102
Quandoquidem inter nos sanctissima divitiarum maiestas Since the majesty of wealth is most sacred with us Juvenal (55–120), Sat. 1, 113 20 Path Integrals and Financial Markets An important field of applications for path integrals are financial markets. The prices of assets fluctuate as a function of time and, if the number of participants in the market is large, the fluctuations are pretty much random. Then the time dependence of prices can be modeled by fluctuating paths. 20.1 Fluctuation Properties of Financial Assets Let S (t) denote the price of a stock or another financial asset. Over long time spans, i.e., if data recording frequency is low, the average over many stock prices has a time behavior that can be approximated by pieces of exponentials. This is why they are usually plotted on a logarithmic scale. This is best illustrated by a plot of the Dow-Jones industrial index over 60 years in Fig. 20.1. The fluctuations of the index have a certain average width called the volatility of the market. Over 1940 1960 1980 2000 Figure 20.1 Logarithmic plot of Dow Jones industrial index over 80 years. There are four roughly linear regimes, two of exponential growth, two of stagnation [1]. 1440

Economic and Path Integrals

Embed Size (px)

Citation preview

Page 1: Economic and Path Integrals

Quandoquidem inter nos sanctissima divitiarum maiestas

Since the majesty of wealth is most sacred with us

Juvenal (55–120), Sat. 1, 113

20

Path Integrals and Financial Markets

An important field of applications for path integrals are financial markets. Theprices of assets fluctuate as a function of time and, if the number of participantsin the market is large, the fluctuations are pretty much random. Then the timedependence of prices can be modeled by fluctuating paths.

20.1 Fluctuation Properties of Financial Assets

Let S(t) denote the price of a stock or another financial asset. Over long timespans, i.e., if data recording frequency is low, the average over many stock priceshas a time behavior that can be approximated by pieces of exponentials. This iswhy they are usually plotted on a logarithmic scale. This is best illustrated by aplot of the Dow-Jones industrial index over 60 years in Fig. 20.1. The fluctuationsof the index have a certain average width called the volatility of the market. Over

1940 1960 1980 2000

10000

100

200

500

1000

2000

5000

Figure 20.1 Logarithmic plot of Dow Jones industrial index over 80 years. There are

four roughly linear regimes, two of exponential growth, two of stagnation [1].

1440

Page 2: Economic and Path Integrals

20.1 Fluctuation Properties of Financial Assets 1441

1.0

5.0

a

b

100

500

S&P 500

Volatility σ × 103

1984 1986 1988 1990 1992 1994 1996

Figure 20.2 (a) Index S&P 500 for 13-year period Jan. 1, 1984 —Dec. 14, 1996, recorded

every minute, and (b) volatility in time intervals 30 min (from Ref. [2]).

long times, the volatility is not constant but changes stochastically, as illustrated bythe data of the S&P 500 index over the years 1984-1997, as shown in Fig. 20.2 [3].In particular, there are strong increases shortly before a market crash.

The theory to be developed will at first ignore these fluctuations and assumea constant volatility. Attempts to include them have been made in the literature[3]–[79] and a promising version will be described in Section 20.4.

The volatilities follow approximately a Gamma distribution, as illustrated inFig. 20.3.

0.000 0.001 0.002 0.0030.0

0.2

0.4

0.6

0.8

1.0

Pro

babil

ity d

istr

ibuti

on x

10

-3

Normalized volatility

Empirical volatility

Gauss distr. fit

Log-normal distr. fit

Gamma distr. fit

Figure 20.3 Comparison of best Gaussian, log-normal, and Gamma distribution fits

to volatilities over 300 min (from Ref. [80]). The normalized log-normal distribution has

the form Dlog−normal(z) = (2πσ2z2)−1/2e−(log z−µ)2/2σ2. The Gamma distribution will be

discussed further in Subsection 20.1.5.

Page 3: Economic and Path Integrals

1442 20 Path Integrals and Financial Markets

An individual stock will in general be more volatile than an averge market index,especially when the associated company is small and only few shares are traded perday.

20.1.1 Harmonic Approximation to Fluctuations

To lowest approximation, the stock price S(t) satisfies a stochastic differential equa-tion for exponential growth

S(t)

S(t)= rS + η(t), (20.1)

where rS is the growth rate, and η(t) is a white noise variable defined by the corre-lation functions

〈η(t)〉 = 0, 〈η(t)η(t′)〉 = σ2δ(t− t′). (20.2)

The standard deviation σ is a precise measure for the volatility of the stock price.The squared volatility v ≡ σ2 is called the variance.

The quantity dS(t)/S(t) is called the return of the asset. From financial data, thereturn is usually extracted for finite time intervals ∆t rather than the infinitesimaldt since prices S(t) are listed for certain discrete times tn = t0 + n∆t. There are,for instance, abundant tables of daily closing prices of the market S(tn), from whichone obtains the daily returns ∆S(tn)/S(tn) = [S(tn+1) − S(tn)]/S(tn). The set ofavailable S(tn) is called the time series of prices.

For a suitable choice of the time scales to be studied, the assumption of a whitenoise is fulfilled quite well by actual fluctuations of asset prices, as illustrated inFig. 20.4.

ω [sec−1]

S(ω)

Figure 20.4 Fluctuation spectrum of exchange rate DM/US$ as function of frequency

in units 1/sec, showing that the noise driving the stochastic differential equation (20.1) is

approximately white (from [13]).

Page 4: Economic and Path Integrals

20.1 Fluctuation Properties of Financial Assets 1443

For the logarithm of the stock or asset price1

x(t) ≡ log S(t) (20.3)

this implies a stochastic differential equation for linear growth [14, 15, 16, 17]

x(t) =S

S− 1

2σ2 = rx + η(t), (20.4)

where

rx ≡ rS − 1

2σ2 (20.5)

is the drift of the process [compare (18.405)]. A typical set of solutions of (20.4) isshown in Fig. 20.5.

x(t) ≡ logS(t)

t2 4 6 8 10

-1

1

2

3

4

5

6

Figure 20.5 Behavior of logarithmic stock price following the stochastic differential

equation (20.3).

The finite differences ∆x(tn) = x(tn+1)−x(tn) and the corresponding differentialsdx are called log-returns .

The extra term σ2/2 in (20.5) is due to Ito’s Lemma (18.413) for functions of astochastic variable x(t). Recall that the formal expansion in powers of dt:

dx(t) =dx

dSdS(t) +

1

2

d2x

dS2dS2(t) + . . .

=S(t)

S(t)dt− 1

2

[

S(t)

S(t)

]2

dt2 + . . . (20.6)

may be treated in the same way as the expansion (18.426) using the mnemonic rule(18.429), according to which we may substitute x2dt→ 〈x2〉dt = σ2, and thus

[

S(t)

S(t)

]2

dt→ x2(t)dt = σ2. (20.7)

The higher powers in dt do not contribute for Gaussian fluctuations since they carryhigher powers of dt. For the same reason the constant rates rS and rx in S(t)/S(t)and x(t) do not show up in [S(t)/S(t)]2 dt= x2(t)dt.

1To form the logarithm, the stock or asset price S(t) is assumed to be dimensionless, i.e. thenumeric value of the price in the relevant currency.

Page 5: Economic and Path Integrals

1444 20 Path Integrals and Financial Markets

In charts of stock prices, relation (20.5) implies that if we fit a straight linethrough a plot of the logarithms of the prices with slope rx, the stock price itselfgrows on the average like

〈S(t)〉 = S(0) erSt = S(0)〈erxt+∫ t

0dt′ η(t′)〉 = S(0) e(rx+σ2/2)t. (20.8)

This result is, of course, a direct consequence of Eq. (18.425).The description of the logarithms of the stock prices by Gaussian fluctuations

around a linear trend is only a rough approximation to the real stock prices. Thevolatilities depend on time. If observed at small time intervals, for instance everyminute or hour, they have distributions in which frequent events have an exponen-tial distribution [see Subsection 20.1.6]. Rare events, on the other hand, have amuch higher probability than in Gaussian distributions. The observed probabil-ity distributions possess heavy tails in comparison with the extremely light tails ofGaussian distributions. This was first noted by Pareto in the 19th century [18],reemphasized by Mandelbrot in the 1960s [19], and investigated recently by severalauthors [20, 22]. The theory needs therefore considerable refinement. As an inter-mediate generalization we shall introduce, beside the heavy power-like tails, alsothe so-called semi-heavy tails, which drop off faster than any power, such as e−xa

xb

with arbitrarily small a > 0 and any b > 0. We shall see later in Section 20.4 thatsemi-heavy tails of financial distribution may be viewed as a consequence of Gaus-sian fluctuations with fluctuating volatilities. Before we come to these we may fitthe data phenomenologically with various non-Gaussian distributions and explorethe consequences.

20.1.2 Levy Distributions

Following Pareto and Mandelbrot we may attempt to approximately fit the distri-butions of the price changes ∆Sn = S(tn+1)−S(tn), the returns ∆Sn/S(tn), and thelog-returns ∆xn = x(tn+1) − x(tn) for a certain time difference ∆t = tn+1 − tn withthe help of Levy distributions [19, 22, 24, 23]. For brevity we shall, from now on, usethe generic variable z to denote any of the above differences. The Levy distributionsare defined by the Fourier transform

Lλσ2(z) ≡

∫ ∞

−∞

dp

2πeipz Lλ

σ2(p), (20.9)

with

Lλσ2(p) ≡ exp

[

−(σ2p2)λ/2/2]

. (20.10)

For an arbitrary distribution D(z), we shall write the Fourier decomposition as

D(z) =∫

dp

2πeipzD(p) (20.11)

Page 6: Economic and Path Integrals

20.1 Fluctuation Properties of Financial Assets 1445

and the Fourier components D(p) as an exponential

D(p) ≡ e−H(p), (20.12)

where H(p) plays a similar role as the Hamiltonian in quantum statistical pathintegrals. By analogy with this we shall also define H(z) so that

D(z) = e−H(z). (20.13)

An equivalent definition of the Hamiltonian is

z/σ

1/x1+λ

z/σ

P (z)

P (z)1 + λ ≈ 2.7

1 + λ ≈ 4

Figure 20.6 Left: Levy tails of the S&P 500 index (1 minute log-returns) plotted

against z/δ. Right: Double-logarithmic plot exhibiting power-like tail regions of the S&P

500 index (1 minute log-returns) (after Ref. [23])

e−H(p) ≡ 〈e−ipz〉. (20.14)

For the Levy distributions (20.9), the Hamiltonian is

H(p) =1

2(σ2p2)λ/2. (20.15)

The Gaussian distribution is recovered in the limit λ → 2 where the Hamiltoniansimply becomes σ2p2/2.

For large z, the Levy distribution (20.9) falls off with the characteristic power-law

Lλσ2(z) → Aλ

σ2

λ

|z|1+λ. (20.16)

This power falloff is the heavy tail of the distribution discussed above. It is alsocalled power tail , Paretian tail , or Levy tail). The size of the tails is found byapproximating the integral (20.9) for large z, where only small momenta contribute,as follows:

Lλσ2(z) ≈

∫ ∞

−∞

dp

2πeipz

[

1 − 1

2(σ2p2)λ/2

]

→z→∞

Aλσ2

λ

|z|1+λ, (20.17)

Page 7: Economic and Path Integrals

1446 20 Path Integrals and Financial Markets

with

Aλσ2 = −σ

λ

∫ ∞

0

dp′

πp′λ cos p′ =

σλ

2πλsin(πλ/2) Γ(1 + λ). (20.18)

The stock market data are fitted best with λ between 1.2 and 1.5 [13], and weshall use λ = 3/2 most of the time, for simplicity, where one has

A3/2σ2 =

1

4

σ3/2

√2π. (20.19)

The full Taylor expansion of the Fourier transform (20.10) yields the asymptoticseries

Lλσ2(z) =

∞∑

n=0

(−1)n

n!

∫ ∞

0

dp

π

σλnpλn

2ncos pz =

∞∑

n=0

(−1)n+1

n!

σλn

2nπΓ(1 + nλ)

sin πλ2

|z|1+λ. (20.20)

This series is not useful for practical calculations since it diverges. In particular, itis unable to reproduce the pure Gaussian distribution in the limit λ→ 2.

There also exists an asymmetric Levy distribution whose Hamiltonian is

Hλ,σ,β(p) =1

2|σp|λ [1 − iβǫ(p)Fλ,σ(p)], (20.21)

where ǫ(p) is the step function (1.316), and

Fλ,σ(p) =

tan(πλ/2) for λ 6= 1,−(1/π) log p2 for λ = 1.

(20.22)

The large-|z| behavior of this distribution is again given by (20.17), except that theprefactor (20.18) is multiplied by a factor (1 + β).

20.1.3 Truncated Levy Distributions

Mathematically, an undesirable property of the Levy distributions is that their fluc-tuation width diverges for λ < 2, since the second moment

σ2 = 〈z2〉 ≡∫ ∞

−∞dz z2 Lλ

σ2(z) = − d2

dp2Lλσ2(p)

p=0

(20.23)

is infinite. If one wants to describe data which show heavy tails for large log-returnsbut have finite widths one must make them fall off at least with semi-heavy tailsat very large returns. Examples are the so-called truncated Levy distributions [22].They are defined by

L(λ,α)σ2 (z) ≡

∫ ∞

−∞

dp

2πeipz L

(λ,α)σ2 (p) =

∫ ∞

−∞

dp

2πeipz−H(p) , (20.24)

Page 8: Economic and Path Integrals

20.1 Fluctuation Properties of Financial Assets 1447

with a Hamiltonian which generalizes the Levy Hamiltonian (20.15) to

H(p) ≡ σ2

2

α2−λ

λ(1 − λ)

[

(α + ip)λ + (α− ip)λ − 2αλ]

= σ2 (α2 + p2)λ/2 cos[λ arctan(p/α)] − αλ

αλ−2λ(1 − λ). (20.25)

The asymptotic behavior of the truncated Levy distributions differs from thepower behavior of the Levy distribution in Eq. (20.17) by an exponential factor e−αz

which guarantees the finiteness of the width σ and of all higher moments. A roughestimate of the leading term is again obtained from the Fourier transform of thelowest expansion term of the exponential function e−H(p):

L(λ,α)σ2 (z) ≈ e2sα

λ∫ ∞

−∞

dp

2πeipz

1 − s[

(α + ip)λ + (α− ip)λ]

→z→∞

e2sαλ

Γ(1 + λ)sin(πλ)

πse−α|z|

|z|1+λ, (20.26)

where

s ≡ σ2

2

α2−λ

λ(1 − λ). (20.27)

The integral follows directly from the formulas [25]

∫ ∞

−∞

dp

2πeipz (α+ ip)λ =

Θ(z)

Γ(−λ)

e−αz

z1+λ,∫ ∞

−∞

dp

2πeipz (α− ip)λ =

Θ(−z)Γ(−λ)

e−α|z|

|z|1+λ,(20.28)

and the identity for Gamma functions2

1

Γ(−z) = −Γ(1 + z) sin(πz)/π. (20.29)

The full expansion is integrated with the help of the formula [26]∫ ∞

−∞

dp

2πeipz (α + ip)λ(α− ip)ν

= (2α)λ/2+ν/2 1

|z|1+λ/2+ν/2

1

Γ(−λ)W(ν−λ)/2,(1+λ+ν)/2(2αz)

1

Γ(−ν)W(λ−ν)/2,(1+λ+ν)/2(2αz)

forz > 0,

z < 0,(20.30)

where the Whittaker functions W(ν−λ)/2,(1+λ+ν)/2(2αz) can be expressed in terms ofKummer’s confluent hypergeometric function 1F1(a; b; x) of Eq. (9.45) as

Wδ,κ(x) =Γ(−2κ)

Γ(1/2 − κ− δ)xκ+1/2e−x/2

1F1(1/2 + κ− δ; 2κ+ 1; x)

+Γ(2κ)

Γ(1/2 + κ− δ)x−κ+1/2e−x/2

1F1(1/2 − κ− δ;−2κ+ 1; x), (20.31)

2M. Abramowitz and I. Stegun, op. cit., Formula 6.1.17.

Page 9: Economic and Path Integrals

1448 20 Path Integrals and Financial Markets

as can be seen from (9.39), (9.46) and Ref. [27]. For ν = 0, only z > 0 gives anonzero integral (20.30), which reduces, with W−λ/2,1/2+λ/2(z) = z−λ/2e−z/2, to theleft equation in (20.28). Setting λ = ν we find

∫ ∞

−∞

dp

2πeipz (α2 + p2)ν = (2α)ν/2

1

|z|1+ν

1

Γ(−ν)W0,1/2+ν(2α|z|). (20.32)

Inserting

W0,1/2+ν(x) =

2z

πK1/2+ν(x/2), (20.33)

we may write

∫ ∞

−∞

dp

2πeipz (α2 + p2)ν =

(

|z|

)1/2+ν1√

πΓ(−ν)K1/2+ν(α|z|). (20.34)

For ν = −1 where K−1/2(x) = K1/2(x) =√

π/2xe−x, this reduces to

∫ ∞

−∞

dp

2πeipz

1

α2 + p2=

1

2αe−α|z|. (20.35)

Summing up all terms in the expansion of the exponential function e−H(p):

L(λ,α)σ2 (z) ≈ e2sα

λ∫ ∞

−∞

dp

1 +∞∑

n=1

(−s)nn!

[

(α + ip)λ + (α− ip)λ]n

eipz (20.36)

yields the true asymptotic behavior

L(λ,α)σ2 (z) →

z→∞e(2−2λ)sαλ

Γ(1 + λ)sin(πλ)

πse−α|z|

|z|1+λ, (20.37)

which differs from the estimate (20.26) by a constant factor (see Appendix 20A fordetails) [28]. Hence the tails are semi-heavy.

In contrast to Gaussian distributions which are characterized completely by theirwidth σ, the truncated Levy distributions contain three parameters σ, λ, and α. Bestfits to two sets of fluctuating market prices are shown in Fig. 20.7. For the S&P 500index we plot the cumulative distributions

P<(z) =∫ z

−∞dz′ L

(λ,α)σ2 (z′), P>(z) =

∫ ∞

zdz′ L

(λ,α)σ2 (z′) = 1 − P<(z), (20.38)

for the price differences z = ∆S over ∆t = 15 minutes. For the ratios of the changesof the currency rates DM/$ we plot the returns z = ∆S/S with the same ∆t. Theplot shows the negative and positive branches P<(−z), and P>(z) both plotted onthe positive z axis. By definition:

P<(−∞) = 0, P<(0) = 1/2, P<(∞) = 1,

P>(−∞) = 1, P>(0) = 1/2, P>(∞) = 0. (20.39)

Page 10: Economic and Path Integrals

20.1 Fluctuation Properties of Financial Assets 1449

z z

P><

(±z)P><

(±z)

Figure 20.7 Best fit of cumulative versions (20.38) of truncated Levy distribution to

financial data. For the S&P 500 index, the fluctuating variable z is directly the index

change ∆S every ∆t=15 minutes (fit with σ2 = 0.280 and κ = 12.7). For the DM/US$

exchange ratio, the variable z is equal to 100∆S/S every fifteen minutes (fit with σ2 =

0.0163 and κ = 20.5). The negative fluctuations lie on a slightly higher curve than the

positive ones. The difference is often neglected. The parameters A and α are the size

and truncation parameters of the distribution. The best value of λ is 3/2 (from [13]).

The dashed curves show the best fits of generalized hyperbolic functions (20.117) (l.h.s.

λ = 1.46, α = 4.93, β = 0, δ = 0.52; r.h.s. λ = 1.59, α = 32.1, β = 0, δ = 0.221).

The fits are also compared with those by other distributions explained in the figurecaptions. A fit to most data sequences is possible with a rather universal parameterλ close to λ = 3/2. The remaining two parameters fix all expansion coefficients ofHamiltonian (20.25):

H(p) =1

2c2 p

2 − 1

4!c4 p

4 +1

6!c6 p

6 − 1

8!c8 p

8 + . . . . (20.40)

The numbers c2n = −(−1)nH(2n)(0) are the cumulants of the truncated Levy dis-tribution [compare (3.587)], also denoted by 〈zn〉c. Here they are equal to

〈z2〉c = c2 = σ2,

〈z4〉c = c4 = σ2(2 − λ)(3 − λ)α−2,

〈z6〉c = c6 = σ2(2 − λ)(3 − λ)(4 − λ)(5 − λ)α−4,...

〈z2n〉c = c2n = σ2Γ(2n− λ)

Γ(2 − λ)α2−2n. (20.41)

Page 11: Economic and Path Integrals

1450 20 Path Integrals and Financial Markets

The first cumulant c2 determines the quadratic fluctuation width

〈z2〉 ≡∫ ∞

−∞dz z2 L

(λ,α)σ2 (z) = − d2

dp2e−H(p)

p=0

= c2 = σ2, (20.42)

the second the expectation of the fourth power of z

〈z4〉 ≡∫ ∞

−∞dz z4 L

(λ,α)σ2 (z) =

d4

dp4e−H(p)

p=0

= c4 + 3c22, (20.43)

and so on:

〈z6〉 = c6 + 15c4c2 + 15c32, 〈z8〉 = c8 + 28c6c2 + 35c24 + 210c4c22 + 105c42, . . . . (20.44)

In a first analysis of the data, one usually determines the so-called kurtosis , whichis the normalized fourth-order cumulant

κ ≡ c4 ≡c4c22

=〈z4〉c〈z2〉2c

=〈z4〉cσ4

. (20.45)

It depends on the parameters σ, λ, α as follows

κ =(2 − λ)(3 − λ)

σ2α2. (20.46)

Given the volatility σ and the kurtosis κ, we extract the Levy parameter α from theequation

α =1

σ

(2 − λ)(3 − λ)

κ. (20.47)

In terms of κ and λ, the normalized expansion coefficients are

-3 -2 -1 0 1 2 3

0.1

0.2

0.3 Lµ,ασ2 (z)

z

Figure 20.8 Change in shape of truncated Levy distributions of width σ = 1 with

increasing kurtoses κ = 0 (Gaussian, solid curve), 1, 2 , 5, 10.

Page 12: Economic and Path Integrals

20.1 Fluctuation Properties of Financial Assets 1451

c4 = κ, c6 = κ2(5 − λ)(4 − λ)

(3 − λ)(2 − λ), c8 = κ2

(7 − λ)(6 − λ)(5 − λ)(4 − λ)

(3 − λ)2(2 − λ)2,

...

cn = κn/2−1 Γ(n− λ)/Γ(4 − λ)

(3 − λ)n/2−2(2 − λ)n/2−2. (20.48)

For λ = 3/2, the second equation in (20.47) becomes simply

α =1

2

3

σ2κ, (20.49)

and the coefficients (20.50):

c4 = κ, c6 =5 · 7

3κ2, c8 = 5 · 7 · 11 κ2,

...

cn =Γ(n− 3/2)/Γ(5/2)

3n/2−2/2n−4κn/2−1. (20.50)

At zero kurtosis, the truncated Levy distribution reduces to a Gaussian distributionof width σ. The change in shape for a fixed width and increasing kurtosis is shownin Fig. 20.8.

From the S&P and DM/US$ data with time intervals ∆t = 15 min one extractsσ2 = 0.280 and 0.0163, and the kurtoses κ = 12.7 and 20.5, respectively. This impliesα ≈ 0.46 and α ≈ 1.50, respectively. The other normalized cumulants (c6, c8, . . .)are then all determined to be (1881.72, 788627.46, . . .) and (−4902.92, 3.3168 × 106

, . . .), respectively. The cumulants increase rapidly showing that the expansion needsresummation.

The higher normalized cumulants are given by the following ratios of expectationvalues

c6 =〈z6〉〈z2〉3 − 15

〈z4〉〈z2〉2 + 30,

c8 =〈z8〉〈z2〉4 − 28

〈z6〉〈z2〉3 − 35

〈z4〉2〈z2〉4 + 420

〈z4〉〈z2〉2 − 630, . . . . (20.51)

In praxis, the high-order cumulants cannot be extracted from the data since theyare sensitive to the extremely rare events for which the statistics is too low to fit adistribution function.

20.1.4 Asymmetric Truncated Levy Distributions

We have seen in the data of Fig. 20.7 that the price fluctuations have a slightasymmetry: Price drops are slightly larger than rises. This is accounted for by anasymmetric truncated Levy distribution. It has the general form [24]

L(λ,α,β)σ2 (p) ≡ e−H(p), (20.52)

Page 13: Economic and Path Integrals

1452 20 Path Integrals and Financial Markets

with a Hamiltonian function

H(p) ≡ σ2

2

α2−λ

λ(1 − λ)

[

(α + ip)λ(1 + β) + (α− ip)λ(1 − β) − 2αλ]

= σ2 (α2 + p2)λ/2 cos[λ arctan(p/α)] + iβ sin[λ arctan(p/α)] − αλ

αλ−2λ(1 − λ). (20.53)

This has a power series expansion

H(p) = ic1p+1

2c2 p

2 − i1

3!c3p

3 − 1

4!c4 p

4 + i1

5!c5p

5 + . . . . (20.54)

There are now even and odd cumulants cn = −inH(n)(0) with the values

cn = σ2Γ(n− λ)

Γ(2 − λ)α2−n

1

βfor

n = even,

n = odd.(20.55)

The even cumulants are the same as before in (20.41). Similarly, the even expecta-tion values (20.42)–(20.44) are extended by the odd expectation values:

〈z〉≡∫ ∞

−∞dz z L

(λ,α,β)σ2 (z) = i

d

dpe−H(p)

p=0

= c1,

〈z2〉≡∫ ∞

−∞dz z2 L

(λ,α,β)σ2 (z) = − d2

d2pe−H(p)

p=0

= c2 + c21,

〈z3〉≡∫ ∞

−∞dz z3 L

(λ,α,β)σ2 (z) = −i d

3

d3pe−H(p)

p=0

= c3 + 3c2c1 + c31,

〈z4〉≡∫ ∞

−∞dz z4 L

(λ,α,β)σ2 (z) =

d4

d4pe−H(p)

p=0

=c4 + 4c3c1 + 3c22 + 6c2c21 + c41,

... . (20.56)

The inverse relations are

c1 = 〈z〉c = 〈z〉,c2 = 〈z2〉c = 〈z2〉 − 〈z〉2 = 〈(z − 〈z〉c)〉2,c3 = 〈z3〉c = 〈z3〉 − 3〈z〉〈z2〉 + 2〈z〉3 = 〈(z − 〈z〉c)〉3,c4 = 〈z4〉c = 〈z4〉 − 3〈z2〉2 − 4〈z〉〈z3〉 + 12〈z〉2〈z2〉 − 6〈z〉4

= 〈(z − 〈z〉c)〉4 − 3〈z2 − 〈z〉2c〉2 = 〈(z − 〈z〉c)〉4 − 3c22. (20.57)

These are, of course, just simple versions of the cumulant expansions (3.585) and(3.587).

The distribution is now centered around a nonzero average value:

µ ≡ 〈z〉 = c1. (20.58)

Page 14: Economic and Path Integrals

20.1 Fluctuation Properties of Financial Assets 1453

The fluctuation width is given by

σ2 ≡ 〈z2〉 − 〈z〉2 = 〈(z − 〈z〉)2 〉 = c2. (20.59)

For large z, the asymmetric truncated Levy distributions exhibit semi-heavytails, obtained by a straightforward modification of (20.28):

L(λ,α)σ2 (z) ≈

∫ ∞

−∞

dp

2πeipz

1−σ2

2

α2−λ

λ(1−λ)

[

(α+ ip)λ(1+β) + (α− ip)λ(1−β) − 2αλ]

→z→∞

σ2 e2sαλ

Γ(1 + λ)sin(πλ)

πse−α|z|

|z|1+λ[1 + β sgn(z)]. (20.60)

In analyzing the data, one uses the skewness

s ≡ 〈(z − 〈z〉)3〉σ3

= c3 =c3

c3/22

. (20.61)

It depends on the parameters σ, λ, β, and α or κ as follows

-4 -2 2 4

0.1

Lµ,α,βσ2 (z)

z − 〈z〉

Figure 20.9 Change in shape of truncated Levy distributions of width σ = 1 and

kurtosis κ = 1 with increasing skewness s = 0 (solid curve), 0.4, 0.8 . The curves are

centered around 〈z〉.

s =(2 − λ)β

σα. (20.62)

The kurtosis can also be defined by [compare (20.45)]3

κ ≡ c4 ≡c4c22

=〈z4〉c〈z2〉2c

=〈z4〉cσ4

=〈(z − 〈z〉)4〉

σ4− 3. (20.63)

From the data one extracts the three parameters volatility σ, skewness s, and kur-tosis κ, which completely determine the asymmetric truncated Levy distribution.

3Some authors call the ratio 〈(z − 〈z〉)〉4/σ4 in (20.63) kurtosis and the quantity κ the excess

kurtosis . Their kurtosis is equal to 3 for a Gaussian distribution, ours vanishes.

Page 15: Economic and Path Integrals

1454 20 Path Integrals and Financial Markets

The data are then plotted against z − 〈z〉 = z − µ, so that they are centered at the

average position. This centered distribution will be denoted by L(λ,α,β)σ2 (z), i.e.

L(λ,α,β)σ2 (z) ≡ L

(λ,α,β)σ2 (z − µ). (20.64)

The Hamiltonian associated with this zero-average distribution is

H(p) ≡ H(p) −H ′(0)p, (20.65)

and its expansion in power of the momenta starts out with p2, i.e. the first term in(20.54) is subtracted.

