Skip to main content
Log in

Commodity Spread Option with Cointegration

  • Published:
Asia-Pacific Financial Markets Aims and scope Submit manuscript

Abstract

We derive the valuation formula of a European call option on the spread of two cointegrated commodity futures prices, based on the Gibson–Schwartz with cointegration (GSC) model. We also analyze the American commodity spread option including the early exercise premium representation and an analytical approximation valuation formulae with cointegration. In the numerical analysis, we compare the spread option values calculated by the GSC model and the Gibson–Schwartz (GS) model that ignores cointegration. Consistent with the intuition that the cointegration prevents the prices from diverging, the GSC model prices the commodity spread option lower than the GS model which have longer maturity of more than 6 years. In other words, the GS model may overprice the commodity spread options for those with longer maturity without taking account of cointegration. Thus, incorporating cointegration is important for valuation and hedging of long-term commodity spread options such as large scale oil refining plant developments.

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

Similar content being viewed by others

Notes

  1. Closely related to but independently of this paper, Casassus et al. (2013) develop a pricing model for commodity spread option with linear relations between commodity prices. To check the validity of their model, they conduct some Monte Carlo simulations to calculate prices of European commodity spread options. However, they do not provide analytical characterizations of the commodity spread option and of the effect of cointegration.

  2. See Gibson and Schwartz (1990), Bjerksund (1991), and Schwartz (1997) for derivation.

  3. We assume a filtered probability space \((\Omega , {\mathcal {F}}, \{{\mathcal {F}}_t\}_{t \ge 0}, P)\).

  4. Note also that since futures prices are equal to spot prices at maturity, the prices of European call options on futures spread are equal to those on spot spread when all maturities of the option and the underlying futures, i.e., \(T_0 = T_1 = T_2\).

  5. Notice that while the GS model only concerns two sets of commodity prices and convenience yields, the GSC model may incorporate n sets of commodity prices and convenience yields linear relations. Here we fixed the n to 2.

  6. Although we treat the case where there is only one linear relation between prices, i.e., the case with one-dimensional z(t), we can extend the model to include several linear relations.

  7. A crack spread option such as heating oil/crude oil and RBOB gasoline/crude oil, which are traded on the NYMEX, are both American types.

  8. \({\varvec{R}}_+\) is a set which only includes nonnegative real values \(x \ge 0\).

  9. See Broadie and Detemple (1997), Proposition 3.1 or Detemple (2006), Proposition 73.

  10. See Broadie and Detemple (1997), Proposition 2.5 or Detemple (2006), Proposition 69, for similar properties on max-option.

  11. See Broadie and Detemple (1997), Proposition 3.2 or Detemple (2006) Proposition 74 for the early exercise premium representation without the cointegration.

  12. The index numbers and therefore the parameters are set as different to Nakajima and Ohashi (2012).

  13. \(h_{1} = 0.42\) and \(h_{2} = 1\) are taken from the NYMEX crack spread between WTI crude oil and heating oil. Note that the payoff of this call spread option at maturity is \(Max[h_{1}G_{1} - h_{2}G_{2} - K, 0]\).

  14. Here, we denote the convenience yield in the GS model by \(\delta _{i}^{\textit{GS}}\) and that in the GSC model by \(\delta _{i}^{\textit{GSC}}\) to clarify that they are different.

  15. In this sense, we may interpret that the GSC model puts an additional structure on the commodity price model by assuming that a part of deviation of the drift of commodity i’s price from the risk-free rate r under the risk-neutral probability is explained by deviation of commodity prices from their long-term relation.

  16. For the GSC model to be cointegrated, we need \(b<0\). This is discussed in Nakajima and Ohashi (2012).

  17. See Eq. (8).

  18. See Nakajima and Ohashi (2012) for derivation.

  19. See for example Wiersema (2008) Section 4.6.

  20. Our model is defined under risk neutral measure. That is, the model have determined its measure to price derivatives, which rationalizes the above Eq. (16). In another point of view, while there are 4 Brownian motions and 2 spot commodities, there are enough number of futures (i.e., 2 futures) for each commodity, which can be used to construct risk neutral portfolio. Thus, we can use the above result provided by Karatzas (1988) under complete market.

  21. Here we define ||x|| to be the Euclidean norm.

  22. The proof is based on Broadie and Detemple (1997), Proposition 3.1, A.2, A.3, A.4, and A.6.

  23. The argument here is similar to Jaillet et al. (1990) and Broadie and Detemple (1997).

  24. If \(C^A\) is convex, then \(x^\top H x \le 0\) for all \(x \in {\varvec{R}}^2\) where H represents the Hessian of \(C^A\).

  25. The argument is based on the proof of Theorem 21, Lemma 29, Lemma 30, Theorem 34 in Detemple (2006).

  26. See Karatzas and Shreve (1991), Section 3.7 and especially equation (7.7) for Tanaka’s formula on continuous semimartingale process.

  27. See Karatzas and Shreve (1998), Theorem D.12 for the existence of the optimal stopping time.

  28. See the proof of Proposition 2.1 in Rutkowski (1994).

  29. See Karatzas and Shreve (1991), Section 3.7, Theorem 7.1 (ii).

  30. They have also analyzed the performance of their approach.

References

  • Barone-Adesi, G., & Whaley, R. E. (1987). Efficient analytic approximation of American option values. The Journal of Finance, 42(2), 301–320.

    Article  Google Scholar 

  • Bjerksund, P. (1991). Contingent claims evaluation when the convenience yield is stochastic: Analytical results. Working paper 1991:1, Norwegian School of Economics and Business Administration. Department of Finance and Management Science.

  • Bjerksund, P., & Stensland, G. (1994). An American call on the difference of two assets. International Review of Economics and Finance, 3(1), 1–26.

    Article  Google Scholar 

  • Broadie, M., & Detemple, J. (1997). The valuation of American options on multiple assets. Mathematical Finance, 7(3), 241–286.

    Article  Google Scholar 

  • Casassus, J., Liu, P., & Tang, K. (2013). Economic linkages, relative scarcity, and commodity futures returns. Review of Financial Studies, 26(5), 1324–1362.

    Article  Google Scholar 

  • Dempster, M. A. H., Medova, E., & Tang, K. (2008). Long term spread option valuation and hedging. Journal of Banking & Finance, 32(12), 2530–2540.

    Article  Google Scholar 

  • Detemple, J. (2006). American-style derivatives: Valuation and computation. Boca Raton, FL: CRC Press.

    Google Scholar 

  • Fezzi, C., & Bunn, D. W. (2009). Structural interactions of European carbon trading and energy prices. The Journal of Energy Market, 2(4), 53–69.

    Google Scholar 

  • Gibson, R., & Schwartz, E. S. (1990). Stochastic convenience yield and the pricing of oil contingent claims. The Journal of Finance, 45(3), 959–976.

    Article  Google Scholar 

  • Girma, P. B., & Paulson, A. S. (1999). Risk arbitrage opportunities in petroleum futures spreads. The Journal of Futures Markets, 19(8), 931–955.

    Article  Google Scholar 

  • Harrison, J. M., & Kreps, D. M. (1979). Martingales and arbitrage in multiperiod security markets. Journal of Economic Theory, 20(3), 381–408.

    Article  Google Scholar 

  • Harrison, J. M., & Pliska, S. R. (1981). Martingales and stochastic integrals in the theory of continuous trading. Stochastic Processes and Their Applications, 11, 215–260.

    Article  Google Scholar 

  • Jaillet, P., Lamberton, D., & Lapeyre, B. (1990). Variational Inequalities and the pricing of American options. Acta Applicandae Mathematicae, 21, 263–289.

    Article  Google Scholar 

  • Karatzas, I. (1988). On the pricing of American options. Applied Mathematics and Optimization, 17, 37–60.

    Article  Google Scholar 

  • Karatzas, I., & Shreve, S. E. (1991). Brownian motion and stochastic calculus. New York: Springer.

    Google Scholar 

  • Karatzas, I., & Shreve, S. E. (1998). Methods of mathematical finance. New York: Springer.

    Book  Google Scholar 

  • Malliaris, A. G., & Urrutia, J. L. (1996). Linkages between agricultural commodity futures contracts. The Journal of Futures Markets, 16(5), 595–609.

    Article  Google Scholar 

  • Nakajima, K., & Maeda, A. (2007). Pricing commodity spread options with stochastic term structure of convenience yields and interest rates. Asia-Pacific Financial Markets, 14(1–2), 157–184.

    Article  Google Scholar 

  • Nakajima, K., & Ohashi, K. (2012). Cointegrated commodity pricing model. Journal of Futures Markets, 32(11), 995–1033.

    Article  Google Scholar 

  • Pearson, N. D. (1995). An efficient approach for pricing spread options. The Journal of Derivatives, 3(1), 76–91.

    Article  Google Scholar 

  • Rutkowski, M. (1994). The early exercise premium representation of foreign market American options. Mathematical Finance, 4(4), 313–325.

    Article  Google Scholar 

  • Schwartz, E. S. (1997). The stochastic behavior of commodity prices: Implications for valuation and hedging. The Journal of Finance, 52(3), 923–973.

    Article  Google Scholar 

  • Shimko, D. C. (1994). Options on futures spreads: Hedging, speculation, and valuation. Journal of Futures Markets, 14(2), 183–213.

    Article  Google Scholar 

  • Wiersema, U. F. (2008). Brownian motion calculus. West Sussex: Wiley.

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Katsushi Nakajima.

Additional information

We are grateful to Fumio Hayashi, Toshiki Honda, Takashi Kanamura, Ryozo Miura, Hidetoshi Nakagawa, Nobuhiro Nakamura, Tatsuyoshi Okimoto, Marcel Prokopczuk, Hiroaki Yamauchi, Toshinao Yoshiba, and all seminar participants at Hitotsubashi University and 2013 JAFEE meeting for their many valuable comments and discussions. We also thank the editor and two anonymous referee for suggesting many ideas to enhance this paper. All remaining errors are our own. Financial support from the Japan Society for the Promotion of Science (KAKENHI Grant Number 26780190) and Ritsumeikan Asia Pacific University (the APU Academic Research Subsidy) are gratefully acknowledged.

Appendix

Appendix

1.1 Expectation and Covariances of Log Futures Return

In this subsection, we derive the futures price and expectation value and covariance of log futures return. We use the future price equation written in terms of spot price and derive the future price process using Ito’s lemma written in terms of futures price. This price process can be explicitly written in terms of futures price levels. Finally, we calculate the expectation and covariance of stochastic terms of futures price using properties of stochastic calculus.

Note that futures price in terms of spot price isFootnote 18

$$\begin{aligned} G_i(t, T) = e^{\mu _{X_i}(t, T) + \frac{\sigma _{X_i}^2(t, T)}{2}},\quad i = 1, 2 \end{aligned}$$

where

$$\begin{aligned} \mu _{X_i}(t, T) = \Bigg [ e^{T {\varvec{\beta }}} \Bigg \{ e^{-t {\varvec{\beta }}} {\varvec{X}}(t) + \int _t^T e^{- s {\varvec{\beta }}} {\varvec{\beta }}_0 (s) ds \Bigg \} \Bigg ]_i, \end{aligned}$$

and

$$\begin{aligned} \sigma _{X_i X_j} (t, T) = \Bigg [\int _t^T (e^{(T-s) {\varvec{\beta }}}) {\varvec{\Sigma }} (e^{(T-s) {\varvec{\beta }}})^\top ds \Bigg ]_{i j},\quad i, j = 1, 2. \end{aligned}$$

The partial derivatives are

$$\begin{aligned} \frac{\partial G_i(t, T)}{\partial S_j(t)}= & {} \frac{\left[ e^{(T-t) {\varvec{\beta }}}\right] _{i,j}}{S_j(t)} G_i(t, T),\quad i, j = 1, 2, \\ \frac{\partial G_i(t, T)}{\partial \delta _j(t)}= & {} \left[ e^{(T-t) {\varvec{\beta }}}\right] _{i,2+j} G_i(t, T), \quad i, j = 1, 2, \end{aligned}$$

where we denote \([A]_{i,j}\) as [ij]th entry of matrix A.

