Skip to main content
Log in

Dynamic Investment and Counterparty Risk

  • Published:
Applied Mathematics & Optimization Submit manuscript

Abstract

We introduce a dynamic optimization framework in which collateral is used to mitigate losses arising at counterparty’s default. The investor faces two sources of risk: the default risk of the entity referencing the traded credit swap security, and counterparty risk generated from the default event of the trading counterparty. We show that the value function of the control problem coincides with the classical solution of a nonlinear dynamic programming equation. We provide an explicit characterization of the optimal investment strategy, and show that the investor does not trade if counterparty risk is sufficiently high. These findings suggest that moving credit swap trades into well-designed clearinghouses may stimulate economic activities.

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
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11

Similar content being viewed by others

References

  1. Assefa, S., Bielecki, T., Crépey, S., Jeanblanc, M.: CVA computation for counterparty risk assesment in credit portfolio. In: Bielecki, T., Brigo, D., Patras, F. (eds.) Credit Risk Frontiers: Subprime Crisis, Pricing and Hedging, CVA, MBS, Ratings and Liquidity. Wiley, Hoboken (2009)

    Google Scholar 

  2. Azizpour, S., Giesecke, K., Schwenkler, G.: Exploring the sources of default clustering. Working paper, Stanford University (2013)

  3. Belanger, A., Shreve, S., Wong, D.: A general framework for pricing credit risk. Math. Finance 14, 317–350 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  4. Bielecki, T., Jeanblanc, M., Rutkowski, M.: Pricing and trading credit default swaps in a hazard process model. Ann. Appl. Probab. 18, 2495–2529 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  5. Bo, L., Wang, Y., Yang, X.: An optimal portfolio problem in a defaultable market. Adv. Appl. Probab. 42, 689–705 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  6. Bo, L., Capponi A.: Optimal investment in credit derivatives portfolio under contagion risk. Math. Finance, 1–50 (2014)

  7. Brigo, D., Capponi, A., Pallavicini, A.: Arbitrage-free bilateral counterparty risk valuation under collateralization and application to credit default swaps. Math. Finance 24, 125–146 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  8. Capponi, A.: Pricing and mitigation of counterparty credit exposure. In: Fouque, J.P., Langsam, J. (eds.) Handbook of Systemic Risk. Cambridge University Press, Cambridge (2013)

    Google Scholar 

  9. Capponi, A., Figueroa-López, J.E.: Dynamic portfolio optimization with a defaultable security and regime-switching. Math. Finance 24, 207–249 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  10. Capponi, A., Figueroa-López, J.E., Pascucci A.: Power utility maximization in hidden regime-switching markets with default risk. Finance Stoch. (2014)

  11. Capponi, A., and C. Frei. 2015. Systemic influences on optimal equity-credit investment. Manage. Sci. Forthcoming

  12. Chow, Y., Teicher, H.: Probability Theory. Springer, New York (1978)

    Book  MATH  Google Scholar 

  13. El Karoui, N., Jeanblanc, M., Jiao, Y.: What happens after a default: the conditional density approach. Stoch. Process. Appl. 120, 1011–1032 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  14. Filipovic, D., Trolle, A.: The term structure of interbank risk. J. Financ. Econ. 109, 707–733 (2013)

    Article  Google Scholar 

  15. Jiao, Y., Kharroubi, I., Pham, H.: Optimal investment under multiple defaults risk: a BSDE-decomposition approach. Ann. Appl. Probab. 23, 455–491 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  16. Frey, R., Backhaus, J.: Pricing and hedging of portfolio credit derivatives with interacting default intensities. Int. J. Theor. Appl. Finance 11, 611–634 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  17. ISDA News Release. 2009. ISDA announces successful implementation of Big Bang? CDS protocol; determinations committees and auction settlement changes take effect. Available at http://www.isda.org/press/press040809.html

  18. “ISDA Credit Support Annex” 1992. “ISDA Guidelines for Collateral Practitioners”(1998), ISDA “Credit Support Protocol” (2002), ISDA “Close-Out Amount Protocol” (2009), ISDA “Margin Survey” (2010), “ISDA Market Review of OTC Derivative Bilateral Collateralization Practices”(2010). Available at www.isda.com

  19. ISDA Margin Survey 2014. International Swaps and Derivatives Association. Available at www.isda.com

  20. Jarrow, R., Yu, F.: Counterparty risk and the pricing of defaultable securities. J. Finance 56, 1765–1799 (2001)

    Article  Google Scholar 

  21. Kraft, H., Steffensen, M.: Portfolio problems stopping at first hitting time with application to defaut risk. Math. Methods Oper. Res. 63, 123–150 (2005)

    Article  MATH  Google Scholar 

  22. Lim, T., Quenez, M.C.: Exponential utility maximization in an incomplete market with defaults. Electron. J. Probab. 16, 1434–1464 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  23. Merton, R.: Optimum consumption and portfolio rules in a continuous-time model. J. Econ. Theory 3, 373–413 (1971)

    Article  MathSciNet  MATH  Google Scholar 

  24. Pham, H.: Stochastic control under progressive enlargment of filtrations and applications to multiple defaults risk management. Stoch. Process. Appl. 120, 1795–1820 (2010)

    Article  MATH  Google Scholar 

  25. Protter, P.: Stochastic Integrations and Differential Equations, 2nd edn. Springer, New York (2004)

    Google Scholar 

Download references

Acknowledgments

The authors would like to thank the two anonymous reviewers for their valuable and constructive comments and suggestions to improve the manuscript greatly. This research was partially supported by NCET-12-0914, NSF of China (No. 11471254) and Fundamental Research Funds for the Central Universities (No. WK3470000008).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Lijun Bo.

