ArticlePDF Available

On the Evaluation of First-Passage-Time Probability Densities via Non-Singular Integral Equations

Authors:

Abstract

The algorithm given by Buonocore et al. [1] to evaluate first-passage-time p.d.f.'s for Wiener and Ornstein-Uhlenbeck processes through a time-dependent boundary is extended to a wide class of time-homogeneous one-dimensional diffusion processes. Several examples are thoroughly discussed along with some computational results.
On the Evaluation of First-Passage-Time Probability Densities via Non-Singular Integral
Equations
Author(s): V. Giorno, A. G. Nobile, L. M. Ricciardi, S. Sato
Source:
Advances in Applied Probability,
Vol. 21, No. 1 (Mar., 1989), pp. 20-36
Published by: Applied Probability Trust
Stable URL: http://www.jstor.org/stable/1427196
Accessed: 19/04/2010 09:53
Your use of the JSTOR archive indicates your acceptance of JSTOR's Terms and Conditions of Use, available at
http://www.jstor.org/page/info/about/policies/terms.jsp. JSTOR's Terms and Conditions of Use provides, in part, that unless
you have obtained prior permission, you may not download an entire issue of a journal or multiple copies of articles, and you
may use content in the JSTOR archive only for your personal, non-commercial use.
Please contact the publisher regarding any further use of this work. Publisher contact information may be obtained at
http://www.jstor.org/action/showPublisher?publisherCode=apt.
Each copy of any part of a JSTOR transmission must contain the same copyright notice that appears on the screen or printed
page of such transmission.
JSTOR is a not-for-profit service that helps scholars, researchers, and students discover, use, and build upon a wide range of
content in a trusted digital archive. We use information technology and tools to increase productivity and facilitate new forms
of scholarship. For more information about JSTOR, please contact support@jstor.org.
Applied Probability Trust is collaborating with JSTOR to digitize, preserve and extend access to Advances in
Applied Probability.
http://www.jstor.org
Adv. Appl. Prob. 21, 20-36 (1989)
Printed
in N. Ireland
? Applied Probability
Trust
1989
ON THE EVALUATION OF FIRST-PASSAGE-TIME
PROBABILITY DENSITIES VIA NON-SINGULAR
INTEGRAL EQUATIONS
V. GIORNO,
*
University of Salerno
A. G. NOBILE,
*
University of Salerno
L. M. RICCIARDI,
**University of Naples
S. SATO, ***Osaka
University
Abstract
The algorithm
given by Buonocore
et al. [1] to evaluate
first-passage-time
p.d.f.'s
for Wiener
and Ornstein-Uhlenbeck
processes
through
a time-dependent
boundary
is extended to a wide class of time-homogeneous
one-dimensional
diffusion
proc-
esses. Several examples are thoroughly
discussed
along with some computational
results.
DIFFUSION PROCESSES; VARYING BOUNDARIES; RESTRICTED PROCESSES
1. Introduction
In a previous
paper (Buonocore et al. [1]) a new second-kind Volterra integral
equation was proposed to determine first-passage-time
p.d.f.'s through time-
dependent boundaries
for time-homogeneous
one-dimensional diffusion processes
with natural
boundaries.
This equation
was seen to be particularly manageable
for
computational
purposes
in the cases of Wiener
and Ornstein-Uhlenbeck
processes.
Indeed, for these the kernel of the integral
equation
can be made continuous,
thus
overcoming
some crucial difficulties
arising
from the singular
nature
of the integral
equations
for the first-passage-time p.d.f. Such continuity
was made possible by a
suitable
choice of two arbitrary
time-dependent
functions that appear
in the kernel
of the integral
equations.
Here, in view of the relevance of first-passage-time
problems
in numerous
applied
fields (see, for instance, Buonocore et al. [1] and references
therein), we intend to
provide
an extension
of the results
of Buonocore et al. [1] in two directions: on the
one hand, we shall prove that the method for regularizing
the kernel of the integral
equation
is valid not only for the Wiener
and the Ornstein-Uhlenbeck
processes
but
can also be used for a more general
class of diffusion
processes
whose free transition
Received
16 September
1987;
revision received
20 January
1988.
* Postal
address:
Dipartimento
di Informatica e Applicazioni,
University
of Salerno,
84100
Salerno,
Italy.
**
Postal
address:
Dipartimento
di Matematica
e Applicazioni,
Universith
di Napoli, Via Mezzocan-
none 8, 80134
Napoli, Italy.
***
Postal address:
Department
of Biophysical
Engineering, Faculty
of Engineering
Science, Osaka
University,
Toyonaka
(Osaka), Japan.
20
Evaluation
offirst-passage-time
probability
densities 21
p.d.f.'s are known; on the other hand, we shall see that the method can also be
employed to make the kernel continuous
if the diffusion
process is restricted
by a
reflecting
boundary
and the transition
p.d.f. in the
presence
of such
boundary
is known.
In the remaining part of this section we shall briefly review some background
results
from Buonocore et al. [1] and then see how they can be adapted to more
general instances. Some straightforward
intuitive considerations are then made in
Section 2 to indicate why the method of Buonocore et al. [1] can be extended
beyond the cases of Wiener and Ornstein-Uhlenbeck
processes. Rigorous proofs
will instead
be given in Sections 3 and 4 by making
heavy use of some fundamental
results
due to Fortet [4]. A few examples
are discussed in Sections 5 and 6.
Let S(t) and k(t) be continuous functions in [to, +oo), toE R, and let {X(t);
to
5 t _ T < +oo} be a time-homogeneous one-dimensional diffusion process with drift
A
1(x)
and infinitesimal variance
A2(x).
Let us denote
the
diffusion
interval
by
I = (rl, r2),
with rl and r2
natural boundaries. Then [1]:
g[S(t), tl xo, to]
=
(1.1) -2W[S(t), t xo,,
to]
+ 2f drg[S(r), r xo, to][S(t), tI
lS(r), r], xo<S(to)
to
2W[S(t),
t xo,
to]
- 2 drg[S(r),
r Ixo,
t
So][),S(
t
I
(Sr),
,], xo>
S(to)
where g[S(t), t I
xo, to] is the first-passage-time
p.d.f. of X(t) through S(t) condi-
tional upon X(to)
= xo and
(1.2) P[S(t), t I
z, -] = d F[S(t), t I
z, -] + k(t)f[S(t), t I
z, -].
(Throughout
this paper we shall take the function r(t) in Buonocore et al. [1] as
identically
zero. Indeed, the case r(t)# 0 does not play any role in what follows.)
