fbpx
Wikipedia

Leibniz integral rule

In calculus, the Leibniz integral rule for differentiation under the integral sign states that for an integral of the form

where and the integrands are functions dependent on the derivative of this integral is expressible as

where the partial derivative indicates that inside the integral, only the variation of with is considered in taking the derivative.[1] It is named after Gottfried Leibniz.

In the special case where the functions and are constants and with values that do not depend on this simplifies to:

If is constant and , which is another common situation (for example, in the proof of Cauchy's repeated integration formula), the Leibniz integral rule becomes:

This important result may, under certain conditions, be used to interchange the integral and partial differential operators, and is particularly useful in the differentiation of integral transforms. An example of such is the moment generating function in probability theory, a variation of the Laplace transform, which can be differentiated to generate the moments of a random variable. Whether Leibniz's integral rule applies is essentially a question about the interchange of limits.

General form: differentiation under the integral sign Edit

Theorem — Let   be a function such that both   and its partial derivative   are continuous in   and   in some region of the  -plane, including     Also suppose that the functions   and   are both continuous and both have continuous derivatives for   Then, for  

 

The right hand side may also be written using Lagrange's notation as:  

Stronger versions of the theorem only require that the partial derivative exist almost everywhere, and not that it be continuous.[2] This formula is the general form of the Leibniz integral rule and can be derived using the fundamental theorem of calculus. The (first) fundamental theorem of calculus is just the particular case of the above formula where   is constant,   and   does not depend on  

If both upper and lower limits are taken as constants, then the formula takes the shape of an operator equation:

 
where   is the partial derivative with respect to   and   is the integral operator with respect to   over a fixed interval. That is, it is related to the symmetry of second derivatives, but involving integrals as well as derivatives. This case is also known as the Leibniz integral rule.

The following three basic theorems on the interchange of limits are essentially equivalent:

  • the interchange of a derivative and an integral (differentiation under the integral sign; i.e., Leibniz integral rule);
  • the change of order of partial derivatives;
  • the change of order of integration (integration under the integral sign; i.e., Fubini's theorem).

Three-dimensional, time-dependent case Edit

 
Figure 1: A vector field F(r, t) defined throughout space, and a surface Σ bounded by curve ∂Σ moving with velocity v over which the field is integrated.

A Leibniz integral rule for a two dimensional surface moving in three dimensional space is[3][4]

 

where:

  • F(r, t) is a vector field at the spatial position r at time t,
  • Σ is a surface bounded by the closed curve ∂Σ,
  • dA is a vector element of the surface Σ,
  • ds is a vector element of the curve ∂Σ,
  • v is the velocity of movement of the region Σ,
  • ∇⋅ is the vector divergence,
  • × is the vector cross product,
  • The double integrals are surface integrals over the surface Σ, and the line integral is over the bounding curve ∂Σ.

Higher dimensions Edit

The Leibniz integral rule can be extended to multidimensional integrals. In two and three dimensions, this rule is better known from the field of fluid dynamics as the Reynolds transport theorem:

 

where   is a scalar function, D(t) and D(t) denote a time-varying connected region of R3 and its boundary, respectively,   is the Eulerian velocity of the boundary (see Lagrangian and Eulerian coordinates) and dΣ = n dS is the unit normal component of the surface element.

The general statement of the Leibniz integral rule requires concepts from differential geometry, specifically differential forms, exterior derivatives, wedge products and interior products. With those tools, the Leibniz integral rule in n dimensions is[4]

 
where Ω(t) is a time-varying domain of integration, ω is a p-form,   is the vector field of the velocity,   denotes the interior product with  , dxω is the exterior derivative of ω with respect to the space variables only and   is the time derivative of ω.

However, all of these identities can be derived from a most general statement about Lie derivatives:

 

Here, the ambient manifold on which the differential form   lives includes both space and time.

  •   is the region of integration (a submanifold) at a given instant (it does not depend on  , since its parametrization as a submanifold defines its position in time),
  •   is the Lie derivative,
  •   is the spacetime vector field obtained from adding the unitary vector field in the direction of time to the purely spatial vector field   from the previous formulas (i.e,   is the spacetime velocity of  ),
  •   is a diffeomorphism from the one-parameter group generated by the flow of  , and
  •   is the image of   under such diffeomorphism.

Something remarkable about this form, is that it can account for the case when   changes its shape and size over time, since such deformations are fully determined by  .

Measure theory statement Edit

Let   be an open subset of  , and   be a measure space. Suppose   satisfies the following conditions:[5][6][2]

  1.   is a Lebesgue-integrable function of   for each  .
  2. For almost all   , the partial derivative   exists for all  .
  3. There is an integrable function   such that   for all   and almost every  .

Then, for all  ,

 

The proof relies on the dominated convergence theorem and the mean value theorem (details below).

Proofs Edit

Proof of basic form Edit

We first prove the case of constant limits of integration a and b.

We use Fubini's theorem to change the order of integration. For every x and h, such that h > 0 and both x and x +h are within [x0,x1], we have:

 

Note that the integrals at hand are well defined since   is continuous at the closed rectangle   and thus also uniformly continuous there; thus its integrals by either dt or dx are continuous in the other variable and also integrable by it (essentially this is because for uniformly continuous functions, one may pass the limit through the integration sign, as elaborated below).

Therefore:

 

Where we have defined:

 
(we may replace x0 here by any other point between x0 and x)

F is differentiable with derivative  , so we can take the limit where h approaches zero. For the left hand side this limit is:

 

For the right hand side, we get:

 
And we thus prove the desired result:
 

Another proof using the bounded convergence theorem Edit

If the integrals at hand are Lebesgue integrals, we may use the bounded convergence theorem (valid for these integrals, but not for Riemann integrals) in order to show that the limit can be passed through the integral sign.

Note that this proof is weaker in the sense that it only shows that fx(x,t) is Lebesgue integrable, but not that it is Riemann integrable. In the former (stronger) proof, if f(x,t) is Riemann integrable, then so is fx(x,t) (and thus is obviously also Lebesgue integrable).

Let

 

 

 

 

 

(1)

By the definition of the derivative,

 

 

 

 

 

(2)

Substitute equation (1) into equation (2). The difference of two integrals equals the integral of the difference, and 1/h is a constant, so

 

We now show that the limit can be passed through the integral sign.

We claim that the passage of the limit under the integral sign is valid by the bounded convergence theorem (a corollary of the dominated convergence theorem). For each δ > 0, consider the difference quotient

 
For t fixed, the mean value theorem implies there exists z in the interval [x, x + δ] such that
 
Continuity of fx(x, t) and compactness of the domain together imply that fx(x, t) is bounded. The above application of the mean value theorem therefore gives a uniform (independent of  ) bound on  . The difference quotients converge pointwise to the partial derivative fx by the assumption that the partial derivative exists.

