Skip to main content
Log in

Asymptotics for fixed transaction costs

  • Published:
Finance and Stochastics Aims and scope Submit manuscript

Abstract

An investor with constant relative risk aversion trades a safe and several risky assets with constant investment opportunities. For a small fixed transaction cost, levied on each trade regardless of its size, we explicitly determine the leading-order corrections to the frictionless value function and optimal policy.

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.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3

Similar content being viewed by others

Notes

  1. Indeed, these schemes asymptotically correspond to constant absolute risk version; see [17] for more details.

  2. For our formal derivations, we consider general utilities like in recent independent work of Alcala and Fahim [1].

  3. Here, both quantities are measured in relative terms, as is customary for investors with constant relative risk aversion. That is, trading boundaries are parameterized by the fractions of wealth held in the risky asset, and the welfare effect is described by the relative certainty equivalent loss, that is, the fraction of the initial endowment the investor would be willing to give up to trade without frictions.

  4. See [27, 1] for corresponding formal asymptotics.

  5. By convention, the value of the integral is set to minus infinity if its negative part is infinite.

  6. To see this, formally let the time horizon tend to infinity in [24, Sects. 4.1 and 4.2] and insert the explicit formulas for the optimal consumption rate and risky weight. This immediately yields that the leading-order no-trade regions coincide; for the corresponding welfare effects, this follows after integrating.

  7. To facilitate comparison, we use the same market parameters μ,σ,r and risk aversion γ as in Muthuraman and Kumar [35]. The fixed cost and the current wealth are chosen so that the one-dimensional no-trade region for each asset corresponds to the one for their 1 % proportional cost.

  8. Compare Nutz [36] for a general frictionless setting and Guasoni and Muhle-Karbe [20] for a model with proportional transaction costs. A similar result for fixed costs is more difficult to formulate because the investor’s wealth does not factor out of the trading policy in this case.

  9. The full multidimensional derivation can be found in [43]. In the no-trade region, the calculations are identical.

  10. This will turn out to be consistent with the results of our calculations below; see Sects. 3.5 and 3.6.

  11. For logarithmic utility (γ=1), this follows similarly by additionally exploiting the estimate (4.5).

References

  1. Alcala, J.V., Fahim, A.: Balancing small fixed and proportional transaction cost in trading strategies (2013). Available online http://arxiv.org/abs/1304.7562.pdf

  2. Atkinson, C., Wilmott, P.: Portfolio management with transaction costs: an asymptotic analysis of the Morton and Pliska model. Math. Finance 5, 357–367 (1995)

    Article  MATH  Google Scholar 

  3. Balduzzi, P., Lynch, A.: Transaction costs and predictability: some utility cost calculations. J. Financ. Econom. 52, 47–78 (1999)

    Article  Google Scholar 

  4. Barles, G., Perthame, B.: Discontinuous solutions of deterministic optimal stopping time problems. RAIRO Modél. Math. Anal. Numér. 21, 557–579 (1987)

    MATH  MathSciNet  Google Scholar 

  5. Bertsimas, D., Kogan, L., Lo, A.W.: When is time continuous? J. Financ. Econom. 55, 173–204 (2000)

    Article  Google Scholar 

  6. Bichuch, M.: Asymptotic analysis for optimal investment in finite time with transaction costs. SIAM J. Financ. Math. 3, 433–458 (2012)

    Article  MATH  MathSciNet  Google Scholar 

  7. Bichuch, M., Shreve, S.E.: Utility maximization trading two futures with transaction costs. SIAM J. Financ. Math. 4, 26–85 (2013)

    Article  MATH  MathSciNet  Google Scholar 

  8. Bouchard, B., Touzi, N.: Weak dynamic programming principle for viscosity solutions. SIAM J. Control Optim. 49, 948–962 (2011)

    Article  MATH  MathSciNet  Google Scholar 

  9. Constantinides, G.: Capital market equilibrium with transaction costs. J. Polit. Econ. 94, 842–862 (1986)

    Article  Google Scholar 

  10. Crandall, M.G., Ishii, H., Lions, P.-L.: User’s guide to viscosity solutions of second order partial differential equations. Bull. Am. Math. Soc. 27, 1–67 (1992)

    Article  MATH  MathSciNet  Google Scholar 

  11. Davis, M.H.A., Norman, A.R.: Portfolio selection with transaction costs. Math. Oper. Res. 15, 676–713 (1990)

    Article  MATH  MathSciNet  Google Scholar 

  12. Dumas, B., Luciano, E.: An exact solution to a dynamic portfolio choice problem under transaction costs. J. Finance 46, 577–595 (1991)

    Article  Google Scholar 

  13. Eastham, J.F., Hastings, K.J.: Optimal impulse control of portfolios. Math. Oper. Res. 13, 588–605 (1988)

    Article  MATH  MathSciNet  Google Scholar 

  14. Evans, L.C.: The perturbed test function method for viscosity solutions of nonlinear PDE. Proc. R. Soc. Edinb. A 111, 359–375 (1989)

    Article  MATH  Google Scholar 

  15. Fleming, W.H., Soner, H.M.: Controlled Markov Processes and Viscosity Solutions, 2nd edn. Springer, New York (2006)

    MATH  Google Scholar 

  16. Fukasawa, M.: Asymptotically efficient discrete hedging. In: Kohatsu-Higa, A., Privault, N., Sheu, S. (eds.) Stochastic Analysis with Financial Applications. Progress in Probability, vol. 65, pp. 331–346. Birkhäuser, Basel (2011)

