Skip to main content
Log in

On the existence and uniqueness of the optimal central bank intervention policy in a forex market with jumps

  • Published:
Journal of the Operational Research Society

Abstract

We study a central bank intervention (CBI) problem in the foreign exchange market when the exchange rate follows a jump-diffusion process and show that the optimal CBI policy is a control-band policy. Our main contribution is a rigorous proof of the existence and uniqueness of the optimal CBI 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.

Similar content being viewed by others

Notes

  1. The number of domestic currency units per unit of foreign currency.

  2. An impulse control is a double sequence

    $$\nu = (\tau _{1},\tau _{2},\ldots ,\tau _{j},\ldots ;\zeta _{1},\zeta _{2},\ldots ,\zeta _{j},\ldots ),$$

    where \(\tau _{1}<\tau _{2}<\cdots \) are \({\mathcal {F}}_{t}\)-stopping times (the intervention times) and \(\zeta _{1},\zeta _{2},\ldots \) are the corresponding impulses at these times such that each \(\zeta _{j}\) is \({\mathcal {F}}_{\tau _{j}}\)-measurable.

  3. For the sake of completeness, we have shown the optimality of \(\{\hat{x},\bar{x}\}\) -policy in this section. However, it should be noted that this result is clearly subsumed in the general framework of Øksendal and Sulem (2007).

References

  • Avesani R and Gallo G (1991). Jumping in the Band: Undeclared Intervention Thresholds in a Target Zone. Working paper, Economics Department, University of Torento.

  • Baccarin S (2009). Optimal impulse control for a multidimensional cash management system with generalized cost functions. European Journal of Operational Research 196:198–206.

    Article  Google Scholar 

  • Bekaert G and Gray S (1998). Target zones and exchange rates: An empirical investigation. Journal of International Economics 45:1–35.

    Article  Google Scholar 

  • Benkherouf L and Bensoussan A (2009). Optimality of an (s,S) policy with compound poisson and diffusion demands: A quasi-variational inequalities approach. SIAM Journal on Control and Optimization 48(2):756–762.

    Article  Google Scholar 

  • Bensoussan A and Lions JL (1984). Impulse Control and Quasi-Variational Inequalities. Gauhiers-Villars: Paris.

    Google Scholar 

  • Bensoussan A, Long H, Perera S and Sethi S (2012). Impulse control with random reaction periods: A central bank intervention problem. Operations Research Letters 40(6):425–430.

    Article  Google Scholar 

  • Bertola G (1994). Continuous-time models of exchange rate and intervention. In: van der Ploeg F (ed). Handbook of International Macroeconomics. Basil Blackwell: London.

    Google Scholar 

  • Bertola G and Caballero R (1992). Target zones and realignments. The American Economic Review 82(3):520–536.

    Google Scholar 

  • Brekke KA and Øksendal B (1997). A verification theorem for combined stochastic control and impulse control. In: Decreasefond L et al (eds). Stochastic Analysis and Related Topics, Vol. 6, Springer Birkhauser: Geilo, Oslo.

  • Cadenillas A, Lakner P and Pinedo M (2010). Optimal control of a mean-reverting inventory. Operations Research 58(6):1697–1710.

    Article  Google Scholar 

  • Cadenillas A and Zapatero F (1999). Optimal central bank intervention in the foreign exchange market. Journal of Economic Theory 87(1):218–242.

    Article  Google Scholar 

  • Dominguez K and Frankel J (1993). Does Foreign Exchange Intervention Work? Institute for International Economics: Washington, DC.

    Google Scholar 

  • Flood R and Garber P (1991). The linkage between speculative attack and target zone models of exchange rates. Quarterly Journal of Economics 106(4):1367–1372.

    Article  Google Scholar 

  • Froot KA and Obstfeld M (1991). Exchange rate dynamics under stochastic regime shifts: A unified approach. Journal of International Economics 31(3):203–230.

    Article  Google Scholar 

  • Garber PM and Svensson L (1996). The operation and collapse of fixed exchange rate regime. In: Rogoff K and Grossman G (eds). Handbook of International Economics. North-Holland: Amsterdam.

    Google Scholar 

  • Jeanblanc-Picqué M (1993). Impulse control method and exchange rate. Mathematical Finance 3(2):161–177.

    Article  Google Scholar 

  • Jorion P (1988). On jump processes in the foreign exchange and stock markets. Review of Financial Studies 1(4):427–445.

    Article  Google Scholar 

  • Korn R (1997). Optimal impulse control when control actions have random consequences. Mathematics of Operations Research 22(3):639–667.

    Article  Google Scholar 

  • Korn R (1999). Some applications of impulse control in mathematical finance. Mathematical Methods of Operations Research 50(3):493–518.

    Article  Google Scholar 

  • Krugman P (1987). Trigger Strategies and Price Dynamics in Equity and Foreign Exchange Markets. NBER working paper # 2459.

