Sei sulla pagina 1di 19

NAME: TERM PROJECT OF MATHS

SECTION-
REG-NO-
ROLLNO-
TERM PROJECT-MATHAMATICS

INTEGRAL

International Gamma-Ray Astrophysics


Laboratory (INTEGRAL)

Artist's illustration of INTEGRAL in orbit (credit: ESA)

General information

Organization ESA / NASA / RKA

Launch date 17 October2002

Mass over 4,000 kg

Orbit height 9,000 km (perigee)


153,000 km (apogee)
Orbit period 72 hr

Telescope style coded mask

Wavelength gamma ray

Diameter 3.7 m

Collecting area 500 cm² (SPI, JEM-X)


3,100 cm² (IBIS)

Focal length ~4 meters

Instruments

SPI SpecTROMETER

IBIS Imager

JEM-X X-ray monitor

OMC optical monitor

The EuropEON SPACE AGENCY INTErnational Gamma-Ray Astrophysics


Laboratory (INTEGRAL) is detecting some of the most energetic radiation that comes
from space. It is the most sensitive gammaRAY observatory ever launched.

INTEGRAL is an ESA mission in cooperation with the Russian and NASA It has had
some notable successes, for example in detecting a mysterious '. It has also had great
success in investigating and evidence for black holes

Instruments:
Four instruments are coaligned to study a target across several ranges. The coded masks
were led by the University of Valencia, Spain.
The INTEGRAL imager, IBIS (Imager on-Board the INTEGRAL Satellite) observes
from 15 keV (hard X-rays) to 10 MeV (gamma rays). Mechanical resolution is 12 arcmin,
but deconvolution can reduce that to as little as 1 arcmin. A 95 x 95 mask of rectangular
tungsten tiles sits 3.2 meters above the detectors. The detector system contains a forward
plane of 128 x 128 Cadmium-Telluride tiles (ISGRI- Integral Soft Gamma-Ray Imager),
backed by a 64 x 64 plane of Caesium-Iodide tiles (PICsIT- Pixellated Caesium-Iodide
Telescope). ISGRI is sensitive up to 500 keV, while PICsIT extends to 10 MeV. Both are
surrounded by passive shields of tungsten and lead.

The primary spectrometer aboard INTEGRAL is SPI, the SPectrometer for INTEGRAL.
It observes radiation between 20 keV and 8 MeV. SPI consists of a coded mask of
hexagonal tungsten tiles, above a detector plane of 19 germanium crystals (also packed
hexagonally). The Ge crystals are actively cooled with a mechanical system, and give an
energy resolution of 2 keV at 1 MeV.

IBIS and SPI need a method to stop background radiation. The SPI ACS
(AntiCoincidence Shield) consists of a mask shield and a detector shield. The mask shield
is a layer of plastic scintillator behind the tungsten tiles. It absorbs secondary radiation
produced by impacts on the tungsten. The rest of the shield consists of BGO scintillator
tiles around the sides and back of the SPI.

The enormous area of the ACS that results makes it an instrument in its own right. Its all-
sky coverage and sensitivity make it a natural gamma-ray burst detector, and a valued
component of the IPN (InterPlanetary Network). Recently, new algorithms allow the
ACS to act as a telescope, through double Compton scattering. Thus ACS can study
objects outside the field of view of the other instruments, with surprising spatial and
energy resolution.

Dual JEM-X units provide additional information on targets. They observe in soft and
hard X-rays, from 3 to 35 keV. Aside from broadening the spectral coverage, imaging is
more precise due to the shorter wavelength. Detectors are gas scintillators (xenon plus
methane) in a microstrip layout, below a mask of hexagonal tiles.

INTEGRAL mounts an Optical Monitor (OM), sensitive from 500 to 850 nm. It acts as
both a framing aid, and can note the activity and state of some brighter targets.

The spacecraft also mounts a radiation monitor, INTEGRAL Radiation Environment


Monitor (IREM), to note the orbital background for calibration purposes. IREM has an
electron and a proton channel, though radiation up to cosmic rays can be sensed. Should
the background exceed a preset threshold, IREM can shut down the instrument.

