This post contains some summary informal notes of key ideas from my reading of Mícheál Ó Searcóid’s Metric Spaces (Springer, 2007). These notes are here as a reference for me, my students, and any others who may be interested. They are by no means exhaustive, but rather cover topics that seemed interesting to me on first reading. By way of a brief book review, it’s worth noting that Ó Searcóid’s approach is excellent for learning a subject. He has a few useful tricks up his sleeve, in particular:
- Chapters will often start with a theorem proving equivalence of various statements (e.g. Theorem 8.1.1, Criteria for Continuity at a Point). Only then will he choose one of these statements as a definition, and he explains this choice carefully, often via reference to other mathematics.
- The usual definition-theorem-proof style is supplemented with ‘question’ – these are relatively informally-stated questions and their answers. They have been carefully chosen to highlight some questions the reader might be wondering about at that point in the text and to demonstrate key (and sometimes surprising) answers before the formal theorem statement.
- The writing is pleasant, even playful at times though never lacking formality. This is a neat trick to pull off.
- There are plenty of exercises, and solutions are provided.
These features combine to produce an excellent learning experience.
1. Some Basic Definitions
A metric on a set X is a function
such that:
- Positivity:
with equality iff 
- Symmetry:

- Triangle inequality:

The combination of such a metric and a the corresponding set is a metric space.
Given a metric space
, the point function at
is
.
A pointlike function
is one where 
For metric spaces
and
,
is a metric subspace of
iff
and
is a restriction of
.
For metric spaces
and
, an isometry
is a function such that
. The metric subspace
is an isometric copy of
.
Some standard constructions of metrics for product spaces:



A conserving metric
on a product space is one where
. Ó Searcóid calls these conserving metrics because they conserve an isometric copy of the individual spaces, recoverable by projection (I don’t think this is a commonly used term). This can be seen because fixing elements of all-but-one of the constituent spaces makes the upper and lower bound coincide, resulting in recovery of the original metric.
A norm on a linear space
over
or
is a real function such that for
and
scalar:
The metric defined by the norm is
.
2. Distances
The diameter of a set
of metric space
is
.
The distance of a point
from a set
is
.
An isolated point
where
is one for which
.
An accumulation point or limit point
of
is one for which
. Note that
doesn’t need to be in
. A good example is
,
,
.
The distance from subset
to subset
of a metric space is defined as
.
A nearest point
of
to
is one for which
. Note that nearest points don’t need to exist, because
is defined via the infimum. If a metric space is empty or admits a nearest point to each point in every metric superspace, it is said to have the nearest-point property.
3. Boundaries
A point
is a boundary point of
in
iff
. The collection of these points is the boundary
.
Metric spaces with no proper non-trivial subset with empty boundary are connected. An example of a disconnected metric space is
as a metric subspace of
, while
itself is certainly connected.
Closed sets are those that contain their boundary.
The closure of
in
is
. The interior is
. The exterior is
.
Interior, boundary, and exterior are mutually disjoint and their union is
.
4. Sub- and super-spaces
A subset
is dense in
iff
, or equivalently if for every
,
. The archetypal example is that
is dense in
.
A complete metric space
is one that is closed in every metric superspace of
. An example is
.
5. Balls
Let
denote an open ball and similarly
denote a closed ball. In the special case of normed linear spaces,
and similarly for closed balls, so the important object is this unit ball – all others have the same shape. A norm on a space
is actually defined by three properties such balls
must have:
- Convexity
- Balanced (i.e.
) - For each
, the set
,
- is nonempty
- must have real supremum


6. Convergence
The
th tail of a sequence
is the set
.
Suppose
is a metric space,
and
is a sequence in
. Sequence
converges to
in
, denoted
iff every open subset of
that contains
includes a tail of
. In this situation,
is unique and is called the limit of the sequence, denoted
.
It follows that for
a metric space,
and
a sequence in
, the sequence
converges to
in
iff the real sequence
converges to
in
.
For real sequences, we can define the:
- limit superior,
and - limit inferior,
.
It can be shown that
iff
.
Clearly sequences in superspaces converge to the same limit – the same is true in subspaces if the limit point is in the subspace itself. Sequences in finite product spaces equipped with product metrics converge in the product space iff their projections onto the individual spaces converge.
