Limit inferior and limit superior

Last updated

In mathematics, the limit inferior and limit superior of a sequence can be thought of as limiting (that is, eventual and extreme) bounds on the sequence. They can be thought of in a similar fashion for a function (see limit of a function). For a set, they are the infimum and supremum of the set's limit points, respectively. In general, when there are multiple objects around which a sequence, function, or set accumulates, the inferior and superior limits extract the smallest and largest of them; the type of object and the measure of size is context-dependent, but the notion of extreme limits is invariant. Limit inferior is also called infimum limit, limit infimum, liminf, inferior limit, lower limit, or inner limit; limit superior is also known as supremum limit, limit supremum, limsup, superior limit, upper limit, or outer limit.

Contents

An illustration of limit superior and limit inferior. The sequence xn is shown in blue. The two red curves approach the limit superior and limit inferior of xn, shown as dashed black lines. In this case, the sequence accumulates around the two limits. The superior limit is the larger of the two, and the inferior limit is the smaller. The inferior and superior limits agree if and only if the sequence is convergent (i.e., when there is a single limit). Lim sup example 5.png
An illustration of limit superior and limit inferior. The sequence xn is shown in blue. The two red curves approach the limit superior and limit inferior of xn, shown as dashed black lines. In this case, the sequence accumulates around the two limits. The superior limit is the larger of the two, and the inferior limit is the smaller. The inferior and superior limits agree if and only if the sequence is convergent (i.e., when there is a single limit).

The limit inferior of a sequence is denoted by

and the limit superior of a sequence is denoted by

Definition for sequences

The limit inferior of a sequence (xn) is defined by

or

Similarly, the limit superior of (xn) is defined by

or

Alternatively, the notations and are sometimes used.

The limits superior and inferior can equivalently be defined using the concept of subsequential limits of the sequence . [1] An element of the extended real numbers is a subsequential limit of if there exists a strictly increasing sequence of natural numbers such that . If is the set of all subsequential limits of , then

and

If the terms in the sequence are real numbers, the limit superior and limit inferior always exist, as the real numbers together with ±∞ (i.e. the extended real number line) are complete. More generally, these definitions make sense in any partially ordered set, provided the suprema and infima exist, such as in a complete lattice.

Whenever the ordinary limit exists, the limit inferior and limit superior are both equal to it; therefore, each can be considered a generalization of the ordinary limit which is primarily interesting in cases where the limit does not exist. Whenever liminfxn and limsupxn both exist, we have

The limits inferior and superior are related to big-O notation in that they bound a sequence only "in the limit"; the sequence may exceed the bound. However, with big-O notation the sequence can only exceed the bound in a finite prefix of the sequence, whereas the limit superior of a sequence like en may actually be less than all elements of the sequence. The only promise made is that some tail of the sequence can be bounded above by the limit superior plus an arbitrarily small positive constant, and bounded below by the limit inferior minus an arbitrarily small positive constant.

The limit superior and limit inferior of a sequence are a special case of those of a function (see below).

The case of sequences of real numbers

In mathematical analysis, limit superior and limit inferior are important tools for studying sequences of real numbers. Since the supremum and infimum of an unbounded set of real numbers may not exist (the reals are not a complete lattice), it is convenient to consider sequences in the affinely extended real number system: we add the positive and negative infinities to the real line to give the complete totally ordered set [−∞,∞], which is a complete lattice.

Interpretation

Consider a sequence consisting of real numbers. Assume that the limit superior and limit inferior are real numbers (so, not infinite).

Properties

In case the sequence is bounded, for all
[?]
>
0
{\displaystyle \epsilon >0}
almost all sequence members lie in the open interval
(
lim inf
n
-
[?]
x
n
-
[?]
,
lim sup
n
-
[?]
x
n
+
[?]
)
.
{\displaystyle (\liminf _{n\to \infty }x_{n}-\epsilon ,\limsup _{n\to \infty }x_{n}+\epsilon ).} Illustration of Limes superior and Limes Inferior.svg
In case the sequence is bounded, for all almost all sequence members lie in the open interval

The relationship of limit inferior and limit superior for sequences of real numbers is as follows:

As mentioned earlier, it is convenient to extend to Then, in converges if and only if

in which case is equal to their common value. (Note that when working just in convergence to or would not be considered as convergence.) Since the limit inferior is at most the limit superior, the following conditions hold

If and , then the interval need not contain any of the numbers but every slight enlargement for arbitrarily small will contain for all but finitely many indices In fact, the interval is the smallest closed interval with this property. We can formalize this property like this: there exist subsequences and of (where and are increasing) for which we have