Here
(1.3) F(x, t
lz, -) = P{X(t)5x I
X(-r)= z}
and
(1.4) f (x,
t z,
I ) - F(x,
tIz, r)
ax
denote the transition distribution
function
and the transition
p.d.f. of X(t). We have
the following
result.
Lemma
1.1. Let S(t) e Cl[to, T]. Then:
P[S(t), tI z, r] = (S'(t) - A1[S(t)] + JA'[S(t)] + k(t)}f[S(t), t I
z, r]
(1.5) a
+ ? A2[S(t)] -f(x, tI z, )
ax x=S(t)
22 V. GIORNO,
A. G. NOBILE, L. M. RICCIARDI
AND S. SATO
d d
where
we have set S'(t) = S(t) and
A2[S(t)] = A2(x)
dt dx x =S(t)
Proof. From (1.2) we have
d rs(t)
TY[S(t),
t I
z, -] = -
f
dxf(x, t I
z, r)
+ k(t)f[S(t), t I
z, -]
(1.6) = [S'(t) + k(t)]f[S(t), t I
z, -r]
+ dx- f(x, t I
z, -r).
The integral
on the right-hand
side can be calculated
by recalling
that
f satisfies
the
Fokker-Planck
equation
af
_ 1 a2
(1.7) af a [Al(x)f]
+ [A2(x)f]
(t 7x 2
aX2
and that no probability
mass flows at the natural
boundary
rl.
We recall that the arbitrariness of k(t) has been used in Buonocore et al. [1] to
make the kernel
of (1.1) continuous at r = t.
Let us now denote by r an arbitrarily
fixed state with rl
- r < r2
and assume
that r2
is a natural
boundary.
Further,
let us denote by f(r)(x, t z, r), (x ? r), the transition
p.d.f. of X(t) in the presence
of a zero-flux
condition
at r, by F(r)(x,
t I
z, -), (x - r),
the distribution
function and by g(r)[S(t),
tI xo, to] the first-passage-time
p.d.f. of
X(t) through S(t) in the presence of the zero-flux
condition at r. Let us define the
function
(1.8) p(r)[S(t), t I
z, -] =d F'r)[S(t), t I
z, r] + k(t)f(r)[S(t), t I
z, r].
We have the following
result.
Lemma 1.2. Let S(t) e Cl[to, T]. Then:
TP(r)[S(t),
t I
z, ] = {S'(t) - A1[S(t)]
+ JA'[S(t)] + k(t)}f(r)[S(t), t I z, T]
(1.9) + 10W a
+ 1A2[S(t)] f(r)(x, tI z, 0)
ax x=S(t)
Proof. It goes as for Lemma 1.1, the difference being that now a zero-flux
condition
at r must hold.
It is convenient
to consider
separately
the cases X(to)
x- o
> S(to) and xo
< S(to).
Lemma
1.3. Let xo
> S(to). Consider the following
two possibilities:
(a) rl is natural
(hence r > rl). Then:
(1.10) g'(r)[S(t),
t xo, to]- g[S(t), t l
xo, to], xo > S(to)
(so that g(r) can be obtained as solution of the second of Equations (1.1)).
(b) rl is regular or entrance, so that r _ rl. Then g(r)[S(t), t xo, to] is the solution
Evaluation
of first-passage-time
probability
densities 23
of the following equation:
g(r)[S(t), t I
xo, to]
= 2W(r)[S(t), t I
xo, to]
- 2 drg('[S(r),
Ixo,
toP(r)[S(t),
t S(r),T, xo
>
S(to)
where W(r)
is obtained from (1.2) by substituting
F and f with the corresponding
functions obtained
in the presence
of a zero-flux
condition
at r.
Lemma
1.4. Let xo
< S(to). Then g(r')
is the solution of the following
equation:
g(r)[s(t), t I
xo, to]
= -2T(r)[S(t), t xo, tol
(1.12)+ 2 drrg(r)[S(r), ,
irtolt(r)[s(t), tI S(r), r], x < S(to)
where the notation
used has the same meaning
as in (b).
The proofs of Lemmas
1.3 and 1.4 goes along similar
lines as in Buonocore
et al.
[1]. Note that Equations
(1.11) and (1.12) exhibit a weakly singular
kernel at r = t.
However, this kernel may be made continuous at t after a suitable choice of the
arbitrary
function
k(t) appearing
in (1.8) is made and once the boundary
S(t) has
been specified. The algorithm of Buonocore et al. [1] can then be applied to
evaluate numerically
the function
g(r)[S(t), t Ixo, to]. Some examples
will be briefly
discussed
in Sections
5 and 6.
2. Determination of the function k(t)
Here we shall outline some straightforward
intuitive
considerations
indicating
that
a determination
of the function k(t) in (1.2) can be achieved for processes other
than Wiener and Ornstein-Uhlenbeck.
Let Al(x) and A2(x) denote the drift and infinitesimal variance of a one-
dimensional time-homogeneous
diffusion process X(t) defined over the interval
I= (rl, r2), with rl and r2 natural boundaries. We may then think of X(t) as
generated
by the stochastic
differential
equation
(2.1) dX(t) = AI(x) dt + \FA2(x)
dB(t),
where B(t) denotes normalized
Brownian motion. To obtain the transition
p.d.f.
fx(x, t I
xo, to) of X(t) conditional
upon X(to) = xo, we argue as follows. We denote
by x' an arbitrarily
fixed point such that r1
< x' < r2 and set
x dO
(2.2) y h(x)=V 2()
so that Y(t)=h[X(t)] is also a diffusion process. We shall further assume that
transformation (2.2) maps I onto (-o, +oo). The process Y(t) satisfies (Ito's
24 V. GIORNO, A. G. NOBILE, L. M. RICCIARDI AND S. SATO
Lemma)
(2.3) dY(t) =
[A(x))
A (x) dt + V dB(t).
A2(x) 4 x=h-'(y)
The increment AY of Y(t) in [-, t] conditional on Y(r) = w h(z) is then
(2.4) AY(t) Y(t) - w A (z) - A
z=h-(w)
Since AB(t) is a zero-mean
stationary
normal
process
with variance
t - r, from (2.4)
we see that, for small t - -, Y(t) has the p.d.f.
