2.6. Real Functions#

In this section, we will deal with functions of type f:RR. Main references are [2, 76].

Our goal is a cursory review of the relevant theory. We will state the main definitions and results and provide a few examples wherever needed. Detailed proofs will be skipped.

Definition 2.62 (Real function)

A (partial) function of type f:RR mapping real values to real values is called a real function.

The domain and range of a real function are both subsets of R.

Definition 2.63 (Arithmetic operators)

Let f,g be real functions with domfdomg.

We define f+g as:

(f+g)(x)=f(x)+g(x)xdomfdomg.

We define fg as:

(fg)(x)=f(x)g(x)xdomfdomg.

We define fg as:

(fg)(x)=f(x)g(x)xdomfdomg.

We define the quotient f/g as:

(fg)(x)=f(x)g(x)xdomfdomg such that g(x)0.

For some cR, we define cf as:

(cf)(x)=cf(x)xdomf.

Example 2.10 (Function sum, product, powers)

Let f1,f2,,fn be real functions.

Then, their sum is defined by:

(f1+f2++fn)(x)=f1(x)+f2(x)++fn(x)xi=1ndomfi

and their product is defined by:

(f1f2fn)(x)=f1(x)f2(x)fn(x)xi=1ndomfi

provided the domain D=i=1ndomfi is nonempty.

If f=f1=f2==fn, then the n-th power of f is defined as:

(fn)(x)=(f(x))nxdomf.

2.6.1. Limits#

Definition 2.64 (Limit)

We say that f approaches the limit a as x approaches x0 and write:

limxx0f(x)=a

if f is defined on some deleted neighborhood of x0 and for every ϵ>0, there is δ>0 such that:

|f(x)a|<ϵ whenever 0<|xx0|<δ.
  • f may or may not be defined at x=x0. But, it must be defined in the neighborhood around x0.

  • If f is defined at x=x0, then f(x0) doesn’t need to be equal to a.

Example 2.11 (Limit outside domain points)

Let

f(x)=xsin1x,x0.

We have domf=R{0}=(,0)(0,). We can compute the limit of f at x=0. We will show that:

limx0f(x)=0.

Assume, ϵ>0 and let δ=ϵ.

Then, for any x satisfying 0<|x0|<δ, we have

|f(x)0|=|xsin1x|=|x||sin1x||x|<δ=ϵ.

Thus, for every ϵ>0, there exists δ>0, given by δ=ϵ, such that:

0<|x0|<δ|f(x)0|<ϵ.

Thus,

limx0f(x)=0.

Theorem 2.35 (Limit uniqueness)

If limxx0f(x) exists, then it is unique.

Proof. Assume that the limit exists and assume that:

limxx0f(x)=a1 and limxx0f(x)=a2

hold true. We will show that a1=a2 must be true.

Let ϵ>0. Since, the limit exists, hence there are positive numbers δ1 and δ2 such that:

|f(x)ai|<ϵ whenever 0<|xx0|<δi,i=1,2.

Choose δ=min(δ1,δ2).

Then:

|a1a2||f(x)a1|+|f(x)a2|2ϵ whenever 0<|xx0|<δ.

Thus, |a1a2|<2ϵ. Since this is valid for every ϵ>0, hence a1=a2 must be true (Proposition 2.5).

Theorem 2.36 (Arithmetic of limits)

Let limxx0f(x)=a and limxx0g(x)=b.

Then, we have the following rules.

Addition of limits:

limxx0(f+g)(x)=a+b.

Subtraction of limits:

limxx0(fg)(x)=ab.

Multiplication of limits:

limxx0(fg)(x)=ab.

Division of limits. If b0, then:

limxx0(fg)(x)=ab.

Definition 2.65 (One sided limits)

We that that f approaches the left hand limit p as x approaches a from the left and write:

limxaf(x)=p

if f is defined on some open interval (ab,a) (with b>0) and for each ϵ>0, there is a δ>0 such that

|f(x)p|<ϵ whenever aδ<x<a.

We that that f approaches the right hand limit p as x approaches a from the right and write:

limxa+f(x)=p

if f is defined on some open interval (a,a+b) (with b>0) and for each ϵ>0, there is a δ>0 such that

|f(x)p|<ϵ whenever a<x<a+δ.

