Sie sind auf Seite 1von 25

wilm003.

qxd 7/26/02 7:05 PM Page 84

Managing
Smile Risk
Patrick S. Hagan*, Deep Kumar, Andrew S. Lesniewski,
and Diana E. Woodward

Abstract the smile. We apply the SABR model to USD interest rate options, and
Market smiles and skews are usually managed by using local volatility find good agreement between the theoretical and observed smiles.
models a la Dupire. We discover that the dynamics of the market smile pre- Key words. smiles, skew, dynamic hedging, stochastic vols, volga,
dicted by local vol models is opposite of observed market behavior: when vanna
the price of the underlying decreases, local vol models predict that the
smile shifts to higher prices; when the price increases, these models pre-
dict that the smile shifts to lower prices. Due to this contradiction between 1 Introduction
model and market, delta and vega hedges derived from the model can be
unstable and may perform worse than naive Black-Scholes hedges. European options are often priced and hedged using Blacks model, or,
To eliminate this problem, we derive the SABR model, a stochastic equivalently, the Black-Scholes model. In Blacks model there is a one-to-
volatility model in which the forward value satisfies one relation between the price of a European option and the volatility
parameter B . Consequently, option prices are often quoted by stating
dF = aF dW 1
the implied volatility B , the unique value of the volatility which yields the
da = a dW 2 options dollar price when used in Blacks model. In theory, the volatility
and the forward F and volatility a are correlated: dW 1 dW 2 = dt . We use B in Blacks model is a constant. In practice, options with different
singular perturbation techniques to obtain the prices of European strikes K require different volatilities B to match their market prices. See
options under the SABR model, and from these prices we obtain explicit, figure 1. Handling these market skews and smiles correctly is critical to
closed-form algebraic formulas for the implied volatility as functions of fixed income and foreign exchange desks, since these desks usually have
todays forward price f = F(0 ) and the strike K. These formulas immedi- large exposures across a wide range of strikes. Yet the inherent contra-
ately yield the market price, the market risks, including vanna and volga diction of using different volatilities for different options makes it diffi-
risks, and show that the SABR model captures the correct dynamics of cult to successfully manage these risks using Blacks model.

*phagan@bear.com; Bear-Stearns Inc, 383 Madison Ave, New York, NY 10179


BNP Paribas; 787 Seventh Avenue; New York NY 10019
BNP Paribas; 787 Seventh Avenue; New York NY 10019
Societe Generale; 1221 Avenue of the Americas; New York NY 10020

84 Wilmott magazine
wilm003.qxd 7/26/02 7:05 PM Page 85

TECHNICAL ARTICLE 4

The development of local volatility models by Dupire [2], [3] and Derman- M99 Eurodollar option
Kani [4], [5] was a major advance in handling smiles and skews. Local 30
volatility models are self-consistent, arbitrage-free, and can be calibrated to
precisely match observed market smiles and skews. Currently these mod- 25
els are the most popular way of managing smile and skew risk. However, as
we shall discover in section 2, the dynamic behavior of smiles and skews
20
predicted by local vol models is exactly opposite the behavior observed in Vol (%)
the marketplace: when the price of the underlying asset decreases, local vol
models predict that the smile shifts to higher prices; when the price increas- 15
es, these models predict that the smile shifts to lower prices. In reality, asset
prices and market smiles move in the same direction. This contradiction
10
between the model and the marketplace tends to de-stabilize the delta and
vega hedges derived from local volatility models, and often these hedges
perform worse than the naive Black-Scholes hedges. 5
92.0 93.0 94.0 95.0 96.0 97.0
To resolve this problem, we derive the SABR model, a stochastic
volatility model in which the asset price and volatility are correlated. Strike
Singular perturbation techniques are used to obtain the prices of Fig. 1.1 Implied volatility for the June 99 Eurodollar options. Shown are
European options under the SABR model, and from these prices we close-of-day values along with the volatilities predicted by the SABR model.
obtain a closed-form algebraic formula for the implied volatility as a Data taken from Bloomberg information services on March 23, 1999
function of todays forward price f and the strike K. This closed-form for-
mula for the implied volatility allows the market price and the market Here the expectation E is over the forward measure, and |F0 can be inter-
risks, including vanna and volga risks, to be obtained immediately from pretted as given all information available at t = 0. Martingale pricing the-
Blacks formula. It also provides good, and sometimes spectacular, fits to ory [6-9] also shows that the forward price F (t ) is a Martingale under this
the implied volatility curves observed in the marketplace. See Figure 1.1. measure, so the Martingale representation theorem shows that F (t ) obeys
More importantly, the formula shows that the SABR model captures the
correct dynamics of the smile, and thus yields stable hedges. dF = C (t, ) dW , F (0 ) = f, (2.1c)

for some coefficient C (t, ), where dW is Brownian motion in this meas-


2 Reprise ure. The coefficient C (t, ) may be deterministic or random, and may
depend on any information that can be resolved by time t. This is as far as
Consider a European call option on an asset A with exercise date tex , settle- the fundamental theory of arbitrage free pricing goes. In particular, one
ment date tset , and strike K. If the holder exercises the option on tex , then on cannot determine the coefficient C (t, ) on purely theoretical grounds.
the settlement date tset he receives the underlying asset A and pays the Instead one must postulate a mathematical model for C (t, ).
strike K. To derive the value of the option, define F(t ) to be the forward European swaptions fit within an indentical framework. Consider a
price of the asset for a forward contract that matures on the settlement European swaption with exercise date tex and fixed rate (strike) Rfix . Let
date tset , and define f = F(0 ) to be todays forward price. Also let D(t ) be Rs (t ) be the swaptions forward swap rate as seen at date t, and let
the discount factor for date t; that is, let D(t ) be the value today of $1 to be R0 = Rs (0 ) be the forward swap rate as seen today. In [9] Jamshidean shows
delivered on date t. Martingale pricing theory [6-9] asserts that under the that one can choose a measure in which the value of a payer swaption is
usual conditions, there is a measure, known as the forward measure,  
under which the value of a European option can be written as the expect- Vpay = L0 E [Rs (tex ) Rfix ]+ |F0 , (2.2a)
ed value of the payoff. The value of a call options is
  and the value of a receiver swaption is
Vcall = D(tset ) E [F(tex ) K ]+ |F0 , (2.1a)  
Vrec = L0 E [Rfix Rs (tex )]+ |F0
and the value of the corresponding European put is (2.2b)
Vpay + L0 [Rfix R0 ].
 
Here the level L0 is todays value of the annuity, which is a known quanti-
Vput = D(tset ) E [K F (tex )]+ |F0
(2.1b) ty, and E is the expectation over the level measure of Jamshidean [9]. In
Vcall + D(tset )[K f ].
W
Appendix A it is also shown that the PV01 of the forward swap; like the

Wilmott magazine 85
wilm003.qxd 7/26/02 7:05 PM Page 86

discount factor rate Rs (t ) is a Martingale in this measure, so once again 0.28

dRs = C(t, ) dW , Rs (0 ) = R0 , (2.2c)


0.26

where dW is Brownian motion. As before, the coefficient C (t, ) may be


1m
deterministic or random, and cannot be determined from fundamental 0.24
theory. Apart from notation, this is identical to the framework provided 3m

Vol
by equations (2.1a2.1c) for European calls and puts. Caplets and floor- 6m
lets can also be included in this picture, since they are just one period 0.22
payer and receiver swaptions. For the remainder of the paper, we adopt 12m
the notation of (2.1a2.1c) for general European options.
2.1 Blacks model and implied volatilities. To go any further 0.20
requires postulating a model for the coefficient C (t, ). In [10], Black pos-
tulated that the coefficient C (t, ) is B F (t ), where the volatilty B is a
0.18
constant. The forward price F (t ) is then geometric Brownian motion: 80 90 100 110 120
Strike
dF = B F (t ) dW , F (0 ) = f. (2.3)
Fig. 2.1 Implied volatility B (K ) as a function of the strike K for 1 month, 3 month,
Evaluating the expected values in (2.1a, 2.1b) under this model then 6 month, and 12 month European options on an asset with forward price 100.
yields Blacks formula,

Vcall = D (tset ){ f N (d1 ) KN (d2 )}, (2.4a) implied volatility at the barrier K2 , or some combination of the two to
price this option? Clearly, this option cannot be priced without a single,
self-consistent, model that works for all strikes without adjustments.
Vput = Vcall + D(tset )[K f ], (2.4b) The second problem is hedging. Since different models are being used
for different strikes, it is not clear that the delta and vega risks calculated at
where one strike are consistent with the same risks calculated at other strikes.
For example, suppose that our 1 month option book is long high strike
log f/K 12 B2 tex
d1,2 = , (2.4c) options with a total risk of +$1MM , and is long low strike options with a
B tex
of $1MM . Is our is our option book really -neutral, or do we have
for the price of European calls and puts, as is well-known [10], [11], [12]. residual delta risk that needs to be hedged? Since different models are
All parameters in Blacks formula are easily observed, except for the used at each strike, it is not clear that the risks offset each other.
volatility B . An options implied volatility is the value of B that needs to Consolidating vega risk raises similar concerns. Should we assume parallel
be used in Blacks formula so that this formula matches the market price or proportional shifts in volatility to calculate the total vega risk of our
of the option. Since the call (and put) prices in (2.4a 2.4c) are increasing book? More explicitly, suppose that B is 20% at K = 100 and 24% at
functions of B , the volatility B implied by the market price of an option K = 90, as shown for the 1m options in Figure 2.1 Should we calculate vega
is unique. Indeed, in many markets it is standard practice to quote prices by bumping B by, say, 0.2% for both options? Or by bumping B by 0.2% for
in terms of the implied volatility B ; the options dollar price is then the first option and by 0.24% for the second option? These questions are
recovered by substituting the agreed upon B into Blacks formula. critical to effective book management, since this requires consolidating
The derivation of Blacks formula presumes that the volatility B is a the delta and vega risks of all options on a given asset before hedging, so
constant for each underlying asset A . However, the implied volatility that only the net exposure of the book is hedged. Clearly one cannot
needed to match market prices nearly always varies with both the strike answer these questions without a model that works for all strikes K.
K and the time-to-exercise tex . See Figure 2.1. Changing the volatility B The third problem concerns evolution of the implied volatility curve
means that a different model is being used for the underlying asset for B (K ). Since the implied volatility B depends on the strike K, it is likely to
each K and tex . This causes several problems managing large books of also depend on the current value f of the forward price: B = B ( f, K ). In
options. this case there would be systematic changes in B as the forward price f of
The first problem is pricing exotics. Suppose one needs to price a call the underlying changes See Figure 2.1. Some of the vega risks of Blacks
option with strike K1 which has, say, a down-and-out knock-out at model would actually be due to changes in the price of the underlying
K2 < K1 . Should we use the implied volatility at the calls strike K1 , the asset, and should be hedged more properly (and cheaply) as delta risks.

