Fact-checked by Grok 2 weeks ago

Multiple integral

A multiple integral is a definite integral of a function of several real variables over a domain in n-dimensional , generalizing the single-variable definite to higher dimensions such as two or three variables. For instance, double integrals apply to functions of two variables over regions in the plane, while triple integrals extend to three variables over volumes in space. The concept of multiple integrals emerged in the 17th century with methods like indivisibles developed by Bonaventura Cavalieri for computing areas and volumes, evolving through contributions from mathematicians like Cauchy on double integrals in the 19th century. Multiple integrals are fundamental in multivariable calculus, where they enable the computation of accumulated quantities across multidimensional regions, such as the volume beneath a surface defined by z = f(x, y) or the mass of a thin plate with variable density. Defined as limits of Riemann sums over partitions of the domain, they reduce computationally to iterated single integrals via Fubini's theorem, which justifies integrating one variable at a time while treating others as constants. This structure allows evaluation over rectangular or general regions, often requiring adjustments for boundaries and coordinate systems. Beyond basic volumes, multiple integrals find broad applications in physics and , including calculating centers of and moments of for rigid bodies, and in probability for computing expectations and probabilities in multivariate distributions. They also support change-of-variables techniques, such as transformations to polar, cylindrical, or spherical coordinates, to simplify integration over symmetric domains. In , multiple integrals underpin theorems like the and , linking surface and volume integrals to flux and circulation.

Introduction and Overview

Definition and Scope

A multiple generalizes the concept of a single to functions of several variables, allowing the computation of over regions in higher-dimensional spaces \mathbb{R}^n. In essence, it extends the idea of measuring signed areas under curves in one to measuring signed volumes under hypersurfaces in multiple ; for instance, when n=2, it corresponds to the area beneath a surface in the , while for n=3, it yields volumes within regions. This framework is fundamental in for quantifying accumulated quantities, such as mass, probability, or total flux, over multi-dimensional domains. The basic notation for a multiple integral of a function f: \mathbb{R}^n \to \mathbb{R} over a domain D \subseteq \mathbb{R}^n is \int_D f(x_1, \dots, x_n) \, dV, where dV denotes the volume element in \mathbb{R}^n, often expressed as dx_1 \dots dx_n in Cartesian coordinates. This integral represents the limit of Riemann sums approximating the function's values multiplied by the volumes of small subregions partitioning D. Unlike line integrals, which accumulate quantities along one-dimensional curves in space, or surface integrals, which do so over two-dimensional surfaces, multiple integrals specifically evaluate volumes over n-dimensional regions, emphasizing the full interior rather than boundaries. The construction builds on the from one dimension by extending partitions to multi-dimensional grids, where each subregion's volume replaces the interval length in the sum. Multiple integrals inherit key properties from their single-variable counterparts, such as , which will be explored further in subsequent sections.

Historical Development

The development of multiple integrals traces its roots to the foundational work on by and in the late 17th century, where they independently formulated the concepts of and single-variable as tools for analyzing motion and areas under curves. Newton's approach, outlined in his 1669 manuscript De analysi per aequationes numero terminorum infinitas, emphasized fluxions and fluents to compute integrals geometrically, while Leibniz's 1684 publication Nova methodus introduced the ∫ and treated as summation of infinitesimals analytically. These single-variable techniques laid the groundwork for extensions to functions of multiple variables, though Newton and Leibniz themselves did not systematically explore multivariable cases. In the , Leonhard Euler advanced the theory by extending integral calculus to multiple dimensions, particularly through his comprehensive three-volume work Institutionum calculi integralis published between 1768 and 1770. There, Euler systematically investigated double integrals, deriving formulas for integrals over rectangular regions and connecting them to and gamma functions, which he had earlier introduced in 1729. A key contribution appears in his 1768 paper "De formulis integralibus duplicatis," later published in 1770, where he explored the evaluation and properties of double integrals, treating them as iterated single integrals while addressing challenges in changing variables. The 19th century brought rigorous foundations to multiple integrals amid growing concerns over convergence and discontinuities. Augustin-Louis Cauchy, in his 1823 Résumé des leçons sur le calcul infinitésimal, defined the definite integral for continuous functions using partitions and upper/lower sums, a framework that influenced multivariable extensions by emphasizing uniformity in limits. Cauchy also examined double integrals, noting in later works that interchanging integration order could yield inconsistencies unless absolute convergence holds, prompting early notions of absolute integrability. Bernhard Riemann, in his 1854 paper "Über die Darstellbarkeit einer Funktion durch eine trigonometrische Reihe" (published in 1868), generalized the integral to bounded functions with discontinuities via tagged partitions and oscillation criteria, providing a basis for multi-dimensional Riemann integrals that accommodated more complex domains. Toward the century's end, Italian mathematicians Giuseppe Peano and Vito Volterra formalized multiple integrals further; Peano's 1887 Calcolo vettoriale generalized measure concepts to one-, two-, and three-dimensional sets, defining internal and external measures for regions with boundaries and linking them to integrability. Volterra, in his 1881 papers, constructed examples of sets with positive outer content but nowhere dense interiors and discontinuous derivatives that failed Riemann integrability, highlighting limitations and paving the way for advanced theories around 1880–1890. The early 20th century saw revolutionize multiple integration through his 1902 doctoral thesis Intégrale, longueur, aire, which introduced measure theory to handle non-continuous functions in higher dimensions. Lebesgue's approach defined the integral via measurable sets and simple functions, extending naturally to \mathbb{R}^n by product measures, thus resolving issues with Riemann's method for functions discontinuous on sets of positive measure and enabling integration over irregular domains. This framework, building on Riemann's foundations, became the standard for modern analysis of multiple integrals.

Mathematical Definition

General Formulation

In the context of , a multiple provides a means to compute the of a over a multi-dimensional domain, extending the one-dimensional Riemann integral. For a f: D \subset \mathbb{R}^n \to \mathbb{R}, where D is a bounded domain in n-dimensional Euclidean space, the multiple is formally defined as the limit of Riemann sums over partitions of D. Specifically, consider a partition P of D into finitely many subregions D_i with volumes \Delta V_i, and choose sample points \xi_i \in D_i. The Riemann sum is \sum_i f(\xi_i) \Delta V_i, and the is \int_D f \, dV = \lim_{\|P\| \to 0} \sum_i f(\xi_i) \Delta V_i, where \|P\| denotes the mesh of the partition, typically the maximum diameter of the subregions D_i, and the limit exists if the upper and lower Riemann sums converge to the same value. For the integral to be well-defined in the Riemann sense, f must be bounded on the compact set D, and the upper sum U(f, P) = \sum_i M_i \Delta V_i (where M_i = \sup_{D_i} f) and lower sum L(f, P) = \sum_i m_i \Delta V_i (where m_i = \inf_{D_i} f) must satisfy \inf_P U(f, P) = \sup_P L(f, P) as the partition is refined. A sufficient condition for Riemann integrability is that f is continuous on D, or more generally, discontinuous only on a set of n-dimensional zero. This ensures the is independent of the choice of sample points and partitions. In the n-dimensional case, the integral is commonly notated as \int_D f(x_1, \dots, x_n) \, dV, where dV represents the volume element, or more abstractly as \int_{\mathbb{R}^n} f \, d\mu with \mu denoting the on \mathbb{R}^n, which coincides with the for Riemann-integrable functions but extends to a broader class via measure-theoretic construction. The Lebesgue approach assigns a measure to subsets of \mathbb{R}^n based on coverings by rectangles, providing a rigorous foundation for over non-rectifiable domains, though the Riemann definition remains central for continuous functions on compact sets. For unbounded domains D \subset \mathbb{R}^n, the multiple integral is defined as an by taking the over an increasing of compact exhaustion sets D_k \uparrow D, such that \int_D f \, dV = \lim_{k \to \infty} \int_{D_k} f \, dV, provided the exists and is finite. This extension allows over regions like \mathbb{R}^n itself, but requires additional convergence conditions on f to ensure the is well-behaved.