Appendices

Appendix 1: Price Function: Closed-Form Expressions and Comparisons

Lemma 7.1

The CDS price function \(\Gamma (t,\mathbf{z})\), \((t,\mathbf{z})\in [0,T_1]\times \mathcal{S}\), admits the following closed-form:

  • \(\Gamma (t,(1,0))=\Gamma (t,(1,1))=L\), for all \(t\in [0,T_1]\).

  • The price function \(\Gamma (t,(0,1))\) is given by

    $$\begin{aligned} \Gamma (t,(0,1)) = \int _t^{T_1} \left( Lh_{1}(s,(0,1))-\nu \right) e^{-\int _t^s(r+h_{1}(u,(0,1)))\mathrm {d}u}\mathrm {d}s,\ \ \ \ t\in [0,T_1].\nonumber \\ \end{aligned}$$
    (58)
  • The price function \(\Gamma (t,(0,0))\) is given by

    $$\begin{aligned} \Gamma (t,(0,0))&= \int _t^{T_1} \big (Lh_{1}(s,(0,0))-\nu + h_{C}(s,(0,0))\Gamma (s,(0,1))\big )\nonumber \\&\quad \times e^{-\int _t^s(r+h_{1}(u,(0,0))+h_{C}(u,(0,0)))\mathrm {d}u}\mathrm {d}s,\ \ \ t\in [0,T_1]. \end{aligned}$$
    (59)

    Here \(\Gamma (t,(0,1))\), \(t\in [0,T_1]\), is given by (58).

Proof

Let \((t,\mathbf{z})\in [0,T_1]\times \mathcal{S}\). We decompose \(\Gamma (t,\mathbf{z})\) as \(\Gamma (t,\mathbf{z}) = L\Gamma ^{(1)}(t,\mathbf{z}) - \nu \Gamma ^{(2)}(t,\mathbf{z})\), where

$$\begin{aligned} \Gamma ^{(1)}(t,\mathbf{z})&:= \mathbb {E} \left[ H_1(T_1)e^{-\int _t^{T_1} r(1-H_1(u))\mathrm {d}u}\Big | \mathbf{H}(t)=\mathbf{z}\right] ,\ \ \ \mathrm{and}\nonumber \\ \Gamma ^{(2)}(t,\mathbf{z})&:= \mathbb {E} \left[ \int _t^{T_1}e^{-\int _t^u r\mathrm {d}s}(1-H_1(u))\mathrm {d}u \Big | \mathbf{H}(t)=\mathbf{z}\right] . \end{aligned}$$

Using the Feymann-Kac’s formula, it follows that \(\Gamma ^{(1)}\) and \(\Gamma ^{(2)}\) satisfy the following equations. For \(\mathbf{z}=(z_1,z_C)\in \mathcal{S}\),

$$\begin{aligned} \left( \frac{\partial }{\partial t} + \mathcal{A}\right) \Gamma ^{(1)}(t,\mathbf{z}) = r(1-z_1)\Gamma ^{(1)}(t,\mathbf{z}),\ \ \ \ \ \Gamma ^{(1)}(T_1,\mathbf{z})=z_1, \end{aligned}$$
(60)

and

$$\begin{aligned} \left( \frac{\partial }{\partial t} + \mathcal{A}\right) \Gamma ^{(2)}(t,\mathbf{z}) + (1-z_1) = r\Gamma ^{(2)}(t,\mathbf{z}),\ \ \ \ \ \Gamma ^{(2)}(T_1,\mathbf{z})=0, \end{aligned}$$
(61)

where the operator \(\mathcal{A}\) is defined by (1). From the representation given in (5), we have \(\Gamma (t,(1,0))=\Gamma (t,(1,1))=L\). Using this, along with Eqs. (60) and (61), we have that \(\Gamma (t,(0,1))\) satisfies the equation:

$$\begin{aligned} \Gamma '(t,(0,1)) + h_{1}(t,(0,1))\left( L-\Gamma (t,(0,1))\right) - r \Gamma (t,(0,1)) -\nu =0 \end{aligned}$$
(62)

with terminal condition \(\Gamma (T_1,(0,1))=0\). Solving (62), we obtain (58). Using such an expression, along with Eqs. (60) and (61), we deduce that \(\Gamma (t,(0,0))\) satisfies

$$\begin{aligned}&\Gamma '(t,(0,0)) + h_{1}(t,(0,0))\left( L-\Gamma (t,(0,0))\right) \nonumber \\&\qquad + h_{C}(t,(0,0))\left( \Gamma (t,(0,1))-\Gamma (t,(0,0))\right) - r \Gamma (t,(0,0)) -\nu =0 \end{aligned}$$
(63)

with terminal condition \(\Gamma (T_1,(0,0))=0\). Solving it, we obtain (59). \(\square \)

