 lt 0 For all points x on the face opposite to v0 fi x gt 0 for at least one i in 1 n The theorem can be proved based on the Knaster Kuratowski Mazurkiewicz lemma In can be used for approximations of fixed points and zeros 13 General metric and topological spaces edit The intermediate value theorem is closely linked to the topological notion of connectedness and follows from the basic properties of connected sets in metric spaces and connected subsets of R in particular If X displaystyle X nbsp and Y displaystyle Y nbsp are metric spaces f X Y displaystyle f colon X to Y nbsp is a continuous map and E X displaystyle E subset X nbsp is a connected subset then f E displaystyle f E nbsp is connected A subset E R displaystyle E subset mathbb R nbsp is connected if and only if it satisfies the following property x y E x lt r lt y r E displaystyle x y in E x lt r lt y implies r in E nbsp In fact connectedness is a topological property and generalizes to topological spaces If X displaystyle X nbsp and Y displaystyle Y nbsp are topological spaces f X Y displaystyle f colon X to Y nbsp is a continuous map and X displaystyle X nbsp is a connected space then f X displaystyle f X nbsp is connected The preservation of connectedness under continuous maps can be thought of as a generalization of the intermediate value theorem a property of real valued functions of a real variable to continuous functions in general spaces Recall the first version of the intermediate value theorem stated previously Intermediate value theorem Version I Consider a closed interval I a b displaystyle I a b nbsp in the real numbers R displaystyle mathbb R nbsp and a continuous function f I R displaystyle f colon I to mathbb R nbsp Then if u displaystyle u nbsp is a real number such that min f a f b lt u lt max f a f b displaystyle min f a f b lt u lt max f a f b nbsp there exists c a b displaystyle c in a b nbsp such that f c u displaystyle f c u nbsp The intermediate value theorem is an immediate consequence of these two properties of connectedness 14 Proof By I a b displaystyle I a b nbsp is a connected set It follows from that the image f I displaystyle f I nbsp is also connected For convenience assume that f a lt f b displaystyle f a lt f b nbsp Then once more invoking f a lt u lt f b displaystyle f a lt u lt f b nbsp implies that u f I displaystyle u in f I nbsp or f c u displaystyle f c u nbsp for some c I displaystyle c in I nbsp Since u f a f b displaystyle u neq f a f b nbsp c a b displaystyle c in a b nbsp must actually hold and the desired conclusion follows The same argument applies if f b lt f a displaystyle f b lt f a nbsp so we are done Q E D The intermediate value theorem generalizes in a natural way Suppose that X is a connected topological space and Y lt is a totally ordered set equipped with the order topology and let f X Y be a continuous map If a and b are two points in X and u is a point in Y lying between f a and f b with respect to lt then there exists c in X such that f c u The original theorem is recovered by noting that R is connected and that its natural topology is the order topology The Brouwer fixed point theorem is a related theorem that in one dimension gives a special case of the intermediate value theorem In constructive mathematics editIn constructive mathematics the intermediate value theorem is not true Instead one has to weaken the conclusion Let a displaystyle a nbsp and b displaystyle b nbsp be real numbers and f a b R displaystyle f a b to R nbsp be a pointwise continuous function from the closed interval a b displaystyle a b nbsp to the real line and suppose that f a lt 0 displaystyle f a lt 0 nbsp and 0 lt f b displaystyle 0 lt f b nbsp Then for every positive number e gt 0 displaystyle varepsilon gt 0 nbsp there exists a point x displaystyle x nbsp in the unit interval such that f x lt e displaystyle vert f x vert lt varepsilon nbsp 15 Practical applications editA similar result is the Borsuk Ulam theorem which says that a continuous map from the n displaystyle n nbsp sphere to Euclidean n displaystyle n nbsp space will always map some pair of antipodal points to the