• Nem Talált Eredményt

Finally, we provide the missing results and proofs of the paper. We start with the following results from Section 2.

Proof of Lemma 2.2. We refer to Section 6.1 of Carassus and R ´asonyi (2015) for the definition and various properties of generalized conditional expectations. In particular sinceE(h+) =R

th+dPt<∞, E(h|Fs)is well-defined (in the generalised sense) for all0≤s≤t(see Lemma 6.2 of Carassus and R ´asonyi (2015) ). Similarly, from Proposition 7.4 we have thatϕ: Ωs→ R∪ {±∞}is well-defined (in the gen-eralised sense) andFs-measurable.

Asϕ(X1, . . . , Xs)isFs-measurable, it remains to prove thatE(gh) =E(gϕ(X1, . . . , Xs))for allg: Ωs→ R+non-negative,Fs-measurable and such thatE(gh)is well-defined in the generalised sense,i.esuch thatE(gh)+ <∞orE(gh)<∞. Recalling the notations of the beginning of Section 2 and using the Fubini Theorem for the third and fourth equality (see Proposition 7.4 and Remark 7.5), we get that

E(gh) = E(g(X1, . . . , Xs)h(X1, . . . , Xt)) = Z

T

g(ω1, . . . , ωs)h(ω1, . . . , ωt)P(dωT)

= Z

t

g(ω1, . . . , ωs)h(ω1, . . . , ωt)qttt−1). . . qs+1s+1s)Ps(dωs)

= Z

s

g(ω1, . . . , ωs) Z

s+1×...×Ωt

h(ω1, . . . , ωs, ωs+1, . . . , ωt)qttt−1). . . qs+1s+1s)

!

Ps(dωs)

= Z

s

g(ω1, . . . , ωs)ϕ(ω1, . . . , ωs)Ps(dωs)

= E(g(X1, . . . , Xs)ϕ(X1, . . . , Xt)),

which concludes the proof. ✷

We give now the proof of results of Section 3.

Proof of Lemma 3.4. We first prove that Det+1 is a non-empty, closed-valued and Ft-measurable random set. It is clear from its definition (see (2)) that for allωt ∈ Ωt,Det+1t) is a non-empty and closed subset of Rd. We now show that Det+1 is measurable. Let O be a fixed open set in Rd and introduce

µOt∈Ωt→µOt) := qt+1 ∆St+1t, .)∈O|ωt

= Z

t+1

1∆St+1(·,·)∈Ot, ωt+1)qt+1(dωt+1t).

We prove thatµOisFt-measurable. As(ωt, ωt+1)∈Ωt×Ωt+1 →∆St+1t, ωt+1)isFt⊗Gt+1-measurable andO ∈ B(Rd),(ωt, ωt+1) →1∆St+1(·,·)∈Ot, ωt+1)isFt⊗ Gt+1-measurable and the result follows from Proposition 7.9.

By definition ofDet+1t)we get that

t∈Ωt, Det+1t)∩O 6=∅}={ωt∈Ωt, µOt)>0} ∈ Ft.

Next we prove that Dt+1 is a non-empty, closed-valued andFt-measurable random set. Using (3), Dt+1 is a non-empty and closed-valued random set. It remains to prove thatDt+1 isFt-measurable.

As Det+1 is Ft-measurable, applying the Castaing representation (see Theorem 2.3 in Chapter 1 of Molchanov (2005) or Theorem 14.5 of Rockafellar and Wets (1998)), we obtain a countable family of Ft-measurable functions(fn)n≥1 : Ωt→Rdsuch that for allωt∈Ωt,Det+1t) ={fnt), n≥1}(where the closure is taken in Rdwith respect to the usual topology). Let ωt ∈ Ωt be fixed. It can be easily shown that

Dt+1t) =Aff(Det+1t)) = (

f1t) + Xp

i=2

λi(fit)−f1t)), (λ2, . . . , λp)∈Qp−1, p≥2 )

