fbpx
Wikipedia

Integration by parts

In calculus, and more generally in mathematical analysis, integration by parts or partial integration is a process that finds the integral of a product of functions in terms of the integral of the product of their derivative and antiderivative. It is frequently used to transform the antiderivative of a product of functions into an antiderivative for which a solution can be more easily found. The rule can be thought of as an integral version of the product rule of differentiation.

The integration by parts formula states:

Or, letting and while and , the formula can be written more compactly:

Mathematician Brook Taylor discovered integration by parts, first publishing the idea in 1715.[1][2] More general formulations of integration by parts exist for the Riemann–Stieltjes and Lebesgue–Stieltjes integrals. The discrete analogue for sequences is called summation by parts.

Theorem

Product of two functions

The theorem can be derived as follows. For two continuously differentiable functions u(x) and v(x), the product rule states:

 

Integrating both sides with respect to x,

 

and noting that an indefinite integral is an antiderivative gives

 

where we neglect writing the constant of integration. This yields the formula for integration by parts:

 

or in terms of the differentials  ,  

 

This is to be understood as an equality of functions with an unspecified constant added to each side. Taking the difference of each side between two values x = a and x = b and applying the fundamental theorem of calculus gives the definite integral version:

 
The original integral ∫ uv′ dx contains the derivative v′; to apply the theorem, one must find v, the antiderivative of v', then evaluate the resulting integral ∫ vu′ dx.

Validity for less smooth functions

It is not necessary for u and v to be continuously differentiable. Integration by parts works if u is absolutely continuous and the function designated v′ is Lebesgue integrable (but not necessarily continuous).[3] (If v′ has a point of discontinuity then its antiderivative v may not have a derivative at that point.)

If the interval of integration is not compact, then it is not necessary for u to be absolutely continuous in the whole interval or for v′ to be Lebesgue integrable in the interval, as a couple of examples (in which u and v are continuous and continuously differentiable) will show. For instance, if

 

u is not absolutely continuous on the interval [1, ∞), but nevertheless

 

so long as   is taken to mean the limit of   as   and so long as the two terms on the right-hand side are finite. This is only true if we choose   Similarly, if

 

v′ is not Lebesgue integrable on the interval [1, ∞), but nevertheless

 
with the same interpretation.

One can also easily come up with similar examples in which u and v are not continuously differentiable.

Further, if   is a function of bounded variation on the segment   and   is differentiable on   then

 

where   denotes the signed measure corresponding to the function of bounded variation  , and functions   are extensions of   to   which are respectively of bounded variation and differentiable.[citation needed]

Product of many functions

Integrating the product rule for three multiplied functions, u(x), v(x), w(x), gives a similar result:

 

In general, for n factors

 

which leads to

 

Visualization

 
Graphical interpretation of the theorem. The pictured curve is parametrized by the variable t.

Consider a parametric curve by (x, y) = (f(t), g(t)). Assuming that the curve is locally one-to-one and integrable, we can define

 
 

The area of the blue region is

 

Similarly, the area of the red region is

 

The total area A1 + A2 is equal to the area of the bigger rectangle, x2y2, minus the area of the smaller one, x1y1:

 

Or, in terms of t,

 

Or, in terms of indefinite integrals, this can be written as

 

Rearranging:

 

Thus integration by parts may be thought of as deriving the area of the blue region from the area of rectangles and that of the red region.

This visualization also explains why integration by parts may help find the integral of an inverse function f−1(x) when the integral of the function f(x) is known. Indeed, the functions x(y) and y(x) are inverses, and the integral ∫ x dy may be calculated as above from knowing the integral ∫ y dx. In particular, this explains use of integration by parts to integrate logarithm and inverse trigonometric functions. In fact, if   is a differentiable one-to-one function on an interval, then integration by parts can be used to derive a formula for the integral of  in terms of the integral of  . This is demonstrated in the article, Integral of inverse functions.

Applications

Finding antiderivatives

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 this single function into a product of two functions u(x)v(x) such that the residual integral from the integration by parts formula is easier to evaluate than the single function. The following form is useful in illustrating the best strategy to take:

 

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

 

Since the derivative of ln(x) is 1/x, one makes (ln(x)) part u; since the antiderivative of 1/x2 is −1/x, one makes 1/x2 dx part dv. The formula now yields:

 

The antiderivative of −1/x2 can be found with the power rule and is 1/x.

Alternatively, one may choose u and v such that the product u′ (∫v dx) simplifies due to cancellation. For example, suppose one wishes to integrate:

 

If we choose u(x) = ln(|sin(x)|) and v(x) = sec2x, then u differentiates to 1/ tan x using the chain rule and v 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.

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.

Polynomials and trigonometric functions

In order to calculate

 

let:

 
 

then:

 

where C is a constant of integration.

For higher powers of x in the form

 

repeatedly using integration by parts can evaluate integrals such as these; each application of the theorem lowers the power of x by one.

Exponentials and trigonometric functions

An example commonly used to examine the workings of integration by parts is

 

Here, integration by parts is performed twice. First let

 
 

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

 

which rearranges to

 

where again C (and C′ = C/2) is a constant of integration.

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

Functions multiplied by unity

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 C is the constant of integration.

The second example is the inverse tangent function arctan(x):

 

Rewrite this as

 

Now let:

 
 

then

 

using a combination of the inverse chain rule method and the natural logarithm integral condition.

