Fact-checked by Grok 2 weeks ago

Fubini's theorem

Fubini's theorem is a cornerstone of and measure theory, providing rigorous justification for evaluating multiple integrals via iterated integrals and interchanging their order under appropriate conditions on the integrand. Named after the mathematician Guido Fubini, who established its general form in 1907, the theorem addresses the equality between a double integral over a product and the corresponding iterated integrals. Specifically, for spaces (X, \mathcal{A}, \mu) and (Y, \mathcal{B}, \nu) with product measure \mu \times \nu on the product σ-algebra, if f: X \times Y \to \mathbb{R} is integrable (i.e., \int_{X \times Y} |f| , d(\mu \times \nu) < \infty), then \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). A related result, often considered a precursor or variant known as Tonelli's theorem, extends this to nonnegative measurable functions f \geq 0, where the iterated integrals always equal the double integral (possibly infinite), without requiring absolute integrability. In the classical setting of multivariable calculus with Riemann integrals, the theorem applies to continuous functions on compact rectangular domains, where the double integral over R = [a,b] \times [c,d] equals both possible iterated integrals. Fubini's original proof in his paper "Sugli integrali multipli" addressed Lebesgue integrals over bounded measurable sets, building on earlier work by Henri Lebesgue in 1904 and special cases known to Leonhard Euler in the 18th century for continuous functions. The theorem's significance lies in its foundational role for computations in higher-dimensional analysis, probability theory (e.g., joint distributions), and applications in physics and engineering involving volume, mass, or expectation calculations.

History

Origins and Development

The concept of interchanging the order of integration in multiple integrals has roots in the 18th century, where Leonhard Euler considered special cases for continuous functions over simple domains, effectively using early forms of what would become Riemann integrals. These ideas were informal and lacked the rigorous foundation provided by later developments in analysis. In the early 20th century, the creation of measure theory by Henri Lebesgue laid the groundwork for generalizing multiple integrals. Lebesgue's 1902 doctoral thesis Intégrale, longueur, aire introduced the Lebesgue measure and integral, enabling treatment of more general functions and sets. By 1904, Lebesgue had extended these ideas to multiple integrals, addressing issues of measurability and integrability that foreshadowed Fubini's theorem.

Key Contributors and Milestones

Guido Fubini, an Italian mathematician born in 1879, made a foundational contribution to multiple integration by proving a theorem allowing the evaluation of double Lebesgue integrals over bounded measurable sets via iterated integrals for integrable functions. This result appeared in his 1907 paper "Sugli integrali multipli," published in the Rendiconti dell'Accademia dei Lincei (series 5, volume 16, pages 608–614). Building on Fubini's work, Leonida Tonelli, another prominent Italian mathematician (1885–1946), extended the theorem in 1909 to apply to non-negative measurable functions, removing the absolute integrability restriction and broadening its applicability in analysis. Tonelli's extension was detailed in his paper "Sull'integrazione per parti," published in the Rendiconti dell'Accademia dei Lincei (series 5, volume 18, pages 246–253). The development of Fubini's theorem was deeply influenced by the earlier foundations of measure theory laid by Henri Lebesgue in his 1902 doctoral thesis , which introduced the Lebesgue integral and measure, enabling rigorous treatment of integration over more general spaces. This work, published in the Annali di Matematica Pura e Applicata (series 3, volume 7, pages 1–163), provided the theoretical groundwork for later generalizations of multiple integrals. In the 1930s, further refinements emerged within functional analysis and abstract measure theory, notably in Stanisław Saks's 1937 monograph Theory of the Integral, which extended Fubini's theorem to abstract product measure spaces and clarified conditions for its validity in broader settings (see sections 8 and 9, pages 76–82). These advancements solidified the theorem's role in modern analysis. Key milestones in the theorem's history include Fubini's initial 1907 formulation for Lebesgue integrable functions over bounded sets, Tonelli's 1909 extension to non-negative functions, and its incorporation into standard measure theory texts during the 1930s, exemplified by Saks's comprehensive treatment.

Foundations

Product Measure Spaces

In measure theory, the product measure space arises from two given measure spaces (X, \mathcal{A}, \mu) and (Y, \mathcal{B}, \nu). The underlying set is the Cartesian product X \times Y, and the product \sigma-algebra \mathcal{A} \otimes \mathcal{B} is the smallest \sigma-algebra on X \times Y that contains all measurable rectangles of the form E \times F, where E \in \mathcal{A} and F \in \mathcal{B}. This construction ensures that the product \sigma-algebra captures the joint measurability structure derived from the individual spaces. The product measure \mu \otimes \nu on \mathcal{A} \otimes \mathcal{B} is defined initially on the collection of measurable rectangles by setting (\mu \otimes \nu)(E \times F) = \mu(E) \nu(F) for E \in \mathcal{A} and F \in \mathcal{B}. The rectangles form a semi-ring, and the set function defined on them is a premeasure. Under the assumption that \mu and \nu are \sigma-finite, Carathéodory's extension theorem guarantees a unique extension of this premeasure to a measure on the product \sigma-algebra. Without \sigma-finiteness, such uniqueness may fail, as multiple extensions could exist. A measure on a space is \sigma-finite if the space can be expressed as a countable union of measurable sets each having finite measure; thus, both X = \bigcup_{n=1}^\infty X_n with \mu(X_n) < \infty for each n, and similarly for Y. This condition is crucial for the well-definedness of the product measure, as it allows the outer measure to coincide with the extension on the product \sigma-algebra and ensures additivity over disjoint unions covering the space. For a measurable rectangle E \times F, the integral of its indicator function with respect to the product measure satisfies \int_{X \times Y} 1_{E \times F} \, d(\mu \otimes \nu) = (\mu \otimes \nu)(E \times F) = \mu(E) \nu(F). This equality follows directly from the definition of the measure on rectangles and the linearity of the integral for simple functions.

