• Keine Ergebnisse gefunden

Probability Theory

N/A
N/A
Protected

Academic year: 2022

Aktie "Probability Theory"

Copied!
20
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

Probability Theory

Wilhelm Stannat

Technische Universität Darmstadt Winter Term 2007/08

Third part - corrected version

This text is a summary of the lecture on Probability Theory held at the TU Darmstadt in Winter Term 2007/08.

Please email all misprints and mistakes to

stannat@mathematik.tu-darmstadt.de

(2)

Bibliography

1. Bauer, H.,Probability theory, de Gruyter, 1996.

2. Bauer, H.,Maß- und Integrationstheorie, de Gruyter, 1996.

3. Billingsley, P.,Probability and Measure, Wiley, 1995.

4. Billingsley, P.,Convergence of probability measures, Wiley, 1999.

5. Dudley, R.M.,Real analysis and probability, Cambridge University Press, 2002.

6. Elstrodt, J.,Maß- und Integrationstheorie, Springer, 2005.

7. Feller, W.,An introduction to probability theory and its applications, Vol. 1 & 2, Wiley, 1950.

8. Halmos, P.R.,Measure Theory, Springer, 1974.

9. Klenke, A.,Wahrscheinlichkeitstheorie, Springer, 2006.

10. Shiryaev, A.N.,Probability, Springer, 1996.

(3)

2 Independence

4 Joint distribution and convolution

LetXi∈L1 i.i.d. Kolmogorov’s law of large numbers implies that 1

n Xn i=1

Xi(ω)

| {z }

=:Sn

n→∞

−−−−→E[X1] P-a.s.

hence Z

f(x)d P◦ Sn

n 1!

(x) =E

"

f Sn

n

#

(Lebesgue)

−−−−→n→∞ f E[X1]

= Z

f(x)dδE[X1](x) ∀f ∈Cb(R)

i.e., the distribution of Snn converges weakly toδE[X1]. This is not surprising, because at least forXi∈L2

var Sn

n

= 1 n2

Xn i=1

var(Xi)

| {z }

=var(X1)

−−−−→n→∞ 0.

We will see later that if we rescaleSnappropriately, namely1

nSn, so thatvar 1 nSn

= var(X1)), the sequence of distributions of 1nSn is asymptotically distributed as a nor- mal distribution.

One problem in this context is: How to calculate the distribution ofSn? SinceSn is a function ofX1, . . . , Xn, we need to consider theirjoint distribution in the sense of the following definition:

Definition 4.1. LetX1, . . . , Xnbe real-valued r.v. on(Ω,A, P). Then the distribution

¯

µ:=P◦X¯1of the transformation

X¯ : Ω→Rn, ω7→ X1(ω), . . . , Xn(ω)

underP is said to be thejoint distributionofX1, . . . , Xn. Note thatµ¯ is a probability measure on Rn,B(Rn)

withµ( ¯¯ A) =P[ ¯X ∈A]¯ for all A¯∈B(Rn).

Remark 4.2. (i) µ¯ is well-defined, becauseX¯ : Ω→Rn is A/B(Rn)-measurable.

(4)

Proof:

B(Rn) =σ

A1× · · · ×An

Ai ∈B(R)

A1× · · · ×An

Ai = ]−∞, xi], xi∈R and

1(A1× · · · ×An) =

\n i=1

{Xi∈Ai}

| {z }

A

∈A

which implies the measurability of the transformationX¯ (see Remark 1.3.2 (ii)) (ii) Proposition 1.11.5 implies thatµ¯is uniquely determined by

¯

µ(A1× · · · ×An) =P\n

i=1

{Xi∈Ai} .

