Skip to main content
Log in

On managerial risk-taking incentives when compensation may be hedged against

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

Abstract

We consider a continuous time principal-agent model where the principal/firm compensates an agent/manager who controls the output’s exposure to risk and its expected return. Both the firm and the manager have exponential utility and can trade in a frictionless market. When the firm observes the manager’s choice of effort and volatility, there is an optimal contract that induces the manager to not hedge. In a two factor specification of the model where an index and a bond are traded, the optimal contract is linear in output and the log return of the index. We also consider a manager who receives exogenous share or option compensation and illustrate how risk taking depends on the relative size of the systematic and firm-specific risk premia of the output and index. Whilst in most cases, options induce greater risk taking than shares, we find that there are also situations under which the hedging manager may take less risk than the non-hedging manager.

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. Other papers considering the effect of specific contracts on portfolio managers, typically without possibility of hedging, include [2, 3, 8, 14], and [16]. Grinblatt and Titman [10], Henderson [13] and Hodder and Jackwerth [15] include hedging possibilities, again only in frameworks with specific contracts.

  2. Whilst [15] obtain numerical results showing that the manager prefers low total volatility when compensated with shares, and higher volatility when compensated with calls, they do not consider separately controlling systematic and firm-specific risks, and do not delineate any dependence on corresponding risk premia. [13] does consider separately systematic and firm-specific risk but only in the case where the output \(X\) is traded and \(\alpha _x = \lambda \).

  3. Panageas and Westerfield [23] show that even the risk-neutral managers need not behave aggressively when paid with high water mark contracts, if the time-horizon of their compensation is not fixed.

  4. Alternatively, we could assume that the firm and the manager enjoy utility from discounted values.

  5. These technical conditions are typically either satisfied, or they are satisfied if instead of \(\mathcal{Z}\) we consider its closure in an appropriate topology. In particular, the condition \(V(H_0)=\bar{V}(H_0)\) means that there is no duality gap between the primal problem of portfolio optimization and the dual problem of finding the optimal dual state-price density. Economically, it means that the standard marginal utility expression holds for the agent’s problem at the optimum: \(U'_A(P_T)= yY_T\) for the agent’s final total wealth \(P_T\), for some constant \(y\) and some SDF \(Y_T\) (or \(Y_T\) in the closure of the set \(\{ \mathcal Z \}\) of SDF’s). For details, see the references mentioned in the proof.

  6. If we modeled \(S\) as a Brownian motion with drift rather than a geometric Brownian motion, the optimal contract would be linear in \(S\), not in \(\log (S)\), in the case of CARA preferences.

  7. Alternatively, we could assume that the market trades a risk-free asset with the interest rate independent of all the risk-neutral densities in the market.

  8. The Excel spreadsheet is available at http://www.hss.caltech.edu/~cvitanic/PAPERS/Excelhedge.xls. In order to interpret this, write now

    $$\begin{aligned} \tilde{W}=\rho W +\sqrt{1-\rho ^2} M \end{aligned}$$
    (4.23)

    for a Brownian Motion \(M\) independent of \(W\). Note that we have

    $$\begin{aligned} dS/S=[r+\sigma \lambda ] dt +\sigma /\rho d\tilde{W}_t-\sigma \sqrt{1-\rho ^2}/\rho dM_t=rdt+\sigma /\rho dW^Q-\sigma \sqrt{1-\rho ^2}/\rho dM_t \end{aligned}$$
    (4.24)

    Thus, \(Q\) is a risk-neutral measure for \(S\). It is actually the projection on \(\tilde{F}\) of the measure which would be the risk-neutral measure if \(S\) was the only traded asset.

  9. [11] consider the case of non-CARA preferences with no hedging, and with linear contracts.

