Fact-checked by Grok 2 weeks ago

Absolute convergence

In mathematics, particularly in the field of real and complex analysis, absolute convergence refers to a strong form of convergence for an infinite series \sum_{n=1}^\infty a_n, where the series formed by the absolute values of its terms, \sum_{n=1}^\infty |a_n|, also converges. This property ensures that the original series converges, since the partial sums of \sum a_n are bounded in absolute value by the partial sums of the convergent positive-term series \sum |a_n|, allowing the application of tests like the comparison test for convergence. A key advantage of absolute convergence is that the sum of the series is independent of the order in which the terms are rearranged, unlike conditionally convergent series where such rearrangements can alter the sum arbitrarily. Additionally, the product of two absolutely convergent series is itself absolutely convergent, providing a useful closure property in series manipulations. In contrast, a series that converges but whose absolute-value series diverges is termed conditionally convergent, highlighting the role of term cancellations in achieving convergence without absolute boundedness.

Basic Concepts

Definition for series of real and complex numbers

In the context of infinite series with real terms, a series \sum_{n=1}^\infty a_n, where a_n \in \mathbb{R}, is said to converge absolutely if the series of absolute values \sum_{n=1}^\infty |a_n| converges. This condition is equivalent to \sum_{n=1}^\infty |a_n| < \infty. Absolute convergence ensures that the original series converges, though the converse does not always hold. A classic example is the p-series \sum_{n=1}^\infty \frac{1}{n^p}, which converges absolutely if p > 1 and diverges if p \leq 1. For instance, the series \sum_{n=1}^\infty \frac{1}{n} (where p=1) diverges, so it does not converge absolutely. In contrast, \sum_{n=1}^\infty \frac{1}{n^2} (where p=2 > 1) converges absolutely. The alternating series \sum_{n=1}^\infty \frac{(-1)^{n+1}}{n} converges but not absolutely, illustrating . For series with complex terms, \sum_{n=1}^\infty z_n where z_n \in \mathbb{C}, absolute convergence is defined analogously: the series converges absolutely if \sum_{n=1}^\infty |z_n| converges, with |z_n| denoting the . The modulus of a z = x + iy, with x, y \in \mathbb{R} and i = \sqrt{-1}, is given by |z| = \sqrt{x^2 + y^2}. This definition aligns with the real case, as absolute convergence of \sum z_n holds the series of the real parts \sum \operatorname{Re}(z_n) and imaginary parts \sum \operatorname{Im}(z_n) both converge absolutely.

Historical development

The concept of absolute convergence emerged in the early as mathematicians sought to rigorize the study of infinite series. played a pivotal role with his 1821 textbook Cours d'analyse de l'Académie Royale des Sciences, where he defined for series and required that the absolute values of sums of subsequent terms remain bounded by arbitrarily small limits as the number of terms increases. This criterion effectively anticipated absolute convergence by ensuring stability regardless of minor variations in order, though Cauchy did not yet employ the modern terminology. A key insight into the limitations of ordinary convergence came from in 1829, during his investigations into the . Dirichlet observed that certain could have their sums altered by rearranging terms, revealing the phenomenon of and the necessity of absolute convergence for order-independent results. In the 1850s, advanced this distinction, particularly through his work around 1852 on trigonometric series. Riemann recognized in alternating series like the harmonic series and proved that such series could be rearranged to yield any desired real , formalizing the contrast with absolutely convergent series that resist such manipulation. Mid-19th-century efforts by and Dirichlet further solidified the theory, emphasizing absolute convergence for rigorous handling of infinite series in contexts like . Weierstrass, through his lectures and publications in the 1860s and 1870s, developed convergence tests that relied on absolute summability to guarantee , enhancing the applicability of series expansions. Twentieth-century developments integrated absolute convergence into , with Stefan Banach's contributions in the 1920s marking a significant extension. In his 1922 dissertation Théorie des opérations linéaires, Banach demonstrated that in complete normed linear spaces—termed Banach spaces—absolute convergence in the norm implies ordinary convergence, broadening the concept to infinite-dimensional settings.

