Fact-checked by Grok 2 weeks ago

Limit of a sequence

In mathematics, the limit of a sequence is a foundational concept in that describes the value L toward which the terms of an infinite sequence \{a_n\} of real numbers approach as the index n tends to , provided such a value exists. Informally, the sequence converges to L if, for all sufficiently large n, the terms a_n can be made arbitrarily close to L. Formally, \lim_{n \to \infty} a_n = L if for every \varepsilon > 0, there exists a positive N such that |a_n - L| < \varepsilon whenever n > N. This \varepsilon-N definition ensures that the tail of the sequence lies within any given neighborhood of L, capturing the intuitive notion of eventual proximity. If no such finite L exists, the sequence may diverge to \pm \infty or oscillate without converging; for instance, a_n = n diverges to +\infty, while a_n = (-1)^n fails to converge due to perpetual alternation between -1 and 1. A convergent sequence has a unique limit, a property proven using the triangle inequality in the real numbers. Limits of sequences underpin key theorems in analysis, such as algebraic properties allowing operations like \lim_{n \to \infty} (a_n + b_n) = \lim_{n \to \infty} a_n + \lim_{n \to \infty} b_n for convergent sequences \{a_n\} and \{b_n\}, and the squeeze theorem, which states that if a_n \leq c_n \leq b_n for large n and both \{a_n\} and \{b_n\} converge to L, then so does \{c_n\}. This concept is among the most subtle and essential in , serving as the basis for defining of functions, , integrals, and more advanced structures like metric spaces and . Sequences without limits, or those diverging in specific ways, are crucial for studying series convergence and asymptotic behavior in applied fields like physics and .

Historical Development

Early Intuitive Notions

The concept of limits in sequences emerged intuitively in through paradoxes that challenged notions of motion and infinity. , around the 5th century BCE, posed the paradox of Achilles and the , where the swift Achilles appears unable to overtake a slower due to an infinite series of ever-diminishing intervals that he must traverse. This puzzle intuitively suggested that infinite processes could converge to a finite outcome, foreshadowing the idea of a without providing a . In the , of Syracuse (c. 287–212 BCE) advanced these ideas through the , a technique for approximating areas by inscribing and circumscribing polygons that increasingly approached the curved boundary. In his treatise , demonstrated that the area of a parabolic segment equals four-thirds the area of the inscribed triangle by iteratively adding triangles whose areas summed in a , effectively bounding the region between lower and upper limits that to the exact value. This approach relied on the principle that if two quantities could be made arbitrarily close without equaling, one must be equal to the other, providing an early rigorous yet intuitive handling of convergence. Medieval and Renaissance mathematics further explored infinite processes, particularly in Indian traditions. While (476–550 CE) contributed rational approximations, such as π ≈ 3.1416 derived from circumference-to-diameter ratios, the Kerala school in the 14th–16th centuries developed infinite series expansions for and π, like the series for arctangent that used to compute precise values through partial sums approaching a . These methods echoed exhaustion by summing infinitely many terms to approximate transcendental quantities. In , Renaissance scholars revisited Archimedean techniques, applying them to volumes and areas in preparation for . By the , intuitive notions of limits underpinned the invention of . Isaac developed fluxions around 1665–1666, treating quantities as flowing variables whose instantaneous rates of change—moments or infinitesimally small increments—approximated tangents and areas through limiting processes. Independently, Gottfried Wilhelm formulated infinitesimals in the 1670s as "inassignable" quantities smaller than any given positive number yet non-zero, using them to derive rules for and as ratios of these evanescent differences. These precursors treated limits as the outcome of infinite approximations in continuous change, setting the stage for 19th-century formalization.

Formalization in the 19th Century

The formalization of the concept for sequences in the emerged as a direct response to longstanding philosophical critiques of infinitesimal methods in , particularly George Berkeley's 1734 attack in , where he derided as "ghosts of departed quantities" lacking logical foundation. This prompted mathematicians to develop rigorous, non-infinitesimal definitions grounded in inequalities, transforming intuitive notions from ancient paradoxes—such as Zeno's—into precise analytical tools. By mid-century, these efforts established the epsilon-based framework that underpins modern . Bernard Bolzano laid early groundwork in his 1817 pamphlet Rein analytischer Beweis des Lehrsatzes, daß zwischen je zwey Werthen, die ein entgegengesetztes Resultat gewähren, wenigstens eine reelle Wurzel der Gleichung vorhanden sey. While primarily proving the for , Bolzano introduced a definition of that implicitly relied on concepts: a is continuous if, for points sufficiently close, the difference in values can be made arbitrarily small. This approach used the theorem to ensure the existence of points in infinite sets, bridging geometric intuition to algebraic precision without invoking infinitesimals. Augustin-Louis Cauchy advanced this rigor in his 1821 textbook Cours d'analyse de l'École Polytechnique, where he provided the first systematic definition of the limit of a . Cauchy stated: "When the successive values attributed to the same variable indefinitely approach a fixed value, so as to end by differing from it by as little as one wishes, this last is called the limit of all the others." For proofs, he operationalized this with an condition, paraphrased as: a converges to L if for every \epsilon > 0, there exists a N such that for all n > N, |a_n - L| < \epsilon. This formulation, applied extensively to series and functions, eliminated reliance on fluxions and established limits as the cornerstone of calculus. Karl Weierstrass further refined these ideas in his Berlin University lectures beginning in the 1850s, culminating in a fully epsilon-N formalization by 1861 that dispelled any residual ambiguity. He defined the limit of a sequence p_n as L if, for every \epsilon > 0, there exists an N such that for all n > N, |p_n - L| < \epsilon, emphasizing arithmetic verification over geometric intuition. Delivered to students like Hermann Amandus Schwarz, these lectures—later disseminated through notes—ensured the epsilon method's adoption, purging infinitesimals entirely and solidifying sequence limits as a discrete, verifiable process.

Limits over the Real Numbers

Formal Definition

A sequence of real numbers is a function a: \mathbb{N} \to \mathbb{R}, where \mathbb{N} denotes the set of positive integers, often denoted as \{a_n\}_{n=1}^\infty or simply \{a_n\}. The real numbers \mathbb{R} are equipped with the standard metric given by the absolute value |x - y| for x, y \in \mathbb{R}, which measures the distance between points. The formal definition of the limit of a sequence in \mathbb{R}, known as the \varepsilon-N definition, is as follows: A sequence \{a_n\} converges to a limit L \in \mathbb{R} if for every \varepsilon > 0, there exists N \in \mathbb{N} such that for all n > N, |a_n - L| < \varepsilon. \lim_{n \to \infty} a_n = L \iff \forall \varepsilon > 0 \, \exists N \in \mathbb{N} \, \forall n > N, \, |a_n - L| < \varepsilon. This definition was introduced by in 1821 and rigorously formalized by in the mid-19th century. Common notations for this convergence include \lim_{n \to \infty} a_n = L or a_n \to L. If a limit exists, it is unique. To see this, suppose \lim_{n \to \infty} a_n = L and \lim_{n \to \infty} a_n = M with L \neq M. Let \varepsilon = |L - M|/2 > 0. Then there exists N_1 \in \mathbb{N} such that for all n > N_1, |a_n - L| < \varepsilon, and N_2 \in \mathbb{N} such that for all n > N_2, |a_n - M| < \varepsilon. For n > \max(N_1, N_2), the yields |L - M| \leq |L - a_n| + |a_n - M| < 2\varepsilon = |L - M|, a contradiction. Thus, L = M. A constant sequence \{a_n\} where a_n = c \in \mathbb{R} for all n converges to c, since |a_n - c| = 0 < \varepsilon holds for any \varepsilon > 0 and any N \in \mathbb{N} (e.g., N = 1).