Since the futures price \(G_i(t, T)\) is a function of \(S_i(t), \delta _i(t)\) and twice differentiable, we can use the Ito’s lemma and the dynamics of future price is

$$\begin{aligned} dG_i(t, T)= & {} \sum _{k=1}^2 \sigma _{S_k} S_k(t) \frac{\partial G_i}{\partial S_k} dW_{S_k}(t) + \sum _{k=1}^2 \sigma _{\delta _k} \frac{\partial G_i}{\partial \delta _k} dW_{\delta _k}(t) \\= & {} G_i(t, T) \left[ \sum _{k=1}^2 \sigma _{S_k} \left[ e^{(T-t) {\varvec{\beta }}}\right] _{i,k} dW_{S_k}(t) \right. \\&\left. + \sum _{k=1}^2 \sigma _{\delta _k} \left[ e^{(T-t) {\varvec{\beta }}}\right] _{i,2+k} dW_{\delta _k}(t) \right] , \end{aligned}$$

where the drift term is 0 since \(G_i(t, T)\) is martingale under the risk-neutral probability.

Again, using Ito’s lemma we have,

$$\begin{aligned} d \log G_i(t, T)= & {} - \frac{1}{2} \Bigg \{ \sum _{k,l=1}^2 \sigma _{S_k S_l} \left[ e^{(T-t) {\varvec{\beta }}}\right] _{i,k} \left[ e^{(T-t) {\varvec{\beta }}}\right] _{i,l} \\&+\,2 \sum _{k,l=1}^2 \sigma _{S_k \delta _l} \left[ e^{(T-t) {\varvec{\beta }}}\right] _{i,k} \left[ e^{(T-t) {\varvec{\beta }}}\right] _{i,2+l}\\&+ \sum _{k,l=1}^2 \sigma _{\delta _k \delta _l} \left[ e^{(T-t) {\varvec{\beta }}}\right] _{i,2+k} \left[ e^{(T-t) {\varvec{\beta }}}\right] _{i,2+l} \Bigg \} dt \\&+ \sum _{k=1}^2 \sigma _{S_k} \left[ e^{(T-t) {\varvec{\beta }}}\right] _{i,k} dW_{S_k}(t) + \sum _{k=1}^2 \sigma _{\delta _k} \left[ e^{(T-t) {\varvec{\beta }}}\right] _{i,2+k} dW_{\delta _k}(t). \end{aligned}$$

The futures price can be expressed as follows.

$$\begin{aligned} G_i(T_0, T_i) = G_i(t, T_i) e^{X_{G_i}(t, T_0, T_i)}, \quad t \le T_0 \le T_i, \end{aligned}$$

where

$$\begin{aligned} X_{G_i}(t, T_0, T_i)\equiv & {} \mu _{X_{G_i}}(t, T_0, T_i) \\&+ \int _t^{T_0} \sum _{k=1}^2 \sigma _{S_k} \left[ e^{(T_i-t) {\varvec{\beta }}}\right] _{i,k} dW_{S_k}(u) \\&+ \int _t^{T_0} \sum _{k=1}^2 \sigma _{\delta _k} \left[ e^{(T_i-t) {\varvec{\beta }}}\right] _{i,2+k} dW_{\delta _k}(u). \\ \end{aligned}$$

The expectation value is

$$\begin{aligned} \mu _{X_{G_i}}(t, T_0, T_i)\equiv & {} E_t[X_{G_i}(t, T_0, T_i)] \nonumber \\= & {} - \frac{1}{2} \Bigg \{ \int _t^{T_0} \sum _{k,l=1}^2 \sigma _{S_k S_l} \left[ e^{(T_i-u) {\varvec{\beta }}}\right] _{i,k} \left[ e^{(T_i-u) {\varvec{\beta }}}\right] _{i,l} du \nonumber \\&+\,2\int _t^{T_0} \sum _{k,l=1}^2 \sigma _{S_k \delta _l} \left[ e^{(T_i-u) {\varvec{\beta }}}\right] _{i,k} \left[ e^{(T_i-u) {\varvec{\beta }}}\right] _{i,2+l} du \nonumber \\&+ \int _t^{T_0} \sum _{k,l=1}^2 \sigma _{\delta _k \delta _l} \left[ e^{(T_i-u) {\varvec{\beta }}}\right] _{i,2+k} \left[ e^{(T_i-u) {\varvec{\beta }}}\right] _{i,2+l} du \Bigg \}. \end{aligned}$$
(14)

The covariance of \(X_{G_i}(t, T_0, T_i)\) and \(X_{G_j}(t, T_0, T_j)\) is

$$\begin{aligned} \sigma _{X_{G_i} X_{G_j}}(t, T_0, T_i, T_j)\equiv & {} {\text{ cov }}_t [X_{G_i}(t, T_0, T_i), X_{G_i}(t, T_0, T_j)] \nonumber \\= & {} \int _t^{T_0} \sum _{k,l=1}^2 \sigma _{S_k S_l} \left[ e^{(T_i-u) {\varvec{\beta }}}\right] _{i,k} \left[ e^{(T_j-u) {\varvec{\beta }}}\right] _{j,l} du \nonumber \\&+ \int _t^{T_0} \sum _{k,l=1}^2 \sigma _{S_k \delta _l} \left[ e^{(T_i-u) {\varvec{\beta }}}\right] _{i,k} \left[ e^{(T_j-u) {\varvec{\beta }}}\right] _{j,2+l} du \nonumber \\&+ \int _t^{T_0} \sum _{k,l=1}^2 \sigma _{S_l \delta _k} \left[ e^{(T_j-u) {\varvec{\beta }}}\right] _{j,l} \left[ e^{(T_i-u) {\varvec{\beta }}}\right] _{i,2+k} du \nonumber \\&+ \int _t^{T_0} \sum _{k,l=1}^2 \sigma _{\delta _k \delta _l} \left[ e^{(T_i-u) {\varvec{\beta }}}\right] _{i,2+k} \left[ e^{(T_j-u) {\varvec{\beta }}}\right] _{j,2+l} du.\nonumber \\ \end{aligned}$$
(15)

1.2 Proof of Proposition 2.4

In this subsection, we prove Proposition 2.4. This is done by the following scheme. What we need to calculate is the expectation which can be expressed in terms of double integrals, since we are only dealing with the bivariate Gaussian processes. We know the expectation and covariance of the stochastic parts as we mentioned previously. The integrals can be calculated using multivariate version of completing squares, decomposing bivariate normal joint distribution in to conditional distribution and marginal distribution, and changing of variables. Finally, collecting all the terms, we have the pricing equation.

From Harrison and Kreps (1979) and Harrison and Pliska (1981), the price of commodity spread option at t, which option maturity is \(T_0\), futures maturity for \(G_1\) and \(G_2\) are \(T_1\) and \(T_2\), respectively, is

$$\begin{aligned} C^E (G_1(t, T_1), G_2(t, T_2), t) = e^{-r(T_0 - t)} E_t[ (h_1 G_1(T_0, T_1) - h_2 G_2(T_0, T_2) - K)^+ ]. \end{aligned}$$

The expectation value can be calculated as follows.

$$\begin{aligned}&E_t [ (h_1 G_1(T_0, T_1) - h_2 G_2(T_0, T_2) - K)^+ ] \\&\quad = \int _D (h_1 G_1(t, T_1) e^{x_1} - h_2 G_(t, T_2) e^{x_2} - K) n({\varvec{x}}| {\varvec{\mu }}_{{\varvec{x}}}, {\varvec{\Sigma }}_{{\varvec{x}}}) d{\varvec{x}}, \end{aligned}$$

where

$$\begin{aligned} {\varvec{x}}= & {} \left[ \begin{array}{c} x_1 \\ x_2 \end{array} \right] , \\ {\varvec{\mu }}_{{\varvec{x}}}= & {} \left[ \begin{array}{cc} \mu _{X_{G_1}}(t, T_0, T_1) \\ \mu _{X_{G_2}}(t, T_0, T_2) \end{array} \right] , \\ {\varvec{\Sigma }}_{{\varvec{x}}}= & {} \left[ \begin{array}{cc} \sigma _{X_{G_1}}^2(t, T_0, T_1) &{} \sigma _{X_{G_1} X_{G_2}}(t, T_0, T_1, T_2) \\ \sigma _{X_{G_1} X_{G_2}}(t, T_0, T_1, T_2) &{} \sigma _{X_{G_2}}^2(t, T_0, T_2) \end{array} \right] , \end{aligned}$$

and

$$\begin{aligned} D= & {} \{ {\varvec{x}} | h_1 G_1(t, T_1) e^{x_1} - h_2 G_2(t, T_2) e^{x_2} - K \ge 0 \} \\= & {} \{ {\varvec{x}} | d(x_2) \le x_1 \}, \\ d(x_2, K)\equiv & {} \ln (h_2 G_2(t, T_2) e^{x_2} + K) - \ln (h_1 G_1(t, T_1)). \end{aligned}$$

We now calculate the integrals. Suppose that \({\varvec{e}}_1\) is unit vector which the 1st entry is 1.

$$\begin{aligned}&\int _D e^{x_1} n({\varvec{x}} | {\varvec{\mu }}_{{\varvec{x}}}, {\varvec{\Sigma }}_{{\varvec{x}}}) d{\varvec{x}} \\&\quad = \int _D (2 \pi )^{-1} |{\varvec{\Sigma }}_{{\varvec{x}}}|^{-\frac{1}{2}} \exp \left\{ {\varvec{e}}_1^\top {\varvec{x}} - \frac{1}{2} ({\varvec{x}} - {\varvec{\mu }}_{{\varvec{x}}})^\top {\varvec{\Sigma }}_{{\varvec{x}}}^{-1} ({\varvec{x}} - {\varvec{\mu }}_{{\varvec{x}}}) \right\} d{\varvec{x}} \\&\quad = \int _D (2 \pi )^{-1} |{\varvec{\Sigma }}_{{\varvec{x}}}|^{-\frac{1}{2}} \exp \left\{ {\varvec{e}}_1^\top {\varvec{\mu }}_{{\varvec{x}}} + {\varvec{e}}_1^\top ({\varvec{x}} - {\varvec{\mu }}_{{\varvec{x}}}) - \frac{1}{2} ({\varvec{x}} - {\varvec{\mu }}_{{\varvec{x}}})^\top {\varvec{\Sigma }}_{{\varvec{x}}}^{-1} ({\varvec{x}} - {\varvec{\mu }}_{{\varvec{x}}}) \right\} d{\varvec{x}} \\&\quad = \int _D (2 \pi )^{-1} |{\varvec{\Sigma }}_{{\varvec{x}}}|^{-\frac{1}{2}} \exp \Biggl \{ {\varvec{e}}_1^\top {\varvec{\mu }}_{{\varvec{x}}} + \frac{1}{2} {\varvec{e}}_1^\top {\varvec{\Sigma }}_{{\varvec{x}}} {\varvec{e}}_1 \\&\qquad - \frac{1}{2} ({\varvec{x}} - {\varvec{\mu }}_{{\varvec{x}}} - {\varvec{\Sigma }}_{{\varvec{x}}} {\varvec{e}}_1)^\top {\varvec{\Sigma }}_{{\varvec{x}}}^{-1} ({\varvec{x}} - {\varvec{\mu }}_{{\varvec{x}}} - {\varvec{\Sigma }}_{{\varvec{x}}} {\varvec{e}}_1) \Biggr \} d{\varvec{x}} \\&\quad = \exp \Biggl \{ \mu _{X_{G_1}}(t, T_0, T_1) + \frac{1}{2} \sigma _{X_{G_1}}^2(t, T_0, T_1) \Biggr \} \\&\qquad \times \int _D (2 \pi )^{-1} |{\varvec{\Sigma }}_{{\varvec{x}}}|^{-\frac{1}{2}} \exp \{ - \frac{1}{2} ({\varvec{x}} - {\varvec{\mu }}_{{\varvec{x}}} - {\varvec{\Sigma }}_{{\varvec{x}}} {\varvec{e}}_1)^\top {\varvec{\Sigma }}_{{\varvec{x}}}^{-1} ({\varvec{x}} - {\varvec{\mu }}_{{\varvec{x}}} - {\varvec{\Sigma }}_{{\varvec{x}}} {\varvec{e}}_1) \} d{\varvec{x}}. \end{aligned}$$

