Integrals-Wikipedia Pages On Integral Definitions
Integrals-Wikipedia Pages On Integral Definitions
PDF generated using the open source mwlib toolkit. See http://code.pediapress.com/ for more information.
PDF generated at: Mon, 30 Nov 2009 15:01:30 UTC
Contents
Articles
Bochner integral
Daniell integral
Darboux integral
HenstockKurzweil integral
Homological integration
11
It calculus
12
Lebesgue integration
19
LebesgueStieltjes integration
27
Motivic integration
30
PaleyWiener integral
31
Pfeffer integral
32
Regulated integral
33
Riemann integral
35
RiemannStieltjes integral
41
RussoVallois integral
44
Skorokhod integral
46
Stratonovich integral
48
References
Article Sources and Contributors
50
51
Article Licenses
License
52
Bochner integral
Bochner integral
In mathematics, the Bochner integral, named for Salomon Bochner, extends the definition of Lebesgue integral to
functions which take values in a Banach space, as the limit of integrals of simple functions.
Definition
Let (X,,) be a measure space and B a Banach space. The Bochner integral is defined in much the same way as the
Lebesgue integral. First, a simple function is any finite sum of the form
where the Ei are disjoint members of the -algebra , the bi are distinct elements of B, and E is the characteristic
function of E. If (Ei) is finite whenever bi 0, then the simple function is integrable, and the integral is then
defined by
Properties
Many of the familiar properties of the Lebesgue integral continue to hold for the Bochner integral. Perhaps the most
striking example is Bochner's criterion for integrability, which states that if (X,,) is a finite measure space, then a
Bochner-measurable function :XB is Bochner integrable if and only if
A function :XB is called Bochner-measurable if it is equal -almost everywhere to a function g taking values in
a separable subspace B0 of B, and such that the inverse image g1(U) of every open set U in B belongs to .
Equivalently, is limit -almost everywhere of a sequence of simple functions.
A version of the dominated convergence theorem also holds for the Bochner integral. Specifically, if n: X B is a
sequence of measurable functions tending almost everywhere to a limit function , and if
for almost every xX, and g L1(), then
as n and
for all E.
Bochner integral
If is Bochner integrable, then the inequality
defines a countably-additive B-valued vector measure on X which is absolutely continuous with respect to .
RadonNikodym property
An important fact about the Bochner integral is that the RadonNikodym theorem fails to hold in general. This
results in an important property of Banach spaces known as the RadonNikodym property. Specifically, if is a
measure on (X,), then B has the RadonNikodym property with respect to if, for every vector measure on
(X,) with values in B which has bounded variation and is absolutely continuous with respect to , there is a
-integrable function g : X B such that
See also
Bochner space
Pettis integral
Vector measure
References
Bochner, Salomon (1933), "Integration von Funktionen, deren Werte die Elemente eines Vectorraumes sind [1]",
Fundamenta Mathematicae 20: 262276
Diestel, Joseph (1984), Sequences and series in Banach spaces. Graduate Texts in Mathematics, Springer-Verlag,
ISBN 0-387-90859-5
Diestel, J.; Uhl, J. J. (1977), Vector measures, Providence, R.I.: American Mathematical Society, ISBN
978-0-8218-1515-1
Lang, Serge (1969), Real analysis, Addison-wesley, ISBN 0-201-04172-3 (now published by springer Verlag)
Sobolev, V. I. (2001), "Bochner integral [2]", in Hazewinkel, Michiel, Encyclopaedia of Mathematics, Kluwer
Academic Publishers, ISBN 978-1556080104
van Dulst, D. (2001), "Vector measures [3]", in Hazewinkel, Michiel, Encyclopaedia of Mathematics, Kluwer
Academic Publishers, ISBN 978-1556080104
Bochner integral
References
[1] http:/ / matwbn. icm. edu. pl/ ksiazki/ fm/ fm20/ fm20127. pdf
[2] http:/ / eom. springer. de/ B/ b016710. htm
[3] http:/ / eom. springer. de/ V/ v096490. htm
Daniell integral
One of the main difficulties with the traditional formulation of the Lebesgue integral is that it requires the initial
development of a workable measure theory before any useful results for the integral can be obtained. However, an
alternative approach is available, developed by Percy J. Daniell in his 1918 paper "A general form of integral" (Ann.
of Math, 19, 279) that does not suffer from this deficiency, and has a few significant advantages over the traditional
formulation, especially as the integral is generalized into higher dimensional spaces and further generalizations such
as the Stieltjes integral. The basic idea involves the axiomatization of the integral.
of bounded real functions (called elementary functions) defined over some set
is a linear space with the usual operations of addition and scalar multiplication.
2. If a function
is in
and
.
2. Nonnegativity. If
3. Continuity. If
, then
) of functions in
that
These elementary functions and their elementary integrals may be any set of functions and definitions of integrals
over these functions which satisfy these axioms. The family of all step functions evidently satisfies the above axioms
for elementary functions. Defining the elementary integral of the family of step functions as the (signed) area
underneath a step function evidently satisfies the given axioms for an elementary integral. Applying the construction
of the Daniell integral described further below using step functions as elementary functions produces a definition of
an integral equivalent to the Lebesgue integral. Using the family of all continuous functions as the elementary
functions and the traditional Riemann integral as the elementary integral is also possible, however, this will yield
an integral that is also equivalent to Lebesgue's definition. Doing the same, but using the RiemannStieltjes
integral, along with an appropriate function of bounded variation, gives a definition of integral equivalent to the
LebesgueStieltjes integral.
Sets of measure zero may be defined in terms of elementary functions as follows. A set
is a set of measure zero if for any
functions
in H such that
which is a subset of
on
.
, is a set of measure zero. We say that if some
property holds at every point of a set of full measure (or equivalently everywhere except on a set of measure zero), it
holds almost everywhere.
Daniell integral
is defined as:
It can be shown that this definition of the integral is well-defined, i.e. it does not depend on the choice of sequence
.
However, the class
is in general not closed under subtraction and scalar multiplication by negative numbers, but
and
can be represented
in the class
. Then the
Again, it may be shown that this integral is well-defined, i.e. it does not depend on the decomposition of
and
into
Properties
Nearly all of the important theorems in the traditional theory of the Lebesgue integral, such as Lebesgue's dominated
convergence theorem, the RieszFischer theorem, Fatou's lemma, and Fubini's theorem may also readily be proved
using this construction. Its properties are identical to the traditional Lebesgue integral.
Daniell integral
See also
Lebesgue integral
Riemann integral
LebesgueStieltjes integration
References
Daniell, Percy John, 1918, "A general form of integral," Annals of Mathematics 19: 27994.
, 1919, "Integrals in an infinite number of dimensions," Annals of Mathematics 20: 28188.
, 1919, "Functions of limited variation in an infinite number of dimensions," Annals of Mathematics 21:
3038.
, 1920, "Further properties of the general integral," Annals of Mathematics 21: 20320.
, 1921, "Integral products and probability," American Journal of Mathematics 43: 14362.
Royden, H. L., 1988. Real Analysis, 3rd. ed. Prentice Hall. ISBN 978-0-02-946620-9.
Shilov, G. E., and Gurevich, B. L., 1978. Integral, Measure, and Derivative: A Unified Approach, Richard A.
Silverman, trans. Dover Publications. ISBN 0-486-63519-8.
Asplund Edgar and Bungart Lutz, 1966 -"A first course in Integration" - Holt, Rinehart and Winston
library of congress catalog card number-66-10122
Taylor A.E , 1965, "General Theory of Functions and Integration" -I edition -Blaisdell Publishing Companylibrary of congress catalog card number- 65-14566
Darboux integral
In real analysis, a branch of mathematics, the Darboux integral or Darboux sum is one possible definition of the
integral of a function. Darboux integrals are equivalent to Riemann integrals, meaning that a function is
Darboux-integrable if and only if it is Riemann-integrable, and the values of the two integrals, if they exist, are
equal. Darboux integrals have the advantage of being simpler to define than Riemann integrals. Darboux integrals
are named after their discoverer, Gaston Darboux.
Definition
A partition of an interval [a,b] is a finite sequence of values xi such that
Each interval [xi1,xi] is called a subinterval of the partition. Let :[a,b]R be a bounded function, and let
be a partition of [a,b]. Let
Darboux integral
is a partition
In other words, to make a refinement, cut the subintervals into smaller pieces and do not remove any existing cuts. If
Darboux integral
is a refinement of
then
and
If P1, P2 are two partitions of the same interval (one need not be a refinement of the other), then
.
It follows that
Riemann sums always lie between the corresponding lower and upper Darboux sums. Formally, if
and
(as in the definition of the Riemann integral), and if the Riemann sum of corresponding to P and T is R, then
From the previous fact, Riemann integrals are at least as strong as Darboux integrals: If the Darboux integral exists,
then the upper and lower Darboux sums corresponding to a sufficiently fine partition will be close to the value of the
integral, so any Riemann sum over the same partition will also be close to the value of the integral. It is not hard to
see that there is a tagged partition that comes arbitrarily close to the value of the upper Darboux integral or lower
Darboux integral, and consequently, if the Riemann integral exists, then the Darboux integral must exist as well.
See also
Regulated integral
Lebesgue integration
HenstockKurzweil integral
HenstockKurzweil integral
In mathematics, the HenstockKurzweil integral, also known as the Denjoy integral (pronounced [dwa]) and
the Perron integral, is a possible definition of the integral of a function. It is a generalization of the Riemann
integral which in some situations is more useful than the Lebesgue integral.
