Measure theory and Perron-Frobenius operators for continued fractions

Jordan Bell
April 18, 2016

1 The continued fraction transformation

For ξ let [x] be the greatest integer ξ, let R(ξ)=ξ-[ξ], and let ξ=min(R(ξ),1-R(ξ)), the distance from ξ to a nearest integer. Let I=[0,1] and define the continued fraction transformation τ:II by

τ(x)={x-1-[x-1]x00x=0.

It is immediate that for xI, xI if and only if τ(x)I. For x, define a0(x)=[x], and for n1 define an(x)1{} by

an(x)=[1τn-1(x-a0(x))].

For example, let x=1371.

τ(x)=7113-[7113]=7113-5=613.
τ2(x)=136-[136]=136-2=16.
τ3(x)=61-[61]=0.

Then τn(x)=0 for n3. Thus, with x=1371,

a0(x)=0,a1(x)=[7113]=5.
a2(x)=[1τ(x)]=[136]=2,a3(x)=[1τ2(x)]=[61]=6.
a4(x)=[1τ3(x)]=,a5(x)=,.

2 Convergents

For xΩ=I write an=an(x), and define

q-1=0,p-1=1,q0=1,p0=0,

and for n1,

qn=anqn-1+qn-2,pn=anpn-1+pn-2.

Thus

q1=a1q0+q-1=a1,p1=a1p0+p-1=1.

One proves

pnqn-1-pn-1qn=(-1)n+1,n0.

Also,11 1 Marius Iosifescu and Cor Kraaikamp, Metrical Theory of Continued Fractions, p. 9, Proposition 1.1.1.

x=pn+τn(x)pn-1qn+τn(x)qn-1,xΩ,n0.

From this,

x-pnqn=(-1)nτn(x)qn(qn+τn(x)qn-1).

Now,

an+1+τn+1(x)=[1τn(x)]+1τn(x)-[1τn(x)]=1τn(x),

and using this,

τn(x)qn(qn+τn(x)qn-1) =1qn(qn(an+1+τn+1(x))+qn-1)
=1qn(qn+1+τn+1(x)qn).

Thus

1qn(qn+qn-1)<|x-pnqn|<1qnqn+1.

For n1 let

rn(x)=1τn-1(x)=an+τn(x)

and

sn=qn-1qn,yn=1sn

and

un =qn-1-2|x-pn-1qn-1|-1
=1qn-12qn-1(qn-1+τn-1(x)qn-2)τn-1(x)
=qn-1+τn-1(x)qn-2τn-1(x)qn-1
=qn-1(an+τn(x))+qn-2qn-1
=an+τn(x)+qn-2qn-1.

Let s0=0. It is worth noting that

y1yn=q1q0qnqn-1=qnq0=qn.
1sn=qnqn-1=an+qn-2qn-1=an+sn-1.
un=an+τn(x)+qn-2qn-1=rn+sn-1.

3 Measure theory

Suppose that (X,𝒜) is a measurable space and μ,ν are probability measures on 𝒜. Let 𝒟={A𝒜:μ(A)=ν(A)}. First, X𝒟. Second, if A,B𝒟 and AB then

μ(BA)=μ(B)-μ(A)=ν(B)-ν(A)=ν(BA),

so BA𝒟. Third, suppose that An𝒟, n1, and AnA. Because 𝒜 is a σ-algebra, A𝒜, and then, setting A0=,

μ(A)=μ(n1(AnAn-1))=n1(μ(An)-μ(An-1)),

whence μ(A)=ν(A). Therefore 𝒟 is a Dynkin system. Dynkin’s theorem says that if 𝒟 is a Dynkin system and 𝒞𝒟 where 𝒞 is a π-system (nonempty and closed under finite intersections), then σ(𝒞)𝒟.22 2 Charalambos D. Aliprantis and Kim C. Border, Infinite Dimensional Analysis: A Hitchhiker’s Guide, third ed., p. 136, Lemma 4.11.

