Lectures On Financial: Statistics, Stochastics, and Optimization
Lectures On Financial: Statistics, Stochastics, and Optimization
Lectures On Financial: Statistics, Stochastics, and Optimization
SHIRYAEV
Steklov Mathematical Institute,
Lomonosov Moscow State University
LECTURES on
Financial
Statistics, Stochastics, and Optimization
e-mail: albertsh@mi.ras.ru
1
TOPIC I. Financial models and innovations in stochastic
economics
1. The classical and neoclassical models of the dynamics of
the prices driven by Brownian motion and Lévy processes
2. Stylized facts
3. Constructions based on the change of time, stochastic
volatility
I-1
The construction of the right probability-statistical models of the
dynamics of prices of the basic financial instruments (bank account,
bonds, stocks, etc.) is undoubtedly one of important steps for successful
application of the results of mathematical finance and financial
engineering.
I-2
THE FIRST CLASSICAL MODELS FOR
PRICE DYNAMICS
In the sequel,
St = S0 + µt + σBt ,
I-3
M. Kendall (1953). The analysis of economic time series.
Part 1. Prices (J. Roy. Statist. Soc., 96, 11–25):
! σ2 "
St = S0eHt , Ht = µ − t + σBt ,
2
I-5
MARTINGALE APPROACH TO STUDYING THE MODELS
S = (Sn)n≥0, Sn = S0eHn , Hn = h1 + · · · + hn
(hn = log(Sn/Sn−1) is a “return”, “logarithmic return”)
AR(p) model:
µn = a0 + a1hn−1 + · · · + aphn−p, σn = const
MA(q) model:
µn = b0 + b1εn−1 + · · · + bq εn−q , σn = const
ARMA(p, q) model:
( )
µn = a0 + a1hn−1 + · · · + aphn−p
( )
+ b0 + b1εn−1 + · · · + bq εn−q , σn = const
I-7
1980s : nonlinear models ARCH, GARCH, CRR
#analysis of day data$
Sn = S0 exp{h1 + · · · + hn}
I-8
1990s : (a) Stochastic processes with discrete
#intraday data analysis$ intervention of chance (piecewise
constant trajectories with jumps at
“close” times τ1, τ2, . . .):
#
Ht = hk I(τk ≤ t)
I-9
WEAKNESS of the MODEL St = S0eHt , Ht = (µ−σ 2/2)t+σBt ,
based on a Brownian motion, i.e., dSt = St(µ dt + σ dBt ) with a
constant volatility σ.
σ −→ σ(t)
σ(t) −→ σ(t, St)
2nd CORRECTION
(B. Dupire, 1994)
Pricing with a smile,
RISK, 7, 18–20
I-11
One- and two-dimensional distributions of H = (Ht )t≥0.
(∆)
The observable properties of ht = log(St/St−∆)
shows that for small n∆ the value ρ(n∆) is negative, while most
of the values of ρ(n∆) are close to zero (noncorrelatedness).
C. Analogous estimators for autocorrelation of absolute values
(∆) (∆)
|ht | and |ht+n∆| show that for small n∆ the autocorrelation
is positive (clustering effect).
I-12
HERE THE PICTURE
I-13
Searching for adequate statistical models which describe
dynamics of the prices S = (St)t≥0 led to
LÉVY PROCESSES.
Now these processes take the central place in modelling the
prices of financial indexes, the latter displaying the jump
character of changes.
I-14
MAIN MODELS
based on a Brownian motion
Exponential
Brownian model
St = S0 exp{µt + σBt}
↓ ↓
Exponential
Exponential
INTEGRAL
TIME-CHANGED
Brownian model:
45 t 5 t 6 Brownian model:
7 8
St = S0 exp µs ds + σs dBs St = S0 exp µT (t) + BT (t)
0 0
I-15
Assuming that µ = 0, one can rewrite these models in a brief form:
S = S0eσB
↓ ↓
S = S0eσ·B S = S0eB◦T
5 ·
where • σ · B is the stochastic integral ( σs dBs ),
0
• B ◦ T is a time change in Brownian motion (BT (t) ).
I-16
A generalization of these “Brownian” models, which have being
predominating in financial modelling for a long time, is based on
the idea to replace
S = S0eσL
↓ ↓
S = S0eσ·L S = S0eL◦T
I-17
LÉVY PROCESS L = (Lt )t≥0 is a process with stationary
increments, L0 = 0, which is continuous in probability.
• Brownian motion,
• Poisson process,
Nt
#
• compound Poisson process Lt = ξk , where
k=1
(Nt )t≥0 is a Poisson process,
(ξk )k≥1 is a sequence of independent and identically
distributed random variables
I-20
In connection with financial econometrics, these are
I-21
For a Lévy process (Ht)t≥0 we have
h = µ + βσ 2 + σε,
where ε is a standard Gaussian random variable, ε ∼ N (0, 1),
σ = σ(ω) is the “volatility” (which does not depend on ε), for
whose square, σ 2, we will construct the special distribution
I-22
Strikingly, this distribution (on R+ ) is infinitely divisible and the
distribution of h = µ + βσ 2 + σε (on R) is infinitely divisible as well.
Hence there exist Lévy processes T = (T (t))t≥0 and H ∗ = (Ht∗)t≥0
such that
Ht = µt + βT (t) + BT (t) ,
where the “time change” T = (T (t))t≥0 and the Brownian motion
B = (Bt )t≥0 are independent.
I-23
Let discuss the details of construction of GIG-distributions for σ 2.
I-24
The formula for the density pT A(B) (s) = dP(T A (B) ≤ s)/ds is well
known:
B 1 2
pT A (B)(s) = ϕs(B − As), где ϕs(x) = √ e−x /(2s). (1)
s 2πs
Herefrom we find the Laplace transform:
7 < 8
−λT A (B) 2
Ee = exp AB(1 − 1 + 2λ/A ) .
I-25
Next important step: one define ad hoc the function
I-26
The distribution on R+ with density
I-27
IMPORTANT PROPERTIES of GIG-DISTRIBUTION (for σ 2):
I-28
Particularly important SPECIAL CASES of GIG-distributions:
I-29
Since GIG-distribution is infinitely divisible, if one take it as the
distribution of squared volatility σ 2,
Law(σ 2) = GIG,
then one can construct a nonnegative nonincreasing Lévy process
T = (T (t))t≥0 (a subordinator) such that
I-30
Form the variable h = µ+βσ 2 +σε, where Law(σ 2) = GIG, Law(ε) =
N (0, 1), σ 2 and ε are independent. It is clear that
Law(h) = Eσ 2 N (µ + βσ 2, σ 2)
is a mixture of normal distributions, i.e., the density ph(x) of h is of
the form
5 ∞ 9 :
1 (x − (µ + βy))2
ph(x) = √ exp − pGIG (y) dy.
0 2πy 2y
Integrating and denoting ph(x) by p∗(x; a, b, µ, β, ν), we find that
! < "
∗
Kν−1/2 α b + (x − µ)2
p (x; a, b, µ, β, ν) = c3(a, b, β, ν) !< "1/2−ν eβ(x−µ) ,
b + (x − µ) 2
< 1
(a/b)ν/2 α 2 −ν
where α = a + β 2 and c3(a, b, β, ν) = √ √ .
2π Kν ( ab)
I-31
The obtained distribution Law(h) with density p∗(x; a, b, µ, β, ν) bears
the name
I-32
SOME PROPERTIES of GH-distribution (for h):
I-33
THREE important SPECIAL CASES of GH-distributions:
I-34
II*. a > 0, b = 0, ν > 0: in this case
VG-distribution
(notation: VG [Variance Gamma]). Density:
∗ aν ν−1/2
p (x; a, 0, µ, β, ν) = √ |x − µ|
πΓ(ν)(2α)ν−1/2 .
× Kν−1/2(α|x − µ|) eβ(x−µ)
I-36
Having GIG-distributions for σ 2 and GH-distributions for h, we turn
to construction of the Lévy process H = (Ht )t≥0 used in representation
of the prices St = S0eHt , t ≥ 0.
TWO POSSIBILITIES
↓ ↓
the fact that h has infinitely using the already constructed
divisible distribution allows process T = (Tt)t≥0, one forms the
one to construct, using the process H = (Ht )t≥0:
general theory, the Lévy
∗ ∗ Ht = µt + βT (t) + BT (t) ,
process H = (Ht )t≥0 such
that where Brownian motion B and
∗
process T are taken to be
Law(H1) = Law(h)
independent.