References

  1. Acharya, V., Bisin, A.: Managerial hedging, equity ownership, and firm value. RAND J. Econ. 40(1), 47–77 (2009)

    Article  Google Scholar 

  2. Basak, S., Pavlova, A., Shapiro, A.: Optimal asset allocation and risk shifting in money management. Rev. Financ. Stud. 20(5), 1583–1621 (2007)

    Article  Google Scholar 

  3. Basak, S., Shapiro, A., Tepla, L.: Risk management with benchmarking. Manag. Sci. 54, 542–557 (2006)

    Article  Google Scholar 

  4. Bettis, J.C., Bizjak, J.M.: Managerial ownership, incentive contracting, and the use of zero-cost collars and equity swaps by corporate insiders. J. Financ. Quant. Anal. 36, 345–370 (2001)

    Article  Google Scholar 

  5. Bisin, A., Gottardi, P.: Managerial hedging and portfolio monitoring. J. Eur. Econ. Assoc. 6(1), 158–209 (2008)

    Article  Google Scholar 

  6. Cadenillas, A., Cvitanić, Zapatero, F.: Optimal risk-sharing with effort and project choice. J. Econ. Theory 133, 403–440 (2007)

    Article  MATH  Google Scholar 

  7. Carpenter, J.N.: Does option compensation increase managerial risk appetite? J. Financ. 55, 2311–2331 (2000)

    Article  Google Scholar 

  8. Cuoco, D., Kaniel, R.: Equilibrium prices in the presence of delegated portfolio management. J. Financ. Econ. 101, 264–296 (2011)

    Article  Google Scholar 

  9. Garvey, G., Milbourn, T.: Incentive compensation when executives can hedge the market: evidence of relative performance evaluation in the cross section. J. Financ. 58, 1557–1581 (2003)

    Article  Google Scholar 

  10. Grinblatt, M., Titman, S.: Adverse risk incentives and the design of performance based contracts. Manag. Sci. 35(7), 807–822 (1989)

    Article  Google Scholar 

  11. Guo, M., Ou-Yang, H.: Incentives and performance in the presence of wealth effects and endogenous risk. J. Econ. Theory 129, 150–191 (2006)

    Article  MATH  MathSciNet  Google Scholar 

  12. Henderson, V.: Valuation of claims on non-traded assets using utility maximization. Math. Financ. 12(4), 351–373 (2002)

    Article  MATH  Google Scholar 

  13. Henderson, V.: The impact of the market portfolio on the valuation, incentives and optimality of executive stock options. Quant. Financ. 5(1), 1–13 (2005)

    Article  Google Scholar 

  14. Hodder, J.E., Jackwerth, J.C.: Incentive contracts and hedge fund management. J. Financ. Quant. Anal. 42(4), 811–826 (2007)

    Article  Google Scholar 

  15. Hodder, J.E., Jackwerth, J.C.: Managerial responses to incentives: control of firm-risk, derivative pricing implications, and outside wealth management. J. Bank. Financ. 35(6), 1507–1518 (2011)

    Article  Google Scholar 

  16. Hugonnier, J., Kaniel, R.: Mutual fund portfolio choice in the presence of dynamic flows. Math. Financ. 20(2), 187–227 (2010)

    Article  MATH  MathSciNet  Google Scholar 

  17. Holmstrom, B.: Moral hazard in teams. Bell J. Econ. 19, 324–340 (1982)

    Article  Google Scholar 

  18. Jin, L.: CEO compensation, diversification and incentives: theory and empirical results. J. Financ. Econ. 66, 29–63 (2002)

    Article  Google Scholar 

  19. Karatzas, I., Shreve, S.: Methods Math. Financ. Springer-Verlag, New York (1998)

    Book  Google Scholar 

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

    Article  MATH  MathSciNet  Google Scholar 

  21. Ofek, E., Yermack, D.: Taking stock: equity-based compensation and the evolution of managerial ownership. J. Financ. 55, 1367–1384 (2000)

    Article  Google Scholar 

  22. Ozerturk, S.: Financial innovations and managerial incentive contracting. Can. J. Econ. 39, 434–454 (2006)

    Article  Google Scholar 

  23. Panageas, S., Westerfield, M.M.: High-water marks: high risk appetites? J. Financ. 64(1), 1–36 (2009)

    Article  Google Scholar 

  24. Ross, S.A.: Compensation, incentives, and the duality of risk aversion and riskiness. J. Financ. 59, 207–225 (2004)

    Article  Google Scholar 

  25. Tehranchi, M.: Explicit solutions of some utility maximization problems in incomplete markets. Stoch. Process. Appl. 114, 109–125 (2004)

    Article  MATH  MathSciNet  Google Scholar 

Download references

Acknowledgments

J. Cvitanić, Research supported in part by NSF Grant DMS 10-08219. A. Lazrak, Research supported in part by the Social Sciences and Humanities Research Council (SSHRC) of Canada.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Ali Lazrak.

