A word about Logic: Sentential theory of inference

Reference: Introduction to Logic: Patrick Suppes, Chapter 2, Affiliated East-West Press Pvt. Ltd., available Amazon India.

Criterion I. Given a set of premises, the rules of logical derivation must permit us to infer ONLY those conclusions which logically follow from the premises.

Criterion II. Given a set of premises, the rules of logical derivation must permit us to infer ALL conclusions which logically follow from the premises.

For our purposes, we may define the restricted notion of sentential interpretation thus:

A senence P is a sentential interpretation of a sentence Q if and only if P can be obtained from Q by replacing the component atomic sentences of Q by other (not necessaily distinct ) sentences.

In other words, a sentential interpretation of a sentence must preserve its sentential form.

It should be understood that if a component atomic sentence seems occurs more than onee in a sentence, any sentential interpretation of that sentence must replace that component atomic sentence by the same thing in both of the occurrences.

A sufficient (but not necessary) condition for one sentence to be a logical consequence of another may be stated as follows:

(I) Q logically follows from P if every sentential interpretation of the implication P \Longrightarrow Q is true.

(II) A tautology is a sentence whose atomic sentential interpretations are all true.

(III) If every atomic sentential interpretation of a sentence is true then every sentential interpretation of the sentence is true.

(IV) Q logically follows from P if Q is tautologically implied by P.

The whole sentential theory of inference is summarized by (IV).

Criterion I says that the rules of inference must be sound, that is, they must not permit a fallacious statement. Combining the criterion with (I) we may obtain a working criterion for testing the validity of a proposed rule of inference. First we see that Q should be derivable from P by use of the rules only if every interpretation of P \Longrightarrow Q is true. The test criterion is then:

If a new proposed rule of inference permits the derivation of a false conclusion from true premises, reject it.

\bf{The \hspace{0.1in} Three \hspace{0.1in} Sentential \hspace{0.1in} Rules \hspace{0.1in}of \hspace{0.1in} Interpretation}

Rule P. We may introduce a premise at any point in a derivation.

Rule T. We may introduce a sentence S in a derivation if there are preceding sentences in the derivation such that their conjunction tautologically implies S.

Rule CP. If we can derivce S from R and a set of premises, then we may derive R \Longrightarrow S from the set of premises alone.

Cheers,

Nalin Pithwa

Topological Preliminaries : K Chandrashekaran

Reference: A Course on Topological Groups by K Chandrashekharan, Hindustan Book Agency, available Amazon Indiia.

Chapter 1. Topological Preliminaries:

Section 1.1 Topological Space:

A topology T in a set X is a class of subsets of X (called open sets) satisfying the following axioms:

(1) the union of any number of open sets is open

(2) the intersection of any two (or finitely many) open sets is open

(3) X and the emtpy set of null set \phi are both open.

A topological space is a set X with a topology T in X. T is called the trivial topology if T = (X, latex \phi$) and is called the discrete topology if T= (A: A \subset X). (PS: the trivial topology has too many limit points; it is like a lumped mass) whereas the discrete topology has no limit points, yet this latter topology is quite useful ! )

Let X be a topological space and A \subset X. We define an induced topology in A as follows: the class of open sets in A is the class of sets of the form U \bigcap A where U runs through all open sets in X.

In a topological space X, a closed set is any set whose complement is an open set.

The closure \overline{E} of any set E is the intersection of all closed sets containing E. By axioms (1) and (3) above, the closure is a closed set.

A neighbourhood of x \in X is any open set containing x. Let X_{1} and X_{2} be two topological spaces and f a mapping

f : X_{1} \rightarrow X_{2}

A mapping is an assignment to each x \in X_{1} of an element f(x) in X_{2}. Then f is continuous if and only if for every open set O_{2} \in X_{2}, the set f^{-1}(O_{2}) is an open set in O_{1}. Here, f^{-1}(O_{2}) = \{ x: x \in X_{1}, f(x) \in O_{2}\}.

Let x \in X_{1}. We say that f is continuous at x, if to every neighbourhood U of f(x) there exists a neighbourhood V of x such that f(V) \subset U. We say that f is continuous, if f is continuous at every point of X_{1}. This definition is equivalent to the preceding one.

The mapping f is said to be open if for every open set O_{1} \in X_{1}, the set f(O_{1}) is an open set in X_{2}.

Let f: X \rightarrow Y and g: Y \rightarrow Z be two continuous mappings. Then the composite mapping g \circ f: X \rightarrow Z is continuous.

Examples

(i) If X is a discrete topological space, Y is a topological space, then every mapping f: X \rightarrow Y is continuous. (PS: this is vacuously true).

(ii) If X is any topological space, and Y a set wtih the trivial topology, then f: X \rightarrow Y is continuous.

(iv) f_{X}: X \rightarrow X is the identity mapping.

The topological product of two topological spaces X_{1} and X_{2} is the topological space X = X_{1} \times X_{2}, whose set is the cartesian product of X_{1} and X_{2}, namely, \{ (x_{1}, x_{2}) : x_{1} \in X_{1}, x_{2} \in X_{2}\} with the open sets being all unions of sets of the form O_{1} \times O_{2} where O_{1} is an open set in X_{1} and O_{2} is an open set in X_{2}. Sets of the form O_{1} \times O_{2} form a basis of open sets in X_{1} \times X_{2}

Note that if p_{1}: X_{1} \times X_{2} \rightarrow X_{1} and p_{2}: X_{1} \times X_{2} \rightarrow X_{2} are the projections defined by p_{1}(x_{1}, x_{2}) = x_{1} and p_{2}(x_{1}, x_{2}) = x_{2} respectively then p_{1} and p_{2} are continuous mappings where x_{1} \in X_{1} and x_{2} \in X_{2}.

COVERING. A family \{ V_{\alpha} \} where \alpha \in I of subsets of a set X is a covering of X, if \bigcup_{\alpha \in I}V_{\alpha} = X. That is to say, each point of X belongs to at least one V_{\alpha}. If further, X is a topological space, and each V_{\alpha} is an open subset of X, we say that \{ V_{\alpha}\} is an open covering of X.

A covering \{V_{\alpha}\}_{\alpha \in I} is a finite covering if I is a finite set.