Remark 7.2

Under assumptions (A1) and (A2), we have \(Lh_{1}(t,(0,1))-\nu>Lh_{1}(t,(0,0))-\nu >0\) for all \(t\in [0,T_1]\). Then, \(\Gamma (t,(0,1))>0\) for all \(t\in [0,T_1)\) and \(\Gamma (T_1,(0,1))=0\) from (58). Moreover, it follows from (59) that \(\Gamma (t,(0,0))>0\) for all \(t\in [0,T_1)\) and \(\Gamma (T_1,(0,0))=0\). Further, using (58), Under assumptions (A1) and (A2), for all \(t\in [0,T_1)\),

$$\begin{aligned} \Gamma (t,(0,1))\ge & {} \frac{L\inf _{u\in [0,T_1]}h_{1}(u,(0,1))-\nu }{r+\sup _{u\in [0,T_1]}{h}_{1}(u,(0,1))}\left[ 1-e^{-(r+(T_1-t)\sup _{u\in [0,T_1]}{h}_{1}(u,(0,1)))}\right] \\> & {} 0 \nonumber . \end{aligned}$$

Then, it holds that, for all \(t\in [0,T_1)\),

$$\begin{aligned} \Gamma (t,(0,0))&\ge \frac{L\inf _{u\in [0,T_1]}{h}_{1}(u,(0,0))-\nu }{r+\sup _{u\in [0,T_1]}{h}_{1}(u,(0,0))+\sup _{u\in [0,T_1]}{h}_{C}(u,(0,0))}\nonumber \\&\quad \times \left[ 1-e^{-(r+\sup _{u\in [0,T_1]}{h}_{1}(u,(0,0))+(T_1-t)\sup _{u\in [0,T_1]}{h}_{C}(u,(0,0)))}\right] >0 \nonumber . \end{aligned}$$

On the other hand, using (58), it follows that, for all \(t\in [0,T_1)\),

$$\begin{aligned} \Gamma (t,(0,1))&\le L\int _t^{T_1} (r+h_{1}(s,(0,1)))e^{-\int _t^s(r+h_{1}(u,(0,1)))\mathrm {d}u}\mathrm {d}s'\nonumber \\&=\left[ 1-e^{-\int _t^{T_1}(r+h_{1}(s,(0,1)))\mathrm {d}s}\right] L\nonumber \\&\le \left[ 1-e^{-(r+T_1\sup _{u\in [0,T_1]}h_{1}(u,(0,1)))}\right] L=:\delta L, \end{aligned}$$
(64)

where \(0<\delta < 1\).

Next, we compare price functions \(\Gamma (t,(0,0))\) and \(\Gamma (t,(0,1))\). Since the first is the price in the state where both CDS reference entity and counterparty are alive, while the second is the price in the state where the counterparty is defaulted, we expect that \(\Gamma (t,(0,1)) > \Gamma (t,(0,0))\). This is because after the counterparty defaults, the default risk of the reference entity increases due to contagion effects. The next lemma shows this formally.

Lemma 7.3

Under the assumption (A1), for all \(T\in (0,T_1)\), it holds that \(\Gamma (t,(0,1))>\Gamma (t,(0,0))\) for \(t\in [0,T]\).

Proof

Recall \(\Gamma (t,(0,1))\) and \(\Gamma (t,(0,0))\) satisfying Eqs. (62) and (63), respectively. Set \(E(t):=\Gamma (t,(0,1))-\Gamma (t,(0,0))\). Then we obtain the following ODE:

$$\begin{aligned}&E'(t) - \big (r + h_{1}(t,(0,0)) + h_{C}(t,(0,0))\big ) E(t) + (h_{1}(t,(0,1))-h_{1}(t,(0,0)))\\&\quad \times \big (L-\Gamma (t,(0,1))\big ) = 0 \end{aligned}$$

with terminal condition \(E(T_1)=\Gamma (T_1,(0,1))-\Gamma (T_1,(0,0))=0\). Solving it, we obtain the explicit solution:

$$\begin{aligned} E(t)= & {} \int _t^{T_1}(h_{1}(s,(0,1))-h_{1}(s,(0,0)))\\&\times \big (L-\Gamma (s,(0,1))\big )e^{-\int _t^{s}(r + h_{1}(u,(0,0)) + h_{C}(u,(0,0)))\mathrm {d}u}\mathrm {d}s. \end{aligned}$$

Since \(\Gamma (t,(0,1))<L\) for all \(t\in [0,T_1]\) and \(h_{1}(t,(0,1))-h_{1}(t,(0,0))>0\) for all \(t\in [0,T_1]\) by the assumption (A1), it implies that \(E(t)>0\) for all \(t\in [0,T]\). This completes the proof. \(\square \)

