Virtual Barrels Quantitative Trading in The Oil Market
Virtual Barrels Quantitative Trading in The Oil Market
Virtual Barrels Quantitative Trading in The Oil Market
Economics
Virtual Barrels
Quantitative Trading in the Oil Market
Ilia Bouchouev
Pentathlon Investments, LLC, Westfield, NJ, USA
This work is subject to copyright. All rights are solely and exclusively
licensed by the Publisher, whether the whole or part of the material is
concerned, specifically the rights of translation, reprinting, reuse of
illustrations, recitation, broadcasting, reproduction on microfilms or in
any other physical way, and transmission or information storage and
retrieval, electronic adaptation, computer software, or by similar or
dissimilar methodology now known or hereafter developed.
The publisher, the authors, and the editors are safe to assume that the
advice and information in this book are believed to be true and accurate
at the date of publication. Neither the publisher nor the authors or the
editors give a warranty, expressed or implied, with respect to the
material contained herein or for any errors or omissions that may have
been made. The publisher remains neutral with regard to jurisdictional
claims in published maps and institutional affiliations.
This Springer imprint is published by the registered company Springer
Nature Switzerland AG
The registered company address is: Gewerbestrasse 11, 6330 Cham,
Switzerland
To our children, Vlad and Valeria
Preface
The public perception of oil trading revolves around counting barrels,
predicting outcomes of OPEC meetings, estimating gasoline and jet fuel
consumption from high-frequency traffic measures, analyzing Middle
Eastern politics, assessing the impact of regulations and trade
sanctions, or hiring meteorologists to forecast the trajectory of a
hurricane that may impact offshore production and refining centers.
These are fascinating topics to discuss at the dinner table where
everyone has an opinion on some of numerous factors that impact the
price of oil. But how relevant is this information for professional oil
trading, given so much noise and almost no reliable facts that the
market does not already know? More importantly, what are the
channels that translate such fundamental information into the price
action in the futures market where the price of oil is determined, and
how do professional oil traders filter out the noise and turn the relevant
bits into consistently profitable trading strategies?
The book attempts to answer these questions following the author’s
own experiences in managing an energy derivatives trading business
for over twenty years. The language of the book is rather quantitative,
reflecting the demands of modern markets for more data, cutting-edge
technology, and advanced analytical modeling. The book tracks the
evolution of quantitative oil trading throughout its entire history.
Options play a special role throughout the book. With somewhat
elevated barriers to entry, the business of options and volatility trading
is where quants tend to shine. But, in contrast to many other financial
markets, the dynamics of oil options can hardly be separated from the
economics of the underlying physical system, the behavior of hedgers,
and the crucial role played by storage.
The need for storage is what led to the beginnings of oil trading in
the 1860s shortly after the process of oil drilling was discovered in
Western Pennsylvania. Oil quickly proved its worth, but the demand for
barrels at the well, constrained by complicated logistics, has not always
moved in unison with supply. To buy some time and buffer fluctuations
between supply and demand, oil had to be kept somewhere. The
cheapest storage facility at the time were the dumps. Not surprisingly,
the pioneers of oil trading, who were buying excess production at
discounts and storing it, were called “dump men.”
To move oil from the dumps to consumers, which was done largely
with the help of horses, oil was poured into empty whiskey barrels. The
whiskey itself helped to negotiate the price. The price of oil was
measured in dollars per barrel. Only a handshake agreement was
needed to secure the deal, as honor and trust were held in higher
regard than written contracts. The business of oil trading was built as
an exclusive private club, largely closed to outsiders.
Little has changed in the first hundred years of the oil business,
other than horses being replaced by pipelines and ships, and whiskey
by a much wider selection of fine spirits. Membership in the private oil
club was not open to newcomers until the 1990s when oil bosses
brought in a breed of young scientists, the quants. Hiring geeks for the
old-fashioned teams of oil traders was considered somewhat
progressive and even entertaining. The mandate for quants was clear.
The oil world was becoming more complex, and in a wider web of
interlinked prices, they must search for a low-hanging fruit, the
strategy called arbitrage.
An opportunity to make small but consistent profits from low-risk
arbitrage trades had a particular allure for quants. Not only did it
provide a more financially lucrative career alternative to teaching
mathematics at a university, but it also gave quants a chance to prove
that mathematics actually works in the real world. One challenge for
quants was the relatively small size of the strategy, which was dwarfed
by riskier directional wagers made by old-timers. In the oil market, the
size of the bet and the amount of risk taken determine the trading
hierarchy, and for a while, quants were largely viewed as second-class
citizens in the oil world. To earn their proper seat at the table, more
arbitrages had to be found. Fortunately, a new era of oil trading was
about to start where quants were destined to play a much larger role.
The world was globalizing and digitalizing. To meet fast-growing oil
demand, new sources of supply had to be found. The shale revolution in
the US led to unprecedented growth in the global petroleum
infrastructure. New energy assets were built with large risks held by
financial investors. In contrast to prior decades, when oil supply growth
was predominantly driven by sovereign-controlled entities, this time
the risks were held by private capital and managed in the market. The
risks were quite complex, requiring more sophisticated hedging
instruments and more advanced quantitative skills in managing them.
At the same time, oil was also rapidly financializing and moving
away from obscure voice negotiations to transparent electronic screens.
The private club of oil trading was finally open to the public. Buying the
new digital barrel of oil has proven to be particularly handy at times of
rising inflation and geopolitical conflicts, when many other financial
assets struggle. With new computer technologies coming in handy,
quants went on a journey to modernize the oil industry. The journey
was fully endorsed by the Wall Street marketing machine, which
smelled the potential and scalability of the new business–the business
of oil derivatives.
This book tells the inside story of this business. This story has been
molded into its current form by the energy trading course taught by the
author at the Courant Institute of Mathematical Sciences at New York
University, and I thank my students for their invaluable feedback. Many
of my former colleagues contributed to the ideas presented in this
book. I thank them all sincerely for being such a critical part of the
trading business that we have successfully run together for over two
decades. And I especially thank my family for their constant
encouragement and support of my project in very unique and special
ways.
Ilia Bouchouev
New York, USA
May 2023
Contents
1 Introduction
1.1 The Ecosystem of Real Options
1.2 The Structure of the Book
Reference
Part I Economic Foundations, Markets, and Participants
2 Oil, Money, and Yields
2.1 Commodities and Money
2.2 The Oil Own Rate of Interest
2.3 Carry and Convenience Yield
2.4 The Roll Yield
References
3 Fundamentals, Storage, and the Model of the Squeeze
3.1 The Invisible Hand of Storage Boundaries
3.2 The Canonical Theory of Storage
3.3 A Stylized Model of the Squeeze
3.4 Cushing:Pipeline Crossroads of the World
3.5 Negative Oil Prices
References
4 Financializationand the Theory of Hedging Pressure
4.1 The Theory of Normal Backwardation
4.2 The Hedging Equilibrium
4.3 The Genesis of Oil Financialization
4.4 Inflation Hedging and Risk Parity
4.5 Inconvenience Yield, or the Theory of Normal Contango
References
Part II Quantitative Futures Strategies
5 Systematic Risk Premia Strategies
5.1 The Evolution of Algos
5.2 Myths and Realities of Oil Momentum
5.3 Carry as a Transmitter of Fundamentals to Prices
5.4 Value and Mean-Reversion
5.5 The Reaction Function
References
6 Quantamentals
6.1 Trading Curve and Convexity
6.2 Time Spreads and Inventories
6.3 WTI-Brent Accordion
6.4 Cointegration and Energy Stat-Arb
6.5 Disentangling Flows and Positioning
References
7 Macro Trading
7.1 Dynamic Systems and Feedback Loops
7.2 Oil, Dollar, and Commodity Terms of Trade Strategies
7.3 Oil and Energy Equities
7.4 Oil and Inflation
7.5 Macro Fair-Value Model
Part III Volatility Trading
8 Options and Volatilities
8.1 Options and “Théorie de la Spéculation”
8.2 Local Volatility and Diffusions
8.3 Delta Hedging and Option Replication
8.4 Realized Volatility
8.5 Implied Volatility and its Skew
References
9 The Hidden Power of Negative Gamma
9.1 Options and Insurance
9.2 The Most Powerful Option Greek
9.3 The Smile of the Volatility Risk Premium
9.4 The Art and Science of Delta Hedging
9.5 The Behavior of Hedgers and Regime Changes
References
10 Volatility Smile Trading
10.1 Producer Hedging and Volatility Market-Making
10.2 Skew Delta and Two Types of Stickiness
10.3 When Black Smirks, Bachelier Smiles
10.4 Fat Tails and the Quadratic Normal Model
References
Part IV Over-the-Counter Options
11 Volatility Term Structure and Exotic Options
11.1 Dark Pools and the Hacienda Hedge
11.2 The Term Structure of Implied and Local Volatilities
11.3 Volatility Discount from Price Averaging
11.4 Early Expiry Options and Swaptions
11.5 Multi-Factor Models and Other Exotics
References
12 Volatility Arbitrage and Model Calibration
12.1 The Inverse Problem of Option Pricing
12.2 Bootstrapping in Time
12.3 Market-Implied Probability Distribution
12.4 Local Volatility Smile
References
13 Spread Options and Virtual Storage
13.1 The Synthetic Storage Strategy
13.2 Triangular Correlation Arbitrage
13.3 The Dichotomy of Spread Option Pricing
13.4 Dealing with Unobservables
References
14 Epilogue
14.1 The Roadmap for Energy Transition and Virtual
Commodities
Appendix A:Diffusions and Probabilities
Appendix B:Option Pricing under Normal and Lognormal
Distributions
Appendix C:The Perturbation Method and the Quadratic Normal
Model
Appendix D:Option Pricing with Time-Dependent Volatility
Appendix E:Average Price Options
Appendix F:The Inverse Diffusion Problem
Glossary
References
Index
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023
I. Bouchouev, Virtual Barrels, Springer Texts in Business and Economics
https://doi.org/10.1007/978-3-031-36151-7_1
1. Introduction
Ilia Bouchouev1
(1) Pentathlon Investments, LLC, Westfield, NJ, USA
Reference
Bouchouev, I. (2023). Testimony to the House Subcommittee on Economic Growth, Energy Policy,
and Regulatory Affairs of the Committee on Oversight and Accountability, March 8. Reprinted in
Commodity Insights Digest (2023), Vol. 1.
Footnotes
1 From the author’s testimony to the U.S. House Subcommittee on Economic Growth, Energy
Policy, and Regulatory Affairs of the Committee on Oversight and Accountability. See Bouchouev
(2023).
Part I
Economic Foundations, Markets, and
Participants
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023
I. Bouchouev, Virtual Barrels, Springer Texts in Business and Economics
https://doi.org/10.1007/978-3-031-36151-7_2
(2.1)
(2.2)
Let us assume that we also know the price of the second money
standard in terms of the first one. In other words, we know spot oil
prices S(t) and S(T) at times t and T, both expressed in USD per barrel
($/bbl)2:
From here, we express barrels as the ratio of dollars to oil prices, and
substitute into (2.2), which allows us to relate the oil rate of interest to
the dollar rate of interest, using the definition (2.1), as follows:
(2.3)
where
Therefore, the futures price can be expressed as the spot price that
grows exponentially at the rate equal to the difference between the
nominal rate of interest and the oil own rate of interest
(2.5)
Alternatively, one can back out the oil own rate of interest from the
futures curve
(2.6)
where the spot price is understood as the futures with the shortest
maturity.
Figure 2.1 illustrates how the own rate relates to backwardation
and contango of the futures curve with different maturities.
Fig. 2.1 If the oil own rate of interest exceeds (falls below) USD rate of interest then the
futures curve is in backwardation (contango)
If b > r and the intrinsic worth of the commodity is relatively high,
the futures curve is in backwardation. In contrast, if b < r and the
commodity own rate is relatively low, the futures curve is in contango.
If b = 0, then the futures curve is in a slight contango, monotonically
increasing at the rate r.
Having defined the basic terminology of the oil economy, we can
now illustrate it with an important case study. We consider an example
of oil loans from the US Strategic Petroleum Reserves (SPR), the world’s
largest supply of emergency oil stocks, held by the US Government. The
SPR was established in 1975 by the Energy Policy and Conservation Act
(EPCA) following the Oil Embargo of 1973–1974 by the Organization of
Arab Petroleum Exporting Countries (OAPEC). The mission of the SPR
is to store petroleum to diminish the impact of disruptions on
petroleum supplies and to carry out US obligations under the
International Energy Program. Oil stocks can be released from the SPR
either via competitive sales, or as oil exchanges. Oil exchanges are loans
where barrels are released to traders in exchange for a larger quantity
of barrels to be returned at some time in the future.6
Consider an example of the large SPR oil loan that was authorized in
November 2021. At that time, oil demand was quickly recovering from
the Covid-19 induced slowdown, while oil supplies from the OPEC+
group of large producers were released to the market at a slower pace.7
To reduce fundamental supply and demand imbalance, the US
Government offered oil from the SPR to traders in exchange for a larger
quantity of oil to be returned at a later time. The rates of return on
these oil loans were contractually set for five designated loan
maturities. The longest maturity loan stipulated 9.1% more barrels in
excess of the originally borrowed quantity to be returned to the SPR in
approximately thirty months. This allows the US Government not only
to replenish the reserve, but also to grow it at no cost.
The question is then why would anyone agree to borrow oil at such
a high rate, especially given that the interest rate on fiat money at that
time was nearly zero? The answer lies in the steeply backwardated
shape of the oil futures curve. To compare loan rates offered by the US
Government with the market-implied rates, one can apply formula (2.6)
to calculate oil own rates prevailing at the same time from WTI futures,
which is the primary US oil futures contract. It turns out that the thirty-
month oil own rate implied by the WTI futures curve at that time was
even higher, closer to 20%. Figure 2.2 illustrates.
Fig. 2.2 The rates on SPR oil loans (grey bars) offered by the US government in November
2021 are compared to WTI market implied rates (orange) computed using (2.6)
If oil markets were frictionless and conducted only using the WTI
grade of oil with all transactions executed at no cost, then the trader
could have theoretically made riskless profits by borrowing barrels
from the US Government at 9%, selling them in the spot market, buying
thirty-month futures at approximately 20% discount and taking the
delivery of physical barrels at that time. Obviously, the real market has
plenty of frictions, including high logistical costs associated with
physical oil delivery and large price uncertainty in the basis between
WTI and specific oil grades that must be returned to the SPR.
Nevertheless, for a physical oil trader, participation in such loan
transactions could be quite lucrative. This is our first example of the
trading strategy where the needs of one market participant, the US
Government, creates an arbitrage opportunity for a professional oil
trader.
Despite the fact that oil has many attributes of an alternative
standard of money, including even having its own rate of interest, we
still do not use oil as money. One reason is the complexity of storing oil,
but perhaps a more important one is its extreme volatility. To illustrate,
Fig. 2.3 shows the history of a one-year own rate of interest implied by
WTI futures.
Fig. 2.3 The history of a one-year oil own rate of interest as implied by WTI futures
The oil own rate of interest fluctuates wildly, as it is driven by the
constantly changing slope of the futures curve. For example, at the
onset of the Covid-19 pandemic when oil demand fell rapidly, storage
costs increased substantially, culminating in an unprecedented episode
of negative prices, which we will discuss in detail in the next chapter.
For a negative price, one cannot compute an own rate, but even with
this abnormal data point removed, the own rate during that time
reached its nadir of negative 100%. However, within two years it swung
to positive 30%, as the world recovered from the pandemic and the
convenience yield of owning physical barrels skyrocketed. With such a
high volatility in a lending rate, it would be extremely challenging to
ensure any stability in the oil-denominated economy.
While the oil own rate of interest may sound like an obscure
theoretical concept, it underpins many commercial transactions and
plays the overall crucial role in the oil market. We will see throughout
the book how large its impact is on different trading strategies. For
many market participants, however, this concept is better known under
different names. In fundamental trading, it is more often associated
with the concept of convenience yield.
2.3 Carry and Convenience Yield
The own rate of interest reflects the combined effect of the intrinsic
value of the commodity and the cost of storing it. It is calculated by
comparing the commodity price today to its price in the future. The
current price and the future price are connected by the carry trade. The
carry arbitrage trade has been well known to commodity merchants
from the early days of trading. It simply refers to merchants buying and
storing a physical commodity, while hoping to sell it later at a higher
price. If the forward price is also locked in, then the carry trade
effectively represents arbitrage in time.
Oil started to trade in the 1860s in Western Pennsylvania shortly
after the first oil well has been drilled there. The first oil traders were
carry arbitrageurs who developed some primitive forms of oil storage.
They were buying excess oil from local producers at discounts and
storing it in dumps. These pioneers of oil trading were called “dump
men”:
(2.12)
The first term in this decomposition is defined as the spot return (SR). It
represents a hypothetical investment return on buying a physical barrel
without incurring any storage costs, which cannot be realized by
trading futures. The second term is the roll return (RR). The roll return
also cannot be generated by means of a futures trading strategy, as it is
calculated by comparing prices for contracts with different maturities.
Both the spot return and the roll return act merely as attribution terms
in the decomposition of the actual return on investment in futures. In
practice, the roll return is calculated as the difference between the
actual realized excess return and the hypothetical spot return.12
It is more common to look at spot and roll returns in annualized
terms, in which case all terms in the Eq. (2.12) must be divided by T − t.
We then define the roll yield as the annualized roll return on this
strategy as follows
(2.13)
In other words,
(2.14)
It follows from (2.5), (2.10), and (2.14) that the roll yield, the own rate,
and the convenience yield are related as
(2.15)
In other words, the roll yield can be thought of as the spread between
the oil own rate of interest and the USD rate of interest. Alternatively, it
can be understood as the convenience yield, net of storage and interest
costs. Thus, we have defined an important variable that characterizes
the shape of the futures curve and translated it into three different
languages used by different market participants.
To illustrate the importance of the roll yield, consider a simple
strategy where a financial investor attempts to synthesize the
continuous ownership of a physical barrel by holding futures with the
nearest maturity. Since futures are listed with monthly expirations, in
this simplified strategy an investor buys one-month expiry futures,
liquidates the contract when it expires, and buys the next maturity
futures at that time. Note that such a strategy is unlikely to be
implemented in practice and it is used only for illustration purposes. In
the real world, most investors must liquidate futures prior to expiration
with more realistic examples of a buy-and-roll strategy analyzed in
Chap. 4.
Let t = T0 and assume that the investor repeats the strategy for N
consecutive months by buying Ti, i = 1, 2, …N maturity futures and
successively liquidating them at the spot price at time Ti. The
cumulative excess return (CER) of this strategy is
In each period, we use (2.12) to decompose the excess return into the
spot return and the roll return. The cumulative spot return (CSR) is
then
References
Bouchouev, I. (2022). The Strategic Petroleum Reserve strategies: Risk-free return or return-free
risk? The Oxford Institute for Energy Studies.
Boyle, P. C. (1899, September 6). Testimony at the Hearing of the US Industrial Commission on
Trusts and Industrial Combination.
Fisher, I. (1896). Appreciation and interest. Publications of the American Economic Association,
11(4), 331–442.
Geman, H. (2005). Commodities and commodity derivatives: Modeling and pricing for
agriculturals, metals and energy. Wiley.
Hull, J. C. (2018). Options, futures, and other derivatives (10th ed.). Pearson.
[zbMATH]
Kaldor, N. (1939). Speculation and economic stability. The Review of Economic Studies, 7(1), 1–
27.
[Crossref]
Keynes, J. M. (1936). The general theory of employment, interest, and money. Macmillan.
Naldi, N. (2015). Sraffa and Keynes on the concept of commodity rates of interest. Contributions
to Political Economy, 34(1), 17–30.
[Crossref]
Smiley, A. W. (1907). A few scraps: Oily and otherwise. The Derrick Publishing Company.
Sraffa, P. (1932). Dr. Hayek on money and capital. The Economic Journal, 42 (165), 42–53.
[Crossref]
Whiteshot, C. A. (1905). The oil-well driller: A history of the world’s greatest enterprise, the oil
industry. Acme Publishing Company.
Footnotes
1 See Fisher (1896). For consistency with contemporary conventions adopted in the economic
literature, our definitions of certain terms differ from Fisher’s original formulation. For
example, we use commodity price appreciation in terms of money to characterize commodity
inflation, in contrast to Fisher’s choice of looking at inflation as money depreciation in terms of
commodities.
2 The term spot price is used only for pedagogical clarity. We will explain shortly that oil spot
prices with instantaneous delivery do not really exist, and the spot price should be understood
as the forward contract with the nearest delivery.
3 Sraffa (1932) only defined the concept by virtue of the following example of borrowing bales
of cotton: “The rate of interest … per hundred bales of cotton, is the number of bales that can be
purchased with the following sum of money: the interest on the money required to buy spot 100
bales, plus the excess (or minus the deficiency) of the spot over the forward prices of the 100
bales”.
4 See Keynes (1936), chapter 17. Keynes restated Sraffa’s definition algebraically, which
triggered a heated debate between the two economists in a series of letters. The debate was
caused by their disagreement on whether to use spot prices or forward prices in defining the
dollar cost of borrowing the commodity. While it may appear to be a small technical nuance, it
led to a more profound philosophical differences in the interpretation of the concept. For
details, see, e.g., Naldi (2015). In this book, we accept the definition as it was stated by Keynes,
which simply mimics the definition of the interest rate on fiat money.
7 The Organization of Petroleum Exporting Countries (OPEC) was formed in 1960. The
membership of OPEC has varied over the years. As of 2022, OPEC included thirteen countries
with the largest producer being Saudi Arabia. A larger but more loosely structured
organization, known as OPEC+, which included several other major sovereign producers, was
formed in 2016.
8 From the testimony of Patrick C. Boyle, proprietor and publisher of The Oil City Derrick at the
Hearing of the US Industrial Commission on Trusts and Industrial Combinations, see Boyle
(1899). The testimony was subsequently reprinted in Whiteshot (1905). It describes the origins
of oil trading and highlights that dump men were the predecessors of the oil refiners. The dump
business is also described in Smiley (1907), who wrote that “the dumps in those days were
practically the exchange, and made the market price for oil each day”. Many other interesting
facts related to the early days of oil drilling in Pennsylvania, including the construction of first
wooden oil tanks, are given by Giddens (1947).
9 Kaldor (1939) writes that “in normal circumstances, stocks of all goods possess a yield,
measured in terms of themselves, and this yield which is a compensation to the holder of stocks,
must be deducted from carrying costs proper in calculating net carrying costs. The latter can,
therefore, be negative or positive.”
11 The term excess return differs from the total return as the latter includes the interest on
collateral. Since in this book, for the most part, we ignore the interest on collateral for futures,
the two terms are often used interchangeably. The distinction will become clearer in Chap. 4 in
the context of fully collateralized commodity indices.
12 Unlike price changes and log-returns, simple percentage returns are not additive. This
makes a similar decomposition for simple returns more complex. For simple returns, the roll
return is formally defined as the difference between the excess return and the spot return.
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023
I. Bouchouev, Virtual Barrels, Springer Texts in Business and Economics
https://doi.org/10.1007/978-3-031-36151-7_3
Fig. 3.1 Working’s representation of the relationship between the price of storage and
inventories. The sign of the spot-futures spread S − F is flipped for consistency with the
convention in the oil market
Even though Working’s graph is sometimes referred to as the first
theory of storage, his argument did not follow any analytical theory or
model. It only conceptually represented the nonlinearity of convenience
yield at low inventories based on empirical observations in several
agricultural markets. The objective of a proper storage model is to
analytically describe the behavior of the system that generates such
nonlinear dynamics of prices versus inventories within the model itself.
(3.2)
The supply is the sum of the local production and imports from
other regions. The total demand D(t) is made up of the local
consumption and exports. We use tilde to highlight that the supply
is uncertain, and, for simplicity, we assume the demand to be
deterministic. The problem remains conceptually similar if, instead, the
uncertainty is attributed to demand, or to both supply and demand. The
Eq. (3.2) is an accounting identity; it does not represent any model or
assumption.
The integration of supply and demand imbalances over a period (t,
T) results in the aggregate level of inventories accumulated over this
period
The presence of the same variable x(t) in both equations for today and
for tomorrow highlights the intertemporal nature of the storage
problem. While this variable represents a carry-out portion for the
availability today, it also acts as a carry-in for the availability tomorrow.
The availabilities on both days are, therefore, linked as follows
(3.3)
In other words, tomorrow’s availability is equal to today’s availability
minus what is consumed today, plus whatever can be replenished
tomorrow, with the latter being uncertain. The theory of storage looks
at the price of oil as an implicit function of its availability, S(t) = S(a(t)).
Both price and availability depend on inventory decisions, and they
must be determined jointly in equilibrium.
We now differentiate between two states of inventories: one normal
state that corresponds to non-zero inventories, and a second stock-out
state when inventories are depleted. In the first state, when some oil is
available, the spot price must be linked to the forward price for the next
period F(a(t + dt)) via the economics of the carry arbitrage trade. In the
second case of no inventories, the price is entirely driven by immediate
supply and demand and is determined by the inverse demand function.
Therefore,
(3.4)
Here, C(x) represents the total storage cost, which can also depend on
the level of inventories. To simplify the problem, we assume that C(x) is
a known nonlinear function that rises without limit as x → Xmax.
Alternatively, one could have considered a third state where inventories
reach the maximum storage capacity, x = Xmax. In this simplified two-
state formulation of the problem, C(x) can be viewed as a mirror image
of the inverse demand function at low inventories. Note that in the
stock-out state, the availability is the same as the demand, and D(t) in
the argument of the inverse demand function is replaced by a(t).
In the traditional formulation of the theory of storage, the futures
price in (3.4) is replaced by the expected spot price at time t + dt, which
depends on the availability a(t + dt).5 The expectation is taken over all
uncertain realizations of future production , which impacts the
availability a(t + dt) via (3.3). This highlights the circular nature of the
problem caused by the feedback loop between prices and inventories.
The price today depends on the availability today via the demand
function. Likewise, the price tomorrow depends on the availability
tomorrow. The prices today and tomorrow are connected via the carry
trade. At the same time, the availability today and the availability
tomorrow are linked via the Eq. (3.3). This problem is intertemporal, as
a similar loop exists for tomorrow, the day after tomorrow, etc. To solve
such a problem, one must construct an iterative process which
converges to some equilibrium state for both price and availability. The
process also needs a boundary condition, which in this case is defined
by the possibility of a stock-out.
This is a complex problem. Think about Crusoe’s decision-making
process as a tree, where at each time period alternative decisions could
be made about food allocation between immediate consumption and its
storage for tomorrow. As time moves forward, the tree of possible
scenarios branches out. Each node represents total supplies that will be
held at that time, which is the sum of inventories carried from the prior
period and whatever else he can find on that day. The process continues
until some random path leads to the case of zero inventories. The stock-
out sets the terminal boundary condition for the process, and the
problem is solved iteratively backward in time to establish the
inventory management rule at each prior step that maximizes the usage
of all inventories.
In general, to solve this problem, techniques of stochastic dynamic
programming must be applied. Fortunately, for the purpose of this
book, a precise solution of this complex problem is not needed. Our
primary interest is twofold. First is to use this framework to illustrate
an important feedback loop that exists between prices and inventories.
Second is to explain its limitations when applied to oil markets by
simply looking at salient properties of a typical solution. We then use
some valuable insights resulting from these limitations to develop a
simpler and more practical alternative approach in the following
section.
A typical solution to this problem shows the equilibrium price as a
function of availability, as illustrated in Fig. 3.2. It is characterized by
the kink that corresponds to the level of zero inventories separating the
two states, like in the Eq. (3.4). This kink symbolizes the dual role of
commodities, as an investment asset and as a product for consumption.
The location of the kink that marks the danger zone of a stock-out is
what the storage problem is effectively solving for. The price in the
region to the left of the kink is solely determined by production and
consumption. It is computed via an inverse demand function, as there is
no storage buffer in this region. To the right of the kink, oil behaves as
an investment asset. In this state, the price is mostly determined by the
cost of storage, which increases nonlinearly near the boundary on the
maximum storage capacity. It should also be noted that conventional
models require prices to remain positive. However, for the oil market
this assumption, as we will see shortly, is too restrictive.
Fig. 3.2 A typical solution of the canonical storage problem has a characteristic kink that
separates the consumption region with zero inventories from the investment region driven by
the economics of the carry trade
Despite its popularity in academic studies, the standard approach to
the storage theory attracted only very limited interest among oil
traders. One practical challenge is the extremely low price elasticity of
oil demand. The solution to the problem at zero inventories effectively
degenerates into a straight vertical line. It means that if we truly run
out of oil stocks, and the supply of oil cannot meet the demand, then the
price must rise towards infinity. Therefore, the optimal solution is to
always store oil in sufficient quantities so that a stock-out is never
reached. However, this solution is impractical because of the high cost
of oil and the limitation on the storage capacity.
A similar problem occurs when inventories reach maximum storage
capacity. Like oil demand, the short-term oil supply is also inelastic
with respect to price, as many producers cannot shut down production
instantaneously without permanently damaging oil reservoirs. In this
case, if a significant portion of the oil demand disappears, like it did, for
example, during the early days of the Covid-19 pandemic, then the
excess oil must go to storage. Once all storage is filled, then the price
must theoretically fall without any boundary to force either production
or consumption to change. We will discuss shortly the infamous
episode of oil falling to negative forty dollars per barrel, which is
probably as close as one can get to negative infinity in financial
markets.
To summarize, given that the price elasticity of oil demand and oil
supply is extremely low, the optimal solution to the oil storage problem
collapses into something that resembles the graph in Fig. 3.3. It is
driven by two extremities. The left tail marks the case of zero
availability, when the price goes to infinity. The right tail indicates that
the storage capacity is full, and the price goes to minus infinity.
Anywhere in between, where inventories are deemed to be normal and
sufficiently far away from either boundary, the price becomes largely
insensitive to the availability of inventories.
Fig. 3.3 The solution to the canonical storage problem degenerates at the boundaries when
the price elasticities of demand and supply approach zero
then the solution to the ordinary differential Eq. (3.5) can be easily
found, as follows
(3.6)
One can easily verify it by direct substitution. As time increases, the
inventories x(t) are pulled from their current state x0 towards the
equilibrium level . The pull occurs at an exponential rate defined by
the parameter k. As t → ∞, inventories converge towards the
equilibrium, i.e., . The larger the parameter k, the faster the
speed of mean-reversion.
The solution (3.6) is illustrated in Fig. 3.4 for two scenarios. In the
first example of slow draws, the initial inventory level x0 is high and k is
relatively small, which could be the result of a prior fall in demand,
perhaps caused by an economic recession. As demand gradually
recovers, inventories decrease and slowly converge to the long-term
equilibrium. In the second scenario of rapid builds, initial inventories
are low but mean-reversion is fast, which could be the result of a short-
term supply disruption caused by a geopolitical event, but the supply is
assumed to be restored relatively quickly.
Fig. 3.4 Examples of slowly drawing and rapidly building inventories that converge to a long-
term equilibrium
The Eq. (3.5) describes an unrealistic dynamic where one can
predict future inventories with certainty. In the real world, the path
towards an inventory equilibrium state is likely to be noisy with some
random fluctuations along the way that can either delay the
convergence to the normal state or accelerate it. To capture this noise, a
second term is added to the Eq. (3.5):
(3.7)
The noise is assumed to come from an increment dz, which represents a
random variable, drawn from a normal distribution with mean of zero
and variance equal to dt. It is scaled by the volatility parameter σ, which
characterizes the magnitude of the uncertainty.