  • Krugman P (1991). Target zones and exchange rate dynamics. Quarterly Journal of Economics 106(3):669–682.

    Article  Google Scholar 

  • Øksendal B and Sulem A (2007). Applied Stochastic Control of Jump Diffusions. Springer: Berlin.

    Book  Google Scholar 

  • Lumley R and Zervos M (2001). A model for investments in the natural resource industry with switching costs. Mathematics of Operations Research 26(4):637–653.

    Article  Google Scholar 

  • Miller M and Weller P (1991). Currency bands, target zones and price flexibility. IMF Economic Review 38(1):184–215.

    Article  Google Scholar 

  • Mundaca G and Øksendal B (1998). Optimal stochastic intervention control with application to the exchange rate. Journal Mathematical Economics 29(2):225–243.

    Article  Google Scholar 

  • Ohnishi M and Tsujimura M (2006). An impulse control of a geometric Brownian motion with quadratic costs. European Journal of Operational Research 168(2):311–321.

    Article  Google Scholar 

  • Perera S, Buckley W and Long H (2016). Market-reaction-adjusted optimal central bank intervention policy in a forex market with jumps. Annals of Operations Research. doi:10.1007/s10479-016-2297-y.

    Google Scholar 

  • Presman E and Sethi SP (2006). Inventory models with continuous and poisson demands and discounted and average costs. Production and Operations Management 15(2):279–293.

    Article  Google Scholar 

  • Svensson L (1991). The term structure of interest rate differentials in a target zone. Journal of Monetary Economics 28(1):87–116.

    Article  Google Scholar 

  • Yao D, Yang H and Wang R (2011). Optimal dividend and capital injection problem in the dual model with proportional and fixed transaction costs. European Journal of Operational Research 211(3):568–576.

    Article  Google Scholar 

Download references

Acknowledgements

The authors thank the editors, Tom Archibald and Jonathan Crook, and the anonymous associate editor for their contribution. Moreover, we are grateful to the reviewer for his/her thoughtful comments and suggestions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Winston Buckley.

Appendix: Further generalization of the problem

Appendix: Further generalization of the problem

In our original problem in Section 2, we are only allowed to give impulses \(\zeta \) with values in \({\mathcal {Z}}:=(0,\infty )\). However, in some applications, we may also have to bring the exchange rate upward by applying impulses with negative values. Therefore, we now assume that we are allowed to give impulses \(\zeta \) with values in \({\mathcal {Z}}:=(-\infty ,\infty )\). Again, if we apply an impulse control \(\nu = (\tau _{1},\tau _{2},\ldots ,\tau _{j},\ldots ;\zeta _{1},\zeta _{2},\ldots ,\zeta _{j},\ldots )\) to \(X_{x}(t)\), the resulting intervened process \(X_{x}^{(\nu )}(t)\) can be defined by \(X_{x}^{(\nu )}(t)=X_{x}(t)+\sum _{\tau _{k} \le t} \zeta _{k}\). The cost of intervention in this case is defined by \(K(\zeta )=C+\lambda |\zeta |\). Then, the expected total discounted cost associated with the given impulse control is

$$\begin{aligned} J^{(\nu )}(x)=E^{x}\left[ \int _{0}^{\infty } \hbox {e}^{-rt}(X_{x}^{(v)})^{2}\hbox {d}t+\sum _{k=1}^{N} \hbox {e}^{-r\tau _{k}}(C+\lambda |\zeta _{k}|)\right] . \end{aligned}$$

We again seek \(\varPhi ^{*}(x)\) and \(\nu ^{*}= (\tau _{1}^{*},\ldots ,\tau _{j}^{*},\ldots ;\zeta _{1}^{*},\ldots ,\zeta _{j}^{*},\ldots )\) such that

$$ \varPhi ^{*}(x)=\inf _{\nu }J^{(\nu )}(x) = J^{(\nu ^{*})}(x). $$

Here, we assume that the corresponding Lévy measure \(\eta \) is symmetric, i.e., \(\eta (H)=\eta (-H)\) for all \(H\subset {\mathbb {R}}\setminus \{0\}\). By symmetry, we expect the continuation region to be of the form \({\mathbb {C}}=(-\bar{x},\bar{x})\), for some \(\bar{x}>0\) to be determined. As soon as \(X_{x}(t)\) reaches level \(\bar{x}\) (or \(-\bar{x}\)), there is an intervention and \(X_{x}(t)\) is brought down (or up) to a certain value \(\hat{x}\) (or \(-\hat{x}\)) where \(-\bar{x}<-\hat{x}<0<\hat{x}<\bar{x}\). We determine \(\bar{x}\) and \(\hat{x}\) in the following way.

