arXiv:1706.03658v4 [math.CA] 31 Jul 2018 tv eeqemaue eaegigt nwrti usino mea on question this answer to going are ground. We theoretic measure? Lebesque itive eeaieodnr en eg h emti en oLbsu m Lebesgue to ) of geometric subsets the (e.g. ordinary generalize aiu eea rpriso uhmas h eain mn th [6]. among and relations [5] the in means, thoroughly studied such were of properties general various eea ocp hr ehv eno oeifiiesbesof subsets infinite m o a some get means on to mean of order a in theory have ways fi the we many (or two in where build of extended concept mean to be the general can started calculating This for we is numbers. mean where many) inves ordinary the [6] An of sets. and continuation infinite natural [5] a in as considered started be can paper This Introduction 1 e od n hae:gnrlzdma,Brlmaue Lebes measure, Borel mean, generalized Phrases: and Words Key 0 nti ae elo o h nwrtefloigqeto.Hwcno can How question. following the answer the for look we paper this In M 21)SbetCasfiain:28A10,28A20 Classifications: Subject (2010) AMS esrsb en,masb measures by means means, by Measures eairo uhgnrlzdmasta r bandb mea a by obtained are that proper to means the mea generalized mean study geometric such also the We the of extend calculate behavior measure. can Lebesgue can we positive we with E.g. set measure well. that as sets Using infinite way. natural very ecntutmauewihdtrie nodnr eni a in mean ordinary an determines which measure construct We R ..cnw aclt h emti eno e ihpos- with set a of mean geometric the calculate we can I.e. . tiaLosonczi Attila 0Arl2018 April 10 Abstract 1 u-tete measure gue-Stieltjes fany of n isand ties sure. s means ose easurable tigations R The . nitely sure ore ne n In the first part of the paper we investigate means that are created by measures on R. We enumerate many properties of such means and we also study uniqueness. In the second part of the paper we fulfill our main aim that is to find a measure that generates a given ordinary mean. We prove that under some basic smoothness conditions the generating measure always exists. Based on that result we provide an interesting application to ordinary means. We show that such ordinary means (a, b) are determined by the function f(x) = (1, x) i.e. (a, b) can be calculatedK by f(x) in a generic way. K K We also investigate some alternative ways how one can generate the given ordinary mean. One may ask whether some properties of an ordinary mean are inherited to the associated generalized mean. In this respect we show that the AM–GM inequality remains valid for the associated generalized means too. In the last section we analyse the behaviour of such means in infinity and show a sufficient condition for a mean approaching the in infinity.

1.1 Basic notions and notations

y +y For K R, y R let us use the notation K− = K ( ,y],K = K [y, ⊂+ ). ∈ ∩ −∞ ∩ ∞ If H R, x R then set H +x = h+x : h H . We use the convention that this⊂ operation∈ + has to be applied{ prior to the∈ set} theoretical operations, e.g. H K L + x = H K (L + x). ∪ ∪ ∪ ∪ Let us recall some very basic notions. Let be an ordinary mean that is just for calculating the mean of two numbersKa, b R. is called symmetric if (a, b) = (b, a). It is strictly ∈ K K K internal if a < (a, b) < b whenever a < b. It is called continuous if it is a continuous 2-variableK function of a and b. A generalized mean is a function : C R where C P (R) consists K → ⊂ of some (finite or infinite) subsets of R and inf H (H) sup H holds for all H C. ≤K ≤ ∈ Let us recall some definitions from [5] and [6] that regards for generalized means. Please note that here we are dealing with bounded sets only.

2 A mean is called internal if inf H (H) sup H. It is strongly K ≤ K ≤ internal if lim H (H) lim H where lim H = min H′, lim H = max H′. ≤K ≤ has property strict strong internality if it is strongly internal and limKH < (H) < lim H whenever H has at least 2 accumulation points. is disjoint-monotoneK if H H = , (H ) (H ) then (H ) K 1 ∩ 2 ∅ K 1 ≤K 2 K 1 ≤ (H1 H2) (H2). K is∪ union-monotone≤K if B C = , (A) (A B), (A) (A C) K ∩ ∅ K ≤K ∪ K ≤K ∪ implies (A) (A B C) and (A B) (A), (A C) (A) implies K(A ≤B KC) ∪ (∪A). MoreoverK if∪ any of≤ the K inequalitiesK ∪ on≤ the K left hand sideK is∪ strict∪ then≤K so is the inequality on the right hand side. +lim H lim H is bi-slice-continuous if H Dom( ) then H ,H− K x ∈+y K ∈ Dom( ) and f(x, y) = (H− H ) is continuous where Dom(f) = K x +y K ∪ (x, y): H− H Dom( ) . { ∪ ∈ K } is Cantor-continuous if H Dom( ),H H , ∞ H Dom( ) K i ∈ K i+1 ⊂ i ∩n=1 i ∈ K implies that (H ) ( ∞ H ). K i →K ∩n=1 i Throughout this paper λ will denote the Lebesgue measure. If H is bounded, Lebesgue measurable, λ(H) > 0 then

x dλ Avg(H)= HR . λ(H)