86 Wilmott magazine
wilm003.qxd 7/26/02 7:05 PM Page 87

TECHNICAL ARTICLE 4

2.2 Local volatility models. An apparent solution to these problems and vega risks for all options, so these risks can be consolidated across
is provided by the local volatility model of Dupire [2], which is also attrib- strikes. Finally, perturbing f and re-calculating the option prices enables
uted to Derman [4], [5]. In an insightful work, Dupire essentially argued one to determine how the implied volatilites change with changes in the
that Black was to bold in setting the coefficient C (t, ) to B F. Instead one underlying asset price. Thus, the local volatility model thus provides a
should only assume that C is Markovian: C = C (t, F ). Re-writing C (t, F ) as method of pricing and hedging options in the presence of market smiles
loc (t, F ) F then yields the local volatility model, where the forward and skews. It is perhaps the most popular method of managing exotic
price of the asset is equity and foreign exchange options. Unfortunately, the local volatility
model predicts the wrong dynamics of the implied volatility curve, which
dF = loc (t, F ) FdW
, F (0 ) = f, (2.5a) leads to inaccruate and often unstable hedges.
To illustrate the problem, consider the special case in which the local
in the forward measure. Dupire argued that instead of theorizing about vol is a function of F only:
the unknown local volatility function loc (t, F ) , one should obtain
loc (t, F ) directly from the marketplace by calibrating the local volatili- ,
dF = loc (F ) FdW F (0 ) = f. (2.7)
ty model to market prices of liquid European options.
In calibration, one starts with a given local volatility function In [13] and [14] singular perturbation methods were used to analyze this
loc (t, F ), and evaluates model. There it was found that European call and put prices are given by
  Blacks formula (2.4a-2.4c) with the implied volatility
Vcall = D(tset ) E [F (tex ) K]+ |F (0 ) = f, (2.5b)   1
1 1 loc (2[f + K])
B (K, f ) = loc [f + K] 1+ (f K )2 + . (2.8)
2 24 loc ( 12 [ f + K ] )
Vput + D(tset )( f K ) (2.5c)
On the right hand side, the first term dominates the solution and the
to obtain the theoretical prices of the options; one then varies the local second term provides a much smaller correction The omitted terms are
volatility function loc (t, F ) until these theoretical prices match the actu- very small, usually less than 1% of the first term.
al market prices of the option for each strike K and exercise date tex . In The behavior of local volatility models can be largely understood by
practice liquid markets usually exist only for options with specific exer- examining the first term in (2.8). The implied volatility depends on both
cise dates tex1 , tex2 , tex3 , . . .; for example, for 1m, 2m, 3m, 6m, and 12m from the strike K and the current forward price f. So supppose that today the
today. Commonly the local vols loc (t, F ) are taken to be piecewise con- forward price is f0 and the implied volatility curve seen in the market-
stant in time: place is B0 (K ) . Calibrating the model to the market clearly requires
choosing the local volatility to be
loc (t, F ) = loc (F )
(1 )
for t < tez1 ,
loc (t, F ) = loc (F )
(j )
for texj1 < t < tezj j = 2, 3, . . . J (2.6) loc (F ) = B0 (2F f0 ){1 + }. (2.9)
loc (t, F ) = loc (F )
(J )
for t > tezJ
Now that the model is calibrated, let us examine its predictions. Suppose
that the forward value changes from f0 to some new value f. From (2.8),
(1 )
One first calibrates loc (F ) to reproduce the option prices at tex1 for all (2.9) we see that the model predicts that the new implied volatility curve
(2 )
strikes K, then calibrates loc (F ) to reproduce the option prices at tex2 , for is
all K, and so forth. This calibration process can be greatly simplified by
B (K, f ) = B0 (K + f f0 ){1 + } (2.10)
using the results in [13] and [14]. There we solve to obtain the prices of
European options under the local volatility model (2.5a2.5c), and from
these prices we obtain explicit algebraic formulas for the implied volatil- for an option with strike K, given that the current value of the forward
ity of the local vol models. price is f. In particular, if the forward price f0 increases to f, the implied
Once loc (t, F ) has been obtained by calibration, the local volatility volatility curve moves to the left; if f0 decreases to f, the implied volatility
model is a single, self-consistent model which correctly reproduces the curve moves to the right. Local volatility models predict that the market
market prices of calls (and puts) for all strikes K and exercise dates tex smile/skew moves in the opposite direction as the price of the underlying asset.
without adjustment. Prices of exotic options can now be calculated This is opposite to typical market behavior, in which smiles and skews
from this model without ambiguity. This model yields consistent delta move in the same direction as the underlying.
^

Wilmott magazine 87
wilm003.qxd 7/26/02 7:05 PM Page 88

To demonstrate the problem concretely, suppose that todays implied 0.28


volatility is a perfect smile
0.26

Implied Vol
B0 (K ) = + [K f0 ] 2
(2.11a) 0.24

around todays forward price f0 . Then equation (2.8) implies that the 0.22
local volatility is
0.20
2
loc (F ) = + 3(F f0 ) + . (2.11b) 0.18
f f0 K
As the forward price f evolves away from f0 due to normal market fluctu-
ations, equation (2.8) predicts that the implied volatility is Fig. 2.3 Implied volatility B (K, f ) if the forward price decreases from
f0 to f (solid line).
 
2 3
B (K, f ) = + K 32 f0 12 f + 4 ( f f0 ) 2 + . (2.11c)

The implied volatility curve not only moves in the opposite direction as predicted by the local volatility model. The first term is clearly the risk
the underlying, but the curve also shifts upward regardless of whether f one would calculate from Blacks model using the implied volatility from
increases or decreases. Exact results are illustrated in Figures 2.2 2.4. the market. The second term is the local volatility models correction to
There we assumed that the local volatility loc (F ) was given by (2.11b), the risk, which consists of the Black vega risk multiplied by the predict-
and used finite difference methods to obtain essentially exact values for ed change in B due to changes in the underlying forward price f. In real
the option prices, and thus implied volatilites. markets the implied volatily moves in the opposite direction as the direc-
Hedges calculated from the local volatility model are wrong. To see tion predicted by the model. Therefore, the correction term needed for
this, let BS( f, K, B , tex ) be Blacks formula (2.4a2.4c) for, say, a call real markets should have the opposite sign as the correction predicted by
option. Under the local volatility model, the value of a call option is the local volatility model. The original Black model yields more accurate hedges
given by Blacks formula than the local volatility model, even though the local vol model is self-consistent
across strikes and Blacks model is inconsistent.
Vcall = BS( f, K, B (K, f ), tex ) (2.12a) Local volatility models are also peculiar theoretically. Using any func-
tion for the local volatility loc (t, F ) except for a power law,
with the volatility B (K, f ) given by (2.8). Differentiating with respect to f
yields the risk C(t, ) = (t ) F , (2.13)
Vcall BS BS B (K, f )
= + . (2.12b)
f f B f
loc (t, F ) = (t ) F /F = (t ) /F1 , (2.14)

0.28 0.28

0.26 0.26
Implied Vol
Implied Vol

0.24 0.24

0.22 0.22

0.20 0.20

0.18 0.18
f0 K f0 f K

Fig. 2.2 Exact implied volatility B (K, f0 ) (solid line) obtained from Fig. 2.4 Implied volatility B (K, f ) if the forward prices increases from
the local volatility loc (F ) (dashed line): f0 to f (solid line).

88 Wilmott magazine
wilm003.qxd 7/26/02 7:05 PM Page 89

TECHNICAL ARTICLE 4

introduces an intrinsic length scale for the forward price F into the hedged by buying and selling options as a matter of routine, so
model. That is, the model becomes inhomogeneous in the forward price whether the market would be complete if these risks were not hedged
F. Although intrinsic length scales are theoretically possible, it is diffi- is a moot question.
cult to understand the financial origin and meaning of these scales [15], The SABR model (2.15a2.15c) is analyzed in Appendix B. There sin-
and one naturally wonders whether such scales should be introduced gular perturbation techniques are used to obtain the prices of European
into a model without specific theoretical justification. options. From these prices, the options implied volatility B (K, f ) is then
2.3 The SABR model. The failure of the local volatility model means obtained. The upshot of this analysis is that under the SABR model, the
that we cannot use a Markovian model based on a single Brownian price of European options is given by Blacks formula,
motion to manage our smile risk. Instead of making the model non-
Markovian, or basing it on non-Brownian motion, we choose to develop a Vcall = D(tset ){ f N (d1 ) KN (d2 )}, (2.16a)
two factor model. To select the second factor, we note that most markets
experience both relatively quiescent and relatively chaotic periods. This
suggests that volatility is not constant, but is itself a random function of Vput = Vcall + D(tset )[K .f ], (2.16a)
time. Respecting the preceding discusion, we choose the unknown coef-
ficient C (t, ) to be F , where the volatility is itself a stochastic with
process. Choosing the simplest reasonable process for now yields the
stochastic- model, which has become known as the SABR model. In
log f/K 12 B2 tex
this model, the forward price and volatility are d1,2 = , (2.16c)
B tex
dF = F dW 1 , F (0 ) = f (2.15a)
where the implied volatility B (f, K ) is given by

d = dW 2 , (0 ) = (2.15b)
B (K , f )
under the forward measure, where the two processes are correlated by:  
z
=  
( f K )(1 ) /2 1 + (1 )2
log 2 f/K + (1 )4
log 4 f/K + x(z )
dW 1 dW 2 = dt. (2.15c) 24 1920


Many other stochastic volatility models have been proposed, for example (1 )2 2 1 2 3 2 2
1+ + + tex + .
[16], [17], [18], [19]. However, the SABR model has the virtue of being the 24 ( f K )1 4 ( f K )(1 ) /2 24
simplest stochastic volatility model which is homogenous in F and . We (2.17a)
shall find that the SABR model can be used to accurately fit the implied
volatility curves observed in the marketplace for any single exercise date Here
tex . More importantly, it predicts the correct dynamics of the implied
z= ( f K)(1 ) /2 log f/K, (2.17b)
volatility curves. This makes the SABR model an effective means to man-
age the smile risk in markets where each asset only has a single exercise and x(z ) is defined by
date; these markets include the swaption and caplet/floorlet markets.
As written, the SABR model may or may not fit the observed volatility  
1 2 z + z2 + z
surface of an asset which has European options at several different exer- x(z ) = log . (2.17c)
1
cise dates; such markets include foreign exchange options and most
equity options. Fitting volatility surfaces requires the dynamic SABR model
which is discussed in an Appendix. For the special case of at-the-money options, options struck at K = f, this
It has been claimed by many authors that stochastic volatility mod- formula reduces to
els are models of incomplete markets, because the stochastic volatility
risk cannot be hedged. This is not true. It is true that the risk to
(1 )2 2
changes in (the vega risk) cannot be hedged by buying or selling the ATM = B ( f, f ) = 1+
f (1 ) 24 f 22
underlying asset. However, vega risk can be hedged by buying or selling (2.18)
1 2 3 2
2
options on the asset in exactly the same way that -hedging is used to + (1 ) + tex + .
^

neutralize the risks to changes in the price F. In practice, vega risks are 4f 24

Wilmott magazine 89
wilm003.qxd 7/26/02 7:05 PM Page 90

These formulas are the main result of this paper. Although it appears 22%
formidable, the formula is explicit and only involves elementary trigno- 20%
=0
metric functions. Implementing the SABR model for vanilla options is very 18%

Implied vol
easy, since once this formula is programmed, we just need to send the 16%
options to a Black pricer. In the next section we examine the qualitative
14%
behavior of this formula, and how it can be used to managing smile risk.
The complexity of the formula is needed for accurate pricing. 12%
Omitting the last line of (2.17a), for example, can result in a relative error 10%
that exceeds three per cent in extreme cases. Although this error term 8%
seems small, it is large enough to be required for accurate pricing. The 4% 6% 8% 10% 12%
omitted terms + are much, much smaller. Indeed, even though we Fig. 3.1 Backbone and smiles for = 0. As the forward f varies, the implied
have derived more accurate expressions by continuing the perturbation volatiliity B ( f, f ) of ATM options traverses the backbone (dashed curve). Shown are
expansion to higher order, (2.17a 2.17c) is the formula we use to value the smiles B (K, f ) for three different values of the forward. Volatility data from 1
and hedge our vanilla swaptions, caps, and floors. We have not imple- into 1 swaption on 4/28/00, courtesy of Cantor-Fitzgerald.
mented the higher order results, believing that the increased precision
of the higher order results is superfluous.
22%
There are two special cases of note: = 1, representing a stochastic
20%
log normal model), and = 0, representing a stochastic normal model. =1
18%
Implied vol