COMPACT SETS. A topological space X is said to be compact if every covering of X contains a finite subcover. That is to say, X = \bigcup_{\alpha \in I}U_{\alpha} where U_{\alpha} is open implies that there exists \alpha_{1}, \alpha_{2}, \ldots, \alpha_{n} all belonging to I such that \bigcup_{1 \leq i \leq n}U_{\alpha_{i}}=X. A subset A \subset X is compact if A is compact in the induced topology.

Examples:

(i) The \Re is not compact.

(ii) Let X be a topological space and A \subset X/ If A consists of finitely many elements of X, then A is compact. (PS: A finite point set has no limit points).

(iii) Let a, b belong to reals. Let a \leq b. Then the set a \leq x \leq b is compact.

An equivalent definition of compactness is the following:

If \{ F_{\alpha}\} is a family of closed sets such that every finite subfamily has a non-empty intersection, then \bigcap_{\alpha}F_{\alpha} \neq \phi. (If \{ F_{\alpha}\} is a family of closed sets with the finite intersection property, then the intersection of the whole class is non-empty).

A closed subset of a compact set is compact. A continuous image of a compact set is compact. (If X_{1} is compact, X_{2} is Hausforff, and f: X_{1} \rightarrow X_{2} is continuous, then f(X_{1}) is closed in X_{2}).

A product of compact spaces is compact (Tychonoff theorem).

A space X is locally compact if for every x \in X, there is a neighbourhood O_{1} of x such that the closure of O_{1} is compact.

Every compact space is locally compact but not vice-versa.

HOMEOMORPHISMS.

Let X and Y be topological spaces and f a mapping f: X \rightarrow Y. We say that f is a homeomorphism if f is one-to-one, onto and both f and f^{-1} are continuous.

Two topological spaces X and Y are said to be homeomorphic if there exists a homeomorphism f from X to Y.

Examples:

(i) X = Y = \Re and f(x) = x^{3}-x for x \in X.

(ii) (a) X = \Re and Y = \{ x \in \Re -1 < x < 1 \}. X and Y are homeomorphic under the mapping f(x) = \frac{x}{1+|x|}

(ii) (b) \Re^{n} and the open unit ball B = \{ x \in \Re^{n}: ||x||<1\}

(iii) If X = \Re and Y = \{ -1 < x \leq 1\} then X and Y are not homeomorphic. Note that Y is compact while \Re is not compact.

(iv) If a mapping is one to one, onto and continuous, it does not follow that it is a homeomorphism. For example, let X = a set with more than one element and T is the discrete topology in X, and T^{'} is the trivial topology in X. Then the identity mapping I_{X}: (X, T) \rightarrow (X, T^{'}) is continuous, but not a homeomorphism. (PS: f^{-1} is not a bijection).

SEPARATION AXIOMS:

Two sets M_{1} and M_{2} can be separated by open sets if there exist open sets O_{1} and O_{2} such that M_{1} \subset O_{1} and M_{2} \subset O_{2} and O_{1} \bigcap O_{2} = \phi.

M_{1} and M_{2} can be separated by a real function if there exists a continuous real function f on X such that 0 \leq f(x) \leq 1 for x \in X and f(x)=0 for x \in M_{1} and f(x)=1 for x \in M_{2}.

T_{1} separation axiom:

For any two disjoint points, there exists a neighbourhood of either point not containing the other. (This implies that the complement of each point is an open set or that each point is a closed set).

T_{2} separation axiom:

Any two distinct points can be separated by open sets (or any two distinct points have disjoint neighbourhoods). A topological space is Hausdorff if it satisfies T_{2}).

Examples:

(i) A set X with the discrete topology (PS: If a topological space is T_{2}, then it is also T_{1} but the converse is not true in general).

(ii) If X is a set with more than one element, then X with the trivial topology is not Hausdorff !

(iii) If X is Hausdorff and A \subset X, then A is Hausdorff with the induced topology.

(iv) Let X_{1} be compact, X_{2} be Hausdorff and f: X_{1} \rightarrow X_{2} be continuous. Then f(X_{1}) is closed in X_{2}.

(v) Let X be Hausdorff with A \subset X and let A be compact. Then A is closed.

AXIOM T_{3}:

A closed set F and a point x \notin F can be separated by open sets.

A topological space is regular if it satisfies separation axioms T_{1} and T_{3}/ A completely regular topological space is one in which a closed set F and a point x \notin F can be separated by a real function.

AXIOM T_{4}:

Two disjoint closed sets can be separated by open sets. A topological space is normal if it satisfies seperation axioms T_{1} and T_{4}.

\bf {Urysohns \hspace{0.1in} \bf {Lemma}}

(Solomon Lefshetz, Algebraic Topology): In a normal space, every two disjoint closed sets can be separated by a real function.

Section 1.2 \bf {Topological \hspace{0.1in} \bf {Groups}}

A topological group is a set G which is both a group and a T_{1} space with the topology and group structure related by the assumption that the functions (x, y) \rightarrow x.y and x \rightarrow x^{-1} are continuous. Here, x, y \in G and x^{-1} is the group inverse element of x in G.

Equivalently, the function (x,y) \rightarrow xy^{-1} (from G \times G to G) is continuous.

Examples:

(i) The additive group of real numbers is the underlying group of a topological group whose underlying topological space is the usual space of real numbers. More generally, the Euclidean n-space under the usual topology.

(ii) Any group with the usual discrete topology (every set is its own closure).

(iii) The n-dimensional torus (that is to say, the product of n circles). Here a circle is the topological group \{ x \in \mathcal{C}: |x| =1\} under multiplication.

(iv) GL(n,C), the general linear group, that is, the group of all non-singular n \times n matrices with complex coefficients. For the topology use that “induced” by considering the n \times n matrices as a subset of \mathcal{C^{n}}.

(v) Any product of topological groups.

\bf{Trivial \hspace{0.1in} properties}

Let e be the identity in the topological group G. If E, F \subset G then EF  \equiv \{ xy : x \in E, y \in F\} (this is the definition, by the way)

(1) If z=xy, and O is a neighbourhood of x, there exist neighbourhoods P of x and Q of y such that PQ \subset O.