Convergence Properties

Relation to ordinary convergence

Ordinary convergence of an infinite series \sum a_n refers to the convergence of its partial sums to a finite , without regard to the signs of the terms. In contrast, absolute convergence requires that the series \sum |a_n| converges. A series that converges ordinarily but not absolutely is said to converge conditionally. When all terms a_n are nonnegative, absolute convergence is equivalent to ordinary convergence, as |a_n| = a_n in this case. Absolute convergence implies ordinary convergence for any series, but the does not hold. For instance, the alternating series \sum_{n=1}^\infty \frac{(-1)^{n+1}}{n} converges ordinarily to \ln 2, yet the series of absolute values \sum_{n=1}^\infty \frac{1}{n} diverges, demonstrating . In the context of series of functions, the Weierstrass M-test provides a criterion linking absolute convergence to : if |f_n(x)| \leq M_n for all x in a set and \sum M_n converges, then \sum f_n(x) converges absolutely and uniformly on that set.

Proofs for numbers and Banach spaces

For a series \sum_{n=1}^\infty a_n of real numbers, absolute convergence means \sum_{n=1}^\infty |a_n| < \infty. The partial s_n = \sum_{k=1}^n a_k satisfy, for m > n, |s_m - s_n| = \left| \sum_{k=n+1}^m a_k \right| \leq \sum_{k=n+1}^m |a_k|, by the . Since \sum |a_k| converges, the tail \sum_{k=n+1}^m |a_k| \to 0 as m, n \to \infty, so \{s_n\} is a in \mathbb{R}, which is complete, and thus converges. The generalization to Banach spaces is addressed in the dedicated subsection below.

Proof for Complex Numbers

Consider a series \sum_{k=1}^\infty a_k of complex numbers where a_k \in \mathbb{C}. The partial sums are defined as s_n = \sum_{k=1}^n a_k. The series converges if the sequence \{s_n\} is Cauchy, meaning that for every \epsilon > 0, there exists N \in \mathbb{N} such that |s_m - s_n| < \epsilon for all m > n > N. Assume the series is absolutely convergent, so \sum_{k=1}^\infty |a_k| < \infty. Then, for m > n, |s_m - s_n| = \left| \sum_{k=n+1}^m a_k \right| \leq \sum_{k=n+1}^m |a_k|, by the for the modulus in \mathbb{C}. Since \sum |a_k| converges, its partial sums form a , so \sum_{k=n+1}^m |a_k| \to 0 as m, n \to \infty. Thus, \{s_n\} is Cauchy and converges in \mathbb{C}, which is complete. An alternative proof decomposes into real and imaginary parts. Let a_k = u_k + i v_k with u_k, v_k \in \mathbb{R}. Then |u_k| \leq |a_k| and |v_k| \leq |a_k|, so \sum |u_k| and \sum |v_k| converge by the comparison test. For the real series \sum u_k, note that $0 \leq u_k + |u_k| \leq 2 |u_k|, so \sum (u_k + |u_k|) converges, implying \sum u_k = \sum (u_k + |u_k|) - \sum |u_k| converges as a difference of convergent series. Similarly for \sum v_k. Hence, \sum a_k = \sum u_k + i \sum v_k converges.

Generalization to Banach Spaces

In a Banach space (X, \|\cdot\|), which is a complete normed vector space, absolute convergence of \sum_{n=1}^\infty x_n means \sum_{n=1}^\infty \|x_n\| < \infty. The partial sums s_n = \sum_{k=1}^n x_k satisfy, for m > n, \|s_m - s_n\| = \left\| \sum_{k=n+1}^m x_k \right\| \leq \sum_{k=n+1}^m \|x_k\|, by the triangle inequality for the norm. Since \sum \|x_k\| converges in \mathbb{R}, the tail \sum_{k=n+1}^m \|x_k\| \to 0 as m, n \to \infty, so \{s_n\} is Cauchy in X. By completeness of X, \{s_n\} converges to some s \in X, and thus \sum x_n converges.

