Fact-checked by Grok 2 weeks ago

Upper and lower bounds

In mathematics, particularly within the field of , an upper bound for a S of the \mathbb{R} is any M such that x \leq M for every x \in S. A lower bound for S is any m such that x \geq m for every x \in S. A set S is said to be bounded if it is both bounded above and bounded below, meaning there exist finite m and M such that m \leq x \leq M for all x \in S. The least upper bound of a nonempty set S bounded above, denoted \sup S, is the smallest number that serves as an upper bound for S; it is unique when it exists and may or may not belong to S itself. Similarly, the greatest lower bound, or infimum \inf S, is the largest number that is a lower bound for S, and it too may lie outside the set. If \sup S \in S, then \sup S is the maximum element of S; analogously, if \inf S \in S, it is the minimum. The existence of suprema and infima for bounded sets is a of the real numbers' structure, encapsulated in the : every nonempty of \mathbb{[R](/page/R)} that is bounded above has a least upper bound in \mathbb{[R](/page/R)}, and every nonempty bounded below has a greatest lower bound in \mathbb{[R](/page/R)}. This property distinguishes the reals from , where bounded sets may lack suprema (e.g., the set of less than \sqrt{2} has no least upper bound in \mathbb{[Q](/page/Q)}), and it underpins the development of limits, , and in . Beyond , upper and lower bounds extend to ordered structures like partially ordered sets in , where they define bounds relative to the order relation, and to approximation theory, where they quantify error estimates in numerical methods. In algorithm analysis, big-O notation provides upper bounds on , while big-Omega offers lower bounds, enabling rigorous performance guarantees. These concepts are indispensable for proving inequalities, optimizing functions, and establishing foundational results across and its applications.

Fundamental Concepts

Upper Bound

In a (poset) (P, \leq), an element M \in P is called an upper bound for a S \subseteq P if every element of S is less than or equal to M under the order relation. This means that M serves as a "ceiling" above all elements in S, limiting the subset from above without necessarily being the tightest such limit. Formally, M is an upper bound for S if \forall s \in S, \ s \leq M. This condition ensures that no element in S exceeds M, capturing the intuitive notion of an upper limit in ordered structures. For instance, in the set of real numbers \mathbb{R} equipped with the standard order \leq, the number 5 acts as an upper bound for the finite subset \{1, 2, 3, 4\}, since $1 \leq 5, $2 \leq 5, $3 \leq 5, and $4 \leq 5. Here, any greater than or equal to 5 would also qualify as an upper bound, illustrating how upper bounds are not unique in general posets. The collection of all upper bounds for a fixed subset S in a poset forms an upset, a U \subseteq P that is upward closed: if M \in U and M \leq N for some N \in P, then N \in U. This property arises because if M bounds S from above and N is greater than or equal to M, then N also bounds S. Multiple upper bounds may exist for the same S. The concept of upper bounds traces back to Richard Dedekind's 1872 essay "Stetigkeit und irrationale Zahlen" (Continuity and Irrational Numbers), where cuts in rely on partitioning sets to define reals via upper and lower classes, laying foundational groundwork for constructing numbers with the .

Lower Bound

In a (poset) (P, \leq), a lower bound for a nonempty S \subseteq P is an element m \in P such that m \leq s for every s \in S. This condition ensures that m lies "below" or is comparable from underneath to all elements of S with respect to the partial order. Formally, m is a lower bound of S if \forall s \in S, m \leq s. The collection of all lower bounds of a fixed subset S in a poset forms a , meaning that if m is a lower bound of S and m' \leq m for some m' \in P, then m' is also a lower bound of S. This property arises because the order relation is reflexive and transitive, preserving the bounding condition downward. Lower bounds are the order-dual of upper bounds: reversing the partial \leq to \geq transforms lower bounds into upper bounds and vice versa. For illustration, consider the poset of real numbers (\mathbb{R}, \leq) under the standard order. The number serves as a lower bound for the \{1, 2, 3, 4\}, since $0 \leq 1, $0 \leq 2, $0 \leq 3, and $0 \leq 4. Similarly, in the poset of integers (\mathbb{Z}, \leq), -10 is a lower bound for the set of positive even integers \{2, 4, 6, \dots\}, as -10 is less than or equal to each element in the set.

