Skip to main content
Log in

Seasonal volatility in agricultural markets: modelling and empirical investigations

  • Original Research
  • Published:
Annals of Operations Research Aims and scope Submit manuscript

Abstract

This paper deals with the issue of modelling the volatility of futures prices in agricultural markets. We develop a multi-factor model in which the stochastic volatility dynamics incorporate a seasonal component. In addition, we employ a maturity-dependent damping term to account for the Samuelson effect. We give the conditions under which the volatility dynamics are well defined and obtain the joint characteristic function of a pair of futures prices. We then derive the state-space representation of our model in order to use the Kalman filter algorithm for estimation and prediction. The empirical analysis is carried out using daily futures data from 2007 to 2019 for corn, cotton, soybeans, sugar and wheat. In-sample, the seasonal models clearly outperform the nested non-seasonal models in all five markets. Out-of-sample, we predict volatility peaks with high accuracy for four of these five commodities.

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

Similar content being viewed by others

Notes

  1. Since it costs nothing to take a position in a futures contract, in the absence of arbitrage opportunities, the futures price process must be a martingale under the risk-neutral measure. See Hull (2012) “The drift of the futures price in a risk-neutral world is therefore zero. [...] This result is a very general one. It is true for all futures prices and does not depend on any assumptions about interest rates, volatilities, etc.” See also Theorem 5.6.5 in Shreve (2004).

  2. For the computation of likelihood via the Kalman filter algorithm we refer to Tsay (2010). For model selection and ranking based on estimation results we refer to Burnham and Anderson (2002) and for likelihood ratio tests we refer to Chapter 7 of Silvey (1975).

  3. See Chap. 5 in Shreve (2004) for a proof and further details. We also refer to Chap. 2 in Lioui and Poncet (2005) for a detailed discussion of the dynamics of futures prices under the risk-neutral measure \(\mathbb {Q}\).

  4. We use the first ten years of the datasets (from November 2007 to November 2017) for analysis and estimation, while the last two years are kept separate and used only for out-of-sample predictions. This is to avoid, as far as possible, any data snooping (i.e. data dredging, see Burnham and Anderson (2002) for details).

  5. Maximisation of likelihood with simulated annealing algorithm is implemented in C++, using Boost libraries www.boost.org. All the rest is done in Matlab with the optimization toolbox and quadgk routine for numerical integration.

  6. We therefore estimate 12 models on each dataset.

References

  • Alexander, C., & Korovilas, D. (2013). Volatility exchange-traded notes: Curse or cure. Journal of Alternative Investments, 16(2), 52–71.

    Google Scholar 

  • Allen, S., & Schuster, E. (2004). Controlling the risk for an agricultural harvest. Manufacturing and Service Operations Management, 6(3), 225–236.

    Google Scholar 

  • Arismendi, J. C., Back, J., Prokopczuk, M., Paschke, R., & Rudolf, M. (2016). Seasonal stochastic volatility: Implications for the pricing of commodity options. Journal of Banking and Finance, 66, 53–65.

    Google Scholar 

  • Back, J., Prokopczuk, M., & Rudolf, M. (2013). Seasonality and the valuation of commodity options. Journal of Banking and Finance, 37, 273–290.

    Google Scholar 

  • Bakshi, G., & Madan, D. (2000). Spanning and derivative-security valuation. Journal of Financial Economics, 55(2), 205–238.

    Google Scholar 

  • Benhamou, E., Gobet, E., & Miri, M. (2010). Time dependent Heston model. SIAM Journal on Financial Mathematics, 1(1), 289–325.

    MathSciNet  Google Scholar 

  • Bollerslev, T., & Zhou, H. (2002). Estimating stochastic volatility diffusion using conditional moments of integrated volatility. Journal of Econometrics, 109, 33–65.

    MathSciNet  Google Scholar 

  • Burnham, K. P., & Anderson, D. R. (2002). Model selection and multimodel inference: A practical information-theoretic approach (2nd ed.). Springer.

  • Caldana, R., & Fusai, G. (2013). A general closed-form spread option pricing formula. Journal of Banking and Finance, 37(12), 4893–4906.

    Google Scholar 

  • Casassus, J., & Collin-Dufresne, P. (2005). Stochastic convenience yield implied from commodity futures and interest rates. Journal of Finance, 60(5), 2283–2331.

    Google Scholar 

  • Cerqueti, R., & Fanelli, V. (2021). Long memory and crude oil’s price predictability. Annals of Operations Research, 299, 895–906.

    MathSciNet  Google Scholar 

  • Chiarella, C., Kang, B., Nikitopoulos, C. S., & Tô, T. D. (2013). Humps in the volatility structure of the crude oil futures market: New evidence. Energy Economics, 40, 989–1000.

    Google Scholar 

  • Chun, Y. (2003). Optimal pricing and ordering policies for perishable commodities. European Journal of Operational Research, 144(1), 68–82.