Generalizations to Vector Spaces

Absolute convergence in normed spaces

In a normed vector space (X, \|\cdot\|), a series \sum_{n=1}^\infty x_n with x_n \in X is said to converge absolutely if the scalar series \sum_{n=1}^\infty \|x_n\| converges (to a finite real number). This definition generalizes the notion from real or complex numbers by replacing the absolute value with the vector space norm, which induces a metric on X. Absolute convergence has key implications for the behavior of the series. Specifically, if \sum \|x_n\| < \infty, then the partial sums s_m = \sum_{n=1}^m x_n form a Cauchy sequence in the norm of X, because for m > k, \|s_m - s_k\| \leq \sum_{n=k+1}^m \|x_n\| \to 0 as k, m \to \infty. In complete normed spaces (Banach spaces), this Cauchy property ensures that the series converges to some limit in X. Moreover, a normed space X is complete if and only if every absolutely convergent series in X converges. A representative example occurs in the Banach spaces \ell^p for $1 < p < \infty, consisting of sequences (a_n) with \sum |a_n|^p < \infty under the p-norm. Consider the coordinate series \sum a_n e_n, where e_n are the standard basis vectors (with 1 in the nth position and 0 elsewhere). This series converges in \ell^p if and only if (a_n) \in \ell^p, but it converges absolutely if and only if (a_n) \in \ell^1, since \|a_n e_n\|_p = |a_n| and \sum |a_n| < \infty. Thus, absolute convergence in \ell^p corresponds to membership of the coefficient sequence in \ell^1. In finite-dimensional normed spaces, the concept aligns closely with that for series of real or complex numbers, as all norms are equivalent and the space is complete, but the definition applies vector norms to the terms.

In topological vector spaces

In locally convex topological vector spaces, the notion of absolute convergence for a series \sum x_n, where x_n belong to the space X, is defined in terms of the topology rather than a single norm. A series converges absolutely if, for every continuous seminorm p on X, the numerical series \sum p(x_n) converges (to a finite value). Equivalently, the series converges absolutely if \sum |f(x_n)| < \infty for every continuous linear functional f on X. This generalization captures the idea that the terms become "small" in a uniform way across the topology, without requiring a norm. In such spaces, the topology is generated by a family of continuous seminorms \{p_\alpha\}_{\alpha \in A}. Here, the series \sum x_n is absolutely convergent if and only if \sum p_\alpha(x_n) < \infty for every \alpha \in A. This condition ensures that the partial sums are controlled by the seminorms defining the topology, extending the normed space case where a single norm suffices. In such spaces, absolute convergence implies ordinary convergence under additional completeness assumptions, but the converse generally fails. Fréchet spaces, which are complete metrizable locally convex spaces defined by a countable family of seminorms, provide concrete examples. For instance, the space of smooth functions with compact support on \mathbb{R}^n, equipped with seminorms involving supremum norms of derivatives, admits absolutely convergent series that represent elements densely. However, counterexamples exist where a series converges ordinarily but not absolutely; in non-normable Fréchet spaces like the space of entire analytic functions, rearrangements of convergent series may diverge, highlighting that absolute convergence is stricter and fails while ordinary convergence holds. The concept relates to barrelled spaces, where every absorbing balanced closed convex set (barrel) is a neighborhood of the origin. In barrelled locally convex spaces, convergence tests for series leverage this property: the uniform boundedness principle applies to families of functionals evaluating partial sums, ensuring that absolute convergence via seminorms implies equicontinuity and thus controlled behavior in absorbing sets. This facilitates proofs of completeness and duality results without norms.

Rearrangement and Unconditional Convergence

For series of numbers

In contrast to conditionally convergent series, where the Riemann rearrangement theorem establishes that rearrangements can yield any real number as the sum or cause divergence, absolutely convergent series of real or complex numbers maintain the same sum under any rearrangement of terms. This invariance arises because the convergence of ∑ |a_n| ensures that the order of summation does not affect the limit, a property absent in conditional convergence. The proof of this invariance proceeds by comparing partial sums. For an absolutely convergent series ∑ a_n with sum S and a rearrangement ∑ b_k = ∑ a_{σ(k)}, the partial sums T_N = ∑{k=1}^N b_k satisfy |T_N - S| ≤ ∑{k=N+1}^∞ |b_k|, which is the tail of ∑ |a_n| and thus tends to 0 as N → ∞ by the convergence of the absolute series. Similarly, the difference between partial sums of the original and rearranged series is bounded by the tail of the absolute series, confirming that all rearrangements converge to S. For series of real or complex numbers, absolute convergence is equivalent to unconditional convergence, the condition that every rearrangement converges to the original sum. This equivalence holds in finite-dimensional spaces like ℝ or ℂ, distinguishing them from higher-dimensional settings. A representative example is the p-series ∑_{n=1}^∞ 1/n^p for p > 1, which converges absolutely since ∑ 1/n^p converges by the p-series test. Consequently, any permutation of its terms sums to the same value, ζ(p), the Riemann zeta function at p.