. (66) So, using again the Castaing representation (see Theorem 14.5 of Rockafellar and Wets (1998)), we ob-tain thatDt+1t)isFt-measurable. From Theorem 14.8 of Rockafellar and Wets (1998),Graph(Dt+1)∈ Ft⊗ B(Rd)(recall thatDt+1is closed-valued). ✷ Proof of Lemma 3.5. Introduce Ct+1t) := Conv(Det+1t)) the closed convex hull generated by Det+1t). As Ct+1t) ⊂ Dt+1t) we will prove that 0 ∈ Ct+1t). Since Ct+1t) ⊂ Dt+1t) by assumption, for allh∈Ct+1t)\{0}

qt+1(h∆St+1t,·)≥0|ωt)<1. (67) Thus if we find someh0 ∈ Ct+1t) such thatqt+1(h0∆St+1t,·) ≥ 0|ωt) = 1 thenh0 = 0. We distin-guish two cases. First assume that for allh∈Rd,h6= 0,qt+1(h∆St+1t, .)≥0|ωt)<1. Then the polar cone ofCt+1t),i.ethe set

Ct+1t)

:={y∈Rd, yx≤0,∀x∈Ct+1t)}

is reduced to{0}. Indeed if this is not the case there exists y0 ∈ Rd such that −y0x ≥ 0 for all x ∈ Ct+1t). AsA:={ωt+1 ∈Ωt+1, ∆St+1t, ωt+1) ∈Det+1t)} ⊂ {ωt+1 ∈Ωt+1, −y0∆St+1t, ωt+1)≥0}

andqt+1(A|ωt) = 1we obtain thatqt+1(−y0∆St+1t,·)≥0|ωt) = 1a contradiction. As Ct+1t)

= cone Ct+1t)

where cone Ct+1t)

denote the cone generated byCt+1t)we get that cone Ct+1t)

= Rd. Let u 6= 0 ∈ cone Ct+1t)

then −u ∈ cone Ct+1t)

and there exist λ1 > 0, λ2 > 0 and v1, v2∈Ct+1t)such thatu=λ1v1and−u=λ2v2. Thus0 = λ1λ1 2v1+λ1λ2 2v2 ∈Ct+1t)by convex-ity ofCt+1t).

Now we assume that there exists someh0∈Rd,h0 6= 0such thatqt+1(h0∆St+1t, .)≥0|ωt) = 1. Note that since h0 ∈ Rd we cannot use (67). Introduce the orthogonal projection on Ct+1t) (recall that Ct+1t)is a closed convex subset ofRd)

p:h∈Rd→p(h)∈Ct+1t).

Then p is continuous and we have(h−p(h)) (x−p(h)) ≤ 0 for all x ∈ Ct+1t). Fixωt+1 ∈ {ωt+1 ∈ Ωt+1, ∆St+1t, ωt+1)∈Det+1t)} ∩ {ωt+1 ∈Ωt+1, h0∆St+1t, ωt+1) ≥0}andλ≥0. Leth =λh0 and x= ∆St+1t, ωt+1)∈Ct+1t)in the previous equation, we obtain (recall thatDet+1t)⊂Ct+1t))

0≤λh0∆St+1t, ωt+1) = (λh0−p(λh0)) ∆St+1t, ωt+1) +p(λh0)∆St+1t, ωt+1)

≤(λh0−p(λh0))p(λh0) +p(λh0)∆St+1t, ωt+1).

As this is true for allλ≥0we may take the limit whenλgoes to zero and use the continuity ofp p(0)∆St+1t, ωt+1)≥ |p(0)|2 ≥0

Asqt+1

n

ωt+1∈Ωt+1, ∆St+1t, ωt+1)∈Det+1t)o

t

= 1by definition ofDet+1t)and as qt+1(h0∆St+1t, .)≥0|ωt) = 1as well we have obtained that

qt+1(p(0)∆St+1t,·)≥0|ωt) = 1.

The fact thatp(0)∈Ct+1t)together with (67) implies thatp(0) = 0and0∈Ct+1t)follows.

The following lemma has been used in the proof of Lemma 3.6. It corresponds to Lemma 2.5 of Nutz (2014)

Lemma 7.18 Let ωt ∈ Ωt be fixed. Recall that Lt+1t) := Dt+1t)

is the orthogonal space of Dt+1t)(see (6)). Then forh∈Rdwe have that

qt+1(h∆St+1t,·) = 0|ωt) = 1 ⇐⇒ h∈Lt+1t).