Properties of Multiple Integrals

Multiple integrals, defined as limits of Riemann sums over partitions of a bounded domain D \subset \mathbb{R}^n, inherit several fundamental algebraic and analytic analogous to those of the single-variable . These properties hold for Riemann-integrable functions, typically continuous functions on compact Jordan-measurable sets, and are derived directly from the definition via sums over subdomains. One key property is . For Riemann-integrable functions f, g: D \to \mathbb{R} and constants a, b \in \mathbb{R}, the multiple integral satisfies \int_D (a f + b g) \, dV = a \int_D f \, dV + b \int_D g \, dV, where dV denotes the volume element. This follows from the linearity of the Riemann sum approximation: if \sigma is a of D into subdomains R_k with tags \mathbf{x}_k \in R_k and \Delta V_k = \mathrm{vol}(R_k), then the for a f + b g is \sum_k (a f(\mathbf{x}_k) + b g(\mathbf{x}_k)) \Delta V_k = a \sum_k f(\mathbf{x}_k) \Delta V_k + b \sum_k g(\mathbf{x}_k) \Delta V_k. Taking the limit as the mesh of the partition approaches zero preserves this relation, yielding the integral equality. Monotonicity is another essential property: if f \leq g on D, where f and g are Riemann-integrable, then \int_D f \, [dV](/page/DV) \leq \int_D g \, [dV](/page/DV). To see this, consider the non-negative function h = g - f \geq 0. By , \int_D g \, dV = \int_D f \, dV + \int_D h \, dV, so it suffices to show \int_D h \, dV \geq 0 for non-negative integrable h. In the Riemann , each h(\mathbf{x}_k) \Delta V_k \geq 0, so the sum is non-negative, and thus the limit (integral) is non-negative. This monotonicity extends to the case where $0 \leq f \leq g. Additivity over disjoint domains also holds: if D = D_1 \cup D_2 with D_1 \cap D_2 = \emptyset and f is Riemann-integrable on D, then \int_D f \, dV = \int_{D_1} f \, dV + \int_{D_2} f \, dV. This property arises from refining partitions of D to separately partition D_1 and D_2; the Riemann sums decompose additively over the subdomains, and the limit respects this decomposition since the integrals over each exist. For more general disjoint unions of Jordan-measurable sets, the property extends by finite additivity of the content (Jordan measure). For continuous functions on a compact D, the for multiple integrals states that if m \leq f(\mathbf{x}) \leq M for all \mathbf{x} \in D, then there exists some c \in [m, M] such that \int_D f \, dV = c \cdot \mathrm{vol}(D). More precisely, for continuous f on a connected compact Jordan-measurable D, there exists \mathbf{c} \in D with \int_D f \, dV = f(\mathbf{c}) \cdot \mathrm{vol}(D). This follows from the applied to the continuous g(t) = \int_D (f - t) \, dV on [m, M], which changes sign or zero at some point, combined with monotonicity to ensure the integral scales with . Basic inequalities derive from these properties, such as the \left| \int_D f \, dV \right| \leq \int_D |f| \, dV for Riemann-integrable f. To prove this, note that |f| \geq f and |f| \geq -f, so by monotonicity, \int_D |f| \, dV \geq \int_D f \, dV and \int_D |f| \, dV \geq -\int_D f \, dV = \int_D (-f) \, dV. Adding these yields $2 \int_D |f| \, dV \geq \int_D f \, dV + \int_D (-f) \, dV = 0, but more directly, \left| \int_D f \, dV \right| = \int_D |f| \, dV if f \geq 0 or by for general f = f^+ - f^- with f^\pm \geq 0, though the holds via the above bounds.

Particular Cases

Double Integrals

A double is a type of multiple that extends the concept of a single to functions of two variables over a in the . It is denoted by \iint_D f(x,y) \, dA, where f(x,y) is the integrand, D is a bounded in \mathbb{R}^2, and dA = dx \, dy represents the area element. This represents the signed volume under the surface z = f(x,y) over the domain D. The domain D for a double is typically a bounded of \mathbb{R}^2, often classified into Type I and Type II regions to facilitate via iterated integrals. A Type I region is described as D = \{(x,y) \mid a \leq x \leq b, \, g(x) \leq y \leq h(x)\}, where g(x) and h(x) are continuous functions with g(x) \leq h(x) for a \leq x \leq b, forming vertical strips bounded by the curves y = g(x) and y = h(x). Similarly, a Type II region is D = \{(x,y) \mid c \leq y \leq d, \, g(y) \leq x \leq h(y)\}, where g(y) and h(y) are continuous with g(y) \leq h(y) for c \leq y \leq d, consisting of horizontal strips. These classifications allow the double to be evaluated as an in the appropriate order. The double integral is defined as the limit of over of the region D. For a of D into subregions with areas \Delta A_{ij}, the is \sum_i \sum_j f(x_i^*, y_j^*) \Delta A_{ij}, where (x_i^*, y_j^*) is a sample point in the ij-th subregion; for rectangular , \Delta A_{ij} = \Delta x_i \Delta y_j. As the mesh of the approaches zero, this sum converges to \iint_D f(x,y) \, dA. Double integrals inherit properties such as from the general theory of multiple integrals. A f(x,y) is Riemann integrable over a closed and bounded region D if it is bounded and on D, except possibly on a set of measure zero; on such a D ensures integrability. This condition guarantees that the limit of the Riemann sums exists and is independent of the choice of partitions and sample points.

Triple Integrals

A triple integral extends the concept of multiple integration to three dimensions, providing a means to compute the of a over a bounded E \subset \mathbb{R}^3. It is denoted by \iiint_E f(x,y,z) \, [dV](/page/DV), where f: E \to \mathbb{R} is the integrand and dV represents the volume element, typically expressed in Cartesian coordinates as dx \, dy \, dz. This quantifies quantities such as the total mass of a with variable f(x,y,z), or the average value of f over E when normalized by of E. The triple integral is rigorously defined as the limit of a triple over a of E. Consider a of E into sub-boxes with volumes \Delta V_{ijk} = \Delta x_i \Delta y_j \Delta z_k, where sample points (\xi_i, \eta_j, \zeta_k) are chosen within each sub-box. The is then \sum_i \sum_j \sum_k f(\xi_i, \eta_j, \zeta_k) \Delta V_{ijk}, and the triple integral is \iiint_E f(x,y,z) \, dV = \lim \sum_i \sum_j \sum_k f(\xi_i, \eta_j, \zeta_k) \Delta V_{ijk} as the mesh of the partition approaches zero. For the integral to exist, f must be bounded and continuous on the compact set E, ensuring of the Riemann sums. To evaluate triple integrals, the solid E is often described by its projections onto the coordinate planes, leading to iterated integrals with specific orders of . For a Type I region, E is defined such that a \leq x \leq b, g(x) \leq y \leq h(x), and u(x,y) \leq z \leq v(x,y), where g, h, u, and v are continuous functions; the integral becomes \int_a^b \int_{g(x)}^{h(x)} \int_{u(x,y)}^{v(x,y)} f(x,y,z) \, dz \, dy \, dx. Similar Type II and Type III descriptions project onto the yz- and xz-planes, respectively, by cycling the variables, allowing flexibility in setup based on the 's . The volume element dV = dx \, dy \, dz facilitates this , with integrability holding under the same assumptions on f and the bounding surfaces of E. Building on integrals over planar regions, these descriptions add a third layer of variable limits dependent on the prior coordinates.