This integral was first defined by Arnaud Denjoy (1912). Denjoy was interested in a definition that would allow one
to integrate functions like
This function has a singularity at 0, and is not Lebesgue integrable. However, it seems natural to calculate its integral
except over [,] and then let , 0.
Trying to create a general theory Denjoy used transfinite induction over the possible types of singularities which
made the definition quite complicated. Other definitions were given by Nikolai Luzin (using variations on the
notions of absolute continuity), and by Oskar Perron, who was interested in continuous major and minor functions. It
took a while to understand that the Perron and Denjoy integrals are actually identical.
Later, in 1957, the Czech mathematician Jaroslav Kurzweil discovered a new definition of this integral elegantly
similar in nature to Riemann's original definition which he named the gauge integral; the theory was developed by
Ralph Henstock. Due to these two important mathematicians, it is now commonly known as Henstock-Kurzweil
integral. The simplicity of Kurzweil's definition made some educators advocate that this integral should replace the
Riemann integral in introductory calculus courses, but this idea has not gained traction.
Definition
Henstock's definition is as follows:
Given a tagged partition P of [a, b], say
-fine if
Given a function
we now define a number I to be the HenstockKurzweil integral of f if for every >0 there exists a gauge
that whenever P is
-fine, we have
such
HenstockKurzweil integral
Cousin's lemma states that for every gauge
9
, such a
satisfied vacuously. The Riemann integral can be regarded as the special case where we only allow constant gauges.
Properties
Let f: [a, b] R be any function.
If a < c < b, then f is HenstockKurzweil integrable on [a,b] if and only if it is HenstockKurzweil integrable on
both [a,c] and [c,b], and then
The HenstockKurzweil integral is linear, i.e., if f and g are integrable, and , are reals, then f + g is integrable
and
If f is Riemann or Lebesgue integrable, then it is also HenstockKurzweil integrable, and the values of the integrals
are the same. The important Hake's theorem states that
whenever either side of the equation exists, and symmetrically for the lower integration bound. This means that if f is
"improperly HenstockKurzweil integrable", then it is properly HenstockKurzweil integrable; in particular,
improper Riemann or Lebesgue integrals such as
are also HenstockKurzweil integrals. This shows that there is no sense in studying an "improper
HenstockKurzweil integral" with finite bounds. However, it makes sense to consider improper HenstockKurzweil
integrals with infinite bounds such as
For many types of functions the HenstockKurzweil integral is no more general than Lebesgue integral. For
example, if f is bounded, the following are equivalent:
f is HenstockKurzweil integrable,
f is Lebesgue integrable,
f is Lebesgue measurable.
In general, every HenstockKurzweil integrable function is measurable, and f is Lebesgue integrable if and only if
both f and |f| are HenstockKurzweil integrable. This means that the HenstockKurzweil integral can be thought of
as a "non-absolutely convergent version of Lebesgue integral". It also implies that the HenstockKurzweil integral
satisfies appropriate versions of the monotone convergence theorem (without requiring the functions to be
nonnegative) and dominated convergence theorem (where the condition of dominance is loosened to g(x) fn(x)
h(x) for some integrable g, h).
If F is differentiable everywhere (or with countable many exceptions), the derivative F is HenstockKurzweil
integrable, and its indefinite HenstockKurzweil integral is F. (Note that F need not be Lebesgue integrable.) In
other words, we obtain a simpler and more satisfactory version of the second fundamental theorem of calculus: each
differentiable function is, up to a constant, the integral of its derivative:
HenstockKurzweil integral
10
Conversely, the Lebesgue differentiation theorem continues to holds for the HenstockKurzweil integral: if f is
HenstockKurzweil integrable on [a,b], and
then F(x) = f(x) almost everywhere in [a,b] (in particular, F is almost everywhere differentiable).
McShane integral
Interestingly, Lebesgue integral on a line can also be presented in a similar fashion.
First of all, change of
to
(here
is a
integral equivalent to one given above. But after this change we can drop condition
and get a definition of McShane integral, which is equivalent to Lebesgue integral.
References
Das, A.G. (2008). The Riemann, Lebesgue, and Generalized Riemann Integrals. Narosa Publishers. ISBN
978-8173199332.
Swartz, Charles W.; Kurtz, Douglas S. (2004). Theories of Integration: The Integrals of Riemann, Lebesgue,
Henstock-Kurzweil, and McShane. Series in Real Analysis. 9. World Scientific Publishing Company. ISBN
978-9812566119.
Kurzweil, Jaroslav (2002). Integration Between the Lebesgue Integral and the Henstock-Kurzweil Integral: Its
Relation to Locally Convex Vector Spaces. Series in Real Analysis. 8. World Scientific Publishing Company.
ISBN 978-9812380463.
Bartle, Robert G. (2001). A Modern Theory of Integration. Graduate Studies in Mathematics. 32. American
Mathematical Society. ISBN 978-0821808450.
Swartz, Charles W. (2001). Introduction to Gauge Integrals. World Scientific Publishing Company. ISBN
978-9810242398.
Leader, Solomon (2001). The Kurzweil-Henstock Integral & Its Differentials. Pure and Applied Mathematics
Series. CRC. ISBN 978-0824705350.
Kurzweil, Jaroslav (2000). Henstock-Kurzweil Integration: Its Relation to Topological Vector Spaces. Series in
Real Analysis. 7. World Scientific Publishing Company. ISBN 978-9810242077.
Lee, Peng-Yee; Vborn, Rudolf (2000). Integral: An Easy Approach after Kurzweil and Henstock. Australian
Mathematical Society Lecture Series. Cambridge University Press. ISBN 978-0521779685.
Bartle, Robert G.; Sherbert, Donald R. (1999). Introduction to Real Analysis (3rd ed.). Wiley. ISBN
978-0471321484.
Gordon, Russell A. (1994). The integrals of Lebesgue, Denjoy, Perron, and Henstock. Graduate Studies in
Mathematics. 4. Providence, RI: American Mathematical Society. ISBN 978-0821838051.
elidze, V G; Dvarevili, A G (1989). The Theory of the Denjoy Integral and Some Applications. Series in
Real Analysis. 3. World Scientific Publishing Company. ISBN 978-9810200213.
Lee, Peng-Yee (1989). Lanzhou Lectures on Henstock Integration. Series in Real Analysis. 2. World Scientific
Publishing Company. ISBN 978-9971508913.
HenstockKurzweil integral
Henstock, Ralph (1988). Lectures on the Theory of Integration. Series in Real Analysis. 1. World Scientific
Publishing Company. ISBN 978-9971504502.
McLeod, Robert M. (1980). The generalized Riemann integral. Carus Mathematical Monographs. 20.
Washington, D.C.: Mathematical Association of America. ISBN 978-0883850213.
External links
The following are additional resources on the web for learning more:
http://www.math.vanderbilt.edu/~schectex/ccc/gauge/
http://www.math.vanderbilt.edu/~schectex/ccc/gauge/letter/
Homological integration
In the mathematical fields of differential geometry and geometric measure theory, homological integration or
geometric integration is a method for extending the notion of the integral to manifolds. Rather than functions or
differential forms, the integral is defined over currents on a manifold.
The theory is "homological" because currents themselves are defined by duality with differential forms. To wit, the
space Dk of k-currents on a manifold M is defined as the dual space, in the sense of distributions, of the space of
k-forms k on M. Thus there is a pairing between k-currents T and k-forms , denoted here by
defined by
for all k. This is a homological rather than cohomological construction.
References
Federer, Herbert (1969), Geometric measure theory, series Die Grundlehren der mathematischen Wissenschaften,
Band 153, New York: Springer-Verlag New York Inc., pp.xiv+676, MR0257325 [1], ISBN 978-3540606567
Whitney, Hassler (1957), Geometric integration theory, Princeton University Press, MR0087148 [2]
References
[1] http:/ / www. ams. org/ mathscinet-getitem?mr=0257325
[2] http:/ / www. ams. org/ mathscinet-getitem?mr=0087148
11
It calculus
12
It calculus
It calculus, named after Kiyoshi It,
extends the methods of calculus to
stochastic processes such as Brownian
motion (Wiener process). It has important
applications in mathematical finance and
stochastic differential equations. The central
concept is the It stochastic integral
where X is a Brownian motion or, more generally, a semimartingale. The paths of Brownian motion fail to satisfy the
requirements to be able to apply the standard techniques of calculus. In particular, it is not differentiable at any point
and has infinite variation over every time interval. As a result, the integral cannot be defined in the usual way (see
RiemannStieltjes integral). The main insight is that the integral can be defined as long as the integrand H is
adapted, which means that its value at time t can only depend on information available up until this time.
The prices of stocks and other traded financial assets can be modeled by stochastic processes such as Brownian
motion or, more often, geometric Brownian motion (see BlackScholes). Then, the It stochastic integral represents
the payoff of a continuous-time trading strategy consisting of holding an amount Ht of the stock at time t. In this
situation, the condition that H is adapted corresponds to the necessary restriction that the trading strategy can only
make use of the available information at any time. This prevents the possibility of unlimited gains through high
frequency trading: buying the stock just before each uptick in the market and selling before each downtick.
Similarly, the condition that H is adapted implies that the stochastic integral will not diverge when calculated as a
limit of Riemann sums.
Important results of It calculus include the integration by parts formula and It's lemma, which is a change of
variables formula. These differ from the formulas of standard calculus, due to quadratic variation terms.
Notation
The integral of a process H with respect to another process X up until a time t is written as
This is itself a stochastic process with time parameter t, which is also written as H X. Alternatively, the integral is
often written in differential form dY = H dX, which is equivalent to YY0 =HX. As It calculus is concerned with
continuous-time stochastic processes, it is assumed that there is an underlying filtered probability space.