Illustrative Examples

To illustrate the concept of limits for sequences in the real numbers, consider the sequence defined by a_n = \frac{1}{n} for n \in \mathbb{N}. This sequence converges to 0, as for any \epsilon > 0, choosing N = \lceil 1/\epsilon \rceil ensures that for all n > N, |a_n - 0| = \frac{1}{n} < \epsilon. The following table shows the first ten terms of this sequence, demonstrating its approach to 0:
na_n = 1/n
11.000
20.500
30.333
40.250
50.200
60.167
70.143
80.125
90.111
100.100
As n increases, the terms decrease monotonically toward 0. In contrast, the sequence a_n = n diverges to +\infty, meaning that for every M > 0, there exists N \in \mathbb{N} such that for all n > N, a_n > M. For instance, selecting N = \lfloor M \rfloor + 1 satisfies this condition. Another example is the sequence a_n = (-1)^n, which alternates between -1 and 1 and does not converge to any , as the terms fail to get arbitrarily close to any single limit point. Finally, the sequence a_n = \sin n is bounded between -1 and 1 but does not converge, because its terms are dense in the interval [-1, 1], visiting every subinterval infinitely often due to the irrationality of \pi.

Fundamental Properties

The algebra of limits for sequences in the real numbers allows for the manipulation of convergent sequences using field operations. Specifically, if \{a_n\} and \{b_n\} are sequences converging to limits A and B respectively, and c \in \mathbb{R} is a constant, then the sequence \{c a_n + b_n\} converges to cA + B. Similarly, the product sequence \{a_n b_n\} converges to AB, and if B \neq 0, the quotient sequence \{a_n / b_n\} converges to A/B. A brief sketch of the proof for the sum rule relies on the : for any \epsilon > 0, there exist N_1, N_2 such that for n > \max(N_1, N_2), |a_n - A| < \epsilon/2 and |b_n - B| < \epsilon/2, so |(a_n + b_n) - (A + B)| \leq |a_n - A| + |b_n - B| < \epsilon. Limits of sequences also preserve order relations when they exist. If \{a_n\} and \{b_n\} converge to A and B, and a_n \leq b_n for all sufficiently large n, then A \leq B. This follows from the contrapositive: assuming A > B leads to a via the definition of and the of the reals. The provides a powerful tool for establishing indirectly. If \{g_n\}, \{f_n\}, and \{h_n\} are sequences such that g_n \leq f_n \leq h_n for all sufficiently large n, and both \{g_n\} and \{h_n\} converge to the same L, then \{f_n\} also converges to L. The proof proceeds by showing that for any \epsilon > 0, the inequalities imply |f_n - L| \leq \max(|g_n - L|, |h_n - L|) < \epsilon for large n. Monotone sequences exhibit particularly nice convergence behavior in the reals. A sequence \{a_n\} is monotone if it is either non-decreasing (a_{n+1} \geq a_n for all n) or non-increasing (a_{n+1} \leq a_n for all n). The monotone convergence theorem states that every bounded monotone sequence converges: if non-decreasing and bounded above, it converges to its supremum; if non-increasing and bounded below, to its infimum./02%3A_Sequences/2.03%3A_Monotone_Sequences) This result is closely tied to the Bolzano-Weierstrass theorem, which guarantees that every bounded sequence has a convergent subsequence, allowing the limit of a monotone sequence to be identified as the least upper bound of its range.

Infinite and Oscillatory Limits

In the context of sequences of real numbers, a sequence \{a_n\} is said to diverge to +\infty if, for every M > 0, there exists N \in \mathbb{N} such that a_n > M for all n > N. Similarly, the sequence diverges to -\infty if, for every M > 0, there exists N \in \mathbb{N} such that a_n < -M for all n > N. This extends the notion of beyond finite limits, capturing unbounded growth in a precise manner. A classic example is the sequence a_n = n^2, which diverges to +\infty since the growth ensures terms exceed any positive bound for sufficiently large n. Another is a_n = -n, diverging to -\infty. These cases illustrate how sequences can "tend to infinity" without converging to a . Oscillatory divergence occurs when a sequence fails to converge to any real limit L, meaning there exists some such that for every N \in \mathbb{N}, the \epsilon-neighborhood of L does not contain all but finitely many terms of the sequence./02:_Sequences/2.01:_Convergence) For instance, the sequence a_n = \sin(n\pi/2) oscillates between -1, 0, and 1 indefinitely, preventing convergence to any single value. A more dramatic case is a_n = n \sin(n), which oscillates with increasing amplitude and thus without bound. One-sided limits for sequences are less common than for functions but arise in the context of , such as even or indices. If the of even terms converges to one value and the terms to another, the full sequence oscillates and diverges. For example, in a_n = (-1)^n, the even terms are constantly while terms are -, yielding distinct "one-sided" behaviors along these . All convergent sequences are bounded, meaning there exists M > 0 such that |a_n| \leq M for all n./02:_Sequences/2.01:_Convergence) To see this, if \{a_n\} converges to L, choose \epsilon = 1; then for n > N, |a_n| < |L| + 1, and bounding the finite initial terms yields an overall bound. However, the converse fails: bounded sequences like \sin(n\pi/2) need not converge./02:_Sequences/2.01:_Convergence) In oscillatory cases, the squeeze theorem can sometimes bound terms to show divergence, as with alternating sequences trapped between constants.

Generalizations to Metric Spaces

Definition in Metric Spaces

In a metric space (X, d), where X is a set and d: X \times X \to [0, \infty) is a metric satisfying the usual axioms of non-negativity, symmetry, and the triangle inequality, the notion of convergence for a sequence generalizes the real-number case. Specifically, the real line \mathbb{R} equipped with the standard metric d(a, b) = |a - b| serves as a prototypical example. A sequence \{x_n\}_{n=1}^\infty in X converges to a limit x \in X, denoted \lim_{n \to \infty} x_n = x or x_n \to x, if for every \epsilon > 0, there exists a positive integer N such that d(x_n, x) < \epsilon for all n > N. This definition captures the intuitive idea that the terms x_n get arbitrarily close to x as n increases, measured via the metric d. The condition d(x_n, x) < \epsilon is equivalent to stating that the sequence terms eventually lie inside the open ball centered at the limit point, defined as B(x, \epsilon) = \{ y \in X \mid d(y, x) < \epsilon \}. These open balls form the basic neighborhoods in the topology induced by the metric, providing a geometric foundation for convergence without relying on any additional structure beyond the distance function. In contrast to the real numbers, where the total order enables properties such as the monotone convergence theorem for bounded increasing sequences, general metric spaces lack a canonical ordering; thus, monotonicity must be defined metrically (e.g., via additive distances along the sequence), and convergence depends purely on the metric-induced topology rather than order-based supremum or infimum principles. A concrete illustration occurs in the Euclidean space \mathbb{R}^m with the standard Euclidean metric d(\mathbf{u}, \mathbf{v}) = \sqrt{\sum_{k=1}^m (u_k - v_k)^2}. Consider the sequence \mathbf{x}_n = \left( \frac{1}{n}, \frac{1}{n^2}, \dots, \frac{1}{n^m} \right), which approaches the origin \mathbf{0} = (0, \dots, 0). For any \epsilon > 0, choose N > m^{1/2} / \epsilon; then for n > N, d(\mathbf{x}_n, \mathbf{0}) = \sqrt{\sum_{k=1}^m \frac{1}{n^{2k}}} < \sqrt{m / n^2} < \epsilon. This componentwise decay ensures convergence in the metric sense. The limit, when it exists, is unique in any metric space. Suppose x_n \to x and x_n \to z; then for any \epsilon > 0, there exists N such that d(x_n, x) < \epsilon/2 and d(x_n, z) < \epsilon/2 for n > N. By the triangle inequality, d(x, z) \leq d(x, x_n) + d(x_n, z) < \epsilon, so d(x, z) = 0 and x = z. This uniqueness holds regardless of the specific metric space structure.