Bounds in Ordered Sets

In ordered sets, including partially ordered sets (posets), an upper bound for a subset S is an element u such that s \leq u for all s \in S, and a lower bound is an element l such that l \leq s for all s \in S. Unlike total orders like \mathbb{R}, where all elements are comparable, posets allow incomparability, so subsets may lack upper or lower bounds. The least upper bound (supremum) and greatest lower bound (infimum), if they exist, are the smallest upper bound and largest lower bound, respectively.

Bounds of Finite Sets

In the real numbers \mathbb{R}, equipped with the standard , any nonempty finite S possesses both upper and lower bounds, with the maximum element \max(S) serving as the least upper bound, or supremum \sup S, and the minimum \min(S) as the greatest lower bound, or infimum \inf S. This property arises because, among a finite collection of real numbers, direct pairwise comparisons identify the largest and smallest elements unequivocally. To compute these bounds, one enumerates the elements of S and identifies the extrema through comparison. For instance, given S = \{1, 3, 4\}, enumeration reveals $4 as the maximum, so \sup S = 4, and $1 as the minimum, so \inf S = 1; consequently, any \geq 4 qualifies as an upper bound, while any \leq 1 serves as a lower bound. This direct method ensures the bounds are tight, meaning the least upper bound is attained within S itself, and likewise for the greatest lower bound. Every nonempty finite subset of \mathbb{R} is bounded above and below, a consequence of the existence of maxima and minima in finite totally ordered sets.

Bounds of Infinite Sets

In the context of subsets of the real numbers \mathbb{R}, a subset S is bounded above if there exists a M such that s \leq M for all s \in S, where M is called an upper bound for S. Similarly, S is bounded below if there exists a m such that s \geq m for all s \in S, with m serving as a lower bound. A subset S is bounded if it is both bounded above and bounded below. For infinite sets, these properties highlight distinctions from finite cases. Consider the open interval (0,1), an infinite of \mathbb{R}; it is bounded above by 1 (since all elements are less than 1) and bounded below by 0 (since all elements are greater than 0). In contrast, the set of natural numbers \mathbb{N} = \{1, 2, 3, \dots\} is bounded below by 1 but unbounded above, as no exceeds all its elements. The set of integers \mathbb{Z} provides an example of a set unbounded both above and below, since its elements extend indefinitely in both positive and negative directions. Bounded infinite subsets of \mathbb{R} can be characterized as those contained within some closed [m, M], where m is a lower bound and M is an upper bound. A key property of such sets is given by the Bolzano-Weierstrass theorem: every bounded infinite subset of \mathbb{R} has at least one . In cases where tight bounds exist, the supremum or infimum may coincide with an actual element of the set, though full exploration of such notions falls under advanced concepts.

Bounds in Sequences and Series

Bounds of Sequences

In , a sequence \{a_n\}_{n=1}^\infty in \mathbb{R} is bounded above if there exists a M \in \mathbb{R} such that a_n \leq M for all n \in \mathbb{N}; it is bounded below if there exists m \in \mathbb{R} such that a_n \geq m for all n \in \mathbb{N}. A sequence is bounded if it is both bounded above and below. This notion of boundedness for sequences is equivalent to the range set \{a_n \mid n \in \mathbb{N}\} being a bounded of \mathbb{R}, meaning it is contained within some finite [m, M]. A fundamental result connecting boundedness and convergence is the : every monotonic bounded in \mathbb{R} converges to a in \mathbb{R}. Conversely, if a monotonic is unbounded above (respectively, below), it diverges to +\infty (respectively, -\infty). For example, consider the sequence a_n = 1 - \frac{1}{n}; it satisfies $0 \leq a_n < 1 for all n \geq 1, so it is bounded, and it is increasing and thus converges to 1 by the monotone convergence theorem. Bounded sequences in \mathbb{R} also exhibit finite oscillation, where the oscillation is defined as \sup_{m,n \in \mathbb{N}} |a_m - a_n|, which remains finite precisely because the range set is bounded and hence has finite diameter.