1
fy(y, tI w, -)= 2
1
(2.5) S1 y
1 -w Al(z) - A'(z)/4
ep 2(t - -) V2(
_
Jz=h-'(w)
Expressing
y and
w in terms
of x and
z via (2.2), from
(2.5)
we obtain
(2.6) fx(x, tIZ,0
z() =
) fy[h(x), tl h(z), z]
A2(x)
or, for small t- r,
1
fx(x, t z, ) = r(t- 'r)A2(x)
(2.7)
Xexp{-2(t - r)A2(z) z - (A)
Let us now assume
that
S(t) is a smooth function.
Then,
making
use of Lemma 1.1
and
of the approximation
af(x, t I
z,
,) x - z - [Al(z) - A'(z)/4](t - r)
ax (t - r)A2(z)
(2.8) A'(x)
xf(x, tI z,)- f(x, tl z,
t)
2A2(x)
we obtain
),t A[S(t)] 1 S(t)- z A2[S(t)]
W[S(t),
tI z, r]
-S'(t) - A[S(t)]
+ 2 t - A2(z)
(2.9)
+ A(z) A[S(t)] +k(t)f[S(t), t I
z, T].
Evaluation
of first-passage-time
probability
densities 25
Hence
uA'[S(t)] 1 S(t)- S(r) AE[S(t)] ([S(t), t I
S(r), C]
- S'(t) -
AI[S(t)] +A~[S(t)] 1 S(t) - Sr) A2[S(t)]
(2.10) 1 fA[r Ai[S(r)] A2[S(t)
+24A AS() A2[S(r)] + it)f[S(t), t S(r), 4.
From
(2.10) we thus see that if
1 Ar[S(t)]
(2.11) k(t) = Al[S(t)] - S'(t)- AS(t)4
2 4
then
(2.12) lim [S(t), t I
S(i), r] = 0.
rTt
Indeed,
f has a singularity
(t - -r)- whereas the expression
in curly
brackets
on the
right-hand
side of (2.10) goes to 0 of order t - r due to the assumed
smoothness of
S(t) if k(t) is taken as (2.11). Hence, with the choice (2.11) for k(t) the kernel 'P of
Equations
(1.1) is continuous
at r = t. Note that (2.11) includes
the results derived
in Buonocore
et al. [1] for the Wiener and the Ornstein-Uhlenbeck
process.
In Sections 3 and 4 under the assumption
that S(t) C2[t0,
T], we shall give a
rigorous proof of (2.11) in the form of a necessary
and sufficient condition
not only
for the case when both endpoints of I are natural boundaries but also when a
zero-flux
condition
holds at one endpoint
of L
3. Two natural
boundaries
In this section use of some well-known
results
by Feller [2] and Fortet [4] will be
made to provide
a rigorous
extension
of the main conclusion drawn in Buonocore et
al. [1]. We shall assume throughout that X(t) is a diffusion process defined in
I= (rl, r2) with ri (i = 1, 2) natural boundaries. Let Al(x) and A2(x) be the
infinitesimal
moments of X(t). We shall assume that transformation
(2.2) maps I
onto (-o, +o). Then, if
(3.1) Bi(y)i 2A(x[A(x) -
=h(y)
and B'(y) exist and are bounded in (-oo, +oo) and if furthermore
B'(y) satisfies a
suitable Lipschitz condition (Fortet [4]), the transition p.d.f. of X(t), with
X(to)
= xo, can be expanded
as (Feller [2])
(3.2) f (x, t xo, to)= H,n,(x,
t xo, to)
n---=O
26 V. GIORNO, A. G. NOBILE, L. M. RICCIARDI
AND S. SATO
where
1 1 x dO 2
(3.3) Ho(x,
t l
xo, to)
= V2;r(t- to)A2(x)
expI 2(t- to)
-
o
VVA()
H,,(x,
t Ixo0, to)
= dr dz Ai(z) z)
(3.4) to r
az
Ho(z, r jXo, to)- H,_•(x, t z, z) (n = 1, 2,...).
As proved
by Fortet [4], the following inequalities
hold:
(3.5) IHn(x,
t I xo, to) a,(t - to)(n-2/2, n = 1,
2, ...
and
(3.6) H(x, t I
xo, to)
-
fn(t - to)(n-2)/2, n = 1, 2,
where ar,,
and
fi, are the terms of convergent sequences, both independent
of x, t, xo
and to. From (3.2)-(3.4) one can prove that
f(x, tI xo, to)
= Ho(x,
t ixo, to) 1 + Q1(x,
t IXo,
to) + 2to)t - to
(3.7) + Q3(x,
tI x, to)
[f
•A
]2 +Q4(x, t I
Xo, to)Vt
- to od
+ Qs(x, tI xo, to)(t - to)J + Q6(x, t Xo, to)I t- to
where Qi(x, t I
xo, to), (i = 1, 2, .- -. , 6) are bounded functions.
In particular,
Q1
and
Q2
are
given
by
(38)1 (x. exp (_- 2/2) - 1 dA2[p(?)]
(3.8) Ql(x, tj Xo, to) dpf d A AI[4() 4
V/A2[4(f)] -4
dpV(A) J'
Q2(x,t3I
Xo,
to)
= - dp d2
9 exp
(_-2/2)
(3.9) x fA[( )]1 dA(0)]
a[)1 4 d)
[(J)
where
(3.10) (5) = h-1{1/2(t - to)p(1 - p) + h(xo) + p[h(x) - h(xo)]}.
Evaluation
of first-passage-time
probability
densities 27
Lemma
3.1. Let S(t) E C[to, T]. Then
limW[S(t), tI S(r), r]
(3.11)
Tt f[S(t), t S(r), r]
iff
1{ C AJ[S(t)] S()
(3.12) k(t) = {Ai[S(t)] - t- S'(t)
Proof. From (3.2)-(3.10) we have
lim
{f[S(t), t I
S(r), r]}l1
-'f [, t I
S(), r]
,rTt ax x=S(t)
A'[S(t)] S'(t) QI[S(t), t l S(r), r]
(3.13) + lim
2A2[S(t)] A2[S(t)] rTt VA2[S(t)]
_ _ 3
A2[S () -S'(t) + A[S(t) A[S(t) .
A2[S(t)] 4
Making
use of Lemma
1.1, from (3.13) we obtain
u[s(t),
t S(r), •] s'(t) AI[S(t)] A'[S(t)]
(3.14) lim + k(t)
+2
rTt f[S(t), t S(T), r] 2 2 8
which
vanishes
if and only if (3.12) holds.
Theorem
3.1. Let S(t) E C2[to,
T]. Then
(3.15) lim
WI[S(t),
tI
S(r), r] = 0
rTt
iff (3.12) holds.
