A Crash Course in P-Adic Analysis: W. H. Schikhof
A Crash Course in P-Adic Analysis: W. H. Schikhof
A Crash Course in P-Adic Analysis: W. H. Schikhof
W. H. Schikhof
It is a pleasure to make many acknowledgements. First to Marı́a
Soledad Alcaı́no A. and Marı́a Eugenia Heckmann G. for typing
the manuscript. My thanks are also due to Carla Barrios R. and
Tonino Costa A., graduate students of the Pontificia Universidad
Católica de Chile, who made a thorough revision of the typewrit-
ten text, and materially assisted in its preparation.
Contents
1 BASICS 4
1.1 Ultrametric Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.1.1 Metrics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.1.2 Ultrametrics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.1.3 Compact Ultrametric Spaces . . . . . . . . . . . . . . . . . . . . . . . . 6
1.1.4 Spherical Completeness . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.2 Non-Archimedean Valued Fields . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.2.1 Valued Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.2.2 Non-Archimedean Valued Fields . . . . . . . . . . . . . . . . . . . . . . 8
1.2.3 Sequences and series in K . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.2.4 Power Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.2.5 Continuous Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.2.6 Differentiability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.3 Normed and Banach spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
1.3.1 Normed Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
1.3.2 Operators with an empty spectrum . . . . . . . . . . . . . . . . . . . . . 21
1.3.3 Commutation Relations . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
1.3.4 Addendum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2 ORTHOGONALITY 23
2.1 Definition of orthogonality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.2 Orthogonal bases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3 COUNTABILITY 32
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
4 DUALITY 36
5 COMPACTOIDS 42
5.1 List of properties of compactoids . . . . . . . . . . . . . . . . . . . . . . . . . . 43
In many branches of mathematics and its applications the fields R and C are playing a
fundamental role. For quite some time one has been discussing the consequences of replacing
in those theories R or C by a non-archimedean valued field. This story started in algebra
and number theory, where the p-adic fields, discovered by Hensel in 1909, could be used
successfully. Around 1930 analysis entered the picture through the work of Schnirlemann who
developed the basic theory of p-adic power series and analytic functions. From 1940 on the
Dutch mathematician A.F. Monna established the basics on p-adic Functional Analysis. At
this moment there are non-archimedean activities in all feasible disciplines such as Elementary
Calculus, Theory of C n - and C ∞ -functions, p-adic Lie Groups, Analytic Functions in one or
several variables, Algebraic and Analytic Number Theory, Algebraic Geometry, Functional
Analysis, ...
Of course, people are asking for applications.
2. In 1987 Igor Volovich raised the question as to whether at Planck distances (10−34 cm)
space should be disordered or disconnected. He suggested the use of p-adic numbers
to build more adequate quantum mechanical models. Right now serious mathematician
are working in this area.
Ultrametrics also appear speculatively in Psychology, Social Sciences, Economy (stock
market). Attempts are made to build a p-adic Probability Theory. A conference on
these applications was held, for the first time in history, in Moscow, October 2003.
Parts in the text, indicated ‘Background’are not needed for the course, but might be interesting
to hear about.
November, 2003.
BASICS
1.1.1 Metrics
A metric on a set X is a map d : X × X −→ [0, ∞) (d = distance) such that for all x, y, z ∈ X
X = (X, d) is called a metric space. (If we relax (i) to just d(x, x) = 0 then we have a
semi-metric d)
For a ∈ X, r > 0 we set
B(a, r) := {x ∈ X : d(x, a) ≤ r},
called the ‘closed’ ball about a with radius r and
diam Y := sup{d(x, y) : x, y ∈ Y }
1.1.2 Ultrametrics
A metric d on a set X is called ultrametric (and (X, d) is called an ultrametric space) if it
satisfies the so-called strong triangle inequality
for all x, y, z ∈ X. (Clearly this implies the ‘ordinary’ triangle inequality). In the spirit of
above one defines semi-ultrametrics. We have the fundamental
ISOSCELES TRIANGLE PRINCIPLE:
Example: Subsets of a non-archimedean valued field (see next Section) with the ultrametric
(x, y) 7→ |x − y|.
[BACKGROUND: In fact we have this way all examples, since each ultrametric space can
isometrically be embedded into a non-archimedean valued field (Indag. Math. 46 (1984),
51-53)]
Exercise 1.C. Let (X, d) be a metric space. Then, among all semi-ultrametric that are ≤ d
there is a largest one. It is called the subdominant semi-ultrametric for d.
(iii) Each point of a ball is a center. Each ball has an empty boundary.
(v) Two balls are either disjoint, or one is contained in the other.
(vi) If two balls B1 , B2 are disjoint, then dist (B1 , B2 ) = d(x, y) for each x ∈ B1 , y ∈ B2 .
(vii) Let ε > 0. The relation d(x, y) < ε (x, y ∈ X) is an equivalence relation and induces
a partition of X into ‘open’ balls of radius ε. A similar story holds for d(x, y) ≤ ε and
‘closed’ balls.
The topology induced by an ultrametric is zerodimensional, i.e. there is a base of the topology
consisting of clopen sets. Hence, an ultrametric space is totally disconnected.