As time t moves forward, the market may experience large,
unexpected shocks to supply and demand that are described by large
random values of dz(t). During such episodes, the random term in (3.7)
dominates the change in inventories. However, over time the steady
gravitational pull of the mean-reverting term ensures that inventories
drift towards an equilibrium level. The further inventories deviate from
the normal level, the stronger the pull towards the long-term
equilibrium. The Eq. (3.7) is one of many examples of stochastic
processes that are commonly used to describe uncertainty in financial
markets. It is known as an Ornstein-Uhlenbeck process, which was
originally developed in physics to analyze Brownian motion with
friction. It has been widely adopted for modeling mean-reversion in
commodity prices.
When noise is introduced to the system, the forward dynamics is no
longer certain. It can only be understood in a probabilistic sense. Each
stochastic process is associated with a probability density function p(x,
t; xT, T) which describes transition probabilities of the random variable
between two states at times t and T. If the variable at time t < T is
known to be located at the point x, then the density function describes
the probability of this variable reaching various points xT at some
future time T. Alternatively, if the variable is known to be at some target
state xT at time T, then the same density function also represents the
probability of having reached this target state starting from different
levels x at an earlier time t < T. The probability density function
satisfies the Kolmogorov forward and backward equations, with the
former also known as the Fokker-Planck equation. These equations and
other basic facts related to transition probabilities and stochastic
processes are summarized in Appendix A.
The most well-known probability density function is a Gaussian
bell-shaped curve that represents the normal distribution, centered at
the point xT = x with variance σ2(T − t):
(3.8)
(3.9)
(3.10)
(3.11)
One can think about infinite prices as the event of default by futures
traders. More practically, the market calls such an event a squeeze, and
we refer to this formulation of the storage problem as a stylized model
of the squeeze.
The solution to the storage problem now becomes simple and
intuitive. It is given by the derivative pricing formula (3.10) with the
boundary condition (3.11):
Note that the integration here is effectively performed over the range
xT ∈ (Xmin, Xmax), outside of which the probability density function is
equal to zero. Applying properties of the Dirac delta function (A.4) and
(A.5), we obtain that the price of the futures spread is given by
(3.12)
This representation explicitly relates the price of the futures spread to
market fundamentals, specifically to the probability distribution
function for stochastic inventories. It follows that the value of the
futures spread at time t is determined by the relative probabilities of
upside and downside squeezes, adjusted by the normal cost of storage.
One squeeze corresponds to the case of zero inventories, and the other
one to the case of zero available storage capacity. The proximity to each
boundary determines the value of the spread within the boundaries at
all times prior to expiration.
Let us now apply the generic formula (3.12) to a mean-reverting
stochastic process (3.7). It is shown in Appendix A that the probability
density for the mean-reverting process is also described by the normal
distribution
(3.13)
(3.15)
The state variable y(t) resembles the function (3.6) that describes the
behavior of inventories without the added noise. The role of the new
time variable is to capture the impact of mean-reversion on the
process variance. It reduces the effective variance via an exponential
dampening factor that depends on the speed of mean-reversion k. For
example, if the mean-reversion is slow and k → 0, then and the
variance of the mean-reverting process converges to the variance of the
normal process. However, for k > 0 the new effective remaining time to
maturity is reduced, as . We should also note that at
time t = T, the new time , y(T) = xT, and the boundary condition
(3.11) remains intact in new variables.
The substitution of the probability density (3.13) into the
representation (3.12) results in the following pricing formula for the
futures spread:
(3.16)
where the variables y and are defined above by (3.14) and (3.15).
The solution (3.16) to our stylized model of the squeeze is
sufficiently flexible to cover the wide range of plausible spread
behaviors versus the level of inventories. Figure 3.6 illustrates that
faster mean-reversion makes the range of spread values more
contained. In this example, the inventories are expressed as a
percentage of total storage capacity, so that 0 ≤ x ≤ 1, and the long-term
equilibrium inventory level is set at . Since is slightly closer
to the upper boundary, this makes the probability of tank-tops higher
than the probability of stock-outs, resulting in wider contango than
backwardation near the boundaries. The opposite effect can be
generated by choosing to be closer to the lower boundary. One can
easily extend this framework and apply (3.12) to the more complex
dynamics of inventories characterized by different stochastic processes
and probability density functions.
Fig. 3.6 Examples of one-month futures spread functions (3.16) for fast (k = 4) and slow
(k = 1) mean-reversions with 0 < x < 1 and
We now move from theory to practice and illustrate the approach
for the world’s most important storage hub, where the futures market
meets the physical barrels. For nearly three decades, all roads in the oil
business have led to the rather nondescript US town of Cushing,
Oklahoma, which has often been called the oil capital of the world.
References
Bouchouev, I. (2020, April 30). Negative oil prices put spotlight on investors, Risk.net.
Brennan, M. J. (1958). The supply of storage. The American Economic Review, 48(1), 50–72.
Brennan, M. J. (1991). The price of convenience and the valuation of commodity contingent
claims. In D. Lund & B. Oksendal (Eds.), Stochastic models and option values. North Holland.
Brennan, M. J., & Schwartz, E. S. (1985). Evaluating natural resource investments. Journal of
Business, 58(2), 135–157.
[Crossref]
Carmona, R., & Ludkovski, M. (2004). Spot convenience yield models for the energy markets.
Contemporary Mathematics, 351, 65–79.
[MathSciNet][Crossref][zbMATH]
Casassus, J., & Collin-Dufresne, P. (2005). Stochastic convenience yield implied from commodity
futures and interest rates. The Journal of Finance, 60(5), 2283–2331.
[Crossref]
Clewlow, L., & Strickland, C. (2000). Energy derivatives: Pricing and risk management. Lacima
Publications.
Deaton, A., & Laroque, G. (1992). On the behavior of commodity prices. The Review of Economic
Studies, 59(1), 1–23.
[Crossref][zbMATH]
Dempster, M. A. H., Medova, E., & Tang, K. (2012). Determinants of oil futures prices and
convenience yields. Quantitative Finance, 12(12), 1795–1809.
[MathSciNet][Crossref][zbMATH]
Dvir, E., & Rogoff, K. (2009). Three epochs of oil, NBER Working Paper, 14927.
Eydeland, A., & Wolyniec, K. (2003). Energy and power risk management: New developments in
modeling, pricing, and hedging. Wiley.
Fernandez-Perez, A., Fuertes, A.-M., & Miffre, J. (2021, Summer). On the negative pricing of WTI
crude oil futures. Global Commodities Applied Research Digest, 6(1), 36–43.
Gibson, R., & Schwartz, E. S. (1990). Stochastic convenience yield and the pricing of oil
contingent claims. The Journal of Finance, 45(3), 959–976.
[Crossref]
Gustafson, R. L. (1958). Carryover levels for grains, U.S. Department of Agriculture, Technical
Bulletin, 1178.
Hamilton, J. D. (2009). Understanding crude oil prices. The Energy Journal, 30(2), 179–206.
[Crossref]
Interim Stuff Report. (2020). Trading in NYMEX WTI crude oil futures contract leading up to, on,
and around April 20, 2020, Commodity Futures Trading Commission, November 23.
Kilian, L. (2020). Understanding the estimation of oil demand and oil supply elasticities, Federal
Reserve Bank of Dallas Working Paper, 2027.
Ma, L. (2022). Negative WTI price: What really happened and what can we learn? The Journal of
Derivatives, 29(3), 9–29.
[MathSciNet][Crossref]
Miltersen, K. R. (2003). Commodity price modelling that matches current observables: A new
approach. Quantitative Finance, 3(1), 51–58.
[MathSciNet][Crossref][zbMATH]
Routledge, B. R., Seppi, D. J., & Pratt, C. S. (2000). Equilibrium forward curves for commodities.
The Journal of Finance, 55(3), 1297–1338.
[Crossref]
Schwartz, E. S. (1997). The stochastic behavior of commodity prices: Implications for valuation
and hedging. The Journal of Finance, 52(3), 923–973.
[Crossref]
Williams, J. C., & Wright, B. D. (1991). Storage and commodity markets. Cambridge University
Press.
[Crossref]
Working, H. (1948). Theory of the inverse carrying charge in futures markets. Journal of Farm
Economics, 30(1), 1–28.
[Crossref]
Working, H. (1949). The theory of price of storage. The American Economic Review, 39(6), 1254–
1262.
Footnotes
1 Working’s original observations on commodity storage were published between 1929 and
1933 in a series of short articles issued by Wheat Studies of the Food Research Institute. The
more complete version of his argument is summarized in Working (1948) and Working (1949).
The approach was further extended by Brennan (1958).
2 For simplicity, we use the terms convenience benefits and convenience yield interchangeably
and refer to the previous chapter for their definitions.
3 The theory of storage was pioneered by Gustafson (1958) and subsequently extended in
multiple directions. The formulation that we present here is largely based on the ideas of
Deaton and Laroque (1992). This method has been applied to the oil market by Dvir and Rogoff
(2009). For other related methods of solving this problem, see Williams and Wright (1991),
Routledge et al. (2000), and Pirrong (2012).
4 More formally, the price elasticity of demand is defined as . While it is difficult to
measure this elasticity precisely, it has been steadily declining over time, as increasing overall
wealth made consumers less sensitive to energy prices. See, for example, Hamilton (2009) and
Kilian (2020).
5 The assumption that the futures price is equal to the expected spot price implies that the
futures price is fair and unbiased. It is equivalent to the so-called risk-neutral pricing that we
will formally introduce in later chapters. In the next chapter, we also consider the case when
futures price differs from the expected spot price, as futures may be distorted by imbalances in
the hedging market.
6 Brennan and Schwartz (1985) applied a one-factor lognormal model for the spot price with
deterministic convenience yield to derive futures prices. The model was further extended in
Brennan (1991). One-factor models, however, have quickly proven to be too restrictive, as they
allow futures across all maturities to move only in the same direction. A popular two-factor
model of Gibson and Schwartz (1990), which assumes a stochastic mean-reverting convenience
yield, generates a much richer dynamics for the futures curve and volatilities. Miltersen (2003)
allowed the equilibrium convenience yield to be time-dependent and showed how to make the
model consistent with the futures curve. Several three-factor models have been proposed, such
as Schwartz (1997), Casassus and Collin-Dufresne (2005), and Dempster et al. (2012). For
surveys of reduced-form models, we refer to Clewlow and Strickland (2000), Eydeland and
Wolyniec (2003), and Carmona and Ludkovski (2004).
7 In Bouchouev (2021), the author applied a slightly different methodology to this problem by
representing the futures spread as the solution to the Black-Scholes-Merton (BSM) partial
differential equation, where inventory is a state variable. Here, we use a somewhat simplified
approach, as BSM equation is formally introduced only later, in Chap. 8.
8 Some storage was already committed but not yet reflected in inventory data, nevertheless,
even including these additional volumes, the total inventory levels were substantially below the
maximum operating capacity.
9 For additional discussions of the episode of negative prices, see Interim Stuff Report (2020),
Bouchouev (2020), Fernandez-Perez et al. (2021), and Ma (2022).
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023
I. Bouchouev, Virtual Barrels, Springer Texts in Business and Economics
https://doi.org/10.1007/978-3-031-36151-7_4
The business model of the risk bearer is very different from the
business of price forecasting. Risk bearing resembles the business of a
casino dealer. The primary requirement for the risk bearer is an ability
to provide capital and withstand short-term losses resulting from
random fluctuations in spot prices. Since the spot price is likely to
mean-revert, as implied by the theory of storage, cumulative spot price
changes in the first term of (4.1) are expected to net out. In the
meantime, the contribution of the second term steadily accumulates as
long as the market remains backwardated. This is the reason why
casinos tend to win; they stay in the game and repeat the same bet with
a tiny statistical edge over and over. For them, wins and losses of
individual gamblers, or aspiring prophets, only represent noise that
eventually cancels out. By the law of large numbers, the casino’s
statistical edge turns into profits over time. For risk bearers in
commodity markets, this casino-like edge comes from the
backwardated shape of the futures curve.
We next quantify the Keynesian hypothesis of normal
backwardation and expected profits for risk absorbers with a simple
analytical framework based on the rational behavior of market
participants.
4.2 The Hedging Equilibrium
Let us consider a simplified model of the futures market which is made
up of three participants: a producer, a consumer, and a speculator. We
assume that they follow a conventional one-period mean-variance
decision-making framework, where participants are maximizing their
expected wealth, , while being penalized for risks measured by
(4.2)
Solving this equation for Np, we obtain that the producer optimal
position in the futures market is given by
(4.3)
Here,
(4.5)
Adding up Eqs. (4.3), (4.4) and (4.5) for optimal futures positions held
by producer, consumer and speculator, we obtain that
(4.6)
The Eq. (4.6) explicitly relates the Keynesian risk premium to the
aggregate hedging imbalance between producers and consumers. Given
that the producer has a much more concentrated price exposure than
the consumer, the producer propensity to hedge is likely to be higher,
and Qp ≫ Qc. This implies that in this simple market model, the
resulting oil risk premium in the Eq. (4.6) is indeed expected to be
positive, as is suggested by the theory of normal backwardation.
The size of the risk premium depends on the volatility of the market
and the risk aversion of all market participants. The risk tolerance of
speculators cannot change the sign of the risk premium, but the
speculators’ lower risk aversion can dilute the magnitude of the
imbalance. The risk in this hedging equilibrium model is shared among
all market participants. The parameter α can be thought of as the
market collective risk aversion coefficient. It is proportional to the
harmonic mean of the individual risk aversions. Remarkably, the idea of
using a harmonic mean in risk sharing goes back to Ancient Greece and
Aristotle, who considered the price of voluntarily exchanged goods to
be fair if it is calculated as the harmonic mean of the buyer’s bid and the
seller’s offer.8
To validate the existence of the structural risk premium in the oil
market, we now test the hypothesis of normal backwardation
empirically.
Fig. 4.5 US inflation measured by year-over-year changes in CPI is largely explained by annual
changes in the price of oil
Fig. 4.7 Cumulative performance of long WTI futures during the regime of normal contango
The futures market is a zero-sum game. For every loser, there must
be a winner. Where did the negative risk premium go to? The answer
has already been given in the previous chapter. Selling futures at a
premium to financial investors and buying physical barrels is the
business of the professional storage trader. From being a provider of a
service to oil producers, investors themselves turned into hedgers of
financial risks. Investors pay for this service with the risk premium by
rolling futures in the contango market with profits accruing to the
storage owner. The investor looking to buy futures must incentivize
new marginal suppliers of short futures to come to the market, if
existing short futures offered by producer hedgers are not sufficient.
These incentives have also provided the impetus for building new
storage capacity, which we discuss later in the context of specific
trading strategies in Chaps. 6 and 13.
Since investors cannot easily buy and store physical barrels
themselves, they have no choice but to pay for the privilege to invest in
oil. These payments are not explicit. Every month when the futures
curve is contango, investors must sell prompt futures before their
expiration, and reset the long position at a higher price by buying the
next available futures contract. Even though no money is exchanged on
the day of the roll, losses are accumulated over time as futures tend to
roll down towards the lower spot price. The convenience yield of
owning a physical commodity, net of storage costs, has turned negative.
The intrinsic yield of oil became much more of an inconvenience. While
the negative roll yield acted as the synthetic cost of storage, it also
compensated storage owners for providing the service of essentially
storing the bulky commodity on behalf of investors.
With the addition of another major market participant, the manager
of the broader financial portfolio, the hedging equilibrium Eq. (4.6)
must be revisited. An investor, such as a risk parity find, now has an
inflation-hedging mandate that counters the producer’s need to sell
futures. Let us assume that such an investor is tasked to manage the
notional inflation exposure Qi in a financial portfolio that generates
income Ui while being exposed to inflation uncertainty . The investor
follows the same mean-variance framework with risk aversion
coefficient αi, and trades Ni oil futures to hedge the inflation risk. The
inflation hedger’s wealth is then given by19
Following the same steps as before, we find the futures position Ni that
maximizes the following expression for the investor wealth function
(4.7)
where
(4.8)
The theory of hedging pressure is now generalized to cover all the main
market participants. The resulting risk premium is a function of the net
imbalance between producers, consumers, and inflation hedgers.
Traditional speculators are incentivized to take either long or short
positions depending on the sign of the net hedging imbalance. Since
financial markets are significantly larger than oil markets and the
demand for inflation hedging has steadily increased over time, the
structural risk premium during the era of financialization has become
negative. In this case, systematically buying and rolling oil futures is a
losing value proposition.
All good things in financial markets eventually come to an end.
Storage operators and short speculators enjoyed a long run of
spectacular performance during the regime of normal contango.
Financial investors, in turn, have learned their lessons and became
more dynamic with their oil investment strategies. By and large, the oil
markets have largely matured, and the structural directional oil risk
premium has vanished. It became nimbler, responding to faster changes
in fundamentals of storage and hedging imbalances. To capture it,
investment strategies must become dynamic as well. These strategies
employed by different types of investors and speculators are presented
in the following three chapters of the book.
References
Acharya, V. V., Lochstoer, L. A., & Ramadorai, T. (2013). Limits to arbitrage and hedging:
Evidence from commodity markets. Journal of Financial Economics, 109(2), 441–465.
[Crossref]
Ashton, M., & Greer, R. (2008). History of commodities as the original real return asset class. In
Inflation risk and products (pp. 85–109). Risk Books.
Bodie, Z., & Rosansky, V. I. (1980, May–June). Risk and return in commodity futures. Financial
Analysts Journal, 36(3), 27–39.
Bouchouev, I. (2020). From risk bearing to propheteering. Quantitative Finance, 20(6), 887–894.
[MathSciNet][Crossref]
Bü yü kşahin, B., & Robe, M. A. (2014). Speculators, commodities and cross-market linkages.
Journal of International Money and Finance, 42, 48–70.
[Crossref]
Cheng, I.-H., Kirilenko, A., & Xiong, W. (2015). Convective risk flows in commodity futures
markets. Review of Finance, 19(5), 1733–1781.
[Crossref]
Erb, C. B., & Harvey, C. R. (2006). The strategic and tactical value of commodity futures.
Financial Analysts Journal, 62(2), 69–97.
[Crossref]
Fama, E. F., & French, K. R. (1987). Commodity futures prices: Some evidence on forecast power,
premiums, and the theory of storage. Journal of Business, 60(1), 55–73.
[Crossref]
Fattouh, B., & Mahadeva, L. (2014). Causes and implications of shifts in financial participation in
commodity markets. The Journal of Futures Market, 34(8), 757–787.
[Crossref]
Gorton, G. B., Hayashi, F., & Rouwenhorst, K. G. (2012). The fundamentals of commodity futures
returns. Review of Finance, 17(1), 35–105.
[Crossref]
Gorton, G., & Rouwenhorst, K. G. (2006). Facts and fantasies about commodity futures. Financial
Analysts Journal, 62(2), 47–68.
[Crossref]
Greer, R. J. (1978, Summer). Conservative commodities: A key inflation hedge. Journal of
Portfolio Management, 4(4), 26–29.
[Crossref]
Hamilton, J. D. (1983). Oil and the macroeconomy since World War II. Journal of Political
Economy, 91(2), 228–248.
[Crossref]
Hamilton, J. D., & Wu, J. C. (2014). Risk premia in crude oil futures prices. Journal of
International Money and Finance, 42, 9–37.
[Crossref]
Hicks, J. R. (1939). Value and capital: An inquiry into some fundamental principles of economic
theory. Oxford University Press.
Hirshleifer, D. (1988). Residual risk, trading costs, and commodity futures risk premia. The
Review of Financial Studies, 1(2), 173–193.
[MathSciNet][Crossref]
Keynes, J. M. (1923, March 29). Some aspects of commodity markets. The Manchester Guardian
Commercial, Reconstruction Supplement.
Kilian, L. (2009). Not all oil price shocks are alike: Disentangling demand and supply shocks in
the crude oil market. American Economic Review, 99(3), 1053–1069.
[Crossref]
Kilian, L., & Murphy, D. P. (2014). The role of inventories and speculative trading in the global
market for crude oil. Journal of Applied Econometrics, 29(3), 454–478.
[MathSciNet][Crossref]
Kilian, L., & Vigfussion, R. J. (2017). The role of oil price shocks in causing U.S. recessions.
Journal of Money, Credit and Banking, 49 (8), 1747–1776.
[Crossref]
Neville, H., Draaisma, T., Funnell, B., Harvey, C. R., & Van Hemert, O. (2021). The best strategies
for inflationary times, SSRN.
Stoll, H. R. (1979). Commodity futures and spot price determination and hedging in capital
market equilibrium. The Journal of Financial and Quantitative Analysis, 14(4), 873–894.
[Crossref]
Tang, K., & Xiong, W. (2012). Index investment and the financialization of commodities.
Financial Analysts Journal, 68(6), 54–74.
[MathSciNet][Crossref]
Till, H., & Eagleeye, J. (Eds.). (2007). Intelligent commodity investing. Risk Books.
Footnotes
1 Among many studies of commodity markets in the 1920s, the newspaper article published
by Keynes (1923) is credited with a particularly significant contribution. In this article, the
author highlighted that the value of agricultural inventories after the harvest is so large relative
to financial resources of producers that they cannot themselves bear the risk of inventory value
to drop. Thus, the imbalance requires the service of professional speculators. Keynes was
careful to distinguish between agricultural and extraction commodities, such as oil, for which
the demand for temporary credit is lower due to the ratable nature of production.
3 The term normal backwardation was introduced in Keynes (1930), volume II, chapter 29.
4 Keynes (1923).
5 Keynes (1923).
6 This approach follows the mean-variance hedging framework originally developed by Stoll
(1979) and Hirshleifer (1988).
7 The theory of hedging pressure has been combined with the canonical theory of storage by
Gorton et al. (2012), Acharya et al. (2013), and Baker (2021).
9 Several attempts have been made to provide individual drivers with hedging instruments,
typically by selling them prepaid gasoline cards to be used at retail gas stations, but such
attempts never gained the economy of scale.
10 The term fully funded means that the entire notional value of the futures contract is held as
collateral. We use log-returns which are additive for the reasons of analytical tractability, as
explained in Chap. 2. While it is not recommended to average simple returns, the annualized
arithmetic average of monthly returns over this period would have been even higher at 15.6%.
These returns do not include any additional return on collateral.
11 A typical initial cash margin requirement for energy futures is 10–20% of the contract’s
notional value. Therefore, the returns on cash required to hold futures are 5–10 times higher
than returns on a fully funded position.
12 The idea of investments in diversified commodity indices has been covered extensively in
the literature. For earlier studies of this subject that predate the introduction of energy futures,
see Greer (1978), Bodie and Rosansky (1980), and Fama and French (1987). The addition of
petroleum futures markedly improved the performance and the overall attractiveness of
commodity indices for investors, as documented in the influential work of Gorton and
Rouwenhorst (2006), and Erb and Harvey (2006). See also Till and Eagleeye (2007), Ashton
and Greer (2008), Tang and Xiong (2012), Fattouh and Mahadeva (2014), Bü yü kşahin and Robe
(2014), Hamilton and Wu (2014), Cheng et al. (2015), and references therein.
13 In the USA, a recession is formally defined by the National Bureau of Economic Research
(NBER) as “a significant decline in economic activity spread across the market, lasting more
than a few months, normally visible in real gross domestic product (GDP), real income,
employment, industrial production, and wholesale-retail sales.” In practice, two consecutive
quarters of negative GDP growth is often taken as an indicator of a recession.
15 Hamilton (2003) developed a nonlinear metric for net oil price increases and applied it to
explain oil price shocks mostly with supply disruptions driven by geopolitical events. In
contrast, Kilian (2009) attributed a much larger role in many oil shocks to growth in global
demand for commodities and to the so-called precautionary demand, which is associated with
speculative buying in anticipation of rising uncertainty and shifts in future expectations. The
latter approach was formalized via a structural vector autoregressive model (VAR) in Kilian and
Murphy (2014) that decomposes contribution of supply disruptions, business cycle-related
demand, and speculative oil-specific demand for inventories to the real price of oil. Another
VAR model was used by Kilian and Vigfussion (2017) to quantify the contribution of oil shocks
to past US recessions.
19 This approach to the derivation of the hedging equilibrium was proposed in Bouchouev
(2020).
Part II
Quantitative Futures Strategies
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023
I. Bouchouev, Virtual Barrels, Springer Texts in Business and Economics
https://doi.org/10.1007/978-3-031-36151-7_5
The momentum trading signal, πM(Ft), buys and sells futures Ft based
on the sign of Mt4
(5.1)
where
Fig. 5.1 Historical returns of a 20-day momentum strategy for WTI futures (1993–2022)
At first glance the average annual return of 10.4% generated by this
simple strategy over such a long time may look appealing. However, it is
less attractive on a risk-adjusted basis, with an information ratio of only
0.27.6 In four years, the strategy drawdowns substantially exceeded
25%, which often marks the maximum loss that investors are willing to
tolerate for any hedge fund investment strategy. While the momentum
signal clearly has some informational content, the strategy is unlikely to
be good enough to be traded on a stand-alone basis.
One of the main objectives of this book is not only to share ideas
that have a proven track record, but also to highlight their pitfalls.
While momentum is a useful trading concept, it is, unfortunately, one
that is often abused. In systematic trading, such abuse comes from data
mining and overfitting when the trading signal is allowed to have too
many degrees of freedom that are fitted to produce better-looking
backtests. The more parameters a trading strategy has, the more likely
it is to fall apart when some of these parameters are modified.
Momentum strategies provide a good case study.
The most common extension of basic momentum is to smooth the
signal sensitivity to the latest data point. Financial markets are full of
noise, and the momentum trader may want to avoid being whipsawed
by false signals and instead to wait for some confirmation of the trend
to be firmly established. Such smoothing is typically done by replacing
the latest price in the definition of momentum with another, shorter-
term moving average. The crossover momentum signal is generated by
the difference between moving averages
and
Fig. 5.2 The weights for the M(m, 20) momentum strategy peak for the price change that
occurred exactly m − 1 days ago
Fig. 5.3 An example of instability of crossover momentum strategies with respect to short-
term moving average. In 2016, M(1,20) lost 40%, M(3,20) made 20%, while returns on M(2,20)
and M(4,20) strategies were close to zero
There is nothing structural about this choice of parameters and one
can easily find another period when the reverse was true. The
difference is typically driven by how well the parameter optimizer
captures a small number of large returns that tend to dominate the
overall strategy performance. On the bright side, the sensitivity to the
longer-term moving average, n, becomes significantly smaller as the
weights on the corresponding price changes monotonically decrease.
Our goal here is not to find the best combination of parameters for
the oil momentum strategy. We believe that this would be an ill-
intentioned task given the non-stationarity of the financial time series.
It is clear that the oil momentum does exist. It contains some useful
information that one can see even in the historical performance of the
basic momentum strategy. A few other formulations of momentum
strategies are also worth mentioning, not because they are better, but
more because they are often cited by researchers who claim to have
found some other magic versions of momentum.
One popular variation of momentum is a class of breakout strategies
where the asset is bought or sold only when the latest price moves
beyond its prior maximum or minimum over some given lookback
period. In contrast to the basic momentum, the breakout strategy is
likely to remain on the sidelines for substantial periods of time, which,
unfortunately, makes it even more prone to overfitting. Further
variations could include filters to keep the moving averages crossover
intact for a certain number of days to avoid being whipsawed before
taking a position. Alternatively, one can impose the filter to enter the
trade only when the spread between two moving averages exceeds a
certain threshold, which becomes an additional parameter as well.
Similar conditions can be imposed around the exit of the position.
There are a few other general techniques that are helpful for
momentum trading regardless of its exact specification. One can often
benefit from combining momentum signals across multiple trading
frequencies. For example, one can use the crossover of the latest price
and its weekly price average as the short-term momentum signal, the
weekly versus monthly averages as the medium-term signal, and the
monthly versus annual moving averages as the longer-term momentum
signal. Then the three signals can be combined into an aggregate
momentum score which determines the overall size of the trading
position. The weights for each of the three momentum frequencies can
also be customized.
Undoubtedly, the more parameters the strategy backtest is allowed
to have, the better the fitted results will inevitably be. However, if one
normalizes numerous permutations of momentum strategies for the
same degrees of freedom, then their long-term performances tested
out-of-sample become broadly similar as well. Any claim of the
discovery of some magic momentum specification that works much
better than any other momentum strategies should be treated with a
great amount of skepticism. If one is sufficiently excited about the
rationale of the oil momentum strategy, then the best advice would be
not to chase the best backtest but to keep the strategy simple and to
avoid overfitting.
One of the reasons that explains the popularity of the oil
momentum strategy is its negative correlation to directional financial
investments during synchronized risk-off events when any benefits of
portfolio diversification are particularly appreciated. For example, a
strategy of buying and rolling oil futures lost over 50% in four years
since 2000, but simultaneous investments in a momentum strategy
would have largely offset these losses. Figure 5.4 highlights these years
with blue dots. Since momentum strategies tend to do well when prices
make large moves in either direction but underperform when prices
are range bound, the graph of momentum returns versus price returns
is sometimes referred to as momentum smile.
Fig. 5.4 Momentum strategies would have mostly offset large losses from directional
investments in oil futures (WTI, 2000–2022)
One important topic for systematic trading that we have not yet
discussed is the sensitivity of the strategy to transaction costs.
Fortunately, for liquid oil futures, such as WTI and Brent, the existence
of the TAS mechanism described in Chap. 3 makes the task much easier
as it allows systematic traders to lock in the settlement price practically
without crossing the bid-ask spread. We should note that similarly set
up minute marker contracts also allow traders to lock in prices during
pricing windows for physical markets in the European and Asian time
zones. The European prices are often used by managers of global
systematic portfolios to capture the best simultaneous liquidity across
the three main time zones. One can also run faster intraday momentum
strategies based on higher-frequency data. Such strategies are outside
the scope of this book, as they depend more on the market
microstructure than on the behavior of participants.
To summarize, one can consider practically infinite permutations
for different momentum strategies for oil futures, but in the long run,
most exhibit similar predictive power once the somewhat artificial
contribution driven by additional degrees of freedom is removed. Oil
momentum does add some useful information content, but its strength
appears to be declining as markets become more liquid and mature.
Practitioners observed that momentum strategies generally work
better in less liquid markets, where the information diffuses more
slowly. However, the execution of such strategies in illiquid markets is
more challenging due to higher transaction costs.
The oil momentum can be used to enhance other strategies with
more solid economic foundations, but it should not be abused by data
mining. As we will see later, the momentum in liquid oil markets can
still add significant value when it is combined with other drivers that
have stronger economic foundation. Momentum is a great optimizer of
other trading signals and should be used as such. As a stand-alone
trading strategy, it is not sufficiently robust and the economic
foundation behind the concept also remains rather weak. As traders
often quip, momentum works until it does not.
and the risk premium is determined by the carry, which is the value of
the spot-futures spread, as observed at time t.
The concept of systematic carry trading came from foreign exchange
markets. It hinges on the same principle of no-arbitrage between two
alternative money standards that Irving Fisher illustrated for the case
of the commodity money, as it was explained in Chap. 2. The argument
applies to fiat money of different countries as well. Each currency pays
its own domestic rate of interest. The forward value of a currency with
a relatively high interest rate must be lower relative to a currency with
a lower interest rate to make investors indifferent in which currency to
hold their money. According to an economic theory, the magnitude of
the forward discount, or an expected currency depreciation, must
negate the value that can be gained from interest rate differentials, the
hypothesis known as an uncovered interest rate parity. In practice,
however, this economic theory does not hold. In the long run, holding
riskier currencies with higher yields is often more profitable, as on
average, expected currency depreciation is not fully realized. Obviously,
in the short-term such a strategy can experience large losses, but over
time investors are generally able to extract positive risk premium from
the foreign exchange carry trade.