Suppose now that σ(𝒞)=𝒜, that 𝒞 is closed under finite intersections, and that μ(A)=ν(A) for all A𝒞. Then 𝒞𝒟, so by Dynkin’s theorem, 𝒜=σ(𝒞)𝒟, hence 𝒟=𝒜. That is, for any A𝒜, μ(A)=ν(A), meaning μ=ν.

We shall apply the above with (I,I), I=[0,1]. For

𝒞={(0,u]:0<u1},

it is a fact that σ(𝒞)=I. Therefore if μ and ν are probability measures on I such that μ((0,u])=ν((0,u]) for every 0<u1, then μ=ν.

Let λ be Lebesgue measure on I=[0,1]. Define

dγ(x)=1(1+x)log2dλ(x),

called the Gauss measure. If μ is a Borel probability measure on I, for measurable T:II and for AI let

T*μ(A)=μ(T-1(A)).

T*μ, called the pushforward of μ by T, is itself a Borel probability measure on I. We prove that γ is an invariant measure for τ.33 3 Marius Iosifescu and Cor Kraaikamp, Metrical Theory of Continued Fractions, p. 17, Theorem 1.2.1; Manfred Einsiedler and Thomas Ward, Ergodic Theory with a view towards Number Theory, p. 77, Lemma 3.5.

Theorem 1.

τ*γ=γ.

Proof.

Let 0<u1. For xI, 0<τ(x)u if and only if 0<1x-[1x]u if and only if [1x]<1xu+[1x] if and only if 1u+[1x]x<1[1x]. Then, as 0τ-1((0,u]),

τ-1((0,u])=i1[1u+i,1i).

We calculate

γ(τ-1((0,u])) =i1γ([1u+i,1i))
=i1[1u+i,1i)1(1+x)log2𝑑λ(x)
=1log2i1(log(1+1i)-log(1+1u+i)).

Using

1+1i1+1u+i=1+ui1+ui+1,

this is

γ(τ-1((0,u])) =1log2i1(log(1+ui)-log(1+ui+1))
=1log2i1ui+1ui11+x𝑑λ(x)
=γ((0,u]).

Because γ(τ-1((0,u]))=γ((0,u]) for every 0<u1, it follows that τ*γ=γ. ∎

We remark that for a set X, X0 is a singleton. For i10 let I0(i)=Ω. For n1 and i1n, let

In(i)={ωΩ:ak(x)=ik,1kn}.

For n1 and for i1n, define

[i1,,in]=1i1+1+1in-1+1in.

For xIn(i),

pn(x)qn(x)=[i1,,in],pn-1(x)qn-1(x)=[i1,,in-1].

The following is an expression for the sets In(i).44 4 Marius Iosifescu and Cor Kraaikamp, Metrical Theory of Continued Fractions, p. 18, Theorem 1.2.2.

Theorem 2.

Let n1, i1n, and define

un(i)={pn+pn-1qn+qn-1n oddpnqnn even

and

vn(i)={pnqnn oddpn+pn-1qn+qn-1n even.

Then

In(i)=Ω(un(i),vn(i)).

From the above, if n is odd and i1 then

λ(In(i)) =vn(i)-un(i)
=pnqn-pn+pn-1qn+qn-1
=pnqn-1-pn-1qnqn(qn+qn-1)
=(-1)n+1qn(qn+qn-1)
=1qn(qn+qn-1),

and if n is even then likewise

λ(In(i))=1qn(qn+qn-1).

Kraaikamp and Iosifescu attribute the following to Torsten Brodén, in a 1900 paper.55 5 Marius Iosifescu and Cor Kraaikamp, Metrical Theory of Continued Fractions, p. 21, Corollary 1.2.6.

Theorem 3.

For n1, in, xI,

λ(τn<x|i)=x(sn+1)snx+1.
Proof.

We have

λ(τn<x|i)=λ((τn<x)In(i))λ(In(i)).