Higher-Dimensional Integrals

The multiple integral in n dimensions, for n \geq 4, generalizes the formulation to integrals over subsets D \subset \mathbb{R}^n and is denoted as \int_D f(x_1, \dots, x_n) \, dx_1 \cdots dx_n, where f: \mathbb{R}^n \to \mathbb{R} is an integrable function and the measure is the standard on \mathbb{R}^n. This notation extends the product structure of lower-dimensional cases, treating the integral as an n-fold over the coordinates. A primary challenge in higher-dimensional integration arises from the curse of dimensionality, where the volume of the unit cube in \mathbb{R}^n remains 1, but the complexity of partitioning the domain into manageable elements grows exponentially with n, rendering standard grid-based methods computationally intractable for n > 3. For instance, approximations require resources scaling as O(m^n) for grid size m, quickly exceeding practical limits even for moderate n around 10. This intractability extends to both analytical evaluation and numerical approximation, necessitating specialized techniques beyond simple iteration. To address limitations of Riemann integrals in higher dimensions, the Lebesgue integral provides a robust extension by defining it over the \sigma-algebra of Lebesgue measurable sets in \mathbb{R}^n, generated as the completion of the Borel \sigma-algebra with respect to the Lebesgue . A function f is measurable if the preimage of Borel sets under f belongs to this \sigma-algebra, enabling the integral to handle a broader class of functions, including those with singularities or on irregular domains, while preserving properties like countable additivity. An important abstract property of the higher-dimensional is its invariance under : for any Q \in O(n), the integral satisfies \int_{\mathbb{R}^n} f(Qx) \, dx = \int_{\mathbb{R}^n} f(x) \, dx for integrable f, reflecting the rotation invariance of the itself. This property ensures that the integral remains unchanged under rigid rotations of the domain, facilitating applications in symmetric problems across \mathbb{R}^n.

Computation Techniques

Iterated Integrals

Iterated integrals provide a practical for computing multiple integrals by reducing them to successive single-variable integrations. For a f(x,y) defined over a D in the that can be described as a Type I region, where D = \{(x,y) \mid a \leq x \leq b, g(x) \leq y \leq h(x)\} with g and h continuous, the is defined as \int_a^b \left( \int_{g(x)}^{h(x)} f(x,y) \, dy \right) dx. This inner integral treats x as fixed and integrates with respect to y, yielding a function of x that is then integrated over the x-interval. Similarly, for a Type II region D = \{(x,y) \mid c \leq y \leq d, p(y) \leq x \leq q(y)\}, the takes the form \int_c^d \left( \int_{p(y)}^{q(y)} f(x,y) \, dx \right) dy. These definitions extend naturally to higher dimensions, such as triple integrals over regions described by nested bounds. For rectangular domains, such as R = [a,b] \times [c,d], the offers flexibility, allowing evaluation as either \int_a^b \int_c^d f(x,y) \, dy \, dx or \int_c^d \int_a^b f(x,y) \, dx \, dy. This interchange simplifies computation when one order yields an easier . The equivalence of these iterated integrals to the multiple integral over the domain is justified by Fubini's theorem. Fubini's theorem states that if f(x,y) is continuous on a compact rectangular R = [a,b] \times [c,d], then the double integral equals both : \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 holds more generally for continuous f on a compact bounded D \subset \mathbb{R}^2, where the multiple integral \iint_D f \, dA equals the appropriate over a description of D. The theorem extends to higher dimensions analogously. A of the proof for the rectangular case relies on the of f on the compact set R, which follows from its . R into a of small rectangles with side lengths \Delta x = (b-a)/m and \Delta y = (d-c)/n. The for the double is \sum_{i=1}^m \sum_{j=1}^n f(x_i^*, y_j^*) \Delta x \Delta y, which can be grouped as a telescoping double : first summing over j for fixed i approximates the inner integral with respect to y, then over i for the outer with respect to x. ensures that the choice of sample points (x_i^*, y_j^*) introduces an error bounded by \epsilon \cdot (b-a)(d-c) for sufficiently fine partitions, making the iterated sums converge to the same limit as the double . This argument generalizes to non-rectangular compact regions by extending f continuously to a containing .

Change of Variables

In multiple integrals, the theorem facilitates the substitution of new coordinates via a , invertible , accounting for the distortion of volumes through the determinant. For an n-dimensional over a D \subset \mathbb{R}^n and a g: U \to D where U \subset \mathbb{R}^n, the formula states that \int_D f(\mathbf{x}) \, d\mathbf{x} = \int_U f(g(\mathbf{u})) \left| \det J_g(\mathbf{u}) \right| \, d\mathbf{u}, with J_g denoting the matrix of partial derivatives of g. This holds under conditions of continuity and invertibility, ensuring the transformation preserves the integral's value up to the scaling factor provided by the absolute . The matrix J_g(\mathbf{u}) is the n \times n with entries (J_g)_{ij} = \partial g_i / \partial u_j, and its measures the local volume scaling. In two dimensions, for coordinates x = x(u,v), y = y(u,v), the simplifies to \left| \frac{\partial(x,y)}{\partial(u,v)} \right| = \left| \frac{\partial x}{\partial u} \frac{\partial y}{\partial v} - \frac{\partial x}{\partial v} \frac{\partial y}{\partial u} \right|, which adjusts the area element dA = dx \, dy to \left| \frac{\partial(x,y)}{\partial(u,v)} \right| du \, dv in the transformed variables. This determinant arises from the linear approximation of the transformation near each point, analogous to the single-variable case where dx = |g'(u)| du. A standard application is the polar coordinate transformation in the plane, defined by x = r \cos \theta, y = r \sin \theta, with parameters r \geq 0 and \theta \in [0, 2\pi). To derive the Jacobian, compute the partial derivatives: \frac{\partial x}{\partial r} = \cos \theta, \quad \frac{\partial x}{\partial \theta} = -r \sin \theta, \quad \frac{\partial y}{\partial r} = \sin \theta, \quad \frac{\partial y}{\partial \theta} = r \cos \theta. The determinant is then \frac{\partial(x,y)}{\partial(r,\theta)} = \begin{vmatrix} \cos \theta & -r \sin \theta \\ \sin \theta & r \cos \theta \end{vmatrix} = (\cos \theta)(r \cos \theta) - (-r \sin \theta)(\sin \theta) = r \cos^2 \theta + r \sin^2 \theta = r, since \cos^2 \theta + \sin^2 \theta = 1 and r \geq 0. Thus, the area element becomes dA = r \, dr \, d\theta. For integration over a disk of radius R, the bounds are $0 \leq r \leq R and $0 \leq \theta \leq 2\pi, simplifying computations for circularly symmetric regions. In three dimensions, cylindrical coordinates extend polar coordinates by including height: x = r \cos \theta, y = r \sin \theta, z = z, with r \geq 0, \theta \in [0, 2\pi), and z varying according to the domain (e.g., a \leq z \leq b). The Jacobian matrix is J = \begin{pmatrix} \cos \theta & -r \sin \theta & 0 \\ \sin \theta & r \cos \theta & 0 \\ 0 & 0 & 1 \end{pmatrix}, and its determinant is r, yielding the volume element dV = r \, dr \, d\theta \, dz. This transformation is useful for regions with cylindrical symmetry, such as infinite cylinders or finite heights. Spherical coordinates provide another common substitution: x = \rho \sin \phi \cos \theta, y = \rho \sin \phi \sin \theta, z = \rho \cos \phi, where \rho \geq 0, \theta \in [0, 2\pi), and \phi \in [0, \pi]. The Jacobian matrix is J = \begin{pmatrix} \sin \phi \cos \theta & -\rho \sin \phi \sin \theta & \rho \cos \phi \cos \theta \\ \sin \phi \sin \theta & \rho \sin \phi \cos \theta & \rho \cos \phi \sin \theta \\ \cos \phi & 0 & -\rho \sin \phi \end{pmatrix}, with determinant \rho^2 \sin \phi (computed by cofactor expansion along the second row or verified via standard results). Since \sin \phi \geq 0 for \phi \in [0, \pi], the volume element is dV = \rho^2 [\sin \phi](/page/Sin) \, d\rho \, d\theta \, d\phi. For a ball of radius R, the bounds are $0 \leq \rho \leq R, $0 \leq \theta \leq 2\pi, $0 \leq \phi \leq \pi, aiding integrals over spherical domains.