Using the broad definition of carry as the strategy return when
nothing changes, the idea of carry trading can be extended to almost
every asset class. In equities, carry is simply given by the dividend yield.
The future stock price is then discounted relative to the current stock
price by the amount of the expected dividend, net of the risk-free rate.
The carry trade in the equity market buys single-stock futures that have
larger discounts relative to the current stock price, which is equivalent
to buying high-dividend stocks. The dividend yield behaves like the
foreign interest rate in the currency market, or like the convenience
yield of holding oil inventories. In the fixed income market, carry is
simply the coupon received from holding the bond, or the roll down in
the interest rate swap curve. The fixed income carry strategy then buys
bonds with higher coupons and steeper interest rate curves, where
short-term rates are typically lower than long-term rates.7
The power of carry in the oil market has already been shown in the
previous chapter. When carry is positive and the oil market is in
backwardation, buying oil futures tends to be a profitable strategy. As
the futures curve moves into contango and the carry becomes negative,
the investor returns deteriorate. Since, in the futures market, selling is
as easy as buying, one can replace the long-only futures investment
with the dynamic strategy of buying and selling based on the direction
of carry. Like momentum, the same idea can be applied across all
commodities to achieve some diversification, but the strategy again
works better for energy commodities. Oil carry is also rooted in the
theory of storage. Unlike momentum, however, the transmission of the
carry theory to the futures market is much more explicit. An oil carry
plays a unique role in the market as it transforms the fundamentals of
supply and demand into pressure on prices via the behavior of
inventory hedgers.
The business model of the storage trader is to generate low-risk
steady returns by providing the service of storing oil. In such a business
model, the risk of volatile oil prices must be eliminated by hedging in
the futures market. As soon as a physical barrel is purchased for
storage, a financial barrel is sold in the futures market to reduce the
risk. This downward pressure on futures price occurs when the
contango is steep enough to cover the cost of storage. If the market flips
to backwardation or the contango becomes too narrow to economically
hold inventories, then the storage hedger is incentivized to pull oil out
of storage and buy back short futures hedges. This creates an upward
pressure on the futures price. The carry transmission channel is
illustrated in Fig. 5.5.
Fig. 5.5 An inventory hedger sells futures when contango covers the cost of storage and buys
them back otherwise
Such buying and selling of futures by inventory hedgers driven by
the shape of the futures curve is rather mechanical and somewhat
predictable. The cost of carry, which includes transportation, logistics,
and borrowing costs, and the convenience yield of holding inventories
depend on the economics of individual traders, which makes hedging
pressure vary with the steepness of the curve. In general, the steeper
the contango, the more widespread hedging becomes among inventory
managers, which increases the selling pressure in the futures market.
However, when carry becomes too steep, most of the hedging is already
finished, and the pressure on price starts waning. We address this
important transition point in more detail in the last section of this
chapter.
To define the systematic carry signal more formally, let
represent the spread between the two futures contracts with M and
N months to expiration. The basic carry strategy buys futures when
carry Ct is positive, and sells futures when it is negative, i.e., the trading
signal πc can be defined as
(5.2)
The definition of the carry signal again does not have to use the same
contract that is traded. Like in the case of the momentum strategy,
multiple carry signals computed from different parts of the futures
curve can be combined to generate a single aggregate carry signal.
However, in the case of crude oil multiple carry signals measured from
different parts of the futures curve tend to have the same sign and
become largely redundant. If the futures curve is monotonically
decreasing or increasing, then the sign of each monthly carry signal is
the same for all M and N.
Figure 5.6 shows the historical returns of the benchmark WTI carry
strategy that buys and sells prompt futures contracts based on the sign
of the annual carry Ct(1, 13). The twelve-month spread is commonly
used to define carry to eliminate seasonal effects, which are more
substantial for refined products and natural gas. As before, the futures
are rolled at the end of the month. Since both momentum and carry
strategies maintain either a long or a short position at all times, the two
strategies have the same volatility. The carry strategy, however,
generated a much higher annualized log-return of 20.3% over the
thirty-year period with impressive Sharpe ratio of 0.53. In addition, the
drawdowns of the carry strategy are significantly smaller than the
drawdowns of the momentum strategy.
Fig. 5.6 Historical returns of the C(1,13) carry strategy for WTI futures (1993–2022)
Carry, or the shape of the futures curve, is the ultimate mechanism
that transmits fundamental information into futures prices. When we
read media reports that the price of oil went up because of stronger
Asian demand, or it went down because of increased OPEC production,
these reports do not tell the full story. OPEC itself does not buy futures
and very few oil consumers do. However, any change in supply and
demand impacts the spot-futures spread via variations in convenience
yields and storage costs. The actual futures are bought and sold by
storage traders, who react to the economics determined by the carry.
We will see in the following chapter that an inventory hedger is indeed
the largest trader in the oil futures market. The systematic carry
strategy described in this section essentially trades ahead of
anticipated behavior of the largest market participant, who is acting in
response to the arrival of new fundamental information.
Compared with the momentum strategy, carry trading has one
tremendous advantage. The strategy is essentially model-free. Unlike
momentum, carry is a forward-looking signal. It does not depend on
history and does not require any parameter estimation. The trading
signal is directly observable from the current shape of the futures
curve. It measures the economic incentives of inventory hedgers and,
by and large, the strategy front-runs the expected flows of hedgers in
the futures market.
The idea of oil carry is probably one of the most powerful and
widely used indicators for directional oil trading. Many physical traders
simply refuse to take any positions that go against the direction of
carry. A similar idea of investing in oil only when the carry is positive
has found some interesting applications in broader financial
portfolios.8
The primary drawback of the carry strategy is its inability to react
fast enough to rapidly changing market conditions. Since the shape of
the futures curve does not flip frequently between contango and
backwardation, the carry signal often retains the same sign for a long
time. Even though it may capture early gains quickly when the shape of
the curve changes, it then often gives up a portion of the gains when the
market starts correcting. By holding and rolling the same directional
position until the time spread crosses zero, or some pre-defined
threshold, the carry strategy often misses early warning signs that
fundamentals are beginning to change. To capture an anticipated
change in fundamentals, traders often look at the change in carry
instead of its level. One can think about carry being a measure of the
current state of supply and demand, and the change in carry as a
measure of expected changes in future supply and demand.
A popular way to make the carry strategy more dynamic is to blend
it with the basic momentum signal. This can be accomplished by
applying momentum not to the price of oil, but instead directly to carry,
or to the shape of the futures curve. The carry-momentum trading
signal πCM is then defined as
(5.3)
The results from such simple signal blending, shown in Fig. 5.7, are
quite remarkable. The strategy, which we call the dynamic carry, or
carry-momentum, generated 24.7% annualized returns with the Sharpe
ratio of 0.64 over the thirty-year period. Out of hundreds of different
blends of systematic signals that the author has traded over many
years, the carry-momentum signal stood out in its long-term
robustness. It highlights how a technical indicator, such as momentum,
can improve a trading concept that has stronger links to the market
fundamentals.
Fig. 5.7 Historical returns of the 20-day carry-momentum CM(1,13) strategy for WTI futures
(1993–2022)
This strategy buys futures when the price falls below its moving
average and sells when it rises above. Obviously, since the momentum
strategy is generally profitable, systematically running the opposite
value strategy would be a loser.
To capture the value risk premium, traders usually run their
strategies on slower frequencies. While it is critical for momentum
traders to react fast and trade as soon as the price crosses some trigger,
value traders generally prefer to remain patient and avoid rushing into
a trade. They often let the market run in the same direction for some
distance and only trade when the deviation exceeds a certain threshold,
ε. Such a threshold-based value signal can be defined as:
(5.4)
In this variation of the value strategy, the trader may remain on the
sidelines for a while and bet only when the deck is rich, i.e., when the
price deviates sufficiently far away from its normal range.
In general, calling the tops and the bottoms for the price of oil is not
a prudent strategy in the futures market, as one must be very precise
on the timing of anticipated price reversals. If one could trade physical
oil and keep some oil in storage without incurring any cost, then selling
high and buying low would undoubtedly work well, as spot price does
eventually mean-revert. However, storage is not free, and it is rarely
accessible by financial traders, so that the closest proxy that one can
trade is the futures contract. Buying low and selling high in the futures
market, however, is very different from doing it in the spot market. In
the futures market, the value trader constantly fights against the
punitive cost of carry, which is determined by the slope and the
convexity of the futures curve.
If the market is oversupplied, then the price of prompt futures is
likely to be lower than the forward price. The slope of the futures curve
reflects the cost of storage, which is typically higher for shorter-
maturity futures, where the fundamental imbalance is the most acute.
Storage buys time either for consumption to increase or for production
to be curtailed. As time moves forward and fundamentals normalize,
the cost of storage decreases, and consequently the slope of the futures
curve flattens. Since the slope, which is the first derivative with respect
to futures maturity, is decreasing, then the futures curve in an
oversupplied market is expected to be concave. We call this a state of
concave contango.
Likewise, at times when inventories are low, the slope of the futures
curve is the steepest in the front-end of the curve, which is dominated
by the highest convenience yield of owning a physical barrel. As
inventories normalize over time, the convenience yield decreases,
which makes the forward slopes flatter. This is a state of convex
backwardation. The futures curves that are shown in Fig. 5.5 are in
their normal states of concave contango and convex backwardation.
The convexity of the futures curve creates a dilemma for the value
trader: whether to use shorter-term futures and lock in better value at
the expense of a steeper carry, or trade further out on the curve where
the carry is less punitive, but the value is also less attractive. Figure 5.8
illustrates the challenge of buying cheap futures when the market is in
concave contango.
Fig. 5.8 Buying cheaper short-term futures in concave contango is offset by steeper carry,
while buying long-term futures has less value
The case of selling futures in convex backwardation is handled
similarly, as the mirror image of concave contango. In either case, for
the value strategy to be profitable the magnitude of the expected price
mean-reversion must dominate the accumulation of carry over the
holding period. One should avoid using short-term futures for
capturing value unless the reversal is deemed to be imminent as
indicated by other fundamental factors. Otherwise, time is not on the
side of the value trader. The apparent cheapness of the futures contract
could quickly become overpowered by the adverse buildup of the
negative carry. These types of strategies where a systematic signal is
best combined with a discretionary overlay based on fundamental and
flow factors are discussed in the next chapter.
In theory, it is possible for both systematic momentum and value
strategies to generate positive returns if the two are traded on different
frequencies. Momentum tends to work better on the shorter time scale,
while the contrarian value signal works better in the longer run,
especially when an additional buffer is provided by a threshold ε. One
can also combine momentum, carry and value into signal-integrated
portfolios. This is typically done on a cross-sectional basis within
broader commodities portfolio.
Overall, momentum and carry signals tend to overpower value in
directional oil trading, but in the next chapter we will show that value
strategies are more robust in trading futures spreads and construct one
of such diversified energy value portfolios. For now, we continue to
focus on single-asset strategies and conclude this chapter by showing
another way of blending multiple signals that can help systematic
traders to determine the optimal size of their bet.
References
Bakshi, G., Gao, X., & Rossi, A. G. (2019). Understanding the sources of risk underlying the cross
section of commodity returns. Management Science, 65(2), 619–641.
[Crossref]
Boons, M., & Prado, M. P. (2019). Basis-momentum. The Journal of Finance, 74(1), 239–279.
[Crossref]
Daskalaki, C., Kostakis, A., & Skiadopoulos, G. (2014). Are there common factors in individual
commodity futures returns? Journal of Banking and Finance, 40, 346–363.
[Crossref]
Fernandez-Perez, A., Frijns, B., Fuertes, A.-M., & Miffre, J. (2018). The skewness of commodity
futures returns. The Journal of Banking and Finance, 86, 143–158.
[Crossref]
Koijen, R. S. J., Moskovitz, T. J., Pedersen, L. H., & Vrugt, E. B. (2018). Carry. Journal of Financial
Economics, 127, 197–225.
[Crossref]
Szymanowska, M., De Roon, F., Nijman, T., & Van Den Goorbergh, R. (2014). An anatomy of
commodity futures risk premia. The Journal of Finance, 69(1), 453–482.
[Crossref]
Till, H. (2022, Winter). Commodities, crude oil, and diversified portfolios. Global Commodities
Applied Research Digest, 7(2), 65–74.
Weymar, F. H. (1965). The dynamics of the world cocoa market. Ph.D. Thesis, Massachusetts
Institute of Technology.
Footnotes
1 See Weymar (1965). In addition to empirical study of cocoa prices using fundamental data,
Weymar developed one of the first theoretical models of storage, which links prices and
inventories via coupled differential equations.
2 The fund’s trading roster included the names of Paul Tudor Jones, Louis Bacon, and Bruce
Kovner, who subsequently became trading legends in their own right, creating multi-billion
dollar hedge funds known to be among the largest commodity traders. Jones founded Tudor
Investment Corporation, Bacon is the founder of Moore Capital Management, and Kovner
established Caxton Associates. More details about the history of Commodity Corp. are provided
in Tully (1981) and Lux (2003).
3 While the literature on long-short commodity risk premia portfolios is broad, we highlight
the following work by Szymanowska et al. (2014), Daskalaki et al. (2014), Fernandez-Perez et
al. (2018), Bakshi et al. (2019), and Boons and Prado (2019). See also Miffre (2016) for a
comprehensive survey of empirical results and additional references on long-short commodity
portfolios.
4 For brevity, when defining the trading signal with the sign function we assume that sign(0) =
+1, which ensures that either a long or a short position is held on each trading day.
5 RBOB stands for Reformulated Blendstock for Oxygenate Blending. It represents a primary
blending component used to create the finished gasoline product. Even though RBOB does not
represent the finished product, it is widely used as the most liquid benchmark for gasoline
prices in the futures market.
6 In general, the term information ratio is used to measure the performance of the strategy
relative to a certain benchmark. It is analogous to the conventional Sharpe ratio if the risk-free
return is replaced with the return on the benchmark. Since trading futures does not require
much initial capital, the benchmark for most strategies is typically set at zero. Therefore, here
the information ratio is simply defined as the ratio of the strategy’s annualized return to its
annualized volatility, and, for convenience, we use the term information ratio interchangeably
with the Sharpe ratio.
7 The empirical performance of the broad cross-asset carry portfolio is presented in Koijen et
al. (2018).
8 Till (2022) showed that a dynamic strategy that replaces a conventional 60–40 equity-bond
portfolio with a 30–40–30 equity-bond-oil futures portfolio when oil carry is positive
significantly outperforms the static 60–40 portfolio.
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023
I. Bouchouev, Virtual Barrels, Springer Texts in Business and Economics
https://doi.org/10.1007/978-3-031-36151-7_6
6. Quantamentals
Ilia Bouchouev1
(1) Pentathlon Investments, LLC, Westfield, NJ, USA
Fig. 6.2 The impact of roll hedging pressure on the shape of the futures curve
Fig. 6.3 Decomposition of short time spread strategy by slope and convexity. The number in
parenthesis indicates the percentage of time that the futures curve was in a given regime
and the spread carry strategy is to sell and roll the prompt spread when
the curve is concave and to buy it when the curve is convex.
This long-short strategy further improves the cumulative strategy
performance, mostly from having long spread positions, prior to the
beginning of financialization. Since 2005 nearly all profits from the
spread carry strategy came from selling spreads when the futures curve
was concave. Is the strategy performance then entirely driven by
financialization, or some other fundamental factors, such as rapid
growth of US shale production, also played a role in the structural
rolldown of WTI time spreads? To assess the impact of fundamental
factors, we next look at another important example of fractionation
analysis and decompose P&L of the time spread strategy by the level of
inventories.
Fig. 6.4 P&L of short WTI time spread strategy (2011–2022) split into deciles based on
utilized storage capacity in Cushing
At first glance, these results may look surprising. There is no visible
correlation between forward P&L of the short time spread strategy and
the previously observed level of inventories. The conventional theory of
storage only explains contemporaneous correlation between time
spreads and the level of inventories. However, contrary to a common
perception, neither inventory level nor recent changes in inventories
show any strong predictive power for the direction of time spreads.
In fact, a deeper econometric analysis shows that the reverse is true,
and time spreads have some predictive power for the future change in
inventories.3 This happens because storage traders often lock in
contango for several months ahead when the steepness of the curve
exceeds the net cost of storage for a longer time period. Consider a
trader who buys a heavily contangoed spread between one-month and
six-month futures, then takes a physical delivery on the long futures
contract and keeps oil in storage for six months. These barrels increase
oil inventory next month and decrease inventory in six months, when
oil is pulled out of storage and delivered against a short futures
contract. Therefore, the storage decision, which is made based on the
slope of the futures curve today, affects the level of inventories six
months from now. Moreover, the level of inventories in six months
depends on all prior storage decisions made based on the prevailing
shape of the futures curve.
P&L fractionation by inventory level does highlight an important
feature of the spread strategy, the impact of inventory boundaries.
When inventories drop to a low level then selling time spreads becomes
particularly dangerous as the probability of an upside squeeze
increases. In such environments, the strategy generated a small loss,
shown in the leftmost bucket of Fig. 6.4. A much larger loss in the case
of high inventories is nearly exclusively explained by an isolated event
that occurred when oil price and time spreads recovered from super-
contango levels set at the onset of the Covid-19 pandemic. In this
episode, discussed in detail in Chap. 3 and illustrated by Fig. 3.11, the
market was extrapolating a trend in inventories that would have
breached the maximum available storage capacity. Since the boundary
on storage capacity cannot be violated, supply and demand have no
choice but to adjust, and the inventory trend must stop and reverse.
This boundary phenomenon resembles the structure of the reaction
function, presented in the previous chapter. The trend in inventories
and time spreads can continue up to a certain inflection point, beyond
which the presence of the hard boundary induces the reversal.
One lesson that a quantamental trader can learn from the
fractionation analysis is that it might be better not to trade the time
spread strategy when Cushing inventories approach either boundary.
The proximity of boundaries clearly introduces additional uncertainty
and makes spread trading particularly risky. Unfortunately, the useful
insight obtained from fractionation analysis is sometimes abused by
systematic traders, who are often driven by an overarching desire to
improve the historical backtest. Imagine a conditional systematic
strategy which is short a time spread for structural reasons but the
strategy switches to a long spread position when inventories are less
than 20% or more than 80% of the storage capacity. Obviously, such a
rule-based strategy conditioned on inventories produces a substantial
improvement to the backtest, as the two negative buckets in Fig. 6.4 are
flipped from losses to gains. If the purpose of this book was to sell a
systematic strategy to an investor, we might have even presented an
optimized backtest for such a systematic strategy. Our goal, however, is
to describe how the sausage is made, which is what the fractionation
analysis reveals.
The fractionation analysis highlights the danger of data mining and
the sensitivity of any systematic strategy to additional parameters,
much like an optimized momentum crossover strategy. For a
quantamental trader, however, such analysis is a valuable tool for
decision making. It confirms the structurally short bias of the naïve
static strategy that results from the joint impact of financial rolls and
the growth of US shale production. When the two forces work in
tandem, the spreads steadily roll down. However, when inventories
decrease towards particularly low levels, then fundamental and flow
forces pull the spread in opposite directions. During such periods, the
strategy performance deteriorates. If a trading strategy makes money
when two primary driving factors point in the same direction, but does
not lose much when factors diverge, then it is definitely an attractive
strategy. In addition, the analysis emphasizes the important role of
storage boundaries.
In this section we only illustrated the curve convexity and inventory
analysis for WTI futures. While similar techniques can be used for other
energy futures, the strategy implementations are more nuanced, which
makes it difficult to apply these concepts across broader systematic
portfolios. Recall from Chap. 3 that the value of the time spread is
driven by the relative probabilities of downside and upside squeezes.
For example, in contrast to crude oil, time spreads for refined products
have an upside skew. In the case of many refined products,
fractionation of spread P&L by inventories often shows more
discernible impact of zero inventory boundary, when time spreads tend
to appreciate. The downside for time spreads on refined products is
more limited as refineries usually can cut product runs quickly in
response to rising inventories and declining profit margins. The upside,
on the other hand, does not have a well-defined boundary due to
potentially unlimited convenience yield if the market runs out of an
essential product during its peak consumption period, such as heating
oil during cold weather. The statistical analysis of refined products
inventories is, however, more challenging, due to their strong
seasonality, which significantly limits the quantity of relevant historical
data.
The case study of inventories is only one of many tools that can be
utilized by a quantamental trader. Additional insights can be gained
from a traditional fundamental analysis that involves so-called barrel
counting. Since the spreads are only contemporaneously correlated to
inventories, to forecast where spreads will be in the future, one needs
to know where inventories will be in the future. Such forecasting of
future inventories is based on counting physical barrels of oil that are
being moved by pipelines and ships and by extrapolating the data
forward using estimates for production and refinery inputs. While this
analysis is outside the scope of the book, a good quantamental trader
would always attempt to incorporate such information in a
discretionary manner to improve a rule-based trading strategy.
We next discuss another popular quantamental strategy that trades
the most important locational spread between two primary oil
benchmarks, WTI and Brent.
(6.1)
Fig. 6.6 The deviation of WTI-Brent (m3) from its moving average
This strategy has performed quite well since the beginning of the
third regime of global interconnectedness, reaching a Sharpe ratio of
nearly 1.0 with a fairly stable combination of model parameters.7 As
before, our preference is to avoid showing backtests for quantamental
strategies as it only creates an urge to improve them by introducing
additional conditions and filters. These filters can be identified by
fractionation analysis with respect to various fundamental factors, an
example of which was presented in the previous section. More
importantly, these strategies thrive only during specific fundamental
regimes, the identification of which is largely discretionary. Since oil
regimes do change rather frequently, any backtest for a quantamental
strategy can quickly become obsolete.
Certain regime features, however, are likely to be more long-lasting
and even irreversible. For example, increasing global
interconnectedness of energy markets is unlikely to disappear once
appropriate infrastructure is built. This latest regime opened a new
trading opportunity not only for WTI-Brent convergence strategies, but
also for many other fundamentally linked petroleum spreads. It allowed
the construction of a much broader mean-reverting spread portfolio
that can diversify idiosyncratic risks associated with individual spreads.
In the equity markets, such portfolios fall into the category of statistical
arbitrage, or stat-arb, strategies. In the next section, we explain how to
construct its analogue in the energy market.
Fig. 6.8 Physical arbitrage relationships actively traded within a global refined-products
portfolio
Our intent is only to provide the generic signal specification but avoid
fitting specific parameters. Given the high sensitivity of momentum
signals to the lookback period, to avoid the risk of overfitting, we would
rather steer away from optimization and let the parameters be
customized by the user.
In practice, any trading based on positioning indicators can only be
implemented with a lag. Each weekly report reflects positions held as of
Tuesday, but the report is released on Friday afternoon after the market
is already closed. If one backtests the strategy using daily closing prices,
then the earliest day when the trader can react to the latest positioning
information would be the following Monday. The existence of such a lag
is not specific to strategies based on positioning. Similar issues arise
with any non-price-based trading signals where fundamental and
economic data become publicly available only with a lag.
Another important concept in trading based on positioning
resembles the idea of an inflection point used in the reaction function
in the previous chapter. While it may pay to follow hedge funds up to a
certain point, if too many funds end up in the same bandwagon, then
their own buying or selling can bring the price to an unsustainable level
and the trend can quickly reverse. As this turning point is reached, it
might be better to enter into a contrarian trade. This strategy is known
as fading extreme positioning or fading the crowded trade. This concept
is popular among physical traders, as crowded trades are often
accompanied by large fundamentally unjustifiable price moves that
may open opportunities for physical arbitrages.
There are numerous ways of measuring the crowdedness of the
trade. One metric that has proven to be particularly useful is the so-
called sentiment index (SI). The index normalizes raw positioning data
for a certain category, for example the NC, by measuring it relative to its
previous minimum and maximum levels over a specified lookback
period, as follows
(6.2)
(6.3)
For example, one can let ε = 0.20 and take long and short positions,
respectively, in lower and upper quintiles of the sentiment index.
Like in previously discussed systematic strategies, multiple signals
can be blended. One can easily combine a follow the flow strategy with
fading extreme positioning using the idea similar to the reaction
function from the previous chapter. Here, one would follow hedge funds
up to a certain inflection point, beyond which the position is reduced
and then switched to a contrarian direction as the trade gradually
becomes crowded. Obviously, the inflection point is hard to pinpoint,
and here it becomes a strategy parameter that can be calibrated to
historical data.
Even more powerful signals can be constructed if positioning
indicators are further confirmed by other signals. For example, one can
strengthen the idea of fading the crowded momentum trade by taking a
contrarian signal only when some combination of positioning and the
price moves is extreme. One way to do this is to apply the algebraic
structure (6.2) of the sentiment index to construct a normalized price
(NP) indicator, as follows
Then the two scaled metrics for positioning and price can be combined,
using, for example, the Euclidean distance formula
and the same trading signal (6.3) can be applied where SIt is replaced
with dt.
Any information that can be deduced from the positioning analysis
is highly sought after by oil traders. It is analogous to playing a poker
game with an educated guess about the opponent’s cards. Complete
knowledge of the opponent’s hand is unlikely to be possible, but even a
quick glimpse could lead to a powerful edge. However, given the
incompleteness of the overall picture, which is contaminated by
overlapping categories in reported data, running a systematic trading
strategy purely based on positioning data is unlikely to be sustainable.
Like any other quantamental strategy, this concept works better when
it is combined with other discretionary inputs. In the next chapter, we
extend the idea of quantamental trading to macro factors that connect
oil to the broader world of financial markets.
References
Alexander, C. (2001). Market models. Wiley.
Bouchouev, I., & Zuo, L. (2020, Winter). Oil risk premia under changing regimes. Global
Commodities Applied Research Digest, 5(2), 49–59.
Ederington, L. H., Fernando, C. S., Holland, K. V., Lee, T. K., & Linn, S. C. (2021). The dynamics of
arbitrage. Journal of Financial and Quantitative Analysis, 56(4), 1350–1380.
[Crossref]
Fattouh, B. (2011). An anatomy of the crude oil pricing system. Oxford Institute for Energy
Studies, Working Paper, 40.
Imsirovic, A. (2021). Trading and price discovery for crude oils: Growth and development of
international oil markets. Palgrave Macmillan.
[Crossref]
Kang, W., Rouwenhorst, K. G., & Tang, K. (2020). A tale of two premiums: The role of hedgers
and speculators in commodity futures markets. The Journal of Finance, 75(1), 377–417.
[Crossref]
Footnotes
1 Many financial speculators are required to roll their position at least five days prior to the
expiration of the futures contract to avoid regulatory position limits and risks of physical
delivery. The static short spread strategy is relatively insensitive to the exact rolling schedule,
provided that a short position is entered prior to the scheduled index rolls.
2 The choice of this metric does shorten the lookback period as Cushing capacity data is only
available since 2011. However, the conclusions are substantially similar if the analysis is
repeated for longer lookbacks using private estimates for the storage capacity prior to 2011.
4 Brent futures are based on the so-called BFOET basket, which includes Brent, Forties,
Oseberg, Ekofisk, and Troll, all produced in the North Sea. US Midland oil produced in the
Permian Basin was added to the basket in 2023. For a detailed description of the price setting
mechanism in the Brent market, we refer to Fattouh (2011). See also Imsirovic (2021).
5 The USA continues to import some oil to better match refinery needs that are optimized to
run on heavier crude oil. While the USA imports heavy oil, the light shale oil is exported to less
complex refineries located mostly in Asia and Latin America. With the growth of shale
production, US oil exports started to exceed its imports.
6 In practice, arbitrage traders often shift one leg of the spread by one month to account for the
shipping time. For example, to approximate the economics of oil exports from the USA to
Europe, traders are more likely to use the spread between Brent futures that expire at time
T + 1 and WTI futures that expire at time T.
7 The WTI-Brent convergence strategy and its sensitivity to model parameters are analyzed in
Bouchouev and Zuo (2020).
8 For a good introductory discussion of cointegration and its application to financial markets,
we refer to Alexander (2001).
9 There are, in fact, several competing US Gulf Coast WTI contracts listed by two major
exchanges, CME and ICE, but for simplicity, we do not differentiate between them.
10 Light and heavy oil correspond to the oil density and sweet and sour to its sulfur content.
11 For illustration, we use holdings of two main futures benchmarks, WTI traded on CME and
Brent traded on ICE. Adding options and other less liquid exchange-traded WTI and Brent
futures does not change the conclusions.
12 To simplify the exposition, we do not include positions held in refined products, such as
diesel, gasoil, or RBOB. The notional size of such positions is smaller than it is for crude oil
futures, even though it is comparable and sometimes even larger if measured relative to the size
of the market for refined products.
13 The question whether speculators or hedgers make money by trading commodity futures
has been studied extensively in the academic literature but with largely inconclusive results.
Given the complexity of differentiating between hedgers and speculators in the oil market and
the lack of available data, establishing such a causality using purely statistical tools is extremely
difficult. One interesting attempt to separate the impact of hedgers and speculators on prices by
the investment horizon was made by Kang et al. (2020) for a broader commodity portfolio.
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023
I. Bouchouev, Virtual Barrels, Springer Texts in Business and Economics
https://doi.org/10.1007/978-3-031-36151-7_7
7. Macro Trading
Ilia Bouchouev1
(1) Pentathlon Investments, LLC, Westfield, NJ, USA
As demand for oil grew globally, particularly outside of the USA, the
dominant role in oil-USD dynamics shifted to the denomination, or the
numeraire effect, which is the consequence of oil being priced
predominantly in USD. The denomination effect leads to an inverse
relationship, as stronger USD implies that it takes fewer USD to buy a
barrel of oil.
This inverse relationship can be visualized from the supply and
demand perspective. A stronger dollar translates into higher domestic
prices in countries outside of the USA, which over the last decade were
the primary drivers of oil demand growth. Higher prices have a
negative impact on demand, and, therefore, create downward pressure
on the price of oil. In addition, USD appreciation also causes a negative
impact on price via increasing supply by oil-producing countries with
free floating currencies. Since a portion of production costs is
denominated in local currencies, stronger USD improves profit margins
and incentivizes oil exports. The supply-side channel is more
complicated though due to secondary factors, such as the value of USD-
denominated debt issued by foreign producers.
There are several macroeconomic factors that counter the
predominantly negative oil-USD relationship. One is driven by oil direct
pass-through into inflation and its impact on changes in monetary
policy. As we have already mentioned, rising inflation increases the
likelihood of interest rate hikes, which, in turn, positively impact
demand for short-term US Treasury bonds. Another positive
contribution to the oil-USD relationship comes from the US growth
effect. More favorable economic prospects in the USA relative to the
rest of the world could result in stronger USD given the higher
likelihood of monetary tightening and stronger oil prices, as the USA is
still the world’s largest petroleum consumer. Likewise, both USD and oil
can falter when the USA experiences an idiosyncratic slowdown. This
happened, for example, in the early days of the Covid-19 pandemic in
the USA when both oil demand and USD fell, and for a short period of
time the correlation between the two falling assets was positive.
Subsequently, as the pandemic escalated around the world, USD
regained its stability, while oil continued to fall because of the reduction
in global demand. As a result, the correlation returned to negative.