    Chapter  Google Scholar 

  17. Fukasawa, M.: Efficient discretization of stochastic integrals. Finance Stoch. 18, 175–208 (2014)

    Article  MATH  MathSciNet  Google Scholar 

  18. Gerhold, S., Guasoni, P., Muhle-Karbe, J., Schachermayer, W.: Transaction costs, trading volume, and the liquidity premium. Finance Stoch. 18, 1–37 (2014)

    Article  MATH  MathSciNet  Google Scholar 

  19. Goodman, J., Ostrov, D.N.: Balancing small transaction costs with loss of optimal allocation in dynamic stock trading strategies. SIAM J. Appl. Math. 70, 1977–1998 (2010)

    Article  MATH  MathSciNet  Google Scholar 

  20. Guasoni, P., Muhle-Karbe, J.: Long horizons, high risk aversion, and endogeneous spreads. Math. Finance (2013). Available online http://onlinelibrary.wiley.com/doi/10.1111/mafi.12046/abstract

  21. Hayashi, T., Mykland, P.A.: Evaluating hedging errors: an asymptotic approach. Math. Finance 15, 309–343 (2005)

    Article  MATH  MathSciNet  Google Scholar 

  22. Janeček, K., Shreve, S.E.: Asymptotic analysis for optimal investment and consumption with transaction costs. Finance Stoch. 8, 181–206 (2004)

    Article  MATH  MathSciNet  Google Scholar 

  23. Kabanov, Y.M., Safarian, M.M.: Markets with Transaction Costs. Springer, Berlin (2009)

    MATH  Google Scholar 

  24. Kallsen, J., Muhle-Karbe, J.: The general structure of optimal investment and consumption with small transaction costs (2013). Available online http://arxiv.org/abs/1303.3148

  25. Kallsen, J., Muhle-Karbe, J.: Option pricing and hedging with small transaction costs. Math. Finance (2013). Available online http://onlinelibrary.wiley.com/doi/10.1111/mafi.12046/abstract

  26. Kelley, J.: General Topology. Van Nostrand-Reinhold, New York (1955)

    MATH  Google Scholar 

  27. Korn, R.: Portfolio optimisation with strictly positive transaction costs and impulse control. Finance Stoch. 2, 85–114 (1998)

    Article  MATH  MathSciNet  Google Scholar 

  28. Liu, H.: Optimal consumption and investment with transaction costs and multiple risky assets. J. Finance 59, 289–338 (2004)

    Article  Google Scholar 

  29. Lo, A.W., Mamaysky, H., Wang, J.: Asset prices and trading volume under fixed transaction costs. J. Polit. Econ. 112, 1054–1090 (2004)

    Article  Google Scholar 

  30. Lynch, A., Balduzzi, P.: Predictability and transaction costs: the impact on rebalancing rules and behavior. J. Finance 55, 2285–2310 (2000)

    Article  Google Scholar 

  31. Lynch, A., Tan, S.: Explaining the magnitude of liquidity premia: the roles of return predictability, wealth shocks, and state dependent transaction costs. J. Finance 66, 1329–1368 (2011)

    Article  Google Scholar 

  32. Magill, M., Constantinides, G.: Portfolio selection with transaction costs. J. Econ. Theory 13, 245–263 (1976)

    Article  MATH  MathSciNet  Google Scholar 

  33. Martin, R.: Optimal multifactor trading under proportional transaction costs (2012). Available online http://arxiv.org/abs/1204.6488

  34. Morton, A.R., Pliska, S.R.: Optimal portfolio management with fixed transaction costs. Math. Finance 5, 337–356 (1995)

    Article  MATH  Google Scholar 

  35. Muthuraman, K., Kumar, S.: Multidimensional portfolio optimization with proportional transaction costs. Math. Finance 16, 301–335 (2006)

    Article  MATH  MathSciNet  Google Scholar 

  36. Nutz, M.: Risk aversion asymptotics for power utility maximization. Probab. Theory Relat. Fields 152, 703–749 (2012)

    Article  MATH  MathSciNet  Google Scholar 

  37. Øksendal, B., Sulem, A.: Optimal consumption and portfolio with both fixed and proportional transaction costs. SIAM J. Control Optim. 40, 1765–1790 (2002)

    Article  MathSciNet  Google Scholar 

  38. Possamaï, D., Soner, H.M., Touzi, N.: Homogenization and asymptotics for small transaction costs: the multidimensional case (2012). Available online http://arxiv.org/abs/1212.6275

  39. Rogers, L.C.G.: Why is the effect of proportional transaction costs O(δ 2/3). In: Yin, G., Zhang, Q. (eds.) Mathematics of Finance, pp. 303–308. Am. Math. Soc., Providence (2004)

    Chapter  Google Scholar 

  40. Rosenbaum, M., Tankov, P.: Asymptotically optimal discretization of hedging strategies with jumps. Ann. Appl. Probab. 23, 1002–1048 (2014)

    Article  MathSciNet  Google Scholar 

  41. Schroder, M.: Optimal portfolio selection with fixed transaction costs: numerical solutions (1995). Available online https://www.msu.edu/~schrode7/numerical.pdf