In terms of σ, s, and κ, the normalized expansion coefficients are

cn =κn/2−1 Γ(n− λ)/Γ(4 − λ)

(3 − λ)n/2−2(2 − λ)n/2−2

1

(3 − λ)/(2 − λ)κ sfor

n = even,

n = odd.(20.66)

The change in shape of the distributions of a fixed width and kurtosis with increasingskewness is shown in Fig. 20.9. We have plotted the distributions centered aroundthe average position z = 〈c1〉 which means that we have removed the linear termic1p from H(p) in (20.52), (20.53), and (20.54). This subtracted Hamiltonian whosepower series expansion begins with the term c2p

2/2 will be denoted by

H(p) ≡ H(p) −H ′(0)p =1

2c2 p

2 − i1

3!c3p

3 − 1

4!c4 p

4 + i1

5!c5p

5 + . . . . (20.67)

20.1.5 Gamma Distribution

For a HamiltonianH+(p) = ν log (1 − ip/µ) (20.68)

one obtains the normalized Gamma distribution of mathematical statistics:

DGammaµ,ν (z) =

1

Γ(ν)µνzν−1e−µz ,

∫ ∞

0dz DGamma

µ,ν (z) = 1, (20.69)

which is restricted to positive variables z.Expanding H+(p) in a power series −∑∞

n=1 inνµ−npn/n = −∑∞

n=1 incnp

n/n!, weidentify the cumulants

cn = (n− 1)! ν/µn, (20.70)

so that the lowest moments are

z = ν/µ, σ2 = ν/µ2, s = 2/√ν, κ = 6/ν. (20.71)

The maximum of the distribution lies slightly below the average value z at

zmax = (ν − 1)/µ = z − 1/µ. (20.72)

Page 16: Economic and Path Integrals

20.1 Fluctuation Properties of Financial Assets 1455

The Gamma distribution was seen in Fig. 20.3 to yield an optimal fit to thefluctuating volatilities when is it is plotted as a distribution of the volatility σ =

√z:

DChiµ,ν (σ) ≡ 2σDGamma

µ,ν (σ2) ,∫ ∞

0dσ DChi

µ,ν (σ) = 1. (20.73)

As a normalized distribution of the volatility σ rather than the variance v = σ2, onecalls this function a Chi distribution.

In the limit ν → ∞ at fixed z = ν/µ, the Gamma distribution becomes a Diracδ-function:

DGammaµ,ν (z) → δ

(

z − ν

µ

)

. (20.74)

If we add to H+(p) the Hamiltonian

H−(p) = ν log(1 + ip/µ) (20.75)

we obtain the distribution of negative z-values:

DGammaµ,ν (z) =

1

Γ(ν)µ−ν |z|ν−1e−µ|z|, z ≤ 0. (20.76)

By combining the two Hamiltonians with different parameters µ one can set up askewed two-sided distribution.

20.1.6 Boltzmann Distribution

The highest-frequency returns ∆S of NASDAQ 100 and S&P 500 indices have aspecial property: they display a purely exponential behavior for positive as well asnegative z, as long as the probability is rather large [32]. The data are fitted by theBoltzmann distribution

-1 -0.5 0 0.5 1-5

-4

-3

-2

-1

-4 -2 0 2 4-5

-4

-3

-2

-1

z in % z in %

logP (z)logP (z)

NASDAQ100 2001-02 (1min)S&P500 2004-05 (1min)

Figure 20.10 Boltzmann distribution of S&P 500 and NASDAQ 100 high-frequency

log-returns recorded by the minute.

B(z) =1

2Te−|z|/T . (20.77)

Page 17: Economic and Path Integrals

1456 20 Path Integrals and Financial Markets

We can see in Fig. 20.10, that only a very small set of rare events of large |z| doesnot follow the Boltzmann law, but displays heavy tails. This allows us to assigna temperature to the stock markets [33]. The temperature depends on the volatil-ity of the selected stocks and changes only very slowly with the general economicand political environment. Near a crash it reaches maximal values, as shown inFig. 20.11.

It is interesting to observe the historic development of Dow Jones temperatureover the last 78 years (1929-2006) in Fig. 20.12. Although the world went througha lot of turmoil and economic development in the 20th century, the temperatureremained almost a constant except for short heat bursts. The hottest temperaturesoccurred in the 1930’s, the time of the great depression. These temperatures werenever reached again. An especially hot burst occurred during the crash year 1987.Thus, in the long run, the market temperature is not related to the value of theindex but indicates the riskiness of the market. Only in bubbles, high temperaturesgo along with high stock prices. An extraordinary increase in market temperaturebefore a crash may be useful to investors as a signal to go short.

The Fourier transform of the Boltzmann distribution is

B(p) =∫ ∞

−∞dz eipz

1

2Te−|z|/T =

1

1 + (Tp)2= e−H(p), (20.78)

so that we identify the Hamiltonian as

H(p) = log[1 + (Tp)2]. (20.79)

This has only even cumulants:

c2n = −(−1)nH(2n)(0) = 2(2n− 1)!T 2n, n = 1, 2, . . . . (20.80)

1990 1995 2000 20050

500

1000

1500 Index

1990 1995 2000 20050

1

2

1990 1995 2000 20050

2000

4000

6000

Index

1990 1995 2000 20050

2

4

104T104T

S&P500 index 1990-2006 NASDAQ100 index 1990-2006

year year

Figure 20.11 Market temperatures of S&P 500 and NASDAQ 100 indices from 1990 to

2006. The crash in the year 2000 occurred at the maximal temperatures TS&P500 ≈ 2×10−4

and TNASDAQ ≈ 4 × 10−4.

Page 18: Economic and Path Integrals

20.1 Fluctuation Properties of Financial Assets 1457

1930 1940 1950 1960 1970 1980 1990 2000

41

100

1000

10000

1930 1940 1950 1960 1970 1980 1990 20000

1

2

3

4

1987

Dow Jones Index 1929–2006

Market Temperature

104 T

year

Figure 20.12 Dow Jones index over 78 years (1929-2006) and the annual market tem-

perature, which is remarkably uniform, except in the 1930’s, in the beginning of the great

depression. An other high extreme temperature occurred in the crash year 1987.

The Boltzmann distribution is a special case of a two-sided Gamma distributiondiscussed in the previous subsection. It has semi-heavy tails.

We now observe that the Fourier transform can be rewritten as an integral [recall(2.499)]

B(p) = e−H(p) =1

1 + (Tp)2=∫ ∞

0dτ e−τ(1+T 2p2). (20.81)

This implies that Boltzmann distribution can be obtained from a superposition ofGaussian distributions. Indeed, the prefactor of p2 is equal to twice the variancev = σ2 of the Gaussian distribution. Hence we change the variable of integrationfrom τ to σ2, substituting τ = σ2/2T 2, and obtain

B(p) = e−H(p) =1

2T 2

∫ ∞

0dσ2 e−σ2/2T 2

e−σ2p2/2. (20.82)

Page 19: Economic and Path Integrals

1458 20 Path Integrals and Financial Markets

The Fourier transform of this is the desired representation of the Boltzmann distri-bution as a volatility integral over Gaussian distributions of different widths:

B(z) =1

2T 2

∫ ∞

0dσ2 e−σ2/2T 2 1√

2πσ2e−z2/2σ2

. (20.83)

The volatility distribution function is recognized to be a special case of the Gammadistribution (20.69) for µ = 1/2T 2, ν = 1.

20.1.7 Student or Tsallis Distribution

Recently, another non-Gaussian distribution has become quite popular through workof Tsallis [58, 59, 60]. It has been proposed as a good candidate for describing heavytails a long time ago under the name Student distribution by Praetz [61] and byBlattberg and Gonedes [62]. This distribution can be written as

Dδ(z) = Nδ1

2πσ2δ

e−z2/2σ2

δδ , (20.84)

where Nδ is a normalization factor

Nδ =

√δΓ(1/δ)

Γ(1/δ − 1/2), (20.85)

and σK ≡ σ√

1 − 3δ/2, where σ is the volatility of the distribution. The function

eδ(z) is an approximation of the exponential function called δ-exponential:

ezδ ≡ [1 − δz]−1/δ . (20.86)

In the limit δ → 0, this reduces to the ordinary exponential function ez.Remarkably, the distribution of a fixed total amount of money W between N

persons of equal earning talents follows such a distribution. The partition functionsis given by

ZN(W ) =

[

N∏

n=1

∫ W

0dwn

]

δ(w1+. . .+wN −W ). (20.87)

After rewriting this as

ZN(W ) =∫ ∞

−∞

[

N∏

n=1

∫ W

0dwn

]

e−iλ(w1+...+wN )eiλW =∫ ∞

−∞

(e−iλW − 1)NeiλW

(−iλ+ ǫ)N,

(20.88)

where ǫ is an infinitesimal positive number, and a binomial expansion of (e−iλW −1)NeiλW , we can perform the integral over λ, using the formula4

∫ ∞

−∞

1

(−iλ + ǫ)νe−ipx =

pν−1

Γ(ν)e−ǫp Θ(p), (20.89)

4See I.S. Gradshteyn and I.M. Ryzhik, op. cit., Formula 3.382.7.

Page 20: Economic and Path Integrals

20.1 Fluctuation Properties of Financial Assets 1459

where Θ(z) is the Heaviside function (1.304), we obtain

ZN(W ) =WN−1

Γ(N)

N−1∑

k=0

(−1)k(

N

N − k

)

(N − k − 1)N−1 (20.90)

The sum over binomial coefficients adds up to unity, due to a well-known identityfor binomial coefficients5, so that

ZN(W ) =WN−1

Γ(N). (20.91)

If we replace W by W − wn, this partition function gives us the unnormalizedprobability that an individual owns the part wn of total wealth. The normalizedprobability is then

PN(wn) = Z−1N

(W − wn)N−2

Γ(N − 1)=N − 1

W

(

1 − wn

E

)N−2

. (20.92)

Defining w ≡W/(N−2), which for large N is the average wealth per person, PN(wn)can be expressed in terms of the δ-exponential (20.86) as

PN(wn) =N − 1

W

[

1 − wn

(N − 2)w

]N−2

=N − 1

We−wn/w−1/(N−2). (20.93)

In the limit of infinitely many individuals, this turns into the normalized Boltzmanndistribution w−1e−wn/w.

The Student-Tsallis distribution (20.84) is simply a normalized δ-exponential.Instead of the parameter δ one often uses the Tsallis parameter q = δ+ 1. A plot ofthese functions for different δ-values is shown on the left hand of Fig. 20.13. FromEq. (20.86) we see that the Student-Tsallis distribution with δ > 0 has heavy tailswith a power behavior 1/z1/δ = 1/z1/(q−1). A fit of the log-returns of 10 NYSEtop-volume stocks is shown in Fig. 20.13.

Remarkably, the δ-exponential in (20.84) can be written as a superposition ofGaussian distributions. With the help of the integral formula (2.499) we find

e−z2/2σ2

δδ =

1

Γ(1/δ)

∫ ∞

0

ds

ss1/δe−se−sδ z2/2σ2

δ . (20.94)

After a change of variables this can be rewritten as

e−z2/2σ2

δδ =

µ1/δ

Γ (1/δ)

∫ ∞

0

dv

vv1/δe−µve−vz2/2 , µ = σ2

δ/δ. (20.95)

This is a superposition of Gaussian functions e−vz2/2 whose inverse variances v =1/σ2 occur with a weight that is precisely the Gamma distribution of the variable vwith ν = 1/δ:

DGammaµ,1/δ (v) =

1

Γ (1/δ)µ1/δvνe−µv. (20.96)

5See I.S. Gradshteyn and I.M. Ryzhik, op. cit., Formula 0.154.6.

Page 21: Economic and Path Integrals

1460 20 Path Integrals and Financial Markets

- 4 - 2 2 4

- 10

- 8

- 6

- 4

- 2

0

Probability

z

δ = 0.43

δ = 0.44

δ = 0.41

log Dδ(z)

z

logDδ(z)

Figure 20.13 Left: Logarithmic plot of the normalized δ-exponential (20.84) (Student-

Tsallis distribution) for δ = 0 (Gaussian), 0.2, 0.4, 0.6, all for σ = 1. Right: Fit of the

log-returns of 20 NYSE top-volume stocks over short time scales from 1 to 3 minutes by

this distribution (after Ref. [60]). Dotted line is Gaussian distribution.

20.1.8 Tsallis Distribution in Momentum Space

When fitting the volatility distributions of the S&P 500 index in Fig. 20.3, weobserved that the best fit was obtained by a Gamma distribution. Inspired by thediscussion in the last section we observe that the δ-exponential (20.86) in momentumspace

e−Hδ,β(p) ≡ e−βp2/2δ =

[

1 + βδ p2/2]−1/δ

, (20.97)

has precisely such a decomposition. We simply rewrite it by analogy with (20.94)as

e−βp2/2δ =

1

Γ(1/δ)

∫ ∞

0

ds

ss1/δe−se−sβδ p2/2 , β ≡ ν/µ = 1/µδ, (20.98)

and change the variable of integration to obtain

e−βp2/2δ =

µν

Γ (ν)

∫ ∞

0

dv

vvνe−µve−vp2/2, ν = 1/δ, µ = ν/β = 1/βδ. (20.99)

This is a superposition of Gaussian distributions whose variances v = σ2 areweighted by the Gamma distribution (20.69):

e−βp2/2δ =

∫ ∞

0dv DGamma

µ,ν (v)e−vp2/2, ν = 1/δ, µ = ν/β = 1/βδ. (20.100)

The average of the Gamma distribution is v = ν/µ = β [recall (20.71)], so that the

left-hand side can also be written as e−vp2/2δ .

The lowest cumulants in the Hamiltonian Hδ,β(p) = 12!c2 p

2 − 14!c4 p

4 + . . . are

c2=ν

µ, c4 =3

ν

µ2, c6=

5!!

µ3

(

2+3ν−2ν2−ν3)

, c8=7!!

µ3

(

2+ν−3ν2+ν3+nu4)

.(20.101)

Page 22: Economic and Path Integrals

20.1 Fluctuation Properties of Financial Assets 1461

For µ = 1/2T 2 and ν = 1, these reduce to those of the Boltzmann distribution inEq. (20.80).

The superposition (20.99) can immediately be Fourier transformed to

Dδ,β(z) =µ1/δ

Γ (1/δ)

∫ ∞

0

dv

vv1/δe−µv 1√

2πve−z2/2v , µ = 1/βδ. (20.102)

which is a superposition of Gaussians whose variances v = σ2 follow a Gammadistribution (20.69) centered around v = 1/δµ of width (v − v)2 = 1/δµ2. Hence δis given by the ratio δ = (v − v)2/v2. Recalling Eq. (20.71) we see that δ determinesthe kurtosis of the distribution to be κ = 6δ.

The integral over v in (20.102) can be done using Formula (2.559), yielding

Dδ,β(z) =

õ

Γ (1/δ)

1√2π

(

µz2

2

)1/2δ−1/4

2K1/δ−1/2(√

2µz), µ = 1/βδ.(20.103)

For δ = 1 and µ = 1/2T 2, we use (1.349) and recover the Boltzmann distribution(20.83)

The small-z behavior of Kν(z) is (1/2)Γ(ν) (z/2)−ν for Re ν > 0 [recallEq. (1.351)]. If we assume δ < 2, we obtain from (20.103) the value of the dis-tribution function at the origin:

Dδ,β(0) =

√µΓ(1/δ − 1/2)

Γ(1/δ). (20.104)

The same result can, of course, be found directly from Eq. (20.102). This valuediverges at δ = 2/(1 − 2n) (n = 0, 1, 2, . . .).

20.1.9 Relativistic Particle Boltzmann Distribution

A physically important distribution in momentum space is the Boltzmann distri-bution e−βE(p) of relativistic particle energies E(p) =

√p2 +M2 , where β is the

inverse temperature 1/kBT . The exponential can be expressed as a superposition ofGaussian distributions

e−β√

p2+M2=∫ ∞

0dv ωβ(v)e−βv(p2+M2)/2 (20.105)

with a weight function

ωβ(v) ≡√

β

2πv3e−β/2v. (20.106)

This is a special case of a Weibull distribution

DW(x) =b

Γ(a/b)xa−1e−xb

, (20.107)

which has the simple moments

〈xn〉 = Γ((a+ n)/b)/Γ(a/b). (20.108)

Page 23: Economic and Path Integrals

1462 20 Path Integrals and Financial Markets

20.1.10 Meixner Distributions

Quite reasonable fits to financial data are provided by the Meixner distributions [34, 35] whichread in configuration and momentum space:

M(z) =[2 cos(b/2)]

2d

2aπΓ(2d)|Γ (d+ iz/a) |2 exp [bz/a] , (20.109)

M(p) =

cos(b/2)

cosh [(ap− ib)/2]

2d

. (20.110)

They have the same semi-heavy tail behavior as the truncated Levy distributions

M(z) → C±|z|ρe−σ±|z| for z → ±∞, (20.111)

with

C± =[2 cos(b/2)]

2d

2aπΓ(2d)

a2d−1e±2πd tan(b/2), ρ = 2d− 1, σ± ≡ (π ± b)/a. (20.112)

The moments are

µ = ad tan(b/2), σ2 = a2d/2 cos2(b/2), s =√

2 sin(b/2)/√d, κ = [2 − cos b]/d, (20.113)

such that we can calculate the parameters from the moments as follows:

a2 = σ2(

2κ− 3s2)

, d =1

κ− s2, b = 2 arcsin

(

s√

d/2)

. (20.114)

As an example for the parameters of the distribution, a good fit of the daily Nikkei-225 indexis possible with

a = 0.029828, b = 0.12716, d = 0.57295, 〈z〉 = −0.0011243. (20.115)

The curve has to be shifted in z by ∆z to make ∆z + µ equal to 〈z〉. Such a Meixner distributionhas been used for option pricing in Ref. [35].

The Meixner distributions can be fitted quite well to the truncated Levy distribution in theregime of large probability. In doing so we observe that the variance σ2 and the kurtosis κ arenot the best parameters to match the two distributions. The large-probability regime of thedistributions can be matched perfectly by choosing, in the symmetric case, the value and thecurvature at the origin to be the same in both curves. This is seen in Fig. 20.14.

In the asymmetric case we have to match also the first and third derivatives. The derivativesof the Meixner distribution are:

M(0) =22d−1Γ2(d)

πaΓ(2d),

M ′(0) = b22d−1Γ2(d) [1 − dψ(d)]

πa2Γ(2d),

M ′′(0) = −22dΓ2(d)ψ(d)

πa3Γ(2d),

M (3)(0) = − b

2

22dΓ2(d)[

6ψ(d) − 6dψ2(d) − dψ(3)(d)]

πa4Γ(2d),

M (4)(0) =22dΓ2(d)

[

6ψ2(d) + ψ(3)(d)]

πa5Γ(2d), (20.116)

where ψ(n)(z) ≡ dn+1 log Γ(z)/dzn+1 are the Polygamma functions.

Page 24: Economic and Path Integrals

20.1 Fluctuation Properties of Financial Assets 1463

0.25 0.5 0.75 1 1.25 1.5 1.75 20

0.25

0.5

0.75

1

1.25

1.5

1.75

z/σ

M(z)

Figure 20.14 Comparison of best fit of Meixner distribution to truncated Levy distri-

butions. One of them (short dashed) has the same volatility σ and kurtosis κ. The other

(long-dashed) has the same size and curvature at the origin. The parameters are σ2 = 0.280

and κ = 12.7 as in the left-hand cumulative distribution in Fig. 20.7. The Meixner distri-

bution with the same σ2 and κ has parameters a = 2.666, d = 0.079, b = 0, the distribution

with the same value and curvature at the origin has a = 0.6145, d = 1.059, b = 0. The very

large σ-regime, however, is not fitted well as can be seen in the cumulative distributions

which reach out to z of the order of 10σ in Fig. 20.7.

20.1.11 Generalized Hyperbolic Distributions

Another non-Gaussian distributions proposed in the literature is the so-called generalized hyperbolic

distributions ..6 As the truncated Levy and Meixner distributions, these have simple analytic formsboth in z- and p-space:

HG(z) =

(

γ2− β2)λ/2

eβ z

γλ−1/2δλ√

2 π

[

δ2+ z2]λ/2−1/4Kλ−1/2

(

γ√δ2+ z2

)

(

δ√

γ2 − β2) (20.117)

and

G(p) =

(

δ√

γ2 − β2)λ

(

δ√

γ2 − β2)

(

δ

γ2 − (β + ip)2

)

[

δ

γ2 − (β + ip)2

]λ, (20.118)

the latter defining another Hamiltonian

HG(p) ≡ − logG(p). (20.119)

But in contrast to the former, this set of functions is not closed under time evolution. Thedistributions at a later time t are obtained from the Fourier transforms e−H(p)t. For truncated Levydistributions and the Meixner distributions the factor t can simply be absorbed in the parametersof the functions (σ2 → tσ2 in the first and d → td in the second case). For the generalized

6See the publications [36]–[74].

Page 25: Economic and Path Integrals

1464 20 Path Integrals and Financial Markets

hyperbolic distributions, this is no longer true since e−HG(p)t = [G(p)]t involves higher powers ofBessel functions whose analytic Fourier transform cannot be found. In order to describe a completetemporal evolution we must therefore close the set of functions by adding all Fourier transformsof e−HG(p)t. In praxis, this is not a serious problem — it merely leads to a slowdown of numericcalculations which always involve a numeric Fourier transformation.

The asymptotic behavior of the generalized hyperbolic distributions has semi-heavy tails. Fromthe large-z behavior of the Bessel function Kν(z) →

π/2ze−z we find

HG(z) →√

π

(

γ2− β2)λ/2

eβ z

γλ−1/2δλ√

2 π

1

(

δ√

γ2 − β2)zλ−1e−γz. (20.120)

Introducing the variable ζ ≡ δ√

γ2 − β2, this can again be expanded in powers of p as inEq. (20.54), yielding the first two cumulants:

c1 = βδ2

ζ

K1+λ(ζ)

Kλ(ζ); (20.121)

c2 =δ2

ζ

K1+λ(ζ)

Kλ(ζ)+β2δ4

ζ2

K2+λ(ζ)

Kλ(ζ)−[

K1+λ(ζ)

Kλ(ζ)

]2

. (20.122)

Using the identity [50]

Kν+1(z) −Kν−1(z) =2ν

zKν(z), (20.123)

the latter equation can be expressed entirely in terms of

ρ = ρ(ζ) =K1+λ(ζ)

Kλ(ζ)(20.124)

as

c2 =δ2

ζρ+

β2δ4

ζ3[

ζ + 2 (1 + λ) ρ− ζρ2]

. (20.125)

Usually, the asymmetry of the distribution is small implying that c1 is small, implying a small β.It is then useful to introduce the symmetric variance

σ2s ≡ δ2ρ/ζ, (20.126)

and write

c1 = βσs, c2 = σ2 = σ2s + β2

[

δ4

ζ2+ 2 (1 + λ)

δ2

ζ2σs − σ2

s

]

. (20.127)

The cumulants c3 and c4 are most compactly written as

c3 = β

[

3δ4

ζ2+ 6 (1 + λ)

δ2

ζ2σ2s − 3σ4

s

]

+ β3

2 (2+λ)δ6

ζ4+[

4 (1+λ) (2+λ)− 2ζ2] δ4

ζ4σ2s − 6 (1 + λ)

δ2

ζ2σ4s + 2σ6

s

(20.128)

and

c4 = κσ4 =3δ4

ζ2+

6δ2

ζ2(1 + λ) σ2

s − 3σ4s

+ 6β2

2 (2+λ)δ6

ζ4+[

4 (1+λ) (2+λ) − 2ζ2] δ4

ζ4σ2s − 6 (1 + λ)

δ2

ζ2σ4s + 2σ6

s

+ β4

[

4 (2+λ) (3 + λ) − ζ2] δ8

ζ6+[

4 (1+λ) (2+λ) (3 + λ) − 2 (5 + 4λ) ζ2] δ6

ζ6σ2s

− 2[

(1 + λ) (11 + 7λ) − 2ζ2] δ4

ζ4σ4s + 12 (1 + λ)

δ2

ζ2σ6s − 3σ8

s

. (20.129)

Page 26: Economic and Path Integrals

20.1 Fluctuation Properties of Financial Assets 1465

The first term in c4 is equal to σ4s times the kurtosis of the symmetric distribution

κs ≡3δ4

ζ2σ4s

+6δ2

ζ2σ2s

(1 + λ) − 3. (20.130)

Inserting here σ2s from (20.126), we find

κs ≡3

r2(ζ)+ (1 + λ)

6

ζ r(ζ)− 3. (20.131)

Since all Bessel functions Kν(z) have the same large-z behavior Kν(z) →√

π/2ze−z and thesmall-z behavior Kν(z) → Γ(ν)/2(z/2)ν, the kurtosis starts out at 3/λ for ζ = 0 and decreasesmonotonously to 0 for ζ → ∞. Thus a high kurtosis can be reached only with a small parameterλ.

The first term in c3 is βκsσ4s , and the first two terms in c4 are κsσ

4s + 6

(

c3/β − κsσ4s

)

. For asymmetric distribution with certain variance σ2

s and kurtosis κs we select some parameter λ < 3/κs,and solve the Eq. (20.131) to find ζ. This is inserted into Eq. (20.126) to determine

δ2 =σ2s ζ

ρ(ζ). (20.132)

If the kurtosis is larger, it is not an optimal parameter to determine generalized hyperbolic dis-tributions. A better fit to the data is reached by reproducing correctly the size and shape of thedistribution near the peak and allow for some deviations in the tails of the distribution, on whichthe kurtosis depends quite sensitively.

For distributions which are only slightly asymmetric, which is usually the case, it is sufficientto solve the above symmetric equations and determine the small parameter β approximately bythe skew s = c3/σ

3 from the first line in (20.128) as

β ≈ s

κsσs. (20.133)

This approximation can be improved iteratively by reinserting β into the second equation in(20.127) to determine, from the variance σ2 of the data, an improved value of σ2

s . Then Eq. (20.129)is used to determine from the kurtosis κ of the data an improved value of κs, and so on.

For the best fit near the origin where probabilities are large we use the derivatives

G(0) =

(

δ

γ

)λ−1/2

ζ−λk−, (20.134)

G′(0) = βG(0), (20.135)

G′′(0) = −(

δ

γ

)λ−3/2

ζ−λk+ +

(

δ

γ

)λ−1/2

δ−2ζ−λ(

1 − 2λ− β2δ2)

k−,

G(3)(0) = −β2

[

3

(

δ

γ

)λ−3/2

ζ−λk+ +

(

δ

γ

)λ−1/2

δ−2ζ−λ(

3 − 6λ− β2δ2)

]

k−, (20.136)

where we have abbreviated

k± ≡ 1√2π

Kλ±1/2(δγ)

Kλ(δγ). (20.137)

The generalized hyperbolic distributions with λ = 1 are called hyperbolic distributions . Theoption prices following from these can be calculated by inserting the appropriate parameters intoan interactive internet page (see Ref. [51]). Another special case used frequently in the literatureis λ = −1/2 in which case one speaks of a normal inverse Gaussian distributions also abbreviatedas NIGs.

Page 27: Economic and Path Integrals

1466 20 Path Integrals and Financial Markets

20.1.12 Debye-Waller Factor for Non-Gaussian Fluctuations

At the end of Section 3.10 we calculated the expectation value of an exponentialfunction 〈ePz〉 of a Gaussian variable which led to the Debye-Waller factor of Braggscattering (3.311). This factor was introduced in solid state physics to describe thereduction of intensity of Bragg peaks due to the thermal fluctuations of the atomicpositions. It is derivable from the Fourier representation

〈ePz〉 ≡∫

dz1√

2πσ2e−z2/2σ2

ePz =∫

dz∫

dp

2πe−σ2p2/2eipz+Pz = eσ

2P 2/2.(20.138)

There exists a simple generalization of this relation to non-Gaussian distributions,which is

〈ePz〉 ≡∫

dz∫ dp

2πe−H(p)eipz+Pz = e−H(iP ). (20.139)

20.1.13 Path Integral for Non-Gaussian Distribution

Let us calculate the properties of the simplest process whose fluctuations are dis-tributed according to any of the general non-Gaussian distributions. We considerthe stochastic differential equation for the logarithms of the asset prices

x(t) = rx + η(t), (20.140)

where the noise variable η(t) is distributed according to an arbitrary distribution.The constant drift rx in (20.140) is uniquely defined only if the average of the noisevariable vanishes: 〈η(t)〉 = 0. The general distributions discussed above can have anonzero average 〈x〉 = c1 which has to be subtracted from η(t) to identify rx. Thesubsequent discussion will be simplest if we imagine rx to have replaced c1 in theabove distributions, i.e. if the power series expansion of the Hamiltonian (20.54) isreplaced as follows:

H(p) → Hrx(p) ≡ H(p) −H ′(0)p+ irxp ≡ H(p) + irxp

≡ irx p +1

2c2 p

2 − i1

3!c3p

3 − 1

4!c4 p

4 + i1

5!c5p

5 + . . . . (20.141)

Thus we may simply work with the original expansion (20.54) and replace at theend

c1 → rx. (20.142)

The stochastic differential equation (20.140) can be assumed to read simply

x(t) = η(t). (20.143)

With the ultimate replacement (20.142) in mind, the probability distribution of theendpoints xb = x(tb) for the paths starting at a certain initial point xa = x(ta) isgiven by a path integral of the form (18.342):

P (xbtb|xata) =∫

Dη∫ x(tb)=xb

x(ta)=xa

Dx exp[

−∫ tb

tadt H(η(t))

]

δ[x− η]. (20.144)

Page 28: Economic and Path Integrals

20.1 Fluctuation Properties of Financial Assets 1467

The function H(η) is the negative logarithm of the chosen distribution of log-returns[recall (20.13)]

H(η) = − log D(η). (20.145)

For example, H(η) is given by − log L(λ,α)σ2 (η) of Eq. (20.24) for the truncated Levy

distribution or by − log B(η) of Eq. (20.77) or the Boltzmann distributions.In the mathematical literature, the measure in the path integral over the noise

Dµ ≡ Dη P [η] = Dη e−∫ tbta

dt H(η(t))(20.146)

of the probability distribution (20.144) is called the measure of the process x(t) =η(t). The path integral

∫ x(tb)=xb

x(ta)=xa

Dx δ[x− η] (20.147)

is called the filter which determines the distribution of xb at time tb for all pathsx(t) starting out at xa at time ta.

