Skip to main content
Log in

Optimal portfolio choice for a behavioural investor in continuous-time markets

  • Symposium
  • Published:
Annals of Finance Aims and scope Submit manuscript

Abstract

The aim of this work consists in the study of the optimal investment strategy for a behavioural investor, whose preference towards risk is described by both a probability distortion and an S-shaped utility function. Within a continuous-time financial market framework and assuming that asset prices are modelled by semimartingales, we derive sufficient and necessary conditions for the well-posedness of the optimisation problem in the case of piecewise-power probability distortion and utility functions. Finally, under straightforwardly verifiable conditions, we further demonstrate the existence of an optimal strategy.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

Notes

  1. Note that it does not necessarily have to be a common stock, it can also be referring to a commodity, a foreign currency, an exchange rate or a market index.

  2. We recall that, for every real number \(x\), the cumulative distribution function of \(\rho \) with respect to the probability measure \(\mathbb Q \) is given by \(F_{\rho }^\mathbb{Q }\!\left(x\right)=\mathbb Q \!\left(\rho \le x\right)\).

  3. Note that here \(x^{+}=\max \left\{ x,0\right\} \) and \(x^{-}=-\min \left\{ x,0\right\} \).

  4. We denote by \(A^{c}\) the complement of \(A\) in \(\varOmega \).

  5. Here \(\fancyscript{B}\!\left(X\right)\) denotes the Borel \(\sigma \)-algebra on the topological space \(X\). A mapping \(K\) from \(\mathbb R \times \fancyscript{B}\!\left(\mathbb R \right)\) into \(\left[0,+\infty \right]\) is called a transition probability kernel on \(\left(\mathbb R ,\fancyscript{B}\!\left(\mathbb R \right)\right)\) if the mapping \(x \mapsto K\!\left(x,B\right)\) is measurable for every set \(B \in \fancyscript{B}\!\left(\mathbb R \right)\), and the mapping \(B \mapsto K\!\left(x,B\right)\) is a probability measure for every \(x \in \mathbb R \).

References

  • Berkelaar, A.B., Kouwenberg, R., Post, T.: Optimal portfolio choice under loss aversion. Rev Econ Stat 86(4), 973–987 (2004)

    Article  Google Scholar 

  • Bernard, C., Ghossoub, M.: Static portfolio choice under cumulative prospect theory. Math Financ Econ 2, 277–306 (2010)

    Article  Google Scholar 

  • Borkar, V.S.: Probability Theory: An Advanced Course. Berlin: Springer (1995)

  • Campi, L., Del Vigna, M.: Weak insider trading and behavioural finance. SIAM J Financ Math 3, 242–279 (2012)

    Article  Google Scholar 

  • Carassus, L., Rásonyi, M.: On optimal investment for a behavioural investor in multiperiod incomplete market models. Preprint version 1. Forthcoming in Math Finance. (2011) Available online at http://arxiv.org/abs/1107.1617

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

    Google Scholar 

  • Harrison, J.M., Pliska, S.R.: Martingales and stochastic integrals in the theory of continuous trading. Stoch Process Appl 11, 215–260 (1981)

    Article  Google Scholar 

  • He, X.D., Zhou, X.Y.: Portfolio choice under cumulative prospect theory: an analytical treatment. Manag Sci 57(2), 315–331 (2011)

    Article  Google Scholar 

  • Jin, H., Zhou, X.Y.: Behavioral portfolio selection in continuous time. Math Financ 18, 385–426 (2008)

    Article  Google Scholar 

  • Kahneman, D., Tversky, A.: Prospect theory: an analysis of decision under risk. Econometrica 47(2), 263–292 (1979)

    Article  Google Scholar 

  • Karatzas, I., Shreve, S.E.: Methods of Mathematical Finance, Applications of Mathematics., vol. 39, 2nd edn. Berlin: Springer (1998)