The implied volatility for these special cases is obtained in the last sec-
tion of Appendix B. 16%
14%
12%
3 Managing Smile Risk 10%
8%
The complexity of the above formula for B (K, f ) obscures the qualita- 4% 6% 8% 10% 12%
tive behavior of the SABR model. To make the models phenomenology
and dynamics more transparent, note that formula (2.17a 2.17c) can Fig. 3.2 Backbone and smiles as above, but for = 1.
be approximated as

1
B (K, f ) = 1 1 (1 ) log K/f Let us now consider the implied volatility B (K, f ) in detail. The first
f 2
(3.1a) factor /f 1 in (3.1a( is the implied volatility for at-the-money (ATM)
1 

+ (1 )2 + (2 3 2 ) 2 log 2 K/f + , options, options whose strike K equals the current forward f. So the back-
12 bone traversed by ATM options is essentially B ( f, f ) = /f 1 for the
provided that the strike K is not too far from the current forward f. Here SABR model. The backbone is almost entirely determined by the expo-
the ratio nent , with the exponent = 0 (a stochastic Gaussian model) giving a
steeply downward sloping backbone, and the exponent = 1 giving a
= f 1 (3.1b)
nearly flat backbone.
The second term 12 (1 ) log K/f represents the skew, the
measures the strength of the volatility of volatility (the volvol) com- slope of the implied volatility with respect to the strike K . The
pared to the local volatility /f 1 at the current forward. Although equa- 12 (1 ) log K/f part is the beta skew, which is downward sloping since
tions (3.1a3.1b) should not be used to price real deals, they are accurate 0 1 . It arises because the local volatility F /F1 = /F1 is a
enough to depict the qualitative behavior of the SABR model faithfully. decreasing function of the forward price. The second part 12 log K/f is
As f varies during normal trading, the curve that the ATM volatility the vanna skew, the skew caused by the correlation between the volatility
B (f, f ) traces is known as the backbone, while the smile and skew refer to and the asset price. Typically the volatility and asset price are negatively
the implied volatility B (K, f ) as a function of strike K for a fixed f. That correlated, so on average, the volatility would decrease (increase) when
is, the market smile/skew gives a snapshot of the market prices for dif- the forward f increases (decreases). It thus seems unsurprising that a
ferent strikes K at a given instance, when the forward f has a specific negative correlation causes a downward sloping vanna skew.
price. Figures 3.1 and 3.2. show the dynamics of the smile/skew predicted It is interesting to compare the skew to the slope of the backbone. As f
by the SABR model. changes to f the ATM vol changes to

90 Wilmott magazine
wilm003.qxd 7/26/02 7:05 PM Page 91

TECHNICAL ARTICLE 4

 1y into 1y
f f
B (f , f ) = 1 (1 ) + . (3.2a) 22%
f 1 f
20%

Implied vol
Near K = f, the component of skew expands as 18%

 16%
1 Kf
B (K , f ) = 1 (1 ) + , (3.2b)
f 1 2 f 14%

12%
so the slope of the backbone B ( f, f ) is twice as steep as the slope of rthe 4% 6% 8% 10% 12%
smile B (K, f ) due to the -component of the skew.
Fig. 3.3 Implied volatilities as a function of strike. Shown are the curves
The last term in (3.1a) also contains two parts. The first part obtained by fitting the SABR model with exponent = 0 and with = 1
1
12
(1 )2 log 2 K/f appears to be a smile (quadratic) term, but it is domi- to the 1y into 1y swaption vol observed on 4/28/00. As usual, both fits are
nated by the downward sloping beta skew, and, at reasonable strikes K, it equally good. Data courtesy of Cantor-Fitzgerald.
just modifies this skew somewhat. The second part 12 1
(2 3 2 )
2 2
log K/f is the smile induced by the volga (vol-gamma) effect. Physically
this smile arises because of adverse selection: unusually large move- The exponent can be determined from historical observations of the
ments of the forward F happen more often when the volatility increas- backbone or selected from aesthetic considerations. Equation (2.18)
es, and less often when decreases, so strikes K far from the money rep- shows that the implied volatility of ATM options is
resent, on average, high volatility environments.
(1 )2 2
3.1 Fitting market data. The exponent and correlation affect log B ( f, f ) = log (1 ) log f + log 1 +
24 f 22
the volatility smile in similar ways. They both cause a downward slop-  (3.3)
ing skew in B (K, f ) as the strike K varies. From a single market snap- 1 2 3 22
+ (1 ) + tex + .
shot of B (K, f ) as a function of K at a given f, it is difficult to distin- 4f 24
guish between the two parameters. This is demonstrated by figure 3.3.
There we fit the SABR parameters , , with = 0 and then re-fit the The exponent can be extracted from a log log plot of historical observa-
parameters , , with = 1 . Note that there is no substantial differ- tions of f, ATM pairs. Since both f and are stochastic variables, this fit-
ence in the quality of the fits, despite the presence of market noise. This ting procedure can be quite noisy, and as the [ ]tex term is typically less
matches our general experience: market smiles can be fit equally well than one or two per cent, it is usually ignored in fitting .
with any specific value of . In particular, cannot be determined by Selecting from aesthetic or other a priori considerations usually
fitting a market smile since this would clearly amount to fitting the results in = 1 (stochastic lognormal), = 0 (stochastic normal), or
noise. = 12 (stochastic CIR) models. Proponents of = 1 cite log normal mod-
Figure 3.3 also exhibits a common data quality issue. Options with els as being more natural. or believe that the horizontal backbone best
strikes K away from the current forward f trade less frequently than at- represents their market. These proponents often include desks trading
the-money and near-the-money options. Consequently, as K moves away foreign exchange options. Proponents of = 0 usually believe that a nor-
from f, the volatility quotes become more suspect because they are more mal model, with its symmetric break-even points, is a more effective tool
likely to be out-of-date and not represent bona fide offers to buy or sell for managing risks, and would claim that = 0 is essential for trading
options. markets like Yen interest rates, where the forwards f can be negative or
Suppose for the moment that the exponent is known or has been near zero. Proponents of = 12 are usually US interest rate desks that
selected. Taking a snapshot of the market yields the implied volatility have developed trust in CIR models.
B (K, f ) as a function of the strike K at the current forward price f. With It is usually more convenient to use the at-the-money volatility
given, fitting the SABR model is a straightforward procedure. The ATM , , , and as the SABR parameters instead of the original parame-
three parameters , , and have different effects on the curve: the ters ,, , . The parameter is then found whenever needed by invert-
parameter mainly controls the overall height of the curve, changing ing (2.18) on the fly; this inversion is numerically easy since the [ ]tex
the correlation controls the curves skew, and changing the vol of vol term is small. With this parameterization, fitting the SABR model
controls how much smile the curve exhibits. Because of the widely seper- requires fitting and to the implied volatility curve, with ATM and
ated roles these parameters play, the fitted parameter values tend to be given. In many markets, the ATM volatilities need to be updated fre-
^

very stable, even in the presence of large amounts of market noise. quently, say once or twice a day, while the smiles and skews need to be

Wilmott magazine 91
wilm003.qxd 7/26/02 7:05 PM Page 92

U99 Eurodollar option Z99 Eurodollar option


30 20

25
18

Vol (%)
Vol (%)

20
16
15

10 14
93.0 94.0 95.0 96.0 97.0 92.0 93.0 94.0 95.0 96.0
Strike Strike

Fig.3.4 Volatility of the Sep 99 EDF options Fig. 3.5 Volatility of the Dec 99 EDF options

H00 Eurodollar option


updated infrequently, say once or twice a month. With the new parame- 24
terization, ATM can be updated as often as needed, with , (and )
updated only as needed. 22
Let us apply SABR to options on US dollar interest rates. There are
Vol (%)

three key groups of European options on US rates: Eurodollar future 20


options, caps/floors, and European swaptions. Eurodollar future options
18
are exchange-traded options on the 3 month Libor rate; like interest rate
futures, EDF options are quoted on 100(1 rL i b o r ) . Figure 1.1 fits the
16
SABR model (with = 1) to the implied volatility for the June 99 con-
tracts, and figures 3.43.7 fit the model (also with = 1) to the implied 14
volatility for the September 99, December 99, and March 00 contracts. 92.0 93.0 94.0 95.0 96.0 97.0
All prices were obtained from Bloomberg Information Services on March Strike
23, 1999. Two points are shown for the same strike where there are
quotes for both puts and calls. Note that market liquidity dries up for the Fig. 3.6 Volatility of the Mar 00 EDF options
later contracts, and for strikes that are too far from the money.
Consequently, more market noise is seen for these options. the skew is overwhelmed by the smile for short-dated options, but is
Caps and floors are sums of caplets and floorlets; each caplet and important for long-dated options. This picture is confirmed tables 3.1
floorlet is a European option on the 3 month Libor rate. We do not con- and 3.2. These tables show the values of the vol of vol and correlation
sider the cap/floor market here because the broker-quoted cap prices obtained by fitting the smile and skew of each m into n swaption,
must be stripped to obtain the caplet volatilities before SABR can be again using the data from April 28, 2000. Note that the vol of vol is very
applied. high for short dated options, and decreases as the time-to-exercise
A m year into n year swaption is a European option with m years to the increases, while the correlations starts near zero and becomes substan-
exercise date (the maturity); if it is exercised, then one receives an n year tially negative. Also note that there is little dependence of the market
swap (the tenor, or underlying) on the 3 month Libor rate. See Appendix skew/smile on the length of the underlying swap; both and are fairly
A. For almost all maturities and tenors, the US swaption market is liquid constant across each row. This matches our general experience: in most
for at-the-money swaptions, but is ill-liquid for swaptions struck away markets there is a strong smile for short-dated options which relaxes as
from the money. Hence, market data is somewhat suspect for swaptions the time-to-expiry increases; consequently the volatility of volatility is
that are not struck near the money. Figures 3.83.11 fits the SABR model large for short dated options and smaller for long-dated options, regard-
(with = 1) to the prices of m into5Y swaptions observed on April 28, less of the particular underlying. Our experience with correlations is less
2000. Data supplied courtesy of Cantor-Fitzgerald. clear: in some markets a nearly flat skew for short maturity options
We observe that the smile and skew depend heavily on the time-to- develops into a strongly downward sloping skew for longer maturities. In
exercise for Eurodollar future options and swaptions. The smile is pro- other markets there is a strong downward skew for all option maturities,
nounced for short-dated options and flattens for longer dated options; and in still other markets the skew is close to zero for all maturities.