Numerous academic studies of macroeconomic links between oil
and USD have been conducted, but they only produced conflicting and
inconclusive results. This highlights the time-varying nature of the
relationship and the high sensitivity of econometric analyses to the
sample selection. Many attempts have also been made to establish an
aggregate direction of causality. However, given so many comingled
channels acting at once, disentangling causality in any robust way using
statistical methods is practically impossible. While slow-moving
macroeconomic forces remain more relevant for a long-run
relationship, in the short-run their combined impact is
indistinguishable from non-tradable noise. In fact, prior to the
financialization of oil markets that began in the 2000s, the dynamics of
short-run oil-USD correlation largely resembled a random walk around
zero mean. Financialization has brought drastic changes to this
relationship. Markets started to move much faster, and the dominance
of fundamental forces has been challenged and superseded by fast-
moving financial flows and cross-asset spillovers.
Perhaps the most important driver of the oil-USD relationship in the
modern financialized world of oil trading is the risk aversion or safe
haven channel, which is driven by risk-on, risk-off sentiment shifts.
When adverse economic news triggers a large selloff in the equity
market, many other risky assets, including oil, are also liquidated by
financial portfolio managers tasked to maintain fixed allocation across
all risky assets. During such a risk-off event, investors reallocate capital
towards safer securities, such as US Treasuries which, in turn, increases
demand for USD.
The financial safe haven channel has become the dominant driver of
the predominantly inverse oil-USD relationship. Since approximately
2004, the correlation between returns on oil and a broad USD index has
had a strong negative bias, as illustrated in Fig. 7.3. The correlation was
the strongest, i.e., the most negative, during the years of quantitative
easing following the global financial crisis. To stimulate investment
demand, large bond purchases were made by the US Federal Reserve
Bank, which led to low interest rates and downward pressure on USD,
pushing investors into riskier assets, including oil.
Fig. 7.3 Rolling one-year and sub-sample WTI-USD correlations during three different regimes
Occasionally, the risk aversion channel also contributes to a positive
relationship between oil and USD. This typically occurs either during
periods of high inflation caused by supply-side constraints, or when
geopolitical tensions jeopardize the stability of global oil production.
During such episodes, oil itself, along with USD, is perceived by the
market to be a safe haven instrument, and the dominant negative
relationship is interrupted by shorter spells of positive correlation. For
example, the inflation factor became more visible when the economy
was recovering from the Covid-19 pandemic and higher inflation was
driven by supply disruptions and rising commodity prices.
Finally, the conventional denomination effect also started to play a
larger role in financial oil markets. Many international investors have
oil allocation mandates set in their local currency, such as EUR. If USD
strengthens, i.e., the EURUSD exchange rate weakens, then the notional
value of oil holdings measured in EUR exceeds the investor’s target
allocation. Oil positions must then be reduced to bring the allocation
back to the target, creating downward pressure on the price of oil. The
numeraire effect is even more pronounced for highly leveraged risk
party funds and CTAs that manage allocations based on the asset
volatility. If oil and USD are negatively correlated, then oil volatility
measured in EUR is generally smaller than oil volatility measured in
USD. Consider, for example, oil trading at 90 USD per barrel and a
EURUSD exchange rate at 1.20, which translates to an oil price of 75
EUR per barrel. If the oil price rises 10% to 99 USD per barrel and EUR
vs USD strengthens 10% to 1.32, then the oil price measured in EUR
remains unchanged at 75 EUR per barrel despite its 10% change in
USD.
With so many overlapping driving factors of the oil-USD
relationship, trying to use one of the two variables to predict the other
rarely results in a viable trading strategy. One is better off trading them
as a spread that does not depend on the direction of causality. For the
spread to be a viable candidate for the convergence strategy, some
cointegration criteria must be satisfied. While such criteria usually fail
for the spread between oil and a broad USD index, they are often met
for a subset of the so-called commodity currencies of producing
countries, such as Canada or Norway. In contrast to petrodollar
recycling by producers whose currencies are pegged to USD, many
commodity exporters with free-floating currencies convert excess USD-
denominated oil revenues into domestic currencies, which makes them
highly sensitive to the price of oil. This transmission channel is
sometimes referred to as a wealth effect.
In fact, oil plays such an important role in the economics of these
countries that the price of oil is frequently used by Central Banks as an
input for modeling the fair value of their currencies. Similar models for
commodity currencies have also been developed and widely adopted by
foreign exchange traders. This opened an opportunity for a cross-asset
stat-arb-like strategy to trade certain commodities as the spread
relative to their currencies whenever the two markets dislocate. We
illustrate this with a popular convergence strategy between CADUSD, a
Canadian dollar expressed in USD, and WTI.
Since the CADUSD exchange rate and WTI represent values of two
assets both denominated in USD, trading one as the spread against the
other eliminates a large portion of the idiosyncratic USD-specific noise.
As a result, the co-movement between CADUSD and WTI is more
visible, as illustrated in Fig. 7.4. In other words, CADUSD and WTI are
generally considered to be cointegrated.
Fig. 7.4 Co-movement between CADUSD and WTI futures prices (m3)
There are effectively two primary factors that drive CADUSD. One
factor is the relative monetary policy in Canada versus the USA. This
factor is typically modeled by interest rate differentials, which steer
financial flows towards the direction of higher interest rates. The other
primary factor is the price of oil. When monetary policies in both
countries are on hold, then the contribution of the first factor weakens,
and the price of oil starts playing the dominant role. However, when
monetary policies are changing, the exchange rate becomes more
driven by interest rate differentials and the relationship with oil
weakens. One can clearly see this during uneven recovery after the
Covid-19 pandemic and rapidly rising inflation.
If the monetary policies in both countries are expected to remain
relatively stable, then the technical implementation of the convergence
strategy becomes very simple, as in the energy convergence strategies
discussed in the previous chapter. We define the cross-asset spread as
and then buy and sell the pair when the spread moves outside of a
certain band. The band can be specified, for example, by the spread
deviation from its moving average, as follows:
(7.1)
Fig. 7.7 Co-movement between XOP and WTI futures prices (m3)
Fig. 7.8 A large portion of CPI monthly variance is explained by retail gasoline prices (1993–
2022)
To describe the trading strategy, we use RBOB futures, which are
most closely related to retail gasoline prices, even though a similar
strategy can be constructed using more liquid WTI futures, or a basket
of petroleum futures that may also include diesel futures and Brent.
The trading opportunity exists largely because energy futures and
inflation markets operate at different speeds. While the electronic
market for futures moves very fast, inflation swaps largely trade OTC,
where transactions are still negotiated in the old-fashioned manner
bilaterally between clients and their bank dealers. If the inflation
market were efficient, then one would expect every move in energy
futures to be instantaneously repriced in CPI swaps. In practice,
however, this does not always happen, as the slow-moving OTC inflation
market often exhibits some inertia. This market is dominated by a
different set of market participants, predominantly by fixed income
portfolio managers whose decisions and investment mandates are set
based on longer-term structural views. They largely ignore short-term
fluctuations in energy prices as many of them are not even authorized
to trade RBOB futures. Energy traders, on the other hand, tend to be
nimbler and more flexible with a better eye on short-term arbitrage-
type profits.
When the inflation market lags the move in energy futures, the
trader can take a position against the dislocation between RBOB
futures and inflation swaps and trade the pair as a cross-asset spread.
The strategy bets on the residual which we can define as ex-gasoline
inflation. The dislocation between the two prices may allow the trade to
be entered at sufficiently advantageous levels, so that other
components of the CPI do not have to be estimated with high precision.
The crux of the strategy is in separation and careful hedging of the CPI
gasoline component. The ex-gasoline component is effectively treated
as residual noise. To implement the strategy, however, one needs to
understand the somewhat intricate plumbing of inflation derivatives.
One important feature of the inflation market that often causes a
considerable amount of confusion to outsiders is the so-called base
effect. Another peculiar quirk of inflation derivatives is the lag between
the valuation date and the period over which inflation is measured. We
will briefly explain these features but without getting too deeply into
details. For simplicity, we use a benchmark one-year inflation swap.
The fixed rate of the swap is quoted as an annual expected rate of
inflation. This rate is swapped for the floating rate of inflation i(T),
which is determined by the ratio of the reference CPI index I(T) at the
future time T > t to the reference index I(t0) at past time t0 < t, as
follows:
The reference index is calculated using monthly CPI prints but with a
lag of between two and three months. If the observation day t is the
first day of a calendar month, then the reference index uses a monthly
CPI print for the third preceding calendar month. In other words, in this
case t0 = t − 3/12 and the annual inflation rate will be determined by
the CPI print nine months forward, i.e., at time T = t + 9/12. If the
observation date t is any other day of the calendar month, then the
reference index is determined by a linear interpolation between CPI
prints two and three months prior.2
If another one-year swap is traded tomorrow at time t + dt, then it
will reference a new base index I(t0 + dt) which could be substantially
different from I(t0) if energy futures happened to experience a large
move at time t0 + dt. This makes it rather difficult to compare market
inflation swap rates on a rolling basis and visualize the price dynamics
in the form of a time series. As the base index constantly rolls in time,
the quoted swap level changes daily even if market expectations about
the future inflation index remain the same.
The skill of trading short-term inflation relative to energy futures
hinges on the trader’s ability to separate and hedge out the volatile
gasoline contribution from the residual ex-gasoline inflation
component of the index, which we denote by ix. Let R(T) represent
retail gasoline prices, and ω denote the weight of gasoline expenditures
in the CPI basket. For simplicity, we assume that ω is a known constant,
even though, more precisely, it can also vary with energy prices. Then
the quoted rate of inflation can be decomposed as
(7.2)
(7.4)
If one were able to trade futures directly on retail gasoline prices, then
(7.4) would have applied with p = 100%. In fact, there exists an OTC
market for retail gasoline swaps, but, unfortunately, it is extremely
illiquid. A similar market for retail diesel swaps is more developed, and
some energy traders use the market forward curve for retail diesel
prices as a guide to estimate the forward price of retail gasoline, which
is the most uncertain variable in the CPI-RBOB convergence strategy.
Let us consider a trading position in CPI swap i hedged with n
contracts of RBOB futures
What makes this strategy appealing is that the crucial hedging ratio can
be calculated algebraically rather than estimated statistically.
Let us assume that the futures price experiences an instantaneous
shock dF. Then from (7.3) we obtain that the value of the inflation swap
contract changes by
Since the idea of the strategy is to trade it only when the deviation
becomes particularly significant, one can avoid the need for high
precision in the estimation of ih. If the dislocation between inflation and
futures markets is very large, then ex-gasoline inflation ix,t can even fall
entirely outside of the range of previous historical observations. This
would suggest that the trading opportunity is so good that it would
have made money under all previous paths of realized inflation.
In contrast to inflation specialists, oil traders like to turn the
problem upside down and map all variables instead into energy terms
that they are more familiar with. The primary uncertain variable in this
strategy is the forward price of retail gasoline Rt(T), which at time t is
approximated using the shape of observable futures curves for RBOB
and, if necessary for longer maturities, crude oil futures. Therefore, by
and large, this strategy amounts to trading the spread between retail
gasoline prices implied by the inflation market and RBOB futures.
Instead of expressing market-implied ix,t via pass-through of
observable futures prices as in (7.4) and comparing it to ih, energy
traders assume that the inflation market already embeds some
consensus about historical average ex-gasoline inflation ih. In this case,
one can replace ix with ih in the Eq. (7.2), and instead back out the
market-implied future retail gasoline price from the headline
inflation rate it as follows
The practical challenge with this more advanced approach lies in the
estimation of the pass-through of agricultural futures into the food sub-
component of inflation. Given the larger diversity of products that make
up the food basket, such pass-through estimates for agricultural futures
are less reliable.
Having developed some tradable linkages between oil and other
main financial asset classes, we conclude this chapter by building a
simple model that aggregates the information from various macro
variables and attempts to use this information to estimate the fair value
for oil.
(7.5)
Fig. 7.10 Additional factors that can be included in the oil fair-value model
The factors are grouped into three categories that represent
tradable market prices of different assets, macroeconomic fundamental
factors, and oil-specific factors. Some non-tradable fundamental factors
may not be available on the daily basis, so a broader macro model can
only be implemented at a lower frequency, which does decrease its
performance characteristics. Another challenge comes from the fact
that very few of these factors are cointegrated with oil, making it
difficult to use price-level regressions of the form (7.5). In this case, one
can switch to using regressions on price changes which makes all
variables stationary.
One can take this idea a step further and make the factor-selection
process more dynamic in an attempt to identify the most relevant
factors during each period. This can be done by running individual
single-factor regressions on price differences and ranking factors based
on corresponding R2. One can then estimate the fair value of oil using a
multidimensional regression against several factors that have the
highest R2 in the previous period. However, we would caution against
the mechanical application of such a factor-selection approach given
the high degree of collinearity among them. At the end, a quantamental
trader still has the final say on which factors to use in any given period.
This problem of dynamic factor selection appears to be a good
candidate for using more sophisticated statistical techniques, such as
the methods of machine learning. While the author spent a
considerable amount of time in trying to apply these methods, so far,
the results produced by more advanced statistical models are only
marginally better as compared to a simple static three-factor model.
However, with recent advances in data science, this area of research
continues to evolve. If this book ever makes it to its second edition, the
chapter on application of machine learning methods will most certainly
be expanded. Until then, we can say that our analysis of relatively
simple linear trading strategies is now completed. We can move to a
more advanced, but also a more lucrative universe of nonlinear option
strategies.
Footnotes
1 As of December 2022, the relative importance of the motor fuel category was 3.275%, out of
which 3.172% was gasoline. Source: US Bureau of Labor Statistics (BLS).
3 Since inflation is measured on a calendar monthly basis, future prices must be interpolated
between two contracts with surrounding maturities. To calculate the base gasoline price, one
must also use linear interpolation similar to that described in the prior footnote for the
reference inflation index.
Part III
Volatility Trading
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023
I. Bouchouev, Virtual Barrels, Springer Texts in Business and Economics
https://doi.org/10.1007/978-3-031-36151-7_8
(8.4)
In the GBM case, asset returns are normally distributed, or
alternatively, the logarithm of prices is normally distributed.
Under these assumptions, the percentage return on a fully funded
futures position is expected to grow with a constant rate of return
μG and uncertainty characterized by a constant percentage volatility of
returns σG. For the most part, when a generic reference is made to the
term volatility, it is typically assumed to be σG. It is important to
highlight that unlike the price of the financial asset, the term volatility
can only be understood in reference to a specific model. Here, volatility
is understood as the standard deviation of percentage returns under
the assumption that prices are lognormally distributed. The
corresponding lognormal probability density function is given by (A.
14).
The lognormal probability density is inherently asymmetric. It is
only defined for positive prices, and it has a much larger right-side tail
compared to the symmetric bell-shaped normal curve. The shapes of
two probability densities are illustrated in Fig. 8.1. While one can justify
using an asymmetric distribution in financial markets that tend to grow
over time, applying it to mean-reverting commodity markets brings
some adverse side-effects. As we explained in the chapters on storage
and hedging pressure, in the oil market, the frequency and the
magnitude of upward and downward price movements tend to be more
symmetric, if not even skewed to the downside.
Fig. 8.1 An example of normal and lognormal probability densities
By and large, the commodity industry took the path of least
resistance and followed the lognormal modeling paradigm prevalent in
the equity markets. Without any doubt, this assumption is extremely
convenient when comparing strategies across different commodities
using a standardized percentage-based volatility metric. No risk
manager would want to deal with volatility measured in dollars per
barrel for oil, in dollars per bushel for corn, or in dollars per ounce for
gold.
The convenience of such a one-size-fits-all measure comes at a
price. It forces traders to look at the oil market through an incorrect
frame of reference imposed by inflexible and highly standardized
operational setups of their trading systems. In contrast, dedicated oil
specialists customize their metrics, which allows them to capture
important salient features of the market and to take advantage of the
framing bias often suffered by generalists. This modeling duality adds
certain pedagogical challenges, as many market phenomena must be
presented in two alternative ways, first as they are seen by the general
crowd through standard but poor-quality lenses, and then, as perceived
by oil professionals equipped with more accurate viewing tools.
Both ABM and GBM assumptions are rather simplistic and chosen
primarily for their analytical tractability. It would be naïve to think that
oil price behavior, which is influenced by a multitude of driving factors,
can be fully characterized by only two parameters, the constant drift,
and the constant volatility. The real world of oil prices is much more
complex. Fortunately, as we will see in the following section, the drift
term in the stochastic equation plays no role in modeling options, and
our entire focus will be on volatility. However, volatility of oil futures is
anything but constant. It varies not only with time, but also with the
level of futures prices. For example, as futures move outside of some
normal price range, absolute volatility tends to rise. This indicates
much higher probabilities of extreme events that are nearly impossible
under either normal or lognormal distributions. In other words, the
distribution of oil prices has fat tails.
To allow ourselves more flexibility in modeling notorious tails of the
price distribution, we assume that the drift and volatility parameters of
the stochastic process are deterministic, but not yet specified, functions
of time and the futures price. Futures then follow a more general
stochastic process described by the following equation:
(8.5)
Such stochastic processes, which arise in many natural sciences and
applications, are called diffusions. In finance, the function σ(F, t) that
characterizes the volatility of the diffusion process is known as the local
volatility function. ABM and GBM processes represent only two special
cases of the more general class of diffusions. The local volatility of the
ABM process is constant σA. For the GBM process, the local volatility
σGF is proportional to the futures prices. All price changes for diffusions
are still driven by the same normally distributed single source of noise
dz.
The dependency of the volatility function on the futures price,
which itself is stochastic, allows diffusions to produce a much wider
universe of probability distributions for future price changes. One
simple way to generate a probability distribution by a diffusion process
is to use Monte Carlo simulations, where multiple sequences of random
variables dz(t) are drawn from the normal distribution. The
corresponding futures prices are then computed using the discretized
version of (8.5) and presented in the form of a histogram.
The science of diffusions largely hinges on one important result,
known as Itô’s lemma. It describes how a function of a stochastic
variable changes for a given change in the variable itself. In the
ordinary calculus, a small change in the value of the function G(F, t) is
approximated by its partial derivatives multiplied by a change in the
function’s arguments, the result known as Taylor’s formula. When one
of the variables F is stochastic, the analogous approximation is given by
the following Itô’s lemma:
(8.6)
(8.7)
Note that both the futures price and the option price are driven by
the same source of uncertainty dz. This allows us to combine the option
and the future in a particular way that eliminates the source of
randomness.
Assume that we bought a call option and want to offset some risks
by selling a yet to be determined quantity of futures, denoted by the
Greek letter delta, Δ. We now have a portfolio Π that consists of a long
call option and a short position in Δ units of futures:
Note that in the last term only the option premium C accrues
interest. The futures contract does not require any initial investment,
besides a small collateral, the impact of which for simplicity is ignored
here. Since it does not cost anything to enter into the futures contract,
there is no interest accrued on the futures position either.
Cancelling the dt term, we obtain the generalized BSM equation for
the option price written on futures, which follows a diffusion process
with local volatility σ(F, t):
(8.8)
The solution to the Eq. (8.8) with the boundary condition (8.1) is
then given by the following Bachelier formula:
(8.9)
Here, standard notations are used for the normal Gaussian
probability density with zero mean and variance equal to one:
The quantity
The solution to the Eq. (8.8) with the boundary condition (8.1) is
then given by the Black formula
(8.10)
The Black formula can also be derived by integrating the payoff with
the lognormal probability density or, alternatively, it can be verified by
the direct substitution of its partial derivatives into (8.8). The details
are provided in Appendix B.
The normalized log-moneyness term mG in (8.10) is defined as the
logarithmic ratio of the futures price to the strike price scaled by the
geometric volatility over the life of the option:
So far, we have only dealt with the valuation of a call option.
However, the Eq. (8.8) applies to any derivative of the futures price. The
fact that it is a call option was only specified by its boundary condition
at maturity. If the boundary condition (8.1) is replaced with (8.2), then
similar formulas can be obtained for a put option. However, an easier
way to derive pricing formulas for puts would be to utilize an important
no-arbitrage relationship that links prices of call and put options.
If we construct a portfolio which is long a call option and short a put
option with the same strike K, then the payoff of this portfolio at
expiration time T can be written as
(8.11)
In other words, holding this portfolio is identical to holding a long
futures position established at the price K. Since the portfolio of a long
call and a short put is itself a financial derivative of the futures, then the
solution to (8.8) with the boundary condition (8.11) is simply the
discounted futures payoff, reflecting the fact that option premia are
typically paid upfront at the initiation of the trade:
(8.12)
This formula, known as a put-call parity, allows one to determine
the price of a European put option given the price of a European call
and the price of futures. Alternatively, one can determine the price of a
call given the price of a put along with futures.
We have chosen to repeat the standard BSM replication algorithm
not because of its mathematical elegance, but rather because it
provides an explicit recipe for trading volatility. This book is not meant
to be a comprehensive reference of derivatives pricing models. Our goal
is to highlight certain features of these models that can be turned into
profitable strategies, specifically in the oil market. Not all models, of
course, present such unique trading opportunities. As such, to gain
more clarity in already complex topics, we make some simplifying
assumptions in areas where we do not see any particularly unique
trading opportunities.
For the most part, we ignore the impact of the interest rate,
assuming it to be zero. Unlike stocks and bonds, the interest rate plays
only a relatively minor role in the valuation of options on oil futures.
For European options that can only be exercised at expiration, the
impact of the interest rate amounts merely to the discounting factor
that appears as the multiplier in pricing formulas. The discounting
comes from the fact that an option premium is assumed to be paid
upfront, but the settlement of the option occurs later, at the expiration.
In the oil market, the premium is paid upfront only for WTI options,
while for exchange-traded Brent options, the premium is netted against
the final settlement. Therefore, Brent options are marked-to-market
like futures, in which case the discounting factor should be removed. It
is interesting that this was also the case for options considered by
Bachelier, which is why the interest rate was also omitted in his study.
Regular exchange-traded oil options are American options that can
be exercised at any time prior to the expiration. Thus, an American
option must be slightly more expensive than an equivalent European
option. The precise calculation of an early exercise premium is rather
complicated as the partial differential equation for the price of an
option turns into an inequality, where one needs to solve a complex free
boundary value problem or to apply other optimization techniques.8 To
simplify this, several analytical approximations have been developed,
and the one developed by Barone-Adesi and Whaley has proven to be
adequate for the oil market.9 These models show that it is optimal to
exercise an American option early only if the benefits from receiving a
non-discounted cash settlement sooner outweigh an additional value
from holding an option longer. The latter is driven by the remaining
volatility. Since the impact of oil volatility on the option price is much
larger than the contribution of the interest rate, it is rarely optimal to
exercise an American option early, except for deep ITM options at times
of high interest rates. Furthermore, under our simplified assumption of
zero interest rates, the early exercise premium is equal to zero.
While the interest rate does matter for oil trading, using it
generically in pricing formulas can cause more harm than it adds value.
The interest rate for all traders is nearly always tied to the cost of
funding specific to the trading company and to customized collateral
agreements with clearing brokers and counterparties. Some strategies
can indeed be enhanced by optimizing financing costs, but they are
ultimately inseparable from the management of the trader’s balance
sheet. For example, when a long-dated ITM option has a significant
intrinsic value that cannot be withdrawn, which is sometimes called the
trapped option value, an auxiliary agreement is often structured to use
it as the collateral for other trades. The true value of the early exercise
premium becomes heavily dependent on the details of such agreements
and remains highly tailored to the needs of individual traders. In other
words, this area is much closer to the field of structured finance than to
the topic of oil trading, and for this reason we do not consider it in the
book, mostly assuming zero interest rate, unless noted otherwise.
Finally, for some theoretical arguments we generally assume that
the maturity of the option coincides with the maturity of the underlying
futures. In practice, however, standard exchange-traded oil options
expire three business days prior to the expiration of the corresponding
futures. To simplify the exposition, in this chapter we ignore the
relatively minor impact of the three-day maturity mismatch between
options and futures and assume that both expire at time T. In Chap. 11,
we will cover a more general case of options whose expiration is
decoupled from the expiration of the underlying futures.
In the remainder of this chapter, we discuss two other types of
volatility, realized and implied, both of which can only be understood in
the context of particular simple modeling assumptions. Our initial goal
is to highlight potential challenges and pitfalls with blind application of
such simplified frameworks to the oil market. The actual recipes for
trading that feature models specifically tailored to opportunities in the
oil markets will be provided in the following five chapters of the book.
Fig. 8.4 Realized 3-month volatility of percentage returns versus third nearby futures (WTI,
2000–2022)
The problem with traditional percentage-based measures of
volatility starts with the definition of an investment return, which is
more ambiguous for trading futures. Since the only investment required
to enter into a futures contract is a relatively small initial margin posted
with the clearing broker, in theory, this posted margin should be used
as an initial investment in the definition of return. In practice, margin
requirements fluctuate along with prevailing risks and using the margin
for the standardized denominator of a return would be rather
cumbersome. Many financial analysts simply ignore this specificity of
futures trading, and for convenience, use the notional value of futures
in the definition of returns and subsequent calculation of volatility.
While such percentage-based metrics make sense for financial assets
that tend to grow over time, applying them to mean-reverting
commodity prices creates an artificial framing bias that could adversely
impact decision making. When oil prices are low, the division by a small
number blows up the magnitude of percentage returns, and even makes
calculation mathematically impossible if the asset price becomes
negative.
Since the profitability of highly leveraged futures trading is
measured by professional oil traders nearly exclusively in dollar terms,
corresponding risks are better measured accordingly. What oil traders
care about is the volatility of profits and losses driven by price changes,
not the volatility of somewhat artificially constructed investment
returns. If realized volatility is calculated instead as the standard
deviation of price changes instead of percentage returns, then this
results in a more accurate measure of uncertainty. We refer to such a
measure as the realized dollar volatility.
As shown in Fig. 8.5, the artificial negative skewness of the volatility
with respect to the price level disappears. The relationship between
volatility of price changes and futures prices becomes more intuitive.
The volatility tends to be relatively low when oil prices remain in some
normal range, but it increases sharply when prices move away from
such a range in either direction. In other words, the relationship
between the realized dollar volatility and prices is somewhat parabolic,
the insight that we will use later in developing an appropriate model for
pricing oil options.
Fig. 8.5 Realized 3-month volatility of price changes measured in dollars per barrel versus
third nearby futures (WTI, 2000–2022)
Figures 8.4 and 8.5 highlight the importance of choosing the correct
lenses through which to look at oil volatility. One measure is rather
poor, but it is often chosen for its operational convenience and for its
easy standardization across asset classes. Another one is a more
customized metric tailored to the specifics of the oil market.
Fortunately, for many practical purposes it is rather straightforward to
go back and forth between the two metrics. Since an investment return
is defined as the ratio of a price change to the initial futures price, the
volatility of returns can often be approximated by the ratio of the
volatility of price changes to the price level.
This distinction between alternative ways of measuring volatility
turns out to be vital for trading oil options. Switching the calculation of
the historical volatility from percentage returns to price changes does
not, of course, resolve any conceptual problems with the realized
volatility concept, but, at least, it removes an artificial skewness of risk
created by volatility of percentage returns. This skewness becomes
even more visible and problematic when statistical estimates of the
volatility of stochastic process are replaced with an alternative method
of estimating volatility by deducing it from the options market.
8.5 Implied Volatility and its Skew
Any model of financial markets is only a theoretical construct designed
to approximate the real behavior of prices. A typical financial model
transforms a certain set of inputs by means of some quantitative
machinery into a description of market prices. Regardless of how good
the machinery is, the output of the model can only be as good as the
quality of its inputs. The primary input to the models for option prices
is the volatility of the stochastic process that governs the dynamics of
the underlying futures contract. As discussed in the previous section,
estimating this input from history leaves a lot to be desired, as
historical realized volatility is all over the place. One can, of course,
build more complex statistical models to forecast volatility, but
regardless of the econometric technique, the resulting option price
based on the volatility estimate is likely to be different from the one
observed in the market.
When a choice is to be made between the model and the market,
practitioners tend to assign higher powers to the market. To reconcile
the model with the market, traders came up with a clever workaround
and turned the option pricing problem upside down. Instead of betting
on the model driven by a noisy input, they use the model in reverse and
back out the missing volatility input from the market price of an option.
In other words, the pricing formula is matched to the observable option
price and then inverted for volatility. The resulting measure of volatility
is known as the market-implied volatility.
The general pricing formulas (8.9) and (8.10) cannot be analytically
inverted for volatility, and implied volatility must be calculated
numerically using a root-finding algorithm. Since the option’s vega,
which is the derivative of an option price with respect to its volatility
input, is strictly positive, and an option price is a monotonically
increasing function of volatility, such numerical inversion is always
possible. Therefore, every option price is uniquely mapped to its
corresponding implied volatility, and one can view implied volatilities
as an alternative metric for option prices.
The market-implied volatility presents an alternative method of
estimating some properties of unobservable local volatility of the
stochastic process. In contrast to a backward-looking realized volatility
estimate, implied volatility is a forward-looking measure. It is simply a
plug to a particular option pricing formula that forces the model to
match the market. Only in a very special and somewhat unrealistic case,
when the futures market happens to behave exactly as prescribed by
the model, can implied and realized volatilities be meaningfully
compared. The implied volatility can then be interpreted as the market
expectation of the future realized volatility. But even in this case, the
two volatilities can still be very different if the option price contains an
additional risk premium caused by hedging imbalances. One can also
think of the implied volatility as the average of the expected local
volatility, adjusted by the volatility risk premium. We will discuss these
topics in more detail in subsequent chapters.
To distinguish the market-implied volatility from other types of
volatility, we denote it by a different letter. For an option with the strike
price K and expiration T, we use v(K, T) to represent the market
standard implied Black volatility (IBV), which is computed by inverting
the Black pricing formula (8.10). We will also use the less common, but
arguably more important measure of implied normal volatility (INV) or,
alternatively, implied dollar volatility, which we denote by vN(K, T). The
INV is backed out from the same market price of an option by inverting
the Bachelier formula (8.9). Regrettably, the market rarely explicitly
attaches Bachelier’s name to the normal volatility, and we reluctantly
follow the market lingo and adopt the term normal volatility. In
contrast to IBV, which is measured in percent, INV is expressed in
dollars per barrel.
The irony of the market standard IBV metric is that it debunks the
main assumption of the model that is responsible for its own creation.
For a given maturity, there are many options with different strikes, but
there is only one futures contract. The lognormal assumption of
constant proportional volatility is a property of the futures contract,
and it has nothing to do with options. If the Black model were correct,
then the inversion of any option price should result in the same
number, the same constant volatility of the futures that the model is
based on. In practice, implied Black volatilities computed for options
with different strikes and maturities are nearly always different.
A typical dependency of implied Black volatilities v(K) on the strike
price is shown in Fig. 8.6 for short-term, medium-term, and long-term
oil options. The graph is called the volatility smile. The reference to a
smile comes from its origins in the foreign exchange market, where the
plot is often more symmetric with similar curvature on both ends,
making it look like a smile. In the oil market, this graph is also known as
the volatility skew to highlight its predominantly negative slope, except
for the utmost right tail for shorter-term options, which makes the
graph look more like a smirk. The curvature of the smile on both ends
indicates that the options market expects a higher frequency and
magnitude of extreme events than can be generated by the lognormal
distribution. The market adjusts the pricing model by using higher
volatility input to capture such events. We will discuss alternative
normal volatility smiles and develop a more accurate model that
captures the fat tails of the price distribution in Chap. 10.