Bounds of Series

In the context of infinite series, the notion of boundedness applies to the sequence of partial sums, which determines the convergence of the series \sum a_n. The partial sum is defined as s_n = \sum_{k=1}^n a_k, and the series is said to be bounded if the sequence \{s_n\} is bounded, meaning there exists some M > 0 such that |s_n| \leq M for all n. This boundedness of partial sums is a necessary condition for convergence, as an unbounded sequence of partial sums implies divergence. Absolute convergence provides a sufficient condition for both convergence and boundedness. If the series \sum |a_n| converges, then \sum a_n is absolutely convergent, which implies that it converges and that its partial sums are bounded. This follows from the triangle inequality, as the partial sums of the absolute series bound those of the original series. Classic examples illustrate these concepts. The harmonic series \sum_{n=1}^\infty \frac{1}{n} diverges because its partial sums s_n are unbounded above; specifically, s_n grows like \ln n + \gamma, where \gamma is the Euler-Mascheroni constant, exceeding any fixed bound for sufficiently large n. In contrast, the geometric series \sum_{n=0}^\infty r^n with |r| < 1 has partial sums s_n = \frac{1 - r^{n+1}}{1 - r}, which are bounded and converge to \frac{1}{1-r}. For series with positive terms, convergence tests often rely on the boundedness of partial sums. If a_n > 0 for all n and the partial sums \{s_n\} are increasing and bounded above, then the series converges by the , as the \{s_n\} converges to some finite . This monotonicity arises naturally for positive terms, reducing the problem to establishing an upper bound on the partial sums, often via comparison with a . Alternating series exhibit bounded partial sums under specific conditions. For an \sum (-1)^{n+1} b_n where b_n > 0, b_n is decreasing, and \lim_{n \to \infty} b_n = 0, the Leibniz test (or ) guarantees convergence, with partial sums bounded between consecutive terms. The even partial sums form a decreasing bounded below, while the odd partial sums form an increasing bounded above, ensuring the overall \{s_n\} is bounded and converges. A more general criterion for convergence involves the Cauchy condition on partial sums. The series \sum a_n converges if and only if for every \epsilon > 0, there exists N \in \mathbb{N} such that for all m > n \geq N, |s_m - s_n| < \epsilon, meaning the partial sums are Cauchy and thus bounded in a uniform sense across tails of the sequence. This strengthens mere boundedness by controlling differences between partial sums, applicable to any series regardless of term signs.

Bounds of Functions

Pointwise Bounds

In real analysis, a function f: D \to \mathbb{R}, where D \subseteq \mathbb{R} is the domain, is said to have a pointwise upper bound M \in \mathbb{R} if f(x) \leq M for every x \in D; similarly, it has a pointwise lower bound m \in \mathbb{R} if f(x) \geq m for every x \in D. The term "pointwise" emphasizes that the inequality is verified individually at each point in the domain, without regard to the uniformity of the bound across D. A classic example is the constant function f(x) = c for c \in \mathbb{R} and x \in D, which satisfies c \leq f(x) \leq c for all x \in D, making c both an upper and lower bound. Another is the sine function f(x) = \sin x on \mathbb{R}, which is pointwise bounded above by 1 and below by -1, since -1 \leq \sin x \leq 1 holds for all real x. A function f is bounded if it admits both a finite pointwise upper bound and a finite pointwise lower bound. Consider the function f(x) = 1/x for x > 0; on the (0,1], it is bounded below by 0 since f(x) > 0 for all x \in (0,1], but it has no upper bound because f(x) \to \infty as x \to 0^+. This example, which is continuous on its but discontinuous at 0 if extended, illustrates how bounds can fail to exist on certain subdomains despite the function being well-defined. The existence of pointwise bounds for a function does not imply or ; for instance, f(x) = \sin(1/x) on (0,1] is bounded by -1 and 1 but fails due to rapid oscillations near 0. Graphically, pointwise bounds appear as lines y = M and y = m that lie entirely above and below the graph of f over D, respectively, without the graph crossing these lines at any point. Unlike uniform bounds, which require a single bound with a controlled supremum difference across the , pointwise bounds allow the "tightness" to vary locally.

