In mathematics, an integral is a fundamental concept in calculus that represents the continuous accumulation of a quantity, such as the area under a curve or the total change in a function, serving as the inverse operation to differentiation for many functions, particularly continuous ones, but not in general, as some Riemann integrable functions (e.g., those lacking the intermediate value property required by Darboux's theorem) cannot be derivatives.[1][2] There are two primary types: the definite integral, which computes a specific numerical value over an interval [a, b] as \int_a^b f(x) \, dx = \lim_{\|P\| \to 0} \sum_{i=1}^n f(x_i^*) \Delta x_i, where the limit is taken as the norm \|P\| (the maximum length of the subintervals) of the partition P approaches zero, using Riemann sums to approximate areas or net changes, and the indefinite integral, which yields the general antiderivative \int f(x) \, dx = F(x) + C, where F'(x) = f(x) and C is the constant of integration. For details on these limitations, see the Formal Definitions section.[3][4][5]The development of integral calculus is credited to Isaac Newton and Gottfried Wilhelm Leibniz in the late 17th century, who independently formulated its principles around 1665–1676, with Leibniz introducing the integral symbol \int in 1675 to denote summation of infinitesimals.[6] Their work, building on earlier ideas from mathematicians like Archimedes and Cavalieri on quadrature methods, culminated in the Fundamental Theorem of Calculus, which states that if f is continuous on [a, b] and F is an antiderivative of f, then \int_a^b f(x) \, dx = F(b) - F(a), linking differentiation and integration as inverse processes.[7][8]Integrals are essential across sciences and engineering for modeling phenomena like motion, probability distributions, and optimization problems; for instance, in physics, the definite integral calculates work as W = \int_a^b F(x) \, dx, where F(x) is force.[9] Modern extensions include multiple integrals for volumes in higher dimensions and improper integrals for infinite domains, enabling applications in fields from economics to machine learning.[1][10][11]
Historical Development
Precursors to Calculus
The method of exhaustion, developed by ancient Greek mathematicians, provided an early framework for computing areas under curves by approximating them with increasingly finer polygons, effectively bounding the area between inscribed and circumscribed figures until the difference vanished. Eudoxus of Cnidus (c. 408–355 BCE) laid the groundwork for this technique, which Archimedes of Syracuse (c. 287–212 BCE) refined and applied rigorously in his work Quadrature of the Parabola. In this treatise, Archimedes demonstrated that the area of a segment of a parabola is \frac{4}{3} times the area of the triangle inscribed in that segment with the same base and height, achieved by iteratively inscribing triangles within the parabolic region and showing the exhaustion process converges to this value.[12]In the 17th century, Italian mathematician Bonaventura Cavalieri (1598–1647) advanced these ideas through his method of indivisibles, introduced in Geometria indivisibilibus continuorum nova quadam ratione promota (1635), where plane figures were treated as stacks of infinitely thin lines and volumes as stacks of infinitely thin planes, allowing comparisons of areas and volumes by equating the sums of these indivisibles. This approach, while not fully rigorous, enabled computations such as the area under a hyperbola, foreshadowing summation techniques. Cavalieri's method built on earlier geometric intuitions but shifted toward a more algebraic handling of infinite sums, influencing subsequent mathematicians.[13][14]English mathematician John Wallis (1616–1703) further developed these summation concepts in Arithmetica infinitorum (1656), where he interpolated areas under curves using patterns of fractions to approximate integrals, most notably deriving an infinite product formula for \frac{\pi}{2} = \prod_{n=1}^{\infty} \frac{(2n)^2}{(2n-1)(2n+1)} = \frac{2}{1} \cdot \frac{2}{3} \cdot \frac{4}{3} \cdot \frac{4}{5} \cdot \frac{6}{5} \cdot \frac{6}{7} \cdots. This work extended Cavalieri's indivisibles by treating areas as limits of discrete sums, providing a bridge from geometric exhaustion to algebraic series. Scottish mathematician James Gregory (1638–1675) similarly explored arclengths in the early 1660s, approximating them as sums of infinitesimal segments along curves, as seen in his unpublished manuscripts and early publications like Geometriae pars universalis (1668), which treated such sums as precursors to integration.[15][16]These pre-1660s developments in exhaustion, indivisibles, and infinite sums laid essential groundwork for the summation processes that Leibniz and Newton would formalize in their invention of calculus.[17]
Leibniz and Newton Contributions
Isaac Newton developed the foundational ideas of calculus during his annus mirabilis in 1665–1666, extending them through 1669, with integration conceived as the inverse process to finding fluxions (derivatives) and as a method of summing infinitesimally small quantities to compute areas under curves and distances from velocities.[18] In this framework, known as the method of fluxions, a fluent represented a varying quantity such as position or area, while its fluxion was the instantaneous rate of change; integration recovered the fluent by accumulating these fluxions as an infinite series, akin to summing rectangular areas in the limit.[18] Newton's approach drew inspiration from earlier summation methods, such as Bonaventura Cavalieri's indivisibles, but formalized them within a dynamic, physical context of motion.[17]Newton documented these concepts in his 1671 manuscript De Analysi per Aequationes Numero Terminorum Infinitas, where he demonstrated how to integrate by expanding functions into infinite power series and term-by-term summation to find areas and tangents, though the work remained unpublished until 1711.[19] He applied fluxions extensively to problems in mechanics, notably in Philosophiæ Naturalis Principia Mathematica (1687), where geometric formulations disguised the underlying calculus to derive planetary orbits under gravitational forces, treating integration as essential for computing trajectories from accelerations.[18]Independently, Gottfried Wilhelm Leibniz began developing his version of calculus in the early 1670s, introducing the integral symbol ∫ in a private manuscript dated November 21, 1675, while in Paris, where it represented the "summa" (sum) of the function values times infinitesimal dx, explicitly linking integration to summation of differences.[20] Leibniz's system integrated this with his differential calculus, viewing ∫ f(x) dx as the antiderivative obtained by reversing the operation of finding differentials (dx as infinitesimals), and he emphasized its utility for solving geometric and physical problems through symbolic manipulation.[20]Leibniz first published his calculus, including rules for integration, in two seminal papers in Acta Eruditorum: "Nova Methodus pro Maximis et Minimis" in October 1684, focusing on differentials, and a follow-up in June 1686 extending to integrals and their applications.[20] These works established his notation and methods publicly, influencing continental mathematicians.The independent inventions sparked a prolonged priority dispute beginning around 1710, fueled by Newton's influence at the Royal Society, which in 1711 issued a report—authored by Newton himself—accusing Leibniz of plagiarism based on alleged similarities from Leibniz's 1676 visit to England; however, subsequent scholarship confirms their parallel developments without direct borrowing.[6] This controversy overshadowed early adoption but highlighted the revolutionary impact of their integration techniques on astronomy and mechanics, such as Newton's orbital computations.[6]
Formalization and Notation Evolution