If (X, d) is an ultrametric space and Y ⊂ X is dense then
Thus, completion of an ultrametric space does not create new values of the ultrametric.
Exercise 1.E. Let (X, d) be a compact ultrametric space. Show that {d(x, y) : x, y ∈ X} is
countable and has only 0 as possible accumulation point.
Let (X, d) be an infinite compact ultrametric space (if X is finite the process below breaks off).
By the above exercise the set of non-zero values of d is a sequence r1 > r2 > . . . tending to 0.
We shall make a picture of the collection of all balls in X as follows. We have X = B(a, r1 )
for any a ∈ X. The relation d(x, y) ≤ r2 (or d(x, y) < r1 if you want) decomposes X into
finitely many ‘closed’ balls of radius r2 :
Each of the closed balls of level 2 decomposes into finitely many ‘closed’ balls of radius r2 :
r.
DA . . . .
DA .
DA
r. r DDr AAr
.. E AB@
.. E
. E
BA@
BA @
r r r r Er r r
E BBrAAr @
@r
min{d(y, x) : x ∈ X}
exists.
Obvious examples are Q, R, C with the ordinary absolute value function. The mapping
(λ, µ) 7→ |λ − µ| is a metric on K making it into a topological field (i.e. the basic arithmetic
operations are continuous).
The metric completion of K is in a natural way again a valued field.
Two valuations on a field K are called equivalent if they induce the same topology.
[BACKGROUND: e.g. G. Bachman, Introduction to p-adic numbers and valuation theory,
Academic Press, New York, 1964: Two valuations | . |1 , | . |2 on K are equivalent if and only
if there is a positive constant c such that | . |1 = | . |2 c .]
The following theorem essentially separates R, C from all other valued fields. It shows the
‘alternative character’ of our non-archimedean analysis.
[BACKGROUND: Bachman, see before, page 127:
Theorem Let (K, | . |) be a valued field. Then, there are only two possibilities. Either
(ii) the valuation | . | is non-archimedean (n.a) i.e. it satisfies the strong triangle inequality
So, by excluding C and its (valued) subfields we obtain in return the strong triangle inequality.
The completion of (Q, |.|p ) is called (Qp , |.|p ), the field of the p-adic numbers.
Its value group is {pn : n ∈ Z}, its residue field is the field Fp of p elements.
(Of course, in the above, p is a prime number)
{pr : r ∈ Q},
so the valuation is dense, and the residue class field is the algebraic closure of Fp , hence
infinite.
Cp is spherically complete! This follows from:
Proof There exist r1 > r2 > . . . in |K ∗ | such that r := limn→∞ rn > 0. Let {s1 , s2 , . . .} be a
countable dense subset. There is a ‘closed’ ball B1 of radius r1 such that s1 ∈ / B1 . Since B1
decomposes into infinitely many ‘closed’ balls of radius r2 (why infinite?) there is a ‘closed’
ball B2 ⊂ B1 such that s2 ∈ / B2 . Going on this way we find a nested sequence B1 ⊃ B2 ⊃ . . .
of closed balls with radius Bn = rn , and sn ∈ / Bn for each n.
T
Let B := n Bn . If B 6= ∅, it would be a ‘closed’ ball of radius r, hence open so the set
{s1 , s2 , . . .} must meet B. But on the other hand, by construction {s1 , s2 , . . .} ∩ B = ∅, a
contradiction.
FROM NOW ON IN THIS COURSE K = (K, |.|) IS A N.A. VALUED COMPLETE FIELD.
WE ASSUME THAT |.| IS NON-TRIVIAL I.E. THERE IS A λ ∈ K WITH |λ| =
6 0, |λ| =
6 1.
Exercise 1.H. Show that {λ ∈ K : |1 − λ| < 1} is a multiplicative subgroup of {λ ∈ K :
|λ| = 1}.
n
X
– Let a1 , a2 , . . . ∈ K. This sequence is called summable if lim ai exists. Then
n→∞
i=1
A STUDENT’S DREAM COME TRUE:
lim an = 0 =⇒ a1 , a2 , . . . is summable.
n→∞
1
= 1 + p + p2 + . . .
1−p
∞
P
Because lim n! = 0 in every Qp the sum n! exists in every Qp . The following problem has
n→∞ i=0
been open since 1971.
P∞
PROBLEM: Can n! be rational for some prime p?
i=0
g.c.d.(!n, n!) = 2.
∞
P
Exercise 1.I. Compute n (n!) in Qp .
n=0
(where ∞−1 := 0 , 0−1 := ∞) is called the radius of convergence of the power series an xn .
P
Just like in the complex case one proves that a0 , a1 x, a2 x2 , . . . is summable for |x| < R, not
summable for |x| > R. The behaviour on the ‘boundary’ |x| = R is much easier that in the
complex case: a0 , a1 x, . . . is either summable everywhere on {x : |x| = R} or nowhere. This is
because the region of convergence {x ∈ K : a0 , a1 x, . . . is summable} = {x ∈ K : lim an xn =
n→∞
0}.