    MathSciNet  Google Scholar 

  • Clark, I. J. (2014). Commodity Option Pricing: A Practitioner’s Guide. Wiley Finance, Wiley.

  • Clewlow, L., & Strickland, C. (1999a). A multi-factor model for energy derivatives, working Paper, 20 pages.

  • Clewlow, L., & Strickland, C. (1999b). Valuing energy options in a one factor model fitted to forward prices, working Paper, 30 pages.

  • Cox, J. C., Ingersoll, J. E., & Ross, S. A. (1985). A theory of the term structure of interest rates. Econometrica, 53, 385–408.

    MathSciNet  Google Scholar 

  • Diebold, F. X. (2015). Comparing predictive accuracy, twenty years later: A personal perspective on the use and abuse of Diebold-Mariano tests. Journal of Business and Economic Statistics, 33(1), 1–8.

    MathSciNet  Google Scholar 

  • Diebold, F. X., & Mariano, R. S. (1995). Comparing predictive accuracy. Journal of Business and Economic Statistics, 13(3), 253–263.

    Google Scholar 

  • Doran, J. S., & Ronn, E. I. (2008). Computing the market price of volatility risk in the energy commodity markets. Journal of Banking and Finance, 32, 2541–2552.

    Google Scholar 

  • Duffie, D., Pan, J., & Singleton, K. (2000). Transform analysis and asset pricing for affine jump-diffusions. Econometrica, 68(6), 1343–1376.

    MathSciNet  Google Scholar 

  • Fanelli, V., & Schmeck, M. D. (2019). On the seasonality in the implied volatility of electricity options. Quantitative Finance, 19(8), 1321–1337.

    MathSciNet  Google Scholar 

  • Fanelli, V., Musti, S., & Maddalena, L. (2015). Electricity market equilibrium model with seasonal volatilities. Procedia Engineering, 118, 1217–1224.

    Google Scholar 

  • Fanelli, V., Musti, S., & Maddalena, L. (2016). Modelling electricity futures prices using seasonal path-dependent volatility. Applied Energy, 173, 92–102.

    ADS  Google Scholar 

  • Galai, D. (1979). A proposal for indexes for traded call options. Journal of Finance, 34(5), 1157–1172.

    Google Scholar 

  • Geman, H., & Nguyen, V. N. (2005). Soybean inventory and forward curve dynamics. Management Science, 51(7), 1076–1091.

    Google Scholar 

  • Geman, H., & Roncoroni, A. (2006). Understanding the fine structure of electricity prices. Journal of Business, 79(3), 1225–1261.

    Google Scholar 

  • Goffe, W. L., Ferrier, G. D., & Rogers, J. (1994). Global optimization of statistical functions with simulated annealing. Journal of Econometrics, 60, 65–99.

    Google Scholar 

  • Gorton, G., & Rouwenhorst, K. G. (2006). Facts and fantasies about commodity futures. Financial Analysts Journal, 62(2), 47–68.

    Google Scholar 

  • Gorton, G., Hayashi, F., & Rouwenhorst, K. G. (2013). The fundamentals of commodity futures returns. Review of Finance, 17(1), 35–105.

    Google Scholar 

  • Heston, S. (1993). A closed-form solution for options with stochastic volatility with applications to bond and currency options. Review of Financial Studies, 6(2), 327–343.

    MathSciNet  Google Scholar 

  • Hull, J. (2012). Options, futures, and other derivatives (8th ed.). Pearson.

  • Hull, J., & White, A. (1990). Pricing interest-rate-derivative securities. The Review of Financial Studies, 3(4), 573–592.

    Google Scholar 

  • Islyaev, S., & Date, P. (2015). Electricity futures price models: Calibration and forecasting. European Journal of Operational Research, 247(1), 144–154.

    MathSciNet  Google Scholar 

  • Jaillet, P., Ronn, E. I., & Tompaidis, S. (2004). Valuation of commodity-based swing options. Management Science, 50(7), 909–921.

    Google Scholar 

  • Kaeck, A., & Alexander, C. (2012). Volatility dynamics for the S&P 500: Further evidence from non-affine, multi-factor jump diffusions. Journal of Banking and Finance, 36(11), 3110–3121.

    Google Scholar 

  • Karatzas, I., & Shreve, S. E. (1988). Brownian motion and stochastic calculus (2nd ed.). Graduate Texts in Mathematics. Springer.

  • Lioui, A., & Poncet, P. (2005). Dynamic Asset Allocation with Forwards and Futures. Springer.

  • Lu, T., Fransoo, J., & Lee, C. Y. (2017). Carrier portfolio management for shipping seasonal products. Operations Research, 65(5), 1250–1266.

    MathSciNet  Google Scholar 

  • Lucia, J. J., & Schwartz, E. S. (2002). Electricity prices and power derivatives: Evidence from the Nordic power exchange. Review of Derivatives Research, 5(1), 5–50.

