Skip to main content
Log in

Increasing risk aversion and life-cycle investing

  • Published:
Mathematics and Financial Economics Aims and scope Submit manuscript

Abstract

We derive the optimal portfolio for an investor with increasing relative risk aversion in a complete continuous-time securities market. The IRRA assumption helps to mitigate the criticism of constant relative risk aversion that it implies an unreasonably large aversion to large gambles, given reasonable aversion to small gambles. The model provides theoretical support for the common recommendation of financial advisors that older investors should reduce their allocations to risky assets, and it is consistent with empirical relations between age, wealth, and portfolios.

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.

Fig. 1
Fig. 2
Fig. 3
Fig. 4

Similar content being viewed by others

Notes

  1. IRRA preferences do not eliminate the concern entirely. Rabin [15] argues that it is a feature of expected utility in general that either aversion to small gambles is too small, or aversion to large gambles is too large.

  2. The empirical relation of equity allocations to age is not entirely clear. Bodie and Crane [2] find that the allocation is decreasing in age. Also, Curcuru et al. [7] find that the allocation is decreasing in age for low-wealth individuals and hump-shaped for others. However, Ameriks and Zeldes [1] find that the allocation is increasing in age for individual investors, with the hump-shaped cross-sectional pattern being due to cohort effects.

  3. There are different results in the literature regarding whether these utility functions are actually representative of individuals’ preferences. [3] argue that they are the most representative within the HARA class, based on experimental evidence. However, [4] argue that decreasing relative risk aversion is a better assumption, based on Swedish household portfolio data.

  4. Huggett and Kaplan [8] document that there is substantial risk in labor income, though they estimate the value of the part that is spanned by asset markets to be less than 35% of the total value of human capital.

References

  1. Ameriks, J., Zeldes, S.P.: How do Household Portfolio Shares Vary with Age. Technical report, working paper, Columbia University (2004)

  2. Bodie, Z., Crane, D.B.: Personal investing: advice, theory, and evidence. Financ. Anal. J. 53, 13–23 (1997)

    Article  Google Scholar 

  3. Brocas, I., Carrillo, J.D., Giga, A., Zapatero, F.: Risk Aversion in a Dynamic Asset Allocation Experiment. Working Paper (2014)

  4. Calvet, L.E., Sodini, P.: Twin picks: disentangling the determinants of risk-taking in household portfolios. J. Finance 69, 867–906 (2014)

    Article  Google Scholar 

  5. Canner, N., Mankiw, N.G., Weil, D.N.: An asset allocation puzzle. Am. Econ. Rev. 87, 181–191 (1997)

    Google Scholar 

  6. Cox, J.C., Huang, C.F.: Optimal consumption and portfolio policies when asset prices follow a diffusion process. J. Econ. Theory 49, 33–83 (1989)

    Article  MathSciNet  MATH  Google Scholar 

  7. Curcuru, S., Heaton, J., Lucas, D., Moore, D.: Heterogeneity and portfolio choice: theory and evidence. In: Aït-Sahalia, Y., Hansen, L.P. (eds.) The Handbook of Financial Econometrics, vol. 1, pp. 337–382. North Holland, Amsterdam, Netherlands (2010)

  8. Huggett, M., Kaplan, G.: How large is the stock component of human capital. Rev. Econ. 22, 21–51 (2016)

    Google Scholar 

  9. Jagannathan, R., Kocherlakota, N.R.: Why should older people invest less stock than younger people? Fed. Reserve Bank Minneap. Q. Rev. 20, 11–23 (1996)

    Google Scholar 

  10. Kandel, S., Stambaugh, R.F.: Asset returns and intertemporal preferences. J. Monet. Econ. 27, 39–71 (1991)

    Article  Google Scholar 

  11. Karatzas, I., Lehoczky, J.P., Shreve, S.E.: Optimal portfolio and consumption decisions for a “small investor” on a finite horizon. SIAM J. Control Optim. 25, 1557–1586 (1987)

    Article  MathSciNet  MATH  Google Scholar 

  12. Lakner, P., Nygren, L.M.: Portfolio optimization with downside constraints. Math. Finance 16, 283–299 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  13. Merton, R.C.: Optimum consumption and portfolio rules in a continuous-time model. J. Econ. Theory 3, 373–413 (1971)