LIATE rule

A rule of thumb has been proposed, consisting of choosing as u the function that comes first in the following list:[4]

Llogarithmic functions:   etc.
Iinverse trigonometric functions (including hyperbolic analogues):   etc.
Aalgebraic functions:   etc.
Ttrigonometric functions (including hyperbolic analogues):   etc.
Eexponential functions:   etc.

The function which is to be dv is whichever comes last in the list. The reason is that functions lower on the list generally have easier antiderivatives than the functions above them. The rule is sometimes written as "DETAIL" where D stands for dv and the top of the list is the function chosen to be 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 dx 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

 

one would set

 

so that

 

Then

 

Finally, this results in

 

Integration by parts is often used as a tool to prove theorems in mathematical analysis.

Wallis product

The Wallis infinite product for  

 

may be derived using integration by parts.

Gamma function identity

The gamma function is an example of a special function, defined as an improper integral for  . Integration by parts illustrates it to be an extension of the factorial function:

 

Since

 

when   is a natural number, that is,  , applying this formula repeatedly gives the factorial:  

Use in harmonic analysis

Integration by parts is often used in harmonic analysis, particularly Fourier analysis, to show that quickly oscillating integrals with sufficiently smooth integrands decay quickly. The most common example of this is its use in showing that the decay of function's Fourier transform depends on the smoothness of that function, as described below.

Fourier transform of derivative

If f is a k-times continuously differentiable function and all derivatives up to the kth one decay to zero at infinity, then its Fourier transform satisfies

 

where f(k) is the kth derivative of f. (The exact constant on the right depends on the convention of the Fourier transform used.) This is proved by noting that

 

so using integration by parts on the Fourier transform of the derivative we get

 

Applying this inductively gives the result for general k. A similar method can be used to find the Laplace transform of a derivative of a function.

Decay of Fourier transform

The above result tells us about the decay of the Fourier transform, since it follows that if f and f(k) are integrable then

 

In other words, if f satisfies these conditions then its Fourier transform decays at infinity at least as quickly as 1/|ξ|k. In particular, if k ≥ 2 then the Fourier transform is integrable.

The proof uses the fact, which is immediate from the definition of the Fourier transform, that

 

Using the same idea on the equality stated at the start of this subsection gives

 

Summing these two inequalities and then dividing by 1 + |2πξk| gives the stated inequality.

Use in operator theory

One use of integration by parts in operator theory is that it shows that the −∆ (where ∆ is the Laplace operator) is a positive operator on L2 (see Lp space). If f is smooth and compactly supported then, using integration by parts, we have

 

Other applications

Repeated integration by parts

Considering a second derivative of   in the integral on the LHS of the formula for partial integration suggests a repeated application to the integral on the RHS:

 

Extending this concept of repeated partial integration to derivatives of degree n leads to

 

This concept may be useful when the successive integrals of   are readily available (e.g., plain exponentials or sine and cosine, as in Laplace or Fourier transforms), and when the nth derivative of   vanishes (e.g., as a polynomial function with degree  ). The latter condition stops the repeating of partial integration, because the RHS-integral vanishes.

In the course of the above repetition of partial integrations the integrals

  and   and  

get related. This may be interpreted as arbitrarily "shifting" derivatives between   and   within the integrand, and proves useful, too (see Rodrigues' formula).

Tabular integration by parts

The essential process of the above formula can be summarized in a table; the resulting method is called "tabular integration"[5] and was featured in the film Stand and Deliver (1988).[6]

For example, consider the integral

  and take  

Begin to list in column A the function   and its subsequent derivatives   until zero is reached. Then list in column B the function   and its subsequent integrals   until the size of column B is the same as that of column A. The result is as follows:

# i Sign A: derivatives u(i) B: integrals v(ni)
0 +    
1    
2 +    
3    
4 +    

The product of the entries in row i of columns A and B together with the respective sign give the relevant integrals in step i in the course of repeated integration by parts. Step i = 0 yields the original integral. For the complete result in step i > 0 the ith integral must be added to all the previous products (0 ≤ j < i) of the jth entry of column A and the (j + 1)st entry of column B (i.e., multiply 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 the given jth sign. This process comes to a natural halt, when the product, which yields the integral, is zero (i = 4 in the example). The complete result is the following (with the alternating signs in each term):

 

This yields

 

The repeated partial integration also turns out useful, when in the course of respectively differentiating and integrating the functions   and   their product results in a multiple of the original integrand. In this case the repetition may also be terminated with this index i.This can happen, expectably, with exponentials and trigonometric functions. As an example consider

 
# i Sign A: derivatives u(i) B: integrals v(ni)
0 +    
1    
2 +    

In this case the product of the terms in columns A and B with the appropriate sign for index i = 2 yields the negative of the original integrand (compare rows i = 0 and i = 2).

 

Observing that the integral on the RHS can have its own constant of integration  , and bringing the abstract integral to the other side, gives

 

and finally:

 

where C = C′/2.

Higher dimensions

Integration by parts can be extended to functions of several variables by applying a version of the fundamental theorem of calculus to an appropriate product rule. There are several such pairings possible in multivariate calculus, involving a scalar-valued function u and vector-valued function (vector field) V.[7]

The product rule for divergence states:

 

Suppose   is an open bounded subset of   with a piecewise smooth boundary  . Integrating over   with respect to the standard volume form  , and applying the divergence theorem, gives:

 

where   is the outward unit normal vector to the boundary, integrated with respect to its standard Riemannian volume form  . Rearranging gives:

 

or in other words

 
The regularity requirements of the theorem can be relaxed. For instance, the boundary   need only be Lipschitz continuous, and the functions u, v need only lie in the Sobolev space H1(Ω).

