This work is licensed under a Creative Commons Attribution-NonCommercial-ShareAlike 4.0 International License.2 Basics of Linear Spaces
A person is solely the concentration of an infinite set of
interrelations with another and others, and to separate a person
from these relations means to take away any real meaning of the
life.
Vl. Soloviev
A space around us could be described as a three dimensional Euclidean
space. To single out a point of that space we need a fixed
frame of references and three real numbers, which are
coordinates of the point. Similarly to describe a pair of
points from our space we could use six coordinates; for three
points—nine, end so on. This makes it reasonable to consider
Euclidean (linear) spaces of an arbitrary finite dimension, which are
studied in the courses of linear algebra.
The basic properties of Euclidean spaces are determined by its
linear and metric structures. The linear
space (or vector space) structure allows
to add and subtract vectors
associated to points as well as
to multiply vectors by
real or complex numbers (scalars).
The metric
space structure assign a
distance—non-negative real
number—to a pair of points or, equivalently, defines a
length of a vector defined by that
pair. A metric (or, more generally a topology) is essential for
definition of the core analytical notions like limit or continuity.
The importance of linear and metric (topological) structure in
analysis sometime encoded in the formula:
Analysis = Algebra +
Geometry .
(8) |
On the other hand we could observe that many sets admit a sort of
linear and metric structures which are linked each other. Just
few among many other examples are:
- The set of convergent sequences;
- The set of continuous functions on [0,1].
It is a very mathematical way of
thinking to declare
such sets to be spaces and call their elements
points.
But shall we lose all information on a particular element (e.g. a
sequence {1/n}) if we represent it by a shapeless and size-less
“point” without any inner configuration? Surprisingly not: all
properties of an element could be now retrieved not from its inner
configuration but from interactions with other elements through linear
and metric structures. Such a “sociological” approach to
all kind of mathematical objects was codified in the abstract
category
theory.
Another surprise is that starting from our three dimensional Euclidean
space and walking far away by a road of abstraction to infinite
dimensional Hilbert spaces we are arriving just to yet another picture
of the surrounding space—that time on the language of
quantum mechanics.
The distance from Manchester to Liverpool is 35 miles—just
about the mileage in the opposite direction!
A tourist guide
to England
2.1 Banach spaces (basic definitions only)
The following definition generalises the notion of distance
known from the everyday life.
Definition 1
A metric
(or distance function
) d on a set M
is a function d:
M×
M →ℝ
+ from the
set of pairs to non-negative real numbers such that:
-
d(x,y)≥0 for all x, y ∈ M, d(x,y)=0
implies x=y .
- d(x,y)=d(y,x) for all x and y in M.
-
d(x,y)+d(y,z)≥ d(x,z) for all x, y, and z
in M (triangle inequality).
Exercise 2
Let M be the set of UK’s cities are the following function are
metrics on M:
-
d(A,B) is the price of 2nd class railway ticket from A
to B.
- d(A,B) is the off-peak driving time from A to B.
The following notion is a useful specialisation of metric adopted to the
linear structure.
Definition 3
Let V be a (real or complex) vector space. A norm
on
V is a real-valued function, written ||
x||
, such that
-
||x||≥ 0 for all x∈ V, and ||x||=0
implies x=0.
- ||λ x|| = | λ | ||x|| for all scalar
λ and vector x.
-
||x+y||≤ ||x||+||y|| (triangle
inequality).
A vector space with a norm is called a normed
space.
The connection between norm and metric is as follows:
Proposition 4
If ||·||
is a norm on V, then it gives a metric on
V by d(
x,
y)=||
x−
y||
.
(a)
(b)
Figure 1: Triangle inequality in metric (a)
and normed (b) spaces. |
Proof.
This is a simple exercise to derive
items
1–
3 of
Definition
1 from corresponding items of
Definition
3. For example, see the
Figure
1 to derive
the triangle inequality.
□
An important notions known from real analysis
are limit and
convergence. Particularly we usually wish to have enough limiting
points for all “reasonable” sequences.
Definition 5
A sequence {
xk}
in a metric space (
M,
d)
is a
Cauchy sequence, if for every
є>0
, there exists an integer n such that k,
l>
n
implies that d(
xk,
xl)<є
.
Definition 6
(
M,
d)
is a complete metric
space if every Cauchy sequence in M
converges to a limit in M.