    Article  MathSciNet  MATH  Google Scholar 

  14. Nielsen, L.T., Vassalou, M.: The instantaneous capital market line. Econ. Theory 28, 651–664 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  15. Rabin, M.: Risk aversion and expected-utility theory: a calibration theorem. Econometrica 68, 1281–1292 (2000)

    Article  Google Scholar 

  16. Samuelson, P.A.: Lifetime portfolio selection by dynamic stochastic programming. Rev. Econ. Stat. 51, 239–246 (1969)

    Article  Google Scholar 

  17. Sethi, S.P., Taksar, M.I.: A note on Merton’s “optimum consumption and portfolio rules in a continuous-time model”. J. Econ. Theory 46, 395–401 (1988)

    Article  MATH  Google Scholar 

  18. Sethi, S.P., Taksar, M.I., Presman, E.L.: Explicit solution of a general consumption/portfolio problem with subsistence consumption and bankruptcy. J. Econ. Dyn. Control 16, 747–768 (1992)

    Article  MathSciNet  MATH  Google Scholar 

  19. Shin, Y.H., Lim, B.H., Choi, U.J.: Optimal consumption and portfolio selection problem with downside consumption constraints. Appl. Math. Comput. 188, 1801–1811 (2007)

    MathSciNet  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Ruomeng Liu.

Appendices

Appendix A. Proof of Proposition 1

The function

$$\begin{aligned} \gamma \mapsto \mathsf {E}\int _0^T M_t (\gamma X_t - \xi )^+\,\mathrm {d}t \end{aligned}$$

is strictly monotone and maps \([0,\infty )\) onto \([0,\infty )\). Thus, there is a unique \(\gamma \) such that

$$\begin{aligned} \mathsf {E}\int _0^T M_t (\gamma X_t - \xi )^+\,\mathrm {d}t = W_0. \end{aligned}$$
(A.1)

Let \(C^*\) denote \((\gamma X_t - \xi )^+\), and let C be any other nonnegative consumption process satisfying the budget constraint

$$\begin{aligned} \mathsf {E}\int _0^T M_t C_t \mathrm {d}t \le W_0. \end{aligned}$$
(A.2)

By concavity,

$$\begin{aligned} u(C) \le u(C^*) + u'(C^*)(C-C^*), \end{aligned}$$

so

$$\begin{aligned} \mathsf {E}\int _0^T \mathrm {e}^{-\delta t}u(C_t) \,\mathrm {d}t \le \mathsf {E}\int _0^T \mathrm {e}^{-\delta t}u(C^*_t) \,\mathrm {d}t + \mathsf {E}\int _0^T \mathrm {e}^{-\delta t}u'(C_t^*)(C_t-C_t^*)\,\mathrm {d}t. \end{aligned}$$
(A.3)

The second term on the right-hand side is the sum of

$$\begin{aligned} \mathsf {E}\int _0^T \mathrm {e}^{-\delta t}u'(C_t^*)(C_t-C_t^*)1_{\{C^*_t>0\}}\,\mathrm {d}t \end{aligned}$$
(A.4)

and

$$\begin{aligned} \mathsf {E}\int _0^T \mathrm {e}^{-\delta t}u'(C_t^*)(C_t-C_t^*)1_{\{C^*_t=0\}}\,\mathrm {d}t. \end{aligned}$$
(A.5)

The term (A.4) equals

$$\begin{aligned} \mathsf {E}\int _0^T \mathrm {e}^{-\delta t}(\gamma X_t)^{-\rho } (C_t-C^*_t)1_{\{C^*_t>0\}} \,\mathrm {d}t = \gamma ^{-\rho } \mathsf {E}\int _0^T M_t (C_t-C^*_t) 1_{\{C^*_t>0\}}\,\mathrm {d}t. \end{aligned}$$
(A.6)

