1,788
Views
5
CrossRef citations to date
0
Altmetric
Features

A structural Heath–Jarrow–Morton framework for consistent intraday spot and futures electricity prices

ORCID Icon, &
Pages 347-357 | Received 29 Jan 2019, Accepted 25 Oct 2019, Published online: 02 Dec 2019

Abstract

The framework is based on stochastic processes with economic interpretations and is consistent with the initial forward price curve


© 2019 iStockphoto LP
1. Introduction

In recent years, the electricity intraday markets have gained increased popularity: the traded volume at the German/Austrian intraday market has grown by 30.3 percent from May 2016 to May 2018 (EPEX Citation2017Citation2018). In the same time period, European day-ahead trading increased by 10.4 percent. Furthermore, comparing the year 2017 with 2018, European power derivatives gained 19 percent in traded volume (EEX Citation2019). Since different electricity contracts exhibit different price behaviour such as spikes in the day-ahead spot but not in futures prices, it is a rising challenge in energy finance to define a single model that allows for a joint simulation of power prices at intraday spot and futures markets.

In this paper, we suggest a Heath–Jarrow–Morton framework for modelling electricity prices. The framework is consistent with the current forward term structure (i.e. the price forward curve) and we motivate each mathematical component by an economic interpretation. Furthermore, we discuss the computation of intraday spot and futures prices within this framework and we show how options on futures contracts can be priced. A new approach is the use of structural models for day-ahead spot price modelling within the Heath–Jarrow–Morton framework.

The starting point for the Heath–Jarrow–Morton (HJMFootnote1) approach for electricity prices is the fictitious forward price or forward kernel.Footnote2 The forward kernel ft(τ), tτ, is the price at time t of a forward contract delivering electricity instantly at time τ. It follows that the price at t of a futures contract delivering from τ1 to τ2 is the averaged forward kernel during the delivery period, i.e. (1) Ft(τ1,τ2)=1τ2τ1τ1τ2ft(u)du,tτ1.(1) In the HJM framework for interest rates the forward rate is modelled instead of the short rate (cf. Brigo and Mercurio Citation2006). Therefore, modelling the forward kernel instead of the day-ahead spot priceFootnote3 makes this an HJM approach for power prices. Furthermore, just like in the HJM framework for interest rates, the forward kernel itself is not a traded product at the market but through equation (1) its derivatives are.

Several models for the forward kernel ft(τ) have been introduced by Clewlow and Strickland (Citation1999), Hinz et al. (Citation2005), Koekebakker and Ollmar (Citation2005), Benth and Koekebakker (Citation2008), Kiesel et al. (Citation2009). They define the forward kernel dynamics driven by Brownian motions. However, since the day-ahead spot prices show spikes, these models have drawbacks. Therefore, there is a need for a forward kernel model that allows for spikes in relatively short delivery periods (day-ahead spot contracts) but smooths these out for longer delivery periods (futures contracts). The theoretical HJM framework of Benth et al. (Citation2019) introduces forward kernel dynamics driven by Brownian motions and pure jump Lévy processes. Their framework is similar to ours but touches on the differences between the real-world and the pricing measure. A difference to our approach is that we have motivated the modelling ingredients by economic arguments and allow for an easy transfer of day-ahead structural models to an HJM setting, which we also show in Section 3.

In the literature, the use of more than one probability measure has been challenged: Lyle and Elliott (Citation2009), Caldana et al. (Citation2017) assume a single probability measure, for example. This is supported by the fact that it is not clear which equivalent measure should be the pricing measure Q. Since electricity is a non-storable commodity and buy-and-hold strategy arguments are not valid, it is not clear what the relation between the price of electricity contracts and the money market account is (Bessembinder and Lemmon Citation2002). This also implies that the market is incomplete and that there are (possibly) infinitely many equivalent pricing measures. Again, this leaves the choice of pricing measure unclear.

We follow the idea of Caldana et al. (Citation2017) that the prices of day-ahead spot and futures contracts both should be computed by equation (Equation1). This actually sounds right intuitively since, for example in the German markets, day-ahead spot contracts are traded at least 12 hours before delivery. In other countries such as the USA the terminology is different: the day-ahead spot price is commonly referred to as the forward price (Longstaff and Wang Citation2004). Even in Europe, with the increasing popularity of the intraday markets, we observe a shift in terminology. Weron (Citation2014) remarks that the term spot is used more and more frequently for the real-time or intraday market. We will always explicitly state to which spot market we refer.

In this paper, we even propose to extend equation (Equation1) to the intraday market. Figure  gives an example of the development of the forward kernel ft(τ) and how it becomes observable at the German/Austrian market. First the forward kernel ft(τ) is only (partly) observable through EEX futures contracts. Then the Austrian EXAA and two German EPEX day-ahead spot auctions are held, after which the EPEX intraday spot market opens.

Figure 1. Observation structure of ft(τ) for the German electricity market and a fixed delivery time τ. The red marked lines and time points are the (indirect) observation moments. The lines with d−1 and d stand for the start of day d−1 and d.

Figure 1. Observation structure of ft(τ) for the German electricity market and a fixed delivery time τ. The red marked lines and time points are the (indirect) observation moments. The lines with d−1 and d stand for the start of day d−1 and d.

Furthermore, we show how the classical models described by Schwartz and Smith (Citation2000), Lucia and Schwartz (Citation2002) fit into our framework. We also show how other more general day-ahead spot price models can be used to fit into our model. A particular new example we introduce in this paper is to use structural models in the context of an HJM framework. We also apply our framework to the setting of multi-factor models.

This paper is structured as follows: section 2 introduces a model for the forward kernel based on the economic intuition that there are two driving components behind the forward kernel. The first component is the equilibrium of supply and demand at delivery time and the second is a general noise from partially informed traders or illiquidity at trading time t. Successively, in section 2.2 and section 2.3 the futures and option prices are computed, respectively. Section 3 contains the above explicitly mentioned examples for the market equilibrium process, while section 4 concludes.

2. HJM framework

In section 2.1, we will define a model for the forward kernel motivated by economic interpretations. Using this model in sections 2.2 and 2.3, we derive the prices of futures contracts and options on futures contracts, respectively. Section 2.4 gives an overview of the prices for different electricity contracts for the example of the German market.