  42. Shreve, S.E., Soner, H.M.: Optimal investment and consumption with transaction costs. Ann. Appl. Probab. 4, 609–692 (1994)

    Article  MATH  MathSciNet  Google Scholar 

  43. Soner, H.M., Touzi, N.: Homogenization and asymptotics for small transaction costs. SIAM J. Control Optim. 51, 2893–2921 (2013)

    Article  MATH  MathSciNet  Google Scholar 

  44. Whalley, A.E., Wilmott, P.: An asymptotic analysis of an optimal hedging model for option pricing with transaction costs. Math. Finance 7, 307–324 (1997)

    Article  MATH  MathSciNet  Google Scholar 

  45. Zhang, R.: Couverture Approchée des Options Européennes. Ph.D. thesis, Ecole Nationale des Ponts et Chaussées (1999). Available online https://tel.archives-ouvertes.fr/tel-00005623/document

Download references

Acknowledgements

The authors thank Bruno Bouchard, Paolo Guasoni, Jan Kallsen, Ludovic Moreau, Mathieu Rosenbaum, and Peter Tankov for fruitful discussions and Martin Forde for pertinent remarks on an earlier version. Moreover, they are very grateful to two anonymous referees and the editor for numerous constructive comments.

The second author was partially supported by the National Centre of Competence in Research “Financial Valuation and Risk Management” (NCCR FINRISK), Project D1 (Mathematical Methods in Financial Risk Management) of the Swiss National Science Foundation (SNF), and the Swiss Finance Institute. The third author gratefully acknowledges partial support by the European Research Council under the grant 228053-FiRM, by the ETH Foundation, and the Swiss Finance Institute.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Johannes Muhle-Karbe.

Appendices

Appendix A: Pointwise estimates

Proposition A.1

There exists K=K(β,μ,r,σ,γ)>0 such that for k=0,1,2 and j=0,1,2,

$$|D^k_{\xi}\partial^j_z w(z,\xi)|\leq K z^{-j-3k/4-\gamma} \quad \textit{for all }(z,\xi) \leftrightarrow(x,y)\in\mathrm {NT}^{\epsilon}. $$

Proof

This follows from tedious but straightforward computations since all the functions and domains involved are known explicitly (cf. [43, Sect. 4.2] for a similar calculation). □

Remark A.2

Proposition A.1 yields the expansion

where the remainder satisfies the bound

$$|{\mathcal{R}}_w(z,\xi)|\leq K \sum_{k=0}^3\epsilon^{k}z^{(1-k)/4-\gamma }\quad \mbox{for all }(z,\xi) \leftrightarrow(x,y)\in\mathrm{NT}^{\epsilon}. $$

In particular, over the same region, we have

We can also expand

with a bound on the remainder of

$$|{\mathcal{R}}_u(z,\xi)|\leq K(\epsilon z^{1/4-\gamma} + \epsilon^2 z^{-\gamma})\quad \mbox{for all }(z,\xi) \leftrightarrow(x,y)\in\mathrm {NT}^{\epsilon}. $$

Proposition A.3

Let ϵ>0 be given and consider \(S:=[z_{0}-r_{0},z_{0}+r_{0}]\times \mathbb {R}^{d}\subset\mathrm{K}_{\epsilon}\) for some z 0>r 0>0. Then, given any ΨC 2(S) for which D ξ Ψ has a compact support, there exists K>0, independent of ϵ, such that

$$\|\varPsi\|_{C^2(S)} \leq K $$

and

Proof

This again follows from a tedious but straightforward calculation. □

Appendix B: Proof of Theorem 2.1

In this section, we prove that for each fixed λ, the value function v λ is a viscosity solution of the corresponding dynamic programming equation (3.7) on the domain

$${\mathcal{O}}_\lambda= \{ (x,y) \in\mathrm{K_\lambda}\ :\ x+ y \cdot{\mathbf{1}}_d > 2\lambda\}. $$

As observed by Bouchard and Touzi [8], a “weak version” of the dynamic programming principle is sufficient to derive the viscosity property. For the convenience of the reader, we present a direct proof of the weak dynamic programming principle in our specific setting using the techniques of [8]. Then we use it to prove that v λ is indeed a viscosity solution of (3.7).

2.1 B.1 Weak dynamic programming principle for v λ

Fixing \((x,y) \in {\mathcal{O}}_{\lambda}\) and δ>0, let \(B_{\delta }(x,y)\subset \mathbb{R}^{d+1}\) denote the ball of radius δ centered at (x,y) and set

$$K(x,y;\delta)_\lambda:= \{ (x^\prime,y^\prime)\ :\ x+ y \cdot{\mathbf{1}}_d -\delta-\lambda\le x^\prime+ y^\prime\cdot{\mathbf{1}}_d \le x+ y \cdot{\mathbf{1}}_d +\delta\}. $$

Take δ>0 sufficiently small so that \(K(x,y;2\delta)_{\lambda}\subset {\mathcal{O}}_{\lambda}\). For any investment–consumption policy ν and initial endowment (x′,y′)∈B δ/2(x,y), define θ:=θ ν as the exit time of the state process (X,Y)ν,x′,y from B δ/2(x,y). Following the standard convention, our notation does not explicitly show the dependence of θ on ν. It is then clear that

$$(X_{\theta^-},Y_{\theta^-}) \in\overline{B_{\delta/2}(x,y)}\quad {\mbox{and}} \quad (X_{\theta},Y_{\theta}) \in K(x,y,\delta)_\lambda. $$