An integral helps to find out how much space is under a graph of something. Integrals
undo derivatives. A derivative helps to find what the steepness is of a graph.

This is the symbol for integration:


Integrals and derivatives are part of a branch of mathematics called calculus.

Integration helps when trying to multiply units into a problem. For example, if a problem

with rate ( ) needs an answer with just distance, one solution is to integrate

with respect to time. This means multiplying in time (to cancel the time in ).
This is done by adding small slices of the rate graph together. The slices are close to zero
in width, but adding them forever makes them add up to a whole. This is called a
Riemann Sum.

Adding these slices together gives the equation that the first equation is the derivative of.
Integrals are kind of like adding machines.

Another time integration is helpful is when finding the volume of a solid. It can add two-
dimensional (without width) slices of the solid together forever until there is a width.
This means the object now has three dimensions: the original two and a width. This gives
the volume of the three-dimensional object described.

Abelian integral:
In mathematics, an abelian integral in Riemann surface theory is a function related to
the indefinite integral of a differential of the first kind. Suppose we are given a Riemann
surface S and on it a differential 1-form ω that is everywhere holomorphic on S, and fix a
point P on S from which to integrate. We can regard

as a multi-valued function f(Q), or (better) an honest function of the chosen path C drawn
on S from P to Q. Since S will in general be multiply-connected, one should specify C,
but the value will in fact only depend on the homology class of C.

In the case of S a compact Riemann surface of genus 1, i.e. an elliptic curve, such
functions are the elliptic integrals. Logically speaking, therefore, an abelian integral
should be a function such as f.

Such functions were first introduced to study hyperelliptic integrals, i.e. for the case
where S is a hyperelliptic curve. This is a natural step in the theory of integration to the
case of integrals involving algebraic functions √A, where A is a polynomial of degree > 4.
The first major insights of the theory were given by Niels Abel; it was later formulated in
terms of the Jacobian variety J(S). Choice of P gives rise to a standard holomorphic
mapping
S → J(S)

of complex manifolds. It has the defining property that the holomorphic 1-forms on J(S),
of which there are g independent ones if g is the genus of S, pull back to a basis for
the differentials of the first kind on S. Elliptic integral

Redirected from Elliptic Integral

An elliptic integral is any function f which can be expressed in the form


<math> f(x) = \int_{c}^{x} R(t,P(t))\ dt </math>

where R is a rational function of its two arguments, P is the square root of a polynomial
of degree 3 or 4 with no repeated roots, and c is a constant.

Particular examples include:

• The complete elliptic integral of the first kind K is defined as

<math> K(x) = \int_{0}^{1} \frac{1}{ \sqrt{(1-t^2)(1-x^2 t^2)} }\ dt </math>


and can be computed in terms of the arithmetic-geometric mean.

• The complete elliptic integral of the second kind E is defined as

<math> E(x) = \int_{0}^{1} \frac{ \sqrt{1-x^2 t^2} }{ \sqrt{1-t^2} }\ dt </math>

A real number is called integral if it is an integer. The integral value of a real number x
is defined to be the largest integer which is less than or equal to x; it is often denoted by
⌊x⌋ and also called the floor function.

Integral of a mathematical function :


In the integral calculus, the integral of a function is informally defined as the size of the
area delimited by the x axis and the graph of the function. In the case of non-negative
functions, the notion of area is the usual one. For functions which take negative values, a
special interpretation is used, and "negative area" is possible.

Let f(x) be a function of the interval [a,b] into the real numbers. For simplicity, assume
that this function is non-negative (it takes no negative values.) The set S=Sf:={(x,y)|
0≤y≤f(x)} is the region of the plane between f and the x axis. Measuring the "area" of S is
desirable, and this area is denoted by ∫f, and it is the (definite) integral of f.