  • Kast, R., Lapied, A., Toquebeuf, P.: Updating choquet integrals, consequentialism and dynamic consistency. In: ICER Working Papers—Applied Mathematics Series. Available at http://ideas.repec.org/p/icr/wpmath/04-2008.html (2008)

  • Prokhorov, Y.V.: Convergence of random processes and limit theorems in probability theory. Theor Probab Appl 1(2), 157–214 (1956)

    Article  Google Scholar 

  • Rásonyi, M., Stettner, L.: On the utility maximization problem in discrete-time financial market models. Ann Appl Probab 15, 1367–1395 (2005)

    Article  Google Scholar 

  • Reichlin, C.: Non-concave utility maximization with a given pricing measure. Available at SSRN http://ssrn.com/abstract=1940277 (2011)

  • Schachermayer, W.: Optimal investment in incomplete markets when wealth may become negative. Ann Appl Probab 11(3), 694–734 (2001)

    Article  Google Scholar 

  • Tversky, A., Kahneman, D.: Advances in prospect theory: cumulative representation of uncertainty. J Risk Uncertain 5, 297–323 (1992)

    Article  Google Scholar 

  • von Neumann, J., Morgenstern, O.: Theory of Games and Economic Behavior, 2nd edn. Princeton: Princeton University Press (1953)

Download references

Acknowledgments

AMR is sponsored by the Doctoral Grant SFRH/BD/69360/2010 from the Portuguese Foundation for Science and Technology (FCT). The authors thank an anonymous referee for his/her useful comments.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Miklós Rásonyi.

Appendix: Proofs

Appendix: Proofs

1.1 Proofs of Section 3

1.1.1 Proof of Lemma 3.6

Let the function \(f:\left[\left.0,+\infty \right)\right. \rightarrow \left[\left.0,+\infty \right)\right.\) be given by \(f\!\left(x\right) \text{:=} \mathbb E \!\left[X\wedge x\right]\). It is clear that \(f\) is a non-decreasing function of \(x\) and that \(\lim _{x \rightarrow +\infty } f\!\left(x\right)=+\infty \). Indeed, to see this let \(M>0\) be arbitrary. Since \(\mathbb E \!\left[X\wedge n\right]\rightarrow \mathbb E \!\left[X\right]=+\infty \) by the Monotone Convergence Theorem, there exists some \(n_{0} \in \mathbb N \) such that \(\mathbb E \!\left[X\wedge n\right]>M\) for all \(n \ge n_{0}\). Given that \(f\) is non-decreasing, we conclude \(f\!\left(x\right)\ge f\!\left(n_{0}\right)>M\) for all \(x\ge n_{0}\), as intended.

Moreover, \(f\) is a continuous function on \(\left[\left.0,+\infty \right)\right.\). In fact, let \(\left\{ x_{n}\right\} _{n \in \mathbb N }\) be a sequence in \(\left[\left.0,+\infty \right)\right.\) converging to some \(x\ge 0\). Being convergent, \(\left\{ x_{n}\right\} _{n \in \mathbb N }\) is bounded, that is, there exists some \(R>0\) such that \(x_{n}\le R\) for all \(n\). Hence, the sequence \(\left\{ X_{n}\right\} _{n \in \mathbb N }\) of random variables \(X_{n} \text{:=} X\wedge x_{n}\) is dominated by the integrable random variable \(X\wedge R\) and converges pointwise to \(X\wedge x\) (we recall that the minimum is a continuous function). Therefore, by Lebesgue’s Dominated Convergence Theorem, \(f\!\left(x_{n}\right)=\mathbb E \!\left[X\wedge x_{n}\right] \xrightarrow [n\rightarrow +\infty ]{} \mathbb E \!\left[X\wedge x\right]=f\!\left(x\right)\) and \(f\) is continuous as claimed.