The following weak version of the DPP is introduced in [8]. Let φ be a smooth and bounded function on K(x,y,2δ) λ satisfying

$$v^{\lambda} \leq\varphi\quad\mbox{on } K(x,y,2\delta)_\lambda. $$

Then we have

$$ v^\lambda(x,y) \le\sup_{\nu\in\varTheta_\lambda(x,y)}\mathbb{E}\left[ \int_0^\theta e^{-\beta t} U(c_t)\,dt + e^{-\beta\theta} \varphi\big( X_{\theta},Y_{\theta}\big) \right]. $$
(B.1)

(The restriction to bounded test functions φ is possible since by (4.12) v λ is bounded on K(x,y;2δ) λ .) Conversely, let φ be a smooth function bounded on K(x,y,2δ) λ , satisfying

$$v^{\lambda} \geq\varphi\quad\mbox{on } K(x,y,2\delta)_\lambda. $$

Then we have

$$ v^\lambda(x,y) \ge\sup_{\nu\in\varTheta_\lambda(x,y)}\mathbb{E}\left[ \int_0^\theta e^{-\beta t} U(c_t)\,dt + e^{-\beta\theta} \varphi\big( X_{\theta},Y_{\theta}\big) \right]. $$
(B.2)

For proving (B.1) and (B.2), without loss of generality, let \(\varOmega= C_{0}([0,\infty), \mathbb{R}^{d})\) be the space of continuous functions starting at zero, equipped with the Wiener measure \(\mathbb{P}\), a standard Brownian motion W, and the completion \(({\mathcal{F}}_{t})_{t\geq0}\) of the filtration generated by W. Given a control νΘ λ (x,y) and the exit time θ:=θ ν from above, fix ωΩ and define

$$\nu^{\theta,\omega}(\omega',t) := \nu\big(\omega\stackrel {\theta}{\oplus }\omega', t + \theta(\omega)\big)\qquad\forall\omega'\in\varOmega ,\ t\geq0, $$

where

$$(\omega\stackrel{\theta}{\oplus}\omega')_t := \left\{ \begin{array}{ll} \omega_t & \quad\mbox{if }t\in[0,\theta(\omega)), \\ \omega'_{t-\theta(\omega)} + \omega_{\theta(\omega)} & \quad \mbox{if }t\geq\theta(\omega). \end{array} \right. $$

We start with the proof of (B.1). By construction,

$$\nu^{\theta,\omega}\in\varTheta_{\lambda}\Big(X_{\theta(\omega )}^{\nu ,x},Y_{\theta(\omega)}^{\nu,y}\Big); $$

in particular, ν θ,ω is a well-defined impulse control. Moreover, note that

$$(X_{\theta(\omega)}^{\nu,x},Y_{\theta(\omega)}^{\nu,y}) \in K(x,y,\delta )_\lambda $$

lies in the set K(x,y,2δ) λ on which φ dominates v λ by definition. Therefore,

$$\begin{aligned} &\left.\mathbb{E}\left[\int_0^{\infty} e^{-\beta t}U(c^{\nu}_t)\,dt \right |{\mathcal{F}}_{\theta} \right](\omega)\\ &\qquad= \int_0^{\theta(\omega)}e^{-\beta t}U\big(c^{\nu }_t(\omega) \big)\,dt + e^{-\beta\theta(\omega)}\int_{\varOmega}\int_0^{\infty} e^{-\beta t}U\big(c_t^{\nu^{\theta,\omega}}(\omega')\big)\,dt \,d\mathbb{P}(\omega ')\\ & \qquad\leq\int_0^{\theta(\omega)}e^{-\beta t}U\big(c^{\nu }_t(\omega) \big)\,dt + e^{-\beta\theta(\omega)}v^{\lambda}\big(X_{\theta (\omega )}^{\nu,x},Y_{\theta(\omega)}^{\nu,y}\big)\\ &\qquad\leq\int_0^{\theta(\omega)}e^{-\beta t}U\big(c^{\nu }_t(\omega) \big)\,dt + e^{-\beta\theta(\omega)}\varphi\big(X_{\theta(\omega )}^{\nu ,x},Y_{\theta(\omega)}^{\nu,y}\big). \end{aligned}$$

As a result, for any νΘ λ (x,y),

$$\begin{aligned} \mathbb{E}\left[\int_0^\infty e^{-\beta t} U(c^{\nu}_t)\,dt\right] \le \mathbb{E}\left[\int_0^\theta e^{-\beta t} U(c^{\nu}_t)\,dt + e^{-\beta\theta}\varphi(X_{\theta}^{\nu,x},Y_{\theta}^{\nu ,y})\right]. \end{aligned}$$

By taking the supremum over all policies ν we arrive at (B.1).