Symmetry and Reduction Methods

Symmetry methods provide powerful tools for simplifying the evaluation of multiple integrals by exploiting inherent properties of the integrand or the , often reducing the computational burden without requiring complete or . These techniques are particularly effective when the possesses reflectional, rotational, or other symmetries that align with the of the being integrated. In multiple variables, a f(\mathbf{x}) is considered odd with to one , say x_i, if f(-x_i, x_1, \dots, x_{i-1}, x_{i+1}, \dots, x_n) = -f(x_1, \dots, x_n) for all points in the . If the D is symmetric with respect to the where x_i = [0](/page/0) (i.e., if \mathbf{x} \in D implies that the point with x_i replaced by -x_i is also in D), then the multiple integral \int_D f(\mathbf{x}) \, d\mathbf{x} = [0](/page/0). This result follows from pairing points symmetric across the , where the contributions cancel due to the oddness in x_i. The property extends analogously if the is odd in multiple variables and the is symmetric in those directions. For instance, in double integrals over a region symmetric about the y-axis, if f(x,y) is odd in x, the integral vanishes. Radial symmetry arises when the integrand depends only on the radial distance r = \sqrt{x_1^2 + \dots + x_n^2} from the origin, and the domain is a ball or disk centered at the origin. For a double integral over the unit disk in the plane, if f(x,y) = g(\sqrt{x^2 + y^2}), the integral simplifies to \iint_{x^2 + y^2 \leq 1} g(\sqrt{x^2 + y^2}) \, dx \, dy = 2\pi \int_0^1 g(r) r \, dr, due to the rotational invariance around the origin, which distributes the area element uniformly in angular directions. This reduction leverages the full $2\pi rotational symmetry to convert the two-dimensional integral into a one-dimensional radial integral, weighted by the Jacobian factor r. Similar reductions apply in higher dimensions, such as for triple integrals over a ball, yielding factors involving the surface area of the unit sphere. Reduction formulas for multiple integrals often employ or differentiation under the integral sign to express a given in terms of simpler ones, particularly for families or functions. For example, consider a I(a) = \iint_D f(x,y,a) \, dx \, dy; differentiating with respect to the parameter a yields I'(a) = \iint_D \frac{\partial}{\partial a} f(x,y,a) \, dx \, dy, assuming suitable continuity and differentiability conditions to justify interchanging derivative and . Integrating back and applying boundary conditions can reduce I(a) to lower-order terms. in one variable, treating others as fixed, similarly reduces powers: for \iint_D x^m y^n \, dx \, dy over a rectangular domain, parts on x^m yields a recurrence relating it to integrals with m-2. These methods are especially useful for evaluating moments or polynomial integrals without full expansion. Domain decomposition exploits symmetry by partitioning an irregular or complex into congruent symmetric subregions, computing the over one subregion, and multiplying by the number of subregions. For a invariant under a group of symmetries (e.g., reflections across coordinate planes), the over the full equals the number of symmetric copies times the over a fundamental . In triple integrals over a of radius R, the decomposes into 8 octants symmetric under sign changes in x, y, z; thus, the volume is $8 times the over the positive octant x \geq 0, y \geq 0, z \geq 0. This approach minimizes setup and computation while preserving accuracy, provided the integrand respects the symmetries or averages appropriately across subregions.

Examples and Illustrations

Rectangular Domains

In rectangular domains, multiple integrals are evaluated over regions that are Cartesian products of intervals, such as a R = [a, b] \times [c, d] in the or a in three dimensions. For a \iint_R f(x,y) \, [dA](/page/DA), where f is continuous on R, the is computed using iterated integrals in the form \int_a^b \int_c^d f(x,y) \, dy \, dx. This approach treats the inner with respect to y as a of the fixed outer x, followed by with respect to x. A concrete example is the double integral \iint_{[0,1] \times [0,1]} (x^2 + y^2) \, dA. Compute it as the \int_0^1 \int_0^1 (x^2 + y^2) \, dy \, dx: \begin{align*} \int_0^1 (x^2 + y^2) \, dy &= \left[ x^2 y + \frac{y^3}{3} \right]_{y=0}^{y=1} = x^2 \cdot 1 + \frac{1}{3} - 0 = x^2 + \frac{1}{3}, \\ \int_0^1 \left( x^2 + \frac{1}{3} \right) \, dx &= \left[ \frac{x^3}{3} + \frac{x}{3} \right]_{x=0}^{x=1} = \frac{1}{3} + \frac{1}{3} = \frac{2}{3}. \end{align*} Thus, \iint_{[0,1] \times [0,1]} (x^2 + y^2) \, dA = \frac{2}{3}. This step-by-step iteration leverages the fixed limits of the rectangular domain, simplifying the evaluation. For triple integrals over a rectangular B = [a, b] \times [c, d] \times [e, f], the \iiint_B f(x,y,z) \, dV is expressed as the \int_a^b \int_c^d \int_e^f f(x,y,z) \, dz \, dy \, dx, assuming f is continuous on B. The innermost treats x and y as constants, followed by successive integrations over the remaining variables with their fixed limits. Over rectangular domains, the is independent, meaning \int_a^b \int_c^d f(x,y) \, dy \, dx = \int_c^d \int_a^b f(x,y) \, dx \, dy for continuous f, allowing flexibility in choosing the sequence that simplifies computation. This property extends to higher dimensions for , where any of the six possible orders yields the same result.