The term (A.5) equals

$$\begin{aligned} \mathsf {E}\int _0^T \mathrm {e}^{-\delta t} \xi ^{-\rho } (C_t-C^*_t)1_{\{C^*_t=0\}} \,\mathrm {d}t&\le \mathsf {E}\int _0^T \mathrm {e}^{-\delta t}(\gamma X_t)^{-\rho } (C_t-C^*_t)1_{\{C^*_t=0\}} \,\mathrm {d}t \nonumber \\&= \gamma ^{-\rho } \mathsf {E}\int _0^T M_t (C_t-C^*_t) 1_{\{C^*_t=0\}}\,\mathrm {d}t, \end{aligned}$$
(A.7)

the inequality being due to the fact that \(\xi \ge \gamma X_t\) when \(C^*_t=0\). Adding (A.6) and (A.7) and using (A.1) and (A.2) yields

$$\begin{aligned} \mathsf {E}\int _0^T \mathrm {e}^{-\delta t}u'(C_t^*)(C_t-C_t^*)\,\mathrm {d}t \le 0. \end{aligned}$$

Hence, (A.3) implies that \(C^*\) is optimal.

Because \(C^*\) is the optimal consumption process, the expression (12)–which equals \(f(t,X_t)\)—is the optimal wealth process. We have

$$\begin{aligned} f(t,x)= & {} \int _t^T \mathrm {e}^{-\delta (u-t) }\mathsf {E}\left[ \left( \frac{X_u}{x}\right) ^{-\rho } (\gamma X_u-\xi )^+\mid X_t=x\right] \,\mathrm {d}u\nonumber \\= & {} \gamma x \int _t^T \mathrm {e}^{-\delta (u-t)}\mathsf {E}\left[ \left( \frac{X_u}{x}\right) ^{1-\rho } 1_{\{X_u>\xi /\gamma \}}\mid X_t = x\right] \,\mathrm {d}u\nonumber \\&- \xi \int _t^T \mathrm {e}^{-\delta (u-t)}\mathsf {E}\left[ \left( \frac{X_u}{x}\right) ^{-\rho } 1_{\{X_u>\xi /\gamma \}}\mid X_t = x\right] \,\mathrm {d}u \end{aligned}$$
(A.8)

Consider any constant \(\alpha \) and dates \(t<u\). Set \(\tau = u-t\). Using the definition of X and the fact that M is the geometric Brownian motion (5), we have

$$\begin{aligned} \log X_u - \log X_t = \frac{(r-\delta +\lambda ^2/2)\tau }{\rho } + \frac{\lambda }{\rho }(B_u-B_t). \end{aligned}$$
(A.9)

Therefore,

$$\begin{aligned} \left( \frac{X_u}{X_t}\right) ^\alpha = \mathrm {e}^{\alpha (r-\delta + \lambda ^2/2) \tau /\rho } \mathrm {e}^{- \alpha \lambda \sqrt{\tau } \varepsilon / \rho }, \end{aligned}$$

where \(\varepsilon =-(B_u-B_t)/\sqrt{u-t}\), which is a standard normal random variable. Furthermore,

$$\begin{aligned} X_u \ \ge \ \frac{\xi }{\gamma }&\quad \Leftrightarrow \quad \log X_t + \frac{(r-\delta +\lambda ^2/2)\tau }{\rho } - \frac{\lambda \sqrt{\tau } \varepsilon }{ \rho } \ \ge \ \log \left( \frac{\xi }{\gamma }\right) \\.&\quad \Leftrightarrow \quad \frac{\lambda \sqrt{\tau } \varepsilon }{ \rho } \ \le \ \log (\gamma X_t) - \log \xi + \frac{(r-\delta +\lambda ^2/2)\tau }{\rho } \\&\quad \Leftrightarrow \quad \varepsilon \ \le \ \omega , \end{aligned}$$

where we set

$$\begin{aligned} \omega = \frac{\rho [\log (\gamma X_t) - \log \xi ]}{\lambda \sqrt{\tau }} + \frac{(r-\delta +\lambda ^2/2)\sqrt{\tau }}{\lambda } . \end{aligned}$$