A measure differing from (20.146) only by the drift

Dµ′ = Dη P [η − r] = Dη e−∫ tbta

dt H(η(t)−r)(20.148)

is called equivalent measure. The ratio

Dµ′/Dµ = e−∫ tbta

[H(η−r)−H(η)] (20.149)

is called the Radon-Nikodym derivative. For a Gaussian noise, this is simply

Dµ′/Dµ = e∫ tbta

[rη(t)−r2t/2]. (20.150)

If one wants to calculate the expectation value of any function f(x(t)), one hasto split the filter into a product

[

∫ x(tb)=xb

x(t)=xDx δ[x− η]

]

×[

∫ x(t)=x

x(ta)=xa

Dx δ[x− η]

]

, (20.151)

and perform an integral over f(x) with this filter in the path integral (20.146). Goingover to the probabilities P (xbtb|xata), we obtain the integral

〈f(x(t))〉 =∫

dxP (xbtb|x t)f(x)P (x t|xata). (20.152)

The correlation functions of the noise variable η(t) in the path integral (20.144)are given by a straightforward functional generalization of formulas (20.56). For this

purpose, we express the noise distribution P [η] ≡ exp[

− ∫ tbta dt H(η(t))]

in (20.144)as a Fourier path integral

P [η] =∫ Dp

2πexp

∫ tb

tadt [ip(t)η(t) −H(p(t))]

, (20.153)

Page 29: Economic and Path Integrals

1468 20 Path Integrals and Financial Markets

and note that the correlation functions can be obtained from the functional deriva-tives

〈η(t1) · · ·η(tn)〉 = (−i)n∫

Dη∫ Dp

[

δ

δp(t1)· · · δ

δp(tn)ei∫ tbta

dt p(t)η(t)

]

e−∫ tbta

dtH(p(t)).

After n partial integrations, this becomes

〈η(t1) · · ·η(tn)〉 = in∫

Dη∫ Dp

2πei∫ tbta

dt p(t)η(t) δ

δp(t1)· · · δ

δp(tn)e−∫ tbta

dtH(p(t))

= in[

δ

δp(t1)· · · δ

δp(tn)e−∫ tbta

dtH(p(t))

]

p(t)≡0

. (20.154)

By expanding the exponential e−∫ tbta

dtH(p(t))in a power series (20.54) we find imme-

diately the lowest correlation functions

〈η(t1)〉 ≡ Z−1∫

Dη η(t1) exp[

−∫ tb

tadt H(η(t))

]

= 0, (20.155)

〈η(t1)η(t2)〉 ≡ Z−1∫

Dη η(t1)η(t2) exp[

−∫ tb

tadt H(η(t))

]

= c2δ(t1 − t2) + c21, (20.156)

〈η(t1)η(t2)η(t3)〉 ≡ Z−1∫

Dη η(t1)η(t2)η(t3) exp[

−∫ tb

tadt H(η(t))

]

= c3δ(t1−t2)δ(t1−t3)+ c2c1[δ(t1−t2) + δ(t2−t3)+δ(t1−t3)] + c31, (20.157)

〈η(t1)η(t2)η(t3)η(t4)〉 ≡ Z−1∫

Dη η(t1)η(t2)η(t3)η(t4) exp[

−∫ tb

tadt H(η(t))

]

= c4δ(t1−t2)δ(t1−t3)δ(t1−t4)+ c3c1[δ(t1−t2)δ(t1−t3) + 3 cyclic perms]

+ c22[δ(t1−t2)δ(t3−t4)+δ(t1−t3)δ(t2−t4)+δ(t1−t4)δ(t2−t3)]+ c2c

21[δ(t1−t2) + 5 pair terms] + c41, (20.158)

where

Z ≡∫

Dη exp[

−∫ tb

tadt H(η(t))

]

. (20.159)

The higher correlation functions are obvious generalizations of (20.44). The differentcontributions on the right-hand side of (20.156)–(20.262) are distinguishable by theirconnectedness structure.

Note that the term proportional to c3 in the three-point correlation function(20.158) and the terms proportional to c3 and to c4 in the four-point correlationfunction (20.262) do not obey Wick’s rule (3.305) since they contain contributionscaused by the non-Gaussian terms −ic3p3/3! −c4p4/4! in the Hamiltonian (20.141).

Page 30: Economic and Path Integrals

20.1 Fluctuation Properties of Financial Assets 1469

20.1.14 Time Evolution of Distribution

The δ-functional in Eq. (20.144) may be represented by a Fourier integral leadingto the path integral

P (xbtb|xata)=∫

Dη∫

Dx∫ Dp

2πexp

∫ tb

tadt[

ip(t)x(t)−ip(t)η(t)−H(η(t))]

. (20.160)

Integrating out the noise variable η(t) amounts to performing the inverse Fouriertransform of the type (20.24) at each instant of time. Then we obtain

P (xbtb|xata) =∫

Dx∫ Dp

2πexp

∫ tb

tadt [ip(t)x(t) −H(p(t))]

. (20.161)

Integrating over all x(t) with fixed endpoints enforces a constant momentum alongthe path, and we remain with the single momentum integral

P (xbtb|xata) =∫ ∞

−∞

dp

2πexp [ip(xb − xa) − (tb − ta)H(p)] . (20.162)

Given an arbitrary noise distribution D(z) extracted from the data of a given fre-quency 1/∆t, we simply identify the Hamiltonian H(p) from the Fourier represen-tation

D(z) =∫ ∞

−∞

dp

2πeikz−H(p), (20.163)

and insert H(p) into (20.170) to find the time dependence of the distribution. Thetime is then measured in units of the time interval ∆t.

For a truncated Levy distribution, the result is

P (xbtb|xata) = L(λ,α)σ2(tb−ta)

(xb − xa). (20.164)

This is a truncated Levy distribution of increasing width. The result for otherdistributions is analogous.

Since the distribution (20.170) depends only on t = tb − ta and x = xb − xa, weshall write it shorter as

P (x, t) =∫ ∞

−∞

dp

2πexp [ipx− tH(p)] . (20.165)

20.1.15 Central Limit Theorem

For large t, the distribution P (x, t) becomes more and more Gaussian (see Fig. 20.18for the Boltzmann distribution). This is a manifestation of the central limit theorem

of statistical mechanics which states that the convolution of infinitely many arbitrarydistribution functions of finite width always approaches a Gaussian distribution.This is easily proved. We simply note that after an integer number t of convolutions,

Page 31: Economic and Path Integrals

1470 20 Path Integrals and Financial Markets

a probability distribution D(z) with Hamiltonian H(p) has the Fourier components[D(p)]t = e−tH(p), so that it is given by the integral

D(z, t) =∫

dp

2πeipze−tH(p). (20.166)

For large t, the integral can be evaluated in the saddle point approximation (4.51).Denoting the momentum at extremum of the exponent by pz, which is determinedimplicitly by tH ′(pz) = iz, and setting σ2 = H ′′(pz), we obtain the Gaussian distri-bution

D(z, t) →t large

eipzz−tH(pz)∫

dp

2πei(p−pz)z−tσ2(p−pz)2/2=

eipzz−tH(pz)

√2πσ2

e−z2/2tσ2

=eσ

2p2z/2−tH(pz)

√2πσ2

e−(z−tσ2pz)2/2tσ2

. (20.167)

The same procedure can be applied to the integral (20.165).The transition from Boltzmann to Gaussian distributions is shown for the S&P

500 index in Fig. 20.15.If a distribution has a heavy power tail ∝ |z|−λ−1 with λ < 2, so that the volatility

σ is infinite, the Hamiltonian starts out for small p like |p|λ, and the saddle pointapproximation is governed by this term rather than the quadratic term p2. Foran asymmetric distribution, the leading terms in the Hamiltonian are those of theasymmetric Levy distribution (20.21) plus a possible linear term irp accounting fora drift, and the asymptotic z-behavior will be given by Eq. (20.17) with an extrafactor (1 + β) [21].

-5 0 5

-4

-3

-2

-1

-5 0 5-4

-3

-2

-1

-5 0 5

-3

-2

-1

z in % z in % x in %

logP (z) logP (z) logP (z)

Figure 20.15 Fits of Gaussian distribution to S&P 500 log-returns recorded in intervals

of 60 min, 240 min, and 1 day.

If a distribution has no second moments, the central limit theorem does notapply. Such distributions tend to another limit, the so-called Pareto-Levy-stable

distribution. Their Hamiltonian has precisely the generic form (20.21), except for apossible additional linear drift term [21]:

H(p) = −irp+Hλ,σ,β(p). (20.168)

Page 32: Economic and Path Integrals

20.1 Fluctuation Properties of Financial Assets 1471

In this context, the Levy parameter λ is also called the stability parameter . Theparameter α has to be in the interval (0, 2]. The parameter β is called skewness

parameter . For plots of D(z, 1) see Ref. [75].The convergence of variable with heavy-tail distributions is the content of the

generalized central limit theorem.

20.1.16 Additivity Property of Noises and Hamiltonians

At this point we make the useful observation that each term in the Hamiltonian(20.141) can be attributed to an independent noise term in the stochastic differentialequation (20.140). Indeed, if this differential equation has two noise terms

x(t) = rx + η1(t) + η2(t) (20.169)

whose fluctuations are governed by two different Hamiltonians, the probability dis-tribution (20.144) is replaced by

P (xbtb|xata)=∫

Dη1∫

Dη2∫

Dx exp[

−∫ tb

tadt[H1(η1(t))+H2(η2(t))

]

δ[x−rx−η1−η2].

After a Fourier decomposition of the δ-functional, this becomes

P (xbtb|xata) =∫

Dp1∫

Dp2∫

Dη1∫

Dη2∫

Dx∫

Dpe−∫ tbta

dt [ip(x−rx−η1−η2)]

× exp

−∫ tb

tadt [H1(p1) − ip1η1(t) +H2(p2) − ip2η2(t)]

,

after which the path integrals over η1(t), η2(t) lead to

P (xbtb|xata)=∫

Dp∫

Dx exp∫ tb

tadt [ipx− irxp−H1(p)−H2(p)]

. (20.170)

This is the integral representation Eq. (20.161) with the combined HamiltonianH(p) = irxp + H1(p) + H2(p), which can be rewritten as a pure integral (20.170).By rewriting this integral trivially as

P (xbtb|xata) =∫ ∞

−∞

dp12π

dp22π

eip1(xb−xc)−(tb−ta)[irxp+H1(p1)]eip2(xc−xa)−(tb−ta)H2(p2),(20.171)

we see that the probability distribution associated with a sum of two Hamiltoniansis the convolution of the individual probability distributions:

P (xbtb|xata) =∫ ∞

−∞dxdxcP1(xbtb|xcta)P2(xctb|xata). (20.172)

Thus we may calculate the probability distributions associated with the noises η1(p)and η2(p) separately, and combine the results at the end by a convolution.

Page 33: Economic and Path Integrals

1472 20 Path Integrals and Financial Markets

20.1.17 Levy-Khintchine Formula

It is sometimes useful to represent the Hamiltonian in the form of a Fourier integral

H(p) =

dz eipzF (z). (20.173)

Due to the special significance of the linear term in H(p) governing the drift, this is usuallysubtracted out of the integral by rewriting (20.173) as

Hr(p) = irp+

dz(

eipz − 1 − ipz)

F (z). (20.174)

The first subtraction ensures the property Hr(0) = 0 which guarantees the unit normalization ofthe distribution. This subtracted representation is known as the Levy-Khintchine formula, andthe function F (z) is the so-called Levy weight of the distribution. Some people also subtract outthe quadratic term and write

Hr(p) = irp+σ2

2p2 +

dz(

eipz − 1)

F (z). (20.175)

They employ a weight F (z) which has no first and second moment, i.e.,∫

dz F (z)z =0,∫

dz F (z)z2 = 0, to avoid redundancy in the representation.Note that according to the central limit theorem (20.167), the large-time probability distribu-

tion of x will become Gaussian for large t. Thus, in the Levy-Khintchine decomposition (20.175)of the Hamiltonian H(p), only the first two terms contribute for large t leading to the distribution:

P (x, t) →t large

e−(x−tr)2/2tσ2

√2πσ2

. (20.176)

The Levy measure has been calculated explicitly for many non-Gaussian distributions. As anexample, the generalized hyperbolic case has for λ ≥ 0 a Levy measure [76]

F (z) =eβz

|z|

1

π2

∫ ∞

0

dy

y

e−√

y+γ2|z|

J2λ(δ

√y) + Y 2

λ (δ√y)

+ λe−γ|z|

, (20.177)

where Jλ(z) and Yλ(z) are standard Bessel functions.The decomposition of the Hamiltonian according the Levy-Khintchine formula and the addi-

tivity of the associated noises form the basis of the Lvy-Ito theorem which states that an arbitrarystochastic differential equation with Hamiltonian (20.175) can be decomposed in the form

x = rxt+ ηG + η≤1 + η>1, (20.178)

where ηG is a Gaussian noise

η≤1 =

|x|≤1

dz(

eipz − 1)

F (z) (20.179)

a superposition of discrete noises called Poisson point process with jumps smaller than or equal tounity, and

η>1 =

|x|≤1

dz(

eipz − 1)

F (z) (20.180)

a noise with jumps larger than unity.Consider the simplest noise of the type η≤1 arising from a Levy weight FZ(z) = τZδ(z + Z)

with 0 < Z ≤ 1 in (20.179) so that the Hamiltonian is HZ(p) = τZ(e−iZp − 1). The associateddistribution function DZ(z) is, according to (20.11),

DZ(z) =

dp

2πeipzeτZ(e−ipZ−1). (20.181)

Page 34: Economic and Path Integrals

20.1 Fluctuation Properties of Financial Assets 1473

Expanding the last exponential in powers of eipZ , we obtain

DZ(z) =

dp

2πeipz

∞∑

n=0

e−τZτnZn!τnZe

−ipnZ =

∞∑

n=0

e−τZτnZn!δ(z − nZ). (20.182)

This function describes jumps by nZ (n = 0, 1, 2, 3, . . .) whose probability follows a Poisson dis-

tribution

P (n, τZ) = e−τZτnZn!, (20.183)

which is properly normalized to unity:∑∞

n=0 P (n, τZ) = 1. The expectation values of powers ofthe jump number are

〈nk〉 =

∞∑

n=0

nkP (n, τZ) = e−τZ (τZ∂τz )keτZ

∞∑

n=0

P (n, τZ) = e−τZ (τZ∂τz)keτZ =

Γ(λZ + k)

Γ(λZ).

(20.184)Thus 〈n〉 = λZ , 〈n2〉 = λZ(λZ +1), 〈n3〉 = λZ(λZ +1)(λZ +2), 〈n4〉 = λZ(λZ +1)(λZ +2)(λZ +3),so that σ2 = λZ , s = 2/

√λZ , and κ = 6/λZ .

As typical noise curve is displayed in Fig. 20.16. An arbitrary Levy weight F (z) in η≤1 may

t

Figure 20.16 Typical Noise of Poisson Process.

always be viewed as a superposition F (z) =∫ 1

−1 dZ F (Z)δ(z−Z) so that the distribution functionbecomes a superposition of Poisson noises:

D(z) =

∫ 1

−1

dZ F (Z)∞∑

n=0

e−τZτnZn!δ(z − nZ). (20.185)

20.1.18 Semigroup Property of Asset Distributions

An important property of the probability (20.144) is that it satisfies the semigroupequation

P (xctc|xata) =∫ ∞

−∞dxb P (xctc|xbtb)P (xbtb|xata). (20.186)

In the stochastic context this property is referred to as Chapman-Kolmogorov equa-

tion or Smoluchowski equation. It is a general property of processes which a without

Page 35: Economic and Path Integrals

1474 20 Path Integrals and Financial Markets

memory. Such processes are referred to a Markovian [12]. Note that the semigroupproperty (20.186) implies the initial condition

P (xcta|xata) = δ(xb − xa). (20.187)

In Fig. 20.17 we show that the property (20.186) is satisfied reasonably well byexperimental asset distributions, except for small deviations in the low-probabilitytails.

z

P><

(z)

Figure 20.17 Cumulative distributions obtained from repeated convolution integrals

of the 15-min distribution (from Ref. [13]). Apart from the far ends of the tails, the

semigroup property (20.186) is reasonably well satisfied.

One may verify the semigroup property also using the high-frequency distribu-tions of S&P 500 and NASDAQ 100 indices in Fig. 20.10 recorded by the minute,which display the Boltzmann distribution (20.77). If we convolute the minute distri-bution an integer number t times, the result aggress very well with the distributionof t-minute data. Only at the heavy tails of rare events do not follow this pattern

Note that due to the semigroup property (20.186) of the probabilities one hasthe trivial identity

〈f(x(tb), tb)〉 =∫

dxb f(xb, tb)P (xbtb|xata)

=∫

dx[∫

dxb f(xb, tb)P (xbtb|x t)]

P (x t|xata). (20.188)

In the mathematical literature, the expectation value on the left-hand side ofEq. (20.188) is written as

EE[f(x(tb), tb)|xata] ≡∫

dxb f(xb, tb)P (xbtb|xata). (20.189)

Page 36: Economic and Path Integrals

20.1 Fluctuation Properties of Financial Assets 1475

With this notation, the second line can be re-expressed in the form∫

dxEE[f(xb, tb)|x t]P (x t|xata) = EE[EE[f(xb, tb)|x t]|xata], (20.190)

so that we obtain the property of expectation values

EE[f(x(tb, tb))|xata] = EE[EE[f(xb, tb)|x t]|xata]. (20.191)

In mathematical finance, this complicated-looking but simple property is called thetowering property of expectation values.

Note that since P (xbtb|xata) is equal to δ(xb − xa) for ta = tb, the expectationvalue (20.189) has the obvious property

EE[f(x(ta), ta)|xata] = f(xa, ta) (20.192)

20.1.19 Time Evolution of Moments of Distribution

From the time-dependent distribution (20.165) it is easy to calculate the timedependence of the moments:

〈xn〉(t) ≡∫ ∞

−∞dx xn P (x, t) (20.193)

Inserting (20.165), we obtain

〈xn〉(t) =∫ ∞

dp

2πe−tH(p)

∫ ∞

−∞dx xneipx =

∫ ∞

∞dpe−tH(p)(−i∂p)nδ(p). (20.194)

After n partial integrations, this becomes

〈xn〉(t) = (i∂p)ne−tH(p)

p=0. (20.195)

All expansion coefficients cn of H(p) in Eq. (20.141) receive the same factor t, sothat the cumulants of the moments all grow linearly in time:

〈xn〉c(t) = −tH(n)(0) = t〈xn〉c(1) = tcn. (20.196)

-10 -5 0 5 10

-3

-2

-1

-50 0 50-3

-2

-1

z in % z in %

logP (z)logP (z)

Figure 20.18 Gaussian distributions of S&P 500 and NASDAQ 100 weekly log-returns.

Page 37: Economic and Path Integrals

1476 20 Path Integrals and Financial Markets

20.1.20 Boltzmann Distribution

As a an example, consider the Boltzmann distribution (20.77) of the minute data.The subsequent time-dependent variance and kurtosis are found by inserting thecumulants Eq. (20.80) into (20.196), yielding

〈x2〉c(t) = t 2T 2, κ(t) =c4tc22

=3

t. (20.197)

The first increases linearly in time, the second decreases like one over time, whichmakes the distribution more and more Gaussian, as required by the central limittheorem (20.167). The two quantities are plotted in Figs. 20.19 and 20.20. Theagreement with the data is seen to be excellent.

0 5 10 15 200

0.4

0.8

1.2

0 5 10 15 20-10

-5

0

5

10

0 5 10 15 200

6

12

18

0 5 10 15 20-10

-5

0

5

10

S&P500 2004-2005

NASDAQ100 2001-02

σ2(t)

σ2(t)

Time lag t (hours)

Time lag t (hours)

S&P500 2004-2005

NASDAQ100 2001-02

deviation in %

deviation in %

Time lag t (hours)

Time lag t (hours)

Figure 20.19 Variance of S&P 500 and NASDAQ 100 indices as a function of time.

The slopes 〈x2〉c(t)/t are roughly 1.1/20 × 1/60 min and 18.1/20 × 1/60 min, respectively,

so that Eq. (20.198) yields the temperatures TSP500 ≈ 0.075 and TNASDAQ100 ≈ 0.15. The

right-hand side shows the relative deviation from the linear shape in percent.

Note that if the minute data are not available, but only data taken with afrequency 1/t0 (in units 1/min) and an expectation 〈x2〉c(t0), then we can find thetemperature of the minute distribution from the formula

T =√

〈x2〉c(t0)/2t0. (20.198)

Since κ goes to zero like 1/t, the distribution becomes increasingly Gaussian asthe time grows, this being a manifestation of the central limit theorem (20.167) ofstatistical mechanics according to which the convolution of infinitely many arbitrarydistribution functions of finite width always approaches a Gaussian distribution.

Page 38: Economic and Path Integrals

20.1 Fluctuation Properties of Financial Assets 1477

0 5 10 15 200

1

2

3

0 5 10 15 20-20

-10

0

10

20

0 5 10 15 200

200

400

600

0 5 10 15 20-10

-5

0

5

10

S&P500 2004-2005

NASDAQ100 2001-02

κ(t)

κ(t)

Time lag t (hours)

Time lag t (hours)

S&P500 2004-2005

NASDAQ100 2001-02

deviation in %

deviation in %

Time lag t (hours)

Time lag t (hours)

Figure 20.20 Kurtosis of S&P 500 and NASDAQ 100 indices as a function of time. The

right-hand side shows the relative deviation from the 1/t behavior in percent.

This result is in contrast to the pure Levy distribution in Subsection 20.1.2 withλ < 2, which has no finite width and therefore maintains its power falloff at largedistances.

If we omit the time argument in the cumulants 〈xn〉c(1), for brevity, and insertthese into the decompositions (20.56) we obtain the time dependence of the mo-ments:

〈x〉(t) = t〈x〉c,

〈x2〉(t) = t〈x2〉c + t2〈x〉2c ,

〈x3〉(t) = t〈x3〉c + 3t2〈x〉c〈x2〉c + t3〈x〉3c ,

〈x4〉(t) = t〈x4〉c + 3t2〈x2〉2c − 4t2〈x〉c〈x3〉c + 6t3〈x〉2c〈x2〉c + t4〈x〉4c , (20.199)... .

Let us now calculate the time evolution of the Boltzmann distribution (20.77) ofthe minute data. The Hamiltonian H(p) was identified in Eq. (20.79), so that

e−tH(p) =1

[1 + (Tp)2]t. (20.200)

The time-dependent distribution is therefore given by the Fourier integral

P (x, t) =∫ ∞

−∞

dp

2πeipx−tH(p) =

∫ ∞

−∞

dp

1

(1 + T 2p2)teipx. (20.201)

Page 39: Economic and Path Integrals

1478 20 Path Integrals and Financial Markets

The calculation proceeds most easily rewriting (20.200), by analogy with (20.81)and using again formula (2.499), as an integral

e−tH(p) =1

[1 + (Tp)2]t=

1

Γ(t)

∫ ∞

0

ττ te−τ(1+T 2p2). (20.202)

The prefactor of p2 is equal to tσ2 = tv a Gaussian distribution in x-space. Wetherefore change the variable of integration from τ to v, substituting τ = tv/2T 2,and obtain [compare (20.82)]

e−tH(p) =(

t

2T 2

)t 1

Γ(t)

∫ ∞

0

dv

vvte−tv/2T 2

e−tvp2/2. (20.203)

The Fourier transform of this yields the time evolution of the Boltzmann distribu-tion as a time-dependent superposition of Gaussian distributions of different widths[compare (20.83)]:

P (x, t) =(

t

2T 2

)t 1

Γ(t)

∫ ∞

0

dv

vvte−tv/2T 2 1√

2πtve−x2/2tv. (20.204)

The integral can be performed using the integral formula (2.559), and we find thetime-dependent distribution

P (x, t) =∫ ∞

−∞

dp

2πeipx−tH(p) =

1

T√πΓ(t)

(

|x|2T

)t−1/2

Kt−1/2(|x|/T ), (20.205)

where t is measured in minutes.For t = 1, we reobtain the fundamental distribution (20.77), recalling the explicit

form of K1/2(z) in Eq. (1.349).In the limit of large t, the integral over v in (20.204) can be evaluated in the

saddle-point approximation of Section 4.2. We expand the weight function in theintegrand around the maximum at vm = 2T 2(1 − 1/t) as

vt−1e−tv/2T 2

=[

2T 2(1 − 1/t)]t−1

e−(t−1)e−tδv2/2[2T 2(1−1/t)]2[

1 + O(δv3)]

,(20.206)

where δv = v−vm, and O(δv3) is equal to −tδv3/3[2T 2(1−1/t)]3. Then we observethat for large t, the Gaussian can be expanded into derivatives of δ-functions asfollows:

e−tz2/2[

1 +a

3z3 + . . .

]

=√

2π/t[

δ(z) +1

2tδ′′(z) +

a

t2δ′(z) . . .

]

. (20.207)

This can easily be verified by multiplying both sides with f(z) = f(0) + zf ′(0) +z2f ′′(0)/2 + . . . and integrating over z. Thus we find to leading order

e−tδv2/2[2T 2(1−1/t)]2 →√

t2T 2(1 − 1/t) δ(δv). (20.208)

Page 40: Economic and Path Integrals

20.1 Fluctuation Properties of Financial Assets 1479

Using the large-t limit (1 − 1/t)t → e−1, and including the first correction from(20.207), we obtain

vt−1e−tv/2T 2 →√

t(2T 2)te−t

δ(δv) +[2T 2(1 − 1/t)]2

2tδ′′(δv) + . . .

. (20.209)

If we now employ Stirling’s formula (17.286) to approximate tt/Γ(t) →√

t/2πet,

we see that the integral (20.204) converges for large t against the Gaussian distri-bution e−x2/2tvm/

√2πtvm with the saddle point variance vm → 2T 2.

The behavior near the peak of the distribution can be calculated from the small-zexpansion7

(

z

2

Kν(z) =π

2 sin πν Γ(1 − ν)

[

1 +Γ(1 − ν)

1!Γ(2 − ν)

z2

4+ O(z4, z2ν)

]

. (20.210)

For large ν and small z, this can be approximated by [recall (20.29)]

(

z

2

Kν(z) ≈ Γ(ν)

2e−z2/4(t−3/2). (20.211)

We now use Stirling’s formula (17.286) to find the large-t limit Γ(t− 1/2)/Γ(t) → 1,leading to the Gaussian behavior near the peak:

P (x, t) ≈small |x|

1√2π 2T 2t

e−x2/2 2T 2(t−3/2). (20.212)

This corresponds to the approximation of the Fourier transform (20.200) by e−tT 2p2.The Gaussian shape reaches out to x ≈ T

√t. For very large t, it holds everywhere,

as it should by virtue of the central limit theorem (20.167).

20.1.21 Fourier-Transformed Tsallis Distribution

For the Hamiltonian defined in Eq. (20.97), the time evolution is given by the Fouriertransform

e−tHδ,β(p)=[

1 − βδ p2/2]−t/δ

=1

Γ(t/δ)

∫ ∞

0

ds

sst/δe−se−sβδ p2/2, β ≡ ν

µ=

1

µδ,(20.213)

which can be rewritten, by analogy with (20.99), as

e−tHδ,β(p) =µt/δ

Γ (t/δ)

∫ ∞

0

dv

vvt/δe−µve−vp2/2 , µ = 1/βδ. (20.214)

Taking the Fourier transform of this yields the time-dependent distribution function

Pδ,β(x, t) =µt/δ

Γ (t/δ)

∫ ∞

0

dv

vvt/δe−µv 1√

2πve−x2/2v , µ = 1/βδ, (20.215)

7M. Abramowitz and I. Stegun, op. cit., Formulas 9.6.2 and 9.6.10.

Page 41: Economic and Path Integrals

1480 20 Path Integrals and Financial Markets

which becomes with the help of Formula (2.499):

Pδ,β(x, t) =µt/δ

Γ (t/δ)

1√2π

(

x2

)t/2δ−1/4

2Kt/δ−1/2(√

2µx), µ =1

βδ. (20.216)

For δ = 1 and µ = 1/2T 2, this reduces to the time-dependent Boltzmann distribution(20.205). At the origin, its value is [compare (20.104)]

Pδ,β(0, t) =√µΓ(t/δ − 1/2)/Γ(t/δ). (20.217)

20.1.22 Superposition of Gaussian Distributions

All time-dependent distributions whose Fourier transforms have the form e−tH(p) pos-sess a path integral representation (20.161) and which obeys the semigroup equation(20.186). In several examples we have seen that the Fourier transforms e−H(p) canbe obtained from a superposition of Gaussian distributions of different variancese−vp2/2 with some weight function w(v):

e−H(p) =∫ ∞

0dv w(v)e−vp2/2. (20.218)

This was true for the Boltzmann distribution in Eq. (20.81), for the Fourier-transformed Tsallis distribution in Eq. (20.100), and for the Boltzmann Distributionof a relativistic particle in Eq. (20.105). In all these cases it was possible to write asimilar superposition for the time-dependent distributions:

e−tH(p) =∫ ∞

0dv′wt(v

′)e−v′p2/2 =∫ ∞

0dv ωt(v)e−tvp2/2, (20.219)

where we have introduced the time-dependent weight function

ωt(v) ≡ t wt(vt). (20.220)

See Eqs. (20.202), (20.214), and (20.105) for the above three cases.The question arises as to general property of the time dependence of the weight

function wt(v) which ensures that that the Fourier transform of the superposition(20.219) obtained as in Eq. (20.164) satisfies the semigroup condition (20.186) [52].For this, the superposition has to depend on time as

e−(t2+t1)H(p) = e−t2H(p)e−t1H(p),

or e−tH(p) = [e−H(p)]t. This implies that the weight factor wt(v) has the property∫

dv12wt1+t2(v12)e−v12p2/2 =

dv2wt2(v2)e−v2p2/2

dv1wt1(v1)e−v1p2/2. (20.221)

For the Laplace transforms wt(v) of wt(v)

wt(pv) ≡∫

dv e−pvvwt(v) (20.222)

Page 42: Economic and Path Integrals

20.1 Fluctuation Properties of Financial Assets 1481

this amounts to the factorization property

wt1+t2(pv) = wt2(pv)wt1(pv), (20.223)

which is fulfilled by the exponential

wt(pv) = e−tHv(pv), (20.224)

with some Hamiltonian Hv(pv). Since the integral over v in (20.222) runs only overthe positive v-axis, the Hamiltonian Hv(pv) is analytic in the upper half-plane of pv.

It is easy to relate Hv(pv) to H(p). For this we form the inverse Laplace trans-formation, the so-called Bromwich integral [53]

wt(v) =∫ γ+i∞

γ−i∞

dpv2πi

epvv−tHv(pv), (20.225)

in which γ is some real number larger than the real parts of all singularities ine−pvv−tHv(pv). Inserting this into (20.219) and performing the v-integral yields

e−tH(p) =∫ γ+i∞

γ−i∞

dpv2πi

1

p2/2 − pve−tHv(pv). (20.226)

Closing the integral over pv in the complex pv-plane and deforming the contour toan anticlockwise circle around pv = p2/2 we find the desired relation:

Hv(p2/2) = H(p). (20.227)

The time-dependent distribution associated with any Hamiltonian H(p) via theFourier integral (20.170) can be represented as a superposition of Gaussian distri-butions with the help of Eq. (20.219) if we merely choose Hv(pv) = H(

√2pv).