If f : R R is an increasing continuous function then let µf be the Lebesgue-Stieltjes→ measure belonging to f. I.e. the Carath´eodory extension of µ([a, b)) = f(b) f(a) (a < b). − Each measure µ considered in this paper is a Borel measure on some interval (finite or infinite) of the real line R that satisfies two conditions: (1) if H R is bounded and measurable then µ(H) < + ⊂ ∞ (2) if I is a non degenerative interval then 0 <µ(H). Let us remark that if H R is bounded then µ H is absolutely continuous with respect to λ iff it is ǫ⊂ δ absolutely continuous| ( ǫ > 0 δ > 0 such − ∀ ∃ that λ(K) < δ implies that µ(K) < ǫ).

2 Means by measures

Definition 2.1. Let I R be an interval (finite or infinite). Let µ be a Borel measure on I such⊂ that H R being bounded and measurable ⊂

3 implies that µ(H) < + . Let H be a bounded µ-measurable set such that 0 <µ(H) < + . Set ∞ ∞ xdµ µ(H)= HR . M µ(H) Definition 2.2. If µ is given, a, b R,a

Proof. By [6] Proposition 2 it is enough to prove finite independence and internality. The condition is equivalent to finite independence.

µ Lemma 2.5. Let H1,H2 Dom( ), µ((H1 H2) (H2 H1))=0. Then µ(H )= µ(H ). ∈ M − ∪ − M 1 M 2 R xdµ R xdµ R xdµ R xdµ+ R xdµ − Proof. H1 H2 = H1 H1 H2 H1 xdµ 1 1 + µ(H1) µ(H2) µ(H1) µ(H2) µ(H1) µ(H2) − − ≤ H − R1 R xdµ R xdµ − − H2 H1 µ(H 2) µ(H1) H2 H1 = xdµ − + = 0. µ(H2) µ(H1)µ(H2) µ(H2) H R1

Proposition 2.6. If µ(H)=0 whenever H is countable then µ is strict strong internal. M

µ a +b Proof. Let H Dom( ), a = lim H, b = lim H. By µ(H− )= µ(H )=0 we get that ∈µ([a, b])M = µ(H). Obviously there is c (a, b) such that M M ∈ µ µ(H [c, b]) > 0. Let H1 = H [a, c),H2 = H [c, b]. Then (H) µ(H1)a∩+µ(H2)c µ(H1) µ(H2) ∩ ∩ M ≥ µ(H) = µ(H) a + µ(H) c > a because it is a weighted and µ(H2) µ(H) > 0. The other inequality can be shown similarly. Exactly the same way one can show:

Proposition 2.7. Let µ is absolutely continuous with respect to λ, H µ x +x ∈ Dom( ). Let a = sup x R : λ(H− )=0 , b = inf x R : λ(H )=0 (cf. [5]M Definition 4). Then{ ∈a< µ(H) < b.} { ∈ } M

4 Proposition 2.8. Let H,H Dom( µ) (i N),H H = (i = i ∈ M ∈ i ∩ j ∅ 6 ∞ j),H = Hi. Then iS=1

∞ µ(H ) µ(H ) i M i µ(H)= iP=1 . M ∞ µ(Hi) iP=1 Obviously it holds for finitely many sets as well.

Proposition 2.9. Let H ,H Dom( µ), H = ∞ H . Then µ(H ) n ∈ M n M n → nT=1 µ(H). I.e. µ is Cantor-continuous. M M R xdµ R xdµ µ µ Hn H Proof. (Hn) (H) = = |M −M | µ(Hn) − µ(H)

µ(H)( xdµ xdµ)+(µ(H) µ(H )) xdµ − − n HRn HR HR = µ(Hn)µ(H)

µ(H) xdµ +(µ(H) µ(H )) xdµ − n HnR H HR − µ(Hn)µ(H) ≤

µ(H)µ(H H) sup H (µ(H) µ(H ))µ(H) sup H n − 1 + − n 0 2 2 µ(H) µ(H) → using that µ(H ) µ(H). n → Proposition 2.10. If H Dom( µ) then µ(H) is determined by ∈ M M all µ((a, b)) (a < b). M

Proof. It is known that there is a sequence (H ) such that H ∞H ,µ(H)= n ⊂ ∩1 n µ( ∞H ) and H is a countable union of disjoint open intervals. By Lemma ∩1 n n 2.5, Proposition 2.8 and 2.9 we get the statement.