Fig. 8.6 Representative implied volatility skews for short-term, medium-term, and long-term
oil options
It is important to emphasize that the implied volatility smile is
nothing more than a collection of option prices presented in a more
convenient coordinate system. Simultaneous tracking of prices for all
options with different strikes and maturities in dollars per barrel would
be next to impossible, as option prices are constantly changing along
with futures based on their corresponding deltas. In contrast to many
statistical forecasting models, the main purpose of the Black model is to
translate market observable option prices to a different and more
convenient scale. Such a transformation does not require any statistical
methods.
Even though the implied volatility smile has a negative skew with
puts generally having higher implied volatilities than calls, this does not
mean that puts are overpriced relative to calls. It is simply another
artefact of looking at the market through incorrect lenses. The
skewness of the IBV curve versus strikes is a side-effect of the same
wrong unit of measure, previously illustrated by Fig. 8.4, where the
realized volatility was plotted versus futures. If the true oil price
distribution happened to be more symmetric, then observing a
phenomenon through skewed lognormal lenses would force us to
adjust our perception of option prices by raising percentage volatility
for lower strikes and lowering it for higher strikes. The lognormal
skewness is akin to prescribing lenses for astigmatism to someone who
does not need them, which will result in blurry vision. In Chap. 10, we
will construct more appropriate lenses to view the oil options market,
but for now, we proceed down the conventional route of expressing
option prices in terms of their IBVs.
Choosing the right metric is critical in trading as the frame of
reference has a strong impact on traders’ decision making. The
volatility smile, shown in Fig. 8.6 as a function of strike K, works well as
a static snapshot of option prices. However, this smile is more difficult
to track dynamically once futures move, as the range of actively traded
strikes also shifts. The most liquid option is typically ATM and the
strikes for many other options are often chosen by traders relative to
ATM, which makes it more practical to maintain the smile as a function
of option moneyness instead of fixed strikes. The simplest way to define
moneyness would be to use either the difference between the strike
price and futures, K − F, measured in dollars per barrel, or their ratio,
K/F, in percent. Such choices are intuitive and indeed both are often
utilized for short-term analysis and for options with the same
expiration. However, simple moneyness metrics become more
problematic when comparing implied volatilities across multiple time
horizons.
In both the Bachelier and Black formulas, the volatility is scaled
with the square root of time to maturity, . Therefore, when we back
out implied volatility from option prices, we always divide by . This
magnifies the resulting smile for shorter-term options and flattens it for
longer-term options, as clearly seen in Fig. 8.6. However, what matters
for option pricing is volatility-normalized moneyness, such as mA for
the normal distribution, and mG for the lognormal one. If we use such a
normalized moneyness as the primary unit of measure, then volatility
smiles will be more accurately compared across all maturities.
We could have stopped here and accepted a normalized moneyness
setup as the base case for tracking the smile, but option traders often
take it one step further and define moneyness instead directly in terms
of their hedging deltas. The deltas for Bachelier and Black models,
which are given in Appendix B, effectively transform volatility-
normalized moneyness to make it vary between minus one and plus
one. Keeping volatility as a function of delta has proven to be handy for
traders as the required hedging ratio comes directly as a biproduct of
the smile setup. For example, to hedge the sale of 100 units of 25-delta
calls, one would need to buy 25 futures. Black deltas are often used as
the market primary communication tool, even though traders typically
make further adjustments to the actual delta hedging, which we discuss
in more detail in Chap. 10.
To illustrate historically observed oil smiles, Fig. 8.7 shows the
average of IBV smiles for 10-, 20-, 30-, and 40-delta OTM puts and calls
and zero-delta straddle since 2000. Note that because of an asymmetry
embedded in the lognormal distribution, the Black delta for an ATM
straddle is counterintuitively not equal to zero. To preserve the uniform
spacing in the graph of the smile, we use a zero-delta straddle instead
of an ATM straddle.
Fig. 8.7 Average IBV versus delta-moneyness for different maturities (WTI, 2000–2022)
Average IBV smiles are shown for different days to maturity (DTM)
that correspond to one-, three-, six-, and twelve-month options. Implied
volatilities are always calculated using more liquid OTM options. Less
liquid ITM options trade infrequently, and their prices are usually
synthetically derived from corresponding OTM options and the put-call
parity relationship (8.12). We should also acknowledge that
maintaining volatilities as a function of delta has its own challenges, as
such a setup is somewhat circular. While volatility is commonly
measured versus delta, the delta itself depends on volatility. In practice,
implied volatilities are first calculated for fixed strikes along with
corresponding deltas, and then subsequently interpolated for the
desired grid of deltas.
It is also clear from Fig. 8.7 that for each moneyness, implied
volatility declines with increasing time to maturity. However, unlike the
contradictory presence of the skew for any given futures, the
decreasing term structure pattern of implied volatilities does not cause
any alarms, as it reflects volatilities of different futures which do not
have to be the same. In fact, the volatility should naturally decline for
longer maturities futures as short-term fundamental uncertainty fades
away while being smoothed out by the balancing role of storage. This
phenomenon is known as the Samuelson effect.11 The Samuelson effect,
which was originally observed for the realized volatility, applies to the
implied volatility as well. Under the normal market conditions, the term
structure of implied volatilities is expected to be a decreasing function
of time to maturity, as illustrated by the middle line in Fig. 8.8.
Fig. 8.8 While oil volatility generally declines with time to maturity, the short-term implied
volatility is more affected by realized volatility
If short-term realized volatility is particularly high, which, for
example, could be driven by falling oil prices, then the slope of the
implied volatility term structure steepens, as long-maturity futures
move less. This effect is exacerbated by a percentage-based volatility
metric, as weaker prices generally result in a contango market when
short-term futures prices fall below long-term futures. To compensate,
the percentage volatility must be raised for contracts with lower prices.
When volatility is measured instead in dollar terms, the volatility term
structure is also typically decreasing but with a flatter slope. During
periods of low realized volatility, the implied volatility term structure
may also exhibit a visible hump. The hump may reflect not only market
expectations of eventual increase in volatility, but also the risk premium
embedded in oil options, which we will study in the following chapter.
In Chap. 11, we will also see how critical the shape of the volatility term
structure is for pricing many exotic options.
To summarize, so far, we have only adapted some standard
methodologies for pricing options in the oil market. We chose to model
futures prices in the setting of general diffusions characterized by the
local volatility function. This framework has proven to be the sweet
spot in the oil market, providing a reasonable trade-off between model
accuracy and complexity. We introduced and discussed two essentially
defective, but, nevertheless, commonly used estimates of the local
volatility of the diffusion process. The realized volatility estimates it by
taking a look at the history, while the market-implied one attempts to
extract it from an unknowable future. Both volatilities can only be
understood in the context of simplified and largely unrealistic
assumptions about the distribution of futures prices.
The next step often taken by many amateur traders and occasionally
by some textbook writers is a comparison of one defective metric to
another, as an attempt to determine whether an option is cheap or
expensive. Our preference is to stay away from making such a naïve
comparison, as in the volatile oil markets, this simplified approach can
potentially bring more harm than value. While the difference between
implied and realized volatilities could indeed provide some information
about the richness of the option, the question of the fairness of the
option price cannot be properly answered without a careful
examination of the option’s gamma, the measure of its convexity.
References
Alexander, C. (2008). Market risk analysis, Vol. III: Pricing, hedging and trading financial
instruments. Wiley.
Barone-Adesi, G., & Whaley, R. E. (1987). Efficient analytic approximation of American option
values. Journal of Finance, 42(2), 301–320.
[Crossref]
Black, F. (1976). The pricing of commodity contracts. Journal of Financial Economics, 3(1/2),
167–179.
[Crossref]
Black, F., & Scholes, M. (1973). The pricing of options and corporate liabilities. The Journal of
Political Economy, 81(3), 637–654.
[MathSciNet][Crossref][zbMATH]
Hull, J. C. (2018). Options, futures, and other derivatives (10th ed.). Pearson.
[zbMATH]
Merton, R. C. (1973). Theory of rational option pricing. The Bell Journal of Economics and
Management Science, 4(1), 141–183.
[MathSciNet][Crossref][zbMATH]
Samuelson, P. A. (1965). Proof that properly anticipated prices fluctuate randomly. Industrial
Management Review, 6(2), 41–49.
Wilmott, P., Dewynne, J., & Howison, S. (1993). Option pricing: Mathematical models and
computation. Oxford Financial Press.
[zbMATH]
Footnotes
1 Bachelier (1900).
3 Some energy commodities, such as natural gas or power prices, are more susceptible to large
short-term spikes where diffusions are often combined with jump processes, but this modeling
paradigm is more complex as options can no longer be dynamically replicated with futures. We
discuss some limitations of diffusions in Chap. 12.
4 Black (1976).
5 For the type of options studied by Bachelier, the discounting was not necessary as the option
premium was netted against settlement at expiry. As a result, in his original derivation the
interest rate was ignored. We will explain shortly why a similar assumption can be made for
pricing many oil options. The discounting factor brings in an additional time dependency
resulting from the present value of money, rather than from the evolution of the variance.
6 Greeks are discussed in many standard derivatives textbooks, such as Alexander (2008) and
Hull (2018). For their practical interpretation, see also Leoni (2014).
7 Vega is not a letter of a Greek alphabet but it was adopted by option traders for its phonetic
similarity.
10 For example, if daily prices are used then realized volatility is the standard deviation of
returns multiplied by given approximately 250 trading days in a year, and for weekly
prices, it is multiplied by .
Fig. 9.1 Frequency distribution of P&L ($/bbl) for the strategy of selling unhedged 3-month
ATM WTI straddles (2000–2022)
Table 9.1 The summary statistics of the P&L distribution for two strategies of selling
unhedged and daily delta-hedged 3-month ATM WTI straddles (2000–2022)
(9.1)
Now we need to make the choice of what delta to use to hedge the
option. Even though we believe that the implied Black volatility v may
not be accurate, we also notice that the only way for us to eliminate the
futures risk dF in (9.1) would be to hedge the option with deltas
calculated using the same implied Black volatility. Thus, we choose our
delta as
The fact that the only delta that instantaneously removes the
futures risk dF happens to be the one that is calculated using the wrong
volatility may sound a bit odd. This subtle but important point deserves
a short diversion. If somehow, we knew in advance what the future
actual volatility σ(F, t) would be and hedged the option continuously
using the corresponding deltas, then the option payoff could be
perfectly replicated. In such a purely hypothetical case, we would not
even care what the market thinks about daily volatility and what our
P&L is on a daily basis, because the terminal P&L would be known with
certainty. The terminal P&L is the difference between the sale price P(v)
and the cost of the option’s replication given by the diffusion (8.8) with
the local volatility σ(F, t). However, even though hedging based on
perfect foresight ensures certainty of the terminal P&L, on a daily basis
P&L will fluctuate along with dF.2 This is because the market does not
know anything about our unique and exclusive foresight of the future
volatility, and it is not pricing the option correctly. We will illustrate this
argument shortly with specific strategy examples that use different
hedging deltas. For now, we take the easiest conventional route and
hedge the option using deltas computed with the implied volatility v.
With the contribution of dF instantaneously removed by the delta
hedge, we now have that
We then substitute this expression into the formula for dΠ, which
leads to the following P&L of the portfolio over period dt:
(9.2)
We should also note that call and put options with the same strike
have the same gamma. This can be easily seen by differentiating the
put-call parity relationship (8.12) twice with respect to the futures
price. For volatility dealers, puts and calls with the same strike carry
essentially the same risk.
We are now in a good position to see why option sellers tremble
when they hear the term gamma. If the actual price moves on a given
day happens to be smaller than the one implied by the option market,
then the seller makes some money, but the gain is limited to the daily
option’s time decay, i.e., its theta. If the market moves exactly one
standard deviation, as measured by the option’s implied volatility,
which is often referred to as the daily breakeven point, then the daily
P&L is zero. However, if futures move N standard deviations, then the
option seller’s losses are proportional to N2 − 1 multiplied by the
option’s gamma. Things could quickly become troublesome for the
option seller if a large market move occurs at a time when the options
gamma is particularly large.
The option’s gamma is very dynamic, as it varies significantly with
the moneyness of the option and time remaining to maturity. To
illustrate, Fig. 9.3 shows Black gamma for one-, three-, and twelve-
month put options struck at $50/bbl.
Fig. 9.3 Black gamma for the $50/bbl put option with v = 0.30 and different times to
expiration
Algebraically, Black gamma is simply the lognormal probability
density function with the mean at the strike price. It is highest for an
option near ATM, and the peak gamma rises as time to maturity
shrinks. In the limiting case, the gamma of an ATM option right before
its expiration approaches infinity. Imagine now the scenario of a large
futures price move that unexpectedly crosses the strike price of the
option at expiration. The losses for the option seller determined by the
difference between realized and implied volatility are magnified by a
large gamma. This is where the big money in option trading is made or
lost, and why short-term volatility trading strategies are deservingly
named after this powerful Greek letter.
In addition to suffering losses from a large price move, an initially
delta-hedged option portfolio quickly loses its directional immunity,
and the residual delta must be promptly rebalanced. Delta hedging can
only keep the portfolio neutral with respect to futures instantaneously.
As soon as futures move, the desired risk neutrality disappears, as delta
itself changes according to its gamma profile. Gamma is the second
derivative of the option’s price, or alternatively, it is the first derivative
of the option’s delta. If the futures price falls and moves towards the
strike of a short OTM put option, then the put delta becomes more
negative, making the portfolio directionally longer futures. Additional
futures must be sold to bring net delta back to zero. Likewise, if the
futures price rises and moves away from the strike, then the put delta
becomes less negative, making the portfolio overall short. Some short
futures hedges which are no longer needed must then be bought to
bring net delta back to zero.
The short gamma trader must always sell futures on the way down
and buy them back on the way up to keep the portfolio delta neutral.
The quantity of futures needed for rebalancing is determined by the
combined impact of the option’s gamma and the magnitude of the
futures price move. Figure 9.4 shows the number of futures contracts
needed for rebalancing the delta of 100 contracts of short put option
struck at K = 50 with different times to maturity. The options have
gamma profiles, as in Fig. 9.3, but taken with the negative sign to
represent the short position. The initial futures price is F = 60 when the
hedged portfolios are delta neutral.
Fig. 9.4 Incremental futures required for rebalancing 100 short put options with K = 50,
v = 0.30 and different times to expiration. Initially, F = 60 and the portfolio is instantaneously
delta neutral
The gamma here is observed as the change in the slope of the delta
function. If the market moves down, then gamma increases as futures
move towards the strike. In this case, dealers have no choice but to sell
more futures to remain delta neutral. Larger gamma leads to more
aggressive selling of futures by volatility dealers as the futures price
drops. On the other hand, if futures move up, drifting further away from
the short strike, then the need for rebalancing becomes more muted.
Such an asymmetric gamma profile is very typical for the oil options
market, which is dominated by producer hedging demand for downside
protection. On the upside, the overall industry gamma profile is more
balanced, as some producers gain leverage by selling options to the
market, a topic that we discuss in more detail in the next chapter.
The rebalancing of short gamma portfolios is one of the most
powerful forces that drives the market for the underlying futures. This
process is completely mechanical. It has little to do with volatility
dealers’ own opinion about the direction of oil prices. When futures fall,
dealers who are short puts to producers must sell futures to remain
within tight volumetric risk limits on residual deltas, prudently
imposed by their risk managers. The larger the gamma, the more
futures they need to sell, pushing the price further down. This creates a
vicious cycle when prices fall in a downward spiral until strikes are
crossed and dealers’ short gamma exposure starts subsiding. The
gamma hedging profile can also be viewed as an alternative to the
reaction function previously introduced for the momentum strategy.
The insurance-like business of selling delta-hedged options, which
mandates selling futures on the lows and buying them back on the
highs with potentially unlimited losses, may appear to be a
questionable-value proposition for investors. This perception is driven
by cognitive biases that cause painful financial losses to be
overweighted relative to their frequency and magnitude. In the
meantime, every day when nothing major happens in the market, and
the futures price moves less than its daily breakeven, the option seller
retains a small portion of the option premium. The positive P&L on
uneventful days could easily add up to a substantial buffer which more
than offsets periodic large losses.
Is this strategy just an example of picking up pennies in front of the
steamroller, or can it be turned into a viable investment with its risk
kept under control? The answer is not so straightforward. It turns out
that one needs to be more selective on which options to sell, and which
ones are better to stay away from. We address this important topic in
the following sections with a comprehensive empirical study of the
volatility risk premium (VRP) strategy in the oil market.3
Fig. 9.6 VRP term structure, as the premium retained ratio versus maturity with daily hedging
for 0.20-delta puts (20P), zero-delta straddles (0S), and 0.20-delta calls (20C) (WTI, 2000–
2022)
While the historical returns of VRP strategies appear to be
attractive, returns only represent one side of the investment analysis.
The other side, which is arguably even more important for short option
strategies, is the amount of risk that must be taken to achieve such
returns. Here, we need to make an important distinction between the
risk characteristics of individual trades, which we define as positions,
and the risks of the portfolio made up of such positions. Since standard
oil options are only available with monthly maturities, this adds some
complexity to the analysis of risk-adjusted returns because of the
overlapping nature of multiple positions within each portfolio that
remain open at the same time. While returns generated by portfolios
with different maturities can be easily annualized and compared to
each other, the annualization of risks resulting from individual trades
with different maturities is prone to statistical anomalies, especially
when trades are highly correlated.4
To properly measure the relative performance of VRP strategies
across different time horizons, we construct continuously run
systematic strategies made up of corresponding VRP positions with the
same moneyness and the same DTM at the time of the option sale. Each
one-month portfolio (20 DTM) consists of only one position, as the new
trade is initiated within a few days of the expiration of the previous one.
However, it is not the same for portfolios with longer DTM. The two-
month (40 DTM) portfolio has two open trades with expirations
approximately twenty days apart, the three-month (60 DTM) portfolio
has three open trades, etc., and the twelve-month (240 DTM) portfolio
consists of twelve open trades. Only one option in a portfolio expires
each month. Therefore, all portfolios are exposed to the same futures
risk at the expiration of this option. However, since a new option is
added to the portfolio every month at the moneyness level prevalent at
the time, the strikes of options within the portfolio are spread out,
providing some valuable risk diversification benefits to the portfolio of
options.
For a proper apples-to-apples comparison of VRP investments
across time horizons, we construct continuous equity lines by
cumulating daily P&L from all open positions within the portfolio. We
can then calculate traditional investment characteristics for each
portfolio, such as the Sharpe ratio.5 Figures 9.7 and 9.8 display Sharpe
ratios for various VRP portfolios by moneyness and maturity.
Fig. 9.7 Sharpe ratios of VRP portfolios by moneyness (WTI, 2000–2022)
Fig. 9.8 Sharpe ratios of VRP portfolios by maturity for 0.20-delta puts (20P), zero-delta
straddles (0S), and 0.20-delta calls (20C) (WTI, 2000–2022)
The main difference between the portfolio VRP curves, shown in
Figs. 9.7 and 9.8, and position VRP curves, shown in Figs. 9.5 and 9.6, is
the materially improved performance of zero-delta straddles. ATM
options, being highly exposed to large gamma risks, benefit the most
from the strike diversification. The benefits of diversification for OTM
options are more muted. The reason again lies in the crucial role of the
option’s gamma, which becomes infinite if the futures cross the strike
price at the option’s expiration. Since strikes near ATM are more likely
to be crossed than strikes further OTM, the benefits from diversification
of the strike risk are more substantial for ATM options.
Similarly to other systematic risk premia strategies, oil VRP clearly
demonstrates the existence of an investment edge, but its actual
implementation is highly customized by professional volatility traders.
In the next section, we provide some examples of such practical
implementations. These implementations explicitly incorporate the
impact of transaction costs and optimize VRP by combining it with
short-term biases in the behavior of the underlying futures.
Fig. 9.9 The performance of VRP strategies for 60 DTM ATM straddle for different hedging
frequencies (WTI, 2000–2022)
It is not surprising that with less frequent hedging, the risks
measured by the annualized standard deviation of the portfolio’s P&L
steadily increase. More interesting is the observation that annualized
profits also increase slightly if the portfolio does not rebalance the delta
for at least two days. This confirms that some additional gains can
indeed be captured from short-term price reversals by leaving the
portfolio unhedged for a few days. If instead of hedging every day, the
trader hedges only every two or three days, then higher profitability
adequately compensates for taking larger risks while simultaneously
saving on transaction costs. However, leaving the strategy unhedged for
more than a week makes it less attractive on a risk-adjusted basis.
While such a strategy allows traders to benefit from short-term price
reversals, it can suffer significant losses if the portfolio is left unhedged
for too long.
The second important driver of VRP profitability is the choice of the
hedging delta. Up until now, VRP strategies have been hedged with
deltas calculated using implied volatilities. This choice was motivated
by the desire to eliminate the stochastic term in (9.1). However, this
choice is somewhat inconsistent with the entire investment concept
behind VRP. We have already highlighted that the option’s payoff can be
perfectly replicated by trading futures only if the hedging delta is
calculated using the future actual volatility, which, of course, is not
known in advance. If the whole reason to trade VRP is driven by the
view that options are overpriced, which means that implied volatility is
too high relative to future realized volatility, then why would we be
hedging using a volatility that we ourselves do not even believe to be
correct?
To illustrate VRP sensitivity to the choice of the hedging delta, we
scale the hedging volatility by multiples of 0.5, 0.75, 1.25, and 1.5 of the
prevalent implied volatility. The multiple of 1.0 represents the base
case. The results of this experiment are summarized in Fig. 9.10.
Fig. 9.10 The performance of VRP strategy for 60 DTM ATM straddle versus scaled implied
volatility used for calculating hedging deltas (WTI, 2000–2022)
The most interesting takeaway from this analysis is P&L
improvement from hedging with low volatility and P&L decline for
hedging with high volatility. To interpret this result, recall from Fig. 9.1
and Table 9.1 how poorly the strategy of selling the naked straddle
performs. Selling unhedged straddles suffers from large losses when
futures periodically deviate too far from the initial ATM strike. This is
consistent with the overall trendiness of oil futures already established
in directional risk premia strategies. In the VRP strategy, when the
option crosses the strike and moves ITM, the sooner the seller hedges
changing delta, the better the protection against losses if the trend
persists. Using the lower volatility to calculate deltas does indeed
induce such a faster reaction. Lower input for the implied volatility
reduces the remaining variance in the price of the option and brings the
option payoff function closer to its terminal hockey stick-like payoff.
The deltas computed in this manner, therefore, increase for calls and
become more negative for puts when options are ITM. This forces the
dealer to hedge more aggressively in the direction of the trend. In
contrast to the reduced-hedging example, however, such aggressive
hedging increases transaction costs.
Even though one might easily jump to the conclusion that hedging
with a lower volatility is a better way to go, this approach has its own
challenges. If we hedge conventionally, using what we perceive to be an
incorrect implied volatility, then the second term in the Eq. (9.1) drops
out and the mark-to-market of the P&L becomes less volatile. In
contrast, if we hedge with any other volatility, even with a more
accurate one, then daily P&L will fluctuate along with dF, and the
standard deviation of P&L increases. This can also be seen from Fig.
9.10. Should we hedge then with a volatility that we believe to be
incorrect and benefit from relatively stable daily P&L, or is it better to
stick with our own opinion about future volatility and accept larger
daily swings?
While we may believe that we are right and the market is wrong,
our risk managers and controllers may have a different opinion. From
their independent perspective, the market is fair. Therefore, whatever
delta is implied by the market volatility should be the right one, and if
our creative delta-hedging strategy results in additional P&L from
futures, then it should be treated as a speculative position. Obviously, if
risk managers view an option price to be fair, then they should not be
letting the volatility trader sell an option to begin with. Some debate
with managers becomes inevitable, and to avoid it, many gamma
traders opt for a less controversial route and use conventional Black
deltas based on exchange-published futures settlement prices for the
base hedging case.
Despite the evidence of systematic biases highlighted by VRP
hedging alternatives presented in this section, actual optimization
decisions are usually made more tactically only when the volatility
trader has particularly strong views about the short-term expected
behavior of the market. To a certain degree, the problem of delta
hedging also has elements of quantamental trading, where a trading
algorithm is combined with human intervention. It would be a
disservice to attempt to provide readers with any more precise
systematic guidance on how frequently to rebalance a short VRP
portfolio, and what volatility to hedge it with, but these case studies
may steer hedgers towards better independent decisions.
The primary subjective decision in trading VRP, like in any other
systematic oil strategy, is the identification of regimes in which the
strategy is more likely to perform. For the oil VRP, such regimes are
mostly driven by the demand for hedging and the behavior of large
market participants.
Fig. 9.11 20 DTM VRP equity history for 0.25-delta puts (25P), zero-delta straddles (0S), and
0.25-delta calls (25C) (WTI, 2000–2022)
The idea of selling oil options as an overpriced insurance has
proven to work remarkably well for a while. Between 2000 and 2014,
VRP strategies generated impressive risk-adjusted returns with Sharpe
ratios exceeding 1.0 for many moneyness and maturity configurations.
After that, by and large, the risk premium contained in oil options
followed the path of Keynesian normal backwardation. It gradually
disappeared, like the structural futures risk premium did
approximately a decade earlier. The opportunity to make easy money in
any competitive market rarely lasts long. At some point, the strategy’s
consistent profitability motivates enough participants to invest in the
capability needed to manage the strategy, which helps to bring the
market towards an equilibrium.
The VRP strategy became a victim of its own success. As the
business of passive commodity investments struggled under the
pressure of contango and punitive rolling costs, capital shifted towards
more dynamic strategies designed to capture alternative risk premia. To
make it easier for investors, the VRP concept was packaged into
investable indices, and the cumbersome task of delta hedging was
effectively outsourced to index providers. Such investable volatility
indices provided large pools of capital held by institutional investors
with access to what used to be an obscure investment opportunity
previously dominated by oil quants. The barriers to entry were lifted,
and once again risk-bearing investors were compensated for providing
capital rather than for having any particularly unique trading skills. As
a consequence, the reward for offering relatively simple services of risk
absorption cratered. Another factor that contributed to the structural
break in VRP is a significant change in hedging strategies run by US
shale producers, a topic that we will discuss in more detail in the next
chapter.
The profitability of the VRP strategy will likely increase again, when
more buyers come to the market to buy oil insurance. This usually
happens shortly after the crisis that often forces insurance writers out
of business, resulting in higher premiums. At the time of writing this
book, there are indeed some indications that the Covid-19 pandemic
could well be another turning point, as the VRP performance in the
following two years improved substantially. This is driven by increasing
demand for hedging and the exodus of many option sellers after the
period of extreme volatility.
Even though the simple VRP strategy became less attractive to
financial investors, the aggregate risk premium in oil options has not
vanished. Instead, it spread out across various strikes and maturities,
but not in a uniform manner. Some options became consistently more
expensive than others. Like futures traders, many professional volatility
traders have also switched from outright directional bets to relative
value trades. In relative VRP strategies, one aims to sell options with
higher risk premia and hedge them by buying options with lower risk
premia.
A professional oil volatility portfolio is typically made up of many
relative value strategies based on the primary VRP building blocks
presented in this chapter. For example, the presence of the VRP smile
suggests that OTM calls, while being cheaper in dollar terms than ATM
calls, appear to contain a larger risk premium. One can then construct a
relative value strategy that takes advantage of the VRP call skew by
selling a larger quantity of OTM calls versus buying a smaller quantity
of ATM calls. Such a trade, known as a ratio call spread, is very common
in the marketplace. The weights between two legs are often chosen for
the trade to be either volatility-neutral or premium-neutral. However,
to successfully trade options across different moneyness, one needs to
better understand important quantitative linkages that connect prices,
which we discuss in the next chapter.
A relative value option portfolio typically includes options with
multiple expirations. For example, the decreasing term structure of VRP
may suggest hedging the sale of a short-term ATM option by buying a
longer-term ATM option that is less overpriced or even underpriced. If
the two options are traded in equal volumes, then the resulting
calendar spread trade is net short gamma, since gamma of the short leg
exceeds gamma of the longer-dated leg. However, the trade is net long
vega, since vega of the longer-dated leg exceeds vega of the short leg.
Buying deferred options can also be used as a valuable risk mitigant to
the base VRP strategy, as it provides significant mark-to-market offsets
when futures experience large moves. The weights between the two
legs can be adjusted to make the trade either gamma or vega neutral.
The proper construction of such strategies requires a more elaborate
modeling of the local volatility term structure, which we focus on in
Chap. 11.
The number of permutations that can be constructed from
individual options and distinct hedging strategies is nearly infinite, but
they are often based on the main features of VRP described in this
chapter. In addition, professional volatility traders rarely even hold
options until expiration. They might enter the trade when the
opportunity caused by end-user flows is particularly attractive and then
attempt to exit the trade once imbalances normalize. In the meantime,
they would delta hedge based on a particular model to preserve the
value of the option until the opportunity comes to exit the trade. This
brings an important additional dimension to volatility trading since the
trader is no longer betting on implied volatility versus gamma-scaled
realized volatility during the entire lifespan of the option. The trader
must also manage P&L fluctuations that come from changes in implied
volatility itself, which is often referred to as vega trading, the topic we
discuss next.
References
Bouchouev, I., & Johnson, B. (2022). The volatility risk premium in the oil market. Quantitative
Finance, 22(8), 1561–1578.
[MathSciNet][Crossref][zbMATH]
Doran, J. S., & Ronn, E. I. (2006). The bias in Black-Scholes/Black implied volatility: An analysis
of equity and energy markets. Review of Derivatives Research, 8(3), 177–198.
[Crossref][zbMATH]
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(12), 2541–2552.
[Crossref]
Ellwanger, R. (2017). On the tail risk premium in the oil market, Bank of Canada Working Paper,
46.
Jacobs, K., & Li, B. (2023). Option returns, risk premiums, and demand pressure in energy
markets. Journal of Banking and Finance, 146, 1–26.
[Crossref]
Kang, S. B., & Pan, X. (2015). Commodity variance risk premia and expected futures returns:
Evidence from the crude oil market, SSRN.
Lo, A. W. (2002). The statistics of Sharpe ratios. Financial Analysts Journal, 58(4), 36–52.
[Crossref]
Prokopczuk, M., Symeonidis, L., & Simen, C. W. (2017). Variance risk in commodity markets.
Journal of Banking and Finance, 81, 136–149.
[Crossref]
Trolle, A. B., & Schwartz, E. S. (2010, Spring). Variance risk premia in energy commodities. The
Journal of Derivatives, 17(3), 15–32.
[Crossref]
Footnotes
1 Note that this analysis is presented on the trade level. Since the trades are entered once a
month and held for three months, multiple trades are open at the same time. The analysis of the
continuously managed strategy of selling options is presented later in this chapter.
4 The challenges that arise with annualization of risk-adjusted returns are well explained by Lo
(2002).
5 As before, we calculate the Sharpe ratio assuming zero risk-free interest rate. Since we do not
include the cost of capital and exchange margin in our calculations, we do not include any
interest that could be received on the initial margin either. Overall, the impact of financing costs
on the performance of VRP strategies is relatively minor. In addition, both the numerator and
the denominator of the Sharpe ratio here are computed in dollars per barrel, not in percent.
6 Another popular reduced-hedging strategy is to trade fewer futures than required by the
model. The details of this strategy can be found in Bouchouev and Johnson (2022).