So let us consider \(b\ge 0\) arbitrary. It is obvious that \(f\!\left(b\right)\le b\). Besides, since \(f\!\left(x\right)\rightarrow +\infty \) as \(x\rightarrow +\infty \), there exists some \(N>0\) so that \(f\!\left(x\right)>b\) for all \(x\ge N\). Hence, because \(f\!\left(b\right)\le b < f\!\left(N\right)\) and \(f\) is continuous, by the Intermediate Value Theorem (note that \(N>b\), otherwise \(f\!\left(N\right)\le f\!\left(b\right)\le b\) by the monotonicity of \(f\)) we conclude that \(b=f\!\left(a\right)\) for some \(a \in \left[b,N\right]\).\(\square \)

1.1.2 Proof of Lemma  3.12

Let \(t>0\) be arbitrary (but fixed). Then, for any nonnegative random variable \(X\), we have by the monotonicity of the integral that

$$\begin{aligned} \int \limits _{0}^{+\infty } \mathbb P \!\left\{ X^{b}>y\right\} ^{a} dy&= \int \limits _{0}^{+\infty } \mathbb P \!\left\{ \left(X^{s}\right)^{\frac{b}{s}}>y\right\} ^{a} dy \ge \int \limits _{0}^{t^{\frac{b}{s}}} \mathbb P \!\left\{ \left(X^{s}\right)^{\frac{b}{s}}>y\right\} ^{a} dy\\&\ge \int \limits _{0}^{t^{\frac{b}{s}}} \mathbb P \!\left\{ \left(X^{s}\right)^{\frac{b}{s}}>t^{\frac{b}{s}}\right\} ^{a} dy=t^{\frac{b}{s}} \mathbb P \!\left\{ \left(X^{s}\right)^{\frac{b}{s}}>t^{\frac{b}{s}} \right\} ^{a}, \end{aligned}$$

where the last inequality follows from the inclusion \(\left\{ \left(X^{s}\right)^{\frac{b}{s}}>t^{\frac{b}{s}}\right\} \subseteq \left\{ \left(X^{s}\right)^{\frac{b}{s}}>y\right\} \) for all \(0 \le y \le t^{\frac{b}{s}}\). Furthermore, \(\mathbb P \!\left\{ \left(X^{s}\right)^{\frac{b}{s}}>t^ {\frac{b}{s}}\right\} =\mathbb P \!\left\{ X^{s}>t\right\} \), so

$$\begin{aligned} \mathbb P \!\left\{ X^{s}>t\right\} \le \frac{1}{t^{\frac{b}{s a}}}\left(\int \limits _{0}^{+\infty } \mathbb P \!\left\{ X^{b}>y\right\} ^{a} dy\right)^{\frac{1}{a}}. \end{aligned}$$

Hence,

$$\begin{aligned} \mathbb E _\mathbb{P }\!\left[X^{s}\right]&= \int \limits _{0}^{\infty } \mathbb P \!\left\{ X^{s}>t\right\} dt\le 1+\int \limits _{1}^{+\infty } \mathbb P \!\left\{ X^{s}>t\right\} dt\\&\le 1+\left(\int \limits _{0}^{\infty } \mathbb P \!\left\{ X^{b}>y\right\} ^{a} dy\right)^{\frac{1}{a}}\int \limits _{1}^{+\infty } \frac{1}{t^{\frac{b}{s a}}} dt, \end{aligned}$$

and setting the constant \(D=\int _{1}^{+\infty } \frac{1}{t^{\frac{b}{s a}}} dt \in \left(0,+\infty \right)\) (recall that \(\frac{b}{s a}>1\) by hypothesis), which depends only on the parameters, completes the proof.\(\square \)