Exercise 1.J. ∞
xn is {x ∈ K : |x| < 1}.
P
(i) Prove that the region of convergence of
n=0
∞
P xn
(ii) Prove that, in Cp , n exists if and only if |x| < 1.
n=1
So, we can define the p-adic logarithm via the formula
∞ n
X x
− logp (1 − x) = (|x| < 1)
n
n=1
n = a0 + a1 p + . . . + as ps
[BACKGROUND: One can prove that the p-adic logarithm maps {x ∈ Cp : |1 − x| < 1} onto
Cp and that logp x = 0 if and only if x is a root of unity.]
Proof (Compare Exercise 1.D (vii)) The relation |f (x)−f (y)| < ε (x, y ∈ X) is an equivalence
S
relation and yields a partition of X into clopen sets: X = Ui .
i∈I
Choose ai ∈ Ui for each i and define
1.2.6 Differentiability
Now we take for X a subset of K. To avoid problems, assume X is without isolated points.
A function f : X −→ K is called differentiable at a ∈ X if
f (x) − f (a)
f 0 (a) := lim
x→a x−a
exists (you can imagine how to define lim ). In the same spirit we define (everywhere) dif-
x→a
ferentiable function, derivative. The well-known rules for differentiation of sums, products,
quotients, compositions (‘chain rule’) hold also in the n.a. case. The proofs are classical.
A differentiable function is continuous. Rational functions without poles in X are differen-
tiable.
To be able to define analytic functions properly we need the following exercise on ‘double
sequences’.
Exercise 1.M. For m, n ∈ N, let amn ∈ K. We say that limm+n→∞ amn = 0 if, for each
ε > 0, the set {(m, n) ∈ N × N : |amn | ≥ ε} is finite. Show that the following statements
(α), (β), (γ) below are equivalent.
and
n
tkn := an (x − b)k (b − a)n−k .
k
Now clearly limk→∞ tkn = 0 for each n and since
n k n−k
|tkn | ≤ |an | r r ≤ |an | rn
k
( nk is an integer, hence nk ≤ 1)
∞ X
X ∞ ∞ X
X ∞ ∞
X
f (x) = tkn = tkn = bk (x − b)k
n=0k=0 k=0n=0 k=0
∞
n
(b − a)n−k .
P
where bk := an k
n=0
Definition Let B be a ‘closed’ ball in K with radius r ∈ |K ∗ |. A function f : B −→ K is
called analytic if there exists an a ∈ B and a0 , a1 , . . . ∈ K such that
∞
an (x − a)n
P
(∗) f (x) = (x ∈ B).
n=0
With the help of Lemma 3 one easily derives that sums and products of analytic functions on
B are analytic. Analytic functions are differentiable. If f is as in (∗) then
∞
X
0
f (x) = nan (x − a)n−1 .
n=1
If f is a complex analytic function defined on some bounded domain the maximum principle
states that |f | takes its maximum on the boundary. A similar result holds also for our n.a.
case; we include a proof since it is completely different from the ‘complex’ proof.
Proof We only consider the case where B = B(0, 1). (The reader can furnish the missing
details). Thus, we have
∞
an xn
P
(∗∗) f (x) = (|x| ≤ 1).
n=0
for some N (since |an | < 1 for large n), for which aN 6= 0. Thus, a nonzero polynomial in k[x]
has each element of k as a root. But k is infinite, a contradiction.
Thus we have max{|f (x)| : |x| ≤ 1} = maxn |an |. To prove that also max{|f (x)| : |x| = 1} =
maxn |an | observe that |f (x)| < 1 for all |x| = 1 yields
N
X
0= an xn
n=0
for all x ∈ k \ {0}. Now reason as above, using that k \ {0} is infinite as well.
[BACKGROUND: Krull Valued Fields.
When looking at the requirements for a non-archimedean valuation:
one notices that, unlike for the archimedean triangle inequality, addition of real numbers
does not play a role; one only needs multiplication and ordering. This leads to the following
generalized concept of valued fields.
Let G be a commutative, multiplicatively written, totally ordered group such that
(Example: (0, ∞) with the usual multiplication and ordering. Observe that R\{0} is a totally
ordered commutative group but that it does not satisfy (*)!)
Augment G with an element 0 and put for all g ∈ G
0·g =g·0=0·0=0
0<g
are two distinct elements of G we say that g < g 0 if nk < mk . Or: if nk = mk and nk−1 < mk−1 .
Or if nk = mk , nk−1 = mk−1 and nk−2 < mk−2 , etc.
Observe that this G is not isomorphic -as an ordered group- to a multiplicative subgroup of
(0, ∞) since it contains bounded non-trivial subgroups like
G1
G1 ⊕ G2
..
.
G1 ⊕ G2 ⊕ . . . ⊕ Gn
..
.
|0| : = 0
|λ| : = 1 if λ ∈ R \ {0}
|Xn | = (1, 1, . . . , gn , 1, 1, . . .)
↑
n-th place
for all x, y ∈ E, λ ∈ K.
(Requirement (iii) is not a logical necessity but there are good reasons for assuming it, see
also Exercise 1.N).