For example, the set of integers ℤ and reals
ℝ with the natural distance functions are complete
spaces, but the set of rationals ℚ is not. The complete
normed spaces
deserve a special name.
Definition 7
A Banach space is a complete normed space.
Exercise* 8
A convenient way to define a norm in a Banach space is as
follows. The unit ball U in a normed space B
is the set of x such that ||
x||≤ 1
. Prove that:
-
U is a convex set, i.e. x,
y∈ U and λ∈ [0,1] the point λ x
+(1−λ)y is also in U.
- ||x||=inf{ λ∈ℝ+ ∣
λ−1x ∈ U}.
- U is closed if and only if the space is Banach.
(i)
(ii)
(iii)
Figure 2: Different unit balls defining norms in ℝ2 from
Example 9. |
Example 9
Here is some examples of normed spaces.
-
l2n is either ℝn or
ℂn with norm defined by
| ⎪⎪
⎪⎪ | (x1,…,xn) | ⎪⎪
⎪⎪ | 2 = | √ | |
⎪
⎪ | x1 | ⎪
⎪ | 2+
| ⎪
⎪ | x2 | ⎪
⎪ | 2+ ⋯+ | ⎪
⎪ | xn | ⎪
⎪ | 2 |
|
| .
(9) |
-
l1n is either ℝn or
ℂn with norm defined by
| ⎪⎪
⎪⎪ | (x1,…,xn) | ⎪⎪
⎪⎪ | 1 = | ⎪
⎪ | x1 | ⎪
⎪ | +
| ⎪
⎪ | x2 | ⎪
⎪ | + ⋯+ | ⎪
⎪ | xn | ⎪
⎪ |
| .
(10) |
-
l∞n is either ℝn or
ℂn with norm defined by
| ⎪⎪
⎪⎪ | (x1,…,xn) | ⎪⎪
⎪⎪ | ∞ = max( | ⎪
⎪ | x1 | ⎪
⎪ | ,
| ⎪
⎪ | x2 | ⎪
⎪ | , ⋯, | ⎪
⎪ | xn | ⎪
⎪ |
| ).
(11) |
-
Let X be a topological space, then Cb(X) is
the space of continuous bounded functions f:
X→ℂ with norm ||f||∞=supX
| f(x) |.
-
Let X be any set, then l∞(X) is the
space of all bounded (not necessarily continuous)
functions f: X→ℂ with norm
||f||∞=supX | f(x) |.
All these normed spaces are also complete and thus are Banach
spaces. Some more examples of both complete and incomplete spaces
shall appear later.
—We need an extra space to accommodate this product!
A
manager to a shop assistant
2.2 Hilbert spaces
Although metric and norm capture important geometric information about
linear spaces they are not sensitive enough to represent such
geometric characterisation as angles (particularly
orthogonality). To
this end we need a further refinements.
From courses of linear algebra known that the scalar product
⟨ x,y
⟩= x1 y1 + ⋯ + xn yn is important in a space
ℝn and defines a norm ||x||2=⟨ x,x
⟩. Here
is a suitable generalisation:
Definition 10
A scalar product (or
inner product) on a real or complex
vector space V is a mapping V×
V →
ℂ
, written ⟨
x,
y
⟩
, that satisfies:
-
⟨ x,x
⟩ ≥ 0 and ⟨ x,x
⟩ =0 implies x=0.
- ⟨ x,y
⟩ = ⟨ y,x
⟩ in complex spaces and
⟨ x,y
⟩ = ⟨ y,x
⟩ in real ones for all x, y∈ V.
- ⟨ λ x,y
⟩=λ ⟨ x,y
⟩, for all x,
y∈ V and scalar λ. (What is
⟨ x,λ y
⟩?).
- ⟨ x+y,z
⟩=⟨ x,z
⟩ + ⟨ y,z
⟩, for all x,
y, and z∈ V. (What is
⟨ x, y+z
⟩?).
Last two properties of the scalar product is oftenly encoded in the
phrase: “it is linear in the first variable if we fix the second
and anti-linear in the second if we fix the first”.
Definition 11
An inner product space V is a
real or complex vector space with a scalar product on it.
Example 12
Here is some examples of inner product spaces which demonstrate that
expression ||
x||=√
⟨ x,x
⟩ defines a norm.