Using

ω=pn+τn(ω)pn-1qn+τn(ω)qn-1,ωΩ,n0,

if n is odd then

(τn<x)In(i) ={ωΩ:pn+pn-1qn+qn-1<ω<pnqn,τn(ω)<x}
={ωΩ:pn+xpn-1qn+xqn-1<ω<pnqn}

and if n is even then

(τn<x)In(i)={ωΩ:pnqn<ω<pn+xpn-1qn+xqn-1}.

Therefore if n is odd,

λ((τn<x)In(i)) =pnqn-pn+xpn-1qn+xqn-1
=xpnqn-1-xpn-1qnqn(qn+xqn-1)
=xqn(qn+xqn-1)

and likewise if n is even then

λ((τn<x)In(i))=xqn(qn+xqn-1).

Therefore for n1,

λ(τn<x|i) =xqn(qn+xqn-1)qn(qn+qn-1)
=x(qn+qn-1)qn+xqn-1.

Using sn+1=qn+qn-1qn and snx+1=xqn-1+qnqn,

λ(τn<x|i) =xqn(sn+1)qn(snx+1)
=x(sn+1)snx+1.

For j1 and sI define

Pj(s)=s+1(s+j)(s+j+1).

We now apply Theorem 3 to prove the following.66 6 Marius Iosifescu and Cor Kraaikamp, Metrical Theory of Continued Fractions, p. 22, Proposition 1.2.7.

Theorem 4.

For j1,

λ(a1=j)=1j(j+1).

For n1 and in,

λ(an+1=j|i)=Pj(sn).
Proof.

By Theorem 2,

{ωΩ:a1(ω)=j}=I1(j)=Ω(u1(j),v1(j)).

In this case, q1=j, so u1(j)=p1+p0q1+q0=1+0j+1=1j+1 and v1(j)=p1q1=1j, so

{ωΩ:a1(ω)=j}=Ω(1j+1,1j).

Now,

an+1(ω)=[1τn(ω)]=a1(τn(ω)).

Thus

{ωΩ:an+1(ω)=j}={ωΩ:τn(ω)(1j+1,1j)}.

Then using Theorem 3,

λ(an+1=j|i) =λ(τn<1j|i)-λ(τn<1j+1|i)
=1j(sn+1)sn1j+1-1j+1(sn+1)sn1j+1+1
=sn+1(sn+1)(sn+j+1).

4 Perron-Frobenius operators

For a probability measure μ on I and for fL1(μ) let dμf=fdμ. If τ*μ is absolutely continuous with respect to μ, check that τ*μf is itself absolutely continuous with respect to μ. Then applying the Radon-Nikodym theorem, let

Pμf=d(τ*μf)dμ.

For gL(μ),

IgPμf𝑑μ=Igd(τ*μf)=Igτ𝑑μf=I(gτ)f𝑑μ.

In particular, for g=1A, AI,

I1APμf𝑑μ=I1τ-1(A)f𝑑μ.

For gL(μ),

IgPγ1𝑑γ=Igτ𝑑γ=Igd(τ*γ),

hence Pγ1=1 if and only if τ*γ.

We shall be especially interested in

U=Pγ,

where γ is the Gauss measure on I. We establish almost everywhere an expression for Uf(x).77 7 Marius Iosifescu and Cor Kraaikamp, Metrical Theory of Continued Fractions, p. 59, Proposition 2.1.2.

Theorem 5.

For fL1(γ), for γ-almost all xI,

Uf(x)=i1Pi(x)f(1x+i).
Proof.

Let Ii=(1i+1,1i] and let τi be the restriction of τ:II to Ii. For uIi, i1u<i+1, hence τi(u)=τ(u)=1u-i, i.e. u=1τi(u)+i, i.e. τi-1(x)=1x+i.

For AI, if 0A then

τ-1(A)=τ-1(i1(AIi))=i1τ-1(AIi),

and the sets τ-1(AIi) are pairwise disjoint, hence

τ-1(A)f𝑑γ=i1τ-1(AIi)f𝑑γ=i1τi-1(A)f𝑑γ.