Appendix

Appendix

Proof of Proposition 2

The manager wants to maximize \(E[U_A(C_T\,-\,G_T\,+\,H_T)]\), which, under our assumptions, is the same as maximizing \(E[U_A(C_T\,-\,G_T\,+\,H_T\,-\,H_0B_T)]\). The standard martingale/duality approach to portfolio selection (see, e.g., [19] for diffusion models, or [20] for general semimartingale models) says that she will hedge so that

$$\begin{aligned} C_T-G_T+H_T-H_0B_T=I_A(yY_T) \end{aligned}$$
(6.1)

where \(y,Y_T\) solves, over constant numbers \(y\) and SDFs \(Y_T\in \mathcal{Z}\), the dual problem

$$\begin{aligned} \bar{V}(H_0):=\min _{y,Y}E[U_A(I_A(yY_T))-yY_TI_A(yY_T)+yY_T(C_T-G_T) +yY_T(H_T-H_0B_T)] \end{aligned}$$
(6.2)

if a solution exists, and if we have \(V(H_0)=\bar{V}(H_0)\). The last term in (6.2) disappears, because \(E[Y_T(H_T-H_0B_T)]=0\). Then, if (3.5) is satisfied, it is easily seen that the above is minimized for \(yY_T=zZ_T\). From (6.1) this implies \(H_T=H_0B_T\). Conversely, if the optimal hedging strategy results in \(H_T=H_0B_T\), we see from (6.1) that \(C_T\) has to be of the form (3.5). \(\square \)

Computations for Sect. 3.2: Consider the more general model where output is given by

$$\begin{aligned} dX_t=[\delta u_t + \alpha _x x_t +\alpha _y y_t]dt+ x_tdW_t+y_tdM_t \end{aligned}$$

and where the manager pays the cost \(G_T= \int _0^T g(u_t) dt\) with \(g(u) = \frac{u^2}{2k}\) for \(k>0\) when exerting the effort \(u\). Under optimal behavior of the firm, the manager is given a contract of the form \(C_T=aX_T+b\log (S_T) +c\). Denote by \(\tilde{c}\) the certainty equivalent (CE) of \(c\) and with \(\tilde{R}\) the CE of the manager’s reservation utility. Also denote by \(\pi _A\) the amount of capital manager invest in \(S\), and by \(\pi _P\) the analogous hedging amount for the firm. Given that at the optimum \(\pi _A,\pi _P\) and effort \(u\) are constant in our framework, the manager’s CE is then

$$\begin{aligned}&\tilde{c}+a[X_0+\alpha _y yT +\delta uT]+[\alpha _xax+ \lambda \sigma \pi _A-g(u)]T\nonumber \\&\quad +\,\,b\left[ \log S_0+\left( \mu -\frac{\sigma ^2}{2}\right) T\right] -\frac{\gamma _A T}{2}\left[ (ax+\pi _A\sigma +b\sigma )^2+a^2y^2\right] \end{aligned}$$
(6.3)

Maximizing over \(u,\pi _A\) and \(y\) we get

$$\begin{aligned}&\displaystyle ax+\sigma \pi _A+b\sigma =\frac{\lambda }{\gamma _A}\end{aligned}$$
(6.4)
$$\begin{aligned}&\displaystyle y= \frac{\alpha _y}{a\gamma _A}\end{aligned}$$
(6.5)
$$\begin{aligned}&\displaystyle g'(u)=a\delta \end{aligned}$$
(6.6)

This means that the manager’s CE is

$$\begin{aligned}&\tilde{R}=\tilde{c}+a\left[ X_0+\left( \frac{\alpha _y^2}{a\gamma _A} +\delta u\right) T\right] -g(u)T\nonumber \\&\quad +\,\,(\alpha _x-\lambda )ax+\lambda \left[ \frac{\lambda }{\gamma _A}-b\sigma \right] T+b[\log S_0+(\mu -\sigma ^2/2)T]\end{aligned}$$
(6.7)
$$\begin{aligned}&\quad -\,\,\frac{T}{2}\gamma _A\left[ \left( \frac{\lambda }{\gamma _A}\right) ^2 +\left( \frac{\alpha _y}{\gamma _A}\right) ^2\right] \end{aligned}$$
(6.8)