Recalling the derivation of the Fourier representation (20.165) from the pathintegral (20.153) and the descendence of that path integral from the stochastic dif-ferential equation (20.140) with the noise Hamiltonian H(p) we conclude that theHamiltonian Hv(ipv) governs the noise in a stochastic differential equation for thevolatility fluctuations.

Take, for example, the superposition of Gaussians (20.105). The Laplace trans-form of the weight function ωβ(v) is

ωβ(pv) =∫

dv e−vpvωβ(v) =∫

dv e−vpv

β

2πv3e−β/2v = e−

√2βpv . (20.228)

The Laplace transform of wβ(v) is related to this by∫

dv′e−v′pvwβ(v′)=∫

dv e−βvpvωβ(v) = ωβ(βpv). (20.229)

Hence it is given by wβ(pv)=e−β√2pv =e−βHv(pv), which satisfies the relation (20.223).

Moreover, Hv(p2/2) is equal to H(p) =

√p2, as required by Eq. (20.227) and in

accordance with Eq. (20.105) for M = 0.Note that instead of the Bromwich integral (20.226) it is sometimes more con-

venient to use Post’s Laplace inversion formula [54]

ωt(v) = limk→∞

(−1)k

k!xk+1 ∂

kωt(x)

∂xk

x=k/v

. (20.230)

Page 43: Economic and Path Integrals

1482 20 Path Integrals and Financial Markets

20.1.23 Fokker-Planck-Type Equation

From the Fourier representation (20.170) it is easy to prove that the probabilitysatisfies a Fokker-Planck-type equation

∂tP (xbtb|xata) = −H(−i∂x)P (xbtb|xata). (20.231)

Indeed, the general solution ψ(x, t) of this differential equation with the initial con-dition ψ(x, 0) is given by the path integral generalizing (20.144)

ψ(x, t) =∫

Dη exp[

−∫ tb

tadt H(η(t))

]

ψ(

x−∫ t

tadt′η(t′)

)

. (20.232)

This satisfies the Fokker-Planck-type equation (20.231). To show this, we takeψ(x, t) at a slightly later time t+ ǫ and expand

ψ(x, t + ǫ) =∫

Dη exp[

−∫ tb

tadt H(η(t))

]

ψ(

x−∫ t

tadt′η(t′) −

∫ t+ǫ

tdt′η(t′)

)

.

=∫

Dη exp[

−∫ tb

tadt H(η(t))

]

ψ(

x−∫ t

tadt′η(t′)

)

− ψ′(

x−∫ t

tadt′η(t′)

)∫ t+ǫ

tdt′η(t′)

+1

2ψ′′

(

x−∫ t

tadt′η(t′)

)∫ t+ǫ

tdt1dt2 η(t1)η(t2) (20.233)

− 1

3!ψ′′′

(

x−∫ t

tadt′η(t′)

)∫ t+ǫ

tdt1dt2dt3 η(t1)η(t2)η(t3)

+1

4!ψ(4)

(

x−∫ t

tadt′η(t′)

)∫ t+ǫ

tdt1dt2dt3dt4 η(t1)η(t2)η(t3)η(t4) + . . .

.

Using the correlation functions (20.155)–(20.262) we obtain

ψ(x, t+ ǫ) =∫

Dη exp[

−∫ tb

tadt H(η(t))

]

×[

−ǫc1∂x +(

ǫc2 + ǫ2c1) 1

2∂2x −

(

ǫc3 + 3ǫ2c2c1) 1

3!∂3x (20.234)

+(

ǫc4 + ǫ24c3c1 + ǫ23c22 + ǫ3c2c21 + ǫ4c21

) 1

4!∂4x + . . .

]

ψ(

x−∫ t

tadt′η(t′)

)

.

In the limit ǫ → 0, only the linear terms in ǫ contribute, which are all due to theconnected parts of the correlation functions of η(t). The differential operators in thebrackets can now be pulled out of the integral and we find the differential equation

∂tψ(x, t) =[

−c1∂x +c22!∂2x −

c33!∂3x +

c44!∂4x + . . .

]

ψ (x, t) . (20.235)

We now replace c1 → rx and express using (20.141) the differential operators inbrackets as Hamiltonian operator −Hrx(−i∂x). This leads to the Schrodinger-likeequation [55]

∂tψ(x, t) = −Hrx(−i∂x)ψ (x, t) . (20.236)

Page 44: Economic and Path Integrals

20.1 Fluctuation Properties of Financial Assets 1483

Due to the many derivatives in H(i∂x), this equation is in general non-local. Thiscan be made explicit with the help of the Levy-Khintchine weight function F (x) inthe Fourier representation (20.173). In this case, the right-hand side

−H(−i∂x)ψ(x, t) =∫

dx′ e−x′∂x F (x′)ψ(x, t) =∫

dx′ F (x′)ψ(x− x′, t), (20.237)

and the Fokker-Planck-like equation (20.236) takes the form of an integro-differentialequation. Some people like to use the subtracted form (20.175) of the Levy-Khintchine formula and arrive at the integro-differential equation

∂tψ(x, t) =[

−c1∂x −c22∂2x

]

ψ(x, t) +∫

dx′ F (x′)ψ(x− x′, t). (20.238)

The integral term can then be treated as a perturbation to an ordinary Fokker-Planck equation.

The initial condition of the probability is, of course, always given by (20.187),as a consequence of the semigroup property (20.186).

20.1.24 Kramers-Moyal Equation

The Fokker-Planck-type equation (20.231) is a special case of the most general timeevolution equation satisfied by a probability distribution P (xb, tb, |xata) which fulfillsthe Chapman-Kolmogorov or Smoluchowski equation (20.186). For a short timeinterval tc − tb = ǫ, that equation can be written as

P (xc tb+ǫ|xata) =∫ ∞

−∞dxb P (xctb + ǫ|xbtb)P (xbtb|xata). (20.239)

We now re-express P (xctb+ǫ|xbtb) as a trivial integral

P (xctb+ǫ|xbtb) =∫

dx δ(x−xb+xb−xc)P (x tb+ǫ|xbtb), (20.240)

and expand the δ-function in powers of x− xb to obtain

P (xctb+ǫ|xbtb) =∫

dx∞∑

n=0

(x− xb)n

n!P (x tb+ǫ|xbtb) ∂nxb

δ(xb − xc). (20.241)

Introducing the moments

Cn(xbtb) ≡∫

dx (x− xb)n∂tbP (x tb|xbtb), (20.242)

and inserting (20.241) into Eq. (20.239) the right-hand side reads

P (xctb|xbtb) + ǫ∞∑

n=1

dxbCn(xbtb)

n![∂nxb

δ(xb − xc)]∂tbP (xbtb|xata) + O(ǫ2). (20.243)

Page 45: Economic and Path Integrals

1484 20 Path Integrals and Financial Markets

Taking the n = 0 -term to the left-hand side of (20.239) and going to the limitǫ→ 0, we obtain the Kramers-Moyal equation:

∂tbP (xbtb|xata) = −H(−i∂xb, xb, tb)P (xbtb|xata). (20.244)

with the Hamiltonian operator

H(−i∂xb, xb, tb) ≡ −

∞∑

n=1

(−∂xb)nCn(xbtb)

n!, (20.245)

which is the generalization of the Hamiltonian operator in Eqs. (20.235) and(20.231).

There is a simple formal solution of the time evolution equation (20.244) withthe initial condition (20.187). It can immediately be written down by recalling thesimilar time evolution equation (1.233):

P (xbtb|xata) = T e−∫ tbta

dtH(−i∂xb ,xb,t)δ(xb − xa), (20.246)

where T is the time-ordering operator (1.241). Using Dirac bra-ket states 〈xb| and|xa〉 with the scalar product 〈xb|xa〉 = δ(xb − xa) as in Eq. (18.462), and the rules(18.463), we can rewrite (20.246) as

P (xbtb|xata) = 〈xb|T e−∫ tbta

dtH(−i∂xb ,xb,t)|xa〉. (20.247)

Due to the initial condition (20.187), the moments (20.242) can also be calculatedfrom the short-time limit

Cn(x t) =ǫ→0

1

ǫ

dx (x′ − x)nP (x′ t+ǫ|x t) =ǫ→0

1

ǫ〈[x(t+ǫ) − x(t)]n〉, n ≥ 1. (20.248)

Another way of writing the expectation value on the right-hand side is

Cn(x t) =ǫ→0

1

ǫ

∫ t+ǫ

tdt1 · · ·

∫ t+ǫ

tdtn 〈x(t1) · · · x(tn)〉, n ≥ 1. (20.249)

If x(t) follows the Langevin equation (20.169) with the correlation functions(20.155)–(20.262), we obtain

Cn(x t) = cn, (20.250)

and the Kramers-Moyal equation (20.248) reduces to out previous Fokker-Planck-type equation (20.231).

For a Langevin equation

x(t) = r(x, t) + σ(x, t)η(t), (20.251)

with a white noise variable η(t), we find instead

C1(x t)=a(x, t)+b′(x, t)b(x, t)=a(x, t)+∂xC2(x t)

2, C2(x t) = b2(x, t). (20.252)

According to a theorem due to Pawula [56], the positivity of the probability inthe Kramers-Moyal equation (20.244) requires that the expansion (20.245) can onlybe truncated after the first or second terms, or must go on to infinity.

Page 46: Economic and Path Integrals

20.2 Ito-like Formula for Non-Gaussian Distributions 1485

20.2 Ito-like Formula for Non-Gaussian Distributions

By a procedure similar to that in Subsection 18.13.3 it is possible to derive a gen-eralization of the Ito-like expansions (18.412) and (18.432) for the fluctuations offunctions containing non-Gaussian noise.

20.2.1 Continuous Time

As in (18.408) we expand f(x(t+ ǫ)):

f(x(t+ ǫ)) = f(x(t)) + f ′(x(t))∆x(t)

+1

2f ′′(x(t))[∆x(t)]2 +

1

3!f (3)[∆x(t)]3 + . . . . (20.253)

where x(t) = η(t) is the stochastic differential equation with a nonzero expectationvalue 〈η(t)〉 = c1. In contrast to the expansion (18.426), which for ǫ → 0 had tobe carried only up to the second order in ∆x(t) ≡ ∫ t+ǫ

t dt′ x(t′), we must now keepall orders. Evaluating the noise averages of the multiple integrals on the right-hand side with the help of the correlation functions (20.155)–(20.262), we find thetime dependence of the expectation value of an arbitrary function of the fluctuatingvariable x(t)

〈f(x(t+ǫ))〉 = 〈f(x(t))〉 + 〈f ′(x(t))〉ǫc1 +1

2〈f ′′(x(t))〉(ǫc2+ǫ2c21)

+1

3!〈f (3)(x(t))〉(ǫc3 + ǫ2c2c1 + ǫ3c31) + . . . (20.254)

= 〈f(x(t))〉 + ǫ[

c1∂x + c21

2∂2x + c3

1

3!∂3x + . . .

]

〈f(x(t))〉 + O(ǫ2).

After the replacement c1 → rx, the function f(x(t)) obeys therefore the followingequation:

〈f(x(t))〉 = −Hrx(i∂x)〈f(x(t))〉. (20.255)

Separating the lowest-derivative term, this takes a form generalizing Eq. (18.412):

〈f(x(t))〉 = 〈f ′(x(t))〉〈x(t)〉 − Hrx(i∂x)〈f(x(t))〉. (20.256)

This might be viewed as the expectation value of the stochastic differential equation

f(x(t)) = f ′(x(t))x(t) − Hrx(i∂x)f(x(t)), (20.257)

which, if valid, would be a simple direct generalization of Ito’s Lemma (18.413).However, this conclusion is not allowed. The reason lies in the increased size of thefluctuations of the higher expansion terms [∆x(t)]n for n ≥ 2. In the harmonic caseof Subsection 18.13.3, these were negligible in comparison with the leading termz1(t) in Eq. (18.413). Let us see what happens in the present case, where all higher

Page 47: Economic and Path Integrals

1486 20 Path Integrals and Financial Markets

cumulants cn in the Hamiltonian (20.141) may be nonzero. For the argument weproceed as in Eq. (18.405) by working with the stochastic differential equation

x(t) = 〈x(t)〉 + η(t) = c1 + η(t). (20.258)

rather that (20.169). Then the correlation functions of η(t) consist only of theconnected parts of (20.155)–(20.262):

〈η(t1)〉 = c1, (20.259)

〈η(t1)η(t2)〉 = c2δ(t1 − t2) (20.260)

〈η(t1)η(t2)η(t3)〉 = c3δ(t1−t2)δ(t1−t3), (20.261)

〈η(t1)η(t2)η(t3)η(t4)〉 = c4δ(t1−t2)δ(t1−t3)δ(t1−t4) (20.262)

+ c22[δ(t1−t2)δ(t3−t4)+δ(t1−t3)δ(t2−t4)+δ(t1−t4)δ(t2−t3)].... .

We now estimate the size of the fluctuations zn of [∆x(t)]n. The first contributionz2,1 to z2 defined in Eq. (18.415) is still negligible since it is smaller than the leadingone z1(t) ≡ ∫ t+ǫ

t dt′ η(t′) by a factor ǫ. The second contribution z2,3 to z2 definedin Eq. (18.415), however, has now a larger variance, due to the c4-term in (20.262),which contains one more δ-function than the c22-term, so that

[z2,2(t)]2⟩

=∫ t+ǫ

tdt1

∫ t+ǫ

tdt2

∫ t+ǫ

tdt3

∫ t+ǫ

tdt4 〈η(t1)η(t2)η(t3)η(t4)〉=ǫc4+ǫ2c22. (20.263)

This is larger than the harmonic estimate (18.420) by a factor 1/ǫ, which makesz2,2(t), in general, as large as the leading fluctuation z1(t) of x1(t). It can thereforenot be ignored. The subtraction of the second term 〈z2,2(t)〉2 = ǫ2c22 in the variancedoes not help since that is ignorable.

A similar estimate holds for the higher powers. Take for instance [∆x(t)]3:

[∆x(t)]3 = ǫ3c31 + 3ǫ2c21z1(t) + ǫc1[z1(t)]2 + [z1(t)]

3, (20.264)

and calculate the variance of the strongest fluctuating last term in this:〈[z1(t)]

32〉 − 〈[z1(t)]3〉2〉. The first contribution is equal to

∫ t+ǫ

tdt1

∫ t+ǫ

tdt2

∫ t+ǫ

tdt3

∫ t+ǫ

tdt4

∫ t+ǫ

tdt5

∫ t+ǫ

tdt6 〈η(t1)η(t2)η(t3)η(t4)η(t5)η(t6)〉,

(20.265)and becomes, after an obvious extension of the expectation values (20.155)–(20.262)to the six-point function:

〈[z1(t)]32〉 = ǫc6 + O(ǫ2). (20.266)

The second contribution 〈[z1(t)]3〉2 in (20.264) is equal to ǫ2c23 and can be ignoredfor ǫ → 0. Thus, due to (20.266), the size of the fluctuations [z1(t)]

3 is of the sameorder as of the leading one x1(t) [compare again (18.421)].

Page 48: Economic and Path Integrals

20.2 Ito-like Formula for Non-Gaussian Distributions 1487

This is an important result which is the reason why we cannot derive a directgeneralization (20.257) of the Ito formula (18.413) to non-Gaussian fluctuations, butonly the weaker formula (20.256) for the expectation value.

For an exponential function f(x) = ePx this implies the relation

d

dt〈ePx(t)〉 =

[

P 〈x(t)〉 − Hrx(iP )] ⟨

ePx(t)⟩

, (20.267)

and it is not allowed to drop the expectation values.A consequence of the weaker Eq. (20.267) for P = 1 is that the rate rS with

which the average of the stock price S(t) = ex(t) grows is given by formula (20.1),where the rate rx is related to rS by

rS = rx − H(i) = rx − [H(i) − iH ′(0)] = −Hrx(i), (20.268)

which replaces the simple Ito relation rS = rx + σ2/2 in Eq. (20.5). Recall thedefinition of H(p) ≡ H(p) − H ′(0)p in Eq. (20.141). The corresponding version ofthe left-hand part of Eq. (20.4) reads⟨

S

S

= 〈x(t)〉 − H(i) = 〈x(t)〉 − [H(i) − iH ′(0)] = 〈x(t)〉 − rx −Hrx(i). (20.269)

The forward price of a stock must therefore be calculated with the generalization offormula (20.8), in which we assume again η(t) to fluctuate around zero rather thanrx:

〈S(t)〉 = S(0)erSt = S(0)〈erxt+∫ t

0dt′ η(t′)〉 = S(0)e−Hrx (i)t = S(0)erxt−[H(i)−iH′(0)]t.

(20.270)If η(t) fluctuates around rx this takes the simpler form

〈S(t)〉 = S(0)erSt = S(0)〈e∫ t

0dt′ η(t′)〉 = S(0)e−Hrx (i)t. (20.271)

This result may be viewed as a consequence of the following generalization of theGaussian Debye-Waller factor (18.425):

eP∫ t

0dt′ η(t′)

= e−Hrx (iP )t. (20.272)

Note that we may derive the differential equation (20.255) of an arbitrary func-tion f(x(t)) from a simple mnemonic rule, expanding sloppily [similar to Eq. (18.426)but restricted to the expectation values]

〈f(x(t+ dt))〉 = 〈f(x(t))〉 + 〈f ′(x(t))〉 〈x(t)〉 dt+1

2〈f ′′(x(t))〉

x2(t)⟩

dt2

+1

3!〈f (3)(x(t))〉〈x3(t)〉dt3 + . . . , (20.273)

and replacing

〈x(t)〉dt→ c1dt, 〈x2(t)〉dt2 → c2dt, 〈x3(t)〉dt3 → c3dt, . . . . (20.274)

In contrast to Eq. (18.429), this replacement holds now only on the average.For the same reason, portfolios containing assets with non-Gaussian fluctuations

cannot be made risk-free in the continuum limit ǫ→ 0, as will be seen in Section 20.7.

Page 49: Economic and Path Integrals

1488 20 Path Integrals and Financial Markets

20.2.2 Discrete Times

The prices of financial assets are recorded in the form of a discrete time seriesx(tn) in intervals ∆t = tn+1 − tn, rather than the continuous function x(t). Forstocks with a large turnover, or for market indices, the smallest time interval istypically ∆t = 1 minute. We have seen at the end of Section 18.13.3 that this makesIto’s Lemma (18.413) an approximate statement. Without the limit ∆t → 0, thefluctuations of the higher expansion terms in (20.253) no longer disappear but aremerely suppressed by a factor σ

√∆tn. In quiet economic periods this is usually

quite small for ∆t = 1 minute, so that the higher-order fluctuations can usually beneglected after all.

While this approximate validity is a disadvantage of the discrete time seriesover the continuous one, it is an advantage with respect to processes with non-Gaussian noise, at least as long as the financial markets are not in turmoil. Thenthe non-Gaussian higher-order fluctuations of the expansion terms in (20.253) aresuppressed by the same order σ

√∆t as the corrections to Ito’s rule in the Gaussian

case [recall (18.432)]. As a typical example, take the Boltzmann distribution (20.77).Its cumulants cn carry a factor T n [see Eq. (20.80)] where the market temperature Tis a small number of the order of a few percent in the natural time units of one minuteused in this context [see Fig. (20.11)]. The smallness of T is, of course, a consequencethat the minute data do not usually possess large volatility, except near a crash. Inthe natural units of minutes, the time interval ǫ in the calculations after Eq. (20.257)is unity, so that powers of ǫ can no longer be used for size estimates. This role isnow taken over by T . In fact, since cn is of order T n, the fluctuations of zn(t) areof the order T n. Thus, in non-Gaussian fluctuations of a discrete time series, thesmallness of T leads to a suppression of the higher fluctuations after all. Moreover,the suppression is just as good as for time series with Gaussian fluctuations, wherethe corrections are of the order (σ

√∆t)n−1 for zn(t) with n ≥ 2. These have the size

of T n−1 for the Boltzmann distribution [recall (20.198)]. A similar estimate holdsfor all non-Gaussian noise distributions with semi-heavy tails as defined on at theend of Subsection 20.1.1.

For all semi-heavy tails for which the cumulants cn decrease with power T n of asmall parameter such as the temperature T , we may drop the expectations valuesof the Ito-like expansion (20.256), and remain with an approximate Ito formula forthe fluctuating difference ∆f(x(tn)) ≡ f(x(tn+1)):

∆f(x(tn)) = f ′(x(tn))∆x(tn) − Hrx(i∂x)f(x(tn)) + O(σ√

∆t), (20.275)

which is a direct generalization of the discrete version (18.432) of Ito’s Lemma(18.413).

A characteristic property of the Boltzmann distribution (20.77) and others withsemi-heavy tails is that their Hamiltonian possesses a power series in p2 of the type(20.141) with finite cumulants cn of the order of σn, where σ is the volatility of theminute data. This property is violated only by power tails of price distributionswhich do exist in the data (see Fig. 20.10). These do not go away upon multiple

Page 50: Economic and Path Integrals

20.3 Martingales 1489

convolution which turns the Boltzmann distribution more and more into a Gaussianas required by the central limit theorem (20.167). These heavy tails are caused bydrastic price changes in short times observed in nervous markets near a crash. Ifthese are taken into account, the discrete version (18.432) of Ito’s Lemma (18.413)is no longer valid. This will be an obstacle to setting up a risk-free portfolio inSection 20.7.

20.3 Martingales

In financial mathematics, an often-encountered concept is that of a martingale. Thename stems from a casino strategy in which a gambler doubles his stake each time abet is lost. A stochastic variable m(t) is called a martingale, if its expectation valueis time-independent.

A trivial martingale is provided by any noise variable m(t) = η(t).

20.3.1 Gaussian Martingales

For a harmonic noise variable, the exponential m(t) = e∫ t

0dt′η(t′)−σ2t/2 is a nontrivial

martingale, due to Eq. (20.8). For the same reason, a stock price S(t) = ex(t) withx(t) obeying the stochastic differential equation (20.140) can be made a martingaleby a time-dependent multiplicative factor associated with the average growth raterS, i.e.,

e−rStS(t) = e−rStex(t) = e−rSterxt+∫ t

dt′ η(t′) (20.276)

is a martingale. The prefactor e−rSt is referred to as a discount factor with therate rS. If we calculate the probability distribution (20.170) associated with thestochastic differential equation (20.140), which for harmonic fluctuations of standarddeviation σ corresponds to a Hamiltonian

H(p) = irxp+σ2

2p2, (20.277)

The integral representation (20.170) for the probability distribution,

P rx(xbtb|xata) =∫ ∞

−∞

dp

2πexp

[

ip(xb − xa) − ∆t(

irxp+ σ2p2/2)]

(20.278)

with ∆t = tb− ta has obviously a time-independent expectation value of e−rStS(t) =e−rStex(t). Indeed, the expectation value at the time tb is given by the integral overxb

e−rStb

dxb exbP rx(xbtb|xata) = e−rStb

dxb exb

∫ ∞

−∞

dp

2πexp

[

ip(xb− xa)−∆t σ2p2/2]

,

which yields

e−rStb

∫ ∞

−∞

dp

2πδ(p− i)e−ipxae∆t σ2p2/2 = e−rStexae(rx−σ2/2)∆t = e−rStaexa , (20.279)

Page 51: Economic and Path Integrals

1490 20 Path Integrals and Financial Markets

where we have used the Ito relation rS = rx +σ2/2 of Eq. (20.5). The result impliesthat

〈e−rStbS(tb)〉rx = 〈e−rStbex(tb)〉rx = e−rStaS(ta). (20.280)

Since this holds for all tb we may drop the subscripts b, thus proving the timeindependence and thus the martingale nature of e−rStS(t).

In mathematical finance, where the expectation value in Eq. (20.280) is writtenaccording to the definition (20.189) as EE[e−rS(tb−ta)exb|xata], the martingale propertyof e−rS(tb−ta)S(t) is expressed as

EE[e−rS(tb−ta)S(tb)|xata] = S(ta) (20.281)

or as EE[e−rS(tb−ta)exb |xata] = exa .For the martingale f(xb, tb) = e−rS(tb−ta)exb , the expectation values on the right-

hand side of the towering formula (20.191) reduces with the help of (20.281) and(20.192) to

EE[EE[f(xb, tb)|x t]|xata] = EE[EE[f(x, t)|x t]|xata] = EE[f(x, t)|xata]. (20.282)

Using once more the relations (20.281) and (20.192) to leads to

EE[EE[f(xb, tb)|x t]|xata] = f(xa, ta). (20.283)

Performing the momentum integral in (20.278) gives the explicit distribution

P rx(xbtb|xata) ≡1

2πσ2(tb − ta)exp

− [xb − xa − rx(tb − ta)]2

2σ2(tb − ta)

. (20.284)

We may incorporate the discount factor e−rS(tb−ta) into the probability distribution(20.284) and define a martingale distribution for the stock price

P (M,rx)(xbtb|xata) = e−rS(tb−ta)P rx(xbtb|xata), (20.285)

whose normalization falls off like e−rS(tb−ta). If we define expectation values withrespect to PM,rx(xbtb|xata) by the integral (without normalization)

〈f(xb)〉(M,rx) ≡∫

dxb f(xb)P(M,rx)(xbtb|xata), (20.286)

then the stock price itself is a martingale:

〈exb〉(M,rx) = exa . (20.287)

Note that there exists an entire family of equivalent martingale distributions

P (M,r)(xbtb|xata) = e−r∆t∫ ∞

−∞

dp

2πexp [ip(xb − xa) − ∆tHr(p)] , (20.288)

with an arbitrary rate r and rx ≡ r + H(i). Indeed, multiplying this with exb andintegrating over xb gives rise to a δ-function δ(p − i) and produces the same resultexa for any time difference ∆t = tb − ta. Performing the integral yields

P (M,r)(xbtb|xata) ≡e−r(tb−ta)

2πσ2(tb − ta)exp

− [xb − xa − r(tb − ta)]2

2σ2(tb − ta)

. (20.289)

These martingale distributions are called risk-neutral.

Page 52: Economic and Path Integrals

20.3 Martingales 1491

20.3.2 Non-Gaussian Martingale Distributions

For S(t) = ex(t) with an arbitrary non-Gaussian noise η(t), there are many ways ofconstructing distributions which make the stock price a martingale.

Natural Martingale Distribution

The relation (20.268) allows us to write down immediately the simplest martingaledistribution. It is given by an obvious generalization of the Gaussian expression(20.276):

e−rStS(t) = e−rSterxt+∫ t

0dt′η(t′) (20.290)

where rS and rx are related by rS = rx − H(i) = −Hrx(i). It is easy to provethis. The distribution function associated with the stochastic differential equation(20.140) with non-Gaussian fluctuations is given by

P rx(xbtb|xata) =∫

Dη∫

Dx exp

−∫ tb

tadt Hrx(η(t))

δ[x− rx(tb − ta)η], (20.291)

where ∆t ≡ (tb − ta) and rx = rS + H(i). As explained in Subsection 20.1.23, thepath integral is solved by the Fourier integral [compare (20.170)]

P rx(xbtb|xata) =∫ ∞

−∞

dp

2πexp [ip(xb − xa) − ∆tHrx(p)] . (20.292)

Using this distribution we can again calculate the time independence of the expecta-tion value (20.280), so that PM,rx(xbtb|xata) = e−rS(tb−ta)P rx(xbtb|xata) is a martin-gale distribution which makes the stock price time-independent, as in Eq. (20.287).

Note that there exists an entire family of equivalent martingale distributions

P (M,r)(xbtb|xata) = e−r∆t∫ ∞

−∞

dp

2πexp [ip(xb − xa) − ∆tHrx(p)] , (20.293)

with an arbitrary rate r and rx ≡ r + H(i). Indeed, multiplying this with exb andintegrating over xb gives rise to a δ-function δ(p− i) and produces the same resultexa for any time difference ∆t = tb − ta.

For non-Gaussian fluctuations there exists an infinite set martingale distributionsfor the stock price of which we are now going to dicuss the one proposed by Esscher.

Esscher Martingales

In the literature on mathematical finance, much attention is paid to another familyof equivalent martingale distributions. It has been used a long time ago to estimaterisks of actuaries [67] and introduced more recently into the theory of option prices[68, 69] where it is now of wide use [70]–[76]. This family is constructed as follows.Let D(z) be an arbitrary distribution function with a Fourier transform

D(z) =∫ ∞

−∞

dp

2πe−H(p) eipz, (20.294)

Page 53: Economic and Path Integrals

1492 20 Path Integrals and Financial Markets

and H(0) = 0, to guarantee a unit normalization∫

dzD(z) = 1. We now introducean Esscher-transformed distribution function. It is obtained by slightly tilting theinitial distribution D(z), multiplying it with an asymmetric exponential factor eθz:

Dθ(z) ≡ eH(iθ) eθzD(z). (20.295)

The constant prefactor eH(iθ) is necessary to conserve the total probability. Thisdistribution can be written as a Fourier transform

Dθ(z) =∫ ∞

−∞

dp

2πe−Hθ(p) eipz, (20.296)

with the Esscher-transformed Hamiltonian

Hθ(p) ≡ H(p+ iθ) −H(iθ). (20.297)

Since Hθ(0) = 0, the transformed distribution is properly normalized:∫

dzDθ(z) =1. We now define the Esscher-transformed expectation value

〈F (z)〉θ ≡∫

dzDθ(z)F (z). (20.298)

It is related to the original expectation value by

〈F (z)〉θ ≡ eH(iθ) 〈eθzF (z)〉. (20.299)

For the specific function F (z) = ez, Eq. (20.299) becomes

〈ez〉θ = e−Hθ(i) ≡ eHθ(iθ) 〈e(θ+1)z〉 = eH

θ(iθ)−Hθ(iθ+i). (20.300)

Applying the transformation (20.296) to each time slice in the general pathintegral (20.144), we obtain the Esscher-transformed path integral

P θ(xbtb|xata)=e−rS∆teHrx (iθ)∆t∫

Dη∫

Dx exp∫ tb

tadt[

θη(t)−Hrx(η(t))]

δ[x− η],

(20.301)

where ∆t ≡ tb − ta is the time interval. This leads to the Fourier integral [compare(20.293)]

P θ(xbtb|xata) = e−rS∆teHrx (iθ)∆t∫ ∞

−∞

dp

2πexp [ip(xb − xa) − ∆tHrx(p+ iθ)] .(20.302)

Let us denote the expectation values calculated with this probability by 〈 . . . 〉θ.Then we find for S(t) = ex(t) the time dependence

〈S(t)〉θ = e−Hθrx

(i)t. (20.303)

Page 54: Economic and Path Integrals

20.4 Origin of Semi-Heavy Tails 1493

This equation shows that the exponential of a stochastic variable x(t) can be madea martingale with respect to any Esscher-transformed distribution if we remove theexponential growth factor exp(rθ∆t) with

rθ ≡ −Hθrx(i) = −Hrx(i+ iθ) +Hrx(iθ). (20.304)

Thus, a family of equivalent martingale distributions for the stock price S(t) = ex(t)

is

PM θ(xbtb|xata) ≡ e−rθtP θ(xbtb|xata), (20.305)

for any choice of the parameter θ.