Key Properties

In metric spaces, a function f: X \to Y between metric spaces is continuous at a point x \in X if and only if, whenever a sequence (x_n) in X converges to x, the image sequence (f(x_n)) in Y converges to f(x). This sequential characterization of continuity holds because the \epsilon-\delta definition of continuity implies preservation of sequential limits, and conversely, sequential continuity implies the \epsilon-\delta condition via the properties of balls in metric spaces. For sequences of functions (f_n) from a metric space X to a metric space Y, uniform convergence to a limit function f: X \to Y occurs if \sup_{x \in X} d_Y(f_n(x), f(x)) \to 0 as n \to \infty. This is stronger than pointwise convergence, where f_n(x) \to f(x) for each fixed x \in X, and implies that the convergence is uniform across the entire domain, preserving properties like boundedness of the functions. In subsets of the real line \mathbb{R} with the standard metric, the Heine-Borel theorem states that a set is compact if and only if it is closed and bounded, and this compactness is equivalent to sequential compactness: every sequence in the set has a convergent subsequence with limit in the set. For example, the closed interval [0, 1] is sequentially compact, as any sequence therein admits a convergent subsequence by the Bolzano-Weierstrass theorem. In a discrete metric space, where the metric d(x, y) = 1 if x \neq y and d(x, y) = 0 if x = y, a sequence converges if and only if it is eventually constant. This follows because, for \epsilon < 1, all terms beyond some index must equal the limit to satisfy d(x_n, L) < \epsilon. Isometries, or distance-preserving bijections between metric spaces, preserve limits of sequences because they are continuous maps (with Lipschitz constant 1) and thus map convergent sequences to convergent sequences with the corresponding limits. For instance, if (x_n) \to x in X, then d(f(x_n), f(x)) = d(x_n, x) \to 0, so (f(x_n)) \to f(x) in the codomain.

Connection to Cauchy Sequences

In metric spaces, a sequence \{x_n\} is defined as a Cauchy sequence if for every \epsilon > 0, there exists a positive N such that for all m, n > N, the d(x_m, x_n) < \epsilon. This condition captures the intuitive notion that the terms of the sequence become arbitrarily close to each other as n increases, independent of any specific limit point. A key connection between Cauchy sequences and limits arises in the property of completeness for metric spaces: a metric space is complete if every Cauchy sequence in it converges to a point within the space. For instance, the real numbers \mathbb{R} with the standard metric form a complete space, ensuring that every Cauchy sequence of real numbers converges to a real limit. In contrast, incomplete spaces like the rational numbers \mathbb{Q} with the standard metric contain Cauchy sequences that do not converge to any point in \mathbb{Q}. A classic example is the sequence of rational approximations to \sqrt{2}, such as x_1 = 1, x_2 = 1.4, x_3 = 1.41, x_4 = 1.414, and so on, obtained by truncating the decimal expansion of \sqrt{2}. This sequence is Cauchy in \mathbb{Q} because the terms get arbitrarily close, but its limit \sqrt{2} lies outside \mathbb{Q}, illustrating how incompleteness prevents convergence within the space. The completeness property underpins applications like the Banach fixed-point theorem, which guarantees that iterative sequences generated by contractions in complete metric spaces converge to a unique fixed point. In non-complete spaces, such limits may fail to exist, necessitating extensions like the completion of \mathbb{Q} to \mathbb{R} via equivalence classes of Cauchy sequences.

Limits in Topological Spaces

Definition via Neighborhoods

In a topological space (X, \tau), where \tau is the collection of open sets, a sequence (x_n)_{n \in \mathbb{N}} in X is said to converge to a point x \in X if for every open neighborhood U of x (that is, U \in \tau and x \in U), there exists a positive integer N such that x_n \in U for all n > N. This definition generalizes the notion, where open neighborhoods are induced by open balls, but applies to arbitrary topologies without requiring a metric. Not all topological spaces behave well with respect to sequential convergence; a space is called sequential if every sequentially open set (one with no sequence from its complement converging to a point inside it) is open. In sequential spaces, sequences suffice to characterize the , but general topological spaces may not be sequential—for instance, the countable complement on an fails to be sequential, as certain closed sets are not sequentially closed. A striking example occurs in the indiscrete topology on a set X with more than one point, where the only open sets are \emptyset and X; here, every sequence in X converges to every point in X, since the sole nontrivial neighborhood X contains all terms after any N. In Hausdorff spaces (also known as T_2 spaces), where any two distinct points admit disjoint open neighborhoods, limits of sequences, if they exist, are unique. This contrasts with non-Hausdorff spaces, where multiple limits are possible, as in the indiscrete case. While sequences capture convergence in first-countable spaces like metric spaces, they are insufficient in general topologies; nets (generalized sequences indexed by directed sets) or filters are needed to fully describe convergence, as every net converges to a point if and only if it converges along every subnet.

Preservation under Continuous Maps

In topological spaces, continuous maps preserve the convergence of sequences. Specifically, if f: X \to Y is a continuous function between topological spaces and a sequence (x_n) in X converges to x \in X, then the sequence (f(x_n)) in Y converges to f(x). This property holds because continuity ensures that the preimage under f of any neighborhood of f(x) is a neighborhood of x; thus, the tail of (x_n) lying in that preimage implies the tail of (f(x_n)) lies in the neighborhood of f(x). Homeomorphisms, being bijective continuous maps with continuous inverses, preserve sequential convergence in both directions. That is, for a homeomorphism f: X \to Y, a sequence (x_n) converges to x in X if and only if (f(x_n)) converges to f(x) in Y. This bidirectional preservation underscores the topological equivalence of the spaces involved. A concrete illustration arises with projection maps in product topologies. For topological spaces X and Y, the product space X \times Y is equipped with the product topology, in which the projection maps \pi_X: X \times Y \to X and \pi_Y: X \times Y \to Y are continuous. Consequently, a sequence ((x_n, y_n)) in X \times Y converges to (x, y) if and only if x_n \to x in X and y_n \to y in Y componentwise. Discontinuous functions fail to preserve sequential limits. For instance, consider the step function f: \mathbb{R} \to \mathbb{R} defined by f(t) = 0 if t \leq 0 and f(t) = 1 if t > 0, which is discontinuous at t = 0. The sequence t_n = 1/n converges to 0 in \mathbb{R}, but f(t_n) = 1 for all n, so (f(t_n)) converges to 1, not to f(0) = 0. While the preservation of sequential limits is a topological property relying on preimages of open sets, in metric subspaces uniform continuity provides stronger guarantees, such as preserving Cauchy sequences, though the core mechanism remains the continuity condition on preimages.