Measurable Functions and Integrability

In the context of a product measure space (X \times Y, \mathcal{A} \otimes \mathcal{B}, \mu \otimes \nu), where (X, \mathcal{A}, \mu) and (Y, \mathcal{B}, \nu) are \sigma-finite measure spaces, a function f: X \times Y \to \mathbb{R} is measurable if the preimage f^{-1}(B) belongs to the product \sigma-algebra \mathcal{A} \otimes \mathcal{B} for every Borel set B \subseteq \mathbb{R}. This ensures that the function respects the measurable structure of the product space, allowing integration to be well-defined. Measurable functions on the product space can be approximated by simple functions, which are finite linear combinations of characteristic functions of measurable sets in \mathcal{A} \otimes \mathcal{B}. For a non-negative measurable function f: X \times Y \to [0, \infty), there exists a sequence of non-negative simple functions \{\phi_n\} such that $0 \leq \phi_1 \leq \phi_2 \leq \cdots \leq f and \phi_n(x,y) \to f(x,y) pointwise for all (x,y) \in X \times Y. If f is bounded on a measurable subset E \subseteq X \times Y, the convergence is uniform on E. This approximation property facilitates the extension of the integral from simple functions to general measurable functions via limits. Absolute integrability on the product space requires that f be measurable and that the integral of its absolute value be finite: \int_{X \times Y} |f| \, d(\mu \otimes \nu) < \infty. This condition defines the space L^1(\mu \otimes \nu) of integrable functions on the product, equipped with the norm \|f\|_{L^1} = \int_{X \times Y} |f| \, d(\mu \otimes \nu). For such functions, the product integral of |f| equals the iterated integral \int_X \left( \int_Y |f(x,y)| \, d\nu(y) \right) d\mu(x), and likewise for the reverse order. For non-negative measurable functions f: X \times Y \to [0, \infty], integrability is extended to allow values in the extended non-negative reals, defined as \int_{X \times Y} f \, d(\mu \otimes \nu) = \sup \left\{ \int_{X \times Y} \phi \, d(\mu \otimes \nu) : 0 \leq \phi \leq f, \, \phi \text{ simple} \right\}, which may be infinite. This framework supports the application of by ensuring the integral is well-defined even when finite integrability fails.

Main Theorems

Fubini's Theorem for Integrable Functions

Fubini's theorem provides a foundational result in measure theory for interchanging the order of integration in double integrals over product spaces, specifically when the integrand is absolutely integrable. Consider two σ-finite measure spaces (X, \mathcal{A}, \mu) and (Y, \mathcal{B}, \nu), with the product measure space (X \times Y, \mathcal{A} \otimes \mathcal{B}, \mu \otimes \nu). A measurable function f: X \times Y \to \overline{\mathbb{R}} belongs to L^1(\mu \otimes \nu) if \int_{X \times Y} |f| \, d(\mu \otimes \nu) < \infty, establishing absolute integrability. Under these conditions, the slice functions f_x: Y \to \overline{\mathbb{R}} defined by f_x(y) = f(x, y) for each fixed x \in X are \nu-integrable for \mu-almost every x, meaning f_x \in L^1(\nu) and \int_Y |f_x| \, d\nu < \infty for \mu-a.e. x. Similarly, the slices f^y: X \to \overline{\mathbb{R}} given by f^y(x) = f(x, y) lie in L^1(\mu) for \nu-a.e. y. The theorem asserts that the double integral equals both iterated integrals: \int_{X \times Y} f \, d(\mu \otimes \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 equality of the double integral and the iterated integrals holds due to the absolute integrability condition, which guarantees the integrability of the slices and justifies the interchange. The σ-finiteness of \mu and \nu ensures the product measure is σ-finite, allowing the theorem to apply to standard spaces such as Euclidean spaces equipped with Lebesgue measure. For signed functions, the result follows from applying the non-negative case () to the positive and negative parts after establishing integrability of |f|. The implications of Fubini's theorem are profound in analysis, enabling the reduction of multiple integrals to sequential single integrals, which simplifies computations in areas like partial differential equations and probability theory. It underscores the robustness of the Lebesgue integral over the Riemann integral by providing conditions under which order interchange is valid without additional assumptions on continuity or boundedness.

Tonelli's Theorem for Non-Negative Functions

Tonelli's theorem addresses the interchange of iterated integrals for non-negative measurable functions on product measure spaces, without the requirement of finite integrability that is necessary in . Let (X, \mathcal{A}, \mu) and (Y, \mathcal{B}, \nu) be \sigma-finite measure spaces, and let f: X \times Y \to [0, \infty] be a measurable function with respect to the product \sigma-algebra \mathcal{A} \otimes \mathcal{B}. Then, \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) = \int_{X \times Y} f \, d(\mu \otimes \nu), where all integrals are understood in the extended non-negative sense and may equal +\infty. The \sigma-finiteness condition on the measures \mu and \nu ensures that the product measure \mu \otimes \nu is well-defined and that the theorem holds; without it, counterexamples exist where the equality fails. The function f must be non-negative and measurable, allowing the integrals to potentially diverge while still equating the iterated and double integrals. Unlike Fubini's theorem, which requires f to be absolutely integrable (i.e., \int |f| \, d(\mu \otimes \nu) < \infty) to guarantee finite equal integrals, Tonelli's theorem permits infinite values but does not demand absolute integrability a priori. The proof of Tonelli's theorem relies implicitly on the monotone convergence theorem, which justifies passing limits through integrals when approximating the non-negative measurable f by an increasing sequence of simple functions. This approach establishes the equality for simple functions first and then extends it to general non-negative measurable functions via monotone limits. Tonelli's theorem, introduced by Leonida Tonelli in 1909, forms the foundation for handling broader classes of functions in multiple integrals. For signed integrable functions, Tonelli's theorem applies after decomposing f = f^+ - f^- into non-negative parts, yielding finite iterated integrals under the absolute integrability condition.