The above argument shows that for every sequence {δn} → 0, the sequence   is uniformly bounded and converges pointwise to fx. The bounded convergence theorem states that if a sequence of functions on a set of finite measure is uniformly bounded and converges pointwise, then passage of the limit under the integral is valid. In particular, the limit and integral may be exchanged for every sequence {δn} → 0. Therefore, the limit as δ → 0 may be passed through the integral sign.

Variable limits form Edit

For a continuous real valued function g of one real variable, and real valued differentiable functions   and   of one real variable,

 

This follows from the chain rule and the First Fundamental Theorem of Calculus. Define

 
and
 
(The lower limit just has to be some number in the domain of  )

Then,   can be written as a composition:  . The Chain Rule then implies that

 
By the First Fundamental Theorem of Calculus,  . Therefore, substituting this result above, we get the desired equation:
 

Note: This form can be particularly useful if the expression to be differentiated is of the form:

 
Because   does not depend on the limits of integration, it may be moved out from under the integral sign, and the above form may be used with the Product rule, i.e.,
 

General form with variable limits Edit

Set

 
where a and b are functions of α that exhibit increments Δa and Δb, respectively, when α is increased by Δα. Then,
 

A form of the mean value theorem,  , where a < ξ < b, may be applied to the first and last integrals of the formula for Δφ above, resulting in

 

Divide by Δα and let Δα → 0. Notice ξ1a and ξ2b. We may pass the limit through the integral sign:

 
again by the bounded convergence theorem. This yields the general form of the Leibniz integral rule,
 

Alternative proof of the general form with variable limits, using the chain rule Edit

The general form of Leibniz's Integral Rule with variable limits can be derived as a consequence of the basic form of Leibniz's Integral Rule, the multivariable chain rule, and the First Fundamental Theorem of Calculus. Suppose   is defined in a rectangle in the   plane, for   and  . Also, assume   and the partial derivative   are both continuous functions on this rectangle. Suppose   are differentiable real valued functions defined on  , with values in   (i.e. for every  ). Now, set

 
and
 

Then, by properties of Definite Integrals, we can write

 

Since the functions   are all differentiable (see the remark at the end of the proof), by the Multivariable Chain Rule, it follows that   is differentiable, and its derivative is given by the formula:

 
 

Now, note that for every  , and for every  , we have that  , because when taking the partial derivative with respect to   of  , we are keeping   fixed in the expression  ; thus the basic form of Leibniz's Integral Rule with constant limits of integration applies. Next, by the First Fundamental Theorem of Calculus, we have that  ; because when taking the partial derivative with respect to   of  , the first variable   is fixed, so the fundamental theorem can indeed be applied.

Substituting these results into the equation for   above gives:

 
as desired.

There is a technical point in the proof above which is worth noting: applying the Chain Rule to   requires that   already be differentiable. This is where we use our assumptions about  . As mentioned above, the partial derivatives of   are given by the formulas   and  . Since   is continuous, its integral is also a continuous function,[7] and since   is also continuous, these two results show that both the partial derivatives of   are continuous. Since continuity of partial derivatives implies differentiability of the function,[8]   is indeed differentiable.

Three-dimensional, time-dependent form Edit

At time t the surface Σ in Figure 1 contains a set of points arranged about a centroid  . The function   can be written as

 
with   independent of time. Variables are shifted to a new frame of reference attached to the moving surface, with origin at  . For a rigidly translating surface, the limits of integration are then independent of time, so:
 
where the limits of integration confining the integral to the region Σ no longer are time dependent so differentiation passes through the integration to act on the integrand only:
 
with the velocity of motion of the surface defined by
 

This equation expresses the material derivative of the field, that is, the derivative with respect to a coordinate system attached to the moving surface. Having found the derivative, variables can be switched back to the original frame of reference. We notice that (see article on curl)

 
and that Stokes theorem equates the surface integral of the curl over Σ with a line integral over ∂Σ:
 

The sign of the line integral is based on the right-hand rule for the choice of direction of line element ds. To establish this sign, for example, suppose the field F points in the positive z-direction, and the surface Σ is a portion of the xy-plane with perimeter ∂Σ. We adopt the normal to Σ to be in the positive z-direction. Positive traversal of ∂Σ is then counterclockwise (right-hand rule with thumb along z-axis). Then the integral on the left-hand side determines a positive flux of F through Σ. Suppose Σ translates in the positive x-direction at velocity v. An element of the boundary of Σ parallel to the y-axis, say ds, sweeps out an area vt × ds in time t. If we integrate around the boundary ∂Σ in a counterclockwise sense, vt × ds points in the negative z-direction on the left side of ∂Σ (where ds points downward), and in the positive z-direction on the right side of ∂Σ (where ds points upward), which makes sense because Σ is moving to the right, adding area on the right and losing it on the left. On that basis, the flux of F is increasing on the right of ∂Σ and decreasing on the left. However, the dot product v × Fds = −F × vds = −Fv × ds. Consequently, the sign of the line integral is taken as negative.

If v is a constant,

 
which is the quoted result. This proof does not consider the possibility of the surface deforming as it moves.

Alternative derivation Edit

Lemma. One has:

 

Proof. From the proof of the fundamental theorem of calculus,

 
and
 

Suppose a and b are constant, and that f(x) involves a parameter α which is constant in the integration but may vary to form different integrals. Assume that f(x, α) is a continuous function of x and α in the compact set {(x, α) : α0αα1 and axb}, and that the partial derivative fα(x, α) exists and is continuous. If one defines:

 
then   may be differentiated with respect to α by differentiating under the integral sign, i.e.,
 

By the Heine–Cantor theorem it is uniformly continuous in that set. In other words, for any ε > 0 there exists Δα such that for all values of x in [a, b],

 

On the other hand,

 

Hence φ(α) is a continuous function.

Similarly if   exists and is continuous, then for all ε > 0 there exists Δα such that:

 

Therefore,

 
where
 

Now, ε → 0 as Δα → 0, so

 

This is the formula we set out to prove.

Now, suppose

 
where a and b are functions of α which take increments Δa and Δb, respectively, when α is increased by Δα. Then,
 

A form of the mean value theorem,   where a < ξ < b, can be applied to the first and last integrals of the formula for Δφ above, resulting in

 

Dividing by Δα, letting Δα → 0, noticing ξ1a and ξ2b and using the above derivation for

 
yields
 

This is the general form of the Leibniz integral rule.

Examples Edit

Example 1: Fixed limits Edit

Consider the function

 

The function under the integral sign is not continuous at the point (x, α) = (0, 0), and the function φ(α) has a discontinuity at α = 0 because φ(α) approaches ±π/2 as α → 0±.