The left and right hand limits are called one sided limits.

We often write simplify the notation as:

limxaf(x)=f(a) and limxa+f(x)=f(a+).

Theorem 2.37

A function f has a limit at x=a if and only if it has left and right hand limits at x=a and they are equal.

limxaf(x)=pf(a)=f(a+)=p.

Definition 2.66 (Limits at infinities)

We say that f approaches the limit p as x approaches and write:

limxf(x)=p

if f is defined on an interval (a,) and for each ϵ>0, there exists a number b such that

|f(x)p|<ϵ whenever x>b.

We say that f approaches the limit p as x approaches and write:

limxf(x)=p

if f is defined on an interval (,a) and for each ϵ>0, there exists a number b such that

|f(x)p|<ϵ whenever x<b.

We sometimes write:

limxf(x)=f() and limxf(x)=f().

Definition 2.67 (Infinite limits)

We say that f approaches as x approaches a from the left, and write:

limxaf(x)=

if f is defined on an interval (ab,a) (with b>0), and, for each real number M, there exists δ>0 such that

f(x)>M whenever aδ<x<a.

We say that f approaches as x approaches a from the right, and write:

limxa+f(x)=

if f is defined on an interval (a,a+b) (with b>0), and, for each real number M, there exists δ>0 such that

f(x)>M whenever a<x<a+δ.

We say that f approaches as x approaches a from the left, and write:

limxaf(x)=

if f is defined on an interval (ab,a) (with b>0), and, for each real number M, there exists δ>0 such that

f(x)<M whenever aδ<x<a.

We say that f approaches as x approaches a from the right, and write:

limxa+f(x)=

if f is defined on an interval (a,a+b) (with b>0), and, for each real number M, there exists δ>0 such that

f(x)<M whenever a<x<a+δ.

If left hand and right hand limits are equal, we say that the limit at x=a given by limxaf(a) exists and is equal to f(a+) or f(a).

Remark 2.15

Theorem 2.35 can be extended for the following too:

  • If a left hand limit exists, it is unique.

  • If a right hand limit exists, it is unique.

  • If a limit at infinity exists, it is unique.

  • If the limit value is infinite, it is unique.

Theorem 2.36 remains valid for the following too:

  • Left hand limits

  • Right hand limits

  • Limits at infinity

Addition, subtraction and multiplication rules remain valid if either or both limits are infinite, provided that the R.H.S. is not indeterminate. E.g., if limf= and limg=, then lim(fg)=limflimg doesn’t make sense as is indeterminate.

Division rule for limits remains valid if limf/limg is not indeterminate and limg0.

2.6.2. Monotonicity#

Definition 2.68 (Monotonic function)

A function f is increasing on an interval I if

f(x1)f(x2) whenever x1<x2x1,x2I.

A function f is decreasing on an interval I if

f(x1)f(x2) whenever x1<x2x1,x2I.

If f is increasing or decreasing on I, then we say that f is monotonic on I.

A function f is strictly increasing on an interval I if

f(x1)<f(x2) whenever x1<x2x1,x2I.

A function f is strictly decreasing on an interval I if

f(x1)<f(x2) whenever x1<x2x1,x2I.

If f is strictly increasing or strictly decreasing on I, then we say that f is strictly monotonic on I.

Theorem 2.38 (Monotonicity and one sided limits)

Let f be monotonic of an interval (a,b). Let

α=infa<x<bf(x) and β=supa<x<bf(x).
  1. If f is increasing, then f(a+)=α and f(b)=β.

  2. If f is decreasing, then f(a+)=β and f(b)=α.

  3. If a<c<b, then f(c+) and f(c) exist and are finite. Moreover, if f is increasing, then:

    f(c)f(c)f(c+)

    and, if f is decreasing, then:

    f(c)f(c)f(c+).

2.6.3. Continuity#

If the limit of a function at a point matches its value at that point, then the function is continuous at that point.