Therefore,

$$\begin{aligned}&\mathsf {E}\left[ \left( \frac{X_u}{x}\right) ^\alpha 1_{\{X_u>\xi /\gamma \}}\mid X_t = x\right] = \mathrm {e}^{\alpha (r-\delta + \lambda ^2/2) \tau /\rho } \frac{1}{\sqrt{2\pi }}\int _{-\infty }^\omega \mathrm {e}^{-y^2/2- \alpha \lambda \sqrt{\tau } y/ \rho } \,\mathrm {d}y\nonumber \\&\qquad = \mathrm {e}^{\alpha (r-\delta + \lambda ^2/2) \tau /\rho +\alpha ^2\lambda ^2\tau /2\rho ^2} \frac{1}{\sqrt{2\pi }}\int _{-\infty }^\omega \mathrm {e}^{-(y+\alpha \lambda \sqrt{\tau }/\rho )^2/2} \,\mathrm {d}y\nonumber \\&\qquad = \mathrm {e}^{\alpha (r-\delta + \lambda ^2/2) \tau /\rho +\alpha ^2\lambda ^2\tau /2\rho ^2} \frac{1}{\sqrt{2\pi }}\int _{-\infty }^{\omega + \alpha \lambda \sqrt{\tau }/\rho } \mathrm {e}^{-y^2/2} \,\mathrm {d}y\nonumber \\&\qquad = \mathrm {e}^{\alpha (r-\delta + \lambda ^2/2) \tau /\rho +\alpha ^2\lambda ^2\tau /2\rho ^2} {{\mathrm{N}}}\left( \omega + \frac{\alpha \lambda \sqrt{\tau }}{\rho }\right) . \end{aligned}$$
(A.10)

Substituting (A.10) with \(\alpha = 1-\rho \) and \(\alpha =-\rho \) into (A.8) verifies (16).

From the formula (16), it is straightforward to verify that f is continuously differentiable in t and twice continuously differentiable in x. Therefore, we can apply Itô’s formula to compute \(\mathrm {d}f\). We compute the optimal portfolio \(\pi \) by matching the stochastic part of \(\mathrm {d}f\) to \(W\pi \sigma \,\mathrm {d}B\), which is the stochastic part of \(\mathrm {d}W\) implied by the intertemporal budget constraint. The drift parts will then match due to the fact that

$$\begin{aligned} \int _0^t M_sC_s\,\mathrm {d}s + M_tf(t,X_t) \end{aligned}$$

is a martingale (which implies that its drift is zero). From (A.9), X is a geometric Brownian motion with volatility \(\lambda /\rho \). Therefore, the stochastic part of \(\mathrm {d}f\) is \((\lambda /\rho )Xf_x \,\mathrm {d}B\). Matching this to \(W\pi \sigma \,\mathrm {d}B\), we see that

$$\begin{aligned} \pi _t = \frac{\lambda }{\rho \sigma } \cdot \frac{X_tf_x(t,X_t)}{W_t} = \frac{\mu -r}{\rho \sigma ^2} \cdot \frac{X_tf_x(t,X_t)}{f(t,X_t)}. \end{aligned}$$

This verifies (17). The formula (16) for f and (17) directly imply (18).

If \(\xi =0\), then the definition (14) simplifies to

$$\begin{aligned} f(t,x) = \gamma x \mathsf {E}\left[ \left. \int _t^T \mathrm {e}^{-\delta (u-t)}\left( \frac{X_u}{x}\right) ^{1-\rho } \,\mathrm {d}u \;\right| \; X_t = x\right] , \end{aligned}$$

which equals xa(t), where \(a(\cdot )\) is the nonrandom function

$$\begin{aligned} a(t) = \mathsf {E}\left[ \left. \int _t^T \mathrm {e}^{-\delta (u-t)}\left( \frac{X_u}{x}\right) ^{1-\rho } \,\mathrm {d}u \;\right| \; X_t = x\right] . \end{aligned}$$

Therefore, \(f(t,x) = xf_x(t,x)\) as claimed.