Let \phi denote the mapping (x,y) \rightarrow xy and \overline{O} = \phi^{-1}(O). Since \phi is continuous and O is open, it follows that \overline{O} is open. Now \overline{O} contains (x,y) so it also contains a set of the form O_{1} \times O_{2} where O_{1} is a neighbourhood of x, and O_{2} is a neighbourhood of y, and \phi(O_{1} \times O_{2}) = O_{1}.O_{2} \subset O. Take P = O_{1} and Q = O_{2}.

(1′) If a \in G, then for every neighbourhood V of a^{-1} there exists a neighbourhood U of a, such that U^{-1} \subset V. This follows again from the fact that a \rightarrow a^{-1} is continuous. (If g is a continuous mapping of a topological space R into R^{'}, then for every point a \in R and every neighbourhood U^{'} of a^{'} = g(a) \in R^{'} there exists a neighbourhood U of a such that g(U) \subset U^{'})

(2) For each x \in G, the mappings y \rightarrow yx and y \rightarrow xy are homeomorphisms or topological mappings.

The mapping f: y \rightarrow xy is one-to-one (because xy = x^{'}y implies that x = x^{'}). The inverse mapping f^{-1}: y \rightarrow x^{-1}y is of the same form. Hence, it suffices to prove that f is continuous. Let O be any neighbourhood of xy. By (1) there exists a neighbourhood O_{1} of x, and a neighbourhood O_{2} of y such that O_{1}O_{2} \subset O. Hence, xO_{2} \subset O, that is to say f(O_{2}) \subset O. This shows that f is continuous (cf. (1′)).

(3) The mapping x \rightarrow x^{-1} is a homeomorphism. For x \rightarrow x^{-1} is one-to-one, and is its own inverse (by group law). It is continuous by definition.

(4) If O is open in G, then O^{-1}, xO, EO, Ox, OE are also open, where x \in G and E \subset G.

By (2) xO and Ox are open so is O^{-1} by (3). Now EO is equal to \bigcup_{x \in E}xO, hence, open, similarly also OE.

(5) If V is any neighbourhood of e, it contains a neighbourhood W of e, such that W.W^{-1} \subset V.

Proof:

Since ee^{-1}=e, there exist neighbourhoods V_{1} of e, and V_{2} of e^{-1}=e, such that W.W^{-1} \subset V by (1).

Let V_{3} = V_{1} \bigcap V_{2}^{-1}. Then V_{3} is open and e \in V_{3} so that V_{3} is a neighbourhood of e.

Next V_{3}V_{3}^{-1} \subset V. For x \in V_{3} implies x \in V_{1} and x \in V_{2}^{-1}, and y^{-1} \in V_{3}^{-1} implies that y \in V_{3}, which in turn implies that y^{-1} \in V_{2}. Hence, xy^{-1} \in V_{2}V_{2}^{-1} implies that xy^{-1} \in V_{1}V_{2} \subset V. Choose W = V_{3}.

(6) A neighbourhood V of e is defined to be symmetric if and only if V=V^{-1}.

Every neighbourhood W of e contains a symmetric neighbourhood (for example, W \bigcap W^{-1}). By (4) W^{-1} is open. Hence, W \bigcap W^{-1} is a symmetric neighbourhood of e.

(7) Every neighbourhood of x \in G is of the form xV as well as of the form Wx, where V and W are neighbourhoods of the identity.

Let U be any neighbourhood of x. Then, V = x^{-1}U \ni e, and x^{-1}U is open by (4). Hence, V is a neighbourhood of e. Similarly, W = Ux^{-1} is a neighbourhood of e. Hence, U=xV=Wx.

(8) The continuity of (x,y) \rightarrow xy^{-1} is equivalent to the continuity of (x,y) \rightarrow xy together with the continuity of x \rightarrow x^{-1}.

(i) If (x,y) \rightarrow xy and x \rightarrow x^{-1} are continuous, then (x,y^{-1}) \rightarrow xy^{-1} and (x,y) \rightarrow (x,y^{-1}) are continuous, hence the composite (x,y) \rightarrow xy^{-1} is continuous.

(ii) Conversely, if (x,y) \rightarrow xy^{-1} is continuous, then (e,y) \rightarrow ey^{-1}=y^{-1} is continuous. Further, y \rightarrow (e,y) is continuous (obviously). Hence, y \rightarrow y^{-1} is continuous. Therefore, (x,y) \rightarrow (x,y^{-1}) is continuous. But by hypothesis, (x,y) \rightarrow xy^{-1} is continuous. Hence, (x,y) \rightarrow xy is continuous.

\bf{Separation \hspace{0.1in} properties}

\bf{Lemma \hspace{0.1in}1}

The topological space of a topological group G is Hausdorff.

\bf{Proof}

Let x,y belong to a group G such that x and y are distinct elements. By the T_{1} property (Given two points of S, each of them lies in an open set not containing the other) we can find a neighbourhood V of e not containing y^{-1}x. By (5) there exists a neighbourhood V_{1} of e, such that V_{1}V_{1}^{-1} \subset V. Then xV_{1}, yV_{1} are neighbourhoods of x and of y respectively. And xV_{1}, yV_{1} are neighbourhoods of x and of y respectively. And we claim that xV_{1} \subset yV_{1} = \phi. This is because of the following: (indirect proof) if xV_{1} \bigcap yV_{1} \neq \phi then there exist points (or elements of G or T) v^{'}, v^{''} such that xv^{'}=yv^{''} so that y^{-1}x contradicting the choice of V. Hence, the Hausdorff axiom is satisfied because xV_{1} \bigcap yV_{1} = \phi.

\bf{Lemma \hspace{0.1in}2}

If E \subset G, then \overline{E} = \bigcap {EV} =\bigcap{VE} where V extends over all neighbourhoods of e. (Note that E could be any, some or run through all subsets of G)

First we prove that (i) \overline{E} \supset \bigcap {EV}

If any arbitrary element x \in EV for all V. we shall see/prove that every neighbourhood of x intersects E and so we would have x \in \overline {E}.

Now, let x \in \bigcap EV and let O be any neighbourhood of x. Then by (7) above, O \in xV, where V is a neighbourhood of e. By hypothesis, x \in EV^{-1}, which implies that x = ay^{-1}, where a \in E, y \in V, or xy=a. Hence xV intersects E (that is, has a non-empty intersection with E). Therefore, O intersects E, hence, x \in \overline{E}.