Definition 2.69 (Continuity)

  1. We say that f is continuous at x=c if f is defined on an interval (a,b) containing c (i.e. a<c<b) and limxcf(x)=f(c).

  2. We say that f is continuous from the left at x=c if f is defined on an interval (a,c] and f(c)=f(c).

  3. We say that f is continuous from the right at x=c if f is defined on an interval [c,b) and f(c+)=f(c).

Theorem 2.39 (Characterization of continuity)

A function f is continuous at x=c if and only if f is defined on an interval (a,b) containing c and for each ϵ>0, there exists δ>0 such that:

|f(x)f(c)|<ϵ whenever |xc|<δ.

A function f is continuous from the left at x=c if and only if f is defined on an interval (a,c] and for each ϵ>0, there exists δ>0 such that:

|f(x)f(c)|<ϵ whenever cδ<xc.

A function f is continuous from the right at x=c if and only if f is defined on an interval [c,b) and for each ϵ>0, there exists δ>0 such that:

|f(x)f(c)|<ϵ whenever cx<c+δ.

f is continuous at x=c if and only if

f(c)=f(c)=f(c+).

This theorem is a restatement of continuity definition in the form of ϵδ rules.

Definition 2.70 (Continuity on an interval)

A function f is continuous on an open interval (a,b) if it is continuous at every point in (a,b).

  • If f(b)=f(b) holds too (i.e. f is continuous from the left at b), then f is continuous on (a,b].

  • If f(a+)=f(a) holds too (i.e. f is continuous from the right at a), then f is continuous on [a,b).

  • If both f(a+)=f(a) and f(b)=f(b) hold true, then f is continuous on the closed interval [a,b].

If S is a subset of domf consisting of finitely or infinitely many disjoint intervals, then f is continuous on S if S is continuous on every interval in S.

When we say that f is continuous on some set Sdomf, we mean that S is a union of finitely or infinitely many disjoint intervals.

Proposition 2.22 (Continuity on a closed interval [a,b])

Let f be continuous on [a,b]. Then, for every t[a,b], and for every ϵ>0, there exists an open interval It=(c,d) containing t such that

|f(x)f(t)|<ϵ whenever xIt[a,b].

Proof. Assume a<t<b. Then, there exists δ>0, such that:

|f(x)f(t)|<ϵ whenever tδ<x<t+δ.

Choose It=(tδ,t+δ).

Now, assume t=a. f must be continuous from the right at t=a. There exists δ>0 such that:

|f(x)f(t)|<ϵ whenever tx<t+δ.

Choose It=(tδ,t+δ). Then It(a,b)=[t,t+δ).

Finally, assume t=b. f must be continuous from the left at t=b. There exists δ>0 such that:

|f(x)f(t)|<ϵ whenever tδ<xt.

Choose It=(tδ,t+δ). Then It(a,b)=(tδ,t].

The set O{Itt[a,b]} forms an open cover of [a,b].

Proposition 2.23

Let f be continuous at x=a

  1. If f(a)>μ, then f(x)>μ in some neighborhood of a.

  2. If f(a)<μ, then f(x)<μ in some neighborhood of a.

Proof. (1) Let ϵ=f(a)μ. ϵ>0 as f(a)>μ. There exists δ>0 such that

|f(x)f(a)|<ϵ whenever |xa|<δ.

Thus, in the neighborhood x(aδ,a+δ):

f(a)ϵ<f(x)<f(a)+ϵμ<f(x)<2f(a)μ.

(2) Let ϵ=μf(a). ϵ>0 as f(a)<μ. There exists δ>0 such that

|f(x)f(a)|<ϵ whenever |xa|<δ.

Thus, in the neighborhood x(aδ,a+δ):

f(a)ϵ<f(x)<f(a)+ϵμ2f(a)<f(x)<μ.

Theorem 2.40 (Continuity and arithmetic)

If f and g are continuous on a set S, then so are f+g, fg, and fg. fg is continuous at each xS such that g(x)0.

2.6.4. Discontinuities#

A function f is discontinuous at some x=c in its domain domf if either limxcf(x) doesn’t exist or limxcf(x)f(c).

Definition 2.71 (Jump discontinuity)

Let f:RR be a real function and let [a,b]domf.