In the 19th century, mathematicians sought to provide a rigorous foundation for integration, building briefly on the 17th-century insights of Newton and Leibniz that linked integrals to antiderivatives.[21]A key advancement came from Augustin-Louis Cauchy, who in 1823 formalized the definite integral in his Calcul Infinitésimal as the limit of sums derived from partitions of the integration interval.[22] He divided the interval [x_0, X] into subintervals via points \{x_0, x_1, \dots, x_{n-1}, X\}, formed sums S = \sum (x_k - x_{k-1}) f(x_{k-1}) using the function value at the left endpoint of each subinterval, and required the maximum subinterval length (mesh) to approach zero through refinements, ensuring the limit exists independently of the partition choice.[22] This approach, resembling modern Riemann sums, applied to continuous functions and emphasized the stability of the sums under finer divisions.[22]Peter Gustav Lejeune Dirichlet contributed conditions for the convergence of Fourier series in 1829, particularly ensuring well-defined integrals for Fourier coefficients of bounded, periodic functions over a period like [0, 2\pi] that are piecewise continuous, with a finite number of discontinuities and extrema per period. Under these conditions, the Fourier series converges pointwise to the function where it is continuous and to the average of the left and right limits at jump discontinuities.[23]Karl Weierstrass elevated the precision of these developments in the 1860s by introducing the epsilon-delta definition in his Berlin lectures.[24] In 1861, he formalized continuity at a point x for a function f as: for every \epsilon > 0, there exists \delta > 0 such that if |h| < \delta, then |f(x + h) - f(x)| < \epsilon, extending this to limits and uniform continuity across intervals.[24] This rigorous framework underpinned integrability by clarifying when functions admit limits of sums without relying on intuitive notions of "infinitesimals," influencing later definitions of integrable functions.[24]The notation for integration also evolved during this period, standardizing from Gottfried Wilhelm Leibniz's 1675 indefinite integral symbol \int f(x) \, dx—representing an infinite sum—to the definite form \int_a^b f(x) \, dx.[21] Joseph Fourier introduced the limits of integration a and b below and above the symbol in his 1822 Théorie Analytique de la Chaleur, a convention subsequently adopted by Cauchy and others to denote the value over a finite interval.[21]
Origin of the Term "Integral"
The term "integral" in mathematics originates from the Latin adjective integralis, derived from integer meaning "whole" or "untouched," signifying the process of combining parts into a complete entity. This etymology underscores the philosophical underpinnings of integration as the inverse operation to differentiation, where infinitesimal differences are summed or "made whole" to reconstruct a function or quantity. Gottfried Wilhelm Leibniz, who developed much of the foundational notation for calculus, initially conceptualized integration in this light during the 1670s, viewing it as a summation of infinitesimals akin to restoring continuity from discrete elements.[25]The first documented use of "integral" with its modern meaning in integration appeared in 1690, in a paper by Jacob Bernoulli published in Acta Eruditorum. Bernoulli employed the term to describe the antiderivative or the solution to a differential equation in his study of the isochrone curve, marking the initial application of "integral" as the operation reversing differentiation. This usage contrasted with earlier geometric terms like "quadrature" and aligned integration explicitly with the emerging calculus framework introduced by Leibniz. Prior to this, Leibniz had referred to the process as calculus summatorius (calculus of summing) in his unpublished manuscripts, emphasizing its summative nature without adopting the specific nomenclature.[26]Philosophically, the term reflected a shift toward viewing integration not merely as area computation but as a constructive process completing the "dismemberment" of differentiation. Leibniz and the Bernoulli brothers elaborated this in correspondence during the mid-1690s; in letters exchanged around 1695–1696, Johann Bernoulli suggested calculus integralis to Leibniz, who accepted it as a more precise descriptor than his preferred summatorius. This adoption highlighted integration's role in philosophical completeness, paralleling Leibniz's monadic metaphysics where complex wholes emerge from simple components. Jacob Bernoulli's 1690 paper and subsequent works by his brother Johann, including applications in 1710 responses to contemporaries like Nicolaus Hermann, further entrenched the term in key texts on variational problems and mechanics. By the early 18th century, calculus integralis gained widespread acceptance across European mathematical discourse, supplanting ad hoc phrases.[27][28]
Core Concepts
Terminology and Notation
The integral sign ∫, introduced by Gottfried Wilhelm Leibniz in 1675, is derived from the elongated form of the letter "s" in the Latin word summa, reflecting the summation aspect of integration.[21]In standard notation, the indefinite integral of a function f(x) is denoted by \int f(x) \, dx, which represents the family of antiderivatives F(x) + C, where F'(x) = f(x) and C is the constant of integration.[29] The definite integral over an interval from a to b is written as \int_a^b f(x) \, dx, yielding a numerical value that depends on the specific limits.[30]Key terms in integration include the integrand, which is the function f(x) being integrated; the antiderivative or primitive F(x), satisfying F'(x) = f(x); and the limits of integration, where a denotes the lower limit and b the upper limit in a definite integral.[29]The variable of integration, such as x in \int f(x) \, dx, functions as a dummy variable, meaning its specific name does not affect the value of the integral and can be replaced by another symbol without changing the result, analogous to an index in a summation.In multiple dimensions, conventions for integrals include the order of differentials in iterated integrals, such as \iint_R f(x,y) \, dx \, dy where integration proceeds first with respect to x then y, and for line or surface integrals, the orientation is specified to ensure consistency, often using the right-hand rule for positive direction in vector calculus contexts.[31]
Interpretations of Integration
The definite integral represents the net signed area between the graph of a continuous function f(x) and the x-axis over the interval from a to b, where areas above the x-axis contribute positively and those below contribute negatively.[3] This geometric interpretation, denoted as \int_a^b f(x) \, dx, provides an intuitive way to quantify regions bounded by curves, such as calculating the area under a velocity-time graph to find displacement.[32]In physics, integrals often model the accumulation of quantities varying continuously along a path or over time. For instance, the work done by a variable force F(x) along a straight line is given by \int_a^b F(x) \, dx, representing the total energy transferred as the force acts over displacement from a to b.[9] Similarly, the total charge Q along a wire with linear charge density \rho(x) is \int_a^b \rho(x) \, dx, summing infinitesimal charges dq = \rho(x) \, dx across the length.[33]Analytically, the indefinite integral \int f(x) \, dx denotes the family of all antiderivative functions F(x) such that F'(x) = f(x), differing only by a constant C, and serves as the solution to the differential equation f' = g where g(x) = f(x).[34] This interpretation views integration as the inverse operation of differentiation, generating a general solution that captures all possible primitives of the integrand.In probability theory, for a continuous random variable X with probability density function f(x) over an interval, the expected value E[X] is computed as \int_{-\infty}^{\infty} x f(x) \, dx, weighting each possible value x by its likelihood to yield the long-run average outcome.[35] This integral interpretation extends the concept of accumulation to measure central tendency in distributions, such as the mean lifetime in reliability analysis.[36]
Formal Definitions
Riemann Integral