(ii) \overline{E} \subset \bigcap{EV}.

If x \in \overline{E}, then every neighbourhood of x intersects E. By (7), xV^{-1} is a neighbourhood of x (where V is a neighbourhood of e). Hence, xV^{-1} intersects E. This implies that x \in EV (for there exists y \in V, such that xy^{-1} \in E, or x \in EY \subset EV. Hence, x \in \bigcap EV

\bf{Remark} A topological group is homogeneous. Given any two elements —- p. q \in G, there exists a topological mapping f of G onto itself which takes p into q.

Take a= p^{-1}q and take f(x)-xa.

It is sufficient for many purposes therefore to verify local properties for a single element only. For example, to show that G is locally compact, it is suffieient to show that its identity e has a neighbourhood U whose closure is compact — so also with regularity.

\bf{Lemma \hspace{0.1in}}

The topological space of a topological group G is regular (Kolmogorov).

Note: A topological space is T_{1} if the following is true: Given two points of a topological space S, each of them is contained in an open set not containing the other. Axiom T_{3} : If C is a closed set in the space S, and if p is a point not contained in C, then there are disjoint open sets in S, one containing C and the other containing p.3

A space that satisfies both T_{1} and T_{3} is called regular.

\bf{Proof}

We can separate e and any closed set F not containing not containing e. Now let O = F^{c}. Then O is a neighbourhood of e. Now there exists a neighbourhood V of e, such that V^{2} \subset O (because of (5) and (6) above). (Remark I think the author has assumed here that V is symmetric)V and \overline{F^{c}} are disjoint open sets. We shall see that F \subset \overline{V^{c}}.

Since V is a neighbourhood of e, that will prove this lemma. Now,

\overline{V} = \bigcap{VW}, where W is a neighbourhood of e (by Lemma 2)

\overline{V} \subset V^{2} (since V is a W, a neighbourhood of e_

\subset \overline{V} \subset O (by the choice of V and O).

(So \overline{V} = F^{c} (by definition of O)

Hence, V \subset F^{c} or \overline{V^{c}} or \overline{V^{c}} \supset F.

\bf{Theorem \hspace{0.1in}} Atopological space which is the underlying space of a topological group G is completely regular.

\bf{Proof} It is sufficient to show that if F is a closed set not containing e, then F and e can be separated by a continuous real function. (cf. the remark above)

Note: Definition: A topological spacce is said to be \bf{completely} \hspace{0.1in} \bf{regular} (also called

\bf{Tychonoff} \hspace{0.1in}\bf{space}) if for every point p of S and for any open set U containing p, there is a continuous function of S to I^{2} such that f(p)=0 and f(x)=1 for all points x in S-U. I is a closed interval [a,b].

Let V=F^{c}. Then V is a neighbourhood of e. Choose a sequence V_{1}, V_{2}, \ldots of neighbourhoods of e, such that V_{1}^{2}\subset V, V_{k+1}^{2}\subset V_{k} where k \geq 1. [This is possible, see the proof of Lemma 3 in which (5) and (6) were used.

Let \alpha be finite dyadic real number (a dyadic rational or binary rational number is a number that can be expressed as a fraction with denominator whose fraction is a power of two)

\alpha = 0.\alpha_{1}\alpha_{2}\ldots 000 where \alpha_{i}=0,1

Define O_{\alpha}=V_{1}^{\alpha_{1}} V_{2}^{\alpha_{2}}\ldots V_{k}^{\alpha_{k}} (group product ) where V_{i}^{0} is e.

We will show that \alpha < \beta implies that \overline{O}_{\alpha}\subset O_{\beta}

If \alpha < \beta then first j digits agree for some j \geq 0 so that

\alpha = 0.\alpha_{1}\alpha_{2}\ldots \alpha_{j}0\alpha_{j+2}\ldots \alpha_{k}000

\beta = 0.\beta_{1}\beta_{2}\ldots \beta_{j}1\beta_{j+2}\ldots \beta{m}000

and \alpha_{l} = \beta_{l} for some l \leq j

Now define

\alpha^{'}=0.\alpha_{1}\ldots 0 111 \ldots 1000 note that here \alpha_{k} is the right most 1.

\beta^{'} = 0.\alpha_{1}\ldots \alpha_{j}1000 \ldots

Then

\alpha \leq \alpha^{'}< \beta^{'} \leq \beta

Clearly, O_{\alpha} \subset O_{\alpha^{'}} \subset O_{\beta^{'}} \leq \O_{\beta} (since V_{i}^{0}=e

We shall see that

{\overline{O}}_{\alpha^{'}} \subset O_{\beta^{'}} (which will imply that {\overline{O}}_{\alpha} \subset O_{\beta}

Let O = V_{1}^{\alpha_{1}}\ldots V_{j}^{\alpha_{j}}

Then

O_{\alpha^{'}} = OV_{j+2}\ldots V_{k} and O_{\beta^{'}} = OV_{j+1}

Hence by Lemma 2,

{\overline{O}}_{\alpha^{'}}V_{k} =  OV_{j+2}\ldots V_{k-1}V_{k}^{2}

\subset OV_{j+2}\ldots V_{k-2}V_{k-1}^{2} (since V_{k}^{2} \subset V_{k-1})\subset OV_{j+2}^{2} \subset OV_{j+1} = O_{\beta^{'}} by definition

Hence, {\overline{O}}_{\alpha^{'}} \subset O_{\beta^{'}} and therefore (\overline{O})_{\alpha} \subset O_{\beta}

Now detine

f(x) = 1 if x \notin O_{\alpha}

f(x) =  \inf \{\alpha : x \in O_{\alpha}\} if x is in some O_{\alpha}

Then we have 0 \leq f(x) \leq 1 for all x \in G. Furtheer

f(x) = 0 for x = e (since Slatex V_{i}^{0}=e$)

f(x) - 1 for x \in F (for x \in F) (as `V = F^{c} (see above)

We shall see that f is continuous. The sets \{ x: f(x)>k\} and \{ x : f(x) <k\}, (where k is a real number) are both open. For,

\{ x : f(x) >k\} This is equal to the following: (\bigcap_{\alpha>k}(\overline{O})_{\alpha})^{c}