    Google Scholar 

  • Maghsoodi, Y. (1996). Solution of the extended CIR term structure and bond option valuation. Mathematical Finance, 6(1), 89–109.

    MathSciNet  Google Scholar 

  • Moreno, M., Novales, A., & Platania, F. (2019). Long-term swings and seasonality in energy markets. European Journal of Operational Research, 279(3), 1011–1023.

    MathSciNet  Google Scholar 

  • Ni, Y., & Sandal, L. K. (2019). Seasonality matters: A multi-season, multi-state dynamic optimization in fisheries. European Journal of Operational Research, 275(2), 648–658.

    MathSciNet  Google Scholar 

  • Øksendal, B. (2003). Stochastic Differential Equations: An Introduction with Applications (6th ed.). Universitext, Springer.

  • Richter, M., & Sørensen, C. (2002). Stochastic volatility and seasonality in commodity futures and options: The case of soybeans. Tech. rep., Copenhagen Business School, working Paper, 45 pages.

  • Roncoroni, A., Fusai, G., & Cummins, M. (2015). Handbook of multi-commodity markets and products: Structuring, trading and risk management. Wiley Finance Series, Wiley.

  • Samuelson, P. A. (1965). Proof that properly anticipated prices fluctuate randomly. Industrial Management Review, 6(2), 41–49.

    Google Scholar 

  • Schneider, L., & Tavin, B. (2018). From the Samuelson volatility effect to a Samuelson correlation effect: An analysis of crude oil calendar spread options. Journal of Banking and Finance, 95, 185–202.

    Google Scholar 

  • Secomandi, N. (2010). Optimal commodity trading with a capacitated storage asset. Management Science, 56(3), 449–467.

    MathSciNet  Google Scholar 

  • Shreve, S. (2004). Stochastic calculus for finance II—Continuous-time models. Springer Finance, Springer.

  • Silvey, S. D. (1975). Statistical inference. Chapman and Hall.

  • Sørensen, C. (2002). Modeling seasonality in agricultural commodity futures. Journal of Futures Markets, 22(5), 393–426.

    Google Scholar 

  • Trolle, A. B., & Schwartz, E. S. (2009). Unspanned stochastic volatility and the pricing of commodity derivatives. Review of Financial Studies, 22(11), 4423–4461.

    Google Scholar 

  • Tsay, R. S. (2010). Analysis of financial time series (3rd ed.). Series in Probability and Statistics, Wiley.

  • Widodo, K., Nagasawa, H., Morizawa, K., & Ota, M. (2006). A periodical flowering-harvesting model for delivering agricultural fresh products. European Journal of Operational Research, 170(1), 24–43.

    Google Scholar 

  • Wiedenmann, S., & Geldermann, J. (2015). Supply planning for processors of agricultural raw materials. European Journal of Operational Research, 242(2), 606–619.

    Google Scholar 

Download references

Acknowledgements

We would like to thank Carol Alexander, Roy Cerqueti, David Criens, Ennio Fedrizzi, René Flacke, Yevhen Havrylenko, Kristofer Jürgensen, Rüdiger Kiesel, François Le Grand, Cassio Neri and participants at the Technical University of Munich 2017 Innovations in Insurance, Risk and Asset-Management Conference, the Nantes 2018 Commodity Markets Winter Workshop, the Duisburg-Essen Energy and Finance Seminar, the 35th Annual Conference of the French Finance Association in Paris, and the Commodity and Energy Markets Association Annual Meeting 2018 at La Sapienza in Rome, for helpful and stimulating comments, discussions and suggestions. We would also like to thank two anonymous referees for their comments and suggestions that helped improve the quality of the paper. All remaining errors are ours. Part of this paper was written while Lorenz Schneider was KPMG Visiting Professor at the Technical University of Munich.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to B. Tavin.

Additional information

Publisher's Note

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

Appendices

Proofs

In this Appendix, we give the proofs of several propositions found in the paper.

Proof of Proposition 1

The proof is an extension of the proof of Proposition 2.1 of Schneider and Tavin (2018) to the case where the variance mean-reversion level \(\theta \) is time-dependent. Going from \(\theta \) to \(\theta (t)\) leads to changes in two places. The first is in Lemma A.1 of Schneider and Tavin (2018), which needs to be modified as follows.