1.1.3 Proof of Lemma 3.13

We start by noticing that the hypothesis \(\alpha <\gamma \le 1\) implies that \(\frac{1}{\gamma }<\frac{1}{\alpha \gamma }\) and \(\frac{1}{\gamma }<\frac{1}{\alpha }\). Moreover, since \(\alpha <\beta \) and \(\delta <\beta \), there exists \(\eta \) so that \(\max \left\{ \alpha ,\delta \right\} <\eta <\beta \). In particular, we deduce that \(\frac{\eta }{\alpha }>1\), and thus \(\frac{\eta }{\alpha \gamma }>\frac{1}{\gamma }\). Hence, we choose \(\lambda \) so that \(\frac{1}{\gamma }<\lambda <\min \left\{ \frac{1}{\alpha },\frac{\eta }{\alpha \gamma }\right\} \). Then, given that \(\frac{1}{\lambda \alpha }>1\), there exists some \(p\) satisfying \(1<p<\frac{1}{\lambda \alpha }\). Finally, we note that \(1<\frac{\eta }{\delta }\) and \(\frac{\alpha \lambda \gamma }{\delta }<\frac{\eta }{\delta }\) (because \(\lambda < \frac{\eta }{\alpha \gamma }\), that is, \(\alpha \gamma \lambda < \eta \)), so we can take \(q\) such that \(\max \left\{ 1,\frac{\alpha \lambda \gamma }{\delta }\right\} <q<\frac{\eta }{\delta }\).

Since for all \(y\ge 1\) we have \(\mathbb{P }\left\{ \left({X}^{+}\right)^{\alpha }>y\right\} \le \frac{\mathbb{E }_{\mathbb{P }}\left[\left({X}^{+}\right)^{\alpha \lambda }\right]}{{y}^{\lambda }}\) by Chebyshev’s inequality, it follows from the monotonicity of the integral that

$$\begin{aligned} \int \limits _{0}^{+\infty } \mathbb{P }\!\left\{ \left(X^{+}\right)^{\alpha }>y\right\} ^{\gamma } dy\le 1+C_{1}\, \mathbb{E }_\mathbb{P }\!\left[\left(X^{+}\right)^{\alpha \lambda }\right]^{\gamma }, \end{aligned}$$

with the strictly positive constant \(C_{1}\) being given by \(C_{1}\text{:=} \int _{1}^{+\infty } 1/y^{\lambda \gamma } dy\) (we recall that \(\lambda \gamma >1\)). Also, applying Hölder’s inequality yields

$$\begin{aligned} \mathbb E _\mathbb{P }\!\left[\left(X^{\!+\!}\right)^{\alpha \lambda }\right]\!=\!\mathbb E _\mathbb{P }\!\left[\frac{1}{\rho ^{1/p}} \rho ^{1/p} \left(X^{\!+\!}\right)^{\alpha \lambda }\right]\le C_{2}\,\mathbb E _\mathbb{P }\!\left[\rho \left(X^{\!+\!}\right)^{\alpha \lambda p}\right]^{\frac{1}{p}}\!=\!C_{2}\,\mathbb E _\mathbb{Q } \!\left[\left(X^{\!+\!}\right)^{\alpha \lambda p}\right]^{\frac{1}{p}}, \end{aligned}$$

where the constant \(C_{2}\text{:=} \mathbb E _\mathbb{P }\!\left[\frac{1}{\rho ^{1/\left(p-1\right)}} \right]^{\frac{p-1}{p}}\) is finite and strictly positive. Thus, combining the previous equation and Jensen’s inequality for concave functions (we note that \(\alpha \lambda p<1\)), we obtain

$$\begin{aligned} \mathbb E _\mathbb{P }\!\left[\left(X^{+}\right)^{\alpha \lambda }\right]^{\gamma }&\le C_{3}\,\mathbb E _\mathbb{Q }\!\left[\left(X^{+}\right)^{\alpha \lambda p}\right]^{\frac{\gamma }{p}}\le C_{3}\,\mathbb E _\mathbb{Q }\!\left[X^{+}\right]^{\alpha \lambda \gamma }\nonumber \\&= C_{3}\left(x_{0}+\mathbb E _\mathbb{Q }\!\left[X^{-}\right] \right)^{\alpha \lambda \gamma } \le C_{4}+C_{3}\,\mathbb E _\mathbb{Q }\!\left[X^{-}\right]^{\alpha \lambda \gamma },\quad \ \ \end{aligned}$$
(7.1)