Proof. Assume thath ∈Lt+1t). Then{ω ∈Ωt, ∆St+1t, ω) ∈Dt+1t)} ⊂ {ω ∈Ωt, h∆St+1t, ω) = 0}. As by definition ofDt+1t),qt+1(∆St+1t, .)∈Dt+1t)|ωt) = 1, we conclude thatqt+1(h∆St+1t, .) = 0|ωt) = 1. Conversely, we assume thath /∈ Lt+1t) and we show thatqt+1(h∆St+1t, .) = 0|ωt) <1.

We first show that there existsv ∈Det+1t)such thathv 6= 0. If not, for allv∈Det+1t),hv = 0and for anyw ∈ Dt+1t) with w = Pm

i=1λivi where λi ∈ R,Pm

i=1λi = 1 andvi ∈ Det+1t), we get that hw= 0, a contradiction. Furthermore there exists an open ball centered invwith radiusε >0,B(v, ε), such thathv 6= 0 for allv ∈ B(v, ε). Assume that qt+1(∆St+1t, .) ∈ B(v, ε)|ωt) = 0or equivalently thatqt+1(∆St+1t, .) ∈Rd\B(v, ε)|ωt) = 1. By definition of the support, Det+1t)⊂Rd\B(v, ε): this contradictsv∈Det+1t). Thereforeqt+1(∆St+1t, .)∈B(v, ε)|ωt)>0. Letω ∈ {∆St+1t, .)∈B(v, ε)}, thenh∆St+1t, ω)6= 0i.e qt+1(h∆St+1t, .) = 0|ωt))<1. ✷

We prove now the following result of Section 5.

Proof of Proposition 5.11. We start with the proof of (25) whenh∈Dx. SinceDis a vectorial subspace ofRdand0 ∈ Hx, the affine hull ofDx is also a vector space that we denote by Aff(Dx). Ifx ≤ 1we have by Assumption 5.4 that for allω∈Ω,h∈Dx,

V+(ω, x+hY(ω))≤V+(ω,1 +hY(ω)). (68) Ifx >1using Assumption 5.7 (see (23) in Remark 5.8) we get that for allω ∈Ω,h∈Dx

V+(ω, x+hY(ω)) =V+

2x 1

2 + h 2xY(ω)

≤(2x)γK

V+

ω,1 + h 2xY(ω)

+C(ω)

. (69) First we treat the case ofDim(Aff(Dx)) = 0,i.e Dx ={0}. For allω ∈Ω,h∈Dx ={0}, using (68) and (69), we obtain that

V+(ω, x+hY(ω))≤V+(ω,1) + (2x)γK V+(ω,1) +C(ω)

≤((2x)γK+ 1)(V+(ω,1) +C(ω)). (70) We assume now thatDim(Aff(Dx))>0. Ifx= 0thenY = 0Q-a.s. If this is not the case then we should haveD0 ={0}a contradiction. Indeed if there exists someh∈D0withh6= 0, thenQ

h

|h|Y(·)<0

>0 by Assumption 5.1 which contradictsh ∈D0. So forx= 0,Y = 0Q-a.s and by Assumption 5.4 we get that for allω∈Ω,h∈D0,

V+(ω,0 +hY(ω))≤V+(ω,1).

From now we assume thatx > 0. Then as for g ∈ Rd, g ∈ Dx if and only if gx ∈ D1, we have that Aff(Dx) = Aff(D1). We set d := Dim(Aff(D1)). Let (e1, . . . , ed) be an orthonormal basis of Aff(D1) (which is a sub-vector space of Rd) and ϕ : (λ1, . . . , λd) ∈ Rd → Σdi=1 λiei ∈ Aff(D1). Then ϕ is an isomorphism (recall that(e1, . . . , ed) is a basis of Aff(D1)). As ϕis linear and the spaces considered are of finite dimension, it is also an homeomorphism betweenRd and Aff(D1). Since D1 is compact by Lemma 5.10,ϕ−1(D1) is a compact subspace ofRd . So there exists some c ≥ 0 such that for all h= Σdi=1 λiei ∈D1,|λi| ≤cfor alli= 1, . . . , d. We complete the family of vector (e1, . . . , ed)in order to obtain an orthonormal basis ofRd, denoted by(e1, . . . , ed, ed+1, . . . ed). For allω ∈Ω, let(yi(ω))i=1,...,d be the coordinate ofY(ω)in this basis.