Uniform Bounds

A function f: D \to \mathbb{R}, where D \subseteq \mathbb{R}^d is a , is uniformly bounded above if there exists M \in \mathbb{R} such that f(x) \leq M for all x \in D; it is uniformly bounded below if there exists m \in \mathbb{R} such that f(x) \geq m for all x \in D. Similarly, f is (uniformly) bounded if it is both uniformly bounded above and below. For a single function, uniform boundedness coincides with ordinary boundedness, but the stresses that the bound holds globally across the domain with a single constant independent of x. The , or , of a f on D is defined as \|f\|_\infty = \sup_{x \in D} |f(x)|, which provides the tightest uniform bound measuring the maximum deviation of |f| from zero over the entire domain. If \|f\|_\infty < \infty, then f is uniformly bounded by \|f\|_\infty. This norm induces the topology of uniform convergence on spaces of and is central to the study of continuity and compactness in function spaces. A key example arises from the extreme value theorem: if f: K \to \mathbb{R} is continuous and K \subseteq \mathbb{R}^d is compact, then f attains its maximum and minimum values on K, implying f is uniformly bounded (with bounds given by these extrema). This result, originally due to Weierstrass, underscores the interplay between continuity, compactness, and uniform boundedness. For families of functions, uniform boundedness requires a single M > 0 such that |f(x)| \leq M for all f in the family and all x in the , or equivalently, \sup_f \|f\|_\infty < \infty. In functional analysis, the uniform boundedness principle (Banach–Steinhaus theorem) asserts that if \mathcal{T} is a pointwise bounded family of continuous linear operators from a Banach space X to a normed space Y—meaning \sup_{T \in \mathcal{T}} \|T x\|_Y < \infty for each x \in X—then \mathcal{T} is uniformly bounded, i.e., \sup_{T \in \mathcal{T}} \|T\| < \infty. This theorem, first proved by Banach and Steinhaus in 1927, has profound implications for operator theory and convergence in Banach spaces. Uniform boundedness also plays a role in compactness criteria for function spaces. Specifically, the Arzelà–Ascoli theorem states that a family \mathcal{F} of continuous real-valued functions on a compact set K is precompact in C(K) (the space of continuous functions on K equipped with the supremum norm) if and only if \mathcal{F} is uniformly bounded and equicontinuous. Equicontinuity ensures controlled variation across the domain, complementing uniform boundedness to guarantee relative compactness. To illustrate the distinction from weaker notions, uniform bounds require a global constant, whereas pointwise bounds allow the constant to depend on x. A classic counterexample of a pointwise bounded family that fails to be uniformly bounded is the sequence f_n defined piecewise on [0,1]: f_n(x)=2n^2 x for $0\leq x\leq \frac{1}{2n}, f_n(x)=2n^2(\frac{1}{n}-x) for \frac{1}{2n}<x<\frac{1}{n}, and f_n(x)=0 for \frac{1}{n}\leq x\leq 1. For each fixed x>0, f_n(x)\to 0 as n\to\infty, so \sup_n |f_n(x)|<\infty, and f_n(0)=0; however, \sup_x f_n(x)=n\to\infty as n\to\infty, so the family is not uniformly bounded. This highlights how localized peaks can evade pointwise control while violating uniform bounds.