In general spaces

In normed linear spaces, a series \sum x_n of vectors is said to be unconditionally convergent if every rearrangement of the terms converges to the same in the norm topology. This notion serves as the appropriate analogue to absolute convergence when considering rearrangements, extending the scalar case where unconditional convergence coincides with absolute convergence. In complete normed linear spaces (Banach spaces), absolute convergence of \sum x_n, defined by \sum \|x_n\| < \infty, implies unconditional convergence, as the partial sums of any rearrangement form a Cauchy sequence in the norm and thus converge. However, the converse fails in infinite-dimensional spaces: unconditional convergence does not imply absolute convergence. Specifically, in every infinite-dimensional , there exist unconditionally convergent series that are not absolutely convergent, as established by the . This equivalence holds only for finite-dimensional spaces. In Hilbert spaces, a canonical example of unconditional convergence arises from expansions with respect to an orthonormal basis (e_n). For any x in the space, the series \sum \langle x, e_n \rangle e_n converges unconditionally to x. Such series highlight the robustness of Fourier representations in Hilbert spaces, where the convergence is independent of the ordering of terms. The space c_0 of real sequences converging to zero under the sup norm provides a counterexample illustrating the distinction: as an infinite-dimensional Banach space, it admits unconditionally convergent series that diverge absolutely, consistent with the Dvoretzky–Rogers result. Explicit constructions, such as those involving blocks of basis vectors with controlled norms, demonstrate series where rearrangements converge but \sum \|x_n\| = \infty. Unconditional convergence plays a key role in the theory of Schauder bases for Banach spaces. A Schauder basis (e_n) is unconditional if, for every x with unique coefficients (a_n(x)), the series \sum a_n(x) e_n converges unconditionally. Such bases require a form of absolute-like summability in their expansions, ensuring stability under permutations, and are exemplified by the standard basis in \ell_p spaces ($1 \leq p < \infty) and orthonormal bases in Hilbert spaces. In incomplete normed spaces, even absolutely convergent series may fail to converge due to lack of completeness; here, unconditional convergence remains tied to rearrangement properties but without the equivalence to absolute convergence, and absolute convergence implies only that partial sums are unconditionally Cauchy.

Key theorems and proofs

A fundamental result in the theory of infinite series states that if a series of complex numbers \sum a_n converges absolutely, then every rearrangement \sum a_{\pi(n)} of the series also converges absolutely to the same sum s = \sum a_n. To prove this, let \sum |a_n| converge, so the tails satisfy \sum_{k=m+1}^\infty |a_k| \to 0 as m \to \infty. Let s_m = \sum_{k=1}^m a_k and s_n' = \sum_{k=1}^n a_{\pi(k)} be the partial sums of the rearrangement, where \pi is a permutation of the natural numbers. Fix \epsilon > 0 and choose m large enough so that |s - s_m| < \epsilon/2 and \sum_{k=m+1}^\infty |a_k| < \epsilon/2. Since \pi is bijective, there exists q \geq m such that the first m terms \{a_1, \dots, a_m\} are included in the first q terms of the rearrangement. For n \geq q, s_n' includes s_m plus the sum of some terms a_k with k > m (the included tail terms). Thus, |s_n' - s_m| \leq \sum_{\substack{k > m \\ \pi(k) \leq n}} |a_k| \leq \sum_{k=m+1}^\infty |a_k| < \frac{\epsilon}{2}. Therefore, |s_n' - s| \leq |s_n' - s_m| + |s_m - s| < \epsilon for all n \geq q, so s_n' \to s. Absolute convergence of the rearrangement follows similarly by applying the result to \sum |a_{\pi(n)}|, since the absolute series is unchanged by permutation. In Banach spaces, unconditional convergence of a series \sum x_n means that \sum \epsilon_n x_n converges for every choice of signs \epsilon_n = \pm 1, with all such series converging to the same sum. A key characterization is that \sum x_n converges unconditionally if and only if \sup \{ \|\sum_{k=1}^n \epsilon_k x_k\| : n \in \mathbb{N}, \epsilon_k = \pm 1 \} < \infty. This supremum bounds the norms of the signed partial sums uniformly. To see the connection to uniform boundedness, consider the family of linear operators T_{n,\epsilon}: X \to X defined by T_{n,\epsilon}(y) = \sum_{k=1}^n \epsilon_k \langle y, x_k^* \rangle x_k or more directly, the partial sum operators over sign choices. Since the series converges pointwise for each fixed signs (to the sum), by the uniform boundedness principle applied to these operators on the Banach space, their norms are uniformly bounded, yielding the finite supremum above. The Dvoretzky–Rogers theorem extends this by showing that in any infinite-dimensional Banach space, unconditional convergence does not imply absolute convergence (\sum \|x_n\| < \infty), but if \sum x_n converges unconditionally, then there exists an equivalent norm \|\cdot\|' on the space such that \sum \|x_n\|' < \infty, making the series absolutely convergent in the new norm. This norm can be taken as \|y\|' = \sup \{ \|\sum_{i \in F} \epsilon_i y_i\| : F \subset \mathbb{N} \text{ finite}, \epsilon_i = \pm 1 \}, where the supremum is finite precisely when the combinations are unconditionally bounded. The theorem highlights that while absolute and unconditional convergence coincide in finite-dimensional spaces, they differ fundamentally in infinite dimensions.