Non-Rectangular Domains

In non-rectangular domains, the boundaries of the integration region are described using inequalities that depend on the variables, often requiring the region to be classified as Type I (vertical strips) or Type II (horizontal strips) for double integrals, or analogous descriptions for higher dimensions. This allows the multiple integral to be expressed as an iterated integral with variable limits, ensuring the integration covers exactly the desired area or volume without overlap or omission. Accurate boundary description is crucial, as it determines the limits of integration and directly affects the result. Consider the double integral \iint_D (x + y) \, dA over the triangular region D defined by $0 \leq x \leq 1 and $0 \leq y \leq x, which lies below the line y = x in the first quadrant. As a Type I region, the iterated integral is set up as \int_0^1 \int_0^x (x + y) \, dy \, dx. Evaluating the inner integral first gives \int_0^x (x + y) \, dy = \left[ x y + \frac{y^2}{2} \right]_0^x = x^2 + \frac{x^2}{2} = \frac{3}{2} x^2. The outer integral then yields \int_0^1 \frac{3}{2} x^2 \, dx = \frac{3}{2} \cdot \frac{1}{3} = \frac{1}{2}. The can be switched for the same region D, now viewed as a Type II region with $0 \leq y \leq 1 and y \leq x \leq 1. The becomes \int_0^1 \int_y^1 (x + y) \, dx \, dy, which evaluates to the same value of \frac{1}{2}, confirming consistency via Fubini's theorem for continuous functions over bounded regions. For a triple integral example, consider the volume of the solid under the paraboloid surface z = x^2 + y^2 and above the unit disk D: x^2 + y^2 \leq 1 in the xy-plane. This volume is given by the triple integral \iiint_E 1 \, dV, where E is the region (x,y,z) with (x,y) \in D and $0 \leq z \leq x^2 + y^2 . As an iterated integral in Cartesian coordinates, it would require describing D with variable limits, such as integrating over y from -\sqrt{1-x^2} to \sqrt{1-x^2} and x from -1 to $1, then z from 0 to x^2 + y^2. However, the curved boundary of the disk suggests a setup in polar coordinates for the base, yielding \int_0^{2\pi} \int_0^1 \int_0^{r^2} r \, dz \, dr \, d\theta = \int_0^{2\pi} \int_0^1 r^3 \, dr \, d\theta, focusing on the boundary description where the radial limit r \leq 1 and angular sweep $0 \leq \theta \leq 2\pi capture the disk precisely.

Volume and Surface Computations

Multiple integrals provide a powerful for computing geometric quantities such as and surface areas in higher dimensions. The beneath a surface defined by z = f(x,y) \geq 0 over a D in the xy- is given by the double V = \iint_D f(x,y) \, dA./03%3A_Multiple_Integrals/3.01%3A_Double_Integrals) A representative example is the under the upper z = \sqrt{4 - x^2 - y^2} over the disk D: x^2 + y^2 \leq 4. Using polar coordinates, where x = r \cos [\theta](/page/Theta), y = r \sin [\theta](/page/Theta), and dA = r \, dr \, d\theta, the becomes V = \int_0^{2\pi} \int_0^2 \sqrt{4 - r^2} \, r \, dr \, d\theta. The inner evaluates to \int_0^2 \sqrt{4 - r^2} \, r \, dr = \frac{8}{3} via the u = 4 - r^2, yielding V = 2\pi \cdot \frac{8}{3} = \frac{16\pi}{3}. For the volume of a solid region E in three dimensions, the triple integral \iiint_E 1 \, dV computes the enclosed volume. Consider the tetrahedron with vertices at (0,0,0), (1,0,0), (0,1,0), and (0,0,1), bounded by the planes x=0, y=0, z=0, and x + y + z = 1. Using iterated integrals over the projection D in the xy-plane (the triangle $0 \leq x \leq 1, $0 \leq y \leq 1 - x), the volume is V = \int_0^1 \int_0^{1-x} \int_0^{1-x-y} dz \, dy \, dx = \int_0^1 \int_0^{1-x} (1 - x - y) \, dy \, dx = \frac{1}{6}. This result matches the known formula for the volume of a tetrahedron with those vertices./15%3A_Multiple_Integration/15.02%3A_Double_Integrals_over_General_Regions) Multiple integrals also determine centroids and moments of solids. For a solid E with volume V, the centroid coordinates are \bar{x} = \frac{1}{V} \iiint_E x \, dV, \bar{y} = \frac{1}{V} \iiint_E y \, dV, and \bar{z} = \frac{1}{V} \iiint_E z \, dV. For the tetrahedron above, symmetry implies \bar{x} = \bar{y} = \bar{z}, and evaluating \bar{x} = \frac{1}{V} \int_0^1 \int_0^{1-x} \int_0^{1-x-y} x \, dz \, dy \, dx = \frac{1}{4}, so the centroid is at (\left( \frac{1}{4}, \frac{1}{4}, \frac{1}{4} \right)./15%3A_Multiple_Integration/15.06%3A_Calculating_Centers_of_Mass_and_Moments_of_Inertia)/15%3A_Multiple_Integration/15.02%3A_Double_Integrals_over_General_Regions) Surface area computations use double integrals for the graph of z = f(x,y) over D, given by \iint_D \sqrt{1 + f_x^2 + f_y^2} \, dA, which accounts for the metric induced by the surface parameterization. This formula arises from the arc length analogy extended to surfaces./13%3A_Multiple_Integration/13.05%3A_Surface_Area)

Advanced Concepts

Improper Multiple Integrals

Improper multiple integrals arise when the domain of integration is unbounded or the integrand has singularities within the domain, extending the concept of proper multiple integrals by taking limits of integrals over bounded approximating regions. Formally, for an unbounded domain D \subset \mathbb{R}^n and a function f: D \to \mathbb{R}, the improper integral is defined as \int_D f \, dV = \lim_{k \to \infty} \int_{D_k} f \, dV, where \{D_k\} is an increasing sequence of bounded domains such that D_k \uparrow D. Similarly, for singularities, the integral excludes a neighborhood of the discontinuity and takes the limit as the neighborhood shrinks to the point. The integral converges if this limit exists and is finite; otherwise, it diverges. A key property is : the improper multiple integral \int_D f \, dV converges absolutely if \int_D |f| \, dV < \infty. Absolute convergence implies ordinary and ensures the value is independent of the choice of approximating sequence \{D_k\}, provided the sequence exhausts D properly. This criterion is particularly useful for testing , as it reduces the problem to the positive integrand |f|, where tests or other methods from single-variable can often be applied after suitable transformations. A classic example is the Gaussian integral over the plane, \iint_{\mathbb{R}^2} e^{-(x^2 + y^2)} \, dA = \pi, which converges due to the rapid decay of the integrand at infinity. This improper double integral can be evaluated using iterated limits: \left( \int_{-\infty}^{\infty} e^{-x^2} \, dx \right)^2 = \pi, where each one-dimensional integral equals \sqrt{\pi}. The rotation invariance of the integrand suggests using polar coordinates, transforming the integral to \int_0^{2\pi} \int_0^{\infty} e^{-r^2} r \, dr \, d\theta = 2\pi \cdot \frac{1}{2} = \pi, confirming the result and illustrating how symmetry aids computation over unbounded domains. In higher dimensions, improper multiple integrals exhibit a difference from their one-dimensional counterparts: conditional convergence does not occur. If the integral converges, it must converge absolutely; there are no cases where \int_D f \, dV exists finitely while \int_D |f| \, dV = \infty. This stems from the nature of the exhaustion process in multiple variables, where the positive and negative contributions cannot cancel in a way that allows without absolute integrability, avoiding pathologies like those in or one-dimensional oscillatory integrals.