The Riemann integral, introduced by the German mathematician Bernhard Riemann in his 1854 habilitation thesis, provides a formal definition of the definite integral for bounded functions on a closed interval [a, b].[37] This construction generalizes earlier notions from Cauchy and others by removing the continuity assumption and focusing on the limit of sums over partitions, allowing integration of certain discontinuous functions.[38]To define the Riemann integral, consider a bounded function f: [a, b] \to \mathbb{R}. A partition P of [a, b] is a finite sequence of points a = x_0 < x_1 < \cdots < x_n = b, dividing the interval into subintervals I_k = [x_{k-1}, x_k] for k = 1, \dots, n. The norm of the partition, denoted \|P\|, is the maximum length of these subintervals, \|P\| = \max_k (x_k - x_{k-1}). For a choice of points \xi_k \in I_k, the corresponding Riemann sum isS(f, P, \xi) = \sum_{k=1}^n f(\xi_k) (x_k - x_{k-1}).The function f is Riemann integrable on [a, b] if there exists a number I such that for every \epsilon > 0, there is \delta > 0 where, for any partition P with \|P\| < \delta and any choice of \xi_k, |S(f, P, \xi) - I| < \epsilon. In this case, I is the Riemann integral, denoted \int_a^b f(x) \, dx.[39]An equivalent formulation, known as the Darboux integral after Gaston Darboux who refined it in 1875, uses upper and lower sums to characterize integrability without relying on specific point choices. For a partition P, the upper sum is U(f, P) = \sum_{k=1}^n M_k (x_k - x_{k-1}), where M_k = \sup_{I_k} f, and the lower sum is L(f, P) = \sum_{k=1}^n m_k (x_k - x_{k-1}), where m_k = \inf_{I_k} f. The upper integral is \overline{\int_a^b} f(x) \, dx = \inf_P U(f, P) and the lower integral is \underline{\int_a^b} f(x) \, dx = \sup_P L(f, P). The function f is Riemann integrable if these coincide, with the integral equal to their common value. This Darboux approach is equivalent to the original Riemann definition, as both yield the same integrable functions and integral values.[38][40]A key criterion for Riemann integrability is continuity: every continuous function on the compact interval [a, b] is Riemann integrable. This follows from the uniform continuity of f on [a, b], which ensures that for fine enough partitions, the difference between upper and lower sums can be made arbitrarily small. More generally, a bounded function on [a, b] is Riemann integrable if its set of discontinuities has Lebesgue measure zero; in particular, functions with only finitely many discontinuities are integrable.[40][39]
Lebesgue Integral
The Lebesgue integral extends the concept of integration to a broader class of functions by incorporating measure theory on a measure space (X, \Sigma, \mu), where \mu is a measure. Introduced by Henri Lebesgue in his 1902 dissertation Intégrale, longueur, aire, it defines the integral for measurable functions, allowing integration over sets of arbitrary measure and functions with substantial discontinuities.[41] The construction begins with simple functions, which are finite sums of the form \phi = \sum_{i=1}^n c_i \chi_{E_i}, where c_i \geq 0 are constants, \chi_{E_i} is the indicator function of a measurable set E_i \in \Sigma, and the E_i are disjoint. The integral of such a \phi is given by\int_X \phi \, d\mu = \sum_{i=1}^n c_i \mu(E_i),with the convention that $0 \cdot \infty = 0.[41]For a non-negative measurable function f: X \to [0, \infty], the Lebesgue integral is defined as the supremum of the integrals of simple functions bounded above by f:\int_X f \, d\mu = \sup\left\{ \int_X \phi \, d\mu : 0 \leq \phi \leq f, \, \phi \text{ simple} \right\}.This definition relies on the monotone convergence theorem: if \{f_n\} is a sequence of non-negative measurable functions with $0 \leq f_n \uparrow f pointwise (i.e., f_n(x) \leq f_{n+1}(x) for all n and x, and f_n(x) \to f(x) as n \to \infty), then\lim_{n \to \infty} \int_X f_n \, d\mu = \int_X f \, d\mu.Such a sequence \{f_n\} of simple functions can always be constructed to approximate f monotonically, ensuring the integral is well-defined and independent of the approximating sequence.[41]An alternative characterization, known as the layer-cake representation, expresses the integral of a non-negative measurable f in terms of the measure of level sets:\int_X f \, d\mu = \int_0^\infty \mu(\{x \in X : f(x) > t \}) \, dt.This formula, derivable via monotone convergence applied to the functions f_n(t) = \min(n, f) \wedge t or directly from Fubini's theorem on the product space, provides an intuitive "slicing" view of the integral by accumulating contributions from horizontal layers at height t.[42]A key tool for interchanging limits and integrals is the dominated convergence theorem: let \{f_n\} be a sequence of measurable functions converging pointwise to a measurable f, and suppose there exists an integrable g: X \to [0, \infty) such that |f_n(x)| \leq g(x) for all n and almost all x \in X; then f is integrable and\lim_{n \to \infty} \int_X f_n \, d\mu = \int_X f \, d\mu.
$$
This theorem enables robust handling of limits in integration, far beyond what is possible with Riemann sums.[](https://www.math.ucdavis.edu/~hunter/m206/ch3_measure_notes.pdf)
The Lebesgue integral's reliance on measure overcomes limitations of the Riemann integral for highly discontinuous functions. For instance, the Dirichlet function $D: [0,1] \to \{0,1\}$, defined by $D(x) = 1$ if $x$ is rational and $D(x) = 0$ if $x$ is irrational, is discontinuous at every point in $[0,1]$ and thus not Riemann integrable. However, since the rational numbers have Lebesgue measure zero, $D = 0$ almost everywhere, so $\int_0^1 D \, d\mu = 0$.[](https://www.stats.ox.ac.uk/~etheridg/integration.pdf) For continuous functions on compact intervals, the Lebesgue integral coincides with the Riemann integral.[](https://www.math.ucdavis.edu/~hunter/m206/ch3_measure_notes.pdf)
### Other Generalizations
The Henstock–Kurzweil integral, independently developed by Jaroslav Kurzweil in 1957 and Ralph Henstock in the early 1960s, extends the Riemann integral by incorporating a gauge—a positive function δ on the interval—to refine partitions more flexibly than a uniform mesh. A bounded function $f$ on $[a, b]$ is Henstock–Kurzweil integrable if there exists $I \in \mathbb{R}$ such that for every $\epsilon > 0$, there is a gauge $\delta$ where, for any $\delta$-fine tagged partition $\{(t_i, [x_{i-1}, x_i])_{i=1}^n\}$ with $t_i \in [x_{i-1}, x_i]$ and $x_i - x_{i-1} < \delta(x_i)$ (or similar gauge condition), $\left| \sum_{i=1}^n f(t_i)(x_i - x_{i-1}) - I \right| < \epsilon$. This gauge-based definition allows integration of functions like $\frac{\sin x}{x}$ over $\mathbb{R}$ without absolute convergence, equating to the Lebesgue integral on Lebesgue-integrable functions while enabling derivatives under the integral sign for a wider class.[](http://www.math.vanderbilt.edu/~schectex/ccc/gauge/)
The Itô stochastic integral, introduced by Kiyosi Itô in his 1944 paper, provides a framework for integrating adapted processes with respect to semimartingales, particularly Brownian motion, addressing the non-differentiability of paths.[](https://djalil.chafai.net/docs/M2/history-brownian-motion/Ito%2520-%25201944.pdf) For a progressively measurable process $\phi$ with $\mathbb{E}\left[\int_0^T \phi_t^2 dt\right] < \infty$, the Itô integral $\int_0^T \phi_t \, dW_t$ is defined as the $L^2$-limit of left-endpoint Riemann sums $\sum \phi_{t_i} (W_{t_{i+1}} - W_{t_i})$ over partitions of $[0, T]$, yielding a martingale. This construction underpins stochastic differential equations like $dX_t = \mu(X_t) dt + \sigma(X_t) dW_t$, where Itô's integral resolves the quadratic variation issue absent in classical calculus, with Itô's work from the 1940s formalizing solutions to such equations via fixed-point arguments.
Distributional integrals arise in the theory of generalized functions, or distributions, where integration is defined by duality against smooth test functions with compact support. Pioneered by Laurent Schwartz in his 1950–1951 treatise, a distribution $T$ acts on a test function $\phi \in \mathcal{D}(\mathbb{R}^n)$ via $\langle T, \phi \rangle$, generalizing the Lebesgue integral as $\langle T, \phi \rangle = \int T(x) \phi(x) \, dx$ for regular distributions $T$. This framework accommodates singular objects like the Dirac delta, $\langle \delta, \phi \rangle = \phi(0)$, and enables formal integration of derivatives or products in contexts where pointwise definitions fail, building on measure-theoretic foundations for broader applicability in partial differential equations.