2 + Example 2.11. Let f(x)= x ,µ = µf defined on subsets of R 0 . 2 2 ∪{ } Then µ(a, b)= 2 a +ab+b . M 3 a+b 5 x µ beb aea Example 2.12. Let f(x)= e ,µ = µf . Then (a, b)= eb−ea 1. M − − Proposition 2.13. µ is disjoint-monotone. M µ µ µ Proof. Let H1 H2 = , (H1) (H2). Then (H1 H2) = µ µ µ µ µ(H1) (H1)+µ(H∩2) (H2) ∅ Mµ(H1) (H2≤)+µ M(H2) (H2) µM ∪ M M M M = (H2). The other µ(H1 H2) ≤ µ(H1)+µ(H2) M inequality is∪ similar. Proposition 2.14. µ is union-monotone. M Proof. Let B C = , µ(A) µ(A B), µ(A) µ(A C). ∩ ∅ M ≤ M ∪ M ≤ M ∪ Obviously we can assume that A B = A C = . We know that µ µ(A) µ(A)+µ(B) µ(B) ∩ µ µ(∩A) µ(A)+∅µ(C) µ(C) (A) M M and (A) M M . M ≤ µ(A)+µ(B) M ≤ µ(A)+µ(C) µ µ(A) µ(A)+µ(B) µ (B)+µ(C) µ (C) (µ(A)+µ(B)) µ (A)+µ(C) µ (C) Then (A B C)= M M M M M = M ∪ ∪ µ(A)+µ(B)+µ(C) ≥ µ(A)+µ(B)+µ(C) µ(A) µ(A)+µ(C) µ (C)+µ(B) µ (A) (µ(A)+µ(C)) µ (A)+µ(B) µ (A) µ M M M M M = (A). µ(A)+µ(B)+µ(C) ≥ µ(A)+µ(B)+µ(C) M Clearly if µ(A) < µ(A B) also holds then we get that µ(A) < µ(A B CM). M ∪ M M ∪ ∪ The opposite inequalities can be handled similarly. Lemma 2.15. Let I be a bounded interval, µ be a Borel measure on I. µ Then H1 Dom( ) ǫ> 0 δ > 0 such that µ((H1 H2) (H2 H1)) < δ, H ∀ Dom∈ ( µ) Mimplies∀ that ∃ µ(H ) µ(H ) <− ǫ. ∪ − 2 ∈ M |M 1 −M 2 | R xdµ R xdµ µ µ H1 H2 Proof. (H1) (H2) = = |M −M | µ(H1) − µ(H2)

R xdµ+ R xdµ R xdµ + R xdµ − ∩ ∩ − H1 H2 H1 H2 H1 H2 H2 H1

µ(H1) − µ(H2) ≤

R xdµ R xdµ − − H1 H2 + xdµ 1 1 + H2 H1 = µ(H1) µ(H1) µ(H2) µ(H2) H1 H2 − ∩R R xdµ R xdµ − − H1 H2 µ(H2) µ(H1) H 2 H1 + xdµ − + . µ(H1) µ(H1)µ(H2) µ(H2) H1 H2 ∩R xdµ < δ H xdµ < µ H H K , xdµ < Clearly sup 1, ( 1) sup 1 = 1 H1 H2 H1 H2 H2 H1 −R ∩R −R δ sup H2 δ sup I. ≤µ(H1) If δ < 2 then δ sup H 2δ 2δ sup I µ H µ H < 1 K < ǫ ( 1) ( 2) + 1 2 + |M −M | µ(H1) µ(H1) µ(H1) showing that δ can be chosen.

6 Corollary 2.16. Let I be a bounded interval, µ be a Borel measure on I. If Dom( µ) is equipped with the pseudo-metric d (H ,H ) = µ((H M µ 1 2 1 − H ) (H H )) then µ is continuous according to d . 2 ∪ 2 − 1 M µ Example 2.17. This is obviuosly not true if I is not bounded. See 1 1 e.g. µ = λ,H1 = [0, 1], ǫ = 0.1,H2 = [0, 1] [ δ , δ + δ]. Then Avg(H1) = 0.5 1+( 1 + δ )δ ∪ · δ 2 0.5,Avg(H2)= 1+δ > 0.75. Proposition 2.18. If µ is absolutely continuous with respect to λ then µ is bi-slice-continuous. M Proof. We know that ǫ > 0 δ > 0 such that λ(H) < δ implies that µ(H) < ǫ. Then apply∀ Lemma 2.15.∃ Our next aim is to investigate inequalities between means.

µ Lemma 2.19. Let H,Hi Dom( ) (i N),Hi Hj = (i = ∈µ n M µ∈ ∩ ∅ 6 j),H = i∞=1Hi. Then limn ( i=1Hi)= (H). ∪ →∞ M ∪ M Proof. It is enough to refer to Lemma 2.15.