Green's first identity

Consider the continuously differentiable vector fields   and  , where  is the i-th standard basis vector for  . Now apply the above integration by parts to each   times the vector field  :

 

Summing over i gives a new integration by parts formula:

 

The case  , where  , is known as the first of Green's identities:

 

See also

Notes

  1. ^ "Brook Taylor". History.MCS.St-Andrews.ac.uk. Retrieved May 25, 2018.
  2. ^ . Stetson.edu. Archived from the original on January 3, 2018. Retrieved May 25, 2018.
  3. ^ "Integration by parts". Encyclopedia of Mathematics.
  4. ^ Kasube, Herbert E. (1983). "A Technique for Integration by Parts". The American Mathematical Monthly. 90 (3): 210–211. doi:10.2307/2975556. JSTOR 2975556.
  5. ^ Thomas, G. B.; Finney, R. L. (1988). Calculus and Analytic Geometry (7th ed.). Reading, MA: Addison-Wesley. ISBN 0-201-17069-8.
  6. ^ Horowitz, David (1990). "Tabular Integration by Parts" (PDF). The College Mathematics Journal. 21 (4): 307–311. doi:10.2307/2686368. JSTOR 2686368.
  7. ^ Rogers, Robert C. (September 29, 2011). "The Calculus of Several Variables" (PDF).

Further reading

  • Louis Brand (10 October 2013). Advanced Calculus: An Introduction to Classical Analysis. Courier Corporation. pp. 267–. ISBN 978-0-486-15799-3.
  • Hoffmann, Laurence D.; Bradley, Gerald L. (2004). Calculus for Business, Economics, and the Social and Life Sciences (8th ed.). pp. 450–464. ISBN 0-07-242432-X.
  • Willard, Stephen (1976). Calculus and its Applications. Boston: Prindle, Weber & Schmidt. pp. 193–214. ISBN 0-87150-203-8.
  • Washington, Allyn J. (1966). Technical Calculus with Analytic Geometry. Reading: Addison-Wesley. pp. 218–245. ISBN 0-8465-8603-7.