If we differentiate φ(α) with respect to α under the integral sign, we get

 
which is, of course, true for all values of α except α = 0. This may be integrated (with respect to α) to find
 

Example 2: Variable limits Edit

An example with variable limits:

 

Applications Edit

Evaluating definite integrals Edit

The formula

 
can be of use when evaluating certain definite integrals. When used in this context, the Leibniz integral rule for differentiating under the integral sign is also known as Feynman's trick for integration.

Example 3 Edit

Consider

 

Now,

 

As   varies from   to  , we have

 

Hence,

 

Therefore,

 

Integrating both sides with respect to  , we get:

 

  follows from evaluating  :

 

To determine   in the same manner, we should need to substitute in a value of   greater than 1 in  . This is somewhat inconvenient. Instead, we substitute  , where  . Then,

 

Therefore,  

The definition of   is now complete:

 

The foregoing discussion, of course, does not apply when  , since the conditions for differentiability are not met.

Example 4 Edit

 

First we calculate:

leibniz, integral, rule, this, article, about, integral, rule, convergence, test, alternating, series, alternating, series, test, this, article, needs, additional, citations, verification, please, help, improve, this, article, adding, citations, reliable, sour. This article is about the integral rule For the convergence test for alternating series see Alternating series test This article needs additional citations for verification Please help improve this article by adding citations to reliable sources Unsourced material may be challenged and removed Find sources Leibniz integral rule news newspapers books scholar JSTOR October 2016 Learn how and when to remove this template message In calculus the Leibniz integral rule for differentiation under the integral sign states that for an integral of the form a x b x f x t d t displaystyle int a x b x f x t dt where lt a x b x lt displaystyle infty lt a x b x lt infty and the integrands are functions dependent on x displaystyle x the derivative of this integral is expressible as d d x a x b x f x t d t displaystyle frac d dx left int a x b x f x t dt right f x b x d d x b x f x a x d d x a x a x b x x f x t d t f big x b x big cdot frac d dx b x f big x a x big cdot frac d dx a x int a x b x frac partial partial x f x t dt where the partial derivative x displaystyle tfrac partial partial x indicates that inside the integral only the variation of f x t displaystyle f x t with x displaystyle x is considered in taking the derivative 1 It is named after Gottfried Leibniz In the special case where the functions a x displaystyle a x and b x displaystyle b x are constants a x a displaystyle a x a and b x b displaystyle b x b with values that do not depend on x displaystyle x this simplifies to d d x a b f x t d t a b x f x t d t displaystyle frac d dx left int a b f x t dt right int a b frac partial partial x f x t dt If a x a displaystyle a x a is constant and b x x displaystyle b x x which is another common situation for example in the proof of Cauchy s repeated integration formula the Leibniz integral rule becomes d d x a x f x t d t f x x a x x f x t d t displaystyle frac d dx left int a x f x t dt right f big x x big int a x frac partial partial x f x t dt This important result may under certain conditions be used to interchange the integral and partial differential operators and is particularly useful in the differentiation of integral transforms An example of such is the moment generating function in probability theory a variation of the Laplace transform which can be differentiated to generate the moments of a random variable Whether Leibniz s integral rule applies is essentially a question about the interchange of limits Contents 1 General form differentiation under the integral sign 2 Three dimensional time dependent case 3 Higher dimensions 4 Measure theory statement 5 Proofs 5 1 Proof of basic form 5 1 1 Another proof using the bounded convergence theorem 5 2 Variable limits form 5 3 General form with variable limits 5 4 Alternative proof of the general form with variable limits using the chain rule 5 5 Three dimensional time dependent form 5 6 Alternative derivation 6 Examples 6 1 Example 1 Fixed limits 6 2 Example 2 Variable limits 7 Applications 7 1 Evaluating definite integrals 7 1 1 Example 3 7 1 2 Example 4 7 1 3 Example 5 7 1 4 Example 6 7 1 5 Other problems to solve 7 2 Infinite series 8 In popular culture 9 See also 10 References 11 Further reading 12 External linksGeneral form differentiation under the integral sign EditTheorem Let f x t displaystyle f x t nbsp be a function such that both f x t displaystyle f x t nbsp and its partial derivative f x x t displaystyle f x x t nbsp are continuous in t displaystyle t nbsp and x displaystyle x nbsp in some region of the x t displaystyle xt nbsp plane including a x t b x displaystyle a x leq t leq b x nbsp x 0 x x 1 displaystyle x 0 leq x leq x 1 nbsp Also suppose that the functions a x displaystyle a x nbsp and b x displaystyle b x nbsp are both continuous and both have continuous derivatives for x 0 x x 1 displaystyle x 0 leq x leq x 1 nbsp Then for x 0 x x 1 displaystyle x 0 leq x leq x 1 nbsp d d x a x b x f x t d t f x b x d d x b x f x a x d d x a x a x b x x f x t d t displaystyle frac d dx left int a x b x f x t dt right f big x b x big cdot frac d dx b x f big x a x big cdot frac d dx a x int a x b x frac partial partial x f x t dt nbsp The right hand side may also be written using Lagrange s notation as f x b x b x f x a x a x a x b x f x x t d t textstyle f x b x b prime x f x a x a prime x displaystyle int a x b x f x x t dt nbsp Stronger versions of the theorem only require that the partial derivative exist almost everywhere and not that it be continuous 2 This formula is the general form of the Leibniz integral rule and can be derived using the fundamental theorem of calculus The first fundamental theorem of calculus is just the particular case of the above formula where a x a R displaystyle a x a in mathbb R nbsp is constant b x x displaystyle b x x nbsp and f x t f t displaystyle f x t f t nbsp does not depend on x displaystyle x nbsp If both upper and lower limits are taken as constants then the formula takes the shape of an operator equation I t x x I t displaystyle mathcal I t partial x partial x mathcal I t nbsp where x displaystyle partial x nbsp is