Applying the change of variables formula, as ddxτi-1(x)=-(x+i)-2,

τi-1(A)f𝑑γ =1log2τi-1(A)f(u)u+1𝑑λ(u)
=1log2Afτi-1(x)τi-1(x)+1(x+i)-2𝑑λ(x)
=1log2Af(1x+i)1(x+i+1)(x+i)𝑑λ(x)
=1log2Af(1x+i)Pi(x)1x+1𝑑λ(x)
=Af(1x+i)Pi(x)𝑑γ(x).

Therefore

τ-1(A)f𝑑γ =i1Af(1x+i)Pi(x)𝑑γ(x)
=Ai1f(1x+i)Pi(x)dγ(x).

Then

APγf𝑑γ=Ai1f(1x+i)Pi(x)dγ(x).

Because this is true for any AI with 0A, it follows that for γ-almost all xI,

Pγf(x)=i1f(1x+i)Pi(x).

The following gives an expression for Pμf(x) under some hypotheses.88 8 Marius Iosifescu and Cor Kraaikamp, Metrical Theory of Continued Fractions, p. 60, Proposition 2.1.3.

Theorem 6.

Let μ be a probability measure on I that is absolutely continuous with respect to λ and suppose that dμ=hdλ with h(x)>0 for μ-almost all xI. Let fL1(μ) and define g(x)=(x+1)h(x)f(x). For μ-almost all xI,

Pμf(x)=1h(x)i1h((x+i)-1)(x+i)2f(1x+i)=Ug(x)(x+1)h(x).

For n1, for μ-almost all xI,

Pμnf(x)=Ung(x)(x+1)h(x).

We prove an expression for μ(τ-n(A)).99 9 Marius Iosifescu and Cor Kraaikamp, Metrical Theory of Continued Fractions, p. 61, Proposition 2.1.5.

Theorem 7.

Let μ be a probability measure on I that is absolutely continuous with respect to λ. Let h=dμdλ and let f(x)=(x+1)h(x). For AI and n1,

μ(τ-n(A))=AUnf(x)x+1𝑑λ(x).
Proof.

For n=0,

μ(A)=A𝑑μ=Ah𝑑λ=Af(x)x+1𝑑λ(x)=AU0f(x)x+1𝑑λ(x).

Suppose by hypothesis that the claim is true for some n0. Then

μ(τ-n-1(A)) =μ(τ-n(τ-1(A)))
=τ-1(A)Unf(x)x+1𝑑λ(x)
=log2τ-1(A)Unf(x)𝑑γ(x)
=log2AUn+1f(x)𝑑γ(x)
=log2AUn+1f(x)x+1𝑑λ(x).

For f(x)=1x+1 and AI,

APλf𝑑λ =τ-1(A)1x+1𝑑λ(x)
=log2τ-1(A)𝑑γ
=log2A𝑑γ
=Af𝑑λ.

Because this is true for all Borel sets A,

Pλ1x+1=1x+1.

For fL1(λ) and xI, let

Π1f(x)=1(x+1)log2If𝑑λ.

Define

T0=Pλ-Π1.

For n1, Π1n=Π1. For fL1(λ),

PλΠ1f=1log2If𝑑λPλ1x+1=1log2If𝑑λ1x+1=Π1f(x)

and

Π1Pλf=1(x+1)log2IPλf𝑑λ=1(x+1)log2If𝑑λ=Π1f(x),

hence

PλΠ1=Π1=Π1Pλ.

Moreover,

T0Π1=(Pλ-Π1)Π1=PλΠ1-Π12=0

and

Π1T0=Π1(Pλ-Π1)=Π1Pλ-Π12=0.

Because Pλ=Π1+T0, using Π12=Π1, T0Π1=0, and Π1T0=0, we have

Pλn=Π1+T0n,n1.