Proposition 2.20. Let µ, ν be Borel measures on an interval I. As- sume that if Ii I (1 i n, n N) are disjoint bounded open in- µ⊂ n ≤ ≤ν n ∈ µ ν tervals then ( i=1Ii) ( i=1Ii). Then (H) (H) H Dom( µ) MDom∪( ν). ≤ M ∪ M ≤ M ∀ ∈ M ∩ M Proof. By Lemma 2.19 it is true for countably many intervals too i.e. it is valid for any bounded open set. Finally if H Dom( µ) Dom( ν) then Cantor-continuity and Lemma 2.5 yield that ∈ µ(H)M ∩ν(H). M M ≤M Let us present a sufficient condition that we will apply later.

Proposition 2.21. Let µ, ν be Borel measures on an interval I and let us assume that the following two conditions hold. (1) If J I is a bounded open interval then µ(J) ν(J). ⊂ M ≤M ν(J) (2)If J, K are bounded open intervals such that sup J inf K then µ(J) ν(K) ≤ ≤ µ(K) . Then µ(H) ν(H) H Dom( µ) Dom( ν). M ≤M ∀ ∈ M ∩ M

7 Proof. First let us observe that condition 2 simply implies that if Ii I (1 i n, n N) are disjoint bounded open intervals and sup I inf⊂K ( ≤i) ≤ ∈ i ≤ ∀ then n ν( i=1Ii) ν(K) ∪n . µ( i=1Ii) ≤ µ(K) ∪ µ n ν n On the same assumptions by 2.20 we have to show ( i=1Ii) ( i=1Ii). It holds for n = 1 by condition 1. We go on by induction.M ∪ Suppose≤M it is true∪ for n 1. Let sup Ii inf In (1 i n 1). Using− Proposition≤ 2.8 we have≤ ≤ − n 1 µ n 1 µ n 1 ν n 1 ν µ n µ( i=1− Ii) ( i=1− Ii)+ µ(In) (In) µ( i=1− Ii) ( i=1− Ii)+ µ(In) (In) ( i=1Ii)= ∪ M n∪1 M ∪ M n∪1 M . M ∪ µ( − I )+ µ(I ) ≤ µ( − I )+ µ(I ) ∪i=1 i n ∪i=1 i n It is enough to prove that n 1 ν n 1 ν n 1 ν n 1 ν µ( i=1− Ii) ( i=1− Ii)+ µ(In) (In) ν( i=1− Ii) ( i=1− Ii)+ ν(In) (In) ∪ M n∪1 M ∪ M n∪1 M µ( − I )+ µ(I ) ≤ ν( − I )+ ν(I ) ∪i=1 i n ∪i=1 i n and that is equivalent to

n 1 ν n 1 ν n 1 µ( − I ) ( − I )ν(I )+ µ(I ) (I )ν( − I ) ∪i=1 i M ∪i=1 i n n M n ∪i=1 i ≤ n 1 ν n 1 ν n 1 ν( − I ) ( − I )µ(I )+ ν(I ) (I )µ( − I ) ∪i=1 i M ∪i=1 i n n M n ∪i=1 i and

ν ν n 1 n 1 n 1 0 (In) ( i=1− Ii) ν(In)µ( i=1− Ii) µ(In)ν( i=1− Ii) ≤ M −M ∪  ∪ − ∪  But the first term is obviously positive and by the consequence of condi- n 1 n 1 tion 2 ν(I )µ( − I ) µ(I )ν( − I ) is valid as well. n ∪i=1 i ≥ n ∪i=1 i Proposition 2.22. Let f,g be increasing differentiable functions. If g′(x) f ′(x) is increasing then condition 2 (in 2.21) holds for µf ,µg. Proof. Let J =(a, b),K =(c,d), b c. We have to show that ≤ g(b) g(a) g(d) g(c) − − . f(b) f(a) ≤ f(d) f(c) − − By Cauchy’s mean value theorem there are α (a, b) and β (c,d) such that ∈ ∈ g(b) g(a) g′(α) g(d) g(c) g′(β) − = , − = f(b) f(a) f (α) f(d) f(c) f (β) − ′ − ′ g′(α) g′(β) and ′ ′ by assumption. f (α) ≤ f (β) 8 Let us investigate uniqueness.

Theorem 2.23. Let µ = ν. Then there is c R,c> 0 such that M M ∈ ν = cµ.

Proof. Let A, B Dom( µ), A B = . Then µ(A B)= ν(A B). ∈ M ∩ ∅ M ∪ M ∪ By 2.8 µ(A) µ(A)+ µ(B) µ(B) ν(A) ν (A)+ ν(B) ν(B) M M = M M . µ(A)+ µ(B) ν(A)+ ν(B)

By µ(A)= ν(A), µ(B)= ν(B) M M M M µ(A) µ(A)+ µ(B) µ(B) ν(A) µ(A)+ ν(B) µ(B) M M = M M . µ(A)+ µ(B) ν(A)+ ν(B)

Then

µ(A) µ(A)ν(B)+µ(B) µ(B)ν(A)= ν(A) µ(A)µ(B)+ν(B) µ(B)µ(A) M M M M and µ(A) µ(B) µ(A)ν(B) ν(A)µ(B) =0. M −M  −  If µ(A) = µ(B), A B = then M 6 M ∩ ∅ ν(A) ν(B) = µ(A) µ(B) has to hold. Let I = (0, 1),c = ν(I) . If H Dom( µ) then let J be an interval µ(I) ∈ M such that I J = , sup H < inf J. Then H J = and µ(I) = ∩ ∅ ∩ ∅ M 6 µ(J), µ(H) = µ(J) hold. We get that ν(I) = ν(J) and ν(J) = ν(H) . M M 6 M µ(I) µ(J) µ(J) µ(H) I.e. ν(H)= cµ(H).

