Skip to main content
Log in

An expansion in the model space in the context of utility maximization

  • Published:
Finance and Stochastics Aims and scope Submit manuscript

Abstract

In the framework of an incomplete financial market where the stock price dynamics are modeled by a continuous semimartingale (not necessarily Markovian), an explicit second-order expansion formula for the power investor’s value function—seen as a function of the underlying market price of risk process—is provided. This allows us to provide first-order approximations of the optimal primal and dual controls. Two specific calibrated numerical examples illustrating the accuracy of the method are also given.

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. Theorem 4.5 in [13] expresses the corresponding value function as an infinite sum of weighted generalized Laguerre polynomials.

References

  1. Altarovici, A., Muhle-Karbe, J., Soner, H.M.: Asymptotics for fixed transaction costs. Finance Stoch. 19, 363–414 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  2. Bick, B., Kraft, H., Munk, C.: Solving constrained consumption-investment problems by simulation of artificial market strategies. Manag. Sci. 59, 485–503 (2013)

    Article  Google Scholar 

  3. Brandt, M.W., Goyal, A., Santa-Clara, P., Stroud, J.R.: A simulation approach to dynamic portfolio choice with an application to learning about return predictability. Rev. Financ. Stud. 18, 831–873 (2005)

    Article  Google Scholar 

  4. Campbell, J.Y.: Intertemporal asset pricing without consumption data. Am. Econ. Rev. LXXXIII, 487–512 (1993)

    Google Scholar 

  5. Campbell, J.Y., Viceira, L.M.: Consumption and portfolio decisions when expected returns are time varying. Q. J. Econ. 114, 433–495 (1999)

    Article  MATH  Google Scholar 

  6. Cannarsa, P., Sinestrari, C.: Semiconcave Functions, Hamilton–Jacobi Equations, and Optimal Control. Progress in Nonlinear Differential Equations and their Applications, vol. 58. Birkhäuser Boston, Boston (2004)

    MATH  Google Scholar 

  7. Chacko, G., Viceira, L.M.: Dynamic consumption and portfolio choice with stochastic volatility in incomplete markets. Rev. Financ. Stud. 18, 1369–1402 (2005)

    Article  Google Scholar 

  8. Cheridito, P., Filipović, D., Kimmel, R.L.: Market price of risk specifications for affine models: theory and evidence. J. Financ. Econ. 83, 123–170 (2007)

    Article  Google Scholar 

  9. 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 

  10. Cvitanić, J., Goukasian, L., Zapatero, F.: Monte Carlo computation of optimal portfolios in complete markets. J. Econ. Dyn. Control 27, 971–986 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  11. Davis, M.H.A.: Optimal hedging with basis risk. In: Kabanov, Yu., et al. (eds.) From Stochastic Calculus to Mathematical Finance. The Shiryaev Festschrift, pp. 169–187. Springer, Berlin (2006)

    Chapter  Google Scholar 

  12. Detemple, J.B., Garcia, R., Rindisbacher, M.: A Monte Carlo method for optimal portfolio. J. Finance LVIII, 401–446 (2003)

    Article  Google Scholar 

  13. Guasoni, P., Robertson, S.: Static fund separation of long-term investments. Math. Finance 25, 789–826 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  14. Haugh, M.B., Kogan, L., Wang, J.: Evaluating portfolio policies: a duality approach. Oper. Res. 54, 405–418 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  15. Henderson, V.: Valuation of claims on nontraded assets using utility maximization. Math. Finance 12, 351–373 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  16. Herrmann, S., Muhle-Karbe, J., Seifried, F.T.: Hedging with small uncertainty aversion. Finance Stoch. 21, 1–64 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  17. Hurd, T., Kuznetsov, A.: Explicit formulas for Laplace transforms of stochastic integrals. Markov Process. Relat. Fields 14, 277–290 (2008)