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023
I. Bouchouev, Virtual Barrels, Springer Texts in Business and Economics
https://doi.org/10.1007/978-3-031-36151-7_10
– Oil volatility traders like to quip that if you do not know who the fool
in the market is, then it is probably you. This does not literally mean
that someone is acting foolishly. Rather, it highlights the importance
of understanding the behavior of large market participants.
– Vega-trading strategies require modeling the evolution of the entire
volatility smile. Common smile heuristics, such as sticky moneyness
and sticky strike, perform poorly in the oil market. Instead, a more
realistic dynamics for the underlying futures contract must be
imposed.
– Empirical study of the relationship between futures prices and
implied volatilities guides model selection. The normal model
removes an artificial skewness embedded in the conventional
lognormal framework, but neither of the two models is capable of
capturing extreme events.
– A novel quadratic normal model is developed based on the
perturbation method for the general diffusion equation. The option
pricing formula incorporates skewness and fat tails and relates
model parameters to market prices for benchmark collars and
strangles.
(10.1)
where VBL is Black vega. In addition to the standard Black delta, the
total delta in (10.1) acquired another component, which is equal to the
product of Black vega and the slope of the implied volatility function
with respect to futures. This second term in (10.1) is colloquially
referred to as a skew delta. It represents the change in the value of the
option resulting from the change in implied volatility, induced by the
change in futures.
Unfortunately, the smile today does not tell us much about its
possible shape tomorrow, so we do not really know the true hedging
delta. What we observe today is the slope of the smile with respect to K.
However, what we need for hedging is the slope of the smile with
respect to F, which arises in the second term of (10.1). To calculate the
latter, one must impose additional assumptions about the evolution of
the smile. Since the smile is predominantly maintained as a function of
option moneyness, the path of least resistance is to assume that implied
volatilities remain the same for a given moneyness.
Let us illustrate the dynamics of the implied volatility in the
simplified case when the smile is linearly decreasing with slope β > 0,
so that
Fig. 10.2 The volatility smile behavior under the sticky moneyness rule
where v0 is ATM volatility for the strike K = F. When futures move up,
then ATM volatility v0 always stays the same, but the implied volatility
for any fixed strike K rises. This is the consequence of a negatively
sloped volatility smile and the positivity of the second term in (10.1)
since
Likewise, when futures fall, then implied volatilities for all strikes
decrease making options somewhat less expensive. Such behavior is
counterintuitive and potentially indicative of a problem with the
assumption of sticky moneyness. Since oil futures tend to have larger
downside gaps and percentage volatility is generally inversely related
to the futures price, it would be strange for the optionality to become
less valuable following the futures move in the direction of greater
uncertainty.
The problem manifests itself more clearly when one considers its
impact on the option’s delta. Since for a negatively sloped smile, the
skew delta is positive, the total hedging delta for sticky moneyness
always exceeds a conventional Black delta:
The second term in (10.1) then vanishes because the smile is static;
it depends only on fixed F0 but does not depend on future price F. The
total hedging delta in the sticky strike case is identical to Black delta for
calls:
The sticky strike heuristic forces the trader to depart from the path
of least operational resistance determined by sticky moneyness.
Instead of doing nothing and letting the smile slide in parallel, the
sticky strike requires the trader to constantly repivot the smile to the
new ATM level and then tweak its shape by moneyness, as required by
operational standards.
At this point, one may wonder why not just keep the entire smile in
terms of fixed strikes? Theoretically, it is doable, especially for short
time periods when the range of popular strikes does not change. In the
long run, however, keeping the skew by strike is impractical. The
market always talks in terms of moneyness, and it uses ATM volatility
as a pivot point for the construction of the smile. Nearly all risk and
accounting systems require smiles to be saved by moneyness to
facilitate standardization across multiple assets. While the trader can
maintain the sticky strike smile for decision making, its system-
required shape by moneyness must be constantly bent. The reward for
this hassle comes from the elimination of undesirable and
counterintuitive skew delta corrections. The improvement, however, is
marginal. The delta of an ATM straddle is still positive, which is caused
by the assumption of lognormality. In the previous numerical example
even with no contribution from the skew delta, the ATM straddle still
has 12% delta. It is half as bad compared to the sticky moneyness case,
but it is still not a viable hedging solution as one should expect ATM
straddle delta to be closer to zero.
The obvious question arises now whether it is possible to come up
with other heuristics that shift artificially skewed Black deltas in the
right direction. Yes, it can be done. Nothing can stop the trader from
specifying any other arbitrary dynamics of the smile. The trader can
draw a certain glide path for an ATM volatility that does not simply
slide along the same shape, but instead increases or decreases with any
chosen slope. An equally arbitrary skew can then be attached to this
glide path of ATM volatility to complete the smile dynamics. However,
all such heuristics amount to more complex but still artificial tweaks to
a base pricing model that does not represent the real market behavior.
The problem can only be solved by treating its cause and not the
symptoms.
One challenge is the natural asymmetry embedded in the lognormal
distribution. While the assumption of lognormality might be reasonable
for some financial markets that tend to grow over time, it becomes
problematic for mean-reverting oil markets, where the underlying price
distribution is much more symmetric. In Chap. 8, we illustrated how the
desired symmetry can be restored if one switches from lognormal to
normal measures of the realized volatility. In the following section, we
take a similar route for the implied volatility and compare the behavior
of volatility smiles, as they are seen from two alternative angles.
Since the function C(F, t) must solve the diffusion equation with
local volatility σ(F), we substitute this decomposition into (8.8), and
make use of the fact that CBC(F, t) also solves the same equation but
with constant coefficient σA. The main idea of the perturbation method
is to retain only the terms of order ε(F) and discard higher-order terms,
whose contribution is smaller for a relatively small ε(F).
It is shown in Appendix C that this substitution leads to the
following equation for U(F, t):
(10.3)
where, as before,
denotes the scaled normal moneyness. It follows that U(F, t) also solves
the BSM equation with constant coefficient σA, but its right-hand side
has an additional term which is proportional to the perturbation
function ε(F) and Bachelier gamma.
To solve (10.3), we need to supplement it with a boundary
condition. At the expiration time, C(F, T) is defined by the option’s
terminal payoff. The same payoff is also produced by the Bachelier
formula CBC(F, T) at expiration. Therefore, the terminal boundary
condition for the skew correction function is zero:
(10.4)
where
denotes Bachelier vega with the constant normal volatility σA, and
(10.5)
References
Andersen, L. (2011). Option pricing with quadratic volatility: A revisit. Finance and Stochastics,
15(2), 191–219.
[MathSciNet][Crossref][zbMATH]
Bouchouev, I. (1998). Derivatives valuation for general diffusion processes, Proceedings of the
Annual Conference of the International Association of Financial Engineers, New York, USA, pp.
91–104.
Bouchouev, I. (2000, August). Black-Scholes with a smile. Energy and Power Risk Management,
pp. 28–29.
Carr, P., Fisher, T., & Ruf, J. (2013). Why are quadratic normal volatility models analytically
tractable? SIAM Journal on Financial Mathematics, 4(1), 185–202.
[MathSciNet][Crossref][zbMATH]
Castagna, A., & Mercurio, F. (2007). The vanna-volga method for implied volatilities. Risk, 20(1),
106–111.
Grunspan, C. (2011). A note on the equivalence between the normal and the lognormal implied
volatility: A model free approach, SSRN.
Ingersoll, J. E., Jr. (1997). Valuing foreign exchange rate derivatives with a bounded exchange
process. Review of Derivatives Research, 1, 159–181.
[Crossref][zbMATH]
Rebonato, R. (2004). Volatility and correlation: The perfect hedger and the fox. Wiley.
[Crossref]
Schachermayer, W., & Teichmann, J. (2008). How close are the option pricing formulas of
Bachelier and Black-Merton-Scholes? Mathematical Finance, 18(1), 155–170.
[MathSciNet][Crossref][zbMATH]
Zü hlsdorff, C. (2001). The pricing of derivatives on assets with quadratic volatility. Applied
Mathematics Finance, 8(4), 235–262.
[Crossref][zbMATH]
Footnotes
1 We only provide a simple illustration of this topic. For a more advanced exposition, we refer
to Rebonato (2004), Gatheral (2006), and Derman and Miller (2016).
2 For brevity, we omit the rigorous derivation of this approximation. It is again based on the
Taylor series expansion for both formulas, but with more cumbersome mathematical details
which are less essential for our goals. We refer to Schachermayer and Teichmann (2008) and
Grunspan (2011) for a more detailed comparison of the two formulas, and more advanced
approximation formulas.
3 Recall that we are using zero interest rate. In the general case, deltas must be discounted.
4 The method of linearization and the related parametrix method for the diffusion equation
were originally applied to option pricing in Bouchouev (1998, 2000).
5 Quadratic parametrizations of the local volatility function have been considered by several
authors, including Ingersoll (1997), Zü hlsdorff (2001), Andersen (2011), and Carr et al. (2013),
but resulting option pricing formulas are considerably more complicated.
6 The use of three parameters that correspond to the first three moments of the price
distributions is often used by market-makers to characterize the dynamics of the volatility
smile in financial markets. For example, Castagna and Mercurio (2007) developed a popular
vanna-volga model for options in foreign exchange markets.
Part IV
Over-the-Counter Options
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023
I. Bouchouev, Virtual Barrels, Springer Texts in Business and Economics
https://doi.org/10.1007/978-3-031-36151-7_11
Fig. 11.1 Profit and Loss of hedging program by the Finance Ministry of Mexico. Sources:
Auditoría Superior de la Federació n, Secretaría de Hacienda y Crédito Pú blico, Oxford Institute
for Energy Studies
The fact that Mexico was able to avoid paying any structural VRP
over the years by timing the execution well does not mean that option
dealers did not make any money. They sure did. The deal highlights the
importance of the difference between the actuarial value of an option
and its fair value determined by delta hedging, as discussed in Chap. 9.
For volatility market-makers, this hedging program turned into a high-
stakes skill competition for a slice of the approximately one-billion-
dollar premium paid by the government annually. The winners and
losers in this competition are often determined by how accurately they
can estimate the APO volatility discount for pricing the deal and
corresponding deltas for managing the risks.
This sovereign hedge by Mexico is only one among many other
examples of a customized OTC derivative, but it illustrates well the
complexity of the task faced by volatility traders. These complex deals
bring unparalleled profit opportunities to the dealers, but they are
inevitably accompanied by equally large risks of mis-hedging if the
price dynamics and resulting hedging ratios are not captured correctly.
The quantitative challenges here are substantial. In addition to tricky
handling of price averaging and basis risks, the trader must also
incorporate the volatility term structure, as the tenor of many OTC
derivatives spans periods covered by multiple futures contracts.
Up until now, we have dealt only with options written on a single
futures contract that allowed us to simplify the problem by fixing the
corresponding maturity. However, even a typical monthly APO volatility
for WTI depends on the volatilities of two different futures due to the
rolls that must occur sometime during the month. Likewise, the
volatility of an annual term APO is determined by volatilities of thirteen
vanilla options that correspond to prompt futures throughout the
calendar year. The volatility of a term APO depends on the slope of the
implied volatility term structure, which is particularly sensitive to
short-term fundamental uncertainty in the physical market. The
analysis of the volatility term structure is a prerequisite for valuing not
only APOs, but also many other OTC options which will be introduced
later in this chapter.
(11.4)
(11.5)
Figure 11.2 illustrates the relationship between σG(τ) and v(T) for
options expiring up to two years forward.
Fig. 11.2 An example of implied and exponentially decreasing local volatility with σ∞ = 0.1,
σ0 = 0.2, k = 1
(11.7)
(11.9)
(11.10)
(11.11)
In (11.11), we omitted the subscript that corresponds to the normal
volatility as we will see shortly that the same formula can also be used
for lognormal volatility.
The local volatility of an APO is the same as local volatility of futures
outside of the pricing period, but it follows from (11.10) that within the
pricing period it is adjusted by the linear multiplier
(11.12)
Since the integration of the local variance must cover the entire
lifespan of the option, one needs to separate two periods when the
option is within and outside of the pricing period. If an APO is already
pricing out, i.e., Ta ≤ t ≤ T, then only a single integral is calculated from
current time t until maturity T. However, if the option is yet to start
pricing and t < Ta, then the integration period (t, T) must be split into
two sub-periods (t, Ta) and (Ta, T) as the local volatilities in the two
periods are different.
In other words, (11.11 and 11.12) can be combined as
(11.13)
What made this transformation possible for the ABM process is the
fact that the average of normally distributed variables is itself normal. If
F(t, T) is normal, then A(t) and x(t) are also normal, and the volatility
coefficient in (11.10) does not depend on any spatial variable. This
property does not hold for the GBM process or other diffusions, and for
the general volatility function a closed-form expression for the price of
an APO is not possible. However, the same volatility transformation
(11.13) can be used as a high-quality analytic approximation for other
diffusion processes.
By using this approximation for other diffusions, one effectively
assumes that the local volatility function behaves similarly with respect
to x and F, i.e., σ(x, t) ≈ σ(F, t), in which case (11.10) turns into a regular
pricing equation with respect to x. For example, under the lognormal
assumption it assumes that the average of lognormal variables is
approximately lognormal. From the practical standpoint, this is a
relatively benign assumption because the real oil price distribution, as
we have seen before, is closer to normal than to lognormal anyway, and
in the normal case the formula holds exactly. It means that one can still
use the Black formula to price APOs with the volatility given by (11.13),
which is now understood in percentage terms.
Let us illustrate this first for the constant local volatility
(11.14)
Fig. 11.3 Volatility reduction for an APO with T − Ta = 1 and constant local volatility
The volatility discount for an APO is not too sensitive to the choice
of normality or lognormality, but it varies substantially with the
steepness of the volatility term structure in either framework. To
illustrate this, let us modify the example above. Instead of using
constant volatility, we calculate the APO volatility discount for an
exponentially decreasing local volatility specified in (11.6) with the
same set of parameters that we used above. The corresponding
volatility behavior is shown in Fig. 11.4.
Fig. 11.4 Volatility reduction for an APO with T − Ta = 1 and exponential decreasing local
volatility
The local volatilities for a vanilla option and for an APO are
represented by the corresponding dashed lines, which coincide prior to
the pricing period. The solid lines are implied volatilities which are the
quadratic means of the corresponding local volatilities for the same
time to maturity. As before, within the pricing period, a volatility
multiplier that linearly decreases with time to maturity applies to an
APO. For an exponential volatility function, however, volatility
attenuation is stronger as compared to the case of constant volatility, as
the multiplier dampens local volatility when it is at its highest near the
option expiration. For example, at the start of the averaging period APO
volatility vAPO(1) is approximately 50% of the corresponding vanilla
volatility in contrast to 60% for the constant volatility assumption.
Likewise, the vAPO(2) discount relative to the volatility of vanilla option
is also larger in comparison to the case of constant volatility. The
steeper the term structure of the local volatility curve, the larger the
discount for APOs relative to vanilla options across all maturities.
If the option dealer can capture even a fraction of the APO price
discrepancy driven by a more accurate modeling of the volatility term
structure, it will definitely lead to a stellar career in the oil market.
Since many less sophisticated market participants use off-the-shelf
software packages for APO pricing that are often constrained to
constant volatility specification, they could easily misprice an APO. The
APO volatility discount is sensitive to the slope of the volatility term
structure, which in turn, is linked to the speed of price mean-reversion.
The impact of the volatility term structure becomes even more
significant for another important type of OTC options that expire prior
to the expiration of the underlying futures.
For clarity, we explicitly show the expirations of both the option and
the underlying futures as two arguments of the volatility function, and,
as before, we omit the subscript that differentiates between lognormal
and normal cases. For a vanilla option, where the option expiration is
tied to the expiration of futures, one can assume that T0 ≈ T and,
therefore,
Fig. 11.5 Volatility reduction for an EEO with T0 − t = 1 and exponential decreasing local
volatility
One can see some similarities between EEOs and APOs. For an APO,
the local volatility is dampened by a linear multiplier once the option
moves into the pricing period. For an EEO, the local volatility for t > T0
vanishes entirely, as the option no longer exists then. However, direct
comparison between the prices of these two types of OTC options is not
straightforward, as prices depend on the aggregate effect of the
different times to maturity and the slope of the volatility term
structure.
While EEOs do trade from time to time in the OTC market, they
primarily function as a building block for a more commonly traded OTC
derivative, known as a swaption, or an option on a swap. The standard
oil swap settles based on the calendar month average of prices for the
prompt futures contract. Like an APO, an oil swap typically trades in the
form of quarterly and annual strips with monthly settlements. This
means that the swap price can effectively be represented as the linear
combination of the M futures that make up the swap, which are also
referred to as swaplets:
(11.15)
(11.16)
(11.17)
References
Blas, J. (2017, April 4). Uncovering the secret history of Wall Street’s largest oil trade,
Bloomberg.
Bouchouev, I. (2000, July). Demystifying Asian options. Energy and Power Risk Management, pp.
26–27.
Geman, H. (2005). Commodities and commodity derivatives: Modeling and pricing for
agriculturals, metals and energy. Wiley.
Kemna, A. G. Z., & Vorst, A. C. F. (1990). A pricing method for options based on average asset
values. Journal of Banking and Finance, 14(1), 113–129.
[Crossref]
Levy, E. (1992). Pricing European average rate currency options. Journal of International Money
and Finance, 11(5), 474–491.
[Crossref]
Lipton, A. (2001). Mathematical methods for foreign exchange: A financial engineer’s approach.
World Scientific.
[Crossref][zbMATH]
Merton, R. C. (1973). Theory of rational option pricing. The Bell Journal of Economics and
Management Science, 4(1), 141–183.
[MathSciNet][Crossref][zbMATH]
Swindle, G. (2014). Valuation and risk management in energy markets. Cambridge University
Press.
[Crossref]
Turnbull, S. M., & Wakeman, L. M. (1991). A quick algorithm for pricing European average price
options. Journal of Financial and Quantitative Analysis, 26(3), 377–389.
[Crossref]
Wilmott, P., Dewynne, J., & Howison, S. (1993). Option pricing: Mathematical models and
computation. Oxford Financial Press.
[zbMATH]
Footnotes
1 It should be noted that oil exchanges did exist in Pennsylvania and New York as early as in
the 1870s, see Giddens (1947). However, all of them were closed within a decade under
monopolistic pressure from the Standard Oil Company.
2 For a colorful description of the history of this program, see Blas (2017).
3 Prior to 2019, the Maya formula was calculated as 40% WTS (West Texas Sour), 40% HSFO
(High-Sulfur Fuel Oil), 10% LLS (Louisiana Light Sweet), 10% Dated Brent, plus a K-factor set
by Mexico. To reflect the growth in US shale and changes in the sulfur specification for marine
fuel, in 2019 the formula changed to 65% WTI Houston, 35% ICE Brent, plus a K-factor.
4 Merton (1973).
5 There is a large body of literature on pricing APOs. Kemna and Vorst (1990) formulated the
partial differential equation for the price of an APO and found a closed-form solution for an
option on the geometric average of prices. Such a solution is possible because the geometric
average of lognormal variables is also lognormal. Subsequently, many authors used the idea of
geometric averages to develop approximations of the probability density function for the
arithmetic average of lognormal variables and corresponding approximations for APO prices,
such as the formulas of Turnbull and Wakeman (1991) and Levy (1992). For other APO pricing
methods, see also Wilmott et al. (1993), Lipton (2001), and Geman (2005). The method
described in this book was originally suggested in Bouchouev (2000).
6 For non-zero interest rates ri that correspond to times Ti, the swap value S is determined by
equating the present value of the future cash flows to zero: Solving it
7 For more details on the time-scale separation for pricing natural gas and power options, see
Swindle (2014).
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023
I. Bouchouev, Virtual Barrels, Springer Texts in Business and Economics
https://doi.org/10.1007/978-3-031-36151-7_12
(12.1)
(12.2)
In the final formula, the implied variance v2(N − 1) for the periods
N − 1 is combined with the local variance σ2(N) in the last period to
produce the variance v2(N) for N periods.
These recursive relationships can be inverted for the local variance
σ2(N), which is expressed as the weighted difference between two
implied variances for consecutive periods:
The last formula is the discrete version of (12.2). This simple
recursive algorithm to recover the local volatility term structure from
the implied volatility curve is often referred to as volatility
bootstrapping.
The local volatility can only be calculated if the implied volatility
does not decline too fast with increasing time to maturity. To illustrate
non-negativity restriction (12.2) with a numerical example, let us
assume that the implied volatilities observed in the market for the first
three monthly futures contracts are
The second row starts with σ(2, 2), which is the volatility of the
contract that is second nearby today, but its volatility is measured
during the following month when this contract becomes the prompt.
Since the total two-month implied variance for the contract is the sum
of its variances during the first month and the second month,
We can then apply the bootstrapping technique and back out the
local volatility σ(2, 2) from the implied volatility v(2) and the local
volatility during the first month σ(1, 2) calculated in the previous step:
The knowledge of σ(2, 2) gives us the pivot spot volatility for the
second row, and we can fill in the rest of the second row for all N > 2,
using the specified exponential function, as follows
Moving on to the third row, the implied variance for the third
contract is the average of the local variances over three periods
Since we already know the local variance for the first two rows and
we also know the implied variance that contains the cumulative
information from all three periods, the local volatility during the third
month can be bootstrapped, as follows:
After obtaining the spot volatility during the third month σ(3, 3), we
again calculate σ(3, N) for all N > 3 using the exponential specification
for the local volatility. The process continues until the entire local
volatility matrix is filled in a unique way.
Having such a local volatility matrix in place is important for pricing
swaptions and other exotic options described in the previous chapter.
To illustrate its application to swaptions, we assume that the volatility
matrix in Fig. 12.1 is constructed for calendar months, rather than for
futures contracts, as one can easily convert between the two. To price a
swaption, the trader must calculate the quadratic mean of local swaplet
volatilities over the life of the swaption. For example, for a three-month
swaption into a six-month swap used in the previous chapter, the
volatility of the swaption spans the highlighted area in Fig. 12.1, which
is made up of eighteen discrete local volatility blocks. The quadratic
average of swaplet volatilities with maturities from four to nine months
over next three months produces the swaption volatility. In this
example, the swaption volatility is only 0.267, which is lower than any
of the implied volatilities for swaplets. It would be a major mistake by
the trader to price a swaption by calculating the quadratic mean of
implied volatilities for swap components. What goes into averaging are
the local variances of swaplets, not the implied ones.
In addition, if some swaption prices are observable in the broker
market, then this flexible format of the local volatility matrix also
allows its parameters to be further calibrated to the market prices of
these swaptions. For example, we still have parameters σ∞ and k that
remain entirely under our control. We can choose them to approximate
market prices of swaptions, EEOs, or some other exotic derivatives that
may be quoted in the market. This is typically done in an ad hoc manner
by trying out different combinations of σ∞ and k, while monitoring how
the entire local volatility matrix responds to different combination of
inputs. If more flexibility is desired, the decay parameter k(t) can also
be made time-dependent.1
A similar volatility matrix can also be built for the two-factor model
specification (11.17). Its construction remains largely identical, except
for an additional degree of freedom introduced by the factor correlation
ρ. Using it, however, often causes more harm than it adds value as many
combinations of parameters may end up producing similar results. For
example, one can achieve a drop in the swaption volatility either by
lowering the factor correlation, or alternatively, by steepening the local
volatility function. The market, however, does not give us enough
observable information to distinguish between these two alternatives.
The problem becomes over-parametrized, and unless some model
parameters are fixed, it becomes prone to instability.
The inverse calibration problem in the time direction is relatively
simple analytically. However, this problem is inherently incomplete, as
there are too many degrees of freedom relative to the number of
available observations. The oil option market does not contain enough
information to allow us to deduce the rich dynamics of the volatility
time-dependency, so the term structure modeling is often
supplemented by econometric analysis of futures prices. In contrast, an
inverse problem in the strike dimension where the local volatility is
allowed to be space-dependent faces the opposite challenge. It is
relatively complete in terms of the availability of granular data for
option prices with different strikes, but it is more challenging from the
mathematical point of view. We turn to this problem next.
The first term comes from differentiation with respect to the lower
limit of integration. It is equal to the value of the integrand evaluated at
the point K, which is zero, and, therefore, this term vanishes.
We then take the second derivative with respect to K, differentiating
it again with respect to the lower limit of integration, which results in
(12.4)
This is a rather remarkable result. The given set of option prices can
be transformed into the set of market-implied risk-neutral probabilities
simply by differentiating the option prices twice with respect to the
strike price. This relationship was once again first discovered by
Bachelier in his seminal doctoral thesis.2
The transformation (12.4) is possible only because of the very
special terminal payoff of the call option, which is given by max(0,
FT − K). The first derivative of the payoff with respect to K taken with
the negative sign is given by the so-called digital, or the Heaviside,
function. This function is equal to either one or zero, depending on
whether FT is above or below K:
(12.5)
(12.6)
The payoff of such call spreads are shown in Fig. 12.2 for decreasing
strike increments.
Fig. 12.2 The payoff of the call spread with strikes K and K + dK approaches the payoff of the
digital option when dK → 0
The maximum payoff of one call spread is dK. Therefore, the
maximum payoff of holding 1/dK units of such infinitesimally tight call
spreads is equal to one dollar. In the limit when dK → 0, the call spread
converges to a digital option D(FT − K), which is defined by (12.5). The
digital option pays one dollar if the futures price is above the strike K at
expiration and zero if the futures price is below the strike:
The terminal payoff of the butterfly call spread is shown in Fig. 12.3
for decreasing strike increments.
Fig. 12.3 The terminal payoff of 1/(dK)2 units of call butterflies with strikes K − dK, K, and
K + dK approaches the Dirac delta function when dK → 0
The butterfly trade is certain to generate a non-negative payoff
which is defined by the triangular region with a width of 2dK and peak
height dK. Therefore, if we hold 1/(dK)2 units of such butterflies, then
the area of the triangle is always equal to one, regardless of the size of
dK. If we let dK → 0, then the second derivative of the option price with
respect to K turns into the Dirac delta function, centered at K:
Note that since the area of the triangle in the butterfly payoff is
equal to one for any dK, the normalization property (A.4) of the Dirac
delta function that integrates to one is also satisfied. The butterfly
payoff is the discrete analogue of (12.6).
Obviously, the second derivative of the option price does not
explicitly trade in the market as it would have an impossible infinite
payoff, but its approximations in the form of tight butterflies with small
dK are quite common even on the exchanges, where the smallest strike
increment is only fifty cents per barrel. The maximum payoff of the
infinitesimally tight butterfly occurs when FT = K, and traders use
butterflies to explicitly bet on the probability of the price reaching the
specific level at expiration. Since at expiration the payoff of the butterfly
approximates the Dirac delta function centered at K, the market prices
of various butterflies with various strikes make up the probability
density function for the futures price to be at the level FT = K at
expiration.
Market-implied probability distributions are very helpful for
volatility traders in spotting arbitrage opportunities in option prices
across different strikes. The probability density function is generally
expected to be smooth. Any visible kink in the density function
indicates that prices for some options with nearby strikes might be
inconsistent. It is rather difficult to identify such an inconsistency by
looking at option prices or at their implied volatilities, which are
cumulative quantities, because averaging tends to smooth the effects of
any mispricing. In contrast, for local quantities, such as probability
densities, there is nowhere to hide, and any distortion in its shape
exposes a pricing anomaly. Looking at the implied probability
distribution is like looking at option prices under the microscope.
The recovery of the market-implied distribution from option prices
is undoubtably useful in identifying potential trading opportunities.
However, to capture such opportunities, additional information is
needed. The market-implied distribution provides us only with a
snapshot of the price distribution taken at a given time, but it tells us
little about the evolution of prices between today and the option’s
maturity. Since the opportunity in trading options is captured by delta
hedging, which must be done throughout the life of the option, to
calculate the deltas, the trader needs to model the evolution of the
entire stochastic process that produces this distribution. This
information is contained in the local volatility function, which is also
needed for pricing and hedging OTC options.
(12.7)
(12.8)
Note that while the actual local volatility function depends on F and
t, when it is reconstructed from option prices, its functional arguments
are replaced with K and T.
To understand this representation better, let us revert to a discrete
approximation of partial derivatives of option prices with respect to
strikes and maturities. We know from the previous section that the
denominator of (12.8) is the probability density function, which can be
approximated by market prices of call butterflies. Similarly, the
numerator of (12.8) can be approximated by an infinitesimally tight call
spread with respect to T.
Consider a calendar call spread which consists of long and short call
options with the same strike K and corresponding expiration times
T + dT and T. If we hold 1/(dT) units of this calendar call spread and let
dT → 0, then such a trading position becomes the partial derivative of
the option price with respect to T:
Therefore,
(12.9)
Furthermore, vQN(K, F) represents the average of the local volatility
between F and K, which is verified by direct integration:
Fig. 12.4 Implied volatilities generated by the local volatility σ(F) = 20 + 0.02 (F − 60)2
Avellaneda, M., Friedman, C., Holmes, R., & Samperi, D. (1997). Calibrating volatility surfaces via
relative entropy minimization. Applied Mathematical Finance, 4(1), 37–64.
[Crossref][zbMATH]
Bouchouev, I., & Isakov, V. (1997). The inverse problem of option pricing. Inverse Problems,
13(5), L11–L17.
[MathSciNet][Crossref][zbMATH]
Bouchouev, I., & Isakov, V. (1999). Uniqueness, stability and numerical methods for the inverse
problem that arises in financial markets. Inverse Problems, 15(3), R95–R116.
[MathSciNet][Crossref][zbMATH]
Bouchouev, I., Isakov, V., & Valdivia, N. (2002). Recovery of volatility coefficient by linearization.
Quantitative Finance, 2(4), 257–263.
[MathSciNet][Crossref][zbMATH]
Breeden, D. T., & Litzenberger, R. H. (1978). Prices of state contingent claims implicit in option
prices. Journal of Business, 51(4), 621–651.
[Crossref]
Carr, P., & Madan, D. (2001). Determining volatility surfaces and option values from an implied
volatility smile. In M. Avellaneda (Ed.), Quantitative analysis in financial markets (Vol. II, pp.
163–191). World Scientific.
[Crossref]
Chiarella, C., Craddock, M., & El-Hassan, N. (2003). An implementation of Bouchouev’s method
for short time calibration of option pricing models. Computational Economics, 22, 113–138.
[Crossref][zbMATH]
Dempster, M. A. H., & Richards, D. G. (2000). Pricing American options fitting the smile.
Mathematical Finance, 10(2), 157–177.
[MathSciNet][Crossref][zbMATH]
Pilipović, D. (2007). Energy risk: Valuing and managing energy derivatives. McGraw-Hill.
Rebonato, R. (2004). Volatility and correlation: The perfect hedger and the fox. Wiley.
[Crossref]
Footnotes
1 Another example of a discrete volatility matrix is constructed in Pilipović (2007), where
instead of imposing the exponential structure on local volatilities, an additional constraint is set
by equating long-term local volatilities to historical realized volatilities. Our preference is to
avoid explicitly tying market-implied volatility matrix to realized volatilities due to hedging
imbalances and the presence of the volatility risk premium documented in Chap. 9.
2 See Bachelier (1900). Bachelier’s brief statement of this formula is often overlooked in the
literature, where the formula is typically attributed to a more comprehensive work on this topic
by Breeden and Litzenberger (1978).
3 The equation was presented at several conferences in 1993 and subsequently published in
Dupire (1994).