In the continuation region \({\mathbb {C}}\), we have \(A\varPhi +f=0\). Therefore we have

$$\begin{aligned} \mu \varPhi '(x)&+\frac{\sigma ^{2}}{2}\varPhi ''(x)-r\varPhi (x)\\&+\int _{{\mathbb {R}}}\{ \varPhi (x+\theta z)-\varPhi (x)-\theta z \varPhi '(x)\}\nu (\hbox {d}z)+x^{2}=0. \end{aligned}$$

Here, we try a solution of the form

$$ \varPhi (x)=C \cosh (\alpha x)+\dfrac{1}{r} x^{2}+\dfrac{1}{r^{2}}(2\mu x+b)+\dfrac{2\mu ^{2}}{r^{3}}, $$

where C is a constant (to be determined), \(b=\sigma ^{2}+\int _{{\mathbb {R}}}\theta ^{2} z^{2} \nu (\hbox {d}z)\), and \(\alpha >0\) is the positive solution of the equation \(F(\alpha )=-r+\mu r+\frac{\sigma ^{2}}{2}\alpha ^{2}+\int _{{\mathbb {R}}}\{\hbox {e}^{\alpha \theta z}-1-\alpha \theta z \}\nu (\hbox {d}z)\). Note that if we make no intervention at all, then the cost is

$$\begin{aligned} \phi (x)&= E^{x}\left[ \int _{0}^{\infty }\hbox {e}^{-rt}(X_{x}^{(\nu )}(t))^{2}\hbox {d}t\right] \\&=E^{x}\left[ \int _{0}^{\infty }\hbox {e}^{-rt}(x+\sigma B(t)+\int _{0}^{t}\int _{{\mathbb {R}}}\theta z \widetilde{N}(\hbox {d}s,\hbox {d}z))^{2}\hbox {d}t\right] \\&=\int _{0}^{\infty }\hbox {e}^{-rt}\left( x^{2}+\mu ^{2}t^{2}+\left( 2\mu x+b\right) t \right) \hbox {d}t\\&=\hbox {e}^{-rs}\left( \dfrac{x^{2}}{r}+\dfrac{b+2\mu x}{r^{2}}+\dfrac{2\mu ^{2}}{r^{3}}\right) . \end{aligned}$$

By the definition of \(\varPhi (x)\), we have \(0 \le \varPhi (x) \le \phi (x)\) for all \(-\bar{x}\le x \le \bar{x}\). Then, we get

$$\begin{aligned} 0\le C \cosh (\alpha x)+\dfrac{1}{r} x^{2}&+\dfrac{1}{r^{2}}(2\mu x+b)+\dfrac{2\mu ^{2}}{r^{3}}\\&\le \left( \dfrac{x^{2}}{r}+\dfrac{b+2\mu x}{r^{2}}+\dfrac{2\mu ^{2}}{r^{3}}\right) \end{aligned}$$

for all \(-\bar{x}\le x\le \bar{x}\). From the right-hand side of the above inequality, we get \(C \cosh (\alpha x)\le 0\) for all \(-\bar{x}\le x\le \bar{x}\). Hence, \(C<0\). We now let \(C=-a\), where \(a>0\), and define

$$ \varPhi _{0} (x)=\dfrac{x^{2}}{r} +\dfrac{b+2\mu x}{r^{2}}+\dfrac{2\mu ^{2}}{r^{3}}-a\cosh (\alpha x), $$

and let \(\varPhi (x):=\varPhi _{0} (x)\) for \(x\in {\mathbb {C}}\). The intervention operator in this case is

$$ {\mathcal {M}}\varPhi (x)=\inf \left\{ \varPhi (x+\zeta )+C+\lambda |\zeta | \; ; \; \zeta \in {\mathbb {R}}\right\} . $$

The first-order condition for a minimum \(\hat{\zeta }=\hat{\zeta }(x)\) of the function

$$\begin{aligned} G(x) = \left\{ \begin{array}{ll} \varPhi (x+\zeta )+C+\lambda \zeta , &{}\quad {\text {if }}\; \zeta >0,\\ \ \varPhi (x+\zeta )+C-\lambda \zeta , &{}\quad {\text {if }}\; \zeta <0, \end{array} \right. \end{aligned}$$

is the following

$$\begin{aligned}&(I) \;\zeta >0 ; \; \varPhi '(x_{\hat{\zeta }}+\hat{\zeta })+\lambda =0 \quad {\text {or}} \quad \varPhi '(x_{\hat{\zeta }}+\hat{\zeta })=-\lambda ,\\&(II) \;\zeta <0; \; \varPhi '(x_{\hat{\zeta }}+\hat{\zeta })-\lambda =0 \quad {\text {or}} \quad \varPhi '(x_{\hat{\zeta }}+\hat{\zeta })=\lambda . \end{aligned}$$