Further Extensions

Products of infinite series

The Cauchy product of two infinite series \sum_{n=0}^\infty a_n and \sum_{n=0}^\infty b_n is defined as the series \sum_{n=0}^\infty c_n, where c_n = \sum_{k=0}^n a_k b_{n-k}. This construction formalizes the multiplication of or corresponding to the original series. When both series converge absolutely, their Cauchy product also converges absolutely, and the sum equals the product of the individual sums. To see the absolute convergence, note that |c_n| \leq \sum_{k=0}^n |a_k| |b_{n-k}| \leq \left( \sum_{m=0}^\infty |a_m| \right) \left( \sum_{m=0}^\infty |b_m| \right). Summing over n, the double sum \sum_{n=0}^\infty \sum_{k=0}^n |a_k| |b_{n-k}| = \left( \sum_{m=0}^\infty |a_m| \right) \left( \sum_{m=0}^\infty |b_m| \right) < \infty by for nonnegative terms, implying \sum |c_n| < \infty. Mertens' theorem provides a related result for convergence (not necessarily absolute): if \sum a_n converges absolutely to A and \sum b_n converges (possibly conditionally) to B, then \sum c_n converges to AB. When both series converge absolutely, this specializes to absolute convergence of the product to AB. If both series converge only conditionally, the Cauchy product may diverge. A standard counterexample is the series \sum_{n=0}^\infty \frac{(-1)^n}{\sqrt{n+1}}, which converges conditionally by the but whose Cauchy product with itself has general term c_n = (-1)^n \sum_{k=0}^n \frac{1}{\sqrt{(k+1)(n-k+1)}}. The inner sum is asymptotically \pi as n \to \infty, so |c_n| \not\to 0 and \sum c_n diverges. Even when one series converges absolutely and the other conditionally, the Cauchy product converges (by ) but need not converge absolutely, as the bound on |c_n| fails due to the divergence of the absolute series for the conditional one.

Absolute convergence over arbitrary sets

In the context of series indexed by an arbitrary set I, a family of scalars (a_i)_{i \in I} is said to converge absolutely if the sum \sum_{i \in I} |a_i| is finite, where the sum over an arbitrary index set is defined as the supremum of sums over all finite subsets of I. This definition extends the standard notion for countable indices by relying on the net of finite partial sums, ensuring that absolute convergence implies the existence of a well-defined sum value. When I is countable, absolute convergence guarantees that the sum is independent of any enumeration of the indices, meaning rearrangements yield the same total. For uncountable I, however, absolute convergence can only occur if at most countably many terms a_i are non-zero, as an uncountable collection of positive terms would yield an infinite sum. This restriction arises because the supremum over finite subsets would otherwise diverge, reducing the uncountable case to an effective countable summation. A concrete example is the double series \sum_{m,n} a_{m,n} over \mathbb{N} \times \mathbb{N}, which converges absolutely if \sum_{m,n} |a_{m,n}| < \infty, allowing the sum to be computed via any enumeration of the pairs (m,n). For non-negative terms a_i \geq 0, absolute convergence (or finiteness of the sum) links directly to the for series, permitting iteration over subsets without regard to order: for instance, \sum_{(m,n) \in M \times N} a_{m,n} = \sum_{m \in M} \sum_{n \in N} a_{m,n} = \sum_{n \in N} \sum_{m \in M} a_{m,n} holds for arbitrary sets M and N, provided the total sum is finite, which again implies at most countably many positive terms.