Proof. Let k(t) be given by (3.12). From (1.5) we then have
W[S(t), I, r] St) A [S(t)] +-AI[S(t)] f[S(t), t Iz, r]
2 2 8
(3.16) A2[S(t)] t I
2 axf Ix=S(t)
Recalling
that Ho[S(t),
t I
S(r), r] exhibits a singularity
of the type (t - r)-1 as
tends to t and using (3.8) and (3.9), due to the boundness
of function
Qi, we obtain
1
i
1 {[S'(t) Al[S(t)]+ A[S(t)]
lim W[S(t), t S(r), r] (limt AS(
xtt -2rA2[S(t)] t (t r) 2 2 4
(3.17) x (t - /) A2S(t)] tiS)
A2[S(t)] Q1[S(t), t S(t), (]-
+v'A2[S(t)] (t- )
28 V. GIORNO,
A. G. NOBILE,
L. M. RICCIARDI
AND
S. SATO
The
right-hand
side of (3.17)
is an indeterminate form.
After a twofold
application
of l'Hospital's
rule and
recalling
that
S(t)
E
C2[t0,
T], we finally
obtain
(3.15).
This
proves
the sufficient
part
of the theorem. To prove
the necessity,
let us assume that
(3.15)
holds,
i.e. that
Y[S(t),
tIS(r),
r]
(3.18) lim I[S(t), tI
S(r), r] limf[S(t), tI
S(r),
'] 0.
rTt rTt f[s(t),
tI
S(t),
r]
Since
f[S(t), t I
S(r), r] has a singularity
of the type (t - r)-I as r grows to t, the
vanishing
of P[S(t), t I
S(r), r] as nrt implies
the vanishing
of the ratio in (3.18) as
rTt.
This, in turn, by Lemma
3.1 implies
that k(t) is given by (3.12).
4. One natural
and one reflecting
or entrance
boundary
In this section we extend Lemma 3.1 and Theorem 3.1 to the case when the
diffusion
process is constrained to the interval
JI= (r, r2), ri? rl, such that r2 is a
natural
boundary
while r is either entrance or regular
with a zero-flux
condition.
With some obvious
changes,
the results of this section can be proved
to hold in the
case when J= (rl, r), r 5 r2. In what follows, we shall assume that the
transformation
(4.1) y = h(x)
maps
J onto (0, +oo).
Furthermore,
we assume
that
(4.2) C(y) - A(x) A 4 x=h'(y)
and C'(x) exist and are bounded in (0, +oo) and that C'(y) satisfies a suitable
Lipschitz
condition
(Fortet [4]).
Let us set
Do(x, t I xo, to) = exp l-1 [X dO ]21
D(x, tX2o, to) r(t - to)A2 ex p- 2(t - to) oA2(0)
1 X
dO
d
2
(4f3)2(t - to)
- If fVA
(432
r-2(t
) L
VA2(z-0)LrVA(0)J
D,,(x, t xo, to)= dr dz Ai(z)- A(z)]
(4.4)r
x Do(z, I
xo,
to) zD
O_(x, tI
zo,
ro) (n=
1,2,-..).
If
(4.5) lim exp2 dz ) Do(x, txo, o)=
o A2(z)J aD0,
Evaluation
offirst-passage-time
probability
densities 29
then the transition
p.d.f. f(r)(x, t I
xo, to) of X(t) conditional on X(to) = xo and in the
presence
of a zero-flux condition
at r can be expanded
as
(4.6) f(r)(x, t l
xo, to)= > D,(x, t Xo, to),
n=O
where
the functions
Dn(x,
t I
xo, to) and - D,(x, t I
xo, to), n >_-
1, are bounded
as in
&xo
(3.5) and (3.6). This can be proved by showing that the right-hand
side of (4.6)
satisfies
the Kolmogorov
backward
equation,
the initial
delta-condition
and, because
of (4.5), also the zero-flux
condition
(4.7)xolimexp {2fdzA f')(x, t Ixo, t) = 0.
xot) A2 XOr), Ot=
One can further
prove that
f(r)(x, t xo,
to)
= Ho(x,
t
x o, to) 1 + Q t(x,
t I
Xo,
to) •dO
+ Q*(x,
t I
Xo,
to)t - to
+ Q3*(x,
t
I
Xo,
to)
(4.8) X[f dO ]2 xo) dO
x fo VA )+ Q (x,
tI
Xo,
to)Vt-
to o
VA
+ Q5*(x,
t I
x, to)(t - to) + Q(xt, t to)
t0,
where Q*(x, t xo, to), (i = 1, 2, - -, 6) are bounded functions. In particular,
we
have:
1
A4() { A1[
Ia[41(0)]}
Q*(x,,
t x0, to)
= dpf A1[1()] - 4
(4.9) 1 4 4
X E(p,,(x, t Ixo,
to)F(p, )(x, t I
x,
to),
Q~f(x, t xo, to) = - dp d?
P A2[1
(4.10) x
A,[(?)]- A2[4,()]
x E(p
?)(x,
t IXo,
to)F(p, )(x,
t lxo,
to)
where
we have set
Ep,(xt I
xo, to)=1+ exp {-2 t o
P 0
(4.11a) xx
2]
Vo dIOO o d }
x
exp
_(t
- to)P
or?
30 V. GIORNO,
A. G. NOBILE, L. M. RICCIARDI
AND S. SATO
-
t
(o -0) _rd
FSg(x,
txo,
to)
=exp(-52/2)-exp
t
-
to
do
2
@
x exp - (t-1(t - to)(1 - ) 2Vd
(4.11b) 1xo do x do rx do
x
V(t - to)p(1 - p) + + )+ p
x (1 - p4)
x (- - (t- to)p(1
- P)]
(4.11c) 1p(j)
= hi'1{• 2(t
- to)p(1
- p) + hi(xo)
+ p[hl(x)
- hl(xo)]}
(4.11d) l= v(t - to)p(1 - p) r 2(0- xo 2(0)
Lemma
4.1. Let S(t) E C1[to,
T]. Then
(4.12)[S(t),
tl S(r), z]
(4.12) lim = 0
,tr
f()[S(t),
t S(t), -]
iff
1{ Aa'[S(t)]
(4.13) k(t) = -2 AA[S(t)] - 4 - S'(t) .