same place Proof for 1 dimensional case Take f displaystyle f nbsp to be any continuous function on a circle Draw a line through the center of the circle intersecting it at two opposite points A displaystyle A nbsp and B displaystyle B nbsp Define d displaystyle d nbsp to be f A f B displaystyle f A f B nbsp If the line is rotated 180 degrees the value d will be obtained instead Due to the intermediate value theorem there must be some intermediate rotation angle for which d 0 and as a consequence f A f B at this angle In general for any continuous function whose domain is some closed convex n displaystyle n nbsp dimensional shape and any point inside the shape not necessarily its center there exist two antipodal points with respect to the given point whose functional value is the same The theorem also underpins the explanation of why rotating a wobbly table will bring it to stability subject to certain easily met constraints 16 See also editMean value theorem On the existence of a tangent to an arc parallel to the line through its endpoints Non atomic measure A measurable set with positive measure that contains no subset of smaller positive measurePages displaying short descriptions of redirect targets Hairy ball theorem Theorem in differential topology Sperner s lemma Theorem on triangulation graph coloringsReferences edit Weisstein Eric W Bolzano s Theorem MathWorld Cates Dennis M 2019 Cauchy s Calcul Infinitesimal p 249 doi 10 1007 978 3 030 11036 9 ISBN 978 3 030 11035 2 S2CID 132587955 Essentially follows Clarke Douglas A 1971 Foundations of Analysis Appleton Century Crofts p 284 Slightly modified version of Abbot Stephen 2015 Understanding Analysis Springer p 123 Sanders Sam 2017 Nonstandard Analysis and Constructivism arXiv 1704 00281 math LO Bos Henk J M 2001 The legitimation of geometrical procedures before 1590 Redefining Geometrical Exactness Descartes Transformation of the Early Modern Concept of Construction Sources and Studies in the History of Mathematics and Physical Sciences New York Springer pp 23 36 doi 10 1007 978 1 4613 0087 8 2 MR 1800805 Russ S B 1980 A translation of Bolzano s paper on the intermediate value theorem Historia Mathematica 7 2 156 185 doi 10 1016 0315 0860 80 90036 1 Grabiner Judith V March 1983 Who Gave You the Epsilon Cauchy and the Origins of Rigorous Calculus PDF The American Mathematical Monthly 90 3 185 194 doi 10 2307 2975545 JSTOR 2975545 Karin Usadi Katz and Mikhail G Katz 2011 A Burgessian Critique of Nominalistic Tendencies in Contemporary Mathematics and its Historiography Foundations of Science doi 10 1007 s10699 011 9223 1 See link O Connor John J Robertson Edmund F Intermediate value theorem MacTutor History of Mathematics Archive University of St Andrews Smorynski Craig 2017 04 07 MVT A Most Valuable Theorem Springer ISBN 9783319529561 Vrahatis Michael N 2016 04 01 Generalization of the Bolzano theorem for simplices Topology and Its Applications 202 40 46 doi 10 1016 j topol 2015 12 066 ISSN 0166 8641 Vrahatis Michael N 2020 04 15 Intermediate value theorem for simplices for simplicial approximation of fixed points and zeros Topology and Its Applications 275 107036 doi 10 1016 j topol 2019 107036 ISSN 0166 8641 Rudin Walter 1976 Principles of Mathematical Analysis New York McGraw Hill pp 42 93 ISBN 978 0 07 054235 8 Matthew Frank July 14 2020 Interpolating Between Choices for the Approximate Intermediate Value Theorem Logical Methods in Computer Science 16 3 arXiv 1701 02227 doi 10 23638 LMCS 16 3 5 2020 Keith Devlin 2007 How to stabilize a wobbly tableExternal links editIntermediate value theorem at ProofWiki Intermediate value Theorem Bolzano Theorem at cut the knot Bolzano s Theorem by Julio Cesar de la Yncera Wolfram Demonstrations Project Weisstein Eric W Intermediate Value Theorem MathWorld Belk Jim January 2 2012 Two dimensional version of the Intermediate Value Theorem Stack Exchange Mizar system proof http mizar org version current html topreal5 html T4 Retrieved from https en wikipedia org w index php title Intermediate value theorem amp oldid 1209685064, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.