Lemma 1

Let \(\theta : \mathbb {R}_0^+ \rightarrow \mathbb {R}^+\) be the seasonal mean-reversion level function, let

$$\begin{aligned} \hat{\theta }_T(\lambda ) := \int _0^T e^{\lambda t} \theta (t) dt \end{aligned}$$

be its transform, and let the function \(f_1\) be given by \(f_1(t) := \sum _{k=1}^2 u_k e^{- \lambda _j(T_k - t)}\). Then:

$$\begin{aligned} \sigma \int _0^T f_1(t) \sqrt{v(t)} d\tilde{B}(t) = \left[ f_1(t) v(t) \right] _0^T - f_1(0) \kappa \hat{\theta }_T(\lambda ) + (\kappa - \lambda ) \int _0^T f_1(t) v(t) dt. \end{aligned}$$
(33)

Proof

Multiplying equation Eq. (10) by \(f_1(t)\) and then integrating from 0 to T gives:

$$\begin{aligned} \int _0^T f_1(t) dv(t) = \int _0^T f_1(t) \kappa (\theta (t) - v(t)) dt + \sigma \int _0^T f_1(t) \sqrt{v(t)} d\tilde{B}(t). \end{aligned}$$
(34)

Using Itô-integration by parts (see Øksendal (2003)), we also have:

$$\begin{aligned} \int _0^T f_1(t) dv(t)&= \left[ f_1(t) v(t) \right] _0^T - \int _0^T v(t) df_1(t) \nonumber \\&= \left[ f_1(t) v(t) \right] _0^T - \lambda \int _0^T f_1(t) v(t) dt. \end{aligned}$$
(35)

Equating the right hand sides of Eqs. (34) and (35) gives:

$$\begin{aligned} \sigma \int _0^T f_1(t) \sqrt{v(t)} d\tilde{B}(t)&= \left[ f_1(t) v(t) \right] _0^T - \lambda \int _0^T f_1(t) v(t) dt - \int _0^T f_1(t) \kappa (\theta (t) - v(t)) dt \\&= \left[ f_1(t) v(t) \right] _0^T - \kappa \int _0^T f_1(t) \theta (t) dt + (\kappa - \lambda ) \int _0^T f_1(t) v(t) dt \\&= \left[ f_1(t) v(t) \right] _0^T - f_1(0) \kappa \int _0^T e^{\lambda t} \theta (t) dt + (\kappa - \lambda ) \int _0^T f_1(t) v(t) dt \\&= \left[ f_1(t) v(t) \right] _0^T - f_1(0) \kappa \hat{\theta }_T(\lambda ) + (\kappa - \lambda ) \int _0^T f_1(t) v(t) dt, \end{aligned}$$

which proves the lemma.

The second change in the proof is due to the appearance of \(\theta \) in the generator of the process v. As in Schneider and Tavin (2018), let the function h be given by:

$$\begin{aligned} h(t,v) = {{\mathbb {E}}} \left[ \exp \left( i \frac{\rho }{\sigma } f_1(T) v(T) + \int _t^T q(s) v(s) ds \right) \right] . \end{aligned}$$

Now h satisfies the partial differential equation (PDE):

$$\begin{aligned} \frac{\partial h}{\partial t} (t,v) + \kappa (\theta (t) - v(t)) \frac{\partial h}{\partial v} (t,v) + \frac{1}{2} \sigma ^2 v(t) \frac{\partial ^2 h}{\partial v^2} (t,v) + q(t) v(t) h(t,v) = 0, \end{aligned}$$
(36)

with terminal condition:

$$\begin{aligned} h(T,v) = \exp \left( i \frac{\rho }{\sigma } f_1(T) v(T) \right) . \end{aligned}$$

Again, we know from Duffie et al. (2000) that h has affine form:

$$\begin{aligned} h(t,v) = \exp \left( A(t,T) v(t) + B(t,T) \right) , \end{aligned}$$
(37)

with \(A(T,T) = i \frac{\rho }{\sigma } f_1(T), B(T,T) = 0.\) Putting (37) in (36) gives:

$$\begin{aligned} B_t + A_t v + \kappa (\theta (t) - v) A + \frac{1}{2} \sigma ^2 v A^2 + q v = 0, \end{aligned}$$

and collecting the terms with and without v leads to the two ordinary differential equations (ODE):

$$\begin{aligned} A_t - \kappa A + \frac{1}{2} \sigma ^2 A^2 + q&= 0, \end{aligned}$$
(38)
$$\begin{aligned} B_t + \kappa \theta (t) A&= 0. \end{aligned}$$
(39)

This completes the proof of the proposition. \(\square \)

Note that \(\theta \) only appears in the second ODE (39), and that the closed-form expression previously given for A in Schneider and Tavin (2018) can still be used. Only the function B changes due to the time-dependence of \(\theta \).

Proof of Proposition 2

It is well known that the stochastic differential equation (SDE) given in Eq. (7) has a unique strong solution.

  1. (i)

    The drift function \(b(t, v_t) := \kappa (\theta (t) - v(t))\) in the SDE given in Eq. (6) is Lipschitz continuous w.r.t the second argument, i.e.

    $$\begin{aligned} | b(t, x) - b(t, y) | \le K | x - y |, \end{aligned}$$

    where we can choose \(K = \kappa \), since \(| b(t, x) - b(t, y) | = \kappa | x - y |\). Proposition 5.2.13 (Yamada and Watanabe) of Karatzas and Shreve (1988) thus guarantees the existence of a unique strong solution to (6) with continuous sample paths.

