Sunteți pe pagina 1din 7

Quasi-Cauchy Sequences

Author(s): David Burton and John Coleman


Source: The American Mathematical Monthly, Vol. 117, No. 4 (April 2010), pp. 328-333
Published by: Mathematical Association of America
Stable URL: http://www.jstor.org/stable/10.4169/000298910X480793 .
Accessed: 30/03/2013 21:48

Your use of the JSTOR archive indicates your acceptance of the Terms & Conditions of Use, available at .
http://www.jstor.org/page/info/about/policies/terms.jsp

.
JSTOR is a not-for-profit service that helps scholars, researchers, and students discover, use, and build upon a wide range of
content in a trusted digital archive. We use information technology and tools to increase productivity and facilitate new forms
of scholarship. For more information about JSTOR, please contact support@jstor.org.

Mathematical Association of America is collaborating with JSTOR to digitize, preserve and extend access to
The American Mathematical Monthly.

http://www.jstor.org

This content downloaded from 171.67.34.69 on Sat, 30 Mar 2013 21:48:11 PM


All use subject to JSTOR Terms and Conditions
Quasi-Cauchy Sequences
David Burton and John Coleman

Beginning students often misunderstand the definition of Cauchy sequences when they
first encounter it in an introductory real analysis course. In particular, many students
fail to understand that it involves far more than that the distance between successive
terms is tending to zero. Nevertheless, sequences which satisfy this weaker property
are interesting in their own right. We call them quasi-Cauchy.

1. ON THE REAL LINE. In this section we investigate quasi-Cauchy sequences in


R, the real number system. For the record:

Definition 1. Suppose that xn  (n = 1, 2, 3, . . .) is a sequence of real numbers.


1. xn  is Cauchy if given any  > 0 there exists an integer K > 0 such that m, n ≥
K implies |xm − xn | < .
2. xn  is quasi-Cauchy if given any  > 0 there exists an integer K > 0 such that
n ≥ K implies |xn+1 − xn | < .

Trivially, Cauchy sequences are quasi-Cauchy. The converse is easily seen to be


false. The most quotable (though not the simplest) counterexample is provided by the
sequence of partial sums of the harmonic series.
Cauchy sequences have the property that any subsequence of a Cauchy sequence
is Cauchy. The analogous property fails for quasi-Cauchy sequences. In fact, it is pre-
cisely the Cauchy sequences which have the property that all of their subsequences
are quasi-Cauchy. The subsequences of a quasi-Cauchy sequence can be somewhat
arbitrary, as the following lemma illustrates.

Lemma 1. Suppose that I is an interval and a1 , b1 , a2 , b2 , . . . is a sequence of


ordered pairs in I with limi→∞ |ai − bi | = 0. Then there exists a quasi-Cauchy se-
quence xi  with the property that for any integer i ≥ 1 there exists a j ≥ 1 such that
ai , bi  = x j , x j +1 .

Proof. For every k ≥ 1, fix y0k , y1k , . . . , ynkk in I with y0k = bk , ynkk = ak+1 , and
|yik − yi−1
k
| < 1k for 1 ≤ i ≤ n k . Then the sequence

a1 , b1 , y01 , y11 , . . . , yn11 , a2 , b2 , y02 , y12 , . . . , yn22 , a3 , b3 , . . .

clearly has the desired property.

There are interesting connections between uniformly continuous functions and


quasi-Cauchy sequences. It is easy to see that if I is a closed interval then a function
f is continuous on I if and only if it is defined on I and preserves Cauchy sequences
from I (xn  Cauchy with {xn } ⊂ I implies  f (xn ) Cauchy). Somewhat similarly:

Theorem 1. Suppose that I is any interval. Then a real-valued function is uniformly


continuous on I if and only if it is defined on I and preserves quasi-Cauchy sequences
from I .
doi:10.4169/000298910X480793