Fubini's Theorem and Iterated vs. Multiple Integrals

Fubini's theorem establishes the equivalence between multiple integrals and iterated integrals under suitable conditions in measure theory. The multiple integral of a measurable function f: X \times Y \to \mathbb{R} is defined intrinsically as the Lebesgue integral with respect to the product measure \mu \times \nu on the product \sigma-algebra of \sigma-finite measure spaces (X, \mathcal{A}, \mu) and (Y, \mathcal{B}, \nu). In contrast, the iterated integral computes f by successive integration: first with respect to one variable, then the other, serving as a practical computational tool rather than a fundamental definition. Fubini's theorem asserts that if f is \mu \times \nu-integrable, meaning \int_{X \times Y} |f| \, d(\mu \times \nu) < \infty, then the iterated integrals exist, are finite and equal to each other, and coincide with the multiple integral: \int_{X \times Y} f \, d(\mu \times \nu) = \int_X \left( \int_Y f(x,y) \, d\nu(y) \right) d\mu(x) = \int_Y \left( \int_X f(x,y) \, d\mu(x) \right) d\nu(y). This holds for \sigma-finite measures, ensuring the product measure is well-defined and the slicing functions are measurable. The absolute integrability condition is crucial, as it implies the inner integrals are well-behaved. For non-negative measurable functions f \geq 0, Tonelli's theorem relaxes the integrability requirement: the iterated integrals always exist (possibly infinite) and equal the multiple integral, regardless of whether \int |f| < \infty. This follows from the applied to simple functions approximating f, allowing interchange without . Tonelli's result is particularly useful for positive integrands in applications like probability, where expectations of non-negative random variables can be computed iteratively. When the absolute integrability condition fails, the iterated integrals may differ, one may diverge while the other converges, or both may fail to match the multiple integral (which may not exist). A standard counterexample is f(x,y) = \frac{x - y}{(x + y)^3} over the unit square [0,1] \times [0,1] with , where \int_0^1 \left( \int_0^1 f(x,y) \, dy \right) dx = 0 but \int_0^1 \left( \int_0^1 f(x,y) \, dx \right) dy = +\infty, showing f is not Lebesgue integrable since \int \int |f| \, dy \, dx = \infty. This discontinuity along x = y causes the pathology, violating Fubini's hypothesis. In the context of Riemann integrals, Fubini's theorem requires stricter conditions, such as continuity of f on a compact rectangle, to ensure equivalence; otherwise, counterexamples exist where Riemann iterated integrals differ despite the function being bounded. The Lebesgue framework succeeds more broadly for integrable functions, as measurability and absolute integrability suffice, encompassing cases where Riemann integration fails due to discontinuities of positive measure. For instance, Lebesgue integration allows Fubini to apply to characteristic functions of non-rectifiable sets, where Riemann definitions break down.

Applications

In Geometry and Physics

In geometry, multiple integrals are essential for computing moments of inertia of solid bodies, which quantify resistance to rotational acceleration about a specified . For a solid body with density function \rho(x, y, z), the moment of inertia I_z about the z-axis is given by the triple integral I_z = \iiint_V (x^2 + y^2) \rho(x, y, z) \, dV, where V is of the body; this formula arises from integrating the squared from the axis over the mass distribution. A classic example is the moment of inertia of a uniform disk of radius R and M about its central axis perpendicular to the plane. Using polar coordinates for the double integral over the disk's area, with constant \rho = M / (\pi R^2), the yields I_z = \iint_D r^2 \rho \, r \, dr \, d\theta = \rho \int_0^{2\pi} d\theta \int_0^R r^3 \, dr = 2\pi \rho \cdot \frac{R^4}{4} = \frac{1}{2} M R^2, demonstrating how multiple integrals simplify such geometric calculations. In physics, multiple integrals determine key quantities like the center of mass for extended objects under gravitational or other fields. The position vector of the center of mass \bar{\mathbf{r}} for a body of total mass M = \iiint_V \rho \, dV is \bar{\mathbf{r}} = \frac{1}{M} \iiint_V \mathbf{r} \, \rho(\mathbf{r}) \, dV, where \mathbf{r} = (x, y, z) is the position vector; the components follow similarly as \bar{x} = (1/M) \iiint_V x \rho \, dV, and so on. Another fundamental application is the electrostatic potential \phi(\mathbf{r}) at a point \mathbf{r} due to a volume charge distribution \rho(\mathbf{r}'), expressed as the triple integral \phi(\mathbf{r}) = \frac{1}{4\pi \epsilon_0} \iiint_{V'} \frac{\rho(\mathbf{r}')}{|\mathbf{r} - \mathbf{r}'|} \, dV', which integrates the contributions from infinitesimal charge elements across the volume V'; this formula underpins the superposition principle in electrostatics. Multiple integrals also play a central role in , particularly for flux calculations like s through surfaces. The through a surface S is the \iint_S \rho \mathbf{v} \cdot \mathbf{n} \, dS, where \rho is fluid density, \mathbf{v} is , and \mathbf{n} is the outward ; this represents the net rate of mass transport across S. By the , this equals the volume integral \iiint_V \nabla \cdot (\rho \mathbf{v}) \, dV over the enclosed volume V, linking local of the to global flow behavior and deriving the for mass conservation in fluids. In the context of work done by vector fields, multiple integrals exploit linearity properties to compute total work via superposition of contributions over regions. For conservative , such as gravitational or electrostatic forces, the work along a path decomposes into volume integrals of field potentials, allowing efficient evaluation without ; this linearity ensures that the total work is the of integrals over disjoint subregions or charge distributions.

In Probability and Statistics

In probability theory, multiple integrals play a central role in defining and computing probabilities for continuous multivariate random variables. For two continuous random variables X and Y with joint probability density function (PDF) f_{X,Y}(x,y), the probability that (X,Y) lies in a region A \subseteq \mathbb{R}^2 is given by the double integral P((X,Y) \in A) = \iint_A f_{X,Y}(x,y) \, dx \, dy. This formulation extends the univariate case, where probabilities are areas under the density curve, to joint distributions where they represent volumes under the surface defined by the joint PDF. The joint PDF must satisfy f_{X,Y}(x,y) \geq 0 for all x,y and normalize to total probability 1, i.e., \iint_{\mathbb{R}^2} f_{X,Y}(x,y) \, dx \, dy = 1. Marginal and conditional densities are derived from the joint PDF using multiple integrals. The marginal PDF of X is obtained by integrating out Y: f_X(x) = \int_{-\infty}^{\infty} f_{X,Y}(x,y) \, dy, and similarly for f_Y(y). This process, often facilitated by Fubini's theorem for iterated integrals, yields the univariate distribution of one variable regardless of the other. A prominent example is the bivariate normal distribution, where the joint PDF is f_{X,Y}(x,y) = \frac{1}{2\pi \sigma_X \sigma_Y \sqrt{1-\rho^2}} \exp\left( -\frac{1}{2(1-\rho^2)} \left[ \frac{(x-\mu_X)^2}{\sigma_X^2} + \frac{(y-\mu_Y)^2}{\sigma_Y^2} - 2\rho \frac{(x-\mu_X)(y-\mu_Y)}{\sigma_X \sigma_Y} \right] \right); integrating over y produces the marginal f_X(x), which is normal with mean \mu_X and variance \sigma_X^2, independent of the correlation \rho. Multiple integrals also compute expectations and related quantities for functions of joint random variables. For a measurable function g: \mathbb{R}^2 \to \mathbb{R}, the expectation is E[g(X,Y)] = \iint_{\mathbb{R}^2} g(x,y) f_{X,Y}(x,y) \, dx \, dy, provided the integral exists. This includes moments like the covariance, which measures linear dependence: \text{Cov}(X,Y) = E[(X - \mu_X)(Y - \mu_Y)] = \iint_{\mathbb{R}^2} (x - \mu_X)(y - \mu_Y) f_{X,Y}(x,y) \, dx \, dy = E[XY] - \mu_X \mu_Y. Transformations of random variables often require change of variables in multiple integrals, incorporating the Jacobian determinant to preserve probabilities. If U = g_1(X,Y) and V = g_2(X,Y) with differentiable inverses X = h_1(U,V), Y = h_2(U,V), the joint PDF of (U,V) is f_{U,V}(u,v) = f_{X,Y}(h_1(u,v), h_2(u,v)) \cdot |J|, where J = \det \begin{pmatrix} \frac{\partial h_1}{\partial u} & \frac{\partial h_1}{\partial v} \\ \frac{\partial h_2}{\partial u} & \frac{\partial h_2}{\partial v} \end{pmatrix} is the Jacobian of the inverse transformation. This technique is essential for deriving distributions of functions of joint variables, such as sums or ratios in statistical models.

