Вы находитесь на странице: 1из 12

Integration by parts

In calculus, and more generally in mathematical analysis, integration by parts is a rule that
transforms the integral of products of functions into other (ideally simpler) integrals. The rule
arises from the product rule of differentiation.
If u = f(x), v = g(x), and the differentials du = f '(x) dx and dv = g'(x) dx, then the product rule in
its simplest form is:

Rule
Suppose f(x) and g(x) are two continuously differentiable functions. The product rule states

Integrating both sides gives

Rearranging terms

From the above one can derive the integration by parts rule, which states that, given an interval
with endpoints a and b,

with the common notation

The rule is shown to be true by using the product rule for derivatives and the fundamental
theorem of calculus. Thus
1

In the traditional calculus curriculum, the rule is often stated using indefinite integrals in the
form

or, if u = f(x), v = g(x) and the differentials du = f (x) dx and dv = g(x) dx, then it is in the form
most often seen:

This formula can be interpreted to mean that the area under the graph of a function u(v) is the
same as the area of the rectangle u v minus the area above the graph.
The original integral contains the derivative of g; in order to be able to apply the rule, the
antiderivative g must be found, and then the resulting integral g f dx must be evaluated.
One can also formulate a discrete analogue for sequences, called summation by parts.
An alternative notation has the advantage that the factors of the original expression are identified
as f and g:

This formula is valid whenever f is continuously differentiable and g is continuous.


More general formulations of integration by parts exist for the Riemann-Stieltjes integral and
Lebesgue-Stieltjes integral.
More complicated forms of the rule are also valid:

Strategy
Integration by parts is a heuristic rather than a purely mechanical process for solving integrals;
given a single function to integrate, the typical strategy is to carefully separate it into a product of
two functions (x)g(x) such that the integral produced by the integration by parts formula is
easier to evaluate than the original one. The following form is useful in illustrating the best
strategy to take:

Note that on the right-hand side, is differentiated and g is integrated; consequently it is useful
to choose as a function that simplifies when differentiated, and/or to choose g as a function that
simplifies when integrated. As a simple example, consider:

Since ln x simplifies to 1/x when differentiated, we make this part of ; since 1/x2 simplifies to
1/x when integrated, we make this part of g. The formula now yields:

The remaining integral of 1/x2 can be completed with the power rule and is 1/x.

Alternatively, we may choose and g such that the product


cancellation. For example, suppose we wish to integrate:

simplifies due to

If we choose (x) = ln(sin x) and g(x) = 1/(cos x)2, then differentiates to 1/tan x using the chain
rule and g integrates to tan x; so the formula gives:

The integrand simplifies to 1, so the antiderivative is x. Finding a simplifying combination


frequently involves experimentation.
3

In some applications, it may not be necessary to ensure that the integral produced by integration
by parts has a simple form; for example, in numerical analysis, it may suffice that it has small
magnitude and so contributes only a small error term. Some other special techniques are
demonstrated in the examples below.

Examples
Integrals with powers of x or ex
In order to calculate:

Let:

Then:

where C is an arbitrary constant of integration.


By repeatedly using integration by parts, integrals such as

can be computed in the same fashion: each application of the rule lowers the power of x by one.
An unusual example commonly used to examine the workings of integration by parts is

Here, integration by parts is performed twice. First let

and

Then:

Now, to evaluate the remaining integral, we use integration by parts again, with:

Then:

Putting these together,

The same integral shows up on both sides of this equation. The integral can simply be added to
both sides to get

where, again, C (and C' = C/2) is an arbitrary constant of integration.


A similar method is used to find the integral of secant cubed.

Interchange of the order of integration

Integration over the triangular area can be done using vertical or horizontal strips as the first step.
This is an overhead view, looking down the z-axis onto the x-y plane. The sloped line is the
curve y = x.
The above formulation includes the technique of interchange of the order of integration, which is
not usually viewed in this manner. Consider the iterated integral:

In the order written above, the strip of width dx is integrated first over the y-direction (a strip of
width dx in the x direction is integrated with respect to the y variable across the y direction) as
shown in the left panel of the figure, which is inconvenient especially when function h(y) is not
easily integrated. The integral can be reduced to a single integration by reversing the order of
integration as shown in the right panel of the figure. To accomplish this interchange of variables,
the strip of width dy is first integrated from the line x = y to the limit x = z, and then the result is
integrated from y = a to y = z, resulting in:

This result can be seen to be an example of the above formula for integration by parts, repeated
below:

Substitute:

and

However, exchange of the order of integration has the merit that it generates the function f in a
natural manner.