Every subsequence of a convergent sequence converges to the same limit as the parent sequence, but the picture for non-convergent parent sequences is more complicated, as we can still have convergent subsequences. There are various equivalent ways of characterising these limits of subsequences, e.g. centres of balls containing an infinite number of terms of the parent sequence.
A sequence
is Cauchy iff for every
, there is a ball of radius
that includes a tail of
. Every convergent sequence is Cauchy. The converse is not true, but only if the what should be the limit point is missing from the space — adding this point and extending the metric appropriately yields a convergent sequence. It can be shown that a space is complete (see above for definition) iff every Cauchy sequence is also a convergent sequence in that space.
7. Bounds
A subset
of a metric space
is a bounded subset iff
or
is included in some ball of
. A metric space
is bounded iff it is a bounded subset of itself. An alternative characterisation of a bounded subset
is that it has finite diameter.
The Hausdorff metric is defined on the set
of all non-empty closed bounded subsets of a set
equipped with metric
. It is given by
.
Given a set
and a metric space
,
is a bounded function iff
is a bounded subset of
. The set of bounded functions from
to
is denoted
. There is a standard metric on bounded functions,
where
is the metric on
.
Let
be a nonempty set and
be a nonempty metric space. Let
be a sequence of functions from
to
and
. Then:
converges pointwise to
iff
converges to
for all 
converges uniformly to
iff
is real for each
and the sequence
converges to zero in
.
It’s interesting to look at these two different notions of convergence because the second is stronger. Every uniformly-convergent sequence of functions converges pointwise, but the converse is not true. An example is the sequence
given by
. This converges pointwise but not uniformly to the zero function.
A stronger notion than boundedness is total boundedness. A subset
of a metric space
is totally bounded iff for each
, there is a finite collection of balls of
of radius
that covers
. An example of a bounded but not totally bounded subset is any infinite subset of a space with the discrete metric. Total boundedness carries over to subspaces and finite unions.
Conserving metrics play an important role in bounds, allowing bounds on product spaces to be equivalent to bounds on the projections to the individual spaces. This goes for both boundedness and total boundedness.
8. Continuity
Given metric spaces
and
, a point
and a function
, the function is said to be continuous at
iff for each open subset
with
, there exists and open subset
of
with
such that
.
Extending from points to the whole domain, the function is said to be continuous on
iff for each open subset
,
is open in
.
Continuity is not determined by the codomain, in the sense that a continuous function is continuous on any metric superspace of its range. It is preserved by function composition and by restriction.
Continuity plays well with product spaces, in the sense that if the product space is endowed with a product metric, a function mapping into the product space is continuous iff its compositions with the natural projections are all continuous.
For
and
metric spaces,
denotes the metric space of continuous bounded functions from
to
with the supremum metric
.
is closed in the space of bounded functions from
to
.
Nicely, we can talk about convergence using the language of continuity. In particular, let
be a metric space, and
. Endow
with the inverse metric
for
,
and
. Let
. Then
is continuous iff the sequence
converges in
to
. In particular, the function extending each convergent sequence with its limit is an isometry from the space of convergent sequences in
to the metric space of continuous bounded functions from
to
.
9. Uniform Continuity
Here we explore increasing strengths of continuity: Lipschitz continuity > uniform continuity > continuity. Ó Searcóid also adds strong contractions into this hierarchy, as the strongest class studied.
Uniform continuity requires the
in the epsilon-delta definition of continuity to extend across a whole set. Consider metric spaces
and
, a function
, and a metric subspace
. The function
is uniformly continuous on
iff for every
there exists a
s.t. for every
for which
, it holds that
.
If
is a metric space with the nearest-point property and
is continuous, then
is also uniformly continuous on every bounded subset of
. A good example might be a polynomial on
.
Uniformly continuous functions map compact metric spaces into compact metric spaces. They preserve total boundedness and Cauchy sequences. This isn’t necessarily true for continuous functions, e.g.
on
does not preserve the Cauchy property of the sequence
.
There is a remarkable relationship between the Cantor Set and uniform continuity. Consider a nonempty metric space
. Then
is totally bounded iff there exists a bijective uniformly continuous function from a subset of the Cantor Set to
. As Ó Searcóid notes, this means that totally bounded metric spaces are quite small, in the sense that none can have cardinality greater than that of the reals.