Example 4.3. (i) LetX, Y be r.v., uniformly distributed on[0,1]. Then

• X,Y independent⇒joint distribution = uniform distribution on[0,1]2

• X =Y ⇒joint distribution = uniform distribution on the diagonal

(ii) LetX, Y be independent,N(m, σ2)distributed. The following Proposition shows that the joint distribution ofX andY has the density

f(x, y) = 1 2πσ2 ·exp

− 1

2 · (x−m)2+ (y−m)2 which is a particular example of a2-dimensional normal distribution.

In the casem= 0it follows that R:=p

X2+Y2, Φ := arctanY

X, are independent and

Φhas a uniform distribution on

π2,π2 ,

R has a density ( r

σ2exp −r22

ifr>0 0 ifr <0.

Definition 4.4. (Products of probability spaces) The product of measurable spaces (Ωi,Ai),i= 1, . . . n, is defined as the measurable space

Ω := Ω1×. . .×Ωn

(5)

endowed with the smallestσ-algebra

A:=σ{A1×. . .×An | Ai∈Ai,1≤i≤n}

generated by measurable cylindrical sets. Ais said to be theproductσ-algebraof Ai (notation: Nn

i=1Ai).

LetPi,i= 1, . . . , n, be probability measures on(Ωi,Ai). Then there exists a unique probability measureP on the product space(Ω,A)satisfying

P(A1× · · · ×An) =P1(A1)·. . .·Pn(An)

for every measurable cylindrical set. P is called theproduct measure ofPi (notation:

Nn i=1Pi).

(Uniqueness ofP follows from 1.11.5, existence later!)

Proposition 4.5. Let X1, . . . , Xn be r.v. on(Ω,A, P) with distributions µ1, . . . , µn

and joint distributionµ. Then¯

X1, . . . , Xn independent ⇔ µ¯= On

i=1

µi,

(i.e.,µ(A¯ 1× · · · ×An) =Q

iµi(Ai)ifAi ∈B(R)).

In this case:

(i) µ¯ is uniquely determined byµ1, . . . , µn. (ii)

Z

ϕ(x1, . . . , xn) d¯µ(x1, . . . , xn)

= Z

· · · Z

ϕ(x1, . . . , xni1(dxi1)

· · ·

!

µin(dxin).

for allB(Rn)-measurable functionsϕ:Rn→R¯ withϕ≥0orϕµ-integrable.¯ (iii) Ifµi is absolutely continuous with densityfi, i= 1, . . . , n, thenµ¯ is absolutely

continuous with density f¯(¯x) :=

Yn i=1

fi(xi).

Proof. The equivalence is obvious.

(i) Obvious from part (ii) of the previous Remark 4.2.

(ii) See text books on measure theory.

(6)

(iii) f¯is nonnegative and measurable onRn w.r.t. B(Rn), and Z

Rn

f¯(¯x) d¯x= Yn i=1

Z

R

fi(xi) dxi= 1.

Hence, ˇ µ(A) :=

Z

A

f¯(¯x) d¯x, A∈B(Rn),

defines a probability measure on(Rn,B(Rn)). ForA1, . . . , An ∈B(R)it follows that

¯

µ(A1× · · · ×An) = Yn i=1

µi(Ai) = Yn i=1

Z

Ai

fi(xi) dxi

(ii)= Z

1A1×···×An(¯x)·f¯(¯x) d¯x= ˇµ(A1× · · · ×An).

Henceµ¯= ˇµby 1.11.5.

LetX1, . . . , Xn be independent,Sn :=X1+· · ·+Xn

How to calculate the distribution ofSn with the help of the distribution ofXi? In the following denote byTx: R1→R1,y7→x+y, the translation byx∈R. Proposition 4.6. LetX1, X2 be independent r.v. with distributionsµ1, µ2. Then:

(i) The distribution ofX1+X2 is given by theconvolution µ1∗µ2:=

Z

µ1(dx12◦Tx11,i.e.

µ1∗µ2(A) = = Z

1A(x1+x21(dx12(dx2)

= Z

µ1(dx12(A−x1) ∀A∈B(R1).