I-37
In the cases I*, II*, and III* mentioned above the corresponding
Lévy processes have the special names:
L(N ◦ IG)-process,
L(N ◦ H+ )- or L(H)-process,
L(N ◦ Gamma)- or L(VG)-process.
It is interesting to note that L(N ◦ IG)- and L(N ◦ Gamma)-processes
have an important property:
I-38
CONCLUDING NOTES.
I-39
These densities are determined by FOUR parameters (a, b0, b1, b2).
infinite divisibility
(this is not the case for distributions from the Pearson system),
which enables us to construct processes H = (Ht )t≥0 which describe
adequately the time dynamics of logarithmic return of the prices
S = (St)t≥0.
I-40
TOPIC II. Technical Analysis. I
1. Kagi and Renko charts
2. Prediction of time of maximum value of prices
observable on time interval [0, T ]
3. Quickest detection of time of appearing of
arbitrage
4. Drawdowns as the characteristics of risk
II-1
The main motivation of this lecture is based on idea to obtain a
mathematical explanation of some practical methods (“when to buy,
when to sell”, etc.) of the Technical Analysis which have as usual
only a descriptive character.
II-2
As to the “technicians” they concentrate on the local peculiarities
of the markets, they emphasize “mass behavior”, “market moods”;
they start their analysis from an idea that stock price movement
is “the product of supply and demand”; their basic concept is the
following: the analysis of past stock prices helps us to see future
prices because past prices take future prices into account; they try
to explain
II-3
1. Kagi and Renko charts
The Japanese “Kagi chart” and “Renko chart” (also called the price
ranges) give methods to forecast price trends from price changes
which exceed either a certain range H or a certain rate H. The price
range or rate H is determined in advance. (In Japan, popular price
ranges are "5, 10, 20, 50, 100, 200.) Greater price ranges are used
for stocks with higher prices because their upward and downward
movements are larger.
II-4
R → |X|, K → max X − X
RENKO construction: Step I: We construct (ρ/i ):
ρ/0 = 0,
7 8
ρ/n+1 = inf t > ρ/n : |Xt − Xρ/n | = H , n ≥ 1.
!
X ρ/7,,
-
4H !
2H !
(
)
! ! !
&% ρ/3 (( #
$
# ρ/9
% ρ/1 ρ/5
H ! !
#
$ +
*
#
ρ/0 ρ/4 *
t"
II-5
Step II: Construction (ρ/n) −→ (ρm, ρ∗m).
We look at all ρ/n such that
! "! "
Xρ/n − Xρ/n−1 Xρ/n−1 − Xρ/n−2 < 0.
ρm is a Markov time
X ρ∗m−1 ρ∗m is a non-Markov time
ρm
ρm−1
ρ∗m
t
II-6
4 6
KAGI construction: κ0 = inf u > 0 : max X − min X = H
[0,u] [0,u]
4 6
inf u ∈ [0, κ0] : Xu = min X
if Xκ0 = max X
[0,κ0] [0,κ0]
κ0∗ = 4 6
inf u ∈ [0, κ0] : Xu = max X
if Xκ0 = min X
[0,κ0] [0,κ0]
X X
H H
t t
κ0∗ κ0 κ0∗ κ0
II-7
Next step: we define by induction
4 6
inf u > κn : max X − Xu = H if Xκn − Xκn∗ = H
[xn,u]
κn+1 = 4 6
inf u > κn : max X − Xu = H
if Xκn − Xκn∗ = −H
[xn,u]
4 6
inf u ∈ [κn, κn+1] : Xu = max X if Xκn − Xκn∗ = H
[κn,κn+1]
∗
κn+1 = 4 6
inf u ∈ [κn, κn+1] : Xu = [κ min if Xκn − Xκn∗ = −H
X
n,κn+1]
II-8
Kagi and Renko variation (on [0, T ]):
N
#
KT (X; H) = |Xκn∗ −Xκ ∗ |, N = NT (X; H),
n−1
n=1
#M
RT (X; H) = |Xρ∗n −Xρ∗ |, M = MT (X; H).
n−1
n=1
KT (X; H)
kT (X; H) = ,
MT (X; H)
R (X; H)
rT (X; H) = T .
MT (X; H)
II-9
THEOREM. If X = σB, then
P
1) kT (σB; H) ∼ 2H, 2) rT (σB; H) ∼ 2H,
T σ2 T σ2
NT ∼ (P-a.s.), MT ∼ (P-a.s.),
H2 2H 2
2 2
P 2T σ P Tσ
KT = kT NT ∼ 2
; RT = rT MT ∼ .
H H
II-10
Results of the statistical analysis
of some stock prices
Almost the same results are valid for Futures on Index Nasdaq 100
(Emini-Nasdag100 Futures), 1 point = $ 20
II-11
3
2.5
Kagi
2
1.5
Renko
1
0.5
II-13
Define Renko strategy γ R = (γtR )t≥0 with
# ! "
γtR = sgn Xρn−1 − Xρ∗ I[ρn−1,ρ∗ (t)
n−1 n−1 )
n≥1
! "
× I(k(H) ≥ 2) − I(k(H) < 2) , t ≥ 0,
and the corresponding capital
5 t 5 t
γR
Ct = γuR dXu − λ |dγuR |.
0 0
Then
γR
C
lim E t = |r(H) − 2| · H − 2λ.
t→∞ Mt
The similar result is valid for the Kagi strategy γ K = (γtK )t≥0.
II-14
X ρ∗m−1
ρm
ρm−1
ρ∗m
t
II-15
2. Prediction of time of maximum value of prices observable
on time interval [0, T ]
!
We shall describe prices
"
10
by a Brownian motion Bθ 1 0
1 0
1 *# !
0*
B = (Bt)0≤t≤1; θ is a 1 # *%
#* %
1
%
point of maximum of B: ! 1 %
/0 1
/ 0 1
Bθ = max Bt . / 01
0≤t≤1 /
/
..#
#/
.
.
.
τ τ t"
θ 1
II-16
Suppose that we begin to observe this process at time t = 0
(“morning time”), and, using only past observations, we stop at
time τ declaring “alarm” about selling. It is very natural to try to
solve the following problem: to find “optimal” times τ ∗ and τ ∗∗ such
that either
inf E |Bθ − Bτ |2 = E |Bθ − Bτ ∗ |2
0≤τ ≤1
or
inf E |θ − τ | = E |θ − τ ∗∗|.
0≤τ ≤1
For us it was a little bit surprising that here the optimal stopping
times coincide: τ ∗∗ = τ ∗. The solution shows that
4 √ 6
τ ∗ = inf t ≤ 1 : max Bs − Bt ≥ z∗ 1 − t ,
s≤t
where z∗ is a certain (known) constant (z∗ = 1.12 . . .).
II-17
This problem belongs to the theory of optimal stopping and method
of its solution is based on reducing to the special free-boundary
problem.
1
t z √1- t
*
t
τ* 1
γ2
#Here C is
the area of
b2
C C
continuation of
D
b1 observations, D
γ1 is the stopping
area.$
0 u t T
* *
II-19
If µ > 0 and µ is close to 0, then the corresponding picture has the
following form:
C
γ2
C
b2
D
γ1 b1
C
u s 0 T
* *
II-20
For µ < 0 and if µ is far from 0, the picture is as follows:
D
b1
C
γ1
0 T
II-21
In the considered problem, the time θ is a “change point” of the
changing of the directions of trend
Solution of the problem
Bt “ inf E |Bτ − Bθ |2 ”
τ
or the problem “ inf τ E |τ −θ|”
depends, of course, on the
construction at any time t
a “good” prediction of the
change point θ. The natural
θ 1 t estimate of θ should be
based on the a posteriori
probability πt = P(θ ≤ t | FtB ), where FtB = σ(Bs, s ≤ t).
II-22
Stochastic analysis shows that
> ?
St − Bt
πt = 2ϕ √ − 1, St = max Bu ,
1−t u≤t
that explains appearing of the expression
St − Bt
√
1−t
which is involved above in the definition of optimal stopping time
9 :
St − Bt
τ ∗ = inf t ≤ 1 : √ ≥ z∗ .
1−t
Statistics St − Bt is appearing in many problems of the financial
mathematics and financial engineering (and, generally, in the mathematical
statistics under name CUSUM statistics).