Details can be found under Riemann integral and Lebesgue integral. The concept of
Riemann integration was developed first, and Lebesgue integrals were developed to deal
with pathological cases for which the Riemann integral was not defined. If a function is
Riemann integrable, then it is also Lebesgue integrable, and the two integrals coincide.

The antiderivative approach occurs when we seek to find a function F(x) whose
derivative F(x) is some given function f(x). This approach is motivated by calculus, and
is the main method used for calculating the area under the curve as described in the
preceding paragraph, for functions given by formulae.

Functions which have antiderivatives are also Riemann integrable (and hence Lebesgue
integrable.) The nonobvious theorem that states that the two approaches ("area under the
curve" and "antiderivative") are in some sense the same as the fundamental theorem of
calculus

The nuance between Riemann and Lebesgue integration :


Both the Riemann and the Lebesgue integral are approaches to integration which seek to
measure the area under the curve, and the overall schema in both cases is the same.

First, we select a family of elementary functions, for which we have an obvious way of
measuring the area under the curve. In the case of the Riemann integral, this choice is so
that the area under the curve can be regarded as a finite union of rectangles, and the
functions are called step functions. For the Lebesgue integral, "rectangle" is replaced by
something more sophisticated, and the resulting functions are called simple functions.

Then we try to impose monotonicity. If 0≤f≤g (and hence Sf is a subset of Sg) then we
should have that ∫f≤∫g. With this monotonicity requirement, for an arbitrary non-negative
function f, we can approximate its area from below using a carefully chosen elementary
function s (in the case of Riemann integration, a step function, and in the case of
Lebesgue integration, a simple function.) We choose s so that s≤f but s is very close to f.
The area under s is a lower bound for the integral of f, and it is called a lower sum. In the
case of the Riemann integral, we also produce upper sums in a similar fashion: we choose
step functions, say s, so that s≥f but s is very close to f, and we regard such an upper sum
as an upper bound for the area under f. The Lebesgue theory does not use upper sums.
Lastly, a limit-taking step is taken to make the elementary functions approach f more and
more closely, and an area is obtained for some functions f. The functions which we can
integrate are said to be integrable. However, the differences begin here; the Riemann
theory was simpler thus far, but its simplicity results in a more limited set of integrable
functions than the Lebesgue theory. In addition, the interaction between limits and the
integral are more difficult to describe in the Riemann setting.

In branch of mathematics known as Real analysis, the


Riemann integral is a simple way of viewing the integral of
a function on an interval as the area under the curve.

The numbers that appear in the upper righthand corner of the animation above give the
sums of the areas of the grey rectangles. As the number of rectangles increases this sum
converges to the Riemann integral for the curve shown.

Let f(x) be a real-valued function of the interval [a,b], so that for all x, f(x)≥0 (f is non-
negative.) Further let S=Sf:={(x,y)|0≤y≤f(x)} (see Figure 2) be the region of the plane
under the function f(x) and above the interval [a,b]. We are interested in measuring the
area of S if that is possible, and we denote this area by ∫abf(x)dx. In case several variables
appear in f, the dx will serve to specify the variable of integration. If the variable of
integration and interval of integration are understood, the notation can be simplified to ∫f.

Once we have succeeded in evaluating the integral of f for certain f which are non-
negative, we can extend the integral to functions which may take negative values by
linearity. Some functions have no clear Riemann integral, but especially, the interactions
of limits and the Riemann integral are difficult to study. An improvement is to use the
Lebesgue integral which both succeeds at integrating a broader variety of functions, as
well as better describing the interactions of limits and integrals.

Historically, Riemann designed this theory first and gave some evidence for the
fundamental theorem of calculus. The theory of Lebesgue integration arrived much later,
when the weaknesses of the Riemann integral were better understood.

The basic idea of the Riemann integral is to use very simple and unambiguous
approximations for the area of S. We find an approximate area which we are certain is
less than the area of S, and we find an approximate area which we are certain is more
than the area of S. If these approximations can be made arbitrarily close to one another,
then we can assign an area to S.
Because of the geometric nature of the Riemann integral, it allows us to formulate many
problems of nature as a problem of integration. It also provides some hints for methods of
numerical integration, for evaluating definite integrals on computers to an acceptable
degree of precision. However, for exact calculations for given formulae, the Riemann
integral does not suggest a suitable approach.