Advanced Notions

Tight Bounds

In mathematics, an upper bound M for a set S is tight if it is the least upper bound, or , of S, meaning M = \sup S = \inf \{ U \mid U \geq s \ \forall s \in S \}, so no smaller value serves as an upper bound. Similarly, a lower bound m is tight if it is the greatest lower bound, or , of S, with m = \inf S = \sup \{ L \mid L \leq s \ \forall s \in S \}. These tight bounds represent the optimal limits without necessarily being attained by elements of S. In asymptotic analysis, tight bounds describe functions that sandwich another function between constant multiples both above and below, denoted by \Theta notation: f(n) = \Theta(g(n)) if there exist constants c_1, c_2 > 0 and n_0 > 0 such that c_1 g(n) \leq f(n) \leq c_2 g(n) for all n \geq n_0. This contrasts with looser O (upper) or \Omega (lower) bounds and was formalized in algorithm analysis by Donald Knuth in the 1970s through his development of rigorous asymptotic notations. Tight asymptotic bounds enable precise characterization of growth rates, as in the case of Stirling's approximation for the factorial, where n! \sim \sqrt{2\pi n} (n/e)^n provides both upper and lower bounds that converge tightly as n \to \infty, facilitating accurate estimates in combinatorics and probability. In optimization, tight bounds arise in relaxations of integer programs, where the relaxed solution yields a value close to the true optimum, minimizing the or integrality gap. For instance, a relaxation is tight if its optimal value equals that of the original problem, offering strong guarantees for algorithms in fields like scheduling and network design. Tight bounds are essential in algorithm analysis for establishing worst-case performance guarantees, as they pinpoint the exact order of complexity without over- or underestimation, aiding in the design and comparison of efficient algorithms.

Exact Bounds

In partially ordered sets, the exact upper bound of a subset S, known as the supremum and denoted \sup S, is defined as the least upper bound of S, or equivalently, the greatest lower bound of the set of all upper bounds of S. Similarly, the exact lower bound, or infimum \inf S, is the greatest lower bound of S, which is the least upper bound of the set of all lower bounds of S. These concepts extend the notions of to cases where such extrema may not belong to S itself. In the real numbers \mathbb{R}, the completeness axiom guarantees that every nonempty subset S \subseteq \mathbb{R} that is bounded above has a supremum \sup S \in \mathbb{R}, and dually, every nonempty subset bounded below has an infimum \inf S \in \mathbb{R}. For instance, \sup [0,1) = 1, where 1 is an upper bound not attained in the set, and \inf (0,1] = 0, a lower bound outside the set. The construction of \mathbb{R} via Dedekind cuts addresses the incompleteness of the rationals \mathbb{Q}, where bounded nonempty subsets may lack suprema; each cut partitions \mathbb{Q} into a lower set A and upper set B with no greatest element in A, and the real number defined by the cut serves as the supremum of A. This ensures that every such cut has a least upper bound in \mathbb{R}, filling historical gaps in \mathbb{Q}. In more general structures, complete lattices are partially ordered sets where every possesses both a supremum and an infimum, providing a for in arbitrary sets. Bounded subsets of \mathbb{R} inherit this property bilaterally, as \inf S = -\sup (-S) for S \neq \emptyset. A key application arises in : for a monotonic increasing \{a_n\} in \mathbb{R} that is bounded above, the equals \sup \{a_n : n \in \mathbb{N}\}, which coincides with the supremum of any tail \sup \{a_n : n \geq k\} for fixed k. The supremum and infimum thus furnish the tightest possible exact bounds for sets.