92 Wilmott magazine
wilm003.qxd 7/26/02 7:05 PM Page 93

TECHNICAL ARTICLE 4

M00 Eurodollar option 1Y into 5Y


24 18%
17%
22
16%
Vol (%)

20 15%

18 14%
13%
16 4% 6% 8% 10% 12%

14 Fig. 3.9 Volatilities of 1 year into 1 year swaptions


92.0 93.0 94.0 95.0 96.0 97.0
Strike
5Y into 5Y
Fig, 3.7 Volatility of the Jun 00 EDF options 17%
16%
15%
3M into 5Y
20% 14%

18% 13%
12%
16% 4% 6% 8% 10% 12%
14%
Fig. 3.10 Volatilities of 5 year into 5 year swaptions
12%
4% 6% 8% 10% 12%

Fig. 3.8 Volatilities of 3 month into 5 year swaption amount as the price f. This predicted dynamics of the smile matches
market experience. If < 1, the backbone is downward sloping, so the
shift in the implied volatility curve is not purely horizontal. Instead, this
3.2. Managing smile risk. After choosing and fitting , , and curve shifts up and down as the at-the-money point traverses the back-
either or ATM , the SABR model bone. Our experience suggests that the parameters and are very sta-
ble ( is assumed to be a given constant), and need to be re-fit only every
dF = F dW 1 , F(0 ) = f (3.4a) few weeks. This stability may be because the SABR model reproduces the
usual dynamics of smiles and skews. In contrast, the at-the-money
volatility ATM , or, equivalently, may need to be updated every few
d = dW 2 , (0 ) = (3.4b) hours in fast-paced markets.
Since the SABR model is a single self-consistent model for all strikes K,
with the risks calculated at one strike are consistent with the risks calculated at
other strikes. Therefore the risks of all the options on the same asset can be
dW 1 dW 2 = dt (3.4c) added together, and only the residual risk needs to be hedged.
Let us set aside the risk for the moment, and calculate the other
risks. Let BS( f, K, B , tex ) be Blacks formula (2.4a2.4c) for, say, a call
fits the smiles and skews observed in the market quite well, especially option. According to the SABR model, the value of a call is
considering the quality of price quotes away from the money . Let us take
for granted that it fits well enough. Then we have a single, self-consistent Vcall = BS( f, K, B (K, f ), tex ) (3.5)
model that fits the option prices for all strikes K without adjustment,
so we can use this model to price exotic options without ambiguity. The where the volatility B (K, f ) B (K, f ; , , , ) is given by equations
SABR model also predicts that whenever the forward price f changes, the (2.17a2.17c). Differentiating @ footnote:{In practice risks are calculated
^

the implied volatility curve shifts in the same direction and by the same by finite differences: valuing the option at , re-valuing the option after

Wilmott magazine 93
wilm003.qxd 7/26/02 7:05 PM Page 94

10Y into 5Y bumping the forward to + , and then subtracting to determine the
13% risk This saves differentiating complex formulas such as (2.17a2.17c).
with respect to yields the vega risk, the risk to overall changes in
12%
volatility:
11%
Vcall BS B (K, f ; , , , )
10% = . (3.6)
B
9%
4% 6% 8% 10% 12%
This risk is the change in value when changes by a unit amount. It is
Fig. 3.11 Volatilities of 10 year into 5 year options traditional to scale vega so that it represents the change in value when
the ATM volatility changes by a unit amount. Since ATM
= ( ATM / ) , the vega risk is

B (K ,f;, , , )
Vcall BS
TABLE 3.1 vega =
ATM (f;, , , )
(3.7a)
B
VOLATILITY OF VOLATILITY FOR EUROPEAN

SWAPTIONS. ROWS ARE TIMETOEXERCISE;


COLUMNS ARE TENOR OF THE UNDERLYING SWAP. where ATM ( f ) = B ( f, f ) is given by (2.18). Note that to leading order,
B / B / and ATM / ATM / , so the vega risk is roughly given
by
1Y 2Y 3Y 4Y 5Y 7Y 10Y
1M 76.2% 75.4% 74.6% 74.1% 75.2% 73.7% 74.1%
BS B (K, f ) BS B (K, f )
3M 65.1% 62.0% 60.7% 60.1% 62.9% 59.7% 59.5% vega = . (3.7b)
B ATM ( f ) B B ( f, f )
6M 57.1% 52.6% 51.4% 50.8% 49.4% 50.4% 50.0%
1Y 59.8% 49.3% 47.1% 46.7% 46.0% 45.6% 44.7%
3Y 42.1% 39.1% 38.4% 38.4% 36.9% 38.0% 37.6% Qualitatively, then, vega risks at different strikes are calculated by bump-
5Y 33.4% 33.2% 33.1% 32.6% 31.3% 32.3% 32.2% ing the implied volatility at each strike K by an amount that is propor-
7Y 30.2% 29.2% 29.0% 28.2% 26.2% 27.2% 27.0% tional to the implied volatiity B (K, f ) at that strike. That is, in using
10Y 26.7% 26.3% 26.0% 25.6% 24.8% 24.7% 24.5% equation (3.7a), we are essentially using proportional, and not parallel,
shifts of the volatility curve to calculate the total vega risk of a book of
options.
Since and are determined by fitting the implied volatility curve
observed in the marketplace, the SABR model has risks to and
changing. Borrowing terminology from foreign exchange desks,
TABLE 3.2 vanna is the risk to changing and volga (vol gamma) is the risk
MATRIX OF CORRELATIONS BETWEEN THE UNDERLY- to changing:
ING AND THE VOLATILITY FOR EUROPEAN SWAPTONS. Vcall BS B (K, f ; , , , )
vanna = = , (3.8a)
B
1Y 2Y 3Y 4Y 5Y 7Y 10Y
Vcall BS B (K, f ; , , , )
1M 4.2% 0.2% 0.7% 1.0% 2.5% 1.8% 2.3% volga = = . (3.8b)
B
3M 2.5% 4.9% 5.9% 6.5% 6.9% 7.6% 8.5%
6M 5.0% 3.6% 4.9% 5.6% 7.1% 7.0% 8.0%
1Y 4.4% 8.1% 8.8% 9.3% 9.8% 10.2% 10.9% Vanna basically expresses the risk to the skew increasing, and
3Y 7.3% 14.3% 17.1% 17.1% 16.6% 17.9% 18.9% volga expresses the risk to the smile becoming more pronounced.
5Y 11.1% 17.3% 18.5% 18.8% 19.0% 20.0% 21.6% These risks are easily calculated by using finite differences on the
7Y 13.7% 22.0% 23.6% 24.0% 25.0% 26.1% 28.7% formula for B in equations (2.17a2.17c). If desired, these risks
10Y 14.8% 25.5% 27.7% 29.2% 31.7% 32.3% 33.7% can be hedged by buying or selling away-from-the-money options.

94 Wilmott magazine
wilm003.qxd 7/26/02 7:05 PM Page 95

TECHNICAL ARTICLE 4

The delta risk expressed by the SABR model depends on whether one dF = C(F ) dW 1 , (A.1a)
uses the parameterization , , , or ATM , , , . Suppose first we use
the parameterization , , , , so that B (K, f ) B (K, f ; , , , ) .
d = dW 2 , (A.1b)
Differentiating respect to f yields the risk
with
Vcall BS BS B (K, f ; , , , )
= + . (3.9)
f f B f dW 1 dW 2 = dt, (A.1c)

The first term is the ordinary risk one would calculate from Blacks in the limit  1. This is the distinguished limit [21], [22] in the language
model. The second term is the SABR models correction to the risk. It of singular perturbation theory. After obtaining the results we replace
consists of the Black vega times the predicted change in the implied volatil- , and to get the answer in terms of the original vari-
ity B caused by the change in the forward f. As discussed above, the pre- ables. We first analyze the model with a general C(F ), and then specialize
dicted change consists of a sideways movement of the volatility curve in the results to the power law F . This is notationally simpler than working
the same direction (and by the same amount) as the change in the for- with the power law throughout, and the more general result may prove
ward price f. In addition, if < 1 the volatility curve rises and falls as the valuable in some future application.
at-the-money point traverses up and down the backbone. There may also We first use the forward Kolmogorov equation to simplify the option
be minor changes to the shape of the skew/smile due to changes in f. pricing problem. Suppose the economy is in state F (t ) = f, (t ) = at
Now suppose we use the parameterization AMT , , , . Then is a date t. Define the probability density p (t, f, ; T, F, A ) by
function of ATM and f defined implicitly by (2.18). Differentiating (3.5)
now yields the risk 
 p (t, f, ; T, F, A ) dF dA = prob F < F(T ) < F + dF, A < (T )
BS BS B (K, f ; , , , ) B (K, f ; , , , ) (ATM , f )   (A.2)
+ + . 
f B f f < A + dA  F (t ) = f, (t ) = .

(3.10)
The delta risk is now the risk to changes in f with ATM held fixed. The As a function of the forward variables T, F, A, the density p satisfies the for-
last term is just the change in needed to keep ATM constant while f ward Kolmogorov equation (the Fokker-Planck equation)
changes. Clearly this last term must just cancel out the vertical compo-
nent of the backbone, leaving only the sideways movement of the pT = 12 2 A2 [C 2 (F ) p]FF
implied volatilty curve. Note that this term is zero for = 1. (A.3a)
Theoretically one should use the from equation (3.9) to risk man- + 2 [A2 C(F ) p]FA + 12 2 2 [A2 p]A A for T > t,
age option books. In many markets, however, it may take several days for
volatilities B to change following significant changes in the forward with
price f. In these markets, using from (3.10) is a much more effective
hedge. For suppose one used from equation (3.9). Then, when the
volatility ATM did not immediately change following a change in f, one p = (F f ) (A ) at T = t, (A.3b)
would be forced to re-mark to compensate, and this re-marking would
change the hedges. As ATM equilibrated over the next few days, one
would mark back to its original value, which would change the as is well-known [24], [25], [26]. Here, and throughout, we use subscripts
hedges back to their original value. This hedging chatter caused by to denote partial derivatives.
market delays can prove to be costly. Let V (t, f, ) be the value of a European call option at date t, when the
economy is in state F (t ) = f, (t ) = .Let tex be the options exercise date,
and let K be its strike. Omitting the discount factor D (tset ), which factors
Appendix A. Analysis of the SABR Model out exactly, the value of the option is
Here we use singular perturbation techniques to price European options
 
under the SABR model. Our analysis is based on a small volatility expan-
V(t, f, ) = E [F (tex ) K ]+ | F (t ) = f, (t ) =
sion, where we take both the volatility and the volvol to be small. To   (A.4)
carry out this analysis in a systematic fashion, we re-write , and
^

= (F K ) p (t, f, ; tex , F, A ) dF dA.


, and analyze K

Wilmott magazine 95
wilm003.qxd 7/26/02 7:05 PM Page 96

See (2.1a). Since


1 2 2 2 1
 tex P = C ( f ) Pff + 2 2 C( f ) Pf + 2 2 2 P , for > 0, (A.13a)
2 2
p (t, f, ; tex , F, A ) = (F f ) (A ) + pT (t, f, ; T, F, A ) dT, (A.5)
t

we can re-write V(t, f, ) as P = 2 ( f K ), for = 0. (A.13b)

 tex   In this appendix we solve (A.13a), (A.13b) to obtain P (, f, ; K ), and


V ( t, f, ) = [ f K ]+ + (F K ) pT (t, f, ; T, F, A ) dA dF dT. then substitute this solution into (A.12) to obtain the option value
t K
(A.6) V (t, f, ). This yields the option price under the SABR model, but the
resulting formulas are awkward and not very useful. To cast the results
We substitute (A.3a) for pT into (A.6). Integrating the A derivatives in a more usable form, we re-compute the option price under the normal
2 [A2 C(F ) p]FA and 12 2 2 [A2 p]A A over all A yields zero. Therefore our model
option price reduces to
d F = N dW , (A.14a)
 tex  
1
V ( t, f, ) = [ f K ]+ + 2 A2 (F K ) [C 2 (F ) p]FF dF dA dT,
2 t K
(A.7) and then equate the two prices to determine which normal volatility N
needs to be used to reproduce the options price under the SABR model.
That is, we find the implied normal volatility of the option under the
where we have switched the order of integration. Integrating by parts SABR model. By doing a second comparison between option prices under
twice with respect to F now yields the log normal model

 tex  dF = B F dW (A.14b)
1 2 2
V ( t, f, ) = [ f K ]+ + C (K ) A2 p(t, f, ; T, K, A ) dA dT. (A.8)
2 t
and the normal model, we then convert the implied normal volatility to
the usual implied log-normal (Black-Scholes) volatility. That is, we quote
The problem can be simplified further by defining the option price predicted by the SABR model in terms of the options
 implied volatility.
P (t, f, ; T, K ) = A2 p(t, f, ; T, K, A ) dA. (A.9) A.1 Singular perturbation expansion. Using a straightforward per-

turbation expansion would yield a Gaussian density to leading order,
Then P satisfies the backwards Kolmogorov equation [24], [25], [26]
1 2 2 2 1
( f K ) 2