Sequential Compactness

In topological spaces, sequential compactness is defined as the property that every in the space has a convergent whose lies within the space. This notion captures a form of "finiteness" in terms of sequence behavior, generalizing the idea that limits of subsequences always exist internally, which is crucial for studying without relying on the full machinery of nets or filters. In metric spaces, sequential compactness is equivalent to compactness, meaning that a subset is sequentially compact if and only if every open cover has a finite subcover. This equivalence holds because metric spaces allow sequences to probe the space effectively, with convergent subsequences corresponding to the exhaustion of covers. Specifically, in Euclidean space \mathbb{R}^n equipped with the standard metric, the Heine-Borel theorem characterizes compact (and thus sequentially compact) subsets as precisely those that are closed and bounded. For instance, the closed interval [0,1] is sequentially compact, as any sequence therein possesses a convergent subsequence by the Bolzano-Weierstrass theorem, whereas the open interval (0,1) is not, since the sequence \{1/n\}_{n=1}^\infty has no convergent subsequence within (0,1), as its limit $0$ lies outside the space. Beyond metric spaces, sequential compactness and compactness diverge, illustrating that the equivalence is not a general topological phenomenon. A notable non-metric example arises in the on ordinals: the space [0, \omega_1), consisting of all countable ordinals with the order topology, is sequentially compact—every sequence, being countable, has a supremum that serves as the limit of a suitable subsequence—but it is not compact, as the open cover \{[0, \alpha) \mid \alpha < \omega_1\} admits no finite subcover. This highlights how sequential compactness can hold independently of compactness in spaces lacking countable bases or metric structure. Conversely, there exist compact spaces that fail to be sequentially compact, such as certain uncountable products of intervals, though ordinal examples underscore the subtleties in ordered topologies where sequential limits may not capture all accumulation behaviors. The Bolzano-Weierstrass theorem provides a key generalization tying sequential compactness to limit points in \mathbb{R}^n: every bounded infinite subset has at least one limit point, ensuring that infinite bounded sequences admit convergent subsequences within closed bounded sets. This property underpins the sequential compactness of closed and bounded subsets in Euclidean spaces, emphasizing the role of limits in identifying accumulation points without exhaustive enumeration.

Non-Standard Approaches

Hyperreal Number Systems

The hyperreal number system, denoted ^*\mathbb{R}, extends the real numbers \mathbb{R} to include infinitesimal quantities smaller than any positive real and infinite quantities larger than any real, forming a non-Archimedean ordered field. This construction proceeds via an ultrapower: sequences of real numbers \mathbb{R}^\mathbb{N} are quotiented by an equivalence relation defined using a free ultrafilter \mathcal{U} on the natural numbers \mathbb{N}, where two sequences (a_n) and (b_n) are equivalent if \{ n \in \mathbb{N} \mid a_n = b_n \} \in \mathcal{U}. The resulting hyperreals include the standard reals via the embedding that maps each real r to the constant sequence (r, r, \dots), and the hypernatural numbers ^*\mathbb{N} extend \mathbb{N} with infinite elements, such as those represented by sequences like (n)_{n \in \mathbb{N}}. In this framework, the limit of a sequence (a_n) in \mathbb{R} is defined using the standard part function \mathrm{st}: ^*\mathbb{R} \to \mathbb{R}, which assigns to each finite hyperreal its unique "shadow" or closest real number. Specifically, \lim_{n \to \infty} a_n = L if, for every infinite hypernatural \nu \in ^*\mathbb{N} \setminus \mathbb{N}, the extended sequence value a_\nu (obtained via the natural extension ^*a) satisfies a_\nu \approx L, meaning \mathrm{st}(a_\nu) = L, or equivalently, a_\nu - L is infinitesimal. The monad of a real L, consisting of all hyperreals infinitely close to L, provides the neighborhood structure for this convergence, allowing evaluation at "large" infinite indices without quantifying over all tails of the sequence. For example, consider the sequence a_n = 1/n. Its limit is 0 because, for any infinite hypernatural \nu, $1/\nu is a positive infinitesimal (smaller than $1/k for all standard natural k), so \mathrm{st}(1/\nu) = 0. This nonstandard approach offers advantages in intuition and proof simplicity, as it permits direct manipulation of infinitesimals and infinite quantities to establish convergence, avoiding the explicit quantification of the epsilon-delta definition in standard analysis.

Transfer Principle Applications

The transfer principle in nonstandard analysis asserts that any first-order logical statement that holds in the real numbers \mathbb{R} also holds in the hyperreal numbers ^*\mathbb{R} when the symbols and quantifiers are appropriately extended, enabling the rigorous importation of standard theorems into the nonstandard framework. This principle, formalized by , relies on the model-theoretic construction of ^*\mathbb{R} as an ultrapower of \mathbb{R}, ensuring that bounded quantifiers over reals correspond to internal sets in ^*\mathbb{R}. It applies specifically to first-order properties, such as those expressible without unbounded existential or universal quantifiers over sets, allowing direct proofs of limit theorems by evaluating expressions at infinite hypernatural indices or infinitesimal values. One key application is to the continuity of functions, where a function f: \mathbb{R} \to \mathbb{R} is continuous at a limit point L if and only if, for any hyperreal sequence (a_n) with infinite hypernatural n such that a_n \approx L (i.e., a_n - L is infinitesimal), it holds that f(a_n) \approx f(L). This nonstandard characterization mirrors the sequential definition of continuity and follows from the transfer principle applied to the first-order statement of the \epsilon-\delta condition, reinterpreted infinitesimally. For sequences specifically, the limit \lim_{n \to \infty} s_n = L holds if ^*s_N \approx L for every infinite hypernatural N, providing an intuitive bridge between standard convergence and infinitesimal approximation. The squeeze theorem for sequences can also be proved nonstandardly using the transfer principle: if real sequences satisfy g(n) \leq f(n) \leq h(n) for all natural n, and \lim g(n) = \lim h(n) = L, then for any infinite hypernatural \omega, ^*g(\omega) \approx L and ^*h(\omega) \approx L, implying ^*f(\omega) \approx L by the transferred order preservation in ^*\mathbb{R}, hence \lim f(n) = L. This approach simplifies the standard \epsilon-N proof by directly leveraging infinitesimal closeness at infinite indices. A classic example is the proof of \lim_{x \to 0} \frac{\sin x}{x} = 1, which uses a geometric interpretation transferred to the hyperreals. The standard geometric argument establishes \cos x < \frac{\sin x}{x} < 1 for $0 < x < \frac{\pi}{2} via areas or lengths in the unit circle; by the transfer principle, this inequality holds in ^*\mathbb{R} for positive infinitesimal \varepsilon \approx 0, yielding \cos \varepsilon < \frac{\sin \varepsilon}{\varepsilon} < 1. Since \cos \varepsilon \approx 1 (as the cosine function is continuous at 0), the squeeze implies \frac{\sin \varepsilon}{\varepsilon} \approx 1, and applying the standard part function st gives the limit as 1. Despite these strengths, the transfer principle has limitations: it applies only to first-order statements, so higher-order assertions involving unbounded quantification over sets or functions do not transfer directly, requiring careful reformulation. Additionally, while effective for sequences via infinite indices, applications to general functions often necessitate extensions like monads or germs, distinguishing sequential limits from functional ones in the nonstandard setting.

Multidimensional Sequences

Definition for Double Sequences

A double sequence in the real numbers is a function a: \mathbb{N} \times \mathbb{N} \to \mathbb{R}, typically denoted by \{a_{m,n}\}_{m,n=1}^\infty, where each term a_{m,n} is indexed by a pair of natural numbers (m, n). The concept of the limit of such a sequence generalizes the single-index case, where the indices approach infinity simultaneously. The index set \mathbb{N} \times \mathbb{N} is equipped with the product metric, for example, d((m_1, n_1), (m_2, n_2)) = \max\{|m_1 - m_2|, |n_1 - n_2|\}, which induces the product topology and allows defining convergence as the indices tend to infinity in this metric space. The double sequence \{a_{m,n}\} is said to converge to a limit L \in \mathbb{R} as (m, n) \to \infty, written \lim_{(m,n) \to \infty} a_{m,n} = L, if for every \varepsilon > 0, there exists N \in \mathbb{N} such that for all m, n > N, |a_{m,n} - L| < \varepsilon.