Therefore, we look for points \(\hat{x}\) and \(\bar{x}\) such that \(0<-\bar{x}<-\hat{x}<0<\hat{x}<\bar{x}\), \(\varPhi '(-\hat{x})=-\lambda \) and \(\varPhi '(\hat{x})=\lambda \). Note that, since \(\hat{x}<\bar{x}\) and \(-\hat{x}>-\bar{x}\), \(\varPhi '(\hat{x})=\varPhi '_{0}(\hat{x})\) and \(\varPhi '(-\hat{x})=\varPhi '_{0}(-\hat{x})\). We also have that, if \(\zeta >0\), then \(x+\hat{\zeta }=-\hat{x}\) or \(\hat{\zeta }=-\hat{x}-x=-(\hat{x}+x)\) and if \(\zeta <0\), then \(x+\hat{\zeta }=\hat{x}\) or \(\hat{\zeta }=\hat{x}-x\).

For \(x > \bar{x}\) or \(x < -\bar{x}\), we have \(\varPhi (x)={\mathcal {M}}\varPhi (x)\) since \(x\notin {\mathbb {C}}\). Therefore, if \(\zeta >0\) (i.e., \(x<-\bar{x}\)), then

$$\varPhi (x)=\varPhi (x+\hat{\zeta })+C+\lambda \hat{\zeta }= \varPhi (-\hat{x})+C-\lambda ( x+\hat{x}).$$

Equivalently, we have

$$ \varPhi (x)=\varPhi _{0}(-\hat{x})+C-\lambda (\hat{x}+x) \quad {\text {if}}\; x<-\bar{x}. $$

Similarly, if \(\zeta <0\) (i.e., \(x>\bar{x}\)), then

$$\varPhi (x)=\varPhi (x+\hat{\zeta })+C-\lambda \hat{\zeta }= \varPhi (\hat{x})+C+\lambda ( x-\hat{x}).$$

Equivalently, we have

$$ \varPhi (x)=\varPhi _{0}(\hat{x})+C+\lambda (x-\hat{x}) \quad {\text {if}}\; x>\bar{x}. $$

Consequently, we have

$$\begin{aligned} \varPhi (x) = \left\{ \begin{array}{ll} \varPhi _{0}(-\hat{x})+C-\lambda (x+\hat{x}); & \quad {\text{if}}\; x <-\bar{x}\\ \varPhi _{0}(x); &{}\quad {\text {if}}\; -\bar{x} \le x \le \bar{x}\\ \varPhi _{0}(\hat{x})+C+\lambda (x-\hat{x}); & \quad {\text{if}}\; x >\bar{x}. \end{array} \right. \end{aligned}$$

Differentiability at \(x=\bar{x}\) gives the equation \(\varPhi '_{0}(\bar{x})=\lambda \). Thus, we have

$$ \dfrac{2\bar{x}}{r}-a\alpha \sinh (\alpha \bar{x})=\lambda -\dfrac{2\mu }{r^{2}}. $$
(15)

We also have that \(\varPhi '(\hat{x})=\lambda \). That is \(\varPhi '_{0}(\hat{x})=\lambda \). Equivalently, we have

$$ \dfrac{2\hat{x}}{r}-a\alpha \sinh (\alpha \hat{x})=\lambda -\dfrac{2\mu }{r^{2}}. $$
(16)

Continuity at \(x=\bar{x}\) gives the equation

$$\varPhi _{0}(\hat{x})+C+\lambda (\bar{x}-\hat{x})=\varPhi _{0}(\bar{x}).$$

Substituting for \(\varPhi _{0}\) in the above equation, we get

$$\begin{aligned}&\dfrac{\hat{x}^{2}}{r} +\dfrac{b+2\mu \hat{x}}{r^{2}}+\dfrac{2\mu ^{2}}{r^{3}}-a\cosh (\alpha \hat{x})+C+\lambda (\bar{x}-\hat{x})\\&\quad =\dfrac{\bar{x}^{2}}{r} +\dfrac{b+2\mu \bar{x}}{r^{2}}+\dfrac{2\mu ^{2}}{r^{3}}-a\cosh (\alpha \bar{x}). \end{aligned}$$

Then, we have

$$ a \cosh (\alpha \bar{x})-a \cosh (\alpha \hat{x})+\frac{1}{r}(\hat{x}^{2}-\bar{x}^{2})+\lambda -\frac{2\mu }{r^{2}} (\bar{x}-\hat{x})+C=0. $$
(17)

Again, the solution to the CBI problem is uniquely determined by the system (15)–(17). We can prove the existence and uniqueness of the solution in this case by using a similar approach to Section 3.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Perera, S., Buckley, W. On the existence and uniqueness of the optimal central bank intervention policy in a forex market with jumps. J Oper Res Soc 68, 877–885 (2017). https://doi.org/10.1057/s41274-017-0208-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1057/s41274-017-0208-5

Keywords

Navigation