The integral can be expanded as follows. We omit the time prameters for simplicity.

$$\begin{aligned}&\int _D (2 \pi )^{-1} |{\varvec{\Sigma }}_{{\varvec{x}}}|^{-\frac{1}{2}} \exp \Biggl \{ - \frac{1}{2} ({\varvec{x}} - {\varvec{\mu }}_{{\varvec{x}}} - {\varvec{\Sigma }}_{{\varvec{x}}} {\varvec{e}}_1)^\top {\varvec{\Sigma }}_{{\varvec{x}}}^{-1} ({\varvec{x}} - {\varvec{\mu }}_{{\varvec{x}}} - {\varvec{\Sigma }}_{{\varvec{x}}} {\varvec{e}}_1) \Biggr \} d{\varvec{x}} \\&\quad = \int _{-\infty }^{\infty } \int _{d(x_2, K)}^{\infty } (2 \pi )^{-1} (\sigma _{X_{G_1}} \sigma _{X_{G_2}} \sqrt{1 - \rho _{X_{G_1} X_{G_2}}^2})^{-1} \exp \Biggl \{ - \frac{1}{2 (1 - \rho _{X_{G_1} X_{G_2}})} \\&\qquad \times \,\Biggl ( \Biggl ( \frac{x_1 - \mu _{X_{G_1}} - \sigma _{X_{G_1}}^2}{\sigma _{X_{G_1}}} \Biggr )^2 - 2 \rho _{X_{G_1} X_{G_2}} \Biggl (\frac{x_1 - \mu _{X_{G_1}} - \sigma _{X_{G_1}}^2}{\sigma _{X_{G_1}}} \Biggr ) \\&\qquad \times \,\Biggl (\frac{x_2 - \mu _{X_{G_2}} - \sigma _{X_{G_1} X_{G_2}}}{\sigma _{X_{G_2}}} \Biggr ) + \Biggl ( \frac{x_2 - \mu _{X_{G_2}} - \sigma _{X_{G_1} X_{G_2}}}{\sigma _{X_{G_2}}} \Biggr )^2 \Biggr )^2 \Biggr \} dx_1 dx_2\\&\quad = \int _{-\infty }^{\infty } \int _{d(x_2)}^{\infty } (2 \pi (1 - \rho _{X_{G_1} X_{G_2}}^2))^{-\frac{1}{2}} \sigma _{X_{G_1}}^{-1} \\&\qquad \times \exp \left\{ - \frac{\Biggl (x_1 - \mu _{X_{G_1}} - \sigma _{X_{G_1}}^2 - \rho _{X_{G_1} X_{G_2}} \sigma _{X_{G_1}} \frac{x_2 - \mu _{X_{G_2}} - \sigma _{X_{G_1} X_{G_2}}}{\sigma _{X_{G_2}}} \Biggr )^2}{2 (1 - \rho _{X_{G_1} X_{G_2}}^2) \sigma _{X_{G_1}}^2} \right\} dx_1 \\&\qquad \times (2\pi )^{-\frac{1}{2}} \sigma _{X_{G_2}}^{-1} \exp \Biggl \{ - \frac{1}{2} \Biggl ( \frac{x_2 - \mu _{X_{G_2}} - \sigma _{X_{G_1} X_{G_2}}}{\sigma _{X_{G_2}}} \Biggr )^2 \Biggr \} dx_2 \\&\quad = \int _{-\infty }^{\infty } \int _{-d_1(x_2)}^{\infty } (2 \pi (1 - \rho _{X_{G_1} X_{G_2}}^2))^{-\frac{1}{2}} \sigma _{X_{G_1}}^{-1} \\&\qquad \times \exp \Biggl \{ - \frac{y^2}{2} \Biggr \} (1 - \rho _{X_{G_1} X_{G_2}}^2 )^{\frac{1}{2}} \sigma _{X_{G_1}} dy \\&\qquad \times (2\pi )^{-\frac{1}{2}} \sigma _{X_{G_2}}^{-1} \exp \Biggl \{ - \frac{1}{2} \Biggl ( \frac{x_2 - \mu _{X_{G_2}} - \sigma _{X_{G_1} X_{G_2}}}{\sigma _{X_{G_2}}} \Biggr )^2 \Biggr \} dx_2 \\&\quad = \int _{-\infty }^{\infty } \Phi (d_1(x_2)) n(x_1 | \mu _{X_{G_2}} + \sigma _{X_{G_1} X_{G_2}}, \sigma _{X_{G_2}}^2) dx_2, \end{aligned}$$

where

$$\begin{aligned} d_1(x_2) = - \frac{d(x_1) - \mu _{X_{G_1}} - \sigma _{X_{G_1}}^2 - \rho _{X_{G_1} X_{G_2}} \sigma _{X_{G_1}} \frac{x_2 - \mu _{X_{G_2}} - \sigma _{X_{G_1} X_{G_2}}}{\sigma _{X_{G_2}}}}{\sigma _{X_{G_1}} \sqrt{(1 - \rho _{X_{G_1} X_{G_2}}^2)}}, \end{aligned}$$

and we used change of variables in the third equation. Other integrals can be derived in the same manner. For the second integral,

$$\begin{aligned}&\int _D (2 \pi )^{-1} |{\varvec{\Sigma }}_{{\varvec{x}}}|^{-\frac{1}{2}} \exp \Biggl \{ - \frac{1}{2} ({\varvec{x}} - {\varvec{\mu }}_{{\varvec{x}}} - {\varvec{\Sigma }}_{{\varvec{x}}} {\varvec{e}}_2)^\top {\varvec{\Sigma }}_{{\varvec{x}}}^{-1} ({\varvec{x}} - {\varvec{\mu }}_{{\varvec{x}}} - {\varvec{\Sigma }}_{{\varvec{x}}} {\varvec{e}}_2) \Biggr \} d{\varvec{x}} \\&\quad = \int _{-\infty }^{\infty } \Phi (d_2(x_2)) n\left( x_2 | \mu _{X_{G_2}} + \sigma _{X_{G_2}}^2, \sigma _{X_{G_2}}^2\right) dx_2, \end{aligned}$$

where

$$\begin{aligned} d_2(x_2) = - \frac{d(x_1) - \mu _{X_{G_1}} - \sigma _{X_{G_1} X_{G_2}} - \rho _{X_{G_1} X_{G_2}} \sigma _{X_{G_1}} \frac{x_2 - \mu _{X_{G_2}} - \sigma _{X_{G_2}}^2}{\sigma _{X_{G_2}}}}{\sigma _{X_{G_1}} \sqrt{\left( 1 - \rho _{X_{G_1} X_{G_2}}^2\right) }}. \end{aligned}$$

And the last integral is

$$\begin{aligned}&\int _D (2 \pi )^{-1} |{\varvec{\Sigma }}_{{\varvec{x}}}|^{-\frac{1}{2}} \exp \Biggl \{ - \frac{1}{2} ({\varvec{x}} - {\varvec{\mu }}_{{\varvec{x}}})^\top {\varvec{\Sigma }}_{{\varvec{x}}}^{-1} ({\varvec{x}} - {\varvec{\mu }}_{{\varvec{x}}}) \Biggr \} d{\varvec{x}} \\&\quad = \int _{-\infty }^{\infty } \Phi (d_2(x_2)) n\left( x_2 | \mu _{X_{G_2}}, \sigma _{X_{G_2}}^2\right) dx_2. \end{aligned}$$

Collecting all terms, we have

$$\begin{aligned}&C^E (G_1(t, T_1), G_2(t, T_2), t) \\&\quad = h_1 G_1(t, T_1) \exp \{ -r(T_0 - t) \} \\&\qquad \times \int _{-\infty }^{\infty } \Phi (d_1(x_2)) n(x_2 | \mu _{X_{G_2}}(t, T_0, T_2) \\&\qquad +\,\sigma _{X_{G_1} X_{G_2}}(t, T_0, T_1, T_2), \sigma _{X_{G_2}}^2(t, T_0, T_2)) dx_2 - h_2 G_2(t, T_2) \exp \{ -r(T_0 - t) \} \\&\qquad \times \int _{-\infty }^{\infty } \Phi (d(x_2)) n(x_2 | \mu _{X_{G_2}}(t, T_0, T_2) \\&\qquad +\,\sigma _{X_{G_2}}^2(t, T_0, T_2), \sigma _{X_{G_2}}^2(t, T_0, T_2)) dx_2 \\&\qquad - K e^{-r(T_0 - t)} \int _{-\infty }^{\infty } \Phi (d(x_2)) n\left( x_2 | \mu _{X_{G_2}}(t, T_0, T_2), \sigma _{X_{G_2}}^2(t, T_0, T_2)\right) dx_2, \end{aligned}$$

where

$$\begin{aligned}&d(x_2) = - \frac{\ln (h_2 G_2(t, T_2) e^{x_2} + K) - \ln (h_1 G_1(t, T_1)) - \mu _{X_{G_1}}(t, t_0, T_1)}{\sigma _{X_{G_1}}(t, T_0, T_1) \sqrt{1 - \rho _{X_{G_1} X_{G_2}}^2 (t, T_0, T_1, T_2)}} \\&\qquad +\,\frac{\rho _{X_{G_1} X_{G_2}}(t, T_0, T_1, T_2) \sigma _{X_{G_1}}(t, T_0, T_1) \frac{x_2 - \mu _{X_{G_2}}(t, T_0, T_2)}{\sigma _{X_{G_2}}(t, T_0, T_2)}}{\sigma _{X_{G_1}}(t, T_0, T_1) \sqrt{1 - \rho _{X_{G_1} X_{G_2}}^2 (t, T_0, T_1, T_2)}}, \\&d_1(x_2) = d(x_2) + \sigma _{X_{G_1}}(t, T_0, T_1) \sqrt{1 - \rho _{X_{G_1} X_{G_2}}^2 (t, T_0, T_1, T_2)}, \\&\rho _{X_{G_1} X_{G_2}}(t, T_0, T_1, T_2) = \frac{\sigma _{X_{G_1} X_{G_2}}(t, T_0, T_1, T_2)}{\sigma _{X_{G_1}}(t, T_0, T_1) \sigma _{X_{G_2}}(t, T_0, T_2)}. \end{aligned}$$