3 Measures by means

Let an ordinary mean be given that is just for calculating the mean of two numbers. Can we extendK this mean somehow to some subsets of R? We may have many options for doing so. But now we are going to approach this from measure theory.

9 We know that b xdλ a + b Ra = λ([a, b]) 2 that is the arithmetic mean. Therefore we can try to look for a measure µ b R xdµ such that a = (a, b) where a, b R and a, b Dom( ). µ([a,b]) K ∈ { } ∈ K Theorem 3.1. Let be an ordinary mean that is symmetric, strictly K∂ (x,y) internal, continuous and K∂y exists for all x, y Dom( ) and it is continuous. Then there exists a measure µ that{ is absolutely} ∈ continuousK to λ such that µ([a, b]) = (a, b). M K Proof. Let us look for µ in the form µ = µf where f is an increasing differ- dµ entiable function. Then µ([a, b]) = f(b) f(a) and f ′ = dλ . Then −

b b xdµf xf ′dλ [xf F ]b bf(b) af(a) (F (b) F (a)) (a, b)= Ra = Ra = − a = − − − K µ ([a, b]) f(b) f(a) f(b) f(a) f(b) f(a) f − − − where F is a primitive function of f. We can assume that there is a point a such that f(a) = F (a) = 0 because f and f + c, F and F + d have the same effect. Let us suppose that a = 1 i.e. f(1) = F (1) = 0. Then we get xf(x) F (x) F (x) − (1, x)= f(x) = x f(x) (x = 1). K Let us observe that µ− is a measure6 hence both f and F are monotone increasing. We can write F (x) = x (1, x). Equivalently f(x) = 1 as is f(x) −K F (x) x (1,x) K strictly internal we do not divide here by 0. −K ′ F (x) f(x) 1 Then (log F (x))′ = F (x) = F (x) = x (1,x) Let ǫ> 0,b> 1+ ǫ. (If b< 1 then we−K can handle that similarly.) b b 1 (log F (x))′ = log F (b) log F (1 + ǫ)+ C = x (1,x) dx 1+Rǫ − 1+Rǫ −K The integral on the right hand side exists because [1 + ǫ, b] is compact, x (1, x) is continuous hence it takes its minimum but it is > 0 since is stricly−K internal. Set C = log F (1 + ǫ). Then K

b 1 R x−K(1,x) dx F (b)= e1+ǫ

10 b 1 dx 1 R x−K(1,x) 1 f(b)= e1+ǫ = F (b) b (1, b) b (1, b) −K −K

b ∂ (1,b) 1 ∂ (1,b) ∂ (1,b) R dx K∂b x−K(1,x) K∂b K∂b f ′(b)= e1+ǫ = F (b)= f(b) (b (1, b))2 (b (1, b))2 b (1, b) −K −K −K We got f and F by using (1, x) only. Therefore we also have to check K whether f and F fulfills our original request i.e. they work for (a, b) as well. K

b b a xdµf xdµf xdµf − µ ([1, b]) (1, b) µ ([1, a]) (1, a) Ra = R1 R1 = f K − f K = µ ([a, b]) f(b) f(a) f(b) f(a) f − −

bf(b) 1f(1) (F (b) F (1)) (af(a) 1f(1) (F (a) F (1))) − − − − − − − = f(b) f(a) − bf(b) af(a) (F (b) F (a)) − − − = (a, b). f(b) f(a) K − Corollary 3.2. Let be an ordinary mean that is symmetric, strictly K∂ (x,y) internal, continuous and K exists for all x, y Dom( ) and it is ∂y { } ∈ K continuous. Then g(x)= (1, x) determines (a, b). K K Proof. Using the constructed f in Theorem 3.1 we get (f(b) f(1)) (1, b) (f(a) f(1)) (1, a) (a, b)= − K − − K K f(b) f(a) − and f is calculated by g(x)= (1, x). K Remark 3.3. If f is an increasing differentiable function, F is one of its primitive functions then bf(b) af(a) (F (b) F (a)) (a, b)= − − − K f(b) f(a) − define a strictly internal, continuous ordinary mean (a < b).