2.1. Forward kernel

The forward kernel ft(τ) is the price at time t of a forward contract delivering 1 MW instantly at time τ. Throughout the rest of this paper, we interpret t as the trading time and τ as the delivery time.

For τ0 let Xτ={Xtτ;t0} and Y={Yt;t0} be two independent, a.s. càdlàg stochastic processes on the complete probability space (Ω,F,P) taking values in R and Rn, respectively. Furthermore, assume that the processes Xτ for each τ0 and Y are adapted to the filtration {Ft;t0}, which satisfies the usual conditions, i.e. {Ft;t0} is right-continuous and F0 contains all P-null sets. The filtration generated by Y and Xτ augmented by all P-null sets automatically fulfils these conditions. Finally, let g:RnR be a function such that g(Yt) is real-valued stochastic process.

We have two strong economic interpretations for these two stochastic processes: we interpret the n-dimensional process Yt as the randomness or the state of the market, where each component of Yt stands for a (random) facet of the market, e.g. demand, load, or weather predictions. The function g maps the state of the market state Yt to its corresponding price. Combining the fact that our inspiration came from the class of structural models for day-ahead spot price modelling and the fact that it gives the basic structure to the forward kernel, we call the pair (g,Yt) the structural component. Often we will also only call Yt the structural component.

The process Xtτ is called the market noise because it accounts for the incomplete market information of all market participants and illiquidity of the market. An example of incomplete market information is the uncertainty of weather predictions: nobody knows with complete certainty about the future weather or temperature. With these interpretations, we define the forward kernel:

Definition 2.1

(Forward kernel)

We define the forward kernel at trading time t and delivery time τ as ft(τ):=XtτEg(Yτ)|Ft, where Xtτ is the market noise at trading time t for the delivery time τ and (g,Yτ) the structural component at delivery time τ.

We use the notation Xtτ to emphasize that the market noise is a stochastic process in the trading time t but can depend on the delivery time τ, whereas the structural component Yτ only depends on delivery time. Economically, this makes sense since the imbalance of supply and demand at delivery time τ determines the price independent of the trading time t at which we predict this imbalance. However, the market noise is the disturbance of this prediction originating from market participants with incomplete market information, which intuitively depends on both the trading time t and the delivery time τ they are trying to predict.Footnote4 Although we call Xtτ the market noise, it can also be interpreted as a measure transformation (or Radon–Nikodym derivative, see Remark 2.7) or as a general additional component that introduces an additional degree of freedom in the modelling process.

Assumption 2.2

(Market noise)

The process Xτ={Xtτ;t0} with its interpretation as market noise for delivery time τ is defined as multiplicative stochastic noise. We assume that it is an a.s. positive càdlàg martingale with expectation one, i.e. EXtτ=1 for all τt0. In particular, we assume that the initial value X0τ=1 a.s. for all τ0.

Assumption 2.3

(Structural component)

We assume that Y={Yt;t0} is a Rn-valued càdlàg stochastic process. In particular, we assume that the initial value equals Y0=y0Rn a.s. such that g(y0)=f0(0), where f0(τ) is the initial price forward curve (PFC) for delivery time τ, which we assume to be known (cf. Remark 2.4). Furthermore, as a technical assumption we need that E|g(Yt)|< for all t0. Finally, although we assume that g(Yt) can take all values in R, including negative values, we assume that its expectation Eg(Yt)>0 is strictly positive. The economic interpretation behind this assumption is that we do not expect negative forward prices to occur.

With these assumptions, the sign of the forward kernel is uniquely determined by the structural component Y and the process Xτ cannot influence it. Furthermore, the expectation Eft(τ) is fully determined by the structural component Yτ and independent of trading time t (cf. Lemma 2.6).

Remark 2.4

(Initial price forward curve f0(τ))

In this framework, the initial price forward curve (PFC), denoted by f0(τ), plays an important role: it determines the expectation of the forward kernel ft(τ). There are many studies that describe how one can construct a PFC from market prices such as Caldana et al. (Citation2017), Kiesel et al. (Citation2018), for example. In practice, every energy utility has an in-house PFC. In the following, we will therefore assume that the PFC is known.

As discussed by Benth and Paraschiv (Citation2018), another interesting possibility is to use a Musiela parametrisation for the forward kernel. This parametrisation is given by the bijective mapping (t,τ)(t,u):=(t,τt) of R2 on itself. In their work, Benth and Paraschiv (Citation2018) propose a spatio-temporal random field model in the context of an HJM framework under the Musiela parametrisation, where they call the time to maturity u the spatial component. They disentangled the temporal from spatial effects on the dynamics of forward prices and found that the temporal noise was non-Gaussian. In our context, we could directly use the Musiela parametrisation by substituting τ=u+t.

Theorem 2.5

For fixed τ0, the forward kernel process f(τ):={ft(τ);t0} is an adapted stochastic process. Furthermore, f(τ) is a.s. càdlàg.

Proof.

By definition, f(τ) is a stochastic process. Moreover, since we assumed Xtτ to be Ft-measurable and since the conditional expectation Ztτ:=E[g(Yτ)|Ft] is always Ft-measurable, the Ft-measurability of ft(τ) follows immediately. Because the filtration satisfies the usual conditions, Ztτ has a càdlàg modification (Karatzas and Shreve Citation1998, Chapter 1, Theorem 3.13). Since the conditional expectation Ztτ is uniquely defined up to null sets, we can choose this modification and the result follows by the assumption that Xtτ is càdlàg.

Since we assume that Xτ and Y both a.s. start at a deterministic value, we assume without loss of generality that F0 is generated by Ω and all P-null sets. This in particular implies that Eg(Yτ)=E[g(Yτ)|F0], a fact we will exploit in the next lemma.

Lemma 2.6

For fixed τ0 the forward kernel process f(τ):={ft(τ);t0} is a martingale. Furthermore, its expectation is given by Eft(τ)=Eg(Yτ)=f0(τ) for all 0tτ.

Proof.