Xt = r(t−θ)+ +σBt
or
σ dB , t ≤ θ,
t
dXt =
r dt + σ dBt , t > θ.
θ
Here a “change point” θ is considered as a time of appearing of
arbitrage. (Brownian motion’s prices correspond to the non-arbitrage
situation. Brownian motion with drift corresponds to a case of
arbitrage.)
One very difficult question here is “what is θ?”. There are two
approaches. In the first one we assume that θ is a random variable.
II-24
Suppose that τ is time of “alarm” θ. Consider two events
{τ < θ} and {τ ≥ θ}.
The set {τ < θ} is the event of a false alarm with a (false alarm)
probability P(τ < θ).
From a financial point of view
{τ <θ} an interesting characteristic of
& '$ %
# #
θ # the event {τ ≥ θ} is a delay
$ %& 'τ
τ time E(τ −θ | τ ≥ θ) or E(τ −θ)+.
{τ ≥θ}
II-25
It turned out that it is not a simple problem if we consider an
arbitrary distribution for θ. However, there exists one case when we
may solve this problem in implicit form. This case is the following.
τα∗ = inf{t : πt ≥ Bα
∗
},
where (for case π = 0, for simplicity)
∗
Bα = 1 − α.
Second formulation of the quickest detection of arbitrage assumes
that θ is simply a parameter from [0, ∞). In this case we denote by
Pθ the distribution of the process X under the assumption that a
change point is occurred at time θ.
II-27
By P∞ we denote the distribution of X under assumption that there
is no change point at all. Denote for given T > 0
MT = {τ : E∞ τ ≤ T }
the class of stopping time for which the mean time E∞ τ before
(false) alarm is less or equal to T .
Put also
( )
+
C(T ) = inf sup ess sup Eθ (τ − θ) | Fθ (ω).
τ ∈MT θ ω
We proved that for each T > 0 in the class MT there exists an
optimal strategy with the following structure: declare alarm at time
4 6
τT∗ = inf t : max Xu − Xt ≥ a∗(T ) ,
u≤t
where a∗(T ) is a certain constant. It is interesting to note that (if
r2/(2σ 2) = 1)
C(T ) ∼ log T, T → ∞.
II-28
The given method, based on the
D(T ) ∼ log T, T → ∞.
II-29
4. Drawdowns as the characteristics of risk
II-30
From the theoretical point of view,
Drawdown is a statistical measure of risk for investments; a
competitor to the standard measure of risk such
as return probability, VaR, Sharpe ration, etc.
There are various definitions of drawdown’s characteristics, which
measure the decline in net asset value from the historic high point.
In one financial paper we read that
. . . Measuring risk through extreme losses is a very
appealing idea. This is indeed how financial companies
perceive risks. This explains the popularity of loss
statistics as the maximum drawdown and maximum
loss. . .
and
. . . it does not seem possible to derive exact results for
the expected maximum drawdown.
II-31
Looking forward:
Under the
Commodity Futures Trading Commission’s
(CFTC)
mandatory disclosure regime managed futures advisors are obliged
to disclose, as part of their capsule performance records, their
“ worst peak-to-valley drawdown”.
We shall demonstrate here some our theoretical calculations related
to drawdowns.
II-32
µ
Let Bt = µt + σWt be a Brownian motion with drift, W0 = 0.
There are several interesting characteristics related to
Range,
Drawdowns, !
Downfalls,. . . max Bs
µ
s≤t !
Range: µ
TH
"
µ
Rt = max Bsµ − min Bsµ
s≤t s≤t H
µ
Statistics TH for B µ: 2
4 6
µ µ
TH = inf t ≥ 0 : max Bsµ − Bt ≥ H
s≤t II-33
If µ = 0:
> ?2
0 = H
E TH , E max Bt0 = H,
σ 0
t≤TH
> ?4 > ?
0 2 H 0
−λTH 1 H√
DTH = , Ee = 2λ .
3 σ cosh σ
If µ 4= 0:
- 4 6 .
µ σ2 2µ 2µ
E TH = exp H −1− 2H ,
2µ2 σ2 σ
- 4 6 .
σ 2 2µ
µ
E max µ
Bt = exp 2
H −1 .
t≤TH 2µ σ
II-34
Towards a problem from Kolmogorov’s diary (1944):
...For free (or not) random walk: How Xt drops when Xt falls for
the first time (on (t − ∆, t)) from above to some level ξ? To all
appearance, certainly very steeply!..
(H) (H)
Bt = H + Bt , B0 = H, Bt = Bt0 !
H
(H) (H)
τ = inf{u : Bu = 0}
! "
(H)
F (t) = P τ ≤ t | mins≤T Bs ≤ 0
dF (t)
f (t) =
dt √
H T 2
= √ t−3/2e−H /2t , t ≤ T, 0 !
T
"
2G(H/ T )
5 ∞ τ (H)
−u2 /2
G(x) = e du
x II-35
The following three characteristics of drawdowns are the most important:
Maximum
1 drawdown
−→ Dt = max (Bs − Bs2 )
0≤s≤s2≤t
B
Dt
" II-36
t
Drawdown from Dt = Bσt − min Bs2
high “water σt≤s2≤t
2 mark” to the −→
next low “water = max Bs − min Bs2
mark” 0≤s≤t σt≤s2≤t
II-37
Drawdown from
previous high
3 “water mark” −→ D t = max Bs − Bσ 2
to the lowest 0≤s≤σt2 t
“water mark”
Dt
" II-38
σt2 t
General results on D , D , D for B :
(1) Dt = Dt
(2) Dt = max(D t, Dt)
law
(3) Dt = max |Bs|
s≤t
law
(4) Dt = max |Bs|
gt≤s≤t
where gt = sup{s ≤ t : Bs = 0}.
II-39
Distributional results on D1, D 1 for a standard Brownian motion
B = B ◦:
D ∞ E 4 6
4 # (−1)n π 2(2n+1)2
(5) P(D1≤x) = P max |Bs|≤x = exp −
s≤1 π n=0 2n+1 8x2
<
E D1 = E max |Bs| = π = 1.2533 . . .
s≤1 2
√ <π
E Dt = σ t 2 (for σB ◦ on [0, t])
D E
(6) P(D 1 ≤ x) = P max |Bs| ≤ x = FD (x)
g1≤s≤1 1
E D1 ≤ ≤ E R1
E D1
= 3 =
8 π 8
log 2 ≤ ≤
π 2 π
1.1061 . . . ≤ 1.2533 . . . ≤ 1.5957 . . .
II-41
LEMMA.
law
(1) Dt = Dt
Dt = Dt on {σt ≤ σt2 }
(2) Dt =
max(Dt, Dt) on {σt > σt2 }
R = R1: t = 1, µ = 0, σ = 1
dP(R ≤ x)
W. Feller (1951) got for fR (x) = , x > 0, the following
dx
formula:
∞ k 2 x2
8 # k−1 2 −
fR (x) = √ (−1) k e 2 .
2π k=1
II-42
REMARK. If b(t) = Bt − tB1, t ≤ 1, !is a Brownian" bridge, then for
Kolmogorov’s distribution FK(x) = P sup |b(t)| ≤ x we have
t≤1
∞ √ ∞
# 2π #
−2k2 x2 −(2k−1)2 π 2 /x2
FK(x) = 1 − 2 (−1)k−1e = e
k=1
x k=1
⇓ (θ-function)
∞
# 2 2
fK(x) = 8x (−1)k−1k2e−2k x .
k=1
∞ k 2 x2
8 # k−1 2 −
Since fR (x) = √ (−1) k e 2 , we get
2π k=1
=
21
fR(x) = fK(x),
π x
so
=
8
ER = (= 1.5957691216 . . .)
π
II-43
THEOREM. (t = 1, µ = 0, σ = 1)
law
(a) D1 = max |Bt|
0≤t≤1
<
(c) E D1 = E max |Bt| = π
0≤t≤1 2
√ < π)
(E DT = σ T 2
II-44
Proof. (a): Denote
5t
1
Mt = max Bs, Lt = lim I(|Bs | ≤ ε) ds.
s≤t ε↓0 2ε
0
By Lévy’s theorem
law
(Mt − Bt, Mt; t ≤ 1) = (|Bt|, Lt ; t ≤ 1).
Hence
! "
D1 = max (Bs −Bs2 ) = max max Bs − Bs2
0≤s≤s2≤1 0≤s2≤1 0≤s≤s2
law
= max (Ms2 − Bs2 ) = max |Bt|.