It calculus
The sigma algebra Ft represents the information available up until time t, and a process X is adapted if Xt is
Ft-measurable. A Brownian motion B is understood to be an Ft-Brownian motion, which is just a standard Brownian
motion with the property that Bt+sBt is independent of Ft for all s,t0.
where, again, the limit can be shown to converge in probability. The stochastic integral satisfies the It isometry
which holds when H is bounded or, more generally, when the integral on the right hand side is finite.
It processes
An It process is defined to be an adapted stochastic process which can be expressed as the sum of an integral with
respect to Brownian motion and an integral with respect to time,
Here, B is a Brownian motion and it is required that is a predictable B-integrable process, and is predictable and
( Lebesgue) integrable. That is,
This is defined for all locally bounded and predictable integrands. More generally, it is required that H be
B-integrable and H be Lebesgue integrable, so that 0t(H22+|H|)ds is finite. Such predictable processes H are
called X-integrable.
13
It calculus
An important result for the study of It processes is It's lemma. In its simplest form, for any twice continuously
differentiable function on the reals and It process X as described above, it states that (X) is itself an It process
satisfying
This is the stochastic calculus version of the change of variables formula and chain rule. It differs from the standard
result due to the additional term involving the second derivative of , which comes from the property that Brownian
motion has non-zero quadratic variation.
Semimartingales as integrators
The It integral is defined with respect to a semimartingale X. These are processes which can be decomposed as
X=M+A for a local martingale M and finite variation processA. Important examples of such processes include
Brownian motion, which is a martingale, and Lvy processes. For a left continuous, locally bounded and adapted
process H the integral HX exists, and can be calculated as a limit of Riemann sums. Let n be a sequence of
partitions of [0,t] with mesh going to zero,
This limit converges in probability. The stochastic integral of left-continuous processes is general enough for
studying much of stochastic calculus. For example, it is sufficient for applications of It's Lemma, changes of
measure via Girsanov's theorem, and for the study of stochastic differential equations. However, it is inadequate for
other important topics such as martingale representation theorems and local times.
The integral extends to all predictable and locally bounded integrands, in a unique way, such that the dominated
convergence theorem holds. That is, if Hn;H and |Hn|J for a locally bounded processJ, then 0t HndX
0tHdX in probability. The uniqueness of the extension from left-continuous to predictable integrands is a result
of the monotone class lemma.
In general, the stochastic integral HX can be defined even in cases where the predictable process H is not locally
bounded. If K=1/(1+|H|) then K and KH are bounded. Associativity of stochastic integration implies that H is
X-integrable, with integral HX =Y, if and only if Y0=0 and KY =(KH)X. The set of X-integrable processes is
denoted by L(X).
Properties
The stochastic integral is a cdlg process. Furthermore, it is a semimartingale.
The discontinuities of the stochastic integral are given by the jumps of the integrator multiplied by the integrand.
The jump of a cdlg process at a time t is XtXt, and is often denoted by Xt. With this notation, (HX)=H
X. A particular consequence of this is that integrals with respect to a continuous process are always themselves
continuous.
Associativity. Let J, K be predictable processes, and K be X-integrable. Then, J is KX integrable if and only if
JK is X integrable, in which case
Dominated convergence. Suppose that Hn H and |Hn| J, where J is an X-integrable process. then HnX
HX. Convergence is in probability at each timet. In fact, it converges uniformly on compacts in probability.
The stochastic integral commutes with the operation of taking quadratic covariations. If X and Y are
semimartingales then any X-integrable process will also be [X,Y]-integrable, and [HX,Y] = H[X,Y]. A
consequence of this is that the quadratic variation process of a stochastic integral is equal to an integral of a
14
It calculus
quadratic variation process,
Integration by parts
As with ordinary calculus, integration by parts is an important result in stochastic calculus. The integration by parts
formula for the It integral differs from the standard result due to the inclusion of a quadratic covariation term. This
term comes from the fact that It calculus deals with processes with non-zero quadratic variation, which only occurs
for infinite variation processes (such as Brownian motion). If X and Y are semimartingales then
It's lemma
It's lemma is the version of the chain rule or change of variables formula which applies to the It integral. It is one
of the most powerful and frequently used theorems in stochastic calculus. For a continuous d-dimensional
semimartingale X = (X1,,Xd) and twice continuously differentiable function f from Rd to R, it states that f(X) is a
semimartingale and,
This differs from the chain rule used in standard calculus due to the term involving the quadratic covariation [Xi,Xj].
The formula can be generalized to non-continuous semimartingales by adding a pure jump term to ensure that the
jumps of the left and right hand sides agree (see It's lemma).
Martingale integrators
Local martingales
An important property of the It integral is that it preserves the local martingale property. If M is a local martingale
and H is a locally bounded predictable process then HM is also a local martingale. For integrands which are not
locally bounded, there are examples where HM is not a local martingale. However, this can only occur when M is
not continuous. If M is a continuous local martingale then a predictable process H is M-integrable if and only if
0tH2d[M] is finite for each t, and HM is always a local martingale.
The most general statement for a discontinuous local martingale M is that if (H2[M])1/2 is locally integrable then
HM exists and is a local martingale.
15
It calculus
16
This equality holds more generally for any martingale M such that H2[M]t is integrable. The It isometry is often
used as an important step in the construction of the stochastic integral, by defining HM to be the unique extension
of this isometry from a certain class of simple integrands to all bounded and predictable processes.
p-Integrable martingales
For any p > 1, and bounded predictable integrand, the stochastic integral preserves the space of p-integrable
martingales. These are cdlg martingales such that E(|Mt|p) is finite for all t. However, this is not always true in the
case where p = 1. There are examples of integrals of bounded predictable processes with respect to martingales
which are not themselves martingales.
The maximum process of a cadlag process M is written as Mt* = supst|Ms|. For any p1 and bounded predictable
integrand, the stochastic integral preserves the space of cadlag martingales M such that E((Mt*)p) is finite for all t. If
p>1 then this is the same as the space of p-integrable martingales, by Doob's inequalities.
The BurkholderDavisGundy inequalities state that, for any given
which depend on
, but not
or on
such that
for all cadlag local martingales M. These are used to show that if (Mt*)p is integrable and H is a bounded predictable
process then
and, consequently, HM is a p-integrable martingale. More generally, this statement is true whenever (H2[M])p/2 is
integrable.
By a continuous linear extension, the integral extends uniquely to all predictable integrands satisfying
E(0tH2ds)< in such way that the It isometry still holds. It can then be extended to all B-integrable processes by
localization. This method allows the integral to be defined with respect to any It process.
For a general semimartingale X, the decomposition X=M+A for a local martingale M and finite variation process A
can be used. Then, the integral can be shown to exist separately with respect to M and A and combined using
It calculus
17
linearity, HX=HM+HA, to get the integral with respect to X. The standard LebesgueStieltjes integral allows
integration to be defined with respect to finite variation processes, so the existence of the It integral for
semimartingales will follow from any construction for local martingales.
For a cadlag square integrable martingale M, a generalized form of the It isometry can be used. First, the
DoobMeyer decomposition theorem is used to show that a decomposition M2=N+<M> exists, where N is a
martingale and <M> is a right-continuous, increasing and predictable process starting at zero. This uniquely defines
<M>, which is referred to as the predictable quadratic variation of M. The It isometry for square integrable
martingales is then
which can be proved directly for simple predictable integrands. As with the case above for Brownian motion, a
continuous linear extension can be used to uniquely extend to all predictable integrands satisfying E(H2<M>t)<.
This method can be extended to all local square integrable martingales by localization. Finally, the DoobMeyer
decomposition can be used to decompose any local martingale into the sum of a local square integrable martingale
and a finite variation process, allowing the It integral to be constructed with respect to any semimartingale.
Many other proofs exist which apply similar methods but which avoid the need to use the DoobMeyer
decomposition theorem, such as the use of the quadratic variation [M] in the It isometry, the use of the Dolans
measure for submartingales, or the use of the BurkholderDavisGundy inequalities instead of the It isometry. The
latter applies directly to local martingales without having to first deal with the square integrable martingale case.
Alternative proofs exist only making use of the fact that X is cadlag, adapted, and the set {HXt:|H|1 is simple
previsible} is bounded in probability for each time t, which is an alternative definition for X to be a semimartingale.
A continuous linear extension can be used to construct the integral for all left-continuous and adapted integrands
with right limits everywhere (caglad or L-processes). This is general enough to be able to apply techniques such as
It's lemma (Protter 2004). Also, a Khintchine inequality can be used to prove the dominated convergence theorem
and extend the integral to general predictable integrands (Bichteler 2002).
where we used the fact that the covariation of Brownian motion B with itself is just its quadratic variation, which is t.
This stochastic derivative turns out to have many of the properties of the usual derivative of elementary calculus. It
leads to a fundamental theorem of stochastic calculus for this stochastic derivative/integral pair:
and
It also leads to a stochastic mean value theorem, stochastic chain rules, as well as other differentiation rules that are
similar to those in elementary calculus. A key difference is that where an indefinite integral (anti-derivative) in the
usual elementary calculus sense is determined only up to an additive constant of integration, an indefinite integral in
this stochastic calculus is determined only up to a process of bounded variation on compacts. These processes are the
It calculus
18
"constants" in this stochastic differentiation theory. Also, if M and B are orthogonal (zero covariation) then
particular, if M and B are independent then
where
used.
If
is a function of the
Stratonovich sde frequently occur in physics as the limit of a stochastic differential equation with colored noise if the
correlation time of the noise term approaches zero. For a recent treatment of different interpretations of stochastic
differential equations see for example Lau, Lubensky: State-dependent diffusion, Phys. Rev. E, 2007.