11 Proof. Let us give a direct proof. By Cauchy’s mean value theorem there is ξf ′(ξ) ξ (a, b) such that (a, b)= ′ = ξ. ∈ K f (ξ) Example 3.4. Let be the : (a, b) = √ab. Then (√x 1)2 K1 1 1 1K F (x) = − , f(x) = (1 ), f ′(x) = . Hence µ([a, b]) = e2 e2 − √x 2e2 x√x µ ([a, b]) = 1 ( 1 1 ). f e2 √a − √b Proof. We know that b 1 R x−K(1,x) dx F (b)= e1+ǫ b b b 1 1 1 hence we have to calculate x √x dx = √x(√x 1) dx = √x 1 dx 1+Rǫ − 1+Rǫ − 1+Rǫ − − b 1 √x dx. Let us apply the following substitution in the first case y = √x 1. 1+Rǫ − 1 1 Then we end up with √x 1 dx = y 2(y + 1)dy = 2y + 2 log y + C = − 2(√x 1) + 2 log(√x 1)+R C. FinallyR 1 dx = 2 + 2 log(√x 1)+ C. − − x √x − − (√x 1)2 R − So we get F (x)= e−2 . Then f, f ′ can be obtained easily from that. b b 1 1 1 √ Let us verify that it works. xf ′ = e2 √x = e2 ( b √a). Then Ra Ra −

b xf ′ 1 (√b √a) Ra = e2 − = √ab. f(b) f(a) 1 ( 1 1 ) − e2 √a − √b It is well known that the (ordinary) geometric mean is a quasi-arithmetic mean i.e. the geometric mean can be derived from the arithmetic mean using log a+log b the log function: (a, b)= e 2 . One might ask whether the generalized geometric mean canG be derived from Avg in the same way i.e. whether (H) = eAvg log H holds where log H = log h : h H and H R+ is a G { ∈ } ⊂ Borel set. The answer is negative as the next proposition shows.

Proposition 3.5. Let µ be the Borel measure associated to the geomet- ric mean. Then there is a Borel set H R+ such that eAvg log H = µ(H). ⊂ 6 M Proof. Let H = [1, e2] [e4, e8]. Then log H = [0, 2] [4, 8] hence Avg log H = 2 2 2 2 ∪ 13 ∪ 1 2 0 +8 4 13 Avg log H 3 2 − 2+4 − = 3 which gives that e = e .

12 2 8 4 4 2 µ √e √1+ √e √e e 1+ e e 13 (H)= − − = − − = e 3 . M 1 1 + 1 1 1 1 + 1 1 6 √1 − √e2 √e4 − √e8 − e e2 − e4 We can go further by proving that the generalized mean (H)= eAvg log H is not a mean by measure. M

Theorem 3.6. There does not exist a Borel measure µ on R+ such that eAvg log H = µ(H) (where H is any Borel set). M Proof. Clearly eAvg log(a,b) = √ab (a, b R+) because Avg(log a, log b) = log a+log b √ ∈ 2 = log ab ((a, b) and (log a, log b) denote open intervals). If we assumed that eAvg log H was a mean by measure then by 2.10 we would get that it would be equal to the mean by measure obtained from the (ordinary) geometric mean. But it is false by 3.5. Now we prove that the inequality between the arithmetic and the geo- metric mean remains valid for the generalized means too.

Theorem 3.7. Let µ be the Borel measure associated to the geometric mean. Then H Dom( µ) implies that µ(H) Avg(H). ∈ M M ≤ g′(x) Proof. By 2.21 and 2.22 we only have to show that f ′(x) is increasing for g(x)= x 1 and f(x)=1 1 . But that is equal to x√x. − − √x

Corollary 3.8. If Ii = (ai, bi) and bi < aj when i < j (1 i, j n) then ≤ ≤ n n 2 2 √bi √ai 1 b a i=1 − i=1 i − i . P n 1 1 n ≤ 2 P i bi ai i=1 √ai − √bi =1 − P P Example 3.9. a, b 2 Let be the : ( ) = 1 + 1 . Then K K a b 1 1 2 F (x) = x 2+ , f(x)=1 , f ′(x) = . Hence µ([a, b]) = µ ([a, b]) = − x − x2 x3 f 1 1 . a2 − b2 1 dx x+1 dx 2 dx 1 dx Proof. We have to calculate x 2 = x(x 1) = x 1 x = 1+ 1 − R − x R − R − R (x 1)2 1 2 log(x 1) log x + C = log − + C hence F (x)= x 2+ . − − x − x

13 b b 2 b a 1 1 (a+b)(b a) Let us verify it. xf ′ = x2 =2 ab− . f(b) f(a)= a2 b2 = a2b2− . Ra Ra − −

b xf ′ b a a2b2 1 Ra =2 − =2 . f(b) f(a) ab (b a)(a + b) 1 + 1 − − a b a b Example 3.10. Let be the : (a, b) = log a−log b . K K − Then one can easily verifies that F (x)= x log x x +1, f(x) = log x, f ′(x)= 1 . Hence µ([a, b]) = µ ([a, b]) = log b log a. − x f − 3.1 Alternative ways First let us present another way to get the arithmetic mean by some integral.