2 4
Proof. Similar
to that of Lemma
3.1. Indeed, we have
lim {f(')[S(t), tI S(T), r]})-1 f(r)[x,
tI S(), r]
(4.14) T =S(t)
A'[S(t)] S'(t) Q
?[S(t), t S(r), c]
+ lim
2A2[S(t)] A2[S(t)] rT A
Since
1
---4A[S(t)]1
lim Q*[S(t), t I
S(r), T] =- ArA tIAI[S(t)]
(4.15) rTt 2 [S(t)]
1o
+= 2 1 A [S(t)]
x dpf d exp - =~/-} A[S(t)] - A
,S(t)
substituting (4.15) in (4.14) and making
use of Lemma 1.2 the proof of Lemma
4.1
follows.
Theorem
4.1. Let S(t) E C2[t0,
T]. Then
(4.16) lim WII(')[S(t),
tI S(t), r] = 0
iff (4.13) holds.
Evaluation
of first-passage-time
probability
densities 31
Proof. Recalling (1.9) and Lemma 4.1, the necessary and sufficient parts of
Theorem 4.1 can be proved as in Theorem 3.1. However, we remark that the
evaluation
of ~Pr)[S(t), t I
S(r), r] as r tends to t requires
a cumbersome
proof that
for brevity
will be omitted.
5. Examples
(a) The
lognormal
process. Let
(5.1) Al(x) = mx, A2(x) = 2 (x 0),
with m and 2 > 0 arbitrary
real numbers. The diffusion
process having moments
(5.1) is defined over the diffusion interval (0, +o), with the endpoints natural
boundaries. Transformation (2.2) thus maps the diffusion interval onto (-oo, +oo),
while condition (3.1) and the other regularity
conditions of Section 3 are satisfied.
Then, from (3.12) and (5.1) we have
1 /
(5.2)
k(t)
=1[(m
-
m
S(t)
-
S'(t)
.
2
2
Since
(5.3) f(x, t z, r)=n exp - r22- r) 1
xu /23r(t - r) 12&(t - r)
from (1.5) and (5.2) we obtain
(5.4) W[S(t),
t I
z, r] = - S'(t)- tIn [St(t), t z, r].
2
t
-'r
z
Since the right-hand
side of (5.4) vanishes identically for z = S(r) and S(t)=
exp (at + b), where a, b E R from (1.1) we have
eat+b
(5.5) g(eat+b, t I
xo, to) = Iln
xo - ato - bl f (eat+b, t I
xo, to) (xo / eat+b).
t - to
(b) The
hyperbolic
process. Let
A - B exp (-2px/Ir2)
(5.6) Al(x) A + B exp (-2x/)' A2(x) A2 = 2,
with A, B real numbers such that AB > 0. We incidentally
note that the Wiener
process is a special case of the process having moments (5.6). Now the diffusion
interval
is the entire real line, with the points at infinity
natural
boundaries.
Clearly,
transformation
(2.2) maps the diffusion
interval onto itself. From (3.12) and (5.6)
we obtain
i a -B exp [-21iS(t)/u2]
(5.7) kt= g2t A + B exp [-2,iS(t)/a2] - JJ.
32 V. GIORNO, A. G. NOBILE, L. M. RICCIARDI
AND S. SATO
Since
(5.8) f(x,tz,) t- A + B exp (-2xz/o2) Cexp
1I- 2 (t
_ r)
(rV2xr(t- r) A + B exp (-2pz/or2) 2or2(t
--
)
_'
from
(1.5)
and
(5.7)
we obtain
1 S(t) - z f[S(t), i
z' T].
(5.9) W[S(t),
t I
z, r]
= s'(t)- S(t)- [S(t) t z, r].
(A detailed
study
of the process (5.8) will be presented
elsewhere.)
We note that the
right-hand
side of (5.9) is identically
0 for z = S(r) and S(t) = at + b. From (1.1) it
then follows that
1xo
- at0 - b
(5.10) g(at + b, t xo, to)
= f(at + b, t I
xo, to) (xo
* ato
+ b).
t - to
(c) Wiener
process with a reflecting boundary. Consider a Wiener process with
infinitesimal
moments
(5.11) A1
=p , A2 = &2
defined in [r, +oo), with r a reflecting boundary.
Since condition
(4.3) and the other
regularity
conditions of Section 4 are satisfied and since (4.5) is also valid, from
(4.13) we have
(5.12) k(t) = ?[p - S'(t)].
Furthermore,
since
1
[
[
x-z
-1i(t-
-)2
f'(r)(x,
ti Z, r) exp
2_(t-_r)
OV2x (tJ- r) I f 2u2(t - r)
(5.13) +exp( 2p(zr) }exp [x + z - 2r - p(t - r)]2
I 222(t - r)
Sexp
2(x - r) x- +z - 2r
+(t - )
from (1.9) we obtain
2 1 S(t)- z
t(r)[S(t), tI z, r]
=2S'(t)-St-
x fexp [S(t) - z - P(t - r)1]2
2o2(t - r) +2aoV2r(t- r)
(5.14) x [S'(t)
S(t) + z - 2r - 2p(t - r) ]exp- 2(z - r)
xexp [S(t)+ z-2r -tp(t- )]2 u [S'(t)+ p]
Sexp
2(t - ) 22
xexp f2[S(t) - r] Erf [S(t) + z-2r + (t - )
xxp1V2(t-Er)
Evaluation of first-passage-time probability densities 33
g(r)
0.25 '
0.20
0.15
0.10
0.05
2 4 6 8 TIME
Figure 1. The oscillating curve is the first-passage-time p.d.f. g(') through S(t) = 3 + 0-1 cos (2.rt) for the
Wiener process having infinitesimal variance equal to 2 and drift y = 0 starting on a reflecting boundary at
x = 0. The first-passage-time p.d.f. through the mean threshold (S(t)) = 3 has also been plotted.
g(r)
0.35
0.28
0.21
0.14
0.076 8 TI
2 4 6 8 TIME
Figure 2. As in Figure 1 except that now M
= 0.5.
g(r)
0.15
0.12-
0.09-
0.06-
0.03 -
2 4 6 8 TIME
Figure 3. As in Figure 1 but with M
= -0-5.
34 V. GIORNO, A. G. NOBILE, L. M. RICCIARDI AND S. SATO
A numerical
evaluation
of g[S(t), t I
xo, to]
can then be obtained
from (1.12), (5.13)
and (5.14) by means of the algorithm
of Buonocore
et al. [1]. Figures
1-3 show the
first-passage-time
p.d.f. through the boundary
S(t)= 3 + 0.1 cos (2xrt)
for Wiener
processes
originating
in a reflecting
boundary
at r = 0.
We remark
that the examples
discussed in this section involve diffusion
processes
all satisfying
conditions
(3.1) or (4.2). In Section 6 an example will be discussed in
which
condition
(3.1) does not hold.