Consider metric spaces
and
and function
. The function is called Lipschitz with Lipschitz constant
iff
for all
.
Note here the difference to uniform continuity: Lipschitz continuity restricts uniform continuity by describing a relationship that must exist between the
s and
s – uniform leaves this open. A nice example from Ó Searcóid of a uniformly continuous non-Lipschitz function is
on
.
Lipschitz functions preserve boundedness, and the Lipschitz property is preserved by function composition.
There is a relationship between Lipschitz functions on the reals and their differentials. Let
be a non-degenerate intervals of
and
. Then
is Lipschitz on
iff
is bounded on
.
A function with Lipschitz constant less than one is called a strong contraction.
Unlike the case for continuity, not every product metric gives rise to uniformly continuous natural projections, but this does hold for conserving metrics.
10. Completeness
Let
be a metric space and
. The function
is called a virtual point iff:
We saw earlier that a metric space
is complete iff it is closed in every metric superspace of
. There are a number of equivalent characterisations, including that every Cauchy sequence in
converses in
.
Consider a metric space
. A subset of
is a complete subset of
iff
is a complete metric space.
If
is a complete metric space and
, then
is complete iff
is closed in
.
Conserving metrics ensure that finite products of complete metric spaces are complete.
A non-empty metric space
is complete iff
is complete, where
denotes the collection of all non-empty closed bounded subsets of
and
denotes the Hausdorff metric.
For
a non-empty set and
a metric space, the metric space
of bounded functions from
to
with the supremum metric is a complete metric space iff
is complete. An example is that the space of bounded sequences in
is complete due to completeness of
.
We can extend uniformly continuous functions from dense subsets to complete spaces to unique uniformly continuous functions from the whole: Consider metric spaces
and
with the latter being complete. Let
be a dense subset of
and
be a uniformly continuous function. Then there exists a uniformly continuous function
such that
. There are no other continuous extensions of
to
.
(Banach’s Fixed-Point Theorem). Let
be a non-empty complete metric space and
be a strong contraction on
with Lipschitz constant
. Then
has a unique fixed point in
and, for each
, the sequence
converges to the fixed point. Beautiful examples of this abound, of course. Ó Searcóid discusses IFS fractals – computer scientists will be familiar with applications in the semantics of programming languages.
A metric space
is called a completion of metric space
iff
is complete and
is isometric to a dense subspace of
.
We can complete any metric space. Let
be a metric space. Define
where
denotes the set of all point functions in
and
denotes the set of all virtual points in
. We can endow
with the metric
given by
. Then
is a completion of
.
Here the subspace
of
forms the subspace isometric to
.
11. Connectedness
A metric space
is a connected metric space iff
cannot be expressed as the union of two disjoint nonempty open subsets of itself. An example is
with its usual metric. As usual, Ó Searcóid gives a number of equivalent criteria:
- Every proper nonempty subset of
has nonempty boundary in 
- No proper nonempty subset of
is both open and closed in 
is not the union of two disjoint nonempty closed subsets of itself- Either
or the only continuous functions from
to the discrete space
are the two constant functions
Connectedness is not a property that is relative to any metric superspace. In particular, if
is a metric space,
is a metric subspace of
and
, then the subspace
of
is a connected metric space iff the subspace
of
is a connected metric space. Moreover, for a connected subspace
of
with
, the subspace
is connected. In particular,
itself is connected.
Every continuous image of a connected metric space is connected. In particular, for nonempty
,
is connected iff
is an interval. This is a generalisation of the Intermediate Value Theorem (to see this, consider the continuous functions
.
Finite products of connected subsets endowed with a product metric are connected. Unions of chained collections (i.e. sequences of subsets whose sequence neighbours are non-disjoint) of connected subsets are themselves connected.
A connected component
of a metric space
is a subset that is connected and which has no proper superset that is also connected – a kind of maximal connected subset. It turns out that the connected components of a metric space
are mutually disjoint, all closed in
, and
is the union of its connected components.