4 This inverse problem has been extensively studied in academic literature. See, among many
others, Avellaneda et al. (1997), Bouchouev and Isakov (1997, 1999), Dempster and Richards
(2000), Carr and Madan (2001), Bouchouev et al. (2002), Chiarella et al. (2003), Alexander
(2008), Lipton and Sepp (2011), and references therein.
5 For a more detailed discussion of this topic, we refer to Gatheral (2006), and Derman and
Miller (2016).
6 For further reading on jump-diffusions and stochastic volatility models, we refer to Rebonato
(2004), Javaheri (2005), Gatheral (2006), and Derman and Miller (2016).
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023
I. Bouchouev, Virtual Barrels, Springer Texts in Business and Economics
https://doi.org/10.1007/978-3-031-36151-7_13
Fig. 13.2 Implied and corresponding normal local volatility for a CSO computed using the
Bachelier formula
The local volatility curve for time spreads is typically very steep. As
a result, the spread volatility rolls up the curve so fast that the gain
from vega negates a large portion of the option theta decay. The
steepness of the curve is driven by a combination of the much higher
fundamental risks embedded in short-term options and discounted
prices offered by storage operators for longer-term deals. Occasionally,
the discount for forward CSOs becomes so large that it even violates the
no-arbitrage restriction (12.2) on how fast implied spread volatility
term structure can decline.
The long-dated CSO that the derivatives trader acquires does not
stay cheap forever. In contrast to a physical storage facility, the lifespan
of the virtual storage is shorter. When the CSO expires, its time value
must decay to zero. It means that at some point prior to its expiration,
the CSO is likely to become expensive relative to its fair value that the
buyer expects to generate by delta hedging. Therefore, one challenge
for a synthetic storage strategy is to liquidate an option before the
escalation of theta turns into a real burden. But why would anyone in
the market at that time buy such an overpriced and rapidly decaying
financial asset?
Recall from Chap. 9 our study of options as insurance contracts that
carry a volatility risk premium paid by hedgers. The same argument
applies to CSOs. Here, the primary demand for hedging comes not from
end-users, but rather from financial investors. We have already seen
how punitive the negative roll yield is for long-only investors holding
futures in contango markets. By paying this roll yield, financial
investors effectively outsource the function of storage to carry traders,
but the magnitude of the cost to roll is subject to the vagaries of the
market. Buying a CSO provides financial investors with an interesting
alternative. Instead of taking the risk of an excessive roll cost in the
event of a super-contango, investors can hedge such a cost in the
financial market. They can pay the premium to buy CSO puts and
eliminate potentially the largest drag on the performance of their long
futures investment. Such an incremental demand for short-term CSO
puts from hedgers of the roll yield turns out to be substantial, especially
during times when the market is in a state of contango.
In addition, professional physical trading houses also like buying
short-term CSOs to place highly leveraged speculative bets on time
spreads. These options are used as effective wagers on the squeeze
probability that the market may either run out of oil or run out of
capacity to store it. CSOs are also routinely used for risk management
by trading desks to protect large speculative positions taken in the
futures market. Physical speculators trade time spreads based on
fundamental models that rarely look beyond just a few months ahead.
Their demand for spread volatility tends to be higher when futures
approach expiration. This hedging imbalance allows CSO holders to exit
the ownership of virtual storage just in time before its theta decay
starts to accelerate.
To summarize, the CSO strategy described here is akin to a three-
step virtual storage warehouse. First, one acquires a long-dated strip of
monthly CSOs from storage operators at a discounted “wholesale” price.
The time value of financial options is then protected by delta hedging.
Given the steepness of the volatility term structure and only a minimal
theta to cover, the bar for delta hedging gains is set rather low. Finally,
the trader parses a long-dated strip into individual month CSOs and
resells them to financial investors looking to hedge against the negative
roll yield, or to short-term physical speculators. An embedded volatility
risk premium allows these short-term sales to be made at higher
“retail” prices.
This strategy may sound like a remarkably simple way to make
money. One challenge is, of course, in knowing how much to pay for
such a long-dated optionality at the initiation of the trade. Another one
is the calculation of delta, which heavily depends on the pricing model
to be covered later in this chapter. But before we get into details of
spread option pricing, an important arbitrage boundary that plays the
central role in many spread option strategies must be introduced. This
boundary is somewhat easier to visualize first in the context of cross-
product and locational spread options.
The size of the angle α affects the length of the opposite side, which
represents the volatility of the spread σ(X − Y). If the two assets are
perfectly correlated, then the triangle collapses into a straight line as
the correlation, or the cosine of a zero angle, is equal to one. In this case
of perfect correlation, the volatility of the spread is determined entirely
by the spread between the two volatilities. If we keep the volatilities of
each leg unchanged, but instead widen the angle, which means
reducing correlation, then the length of the opposite side, which
corresponds to the volatility of the spread, increases. The lower the
correlation, the more volatile the spread, and, therefore, the more
expensive the option on the spread is.
This triangular relationship sets up an important boundary not only
for volatilities, but for option prices as well. It follows, for example,
from the Bachelier formula that prices of ATM options with the same
maturity are directly proportional to their dollar volatilities, and,
therefore, the same inequality must also hold for ATM call prices:
(13.5)
In other words, an option on the spread CSP cannot be worth less
than the spread of two options with the same moneyness and maturity.
If market prices for three options violate this inequality, then it is an
indication of a potential arbitrage opportunity.
Let us illustrate this with a simple numerical example. Let
X(t) = 60 and Y(t) = 55 represent diesel and crude oil prices in dollars
per barrel ($/bbl), as observed today at time t. Assume that their
implied ATM Black volatilities for options with one-year maturity are,
respectively, 30% and 29%. Then call options prices, obtained from the
standard Black formula, are $7.18/bbl for diesel and $6.36/bbl for
crude oil. By selling the diesel option and buying the crude oil option,
the dealer collects $0.82/bbl in exchange for taking unforeseen diesel-
specific risks. It is quite possible that the net hedging cost could exceed
the premium collected if diesel prices happen to be highly volatile. To
mitigate this risk, a prudent dealer would attempt to buy the same
maturity ATM option on the crack spread from a refinery.
If the dealer does buy the spread option, then the overall portfolio
consists of three options, one short option and two long:
Scenario 1 Both diesel and oil markets rally, and diesel outperforms
crude oil, so that X(T) = 70, Y(T) = 60. Then a ten-dollar loss from the
short diesel call is precisely offset by the sum of a five-dollar gain from
owning the oil call, and another five-dollar gain from owning an ATM
crack call struck at $5/bbl:
Scenario 2 Both diesel and oil markets rally, and the oil price
outperforms diesel, so that X(T) = 70, Y(T) = 67. Here, the gain from
owning the oil option exceeds the ten-dollar liability on the diesel call,
while the crack call expires worthless:
Only in the first scenario, which is, admittedly, more likely than the
other three, the dealer’s payoff at the options expiration is precisely
equal to zero. This scenario corresponds to the case of equal dollar
variance and perfect correlation between diesel and crude oil. In the
other three scenarios, which are less likely, the dealer receives
additional benefits from options payoffs. In the second scenario, the
positive payoff comes from a favorable move in relative realized
volatilities when crude oil moves more than diesel. In the last two
scenarios, the profit is driven by decorrelation of the two commodities.
This portfolio provides an example of an arbitrage boundary which
is independent of any models or assumptions. If one can execute a
triangular trade in the market at net zero cost, it would genuinely
represent a free option. The only thing to worry about would be the
creditworthiness of the option sellers. The dealer can only profit from
an OTC trade if neither the refiner, nor the producer, default on their
short option obligations. A popular joke, or an interview question for a
trading job, is “how much to pay for an OTC option to a counterparty
that has an option not to pay you back? The answer is, of course, zero.
Some energy volatility dealers, however, learned painful lessons by
buying “cheap” options from asset owners with poor credit, who ended
up defaulting when their derivatives obligations became particularly
large. In one case, the dealer even ended up owning an entire power
plant which was placed as collateral against the financial derivative sold
by an asset owner. Unfortunately for the dealer, the value of the plant
covered only a fraction of the payoff due on the derivatives trade.
A perfect triangular arbitrage is unlikely to be found in competitive
markets. However, the pricing boundary created by options on two
correlated assets and an option on the spread forms the backbone of
spread option and correlation trading. We now discuss how this
relationship between the volatility of the spread and the correlation is
captured in two alternative modeling paradigms.
(13.6)
Assuming that both futures prices and their implied volatilities are
observable in the market, this transformation allows the trader to
calculate the volatility of the spread given some assumption about the
correlation. Alternatively, if the price for the spread option is observed
in the market, then the trader can back out the implied correlation as
follows:
(13.7)
where, for simplicity, we kept the same letter to denote implied and
local geometric volatility.
When the spread is modeled as an independent asset, its delta and
other Greeks are reported with respect to the spread itself. In other
words, the delta with respect to the first leg of the spread is the
negative of the delta with respect to the second leg. This makes perfect
sense for closely related spreads between assets with similar
volatilities, as hedging is typically done by transacting directly in the
spread contract. For example, an option on the WTI-Brent spread will
always be hedged by trading in the spread itself. However, when one
asset is significantly more volatile than the other, the trader will most
likely be looking to hedge the risk of the more volatile asset differently
by delta hedging it outright rather than as a spread to something that is
moving a lot less. The problem becomes more evident when the two
legs of the spread trade in different units. This is more common for
cross-product spreads that arise in the natural gas and power markets
where the two legs of the spread are more disjoint.
An alternative valuation method attempts to mold spread options
into the standard lognormal BSM pricing framework. Its main idea is
based on the observation that the ratio of lognormal variables is also
lognormal. Furthermore, the variable
Both the spread option price CSP and the underlying variable x are
scaled by the same quantity F2 + K. They are effectively expressed in
terms of the new scaled numeraire. The payoff of the scaled spread
option price is then the same as the payoff of the regular call option
written on the variable x with the strike price equal to one.
Unfortunately, the spread option admits an analytic solution in a
two-dimensional lognormal setting only in a very special case when the
strike price K = 0 and the variable x becomes lognormal. Such a solution
is known as the Margrabe formula, which was initially referred to as an
option to exchange one asset for another. For K ≠ 0, there exist several
convenient approximations, among which the following Kirk
approximation is the most widely used6:
(13.8)
where
Figure 13.4 shows how the value of an ATM spread option changes
for a wide range of correlations.
Fig. 13.4 The price sensitivity of a one-year ATM spread option to correlation for F1 = F2 = 60,
and σ1, G = σ2, G = 0.30
Higher correlation means lower spread option value. It is easy to
see that the relationship becomes highly nonlinear for extremely high
levels of correlation. In fact, since
Fig. 13.5 Realized correlations for many petroleum spreads are extremely high and unstable
(6-month rolling correlations)
References
Carmona, R., & Durrleman, V. (2003). Pricing and hedging spread options. SIAM Review, 45(4),
627–685.
[MathSciNet][Crossref][zbMATH]
Considine, J., Galkin, P., & Aldayel, A. (2022). Inventories and the term structure of oil prices: A
complex relationship. Resources Policy, 77, 1–18.
[Crossref]
Dempster, M. A. H., & Hong, S. S. G. (2002). Spread option valuation and the fast Fourier
transform. In H. Geman, D. Madan, S. R. Pliska, & T. Vorst (Eds.), Mathematical finance, Bachelier
congress (Vol. 1, pp. 203–220). Springer.
[Crossref]
Eydeland, A., & Wolyniec, K. (2003). Energy and power risk management: New developments in
modeling, pricing, and hedging. Wiley.
Johnson, O. (2022). 40 classic crude oil trades: Real-life examples of innovative trading.
Routledge.
Kirk, E. (1995). Correlation in the energy markets. In Managing energy price risk (pp. 71–78).
Risk Publications.
Margrabe, W. (1978). The value of an option to exchange one asset for another. The Journal of
Finance, 33(1), 177–186.
[Crossref]
Shimko, D. C. (1994). Options on futures spreads: Hedging, speculation, and valuation. The
Journal of Futures Markets, 14(2), 183–213.
[Crossref]
Swindle, G. (2014). Valuation and risk management in energy markets. Cambridge University
Press.
[Crossref]
Venkatramanan, A., & Alexander, C. (2011). Closed form approximations for spread options.
Applied Mathematical Finance, 18(5), 447–472.
[MathSciNet][Crossref][zbMATH]
Footnotes
1 The story of how the strategy was originated and developed is described in Johnson (2022).
2 In this chapter, we ignore blending optionality of the storage asset, which allows the owner
to mix different grades of crude oil and to sell the blend at a premium by customizing it to the
rigorous specifications of a refinery.
3 Likewise, the cost of freight determines the strike price of the locational spread option on the
difference between oil prices in two different regions. Considine et al. (2022) relates the value
of such an option to inventories.
4 Formally, this connection follows from reduced-form mean-reverting models, the reference
to which are provided in Chap. 3. Less formally and more intuitively, one can see that faster
mean-reversion precludes the spread from deviating too far from its mean, which generally
results in volatility that declines with time to maturity.
5 For more background on spread options that arise in natural gas and power markets, see, for
example, Eydeland and Wolyniec (2003) and Swindle (2014).
6 Margrabe (1978) derived two-dimensional partial differential equation for the price of a
spread option and solved it for K = 0. Kirk (1995) proposed the extension for K ≠ 0 but did not
publish its derivation. This formula can be formally derived by applying the method of
perturbation introduced in Appendix C. The technical details, however, are rather cumbersome,
and given the limited usage of correlation-based pricing formulas in the oil market, we chose
not to present them.
7 Similar to (A.10), the solution to the partial differential equation for the price of a spread
option can be represented as a double integral of the option payoff multiplied by the risk
neutral joint probability density for F1 and F2. Advanced methods for pricing spread options in
two-dimensional lognormal setting tend to focus on developing efficient integration techniques
and more advanced analytic approximations. See, for example, Shimko (1994), Dempster and
Hong (2002), Carmona and Durrleman (2003), and Venkatramanan and Alexander (2011). For
petroleum markets, a simple approximation (13.8) is generally sufficient.
8 The day when oil prices went negative is removed as neither percentage returns nor
standard correlation metrics can even be computed.
9 Technically, the volatility of the first leg is also not observable as a vanilla option expires
three days prior to futures but the corresponding CSO expires one day before futures. Traders
often make an additional ad hoc adjustment to the implied volatility of a vanilla option to
incorporate this effect.
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023
I. Bouchouev, Virtual Barrels, Springer Texts in Business and Economics
https://doi.org/10.1007/978-3-031-36151-7_14
14. Epilogue
Ilia Bouchouev1
(1) Pentathlon Investments, LLC, Westfield, NJ, USA
(A.2)
(A.3)
(A.5)
(A.6)
(A.8)
This equation is with respect to x and it moves backward in time for
t < T. The terminal boundary condition at t = T is given by the Dirac
delta function centered at x = xT:
(A.9)
In applied mathematics, Eqs. (A.6) and (A.8) are known,
respectively, as the forward and backward parabolic differential
equations that describe the diffusion of heat in a medium, and the
function p(x, t, xT, T) is called the fundamental solution to these
equations.
In this book, we are mainly interested in the backward Eq. (A.8),
which applies to the pricing of a financial derivative at time t if its
payoff is specified at a later time T > t. The function p(x, t, xT, T)
represents only one particular solution to the Eq. (A.8) with a very
special boundary condition given by the Dirac delta function. Any other
solution G(x, t) to (A.8) with a different boundary condition specified at
time T by
(A.10)
(A.12)
(A.14)
In other words, the shifted variable y(t) follows the ABM process,
where for brevity, we no longer explicitly show its dependence on x.
The drift of this process is zero, but its variance is reduced by the
time-dependent exponential factor. This factor can be eliminated by
switching to the new integrated-time variable defined by
(A.16)
(A.17)
The normal delta is the first derivative of the Bachelier formula with
respect to the futures price:
(B.3)
(B.4)
(B.5)
(B.6)
Similarly, for GBM process (A.13) the pricing formula is given by the
integral of the payoff function with lognormal probability density with
μG = 0
where
(B.8)
(B.9)
(B.10)
(B.11)
is satisfied. Partial derivatives for put options can be obtained from the
Greeks for the call option by differentiating the put-call parity
relationship (8.12).
Appendix C: The Perturbation Method and the
Quadratic Normal Model
Let us assume that the local volatility function is defined by a small
perturbation ε(F) of the constant volatility σA
(C.1)
Then the partial differential Eq. (8.8) for the option price is
(C.2)
The sum of the two terms grouped in the first parenthesis is zero
because CBC solves (C.2) with constant dollar volatility σA and ε = 0. The
terms grouped in the last parenthesis are of a higher order with respect
to ε. The method of perturbation or linearization assumes that for small
perturbation ε, these terms can be omitted.
For small ε, we are left with only the terms grouped in the middle
parenthesis, which leads to the following equation for U(F, t):
(C.5)
and substitute them into (C.4). After cancelling the common multiplier
n and multiplying each term by , we arrive at the following
equation for the function p(F, τ)
(C.6)
and representation (C.5) leads to the solution for the skew correction
function
(C.8)
Also, since both C(F, T) and CBC(F, T) at expiration are equal to the
option payoff, max(0, F − K), the skew correction function at time t = T
must be equal to zero. This boundary condition for the skew correction
function at τ = 0
is clearly satisfied by the formula (C.8).
If necessary, this method can be easily extended to a higher-order
polynomial volatility perturbation function, in which case p(F, τ) must
also be taken to be a polynomial of the same order with coefficients
determined by matching corresponding terms of the Eq. (C.6).
(D.1)
(D.2)
where v is constant.
Let represent the call option price using this new time
variable
We calculate the derivative with respect to time using the chain rule
The substitution of this partial derivative into Eq. (D.1) with time
variable volatility σG(t, T) replaces it with an identical equation but with
constant volatility v
The price of the call option is simply given by the Black formula
with respect to time and with constant volatility v:
It follows from (D.2) that the total variance in the Black formula in
the new variables is equal to the integrated local variance in the
original variables
(E.2)
(F.1)
We next consider a more practical case, where the local volatility
only depends on the spatial variable and option prices are available for
the continuum of strikes but only for a fixed maturity T, which is the
case for the oil market.
Let τ = T − t represent the time remaining to maturity. Then the
Dupire Eq. (F.1) with time-independent volatility is
(F.2)
Glossary
Actuarial Valuation A method of pricing options based on their average
historical payoffs.
Arbitrage A strategy that involves buying and selling similar contracts
to capture discrepancies in their prices.
Average Price Option (APO) An option that settles based on the average
price of underlying futures contracts over a given period.
Backwardation A situation where the futures curves declines with
increasing time to maturity.
Barrel Counting Fundamental analysis of supply and demand.
Basis The spread between the price of a physical or a less liquid
financial contract and the benchmark futures price.
Bid-Ask Spread The spread between the price to buy (bid) and the price
to sell (ask).
Black Volatility Implied volatility computed using the Black (1976)
option pricing formula.
Butterfly Spread A three-legged option trade that buys options with
strikes K1 and K2 and sells twice as many options with the strike price
set half-way between K1 and K2, all with the same expiration.
Calendar Spread Option (CSO) An option on the spread between two
futures on the same commodity with different maturities.
Calibration A process of fitting volatility models to market prices of
traded options.
Carry The spread between futures on the same commodity with
different maturities, which is used to measure the expected price roll
up (down) the curve.
Cash Price The price of an over-the-counter contract with physical
delivery.
Cash-Settled Option An option whose payoff at expiration is calculated
financially as the difference between the futures settlement price and
the strike price.
Commodity Currencies Currencies of commodity-exporting countries.
Commitments of Traders (CoT) Report A weekly report on futures and
option positions held by different categories of market participants.
Commodity Trading Advisors (CTAs) In general, anyone who advises on
commodity trading, but in practice, the term is often used to describe
hedge funds that trade based on quantitative algorithms.
Contango A situation when the futures curve increases with time
remaining to maturity.
Convenience Yield Consumption benefits accrued to the owner of a
physical commodity, but not to the owner of a futures contract.
Crack Spread The spread between the price of a refined product and
the price of crude oil.
Cushing A reference to the storage hub and inventories at the delivery
location for WTI futures contract.
Digital (Binary) Option An option with a discontinuous payoff that pays
a fixed amount or zero depending on the futures price.
Denomination Effect An impact on oil price, resulting from oil being
denominated in US dollars (USD).
Excess Return (ER) A return on buying and rolling futures.
Fading the Crowded Trade A contrarian position taken when hedge
funds’ futures holdings are perceived to be excessive.
Follow the Flow Strategy A strategy of mimicking futures position held
by hedge funds.
Fractionation Analysis Decomposition of profit-and-loss (P&L) of a
systematic strategy by the value of a certain explanatory variable.
Gamma Hedging The process of option dealers rebalancing their
directional price exposure caused by shifts in the futures price.
Grades Various types of crude oil that trade as a differential to main
futures benchmarks, such as WTI and Brent.
GSCI Rolls The process of rolling futures between the fifth and the ninth
business days of each month by major commodity indices, such as
Goldman Sachs Commodity Index (GSCI).
Hacienda Hedge A reference to a large-scale hedging program by the
Government of Mexico.
Hedging Pressure An imbalance between buyers and sellers of a
particular futures or options contract.
Implied Volatility An input into an option pricing model that makes the
model price of an option to match its market price.
Local Volatility A volatility function that depends on futures prices and
time and characterizes uncertainty of the random variable in a
stochastic diffusion process.
Market-Implied Probability Distribution A probability distribution
reconstructed from market prices of options with various strikes and
the fixed expiration.
Mean-Reversion A tendency of prices to revert to a long-term
equilibrium.
Naked Option An option position (typically a sale) which is not delta
hedged.
Normal Backwardation (Contango) A situation where the futures price
is below (above) the expected spot price. The term is equivalent to a
risk premium.
Normal (Dollar) Volatility Implied volatility computed using the
Bachelier option pricing formula.
Quantamentals A trading style that combines rule-based decision-
making with discretionary overlay.
Penultimate Expiration An expiration of derivatives contracts one day
prior to the expiration of the futures.
Physically-Settled Option An option, which when it is in-the-money, is
exercised into futures at expiration.
Positioning A reference to positions held by hedge funds and other
large market participants.
Producer Collar (Fence) A derivatives structure where a producer buys
a put option and finances it by selling a call option of the same value.
Producer Three-Way Collar A derivatives structure where a producer
buys a put spread and finances it by selling a call option of the same
value.
Prompt (Nearby) Futures A futures contract with the shortest time to
expiration.
Reaction Function A function that maps the strength of the systematic
signal to the position size.
Realized Volatility A statistical measure of volatility calculated as the
annualized standard deviation of percentage changes in prices over a
given period.
Regime Change A structural shift in dominant market-driving factors.
Risk Parity An asset allocation framework that weighs assets inversely
to their volatility and invests in commodity futures as an inflation
hedge.
Roll Return (RR) The difference between the excess return (ER) and the
spot return (SR).
Signal Blending A combination of multiple systematic trading signals.
Skew Delta A correction to the hedging delta that results from the
change in option’s vega caused by the move in futures.
Spot Return (SR) A return on a hypothetical investment in a spot
futures contract with no convenience benefits or storage costs.
Spread Option An option on the difference between the prices of two
futures.
Squeeze A large price move caused by futures traders exiting positions
near the expiration of the futures contract.
Sticky Moneyness A heuristic rule that assumes that implied volatilities
for all options remain unchanged for a given moneyness.
Sticky Strike A heuristic rule that assumes that implied volatilities for
all options remain unchanged for a given strike.
Stock-Out A situation of zero inventories in storage.
Swap Data Repository (SDR) A registered entity that collects and
disseminates information about over-the-counter transactions on a
nearly real-time basis.
Synthetic Spread Option The spread between an option on a refined
product and an option on crude oil with the same expiration.
Synthetic Storage A portfolio of calendar spread options (CSOs)
designed to mimic cash flows of physical storage.
Swaption An option to enter into a swap.
Tank-Tops A situation where inventories reach the maximum storage
capacity.
Trading-at-Settlement (TAS) A futures contract that allows parties to
trade during a trading day at a settlement price which will only be
determined later after the market closes.
Underlying Futures A futures contract on which the price of an option
contract depends.
Vanilla Options Exchange-traded options, excluding spread options and
APOs.
Volatility Risk Premium An investment return of the systematic strategy
of selling and delta hedging short term options.
Volatility Skew (Smile) A graph of implied volatilities for options versus
their strikes or moneyness for the same underlying futures contract.
Volatility Targeting A strategy that adjusts a notional position size
required to bring expected portfolio volatility to a given target.
Volatility Term Structure A graph of implied volatilities versus time to
maturity.
References
Acharya, V. V., Lochstoer, L. A., & Ramadorai, T. (2013). Limits to arbitrage and hedging:
Evidence from commodity markets. Journal of Financial Economics, 109(2), 441–465.
Alexander, C. (2008). Market risk analysis, Vol. III: Pricing, hedging and trading financial
instruments. Wiley.
Andersen, L. (2011). Option pricing with quadratic volatility: A revisit. Finance and Stochastics,
15(2), 191–219.
[MathSciNet][zbMATH]
Ashton, M., & Greer, R. (2008). History of commodities as the original real return asset class. In
Inflation risk and products (pp. 85–109). Risk Books.
Avellaneda, M., Friedman, C., Holmes, R., & Samperi, D. (1997). Calibrating volatility surfaces via
relative entropy minimization. Applied Mathematical Finance, 4(1), 37–64.
[zbMATH]
Bakshi, G., Gao, X., & Rossi, A. G. (2019). Understanding the sources of risk underlying the cross
section of commodity returns. Management Science, 65(2), 619–641.
Barone-Adesi, G., & Whaley, R. E. (1987). Efficient analytic approximation of American option
values. Journal of Finance, 42(2), 301–320.
Black, F. (1976). The pricing of commodity contracts. Journal of Financial Economics, 3(1/2),
167–179.
Black, F., & Scholes, M. (1973). The pricing of options and corporate liabilities. The Journal of
Political Economy, 81(3), 637–654.
[MathSciNet][zbMATH]
Blas, J. (2017, April 4). Uncovering the secret history of Wall Street’s largest oil trade,
Bloomberg.
Bodie, Z., & Rosansky, V. I. (1980, May–June). Risk and return in commodity futures. Financial
Analysts Journal, 36(3), 27–39.
Boons, M., & Prado, M. P. (2019). Basis-momentum. The Journal of Finance, 74(1), 239–279.
Bouchouev, I. (1998). Derivatives valuation for general diffusion processes, Proceedings of the
Annual Conference of the International Association of Financial Engineers, New York, USA, pp.
91–104.
Bouchouev, I. (2000a, July). Demystifying Asian options. Energy and Power Risk Management,
pp. 26–27.
Bouchouev, I. (2000b, August). Black-Scholes with a smile. Energy and Power Risk Management,
pp. 28–29.
Bouchouev, I. (2020a, April 30). Negative oil prices put spotlight on investors, Risk.net.
Bouchouev, I. (2020b). From risk bearing to propheteering. Quantitative Finance, 20(6), 887–
894.
[MathSciNet]
Bouchouev, I. (2022). The Strategic Petroleum Reserve strategies: Risk-free return or return-free
risk? The Oxford Institute for Energy Studies.
Bouchouev, I. (2023). Testimony to the US House Subcommittee on Economic Growth, Energy
Policy, and Regulatory Affairs of the Committee on Oversight and Accountability, March 8.
Reprinted in Commodity Insights Digest (2023), Vol. 1.
Bouchouev, I., & Isakov, V. (1997). The inverse problem of option pricing. Inverse Problems,
13(5), L11–L17.
[MathSciNet][zbMATH]
Bouchouev, I., & Isakov, V. (1999). Uniqueness, stability and numerical methods for the inverse
problem that arises in financial markets. Inverse Problems, 15(3), R95–R116.
[MathSciNet][zbMATH]
Bouchouev, I., Isakov, V., & Valdivia, N. (2002). Recovery of volatility coefficient by linearization.
Quantitative Finance, 2(4), 257–263.
[MathSciNet][zbMATH]
Bouchouev, I., & Johnson, B. (2022). The volatility risk premium in the oil market. Quantitative
Finance, 22(8), 1561–1578.
[MathSciNet][zbMATH]
Bouchouev, I., & Zuo, L. (2020, Winter). Oil risk premia under changing regimes. Global
Commodities Applied Research Digest, 5(2), 49–59.
Boyle, P. C. (1899, September 6). Testimony at the Hearing of the US Industrial Commission on
Trusts and Industrial Combinations.
Breeden, D. T., & Litzenberger, R. H. (1978). Prices of state contingent claims implicit in option
prices. Journal of Business, 51(4), 621–651.
Brennan, M. J. (1958). The supply of storage. The American Economic Review, 48(1), 50–72.
Brennan, M. J. (1991). The price of convenience and the valuation of commodity contingent
claims. In D. Lund & B. Oksendal (Eds.), Stochastic models and option values. North Holland.
Brennan, M. J., & Schwartz, E. S. (1985). Evaluating natural resource investments. Journal of
Business, 58(2), 135–157.
Bü yü kşahin, B., & Robe, M. A. (2014). Speculators, commodities and cross-market linkages.
Journal of International Money and Finance, 42, 48–70.
Carmona, R., & Durrleman, V. (2003). Pricing and hedging spread options. SIAM Review, 45(4),
627–685.
[MathSciNet][zbMATH]
Carmona, R., & Ludkovski, M. (2004). Spot convenience yield models for the energy markets.
Contemporary Mathematics, 351, 65–79.
[MathSciNet][zbMATH]
Carr, P., Fisher, T., & Ruf, J. (2013). Why are quadratic normal volatility models analytically
tractable? SIAM Journal on Financial Mathematics, 4(1), 185–202.
[MathSciNet][zbMATH]
Carr, P., & Madan, D. (2001). Determining volatility surfaces and option values from an implied
volatility smile. In M. Avellaneda (Ed.), Quantitative analysis in financial markets (Vol. II, pp.
163–191). World Scientific.
Casassus, J., & Collin-Dufresne, P. (2005). Stochastic convenience yield implied from commodity
futures and interest rates. The Journal of Finance, 60(5), 2283–2331.
Castagna, A., & Mercurio, F. (2007). The vanna-volga method for implied volatilities. Risk, 20(1),
106–111.
Cheng, I.-H., Kirilenko, A., & Xiong, W. (2015). Convective risk flows in commodity futures
markets. Review of Finance, 19(5), 1733–1781.
Chiarella, C., Craddock, M., & El-Hassan, N. (2003). An implementation of Bouchouev’s method
for short time calibration of option pricing models. Computational Economics, 22, 113–138.
[zbMATH]
Clewlow, L., & Strickland, C. (2000). Energy derivatives: Pricing and risk management. Lacima
Publications.
Considine, J., Galkin, P., & Aldayel, A. (2022). Inventories and the term structure of oil prices: A
complex relationship. Resources Policy, 77, 1–18.
Daskalaki, C., Kostakis, A., & Skiadopoulos, G. (2014). Are there common factors in individual
commodity futures returns? Journal of Banking and Finance, 40, 346–363.
Deaton, A., & Laroque, G. (1992). On the behavior of commodity prices. The Review of Economic
Studies, 59(1), 1–23.
[zbMATH]
Dempster, M. A. H., & Hong, S. S. G. (2002). Spread option valuation and the fast Fourier
transform. In H. Geman, D. Madan, S. R. Pliska, & T. Vorst (Eds.), Mathematical finance, Bachelier
congress (Vol. 1, pp. 203–220). Springer.