For integrals

In the context of improper integrals, absolute convergence refers to the condition that the integral of the absolute value of the function is finite. Specifically, for a function f continuous on [a, b) and the improper integral \int_a^\infty f(x) \, dx, the integral converges absolutely if \int_a^\infty |f(x)| \, dx < \infty. This implies the convergence of \int_a^\infty f(x) \, dx itself, as the absolute integrability dominates any oscillatory behavior that might allow conditional convergence. Absolute convergence has significant implications for multidimensional integrals, ensuring that the value of the integral is independent of the order of integration. For instance, Fubini's theorem requires absolute integrability over the domain to justify interchanging the order of iterated integrals; without it, the iterated integrals may differ or fail to exist even if one converges. This property aligns improper integrals with broader measure-theoretic frameworks, where absolute convergence guarantees robustness under rearrangements or slicing of the domain. A representative example of absolute convergence is the p-integral \int_1^\infty x^{-p} \, dx, which converges if and only if p > 1, as the antiderivative evaluation yields \lim_{b \to \infty} \frac{b^{1-p}}{1-p} - \frac{1}{1-p} finite precisely when the limit term vanishes. In contrast, conditional convergence occurs for integrals like \int_0^1 \frac{\sin(1/x)}{x} \, dx, which converges via the substitution u = 1/x reducing it to the Dirichlet integral \int_1^\infty \frac{\sin u}{u} \, du < \infty, but fails absolute convergence since \int_0^1 \left| \frac{\sin(1/x)}{x} \right| \, dx diverges, behaving asymptotically like \int_0^1 \frac{1}{x} \, dx near zero due to the bounded oscillation of \sin(1/x). Regarding integration theories, absolute convergence bridges Riemann and Lebesgue integrals effectively. In Riemann integration, improper integrals converge absolutely only if the limit of the absolute integral exists finitely, but Lebesgue integration defines integrability via f \in L^1 precisely when \int |f| \, d\mu < \infty, accommodating a wider class of functions while preserving the absolute convergence criterion as the foundation for L^1 spaces. This alignment ensures that absolutely convergent Riemann improper integrals are Lebesgue integrable, enhancing analytical tools for discontinuous or unbounded functions.