References

  1. [1]
    [PDF] Upper and Lower Bounds - Mathematical and Statistical Sciences
    Definition. We say a set S is bounded if S is bounded above and bounded below. That is, there are finite numbers m and M so that m ≤ x ≤ M for all.
  2. [2]
    [PDF] The supremum and infimum - UC Davis Math
    A set is bounded if it is bounded both from above and below. The supremum of a set is its least upper bound and the infimum is its greatest upper bound.
  3. [3]
    [PDF] Completeness of R Math 122 Calculus III
    Definition 1. An upper bound of a set S of real numbers is any real number which is greater or equal to all numbers in S. A lower bound is any.
  4. [4]
    Lower and Upper Bound Theory - GeeksforGeeks
    Jul 11, 2025 · Lower and upper bound theory is a mathematical concept that involves finding the smallest and largest possible values for a quantity, given certain constraints ...
  5. [5]
    Partially Ordered Set -- from Wolfram MathWorld
    An element in a partially ordered set is said to be an upper bound for a subset of if for every , we have . Similarly, a lower bound for a subset is an element ...
  6. [6]
    [PDF] Philosophy of Vagueness: A Topological Perspective
    on any set is denoted by ≤, a partially ordered set(poset) is denoted by (X,≤) ... the set of upper bounds {x ∈ X|∀y ∈ Y, y ≤ x}. Similarly, the greatest ...<|separator|>
  7. [7]
    [PDF] Notes on Ordered Sets
    Sep 22, 2009 · Definition 2.1 We say that a partially ordered set (S, ) has the largest- lower-bound property if inf E exists for every subset E ⊆ S which is ...
  8. [8]
    [PDF] partially ordered sets and lattices
    If a ≤ c, b ≤ c in a partially ordered set P = (X,≤), we say that c is an upper bound of a and b. If d ≤ a, d ≤ b we say d is a lower bound of a and b. We say ...Missing: definition | Show results with:definition
  9. [9]
    [PDF] Lecture 17 - Section 10.1 Least Upper Bound Axiom Section 10.2 ...
    Mar 11, 2008 · Least Upper Bound Theorem. Every nonempty subset S of R with an upper bound has a least upper bound (also called supremum). Jiwen He ...
  10. [10]
    [PDF] Math 127: Posets
    We define upper bound and supremum symmetrically. Note that this definition of lower/upper bound and infimum/supremum is identical to the definition given for ...
  11. [11]
    Bolzano-Weierstrass Theorem -- from Wolfram MathWorld
    Every bounded infinite set in R^n has an accumulation point. For n=1, an infinite subset of a closed bounded set S has an accumulation point in S.
  12. [12]
    Bounded Sequences
    A sequence sn s n of real numbers is called bounded if there exists a real number M∈R M ∈ R such that, for all n∈N, n ∈ N , we have |sn|≤M. | s n | ≤ M ...
  13. [13]
    9.1 Sequences
    Definition 9.1.​​ A sequence is said to be bounded if there exist real numbers and such that m < a n < M for all in . A sequence is said to be unbounded if it is ...
  14. [14]
    [PDF] Summary of Topics: Real Analysis (127A) - UC Davis Math
    Definition 10 (Bounded sequence). A sequence (xn) is bounded if there exists. M ≥ 0 such that |xn| ≤ M for every n ∈ N.
  15. [15]
    [PDF] Chapter 2. Sequences of Real Numbers
    Feb 17, 2024 · Definition. A sequence is bounded if the terms of the sequence form a bounded set. Page 4. 2.1.
  16. [16]
    Monotonic Convergence
    Monotonic Convergence Theorem: If a sequence is monotonic and bounded, if converges. Unboundedness Theorem: If a sequence is not bounded, it diverges.
  17. [17]
    [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 ...
  18. [18]
    [PDF] Math 341 Lecture #8 §2.4: The Monotone Convergence Theorem ...
    Theorem 2.4.2 (Monotone Convergence Theorem). If a sequence is monotone and bounded, then it converges. Proof. Suppose (an) is monotone and bounded.
  19. [19]
    Monotonic Sequences
