Skip to main content

Advertisement

Log in

A new approach to wind power futures pricing

  • Published:
Decisions in Economics and Finance Aims and scope Submit manuscript

Abstract

We propose a new model for the pricing of wind power futures written on the wind power production index. Our approach is based on an arithmetic multi-factor pure-jump Ornstein–Uhlenbeck setup with time-dependent coefficients. We express the wind power production index and the corresponding futures price in terms of Fourier integrals and derive the related time dynamics. We conclude the paper by an investigation of the risk premium associated with our wind power model.

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

References

  • Applebaum, D.: Lévy Processes and Stochastic Calculus. Cambridge University Press, Cambridge (2009)

    Book  Google Scholar 

  • Benth, F., Christensen, T., Rohde, V.: Multivariate continuous-time modeling of wind indexes and hedging of wind risk. Quant. Finance 21(1), 165–183 (2021)

    Article  Google Scholar 

  • Benth, F., Di Persio, L., Lavagnini, S.: Stochastic modeling of wind derivatives in energy markets. Risks 6(2), 56 (2018)

    Article  Google Scholar 

  • Benth, F., Kallsen, J., Meyer-Brandis, T.: A non-Gaussian Ornstein-Uhlenbeck process for electricity spot price modeling and derivatives pricing. Appl. Math. Finance 14(2), 153–169 (2007)

    Article  Google Scholar 

  • Benth, F., Meyer-Brandis, T.: The information premium for non-storable commodities. J. Energy Mark. 2(3), 111–140 (2009)

    Article  Google Scholar 

  • Benth, F., Pircalabu, A.: A non-Gaussian Ornstein-Uhlenbeck model for pricing wind power futures. Appl. Math. Finance 25(1), 36–65 (2018)

    Article  Google Scholar 

  • Benth, F., Saltyte-Benth, J.: Dynamic pricing of wind futures. Energy Econom. 31(1), 16–24 (2009)

    Article  Google Scholar 

  • Benth, F., Saltyte-Benth, J.: Modeling and Pricing in Financial Markets for Weather Derivatives. World Scientific, Singapore (2012)

    Book  Google Scholar 

  • Benth, F., Saltyte-Benth, J., Koekebakker, S.: Stochastic Modeling of Electricity and Related Markets, 1st edn. World Scientific, Singapore (2008)

    Book  Google Scholar 

  • Carr, P., Madan, D.: Option valuation using the fast Fourier transform. J. Comput. Finance 2(4), 61–73 (1999)

    Article  Google Scholar 

  • Cont, R., Tankov, P.: Financial Modeling with Jump Processes, 1st edn. Chapman & Hall/CRC, London (2004)

    Google Scholar 

  • Di Nunno, G., Øksendal, B., Proske, F.: Malliavin Calculus for Lévy Processes with Applications to Finance, 1st edn. Springer, Berlin (2009)

    Book  Google Scholar 

  • Gersema, G., Wozabal, D.: An equilibrium pricing model for wind power futures. Energy Econom. 65, 64–74 (2017)

    Article  Google Scholar 

  • Hess, M.: Pricing Energy, Weather and Emission Derivatives under Future Information, PhD thesis, University Duisburg-Essen, Germany. https://duepublico.uni-duisburg-essen.de/servlets/DocumentServlet?id=31060. (2013)

  • Hess, M.: Modeling and pricing precipitation derivatives under weather forecasts. Int. J. Theor. Appl. Finance 19(7), 1650051 (2016)

    Article  Google Scholar 

  • Hess, M.: Pricing temperature derivatives under weather forecasts. Int. J. Theor. Appl. Finance 21(5), 1850031 (2018)

    Article  Google Scholar 

  • Huisman, R., Kilic, M.: Electricity futures prices: indirect storability, expectations, and risk premiums. Energy Econ. 34(4), 892–898 (2012)

    Article  Google Scholar 

  • Jacod, J., Shiryaev, A.: Limit Theorems for Stochastic Processes, 2nd edn. Springer, Berlin (2003)

    Book  Google Scholar 

  • Pircalabu, A., Jung, J.: A mixed C-vine copula model for hedging price and volumetric risk in wind power trading. Quant. Finance 17(10), 1583–1600 (2017)