To prove (B.2), set \(\mathbb{V}\) to be the right-hand side of (B.2), that is,

$$\mathbb{V}:= \sup_{\nu\in\varTheta_\lambda(x,y)}\mathbb{E}\left[\int_0^\theta e^{-\beta t} U(c^{\nu}_t)\,dt + e^{-\beta\theta}\varphi(X_{\theta}^{\nu,x},Y_{\theta}^{\nu ,y})\right]. $$

For any η>0, we can choose ν ηΘ λ (x,y) satisfying

$$ \mathbb{V}\le\eta+\mathbb{E}\left[ \int_0^\theta e^{-\beta t} U(c^{\nu^\eta}_t)\,dt + e^{-\beta\theta}\varphi\big(X_{\theta}^{\nu{^\eta},x},Y_{\theta }^{\nu ^{\eta},y}\big) \right]. $$
(B.3)

We have already argued that \((X_{\theta}^{\nu^{\eta},x} ,Y_{\theta}^{\nu ^{\eta},y} )\in K(x,y,\delta)_{\lambda}\). The next step is to construct a countable open cover of K(x,y,δ) λ . For every point \(\zeta= (\tilde{x}, \tilde{y})\) in K(x,y,2δ) λ , set

$$\begin{aligned} R(\zeta):=\ & R_\eta(\tilde{x}, \tilde{y})\\ =\ &\{ (x^\prime,y^\prime) \in K(x,y,2\delta)_\lambda\ :\ x^\prime>\tilde{x}, \ y^\prime> \tilde{y}, \ \varphi(x^\prime,y^\prime) < \varphi(\tilde{x}, \tilde{y}) + \eta\}. \end{aligned}$$

By monotonicity of the value function,

$$ v^\lambda(\zeta) \le v^\lambda(x^\prime,y^\prime) \quad\forall\ (x^\prime,y^\prime) \in R(\zeta). $$

Also, since φ is smooth, each R(ζ) is open, and

$$K(x,y,\delta)_\lambda\subset\bigcup_{\zeta\in K(x,y,2\delta )_\lambda } R(\zeta). $$

Hence, by the Lindelöf covering lemma [26, Theorem 15], we can extract a countable subcover

$$K(x,y,\delta)_\lambda\subset\bigcup_{n\in \mathbb{N}}\ R(\zeta_n). $$

Now define a mapping \({\mathcal{I}}: K(x,y,\delta)_{\lambda}\to \mathbb{N}\) that assigns to each point one of the neighborhoods in the subcover to which it belongs by

$${\mathcal{I}}(x^\prime,y^\prime):= \min\{n: (x^\prime,y^\prime) \in R(\zeta _n)\} $$

and set

$$\zeta(x^\prime,y^\prime):= \zeta_{{\mathcal{I}}(x^\prime,y^\prime)}. $$

By definition,these constructions imply

$$ \varphi(x^\prime,y^\prime) \le\varphi\big(\zeta(x^\prime ,y^\prime)\big) + \eta\quad \forall\ (x^\prime,y^\prime) \in K(x,y,\delta)_\lambda. $$
(B.4)

As a final step, for each positive integer n, we choose a control ν nΘ λ (ζ n ) so that

$$ v^\lambda(\zeta_n) \le \mathbb{E}\left[ \int_0^\infty e^{-\beta t} U(c^{\nu^n}_t)\,dt \right] + \eta. $$
(B.5)

By monotonicity, ν nΘ λ (x′,y′) for every (x′,y′)∈R(ζ n ). We now define a composite strategy ν that follows the policy η satisfying (B.3) until the corresponding state process \((X,Y)^{\nu ^{\eta },x,y}\) leaves B δ/2(x,y) at time \(\theta= \theta^{\nu ^{\eta }}\). It then switches to the policy ν n corresponding to the index n that the state process is assigned to by the mapping \(\mathcal{I}\), that is,

$$\nu^*(\omega\stackrel{\theta}{\oplus}\omega', t):= \left\{ \begin{array}{ll} \nu^\eta(\omega,t) &\quad{\mbox{if}}\ t \in[0,\theta(\omega)],\\ \nu^{{\mathcal{N}}(\omega)}(\omega', t-\theta(\omega))&\quad{\mbox{if}}\ t >\theta(\omega),\ \end{array} \right. $$

with \({\mathcal{N}}(\omega)= {\mathcal{I}}(X_{\theta(\omega)}^{\nu^{\eta },x},Y_{\theta (\omega)}^{\nu^{\eta},y})\). This construction ensures that we have ν Θ λ (x,y). Hence, it follows from the definitions of the value function and ν , inequalities (B.5), v λφ (which holds on K(x,y,2δ) λ by the definition of φ), (B.4), and (B.3) that

$$\begin{aligned} v^\lambda(x,y) &\geq \mathbb{E}\left[\int_0^{\infty}e^{-\beta t}U(c_t^{\nu ^*})\,dt \right]\\ &= \mathbb{E}\left[ \int_0^\theta e^{-\beta t} U(c^\eta_t)\,dt + e^{-\beta\theta} \int_0^{\infty}e^{-\beta t}U(c_t^{{\mathcal{N}}})\,dt \right ]\\ &\ge \mathbb{E}\left[ \int_0^\theta e^{-\beta t} U(c^\eta_t)\,dt + e^{-\beta\theta} \bigg(\varphi\Big(\zeta\big(X_{\theta}^{\nu ^{\eta },x},Y_{\theta}^{\nu^{\eta},y}\big)\Big) - \eta\bigg)\right]\\ &\ge \mathbb{E}\left[ \int_0^\theta e^{-\beta t} U(c^\eta_t)\,dt + e^{-\beta\theta} \Big(\varphi\big(X_{\theta}^{\nu^{\eta },x},Y_{\theta }^{\nu^{\eta},y}\big) -2 \eta\Big)\right]\\ &\ge \mathbb{V}- 3 \eta. \end{aligned}$$

Since η was arbitrary, this establishes (B.2), thereby completing the proof.