## Fundamental Properties
### Linearity and Basic Operations
One of the fundamental properties of the Riemann integral is its linearity, which asserts that if functions $f$ and $g$ are Riemann integrable on $[a, b]$ and $c \in \mathbb{R}$, then $cf$ and $f + g$ are also Riemann integrable on $[a, b]$, with\int_a^b (cf + dg) , dx = c \int_a^b f , dx + d \int_a^b g , dx.This property follows directly from the definition using upper and lower sums, which are limits of Riemann sums. To prove scalar multiplication, note that for $c \geq 0$, the upper sum $U(cf; P) = c U(f; P)$ and lower sum $L(cf; P) = c L(f; P)$ for any partition $P$, so the infimum of upper sums and supremum of lower sums scale by $c$, yielding the result; for $c < 0$, the inequalities reverse but the equality holds similarly. For addition, $L(f + g; P) \geq L(f; P) + L(g; P)$ and $U(f + g; P) \leq U(f; P) + U(g; P)$, and since $f$ and $g$ are integrable, the upper and lower integrals coincide for $f + g$, equaling the sum of the individual integrals.[](https://www.math.ucdavis.edu/~hunter/m125b/ch1.pdf)
The Riemann integral also exhibits additivity over adjacent intervals: if $f$ is Riemann integrable on $[a, c]$ with $a \leq b \leq c$, then\int_a^c f(x) , dx = \int_a^b f(x) , dx + \int_b^c f(x) , dx.This follows from the definition by combining partitions of $[a, b]$ and $[b, c]$ into a partition of $[a, c]$, where the Riemann sums add correspondingly, and the limits preserve the equality as the mesh approaches zero; the property holds for arbitrary ordering via the convention $\int_b^a = -\int_a^b$.[](http://educ.jmu.edu/~waltondb/MA2C/pdf/section.6.2.pdf)
Integration by parts provides a basic operation for products of functions, derived from the product rule for differentiation. Consider differentiable functions $u(x)$ and $v(x)$ on $[a, b]$, where $u'$ and $v'$ lead to integrable $u \, dv$ and $v \, du$. The product rule states\frac{d}{dx} [u(x) v(x)] = u(x) v'(x) + v(x) u'(x).Integrating both sides over $[a, b]$ gives\int_a^b \frac{d}{dx} [u v] , dx = \int_a^b u , dv + \int_a^b v , du,and the left side evaluates to $u(b) v(b) - u(a) v(a)$ by the fundamental theorem of calculus, yielding the formula\int_a^b u , dv = \left[ u v \right]_a^b - \int_a^b v , du.This holds under Riemann integrability assumptions for the involved functions.[](https://tutorial.math.lamar.edu/classes/calcii/integrationbyparts.aspx)
The substitution rule, or change of variables, simplifies integrals of composite functions and derives from the chain rule. Suppose $f$ is continuous on an interval containing the range of differentiable $g$ on $[a, b]$, with $g'$ Riemann integrable, and let $u = g(x)$, so $du = g'(x) \, dx$. Then\int_a^b f(g(x)) g'(x) , dx = \int_{g(a)}^{g(b)} f(u) , du,provided the integral on the right exists; the formula holds even if $g$ is not monotonic, by applying the chain rule to antiderivatives. This adjusts limits directly in the definite case, avoiding back-substitution.[](https://tutorial.math.lamar.edu/classes/calci/SubstitutionRuleDefinite.aspx)
### Key Inequalities and Conventions
One fundamental inequality in integration theory is the monotonicity property: if $f$ and $g$ are integrable functions on $[a, b]$ with $f(x) \leq g(x)$ for all $x \in [a, b]$, then $\int_a^b f(x) \, dx \leq \int_a^b g(x) \, dx$.[](https://tutorial.math.lamar.edu/classes/calci/ProofIntProp.aspx) This follows from the non-negativity of the integral for non-negative functions, applied to $g - f \geq 0$.
Another key inequality is the triangle inequality for integrals, which states that for an integrable function $f$ on $[a, b]$, $\left| \int_a^b f(x) \, dx \right| \leq \int_a^b |f(x)| \, dx$.[](https://tutorial.math.lamar.edu/classes/calci/ProofIntProp.aspx) This bound arises from the properties of absolute values and ensures that the integral's magnitude is controlled by the total variation of $|f|$.
The mean value theorem for integrals provides a precise connection between the average value of a function and its integral: if $f$ is continuous on the closed interval $[a, b]$, then there exists some $c \in [a, b]$ such that $\int_a^b f(x) \, dx = f(c) (b - a)$.[](https://tutorial.math.lamar.edu/classes/calci/ProofIntProp.aspx) This theorem, a consequence of the extreme value theorem and the fundamental theorem of calculus, implies that the integral equals the function's value at some point times the interval length.
For convex functions, Jensen's inequality extends this idea: if $\phi$ is convex on an interval containing the range of the Riemann integrable function $f$ on $[a, b]$, then $\phi\left( \frac{1}{b-a} \int_a^b f(x) \, dx \right) \leq \frac{1}{b-a} \int_a^b \phi(f(x)) \, dx$.[](https://www2.math.upenn.edu/~gressman/analysis/13-jensen.html) This inequality highlights how convexity preserves inequalities under averaging via integration.
Standard conventions in definite integration include positive orientation, where the integral $\int_a^b f(x) \, dx$ with $a < b$ traverses the interval from left to right, reflecting the geometric interpretation of net area accumulation in the positive direction.[](https://mathresearch.utsa.edu/wiki/index.php?title=Properties_of_the_Integral) Absolute integrability requires that $\int |f| < \infty$ for the function to be integrable in the Lebesgue sense, ensuring the integral exists even for functions with potential sign changes or singularities, though Riemann integrability on compact intervals typically assumes boundedness.[](https://ocw.mit.edu/courses/18-102-introduction-to-functional-analysis-spring-2021/resources/lecture-12-lebesgue-integrable-functions-the-lebesgue-integral-and-the-dominated-convergence-theorem/)
## Fundamental Theorem of Calculus
### First Fundamental Theorem
The first fundamental theorem of calculus establishes that differentiation and integration are inverse operations under suitable conditions. Specifically, if $ f $ is a continuous function on an interval $[a, b]$, then the function defined by $ F(x) = \int_a^x f(t) \, dt $ for $ a \leq x \leq b $ is differentiable on $ (a, b) $, and its derivative satisfies $ F'(x) = f(x) $.[](https://mathworld.wolfram.com/FirstFundamentalTheoremofCalculus.html)[](https://www3.nd.edu/~apilking/Math10550/Lectures/26.%20Fundamental%20Theorem.pdf) This result demonstrates that $ F(x) $ serves as an antiderivative of $ f $, meaning $ F $ is a function whose derivative recovers the original integrand $ f $.[](https://mathworld.wolfram.com/FirstFundamentalTheoremofCalculus.html)
A rigorous proof of this theorem relies on the continuity of $ f $ and properties of definite integrals. To show $ F'(x) = f(x) $, consider the difference quotient:\frac{F(x + h) - F(x)}{h} = \frac{1}{h} \int_x^{x+h} f(t) , dtfor small $ h \neq 0 $. By the mean value theorem for integrals, which applies because $ f $ is continuous on the compact interval $[x, x+h]$, there exists some $ c $ between $ x $ and $ x+h $ such that $ \int_x^{x+h} f(t) \, dt = f(c) \cdot h $. Thus, the difference quotient simplifies to $ f(c) $. Taking the limit as $ h \to 0 $, $ c \to x $, and by continuity of $ f $, $ f(c) \to f(x) $, yielding $ F'(x) = f(x) $.[](https://www3.nd.edu/~apilking/Math10550/Lectures/26.%20Fundamental%20Theorem.pdf)
This theorem was first proved rigorously by Augustin-Louis Cauchy in his 1823 work *Résumé des leçons sur le calcul infinitésimal*, where he employed the mean value theorem for integrals to link the processes of differentiation and integration within a limit-based framework.[](https://link.springer.com/book/10.1007/978-3-030-11036-9)[](https://www3.nd.edu/~apilking/Math10550/Lectures/26.%20Fundamental%20Theorem.pdf) Cauchy's contribution provided the foundational rigor that unified the two branches of calculus, resolving earlier intuitive approaches.[](https://link.springer.com/book/10.1007/978-3-030-11036-9)