-
The inner product for ℝn was defined in the beginning
of this section. The inner product for ℂn is given
by ⟨ x,y
⟩=∑1n xj ȳj. The norm
||x||=√∑1n | xj |2 makes it
l2n from Example 1.
-
The extension for infinite vectors: let l2
be
l2={ sequences {xj}1∞ ∣
| | ⎪
⎪ | xj | ⎪
⎪ | 2 < ∞}.
(12) |
Let us equip this set with operations of term-wise addition and
multiplication by scalars, then l2 is closed under
them. Indeed it follows from the
triangle inequality
and properties of absolutely
convergent series. From the standard
Cauchy–Bunyakovskii–Schwarz
inequality follows that the series ∑1∞xjȳj
absolutely
converges and its sum defined to be ⟨ x,y
⟩.
- Let Cb[a,b] be a space of continuous functions
on the interval [a,b]∈ℝ. As we learn from
Example 4 a normed space it is a normed space
with the norm ||f||∞=sup[a,b]| f(x) |. We
could also define an inner product:
⟨ f,g
⟩= | | f(x)ḡ(x) d x
and
| ⎪⎪
⎪⎪ | f | ⎪⎪
⎪⎪ | 2= | ⎛
⎜
⎜
⎝ | | | ⎪
⎪ | f(x) | ⎪
⎪ | 2 d x | ⎞
⎟
⎟
⎠ | | .
(13) |
Now we state, probably, the most important inequality in analysis.
Theorem 13 (Cauchy–Schwarz–Bunyakovskii inequality)
For vectors x and y in an inner product space V let us
define ||
x||=√
⟨ x,x
⟩ and
||
y||=√
⟨ y,y
⟩ then we have
| ⎪
⎪ | ⟨ x,y
⟩ | ⎪
⎪ | ≤ | ⎪⎪
⎪⎪ | x | ⎪⎪
⎪⎪ | ⎪⎪
⎪⎪ | y | ⎪⎪
⎪⎪ | ,
(14) |
with equality if and only if x and y are scalar multiple
each other.
Proof.
For simplicity we start from a real vector space.
Let we have two vectors
u and
v and want to define an inner product on
the two-dimensional vector space spanned by them. That is we need to
know a value of ⟨
au+
bv,
cu+
dv
⟩ for all possible scalars
a,
b,
c,
d.
By the linearity ⟨ au+bv, cu+dv
⟩ = ac⟨ u,u
⟩ + (bc+ad)⟨ u,v
⟩ + db⟨ v,v
⟩,
thus everything is defined as soon as we know three inner products
⟨ u,u
⟩, ⟨ u,v
⟩ and ⟨ v,v
⟩. First of all we need to demand ⟨ u,u
⟩ ≥ 0 and
⟨ v,v
⟩ ≥ 0.
Furthermore, they shall be such that ⟨ au+bv, au+bv
⟩ ≥ 0 for all
scalar a and b. If a=0, that is reduced to the previous case ⟨ v,v
⟩ ≥ 0.
If a is non-zero we note ⟨ au+bv, au+bv
⟩ = a2 ⟨ u+(b/a)v, u+(b/a)v
⟩
and letting λ = b/a we reduce our consideration to the quadratic expression
⟨ u+λ v, u+λ v
⟩ = λ 2⟨ v,v
⟩+2λ ⟨ u,v
⟩+⟨ u,u
⟩.
|
The graph of this function of λ is an upward parabolabecause ⟨ v,v
⟩ ≥ 0. Thus, it will be non-negative for all λ if its
lowest value is non-negative. From the theory of quadratic
expressions, the latter is achieved at λ =−⟨ u,v
⟩/⟨ v,v
⟩
and is equal to
| | ⟨ v,v
⟩ − 2 | | ⟨ u,v
⟩+⟨ u,u
⟩=− | | +⟨ u,u
⟩
|
If −⟨ u,v
⟩2/⟨ v,v
⟩+⟨ u,u
⟩ ≥ 0 then ⟨ v,v
⟩⟨ u,u
⟩ ≥ ⟨ u,v
⟩2.
Therefore, the Cauchy-Schwarz inequality is necessary and sufficient
condition for the non-negativity of the inner product defined by the
three values ⟨ u,u
⟩, ⟨ u,v
⟩ and ⟨ v,v
⟩.