2.2 B.2 v λ Is a viscosity solution of (3.7)

We first state and prove some facts about the intervention operator M from (3.8), which are needed in the subsequent proofs. Similar observations appear in [37] for the case of proportional and fixed transaction costs. The modifications below are required to extend them to the case of pure fixed costs, for which the set of attainable portfolios at a fixed wealth level is no longer compact.

Throughout, \(\underline{\psi}\) and \(\overline{\psi}\) will denote the lower- and upper-semicontinuous envelopes of a locally bounded function ψ, respectively.

Lemma B.1

Suppose that \(\varphi:\mathrm{K}_{\lambda}\to \mathbb{R}\) satisfies sup zK φ(z,⋅)∥<∞ for every nonempty compact set \(K\subset \mathbb{R}_{+}\).

  1. (i)

    If φ is lower-semicontinuous, then M φ is lower-semicontinuous. In particular, if φM φ, then \(\underline{\varphi} \geq\mathbf{M}\underline {\varphi}\).

  2. (ii)

    Let φC 1(K λ ). If (z,ξ)↦D ξ φ(z,ξ) is compactly supported on \(C\times \mathbb{R}^{d}\) for any compact set CR +, then M φ is upper-semicontinuous.

Proof

(i) Assume to the contrary that there exist ζ 0=(z 0,ξ 0)∈K λ , a constant η>0, and a sequence K λ ζ n =(z n ,ξ n )→ζ 0 for which

$$\mathbf{M}\varphi(\zeta_0) > \liminf_{n\to\infty}\mathbf {M}\varphi(\zeta _n) + 2\eta. $$

Choose \(\zeta_{0}^{*} = (z_{0}-\lambda, \hat{\xi}_{0})\) such that \(\varphi (\zeta _{0}^{*}) + \eta/2 \geq\mathbf{M}\varphi(\zeta_{0})\) and \(\zeta_{n}^{*} = (z_{n}-\lambda, \hat{\xi}_{n})\) with \(\varphi(\zeta_{n}^{*})+1/n \geq \mathbf {M}\varphi(\zeta_{n})\). Then, up to choosing a subsequence,

$$\varphi(\zeta_0^*) \geq\lim_{n\to\infty}\varphi(\zeta_n^*) + \eta. $$

By the lower-semicontinuity of φ, there exist an open neighborhood O of \(\xi_{0}^{*}\) and an integer N>0 such that for all nN and all ζO,

$$ \varphi(\zeta)\geq\varphi(\zeta_n^*) + \eta. $$
(B.6)

Observe that since z n z 0 as n→∞, we have

$$ A_n:=\{z_n-\lambda\}\times \mathbb{R}\cap O \neq\emptyset. $$
(B.7)

Combining (B.7) and (B.6) yields

$$\varphi(\zeta_n^*) + \frac{1}{n} \geq\mathbf{M}\varphi(\zeta_n) \geq \sup_{\zeta'\in A_n}\varphi(\zeta')\geq\varphi(\zeta_n^*)+ \eta, $$

which is a contradiction for n large enough. Finally, observe that if φ is lower-semicontinuous and φM φ, then \(\underline{\varphi} \geq\underline{\mathbf{M}\varphi} = \mathbf {M}\underline{\varphi}\) by the previous discussion.

(ii) This follows similarly as in the proof of [37, Lemma 3.2(i)] because the requirements we place on the gradient D ξ φ ensure the existence of optimizers and accumulation points also in our setting. □

We are now ready to tackle the proof of Theorem 2.1, which we split into two lemmas.

Lemma B.2

The value function v λ is a viscosity supersolution of the dynamic programming equation (3.7) on \({\mathcal{O}}_{\lambda}\).

Proof

Let \((x_{0},y_{0}){\,\in\,}{\mathcal{O}}_{\lambda}\), and let φ be a smooth and bounded function on K(x 0,y 0,δ) λ satisfying

$$0 = (\underline{v}^{\lambda} - \varphi)(x_0,y_0) = \min\{ (\underline {v}^{\lambda} - \varphi)(x',y'):\ (x',y')\in K(x_0,y_0,\delta )_\lambda \}. $$

Using Lemma B.1 and the inequality \(\underline{v}^{\lambda }\geq\varphi\) on K(x 0,y 0,δ) λ , we obtain

$$\varphi(x_0,y_0) = \underline{v}^{\lambda}(x_0,y_0)\geq\mathbf {M}\underline{v}^{\lambda}(x_0,y_0) \geq\mathbf{M}\varphi(x_0,y_0). $$

Therefore, it remains to show that

Assume to the contrary that (βφU(c )+c φ x −ℒφ)(x 0,y 0)<0 for some c >0 and set ϕ(x,y):=φ(x,y)−ϵ(|xx 0|4+∥yy 04). Then for ϵ>0 and r>0 small enough, continuity yields