References

  1. [1]
    Calculus II - Absolute Convergence - Pauls Online Math Notes
    Nov 16, 2022 · Series that are absolutely convergent are guaranteed to be convergent. However, series that are convergent may or may not be absolutely ...
  2. [2]
    Absolute Convergence -- from Wolfram MathWorld
    If a series is absolutely convergent, then the sum is independent of the order in which terms are summed. Furthermore, if the series is multiplied by another ...Missing: definition | Show results with:definition
  3. [3]
    Absolute and Conditional Convergence
    ### Summary of Absolute Convergence Content
  4. [4]
    3.4 Absolute and Conditional Convergence
    If the series ∑ n = 1 ∞ | a n | converges then the series ∑ n = 1 ∞ a n also converges. That is, absolute convergence implies convergence. 🔗. Recall that some ...
  5. [5]
    [PDF] 18.100C Real Analysis: Lecture 10 Summary - MIT OpenCourseWare
    Absolute convergence of series (of real or complex numbers). Theorem 10.2. Absolute convergence implies convergence. Theorem 10.3. Suppose that. Then, for ...
  6. [6]
    p-Series - Oregon State University
    A p-series is a series of the form sum over n from 1 to infinity of 1/(n^p), convergent if p > 1, divergent otherwise.
  7. [7]
    [PDF] Series Convergence Tests Math 122 Calculus III
    By use of the integral test, you can determine which p-series converge. Theorem 7 (p-series). A p-series X 1 np converges if and only if p > 1. Proof.
  8. [8]
    [PDF] Alternating Series, Absolute Convergence and Conditional ... - Math
    Conditional Convergence is conditionally convergent if converges but does not. EX 5 Classify as absolutely convergent, conditionally convergent or divergent.
  9. [9]
    11.4 Absolute Convergence
    11.39 Definition (Absolute Convergence.) Let be a complex sequence. We say that is absolutely summable if and only if is summable; i.e., if and only if ...
  10. [10]
    Complex Number Primer - Pauls Online Math Notes
    The modulus of a complex number is always a real number and in fact it will never be negative since square roots always return a positive number or zero.
  11. [11]
    [PDF] Section 5.56. Convergence of Series
    Jan 29, 2020 · The absolute convergence of a series of complex numbers im- plies the convergence of that series. Definition. Let P. ∞ n=1 zn be a series ...
  12. [12]
    Cauchy's Calculus - MacTutor History of Mathematics
    Cauchy wrote Cours d'Analyse (1821) based on his lecture course at the École Polytechnique. In it he attempted to make calculus rigorous.
  13. [13]
    Series with Commuting Terms in Topologized Semigroups - MDPI
    In 1827, Peter Lejeune-Dirichlet was the first to notice that it is possible to rearrange the terms of certain convergent series of real numbers so that the ...
  14. [14]
    [PDF] THE CLASSICAL THEORY OF REARRANGEMENTS - ScholarWorks
    A series ∑ fn is conditionally convergent if and only if for each real number α, there is a rearrangement of ∑ fn that converges to α. In Chapter 3, we analyze ...Missing: dissertation | Show results with:dissertation
  15. [15]
    Karl Weierstrass (1815 - 1897) - Biography - MacTutor
    Known as the father of modern analysis, Weierstrass devised tests for the convergence of series and contributed to the theory of periodic functions ...
  16. [16]
    Stefan Banach (1892 - 1945) - Biography - MacTutor
    Banach founded modern functional analysis and made major contributions to the theory of topological vector spaces. In addition, he contributed to measure theory ...Missing: absolute | Show results with:absolute
  17. [17]
    3.4: Absolute and Conditional Convergence - Mathematics LibreTexts
    Jan 20, 2022 · If the series ∑ n = 1 ∞ | a n | converges then the series ∑ n = 1 ∞ a n also converges. That is, absolute convergence implies convergence.Missing: ordinary | Show results with:ordinary
  18. [18]
    8.5: Alternating Series and Absolute Convergence
    Dec 28, 2020 · The theorem states that the terms of an absolutely convergent series can be rearranged in any way without affecting the sum. theorem 72: ...Missing: ordinary | Show results with:ordinary
  19. [19]
    Weierstrass M-Test -- from Wolfram MathWorld
    x in E , then the series exhibits absolute convergence for each x in E as well as uniform convergence in E . See also. Absolute Convergence, Uniform Convergence ...
  20. [20]
    None
    ### Theorem and Proof: Absolute Convergence Implies Convergence in Banach Spaces
  21. [21]
    [PDF] Chapter 1: Metric and Normed Spaces - UC Davis Mathematics
    A useful property of an absolutely convergent series of real (or complex) numbers is that any series obtained from it by a permutation of its terms converges to ...
  22. [22]
    [PDF] Lecture 4. Series in Normed Vector Spaces - UCSD Math