On the other hand, there exists a so that for all

To recapitulate:

Conversely, it can also be shown that:

In general,

The liminf and limsup of a sequence are respectively the smallest and greatest cluster points. [3]

Analogously, the limit inferior satisfies superadditivity:

In the particular case that one of the sequences actually converges, say then the inequalities above become equalities (with or being replaced by ).

hold whenever the right-hand side is not of the form

If exists (including the case ) and then provided that is not of the form

Examples

  • As an example, consider the sequence given by the sine function: Using the fact that π is irrational, it follows that
    and
    (This is because the sequence is equidistributed mod 2π, a consequence of the equidistribution theorem.)
  • An example from number theory is
    where is the -th prime number.
The value of this limit inferior is conjectured to be 2 – this is the twin prime conjecture – but as of April 2014 has only been proven to be less than or equal to 246. [4] The corresponding limit superior is , because there are arbitrarily large gaps between consecutive primes.

Real-valued functions

Assume that a function is defined from a subset of the real numbers to the real numbers. As in the case for sequences, the limit inferior and limit superior are always well-defined if we allow the values +∞ and −∞; in fact, if both agree then the limit exists and is equal to their common value (again possibly including the infinities). For example, given , we have and . The difference between the two is a rough measure of how "wildly" the function oscillates, and in observation of this fact, it is called the oscillation of f at 0. This idea of oscillation is sufficient to, for example, characterize Riemann-integrable functions as continuous except on a set of measure zero. [5] Note that points of nonzero oscillation (i.e., points at which f is "badly behaved") are discontinuities which, unless they make up a set of zero, are confined to a negligible set.

Functions from topological spaces to complete lattices

Functions from metric spaces

There is a notion of limsup and liminf for functions defined on a metric space whose relationship to limits of real-valued functions mirrors that of the relation between the limsup, liminf, and the limit of a real sequence. Take a metric space , a subspace contained in , and a function . Define, for any limit point of ,

and

where denotes the metric ball of radius about .

Note that as ε shrinks, the supremum of the function over the ball is non increasing ( strictly decreasing or remaining the same), so we have

and similarly

Functions from topological spaces

This finally motivates the definitions for general topological spaces. Take X, E and a as before, but now let X be a topological space. In this case, we replace metric balls with neighborhoods:

(there is a way to write the formula using "lim" using nets and the neighborhood filter). This version is often useful in discussions of semi-continuity which crop up in analysis quite often. An interesting note is that this version subsumes the sequential version by considering sequences as functions from the natural numbers as a topological subspace of the extended real line, into the space (the closure of N in [−∞,∞], the extended real number line, is N  {∞}.)

Sequences of sets

The power set ℘(X) of a set X is a complete lattice that is ordered by set inclusion, and so the supremum and infimum of any set of subsets (in terms of set inclusion) always exist. In particular, every subset Y of X is bounded above by X and below by the empty set ∅ because ∅ ⊆ YX. Hence, it is possible (and sometimes useful) to consider superior and inferior limits of sequences in ℘(X) (i.e., sequences of subsets of X).

There are two common ways to define the limit of sequences of sets. In both cases:

The difference between the two definitions involves how the topology (i.e., how to quantify separation) is defined. In fact, the second definition is identical to the first when the discrete metric is used to induce the topology on X.

General set convergence

A sequence of sets in a metrizable space approaches a limiting set when the elements of each member of the sequence approach the elements of the limiting set. In particular, if is a sequence of subsets of then:

  • which is also called the outer limit, consists of those elements which are limits of points in taken from (countably) infinitely many That is, if and only if there exists a sequence of points and a subsequence of such that and
  • which is also called the inner limit, consists of those elements which are limits of points in for all but finitely many (that is, cofinitely many ). That is, if and only if there exists a sequence of points such that and

The limit exists if and only if and agree, in which case [6] The outer and inner limits should not be confused with the set-theoretic limits superior and inferior, as the latter sets are not sensitive to the topological structure of the space.

Special case: discrete metric

This is the definition used in measure theory and probability. Further discussion and examples from the set-theoretic point of view, as opposed to the topological point of view discussed below, are at set-theoretic limit.

By this definition, a sequence of sets approaches a limiting set when the limiting set includes elements which are in all except finitely many sets of the sequence and does not include elements which are in all except finitely many complements of sets of the sequence. That is, this case specializes the general definition when the topology on set X is induced from the discrete metric.