The firm’s CE is

$$\begin{aligned}&-\tilde{c}+(1-a)\left[ X_0+ \left( \frac{\alpha _y^2}{a\gamma _A} +\delta u\right) T\right] +[\alpha _x(1-a)x+\lambda \sigma \pi _P]T -b[\log S_0\nonumber \\&\quad +(\mu -\sigma ^2/2)T]\end{aligned}$$
(6.9)
$$\begin{aligned}&-\frac{T}{2}\gamma _P\left[ ((1-a)x+\pi _P\sigma -b\sigma )^2+(1-a)^2 \left( \frac{\alpha _y}{a\gamma _A}\right) ^2\right] \end{aligned}$$
(6.10)

Fixing \(\tilde{R}\) and computing \(\tilde{c}\) from the previous expression we get the firm’s CE as

$$\begin{aligned}&-\tilde{R}-g(u)T + (\alpha _x-\lambda )ax +\lambda \left[ \frac{\lambda }{\gamma _A}-b\sigma \right] T-\frac{T}{2}\gamma _A \left[ \left( \frac{\lambda }{\gamma _A}\right) ^2 +\left( \frac{\alpha _y}{\gamma _A}\right) ^2\right] \nonumber \\&\quad +\left[ X_0+ \left( \frac{\alpha _y^2}{a\gamma _A} +\delta u\right) T\right] \end{aligned}$$
(6.11)
$$\begin{aligned}&+[\alpha _x(1-a)x+\lambda \sigma \pi _P]T -\frac{T}{2}\gamma _P\left[ ((1-a)x+\pi _P\sigma -b\sigma )^2+(1-a)^2 \left( \frac{\alpha _y}{a\gamma _A}\right) ^2\right] \nonumber \\ \end{aligned}$$
(6.12)

This means that the hedging portfolio \(\pi _P\) is chosen so that

$$\begin{aligned} \gamma _P[(1-a)x+\pi _P\sigma -b\sigma ]=\lambda \end{aligned}$$
(6.13)

and hence the firm needs to maximize

$$\begin{aligned} \delta u-g(u)-\lambda b\sigma + \frac{\alpha _y^2}{a\gamma _A} +(\alpha _x-\lambda )x+\lambda \sigma b -\frac{\gamma _P}{2}(1-a)^2 (\frac{\alpha _y}{a\gamma _A})^2 \end{aligned}$$
(6.14)

We see that the principal is indifferent with respect to the choice of \(b\). In case there is no effort \(u\), \(u=g(u)=0\), we get

$$\begin{aligned} a=\frac{\gamma _P}{\gamma _A+\gamma _P} \end{aligned}$$
(6.15)

In case \(g(u)=u^2/(2k)\) is quadratic, taking into account that \(g'(u)=\delta a\), we can check that the derivative of the firm’s CE is

$$\begin{aligned} \frac{1}{a^3}[a^3(1-a)k\delta ^2-\frac{1}{\gamma _A}\alpha _y^2[a(1+\gamma _P/\gamma _A)-\gamma _P/\gamma _A] \end{aligned}$$

This is positive for \(a\) positive and close to zero and negative for \(a\) close to one, thus there is exactly one maximizer \(\hat{a}\in [0,1]\). Moreover, it is seen that increasing \(\alpha _y^2\) also increases the derivative in absolute value. Note also that the firm’s CE converges to \(-\infty \) at \(a=0\), and has a higher value at \(a=1\) for higher \(\alpha _y^2\). A combination of all these properties is possible only if the maximizer \(\hat{a}\) moves to higher values when \(\alpha _y^2\) is increased.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Cvitanić, J., Henderson, V. & Lazrak, A. On managerial risk-taking incentives when compensation may be hedged against. Math Finan Econ 8, 453–471 (2014). https://doi.org/10.1007/s11579-014-0123-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11579-014-0123-3

Keywords

JEL Classification

Navigation