0≤s2≤1 0≤t≤1
II-45
Proof. (c): We give two proofs. Let β = (βt)t≥0 be a Brownian
motion. From self-similarity
law
(βat; t ≥ 0) = (a1/2βt; t ≥ 0).
II-46
The normal distribution property
=
5∞ 2
2 − x2
e 2σ dx = σ
π
0
⇓
=
5∞ x2 s1
1 2
E D = E sup |βt| = E √ = E e− 2 dx
0≤t≤1 s1 π
0
II-47
1
We have E e−λs1 = √ . Hence
cosh 2λ
= =
5 5∞
2 ∞ dx 2 ex dx
ED= =2
π 0 cosh x π e2x + 1
0
= = F∞
5∞ F
2 dy 2 F
=2 2
=2 arctan(x)F
π 1+y π F
1 1
= 3 3
2π π π
=2 = ⇒ ED=
π4 2 2
II-48
3
π
Second proof of the equality E D = is based on the fact that
2
5 1
law 1 du
sup |βt| = ,
t≤1 2 0 (2)
Ru
(2)
where Rs is a Bessel-2:
5 s
(2) 1 du
Rs = βGs + .
2 0 Ru(2)
Thus,
< 3
π
E D = E sup |βt| = E R1(2) = E ξ12 + ξ22 = ,
2
ξ1⊥
⊥ ξ2, ξi ∼ N (0, 1).
II-49
THEOREM. (t = 1, µ = 0, σ = 1, D1 = Bσ1 − minσ1≤s2≤1 Bs2 )
law
(a) D1 = sup |Bs|, where g1 = sup{t ≤ 1 : Bt = 0}.
g1 ≤s≤1
=
∞ 2 2
8 # k−1 − k 2x
(b) fD (x) = (−1) ke , x > 0.
1 π k=1
=
8
(c) E D1 = log 2 (= 1.1061 . . .),
π
E D1 ≤ E D1 ≤ ER
< < <
8 log 2 ≤ π ≤ 8
π 2 π
II-50
Proof: By Lévy’s theorem
> ? > ?
Mt − Bt , Mt, Bt ; law |Bt|, Lt, Lt − |Bt |;
=
t≤1 t≤1
⇓
> ? > ?
Mt − Bt , Mt, Bt ; law |Bt|, Lt, Lt − |Bt |;
=
σ1 ≤ t ≤ 1 g1 ≤ t ≤ 1
7 8
where σ1 = min s ≤ 1 : Bs = max Bu .
u≤1
II-51
Therefore
! "
Bσ1 , max (Mt − Bt − Mt)
σ1≤t≤1
law ! "
= Lg1 − |Bg1 |, max (|Bt| − Lt)
g1≤t≤1
! "
= Lg1 , max |Bt| − Lg1
g1≤t≤1
Finally,
II-52
TOPIC III. Technical Analysis. II
1. Buy and Hold
III-1-1
TOPIC III. 1: BUY & HOLD
PROBLEMS:
D E D E
Sτ Sτ
Buying: (1) inf E U Selling: (1) sup E U
τ ≤T MT τ ≤T MT
D E D E
Sτ Sτ
(2) inf E U (2) sup E U
τ ≤T mT τ ≤T mT
where U = U (x) is a “utility function”.
III-1-2
Interesting cases: U (x) = x, U (x) = log x
For U (x) = x :
D E D E
Sτ M T − Sτ
(1): inf EU ∼ sup EU
τ ≤T MT τ ≤T MT
↑
Maximization of the expected relative
error between the buying price and the
highest possible stock price by choosing
a proper time to buy
III-1-3
For U (x) = log x :
Pτ Pτ ∗
T
E log = E log Pτ − E log MT , E log = E log Pτ ∗ − E log MT .
MT MT T
4D E 6
σ2
Take Pt = exp µ−r− t + σWt . Then
2
0D E 1
σ2
E log Pτ =E µ−r− τ + σWτ ,
2
0D E 1
σ2
sup E log Pτ = sup E µ − r − τ + σWτ
τ ≤T τ ≤T 2
0D E 1
σ2
= sup E µ − r − τ .
τ ≤T 2
0, µ − r − σ 2/2 ≤ 0,
Thus, τT∗ = (∗)
T, µ − r − σ 2/2 > 0.
III-1-4
Problem (1) for U (x) = x is more difficult.
However, it is remarkable that answer is the same:
Pτ Pτ ∗
sup E =E T, where τT∗ is given by (∗).
τ ≤T MT MT
Pτ
The first step of solving the nonstandard problem sup E is its
τ ≤T M T
reduction to the standard problem of the type
V = sup EG(Xτ )
τ ≤T
for some Markov process X and Markov time τ (with respect to
(FtX )t≥0, FtX = σ(Xs , s ≤ t)).
III-1-5
Taking E( · | Fτ ), Fτ = FτS , where FtS = σ(Su, u ≤ t), we find
D F E
Pτ Pτ FF ν − W ν ),
E = EE Fτ = E G(τ, M
MT MT F τ τ
where
III-1-6
So the problem sup E(Pτ /MT ) is reduced to sup EG(τ, Xτ ).
τ ≤T τ ≤T
III-1-8
To find V = V (t, x) and sets C and D we usually consider the Stefan
(free-boundary) problem: to find VG = VG (t, x), C,
G and DH such that
LxVG (t, x) = 0, G
(t, x) ∈ C, VG (t, x) = G(t, x), H
(t, x) ∈ D,
where Lx is the infinitesimal operator of the process X x = (Xtx )t≥0,
x ≥ 0. We know that
Law(Xtx, t ≤ T ) = Law(|Ytx|, t ≤ T ),
where Y x is a “bang-bang” process:
III-1-9
By the Tanaka formula
5 t 5 t
|Ytx| = |x|+ sgn Ysx dYsx +Lt(Y x) = |x|−νt+ sgn Ysx dW
I +L (Y x),
s t
0 0
where Lt(Y x) is the local time of Y x at zero over time interval [0, t].
Since Law(X u ) = Law(|Y x|), from the previous formula we find that
∂f ∂f 1 ∂ 2f
Lxf (t, x) = −ν + 2
, t ∈ (0, T ), x ∈ (0, ∞),
∂t ∂x 2 ∂x
1,2 ∂f
for f ∈ C with reflection condition (t, 0+) = 0.
∂x
III-1-10
If G(t, x) ∈ C 1,2, then, by the Itô formula,
5 s
G(t + s, |Ysx|) = G(t, s) + LxG(t + u, |Yux|) du
0
5 s 5 s
∂G ∂G
+ (t + u, |Yux|) dW
I +
u (t + u, |Yux|) dLu (Y x) .
0 ∂x 0 ∂x
$ %& ' $ %& '
is a martingale, = 0,
where −1 ≤ ∂G
∂x ≤ 0 since ∂f
∂x (t, 0+) = 0
∂G ∂G 1 ∂ 2G
where H(t, x) ≡ LxG(t, x) = −ν + 2
.
∂t ∂x 2 ∂x
III-1-11
From explicit formulae for G = G(t, x) we find that
D E
1 ∂G
H(t, x) = ν − G(t, x) − (t, x).
2 ∂x
If ν ≥ 1/2, then ∂G/∂x ≤ 0 and H(t, x) ≥ 0.
5 τ
From V (t, x) = G(t, x)+ sup Ex H(t+u, Xux) du we conclude that
τ ≤T −t 0
III-1-12
As a result we see that
if ν > 1/2, then the optimal τT∗ equals T ;
if ν = 1/2, then any time τT∗ = t is optimal.
The case −1/2 < ν < 1/2 is more complicated and can be investigated
by the “free-boundary methods”.
III-1-13
So, we have the following result:
III-1-14
POPULAR STOCK MODELS IN FINANCE:
III-1-15
For example, let’s take Gaussian-Inverse Gaussian process
Ht = µt + βT (t) + BT (t) ,
√
where / + √as ≥ bt},
T (t) = inf{s > 0 : Bs
(B/ ) and (B ) are independent Brownian motions.
s s
For case
4 6
St
sup E , St = eHt , MT = exp sup Ht
τ ≤T M T t≤T
we find that
III-1-16
TOPIC III. 2: STOCHASTIC “BUY & HOLD”
(PORTFOLIO REBALANCING)
III-2-1
TOPIC IV. General Theory of Optimal Stopping
Lecture 1. Introduction, pp. 2–46.