(ii) If one of the distributionsµ1, µ2is absolutely continuous, e.g. µ2with densityf2, thenµ1∗µ2 is absolutely continuous again with density

f(x) :=

Z

µ1(dx1)f2(x−x1)

= Z

f1(x1)·f2(x−x1) dx1=: (f1∗f2)(x) ifµ1=f1dx1.

(7)

Proof. (i) LetA∈B(R), and define A¯:=

(x1, x2)∈R2

x1+x2∈A ). Then P[X1+X2∈A] =P

(X1, X2)∈A¯

= (µ1⊗µ2)( ¯A)

= Z Z

1A¯(x1, x2) d(µ1⊗µ2)(x1, x2)

= Z Z

1A(x1+x2) d(µ1⊗µ2)(x1, x2)

= Z Z

1A−x1(x22(dx2)

µ1(dx1)

= Z

µ2(A−x11(dx1) = (µ1∗µ2)(A).

(ii)

1∗µ2)(A) = Z

µ1(dx12(A−x1) = Z

µ1(dx1) Z

Ax1

f2(x2) dx2 change of variable

xx1=x2

= Z

µ1(dx1) Z

A

f2(x−x1) dx

4.5= Z

A

Z

µ1(dx1)f2(x−x1)

dx.

Example 4.7.

(i) Let X1, X2 be independent r.v. with Poisson-distribution πλ1 and πλ2. Then X1+X2 has Poisson-distributionπλ12, because

λ1∗πλ2)(n) = Xn k=0

πλ1(k)·πλ2(n−k) =e12) Xn k=0

λk1

k! · λn−k2 (n−k)!

=e12)1 n!

Xn k=0

n k

·λk1λn−k2 =e12)·(λ12)n n! .

(ii) Let X1, X2 be independent r.v. with normal distributions N(mi, σ2i), i = 1,2.

ThenX1+X2has normal distributionN(m1+m2, σ1222), becausefm1+m22

122 = fm121∗fm222 (Exercise!)

(iii) TheGamma distributionΓα,pis defined through its densityγα,pgiven by γα,p(x) =

( 1

Γ(p)·αpxp1eαx ifx >0

0 ifx60

If X1, X2 are independent with distribution Γα,pi, i = 1,2, then X1+X2 has distributionΓα,p1+p2. (Exercise!)

(8)

In the particular casepi= 1: The sumSn=T1+. . .+Tn of independent r.v. Ti

with exponential distribution with parameterαhas Gamma-distributionΓn,α, i.e.

γα,n(x) = ( αn

(n1)!·e−αxxn−1 ifx >0

0 ifx60.

Example 4.8(The waiting time paradox). LetT1, T2, . . . be independent, exponentially distributed waiting times with parameterα >0, so that in particular

E[Ti] = Z

0

x·αeαxdx=· · ·= 1 α.

T1

z }| { T2

z }| { . . . t

| {z } X

| {z } Y

Question: Fix some timet. How long on average is the remaining waiting time until the next event, i.e., how big isE[Y]?

Answer: E[Y] = α1, and E[X+Y] = 1

α(1−e−αt)≈ 1

α for larget . More precisely:

(i) X, Y are independent.

(ii) Y has exponential distribution with parameterα.

(iii) X has exponential distribution with parameterα, "conditioned on"[0, t], i.e.:

P[X >s] =eαs ∀06s6t, P[X =t] =e−αt;

In particular, E[X] =

Z t 0

s·αeαs ds+t·eαt=· · ·= 1

α(1−eαt).