There are many equivalent tools to treat Wiener processes other than ItoLangevin, including Stratonovich. The
choice depends on the details of the problem. The traditional tool is actually FokkerPlanck partial differential
equations, with its perfect alternative, Ito stochastic differential equations. The integral methods that are based on
functional probability distributions, known as path or line integrals in physics, are equivalent to Ito stochastic
integral calculus. All the Ito stuff is about the non-linear Wiener (diffusion) processes which forms a sub-class of
Markovian processes. A process is Markovian if the probability distribution of current value and its delta for all
positive deltas depends on the current value and does not depend on the previous values (i.e. no memory).
See also
Stochastic calculus
Wiener process
It's lemma
Stratonovich integral
Semimartingale
References
Allouba, Hassan (2006). "A Differentiation Theory for It's Calculus". Stochastic Analysis and Applications 24:
367380. doi:10.1080/07362990500522411 [1].
Bichteler, Klaus (2002), Stochastic Integration With Jumps (1st ed.), Cambridge University Press, ISBN
0-521-81129-5
Hagen Kleinert (2004). Path Integrals in Quantum Mechanics, Statistics, Polymer Physics, and Financial
Markets, 4th edition, World Scientific (Singapore); Paperback ISBN 981-238-107-4. Fifth edition available
online: PDF-files [2], with generalizations of It's lemma for non-Gaussian processes.
He, Sheng-wu; Wang, Jia-gang; Yan, Jia-an (1992), Semimartingale Theory and Stochastic Calculus, Science
Press, CRC Press Inc., ISBN 7-03-003066-4,0-8493-7715-3
; in
It calculus
19
Karatzas, Ioannis; Shreve, Steven (1991), Brownian Motion and Stochastic Calculus (2nd ed.), Springer, ISBN
0-387-97655-8
Protter, Philip E. (2004), Stochastic Integration and Differential Equations (2nd ed.), Springer, ISBN
3-540-00313-4
ksendal, Bernt K. (2003). Stochastic Differential Equations: An Introduction with Applications. Berlin:
Springer. ISBN 3-540-04758-1.
Mathematical Finance Programming in TI-Basic, which implements Ito calculus for TI-calculators.
References
[1] http:/ / dx. doi. org/ 10. 1080%2F07362990500522411
[2] http:/ / www. physik. fu-berlin. de/ ~kleinert/ b5
Lebesgue integration
In mathematics, Lebesgue integration refers to both the general
theory of integration of a function with respect to a general measure,
and to the specific case of integration of a function defined on a
sub-domain of the real line or a higher dimensional Euclidean space
with respect to the Lebesgue measure. This article focuses on the more
general concept.
Lebesgue integration plays an important role in real analysis, the
axiomatic theory of probability, and many other fields in the
mathematical sciences.
Introduction
The integral of a function f between limits a and b can be interpreted as the area under the graph of f. This is easy to
understand for familiar functions such as polynomials, but what does it mean for more exotic functions? In general,
what is the class of functions for which "area under the curve" makes sense? The answer to this question has great
theoretical and practical importance.
As part of a general movement toward rigour in mathematics in the nineteenth century, attempts were made to put
the integral calculus on a firm foundation. The Riemann integral, proposed by Bernhard Riemann (18261866), is
a broadly successful attempt to provide such a foundation. Riemann's definition starts with the construction of a
sequence of easily-calculated areas which converge to the integral of a given function. This definition is successful
in the sense that it gives the expected answer for many already-solved problems, and gives useful results for many
other problems.
However, Riemann integration does not interact well with taking limits of sequences of functions, making such
limiting processes difficult to analyze. This is of prime importance, for instance, in the study of Fourier series,
Lebesgue integration
Fourier transforms and other topics. The Lebesgue integral is better able to describe how and when it is possible to
take limits under the integral sign. The Lebesgue definition considers a different class of easily-calculated areas than
the Riemann definition, which is the main reason the Lebesgue integral is better behaved. The Lebesgue definition
also makes it possible to calculate integrals for a broader class of functions. For example, the Dirichlet function,
which is 0 where its argument is irrational and 1 otherwise, has a Lebesgue integral, but it does not have a Riemann
integral.
Measure theory
Measure theory was initially created to provide a useful abstraction of the notion of length of subsets of the real line
and, more generally, area and volume of subsets of Euclidean spaces. In particular, it provided a systematic answer
to the question of which subsets of R have a length. As was shown by later developments in set theory (see
non-measurable set), it is actually impossible to assign a length to all subsets of R in a way which preserves some
natural additivity and translation invariance properties. This suggests that picking out a suitable class of measurable
subsets is an essential prerequisite.
The Riemann integral uses the notion of length explicitly. Indeed, the element of calculation for the Riemann integral
is the rectangle [a,b][c,d], whose area is calculated to be (ba)(dc). The quantity ba is the length of the
base of the rectangle and dc is the height of the rectangle. Riemann could only use planar rectangles to
approximate the area under the curve because there was no adequate theory for measuring more general sets.
In the development of the theory in most modern textbooks (after 1950), the approach to measure and integration is
axiomatic. This means that a measure is any function defined on a certain class X of subsets of a set E, which
satisfies a certain list of properties. These properties can be shown to hold in many different cases.
Integration
We start with a measure space (E,X,) where E is a set, X is a -algebra of subsets of E and is a (non-negative)
measure on X of subsets of E.
For example, E can be Euclidean n-space Rn or some Lebesgue measurable subset of it, X will be the -algebra of all
Lebesgue measurable subsets of E, and will be the Lebesgue measure. In the mathematical theory of probability,
we confine our study to a probability measure, which satisfies
.
In Lebesgue's theory, integrals are defined for a class of functions called measurable functions. A function is
measurable if the pre-image of every closed interval is in X:
It can be shown that this is equivalent to requiring that the pre-image of any Borel subset of R be in X. We will make
this assumption henceforth. The set of measurable functions is closed under algebraic operations, but more
importantly the class is closed under various kinds of pointwise sequential limits:
are measurable if the original sequence (k)k, where k N, consists of measurable functions.
We build up an integral
20
Lebesgue integration
where the coefficients ak are real numbers and the sets Sk are measurable, is called a measurable simple function. We
extend the integral by linearity to non-negative measurable simple functions. When the coefficients ak are
non-negative, we set
The convention 0= 0 must be used, and the result may be infinite. Even if a simple function can be written in
many ways as a linear combination of indicator functions, the integral will always be the same.
Some care is needed when defining the integral of a real-valued simple function, in order to avoid the undefined
expression : one assumes that the representation
is such that (Sk)< whenever ak0. Then the above formula for the integral of makes sense, and the result does
not depend upon the particular representation of satisfying the assumptions.
If B is a measurable subset of E and s a measurable simple function one defines
Non-negative functions: Let be a non-negative measurable function on E which we allow to attain the value +, in
other words, takes non-negative values in the extended real number line. We define
We need to show this integral coincides with the preceding one, defined on the set of simple functions. When E is a
segment [a,b], there is also the question of whether this corresponds in any way to a Riemann notion of integration.
It is possible to prove that the answer to both questions is yes.
We have defined the integral of for any non-negative extended real-valued measurable function onE. For some
functions, this integral Ed will be infinite.
Signed functions: To handle signed functions, we need a few more definitions. If is a measurable function of the
set E to the reals (including ), then we can write
where
21
Lebesgue integration
Note that both + and are non-negative measurable functions. Also note that
If
It turns out that this definition gives the desirable properties of the integral.
Complex valued functions can be similarly integrated, by considering the real part and the imaginary part
separately.
Intuitive interpretation
To get some intuition about the different approaches to integration, let us imagine that it is desired to find a
mountain's volume (above sea level).
The Riemann-Darboux approach: Divide the base of the mountain into a grid of 1 meter squares. Measure the
altitude of the mountain at the center of each square. The volume on a single grid square is approximately
1x1x(altitude), so the total volume is the sum of the altitudes.
The Lebesgue approach: Draw a contour map of the mountain, where each contour is 1 meter of altitude apart. The
volume of earth contained in a single contour is approximately that contour's area times its height. So the total
volume is the sum of these volumes.
Folland [1] summarizes the difference between the Riemann and Lebesgue approaches thus: "to compute the
Riemann integral of f, one partitions the domain [a,b] into subintervals", while in the Lebesgue integral, "one is in
effect partitioning the range of f".
See also Properties of simple functions.
Example
Consider the indicator function of the rational numbers, 1Q. This function is nowhere continuous.
is not Riemann-integrable on [0,1]: No matter how the set [0,1] is partitioned into subintervals, each
partition will contain at least one rational and at least one irrational number, since rationals and irrationals are
both dense in the reals. Thus the upper Darboux sums will all be one, and the lower Darboux sums will all be
zero.
is Lebesgue-integrable on [0,1] using the Lebesgue measure: Indeed it is the indicator function of the
rationals so by definition
since
is countable.
22
Lebesgue integration
23
The function gk is zero everywhere except on a finite set of points, hence its Riemann integral is zero. The sequence
gk is also clearly non-negative and monotonically increasing to 1Q, which is not Riemann integrable.
Unsuitability for unbounded intervals. The Riemann integral can only integrate functions on a bounded interval. It
can however be extended to unbounded intervals by taking limits, so long as this doesn't yield an answer such as
.
What about integrating on structures other than Euclidean space? The Riemann integral is inextricably linked to
the order structure of the line. How do we free ourselves of this limitation?