For certain functions, the theory of antiderivatives provides exact results for definite
integrals. While the Riemann integration theory justifies taking limits and provides a
geometric point of view, the antiderivative theory of integration gives tools for
integrating certain formulae precisely.

The fact that the seemingly disparate theories of Riemann integration and antiderivatives
are essentially talking about the same subject is contained in the fundamental theorem of
calculus.

Step functions :
Let E be any subset of [a,b]. Let XE(x) be the function which is 1 if x is in E and 0 if x is
not in E. XE is called the indicating function of E, or the characteristic function of E.

These functions are our starting point, and we should agree that

∫z X[c,d](x)dx=z(d-c)

for any interval [c,d] in [a,b] and any constant z≥0. Indeed, in this case, the area under the
curve is a rectangle with base [c,d] and height z.

Likewise, some geometric experimentation with such functions suggests that if f1, f2, ..., fn
are n indicating functions of intervals, <math>a_1, a_2, ..., a_n</math> are scalars, then
the area under

<math>f=a_1f_1+a_2f_2+...+a_nf_n</math>

should be

<math>\int f = a_1 \int f_1 + a_2 \int f_2 + ... + a_n \int f_n</math>

A function of this form is called a linear combination of indicating functions, or more


simply a step function. We note now that we've decided what the integral of step
functions ought to be.

We will take a shortcut now by stating that the preceding formula will be used even if
some (or all) of the coefficients aj are negative.
A crucial difference between the Riemann integral and the Lebesgue integral is that the
step functions of the Lebesgue integral are linear combinations of indicating functions of
sets which are not necessarely intervals. Of course, work is then required to calculate the
integral of this larger class of step functions. Also note that the Lebesgue integral does
not use upper sums, and that non-negative functions are dealt with first, before extending
to functions which may take negative values.

Lower and upper sums :


From the geometry of the problem, we impose that if f(x)≤g(x) for all x in [a,b] then we
really ought to have

∫f≤∫g

simply by seeing that the area Sf is a subset of Sg (at least in the case of non-negative
functions, this is clear.) We call this requirement monotonicity.

Given the integral of step functions and the monotonicity requirement, we can get a first
stab at integrating arbitrary non-negative functions. Let f(x) be a real-valued function of
[a,b] and let l(x) be a step function such that l(x)≤f(x) for all x. Furthermore, let u(x) be a
step function such that u(x)≥f(x) for all x. If we are to assign a value to ∫f consistent with
the monotonicity requirement, then we need that

∫l≤∫f≤∫u

The integral ∫l is then called a lower sum for f and the integral ∫u is then called an upper
sum for f. The preceding inequality must hold for all lower and upper sums of f, so we
can deduce another inequality:

supl∫l≤∫f≤infu∫u

where supl∫l is the smallest upper bound for all lower sums, and infu∫u is the largest lower
bound for all upper sums (see supremum and infimum.) The number supl∫l is sometimes
called the lower sum; likewise, the number infu∫u is the upper sum.

If the supremum and infimum are equal, then there is only one choice left for ∫f. It may
not happen that the supremum is larger than the infimum (this is by our construction, as
the reader may check.) However, it may happen that the supremum is less, and not equal
to, the infimum. For instance, the reader may check that, for the indicating function

XQ
where Q is the set of rational numbers in [a,b], a<b, the lower sum is 0 and the upper
sum is b-a>0.

The collection of functions whose lower sum and upper sum are equal and finite is the set
of Riemann integrable functions. By contrast, functions that have differing upper and
lower sums are said to be non-Riemann integrable. In the context of this article, we will
say integrable or non-integrable with the understanding that we are speaking of Riemann
integrability.

One also checks that a step function's integral is equal to is lower and upper sums.