Examples of Convergence Behavior

In the context of double sequences, convergence behavior often reveals path dependence, where the limiting value varies depending on the relative growth rates of the indices m and n as both tend to infinity. Consider the double sequence defined by a_{m,n} = \frac{m - n}{m + n}. Along the path where m = n, a_{n,n} = 0, suggesting a potential limit of 0. However, along the path m = 2n, a_{2n,n} = \frac{n}{3n} = \frac{1}{3}. This discrepancy shows that the double limit does not exist, as the value depends on the chosen path. A pathological example highlighting selective convergence along certain paths is the double sequence a_{m,n} = 1 if m = n and 0 otherwise. Off the diagonal (where m ≠ n), the terms are 0 for sufficiently large indices, approaching 0 along such paths. Along the diagonal m = n, however, a_{n,n} = 1, so the sequence approaches 1. Consequently, the iterated limits both equal 0 (since for fixed m, as n → ∞, most terms are 0, and similarly for fixed n), but the double limit does not exist due to the persistent 1's on the diagonal for arbitrarily large indices. Another illustration of path dependence is the double sequence a_{m,n} = \frac{m}{n}, reminiscent of ratios in recursive sequences like Fibonacci generalizations. The limiting behavior varies with the asymptotic ratio r = lim (m/n); if m/n → r, then a_{m,n} → r. For instance, along m = n, the limit is 1, while along m = 2n, it is 2. This variability underscores that no unique double limit exists unless the path is restricted. Graphical representations aid in visualizing these patterns. Heat maps, where the color or intensity at grid point (m, n) corresponds to the value of a_{m,n}, can reveal regions of stability or divergence as m and n increase. Path traces—lines along specific ratios like m = kn for fixed k—highlight how values approach different limits, emphasizing the multidimensional nature of convergence. For non-convergence without path-specific limits, consider the oscillating double sequence a_{m,n} = (-1)^{m+n}. The terms alternate between -1 and 1 along any path, preventing approach to a single value; neither iterated limits nor the double limit exists.

Pointwise and Uniform Limits

In the context of sequences of functions from \mathbb{R}^d to \mathbb{R}, pointwise convergence requires that for every fixed point x \in \mathbb{R}^d, the real-valued sequence f_k(x) converges to f(x) as k \to \infty. This means that, given \epsilon > 0, there exists N = N(x, \epsilon) such that for all k > N, |f_k(x) - f(x)| < \epsilon. Uniform convergence strengthens this condition by requiring the convergence to occur simultaneously across the entire domain, independent of x: \sup_{x \in \mathbb{R}^d} |f_k(x) - f(x)| \to 0 as k \to \infty. Equivalently, for every \epsilon > 0, there exists N = N(\epsilon) such that for all k > N and all x, |f_k(x) - f(x)| < \epsilon. For double sequences of functions f_{m,n}: \mathbb{R}^2 \to \mathbb{R}, the notions extend analogously. Pointwise convergence holds if, for every (x,y) \in \mathbb{R}^2, \lim_{m,n \to \infty} f_{m,n}(x,y) = f(x,y), meaning for each fixed (x,y) and \epsilon > 0, there exists N((x,y), \epsilon) such that |f_{m,n}(x,y) - f(x,y)| < \epsilon whenever m, n > N. Uniform convergence over the domain requires \sup_{(x,y) \in \mathbb{R}^2} |f_{m,n}(x,y) - f(x,y)| \to 0 as m,n \to \infty, so the same N(\epsilon) works for all points. Uniform convergence implies pointwise convergence, but the converse fails in general, as the rate of convergence may vary significantly across the domain. A representative example illustrates this distinction for double sequences on the unit square [0,1]^2. Consider f_{m,n}(x,y) = x^m y^n. For each fixed (x,y) \in [0,1]^2 with at least one of x < 1 or y < 1, f_{m,n}(x,y) \to 0 as m,n \to \infty, while at (1,1), f_{m,n}(1,1) = 1 \to 1; thus, the pointwise limit is the function f(x,y) = 0 except at (1,1), where f(1,1) = 1. However, the convergence is not uniform, as near (1,1), \sup_{(x,y) \in [0,1]^2} |f_{m,n}(x,y) - f(x,y)| remains bounded away from 0 (approaching 1 along paths close to the boundary). This highlights how pointwise convergence can fail to control behavior globally, particularly near points where the limit function is discontinuous. A key advantage of uniform convergence is its preservation of important properties, such as continuity. If each f_k: \mathbb{R}^d \to \mathbb{R} is continuous and the sequence converges uniformly to f, then f is continuous on \mathbb{R}^d. The proof relies on the \epsilon-\delta definition: for fixed x_0 and \epsilon > 0, choose N so that \sup |f_k - f| < \epsilon/3 for k > N, then use continuity of f_N to find \delta > 0 such that if \|x - x_0\| < \delta, then |f_N(x) - f_N(x_0)| < \epsilon/3, yielding |f(x) - f(x_0)| < \epsilon by the triangle inequality. In contrast, pointwise limits need not preserve continuity, as seen in the example above where the individual f_{m,n} are continuous (polynomials) but the pointwise limit is discontinuous at (1,1). This theorem underscores uniform convergence's role in maintaining structural properties in multivariable settings.

Iterated Limits and Their Pitfalls

In the context of double sequences \{a_{m,n}\}_{m,n=1}^\infty, the iterated limits are defined as follows: the first iterated limit is \lim_{m \to \infty} \lim_{n \to \infty} a_{m,n}, provided the inner limit \lim_{n \to \infty} a_{m,n} exists for each fixed m and the resulting sequence in m converges; the second iterated limit is \lim_{n \to \infty} \lim_{m \to \infty} a_{m,n}, analogously. These differ from the double limit \lim_{(m,n) \to \infty} a_{m,n} = L, which requires that for every \epsilon > 0, there exists N \in \mathbb{N} such that |a_{m,n} - L| < \epsilon whenever m > N and n > N (Pringsheim sense of convergence). A key pitfall is that the iterated limits may exist but differ from each other, implying the limit does not exist. Consider the sequence a_{m,n} = \frac{m}{m + n}. For fixed m, \lim_{n \to \infty} a_{m,n} = 0, so \lim_{m \to \infty} \lim_{n \to \infty} a_{m,n} = 0. However, for fixed n, \lim_{m \to \infty} a_{m,n} = 1, so \lim_{n \to \infty} \lim_{m \to \infty} a_{m,n} = 1. Along the path m = n, a_{m,m} = \frac{1}{2}, confirming the limit fails to exist. Under additional conditions, such as uniform convergence of the inner limits, the iterated and double limits coincide. By an analogue of Osgood's theorem for sequences in complete metric spaces, if \lim_{n \to \infty} a_{m,n} = b_m uniformly in m and \lim_{m \to \infty} b_m = L, then both iterated limits and the double limit exist and equal L. These pitfalls underscore the need for caution in multivariable settings, where can invalidate naive interchanges of limits.