Fubini-Tonelli Theorem

The Fubini-Tonelli theorem unifies the results of Fubini's and Tonelli's theorems, providing conditions under which the order of integration can be interchanged for double integrals over product measure spaces. It applies to measurable functions that are either non-negative or integrable, assuming the underlying measure spaces are σ-finite to ensure the product measure is well-defined and the integrals are comparable. This theorem is fundamental in measure theory for computing multiple integrals by iteration, avoiding direct evaluation of the double integral. For a non-negative measurable function f: X \times Y \to [0, \infty], where (X, \mathcal{A}, \mu) and (Y, \mathcal{B}, \nu) are σ-finite measure spaces equipped with the product σ-algebra \mathcal{A} \otimes \mathcal{B} and product measure \mu \times \nu, the theorem asserts that the double integral equals both iterated integrals whenever they are defined: \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). Here, the inner integrals may be infinite, but the equality holds by Tonelli's theorem for non-negative functions, which guarantees the measurability of the iterated integral functions and the equality of the integrals without requiring finiteness. The σ-finiteness of \mu and \nu is essential, as it allows the product space to be covered by countably many finite-measure rectangles, facilitating the extension of the product measure and ensuring no pathologies arise in the integration process. For a signed measurable function f that is integrable over X \times Y, meaning \int_{X \times Y} |f| \, d(\mu \times \nu) < \infty so f \in L^1(\mu \times \nu), the Fubini-Tonelli theorem extends the result by decomposition. Write f = f^+ - f^-, where f^+ = \max(f, 0) and f^- = \max(-f, 0) are non-negative, and both are integrable since |f| = f^+ + f^- is integrable. Applying the non-negative case to f^+ and f^- yields finite iterated integrals for each, and thus: \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). The inner integrals now exist as finite Lebesgue integrals almost everywhere, and the equality in both orders follows directly from the non-negative result applied to the decomposition. This integrability condition ensures that the iterated integrals are equal and finite, distinguishing it from the non-negative case where infinity is permitted.

Extensions and Variations

For Complete Measure Spaces

A measure space (X, \mathcal{M}, \mu) is complete if every subset of a set of \mu-measure zero is measurable and has measure zero. The completion of an incomplete measure space (X, \mathcal{M}, \mu), denoted (X, \mathcal{M}^*, \mu^*), extends the \sigma-algebra \mathcal{M}^* to include all subsets of \mu-null sets, with \mu^* defined by \mu^*(E) = \inf \{ \mu(A) : A \in \mathcal{M}, E \subset A \} for E \subset X, ensuring \mu^* agrees with \mu on \mathcal{M} and assigns measure zero to all null sets and their subsets. This construction incorporates null sets more robustly, enlarging the class of measurable sets without altering the measure on the original \sigma-algebra. For complete product measure spaces, Fubini's theorem extends to the completed product \mu^* \otimes \nu^* on the completed product \sigma-algebra \mathcal{M}^* \otimes \mathcal{N}^*, where (Y, \mathcal{N}, \nu) is another measure space with completion \nu^*. Specifically, if (X, \mathcal{M}^*, \mu^*) and (Y, \mathcal{N}^*, \nu^*) are \sigma-finite complete measure spaces and f: X \times Y \to [-\infty, \infty] is measurable with respect to \mathcal{M}^* \otimes \mathcal{N}^* and \int_{X \times Y} |f| \, d(\mu^* \otimes \nu^*) < \infty, then f_x(y) = f(x,y) is \mathcal{N}^*-measurable and integrable for \mu^*-almost every x \in X, f^y(x) = f(x,y) is \mathcal{M}^*-measurable and integrable for \nu^*-almost every y \in Y, and the iterated integrals equal the double integral: \int_X \left( \int_Y f(x,y) \, d\nu^*(y) \right) d\mu^*(x) = \int_{X \times Y} f \, d(\mu^* \otimes \nu^*) = \int_Y \left( \int_X f(x,y) \, d\mu^*(x) \right) d\nu^*(y). A similar statement holds for non-negative measurable functions via , replacing integrability with the requirement that at least one iterated integral is finite. This extension differs from the standard Fubini-Tonelli theorem for incomplete spaces by permitting integration over a broader class of functions that are measurable only with respect to the completed product \sigma-algebra, which may include sets not in the original \mathcal{M} \otimes \mathcal{N}. However, it necessitates caution with cross-sections (slices), as these may fail to be measurable with respect to the original \sigma-algebras \mathcal{M} or \mathcal{N}, though they remain measurable almost everywhere with respect to the completions \mu^* and \nu^*. The completed product measure \mu^* \otimes \nu^* is generally larger than the completion of the incomplete product \mu \otimes \nu, ensuring the theorem applies directly without additional modifications. An illustrative example arises in the context of on \mathbb{R}, which is complete. The product \lambda^2 on \mathbb{R}^2 has completion \lambda^{*2} = \lambda^* \otimes \lambda^*, resolving measurability issues for sets like the product of a V \subset [0,1] with a singleton \{0\}, which has outer measure zero in the product but is not measurable in the incomplete product \sigma-algebra; under the completion, such sets become measurable with measure zero, allowing to apply to indicator functions of these sets without violating integrability conditions.