External links

integration, parts, calculus, more, generally, mathematical, analysis, integration, parts, partial, integration, process, that, finds, integral, product, functions, terms, integral, product, their, derivative, antiderivative, frequently, used, transform, antid. In calculus and more generally in mathematical analysis integration by parts or partial integration is a process that finds the integral of a product of functions in terms of the integral of the product of their derivative and antiderivative It is frequently used to transform the antiderivative of a product of functions into an antiderivative for which a solution can be more easily found The rule can be thought of as an integral version of the product rule of differentiation The integration by parts formula states a b u x v x d x u x v x a b a b u x v x d x u b v b u a v a a b u x v x d x displaystyle begin aligned int a b u x v x dx amp Big u x v x Big a b int a b u x v x dx amp u b v b u a v a int a b u x v x dx end aligned Or letting u u x displaystyle u u x and d u u x d x displaystyle du u x dx while v v x displaystyle v v x and d v v x d x displaystyle dv v x dx the formula can be written more compactly u d v u v v d u displaystyle int u dv uv int v du Mathematician Brook Taylor discovered integration by parts first publishing the idea in 1715 1 2 More general formulations of integration by parts exist for the Riemann Stieltjes and Lebesgue Stieltjes integrals The discrete analogue for sequences is called summation by parts Contents 1 Theorem 1 1 Product of two functions 1 2 Validity for less smooth functions 1 3 Product of many functions 2 Visualization 3 Applications 3 1 Finding antiderivatives 3 1 1 Polynomials and trigonometric functions 3 1 2 Exponentials and trigonometric functions 3 1 3 Functions multiplied by unity 3 1 4 LIATE rule 3 2 Wallis product 3 3 Gamma function identity 3 4 Use in harmonic analysis 3 4 1 Fourier transform of derivative 3 4 2 Decay of Fourier transform 3 5 Use in operator theory 3 6 Other applications 4 Repeated integration by parts 4 1 Tabular integration by parts 5 Higher dimensions 5 1 Green s first identity 6 See also 7 Notes 8 Further reading 9 External linksTheorem EditProduct of two functions Edit The theorem can be derived as follows For two continuously differentiable functions u x and v x the product rule states u x v x v x u x u x v x displaystyle Big u x v x Big v x u x u x v x Integrating both sides with respect to x u x v x d x u x v x d x u x v x d x displaystyle int Big u x v x Big dx int u x v x dx int u x v x dx and noting that an indefinite integral is an antiderivative givesu x v x u x v x d x u x v x d x displaystyle u x v x int u x v x dx int u x v x dx where we neglect writing the constant of integration This yields the formula for integration by parts u x v x d x u x v x u x v x d x displaystyle int u x v x dx u x v x int u x v x dx or in terms of the differentials d u u x d x displaystyle du u x dx d v v x d x displaystyle dv v x dx quad u x d v u x v x v x d u displaystyle int u x dv u x v x int v x du This is to be understood as an equality of functions with an unspecified constant added to each side Taking the difference of each side between two values x a and x b and applying the fundamental theorem of calculus gives the definite integral version a b u x v x d x u b v b u a v a a b u x v x d x displaystyle int a b u x v x dx u b v b u a v a int a b u x v x dx The original integral uv dx contains the derivative v to apply the theorem one must find v the antiderivative of v then evaluate the resulting integral vu dx Validity for less smooth functions Edit It is not necessary for u and v to be continuously differentiable Integration by parts works if u is absolutely continuous and the function designated v is Lebesgue integrable but not necessarily continuous 3 If v has a point of discontinuity then its antiderivative v may not have a derivative at that point If the interval of integration is not compact then it is not necessary for u to be absolutely continuous in the whole interval or for v to be Lebesgue integrable in the interval as a couple of examples in which u and v are continuous and continuously differentiable will show For instance ifu x e x x 2 v x e x displaystyle u x e x x 2 v x e x u is not absolutely continuous on the interval 1 but nevertheless 1 u x v x d x u x v x 1 1 u x v x d x displaystyle int 1 infty u x v x dx Big u x v x Big 1 infty int 1 infty u x v x dx so long as u x v x 1 displaystyle left u x v x right 1 infty is taken to mean the limit of u L v L u 1 v 1 displaystyle u L v L u 1 v 1 as L displaystyle L to infty and so long as the two terms on the right hand side are finite This is only true if we choose v x e x displaystyle v x e x Similarly ifu x e x v x x 1 sin x displaystyle u x e x v x x 1 sin x v is not Lebesgue integrable on the interval 1 but nevertheless 1 u x v x d x u x v x 1 1 u x v x d x displaystyle int 1 infty u x v x dx Big u x v x Big 1 infty int 1 infty u x v x dx with the same interpretation One can also easily come up with similar examples in which u and v are not continuously differentiable Further if f x displaystyle f x is a function of bounded variation on the segment a b displaystyle a b and f x displaystyle varphi x is differentiable on a b displaystyle a b then a b f x f x d x f x d x a b x f x displaystyle int a b f x varphi x dx int infty infty widetilde varphi x d widetilde chi a b x widetilde f x where d x a b x f x displaystyle d chi a b x widetilde f x denotes the signed measure corresponding to the function of bounded variation x a b x f x displaystyle chi a b x f x and functions f f displaystyle widetilde f widetilde varphi are extensions of f f displaystyle f varphi to R displaystyle mathbb R which are respectively of bounded variation and differentiable citation needed Product of many functions Edit Integrating the product