Pt + C ( f ) Pff + 2 2 C( f ) Pf + 2 2 2 P = 0, for t < T P=  e 2 2 2 C 2 ( K ) {1 + }. (A.15a)


2 2 2 2 C2 (K )
(A.10a)

Since the + involves powers of ( f K ) / C (K ) , this expansion


P = 2 ( f K ), for t = T. (A.10b) would become inaccurate as soon as ( f K ) C (K ) /C (K ) becomes a sig-
nificant fraction of 1; i.e., as soon as C ( f ) and C (K ) are significantly dif-
Since t does not appear explicitly in this equation, P depends only on the ferent. Stated differently, small changes in the exponent cause much
combination T t, and not on t and T separately. So define greater changes in the probability density. A better approach is to re-cast
the series as
= T t, ex = tex t. (A.11)

( f K ) 2
{1+ }
P=  e 2 2 2 C 2 ( K ) (A.15b)
Then our pricing formula becomes 2 2 C2 (K )

 ex
1 2 2 and expand the exponent, since one expects that only small changes to
V ( t, f, ) = [ f K ]+ + C (K ) P (, f, ; K ) d (A.12)
2 0 the exponent will be needed to effect the much larger changes in the
density. This expansion also describes the basic physics better - P is
where P (, f, ; K ) is the solution of the problem essentially a Gaussian probability density which tails off faster or slower

96 Wilmott magazine
wilm003.qxd 7/26/02 7:05 PM Page 97

TECHNICAL ARTICLE 4

depending on whether the diffusion coefficient C ( f ) decreases or


2 2 2z 2 z2 2 2z
increases. 2
2
+ 2 2 + 2 . (A.19d)
z z z
We can refine this approach by noting that the exponent is the
integral
Also,
  2
(f K) 2
1 1 f
df 1
{1 + } = {1 + }. (A.16) ( f K ) = ( zC(K ) ) = (z ) . (A.19e)
2 2 2 C 2 (K ) 2 K C (f ) C(K )

Suppose we define the new variable Therefore, (A.12) through (A.13b) become

 
1 f
df 1 2 2 ex
z= . (A.17) V ( t, f, a ) = [ f K ]+ + C (K ) P (, z, ) d, (A.20)
K C(f ) 2 0

so that the solution P is essentially e z /2 . To leading order, the density is


2
where P(, z, ) is the solution of
Gaussian in the variable z, which is determined by how easy or hard
it is to diffuse from K to f, which closely matches the underlying physics.
1 1 B
The fact that the Gaussian changes by orders of magnitude as P = 1 2 z + 2 2 z2 Pzz a Pz + ( 2 2 z )( Pz Pz )
z 2 increases should be largely irrelevent to the quality of the expansion. 2 2 B
1 2 2 2
This approach is directly related to the geometric optics technique that + P a for > 0
2
is so successful in wave propagation and quantum electronics [27], [22]. (A.21a)
To be more specific, we shall use the near identity transform method to
carry out the geometric optics expansion. This method, pioneered in

[28], transforms the problem order-by-order into a simple canonical P= (z ) at = 0. (A.21b)
C (K )
problem, which can then be solved trivially. Here we obtain the solution
only through O ( 2 ), truncating all higher order terms.
Let us change variables from f to Accordingly, let us define P(, z, ) by
 f
1 df
z= , (A.18a) P = C(K ) P. (A.22)
K C( f )
and to avoid confusion, we define In terms of P, we obtain
 ex
1
B ( z ) = C ( f ) . (A.18b) V ( t, f, a ) = [ f K ]+ + C(K ) P (, z, ) d, (A.23)
2 0

Then where P (, z, ) is the solution of


1 1 z 1 1 B
= , , (A.19a) P = 1 2 z + 2 2 z 2 Pz z a Pz + ( 2 2 z ) Pz
f C ( f ) z B( z ) z z 2 2 B (A.24a)
1 2 2 2
+ ( P + 2 P ) for > 0,
and 2
P = (z ) at = 0. (A.24b)

2 1 2 B ( z )
2
2 2 2 2
, (A.19b) To leading order P is the solution of the standard diffusion problem
f B ( z ) z B( z ) z
P = 12 Pzz with P = (z ) at = 0. So it is a Gaussian to leading order. The
next stage is to transform the problem to the standard diffusion prob-
lem through O ( ), and then through O ( 2 ), . . . . This is the near identi-
 fy transform method which has proven so powerful in near-
^

2 1 2 z 2 1
, (A.19c) Hamiltonian systems [28].
f B( z ) z z2 z

Wilmott magazine 97
wilm003.qxd 7/26/02 7:05 PM Page 98

Note that the variable does not enter the problem for P until O( ), so of until O( 2 ) and the derivatives are actually no larger than O( 2 ).
Thus, the last term is actually only O( 3 ), and can be neglected since we
P (, z, ) = P0 (, z ) +P1 (, z, ) + (A.25) are only working through O( 2 ). So,
1 1 B
Consequently, the derivatives Pz , P , and P are all O ( ). Recall that we H = 1 2 z + 2 2 z2 Hz z 2 (zHz H )
2 2 B
are only solving for P through O ( 2 ). So, through this order, we can re-   (A.31a)
1 B 3 B 2
write our problem as +
2 2
H for > 0
4 B 8 B2

1 1 B
P = 1 2 z + 2 2 z2 Pz z a Pz + Pz for > 0 H = (z ) at = 0. (A.31b)
2 2 B
(A.26a) There are no longer any derivatives, so we can now treat as a parame-
ter instead of as an independent variable. That is, we have succeeded in
P " = "(z ) at = 0. (A.26b)
effectively reducing the problem to one dimension.
Let us now eliminate the 1
2
a(B /B ) Pz term. Define H(, z, ) by Let us now remove the Hz term through O( 2 ) . To leading order,
  B ( z ) /B( z ) and B ( z ) /B( z ) are constant. We can replace these

P = C ( f ) /C (K )H B( z ) /B(0 )H. (A.27) ratios by

Then b1 = B ( z0 ) /B( z0 ), b2 = B ( z0 ) /B ( z0 ), (A.32)



 1 B
Pz = B( z ) /B(0 ) Hz + H , (A.28a) commiting only an O( ) error, where the constant z0 will be chosen later.
2 B
We now define H by

 B B B 2
Pzz = H.
2
b 1 z 2 /4
B( z ) /B(0 ) H zz + Hz + 2 2 H , (A.28b) H = e (A.33)
B 2B 4B 2
 Then our option price becomes
 1 B 1 B 1 B
Pz = B( z ) /B(0 ) H z + z Hz + H + H + O( 2 ) 
2 B 2 B 2 B 1  ex
V ( t, f, a ) = [ f K ]+ + H(, z ) d,
2 2
B(0 ) B( z )e b1 z /4
2 0
(A.28c) (A.34)

The option price now becomes where H is the solution of


  
1  ex
1 1 3
V ( t, f, a ) = [ f K ]+ + B(0 ) B( z ) H(, z, ) d, (A.29) H = 1 2 z + 2 2 z2 Hz z + 2 2 b2 b21 H
2 0 2 4 8
(A.35a)
3
where + 2 b1 H for > 0
4
1 1 B
H = 1 2 z + 2 2 z2 Hzz 2 (zHz H ) H = (z ) at = 0. (A.35b)
2 2 B
 
1 B 3 B 2 1 B
+
2 2
2
H + (Hz + H ) for > 0 Weve almost beaten the equation into shape. We now define
4 B 8B 2 B
(A.30a)
 z
1 d
x= 
1 2 + 2
0
H = (z ) at = 0. (A.30b)   (A.36a)
1 1 2 z + 2 2 z2 + z
= log ,
Equations (A.30a), (A.30b) are independent of to leading order, and at 1

O( ) they depend on only through the last term (Hz + 12 BB ; H ) .
As above, since A.30a is independent of to leading order, we can con- which can be written implicitly as
clude that the derivatives H and Hz are no larger than O( ), and so
the last term is actually no larger than O( 2 ). Therefore H is independent z = sinh x (cosh x 1 ) . (A.36b)

98 Wilmott magazine
wilm003.qxd 7/26/02 7:05 PM Page 99

TECHNICAL ARTICLE 4

In terms of x, our problem is where z0 will be chosen later. Then through O( 2 ), we can simplify our
 equation to
1  ex
V ( t, f, a ) = [ f K ]+ + H(, x ) d, (A.37) 1
2 2
B (0 ) B ( z )e b1 z /4 Q = Qxx + 2 Q for > 0, (A.45a)
2 0 2
with  
1 1 1 3 Q = (x ) at = 0.
H = Hxx I ( z ) Hx + 2 2 b2 b21 H (A.45b)
2 2 4 8
(A.38a)
3 2
+ b1 H for > 0 The solution of A.45a, A.45b is clearly
4
H = (x ) at = 0. (A.38b)
1 1 1
e x e x
2 2 2
Here Q= /2
e
= /2
 3/2 (A.46)
2 2 1 2
2 +
 3
I( ) = 1 2 + 2 . (A.39)
through O( 2 ).
The final step is to define Q by This solution yields the option price
 1/4 1 
H = I 1/2 ( z (x ) ) Q = 1 2 z + 2 2 z 2 V(t, f, a ) = [ f K]+ + B(0 ) B( z )I1/2 ( z ) e 4 b 1 z
1 2 2
Q. (A.40)
 ex 2
1
(A.47)
e x /2 e d.
2 2
Then
0 2
1
Hx = I 1/2 ( z ) Q x + I ( z ) Q , (A.41a) Observe that this can be written as
2

1fK ex
1
  V ( t, f, a ) = [ f K ]+ + e x
2 2 2
/2
e e
d, (A.48a)
1 1 2 x 2
Hxx = I 1/2 ( z ) Qx x + I Q x + 2 2 I I + I I Q , (A.41b) 0
2 4
where
and so    1/2 
z  xI ( z ) 1
2 = log B(0 ) B( z ) + log + 2 b1 z 2
 ex fK z 4
1  (A.48b)
V ( t, f, a ) = [ f K ]+ +
1 2 2
B(0 ) B( z )I 1/2 ( z ) e 4 b1 z Q (, x ) d,
2 0
(A.42) Moreover, quite amazingly,

2 1 1
where Q is the solution of e
=  3/2 =  3/2 + O( ),
4
(A.48c)
1 2
3
2 1 2 2 x2
   
1 1 1 1 3 3
Q = Qxx + 2 2 I I I I Q + 2 2 b2 b21 Q + 2 b1 Q
2 4 8 4 8 4 through O( 2 ). This can be shown by expanding 2 through O( 2 ), and
noting that 2 /x2 matches /3. Therefore our option price is
(A.43a)
for > 0, with 
1fK ex
1 d
V ( t, f, a ) = [ f K ]+ + e x
2 2

Q = (x ) at = 0. (A.43b) /2
e  3/2 , (A.49)
2 x 0 2 1 2
2
x2
As above, we can replace I ( z ), I ( z ), I ( z ) by the constants and changing integration variables to
I ( z0 ), I ( z0 ), I ( z0 ) , and commit only O( ) errors. Define the con-
x2
stant by q= , (A.50)
  2
1 1
2
= 2
I ( z0 ) I( z0 ) I ( z0 )
4 8 reduces this to
  (A.44) 
1 3 3 |f K|
e q+
2