\Longleftrightarrow \{ x: f(x) \leq k\} So we get the following now:

\{ x: f(x) >k\} = (\bigcap_{\alpha>k}{\overline{O}}_{\alpha}) (this is a closed set)

Note that \alpha < \beta \Longrightarrow O_{\alpha} \subset O_{\beta} hence

\{ x: f(x)\leq k\} \subset \{ x: x \in \bigcap_{\alpha >k} O_{\alpha}\}

for if x is such that f(x) \leq k, and \beta>k such that x \notin O_{\beta}, then x \notin O_{\alpha} for all \alpha < \beta (since O_{\alpha} \subset O_{\beta}), hence f(x) \geq \beta > k, a contradiction. The opposite inclusion is trivial. Thus,

\{ x: f(x) \leq k\} = \{ x: x \in \bigcap_{\alpha >k}O_{\alpha}\}

But \bigcap_{\alpha >k}{\overline{O}}_{\alpha} since on the one hand, O_{\alpha} \subset{\overline{O}}_{\alpha} and on the other \alpha < \beta \Longrightarrow {\overline{O}}_{\alpha} \subset O_{\beta} so that x \in \bigcap_{\alpha>k}{\overline{O}}_{\alpha} implies that x \in \bigcap_{\alpha>k}O_{\alpha} (as \alpha is a dyadic rational), hence \{ x: f(x)>k\} is open.

Similarly, \{ x: f(x)<k\} is also open. For if the set contains the point x, it contains a whole neighbourhood of it; for, if x is such that f(x)<k, then there exists \alpha < k, such that x \in O_{\alpha}, which is neighbourhood of x all of which is contained in \{ x : f(x)<k\} note (y \in {O_{\alpha}} \Longrightarrow f(y) \leq \alpha <k)

Similarly, \{ x: f(x) < k\} is also open. For if the set contains the point x, it contains a whole neighbourhood of it; for, if x is such that f(x)<k, then there exists \alpha <k such that x \in O_{\alpha} which is a neighbourhood of x all of which is contained in \{ x: f(x)<k\}, y \in O_{\alpha} \Longrightarrow f(y) \leq \alpha <k.

\bf{Lemma \hspace{0.1in}4}

If C_{1}, C_{2} are compact subsets of G, then C_{1}C_{2} is compact

Proof: Consider the mapping G \times G \rightarrow G which takes (x,y) into xy. This is continuous. The product C_{1} \times C_{2} is compact (Tychonoff). The proof follows from the fact that the continuous image of a compact set is compact.

\bf{Remark}

If F_{1}, F_{2} are closed, it does not follow that F_{1}.F_{2} is closed. In \Re^{1} let F_{1}= \{ n \in \mathcal{Z}: n \geq 1\} and F_{2} = \{ 0\} \bigcup \{ \frac{1}{n}: n \in F_{1}\}. Then F_{1}.F_{2}=Q \subset \Re^{1}. Here by Q we mean the set of rational numbers greater than or equal to zero.

\bf{Theorem \hspace{0.1in} 2}

A locally compact group is normal.

Note: we recall here the definition of normal: Axiom T_{1}: Given two points of a topological space S, each of them lies in an open set not containing the other. Axiom T_{4}: If H and K are disjoint closed sets in the topological space S, then there exist disjoint open sets, one containing H and the other containing K. A T_{4} space or a normal space is one that satisfies both Axiom T_{1} and Axiom T_{4}.

If the group is compact, the proof is easy since compactness together with regularity (or Hausdorff ) implies normality.

Otherwise take a symmetric neighbourhood U of e with compact closure, and consider G^{'} = \bigcup_{n=1}^{\infty} U^{n}. Then G^{'} is an open and closed subgroup of G, and it is sufficient to prove normality for G^{'}. To do that, use the fact that G^{'} is \sigma compact. That is, G^{'} = \bigcup_{n=1}^{\infty}K_{n} where K_{n} is compact. K_{n}=\bigcup_{l=1}^{n}{\overline{U}}^{l}.

Note: A locally compact T_{1} group is para-compact, hence normal. (Reference: Hewitt and Ross, Abstract Harmonic Analysis. Volume I. page 76, Theorem 8.13

\bf{1.3} \bf{Subgroups} \bf{Quotient} \bf{groups}

Let G be a topological group, and H a subset of G. Then H is, by definition, a subgroup of the topological group G if and only if H is a subgroup of the abstract group G, and H is a closed set in the topological space G.

Let G be a topological group, and H a subset of it which is a subgroup of G considered as an abstract group. Then H is also a topological subgroup with the induced topology. In particular, a subgroup of an abstract group which is a topological group is itself a topological group.

A subgroup N of the topological group G is defined to be a normal subgroup if N is the normal subgroup of the abstract group G.

Let G be a topological group, and let H be a subgroup of the abstract group G. Then \overline{H} is a subgroup of the topological group G. If H is a normal subgroup of the abstract group G, then \overline{H} is also normal, that is, a normal subgroup of the topological group G.

Let us recall that if G is any group, and H a subgroup of G, a left coset of H is a subset of G of the form xH, where x is an element of G. The left coset set is the set of all left cosets of H, denoted by G/H. We have a natural map or projection \pi: G \rightarrow G/H where x \mapsto xH defined by \pi (x) equal to the left coset of H which contains x.

If G is a topological group, we shall topologize G/H, assuming that H is closed. A set O \subset G/H is open if and only if \pi^{-1}(O) is open in G. This means that we require \pi to be a continuous map.

\bf{Lemma \hspace{0.1in}5} G/H is a T_{1} space, and \pi is open.

Again we first recall definitions of T_{1} and open: Axiom T_{1} says: Given two points of a topological space S, each of them lies in an open set not containing the other. Also, a mapping f is said to be open if for every open set O_{1} in X_{1} the set $latex $f(O_{1}) is an open set in X_{2}.. Note: f: X_{1} \rightarrow X_{2}.

\bf{Proof}

Since H is closed, xH is also closed. (homeomorphism) so that (xH)^{c} is open in G. Write \overline{x}=xH, coset containing x. Now \pi^{-1}(G/H - \overline{x}) =(xH)^{c}, which is open. Therefore, G/H is T_{1} (the complement of each point is an open set). We know that \pi is continuous, we have to show that it is also open. Let O be open, where O \subset G. Then \pi O is in the coset space; it is open if and only if \pi^{-1}(\pi O) is open. But \pi^{-1}(\pi O) is OH, which is open since OH = \bigcup_{x \in H}Ox, where Ox is open. Thus, O open \Longrightarrow \pi O is also open. QED.