    MathSciNet  MATH  Google Scholar 

  18. Kallsen, J., Muhle-Karbe, J.: Option pricing and hedging with small transaction costs. Math. Finance 25, 702–723 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  19. Karatzas, I., Kardaras, C.: The numéraire portfolio in semimartingale financial models. Finance Stoch. 11, 447–493 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  20. 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 

  21. Karatzas, I., Lehoczky, J.P., Shreve, S.E., Xu, G.L.: Martingale and duality methods for utility maximization in an incomplete market. SIAM J. Control Optim. 29, 702–730 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  22. Karatzas, I., Shreve, S.E.: Methods of Mathematical Finance, 1st edn. Applications of Mathematics (New York), vol. 39. Springer, New York (1998)

    Book  MATH  Google Scholar 

  23. Kim, T.S., Omberg, E.: Dynamic nonmyopic portfolio behavior. Rev. Financ. Stud. 9, 141–161 (1996)

    Article  Google Scholar 

  24. Kogan, L., Uppal, R.: Risk aversion and optimal policies in partial and general equilibrium economies. Working paper (2001). Available online at http://www.nber.org/papers/w8609

  25. Kraft, H.: Optimal portfolios and Heston’s stochastic volatility model: an explicit solution for power utility. Quant. Finance 5, 303–313 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  26. Kraft, H., Seiferling, T., Seifried, F.T.: Optimal consumption and investment with Epstein–Zin recursive utility. Finance Stoch. 21, 187–226 (2017)

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  28. Kramkov, D., Sîrbu, M.: On the two-times differentiability of the value functions in the problem of optimal investment in incomplete markets. Ann. Appl. Probab. 16, 1352–1384 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  29. Kramkov, D., Sîrbu, M.: Sensitivity analysis of utility-based prices and risk-tolerance wealth processes. Ann. Appl. Probab. 16, 2140–2194 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  30. Larsen, K.: A note on the existence of the power investor’s optimizer. Finance Stoch. 15, 183–190 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  31. Larsen, K., Žitković, G.: Stability of utility-maximization in incomplete markets. Stoch. Process. Appl. 117, 1642–1662 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  32. Larsen, L., Munk, C.: The costs of suboptimal dynamic asset allocation: general results and applications to interest rate risk, stock volatility risk, and growth/value tilts. J. Econ. Dyn. Control 36, 266–293 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  33. Liu, J.: Portfolio selection in stochastic environments. Rev. Financ. Stud. 20, 1–39 (2007)

    Article  Google Scholar 

  34. Monoyios, M.: Malliavin calculus method for asymptotic expansion of dual control problems. SIAM J. Financ. Math. 4, 884–915 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  35. Mostovyi, O.: Necessary and sufficient conditions in the problem of optimal investment with intermediate consumption. Finance Stoch. 19, 135–159 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  36. Munk, C.: Financial Asset Pricing Theory, vol. 6. Oxford University Press, London (2013)

    Book  Google Scholar 

  37. Oliver, H.W.: The exact Peano derivative. Trans. Am. Math. Soc. 76, 444–456 (1954)

    Article  MathSciNet  MATH  Google Scholar 

  38. Rogers, L.C.G.: The relaxed investor and parameter uncertainty. Finance Stoch. 5, 131–154 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  39. Szpiro, G.G.: Measuring risk aversion: an alternative approach. Rev. Econ. Stat. 68, 156–159 (1986)

    Article  Google Scholar 

  40. Wachter, J.: Portfolio and consumption decisions under mean-reverting returns: an exact solution for complete markets. J. Financ. Quant. Anal. 37, 63–91 (2002)

    Article  Google Scholar 

Download references

Acknowledgements

The authors would like to thank the anonymous referees, the anonymous Associate Editor, the Editor Martin Schweizer, Milica Čudina, Claus Munk, Mihai Sîrbu, and Kim Weston for useful comments.