Theorem 6 tells us that for fL1(λ), for λ-almost all xI,

Pλf(x)=i11(x+i)2f(1x+i).

With h(x)=x+1 and g=hf, for n1, for λ-almost all xI,

Pλnf(x)=Ung(x)x+1.

Thus

Ung =hPλnf
=hΠ1f+hT0nf
=1log2If𝑑λ+hT0nf
=Ig𝑑γ+hT0n(g/h).

Define Iγ:L1(γ)L1(γ) by

Iγf=1If𝑑γ.

We have

IγUf=IPγf𝑑γ=If𝑑γ=Iγf,

meaning IγU=Iγ. Furthermore, because τ*γ=γ we have Pγ1=1, so

UIγf=If𝑑γU1=If𝑑γ1=Iγf,

meaning UIγ=Iγ.

Let h(x)=x+1. h,1hL(γ). Now define T:L1(γ)L1(γ) by

Tg=hT0(g/h),

which makes sense because 1hL(γ). Then

T2g =T(hT0(g/h))
=hT0(hT0(g/h)h)
=hT02(g/h).

For n1,

Tng=hT0n(g/h).

Recapitulating the above, for n1 and gL1(γ),

Ung=Iγg+hT0n(g/h)=Iγg+Tng,

meaning

Un=Iγ+Tn,n1.

It is a fact that Tn converges to 0 in the strong operator topology on (L1(γ)), the bounded linear operators L1(γ)L1(γ), that is, for each fL1(γ), Tnf0 in L1(γ), i.e. TnfL10.1010 10 Marius Iosifescu and Cor Kraaikamp, Metrical Theory of Continued Fractions, p. 63, Proposition 2.1.7. Then UnIγ in the strong operator topology: for fL1(γ),

I|Unf(x)-If𝑑γ|𝑑λ0.

Iosifescu and Kraaikamp state that has not been determined whether for γ-almost all xI, Unf(x)Iγf.

Let B(I) be the set of bounded Borel measurable functions f:I and write f=supxI|f(x)|. For fB(I), define for xI,

Uf(x)=i1Pi(x)f(1x+i)=i1x+1(x+i)(x+i+1)f(1x+i).

1B(I), and for xI,

1imx+1(x+i)(x+i+1)=mm+x+1,

hence

U1(x)=i1x+1(x+i)(x+i+1)=1.

For fB(I) and xI,

|Uf(x)|fU1(x),

hence

UB(I)B(I)=1.

Say that f:I is increasing if xy implies f(x)f(y). An increasing function f:I belongs to B(I). We prove that if f is increasing then Uf is decreasing.1111 11 Marius Iosifescu and Cor Kraaikamp, Metrical Theory of Continued Fractions, p. 65, Proposition 2.1.11.

Theorem 8.

If f:I is increasing then Uf is decreasing.

Proof.

Take x<y and let

S1=i1Pi(y)(f(1y+i)-f(1x+i))

and

S2=i1(Pi(y)-Pi(x))f(1x+i).

Then

Uf(y)-Uf(x) =i1(Pi(y)f(1y+i)-Pi(x)f(1x+i))
=S1+S2.

Because f is increasing, S10. Using i1Pi(u)=1 for any uI,

i1(Pi(y)-Pi(x))f(1x+1)=0,

and therefore

S2 =i1(f(1x+i)-f(1x+1))(Pi(y)-Pi(x))
=(f(1x+2)-f(1x+1))(P2(y)-P2(x))
+i3(f(1x+i)-f(1x+1))(Pi(y)-Pi(x)).

For i2, using that f is increasing,

f(1x+i)-f(1x+1)f(1x+2)-f(1x+1)0.

We calculate

Pi(u)=--i2+i+(u+1)2(u+i)2(u+i+1)2.