Proof. Let us first determine the joint distribution of X andY: Fix 0 6x 6t and

(9)

y>0. Then forSn:=T1+· · ·+Tn,S0:= 0:

P[X >x, Y >y]

= P[

nN

{Sn+x6t, Sn+1−y>t}

= P[T1>y+t] + X n=1

P

Sn6t−x, Tn+1>y+t−Sn

=eα(t+y)+ X n=1

Z Z

{(s,r)|s6tx, r>y+ts}

γα,n(s)·αeαrdsdr

=eα(t+y)+ X n=1

Z

{s|s6t−x}

γα,n(s)·eα(y+ts)ds

=eα(t+y)

1 + Z t−x

0

eαs X n=1

γα,n(s)

| {z }

()

=α

ds

=eα(t+y)

1 + Z t−x

0

αeαs ds

=eα(t+y)·eα(tx)=eαy·eαx. Consequently:

(i) If x= 0: Y ist exponentially distributed with parameterα.

(ii) Ify= 0: X ist exponentially distributed with parameterα, conditioned on[0, t].

(iii) X, Y are independent.

We have used in line six that:

X n=1

γα,n(s) = X n=1

αn

(n−1)!·sn1eαs =αeαs X n=1

(αs)n−1

(n−1)! =αeαseαs =α.

5 Characteristic functions

LetM1+(Rn)be the set of all probability measures on(Rn,B(Rn)).

For givenµ∈M1

+(Rn)define itscharacteristic functionas the complex-valued func- tionµˆ:Rn→Cdefined by

ˆ µ(u) :=

Z

eihu,yiµ(dy) :=

Z

cos(hu, yi)µ(dy) +i Z

sin(hu, yi)µ(dy). Proposition 5.1. Letµ∈M1+(Rn). Then

(10)

(i) µ(0) = 1.ˆ (ii) |µˆ|61.

(iii) µ(ˆ −u) = ˆµ(u).

(iv) µˆ is uniformly continuous.

(v) µˆ is positive definite, i.e. for all c1, . . . , cm∈C,u1, . . . , um∈Rn,m>1:

Xm j,k=1

cj¯ck·µ(uˆ j−uk)>0.

Proof. Exercise.

Proposition 5.2 (Uniqueness theorem). Let µ1, µ2 ∈M1+(Rn) with µˆ1 = ˆµ2. Then µ12.

Proposition 5.3 (Bochner’s theorem). Let ϕ : Rn → C be a continuous, positive definite function withϕ(0) = 1. Then there exists one (and only one) µ ∈M1

+(Rn) withµˆ=ϕ.

Proposition 5.4(Lévy’s continuity theorem). Let(µm)mNbe a sequence inM1+(Rn).

Then

(i) limm→∞µm =µ weakly implies limm→∞µˆm = ˆµ uniformly on every compact subset ofRn.

(ii) Conversely, if(ˆµm)mNconverges pointwise to some functionϕ:Rn→Cwhich is continuous inu= 0, then there exists a uniqueµ∈M1+(Rn)such thatµˆ =ϕ andlimm→∞µm=µweakly.

Proof. See Satz 15.23 in Klenke.

Let(Ω,A, P)be a probability space andX : Ω→Rn beA/B(Rn)-measurable. Let PX (:=P◦X1)be the distribution ofX. Then

ϕX(u) := ˆPX(u) = Z

eihu,yiPX(dy) = Z

eihu,XidP =Eh eihu,Xii is said to be the characteristic function ofX.

Remark 5.5. X1, . . . , Xn are independent if and only if Pˆ(X1,...,Xn)

| {z }

(X1,...,Xn)

(u1, . . . , un) = Yn j=1

Xj(uj)

| {z }

Xj(uj)

= (PX1⊗ · · · ⊗ˆ PXn)(u1, . . . , un) ,

i.e.: Pˆ(X1,...,Xn)= Yn j=1

Xj.

(11)

Proposition 5.6. Let X1, . . . , Xn be independent r.v.,α∈RandS :=αPn k=1Xk. Then for allu∈R:

ϕS(u) = Yn k=1

ϕXk(αu).

Proof.

ϕS(u) = Z

eiuS dP = Z Yn

k=1

eiαuXk dP Indep.= Yn k=1

Z

eiαuXk dP = Yn k=1

ϕXk(αu).