328
c THE MATHEMATICAL ASSOCIATION OF AMERICA [Monthly 117

This content downloaded from 171.67.34.69 on Sat, 30 Mar 2013 21:48:11 PM


All use subject to JSTOR Terms and Conditions
Proof. It is an easy exercise to check that uniformly continuous functions preserve
quasi-Cauchy sequences. Conversely, suppose that f defined on I is not uniformly
continuous. Then there exists an  > 0 such that for any δ > 0 there exist a, b ∈ I
with |a − b| < δ but | f (a) − f (b)| ≥ . For every integer n ≥ 1 fix an , bn ∈ I with
|an − bn | < 1/n and | f (an ) − f (bn )| ≥ . By the lemma, there exists a quasi-Cauchy
sequence xi  such that for any integer i ≥ 1 there exists a j with ai = x j and bi =
x j +1 . This implies that | f (x j ) − f (x j +1 )| ≥ ; hence  f (xi ) is not quasi-Cauchy.
Thus f does not preserve quasi-Cauchy sequences.

The preceding theorem can be strengthened in the special case that I is bounded:

Theorem 2. Suppose that f is a function defined on a bounded interval I . Then f is


uniformly continuous on I if and only if the image under f of any Cauchy sequence in
I is quasi-Cauchy.

Proof. By Theorem 1, if f is uniformly continuous on I then the image of any


quasi-Cauchy sequence in I is quasi-Cauchy, and hence in particular the image of
any Cauchy sequence is quasi-Cauchy. For the converse, suppose that the image of
every Cauchy sequence is quasi-Cauchy but that f fails to be uniformly continu-
ous. Then there is an  > 0 such that for any δ > 0 there exist points x, y ∈ I with
|x − y| < δ but | f (x) − f (y)| ≥ . For each n ≥ 1, fix xn , yn ∈ I with |xn − yn | < n1
but | f (xn ) − f (yn )| ≥ . Since I is bounded, xn  has a convergent subsequence by
the Bolzano-Weierstrass theorem. Thus by passing to a subsequence if necessary, we
can without loss of generality assume that xn  converges. Then x1 , y1 , x2 , y2 , x3 , y3 . . .
is also convergent, hence Cauchy, but the image sequence f (x1 ), f (y1 ), f (x2 ), f (y2 ),
f (x3 ), f (y3 ), . . . fails to be quasi-Cauchy, a contradiction.

2. IN GENERAL METRIC SPACES.

Definition 2. Suppose that X is a set and d : X 2 → [0, ∞] is a function.


1. d is called a pseudometric if it satisfies d(x, x) = 0, d(x, y) = d(y, x), and
d(x, y) ≤ d(x, z) + d(z, y) for all x, y, z ∈ X .
2. d is called a metric if it also satisfies d(x, y) < ∞ and d(x, y) = 0 ⇒ x = y
for all x, y ∈ X .

The notion of quasi-Cauchy makes perfect sense in any metric space (X, d) (af-
ter the purely typographical substitution of d(x, y) for |x − y|). The main difference
is that, for some important metric spaces, the distinction between quasi-Cauchy and
Cauchy vanishes. Let us coin a term for this phenomenon:

Definition 3. A metric space (X, d) is called nonincremental if every quasi-Cauchy


sequence in X is Cauchy. In this case d itself is called a nonicremental metric.

The intuition is that in a nonincremental space you can’t travel far via the arbitrarily
small increments between successive terms in a quasi-Cauchy sequence. Nonincre-
mental spaces exist: take the set of integers along with the Euclidean metric. Are there
any interesting examples? Yes: in [2] they are shown to arise naturally in certain classes
of topological algebras. Furthermore, they arise in the theory of ultrametric spaces:

April 2010] QUASI-CAUCHY SEQUENCES 329

This content downloaded from 171.67.34.69 on Sat, 30 Mar 2013 21:48:11 PM


All use subject to JSTOR Terms and Conditions
Definition 4. An ultrametric space is a metric space (X, d) which satisfies the follow-
ing strengthening of the triangle inequality (called the ultrametric inequality):

d(x, y) ≤ sup{d(x, z), d(z, y)} for all x, y, z ∈ X.

Ultrametric spaces are also called non-Archimedean spaces or isosceles spaces. The
latter is because they have the interesting property that in them all triangles are isosce-
les. They arise naturally in the study of p-adic numbers and in other areas of analysis
(e.g., [1]). More recently, they have found applications in theoretical computer science
(e.g., [5]) and other areas. The following theorem is part of the folklore:

Theorem 3. Ultrametric spaces are nonincremental.