After the previous discussion it is easy to get the result for complex vector space as well. For any x, y∈ V and any t∈ℝ we have:
0< ⟨ x+t y,x+t y
⟩= ⟨ x,x
⟩+2t ℜ ⟨ y,x
⟩+t2⟨ y,y
⟩),
|
Thus, the discriminant of this quadratic expression in t is
non-positive: (ℜ ⟨ y,x
⟩)2−||x||2||y||2≤ 0,
that is | ℜ ⟨ x,y
⟩ |≤||x||||y||. Replacing y
by eiαy for an arbitrary α∈[−π,π] we get | ℜ
(eiα⟨ x,y
⟩) | ≤||x||||y||, this
implies the desired inequality.
□
Corollary 14
Any inner product space is a normed space with norm
||
x||=√
⟨ x,x
⟩ (hence also a metric space,
Prop. 4).
Proof.
Just to check
items
1–
3 from
Definition
3.
□
Again complete inner product spaces deserve a special name
Definition 15
A complete inner product space is Hilbert
space.
The relations between spaces introduced so far are as follows:
Hilbert spaces | ⇒ | Banach spaces | ⇒ | Complete
metric spaces |
⇓ | | ⇓ | | ⇓ |
inner product spaces | ⇒ | normed spaces | ⇒ | metric spaces.
|
How can we tell if a given norm comes from an inner product?
Figure 3: To the parallelogram identity. |
Theorem 16 (Parallelogram identity)
In an inner product space H we have for all x and y∈
H (see
Figure 3):
| ⎪⎪
⎪⎪ | x+y | ⎪⎪
⎪⎪ | 2+ | ⎪⎪
⎪⎪ | x−y | ⎪⎪
⎪⎪ | 2=2 | ⎪⎪
⎪⎪ | x | ⎪⎪
⎪⎪ | 2+2 | ⎪⎪
⎪⎪ | y | ⎪⎪
⎪⎪ | 2.
(15) |
Proof.
Just by linearity of inner product:
⟨ x+y,x+y
⟩+⟨ x−y,x−y
⟩=2⟨ x,x
⟩+2⟨ y,y
⟩,
|
because the cross terms cancel out.
□
Exercise 17
Show that (15) is also a
sufficient condition for a norm to
arise from an inner product. Namely, for a norm on a complex Banach
space satisfying
to (15) the formula
| ⟨ x,y
⟩ | = | | ⎛
⎝ | ⎪⎪
⎪⎪ | x+y | ⎪⎪
⎪⎪ | 2− | ⎪⎪
⎪⎪ | x−y | ⎪⎪
⎪⎪ | 2+i | ⎪⎪
⎪⎪ | x+iy | ⎪⎪
⎪⎪ | 2
−i | ⎪⎪
⎪⎪ | x−iy | ⎪⎪
⎪⎪ | 2 | ⎞
⎠ | |
| (16) |
| = | | |
|
defines an inner product. What is a suitable formula for a real
Banach space?
Divide and rule!
Old but still much used recipe
2.3 Subspaces
To study Hilbert spaces we may use the traditional mathematical
technique of analysis and synthesis: we split the
initial Hilbert spaces into smaller and probably simpler subsets,
investigate them separately, and then reconstruct the entire picture
from these parts.
As known from the linear algebra, a linear subspace is a subset
of a linear space is its subset, which inherits the linear structure,
i.e. possibility to add vectors and multiply them by scalars. In this
course we need also that subspaces inherit topological structure
(coming either from a norm or an inner product) as well.
Definition 18
By a subspace of a normed space (or inner product space) we
mean a linear subspace with the same norm (inner product
respectively). We write X⊂ Y or X ⊆ Y.
Example 19
-
Cb(X) ⊂ l∞(X) where X
is a metric space.
- Any linear subspace of ℝn or ℂn
with any norm given in
Example 1–3.
-
Let c00 be the space of finite sequences,
i.e. all sequences (xn)
such that exist N with xn=0 for n>N. This is a
subspace of l2 since ∑1∞| xj |2
is a finite sum, so finite.
We also wish that the both inhered structures (linear and
topological) should be in agreement, i.e. the subspace should be
complete. Such inheritance is linked to the property be closed.