  1. f has a jump discontinuity at a point c(a,b) if both the left hand limit f(c) and the right hand limit f(c+) exist but f(c)f(c+). In this case, limxcf(x) doesn’t exist and f is not continuous at x=c.

  2. f has a jump discontinuity at x=a if the right hand limit f(a+) exists but f(a)f(a+). In this case, f is not continuous from the right at x=a.

  3. f has a jump discontinuity at x=b if the left hand limit f(b) exists but f(b)f(b). In this case, f is not continuous from the left at x=b.

Definition 2.72 (Piecewise continuity)

A function f is piecewise-continuous on [a,b] if

  1. f(x+) exists for all ax<b;

  2. f(x) exists for all a<xb;

  3. f(x+)=f(x)=f(x) for all but finitely many points in (a,b).

Jump discontinuities:

  • If (3) fails to hold at some x=c in (a,b), f has a jump discontinuity at x=c.

  • f has a jump discontinuity at a if f(a+)f(a).

  • f has a jump discontinuity at b if f(b)f(b).

In other words:

  • Left hand limits exist everywhere (except at x=b).

  • Right hand limits exist everywhere (except at x=a).

  • There are only a finite number of points where these two limits don’t match.

  • f is not continuous at those finite number of points.

  • f is continuous everywhere else in the open interval (a,b).

  • f may not be continuous at the boundaries x=a and x=b.

  • At a jump discontinuity, f may be continuous from the right or continuous from the left or neither.

Definition 2.73 (Removable discontinuity)

Let f be discontinuous at some x=a. If limxaf(x) exists, then we say that the discontinuity at x=a is a removable discontinuity. Moreover, the function:

g(x){f(x)xdomf{a}limxaf(x)x=a

is continuous at x=a.

Definition 2.74 (Essential discontinuity)

Let x be an interior point of domf. If either of the one sided limits f(x+), or f(x) don’t exist, then f has an essential discontinuity at x.

If xdomf is a boundary point, then only a one sided limit is possible (left or right hand limit). If such a limit doesn’t exist, then f has an essential discontinuity at the boundary point x.

2.6.5. Continuity with Function Composition#

Next, we look at continuity w.r.t. function composition.

Theorem 2.41

Suppose f is continuous at x=a; f(a) is an interior point of domg and g is continuous at f(a). Then gf is continuous at x=a.

Proof. We proceed as follows:

  1. Let ϵ>0 be arbitrary.

  2. Since g is continuous at f(a), there is δ1>0 such that

    |g(t)g(f(a))|<ϵ whenever |tf(a)|<δ1.
  3. Since f is continuous at a, hence, there exists δ>0 such that

    |f(x)f(a)|<δ1 whenever |xa|<δ.
  4. Together, they imply that

    |g(f(x))g(f(a))|<ϵ whenever |xa|<δ.
  5. Therefore, gf is continuous at x=a.

2.6.6. Boundedness#

Definition 2.75 (Bounded function)

A real function f is bounded below on a set Sdomf if there is a real number m such that

f(x)mxS.

Then, the set

V={f(x)|xS}

has a infimum (due to Corollary 2.2), and we write:

α=infxSf(x).

If there is a point cS such that α=f(c), we say that α is the minimum of f on S and f attains the minimum at x=c.

A real function f is bounded above on a set Sdomf if there is a real number M such that

f(x)MxS.

Then, the set V has a supremum (due to Axiom 2.1), and we write:

β=supxSf(x).

If there is a point dS such that β=f(d), we say that β is the maximum of f on S and f attains the maximum at x=d.

If f is bounded above and below on a set S, we say that f is bounded on S.

Theorem 2.42

If f is continuous on a finite closed interval [a,b], then f is bounded on [a,b].

Proof. Let f be continuous on [a,b]. Let t[a,b]. Choose ϵ=1. Then, due to Proposition 2.22, there exists an interval It such that

|f(x)f(t)|<1 whenever xIt[a,b].

The set O{Itt[a,b]} forms an open cover of [a,b].

Due to Heine-Borel theorem, [a,b] has a finite subcover contained in O.