Select a convergent sequence of points \((x_{n},y_{n},v^{\lambda }(x_{n},y_{n}))\to(x_{0},y_{0},\underline{v}^{\lambda}(x_{0},y_{0}))\) and denote by \((X^{n}_{t},Y^{n}_{t}):=(X_{t}^{x_{n}},Y_{t}^{y_{n}})\) the portfolio process starting at (x n ,y n ) under the consumption-only strategy c t c . Define

$$H^n: = \inf\{t\geq0:\ (X^n_t, Y^n_t) \notin\overline{B}_r(x_0,y_0) \} $$

and note that \(\liminf_{n\to\infty} \mathbb{E}[H^{n}] > 0 \). Hence, there exists δ>0 with \(\mathbb{E}[e^{-\beta H^{n}}] > \delta\) for all n sufficiently large. Itô’s formula gives

By construction of ϕ there exists η>0 with φϕ+η on \(K(x_{0},y_{0},\delta)\backslash\overline {B}_{r}(x_{0},y_{0})\). Hence,

$$\phi(x_n,y_n) \leq \mathbb{E}\bigg[ e^{-\beta H^n}\varphi(X^n_{H^n}, Y^n_{H^n}) + \int_0^{H^n}e^{-\beta s}U(c^*)\,ds \bigg] -\delta\eta. $$

Taking into account (v λϕ)(x n ,y n )→0, we note that for n large enough,

$$v^{\lambda}(x_n,y_n) \leq \mathbb{E}\bigg[ e^{-\beta H^n}\varphi(X^n_{H^n}, Y^n_{H^n}) + \int_0^{H^n}e^{-\beta s}U(c^*)\,ds \bigg] -\frac{\delta \eta }{2}. $$

This contradicts the weak dynamic programming principle (B.2) for v λ, thereby completing the proof. □

The image of an arbitrary smooth function under M is upper-semicontinuous only under additional assumptions (cf. Lemma B.1(ii)). As is customary in the theory of viscosity solutions (see, e.g., Sect. 9 of [10]), the viscosity subsolution property in the following lemma is therefore formulated in terms of the lower-semicontinuous envelope of the DPE.

Lemma B.3

The value function v λ is a viscosity subsolution of

Proof

Step 1. Throughout this proof, C>0 denotes a generic constant that may vary from line to line. We argue by contradiction. Let \((x_{0},y_{0})\in {\mathcal{O}}_{\lambda}\), and let φ be a smooth and bounded function on K(x 0,y 0,δ) λ satisfying

$$0 = (\overline{v}^{\lambda} - \varphi)(x_0,y_0) = \max\{(\overline {v}^{\lambda} - \varphi)(x',y'):\ (x',y')\in K(x_0,y_0,\delta )_\lambda \}. $$

Suppose that for some η>0, we have

By continuity, there is a small rectangular neighborhood

$$N = N(x_0,y_0,\rho) := \Big\{ (x,y)\in \mathbb{R}\times \mathbb{R}^d:\ \max _{i=1,\ldots, d}\big(|x-x_0|, |y^{i} - y^i_0|\big)< \rho\Big\} $$

such that

(B.8)

for all c>0 and (x,y)∈N.

Step 2. Choose a sequence N∋(x n ,y n )→(x 0,y 0) for which v λ(x n ,y n ) converges to \(\overline{v}^{\lambda}(x_{0},y_{0})\). At each of these points, choose a \(\frac{1}{n}\)-optimal control ν n in Θ λ (x n ,y n ). We denote by \((c_{t}^{n})\) and τ n the consumption process and first impulse time of ν n, respectively, and write \((X^{n}_{t},Y^{n}_{t}):= (X_{t}^{\nu^{n}, x_{n}}, Y_{t}^{\nu^{n},y_{n}})\) for the corresponding controlled process. Define the stopping times

$$H^{n} := \inf\{t\geq0: (X^n_t,Y^n_t)\notin N \}\wedge1 + \infty1_{\{ \theta^n=\tau^n\}} $$

and

$$\theta^n := H^n\wedge\tau^n. $$

We can further decompose \(H^{n} = \underline{H}^{n}\wedge\overline{H}^{n} \wedge1\), where

$$\underline{H}^n:= \inf\{t\geq0:\ (X^n_t,Y^n_t)\in\partial N\cap\{x_0 - \rho\}\times \mathbb{R}^d \} + \infty1_{\{\theta^n=\tau^n\}} $$

and

$$\overline{H}^n:= \inf\{t\geq0:\ (X^n_t,Y^n_t)\in\partial N\cap\{x: x > x_0-\rho\}\times \mathbb{R}^d \}+ \infty1_{\{\theta^n=\tau^n\}}. $$

Then there exists δ>0 such that \(\mathbb{E}[\overline{H}^{n}] > \delta\) for all n sufficiently large.