A subspace need not be closed—for example the sequence
x=(1, 1/2, 1/3, 1/4, …)∈ l2
because
∑1/k2
< ∞
|
and xn=(1, 1/2,…, 1/n, 0, 0,…)∈
c00 converges to x thus x∈
c00 ⊂ l2.
Proposition 20
-
Any closed subspace of a Banach/Hilbert space is complete,
hence also a Banach/Hilbert space.
- Any complete subspace is closed.
- The closure of subspace is again a subspace.
Proof.
- This is true in any metric space X: any Cauchy sequence from
Y has a limit x ∈ X belonging to Ȳ, but if Y
is closed then x ∈ Y.
- Let Y is complete and x∈ Ȳ, then there is
sequence xn→ x in Y and it is a Cauchy sequence.
Then completeness of Y implies x∈ Y.
- If x, y∈ Ȳ then there are xn and yn
in Y such that xn→ x and yn→ y.
From the triangle
inequality:
| ⎪⎪
⎪⎪ | (xn+yn)−(x+y) | ⎪⎪
⎪⎪ | ≤ | ⎪⎪
⎪⎪ | xn−x | ⎪⎪
⎪⎪ | + | ⎪⎪
⎪⎪ | yn−y | ⎪⎪
⎪⎪ | →
0,
|
so xn+yn→ x+y and x+y∈ Ȳ.
Similarly x∈Ȳ implies λ x ∈Ȳ for any
λ.
□
Hence c00 is an incomplete inner product space,
with inner product ⟨ x,y
⟩=∑1∞xk ȳk (this
is a finite sum!) as it is not closed in l2.
(a)
(b)
Figure 4: Jump function on (b) as a L2 limit of
continuous functions from (a). |
Similarly C[0,1] with inner product norm
||f||=(∫01 | f(t) |2 dt)1/2 is
incomplete—take the large space X of functions continuous on
[0,1] except for a possible jump at 1/2 (i.e. left and
right limits exists but may be unequal and
f(1/2)=limt→1/2+ f(t).
Then the sequence of functions defined on
Figure 4(a) has the limit shown on
Figure 4(b) since:
| ⎪⎪
⎪⎪ | f−fn | ⎪⎪
⎪⎪ | = | |
| ⎪
⎪ | f−fn | ⎪
⎪ | 2 dt < | | → 0.
|
Obviously
f∈C[0,1]∖C[0,1].
Exercise 21
Show alternatively that the sequence of function fn from
Figure 4(a) is a Cauchy sequence in
C[0,1]
but has no continuous limit.
Similarly the space C[a,b] is
incomplete for any a<b if equipped by the inner product
and the corresponding norm:
| ⟨ f,g
⟩ | = | | (17) |
| = | ⎛
⎜
⎜
⎝ | | | ⎪
⎪ | f(t) | ⎪
⎪ | 2 d t | ⎞
⎟
⎟
⎠ | | .
|
| (18) |
|
Definition 22
Define a Hilbert space L2[
a,
b]
to be the
smallest complete inner product space containing space
C[
a,
b]
with the restriction of inner product given
by (17).
It is practical to realise L2[a,b] as a certain space
of “functions” with the inner product defined via an integral. There are
several ways to do that and we mention just two:
- Elements of L2[a,b] are equivalent classes of
Cauchy sequences f(n) of functions from
C[a,b].
- Let integration be extended from the
Riemann definition
to the wider Lebesgue
integration (see
Section 13). Let L be a set of square
integrable in Lebesgue sense
functions on [a,b] with a finite norm (18). Then
L2[a,b] is a quotient space of L with respect to
the equivalence relation f∼ g ⇔ ||f−g||2=0
.
Example 23
Let the Cantor function on [0,1]
be defined as follows:
This function is not
integrable in the Riemann sense but
does
have the Lebesgue integral. The later however is equal
to 0
and as an L2-function the Cantor function
equivalent to the function identically equal to 0
.
- The third possibility is to map L2(ℝ)
onto a space of “true” functions but with an additional
structure. For example, in quantum mechanics it is useful to
work with the Segal–Bargmann
space of
analytic functions on
ℂ with the inner product [, , ]:
⟨ f1,f2
⟩= | ∫ | | f1(z) f2(z)e | | d z.
|
Theorem 24
The sequence space l2 is complete, hence a Hilbert space.