Appendix 2: Proofs Related to Section 3

Proof of Lemma 3.2

Notice that CDS reference entity and counterparty cannot default simultaneously in our model. By virtue of (4), we have

$$\begin{aligned} C_{\tau _C}&= (1-H_1(\tau _C))\Gamma \big (\tau _C,(H_1(\tau _C),H_C(\tau _C))\big )\\&= (1-H_1(\tau _C-))\Gamma \big (\tau _C,(H_1(\tau _C-),1)\big )\nonumber \\&= (1-H_1(\tau _C-))\Gamma \big (\tau _C-,(H_1(\tau _C-),1)\big )=\tilde{C}_{\tau _C-}, \end{aligned}$$

where the last step follows from the continuity of the price function \(t-\!\!\!\rightarrow \Gamma (t,(z_1,1))\) for each fixed default state \(z_1\in \{0,1\}\). This completes the proof of the lemma. \(\square \)

Proof of Lemma 3.3

Using the price dynamics of CDS given by (6), it follows from (17) that, for \(\psi \in \mathcal{U}_0\),

$$\begin{aligned} \frac{\mathrm {d}V_t^{\psi }}{V_{t-}^{\psi }}&= \bigg \{\pi _{B}(t)r + \pi _M(t)r_M + \psi (t)\Big [r\Gamma (t,\mathbf{H}(t)) - h_{1}(t,\mathbf{H}(t))\big (L-\Gamma (t,\mathbf{H}(t))\big )\nonumber \\&\qquad -\big (\Gamma (t,\mathbf{H}^C(t))-\Gamma (t,\mathbf{H}(t))\big )h_{C}(t,\mathbf{H}(t))\Big ]\bigg \}\mathrm {d}t\nonumber \\&\qquad + \psi (t)\big (L-\Gamma (t,\mathbf{H}(t-))\big )\mathrm {d}H_1(t)\\&\quad \quad + \psi (t)\big [\Gamma (t,\mathbf{H}^C(t-))-\Gamma (t,\mathbf{H}(t-))\big ]\mathrm {d}H_C(t)\nonumber \\&\qquad -\big \{\psi (t) \tilde{C}_{t-} - \alpha \psi ^+(t)C_{t-}\big \}^+\mathrm {d}H_C(t). \end{aligned}$$

Since, for \(\psi \in \mathcal{U}_0\), \(r\left[ \psi (t)\Gamma _{}(t,\mathbf{H}(t-)) + \pi _B(t) + \pi _M(t)\right] = r\), in terms of CDS price (4), we may rewrite the admissible control \(\pi _M(t)\) as \(\pi _M(t) = -\alpha \psi ^+(t)\Gamma (t,\mathbf{H}(t-))\), and hence

$$\begin{aligned} \pi _B(t) = 1 - \psi (t)\Gamma (t,\mathbf{H}(t-)) + \alpha \psi ^+(t)\Gamma (t,\mathbf{H}(t-)). \end{aligned}$$
(65)

Thus we obtain

$$\begin{aligned}&r\left[ \psi (t)\Gamma (t,\mathbf{H}(t-))+\pi _{B}(t)\right] + \pi _M(t)r_M&\\&= r + \pi _M(t)(r_M-r)= r- \psi ^+(t)\alpha \Gamma (t,\mathbf{H}(t-))(r_M-r), \end{aligned}$$

and hence yielding that the dynamics of the wealth is given by (18) using Lemma 3.2. \(\square \)

Appendix 3: Proofs Related to Section 4

Proof of Lemma 4.1

Recall (22). Notice that \(\Gamma (t,(0,0))>0\) by Remark 7.2, and Lemma 7.3 yields that

$$\begin{aligned}&{\Gamma }(t,(0,1))-\alpha \Gamma (t,(0,0))\\&\quad ={\Gamma }(t,(0,1))-\Gamma (t,(0,0)) + (1-\alpha )\Gamma (t,(0,0))>0,\ \ \ \forall \ t\in [0,T], \end{aligned}$$

since \(\alpha \in [0,1]\). Then, for \((\psi ,t)\in \mathbb {R}\times [0,T]\), it holds that

$$\begin{aligned}&\big \{\psi {\Gamma }(t,(0,1)) - \psi ^+\alpha \Gamma (t,(0,0))\big \}^+\nonumber \\&\quad =\left\{ \begin{array}{l@{\quad }l} 0, &{} \mathrm{if}\ \psi \le 0,\\ \psi \big ({\Gamma }(t,(0,1))-\alpha \Gamma (t,(0,0))\big ), &{} \mathrm{if}\ \psi >0 \end{array}\right. \nonumber \\&\quad = \psi ^+\big ({\Gamma }(t,(0,1))-\alpha \Gamma (t,(0,0))\big ). \end{aligned}$$
(66)