A complete normed space is, as customary, called a Banach space.
Two norms k.k1 and k.k2 on a K-vector space E are called equivalent if the metrics (x, y) 7−→
kx − yk1 and (x, y) 7−→ kx − yk2 induce the same topology. The following Proposition can be
proved as in the archimedean case.
Proposition 5 Two norms on a K-vector space E, say k · k1 and k · k2 , are equivalent if and
only if there exist constants 0 < c < C such that
ck · k1 ≤ k · k2 ≤ Ck · k1 .
On a finite-dimensional space all norms are equivalent, and the space is a Banach space with
respect to each norm.
4. Let X be a topological space, let BC(X) be the space of all f ∈ C(X) that are bounded.
It is a Banach space with respect to the supremum norm
PROBLEM (1960) Let (E, k·k) be a Banach space over K. Does there exist an equivalent
norm k · k0 such that kEk0 ⊂ |K|, i.e. for each nonzero x ∈ E there exists a λ ∈ K such
that kλxk = 1?
There are some partial answers and reductions, but the answer is so far unknown! A
good test case would be l∞ over Cp .
Clearly x ∈ B(0, r), λ ∈ K, |λ| ≤ 1 =⇒ λx ∈ B(0, r), but also because of the strong triangle
inequality, x, y ∈ B(0, r) =⇒ x + y ∈ B(0, r). So, in algebraic language, B(0, r) is a module
over the ring {λ ∈ K : |λ| ≤ 1}! Such submodules are called absolutely convex.
[BACKGROUND: a seminorm on a K-vector space E is a map q : E −→ [0, ∞) satisfying
q(λx) = |λ|q(x), q(x + y) ≤ max(q(x), q(y)) but q(x) is allowed to be 0 for nonzero x. Like
in the classical case one can define a locally convex topology on E as the weakest topology
for which each member of a collection P of seminorms is continuous. A subbase of zero
neighbourhoods is formed by the sets {x ∈ E : q(x) < ε} where q ∈ P, ε > 0. By the above
observation these sets are absolutely convex. There is an extensive theory on locally convex
spaces on K.]
[BACKGROUND: Normed spaces over Krull valued fields?
Let (K, | · |) be a Krull valued field with value group G. A norm k · k on a K-vector space E
should satisfy (i), (ii), (iii) of page 17 but it is not at once clear what a ‘natural home’ for norm
values should be. Only admitting G ∪ {0} turns out to be too restrictive. This leads to the
concept of a so-called G-module; this is an ordered set X on which G acts monotonically. For
full details see p-adic Functional Analysis, Lecture Notes in Pure and Applied Mathematics
207, 233-293. Marcel Dekker. New York (1999).
Remark So far, nobody has attempted constructing a theory of locally convex spaces over
Krull valued fields!]
Let us return to the basic theory.
(take the trouble to verify this!). Then kT (λx)k ≤ 1, so kT xk ≤ |λ|−1 which is, by (*),
≤ |π|−1 δ −1 kxk, and we have
kT xk ≤ M kxk (x ∈ E)
Prove that, indeed, T y ∈ c00 and that T maps l∞ linearly and isometrically onto c00 . Hence,
in popular form: ‘the dual of c0 is l∞ ’.
Like in the classical theory one can prove that L(E, F ) is a Banach space as soon as F is a
Banach space. Hence, E 0 , E 00 , E 000 , . . . are all Banach spaces.
Here is a construction that works in both classical and n.a. context, but that you may not
have seen yet: forming of quotients.
Let E be a normed space over K, let D ⊂ E be a closed subspace. We define a natural norm
on E/D as follows. Let π : E −→ E/D be the canonical map (assigning to each x ∈ E the
coset x + D). Set
kπ(x)k := dist(x, D) = inf{kx − dk : d ∈ D}.
Straightforward verification shows that kλzk = |λ|kzk, kz + uk ≤ max(kzk, kuk) for all z, u ∈
E/D, λ ∈ K. Suppose kzk = 0; we prove that z = 0.
z has the form x + D, so kzk = inf{kx − dk : d ∈ D} = 0. Hence, there are d1 , d2 , . . . ∈ D
with limn→∞ kx − dn k = 0 i.e., x = limn→∞ dn . Since D is closed, x ∈ D which implies that
x + D is the zero element of E/D.
The Uniform Boundedness Principle, the Banach Steinhaus Theorem, the Closed Graph The-
orem and the Open Mapping Theorem all rest on the Baire Category Theorem and some
linearity considerations and therefore remain valid in our n.a. theory.
We like to conclude this introductory chapter by signalling two striking differences with the
classical case.
(so the sequence a0 , a1 , . . . is uniquely determined by f ) and the space Hr is (linearly) iso-
metrically isomorphic to {(a0 , a1 , . . .) : |an |rn −→ 0}, with the max norm.
For r = 1 we see that H1 ∼ = c0 . (Show that also Hr ∼ = c0 if r ∈ |K ∗ |.) So, Hr is a Banach
space.
COROLLARY. The uniform limit of a sequence of analytic functions on B(0, r) (r ∈ |K ∗ |) is
again analytic.