### Second Fundamental Theorem
The second fundamental theorem of calculus provides a method for evaluating definite integrals by relating them directly to antiderivatives. Specifically, if $ f $ is continuous on the closed interval $[a, b]$ and $ F $ is any antiderivative of $ f $ (that is, $ F'(x) = f(x) $ for all $ x \in [a, b] $), then\int_a^b f(x) , dx = F(b) - F(a).[](http://aleph0.clarku.edu/~ma121/FTCproof.pdf) This result holds under the continuity assumption on $ f $, ensuring the existence and uniqueness of the antiderivative up to a constant.[](https://people.math.harvard.edu/~knill/teaching/math1a_2011/handouts/29-fundamental.pdf)
The proof relies on the first fundamental theorem of calculus. Define the function $ G(x) = \int_a^x f(t) \, dt $ for $ x \in [a, b] $. By the first fundamental theorem, $ G $ is differentiable on $[a, b]$ with $ G'(x) = f(x) $, so $ G $ is itself an antiderivative of $ f $. Since antiderivatives differ only by a constant, there exists a constant $ C $ such that $ G(x) = F(x) + C $ for all $ x \in [a, b] $. Evaluating at $ x = a $ gives $ G(a) = F(a) + C = 0 $, so $ C = -F(a) $. Thus, $ G(x) = F(x) - F(a) $, and substituting $ x = b $ yields $ \int_a^b f(t) \, dt = G(b) = F(b) - F(a) $.[](http://aleph0.clarku.edu/~ma121/FTCproof.pdf)[](https://mathbooks.unl.edu/Calculus/sec-5-4-FTC2.html)
This theorem interprets the definite integral as the net change in the antiderivative function over the interval, often called the net change theorem. For instance, if $ f(x) $ represents velocity, then $ \int_a^b f(x) \, dx = F(b) - F(a) $ gives the net displacement from position $ F(a) $ at time $ a $ to $ F(b) $ at time $ b $.[](https://mathbooks.unl.edu/Calculus/sec-5-4-FTC2.html)[](https://people.math.harvard.edu/~knill/teaching/math1a_2011/handouts/29-fundamental.pdf)
As a corollary, the second fundamental theorem is essential for practical computation of definite integrals, allowing evaluation via antiderivatives rather than direct approximation methods like Riemann sums. For example, to compute $ \int_0^1 x^2 \, dx $, note that $ F(x) = \frac{x^3}{3} $ satisfies $ F'(x) = x^2 $, so the integral equals $ F(1) - F(0) = \frac{1}{3} $. This approach leverages linearity for more complex integrals but remains foundational for analytical evaluation.[](http://aleph0.clarku.edu/~ma121/FTCproof.pdf)[](https://people.math.harvard.edu/~knill/teaching/math1a_2011/handouts/29-fundamental.pdf)
## Extensions and Generalizations
### Improper Integrals
Improper integrals extend the concept of the Riemann integral to functions defined on unbounded intervals or with discontinuities within the interval of integration, defined through limiting processes. For an integral over an unbounded domain, such as $\int_a^\infty f(x) \, dx$, it is defined as $\lim_{b \to \infty} \int_a^b f(x) \, dx$, where the integral converges if the limit exists and is finite, and diverges otherwise. Similarly, for a discontinuity at a point $c$ in $[a, b]$, the improper integral $\int_a^b f(x) \, dx$ is $\lim_{\epsilon \to 0^+} \left( \int_a^{c - \epsilon} f(x) \, dx + \int_{c + \epsilon}^b f(x) \, dx \right)$, converging if this limit is finite. Convergence criteria include the comparison test: if $0 \leq f(x) \leq g(x)$ for $x \geq a$ and $\int_a^\infty g(x) \, dx$ converges, then $\int_a^\infty f(x) \, dx$ converges; if $\int_a^\infty f(x) \, dx$ diverges and $g(x) \leq f(x)$, then $\int_a^\infty g(x) \, dx$ diverges.
For integrals over symmetric unbounded intervals like $(-\infty, \infty)$, the Cauchy principal value addresses cases where separate limits from left and right may diverge but a symmetric limit exists, defined as $\mathrm{P.V.} \int_{-\infty}^\infty f(x) \, dx = \lim_{R \to \infty} \int_{-R}^R f(x) \, dx$ if the limit is finite. A classic example is $\mathrm{P.V.} \int_{-\infty}^\infty \frac{dx}{x} = 0$, where the principal value symmetrizes the divergent contributions from positive and negative infinities. In contrast, the Gaussian integral $\int_{-\infty}^\infty e^{-x^2} \, dx = \sqrt{\pi}$ converges absolutely, as the rapid decay of the exponential ensures the limit exists without needing principal value adjustments. The value $\sqrt{\pi}$ is obtained via techniques like polar coordinate substitution or differentiation under the integral sign.
The harmonic integral $\int_1^\infty \frac{dx}{x}$ diverges, mirroring the divergence of the harmonic series, as the antiderivative $\ln x$ grows without bound as the upper limit approaches infinity. This divergence highlights the logarithmic growth's insufficiency for convergence over unbounded domains. Distinguishing absolute and conditional convergence is crucial: an improper integral converges absolutely if $\int |f(x)| \, dx$ converges, implying ordinary convergence; it converges conditionally if $\int f(x) \, dx$ converges but $\int |f(x)| \, dx$ diverges, often due to oscillatory behavior canceling divergences. For instance, $\int_0^\infty \frac{\sin x}{x} \, dx$ converges conditionally to $\frac{\pi}{2}$, while $\int_0^\infty |\frac{\sin x}{x}| \, dx$ diverges.
### Multiple and Vector Integrals
Multiple integrals extend the concept of the single-variable integral to functions over regions in higher-dimensional Euclidean spaces, such as $\mathbb{R}^n$. For a function $f: D \to \mathbb{R}$ where $D \subset \mathbb{R}^n$ is a bounded region, the multiple integral $\int_D f(\mathbf{x}) \, d\mathbf{x}$ represents the signed volume under the graph of $f$ over $D$, analogous to the area under a curve in one dimension. These integrals are foundational in vector calculus for computing quantities like mass, center of mass, and flux in multi-dimensional settings.[](https://tutorial.math.lamar.edu/classes/calciii/IteratedIntegrals.aspx)
In two dimensions, the double integral $\iint_D f(x,y) \, dA$ over a region $D$ can be evaluated as an iterated integral, typically $\int_a^b \int_{g_1(x)}^{g_2(x)} f(x,y) \, dy \, dx$ for Type I regions where $D = \{(x,y) \mid a \leq x \leq b, g_1(x) \leq y \leq g_2(x)\}$, or similarly for Type II regions by integrating with respect to $x$ first. This reduction to nested single integrals leverages the structure of rectangular or simple bounded domains. Fubini's theorem states that if $f$ is continuous on the closed rectangle $[a,b] \times [c,d]$, then the double integral equals the iterated integral in either order: $\iint_R f(x,y) \, dA = \int_a^b \int_c^d f(x,y) \, dy \, dx = \int_c^d \int_a^b f(x,y) \, dx \, dy$. This interchange simplifies computations by allowing the order of integration to be chosen based on convenience.[](https://web.ma.utexas.edu/users/m408m/Display15-2-3.shtml)[](https://tutorial.math.lamar.edu/classes/calciii/IteratedIntegrals.aspx)
Line integrals arise in vector calculus to integrate vector fields along curves in $\mathbb{R}^n$. For a vector field $\mathbf{F}: \mathbb{R}^n \to \mathbb{R}^n$ and a piecewise smooth curve $C$ parametrized by $\mathbf{r}(t)$ for $t \in [a,b]$, the line integral is $\int_C \mathbf{F} \cdot d\mathbf{r} = \int_a^b \mathbf{F}(\mathbf{r}(t)) \cdot \mathbf{r}'(t) \, dt$, measuring the work done by $\mathbf{F}$ along $C$. Surface integrals generalize this to integration over surfaces in $\mathbb{R}^3$. For a scalar function $g$ over an oriented surface $S$ parametrized by $\mathbf{r}(u,v)$ for $(u,v) \in D$, the surface integral is $\iint_S g \, dS = \iint_D g(\mathbf{r}(u,v)) \|\mathbf{r}_u \times \mathbf{r}_v\| \, du \, dv$; for a vector field $\mathbf{F}$, the flux integral is $\iint_S \mathbf{F} \cdot d\mathbf{S} = \iint_D \mathbf{F}(\mathbf{r}(u,v)) \cdot (\mathbf{r}_u \times \mathbf{r}_v) \, du \, dv$. These quantify flow or total scalar accumulation across the surface.[](https://tutorial.math.lamar.edu/classes/calciii/LineIntegralsVectorFields.aspx)[](https://tutorial.math.lamar.edu/classes/calciii/surfaceintegrals.aspx)
Green's theorem connects line integrals to double integrals over planar regions. For a positively oriented, piecewise smooth, simple closed curve $C$ bounding a region $D$ in $\mathbb{R}^2$, and vector field $\mathbf{F} = (P, Q)$ with continuous partial derivatives in an open region containing $D$,
\int_C P , dx + Q , dy = \iint_D \left( \frac{\partial Q}{\partial x} - \frac{\partial P}{\partial y} \right) dA.