During the preparation of this work, the first author has been supported by the National Science Foundation under Grant No. DMS-1411809 (2014–2017). The second author has been supported by the National Science Foundation under grant No. DMS-1600307 (2015–2018). The third author has been supported by the National Science Foundation under Grant No. DMS-1107465 (2012–2017) and Grant No. DMS-1516165 (2015–2018). Any opinions, findings and conclusions or recommendations expressed in this material are those of the author(s) and do not necessarily reflect the views of the National Science Foundation (NSF).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Kasper Larsen.

Appendices

Appendix A: Details on the Kim–Omberg model

The following result summarizes the main properties in [23].

Theorem A.1

(Kim and Omberg 1996)

Let the market price of risk process be defined by (5.3), \(M:=B\), and let \(p<0\). Then there exist continuously differentiable functions \(b,c:[0,\infty)\to{\mathbb{R}}\) such that for \(t\in[0,T)\), we have

$$ \textstyle\begin{array}{rlrl} b'(t) &= -\alpha_{4}\, b(t)- \alpha_{1}\, c(t) + \alpha_{2}\, b(t) c(t), \qquad & b(T)&=0, \\ c'(t) &= q - 2 \alpha_{4}\, c(t) + \alpha_{2}\, c^{2}(t), & c(T)&=0, \end{array} $$

where \(\alpha_{1}:=\theta\kappa\), \(\alpha_{2} := (1+q) \beta^{2}+ \gamma^{2}\), \(\alpha_{3} := \beta^{2}+\gamma^{2}\) and \(\alpha_{4} := q \beta- \kappa\). Furthermore, the primal optimizer is given by

$$\begin{aligned} \hat{\pi}^{\mathit{KO}}_{t} = \frac{b(t)\beta+ (c(t)\beta-1) \lambda^{\mathit{KO}}_{t}}{p-1},\quad t\in[0,T]. \end{aligned}$$

For \(p<0\), the above Riccati equation describing \(c\) has a “normal non-exploding solution” as defined in the Appendix of [23]. Therefore, all three functions \(a\), \(b\) and \(c\) are bounded on any finite subinterval \([0,T]\) of \([0,\infty)\).

Lemma A.2

Let \((\lambda,\lambda')\) be defined by (5.4) and (5.5) and let \(p<0\). For the basis model \(\lambda\), the primal and dual optimizers are given by

$$\begin{aligned} \hat{\pi}^{(0)}_{t} = \frac{\lambda_{t}}{1-p}, \qquad \hat{\nu}^{(0)}_{t} = \gamma\big(b(t)+c(t)\lambda_{t}\big),\quad \qquad t\in[0,T]. \end{aligned}$$

Furthermore, the processes \((\gamma^{B},\gamma^{W})\) appearing in the representation (3.6) of \(\varPhi\) are given by (5.6) and (5.7), where the functions \(C_{j}\) satisfy the ODEs

$$ \renewcommand{\arraystretch}{1.3} \textstyle\begin{array}{rlrl} -C_{1}'(t)&= \tilde{b}(t) \, C_{4}(t)+ \gamma^{2} \, C_{5}(t), \qquad & C_{1}(T)&=0, \\ -C_{2}'(t)&= \tilde{b}(t) \, C_{6}(t) - \kappa\, C_{2}(t),& C_{2}(T)&=0, \\ -C_{4}'(t) &= q\, C_{2}(t) - \tilde{c}(t) \, C_{4}(t) + 2\tilde{b}(t) \, C_{5}(t), \qquad & C_{4}(T)&=0, \\ -C_{5}'(t) &= q\, C_{6}(t) -2 \tilde{c}(t) \, C_{5}(t),& C_{5}(T)&=0, \\ -C'_{6}(t) &= - \big(\kappa+\tilde{c}(t)\big) \, C_{6}(t) -1,& C_{6}(T)&=0 \end{array} $$

on \([0,T)\), with \((a,b,c)\) as in Theorem  A.1 with \(\beta:=0\), \(\tilde{b}(t):=\kappa\theta-\gamma^{2} b(t)\) and \(\tilde{c}(t) := \kappa+\gamma^{2} c(t) \). Furthermore, for the measure \(\tilde{\mathbb{P}}^{(0)}\) defined by (2.6) and for all \(T>0\), we have

$$\begin{aligned} \Delta^{(0)} &:= {\mathbb{E}}^{\tilde{\mathbb{P}}^{(0)}}\left[ \int _{0}^{T} \lambda'_{s}\hat{\pi}_{s}^{(0)} ds\right] = -\frac{1}{1-p} \big(C_{1}(0) + C_{4}(0) \lambda_{0} +C_{5}(0) \lambda_{0}^{2}\big). \end{aligned}$$