Results about the Riemann integral :


Lemma 1: Let [a,b] be an interval. The map I:f→∫f which maps f to its integral from a
to b is a linear map. That is, for any integrable functions f and g, and any real number a,
I(af+g)=aI(f)+I(g).

This can be shown from first principles, from the construction of the Riemann integral.

Theorem 2: Any real-valued continuous function of the interval [a,b] is integrable. The
proof relies on the fact that any continuous function of an interval is necessarely
uniformly continuous.

Corollary 3: If f is continuous everywhere in [a,b] except perhaps for finitely many


points of discontinuity, and f is bounded, then f is integrable.

The boundedness requirement can not be dropped.

Theorem 4: If fk is a sequence of integrable functions over [a,b], and if fk converge


uniformly to a function f, then f is integrable, and the integrals ∫fk converge to ∫f.

Corollary 5: Let C(a,b) be the Banach space of continuous functions over [a,b] with the
uniform norm[?]. Then I:f→∫f is continuous. Together with Lemma 1, this says that the
integral is a continuous functional of C(a,b).

The hypotheses of theorem 4 (uniform convergence on a fixed, bounded interval) are


very strong. A primary failing of the Riemann integral is the difficulty we face when
attempting to relax these hypotheses. In fact, the numerical sequence ∫fk will converge to
the number ∫f a lot more often than is suggested by the theorem, but it is very difficult to
prove so in this setting. The correct way of getting a stronger theorem is to use the
Lebesgue integral.
Another problem with the Riemann integral is that it does not extend to unbounded
intervals very succesfully. If we wish to integrate a function f from -∞ to +∞, we can
naively calculate

limn→+∞ ∫-nnf(x)dx

However, certain properties (such as translation invariance, the fact that the Riemann
integral does not change if we translate the integrand f) are lost. In fact, Theorem 4
becomes false for such an integral, and it becomes very difficult to use limits in
conjunctions with integrals. Such an integral is called an improper integral, for it is not
deemed to be a Riemann integral, strictly speaking. Again, the Lebesgue integral
alleviates these difficulties

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 Riemann-Stieltjes integral, and most of them disappear in the Lebesgue integral.

The integral as the area of a region under a curve.

INTRODUCTION:
Let f(x) be a non-negative real-valued function of the interval [a,b], and let S = {(x,y) | 0
< y < f(x)} be the region of the plane under the function f(x) and above the interval [a,b]
(see the figure on the top right). We are interested in measuring the area of S. Once we
have measured it, we will denote the area by:
The basic idea of the Riemann integral is to use very simple approximations for the area
of S. By taking better and better approximations, we can say that "in the limit" we get
exactly the area of S under the curve.

A sequence of Riemann sums. The numbers in the upper right are the areas of the grey
rectangles. They converge to the integral of the function.

Definition of the Riemann integral Partitions of an


interval:
A partition of an interval [a,b] is a finite sequence
. Each [xi,xi + 1] is called a subinterval of the
partition. The mesh of a partition is defined to be the length of the longest subinterval
[xi,xi + 1], that is, it is max(xi + 1 − xi) 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 i,
. In other words, it is a partition together with a distinguished point of
every subinterval. The mesh of a tagged partition is defined the same as for an ordinary
partition.

Suppose that together with are a tagged partition of [a,b], and


that together with are another tagged partition of [a,b]. We
say that and together are a refinement of
together with if for each integer i with , there is an integer r(i)
such that xi = yr(i) and such that ti = sj for some j with . Said more
simply, a 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 f which is defined on the interval [a,b]. The Riemann sum
of f with respect to the 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 f(ti) and length xi + 1 − xi. The Riemann sum is the signed area under all the
rectangles.

The Riemann integral:


Loosely speaking, the Riemann integral is the limit of the Riemann sums of a function as
the partitions get finer and finer. However, being precise about what is meant by "finer
and finer" is somewhat tricky.