1.3 Proofs for Characterization and Valuation Formulae of American Commodity Spread Option

In this section, we provide proofs for propositions in Sect. 2.3. As we have seen before,

$$\begin{aligned} d \ln G_i(t, T)= & {} - \frac{1}{2} \Bigg \{ \int _t^{T_0} \sum _{k,l=1}^2 \sigma _{S_k S_l} \left[ e^{(T_i-u) {\varvec{\beta }}}\right] _{i,k} \left[ e^{(T_i-u) {\varvec{\beta }}}\right] _{i,l} du \nonumber \\&\quad +\,2\int _t^{T_0} \sum _{k,l=1}^2 \sigma _{S_k \delta _l} \left[ e^{(T_i-u) {\varvec{\beta }}}\right] _{i,k} \left[ e^{(T_i-u) {\varvec{\beta }}}\right] _{i,2+l} du \nonumber \\&\quad + \int _t^{T_0} \sum _{k,l=1}^2 \sigma _{\delta _k \delta _l} \left[ e^{(T_i-u) {\varvec{\beta }}}\right] _{i,2+k} \left[ e^{(T_i-u) {\varvec{\beta }}}\right] _{i,2+l} du \Bigg \} \\&\quad + \Bigg \{ \sum _{k=1}^2 \sigma _{S_k} \left[ e^{(T-t) {\varvec{\beta }}}\right] _{i,k} dW_{S_k}(t) {+} \sum _{k=1}^2 \sigma _{\delta _k} \left[ e^{(T-t) {\varvec{\beta }}}\right] _{i,2+k} dW_{\delta _k}(t) \Bigg \}. \end{aligned}$$

Since the volatility term is only a linear combination of Brownian motion with time varying parameter, we can redefine this sde asFootnote 19

$$\begin{aligned} d\ln G_i(t, T)= & {} \mu _{G_i}(t) dt + \sigma _{G_i}(t) dW_{G_i}(t) \end{aligned}$$

with a new \(\sigma _{G_i}(t)\) and 1-dimensional Brownian motion \(dW_{G_i}(t)\). The covariance of \(\ln G_i(t,T)\) and \(\ln G_j(t,T)\) is

$$\begin{aligned} \sigma _{G_i,G_j}(t)= & {} {\varvec{z}}_i^{\top } {\varvec{\rho }} {\varvec{z}}_j\\ {\varvec{z}}_i= & {} \left[ \sigma _{S_1} \left[ e^{(T_i-t){\varvec{\beta }}} \right] _{i,1}, \sigma _{S_2} \left[ e^{(T_i-t){\varvec{\beta }}} \right] _{i,2}, \sigma _{\delta _1} \left[ e^{(T_i-t){\varvec{\beta }}} \right] _{i,3}, \sigma _{\delta _2} \left[ e^{(T_i-t){\varvec{\beta }}} \right] _{i,4} \right] ^{\top }\\ {\varvec{\rho }}= & {} \left[ \begin{array}{cccccc} 1 &{} \rho _{S_1 S_2} &{} \rho _{S_1} &{} \rho _{S_1 \delta _2}\\ \rho _{S_2 S_1} &{} 1 &{} \rho _{S_2 \delta _1} &{} \rho _{S_2 \delta _2}\\ \rho _{\delta _1 S_1} &{} \rho _{\delta _1 S_2} &{} 1 &{} \rho _{\delta _1 \delta _2}\\ \rho _{\delta _2 S_1} &{} \rho _{\delta _2 S_2} &{} \rho _{\delta _2 \delta _1} &{} 1 \end{array} \right] \end{aligned}$$

Suppose \(T_0 \le T_1, T_2\). From Karatzas (1988), we have \(C^A (G_1, G_2, t)\) which is the value of American call spread option. For \(t \le T_0\),

$$\begin{aligned} C^A (G_1, G_2, t) = \sup _{\tau \in \mathcal {S}_{t, T_0}} E_t [ e^{-r(\tau -t)} (h_1 G_1(\tau , T_1) - h_2 G_2(\tau , T_2) - K)^+ ], \end{aligned}$$
(16)

where \(\mathcal {S}_{t, T_0}\) is a set of stopping time taking values in \([t, T_0]\) almost surely.Footnote 20

Before showing some fundamental properties of the commodity spread options, we need the following lemma which is basically same as Lemma 3.9 in Jaillet et al. (1990) but is generalized for a function \(\psi \) for two variables.

Lemma 4.1

(Jaillet et al. 1990) Let X(t) be the 2-dimesional vector which satisfies the following stochastic differential equation.

$$\begin{aligned} X(T) = x + \int _t^T \mu (s) ds + \int _t^T \sigma (s) dW(s) \end{aligned}$$

Suppose \(\psi (x)\) is a function on \({\varvec{R}}^2\). Suppose there exists \(\lambda > 0\) such that \(\psi (x) exp(-\lambda ||x||)\) is a bounded, continuous function and admits a bounded weak derivative on \({\varvec{R}}^2\).Footnote 21 Define

$$\begin{aligned} C(x, a)\equiv & {} \sup _{\tau \in \mathcal {S}_{t,T}} E \left[ e^{-r(\tau -t)} \psi \left( x + \int _t^\tau \mu (s)ds + \int _t^\tau \sigma (s)dW(s) \right) \right] . \end{aligned}$$

Then

$$\begin{aligned} C(x, a)= & {} \sup _{\tau _0 \in \mathcal {S}_{0,1}} E \left[ e^{-r(T-t)\tau _0} \psi \left( x + \int _0^{\tau _0} \mu _0(t,a) da + \int _0^{\tau _0} \sigma _0(t,a) dW(a) \right) \right] \end{aligned}$$

with

$$\begin{aligned} \mu _0(t,a)= & {} (T-t)\mu (t+a(T-t)) \\ \sigma _0(t,a)= & {} \sqrt{T-t}\sigma (t+a(T-t)) \end{aligned}$$

Proof

The proof is similar to that of Jaillet et al. (1990), Lemma 3.9. \(\square \)

Notice that \(\psi (x_1, x_2) = (h_1 e^{x_1} - h_2 e^{x_2} - K)^+\) satisfies the condition of the lemma.

1.3.1 Proof of Proposition 2.5

Here we prove Proposition 2.5.Footnote 22

Proof

If \(h_1 G_1(t, T_1) \le h_2 G_2(t, T_2) + K\) then immediate exercise is suboptimal and thus we have the first assertion.

We prove the second assertion. Using Lemma 4.1, we have

$$\begin{aligned} C^A(G_1, G_2, t)= & {} \sup _{\tau _t \in \mathcal {S}_{0,1}} E\left[ e^{-r \tau _t (T-t)} (h_1 G_1 e^{a_1(t, \tau _t)} - h_2 G_2 e^{a_2(t, \tau _t)} - K)^+\right] \end{aligned}$$

where

$$\begin{aligned} a_i(t, \tau _t)= & {} \int _0^{\tau _t} \mu _{i}(t,a) da + \int _0^{\tau _t} \sigma _{i}(t,a) dW_{G_i}(a)\\ \mu _i(t,a)= & {} (T_0-t)\mu _{G_i}(t+a(T_0-t))\\ \sigma _i(t,a)= & {} \sqrt{T_0-t}\sigma _{G_i}(t+a(T_0-t)), i=1,2. \end{aligned}$$

Consider a new stopping time \(\tau _s = \tau _t (T-t)/(T-s)\) for \(s>t\). Observe that \(\tau _t \in \mathcal {S}_{0,1}\) if and only if \(\tau _s \in \mathcal {S}_{0,(T-t)/(T-s)}\) and \(\tau _s(T-s)=\tau _t(T-t)\). Now we have

$$\begin{aligned}&C^A(G_1, G_2, t)\\&\quad = \sup _{\tau _t \in \mathcal {S}_{0,1}} E\left[ e^{-r \tau _t (T-t)} (h_1 G_1 e^{a_1(t, \tau _t)} - h_2 G_2 e^{a_2(t, \tau _t)} - K)^+\right] \\&\quad = \sup _{\tau _s \in \mathcal {S}_{0,(T-t)/(T-s)}} E\left[ e^{-r \tau _s (T-s)} (h_1 G_1 e^{a_1(s, \tau _s)} - h_2 G_2 e^{a_2(s, \tau _s)} - K)^+\right] \\&\quad \ge \sup _{\tau _s \in \mathcal {S}_{0,1}} E\left[ e^{-r \tau _s (T-s)} (h_1 G_1 e^{a_1(s,\tau _s)} - h_2 G_2 e^{a_2(s,\tau _s)} - K)^+\right] \\&\quad = C^A(G_1, G_2, s) \end{aligned}$$

where the inequality follows from \(\mathcal {S}_{0,1} \subset \mathcal {S}_{0,(T-t)/(T-s)}\). Therefore, if \((G_1, G_2, t) \in \mathcal {E}\) which is equivalent to \(C^A(G_1, G_2, t) \le (h_1 G_1 - h_2 G_2 - K)^+\), then \((G_1, G_2, s) \in \mathcal {E}\).

We prove the third assertion. It is easy to see that

$$\begin{aligned} (h_1 \lambda G_1 - h_2 G_2 - K)^+= & {} (h_1 G_1 - h_2 G_2 - K + (\lambda - 1) h_1 G_1)^+\\\le & {} (h_1 G_1 - h_2 G_2 - K)^+ + (\lambda - 1) h_1 G_1. \end{aligned}$$

If \((G_1, G_2, t) \in \mathcal {E}\) which implies \(h_1 G_1 - h_2 G_2 - K > 0\) and \(\lambda > 1\), then

$$\begin{aligned} (h_1 \lambda G_1 - h_2 G_2 - K)^+= & {} h_1 \lambda G_1 - h_2 G_2 - K\\= & {} h_1 G_1 - h_2 G_2 - K + (\lambda - 1) h_1 G_1\\= & {} (h_1 G_1 - h_2 G_2 - K)^+ + (\lambda - 1) h_1 G_1. \end{aligned}$$

Using the above facts, if \((G_1, G_2, t) \in \mathcal {E}\) then

$$\begin{aligned}&C^A(\lambda G_1, G_2, t)\\&\quad = \sup _{\tau \in \mathcal {S}_{0,1}} E\left[ e^{-r \tau (T-t)} (h_1 \lambda G_1 e^{a_1(t, \tau )} - h_2 G_2 e^{a_2(t, \tau )} - K)^+\right] \\&\quad \le \sup _{\tau \in \mathcal {S}_{0,1}} E\left[ e^{-r \tau (T-t)} (h_1 G_1 e^{a_1(t, \tau )} - h_2 G_2 e^{a_2(t, \tau )} - K)^+ + (\lambda - 1) h_1 G_1 e^{a_1(t, \tau )}\right] \\&\quad \le C^A(G_1, G_2, t) + (\lambda - 1) h_1 G_1\\&\quad = (h_1 G_1 - h_2 G_2 - K)^+ + (\lambda - 1) h_1 G_1\\&\quad = (h_1 \lambda G_1 - h_2 G_2 - K)^+ \end{aligned}$$

Therefore \((\lambda G_1, G_2, t) \in \mathcal {E}\).