Proposition 5.7. For allu∈Rn: 1

n2 Z

eihu,yie12|y|2 dy=e12kuk2. Proof. See Satz 15.12 in Klenke.

Example 5.8. (i) δˆa(u) =eiua. (ii) Letµ:=P

i=1αiδaii≥0,P

i=1αi= 1). Then ˆ

µ(u) = X i=1

αieiuai, u∈Rn. Examples:

a) Binomial distributionβnp =Pn k=1

n k

pkqnkδk Then for allu∈R: βˆnp(u) =

Xn k=0

n k

pkqnk·eiuk= (q+peiu)n.

b) Poisson distributionπα=P

n=0eα αn!nδn. Then for allu∈R: ˆ

πα(u) =e−α X n=0

αn n! ·eiun

| {z }

=(αeiu)nn!

=eα(eiu1).

6 Central limit theorem

Definition 6.1. LetX1, X2, . . .∈L2 be independent r.v.,Sn:=X1+· · ·+Xn and Sn:= Sn−E[Sn]

pvar(Sn) ("standardization")

(so that in particularE[Sn] = 0andvar(Sn= 1)). The sequenceX1, X2, . . . of r.v. is said to have thecentral limit property (CLP), if

n→∞lim PSn=N(0,1) weakly,

(12)

or equivalently (by the Portmanteau theorem)

nlim→∞P[Sn6b] = 1

√2π Z b

−∞

ex

2

2 dx= Φ(b) ∀b∈R.

Proposition 6.2. (Central limit theorem) Let X1, X2, . . . ∈ L2 be independent r.v., σn2 := var(Xn)>0and

sn :=Xn

k=1

σ2k12 .

Assume that(Xn)nNsatisfies Lindeberg’s condition

nlim→∞

Xn k=1

Z

n|Xk−E[Xk]|

sn o

Xk−E[Xk] sn

2

dP

| {z }

=:Ln(ε)

= 0 ∀ε >0. (L)

Then(Xn)nNhas the CLP.

Remark 6.3. (i) (Xn)nNi.i.d. ⇒(Xn)nNsatisfies (L).

Proof: Letm:=E[Xn],σ2:= var(Xn). Thens2n=nσ2, so that

Ln(ε) =σ2 Z

{|X1−m| nσ}

(X1−m)2dP

Lebesgue

−−−−→n→∞ 0.

(ii) The following stronger condition, known as Lyapunov’s condition, is often easier to check in applications:

∃δ >0 : lim

n→∞

Xn k=1

Eh

Xk−E[Xk] 2+δ

i

s2+δn = 0. (Lya)

To see that Lyapunov’s condition implies Lindeberg’s condition note that for all ε >0:

Xk−E[Xk]

>εsn

Xk−E[Xk] 2+δ >

Xk−E[Xk]

2·(εsn)δ and therefore

Ln(ε)6 1 εδs2+δn

Xn k=1

Eh

Xk−E[Xk] 2+δ

i .

(13)

(iii) Let(Xn)be bounded and suppose thatsn→ ∞. Then(Xn)satisfies Lyapunov’s condition for anyδ >0, because

|Xk|6 α 2

Xk−E[Xk] 6α

⇒ 1

s2+δn

Xn k=1

Eh

Xk−E[Xk] 2+δ

i

6 αδ s2+δn

Xn k=1

Eh

Xk−E[Xk] 2

i

| {z }

=s2n

= α

sn

δ

.

Lemma 6.4. Suppose that (Xn)satisfies Lindeberg’s condition. Then

nlim→∞ max

16k6n

σk

sn

= 0. (2.1)

Proof. For all16k6n σk

sn

2

=

Z Xk−E[Xk] sn

2

dP 6 Z

n|Xk−E[Xk]|

sn o

Xk−E[Xk] sn

2

dP+ε2

6Ln(ε) +ε2.

The proof of Proposition 6.2 requires some further preparations.