Let t1<t2<<tn index the finite open subcover. We have:

[a,b]i=1n(Iti[a,b]).

Then, for each ti:

|f(x)f(ti)|<1 whenever xIti[a,b].

Therefore,

|f(x)|=|f(x)f(ti)+f(ti)||f(x)f(ti)|+|f(ti)|1+|f(ti)| whenever xIti[a,b].

Thus, for every x[a,b], there exists a ti[a,b] such that,

|f(x)|1+|f(ti)|.

Taking the maximum on the R.H.S. over all the inequalities, we get:

|f(x)1+max1in|f(ti)|.

Thus, f is bounded with a bound:

M=1+max1in|f(ti)|.

Corollary 2.8

If f is continuous on a finite closed interval [a,b], then f has an infimum and a supremum.

Proof. By Theorem 2.42, f is bounded. Let

V={f(x)|x[a,b]}.

Since f is bounded, hence V is bounded, hence V has an infimum as well as a supremum.

Theorem 2.43

Let f be continuous on a finite closed interval [a,b]. Let

α=infaxbf(x) and β=supaxbf(x).

Then, α and β are respectively, the minimum and maximum values of f on [a,b] and f attains these values at some points in [a,b].

I.e., there exists x1,x2[a,b] such that:

f(x1)=α and f(x2)=β.

Proof. Assume that α is the infimum of f over [a,b] and there is no x1[a,b] such that f(x1)=α. Then f(x)>α for all x[a,b].

Let t[a,b]. Then, f(t)>α. Thus,

f(t)>f(t)+α2>α.

By Proposition 2.22 and Proposition 2.23, there is a open interval It at t such that

f(x)>f(t)+α2 whenever xIt[a,b].

The set O={It|atb} is an open cover of [a,b].

Due to Heine-Borel theorem, [a,b] has a finite subcover contained in O. Thus, there are finitely many points t1,t2,,tn such that:

[a,b]i=1n(Iti[a,b]).

Define:

α1=min1inf(ti)+α2.

Due to the finite cover; for every x[a,b], there exists ti such that xIti[a,b] and thus, f(x)>f(ti)+α2α1. Thus,

f(x)>α1x[a,b].

But α1>α. Thus, α cannot be the infimum of f over [a,b]. We arrive at the contradiction.

Thus, there exists some x=x1 such that f(x1)=α.

A similar argument shows that f attains β at some x2[a,b].

Theorem 2.44 (Intermediate value theorem)

Let f be continuous on a finite closed interval [a,b]. Assume that f(a)f(b). Let μ be in between f(a) and f(b). Then f attains the value of μ at some x=c(a,b).

In other words, there exists c(a,b) such that f(c)=μ.

Proof. Let us assume that f(a)<μ<f(b).

Define the set:

A={x[a,b]|f(x)μ}.
  1. The set is bounded since axb.

  2. The set is nonempty since f(a)<μ. And since f is continuous from the right at a, hence there exists an interval [a,a+δ) such that f(x)<μ in this interval.

  3. Thus, A has an infimum(obviously a) and a supremum.

  4. Let c=supA.

  5. We will claim that f(c)=μ.

If f(c)>μ, then:

  1. c>a.

  2. f is continuous at x=c.

  3. Thus, there exists an ϵ>0 such that f(x)>μ whenever cϵ<xc.

  4. Therefore, the interval (cϵ,c] is not included in A.

  5. Therefore, cϵ is an upper bound for A.

  6. This contradicts the assumption that c=supA.

If f(c)<μ, then:

  1. c<b.

  2. f is continuous at x=c.

  3. Thus, there exists an ϵ>0 such that f(x)<μ whenever cx<c+ϵ.

  4. Therefore, the interval [c,c+ϵ) is included in A.

  5. Therefore, c is not an upper bound for A.

  6. This also contradicts the assumption that c=supA.

Therefore f(c)=μ must be true.

A similar argument can be pursued when f(b)<μ<f(a).

Note that the proof picks up just one possible value of x such that f(x)=μ. It is quite possible that f attains μ more than one times in the interval [a,b]. The proof doesn’t claim to find all such values of x.