    May 9, 2010 · First define what we mean by convergence of a series in a normed vector space; i.e., convergence of the sequence of partial sums. Definition 1 ...
  23. [23]
    [PDF] Banach spaces
    Def A series 5xx is absolutely convergent if all coo. THM (Completeness criterion). A normed space X is a Banach space iff every absolutely convergent series in ...
  24. [24]
    Topological Vector Spaces
    **Summary of Schaefer's Topological Vector Spaces (Sections on Series, Absolute Convergence, Summability)**
  25. [25]
    A study on absolutely convergent series in locally convex spaces
    The concept of absolute convergence for series, a previously simple concept in normed spaces, was generalized to locally convex spaces.
  26. [26]
    [PDF] Calculus 101 - GitHub Pages
    Prove this by using the Cauchy criteria: show that if the partial sums of Pk |ak| are Cauchy, then so are the partial sums of Pk ak. Then, rearrangement works ...
  27. [27]
    [PDF] Notes on Unordered Sums - UC Davis Math
    Jan 13, 2007 · Thus, absolute convergence and unconditional convergence are equivalent in finite-dimensional Banach spaces. In infinite-dimensional Banach ...
  28. [28]
    Absolute and Unconditional Convergence in Normed Linear Spaces
    is called absolutely convergent if E Ix,j < 00; it is called unconditionally convergent if the series E y, converges whenever the sequence (y,) is a.<|separator|>
  29. [29]
    unconditional convergence - PlanetMath.org
    Mar 22, 2013 · When X=Rn X = ℝ n then by a famous theorem of Riemann (∑xn) ( ∑ x n ) is unconditionally convergent if and only if it is absolutely convergent.
  30. [30]
    Absolute and Unconditional Convergence in Normed Linear Spaces
    Oct 24, 2008 · In 1933 Orlicz proved various results concerning unconditional convergence in Banach spaces (4), which were noted by Banach ((l), p.
  31. [31]
    [PDF] Let H be a Hilbert space and E an orthonormal set in H, ie ...
    For each x Є H, x = ΣaЄA < X, Ua > ua and this series is unconditionally convergent. ... Theorem: Every nonzero Hilbert space has an orthonormal basis.<|separator|>
  32. [32]
    [PDF] Chapter 6: Hilbert Spaces - UC Davis Math
    By the definition of convergence, there is a finite set3:?B such that è <> • xè. B. ٌ,ô2" for all finite setsW: that containX:?B . IfYS is any finite subset ...
  33. [33]
    [PDF] Schauder basis
    Dec 4, 2012 · The standard bases of the sequence spaces and for 1 ≤ p < ∞, as well as every orthonormal basis in a. Hilbert space, are unconditional. The ...
  34. [34]
    [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 ...
  35. [35]
    [PDF] Introduction to Bases in Banach Spaces - Matthew Daws
    Jun 5, 2005 · Firstly, we need to explore some properties of unconditional convergence in Banach spaces. Proposition 4.1. Let E be a Banach space, and let (xn) ...
  36. [36]
    [PDF] An Introduction to Measure Theory - Terry Tao
    Exercise 0.0.2 (Tonelli's theorem for series over arbitrary sets). Let. A, B be sets (possibly infinite or uncountable), and (xn,m)n∈A,m∈B be a doubly ...<|control11|><|separator|>
  37. [37]
    Sums of uncountably many real numbers [closed] - MathOverflow
    May 10, 2011 · The sum of uncountably-many non-negative real numbers is finite only if all but countably many of those real numbers are 0.
  38. [38]
    [PDF] Overview of Improper Integrals MAT 104 - Math (Princeton)
    Absolute convergence test: If R |f(x)|dx converges, then R f(x)dx converges as well. Note that this test is only useful for showing convergence; it's often used ...
  39. [39]
    [PDF] Improper Integrals
    Theorem 2 (Absolute convergence implies convergence.). If the improper integral (1) con- verges absolutely then it converges. Proof. We make use of the Cauchy ...
  40. [40]
    [PDF] Real Analysis MAA 6616 Lecture 18 Tonelli's Theorem and ...
    Fubini's Theorem asserts that the integral of an integrable function in Rr+s is the same as the iterated integrals and interchange of order of integration ...
  41. [41]
    Calculus II - Improper Integrals - Pauls Online Math Notes
    Nov 16, 2022 · We will call these integrals convergent if the associated limit exists and is a finite number (i.e. it's not plus or minus infinity) and ...
  42. [42]
    [PDF] 11. Absolute and conditional convergence of improper integrals
    It is clear that conditional convergence may hold only for integrals whose integrands change sign. Example. Consider the integral /. +∞. 1 sin x x.<|separator|>
  43. [43]
    [PDF] The Riemann Integral - UC Davis Math
    The Lebesgue integral allows one to integrate unbounded or highly discontinuous functions whose Riemann integrals do not exist, and it has better mathematical ...
  44. [44]
    [PDF] The Lebesgue integral
    Note that the 'improper integral' without the absolute value can converge without u being Lebesgue integrable. ... Lebesgue's Dominated Convergence, |f|2 ∈ L1(R).