+ 2 b2 b21 + b1 , V ( t, f, a ) = [ f K ]+ +  3/2 dq. (A.51)
4 8 4
^

4 x2
2 ex q 2

Wilmott magazine 99
wilm003.qxd 7/26/02 7:05 PM Page 100


That is, the value of a European call option is given by fK
N = 1 + 2 2 ex + (A.56)
x x

|f K| e q through O( 2 ).
V ( t, f, a ) = [ f K ]+ + dq, (A.52a)
4 x2
2 ex
2 q3/2 Before continuing to the implied log normal volatility, let us seek the
simplest possible way to re-write this answer which is correct through
with O( 2 ). Since x = z [1 + O( )] , we can re-write the answer as
  
    fK z  
z  xI 1/2 ( z ) 1 2 N = 1 + 2 (1 + 2 + 3 ) ex + , (A.57a)
2 = log B(0 ) B( z ) + log + b1 z 2 , z x(z )
fK z 4
(A.52b)
where
through O( ). 2   1
A.2. Equivalent normal volatility. Equations (A.52a) and (A.52a) are a fK ( f K ) 1 f
df
=  f df = .
formula for the dollar price of the call option under the SABR model. z
fK K C(f )
K C( f )
The utility and beauty of this formula is not overwhelmingly apparent.
To obtain a useful formula, we convert this dollar price into the equiva- This factor represents the average difficulty in diffusing from todays for-
lent implied volatilities. We first obtain the implied normal volatility, and ward f to the strike K, and would be present even if the volatility were
then the standard log normal (Black) volatility. not stochastic.
Suppose we repeated the above analysis for the ordinary normal model The next factor is
z
dF = N dW , F(0 ) = f. (A.53a) =  , (A.57b)
x(z ) 12 + 2 +
log 1
where the normal volatily N is constant, not stochastic. (This model is where
also called the absolute or Gaussian model). We would find that the option 
value for the normal model is exactly f
df fK  
= z = = 1 + O( 2 ) . (A.57c)

C (f ) C ( fa )
| f K | e q K
V ( t, f ) = [ f K ]+ + dq (A.53b) 
4 ( f K ) 2
2
q3/2 Here fav = f K is the geometric average of f and K. (The arithmetic aver-
2 ex
N
age could have been used equally well at this order of accuracy). This fac-
This can be seen by setting C (f ) to 1, setting to N and setting to 0 in tor represents the main effect of the stochastic volatility.
A.52a, A.52b. Working out this integral then yields the exact European The coefficients 1 , 2 , and 3 provide relatively minor corrections.
option price Through O( 2 ) these corrections are
   
fK fK  
V (t, f ) = ( f K ) N + N ex G , (A.54a) 1 z 
N ex N ex 1 = 2 log
2
C ( f ) C (K )
z fK
for the normal model, where N is the normal distribution and G is the (A.57d)
22 12 2 2 2
Gaussian density = C ( fav ) +
24
1
G(q ) = e q /2 .
2
(A.54b)
1 x 
 2 3 2 2 2
2 2 2 = z + 2 2 2 1/4
z = +
2
log 1 2 (A.57e)
z z 24
From A.53b it is clear that the option price under the normal model
matches the option price under the SABR model A.52a, A.52a if and only
1 2 B ( z0 ) 1
if we choose the normal volatility N to be 2 3 = = 2 1 C( fav ) + (A.57f)

4 B( z0 ) 4
1 x2 2
= 1 2 ex . (A.55) where
N2 (f K )2 x2
C (fav ) C ( fav )
Taking the square root now shows the options implied normal (absolute) 1 = , 2 = . (A.57g)
volatility is given by C(fav ) C( fav )

100 Wilmott magazine


wilm003.qxd 7/26/02 7:05 PM Page 101

TECHNICAL ARTICLE 4

Let us briefly summarize before continuing. Under the normal model, dF = B FdW
, F (0 ) = f, (A.61)
the value of a European call option with strike K and exercise date ex is
given by (A.54a), (A.54b). For the SABR model, where we have written the volatility as B to stay consistent with the
preceding analysis. For Blacks model, the value of a European call with
dF = C(F ) dW 1 , F(0 ) = f (A.58a) strike K and exercise date ex is

d = dW 2 , (0 ) = (A.58b) Vcall = f N (d1 ) K N (d2 ), (A.62a)

dW 1 dW 2 = dt, (A.58c) Vput = Vcall + D (tset )[K f ], (A.62b)


the value of the call option is given by the same formula, at least
through O( 2 ), provided we use the implied normal volatility with

  log f/K 12 2 B2 ex
( f K ) d1,2 = , (A.62c)
N (K ) =  f df B ex
x( )
K C (f )
where we are omitting the overall factor D (tset ) as before.
22 12 2 2 1 We can obtain the implied normal volatility for Blacks model by
1+ C ( fav ) + 1 C ( fav ) (A.59a)
24 4 repeating the preceding analysis for the SABR model with C ( f ) = f and

2 3 2 2
2
= 0. Setting C ( f ) = f and = 0 in (A.59a A.59c) shows that the nor-
+ ex + .
24 mal volatility is

Here B ( f K ) 1 2 2
N (K ) = 1 B ex + . (A.63)
log f/K 24

C ( fav ) C ( fav )
fav = f K, 1 = , 2 = , (A.59b)
C ( fav ) C ( fav ) through O( 2 ) . Indeed, in [14] it is shown that the implied normal
  volatility for Blacks model is
fK 1 2 + 2 +
= ,
x( ) = log . (A. 59c)  1+ 1
log 2 f/K + 1920
1
log 4 f/K +
C ( fav ) 1 N (K ) = B f K  24

1+ 1
24
1 1
120
log 2 f/K 2 B2 ex + 5760
1
4 B4 ex2 +

The first two factors provide the dominant behavior, with the remaining (A.64)
factor 1 + [ ] 2 ex usually provideing corrections of around 1% or so.
through O ( 4 ). We can find the implied Black vol for the SABR model by
One can repeat the analysis for a European put option, or simply use
setting N obtained from Blacks model in equation (A.63) equal to N
call/put parity. This shows that the value of the put option under the
obtained from the SABR model in (A.59a B.59c). Through O( 2 ) this
SABR model is
yields
     
Kf Kf log f/K
Vput ( f, , K ) = (K f ) N + N ex G (A.60) B (K ) =  f df
N ex N ex
x( )
K C (f )


where the implied normal volatility N is given by the same formulas 22 12 + 1/fav2 2 2 1 (A.65)
1+ C (fav ) + 1 C (fav )
(A.59a A.59c) as the call. 24 4
We can revert to the original units by replacing , 
everywhere in the above formulas; this is equivalent to setting to 1 2 3 2 2 2
+ ex +
everywhere. 24
A. 3. Equivalent Black vol. With the exception of JPY traders, most
traders prefer to quote prices in terms of Black (log normal) volatilities, This is the main result of this article. As before, the implied log normal
rather than normal volatilities. To derive the implied Black volatility, volatility for puts is the same as for calls, and this formula can be re-cast
consider Blacks model in terms of the original variables by simpley setting to 1.
^

Wilmott magazine 101


wilm003.qxd 7/26/02 7:05 PM Page 102

A.4. Stochastic model. As originally stated, the SABR model con- where
sists of the special case C ( f ) = f :
 
1 2 + 2 +

dF = F dW 1 , F(0 ) = f (A.66a) = ( f K)(1 ) /2 log f/K, x ( ) = log .
1
(A.69b)
d = dW 2 , (0 ) = (A.66b)
This is the formula we use in pricing European calls and puts.
dW 1 dW 2 = dt. (A.66c) To obtain the implied Black volatility, we equate the implied normal
volatility N (K ) for the SABR model obtained in (A.69a A.69b) to the
Making this substitution in (A.58aA.58b) shows that the implied normal implied normal volatility for Blacks model obtained in (A.63). This
volatility for this model is shows that the implied Black volatility for the SABR model is

   
(1 )( f K ) 1
N (K ) = B (K ) =
f 1 K 1 x( ) ( f K )(1 ) /2 1 + (1 )2
log 2 f/K + (1 )4
log 4 f/K + x( )
24 1920
    
(2 ) 2 2 3 2 2 2 (1 )2 2 2 3 2 2 2
1+ + 1 + ex + 1+ + + ex + ,
24fav
22
4fav 24 24( f K )1 4( f K )(1 ) /2 24
(A.67a) (A.69c)

 through O( 2 ), where and x( ) are given by (A.69b) as before. Apart


through O( ), where fav =
2
f K as before and from setting to 1 to recover the original units, this is the formula quot-
ed in section 2, and fitted to the market in section 3.
  A.5. Special cases Two special cases are worthy of special treatment:
fK 1 2 + 2 + the stochastic normal model ( = 0) and the stochastic log normal
= , x( ) = log . (A.67b)
fav 1 model ( = 1). Both these models are simple enough that the expansion
can be continued through O( 4 ). For the stochastic normal model ( = 0)
the implied volatilities of European calls and puts are
We can simplify this formula by expanding 
2 3 2 2 2
N (K ) = 1 + ex + (A.70a)
24

 1 1
fK= f K log f/K 1 + log 2 f/K + log 4 f/K + , (A68a)
24 1920
  2 
log f/K 2 3 2 2 2
B (K ) = 1+ + ex +
(1 ) /2 fK x( ) 24f K 24
f 1 K 1 = (1 )(f K ) log f/K
(A.68b) (A.70b)
(1 )2 (1 )4
1+ log 2 f/K + log 4 f/K +
24 1920
through O( 4 ), where
 
 1 2 + 2 +
and neglecting terms higher than fourth order. This expansion reduces = f K log f/K, x( ) = log . (A.70c)
1
the implied normal volatility to
 
/2 1+ 1
log 2 f/K + 1920
1
log 4 f/K + For the stochastic log normal model ( = 1 ) the implied volatilities are
N (K ) = ( f K ) 24
 
1+ (1 ) 2
log 2 f/K + (1 )4
log 4 f/K + x( ) fK 1 1
24 1920 N (K ) = 1 + 2 +
 log f/K x( ) 24 4
(2 ) 2 2 3 2 2 2  (A.71a)
1+ + + ex + , 1
24( f K )1 4( f K )(1 ) /2 24 + (2 3 ) ex +
2 2 2
(A.69a) 24

102 Wilmott magazine


wilm003.qxd 7/26/02 7:05 PM Page 103

TECHNICAL ARTICLE 4

   P = 2 (f K ), for t = T. (B.4b)
1 1
B (K ) = 1+ + (2 3 2 ) 2 2 ex +
x( ) 4 24
We eliminate (t ) by defining the new time variable
(A.71b)  t  T  tex
s= 2 (t ) dt , s = 2 (t ) dt , sex = 2 (t ) dt . (B.5)
4 0 0 0
through O( ), where
  Then the option price becomes
1 2 + 2 +
= log f/K, x( ) = log . (A.71c) 
1 1 2 2 sex
V ( t, f, a ) = [ f K ]+ + C (K ) P (s, f, ; s , K ) ds , (B.6)
2 s

Appendix B. Analysis of the Dynamic where P(s, f, ; s , K ) solves the forward problem