where the last inequality follows from the trivial inequality \(\left|x+y\right|^{a}\le \left|x\right|^{a}+\left|y\right|^{a}\) for \(0<a\le 1\) (notice that \(\lambda <\frac{1}{\alpha }\le \frac{1}{\alpha \gamma }\) implies \(\alpha \lambda \gamma <1\)), and \(C_{3},C_{4} \in \left(0,+\infty \right)\). Now, we again use Hölder’s inequality to see that \(\mathbb E _\mathbb{Q }\!\left[X^{-}\right]^{\alpha \lambda \gamma }\le C_{5}\,\mathbb E _\mathbb{P }\!\left[\left(X^{-}\right)^{q}\right] ^{\frac{\alpha \lambda \gamma }{q}}\) (here \(C_{5}\text{:=} \mathbb E _\mathbb{P }\!\left[\rho ^{q/\left(q-1\right)}\right]^{\alpha \lambda \gamma \left(q-1\right)/q}\) is also strictly positive and finite). Moreover, since \(\frac{\alpha \lambda \gamma }{q}<\delta <1\), we have by the trivial inequality mentioned above that \(\mathbb E _\mathbb{P }\!\left[\left(X^{-}\right)^{q}\right]^ {\frac{\alpha \lambda \gamma }{q}}\le 2+\mathbb E _\mathbb{P }\!\left[\left(X^{-}\right)^{q}\right]^ {\delta }\). Therefore, these inequalities combined with (7.1) yield

$$\begin{aligned} \mathbb E _\mathbb{P }\!\left[\left(X^{+}\right)^{\alpha \lambda }\right]^{\gamma }&\le C_{6}+C_{7}\,\mathbb E _\mathbb{P }\!\left[\left(X^{-}\right)^ {q}\right]^{\delta }\nonumber \\&\le C_{6}+C_{7}\left[1+D_{1}\left(\int \limits _{0}^{+\infty } \mathbb P \!\left\{ \left(X^{-}\right)^{\eta }>y\right\} ^{\delta } dy\right)^{\frac{1}{\delta }}\right]^{\delta }\nonumber \\&\le M_{1}+M_{2}\int \limits _{0}^{+\infty } \mathbb P \!\left\{ \left(X^{-}\right)^{\eta }>y\right\} ^{\delta } dy, \end{aligned}$$
(7.2)

where to obtain the second inequality we apply Lemma 3.12 above (note that \(\frac{\eta }{\delta q}>1\)), and the last inequality is again due to the trivial inequality we referred to previously in this proof (with \(0<\delta <1\)). Furthermore, all constants \(C_{6},C_{7},M_{1},M_{2}\) belong to \(\left(0,+\infty \right)\).

Hence,

$$\begin{aligned} \int \limits _{0}^{+\infty } \mathbb P \!\left\{ \left(X^{+}\right)^{\alpha }>y\right\} ^{\gamma } dy \le L_{1}+L_{2}\int \limits _{0}^{+\infty } \mathbb P \!\left\{ \left(X^{-}\right)^{\eta }>y\right\} ^{\delta } dy, \end{aligned}$$

with \(L_{1}\) and \(L_{2}\) positive constants that do not depend on the r.v. X (only on the parameters), as intended.\(\square \)

1.1.4 Proof of Lemma  3.14

We start by fixing some \(\chi \) satisfying \(\frac{1}{s}<\chi <\frac{b}{s a}\). Such a \(\chi \) exists since \(1<\frac{b}{a}\) implies that \(\frac{1}{s}<\frac{b}{s a}\). We also note that, because \(\chi a<\frac{b}{s}\), we can choose \(\xi \) so that \(\chi a<\xi <\frac{b}{s}\).