the partial derivative with respect to x displaystyle x nbsp and I t displaystyle mathcal I t nbsp is the integral operator with respect to t displaystyle t nbsp over a fixed interval That is it is related to the symmetry of second derivatives but involving integrals as well as derivatives This case is also known as the Leibniz integral rule The following three basic theorems on the interchange of limits are essentially equivalent the interchange of a derivative and an integral differentiation under the integral sign i e Leibniz integral rule the change of order of partial derivatives the change of order of integration integration under the integral sign i e Fubini s theorem Three dimensional time dependent case EditSee also Higher dimensions nbsp Figure 1 A vector field F r t defined throughout space and a surface S bounded by curve S moving with velocity v over which the field is integrated A Leibniz integral rule for a two dimensional surface moving in three dimensional space is 3 4 d d t S t F r t d A S t F t r t F r t v d A S t v F r t d s displaystyle frac d dt iint Sigma t mathbf F mathbf r t cdot d mathbf A iint Sigma t left mathbf F t mathbf r t left nabla cdot mathbf F mathbf r t right mathbf v right cdot d mathbf A oint partial Sigma t left mathbf v times mathbf F mathbf r t right cdot d mathbf s nbsp where F r t is a vector field at the spatial position r at time t S is a surface bounded by the closed curve S dA is a vector element of the surface S ds is a vector element of the curve S v is the velocity of movement of the region S is the vector divergence is the vector cross product The double integrals are surface integrals over the surface S and the line integral is over the bounding curve S Higher dimensions EditThe Leibniz integral rule can be extended to multidimensional integrals In two and three dimensions this rule is better known from the field of fluid dynamics as the Reynolds transport theorem d d t D t F x t d V D t t F x t d V D t F x t v b d S displaystyle frac d dt int D t F mathbf x t dV int D t frac partial partial t F mathbf x t dV int partial D t F mathbf x t mathbf v b cdot d mathbf Sigma nbsp where F x t displaystyle F mathbf x t nbsp is a scalar function D t and D t denote a time varying connected region of R3 and its boundary respectively v b displaystyle mathbf v b nbsp is the Eulerian velocity of the boundary see Lagrangian and Eulerian coordinates and dS n dS is the unit normal component of the surface element The general statement of the Leibniz integral rule requires concepts from differential geometry specifically differential forms exterior derivatives wedge products and interior products With those tools the Leibniz integral rule in n dimensions is 4 d d t W t w W t i v d x w W t i v w W t w displaystyle frac d dt int Omega t omega int Omega t i mathbf v d x omega int partial Omega t i mathbf v omega int Omega t dot omega nbsp where W t is a time varying domain of integration w is a p form v x t displaystyle mathbf v frac partial mathbf x partial t nbsp is the vector field of the velocity i v displaystyle i mathbf v nbsp denotes the interior product with v displaystyle mathbf v nbsp dxw is the exterior derivative of w with respect to the space variables only and w displaystyle dot omega nbsp is the time derivative of w However all of these identities can be derived from a most general statement about Lie derivatives d d t t 0 im ps t W w W L PS w displaystyle left frac d dt right t 0 int operatorname im psi t Omega omega int Omega mathcal L Psi omega nbsp Here the ambient manifold on which the differential form w displaystyle omega nbsp lives includes both space and time W displaystyle Omega nbsp is the region of integration a submanifold at a given instant it does not depend on t displaystyle t nbsp since its parametrization as a submanifold defines its position in time L displaystyle mathcal L nbsp is the Lie derivative PS displaystyle Psi nbsp is the spacetime vector field obtained from adding the unitary vector field in the direction of time to the purely spatial vector field v displaystyle mathbf v nbsp from the previous formulas i e PS displaystyle Psi nbsp is the spacetime velocity of W displaystyle Omega nbsp ps t displaystyle psi t nbsp is a diffeomorphism from the one parameter group generated by the flow of PS displaystyle Psi nbsp and im ps t W displaystyle text im psi t Omega nbsp is the image of W displaystyle Omega nbsp under such diffeomorphism Something remarkable about this form is that it can account for the case when W displaystyle Omega nbsp changes its shape and size over time since such deformations are fully determined by PS displaystyle Psi nbsp Measure theory statement EditLet X displaystyle X nbsp be an open subset of R displaystyle mathbf R nbsp and W displaystyle Omega nbsp be a measure space Suppose f X W R displaystyle f colon X times Omega to mathbf R nbsp satisfies the following conditions 5 6 2 f x w displaystyle f x omega nbsp is a Lebesgue integrable function of w displaystyle omega nbsp for each x X displaystyle x in X nbsp For almost all w W displaystyle omega in Omega nbsp the partial derivative f x displaystyle f x nbsp exists for all x X displaystyle x in X nbsp There is an integrable function 8 W R displaystyle theta colon Omega to mathbf R nbsp such that f x x w 8 w displaystyle f x x omega leq theta omega nbsp for all x X displaystyle x in X nbsp and almost every w W displaystyle omega in Omega nbsp Then for all x X displaystyle x in X nbsp d d x W f x w d w W f x x w d w displaystyle frac d dx int Omega f x omega d omega int Omega f x x omega d omega nbsp The proof relies on the dominated convergence theorem and the mean value theorem details below Proofs EditProof of basic form Edit We first prove the case of constant limits of integration a and b We use Fubini s theorem to change the order of integration For every x and h such that h gt 0 and both x and x h are within x0 x1 we have x x h a b f x x t d t d x a b x x h f x x t d x d t a b f x h t f x t d t a b f x h t d t a b f x t d t displaystyle int x x h int a b f x x t dt dx int a b int x x h f x x t dx dt int a b left f x h t f x t right dt int a b f x h t dt int a b f x t dt nbsp Note that the integrals at hand are well defined since f x x t displaystyle f x x t nbsp is continuous at the closed rectangle x 0 x 1 a b displaystyle x 0 x 1 times a b nbsp and thus also uniformly continuous there thus its integrals by either dt