Dempster, M. A. H., Medova, E., & Tang, K. (2012). Determinants of oil futures prices and
convenience yields. Quantitative Finance, 12(12), 1795–1809.
[MathSciNet][zbMATH]
Dempster, M. A. H., & Richards, D. G. (2000). Pricing American options fitting the smile.
Mathematical Finance, 10(2), 157–177.
[MathSciNet][zbMATH]
Doran, J. S., & Ronn, E. I. (2006). The bias in Black-Scholes/Black implied volatility: An analysis
of equity and energy markets. Review of Derivatives Research, 8(3), 177–198.
[zbMATH]
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(12), 2541–2552.
Dupire, B. (1994). Pricing with a smile. Risk, 7(1), 18–20.
Dvir, E., & Rogoff, K. (2009). Three epochs of oil, NBER Working Paper, 14927.
Ederington, L. H., Fernando, C. S., Holland, K. V., Lee, T. K., & Linn, S. C. (2021). The dynamics of
arbitrage. Journal of Financial and Quantitative Analysis, 56(4), 1350–1380.
Ellwanger, R. (2017). On the tail risk premium in the oil market, Bank of Canada Working Paper,
46.
Erb, C. B., & Harvey, C. R. (2006). The strategic and tactical value of commodity futures.
Financial Analysts Journal, 62(2), 69–97.
Eydeland, A., & Wolyniec, K. (2003). Energy and power risk management: New developments in
modeling, pricing, and hedging. Wiley.
Fama, E. F., & French, K. R. (1987). Commodity futures prices: Some evidence on forecast power,
premiums, and the theory of storage. Journal of Business, 60(1), 55–73.
Fattouh, B. (2011). An anatomy of the crude oil pricing system. Oxford Institute for Energy
Studies, Working Paper, 40.
Fattouh, B., & Mahadeva, L. (2014). Causes and implications of shifts in financial participation in
commodity markets. The Journal of Futures Market, 34(8), 757–787.
Fernandez-Perez, A., Frijns, B., Fuertes, A.-M., & Miffre, J. (2018). The skewness of commodity
futures returns. The Journal of Banking and Finance, 86, 143–158.
Fernandez-Perez, A., Fuertes, A.-M., & Miffre, J. (2021, Summer). On the negative pricing of WTI
crude oil futures. Global Commodities Applied Research Digest, 6(1), 36–43.
Fisher, I. (1896). Appreciation and interest. Publications of the American Economic Association,
11(4), 331–442.
Geman, H. (2005). Commodities and commodity derivatives: Modeling and pricing for
agriculturals, metals and energy. Wiley.
Gibson, R., & Schwartz, E. S. (1990). Stochastic convenience yield and the pricing of oil
contingent claims. The Journal of Finance, 45(3), 959–976.
Gorton, G. B., Hayashi, F., & Rouwenhorst, K. G. (2012). The fundamentals of commodity futures
returns. Review of Finance, 17(1), 35–105.
Gorton, G., & Rouwenhorst, K. G. (2006). Facts and fantasies about commodity futures. Financial
Analysts Journal, 62(2), 47–68.
Greer, R. J. (1978, Summer). Conservative commodities: A key inflation hedge. Journal of
Portfolio Management, 4(4), 26–29.
Grunspan, C. (2011). A note on the equivalence between the normal and the lognormal implied
volatility: A model free approach, SSRN.
Gustafson, R. L. (1958). Carryover levels for grains, U.S. Department of Agriculture, Technical
Bulletin, 1178.
Hamilton, J. D. (1983). Oil and the macroeconomy since World War II. Journal of Political
Economy, 91(2), 228–248.
Hamilton, J. D. (2009). Understanding crude oil prices. The Energy Journal, 30(2), 179–206.
Hamilton, J. D., & Wu, J. C. (2014). Risk premia in crude oil futures prices. Journal of
International Money and Finance, 42, 9–37.
Hicks, J. R. (1939). Value and capital: An inquiry into some fundamental principles of economic
theory. Oxford University Press.
Hirshleifer, D. (1988). Residual risk, trading costs, and commodity futures risk premia. The
Review of Financial Studies, 1(2), 173–193.
[MathSciNet]
Hull, J. C. (2018). Options, futures, and other derivatives (10th ed.). Pearson.
[zbMATH]
Imsirovic, A. (2021). Trading and price discovery for crude oils: Growth and development of
international oil markets. Palgrave Macmillan.
Ingersoll, J. E., Jr. (1997). Valuing foreign exchange rate derivatives with a bounded exchange
process. Review of Derivatives Research, 1, 159–181.
[zbMATH]
Interim Stuff Report. (2020). Trading in NYMEX WTI crude oil futures contract leading up to, on,
and around April 20, 2020, Commodity Futures Trading Commission, November 23.
Jacobs, K., & Li, B. (2023). Option returns, risk premiums, and demand pressure in energy
markets. Journal of Banking and Finance, 146, 1–26.
Johnson, O. (2022). 40 classic crude oil trades: Real-life examples of innovative trading.
Routledge.
Kaldor, N. (1939). Speculation and economic stability. The Review of Economic Studies, 7(1), 1–
27.
Kang, S. B., & Pan, X. (2015). Commodity variance risk premia and expected futures returns:
Evidence from the crude oil market, SSRN.
Kang, W., Rouwenhorst, K. G., & Tang, K. (2020). A tale of two premiums: The role of hedgers
and speculators in commodity futures markets. The Journal of Finance, 75(1), 377–417.
Kemna, A. G. Z., & Vorst, A. C. F. (1990). A pricing method for options based on average asset
values. Journal of Banking and Finance, 14(1), 113–129.
Keynes, J. M. (1936). The general theory of employment, interest, and money. Macmillan.
Kilian, L. (2009). Not all oil price shocks are alike: Disentangling demand and supply shocks in
the crude oil market. American Economic Review, 99(3), 1053–1069.
Kilian, L. (2020). Understanding the estimation of oil demand and oil supply elasticities, Federal
Reserve Bank of Dallas Working Paper, 2027.
Kilian, L., & Murphy, D. P. (2014). The role of inventories and speculative trading in the global
market for crude oil. Journal of Applied Econometrics, 29(3), 454–478.
[MathSciNet]
Kilian, L., & Vigfussion, R. J. (2017). The role of oil price shocks in causing U.S. recessions.
Journal of Money, Credit and Banking, 49 (8), 1747–1776.
Kirk, E. (1995). Correlation in the energy markets. In Managing energy price risk (pp. 71–78).
Risk Publications.
Koijen, R. S. J., Moskovitz, T. J., Pedersen, L. H., & Vrugt, E. B. (2018). Carry. Journal of Financial
Economics, 127, 197–225.
Levy, E. (1992). Pricing European average rate currency options. Journal of International Money
and Finance, 11(5), 474–491.
Lipton, A. (2001). Mathematical methods for foreign exchange: A financial engineer’s approach.
World Scientific.
[zbMATH]
Lipton, A., & Sepp, A. (2011, October). Filling the gaps. Risk, 24(10), 78–83.
Lo, A. W. (2002). The statistics of Sharpe ratios. Financial Analysts Journal, 58(4), 36–52.
Ma, L. (2022). Negative WTI price: What really happened and what can we learn? The Journal of
Derivatives, 29(3), 9–29.
[MathSciNet]
Margrabe, W. (1978). The value of an option to exchange one asset for another. The Journal of
Finance, 33(1), 177–186.
Merton, R. C. (1973). Theory of rational option pricing. The Bell Journal of Economics and
Management Science, 4(1), 141–183.
[MathSciNet][zbMATH]
Miltersen, K. R. (2003). Commodity price modelling that matches current observables: A new
approach. Quantitative Finance, 3(1), 51–58.
[MathSciNet][zbMATH]
Naldi, N. (2015). Sraffa and Keynes on the concept of commodity rates of interest. Contributions
to Political Economy, 34(1), 17–30.
Neville, H., Draaisma, T., Funnell, B., Harvey, C. R., & Van Hemert, O. (2021). The best strategies
for inflationary times, SSRN.
Pilipović, D. (2007). Energy risk: Valuing and managing energy derivatives. McGraw-Hill.
Prokopczuk, M., Symeonidis, L., & Simen, C. W. (2017). Variance risk in commodity markets.
Journal of Banking and Finance, 81, 136–149.
Rebonato, R. (2004). Volatility and correlation: The perfect hedger and the fox. Wiley.
Routledge, B. R., Seppi, D. J., & Pratt, C. S. (2000). Equilibrium forward curves for commodities.
The Journal of Finance, 55(3), 1297–1338.
Samuelson, P. A. (1965). Proof that properly anticipated prices fluctuate randomly. Industrial
Management Review, 6(2), 41–49.
Schachermayer, W., & Teichmann, J. (2008). How close are the option pricing formulas of
Bachelier and Black-Merton-Scholes? Mathematical Finance, 18(1), 155–170.
[MathSciNet][zbMATH]
Schwartz, E. S. (1997). The stochastic behavior of commodity prices: Implications for valuation
and hedging. The Journal of Finance, 52(3), 923–973.
Shimko, D. C. (1994). Options on futures spreads: Hedging, speculation, and valuation. The
Journal of Futures Markets, 14(2), 183–213.
Smiley, A. W. (1907). A few scraps: Oily and otherwise. The Derrick Publishing Company.
Sraffa, P. (1932). Dr. Hayek on money and capital. The Economic Journal, 42 (165), 42–53.
Stoll, H. R. (1979). Commodity futures and spot price determination and hedging in capital
market equilibrium. The Journal of Financial and Quantitative Analysis, 14(4), 873–894.
Swindle, G. (2014). Valuation and risk management in energy markets. Cambridge University
Press.
Szymanowska, M., De Roon, F., Nijman, T., & Van Den Goorbergh, R. (2014). An anatomy of
commodity futures risk premia. The Journal of Finance, 69(1), 453–482.
Tang, K., & Xiong, W. (2012). Index investment and the financialization of commodities.
Financial Analysts Journal, 68(6), 54–74.
[MathSciNet]
Till, H. (2022, Winter). Commodities, crude oil, and diversified portfolios. Global Commodities
Applied Research Digest, 7(2), 65–74.
Till, H., & Eagleeye, J. (Eds.). (2007). Intelligent commodity investing. Risk Books.
Trolle, A. B., & Schwartz, E. S. (2010, Spring). Variance risk premia in energy commodities. The
Journal of Derivatives, 17(3), 15–32.
Turnbull, S. M., & Wakeman, L. M. (1991). A quick algorithm for pricing European average price
options. Journal of Financial and Quantitative Analysis, 26(3), 377–389.
Venkatramanan, A., & Alexander, C. (2011). Closed form approximations for spread options.
Applied Mathematical Finance, 18(5), 447–472.
[MathSciNet][zbMATH]
Weymar, F. H. (1965). The dynamics of the world cocoa market. Ph.D. Thesis, Massachusetts
Institute of Technology.
Whiteshot, C. A. (1905). The oil-well driller: A history of the world’s greatest enterprise, the oil
industry. Acme Publishing Company.
Williams, J. C., & Wright, B. D. (1991). Storage and commodity markets. Cambridge University
Press.
Wilmott, P., Dewynne, J., & Howison, S. (1993). Option pricing: Mathematical models and
computation. Oxford Financial Press.
[zbMATH]
Working, H. (1948). Theory of the inverse carrying charge in futures markets. Journal of Farm
Economics, 30(1), 1–28.
Working, H. (1949). The theory of price of storage. The American Economic Review, 39(6), 1254–
1262.
Zü hlsdorff, C. (2001). The pricing of derivatives on assets with quadratic volatility. Applied
Mathematics Finance, 8(4), 235–262.
[zbMATH]
Index
A
Actuarial valuation 188, 331
Agricultural markets 30, 56, 64, 65, 129, 155, 308
Airline hedging 65, 234
All Weather strategy 74
Aluminum 308
American options 174
Arbitrage
model 214, 261
paper 118, 120, 121
statistical 5, 109, 122–124, 139, 148, 306
triangular correlation 7, 281, 287–291
volatility 230, 257
Aristotle 64, 161
Arithmetic Brownian motion (ABM) 40, 165, 171, 238, 293, 311
Asian options 234
At-the-money (ATM) options 162, 163, 172, 183, 188–190, 194, 199,
200, 203–205, 207, 214–225, 289–291, 298, 300
Autocorrelation 151
Availability 34–37
Average price options (APOs) 7, 233–237, 241–247, 325–326, 331
B
Bachelier formula 6, 171, 222, 227, 295, 315
Bachelier, L. 6, 162
Backtesting 85
Backwardation 15–17, 57, 58, 98, 103, 110–113, 331
Bank of China 52
Barone-Adesi, G. 174
Barrel counting 117, 331
Basis risk 62
Basis spread 2, 126, 292, 331
Batteries 282, 305
Beta hedge ratio 62, 77, 144, 146
Biodiesel 308
Black, F. 164, 169
Black formula 173, 222, 316
Black-Scholes-Merton (BSM) model 6, 161, 163, 169, 171, 193, 273
Black volatility 181, 192, 245, 331
Bootstrapping 7, 257, 261–268
Breakout strategy 85
Brent 3, 117, 118, 123–125, 130, 132, 145, 174, 236, 287, 293, 295,
306
Bridgewater Associates 74
Brownian motion 40, 162
Bureau of Labor Statistics (BLS) 149
Butane 127
Butterfly spread 271, 272, 331
C
Calendar spread options (CSOs) 7, 281–286, 293, 299–301, 331
Calibration 7, 257, 259–279
Call options 162
Call spreads 270, 271, 274
Canadian dollar (CAD) 143–145, 156, 157
Capital Asset Pricing Model (CAPM) 69, 74
Carry 5, 83, 86, 97–101, 103–106, 110, 111, 114, 118, 128, 331
Carry-momentum 101, 102, 106
Carry trade 5, 12, 19–21, 30, 35, 36, 42, 43
Causality 132, 137, 139, 141, 143, 148, 305
Chapman-Kolmogorov equation 163
Chicago Mercantile Exchange (CME) 4, 125, 130, 131
Chilean peso (CLP) 145
Coal 127, 306
Coefficient of instability 163
Cointegration 122, 123, 139, 143–146, 148, 156, 157
Collateral 24, 65, 67, 170, 174, 211, 234, 255, 291, 307
Commitments of Traders (CoT) report 129, 331
Commodities Corp. 84
Commodity currencies 143, 145, 148, 331
Commodity Futures Trading Commission (CFTC) 131
Commodity indices 67, 68
Commodity own rate of interest 5, 14, 22, 23, 26
Commodity terms of trade (CToT) 139–146
Commodity trading advisors (CTAs) 5, 84, 332
Concave backwardation 112, 113
Concave contango 103, 104, 110–113
Consumer hedging 57, 62–64, 77, 130, 234
Consumer Price Index (CPI) 71, 73, 149–152, 154, 157
Contango 15–17, 21, 26, 58, 75, 76, 98, 103, 107, 110–116, 282, 286,
332
Convenience yield 5, 11, 19, 22, 23, 25, 31, 32, 41, 76, 97, 98, 100, 103,
110, 117, 332
Convex backwardation 103, 110–112
Convex contango 112
Convexity 103, 110, 162, 172, 186, 191
Cootner, P. 84
Copper 11, 145, 308
Corn 13, 167, 308
Correlation
frown 297, 301
implied 295, 297, 301
realized 299
Costless collars 211, 212, 229, 276
Crack options 287, 290–292
Crack spreads 287, 292, 332
Credit risks 212, 214, 234, 255, 291
Crisis alpha 85
Cross-sectional momentum 87
Cushing 46, 114–117, 119, 124, 125, 127, 332
D
Dalio, R. 74
Dated Brent 125, 236
Delta 169, 172, 183, 184, 189–196, 202–206, 208, 216, 218–220, 225,
226, 229, 259, 271, 279, 295, 315–317
Delta hedging 6, 169, 183, 187, 189–196, 202–206, 218, 259, 273, 281,
284–288, 294, 296, 305
Denomination effect 140, 143, 332
Derivatives market 1, 3, 4
Diesel 131, 149, 152, 154, 288, 290–292, 294
Diffusion process 6, 7, 161–171, 227, 228, 257–259, 276–278, 309, 328
Digital options 270, 271, 332
Dirac delta function 41, 43, 44, 268–272, 309, 310, 328
Direct problems 257
Disaggregated reports 129
Domestic sweet oil (DSW) 125
Dubai oil 125
Dump men 20
Dupire equation 273–275, 327
Dynamic systems 1, 3, 29, 30, 38, 105, 137–139, 148, 305
E
Early exercise premium 174
Early expiry options (EEOs) 233, 247–252, 300
Einstein, A. 163
Emerging markets 138
Emission credits 127, 306
Energy Information Administration (EIA) 4, 47
Energy Policy and Conservation Act (EPCA) 16
Energy transition 282, 305–308
Ethane 127
Ethanol 308
Euclidean distance 134
Euro (EUR) 143
Eurobob 126
European options 174
Excess return (ER) 24, 25, 65, 66, 332
Exchange-traded options 174, 233
Expandable options 255
Extendable options 254
F
Fading extreme positioning 133
Fading the crowded trade 133, 332
Fat tails 167, 182, 209, 225, 226, 228, 293, 307
Federal Reserve Economic Data (FRED) 4
Feedback loops 1, 3, 5, 29, 30, 35, 52, 137, 138, 305
Fence 211, 333
Financialization 5, 55, 64–68, 74, 78, 113, 114, 141, 206, 221, 223
Finite difference methods 171, 227
Fisher, I. 5, 12, 97
Fisher inflation law 5, 11, 12, 16
Floating storage 21, 50
Fokker-Planck equation 40, 163, 273, 310, 327
Follow the flow strategy 132, 332
Fourier equation 163
Fractionation analysis 5, 113, 114, 116, 122, 128, 332
Freight 120, 121, 282, 293
Fuel oil 126
Fundamental solution 268, 310
Futures margin requirements 66
G
Gamma 6, 172, 187, 191–196, 199, 200, 207, 215, 218, 225, 228, 307,
316, 317, 320, 332
Gasoil 126, 131, 294
Gasoline 64, 68, 71, 89, 126–128, 137, 148–154, 292, 294
Geometric Brownian motion (GBM) 166, 311
Gold 11–13, 73, 140, 167, 308
Goldman Sachs Commodity Index (GSCI) 67, 332
H
Hacienda hedge 233–237, 332
Harmonic mean 64
Hayek, F. 14
Heat equation 161, 171, 258, 310, 328
Heat rate 294
Heaviside function 269
Hedging
airline 65, 234
consumer 57, 62–64, 77, 130, 234
inflation 76
Mexico sovereign 7, 210, 235–237
pressure 4, 5, 55–64, 76–78, 112, 332
producer 57, 60–67, 209–214, 234
Hicks, J.R. 57
High-Sulfur Fuel Oil (HSFO) 236
I
Ill-posed problems 260
Implied Black volatility (IBV) 181, 221–225
Implied correlation 295, 297, 301
Implied normal volatility (INV) 181, 222–224
Implied volatility 6, 161, 164, 180–186, 215–226, 239–241, 261–268,
276, 277, 332
Impulse function 269, 309
Incomplete markets 257, 267
Inconvenience yield 76
Inflation 11–14, 16, 55, 69–78, 137–139, 148–154, 308
Inflation base effect 150
Inflation breakeven rate 72
Inflation hedging 76
Inflation pass-through 71, 138, 141, 148–154
Inflation swaps 139, 148–154
Inflection point 105, 106, 116, 133, 134, 139
Information ratio 91
Intercontinental Exchange (ICE) 4, 125, 130
In-the-money (ITM) options 162
Inventories 29–53, 83, 87, 98, 100, 102, 103, 114–117
Inverse demand function 33–36, 38
Inverse diffusion problem 327–328
Inverse problem of option pricing 7, 257–260
Inverse problems 258–260
Itô’s lemma 168, 309
J
Jet fuel 64, 126, 288
K
Kaldor, N. 22
Keynes, J.M. 11, 14, 56–60
K-factor 236
Kirk approximation 296
Kolmogorov equations 40, 274, 310
Kurtosis 6, 190, 214, 276
L
Law of cosines 289
Law of radiation of probability 163
Linearization methods 227, 319
Liquefied natural gas (LNG) 127, 145, 306
Liquidity preference theory of money 23
Local volatility 6, 161, 164, 167, 171, 172, 175, 177, 181, 185, 226–
230, 237–255, 261, 263–267, 273–279, 285, 292–294, 300, 301, 309,
319, 326, 327, 332
Lognormal distribution 166, 167
Louisiana Light Sweet (LLS) 236
M
Macro fair-value model 155–158
Managed money (MM) 131
Margrabe formula 296, 297
Market-implied diffusion 7
Market-implied probability distribution 7, 257, 268–273, 332
Mars oil 125
Maya oil 235
Mean-reversion 39, 40, 44, 45, 83, 101–104, 107, 109, 118, 120–123,
127, 132, 138, 247, 254, 293, 312, 332
Medium of exchange channel 140
Mental models 1, 4, 5
Merton, R.C. 164, 169, 238
Metals markets 145, 307
Model arbitrage 214, 261
Model of the squeeze 5, 38–46, 305
Momentum 5, 83–97, 132, 133, 138
Momentum smile 95
Monte Carlo simulation 168, 175, 255, 259, 276, 278
Multi-factor models 252–255
N
Naphtha 127
National Bureau of Economic Research (NBER) 69
Natural gas 68, 99, 127, 131, 148, 168, 219, 239, 254, 265, 294, 296,
306
Natural gas liquids (NGLs) 123, 127
Negative oil prices 5, 49–53
Normal backwardation 5, 55–60, 63–68, 74, 75, 85, 206, 332
Normal contango 5, 55, 75–78, 113, 332
Normal distribution 40, 164–167, 171, 225, 311
Normal volatility 165
Norwegian krone (NOK) 145
Numeraire effect 140
O
Oil
busts 177
grades 124, 332
heavy 125
light 125
loans 16–18
own rate of interest 15–19
sour 125
swap 249, 250
sweet 125
Open interest 86, 124, 233, 302
Options
American 174
Asian 234
at-the-money (ATM) 162, 163, 172, 183, 188–190, 194, 199, 200,
203–205, 207, 214–225, 289–291, 298, 300
average price (APOs) 7, 233–237, 241–247, 325–326, 331
calendar spread (CSOs) 7, 281–286, 293, 299–301, 331
call 162
crack 287, 290–292
digital 270, 271, 332
early expiry (EEOs) 233, 247–252, 300
European 174
exchange-traded 174, 233
expandable 255
extendable 254
Greeks 172
as insurance 187–191, 197, 198
in-the-money (ITM) 162
moneyness 172, 183, 315, 316
out-of-the-money (OTM) 162
over-the-counter (OTC) 7, 233, 234, 237, 247, 249, 252, 254, 255,
291
put 162
real 2–3, 6, 109, 118, 123, 125, 210–213, 282, 287, 305, 307
simple 163, 172
spread 2, 118, 281–302, 333
synthetic spread 288, 333
vanilla 233, 334
Organization of Arab Petroleum Exporting Countries (OAPEC) 16, 70,
140
Organization of the Petroleum Exporting Countries (OPEC) 17, 47, 100,
157, 177, 264, 291, 306
Ornstein-Uhlenbeck process 40, 312
Other reportables (OTH) 131
Out-of-the-money (OTM) options 162
Over-the-counter (OTC) options 7, 233, 234, 237, 247, 249, 252, 254,
255, 291
P
Paper arbitrage 118, 120, 121
Parametrix method 227
Penultimate expiration 282, 333
Perturbation method 6, 209, 227–229, 319–322
Petrodollar recycling 140
Petroleum Administration for Defense Districts (PADDs) 47, 114
Pipelines 2, 21, 23, 46, 117, 119, 123, 125, 127, 287, 293, 302
Positioning 128–134
Position limits 307
Power markets 254, 294, 296, 305
Premium collars 214
Premium retained ratio 197, 199
Price elasticities of demand and supply 29, 34, 36, 37
Price of storage 30, 31
Principle of zero expectations 163
Probability density 40, 309
Producer hedging 57, 60–67, 209–214, 234
Producers, merchants, processors, and users (PMPUs) 129
Prompt futures 2, 333
Propane 127
Put-call parity 173
Put option 162
Q
Quadratic mean 233, 238–241
Quadratic normal (QN) model 6, 228, 229, 276, 277, 293, 319–322
Quadratic utility function 60
Quantamentals 5, 6, 109, 333
R
Ratio call spread 207
Reaction function 5, 83, 105–107, 116, 133, 134, 139, 196, 333
Realized correlation 299
Realized volatility 6, 161, 164, 175–181, 333
Real options 2–3, 6, 109, 118, 123, 125, 210–213, 282, 287, 305, 307
Real rate 11, 16
Rebalancing effect 68
Recessions 55, 69–71, 140
Reduced-form models 41, 293
Refined products 3, 62, 77, 99, 125–128, 219, 239, 292, 294
Refineries 3, 23, 46, 51, 112, 119, 123, 287, 288, 290, 294
Reformulated gasoline blendstock for oxygenate blending (RBOB) 89,
126, 131, 149–154
Regularization 260
Relative vega trading 210
Reverse carry trade 21
Risk aversion channel 142
Risk aversion coefficient 60, 62, 64, 76
Risk-neutral pricing 35, 164, 170, 250
Risk-neutral probabilities 43, 268, 269, 311, 327
Risk parity 55, 69–76, 105, 148, 333
Risk premium 55, 57, 59, 63, 64, 75, 77, 78, 86
Rockefeller, J.D. 3
Roll return 24–26, 66, 75, 333
Roll yield 5, 11, 23, 25, 26, 55, 58, 68, 75, 76, 97, 110, 132, 286
S
Safe haven channel 142
Samuelson effect 185, 240
Samuelson, P. 84
Scholes, M. 164, 169
Seasonality 88, 89, 99, 111, 117, 151, 153, 239, 254, 265
Sentiment index 133
Shale oil 47, 114, 116, 119, 120, 123, 125, 140, 177, 206, 210, 211, 221,
223, 236, 284
Shanghai International Energy Exchange (INE) 124
Signal blending 5, 101, 333
Signal transformation function 105
Silver 11, 12, 308
Simple options 163, 172
Skew correction function 227, 228, 319, 321
Skew delta 215–221, 333
Skewness 6, 86, 128, 179, 180, 183, 187, 190, 209, 212–214, 225, 226,
228, 229, 276
Spot prices 1
Spot return 24–26, 66, 68, 333
Spread options 2, 118, 281–302, 333
Squeeze 5, 29, 38–46, 48, 52, 116, 117, 286, 302, 333
Sraffa, P. 14
Stationarity 47, 48, 73, 122, 156, 158
Statistical arbitrage 5, 109, 122–124, 139, 148, 306
Sticky moneyness 6, 209, 217–220, 333
Sticky strike 6, 209, 219, 220, 333
Stochastic dynamic programming 35
Stochastic volatility 279
Stock-out 33–36, 45, 333
Storage
boundaries 29, 32–38, 43, 50, 115–117
capacity 33, 35–37, 51, 76, 115–116
capacity utilization 48, 115
costs 2, 21, 22, 31, 35, 76, 282
floating 21, 50
optimization 29
price of 30, 31
synthetic 281, 283, 286, 333
tanks 21, 123, 125
theory of 5, 29–37, 83, 87, 305
virtual 7, 281, 282, 286, 287, 305
Straddles 184, 188–190, 214, 218, 229, 276
Strangles 213–214, 229, 276
Super-backwardation 112
Super-contango 112
Swap Data Repositories (SDRs) 148, 333
Swap dealers (SDs) 129
Swaplets 249–251, 253, 267
Swaption 7, 233, 249–254, 266, 267, 333
Synthetic spread options 288, 333
Synthetic storage 281, 283, 286, 333
T
Tank-tops 33, 45, 334
Target redemption swap 255
Taylor formula 168, 309
Thales of Miletus 161
Theory of hedging pressure 4, 5, 55–64, 76–78, 112, 332
Theory of storage 5, 29–37, 83, 87, 305
Theta 172, 193, 283, 286, 287, 316, 317
Three-way collars 212–214, 333
Time spreads 109–117
Trading at settlement (TAS) contract 52, 95, 202, 283, 334
Transaction costs 95, 96, 145, 191, 200–204
Trapped option value 174
Treasury Inflation-Protected Securities (TIPS) 72, 148
Trend following 87
Triangular correlation arbitrage 7, 281, 287–291
Two-factor models 42, 253, 267
U
Ultra-low-sulfur diesel (ULSD) 126
Uncovered interest rate parity 97
US Department of Energy (DOE) 47
US dollar (USD) 5, 12, 13, 137–146, 156, 157
US Strategic Petroleum Reserve (SPR) 5, 16–18, 21
US terms of trade 140
V
Value risk premium 83, 101, 102
Vanilla options 233, 334
Vector autoregressive models (VAR) 71
Vega 6, 172, 181, 208–218, 229, 316, 317
Virtual rate of interest in commodities 14
Virtual storage 7, 281, 282, 286, 287, 305
VIX index 156
Volatility
arbitrage 230, 257
Black 181, 192, 245, 331
exponential 239, 246–247, 251, 254, 266
ghost 176
heuristics 226
implied 6, 161, 164, 180–186, 215–226, 239–241, 261–268, 276,
277, 332
implied Black (IBV) 181, 221–225
implied normal (INV) 181, 222–224
local 6, 161, 164, 167, 171, 172, 175, 177, 181, 185, 226–230, 237–
255, 261, 263–267, 273–279, 285, 292–294, 300, 301, 309, 319, 326,
327, 332
matrix 264
normal 165
realized 6, 161, 164, 175–181, 333
risk premium (VRP) 6, 187, 196–208, 286, 307, 334
risk premium (VRP) smile 6, 187, 198
risk premium (VRP) term structure 187, 199
skew 182, 334
smile 6, 161, 182, 183, 209, 214–221, 226, 276–278, 334
stochastic 279
targeting 105, 334
term structure 7, 185, 237–241, 261, 334
W
Wealth effect 143
Well-posed problems 258
Western Canadian Select (WCS) 125, 127, 145
West Texas Intermediate (WTI) 3, 18, 19, 26, 46, 47, 49–52, 112–115,
117–121, 124, 125, 127, 130, 132
West Texas Sour (WTS) 236
Weymar, H. 84
Whaley, R.E. 174
Working, H. 30
WTI-Brent accordion 5, 120
WTI Houston 125, 236
WTI Midland 125
X
XOP ETF 146, 147, 157
Y
YuanYouBao 52
Footnotes
1 The integration details are presented in many standard derivatives textbooks, such as
Wilmott et al. (1993) and Hull (2018).
2 See, for example, Kemna and Vorst (1990) and Wilmott et al. (1993) for a more detailed
derivation.
3 Some technical regularity conditions must be imposed to make sure that boundary terms at
K → 0 and K → ∞ vanish, which for simplicity, we omit here.
4 This problem was formulated in Bouchouev and Isakov (1997, 1999) and solved analytically
only under additional assumptions. For numerical solutions, see also references in the footnote
in Chap. 12.
5 The drift of the diffusion process must be fixed, and here it is assumed to be zero. It is well
known that two diffusions with different drifts and time-independent volatilities can generate
the same probability density. For example, the mean-reverting process and arithmetic
Brownian motion represent different diffusion processes, but they can produce the same
normal distribution, as shown in Appendix A.