rule for three multiplied functions u x v x w x gives a similar result a b u v d w u v w a b a b u w d v a b v w d u displaystyle int a b uv dw Big uvw Big a b int a b uw dv int a b vw du In general for n factors i 1 n u i x j 1 n u j x i j n u i x displaystyle left prod i 1 n u i x right sum j 1 n u j x prod i neq j n u i x which leads to i 1 n u i x a b j 1 n a b u j x i j n u i x displaystyle left prod i 1 n u i x right a b sum j 1 n int a b u j x prod i neq j n u i x Visualization Edit Graphical interpretation of the theorem The pictured curve is parametrized by the variable t Consider a parametric curve by x y f t g t Assuming that the curve is locally one to one and integrable we can define x y f g 1 y displaystyle x y f g 1 y y x g f 1 x displaystyle y x g f 1 x The area of the blue region is A 1 y 1 y 2 x y d y displaystyle A 1 int y 1 y 2 x y dy Similarly the area of the red region is A 2 x 1 x 2 y x d x displaystyle A 2 int x 1 x 2 y x dx The total area A1 A2 is equal to the area of the bigger rectangle x2y2 minus the area of the smaller one x1y1 y 1 y 2 x y d y A 1 x 1 x 2 y x d x A 2 x y x x 1 x 2 y x y y 1 y 2 displaystyle overbrace int y 1 y 2 x y dy A 1 overbrace int x 1 x 2 y x dx A 2 biggl x cdot y x biggl x 1 x 2 biggl y cdot x y biggl y 1 y 2 Or in terms of t t 1 t 2 x t d y t t 1 t 2 y t d x t x t y t t 1 t 2 displaystyle int t 1 t 2 x t dy t int t 1 t 2 y t dx t biggl x t y t biggl t 1 t 2 Or in terms of indefinite integrals this can be written as x d y y d x x y displaystyle int x dy int y dx xy Rearranging x d y x y y d x displaystyle int x dy xy int y dx Thus integration by parts may be thought of as deriving the area of the blue region from the area of rectangles and that of the red region This visualization also explains why integration by parts may help find the integral of an inverse function f 1 x when the integral of the function f x is known Indeed the functions x y and y x are inverses and the integral x dy may be calculated as above from knowing the integral y dx In particular this explains use of integration by parts to integrate logarithm and inverse trigonometric functions In fact if f displaystyle f is a differentiable one to one function on an interval then integration by parts can be used to derive a formula for the integral of f 1 displaystyle f 1 in terms of the integral of f displaystyle f This is demonstrated in the article Integral of inverse functions Applications EditFinding antiderivatives Edit 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 this single function into a product of two functions u x v x such that the residual integral from the integration by parts formula is easier to evaluate than the single function The following form is useful in illustrating the best strategy to take u v d x u v d x u v d x d x displaystyle int uv dx u int v dx int left u int v dx right dx On the right hand side u is differentiated and v is integrated consequently it is useful to choose u as a function that simplifies when differentiated or to choose v as a function that simplifies when integrated As a simple example consider ln x x 2 d x displaystyle int frac ln x x 2 dx Since the derivative of ln x is 1 x one makes ln x part u since the antiderivative of 1 x2 is 1 x one makes 1 x2 dx part dv The formula now yields ln x x 2 d x ln x x 1 x 1 x d x displaystyle int frac ln x x 2 dx frac ln x x int biggl frac 1 x biggr biggl frac 1 x biggr dx The antiderivative of 1 x2 can be found with the power rule and is 1 x Alternatively one may choose u and v such that the product u v dx simplifies due to cancellation For example suppose one wishes to integrate sec 2 x ln sin x d x displaystyle int sec 2 x cdot ln Big bigl sin x bigr Big dx If we choose u x ln sin x and v x sec2x then u differentiates to 1 tan x using the chain rule and v integrates to tan x so the formula gives sec 2 x ln sin x d x tan x ln sin x tan x 1 tan x d x displaystyle int sec 2 x cdot ln Big bigl sin x bigr Big dx tan x cdot ln Big bigl sin x bigr Big int tan x cdot frac 1 tan x dx The integrand simplifies to 1 so the antiderivative is x Finding a simplifying combination frequently involves experimentation 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 Polynomials and trigonometric functions Edit In order to calculate I x cos x d x displaystyle I int x cos x dx let u x d u d x displaystyle u x Rightarrow du dx d v cos x d x v cos x d x sin x displaystyle dv cos x dx Rightarrow v int cos x dx sin x then x cos x d x u d v u v v d u x sin x sin x d x x sin x cos x C displaystyle begin aligned int x cos x dx amp int u dv amp u cdot v int v du amp x sin x int sin x dx amp x sin x cos x C end aligned where C is a constant of integration For higher powers of x in the form x n e x d x x n sin x d x x n cos x d x displaystyle int x n e x dx int x n sin x dx int x n cos x dx repeatedly using integration by parts can evaluate integrals such as these each application of the theorem lowers the power of x by one Exponentials and trigonometric functions Edit See also Integration using Euler s formula An example commonly used to examine the workings of integration by parts is I e x cos x d x displaystyle I int e x cos x dx Here integration by parts is performed twice First let u cos x d u sin x d x displaystyle u cos x Rightarrow du sin x dx d v e x d x v e x d x e x displaystyle dv e x dx Rightarrow v int e x dx e x then e x cos x d x e x cos x e x sin x d x displaystyle int e x cos x dx e x cos x int e x sin x dx Now to evaluate the remaining integral we use integration by parts again with u sin x d u cos x d x displaystyle u sin x Rightarrow du cos x dx d v e x d x v e x d x e x displaystyle dv e x dx Rightarrow v int e x dx e x Then e x sin x d x e x sin x e x cos x d x displaystyle int e x sin x dx e x sin x int e x cos x dx Putting these together e x cos x d x e x cos x e x sin x e x cos x d x displaystyle int e x cos x dx e x