Specifically, for points x, yX, the discrete metric is defined by

under which a sequence of points (xk) converges to point xX if and only if xk = x for all but finitely many k. Therefore, if the limit set exists it contains the points and only the points which are in all except finitely many of the sets of the sequence. Since convergence in the discrete metric is the strictest form of convergence (i.e., requires the most), this definition of a limit set is the strictest possible.

If (Xn) is a sequence of subsets of X, then the following always exist:

Observe that x ∈ limsupXn if and only if x ∉ liminfXnc.

In this sense, the sequence has a limit so long as every point in X either appears in all except finitely many Xn or appears in all except finitely many Xnc. [7]

Using the standard parlance of set theory, set inclusion provides a partial ordering on the collection of all subsets of X that allows set intersection to generate a greatest lower bound and set union to generate a least upper bound. Thus, the infimum or meet of a collection of subsets is the greatest lower bound while the supremum or join is the least upper bound. In this context, the inner limit, liminfXn, is the largest meeting of tails of the sequence, and the outer limit, limsupXn, is the smallest joining of tails of the sequence. The following makes this precise.

The sequence (In) is non-decreasing (i.e. InIn+1) because each In+1 is the intersection of fewer sets than In. The least upper bound on this sequence of meets of tails is
So the limit infimum contains all subsets which are lower bounds for all but finitely many sets of the sequence.
The sequence (Jn) is non-increasing (i.e. JnJn+1) because each Jn+1 is the union of fewer sets than Jn. The greatest lower bound on this sequence of joins of tails is
So the limit supremum is contained in all subsets which are upper bounds for all but finitely many sets of the sequence.

Examples

The following are several set convergence examples. They have been broken into sections with respect to the metric used to induce the topology on set X.

Using the discrete metric
Using either the discrete metric or the Euclidean metric
The "odd" and "even" elements of this sequence form two subsequences, ({0}, {0}, {0}, ...) and ({1}, {1}, {1}, ...), which have limit points 0 and 1, respectively, and so the outer or superior limit is the set {0,1} of these two points. However, there are no limit points that can be taken from the (Xn) sequence as a whole, and so the interior or inferior limit is the empty set { }. That is,
  • limsupXn = {0,1}
  • liminfXn = { }
However, for (Yn) = ({0}, {0}, {0}, ...) and (Zn) = ({1}, {1}, {1}, ...):
  • limsupYn = liminfYn = limYn = {0}
  • limsupZn = liminfZn = limZn = {1}
As in the previous two examples,
  • limsupXn = {0,1}
  • liminfXn = { }
That is, the four elements that do not match the pattern do not affect the liminf and limsup because there are only finitely many of them. In fact, these elements could be placed anywhere in the sequence. So long as the tails of the sequence are maintained, the outer and inner limits will be unchanged. The related concepts of essential inner and outer limits, which use the essential supremum and essential infimum, provide an important modification that "squashes" countably many (rather than just finitely many) interstitial additions.
Using the Euclidean metric
The "odd" and "even" elements of this sequence form two subsequences, ({0}, {1/2}, {2/3}, {3/4}, ...) and ({1}, {1/2}, {1/3}, {1/4}, ...), which have limit points 1 and 0, respectively, and so the outer or superior limit is the set {0,1} of these two points. However, there are no limit points that can be taken from the (Xn) sequence as a whole, and so the interior or inferior limit is the empty set { }. So, as in the previous example,
  • limsupXn = {0,1}
  • liminfXn = { }
However, for (Yn) = ({0}, {1/2}, {2/3}, {3/4}, ...) and (Zn) = ({1}, {1/2}, {1/3}, {1/4}, ...):
  • limsupYn = liminfYn = limYn = {1}
  • limsupZn = liminfZn = limZn = {0}
In each of these four cases, the elements of the limiting sets are not elements of any of the sets from the original sequence.
  • For example, an LTI system that is the cascade connection of several stable systems with an undamped second-order LTI system (i.e., zero damping ratio) will oscillate endlessly after being perturbed (e.g., an ideal bell after being struck). Hence, if the position and velocity of this system are plotted against each other, trajectories will approach a circle in the state space. This circle, which is the Ω limit set of the system, is the outer limit of solution trajectories of the system. The circle represents the locus of a trajectory corresponding to a pure sinusoidal tone output; that is, the system output approaches/approximates a pure tone.

Generalized definitions

The above definitions are inadequate for many technical applications. In fact, the definitions above are specializations of the following definitions.

