Skip to main content
Log in

Robo-advising: a dynamic mean-variance approach

  • Original Article
  • Published:
Digital Finance Aims and scope Submit manuscript

Abstract

In contrast to traditional financial advising, robo-advising needs to elicit investors’ risk profile via several simple online questions and provide advice consistent with conventional investment wisdom, e.g., rich and young people should invest more in risky assets. To meet the two challenges, we propose to do the asset allocation part of robo-advising using a dynamic mean-variance criterion over the portfolio’s log returns. We obtain analytical and time-consistent optimal portfolio policies under jump-diffusion models and regime-switching models.

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. Of course, for the last two rules to hold, one needs some reasonable assumptions.

  2. Here we assume the square integrability of \(\hat{\pi }\) under some mild conditions on \(\mu (\cdot , \cdot ;\, i)\) and \(\sigma (\cdot ,\cdot ; \, i)\).

  3. We assume that the investor using our system knows the meaning of the average and the standard deviation and understands the basic trade-off between risk and return; otherwise, it is perhaps better for the investor to seek advice from a professional financial advisor directly or to buy other financial products such as fixed annuity contracts, rather than using the robo-advising.

References

  • Ang, A., & Bekaert, G. (2002). International asset allocation with regime shifts. Review of Financial Studies, 15, 1137–1187.

    Article  Google Scholar 

  • Basak, S., & Chabakauri, G. (2010). Dynamic mean-variance asset allocation. Review of Financial Studies, 23, 2970–3016.

    Article  Google Scholar 

  • Björk, T., Khapko, M., & Murgoc, A. (2017). On time-inconsistent stochastic control in continuous time. Finance and Stochastics, 21(2), 331–360.

    Article  Google Scholar 

  • Björk, T., Murgoci, A., & Zhou, X. Y. (2014). Mean-variance portfolio optimization with state-dependent risk aversion. Mathematical Finance, 24, 1–24.

    Article  Google Scholar 

  • Capponi, A., Olafsson, S., & Zariphopoulou, T. (2019). Personalized robo-advising: Enhancing investment through client interaction. arXiv:1911.01391.

  • Cai, J., Chen, X., & Dai, M. (2017). Portfolio selection with capital gains tax, recursive utility, and regime switching. Management Science,64(5), 2308–2324.

    Article  Google Scholar 

  • Chacko, G., & Viceira, L. M. (2005). Dynamic consumption and portfolio choice with stochastic volatility in incomplete markets. Review of Financial Studies,18, 1369–402.

    Article  Google Scholar 

  • Dai, M., Jin, H. Q., Kou, S., & Xu, Y. H. (2020). A dynamic mean-variance analysis for log returns. Management Science. https://doi.org/10.1287/mnsc.2019.3493.

  • Campbell, J. Y., & Viceira, L. M. (1999). Consumption and portfolio decisions when expected returns are time varying. Quarterly Journal of Economics,114, 433–495.

    Article  Google Scholar 

  • Kim, T. S., & Omberg, E. (1996). Dynamic nonmyopic portfolio behavior. Review of Financial Studies,9, 141–161.

    Article  Google Scholar 

  • Liu, J., Longstaff, F. A., & Pan, J. (2003). Dynamic asset allocation with event risk. The Journal of Finance,58(1), 231–259.

    Article  Google Scholar 

  • Liu, J. (2001). Dynamic portfolio choice and risk aversion. Working Paper, UCLA, SSRN-id287095.

  • Markowitz, H. (1952). Portfolio selection. The Journal of Finance, 7, 77–91.

    Google Scholar 

  • Merton, R. C. (1971). Optimum consumption and portfolio rules in a continuous-time model. Journal of Economic Theory,3, 373–413.

    Article  Google Scholar 

  • Strub, M. S., Li, D., & Cui, X. Y. (2019). An enhanced mean-variance framework for robo-advising applications. Available at SSRN: https://ssrn.com/abstract=3302111.

  • Wachter, J. A. (2002). Portfolio and consumption decisions under mean-reverting returns: An explicit solution for complete market. Journal of Financial and Quantitative Analysis, 37, 63–91.