Proof.
Take a Cauchy sequence
x(n)∈
l2, where
x(n)=(
x1(n),
x2(n),
x3(n), … ). Our proof
will have three steps: identify the limit
x; show it is in
l2; show
x(n)→
x.
- If x(n) is a Cauchy sequence in l2 then
xk(n) is also a Cauchy sequence of numbers for any fixed
k:
| ⎪
⎪ | xk(n)−xk(m) | ⎪
⎪ | ≤ | ⎛
⎜
⎜
⎝ | | ⎪
⎪ | xk(n)−xk(m) | ⎪
⎪ | 2 | ⎞
⎟
⎟
⎠ | | =
| ⎪⎪
⎪⎪ | x(n)−x(m) | ⎪⎪
⎪⎪ | → 0.
|
Let xk be the limit of xk(n).
- For a given є>0 find n0 such that
||x(n)−x(m)||<є for all n,m>n0. For any
K and m:
| | | ⎪
⎪ | xk(n)−xk(m) | ⎪
⎪ | 2 ≤ | ⎪⎪
⎪⎪ | x(n)−x(m) | ⎪⎪
⎪⎪ | 2<є2.
|
Let m→ ∞ then ∑k=1K
| xk(n)−xk |2 ≤ є2.
Let K→ ∞ then ∑k=1∞| xk(n)−xk |2 ≤ є2. Thus
x(n)−x∈l2 and because l2 is a
linear space then x = x(n)−(x(n)−x) is also in
l2.
- We saw above that for any є >0 there is n0
such that ||x(n)−x||<є for all n>n0. Thus
x(n)→ x.
Consequently
l2 is complete.
□
All good things are covered by a thick layer
of chocolate (well, if something is not yet–it certainly will)
2.4 Linear spans
As was explained into introduction 2, we
describe “internal” properties of a vector through its relations to
other vectors. For a detailed description we need sufficiently many
external reference points.
Let A be a subset (finite or infinite) of a normed space V. We
may wish to upgrade it to a linear subspace in order to make it
subject to our theory.
Definition 25
The linear span of A, write Lin(
A)
, is
the intersection of all linear subspaces of V containing A,
i.e. the smallest subspace containing A, equivalently the set of all finite
linear combination of elements of A.
The closed linear span of A write
CLin(
A)
is
the intersection of all closed
linear subspaces of V containing A,
i.e. the smallest closed
subspace containing A.
Exercise* 26
-
Show that if A is a subset of finite dimension space then Lin(A)=CLin(A).
- Show that for an infinite A spaces Lin(A) and
CLin(A)could be different. (Hint: use
Example 3.)
Proposition 27
Lin(A)=CLin(A).
Proof.
Clearly Lin(A) is a closed subspace containing A thus it
should contain CLin(A). Also Lin(A)⊂ CLin(A) thus
Lin(A)⊂
CLin(A)=CLin(A). Therefore
Lin(A)= CLin(A).
□
Consequently CLin(A) is the set of all limiting points of finite
linear combination of elements of A.
The following simple result will be used later many times without
comments.
Lemma 30 (about Inner Product Limit)
Suppose H is an inner product space and sequences xn and
yn have limits x and y correspondingly. Then
⟨
xn,
yn
⟩→⟨
x,
y
⟩
or equivalently:
Proof.
Obviously by the
Cauchy–Schwarz inequality:
| = | | ⎪
⎪ | ⟨ xn−x,yn
⟩+⟨ x,yn−y
⟩ | ⎪
⎪ |
|
| ≤ | ⎪
⎪ | ⟨ xn−x,yn
⟩ | ⎪
⎪ | + | ⎪
⎪ | ⟨ x,yn−y
⟩ | ⎪
⎪ |
|
| ≤ | ⎪⎪
⎪⎪ | xn−x | ⎪⎪
⎪⎪ | ⎪⎪
⎪⎪ | yn | ⎪⎪
⎪⎪ | + | ⎪⎪
⎪⎪ | x | ⎪⎪
⎪⎪ | ⎪⎪
⎪⎪ | yn−y | ⎪⎪
⎪⎪ |
→ 0,
|
|
|
since ||
xn−
x||→ 0, ||
yn−
y||→ 0,
and ||
yn|| is bounded.
□
Last modified: November 6, 2024.