Let \(X\) be an arbitrary nonnegative random variable. Given that \(\frac{b}{s \xi }>1\), we know from Lemma 3.12 that \(\mathbb E _\mathbb{P }\!\left[X^{\xi }\right]\le 1+D \left(\int _{0}^{+\infty } \mathbb P \!\left\{ X^{b}>y\right\} ^{s} dy\right)^{1/s}\) for some strictly positive finite constant \(D\) (not depending on \(X\), but only on the parameters). Therefore, recalling that \(s\le 1\), it follows from the trivial inequality \(\left|x+y\right|^{s}\le \left|x\right|^{s}+\left|y\right|^{s}\) that

$$\begin{aligned} \mathbb E _\mathbb{P }\!\left[X^{\xi }\right]^{s}\le 1+C_{1}\int \limits _{0}^{+\infty } \mathbb P \!\left\{ X^{b}>y\right\} ^{s} dy, \end{aligned}$$
(7.3)

with \(C_{1} \in \left(0,+\infty \right)\). Now, by Jensen’s inequality for concave functions (note that \(\frac{a \chi }{\xi }<1\)), we obtain

$$\begin{aligned} \mathbb E _\mathbb{P }\!\left[X^{a \chi }\right]=\mathbb E _\mathbb{P }\!\left[\left(X^{\xi }\right) ^{\frac{a \chi }{\xi }}\right]\le \mathbb E _\mathbb{P }\!\left[X^{\xi }\right]^{\frac{a \chi }{\xi }}. \end{aligned}$$
(7.4)

Moreover, using Chebyshev’s inequality, we get

$$\begin{aligned} \int \limits _{0}^{+\infty } \mathbb P \!\left\{ X^{a}>y\right\} ^{s} dy\le 1+C_{2}\,\mathbb E _\mathbb{P }\!\left[X^{a \chi }\right]^{s}, \end{aligned}$$
(7.5)

where the strictly positive constant \(C_{2}\) is given by \(C_{2}\text{:=} \int _{1}^{+\infty } \frac{1}{y^{s \chi }} dy\) (note that \(s \chi >1\)).

Thus, combining the inequalities (7.3), (7.4) and (7.5) above yields

$$\begin{aligned} \int \limits _{0}^{+\infty } \mathbb P \!\left\{ X^{a}>y\right\} ^{s} dy&\le 1+C_{2}\left(\mathbb E _\mathbb{P }\!\left[X^{\xi }\right]^{s}\right) ^{\frac{a \chi }{\xi }}\\&\le 1+C_{2}\left(1+C_{1}\int \limits _{0}^{+\infty } \mathbb P \!\left\{ X^{b}>y\right\} ^{s} dy\right)^{\!\!\frac{a \chi }{\xi }}\\&\le R_{1}+R_{2}\left(\int \limits _{0}^{+\infty } \mathbb P \!\left\{ X^{b}>y\right\} ^{s} dy\right)^{\!\!\frac{a \chi }{\xi }} \end{aligned}$$

where the last inequality is due to the trivial inequality mentioned above (again we recall that \(\frac{a \chi }{\xi }<1\)), and the positive constants \(R_{1}\), \(R_{2}\) depend only on the parameters. Setting \(\zeta =\frac{a \chi }{\xi }\) completes the proof.\(\square \)

1.2 Proofs of Section 4

1.2.1 Proof of Lemma 4.2

Let \(\lambda >0\) be as in the proof of Lemma 3.13. We begin by showing that

$$\begin{aligned} \sup _{n \in \mathbb N } \mathbb E _\mathbb{P }\!\left[\left(X_{n}^{+}\right)^{\alpha \lambda }\right]<+\infty . \end{aligned}$$
(7.6)