Substitute (66) into the second operator in (22) to conclude that

$$\begin{aligned}&\mathcal{H}(\psi ;t,B):=\gamma \bigg \{r -\psi ^+\alpha \Gamma (t,(0,0))(r_M-r)- \psi \Big [(L-\Gamma (t,(0,0)))h_1(t,(0,0))\nonumber \\&\qquad +(\Gamma (t,(0,1))-\Gamma (t,(0,0)))h_{C}(t,(0,0))\Big ]\bigg \}B\nonumber \\&\qquad + \Big \{\big [1+\psi (L-\Gamma (t,(0,0)))\big ]^{\gamma }B(t,(1,0)) - B(t,(0,0))\Big \}h_{1}^{ \mathbb {P} }(t,(0,0))\nonumber \\&\qquad + \bigg \{\Big [1+\psi \left( \Gamma (t,(0,1))-\Gamma (t,(0,0))\right) -\psi ^+\big ({\Gamma }(t,(0,1)) - \alpha \Gamma (t,(0,0))\big )\Big ]^{\gamma }\nonumber \\&\quad \quad \times B(t,(0,1))-B\bigg \}h_{C}^{ \mathbb {P} }(t,(0,0)). \end{aligned}$$
(67)

Then the decomposition (25) follow from (67) by considering \(\psi \le 0\) and \(\psi >0\) respectively. \(\square \)

Proof of Lemma 4.2

From (26) and (27), we have

$$\begin{aligned} {g}_1(\psi ;t,B)&= - \left[ (L-\Gamma (t,(0,0))) h_{1}(t,(0,0))+(\Gamma (t,(0,1))\right. \nonumber \\&\quad \left. -\Gamma (t,(0,0)))h_{C}(t,(0,0))\right] B+ \left[ 1+\psi (L-\Gamma (t,(0,0)))\right] ^{\gamma -1}\nonumber \\&\quad \times (L-\Gamma (t,(0,0)))h_{1}^{ \mathbb {P} }(t,(0,0))B(t,(1,0))\nonumber \\&\quad + \left[ 1+\psi \left( \Gamma (t,(0,1))-\Gamma (t,(0,0))\right) \right] ^{\gamma -1}\nonumber \\&\quad \times \left( \Gamma (t,(0,1))-\Gamma (t,(0,0))\right) h_{C}^{ \mathbb {P} }(t,(0,0))B(t,(0,1)),\nonumber \\ {g}_2(\psi ;t,B)&= - \left[ \alpha \Gamma (t,(0,0))(r_M-r) + (L-\Gamma (t,(0,0)))h_{1}(t,(0,0))\right. \nonumber \\&\quad \left. +(\Gamma (t,(0,1))-\Gamma (t,(0,0)))h_{C}(t,(0,0))\right] B\nonumber \\&\quad + \left[ 1+\psi (L-\Gamma (t,(0,0)))\right] ^{\gamma -1}(L-\Gamma (t,(0,0)))h_{1}^{ \mathbb {P} }(t,(0,0))B(t,(1,0))\nonumber \\&\quad - \left[ 1-\psi (1-\alpha )\Gamma (t,(0,0))\right] ^{\gamma -1}(1-\alpha )\Gamma (t,(0,0))h_{C}^{ \mathbb {P} }(t,(0,0))B(t,(0,1)). \end{aligned}$$
(68)

For fixed \((t,B)\in [0,T]\times \mathbb {R}_+\), we deduce from (68) that for all \(\psi >-\frac{1}{L-\Gamma (t,(0,0))}\), the function \(\psi -\!\!\!\rightarrow {g}_{1}(\psi ;t,B)\) is continuous and decreasing. Notice that the following limits hold:

$$\begin{aligned}&\lim _{\psi \downarrow -\frac{1}{L-\Gamma (t,(0,0))}}{g}_{1}(\psi ;t,B) =+\infty ,\nonumber \\&\lim _{\psi \uparrow +\infty } {g}_{1}(\psi ;t,B) = - \Big [(L-\Gamma (t,(0,0)))h_{1}(t,(0,0))+(\Gamma (t,(0,1))\nonumber \\&-\Gamma (t,(0,0)))h_{C}(t,(0,0))\Big ]B<0,\nonumber \end{aligned}$$

with the above inequality following from Lemma 7.3. Applying Intermediate Value Theorem we then obtain a unique finite solution \(\psi _{1}^{foc}>-\frac{1}{L-\Gamma (t,(0,0))}\) satisfying Eq. (28). Further, notice that the derivative \(\frac{\partial {g}_{1}(\psi ;t,B)}{\partial \psi }<0\) in the desired domain of \(\psi \). Then, in light of Kumagai (1980)’s implicit function theorem, we also have that \(\psi _{1}^{foc}\), viewed as a function of (tB) is \(C^1\) in (tB).

We next consider \({g}_{2}(\psi ;t,B)\). For all \(\psi \) satisfying \(-\frac{1}{L-\Gamma (t,(0,0))}<\psi <\frac{1}{(1-\alpha )\Gamma (t,(0,0))}\), we deduce from (68) that \(\psi -\!\!\!\rightarrow {g}_{2}(\psi ;t,B)\) is continuous and decreasing. Moreover, the following limits hold:

$$\begin{aligned} \lim _{\psi \downarrow -\frac{1}{L-\Gamma (t,(0,0))}} {g}_{2}(\psi ;t,B) =+\infty ,\ \ \mathrm{and}\ \ \lim _{\psi \uparrow \frac{1}{(1-\alpha )\Gamma (t,(0,0))}}{g}_{2}(\psi ;t,B) = -\infty . \end{aligned}$$

Applying Intermediate Value Theorem, we deduce that there exists a unique finite solution

$$\begin{aligned} \psi _2^{foc}\in \left( -\frac{1}{L-\Gamma (t,(0,0))},\frac{1}{(1-\alpha )\Gamma (t,(0,0))}\right) \end{aligned}$$

satisfying Eq. (28). Further, in light of Kumagai (1980)’s implicit function theorem, we also have that \(\psi _{2}^{foc}\) viewed as a function of (tB) is \(C^1\) in (tB), given that the derivative \(\frac{\partial {g}_{2}(\psi ;t,B)}{\partial \psi }<0\) in the desired domain of \(\psi \).