Step 3. Write

$$h(c,x,y) := I(c,x,y) - \sup_{\hat{c} > 0}I(\hat{c},x,y), $$

where

Note that I(c,x,y)<0 for all \(c\in \mathbb{R}_{+}\) and (x,y)∈N by (B.8). If we now set c (x,y)=(U′)−1(φ x (x,y)), then it follows that

$$h(c,x,y)= I(c,x,y) - I\big( c^*(x,y),x,y\big)\leq0. $$

By smoothness of φ and c and compactness of N, there exists a constant L ρ >0 with |I(c (x,y),x,y)|≤L ρ for all (x,y)∈N. On the other hand, there is α>0 such that I(c,x,y)≤−αc for all c>0. This leads to the upper bound

$$ h(c,x,y) \leq(-\alpha c + L_{\rho})\wedge0\qquad\mbox{for all } c>0, (x,y)\in N. $$
(B.9)

Since we only consider times t up to θ n, we can assume without loss of generality that \(c^{n}_{t} = c^{*}(X^{n}_{t},Y^{n}_{t})\) for t∈(θ n,H n]. Together with (B.9), we obtain

$$\begin{aligned} \mathbb{E}\bigg[ \int_0^{\theta^n}-e^{-\beta t}h(c_t,X_t,Y_t)\,dt\bigg] =& \mathbb{E}\bigg[\int_0^{H^n}-e^{-\beta t}h(c_t,X_t,Y_t)\,dt\bigg]\\ \geq& C\alpha \mathbb{E}\bigg[\int_0^{H^n} e^{-rt}c_t\,dt\bigg] - L_{\rho} \mathbb{E}[ H^n]. \\ \geq& C\alpha \mathbb{E}\bigg[\int_0^{\underline{H}^n\wedge1} e^{-rt}c_t 1_{\{\theta^n = \underline{H}^n\}}\,dt\bigg] - L_{\rho} \mathbb{E}[H^n], \end{aligned}$$
(B.10)

where the first inequality uses (B.9) to change the discount factor.

Step 4. Set \(\zeta^{n}_{t}:= (X^{n}_{t}, Y^{n}_{t})\). Weak dynamic programming in (B.1) implies

$$\begin{aligned} v^{\lambda}(x_n,y_n) \leq& \frac{1}{n} + \mathbb{E}\bigg[\int_0^{\theta^n} e^{-\beta t} U(c^n_t)\,dt + e^{-\beta\theta^n}\varphi(\zeta^n_t) \bigg]\\ \leq& \frac{1}{n} + \varphi(x_n,y_n) + \mathbb{E}\bigg[\int_0^{\theta^n} e^{-\beta t}I(c^n_t,\zeta^n_t)\,dt\bigg]\\ &{}+ \mathbb{E}\big[ e^{-\beta\theta^n} \big(\varphi(\zeta^n_{\theta ^n}) - \varphi(\zeta^n_{\theta^n-})\big)1_{\{\theta^n=\tau^n\}}\big]\\ \leq& \frac{1}{n} + \varphi(x_n,y_n) + \mathbb{E}\bigg[\int_0^{\theta^n} e^{-\beta t}I\big(c_t^*(\zeta^n_t),\zeta^n_t\big)\,dt\bigg] \\ &{} + \mathbb{E}\bigg[\int_0^{\theta^n} e^{-\beta t}h(c_t^n,\zeta ^n_t)\,dt\bigg] - C\eta \mathbb{P}[\theta^n = \tau^n]\\ \leq&\frac{1}{n} + \varphi(x_n,y_n) - CL_{\rho}\eta \mathbb{E}[\theta^n] - C\eta \mathbb{P}[\theta^n = \tau^n] \\ &{}+ \mathbb{E}\bigg[\int_0^{\theta^n} e^{-\beta t}h(c_t^n,\zeta ^n_t)\,dt\bigg]. \end{aligned}$$

Since \(v^{\lambda}(x_{n},y_{n}) - \varphi(x_{n},y_{n}) - \frac{1}{n}\to0\) as n→∞ and since the other terms on the right-hand side are negative, they must each vanish as n tends to infinity.

Step 5. We derive a contradiction using that

$$ \lim_{n\to\infty}\max\bigg(\mathbb{E}[\theta ^n], \mathbb{P}[\theta^n = \tau^n], \mathbb{E}\bigg[\int_0^{\theta^n} -e^{-\beta t}h(c_t^n,\zeta^n_t)\,dt\bigg]\bigg) = 0. $$
(B.11)

Observe that since the first two terms vanish, \(\mathbb{E}[{H}^{n}]\to0\) and \(\mathbb{P}[H^{n} = \theta^{n}]\to1\). Since \(\mathbb{E}[\overline{H}^{n}] > \delta\) for all n sufficiently large, we must therefore have \(\mathbb{E}[\underline {H}^{n}]\to 0\) and \(\mathbb{P}[\underline{H}^{n} = \theta^{n}]\to1\). As a consequence,

$$\mathbb{E}\bigg[\int_0^{\underline{H}^n}e^{-rt}c^n_t 1_{\{\theta^n = \underline {H}^n\}} \,dt\bigg]\to\rho, $$

which follows from the simple observation that for any fixed n, the term inside the expectation represents the amount of discounted consumption needed for cash in the bank account to decrease from x n to x 0ρ. However, by (B.11) and (B.10) we must have

$$0 = \lim_{n\to\infty} \mathbb{E}\bigg[\int_0^{\theta^n}-e^{-\beta t}h(c_t,X_t,Y_t)\,dt\bigg] \geq C\alpha\rho- L_{\rho}\lim_{n\to \infty} \mathbb{E}[ H^n] = C\alpha\rho> 0, $$

which is a contradiction. □

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Altarovici, A., Muhle-Karbe, J. & Soner, H.M. Asymptotics for fixed transaction costs. Finance Stoch 19, 363–414 (2015). https://doi.org/10.1007/s00780-015-0261-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00780-015-0261-3

Keywords

Mathematics Subject Classification

JEL Classification

Navigation