Assume by contradiction that \(\sup _{n \in \mathbb N } \mathbb E _\mathbb{P }\!\left[\left(X_{n}^{+}\right)^{\alpha \lambda }\right]=+\infty \). Then we can take a subsequence of \(\left\{ \mathbb E _\mathbb{P }\!\left[\left(X_{n}^{+}\right)^{\alpha \lambda }\right]\right\} _{n \in \mathbb N }\) such that \(\mathbb E _\mathbb{P }\!\left[\left(X_{n_{l}}^{+}\right)^{\alpha \lambda }\right] \rightarrow +\infty \) as \(l \rightarrow +\infty \). By Eq. (7.2) in the proof of Lemma 3.13, we conclude that \(\int _{0}^{+\infty } \mathbb P \!\left\{ \left(X_{n_{l}}^{-}\right)^{\eta }>y\right\} ^{\delta } dy \xrightarrow [l\rightarrow +\infty ]{} +\infty \), where \(\eta \) is as defined in the proof. Therefore, using Lemma 3.14 we also obtain that \(V_{-}\!\left(X_{n_{l}}^{-}\right)=\int _{0}^{+\infty } \mathbb P \!\left\{ \left(X_{n_{l}}^{-}\right)^{\beta }>y\right\} ^{\delta } dy \xrightarrow [l\rightarrow +\infty ]{} +\infty \), and hence

$$\begin{aligned} V\left(X_{n_{l}}\right)&= V_{+}\!\left(X_{n_{l}}^{+}\right)-V_{-}\!\left(X_{n_{l}}^{-}\right) \le L_{1}+L_{2}\int \limits _{0}^{+\infty } \mathbb P \!\left\{ \left(X_{n_{l}}^{-}\right)^{\eta }>y\right\} ^{\delta } dy-V_{-}\!\left(X_{n_{l}}^{-}\right)\\&\le C_{1}+C_{2} \left(V_{-}\!\left(X_{n_{l}}^{-}\right)\right)^{\zeta }-V_{-}\!\left(X_{n_{l}}^{-}\right) \xrightarrow [l\rightarrow +\infty ]{} -\infty , \end{aligned}$$

where the first and second inequalities follow, respectively, from Lemma 3.13 and Lemma 3.14 (\(C_{1}\) and \(C_{2}\) are strictly positive constants depending only on the parameters.), and \(0<\zeta <1\). But this is absurd, because \(\lim _{n \in \mathbb N } V\!\left(X_{n}\right)=V^{*}>-\infty \) and therefore any subsequence of \(V\!\left(X_{n}\right)\) must also converge to \(V^{*}\).

Now we show that (7.6) implies that \(\sup _{n \in \mathbb N } V_{-}\!\left(X_{n}^{-}\right)<+\infty \). Indeed, by Chebyshev’s inequality, for some \(C_{3}\in \left(0,+\infty \right)\), \(V_{+}\!\left(X_{n}^{+}\right)\le 1+C_{3}\,\mathbb E _\mathbb{P }\!\left[\left(X_{n}^{+}\right)^{\alpha \lambda }\right]^{\gamma }\) for all \(n\). Using (7.6), we thus get \(\sup _{n\in \mathbb N } V_{+}\!\left(X_{n}^{+}\right)<+\infty \). Furthermore, since \(V\!\left(X_{n}\right) \rightarrow V^{*}\) as \(n \rightarrow +\infty \) and any convergent sequence is bounded, we have that \(\inf _{n \in \mathbb N } V\!\left(X_{n}\right)>-\infty \). Thus,

$$\begin{aligned} \sup _{n \in \mathbb N } V_{-}\!\left(X_{n}^{-}\right)\le \sup _{n \in \mathbb N } V_{+}\!\left(X_{n}^{+}\right)-\inf _{n \in \mathbb N } V\!\left(X_{n}\right)<+\infty , \end{aligned}$$

as intended.