    Article  Google Scholar 

  • Protter, P.: Stochastic Integration and Differential Equations, 2nd edn. Springer, Berlin (2005)

    Book  Google Scholar 

  • Ramírez-Rosado, I., Fernández-Jiménez, L.: An advanced model for short-term forecasting of mean wind speed and wind electric power. Control. Intell. Syst. 32(1), 21–26 (2004)

    Google Scholar 

  • Sato, K.: Lévy Processes and Infinitely Divisible Distributions, Cambridge studies in advanced mathematics, No. 68, Cambridge University Press, Cambridge (1999)

  • Schoutens, W.: Lévy Processes in Finance: Pricing Financial Derivatives. Wiley, Chichester (2003)

    Book  Google Scholar 

  • Tan, Z., Tankov, P.: Optimal trading policies for wind energy producer. SIAM J. Financ. Math. 9(1), 315–346 (2018)

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Markus Hess.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix

Appendix

In what follows, we present an alternative modeling approach for the wind power production index. First of all, note that for \(x \ge 0\) the function \(\eta \left( x \right): = e^{ - x}\) takes values in the interval \(\left] {0,1} \right]\). Hence, instead of (2.8), we could alternatively define the wind power production index \(P_{t}\) via

$$ P_{t} : = \eta \left( {Y_{t} } \right) = e^{{ - Y_{t} }} $$
(A.1)

where \(t \ge 0\). If we do so, we find \(0 < P_{t} \le 1\) \({\mathbb{P}}\)-a.s. \(\forall t \ge 0\), because \(Y_{t}\) in (2.7) is non-negative. Substituting (A.1) and (2.7) into (3.1), we get

$$ f_{t} \left( T \right) = {\mathbb{E}}_{{\mathbb{Q}}} \left( {\exp \left\{ { - \mathop \sum \limits_{k = 1}^{n} X_{T}^{k} } \right\}{|}{\mathcal{F}}_{t} } \right) $$
(A.2)

for all \(0 \le t \le T\). We next put (2.6) [with \(t\) therein replaced by \(T\)] into (A.2) and obtain

$$ f_{t} \left( T \right) = \exp \left\{ { - \mathop \sum \limits_{k = 1}^{n} \left( {x_{k} e^{{ - \lambda_{k} T}} + \mathop \int \limits_{0}^{t} \mathop \int \limits_{{D_{k} }}^{ } \theta_{k} \left( {s,z,T} \right) N_{k} \left( {{\text{d}}s,{\text{d}}z} \right)} \right)} \right\} \times {\mathbb{E}}_{{\mathbb{Q}}} \left[ {\exp \left\{ { - \mathop \sum \limits_{k = 1}^{n} \mathop \int \limits_{t}^{T} \mathop \int \limits_{{D_{k} }}^{ } \theta_{k} \left( {s,z,T} \right) N_{k} \left( {{\text{d}}s,{\text{d}}z} \right)} \right\}} \right] $$
(A.3)

where \(\theta_{k} \left( {s,z,T} \right)\) is such as defined in (3.3). In the last step, we used the independent increment property of the processes \(L^{1} , \ldots ,L^{n}\). An application of the Lévy–Khinchin formula yields

$$ {\mathbb{E}}_{{\mathbb{Q}}} \left[ {\exp \left\{ { - \mathop \sum \limits_{k = 1}^{n} \mathop \int \limits_{t}^{T} \mathop \int \limits_{{D_{k} }}^{ } \theta_{k} \left( {s,z,T} \right) N_{k} \left( {{\text{d}}s,{\text{d}}z} \right)} \right\}} \right] = \exp \left\{ {\mathop \sum \limits_{k = 1}^{n} \mathop \int \limits_{t}^{T} \mathop \int \limits_{{D_{k} }}^{ } \left[ {e^{{ - \theta_{k} \left( {s,z,T} \right)}} - 1} \right] e^{{h_{k} \left( {s,z} \right)}} \rho_{k} \left( s \right) \nu_{k} \left( {{\text{d}}z} \right){\text{d}}s} \right\} . $$