∆u = 0, x ∈ C, (∗)
and the boundary condition
IV-1-2
Let
τD = inf{t : Btx ∈ D},
where
Btx = x + Bt
and B = (Bt )t≥0 is a d-dimensional standard Brownian motion.
u(x) = EG(BτxD ), x ∈ C ∪ ∂D
D E
u(x) = ExG(BτD ) .
IV-1-3
The optimal stopping theory operates with the optimization
problems, where
7 8
• we have a set of domains C = C : C ⊆ Rd and
• we want to find the function
IV-1-4
2. The following scheme illustrates the kind of concrete problems
of general interest that will be studied in the courses of lectures:
C. Financial mathematics
stochastic equilibria
IV-1-5
A, B, C
#$ #$
" $
1 4
!" !"
#!
Optimal stopping problems
#$ #$
" $
2 3
!" !"
#!
Free-boundary problems
IV-1-6
3. To get some idea of the character of problems A, B, C that will
be studied, let us begin with the following remarks.
√
B. Davis (1976): E|Bτ | ≤ z1∗ E τ , z1∗ = 1.30693 . . .
IV-1-7
Now our main interest relates with the estimation of the expectations
IV-1-8
The case of max |B| is more difficult. We know that
> ? ∞ > ?
4 # (−1)n π 2(2n + 1)2
P max |Bt| ≤ x = exp − .
t≤T π n=0 2n + 1 8x2
3 ! "
π
E max |Bt| = T = 1.25331 . . . .
t≤T 2
<
(Recall that E|BT | = 2T (= 0.79788 . . .).)
π
IV-1-9
SIMPLE PROOF:
law √
(Bat ; t ≥ 0) = ( aBt; t ≥ 0).
IV-1-10
The normal distribution property:
=
5
2 ∞ − x22
Ee 2a dx = a , a > 0. (∗)
π 0
So,
=
5
1 (∗) 2 ∞ − x2 σ
E sup |Bt| = E √ = Ee 2 dx.
0≤t≤1 σ π 0
IV-1-11
3 3
π π
E sup |Bt| = E sup |Bt| = T
0≤t≤1 2 0≤t≤T 2
E max |Bt| = ?
0≤t≤τ
which is valid
√ for all stopping times τ of B with the best constant
C equal to 2.
We will see that the problem A can be solved in the form (3) by
REFORMULATION to the following optimal stopping problem:
D E
V∗ = sup E max |Bt| − cτ , (4)
τ 0≤t≤τ
where
• the supremum is taken over all stopping times τ of B
satisfying Eτ < ∞, and
• the constant c > 0 is given and fixed.
It constitutes Step 1 in the diagram above.
IV-1-14
If V∗ = V∗(c) can be computed, then from (4) we get
D E
E max |Bt| ≤ V∗(c) + c Eτ (5)
0≤t≤τ
√
This is exactly the inequality (3) above with C = 2.
√
The constant 2 is the best possible in (8).
IV-1-16
In the lectures we consider similar sharp inequalities for other stochastic
processes using ramifications of the method just exposed.
IV-1-17
(B) Classic examples of problems in SEQUENTIAL ANALYSIS:
H0 : µ = µ0 and H1 : µ = µ1 (9)
about the drift parameter µ ∈ R of the observed process
H0 : λ = λ0 and H1 : λ = λ1 (11)
IV-1-18
about the intensity parameter λ > 0 of the observed process
such that the probability errors of the first and second kind satisfy:
IV-1-20
In our lectures we study these as well as closely related problems of
QUICKEST DETECTION.
(The story of creating of the quickest detection problem of randomly
appearing signal, its mathematical formulation, and the route of
solving the problem (1961) are also interesting.)
Two of the prime findings, which also reflect the historical development
of these ideas, are the
IV-1-21
C) One of the best-known specific problems of
MATHEMATICAL FINANCE,
that has a direct connection with optimal stopping problems, is the
problem of determining the
IV-1-23
It turns out that the optimal stopping problem (19):
V∗ = sup Ee−rτ (K − Xτ )+
τ
can be reduced again to a free-boundary problem which can be
solved explicitly. It yields the existence of a constant b∗ such that
the stopping time
τ∗ = inf { t ≥ 0 | Xt ≤ b∗ } (20)
is optimal in (19).
V∗ = sup Ee−rτ (K − Xτ )+
τ
no restriction was imposed on the class of admissible stopping
times, i.e. for certain reasons of simplicity it was assumed there
that
τ belongs to the class of stopping times
IV-1-25
A more realistic requirement on a stopping time in search for the
arbitrage-free price leads to the following optimal stopping problem:
IV-1-26
Its study yields that the stopping time
In our lectures we study these and other similar problems that arise
from various financial interpretations of options.
IV-1-27
5. So far we have only discussed problems A, B, C and their reformulations
as optimal stopping problems. Now we want to address the methods
of solution of optimal stopping problems and their reduction to free-
boundary problems.
or the problem
IV-1-28
In this formulation it is important to realize that
Since the method of solution to the problems (25) and (26) is based
on results from the theory of martingales (Snell’s envelope, 1952),
the method itself is often referred to as the
MARTINGALE METHOD.
IV-1-29
On the other hand, if we are to take a state space (E, B) large
enough, then one obtains the
IV-1-30
Likewise, it is a profound attempt, developed in stages, to study
optimal stopping problems through functions of initial points in the
state space.
In this way we have arrived to the second approach which deals with
the problem
V (x) = sup ExG(Xτ ) (28)
τ
where the supremum is taken over M or MT as above (Dynkin’s
formulation, 1963).
IV-1-31
6. To make the exposed facts more transparent, let us consider the
optimal stopping problem
D E
V∗ = sup E max |Bt| − cτ
τ 0≤t≤τ
in more detail.
Denote
Xt = |x + Bt| (29)
for x ≥ 0, and enable the maximum process to start at any point by
setting for s ≥ x
D E
St = s ∨ max Xr . (30)
0≤r≤t
IV-1-32
D E
St = s ∨ max Xr
0≤r≤t
D E
The value V∗ from (4) above: V∗ = sup E max |Bt| − cτ coincides
τ 0≤t≤τ
with the value function
! "
V∗(x, s) = sup Ex,s Sτ − cτ (31)
τ
when x = s = 0. The problem thus needs to be solved in this more
general form.
IV-1-33
The general theory of optimal stopping for Markov processes makes
it clear that the optimal stopping time in (31) can be written in the
form
τ∗ = inf { t ≥ 0 | (Xt , St) ∈ D∗ } (32)
In other words,
IV-1-34
"
s=x
s Heuristic considerations about the
$
$
D∗ C∗ $ shape of the sets C∗ and D∗
$
$
$ makes it plausible to guess that
$
$
$ there exist a point s∗ ≥ 0 and
$
$
$ a continuous increasing function
% # $
$
%
$
s :→ g∗(s) with g∗(s∗) = 0 such
(xt, st)
$
$ that
$
$
D∗ = { (x, s) ∈ R2 | 0 ≤ x ≤ g∗(s) , s ≥ s∗ } (33)
$
$
$
$ #
x
Note that such a guess about the shape of the set D∗ can be made
using the following intuitive arguments. If the process (X, S) starts
from a point (x, s) with small x and large s, then it is reasonable to
stop immediately because to increase the value s one needs a large
time τ which in the formula (31) appears with a minus sign.
IV-1-35
At the same time it is easy to see that
art of GUESSING
in this context very often plays a crucial role in solving the problem.
IV-1-36
Having guessed that the stopping set D∗ in the optimal stopping
problem V∗(x, s) = supτ Ex,s(Sτ − cτ ) takes the form
D∗ = { (x, s) ∈ R2 | 0 ≤ x ≤ g∗(s) , s ≥ s∗ },
it follows that τ∗ attains the supremum, i.e.,
! "
V∗(x, s) = Ex,s Sτ∗ − cτ∗ for all (x, s) ∈ E. (34)
Consider V∗(x, s) for (x, s) in the continuation set
IV-1-37
Denote by
1 ∂2
LX =
2 ∂x2
the infinitesimal operator of the process X. By the strong Markov
property one finds that V∗ solves
If the process (X, S) starts at a point (x, s) with x < s, then during
a positive time interval the second component S of the process
remains equal to s.