The proof of the fourth assertion proceeds as before. Since \(0 \le \lambda \le 1\),

$$\begin{aligned} (h_1 G_1 - h_2 \lambda G_2 - K)^+= & {} (h_1 G_1 - h_2 G_2 - K + (1 - \lambda ) h_1 G_1)^+\\\le & {} (h_1 G_1 - h_2 G_2 - K)^+ + (1 - \lambda ) h_2 G_2. \end{aligned}$$

If \((G_1, G_2, t) \in \mathcal {E}\) and \(0 \le \lambda \le 1\), then we have

$$\begin{aligned} (h_1 G_1 - h_2 \lambda G_2 - K)^+= & {} h_1 G_1 - h_2 \lambda G_2 - K\\= & {} h_1 G_1 - h_2 G_2 - K + (1 - \lambda ) h_2 G_2\\= & {} (h_1 G_1 - h_2 G_2 - K)^+ + (1 - \lambda ) h_2 G_2. \end{aligned}$$

With this inequality, if \((G_1, G_2, t) \in \mathcal {E}\) then

$$\begin{aligned}&C^A(G_1, \lambda G_2, t)\\&\quad = \sup _{\tau \in \mathcal {S}_{0,1}} E\left[ e^{-r \tau (T-t)} (h_1 G_1 e^{a_1(t, \tau )} - h_2 \lambda G_2 e^{a_2(t, \tau )} - K)^+\right] \\&\quad \le \sup _{\tau \in \mathcal {S}_{0,1}} E\left[ e^{-r \tau (T-t)} (h_1 G_1 e^{a_1(t, \tau )} - h_2 G_2 e^{a_2(t, \tau )} - K)^+ + (1 - \lambda ) h_2 G_2 e^{a_2(t, \tau )}\right] \\&\quad \le C^A(G_1, G_2, t) + (1 - \lambda ) h_2 G_2\\&\quad = (h_1 G_1 - h_2 G_2 - K)^+ + (1 - \lambda ) h_2 G_2\\&\quad = (h_1 G_1 - h_2 \lambda G_2 - K)^+ \end{aligned}$$

Thus, \((\lambda G_1, G_2, t) \in \mathcal {E}\).

If \(G_2(t, T_2)=0\) then \(G_2(u, T_2) = 0\) for all \(u \ge t\). This implies that the commodity spread option reduces to a plain option on a single asset \(G_1\), with exercise boundary \(B_1\) and we have the fifth assertion.

We prove the last assertion which follows from the following inequality.

$$\begin{aligned}&C^A(G(\lambda ), t)\\&\quad = \sup _{\tau \in \mathcal {S}_{0,1}} E_t\left[ e^{-r \tau (T-t)} h_1 \left( \lambda G_1 e^{a_1(t, \tau )} + (1-\lambda ) \tilde{G}_1 e^{a_1(t, \tau )} \right) \right. \\&\left. \qquad -\,h_2 \left( \lambda G_2 e^{a_2(t, \tau )} + (1-\lambda ) \tilde{G}_2 e^{a_2(t, \tau )} - K\right) ^+ \right] \\&\quad \le \sup _{\tau \in \mathcal {S}_{0,1}} E_t\left[ e^{-r \tau (T-t)} \lambda \left( h_1 G_1 e^{a_1(t, \tau )} - h_2 G_2 e^{a_2(t, \tau )} - K\right) ^+\right. \\&\left. \qquad +\,(1-\lambda ) \left( h_1 \tilde{G}_1 e^{a_1(t, \tau )} - h_2 \tilde{G}_2 e^{a_2(t, \tau )} - K\right) ^+ \right] \\&\quad \le \sup _{\tau \in \mathcal {S}_{0,1}} E_t\left[ e^{-r \tau (T-t)} \lambda \left( h_1 G_1 e^{a_1(t, \tau )} - h_2 G_2 e^{a_2(t, \tau )} - K\right) ^+\right] \\&\qquad + \sup _{\tau \in \mathcal {S}_{0,1}} E_t\left[ e^{-r \tau (T-t)} (1-\lambda ) \left( h_1 \tilde{G}_1 e^{a_1(t, \tau )} - h_2 \tilde{G}_2 e^{a_2(t, \tau )} - K\right) ^+ \right] \\&\quad = \lambda C^A(G, t) + (1-\lambda ) C^A(\tilde{G}, t) \end{aligned}$$

where we used the convexity of the payoff function \((h_1 G_1 - h_2 G_2 - K)^+\) in the first inequality. \(\square \)

1.3.2 Proof of Proposition 2.6

The proof for Proposition 2.6 is provided below.

Proof

The first assertion follows from the continuity of the payoff function \((h_1 G_1(\tau , T_1) - h_2 G_2(\tau , T_2) - K)^+\) and the continuity of the flow of the sdes. Since \((h_1 G_1 - h_2 G_2 - K)^+\) is nondecreasing in \(G_1\) and nonincreasing in \(G_2\) and applying Proposition 2.18 of Karatzas and Shreve (1991), the second assertion follows. Properties 3 and 4 are proved using the results of Proposition 2.5, 2 and 6, respectively. \(\square \)

1.3.3 Proof of Proposition 2.7

We prove the variation inequality characterization for spread option. The proof is based on Jaillet et al. (1990), Theorem 3.6 and Broadie and Detemple (1997), Proposition 2.6. However, we emphasize that since the model includes 2 commodities and the parameters depends on time, the proof is enhanced from those above.

Proof

First, let us prove the uniform boundedness of \(\partial C^A / \partial G_i, i = 1, 2\).Footnote 23 We focus on the derivative relative to \(G_1\). The argument for \(G_2\) is similar. Consider two asset values \((G_1, G_2, t)\) and \((\tilde{G}_1, G_2, t)\). Here we slightly abuse the notation and define

$$\begin{aligned} \ln \tilde{G}_1(s, T_1)= & {} \ln \tilde{G}_1(t, T_1) + \int _t^s \mu _{G_1}(u) du + \int _t^s \sigma _{G_1}(t) dW_{G_1}(t)\\\equiv & {} \ln \tilde{G}_1(t, T_1) + X_{G_1}(t, s, T_1) \end{aligned}$$

For any stopping time \(\tau \in \mathcal {S}_{t, T_0}\) we have

$$\begin{aligned}&|(h_1 G_1(\tau , T_1) - h_2 G_2(\tau , T_2) - K)^+ - (h_1 \tilde{G}_1(\tau , T_1) - h_2 G_2(\tau , T_2) - K)^+| \nonumber \\&\quad \le |(h_1 G_1(\tau , T_1) - h_2 G_2(\tau , T_2)) - (h_1 \tilde{G}_1(\tau , T_1) - h_2 G_2(\tau , T_2))| \nonumber \\&\quad = |h_1 G_1(\tau , T_1) - h_1 \tilde{G}_1(\tau , T_1)| \nonumber \\&\quad = |h_1 G_1(t, T_1) - h_1 \tilde{G}_1(t, T_1)| e^{X_{G_1}(t, \tau , T_1)} . \end{aligned}$$
(17)

Without loss of generality, suppose \(G_1(t, T_1) > \tilde{G}_1(t, T_1)\). Let \(\tau \in \mathcal {S}_{t, T_0}\) represent the optimal stopping time for \((G_1, G_2, t)\). We have

$$\begin{aligned}&|C^A(G_1, G_2, t) - C^A(\tilde{G}_1, G_2, t)|\\&\quad \le \left| E_t\left[ e^{-r(T_0-t)} \left( h_1 G_1(\tau , T_1) - h_2 G_2(\tau , T_2) - K\right) ^+\right] \right. \\&\qquad \left. - E_t\left[ e^{-r(T_0-t)} \left( h_1 \tilde{G}_1(\tau , T_1) - h_2 G_2(\tau , T_2) - K\right) ^+ \right] \right| \\&\quad \le e^{-r(T_0-t)} |h_1 G_1(t, T_1) - h_1 \tilde{G}_1(t, T_1)| E_t[ e^{X_{G_1}(t, \tau , T_1)} ]\\&\quad = e^{-r(T_0-t)} |h_1 G_1(t, T_1) - h_1 \tilde{G}_1(t, T_1)|\\&\quad \le |h_1 G_1(t, T_1) - h_1 \tilde{G}_1(t, T_1)|. \end{aligned}$$

where the first inequality is obtained by definition of the optimal stopping time, the second inequality is obtained by (17), and the first equality is by the expectation of log normal distribution. Hence

$$\begin{aligned} \frac{|C^A(G_1, G_2, t) - C^A(\tilde{G}_1, G_2, t)|}{|h_1 G_1(t, T_1) - h_1 \tilde{G}_1(t, T_1)|} \le 1. \end{aligned}$$

This implies Lipschitz continuous on \({\varvec{R}}_+\). Therefore there exist partial differentiation almost everywhere which is uniformly bounded.

Second, we prove the local boundedness of the time derivative. We have

$$\begin{aligned}&|C^A(G_1(t, T_1), G_2(t, T_2), t) - C^A(G_1(t, T_2), G_2(t, T_2), s)|\\&\quad = \left| E_t\left[ e^{-r \tau (T_0-t)} \left( h_1 G_1(t, T_1) e^{a_1(t)} - h_2 G_2(t, T_2) e^{a_2(t)} - K\right) ^+\right. \right. \\&\left. \left. \qquad -\,e^{-r \tau (T_0-s)} \left( h_1 G_1(t, T_1) e^{a_2(s)} - h_2 G_2(t, T_2) e^{a_2(s)} - K\right) ^+ \right] \right| \\&\quad \le E_t\left[ \left| e^{-r \tau (T_0-t)} - e^{-r \tau (T_0-s)} \right| \left( h_1 G_1(t, T_1) e^{a_1(t)} - h_2 G_2(t, T_2) e^{a_2(t)} - K\right) ^+ \right. \\&\qquad +\,e^{-r \tau (T_0-s)} \left| \left( h_1 G_1(t, T_1) e^{a_1(t)} - h_2 G_2(t, T_2) e^{a_2(t)} - K\right) ^+\right. \\&\left. \left. \qquad -\,\left( h_1 G_1(t, T_1) e^{a_1(s)} - h_2 G_2(t, T_2) e^{a_2(s)} - K\right) ^+\right| \right] . \end{aligned}$$

where we denoted

$$\begin{aligned} a_i(u)= & {} \int _0^{\tau } \mu _{i}(u,a) da + \int _0^{\tau } \sigma _{i}(u,a) dW_{G_i}(a)\\ \mu _i(t,a)= & {} (T_0-t)\mu _{G_i}(t+a(T_0-t))\\ \sigma _i(t,a)= & {} \sqrt{T_0-t}\sigma _{G_i}(t+a(T_0-t)),\quad i=1,2. \end{aligned}$$

and used Lemma 4.1 in the first equality. Since \(e^{x}\) is convex which implies \((e^{x})^\prime \) is monotonically increasing, we have

$$\begin{aligned} |e^{-r \tau (T-t)} - e^{-r \tau (T-s)}| \le \max _{u \in [0,1], v \in [0, T]} (r u e^{-ru(T-v)}) |t-s| \equiv M_1 |t-s| \end{aligned}$$
(18)