Lemma 6.5. For allt∈Randn∈N:

eit−1−it 1!−(it)2

2! − · · · − (it)n−1 (n−1)!

6|t|n

n! .

Proof. Definef(t) :=eit, thenf(k)(t) =ikeit, and the Taylor series expansion, applied to real and imaginary part, implies that

eit−1− · · · − (it)n1 (n−1)!

=

Rn(t) with

Rn(t) =

1 (n−1)!

Z t 0

(t−s)n1ineisds 6 1

(n−1)!

Z |t| 0

sn1ds= |t|n n! . Proposition 6.6. Let X ∈ L2. Then ϕX(u) =R

eiuXdP is two times continuously differentiable with

ϕX(u) =i Z

XeiuX dP , ϕ′′X(u) =− Z

X2eiuXdP .

(14)

In particular

ϕX(0) =i·E[X], ϕ′′X(0) =−E[X2], |ϕ′′X|6E[X2].

Moreover, for allu∈R

ϕX(u) = 1 +iu·E[X] + 1

2·θ(u)u2·E[X2] with

θ(u)

61 andθ(u)∈C. Proof. Clearly,

(eiuX) =iX·eiuX, (eiuX)′′=−X2eiuX, |eiuX|= 1.

Now, Lebesgue’s dominated convergence theorem implies all assertions up to the last one. For the proof of the last assertion note that the previous lemma implies in the casen= 2that

|eiuX−1−iuX|6 1

2 ·u2X2. Integration w.r.t. P now implies that

Z

eiuX−1−iuX dP =

ϕX(u)−1−iu·E[X] 6 1

2·u2·E[X2].

From now on assume thatX1, X2,· · · ∈L2are independent and E[Xn] = 0∀n, σ2n:= var(Xn)>0, sn=Xn

k=1

σk212 .

Proposition 6.7. Suppose that

(a) lim

n→∞ max

1≤k≤n

σk

sn

= 0 and

(b) lim

n→∞

Xn k=1

ϕXk

u sn

−1

!

=−1

2u2 ∀u∈R. Then(Xn)has the CLP.

Proof. It is sufficient to show that

n→∞lim Yn k=1

ϕXk

u sn

=e12u2. (2.2)

because forSn =s1

n

Pn

k=1Xk we have that ϕSn(u) =

Yn k=1

ϕXk

u sn

,

(15)

andϕSn(u)−−−−→n→∞ e12u2=N\(0,1)(u)pointwise, implies by Lévy’s continuity theorem thatlimn→∞PSn =N(0,1)weakly.

For the proof of (2.6) we need to show that for allu∈R

n→∞lim Yn k=1

ϕXk

u sn

− Yn k=1

exp

ϕXk

u sn

−1

| {z }

=exp[P

···]exp[12u2]

!

= 0.

To this end fixu∈Rand note that|ϕXk|61, hence

exp

ϕXk

u sn

−1

= exp

ReϕXk

u sn

−1

61.

Note that fora1, . . . , an, b1, . . . , bn ∈ z∈C

|z|61

Yn k=1

ak− Yn k=1

bk

=

(a1−b1)·a2· · ·an+b1·(a2−b2)·a3· · ·an+. . . +b1· · ·bn−1·(an−bn)

6

Xn k=1

|ak−bk|.

Consequently,

Yn k=1

ϕXk

u sn

− Yn k=1

exp

ϕXk

u sn

−1 6

Xn k=1

ϕXk

u sn

−exp

ϕXk

u sn

−1

=:Dn.

If we definezk:=ϕXk(sun)−1, we can write Dn =

Xn k=1

|zk+ 1−ezk|.

Note thatE[Xk] = 0andE[Xk2] =σ2k. The previous proposition now implies that for allk

|zk|61 2

u sn

2

σ2k.

Forε >0we can find δ >0such that

|z+ 1−ez|6ε|z| ∀ |z|< δ .