Now leth∈Dxbe fixed. Then 2xh ∈D1

2 ⊂D1 and 2xh = Σdi=1 λiei for some(λ1, . . . λd)∈Rd with|λi| ≤c for alli= 1, . . . , d. Note that as 2xh ∈ D1i = 0fori≥d+ 1. Then as(e1, . . . , ed)is an orthonormal basis ofRd, we obtain for allω∈Ω

1 + h

2xY(ω) = 1 + Σdi=1 λiyi(ω)

≤1 + Σdi=1i||yi(ω)|

≤1 +cΣdi=1 |yi(ω)|.

Thus from Assumption 5.4 for allω∈Ωwe get that V+

ω,1 + h 2xY(ω)

≤V+

ω,1 +cΣdi=1 |yi(ω)|

. We set

L(·) :=V+

ω,1 +cΣdi=1 |yi(ω)|

1d>0+V+(·,1) +C(·).

Asd =Dim(Aff(D1))it is clear thatLdoes not depend onx. It is also clear thatLisH-measurable.

Then using (68), (69) and (70) we obtain that for allω∈Ω

V+(ω, x+hY(ω))≤((2x)γK+ 1)L(ω).

Note that the first term in L is used in the above inequality if x 6= 0 andDim(Aff(Dx)) > 0. The second and the third one are there for both the case ofDim(Aff(Dx)) = 0and the case of x = 0 and Dim(Aff(Dx))> 0. As by Assumptions 5.7 and 5.9, E(V+(·,1) +C(·))< ∞, it remains to prove that d >0impliesE

V+

·,1 +cΣdi=1 |yi(·)|

<∞.

IntroduceW, the finite set of Rd whose coordinates on(e1, . . . , ed) are1 or−1and0on (ed+1, . . . ed).

ThenW ⊂Aff(D1) and the vectors ofW will be denoted byθj forj ∈ {1, . . . ,2d}. Letθω be the vector whose coordinates on(e1, . . . , ed) are(sign(yi(ω)))i=1...d and0on (ed+1, . . . ed). Thenθω ∈W and we get that

V+

ω,1 +cΣdi=1 |yi(ω)|

=V+(ω,1 +cθωY(ω))≤

2d

X

j=1

V+(ω,1 +cθjY(ω)).

So to prove thatEL <∞it is sufficient to prove that ifd >0for all1≤j≤2d,EV+(·,1+cθjY(·))<∞.

Recall thatθj ∈Aff(D1). Letri(D1) ={y ∈D1,∃α >0s.tAff(D1)∩B(y, α)⊂D1}5denote the relative interior ofD1. AsD1 is convex and non-empty (recalld >0),ri(D1)is also non-empty and convex and we fix somee ∈ ri(D1). We prove that e2 ∈ ri(D1). Let α > 0 be such that Aff(D1)∩B(e, α) ⊂ D1 andg∈Aff(D1)∩B(e2,α2). Then2g∈Aff(D1)∩B(e, α)(recall that Aff(D1)is actually a vector space) and thus2g∈D1. AsD1 is convex and0∈D1, we get thatg∈D1 and Aff(D1)∩B(e2,α2)⊂D1 which proves that e2 ∈ ri(D1). Now letεj be such thatεj(c2θje2) ∈B(0,α2). It is easy to see that one can

5HereB(y, α)is the ball ofRdcentered atyand with radiusα.

choseεj ∈(0,1). Then as¯ej := e2+ε2j(cθj−e)∈Aff(D1)∩B(e2,α2)(recall thatθj ∈W ⊂Aff(D1)), we the monotonicity property ofV in Assumption 4.1. Note that the above inequalities are true even if 1 +cθjY(ω)<0since (23) (see remark 5.8) and the monotonicity property ofV hold true for allx∈R.