This equates the circulation of $\mathbf{F}$ around $C$ to the double integral of its curl over $D$, enabling efficient evaluation of one form using the other.[](https://tutorial.math.lamar.edu/classes/calciii/GreensTheorem.aspx)
Change of variables in multiple integrals accounts for nonlinear transformations of the domain. For a differentiable, one-to-one transformation $\mathbf{T}: D^* \to D$ with continuous partial derivatives and nonzero Jacobian determinant, the double integral transforms as $\iint_D f(x,y) \, dA = \iint_{D^*} f(\mathbf{T}(u,v)) \left| \det \frac{\partial(x,y)}{\partial(u,v)} \right| \, du \, dv$, where the Jacobian determinant $\det \frac{\partial(x,y)}{\partial(u,v)} = \frac{\partial x}{\partial u} \frac{\partial y}{\partial v} - \frac{\partial x}{\partial v} \frac{\partial y}{\partial u}$ scales the area element. This formula, generalizable to higher dimensions, simplifies integrals over irregular regions by mapping to standard coordinates, such as polar or spherical systems. Linearity of the integral extends naturally to multiple integrals, preserving additivity over disjoint domains.[](https://tutorial.math.lamar.edu/classes/calciii/changeofvariables.aspx)
### Advanced Forms and Summations
In complex analysis, the contour integral of a function $f(z)$ along a curve $\gamma$ in the complex plane is defined as $\int_\gamma f(z) \, dz$, where $z$ is parameterized along $\gamma$. This integral generalizes the real line integral and plays a crucial role in evaluating integrals that may be difficult in the real domain. A fundamental result is Cauchy's integral theorem, which states that if $f$ is analytic (holomorphic) everywhere inside and on a simple closed contour $\gamma$, then $\int_\gamma f(z) \, dz = 0$. This theorem implies that the integral depends only on the singularities of $f$ enclosed by $\gamma$, enabling powerful residue calculus techniques for computation.[](https://ocw.mit.edu/courses/18-04-complex-variables-with-applications-spring-2018/b1a90df19f643b974555bbbb93138a48_MIT18_04S18_topic4.pdf)
Integrals of differential forms extend the concept to higher dimensions on smooth manifolds. A differential $k$-form $\omega$ on an oriented $k$-dimensional manifold $M$ can be integrated as $\int_M \omega$, capturing geometric and topological properties of $M$. The generalized Stokes' theorem relates this to the boundary $\partial M$ via $\int_{\partial M} \omega = \int_M d\omega$, where $d$ is the exterior derivative. This theorem unifies classical results like the divergence theorem and Green's theorem, providing a coordinate-free framework for integration on manifolds, with applications in geometry and topology.[](https://maths-people.anu.edu.au/andrews/DG/DG_chap14.pdf)
In quantum mechanics, the path integral formulation introduces functional integrals over infinite-dimensional spaces of paths. Proposed by Richard Feynman, it expresses the transition amplitude from an initial state to a final state as $\int \mathcal{D}\phi \, e^{i S[\phi]/\hbar}$, where the integral is over all possible paths $\phi$ in configuration space, weighted by the exponential of the action functional $S[\phi]$. This approach reformulates quantum mechanics in terms of a sum over histories, bridging classical and quantum descriptions and facilitating quantization of fields.
Discrete analogs of integrals appear in summations, where the Euler-Maclaurin formula approximates sums by integrals with correction terms. It states that for a smooth function $f$ on $[a, b]$,\sum_{n=a}^b f(n) = \int_a^b f(x) , dx + \frac{f(a) + f(b)}{2} + \sum_{k=1}^m \frac{B_{2k}}{(2k)!} \left( f^{(2k-1)}(b) - f^{(2k-1)}(a) \right) + R_m,where $B_{2k}$ are Bernoulli numbers and $R_m$ is a remainder term. Discovered independently by Euler and Maclaurin, this formula links discrete sums to continuous integrals, aiding asymptotic analysis and numerical approximations.[](https://people.csail.mit.edu/kuat/courses/euler-maclaurin.pdf)
## Applications
### In Physics and Engineering
In classical mechanics, integrals are essential for calculating work and energy in systems where forces vary along a path. The work $ W $ done by a force field $ \mathbf{F} $ on a particle moving along a curve $ C $ is given by the line integral $ W = \int_C \mathbf{F} \cdot d\mathbf{r} $, which quantifies the energy transferred to the object. This formulation arises from the definition of infinitesimal work $ dW = \mathbf{F} \cdot d\mathbf{r} $, summed over the path. For conservative forces, this integral is path-independent, linking directly to potential energy differences.[](https://openstax.org/books/calculus-volume-3/pages/6-2-line-integrals)
The work-energy theorem further demonstrates integrals' role by deriving kinetic energy. Starting from Newton's second law, the work done equals the change in kinetic energy: $ W = \int \mathbf{F} \cdot d\mathbf{r} = \int m \frac{dv}{dt} \cdot v \, dt = \int m v \, dv = \frac{1}{2} m v^2 - \frac{1}{2} m v_0^2 $, where the integral transforms momentum changes into the familiar kinetic energy expression. This derivation is foundational for analyzing variable acceleration in mechanics, such as in projectile motion or orbital dynamics.[](https://phys.libretexts.org/Courses/Berea_College/Introductory_Physics%253A_Berea_College/07%253A_Work_and_energy/7.02%253A_Kinetic_energy_and_the_work_energy_theorem)
In fluid dynamics, surface integrals compute mass flow rates across boundaries, crucial for continuity and conservation principles. The mass flow rate $ \dot{m} $ through a surface $ S $ is $ \dot{m} = \iint_S \rho \mathbf{v} \cdot d\mathbf{A} $, where $ \rho $ is density, $ \mathbf{v} $ is velocity, and $ d\mathbf{A} $ is the outward normal area element; this ensures mass balance in pipes, nozzles, or control volumes. For heat transfer, Fourier's law in integral form calculates total heat flux $ Q = \iint_S \mathbf{q} \cdot d\mathbf{A} $, with $ \mathbf{q} = -k \nabla T $, applying to conduction through walls or fins in engineering designs like heat exchangers.[](https://web.mit.edu/16.unified/www/FALL/fluids/Lectures/f06.pdf)
Engineering applications leverage integrals for structural analysis and signal processing. The center of mass coordinates for a continuous body, such as $ \bar{x} = \frac{1}{M} \int x \, dm $, determine balance points in beams or vehicles, where $ dm = \rho \, dV $ integrates over volume for irregular shapes. In signal processing, the Fourier transform decomposes time-domain signals into frequency components via $ \hat{f}(\omega) = \int_{-\infty}^{\infty} f(t) e^{-i \omega t} \, dt $, enabling filtering and analysis in communications systems or vibration control.[](https://phys.libretexts.org/Workbench/PH_245_Textbook_V2/03%253A_Module_2_-_Multi-Dimensional_Mechanics/3.04%253A_Objective_2.d./3.4.02%253A_Center_of_Mass_%28Part_2%29)[](https://web.stanford.edu/class/ee102/lectures/fourtran)
In electromagnetism, Gauss's law expresses electric flux through a closed surface as a surface integral: $ \oint_S \mathbf{E} \cdot d\mathbf{A} = \frac{Q_{\text{enc}}}{\epsilon_0} $, where $ \mathbf{E} $ is the electric field and $ Q_{\text{enc}} $ is enclosed charge; this simplifies field calculations for symmetric charge distributions like spheres or cylinders in capacitor design.[](https://ocw.mit.edu/courses/8-02-physics-ii-electricity-and-magnetism-spring-2007/b6306257b5c818e4034919c3d8680a70_chapte4gauss_law.pdf)
### In Probability and Statistics
In probability theory, the integral plays a central role in defining the behavior of continuous random variables through their probability density functions (PDFs). For a continuous random variable $X$ with PDF $f(x)$, the probability that $X$ falls within an interval $(a, b)$ is given by $P(a < X < b) = \int_a^b f(x) \, dx$.[](https://stats.libretexts.org/Courses/Saint_Mary%27s_College_Notre_Dame/MATH_345__-_Probability_%28Kuter%29/4%253A_Continuous_Random_Variables/4.1%253A_Probability_Density_Functions_%28PDFs%29_and_Cumulative_Distribution_Functions_%28CDFs%29_for_Continuous_Random_Variables) This function $f(x)$ must satisfy non-negativity, $f(x) \geq 0$ for all $x$, and normalization, ensuring the total probability over the entire support is 1, i.e., $\int_{-\infty}^{\infty} f(x) \, dx = 1$.[](https://online.stat.psu.edu/stat414/lesson/14/14.1) These properties allow the PDF to describe the distribution's shape and density, where the integral over any subinterval quantifies the relative likelihood of outcomes in that range.