If f, g are non-negative measurable functions (possibly assuming the value +) such that f= g almost everywhere,
then
Monotonicity: If f g, then
Monotone convergence theorem: Suppose {fk}kN is a sequence of non-negative measurable functions such that
Lebesgue integration
Then
Proof techniques
To illustrate some of the proof techniques used in Lebesgue integration theory, we sketch a proof of the above
mentioned Lebesgue monotone convergence theorem. Let {fk}kN be a non-decreasing sequence of non-negative
measurable functions and put
and the limit on the right exists, since the sequence is monotonic. We now prove the inequality in the other direction.
It follows from the definition of integral that there is a non-decreasing sequence (gn) of non-negative simple
functions such that gn f and
By breaking up the function g into its constant value parts, this reduces to the case in which g is the indicator
function of a set. The result we have to prove is then
Suppose A is a measurable set and {fk}kN is a nondecreasing sequence of non-negative measurable functions
on E such that
24
Lebesgue integration
To prove this result, fix > 0 and define the sequence of measurable sets
Because of the fact that almost every x will be in Bk for large enough k, we have
Alternative formulations
It is possible to develop the integral with respect to the Lebesgue measure without relying on the full machinery of
measure theory. One such approach is provided by Daniell integral.
There is also an alternative approach to developing the theory of integration via methods of functional analysis. The
Riemann integral exists for any continuous function f of compact support defined on Rn (or a fixed open subset).
Integrals of more general functions can be built starting from these integrals. Let Cc be the space of all real-valued
compactly supported continuous functions of R. Define a norm on Cc by
Then Cc is a normed vector space (and in particular, it is a metric space.) All metric spaces have Hausdorff
completions, so let L1 be its completion. This space is isomorphic to the space of Lebesgue integrable functions
modulo the subspace of functions with integral zero. Furthermore, the Riemann integral is a uniformly continuous
functional with respect to the norm on Cc, which is dense in L1. Hence has a unique extension to all of L1. This
integral is precisely the Lebesgue integral.
This approach can be generalised to build the theory of integration with respect to Radon measures on locally
compact spaces. It is the approach adopted by Bourbaki (2004); for more details see Radon measures on locally
compact spaces.
See also
25
Lebesgue integration
References
Bartle, Robert G. (1995). The elements of integration and Lebesgue measure. Wiley Classics Library. New York:
John Wiley & Sons Inc.. xii+179. ISBN 0-471-04222-6. MR1312157 [2]
Bourbaki, Nicolas (2004). Integration. I. Chapters 16. Translated from the 1959, 1965 and 1967 French
originals by Sterling K. Berberian. Elements of Mathematics (Berlin). Berlin: Springer-Verlag. xvi+472. ISBN
3-540-41129-1. MR2018901 [3]
Dudley, Richard M. (1989). Real analysis and probability. The Wadsworth &ammp; Brooks/Cole Mathematics
Series. Pacific Grove, CA: Wadsworth & Brooks/Cole Advanced Books & Software. xii+436. ISBN
0-534-10050-3. MR982264 [4] Very thorough treatment, particularly for probabilists with good notes and
historical references.
Folland, Gerald B. (1999). Real analysis: Modern techniques and their applications. Pure and Applied
Mathematics (New York) (Second ed.). New York: John Wiley & Sons Inc.. xvi+386. ISBN 0-471-31716-0.
MR1681462 [5]
Halmos, Paul R. (1950). Measure Theory. New York, N. Y.: D. Van Nostrand Company, Inc.. pp.xi+304.
MR0033869 [6] A classic, though somewhat dated presentation.
Lebesgue, Henri (1904), Leons sur l'intgration et la recherche des fonctions primitives, Paris: Gauthier-Villars
Lebesgue, Henri (1972) (in French). Oeuvres scientifiques (en cinq volumes). Geneva: Institut de Mathmatiques
de l'Universit de Genve. pp.405. MR0389523 [7]
Loomis, Lynn H. (1953). An introduction to abstract harmonic analysis. Toronto-New York-London: D. Van
Nostrand Company, Inc.. pp.x+190. MR0054173 [8] Includes a presentation of the Daniell integral.
Munroe, M. E. (1953). Introduction to measure and integration. Cambridge, Mass.: Addison-Wesley Publishing
Company Inc.. pp.x+310. MR0053186 [9] Good treatment of the theory of outer measures.
Royden, H. L. (1988). Real analysis (Third ed.). New York: Macmillan Publishing Company. pp.xx+444. ISBN
0-02-404151-3. MR1013117 [10]
Rudin, Walter (1976). Principles of mathematical analysis. International Series in Pure and Applied Mathematics
(Third ed.). New York: McGraw-Hill Book Co.. pp.x+342. MR0385023 [11] Known as Little Rudin, contains the
basics of the Lebesgue theory, but does not treat material such as Fubini's theorem.
Rudin, Walter (1966). Real and complex analysis. New York: McGraw-Hill Book Co.. pp.xi+412. MR0210528
[12]
Known as Big Rudin. A complete and careful presentation of the theory. Good presentation of the Riesz
extension theorems. However, there is a minor flaw (in the first edition) in the proof of one of the extension
theorems, the discovery of which constitutes exercise 21 of Chapter 2.
Shilov, G. E.; Gurevich, B. L. (1977). Integral, measure and derivative: a unified approach. Translated from the
Russian and edited by Richard A. Silverman. Dover Books on Advanced Mathematics. New York: Dover
Publications Inc.. xiv+233. ISBN 0-486-63519-8. MR0466463 [13] Emphasizes the Daniell integral.
26
Lebesgue integration
References
[1] Gerald B. Folland, Real Analysis: Modern Techniques and Their Applications, 1984, p. 56.
[2] http:/ / www. ams. org/ mathscinet-getitem?mr=1312157
[3] http:/ / www. ams. org/ mathscinet-getitem?mr=2018901
[4] http:/ / www. ams. org/ mathscinet-getitem?mr=982264
[5] http:/ / www. ams. org/ mathscinet-getitem?mr=1681462
[6] http:/ / www. ams. org/ mathscinet-getitem?mr=0033869
[7] http:/ / www. ams. org/ mathscinet-getitem?mr=0389523
[8] http:/ / www. ams. org/ mathscinet-getitem?mr=0054173
[9] http:/ / www. ams. org/ mathscinet-getitem?mr=0053186
[10] http:/ / www. ams. org/ mathscinet-getitem?mr=1013117
[11] http:/ / www. ams. org/ mathscinet-getitem?mr=0385023
[12] http:/ / www. ams. org/ mathscinet-getitem?mr=0210528
[13] http:/ / www. ams. org/ mathscinet-getitem?mr=0466463
LebesgueStieltjes integration
In measure-theoretic analysis and related branches of mathematics, LebesgueStieltjes integration generalizes
RiemannStieltjes and Lebesgue integration, preserving the many advantages of the former in a more general
measure-theoretic framework. The LebesgueStieltjes integral is the ordinary Lebesgue integral with respect to a
measure known as the LebesgueStieltjes measure, which may be associated to any function of bounded variation on
the real line. The LebesgueStieltjes measure is a regular Borel measure, and conversely every regular Borel
measure on the real line is of this kind.
LebesgueStieltjes integrals, named for Henri Leon Lebesgue and Thomas Joannes Stieltjes, are also known as
LebesgueRadon integrals or just Radon integrals, after Johann Radon, to whom much of the theory is due. They
find common application in probability and stochastic processes, and in certain branches of analysis including
potential theory.
Definition
The LebesgueStieltjes integral
is defined when :[a,b]R is Borel-measurable and bounded and g:[a,b]R is of bounded variation in [a,b]
and right-continuous, or when is non-negative and g is monotone and right-continuous. To start, assume that is
non-negative and g is monotone non-decreasing and right-continuous, and define w((s,t]):=g(t)g(s).
(Alternatively, the construction works for g left-continuous and w([s,t)):=g(t)g(s).)
By Carathodory's extension theorem, there is a unique Borel measure g on [a,b] which agrees with w on every
interval I. The measure g arises from an outer measure (in fact, a metric outer measure) given by
the infimum taken over all coverings of E by countably many semiopen intervals. This measure is sometimes
called[1] the LebesgueStieltjes measure associated with g.
The LebesgueStieltjes integral
27
LebesgueStieltjes integration
28
is defined as the Lebesgue integral of with respect to the measure g in the usual way. If g is non-increasing, then
define
where g1(x) := Vxag is the total variation of g in the interval [a,x], and g2(x)=g1(x)g(x). Both g1 and g2 are
monotone non-decreasing. Now the LebesgueStieltjes integral with respect to g is defined by
where the latter two integrals are well-defined by the preceding construction.
Daniell integral
An alternative approach (Hewitt & Stromberg 1965) is to define the LebesgueStieltjes integral as the Daniell
integral that extends the usual RiemannStieltjes integral. Let g be a non-increasing right-continuous function on
[a,b], and define I() to be the RiemannStieltjes integral
for all continuous functions . The functional I defines a Radon measure on [a,b]. This functional can then be
extended to the class of all non-negative functions by setting
and
and either side of the identity then defines the LebesgueStieltjes integral of h. The outer measure g is defined via
where A is the indicator function of A.
Integrators of bounded variation are handled as above by decomposing into positive and negative variations.
Example
Suppose that
we may define the length of
where
to
,
. This notion
is quite useful for various applications: for example, in muddy terrain the speed in which a person can move may
depend on how deep the mud is. If
denotes the inverse of the walking speed at or near , then the -length
of
-length of
LebesgueStieltjes integration
29
Integration by parts
A function
and
where
and
or
is continuous, or if both
are regular, then there is an integration by parts formula for the LebesgueStieltjes integral:
and
can be dropped.[2]
Related concepts
Lebesgue integration
When g(x)=x for all real x, then g is the Lebesgue measure, and the LebesgueStieltjes integral of f with respect to
g is equivalent to the Lebesgue integral of f.
for the LebesgueStieltjes integral, letting the measure v remain implicit. This is particularly common in probability
theory when v is the cumulative distribution function of a real-valued random variable X, in which case
(See the article on RiemannStieltjes integration for more detail on dealing with such cases.)