    Article  Google Scholar 

  • Zhou, X. Y., & Yin, G. (2003). Markowitz’s mean-variance portfolio selection with regime switching: A continuous-time model. SIAM Journal on Control and Optimization, 42(4), 1466–1482.

    Article  Google Scholar 

Download references

Funding

Funded by National Natural Science Foundation of China 11671292, 11871050, and 12071333, Singapore Academic Research Fund R-703-000-032-112, R-146-000-306-114, and R-146-000-311-114.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Min Dai.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Proofs for results in section 2

Proofs for results in section 2

Proofs of proposition 1.

The value function is defined as \(V\left( t,R_{t}\right) =\mathrm {E} _{t}(R_{T}^{*})-\frac{\gamma }{2}\mathrm {Var}_{t}(R_{T}^{*})\), where \(R^{*}\) is the log return with equilibrium \(\hat{\pi }\). Let \(g\left( t,R_{t}\right) :=\mathrm {E}_{t}\left[ R_{T}^{*}\right]\). By the structures of \(R_{T}^{*}\) and the conditional expectation and the conditional variance, we consider the forms

$$\begin{aligned} V\left( t,R_{t}\right)&=R_{t}+A\left( t\right) \text {, }A\left( T\right) =0; \end{aligned}$$
(18)
$$\begin{aligned} g\left( t,R_{t}\right)&=R_{t}+a\left( t\right) \text {, }a\left( T\right) =0. \end{aligned}$$
(19)

Furthermore, we have, via the general extended HJB in Björk et al. (2017), that

$$\begin{aligned} \underset{\pi _{t}}{\max }&\{\mathcal {D}V(t,R_{t})-\frac{\gamma }{2} \sigma ^{2}\pi _{t}^{2}\left( \frac{\partial g}{\partial R}\right) ^{2} +\int _{\mathbb {R}}\left[ g\left( t,R_{t^{-}}+\ln \left( 1+\beta \pi _{t}\right) \right) -g\left( t,R_{t^{-}}\right) \right] \lambda \left( dz\right) \\&+\int _{\mathbb {R}}\left[ g^{2}\left( t,R_{t^{-}}+\ln \left( 1+\beta \pi _{t}\right) \right) -g^{2}\left( t,R_{t^{-}}\right) \right] \lambda \left( \mathrm{d}z\right) \}=0, \end{aligned}$$

subject to \(V(T,R)=R\), where \(\mathcal {D}\) is the Dynkin operator. For a function F of \(\left( t,R\right)\), \(\mathcal {D}F\left( t,R\right) =\frac{\partial F}{\partial t}+\left( r+\left( \mu -r\right) \pi -\frac{1}{2}\sigma ^{2}\pi ^{2}\right) \frac{\partial F}{\partial R}+\frac{1}{2}\sigma ^{2}\pi ^{2}\frac{\partial ^{2} F}{\partial R^{2}}+\int _{\mathbb {R}}\left[ F\left( t,R_{t^{-}}+\ln \left( 1+\beta \pi \right) \right) -F\left( t,R_{t^{-}}\right) \right] \lambda \left( \mathrm{d}z\right)\). Incorporating the special forms (18) and (19), we have

$$\begin{aligned} \underset{\pi _{t}}{\max }&\{\frac{\mathrm{d}A}{\mathrm{d}t}+\left( r+\left( \mu -r\right) \pi _{t}-\frac{1}{2}\sigma ^{2}\pi _{t}^{2}\right) -\frac{\gamma }{2}\sigma ^{2}\pi _{t}^{2}+\int _{\mathbb {R}}\ln \left( 1+\beta \pi _{t}\right) \lambda \left( \mathrm{d}z\right) \\&-\frac{\gamma }{2}\int _{\mathbb {R}}\ln ^{2}\left( 1+\beta \pi _{t}\right) \lambda \left( \mathrm{d}z\right) \}=0. \end{aligned}$$

We then obtain the equilibrium strategy (5) by the first-order condition. \(\square\)

Proof of Proposition 2.