IV-1-39
This analysis indicates that the value function V∗ and the optimal
stopping boundary g∗ can be obtained by searching for the pair of
functions (V, g) solving the following free-boundary problem:
which yield
1
g 2(s) = , for s ≥ s0. (49)
2(s − g(s))
IV-1-41
It is easily verified that the linear function
1
g(s) = s − (50)
2c
solves (49). In this way a candidate for the optimal stopping boundary
g∗ is obtained.
IV-1-42
For other points (x, s) ∈ E when s < 1/2c one can determine V (x, s)
using that the observation must be continued. In particular for x =
s = 0 this yields that
IV-1-43
The key role in the proof of the fact that
V = V∗ and g = g∗
is played by
IV-1-44
7. The important point to be made in this context is that the
verification theorem is usually not difficult to prove in the cases
when a candidate solution to the free-boundary problem is obtained
explicitly.
IV-1-45
8. From the material exposed above it is clear that our basic interest
concerns the case of continuous time.
However, since the former theory uses many basic ideas from the
latter, we have chosen to present the case of discrete time first, both
in the martingale and Markovian setting, which is then likewise
followed by the case of continuous time. The two theories form
several my lectures.
IV-1-46
LECTURE 2–3:
Theory of optimal stopping for discrete time.
A. Martingale approach.
1. Definitions
(Ω, F , (Fn)n≥0, P), F0 ⊆ F1 ⊆ · · · ⊆ Fn ⊆ · · · ⊆ F , G = (Gn)n≥0.
Gain Gn is Fn -measurable
Stopping (Markov ) time τ = τ (ω):
τ : Ω → {0, 1, . . . , ∞}, {τ ≤ n} ∈ Fn for all n ≥ 0.
IV-2/3-1
The optimal stopping problem to be studied seeks to solve
V∗ = sup E Gτ . (53)
τ
In the class MN
n we consider
VnN = sup E Gτ
n≤τ ≤N
Stop at time N
*+ ,,
*
** ,,
,,
&
** ,,
&'(
& ,, * -!
,
&&& ( ,,
, *
&& ( *+
* -*
,
&& ( **
(
( ** #
( **
()*
(
0 1 2 N −2 N −1 N
IV-2/3-3
For n = N − 1 we can either stop or continue. If we stop, our gain
N N
SN −1, equals GN −1, and if we continue our gain SN −1 will be equal
to E(SNN |F
N −1).
0 1 2 N −2 N −1 N
So,
N N
SN −1 = max{GN −1, E(SN | FN −1)}
and optimal stopping time is
N N
τN −1 = min{N − 1 ≤ k ≤ N : S k = Gk }.
IV-2/3-4
Define now a sequence (SnN )0≤n≤N recursively as follows:
SnN = GN , n = N,
SnN = max{Gn, E(Sn+1
N
| Fn)}, n = N −1, . . . , 0.
The first part of the following theorem shows that SnN and τnN solve
the problem in a stochastic sense.
The second part of the theorem shows that this leads also to a
solution of the initial problem
VnN = sup E Gτ for each n = 0, 1, . . . , N.
n≤τ ≤N
IV-2/3-5
Theorem 1. (Finite horizon)
I. For all 0 ≤ n ≤ N we have:
IV-2/3-6
Proof of Theorem 1.
Conditions
and
Suppose that (a) and (b) are satisfied for n = N, N −1, . . . , k, where
k ≥ 1, and let us show that they must then also hold for n = k−1.
IV-2/3-7
! "
(a) SnN ≥ E(Gτ | Fn), ∀τ ∈ MN
n : Take τ ∈ MN
k−1 and set τ̄ = τ ∨ k;
then τ̄ ∈ MN
k , and since {τ ≥ k} ∈ Fk−1 it follows that
IV-2/3-8
Using (56)–(58) in (??) we get
N
E(Gτ | Fk−1) ≤ I(τ = k−1) Sk−1 + I(τ ≥ k) E(SkN | Fk−1)
N + I(τ ≥ k) S N
≤ I(τ = k−1) Sk−1 N (59)
k−1 = Sk−1.
IV-2/3-9
Then we get
( )
N = k − 1) G
E Gτ N | Fk−1 = I(τk−1 k−1
k−1
( )
N ≥ k)
+ I(τk−1 E E(Gτ N | Fk ) | Fk−1
k
N = k − 1) G
= I(τk−1 + N ≥ k) E(S N | F
k−1 I(τk−1 k k−1)
N N N N N
= I(τk−1 = k − 1) Sk−1 + I(τk−1 ≥ k) Sk−1 = Sk−1 .
Thus
! "
SnN = E Gτ N | Fn
n
holds for n = k − 1. (We supposed by induction that (b) holds for
n = N, . . . , k.)
IV-2/3-10
! "
(c) τnN is optimal in VnN = sup E Gτ :
n≤τ ≤N
Take expectation E in SnN ≥ E(Gτ | Fn), τ ∈ Mn
n. Then
IV-2/3-11
So,
E SnN = VnN
and since E SnN = E Gτ N , we see that
n
VnN = E Gτ N
n
IV-2/3-12
It thus follows that
(α) (β)
E Gτ∗ < E SτN∗ ≤ E Sn = VnN ,
N
where
The strict inequality E Gτ∗ < VnN , however, contradicts the fact that
τ∗ is optimal.
Hence SτN∗ = Gτ∗ (P-a.s.) and the fact that τnN ≤ τ∗ (P-a.s.) follows
from the definition
From
SkN ≥ E(Sk+1
N | Fk ).
IV-2/3-15
! "
(f) the stopped sequence (S N ) is a martingale :
k∧τnN n≤k≤N
To verify the martingale property
( )
N N
E S(k+1)∧τ N | Fk = Sk∧τ N
n n
with n ≤ k ≤ N − 1 given and fixed, note that
( ) ( )
E N
S(k+1)∧τ N | Fk = E N N
I(τn ≤ k) Sk∧τ N | Fk
n ( n )
N N
+ E I(τn ≥ k + 1) Sk+1 | Fk
= I(τnN ≤ k)Sk∧τ
N N N
N + I(τn ≥ k + 1) E(Sk+1 | Fk )
n
= I(τnN ≤ k)Sk∧τ
N N N N
N + I(τn ≥ k + 1)Sk = Sk∧τ N
n n
SkN = E(Sk+1
N
| Fk ) on { τnN ≥ k + 1 }
and { τnN ≥ k + 1 } ∈ Fk since τnN is a stopping time.
IV-2/3-16
Summary
V0N = sup E Gτ
τ ∈MN
0
is solved inductively by solving the problems
if the stopping rule τ0N is optimal for V0N and it was not
optimal to stop within the time set {0, 1, . . . , n − 1}, then
starting the observation at time n and being based on the
information Fn, the same stopping rule is still optimal for
the problem VnN .
" 4
$
$$ $
$ $ $$ $$ $ $$ $$ $$ $$ $
$ $ $ $ $ $ $ D
$ $ $
$
$
$ $ $ $ $ $ $ $ $ $$
$ $ $ $ 000
122
222 $
000 3
2$
+
*
**
**
0000
1222
0 000
1222
00 1*
00 C
00 2 300
22 2 300
22
$ $ $ $
$
$ $ $ $ $
$ $ $ $ $
$
$
$
$
$
$
$
$
$
$
$
$
$
$
D
$
$
$
$
$
$
$$
$ $ $ $
$ $ $ $ $ $ $ $ $ $ $$ $ $ $$ $ $ $
$ $ $ $ $ $ $ $ $ $ $ $ $ $ $ $ $ #
0 n−1 n
IV-2/3-18
3. The method of ESSENTIAL SUPREMUM
It turns out, however, that the random variables SnN defined by the
recurrent relations
SnN = GN , n = N,
SnN = max{Gn, E(Sn+1
N
| Fn)}, n = N −1, . . . , 0,
admit a different characterization which can be directly extended to
the case of infinite horizon N .
ESSENTIAL SUPREMUM
proves useful.
IV-2/3-20
Lemma (about Essential Supremum).
Z ∗ = sup Zα
α∈J
satisfies the following two properties:
(a) P(Zα ≤ Z ∗) = 1, ∀α ∈ A;
(b) If Z̃ : Ω → R is another random variable
satisfying P(Zα ≤ Z ∗) = 1, ∀α ∈ A, then
P(Z ∗ ≤ Z̃) = 1.