Next, we prove that \(\psi _{2}^{foc}(t,B)<\psi _{1}^{foc}(t,B)\). Notice that \(-\alpha \Gamma (t,(0,0))(r_M-r)\le 0\) since \(r_M \ge r\). Moreover, for \(\psi \le \frac{1}{(1-\alpha )\Gamma (t,(0,0))}\), it holds that

$$\begin{aligned} - \Big [1-\psi (1-\alpha )\Gamma (t,(0,0))\Big ]^{\gamma -1}(1-\alpha )\Gamma (t,(0,0))h_{C}^{ \mathbb {P} }(t,(0,0))B(t,(0,1))\le 0, \end{aligned}$$

since \(\alpha \in [0,1]\), where we set \(\frac{1}{0}=\infty \). Additionally, for all \(\psi >-\frac{1}{L-\Gamma (t,(0,0))}\) using Lemma 7.3 we obtain

$$\begin{aligned}&\Big [1+\psi \left( \Gamma (t,(0,1))-\Gamma (t,(0,0))\right) \Big ]^{\gamma -1}\\&\quad \times \left( \Gamma (t,(0,1))-\Gamma (t,(0,0))\right) h_{C}^{ \mathbb {P} }(t,(0,0))B(t,(0,1))>0. \end{aligned}$$

Comparing the expressions of \({g}_1(\psi ;t,B)\) and \({g}_2(\psi ;t,{B})\) defined in (68), and taking the above established inequalities into account, it follows that for all \(-\frac{1}{L-\Gamma (t,(0,0))}<\psi \le \frac{1}{(1-\alpha )\Gamma (t,(0,0))}\) and \((t,{B})\in [0,T]\times \mathbb {R}_+\), \({g}_2(\psi ;t,{B})<{g}_1(\psi ;t,{B})\). This implies that \(\psi _{2}^{foc}(t,{B})<\psi _{1}^{foc}(t,{B})\) for all \((t,{B})\in [0,T]\times \mathbb {R}_+\). If \(\alpha =1\), then \({g}_2(\psi ;t,{B})\) is reduced to

$$\begin{aligned} {g}_2(\psi ;t,{B})&= - \Big [\Gamma (t,(0,0))(r_M-r) + (L-\Gamma (t,(0,0)))h_{1}(t,(0,0))\\&\quad +(\Gamma (t,(0,1))-\Gamma (t,(0,0)))h_{C}(t,(0,0))\Big ]B\\&\quad + \big [1+\psi (L-\Gamma (t,(0,0)))\big ]^{\gamma -1}\nonumber \\&\quad \times (L-\Gamma (t,(0,0)))h_{1}^{ \mathbb {P} }(t,(0,0))B(t,(1,0)). \end{aligned}$$

Solving for \({g}_2(\psi _2^{foc};t,{B})=0\) yields (29). \(\square \)

Proof of Lemma 4.3

From (68), we deduce that \(\psi -\!\!\!\rightarrow {g}_1(\psi ;t,{B})\) is continuous and decreasing if \(\psi >-\frac{1}{L-\Gamma (t,(0,0))}\). Further, we may decompose \({g}_1(\psi ;t,{B})\) as \({g}_1(\psi ;t,{B}) = \tilde{g}_1(\psi ;t,{B}) + \hat{g}_1(\psi ;t,{B})\), where the functions in \((\psi ,t,{B})\in \mathbb {R}\times [0,T]\times \mathbb {R}_+\) are defined by

$$\begin{aligned} \tilde{g}_1(\psi ;t,{B})&:= - \left( L-\Gamma (t,(0,0))\right) h_{1}(t,(0,0)){B}+\left[ 1+\psi \left( L-\Gamma (t,(0,0))\right) \right] ^{\gamma -1}\nonumber \\&\quad \quad \times \left( L-\Gamma (t,(0,0))\right) h_{1}^{ \mathbb {P} }(t,(0,0))B(t,(1,0)),\ \ \ \ \mathrm{and}\nonumber \\ \hat{g}_1(\psi ;t,{B})&:=- \left( \Gamma (t,(0,1))-\Gamma (t,(0,0))\right) h_{C}(t,(0,0))B\nonumber \\&\qquad +\left[ 1+\psi \left( \Gamma (t,(0,1))-\Gamma (t,(0,0))\right) \right] ^{\gamma -1}\nonumber \\&\quad \quad \times \left( \Gamma (t,(0,1))-\Gamma (t,(0,0))\right) h_{C}^{ \mathbb {P} }(t,(0,0))B(t,(0,1)). \end{aligned}$$