(16)

Letn0∈Nbe such that for alln>n0

u2 2

σk

sn

2

< δ

for alln≥n0. Then, for all n>n0

Dn 6ε Xn k=1

|zk|6εu2 2

Xn k=1

σk2

s2n =ε· u2 2 . Consequently,limn→∞Dn= 0.

Proof of Proposition 6.2. It remains to show that Lindeberg’s condition implies

nlim→∞

Xn k=1

ϕXk

u sn

−1

!

=−1 2·u2.

W.l.o.g. assume that E[Xn] = 0 for all n ∈ N. Let u ∈ R, n ∈ N, 1 6 k 6 n.

Lemma 6.5 implies that Yk:=

exp

i· u

sn ·Xk

−1−i· u sn ·Xk

| {z }

E[...]=0

+1 2·u2

s2n ·Xk2

6 1 6

u sn ·Xk

3

,

and Yk6

exp

i· u sn ·Xk

−1−i· u sn ·Xk

+1 2 ·u2

s2n ·Xk26 u2 s2n ·Xk2. With these notations

Xn k=1

ϕXk

u sn

−1

! +1

2·u2

=

Xn k=1

ϕXk

u sn

−1 +1 2 ·u2

s2n ·σk2

! 6

Xn k=1

E[Yk].

Forδ >0 E[Yk] =

Z

{|Xk|>δsn}

Yk dP+ Z

{|Xk|<δsn}

Yk dP

6u2 s2n

Z

{|Xk|>δsn}

Xk2dP+|u|3 6s3n

Z

{|Xk|<δsn}|Xk|3dP.

Note that 1 s3n

Z

{|Xk|<δsn}|Xk|3dP 6 δ s2n

Z

Xk2dP =δ· σk2 s2n,

(17)

so that forε >0andδ >0 with |u6|3δ < ε2, we obtain that Xn

k=1

E[Yk]6u2 Xn k=1

Z

{|Xksn|}

Xk

sn

2

dP+|u|3 6 ·δ

Xn k=1

σ2k s2n

| {z }

=1

6u2Ln(δ) +ε 2.

Note thatu2Ln(δ)< ε2 for largen, so that

nlim→∞

Xn k=1

E[Yk] = 0,

and thus

nlim→∞

Xn k=1

ϕXk

u sn

−1

+1 2·u2

!

= 0.

Now the assertion follows from Proposition 6.7.

Example 6.8(Applications). (i) "Ruin probability"

Consider a portfolio of n contracts of a risc insurance (e.g. car insurance, fire insurance, health insurance, ...). LetXi>0 be the claim size (or claim severity) of the ithcontract,16i6n. We assume thatX1, . . . , Xn∈L2 are i.i.d. with m:=E[Xi] andσ2:= var(Xi).

Suppose the insurance holder has to pay he following premium Π :=m+λσ2

=average claim size+safety loading.

After some fixed amount of time:

Income: nΠ

Expenditures:Sn= Xn i=1

Xi.

Suppose that K is the initial capital of the insurance company. What is the probabilityP(R), where

R:={Sn > K+nΠ} denotes the ruin ? We assume here that:

• No interest rate.

• Payments due only at the end of the time period.

(18)

Let

Sn :=Sn−nm

√nσ .

The central limit theorem implies for largenthatSn∼N(0,1), so that P(R) =P

Sn> K+nΠ−nm

√nσ

=P

Sn >K+nλσ2

√nσ

CLT≈ 1−Φ

K+nλσ2

√nσ

| {z }

−−−−→n→∞

,

where Φdenotes the distribution of the standard normal distribution. Note that the ruin probability decreases with an increasing number of contracts.

Example

Assume thatn= 2000,σ= 60,λ= 0.5‰.

(a) K= 0 ⇒P(R)≈1−Φ(1.342)≈9%.

(b) K= 1500⇒P(R)≈3%.