Proof. Suppose that (X, d) is an ultrametric space. Suppose that xn  is a quasi-
Cauchy sequence in X . Suppose  > 0 is given. Fix K > 0 such that n ≥ K implies
d(xn , xn+1 ) < . Suppose that m, n ≥ K with m ≤ n. Then repeated application of
the ultrametric inequality yields

d(xm , xn ) ≤ sup{d(xm , xm+1 ), d(xm+1 , xm+2 ), . . . , d(xn−1 , xn )}


< sup{, , . . . , } = .

Thus, xn  is Cauchy as well.

The idea behind the proof is that for ultrametric spaces the relation between points
of being within  of each other is a transitive relation. Thus for any  > 0 the collection
of -balls forms a partition of the space.
A natural question is if the converse of this last theorem is true. The answer is no.
For a simple example, let X be a three-element set consisting of the vertices of any
nonisosceles triangle, and let d be the distance between the vertices. Then (X, d) is
not an ultrametric space, but it is nonincremental since any quasi-Cauchy sequence is
eventually constant and hence Cauchy as well. There is an interesting partial converse,
involving notions of metric equivalence.

Definition 5. Suppose that (X 1 , d1 ) and (X 2 , d2 ) are pseudometric spaces. We say that


they are
1. topologically equivalent if there exists a homeomorphism h : X 1 → X 2 ;
2. uniformly equivalent if there exists a bijection h : X 1 → X 2 such that both h and
h −1 are uniformly continuous with respect to the given pseudometrics. Such an
h is called a uniform equivalence.

In the special case that X 1 = X 2 = X (so that d1 and d2 are pseudometrics on the
same space) we will simply say that d1 and d2 are topologically or uniformly equivalent
if the identity mapping is either a homeomorphism or a uniform equivalence. Note
that all 3-element metric spaces (including equilateral ones) are uniformly equivalent
to each other. Thus the property of being an ultrametric space is not preserved by
uniform equivalence. On the other hand, it is easy to see that the property of being
nonincremental is preserved by uniform equivalence.
There is a partial converse to the previous theorem:

Theorem 4. Suppose that (X, d) is a nonincremental space. Then it is topologically


equivalent to an ultrametric space.

330
c THE MATHEMATICAL ASSOCIATION OF AMERICA [Monthly 117

This content downloaded from 171.67.34.69 on Sat, 30 Mar 2013 21:48:11 PM


All use subject to JSTOR Terms and Conditions
Proof. If x, y ∈ X , say that x and y are -connected if there exist points x0 , x1 , . . . , xn
with x0 = x, xn = y, and d(xi , xi+1 ) <  for i < n. Such a sequence of points is called
an -chain connecting x and y. For x, y ∈ X let

d ∗ (x, y) = inf{ | x and y are -connected}.

Note that d ∗ (x, y) <  implies that x and y are -connected (with respect to d). It
is easy to check that d ∗ (x, y) ≥ 0, d ∗ (x, y) = d ∗ (y, x), and d ∗ (x, y) ≤ sup{d ∗ (x, z),
d ∗ (z, y)} for all x, y, z ∈ X . This last inequality follows from the observation that
the relation of being -connected is transitive. Thus d ∗ is an ultra-pseudometric (a
pseudometric which satisfies the ultrametric inequality). We claim that d ∗ and d are
equivalent. Note that establishing this claim automatically yields that d ∗ is in fact a
metric since any pseudometric equivalent to a metric is itself a metric.
Let Bd (x, ) = {y | d(x, y) < }, the -ball with respect to d centered at x. Define
Bd ∗ (x, ) in the analogous way. Note that d ∗ (x, y) ≤ d(x, y) for all x, y. This implies
that for any x ∈ X and any  > 0,

Bd (x, ) ⊆ Bd ∗ (x, ).