\bf{Lemma \hspace{0.1in}6} G/H is a T_{2}

Let us recall definition of T_{2} or Hausdorff space: Given two distinct points of a topological space S, there are disjoint open sets each containing just one of the two points.

\bf{Proof}

Let x_{1} and y_{1} be two distinct points of G/H = Q and x,y \in G such that \pi(x)=x_{1} and \pi(y)=y_{1}. Choose a neighbourhood v of e such that Vx \bigcap yH = \phi. This is possible because x \notin yH = \overline{yH} (H is closed as yH is closed. Vx is a neighbourhood of x, use the definition of \overline{yH}.) It follows that VxH \bigcap yH=\phi. For if VxH \bigcap yh \neq \phi, then vxh_{1}=yh_{2}, say,, where v \in V, h_{1}, h_{2} \in H. Hence vx = yh_{2}h_{1}^{-1}=yh_{3}, which contradicts Vx \bigcap yH = \phi.

Let V_{1} be a neighbourhood of e such that V_{1}^{-1}V_{1} \subset V. Then V_{1}^{-1}V_{1}xH \bigcap yH = \phi, hence V_{1}xH \bigcap V_{1}yH= \phi. Therefore,

\pi(V_{1}x) \pi(V_{1}y) = \phi

Now x_{1} \in \pi (V_{1}x) since e \in V_{1} and y_{1} \in \pi (V_{1}y) and \pi (V_{1}x), \pi (V_{1}y) are open (by Lemma 1, V_{1}x and V_{1}y are open, and by Lemma 5 \pi is open mapping. Hence, x_{1}, y_{1} are separated by disjoint open sets. QED.

\bf{Remark}

If G is any topological group, H a closed subgroup, we have topologized the coset set G/H in such a way that it is a T_{2} space. We may call G/H the quotient space, and the given topology the quotient space topology. On the other hand, if G is any group, and H a normal subgroup (that is, \forall {x} \in H, \forall a \in G, axa^{-1} \in H or aHa^{-1} \subset H), then the coset G/H is, in fact, a group known as the quotient group. The next lemma shows that the quotient group, with the quotient space topology is a topological group, if, to start with, G is a topological group.

\bf{Lemma \hspace{0.1in}7} If G is a topological group, and H a normal subgroup, then the quotient group G/H with the quotient space topology is a topological group.

\bf{Proof}

We have only to show that the mapping \psi_{1}: G/H \times G/H \rightarrow G/H given by \psi_{1}(x_{1}, y_{1}=x_{1}y_{1}^{-1}) where x_{1}, y_{1} \in G/H is continuous.

Let \psi: G \times G \rightarrow G be given by (x,y) \mapsto xy^{-1} where x, y are elements of G. We then have the following “loop”:

\pi \times \pi : G \times G \rightarrow G/H \times G/H

\psi_{1}: G/H \times G/H \rightarrow G/H

\psi : G \times G \rightarrow G

\pi  G \rightarrow G/H

Trivially, we have \pi \psi = \psi_{1}(\psi \times \psi)

Since \pi and \psi are continuous, \pi\psi is also continuous. Hence, \psi_{1}(\pi \times \pi) is continuous.

Let O_{1} be an open set in G/H. Then, [\psi_{1}(\pi \times \pi)]^{-1}O_{1} is open in G \times G. But \pi \times \pi is open. Hence,

(\pi \times \pi)(\pi \times \pi)^{-1}\psi_{1}^{-1}(O_{1}) is open in G/H \times G/H; that is to say, \psi_{1}^{-1}(O_{1}) is open, hence \psi_{1} is continuous. QED.

\bf{Lemma \hspace{0.1in}8}

Let G be a topological group, and H a subgroup.

(a) If G is compact, then H and G/H are both compact.

(b) If G is locally compact, then both H and G/H are locally compact.

\bf{Proof}

(a) H is a closed subset of a compact set, hence, compact, G/H is the continuous image of a compact set, hence, compact.

(b) H is locally compact since it is a closed subset of a locally compact space. [Let S be a locally compact space, T \subset S, T = \overline{T}. Then T is locally compact. For, p \in T \Longrightarrow p \in S \Longrightarrow \exists U_{p} \subset S so that p \in U_{p} and (\overline{U})_{p} is compact. Now, T \bigcap (\overline{U})_{p} is a compact neighbourhood of p.]

To prove that G/H is locally compact, let q, let q \in G/H, and U_{1} a neighbourhood of q, that is, an open set containing q.) Let

U  = \pi^{-1}(U_{1})

and let x \in U. so that \pi(x)=q. Since G is locally compact, and U is open, there exists a neighbourhood O of x, such that \overline{O} \subset U, with \overline{O} compact. [Let G be any topological group. To every neighbourhood U of e, there exists a neighbourhood V of e, such that \overline{V} \subset U. For let V be a symmetric neighbourhood of e, such that V^{2} \subset U]. (See (5) and (6) above). Now x \in \overline{V} \Longrightarrow (xV) \bigcap V \neq \phi. Hence, xv_{1}=v_{2}, where v_{1}, v_{2} \in V. Therefore x=v_{2}v_{1}^{-1} \in V.V^{-1}\subset V^{2}\subset U. Hence, \overline{V} \subset U. Then we have

\pi(\overline{O}) \subset \pi(U) = U_{1}.

Since O_{1} = \pi{O} a neighbourhood of q(O is a neighbourhood of x). And, \pi(\overline{O}) is compact. (since \overline{O}). Since G/H is a T_{2} space, \pi(\overline{O}) is closed.

We have O_{1} \subset \pi(\overline{O})…..see above O_{1} = \pi(O) which implies that (\overline{O})_{1} \subset \overline{\pi{O}} = \pi(\overline(O)) (compact).

Hence, O_{1} is compact (closed subset of a compact set of a compact set). Thus, {O}_{1} is a neighbourhood of q with compact closure. It follows that G/H is locally compact.