For a harmonic distribution function (20.284), the Esscher martingales and theprevious ones are equivalent. Indeed, starting from (20.284) in which rS = rx+σ2/2,the Esscher transform leads, after a quadratic completion, to the family of naturalmartingales (20.284) with the rate parameter r = rx + θσ2.

Other Non-Gaussian Martingales

Many other non-Gaussian martingales have been discussed in the literature. Mathe-maticians have invented various sophisticated criteria under which one would bepreferable over the others for calculating financial risks. Davis, for instance, hasintroduced a so-called utility function [77] which is supposed to select optimal mar-tingales for different purposes. There exists also a so-called minimal martingale [83],but the mathematical setup in these discussions is hard to understand.

For the upcoming development of a theory of option pricing, only the initialnatural martingale will be relevant.

20.4 Origin of Semi-Heavy Tails

We have indicated at the end of Subsection 20.1.1 that semi-heavy tails of the typeused in the previous discussions may be viewed as a phenomenological descriptionof a nontrivial Gaussian process with fluctuating volatility. This has been confirmedin Subsection 20.1.22 where we have expressed a number of non-Gaussian distribu-tions as a superposition of Gaussian distributions. We have further seen that theweight functions wt(v) for the superpositions can be Laplace-transformed to find aHamiltonian Hv(pv) from which we can derive which governs the noise distributiona stochastic differential equation for the volatility whose noise distribution is whichgoverned by Hv(ipv).

There exist a simple solvable model due to Heston [78] which shows thisexplicitly.8

8The discussion in this section follows a paper by Dragulescu and Yakovenko [79].

Page 55: Economic and Path Integrals

1494 20 Path Integrals and Financial Markets

20.4.1 Pair of Stochastic Differential Equations

Starting point is a coupled pair of ordinary stochastic differential equations withGaussian noise, one for the fluctuating logarithm x(t) of the stock price, and onefor the fluctuating time-dependent variance σ2(t). Since this will appear frequentlyin the upcoming discussion, we shall name it v(t). For simplicity, we remove thegrowth rate of the share price. Replacing σ2 in the stochastic differential equation(20.4) by the time-dependent variance v(t) we obtain

x(t) = −v(t)

2+√

v(t)η(t), (20.306)

where the noise variable η(t) has zero average and unit volatility

〈η(t)〉 = 0, 〈η(t)η(t′)〉 = δ(t− t′). (20.307)

The variance is assumed to satisfy an equation with another noise variable ηv(t) [7]

v(t) = −γ[v(t) − v] + ε√

v(t)ηv(t). (20.308)

The time in√

v(t) is supposed to lie slightly before the time in the noise ηv(t). Theparameter ε determines the volatility of the variance. The parameter v will becomethe average of the variance at large times. It is proportional to the variance σ2 ofthe Gaussian distribution to which the distribution converges in the long-time limitby the central limit theorem. The precise relation will be given in Eq. (20.363).

In general, the noise variables η(t) and ηv(t) are expected to exhibit correlationswhich are accounted for by introducing an independent noise variable η′(t) andsetting

ηv(t) ≡ ρ η(t) +√

1 − ρ2η′(t). (20.309)

The pair correlation functions are then

〈η(t)η(t′)〉 = δ(t− t′), 〈η(t)ηv(t′)〉 = ρδ(t− t′), 〈ηv(t)ηv(t′)〉 = δ(t− t′). (20.310)

20.4.2 Fokker-Planck Equation

If (t) denotes the pair of noise variables (η(t), ηv(t)), the probability distributionanalogous to (18.324) for paths x(t), v(t) starting out at xa, va and arriving at x, vis

P(x v t |xavata) = δ(xη(t) − x)δ(vηv(t) − v). (20.311)

The time evolution of this is calculated using the differentiation rule (18.381) as

∂tP(x v t |xavata) = − [∂xxη(t) + ∂vvηv(t)]P(x v t |xavata), (20.312)

Page 56: Economic and Path Integrals

20.4 Origin of Semi-Heavy Tails 1495

or more explicitly

∂tP(x v t |xavata)=

∂x

[

v(t)

2−√

v(t)η(t)

]

+ ∂v

[

γ (v(t) − v) − ε√

v(t)ηv(t)]

P(x v t |xavata). (20.313)

We now take the noise average (18.325) with the distribution

P []=exp

−1

2

dt[

η2(t)+η′2(t)]

=exp

− 1

1−ρ2∫

dt[

η2(t)+η2v(t)−2ρ ηηv]

.

(20.314)Applying the rules (18.385) and (18.386), we may replace η(t) and ηv(t) on theright-hand side of (20.313):

η(t) → −δ/δη(t) − ρ δ/δηv(t) (20.315)

ηv(t) → −ρ δ/δη(t) − δ/δηv(t). (20.316)

The functional derivatives are evaluated by analogy with (18.388) as

δ

δη(t′)δ(xη(t) −x)δ(vηv(t) −v) = −

[

δxη(t)

δη(t′)∂x +

δvηv(t)

δη(t′)∂v

]

δ(xη(t) −x)δ(vηv(t) −v),

(20.317)

with a similar equation for δ/δηv(t). Under the above-made assumption that√

v(t)

lies before ηv(t), we find the evolution equation

∂tP (x v t |xavata) = −H P (x v t |xavata), (20.318)

where H is the Hamiltonian operator

H = −1

2∂2x v −

1

2∂x v −

ε2

2∂2v v − γ∂v(v − v) − ρε∂x∂v v. (20.319)

The probability distribution can thus be written as a path integral

P (xbvbtb|xavata) =∫

Dx∫ Dp

Dv∫ Dpv

× exp∫ tb

tadt [i(px + pvv) −H(p, pv, v)]

, (20.320)

with the Hamiltonian

H(p, pv, v)=p2

2v−i1

2pv +ε2

p2v2v−iγpv(v − v)+ρε ppv v−

3ε2

4ipv−

γ

2− i

2ρǫp. (20.321)

The last two terms arise from the fact that the order of the operators in H obtainedfrom the path integral is always symmetric in v and pv, such that the order inEq. (20.319) is reached only after the replacement of pvv → (1/2)pv, v = pvv+ i/2

Page 57: Economic and Path Integrals

1496 20 Path Integrals and Financial Markets

in the fourth and fifth terms of (20.321). Recall the discussions in Sections 10.5,11.3, and Subsection 18.9.2.

The third-last term in (20.321) is present to ensure the appearance of the operator∂2v v in (20.319) with no extra ∂v due to operator ordering. The term p2vv in (20.321)can be understood as a kinetic term gµνp

µpν in a one-dimensional Riemannian spacewith a metric gµν = δµν/v. According to Subsection 11.1.1, this turns into theLaplace-Beltrami operator ∆ =

√v∂2v

√v∂v= v∂2v + 1

2∂v= ∂2vv − 3

2∂v. The extra ∂v

is canceled by the third-last term in (20.319).The stochastic differential equation (20.308) can, in principle, lead to negative

variances. Since this would be unphysical, we must guarantee that this does nothappen. The condition for this is that the fluctuation width of the variance issufficiently small, satisfying

ε2

2γv≤ 1. (20.322)

Then an initially positive v(t) will always stay positive. Indeed, consider the partialdifferential equation (20.318) for v = 0:

(∂t + α∂v)P (x v t |xavata) = (γ + ρε∂x)P (x v t |xavata), (20.323)

where α ≡ γv − ε2/2. It is solved by some function

P (x v t |xavata) = f(v − αt, x), (20.324)

which shows that a nonvanishing f for positive v can never propagate to negative-v.9

The time dependence of the variance is given by the integral

P (v t |vata) ≡∫

dxP (x v t |xavata). (20.325)

It satisfies the equation

∂tP (v t |vata) =

[

ε2

2∂2vv + γ∂v(v − v)

]

P (v t |vata), (20.326)

which follows from (20.318) by integration over x. After a long time, the solutionbecomes stationary and reads

P ∗(v) =µν

Γ(ν)vν−1e−µv, where µ ≡ 2γ

ε2, ν ≡ µv. (20.327)

This is the Gamma distribution found in Fig. 20.3. Hence the pair of stochasticdifferential equations (20.306) and (20.308) produces precisely the type of volatilityfluctuations observed in the previous financial data.

The maximum of P ∗(v) lies slightly below v at vmax = (ν − 1)/µ = v − ε2/2γ[recall (20.72)]. If we define the w of P ∗(v) by the curvature at the maximum, we

9See also p. 67 in the textbook [8].

Page 58: Economic and Path Integrals

20.4 Origin of Semi-Heavy Tails 1497

find w =√

vmaxε2/2γ. The shape of P ∗(v) is characterized by the dimensionlessratio

vmax

w=

2γv

ε2− 1. (20.328)

The size of the fluctuations of the variance are restricted by the condition (20.322)which guarantees positive v(t) for all times. The distribution (20.327) is plotted inFig. 20.21. As a cross check, we go to the limit ε2/2γv → 0 where the fluctuationsof the variance are frozen out. The ratio vmax/w diverges and the distribution P ∗(v)tends to δ(v − v), as expected.

0 1 2 3 4 5

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

P ∗(v)

v/v

Figure 20.21 Stationary distribution (20.327) of variances for parameters of fits to the

Dow-Jones data in Fig. 20.22 listed in Table 20.1. Compare with Fig. 20.3.

20.4.3 Solution of Fokker-Planck Equation

Since the Hamiltonian operator (20.319) does not depend explicitly on x, we cantake advantage of translational invariance and write P (x v t |xavata) as a Fourierintegral

P (x v t |xavata) =∫

dp

2πeip(x−xa) Pp(v t |vata). (20.329)

The fixed-momentum probability satisfies the Fokker-Planck equation

∂tPp(v t |vata)=

[

γ∂

∂v(v−v) − p2−ip

2v − iρεp

∂vv − ε2

2

∂2

∂v2v

]

Pp(v t |vata). (20.330)

This partial differential equation is of second-order. The variable v occurs onlylinearly. It is therefore convenient to go to Fourier space also in v and write

Pp(v t |vata) =∫

dpv2π

eipvv Pp(pv t |vata), (20.331)

Page 59: Economic and Path Integrals

1498 20 Path Integrals and Financial Markets

which solves the first-order partial differential equation[

∂t+

(

Γpv +iε2

2p2v +

ip2 + p

2

)

∂pv

]

Pp(pv t |vata) = −iγv pv Pp(pv t |vata), (20.332)

where we have used the abbreviation

Γ(p) ≡ γ + iρεp. (20.333)

Since the initial condition for Pp(v t |vata) is δ(v−va), the Fourier transform has theinitial condition

Pp(pv ta|vata) = e−ipvva . (20.334)

The solution of the first-order partial differential equation (20.332) is found by themethod of characteristics [9]:

Pp(pv t|vata) = exp[

−ipv(ta)va − iγv∫ t

tadt′ pv(t

′)]

, (20.335)

where the function pv(t) is the solution of the characteristic differential equation

dpv(t)

dt= Γ(p)pv(t) +

iε2

2p2v(t) +

i

2(p2 − ip), (20.336)

with the boundary condition pv(tb) = pv. The differential equation is of the Riccatitype with constant coefficients [10], and its solution is

pv(t) = −i2Ω(p)

ε21

ζ(p, pv)eΩ(p)(tb−t) − 1+ i

Γ(p) − Ω(p)

ε2, (20.337)

where we introduced the complex frequency

Ω(p) =√

Γ2(p) + ε2(p2 − ip), (20.338)

and the coefficient are

ζ(p, pv) = 1 − i2Ω(p)

ε2pv − i[Γ(p) − Ω(p)]. (20.339)

Taking the Fourier transform we obtain the solution of the original Fokker-Planckequation (20.318):

P (x v t |xavata) =∫ ∫ +∞

−∞

dp

dpv2π

eipx+ipvv

× exp

−ipv(ta)va +γv[Γ(p)−Ω(p)]

ε2∆t− 2γv

ε2lnζ(p, pv)−e−Ω(p)∆t

ζ(p, pv) − 1

, (20.340)

where ∆t ≡ t− ta is the time interval.

Page 60: Economic and Path Integrals

20.4 Origin of Semi-Heavy Tails 1499

20.4.4 Pure x-Distribution

We now show that upon averaging over the variance we obtain a non-Gaussiandistribution of x with semi-heavy tails of the type observed in actual financial data.Thus we go over to the probability distribution

P (x t |xavata) ≡∫ +∞

−∞dv P (x v t |xavata), (20.341)

where the final variable v is integrated out. The integration of (20.340) over vgenerates the delta-function δ(pv) which enforces pv = 0. Thus we obtain

P (x t |xavata) =∫ +∞

−∞

dp

2πeipx−

p2−ipΓ+Ωcoth (Ω∆t/2)

va+γΓv

ε2∆t− 2γv

ε2ln(cosh Ω∆t

2+ Γ

Ωsinh Ω∆t

2 ), (20.342)

where we have omitted the arguments of Γ(p) and Ω(p), for brevity. As a crosscheck of this result we consider the special case ε = 0 where the time evolution ofthe variance is deterministic:

v(t) = v + (va − v)e−γ∆t. (20.343)

Performing the integral over p in Eq. (20.342) we obtain

P (x t |xatava) =1

2π(t− ta)v(t)exp

− [x + v(t)(t− ta)/2]2

2v(t)

, (20.344)

where v(t) denotes the time-averaged variance

v(t) ≡ 1

∆t

∫ t

tadt′ v(t′). (20.345)

The distribution becomes Gaussian in this limit. The same result could, of course,have been obtained directly from the stochastic differential equation (20.306).

The result (20.342) cannot be compared directly with financial time series data,because it depends on the unknown initial variance va. Let us assume that va hasinitially a stationary probability distribution P ∗(va) of Eq. (20.327).10 Then weevaluate the probability distribution P (x t |xata) by averaging (20.342) over va withthe weight P ∗(va):

P (x t |xata) =∫ ∞

0dva P

∗(va)P (x t |xatava). (20.346)

The final result has the Fourier representation

P (x t |xata) =∫ +∞

−∞

dp

2πeip∆x−∆tH(p,∆t), (20.347)

10For a determination of the empirical probability distribution of volatilities for the S&P 500index see Ref. [2].

Page 61: Economic and Path Integrals

1500 20 Path Integrals and Financial Markets

where ∆x ≡ x− xa and H(p,∆t) is the Hamiltonian

H(p,∆t)=−γΓ(p)v

ε2+

2γv

ε2∆tln

[

coshΩ(p)∆t

2+

Ω2(p)−Γ2(p)+2γΓ(p)

2γΩ(p)sinh

Ω(p)∆t

2

]

.

(20.348)

In the absence of correlations between the fluctuations of stock price and variance,i.e. for ρ = 0, the second term in the right-hand side of Eq. (20.348) vanishes. Thesimplified result is discussed in Appendix 20B.

The general Hamiltonian (20.348) vanishes for p = 0. This guarantees the unitnormalization of the distribution (20.347) at all times:

dxP (x t |xata) = 1. Thefirst expansion coefficient of H(p,∆t) in powers of p is [recall the definition inEq. (20.54)]

c1(∆t) = − v2− ρ

ǫ∆t

(

1 − e−γ∆t)

. (20.349)

We have added the argument ∆t to emphasize the time dependence. By addingic1(∆t)p to H(p,∆t), we obtain the Hamiltonian H(p,∆t) which starts out quadrat-ically in p in accordance with our general definition in Eq. (20.65):

H(p,∆t) ≡ H(p,∆t) − ic1(∆t)p = H(p,∆t) + i[

v

2+

ρ

ǫ∆t

(

1 − e−γ∆t)

]

p. (20.350)

The distribution P (x t |xata) is real since ReH is an even function and ImH is anodd function of p.

In general, the integral in (20.347) must be calculated numerically. The inter-esting limit regimes ∆t ≫ 1 and x≫ 1, however, can be understood analytically.These will be treated in Subsections 20.4.5 and 20.4.6. In Fig. 20.22, the calcu-lated distributions P (x t |xata) with increasing time intervals are displayed as solidcurves and compared with the corresponding Dow-Jones data indicated by dots.The technical details of the data analysis are explained in Appendix 20C. The fig-ure demonstrates that with a fixed set of the five parameters γ, v, ε, µ, and ρ, thedistribution (20.347) with (20.348) reproduce extremely well the data for all timescales ∆t. In the logarithmic plot the far tails of the distributions fall off linearly.

20.4.5 Long-Time Behavior

According to Eq. (20.308), the variance approaches the equilibrium value v withinthe characteristic relaxation time 1/γ. Let us study the limit in which the timeinterval ∆t is much longer than the relaxation time: γ∆t ≫ 1. According toEqs. (20.333) and (20.338), this condition also implies that Ω(p)∆t ≫ 1. ThenEq. (20.348) reduces to

H(p,∆t) ≈ γv

ε2[Ω(p) − Γ(p)]. (20.351)

Page 62: Economic and Path Integrals

20.4 Origin of Semi-Heavy Tails 1501

-0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3

100

101

102

103

104

Dow Jones data, 1982-2001

1 day

5 days

20 days

40 days

250 days

-0.4 -0.2 0 0.2

1

10

100

∆t = t−ta

∆x = x− xa

P (x t |xata)

Figure 20.22 Probability distribution of logarithm of stock price for different time scales

(from Ref. [79]). For a better comparison of the shapes, the data points for increasing ∆t

are shifted upwards by a factor 10 each. The unshifted positions are shown in the insert.

The Fourier integral (20.347) can easily be performed after changing the variable ofintegration to

p = C p + ip0, (20.352)

where

C ≡ ω0

ε√

1 − ρ2, ω0 ≡

γ2 + ε2(1 − ρ2)p20, p0 ≡ε− 2ργ

2ε(1 − ρ2). (20.353)

Then the integral (20.347) takes the form

P (x t |xata) =C

πe−p0∆x+Λ∆t

∫ ∞

0dp cos(Ap)e−B

√1+p2, (20.354)

where

A = C(

∆x + ργv

ε∆t)

, B =γvω0

ε2∆t, (20.355)

and

Λ =γv

2ε22γ − ρε

1 − ρ2. (20.356)

The integral in (20.354) is equal to11 BK1(√A2 +B2)/

√A2 +B2, where K1(y) is a

modified Bessel function, such that the probability distribution (20.347) for γ∆t≫ 1can be represented in the scaling form

P (x t |xata) = N(∆t) e−p0∆xF ∗(y), F ∗(y) = K1(y)/y, (20.357)

11See I.S. Gradshteyn and I.M. Ryzhik, op. cit., Formula 3.914.

Page 63: Economic and Path Integrals

1502 20 Path Integrals and Financial Markets

with the argument

y ≡√A2 +B2 =

ω0

ε

(∆x + ργv∆t/ε)2

1 − ρ2+(

γv∆t

ε

)2

, (20.358)

and the time-dependent normalization factor

N(∆t) =ω20γv

πε3√

1 − ρ2∆t eΛ∆t. (20.359)

Thus, up to the factors N(∆t) and e−p0∆x, the dependence of P (x t |xata) on thearguments ∆x and ∆t is given by the function F ∗(y) of the single scaling argumenty. When plotted as a function of y, the data for different ∆x and ∆t should collapseon the single universal curve F ∗(y). This does indeed happen as illustrated inFig. 20.23, where the Dow-Jones data for different time differences ∆t follows thecurve F ∗(y) for seven decades.

-10 -8 -6 -4 -2 0 2 4 6 8 10

10-6

10-4

10-2

100

102

Dow-Jones data, 1982-2001

10 days20 days30 days40 days80 days100 days150 days200 days250 days

y

RenormalizedDistribution

Figure 20.23 Renormalized distribution function P (x t |xata)ep0∆x/N(∆t) plotted as a

function of the scaling argument y defined in Eq. (20.358). The solid curve shows the

universal function F ∗(y) = K1(y)/y (from Ref. [79]).

In the limit y ≫ 1, we can use the asymptotic expression K1(y) ≈ e−y√

π/2yand find the asymptotic behavior

lnP (x t |xata)N(∆t)

≈ −p0∆x− y. (20.360)

Let us examine this expression for large and small |∆x|. In the first case |∆x| ≫γv∆t/ε, and Eq. (20.358) shows that y ≈ ω0|∆x|/ε

√1 − ρ2, so that (20.360) be-

comes

lnP (x t |xata)N(∆t)

≈ −p0∆x− c|∆x|. (20.361)

Page 64: Economic and Path Integrals

20.4 Origin of Semi-Heavy Tails 1503

0 10 20 30 40 50 60 70 800

20

40

60

80

100

120

140

160

1800 1 2 3

∆t

−d logP (x t |xata))/dx

γ∆t

Figure 20.24 Solid curve showing slope −d log P (x t |xata)/dx of exponential tail of

distribution as a function of time. The dots indicate the analytic short-time approximation

(20.369) to the curve (from [79]).

This shows that the probability distribution P (x t |xata) has exponential tails(20.361) for large |∆x|. Note that in the present limit γ∆t ≫ 1 the slopes ofthe exponential tails are independent of ∆t. The presence of p0 causes the slopesfor positive and negative ∆x to be different, so that the distribution P (x t |xata) isnot up-down symmetric with respect to price changes. From the definition of p0 inEq. (20.353) we see that this asymmetry increases for a negative correlation ρ < 0between stock price and variance.

In the second case |∆x| ≪ γv∆t/ε, by Taylor-expanding y in (20.358) near itsminimum and substituting the result into (20.360), we obtain

lnP (x t |xata)N ′(∆t)

≈ −p0∆x−ω0(∆x + ργv∆t/ε)2

2(1 − ρ2)γv∆t, (20.362)

where N ′(∆t) = N(∆t) exp(−ω0γv∆t/ε2). Thus, for small |∆x|, the probabilitydistribution P (x t |xata) is a Gaussian, whose width

σ2 =(1 − ρ2)γv

ω0

∆t (20.363)

grows linearly with ∆t. The maximum of P (x t |xata) lies at

∆xm(t) = ∆rS∆t, with ∆rS ≡ − γv

2ω0

[

1 + 2ρ(ω0 − γ)

ε

]

. (20.364)

The position moves with a constant rate ∆rS which adds to the average growth raterS of S(t) removed at the beginning of the discussion. The true final growth rate ofS(t) is rS = rS + ∆rS.

Page 65: Economic and Path Integrals

1504 20 Path Integrals and Financial Markets

The above discussion explains the property of the data in Fig. 20.22 that thelogarithmic plots of P (x t |xata) are linear in the tails and quadratic near the peakswith the parameters specified in Eqs. (20.361) and (20.362).

As time progresses, the distribution broadens in accordance with the scaling form(20.357) and (20.358). In the limit ∆t → ∞, the asymptotic expression (20.362) isvalid for all ∆x and the distribution becomes a pure Gaussian, as required by thecentral limit theorem [22].

It is interesting to quantify the fraction f(∆t) of the total probability contained inthe Gaussian portion of the curve. This fraction is plotted in Fig. 20.25. The preciseway of defining and calculating the fraction f(∆t) is explained in Appendix 20B.The inset in Fig. 20.25 illustrates that the time dependence of the probability densityat the maximum xm approaches ∆t−1/2 for large time, a characteristic property ofevolution of Gaussian distributions.

0 40 80 120 160 200 2400

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

10 1 2 3 4 5 6 7 8 9 10 11

1 10 1001

10

100

f(∆)

P (xmt |xata)

∆t

∆t

Figure 20.25 Fraction f(∆t) of total probability contained in Gaussian part of

P (x t |xata) as function of time interval ∆t. The inset shows the time dependence of the

probability density at maximum P (xmt |xata) (points), compared with the falloff ∝ ∆t−1/2

of a Gaussian distribution (solid curve).

20.4.6 Tail Behavior for all Times

For large |∆x|, the integrand in (20.347) oscillates rapidly as a function of p, so thatthe integral can be evaluated in the saddle point approximation of Section 4.2. Asin the evaluation of the integral (17.9) we shift the contour of integration in thecomplex p-plane until it passes through the leading saddle point of the exponentip∆x − ∆tH(p,∆t). To determine its position we note that the function H(p,∆t)in Eq. (20.348) has singularities in the complex p-plane, where the argument of the

Page 66: Economic and Path Integrals

20.4 Origin of Semi-Heavy Tails 1505

ps

p1

+

p2

+

p3

+

p1

-

p2

00

-iq-

-

iq+

Im p

Re p

Figure 20.26 Singularities of H(p,∆t) in complex p-plane (dots). Circled crosses in-

dicate the limiting positions ±iq±∗ of the singularities for γ∆t ≫ 1. The cross shows the

saddle point ps located in the upper half-plane for ∆x > 0. The dashed line is the shifted

contour of integration to pass through the saddle point ps.

logarithm vanish. These points are located on the imaginary p-axis and are shownby dots in Fig. 20.26. The relevant singularities are those lying closest to the realaxis. They are located at the points p+1 and p−1 , where the argument of the secondlogarithm in (20.348) vanishes. Near these zeros, we can approximate H(p,∆t) bythe singular term:

H(p,∆t) ≈ 2γv

ε2∆tlog(p− p±1 ). (20.365)

With this approximation, the position of the saddle point ps = ps(∆x) is determinedby the equation

i∆x = ∆tdH(p,∆t)

dp

p=ps

≈ 2γv

ε2×

1

ps − p+1, ∆x > 0,

1

ps − p−1, ∆x < 0.

(20.366)

The solutions are indicated in Fig. 20.26 by the cross. The approximation (20.366)is obviously applicable since for a large |∆x| satisfying the condition |∆xp±1 | ≫γv/ε2, the saddle point ps is very close to one of the singularities. Inserting theapproximation (20.366) into the Fourier integral (20.347), we obtain the asymptoticexpression for the probability distribution

P (x t |xata) ∼

e−∆x q+t , ∆x > 0,

e−|∆x| q−t , ∆x < 0,(20.367)

Page 67: Economic and Path Integrals

1506 20 Path Integrals and Financial Markets

where q±t ≡ ∓ip±1 (t) are real and positive. Thus the tails of the probability dis-tribution P (x t |xata) for large |∆x| are exponential for all times t. The slopes ofthe logarithmic plots in the tails q± ≡ ∓ d log(x t |xata)/dx are determined by thepositions p±1 of the singularities closest to the real axis.

These positions p±1 depend on the time interval ∆t. For times much shorterthan the relaxation time (γt ≪ 1), the singularities lie far away from the real axis.As time increases, the singularities approach the real axis. For times much longerthan the relaxation time (γ∆t ≫ 1), the singularities approach the limiting pointsp±1 → ±iq±∗ , marked in Fig. 20.26 by circled crosses. The limiting values are

q∗± = ±p0 +

ω0

ε√

1 − ρ2for γ∆t ≫ 1. (20.368)

The slopes q±(∆t) approach these limiting slopes monotonously from above. Thebehavior is shown in Fig. 20.24. The slopes (20.368) are of course in agreement withEq. (20.361) in the limit γ∆t ≫ 1. In the opposite limit of short time (γ∆t ≪ 1),we find the analytic time behavior

q±(∆t) ≈ ±p0 +

p20 +4γ

ε2(1 − ρ2)∆tfor γ∆t≪ 1. (20.369)

This approximation is shown in Fig. 20.24 as dots.

20.4.7 Path Integral Calculation

Instead of solving the Fokker-Planck equation (20.330) we may also study directlythe path integral for the probability distribution using the Hamiltonian (20.321).Note the extra two terms at the end in comparison with the operator expression(20.319). They account for the symmetric operator order implied by the path inte-gral. The distribution Pp(vb, tb|vata) at fixed momentum introduced in Eq. (20.329)has the path integral representation

Pp(vb, tb|vata) =∫

Dv Dpv2π

eAp[pv,v], (20.370)

with the action

Ap[pv, v] =∫ tb

tadt [ipvv −H (p, pv, v)] . (20.371)

The path integral (20.370) sums over all paths pv(t) and v(t) with the boundaryconditions v(ta) = va and v(tb) = vb.

It is convenient to integrate the first term in the (20.371) by parts, and toseparate H (p, pv, v) into a v-independent part iγvpv − iγ/2 − iρǫp/2 and a linearterm [∂H (p, pv, v) /∂v]v. Thus we write

Ap[pv, v] = i[pv(tb)vb − pv(ta)va] − iγv∫ tb

tadt pv(t)(tb − ta)

−∫ tb

tadt

[

ipv(t) +∂H

∂v(t)

]

v(t). (20.372)

Page 68: Economic and Path Integrals

20.4 Origin of Semi-Heavy Tails 1507

Since the path v(t) appears linearly in this expression, we can integrate it out toobtain a delta-functional δ [pv(t) − pv(t)], where pv(t) is the solution of the Hamiltonequation pv(t) = i∂H/∂v(t). This, however, coincides exactly with the characteristicdifferential equation (20.336) which was solved by Eq. (20.337) with a boundary con-dition pv(tb) = pv. Taking the path integral over Dpv removes the delta-functional,and we find

Pp(vb, tb|vata) =

+∞∫

−∞

dpv2π

J ei[pvv−pv(ta)va]−iγθ∫ t

0dt pv(t)+(γ−iρǫp)(tb−ta)/2, (20.373)

where J denotes the Jacobian

J = Det−1

(

i∂t +∂2H(p, pv, v)

∂pv∂v

)

. (20.374)

From (20.321) we see that

∂2H(p, pv, v)

∂pv∂v= −iγ + ρǫp. (20.375)

According to Formula (18.254), this is equal to

J = e−(γ−iρǫp)(tb−ta)/2,, (20.376)

thus canceling the last term in the exponent of the integrand in (20.373). The resultfor Pp(vb, tb|vata) is therefore the same as the one obtained from the Fokker-Planckequation in Eq. (20.331) with the Fourier transform (20.335).