References

  1. [1]
    Multiple Integrals - Department of Mathematics at UTSA
    Jan 20, 2022 · A multiple integral is a definite integral of a function of several real variables, for instance, f(x, y) or f(x, y, z).
  2. [2]
    Calculus III - Multiple Integrals - Pauls Online Math Notes
    Nov 16, 2022 · In this chapter will be looking at double integrals, i.e. integrating functions of two variables in which the independent variables are from ...
  3. [3]
    [PDF] Multiple Integral
    This chapter shows how to integrate functions of two or more variables. First, a double integral is defined as the limit of sums. Second, we find a fast way ...
  4. [4]
    [PDF] Contents 3 Multiple Integration - Evan Dummit
    In this chapter we develop the theory of integration in multiple variables. Our focus is on double and triple integrals, as they are the ones that show up in ...
  5. [5]
    [PDF] Integration on Rn These are somewhat condensed notes ...
    It allows one to write a multi-variable integral as iterated one-variable integrals which can then be treated using more familiar tools such as the fundamental ...
  6. [6]
    The integrals of multivariable calculus - Math Insight
    ### Summary of Integrals in Multivariable Calculus
  7. [7]
    Calculus history - MacTutor - University of St Andrews
    For Newton the calculus was geometrical while Leibniz took it towards analysis. Leibniz was very conscious that finding a good notation was of fundamental ...
  8. [8]
    Leonhard Euler (1707 - 1783) - Biography - MacTutor
    In Institutiones calculi integralis (1768-70) Euler ... As well as investigating double integrals, Euler considered ordinary and partial differential equations in ...
  9. [9]
    "De formulis integralibus duplicatis" by Leonhard Euler
    Sep 25, 2018 · Euler investigates the theory of double integrals. Published as. Journal article. Published Date. 1770. Written Date. 1768. Original Source ...
  10. [10]
    [PDF] The Definite Integrals of Cauchy and Riemann
    Nov 30, 2022 · While the study of Fourier series raged on for the next couple decades, it wasn't until 1854 that. Bernard Riemann developed a more general ...
  11. [11]
    Did Cauchy ever deal with double or triple integrals?
    Oct 5, 2019 · Cauchy studied why switching the order of integration in a double integral can sometimes lead to different results. This led him to introduce the notion of ...
  12. [12]
    Contribution of Italian Mathematicians to Real Analysis in the last ...
    In this paper, I consider in particular the contribution to real analysis theory done by some Italian mathematicians as Volterra, Peano, Ascoli, Arzelà, etc., ...
  13. [13]
    [PDF] The Lebesgue Integral - arXiv
    May 8, 2023 · The theory of Lebesgue integration was developed by H.-L. Lebesgue in his doctoral dissertation published in 1902 in the italian ...
  14. [14]
    [PDF] Riemann integral in higher dimensions - MIT Mathematics
    Let B = n x : f is not continuous at x o . Then f is integrable if and only if B is a set of measure 0. In particular, continuity of f implies its integrability ...
  15. [15]
    [PDF] Multiple Integrals - People
    MULTIPLE INTEGRALS. 40. Moments of Inertia and Center of Mass. An important application of multiple integrals is finding the center of mass and moments of ...<|control11|><|separator|>
  16. [16]
    [PDF] Measure Theory John K. Hunter - UC Davis Math
    where the integral is with respect n-dimensional Lebesgue measure. The ... The notation Lp(X) assumes that the measure µ on X is understood. We say.
  17. [17]
    [PDF] Analysis on Manifolds
    This book is intended as a text for a course in analysis, at the senior or first-year graduate level. A year-long course in real analysis is an essential ...
  18. [18]
    [PDF] Rudin (1976) Principles of Mathematical Analysis.djvu
    This book is intended to serve as a text for the course in analysis that is usually taken by advanced undergraduates or by first-year students who study ...
  19. [19]
    Calculus III - Double Integrals - Pauls Online Math Notes
    Nov 16, 2022 · A double integral integrates a function of two variables over a 2D region, like a rectangle, and is defined as ∬Rf(x,y)dA.<|control11|><|separator|>
  20. [20]
    Type I and Type II regions
    Type I regions are regions that are bounded by vertical lines x=a and x=b, and curves y=g(x) and y=h(x), where we assume that g(x)<h(x) and
  21. [21]
    [PDF] Evaluation of Double Integrals
    If the set of boundary points of Ω has area zero, then the area of Ω is defined. ... But (8) is a Riemann sum for the double integral I, and is therefore close ...
  22. [22]
    3.6 Triple Integrals
    By simply adding a z -coordinate to our earlier work, we can define both a triple Riemann sum and the corresponding triple integral. 🔗. 🔗. Definition 3.6.2.
  23. [23]
    Introduction to triple integrals - Math Insight
    We want to define the triple integral of f over W to be the total mass of W. As with double integrals, we define the integral with Riemann sums. We chop up the ...
  24. [24]
    [PDF] Unit 17: Triple integrals
    Jul 30, 2024 · MULTIVARIABLE CALCULUS. MATH S-21A. Unit 17: Triple integrals. Lecture ... f(x, y, z) dxdydz is defined as the n → ∞ limit of the Riemann sum.
  25. [25]
    [PDF] Measure Theory - University of Waterloo
    the variables, in multiple integrals. So we write. ZZ f dν dµ or. ZZ f(x, y)dν ... algebra of Lebesgue measurable sets in Rn. Even though m is a complete ...Missing: challenges curse
  26. [26]
    Challenging the Curse of Dimensionality in Multidimensional ... - MDPI
    To calculate such integrals accurately, we need to overcome the curse of dimensionality, which is bad scaling of the computational complexity in terms of ...
  27. [27]
    [PDF] Chapter 2: Lebesgue Measure - UC Davis Math
    coordinate axes and is invariant under orthogonal transformations. We also show that Lebesgue measure transforms under a linear map by a factor equal to the.
  28. [28]
    Calculus III - Iterated Integrals - Pauls Online Math Notes
    Nov 16, 2022 · In this section we will show how Fubini's Theorem can be used to evaluate double integrals where the region of integration is a rectangle.
  29. [29]
    [PDF] Theorems of Fubini and Clairaut In this note we'll prove that, for ...
    In this note we'll prove that, for uniformly continuous functions on a rectangle, the Riemann integral is given by two iterated one variable integrals (Fubini) ...
  30. [30]
    Fubini's Theorem