  2. (ii)

    The comparison result given in Proposition 5.2.18 of Karatzas and Shreve (1988) establishes \(v_t \ge \tilde{v}_t\) a.s. for all \(t \ge 0\) under the hypothesis that the drift function \(b(t, v_t)\) is continuous. Now, if \(\theta \) has a discontinuity at time \(t_1\), we know from applying this argument to the interval \([0, t_1[\) that \(\tilde{v}_{t} \le v_{t} \forall t \in [0, t_1[\) (a.s.). It then follows from the continuity of the sample paths that \(\tilde{v}_{t_1} \le v_{t_1}\) (a.s.), and we can apply the argument again to the interval \(]t_1, t_2[\) to obtain \(\tilde{v}_{t} \le v_{t} \forall t \in ]t_1, t_2[\) (a.s.). Since by assumption the set \(\mathcal {T}\) of times where \(\theta \) has discontinuities has no limit points, we can proceed in this manner to cover all of \(\mathbb {R}_0^+\).

  3. (iii)

    The Feller condition \(\sigma ^2 < 2 \kappa \theta _{{\textit{min}}}\) for \(\theta _{{\textit{min}}}\) implies the strict positivity a.s. of \(\tilde{v}\). The strict positivity of v itself therefore follows immediately from (ii).

\(\square \)

Transforms of the seasonality functions

In this appendix we show how the integral transforms \(\hat{\theta }_T(\lambda )\) can be calculated for the sinusoidal, sawtooth and triangle patterns. To the best of our knowledge, there are no closed-form expressions for the transforms of the exponential-sinusoidal and spiked patterns; we have also tried to solve these two integrals with the computer algebra software Maple, but without success.

The transform of the sinusoidal pattern is given by:

$$\begin{aligned} \hat{\theta }_T(\lambda )&= \frac{b e^{\lambda T}}{\lambda ^2+4\pi ^2}\left( 2 \pi \sin {\left( 2 \pi (T-t_0)\right) } + \lambda \cos {\left( 2 \pi (T-t_0)\right) } \right) \nonumber \\&\quad + \frac{b}{\lambda ^2+4\pi ^2}\left( 2 \pi \sin {\left( 2\pi t_0 \right) - \lambda \cos {\left( 2\pi t_0 \right) }} \right) + \frac{a}{\lambda }\left( e^{\lambda T}-1\right) . \end{aligned}$$
(40)

The transform of the sawtooth pattern is given by:

$$\begin{aligned} \hat{\theta }_T(\lambda )&= \frac{1}{\lambda }\left( a+b\left( \frac{1}{\lambda }-t_0 \right) \right) -\frac{e^{-\lambda T}}{\lambda }\left( a+b\left( T+\frac{1}{\lambda }-t_0 \right) \right) \nonumber \\&\quad -\frac{be^{\lambda t_0}}{\lambda } \left( \left\lfloor T-t_0 \right\rfloor e^{\lambda (T-t_0)} - \left( \sum ^{\left\lfloor T-t_0 \right\rfloor }_{k=1}{e^{\lambda k}}\right) \mathbb {1}_{\left\{ T \ge t_0\right\} } + \mathbb {1}_{\left\{ T<t_0\right\} }+e^{-\lambda t_0}-1 \right) , \end{aligned}$$
(41)

where \(\left\lfloor . \right\rfloor \) denotes the floor function, \(\mathbb {1}\) is the indicator function and, by convention, \(\sum ^{p}_{k=1}{e^{\lambda k}}=0\) if \(p<1\).

The transform of the triangle pattern is given by

$$\begin{aligned}&\hat{\theta }_T(\lambda ) = \frac{a}{\lambda }\left( e^{\lambda T}-1 \right) + \frac{b e^{\lambda t_0}}{\lambda }\left[ \left( z_2 + \left( \frac{2}{\lambda }e^{-\frac{\lambda }{2}}+e^{-\lambda t_0}(z_2-t_0) \right) \mathbb {1}_{\left\{ t_0> \frac{1}{2} \right\} } - e^{-\lambda t_0}(z_2-t_0)\mathbb {1}_{\left\{ t_0 \le \frac{1}{2} \right\} } \right) \right. \nonumber \\&\quad + \left( \left( \frac{2}{\lambda } e^{\frac{\lambda }{2}} + z_2 e^{\lambda }-z_1 \right) \sum ^{n-1}_{k=0}{e^{\lambda k}}+ e^{\lambda n}\left( \left( \frac{2}{\lambda }e^{\frac{\lambda }{2}}-z_3 e^{\lambda \alpha } \right) \mathbb {1}_{\left\{ \alpha > \frac{1}{2} \right\} } + z_3 e^{\lambda \alpha }\mathbb {1}_{\left\{ \alpha \le \frac{1}{2} \right\} } -z_1 \right) \right) \mathbb {1}_{\left\{ T \ge t_0\right\} } \nonumber \\&\quad + \left( e^{\lambda (T-t_0)}\left( z_2+T-t_0 \right) -z_2 \right) \mathbb {1}_{\left\{ T-t_0 \in [-\frac{1}{2},0[ \right\} } \nonumber \\&\quad - \left. \left( \frac{2}{\lambda }e^{-\frac{\lambda }{2}} + e^{\lambda (T-t_0)}\left( z_2+T-t_0 \right) + z_2 \right) \mathbb {1}_{\left\{ T-t_0 \in [-1,-\frac{1}{2}[ \right\} } \right] , \end{aligned}$$
(42)