Definition for a set

The limit inferior of a set X Y is the infimum of all of the limit points of the set. That is,

Similarly, the limit superior of X is the supremum of all of the limit points of the set. That is,

Note that the set X needs to be defined as a subset of a partially ordered set Y that is also a topological space in order for these definitions to make sense. Moreover, it has to be a complete lattice so that the suprema and infima always exist. In that case every set has a limit superior and a limit inferior. Also note that the limit inferior and the limit superior of a set do not have to be elements of the set.

Definition for filter bases

Take a topological space X and a filter base B in that space. The set of all cluster points for that filter base is given by

where is the closure of . This is clearly a closed set and is similar to the set of limit points of a set. Assume that X is also a partially ordered set. The limit superior of the filter base B is defined as

when that supremum exists. When X has a total order, is a complete lattice and has the order topology,

Similarly, the limit inferior of the filter base B is defined as

when that infimum exists; if X is totally ordered, is a complete lattice, and has the order topology, then

If the limit inferior and limit superior agree, then there must be exactly one cluster point and the limit of the filter base is equal to this unique cluster point.

Specialization for sequences and nets

Note that filter bases are generalizations of nets, which are generalizations of sequences. Therefore, these definitions give the limit inferior and limit superior of any net (and thus any sequence) as well. For example, take topological space and the net , where is a directed set and for all . The filter base ("of tails") generated by this net is defined by

Therefore, the limit inferior and limit superior of the net are equal to the limit superior and limit inferior of respectively. Similarly, for topological space , take the sequence where for any . The filter base ("of tails") generated by this sequence is defined by

Therefore, the limit inferior and limit superior of the sequence are equal to the limit superior and limit inferior of respectively.

See also

Related Research Articles

In mathematics, the infimum of a subset of a partially ordered set is the greatest element in that is less than or equal to each element of if such an element exists. In other words, it is the greatest element of that is lower or equal to the lowest element of . Consequently, the term greatest lower bound is also commonly used. The supremum of a subset of a partially ordered set is the least element in that is greater than or equal to each element of if such an element exists. In other words, it is the least element of that is greater than or equal to the greatest element of . Consequently, the supremum is also referred to as the least upper bound.

In probability theory, the Borel–Cantelli lemma is a theorem about sequences of events. In general, it is a result in measure theory. It is named after Émile Borel and Francesco Paolo Cantelli, who gave statement to the lemma in the first decades of the 20th century. A related result, sometimes called the second Borel–Cantelli lemma, is a partial converse of the first Borel–Cantelli lemma. The lemma states that, under certain conditions, an event will have probability of either zero or one. Accordingly, it is the best-known of a class of similar theorems, known as zero-one laws. Other examples include Kolmogorov's zero–one law and the Hewitt–Savage zero–one law.

In probability theory, there exist several different notions of convergence of sequences of random variables. The different notions of convergence capture different properties about the sequence, with some notions of convergence being stronger than others. For example, convergence in distribution tells us about the limit distribution of a sequence of random variables. This is a weaker notion than convergence in probability, which tells us about the value a random variable will take, rather than just the distribution.

In the mathematical field of real analysis, the monotone convergence theorem is any of a number of related theorems proving the convergence of monotonic sequences that are also bounded. Informally, the theorems state that if a sequence is increasing and bounded above by a supremum, then the sequence will converge to the supremum; in the same way, if a sequence is decreasing and is bounded below by an infimum, it will converge to the infimum.

In mathematics, Fatou's lemma establishes an inequality relating the Lebesgue integral of the limit inferior of a sequence of functions to the limit inferior of integrals of these functions. The lemma is named after Pierre Fatou.

In mathematics, the limit of a sequence of sets is a set whose elements are determined by the sequence in either of two equivalent ways: (1) by upper and lower bounds on the sequence that converge monotonically to the same set and (2) by convergence of a sequence of indicator functions which are themselves real-valued. As is the case with sequences of other objects, convergence is not necessary or even usual.

<span class="mw-page-title-main">Oscillation (mathematics)</span> Amount of variation between extrema

In mathematics, the oscillation of a function or a sequence is a number that quantifies how much that sequence or function varies between its extreme values as it approaches infinity or a point. As is the case with limits, there are several definitions that put the intuitive concept into a form suitable for a mathematical treatment: oscillation of a sequence of real numbers, oscillation of a real-valued function at a point, and oscillation of a function on an interval.

In number theory, natural density, also referred to as asymptotic density or arithmetic density, is one method to measure how "large" a subset of the set of natural numbers is. It relies chiefly on the probability of encountering members of the desired subset when combing through the interval [1, n] as n grows large.

In mathematics and, specifically, real analysis, the Dini derivatives are a class of generalizations of the derivative. They were introduced by Ulisse Dini, who studied continuous but nondifferentiable functions.