In this sense, this theorem is a weak result as it claims the existence of just one point at which f(x)=μ. It doesn’t claim to characterize the set of points at which f(x)=μ.

2.6.7. Uniform Continuity#

Definition 2.76 (Continuity over a set)

Let Adomf. We say that f is continuous over the set A if for each ϵ>0 and each x0A, there exists δ>0 (depending on x0 and ϵ) such that

|f(x)f(x0)|<ϵ whenever |xx0|<δ and xdomf.

The clause xdomf in the definition is important.

  1. If x0 is an interior point of domf, we can pick an interval (x0δ,x0+δ).

  2. If x0 is a non-interior point of domf, we can pick up a half-open interval (x0δ,x0] or [x0,x0+δ) whichever is applicable.

  3. If x0 is an isolated point, we pick the degenerate interval [x0,x0] with suitable choice of δ.

  4. Thus, on the non-interior points, f is either continuous from the left or right while on the interior points, f is continuous.

The key issue here is that the size of the interval (decided by δ) may depend on both ϵ as well as the point x0. This is not always desirable. Uniform continuity addresses this concern.

Definition 2.77 (Uniform continuity)

Let f:RR be a real function. Let Adomf. We say that f is uniformly continuous over the set A if for every ϵ>0, there exists δ>0 (depending on ϵ) such that

|f(x)f(y)|<ϵ whenever |xy|<δ and x,yA.

Few observations on this definition:

  1. δ depends on ϵ.

  2. δ is independent of the choice of x and y.

  3. δ might depend on the set A. E.g., if A is a bounded set, it may depend on its size.

  4. The definition is restricted to points in A. It doesn’t consider points in domfA.

Remark 2.16

If f is not uniformly continuous on a set A, then there is an ϵ0>0 such that for any δ>0, there are points x,y in A such that:

|xy|<δ but |f(x)f(y)|ϵ0.

While a continuous function may not be uniformly continuous in general, it is so on a compact subset.

Theorem 2.45

If f is continuous on a closed and bounded (compact) interval [a,b], then f is uniformly continuous on [a,b].

Proof. Let ϵ>0 be arbitrary. Since f is continuous on [a,b], for every t[a,b], there exists δt>0 such that

|f(x)f(t)|<ϵ2 whenever |xt|<2δt and x[a,b].

Choose an open interval It=(tδt,t+δt) for every t[a,b].

The collection O={It|t[a,b]} is an open cover for [a,b]. Since [a,b] is closed and bounded (compact), hence due to Heine-Borel theorem, there are finitely many points t1,t2,,tn in [a,b] such that P={It1,It2,,Itn} form a finite open cover of [a,b].

Define:

δ=min{δt1,δt2,,δtn}.

Assume that |xy|<δ and x,y[a,b]. Assume that xItr for some r. This is true since P is a cover for [a,b].

Now, from triangle inequality:

|f(x)f(y)|=|f(x)f(tr)+f(tr)f(y)||f(x)f(tr)|+|f(tr)f(y)|.

Since xItr, hence |xtr|<δtr, hence:

|f(x)f(tr)|<ϵ2.

On the other hand:

|ytr||yx|+|xtr|<δ+δtr2δtr.

Thus,

|f(tr)f(y)|<ϵ2.

Together, we have:

|f(x)f(y)|<ϵ.

We have shown that for any ϵ>0, a δ>0 can be chosen such that, |xy|<δ with x,y[a,b] implies |f(x)f(y)|<ϵ. Thus, f is uniformly continuous over [a,b].

Corollary 2.9

If f is continuous on a set A, then f is uniformly continuous on any finite closed interval contained in A.

2.6.8. Continuity and Monotonic Functions#

Proposition 2.24

If f is monotonic on an interval I, then f is either continuous or has a jump discontinuity at each xI.

Proof. Let I be an open interval (a,b). Then, due to Theorem 2.38, both left hand and right hand limits exist at each point in (a,b).

Consider some c(a,b).

  1. If f(c)=f(c+), then by monotonicity, f(c)=f(c)=f(c+), thus f is continuous at c.

  2. If f(c)f(c+), then we have a jump discontinuity.

Next, the boundary points. Consider x=a.