IV-2/3-21
II. Moreover, if the family {Zα, α ∈ A} is upwards directed in the
sense that
for any α and β in A there exists γ in A
such that max(Zα, Zβ ) ≤ Zγ (P-a.s.),
then the countable set J = {αn, n ≥ 1} can be chosen so that
Z ∗ = lim Zαn (P-a.s.)
n→∞
where Zα1 ≤ Zα2 ≤ · · · (P-a.s.).
IV-2/3-22
def L∞
Then J = n=1 Cn is a countable subset of A and we claim that
def
Z ∗ = sup Zα
α∈J
satisfies the properties (a) and (b).
SnN ≥ E(Gτ | Fn ) ∀τ ∈ MN
n; SnN = E(Gτ N | Fn)
n
in Theorem 53 above as follows:
IV-2/3-24
This ess sup identity provides an additional characterization of the
sequence of r.v.’s (SnN )0≤n≤N introduced initially by means of the
recurrent relations
SnN = GN , n = N,
SnN = max{Gn, E(Sn+1
N
| Fn)}, n = N −1, . . . , 0.
Vn = sup E Gτ .
τ ∈M∞
n
IV-2/3-25
To solve this problem we will consider the sequence of r.v.’s (Sn)n≥0
defined as follows:
Sn = ess sup E(Gτ | Fn)
τ ≥n
as well as the following stopping time:
τn = inf{k ≥ n | Sk = Gk } for n ≥ 0,
where inf ∅ = ∞ by definition.
The first part (I) of the next theorem shows that (Sn)n≥0 satisfies
the same recurrent relations as (SnN )0≤n≤N .
The second part (II) of the theorem shows that Sn and τn solve the
problem in a stochastic sense.
The third part (III) shows that this leads to a solution of the initial
problem Vn = supτ ≥n E Gτ .
The fourth part (IV) provides a supermartingale characterization of
the solution.
IV-2/3-26
Theorem 2 (Infinite horizon).
P(τn < ∞) = 1.
Then for all n ≥ 0 we have:
IV-2/3-27
III. Moreover, if n ≥ 0 is given and fixed, then we have:
IV-2/3-28
Proof. I. We need prove the recurrent relations
IV-2/3-29
From this inequality it follows that
Sn ≥ E(Sn+1 | Fn )
which is the supermartingale property of (Sn)n≥0. To verify this
inequality, let us first show that the family {E(Gτ | Fn+1); τ ∈ Mn+1}
is upwards directed in the sense that
for any α and β in A there exists γ in A
(∗)
such that Zα ∨ Zβ ≤ Zγ .
IV-2/3-30
For this, note that if σ1 and σ2 are from Mn+1 and we set σ3 =
σ1IA + σ2IĀ where
IV-2/3-31
Since
Sn+1 = ess sup E(Gτ | Fn+1),
τ ≥n+1
by the conditional monotone convergence theorem we get
( )
E(Sn+1 | Fn) = E lim E(Gσk | Fn+1) | Fn
k→∞( )
= lim E E(Gσk | Fn+1) | Fn
k→∞
= lim E(Gσk | Fn ) ≤ Sn.
k→∞
So, Sn = max{Gn, E(Sn+1 | Fn)} and the proof if I is complete.
IV-2/3-34
The statement
IV-2/3-35
Remark. From the definition
IV-2/3-36
Note also that from
VnN = sup E Gτ
n≤τ ≤N
From SnN = ess supn≤τ ≤N E(Gτ | Fn ) and Sn = ess supτ ≥n E(Gτ | Fn)
we see that
Sn∞ ≤ Sn and τn∞ ≤ τn. (∗)
Similarly,
! "
Vn∞ ≤ Vn = sup E Gτ . (∗∗)
τ ≥n
If condition E supn≤k<∞ |Gk | < ∞ does not hold then the inequalities
in (∗) and (∗∗) can be strict.
IV-2/3-37
Theorem 3 (From finite to infinite horizon).
Proof. From
(Sn∞)− ≤ G− −
n ≤ sup Gn , n ≥ 0.
n≥0
So, ((Sn∞ )−)n≥0 is uniformly integrable.
IV-2/3-38
Then by the optional sampling theorem we get
IV-2/3-40
It is assumed that the chain X starts at x under Px for x ∈ E.
IV-2/3-41
Given a measurable function G : E → R satisfying the following
condition (with G(XN ) = 0 if N = ∞):
D E
Ex sup |G(Xn)| < ∞
0≤n≤N
for all x ∈ E, we consider the optimal stopping problem
IV-2/3-42
Since the same results remain valid if we take the supremum in
IV-2/3-43
To solve
V N (x) = sup Ex G(Xτ ) (∗)
0≤τ ≤N
when N < ∞, we may note that by setting Gn = G(Xn) for n ≥ 0
the problem reduces to the problem
VnN = sup Ex Gτ . (∗∗)
n≤τ ≤N
T F (x) = Ex F (X1)
for x ∈ E whenever F : E → R is a measurable function so that
F (X1 ) is integrable w.r.t. Px for all x ∈ E.
IV-2/3-45
Theorem 4 (Finite horizon: The time-homogeneous case)
IV-2/3-46
IV. The sequence (V N −n(Xn))0≤n≤N is the smallest
supermartingale which dominates (G(Xn ))0≤n≤N under Px
for x ∈ E given and fixed.
V. The stopped sequence (V N −n(Xn∧τD ))0≤n≤N is a
martingale under Px for every x ∈ E.
IV-2/3-47
Inserting (ii) into (i) and using the Markov property we obtain
( ) ( )
SnN = Ex G(Xn+τ N −n◦θ ) | Fn = Ex G(Xτ N −n ) ◦ θn | Fn
0 n 0
(α)
(iii)
= EXn G(Xτ N −n ) = V N −n(Xn )
0
IV-2/3-48
To verify the “Wald–Bellman equation”, note that the equality
IV-2/3-49
The “Wald–Bellman equation” can be written in a more compact
form as follows. Introduce the operator Q by setting
IV-2/3-50
TIME-INHOMOGENEOUS MARKOV CHAINS X = (Xn )n≥0
We assume
D E
(∗∗) En,x sup |G(n + k, Xn+k )| < ∞, 0 ≤ n ≤ N.
0≤k≤N −n
IV-2/3-51
Theorem 5 (Finite horizon: The time-inhomogeneous case)
Consider the optimal stopping problem (∗) upon assuming that the
condition (∗∗) holds. Then:
IV-2/3-52
II. The stopping time
N
τD = inf{n ≤ k ≤ N : (n + k, Xn+k ) ∈ D}
with
7 8
D = (n, x) ∈ {0, 1, . . . , N }×E : V (n, x) = G(n, x)
is optimal in the problem (∗):
IV-2/3-53
IV. The value function V N is the smallest superharmonic
function which dominates the gain function G on
{0, . . . , N }×E,
The proof is carried out in exactly the same way as the proof of
Theorem 4.
IV-2/3-54
Optimal stopping for infinite horizon (N = ∞):
Theorem 6
τD = inf{t ≥ 0 : Xt ∈ D}
with D = {x ∈ E : V (x) = G(x)}. Then the stopping time τD
is optimal.
IV-2/3-55
III. If τ∗ is an optimal stopping time then τD ≤ τ∗ (Px-a.s. for
every x ∈ E).
IV-2/3-56
Corollary (Iterative method). We have
IV-2/3-57
2. Given α ∈ (0, 1] and bounded g : E → R and c : E → R+ , consider
the optimal stopping problem
D τ
# E
V (x) = sup Ex ατ g(Xτ ) − αk−1c(Xk−1) .
τ
k=1
I = (X
Let X I )
n n≥0 denote the Markov chain X killed at rate α. It
means that
T/ F (x) = α T F (x).
Then
D τ
# E
I )−
V (x) = sup Ex g(X I
τ c(Xk−1) .
τ
k=1
IV-2/3-58
LECTURES 4–5.
Theory of optimal stopping for continuous time
A. Martingale approach
Let (Ω, F , (Ft )t≥0, P) be a stochastic basis (a filtered probability
space with right-continuous family (Ft )t≥0 where each Ft contains
all P-null sets from F .
DEFINITION.
A random variable τ : Ω → [0, ∞] is called a Markov time
if {τ ≤ t} ∈ Ft for all t ≥ 0.
A Markov time is called a stopping time if τ < ∞ P-a.s.