Generalizations to Multiple Integrals

The generalization of the Fubini-Tonelli theorem to multiple integrals extends the two-dimensional case through iterative application to n-fold product measures. Consider σ-finite measure spaces (X_i, \mathcal{A}_i, \mu_i) for i = 1, \dots, n. The n-fold product measure \mu = \mu_1 \otimes \cdots \otimes \mu_n is defined on the product space X = X_1 \times \cdots \times X_n equipped with the product \sigma-algebra \mathcal{A} = \mathcal{A}_1 \otimes \cdots \otimes \mathcal{A}_n. For a non-negative measurable function f: X \to [0, \infty] or an integrable function f \in L^1(\mu), the multiple integral equals the iterated integral in any order: \int_X f \, d\mu = \int_{X_1} \left( \cdots \int_{X_{n-1}} \left( \int_{X_n} f(x_1, \dots, x_n) \, d\mu_n(x_n) \right) d\mu_{n-1}(x_{n-1}) \cdots \right) d\mu_1(x_1), and all such iterated integrals coincide. This holds under the condition that all measures \mu_i are \sigma-finite. These generalizations find key applications in probability theory, where they enable the computation of expectations for functions of several random variables by iterating over joint densities on product probability spaces. In partial differential equations, they support the evaluation of higher-dimensional multiple integrals, such as those appearing in convolution representations of solutions.

Proofs

Outline for Lebesgue Integrals

The proof of the Fubini-Tonelli theorem in the Lebesgue integration setting proceeds by leveraging the approximation of measurable functions by simple functions and employing key convergence theorems from measure theory. In a product measure space (X \times Y, \mathcal{M} \otimes \mathcal{N}, \mu \times \nu), where \mu and \nu are \sigma-finite measures, the strategy begins with establishing the product measure via the Carathéodory extension theorem applied to the semi-ring of measurable rectangles A \times B with A \in \mathcal{M}, B \in \mathcal{N}, and measure \mu(A)\nu(B). This construction ensures the double integral is well-defined for non-negative measurable functions, with \sigma-finiteness guaranteeing countable additivity and avoiding issues in non-finite spaces. A foundational step is verifying the measurability of slices: for a measurable set E \subset X \times Y, the x-slice E_x = \{ y \in Y \mid (x,y) \in E \} is \mathcal{N}-measurable for \mu-almost every x \in X, and the measure of the slice \nu(E_x) is measurable in x. Similarly for y-slices. This relies on the \sigma-finiteness to approximate E by unions of rectangles and apply properties of the product \sigma-algebra. For indicator functions of rectangles, $1_{A \times B}, the equality holds explicitly: \iint 1_{A \times B} \, d(\mu \times \nu) = \mu(A) \nu(B) = \int_X \nu(B) 1_A(x) \, d\mu(x) = \int_X \left( \int_Y 1_{A \times B}(x,y) \, d\nu(y) \right) d\mu(x), with the iterated integral in the reverse order matching by symmetry. \sigma-finiteness ensures this extends to indicators of finite disjoint unions of rectangles via linearity and countable additivity. The result extends to simple functions, which are finite non-negative linear combinations \sum c_i 1_{R_i} of disjoint rectangles R_i = A_i \times B_i, by linearity of the integral, yielding equality between the double integral and both iterated integrals. For general non-negative measurable functions f \geq 0, approximate f by an increasing sequence of simple functions f_n \uparrow f pointwise; the monotone convergence theorem then implies \iint f \, d(\mu \times \nu) = \lim_{n \to \infty} \iint f_n \, d(\mu \times \nu) = \lim_{n \to \infty} \int_X \int_Y f_n(x,y) \, d\nu(y) \, d\mu(x) = \int_X \int_Y f(x,y) \, d\nu(y) \, d\mu(x), with the reverse iterated integral following analogously, provided the slices f_x(y) = f(x,y) are measurable for almost every x. This uses the monotone convergence theorem applied to the sequence of inner integrals \int_Y f_n(x,y) \, d\nu(y) \uparrow \int_Y f(x,y) \, d\nu(y). For integrable functions f with \iint |f| \, d(\mu \times \nu) < \infty, decompose f = f^+ - f^- into non-negative parts and apply to each, ensuring the iterated integrals converge absolutely. Alternatively, approximate by simple functions dominated by an integrable g (e.g., |f|) and invoke the to justify the limits, confirming \iint f \, d(\mu \times \nu) = \int_X \int_Y f(x,y) \, d\nu(y) \, d\mu(x) = \int_Y \int_X f(x,y) \, d\mu(x) \, d\nu(y). The role of \sigma-finiteness is pivotal throughout, as it facilitates the measurability of slices for indicators (via exhaustion by finite-measure sets) and enables the approximations without divergence. Unlike proofs for Riemann integrals on compact sets, which depend on uniform continuity, this measure-theoretic approach handles general \sigma-finite spaces.