Proof Techniques

Epsilon-Delta Proofs for Real Limits

The ε-N definition provides a rigorous foundation for proving limit properties of sequences in the real numbers. For sequences \{a_n\} and \{b_n\} converging to limits A and B respectively, the sum rule states that \{a_n + b_n\} converges to A + B. To prove this, let \epsilon > 0. Since \lim_{n \to \infty} a_n = A, there exists N_1 \in \mathbb{N} such that for all n > N_1, |a_n - A| < \epsilon/2. Similarly, there exists N_2 \in \mathbb{N} such that for all n > N_2, |b_n - B| < \epsilon/2. Let N = \max\{N_1, N_2\}. Then, for all n > N, |(a_n + b_n) - (A + B)| \leq |a_n - A| + |b_n - B| < \epsilon/2 + \epsilon/2 = \epsilon. This establishes the sum rule using the triangle inequality. The limit of a convergent sequence is unique. Suppose \lim_{n \to \infty} a_n = L and \lim_{n \to \infty} a_n = M with L \neq M. Let \epsilon = |L - M|/2 > 0. There exists N_1 \in \mathbb{N} such that for all n > N_1, |a_n - L| < \epsilon. Similarly, there exists N_2 \in \mathbb{N} such that for all n > N_2, |a_n - M| < \epsilon. Let N = \max\{N_1, N_2\}. For n > N, |L - M| \leq |L - a_n| + |a_n - M| < \epsilon + \epsilon = |L - M|, which is a contradiction. Thus, the limit is unique. The squeeze theorem asserts that if \{g_n\} and \{h_n\} converge to the same limit L and g_n \leq f_n \leq h_n for all sufficiently large n, then \lim_{n \to \infty} f_n = L. To prove this, let \epsilon > 0. There exists N_1 \in \mathbb{N} such that for all n > N_1, |g_n - L| < \epsilon. Similarly, there exists N_2 \in \mathbb{N} such that for all n > N_2, |h_n - L| < \epsilon. Let N = \max\{N_1, N_2\}. For n > N, L - \epsilon < g_n \leq f_n \leq h_n < L + \epsilon, so |f_n - L| < \epsilon. This follows from the ordering of the sequences. A monotone increasing sequence \{a_n\} that is bounded above converges to its least upper bound. Let s = \sup\{a_n : n \in \mathbb{N}\}. For \epsilon > 0, there exists k \in \mathbb{N} such that s - a_k < \epsilon by the definition of the supremum. Since the sequence is increasing, for all n \geq k, a_k \leq a_n \leq s, so |a_n - s| < \epsilon. The least upper bound property of \mathbb{R} ensures s exists and is finite. A similar argument applies to monotone decreasing bounded sequences. In ε-N proofs, the Archimedean property—that for any \epsilon > 0, there exists N \in \mathbb{N} such that N > 1/\epsilon—is often used to bound indices when dealing with terms like $1/n. This property guarantees the existence of sufficiently large N to satisfy the inequality for all subsequent terms.

Proofs of Sequential Criteria in Topology

In topological spaces, a function f: X \to Y between topological spaces X and Y is continuous at a point x \in X if for every neighborhood V of f(x), the preimage f^{-1}(V) is a neighborhood of x. To prove that continuity implies sequential continuity, suppose x_n \to x in X. Let V be any neighborhood of f(x) in Y. Since f is continuous, f^{-1}(V) is a neighborhood of x in X, so there exists N such that x_n \in f^{-1}(V) for all n \geq N, implying f(x_n) \in V for n \geq N. Thus, f(x_n) \to f(x). Sequential compactness in a topological space X means every sequence in X has a convergent subsequence. To show sequential compactness implies compactness, consider an open cover \{U_i\}_{i \in I} of X. In general topological spaces, this implication requires additional structure like second-countability, but it holds in metric spaces. For a metric space (X, d), assume X is sequentially compact. For any \epsilon > 0, the collection of open balls B(x, \epsilon) for x \in X forms an open cover. By sequential compactness, X is totally bounded: if not, there exists \epsilon > 0 and a sequence with no Cauchy subsequence, contradicting the existence of convergent subsequences. Moreover, sequential compactness implies completeness, as any Cauchy sequence has a convergent subsequence, and the limit is the Cauchy limit. A complete and totally bounded metric space is compact by the Heine-Borel-like theorem for metrics. Specifically, for total boundedness, cover X by finitely many \epsilon-balls; for completeness, any Cauchy sequence converges. Thus, every open cover has a finite subcover. In first-countable topological spaces, sequential limits characterize closed sets. A space X is first-countable if every point has a countable neighborhood basis. A subset A \subseteq X is closed if and only if it contains all limits of convergent sequences in A. First, if A is closed and x_n \to x with x_n \in A, then x \in \overline{A} = A. Conversely, if x \in \overline{A}, let \{B_n\} be a countable decreasing neighborhood basis at x. Choose a_n \in A \cap B_n; then a_n \to x since for any neighborhood U of x, some B_k \subseteq U, so eventually a_n \in U. By assumption, x \in A. This fails in non-first-countable spaces, where points in the closure may lack sequential witnesses. In metric spaces, sequential compactness is equivalent to compactness via total boundedness and completeness. To prove sequential compactness implies these: For completeness, let \{x_n\} be Cauchy; by sequential compactness, a subsequence x_{n_k} \to L, so the original converges to L. For total boundedness, suppose not; then for \epsilon = 1/k, there is an infinite \epsilon-separated set, yielding a sequence with no convergent subsequence, contradiction. Conversely, in a complete totally bounded metric space, any sequence has a Cauchy subsequence (by finite \epsilon-net and pigeonhole), hence convergent. A illustrating the need for nets beyond is the cocountable topology on an X, where open sets are \emptyset and complements of countable sets. This space is not sequential: consider A = X \setminus C where C is a countable . Then A is open (complement countable) and hence not closed, but A is sequentially closed because the only convergent are eventually constant, so any convergent sequence in A has its limit in A. Thus, sequentially closed sets properly contain all closed sets (which are the countable subsets and X), showing that do not suffice to characterize the . Nets (generalized ) are required to characterize and here, as suffice only in sequential spaces like first-countable ones.