6. Process with linear drift and linear infinitesimal
variance
Let X(t) be the diffusion
process
with infinitesimal
moments
(Feller [3])
(6.1) Al(x) =px + q, A2(x) = 2rx,
with p, q, r E R and r > 0. The diffusion interval is I = [0, +oo). Hereafter we shall
assume q > 0 so that the point 0 is either a regular
or an entrance boundary.
As
proved in Giorno et al. [5], the expression
of the transition
p.d.f. of X(t) in the
presence
of a zero-flux
condition
at zero is given by
f(r)(x, t I
Z, )
r)=[eP <-) - 1] exp ze" -
(6.2) r[ -- (q-r)2rp
p xze
X[xe-
P
-
T)
(q-r)/2r
I[
2p
V
xz
eP
t -
cz
e
*
zLr[eP
'-')
-
1]
p
0;
f(r)(x, t z, r) = 1 (X\(qr)/2r exp{ +>
(6.3)x, r(t
-
e
r(t
- )
(6.3)
[ 2V ]
Ir[(t
- r)' p =0.
Using (1.9), we obtain
Wr)[S(t)z, t z, ] =[e _ exp p[S(t) +
zeP(t-]p
S(t)zpt)(q-r)2r
(6.4)
x[S'(t)pS(t)eP(t-) )[2pVS(t)_zept_-
(6.4) x S'(t) eP('T) -1 + k(t)I lr-1 r[eP(t -
1- ]
p
V/S(t)ze"t- I [2pVS
(t)zePt-}) P
+ ep"-) - 1 q/rL r[e(-)- 1] 1, p =;
W(r)[S(t), ti z, r]= r(t-
1) exp1
r(t)+z-
) (q-r)/2r ( S(t) + k(t)
(
r(t-r) r(t-) z(t
(6.5)
2lir [zS$]+ \7St) iq/r,[2p"S7) = 0.
Cay(1dn
tty- r(tt
- ()
Clearly, (6.1) do not satisfy condition (4.2) so that the procedure of Section 4 cannot
Evaluation
of first-passage-time
probability
densities 35
be used. However, in this case a direct calculation leads to the expressions
of the
functions
W(r)[S(t), t I
z, r] and k(t) such that the kernel of integral
equation (1.11)
is continuous
at r = t. Indeed, we have the following
result.
Theorem
6.1. Let S(t) e C2[to, T]. Then:
(6.6) lim W(r)[s(t),
t I
S(r), ] = 0
rTt
iff
(6.7) k(t) = - pS(t) + q - S'(t)].
For brevity,
we omit the proof.
We conclude with a special case of the process considered in this section, that
leads us to a new closed form result for the first-passage-time p.d.f. when S(t)
reduces
to an arbitrary
constant
S. Let p = 0, q = 3r/2, S(t) = S and X(0) xo
> S.
Making use of well-known properties of the Bessel function (Gradshteyn and
Ryzhik
[6]):
(6.8) II/2(z) = (eZ
- e-),
1 [1 )]
(6.9) I3/2(z)
- + e-z (ez - e-z)
from (6.5) with the choice (6.7) we obtain
(6.10) W(r)(S,
t I
z, r)
= 'I1(S,
t I
z, ) + P2(S, t I
z, r),
where we have set
1 S [
-_
]2
(6.11) 1,(S,
tl z, )=
1-•r / exp -
r-_)12
2 xrr(t-r) Lt-,r(t-r)/r(t-r)
1 S
[V•
+
V]2
(6.12) T 2(S, t ] z, r) = 1
_
+
(
S
-exp +-
r.t
2V rr(t--r) Lt- (t-r)/ r(t - )
The following
theorem
then holds.
Theorem
6.3. For
xo > S one has
(6.13) g(r)(S,
t I
xo,
to)
= 2T,(S,
t I
xo,
to)
where IPl(S, t I
xo, to) is obtained
from (6.11).
Proof. We note that IIl(S, t I
S, r) = 0, for all t, r (r < t). From Equation
(1.11),
we thus have
g(r')(s, t I
xo, to)= 2[WI(S, t I
xo, to) + 2(S, t I
xo, to)]
(6.14) - drW2(S, tI S, r)g(r)(s, r xo, to).
tO
36 V. GIORNO, A. G. NOBILE, L. M. RICCIARDI AND S. SATO
To prove (6.13), we substitute
g(r)(S,
r xo, to) with 2WI(S,
r I
xo, to) in the integral
on the right-hand
side of (6.14). This integral
can be explicitly
calculated.
Indeed,
with the change of variable (r - to)/(t - r) = p one finds
2 drW2(S,
t S, r)WI(S, xo,to)= S 1- Sexp (V- o)
to I rr r(t - to)
1
_ ((fs-Vo)2 4Sp 1
x
120
dp
(p-
+
Vpi)
exp
(VS-
_)Sp
(t to) PP+Pexp r(t - to)p r(t - to)
= t2(S,
tI
xo,
to).
Equation
(6.14) thus identifies with (6.13), which
proves the theorem.
Note that, differently
from the closed form
solutions
of Section 5, the closed form
result (6.13) holds even though the kernel of the integral equation for the
first-passage-time
p.d.f. is non-vanishing
for all t and r.
Acknowledgements
Research carried out under CNR contracts
Nos. 86.02127, 86.02115, 87.01005,
and under
MPI financial
support.
References
[1] BUONOCORE, A., NOBILE,
A. G. AND
RICCIARDI,
L. M. (1987) A new integral equation for the
evaluation of first-passage-time probability densities. Adv. Appl. Prob. 19, 784-800.
[2] FELLER,
W. (1936) Zur Theorie der stochastichen Prozesse. Math. Ann. 113, 113-160.
[3] FELLER,
W. (1951) Two singular diffusion processes. Ann. Math. 54, 173-182.
[4] FORTET,
R. (1943) Les fonctions al6atoires du type de Markoff associ6es A
certaines equations
lin6aires aux d6riv6es partielles du type parabolique. J. Math. Pures Appl. 22, 177-243.
[5] GIORNO,
V., NOBILE,
A. G., RICCIARDI, L. M. AND
SACERDOTE,
L. (1986) Some remarks on the
Rayleigh process. J. Appl. Prob. 23, 398-408.
GARDSHTEYN, I. S. AND RYZHIK,
I. M. (1965) Tables of Integrals, Series and Products. Academic
Press, New York.