SABR Model. Ps +
1 2 2 2 
C ( f ) Pff + 2(s ) 2 C ( f ) Pf + 2 (s ) 2 P = 0 for s < s
2
We use effective medium theory [23] to extend the preceding analysis to the (B.7a)
dynamic SABR model. As before, we take the volatility (t ) and volvol
(t ) to be small, writng (t ) (t ), and (t ) (t ) , and analyze P = 2 ( f K ), for s = s . (B.7b)

dF = (t ) C(F ) dW 1 , (B.1a) Here

d = (t ) dW 2 , (B.1b) (s ) = (t ) (t ) / (t ), (s ) = (t ) / (t ) . (B.8)

with We solve this problem by using an effective media strategy [23]. In this
strategy our objective is to determine which constant values and
dW 1 dW 2 = (t ) dt, (B.1c)
yield the same option price as the the time dependent coefficients (s )
in the limit  1. We obtain the prices of European options, and from and (s ). If we could find these constant values, this would reduce the
these prices we obtain the implied volatity of these options. After obtain- problem to the non-dynamic SABR model solved in Appendix B.
ing the results, we replace (t ) (t ) .and (t ) (t ) to get the We carry out this strategy by applying the same series of time-independ-
answer in terms of the original variables. ent transformations that was used to solve the non-dynamic SABR model
Suppose the economy is in state F (t ) = f, (t ) = at date t. Let in Appendix B, defining the transformations in terms of the (as yet
V (t, f, ) be the value of, say, a European call option with strike K and exer- unknown) constants and . The resulting problem is relatively complex,
cise date tex . As before, define the transition density p(t, f, ; T, F, A ) by more complex than the canonical problem obtained in Appedix B. We use
 a regular perturbation expansion to solve this problem, and once we have
p (t, f, ; T, F, A ) dF dA prob F < F(T ) < F + dF, A < (T ) solved this problem, we choose and so that all terms arising from the
 (B.2a) time dependence of (t ) and (t ) cancel out. As we shall see, this simul-
< A + dA | F(t ) = f, (t ) = taneously determines the effective parameters and allows us to use the
and define analysis in Appendix B to obtain the implied volatility of the option.
 B.1 Transformation. As in Appendix B, we change independent vari-
P(t, f, ; T, K ) = A2 p(t, f, ; T, K, A ) dA. (B.2b) ables to
 f
1 df
Repeating the analysis in Appendix B through equation (A.10a), (A.10b) z= , (B.9a)
K C( f )
now shows that the option price is given by
 tex and define
1
V ( t, f, a ) = [ f K ]+ + 2 C 2 ( K ) 2 (T ) P(t, f, ; T, K ) dT, (B.3)
2 t B( z ) = C( f ) . (B.9b)
where P(t, f, ; T, K ) is the solution of the backwards problem
We then change dependent variables from P to P, and then to H:
1 2 2 2 2 
Pt + C (f ) Pff + 2 2 C( f ) Pf + 2 2 P = 0, for t < T
^

2
(B.4a) P = C(K ) P, (B.9c)

Wilmott magazine 103


wilm003.qxd 7/26/02 7:05 PM Page 104

 
H= C(K ) /C( f )P B(0 ) /B( z )P. (B.9d) change in price due to 12 2 b1 zHz term. In this way to the transforma-
tion cancels out the zHz term on average.
Following the reasoning in Appendix B, we obtain In a similar vein we define
 sex
1  
V (t, f, a ) = [ f K ]+ + B(0 ) B( z ) H(s, z, ; s ) ds , (B.10) I( z ) = 1 2 z + 2 2 z 2 , (B.16a)
2 s

where H(s, z, ; s ) is the solution of and


1 1 B
  
Hs + 1 2 z + 2 2 z 2 Hzz 2 (zHz H ) 1 z
d 1 1 2 z + 2 2 z2 / + z
2
 
2 B x= = log ,
1 B 3 B 2
(B.11a) 0 I( ) 1 /
+
2 2
H=0 (B.16b)
4 B 8 B2
for s < s , and where the constants and will be chosen later. This yields

H = (z ) at s = s (B.11b)
1 1 2 z + 2 2 z 2 1
Hs + (Hxx I ( z ) Hx ) 2 b1 ( ) xHx
2 1 2 z + 2 2 z 2 2
through O( 2 ). See (A.29), (A.31a), and (A.31b). There are no derivatives  
1 1 3
in equations (B.11a), (B.11b), so we can treat as a parameter instead of a + 2 b1 (2 + ) H + 2 2 b2 b21 H = 0 for s < s ,
variable. Through O( 2 ) we can also treat B /B and B /B as constants: 4 4 8
(B.17a)
B ( z0 ) B ( z0 )
b1 , b2 , (B.12)
B( z0 ) B( z0 )
H = (x ) at s = s , (B.17b)

where z0 will be chosen later. Thus we must solve through O( 2 ). Here we used z = x + and zHz = xHx + to leading
order to simplify the results. Finally, we define
1 1
Hs + 1 2 z + 2 2 z 2 Hzz 2 b1 (zHz H )
2 2
  (B.13a) H = I 1/2 ( z ) Q. (B.18)
1 3
+ 2 2 b2 b21 H = 0 for s < s ,
4 8
Then the price of our call option is
H = (z ) at s = s . (B.13b)
1 
V ( t, f, a ) = [ f K ]+ + B(0 ) B( z )I 1/2
2
At this point we would like to use a time-independent transformation  sex (B.19)
Q(s, x; s ) ds ,
1 2 2
to remove the zHz term from equation (B.13a). It is not possible to cancel ( z ) e 4 b1 z
s
this term exactly, since the coefficient (s ) is time dependent. Instead we

use the transformation where Q(s, x; s ) is the solution of

H,
1 2
b1 z2
H = e4 (B.14)
1 1 2 z + 2 2 z 2 1 1
Qs + Qxx 2 b1 ( ) xQx + 2 b1 (2 + ) Q
2 1 2 z + 2 2 z 2 2 4
where the constant will be chosen later. This transformation yields    
1 1 1 3
1 1 + 2 2 I I I I Q + 2 2 b2 b21 Q = 0 for s < s ,
Hs + 1 2 z + 2 2 z 2 Hzz 2 b1 ( ) zHz 4 8 4 8 (B.20a)
2  2 
1 1 3
+ 2 b1 (2 + ) H + 2 2 b2 b21 H = 0 for s < s ,
4 4 8 Q = (x ) at s = s , (B.20b)

H = (z ) at s = s , (B.15a)
Using
through O( 2 ). Later the constant will be selected so that the change in 1
the option price caused by the term 12 2 b1 zHz is exactly offset by the z=x x2 + , (B.21)
2

104 Wilmott magazine


wilm003.qxd 7/26/02 7:05 PM Page 105

TECHNICAL ARTICLE 4

we can simplify this to where


1 1 
1 (2a ) 1 
Qs + Q xx = ( ) xQ xx 2 2 2 3( ) x2 Q xx Q (2
s
a)
+ Q xx = b1 ( ) xQ (0
x Q
) (0 )
for s < s , (B.28a)
2 2 2 2
1
+ 2 b1 ( )(xQ x Q )
2  
3 1 1 Q (2a ) = 0 at s = s , (B.28b)
2 b1 Q 2 2 I I I I Q
4 4 8
  and where
1 3 (B.22a)
2 2 b2 b21 Q for s < s , 1 (2b ) 1 2
4 8 Q (2
s
b)
+ Q xx = ( ) xQ (1
xx
)
2
2 2 (B.29a)
3( )] x2 Q (0
xx
)
for s < s ,
Q = (x ) at s = s , (B.22b)
Q (2b ) = 0 at s = s . (B.29b)
2
through O( ). Note that I, I , and I can be replaced by the constants
I( z0 ), I ( z0 ) , and I ( z0 ) through O( 2 ). Once we have solved these equations, then the option price is then given by
B. 2. Perturbation expansion. Suppose we were to expand Q(s, x ; s )
1 
V ( t, f, a ) = [ f K ]+ +
1 2 2
as a power series in : B(0 ) B( z )I 1/2 ( z ) e 4 ab 1 z J, (B.30a)
2
Q(s, x ; s ) = Q (0 ) (s, x ; s ) + Q (1 ) (s, x ; s ) + 2 Q (2 ) (s, x ; s ) + . (B.23) where
 sex  sex  sex
J= Q (0 ) (s, x ; s ) ds + Q (1 ) (s, x ; s ) ds + 2 Q (2s ) (s, x ; s ) ds
Substituting this expansion into (B.22a), (B.22b) yields the following hier- s s s
archy of equations. To leading order we have  
sex sex
1
+ 2 Q (2a ) (s, x ; s ) ds + 2 Q (2b ) (s, x ; s ) ds + .
Q (0
s
)
+ Q (0 )
=0 fors < s , (B.24a) s s (B.30b)
2 xx
The terms Q (1 ) , Q (2a ) , and Q (2b ) arise from the time-dependence of the

Q (0 )
= (x ) at s = s . (B.24a) coefficients (s ) and (s ). Indeed, if (s ) and (s ) were constant in time,
we would have Q (1 ) Q (2a ) Q (2b ) 0 , and the solution would be just
At O( ) we have Q (s ) Q (0 ) + 2 Q (2s ) . Therefore, we will first solve for Q (1 ) , Q (2a ) ,and Q (2b ) ,
and then try to choose the constants , , and so that the last three inte-
1 (1 )
Q (1
s
)
+ Q = ( ) x Q (0 )
for s < s , (B.25a) grals are zero for all x. In this case, the option price woud be given by
2 xx xx

1 
V(t, f, a ) = [ f K ]+ + B(0 ) B( z )I 1/2 ( z )
Q (1 ) = 0 at s = s . 2
(B.25b)  sex (B.31a)
Q(s ) (s, x ; s ) ds ,
1 2 2
e4 ab 1 z

At O( 2 ) we can break the solution into s

Q (2 ) = Q (2s ) + Q (2a ) + Q (2b ) , (B.26) and, through O( 2 ), Q (s ) would be the solution of the static problem
 
1 (s ) 3 2 1 1
where Q s + Q xx = ab1 Q
(s ) (s ) 2 2
I I I I Q (s )
2 4 4 8
    (B.31b)
1 (2s ) 3 1 1 1 3
Qs(2s ) + Q xx = ab1 Q
(0 ) 2
I I I I Q (0 ) 2 2 b2 b21 Q (s ) for s < s ,
2 4 4 8 4 8
 
1 3
2 b2 b21 Q (0 ) for s < s , Q (s ) = (x ) at s = s . (B.31c)
4 8
(B.27a) This is exactly the time-independent problem solved in Appendix B. See
equations (A.42), (A.43a), and (A.43b). So if we can carry out this strategy,
we can obtain option prices for the dynamic SABR model by reducing
^

Q (2s ) = 0 at s = s , (B.27b) them to the previously-obtained prices for the static model.