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 f equals s if the following
condition holds:

For all ε > 0, there exists δ > 0 such that for any tagged partition and
whose mesh is less than δ, we have

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 f equals s if the following condition holds:

For all ε > 0, there exists a tagged partition and such


that for any refinement and of and
, we have
Both of these mean that eventually, the Riemann sum of f 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 ε, and choose a δ that satisfies the condition.
Choose any tagged partition whose mesh is less than δ. Its Riemann sum is within ε of s,
and any refinement of this partition will also have mesh less than δ, so the Riemann sum
of the refinement will also be within ε 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. Choose a partition such that the lower and upper Darboux sums

with respect to this partition are within of the value s of the Darboux integral. Let r

equal , where Mi and mi are the supremum and infimum,

respectively, of f on [xi,xi + 1], and let δ be less than both and .


Then it is not hard to see that the Riemann sum of f with respect to any tagged partition of

mesh less than δ will be within of the upper or lower Darboux sum, so it will be within
ε of s.

Examples:
Let be the function which takes the value 1 at every point. Any
Riemann sum of f on [0,1] will have the value 1, therefore the Riemann integral of f on
[0,1] is 1.

Let be the indicator function of the rational numbers in [0,1]; that is,
takes the value 1 on 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 ti is between xi


and xi + 1). Choose ε > 0. The ti have already been chosen, and we can't change the value
of f at those points. But if we cut the partition into tiny pieces around each ti, we can
minimize the effect of the ti. Then, by carefully choosing the new tags, we can make the
value of the Riemann sum turn out to be within ε of either zero or one—our choice!
Our first step is to cut up the partition. There are n of the ti, and we want their total effect
to be less than ε. If we confine each of them to an interval of length less than ε / n, then
the contribution of each ti to the Riemann sum will be at least and at most. This makes the
total sum at least zero and at most ε. So let δ be a positive number less than ε / n. If it
happens that two of the ti are within δ of each other, choose δ smaller. If it happens that
some ti is within δ of some xj, and ti is not equal to xj, choose δ smaller. Since there are
only finitely many ti and xj, we can always choose δ sufficiently small.

Now we add two cuts to the partition for each ti. One of the cuts will be at ti − δ / 2, and
the other will be at ti + δ / 2. If one of these leaves the interval [0,1], then we leave it out.
ti will be the tag corresponding to the subinterval [ti − δ / 2,ti + δ / 2]. If ti is directly on
top of one of the xj, then we let ti be the tag for both [ti − δ / 2,xj] and [xj,ti + δ / 2]. 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 1 − ε. The second
way is to always choose an irrational point, so that the Riemann sum is as small as
possible. This will make the value of the Riemann sum at most ε.

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 s, 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.

Other concepts similar to the Riemann integral:


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, ti = xi for all i, and in a right-hand Riemann sum, ti = xi + 1 for all
i. 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 ti. 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 n'th regular subdivision of [0,1] consists of the intervals. Again, alone this 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
[0,1] 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. The Lebesgue integral is defined in such a
way that all these integrals are 0.

Facts about the Riemann integral:


The Riemann integral is a linear transformation; that is, if f and g are Riemann-integrable
on [a,b] and α and β are constants, then A real-valued function f on [a,b] is Riemann-
integrable if and only if it is bounded and continuous almost everywhere. If a real-valued
function on [a,b] is Riemann-integrable, it is Lebesgue-integrable.

If fn is a uniformly convergent sequence on [a,b] with limit f, then If a real-valued


function is monotone on the interval [a,b], it is Riemann-integrable.

Generalizations of the Riemann integral:


It is easy to extend the Riemann integral to functions with values in the Euclidean vector
space for any n. The integral is defined by linearity; in other words, if, 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 f(x) = 1 for all x > 0, f(0) = 0, and f(x) = − 1 for all x < 0. Then, for all x.
But if we shift f(x) to the right by one unit to get f(x − 1), we get for all x > 1.

Since this is unacceptable, we could try the definition: Then if we attempt to integrate the
function f above, we get, because we take the limit as b tends to first. If we reverse the
order of the limits, then we get.