20.4.8 Natural Martingale Distribution

Let us calculate the natural martingales associated with the Hamiltonian (20.348).Reinserting the initially removed drift rS, the total Hamiltonian reads

Htot(p,∆t) = H(p,∆t) + irSp. (20.377)

To construct the martingale according to the rule of Subsection 20.1.11, we need toknow the value of H(p,∆t) at the momentum p = i. The expression is somewhatcomplicated, but the analysis of the Dow-Jones data in Appendix 20C shows thatthe parameter ρ determining the strength of correlations between the noise functionsη(t) and ηv(t) [see (20.310)] is small. It is therefore sufficient to find only the firsttwo terms of the Taylor series of H(i,∆t) in powers of ρ:

H(i,∆t) ≈ v

ǫ∆t

(

1−e−γ∆t)

ρ+v

4γ∆t

[

3 − 2e−γ∆t (1+2γ∆t) − e−2γ∆t]

ρ2. (20.378)

From this we obtain

H(i,∆t) = H(i,∆t) + c1(∆t). (20.379)

Page 69: Economic and Path Integrals

1508 20 Path Integrals and Financial Markets

The analog of the martingale (20.291) is given by a Fourier integral (20.292) withHrx(p) replaced by Htot

rx(∆t)(p,∆t) of Eq. (20.350):

P (M,rS)(xbtb|xata) = e−r(∆t)∆t∫ ∞

−∞

dp

2πexp

[

ip(xb − xa) − ∆tHtotrx(∆t)(p,∆t)

]

.(20.380)

The linear term irx(∆t)p in Htotrx(∆t)(p,∆t) is now time-dependent:

rx(∆t) = rS + c1(∆t) + H(i,∆t). (20.381)

This makes the rate in the exponential prefactor, by which the distribution becomesa martingale distribution, a time-dependent quantity:

r(∆t) = rS + c1(∆t). (20.382)

By analogy with (20.293), there exists again an entire family of natural martingales,which is obtained by replacing rS by an arbitrary growth rate r in which case r(∆t)becomes equal to r + c1(∆t).

20.5 Time Series

True market prices are not continuous functions of time but recorded in discrete timeintervals. For stocks which have a low turnover, only the daily prices are recorded.For others which are traded in large amounts, the prices change by the minute.They are listed as time series S(tn). In the year 2003, Robert Engle received theNobel prize for his description of time series with the help of the so called ARCH

models (autoregressive conditional heteroskedasticity model) and its generalizationproposed by Tim Bollerslev [85], the GARCH models. The first contains an inte-ger parameter q and is defined by a discrete modification of the pair of stochasticdifferential equations (20.306) and (20.308) for log-return and variance

x(tn) =√

v(tn−1)η(tn), v(tn) = v0 +q∑

k=1

αk(tn−k)x2(tn−k), (20.383)

with a white noise variable η(tn). The drift is omitted so that 〈x(tn)〉 = 0. Thisis called the ARCH(q) model. The second is a generalization of this in which theequation for the variance contains extra terms involving the past p values of σ2:

v(tn) = v0 +q∑

k=1

αk(tn−k)x2(tn−k) +

p∑

k=1

βk(tn−k)v(tn−k). (20.384)

The ARCH(1) process has the time-independent expectations of variance andkurtosis

σ2 = 〈v(tn)〉 =v0

1 − α1, κ =

〈x4(tn)〉c〈x2(tn)〉2c

=6α2

1

1 − 3α21

. (20.385)

For the GARCH(1,1) process, these quantities are

σ2 = 〈v(tn)〉 =v0

1 − α1 − β1, κ =

〈x4(tn)〉c〈x2(tn)〉2c

=6α2

1

1 − 3α21 − 2α1β1 − β2

1

. (20.386)

Page 70: Economic and Path Integrals

20.6 Spectral Decomposition of Power Behaviors 1509

-0.2 -0.1 0 0.1 0.210-2

10-1

100

101

102

Real dataGARCH process

Heston model

Gaussian

x/σx

P (x)log10 P (x)

Figure 20.27 Left: Comparison of GARCH(1,1) process with v0 = 2.3 × 10−5, α1 =

0.091, β = 0.9 with S&P 500 index (minute data). Right: Comparison of GARCH(1,1),

Heston model, and Gaussian of the same σ with daily market data. Parameters of GARCH

process are v0 = 7.7 × 10−6, α1 = 0.07906, β1 = 0.90501 and of Heston model γ =

4.5 × 10−2, v = 8.62 × 10−8, rS = 5.67 × 10−4, ε = 10.3 × 10−3 [79, 81, 82].

20.6 Spectral Decomposition of Power Behaviors

Plots of the log return curves taken over short time intervals reveal a richer structurethan the simple model functions discussed so far. In fact, it seems reasonable todistinguish different types of stocks according to the characters of the investors. Forinstance, the bubble of the computer stocks was partly caused by quite inexperiencedpeople who were not investing mainly to develop an industry but who wanted to earnfast money. There was, on the other hand also a substantial portion of institutionalinvestors who poured in money more steadily. It appears that one should distinguishthe sources of noise coming from the different groups of investors.

We therefore improve the stochastic differential equation (20.4) by extending itto

x(t) = rx +∑

λ

ηλ(t), (20.387)

where the sum runs over different groups of investors, each producing a different anoise ηλ(t) with Levy distributions falling off with different powers |x|−1−λi . Theirprobability distributions are

Pλ[ηλ] = exp[

−∫ tb

tadt Hλ(ηλ(t))

]

=∫ Dp

2πexp

∫ tb

tadt [ip(t)ηλ(t) −Hλ(p(t))]

,

(20.388)with a Hamiltonian [compare (20.10)]

Hλ(p) ≡ σ2λ |p|λ2

. (20.389)

The probability (20.144) of returns becomes now

P (xbtb|xata)=∏

λ

Dηλ∫

Dx exp[

−∫ tb

tadt Hλ(ηλ(t))

]

δ[x−∑

λ

ηλ]. (20.390)

Page 71: Economic and Path Integrals

1510 20 Path Integrals and Financial Markets

If we insert here a Fourier representation of the δ-functional, we obtain

P (xbtb|xata) =∫ ∞

−∞

Dp2π

λ

[∫

Dηλ]∫

Dx e∫ tbta

dt [ip(t)x(t)−Hλ(ηλ(t))]e−i∑

λ

−∞dt p(t)ηλ(t).

(20.391)Performing the path integrals over ηλ(t), this becomes

P (xbtb|xata) =∫ ∞

−∞

Dp2π

Dx e∫ tbta

dt [ip(t)x(t)−Hλ(p)], (20.392)

where the Hamiltonian is the sum of the group Hamiltonians

H(p) ≡∑

λ

Hλ(p). (20.393)

The continuous generalization of this is

H(p) = H(|p|) ≡∫ ∞

0dλσ2

λ|p|λ. (20.394)

The spectral function σ2λ must be extracted from the data by forming the integral

σ2λ =

∫ i∞

−i∞

d log |p|2πi

p−λH(p). (20.395)

20.7 Option Pricing

Historically, the most important use of path integrals in financial markets was madein the context of determining a fair price of financial derivatives, in particularoptions.12Options are an ancient financial tool. They are used for speculative pur-poses or for hedging major market transactions against unexpected changes in themarket environment. These can sometimes produce dramatic price explosions orerosions, and options are supposed to prevent the destruction of large amounts ofcapital. Ancient Romans, Grecians, and Phoenicians traded options against outgo-ing cargos from their local seaports. In financial markets, options are contractedbetween two parties in which one party has the right but not the obligation to dosomething, usually to buy or sell some underlying asset. Having rights without obli-gations has a value, so option holders must pay a price for acquiring them. Theprice depends on the value of the associated asset, which is why they are also calledderivative assets or briefly derivatives . Call options are contracts giving the optionholder the right to buy something, while put options entitle the holder to sell some-thing. The price of an option is called premium. Usually, options are associatedwith stock, bonds, or commodities like oil, metals or other raw materials. In thesequel we shall consider call options on stocks, to be specific.

Modern option pricing techniques have their roots in early work by CharlesCastelli who published in 1877 a book entitled The Theory of Options in Stocks

12An introduction is found on the site http://bradley.bradley.edu/~arr/bsm/model.html.

Page 72: Economic and Path Integrals

20.7 Option Pricing 1511

and Shares . This book presented an introduction to the hedging and speculationaspects of options. Twenty three years later, Louis Bachelier offered the earliestknown analytical valuation for options in his dissertation at the Sorbonne [86]. Re-markably, Bachelier discovered the treatment of stochastic phenomena five yearsbefore Einstein’s related but much more famous work on Brownian motion [87], andtwenty-three years before Wiener’s mathematical development [88]. The stochasticdifferential equations considered by him still had an important defect of allowingfor negative security prices, and for option prices exceeding the price of the under-lying asset. Bachelier’s work was continued by Paul Samuelson, who wrote in 1955an unpublished paper entitled Brownian Motion in the Stock Market . During thatsame year, Richard Kruizenga, one of Samuelson’s students, cited Bachelier’s workin his dissertation Put and Call Options: A Theoretical and Market Analysis . In1962, a dissertation by A. James Boness entitled A Theory and Measurement of

Stock Option Value developed a more satisfactory pricing model which was furtherimproved by Fischer Black and Myron Scholes. In 1973 they published their famousBlack and Scholes Model [89] which, together with the improvements introduced byRobert Merton, earned them the Nobel prize in 1997.13

As discussed before, the Gaussian distribution severely underestimates the prob-ability of large jumps in asset prices and this was the main reason for the catastrophicfailure in the early fall of 1998 of the hedge fund Long Term Capital Management ,which had Scholes and Merton on the advisory board (and as shareholders). Thefund contained derivatives with a notional value of 1,250 Billion US$. The fundcollected 2% for administrative expenses and 25% of the profits, and was initiallyextremely profitable. It offered its shareholders returns of 42.8% in 1995, 40.8%in 1996, and 17.1% even in the disastrous year of the Asian crisis 1997. But inSeptember 1998, after mistakenly gambling on a convergence in interest rates, italmost went bankrupt. A number of renowned international banks and Wall Streetinstitutions had to bail it out with 3.5 Billion US$ to avoid a chain reaction of creditfailures.

In spite of this failure, the simple model is still being used today for a rough butfast orientation on the fairness of an option price.

20.7.1 Black-Scholes Option Pricing Model

In the early seventies, Fischer Black was working on a valuation model for stockwarrants and observed that his formulas resembled very much the well-known equa-tions for heat transfer. Soon after this, Myron Scholes joined Black and togetherthey discovered an approximate option pricing model which is still of wide use.

The Black and Scholes Model is based on the following assumptions:

1. The returns are normally distributed.

The shortcomings of this assumption have been discussed above in Sec-tion 20.1. The appropriate improvement of the model will be developed below.

13For F. Black the prize came too late — he had died two years earlier.

Page 73: Economic and Path Integrals

1512 20 Path Integrals and Financial Markets

2. Markets are efficient.

This assumption implies that the market operates continuously with shareprices following a continuous stochastic process without memory. It also im-plies that different markets have the same asset prices.

This is not quite true. Different markets do in general have slightly differentprices. Their differences are kept small by the existence of arbitrage dealers.There also exist correlations over a short time scale which make it possible, inprinciple, to profit without risk from the so-called statistical arbitrage. Thispossibility is, however, strongly limited by transaction fees.

3. No commissions are charged.

This assumption is not satisfied. Usually, market participants have to pay acommission to buy or sell assets. Even floor traders pay some kind of fee,although this is usually very small. The fees payed by individual investors ismore substantial and can distort the output of the model.

4. Interest rates remain constant and known.

The Black and Scholes model assumes the existence of a riskfree interest rate

to represent this constant and known rate. In reality there is no such thingas the riskfree rate. As an approximation, one uses the discount rate on U.S.Government Treasury Bills with 30 days left until maturity. During periods ofrapidly changing interest rates, these 30 day rates are often subject to change,thereby violating one of the assumptions of the model.

5. The stock pays no dividends during the option’s life.

Most companies pay dividends to their share holders, so this is a limitationto the model since higher dividends lead to lower call premiums. There is,however, a simple possibility of adjusting the model to the real situation bysubtracting the discounted value of a future dividend from the stock price.

6. European exercise terms are used.

European exercise terms imply the exercise of an option only on the expirationdate. This is in contrast to the American exercise terms which allow for thisat any time during the life of the option. This greater flexibility makes anAmerican option more valuable than the European one.

The difference is, however, not dramatic in praxis because very few calls areever exercised before the last few days of their life, since an early exercisemeans giving away the remaining time value on the call. Different exercisetimes towards the end of the life of a call are irrelevant since the remainingtime value is very small and the intrinsic value has a small time dependence,barring a dramatic event right before expiration date.

Page 74: Economic and Path Integrals

20.7 Option Pricing 1513

Since 1973, the original Black and Scholes Option Pricing Model has been im-proved and extended considerably. In the same year, Robert Merton [90] includedthe effect of dividends. Three years later, Jonathan Ingersoll relaxed the assumptionof no taxes or transaction costs, and Merton removed the restriction of constant in-terest rates. In recent years, the model has been generalized to determine the pricesof options with many different properties.

The relevance of path integrals to this field was recognized in 1988 by the theoret-ical physicist J.W. Dash, who wrote two unpublished papers on the subject entitledPath Integrals and Options I and II [91]. Since then many theoretical physicistshave entered the field, and papers on this subject have begun appearing on the LosAlamos server [13, 92, 93].

20.7.2 Evolution Equations of Portfolios with Options

The option price O(t) has larger fluctuations than the associated stock price. It usu-ally varies with a slope ∂O(S(t), t)/∂S(t) which is commonly denoted by ∆(S(t), t)and called the Delta of the option. If ∆(S(t), t) depends only weakly on S(t) andt it is possible, in the ideal case of Gaussian price fluctuations, to guarantee asteady growth of a portfolio. One merely has to mix a suitable number NS(t)stocks with NO(t) options and short-term bonds whose number is denoted by NB(t).As mentioned before, these are typically U.S. Government Treasury Bills with 30days left to maturity which have only small price fluctuations. The composition[NS(t), NO(t), NB(t)] is referred to as the strategy of the portfolio manager. Thetotal wealth has the value

W (t) = NS(t)S(t) +NO(t)O(S, t) +NB(t)B(t). (20.396)

The goal is to make W (t) grow with a smooth exponential curve without fluctuations

W (t) ≈ rWW (t). (20.397)

As an idealization, the short-term bonds are assumed to grow deterministicallywithout any fluctuations:

B(t) ≈ rBB(t). (20.398)

The rate rB is the earlier-introduced riskfree interest rate encountered in true mar-kets only if there are no events changing excessively the value of short-term bonds.

The existence of arbitrage dealers will ensure that the growth rate rW is equalto that of the short-term bonds

rW ≈ rB. (20.399)

Otherwise the dealers would change from one investment to the other.In the decomposition (20.396), the desired growth (20.397) reads

NS(t)S(t) +NO(t)O(S, t) +NB(t)B(t) + NS(t)S(t) + NO(t)O(S, t) + NB(t)B(t)

= rW [NS(t)S(t) +NO(t)O(S, t) +NB(t)B(t)] . (20.400)

Page 75: Economic and Path Integrals

1514 20 Path Integrals and Financial Markets

Due to (20.398) and (20.399), the terms containing NB(t) without a dot drop out.Moreover, if no extra money is inserted into or taken from the system, i.e., if stocks,options, and bonds are only traded against each other, this does not change the totalwealth, assuming the absence of commissions. This so-called self-financing strategy

is expressed in the equation

NS(t)S(t) + NO(t)O(S, t) + NB(t)B(t) = 0. (20.401)

Thus the growth equation (20.397) translates into

W (t) = NSS +NOO +NBB = rW (NSS +NOO +NBB) . (20.402)

Due to the equality of the rates rW = rB and Eq. (20.398), the entire contributionof B(t) cancels, and we obtain

NSS +NOO = rW (NSS +NOO) . (20.403)

The important observation is now that there exists an optimal ratio between thenumber of stocks NS and the number of options NO, which is equal to the negativeslope ∆(S(t), t):

NS(t)

NO(t)= −∆(S(t), t) = −∂O(S(t), t)

∂S(t). (20.404)

Then Eq. (20.403) becomes

NSS +NOO = NOrW

(

−∂O∂x

+O

)

. (20.405)

The two terms on the left-hand side are treated as follows: First we use the relation(20.404) to rewrite

NSS = −NO∂O(S, t)

∂SS = −NO

∂O(S, t)

∂x

S

S. (20.406)

In the second term on the left-hand side of (20.405), we expand the total timedependence of the option price in a Taylor series

dO

dt=

1

dt

[

O(x(t) + x(t) dt, t+ dt) − O(x(t), t)]

=∂O

∂t+∂O

∂xx +

1

2

∂2O

∂x2x2 dt+

1

3!

∂3O

∂x3x3 dt2 + . . . . (20.407)

We have gone over to the logarithmic stock price variable x(t) rather than S(t)itself. In financial mathematics, the lowest derivatives on the right-hand side areall denoted by special Greek symbols. We have already introduced the name Deltafor the slope ∂O/∂S. The curvature Γ ≡ ∂2O/∂S2 = (∂2O/∂x2 − ∂O/∂x)/S2 iscalled the Gamma of an option. Another derivative with a standard name is the

Page 76: Economic and Path Integrals

20.7 Option Pricing 1515

Vega V ≡ ∂O/∂σ. The partial time derivative ∂O/∂t is denoted by Θ. The set ofthese quantities is collectively called the Greeks [94].

In general, the expansion (20.407) is carried to arbitrary powers of x as in(20.273). It is, of course, only an abbreviated notation for the proper expansionin powers of a stochastic variable to be performed as in Eq. (20.273). After insert-ing (20.407) and (20.406) on the left-hand side of Eq. (20.405), this becomes

NSS +NOO=− NO∂O

∂x

S

S

+ NO

(

∂O

∂t+∂O

∂xx+

1

2

∂2O

∂x2x2dt+

1

3!

∂O

∂xx3dt+ . . .

)

(20.408)

=NO

[

∂O

∂t+

(

x− S

S

)

∂O

∂x+

1

2

∂2O

∂x2x2dt+

1

3!

∂3O

∂x3x3dt+ . . .

]

.

Replacing further the left-hand side by the right-hand side of (20.405) we obtain

∂O

∂t= −rWO −

(

x− S

S+ rW

)

∂O

∂x− 1

2

∂2O

∂x2x2dt− 1

3!

∂3O

∂x3x3dt+ . . . = 0. (20.409)

The crucial observation which earned Black and Scholes the Nobel prize is thatfor Gaussian fluctuations with a Hamiltonian H(p) = σ2p2/2, the equation (20.409)becomes very simple. First, due to the Ito relation (20.4), the prefactor of −∂O/∂xbecomes a constant

rW − σ2

2≡ rxW

, (20.410)

where the notation rxWis chosen by analogy with rx in the Ito relation (20.5). Thus

there are no more fluctuations in the prefactor of ∂O/∂x.Moreover, also the fluctuations of all remaining terms

−1

2

∂2O

∂x2x2 dt− 1

6

∂3O

∂x3x3 dt2+. . . (20.411)

can be neglected due to the result (18.429) and the estimates (18.431), which maketruncate the expansion (20.411) and make it equal to −(σ2/2) ∂2O/∂x2. ThusEq. (20.409) loses its stochastic character, and the fair option price O(x, t) is foundto obey the Fokker-Planck-like differential equation

∂O

∂t= rWO − rxW

∂O

∂x− σ2

2

∂2O

∂x2, (20.412)

At the same time, the total wealth (20.402) loses its stochastic character andgrows with the riskfree rate rW following the deterministic Eq. (20.398). The can-cellation of the fluctuations is a consequence of choosing the ratio between optionsand stocks according to Eq. (20.404). The portfolio is now hedged against fluctua-tion.

Page 77: Economic and Path Integrals

1516 20 Path Integrals and Financial Markets

The Fokker-Planck equation (20.412) can be expressed completely in terms ofthe Greeks of the option, using the relation (20.410):

Θ = rWO−rxWS∆−σ2

2

(

S∂O

∂S+S2∂

2O

∂S2

)

= rWO−rWS∆−σ2

2S2Γ. (20.413)

The strategy to make a portfolio riskfree by balancing the fluctuations of NS

stocks by NO options to satisfy Eq. (20.404) is called Delta-hedging. Hedging is ofcourse imperfect. Since ∆(S(t), t) depends on the stock price, a Delta hedge requiresfrequent re-balancing of the portfolio, even several times per day. After a time spanδt, the Delta has changed by

δ∆ =d

dt

∂2O(S(t), t)

∂S(t)δt =

[

∂t

∂O(S(t), t)

∂S(t)+∂2O(S(t), t)

∂S(t)2

]

δt =

[

∂Θ

∂S+ Γ

]

δt.

(20.414)So one has to readjust the ratio of stocks and options in Eq. (20.404) by

δNS

NO= −δ∆ = −

(

∂Θ

∂S+ Γ

)

δt. (20.415)

This procedure is called dynamical ∆-hedging . Since buying and selling costs money,δt cannot be made too small, otherwise dynamical hedging becomes too expensive.

In principle it is possible to buy a third asset to hedge also the curvature Gamma.This procedure is not so popular, again because of the transaction costs.

For non-Gaussian noise, the differential equation (20.409) is still stochastic dueto remaining fluctuations in the expansion terms (20.411). This is an obstacle tobuilding a riskfree portfolio with deterministically growing total wealth W (t). As inEq. (20.255) we can only derive the equation of motion for the average value

1

2

∂2O

∂x2x2 dt+

1

6

∂3O

∂x3x3 dt2+. . .

= −H(i∂x)O. (20.416)

Hence a fair option price can only be calculated on the average from the Fokker-Planck-like differential equation

∂t〈O〉 =

[

rW − rxW

∂x+ H(i∂x)

]

〈O〉 , (20.417)

where we have defined, by analogy with (20.268) and (20.410), an auxiliary rateparameter

rxW≡ rW + H(i). (20.418)

However, as pointed out in Subsection 20.2.2, this limitation is not really strin-gent for discrete short-time data if the tails of the noise distribution are non-Gaussian

Page 78: Economic and Path Integrals

20.7 Option Pricing 1517

with semi-heavy tails. Then the higher expansion terms in Eq. (20.411) are sup-pressed by a small factor σ

√∆t, and the expectation values in Eqs. (20.416) and

(20.417) can again be dropped approximately.For Gaussian fluctuations, the Fokker-Planck equation (20.412) can easily be

solved. If we rename t→ ta, and x→ xa, for symmetry reasons, the solution whichstarts out, at some time tb, like δ(xa − xb), has the Fourier representation

P (M,rx)(xbtb|xata)=e−rW (tb−ta)∫ ∞

−∞

dp

2πeip(xb−xa)e−(σ2p2/2+irxW p)(tb−ta). (20.419)

A convergent integral exists for tb > ta.For non-Gaussian fluctuations with semi-heavy tails, there exists an approximate

solution whose Fourier representation is

P (M,rx)(xbtb|xata)=e−rW (tb−ta)∫ ∞

−∞

dp

2πeip(xb−xa)e−[H(p)+irxW p](tb−ta), (20.420)

with H(p) of Eq. (20.141).Recalling the discussion in Section 20.3, this distribution function is recognized

as a member of the equivalent family of martingale distributions (20.293) for thestock price S(t) = ex(t). It is the particular distribution in which the discount factorr coincides with the riskfree interest rate rW .

20.7.3 Option Pricing for Gaussian Fluctuations

For Gaussian fluctuations where H(p) = σ2p2/2, the integral in (20.419) can easilybe performed and yields for tb > ta the martingale distribution

P (M,rx)(xbtb|xata) =e−rW (tb−ta)

2πσ2(tb − ta)exp

− [xb − xa− rxW(tb − ta)]

2

2σ2(tb − ta)

. (20.421)

This is the risk-neutral martingale distribution of Eq. (20.289) for r = rx.This distribution is obviously the solution of the path integral

P (M,rx)(xbtb|xata) = e−rW (tb−ta)∫

Dx exp

− 1

2σ2

∫ tb

ta[x− rxW

]2

. (20.422)

An option is written for a certain strike price E of the stock. The value of theoption at its expiration date tb is given by the difference between the stock price onexpiration date and the strike price:

O(xb, tb) = Θ(xb − xE)(exb − exE), (20.423)

where

xE ≡ logE. (20.424)

Page 79: Economic and Path Integrals

1518 20 Path Integrals and Financial Markets

The Heaviside function in (20.423) accounts for the fact that only for Sb > E it isworthwhile to execute the option.

From (20.423) we calculate the option price at an arbitrary earlier time usingthe time evolution probability (20.421)

O(xa, ta) =∫ ∞

−∞dxbO(xb, tb)P

(M,rW )(xb tb|xata). (20.425)

Inserting (20.423) we obtain the sum of two terms

O(xa, ta) = OS(xa, ta) −OE(xa, ta), (20.426)

where

OS(xa, ta) =e−rW (tb−ta)

2πσ2(tb−ta)

∫ ∞

xE

dxb exb exp

− [xb − xa − rxW(tb − ta)]

2

2σ2(tb − ta)

, (20.427)

and

OE(xa, ta) = Ee−rW (tb−ta)1

2πσ2(tb−ta)

∫ ∞

xE

dxb exp

− [xb − xa − rxW(tb − ta)]

2

2σ2(tb − ta)

.

(20.428)

In the second integral we set

x− ≡ xa + rxW(tb − ta) = xa +

(

rW − 1

2σ2)

(tb − ta), (20.429)

and obtain

OE(xa, ta) = Ee−rW (tb−ta)

2πσ2(tb − ta)

∫ ∞

xE−x−

dxb exp

− x2b2σ2(tb − ta)

. (20.430)

After rescaling the integration variable xb → −ξσ√tb − ta, this can be rewritten as

OE(xa, ta) = e−rW (tb−ta)E N(y−), (20.431)

where N(y) is the cumulative Gaussian distribution function

N(y) ≡∫ y

−∞

dξ√2πe−ξ2/2 =

1

2

[

1 + erf

(

y√2

)]

, (20.432)

evaluated at

y− ≡ x− − xE√

σ2(tb − ta)=

log[S(ta)/E] + rxW(tb − ta)

σ2(tb − ta)

=log[S(ta)/E] +

(

rW − 12σ2)

(tb − ta)√

σ2(ta − tb). (20.433)

Page 80: Economic and Path Integrals

20.7 Option Pricing 1519

The integral in the first contribution (20.427) to the option price is found aftercompleting the exponent in the integrand quadratically as follows:

xb −[xb − xa − rxW

(tb − ta)]2

2σ2(tb − ta)

= − [xb − xa − (rxW+ σ2)(tb − ta)]

2 − 2rWσ2(tb − ta) − 2xaσ

2(tb − ta)

2σ2(tb − ta).(20.434)

Introducing now

x+ ≡ xa +(

rxW+ σ2

)

(tb − ta) = xa +(

rW +1

2σ2)

(tb − ta), (20.435)

and rescaling xb as before, we obtain

OS(xa, ta) = S(ta)N(y+), (20.436)

with

y+ ≡ x+ − xE√

σ2(ta − tb)=

log[S(ta)/E] + (rxW+ σ2) (tb − ta)

σ2(ta − tb)

=log[S(ta)/E] +

(

rW + 12σ2)

(tb − ta)√

σ2(ta − tb). (20.437)

The combined result

O(xa, ta) = S(ta)N(y+) − e−rW (tb−ta)E N(y−) (20.438)

is the celebrated Black-Scholes formula of option pricing.In Fig. 20.28 we illustrate how the dependence of the call price on the stock price

varies with different times to expiration tb − ta and with different volatilities σ.Floor dealers of stock markets use the Black-Scholes formula to judge how ex-

pensive options are, so they can decide whether to buy or to sell them. For a givenriskfree interest rate rW and time to expiration tb − ta, and a set of option, stock,and strike prices O, S, E, they calculate the volatility from (20.438). The result iscalled the implied volatility , and denoted by Σ(x−xE). As typical plot as a functionof x− xE is shown in Fig. 20.29.

If the Black-Scholes formula were exactly valid, the data should lie on a horizontalline Σ(x − xE) = σ. Instead, they scatter around a parabola which is called thesmile of the option. The smile indicates the presence of a nonzero kurtosis in thedistribution of the returns, as we shall see in Subsection 20.7.8.

It goes without saying that if the integral (20.427) is carried out over the entirexb axis, it becomes independent of time, due to the martingale character of therisk-neutral distribution P (M,rW )(xbtb|xata).

Page 81: Economic and Path Integrals

1520 20 Path Integrals and Financial Markets

O(S)

S20 40 60 80

5

10

15

20

25

30

35

40

O(E)

E20 40 60 80 100

5

10

15

20

25

30

35

O(S)

S20 40 60 80

5

10

15

20

25

30

35

40

Figure 20.28 Left: Dependence of call price O on stock price S for different times

before expiration date (increasing dash length: 1, 2, 3, 4, 5 months). The parameters are

E = 50 US$, σ = 40%, rW = 6% per month. Right: Dependence on the strike price E for

fixed stock price 35 US$ and the same times to expiration (increasing with dash length).

Bottom: Dependence on the volatilities (from left to right: 80%, 60%, 20%, 10%, 1%) at a

fixed time tb − ta = 3 months before expiration.

Σ(xb − xa)

xb − xa

Figure 20.29 Smile deduced from options (see [13]).

20.7.4 Option Pricing for Asian Option

In an Asian option one does not bet on an option price at the expiration date buton an average option price over the option life. The calculation is very simple if werecall the time amplitude amplitude (3.200) of a particle at a fixed average x0. Thestochastic version of the quantum-mechanical expression (3.200) is

P (xbtb|xata)x0 =

√3

πσ2(tb−ta)exp

− 1

2σ2(tb−ta)

[

(xb−xa)2+12(

x0−xb+xa

2

)2]

.