Merging the latter equation into (A.3), we deduce the following representation for the wind power futures price

$$ f_{t} \left( T \right) = \exp \left\{ { - \mathop \sum \limits_{k = 1}^{n} \left( {x_{k} e^{{ - \lambda_{k} T}} + \mathop \int \limits_{t}^{T} \mathop \int \limits_{{D_{k} }}^{ } \left[ {1 - e^{{ - \theta_{k} \left( {s,z,T} \right)}} } \right] e^{{h_{k} \left( {s,z} \right)}} \rho_{k} \left( s \right) \nu_{k} \left( {{\text{d}}z} \right){\text{d}}s + \mathop \int \limits_{0}^{t} \mathop \int \limits_{{D_{k} }}^{ } \theta_{k} \left( {s,z,T} \right) N_{k} \left( {{\text{d}}s,{\text{d}}z} \right)} \right)} \right\} $$
(A.4)

which corresponds to Eq. (3.4). Applying Itô’s formula on (A.4), for all \(0 \le t \le T\) we infer the geometric \(\left( {{\mathcal{F}},{\mathbb{Q}}} \right)\)-martingale dynamics

$$ {\text{d}}f_{t} \left( T \right) = f_{t - } \left( T \right)\mathop \sum \limits_{k = 1}^{n} \mathop \int \limits_{{D_{k} }}^{ } \left[ {e^{{ - \theta_{k} \left( {t,z,T} \right)}} - 1} \right] \tilde{N}_{k}^{\mathbb{Q}} \left( {{\text{d}}t,{\text{d}}z} \right) $$
(A.5)

where the initial value \(f_{0} \left( T \right)\) can immediately be derived from (A.4). We recall that (A.5) corresponds to Eq. (3.13). We stress that the wind power production index \(P\) defined in (A.1) is always strictly positive which ought to be regarded as a disadvantage from a practitioner’s perspective, as in reality the wind power production index can be equal to zero during certain time periods with no or little wind. We recall that, on the contrary, the wind power production index defined in (2.8) also takes zero values yielding a more realistic modeling setup.

In the next step, we derive the risk premium formula associated with the geometric model (A.1)–(A.5). Parallel to Sect. 3, we define

$$ f_{t}^{{\mathbb{P}}} \left( T \right) = {\mathbb{E}}_{\mathbb{P}} \left( P_T| {\mathcal{F_t}} \right). $$

Using (A.1), (2.7) and (2.6), we then get

$$ f_{t}^{{\mathbb{P}}} \left( T \right) = \exp \left\{ { - \mathop \sum \limits_{k = 1}^{n} \left( {x_{k} e^{{ - \lambda_{k} T}} + \mathop \int \limits_{0}^{t} \mathop \int \limits_{{D_{k} }}^{ } \theta_{k} \left( {s,z,T} \right) N_{k} \left( {{\text{d}}s,{\text{d}}z} \right)} \right)} \right\} \times {\mathbb{E}}_{{\mathbb{P}}} \left[ {\exp \left\{ { - \mathop \sum \limits_{k = 1}^{n} \mathop \int \limits_{t}^{T} \mathop \int \limits_{{D_{k} }}^{ } \theta_{k} \left( {s,z,T} \right) N_{k} \left( {{\text{d}}s,{\text{d}}z} \right)} \right\}} \right] $$

which closely resembles (A.3), but with \({\mathbb{Q}}\) therein replaced by \({\mathbb{P}}\). The appearing \({\mathbb{P}}\)-expectation can be computed by the Lévy–Khinchin formula yielding

$$ f_{t}^{{\mathbb{P}}} \left( T \right) = \exp \left\{ { - \mathop \sum \limits_{k = 1}^{n} \left( {x_{k} e^{{ - \lambda_{k} T}} + \mathop \int \limits_{t}^{T} \mathop \int \limits_{{D_{k} }}^{ } \left[ {1 - e^{{ - \theta_{k} \left( {s,z,T} \right)}} } \right] \rho_{k} \left( s \right) \nu_{k} \left( {{\text{d}}z} \right){\text{d}}s + \mathop \int \limits_{0}^{t} \mathop \int \limits_{{D_{k} }}^{ } \theta_{k} \left( {s,z,T} \right) N_{k} \left( {{\text{d}}s,{\text{d}}z} \right)} \right)} \right\} $$