Proof

The first part follows from Theorem A.1 applied to the case \(\beta:=0\). To find the representation (3.6), we define the function

$$f(t,x,\lambda):= \frac{x^{p}}{p}e^{-a(t) - b(t)\lambda- \frac{1}{2} c(t)\lambda^{2}},\quad t\in[0,T], x>0, \lambda\in{\mathbb{R}}, $$

where the functions \((b,c)\) are as in Theorem A.1 and \(a\) is given by

$$ a'(t) = -\alpha_{1}\, b(t) - \frac{1}{2}\alpha_{3}\, c(t) + \frac{1}{2}\alpha_{2}\, b^{2}(t), \qquad a(T)=0. $$

The martingale properties of \((f(t,\hat{X}^{(0)}_{t},\lambda_{t}))\) and \((\hat{X}^{(0)}_{t}\hat{Y}^{(0)}_{t})\) as well as the proportionality property \((\hat{X}^{(0)}_{T})^{p}\propto\hat{X}^{(0)}_{T} \hat{Y} ^{(0)}_{T}\) produce

$$pf(t, \hat{X}^{(0)}_{t},\lambda_{t})=p{\mathbb{E}}[f(T, \hat{X}^{(0)} _{T},\lambda_{T})|{\mathcal{F}}_{t}]\propto{\mathbb{E}}[ \hat{Y}^{(0)} _{T}\hat{X}^{(0)}_{T}|{\mathcal{F}}_{t}] = \hat{X}^{(0)}_{t}\hat{Y} ^{(0)}_{t}. $$

By computing the dynamics of the left-hand side, we see from Girsanov’s theorem that the two processes (see (4.13))

$$\begin{aligned} \begin{aligned} dB^{\tilde{\mathbb{P}}^{(0)}}_{t} &= - q\lambda_{t}dt + dB_{t}, \\ \quad dW^{\tilde{\mathbb{P}}^{(0)}}_{t} &= \big(b(t)+c(t)\lambda_{t} \big)\gamma dt + dW_{t}, \end{aligned} \end{aligned}$$

are independent Brownian motions under \(\tilde{\mathbb{P}}^{(0)}\). These dynamics and Itô’s lemma ensure that

$$\begin{aligned} N_{t} := \int_{0}^{t} &\lambda'_{s}\lambda_{s} ds -C_{1}(t) -C_{2}(t) \lambda'_{t}- C_{4}(t) \lambda_{t} -C_{5}(t) \lambda_{t}^{2}-C_{6}(t) \lambda_{t}\lambda'_{t} \end{aligned}$$

is a \(\tilde{\mathbb{P}}^{(0)}\)-local martingale. Because the processes \((\lambda,\lambda')\) remain Ornstein–Uhlenbeck processes under \(\tilde{\mathbb{P}}^{(0)}\) and the functions \(C_{1},\dots,\! C_{6}\) are bounded, \(N\) is indeed a \(\tilde{\mathbb{P}}^{(0)}\)-martingale. Furthermore, thanks to the zero terminal conditions imposed on \(C_{1},\dots,C_{6}\), we see that

$$\begin{aligned} \varPhi=\frac{1}{1-p}\int_{0}^{T} \!\! \lambda_{t} \lambda'_{t}dt &= \frac{1}{1-p}N_{T} = \frac{1}{1-p}N _{0} +\int_{0}^{T} \!\! \gamma^{B}_{t} dB^{\tilde{{\mathbb{P}}}^{(0)}}_{t} +\int_{0}^{T} \! \! \gamma^{W}_{t} dW^{\tilde{{\mathbb{P}}}^{(0)}}_{t} \end{aligned}$$

for \((\gamma^{B},\gamma^{W})\) defined by (5.6) and (5.7). □

Appendix B: Details on the extended affine models

The following result is from [25].