Exercise 1.Q. Show that k · k∞ on Hr is multiplicative i.e. for f, g ∈ Hr
kf gk∞ = kf k∞ kgk∞
ORTHOGONALITY
Exercise 2.A. Requirement (iii)’ seems to be a kind of arbitrary relaxation of (iii), born out
of emergency. But it is not such a ‘wild’ generalization as one may think. In fact, prove the
following. In a vector space over C, any form ( · , · ) satisfying (i), (ii) and (iii)’ is either an
inner product or else −( · , · ) is an inner product!
For E as above and x ∈ E we put
p
kxk = |(x, x)|.
Theorem 8 Let ( · , · ) be an inner product on a Banach space E over K such that |(x, x)| =
kxk2 for all x ∈ E. Suppose that each closed subspace has an orthogonal complement. Then
dim E < ∞.
[BACKGROUND: Theorem 8 is not the last word that is to be said about the subject. The
picture changes completely if we allow Krull valuations on the scalar field. In fact, let L be
the completion of the Krull valued field R(X1 , X2 , . . .) as constructed on pages 16–17. Put
X0 = 1. Let E be the L- vector space of all sequences (ξ0 , ξ1 , . . .) ∈ LN for which ∞ 2
P
i=0 ξi Xi
converges in L. For x, y ∈ E, say x = (ξ1 , ξ2 , . . .) and y = (η1 , η2 , . . .) put
∞
X
(x, y) := ξi ηi Xi .
i=0
H. Keller showed in 1980 that ( · , · ) is an inner product in the sense of (i)0 , (ii)0 , (iii)0 of page 23
(where λ∗ = λ for all λ ∈ L), but, what is more, he proved has every closed subspace D of E has
an orthogonal complement, i.e. D + D⊥ = E when D⊥ = {x ∈ E : (x, y) = 0 for all y ∈ D}.
The study those ‘non-classical Hilbert spaces’ is in full progress.]
So, for our theory to develop, inner products do not seem appropriate. But we do have a
powerful non-archimedean concept, valid in any normed space:
Definition Let x, y be elements of a normed space E over K. We say that the vector x is
(norm-)orthogonal to y if the distance of x to the space Ky is precisely kxk:
Proposition 9 If x ⊥ y then y ⊥ x.
Prove that (1, 0) ⊥ (0, 1) (not surprising), but also that (1, 1) ⊥ (1, 0) and (1, 1) ⊥ (0, 1).
Exercise 2.C. Let E be a normed space. Prove that 6⊥ (‘not orthogonal’) is an equivalence
relation on E \ {0}. Thus, if x ⊥ y and x 6⊥ z then y ⊥ z!
Inspired by Proposition 10 we now define:
Definition A sequence e1 , e2 , . . . of non-zero vectors in a normed space E is called orthogonal
if for each n ∈ N, λ1 , . . . , λn ∈ K
n
X
λi ei
= max{kλi ei k : 1 ≤ i ≤ n}
i=1
(i.e. each ei is orthogonal to every vector in the linear span of {ej : j 6= i}). The sequence is
orthonormal if, in addition, ken k = 1.
The difference with classical orthogonality is demonstrated again in the following
∞
X
Definition e1 , e2 , . . . is called an orthogonal base for E if x ∈ E has an expansion x = λi ei
i=1
for some suitable λi ∈ K.
Clearly such an expansion is unique.
Proof Let π ∈ K, 0 < |π| < 1. Then there exists µ1 , µ2 , . . . ∈ K such that |π| ≤ kµn en k ≤ 1
∞
X
for all n ∈ N. For (λ1 , λ2 , . . .) ∈ c0 , put T (λ1 , λ2 , . . .) = λn µn en .
n=1
One verifies immediately that T is a linear map c0 −→ E, that, for x = (λ1 , λ2 , . . .) ∈ c0 :
so that T is a homeomorphism, hence T c0 is complete. But also T c0 contains all finite linear
combinations of e1 , e2 , . . ., hence T c0 is dense. It follows that T c0 = E, which is what we
wanted to prove.
The most famous example of an orthogonal base is the so-called Mahler base of C(Zp −→ Cp ):
For x ∈ Zp and n ∈ {0, 1, 2, . . .} we set
x x(x − 1) . . . (x − n + 1)
en (x) := :=
n n!
Then we have
3. en (−1) = (−1)n ,
5. e0 , e1 , . . . is an orthonormal sequence.
kλ0 e0 + . . . + λn en k∞ ≥ kλ1 e1 + . . . + λn en k∞ .
kλ0 e0 + . . . + λn en k∞ ≥ max{kλi ei k : 0 ≤ i ≤ n}
= max{|λi | : 0 ≤ i ≤ n}).
To conclude the proof that e0 , e1 , . . . is an orthonormal base some work has to be done. First,
we need
Exercise 2.D. (Continuation of Exercise 1.K)
(ii) Show that spj = sj for each j ∈ N (here again sm is the sum of digits of n in base p)
To get an idea for the proof we imagine that f ∈ C(Zp −→ Cp ) has an expansion. Then what
would be the candidates for the coefficients? So suppose f (x) = a0 x0 + a1 x1 + . . .