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 fn(x) = 1 / n
on (0,n) and 0 everywhere else. For all n,. But fn converges uniformly to zero, so the
integral of is zero. Consequently. 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 bounded Lebesgue-integrable function f defined on a bounded interval is
Riemann-integrable if and only if the set of points where f is discontinuous has Lebesgue
measure zero. An integral which is in fact a direct generalization of the Riemann integral
is the Henstock-Kurzweil integral. Another way of generalizing the Riemann integral is
to replace the factors xi + 1 − xi 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 Riemann-Stieltjes integral.

Višestruki integral:
ACKNOWLEDGEMENT:
Just as the definite integral of a positive function of one variable represents the area of the
region between the graph of the function and the x-axis, the double integral of a positive
function of two variables represents the volume of the region between the surface defined
by the function and the plane which contains its domain. (Note that the same volume can
be obtained via the triple integral — the integral of a function in three variables — of
the constant function f(x, y, z) = 1 over the above-mentioned region between the surface
and the plane.) If there are more variables, a multiple integral will yield hypervolumes of
multi-dimensional functions.

Multiple integration of a function in variables: over a domain is


most commonly represented by nesting integral signs in the reverse order of execution
(the leftmost integral sign is computed last) proceeded by the function and integrand
arguments in proper order (the rightmost argument is computed last). The domain of
integration is either represented symbolically for every integrand over each integral sign,
or is often abbreviated by a variable at the rightmost integral sign:

Since it is impossible to calculate the antiderivative of a function of more than one


variable, indefinite multiple integrals do not exist. Therefore all multiple integrals are
definite integrals.
SCOPE:
Because gamma rays and X-rays cannot penetrate Earth's atmosphere, direct observations
must be made from space. INTEGRAL was launched from Baikonur spaceport, in
Kazakhstan. The 2002 launch aboard a Proton-DM2 rocket achieved a 700 km perigee.
The onboard thrusters then raised the perigee out of the residual atmosphere, and the
worst regions of the radiation belts. The apogee was trimmed with the thrusters to
synchronize with Earth's rotation, and thus, the satellite's ground stations.

INTEGRAL's operational orbit has a period of 72 hours, and has a high eccentricity, with
perigee close to the Earth at 10,000 km, within the magnetospheric radiation belt.
However, most of each orbit is spent outside this region, where scientific observations
may take place. It reaches a furthest distance from Earth (apogee) of 153,000 km. The
apogee was placed in the northern hemisphere, to reduce time spent in damaging eclipses,
and maximize contact time over the ground stations in the northern hemisphere. It is
controlled from ESOC in Darmstadt, Germany, ESA's control centre, through ground
stations in Belgium (Redu) and California (Goldstone).Fuel usage is within predictions.
INTEGRAL has already exceeded its 2.2-year planned lifetime; barring mechanical
failures, it should continue to function for six years or more. The spacecraft body
("service module") is a copy of the XMM-Newton body. This saved development costs
and simplified integration with infrastructure and ground facilities. (An adapter was
necessary to mate with the different booster, though.) However, the denser instruments
used for gamma rays and hard X-rays make INTEGRAL the heaviest scientific payload
ever flown by ESA. The body is constructed largely of composites. Propulsion is by a
hydrazine monopropellant system, containing 544 kg of fuel in four exposed tanks. The
titanium tanks were charged with gas to 24 bar (2.4 MPa) at 30 °C, and have tank
diaphragms. Attitude control is via a star tracker, multiple Sun sensors, and multiple
momentum wheels. The dual solar arrays, spanning 16 meters when deployed and
producing 2.4 kW BoL, are backed up by dual nickel-cadmium battery sets. The
instrument structure ("payload module") is also composite. A rigid base supports the
detector assemblies, and an H-shaped structure holds the coded masks approximately 4
meters above their detectors. The payload module can be built and tested independently
from the service module, reducing cost. Alenia Spazio was the spacecraft prime
contractor.

Potrebbero piacerti anche