or dx are continuous in the other variable and also integrable by it essentially this is because for uniformly continuous functions one may pass the limit through the integration sign as elaborated below Therefore a b f x h t d t a b f x t d t h 1 h x x h a b f x x t d t d x F x h F x h displaystyle frac int a b f x h t dt int a b f x t dt h frac 1 h int x x h int a b f x x t dt dx frac F x h F x h nbsp Where we have defined F u x 0 u a b f x x t d t d x displaystyle F u equiv int x 0 u int a b f x x t dt dx nbsp we may replace x0 here by any other point between x0 and x F is differentiable with derivative a b f x x t d t textstyle int a b f x x t dt nbsp so we can take the limit where h approaches zero For the left hand side this limit is d d x a b f x t d t displaystyle frac d dx int a b f x t dt nbsp For the right hand side we get F x a b f x x t d t displaystyle F x int a b f x x t dt nbsp And we thus prove the desired result d d x a b f x t d t a b f x x t d t displaystyle frac d dx int a b f x t dt int a b f x x t dt nbsp Another proof using the bounded convergence theorem Edit If the integrals at hand are Lebesgue integrals we may use the bounded convergence theorem valid for these integrals but not for Riemann integrals in order to show that the limit can be passed through the integral sign Note that this proof is weaker in the sense that it only shows that fx x t is Lebesgue integrable but not that it is Riemann integrable In the former stronger proof if f x t is Riemann integrable then so is fx x t and thus is obviously also Lebesgue integrable Let u x a b f x t d t displaystyle u x int a b f x t dt nbsp 1 By the definition of the derivative u x lim h 0 u x h u x h displaystyle u x lim h to 0 frac u x h u x h nbsp 2 Substitute equation 1 into equation 2 The difference of two integrals equals the integral of the difference and 1 h is a constant sou x lim h 0 a b f x h t d t a b f x t d t h lim h 0 a b f x h t f x t d t h lim h 0 a b f x h t f x t h d t displaystyle begin aligned u x amp lim h to 0 frac int a b f x h t dt int a b f x t dt h amp lim h to 0 frac int a b left f x h t f x t right dt h amp lim h to 0 int a b frac f x h t f x t h dt end aligned nbsp We now show that the limit can be passed through the integral sign We claim that the passage of the limit under the integral sign is valid by the bounded convergence theorem a corollary of the dominated convergence theorem For each d gt 0 consider the difference quotientf d x t f x d t f x t d displaystyle f delta x t frac f x delta t f x t delta nbsp For t fixed the mean value theorem implies there exists z in the interval x x d such that f d x t f x z t displaystyle f delta x t f x z t nbsp Continuity of fx x t and compactness of the domain together imply that fx x t is bounded The above application of the mean value theorem therefore gives a uniform independent of t displaystyle t nbsp bound on f d x t displaystyle f delta x t nbsp The difference quotients converge pointwise to the partial derivative fx by the assumption that the partial derivative exists The above argument shows that for every sequence dn 0 the sequence f d n x t displaystyle f delta n x t nbsp is uniformly bounded and converges pointwise to fx The bounded convergence theorem states that if a sequence of functions on a set of finite measure is uniformly bounded and converges pointwise then passage of the limit under the integral is valid In particular the limit and integral may be exchanged for every sequence dn 0 Therefore the limit as d 0 may be passed through the integral sign Variable limits form Edit For a continuous real valued function g of one real variable and real valued differentiable functions f 1 displaystyle f 1 nbsp and f 2 displaystyle f 2 nbsp of one real variable d d x f 1 x f 2 x g t d t g f 2 x f 2 x g f 1 x f 1 x displaystyle frac d dx left int f 1 x f 2 x g t dt right g left f 2 x right f 2 x g left f 1 x right f 1 x nbsp This follows from the chain rule and the First Fundamental Theorem of Calculus DefineG x f 1 x f 2 x g t d t displaystyle G x int f 1 x f 2 x g t dt nbsp and G x 0 x g t d t displaystyle Gamma x int 0 x g t dt nbsp The lower limit just has to be some number in the domain of g displaystyle g nbsp Then G x displaystyle G x nbsp can be written as a composition G x G f 2 x G f 1 x displaystyle G x Gamma circ f 2 x Gamma circ f 1 x nbsp The Chain Rule then implies thatG x G f 2 x f 2 x G f 1 x f 1 x displaystyle G x Gamma left f 2 x right f 2 x Gamma left f 1 x right f 1 x nbsp By the First Fundamental Theorem of Calculus G x g x displaystyle Gamma x g x nbsp Therefore substituting this result above we get the desired equation G x g f 2 x f 2 x g f 1 x f 1 x displaystyle G x g left f 2 x right f 2 x g left f 1 x right f 1 x nbsp Note This form can be particularly useful if the expression to be differentiated is of the form f 1 x f 2 x h x g t d t displaystyle int f 1 x f 2 x h x g t dt nbsp Because h x displaystyle h x nbsp does not depend on the limits of integration it may be moved out from under the integral sign and the above form may be used with the Product rule i e d d x f 1 x f 2 x h x g t d t d d x h x f 1 x f 2 x g t d t h x f 1 x f 2 x g t d t h x d d x f 1 x f 2 x g t d t displaystyle frac d dx left int f 1 x f 2 x h x g t dt right frac d dx left h x int f 1 x f 2 x g t dt right h x int f 1 x f 2 x g t dt h x frac d dx left int f 1 x f 2 x g t dt right nbsp General form with variable limits Edit Setf a a b f x a d x displaystyle varphi alpha int a b f x alpha dx nbsp where a and b are functions of a that exhibit increments Da and Db respectively when a is increased by Da Then D f f a D a f a a D a b D b f x a D a d x a b f x a d x a D a a f x a D a d x a b f x a D a d x b b D b f x a D a d x a b f x a d x a a D a f x a D a d x a b f x a D a f x a d x b b D b f x a D a d x displaystyle begin aligned Delta varphi amp varphi alpha Delta alpha varphi alpha 4pt amp int a Delta a b Delta b f x alpha Delta alpha dx int a b f x alpha dx 4pt amp int a Delta a a f x alpha Delta alpha dx int a b f x alpha Delta alpha dx int b b Delta b f x alpha Delta alpha dx int a b f x alpha dx 4pt amp int a a Delta a f x alpha Delta alpha dx int a b f x alpha Delta alpha f x alpha dx int b b Delta b f x alpha Delta alpha dx end aligned nbsp A form of the mean value theorem a b f x d x b a f 3 textstyle int a b f x dx b a f xi nbsp where a lt 3 lt b may be applied to the first and last integrals of the formula