References
Halmos, Paul R. (1974), Measure Theory, Berlin, New York: Springer-Verlag, ISBN 978-0-387-90088-9
Hewitt, Edwin; Stromberg, Karl (1965), Real and abstract analysis, Springer-Verlag.
Saks, Stanislaw (1937) Theory of the Integral. [3]
Shilov, G. E., and Gurevich, B. L., 1978. Integral, Measure, and Derivative: A Unified Approach, Richard A.
Silverman, trans. Dover Publications. ISBN 0-486-63519-8.
LebesgueStieltjes integration
External links
www.probability.net Probability and foundations tutorial. [4]
References
[1] Halmos (1974), Sec. 15
[2] Hewitt, Edwin (5 1960). " Integration by Parts for Stieltjes Integrals (http:/ / www. jstor. org/ pss/ 2309287)". The American Mathematical
Monthly 67 (5): 419423. doi: 10.2307/2309287 (http:/ / dx. doi. org/ 10. 2307/ 2309287). . Retrieved 2008-04-23.
[3] http:/ / matwbn. icm. edu. pl/ kstresc. php?tom=7& wyd=10
[4] http:/ / www. probability. net/
Motivic integration
Motivic integration is a branch of algebraic geometry which was invented by Maxim Kontsevich in 1995 and was
developed by Jan Denef and Franois Loeser. Since its introduction it has proved to be quite useful in various
branches of algebraic geometry, most notably birational geometry and singularity theory. Roughly speaking, motivic
integration assigns to subsets of the arc space of an algebraic geometry a volume living in the Grothendieck ring of
algebraic varieties. The naming 'motivic' mirrors the fact that unlike ordinary integration, for which the values are
real numbers, in motivic integration the values are geometric in nature.
External links
AMS Bulletin Vol. 42 [1]Tom Hales
What is motivic measure?, an excellent introduction.
math.AG/9911179 [2] A.Craw
An introduction to motivic integration
Lecture Notes [3] Franois Loeser
Seattle lecture notes on motivic integration
Lecture Notes [4] W.Veys
Arc spaces, motivic integration and stringy invariants
References
[1]
[2]
[3]
[4]
30
PaleyWiener integral
PaleyWiener integral
In mathematics, the PaleyWiener integral is a simple stochastic integral. When applied to classical Wiener space,
it is less general than the It integral, but the two agree when they are both defined.
The integral is named after its discoverers, Raymond Paley and Norbert Wiener.
Definition
Let i:HE be an abstract Wiener space with abstract Wiener measure on E. Let j:EH be the adjoint of i.
(We have abused notation slightly: strictly speaking, j:EH, but since H is a Hilbert space, it is isometrically
isomorphic to its dual space H, by the Riesz representation theorem.)
It can be shown that j is an injective function and has dense image in H. Furthermore, it can be shown that every
linear functional fE is also square-integrable: in fact,
This defines a natural linear map from j(E) to L2(E,;R), under which j(f)j(E)H goes to the equivalence class
[f] of f in L2(E,;R). This is well-defined since j is injective. This map is an isometry, so it is continuous.
However, since a continuous linear map between Banach spaces such as H and L2(E,;R) is uniquely determined by
its values on any dense subspace of its domain, there is a unique continuous linear extension I:HL2(E,;R) of
the above natural map j(E)L2(E,;R) to the whole of H.
This isometry I:HL2(E,;R) is known as the PaleyWiener map. I(h), also denoted h,, is a function on E
and is known as the PaleyWiener integral (with respect to hH).
It is important to note that the PaleyWiener integral for a particular element hH is a function on E. The notation
h,x does not really denote an inner product (since h and x belong to two different spaces), but is a convenient
abuse of notation in view of the CameronMartin theorem. For this reason, many authors prefer to write h,(x) or
I(h)(x) rather than using the more compact but potentially confusing h,x notation.
See also
Other stochastic integrals:
It integral
Skorokhod integral
Stratonovich integral
31
Pfeffer integral
32
Pfeffer integral
In mathematics, the Pfeffer integral is an integration technique created by Washek Pfeffer as an attempt to extend
the Henstock integral to a multidimensional domain. This was to be done in such a way that the fundamental
theorem of calculus would apply analogously to the theorem in one dimension, with as few preconditions on the
function under consideration as possible. The integral also permits analogues of the chain rule and other theorems of
the integral calculus for higher dimensions.
Definition
The construction is based on the Henstock or gauge integral, however Pfeffer proved that the integral, at least in the
one dimensional case, is less general than the Henstock integral. It relies on what Pfeffer refers to as a set of
bounded variation, this is equivalent to a Caccioppoli set. The Riemann sums of the Pfeffer integral are taken over
partitions made up of such sets, rather than intervals as in the Riemann or Henstock integrals. A gauge is used,
exactly as in the Henstock integral, except that the gauge function may be zero on a negligible set.
Properties
Pfeffer defined a notion of generalized absolute continuity
function being
function. He also proved a chain rule for the Pfeffer integral. In one dimension his work as well as similarities
between the Pfeffer integral and the McShane integral indicate that the integral is more general than the Lebesgue
integral and yet less general than the Henstock integral.
Bibliography
Bongiorno, Benedetto & Pfeffer, Washek (1992). A concept of absolute continuity and a Riemann type integral.
Comment. Math. Univ. Carolinae 33.2:189196
Pfeffer, Washek (1992). A Riemann type definition of a variational integral. Proc. American Math. Soc.
114:99106
Regulated integral
Regulated integral
In mathematics, the regulated integral is a definition of integration for regulated functions, which are defined to be
uniform limits of step functions. The use of the regulated integral instead of the Riemann integral has been
advocated by Nicolas Bourbaki and Jean Dieudonn.
Definition
Definition on step functions
Let [a, b] be a fixed closed, bounded interval in the real line R. A real-valued function : [a, b] R is called a step
function if there exists a finite partition
of [a, b] such that is constant on each open interval (ti, ti+1) of ; suppose that this constant value is ci R. Then,
define the integral of a step function to be
It can be shown that this definition is independent of the choice of partition, in that if 1 is another partition of [a, b]
such that is constant on the open intervals of 1, then the numerical value of the integral of is the same for 1 as
for .
exists as well.
Define the integral of a regulated function f to be
33
Regulated integral
The integral is also a bounded operator: every regulated function f is bounded, and if m f(t) M for all t [a,
b], then
In particular:
Since step functions are integrable and the integrability and the value of a Riemann integral are compatible with
uniform limits, the regulated integral is a special case of the Riemann integral.
References
Berberian, S.K. (1979). "Regulated Functions: Bourbaki's Alternative to the Riemann Integral". The American
Mathematical Monthly 86: 208. doi:10.2307/2321526 [1].
Gordon, Russell A. (1994). The integrals of Lebesgue, Denjoy, Perron, and Henstock. Graduate Studies in
Mathematics, 4. Providence, RI: American Mathematical Society. ISBN 0-8218-3805-9.
See also
Lebesgue integral
Riemann integral
References
[1] http:/ / dx. doi. org/ 10. 2307%2F2321526
34
Riemann integral
35
Riemann integral
In the branch of mathematics known as real analysis, the Riemann integral, created by Bernhard Riemann, was the
first rigorous definition of the integral of a function on an interval. While the Riemann integral is unsuitable for
many theoretical purposes, it is one of the easiest integrals to define. Some of these technical deficiencies can be
remedied by the RiemannStieltjes integral, and most of them disappear in the Lebesgue integral.
Overview
Let
, and let
be the
The basic idea of the Riemann integral is to use very simple approximations for the area of
better approximations, we can say that "in the limit" we get exactly the area of
Note that where can be both positive and negative, the integral corresponds to signed area under the graph of ; that
is, the area above the x-axis minus the area below the x-axis.
Definition
Riemann integral
36
Partitions of an interval
A partition of an interval
is a finite sequence
. Each
is
called a subinterval of the partition. The mesh of a partition is defined to be the length of the longest subinterval
, that is, it is
where
. It is also called the norm of the partition.
A tagged partition of an interval is a partition of an interval together with a finite sequence of numbers
subject to the conditions that for each ,
. In other words, it is a partition together
with a distinguished point of every subinterval. The mesh of a tagged partition is the same as that of an ordinary
partition.
Suppose that
together with
together with
. We say that
together with
such that
and
to equal
, and that
for some
because
with
with
, there
. (It is
refinement of a tagged partition takes the starting partition and adds more tags, but does not take any away.
We can define a partial order on the set of all tagged partitions by saying that one tagged partition is bigger than
another if the bigger one is a refinement of the smaller one.
Riemann sums
Choose a real-valued function
tagged partition
together with
is:
Each term in the sum is the product of the value of the function at a given point and the length of an interval.
Consequently, each term represents the area of a rectangle with height
and length
. The Riemann
sum is the signed area under all the rectangles.
Riemann integral
Loosely speaking, the Riemann integral is the limit of the Riemann sums of a function as the partitions get finer. If
the limit exists then the function is said to be integrable (or more specifically Riemann-integrable). The Riemann
sum can be made as close as desired to the Riemann integral by making the partition fine enough.
One important fact is that the mesh of the partitions must become smaller and smaller, so that in the limit, it is zero.