Thus the d topology is finer than the d ∗ topology. On the other hand, suppose x ∈ X
and  > 0. We need to show that there exists a δ > 0 with Bd ∗ (x, δ) ⊆ Bd (x, ).
Suppose no such δ exists. Then for every integer n ≥ 1 we can find a point an ∈
Bd ∗ (x, 1/n) with an  ∈ Bd (x, ). Since d ∗ (x, an ) < 1/n we can fix a (1/n)-chain
x0 , x1 , . . . xk connecting x to an . Thus x0 = x, xk = an , and d(xi , xi+1 ) < 1/n for
i < k. Call

x0 , x1 , . . . , xk−1 , xk , xk−1 , . . . , x1 , x0

the (1/n)-circuit. It leads from x to a point outside of Bd (x, ) and then back again
with steps of length < 1/n. Now define a sequence yi  by starting with the 1-circuit,
following it by the (1/2)-circuit, then the (1/3)-circuit, etc. Since the circuits begin
and end at x, d(yi , yi+1 ) = 0 if yi and yi+1 straddle the boundary between successive
circuits. On the other hand, if yi and yi+1 are both inside a circuit (say the (1/n)-
circuit) then d(yi , yi+1 ) < 1/n. Note that n → ∞ as i → ∞. Thus yi  is a quasi-
Cauchy sequence in (X, d). It is not Cauchy since for any K > 0 you can find i, j ≥
K with d(yi , y j ) ≥ : pick i ≥ K so that yi = x is the start of a circuit and pick
j > i so that y j is the midpoint of that circuit. This contradicts the fact that (X, d) is
nonincremental, so the required δ must exist after all.

At this stage we know that if (X, d) is a metric space then d uniformly equivalent to
an ultrametric implies that d is nonincremental which in turn implies that d is equiva-
lent to an ultrametric. It is natural to ask if either of these implications can be reversed.
We will give two examples to show that they can’t.
For the first example, note that any discrete metric space is equivalent to an ultramet-
ric space: if (X, d) is discrete (for any point x there is an  > 0 with Bd (x, ) = {x})
then d ∗ given by x  = y implies d ∗ (x, y) = 1 is an equivalent ultrametric on X . If we
let X be the set of partial sums of the harmonic series with d the induced Euclidean
metric on X , then X is clearly discrete but fails to be nonincremental. Thus being
equivalent to an ultrametric does not imply being nonincremental.
The second example is more involved. Suppose that (X, d) is uniformly equiva-
lent to an ultrametric space. Then there exists an ultrametric d ∗ on X with d and d ∗
uniformly equivalent. Suppose that  > 0 is given. Then there exists γ > 0 such that

April 2010] QUASI-CAUCHY SEQUENCES 331

This content downloaded from 171.67.34.69 on Sat, 30 Mar 2013 21:48:11 PM


All use subject to JSTOR Terms and Conditions
d ∗ (x, y) < γ implies d(x, y) < . Similarly we can find δ > 0 such that d(x, y) < δ
implies d ∗ (x, y) < γ (and hence d(x, y) < ). Suppose that x and y are δ-connected
in (X, d). Pick x0 , x1 , . . . , xn in X with x0 = x, xn = y, and d(xi , xi+1 ) < δ for i < n.
This implies d ∗ (xi , xi+1 ) < γ , so x and y are γ -connected in (X, d ∗ ). But d ∗ is an
ultrametric, so repeated application of the ultrametric inequality yields d ∗ (x, y) < γ .
Thus d(x, y) < . We can thus infer that if d is uniformly equivalent to an ultrametric
then

∀ > 0 ∃ δ > 0 (x, y δ-connected ⇒ d(x, y) < ). (∗)

Now let

X = {(n, m/n) | m, n integers with n ≥ 1 and 0 ≤ m ≤ n} (see Figure 1).

6
1 t t t

t
t q q q
t

t t t -
1 2 3 q q q
Figure 1.