For any proportional portfolio \(\pi\), by the dynamics of the log return process \(\{R(t;\pi )\}\), we have, given \(R(t,\pi )=R\), that

$$\begin{aligned} R(T;\pi )= & {} \int _{t}^{T}\left( r(s, X_s;\alpha (s))+\left( \mu (s, X_s;\alpha (s))-r(s, X_s;\alpha (s))\right) \pi _{s}-\frac{1}{2}(\sigma (s, X_s;\alpha (s))\pi _{s})^{2}\right) \mathrm{d}s\\&+\int _{t}^{T}\sigma _{s}^{i}\pi _{s}^{i}dB_{t}+R, \end{aligned}$$

and the objective function is to maximize

$$\begin{aligned} J(t,R,x;\pi , \alpha (t))=\mathbb {E}_{t}\left[ R(T;\pi )\ |X_t=x, \alpha (t)=i \right] -\frac{\gamma }{2}\mathrm {Var}_{t}(R(T;\pi )\ |X_t=x, \alpha (t)=i ). \end{aligned}$$

Let \(\hat{\pi }(t, R, x, i)\) be an equilibrium feedback proportional allocation, \(\{R(s)\}\) be the resulted log return process, \(\{\hat{\pi }_s\}\) be the proportional portfolio process, and V(tRxi) be the value function. Define

$$\begin{aligned} f(t,x,R,i)=\mathrm {E}_{t }\left[ \int _{t}^{T}\left( \left( \mu (s, X_s; \alpha (s))-r(s, X_s; \alpha (s))\right) \hat{\pi }_s-\frac{1}{2}(\sigma (s, X_s; \alpha (s))\hat{\pi }_s)^{2}\right) \mathrm{d}s \ |X_t=x, \alpha (t)=i \right] . \end{aligned}$$

It is easy to see that \(J(t,R,X_t;\pi , \alpha (t))-R\) is not explicitly dependent on R, this gives us the conjecture that V(tRxi) admits the form

$$\begin{aligned} V(t,R,x,i)=\tilde{V}(t,x,i)+R+{\mathbb {E}}_t\left[ \int _t^T r(s, X_s;\alpha (s))\mathrm{d}s\ |X_t=x, \alpha (t)=i \right] . \end{aligned}$$

By this form and the definition of equilibrium solution, we then obtain the PDE

$$\begin{aligned} \underset{\pi }{\max }\left\{ D\tilde{V}(t,x,i)+\left( \mu ^{i} -r^{i}\right) \pi -\frac{1}{2}(\sigma ^{i}\pi )^{2}-\frac{\gamma }{2}\left[ (\nu ^{i}\frac{\partial f^{i}}{\partial x})^{2}+2\rho \nu ^{i} \sigma ^{i}\pi \frac{\partial f^{i}}{\partial x}+(\sigma ^{i}\pi )^{2}+\sum _{j=1}^N q_{ij} f(t,x;j)^2\right] \right\} =0, \end{aligned}$$
(20)

subject to \(\tilde{V}(T,x,i)=0\), where we used the simplified notations \(\mu ^i:=\mu (t, x; i), \sigma ^i:=\sigma (t, x; i), m_i:=m( x; i), \nu _i=\nu (x;i)\), and

$$\begin{aligned} D\tilde{V}(t,x,i)=\frac{\partial \tilde{V}}{\partial t}+m_{i}\frac{\partial \tilde{V}}{\partial x}+\frac{1}{2}\nu _{i}^{2}\frac{\partial ^{2}\tilde{V}}{\partial x^{2}}+\sum _{j=1} ^{N}q_{ij}\tilde{V}(t,x,j). \end{aligned}$$

If \(\tilde{V}\) is a solution to this equation, then the optimal \(\pi ^*=\hat{\pi }(t,x;i)\) will be the equilibrium feedback strategy.