with \(n = \left\lfloor T-t_0 \right\rfloor \), \(\alpha = T- t_0 - \left\lfloor T-t_0 \right\rfloor \), \(z_1 = \frac{1}{2} + \frac{1}{\lambda }\), \(z_2 = \frac{1}{2} - \frac{1}{\lambda }\) and \(z_3 = z_1 - \alpha \), and with the convention \(\sum ^{p}_{k=0}{e^{\lambda k}}=0\) if \(p<0\).

The proof of (40) is straightforward. The proofs of (41) and (42) are lengthy and omitted here for brevity, but available from the authors upon request.

Expected integrated variance

We provide the proofs of the two results given in Sect. 3.2 of the main manuscript. These results correspond to the expected integrated variance of a futures with maturity \(T_m\). For more details on these developments we refer to Bollerslev and Zhou (2002) and Kaeck and Alexander (2012).

The Cox–Ingersoll–Ross square-root process is described by the following SDE:

$$\begin{aligned} dv_t = \kappa (\theta - v_t) dt + \sigma \sqrt{v_t} dB_t, \quad v_0 > 0. \end{aligned}$$

We know that for \(T > t\):

$$\begin{aligned} E_t[v_T] = e^{-\kappa (T - t)} v_t + \theta \left( 1 - e^{-\kappa (T - t)} \right) . \end{aligned}$$
(43)

It follows, by exchanging the integrals, that:

$$\begin{aligned} E_t \left[ \int _t^T v_s ds \right]&= \int _t^T E_t[v_s] ds \\&= \int _t^T e^{-\kappa (s - t)} v_t + \theta \left( 1 - e^{-\kappa (s - t)} \right) ds \\&= \frac{1 - e^{-\kappa (T - t)}}{\kappa } v_t + \theta (T - t) - \theta \frac{1 - e^{-\kappa (T - t)}}{\kappa }. \end{aligned}$$

When the mean-reversion level is time-dependent, the dynamics and the expected variance at time \(T>t\) become:

$$\begin{aligned}&dv_t = \kappa (\theta (t) - v_t) dt + \sigma \sqrt{v_t} dB_t, \quad v_0 > 0.\nonumber \\&E_t[v_T] = e^{-\kappa (T - t)} v_t + \kappa e^{-\kappa T} \int _t^T e^{\kappa u} \theta (u) du. \end{aligned}$$
(44)

It follows, by exchanging the integrals twice, that:

$$\begin{aligned} E_t \left[ \int _t^T v_s ds \right]&= \int _t^T E_t[v_s] ds \\&= \int _t^T \left[ e^{-\kappa (s - t)} v_t + \kappa e^{-\kappa s} \int _t^s e^{\kappa u} \theta (u) du \right] ds \\&= \frac{1 - e^{-\kappa (T - t)}}{\kappa } v_t + \kappa \int _t^T \int _t^s e^{-\kappa s} e^{\kappa u} \theta (u) du \; ds \\&= \frac{1 - e^{-\kappa (T - t)}}{\kappa } v_t + \kappa \int _t^T e^{\kappa u} \theta (u) \int _u^T e^{-\kappa s} ds \; du \\&= \frac{1 - e^{-\kappa (T - t)}}{\kappa } v_t + \int _t^T \left( 1 - e^{-\kappa (T - u)} \right) \theta (u) du. \end{aligned}$$

It can be easily verified that this expression is equal to the expression obtained above when the function \(\theta \) is constant.