Furthermore,

$$\begin{aligned}&E_t\left[ \left( h_1 G_1(t, T_1) e^{a_1(t)} - h_2 G_2(t, T_2) e^{a_2(t)} - K\right) ^+\right] \nonumber \\&\quad \le E_t\left[ h_1 G_1(t, T_1) e^{a_1(t)} + h_2 G_2(t, T_2) e^{a_2(t)}\right] \nonumber \\&\quad = h_1 G_1(t, T_1) + h_2 G_2(t, T_2) \end{aligned}$$
(19)

We also have

$$\begin{aligned}&\left| \left( h_1 G_1(t, T_1) e^{a_1(t)} - h_2 G_2(t, T_2) e^{a_2(t)} - K\right) ^+ \right. \nonumber \\&\left. \quad - \left( h_1 G_1(t, T_1) e^{a_1(s)} - h_2 G_2(t, T_2) e^{a_2(s)} - K\right) ^+ \right| \nonumber \\&\quad \le \left| \left( h_1 G_1(t, T_1) e^{a_1(t)} - h_2 G_2(t, T_2) e^{a_2(t)}\right) \right. \nonumber \\&\left. \qquad - \left( h_1 G_1(t, T_1) e^{a_1(s)} - h_2 G_2(t, T_2) e^{a_2(s)}\right) \right| \nonumber \\&\quad \le \left| h_1 G_1(t, T_1) e^{a_1(t)} - h_1 G_1(t, T_1) e^{a_1(s)} \right| \nonumber \\&\qquad + \left| h_2 G_2(t, T_2) e^{a_2(t)} - h_2 G_2(t, T_2) e^{a_2(s)} \right| \nonumber \\&\quad \le h_1 G_1(t, T_1) e^{|a_1(t)| + |a_1(s)|} |a_1(t) - a_1(s) | \nonumber \\&\qquad + h_2 G_2(t, T_2) e^{|a_2(t)| + |a_2(s)|} |a_2(t) - a_2(s) | \nonumber \\&\quad \le \left( h_1 G_1(t, T_1) + h_2 G_2(t, T_2)\right) e^{|a_1(t)| + |a_1(s)| + |a_2(t)| + |a_2(s)|} \nonumber \\&\qquad \times (|a_1(t) - a_1(s) | + |a_2(t) - a_2(s) |) \end{aligned}$$
(20)

where the third inequality follows from the convexity of the exponential function. Moreover,

$$\begin{aligned}&\exp \left( |a_1(t)| + |a_1(s)| + |a_2(t)| + |a_2(s)|\right) \nonumber \\&\quad \le \exp \Bigg ( \int _0^{\tau _0} |\mu _1(t, a)| da + \left| \int _0^{\tau _0} \sigma _1(t, a) dW_{G_1}(a) \right| \nonumber \\&\quad \qquad + \int _0^{\tau _0} |\mu _1(s, a)| da + \left| \int _0^{\tau _0} \sigma _1(s, a) dW_{G_1}(a) \right| \nonumber \\&\quad \qquad + \int _0^{\tau _0} |\mu _2(t, a)| da + \left| \int _0^{\tau _0} \sigma _2(t, a) dW_{G_2}(a) \right| \nonumber \\&\quad \qquad + \int _0^{\tau _0} |\mu _2(s, a)| da + \left| \int _0^{\tau _0} \sigma _2(s, a) dW_{G_2}(a) \right| \Bigg ) \nonumber \\&\quad \le M_2 \exp \Bigg ( \left| \int _0^{\tau _0} \sigma _1(t, a) dW_{G_1}(a) \right| + \left| \int _0^{\tau _0} \sigma _1(s, a) dW_{G_1}(a) \right| \nonumber \\&\quad \qquad + \left| \int _0^{\tau _0} \sigma _2(t, a) dW_{G_2}(a) \right| + \left| \int _0^{\tau _0} \sigma _2(s, a) dW_{G_2}(a) \right| \Bigg ) . \end{aligned}$$
(21)

where we used the boundedness of \(\mu _i(u,a)\). Since \(\mu _i(u,a)\) have a bounded derivative, it is Lipschitz continuous which implies

$$\begin{aligned}&|a_1(t) - a_1(s) | + |a_2(t) - a_2(s)| \nonumber \\&\quad \le \left| \int _0^{\tau _0} \mu _1(t, a) - \mu _1(s, a) da \right| + \left| \int _0^{\tau _0} \sigma _1(t, a) - \sigma _1(s, a) dW_{G_1}(a) \right| \nonumber \\&\qquad + \left| \int _0^{\tau _0} \mu _2(t, a) - \mu _2(s, a) da \right| + \left| \int _0^{\tau _0} \sigma _2(t, a) - \sigma _2(s, a) dW_{G_2}(a) \right| \nonumber \\&\quad \le \int _0^{\tau _0} |\mu _1(t, a) - \mu _1(s, a)| da + \left| \int _0^{\tau _0} \sigma _1(t, a) - \sigma _1(s, a) dW_{G_1}(a) \right| \nonumber \\&\qquad + \int _0^{\tau _0} |\mu _2(t, a) - \mu _2(s, a)| da + \left| \int _0^{\tau _0} \sigma _2(t, a) - \sigma _2(s, a) dW_{G_2}(a) \right| \nonumber \\&\quad \le M_3 |t-s| + \left| \int _0^{\tau _0} \sigma _1(t, a) - \sigma _1(s, a) dW_{G_1}(a) \right| \nonumber \\&\qquad + \left| \int _0^{\tau _0} \sigma _2(t, a) - \sigma _2(s, a) dW_{G_2}(a) \right| . \end{aligned}$$
(22)

We use the following two inequalities.

$$\begin{aligned} E \Bigg [ \exp \Bigg ( A \left| \int _0^{\tau _0} \sigma _i(u, a) dW_{G_i}(a) \right| \Bigg ) \Bigg ]\le & {} E \Bigg [ \exp \Bigg ( A \left| \int _0^1 \sigma _i(u, a) dW_{G_i}(a) \right| \Bigg ) \Bigg ] \nonumber \\\le & {} 2 \exp \Bigg ( \frac{A^2 \int _0^1 \sigma _i^2(u, a) da}{2} \Bigg ) \nonumber \\\le & {} M_4. \end{aligned}$$
(23)

where we used the submartingale property of \(\exp \left( A \left| \int _s^t \sigma _i(u, a) dW_{G_i}(a) \right| \right) _{s \ge t}\). For the other inequality,

$$\begin{aligned} |\sigma _1(t, a) - \sigma _1(s, a)|= & {} \left| \frac{\partial \sigma _1(u,a)}{\partial t} (t-s) \right| \nonumber \\\le & {} \sup _{u \in [s,t], a \in [0,1]} \left| \frac{\partial \sigma _1(u,a)}{\partial t} \right| |t - s| = M_5 |t-s| \nonumber \\\Rightarrow & {} \int _0^{1} (\sigma _1(t, a) - \sigma _1(s, a))^2 da \le M_5^2 |t-s|^2. \end{aligned}$$

Therefore, we have

$$\begin{aligned} E \Bigg [ \left| \int _0^{\tau _0} \sigma _i(t, a) - \sigma _i(s, a) dW_{G_i}(a) \right| ^2 \Bigg ]= & {} E \Bigg [ \int _0^{\tau _0} (\sigma _i(t, a) - \sigma _i(s, a))^2 da \Bigg ] \nonumber \\\le & {} \int _0^{1} (\sigma _i(t, a) - \sigma _i(s, a))^2 da \nonumber \\\le & {} M_5^2 |t-s|^2. \end{aligned}$$
(24)

Using inequalities (18)–(24) and the Cauchy–Schwartz inequality, we have

$$\begin{aligned}&|C^A(G_1, G_2, t) - C^A(G_1, G_2, s)|\\&\quad \le (h_1 G_1(t, T_1) + h_2 G_2(t, T_2)) \times ( M_1 |t-s|\\&\qquad +\,E \Bigg [ M_2 \exp \Bigg ( \left| \int _0^{\tau _0} \sigma _1(t, a) dW_{G_1}(a) \right| + \left| \int _0^{\tau _0} \sigma _1(s, a) dW_{G_1}(a) \right| \\&\qquad + \left| \int _0^{\tau _0} \sigma _2(t, a) dW_{G_2}(a) \right| + \left| \int _0^{\tau _0} \sigma _2(s, a) dW_{G_2}(a) \right| \Bigg )\\&\qquad \times \Bigg ( M_3 |t-s| + \left| \int _0^{\tau _0} \sigma _1(t, a) - \sigma _1(s, a) dW_{G_1}(a) \right| \\&\qquad + \left| \int _0^{\tau _0} \sigma _2(t, a) - \sigma _2(s, a) dW_{G_2}(a) \right| \Bigg ) \Bigg ]\le M_6 |t-s|\\&\qquad +\,M_7 |t-s| E \Bigg [ \exp \Bigg ( \left| \int _0^{\tau _0} \sigma _1(t, a) dW_{G_1}(a) \right| + \left| \int _0^{\tau _0} \sigma _1(s, a) dW_{G_1}(a) \right| \\&\qquad + \left| \int _0^{\tau _0} \sigma _2(t, a) dW_{G_2}(a) \right| + \left| \int _0^{\tau _0} \sigma _2(s, a) dW_{G_2}(a) \right| \Bigg )\Bigg ]\\&\qquad +\,M_8 E \Bigg [ \exp \Bigg ( \left| \int _0^{\tau _0} \sigma _1(t, a) dW_{G_1}(a) \right| + \left| \int _0^{\tau _0} \sigma _1(s, a) dW_{G_1}(a) \right| \\&\qquad + \left| \int _0^{\tau _0} \sigma _2(t, a) dW_{G_2}(a) \right| + \left| \int _0^{\tau _0} \sigma _2(s, a) dW_{G_2}(a) \right| \Bigg )\\&\qquad \times \Bigg ( \left| \int _0^{\tau _0} \sigma _1(t, a) - \sigma _1(s, a) dW_{G_1}(a) \right| \\&\qquad + \left| \int _0^{\tau _0} \sigma _2(t, a) - \sigma _2(s, a) dW_{G_2}(a) \right| \Bigg ) \Bigg ] \end{aligned}$$
$$\begin{aligned}&\le M_6 |t-s| + M_7 |t-s| \Bigg [ S(1,t,2) S(1,s,4) S(2,t,8) S(2,s,8)\\&\quad + M_8 \Bigg ( E \Bigg [ \left| \int _0^{\tau _0} \sigma _1(t, a) - \sigma _1(s, a) dW_{G_1}(a) \right| ^2 \Bigg ]^{1/2}\\&\quad \times \,S(1,t,4) S(1,s,8) S(2,t,16) S(2,s,16)\\&\quad + E \Bigg [ \left| \int _0^{\tau _0} \sigma _2(t, a) - \sigma _2(s, a) dW_{G_2}(a) \right| ^2 \Bigg ]^{1/2}\\&\quad \times \,S(1,t,4) S(1,s,8) S(2,t,16) S(2,s,16) \Bigg )\\&\le M_9 |t-s| + M_{10} \Bigg [ \Bigg ( \int _0^{1} (\sigma _1(t, a) - \sigma _1(s, a))^2 da \Bigg )^{1/2}\\&\quad + \Bigg ( \int _0^{1} (\sigma _2(t, a) - \sigma _2(s, a))^2 da \Bigg )^{1/2} \Bigg ]\le M_{11} |t-s| \end{aligned}$$

where we denoted

$$\begin{aligned} S(i,u,n) = E \Bigg [ \exp \Bigg ( n \left| \int _0^{\tau _0} \sigma _i(u, a) dW_{G_i}(a) \right| \Bigg ) \Bigg ]^{1/n}. \end{aligned}$$

Therefore \(C^A(G_1, G_2, t)\) is Lipschitz continuous and there exist \(\partial C^A/ \partial t\) almost everywhere which is locally bounded.