Proposition 3.11. One can easily show that

b d x + y (b2 a2)(d c)+(d2 c2)(b a) dxdy = − − − − . Z Z 2 4 a c Corollary 3.12.

b d x+y dxdy 2 a + b + c + d Ra Rc = . λ([a, b])λ([c,d]) 4 Corollary 3.13.

b b x+y dxdy 2 a + b Ra Ra = . λ([a, b])2 2

For a given ordinary mean one can try to find a measure µ on R such that K b b b b x+y dµ µ x+y dµ(x)dµ(y) 2 × 2 Ra Ra = Ra Ra = (a, b). µ µ([a, b] [a, b]) µ([a, b])2 K × ×

14 Theorem 3.14. Let be an ordinary mean that is symmetric, strictly ∂K (x,y) internal, continuous and K exists for all x, y Dom( ) and it is ∂y { } ∈ K continuous. Then there exists a measure µ that is absolutely continuous to λ b b x+y R R 2 dµ(x)dµ(y) such that a a = (a, b). µ([a,b])2 K Proof. We follow exactly the same way than in Theorem 3.1. Let us look for µ in the form µ = µf where f is an increasing differentiable dµ function. Then µ([a, b]) = f(b) f(a) and f ′ = dλ . Let F be a primitive function of f. − Then b b b b x+y x+y 2 dµ(x)dµ(y) 2 f ′(x)f ′(y)dxdy Ra Ra = Ra Ra = µ([a, b])2 (f(b) f(a))2 −

b 1 f ′(y)[bf(b) af(a) (F (b) F (a)) + y(f(b) f(a))]dy 2 − − − − Ra = (f(b) f(a))2 − 1 [bf(b) af(a) (F (b) F (a))](f(b) f(a)) 2 − − − − + (f(b) f(a))2 − 1 (f(b) f(a))[bf(b) af(a) (F (b) F (a))] 2 − − − − = (f(b) f(a))2 −

[bf(b) af(a) (F (b) F (a))](f(b) f(a)) bf(b) af(a) (F (b) F (a)) − − − − = − − − . (f(b) f(a))2 f(b) f(a) − − That is exactly the same formula that we got in Theorem 3.1. Therefore the same measure will work here as well.

Problem 1. One might ask the following question. For a given mean b b R R (x,y)dµ(x)dµ(y) can we find a measure µ on R such that a a K = (a, b)? For K µ([a,b])2 K which means can we expect such measure?

15 4 Behaviour in infinity

It is known that a + b (a + x)+(b + x) lim ( (a + x)(b + x) x) = lim (a + x)(b + x)=0 x + 2 x + 2 → ∞ − p − → ∞ −p i.e. in the far distance the geometric mean starts to behave as the arithmetic mean. Similarly we can ask when a mean by measure µ behaves in the same way, namely M lim µ(H + x) Avg(H + x) =0(H Dom( µ)). x + → ∞ |M − | ∈ M We are going to present a sufficient condition for that. In this section µ will denote a Borel measure on R+. Definition 4.1. Let I R+ be a finite interval. ⊂ µ(H) µ mI = inf : H I,H Dom( ) nλ(H) ⊂ ∈ M o

µ(H) µ MI = sup : H I,H Dom( ) . nλ(H) ⊂ ∈ M o Theorem 4.2. Let µ be a Borel measure on R+ such that if I R+ is a finite interval then 0 < m M < + and ⊂ I ≤ I ∞ M lim I+x =1. x + m → ∞ I+x Moreover if H is µ-measurable then so is H + x x> 0. ∀ Then H Dom( µ) implies that ∈ M lim µ(H + x) Avg(H + x) =0. x + → ∞ |M − | Proof. Let H Dom( µ),x> 0. Let I R+ be a finite interval such that ∈ M ⊂ H I. ⊂First let us observe that H Dom( µ) implies that H+x Dom( µ) x> ∈ M ∈ M ∀ 0 by the first condition. Then we get the statement by m xdλ xdµ M xdλ m I+x I+x M I+x Avg(H) HR HR HR I+x Avg(H). MI+x ≤ MI+xλ(H) ≤ µ(H) ≤ mI+xλ(H) ≤ mI+x

16 Now our aim is to prove that the geometric mean satisfies these conditions.

Lemma 4.3. Let (Hi) is a sequence of µ-measurable sets such that H H in the pseudo-metric d . Then µ(H ) µ(H). i → µ i → Lemma 4.4. Let µ be absolutely continuous with respect to λ. Let (Hi) be a sequence of bounded λ-measurable sets such that Hi H in the µ(Hi) µ→(H) pseudo-metric dλ. Let λ(H) > 0,H be bounded. Then . λ(Hi) → λ(H) Proof. By absolute continuity H H according to d as well. Then by i → µ Lemma 4.3 µ(H ) µ(H) and λ(H ) λ(H). i → i → Lemma 4.5. Let µ is absolutely continuous with respect to λ and vica versa. Let I R+ be a finite interval. Then ⊂ µ(K) n µ N mI = inf : K = i=1Ii,Ii I,Ii Ij = (i = j),Ii Dom( )(i, j ) nλ(K) ∪ ⊂ ∩ ∅ 6 ∈ M ∈ o where Ii denotes an interval.