The roots of the above rational function are u=-(i-1)i-1,(i-1)i-1. Thus, Pi(u)=0 if and only if u=(i-1)i-1. But (i-1)i-1I if and only if i2-i-10 and i2-i-40. This is possible if and only if i=2. And

Pi(0)=i2-i-1i2(i+1)2,

so P1(u)0 for all uI and for i3, Pi(u)0 for all uI. For i=2, check that if 0u2-1 then P2(u)0 and if 2-1u1 then P2(u)0. Then

S2 (f(1x+2)-f(1x+1))(P2(y)-P2(x))
+i3(f(1x+2)-f(1x+1))(Pi(y)-Pi(x))
=(f(1x+2)-f(1x+1))(P2(y)-P2(x))
+(f(1x+2)-f(1x+1))(-P1(y)-P2(y)-(-P1(x)-P2(x)))
=(f(1x+2)-f(1x+1))(P1(x)-P1(y))
0.

We have shown that S10 and S20, so

Uf(y)-Uf(x)=S1+S20,

which means that Uf:I is decreasing. ∎

For J=[a,b]I, a partition of J is a sequence P=(t0,,tn) such that a=t0<<tn=b. For f:I define

V(f,P)=1in|f(ti)-f(ti-1)|.

Define

VJf=sup{V(f,P):P is a partition of J}.

Let vf(x)=V[0,x]f, the variation of f. vf(1)=V[0,1]f. We say that f has bounded variation if vf(1)<, and denote by BV(I) the set of functions f:I with bounded variation. It is a fact that with the norm

fBV=|f(0)|+VIf,

BV(I) is a Banach algebra.

If f is increasing then VIf=f(1)-f(0). We will use the following to prove the theorem coming after it.1212 12 Marius Iosifescu and Cor Kraaikamp, Metrical Theory of Continued Fractions, p. 66, Proposition 2.1.12.

Lemma 9.

If f:I is increasing then

VI(Uf)12VIf.
Proof.

Because Uf is decreasing,

VI(Uf)=Uf(0)-Uf(1)=i1(Pi(0)f(1i)-Pi(1)f(11+i)).

As Pi(u)=u+1(u+i)(u+i+1),

Pi(1)=2(i+1)(i+2)=2Pi+1(0),

hence

VI(Uf) =i1(Pi(0)f(1i)-Pi(1)f(11+i))
=i1(Pi(0)f(1i)-Pi+1(0)f(11+i))
-i1Pi+1(0)f(11+i)
=P1(0)f(1)-i1Pi+1(0)f(11+i)
=12f(1)-i1Pi+1(0)f(11+i).

Because f(11+i)f(0) we have -f(11+i)-f(0), hence

VI(Uf)12f(1)-f(0)i1Pi+1(0)=12f(1)-12f(0),

using i1Pi(0)=1 and P1(0)=12. As f is increasing this means

VI(Uf)12(f(1)-f(0))=12VIf.

Theorem 10.

If fBV(I) then

VI(Uf)12VIf.
Proof.

Let

pf(x)=vf(x)+f(x)-f(0)2,nf(x)=vf(x)-f(x)+f(0)2,

the positive variation of f and the negative variation of f. It is a fact that 0pfvf, 0nfvf, and pf and nf are increasing. Using this,

VI(Uf) =VI(Upf+Unf)
12VIpf+12VInf
=12(pf(1)-pf(0))+12(nf(1)-nf(0))
=12(vf(1)-vf(0))
=12VIf.

For f:I, let

s(f)=supx,yI,xy|f(x)-f(y)||x-y|.

We denote by Lip(I) the set of f:I such that s(f)<.1313 13 Marius Iosifescu and Cor Kraaikamp, Metrical Theory of Continued Fractions, p. 67, Proposition 2.1.14.

Theorem 11.

For fLip(I),

s(Uf)(2ζ(3)-ζ(2))s(f).
Proof.

Suppose x,yI, x>y. We calculate

Uf(y)-Uf(x)y-x=1y-xi1(Pi(y)f(1y+i)-Pi(y)f(1x+i))+1y-xi1(Pi(y)f(1x+i)-Pi(x)f(1x+i))=i1Pi(y)f(1y+i)-f(1x+i)y-x+i1Pi(y)-Pi(x)y-xf(1x+i).