The product of two independent martingales clearly is a martingale. Furthermore, it follows immediately from Assumption 2.2 and 2.3 that Eft(τ)=EXtτEEg(Yτ)|Ft=Eg(Yτ)=X0τEg(Yτ)|F0=f0(τ) by the independence of Xtτ and Yt.

Lemma 2.6 also imposes a condition for the expectation Eg(Yτ) of the structural component, which can be used to calibrate the structural component Y and function g after the PFC f0(τ) has been determined. If one wants to obtain a model that is consistent with an existing PFC f0(τ), one needs to choose and calibrate g and Y such that Eg(Yτ)=f0(τ).

Remark 2.7

(Risk-neutral measure)

In the previous discussion, we considered the measure space (Ω,F,P) equipped with the measure P. However, as seen in Lemma 2.6 and will be seen in Lemma 2.12, the forward kernel and its induced futures prices are martingales under the measure P. This is an argument in favour of viewing P as the risk-neutral measure or pricing measure in this framework, making this framework especially suitable for pricing derivatives. However, if one wants to simulate market prices through this framework, one needs to derive the dynamics of the market prices under the real-world measure which can be done by a suitable measure change, as is discussed by Benth et al. (Citation2019). Under the real-world measure the expectation is then no longer constant in trading time t, which is a phenomenon that is supported by plenty of empirical evidence. In particular in the case of the continuous intraday market this is studied by the findings of Kiesel and Paraschiv (Citation2017).

Another view on this framework is achieved by defining the following measure: the τ-forward measure Pτ defined by its Radon–Nikodym derivative (2) dPτdP|Ft=1Xtτ(2) could be used for this purpose. Using the τ-forward measure and Bayes' theorem for conditional expectations, we can rewrite Definition 2.1 ft(τ)=Xtτ EP[g(Yτ)|Ft]=EPτ[Xττg(Yτ)|Ft]. The latter term can be defined as St:=Xttg(Yt), which yields a general spot price model. This is another argument in favour of viewing P and Pτ as equivalent pricing measures (where Pτ is viewed as a forward pricing measure). As discussed in the introduction electricity markets are incomplete and therefore it is possible that multiple equivalent pricing measures exist. In this setting, the choice of the stochastic process Xtτ can be viewed as the choice of delivery time specific pricing measure Pτ in light of equation (Equation2). If the noise Xt:=Xtτ is chosen to be independent of the delivery time τ, so is the τ-forward measure Pτ.

2.2. Futures contracts

As discussed in section 1, the forward kernel can be used to compute the price of futures contracts. In the following, we assume the interest rate to equal r = 0 for notational convenience. Of course, when one assumes r0, discounting has to be taken into account. In Remark 2.13, we have some notes on how to change our framework to include discounting. Furthermore, we assume that all prices are normalized, meaning that we assume all prices to be in Euro/MWh as usual.

Definition 2.8

(Futures contract price)

For 0tτ1<τ2, we call Ft(τ1,τ2):=1τ2τ1τ1τ2ft(u)du the price of a futures contract at time t delivering 1 MW continuously from τ1 to τ2.

Since we denote all prices in Euro/MWh, the price that one pays at time t when one buys a futures contract delivering 1 MW from τ1 to τ2 is given by (τ2τ1)Ft(τ1,τ2), where we assume that τ2τ1 is measured in hours.

Example 2.9

(Day-ahead spot price)

We compute the day-ahead spot price as a futures contract. It is auctioned at day d−1 at hour a and delivered at day d from h:00 until (h+1):00 o'clock, i.e. S(d,h):=Ftd1atdh,tdh+1. Here tdh denotes the time at day d and hour h.

The next theorem shows that the framework is consistent with cascading.Footnote5 It also shows that there are no arbitrage opportunities in the sense that the cost of a futures contract delivering for 1 year is the same as the cost of its four quarters, for example.

Proposition 2.10

(Consistency of cascading)

Let 0τ0<τ1<τ2<<τn be delivery times, then we have (τnτ0)Ft(τ0,τn)=i=1n(τiτi1)Ft(τi1,τi) for all t0.

Proof.

This follows directly from Definition 2.8 and the countable additivity of the Lebesgue integral.

Lemma 2.11

Fix 0t<τ. If uft(u) is almost surely continuous on (τϵ,τ] for some ϵ>0, then we have limsτFt(s,τ)=ft(τ) almost surely.

Proof.

We compute limsτFt(s,τ)=limsτsτft(u)dulimsττs=limsτft(s)1=ft(τ) where we used L'Hôpital's rule for the second equality.

The previous lemma shows that the price of a futures contract delivering for just an instant equals the forward kernel. This supports the naming of the quantity ft(τ) as forward kernel.

Lemma 2.12

Assume that the price forward curve τf0(τ) is continuous. Then the futures price process F(τ1,τ2):={Ft(τ1,τ2);t0} is a martingale. Its expectation is given by EFt(τ1,τ2)=F0(τ1,τ2)=1τ2τ1τ1τ2f0(u)du for all 0tτ1<τ2.

Proof.

Since the price forward curve is continuous, it is bounded on any compact set, in particular intervals of the form [τ1,τ2], and therefore integrable on compacts. Direct computation with Fubini's Theorem shows that for 0t<s E[Fs(τ1,τ2)|Ft]=E1τ2τ1τ1τ2fs(u)du|Ft=1τ2τ1τ1τ2E[fs(u)|Ft]du, where the latter exists and therefore all integrals exist. Combination with Lemma 2.6 now proves the theorem.

Remark 2.13

(r0)

If we assume that r0, the futures price depends on the settlement date. There are two possibilities: settlement takes place either through continuous paymentsFootnote6 during the delivery period or at once at the end of the delivery period. If dt(τ) denotes the discount factor of a future payment at time τ to an earlier time t, the price of a futures contract is given by Ft(τ1,τ2)=1τ1τ2dt(u)duτ1τ2dt(u)ft(u)du for continuous settlement and by Ft(τ1,τ2)=1(τ2τ1)dt(τ2)τ1τ2dt(u)ft(u)du for settlement at the end of delivery.