Let \(T_m\) be the maturity of a futures contract, with \(T_m> T > t\). We now calculate the expression (43) with an exponential damping factor:

$$\begin{aligned} E_t[e^{-\lambda (T_m - T)} v_T]&= e^{-\lambda (T_m - T)} E_t[v_T] \\&= e^{-\lambda (T_m - T)} \left[ e^{-\kappa (T - t)} v_t + \theta \left( 1 - e^{-\kappa (T - t)} \right) \right] . \end{aligned}$$

It follows, by exchanging the integrals, that:

$$\begin{aligned} E_t \left[ \int _t^T e^{-\lambda (T_m - s)} v_s ds \right]&= \int _t^T E_t[e^{-\lambda (T_m - s)} v_s] ds \\&= \int _t^T e^{-\lambda (T_m - s)} \left[ e^{-\kappa (s - t)} v_t + \theta \left( 1 - e^{-\kappa (s - t)} \right) \right] ds \\&= v_t \int _t^T e^{(\lambda - \kappa ) s - \lambda T_m + \kappa t} ds + \theta \int _t^T \left( e^{-\lambda (T_m - s)} - e^{(\lambda - \kappa ) s - \lambda T_m + \kappa t} \right) ds \\&= v_t \frac{1}{\lambda - \kappa } \left( e^{-\lambda (T_m - T) - \kappa (T - t)} - e^{-\lambda (T_m - t)} \right) \\&\quad + \theta \frac{1}{\lambda } \left( e^{-\lambda (T_m - T)} - e^{-\lambda (T_m - t)} \right) \\&\quad - \theta \frac{1}{\lambda - \kappa } \left( e^{-\lambda (T_m - T) - \kappa (T - t)} - e^{-\lambda (T_m - t)} \right) . \end{aligned}$$

We now calculate the expression (44) with an exponential damping factor:

$$\begin{aligned} E_t[e^{-\lambda (T_m - T)} v_T]&= e^{-\lambda (T_m - T)} E_t[v_T] \\&= e^{-\lambda (T_m - T)} \left[ e^{-\kappa (T - t)} v_t + \kappa e^{-\kappa T} \int _t^T e^{\kappa u} \theta (u) du \right] . \end{aligned}$$

It follows, by exchanging the integrals twice, that:

$$\begin{aligned} E_t \left[ \int _t^T e^{-\lambda (T_m - s)} v_s ds \right]&= \int _t^T E_t[e^{-\lambda (T_m - s)} v_s] ds \\&= \int _t^T e^{-\lambda (T_m - s)} \left[ e^{-\kappa (s - t)} v_t + \kappa e^{-\kappa s} \int _t^s e^{\kappa u} \theta (u) du \right] ds \\&= v_t \int _t^T e^{(\lambda - \kappa ) s - \lambda T_m + \kappa t} ds \\&\quad + \int _t^T \int _t^s \kappa e^{-\lambda T_m + \lambda s - \kappa s + \kappa u} \theta (u) du \; ds \\&= v_t \frac{1}{\lambda - \kappa } \left( e^{-\lambda (T_m - T) - \kappa (T - t)} - e^{-\lambda (T_m - t)} \right) \\&\quad + \int _t^T \int _u^T \kappa e^{(\lambda - \kappa ) s -\lambda T_m + \kappa u} \theta (u) ds \; du \\&= v_t \frac{1}{\lambda - \kappa } \left( e^{-\lambda (T_m - T) - \kappa (T - t)} - e^{-\lambda (T_m - t)} \right) \\&\quad + \int _t^T \frac{\kappa }{\lambda - \kappa } \theta (u) \left( e^{(\lambda - \kappa ) T -\lambda T_m + \kappa u} - e^{(\lambda - \kappa ) u -\lambda T_m + \kappa u} \right) du \\&= v_t \frac{1}{\lambda - \kappa } \left( e^{-\lambda (T_m - T) - \kappa (T - t)} - e^{-\lambda (T_m - t)} \right) \\&\quad + \int _t^T \frac{\kappa }{\lambda - \kappa } \theta (u) \left( e^{-\lambda (T_m - T) - \kappa (T - u)} - e^{-\lambda (T_m - u)} \right) du \\&= v_t \frac{1}{\lambda - \kappa } \left( e^{-\lambda (T_m - T) - \kappa (T - t)} - e^{-\lambda (T_m - t)} \right) \\&\quad + \frac{\kappa }{\lambda - \kappa } e^{-\lambda (T_m - T)} \int _t^T e^{-\kappa (T - u)} \theta (u) du \\&\quad - \frac{\kappa }{\lambda - \kappa } \int _t^T e^{-\lambda (T_m - u)} \theta (u) du. \end{aligned}$$

Note that \(\frac{-\kappa }{\lambda - \kappa } = 1 - \frac{\lambda }{\lambda - \kappa }\). When setting \(\lambda = 0\) and/or \(\theta (u) = \theta = {\textit{const}}\), we find the previous results obtained above.

Average volatilities

This appendix provides the average volatilities computed from our datasets, with a breakdown per contract and per calendar month.

In Table 8, we report, for each commodity, the average volatility of each futures series in our sample. In Table 9, we report, for each commodity, the average volatility per calendar month of the first futures contract. In addition, we identify the two calendar months with the highest volatilities.