    Fubini's Theorem: If f(x,y) is a continuous function on a rectangle R=[a,b]×[c,d], then the double integral ∬Rf(x,y)dA is equal to the iterated integral ...
  31. [31]
  32. [32]
    Calculus III - Change of Variables - Pauls Online Math Notes
    Nov 16, 2022 · In order to change variables in a double integral we will need the Jacobian of the transformation. Here is the definition of the Jacobian.
  33. [33]
    [PDF] 18.022: Multivariable calculus — The change of variables theorem
    This determinant is called the Jacobian of F at x. The change-of- variables theorem for double integrals is the following statement. Theorem. Let F: U → V ...
  34. [34]
    [PDF] 16.7 Change of Variables in Multiple Integrals - CSUN
    The Jacobian is evaluated as a 3×3 determinant and is a function of u, v, and w. A change of variables with respect to three variables proceeds in analogy to ...
  35. [35]
    Jacobians
    Change-of-variable formula: If a 1-1 mapping Φ sends a region D∗ in uv-space to a region D in xy-space, then ∬Df(x,y)dxdy = ∬D∗f(Φ(u,v))|∂(x,y)∂(u,v)|dudv. Note ...
  36. [36]
    [PDF] Change of Variables (Jacobian Method)
    (Jacobian ... Change of variables in 3 dimensions. If x = g(u,v,w) y = h(u,v,w) z = j(u,v,w) then where J(u,v,w) = EX 5 Let's check the Jacobian for spherical.
  37. [37]
    Cylindrical and spherical coordinates
    Our Jacobian is then the 3×3 determinant ∂(x,y,z)∂(r,θ,z) = |cos(θ)−rsin(θ)0sin(θ)rcos(θ)0001| = r,. and our volume element is dV=dx ...
  38. [38]
    [PDF] Jacobian - Physics
    Example: Volume element in cylindrical coordinates. • We know that dV = dxdydz in Cartesian coordinates, and also. dV = rdrd cosθdz in cylindrical coordinates ...
  39. [39]
    5.3 Double Integrals in Polar Coordinates - Calculus Volume 3
    Mar 30, 2016 · By symmetry, the total area is twice the area above the polar axis. ... A radial function f f is a function whose value at each point ...
  40. [40]
    A Reduction Formula for Normal Multivariate Integrals - jstor
    The reduction formula expresses (D'3 and 4 as finite sums of single integrals of tabulated functions, which are generally evaluated numerically.
  41. [41]
    Calculus III - Triple Integrals - Pauls Online Math Notes
    Nov 16, 2022 · In this section we will define the triple integral. We will also illustrate quite a few examples of setting up the limits of integration ...Missing: Riemann | Show results with:Riemann
  42. [42]
    Calculus III - Double Integrals over General Regions
    Nov 16, 2022 · In this section we will start evaluating double integrals over general regions, i.e. regions that aren't rectangles.
  43. [43]
    Examples of changing the order of integration in double integrals
    Examples illustrating how to change the order of integration (or reverse the order of integration) in double integrals.
  44. [44]
    [PDF] 18.02SC Notes: Integration in polar coordinates
    In polar coordinates we have z = 1 − r2 and we want the volume under the graph and above the inside of the unit disk. ∫ 2π ∫ 1. ⇒ volume V = 0. 0. (1 − r 2) ...Missing: _0^{ _0^
  45. [45]
    5.2 Double Integrals over General Regions - Calculus Volume 3
    Mar 30, 2016 · 5.2.5 Solve problems involving double improper integrals. In Double Integrals over Rectangular Regions, we studied the concept of double ...
  46. [46]
    Gaussian Integral -- from Wolfram MathWorld
    Gaussian Integral ; int_(-infty)^inftye^(-x^2)dx, = sqrt((int_(-infty)^inftye^(-x^2 ; = sqrt((int_(-infty)^inftye^(-y^2 ; = sqrt(int_(-infty)^inftyint_(-infty)^ ...
  47. [47]
    Improper Multiple Integrals - jstor
    1. Absolute convergence of the integrals defined by thefirst method. Before attacking the main problem we must prove some preliminary theorems. When the proof ...
  48. [48]
    [PDF] An Introduction to Measure Theory - Terry Tao
    tant example of a measure constructed by Carathéodory's theorem is given ... Theorem 1.7.21 (Fubini's theorem). Let (X, BX,µX) and (Y, BY ,µY ) be ...
  49. [49]
    [PDF] Product Measure and Fubini's Theorem - MIT OpenCourseWare
    Fubini's theorem holds under two different sets of conditions: (a) nonnega- tive functions g (compare with the MCT); (b) functions g whose absolute value has a ...
  50. [50]
    [PDF] 8 # » » » » ¹ - LSU Math
    We remark that a Fubini theorem for the Riemann integral is much less general and more cumbersome in its statement than is the case for the Lebesgue integral.
  51. [51]
    15.4 Applications of Double Integrals - Courses
    Compute the moment about the x-axis Ix, the moment about the y-axis Iy, and the moment about the origin I0 for a uniform density disk ρ(x,y)=ρ of radius a ...
  52. [52]
    Triple integrals in physics
    The location of the center-of-mass is as (ˉx,ˉy,ˉz), where ˉx=1M∭Sxρ(x,y,z)dV,ˉy=1M∭Syρ(x,y,z)dV,ˉz=1M∭Szρ(x,y,z)dV.
  53. [53]
    [PDF] Chapter 4 The Electric Potential
    In the last expression we are using the charge density ρ(r) of the distribution to get the element of charge dq for the volume element dτ. 4.1.7 Potential ...
  54. [54]
    [PDF] 11–Applications of the Divergence Theorem - UC Davis Math
    Conclude: The continuity equation implies that the total rate of change of mass inside any volume is equal to minus the flux of mass out through the boundary, a ...
  55. [55]
    [PDF] Lecture 7 Gauss' and Stokes' Theorems
    “The net rate of mass flow of fluid out of the control volume must equal the rate of decrease of the mass of fluid within the control volume" q. S d. Control ...
  56. [56]
    Calculus III - Conservative Vector Fields - Pauls Online Math Notes
    Nov 16, 2022 · In this section we will take a more detailed look at conservative vector fields than we've done in previous sections.
  57. [57]
    Double integrals in probability
    Example: Suppose that two numbers are chosen uniformly from the interval [0,1]. We call the larger of the two numbers X and the smaller one Y. Their joint pdf ...
  58. [58]
    [PDF] Multiple Random Variables
    Definition 4.1.3 A function f(x, y) from R2 into R is called a joint probability density function or joint pdf of the continuous bivariate random vector (X ...
  59. [59]
    [PDF] INTRODUCTION TO STOCHASTIC PROCESSES. BIVARIATE ...
    If u = g(x) and u + ∆u = g(x + ∆x), the shaded areas in this figure represent the same probability. Thus, in the limit ∆x → 0, we must have fU (u)∆u ...
  60. [60]
    23.1 - Change-of-Variables Technique | STAT 414
    This pdf is known as the double exponential or Laplace pdf. « PreviousLesson 23: Transformations of Two Random Variables · Next23.2 - Beta Distribution ...<|control11|><|separator|>