IV-4/5-1
We assume that G = (Gt)t≥0 is right-continuous and left-continuous
over stopping times (if τn ↑ τ then Gτn → Gτ P-a.s.).
VtT = sup E Gτ .
t≤τ ≤T
IV-4/5-2
Two ways to tackle the problem VtT = supt≤τ ≤T E Gτ :
IV-4/5-3
Consider the process S = (St)t≥0 defined as follows:
Introduce
τt = inf {u ≥ t | Su = Gu} where inf ∅ = ∞ by definition.
However,
St = max{Gt, E (Sσ∧τt | Ft)}
for every stopping time σ ≥ t and τt given above.
IV-4/5-4
Theorem 1. Consider the optimal stopping problem
Vt = sup E Gτ , t ≥ 0,
τ ≥t
upon assuming E supt≥0 |Gt| < ∞. Assume moreover when required
below that
P(τt < ∞) = 1, t ≥ 0.
(Note that this condition is automatically satisfied when the horizon
T is finite.) Then:
IV-4/5-5
II. The stopping time τt = inf{u ≥ t : Su = Gu} is
optimal (for the problem Vt = supτ ≥t E Gτ ).
III. If τt∗ is an optimal stopping time as well then
τt ≤ τt∗ P-a.s.
IV. The process (Su)u≥t is the smallest right-
continuous supermartingale which dominates
(Gs)s≥t.
V. The stopped process (Su∧τt )u≥t is a right-
continuous martingale.
VI. If the condition P(τt < ∞) = 1 fails so that
P(τt = ∞) > 0, then there is no optimal stopping
time.
IV-4/5-6
Proof. 1◦. Let us first prove that S = (St)t≥0 defined by
E(Gσ3 | Ft) = E (Gσ1 IA + Gσ2 IĀ | Ft) = IA E (Gσ1 | Ft) + IĀ E (Gσ2 | Ft)
= E (Gσ1 | Ft ) ∨ E (Gσ2 | Ft ).
IV-4/5-7
Hence there exists a sequence {σk ; k ≥ 1} in Mt such that
IV-4/5-8
Note that from E supt≥0 |Gt| < ∞ and
E St = sup E Gτ .
τ ≥t
t % E St is right-continuous on R+ .
IV-4/5-9
By the supermartingale property of S
E S t ≥ · · · ≥ E S t2 ≥ E S t1 , tn ↑ t.
So, L := limn→∞ E Stn exists and
E St ≥ L.
To prove the reverse inequality, fix ε > 0 and by means of E St =
supτ ≥t E Gτ choose σ ∈ Mt such that
E Gσ ≥ E St − ε.
Fix δ > 0 and note that there is no restriction to assume that
tn ∈ [t, t + δ] for all n ≥ 1. Define
σ if σ > tn,
σn =
t+σ if σ ≤ tn.
IV-4/5-10
Then for all n ≥ 1 we have
The inequality (∗) follows from the definition St = ess supτ ≥t E(Gτ | Ft).
The proof of (∗∗) is the most difficult part of the proof of the
Theorem.
IV-4/5-13
Assume that Gt ≥ 0 for all t ≥ 0.
St ≤ E (Gτ 1 | Ft)
t
IV-4/5-14
For the proof of property V:
E Sσ∧τt = E St
for all bounded stopping times σ ≥ t.
IV-4/5-16
We consider the optimal stopping problem
IV-4/5-17
V (x) = sup Ex G(Xτ )
CASE T = ∞: τ
Px(X0 = x) = 1
Introduce
NOTICE! If
V is lsc (lower semicontinuous) G is usc (upper semicontinuous)
$
& %
&
% &
then
C is open and D is closed
IV-4/5-18
The first entry time
τD = inf{t ≥ 0 : Xt ∈ D}
for closed D is a stopping time since both X and (Ft )t≥0 are right-
continuous.
Ex F (Xσ ) ≤ F (x)
for all stopping times σ and all x ∈ E. (It is assumed that F (Xσ ) ∈
L1(Px) for all x ∈ E whenever σ is a stopping time.)
We have:
(F (Xt ))t≥0 is a supermartingale
F is superharmonic iff
under Px for every x ∈ E.
IV-4/5-19
The following theorem presents
NECESSARY CONDITIONS
for the existence of an optimal stopping time.
IV-4/5-20
Let us in addition to “ V (x) = Ex F (Xτ∗ )” assume that
τD ≤ τ∗ (Px -a.s., x ∈ E )
and is optimal;
(III) The stopped process (V (Xt∧τD ))t≥0 is a right-continuous
martingale under Px for every x ∈ E.
IV-4/5-21
Now we formulate
SUFFICIENT CONDITIONS
for the existence of an optimal stopping time.
IV-4/5-22
Let us assume that there exists the smallest superharmonic function
VG which dominates the gain function G on E.
We then have:
IV-4/5-23
Corollary (The existence of an optimal stopping time).
Finite horizon (T < ∞). Suppose that V is lsc and G is usc. Then
τD is optimal.
IV-4/5-24
If so, then evidently V is the smallest superharmonic function
which dominates G on E. Then the claims of the corollary follow
directly from the Theorem (on sufficient conditions) above.
IV-4/5-26
By the monotone convergence theorem using E supt≥0 |Gt| < ∞ we
can conclude
Ex V (Xσ ) = lim
n
Ex G(Xσ+τn◦θσ ) ≤ V (x)
for all stopping times σ and all x ∈ E. This proves that V is
superharmonic.
x :→ Ex G(Xτ )
is continuous (or lsc) for every stopping time τ , then x :→ V (x) is lsc
and the results of the Corollary are applicable. This yields a powerful
existence result by simple means.
IV-4/5-27
REMARK 2. The above results have shown that the optimal stopping
problem
V (x) = sup Ex G(Xτ )
τ
is equivalent to the problem of finding the smallest superharmonic
function VG which dominates G on E. Once VG is found it follows
that V = VG and τD = inf{t : G(Xt) = VG (Xt )} is optimal.
IV-4/5-28
For (i), e.g., it is known that if G is lsc and
and QN
n is the N -th power of Qn .
IV-4/5-29
The basic idea (ii) is that
VG and C (or D)
should solve the free-boundary problem:
(∗) LX VG ≤ 0
(∗∗) VG ≥ G (VG > G on C & VG = G on D)
where LX is the characteristic (infinitesimal) operator of X.
IV-4/5-30
If X after starting at ∂C enters immediately into int(D) (e.g. when X
is a diffusion process and ∂C is sufficiently nice) then the condition
LX VG ≤ 0 under (∗∗) splits into the two conditions:
LX VG = 0 in C
F F
∂ VG FF ∂G FF
F = F (smooth fit).
∂x ∂C ∂x ∂C
LX VG = 0 in C
F F
G F F
VF = GF (continuous fit).
∂C ∂C
IV-4/5-31
Proof of the Theorem on NECESSARY conditions
Basic lines
Thus V is superharmonic.
IV-4/5-32
Let F be a superharmonic function which dominates G on E. Then
V (x) = Ex G(Xτ∗ ), x ∈ E.
IV-4/5-33
We claim that V (Xτ∗ ) = G(Xτ∗ ) Px-a.s. for all x ∈ E.
IV-4/5-34
By (I) the value function V is the superharmonic (Ex V (Xσ ) ≤ V (X)
for all stopping time σ and x ∈ E). Setting σ ≡ s and using the
Markov property we get for all t, s ≥ 0 and all x ∈ E
IV-4/5-35
In particular, since τD ≤ τ∗ we get
IV-4/5-36
If V is only lsc, then again (see the lemma below) the process
(V (Xt))t≥0 is right-continuous (Px-a.s. for each x ∈ E), and the
proof can be completed as above.
IV-4/5-37
(III) The stopped process (V (Xt∧τD ))t≥0 is a right-continuous
martingale under Px for every x ∈ E.
IV-4/5-38
REMARK. The result and proof of the Theorem extend in exactly
the same form (by slightly changing the notation only) to the finite
horizon problem
VT (X) = sup Ex G(Xτ ).
0≤τ ≤T
sufficient condition
for the existence of an optimal stopping time.
IV-4/5-39
THEOREM. Consider the optimal stopping problem
We then have:
VG (x) ≤ V (x)
and optimality of time τD .
IV-4/5-41
(II) If Px(τD < ∞) < 1 for some x ∈ E then there is no optimal
stopping time.
IV-4/5-42