\bf{I.4 \hspace{0.1in} Examples}

Let \mathcal{C}^{n} denote the n-dimensional complex cartesian space. It is a vector space of dimension n over the field \mathcal{C} of complex numbers.

Let (\overline{e})_{i}=\{ \}0,0,0, \ldots ,1,0,0,0\} where 1 is in the ith position. Then, (\overline{e})_{1}, (\overline{e})_{2}, \ldots, (\overline{e})_{n} form a basis of \mathcal{C}^{n} over \mathcal{C}.

An endomorphism \alpha of \mathcal{C}^{n} is defined when the elements \alpha \overline{e}_{i} is equal to \Sigma_{j=1}^{n}a_{ji}(\overline{e})_{j} are given. To \alpha corresponds the matrix \{ a_{ij}\} of degree n, and conversely.

We use the same letter \alpha for the matrix as well as the endomorphism.

We define a multiplication \alpha \circ \beta of two endos, \alpha, \beta with matrics \{ a_{ij}\}, \{ b_{ij} tr \} respectively by defining the corresponding matrix \{ e_{ij}\} as the product of the matrices, namely

e_{ij} = \Sigma_{k=1}^{n}a_{ik}b_{kj}

Now let M_{n}(\mathcal{C}) denote the set of all matrices of degree n with coefficients in \mathcal{C}. If \{a_{ij} \} \in M_{n}(\mathcal{C}), put b_{i+(j-1)n}=a_{ij} (as j goes from 1 to n, j-1 goes from 0 to n-1, and (j-1)n from zero to n^{2}-n in steps of n, while i goes from 1 to n; so that i+(j-1)n goes from one to n^{2})

To (a_{ij}) we associate the point with the coordinates b_{1}, b_{2}, \ldots, b_{n^{2}} in \mathcal{C}^{n^{2}}. In this way, we get a one to one correspondence between M_{n}(\mathcal{C}) and \mathcal{C}^{n^{2}}. Since \mathcal{C}^{n^{2}} is a topological space, we define a topology in M_{n}(\mathcal{C}) by requiring the correspondence to be a homeomorphism.

Let T be any topological space, and let \phi map T into M_{n}(\mathcal{C}); \phi: T \rightarrow M_{n}(\mathcal{C}). If t \in T, \phi(t) is a matrix with coefficients a_{ij}, say. Clearly, \phi is continuous if and only if each function a_{ij}(t) is continuous. Note that:

(\underbrace{T \stackrel{\phi}\rightarrow M_{n}{(C)}\stackrel{\psi}\rightarrow \mathcal{C}^{n} \stackrel{\pi_{ij}}\rightarrow \mathcal{C}}_{\phi_{ij}}), where \phi_{ij}(t)=a_{ij}(t) (In general, the following situation holds: If T \stackrel{I}\rightarrow \prod_{a \in A}X_{a} \stackrel{\pi_{a}}\rightarrow X_{a}, then ” f is continuous iff \pi_{a} \circ f is continuous for *every* a \in A^{n}. On the one hand, it is trivial that : f continuous implies \pi_{a} \circ f is continuous”. On the other, if \pi_{a} \circ f is continuous and *open*, with U \subset X_{a}, then (\pi_{a} \circ f)^{-1}U is open, that is, f^{-1}(\pi_{a}^{-1}(U)) is open. But \pi_{a}^{-1}(U) is open, since sets of the form \pi_{a}^{-1}(U), U open, form a sub-basis of open sets of \prod_{a \in A}X_{a}. So f is continuous).

It follows from this remark, and (4.1) that the product \sigma\tau of two matrices \sigma and \tau is continuous funcction of the pair (\sigma, \tau) considered as a point of the space M_{n}(\mathcal{C}) \times M_{n}(\mathcal{C}).

\bf{Notation} We denote by ^{t}a the transpose of the matrix \alpha = (a_{ij}); ^{t}a = (a^{'}_{ij})=a_{ji}. We denote by \overline{\alpha} the complex conjugate of \alpha; \overline{\alpha}= (\overline{a}_{ij}).

Clearly, \alpha \mapsto ^{t}\alpha and \alpha \mapsto \overline{\alpha} are homeomorphisms, of order 2, of M_{n}(\mathcal{C}) onto itself.

If \alpha and \beta are any two matrices, then ^{t}(\alpha\beta) = ^{t}\alpha^{t}\beta and \overline{\alpha\beta} = \overline{\alpha}\cdot \overline{\beta}

An n \times n matrix \sigma is regular (or non-singular), if it has an inverse, that is, if there exists a matrix \sigma^{-1}, such that \sigma\sigma^{-1}=\sigma^{-1}\sigma = \varepsilon, where \varepsilon is the unit matrix of degree n.

A necessary and sufficient condition for \sigma to be regular is that its determinant is non zero.

If an endomorphism \sigma of \mathcal{C}^{n} maps \mathcal{C}^{n} onto itself (and not some subspace of lower dimension) the corresponding matrix \sigma is regular, and \sigma has a reciprocal endomorphism \sigma^{-1}.

If \sigma is a regular matrix, we have

^{t}(\sigma^{-1}) = (^{t}\sigma)^{-1} and (\overline{\sigma})^{-1}=\overline{(\sigma^{-1})}

If \sigma and \tau are regular matrices, \sigma\tau is also regular, and we have

(\sigma\tau)^{-1}=(\tau)^{-1}\sigma^{-1}.

Hence, the regular matrices od degree n form a group with respect to multiplication, which is called the general linear group GL(n, \mathcal{C}).

Since the determinant of a matrix is obviously a continuous function (being a polynomial) of the matrix, GL_{n, \mathcal{C}} is an open subset of M_{n}(\mathcal{C}) (GL(n, \mathcal{C}) = \{ \sigma: det \sigma \neq 0\}), det is a continuous function. ) The elements of GL{(n, \mathcal{C})} may be considered as points of a topological space which is a subspace of the topological space M_{n}(\mathcal{C}).

If \sigma = (a_{ij}) is a regular matrix, the coefficients b_{ij} of \sigma^{-1} are given by b_{ij}=A_{ij}(\det \sigma)^{-1}, where the A_{ij} are polynomials in the coefficients f \sigma. Hence, the mapping of GL(n, \mathcal{C}) onto itself is continuous. Since the mapping coincides with its reciprocal mapping, it is a homeomorphism of GL(n, \mathcal{C}) with itself.