Integrals are essential for computing key summary statistics of continuous distributions, particularly moments like expectation and variance. The expected value, or mean $\mu = E[X]$, of a continuous random variable $X$ is defined as $E[X] = \int_{-\infty}^{\infty} x f(x) \, dx$, representing the long-run average value under repeated sampling.[](https://stats.libretexts.org/Courses/Saint_Mary%27s_College_Notre_Dame/DSCI_500B_Essential_Probability_Theory_for_Data_Science_%28Kuter%29/04%253A_Continuous_Random_Variables/4.02%253A_Expected_Value_and_Variance_of_Continuous_Random_Variables) Similarly, the variance $\operatorname{Var}(X)$, which measures the spread around the mean, is $\operatorname{Var}(X) = \int_{-\infty}^{\infty} (x - \mu)^2 f(x) \, dx = E[(X - \mu)^2]$.[](https://ocw.mit.edu/courses/18-05-introduction-to-probability-and-statistics-spring-2022/mit18_05_s22_class06-prep-a.pdf) These integral expressions extend the discrete case by replacing sums with integrals weighted by the density, providing a foundation for understanding central tendency and dispersion in continuous settings.
Moment-generating functions (MGFs) offer a powerful integral-based tool for characterizing entire distributions and deriving their moments. The MGF of a random variable $X$ is defined as $M(t) = E[e^{tX}] = \int_{-\infty}^{\infty} e^{tx} f(x) \, dx$, where the integral converges for $t$ in some neighborhood of 0. By differentiating the MGF and evaluating at $t = 0$, higher-order moments such as mean, variance, and skewness can be obtained systematically, facilitating proofs of properties like the uniqueness of distributions under certain conditions.[](https://ocw.mit.edu/courses/6-436j-fundamentals-of-probability-fall-2018/1a592ed184fb4c444547f67c9bcdd8ec_MIT6_436JF18_lec13.pdf)
In Bayesian statistics, integrals are crucial for updating beliefs via the posterior distribution, which incorporates observed data. The posterior density for parameters $\theta$ given data $x$ is proportional to the likelihood times the prior, $\pi(\theta | x) \propto L(x | \theta) \pi(\theta)$, with the normalizing constant being the marginal likelihood $\int L(x | \theta) \pi(\theta) \, d\theta$.[](https://pages.stat.wisc.edu/~larget/stat302/bayes.pdf) This integral, often challenging to compute directly, ensures the posterior integrates to 1 and enables inference on $\theta$, such as credible intervals or hypothesis testing.[](https://www.stat.cmu.edu/~larry/=sml/Bayes.pdf) For more general probability spaces, these concepts extend to the Lebesgue integral over measures beyond the Riemann sense.[](https://math.arizona.edu/~jwatkins/h-expectedvalue.pdf)
## Computation Methods
### Analytical Techniques
Analytical techniques for computing antiderivatives, also known as indefinite integrals, rely on algebraic manipulations and known differentiation rules to express integrals in closed form. These methods are essential for solving a wide range of integrals that arise in calculus, particularly when the integrand can be transformed into a more recognizable form through substitution, integration by parts, or decomposition. Unlike numerical approximations, these approaches yield exact expressions, often leveraging the inverse nature of integration and differentiation.
One fundamental technique is integration by substitution, which reverses the chain rule of differentiation. To evaluate $\int f(g(x)) g'(x) \, dx$, set $u = g(x)$, so $du = g'(x) \, dx$, transforming the integral into $\int f(u) \, du$. For instance, consider $\int 2x \sin(x^2) \, dx$; substituting $u = x^2$ gives $du = 2x \, dx$, yielding $\int \sin u \, du = -\cos u + C = -\cos(x^2) + C$.[](https://openstax.org/books/calculus-volume-1/pages/5-5-substitution)
Integration by parts provides a method for products of functions, derived from the product rule: $\int u \, dv = uv - \int v \, du$. Choosing $u$ and $dv$ appropriately simplifies the resulting integral; typically, select $u$ to decrease in complexity upon differentiation. As an example, for $\int x e^x \, dx$, let $u = x$ (so $du = dx$) and $dv = e^x \, dx$ (so $v = e^x$), resulting in $x e^x - \int e^x \, dx = x e^x - e^x + C = e^x (x - 1) + C$.[](https://openstax.org/books/calculus-volume-2/pages/3-1-integration-by-parts)
For integrals involving trigonometric functions, identities such as $\sin^2 \theta + \cos^2 \theta = 1$ or double-angle formulas are used to rewrite the integrand into integrable forms, often reducing powers or products. These identities facilitate direct integration or application of other techniques like substitution.[](https://openstax.org/books/calculus-volume-2/pages/3-2-trigonometric-integrals)
Rational functions, which are ratios of polynomials, are integrated by decomposing the integrand into partial fractions after polynomial long division if necessary. For a proper fraction $\frac{P(x)}{Q(x)}$ where $\deg P < \deg Q$, factor $Q(x)$ into linear or irreducible quadratic factors and express $\frac{P(x)}{Q(x)}$ as a sum of simpler fractions, each integrable using logarithms or arctangents. This method ensures the integral reduces to elementary functions.[](https://openstax.org/books/calculus-volume-2/pages/3-4-partial-fractions)
Reduction formulas offer a recursive approach for integrals of powers, such as $\int \sin^n x \, dx$, by relating the integral to one with a lower power. Using integration by parts, the formula is $\int \sin^n x \, dx = -\frac{\sin^{n-1} x \cos x}{n} + \frac{n-1}{n} \int \sin^{n-2} x \, dx$ for $n > 1$. Repeated application eventually yields a solvable form, often involving basic trigonometric integrals. Similar formulas exist for $\cos^n x$ and other powered functions.[](https://openstax.org/books/calculus-volume-2/pages/3-2-trigonometric-integrals)
### Numerical and Symbolic Approaches
Numerical integration methods approximate definite integrals by discretizing the interval of integration and summing weighted function evaluations, often building on the concept of Riemann sums as a foundational limit process. These techniques are essential when analytical antiderivatives are unavailable or impractical to compute.[](https://math.mit.edu/~stevenj/trap-iap-2011.pdf)
The trapezoidal rule provides a basic approximation by treating each subinterval as a trapezoid. For a definite integral $\int_a^b f(x) \, dx$ over $[a, b]$ divided into $n$ equal subintervals of width $h = (b - a)/n$, the formula is
\frac{h}{2} \left[ f(a) + 2 \sum_{i=1}^{n-1} f(a + i h) + f(b) \right].