Calculating further,

Uf(y)-Uf(x)y-x =-i1Pi(y)f(1y+i)-f(1x+i)1y+i-1x+i1(x+i)(y+i)
+i1Pi(y)-Pi(x)y-xf(1x+i).

Now,

Pi(u)=u+1(u+i)(u+i+1)=iu+i+1-i-1u+i,

whence

Pi(y)-Pi(x)=(x-y)i(x+i+1)(y+i+1)+(y-x)(i-1)(x+i)(y+i),

therefore

i1Pi(y)-Pi(x)y-xf(1x+i)=i1(i-1(x+i)(y+i)-i(x+i+1)(y+i+1))f(1x+i).

Summation by parts tells us

i1fi(gi+1-gi)=-f1g1-i1gi+1(fi+1-fi),

and here this yields, for gi=i-1(x+i)(y+i) and fi=f(1x+i),

i1(i-1(x+i)(y+i)-i(x+i+1)(y+i+1))f(1x+i)=i1gi+1(fi+1-fi)=i1i(x+i+1)(y+i+1)(f(1x+i+1)-f(1x+i))=i1i(x+i+1)(y+i+1)f(1x+i+1)-f(1x+i)1x+i+1-1x+i-1(x+i)(x+i+1).

Recapitulating the above,

Uf(y)-Uf(x)y-x=-i1Pi(y)f(1y+i)-f(1x+i)1y+i-1x+i1(x+i)(y+i)-i1i(x+i)(x+i+1)2(y+i+1)f(1x+i+1)-f(1x+i)1x+i+1-1x+i.

Then

|Uf(y)-Uf(x)y-x| s(f)i1Pi(y)1(x+i)(y+i)
+s(f)i1i(x+i)(x+i+1)2(y+i+1).

Then, using that x>y,

|Uf(y)-Uf(x)y-x|s(f)i1(Pi(y)1(y+i)2+i(y+i)(y+i+1)3).

Because yI=[0,1], y0 so

i1i(y+i)(y+i+1)3i11(i+1)3=-1+ζ(3).

Let h(u)=u2, with which

i1Pi(y)1(y+i)2=Uh(y).

h:I is increasing, so Uh is decreasing. Because Pi(0)=1i(i+1),

i1Pi(y)1(y+i)2=Uh(y)Uh(0)=i1Pi(0)1i2=i11i3(i+1).

Doing partial fractions,

1i3(i+1)=1i3-1i2+1i-11+i,

so

i11i3(i+1)=ζ(3)-ζ(2)+1.

Therefore

|Uf(y)-Uf(x)y-x|s(f)(ζ(3)-ζ(2)+1-1+ζ(3))=s(f)(2ζ(3)-ζ(2)).

For example, let f(x)=x, for which s(f)=1. Now,

Uf(x)=i1Pi(x)1x+i.

We remind ourselves that

Pi(x)=x+1(x+i)(x+i+1),Pi(x)=i2-i-(x+1)2(x+i)2(x+i+1)2.

Then

(Uf)(x) =i1(Pi(x)1x+i-Pi(x)1(x+i)2)
=i1(i2-i-(x+1)2(x+i)3(x+i+1)2-x+1(x+i)3(x+i+1))
=i1i2-i-(x+1)2-(x+1)(x+i+1)(x+i)3(x+i+1)2
=i1-2x2-ix-4x+i2-2i-2(x+i)3(x+i+1)2.

Check that x(Uf)(x) is increasing and negative. Then (Uf)|(Uf)(0)|, with

(Uf)(0)=i1i2-2i-2i3(i+1)2=-2ζ(3)+ζ(2).

Therefore for f(x)=x,

s(f)=(Uf)=2ζ(3)-ζ(2),

which shows that the above theorem is sharp.