If this were not so, then we would not be getting a good approximation to the function on certain subintervals. In
fact, this is enough to define an integral. To be specific, we say that the Riemann integral of equals s if the
following condition holds:
For all >0, there exists >0 such that for any tagged partition
and
whose mesh
However, there is an unfortunate problem with this definition: it is very difficult to work with. So we will make an
alternate definition of the Riemann integral which is easier to work with, then prove that it is the same as the
definition we have just made. Our new definition says that the Riemann integral of equals s if the following
condition holds:
For all >0, there exists a tagged partition
and
of
and
and
Riemann integral
37
Both of these mean that eventually, the Riemann sum of with respect to any partition gets trapped close to s. Since
this is true no matter how close we demand the sums be trapped, we say that the Riemann sums converge to s. These
definitions are actually a special case of a more general concept, a net.
As we stated earlier, these two definitions are equivalent. In other words, s works in the first definition if and only if
s works in the second definition. To show that the first definition implies the second, start with an , and choose a
that satisfies the condition. Choose any tagged partition whose mesh is less than . Its Riemann sum is within of s,
and any refinement of this partition will also have mesh less than , so the Riemann sum of the refinement will also
be within of s. To show that the second definition implies the first, it is easiest to use the Darboux integral. First
one shows that the second definition is equivalent to the definition of the Darboux integral; for this see the page on
Darboux integration. Now we will show that a Darboux integrable function satisfies the first definition. Fix , and
choose a partition
such that the lower and upper Darboux sums with respect to this partition are within
/2 of the value s of the Darboux integral. Let r equal the supremum of |(x)| on [a,b]. If r=0, then is the zero
function, which is clearly both Darboux and Riemann integrable with integral zero. Therefore we will assume that
r>0. If m>1, then we choose to be less than both /2r(m1) and
. If m=1, then we
choose to be less than one. Choose a tagged partition
and
(We may assume that all the inequalities are strict because otherwise we are in the previous case by our assumption
on the length of .) This can happen at most m1 times. To handle this case, we will estimate the difference
between the Riemann sum and the Darboux sum by subdividing the partition
at yj+1. The term
(ti)(xixi+1) in the Riemann sum splits into two terms:
Suppose that ti[xi,xi+1]. Then mj(ti)Mj, so this term is bounded by the corresponding term in the Darboux
sum for yj. To bound the other term, notice that yj+1xi+1 is smaller than , and is chosen to be smaller than
/2r(m1), where r is the supremum of |(x)|. It follows that the second term is smaller than /2(m1). Since this
happens at most m1 times, the total of all the terms which are not bounded by the Darboux sum is at most /2.
Therefore the distance between the Riemann sum and s is at most .
Riemann integral
38
Examples
Let
be the function which takes the value 1 at every point. Any Riemann sum of
on
on
will
is 1.
; that is,
rational numbers and 0 on irrational numbers. This function does not have a Riemann integral. To prove this, we will
show how to construct tagged partitions whose Riemann sums get arbitrarily close to both zero and one.
To start, let
and
be a tagged partition (each
is between
and
). Choose
. The
new tags, we can make the value of the Riemann sum turn out to be within of either zero or oneour choice!
Our first step is to cut up the partition. There are of the , and we want their total effect to be less than . If we
confine each of them to an interval of length less than
will be at least
and at most
is within
of some
, and
are within
is not equal to
, choose
. We still have to choose tags for the other subintervals. We will choose them in
two different ways. The first way is to always choose a rational point, so that the Riemann sum is as large as
possible. This will make the value of the Riemann sum at least
irrational point, so that the Riemann sum is as small as possible. This will make the value of the Riemann sum at
most .
Since we started from an arbitrary partition and ended up as close as we wanted to either zero or one, it is false to say
that we are eventually trapped near some number , so this function is not Riemann integrable. However, it is
Lebesgue integrable. In the Lebesgue sense its integral is zero, since the function is zero almost everywhere. But this
is a fact that is beyond the reach of the Riemann integral.
There are even worse examples.
function, but there are non-Riemann integrable bounded functions which are not equivalent to any Riemann
integrable function. For example, let C be the SmithVolterraCantor set, and let IC be its indicator function.
Because C is not Jordan measurable, IC is not Riemann integrable. Moreover, no function g equivalent to IC is
Riemann integrable: g, like IC, must be zero on a dense set, so as in the previous example, any Riemann sum of g has
a refinement which is within of 0 for any positive number . But if the Riemann integral of g exists, then it must
equal the Lebesgue integral of IC, which is 1/2. Therefore g is not Riemann integrable.
Riemann integral
39
Similar concepts
It is popular to define the Riemann integral as the Darboux integral. This is because the Darboux integral is
technically simpler and because a function is Riemann-integrable if and only if it is Darboux-integrable.
Some calculus books do not use general tagged partitions, but limit themselves to specific types of tagged partitions.
If the type of partition is limited too much, some non-integrable functions may appear to be integrable.
One popular restriction is the use of "left-hand" and "right-hand" Riemann sums. In a left-hand Riemann sum,
for all , and in a right-hand Riemann sum,
for all . Alone this restriction does not impose a
problem: we can refine any partition in a way that makes it a left-hand or right-hand sum by subdividing it at each
. In more formal language, the set of all left-hand Riemann sums and the set of all right-hand Riemann sums is
cofinal in the set of all tagged partitions.
Another popular restriction is the use of regular subdivisions of an interval. For example, the
subdivision of
th regular
restriction does not impose a problem, but the reasoning required to see this fact is more difficult than in the case of
left-hand and right-hand Riemann sums.
However, combining these restrictions, so that one uses only left-hand or right-hand Riemann sums on regularly
divided intervals, is dangerous. If a function is known in advance to be Riemann integrable, then this technique will
give the correct value of the integral. But under these conditions the indicator function
will appear to be
integrable on
with integral equal to one: Every endpoint of every subinterval will be a rational number, so the
function will always be evaluated at rational numbers, and hence it will appear to always equal one. The problem
with this definition becomes apparent when we try to split the integral into two pieces. The following equation ought
to hold:
If we use regular subdivisions and left-hand or right-hand Riemann sums, then the two terms on the left are equal to
zero, since every endpoint except 0 and 1 will be irrational, but as we have seen the term on the right will equal 1.
As defined above, the Riemann integral avoids this problem by refusing to integrate
Properties
The Riemann integral is a linear transformation; that is, if
and
are Riemann-integrable on
and
and
Because the Riemann integral of a function is a number, this makes the Riemann integral a linear functional on the
vector space of Riemann-integrable functions. It can be shown that a real-valued function
on
is
Riemann-integrable if and only if it is bounded and continuous almost everywhere in the sense of Lebesgue measure.
If a real-valued function on
If
is Riemann-integrable, it is Lebesgue-integrable.
Riemann integrability of
with limit
implies
, and
Riemann integral
40
Generalizations
It is easy to extend the Riemann integral to functions with values in the Euclidean vector space
integral is defined by linearity; in other words, if
for any
. The
, then
. In
particular, since the complex numbers are a real vector space, this allows the integration of complex valued
functions.
The Riemann integral is only defined on bounded intervals, and it does not extend well to unbounded intervals. The
simplest possible extension is to define such an integral as a limit, in other words, as an improper integral. We could
set:
Unfortunately, this does not work well. Translation invariance, the fact that the Riemann integral of the function
should not change if we move the function left or right, is lost. For example, let
for all
,
, and
for all
for all
for all
. But if we shift
. Then,
, we get
above, we get
tends to
This is also unacceptable, so we could require that the integral exists and gives the same value regardless of the
order. Even this does not give us what we want, because the Riemann integral no longer commutes with uniform
limits. For example, let
on
. But
. Even
though this is the correct value, it shows that the most important criterion for exchanging limits and (proper)
integrals is false for improper integrals. This makes the Riemann integral unworkable in applications.
A better route is to abandon the Riemann integral for the Lebesgue integral. The definition of the Lebesgue integral
is not obviously a generalization of the Riemann integral, but it is not hard to prove that every Riemann-integrable
function is Lebesgue-integrable and that the values of the two integrals agree whenever they are both defined.
Moreover, a function defined on a bounded interval is Riemann-integrable if and only if it is bounded and the set
of points where
An integral which is in fact a direct generalization of the Riemann integral is the HenstockKurzweil integral.
Another way of generalizing the Riemann integral is to replace the factors
in the definition of a
Riemann sum by something else; roughly speaking, this gives the interval of integration a different notion of length.
This is the approach taken by the RiemannStieltjes integral.
Riemann integral
See also
Antiderivative
RiemannStieltjes integral
HenstockKurzweil integral
Lebesgue integral
Darboux integral
References
Shilov, G. E., and Gurevich, B. L., 1978. Integral, Measure, and Derivative: A Unified Approach, Richard A.
Silverman, trans. Dover Publications. ISBN 0-486-63519-8.
References
[1] http:/ / planetmath. org/ encyclopedia/ Volume. html
RiemannStieltjes integral
In mathematics, the RiemannStieltjes integral is a generalization of the Riemann integral, named after
Bernhard Riemann and Thomas Joannes Stieltjes.
Definition
The RiemannStieltjes integral of a real-valued function f of a real variable with respect to a real function g is
denoted by
where ci is in the i-th subinterval [xi, xi+1]. The two functions f and g are respectively called the integrand and the
integrator.