By the first-order condition, an equilibrium optimal trading strategy is as given in (7). Substituting \(\hat{\pi }\) into \(f^{i}\) gives

$$\begin{aligned} f(t,x,i)= & \mathrm {E}_{t}\left[ \int _{t}^{T}\left( \frac{( 2\gamma +1) ( \mu _s-r_s) ^{2}}{2(\gamma +1)^{2}\sigma _s^{2}}-\frac{\rho \gamma ^{2}\left( \mu _s-r_s\right) \nu _s}{\left( 1+\gamma \right) ^{2}\sigma _s}\frac{\partial f(s, X_s; \alpha (s))}{\partial x} \right. \right. \\&\left. \left. -\frac{1}{2}\frac{\left( \gamma \rho \nu _s \right) ^{2}}{\left( 1+\gamma \right) ^{2}}\left( \frac{\partial f(s, X_s; \alpha (s))}{\partial x}\right) ^{2}\right) \mathrm{d}s\right] . \end{aligned}$$

with some abusing of notations \(r_s=r(s,X_s;\alpha (s)), \mu _s=\mu (s, X_s;\alpha (s)), \sigma _s=\sigma (s,X_s;\alpha (s))\) and \(m_s=m(X_s;\alpha (s)), \nu _s=\nu (X_s;\alpha (s))\).

By the Feynman-Kac formula, we infer that \(f^{i}\) satisfies (8). \(\square\)


Proof of Corollary 1: For conciseness, we use \(f^i\) to mean the function \(f(\cdot , \cdot ; i)\). For the Gaussian mean returns model, the PDE (8) becomes

$$\begin{aligned}&\frac{\partial f^{i}}{\partial t} +(\lambda (i)\bar{X}(i)-\lambda (i) x-\frac{\rho \gamma ^{2}\nu (i)}{\left( 1+\gamma \right) ^{2} }x)\frac{\partial f^{i}}{\partial x}+\frac{1}{2}\nu (i)^{2}\left( \frac{\partial ^{2}f^{i}}{\partial x^{2}}-\frac{\rho ^{2}\gamma ^{2}}{\left( 1+\gamma \right) ^{2}}(\frac{\partial f^{i}}{\partial x})^{2}\right) \nonumber \\&\quad +\frac{\left( 2\gamma +1\right) }{2\left( \gamma +1\right) ^{2} }x^{2}+\sum _{j=1}^{N}q_{ij}f(t,x,j)=0,\nonumber \\&\qquad f^{i} \left( T,x\right) =0. \end{aligned}$$
(21)

Consider a solution of the following quadratic form:

$$\begin{aligned} f(t,x,i)=A\left( t,i\right) x^{2}+B\left( t,i\right) x+C\left( t,i\right) . \end{aligned}$$

Substituting this into (21) and by separating variables, we obtain the following ODE system:

$$\begin{aligned}&\left\{ \begin{array} [c]{l} \frac{\mathrm{d}A}{\mathrm{d}t}\left( t,i\right) -2\left( \lambda (i)+\frac{\rho \gamma ^{2}\nu _{i}}{\left( 1+\gamma \right) ^{2}}\right) A\left( t,i\right) +\frac{2\rho ^{2}\gamma ^{2}\nu (i)^{2}}{\left( 1+\gamma \right) ^{2}}A^{2}\left( t,i\right) +\frac{\left( 2\gamma +1\right) }{2\left( \gamma +1\right) ^{2}}+\sum _{j=1}^{N} q_{ij}A\left( t,j\right) =0\text {,}\\ A\left( T,i\right) =0\text {,} \end{array} \right. \\&\left\{ \begin{array} [c]{l} \frac{\mathrm{d}B}{\mathrm{d}t}\left( t,i\right) -\left( \lambda (i)+\frac{\rho \gamma ^{2}\nu (i)}{\left( 1+\gamma \right) ^{2}}+\frac{2\rho ^{2}\gamma ^{2}\nu (i)^{2}}{\left( 1+\gamma \right) ^{2}}A\left( t,i\right) \right) B\left( t,i\right) +2\lambda (i)\bar{X}^{i}A\left( t,i\right) +\sum _{j=1}^{N}q_{ij}B\left( t,j\right) =0\text {,}\\ B\left( T,i\right) =0\text {,} \end{array} \right. \\&\left\{ \begin{array} [c]{l} \frac{\mathrm{d}C}{\mathrm{d}t}\left( t,i\right) +\nu (i)^{2}A\left( t,i\right) +\lambda (i)\bar{X}(i)B\left( t,i\right) -\frac{\rho ^{2}\gamma ^{2}\nu _{i}^{2} }{2\left( 1+\gamma \right) ^{2}}B^{2}\left( t,i\right) +\sum _{j=1} ^{N}q_{ij}C\left( t,j\right) =0\text {,}\\ C\left( T,i\right) =0\text {,} \end{array} \right. \end{aligned}$$