Applying Theorem 3.2 in Jaillet et al. (1990), we can see that \(C^A\) satisfies the variational inequalities (9).

Using the transformation \(G_1 = e^{x_1}\) and \(G_2 = e^{x_2}\) we can rewrite Eq. (9) as

$$\begin{aligned}&\frac{1}{2} \Bigg ( \frac{\partial ^2 C^A}{\partial x_1^2} \sigma _{G_1}^2(t) + 2 \frac{\partial ^2 C^A}{\partial x_1 \partial x_2} \rho _{G_1, G_2}(t) \sigma _{G_1}(t) \sigma _{G_2}(t) + \frac{\partial ^2 C^A}{\partial x_2^2} \sigma _{G_2}^2(t) \Bigg )\nonumber \\&\quad \le rC^A + \frac{\sigma _{G_1}^2(t)}{2} \frac{\partial C^A}{\partial x_1} + \frac{\sigma _{G_2}^2(t)}{2} \frac{\partial C^A}{\partial x_2} - \frac{\partial C^A}{\partial t}. \end{aligned}$$
(25)

From Proposition 2.6, \(C^A\) is convex and therefore we haveFootnote 24

$$\begin{aligned} 0\le & {} (\rho _{G_1, G_2}(t) \sigma _{G_1}(t), \sigma _{G_2}(t)) \left( \begin{array}{cc} \frac{\partial ^2 C^A}{\partial x_1^2} &{} \frac{\partial ^2 C^A}{\partial x_1 \partial x_2} \\ \frac{\partial ^2 C^A}{\partial x_1 \partial x_2} &{} \frac{\partial ^2 C^A}{\partial x_2^2} \end{array}\right) \left( \begin{array}{c} \rho _{G_1, G_2}(t) \sigma _{G_1}(t) \\ \sigma _{G_2}(t) \end{array}\right) \\= & {} \rho _{G_1, G_2}(t) \sigma _{G_1}^2(t) \frac{\partial ^2 C^A}{\partial x_1^2} + 2 \rho _{G_1, G_2}(t) \sigma _{G_1}(t) \sigma _{G_2}(t) \frac{\partial ^2 C^A}{\partial x_1 \partial x_2} + \sigma _{G_2}^2 \frac{\partial ^2 C^A}{\partial x_2^2} \end{aligned}$$

This implies

$$\begin{aligned} 0\le & {} \left( 1 - \rho _{G_1, G_2}^2(t)\right) \sigma _{G_1}^2(t) \frac{\partial ^2 C^A}{\partial x_1^2} \nonumber \\\le & {} \sigma _{G_1}^2(t) \frac{\partial ^2 C^A}{\partial x_1^2} + 2 \rho _{G_1, G_2}(t) \sigma _{G_1}(t) \sigma _{G_2}(t) \frac{\partial ^2 C^A}{\partial x_1 \partial x_2} + \sigma _{G_2}^2 \frac{\partial ^2 C^A}{\partial x_2^2}. \end{aligned}$$
(26)

Inequalities (25) and (26) implies

$$\begin{aligned} 0\le & {} \frac{1}{2} \left( 1 - \rho _{G_1, G_2}^2(t)\right) \sigma _{G_1}^2(t) \frac{\partial ^2 C^A}{\partial x_1^2} \nonumber \\\le & {} r C^A + \frac{\sigma _{G_1}^2(t)}{2} \frac{\partial C^A}{\partial x_1} + \frac{\sigma _{G_2}^2(t)}{2} \frac{\partial C^A}{\partial x_2} - \frac{\partial C^A}{\partial t} \end{aligned}$$
(27)

Since we have proved the local boundedness of \(\partial C^A / \partial t\) and the uniform boundedness of \(\partial C^A / \partial G_i, i = 1, 2\), the second partial derivative \(C_{11}^A\) are locally bounded. Similar argument derives the local boundedness of other second partial derivatives. \(\square \)

1.3.4 Proof of Proposition 2.8

With the same argument provided by Detemple (2006), we decompose the valuation of American spread option using the early exercise premium.Footnote 25

Proof

Since

$$\begin{aligned}&h_1 G_1(s, T_1) - h_2 G_2(s, T_2)\\&\quad = h_1 G_1(t, T_1) - h_2 G_2(t, T_2)\\&\qquad +\,h_1 \int _t^s G_1(u, T_1) \sum _{k=1}^2 \sigma _{S_k} \left[ e^{(T-u) {\varvec{\beta }}}\right] _{1,k} dW_{S_k}(u)\\&\qquad +\,h_1 \int _t^s G_1(u, T_1) \sum _{k=1}^2 \sigma _{\delta _k} \left[ e^{(T-u) {\varvec{\beta }}}\right] _{1,2+k} dW_{\delta _k}(u)\\&\qquad -\,h_2 \int _t^s G_2(u, T_2) \sum _{k=1}^2 \sigma _{S_k} \left[ e^{(T-u) {\varvec{\beta }}}\right] _{2,k} dW_{S_k}(u)\\&\qquad -\,h_2 \int _t^s G_2(u, T_2) \sum _{k=1}^2 \sigma _{\delta _k} \left[ e^{(T-u) {\varvec{\beta }}}\right] _{2,2+k} dW_{\delta _k}(u) \end{aligned}$$

holds, \(h_1 G_1(t, T_1) - h_2 G_2(t, T_2)\) is a continuous semimartingale process and thus we can apply Tanaka’s formula.Footnote 26

$$\begin{aligned}&(h_1 G_1(s, T_1) - h_2 G_2(s, T_2) - K)^+\\&\quad = (h_1 G_1(t, T_1) - h_2 G_2(t, T_2) - K)^+\\&\qquad + \int _t^s 1_{h_1 G_1(u, T_1) - h_2 G_2(u, T_2) > K} dM(u) + L_K(t, G) \end{aligned}$$

where

$$\begin{aligned} M(u)= & {} h_1 \int _t^u G_1(v, T_1) \sum _{k=1}^2 \sigma _{S_k} \left[ e^{(T-v) {\varvec{\beta }}}\right] _{1,k} dW_{S_k}(v)\\&+\,h_1 \int _t^u G_1(v, T_1) \sum _{k=1}^2 \sigma _{\delta _k} \left[ e^{(T-v) {\varvec{\beta }}}\right] _{1,2+k} dW_{\delta _k}(v)\\&-\,h_2 \int _t^u G_2(v, T_2) \sum _{k=1}^2 \sigma _{S_k} \left[ e^{(T-v) {\varvec{\beta }}}\right] _{2,k} dW_{S_k}(v)\\&-\,h_2 \int _t^u G_2(v, T_2) \sum _{k=1}^2 \sigma _{\delta _k} \left[ e^{(T-v) {\varvec{\beta }}}\right] _{2,2+k} dW_{\delta _k}(v) \end{aligned}$$

and \(L_K(t, G)\) is the local time associated with \(G(t) = h_1 G_1(t, T_1) - h_2 G_2(t, T_2)\). Let us define

$$\begin{aligned} D(s) = e^{-r(s-t)} (h_1 G_1(s, T_1) - h_2 G_2(s, T_2) - K)^+. \end{aligned}$$

Therefore, our concern is the following Snell envelope.

$$\begin{aligned} C^A(G_1, G_2, t)= & {} \sup _{\tau \in \mathcal {S}_{t,T_0}} E_t [ D(\tau ) ] \end{aligned}$$

Using Ito’s lemma, we have

$$\begin{aligned} dD(s)= & {} d \left( e^{-r(s-t)} (h_1 G_1(s, T_1) - h_2 G_2(s, T_2) - K)^+ \right) \\= & {} e^{-r(s-t)} \left( - r (h_1 G_1(s, T_1) - h_2 G_2(s, T_2) - K)^+ ds + dL_K(s, G) + dM_s\right) . \end{aligned}$$

Denote

$$\begin{aligned} \tau _t = \inf \left\{ s \in [t,T]: C^A(G_1, G_2, s) = e^{-r(s-t)} (h_1 G_1(s, T_1) - h_2 G_2(s, T_2) - K)^+ \right\} \end{aligned}$$

where the following condition holds.Footnote 27

$$\begin{aligned} C^A(G_1, G_2, \tau _t) = e^{-r(\tau _t-t)} (h_1 G_1(\tau _t, T_1) - h_2 G_2(\tau _t, T_2) - K)^+ \end{aligned}$$

This implies

$$\begin{aligned} C^A(G_1, G_2, t)= & {} E_t \left[ e^{-r(\tau _t,t)} (h_1 G_1(\tau _t, T_1) - h_2 G_2(\tau _t, T_2) - K)^+\right] \\= & {} E_t [ C^A(G_1, G_2, \tau _t) ]. \end{aligned}$$

Furthermore, we have \(\tau _s=\tau _t\) for \(t \le s < \tau _t\) or otherwise \(\tau _t\) is dominated by some other stopping time \(\tau _s\) on \(A \in \mathcal {F}_s\), that is,

$$\begin{aligned} E_s[ C^A(G_1, G_2, \tau _t) ] < E_s[ C^A(G_1, G_2, \tau _s) ], A \in \mathcal {F}_s \end{aligned}$$

which yields

$$\begin{aligned}&E_s[ C^A(G_1, G_2, \tau _t) 1_A ] < E_s[ C^A(G_1, G_2, \tau _s) 1_A ]\\&\quad \Rightarrow E_t[ C^A(G_1, G_2, \tau _t) ] < E_t[ C^A(G_1, G_2, \tau _s) 1_A + C^A(G_1, G_2, \tau _t) (1 - 1_A) ] \end{aligned}$$

and this contradicts the optimality of \(\tau _t\). Therefore,

$$\begin{aligned} C^A(G_1, G_2, s)= & {} E_s[ e^{-r(\tau _s-t)} (h_1 G_1(\tau _s, T_1) - h_2 G_2(\tau _s, T_2) - K)^+ ]\\= & {} E_s[ C^A(G_1, G_2, \tau _t) ], s \in [t, \tau _t). \end{aligned}$$

Thus, \(C^A(G_1, G_2, s)\) is a martingale for \(s \in [t, \tau _t)\). Now define

$$\begin{aligned} N(s)= & {} C^A(G_1, G_2, s)+ \int _t^s 1_{\{\tau _u=u\}} e^{-r(u-t)} (r (h_1 G_1(u, T_1) \\&- h_2 G_2(u, T_2) - K)^+ du - dL_K(u, G)). \end{aligned}$$

This is a martingale which can be proved by the following argument. Since \(C^A(G_1, G_2, \tau _t) = E_t[ (h_1 G_1(\tau _t, T_1) - h_2 G_2(\tau _t, T_2) - K)^+]\),

$$\begin{aligned} \int _t^{T_0} 1_{\{\tau _u=u\}} dC^A(G_1, G_2, u) = \int _t^{T_0} 1_{\{\tau _u=u\}} dD(u). \end{aligned}$$