References

  1. [1]
    Limit of a sequence in the Real Numbers - UTSA
    Oct 26, 2021 · The limit of a sequence is said to be the fundamental notion on which the whole of mathematical analysis ultimately rests.
  2. [2]
    Sequences - Calculus II - Pauls Online Math Notes
    Jul 11, 2023 · Working Definition of Limit · We say that limn→∞an=L lim n → ∞ ⁡ a n = L if we can make an as close to L L as we want for all sufficiently large ...
  3. [3]
    Limit of a sequence - StatLect
    In this lecture we introduce the notion of limit of a sequence [eq1]. We start from the simple case in which [eq2] is a sequence of real numbers.Informal definition of the limit of... · Formal definition of the limit of...
  4. [4]
    Sequences, Their Limits and Convergence | CK-12 Foundation
    The limit of a sequence a n is the number L such that if for each ε > 0 , there exists an integer N such that | a n − L | < ε for all n > N .<|control11|><|separator|>
  5. [5]
    [PDF] chapter ii the limit of a sequence of numbers definition of the number
    This chapter contains the beginnings of the most important, and probably the most subtle, notion in mathematical analysis, i.e., the concept of a limit.
  6. [6]
    Zeno's paradoxes - Stanford Encyclopedia of Philosophy
    Apr 30, 2002 · One aspect of the paradox is thus that Achilles must traverse the following infinite series of distances before he catches the tortoise ...
  7. [7]
    Zeno's Paradoxes | Internet Encyclopedia of Philosophy
    The Achilles Paradox is reconstructed from Aristotle (Physics Book VI, Chapter 8, 239b14-16) and some passages from Simplicius in the fifth century C.E. There ...
  8. [8]
    [PDF] Archimedes' Quadrature of the Parabola
    Jun 12, 2013 · One using the “method of exhaustion” – part of standard repertoire of Euclidean mathematics (see Book 12 of. Elements), based on work of ...
  9. [9]
    [PDF] Archimedes' quadrature of the parabola and the method of exhaustion
    By the time of Archimedes, the following facts were known concerning an arbitrary parabolic segment S with base AB and vertex P (see the figure below). P1. The ...
  10. [10]
    Madhava (1350 - 1425) - Biography - MacTutor History of Mathematics
    Madhava was a mathematician from South India. He made some important advances in infinite series including finding the expansions for trigonometric functions.
  11. [11]
    Calculus history - MacTutor - University of St Andrews
    Newton wrote a tract on fluxions in October 1666. This was a work which was not published at the time but seen by many mathematicians and had a major influence ...
  12. [12]
    A Brief History of Infinitesimals: The Idea That Gave Birth to Modern ...
    Apr 1, 2014 · By 1700 Isaac Newton and Gottfried Leibniz had turned this approach into the powerful algorithm we know as “the calculus,” capable of being ...
  13. [13]
    Continuity and Infinitesimals - Stanford Encyclopedia of Philosophy
    Jul 27, 2005 · Continuity means unbroken, uninterrupted, with no gaps. Infinitesimals are quantities so small that some powers, like the square, vanish.
  14. [14]
    [PDF] Bolzano on Continuity and the Intermediate Value Theorem
    Feb 25, 2023 · The Bounded Set Theorem that Bolzano used to prove his version of the Intermediate Value Theorem was a highly original idea, and is closely ...
  15. [15]
    [PDF] Teaching Analysis with Bolzano's Original Text
    Jun 17, 2024 · 1817 monograph1 by Bernard Bolzano on the Intermediate Value Theorem. ... Theorem asserts that every infinite set contains a limit point ...
  16. [16]
    [PDF] Who Gave You the Epsilon? Cauchy and the Origins of Rigorous ...
    Cauchy's foundations of analysis appear in the books based on his lectures at the Ecole Polytechnique; his book of 1821 was the first example of the great ...
  17. [17]
    [PDF] Cauchy's Cours d'analyse
    Not only did Cauchy provide a workable definition of limits and a means to make them the basis of a rigorous theory of calculus, but also he revitalized the ...
  18. [18]
    [PDF] On the history of epsilontics - arXiv
    In 1821, when Cauchy was writing his 'Course of Analysis', E. Heine was born in Berlin. Fifty-one years later, the latter stated the concept of uniform.<|separator|>
  19. [19]
    (PDF) On the history of epsilontics - ResearchGate
    Jun 5, 2015 · It was only in 1861 that the epsilon-delta method manifested itself to the full in Weierstrass' definition of a limit.
  20. [20]
  21. [21]
    [PDF] The Origins of Cauchy's Rigorous Calculus
    Augustin-Louis Cauchy gave the first reasonably success- ful rigorous foundation for the calculus. Beginning with a precise definition of limit, ...
  22. [22]
    [PDF] An Introduction to Real Analysis John K. Hunter - UC Davis Math
    A sequence of real numbers converges if and only if it is a Cauchy sequence. The fact that real Cauchy sequences have a limit is an equivalent way to formu-.
  23. [23]
    [PDF] An Introduction to Real Analysis John K. Hunter - UC Davis Math
    Abstract. These are some notes on introductory real analysis. They cover the properties of the real numbers, sequences and series of real numbers, limits.
  24. [24]
    [PDF] The sequence sin(n) - Concordia University
    Oct 30, 2009 · Fact. The function sin : R → [−1, 1] is periodic of period 2π, continuous and surjective. The number π is irrational. We recall that. Definition.
  25. [25]
    [PDF] Basic Analysis: Introduction to Real Analysis
    May 23, 2025 · ... ANALYSIS. 7. 0.2 About analysis. Analysis is the branch of mathematics that deals with inequalities and limits. The present course deals with ...
  26. [26]
    [PDF] 18.100A Fall 2020 Lecture 8: The Squeeze Theorem and ...
    The Squeeze Theorem and Operations Involving Convergent Sequences. Facts ... Proof: Let > 0. Since limn→∞ an = x, there exists an M0 ∈ N such that ...
  27. [27]
    [PDF] 3.4 Monotone Sequences
    Mar 3, 2018 · Theorem 3.4.4 (Monotone Convergence Theorem). Suppose hsni is a monotone sequence. Then hsni is convergent if and only if hsni is bounded ...
  28. [28]
    [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 ...
  29. [29]
    Metric Spaces: Limits and Continuity
    We say that a sequence in M is convergent if it converges to some point in M . This definition is given in terms of open balls in the metric space. But note ...
  30. [30]
    [PDF] monotone sequences in metric spaces - Journal Of Science and Arts
    Mar 30, 2025 · The aim of this paper is to introduce the notion of monotone sequence in metric spaces. Some properties and examples are given. Keywords: metric ...
  31. [31]
    [PDF] Chapter 5. Limits and Continuity in Metric Spaces
    5.2 Theorem: (Basic Properties of Limits of Sequences) Let (xn)n≥p be a sequence in a metric space X, and let a ∈ X. (1) If (xn)n≥p converges then its limit is ...
  32. [32]
    [PDF] Sequences and Continuity in Metric Spaces
    Our first characterization of continuity is in terms of convergent sequences: A function is con- tinuous if and only if it preserves convergence of sequences.
  33. [33]
    Math 521 Uniform Convergence
    Definition. If X is a metric space, and fn:X→R (n∈N) is a sequence of functions, then fn converges pointwise to f if for every x∈X one has limn→∞fn(x)=f(x).
  34. [34]
    [PDF] 4 Sequences of functions - People
    Definition 4.1.1 (Pointwise convergence). Let (X, dX) and (Y, dY) be metric spaces. The sequence (fn) of functions from X to Y converges.
  35. [35]
    [PDF] Advanced Calculus II - People
    This proves that K is sequentially compact. The Heine–Borel theorem is a special property of Rd (under the usual metrics) and does not hold in general. For ...
  36. [36]
    [PDF] METRIC SPACES 1. Introduction As calculus developed, eventually ...
    A subset S ⊂ X is called dense if every element of X is the limit of a sequence in S or equivalently every open ball around an element of X contains an element.
  37. [37]
    Math 521 Sequences in Metric spaces