Proof for Riemann Integrals

Fubini's theorem for Riemann integrals applies to bounded continuous functions defined on a compact rectangular domain in the plane. Specifically, let f: [a, b] \times [c, d] \to \mathbb{R} be continuous on the compact set [a, b] \times [c, d]. Continuity implies that f is bounded and uniformly continuous, ensuring that the double Riemann integral \iint_{[a,b] \times [c,d]} f(x,y) \, d(x,y) exists and is finite, where the integral is defined via upper and lower Darboux sums over partitions of the rectangle into subrectangles. The theorem states that this double integral equals the iterated Riemann integrals in either order: \iint_{[a,b] \times [c,d]} f(x,y) \, d(x,y) = \int_a^b \left( \int_c^d f(x,y) \, dy \right) dx = \int_c^d \left( \int_a^b f(x,y) \, dx \right) dy. For fixed x \in [a, b], the function y \mapsto f(x, y) is continuous on the compact interval [c, d], so the inner integral \int_c^d f(x,y) \, dy exists as a Riemann integral, yielding a function g(x) = \int_c^d f(x,y) \, dy that is continuous on [a, b] due to the uniform continuity of f. Thus, the outer integral \int_a^b g(x) \, dx exists. The argument for the reverse order is symmetric. To establish equality with the double integral, consider product partitions: partition [a, b] into subintervals of lengths \Delta x_i with points x_{i-1} < x_i for i = 1, \dots, m, and [c, d] into subintervals of lengths \Delta y_j with points y_{j-1} < y_j for j = 1, \dots, n. A Riemann sum for the double integral is S = \sum_{i=1}^m \sum_{j=1}^n f(\xi_i, \eta_j) \Delta x_i \Delta y_j, where \xi_i \in [x_{i-1}, x_i] and \eta_j \in [y_{j-1}, y_j]. This sum can be rewritten as S = \sum_{i=1}^m \left( \sum_{j=1}^n f(\xi_i, \eta_j) \Delta y_j \right) \Delta x_i. The inner sum approximates \int_c^d f(\xi_i, y) \, dy, and uniform continuity of f ensures that the approximation error is bounded uniformly in i by \epsilon \cdot (d - c) as the norm of the partition of [c, d] approaches zero, for any \epsilon > 0. Thus, as the mesh norms of both partitions tend to zero, S converges to \int_a^b \left( \int_c^d f(x,y) \, dy \right) dx. The proof for the other follows analogously by regrouping the sums. The historical development traces back to Paul du Bois-Reymond, who proved the result for continuous functions in 1882 by reducing the double integral to one-dimensional estimates via and bounds on partial sums. Guido Fubini's 1907 work extended the theorem to Lebesgue integrals, but his approach similarly reduced multiple integrals to iterated forms through limiting processes on measurable sets, influencing later Riemann treatments. This proof relies crucially on and , which guarantee and the existence of all . However, the result fails for merely Riemann integrable (possibly discontinuous) functions, as counterexamples exist where the double is finite but one or both iterated integrals diverge or yield unequal values, highlighting the need for the more robust Lebesgue theory to handle broader classes of functions.

Counterexamples and Limitations

Failures in Non-σ-Finite Spaces

In non-σ-finite measure spaces, Tonelli's theorem fails to hold in general, as the double of a non-negative with respect to the may differ from the iterated integrals. This occurs because the is not uniquely determined without σ-finiteness, and the standard construction leads to discrepancies in how sets or functions are assigned measure. Specifically, sets that appear "thin" in slices can have positive or infinite if they span the entire non-σ-finite space, violating the equality guaranteed by the theorem under σ-finite assumptions. A canonical counterexample uses an uncountable set with a measure that vanishes on countable subsets. Let X be an , \mathcal{A} = 2^X the power set of X, and \mu: \mathcal{A} \to [0,\infty] the measure defined by \mu(A) = \begin{cases} 0 & \text{if $A$ is countable}, \\ \infty & \text{if $A$ is uncountable}. \end{cases} This defines a complete, σ-additive measure on (X, \mathcal{A}), but it is not σ-finite, as any countable collection of finite-measure sets (i.e., countable subsets) covers at most a countable portion of X. Consider the product measure space (X \times X, \mathcal{A} \otimes \mathcal{A}, \mu \times \mu), where the product \mathcal{A} \otimes \mathcal{A} is the power set $2^{X \times X} (generated by all rectangles A \times B with A, B \in \mathcal{A}), and the satisfies (\mu \times \mu)(A \times B) = \mu(A) \mu(B) with the convention that $0 \cdot \infty = 0 and \infty \cdot 0 = 0. Now define the non-negative measurable function f: X \times X \to [0,\infty) as the indicator of the diagonal, f(x,y) = \mathbf{1}_D(x,y) = \begin{cases} 1 & \text{if $x = y$}, \\ 0 & \text{otherwise}, \end{cases} where D = \{(x,x) \mid x \in X\}&#36;. The set D$ is measurable in the product σ-algebra, as it belongs to the power set. For the iterated integrals, fix x \in X; the y-section is f_x(y) = \mathbf{1}_{\{x\}}(y), so \int_X f(x,y) \, d\mu(y) = \mu(\{x\}) = 0, since \{x\} is countable. Thus, the iterated integral is \int_X \left( \int_X f(x,y) \, d\mu(y) \right) d\mu(x) = \int_X 0 \, d\mu(x) = 0. By symmetry, the reverse iteration \int_Y \left( \int_X f(x,y) \, d\mu(x) \right) d\mu(y) also equals 0. However, the double integral with respect to the product measure is \int_{X \times X} f \, d(\mu \times \mu) = (\mu \times \mu)(D) = \infty. This holds because any set E \subset X \times X of finite product measure must satisfy \mu(E_x) < \infty for \mu-almost every x \in X, implying E_x is countable (hence measure 0) for almost all x; the only such sets of positive measure are effectively empty in the uncountable directions. Since every section D_x = \{x\} is non-empty (though measure 0), D cannot have finite product measure and must have infinite measure under this construction. This illustrates that without σ-finiteness, the assigns infinite measure to the diagonal despite the slices having measure zero, leading to iterated integrals of 0 but a double integral of \infty. Consequently, Tonelli's theorem requires the σ-finiteness condition to ensure that non-negative measurable functions have equal iterated and double integrals, preventing such pathologies in the product construction.