Rearraging this equation, we have

$$\begin{aligned} \int _t^{T_0} 1_{\{\tau _u=u\}} dM(u)= & {} \int _t^{T_0} 1_{\{\tau _u=u\}} dC^A(G_1, G_2, u)\\&- \int _t^{T_0} 1_{\{\tau _u=u\}} e^{-r(u-t)} (- r (h_1 G_1(u, T_1)\\&-\,h_2 G_2(u, T_2) - K)^+ du + dL_K(u, G)). \end{aligned}$$

Applying Ito’s lemma, we have

$$\begin{aligned} N(s)= & {} C^A(G_1, G_2, s)\\&+ \int _t^s 1_{\{\tau _u=u\}} e^{-r(u-t)} r (h_1 G_1(u, T_1) - h_2 G_2(u, T_2) - K)^+ du\\= & {} C^A(G_1, G_2, t) + \int _t^{s} 1_{\{\tau _u=u\}} dC^A(G_1, G_2, u)\\&+ \int _t^{s} 1_{\{\tau _u>u\}} dC^A(G_1, G_2, u)\\&+ \int _t^s 1_{\{\tau _u=u\}} e^{-r(u-t)} r (h_1 G_1(u, T_1) - h_2 G_2(u, T_2) - K)^+ du\\= & {} C^A(G_1, G_2, t) + \int _t^{s} 1_{\{\tau _u>u\}} dC^A(G_1, G_2, u) + \int _t^{T} 1_{\{\tau _u=u\}} dM(u). \end{aligned}$$

Therefore, N(t) is a martingale. By definition

$$\begin{aligned} C^A(G_1, G_2, T_0)= & {} \sup _{\tau \in \mathcal {S}_{T_0, T_0}} E_{T_0} [ D(\tau ) ] = E_{T_0} [ D_{T_0} ] = D_{T_0} \end{aligned}$$

and

$$\begin{aligned} C^A(G_1, G_2, t)= & {} \sup _{\tau \in \mathcal {S}_{t, T_0}} E_{t} [ D_{\tau } ] = E_t [ D_{\tau _t} ] \end{aligned}$$

Using that N(t) is a martingale and the above two equalities, we have

$$\begin{aligned}&E_{t} \left[ e^{-r(\tau _t-t)} (h_1 G_1(\tau _t, T_1) - h_2 G_2(\tau _t, T_2) - K)^+ \right] \\&\quad = C^A(G_1, G_2, t) = N_t = E_t [ N_{T_0} ]\\&\quad =E_t \Bigg [ C^A(G_1, G_2, T_0)+ \int _t^{T_0} 1_{\{\tau _u=u\}} e^{-r(u-t)} (r (h_1 G_1(u, T_1) \\&\quad \qquad - h_2 G_2(u, T_2) - K)^+ du - dL_K(u, G)) \Bigg ]\\&\quad =E_t \Bigg [ e^{-r(T_0-t)} (h_1 G_1(T_0, T_1) - h_2 G_2(T_0, T_2) - K)^+\\&\quad \qquad + \int _t^{T_0} 1_{\{\tau _u=u\}} e^{-r(u-t)} (r (h_1 G_1(u, T_1) \\&\quad \qquad - h_2 G_2(u, T_2) - K)^+ du - dL_K(u, G)) \Bigg ]. \end{aligned}$$

As indicated by Rutkowski (1994),Footnote 28

$$\begin{aligned}&\int _t^{T_0} 1_{\{\tau _u=u\}} e^{-r(u-t)} dL_K(u, G)\\&\quad = \int _t^{T_0} 1_{\{\tau _u=u\}} 1_{\{h_1 G_1(u, T_1) - h_2 G_2(u, T_2) = K\}} e^{-r(u-t)} dL_K(u, G)\\&\quad = 0. \end{aligned}$$

where the first equality is due to the properties of the local timeFootnote 29 and the second equality is from the fact \(\{ \tau _u=u \} \subset \{h_1 G_1(u, T_1) > h_2 G_2(u, T_2) + K \}\) where Proposition 2.5 implies. Moreover, since

$$\begin{aligned} \tau _u=u \iff (G_1(u, T_1), G_2(u, T_2), u) \in \mathcal {E} \iff G_1(u, T_1) \ge B(G_2(u, T_2), u), \end{aligned}$$

we have the early exercise premium representation for American commodity spread options.

The boundary condition \(C^A(B(G_2, t), G_2, t) = B(G_2, t) - G_2 - K\) imply the recursive Eq. (11). The boundary condition (12) is the value of spread option at maturity. The boundary condition (13) hold since the spread option is a standard option on a single asset when price \(G_2\) is zero. \(\square \)

1.3.5 Proof of Proposition 2.9

Finally, we provide an analytical approximation pricing formula for American commodity spread option. The difficulty is the calculation of early exercise premium, more precisely the domain of integration or the condition inequality of exercise which are not analytically tractable. Therefore, we use the scheme of Bjerksund and Stensland (1994) to approximate the condition inequality which split spread option in to two call option that the first option has stochastic exercise price and then use Barone-Adesi and Whaley (1987) framework to approximate the two American option. The formula can be calculated as we did in European call option which derives the analytical approximated pricing formula for American commodity spread options.

Proof

As Bjerksund and Stensland (1994)Footnote 30 proposed, we approximate the early exercise premium as follows.

$$\begin{aligned}&E_t \left[ (h_1 G_1(u, T_1) - h_2 G_2(u, T_2) - K) 1_{ \{ G_1(u, T_1) \ge B(G_2(u, T_2), u) \} } \right] \\&\quad \approx E_t \left[ (h_1 G_1(u, T_1) - h_2 G_2(u, T_2) - K) 1_{ \{ h_1 G_1(u, T_1) \ge h_2 B_2 G_2(u, T_2) + B_K K \}} \right] , \end{aligned}$$

where

$$\begin{aligned} B_2= & {} B_0\left( T_0 - u, \sigma _{X_{G_1}}^2(u, T_0) - 2 \sigma _{X_{G_1} X_{G_2}} (u, T_0) + \sigma _{X_{G_2}}^2(u, T_0)\right) ,\\ B_K= & {} B_0\left( T_0 - u, \sigma _{X_{G_1}}^2 (u, T_0)\right) ,\\ B_0(t, \sigma ^2)= & {} e^{g(t, \sigma ^2)} + \left( 1 - e^{g(t, \sigma ^2)}\right) B_{\infty }(\sigma ^2),\\ B_{\infty }(\sigma ^2)= & {} \frac{\gamma (\sigma ^2)}{\gamma (\sigma ^2) - 1},\\ \gamma (\sigma ^2)= & {} \frac{1}{2} + \sqrt{\frac{1}{4} + \frac{2r}{\sigma ^2}}, \end{aligned}$$

and

$$\begin{aligned} g(t, \sigma ^2) = - 2 \sigma \sqrt{t} (\gamma (\sigma ^2) - 1). \end{aligned}$$

The approximation is constructed in two steps. The first step is to split the spread option into an exchange option and vanilla type option. And the second step is due to Barone-Adesi and Whaley (1987) framework of approximating American option.

Note that the exercise region is

$$\begin{aligned}&h_1 G_1 (u, T_1) \ge h_2 B_2 G_2(u, T_2) + B_K K \\&\quad \Leftrightarrow x_1 \le d_{eep0}(x_2, u) \equiv \ln (h_2 B_2 G_2 (t, T_2) e^{x_2} + B_K K) - \ln (h_1 G_1(t, T_1)). \end{aligned}$$

The integrals of early exercise premium can be calculated just as the integral of European commodity spread option which we now have

$$\begin{aligned}&a^A (G_1, G_2, t; B)\\&\quad = r \int _t^{T_0} e^{-r(u-t)} E_t \left[ (h_1 G_1(u, T_1) - h_2 G_2(u, T_2) - K)\right. \\&\left. \qquad \times 1_{ \{ G_1(u, T_1) \ge B(G_2(u, T_2), u) \} } \right] du\\&\quad \approx r \int _t^{T_0} e^{-r(u-t)} E_t \left[ (h_1 G_1(u, T_1) - h_2 G_2(u, T_2) - K)\right. \\&\left. \qquad \times 1_{ \{ x_1 \le d_{eep0}(x_2, u) \} } \right] du\\&\quad = r \Biggl [ h_1 G_1(t, T_1) \int _t^{T_0} \exp \Biggl \{ -r (u - t) + \mu _{X_{G_1}}(t, u, T_1) + \frac{1}{2} \sigma _{X_{G_1}}^2(t, u, T_1) \Biggr \}\\&\quad \qquad \times \int _{-\infty }^{\infty } \Phi (d_{eep2}(x_2, u))\\&\quad \qquad \times n\left( x_2 | \mu _{X_{G_2}}(t, u, T_2) + \sigma _{X_{G_1} X_{G_2}}(t, u, T_1, T_2), \sigma _{X_{G_2}}^2(t, u, T_2)\right) dx_2 du\\&\quad \qquad - h_2 G_2(t, T_2) \int _t^{T_0} \exp \Biggl \{ -r (u - t) + \mu _{X_{G_2}}(t, u, T_2) + \frac{1}{2} \sigma _{X_{G_2}}^2(t, u, T_2)\Biggr \}\\&\quad \qquad \times \int _{-\infty }^{\infty } \Phi (d_{eep1}(x_2, u))\\&\quad \qquad \times n\left( x_2 | \mu _{X_{G_2}}(t, u, T_2) + \sigma _{X_{G_2}}^2(t, u, T_2), \sigma _{X_{G_2}}^2(t, u, T_2)\right) dx_2 du\\&\quad \qquad - K \int _t^{T_0} e^{-r(u-t)} \int _{-\infty }^{\infty } \Phi (d_{eep1}(x_2, u))\\&\quad \qquad \times n\left( x_2 | \mu _{X_{G_2}}(t, u, T_2), \sigma _{X_{G_2}}^2(t, u, T_2)\right) dx_2 du \Biggr ], \end{aligned}$$

where

$$\begin{aligned} d_{eep1}(x_2, u)= & {} - \frac{\ln (h_2 B_2 G_2(t, T_2) e^{x_2} + B_K K) - \ln (h_1 G_1(t, T_1)) - \mu _{X_{G_1}}(t, u, T_1)}{\sigma _{X_{G_1}}(t, u, T_1) \sqrt{1 - \rho _{X_{G_1} X_{G_2}}^2(t, u, T_1, T_2)}} \\&+ \frac{\rho _{X_{G_1} X_{G_2}}(t, u, T_1, T_2) \sigma _{X_{G_1}}(t, u, T_1) \frac{x_2 - \mu _{X_{G_2}}(t, u, T_2)}{\sigma _{X_{G_2}}^2(t, u, T_2)}}{\sigma _{X_{G_1}}(t, u, T_1) \sqrt{1 - \rho _{X_{G_1} X_{G_2}}^2(t, u, T_1, T_2)}}, \\ \end{aligned}$$

and

$$\begin{aligned} d_{eep2}(x_2, u)= & {} d_{eep}(x_2, u) + \sigma _{X_{G_1}}(t, u, T_1) \sqrt{1 - \rho _{X_{G_1} X_{G_2}}^2(t, u, T_1, T_2)}. \end{aligned}$$

This is the approximation formula. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Nakajima, K., Ohashi, K. Commodity Spread Option with Cointegration. Asia-Pac Financ Markets 23, 1–44 (2016). https://doi.org/10.1007/s10690-015-9207-1

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10690-015-9207-1

Keywords

JEL Classification

Navigation