Theorem B.1

(Kraft 2005)

Let \(p<0\) and consider the model (5.8) with \(\lambda_{t}:=1\). The primal optimal control is \(\hat{\pi}^{(0)}_{t} = \frac{f(t)-1}{p-1}\), the dual optimal control corresponding to \(W\) is \(\hat{\nu}^{(0)}_{t} =f(t) \sqrt{F_{t}}\), and the deterministic function \(f\) is given by

$$\begin{aligned} f'(t) &= \frac{-f(t)(f(t)(\beta^{2}+\gamma^{2})+2\kappa)+p(f(t)^{2} \gamma^{2}-1+2f(t)(\beta+\kappa))}{2 (p-1)}, \\ f(T) &=0. \end{aligned}$$

For the model (5.10) with \(\lambda_{t}:=\lambda_{21}\), the primal optimal control is \(\hat{\pi}^{(0)}_{t} = \frac{\lambda_{21}}{1-p}\), the dual optimal controls corresponding to \((W^{(1)},W^{(3)})\) are

$$\hat{\nu}^{(0)}_{t} =\Big(f(t) \sqrt{F^{(1)}_{t}},0\Big), $$

and the deterministic function \(f\) is given by

$$ f'(t)=\frac{1}{2} \bigg(-2 b_{11} f(t) + f(t)^{2} + \frac{\lambda_{21} ^{2} p}{1 - p}\bigg), \qquad f(T)=0. $$

We then turn to the representation (3.6) of \(\varPhi:= \int _{0}^{T} \hat{\pi}^{(0)}_{t} \lambda_{t}' \sigma_{t}^{2} dt\). For the model (5.8), this is trivial because \(\int_{0}^{T} \hat{\pi}^{(0)}_{t} dt\) is deterministic (implying that \(\gamma^{B}= \gamma^{W}=0\)). The next lemma provides the representation for the model in (5.11).

Lemma B.2

Let \(p<0\), let \((\lambda,\lambda')\) be defined by (5.10) and (5.11), and define \(\hat{\pi} ^{(0)}_{t} := \frac{\lambda_{21}}{1-p}\). The integrands \((\gamma^{B}, \gamma^{W})\) appearing in the representation (3.6) of

$$\varPhi:= \int_{0}^{T} \hat{\pi}^{(0)}_{t} (\lambda_{20} - \lambda_{21} +\lambda_{22}F_{t}^{(2)} + \lambda_{23}F_{t}^{(3)})dt $$

are given by

$$\begin{aligned} \gamma^{B}_{t} = -C_{2}(t),\quad\gamma^{W}_{t} = -\Big(C_{1}(t) \sqrt{F ^{(1)}_{t}},C_{3}(t) \sqrt{1+F^{(1)}_{t}}\Big), \end{aligned}$$

where the functions \(C_{j}\) satisfy on \([0,T]\) the ODEs

$$\begin{aligned} C_{1}'(t) &=-b_{11} C_{1} - b_{31} C_{3}(t)+ C_{1}(t) f(t) - C_{2}(t)(b _{21} -\lambda_{21} + \hat{\pi}^{(0)}_{t}), \\ C_{1}(T) &=0, \\ C_{2}'(t) &= -b_{22} C_{2}(t) - b_{32} C_{3}(t) + \lambda_{22} \hat{\pi}^{(0)}_{t}, \\ C_{2}(T) &=0, \\ C_{3}'(t) &= -b_{23} C_{2}(t) - b_{33} C_{3}(t) + \lambda_{23} \hat{\pi}^{(0)}_{t}, \\ C_{3}(T) &=0. \end{aligned}$$