Consider X as a subset of the Cartesian plane and let d be the induced Euclidean
metric. X consists of infinitely many columns where each column is distance 1 from
its neighboring columns. Thus any quasi-Cauchy sequence is eventually confined
to a single column. But each column consists of only finitely many points. So any
quasi-Cauchy sequence in eventually constant. Thus every quasi-Cauchy sequence is
Cauchy, so (X, d) is nonincremental. On the other hand, (X, d) fails to satisfy (∗) with
 = 1/2: For any δ > 0 you can pick n with 1/n < δ. Then the points p1 = (n, 0)
and p2 = (n, 1) are clearly δ-connected but d( p1 , p2 ) = 1 > . Thus, (X, d) is not
uniformly equivalent to an ultrametric space.
To put these results in context: papers have appeared charactering ultrametric spaces
up to topological equivalence [3] and up to uniform equivalence [4]. Nonincremental
spaces are strictly in between. Whether this is a useful concept or a curiosity remains
to be seen.

3. EXERCISES. There are a number of interesting exercises involving quasi-Cauchy


sequences that can be given to undergraduate analysis or topology students:
1. Suppose that f is uniformly continuous on some interval I . Prove that the image
under f of any quasi-Cauchy sequence in I is quasi-Cauchy.
2. Suppose that xn  and yn  are quasi-Cauchy sequences of real numbers. Prove
or give a counterexample to the claim that the product sequence xn yn  is also
quasi-Cauchy.
3. Prove that a sequence xn  of real numbers is Cauchy if and only if every subse-
quence is quasi-Cauchy.

332
c THE MATHEMATICAL ASSOCIATION OF AMERICA [Monthly 117

This content downloaded from 171.67.34.69 on Sat, 30 Mar 2013 21:48:11 PM


All use subject to JSTOR Terms and Conditions
4. Prove that a quasi-Cauchy sequence of real numbers is Cauchy if and only if it
has exactly one cluster point.
5. Prove that the set of cluster points of a quasi-Cauchy sequence in R is closed
and connected.
6. Prove that any closed, connected set in R is the cluster point set of some quasi-
Cauchy sequence.
7. Call a set A in a metric space (X, d) pseudoconnected if it has the property that
any two points in A are -connected in (A, d  ) for all  > 0 (where the d  in
(A, d  ) is the restriction of d to A). Show that a set A is the cluster point set of
a quasi-Cauchy sequence in X if and only if A is closed and pseudoconnected.
Show by example that A need not be connected.

ACKNOWLEDGMENTS. We would like to thank the referees for their many helpful suggestions and com-
ments.

REFERENCES

1. S. Bosch, U. Güntzer, and R. Remmert, Non-Archimedean Analysis, Springer-Verlag, Berlin, 1984.


2. J. P. Coleman, Nonexpansive algebras, Algebra Universalis 55 (2006) 479–494. doi:10.1007/
s00012-006-1997-6
3. J. de Groot, Non-Archimedean metrics in topology, Proc. Amer. Math. Soc. 7 (1956) 948–956. doi:10.
2307/2033568
4. A. J. Lemin, Proximity on isosceles spaces, Russian Math. Surveys 39 (1984) 143–144.
5. S. Priess-Crampe and P. Ribenboim, Ultrametric spaces and logic programming, J. Logic Programming
42 (2000) 59–70. doi:10.1016/S0743-1066(99)00002-3

DAVID BURTON graduated from the U.S. Naval Academy in 1981. After serving in the submarine fleet for
several years, he returned to school and earned a Ph.D. in mathematics from Vanderbilt in 1997, specializing
in functional analysis. He is currently chair of the Mathematics and Computer Science Department at the
Franciscan University of Steubenville.
Department of Mathematics and Computer Science, Franciscan University of Steubenville,
Steubenville OH 43952
dburton@franciscan.edu

JOHN COLEMAN received his B.S. from Kent State University in 1985 and his Ph.D. from the University
of Colorado at Boulder in 1992, specializing in universal algebra. His main research interest is in the interplay
between algebra and topology in topological algebras. He currently teaches mathematics and occasionally
computer science at the Franciscan University of Steubenville.
Department of Mathematics and Computer Science, Franciscan University of Steubenville,
Steubenville OH 43952
jcoleman@franciscan.edu

April 2010] QUASI-CAUCHY SEQUENCES 333

This content downloaded from 171.67.34.69 on Sat, 30 Mar 2013 21:48:11 PM


All use subject to JSTOR Terms and Conditions

S-ar putea să vă placă și