Finally, recalling that \(\frac{\beta }{\delta }>1\), we can choose \(\xi \in \left(1,\frac{\beta }{\delta }\right)\). Therefore \(\frac{\beta }{\delta \xi }>1\), and it follows from Lemma 3.12 that there exists some \(D\in \left(0,+\infty \right)\) such that

$$\begin{aligned} \mathbb E _\mathbb{P }\!\left[\left(X_{n}^{-}\right)^{\xi }\right]\le 1+D\left(V_{-}\!\left(X_{n}^{-}\right)\right)^{\frac{1}{\delta }}, \end{aligned}$$

for all \(n \in \mathbb N \), which implies that \(\sup _{n \in \mathbb N } \mathbb E _\mathbb{P }\!\left[\left(X_{n}^{-}\right)^{\xi }\right] <+\infty \). In particular, we note that, because \(g:\left[\left.0,+\infty \right)\right. \rightarrow \left[\left.0,+\infty \right)\right.\) given by \(g\!\left(t\right)\text{:=} t^{\xi }\) is a nonnegative, strictly increasing and strictly convex function on \(\left[\left.0,+\infty \right)\right.\) satisfying \(\lim _{t \rightarrow +\infty } \frac{g\!\left(t\right)}{t}=+\infty \), as well as \(\sup _{n \in \mathbb N } \mathbb E _\mathbb{P }\!\left[g\!\left(X_{n}^{-}\right)\right]<+\infty \), by de la Vallée–Poussin theorem we conclude that the family \(\left\{ X_{n}^{-}\right\} _{n \in \mathbb N }\) is uniformly integrable.

So we set \(\tau =\alpha \lambda \in \left(0,1\right)\). A straightforward application of Hölder’s inequality (with \(p=\frac{\xi }{\tau }>1\)) and of the trivial inequality \(\left|x+y\right|^{\tau }\le \left|x\right|^{\tau }+\left|y\right|^{\tau }\) gives

$$\begin{aligned} \mathbb E _\mathbb{P }\!\left[\left|X_{n}\right|^{\tau }\right]\le \mathbb E _\mathbb{P }\!\left[\left(X_{n}^{+}\right)^{\tau }\right] +\mathbb E _\mathbb{P }\!\left[\left(X_{n}^{-}\right)^{\tau }\right] \le \mathbb E _\mathbb{P }\!\left[\left(X_{n}^{+}\right)^{\tau }\right] +\mathbb E _\mathbb{P }^{\frac{\tau }{\xi }}\!\left[\left(X_{n}^{-} \right)^{\xi }\right], \end{aligned}$$

hence \(\sup _{n \in \mathbb N } \mathbb E _\mathbb{P }\!\left[\left|X_{n}\right|^{\tau }\right] <+\infty \).\(\square \)

1.2.2 Proof of Corollary 4.3

Let \(\epsilon >0\) be arbitrary. By Lemma 4.2 above, \(\sup _{n \in \mathbb N } \mathbb{E }_\mathbb{P }\!\left[\left|X_{n}\right|^{\tau }\right]=S \in \left[\left.0,+\infty \right)\right.\), for some \(\tau \in \left(0,1\right)\). So choosing \(M=M\!\left(\epsilon \right)\) such that \(M>\left(\frac{S}{\epsilon }\right)^{\frac{1}{\tau }}\ge 0\), and setting \(K=\left[-M,M\right]\), by Chebyshev’s inequality we obtain

$$\begin{aligned} \mathbb{P }\left\{ {{X}_{n}}\in K^{c}\right\} =\mathbb{P }\left\{ \left|{X}_{n}\right|{>}M\right\} \le \frac{\mathbb{E }_\mathbb{P }\left[\left|{X}_{n}\right|^{\tau }\right]}{M^{\tau }}<\epsilon , \end{aligned}$$

which completes the proof.\(\square \)

Rights and permissions

Reprints and permissions

About this article

Cite this article

Rásonyi, M., Rodrigues, A.M. Optimal portfolio choice for a behavioural investor in continuous-time markets. Ann Finance 9, 291–318 (2013). https://doi.org/10.1007/s10436-012-0211-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10436-012-0211-4

Keywords

JEL Classification

Navigation