Failures for Non-Measurable or Non-Integrable Functions

Fubini's theorem relies on the function being measurable with respect to the product and absolutely integrable over the product ; violations of these conditions can lead to the double being undefined or the iterated integrals failing to exist or coinciding with it. A prominent for non-measurable functions involves the , a non-Lebesgue measurable V of [0,1] constructed via the by selecting one representative from each of the in the reals 1. Define the function f: [0,1] × [0,1] → {0,1} by f(x,y) = 1 if x - y ∈ V and 0 otherwise. The set {(x,y) | f(x,y) = 1} is not product-measurable, as its projection onto either axis would imply V is measurable if the set were, leading to a contradiction with the properties of V. Consequently, the double ∬ f d(μ × ν) is undefined, where μ and ν are Lebesgue measures on [0,1]. However, the iterated integrals yield different values: one order computes to 0 and the other to 1, due to the disjoint countable covering of [0,1] by rational translates of V, which forces the vertical sections to have measure 0 while the horizontal sections have measure 1 in an sense. This discrepancy highlights that without product measurability, Fubini's theorem cannot guarantee equality between the double and iterated integrals. For non-integrable functions, consider f(x,y) = (x - y)^{-2} on (0,1) × (0,1), which is Borel measurable but fails absolute integrability over the product space, as ∬_{(0,1)^2} |f(x,y)| , dμ dν = ∞ due to the non-integrable singularity along the diagonal x = y, where the behavior resembles ∫ 1/z^2 dz near 0, which diverges. The double integral of f is thus undefined in the Lebesgue sense because the positive and negative parts (though f is positive here, the divergence prevents definition). The iterated integrals ∫_0^1 \left( ∫_0^1 f(x,y) , dy \right) dx and the reverse also diverge in the Lebesgue sense, as the inner integral ∫0^1 (x - y)^{-2} dy = \left[ \frac{1}{x - y} \right]{y=0}^{y=1} formally yields 1/(x-1) - 1/x = -(1/x + 1/(1-x)), but the improper nature causes divergence to +∞ from the singularity at y = x, preventing the inner function from being Lebesgue integrable for any x ∈ (0,1). Without the integrability condition, the theorem fails because the orders cannot be interchanged reliably, and conditional convergence does not hold in the Lebesgue framework—specifically, ∫_0^1 ∫_0^1 |f| , dy , dx = ∞, while naive computations might suggest finite values via symmetric limits or principal values, underscoring the need for absolute integrability to ensure the integrals exist and coincide.

Issues with Non-Maximal Product Measures

In measure theory, the standard construction of the space for two measure spaces (X, \mathcal{A}, \mu) and (Y, \mathcal{B}, \nu) begins with the product \sigma-algebra \mathcal{A} \otimes \mathcal{B}, which is the smallest \sigma-algebra on X \times Y containing all measurable rectangles of the form A \times B where A \in \mathcal{A} and B \in \mathcal{B}. The \mu \times \nu is then defined uniquely on this \sigma-algebra under \sigma-finiteness assumptions. However, this product \sigma-algebra is often referred to as non-maximal because it does not include all subsets of null sets under the , unlike the completion \overline{\mu \times \nu} of the , which enlarges the \sigma-algebra to include all subsets of null sets and is essential for applications like on \mathbb{R}^n. When Fubini's theorem is applied using only the non-maximal product \sigma-algebra \mathcal{A} \otimes \mathcal{B}, certain functions that are integrable with respect to the completed may fail to be measurable, rendering the double undefined. For instance, consider the spaces ([0,1], \mathcal{B}([0,1]), m) and ([0,1], \mathcal{B}([0,1]), m), where \mathcal{B}([0,1]) is the Borel \sigma-algebra and m is (incomplete on Borel sets). Let N \subset [0,1] be a Borel null set (e.g., a fat of measure zero), and let V \subset N be a non-Borel subset of N. The set E = V \times [0,1] \subset [0,1]^2 has product measure zero since m(V) = 0, and in the completed product space (corresponding to on [0,1]^2), E is measurable with m \times m (E) = 0. Thus, the \chi_E is integrable over the completed space with zero. However, \chi_E is not measurable with respect to the product Borel \sigma-algebra \mathcal{B}([0,1]) \otimes \mathcal{B}([0,1]), because if it were, the projection onto the first coordinate would imply V is Borel, a contradiction. Consequently, the double \iint \chi_E \, dm \, dm is undefined in the non-maximal setting, even though the iterated integrals exist and equal zero: \int_0^1 \left( \int_0^1 \chi_E(x,y) \, dy \right) dm(x) = \int_V 1 \, dm(x) = 0 and similarly for the reverse order. This example illustrates the implication for Fubini's theorem: to ensure that measurable functions in the completed product space have measurable slices almost everywhere and allow the interchange of integrals for a broader class of integrands, the theorem must be formulated using the completed product \sigma-algebra \overline{\mathcal{A} \otimes \mathcal{B}}. Without completion, the sections of sets (or functions) measurable in the full space may not be measurable with respect to the original \sigma-algebras \mathcal{A} and \mathcal{B}, violating the hypotheses of Fubini even when the iterated integrals formally exist. In extensions to complete measure spaces, the completed product aligns the theorem with the product of the individual completions, preserving the equality of iterated and double integrals almost everywhere. In the context of , where measure spaces are often taken to be complete, a related issue arises with the " algebra"—the generated by finite disjoint unions of measurable rectangles, which is coarser than the full product \sigma-. Independent events generated by independent \sigma- may not belong to this algebra, as they can require infinite operations to construct (e.g., limits of approximations for events like \{X = Y\} for independent random variables X, Y on atomless spaces). Consequently, integrals over such events cannot be computed using only -based approximations without passing to the full product \sigma-, underscoring the need for the maximal structure in Fubini's application to joint distributions.

