Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Integration by substitution

In calculus, integration by substitution, also known as u-substitution, reverse chain rule or change of
variables,[1] is a method for evaluating integrals and antiderivatives. It is the counterpart to the chain rule
for differentiation, and can loosely be thought of as using the chain rule "backwards."

Substitution for a single variable

Introduction (indefinite integrals)


Before stating the result rigorously, consider a simple case using indefinite integrals.

Compute [2]

Set This means or as a differential form, Now:

where is an arbitrary constant of integration.

This procedure is frequently used, but not all integrals are of a form that permits its use. In any event, the
result should be verified by differentiating and comparing to the original integrand.

For definite integrals, the limits of integration must also be adjusted, but the procedure is mostly the same.

Statement for definite integrals


Let be a differentiable function with a continuous derivative, where is an interval.
Suppose that is a continuous function. Then:[3]
In Leibniz notation, the substitution yields:

Working heuristically with infinitesimals yields the equation

which suggests the substitution formula above. (This equation may be put on a rigorous foundation by
interpreting it as a statement about differential forms.) One may view the method of integration by
substitution as a partial justification of Leibniz's notation for integrals and derivatives.

The formula is used to transform one integral into another integral that is easier to compute. Thus, the
formula can be read from left to right or from right to left in order to simplify a given integral. When used in
the former manner, it is sometimes known as u-substitution or w-substitution in which a new variable is
defined to be a function of the original variable found inside the composite function multiplied by the
derivative of the inner function. The latter manner is commonly used in trigonometric substitution, replacing
the original variable with a trigonometric function of a new variable and the original differential with the
differential of the trigonometric function.

Proof
Integration by substitution can be derived from the fundamental theorem of calculus as follows. Let and
be two functions satisfying the above hypothesis that is continuous on and is integrable on the closed
interval . Then the function is also integrable on . Hence the integrals

and

in fact exist, and it remains to show that they are equal.

Since is continuous, it has an antiderivative . The composite function is then defined. Since is
differentiable, combining the chain rule and the definition of an antiderivative gives:

Applying the fundamental theorem of calculus twice gives:


which is the substitution rule.

Examples: Antiderivatives (indefinite integrals)


Substitution can be used to determine antiderivatives. One chooses a relation between and determines
the corresponding relation between and by differentiating, and performs the substitutions. An
antiderivative for the substituted function can hopefully be determined; the original substitution between
and is then undone.

Example 1
Consider the integral:

Make the substitution to obtain meaning Therefore:

where is an arbitrary constant of integration.

Example 2: Antiderivatives of tangent and cotangent


The tangent function can be integrated using substitution by expressing it in terms of the sine and cosine:
.

Using the substitution gives and


The cotangent function can be integrated similarly by expressing it as and using the
substitution :

Examples: Definite integrals


When evaluating definite integrals by substitution, one may calculate the antiderivative fully first, then apply
the boundary conditions. In that case, there is no need to transform the boundary terms. Alternatively, one
may fully evaluate the indefinite integral (see above) first then apply the boundary conditions. This becomes
especially handy when multiple substitutions are used.

Example 1
Consider the integral:

Make the substitution to obtain meaning Therefore:

Since the lower limit was replaced with and the upper limit with a
transformation back into terms of was unnecessary.

Example 2: Trigonometric substitution


For the integral

a variation of the above procedure is needed. The substitution implying is useful


because We thus have:

The resulting integral can be computed using integration by parts or a double angle formula,
followed by one more substitution. One can also note that the function being
integrated is the upper right quarter of a circle with a radius of one, and hence integrating the upper right
quarter from zero to one is the geometric equivalent to the area of one quarter of the unit circle, or

Substitution for multiple variables


One may also use substitution when integrating functions of several variables.

Here, the substitution function (v1,...,vn) = φ(u 1, ..., u n) needs to be injective and continuously
differentiable, and the differentials transform as:

where det(Dφ)(u 1, ..., u n) denotes the determinant of the Jacobian matrix of partial derivatives of φ at
the point (u 1, ..., u n). This formula expresses the fact that the absolute value of the determinant of a matrix
equals the volume of the parallelotope spanned by its columns or rows.

More precisely, the change of variables formula is stated in the next theorem:

Theorem — Let U be an open set in Rn and φ : U → Rn an injective differentiable


function with continuous partial derivatives, the Jacobian of which is nonzero for every x in
U. Then for any real-valued, compactly supported, continuous function f, with support
contained in φ(U):
The conditions on the theorem can be weakened in various ways. First, the requirement that φ be
continuously differentiable can be replaced by the weaker assumption that φ be merely differentiable and
have a continuous inverse.[4] This is guaranteed to hold if φ is continuously differentiable by the inverse
function theorem. Alternatively, the requirement that det(Dφ) ≠ 0 can be eliminated by applying Sard's
theorem.[5]

For Lebesgue measurable functions, the theorem can be stated in the following form:[6]

Theorem — Let U be a measurable subset of Rn and φ : U → Rn an injective function,


and suppose for every x in U there exists φ′(x) in Rn,n such that
φ(y) = φ(x) + φ′(x)(y − x) + o(‖y − x‖) as y → x (here o is little-o notation). Then
φ(U) is measurable, and for any real-valued function f defined on φ(U):

in the sense that if either integral exists (including the possibility of being properly infinite),
then so does the other one, and they have the same value.