for Df above resulting inD f D a f 3 1 a D a a b f x a D a f x a d x D b f 3 2 a D a displaystyle Delta varphi Delta af xi 1 alpha Delta alpha int a b f x alpha Delta alpha f x alpha dx Delta bf xi 2 alpha Delta alpha nbsp Divide by Da and let Da 0 Notice 31 a and 32 b We may pass the limit through the integral sign lim D a 0 a b f x a D a f x a D a d x a b a f x a d x displaystyle lim Delta alpha to 0 int a b frac f x alpha Delta alpha f x alpha Delta alpha dx int a b frac partial partial alpha f x alpha dx nbsp again by the bounded convergence theorem This yields the general form of the Leibniz integral rule d f d a a b a f x a d x f b a d b d a f a a d a d a displaystyle frac d varphi d alpha int a b frac partial partial alpha f x alpha dx f b alpha frac db d alpha f a alpha frac da d alpha nbsp Alternative proof of the general form with variable limits using the chain rule Edit The general form of Leibniz s Integral Rule with variable limits can be derived as a consequence of the basic form of Leibniz s Integral Rule the multivariable chain rule and the First Fundamental Theorem of Calculus Suppose f displaystyle f nbsp is defined in a rectangle in the x t displaystyle x t nbsp plane for x x 1 x 2 displaystyle x in x 1 x 2 nbsp and t t 1 t 2 displaystyle t in t 1 t 2 nbsp Also assume f displaystyle f nbsp and the partial derivative f x textstyle frac partial f partial x nbsp are both continuous functions on this rectangle Suppose a b displaystyle a b nbsp are differentiable real valued functions defined on x 1 x 2 displaystyle x 1 x 2 nbsp with values in t 1 t 2 displaystyle t 1 t 2 nbsp i e for every x x 1 x 2 a x b x t 1 t 2 displaystyle x in x 1 x 2 a x b x in t 1 t 2 nbsp Now setF x y t 1 y f x t d t for x x 1 x 2 and y t 1 t 2 displaystyle F x y int t 1 y f x t dt qquad text for x in x 1 x 2 text and y in t 1 t 2 nbsp and G x a x b x f x t d t for x x 1 x 2 displaystyle G x int a x b x f x t dt quad text for x in x 1 x 2 nbsp Then by properties of Definite Integrals we can writeG x t 1 b x f x t d t t 1 a x f x t d t F x b x F x a x displaystyle begin aligned G x amp int t 1 b x f x t dt int t 1 a x f x t dt 4pt amp F x b x F x a x end aligned nbsp Since the functions F a b displaystyle F a b nbsp are all differentiable see the remark at the end of the proof by the Multivariable Chain Rule it follows that G displaystyle G nbsp is differentiable and its derivative is given by the formula G x F x x b x F y x b x b x F x x a x F y x a x a x displaystyle G x left frac partial F partial x x b x frac partial F partial y x b x b x right left frac partial F partial x x a x frac partial F partial y x a x a x right nbsp Now note that for every x x 1 x 2 displaystyle x in x 1 x 2 nbsp and for every y t 1 t 2 displaystyle y in t 1 t 2 nbsp we have that F x x y t 1 y f x x t d t textstyle frac partial F partial x x y int t 1 y frac partial f partial x x t dt nbsp because when taking the partial derivative with respect to x displaystyle x nbsp of F displaystyle F nbsp we are keeping y displaystyle y nbsp fixed in the expression t 1 y f x t d t textstyle int t 1 y f x t dt nbsp thus the basic form of Leibniz s Integral Rule with constant limits of integration applies Next by the First Fundamental Theorem of Calculus we have that F y x y f x y textstyle frac partial F partial y x y f x y nbsp because when taking the partial derivative with respect to y displaystyle y nbsp of F displaystyle F nbsp the first variable x displaystyle x nbsp is fixed so the fundamental theorem can indeed be applied Substituting these results into the equation for G x displaystyle G x nbsp above gives G x t 1 b x f x x t d t f x b x b x t 1 a x f x x t d t f x a x a x f x b x b x f x a x a x a x b x f x x t d t displaystyle begin aligned G x amp left int t 1 b x frac partial f partial x x t dt f x b x b x right left int t 1 a x dfrac partial f partial x x t dt f x a x a x right 2pt amp f x b x b x f x a x a x int a x b x frac partial f partial x x t dt end aligned nbsp as desired There is a technical point in the proof above which is worth noting applying the Chain Rule to G displaystyle G nbsp requires that F displaystyle F nbsp already be differentiable This is where we use our assumptions about f displaystyle f nbsp As mentioned above the partial derivatives of F displaystyle F nbsp are given by the formulas F x x y t 1 y f x x t d t textstyle frac partial F partial x x y int t 1 y frac partial f partial x x t dt nbsp and F y x y f x y textstyle frac partial F partial y x y f x y nbsp Since f x textstyle dfrac partial f partial x nbsp is continuous its integral is also a continuous function 7 and since f displaystyle f nbsp is also continuous these two results show that both the partial derivatives of F displaystyle F nbsp are continuous Since continuity of partial derivatives implies differentiability of the function 8 F displaystyle F nbsp is indeed differentiable Three dimensional time dependent form Edit See also Higher dimensions At time t the surface S in Figure 1 contains a set of points arranged about a centroid C t displaystyle mathbf C t nbsp The function F r t displaystyle mathbf F mathbf r t nbsp can be written asF C t r C t t F C t I t displaystyle mathbf F mathbf C t mathbf r mathbf C t t mathbf F mathbf C t mathbf I t nbsp with I displaystyle mathbf I nbsp independent of time Variables are shifted to a new frame of reference attached to the moving surface with origin at C t displaystyle mathbf C t nbsp For a rigidly translating surface the limits of integration are then independent of time so d d t S t d A r F r t S d A I d d t F C t I t displaystyle frac d dt left iint Sigma t d mathbf A mathbf r cdot mathbf F mathbf r t right iint Sigma d mathbf A mathbf I cdot frac d dt mathbf F mathbf C t mathbf I t nbsp where the limits of integration confining the integral to the region S no longer are time dependent so differentiation passes through the integration to act on the integrand only d d t F C t I t F t C t I t v F C t I t F t r t v F r t displaystyle frac d dt mathbf F mathbf C t mathbf I t mathbf F t mathbf C t mathbf I t mathbf v cdot nabla F mathbf C t mathbf I t mathbf F t mathbf r t mathbf v cdot nabla mathbf F mathbf r t nbsp with the velocity of motion of the surface defined by v d d t C t displaystyle mathbf v frac d dt mathbf C t nbsp This equation expresses the material derivative of the field that is the derivative with respect to a coordinate system attached to