Appendix B. Negative consumption

Here we derive the optimum for an IRRA investor when negative consumption is allowed. The first order condition is

$$\begin{aligned} \mathrm {e}^{-\delta t}(\xi +C_t)^{-\rho } = \eta M_t \end{aligned}$$

As before, define \(\gamma = \eta ^{-1/\rho }\) and define X as in (10). Then, the first order condition can be expressed as: \(C_t = \gamma X_t-\xi \). Optimal wealth is

$$\begin{aligned} W_t&= \mathsf {E}_t\int _t^T\frac{M_u}{M_t}C_u\,\mathrm {d}u \\&= \gamma X_t\mathsf {E}_t\int _t^T\frac{M_u}{M_t}\frac{X_u}{X_t}\,\mathrm {d}u - \xi \mathsf {E}_t\int _t^T\frac{M_u}{M_t}\,\mathrm {d}u\\&= \gamma X_t\mathsf {E}_t\int _t^T\mathrm {e}^{-\delta (u-t)/\rho }\left( \frac{M_u}{M_t}\right) ^{(\rho -1)/\rho }\,\mathrm {d}u - \frac{\xi }{r}\left[ 1 - \mathrm {e}^{-r(T-t)}\right] \end{aligned}$$

Itô’s formula and the dynamics of M imply

$$\begin{aligned} \frac{\mathrm {d}M^{(\rho -1)/\rho }}{M^{(\rho -1)/\rho }}&= \frac{\rho -1}{\rho }\frac{\mathrm {d}M}{M} + \frac{1-\rho }{2\rho ^2}\left( \frac{\mathrm {d}M}{M}\right) ^2\\&= \frac{1-\rho }{\rho } r\,\mathrm {d}t + \frac{1-\rho }{\rho }\lambda \,\mathrm {d}B + \frac{1-\rho }{2\rho ^2}\lambda ^2\,\mathrm {d}t \end{aligned}$$

Therefore,

$$\begin{aligned} \mathsf {E}_t\left[ \left( \frac{M_u}{M_t}\right) ^{(\rho -1)/\rho }\right] = \mathrm {e}^{[(1-\rho )r/\rho + (1-\rho )\lambda ^2/2\rho ^2](u-t)}. \end{aligned}$$

It follows that

$$\begin{aligned} W_t = \frac{\gamma X_t}{\phi }\left[ 1-\mathrm {e}^{-\phi (T-t)}\right] - \frac{\xi }{r}\left[ 1 - \mathrm {e}^{-r(T-t)}\right] \end{aligned}$$
(B.1)

where

$$\begin{aligned} \phi \ = \ r \ - \ \frac{r - \delta }{\rho } \ - \ \frac{(1 - \rho )\lambda ^2}{2\rho ^2}. \end{aligned}$$

Define

$$\begin{aligned} \widehat{W}_t = W_t + \frac{\xi }{r}\left[ 1 - \mathrm {e}^{-r(T-t)}\right] . \end{aligned}$$

This is the investor’s wealth plus the proceeds that would be obtained by selling a claim that pays \(\xi \) per unit of time from t to T. Equation (B.1) allows us to calculate \(\gamma \) from \(\widehat{W}_0\) as

$$\begin{aligned} \gamma = \frac{\phi \widehat{W}_0}{1-\mathrm {e}^{-\phi T}} \end{aligned}$$
(B.2)

The optimal consumption satisfies

$$\begin{aligned} C_t + \xi = \gamma X_t = \frac{\phi \widehat{W}_t}{1-\mathrm {e}^{-\phi (T-t)}} \end{aligned}$$

The optimal portfolio \(\pi \) is such that \(W\pi \sigma \,\mathrm {d}B\) equals the stochastic part of \(\mathrm {d}W\), which is the stochastic part of \(\mathrm {d}\widehat{W}\), which is the stochastic part of

$$\begin{aligned} \frac{\gamma }{\phi }\left[ 1-\mathrm {e}^{-\phi (T-t)}\right] \,\mathrm {d}X \end{aligned}$$

Thus,

$$\begin{aligned} W\pi \sigma&=\frac{\gamma \lambda }{\phi \rho }\left[ 1-\mathrm {e}^{-\phi (T-t)}\right] X_t\\&= \frac{\lambda }{\rho }\widehat{W}_t \end{aligned}$$

This implies

$$\begin{aligned} \pi = \frac{\widehat{W}_t}{W_t}\cdot \frac{\mu -r}{\rho \sigma ^2} \end{aligned}$$

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Back, K., Liu, R. & Teguia, A. Increasing risk aversion and life-cycle investing. Math Finan Econ 13, 287–302 (2019). https://doi.org/10.1007/s11579-018-0228-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11579-018-0228-1

Keywords

JEL Classification

Navigation