Proof

This is similar to the proof of Lemma A.2. Because the functions \(C_{j}\) satisfy the above ODEs, we have that

$$\begin{aligned} N_{t} &:= \int_{0}^{t} \hat{\pi}^{(0)}_{u} (\lambda_{20} - \lambda _{21} +\lambda_{22}F_{u}^{(2)} + \lambda_{23}F_{u}^{(3)})du \\ & \phantom{=::} - C_{0}(t) - C_{1}(t)F^{(1)}_{t}- C_{2}(t)F^{(2)}_{t}- C_{3}(t)F^{(3)} _{t} \end{aligned}$$

is a \(\tilde{{\mathbb{P}}}^{(0)}\)-martingale (here we use the \(\tilde{\mathbb{P}}^{(0)}\)-Brownian motions defined in (4.13)). The function \(C_{0}\) is given by

$$\begin{aligned} C_{0}'(t) &= -a_{1} C_{1}(t) - a_{2} C_{2}(t) - a_{3} C_{3}(t) + C_{2}(t) \lambda_{21} - \big(C_{2}(t) - \lambda_{20} + \lambda_{21}\big) \hat{\pi}^{(0)}_{t} \end{aligned}$$

with \(C_{0}(T)=0\). Then we have

$$\begin{aligned} {\mathbb{E}}^{\tilde{{\mathbb{P}}}^{(0)}}[\varPhi|{\mathcal{F}}_{t}] = {\mathbb{E}}^{\tilde{{\mathbb{P}}}^{(0)}}[N_{T}|{\mathcal{F}}_{t}] =N _{t}. \end{aligned}$$

Hence, we find the integrands appearing in the representation (3.6) of \(\varPhi\) to be as in the statement. □

Finally, we verify the integrability conditions (3.7) and (3.8) in Theorem 3.4, provided that \(T>0\) is small. We can define the second-order optimal controls \((\tilde{\pi}, \tilde{\nu})\) by (3.12) and (3.13). We start by considering the model (5.8). Theorem 3.1 in [17] states that the joint Laplace transform of

$$\varLambda:=\int_{0}^{T} \frac{1}{F_{s}}\, ds,\qquad Q:=\int_{0}^{T} F _{s}\, ds $$

is finite for all \(T>0\) in some neighborhood of 0 as soon as the strict Feller condition \(2\kappa\theta> 1\) holds. This implies that both \(\varLambda\) and \(Q\) have some finite exponential positive moments (and consequently, \(\varLambda\) and \(Q\) have all moments). Because the integrand \(\gamma^{B}\) appearing in the representation (3.6) of \(\varPhi\) is zero in this model, the integrand appearing inside the exponential in (3.8) is deterministic. Consequently, (3.8) holds for all \(\varepsilon>0\).

For the model (5.11), we have the estimate

$$\int_{0}^{T} (\lambda'_{t})^{2} (1+F^{(1)}_{t}) dt\le\int_{0}^{T} ( \lambda_{20} - \lambda_{21} +\lambda_{22}F_{t}^{(2)} + \lambda_{23}F _{t}^{(3)})^{2}dt, $$

which shows that the moment requirements in (3.7) hold. Furthermore, by inserting \(\lambda'\) defined in (5.11) and \(\sigma^{2}_{t} = 1+F^{(1)}_{t}\) into the exponent in (3.8), we see that this exponent is affine in \(F^{(1)}\), \(F^{(2)}\) and \(F^{(3)}\) (with time-dependent coefficients). Consequently, the expectation in (3.8) is finite for small positive values of \(\varepsilon\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Larsen, K., Mostovyi, O. & Žitković, G. An expansion in the model space in the context of utility maximization. Finance Stoch 22, 297–326 (2018). https://doi.org/10.1007/s00780-017-0353-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00780-017-0353-3

Keywords

Mathematics Subject Classification (2010)

JEL Classification

Navigation