Another very general version in measure theory is the following:[7]

Theorem — Let X be a locally compact Hausdorff space equipped with a finite Radon
measure μ , and let Y be a σ-compact Hausdorff space with a σ-finite Radon measure ρ. Let
φ : X → Y be an absolutely continuous function (where the latter means that ρ(φ(E)) = 0
whenever μ(E) = 0 ). Then there exists a real-valued Borel measurable function w on X
such that for every Lebesgue integrable function f : Y → R, the function (f ∘ φ) ⋅ w is
Lebesgue integrable on X, and

Furthermore, it is possible to write

for some Borel measurable function g on Y .

In geometric measure theory, integration by substitution is used with Lipschitz functions. A bi-Lipschitz
function is a Lipschitz function φ : U → Rn which is injective and whose inverse function
φ−1 : φ(U) → U is also Lipschitz. By Rademacher's theorem, a bi-Lipschitz mapping is differentiable
almost everywhere. In particular, the Jacobian determinant of a bi-Lipschitz mapping det Dφ is well-
defined almost everywhere. The following result then holds:

Theorem — Let U be an open subset of Rn and φ : U → Rn be a bi-Lipschitz mapping.


Let f : φ(U) → R be measurable. Then

in the sense that if either integral exists (or is properly infinite), then so does the other one,
and they have the same value.

The above theorem was first proposed by Euler when he developed the notion of double integrals in 1769.
Although generalized to triple integrals by Lagrange in 1773, and used by Legendre, Laplace, and Gauss,
and first generalized to n variables by Mikhail Ostrogradsky in 1836, it resisted a fully rigorous formal
proof for a surprisingly long time, and was first satisfactorily resolved 125 years later, by Élie Cartan in a
series of papers beginning in the mid-1890s.[8][9]

Application in probability
Substitution can be used to answer the following important question in probability: given a random variable
X with probability density pX and another random variable Y such that Y= ϕ(X) for injective (one-to-one)
ϕ, what is the probability density for Y?
It is easiest to answer this question by first answering a slightly different question: what is the probability
that Y takes a value in some particular subset S ? Denote this probability P(Y ∈ S). Of course, if Y has
probability density p Y, then the answer is:

but this is not really useful because we do not know p Y; it is what we are trying to find. We can make
progress by considering the problem in the variable X. Y takes a value in S whenever X takes a value in
so:

Changing from variable x to y gives:

Combining this with our first equation gives:


so:

In the case where X and Y depend on several uncorrelated variables (i.e., and
), can be found by substitution in several variables discussed above. The result is:

See also
Mathematics
portal

Probability density function


Substitution of variables
Trigonometric substitution
Weierstrass substitution
Euler substitution
Glasser's master theorem
Pushforward measure

Notes
1. Swokowski 1983, p. 257
2. Swokowski 1983, p. 258
3. Briggs & Cochran 2011, p. 361
4. Rudin 1987, Theorem 7.26
5. Spivak 1965, p. 72
6. Fremlin 2010, Theorem 263D
7. Hewitt & Stromberg 1965, Theorem 20.3
8. Katz 1982
9. Ferzola 1994

References
Briggs, William; Cochran, Lyle (2011), Calculus /Early Transcendentals (Single
Variable ed.), Addison-Wesley, ISBN 978-0-321-66414-3
Ferzola, Anthony P. (1994), "Euler and differentials" (http://mathdl.maa.org/mathDL/22/?pa=c
ontent&sa=viewDocument&nodeId=2688), The College Mathematics Journal, 25 (2): 102–
111, doi:10.2307/2687130 (https://doi.org/10.2307%2F2687130), JSTOR 2687130 (https://w
ww.jstor.org/stable/2687130)
Fremlin, D.H. (2010), Measure Theory, Volume 2, Torres Fremlin, ISBN 978-0-9538129-7-4.
Hewitt, Edwin; Stromberg, Karl (1965), Real and Abstract Analysis, Springer-Verlag,
ISBN 978-0-387-04559-7.
Katz, V. (1982), "Change of variables in multiple integrals: Euler to Cartan", Mathematics
Magazine, 55 (1): 3–11, doi:10.2307/2689856 (https://doi.org/10.2307%2F2689856),
JSTOR 2689856 (https://www.jstor.org/stable/2689856)
Rudin, Walter (1987), Real and Complex Analysis, McGraw-Hill, ISBN 978-0-07-054234-1.
Swokowski, Earl W. (1983), Calculus with analytic geometry (alternate ed.), Prindle, Weber
& Schmidt, ISBN 0-87150-341-7
Spivak, Michael (1965), Calculus on Manifolds, Westview Press, ISBN 978-0-8053-9021-6.

External links
Integration by substitution (https://www.encyclopediaofmath.org/index.php/Integration_by_su
bstitution) at Encyclopedia of Mathematics
Area formula (https://www.encyclopediaofmath.org/index.php/Area_formula) at Encyclopedia
of Mathematics

Retrieved from "https://en.wikipedia.org/w/index.php?title=Integration_by_substitution&oldid=1227209412"

You might also like