It is easy to check that \(\tilde{g}_1(\tilde{\psi }_{1}(t,{B});t,{B})=\hat{g}_1(\hat{\psi }_{1}(t,{B});t,{B})=0\). Next, we prove that

$$\begin{aligned} {g}_{1}\left( \max \left\{ \tilde{\psi }_{1}(t,{B}),\hat{\psi }_{1}(t,{B})\right\} ;t,{B}\right) \le 0. \end{aligned}$$
(69)

First we consider the case where \(\tilde{\psi }_{1}(t,{B})\le \hat{\psi }_{1}(t,{B})\). In this case, \({g}_{1}(\hat{\psi }_{1}(t,{B});t,{B})=\tilde{g}_{1}(\hat{\psi }_{1}(t,{B});t,{B})\), since \(\hat{g}_{1}(\hat{\psi }_{1}(t,{B});t,{B})=0\). Notice that the function \(\psi -\!\!\!\rightarrow \tilde{g}_{1}(\psi ;t,{B})\) is continuous and decreasing on \(\psi >-\frac{1}{L-\Gamma (t,(0,0))}\). Then by the assumption that \(\tilde{\psi }_{1}(t,{B})\le \hat{\psi }_{1}(t,{B})\), it follows that \(\tilde{g}_{1}(\tilde{\psi }_{1}(t,{B});t,{B})\ge \tilde{g}_{1}(\hat{\psi }_{1}(t,{B});t,{B})\). Since \(\tilde{g}_{1}(\tilde{\psi }_{1}(t,{B});t,{B})=0\), we obtain that

$$\begin{aligned} {g}_{1}\left( \max \left\{ \tilde{\psi }_{1}(t,{B}),\hat{\psi }_{1}(t,{B})\right\} ;t,{B}\right) =\tilde{g}_{1}(\hat{\psi }_{1}(t,{B});t,{B})\le 0, \end{aligned}$$

yielding the inequality (69). Using a symmetric argument, we can establish the validity of (69) when \(\tilde{\psi }_{1}(t,{B})>\hat{\psi }_{1}(t,{B})\) and we omit the proof details here. Then the estimate (30) follows immediately from the inequality (69).

Proof of Lemma 4.7

A direct computation results in the limits (31) and (32). Using the implicit function theorem and Lemma 4.2, we have

since \(0<\gamma <1\). Similarly, it also holds that \(\frac{\partial \psi _{2}^{foc}(t,B)}{\partial B}<0\), where we used \(B(t,(1,0))=B(t,(0,1))=\frac{1}{\gamma }e^{\gamma r(T-t)}\). As for \(\psi _{2}^{foc}\), the expressions in (32) follow from the definition of \({g}_2(\psi ;t,{B})\) given in (68). This ends the proof of the lemma. \(\square \)

Appendix 3: Proofs Related to Section 5

Proof of Lemma 5.1

Using (64) and Lemma 7.3, we have \(0<\Gamma (t,(0,0))<\Gamma (t,(0,1))<\delta L\) for all \(t\in [0,T]\), where \(\delta \in (0,1)\). This yields \(0<\frac{1}{L-\Gamma (t,(0,0))}<\frac{1}{L-\Gamma (t,(0,1))}<\frac{1}{(1-\delta )L}\), for all \(t\in [0,T]\). Further using Theorem 4.4, the inequality (30) proved in Lemma 4.3, the estimates of \(\Gamma (t,(0,1))\) and \(\Gamma (t,(0,0))\) given in Remark 7.2 and that \(B(t,(1,1))=\frac{1}{\gamma }e^{\gamma r(T-t)}\), \(t\in [0,T]\), it follows that there exist constants \(K_1,\bar{K}_1>0\) so that

$$\begin{aligned} \left| \psi ^*(t,B)\right| \le K_1 + \bar{K}_1 B^{\frac{1}{\gamma -1}},\ \ \ \ \ \ (t,B)\in [0,T]\times \mathbb {R}_+. \end{aligned}$$
(70)

Using the inequality \(0<\Gamma (t,(0,0))<\Gamma (t,(0,0))< L\) for all \(t\in [0,T]\), we obtain that for all \((t,B)\in [0,T]\times \mathbb {R}_+\),

for some constants \(K,\bar{K}>0\) following from the estimate (70). Similarly, using the estimate (70), we deduce the existence of constants \(K_1,K,\bar{K}>0\) so that

$$\begin{aligned} {C}(t,\psi ^*(t,B))&\le \big [1+|\psi ^*(t,B)|(L-\Gamma (t,(0,0)))\big ]^{\gamma }h_{1}^{ \mathbb {P} }(t,(0,0))B(t,(1,1))\nonumber \\&\qquad +\big [1+|\psi ^*(t,B)|(\Gamma (t,(0,1))-\Gamma (t,(0,0)))\nonumber \\&\quad \quad +|\psi ^*(t,B)|(\Gamma (t,(0,1))-\alpha \Gamma (t,(0,0)))\big ]^{\gamma }\nonumber \\&\qquad \times h_{C}^{ \mathbb {P} }(t,(0,0))B(t,(1,1))\nonumber \\&\le \bigg \{\big [1+2L|\psi ^*(t,B)|\big ]^{\gamma }\sup _{t\in [0,T]}h_1^{ \mathbb {P} }(t,(0,0)) +\big [1+4L|\psi ^*(t,B)|\big ]^{\gamma }\nonumber \\&\quad \quad \times \sup _{t\in [0,T]}h_1^{ \mathbb {P} }(t,(0,0))\bigg \}B(0,(1,1))\nonumber \\&\le K_1 + K_1\left| \psi ^*(t,B)\right| ^{\gamma }\le K + \bar{K}B^{\frac{\gamma }{\gamma -1}}, \end{aligned}$$