... For the Wiener process, r, r < z, is a regular boundary; for the LDLIV process, 0 is a regular boundary. 3 Although the general form of the kernel function derived by Buonocore et al. does not apply to either of these processes, Giorno et al. (1989a) showed that it is nevertheless possible to derive kernel functions for Eq. 3 in which the resulting discrete approximations are stable. I draw heavily on their results throughout the remainder of this article. ...
... In Eq. 4, I ν (·), is a modified Bessel function of the first kind of order ν (Abramowitz and Stegun 1965, p. 374), and k(t) is a function that is chosen to make the kernel of the integral equation nonsingular (Giorno et al. 1989a, Eq. 6.7), which is of the form ...
... Models of the form of Eq. 6 have a long history in the RT literature. An advantage of this representation is that it carries over easily to n-choice decisions, but I restrict myself 4 I have corrected a typographical error in the expression for the kernel function in Giorno et al. (1989a) (a missing "e" denoting exponentiation in the numerator of the third term on the right-hand side in their Eq. 6.4.) ...
Article
Full-text available
Diffusion models of decision making, in which successive samples of noisy evidence are accumulated to decision criteria, provide a theoretical solution to von Neumann’s (1956) problem of how to increase the reliability of neural computation in the presence of noise. I introduce and evaluate a new neurally-inspired dual diffusion model, the linear drift, linear infinitesimal variance (LDLIV) model, which embodies three features often thought to characterize neural mechanisms of decision making. The accumulating evidence is intrinsically positively-valued, saturates at high intensities, and is accumulated for each alternative separately. I present explicit integral-equation predictions for the response time distribution and choice probabilities for the LDLIV model and compare its performance on two benchmark sets of data to three other models: the standard diffusion model and two dual diffusion model composed of racing Wiener processes, one between absorbing and reflecting boundaries and one with absorbing boundaries only. The LDLIV model and the standard diffusion model performed similarly to one another, although the standard diffusion model is more parsimonious, and both performed appreciably better than the other two dual diffusion models. I argue that accumulation of noisy evidence by a diffusion process and drift rate variability are both expressions of how the cognitive system solves von Neumann’s problem, by aggregating noisy representations over time and over elements of a neural population. I also argue that models that do not solve von Neumann’s problem do not address the main theoretical question that historically motivated research in this area.
... Using (16) and the elementary inequality e −a 2 ≤ e −a , ≥ 1, a > 0, we obtain ...
Preprint
We propose a discrete time discrete space Markov chain approximation with a Brownian bridge correction for computing curvilinear boundary crossing probabilities of a general diffusion process on a finite time interval. For broad classes of curvilinear boundaries and diffusion processes, we prove the convergence of the constructed approximations in the form of products of the respective substochastic matrices to the boundary crossing probabilities for the process as the time grid used to construct the Markov chains is getting finer. Numerical results indicate that the convergence rate for the proposed approximation with the Brownian bridge correction is $O(n^{-2})$ in the case of $C^2$-boundaries and a uniform time grid with $n$ steps.
... Subsequently, the class of diffusion processes where the problem was studied, as well as their associated class of boundaries, were generalized. In this way, in the context of time-homogeneous diffusion processes, Giorno et al. [18] proved that the probability density function of the f.p.t. variable (2), through a C 1 [t 0 , T]-class function S, verifies a second kind Volterra integral equation with a non-singular kernel. ...
Article
Full-text available
A central aspect of in vivo experiments with anticancer therapies is the comparison of the effect of different therapies, or doses of the same therapeutic agent, on tumor growth. One of the most popular clinical endpoints is tumor growth delay, which measures the effect of treatment on the time required for tumor volume to reach a specific value. This effect has been analyzed through a variety of statistical methods: conventional descriptive analysis, linear regression, Cox regression, etc. We propose a new approach based on stochastic modeling of tumor growth and the study of first-passage time variables. This approach allows us to prove that the time required for tumor volume to reach a specific value must be determined empirically as the average of the times required for the volume of individual tumors to reach said value instead of the time required for the average volume of the tumors to reach the value of interest. In addition, we define several measures in random environments to compare the time required for the tumor volume to multiply k times its initial volume in control, as well as treated groups, and the usefulness of these measures is illustrated by means of an application to real data.
... Stochastic processes and related first-passage-time (FPT) problems have received increasing attention in the literature within the context of model building in biology (see, for instance, Ricciardi et al., 1999, and references therein). Consequently, efforts have also been directed towards the design of computational methods suitable to estimate FPT densities, in particular for general Gaussian processes (see Di Nardo et al., 2001) and for diffusion processes (see references in Ricciardi et al., 1999;in particular, Buonocore et al., 1987, andGiorno et al., 1989). ...
Preprint
The first-passage-time problem for a Brownian motion with alternating infinitesimal moments through a constant boundary is considered under the assumption that the time intervals between consecutive changes of these moments are described by an alternating renewal process. Bounds to the first-passage-time density and distribution function are obtained, and a simulation procedure to estimate first-passage-time densities is constructed. Examples of applications to problems in environmental sciences and mathematical finance are also provided.
Article
We propose a discrete-time discrete-space Markov chain approximation with a Brownian bridge correction for computing curvilinear boundary crossing probabilities of a general diffusion process on a finite time interval. For broad classes of curvilinear boundaries and diffusion processes, we prove the convergence of the constructed approximations in the form of products of the respective substochastic matrices to the boundary crossing probabilities for the process as the time grid used to construct the Markov chains is getting finer. Numerical results indicate that the convergence rate for the proposed approximation with the Brownian bridge correction is $O(n^{-2})$ in the case of $C^2$ boundaries and a uniform time grid with n steps.
Article
This paper analyzes a method to approximate the first passage time probability density function which turns to be particularly useful if only sample data are available. The method relies on a Laguerre-Gamma polynomial approximation and iteratively looks for the best degree of the polynomial such that a normalization condition is preserved. The proposed iterative algorithm relies on simple and new recursion formulae involving first passage time moments. These moments can be computed recursively from cumulants, if they are known. In such a case, the approximated density can be used also for the maximum likelihood estimates of the parameters of the underlying stochastic process. If cumulants are not known, suitable unbiased estimators relying on κ-statistics might be employed. To check the feasibility of the method both in fitting the density and in estimating the parameters, the first passage time problem of a geometric Brownian motion is considered.