the moving surface Having found the derivative variables can be switched back to the original frame of reference We notice that see article on curl v F F F v v v F displaystyle nabla times left mathbf v times mathbf F right nabla cdot mathbf F mathbf F cdot nabla mathbf v nabla cdot mathbf v mathbf v cdot nabla mathbf F nbsp and that Stokes theorem equates the surface integral of the curl over S with a line integral over S d d t S t F r t d A S t F t r t F v F v v F d A S t v F d s displaystyle frac d dt left iint Sigma t mathbf F mathbf r t cdot d mathbf A right iint Sigma t big mathbf F t mathbf r t left mathbf F cdot nabla right mathbf v left nabla cdot mathbf F right mathbf v nabla cdot mathbf v mathbf F big cdot d mathbf A oint partial Sigma t left mathbf v times mathbf F right cdot d mathbf s nbsp The sign of the line integral is based on the right hand rule for the choice of direction of line element ds To establish this sign for example suppose the field F points in the positive z direction and the surface S is a portion of the xy plane with perimeter S We adopt the normal to S to be in the positive z direction Positive traversal of S is then counterclockwise right hand rule with thumb along z axis Then the integral on the left hand side determines a positive flux of F through S Suppose S translates in the positive x direction at velocity v An element of the boundary of S parallel to the y axis say ds sweeps out an area vt ds in time t If we integrate around the boundary S in a counterclockwise sense vt ds points in the negative z direction on the left side of S where ds points downward and in the positive z direction on the right side of S where ds points upward which makes sense because S is moving to the right adding area on the right and losing it on the left On that basis the flux of F is increasing on the right of S and decreasing on the left However the dot product v F ds F v ds F v ds Consequently the sign of the line integral is taken as negative If v is a constant d d t S t F r t d A S t F t r t F v d A S t v F d s displaystyle frac d dt iint Sigma t mathbf F mathbf r t cdot d mathbf A iint Sigma t big mathbf F t mathbf r t left nabla cdot mathbf F right mathbf v big cdot d mathbf A oint partial Sigma t left mathbf v times mathbf F right cdot d mathbf s nbsp which is the quoted result This proof does not consider the possibility of the surface deforming as it moves Alternative derivation Edit Lemma One has b a b f x d x f b a a b f x d x f a displaystyle frac partial partial b left int a b f x dx right f b qquad frac partial partial a left int a b f x dx right f a nbsp Proof From the proof of the fundamental theorem of calculus b a b f x d x lim D b 0 1 D b a b D b f x d x a b f x d x lim D b 0 1 D b b b D b f x d x lim D b 0 1 D b f b D b O D b 2 f b displaystyle begin aligned frac partial partial b left int a b f x dx right amp lim Delta b to 0 frac 1 Delta b left int a b Delta b f x dx int a b f x dx right 6pt amp lim Delta b to 0 frac 1 Delta b int b b Delta b f x dx 6pt amp lim Delta b to 0 frac 1 Delta b left f b Delta b O left Delta b 2 right right 6pt amp f b end aligned nbsp and a a b f x d x lim D a 0 1 D a a D a b f x d x a b f x d x lim D a 0 1 D a a D a a f x d x lim D a 0 1 D a f a D a O D a 2 f a displaystyle begin aligned frac partial partial a left int a b f x dx right amp lim Delta a to 0 frac 1 Delta a left int a Delta a b f x dx int a b f x dx right 6pt amp lim Delta a to 0 frac 1 Delta a int a Delta a a f x dx 6pt amp lim Delta a to 0 frac 1 Delta a left f a Delta a O left Delta a 2 right right 6pt amp f a end aligned nbsp Suppose a and b are constant and that f x involves a parameter a which is constant in the integration but may vary to form different integrals Assume that f x a is a continuous function of x and a in the compact set x a a0 a a1 and a x b and that the partial derivative fa x a exists and is continuous If one defines f a a b f x a d x displaystyle varphi alpha int a b f x alpha dx nbsp then f displaystyle varphi nbsp may be differentiated with respect to a by differentiating under the integral sign i e d f d a a b a f x a d x displaystyle frac d varphi d alpha int a b frac partial partial alpha f x alpha dx nbsp By the Heine Cantor theorem it is uniformly continuous in that set In other words for any e gt 0 there exists Da such that for all values of x in a b f x a D a f x a lt e displaystyle f x alpha Delta alpha f x alpha lt varepsilon nbsp On the other hand D f f a D a f a a b f x a D a d x a b f x a d x a b f x a D a f x a d x e b a displaystyle begin aligned Delta varphi amp varphi alpha Delta alpha varphi alpha 6pt amp int a b f x alpha Delta alpha dx int a b f x alpha dx 6pt amp int a b left f x alpha Delta alpha f x alpha right dx 6pt amp leq varepsilon b a end aligned nbsp Hence f a is a continuous function Similarly if a f x a displaystyle frac partial partial alpha f x alpha nbsp exists and is continuous then for all e gt 0 there exists Da such that x a b f x a D a f x a D a f a lt e displaystyle forall x in a b quad left frac f x alpha Delta alpha f x alpha Delta alpha frac partial f partial alpha right lt varepsilon nbsp Therefore D f D a a b f x a D a f x a D a d x a b f x a a d x R displaystyle frac Delta varphi Delta alpha int a b frac f x alpha Delta alpha f x alpha Delta alpha dx int a b frac partial f x alpha partial alpha dx R nbsp where R lt a b e d x e b a displaystyle R lt int a b varepsilon dx varepsilon b a nbsp Now e 0 as Da 0 solim D a 0 D f D a d f d a a b a f x a d x displaystyle lim Delta alpha to 0 frac Delta varphi Delta alpha frac d varphi d alpha int a b frac partial partial alpha f x alpha dx nbsp This is the formula we set out to prove Now suppose a b f x a d x f a displaystyle int a b f x alpha dx varphi alpha nbsp where a and b are functions of a which take increments Da and Db respectively when a is increased by Da Then D f f a D a f a a D a b D b f x a D a d x a b f x a d x a D a a f x a D a d x a b f x a D a d x b b D b f x a D a d x a b f x a d x a a D a f x a D a d x a b f x a D a f x a d x b b D b f x a D a d x displaystyle begin aligned Delta varphi amp varphi alpha Delta alpha varphi alpha 6pt amp int a Delta a b Delta b f x alpha Delta alpha dx int a b f x alpha dx 6pt amp int a Delta a a f x alpha Delta alpha dx int a b f x alpha Delta alpha dx int b b Delta b f x alpha Delta alpha dx int a b f x alpha dx 6pt amp int a a Delta a f x alpha Delta alpha dx int a b f