Wilmott magazine 105


wilm003.qxd 7/26/02 7:05 PM Page 106

B.2.1. Leading order analysis The solution of (B.24a), (B.24b) is 1 (2a ) 1 x2 +


Q (2
s +
Q xx = b1 ( )
a)
G
Gaussian: 2 2 (B.38)
1
= b1 ( ) Gx x b1 ( ) G
Q (0 ) = G(x/ ) (B.32a) 2
for s < s , with Q (2a ) = 0 at s = s . Solving then yields
where   s 


Q (2 a )
= b1 (s s)[(s) ]dsG(x/ s s ) .
(B.39)
1 s
e x /2 , = s s.
2
G (x/ ) = (B.32b) s
2 This term makes a contribution of
For future reference, note that
  
se x se x
x x2 x3 3 x Q (2 a )
(s, x; s ) ds = b1
(se x s)[(s) ]ds G(x/ sex s ) (B.40)
Gx = G ; G xx = G; G xxx = G; (B.33a) s s
2 3
to the option price, so we choose
x4 6 x2 + 3 2 x5 10 x3 + 15 2 x  sex
G xxxx = G ; G xxxxx = G, (B.33b)
4 5 (sex s)[(s) ]ds
= = s . (B.41)
1
(s s)2
2 ex
x6 15 x4 + 45 2 x2 15 3
G xxxxxx = G. (B.33c) to eliminate this contribution.
6
B.2.4. The 2 Q (2b ) term Substituting Q (1 ) and Q (0 ) into equation
B.2.2. Order Substituting Q (0 ) into the equation for Q (1 ) and using (B.29a), we obtain
(B.33a) yields
1 (2b ) 1
Q (2
s
b)
+ Qx x = ( ) AxG x x x x x + 2( ) As xGx x x x2 Gx x , (B.42a)
1 (1 ) x x 3 2 2
Q (1
s
)
+ Q = ( ) G
2 xx 2 for s < s , where
= (s s )( ) G xxx 2( ) Gx for s < s ,
(B.34) = 2 (s ) 2 3[(s ) ]. (B.42b)

This can be re-written as


with the initial condition Q (1 ) = 0 at s = s . The solution is
1 (2b )
Q (2
s
b)
+ Q = ( ) A[ Gxxxxxx + 5Gxxxx ] 2( ) As [ Gxxxx
Q (1 )
= A(s, s ) Gxxx + 2As (s, s ) Gx
(B.35a) 2 xx
1
+ 3Gxx ] [ 2 Gxxxx + 5 Gxx + 2G] (B.43)
2


= 2 A ( s , s ) G x ( x / s s , (B.35b) Solving this with the initial condition Q (2b ) = 0 at s = s yields
s
where 1 2
Q(2b ) = A (s, s ) Gxxxxxx + 2A(s, s ) As (s, s ) Gxxxx
  2
s s  s
A (s, s ) = (s s)[(s) ] ds; A s (s, s ) = [(s) ] ds. (B.35c)
s s
+3 [(s) ] A(s, s ) dsGxxxx + 3A2s (s, s ) Gxx
s
 s  s  s
This term contributes 1 5
 se x + [s s]2 (s) dsGxxxx + [s s](s) dsGxx + (s) dsG.
2 s 2 s s
Q (1 ) (s, x; s ) ds = 2A(s, sex ) Gx (x/ sex s ) (B.36)
s

to the option price. See equations (B.30a), (B.30b). To eliminate this con- This can be written as
tribution, we chose so that A(s, se x ) = 0 :   s

 sex Q (2 b )
= 4A2 (s, s ) Gss 12 [(s) ] A(s, s ) dsGs
(sex s) (s) ds s s
= s
. (B.37)  s  s  (B.45)
1
(s s )2
2 ex 2

2 (s s) (s) dsGs + (s s) (s) dsG
s s
B.2.3. The 2 Q (2a ) term From equation (B.28a) we obtain

106 Wilmott magazine


wilm003.qxd 7/26/02 7:05 PM Page 107

TECHNICAL ARTICLE 4

Recall that was chosen above so that A(s, sex ) = 0 . Therefore the contri- and where we have use = . We chose the effective parameters and
bution of Q (2b ) to the option price is so that the integrals of Q (1 ) , Q (2a ) contribute nothing to J. The integral
of Q (2b ) then contributed 12 2 (sex s )2 Q (0 ) (s, x; sex ) . The option price is
 sex   sex
Q (2b ) (s, x; s ) ds = 12 [(s) ]A(s, sex ) ds  sex
s s J= {Q (0 ) (s, x ; s ) + 2 Q (2s ) (s, x ; s )} ds
  s
sex 2 1
+2 (sex s) (s) ds Gs (x/ sex s ) (B.46) + . 2 (sex s)2 Q (0 ) (s, x ; sex ) + (B.50a)
s 2
   sex
sex
+ (sex s) (s) ds G(x/ sex s ), = {Q (0 ) (s, x ; s ) + 2 Q (2s ) (s, x ; s )} ds +
s s

where = 2 (s ) 2 3[(s ) ] . through O( 2 ), where


We can choose the remaining effective media parameter to set
1 2 2
either the coefficient of Gs (x/ se x s ) or the coefficient of G(x/ se x s ) sex = sex + (sex s ) + (B.50b)
2
to zero, but cannot set both to zero to completely eliminate the contri-
bution of the term Q(2b ) . We choose to set the coefficient of Through O( 2 ) we can combine Q (s ) = Q (0 ) (s, x ; s ) + 2 Q (2s ) (s, x ; s ) ,

Gs (x/ se x s ) to zero, for reasons that will become apparent in a where Q (s ) solves the static problem
moment:  
1 3 1 1
 sex  sex Q (s s ) + Q (xxs ) = 2 ab1 Q (s ) 2 2 I I I I
1 2 2 2 4 4 8
= 1
2
(sex s) (s) ds 3
2
(sex s)[(s) ]ds  
( s s) 3
s s 1 3
3 ex
Q( s ) 2 2 b2 b21 Q(s ) for s < s ,
 se x  s1 4 8
6 s2 [(s1 ) ][(s2 ) ]ds2 ds1 . (B.51a)
s s (B.47)
Q(s ) = (s s ) at s = s , (B.51b)
Then the remaining contribution to the option price is
This problem is homogeneous in the time s, so its solution Q (s ) depends
 only on the time difference = s s. The option price is therefore
sex
1
Q (2b ) (s, x; s ) ds = (sex s)2 G(x/ sex s )
s 2
(B.48a)
1 1 
= (sex s)2 Q (0 ) (s, x ; sex ), V ( t, f, a ) = [ f K ]+ + B(0 ) B( z )I 1/2
2 2
 sex s
( z ) e 4 ab 1 z
1 2 2

where Q (s ) (, x ) d, (B.52)
0
 sex where Q s (, x ) is the solution of
1
= (sex s)[ 2 (s) 2 ]ds. (B.48b)  
1
(s
2 ex
s )2 s 1 (s ) 3 2 1 1
Q (s ) Q xx = ab1 Q +
(s ) 2 2
I I I I Q (s )
s 2 4 4 8
Here we have used s ex (sex s)((s) ) ds = 0 to simplify (B.48b).   (B.53a)
1 3
B.3. Equivalent volatilities. We can now determine the implied volatili- + 2 2 b2 b21 Q (s ) for > 0,
4 8
ty for the dynamic model by mapping the problem back to the static model
of Appendix B. Recall from (B.30a), (B.30b) that the value of the option is Q s = (x ) at = 0. (B.53b)
1 
V ( t, f, a ) = [ f K ]+ +
1 2 2
B(0 ) B( z )I 1/2 ( z ) e 4 ab 1 z J, (B.49a) The option price defined by (B.52), (B.53a), and (B.53b) is identical to
2
the static models option price defined by (A.42), (A.43a), and (A.43b), pro-
where vided we make the identifications
 sex  sex  sex
J= Q (0 ) (s, x; s ) ds + Q (1 ) (s, x; s ) ds + 2 Q (2s ) (s, x; s ) ds , / (B.54)
s s s
 sex  sex
^

+ 2 Q (2a ) (s, x; s ) ds + 2 Q (2b ) (s, x; s ) ds + , (B.49b) ex sex s = sex s + 12 2 (sex s)2 (B.55)
s s

Wilmott magazine 107


wilm003.qxd 7/26/02 7:05 PM Page 108

in Appendix B for the original non-dynamic SABR model, provided we REFERENCES


make the identifications
D. T. Breeden and R. H. Litzenberger, (1994), Prices of state-contingent claims implicit

in option prices, J. Business, 51 pp. 621651.
ex = + 2 [ 2 ( ) 2 ]d , (B.56a) Bruno Dupire, (1994), Pricing with a smile, Risk, Jan. pp. 1820.
0
Bruno Dupire, (1997), Pricing and hedging with
smiles, in Mathematics of Derivative Securities, M.A. H. Dempster and S. R. Pliska, eds.,
/, . (B.56b) Cambridge University Press, Cambridge, pp. 103111.
E. Derman and I. Kani, (1994), Riding on a smile, Risk, Feb. pp. 3239.
See equations (A.42 A.43b). Following the reasoning in the preceding E. Derman and I. Kani, (1998), Stochastic implied trees: Arbitrage pricing with stochas-
Appendix now shows that the European call price is given by the formula tic term and strike structure of volatility, Int J. Theor Appl Finance, 1 pp. 61110.
    J.M. Harrison and D. Kreps, (1979), Martingales and arbitrage in multiperiod securities
fK fK markets, J. Econ. Theory, 20 pp. 381408
V ( t, f, K ) = ( f K ) N + N ex G , (B.57)
N ex N ex J.M. Harrison and S. Pliska, (1981), Martingales and stochastic integrals in the theory
of continuous trading,Stoch. Proc. And Their Appl., 11 pp. 215260
with the implied normal volatility I. Karatzas, J.P. Lehoczky, S.E. Shreve, and G.L. XU, (1991), Martingale and duality
  methods for utility maximization in an incomplete market, SIAM J. Control Optim, 29 pp.
(f K )
N (K ) =  f df 702730

x( ) J. Michael Steele, (2001) Stochastic Calculus and Financial Applications, Springer,
K C (f )
F. Jamshidean, (1997), Libor and swap market models and measures, Fin. Stoch. 1 pp.
22 12 2 2 (B.58a) 293330
1+ C (fav ) + 14
1 C (fav )
24 F. Black, (1976), The pricing of commodity contracts, Jour. Pol. Ec., 81 pp. 167179
J. C. Hull, (1997), Options, Futures and Other Derivatives, Prentice-Hall.
22 32 1
+ + 2 ex + . Paul Wilmott, (2000), Paul Wilmott on Quantitative Finance, John Wiley & Sons, .
24 2 Patrick S. Hagan and Diana E. Woodward, (1999),Equivalent Black volatilities, App.
where
Math. Finance, 6 pp. 147157.
  P.S. Hagan, A. Lesniewski, and D. Woodward, Geometrical optics in finance, in prepara-
f K 1 2 / + 2 / + tion
= , x( ) = log , (B.58b) Fred Wan, (1991), A Beginners Book of Modeling, Springer-Verlag, Berlin, New York, .
C( fav ) 1 /
J. Hull and A. White, (1987), The pricing of options on assets with stochastic
 C ( fav ) C ( fav ) volatilities, J. of Finance, 42 pp. 281300
fav = f K, 1 = , 2 = , (B.58c) S. L. Heston, (1993), A closed-form solution for options with stochastic volatility with
C( fav ) C( fav )
applications to bond and currency options, The Review of Financial Studies, 6 pp.
 327343
[ 2 ( ) 2 ]d
= 0
. (B.58d) A. Lewis, (2000), Option Valuation Under Stochastic Volatility, Financial Press,
1 2
2
J.P. Fouque, G. Papanicolaou, K.R. Sirclair, (2000), Derivatives in Financial Markets with
Stochastic Volatility, Cambridge Univ Press,
Equivalently, the option prices are given by Blacks formula with the
N. A. Berner, BNP Paribas, (2000), Private communication,
effective Black volatility of
J.D. Cole, (1968), Perturbation Methods in Applied Mathematics, Ginn-Blaisdell,
  J. Kevorkian and J.D. Cole, (1985), Perturbation Methods in Applied Mathematics,
log f /K
B (K ) =  f df Springer-Verlag,
x( )

K C (f ) J. F. Clouet, (1998), Diffusion Approximation of a Transport Process in Random Media,
SIAM J. Appl. Math, 58 pp. 16041621
22 12 + 1/fav 2
1+ 2 C 2 ( fav ) + 14 1 C (fav ) I. Karatzas and S. Shreve, (1988), Brownian Motion and Stochastic Calculus, Springer.
24 B. Okdendal, (1998), Stochastic Differential Equations, Springer
M. Musiela and M. Rutkowski, (1998), Martingale Methods in Financial Modelling,
2 2 3 2
+ + 12 2 ex + Springer
24
G.B. Whitham, (1974), Linear and Nonlinear Waves, Wiley
J.C. Neu, (1978), Doctoral Thesis, California Institute of Technology

108 Wilmott magazine

Das könnte Ihnen auch gefallen