where \(B(0,(1,1))=\frac{1}{\gamma }e^{\gamma r T}\), and we also used the inequality \((x+y)^{\gamma }\le x^{\gamma }+y^{\gamma }\) for all \(x,y\in \mathbb {R}_+\) and \(\gamma \in (0,1)\). This concludes the proof of the lemma. \(\square \)

Proof of Lemma 5.2

First we notice that \(\psi ^*(t,B)\) is the optimum. Then, we have, for all \((t,B)\in [0,T]\times \mathbb {R}_+\),

$$\begin{aligned} A(t,\psi ^*(t,B))B + C(t,\psi ^*(t,B))\ge A(t,0)B + C(t,0). \end{aligned}$$
(71)

Correspondingly, we introduce the following equation:

$$\begin{aligned} u'(t) + A(t,0)u(t) + C(t,0) =0,\ \ \ t\in [0,T) \end{aligned}$$
(72)

with terminal condition \(u(T)=a_0\in (0,\frac{1}{\gamma }]\). Recall \(A(t,\psi )\) and \(C(t,\psi )\) defined in (41). Since \(\psi -\!\!\!\rightarrow \psi ^+\) is Lipschitz-continuous, we have that \(\psi -\!\!\!\rightarrow A(t,\psi )\) is Lipschitz-continuous and \(\psi -\!\!\!\rightarrow C(t,\psi )\) is locally Lipschitz-continuous uniformly in \(t\in [0,T]\). By Lemma 4.9, we have that the continuous coefficients \(B-\!\!\!\rightarrow {A}(t,\psi ^*(t,B))\) and \(B-\!\!\!\rightarrow {C}(t,\psi ^*(t,B))\) are locally Lipschitz-continuous uniformly in \(t\in [0,T]\). Since B(t) is a solution to Eq. (40), then it must be \(C^1\) on [0, T], and hence it is bounded on [0, T]. Using the comparison theorem for first-order ODEs along with the inequality (71) and observing that \(u(T)=a_0\le \frac{1}{\gamma }=B(T)\), we have that \(B(t)\ge u(t)\) for all \(t\in [0,T]\).

We next prove the existence of a constant \(\eta >0\) so that \(u(t)\ge \eta \) for all \(t\in [0,T]\). First, we have

$$\begin{aligned} A(t,0)= & {} \gamma r -h_{1}^{ \mathbb {P} }(t,(0,0)) - h_{C}^{ \mathbb {P} }(t,(0,0)),\ \ \ \mathrm{and}\\ C(t,0)= & {} \big (h_{1}^{ \mathbb {P} }(t,(0,0))+h_{C}^{ \mathbb {P} }(t,(0,0))\big )B(t,(1,1)). \end{aligned}$$

Then, for all \(t\in [0,T]\), it holds that \(C(t,0)>0\), and

$$\begin{aligned} A(t,0) \ge \gamma r-\sup _{t\in [0,T]}h_{1}^{ \mathbb {P} }(t,(0,0)) - \sup _{t\in [0,T]}h_{C}^{ \mathbb {P} }(t,(0,0))=:\eta _{0}. \end{aligned}$$
(73)

Recall Eq. (72). We have its solution \((u(t);\ t\in [0,T])\) admits the following form, for all \(t\in [0,T]\),

$$\begin{aligned} u(t)&= a_0e^{\int _t^TA(s,0)\mathrm {d}s} + \int _t^T C(s,0)e^{\int _t^sA(\iota ,0)\mathrm {d}\iota }\mathrm {d}s \ge a_0e^{\int _t^TA(s,0)\mathrm {d}s}\ge a_0e^{\eta _0(T-t)}. \end{aligned}$$

In what follows, define the positive constant

$$\begin{aligned} \eta :=\left\{ \begin{array}{ll} a_0, &{}\quad \mathrm{if}\ \eta _{0}\ge 0,\\ a_0e^{\eta _{0}T}, &{}\quad \mathrm{if}\ \eta _{0}<0. \end{array}\right. \end{aligned}$$

Then, it holds that \(u(t)\ge \eta \) for all \(t\in [0,T]\). This implies that \(B(t)\ge u(t)\ge \eta \), for all \(t\in [0,T]\). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bo, L., Capponi, A. Dynamic Investment and Counterparty Risk. Appl Math Optim 77, 1–45 (2018). https://doi.org/10.1007/s00245-016-9364-2

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00245-016-9364-2

Keywords

Mathematics Subject Classifications

Navigation