    The monotone convergence theorem provides a powerful one-two punch that is sufficient to prove that a sequence converges by proving two (probably) ...
  20. [20]
    [PDF] Sequences - UC Davis Math
    A sequence (xn) of real numbers is bounded from above if there exists M ∈ R such that xn ≤ M for all n ∈ N, and bounded from below if there exists m ∈ R such ...
  21. [21]
    Calculus II - Convergence/Divergence of Series
    Aug 13, 2024 · A series is convergent if its partial sums have a finite limit, and divergent if the limit doesn't exist or is infinity. If the limit of the  ...
  22. [22]
    [PDF] Series - UC Davis Math
    This rate of divergence is very slow. It takes 12367 terms for the partial sums of harmonic series to exceed 10, and more than 1.5 × 1043 terms for the partial ...Missing: unbounded | Show results with:unbounded
  23. [23]
    7.6 Absolute Convergence and Error Bounds
    We have shown that if ∑ ( − 1 ) k + 1 a k is a convergent alternating series, then the sum S of the series lies between any two consecutive partial sums . S n .
  24. [24]
    [PDF] MATH1120 Calculus II
    that sn ≤ M for all n. Example 2.3. The harmonic series. ∞. ∑ n=1. 1 n diverges because the sequence of partial sums is unbounded. One can see this by ...
  25. [25]
    [PDF] Alternating Series and Leibniz's Test Let a1,a2,a3,... be a sequence ...
    The partial sums Sn of an alternating series are evidently not monotone,. S1 ... Also S2n+1 = S2n + a2n+1 > S2n for every n, implying that both sequences are.
  26. [26]
    [PDF] The Cauchy criterion 1. Defn: A real sequence {x
    Sep 23, 2021 · Cauchy Convergence Criterion: A real sequence is convergent iff it is Cauchy. ... The series converges if the sequence of partial sums converge.
  27. [27]
    [PDF] Chapter 6: Limits of Functions - UC Davis Math
    The function g : R → R defined by g(x) = x2 is unbounded, but it is bounded on every finite interval [a, b]. We also introduce a notion of being bounded near a ...
  28. [28]
    Bounded Functions - Department of Mathematics at UTSA
    Nov 17, 2021 · A function f defined on some set X with real or complex values is called bounded if the set of its values is bounded.
  29. [29]
    [PDF] Continuous Functions - UC Davis Math
    The non-uniformly continuous functions in the last two examples were un- bounded. However, even bounded continuous functions can fail to be uniformly continuous ...<|control11|><|separator|>
  30. [30]
    bounded function - PlanetMath
    Mar 22, 2013 · supremum norm. Defines, sup norm. Defines, sup-norm. Defines, uniform norm. Defines, bounded function. Defines, unbounded function. Generated on ...
  31. [31]
    Supremum Norm -- from Wolfram MathWorld
    The supremum norm is the norm defined on F by ||f||=sup_(x in K)|f(x)|. Then F is a commutative Banach algebra with identity.
  32. [32]
    [PDF] arXiv:2101.04555v1 [math.FA] 12 Jan 2021
    12 ene 2021 · The Uniform Boundedness Principle, also known as Banach-Steinhaus theorem, is one of the fundamental results in functional analysis which ...<|separator|>
  33. [33]
    11.6 Equicontinuity and the Arzelà–Ascoli theorem
    Basically, a uniformly equicontinuous sequence in the metric space C ( X , C ) that is pointwise bounded is bounded (in C ( X , C ) ) and furthermore contains a ...Missing: precompact | Show results with:precompact
  34. [34]
    [PDF] Chapter 9: Sequences and Series of Functions - UC Davis Math
    Thus, a pointwise convergent sequence (fn) of functions need not be uniformly bounded. (that is, bounded independently of n), even if it converges to zero.Missing: counterexample | Show results with:counterexample
  35. [35]
    [PDF] Project Gutenberg's Essays on the Theory of Numbers, by Richard ...
    ON CONTINUITY AND IRRATIONAL NUMBERS, and ON THE NATURE AND. MEANING OF NUMBERS. By R. Dedekind. From the German by W. W.. Beman. Pages, 115.