The mappings \sigma \mapsto \overline{\sigma} and \sigma \mapsto ^{t}\sigma are also homeomorphisms of GL(n, \mathcal{C}) with itself. The first is an automorphism of the group, not the second (preserves sums and inverts the order of the products)

If \sigma \in GL(n, \mathcal{C}) define \sigma^{*} = (^{t}\sigma)^{-1}

Then, we have (\sigma\tau)^{*}= \sigma^{*}\tau^{*} and \sigma_{}

Thus, we have: (\sigma\tau)^{*}=(\sigma\tau)^{*}=\sigma^{*}\tau^{*} and (\sigma^{*})^{-1}. Hence, \sigma \mapsto \sigma^{*} is a homeomorphism, and an automorphism of order 2 of GL(n, \mathcal{C}).

\bf{The \hspace{0.1in} subgroups} namely, O(n), O(n, \mathcal{C}, U(n) \hspace{0.1in} of GL(n, \mathcal{C}))

Let \sigma \in GL(n, \mathcal{C}). We say that \sigma is orthogomal if \sigma = \overline{\sigma}=\sigma^{*}. The set of all orthogonal matrices of degree n we denote by O(n). If only \sigma= \sigma^{*}, then \sigma is called complex orthogonal, and the set of all such \sigma we denote by O(n, \mathcal{C}). If \overline{\sigma}=\sigma^{*}, then \sigma is called unitary, and we denote by U(n) the set of all such \sigma.

Since \sigma \mapsto \overline{\sigma} and \sigma \mapsto \sigma^{*} are continuous, the sets O(n), O(n, \mathcal{C}, U(n)) are closed subsets of GL(n, \mathcal{C}). [Note that \{ x: f(x)=\sigma\} is closed if f is real and continuous. \phi_{ij}(\sigma)=(\overline{a})_{ij}-a_{ij}^{*} is continuouz for all i,j and \{ \sigma: \phi_{ij}(\sigma)=0\} is closed]. Because these mappings are automorphisms, O(n), O(n, \mathcal{C}, U(n)) are subgroups of GL(n,\mathcal{C}).

[Note, in parenthesis, that if X is any topological space, and Y a Hausdorff space, and f,g are continuous mappings of X into Y, then the set E=\{ x: x \in X, f(x)=g(x0\} is closed. One can see then the set E – \{ x : x \in X, f(x) \neq g(x)\} is open in X. Let x_{0}\in F. Since F = \{ x : x \in X, f(x) \neq g(x)\} is open in X and …f(x_{0}\neq g(x_{0}), and Y is Hausdorff, there exist neighbourhoods U_{1}, U_{2} of f(x_{0}, g(x_{0})) respectively so that U_{1}\bigcap U_{2}=\phi). Since f and g are are continuous, there exist neighbourhoods V_{1}, V_{2} of x_{0} in X such that f(V_{1}) \subset U_{1}, g(V_{2} \subset U_{2}). Let V = V_{1} \bigcap V_{2}. Then V is a neighbourhood of x_{0}, and f(x) \neq g(x) for x \in V, hence V \subset F. It follows that F is open.]

Clearly, O(n) = O(n, \mathcal{C}) \bigcap U(n)….call this (i)

\sigma is real, if its coefficients are real, that is, if \sigma = \overline{\sigma}. The set of all real matrices of degree n we denote by M_{n}(\Re), and we define GL(n, \Re) = M_{n}(\Re) \bigcap GL (n, \mathcal{C}). Hence, O(n) = GL(n, \Re) \bigcap O(n, \mathcal{C}).

Since the determinant of the product of two matrices is the product of their determinants, the matrices of determinant I form a subgroup of GL(n, \mathcal{C}). The group of all matrices with determinant I in GL(n, \mathcal{C}) is called the special linear group SL(n, \mathcal{C}).

We set

SL(n, \Re) = SL(n,\mathcal{C}) \bigcap GL(n, \Re)

SU(\overline{n}) = SL(\overline{n}, \mathcal{C}) \bigcap U(n) ….call this (ii)

O(n) = SL(n, \mathcal{C}) \bigcap O(n)….call this (iii)

Clearly, SL(n, \mathcal{C}), SL(n, \Re), SU(n), SO(n) are subgroups and closed subsets of GL(n, \mathcal{C}). They may be considered as subspaces of GL(n, \mathcal{C}).

\bf{Theorem 3} U(n), O(n), SU(n), SO(n) are compact.

\bf{Proof} We have only to show that U(n) is compact, since O(n), SU(n), SO(n) are closed subsets of U(n). We shall see that U(n) is homeomorphis to a bounded, closed subset of \mathcal{C}^{n^{2}}.

A matrix \sigma is unitary if and only if ^{t}\sigma\overline{\sigma}=\varepsilon, where \varepsilon is the unit matrix. [\overline{\sigma}= \sigma^{*}=(^{t}\sigma)^{-1}]

If \sigma = \{ a_{ij}\}. then

(^{t}\sigma)\overline{\sigma}=e \Longleftrightarrow \Sigma_{j}(a_{ji}).(\overline{a_{jk}})=\delta_{ik}

[\sigma is regular, that is, \sigma\sigma^{-1}=\varepsilon]

The left hand side of the last equations are continuous functions of \sigma. U(n) is not only a closed subset of GL(n, \mathcal{C}) but also of M_{n}(\mathcal{C}). [For \{ \sigma: \Sigma a_{ji}\overline{a}_{jk}=0, i \neq k\} \bigcap \{ \sigma: \Sigma a_{jk}\cdot \overline{a}_{ji} = I\} is an intersection of closed sets].

Further,

\Sigma_{j}a_{ji}\overline{a}_{ji}=1 \Longleftarrow  |a_{ij}| \leq 1 for 1 \leq i,j \leq n.

Therefore the coefficients of the matrix \sigma \in U(n) are bounded. Since f: M_{n}(C) \mapsto \mathcal{C}^{n^{2}} is a homeomorphism, f(U(n)) is closed and bounded, and a subset of \mathcal{C}^{n^{2}} and hence compact.

Cheere,

Nalin Pithwa