The "limit" is here understood in the following sense: there exists a certain number A (the value of the
Riemann-Stieltjes integral) such that for every > 0 there exists a partition P such that for every partition P with
mesh(P) < mesh(P), and for every choice of points ci in [xi, xi+1],
41
RiemannStieltjes integral
Darboux sums
The RiemannStieltjes integral can be efficiently handled using an appropriate generalization of Darboux sums. For
a partition P define the upper Darboux sum of f with respect to g by
If g is a nondecreasing function on [a,b], then f is RiemannStieltjes integrable with respect to g if and only if, for
every > 0, there exists a partition P such that
for example, if the derivative is unbounded. But if the derivative is continuous, they will be the same. This condition
is also satisfied if g is the (Lebesgue) integral of its derivative; in this case g is said to be absolutely continuous.
However, g may have jump discontinuities, or may have derivative zero almost everywhere while still being
continuous and increasing (for example, g could be the Cantor function), in either of which cases the
RiemannStieltjes integral is not captured by any expression involving derivatives of g.
The RiemannStieltjes integral admits integration by parts in the form
and the existence of the integral on the left implies the existence of the integral on the right.
42
RiemannStieltjes integral
But this formula does not work if X does not have a probability density function with respect to Lebesgue measure.
In particular, it does not work if the distribution of X is discrete (i.e., all of the probability is accounted for by
point-masses), and even if the cumulative distribution function g is continuous, it does not work if g fails to be
absolutely continuous (again, the Cantor function may serve as an example of this failure). But the identity
holds if g is any cumulative probability distribution function on the real line, no matter how ill-behaved.
Generalization
An important generalization is the LebesgueStieltjes integral which generalizes the RiemannStieltjes integral in a
way analogous to how the Lebesgue integral generalizes the Riemann integral. If improper RiemannStieltjes
integrals are allowed, the Lebesgue integral is not strictly more general than the RiemannStieltjes integral.
The RiemannStieltjes integral also generalizes to the case when either the integrand or the integrator g take values
in a Banach space. If g : [a,b] X takes values in the Banach space X, then it is natural to assume that it is of
strongly bounded variation, meaning that
of the interval [a,b]. This generalization plays a role in the study of semigroups, via the LaplaceStieltjes transform.
43
RiemannStieltjes integral
44
References
[1] Partial orderings & Moore-Smith limit (http:/ / mathdl. maa. org/ images/ upload_library/ 22/ Chauvenet/ Mcshane. pdf) Retrieved on
03-05-2009
Hildebrandt, T. H. (1938), " Definitions of Stieltjes Integrals of the Riemann Type (http://www.jstor.org/
stable/2302540)", The American Mathematical Monthly 45 (5): 265278, MR 1524276 (http://www.ams.org/
mathscinet-getitem?mr=1524276), ISSN 0002-9890 (http://worldcat.org/issn/0002-9890)
Pollard, Henry (1920), "The Stieltjes integral and its generalizations", Quarterly Journal of Pure and Applied
Mathematics 19
Riesz, F.; Sz. Nagy, B. (1955), Functional Analysis, F. Ungar Publishing.
Shilov, G. E.; Gurevich, B. L. (1978), Integral, Measure, and Derivative: A Unified Approach, Dover
Publications, ISBN 0-486-63519-8, Richard A. Silverman, trans.
Stroock, Daniel W. (1998), A Concise Introduction to the Theory of Integration (3rd ed.), Birkhauser, ISBN
0-8176-4073-8.
RussoVallois integral
In mathematical analysis, the RussoVallois integral is an extension of the classical RiemannStieltjes integral
and
and to pull the limit out of the integral. In addition one changes the type of convergence.
Definition: A sequence
and
On sets:
and
RussoVallois integral
45
In this case the generalised bracket is equal to the classical covariation. In the special case, this means that the
process
then
By a duality result of Triebel one can provide optimal classes of Besov spaces, where the Russo-Vallois integral can
be defined. The norm in the Besov space
is given by
Theorem: Suppose
one has
Notice that in this case the RussoVallois integral coincides with the RiemannStieltjes integral and with the
Young integral for functions with finite p-variation.
References
Russo, Vallois: Forward, backward and symmetric integrals, Prob. Th. and rel. fields 97 (1993)
Russo, Vallois: The generalized covariation process and Ito-formula, Stoch. Proc. and Appl. 59 (1995)
Zhle; Forward Integrals and SDE, Progress in Prob. Vol. 52 (2002)
Fournier, Adams: Sobolev Spaces, Elsevier, second edition (2003)
Skorokhod integral
Skorokhod integral
In mathematics, the Skorokhod integral, often denoted , is an operator of great importance in the theory of
stochastic processes. It is named after the Ukrainian mathematician Anatoliy Skorokhod. Part of its importance is
that it unifies several concepts:
is an extension of the It integral to non-adapted processes;
is the adjoint of the Malliavin derivative, which is fundamental to the stochastic calculus of variations
(Malliavin calculus);
is an infinite-dimensional generalization of the divergence operator from classical vector calculus.
Definition
Preliminaries: the Malliavin derivative
Consider a fixed probability space (,,P) and a Hilbert space H; E denotes expectation with respect to P:
Intuitively speaking, the Malliavin derivative of a random variable F in Lp() is defined by expanding it in terms of
Gaussian random variables that are parametrized by the elements of H and differentiating the expansion formally;
the Skorokhod integral is the adjoint operation to the Malliavin derivative.
Consider a family of R-valued random variables W(h), indexed by the elements h of the Hilbert space H. Assume
further that each W(h) is a Gaussian (normal) random variable, that the map taking h to W(h) is a linear map, and that
the mean and covariance structure is given by
for all g and h in H. It can be shown that, given H, there always exists a probability space (,,P) and a family of
random variables with the above properties. The Malliavin derivative is essentially defined by formally setting the
derivative of the random variable W(h) to be h, and then extending this definition to smooth enough random
variables. For a random variable F of the form
where f:RnR is smooth, the Malliavin derivative is defined using the earlier formal definition and the chain
rule:
In other words, whereas F was a real-valued random variable, its derivative DF is an H-valued random variable, an
element of the space Lp(;H). Of course, this procedure only defines DF for smooth random variables, but an
approximation prcedure can be employed to define DF for F in a large subspace of Lp(); the domain of D is the
closure of the smooth random variables in the seminorm
46
Skorokhod integral
Just as the Malliavin derivative D was first defined on simple, smooth random variables, the Skorokhod integral has
a simple expression for simple processes: if u is given by
Properties
The isometry property: for any process u in L2(;H) that lies in the domain of ,
If u is an adapted process, then the second term on the right-hand side is zero, the Skorokhod and It integrals
coincide, and the above equation becomes the It isometry.
The derivative of a Skorokhod integral is given by the formula
where DhX stands for (DX)(h), the random variable that is the value of the process DX at time h in H.
The Skorokhod integral of the product of a random variable F in D1,2 and a process u in dom() is given by the
formula
References
Ocone, Daniel L. (1988). "A guide to the stochastic calculus of variations". Stochastic analysis and related topics
(Silivri, 1986). Lecture Notes in Math. 1316. Berlin: Springer. pp.179. MR953793 [1]
Sanz-Sol, Marta (2008). "Applications of Malliavin Calculus to Stochastic Partial Differential Equations
(Lectures given at Imperial College London, 711 July 2008) [2]". Retrieved 2008-07-09.
References
[1] http:/ / www. ams. org/ mathscinet-getitem?mr=953793
[2] http:/ / www. ma. ic. ac. uk/ ~dcrisan/ lecturenotes-london. pdf
47
Stratonovich integral
48
Stratonovich integral
In stochastic processes, the Stratonovich integral (developed simultaneously by Ruslan L. Stratonovich and D. L.
Fisk) is a stochastic integral, the most common alternative to the It integral. While the Ito integral is the usual
choice in applied math, the Stratonovich integral is frequently used in physics.
In some circumstances, integrals in the Stratonovich definition are easier to manipulate. Unlike the It calculus,
Stratonovich integrals are defined such that the chain rule of ordinary calculus holds.
Perhaps the most common situation in which these are encountered is as the solution to Stratonovich stochastic
differential equations (SDE). These are equivalent to It SDEs and it is possible to convert between the two
whenever one definition is more convenient.
Definition
The Stratonovich integral can be defined in a manner similar to the Riemann integral, that is as a limit of Riemann
sums. Suppose that
is a Wiener process and
is a semimartingale
adapted to the natural filtration of the Wiener process. Then the Stratonovich integral
of
Riemann-Stieltjes integral).
at
where is a continuously differentiable function and the last integral is an It integral (Kloeden & Platen 1992,
p.101).
It follows that if Xt is a time-homogeneous It diffusion with continuously differentiable diffusion coefficient (i.e.
it satisfies the SDE
), we have
where
is the continuous part of the quadratic covariation. With probability 1, a general stochastic process
does not satisfy the criteria for convergence in the Riemann sense. If it did, then the It and Stratonovich definitions
would converge to the same solution. As it is, for integrals with respect to Wiener processes, they are distinct.
Stratonovich integral
References
ksendal, Bernt K. (2003). Stochastic Differential Equations: An Introduction with Applications. Springer,
Berlin. ISBN 3-540-04758-1.
Gardiner, Crispin W. Handbook of Stochastic Methods Springer, (3rd ed.) ISBN 3-540-20882-8.
Jarrow, Robert and Protter, Philip, "A short history of stochastic integration and mathematical finance: The early
years, 18801970," IMS Lecture Notes Monograph, vol. 45 (2004), pages 117.
Kloeden, Peter E.; Platen, Eckhard (1992), Numerical solution of stochastic differential equations, Applications
of Mathematics, Berlin, New York: Springer-Verlag, ISBN 978-3-540-54062-5.
49
50
51
License
License
Creative Commons Attribution-Share Alike 3.0 Unported
http:/ / creativecommons. org/ licenses/ by-sa/ 3. 0/
52