  1. The right hand limit f(a+) exists due to monotonicity.

  2. If f(a)=f(a+), then f is continuous from the right at x=a.

  3. Otherwise, we have a jump discontinuity at x=a.

Similarly, for x=b:

  1. The left hand limit f(b) exists due to monotonicity.

  2. If f(b)=f(b), then f is continuous from the left at x=b.

  3. Otherwise, we have a jump discontinuity at x=b.

Theorem 2.46

If f is monotonic on [a,b], then f is continuous on [a,b] if and only if its range Rf=f([a,b])={f(x)|axb} is a closed interval with endpoints f(a) and f(b).

In other words, f is continuous on [a,b] if and only if:

f([a,b])=[f(a),f(b)] if f(a)f(b) else [f(b),f(a)].

Proof. If f is constant over [a,b], there is nothing to prove. Hence, we shall restrict our attention to the case where f is non-constant. Then f(a)f(b) since f is monotonic. Without loss of generality, assume that f(a)<f(b) (f is increasing). If f(a)>f(b), we can replace f by f and proceed.

Consider the set

Sf={f((a,b))}={f(x)|a<x<b}.

Due to Theorem 2.38:

Sf[f(a+),f(b)].

Thus,

Rf={f(a)}Sf{f(b)}{f(a)}[f(a+),f(b)]{f(b)}.

If f is continuous on [a,b], then f(a)=f(a+) and f(b)=f(b). Thus, we have: Rf[f(a),f(b)].

Further, due to intermediate value theorem, for every f(a)<μ<f(b), there exists x(a,b), such that μ=f(x). Thus, Rf=[f(a),f(b)].

We now assume that Rf=[f(a),f(b)] and show that f must be continuous.

  1. Since f is increasing, hence f(a)f(a+) and f(b+)f(b).

  2. We also have: [f(a),f(b)]{f(a)}[f(a+),f(b)]{f(b).

  3. If f(a)<f(a+) were true, then the subset relationship above will be invalid. Similar case with f(b)<f(b).

  4. Thus, we both f(a)=f(a+) and f(b)=f(b) must be true.

  5. f is continuous from the right at a and from the left at b.

  6. Also, since f is increasing, hence for any c(a,b), we have f(c)f(c)f(c+).

  7. If f(c)<f(c), then (f(c),f(c)) cannot be part of Rf. Thus, f(c)=f(c)

For strictly monotonic functions which are continuous on an interval [a,b], it is possible to find an inverse function on the same interval.

Theorem 2.47

Let f be a strictly increasing and continuous function on an interval [a,b]. Let f(a)=c and f(b)=d. Then, there is a unique function g defined on [c,d] such that:

g(f(x))=xaxb

and

f(g(y))=ycyd.

Moreover, g is continuous and strictly increasing on [c,d].

Proof. We first show that such a function g can be defined:

  1. Since f is (strictly) monotone, hence due to Theorem 2.46, f([a,b])=[c,d].

  2. In other words, f restricted to [a,b] as f:[a,b][c,d] is total and surjective.

  3. Thus, for each y[c,d], there exists x[a,b] such that y=f(x).

  4. Since f is strictly increasing, hence f(x1)f(x2) for any x1,x2[a,b].

  5. Hence, f:[a,b][c,d] is injective.

  6. Thus, f:[a,b][c,d] is bijective.

  7. Thus, we can introduce an inverse function g:[c,d][a,b] with the rule g(y)=x whenever f(x)=y.

Next, we show that g is strictly increasing.

  1. Let y1,y2[c,d] such that y1<y2.

  2. Let x1,x2[a,b] such that y1=f(x1) and y2=f(x2).

  3. Thus, x1=g(y1) and x2=g(y2).

  4. Since f is strictly increasing, hence f(x1)<f(x2) implies x1<x2.

  5. Thus, y1<y2 implies g(y1)=x1<g(y2)=x2.

  6. Thus, g is strictly increasing.

Finally, notice that g is monotonic and the range g([c,d])=[a,b] is a closed interval with the end points a=g(c) and b=g(d). Thus, due to Theorem 2.46, g is continuous.