Applications

Iterated Integrals and Product Formulas

Fubini's theorem enables the evaluation of integrals over spaces by expressing the double integral as an . For an integrable f: X \times Y \to \mathbb{R} on \sigma-finite measure spaces (X, \mathcal{A}, \mu) and (Y, \mathcal{B}, \nu), with product measure \mu \times \nu, \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 equality justifies computing the double integral by successive single-variable integrations, provided f is integrable with respect to the . Tonelli's theorem extends this result to non-negative measurable functions, ensuring the iterated integrals coincide (possibly infinite) without requiring prior integrability. Specifically, for a non-negative measurable f, \int_X \left( \int_Y f(x,y) \, d\nu(y) \right) d\mu(x) = \int_{X \times Y} f \, d(\mu \times \nu) = \int_Y \left( \int_X f(x,y) \, d\mu(x) \right) d\nu(y). This facilitates applications where functions may not be integrable but yield finite iterated integrals. A key application arises for product functions f(x,y) = g(x) h(y), where g: X \to [0,\infty) is \mathcal{A}-measurable and h: Y \to [0,\infty) is \mathcal{B}-measurable. Tonelli's theorem implies \int_{X \times Y} g(x) h(y) \, d(\mu \times \nu) = \left( \int_X g \, d\mu \right) \left( \int_Y h \, d\nu \right). The proof sketch proceeds by evaluating the : the inner integral \int_Y g(x) h(y) \, d\nu(y) = g(x) \int_Y h \, d\nu, and the outer integral then factors as \left( \int_Y h \, d\nu \right) \int_X g \, d\mu, with equality in the reverse order by symmetry. For general integrable product functions (not necessarily positive), the result follows from Fubini's theorem applied to the separately.

Computational Examples

Fubini's theorem facilitates the evaluation of double integrals by allowing the interchange of the order of integration, often simplifying computations that would be challenging as a single double integral. One classic example is the computation of the normalizing constant for the bivariate normal distribution. Consider the double integral over \mathbb{R}^2: \iint_{\mathbb{R}^2} e^{-(x^2 + y^2)/2} \, dx \, dy. By Fubini's theorem, since the integrand is positive and the integral is absolutely convergent, this equals \int_{-\infty}^{\infty} \left( \int_{-\infty}^{\infty} e^{-x^2/2} \, dx \right) e^{-y^2/2} \, dy = \left( \int_{-\infty}^{\infty} e^{-x^2/2} \, dx \right)^2 = (\sqrt{2\pi})^2 = 2\pi. The inner Gaussian integral is a standard result, and the squaring step leverages the separability enabled by Fubini. This evaluation is crucial in probability theory for normalizing the standard bivariate normal density. Another illustrative example is the double integral representation of the dilogarithm function at 1, which equals the Riemann zeta function at 2: \int_0^1 \int_0^1 \frac{-\ln(1 - xy)}{xy} \, dx \, dy = \frac{\pi^2}{6}. To compute this using Fubini's theorem, first note the series expansion -\ln(1 - xy) = \sum_{n=1}^\infty \frac{(xy)^n}{n}, so the integrand becomes \sum_{n=1}^\infty \frac{(xy)^{n-1}}{n}. Integrating term by term, justified by Fubini for the positive terms, \int_0^1 \int_0^1 \sum_{n=1}^\infty \frac{(xy)^{n-1}}{n} \, dx \, dy = \sum_{n=1}^\infty \frac{1}{n} \left( \int_0^1 x^{n-1} \, dx \right) \left( \int_0^1 y^{n-1} \, dy \right) = \sum_{n=1}^\infty \frac{1}{n^2} = \zeta(2) = \frac{\pi^2}{6}. This approach highlights how Fubini allows separation into iterated integrals, leading to the Basel problem solution.