In mathematics, the Stolz–Cesàro theorem is a criterion for proving the convergence of a sequence. The theorem is named after mathematicians Otto Stolz and Ernesto Cesàro, who stated and proved it for the first time.

In the field of mathematical analysis for the calculus of variations, Γ-convergence (Gamma-convergence) is a notion of convergence for functionals. It was introduced by Ennio de Giorgi.

In mathematics, the Fatou–Lebesgue theorem establishes a chain of inequalities relating the integrals of the limit inferior and the limit superior of a sequence of functions to the limit inferior and the limit superior of integrals of these functions. The theorem is named after Pierre Fatou and Henri Léon Lebesgue.

In mathematics, more specifically measure theory, there are various notions of the convergence of measures. For an intuitive general sense of what is meant by convergence of measures, consider a sequence of measures μn on a space, sharing a common collection of measurable sets. Such a sequence might represent an attempt to construct 'better and better' approximations to a desired measure μ that is difficult to obtain directly. The meaning of 'better and better' is subject to all the usual caveats for taking limits; for any error tolerance ε > 0 we require there be N sufficiently large for nN to ensure the 'difference' between μn and μ is smaller than ε. Various notions of convergence specify precisely what the word 'difference' should mean in that description; these notions are not equivalent to one another, and vary in strength.

In mathematical analysis, Mosco convergence is a notion of convergence for functionals that is used in nonlinear analysis and set-valued analysis. It is a particular case of Γ-convergence. Mosco convergence is sometimes phrased as “weak Γ-liminf and strong Γ-limsup” convergence since it uses both the weak and strong topologies on a topological vector space X. In finite dimensional spaces, Mosco convergence coincides with epi-convergence, while in infinite-dimensional ones, Mosco convergence is strictly stronger property.

In mathematics, the limit comparison test (LCT) (in contrast with the related direct comparison test) is a method of testing for the convergence of an infinite series.

In mathematics, Schilder's theorem is a generalization of the Laplace method from integrals on to functional Wiener integration. The theorem is used in the large deviations theory of stochastic processes. Roughly speaking, out of Schilder's theorem one gets an estimate for the probability that a (scaled-down) sample path of Brownian motion will stray far from the mean path. This statement is made precise using rate functions. Schilder's theorem is generalized by the Freidlin–Wentzell theorem for Itō diffusions.

In mathematics, the Cauchy–Hadamard theorem is a result in complex analysis named after the French mathematicians Augustin Louis Cauchy and Jacques Hadamard, describing the radius of convergence of a power series. It was published in 1821 by Cauchy, but remained relatively unknown until Hadamard rediscovered it. Hadamard's first publication of this result was in 1888; he also included it as part of his 1892 Ph.D. thesis.

In mathematics, specifically in number theory, the extremal orders of an arithmetic function are best possible bounds of the given arithmetic function. Specifically, if f(n) is an arithmetic function and m(n) is a non-decreasing function that is ultimately positive and

This article is supplemental for “Convergence of random variables” and provides proofs for selected results.

In probability theory, Kolmogorov's two-series theorem is a result about the convergence of random series. It follows from Kolmogorov's inequality and is used in one proof of the strong law of large numbers.

References

  1. Rudin, W. (1976). Principles of Mathematical Analysis. New York: McGraw-Hill. p. 56. ISBN   007054235X.
  2. Gleason, Andrew M. (1992). Fundamentals of abstract analysis. Boca Raton, FL. pp. 176–177. ISBN   978-1-4398-6481-4. OCLC   1074040561.{{cite book}}: CS1 maint: location missing publisher (link)
  3. Gleason, Andrew M. (1992). Fundamentals of abstract analysis. Boca Raton, FL. pp. 160–182. ISBN   978-1-4398-6481-4. OCLC   1074040561.{{cite book}}: CS1 maint: location missing publisher (link)
  4. "Bounded gaps between primes". Polymath wiki. Retrieved 14 May 2014.[ unreliable source? ]
  5. "Lebesgue's Criterion for Riemann integrability (MATH314 Lecture Notes)" (PDF). University of Windsor. Archived from the original (PDF) on 2007-03-03. Retrieved 2006-02-24.
  6. 1 2 Goebel, Rafal; Sanfelice, Ricardo G.; Teel, Andrew R. (2009). "Hybrid dynamical systems". IEEE Control Systems Magazine. 29 (2): 28–93. doi:10.1109/MCS.2008.931718.
  7. Halmos, Paul R. (1950). Measure Theory. Princeton, NJ: D. Van Nostrand Company, Inc.