The global error for this approximation is bounded by $-\frac{(b - a) h^2}{12} f''(\xi)$ for some $\xi \in [a, b]$, assuming $f''$ is continuous, yielding an order of convergence $O(h^2)$.[](http://homepage.math.uiowa.edu/~atkinson/ftp/ENA_Materials/Overheads/sec_5-2.pdf)[](https://mathweb.ucsd.edu/~ebender/20B/77_Trap.pdf)
Simpson's 1/3 rule improves accuracy by fitting a quadratic polynomial over pairs of subintervals. For an even number of subintervals $n = 2m$, with width $h = (b - a)/(2m)$, the approximation is
\frac{h}{3} \left[ f(a) + 4 \sum_{i=1}^{m} f(a + (2i - 1) h) + 2 \sum_{i=1}^{m-1} f(a + 2 i h) + f(b) \right].
The error is bounded by $-\frac{(b - a) h^4}{180} f^{(4)}(\xi)$ for some $\xi \in [a, b]$, assuming $f^{(4)}$ is continuous, resulting in $O(h^4)$ convergence.[](http://numericalmethods.eng.usf.edu/mws/gen/07int/mws_gen_int_txt_simpson13.pdf)[](https://scholarworks.lib.csusb.edu/cgi/viewcontent.cgi?article=2564&context=etd)
Symbolic integration seeks exact antiderivatives in closed form, particularly for elementary functions. The Risch algorithm, developed in the late 1960s, provides a decision procedure to determine whether an elementary antiderivative exists and to compute it if possible, handling expressions involving rational, exponential, logarithmic, and algebraic functions through differential algebra techniques.
Computer algebra systems implement advanced symbolic integration capabilities, often extending the Risch algorithm with heuristics for non-elementary cases. For instance, Wolfram Mathematica's `Integrate` function employs pattern matching, cylindrical algebraic decomposition, and Risch-based methods to compute indefinite and definite integrals symbolically. Similarly, Maplesoft Maple uses a combination of Risch integration and table-based methods for symbolic antiderivatives.
In the 19th century, before widespread digital computation, mechanical integrators like the planimeter enabled area measurement by tracing boundaries. Invented by Jakob Amsler in 1854, the polar planimeter uses a pivoting arm and integrating wheel to compute enclosed areas via a mechanical application of Green's theorem, achieving practical accuracy for engineering drawings.[](http://persweb.wabash.edu/facstaff/footer/papers/foote.sandifer.reprint.pdf)
## Illustrative Examples
### Basic Antiderivative Computations
Basic antiderivative computations rely on fundamental integration rules derived from the inverse of differentiation, allowing the indefinite integral of simple functions to be found directly. These rules include the power rule for polynomials, as well as standard forms for trigonometric, exponential, and logarithmic functions. The linearity property of integration permits the antiderivative of a sum to be the sum of the antiderivatives, facilitating computations for combined terms.[](https://openstax.org/books/calculus-volume-1/pages/5-4-integration-formulas-and-the-net-change-theorem)
For polynomials, the power rule states that the antiderivative of $x^n$ (where $n \neq -1$) is $\frac{x^{n+1}}{n+1} + C$. Applying this to a quadratic example, consider $\int (3x^2 + 2x) \, dx$. First, integrate $3x^2$ to obtain $3 \cdot \frac{x^3}{3} = x^3$, and $2x$ to obtain $2 \cdot \frac{x^2}{2} = x^2$, yielding $x^3 + x^2 + C$.[](https://openstax.org/books/calculus-volume-1/pages/5-4-integration-formulas-and-the-net-change-theorem)
Common non-polynomial antiderivatives include those for trigonometric and exponential functions. The integral of $\sin x$ is $-\cos x + C$, while the integral of $e^x$ is $e^x + C$. For rational functions like reciprocals, $\int \frac{1}{x} \, dx = \ln |x| + C$. These formulas arise as the functions whose derivatives match the integrands.[](https://openstax.org/books/calculus-volume-1/pages/5-4-integration-formulas-and-the-net-change-theorem)
Verification of antiderivatives involves differentiation: if $F'(x) = f(x)$, then $F(x)$ is an antiderivative of $f(x)$. For instance, differentiating $x^3 + x^2 + C$ returns $3x^2 + 2x$, confirming the result; similarly, the derivative of $-\cos x + C$ is $\sin x$, and of $\ln |x| + C$ is $\frac{1}{x}$. This process ensures the correctness of computed antiderivatives.[](https://openstax.org/books/calculus-volume-1/pages/4-10-antiderivatives)
### Applications of the Fundamental Theorem
The Fundamental Theorem of Calculus (FTC) bridges differentiation and integration, enabling practical computations and insights across mathematics and applied fields. Its first part states that if $ f $ is continuous on an interval and $ F(x) = \int_a^x f(t) \, dt $, then $ F'(x) = f(x) $, allowing the derivative of an integral to be found directly. This is applied in differentiating functions defined by integrals, such as in the error function where the derivative of $ \int_0^x e^{-t^2} \, dt $ is $ e^{-x^2} $.[](https://www.math.stonybrook.edu/Videos/MAT131Online/Handouts/Lecture-32-Handout.pdf) Similarly, for more complex limits, the chain rule combines with FTC part 1: the derivative of $ \int_{u(x)}^{v(x)} f(t) \, dt $ is $ f(v(x)) v'(x) - f(u(x)) u'(x) $, as seen in evaluating $ \frac{d}{dx} \int_{\cos x}^{\sin x} \frac{dt}{1 - t^2} = \frac{1}{\cos x} + \frac{1}{\sin x} $.
The second part of the FTC, $ \int_a^b f(x) \, dx = F(b) - F(a) $ where $ F' = f $, facilitates evaluating definite integrals via antiderivatives, representing net change in $ F $. In physics, this computes displacement from velocity: if velocity is $ v(t) = 3t $, displacement from $ t = a $ to $ t = b $ is $ \int_a^b 3t \, dt = \frac{3}{2}(b^2 - a^2) $.[](https://www.whitman.edu/mathematics/calculus_online/section07.02.html) For variable forces, work is $ W = \int_a^b F(x) \, dx $; lifting a leaking bucket with force $ w(x) = 100 - \frac{2}{5}x $ pounds from a 50-foot well yields $ W = 4,500 $ foot-pounds.[](https://mathbooks.unl.edu/Calculus/sec-6-5-physics.html) In fluid mechanics, force on a dam due to water pressure $ P(y) = 62.4 y $ over a region is $ F = \int_0^{20} 62.4 y (60 + \frac{2}{5} y) \, dy = 815,360 $ pounds.[](https://mathbooks.unl.edu/Calculus/sec-6-5-physics.html)
In economics, FTC part 2 aggregates marginal quantities: total revenue from marginal revenue $ MR(x) $ over an interval is $ \int_a^b MR(x) \, dx $, such as $ \int_1^4 400 x^{-1/2} \, dx = 800 $ (hundreds of dollars) over four weeks.[](https://www2.math.uconn.edu/ClassHomePages/Math1071/Textbook/sec_Ch5Sec3.html) For probability, FTC links the cumulative distribution function (CDF) $ F_X(x) = \int_{-\infty}^x f_X(u) \, du $ to the probability density function (PDF) $ f_X(x) $, where differentiation gives $ f_X(x) = F_X'(x) $, enabling probability computations like $ P(c < X \leq d) = \int_c^d f_X(x) \, dx $.[](https://www.stat.uchicago.edu/~stigler/Stat244/ch1withfigs.pdf) These applications underscore the FTC's role in modeling accumulation and rates of change.[](https://sites.und.edu/timothy.prescott/apex/web/apex.Ch5.S4.html)