References

  1. [1]
    Fubini Theorem -- from Wolfram MathWorld
    Fubini's theorem, sometimes called Tonelli's theorem, establishes a connection between a multiple integral and a repeated one.
  2. [2]
    [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 ...
  3. [3]
    Fubini's Theorem - FOSSEE Animations
    History. The special case of Fubini's theorem was known to Leonhard Euler in the 18th century. In 1904 Henri Lebesgue extended this to bounded measurable ...
  4. [4]
    [PDF] 6.2 Fubini's Theorem - LSU Math
    We remark that the first form of Fubini's theorem expresses the product measure µ×ν of a set C ∈ C as the integral with respect to µ of the ν-measures of the x ...
  5. [5]
    Integration (Chapter 4) - Real Analysis and Probability
    *Fubini, Guido (1907). Sugli integrali multipli. Rendiconti Accad. Nazionale dei Lincei (Rome) (Ser. 5) 16: 608–614. Halmos, Paul R. (1950). Measure Theory ...
  6. [6]
    [PDF] 5 — Product Measures and the Fubini-Tonelli Theorem
    Nov 15, 2019 · Our first task is to define the product. Definition 5.1 (Product Measure). Let (X, M,µ) and (Y, N,ν) be measure spaces.
  7. [7]
    Intégrale, Longueur, aire : Lebesgue, Henri Leon, 1875-1941
    Jun 25, 2018 · Intégrale, Longueur, aire : 129 p. ; 30 cm Thesis--Université de Paris, 1902 Bibliographical foot-notes Notes No copyright page.
  8. [8]
    Theory of the integral - EuDML
    Theory of the integral. Stanisław Saks. 1937. Access Full Book. top. icon ... Fubini...................................................... 76 § 9. Fubini's ...
  9. [9]
    [PDF] Chapter 5. Product Measures - UC Davis Math
    The integral of a measurable function on the product space may be evaluated as iterated integrals on the individual spaces provided that the function is ...
  10. [10]
    [PDF] Notes on product measures for Math 501, Fall 2010
    Nov 9, 2010 · measure µ ⊗ ν on the product sigma algebra M⊗N that extends m on A. ... be any rectangle with finite product measure. Then for every > 0 ...
  11. [11]
    [PDF] Lecture 7: February 20 7.1 Measure on Product Spaces
    Definition 7.1 (Product σ-Field) Let (Ω1, F1), (Ω2, F2) be two measurable spaces. The product σ-field F1 ⊗ F2 on Ω1 × Ω2 is defined as the σ-field generated by ...
  12. [12]
    [PDF] 4.1 Approximation by Simple Functions - Christopher Heil
    Often, the easiest way to deal with a generic measurable function is to approx- imate it by simpler functions. Of course, the meaning of “simpler” is in the.
  13. [13]
    [PDF] Chapter 3. Measurable Functions - UC Davis Math
    In defining the Lebesgue integral of a measurable function, we will approximate it by simple functions. By contrast, in defining the Riemann integral of a ...
  14. [14]
    [PDF] Chapter 6 - Product Measures and Fubini's Theorem
    The main step in the proof of Fubini's theorem is to extend Lemma 6.14 to the case in which A is an arbitrary element of S T. But it is sufficient to still work ...
  15. [15]
    Fubini's theorem for the Lebesgue integral - PlanetMath
    Mar 22, 2013 · This is the version of the Fubini's Theorem for the Lebesgue integral Mathworld Planetmath. For the Riemann integral, see the standard calculus version.
  16. [16]
    [PDF] Alternate treatment of Fubini's theorem
    repeated integration to the case of Lebesgue integrable functions. The las section contains some applications. 1. Fubini's Theorem. We shall use the ...
  17. [17]
    [PDF] An Introduction to Measure Theory - Terry Tao
    Tonelli's theorem can fail if the σ-finite hypothesis is removed, and also that product measure need not be unique. Let X is the unit interval [0, 1] with ...<|control11|><|separator|>
  18. [18]
    [PDF] Lecture 5 Theorems of Fubini-Tonelli and Radon-Nikodym
    We have seen that it is possible to define products of arbitrary collec- tions of measurable spaces - one generates the σ-algebra on the prod-.<|separator|>
  19. [19]
    [PDF] Tonelli Fubini Fubini in practice Sums and integrals
    Fubini's Theorem: If (X, E,µ) and (Y, K,ν) are two σ-finite measure spaces and if f ∈ M(X ×Y, E ⊗ K) is integrable w.r.t. µ ⊗ ν then.
  20. [20]
    [PDF] Mathematics Department Stanford University
    We can now state Fubini's Theorem. In the statement we require that the measure spaces (X,A,µ) and (Y,B,ν) be complete; a ...
  21. [21]
    [PDF] Measure Theory John K. Hunter - UC Davis Math
    In that case, it is convenient to deal with complete measure spaces. ... Theorem 5.18 (Fubini's Theorem). Suppose that (X, A,µ) and (Y, B,ν) are σ ...
  22. [22]
    [PDF] Real Analysis
    Folland, Gerald B. Real analysis : modern techniques and their applications I Gerald B. Folland. - 2nd ed. p. em. - (Pure and applied mathematics).
  23. [23]
    [PDF] 3.6. Product measure and Fubini's theorem. Let (E 1, E1,µ1) and (E2 ...
    The existence of product measure and Fubini's theorem extend easily to σ-finite ... taking the n-fold product of Lebesgue measure on R is called Lebesgue measure ...
  24. [24]
    [PDF] Math 581 Exam 2
    Fubini's theorem for product probability measures shows double integrals can be calculated with iterated integrals if X1 X2, and the theorem is sometimes ...
  25. [25]
    [PDF] Notes on Partial Differential Equations John K. Hunter - UC Davis Math
    which follows immediately from Fubini's theorem. If n = 3, the theorem states that ... [5] L. C. Evans, Partial Differential Equations, Amer. Math. Soc ...Missing: multiple | Show results with:multiple
  26. [26]
    [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) ...
  27. [27]
    [PDF] Fubini's Theorem for Riemann Integrals - Yizhen Chen
    f(t, y)dy as h → 0 because f is uniformly continuous. So the two integrals have the same derivative, and same value when t = 0, thus they are equal.
  28. [28]
    [PDF] Notes on Riemann Integral - UC Berkeley math
    Dec 2, 2010 · As a corollary of inequality (59), we obtain so called Fubini's Theorem. (for continuous functions proven by du Bois-Reymond already in 1872,.<|control11|><|separator|>
  29. [29]
    [PDF] Version of 31.5.03 Chapter 25 Product Measures I come now to ...
    The first two sections of this chapter are devoted to an analysis of the relationship between one- and two-dimensional Lebesgue measure which makes these.<|control11|><|separator|>
  30. [30]
    [PDF] Measure Theory John K. Hunter - UC Davis Math
    Measures are a generalization of volume; the fundamental example is Lebesgue measure on Rn, which we discuss in detail in the next Chapter. Moreover, as.
  31. [31]
    [PDF] MIRA.pdf - Measure, Integration & Real Analysis
    In addition to publishing numerous research papers, he is the author of six mathematics textbooks, ranging from freshman to graduate level. His book Linear ...
  32. [32]
    10.2 Iterated integrals and Fubini theorem
    The Fubini theorem is commonly thought of as the theorem that allows us to swap the order of iterated integrals, although there are many variations on Fubini, ...
  33. [33]
    [PDF] THE GAUSSIAN INTEGRAL Let I = ∫ ∞ e dx, J ... - Keith Conrad
    The calculation in Section 2 that the iterated integral on the right is π/4 does not need Fubini's theorem in any form. It is going from the iterated integral ...
  34. [34]
    [PDF] series involving dirichlet eta function - ResearchGate
    Abstract. In this article, we obtain an integral representation for a remainder sum of the Dirichlet Eta function. We then obtain numerous generating func-.
  35. [35]
    [PDF] Expression as a Legendre Function, of an Elliptic Integral ... - DTIC
    RELATION TO LEGENDRE FUNCTIONS. T"he integral +t, can ce transforrmed into a double integral by using the Laplace transform of the. Bessel function of first.<|separator|>
  36. [36]
    DLMF: §6.7 Integral Representations ‣ Properties ‣ Chapter 6 ...
    §6.7(i) Exponential Integrals. ⓘ. Keywords: exponential integrals, integral representations; Notes: (6.7.1) and (6.7.2) follow from the definitions (§6.2(i) ...<|separator|>