Hence, by Proposition 2, the equilibrium strategy is as given in (11). \(\square\)

Proof of Corollary 2: Note that in the stochastic volatility model, \(\mu (i)-r_{i}=\delta _{i} x^{\frac{1+\beta }{2\beta }},\ \sigma (i)=x^{\frac{1}{2\beta }},\ m_{i} =\lambda (i)\bar{X}(i)-\lambda (i)x\), and \(\nu (i)=\bar{\nu }(i)x^{\frac{1}{2}}\). Then PDE (8) becomes

$$\begin{aligned}&\frac{\partial f^{i}}{\partial t} +(\lambda (i)\bar{X}(i)-\lambda (i)x-\frac{\gamma ^{2}\rho \delta _{i}\bar{\nu }(i)}{\left( 1+\gamma \right) ^{2}}x)\frac{\partial f^{i}}{\partial x}+\frac{1}{2}\bar{\nu }(i)^{2}x\left( \frac{\partial ^{2}f^{i}}{\partial x^{2}}-\frac{\gamma ^{2}\rho ^{2}}{\left( 1+\gamma \right) ^{2}}(\frac{\partial f^{i} }{\partial x})^{2}\right) \nonumber \\&\quad +\frac{\left( 2\gamma +1\right) \delta (i)^{2}}{2\left( \gamma +1\right) ^{2}}x+\sum _{j=1}^{N}q_{ij}f(t,x,j)=0,\nonumber \\&\qquad f^{i} \left( T,x\right) =0. \end{aligned}$$
(22)

We try a solution of the following linear form:

$$\begin{aligned} f(t,x,i)=A\left( t,i\right) \cdot x+B\left( t,i\right) . \end{aligned}$$

Substituting this into (22) and by separating variables, we obtain the following ODE system:

$$\begin{aligned}&\left\{ \begin{array} [c]{l} \frac{\mathrm{d}A}{\mathrm{d}t}\left( t,i\right) -\left( \lambda (i)+\frac{\rho \gamma ^{2}\delta (i)\bar{\nu }(i)}{\left( 1+\gamma \right) ^{2}}\right) A\left( t,i\right) +\frac{\rho ^{2}\gamma ^{2}\bar{\nu }(i)^{2}}{2\left( 1+\gamma \right) ^{2}}A^{2}\left( t,i\right) +\frac{\left( 2\gamma +1\right) \delta (i)^{2}}{2\left( \gamma +1\right) ^{2}}+\sum _{j=1}^{N}q_{ij}A\left( t,j\right) =0\text {,}\\ A\left( T,i\right) =0\text {,} \end{array} \right. \\&\left\{ \begin{array} [c]{l} \frac{\mathrm{d}B}{\mathrm{d}t}\left( t,i\right) +\lambda (i)\bar{X}(i)A\left( t,i\right) +\sum _{j=1}^{N}q_{ij}B\left( t,j\right) =0\text {,}\\ B\left( T,i\right) =0\text {,} \end{array} \right. \end{aligned}$$

Hence, by Proposition 2, the equilibrium strategy is as given in (16). \(\square\)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Dai, M., Jin, H., Kou, S. et al. Robo-advising: a dynamic mean-variance approach. Digit Finance 3, 81–97 (2021). https://doi.org/10.1007/s42521-021-00028-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s42521-021-00028-4

Keywords

JEL Classification

Navigation