x alpha Delta alpha f x alpha dx int b b Delta b f x alpha Delta alpha dx end aligned nbsp A form of the mean value theorem a b f x d x b a f 3 textstyle int a b f x dx b a f xi nbsp where a lt 3 lt b can be applied to the first and last integrals of the formula for Df above resulting inD f D a f 3 1 a D a a b f x a D a f x a d x D b f 3 2 a D a displaystyle Delta varphi Delta a f xi 1 alpha Delta alpha int a b f x alpha Delta alpha f x alpha dx Delta b f xi 2 alpha Delta alpha nbsp Dividing by Da letting Da 0 noticing 31 a and 32 b and using the above derivation ford f d a a b a f x a d x displaystyle frac d varphi d alpha int a b frac partial partial alpha f x alpha dx nbsp yields d f d a a b a f x a d x f b a b a f a a a a displaystyle frac d varphi d alpha int a b frac partial partial alpha f x alpha dx f b alpha frac partial b partial alpha f a alpha frac partial a partial alpha nbsp This is the general form of the Leibniz integral rule Examples EditExample 1 Fixed limits Edit Consider the functionf a 0 1 a x 2 a 2 d x displaystyle varphi alpha int 0 1 frac alpha x 2 alpha 2 dx nbsp The function under the integral sign is not continuous at the point x a 0 0 and the function f a has a discontinuity at a 0 because f a approaches p 2 as a 0 If we differentiate f a with respect to a under the integral sign we getd d a f a 0 1 a a x 2 a 2 d x 0 1 x 2 a 2 x 2 a 2 2 d x x x 2 a 2 0 1 1 1 a 2 displaystyle frac d d alpha varphi alpha int 0 1 frac partial partial alpha left frac alpha x 2 alpha 2 right dx int 0 1 frac x 2 alpha 2 x 2 alpha 2 2 dx left frac x x 2 alpha 2 right 0 1 frac 1 1 alpha 2 nbsp which is of course true for all values of a except a 0 This may be integrated with respect to a to find f a 0 a 0 arctan a p 2 a 0 displaystyle varphi alpha begin cases 0 amp alpha 0 arctan alpha frac pi 2 amp alpha neq 0 end cases nbsp Example 2 Variable limits Edit An example with variable limits d d x sin x cos x cosh t 2 d t cosh cos 2 x d d x cos x cosh sin 2 x d d x sin x sin x cos x x cosh t 2 d t cosh cos 2 x sin x cosh sin 2 x cos x 0 cosh cos 2 x sin x cosh sin 2 x cos x displaystyle begin aligned frac d dx int sin x cos x cosh t 2 dt amp cosh left cos 2 x right frac d dx cos x cosh left sin 2 x right frac d dx sin x int sin x cos x frac partial partial x cosh t 2 dt 6pt amp cosh cos 2 x sin x cosh sin 2 x cos x 0 6pt amp cosh cos 2 x sin x cosh sin 2 x cos x end aligned nbsp Applications EditEvaluating definite integrals Edit The formulad d x a x b x f x t d t f x b x d d x b x f x a x d d x a x a x b x x f x t d t displaystyle frac d dx left int a x b x f x t dt right f big x b x big cdot frac d dx b x f big x a x big cdot frac d dx a x int a x b x frac partial partial x f x t dt nbsp can be of use when evaluating certain definite integrals When used in this context the Leibniz integral rule for differentiating under the integral sign is also known as Feynman s trick for integration Example 3 Edit Considerf a 0 p ln 1 2 a cos x a 2 d x a 1 displaystyle varphi alpha int 0 pi ln left 1 2 alpha cos x alpha 2 right dx qquad alpha neq 1 nbsp Now d d a f a 0 p 2 cos x 2 a 1 2 a cos x a 2 d x 1 a 0 p 1 1 a 2 1 2 a cos x a 2 d x p a 2 a arctan 1 a 1 a tan x 2 0 p displaystyle begin aligned frac d d alpha varphi alpha amp int 0 pi frac 2 cos x 2 alpha 1 2 alpha cos x alpha 2 dx 6pt amp frac 1 alpha int 0 pi left 1 frac 1 alpha 2 1 2 alpha cos x alpha 2 right dx 6pt amp left frac pi alpha frac 2 alpha left arctan left frac 1 alpha 1 alpha tan left frac x 2 right right right right 0 pi end aligned nbsp As x displaystyle x nbsp varies from 0 displaystyle 0 nbsp to p displaystyle pi nbsp we have 1 a 1 a tan x 2 0 a lt 1 1 a 1 a tan x 2 0 a gt 1 displaystyle begin cases frac 1 alpha 1 alpha tan left frac x 2 right geq 0 amp alpha lt 1 frac 1 alpha 1 alpha tan left frac x 2 right leq 0 amp alpha gt 1 end cases nbsp Hence arctan 1 a 1 a tan x 2 0 p p 2 a lt 1 p 2 a gt 1 displaystyle left arctan left frac 1 alpha 1 alpha tan left frac x 2 right right right 0 pi begin cases frac pi 2 amp alpha lt 1 frac pi 2 amp alpha gt 1 end cases nbsp Therefore d d a f a 0 a lt 1 2 p a a gt 1 displaystyle frac d d alpha varphi alpha begin cases 0 amp alpha lt 1 frac 2 pi alpha amp alpha gt 1 end cases nbsp Integrating both sides with respect to a displaystyle alpha nbsp we get f a C 1 a lt 1 2 p ln a C 2 a gt 1 displaystyle varphi alpha begin cases C 1 amp alpha lt 1 2 pi ln alpha C 2 amp alpha gt 1 end cases nbsp C 1 0 displaystyle C 1 0 nbsp follows from evaluating f 0 displaystyle varphi 0 nbsp f 0 0 p ln 1 d x 0 p 0 d x 0 displaystyle varphi 0 int 0 pi ln 1 dx int 0 pi 0 dx 0 nbsp To determine C 2 displaystyle C 2 nbsp in the same manner we should need to substitute in a value of a displaystyle alpha nbsp greater than 1 in f a displaystyle varphi alpha nbsp This is somewhat inconvenient Instead we substitute a 1 b textstyle alpha frac 1 beta nbsp where b lt 1 displaystyle beta lt 1 nbsp Then f a 0 p ln 1 2 b cos x b 2 2 ln b d x 0 p ln 1 2 b cos x b 2 d x 0 p 2 ln b d x 0 2 p ln b 2 p ln a displaystyle begin aligned varphi alpha amp int 0 pi left ln left 1 2 beta cos x beta 2 right 2 ln beta right dx 6pt amp int 0 pi ln left 1 2 beta cos x beta 2 right dx int 0 pi 2 ln beta dx 6pt amp 0 2 pi ln beta 6pt amp 2 pi ln alpha end aligned nbsp Therefore C 2 0 displaystyle C 2 0 nbsp The definition of f a displaystyle varphi alpha nbsp is now complete f a 0 a lt 1 2 p ln a a gt 1 displaystyle varphi alpha begin cases 0 amp alpha lt 1 2 pi ln alpha amp alpha gt 1 end cases nbsp The foregoing discussion of course does not apply when a 1 displaystyle alpha pm 1 nbsp since the conditions for differentiability are not met Example 4 Edit I 0 p 2 1 a cos 2 x b sin 2 x 2 d x a b gt 0 displaystyle mathbf I int 0 pi 2 frac 1 a cos 2 x b sin 2 x 2 dx qquad a b gt 0 nbsp First we calculate J 0 p 2 1 a cos 2 x b sin 2 x d x 0 p 2 1 cos 2 x a b sin 2 x cos 2 x d x 0 p 2 sec 2 x a b tan 2 x d x 1 b 0 p 2 1 a b 2 tan 2 x d tan x 1 a b arctan b a tan x 0 p 2 p 2 a b displaystyle begin aligned mathbf J amp int 0 pi 2 frac 1 a cos 2 x b sin 2 x dx 6pt amp int 0 pi 2 frac frac 1 cos 2 x a b frac sin 2 x cos 2 x dx 6pt amp int 0 pi 2 frac sec 2 x a b tan 2 x dx 6pt amp frac 1 b int 0 pi 2 frac 1 left sqrt frac a b right 2 tan 2 x d tan x 6pt amp left frac 1 sqrt ab arctan left sqrt frac b a tan x right right 0 pi 2 6pt amp frac pi 2 sqrt ab end aligned img class, 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.