    ... nk≥k. Definition of limit. A sequence of points xn in a metric space X converges to a point a∈X if for every ε>0 there is an Nε∈N such that for ...
  38. [38]
    Complete Metric Spaces - Advanced Analysis
    Jan 17, 2024 · A sequence { x n } n = 1 ∞ in a metric space ( X , d ) is called a Cauchy sequence when for all ϵ > 0 , there is some N ∈ N such that every pair ...
  39. [39]
    [PDF] 18.100A Fall 2020 Lecture 10: The Completeness of the Real ...
    Theorem 6. If {xn} is Cauchy and a subsequence {xnk } converges, then {xn} converges. Proof: Suppose that {xnk } is a subsequence of {xn} such that limk→∞ xnk.
  40. [40]
    [PDF] Cauchy's Construction of R - UCSD Math
    The real numbers will be constructed as equivalence classes of Cauchy sequences. Let CQ denote the set of all Cauchy sequences of rational numbers. We must ...
  41. [41]
    Real Analysis: 3.2. Cauchy Sequences - MathCS.org
    For example, the sequence 1, 1.4, 1.41, 1.414, ... (convergent to the square root of 2) is Cauchy, but does not converge to a rational number. Therefore ...
  42. [42]
    The Banach Fixed Point Theorem: selected topics from its hundred ...
    Jul 9, 2024 · Its important feature is that it gives the existence and uniqueness of a fixed point and convergence of the sequence of successive ...<|control11|><|separator|>
  43. [43]
    [PDF] General Topology
    General Topology. Tom Leinster. 2014–15. Page 2. Contents. A Topological spaces. 2. A1 Review of metric spaces . . . . . . . . . . . . . . . . . . . . . . . 2.Missing: textbook | Show results with:textbook
  44. [44]
    [PDF] Sequences and nets in topology - arXiv
    Jun 23, 2010 · Bourbaki's book on General Topology [3] was the first to fully adopt the use of filters. The connections between nets and filters were ...
  45. [45]
    [PDF] Chapter III Topological Spaces
    (This half of the proof is valid for any topological space : continuous functions always. \. “preserve convergent sequences.”) (. ) Let be a shrinking ...
  46. [46]
    [PDF] A.7 Convergence and Continuity in Topological Spaces
    Let X, Y be topological spaces and let x ∈ X be given. Then a function f : X → Y is continuous at x if for each open neighborhood V of f(x) in Y , there ...
  47. [47]
    [PDF] math 201, april 13, 2020
    Apr 13, 2020 · 0 for x ≤ 0 is discontinuous at 0. Have 1/n → 0, but f(1/n)=1 → 1 6=0= f(0). Proposition. A function f : D → R is continuous at c in D if and.Missing: image | Show results with:image
  48. [48]
    Sequential definitions of compactness - ScienceDirect.com
    A subset of a topological space is sequentially compact if any sequence of points in has a convergent subsequence whose limit is in .
  49. [49]
    [PDF] A subset of a metric space is compact if and only if it is sequentially ...
    Theorem: A subset of a metric space is compact if and only if it is sequentially compact. Proof: ⇒ Suppose that (X,d) is a compact metric space. Further, ...
  50. [50]
    [PDF] Closed Sets and Sequential Compactness Definition 1. Let X be a ...
    A topological space X is sequentially compact iff every se- quence {xn} ⊂ X has a convergent subsequence. Theorem 8. If a metric space (X,d) is compact then it ...
  51. [51]
    [PDF] COMPACTNESS VS. SEQUENTIAL - MIT OpenCourseWare
    The aim of this handout is to provide a detailed proof of the equivalence between the two definitions of compactness: existence of a finite subcover of any ...
  52. [52]
    [PDF] Section 2.3: Bolzano-Weierstrass Theorem
    Oct 19, 2011 · Theorem 2-12 (Bolzano-Weierstrass): Every bounded infinite set of real numbers has at least one limit point. Note: Clearly some bounded infinite ...Missing: generalization | Show results with:generalization
  53. [53]
    [PDF] An introduction to nonstandard analysis - UChicago Math
    Aug 14, 2009 · In section 2 we will introduce the notion of an ultrafilter, which will allow us to do a typical ultrapower construction of the hyperreal ...
  54. [54]
    nonstandard analysis in nLab
    ### Summary of Nonstandard Analysis from nLab
  55. [55]
    None
    Summary of each segment:
  56. [56]
    [PDF] arXiv:1306.5973v1 [math.HO] 25 Jun 2013
    Jun 25, 2013 · sin x = x 1 − ... Euler's argument in fa- vor of (3.5) and (3.6) was formalized in terms of a “nonstandard” proof in (Luxemburg 1973 [88]).
  57. [57]
    [PDF] 8. Doubles Sequences and Their Limits - Benjamin D. Fedoruk
    A double sequence is a function a : N × N → R denoted by (an,m). For each (n, m) ∈ N × N, an,m represents a real number. Definition 2 (Double Limit).
  58. [58]
    [PDF] Analysis I: Example Sheet 1 - DPMMS
    A double sequence is a function from N×N to (say) R, where N × N = {(m, n) : m, n ∈ N}. For example, ap,q is a double sequence, where ap,q = p p + q. , p, q ...
  59. [59]
    [PDF] THE N¨ORLUND ORLICZ SPACE OF DOUBLE ENTIRE RATE ...
    Consider a double sequence x = (xij). The (m, n)th section x[m,n] of the ... 1; if m = n. 0; otherwise. An FK-space (or a metric space) X is said to have ...
  60. [60]
    [PDF] Shanghai Lectures on Multivariable Analysis - Arizona Math
    2.5 Uniform convergence . . . . . . . . . . . . . . . . . . . . . . . . . 54 ... This book originated in lectures given in Fall 2014 at NYU Shanghai for an.
  61. [61]
    [PDF] Statistical Convergence of Double Sequences of Real-Valued ...
    The necessity is trivial. It is clear that uniform convergence implies pointwise convergence with the ... Let {fjk} be a double sequence of functions defined on a.
  62. [62]
  63. [63]
    [PDF] Infinite series, infinite products, and infinite fractions - People
    smn = mn/(m + n)2. We know that lim smn does not exist, but observe that lim n→∞ smn = lim n→∞ mn. (m + n)2. = 0. =⇒ lim m→∞ lim n→∞ smn = lim m→∞. 0=0 ...
  64. [64]
    [PDF] Supplement on lim sup and lim inf
    By property 3, we can give definitions on the upper limit and the lower limit of a given sequence as follows. Definition Given a real sequence an and let bn ...
  65. [65]
    [PDF] TOPIC. Sums and limits. This section considers the question
    Jan 4, 2001 · If the double sum exists, then so do the iterated sums r and c, and ... For example, consider the sequences x1 = (1, −1, 0, 0, 0, ...), x2 ...<|control11|><|separator|>
  66. [66]
    [PDF] 12 Proofs in calculus - UBC Math
    Theorem (Archimedean property). The set N is unbounded above in R. It is clearly bounded below by 1. Proof. Assume that N is bounded above. Then by the ...
  67. [67]
    general topology - sequential continuity vs. continuity
    Jul 1, 2014 · Let be X and Y topological spaces and f:X→Y is a map. Then it holds: 1) If f is continuous then it is sequentially continuous.Missing: "topology | Show results with:"topology
  68. [68]
    [PDF] Real Analysis
    Dec 13, 2022 · If a metric space X is sequentially compact, then it is complete and totally bounded. ع. Proof. Let metric space X be sequentially compact.
  69. [69]
    [PDF] compact sets in metric spaces notes for math 703
    Let (X, d) be a metric space and assume A ⊂ X. Then. A is compact if and only if A is a complete and totally bounded. Proof. Assume first that A is compact.Missing: total | Show results with:total
  70. [70]
    [PDF] MAT327 - Lecture 7
    May 29, 2019 · Let (X,T ) be a first countable topological space and let A ⊆ X, and let a ∈ A. Then a ∈ A is and only if there is a sequence in A converging to ...
  71. [71]
    [PDF] characterizations of compactness for metric spaces
    A metric space is sequentially compact if and only if every infinite subset has an accumulation point. Lemma 3. A compact metric space is sequentially compact.
  72. [72]
    Uncountable Set with Co-countable topology is not a Sequential ...
    Dec 30, 2018 · Prove that a topological space is a sequential space if and only if every sequentially open set is open.non sequential spaces - general topology - Math Stack ExchangeProperties of the co-countable topology on $[0,1]More results from math.stackexchange.com