A path in metric space
is a continuous function
. (These functions turn out to be uniformly continuous.) This definition allows us to consider a stronger notion of connectedness: a metric space
is pathwise connected iff for each
there is a path in
with endpoints
and
. An example given by Ó Searcóid of a space that is connected but not pathwise connected is the closure in
of
. From one of the results above,
is connected because
is connected. But there is no path from, say,
(which nevertheless is in
) to any point in
.
Every continuous image of a pathwise connected metric space is itself pathwise connected.
For a linear space, an even stronger notion of connectedness is polygonal connectedness. For a linear space
with subset
and
, a polygonal connection from
to
in
is an
-tuple of points
s.t.
,
and for each
,
. We then say a space is polygonally connected iff there exists a polygonal connection between every two points in the space. Ó Searcóid gives the example of
as a pathwise connected but not polygonally connected subset of
.
Although in general these three notions of connectedness are distinct, they coincide for open connected subsets of normed linear spaces.
12. Compactness
Ó Searcóid gives a number of equivalent characterisations of compact non-empty metric spaces
, some of the ones I found most interesting and useful for the following material include:
- Every open cover for
has a finite subcover
is complete and totally bounded
is a continuous image of the Cantor set- Every real continuous function defined on
is bounded and attains its bounds
The example is given of closed bounded intervals of
as archetypal compact sets. An interesting observation is given that ‘most’ metric spaces cannot be extended to compact metric spaces, simply because there aren’t many compact metric spaces — as noted above in the section on bounds, there are certainly no more than
, given they’re all images of the Cantor set.
If
is a compact metric space and
then
is compact iff
is closed in
. This follows because
inherits total boundedness from
, and completeness follows also if
is closed.
The Inverse Function Theorem states that for
and
metric spaces with
compact, and for
injective and continuous,
is uniformly continuous.
Compactness plays well with intersections, finite unions, and finite products endowed with a product metric. The latter is interesting, given that we noted above that for non conserving product metrics, total boundedness doesn’t necessarily carry forward.
Things get trickier when dealing with infinite-dimension spaces. The following statement of the Arzelà-Ascoli Theorem is given, which allows us to characterise the compactness of a closed, bounded subset of
for compact metric spaces
and
:
For each
, define
by
for each
. Let
. Then:
and
is compact iff
from
to
is continuous
13. Equivalence
Consider a set
and the various metrics we can equip it with. We can define a partial order
on these metrics in the following way.
is topologically stronger than
,
iff every open subset of
is open in
. We then get an induced notion of topological equivalence of two metrics, when
and
.
As well as obviously admitting the same open subsets, topologically equivalent metrics admit the same closed subsets, dense subsets, compact subsets, connected subsets, convergent sequences, limits, and continuous functions to/from that set.
It turns out that two metrics are topologically equivalent iff the identity functions from
to
and vice versa are both continuous. Following the discussion above relating to continuity, this hints at potentially stronger notions of comparability – and hence of equivalence – of metrics, which indeed exist. In particular
is uniformly stronger than
iff the identify function from
to
is uniformly continuous. Also,
is Lipschitz stronger than
iff the identity function from
to
is Lipschitz.
The stronger notion of a uniformly equivalent metric is important because these metrics additionally admit the same Cauchy sequences, totally bounded subsets and complete subsets.
Lipschitz equivalence is even stronger, additionally providing the same bounded subsets and subsets with the nearest-point property.
The various notions of equivalence discussed here collapse to a single one when dealing with norms. For a linear space
, two norms on
are topologically equivalent iff they are Lipschitz equivalent, so we can just refer to norms as being equivalent. All norms on finite-dimensional linear spaces are equivalent.
Finally, some notes on the more general idea of equivalent metric spaces (rather than equivalent metrics.) Again, these are provided in three flavours:
- topologically equivalent metric spaces
and
are those for which there exists a continuous bijection with continuous inverse (a homeomorphism) from
to
. - for uniformly equivalent metric spaces, we strengthen the requirement to uniform continuity
- for Lipschitz equivalent metric spaces, we strengthen the requirement to Lipschitz continuity
- strongest of all, isometries are discussed above
Note that given the definitions above, the metric space
is equivalent to the metric space
if
and
are equivalent, but the converse is not necessarily true. For equivalent metric spaces, we require existence of a function — for equivalent metrics this is required to be the identity.