cos x e x sin x int e x cos x dx The same integral shows up on both sides of this equation The integral can simply be added to both sides to get 2 e x cos x d x e x sin x cos x C displaystyle 2 int e x cos x dx e x bigl sin x cos x bigr C which rearranges to e x cos x d x 1 2 e x sin x cos x C displaystyle int e x cos x dx frac 1 2 e x bigl sin x cos x bigr C where again C and C C 2 is a constant of integration A similar method is used to find the integral of secant cubed Functions multiplied by unity Edit 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 I ln x 1 d x displaystyle I int ln x cdot 1 dx Let u ln x d u d x x displaystyle u ln x Rightarrow du frac dx x d v d x v x displaystyle dv dx Rightarrow v x then ln x d x x ln x x x d x x ln x 1 d x x ln x x C displaystyle begin aligned int ln x dx amp x ln x int frac x x dx amp x ln x int 1 dx amp x ln x x C end aligned where C is the constant of integration The second example is the inverse tangent function arctan x I arctan x d x displaystyle I int arctan x dx Rewrite this as arctan x 1 d x displaystyle int arctan x cdot 1 dx Now let u arctan x d u d x 1 x 2 displaystyle u arctan x Rightarrow du frac dx 1 x 2 d v d x v x displaystyle dv dx Rightarrow v x then arctan x d x x arctan x x 1 x 2 d x x arctan x ln 1 x 2 2 C displaystyle begin aligned int arctan x dx amp x arctan x int frac x 1 x 2 dx 8pt amp x arctan x frac ln 1 x 2 2 C end aligned using a combination of the inverse chain rule method and the natural logarithm integral condition LIATE rule Edit A rule of thumb has been proposed consisting of choosing as u the function that comes first in the following list 4 L logarithmic functions ln x log b x displaystyle ln x log b x etc I inverse trigonometric functions including hyperbolic analogues arctan x arcsec x arsinh x displaystyle arctan x operatorname arcsec x operatorname arsinh x etc A algebraic functions x 2 3 x 50 displaystyle x 2 3x 50 etc T trigonometric functions including hyperbolic analogues sin x tan x sech x displaystyle sin x tan x operatorname sech x etc E exponential functions e x 19 x displaystyle e x 19 x etc The function which is to be dv is whichever comes last in the list The reason is that functions lower on the list generally have easier antiderivatives than the functions above them The rule is sometimes written as DETAIL where D stands for dv and the top of the list is the function chosen to be dv To demonstrate the LIATE rule consider the integral x cos x d x displaystyle int x cdot cos x dx Following the LIATE rule u x and dv cos x dx hence du dx and v sin x which makes the integral become x sin x 1 sin x d x displaystyle x cdot sin x int 1 sin x dx which equals x sin x cos x C displaystyle x cdot sin x cos x C 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 dx as dv we would have the integral x 2 2 cos x x 2 2 sin x d x displaystyle frac x 2 2 cos x int frac x 2 2 sin x dx 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 x 3 e x 2 d x displaystyle int x 3 e x 2 dx one would set u x 2 d v x e x 2 d x displaystyle u x 2 quad dv x cdot e x 2 dx so that d u 2 x d x v e x 2 2 displaystyle du 2x dx quad v frac e x 2 2 Then x 3 e x 2 d x x 2 x e x 2 d x u d v u v v d u x 2 e x 2 2 x e x 2 d x displaystyle int x 3 e x 2 dx int left x 2 right left xe x 2 right dx int u dv uv int v du frac x 2 e x 2 2 int xe x 2 dx Finally this results in x 3 e x 2 d x e x 2 x 2 1 2 C displaystyle int x 3 e x 2 dx frac e x 2 left x 2 1 right 2 C Integration by parts is often used as a tool to prove theorems in mathematical analysis Wallis product Edit The Wallis infinite product for p displaystyle pi p 2 n 1 4 n 2 4 n 2 1 n 1 2 n 2 n 1 2 n 2 n 1 2 1 2 3 4 3 4 5 6 5 6 7 8 7 8 9 displaystyle begin aligned frac pi 2 amp prod n 1 infty frac 4n 2 4n 2 1 prod n 1 infty left frac 2n 2n 1 cdot frac 2n 2n 1 right 6pt amp Big frac 2 1 cdot frac 2 3 Big cdot Big frac 4 3 cdot frac 4 5 Big cdot Big frac 6 5 cdot frac 6 7 Big cdot Big frac 8 7 cdot frac 8 9 Big cdot cdots end aligned may be derived using integration by parts Gamma function identity Edit The gamma function is an example of a special function defined as an improper integral for z gt 0 displaystyle z gt 0 Integration by parts illustrates it to be an extension of the factorial function G z 0 e x x z 1 d x 0 x z 1 d e x e x x z 1 0 0 e x d x z 1 0 0 z 1 x z 2 e x d x z 1 G z 1 displaystyle begin aligned Gamma z amp int 0 infty e x x z 1 dx 6pt amp int 0 infty x z 1 d left e x right 6pt amp Biggl e x x z 1 Biggl 0 infty int 0 infty e x d left x z 1 right 6pt amp 0 int 0 infty left z 1 right x z 2 e x dx 6pt amp z 1 Gamma z 1 end aligned Since G 1 0 e x d x 1 displaystyle Gamma 1 int 0 infty e x dx 1 when z displaystyle z is a natural number that is z n N displaystyle z n in mathbb N applying this formula repeatedly gives the factorial G n 1 n displaystyle Gamma n 1 n Use in harmonic analysis Edit Integration by parts is often used in harmonic analysis particularly Fourier analysis to show that quickly oscillating integrals with sufficiently smooth integrands decay quickly The most common example of this is its use in showing that the decay of function s Fourier transform depends on the smoothness of that function as described below Fourier transform of derivative Edit If f is a k times continuously differentiable function and all derivatives up to the kth one decay to zero at infinity then its Fourier transform satisfies F f k 3 2 p i 3 k F f 3 displaystyle mathcal F f k xi 2 pi i xi k mathcal F f xi where f k is the kth derivative of f The exact constant on the right depends on the convention of the Fourier transform used This is proved by noting that d d y e 2 p i y 3 2 p i 3 e 2 p i y 3 displaystyle frac d dy e 2 pi iy xi 2 pi i xi e 2 pi iy xi so using integration by parts on the Fourier transform of the derivative we get F f 3 e 2 p i y 3 f y d y e 2 p i y 3 f y 2 p i 3 e 2 p i y 3 f y d y 2 p i 3 e 