How large do we have to choosenin order to let the probability of ruinP(R)fall below1‰?

Answer: Φ(. . .)>0.999, hencen>10 611.

(ii) Stirling’s formula

Remark: Stirling proved the following formula n!≈√

2πnn+12en (2.3)

in the year 1730 and De Moivre used it in his proof of the CLT for Bernoulli experiments.

Conversely, in 1977, Weng provided an independent proof of the formula, using the CLT (note that we did not use Stirling’s formula in our proof of the CLT).

Here is Weng’s proof:

LetX1, X2, . . . be i.i.d. with distribution π1, i.e., PXn =e1

X k=0

1 k!δk.

ThenSn :=X1+· · ·+Xn has Poisson distributionπn, i.e., PSn=e−n

X k=0

nk k!δk,

(19)

and in particularE[Sn] = var(Sn) =n. As usual, define Sn :=Sn−n

√n ,

so thatSn=tn◦Sn fortn(x) := xn n. Then Z

f dPSn=E f(Sn)

=E

(f◦tn)(Sn)

= Z

f◦tn dPSn

|{z}

n

In particular, for

f(x) :=x = (−x)∨0 it follows that

Z

fdPSn= Z

R

f x−n

√n

| {z }

(= 0ifx>n

=nnx ifx6n

πn(dx) =en Xn k=0

nk

k! · n−k

√n

| {z }

=f(k)

= en

√n ·

n+ Xn k=1

nk(n−k) k!

= en

√n ·

n+ Xn k=1

nk+1

k! − nk (k−1)!

| {z }

=nnn!+1n0!1

= en·nn+12 n! .

Moreover, Z

fdN(0,1) = 1

√2π Z 0

−∞

(−x)·ex

2

2 dx= 1

√2π·ex

2 2

0

−∞

= 1

√2π.

Hence, Stirling’s formula (2.7) would follow, once we have shown that Z

fdPSn n→∞

−−−−→

Z

fdN(0,1). (2.4)

Note that this is not implied by the weak convergence in the CLT since f is continuous but unbounded. Hence, we consider for givenm∈N

fm:=f∧m ∈Cb(R) . The CLT now implies that

Z

fmdPSn n→∞

−−−−→

Z

fmdN(0,1).

(20)

Definegm :=f−fm (≥0). (2.8) then follows from a "3ε-argument", once we have shown that

(06) Z

gmdPSn6 1

m ∀m,

(06) Z

gmdN(0,1)6 1

m ∀m.

The first inequality follows from Z

gmdPSn = Z

]−∞,m[ |x| −m

dPSn6 Z

]−∞,m]|x|dPSn

|x|

m>1

6 1

m Z

]−∞,m]

x2dPSn 6 1

m ·var(Sn)

| {z }

=1

,

the second inequality can be shown similarly.

Referenzen

ÄHNLICHE DOKUMENTE

Elstrodt, Maß- und Integrationstheorie, Springer, Berlin, first edition 1996, fifth edition 2007J. Stute, Wahrscheinlichkeitstheorie, Springer,

Motivated by the Examples II.5.2 and II.6.1 we introduce a notion of convergence that is weaker than convergence in mean and convergence almost surely.. In the sequel, X, X n

Billingsley, Probability and Measure, Wiley, New York, first edition 1979, third edition 1995.. Elstrodt, Maß- und Integrationstheorie, Springer, Berlin, first edition 1996,

TU Darmstadt Fachbereich Mathematik.

This text is a summary of the lecture on Probability Theory held at the TU Darmstadt in Winter Term 2007/08.. Please email all misprints and

and observe that male students mainly applied at faculties with a high probability of admission, whereas female students mainly applied at faculties with a low probability

The Radon-Nikodym theorem will be used to obtain a second, independent proof for the existence of the conditional expectation. We will prove the theorem in the case of

‘⇒’ holds trivially. Analogously to the construction of the Lebesgue measure; see ¨