More examples
Two other well-known examples are when integration by parts is applied to a function expressed
as a product of 1 and itself. This works if the derivative of the function is known, and the integral
of this derivative times x is also known.
The first example is ln(x) dx. We write this as:

Let:

Then:

where, again, C is the constant of integration.


The second example is arctan(x) dx, where arctan(x) is the inverse tangent function. Rewrite
this as

Now let:

Then

using a combination of the inverse chain rule method and the natural logarithm integral
condition.
Here is an example:

LIATE rule
A rule of thumb proposed by Herbert Kasube of Bradley University advises that whichever
function comes first in the following list should be u:[1]
L: Logarithmic functions: lnx, logx, etc.
I: Inverse trigonometric functions: arctan x, arcsec x, etc.
A: Algebraic functions: ,
, etc.
T: Trigonometric functions: sin x, tan x, etc.
E: Exponential functions: ,
, etc.
The function which is to be dv is whichever comes last in the list: functions lower on the list
have easier antiderivatives than the functions above them. The rule is sometimes written as
"DETAIL", where D stands for dv.
To demonstrate the LIATE rule, consider the integral

Following the LIATE rule, u = x and dv = cos x dx , hence du = dx and v = sin x , which makes
the integral become

which equals

In general, one tries to choose u and dv such that du is simpler than u and dv is easy to integrate.
If instead cos x was chosen as u and x as dv, we would have the integral

which, after recursive application of the integration by parts formula, would clearly result in an
infinite recursion and lead nowhere.
Although a useful rule of thumb, there are exceptions to the LIATE rule. A common alternative is
to consider the rules in the "ILATE" order instead. Also, in some cases, polynomial terms need to
be split in non-trivial ways. For example, to integrate

we would set

This results in

Recursive integration by parts


Integration by parts can often be applied recursively on the
following formula

term to provide the

Here, is the first derivative of and is the second derivative of u. Further,


is a notation
to describe its nth derivative (with respect to the variable u and v are functions of). Another
notation has been adopted:

There are n + 1 integrals.


Note that the integrand above (uv) differs from the previous equation. The dv factor has been
written as v purely for convenience.
The above mentioned form is convenient because it can be evaluated by differentiating the first
term and integrating the second (with a sign reversal each time), starting out with uv1. It is very
useful especially in cases when u(k+1) becomes zero for some k + 1. Hence, the integral evaluation
can stop once the u(k) term has been reached.

Tabular integration by parts


While the aforementioned recursive definition is correct, it is often tedious to remember and
implement. A much easier visual representation of this process is often taught to students and is
dubbed either "the tabular method",[2] "the Stand and Deliver method",[3] "rapid repeated
integration" or "the tic-tac-toe method". This method works best when one of the two functions
in the product is a polynomial, that is, after differentiating it several times one obtains zero. It
may also be extended to work for functions that will repeat themselves.
For example, consider the integral

Let u = x3. Begin with this function and list in a column all the subsequent derivatives until zero
is reached. Secondly, begin with the function v (in this case cos(x)) and list each integral of v
until the size of the column is the same as that of u. The result should appear as follows.
Derivatives of u (Column A)

Integrals of v (Column B)
10

Now simply pair the 1st entry of column A with the 2nd entry of column B, the 2nd entry of
column A with the 3rd entry of column B, etc... with alternating signs (beginning with the
positive sign). Do so until further pairing is impossible. The result is the following (notice the
alternating signs in each term):

Which, with simplification, leads to the result

With proper understanding of the tabular method, it can be extended. Consider

Derivatives of u (Column A)

Integrals of v (Column B)

In this case in the last step it is necessary to integrate the product of the two bottom cells
obtaining:

which leads to

and yields the result:

11

Higher dimensions
The formula for integration by parts can be extended to functions of several variables. Instead of
an interval one needs to integrate over a n-dimensional set. Also, one replaces the derivative with
a partial derivative.
More specifically, suppose is an open bounded subset of
with a piecewise smooth
boundary . If u and v are two continuously differentiable functions on the closure of , then the
formula for integration by parts is

where is the outward unit surface normal to , is its i-th component, and i ranges from 1 to n.
We can obtain the most general form of the integration by parts by replacing v in the above
formula with vi and summing over i gives the vector formula

where v is a vector-valued function with components v1, ..., vn.


Setting u equal to the constant function 1 in the above formula gives the divergence theorem. For
where

, one gets

which is the first Green's identity.


The regularity requirements of the theorem can be relaxed. For instance, the boundary need
only be Lipschitz continuous. In the first formula above, only
is necessary
(where H1 is a Sobolev space); the other formulas have similarly relaxed requirements.

12

Вам также может понравиться