(20.439)

Page 82: Economic and Path Integrals

20.7 Option Pricing 1521

For the growth rate rW , the associated martingale becomes

P (M,rx)(xbtb|xata)x0 =e−rW (tb−ta)

√3

πσ2(tb−ta)

× exp

− 1

2σ2(tb−ta)

[

(xb−xa − rxW(tb − ta))

2+ 12(

x0−xb+xa

2

)2]

, (20.440)

whose integral over x0 yields back the previous martingale distribution (20.421)∫ ∞

−∞dx0 P

(M,rx)(xbtb|xata)x0 = P (M,rx)(xbtb|xata). (20.441)

Instead of (20.423), the option price at expiration is now

O(xb, tb) = Θ(xb − x0)(exb − ex0), (20.442)

so that stead of (20.426) is replaced by the difference between

OS(xa, ta) =∫ ∞

−∞dx0

∫ ∞

x0

dxb exbP (M,rx)(xbtb|xata)x0 , (20.443)

which is the same as (20.427), due to (20.441), and

Ox(xa, ta) =∫ ∞

−∞dx0 e

x0

∫ ∞

x0

dxb P(M,rx)(xbtb|xata)x0 , (20.444)

which replaces (20.430). Since both expressions involve integrals over errorfunctionsit is useful to represent the δ-function in (20.442) as a Fourier integral and rewritethe option price at ta as

O(xa, ta) =∫ ∞

−∞

dt

2πi

ei(xb−x0)t

t− iη

×∫ ∞

−∞dx0

∫ ∞

−∞dxb (exb − ex0)P (M,rx)(xbtb|xata)x0, (20.445)

The result of the double integral in the second line is simply

e−A(t) − e−A0(t), (20.446)

with

A(t) =−rW (tb − ta) −i

2

(

rW +σ2

2

)

t(tb − ta) +σ2t2(tb − ta)

6,

A0(t) =

(

−rW +σ2

6

)

tb − ta2

− i

2

(

rW−σ2

6

)

t(tb − ta) +σ2t2(tb − ta)

6. (20.447)

Now we use the integral formula

∫ ∞

−∞

dt

2πi

e−a2t2/2+ibt

t− iη= N [b/a]. (20.448)

Page 83: Economic and Path Integrals

1522 20 Path Integrals and Financial Markets

and find [95]

O(xa, ta) = S(ta)[N(z+) − e−(rW+σ2/6)(tb−ta)/2N(z−)], (20.449)

where

z+ =

3(tb − ta)

4σ2

(

rW +σ2

2

)

, z− =

3(tb − ta)

4σ2

(

rW − σ2

6

)

. (20.450)

20.7.5 Option Pricing for Boltzmann Distribution

The above result can easily be extended to price the options for assets whose returnsobey the Boltzmann distribution (20.204). Since this is a superposition of Gaussiandistributions, we merely have to perform the same superposition over the Black-Scholes formula (20.438). Thus we insert (20.438) into the integral (20.204) andobtain the option price for the Boltzmann-distributed assets of variance σ2:

OB(xa, ta)=(

t

σ2

)t 1

Γ(t)

∫ ∞

0

dv

vvte−tv/σ2

Ov(xa, ta), (20.451)

where Ov(xa, ta) is the Black-Scholes option price (20.438) of variance v. The super-script indicates that the variance σ2 in the variables of y+ and y− in (20.433) and(20.437) is now exchanged by the integration variable v.

Since the Boltzmann distribution for the minute data turns rapidly into a Gaus-sian (recall Fig. 20.15), the price changes with respect to the Black-Scholes formulaare relevant only for short-term options running less than a week. The changes canmost easily be estimated by using the expansion (20.209) to write

OB(xa, ta)=Ov(xa, ta) +[2T 2(1 − 1/t)]2

2t∂2vO

v(xa, ta) + . . . . (20.452)

20.7.6 Option Pricing for General Non-Gaussian Fluctuations

For general non-Gaussian fluctuations with semi-heavy tails, the option price mustbe calculated numerically from Eqs. (20.425) and (20.423). Inserting the Fourierrepresentation (20.419) and using the Hamiltonian

HrxW(p) ≡ H(p) + irxW

p (20.453)

defined as in (20.141), this becomes

O(xa, ta) =∫ ∞

xE

dxb (exb − exE)P (xbtb|xata)

= e−rW (tb−ta)∫ ∞

xE

dxb (exb − exE)∫ ∞

−∞

dp

2πeip(xb−xa)−HrxW

(p)(tb−ta). (20.454)

Page 84: Economic and Path Integrals

20.7 Option Pricing 1523

The integrand can be rearranged as follows:

O(xa, ta)=e−rW (tb−ta)∫ ∞

xE

dxb

∫ ∞

−∞

dp

[

exaei(p−i)(xb−xa)− exEeip(xb−xa)]

e−HrxW(p)(tb−ta).

(20.455)Two integrations are required. This would make a numerical calculation quite timeconsuming. Fortunately, one integration can be done analytically. For this purposewe change the momentum variable in the first part of the integral from p to p + iand rewrite the integral in the form

O(xa, ta) = e−rW (tb−ta)∫ ∞

xE

dxb

∫ ∞

−∞

dp

2πeip(xb−xa)f(p), (20.456)

with

f(p) ≡ exae−HrxW(p+i)(tb−ta) − exEe−HrxW

(p)(tb−ta). (20.457)

We have suppressed the arguments xa, xE , tb − ta in f(p), for brevity. The integralover xb in (20.456) runs over the Fourier transform

f(xb − xa) =∫ ∞

−∞

dp

2πeip(xb−xa)f(p) (20.458)

of the function f(p). It is then convenient to express the integral∫∞xEdxb in terms

of the Heaviside function Θ(xb − xE) as∫∞−∞ dxb Θ(xb − xE) and use the Fourier

representation (1.312) of the Heaviside function to write

∫ ∞

xE

dxb f(xb − xa) =∫ ∞

−∞dxb

∫ ∞

−∞

dq

i

q + iηe−iq(xb−xE)f(xb − xa). (20.459)

Inserting here the Fourier representation (20.458), we can perform the integral overxb and obtain the momentum space representation of the option price

O(xa, ta) = e−rW (tb−ta)∫ ∞

−∞

dp

2πeip(xE−xa)

i

p+ iηf(p). (20.460)

For numerical integrations, the singularity at p = 0 is inconvenient. We thereforeemploy the decomposition (18.54)

1

p+ iη=

Pp− iπδ(p), (20.461)

to write

O(xa, ta) = e−rW (tb−ta)

[

1

2f(0) + i

∫ ∞

−∞

dp

eip(xE−xa)f(p) − f(0)

p

]

. (20.462)

We have used the fact that the principal value of the integral over 1/p vanishes tosubtract the constant f(0) from eip(xE−xa)f(p). After this, the integrand is regu-lar and does not need any more the principal-value specification, and allows for anumerical integration.

Page 85: Economic and Path Integrals

1524 20 Path Integrals and Financial Markets

O(S, t) −OBS(S, t)

S20 40 60

-0.4

-0.2

0.2

0.4

Figure 20.30 Difference between call price O(S, t) obtained from truncated Levy distri-

bution with kurtosis κ = 4 and Black-Scholes price OBS(S, t) with σ2 = v as function of

stock price S for different times before expiration date (increasing dash length: 1, 2, 3, 4, 5

months). The parameters are E = 50 US$, σ = 40%, rW = 6% per month.

For xa very much different from xE , we may approximate

∫ ∞

−∞

dp

eip(xE−xa)f(p) − f(0)

p≈ 1

2ǫ(xa − xE)f(0), (20.463)

where ǫ(x) ≡ 2Θ(x) − 1 is the step function (1.315), and obtain

O(xa, ta) ≈e−rW (tb−ta)

2[1 + Θ(xa − xE)] f(0). (20.464)

Using (20.418) we have e−HrxW(i) = erW , and since e−HrxW

(0) = 1 we see thatO(xa, ta) goes to zero for xa → −∞ and has the large-xa behavior

O(xa, ta) ≈ exa − exEe−rW (tb−ta) = S(ta) − e−rW (tb−ta)E. (20.465)

This is the same behavior as in the Black-Scholes formula (20.438).In Fig. 20.30 we display the difference between the option prices emerging from

our formula (20.462) with a truncated Levy distribution of kurtosis κ = 4, and theBlack-Scholes formula (20.438) for the same data as in the upper left of Fig. 20.28.

For truncated Levy distributions, the Fourier integral in Eq. (20.454) can beexpressed directly in terms of the original distribution function which is the Fouriertransform of (20.52):

L(λ,α,β)σ2 (x) =

∫ ∞

−∞

dp

2πeipx−H(p). (20.466)

By inspecting Eq. (20.53) we see that the factor tb − ta multiplying HrxW(p) in

(20.454) can be absorbed into the parameters σ, λ, α, β of the truncated Levy dis-tributions by replacing

σ2 → σ2(tb − ta), rxW→ rxW

(tb − ta). (20.467)

Let us denote the truncated Levy distribution with zero average by L(λ,α,β))σ2 (x). It

is the Fourier transform of e−H(p):

L(λ,α,β)σ2 (x) =

∫ ∞

−∞

dp

2πeipx−H(p). (20.468)

Page 86: Economic and Path Integrals

20.7 Option Pricing 1525

The Fourier transform of e−H(p)(tb−ta) is then simply given by L(λ,α,β))σ2(tb−ta)

(x). The

additional term rxWin the exponent of the integral (20.454) via (20.453) leads to a

drift rxWin the distribution, and we obtain∫ ∞

−∞

dp

2πeipx−[H(p)+irxW p](tb−ta) = L

(λ,α,β))σ2(tb−ta)

(x− rxW(tb − ta)). (20.469)

Inserting this into (20.419), we find the riskfree martingale distribution to be insertedinto (20.454):

P (xbtb|xata) = e−rW (tb−ta)L(λ,α,β)σ2(tb−ta)

(xb − xa − rxW(tb − ta)). (20.470)

The result is therefore a truncated Levy distribution of increasing width and uni-formly moving average position. Since all expansion coefficients cn of H(p) inEq. (20.40) receive the same factor tb− ta, the kurtosis κ = c4/c

22 decreases inversely

proportional to tb − ta. As time proceeds, the distribution becomes increasinglyGaussian, this being a manifestation of the central limit theorem of statistical me-chanics. This is in contrast to the pure Levy distribution which has no finite widthand therefore maintains its power falloff at large distances.

Explicitly, the formula (20.454) for the option price becomes

O(xa, ta) = e−rW (tb−ta)∫ ∞

xE

dxb (exb − exE)L(λ,α,β)σ2(tb−ta)

(xb − xa − rxW(tb − ta)).(20.471)

This and similar equations derived from any of the other non-Gaussian models leadto fairer formulas for option prices.

20.7.7 Option Pricing for Fluctuating Variance

If the fluctuations of the variance are taken into account, the dependence of the priceof an option on v(t) needs to be considered in the derivation of a time evolutionequation for the option price. Instead of Eq. (20.407) we write the time evolutionas

dO

dt=

1

dt

[

O(x(t) + x(t) dt, v(t) + v(t) dt, t+ dt) − O(x(t), v(t), t)]

=∂O

∂t+∂O

∂xx+

∂O

∂vv +

1

2

∂2O

∂x2x2 dt+

∂2O

∂x∂vxv dt+

1

2

∂2O

∂v2v2dt+ . . . .(20.472)

The expansion can be truncated after the second derivative due to the Gaussiannature of the fluctuations. We use Ito’s rule to replace

x2 −−−→ v(t), v2 −−−→ ǫ2v(t), xv −−−→ ρǫv(t). (20.473)

These replacements follow directly from Eqs. (20.306) and (20.308) and the correla-tion functions (20.310). Thus we obtain

dO

dt=

1

dt

[

O(x(t) + x(t) dt, v(t) + v(t) dt, t+ dt) − O(x(t), v(t), t)]

=∂O

∂t+∂O

∂xx +

∂O

∂vv +

1

2

∂2O

∂x2v +

∂2O

∂x∂vρǫv +

1

2

∂2O

∂v2ǫ2v. (20.474)

Page 87: Economic and Path Integrals

1526 20 Path Integrals and Financial Markets

This is inserted into Eq. (20.403). If we adjust the portfolio according to the rule(20.404), and use the Ito relation S/S = x + v/2, we obtain the equation [compare(20.409)]

NOrW

(

−∂O∂x

+O

)

= − NO∂O

∂x

(

x +v2

2

)

+NOO

= NO

[

−v2

2

∂O

∂x+∂O

∂vv +

∂O

∂t+

1

2

∂2O

∂x2v +

∂2O

∂x∂vρǫv +

1

2

∂2O

∂v2ǫ2v + . . .

]

. (20.475)

As before in Eq. (20.409), the noise in x has disappeared. In contrast to the single-variable treatment, however, the noise variable ηv remains in the equation. It canonly be removed if we trade a financial asset V whose price is equal to the variancedirectly on the markets. Then we can build a riskfree portfolio containing four assets

W (t) = NS(t)S(t) +NO(t)O(S, t) +NV (t)V (t) +NB(t)B(t), (20.476)

instead of (20.396). Indeed, by adjusting

NV (t)

NO(t)= −∂O(S(t), v(t), t)

∂v(t), (20.477)

we could cancel the term v in Eq. (20.475). There is definitely need to establishtrading in such an asset. Without this, we can only reach an approximate freedomof risk by ignoring the noise ηv(t) in the term v and replacing v by the deterministicfirst term in the stochastic differential equation (20.308):

v(t) −−−→ −γ[v(t) − v]. (20.478)

In addition, we can account for the fact that the option price rises with the varianceas in the Black-Scholes formula by adding on the right hand side of (20.478) aphenomenological correction term −λv called price of volatility risk [78, 11]. Sucha term has simply the effect of renormalizing the parameters γ and v to

γ∗ = γ + λ, and v∗ = γv/γ∗. (20.479)

Thus we find the Fokker-Planck-like differential equation [compare (20.417)]

∂O

∂t= rWO −

(

rW − v

2

)

∂O

∂x+ γ∗[v(t) − v∗]

∂O

∂v− v

2

∂2O

∂x2− ρǫv

∂2O

∂x∂v− ǫ2v

2

∂2O

∂v2.

(20.480)On the right-hand side we recognize the Hamiltonian operator (20.319), with γ andv replaced by γ∗ and v∗, in terms of which we can write

∂O

∂t= rW (O − ∂xO) +

(

H∗ + γ∗ + ρǫ∂x + ǫ2∂v)

O. (20.481)

The solution of this equation can easily be expressed as a slight modification of

Page 88: Economic and Path Integrals

20.7 Option Pricing 1527

the solution P (xb, vb, tb|xavata) in Eq. (20.340) of the differential equation (20.318).Since it contains only additional first-order derivatives with respect to (20.318), wesimply find, with x ≡ xa and t ≡ ta, the solution P v(xb, vb, tb|xavata) satisfying theinitial condition

P v(xb, vb, tb|xavata) = δ(xb − xa)δ(vb − va) (20.482)

as follows:

P v(xb, vb, tb|xavata) = e−(rW+γ∗)∆tPsh(xb, vb, tb|xavata), (20.483)

where the subscript sh indicates that the arguments xb − xa and vb − va are shifted:

xb − xa → xb − xa − (rW − ρǫ), vb − va → vb − va − ǫ2. (20.484)

The distribution (20.482) may be inserted into an equation of the type (20.425)to find the option price at the time ta from the price (20.423) at the expiration datetb. If we assume the variance va to be equal to v, and the remaining parameters tobe

γ∗ = 2, v = 0.01, ǫ = 0.1, rW = 0, (20.485)

the price of an option with strike price E = 100 one half year before expirationwith the stock price S = E (this is called an option at-the-money) is 2.83 US$ forρ = −0.5 and 2.81 US$ for ρ = 0.5. The difference with respect to the Black-Scholesprice is shown in Fig. 20.31.

S

ρ = −0.5ρ = 0.5O(S, v, t) −OBS(S, t)

0.1

0.05

−0.05

0

−0.1

80 90 100 120 130 140

Figure 20.31 Difference between option price O(S, v, t) with fluctuating volatility and

Black-Scholes price OBS(S, t) with σ2 = v for option of strike price 100 US$. The param-

eters are given in Eq. (20.485). The noise correlation parameter is once ρ = −0.5 and

once ρ = 0.5. For an at-the-money option the absolute value is 2.83 US$ for ρ = −0.5 and

2.81 US$ for ρ = 0.5 (after Ref. [78]).

Page 89: Economic and Path Integrals

1528 20 Path Integrals and Financial Markets

20.7.8 Perturbation Expansion and Smile

A perturbative treatment of any non-Gaussian distributions D(x), which we assume to be sym-metric, for simplicity, starts from the expansion

D(p) =[

1 +a44!p4 − a6

6!p6 +

c88!p8 − . . .− . . .

]

e−σ2p2/2, (20.486)

where

a4 = c4, a6 = c6, a8 = c8 + 35c24, a10 = c10 + 210c4c6, . . . , (20.487)

which can also be expressed as a series

D(p)=

[

1 +a44!

22(

∂σ2

)2

+a66!

23(

∂σ2

)3

+a88!

24(

∂σ2

)4

+ . . .

]

e−σ2p2/2. (20.488)

By taking the Fourier transform we obtain the expansion of the distributions in x-space

D(x) =

[

1 +a44!

22(

∂σ2

)2

+a66!

23(

∂σ2

)3

+a88!

24(

∂σ2

)4

+ . . .

]

e−x2/2σ2

√2πσ2

=

[

1 +c48

− c648

+35c24384

+c8384

+x2

σ2

(

− c44

+c616

− 35c2496

− c896

)

+x4

σ4

(

c424

− c648

+35c24192

+c8192

)

+x6

σ6

(

c6720

− 35c241440

− c81440

)

+x8

σ8

(

35c2440320

+c8

40320

)

+ . . .

]

e−x2/2σ2

√2πσ2

. (20.489)

The quantities cn contain the kurtosis κ, in the case of a truncated Levy distribution with thepowers εn/2−1. If the distribution is close to a Gaussian, we may re-expand all expressions in powersof the higher cumulants. In the case of a truncated Levy distribution, we may keep systematicallyall terms up to a certain maximal power of ε and find

D(x) =

1 − x2

2σ2

(

c42

− c68

+2c243

+c848

+5c4c6192

− 33c34256

)

+x4

σ4

(

c424

− c648

+17c2496

+c8192

+c4c6288

− 239c349216

)

+x6

σ6

(

c6720

− 7c24288

− c81440

− c4c65760

+7c342304

)

+x8

σ8

(

c241152

+c8

40320− c34

9216

)

+ . . .

e−x2/2σ2

2πσ21

, (20.490)

where we have introduced the modified width

σ21 ≡ σ2

(

1 − c44

+c624

− 13c2496

− c8192

− c4c664

− 31c34512

+ . . .

)

. (20.491)

The prefactor can be taken into the exponent yielding

D(x) = exp

− x2

2σ2

(

1 +c42

+2 c243

− 33 c34256

− c68

+5 c4 c6192

+c848

)

+x4

σ4

(

c424

+7 c2448

− 1007 c349216

− c648

+11 c4 c6

576+

c8192

)

+x6

σ6

(

− c2472

+43 c34768

+c6720

− 23 c4 c62880

− c81440

)

+x8

σ8

(

−101 c349216

+7 c4 c65760

+c8

40320

)

+ . . .

1√

2πσ21

. (20.492)

Page 90: Economic and Path Integrals

20.7 Option Pricing 1529

Introducing a second modified width

σ22 ≡ σ2

(

1 − c42

+c68

− 5c2412

− c848

− 29c4c6192

+515c34768

+ . . .

)

, (20.493)

the exponential can be brought to the form

D(x) = exp

−x2

σ22

+x4

σ42

(

c424

− c648

+5c2448

+c8192

+29c4c6576

− 2576c349216

)

+x6

σ62

(

c6720

− c2472

− 8c81440

− 29c4c62880

+59c34768

)

+x8

σ82

(

c840320

+7c4c65760

− 101c349216

)

+ . . .

1√

2πσ21

. (20.494)

The exponential can be re-expressed compactly in the quasi-Gaussian form

D(x) =1

2πσ21

exp

[

− x2

2Σ2(x)

]

, (20.495)

where we have defined an x-dependent width

Σ(x) ≡ σ22

[

1 +x2

σ22

(

c412

+5 c2424

− 2575 c344608

− c624

+29 c4 c6

288+c896

)

+x4

σ42

(−c2448

+217 c341152

− 1375 c4427648

+c6360

− 13 c4 c6480

− c24 c61728

+c26576

− c8720

+c4 c8576

)

+x6

σ62

(

−359 c3413824

+127 c446912

+5 c4 c61728

− 13 c24 c617280

− c264320

+c8

20160− c4 c8

4320

)

+ . . .

]

.

(20.496)

For small x the deviation of Σ2(x) from a constant σ2 is dominated by the quadratic term. Ifone plots the fluctuation width Σ(x) for the logarithms of the observed option prices one finds thesmile parabola shown before in Fig. 20.29. On the basis of the expansion (20.488) it is possible toderive an approximate option price formula for assets fluctuating according to the truncated Levydistribution. We simply apply the differential operator in front of the Gaussian distribution

O ≡[

1 +a44!

22(

∂σ2

)2

+a66!

23(

∂σ2

)3

+a88!

24(

∂σ2

)4

+ . . .

]

(20.497)

to the Gaussian distribution (20.421). For λ = 3/2, the coefficients are

a4 =ε

M2, a6 =

5 · 7 ε23M3

, a8 =5 · 7 · 11 ε2

M4+

5 · 7 ε2M4

,

a10 =52 · 7 · 11 · 13 ε4

3M5+

22 · 5 · 7 · 17 ε4

3M5, . . . . (20.498)

The operator O winds up in front of the Black-Scholes expression (20.438), such that we obtainthe formal result

OL(xa, ta) = OO(xa, ta) = O[

S(ta)N(y+) − e−(rxW+σ2/2)(tb−ta)EN(y−)

]

, (20.499)

where we have used (20.418) to exhibit the full σ-dependence on the right-hand side. This expres-sion may now be expanded in powers of the kurtosis κ. The term proportional to a4 yields a firstcorrection to the Black-Scholes formula, linear in ε,

O1(xa, ta) = − ε

12M2

(

Se−y2+/2 − e−rW (tb−ta)Ee−y2

−/2) ym√

2πσ2

− e−rW (tb−ta)(tb − ta)EN(y−)

. (20.500)

Page 91: Economic and Path Integrals

1530 20 Path Integrals and Financial Markets

The term proportional to a6 adds a correction proportional to ε2:

O2(xa, ta) =35ε2

3M3

1

23 · 32 · 5[

Se−y2+/2(

y+y2− − 3y+ + 4

σ2(tb − ta))

− e−rW (tb−ta)Ee−y2−/2

(

y3− − 3y− − 2σ2(tb − ta)y−)

] 1√2πσ4

+ e−rW (tb−ta)(tb − ta)2EN(y−)

. (20.501)

The next term proportional to a8 adds corrections proportional to ε2 and ε3:

O3(xa, ta) =385ε3 + 35ε2

M4

1

26 · 32 · 5 · 7×[

Se−y2+/2(

−y3−y2+ + 9y−y2+ + y3− − 15y+ + 18 +

σ2(tb − ta)(12y−y+ + 18))

− e−rW (tb−ta)Ee−y2−/2

(

y5− − 10y3− + 15y− + σ2(tb − ta)(3y3− − 9y−)

+σ4(tb − ta)2 3y−

)]

1√2πσ6

+e−rW (tb−ta)(tb − ta)3EN(y−)

. (20.502)

Since the expansion is asymptotic, an efficient resummation scheme will be needed for practicalapplications.

Appendix 20A Large-x Behavior of TruncatedLevy Distribution

Here we derive the divergent asymptotic expansion in the large-x regime. Using the variabley ≡ p/α and the constant a ≡ −sαλ, we may write the Fourier integral (20.24) as

L(λ,α)σ2 (x) = αe−2a

∞∫

−∞

dy

2πeiαyxea[(1−iy)λ+(1+iy)λ]. (20A.1)

Expanding the last exponential in a Taylor series, we obtain

L(λ,α)σ2 (x) = αe−2a

∞∫

−∞

dy

2πeiyαx

∞∑

n=0

an

n![(1 − iy)λ + (1 + iy)λ]n =

= αe−2a

∞∫

−∞

dy

2πeiαxy

∞∑

n=0

an

n!

n∑

m=0

( n

m

)

(1 − iy)λ(n−m)(1 + iy)λm (20A.2)

containing binominal coefficients. Changing the order of summations yields

L(λ,α)σ2 (x) = αe−2a

∞∑

n=0

an

n!

n∑

m=0

( n

m

)

∞∫

−∞

dy

2πeiαxy(1 − iy)λ(n−m)(1 + iy)λm. (20A.3)

This can be written with the help of the Whittaker functions (20.30) as

L(λ,α)σ2 (x) = αe−2a

∞∑

n=0

an

n!

n∑

m=0

( n

m

) (αx)−1−λn/22λn/2

Γ(−λm)Wλn/2−λm,(λn+1)/2(2αx). (20A.4)

Page 92: Economic and Path Integrals

Appendix 20A Large-x Behavior of Truncated Levy Distribution 1531

After converting the Gamma functions of negative arguments into those of positive arguments,this becomes

L(λ,α)σ2 (x) = −α

πe−2a

∞∑

n=1

ann∑

m=0

2λn/2Γ(1+λm) sin(πλm)

(αx)1+λn/2m!(n−m)!Wλn/2−λm,(λn+1)/2(2αx).

(20A.5)

The Whittaker functions Wλ,γ(x) have the following asymptotic expansion

Wλ,γ(x) = e−x/2xλ

1 +

∞∑

k=1

1

k!xk

k∏

j=1

[

γ2 − (λ− j + 1/2)2]

. (20A.6)

For γ ≡ (λn+ 1)/2 and λ ≡ λ(n− 2m)/2, the product takes the form

k∏

j=1

[γ2 − (λ− j + 1/2)2] =

k∏

j=1

(

λn+ 1

2

)2

−[

λ(n− 2m)

2− j +

1

2

]2

=

k∏

j=1

(λm+ j)(λn+1 − λm− j). (20A.7)

Inserting this into (20A.6) and the result into (20A.5), we obtain the asymptotic expansion forlarge x:

L(λ,α)σ2 (x) = − 1

πe−2a e

−αx

x

∞∑

n=1

an 2λnn∑

m=1

Γ(1 + λm) sin(πλm)

m!(n−m)!(2αx)−λm

×[

1 +

∞∑

k=1

∏kj=1(λm+ j)(λn + 1 − λm− j)

k!(2αx)k

]

. (20A.8)

We have raised the initial value of the index of summation m by one unit since sin(πλm) vanishes

for m = 0. If we define a product of the form∏0

j=1 ... to be equal to unity, we can write the termin the last bracket as

∞∑

k=0

1

k!(2αx)k

k∏

j=1

(j + λm)(1 − λm− j + λn). (20A.9)

Rearranging the double sum in (20A.8), we write L(λ,α)σ2 (x) as

L(λ,α)σ2 (x) = − 1

πe−2a e

−αx

x

∞∑

m=1

Γ(1 + λm) sin(πλm)

m!(2αx)λm

∞∑

n=m

an 2λn

(n−m)!

×∞∑

k=0

1

k!(2αx)k

k∏

j=1

(j + λm)(1 − λm− j + λn), (20A.10)

and further as

L(λ,α)σ2 (x) = − 1

πe−2a e

−αx

x

∞∑

m=1

Γ(1 + λm) sin(πλm)

m!(2αx)−λm

×∞∑

n=0

(2λa)n+m 1

n!

∞∑

k=0

1

k!(2αx)k

k∏

j=1

(j + λm)(1 − j + λn) =

Page 93: Economic and Path Integrals

1532 20 Path Integrals and Financial Markets

= − 1

πe−2a e

−αx

x

∞∑

k=0

1

k!(2αx)k

∞∑

m=1

Γ(1 + λm) sin(πλm)(2λa)m

(2αx)λmm!

×k∏

j′=1

(j′ + λm)

∞∑

n=0

(2λa)n

n!

k∏

j=1

(λn+ 1 − j). (20A.11)

The last sum over n in this expression can be re-expressed more efficiently with the help of agenerating function

f (k)(yλ) ≡ dk

dykey

λ

, (20A.12)

whose Taylor series is

f (k)(yλ) =dk

dyk

∞∑

n=0

yλn

n!=

∞∑

n=0

1

n!

k∏

j=1

(λn− j + 1)yλn−k, (20A.13)

leading to

L(λ,α)σ2 (x) =

1

πe−2a e

−αx

x

∞∑

k=0

ak/λ2kf (k)(2λa)

k!(2αx)k

×∞∑

m=1

Γ(1 + λm) sin(πλm)(2λa)m

m!(2αx)−λm

k∏

j=1

(λm+ j).

(20A.14)

This can be rearranged to

L(λ,α)σ2 (x) = −e

−2a

πe−αx

∞∑

k=0

(−s)k/λf (k)(−s(2α)λ)

k!

×∞∑

m=1

Γ(1 + λm+ k) sin(πλm)(−s)mm!x1+λm+k

.

(20A.15)

Thus we obtain the asymptotic expansion

1 1.2 1.4 1.6 1.8 2

0.02

0.04

0.06

0.08

0.1

x

L(µ,α)σ2

(K)(x)

Figure 20.32 Comparison of large-x expansions containing different numbers of terms

(with K = 0, 1, 2, 3, 4, 5, with increasing dash length) with the tails of the truncated

Levy distribution for λ =√

2, σ = 0.5, α = 1.

Page 94: Economic and Path Integrals

Appendix 20B Gaussian Weight 1533

L(λ,α)σ2 (x) = −e

−2a

π

e−αx

x

∞∑

k=0

Ak

∞∑

m=1

Bkm(−s)mxλm+k

, (20A.16)

where

Ak =(−s)k/λf (k)(−s(2α)λ)

k!, Bkm =

Γ(1 + λm+ k) sin(πλm)

m!. (20A.17)

We shall denote by L(λ,α)σ2

(K)(x) the approximants in which the sums over K are truncated afterthe Kth term. To have a definite smallest power in |x|, the sum over m is truncated after thesmallest integer larger than (K − k + λ)/λ. The leading term is

L(λ,α)σ2

(0)(x) = −e−2a

πe−αxf (0)(−s(2α)λ)

Γ(1 + λ) sin(πλ)(−s)x1+λ

= seα

λs(2−2λ)

π

Γ(1 + λ) sin(πλ)

x1+λe−αx. (20A.18)

The large-x approximations L(λ,α)σ2

(0)(x) are compared with the numerically calculated truncatedLevy distribution in Fig. 20.32.