By using
x+1 x x
= + (n ≥ 1)
n n n−1
we find
∞ X ∞
X x x
(Ef )(x) = an + an
n n−1
n=0 n=1
∞
X x
= f (x) + an+1
n
n=0
so that
∞
X x
(E − I)f (x) = an+1
n
n=0
and by putting x = 0:
(E − I)k f (0) = ak .
n
X n
Because (E − I)n = (−1)n−k E k we have
k
k=0
n
X n
(*) an = (−1)n−k f (k)
k
k=0
and
n
k
X n
(E − I) f (x) = (−1)n−k f (x + k).
k
k=0
then
n
X n
an = (−1)n−k f (k).
k
k=0
Corollary 15 (No p-adic Haar integral) Let ϕ ∈ C(Zp −→ Cp )0 have the property that
ϕ(f1 ) = ϕ(f ) (f ∈ C(Zp −→ Cp ), where f1 is the ‘shift’ f1 (x) = f (x + 1). Prove that ϕ = 0.
1
Exercise 2.F. Let α ∈ Cp , |α| < p 1−p . Then exp αx is defined for all x ∈ Zp . Show that its
Mahler expansion is
∞
X x
(exp α − 1)n .
n
n=0
[BACKGROUND: One can prove that if a Banach space E has an orthogonal base then so has
every closed subspace D. However it is not always true that D has an orthogonal complement
i.e. a closed subspace S with D + S = E and D ⊥ S. It depends on K and kEk.]
In the next Chapter we will use the following version of ‘approximate orthogonality’.
Definition Let t ∈ (0, 1]. A sequence e1 , e2 , . . . in a normed space is called t-orthogonal if for
all n ∈ N, λ1 , . . . , λn ∈ K
Xn
λi ei
≥ t max{kλi ei k : i ∈ {1, . . . , n}}.
i=1
In the same spirit we have the notion of t-orthogonal base of a Banach space.
1- orthogonal = orthogonal.
If e1 , e2 , . . . is a t-orthogonal base of E and we introduce a new norm k k∼ by
∼
∞
X
λ n n
= max kλn en k
e
n
n=0
COUNTABILITY
3.1 Introduction
A topological space is called separable if it has a countable dense subset. If we have a metric
space that is separable then any subset is also separable.
In classical Functional Analysis some role is played by the separable Banach spaces.
In our n.a. case separability is not the right notion to work with. In fact, if K happens to be
non-separable then each onedimensional space has the same property.
To overcome this problem one has “linearized”the definition of separability as follows.
Definition A normed space E is called of countable type if it has a countable subset whose
linear hull is dense.
It is an easy exercise to show that if K itself is separable then a normed space over K is of
countable type if and only if it is separable.
It has been an open question for quite some time as to whether each separable Banach space
∞
X
over C has a Schauder base e1 , e2 , . . . i.e. each vector has a unique expansion λn en , where
n=1
λn ∈ C. Only in 1974 Enflo proved that the answer was negative.
Further, in the complex case there are many non-isomorphic separable Banach spaces e.g. lp
for 1 ≤ p < ∞. In the non-archimedean case the situation is much simpler as we will show
now.
Lemma 16 Every finite-dimensional space has, for each t ∈ (0, 1), a t-orthogonal base.
Proof Let E have dimension n, let x1 , . . . , xn be an algebraic base for E. Choose t1 , . . . , tn−1 ∈
√
(0, 1) such that their product t1 t2 . . . tn−1 ≥ t (e.g. ti = n−1 t).
Set e1 = x1 , D1 := [e1 ] = [x1 ] ([·] indicates linear span). Now D1 is closed, x2 ∈ / D so
dist(x2 , D2 ) > 0.
kx2 − d1 k ≤ t−1
1 dist(x2 , D1 ).
ke2 k ≤ t−1
1 dist(e2 , D1 ).
kx3 − d2 k ≤ t−1
2 dist(x3 , D2 ).
ke3 k ≤ t−1
2 dist(e3 , D2 ).
ke2 k ≤ t−1
1 dist(e2 , [e1 ])
ke3 k ≤ t−1 dist(e3 , [e1 , e2 ])
2
(∗) .
.
.
ke k ≤ t−1 dist(e , [e , . . . , e
n n n 1 n−1 ]).
so we have
kλ1 e1 + . . . + λn en k ≥ tn kλn en k
(this formula is also true for λn = 0). By the extension of van Rooij’s Principle, Exercise 2.G,
we have
kλ1 e1 + . . . + λn en k ≥ tn kλ1 e1 + . . . + λn−1 en−1 k.
Now with the same trick as above we can prove
kλ1 e1 + . . . + λn en k ≥ tn kλn en k
kλ1 e1 + . . . + λn en k ≥ tn tn−1 kλn−1 en−1 k
..
.
kλ1 e1 + . . . + λn en k ≥ tn tn−1 . . . t1 kλ1 e1 k.