2.3. Options on futures contracts

In this section, we assume that the market noise is given by a geometric Brownian motion (GBM) without drift, i.e. dXtτ=XtτΣ(t,τ)TdWt where Σ(t,τ) is a deterministic m-dimensional volatility vector and Wt is an m-dimensional Brownian motion. The strong solution of Xtτ is given by Xtτ=exp0tΣ(u,τ)TdWu120tΣ(u,τ)TΣ(u,τ)du. In this case, Xtτ satisfies Assumption 2.2 if Σ(u,τ) is square integrable in u. But this is already a requirement for the stochastic integral to be defined.

Example 2.14

(Hull–White market noise dynamics)

A possible choice for Σ is a two-factor forward dynamic similar to Kiesel et al. (Citation2009), which is also discussed in a geometric setting by Fanelli and Schmeck (Citation2018) for pricing options on futures. This volatility structure is extended by Latini et al. (Citation2018) in an additive setting. They discussed a two-factor volatility structure comparable to the two-factor Hull–White model for interest rate modelling (Brigo and Mercurio Citation2006, Section 4.2.5). It is given by Σ(t,τ)T:=(eκ(τt)σ1,σ2(τ)), where σ1>0 is the additional short-term volatility, κ>0 is the rate of decay of the short-term volatility, and σ2(τ)>0 is the long-term volatility at delivery time τ. A convenient choice for σ2 is a piecewise constant function, being constant on delivery periods of tradable futures contracts. An advantage of this choice is that we can use the calibration methods for Xtτ as discussed by Kiesel et al. (Citation2009), Latini et al. (Citation2018), Fanelli and Schmeck (Citation2018).

Throughout the rest of this section, we assume that the conditional expectation of the structural component decomposes into an affine structure:

Definition 2.15

(Affine structural component decomposition)

We say the structural component (g,Yt) allows for the affine structural component decomposition, if there exist deterministic functions (t,τ)AtτRn×n and (t,τ)BtτRn such that the following decomposition holds: (3) E[g(Yτ)|Ft]=g(AtτYt+Btτ)(3) a.s. for all τt0.

This decomposition can be motivated by the fact that our best guess at time t for the state of market Yτ at time τ is an affine transformation of the current state of the market Yt. This is also the main idea behind Kalman filtering, for example. If the decomposition holds, this merely states that this best guess should hold under the transformation g, which transforms the market state into a price.

It follows immediately that the forward kernel is given by (4) ft(τ)=Xtτg(AtτYt+Btτ),(4) when the affine structural component decomposition assumption is satisfied. Furthermore, the futures price of Definition 2.8 can be rewritten as Ft(τ1,τ2)=1τ2τ1τ1τ2Xtug(AtuYt+Btu)du for all 0tτ1<τ2. As immediate consequences we obtain:

Lemma 2.16

If (g,Yt) allows for the affine structural component decomposition, then E[g(Yτ)|Ft]=E[g(Yτ)|Yt].

Lemma 2.17

Under assumption of the decomposition of Definition 2.15, the forward kernel conditioned on Yt is lognormally distributed, i.e. (ft(τ)|Yt=y)LNln[g(Atτy+Btτ)],0tΣ(u,τ)TΣ(u,τ)du.

Proof.

Using equation (Equation4), we compute Pft(τ)x|Yt=y=PXtτg(Atτy+Btτ)x, which shows the result since XtτLN(0,0tΣ(u,τ)TΣ(u,τ)du).

Theorem 2.18

If (g,Yt) allows for the affine structural component decomposition, then the first two moments of the futures price Ft(τ1,τ2) exist and are given by E[Ft(τ1,τ2)|Yt=y]=1τ2τ1τ1τ2g(Atuy+Btu)du and E[Ft(τ1,τ2)2|Yt=y]=1(τ2τ1)2τ1τ2τ1τ2wtX(u,s)wtY(u,s,y)duds, where (5) wtX(u,s):=exp120tΣ(v,u)TΣ(v,s)+Σ(v,s)TΣ(v,u)dv(5) and (6) wtY(u,s,y):=g(Atuy+Btu)g(Atsy+Bts).(6)

Proof.

We see that the expectation follows immediately by an Fubini argument combined with the fact that EXtτ=1 for all τ0. Applying Fubini twice, we find E[Ft(τ1,τ2)2|Yt=y]=τ1τ2τ1τ2E[XtuXts]E[YuYs|Yt=y]duds(τ2τ1)2, where it is easy to verify that the expectations equal E[XtuXts]=wtX(u,s) and E[YuYs|Yt=y]=wtY(u,s,y) using equation (Equation3).

Corollary 2.19

If (g,Yt) allows for the affine structural component decomposition, then the conditional variance of the futures price Ft(τ1,τ2) is given by Var[Ft(τ1,τ2)|Yt=y]=1(τ2τ1)2τ1τ2τ1τ2wtX(u,s)1wtY(u,s,y)duds, where wX and wY are given by equation (Equation5) and equation (Equation6), respectively.

Proof.

We directly compute Var[Ft(τ1,τ2)|Yt=y]=E[Ft(τ1,τ2)2|Yt=y]E[Ft(τ1,τ2)|Yt=y]2. Using Theorem 2.18 the first term is immediately given and the second term can be computed using Fubini's Theorem E[Ft(τ1,τ2)|Yt=y]2=1τ2τ1τ1τ2g(Atuy+Btu)du2=1(τ2τ1)2τ1τ2τ1τ2g(Atuy+Btu)g(Atsy+Bts)duds, from which the result follows.

Remark 2.20

(Lognormal approximation)

Similar to the discrete approach used by Kiesel et al. (Citation2009), we have that the futures price is an integral of lognormally distributed variables, which can be approximated by a lognormal random variable with the same mean and standard deviation. Since there is no simple expression for the convolution of lognormal distributions, this approximation of the integral (or sum) of lognormal random variables is widely used in finance, e.g. in the context of LIBOR market models by Brigo and Mercurio (Citation2006). An analysis of this approximation, also with regard to Asian options (which may be compared to an option on a futures with delivery period), is found in Dufresne (Citation2004), for example.