Appendix 20B Gaussian Weight

For simplicity, let us study the Gaussian content in the final distribution (20.347) only for thesimpler case ρ = 0. Then we shift the contour of integration in (20.347) to run along p+ i/2, andwe study the Fourier integral

P (x t |xata) = e−∆x/2

∫ +∞

−∞

dp

2πeip∆x−H(p,∆t), (20B.1)

where

H(p,∆t) = −γ2vt

κ2+

2γv

κ2ln

[

coshΩt

2+

Ω2 + γ2

2γΩsinh

Ωt

2

]

, (20B.2)

and

Ω =√

γ2 + κ2(p2 + 1/4). (20B.3)

The function H(p,∆t) is real and symmetric in p. The integral (20B.1) is therefore a symmetricfunction of ∆x. The only source of asymmetry of P (x t |xata) in ∆x is the exponential prefactorin (20B.1).

Let us expand the integral in (20B.1) for small ∆x:

P (x t |xata) ≈ e−∆x/2

[

µ0 −1

2µ2(∆x)

2

]

≈ µ0 e−∆x/2e−µ2(∆x)2/2µ0 , (20B.4)

where the coefficients are the first and the second moments of exp[H(p,∆t)]

µ0(∆t) =

+∞∫

−∞

dp

2πe−H(p,∆t), µ2(∆t) =

+∞∫

−∞

dp

2πp2e−H(p,∆t). (20B.5)

If we ignore the existence of semi-heavy tails and extrapolate the Gaussian expression on the right-hand side to ∆x ∈ (−∞,∞), the total probability contained in such a Gaussian extrapolation willbe the fraction

f(∆t) =

+∞∫

−∞

d∆xµ0 e−∆x/2−µ2∆x2/2µ0 =

2πµ30

µ2eµ0/8µ2 . (20B.6)

Page 95: Economic and Path Integrals

1534 20 Path Integrals and Financial Markets

This is always less than 1 since the integral (20B.6) ignores the probability contained in the semi-heavy tails. The difference 1 − f(∆t) measures the relative contribution of the semi-heavy tails.The parameters µ0(∆t) and µ2(∆t) are calculated numerically and the resulting fraction f(∆t)is plotted in Fig. 20.25 as a function of ∆t. For ∆t → ∞, the distribution becomes Gaussian,whereas for small ∆t, it becomes a broad function of p.

Appendix 20C Comparison with Dow-Jones Data

For the comparison of the theory in Section 20.4 with actual financial data shown in Fig. 20.22,the authors of Ref. [79] downloaded the daily closing values of the Dow-Jones industrial index forthe period of 20 years from 1 January 1982 to 31 December 2001 from the Web site of Yahoo[110]. The data set contained 5049 points S(tn), where the discrete time variable tn parametrizesthe days. Short days before holidays were ignored. For each tn, they compiled the log-returns∆x(tn) = lnS(tn+1)/S(tn). Then they partitioned the x-axis into equally spaced intervals ofwidth ∆x and counted the number of log-returns ∆x(tn) falling into each interval. They omittedall intervals with occupation numbers less than five, which they considered as too few to rely on.Only less than 1% of the entire data set was omitted in this way. Dividing the occupation numberof each bin by ∆r and by the total occupation number of all bins, they obtained the probabilitydensity for a given time interval ∆t = 1 day. From this they found P (DJ)(x t |xata) by replacing∆x→ ∆x − rS∆t.

Assuming that the system is ergodic, so that ensemble averaging is equivalent to timeaveraging, they compared P (DJ)(x t |xata) with the calculated P (x t |xata) in Eq. (20.347).The parameters of the model were determined by minimizing the mean-square deviation∑

∆x,∆t | logP (DJ)(x t |xata) − logP (x t |xata)|2, with the sum taken over all available ∆x andover ∆t = 1, 5, 20, 40, and 250 days. These values of ∆t were selected because they representdifferent regimes: γ∆t≪ 1 for t = 1 and 5 days, γ∆t ≈ 1 for t = 20 days, and γ∆t≫ 1 for t = 40and 250 days. As Figs. 20.22 and 20.23 illustrate, the probability density P (x t |xata) calculatefrom the Fourier integral (20.347) with components (20.348) agrees with the data very well, notonly for the selected five values of time t, but for the whole time interval from 1 to 250 tradingdays. The comparison cannot be extended to ∆t longer than 250 days, which is approximately1/20 of the entire range of the data set, because it is impossible to reliably extract P (DJ)(x t |xata)from the data when ∆t is too long.

The best fits for the four parameters γ, v, ε, µ are given in Table 20.1. Within the scatteringof the data, there are no discernible differences between the fits with the correlation coefficientρ being zero or slightly different from zero. Thus the correlation parameter ρ between the noiseterms for stock price and variance in Eq. (20.310) is practically zero. This conclusion is in contrastwith the value ρ = −0.58 found in [111] by fitting the leverage correlation function introduced in[112]. Further study is necessary to understand this discrepancy. All theoretical curves shown inthe above figures are calculated for ρ = 0, and fit the data very well.

The parameters γ, v, ε, µ have the dimensionality of 1/time. One row in Table 20.1 gives theirvalues in units of 1/day, as originally determined in our fit. The other row shows the annualizedvalues of the parameters in units of 1/year, where one year is here equal to the average numberof 252.5 trading days per calendar year. The relaxation time of variance is equal to 1/γ = 22.2trading days = 4.4 weeks ≈ 1 month, where 1 week = 5 trading days. Thus one finds that thevariance has a rather long relaxation time, of the order of one month, which is in agreement withan earlier conclusion in Ref. [111].

Using the numbers given in Table 20.1, the value of the parameter ε2/2γv is ≈ 0.772 thussatisfying the smallness condition Eq. (20.322) ensuring that v never reaches negative values.

The stock prices have an apparent growth rate determined by the position xm(t) where theprobability density is maximal. Adding this to the initially subtracted growth rate rS we findthat the apparent growth rate is rS = rS − γv/2ω0 = 13% per year. This number coincides withthe apparent average growth rate of the Dow-Jones index obtained by a simple fit of the datapoints Stn with an exponential function of tn. The apparent growth rate rS is comparable to the

Page 96: Economic and Path Integrals

Notes and References 1535

Table 20.1 Parameters of equations with fluctuating variance obtained from fits to Dow-

Jones data. The fit yields ρ ≈ 0 for the correlation coefficient and 1/γ = 22.2 trading days

for the relaxation time of variance.

Units γ v ε µ1/day 4.50 × 10−2 8.62 × 10−5 2.45 × 10−3 5.67 × 10−4

1/year 11.35 0.022 0.618 0.143

average stock volatility after one year σ =√v = 14.7%. Moreover, the parameter (20.328) which

characterizes the width of the stationary distribution of variance is equal to vmax/w = 0.54. Thismeans that the distribution of variance is broad, and variance can easily fluctuate to a value twicegreater than the average value v. As a consequence, even though the average growth rate of the

stock index is positive, there is a substantial probability of∫ 0

−∞ d∆xP (x t |xata) ≈ 17.7% to havenegative growth for ∆t = 1 year.

According to (20.368), the asymmetry between the slopes of exponential tails for positiveand negative ∆x is given by the parameter p0, which is equal to 1/2 when ρ = 0 [see also thediscussion of Eq. (20B.1) in Appendix 20B]. The origin of this asymmetry can be traced backto the transformation from S(t)/S(t) to x(t) using Ito’s formula. This produces a term v(t)/2 inEq. (20.306), which leads to the first term in the Hamiltonian operator (20.319). For ρ = 0 thisis the only source of asymmetry in ∆x of P (x t |xata). In practice, the asymmetry of the slopesp0 = 1/2 is quite small (about 2.7%) compared to the average slope q±∗ ≈ ω0/ε = 18.4.

Notes and References

Option pricing beyond Black and Scholes via path integrals was discussed in Ref. [113], wherestrategies are devised to minimize risks in the presence of extreme fluctuations, as occur on realmarkets. See also Ref. [114].Recently, a generalization of path integrals to functional integrals over surfaces has been proposedin Ref. [115] as an alternative to the Heath-Jarrow-Morton approach of modeling yield curves (seehttp://risk.ifci.ch/00011661.htm).

The individual citations refer to

[1] The development of the Dow Jones industrial index up to date can be found on inter-net sites such as http://stockcharts.com/charts/historical/djia1900.html. Alter-natively they can be plotted directly using Stephen Wolfram’s program Mathematica (v.6)using the small programt=FinancialData["^DJI",All];l = Length[t]

tt = Table[t[[k]][[1]][[1]]+(t[[k]][[1]][[2]]-1)/12,t[[k]][[2]], k,1,l]ListLogPlot[tt]

Note that the last stagnation period was predicted in the 3rd edition of this book in 2004.

[2] P. Fizeau, Y. Liu, M. Meyer, C.-K. Peng, and H.E. Stanley, Volatility Distribution in the

S&P500 Stock Index , Physica A 245, 441 (1997) (cond-mat/9708143).

[3] B.E. Baaquie, A Path Integral Approach to Option Pricing with Stochastic Volatility: Some

Exact Results , J. de Physique I 7, 1733 (1997) (cond-mat/9708178).

[4] J.P. Fouque, G. Papanicolaou, and K.R. Sircar, Derivatives in Financial Markets with

Stochastic Volatility (Cambridge University Press, Cambridge, 2000); International Jour-nal of Theoretical and Applied Finance, 3, 101 (2000).

[5] J. Hull and A. White, Journal of Finance 42, 281 (1987); C.A. Ball and A. Roma, Journalof Financial and Quantitative Analysis 29, 589 (1994); R. Schobel and J. Zhu, EuropeanFinance Review 3, 23 (1999).

Page 97: Economic and Path Integrals

1536 20 Path Integrals and Financial Markets

[6] E.M. Stein and J.C. Stein, Review of Financial Studies 4, 727 (1991).

[7] J.C. Cox, J.E. Ingersoll, and S.A. Ross, Econometrica 53, 385 (1985).In the mathematical literature such an equation runs under the name Feller process. SeeRef. [4] andW. Feller, Probability Theory and its Applications , sec. ed., Vol. II., John Wiley & Sons,1971.See also related work in turbulence byB. Holdom, Physica A 254, 569 (1998) (cond-mat/9709141).

[8] P. Wilmott, Derivatives, John Wiley & Sons, New York, 1998.

[9] R. Courant and D. Hilbert, Methods of Mathematical Physics, vol. 2 (John Wiley & Sons,New York, 1962).

[10] C.M. Bender and S.A. Orszag, Advanced Mathematical Methods for Scientists and Engineers,Springer, New York, 1999.

[11] C.G. Lamoureux and W.D. Lastrapes, Forecasting Stock-Return Variance: Toward an Un-

derstanding of Stochastic Implied Volatilities, Rev. of Financial Studies 6, 293 (1993). SeealsoD.T. Breeden, An Intertemporal Asset Pricing Model with Stochastic Consumption and In-

vertment Opportunities, Jour. of Financial Economics 7, 265 (1979).

[12] This is named after the Russian mathematician Andrei Andrejewitsch Markov who publishedin 1912 the German book on Probalility Theory cited on p. 1093 with the German spellingMarkoff.

[13] J.-P. Bouchaud and M. Potters, Theory of Financial Risks , From Statistical Physics to Risk

Management , Cambridge University Press, 2000.

[14] J. Hull, Options, Futures and Other Derivatives , Prentice-Hall Int., 1997.

[15] R. Rebonato, Interest-Rate Option Models , John Wiley & Sons, Chichester, 1996.

[16] M.W. Baxter and A.J.O. Rennie, Financial Calculus , Cambridge Univ. Press, Cambridge,1996.

[17] J. Voigt, The Statistical Mechanics of Financial Markets , Springer, Berlin, 2001.

[18] V. Pareto, Giornale degli Economisti, Roma, January 1895; and Cours d’economie politique,F. Rouge Editeur, Lausanne and Paris, 1896; reprinted in an edition of his complete works(Vol III) under the title Ecrits sur la courbe de la repartition de la richesse, Librairie Droz,Geneva, 1965 (http://213.39.120.146:8200/droz/FMPro).

[19] B.B. Mandelbrot, Fractals and Scaling in Finance, Springer, Berlin, 1997; J. of Business36, 393 (1963).

[20] T. Lux, Appl. Financial Economics 6, 463 (1996);M. Loretan and P.C.B. Phillips, J. Empirical Finance 1, 211 (1994).

[21] J.P. Nolan, Stable Distributions, American University (Wahington D.C.) lecture notes 2004http://academic2.american.edu/~jpnolan/stable/chap1.pdf.

[22] R.N. Mantegna and H.E. Stanley, Stochastic Process with Ultraslow Convergence to a Gaus-

sian: The Truncated Levy Flight , Phys. Rev. Lett. 73, 2946 (1994).

[23] P. Gopikrishnan, M. Meyer, L.A.N. Amaral, and H.E. Stanley, Europ. Phys. Journ. B 3,139 (1998);P. Gopikrishnan, V. Plerou, L.A.N. Amaral, M. Meyer, and H.E. Stanley, Scaling of the

distribution of fluctuations of financial market indices , Phys. Rev. E 60, 5305 (1999);V. Plerou, P. Gopikrishnan, L.A.N. Amaral, M. Meyer, and H.E. Stanley, Scaling of the

Page 98: Economic and Path Integrals

Notes and References 1537

distribution of price fluctuations of individual companies , Phys. Rev. E 60, 6519 (1999);V. Plerou, P. Gopikrishnan, X. Gabaix, H.E. Stanley, On the Origin of Power-Law Fluctu-

ations in Stock Prices , Quantitative Finance 4, C11 (2004).

[24] I. Koponen, Analytic Approach to the Problem of Convergence of Truncated Levy Flights

Towards the Gaussian Stochastic Process , Phys. Rev. E 52, 1197-1199 (1995).

[25] I.S. Gradshteyn and I.M. Ryzhik, op. cit., Formulas 3.382.6 and 3.382.7.

[26] ibid., op. cit., Formula 3.384.9.

[27] ibid., op. cit., Formula 9.220.3 and 9.220.4.

[28] This result was calculated by A. Lyashin when he visited my group in Berlin. Note thatother authors have missed the prefactor in the asymptotic behavior (20.37), for exampleA. Matacz, Financial Modeling and Option Theory with the Truncated Levy Process , cond-mat/9710197. The prefactor can be dropped only for α = 0.

[29] ibid., op. cit., Formulas 3.462.3.

[30] ibid., op. cit., Formula 9.246.

[31] ibid., op. cit., Formula 9.247.2.

[32] Exponential short-time behavior has been observed in Ref. [13] and numerous other authors:L.C. Miranda and R. Riera, Physica A 297, 509 (2001);J.L. McCauley and G.H. Gunaratne, Physica A 329, 178 (2003);T. Kaizoji, Physica A 343, 662 (2004);R. Remer and R. Mahnke, Physica A 344, 236 (2004);D. Makowiec, Physica A 344, 36 (2004);K. Matia, M. Pal, H. Salunkay, and H.E. Stanley, Europhys. Lett. 66, 909 (2004);A.C. Silva, R.E. Prange, and V.M. Yakovenko, Physica A 344, 227 (2004);R. Vicente, C.M. de Toledo, V.B.P. Leite, and N. Caticha, Physica A 361, 272 (2006);A.C. Silva and V.M. Yakovenko, (physics/0608299).

[33] H. Kleinert and T. Xiao-Jiang to be published.

[34] B. Grigelionis, Processes of Meixner Type, Lith. Math. J. 39, 33 (1999).

[35] W. Schoutens, Meixner Processes in Finance, Report 2001-002, EURANDOM, Eindhoven(www.eurandom.tue.nl/reports/2001/002wsreport.ps).

[36] O. Barndorff-Nielsen, T. Mikosch, S. Resnick, eds., Levy Processes — Theory and Applica-

tions Birkhauser, 2001.

[37] O. Barndorff-Nielsen, Infinite Divisibility of the Hyperbolic and Generalized Inverse Gaus-

sian Distributions , Zeitschrift fur Wahrscheinlichkeitstheorie und verwandte Gebiete 38,309-312 (1977).

[38] O. Barndorff-Nielsen, Processes of Normal Inverse Gaussian Type, Finance & Stochastics,2, No. 1, 41-68 (1998).

[39] O. Barndorff-Nielsen, Normal Inverse Gaussian Distributions and Stochastic Volatility Mod-

eling, Scandinavian Journal of Statistics 24, 1-13 (1977).

[40] O. Barndorff-Nielsen, N. Shephard, Modeling with Levy Processes for Financial Economet-

rics , MaPhySto Research Report No. 16, University of Aarhus, (2000).

[41] O. Barndorff-Nielsen, N. Shephard, Incorporation of a Leverage Effect in a Stochastic

Volatility Model , MaPhySto Research Report No. 18, University of Aarhus, (1998).

[42] O. Barndorff-Nielsen, N. Shephard, Integrated Ornstein Uhlenbeck Processes , Research Re-port, Oxford University, (2000).

Page 99: Economic and Path Integrals

1538 20 Path Integrals and Financial Markets

[43] J. Bertoin, (1996) Levy Processes, Cambridge University Press.

[44] J. Bretagnolle, Processus a increments independants , Ecole d’Ete de Probabilites, LectureNotes in Mathematics, Vol. 237, pp 1-26, Berlin, Springer, (1973).

[45] P.P. Carr, D. Madan, Option Valuation using the Fast Fourier Transform, Journal of Com-putational Finance 2, 61-73 (1998).

[46] P.P. Carr, H. Geman, D. Madan, M. Yor, The Fine Structure of Asset Returns: an Empirical

Investigation, Working Paper, (2000).

[47] T. Chan, Pricing Contingent Claims on Stocks Driven by Levy Processes , Annals of AppliedProbability 9, 504-528, (1999).

[48] R. Cont, Empirical Properties of Asset Returns: Stylized Facts and Statistical Issues , Quan-titative Finance 1, No. 2, (2001).

[49] R. Cont, J.-P. Bouchaud, M. Potters, Scaling in Financial Data: Stable Laws and Beyond , inB. Dubrulle, F. Graner & D. Sornette (eds.): Scale invariance and beyond, Berlin, Springer,(1997).

[50] ibid., op. cit., Formula 8.486.10.

[51] See http://www.fdm.uni-freiburg.de/UK.

[52] P. Jizba, H. Kleinert, Superposition of Probability Distributions , Phys. Rev. E 78, 031122(arXiv:0802.069).

[53] G. Arfken, Mathematical Methods for Physicists , 3rd ed., Academic Press, Orlando, FL. Seet15.12 Inverse Laplace Transformation”, pp. 853-861, 1985.

[54] E. Post, Trans. Amer. Math. Soc. 32 (1930) 723.

[55] H. Kleinert, Stochastic Calculus for Assets with Non-Gaussian Price Fluctuations, PhysicaA 311, 538 (2002) (cond-mat/0203157).

[56] R.F. Pawula, Phys. Rev. 162, 186 (1967).

[57] See http://www.physik.fu-berlin.de/~kleinert/b5/files .

[58] L. Borland, A Theory of Non-Gaussian Option Pricing, Quantitative Finance 2, 415 (2002)(cond-mat/0205078).

[59] C. Tsallis, J. Stat. Phys. 52, 479 (1988); E.M.F. Curado and C. Tsallis, J. Phys. A 24, L69(1991); 3187 (1991); A 25, 1019 (1992).

[60] C. Tsallis, C. Anteneodo, L. Borland, R. Osorio, Nonextensive Statistical Mechanics and

Economics , Physica A 324, 89 (2003) (cond-mat/030130).

[61] P. Praetz The Distribution of Share Price Changes , Journal of Business 45, 49 (1972).

[62] R. Blattberg and N. Gonedes, A Comparison of the Stable and Student Distributions as

Statistical Models of Stock Prices , Journal of Business 47, 244 (1972).

[63] R.S. Liptser and A.N. Shiryaev, Theory of Martingales , Kluwer, 1989.

[64] D. Duffie, Dynamic asset pricing theory , Princeton University Press, 2001, p. 22.

[65] J.M. Steele, Stochastic Calculus and Financial Applications , Springer, New York, 2001, p.50.

[66] H. Kleinert, Option Pricing from Path Integral for Non-Gaussian Fluctuations. Natural

Martingale and Application to Truncated Levy Distributions , Physica A 312, 217 (2002)(cond-mat/0202311).

Page 100: Economic and Path Integrals

Notes and References 1539

[67] F. Esscher, On the Probability Function in the Collective Theory of Risk , SkandinaviskAktuarietidskrift 15, 175 (1932).

[68] H.U. Gerber and E.S.W. Shiu, Option Pricing by Esscher Transforms , Trans. Soc. Ac-turaries 46, 99 (1994).

[69] J.M. Harrison and S.R. Pliska, Martingales and Stochastic Integrals in the Theory of Contin-

uous Trading, Stoch. Proc. Appl. 11, 215 (1981); A Stochastic Calculus Model of Continuous

Trading Complete Markets, 13, 313 (1981);

[70] E. Eberlein, J. Jacod, On the Range of Options Prices , Finance and Stochastics 1, 131,(1997).

[71] E. Eberlein, U. Keller, Hyperbolic Distributions in Finance, Bernoulli 1, 281-299, (1995).

[72] K. Prause, The Generalized Hyperbolic Model: Estimation, Financial Derivatives,

and Risk Measures , Universitat Freiburg Dissertation, 1999 (http://www.freidok.uni-freiburg.de/volltexte/15/pdf/15 1.pdf).

[73] E. Eberlein, U. Keller, K. Prause, New Insights into Smile, Mispricing and Value at Risk:

the Hyperbolic Model , Journal of Business 71, No. 3, 371-405, (1998).

[74] E. Eberlein, S. Raible, Term Structure Models Driven by General Levy Processes , Mathe-matical Finance 9, 31-53, (1999).

[75] See the Wikipedia pageen.wikipedia.org/wiki/Levy skew alpha-stable distribution.

[76] S. Raible, Levy Processes in Finance, Ph.D. Thesis, Univ. Freiburg, 2000(http://www.freidok.uni-freiburg.de/volltexte/15/pdf/51 1.pdf).

[77] M.H.A. Davis, A General Option Pricing Formula, Preprint, Imperial College, London(1994). See also Ref. [70] andT. Chan, Pricing Contingent Claims on Stocks Driven by Levy Processes, Ann. Appl.Probab. 9, 504 (1999).

[78] S.L. Heston, A Closed-Form Solution for Options with Stochastic Volatility with Applications

to Bond and Currency Options, Review of Financial Studies 6, 327 (1993).

[79] A.A. Dragulescu and V.M. Yakovenko, Quantitative Finance 2, 443 (2002) (cond-mat/0203046).

[80] P. Jizba, H. Kleinert, and P. Haener, Perturbation Expansion for Option Pricing with

Stochastic Volatility, Berlin preprint 2007 (arXiv:0708.3012).

[81] S. Micciche, G. Bonanno, F. Lillo, R.N. Mantegna, Physica A 314, 756 (2002).

[82] D. Valenti, B. Spagnolo, and G. Bonanno, Physica A 314, 756 (2002).

[83] H. Follmer and M. Schweizer, Hedging and Contingent Claims under Incomplete Informa-

tion, in Applied Stochastic Analysis , edited by M.H.A. Davis and R.J. Elliot, 389 Gordonand Breach 1991.

[84] R.F. Engle, Autoregressive conditional heteroskedasticity with estimates of the variance

of UK inflation, Econometrica 50, 987 (1982); Dynamic conditional correlation–A sim-

ple class of multivariate GARCH models , J. of Business and Econ. Stat., 20, 339 (3002)(http://www.physik.fu-berlin.de/~kleinert/finance/engle1.pdf);R.F. Engle and K.F. Kroner, Multivariate simultaneous generalized ARCH , EconometricTheory 11, 122 (1995);R.F. Engle and K. Sheppard, Theoretical and empirical properties of dynamic conditional

correlation multivariate GARCH , Nat. Bur. of Standards working paper No. 8554 (2001)

Page 101: Economic and Path Integrals

1540 20 Path Integrals and Financial Markets

(http://www.nber.org/papers/w8554.pdf).See also the GARCH toolbox for calculations at http://www.kevinsheppard.com/

research/ucsd garch/ucsd garch.aspx.

[85] T. Bollerslev Generalized autoregressive conditional heteroskedasticity, J. of Econometrics31, 307 (1986); Modelling the coherence in short-run nominal exchange rates: A multivariate

generalized ARCH model , Rev. Economics and Statistics 72, 498 (1990).

[86] L. Bachelier, Theorie de la Speculation, L. Gabay, Sceaux, 1995 (reprinted in P. Cootner(ed.), The Random Character of Stock Market Prices , MIT Press, Cambridge, Ma, 1964,pp. 17–78.

[87] A. Einstein, Uber die von der molakularkinetischen Theorie der Warme geforderte Bewegung

von in ruhenden Flussigkeiten suspendierten Teilchen, Annalen der Physik 17, 549 (1905).

[88] N. Wiener, Differential-Space, J. of Math. and Phys. 2, 131 (1923).

[89] F. Black and M. Scholes, J. Pol. Economy 81, 637 (1973).

[90] R.C. Merton, Theory of Rational Option Pricing, Bell J. Econ. Management Sci. 4, 141(1973).

[91] These papers are available as CNRS preprints CPT88/PE2206 (1988) and CPT89/PE2333(1989). Since it takes some effort to obtain them I have placed them on the internetwhere they can be downloaded as files dash1.pdf and dash2.pdf from http://www.phy-

sik.fu-berlin.de/~kleinert/b3/papers.

[92] S. Fedotov and S. Mikhailov, preprint (cond-mat/9806101).

[93] M. Otto, preprints (cond-mat/9812318) and (cond-mat/9906196).

[94] See the Wikipedia entry: http://en.wikipedia.org/wiki/Greeks (finance).

[95] J.P.A. Devreese, D. Lemmens, and J. Tempere, Physica A 389, 780 (2009).

B.E. Baaquie, L.C. Kwek, and M. Srikant, Simulation of Stochastic Volatility using Path

Integration: Smiles and Frowns , cond-mat/0008327

[96] R. Cond, Scaling and Correlation in Financial Data, (cond-mat/9705075).

[97] A.J. McKane, H.C. Luckock, and A.J. Bray, Path Integrals and Non-Markov Processes. I.

General Formalism, Phys. Rev. A 41, 644 (1990);A.N. Drozdov and J.J. Brey, Accurate Path Integral Representations of the Fokker-Planck

Equation with a Linear Reference System: Comparative Study of Current Theories , Phys.Rev. E 57, 146 (1998);V. Linetsky, The Path Integral Approach to Financial Modeling and Options Pricing, Com-putational Economics 11 129 (1997);E.F. Fama, Efficient Capital Markets , Journal of Finance 25, 383 (1970).A. Pagan, The Econometrics of Financial Markets , Journal of Empirical Finance 3, 15(1996).C.W.J. Granger, Z.X. Ding, Stylized Facts on the Temporal Distribution Properties of Daily

Data from Speculative Markets , University of San Diego Preprint, 1994.

[98] H. Geman, D. Madan, M. Yor, Time Changes in Subordinated Brownian Motion, Preprint,(2000).

[99] H. Geman, D. Madan, M. Yor, Time Changes for Levy Processes , Preprint (1999).

[100] J. Jacod, A.N. Shiryaev, Limit Theorems for Stochastic Processes , Berlin, Springer, (1987).

[101] P. Levy, Theorie de l’addition des variables aliatoires , Paris, Gauthier Villars, (1937).

Page 102: Economic and Path Integrals

Notes and References 1541

[102] D. Madan, F. Milne, Option Pricing with Variance Gamma Martingale Components , Math-ematical finance, 1, 39-55, (1991).

[103] D. Madan, P.P. Carr, E.C. Chang, The Variance Gamma Process and Option Pricing,European Finance Review 2, 79-105, (1998).

[104] D. Madan, E. Seneta, The Variance Gamma Model for Share Market Returns , Journal ofBusiness 63, 511-524, (1990).

[105] B.B. Mandelbrot, Fractals and Scaling in Finance, Berlin, Springer, (1997).

[106] P. Protter, Stochastic Integration and Differential Equations: a new approach, Berlin,Springer, (1990).

[107] T.H. Rydberg, The Normal Inverse Gaussian Levy Process: Simulation and Approximation,

Commun. Stat., Stochastic Models 13(4), 887-910 (1997).

[108] G. Samorodnitsky, M. Taqqu, Stable Non-Gaussian Random Processes , New York, Chap-man and Hall (1994).

[109] K. Sato, Levy Processes and Infinitely Divisible Distributions , Cambridge University Press,(1999).

[110] Yahoo Finance http://finance.yahoo.com. To download data, enter in the symbol box:^DJI, and then click on the link: Download Spreadsheet .

[111] J. Masoliver and J. Perello, Physica A 308, 420 (2002) (cond-mat/0111334); Phys. Rev. E67, 037102 (2003) (cond-mat/0202203); (physics/0609136).

[112] J.-P. Bouchaud, A. Matacz, and M. Potters, Phys. Rev. Letters 87, 228701 (2001).

[113] J.-P. Bouchaud and D. Sornette, The Black-Scholes Option Pricing Problem in mathematical

finance: Generalization and extensions for a large class of stochastic processes , J. de Phys.4, 863 (1994); Reply to Mikheev’s Comment on the Black-Scholes Pricing Problem, J. dePhys. 5, 219 (1995);

[114] J.-P. Bouchaud, G. Iori, and D. Sornette, Real-World Options , Risk 9, 61 (1996)(http://xxx.lanl.gov/abs/cond-mat/9509095);J.-P. Bouchaud, D. Sornette, and M. Potters, Option Pricing in the Presence of Extreme

Fluctuations , in Mathematics of Derivative Securities , ed. by M.A.H. Dempster and S.R.Pliska, Cambridge University Press, 1997, pp. 112-125; The Black-Scholes Option Pric-

ing Problem in Mathematical Finance: Generalization and Extensions for a Large Class of

Stochastic Processes , J. de Phys. 4, 863 (1994).

[115] P. Santa-Clara and D. Sornette, The Dynamics of the Forward Interest Rate Curve with

Stochastic String Shocks , Rev. of Financial Studies 14, 149 (2001) (cond-mat/9801321).