2 p i y 3 f y d y 2 p i 3 F f 3 displaystyle begin aligned mathcal F f xi amp int infty infty e 2 pi iy xi f y dy amp left e 2 pi iy xi f y right infty infty int infty infty 2 pi i xi e 2 pi iy xi f y dy 5pt amp 2 pi i xi int infty infty e 2 pi iy xi f y dy 5pt amp 2 pi i xi mathcal F f xi end aligned Applying this inductively gives the result for general k A similar method can be used to find the Laplace transform of a derivative of a function Decay of Fourier transform Edit The above result tells us about the decay of the Fourier transform since it follows that if f and f k are integrable then F f 3 I f 1 2 p 3 k where I f f y f k y d y displaystyle vert mathcal F f xi vert leq frac I f 1 vert 2 pi xi vert k text where I f int infty infty Bigl vert f y vert vert f k y vert Bigr dy In other words if f satisfies these conditions then its Fourier transform decays at infinity at least as quickly as 1 3 k In particular if k 2 then the Fourier transform is integrable The proof uses the fact which is immediate from the definition of the Fourier transform that F f 3 f y d y displaystyle vert mathcal F f xi vert leq int infty infty vert f y vert dy Using the same idea on the equality stated at the start of this subsection gives 2 p i 3 k F f 3 f k y d y displaystyle vert 2 pi i xi k mathcal F f xi vert leq int infty infty vert f k y vert dy Summing these two inequalities and then dividing by 1 2p 3k gives the stated inequality Use in operator theory Edit One use of integration by parts in operator theory is that it shows that the where is the Laplace operator is a positive operator on L2 see Lp space If f is smooth and compactly supported then using integration by parts we have D f f L 2 f x f x d x f x f x f x f x d x f x 2 d x 0 displaystyle begin aligned langle Delta f f rangle L 2 amp int infty infty f x overline f x dx 5pt amp left f x overline f x right infty infty int infty infty f x overline f x dx 5pt amp int infty infty vert f x vert 2 dx geq 0 end aligned Other applications Edit Determining boundary conditions in Sturm Liouville theory Deriving the Euler Lagrange equation in the calculus of variationsRepeated integration by parts EditConsidering a second derivative of v displaystyle v in the integral on the LHS of the formula for partial integration suggests a repeated application to the integral on the RHS u v d x u v u v d x u v u v u v d x displaystyle int uv dx uv int u v dx uv left u v int u v dx right Extending this concept of repeated partial integration to derivatives of degree n leads to u 0 v n d x u 0 v n 1 u 1 v n 2 u 2 v n 3 1 n 1 u n 1 v 0 1 n u n v 0 d x k 0 n 1 1 k u k v n 1 k 1 n u n v 0 d x displaystyle begin aligned int u 0 v n dx amp u 0 v n 1 u 1 v n 2 u 2 v n 3 cdots 1 n 1 u n 1 v 0 1 n int u n v 0 dx 5pt amp sum k 0 n 1 1 k u k v n 1 k 1 n int u n v 0 dx end aligned This concept may be useful when the successive integrals of v n displaystyle v n are readily available e g plain exponentials or sine and cosine as in Laplace or Fourier transforms and when the n th derivative of u displaystyle u vanishes e g as a polynomial function with degree n 1 displaystyle n 1 The latter condition stops the repeating of partial integration because the RHS integral vanishes In the course of the above repetition of partial integrations the integrals u 0 v n d x displaystyle int u 0 v n dx quad and u ℓ v n ℓ d x displaystyle quad int u ell v n ell dx quad and u m v n m d x for 1 m ℓ n displaystyle quad int u m v n m dx quad text for 1 leq m ell leq n get related This may be interpreted as arbitrarily shifting derivatives between v displaystyle v and u displaystyle u within the integrand and proves useful too see Rodrigues formula Tabular integration by parts Edit The essential process of the above formula can be summarized in a table the resulting method is called tabular integration 5 and was featured in the film Stand and Deliver 1988 6 For example consider the integral x 3 cos x d x displaystyle int x 3 cos x dx quad and take u 0 x 3 v n cos x displaystyle quad u 0 x 3 quad v n cos x Begin to list in column A the function u 0 x 3 displaystyle u 0 x 3 and its subsequent derivatives u i displaystyle u i until zero is reached Then list in column B the function v n cos x displaystyle v n cos x and its subsequent integrals v n i displaystyle v n i until the size of column B is the same as that of column A The result is as follows i Sign A derivatives u i B integrals v n i 0 x 3 displaystyle x 3 cos x displaystyle cos x 1 3 x 2 displaystyle 3x 2 sin x displaystyle sin x 2 6 x displaystyle 6x cos x displaystyle cos x 3 6 displaystyle 6 sin x displaystyle sin x 4 0 displaystyle 0 cos x displaystyle cos x The product of the entries in row i of columns A and B together with the respective sign give the relevant integrals in step i in the course of repeated integration by parts Step i 0 yields the original integral For the complete result in step i gt 0 the i th integral must be added to all the previous products 0 j lt i of the j th entry of column A and the j 1 st entry of column B i e multiply 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 the given j th sign This process comes to a natural halt when the product which yields the integral is zero i 4 in the example The complete result is the following with the alternating signs in each term 1 x 3 sin x j 0 1 3 x 2 cos x j 1 1 6 x sin x j 2 1 6 cos x j 3 1 0 cos x d x i 4 C displaystyle underbrace 1 x 3 sin x j 0 underbrace 1 3x 2 cos x j 1 underbrace 1 6x sin x j 2 underbrace 1 6 cos x j 3 underbrace int 1 0 cos x dx i 4 to C This yields x 3 cos x d x step 0 x 3 sin x 3 x 2 cos x 6 x sin x 6 cos x C displaystyle underbrace int x 3 cos x dx text step 0 x 3 sin x 3x 2 cos x 6x sin x 6 cos x C The repeated partial integration also turns out useful when in the course of respectively differentiating and integrating the functions u i displaystyle u i and v n i displaystyle