Article
Full-text available
Diffusion models of evidence accumulation have successfully accounted for the distributions of response times and choice probabilities from many experimental tasks, but recently their assumption that evidence is accumulated at a constant rate to constant decision boundaries has been challenged. One model assumes that decision-makers seek to optimize their performance by using decision boundaries that collapse over time. Another model assumes that evidence does not accumulate and is represented by a stationary distribution that is gated by an urgency signal to make a response. We present explicit, integral-equation expressions for the first-passage time distributions of the urgency-gating and collapsing-bounds models and use them to identify conditions under which the models are equivalent. We combine these expressions with a dynamic model of stimulus encoding that allows the effects of perceptual and decisional integration to be distinguished. We compare the resulting models to the standard diffusion model with variability in drift rates on data from three experimental paradigms in which stimulus information was either constant or changed over time. The standard diffusion model was the best model for tasks with constant stimulus information; the models with time-varying urgency or decision bounds performed similarly to the standard diffusion model on tasks with changing stimulus information. We found little support for the claim that evidence does not accumulate and attribute the good performance of the time-varying models on changing-stimulus tasks to their increased flexibility and not to their ability to account for systematic experimental effects. (PsycInfo Database Record (c) 2021 APA, all rights reserved).
Article
Full-text available
Departing from a general stochastic model for a moving boundary problem, we consider the density function of probability for the first passing time. It is well known that the distribution of this random variable satisfies a problem ruled by an advection-diffusion system for which very few solutions are known in exact form. The model considers also a deterministic source, and the coefficients of this equation are functions with sufficient regularity. A numerical scheme is designed to estimate the solutions of the initial-boundary-value problem. We prove rigorously that the numerical model is capable of preserving the main characteristics of the solutions of the stochastic model, that is, positivity, boundedness and monotonicity. The scheme has spatial symmetry, and it is theoretically analyzed for consistency, stability and convergence. Some numerical simulations are carried out in this work to assess the capability of the discrete model to preserve the main structural features of the solutions of the model. Moreover, a numerical study confirms the efficiency of the scheme, in agreement with the mathematical results obtained in this work.
Article
We investigate the joint distribution and the multivariate survival functions for the maxima of an Ornstein-Uhlenbeck (OU) process in consecutive time-intervals. A PDE method, alongside an eigenfunction expansion is adopted, with which we first calculate the distribution and the survival functions for the maximum of a homogeneous OU-process in a single interval. By a deterministic time-change and a parameter translation, this result can be extended to an inhomogeneous OU-process. Next, we derive a general formula for the joint distribution and the survival functions for the maxima of a continuous Markov process in consecutive periods. With these results, one can obtain semi-analytical expressions for the joint distribution and the multivariate survival functions for the maxima of an OU-process, with piecewise constant parameter functions, in consecutive time periods. The joint distribution and the survival functions can be evaluated numerically by an iterated quadrature scheme, which can be implemented efficiently by matrix multiplications. Moreover, we show that the computation can be further simplified to the product of single quadratures if the filtration is enlarged. Such results may be used for the modeling of heatwaves and related risk management challenges.
Article
The first-passage-time p.d.f. through a time-dependent boundary for one-dimensional diffusion processes is proved to satisfy a new Volterra integral equation of the second kind involving two arbitrary continuous functions. Use of this equation is made to prove that for the Wiener and the Ornstein-Uhlenbeck processes the singularity of the kernel can be removed by a suitable choice of these functions. A simple and efficient numerical procedure for the solution of the integral equation is provided and its convergence is briefly discussed. Use of this equation is finally made to obtain closed-form expressions for first-passage-time p.d.f.'s in the case of various time-dependent boundaries.
Article
The transition p.d.f. for a one-dimensional Rayleigh process in the presence of an absorption condition or a zero-flux condition in the origin is obtained in closed form. The first-passage-time problem through an arbitrary constant boundary is then considered and the moment-generating function is determined. In some particular cases the first-passage-time p.d.f. is explicitly derived. Use of some of these results is finally made to obtain the transition p.d.f. of the affine drift-linear infinitesimal-variance diffusion process when the origin is an entrance or a regular boundary in the presence of a reflection condition.
Article
The transition p. d. f. for a one-dimensional Rayleigh process in the presence of an absorption condition or a zero-flux condition in the origin is obtained in closed form. The first-passage-time problem through an arbitrary constant boundary is then considered and the moment-generating function is determined. In some particular cases the first-passage-time p. d. f. is explicitly derived. Use of some of these results is finally made to obtain the transition p. d. f. of the affine drift-linear infinitesimal-variance diffusion process when the origin is an entrance or a regular boundary in the presence of a reflection condition.
Article
The first-passage-time p.d.f. through a time-dependent boundary for one-dimensional diffusion processes is proved to satisfy a new Volterra integral equation of the second kind involving two arbitrary continuous functions. Use of this equation is made to prove that for the Wiener and the Ornstein-Uhlenbeck processes the singularity of the kernel can be removed by a suitable choice of these functions. A simple and efficient numerical procedure for the solution of the integral equation is provided and its convergence is briefly discussed. Use of this equation is finally made to obtain closed-form expressions for first-passage-time p.d.f.'s in the case of various time-dependent boundaries.
A new integral equation for the evaluation of first-passage-time probability densities Zur Theorie der stochastichen Prozesse Two singular diffusion processes
  • A Nobile
  • A G And Ricciardi
  • L M Feller
  • W Feller
  • W Fortet
BUONOCORE, A., NOBILE, A. G. AND RICCIARDI, L. M. (1987) A new integral equation for the evaluation of first-passage-time probability densities. Adv. Appl. Prob. 19, 784-800. [2] FELLER, W. (1936) Zur Theorie der stochastichen Prozesse. Math. Ann. 113, 113-160. [3] FELLER, W. (1951) Two singular diffusion processes. Ann. Math. 54, 173-182. [4] FORTET, R. (1943) Les fonctions al6atoires du type de Markoff associ6es A certaines equations lin6aires aux d6riv6es partielles du type parabolique. J. Math. Pures Appl. 22, 177-243.
Les fonctions al6atoires du type de Markoff associ6es A certaines equations lin6aires aux d6riv6es partielles du type parabolique Some remarks on the Rayleigh process
  • R Giorno
  • V Nobile
  • A G Ricciardi
  • L M And
  • L Sacerdote
FORTET, R. (1943) Les fonctions al6atoires du type de Markoff associ6es A certaines equations lin6aires aux d6riv6es partielles du type parabolique. J. Math. Pures Appl. 22, 177-243. [5] GIORNO, V., NOBILE, A. G., RICCIARDI, L. M. AND SACERDOTE, L. (1986) Some remarks on the Rayleigh process. J. Appl. Prob. 23, 398-408.