which equals (A.4), but with \(h_{k} \left( {s,z} \right) \equiv 0\) therein. We now substitute the latter equation as well as (A.4) into (3.14) which leads us to the risk premium formula

$$ R_{t}^{{{\mathbb{P}},{\mathbb{Q}}}} \left( T \right) = \exp \left\{ { - \mathop \sum \limits_{k = 1}^{n} \left( {x_{k} e^{{ - \lambda_{k} T}} + \mathop \int \limits_{0}^{t} \mathop \int \limits_{{D_{k} }}^{ } \theta_{k} \left( {s,z,T} \right) N_{k} \left( {ds,dz} \right)} \right)} \right\} \times \left[ {\exp \left\{ {\mathop \sum \limits_{k = 1}^{n} \mathop \int \limits_{t}^{T} \mathop \int \limits_{{D_{k} }}^{ } \left[ {e^{{ - \theta_{k} \left( {s,z,T} \right)}} - 1} \right] e^{{h_{k} \left( {s,z} \right)}} \rho_{k} \left( s \right) \nu_{k} \left( {{\text{d}}z} \right){\text{d}}s} \right\} - \exp \left\{ {\mathop \sum \limits_{k = 1}^{n} \mathop \int \limits_{t}^{T} \mathop \int \limits_{{D_{k} }}^{ } \left[ {e^{{ - \theta_{k} \left( {s,z,T} \right)}} - 1} \right] \rho_{k} \left( s \right) \nu_{k} \left( {{\text{d}}z} \right){\text{d}}s} \right\}} \right] $$

where \(0 \le t \le T\). Note that the latter equation corresponds to (3.15). By the Itô formula we eventually deduce the geometric \(\left( {{\mathcal{F}},{\mathbb{P}}} \right)\)-martingale dynamics

$$ {\text{d}}f_{t}^{{\mathbb{P}}} \left( T \right) = f_{t - }^{{\mathbb{P}}} \left( T \right)\mathop \sum \limits_{k = 1}^{n} \mathop \int \limits_{{D_{k} }}^{ } \left[ {e^{{ - \theta_{k} \left( {t,z,T} \right)}} - 1} \right] \tilde{N}_{k}^{{\mathbb{P}}} \left( {{\text{d}}t,{\text{d}}z} \right) $$

where we used (2.4). The latter SDE corresponds to (A.5).

In the sequel, we show that the model proposed in Benth and Pircalabu (2018) is contained as a sub-case in our current geometric model associated with (A.1). If we take \({n = 1}\), \(w_{1} = 1\) and \(\mu \left( t \right) = - \ln \Lambda \left( t \right)\) in (2.1), we get

$$ Y_{t} = X_{t}^{1} - \ln \Lambda \left( t \right) $$

and hence,

$$ P_{t} = \Lambda \left( t \right) e^{{ - X_{t}^{1} }} $$

due to (A.1). This representation for \(P_{t}\) possesses the same structure as Eq. (1) in Benth and Pircalabu (2018). If we further take \(\lambda_{1} = \alpha\) and \(\sigma_{1} \left( t \right) \equiv 1\), then (2.2) translates into

$$ {\text{d}}X_{t}^{1} = - \alpha X_{t}^{1} {\text{d}}t + {\text{d}}L_{t}^{1} $$

which implies

$$ {\text{d}}Y_{t} = \left( { - \frac{{\Lambda^{{\prime }} \left( t \right)}}{\Lambda \left( t \right)} - \alpha X_{t}^{1} } \right){\text{d}}t + {\text{d}}L_{t}^{1} . $$
(A.6)

Comparing (A.6) with Eq. (2) in Benth and Pircalabu (2018), we see that we get correspondence between these SDEs, if we take

$$ \Lambda \left( t \right) = e^{ - \alpha \mu t},\quad X_{t}^{1} = X_{t},\quad L_{t}^{1} = L_{t}. $$

We conclude that the model proposed in Benth and Pircalabu (2018) is a sub-case of the geometric model presented in (A.1)–(A.5) above.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hess, M. A new approach to wind power futures pricing. Decisions Econ Finan 44, 1235–1252 (2021). https://doi.org/10.1007/s10203-021-00345-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10203-021-00345-8

Keywords

Mathematics Subject Classification

JEL Classification

Navigation