v n i their product results in a multiple of the original integrand In this case the repetition may also be terminated with this index i This can happen expectably with exponentials and trigonometric functions As an example consider e x cos x d x displaystyle int e x cos x dx i Sign A derivatives u i B integrals v n i 0 e x displaystyle e x cos x displaystyle cos x 1 e x displaystyle e x sin x displaystyle sin x 2 e x displaystyle e x cos x displaystyle cos x In this case the product of the terms in columns A and B with the appropriate sign for index i 2 yields the negative of the original integrand compare rows i 0 and i 2 e x cos x d x step 0 1 e x sin x j 0 1 e x cos x j 1 1 e x cos x d x i 2 displaystyle underbrace int e x cos x dx text step 0 underbrace 1 e x sin x j 0 underbrace 1 e x cos x j 1 underbrace int 1 e x cos x dx i 2 Observing that the integral on the RHS can have its own constant of integration C displaystyle C and bringing the abstract integral to the other side gives 2 e x cos x d x e x sin x e x cos x C displaystyle 2 int e x cos x dx e x sin x e x cos x C and finally e x cos x d x 1 2 e x sin x cos x C displaystyle int e x cos x dx frac 1 2 left e x sin x cos x right C where C C 2 Higher dimensions EditIntegration by parts can be extended to functions of several variables by applying a version of the fundamental theorem of calculus to an appropriate product rule There are several such pairings possible in multivariate calculus involving a scalar valued function u and vector valued function vector field V 7 The product rule for divergence states u V u V u V displaystyle nabla cdot u mathbf V u nabla cdot mathbf V nabla u cdot mathbf V Suppose W displaystyle Omega is an open bounded subset of R n displaystyle mathbb R n with a piecewise smooth boundary G W displaystyle Gamma partial Omega Integrating over W displaystyle Omega with respect to the standard volume form d W displaystyle d Omega and applying the divergence theorem gives G u V n d G W u V d W W u V d W W u V d W displaystyle int Gamma u mathbf V cdot hat mathbf n d Gamma int Omega nabla cdot u mathbf V d Omega int Omega u nabla cdot mathbf V d Omega int Omega nabla u cdot mathbf V d Omega where n displaystyle hat mathbf n is the outward unit normal vector to the boundary integrated with respect to its standard Riemannian volume form d G displaystyle d Gamma Rearranging gives W u V d W G u V n d G W u V d W displaystyle int Omega u nabla cdot mathbf V d Omega int Gamma u mathbf V cdot hat mathbf n d Gamma int Omega nabla u cdot mathbf V d Omega or in other words W u div V d W G u V n d G W grad u V d W displaystyle int Omega u operatorname div mathbf V d Omega int Gamma u mathbf V cdot hat mathbf n d Gamma int Omega operatorname grad u cdot mathbf V d Omega The regularity requirements of the theorem can be relaxed For instance the boundary G W displaystyle Gamma partial Omega need only be Lipschitz continuous and the functions u v need only lie in the Sobolev space H1 W Green s first identity Edit Consider the continuously differentiable vector fields U u 1 e 1 u n e n displaystyle mathbf U u 1 mathbf e 1 cdots u n mathbf e n and v e 1 v e n displaystyle v mathbf e 1 ldots v mathbf e n where e i displaystyle mathbf e i is the i th standard basis vector for i 1 n displaystyle i 1 ldots n Now apply the above integration by parts to each u i displaystyle u i times the vector field v e i displaystyle v mathbf e i W u i v x i d W G u i v e i n d G W u i x i v d W displaystyle int Omega u i frac partial v partial x i d Omega int Gamma u i v mathbf e i cdot hat mathbf n d Gamma int Omega frac partial u i partial x i v d Omega Summing over i gives a new integration by parts formula W U v d W G v U n d G W v U d W displaystyle int Omega mathbf U cdot nabla v d Omega int Gamma v mathbf U cdot hat mathbf n d Gamma int Omega v nabla cdot mathbf U d Omega The case U u displaystyle mathbf U nabla u where u C 2 W displaystyle u in C 2 bar Omega is known as the first of Green s identities W u v d W G v u n d G W v 2 u d W displaystyle int Omega nabla u cdot nabla v d Omega int Gamma v nabla u cdot hat mathbf n d Gamma int Omega v nabla 2 u d Omega See also EditIntegration by parts for the Lebesgue Stieltjes integral Integration by parts for semimartingales involving their quadratic covariation Integration by substitution Legendre transformationNotes Edit Brook Taylor History MCS St Andrews ac uk Retrieved May 25 2018 Brook Taylor Stetson edu Archived from the original on January 3 2018 Retrieved May 25 2018 Integration by parts Encyclopedia of Mathematics Kasube Herbert E 1983 A Technique for Integration by Parts The American Mathematical Monthly 90 3 210 211 doi 10 2307 2975556 JSTOR 2975556 Thomas G B Finney R L 1988 Calculus and Analytic Geometry 7th ed Reading MA Addison Wesley ISBN 0 201 17069 8 Horowitz David 1990 Tabular Integration by Parts PDF The College Mathematics Journal 21 4 307 311 doi 10 2307 2686368 JSTOR 2686368 Rogers Robert C September 29 2011 The Calculus of Several Variables PDF Further reading EditLouis Brand 10 October 2013 Advanced Calculus An Introduction to Classical Analysis Courier Corporation pp 267 ISBN 978 0 486 15799 3 Hoffmann Laurence D Bradley Gerald L 2004 Calculus for Business Economics and the Social and Life Sciences 8th ed pp 450 464 ISBN 0 07 242432 X Willard Stephen 1976 Calculus and its Applications Boston Prindle Weber amp Schmidt pp 193 214 ISBN 0 87150 203 8 Washington Allyn J 1966 Technical Calculus with Analytic Geometry Reading Addison Wesley pp 218 245 ISBN 0 8465 8603 7 External links Edit The Wikibook Calculus has a page on the topic of Integration by parts Integration by parts Encyclopedia of Mathematics EMS Press 2001 1994 Integration by parts from MathWorld Retrieved from https en wikipedia org w index php title Integration by parts amp oldid 1135330725, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.