Proof. First let us observe that Dom( µ)= Dom( λ). Let m denote the right hand side ofM the above equality.M Obviously m I ≤ m. Let H I,H Dom( µ),ǫ > 0. Then there are countably many ⊂ ∈ M disjoint intervals (Ii) such that H 1∞Ii and 1∞ λ(Ii) < λ(H)+ ǫ. Then ⊂ ∪ n we can choose n N such that λ(H) ǫ< Pλ(I ) <λ(H)+ ǫ. Therefore ∈ − 1 i we can construct a sequence (Ki) such that PKi is a disjoint union of finitely µ(Ki) µ(H) many intervals and λ(Ki) λ(H). By Lemma 4.4 . → λ(Ki) → λ(H) Proposition 4.6. The Borel measure µ associated to the geometric mean satisfies the conditions of Theorem 4.2.

Proof. Let I = [c,d]. Let us calculate mI+x. 1 1 1 1 b a If a, b I + x, a < b then µ([a, b]) = ( ) = ( − ). We ∈ e2 √a − √b e2 √ab(√a+√b) get that µ([a,b]) = 1 ( 1 ). If we want its infimum for a, b then a, b λ([a,b]) e2 √ab(√a+√b) d x 1 1 have to tend to + . Therefore for one interval the infimum is e2 2(d+x)√d+x . We want to show that if I1,...,In I, sup Ii < inf Ij (i < j) then µ(I1)+ +µ(In) µ(In) ⊂ ··· > . We go by induction: Let I ,I I, sup I < inf I . λ(I1)+ +λ(In) λ(In) 1 2 1 2 ··· ⊂ µ(I1)+µ(I2) µ(I2) µ(I1)+ +µ(In−1) Then evidently > . Let us assume that ··· > λ(I1)+λ(I2) λ(I2) λ(I1)+ +λ(In−1) ···

17 µ(In−1) µ(I1)+ +µ(In) µ(In) µ(I1)+ +µ(In−1) µ(In) . Clearly ··· > is equivalent to ··· > λ(In−1) λ(I1)+ +λ(In) λ(In) λ(I1)+ +λ(In−1) λ(In) ··· ··· but by µ(In−1) > µ(In) it holds. λ(In−1) λ(In) m 1 1 M Hence by Lemma 4.5 we get that I+x = e2 2(d+x)√d+x . Similarly I+x = 1 1 e2 2(c+x)√c+x . Finally

d d M (d + x)√d + x ( x + 1) x +1 lim I+x = lim = lim q =1. x + m x + x + c c → ∞ I+x → ∞ (c + x)√c + x → ∞ ( x + 1) x +1 p Corollary 4.7. Let µ be the Borel measure associated to the geometric mean. Then H Dom( µ) implies that ∈ M lim µ(H + x) Avg(H + x) =0. x + → ∞ |M − | References

[1] J. M. Borwein, P. B. Borwein, The way of all means, Amer. Math. Monthly 94 (1987), 519-522.

[2] P. Billingsley, Probability and Measure, vol. 939 John Wiley & Sons (2012).

[3] P. S. Bullen, Handbook of means and their inequalities, vol. 260 Kluwer Academic Publisher, Dordrecht, The Netherlands (2003).

[4] M. Hajja, Some elementary aspects of means, International Journal of Mathematics and Mathematical Sciences, Means and Their Inequali- ties, Volume 2013, Article ID 698906, 1–9.

[5] A. Losonczi, Means of infinite sets I, arXiv prepint

[6] A. Losonczi, Means of infinite sets II, arXiv prepint

[7] A. Losonczi, Measuring sets by means, arXiv prepint

[8] L. Losonczi and Zs. P´ales, Comparison of means generated by two functions and a measure, Journal of Mathematical Analysis and Ap- plications 345 vol.1 (2008), 135–146.

18 [9] L. Losonczi, Inequalities for integral mean values, Journal of Mathe- matical Analysis and Applications 61 vol.3 (1977), 586–606.

[10] J. S´andor, G. Toader, Inequalities for general integral means, Journal of Inequalities in Pure and Applied Mathematics 7.1 (2006), 1–5.

Dennis G´abor College, Hungary 1119 Budapest Fej´er Lip´ot u. 70. E-mail: [email protected], [email protected]

19