From Assumption 5.9 we get thatEV+(·,1 + ¯ejY(·)) < ∞ (recall that e¯j ∈ D1) and Assumption 5.7 impliesEC <∞, thereforeEV+(·,1 +cθjY(·))<∞and (25) is proven forh∈Dx. Now leth∈ Hx and h its orthogonal projection on D, thenhY(·) = hY(·)Q-a.s (see Remark 5.3). It is clear thath ∈ Dx

thusV+(·, x+hY(·)) =V+(·, x+hY(·))Q-a.s and (25) is true also forh∈ Hx. ✷ To conclude, the following lemma was used in the proof of Theorem 4.16.

Lemma 7.19 Assume that (NA) holds true. Letφ∈Φsuch thatVTx,φ≥0P-a.s, thenVtx,φ≥0Pt-a.s.

L. Carassus thanks LPMA (UMR 7599) for support. M. R ´asonyi was supported by the “Lend ¨ulet”

grant LP2015-6 of the Hungarian Academy of Sciences.

References

C. D. Aliprantis and K. C. Border. Infinite Dimensional Analysis : A Hitchhiker’s Guide.

Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sci-ences]. Springer-Verlag, Berlin, 3rd edition, 2006.

Alain Bensoussan, Abel Cadenillas, and Hyeng Keun Koo. Entrepreneurial decisions on effort and project with a nonconcave objective function. Math. Oper. Res., 40(4):902–914, 2015. doi: 10.1287/

moor.2014.0702. URLhttp://dx.doi.org/10.1287/moor.2014.0702.

V.I. Bogachev. Measure Theory, vol 2. Springer-Verlag, Berlin, 2007.

L. Carassus and M. R ´asonyi. Maximization of non-concave utility functions in discrete-time financial market. Mathematics of Operations Research, Published Online ISSN 1526-5471, 2015.

L. Carassus, R ´asonyi M., and A. M. Rodrigues. Non-concave utility maximisation on the positive real axis in discrete time. Mathematics and Financial Economics, 9(4):325–349, 2015.

G. Carlier and R.-A. Dana. Optimal demand for contingent claims when agents have law invariant utilities. Math. Finance, 21:169–201, 2011.

C. Castaing and M. Valadier.Convex Analysis and Measurable Multifunctions, volume 580. Springer, Berlin, 1977.

R.C. Dalang, A. Morton, and W. Willinger. Equivalent martingale measures and no-arbitrage in stochastic securities market models. Stochastics Stochastics Rep., 29:185–201, 1990.

C. Dellacherie and P.-A. Meyer. Probability and potential. North-Holland, Amsterdam, 1979.

H. F¨ollmer and A. Schied. Stochastic Finance: An Introduction in Discrete Time. Walter de Gruyter

& Co., Berlin, 2002.

X. He and X.Y. Zhou. Portfolio choice under cumulative prospect theory: An analytical treatment.

Management Science, 57:315–331, 2011.

J. Jacod and A. N. Shiryaev. Local martingales and the fundamental asset pricing theorems in the discrete-time case. Finance Stoch., 2:259–273, 1998.

H. Jin and X.Y. Zhou. Behavioural portfolio selection in continuous time. Math. Finance, 18:385–426, 2008.

D. O. Kramkov and W. Schachermayer. The asymptotic elasticity of utility functions and optimal investment in incomplete markets. Ann. Appl. Probab., 9:904–950, 1999.

I. Molchanov. Theory of random sets. Springer-Verlag, London, 2005.

M. Nutz. Utility maximisation under model uncertainty in discrete time. Math. Finance. Published Online DOI: 10.1111/mafi.12068, 2014.

M. R ´asonyi and L. Stettner. On the utility maximization problem in discrete-time financial market models. Ann. Appl. Probab., 15:1367–1395, 2005.

M. R ´asonyi and L. Stettner. On the existence of optimal portfolios for the utility maximization problem in discrete time financial models. In: Kabanov, Y.; Lipster, R.; Stoyanov,J. (Eds), From Stochastic Calculus to Mathematical Finance, Springer., pages 589–608, 2006.

R. T. Rockafellar and R. J.-B. Wets. Variational analysis, volume 317 ofGrundlehren der Mathematis-chen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, Berlin, 1998. ISBN 3-540-62772-3.

M.-F. Sainte-Beuve. On the extension of von Neumann-Aumann’s theorem. J. Functional Analysis, 17(1):112–129, 1974.

KAPCSOLÓDÓ DOKUMENTUMOK