Table 8 Average volatilities for each futures contract in our dataset
Table 9 Average volatilities, per calendar month, for the first futures series (c1) in our dataset

Estimated parameters

In this appendix we report, for the six seasonality patterns examined, the results of the maximum likelihood estimation procedure for Corn (Table 10), Cotton (Table 11), Soybeans (Table 12), Sugar (Table 13) and Wheat (Table 14).

Table 10 Estimated parameters for Corn
Table 11 Estimated parameters for Cotton
Table 12 Estimated parameters for Soybeans
Table 13 Estimated parameters for Sugar
Table 14 Estimated parameters for Wheat

Additional plots

See Fig. 5.

Results with alternative datasets

The results obtained with the alternative datasets are found in this appendix.

Fig. 5
figure 5

Results of the likelihood-ratio tests for each commodity: seasonal models versus the non-seasonal model. The bars represent the values taken by the D statistic of the test. The horizontal lines correspond to the significance thresholds where the (null) hypothesis of a non-seasonal model is rejected at 99% (green), 99.9% (red) and 99.99% (black) confidence levels. Tested seasonal models are numbered as: 1. sinusoidal, 2. exp-sinusoidal, 3. triangle, 4. sawtooth and 5. spiked. Upper left Corn. Upper right Cotton. Centre left Soybeans. Centre right Sugar. Bottom Wheat. (Color figure online)

Table 15 Summary of the results obtained with the considered models estimated with the alternative datasets: log-likelihood, AIC, BIC, Statistics D1 and D2 of the likelihood-ratio tests and their p-values, AIC differences \(\varDelta _{{\textit{aic}}}\) and Akaike weights \(\omega _i\)
Table 16 Model rankings for each commodity
Table 17 Summary of the results obtained with the alternative datasets for the volatility peaks prediction: the selected seasonal models, the majority vote of these models and the non-seasonal model

The estimation results obtained with the alternative datasets, described in Sect. 4.7 of the main text, are presented in Table 15. This table shows, for each commodity and model, the log-likelihood obtained, the AIC, the BIC, the value taken by the statistic D1 of the first likelihood ratio test, its p-value and the AIC difference and Akaike weight. In addition, it provides the log-likelihood obtained after estimating the nested versions of the models considered, obtained when setting \(\lambda =0\), and the value taken by the statistic D2 of the second likelihood ratio test and its p-value. We provide the model rankings for each commodity in Table 16. These rankings are based on the AIC differences described and discussed in Sect. 5 of the main text. In this table, we also provide the ranking of models obtained across commodities by computing the sum of their ranks. Table 17 presents the results of this volatility peak prediction in the alternative datasets. For each commodity, we provide the results obtained using the three seasonal models selected and the non-seasonal model. We also provide the results for a meta-predictor that combines the three seasonal predictors via a majority vote. For each model we provide: accuracy rate, RMSE ratio against a random walk and the Diebold-Mariano statistic, \(S_{{\textit{DM}}}\), for the model against the non-seasonal model.

  • Table 15: the likelihood scores are higher than those obtained with the initial datasets, except for cotton. Likelihood-ratio tests confirm the importance of modelling the seasonality and the Samuelson effect. For seasonality, the obtained p-values are slightly higher for corn, sugar and wheat, compared to the results with the initial datasets. However, for all commodities, the majority of confidence levels associated with the seasonality test remain above 99.99%, and all of them are above this value for the Samuelson effect.

  • Table 16: the ranking of models is now different, but the three best-ranked models remain the same for each commodity and in the overall ranking. The non-seasonal model is always ranked sixth, i.e. last. The best-performing model in the overall ranking now changes: while the sinusoidal pattern performed best with the initial datasets, the spiked pattern now performs best with the alternative datasets. This remark points toward model uncertainty with respect to the best seasonality pattern, and supports the selection of the top three models for the out-of-sample analysis. The in-sample performance spread between best and worst models, as measured by \(\varDelta _{{\textit{aic}}}\) of the sixth model, is tighter for corn and cotton. This indicates a slightly weaker evidence of seasonality for these commodities in the alternative datasets, in line with the above comments.

  • Table 17: when predicting volatility peaks during the 2017–2019 period, results are similar to those obtained with the initial datasets. The seasonal models always perform better than the non-seasonal benchmark. Accuracy rates near or above 80% are reached with seasonal models for corn, soybeans and wheat. For cotton, accuracy rates obtained with seasonal models are near 70%. The p-values of the Diebold-Mariano tests are lower with the alternative datasets, indicating a stronger evidence in favor of the model against a random walk predictor. In line with the initial dataset, sugar is the only commodity where the futures-based model developed in the paper fails to perform better than a random predictor.

Conclusion: in this robustness check, we clearly confirm the results already obtained with the initial datasets.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Schneider, L., Tavin, B. Seasonal volatility in agricultural markets: modelling and empirical investigations. Ann Oper Res 334, 7–58 (2024). https://doi.org/10.1007/s10479-021-04241-7

Download citation

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10479-021-04241-7

Keywords

JEL Classification

Navigation