(show first that this limit exists!). Verify that k · k is a norm in K 2 and that k(α1 , α2 )k ∈ |K|
for all (α1 , α2 ) ∈ K 2 . Show that (α1 , α2 ) ⊥ (1, 0) =⇒ α1 = α2 = 0. Finally, show that two
arbitrary non-zero vectors in K 2 are not orthogonal.]
Theorem 17 Let t ∈ (0, 1). Then each Banach space of countable type has a t-orthogonal
base.
Proof By Theorem 7 the space E/D is Banach. That it is of countable type is obvious. By
p
Theorem 17 it has a t0 /t-orthogonal base g1 , g2 , . . .. Let π : E −→ E/D be the canonical
map. By definition of the quotient norm there are f1 , f2 , . . . ∈ E with π(fi ) = gi and kfi k ≤
p
(t/t0 )kgi k for each i. It is easily seen that the map
∞
X ∞
X
Q: λi gi 7−→ λi fi (λi gi → 0)
i=1 i=1
is in L(E/D, E) and that π ◦ Q is the identity on E/D, and that kQk ≤ (t0 )−1 t. Then put
P := I − Q ◦ π; one verifies that kP k ≤ (t0 )−1 , that P is a projection onto D. Finally, if
e1 , e2 , . . . is a t-orthogonal base of D then e1 , f1 , e2 , f2 , . . . is easily seen to be a t0 -orthogonal
base of E.
DUALITY
Proof Like in the classical case, through Zorn’s Lemma it suffices to prove it for the case
E = D + Ka where a ∈ E \ D. By linearity, if we fix η := f (a) then f is determined:
We notice that (*) is true for λ = 0. Then, to have (*) for λ 6= 0 it suffices to have (*) for
λ = 1:
|η + f (d)| ≤ kf kka + dk (d ∈ D).
Thus
η ∈ B(−f (d), kf kka + dk) (d ∈ D).
To be able to choose η this way we need that the balls have a non-empty intersection. By
spherical completeness it suffices to prove that each two of them have a non-empty intersection,
i.e. we have to see if two ‘centers’−f (d1 ), −f (d2 ) have distance ≤ the maximum of kf kka +
d1 k, kf kka + d2 k. But that is true:
For each normed space E over K we have, as in the complex case, the canonical map jE :
E −→ E 00 given by
jE (x)(f ) = f (x) (f ∈ E 0 , x ∈ E).
E T -F
jE jF
? ?
E 00 - F 00
T 00
is commutative.
Exercise 4.B. Apply Corollary 20 and Theorem 21 for one-dimensional subspaces D to prove
the following: Let E be a Banach space of countable type or let K be spherically complete.
Then the canonical map jE : E −→ E 00 is isometrical.
A normed space E is called reflexive if jE is an isometrical bijection E −→ E 00 . Reflexive
spaces are complete.
One proves easily that finite - dimensional spaces are reflexive. We first study reflexivity for
spherically complete K.
Lemma 22 Let K be spherically complete, let E be a reflexive Banach space over K. Then
every closed subspace is also reflexive.
Proof Let D be a closed subspace, let i : D −→ E be the inclusion and π : E −→ E/D the
quotient map.
By Exercise 4.A. we have the commutative diagram:
D i -E π - E/D
jD jE jE/D
? ? ?
D00 - E 00 - (E/D)00
i00 π 00
Proof We will construct a θ ∈ c000 that is not in jc0 (c0 ). Let, for n ∈ N, δn be the n-th
coordinate function c0 −→ K :
(ξ1 , ξ2 , . . .) 7−→ ξn .
∞
X
Then δn ∈ c00 . Let D = [δ1 , δ2 , . . .] ⊂ c00 . The function (ξ1 , ξ2 , . . .) 7−→ ξn is in c00 but not
n=1
in D, so we have D 6= c00 . Hence c00 /D 6= {0} and there is a nonzero ω ∈ (c00 /D)0 (Here we
used that K is spherically complete). Combining ω with the quotient map c00 −→ c00 /D:
ω
θ : c00 −→ c00 /D −→ K
we find a θ ∈ c000 , θ 6= 0 on D. We claim that this θ is not in Im jc0 . In fact, suppose we would
have an x ∈ c0 with jc0 (x) = θ i.e. θ(f ) = f (x) for all f ∈ c00 , say x = (ξ1 , ξ2 , . . .). By taking
for f = δn we find
ξn = δn (x) = θ(δn ) = 0,
implying x = 0 hence θ = 0, a contradiction.
This leads to the disappointing result:
Theorem 24 Let K be spherically complete. Then, the only reflexive spaces are the finite-
dimensional ones.
Lemma 25 Let E be a spherically complete Banach space (no condition on K). Suppose D
is a closed subspace of E. Then E/D is also spherically complete.
Proof Let B(y1 , r1− ) ⊃ B(y2 , r2− ) ⊃ . . . be a nested sequence of ‘open’ balls in E/D. It suffices
to show that their intersection is non-empty. Let π : E −→ E/D be the quotient map. Choose
an x1 ∈ E with π(x1 ) = y1 . Then, from the way the norm on E/D was defined we infer that
f -
E K (On E/Ker f the canonical norm.