Assumption 2.21

(Lognormal approximation)

Assume that the first two moments of the futures price Ft(τ1,τ2) exist. Justified by Remark 2.20, we then assume that (Ft(τ1,τ2)|Yt=y)(F~t(τ1,τ2)|Yt=y)LNμF(y),σF2(y), i.e. the futures price is approximately lognormally distributed.

As stated in Remark 2.20, we need that the first two moments of F and F~ match, which is resolved by the following lemma:

Lemma 2.22

If (g,Yt) allows for the affine structural component decomposition and Assumption 2.21 holds, then the mean and standard deviation of the lognormal distribution are given by μF(y):=lnτ1τ2g(Atuy+Btu)duln(τ2τ1)12σF2(y) and σF2(y):=ln1+τ1τ2τ1τ2wtX(u,s)1wtY(u,s,y)dudsτ1τ2τ1τ2wtY(u,s,y)duds, where wX and wY are given by equation (Equation5) and equation (Equation6), respectively.

Proof.

For a lognormal random variable ZLN(m,s), the expectation and variance are given by EZ=exp(m+s2/2) and VarZ=(EZ)2(exp(s2)1). Using Theorem 2.18 and Corollary 2.19, the result is found by inverting these equations.

Using this lemma, we can compute the price (conditioned on Yt) of call (and put) options on futures contracts by the Black-Scholes formula. A call option with strike price K and maturity T<τ1 has a pay-off equal to (7) (τ2τ1)(FT(τ1,τ2)K)+.(7) Recall that, as stated in section 2.2, the price one has to pay for a futures contract at time T equals (τ2τ1)FT(τ1,τ2), since we consider normalized prices.

Proposition 2.23

(Conditional call option price)

Assume that (g,Yt) allows for the affine structural component decomposition and let assumption 2.21 hold. Denote the futures price at maturity by F:=FT(τ1,τ2). Let μF and σF be given by Lemma 2.22. The price of a call option at t = 0 with pay-off given by (Equation7) conditioned on YT=y equals C0(T,K,τ1,τ2;y)=Φ(δ1(y))τ1τ2g(ATuy+BTu)du(τ2τ1)KΦ(δ2(y)), where Φ is the cumulative distribution function of the standard normal distribution, and with μF(y) and σF(y) as defined in lemma 2.22 the auxiliary variables δ1(y) and δ2(y) are given by δ2(y):=μF(y)lnKσF(y), and δ1(y):=δ2(y)+σF(y).

Proof.

Using the discounted conditional expectation of the pay-off given in (Equation7) yields C0(T,K,τ1,τ2;y)τ2τ1=E(FK)+|YT=y=E[F𝟙{FK}|YT=y]KP(FK|YT=y), where noting that we have (F|YT=y)LN(μF(y),σF2(y)), yields the result by direct computation.

As an immediate consequence we have:

Corollary 2.24

(Call option price)

Assume that (g,Yt) allows for the affine structural component decomposition and let assumption 2.21 hold. Let μF and σF be given by lemma 2.22. The price of a call option at t = 0 with pay-off given by (Equation7) equals (8) C0(T,K,τ1,τ2)=EC0(T,K,τ1,τ2;YT),(8) where the conditional call option price C0(T,K,τ1,τ2;y) is given in proposition 2.23.

When the distribution of YT is specified, the price of a call option given by equation (Equation8) might be evaluated analytically, numerically, or through simulative methods such as Monte Carlo estimation. Alternatively, with further assumptions on the distribution of YT this expectation could also be approximated differently.

2.4. Model representation of exchange traded products

In this section, we give an overview of the prices of several different electricity contracts in this HJM framework. Although there is not a single unique quoted continuous electricity price we regard Ft(τ1,τ2) as the true fair price for the delivery period from τ1 to τ2 at any trading time t.

Futures price. The price of a futures contract at time t delivering 1 MW continuously from τ1 to τ2 is given by Definition 2.8 and denoted by Ft(τ1,τ2).

Options on futures. In the setting of section 2.3 the price of call and put options on futures contracts can be computed by the Black–Scholes formula as given by proposition 2.23 or corollary 2.24.

Day-ahead spot prices. The day-ahead spot price equals the futures price within this framework as discussed in example 2.9.

ID1 and ID3 price. The ID1 and ID3 price indices on the German intraday market are given as the 1- and 3-hour volume-weighted average of all intraday trades before delivery. Therefore, we suggest the IDn price for the delivery period from τ1 to τ2 to equal IDn(τ1,τ2):=22n1τ1nτ10.5Fu(τ1,τ2)du, where n=1 or n = 3 and the subtraction of τ1 is meant in hours.

3. Examples of the structural component

First we show how two classical day-ahead spot price models can be used in this HJM framework. Then we also introduce a structural model approach as well as a multi-factor model approach for Y. To make the choice of an explicit model easier in this framework, we introduce the relative structural component, which can be used to set the initial price forward curve (PFC) to an existing one:

Definition 3.1

(Relative structural component)

The additive mean-normalized version of g(Yτ) Iτa:=g(Yτ)Eg(Yτ) is called the additive relative structural component and its multiplicative mean-normalized version Iτm:=g(Yτ)Eg(Yτ) is called the multiplicative relative structural component.

We directly obtain from these definitions:

Corollary 3.2

The relative structural components Ia and Im are stochastic processes with constant expectation EIτa=0 and EIτm=1 for all τ0.

Corollary 3.3

(Arithmetic PFC decomposition)

For a given initial price forward curve f0(τ), the forward kernel equals ft(τ)=Xtτf0(τ)+E[Iτa|Ft], where Iτa is the arithmetic relative structural component given in Definition 3.1.

Proof.

Define an extended structural component Y~τ=(Yτ,f0(τ))Rn+1, where f0(τ) is the constructed PFC, and another function g~(y,x)=x+g(y)Eg(y). It is clear that Y~ and g~ satisfy Assumption 2.3. It follows immediately that g~(Y~(τ))=f0(τ)+Iτa, which proves the result.

Corollary 3.4

(Geometric PFC decomposition)

For a given initial price forward curve f0(τ) the forward kernel equals ft(τ)=f0(τ)XtτE[Iτm|Ft], where Iτm is the geometric relative structural component given in Definition 3.1.

Proof.

The result can be shown analogously to the proof of Corollary 3.3.

The interpretation of these decompositions is that today's price forward curve is the expectation of the forward kernel that is being disturbed by the market noise Xtτ in trading time t and by the structural component in delivery time τ. Depending on the choice of the structural component (g,Yτ) this disturbance can be chosen to be multiplicatively in case of the geometric PFC decomposition or additively in case of the arithmetic PFC decomposition.

3.1. Classical spot models

We can use classical day-ahead spot price models in our framework by choosing g(Yt)=St, where St denotes the spot price at time t. Two examples of spot price models that we explicitly compute in this section are the spot price models by Schwartz and Smith (Citation2000) and Lucia and Schwartz (Citation2002).

For both examples, we need the same structural component and therefore we assume in this section that it is given by Yt=(yt1,yt2)R2. The first process is an Ornstein–Uhlenbeck process, i.e. (9) dyt1=κyt1dt+σ1dWt1,y01=0,(9) and the second (10) yt2=μ2t+σ2ρWt1+σ21ρ2Wt2(10) is a (correlated) Brownian motion with drift. The standard one-dimensional Brownian motions W1 and W2 are assumed to be independent. The parameters κ>0, σ1,σ2>0, 1ρ1, and μR are assumed to be real-valued.

Example 3.5

(Schwartz and Smith)

Schwartz and Smith (Citation2000) define the day-ahead spot price using the function g(y1,y2)=ey1+y2, i.e. they chose the price to equal St:=g(Yt)=exp(yτ1+yτ2). In the HJM framework, this transfers to the following forward kernel: ft(τ)=XtτE[eyτ1+yτ2|Ft], where we do not assume any extra conditions on Xτ apart from assumption 2.2.

In this setting we can explicitly compute the conditional expectation on g(Yτ) and we find lnE[eyτ1+yτ2|Ft]=eκ(τt)yt1+yt2+μ2+σ222(τt)+σ124κ1e2κ(τt)+ρσ1σ2κ1eκ(τt). This implies that this model for g and Yτ satisfies the affine structural component decomposition of definition 2.15. The coefficient Atτ of the decomposition is given by (11) Atτ=eκ(τt)001(11) and Btτ can be chosen to be any vector in R2 such that lng(Btτ)=μ2+σ222(τt)+σ124κ1e2κ(τt)+ρσ1σ2κ1eκ(τt) holds.

Since the function g is multiplicative in nature, the geometric PFC decomposition, corollary 3.4, is especially suited for this model. The conditional expectation of the multiplicative relative structural component is given by lnE[Iτm|Ft]=lng(AtτYt+Btτ)Eg(Yτ)=eκ(τt)yt1+yt2μ2+σ222t+σ12e2κτ4κ1e2κt+ρσ1σ2eκτκ1eκt, and the forward kernel decomposes to ft(τ)=f0(τ)Xtτeeκ(τt)yt1+yt2μ2+σ222t+σ12e2κτ4κ1e2κt+ρσ1σ2eκτκ1eκt, where any initial price forward curve f0(τ) can be used.

Example 3.6

(Lucia and Schwartz)

Lucia and Schwartz (Citation2002) discuss four different models. Here, we highlight the arithmetic two factor model for the spot price. This model is defined by the function g(y1,y2)=y1+y2 and the forward kernel equals ft(τ)=XtτE[yτ1+yτ2|Ft]. Again, apart from assumption 2.2 the process Xτ can be chosen freely.

The conditional expectation can easily be computed as E[yτ1+yτ2|Ft]=eκ(τt)yt1+yt2+μ2(τt) and the affine structural component decomposition of definition 2.15 follows immediately with the coefficient Atτ given by equation (Equation11) and Btτ can be any vector in R2 such that g(Btτ)=μ2(τt).

The additive nature of g makes the arithmetic PFC decomposition, corollary 3.3, the best suited candidate for this model. It follows that ft(τ)=Xtτf0(τ)+eκ(τt)yt1+yt2μ2t for any initial price forward curve f0(τ). We continue the study of this type of forward kernel in section 3.3 with a factor model approach.

In the rest of this section, we will give two further examples of the structural component Y. The first is based on the structural model approach for day-ahead spot prices and the other uses multi-factor models, which are the sum of Ornstein–Uhlenbeck type processes, cf. Benth et al. (Citation2008).

3.2. Structural model approach

We will use the HJM framework to model the structural component by a structural model approach: a spot price modelling technique started by Barlow (Citation2002) which uses the idea of equilibrium of supply and demand to derive a spot price. In contrast to reduced-form models which need to implement a jump component to model spikes, structural models use a non-linear transformation of a (Gaussian) diffusion process to reach this goal. This method has been developed further by many authors, e.g. Aïd et al. (Citation2009), Wagner (Citation2014).

For the real-valued demand process D we use a Gaussian Ornstein–Uhlenbeck process, i.e. dDt=λDtdt+σdWt,D0=0. We choose the structural component to equal Yt:=β(t)Dt, where β(t) is a real-valued deterministic function. Furthermore, we define the function g as follows: g(y1,y2)=γ+y1sinh(αy2)=γ+y1eαy2eαy22 for α>0 and γ>0. Through the first coordinate of Yt, i.e. β(t), we associate y1 with the evolution of time and y2 through the second coordinate of Yt, namely Dt, with the demand. Therefore, g(Yt) represents the price at time t for a load of Dt through the merit order curve.

Remark 3.7

(Extension of the model)

It might be convenient to use more realistic models, such as described by Wagner (Citation2014). This is an extension of the OU model, where stochastic processes for wind and solar infeed are subtracted from the demand process D. This difference is seen to model power prices even more accurately. It can easily be seen that the structural component Yt and function g can be extended for these processes.

Using the auxiliary function ν2(s):=(σ2/2λ)(1e2λs) the affine structural component decomposition of Definition 2.15 can be derived from the following theorem:

Theorem 3.8

The conditional expectation of the structural component is given by E[g(Yτ)|Ft]=γ+β(τ)eα22ν2(τt)sinh(αeλ(τt)Dt) for all τt0.

Proof.

For Gaussian OU processes, we have the following decomposition: Dτ=deλ(τt)Dt+ν(τt)ε,εN(0,1). Now, exploiting the decomposition and plugging it into the definition, we get E[g(Yτ)|Ft]=γ+β(τ)Esinh(αDτ)|Ft=γ+β(τ)sinh(αeλ(τt)Dt)Eeαν(τt)ε=γ+β(τ)eα22ν2(τt)sinh(αeλ(τt)Dt) by symmetry of the normal distribution.

Corollary 3.9

(Affine structural component decomposition)

With coefficients given by Atτ=β(τ)β(t)eα22ν2(τt)00αeλ(τt) and Btτ=0R2 the affine structural component decomposition of definition 2.15 holds.

By Theorem 3.8, it follows immediately by taking t = 0 that the expectation Eg(Yτ)=γ>0 for all τ0. Therefore we can use both the additive and geometric PFC decomposition, i.e. corollary 3.3 and corollary 3.4, respectively. In the additive case the forward kernel equals ft(τ)=Xtτf0(τ)+g(AtτYt)γ=Xtτf0(τ)+β(τ)eα22ν2(τt)sinh(αeλ(τt)Dt), whereas in the multiplicative case it equals ft(τ)=f0(τ)Xtτg(AtτYt)γ=f0(τ)Xtτ1+β(τ)γeα22ν2(τt)sinh(αeλ(τt)Dt). For both decompositions, any initial price forward kernel can be used.

3.3. Arithmetic factor model approach

In this section, we use an arithmetic factor model approach for the structural component in the HJM framework. More precisely, the structural component is given by an n-dimensional Lévy driven Ornstein–Uhlenbeck process dYt=ΛYtdt+dLt,Y0=y0, where Λ=diag(λ1,λ2,,λn)Rn×n with λ1,λ2,,λn>0 and L is an n-dimensional Lévy process. For more information on this type of moving average process, we refer the interested reader to  Wolfe (Citation1982), Jurek and Vervaat (Citation1983), Barndorff-Nielsen and Shephard (Citation2001), Applebaum (Citation2009), Sato (Citation2013). For an application of OU processes in the form of multi-factor models for energy prices, we refer to Benth et al. (Citation2008).

The function g is given by the summation of all the coefficients, i.e. we assume that g(y)=i=1nyi. If Yt satisfies assumption 2.3, we can explicitly compute the conditional expectation:

Theorem 3.10

The conditional expectation of the structural component is given by E[g(Yτ)|Ft]=geΛ(τt)Yt+EtτeΛ(τu)dLu for all τt0.

Proof.

For general OU processes, the same decomposition holds as was used in the proof of theorem 3.8, i.e. Yτ=eΛ(τt)Yt+tτeΛ(τu)dLu. Noting that the first term is Ft-measurable and the second term is independent of Ft yields the result, as the sum g and E commute.

As a direct consequence we obtain:

Corollary 3.11

(Affine structural component decomposition)

With coefficients given by Atτ=eΛ(τt) and Btτ=EtτeΛ(τu)dLu the affine structural component decomposition of definition 2.15 holds.

Due to the additive structure of g the logical PFC decomposition to choose in this setting is the arithmetic one, i.e. corollary 3.3. From theorem 3.10 we find that the expectation is given by Eg(Yτ)=geΛτy0+E0τeΛ(τu)dLu. It follows that the forward kernel is given by ft(τ)=XtτE0teΛ(τu)dLuf0(τ)+gE0teΛ(τu)dLueΛ(τt)YteΛτy0E0teΛ(τu)dLu, where f0(τ) can be any initial price forward curve.

4. Conclusion

In this paper, we have developed a unifying HJM framework that

  • models intraday spot and futures prices,

  • is based on two stochastic processes motivated by economic interpretations,

  • separates the stochastic dynamics in trading and delivery time,

  • is consistent with the initial term structure (i.e. the price forward curve),

  • is able to price options on futures by means of the Black–Scholes formula,

  • allows for the use of classical day-ahead spot price models such as Schwartz and Smith (Citation2000), Lucia and Schwartz (Citation2002),

  • includes many model classes such as structural models and factor models.

To further the development of this framework, empirical studies are needed: statistical evaluations but also calibration methods need to be discussed. The theoretical applications of section 3 need to be specified and calibrated to real data from intraday, spot, futures, and option prices. This is subject of future research.

Acknowledgments

The authors thank the anonymous reviewers for their careful reading of the manuscript and their insightful comments and suggestions.

Disclosure statement

No potential conflict of interest was reported by the authors.

Additional information

Funding

WJH is grateful for the financial support from the Fraunhofer Institute for Industrial Mathematics ITWM, www.itwm.fraunhofer.de.

Notes

1 See Heath et al. (Citation1992) for the original paper introducing this framework for interest rate modelling.

2 Forward kernel is the name used by Caldana et al. (Citation2017).

3 Modelling of the day-ahead spot price is a common approach, for which several different approaches have been developed, cf. Weron (Citation2014).

4 This also allows for seasonal volatility in the market noise.

5 By cascading we mean the way how futures with a longer delivery period are settled. For example, a calendar year futures contract cascades (or splits up) into three monthly futures (January, February, and March) and three quarterly futures (Q2, Q3, and Q4) upon start of delivery. This way, these can be traded independently again. In the German market monthly futures do not cascade. However, the settlement price at the end of the delivery is exactly the average of the day-ahead spot prices during delivery. This could be interpreted that also monthly futures are cascading to the hourly (day-ahead) spot contracts, since their price converges to this average.

6 Continuous settlement of the futures contract makes it more like a swap contract on the forward kernel.

References

  • Aïd, R., Campi, L., Huu, A.N. and Touzi, N., A structural risk-neutral model of electricity prices. Int. J. Theor. Appl. Finance, 2009, 12, 925–947. doi: 10.1142/S021902490900552X
  • Applebaum, D., Lévy Processes and Stochastic Calculus, 2 ed. Cambridge studies in advanced mathematics, Vol. 116, 2009, (Cambridge University Press: Cambridge).
  • Barlow, M.T., A diffusion model for electricity prices. Math. Finance, 2002, 12, 287–298. doi: 10.1111/j.1467-9965.2002.tb00125.x
  • Barndorff-Nielsen, O.E. and Shephard, N., Non-Gaussian Ornstein–Uhlenbeck-based models and some of their uses in financial economics. J. Royal Statist. Soc. Ser. B (Statist. Methodol.), 2001, 63, 167–241. doi: 10.1111/1467-9868.00282
  • Benth, F.E., Benth, J.Š and Koekebakker, S, Stochastic Modelling of Electricity and Related Markets, 2008 (World Scientific: Singapore).
  • Benth, F.E. and Koekebakker, S., Stochastic modeling of financial electricity contracts. Energ. Econ., 2008, 30, 1116–1157. doi: 10.1016/j.eneco.2007.06.005
  • Benth, F.E. and Paraschiv, F., A space-time random field model for electricity forward prices. J. Bank. Finance, 2018, 95, 203–216. doi: 10.1016/j.jbankfin.2017.03.018
  • Benth, F.E., Piccirilli, M. and Vargiolu, T., Mean-reverting additive energy forward curves in a Heath–Jarrow–Morton framework. Math. Financ. Econ., 2019,
  • Bessembinder, H. and Lemmon, M.L., Equilibrium pricing and optimal hedging in electricity forward markets. J. Finance, 2002, 57, 1347–1382. doi: 10.1111/1540-6261.00463
  • Brigo, D. and Mercurio, F., Interest Rate Models–Theory and Practice, Vol. 2, 2006 (Springer: Berlin/Heidelberg).
  • Caldana, R., Fusai, G. and Roncoroni, A., Electricity forward curves with thin granularity: Theory and empirical evidence in the hourly EPEXspot market. Eur. J. Oper. Res., 2017, 261, 715–734. doi: 10.1016/j.ejor.2017.02.016
  • Clewlow, L. and Strickland, C., Valuing energy options in a one factor model fitted to forward prices. SSRN Electron. J., 1999.
  • Dufresne, D., The log-Normal approximation in financial and other computations. Adv. Appl. Probab., 2004, 36, 747–773. doi: 10.1239/aap/1093962232
  • EEX, Annual Report 2018. Visited 18-06-2019, 2019.
  • EPEX, EPEX SPOT power trading results of May 2017: Year-on-year increase in Intraday trading in May. Visited 27-06-2018, 2017.
  • EPEX, EPEX SPOT power trading results of May 2018: Monthly volumes exceed 50 TWh for the first time since 2015. Visited 27-06-2018, 2018.
  • Fanelli, V. and Schmeck, M.D., On the seasonality in the implied volatility of electricity options. SSRN Electron. J., 2018.
  • Heath, D., Jarrow, R. and Morton, A., Bond pricing and the term structure of interest rates: A new methodology for contingent claims valuation. Econometrica, 1992, 60, 77–105. doi: 10.2307/2951677
  • Hinz, J., von Grafenstein, L., Verschuere, M. and Wilhelm, M., Pricing electricity risk by interest rate methods. Quant. Finance, 2005, 5, 49–60. doi: 10.1080/14697680500040876
  • Jurek, Z.J. and Vervaat, W., An integral representation for selfdecomposable Banach space valued random variables. Zeitschrift Für Wahrscheinlichkeitstheorie Und Verwandte Gebiete, 1983, 62, 247–262. doi: 10.1007/BF00538800
  • Karatzas, I. and Shreve, S.E., Brownian Motion and Stochastic Calculus, 2 ed, 1998 (Springer New York: New York).
  • Kiesel, R. and Paraschiv, F., Econometric analysis of 15-minute intraday electricity prices. Energy Econ., 2017, 64, 77–90. doi: 10.1016/j.eneco.2017.03.002
  • Kiesel, R., Paraschiv, F. and Sætherø, A., On the construction of hourly price forward curves for electricity prices. Comput. Manag. Sci., 2018.
  • Kiesel, R., Schindlmayr, G. and Börger, R.H., A two-factor model for the electricity forward market. Quant. Finance, 2009, 9, 279–287. doi: 10.1080/14697680802126530
  • Koekebakker, S. and Ollmar, F., Forward curve dynamics in the nordic electricity market. Managerial Finance, 2005, 31, 73–94. doi: 10.1108/03074350510769703
  • Latini, L., Piccirilli, M. and Vargiolu, T., Mean-reverting no-arbitrage additive models for forward curves in energy markets. Energy Econ., 2018, in press.
  • Longstaff, F.A. and Wang, A.W., Electricity forward prices: A high-frequency empirical analysis. J. Finance., 2004, 59, 1877–1900. doi: 10.1111/j.1540-6261.2004.00682.x
  • Lucia, J.J. and Schwartz, E.S., Electricity prices and power derivatives: evidence from the nordic power exchange. Rev. Deriv. Res., 2002, 5, 5–50. doi: 10.1023/A:1013846631785
  • Lyle, M.R. and Elliott, R.J., A ‘simple’ hybrid model for power derivatives. Energy Econ., 2009, 31, 757–767. doi: 10.1016/j.eneco.2009.05.007
  • Sato, K.I., Lévy Processes and Infinitely Divisble Distributions. Cambridge studies in advanced mathematics, Vol. 68, 2013 (Cambridge University Press: Cambridge). Revised.
  • Schwartz, E. and Smith, J.E., Short-Term variations and long-term dynamics in commodity prices. Manage. Sci., 2000, 46, 893–911. doi: 10.1287/mnsc.46.7.893.12034
  • Wagner, A., Residual demand modeling and application to electricity pricing. Energy J., 2014, 35. doi: 10.5547/01956574.35.2.3
  • Weron, R., Electricity price forecasting: A review of the state-of-the-art with a look into the future. Int. J. Forecast., 2014, 30, 1030–1081. doi: 10.1016/j.ijforecast.2014.08.008
  • Wolfe, S.J., On a continuous analogue of the stochastic difference equation Xn=ρXn−1+Bn. Stoch. Process. Their Appl., 1982, 12, 301–312. doi: 10.1016/0304-4149(82)90050-3