A ρ is the unique map making the
A f1
ρ
diagram conmute.)
U
A
E/Ker f
By Lemma 25, the space E/Ker f is spherically complete. But ρ is a linear homeomorphism
between two one-dimensional spaces viz. E/Ker f and K. Then there is a constant c > 0
such that |ρ(x)| = ckxk for all x ∈ E/Ker f . Hence, ρ maps balls onto balls and so K must
be spherically complete as well. Contradiction.
(verify this!) Like in the proof of Lemma 25 we can find, for given balls B(y1 , r1− ) ⊃
B(y2 , r2− ) ⊃ . . . in l∞ /c0 , balls B(x1 , r1− ) ⊃ B(x2 , r2− ) ⊃ . . . in l∞ such that π(xn ) = yn ,
π(B(xn , rn− )) = B(yn , rn− ) for each n. Now write
x1 = (x11 , x12 , . . .)
x2 = (x21 , x22 , . . .)
..
.
Corollary 28 Let K be not spherically complete. Then (l∞ /c0 )0 = {0}. If f ∈ (l∞ )0 and
f = 0 on c0 then f is identically zero.
Proof Combine Lemma 26 and Proposition 27. For the second statement, observe that f = 0
on c0 implies that (with π : l∞ −→ l∞ /c0 the canonical map)
Proof With an ingenious trick. Suppose we could extend the given functional to a h ∈ (l∞ )0 .
Define the shift operator Ω : l∞ −→ l∞ by the formula
s = (ξ1 , ξ1 + ξ2 , . . .)
where (
0 if n ∈
/X
bn =
a−1
ni ξi if n = ni .
We have for (ξ1 , ξ2 , . . .) ∈ c0
P P∞ P∞
(∗) (f ◦ T )(ξ1 , ξ2 , . . .) = f (b1 , b2 , . . .) = n∈X an bn = i=1 ani bni = i=1 ξi .
Theorem 31 (c0 and l∞ are reflexive) Let K be not spherically complete. For each x ∈ c0
define fx ∈ (l∞ )0 by
∞
X
fx (y) =< x, y >= ξn ηn (y ∈ l∞ )
n=1
≤ kxk∞ kyk∞
COMPACTOIDS
This little chapter treats a third important notion (next to ‘orthogonality’and ‘of countable
type’), namely that of a compactoid set. We do not have much time to explain it in the
lectures, so the only purpose of this chapter is to give you an idea about these notions, its
impact on the theory but without details and without proofs.
A subset X of a Banach space E (over K or over R or C) is called precompact if, for any ε > 0,
X can be covered by finitely many ε-balls. In other words, X is precompact if for every ε > 0
there is a finite set F ⊂ E such that
X ⊂ B(0, ε) + F
One proves by classical means that X is precompact if and only its closure X in E is compact.
(For this, it is necessary that E is complete).
In classical theory convex compact sets play an important role in Functional Analysis
(e.g. Alaoglu-Bourbaki Theorem, Krein-Milman Theorem, Choquet Theory). How are the
prospects in the non-archimedean case? Recall that a subset C of a K-vector space is absolutely
convex if 0 ∈ C and for x, y ∈ C, λ, µ ∈ K, |λ| ≤ 1 |µ| ≤ 1 we have λx + µy ∈ C. So, if K is
not locally compact the only absolutely convex compact subset of a Banach space is {0}. This
difficulty is similar to the separability problem on page 32, and we will solve it in a similar
way; this time by ‘convexifying’ the definition of precompactness.
Definition A subset X of a Banach space E over K is called compactoid if
for each ε > 0 there is a finite set F ⊂ E such that
X ⊂ B(0, ε) + aco F,
where ‘aco’ stands for ‘absolutely convex hull’, the smallest absolutely convex set containnig
the given set. In our case if F = {a1 , . . . , an } then aco F = {λ1 a1 +. . .+λn an : λi ∈ K, |λi | ≤ 1
for each i}.
A ⊂ aco {e1 , e2 , . . .} ⊂ λ A
– Let X ⊂ E be bounded. Then X is a compactoid if and only if for every t ∈ (0, 1] each
t-orthogonal sequence in X tends to 0.
(Here, Vol(x1 , . . . , xn ) := kx1 k dist(x2 , [x1 ]) dist(x3 , [x1 , x2 ]) . . . dist(xn , [x1 , . . . xn−1 ])).
An A ∈ L(E) is called compact operator if it maps the unit ball into a compactoid.
One can derive a theory a la Riesz for compact operators:
– Alternative of Fredholm.
..
.
One may think of alternative ways of ‘convexifying’ notions of compactness. The following
was introduced by Springer.
Let A ⊂ E be convex. We call A c-compact if the following holds. A is closed and for each
collection of closed convex sets {Ci : i ∈ I} in A such that Ci,1 ∩ . . . ∩ Ci,n 6= ∅ for any
T
i1 , . . . , in ∈ I we have Ci 6= ∅ (This translates a well-known property of closed subset of a
i
compact set).
However:
– If K is not spherically complete each c-compact set is either empty or a singleton set.
But: