This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

\lastname

Hofmann and Kramer \msc22E15, 22E65, 22E99 \annum2022 \editorKarl-Hermann Neeb \finalJanuary 17, 2022

On Weakly Complete Universal Enveloping Algebras:
A Poincaré–Birkhoff–Witt Theorem

Karl Heinrich Hofmann and Linus Kramer Both authors were supported by Mathematisches Forschungsinstitut Oberwolfach in the program RiP (Research in Pairs). Linus Kramer is funded by the Deutsche Forschungsgemeinschaft under Germany’s Excellence Strategy EXC 2044-390685587, Mathematics Münster: Dynamics-Geometry-Structure. Karl Heinrich Hofmann
Fachbereich Mathematik
Technische Universität Darmstadt
Schlossgartenstraße 7
64289 Darmstadt, Germany
[email protected]
Linus Kramer
Mathematisches Institut
Universität Münster
Einsteinstraße 62
48149 Münster, Germany
[email protected]
(December 3, 2021)
Abstract

The Poincaré-Birkhoff-Witt Theorem deals with the structure and universal property of the universal enveloping algebra U(L)U(L) of a Lie algebra LL, e.g., over \mathbb{R} or \mathbb{C}. In 1\langle 1\rangle“K.H. Hofmann and L. Kramer, On Weakly Complete Group Algebras of Compact Groups, J. Lie Theory 30 (2020), 407–426”, the weakly complete universal enveloping algebra 𝐔(𝔤)\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}) of a profinite dimensional topological Lie algebra 𝔤\mathfrak{g} was introduced. Here it is shown that the classical universal enveloping algebra U(|𝔤|)U(|\mathfrak{g}|) of the abstract Lie algebra underlying 𝔤\mathfrak{g} is a dense subalgebra of 𝐔(𝔤)\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}), algebraically generated by 𝔤𝐔(𝔤)\mathfrak{g}\subseteq\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}). It is further shown that, inspite of 𝐔\mathop{\bf U\hphantom{}}\nolimits being a left adjoint functor, it nevertheless preserves projective limits in the form 𝐔(lim𝔦𝔤/𝔦)lim𝔦𝐔(𝔤/𝔦)\mathop{\bf U\hphantom{}}\nolimits(\lim_{\mathfrak{i}}\mathfrak{g}/\mathfrak{i})\cong\lim_{\mathfrak{i}}\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}/\mathfrak{i}), for profinite-dimensional Lie algebras 𝔤\mathfrak{g} represented as projective limits of their finite-dimensional quotients. The required theory is presented in an appendix which is of independent interest.—In a natural way, a weakly complete enveloping algebra 𝐔(𝔤)\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}) is a weakly complete symmetric Hopf algebra with a Lie subalgebra (𝐔(𝔤))\mathbb{P}(\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})) of primitive elements containing 𝔤\mathfrak{g} (indeed properly if 𝔤{0}\mathfrak{g}\neq\{0\}), and with a nontrivial multiplicative pro-Lie group 𝔾(𝐔(𝔤))\mathbb{G}(\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})) of grouplike units, having (𝐔(𝔤))\mathbb{P}(\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})) as its Lie algebra–in contrast with the classical Poincaré-Birhoff-Witt environment of U(L)U(L), thus providing a new aspect of Lie’s Third Fundamental Theorem: Indeed a canonical pro-Lie subgroup Γ(𝔤)\Gamma^{*}(\mathfrak{g}) of 𝔾(𝐔(𝔤))\mathbb{G}(\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})) is identified whose Lie algebra is naturally isomorphic to 𝔤\mathfrak{g}. The structure of 𝐔(𝔤)\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}) is described in detail for dim𝔤=1\dim\mathfrak{g}=1. The primitive and grouplike components and their mutual relationship are evaluated precisely.—In 1\langle 1\rangle and 2\langle 2\rangle“R. Dahmen and K.H. Hofmann, The Pro-Lie Group Aspect of Weakly Complete Algebras()(\dots), J. of Lie Theory 29 (2019), 413–455”, the real weakly complete group Hopf algebra [G]\mathbb{R}[G] of a compact group GG was described. In particular, the set ([G]))\mathbb{P}(\mathbb{R}[G])) of primitive elements of [G]\mathbb{R}[G] was identified as the Lie algebra 𝔤\mathfrak{g} of GG. It is now shown that for any compact group GG with Lie algebra 𝔤\mathfrak{g} there is a natural morphism of weakly complete symmetric Hopf algebras ω𝔤:𝐔(𝔤)[G]\omega_{\mathfrak{g}}\colon\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})\to\mathbb{R}[G], implementing the identity on 𝔤\mathfrak{g} and inducing a morphism of pro-Lie groups Γ(G)𝔾([G])G\Gamma^{*}(G)\to\mathbb{G}(\mathbb{R}[G])\cong G: yet another aspect of Sophus Lie’s Third Fundamental Theorem!

keywords:
Associative algebra, Lie algebra, universal enveloping algebra, weakly complete vector space, projective limit, pro-Lie group, profinite-dimensional Lie algebra, power series algebra, symmetric Hopf algebra, primitive element, grouplike element, Poincaré-Birkhoff-Witt theorem
volume: 32

1 The Weakly Complete Enveloping Algebra
of a Profinite-Dimensional Lie Algebra

In [7] we have initiated the theory of weakly complete universal enveloping algebras over 𝕂\mathbb{K} hoping that in some fashion this concept would resemble the classical universal enveloping algebra of a Lie algebra such as it is presented in the famous Poincaré-Birkhoff-Witt-Theorem (see e.g. [1], Chap. 1, Paragraph 2, no 7, Théorème 1., p.30). While this was not exactly the case, we shall discuss now how close we come to that theorem.

So we let 𝕂\mathbb{K} denote one of the topological fields \mathbb{R} or \mathbb{C}. For a topological Lie algebra 𝔤\mathfrak{g} over 𝕂\mathbb{K} we let (𝔤){\cal I}(\mathfrak{g}) denote the filter basis of all closed ideals 𝔦𝔤\mathfrak{i}\subseteq\mathfrak{g} such that dim𝔤/𝔦<\dim\mathfrak{g}/\mathfrak{i}<\infty.

{Definition}

A topological Lie algebra 𝔤\mathfrak{g} over 𝕂\mathbb{K} is called profinite-dimensional if 𝔤=lim𝔦(𝔤)𝔤/𝔦\mathfrak{g}=\lim_{\mathfrak{i}\in{\cal I}(\mathfrak{g})}\mathfrak{g}/\mathfrak{i}. Let 𝒲{\cal W}\hskip-1.0pt{\cal L} denote the category of profinite-dimensional Lie algebras (over 𝕂\mathbb{K}) and continuous Lie algebra morphisms between them.

Notice that by its definition every profinite-dimensional Lie algebra is weakly complete. A comment following Theorem 3.12 of [2] exhibits an example of a weakly complete 𝕂\mathbb{K}-Lie algebra which is not a profinite-dimensional Lie algebra.

Let 𝒲𝒜{\cal W}\hskip-2.0pt{\cal A} denote the category of weakly complete associative unital algebras over 𝕂\mathbb{K}. However, instead of considering the full category of weakly complete Lie algebras over 𝕂\mathbb{K}, in the following we consider 𝒲{\cal W}\hskip-1.0pt{\cal L}, the category of profinite-dimensional Lie algebras over 𝕂\mathbb{K} and continuous 𝕂\mathbb{K}-Lie algebra morphisms. The reason for this restriction is Theorem 7.1 stating that every weakly complete unital 𝕂\mathbb{K}-algebra is the projective limit of its finite-dimensional quotient algebras. This implies at once the following {Proposition} Let AA be any weakly complete unital 𝕂\mathbb{K}-algebra and ALieA_{\rm Lie} the weakly complete Lie algebra obtained by considering on the weakly complete vector space AA the Lie algebra obtained with the Lie bracket [x,y]=xyyx[x,y]=xy-yx. Then ALieA_{\rm Lie} is profinite-dimensional.

The functor which associates with a weakly complete associative algebra AA the profinite-dimensional Lie algebra ALieA_{\rm Lie} is called the underlying Lie algebra functor.

For the complete proof of the following existence theorem, we shall invoke a considerable portion of a bulk category theoretical arguments. We shall collect these in an appendix, since, firstly, they reach far beyond the current application and, secondly, their full presentation might have led the reader astray from the present line of thought had we presented them at this point.

{Theorem}

(The Existence Theorem of 𝐔\mathop{\bf U\hphantom{}}\nolimits) The underlying Lie algebra functor AALieA\mapsto A_{\rm Lie} from 𝒲𝒜{\cal W}\hskip-2.0pt{\cal A} to 𝒲{\cal W}\hskip-1.0pt{\cal L} has a left adjoint 𝐔:𝒲𝒲𝒜\mathop{\bf U\hphantom{}}\nolimits\colon{\cal W}\hskip-1.0pt{\cal L}\to{\cal W}\hskip-2.0pt{\cal A}.

The front adjunction λ𝔤:𝔤𝐔(𝔤)Lie\lambda_{\mathfrak{g}}\colon\mathfrak{g}\to\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})_{\rm Lie} is an embedding of profinite-dimensional Lie algebras.

{Proof}

The category 𝒲{\cal W}\hskip-1.0pt{\cal L} is complete. (Exercise. Cf. Theorem A3.48 of [11], p. 819.) The “Solution Set Condition” (of Definition A3.58 in [11], p. 824) holds. (Exercise: Cf. the proof Lemma 3.58 of [11], p. 91.) Hence 𝐔\mathop{\bf U\hphantom{}}\nolimits exists by the Adjoint Functor Existence Theorem (i.e., Theorem A3.60 of [11], p. 825).

The assertion about λ𝔤\lambda_{\mathfrak{g}} being an embedding follows from Proposition 7.5 (ii) in the Appendix.

In other words, each profinite-dimensional Lie algebra 𝔤\mathfrak{g} may be considered as a closed Lie subalgebra of 𝐔(𝔤)Lie\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})_{\rm Lie} with the property that each continuous Lie algebra morphism f:𝔤ALief\colon\mathfrak{g}\to A_{\rm Lie} for some weakly complete associative unital algebra AA extends uniquely to a 𝒲𝒜{\cal W}\hskip-2.0pt{\cal A}-morphism f:𝐔(𝔤)Af^{\prime}\colon\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})\to A.

𝒲𝒲𝒜𝔤λ𝔤𝐔(𝔤)Lie𝐔(𝔤)f(f)Lie!fALieidALieA.\begin{matrix}&{\cal W}\hskip-1.0pt{\cal L}&&\hbox to19.91692pt{}&{\cal W}\hskip-2.0pt{\cal A}\cr\vskip 3.0pt\cr\hrule\cr\cr\vskip 3.0pt\cr\mathfrak{g}&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\lambda_{\mathfrak{g}}}}&\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})_{\rm Lie}&\hbox to19.91692pt{}&\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})\\ \hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle\forall f$}}$}\Big{\downarrow}&&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle(f^{\prime})_{\rm Lie}$}}$\hss}&\hbox to19.91692pt{}&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle\exists!f^{\prime}$}}$\hss}\\ A_{\rm Lie}&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits_{\mathop{\rm id}\nolimits}}&A_{\rm Lie}&\hbox to19.91692pt{}&A.\end{matrix}

If necessary we shall write 𝐔𝕂\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{K}} instead of 𝐔\mathop{\bf U\hphantom{}}\nolimits whenever the ground field should be emphasized.

{Definition}

For each profinite-dimensional 𝕂\mathbb{K}-Lie algebra, we shall call 𝐔𝕂(𝔤)\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{K}}(\mathfrak{g}) the weakly complete enveloping algebra of 𝔤\mathfrak{g} (over 𝕂\mathbb{K}).

{Remark}

For every profinite-dimensional Lie algebra 𝔤\mathfrak{g}, a morphism f:𝔤𝕂f\colon\mathfrak{g}\to\mathbb{K}, f(x)=0f(x)=0, according to the definition of 𝐔𝕂(𝔤)\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{K}}(\mathfrak{g}) induces a natural 𝒲𝒜{\cal W}\hskip-2.0pt{\cal A}-morphism α𝔤:𝐔𝕂(𝔤)𝕂\alpha_{\mathfrak{g}}\colon\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{K}}(\mathfrak{g})\to\mathbb{K} such that α𝔤(𝔤)={0}\alpha_{\mathfrak{g}}(\mathfrak{g})=\{0\} and that α𝔤ι𝐔(𝔤)=id𝕂\alpha_{\mathfrak{g}}\circ\iota_{{\mathop{\bf U\hphantom{}}\nolimits}(\mathfrak{g})}=\mathop{\rm id}\nolimits_{\mathbb{K}}.

The retraction α𝔤\alpha_{\mathfrak{g}} is also called the augmentation of 𝐔𝕂(𝔤)\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{K}}(\mathfrak{g}).

In the Appendix we shall also introduce for any weakly complete vector space WW its weakly complete tensor algebra 𝐓(W)\mathop{\bf T\hphantom{}}\nolimits(W) (cf. paragraph (C) preceding Proposition 7.5 and Theorem 8) and show that 𝐔(𝔤)\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}) is a quotient algebra of 𝐓(|𝔤|)\mathop{\bf T\hphantom{}}\nolimits(|\mathfrak{g}|) if |𝔤||\mathfrak{g}| is the underlying weakly complete vector space underlying 𝔤\mathfrak{g}. There is a commutative diagram

𝐓(|𝔤|)α|𝔤|𝕂f=𝐔𝕂(𝔤)α𝔤𝕂.\begin{matrix}\mathop{\bf T\hphantom{}}\nolimits(|\mathfrak{g}|)&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\alpha_{|\mathfrak{g}|}}}&\mathbb{K}\\ \hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle f^{\prime}$}}$}\Big{\downarrow}&&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle=$}}$\hss}\\ \mathop{\bf U\hphantom{}}\nolimits_{\mathbb{K}}(\mathfrak{g})&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits_{\alpha_{\mathfrak{g}}}}&\mathbb{K}.\end{matrix}

Moreover, we have the following corollary to our existence theorem:

{Corollary}

For any profinite-dimensional Lie algebra 𝔤\mathfrak{g}, the unital associative subalgebra 𝔤\langle\mathfrak{g}\rangle generated algebraically in 𝐔(𝔤)\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}) by 𝔤\mathfrak{g} is dense in 𝐔(𝔤)\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}).

{Proof}

The assertion follows from Proposition 7.5 in the Appendix.


Of course we would like to have a better insight into the structure of the algebra 𝔤\langle\mathfrak{g}\rangle. This information we provide in the following section and thereby close the gap between the concepts of the weakly complete enveloping algebra and the classical universal enveloping abstract algebra dealt with in the Poincaré-Birkhoff-Witt Theorem.

2 The Abstract Enveloping Algebra U(L)U(L) of a Lie Algebra LL

We briefly recall that the functor which assigns to a unital 𝕂\mathbb{K}-algebra XX the underlying Lie algebra XLieX_{\rm Lie} (with the underlying 𝕂\mathbb{K} vector space of XX as vector space structure endowed with the bracket operation (x,y)[x,y]:=xyyx(x,y)\mapsto[x,y]:=xy-yx as Lie bracket) has a left adjoint functor UU which assigns to a Lie algebra LL a unital associative algebra U(L)U(L) and a natural Lie algebra morphism ρL:LU(L)Lie\rho_{L}\colon L\to U(L)_{\rm Lie} such that for each Lie algebra morphim f:LXLief\colon L\to X_{\rm Lie} for a unital algebra XX there is a unique morphism of unital algebras f:U(L)Xf^{\prime}\colon U(L)\to X such that f=fLieρLf=f^{\prime}_{\rm Lie}\circ\rho_{L}. The algebra U(L)U(L) is called the universal enveloping algebra of LL. A large body of text book literature is available on it. A prominent result is the Poincaré-Birkhoff-Witt Theorem on the structure of U(L)U(L) which implies in particular that ρL:LU(L)Lie\rho_{L}\colon L\to U(L)_{\rm Lie} is injective.


From the Theorem of Poincaré, Birkhoff and Witt it is known that ρL\rho_{L} is injective. One may therefore assume that LU(L)L\subseteq U(L) such that ρL\rho_{L} is the inclusion function. (See [1] or [3].) In this parlance the universal property reads as follows:

For each unital algebra AA, each Lie algebra morphism f:LALief\colon L\to A_{\rm Lie} extends uniquely to an algebra morphism f:U(L)Af^{\prime}\colon U(L)\to A.

Also from the Theorem of Poincaré, Birkhoff and Witt we know that

U(L)U(L) is the unital algebra generated by LL, i.e., U(L)=LU(L)=\langle L\rangle.


In the present section we shall now denote by |𝔤||\mathfrak{g}| the abstract Lie algebra underlying the profinite-dimensional Lie algebra 𝔤\mathfrak{g}. Then the main result of this section will be a complete clarification of the relation of the weakly complete enveloping algebra 𝐔(𝔤)\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}) of a profinite-dimensional Lie algebra 𝔤\mathfrak{g} and the universal enveloping algebra U(|𝔤|)U(|\mathfrak{g}|) of |𝔤||\mathfrak{g}|.

{Lemma}

For a profinite-dimensional Lie algebra 𝔤\mathfrak{g} there is a natural morphism ε𝔤:U(|𝔤|)|𝐔(𝔤)|\varepsilon_{\mathfrak{g}}\colon U(|\mathfrak{g}|)\to|\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})| of unital algebras such that

  1. (i)

    the following diagram is commutative:

    |𝔤|inclU(|𝔤|)id|𝔤|ε𝔤|𝔤||incl||𝐔(𝔤)|.\begin{matrix}|\mathfrak{g}|&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\mathop{\rm incl}\nolimits}}&U(|\mathfrak{g}|)\\ \hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle\mathop{\rm id}\nolimits_{|\mathfrak{g}|}$}}$}\Big{\downarrow}&&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle\varepsilon_{\mathfrak{g}}$}}$\hss}\\ |\mathfrak{g}|&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits_{|\mathop{\rm incl}\nolimits|}}&|\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})|.\end{matrix}
  2. (ii)

    The image of ε𝔤\varepsilon_{\mathfrak{g}} is dense in 𝐔(𝔤)\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}).

  3. (iii)

    The morphism ε𝔤\varepsilon_{\mathfrak{g}} is injective if 𝔤\mathfrak{g} is finite-dimensional.

{Proof}

(i) The claim is a direct consequence of the universal property of the functor UU.

(ii) We have im(ε𝔤)=ε𝔤(U(|𝔤|))=ε𝔤(|𝔤|=ε𝔤(𝔤)=|𝔤|\mathop{\rm im}\nolimits(\varepsilon_{\mathfrak{g}})=\varepsilon_{\mathfrak{g}}(U(|\mathfrak{g}|))=\varepsilon_{\mathfrak{g}}(\langle|\mathfrak{g}|\rangle=\langle\varepsilon_{\mathfrak{g}}(\mathfrak{g})\rangle=\langle|\mathfrak{g}|\rangle in |𝐔(𝔤)||\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})|. From Corollary 1 we know that |𝔤|\langle|\mathfrak{g}|\rangle is dense in 𝐔(𝔤)\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}).

(iii) If 𝔤\mathfrak{g} is finite-dimensional, then every finite-dimensional Lie algebra representation ρ:𝔤ALie=End(V)Lie\rho\colon\mathfrak{g}\to A_{\rm Lie}=\mathrm{End}(V)_{\rm Lie} for a finite-dimensional vector space VV extends to an associative representation ρ:𝐔(𝔤)End(V)\rho^{\prime}\colon\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})\to\mathrm{End}(V) . Then ρε𝔤:U(𝔤)End(V)\rho\circ\varepsilon_{\mathfrak{g}}\colon U(\mathfrak{g})\to\mathrm{End}(V) is an extension to an associative representation of U(𝔤)U(\mathfrak{g}) which is unique. By Harish-Chandra’s Lemma (see Dixmier [3], 2.5.7), the extensions of associative representations of U(𝔤)U(\mathfrak{g}) of finite-dimensional Lie algebra representations of 𝔤\mathfrak{g} separate the points of U(𝔤)U(\mathfrak{g}) and so the claim follows.

The remainder of this section now is devoted to removing the restriction to finite-dimensionality in Lemma 2(iii). That is, we want to show

{Lemma}

For a profinite-dimensional Lie algebra 𝔤\mathfrak{g}, the algebra morphism ε𝔤:U(|𝔤|)|𝐔(𝔤)|\varepsilon_{\mathfrak{g}}\colon U(|\mathfrak{g}|)\to|\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})| is injective.

The proof will occupy the remainder of this section. We shall resort to the existing literature on U(L)U(L) such as [1] or [3]. We are given the profinite-dimensional Lie algebra 𝔤\mathfrak{g} and we write L:=|𝔤|L:=|\mathfrak{g}| for the underlying Lie algebra. So LU(L)LieL\subseteq U(L)_{\rm Lie}. Let BB be a totally ordered basis of LL. We begin by recalling the following basic fact from the Poincaré-Birkhoff-Witt Theorem (see [1], Corollary 3, Section 7 of Paragraph 2):

B~:={b1b2bm|1m,b1,,bmB,b1b2bm} is a basis of U(L).\widetilde{B}{:=}\{b_{1}b_{2}{\cdots}b_{m}|1{\leq}m,\ b_{1},\dots,b_{m}\in B,\ b_{1}{\leq}b_{2}{\leq}\cdots{\leq}b_{m}\}\mbox{ is a basis of $U(L)$.} (PBW)

Now assume that

0uU(L).0\neq u\in U(L).

Then there is a finite subset FBF\subseteq B such that uspan(F~)u\in\mathop{\rm span}\nolimits(\widetilde{F}) for

F~={b1b2bm|1m,b1,,bmF,b1b2bm}B~.\widetilde{F}=\{b_{1}b_{2}\cdots b_{m}|1\leq m,\quad b_{1},\dots,b_{m}\in F,\quad b_{1}\leq b_{2}\leq\cdots\leq b_{m}\}\subseteq\widetilde{B}.
{Lemma}

There is a closed ideal JJ of 𝔤\mathfrak{g} so that JspanF={0}J\cap\mathop{\rm span}\nolimits F=\{0\}. {Proof} The vector space V=spanFV=\mathop{\rm span}\nolimits F is finite-dimensional. Let CC be the boundary of a compact 0-neighborhood in VV. Then V=𝕂CV=\mathbb{K}{\cdot}C.

Returning at this point to the fact that 𝔤\mathfrak{g} is a profinite-dimensional Lie algebra, we conclude that there is a filterbasis {\cal I} of closed ideals II of 𝔤\mathfrak{g} such that dim𝔤/I<\dim\mathfrak{g}/I<\infty for II\in{\cal I} such that 𝔤limI𝔤/I\mathfrak{g}\cong\lim_{I\in{\cal I}}\mathfrak{g}/I. In particular, ={0}\bigcap{\cal I}=\{0\}. Therefore I(CI)=C{0}=\bigcap_{I\in{\cal I}}(C\cap I)=C\cap\{0\}=\emptyset. Since CC is compact, the filter basis {CI:I}\{C\cap I:I\in{\cal I}\} with empty intersection consists of compact sets and therefore must contain the empty set. Thus there is a JJ\in{\cal I} such that CJ=C\cap J=\emptyset. In fact, we have JspanF={0}J\cap\mathop{\rm span}\nolimits F=\{0\}. Indeed, suppose there were a nonzero t𝕂t\in\mathbb{K} and a cCc\in C such that tcJt{\cdot}c\in J, then c=t1(tc)t1J=Jc=t^{-1}{\cdot}(t{\cdot}c)\in t^{-1}{\cdot}J=J which is impossible.

{Lemma}

Assume that there is an ideal JJ of LL such that JspanF={0}J\cap\mathop{\rm span}\nolimits F=\{0\}. Then the image of uu under the morphism U(L)U(L/J)U(L)\to U(L/J) is nonzero.

{Proof}

We choose a finite dimensional vector subspace HH of LL containing FF such that L=HJL=H\oplus J. Let EE be a totally ordered basis for HH such that the order of EE extends that of FF, choose a totally ordered basis DD of JJ and make sure that EDE\cup D has a total order extending the orders of EE and DD, thus yielding a totally ordered basis of LL.

We consider the quotient morphism of Lie algebras qJ:LL/Jq_{J}\colon L\to L/J. Then qJq_{J} maps EE bijectively onto a basis E=qJ(E)E^{\prime}=q_{J}(E) of L/JL/J. Let

E~={b1bm|m1,b1,,bmE}.\widetilde{E}=\{b_{1}\cdots b_{m}\,|\quad m\geq 1,b_{1},\dots,b_{m}\in E\}.

Then U(qJ):U(L)U(L/J)U(q_{J})\colon U(L)\to U(L/J) maps E~\widetilde{E} bijectively onto a basis E~\widetilde{E^{\prime}} of U(L/J)U(L/J) by (PBW). If we now write u=SE~cSSu=\sum_{S\in\widetilde{E}}c_{S}{\cdot}S with cS𝕂c_{S}\in\mathbb{K}, then U(qJ)(u)=qJ(S)E~cSqJ(S)0U(q_{J})(u)=\sum_{q_{J}(S)\in\widetilde{E^{\prime}}}c_{S}{\cdot}q_{J}(S)\neq 0, since E~\widetilde{E^{\prime}} is a basis of U(L/J)U(L/J).

As the kernel of the quotient map U(L)U(L/J)U(L)\to U(L/J) is U(L)JU(L)J, the claim of the preceding lemma may be expressed equivalently in the form uU(L)Ju\notin U(L)J.

Now we recall that qJ:LL/JU(L/J)q_{J}\colon L\to L/J\subseteq U(L/J) is in fact the underlying Lie algebra morphism |pJ||p_{J}| of a quotient morphism pJ:𝔤𝔤/Jp_{J}\colon\mathfrak{g}\to\mathfrak{g}/J of profinite Lie algebras and that qJq_{J} extends uniquely to an algebra morphism U(|pJ|):U(L)U(L/J)U(|p_{J}|)\colon U(L)\to U(L/J) with kernel U(L)JU(L)J. Then from Lemma 2 we know that U(|pJ|)(u)U(|p_{J}|)(u) is nonzero in U(L/J)U(L/J) and from Lemma 2 we infer that ε𝔤/J\varepsilon_{\mathfrak{g}/J} is injective. The commutative diagram

|𝔤|inclU(|𝔤|)ε𝔤|𝐔(𝔤)||pJ|U(|pJ|)|𝐔(pJ)||𝔤/J|inclU(|𝔤/J|)ε𝔤/J|𝐔(𝔤/J)|\begin{matrix}|\mathfrak{g}|&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\mathop{\rm incl}\nolimits}}&U(|\mathfrak{g}|)&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\varepsilon_{\mathfrak{g}}}}&|\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})|\\ \hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle|p_{J}|$}}$}\Big{\downarrow}&&\hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle U(|p_{J}|)$}}$}\Big{\downarrow}&&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle|\mathop{\bf U\hphantom{}}\nolimits(p_{J})|$}}$\hss}&&\\ |\mathfrak{g}/J|&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\mathop{\rm incl}\nolimits}}&U(|\mathfrak{g}/J|)&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\varepsilon_{\mathfrak{g}/J}}}&|\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}/J)|\end{matrix}

then shows that ε𝔤(u)0\varepsilon_{\mathfrak{g}}(u)\neq 0. Therefore, since uU(L){0}u\in U(L)\setminus\{0\} was arbitrary, ε𝔤\varepsilon_{\mathfrak{g}} is injective, leaving the elements of L=|𝔤|L=|\mathfrak{g}| fixed. This completes the proof of Lemma 2. Thus U(|𝔤|)U(|\mathfrak{g}|) may be considered as a subalgebra of |𝐔(𝔤)||\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})|, containing |𝔤||𝐔(𝔤)||\mathfrak{g}|\subseteq|\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})|.

This may be rephrased in the following Theorem which summarizes our efforts to elucidate the close relation between U(|𝔤|)U(|\mathfrak{g}|) and 𝐔(𝔤)\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}):

{Theorem}

(The Relation of U()U(-) and 𝐔()\mathop{\bf U\hphantom{}}\nolimits(-)) For any profinite-dimensional real or complex Lie algebra 𝔤\mathfrak{g} considered as a closed Lie subalgebra of 𝐔(𝔤)Lie\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})_{\rm Lie}, the associative unital subalgebra 𝔤\langle\mathfrak{g}\rangle generated algebraically by 𝔤\mathfrak{g} in 𝐔(𝔤)\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}) is naturally isomorphic to U(|𝔤|)U(|\mathfrak{g}|) (under an isomorphism fixing the elements of 𝔤\mathfrak{g}) and is dense in 𝐔(𝔤)\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}).

In a slightly careless sense we may memorize this as saying:

For a profinite-dimensional Lie algebra 𝔤\mathfrak{g}, the weakly complete topological enveloping algebra 𝐔(𝔤)\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}) is “a completion of U(|𝔤|)U(|\mathfrak{g}|)”, and we have

𝔤𝔤=U(𝔤)U(𝔤)¯=𝐔(𝔤).\mathfrak{g}\subseteq\langle\mathfrak{g}\rangle=U(\mathfrak{g})\subseteq\overline{U(\mathfrak{g})}=\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}). ()

3 The Projective Limit Preservation
of the Weakly Complete Enveloping Functor U

Since every weakly complete unital algebra is a strict projective limit of all finite-dimensional quotient algebras, it will now turn out to be sufficient to test the universal property of the functor 𝐔\mathop{\bf U\hphantom{}}\nolimits only for finite-dimensional unital associative algebras:

{Proposition}

Assume that the profinite-dimensional Lie algebra 𝔤\mathfrak{g} is contained functorially in a weakly complete unital algebra 𝐕(𝔤)\mathop{\bf V\hphantom{}}\nolimits(\mathfrak{g}) such that for each finite-dimensional unital algebra AA and each morphism of profinite-dimensional Lie algebras f:𝔤ALief\colon\mathfrak{g}\to A_{\rm Lie} there is a unique morphism of weakly complete unital algebras f:𝐕(𝔤)Af^{\prime}\colon\mathop{\bf V\hphantom{}}\nolimits(\mathfrak{g})\to A extending ff. Then 𝐕(𝔤)𝐔(𝔤)\mathop{\bf V\hphantom{}}\nolimits(\mathfrak{g})\cong\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}) naturally.

{Proof}

We apply the Density and Adjunction Theorem 7.2 in the Appendix with 𝒜{\cal A} as the category of weakly complete associative unital algebras, and {\cal B} as the category of profinite-dimensional Lie algebras with the full subcategory d{\cal B}_{d} of finite-dimensional Lie algebras which is topologically dense in {\cal B}. Then, by hypothesis, the function 𝐕:obob𝒜\mathop{\bf V\hphantom{}}\nolimits\colon\mathop{\rm ob}\nolimits{\cal B}\to\mathop{\rm ob}\nolimits{\cal A} is conditionally left adjoint to the functor ()Lie:𝒜(\cdot)_{\rm Lie}\colon{\cal A}\to{\cal B} which maps an associative algebra to the Lie algebra with the Lie bracket [a,b]=abba[a,b]=ab-ba (see Definition 7.2). Then by Theorem 7.2, 𝐕\mathop{\bf V\hphantom{}}\nolimits is naturally isomorphic to the left adjoint 𝐔\mathop{\bf U\hphantom{}}\nolimits of ()Lie(\cdot)_{\rm Lie}.

Perhaps more deeply we shall see now that, while as a left-adjoint functor, 𝐔\mathop{\bf U\hphantom{}}\nolimits preserves colimits, is also preserve certain limits, namely, the projective limits 𝔤=lim𝔦(𝔤)𝔤/𝔦\mathfrak{g}=\lim_{\mathfrak{i}\in{\cal I}(\mathfrak{g})}\mathfrak{g}/\mathfrak{i} of Definition 1. Indeed in Theorem 7.5 in the Appendix we show:

{Theorem}

(𝐔\mathop{\bf U\hphantom{}}\nolimits preserves some projective limits) For a profinite-dimensional Lie algebra 𝔤\mathfrak{g} with its filter basis (𝔤){\cal I}(\mathfrak{g}) of cofinite-dimensional ideals 𝔦\mathfrak{i} we have

𝔤lim𝔦(𝔤)𝔤/𝔦 in 𝒲 and 𝐔(𝔤)lim𝔦(𝔤)𝐔(𝔤/𝔦) in 𝒲𝒜.\mathfrak{g}\cong\lim_{\mathfrak{i}\in{\cal I}(\mathfrak{g})}\mathfrak{g}/\mathfrak{i}\mbox{ in }{\cal W}\hskip-1.0pt{\cal L}\mbox{ and }\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})\cong\lim_{\mathfrak{i}\in{\cal I}(\mathfrak{g})}\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}/\mathfrak{i})\mbox{ in }{\cal W}\hskip-2.0pt{\cal A}.

The argument in the Appendix shows, that while the assertion of the theorem is natural and easy to absorb, its proof is deeper than one would expect initially.

4 The Weakly Complete Universal Enveloping Algebra
as a Hopf Algebra

We now address the important aspect of enveloping algebras from their beginning, namely, the fact that they are symmetric Hopf algebras. For some of the proofs in this section we refer to our predecessor paper [7].

{Proposition}

The universal enveloping functor 𝐔\mathop{\bf U\hphantom{}}\nolimits is multiplicative, that is, there is a natural isomorphism 𝐔(𝔤1×𝔤2)𝐔(𝔤1)𝒲𝐔(𝔤2)\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}_{1}\times\mathfrak{g}_{2})\to\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}_{1})\otimes_{\cal W}\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}_{2}).

For a proof see [7], Proposition 6.3.

{Lemma}

For any weakly complete unital algebra AA, the vector space morphism ΔA:AA𝒲A\Delta_{A}\colon A\to A\otimes_{\cal W}A, ΔA(a)=a1+1a\Delta_{A}(a)=a\otimes 1+1\otimes a is a morphism of weakly complete Lie algebras ALie(AWA)LieA_{\rm Lie}\to(A\otimes_{W}A)_{\rm Lie}.

Cf. [7], Lemma 6.4.

Recall the natural morphism λ𝔤:𝔤𝐔(𝔤)Lie\lambda_{\mathfrak{g}}\colon\mathfrak{g}\to\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})_{\rm Lie} which we consider as an inclusion morphism. By Lemma 4,

p𝔤=δ𝐔(𝔤)λ𝔤:𝔤(𝐔(𝔤)𝒲𝐔(𝔤))Liep_{\mathfrak{g}}=\delta_{\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})}\circ\lambda_{\mathfrak{g}}\colon\mathfrak{g}\to(\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})\otimes_{\cal W}\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}))_{\rm Lie}

is a morphism of weakly complete Lie algebras. By the universal property of 𝐔\mathop{\bf U\hphantom{}}\nolimits, p𝔤p_{\mathfrak{g}} yields a unique natural morphism of weakly complete associative unital algebras γ𝔤:𝐔(𝔤)𝐔(𝔤)𝒲𝐔(𝔤)\gamma_{\mathfrak{g}}\colon\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})\to\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})\otimes_{\cal W}\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}) such that p𝔤=(γ𝔤)Lieλ𝔤p_{\mathfrak{g}}=(\gamma_{\mathfrak{g}})_{\rm Lie}\circ\lambda_{\mathfrak{g}}. Recall the augmentation α𝔤:𝐔𝕂(𝔤)𝕂\alpha_{\mathfrak{g}}\colon\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{K}}(\mathfrak{g})\to\mathbb{K} (see Remark 1) and the inclusion morphism ι𝐔(𝔤):𝕂𝐔𝕂(𝔤)\iota_{\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})}\colon\mathbb{K}\to\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{K}}(\mathfrak{g}) (see Remark 8). Accordingly, we have an idempotent endomorphism

ι𝐔K(𝔤)α𝔤:𝐔𝕂(𝔤)𝐔𝕂(𝔤).\iota_{\mathop{\bf U\hphantom{}}\nolimits_{K}(\mathfrak{g})}\circ\alpha_{\mathfrak{g}}\colon\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{K}}(\mathfrak{g})\to\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{K}}(\mathfrak{g}).

Further, the augmentation α𝔤\alpha_{\mathfrak{g}} acts as coidentity, and the function xx:𝐔(𝔤)𝐔(𝔤)x\mapsto-x:\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})\to\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}) as symmetry as is readily checked for x𝔤x\in\mathfrak{g}, and 𝔤\mathfrak{g} generates 𝐔𝕂(𝔤)\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{K}}(\mathfrak{g}) as topological algebra by Corollary 1.

Now we have

{Proposition}

(𝐔(𝔤)\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}) as a Hopf algebra)

(a) Each weakly complete enveloping algebra 𝐔(𝔤)\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}) is a weakly complete symmetric Hopf algebra with the comultiplication γ𝔤\gamma_{\mathfrak{g}} and the augmentation α𝔤:𝐔(𝔤)𝕂\alpha_{\mathfrak{g}}\colon\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})\to\mathbb{K} as coidentity.

(b) If f:𝔤𝔥f\colon\mathfrak{g}\to\mathfrak{h} is a morphism of profinite-dimensional Lie algebras, then the morphism 𝐔𝕂(f):𝐔K(𝔤)𝐔K(𝔥)\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{K}}(f)\colon\mathop{\bf U\hphantom{}}\nolimits_{K}(\mathfrak{g})\to\mathop{\bf U\hphantom{}}\nolimits_{K}(\mathfrak{h}) respects comultiplication, coidentity, and symmetry, that is, 𝐔𝕂(f)\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{K}}(f) is a morphism of symmetric Hopf algebras.

{Proof}

For (a), see [7], Corollary 6.5.

For (b) we consider a morphism f:𝔤𝔥f\colon\mathfrak{g}\to\mathfrak{h} a morphism of profinite-dimensional Lie algebras and for the functoriality of 𝐔\mathop{\bf U\hphantom{}}\nolimits (short for 𝐔K()\mathop{\bf U\hphantom{}}\nolimits_{K}(-)) as regards to comultiplication γ𝔤:𝐔(𝔤)𝐔(𝔤)𝐔(𝔤)\gamma_{\mathfrak{g}}\colon\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})\to\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})\otimes\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}) we verify the commutativity of the following diagram (with =𝒲\otimes=\otimes_{\cal W})

𝐔(𝔤)γ𝔤𝐔(𝔤)𝐔(𝔤)𝐔(𝔤×𝔤)𝐔(f)𝐔(f)𝐔(f)𝐔(f×f)𝐔(𝔥)γ𝔥𝐔(𝔥)𝐔(𝔥)𝐔(𝔥×𝔥).\begin{matrix}\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\gamma_{\mathfrak{g}}}}&\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})\otimes\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\cong}}&\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}\times\mathfrak{g})\\ \hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle\mathop{\bf U\hphantom{}}\nolimits(f)$}}$}\Big{\downarrow}&&\hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle\mathop{\bf U\hphantom{}}\nolimits(f)\otimes\mathop{\bf U\hphantom{}}\nolimits(f)$}}$}\Big{\downarrow}&&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle\mathop{\bf U\hphantom{}}\nolimits(f\times f)$}}$\hss}\\ \mathop{\bf U\hphantom{}}\nolimits(\mathfrak{h})&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits_{\gamma_{\mathfrak{h}}}}&\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{h})\otimes\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{h})&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits_{\cong}}&\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{h}\times\mathfrak{h}).\end{matrix}

(See also Proposition 4.) Coidentity and symmetry are treated similarly.

This proposition expresses the fact that 𝐔𝕂\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{K}} is a functor from the category of profinite-dimensional Lie algebras to the category of weakly complete symmetric Hopf algebras. Its significance is emphasised by the fact that essential portions of the noteworthy theory of weakly complete symmetric Hopf algebras have meanwhile entered the textbook literature. (See [11], Appendix A3, Appendix A7, Chapter 3–Part 3.) We have collected some essential features in our Appendix such as Theorems 9 and 9.

Now we specialize these to the case of A=𝐔𝕂(𝔤)A=\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{K}}(\mathfrak{g}). We use the notation <={r:0<r}\mathbb{R}_{<}=\{r\in\mathbb{R}:0<r\} and recall that 𝔤A\mathfrak{g}\subseteq A and that A×A^{\times} denotes the group of units of AA. For the exponential function exp:ALieA×\exp\colon A_{\rm Lie}\to A^{\times} as in Theorem 9 we define the closed subgroup

Γ(𝔤):=exp𝔤¯A×.\Gamma^{*}(\mathfrak{g}):=\overline{\langle\exp\mathfrak{g}\rangle}\subseteq A^{\times}.

The following theorem now is a principal result in the theory of weakly complete enveloping algebras of profinite-dimensional real or complex Lie algebras.

{Theorem}

(The Weakly Complete Enveloping Hopf Algebra) Let 𝔤\mathfrak{g} be a profinite-dimensional Lie algebra and 𝐔(𝔤)\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}) its weakly complete enveloping algebra containing 𝔤\mathfrak{g} according to Theorem 2. Then the following statements hold:

  1. (a)

    The group of units 𝐔(𝔤)×\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})^{\times} is dense in 𝐔(𝔤)\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}). It is an almost connected pro-Lie group, connected in the case of 𝕂=\mathbb{K}=\mathbb{C}. The algebra 𝐔(𝔤)\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}) has an exponential function exp:𝐔(𝔤)Lie𝐔(𝔤)×\exp\colon{\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})_{\rm Lie}}\to\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})^{\times}. The Lie algebra 𝔏(𝐔(𝔤)×)\mathfrak{L}(\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})^{\times}) of 𝐔(𝔤)×\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})^{\times} is (naturally isomorphic to) 𝐔(𝔤)Lie\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})_{\rm Lie}.

  2. (b)

    The pro-Lie algebra (𝐔(𝔤))\mathbb{P}(\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})) is the Lie algebra of the pro-Lie group 𝔾(𝐔(𝔤))\mathbb{G}(\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})) of grouplike elements and the restriction and corestriction of exp\exp is the exponential function for this group.

  3. (c)

    The profinite-dimensional Lie algebra (𝐔(𝔤))\mathbb{P}(\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})) contains

    𝔤=(U(|𝔤|))=(𝐔(𝔤))U(|𝔤|).\mathfrak{g}=\mathbb{P}(U(|\mathfrak{g}|))=\mathbb{P}(\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}))\cap U(|\mathfrak{g}|).
  4. (d)

    For 𝕂=\mathbb{K}=\mathbb{R}, the restriction and corestriction of exp\exp yields the exponential function

    expΓ(𝔤):𝔤=𝔏(Γ(𝔤))Γ(𝔤)\exp_{\Gamma^{*}(\mathfrak{g})}\colon\mathfrak{g}=\mathfrak{L}(\Gamma^{*}(\mathfrak{g}))\to\Gamma^{*}(\mathfrak{g})

    of that pro-Lie subgroup Γ(𝔤)\Gamma^{*}(\mathfrak{g}) of 𝐔(𝔤)×\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{R}}(\mathfrak{g})^{\times} whose Lie algebra is precisely 𝔤\mathfrak{g}.

  5. (e)

    Define the hyperplane ideal 𝕀{\mathbb{I}} as the kernel of the augmentation α𝔤\alpha_{\mathfrak{g}}. Then we have

    1. (i)

      for 𝕂=\mathbb{K}=\mathbb{R}exp(𝐔(𝔤))=(<1)𝕀\exp(\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{R}}(\mathfrak{g}))=(\mathbb{R}_{<}{\cdot}1)\oplus{\mathbb{I}}, an open half space,

    2. (ii)

      for 𝕂=\mathbb{K}=\mathbb{C}exp(𝐔(𝔤))=𝐔(𝔤)𝕀\exp(\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{C}}(\mathfrak{g}))=\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{C}}(\mathfrak{g})\setminus{\mathbb{I}}.

{Proof}

For the proofs of (a) and (b) see Theorem 9 in the Appendix.

The proof of (c) follows from [7], Theorem 3.4 and Theorem 2. Cf. also [11], Theorem A3.102 and its proof for 𝕂{,}\mathbb{K}\in\{\mathbb{R},\mathbb{C}\}.

The proof of (d) must verify that 𝔏(𝔤¯)𝔤\mathfrak{L}(\overline{\langle\mathfrak{g}\rangle})\subseteq\mathfrak{g}. This conclusion we derive from [10], Corollary 4.22 and its proof.

The proof of (e) follows from Theorem 9 in the Appendix.

Here are some immediate consequences:

{Corollary}

The weakly complete enveloping algebra 𝐔(𝔤)\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}) of a nonzero profinite-dimensional weakly complete Lie algebra 𝔤\mathfrak{g} has nontrivial grouplike elements contained in 𝐔(𝔤)×\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})^{\times}. Specifically, there is a pro-Lie subgroup Γ(𝔤)\Gamma^{*}(\mathfrak{g}) of grouplike elements whose Lie algebra is isomorphic to 𝔤\mathfrak{g} and whose exponential function is induced by that of 𝐔(𝔤)\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}).

By contrast, on the purely algebraic side, the universal enveloping Hopf algebra U(L)U(L) of a Lie algebra LL shows no visible nontrivial grouplike elements while a nontrivial weakly complete enveloping algebra always does.

We shall see that even in the case of the smallest possible nonzero candidate 𝔤=𝕂\mathfrak{g}=\mathbb{K}, the space (𝐔(𝔤))\mathbb{P}(\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})) is substantially larger than 𝔤\mathfrak{g} (see Theorem 5.2 below). In the classical setting of the discrete enveloping Hopf algebra in characteristic 0 we have (U(L))=L\mathbb{P}(U(L))=L: see e.g. [18], Theorem 5.4 on p. LA 3.10.

{Corollary}

For any profinite-dimensional Lie algebra 𝔤\mathfrak{g} there is a pro-Lie group GG whose Lie algebra 𝔏(G)\mathfrak{L}(G) may be identified with 𝔤\mathfrak{g}.

Indeed the theorem provides a weakly complete unital algebra AA with an exponential function expA\exp_{A} such that expG:𝔏(G)G\exp_{G}\colon\mathfrak{L}(G)\to G may be identified with a restriction and corestriction of expA\exp_{A}.

This is indeed much more than what is historically known as Sophus Lie’s Third Fundamental Theorem.

4.1 Lie’s Third Fundamental Theorem for profinite-dimensional Lie algebras

It is worthwhile to elucidate the insight that our present context throws a new light on Lie’s Third Fundamental Theorem. Therefore we recall the contemporary aspect of this background:

{Theorem}

(Sophus Lie’s Third Principal Theorem) For every profinite-dimensional real Lie algebra 𝔤\mathfrak{g} there is a simply connected pro-Lie group Γ(𝔤)\Gamma(\mathfrak{g}), whose Lie algebra 𝔏(Γ(𝔤))\mathfrak{L}(\Gamma(\mathfrak{g})) is (isomorphic to) 𝔤\mathfrak{g}. For any pro-Lie group GG with Lie algebra 𝔤\mathfrak{g} there is a quotient morphism α𝔤:Γ(𝔤)G\alpha_{\mathfrak{g}}\colon\Gamma(\mathfrak{g})\to G such that the following diagram commutes:

𝔤=𝔤expΓ(𝔤)expGΓ(𝔤)α𝔤G.\begin{matrix}\mathfrak{g}&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{=}}&\mathfrak{g}\\ \hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle\exp_{\Gamma(\mathfrak{g})}$}}$}\Big{\downarrow}&&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle\exp_{G}$}}$\hss}\\ \Gamma(\mathfrak{g})&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits_{\alpha_{\mathfrak{g}}}}&G.\end{matrix}

For a systematic proof see [9], or e.g. [10], Chapter 6, p. 249, see notably Theorem 6.4, p. 232. Our Theorem 4.1 is also cited in [11], Theorem A7.29. For the definition of simple connectivity see [11], Definition A2.6. Let us recall here that for an abelian 𝔤\mathfrak{g} (that is, a weakly complete real vector space), the underlying vector space of 𝔤𝔏(Γ(𝔤))\mathfrak{g}\cong\mathfrak{L}(\Gamma(\mathfrak{g})) is isomorphic to Γ(𝔤)\Gamma(\mathfrak{g}) via exp𝔏(Γ(𝔤)):𝔏(Γ(𝔤))Γ(𝔤)\exp_{\mathfrak{L}(\Gamma(\mathfrak{g}))}\colon\mathfrak{L}(\Gamma(\mathfrak{g}))\to\Gamma(\mathfrak{g}).

Theorem 4.1 applies at once to G=Γ(𝔤)G=\Gamma^{*}(\mathfrak{g}) as follows:

{Corollary}

For each profinite-dimensional real Lie algebra 𝔤\mathfrak{g} there is a natural morphism α𝔤:Γ(𝔤)Γ(𝔤)\alpha_{\mathfrak{g}}\colon\Gamma(\mathfrak{g})\to\Gamma^{*}(\mathfrak{g}) such that the diagram

𝔤=𝔤expΓ𝔤expΓ(𝔤)Γ(𝔤)α𝔤Γ(𝔤)\begin{matrix}\mathfrak{g}&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{=}}&\mathfrak{g}\\ \hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle\exp_{\Gamma_{\mathfrak{g}}}$}}$}\Big{\downarrow}&&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle\exp_{\Gamma^{*}(\mathfrak{g})}$}}$\hss}\\ \Gamma(\mathfrak{g})&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits_{\alpha_{\mathfrak{g}}}}&\Gamma^{*}(\mathfrak{g})\end{matrix}

is commutative.

The pro-Lie group A×A^{\times} of units of a weakly complete unital associative algebra AA has the property that finite-dimensional continuous representations separate the points, and so any pro-Lie group injected into such a group A×A^{\times} shares this property. Consider the Lie algebra 𝔤=sl(2,))\mathfrak{g}={\rm sl}(2,\mathbb{R})) of the Lie group G=SL(2,)G={\rm SL}(2,\mathbb{R}), and let G~=Γ(𝔤)\widetilde{G}=\Gamma(\mathfrak{g}) be the universal covering group of GG. Every continuous linear representation of G~\widetilde{G}, however, factorizes through GG (see [6], p.590, Example 16.1.8), and therefore Γ(𝔤)\Gamma(\mathfrak{g}) cannot be injected into any group of the form A×A^{\times}. Hence for 𝔤=sl(2,)\mathfrak{g}={\rm sl}(2,\mathbb{R}) the morphism α𝔤:Γ(𝔤)Γ(𝔤)\alpha_{\mathfrak{g}}\colon\Gamma(\mathfrak{g})\to\Gamma^{*}(\mathfrak{g}) cannot be injective and thus certainly cannot be an isomorphism.


5 The Abelian Case

For an abelian Lie algebra 𝔤\mathfrak{g}, the weakly complete unital algebra 𝐔(𝔤)\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}) is commutative. In various special aspects we considered this situation in [2], Lemmas 3.4, 3.5, and 3.10ff., and in [7], Section 5 and Example 6.2. We now return to the commutative situation more systematically now and discuss the structure of 𝐔𝕂(𝔤)\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{K}}(\mathfrak{g}) completely for dim𝔤=1\dim\mathfrak{g}=1, and derive consequences for the abelian case in general.

5.1 The power series algebra

A first and simplest step is the discussion of the power series algebra. We recall the notation ={1,2,3,}\mathbb{N}=\{1,2,3,\dots\} and 0={0,1,2,3}\mathbb{N}_{0}=\{0,1,2,3\dots\}. The set 0\mathbb{N}_{0} is a semiring for addition and multiplication (that is, an addition and multiplicative commutative monoid with distributivity).

{Lemma}

The weakly complete vector space 𝕊:=𝕂0\mathbb{S}:=\mathbb{K}^{\mathbb{N}_{0}} supports a monoid multiplication called convolution as follows: Let 𝐚=(ak)k0𝕊{\bf a}{=}(a_{k})_{k\in\mathbb{N}_{0}}{\in}\mathbb{S} and 𝐛=(bm)m0𝕊{\bf b}{=}(b_{m})_{m\in\mathbb{N}_{0}}{\in}\mathbb{S}. Then

𝐚𝐛=(k+m=nakbm)n0.{\bf a}*{\bf b}=\Big{(}\sum_{k+m=n}a_{k}b_{m}\Big{)}_{n\in\mathbb{N}_{0}}. (1)

With pointwise addition and convolution, 𝕊=(𝕊,+,)\mathbb{S}=(\mathbb{S},+,*) is a weakly complete topological algebra.

{Proof}

The verification is an exercise:

We write X=(0,1,0,0,)X=(0,1,0,0,\dots) and observe

X=(0,1,0,0,0,)X2=(0,0,1,0,0,)X3=(0,0,0,1,0,)\begin{matrix}X&=&(0,1,0,0,0,\dots)\\ X^{2}&=&(0,0,1,0,0,\dots)\\ X^{3}&=&(0,0,0,1,0,\dots)\\ \vdots&&\vdots\end{matrix}

Then

𝐚X=(an)nX=a0X+a1X2+a2X3+.{\bf a}*X=(a_{n})_{n\in\mathbb{N}}*X=a_{0}{\cdot}X+a_{1}{\cdot}X^{2}+a_{2}{\cdot}X^{3}+\cdots.

The projective limit representation 𝕊limn𝕂[X]Xn+1𝕂[X]\mathbb{S}{\cong}\lim_{n\in\mathbb{N}}\frac{\mathbb{K}[X]}{X^{n+1}\mathbb{K}[X]} completes the proof.

Accordingly, 𝕊\mathbb{S} is called the power series algebra in one variable, usually written as 𝕂[[X]]\mathbb{K}[[X]].

It is useful to recall that in the category 𝒲{\cal W} of weakly complete 𝕂\mathbb{K}-vector spaces, for any pair of sets X and Y we have a natural isomorphism

𝕂𝐗𝒲𝕂𝐘𝕂𝐗×𝐘\mathbb{K}^{\bf X}\otimes_{\cal W}\mathbb{K}^{\bf Y}\to\mathbb{K}^{{\bf X}\times{\bf Y}} ()

induced by the bijection (ax)x𝐗(by)y𝐘)(axby)(x,y)𝐗×𝐘(a_{x})_{x\in\bf X}\otimes(b_{y})_{y\in\bf Y})\mapsto(a_{x}b_{y})_{(x,y)\in{\bf X}\times{\bf Y}}.

We have a multiplication μ:𝕊𝒲𝕊𝕊\mu\colon\mathbb{S}\otimes_{\cal W}\mathbb{S}\to\mathbb{S} according to

μ(𝐚𝐛)=(k+m=nakbm)n0.\mu({\bf a}\otimes{\bf b})=\Big{(}\sum_{k+m=n}a_{k}b_{m}\Big{)}_{n\in\mathbb{N}_{0}}.

We write 𝟏:=(1,0,0,){\bf 1}:=(1,0,0,\dots) and set

X1:=X𝟏=(0,1,0,)(1,0,0,)𝕊𝒲𝕊,X_{1}:=X\otimes{\bf 1}=(0,1,0,\dots)\otimes(1,0,0,\dots)\in\mathbb{S}\otimes_{\cal W}\mathbb{S},

and

X2:=𝟏X=(1,0,0,)(0,1,0,)𝕊𝒲𝕊.X_{2}:={\bf 1}\otimes X=(1,0,0,\dots)\otimes(0,1,0,\dots)\in\mathbb{S}\otimes_{\cal W}\mathbb{S}.

So we obtain X1X2=X2X1X_{1}X_{2}=X_{2}X_{1} in 𝕊𝒲𝕊\mathbb{S}\otimes_{\cal W}\mathbb{S} and compute

(m0amX1m)(n0bnX2n)=m,n0ambnX1mX2n\Big{(}\sum_{m\in\mathbb{N}_{0}}a_{m}X_{1}^{m}\Big{)}\Big{(}\sum_{n\in\mathbb{N}_{0}}b_{n}X_{2}^{n}\Big{)}=\sum_{m,n\in\mathbb{N}_{0}}a_{m}b_{n}X_{1}^{m}X_{2}^{n}

for 𝐚=(am)m0{\bf a}=(a_{m})_{m\in\mathbb{N}_{0}} and 𝐛=(bn)n0{\bf b}=(b_{n})_{n\in\mathbb{N}_{0}} in 𝕊\mathbb{S}. Thus

𝕊𝒲𝕊={(m,n)0×0cmnX1mX2n:cmn𝕂}\mathbb{S}\otimes_{\cal W}\mathbb{S}=\Big{\{}\sum\nolimits_{(m,n)\in\mathbb{N}_{0}\times\mathbb{N}_{0}}c_{mn}X_{1}^{m}X_{2}^{n}:c_{mn}\in\mathbb{K}\Big{\}}

is the ring of power series in two commuting variables. Write 𝐚=n0anXn𝕊{\bf a}=\sum_{n\in\mathbb{N}_{0}}a_{n}X^{n}\in\mathbb{S} and note 𝐚𝟏=m0amX1m{\bf a}\otimes{\bf 1}=\sum_{m\in\mathbb{N}_{0}}a_{m}X_{1}^{m} and 𝟏𝐚=n0anX2n{\bf 1}\otimes{\bf a}=\sum_{n\in\mathbb{N}_{0}}a_{n}X_{2}^{n} in 𝕊𝒲𝕊\mathbb{S}\otimes_{\cal W}\mathbb{S}. We then have two morphisms of vector spaces

Δ,γ:𝕊𝕊𝒲𝕊 as follows:\Delta,\gamma\colon\mathbb{S}\to\mathbb{S}\otimes_{\cal W}\mathbb{S}\mbox{ as follows:}

For 𝐚=n0anXn{\bf a}=\sum_{n\in\mathbb{N}_{0}}a_{n}X^{n}

Δ(𝐚):=𝐚𝟏+𝟏𝐚=m,n0(amX1m+anX2n)\Delta({\bf a}):={\bf a}\otimes{\bf 1}+{\bf 1}\otimes{\bf a}=\sum_{m,n\in\mathbb{N}_{0}}(a_{m}X_{1}^{m}+a_{n}X_{2}^{n}) (2)

and

γ(𝐚):=γ(n0anXn)=n0an(X1+X2)n,\gamma({\bf a}):=\gamma\Big{(}\sum_{n\in\mathbb{N}_{0}}a_{n}{\cdot}X^{n}\Big{)}=\sum_{n\in\mathbb{N}_{0}}a_{n}(X_{1}+X_{2})^{n}, (3)

where γ\gamma is in fact a morphism of weakly complete algebras. Also, there is an identity ϵ:𝕂𝕊\epsilon\colon\mathbb{K}\to\mathbb{S}, ϵ(t)=t𝟏.\epsilon(t)=t{\cdot}{\bf 1}. and a coidentity (or augmentation κ:𝕊𝕂\kappa\colon\mathbb{S}\to\mathbb{K} given by κ(𝐚)=κ((an)n0):=a0\kappa({\bf a})=\kappa((a_{n})_{n\in\mathbb{N}_{0}}):=a_{0}, and a symmetry σ:𝕊𝕊\sigma\colon\mathbb{S}\to\mathbb{S} given by σ(X):=X\sigma(X):=-X, that is σ(𝐚)=σ((an)n0):=((1)nan)n\sigma({\bf a})=\sigma((a_{n})_{n\in\mathbb{N}_{0}}):=((-1)^{n}a_{n})_{n\in\mathbb{N}}. The following diagram is commutative:

𝕊𝒲𝕊σid𝕊𝕊γμ𝕊κϵ𝕊.\begin{matrix}\mathbb{S}\otimes_{\cal W}\mathbb{S}&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\sigma\otimes\mathop{\rm id}\nolimits}}&\mathbb{S}\otimes\mathbb{S}\\ \hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle\gamma$}}$}\Big{\uparrow}&&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle\mu$}}$\hss}\\ \mathbb{S}&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\kappa\circ\epsilon}}&\mathbb{S}.\end{matrix}

Thus 𝕊\mathbb{S} is a weakly complete commutative symmetric Hopf algebra.

Let us discuss its primitive and grouplike elements:

An element 𝐚=(an)n0𝕊{\bf a}=(a_{n})_{n\in\mathbb{N}_{0}}\in\mathbb{S} is primitive if and only if Δ(𝐚)=γ(𝐚)\Delta({\bf a})=\gamma({\bf a}), that is, by (2) and (3), if and only if

m,n0(amX1m+anX2n)=n0an(X1+X2)n,\sum_{m,n\in\mathbb{N}_{0}}(a_{m}X_{1}^{m}+a_{n}X_{2}^{n})=\sum_{n\in\mathbb{N}_{0}}a_{n}(X_{1}+X_{2})^{n}, (4)

if and only if n1an=0n\neq 1\implies a_{n}=0 if and only if 𝐚=tX{\bf a}=t{\cdot}X for some t𝕂t\in\mathbb{K}. Thus

(𝕊)=𝕂X.\mathbb{P}(\mathbb{S})=\mathbb{K}{\cdot}X. (PR)

An element 𝐚{\bf a} is grouplike if and only if it is nonzero and satisfies γ(𝐚)=𝐚𝐚\gamma({\bf a})={\bf a}\otimes{\bf a} if and only if

nN0an(X1+X2)n=m,n0amX1manX2n,\sum_{n\in N_{0}}\,a_{n}(X_{1}+X_{2})^{n}=\sum_{m,n\in\mathbb{N}_{0}}a_{m}X_{1}^{m}{\cdot}a_{n}X_{2}^{n}, (5)

which is the case if and only if

(n0)an=1n!.(\forall n\in\mathbb{N}_{0})\,a_{n}=\frac{1}{n!}. (6)

Thus

𝔾(𝕊)=exp𝕂X={expX,if 𝕂=,(expX)(expiX),if 𝕂=.\mathbb{G}(\mathbb{S})=\exp\mathbb{K}{\cdot}X=\begin{cases}\exp{\mathbb{R}{\cdot}X},\quad\mbox{if $\mathbb{K}=\mathbb{R}$},\\ (\exp\mathbb{R}{\cdot}X)(\exp\mathbb{R}i{\cdot}X),\quad\mbox{if $\mathbb{K}=\mathbb{C}$}.\end{cases} (GR)

This confirms the general result in [2], Theorem 6.15.


It is now urgent that we precisely describe the exponential function exp:𝕊𝕊×\exp\colon\mathbb{S}\to\mathbb{S}^{\times}:

For any field 𝕂\mathbb{K} we write 𝕂×\mathbb{K}^{\times} for 𝕂{0}\mathbb{K}\setminus\{0\}. We set 𝐈:={𝐚:a0=0}{\bf I}:=\{{\bf a}:a_{0}=0\}. Then 𝐈{\bf I} is the maximal ideal of 𝕊\mathbb{S} with 𝕊/𝐈𝕂\mathbb{S}/{\bf I}\cong\mathbb{K}. Notice that 𝕊=𝕂𝟏𝐈\mathbb{S}=\mathbb{K}{\cdot}{\bf 1}\oplus{\bf I} and exp(t𝟏+x)=(expt)(expx)\exp(t{\cdot}{\bf 1}+x)=(\exp t){\cdot}(\exp x) for t𝕂t\in\mathbb{K} and x𝐈x\in{\bf I}. In particular,

exp𝕊=(exp𝕂)exp(𝟏+𝐈), whereexp𝕂={<={r:0<r}if 𝕂=,×if 𝕂=.\exp\mathbb{S}=(\exp\mathbb{K}){\cdot}\exp({\bf 1}+{\bf I}),\mbox{ where}\exp\mathbb{K}=\begin{cases}\mathbb{R}_{<}=\{r\in\mathbb{R}:0<r\}&\mbox{if $\mathbb{K}=\mathbb{R}$},\\ \mathbb{C}^{\times}&\mbox{if $\mathbb{K}=\mathbb{C}$}.\end{cases}
{Lemma}

(i) For 𝐚=(an)n𝕊{\bf a}=(a_{n})_{n\in\mathbb{N}}\in\mathbb{S} we have 𝐚𝕊×{\bf a}\in\mathbb{S}^{\times} if and only if a00a_{0}\neq 0 and so 𝕊×=𝕊𝐈\mathbb{S}^{\times}=\mathbb{S}\setminus{\bf I}.

(ii) The function exp:𝕂𝕂×\exp\colon\mathbb{K}\to\mathbb{K}^{\times} maps \mathbb{R} bijectively onto <\mathbb{R}_{<} if 𝕂=\mathbb{K}=\mathbb{R} and \mathbb{C} surjectively onto ×\mathbb{C}^{\times} with kernel 2πi2\pi i\mathbb{Z} if 𝕂=\mathbb{K}=\mathbb{C}.

The function exp:𝐈𝟏+𝐈\exp\colon{\bf I}\to{\bf 1}+{\bf I} is bijective with inverse log:𝟏+𝐈𝐈\log\colon{\bf 1}+{\bf I}\to{\bf I}.

(iii) exp:𝕊=𝕂𝟏𝐈𝕊×=𝕂××(1+𝐈)\exp\colon\mathbb{S}=\mathbb{K}{\cdot}{\bf 1}\oplus{\bf I}\to\mathbb{S}^{\times}=\mathbb{K}^{\times}\times(1+{\bf I}) is injective for 𝕂=\mathbb{K}=\mathbb{R} and surjective for 𝕂=\mathbb{K}=\mathbb{C}.

(iv) The function exp:(𝕊)𝔾(𝕊)\exp\colon\mathbb{P}(\mathbb{S})\to\mathbb{G}(\mathbb{S}) is bijective, the inverse function being the logarithm log\log.

{Proof}

(i) If t:=a00t:=a_{0}\neq 0 then 𝐚=t(𝟏Y){\bf a}=t{\cdot}({\bf 1}-Y) where Y=(0,a1/t,a2/t,)Y=(0,-a_{1}/t,a_{2}/t,\dots) and (𝟏Y)1=𝟏+Y+Y2+({\bf 1}-Y)^{-1}={\bf 1}+Y+Y^{2}+\cdots. So 𝐚{\bf a} is invertible. On the other hand, if a0=0a_{0}=0, then 𝐚𝐛=(0,b0,)=(0,){\bf a}{\bf b}=(0,b_{0},\dots)=(0,\dots) and so 𝐚{\bf a} fails to be invertible.

(ii) and (iii) were shown above.

(iv) This is immediate from the preceding, since (𝕊)=𝕂X\mathbb{P}(\mathbb{S})=\mathbb{K}{\cdot}X and 𝔾(𝕊)=exp((𝕊))\mathbb{G}(\mathbb{S})=\exp(\mathbb{P}(\mathbb{S})).

We summarize our results on the power series algebra in one variable:

{Proposition}

(The power series algebra 𝕂[[X]]\mathbb{K}[[X]]) (i) The weakly complete power series algebra 𝕊:=𝕂[[X]]=(𝕂0,+,)\mathbb{S}:=\mathbb{K}[[X]]=(\mathbb{K}^{\mathbb{N}_{0}},+,*) is a singly generated weakly complete symmetric Hopf algebra generated by the element XX.

(ii) 𝕊\mathbb{S} is a local weakly complete algebra with maximal ideal

𝐈={n=1anXn:𝐚=(0,a1,a2,)𝕂0} and 𝕊×=𝕊𝐈.{\bf I}=\Big{\{}\sum_{n=1}^{\infty}a_{n}X^{n}:{\bf a}{=}(0,a_{1},a_{2},\dots){\in}\mathbb{K}^{\mathbb{N}_{0}}\Big{\}}\mbox{ and }\mathbb{S}^{\times}{=}\mathbb{S}\setminus{\bf I}.

Further,

𝕊=𝕂𝟏𝐈 and (t𝕂,x𝐈)exp(t𝟏+x)=etexpx.\mathbb{S}=\mathbb{K}{\cdot}{\bf 1}\oplus{\bf I}\mbox{ and }(\forall t\in\mathbb{K},x\in{\bf I})\,\exp(t{\cdot}{\bf 1}+x)=e^{t}{\cdot}\exp x.

(iii) exp:𝐈𝟏+𝐈\exp\colon{\bf I}\to{\bf 1}+{\bf I} has the inverse log\log and therefore implements an isomorphism of pro-Lie groups (𝐈,+)(𝟏+𝐈,)({\bf I},+)\cong({\bf 1}+{\bf I},{\cdot})

(iv) The additive group (𝕊)\mathbb{P}(\mathbb{S}) of primitive elements is 𝕂X\mathbb{K}\cdot X, the multiplicative group 𝔾(A)\mathbb{G}(A) of grouplike elements is (exp𝕂X,)(\exp\mathbb{K}{\cdot}X,*).

Recall that on 𝕂\mathbb{K} (with 𝕋=/\mathbb{T}=\mathbb{R}/\mathbb{Z}) we have

exp𝕂={<={r:0<r}(,+), if 𝕂=, and×(𝕋,+), if 𝕂=.\exp\mathbb{K}=\begin{cases}\mathbb{R}_{<}=\{r\in\mathbb{R}:0<r\}\cong(\mathbb{R},+),\mbox{ if $\mathbb{K}=\mathbb{R}$, and}\\ \mathbb{C}^{\times}\cong(\mathbb{R}\oplus\mathbb{T},+),\mbox{ if $\mathbb{K}=\mathbb{C}$}.\end{cases}

By Proposition 5.1 we have the following isomorphisms of abelian pro-Lie groups

(𝕊,+)=𝕂1𝐈(𝕂,+)×(𝐈,+)(𝕂,+)0,(\mathbb{S},+)=\mathbb{K}{\cdot}1\oplus{\bf I}\cong(\mathbb{K},+)\times({\bf I},+)\cong(\mathbb{K},+)^{\mathbb{N}_{0}},

and

(𝕊×,)=(𝕊𝐈,)=(𝕂×,)(𝟏+𝐈,)(𝕂×,)×(𝕂,+),(\mathbb{S}^{\times},{\cdot}){=}(\mathbb{S}\setminus{\bf I},{\cdot}){=}(\mathbb{K}^{\times},{\cdot}){\cdot}({\bf 1}+{\bf I},{\cdot}){\cong}(\mathbb{K}^{\times},{\cdot})\times(\mathbb{K},+)^{\mathbb{N}},
(e𝕂,){(<,), if 𝕂=,(×,), if 𝕂=.(e^{\mathbb{K}},{\cdot}){\cong}\begin{cases}(\mathbb{R}_{<},{\cdot}),\mbox{ if $\mathbb{K}{=}\mathbb{R}$},\\ (\mathbb{C}^{\times},\cdot),\mbox{ if $\mathbb{K}{=}\mathbb{C}$}.\end{cases}

Note also that on the level of primitive and grouplike elements we have simply

(𝕊)=𝕂X(𝕂,+),𝔾(𝕊)=exp(𝕂X){(,+),𝕂=,×𝕋,𝕂=..\mathbb{P}(\mathbb{S})=\mathbb{K}{\cdot}X\cong(\mathbb{K},+),\quad\mathbb{G}(\mathbb{S})=\exp(\mathbb{K}{\cdot}X)\cong\begin{cases}(\mathbb{R},+),\quad\mbox{$\mathbb{K}=\mathbb{R}$},\\ \mathbb{R}\times\mathbb{T},\quad\mbox{$\mathbb{K}=\mathbb{C}$}.\end{cases}.

Here one should keep in mind the example of the power series algebra over 𝕂\mathbb{K}:

𝕊=𝕂[[X]](𝕂0in𝒲).\mathbb{S}=\mathbb{K}[[X]]\quad(\cong\mathbb{K}^{\mathbb{N}_{0}}\quad\mbox{in}\ {\cal W}).

5.2 The universal monothetic algebra

We know from [2] that there is a singly generated universal weakly complete algebra 𝕂X=𝐔𝕂(𝔤)\mathbb{K}\langle X\rangle=\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{K}}(\mathfrak{g}), where 𝔤=𝕂\mathfrak{g}=\mathbb{K} is the one-dimensional Lie algebra. At his point we shall also discuss the weakly complete symmetric Hopf algebra structure of 𝕂X\mathbb{K}\langle X\rangle For the following we refer to [2], Corollary 3.3ff. The defining fact of 𝕂X\mathbb{K}\langle X\rangle is the following universal property:

\bulletFor each weakly complete unital algebra AA in 𝒲𝒜{\cal W}\hskip-2.0pt{\cal A} and each element aAa\in A there is a unique 𝒲𝒜{\cal W}\hskip-2.0pt{\cal A}-morphism φ:𝕂XA\varphi\colon\mathbb{K}\langle X\rangle\to A such that φ(X)=a\varphi(X)=a.

We shall see that the internal structure of 𝕂X\mathbb{K}\langle X\rangle is more complicated overall than one might expect initially.

It is clear that without loss of generality we may assume that AA is abelian. By Theorem 7.1, we have AlimI𝒥(A)A/IA\cong\lim_{I\in{\cal J}(A)}A/I and so \bullet holds if and only if it holds for all finite-dimensional commutative algebras AA. However, this universal property is satisfied exactly by the weakly complete algebra limJ𝐈(𝕂[x])𝕂[x]/J\lim_{J\in{\bf I}(\mathbb{K}[x])}\mathbb{K}[x]/J for the polynomial ring 𝕂[x]\mathbb{K}[x] in one variable xx over 𝕂\mathbb{K} and the filter basis of all of its ideals 𝐈(𝕂[x]){\bf I}(\mathbb{K}[x]). Since 𝕂[x]\mathbb{K}[x] is a principal ideal domain, every J𝐈(𝕂[x])J\in{\bf I}(\mathbb{K}[x]) is of the form J=(f)=f𝕂[x]J=(f)=f\mathbb{K}[x] for some polynomial f𝕂[x]f\in\mathbb{K}[x]. We may assume

𝕂X=limJ𝐈(𝕂[x])𝕂[x]J=limf𝕂[x]𝕂[x](f)f𝕂[x]𝕂[x]f𝕂[x],\mathbb{K}\langle X\rangle=\lim_{J\in{\bf I}(\mathbb{K}[x])}\frac{\mathbb{K}[x]}{J}=\lim_{f\in\mathbb{K}[x]}\frac{\mathbb{K}[x]}{(f)}\subseteq\prod_{f\in\mathbb{K}[x]}\frac{\mathbb{K}[x]}{f\mathbb{K}[x]}, (1)

generated by X:=(x+f𝕂[x])f𝕂[x]𝕂XX:=(x+f\mathbb{K}[x])_{f\in\mathbb{K}[x]}\in\mathbb{K}\langle X\rangle. (See also [2], Lemma 3.4.)

We let 𝔓=𝔓𝕂\mathfrak{P}=\mathfrak{P}_{\mathbb{K}} denote the set of the irreducible polynomials pp over 𝕂\mathbb{K} with leading coefficient 11 from the polynomial ring 𝕂[x]\mathbb{K}[x]. Then f𝕂[x]f\in\mathbb{K}[x], by the Chinese Remainder Theorem, is of the form

f=tp𝔓pkp some t𝕂 and (kp)p𝔓0(0),f=t{\cdot}\prod_{p\in\mathfrak{P}}p^{k_{p}}\mbox{ some $t\in\mathbb{K}$ and $(k_{p})_{p\in\mathfrak{P}}\in{\mathbb{N}_{0}}^{(\mathbb{N}_{0})}$},

where 0(0){\mathbb{N}_{0}}^{(\mathbb{N}_{0})} denotes the set of all families of nonnegative integers vanishing with the exception of indices pp from a finite subset of 𝔓\mathfrak{P}. Then for all f𝕂[x]f\in\mathbb{K}[x] we have.

𝕂[x](f)p𝔓𝕂[x](pkp).\frac{\mathbb{K}[x]}{(f)}\cong\prod_{p\in\mathfrak{P}}\frac{\mathbb{K}[x]}{(p^{k_{p}})}. (2)

Now {𝕂[x]/pn𝕂[x]:n}\{\mathbb{K}[x]/p^{n}\mathbb{K}[x]:n\in\mathbb{N}\} is a projective system and we introduce the notation

𝕊p:=limn𝕂[x]pn𝕂[x]\mathbb{S}_{p}:=\lim_{n\in\mathbb{N}}\frac{\mathbb{K}[x]}{p^{n}\mathbb{K}[x]} (3)

which is is a weakly complete commutative algebra generated by

Xp:=(x+pn𝕂[x])nlimn𝕂[x]pn𝕂[x].X_{p}:=(x+p^{n}\mathbb{K}[x])_{n\in\mathbb{N}}\in\lim_{n\in\mathbb{N}}\frac{\mathbb{K}[x]}{p^{n}\mathbb{K}[x]}.

In 𝕊p\mathbb{S}_{p} we have a maximal ideal 𝐈p:=Xp𝕊p{\bf I}_{p}:=X_{p}\mathbb{S}_{p} so that 𝕊p/𝐈p𝕂\mathbb{S}_{p}/{\bf I}_{p}\cong\mathbb{K} and

𝕊p=𝕂𝟏𝐈p.\mathbb{S}_{p}=\mathbb{K}{\cdot}{\bf 1}\oplus{\bf I}_{p}.

We conclude

𝕂X=p𝔓𝕂𝕊p,X=(Xp)p𝔓𝕂.\mathbb{K}\langle X\rangle=\prod_{p\in\mathfrak{P}_{\mathbb{K}}}\mathbb{S}_{p},\quad X=(X_{p})_{p\in\mathfrak{P}_{\mathbb{K}}}. (4)

In the interest of brevity again, we shall also write SS\SS in the place of 𝕂X\mathbb{K}\langle X\rangle. For easy reference we summarize the preceding discussion in the following lemma:

{Lemma}

(i) We have SS=p𝔓𝕂𝕊p\SS=\prod_{p\in\mathfrak{P}_{\mathbb{K}}}\mathbb{S}_{p}. For each p𝔓p\in\mathfrak{P} the algebra 𝕊p\mathbb{S}_{p} is generated algebraically and topologically by XpX_{p}, and SS\SS is generated algebraically and topologically by X=(Xp)p𝔓SSX=(X_{p})_{p\in\mathfrak{P}}\in\SS.

(ii) For each weakly complete algebra AA and each aAa\in A there is a unique 𝒲𝒜{\cal W}\hskip-2.0pt{\cal A}-morphism φa:SSA\varphi_{a}\colon\SS\to A such that φa(X)=a\varphi_{a}(X)=a.


The remainder of this section is devoted to a clarification of the structure of 𝕊p\mathbb{S}_{p} defined in (3).

{Lemma}

If p𝔓p\in\mathfrak{P} is of degree 1, then

𝕊p𝕂[[Xp]].\mathbb{S}_{p}\cong\mathbb{K}[[X_{p}]]. (5)
{Proof}

If c𝕂c\in\mathbb{K} and p=xcp=x-c, abbreviate R:=𝕂[x]𝕂[xc]R:=\mathbb{K}[x]\cong\mathbb{K}[x-c]. Then xxcx\mapsto x-c induces an automorphism of RR and 𝕊p=limnR(xc)nRlimnRxnR𝕂[[Xp]]\mathbb{S}_{p}=\lim_{n\in\mathbb{N}}\frac{R}{(x-c)^{n}R}\cong\lim_{n\in\mathbb{N}}\frac{R}{x^{n}R}\cong\mathbb{K}[[X_{p}]].

Since for 𝕂=\mathbb{K}=\mathbb{C} every p𝔓p\in\mathfrak{P}_{\mathbb{C}} is of degree 1, we know that in this case 𝕊p[[Xp]]\mathbb{S}_{p}\cong\mathbb{C}[[X_{p}]] for all pp\in\mathbb{P}.

Now we assume 𝕂=\mathbb{K}=\mathbb{R}\subseteq\mathbb{C}. Then there are two cases:

(a) p=xrp=x-r for rr\in\mathbb{R}. Then

𝕊p[[Xp]].\mathbb{S}_{p}\cong\mathbb{R}[[X_{p}]]. (5a)

(b) There is a cc\in\mathbb{C}\setminus\mathbb{R} such that

p(x)=(xc)(xc¯)=x2(c+c¯)x+cc¯=x22Re(c)x+|c|2,p𝔓.p(x)=(x-c)(x-\overline{c})=x^{2}-(c+\overline{c})x+c\overline{c}=x^{2}-2{\rm Re}(c)x+|c|^{2},\quad p\in\mathfrak{P}_{\mathbb{R}}.

In this case we write p1=xcp_{1}=x-c, Im c>0c>0, and p2=xc¯p_{2}=x-\overline{c}pn𝔓p_{n}\in\mathfrak{P}_{\mathbb{C}}, n=1,2n=1,2.

{Lemma}

In the case of (b) above, the real algebra [x]pn[x]\frac{\mathbb{R}[x]}{p^{n}\mathbb{R}[x]} is isomorphic to the real algebra underlying [x]p1n[x]\frac{\mathbb{C}[x]}{p_{1}^{n}\mathbb{C}[x]}. {Proof} We use the abbreviations Rn=[x]pn[x]R_{n}=\frac{\mathbb{R}[x]}{p^{n}\mathbb{R}[x]}, Cn=[x]pn[x]C_{n}=\frac{\mathbb{C}[x]}{p^{n}\mathbb{C}[x]}, and Ckn=[x]pkn[x]C_{kn}=\frac{\mathbb{C}[x]}{p_{k}^{n}\mathbb{C}[x]}, k=1,2k=1,2. By the Chinese Remainder Theorem we have an isomorphism

ρ:CnC1n×C2n\rho\colon C_{n}\to C_{1n}\times C_{2n} (i)

such that ρ(u+pn[x])=(u+p1n[x],u+p2n[x])\rho(u+p^{n}\mathbb{C}[x])=(u+p_{1}^{n}\mathbb{C}[x],u+p_{2}^{n}\mathbb{C}[x]). The real algebra underlying the right hand side of (i) has an involution σ\sigma defined by

σ(u+p1n[x],v+p2n[x])=(v¯+p1n[x],u¯+p2n[x]),\sigma(u+p_{1}^{n}\mathbb{C}[x],v+p_{2}^{n}\mathbb{C}[x])=(\overline{v}+p_{1}^{n}\mathbb{C}[x],\overline{u}+p_{2}^{n}\mathbb{C}[x]), (ii)

such that the elements of the real fixed point algebra FF of σ\sigma are the elements (u+p1n[x],u¯+p2n[x])(u+p_{1}^{n}\mathbb{C}[x],\overline{u}+p_{2}^{n}\mathbb{C}[x]) with u[x]u\in\mathbb{C}[x]. The restriction of the projection C1n×C2nC1nC_{1n}\times C_{2n}\to C_{1n} to FF is an isomorphism. Thus FC1nF\cong C_{1n} as real algebra. In particular, dimF=dimC1n=2n=dimRn\dim_{\mathbb{R}}F=\dim_{\mathbb{R}}C_{1n}=2n=\dim_{\mathbb{R}}R_{n}. Thus the injection RnFR_{n}\to F via ρ\rho is in fact surjective. Hence [x]pn[x]=RnF=[x]p1n[x]\frac{\mathbb{R}[x]}{p^{n}\mathbb{R}[x]}=R_{n}\cong F=\frac{\mathbb{C}[x]}{p_{1}^{n}\mathbb{C}[x]} as real algebras.

As a consequence we conclude that

𝕊p=limn[x]pn[x]limn[x]p1n[x]\mathbb{S}_{p}=\lim_{n\in\mathbb{N}}\frac{\mathbb{R}[x]}{p^{n}\mathbb{R}[x]}\cong\lim_{n\in\mathbb{N}}\frac{\mathbb{C}[x]}{p_{1}^{n}\mathbb{C}[x]}

and thus

𝕊p[[Xp1]] as real algebras.\mathbb{S}_{p}\cong\mathbb{C}[[X_{p_{1}}]]\mbox{ as real algebras}. (5b)

For 𝕂=\mathbb{K}=\mathbb{R} or 𝕂=\mathbb{K}=\mathbb{C}, there is an injection pcp:𝔓𝕂p\mapsto c_{p}:\mathfrak{P}_{\mathbb{K}}\to\mathbb{C}, where

p(x)={xcp,if deg(p)=1,xcp1,Imcp1>0,if deg(p)=2.p(x)=\begin{cases}x-c_{p},\quad\mbox{if $\deg(p)=1$,}\\ x-c_{p_{1}},\ {\rm Im\ }c_{p_{1}}>0,\quad\mbox{if $\deg(p)=2$.}\end{cases} (6)

If 𝕂=\mathbb{K}=\mathbb{C}, then we are in the first case and cpc_{p} ranges through all of \mathbb{C}. If 𝕂=\mathbb{K}=\mathbb{R}, then both cases occur, and in the first case cpc_{p} ranges through 𝕂=\mathbb{K}=\mathbb{R} and in the second case cp1c_{p_{1}} ranges through the open upper complex half-plane.

The different cases now sum up to the following statement:

{Lemma}
𝕂X=p𝔓𝕂𝕊p{p𝔓,degp=1[[Xp]]×p𝔓,degp=2[[Xp1]],if 𝕂=,p𝔓[[Xp]],if 𝕂=,\mathbb{K}\langle X\rangle{=}\prod_{p\in\mathfrak{P}_{\mathbb{K}}}\mathbb{S}_{p}{\cong}\begin{cases}\prod_{p\in\mathfrak{P}_{\mathbb{R}},\deg p=1}\mathbb{R}[[X_{p}]]\times\prod_{p\in\mathfrak{P}_{\mathbb{R}},\deg p=2}\mathbb{C}[[X_{p_{1}}]],\,\mbox{if $\mathbb{K}=\mathbb{R}$},\\ \prod_{p\in\mathfrak{P}_{\mathbb{C}}}\mathbb{C}[[X_{p}]],\quad\mbox{if $\mathbb{K}=\mathbb{C}$},\end{cases} (6′′)

where all algebras in the top line are real algebras.

For p𝔓𝕂p\in\mathfrak{P}_{\mathbb{K}} we write

𝕂p={,if 𝕂= and degp=1,,if either 𝕂= and degp=2, or 𝕂=.\mathbb{K}_{p}=\begin{cases}\mathbb{R},\quad\mbox{if $\mathbb{K}=\mathbb{R}$ and $\deg p=1$,}\\ \mathbb{C},\quad\mbox{if either $\mathbb{K}=\mathbb{R}$ and $\deg p=2$, or $\mathbb{K}=\mathbb{C}$.}\end{cases} ()

Moreover, elements in 𝕊p\mathbb{S}_{p} with p𝔓𝕂p\in\mathfrak{P}_{\mathbb{K}} we denote by

𝐚p=n0anpXpn,anp𝕂p.{\bf a}_{p}=\sum_{n\in\mathbb{N}_{0}}a_{np}{X_{p}}^{n},\quad a_{np}\in\mathbb{K}_{p}.

Finally we have

SS:=𝕂X=p𝔓𝕂𝕊p={𝔞:=(𝐚p)p𝔓𝕂:𝐚p𝕊p},\SS:=\mathbb{K}\langle X\rangle=\prod_{p\in\mathfrak{P}_{\mathbb{K}}}\mathbb{S}_{p}=\{{\mathfrak{a}}:=({\bf a}_{p})_{p\in\mathfrak{P}_{\mathbb{K}}}:{\bf a}_{p}\in\mathbb{S}_{p}\}, (7)

a weakly complete commutative symmetric Hopf algebra, with componentwise operations and co-operations. By Lemma 5.2(i), the weakly complete algebra SS\SS is the algebraically and topologically singly generated weakly complete algebra with generator

X:=(Xp)p𝔓𝕂p𝔓𝕂𝕊p,X:=(X_{p})_{p\in\mathfrak{P}_{\mathbb{K}}}\in\prod_{p\in\mathfrak{P}_{\mathbb{K}}}\mathbb{S}_{p}, (7X)

For each p𝔓𝕂p\in\mathfrak{P}_{\mathbb{K}} we define 𝐈p𝕊p{\bf I}_{p}\subseteq\mathbb{S}_{p} to be the maximal ideal of 𝕊p\mathbb{S}_{p}, where exp:𝐈p1+𝐈p\exp\colon{\bf I}_{p}\to 1+{\bf I}_{p} has the inverse log:1+𝐈p𝐈p\log\colon 1+{\bf I}_{p}\to{\bf I}_{p}. In view of Lemma 5.1 we observe the following fact:

{Remark}

The exponential function exp:SSSS×\exp\colon\SS\to\SS^{\times} is given componentwise for 𝔞=(𝐚p)p𝔓{\mathfrak{a}}=({\bf a}_{p})_{p\in\mathfrak{P}} as

exp𝔞=(exp𝐚p)p𝔓.\exp{\mathfrak{a}}=(\exp{\bf a}_{p})_{p\in\mathfrak{P}}.

The exponential function exp:𝕊p𝕊p\exp\colon\mathbb{S}_{p}\to\mathbb{S}_{p} is surjective if either 𝕂=\mathbb{K}=\mathbb{R} and degp=2\deg p=2, or 𝕂=\mathbb{K}=\mathbb{C}, and it is injective if 𝕂=\mathbb{K}=\mathbb{R} and degp=1\deg p=1.


We now aim to discuss the restriction of the exponential function to the set of primitive elements. First recall that on SS=𝕂X\SS=\mathbb{K}\langle X\rangle we have the operations

𝔞+𝔟=(𝐚p+𝐛p)p𝔓 and 𝔞𝔟=(𝐚p𝐛p)p𝔓=(n0(k+m=nakpbmp)Xpn)p𝔓.{\mathfrak{a}}+{\mathfrak{b}}=({\bf a}_{p}+{\bf b}_{p})_{p\in\mathfrak{P}}\mbox{ and }{\mathfrak{a}}*{\mathfrak{b}}=({\bf a}_{p}*{\bf b}_{p})_{p\in\mathfrak{P}}=\bigg{(}\sum_{n\in\mathbb{N}_{0}}\Big{(}\sum\nolimits_{k+m=n}a_{kp}b_{mp}\Big{)}{X_{p}}^{n}\bigg{)}_{p\in\mathfrak{P}}.

In SS𝒲SS\SS\otimes_{\cal W}\SS, with 𝔓=𝔓𝕂\mathfrak{P}=\mathfrak{P}_{\mathbb{K}} we write

X1p:=(Xp)p𝔓𝟏=(Xp𝟏)p𝔓SS𝒲SS,X_{1p}:=(X_{p})_{p\in\mathfrak{P}}\otimes{\bf 1}=(X_{p}\otimes{\bf 1})_{p\in\mathfrak{P}}\in\SS\otimes_{\cal W}\SS,

and

X2p:=𝟏(Xp)p𝔓=(𝟏Xp)p𝔓SS𝒲SS,X_{2p}:={\bf 1}\otimes(X_{p})_{p\in\mathfrak{P}}=({\bf 1}\otimes X_{p})_{p\in\mathfrak{P}}\in\SS\otimes_{\cal W}\SS,

Again we may consider SS𝒲SS\SS\otimes_{\cal W}\SS as weakly complete power series algebra with commuting variables X1pX_{1p} and X2pX_{2p} with pp ranging through 𝔓\mathfrak{P}.


The identity and coidentity ϵ:𝕂SS\epsilon\colon\mathbb{K}\to\SS, κ:SS𝕂\kappa\colon\SS\to\mathbb{K} (augmentation), and symmetry σ:SSSS\sigma\colon\SS\to\SS are straightforward from the respective operations in 𝕊\mathbb{S}, but let us also consider the diagonal vector space morphism Δ\Delta and the algebra comultiplication γ\gamma:

Δ,γ:SSSS𝒲SS defined as follows:\Delta,\gamma:\SS\to\SS\otimes_{\cal W}\SS\mbox{\quad defined as follows:}

For 𝔞=(n,p)0×𝔓anpXpn{\mathfrak{a}}=\sum_{(n,p)\in\mathbb{N}_{0}\times\mathfrak{P}}a_{np}{X_{p}}^{n} we have

Δ(𝔞):=𝔞𝟏+𝟏𝔞=m,n0,f,g𝔓(ampX1pm+angX2gn)\Delta({\mathfrak{a}}){:=}{\mathfrak{a}}\otimes{\bf 1}+{\bf 1}\otimes{\mathfrak{a}}{=}\sum_{m,n\in\mathbb{N}_{0},f,g\in\mathfrak{P}}(a_{mp}{X_{1p}}^{m}{+}a_{ng}{X_{2g}}^{n}) (4)

and

γ(𝔞)=n0,p𝔓anp(X1p+X2p)n.\gamma({\mathfrak{a}})=\sum_{n\in\mathbb{N}_{0},p\in\mathfrak{P}}a_{np}(X_{1p}+X_{2p})^{n}. (5)

We have the commutative diagram

SS𝒲SSσ𝒲idSS𝒲SSγμSSκϵSS\begin{matrix}\SS\otimes_{\cal W}\SS&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\sigma\otimes_{\cal W}\mathop{\rm id}\nolimits}}&\SS\otimes_{\cal W}\SS\\ \hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle\gamma$}}$}\Big{\uparrow}&&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle\mu$}}$\hss}\\ \SS&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\kappa\circ\epsilon}}&\SS\end{matrix}

identifying SS\SS as weakly complete symmetric Hopf algebra allowing us now to turn to the determination of the primitive and grouplike elements of SS\SS. Indeed an element 𝔞{\mathfrak{a}} is primitive in SS\SS if γ(𝔞)=Δ(𝔞)\gamma({\mathfrak{a}})=\Delta({\mathfrak{a}}), that is, if and only if

m,n0,p𝔓(ampX1pm+anfX2pn)=n0,p𝔓anp(X1p+X2p)n\sum_{m,n\in\mathbb{N}_{0},\ p\in\mathfrak{P}}(a_{mp}{X_{1p}}^{m}+a_{nf}{X_{2p}}^{n})=\sum_{n\in\mathbb{N}_{0},\ p\in\mathfrak{P}}a_{np}(X_{1p}+X_{2p})^{n}

if and only if

(p𝔓)n1anp=0(\forall p\in\mathfrak{P})\,n\neq 1\implies a_{np}=0

if and only if

(p𝔓)(tp𝕂p)𝔞=(tpXp)p𝔓.(\forall p\in\mathfrak{P})(\exists t_{p}\in\mathbb{K}_{p})\,{\mathfrak{a}}=(t_{p}X_{p})_{p\in\mathfrak{P}}.

Thus we have

(SS)=p𝔓𝕂𝕂pXp.\mathbb{P}(\SS)=\prod_{p\in\mathfrak{P}_{\mathbb{K}}}\mathbb{K}_{p}\cdot X_{p}. (PR)

On the other hand, an element 𝔞{\mathfrak{a}} is grouplike if it is nonzero and satisfies γ(𝔞)=𝔞𝔞\gamma({\mathfrak{a}})={\mathfrak{a}}\otimes{\mathfrak{a}}, that is,

n0,p𝔓anp(X1p+X2p)n=m,n0,p𝔓am,pX1mangX2n,\sum_{n\in\mathbb{N}_{0},p\in\mathfrak{P}}a_{np}(X_{1p}+X_{2p})^{n}=\sum_{m,n\in\mathbb{N}_{0},p\in\mathfrak{P}}a_{m,p}X_{1}^{m}{\cdot}a_{ng}X_{2}^{n},

which is the case if and only if

((n,p)0×𝔓)anp=1n!.(\forall(n,p)\in\mathbb{N}_{0}\times\mathfrak{P})\ a_{np}=\frac{1}{n!}.

Thus we have

𝔾(SS)=p𝔓𝕂exp(𝕂pXp).\mathbb{G}(\SS)=\prod_{p\in\mathfrak{P}_{\mathbb{K}}}\exp(\mathbb{K}_{p}{\cdot}X_{p}). (GR)
{Definition}

The algebraically and topologically singly generated universal weakly complete algebra SS=𝕂X\SS=\mathbb{K}\langle X\rangle is called the universal monothetic algebra.

Lemma 5.2 (i) justifies the name. Recall from Proposition 5.1 that for each p𝔓𝕂p\in\mathfrak{P}_{\mathbb{K}} the algebra 𝕊p\mathbb{S}_{p} is a local weakly complete algebra with a maximal ideal 𝐈p{\bf I}_{p} and that 𝕊p×=𝕂p×(𝕊p𝐈p){\mathbb{S}_{p}}^{\times}=\mathbb{K}_{p}^{\times}{\cdot}(\mathbb{S}_{p}\setminus{\bf I}_{p}) where 𝕂p\mathbb{K}_{p} was defined in ()({\bf*}). Now we are prepared to summarize the structure theorem for SS\SS: {Theorem} (The universal monothetic algebra 𝕂X\mathbb{K}\langle X\rangle)

(i) The universal monothetic algebra

SS:=𝕂X=p𝔓𝕂𝕊p\SS:=\mathbb{K}\langle X\rangle=\prod_{p\in\mathfrak{P}_{\mathbb{K}}}\mathbb{S}_{p}

is a weakly complete symmetric Hopf algebra generated by the element X=(Xp)p𝔓𝕂X=(X_{p})_{p\in\mathfrak{P}_{\mathbb{K}}}.

(ii) The group of units SS×\SS^{\times} is dense in SS\SS, where SS×=p𝔓𝕂𝕊p×=\SS^{\times}=\prod_{p\in\mathfrak{P}_{\mathbb{K}}}{\mathbb{S}_{p}}^{\times}=

{(n0anpXpn)p𝔓𝕂:anp𝕂p and a0,p0}\Big{\{}\big{(}\sum\nolimits_{n\in\mathbb{N}_{0}}a_{np}{X_{p}}^{n}\big{)}_{p\in\mathfrak{P}_{\mathbb{K}}}:\ a_{np}\in\mathbb{K}_{p}\mbox{ and }a_{0,p}{\neq}0\Big{\}}
=p𝔓𝕂𝕂p×(𝕊p𝐈p).=\prod_{p\in\mathfrak{P}_{\mathbb{K}}}\mathbb{K}^{\times}_{p}(\mathbb{S}_{p}\setminus{\bf I}_{p}). (#)

(iii) The exponential function exp:SSSS×\exp\colon\SS\to\SS^{\times} operates componentwise on p𝔓𝕂𝕊p\prod_{p\in\mathfrak{P}_{\mathbb{K}}}\mathbb{S}_{p} and induces an isomorphism of of topological groups p𝔓K𝐈pp𝔓𝕂(1+𝐈p)\prod_{p\in\mathfrak{P}_{K}}{\bf I}_{p}\to\prod_{p\in\mathfrak{P}_{\mathbb{K}}}(1+{\bf I}_{p}), whose inverse is given by the componentwise logarithm.

(iv) The additive group (SS)\mathbb{P}(\SS) of primitive elements is

(SS)=p𝔓𝕂𝕂pXpp𝔓𝕂𝕊p.\mathbb{P}(\SS)=\prod_{p\in\mathfrak{P}_{\mathbb{K}}}\mathbb{K}_{p}\cdot X_{p}\subseteq\prod_{p\in\mathfrak{P}_{\mathbb{K}}}\mathbb{S}_{p}.

In particular, the element XX is primitive. The image of 𝔤\mathfrak{g} in 𝕂X\mathbb{K}\langle X\rangle is 𝕂X\mathbb{K}{\cdot}X, i.e. 𝔤(𝐔(𝔤))\mathfrak{g}\subset\mathbb{P}(\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})).

Let 𝕋=/\mathbb{T}=\mathbb{R}/\mathbb{Z} denote the additive circle group again and 𝔠\mathfrak{c} the cardinality 202^{\aleph_{0}} of the continuum. The multiplicative group 𝔾(SS)\mathbb{G}(\SS) of grouplike elements is

𝔾(SS)=p𝔓𝕂exp(𝕂pXp){𝔓𝕋{p𝔓:degp=2},if 𝕂=,(𝕋)𝔓,if 𝕂=}(×)𝔠.\mathbb{G}(\SS)=\prod_{p\in\mathfrak{P}_{\mathbb{K}}}\exp(\mathbb{K}_{p}{\cdot}X_{p})\cong\begin{cases}\mathbb{R}^{\mathfrak{P}_{\mathbb{R}}}\oplus\mathbb{T}^{\{p\in\mathfrak{P}_{\mathbb{R}}:\ \deg p=2\}},\,\mbox{if $\mathbb{K}=\mathbb{R}$},\\ (\mathbb{R}\oplus\mathbb{T})^{\mathfrak{P}_{\mathbb{C}}},\,\mbox{if $\mathbb{K}=\mathbb{C}$}\end{cases}\hskip-15.0pt\mbox{$\Bigg{\}}$}\cong(\mathbb{C}^{\times})^{\mathfrak{c}}.

The exponential function exp:(SS)𝔾(SS)\exp\colon\mathbb{P}(\SS)\to\mathbb{G}(\SS) is a quotient morphism of topological abelian groups onto its image.

In particular we derive the following (with 𝔠=20\mathfrak{c}=2^{\aleph_{0}}):

{Corollary}

The abelian pro-Lie group 𝔾(𝕂X)\mathbb{G}(\mathbb{K}\langle X\rangle) is connected and, for 𝕂=\mathbb{K}=\mathbb{R}, is isomorphic to 𝔠𝕋𝔠(×𝕋)𝔠(×)𝔠\mathbb{R}^{\mathfrak{c}}\oplus\mathbb{T}^{\mathfrak{c}}\cong(\mathbb{R}\times\mathbb{T})^{\mathfrak{c}}\cong(\mathbb{C}^{\times})^{\mathfrak{c}}.

We do not know whether in general the pro-Lie group 𝔾(𝐔(𝔤))\mathbb{G}(\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})) is connected. Theorem 5.2 (iv)) shows that 𝔤A:=𝐔𝕂(𝔤)\mathfrak{g}\subseteq A:=\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{K}}(\mathfrak{g}) is considerably smaller than (A)\mathbb{P}(A). The discrepancy between 𝔤\mathfrak{g} and (A)\mathbb{P}(A) arises in the detailed description of the universal monothetic algebra 𝕂X\mathbb{K}\langle X\rangle. The origin of this complication is the Galois theory of the polynomial ring 𝕂[x]\mathbb{K}[x].

5.3 Comments

We discussed extensively the “smallest possible” nontrivial weakly complete enveloping algebra 𝐔𝕂(𝔤)\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{K}}(\mathfrak{g}), namely, the one arising for dim𝔤=1\dim\mathfrak{g}=1. Any abelian profinite-dimensional Lie algebra 𝔤\mathfrak{g} is isomorphic to 𝕂J\mathbb{K}^{J} for some set JJ. We have a pair of adjoint functors between the category 𝒲{\cal W} of weakly complete vector spaces over 𝕂\mathbb{K} and the category 𝒲𝒜𝒞{\cal W}\hskip-2.0pt{\cal A}\hskip-0.5pt{\cal C} of weakly complete commutative unital algebras, namely, the functor A|A|:𝒲𝒜𝒞𝒲A\mapsto|A|:{\cal W}\hskip-2.0pt{\cal A}\hskip-0.5pt{\cal C}\to{\cal W} assigning to a weakly complete commutative algebra its underlying weakly complete vector space and 𝔤𝐔𝕂(𝔤):𝒲𝒲𝒜𝒞\mathfrak{g}\mapsto\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{K}}(\mathfrak{g}):{\cal W}\to{\cal W}\hskip-2.0pt{\cal A}\hskip-0.5pt{\cal C}, the restriction of the universal enveloping functor. Then 𝐔𝕂|𝒲\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{K}}|{\cal W} is left adjoint to |||\cdot|, and therefore it preserves colimits. For a finite set JJ of nn elements we note that in the category 𝒲{\cal W} we have 𝔤=𝔤1𝔤n\mathfrak{g}=\mathfrak{g}_{1}\oplus\cdots\oplus\mathfrak{g}_{n} with 𝔤k𝕂\mathfrak{g}_{k}\cong\mathbb{K} for k=1,,nk=1,\dots,n and so 𝔤\mathfrak{g} is the coproduct of nn cofactors of dimension 1. Accordingly, in the category 𝒲𝒜𝒞{\cal W}\hskip-2.0pt{\cal A}\hskip-0.5pt{\cal C} we observe

𝐔𝕂(𝔤)j=1n𝐔𝕂(𝔤j),𝐔K(𝔤j)SS𝕊𝔓𝕂.\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{K}}(\mathfrak{g})\cong\coprod_{j=1}^{n}\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{K}}(\mathfrak{g}_{j}),\quad\mathop{\bf U\hphantom{}}\nolimits_{K}(\mathfrak{g}_{j})\cong\SS\cong\mathbb{S}^{\mathfrak{P}_{\mathbb{K}}}. (1)

To the extent that finite coproducts in the category 𝒲𝒜𝒞{\cal W}\hskip-2.0pt{\cal A}\hskip-0.5pt{\cal C} are understood, one knows 𝐔𝕂(𝔤)\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{K}}(\mathfrak{g}) for finite dimensional abelian Lie algebras 𝔤\mathfrak{g}.

Let now 𝔤=𝕂J\mathfrak{g}=\mathbb{K}^{J} in 𝒲{\cal W} for some nonempty set JJ. If JJ is finite, then 𝔤\mathfrak{g} is a finite coproduct and the dual of a finite product. If JJ is infinite, then let JfinJ_{\rm fin} denote the directed family of finite subsets FJF\subseteq J and recall

𝔤=𝕂JlimFJfin𝕂F.\mathfrak{g}=\mathbb{K}^{J}\cong\lim_{F\in J_{\rm fin}}\mathbb{K}^{F}.

We may then apply Theorem 3 and deduce

𝐔𝕂(𝔤)limFJfin𝐔𝕂(𝕂F),\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{K}}(\mathfrak{g})\cong\lim_{F\in J_{\rm fin}}\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{K}}(\mathbb{K}^{F}), (2)

where 𝐔𝕂(𝕂F)\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{K}}(\mathbb{K}^{F}) is known by (1) if we know finite coproducts in the category 𝒲𝒜𝒞{\cal W}\hskip-2.0pt{\cal A}\hskip-0.5pt{\cal C}.

In Theorem 8 in the appendix we argued that for any profinite-dimensional Lie algebra 𝔤\mathfrak{g} with the underlying weakly complete vector space |𝔤||\mathfrak{g}| there is a canononical quotient morphism of weakly complete unital algebras q𝔤:𝐓(|𝔤|)𝐔(𝔤)q_{\mathfrak{g}}\colon\mathop{\bf T\hphantom{}}\nolimits(|\mathfrak{g}|)\to\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}) from the weakly complete tensor algebra 𝐓(|𝔤|)\mathop{\bf T\hphantom{}}\nolimits(|\mathfrak{g}|) onto 𝐔(𝔤)\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}). In the present situation of abelian Lie algebras we may write |𝔤|=𝔤|\mathfrak{g}|=\mathfrak{g} and conclude that for each set JJ and for 𝔤=𝕂J\mathfrak{g}=\mathbb{K}^{J} we have a natural quotient morphism of weakly complete algebras

q𝔤:𝐓𝕂(𝔤)𝐔𝕂(𝔤)q_{\mathfrak{g}}\colon\mathop{\bf T\hphantom{}}\nolimits_{\mathbb{K}}(\mathfrak{g})\to\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{K}}(\mathfrak{g}) (3)

whose kernel is the closed ideal generated in 𝐓𝕂(𝔤)\mathop{\bf T\hphantom{}}\nolimits_{\mathbb{K}}(\mathfrak{g}) by the elements xyyxxy-yx, x,y𝔤𝐓(𝔤)x,y\in\mathfrak{g}\subseteq\mathop{\bf T\hphantom{}}\nolimits(\mathfrak{g}). The structure theory of SS\SS for the case 𝔤=𝕂\mathfrak{g}=\mathbb{K} shows that the presentation (3) conceals more than it reveals. Indeed, the universal property of 𝐔𝕂(𝕂J)\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{K}}(\mathbb{K}^{J}) yields for 𝔤=𝕂J\mathfrak{g}=\mathbb{K}^{J} a surjective morphism of weakly complete unital algebras

A:=𝐔𝕂(𝔤)=𝐔𝕂(𝕂J)𝐔𝕂(𝕂)JSSJ(𝕂0×𝔓𝕂×J in 𝒲).A:=\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{K}}(\mathfrak{g})=\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{K}}(\mathbb{K}^{J})\to\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{K}}(\mathbb{K})^{J}\cong\SS^{J}\quad(\cong\mathbb{K}^{\mathbb{N}_{0}\times\mathfrak{P}_{\mathbb{K}}\times J}\mbox{ in }{\cal W}). (4)

Since (SSJ)=(SS)J𝕂𝔓×J\mathbb{P}(\SS^{J})=\mathbb{P}(\SS)^{J}\cong\mathbb{K}^{\mathfrak{P}\times J} the quotient morphism in (4) shows that the vector space of primitive elements of SSJ\SS^{J} is considerably larger than 𝔤=𝕂J\mathfrak{g}=\mathbb{K}^{J}. This information indicates that for the weakly complete symmetric Hopf algebra A:=𝐔(𝔤)A:=\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}), the subspace (A)\mathbb{P}(A) of primitive elements is likely to be large by comparison with 𝔤\mathfrak{g}. Clearly 𝔾(SSJ)\mathbb{G}(\SS^{J}) is a connected abelian group whose structure is known by Corollary 5.2. The simplest example along this line is the power series Hopf algebra 𝕂[[X]]\mathbb{K}[[X]] (cf. Proposition 5.1). Accordingly, one expects the group 𝔾(A)\mathbb{G}(A) to be considerable.

However, some caution is in order:

{Example}

(i) Let A=[^]A=\mathbb{R}[\widehat{\mathbb{Q}}] be the real group algebra of ^\widehat{\mathbb{Q}}, the universal solenoidal compact abelian group. Then 𝔾(A)=^A\mathbb{G}(A)=\widehat{\mathbb{Q}}\subseteq A, while (A)=𝔏(^)\mathbb{P}(A)=\mathfrak{L}(\widehat{\mathbb{Q}})\cong\mathbb{R}. Then exp^:(A)𝔾(A)\exp_{\widehat{\mathbb{Q}}}\colon\mathbb{P}(A)\to\mathbb{G}(A) is a morphism of locally compact abelian groups with a dense image, but it is not surjective.

(ii) If we take A=[p]A=\mathbb{R}[\mathbb{Z}_{p}] for a prime number pp, where p\mathbb{Z}_{p} is the additive group of the pp-adic integers, then 𝔾(A)p\mathbb{G}(A)\cong\mathbb{Z}_{p} and (A)=𝔏(𝔾(A))={0}\mathbb{P}(A)=\mathfrak{L}(\mathbb{G}(A))=\{0\}, since p\mathbb{Z}_{p} is totally disconnected. The exponential map expA:(A)𝔾(A)\exp_{A}\colon\mathbb{P}(A)\to\mathbb{G}(A) is the zero morphism.

These examples show that even on the abelian level, the weakly complete symmetric Hopf algebra structure of the weakly complete enveloping algebras and that of the weakly complete group algebras behave rather differently. Yet they are related in a natural way as we shall observe in the following section.

6 Enveloping Algebras Versus Group Algebras

The class of compact groups and their Lie algebras are distinguished domains for which the relationship between weakly complete enveloping algebras and weakly complete group algebras is particularly lucid. Hence we focus on these classes.

6.1 The case of compact groups

A particularly appropriate situation is that of a compact topological group GG. Our level of information regarding the associated real group algebra is particularly advanced in that situation. Indeed recall that for a compact group we may naturally identify GG with the group of grouplike elements of [G]\mathbb{R}[G] (cf. [2], Theorems 8.7, 8.9 and 8.12), and we may further identify 𝔤:=𝔏(G)\mathfrak{g}:=\mathfrak{L}(G) with the pro-Lie algebra ([G])\mathbb{P}(\mathbb{R}[G]) of primitive elements. (Cf.  also Theorem 9 in the Appendix.) We may also assume that the Lie algebra 𝔤\mathfrak{g} of GG is contained in the set (𝐔(𝔤))\mathbb{P}(\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{R}}(\mathfrak{g})) of primitive elements of 𝐔(𝔤)\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{R}}(\mathfrak{g}).

{Theorem}

(i) Let GG be a compact group and 𝔤\mathfrak{g} its Lie algebra. Then there is a natural morphism of weakly complete algebras ωG:𝐔(𝔤)[G]\omega_{G}\colon\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{R}}(\mathfrak{g})\to\mathbb{R}[G] fixing the elements of 𝔤\mathfrak{g} elementwise.

(ii) The image of ωG\omega_{G} is the closed subalgebra [G0]\mathbb{R}[G_{0}] of [G]\mathbb{R}[G].

(iii) The pro-Lie group 𝔾(𝐔(𝔤))\mathbb{G}(\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{R}}(\mathfrak{g})) is mapped into G0=𝔾([G0])[G]G_{0}=\mathbb{G}(\mathbb{R}[G_{0}])\subseteq\mathbb{R}[G]. The connected pro-Lie group 𝔾(𝐔(𝔤))0\mathbb{G}(\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{R}}(\mathfrak{g}))_{0} maps epimorphically to G0G_{0} and (𝐔(𝔤))\mathbb{P}(\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{R}}(\mathfrak{g})) maps surjectively onto ([G])=𝔤\mathbb{P}(\mathbb{R}[G])=\mathfrak{g}.

{Proof}

(i) follows at once from the universal property of 𝐔\mathop{\bf U\hphantom{}}\nolimits.

(ii) As a morphism of weakly complete Hopf algebras, ωG\omega_{G} has a closed image which is generated as a weakly complete subalgebra by 𝔤\mathfrak{g} which is [G0]\mathbb{R}[G_{0}] by Corollary 3.3 (ii) of [7].

(iii) The morphism ωG\omega_{G} of weakly complete Hopf algebras maps grouplike elements to grouplike elements, whence we have the commutative diagram

𝔤(𝐔(𝔤))(ωG)((G))=𝔤exp𝔾(𝐔(𝔤))expG𝔾(𝐔(𝔤))𝔾(ωG)𝔾([G])=G.\begin{matrix}\mathfrak{g}\subseteq\mathbb{P}(\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{R}}(\mathfrak{g}))&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\mathbb{P}(\omega_{G})}}&\mathbb{P}(\mathbb{R}(G)){=}\mathfrak{g}\\ \hskip 40.0pt\hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle\exp_{\mathbb{G}(\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{R}}(\mathfrak{g}))}$}}$}\Big{\downarrow}&&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle\exp_{G}$}}$\hss}\\ \hfill\mathbb{G}(\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{R}}(\mathfrak{g}))&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits_{\mathbb{G}(\omega_{G})}}&\mathbb{G}(\mathbb{R}[G])=G.\end{matrix}

Since (ωG)\mathbb{P}(\omega_{G}) is a retraction and the image of expG\exp_{G} topologically generates G0G_{0}, the image of 𝔾(ωG)exp𝐔(𝔤)\mathbb{G}(\omega_{G})\circ\exp_{\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{R}}(\mathfrak{g})} topologically generates G0G_{0}. Since the image of the exponential function of the pro-Lie group 𝔾(𝐔(𝔏(G)))\mathbb{G}(\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{R}}(\mathfrak{L}(G))) generates topologically its identity component, 𝔾(ωG)\mathbb{G}(\omega_{G}) maps this identity component onto G0G_{0}.

Since 𝔤(𝐔(𝔤))\mathfrak{g}\subseteq\mathbb{P}(\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{R}}(\mathfrak{g})), and since also any morphism of Hopf algebras maps a primitive element onto a primitive element we know ωG((𝐔(𝔤)))=([G])\omega_{G}(\mathbb{P}(\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{R}}(\mathfrak{g})))=\mathbb{P}(\mathbb{R}[G]).

The following overview of the situation may be helpful:

[G]|𝐔(𝔤)ωG,onto[G0]||||𝔾(𝐔(𝔤))|𝔾([G])=G|||𝔾(𝐔(𝔤))0onto𝔾([G0])=G0=G0exp𝔾(𝐔(𝔤))exp𝔾([G])=expG𝔤(𝐔(𝔤))retract([G0])=([G])=𝔤.\begin{matrix}&&&&\mathbb{R}[G]\\ &&&&\Big{|}\\ \mathop{\bf U\hphantom{}}\nolimits_{\mathbb{R}}(\mathfrak{g})&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\omega_{G},{\rm onto}}}&\mathbb{R}[G_{0}]&&|\\ \Big{|}&&\Big{|}&&\Big{|}\\ \mathbb{G}(\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{R}}(\mathfrak{g}))&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits}&\big{|}&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits}&\mathbb{G}(\mathbb{R}[G])=G\\ \Big{|}&&\Big{|}&&\Big{|}\\ \mathbb{G}(\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{R}}(\mathfrak{g}))_{0}&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\rm onto}}&\mathbb{G}(\mathbb{R}[G_{0}]){=}G_{0}&=&G_{0}\\ \hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle\exp_{\mathbb{G}(\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{R}}(\mathfrak{g}))}$}}$}\Big{\uparrow}&&\Big{\uparrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle\exp_{\mathbb{G}(\mathbb{R}[G])}$}}$\hss}&=&\Big{\uparrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle\exp_{G}$}}$\hss}\\ \mathfrak{g}{\subseteq}\mathbb{P}(\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{R}}(\mathfrak{g}))&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits_{\rm retract}}&\mathbb{P}(\mathbb{R}[G_{0}]){=}\mathbb{P}(\mathbb{R}[G])&=&\mathfrak{g}.\\ \end{matrix} (D)
{Example}

Let 𝔤\mathfrak{g} be a compact semisimple Lie algebra. Then 𝔤=𝔏(G)\mathfrak{g}=\mathfrak{L}(G) for the compact projective group G=Pr(𝔤)G=\Pr(\mathfrak{g}). In this case, G=Γ(𝔤)G=\Gamma(\mathfrak{g}), and we have a commutative diagram

𝐔(𝔤)ωG,surjective[G]||𝔾(𝐔(𝔤))0retractG=𝔾([G])exp𝔾expG𝔤(𝐔(𝔤))retract𝔤.\begin{matrix}\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\omega_{G},surjective}}&\mathbb{R}[G]\\ \Big{|}&&\Big{|}\\ \mathbb{G}(\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}))_{0}&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\rm retract}}&G=\mathbb{G}(\mathbb{R}[G])\\ \hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle\exp_{\mathbb{G}}$}}$}\Big{\uparrow}&&\Big{\uparrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle\exp_{G}$}}$\hss}\\ \mathfrak{g}\subseteq\mathbb{P}(\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}))&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits_{\rm retract}}&\mathfrak{g}.\\ \end{matrix} (D1)

We do not precisely know what 𝔾(𝐔(𝔤))\mathbb{G}(\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})) and (𝐔(𝔤))\mathbb{P}(\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})) are even if 𝔤=so(3)\mathfrak{g}={\rm so}(3) in which case Γ(𝔤)SU(2)\Gamma^{*}(\mathfrak{g})\cong{\rm SU}(2). Still, in this case expΓ(𝔤):𝔤Γ(𝔤)\exp_{\Gamma(\mathfrak{g})}\colon\mathfrak{g}\to\Gamma(\mathfrak{g}) is surjective (cf. [11], Theorems 6.30, 9.19(ii) and Theorem 9.32(ii)).

The group 𝔾(𝐔(𝔤))\mathbb{G}(\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})) of grouplike elements of 𝐔(𝔤)\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}) is a semidirect product of some unknown closed normal subgroup NN by GG. From the content of Diagram (D1)(D_{1}) we do not know anything about NN.

The following example is the opposite to the preceding one:

{Example}

Let 𝔤=X\mathfrak{g}=\mathbb{R}^{X} for some set XX. Then G=Pr(𝔤)=(^)XG=\Pr(\mathfrak{g})=(\widehat{\mathbb{Q}})^{X} and Γ(𝔤)=Γ(G)=X\Gamma(\mathfrak{g})=\Gamma^{*}(G)=\mathbb{R}^{X}.

In our discussion of abelian profinite-dimensional Lie algebras 𝔤\mathfrak{g} we have obtained more information on 𝐔(𝔤)\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}). Here we have our standard diagram:

𝐔(𝔤)ω𝔤,surjective[G](X)||𝔾(𝐔(𝔤))0ontoG=(^)Xexp𝔾expG𝔤(𝐔(𝔤))retract𝔤.\begin{matrix}\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\omega_{\mathfrak{g}},\rm surjective}}&\mathbb{R}[G]{\subset}\mathbb{C}^{\mathbb{Q}^{(X)}}\\ \Big{|}&&\Big{|}\hfill\\ \mathbb{G}(\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}))_{0}&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\rm onto}}&G=(\widehat{\mathbb{Q}})^{X}\hfill\\ \hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle\exp_{\mathbb{G}}$}}$}\Big{\uparrow}&&\Big{\uparrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle\exp_{G}$}}$\hss}\hfill\\ \mathfrak{g}\subseteq\mathbb{P}(\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}))&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits_{\rm retract}}&\mathfrak{g}.\hfill\\ \end{matrix} (D2)

Recall that a finite-dimensional real Lie algebra 𝔤\mathfrak{g} is called “compact” if it is isomorphic to the Lie algebra of a compact group (apologetically defined in [11] Definition 6.1 in that fashion). We now expand this definition to read as follows:

{Definition}

A Lie algebra is called compact if it is profinite-dimensional and is isomorphic to the Lie algebra of a compact group.

We know a real Lie algebra to be compact if and only if there exists a set XX and a family 𝒮{\cal S} of compact simple Lie algebras 𝔰\mathfrak{s} such that 𝔤X×𝒮\mathfrak{g}\cong\mathbb{R}^{X}\times\prod{\cal S}, where we wrote 𝒮\prod{\cal S} for 𝔰𝒮𝔰\prod_{\mathfrak{s}\in{\cal S}}\mathfrak{s}. Now from [11], Theorem 9.76 we obtain the following statement:

{Theorem}

(Sophus Lie’s Third Principal Theorem for Compact Lie Algebras) For every compact real Lie algebra 𝔤\mathfrak{g} there is a projective connected compact group Pr(𝔤)\Pr(\mathfrak{g}) whose Lie algebra 𝔏(Pr(𝔤))\mathfrak{L}(\Pr(\mathfrak{g})) is (isomorphic to) 𝔤\mathfrak{g}.

Every compact connected group GG with 𝔏(G)𝔤\mathfrak{L}(G)\cong\mathfrak{g} is a quotient of Pr(𝔤)\Pr(\mathfrak{g}). modulo some central 0-dimensional subgroup. For details see [11], discussion following Lemma 9.72, notably Theorem 9.76 and Theorem 9.76bis. For the abelian case see [11], Theorem 8.78ff. Notice that for a compact Lie algebra 𝔤\mathfrak{g} the projective compact connected group Pr(𝔤)\Pr(\mathfrak{g}) is simply connected if and only if 𝔤\mathfrak{g} is semisimple. By contrast, if 𝔤=X\mathfrak{g}=\mathbb{R}^{X} for some set then Pr(𝔤)=(^)X\Pr(\mathfrak{g})=(\widehat{\mathbb{Q}})^{X} (see [11], Proposition 8.81), a compact connected abelian group that fails to be simply connected while π1(Pr(𝔤))={0}\pi_{1}(\Pr(\mathfrak{g}))=\{0\} (see [11], Theorem 8.62).

It is very important here to distinguish between the prosimply connected pro-Lie group Γ(𝔤)\Gamma(\mathfrak{g}) and, in the case of a compact Lie algebra 𝔤\mathfrak{g}, the projective compact group Pr(𝔤)\Pr(\mathfrak{g}).


The present concept of weakly complete enveloping algebras now belongs to the circle of ideas of Lie’s Third Fundamental Theorem.

Let 𝔤\mathfrak{g} be a profinite-dimensional Lie algebra over \mathbb{R}. By Theorem 4 above, Γ(𝔤)𝔾(𝐔(𝔤))\Gamma^{*}(\mathfrak{g})\subseteq\mathbb{G}(\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})) is a pro-Lie group whose Lie algebra is 𝔤\mathfrak{g} and the exponential function exp:𝐔(𝔤)Lie𝐔(𝔤)×\exp\colon\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})_{\rm Lie}\to\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})^{\times} of 𝐔(𝔤)\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}) induces the exponential function

exp:𝔏(Γ(𝔤))=𝔤Γ(𝔤).\exp:\mathfrak{L}(\Gamma^{*}(\mathfrak{g}))=\mathfrak{g}\to\Gamma^{*}(\mathfrak{g}).

If G=G0=Γ(𝔤)G=G_{0}=\Gamma(\mathfrak{g}) embeds into its weakly complete group algebra [G]\mathbb{R}[G], then the diagram (D)(D) above shows that α𝔤\alpha_{\mathfrak{g}} is an isomorphism.

We summarize for 𝕂=\mathbb{K}=\mathbb{R}, recalling that we consider 𝔤\mathfrak{g} as a Lie subalgebra of (𝐔(𝔤))𝐔(𝔤)\mathbb{P}(\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}))\subseteq\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}). Indeed, in the context of Lie’s Third Fundamental Theorem, there are two basic pro-Lie groups Γ(𝔤)\Gamma(\mathfrak{g}) and Γ(𝔤)\Gamma^{*}(\mathfrak{g}) attached, and, in the case of a compact profinite-dimensional Lie algebra 𝔤\mathfrak{g}, a third one, Pr(𝔤)\Pr(\mathfrak{g}), and for these we have:

{Theorem}

Let 𝔤\mathfrak{g} be a profinite-dimensional real Lie algebra. Then the pro-Lie group of grouplike elements in the weakly complete enveloping algebra 𝐔(𝔤)\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}) contains the pro-Lie group Γ(𝔤)\Gamma^{*}(\mathfrak{g}), having 𝔤\mathfrak{g} as Lie algebra with the exponential function exp:𝔤Γ(𝔤)\exp\colon\mathfrak{g}\to\Gamma^{*}(\mathfrak{g}) induced by the exponential function of 𝐔(𝔤)\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}). There is a natural quotient morphism α𝔤:Γ(𝔤)Γ(𝔤)\alpha_{\mathfrak{g}}\colon\Gamma(\mathfrak{g})\to\Gamma^{*}(\mathfrak{g}). If the natural morphism Γ(𝔤)[Γ(𝔤)]\Gamma(\mathfrak{g})\to\mathbb{R}[\Gamma(\mathfrak{g})] of Γ(𝔤)\Gamma(\mathfrak{g}) into its weakly complete group algebra is an embedding as is the case if 𝔤\mathfrak{g} is a compact Lie algebra, then α𝔤\alpha_{\mathfrak{g}} is an isomorphism, and in the latter case, there is a natural injective morphism Γ(𝔤)Pr(𝔤)\Gamma(\mathfrak{g})\to\Pr(\mathfrak{g}) with dense image.


7 Appendix: The Category Theoretical Background

For a category 𝒯𝒜{\cal T}\hskip-3.0pt{\cal A} of topological algebraic structures—in the simplest case the category 𝒲{\cal W} of weakly complete vector spaces, and for the category 𝒲𝒜{\cal W}\hskip-2.0pt{\cal A} of weakly complete associative unital algebras, we shall repeatedly discuss an adjoint pair of functors R:𝒲𝒜𝒯𝒜R\colon{\cal W}\hskip-2.0pt{\cal A}\to{\cal T}\hskip-3.0pt{\cal A} and L:𝒯𝒜𝒲𝒜L\colon{\cal T}\hskip-3.0pt{\cal A}\to{\cal W}\hskip-2.0pt{\cal A}. As an example on the simplest level, in the case of 𝒯𝒜=𝒲{\cal T}\hskip-3.0pt{\cal A}={\cal W}, for a weakly complete algebra AA, the 𝒲{\cal W}-object R(A)R(A) will simply be the weakly complete vector space underlying AA, while for a weakly complete vector space WW, the weakly complete algebra L(W)L(W) will be the weakly complete tensor algebra of WW in the category 𝒲{\cal W}.


7.1 Limits and topologically dense subcategories

Since the category 𝒲𝒜{\cal W}\hskip-2.0pt{\cal A} of weakly complete associative unital algebras is at the focus of our considerations, let us point to one important property of the objects in this category, which was expressed in Appendix 7 of [11], Theorem A7.34.

{Theorem}

For every weakly complete unital topological 𝕂\mathbb{K}-algebra AA, the set 𝒥(A){\cal J}(A) of closed two-sided ideals II with finite-dimensional quotient algebras A/IA/I is a filterbasis converging to 0 in AA, and AA is (naturally isomorphic to) the projective limit limI𝒥(A)A/I\lim_{I\in{\cal J}(A)}A/I of these finite-dimensional unital quotient algebras.


This theorem says that any weakly complete unital associative algebra “is approximated by finite-dimensional 𝕂\mathbb{K}-algebras”. Let us briefly recall our approach to projective limits in a category 𝒜{\cal A}. Each directed set JJ is a category with the elements of JJ as objects and for each pair (j,k)(j,k) satisfying jkj\leq k an arrow (JJ-morphism) kjk\to j. A projective (or inverse) system is a functor J𝒜J\to{\cal A}, usually written jAjj\mapsto A_{j} and (kj)(fjk:AkAj)(k\to j)\mapsto(f_{jk}\colon A_{k}\to A_{j}). The projective limit of this system is an object limjJAj\lim_{j\in J}A_{j} together with a family of morphisms fk:limjJAjAkf_{k}\colon\lim_{j\in J}A_{j}\to A_{k}, kJk\in J such that fk=fknfnf_{k}=f_{kn}f_{n} for all arrows nkn\to k. The limit has the universal property that for any system of morphisms φk:AAk\varphi_{k}\colon A\to A_{k} of 𝒜{\cal A}-morphisms satisfying φk=fknφn\varphi_{k}=f_{kn}\varphi_{n} for all arrows nkn\to k there is a unique morphism φ:AlimjJAj\varphi\colon A\to\lim_{j\in J}A_{j} satisfying φk=fkφ\varphi_{k}=f_{k}\varphi for all kJk\in J. The morphisms fkf_{k} are called limit morphisms. (For the example of the category of compact groups see e.g. [11], Definitions 1.25 and 1.27, or see Chapter 1 of [10]. For the general concept of a limit see [11], Definition A3.41, or go to MacLane’s general source book [15].) We have already seen a concrete example of a projective limit in Theorem 7.1. In fact, that example was particular insofar the limit morphisms fkf_{k} were all quotient morphisms. To mathematicians working on the topological algebra of locally compact groups, projective limits are utterly familiar by the Theorem of Yamabe saying that

every locally compact topological group GG with the identity component G0G_{0} is a projective limit of Lie groups provided that G/G0G/G_{0} is compact.

(See the classic of 1955 by Montgomery and Zippin [16].)

In particular, this says that every connected locally compact group is approximated by connected Lie groups. Therefore we need to pinpoint in functorial terms what important theorems like these say on the principle of “approximating complicated topological algebraic structures” by simpler ones.

Topologists like to use the concept of a net on a set XX generalizing that of a sequence [14]: A net (xj)jJ(x_{j})_{j\in J} is a function jxj:JXj\mapsto x_{j}:J\to X for a directed poset JJ. If XX is a topological space and YY a subset of XX such that for every xXx\in X there is a net (yj)jJ(y_{j})_{j\in J} of elements in YY such that x=limjJyjx=\lim_{j\in J}y_{j}, then we say that YY is dense in XX.

So let us now look at a category {\cal B} with a subcategory d{\cal B}_{d}.

{Definition}

We call d{\cal B}_{d} topologically dense in {\cal B} if it is a full subcategory of {\cal B} such that for each object BB in {\cal B} there is a directed set JJ and some projective system

{fjk:BkBj;(j,k)J×J,jk}\{f_{jk}:B_{k}\to B_{j};(j,k)\in J\times J,j\leq k\}

of morphisms in d{\cal B}_{d} such that in {\cal B} the object BB is (isomorphic to) the projective limit limjJBj\lim_{j\in J}B_{j} of this system with suitable limit morphisms

BlimkJBkqjBjB\cong\lim_{k\in J}B_{k}\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{q_{j}}}B_{j}, jJj\in J.

As an example we have seen in Theorem 7.1 that the full subcategory of finite-dimensional unital algebras 𝒲𝒜d{\cal W}\hskip-2.0pt{\cal A}_{d} is topologically dense in the category of weakly complete unital algebras 𝒲𝒜{\cal W}\hskip-2.0pt{\cal A}. In the same spirit, by Peter and Weyl, the category of compact Lie groups is topologically dense in the category of compact groups and continuous group morphisms (see [11],Corollary 2.43). In [11] this Density Theorem is exploited widely.

We owe our readers an explanation of our choice of terminology of a topologically dense subcategory which, as we have argued intuitively, is indeed close to the geometric idea of a dense subspace in a topological space. The necessity of a comment arises from the fact that in category theoretical circles, the choice of the terminology of a “dense subcategory of a category” is half a century old or older as can be seen from MacLane’s standard text of 1971, where the terminology is introduced close to the end of the book [15] on pp. 241, 242, 243. However, that generation of ground breaking category theoreticians had a distinct leaning towards examples supplied by combinatorics and algebra. Therefore, in their eyes, a category DD is, firstly, dense in a category CC if every object of CC is a colimit of a subsystem of objects from DD. We would accordingly suggest to call their approach an approach to codensity. However, secondly, their formation of colimits is not restricted to directed systems (in the way we insist to use projective limits when we (truly!) use limits). As a consequence in their terminology, in the category of sets a category consisting of one singleton object is codense in the whole category, and the category consisting of the object 2\mathbb{Z}^{2} is codense in the category of abelian groups. So dualizing their approach via Pontryagin would yield that the category consisting of the single object of the traditional torus 𝕋2\mathbb{T}^{2} would be dense in the category of all compact abelian groups. — At any rate, this predicament causes us to set off our own terminology of “topologically dense subcategories.”


7.2 Density and Adjunction

For the class of objects of a category 𝒜{\cal A} we write ob(𝒜\mathop{\rm ob}\nolimits({\cal A}). Now let Lo:ob(𝒜)ob()L_{o}\colon\mathop{\rm ob}\nolimits({\cal A})\to\mathop{\rm ob}\nolimits({\cal B}) be a function and R:𝒜R\colon{\cal B}\to{\cal A} a functor and assume that {\cal B} has a subcategory d{\cal B}_{d}.

{Definition}

We say that LoL_{o} is conditionally left adjoint to RR with respect to a subcategory d{\cal B}_{d} of {\cal B} if for each Aob(𝒜)A\in\mathop{\rm ob}\nolimits({\cal A}) there is an 𝒜{\cal A}-morphism ηA:ARLo(A)\eta_{A}\colon A\to RL_{o}(A) such that for each Bob(d)B\in\mathop{\rm ob}\nolimits({\cal B}_{d}) and each morphism f:AR(B)f\colon A\to R(B) in 𝒜{\cal A} there is a unique morphism f:Lo(A)Bf^{\prime}\colon L_{o}(A)\to B in {\cal B} such that f=R(f)ηAf=R(f^{\prime})\circ\eta_{A}.

A special case illustrates this technical concept:

{Remark}

If LoL_{o} is conditionally left adjoint to RR with respect to {\cal B} itself (in place of d{\cal B}_{d}), then LoL_{o} is the restriction to the objects of a functor L:𝒜L\colon{\cal A}\to{\cal B} which is left adjoint to the functor RR. {Proof} [11], Theorem A3.28.

But now we show that the much weaker condition in Definition 7.2 suffices frequently for LoL_{o} to extend to a left adjoint of RR.

{Theorem}

(The Density and Adjunction Theorem) Assume that 𝒜{\cal A} and {\cal B} are two categories and that {\cal B} has a topologically dense subcategory d{\cal B}_{d}. Further assume that

Lo:ob(𝒜)ob() is a function andR:𝒜 is a functor.\begin{matrix}L_{o}&{\colon}&\mathop{\rm ob}\nolimits({\cal A})&\to&\mathop{\rm ob}\nolimits({\cal B})&\mbox{ is a function and}\\ R&{\colon}&{\cal B}&\to&{\cal A}&\mbox{ is a functor.}\hfill\\ \end{matrix}

Then the following conditions are equivalent:

  1. (a)

    LoL_{o} is conditionally left adjoint to RR with respect to d{\cal B}_{d}, and RR preserves projective limits.

  2. (b)

    LoL_{o} extends to a left adjoint LL of RR.

{Proof}

For (b) \Rightarrow (a) we refer to Remark 7.2 and to [11], Theorem A3.52, saying that right adjoints are continuous, that is, preserve all limits.

Now we prove (a) \Rightarrow (b): In view of Remark 7.2 it suffices to show that LoL_{o} is conditionally left adjoint to RR with respect to {\cal B} (in place of merely to d{\cal B}_{d}). So assume now that AA and BB are objects of 𝒜{\cal A} and {\cal B}, respectively, and that f:AR(B)f\colon A\to R(B) is a morphism in 𝒜{\cal A}. Then since d{\cal B}_{d} is topologically dense in {\cal B} we know that there exists a projective system

{fjk:BkBj;(j,k)J×J,jk}\{f_{jk}\colon B_{k}\to B_{j};\ (j,k)\in J\times J,\ j\leq k\} of morphisms in d{\cal B}_{d}

for some directed set JJ in d{\cal B}_{d} such that

B=limjJBj.B=\lim_{j\in J}B_{j}. (1)

Then we obtain a projective system

{R(fjk):R(Bk)R(Bj);(j,k)J×J,jk}\{R(f_{jk})\colon R(B_{k})\to R(B_{j});\ (j,k)\in J\times J,\ j\leq k\} of morphisms in 𝒜{\cal A}

for our directed set JJ. Since LoL_{o} is conditionally adjoint to RR with respect to d{\cal B}_{d}, for each jJj\in J, then there is a unique {\cal B}-morphism (Rfjf):LoABj(Rf_{j}\circ f)^{\prime}\colon L_{o}A\to B_{j} such that

Rfjf=R((Rfjf))ηA.Rf_{j}\circ f=R\big{(}(Rf_{j}\circ f)^{\prime}\big{)}\circ\eta_{A}. (2)

We claim that for jkj\leq k in JJ we have

(Rfjf)=fjk(Rfkf).(Rf_{j}\circ f)^{\prime}=f_{jk}\circ(Rf_{k}\circ f)^{\prime}. (3)

For a proof of this claim, we recall from (2) that (Rfkf)(Rf_{k}\circ f)^{\prime} is the unique {\cal B}-morphism for which R((Rfkf))ηA=RfkfR\big{(}(Rf_{k}\circ f)^{\prime}\big{)}\circ\eta_{A}=Rf_{k}\circ f. Now R((fjk(Rfkf))ηAR\big{(}(f_{jk}\circ(Rf_{k}\circ f)^{\prime}\big{)}\circ\eta_{A}
=RfjkR((Rfkf))ηA(since R is a functor)=RfjkRfkf(by (2))=R(fjkfk)f=Rfjf(since R is a functor)=R(R(fjf))ηA(by (2)).\begin{matrix}=Rf_{jk}\circ R\big{(}(Rf_{k}\circ f)^{\prime}\big{)}\circ\eta_{A}&\mbox{(since $R$ is a functor)}\\ =Rf_{jk}\circ Rf_{k}\circ f\hfill&\mbox{(by (2))}\hfill\\ =R(f_{jk}\circ f_{k})\circ f=Rf_{j}\circ f&\mbox{(since $R$ is a functor)}\\ =R\big{(}R(f_{j}\circ f)^{\prime}\big{)}\circ\eta_{A}\hfill&\mbox{(by (2)).}\hfill\end{matrix}

By the uniqueness in the definition of (Rfjf)(Rf_{j}\circ f)^{\prime} this proves the Claim.

By the universal property of the limit, there is now a unique {\cal B}-morphism f:LoABf^{\prime}\colon L_{o}A\to B so that

(jJ)(Rfjf)=fjf.(\forall j\in J)\,(Rf_{j}\circ f)^{\prime}=f_{j}\circ f^{\prime}. (4)

Consequently, since R(Rfjf)ηA=RfjfR(Rf_{j}\circ f)^{\prime}\circ\eta_{A}=Rf_{j}\circ f by (2), we have

(jJ)Rfjf=Rfj(RfηA).(\forall j\in J)\,Rf_{j}\circ f=Rf_{j}\circ(Rf^{\prime}\circ\eta_{A}). (5)

By (a) we know that we may write RB=limjJRBjRB=\lim_{j\in J}RB_{j} with Rfj:RBRBjRf_{j}\colon RB\to RB_{j} as limit morphisms. By the uniqueness in the universal property of the limit (as specified in great generality in [11], Definition A3.41), from (5) we conclude

(f:ARB)(!f:LoAB)f=RfηA.(\forall f\colon A\to RB)(\exists!f^{\prime}\colon L_{o}A\to B)\quad f=Rf^{\prime}\circ\eta_{A}. (6)

This completes the proof of (b)


7.3 An application: The weak completion of a 𝕂\mathbb{K}-vectorspace

As an example, consider the functor WWu:𝒲𝒱W\to W_{\rm u}\colon{\cal W}\to{\cal V} which assigns to a weakly complete vector space WW the underlying 𝕂\mathbb{K}-vector space WuW_{\rm u}. This functor has a left adjoint L:𝒱𝒲L\colon{\cal V}\to{\cal W} characterized by the usual universal property recognized in the usual diagram:

𝒱𝒲VϵVL(V)uL(V)fL(f)u!fWuidWuW.\begin{matrix}&{\cal V}&&\hbox to19.91692pt{}&{\cal W}\cr\vskip 3.0pt\cr\hrule\cr\cr\vskip 3.0pt\cr V&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\epsilon_{V}}}&L(V)_{\rm u}&\hbox to19.91692pt{}&L(V)\\ \hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle\forall f$}}$}\Big{\downarrow}&&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle L(f^{\prime})_{\rm u}$}}$\hss}&\hbox to19.91692pt{}&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle\exists!f^{\prime}$}}$\hss}\\ W_{\rm u}&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\mathop{\rm id}\nolimits}}&W_{\rm u}&\hbox to19.91692pt{}&W.\end{matrix}

The function ff:𝒱(V,Wu)𝒲(L(V),W)f\mapsto f^{\prime}:{\cal V}(V,W_{\rm u})\to{\cal W}(L(V),W) is a natural bijection.


{Proposition}

For a 𝕂\mathbb{K}-vector space VV we have

L(V)=(V)u and (ωV)ϵV(v)(ω)=ω(v)).L(V)={(V^{*})_{\rm u}}^{*}\mbox{ and }\qquad(\forall\omega\in V^{*})\,\epsilon_{V}(v)(\omega)=\omega(v)).

Note: In a loose fashion we might write L(V)=VL(V)=V^{**} and say: the weak completion of a 𝕂\mathbb{K}-vector space VV is its bidual VV^{**}.

{Proof}

First we test the universal property of L(V)L(V) for Wob(𝒲)W\in\mathop{\rm ob}\nolimits({\cal W}) with dimW<\dim W<\infty. Then the natural morphism ϵW:WW\epsilon_{W}\colon W\to W^{**} is an isomorphism and for f:VWf\colon V\to W we have a commutative diagram

VϵVVffWϵWW.\begin{matrix}V&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\epsilon_{V}}}&V^{**}\\ \hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle f$}}$}\Big{\downarrow}&&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle f^{**}$}}$\hss}\\ W&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits_{\epsilon_{W}}}&W^{**}.\end{matrix}

Any 𝒱{\cal V}-morphism f:VWf\colon V{\to}W yields a unique morphism f=ϵW1f:VWf^{\prime}=\epsilon_{W}^{-1}\circ f^{**}\colon V^{**}\to W. The equation f=(f)uϵVf{=}(f^{\prime})_{\rm u}\circ\epsilon_{V} is now clear. Thus the function V(V)u:ob𝒱ob𝒲V{\mapsto}{(V^{*})_{\rm u}}^{*}:\mathop{\rm ob}\nolimits{\cal V}{\to}\mathop{\rm ob}\nolimits{\cal W} has the universal property of a conditional left adjoint of the functor WWuW\mapsto W_{\rm u} with respect to the topologically dense subcategory 𝒲d{\cal W}_{d} of 𝒲{\cal W} consisting of all finite-dimensional vector spaces. So Theorem 7.2 applies and proves that V(V)uV\mapsto{(V^{*})_{\rm u}}^{*} is left adjoint to WWuW\mapsto W_{\rm u}.

We note that the necessity of invoking Theorem 7.2 indicates that the proof is not entirely trivial. For 𝕂=\mathbb{K}=\mathbb{R} we have seen that VV^{*} for Vob(𝒱)V\in\mathop{\rm ob}\nolimits({\cal V}) is naturally isomorphic to the Pontryagin dual V^=Homcontinuous(V,𝕋)\widehat{V}=\mathop{\rm Hom}\nolimits_{\rm continuous}(V,\mathbb{T}) with 𝕋=/\mathbb{T}=\mathbb{R}/\mathbb{Z} when VV is endowed with the finest locally convex topology. (See [11], Theorem A7.10). If, for a Wob(𝒲)W\in\mathop{\rm ob}\nolimits({\cal W}) we let WfW_{f} denote the underlying vector space of WW endowed with its finest locally convex topology. Then we have

L(V)=(V)u(V^f)^.L(V)={(V^{*})_{\rm u}}^{*}\cong(\widehat{V}_{f})\widehat{\phantom{x}}.

If AA is any abelian topological group and A^d{\widehat{A}}_{d} is its character group endowed with the discrete topology, then the compact “bidual” α(A):=(A^d)^\alpha(A):=(\widehat{A}_{d})\widehat{\phantom{f}} together with natural continuous morphism Aα(A)A\to\alpha(A) is the so called almost periodic compactification of AA. We mention this here in order to exhibit the analogy between the weak completion and the almost periodic compactification.


7.4 Strict density and the preservation of projective limits

We continue with categories 𝒜{\cal A} and {\cal B} having topologically dense subcategories 𝒜d{\cal A}_{d}, respectively, d{\cal B}_{d}, and we consider a pair of adjoint functors L:𝒜L\colon{\cal A}\to{\cal B} and R:𝒜R\colon{\cal B}\to{\cal A} between them. Thus we have the following situation

Strict Density. For each object AA of 𝒜{\cal A} we have some family qj:AAjq_{j}\colon A\to A_{j}, jQ(A)j\in Q(A) of morphisms with AjA_{j} in 𝒜d{\cal A}_{d} with a directed set Q(A)Q(A) of indices, together with a projective system in 𝒜d{\cal A}_{d}, say, qjk:AkAjq_{jk}\colon A_{k}\to A_{j} for jkj\leq k in Q(A)Q(A) such that qj=qjkqkq_{j}=q_{jk}\circ q_{k} for jkj\leq k, giving us a unique isomorphism qA:AlimjQ(A)Ajq_{A}\colon A\to\lim_{j\in Q(A)}A_{j} such that

AqAlimkQ(A)AkqjρjAjqjj=idAj\begin{matrix}A&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{q_{A}}}&\lim_{k\in Q(A)}A_{k}\\ \hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle q_{j}$}}$}\Big{\downarrow}&&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle\rho_{j}$}}$\hss}\hfill\\ A_{j}&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits_{q_{jj}=id}}&A_{j}\hfill\\ \end{matrix}

commutes for each jQ(A)j\in Q(A) for the limit morphisms ρj\rho_{j}.

The universal property of the limit will now provide us with the existence of a crucial morphism

φA:L(A)limjQ(A)L(Aj)\varphi_{A}\colon L(A)\to\lim_{j\in Q(A)}L(A_{j}) ()

We shall investigate this situation in more detail in the remainder of the chapter.

{Lemma}

If L:𝒜L\colon{\cal A}\to{\cal B} is any functor into a complete category {\cal B}, then {L(qjk):L(Ak)L(Aj);(j,k)Q(A)×Q(A),jk}\{L(q_{jk})\colon L(A_{k})\to L(A_{j});\ (j,k)\in Q(A)\times Q(A),\ j\leq k\} is a projective system in {\cal B}, which has a limit

L#(A):=limjQ(A)L(Aj)L^{\#}(A):=\lim_{j\in Q(A)}L(A_{j})

and which provides a morphism φA:L(A)L#(A)\varphi_{A}\colon L(A)\to L^{\#}(A) such that

L(A)φAL#(A)LqjρjL(Aj)=L(Aj)\begin{matrix}L(A)&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\varphi_{A}}}&L^{\#}(A)\\ \hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle Lq_{j}$}}$}\Big{\downarrow}&&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle\rho_{j}$}}$\hss}\\ L(A_{j})&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits_{=}}&L(A_{j})\\ \end{matrix} (7)

commutes for all jQ(A)j\in Q(A) for the limit morphisms ρj\rho_{j}.

If 𝒜{\cal A}, for example, is the category 𝒲{\cal W} of weakly complete vector spaces VV over 𝕂=\mathbb{K}=\mathbb{R} or 𝕂=\mathbb{K}=\mathbb{C}, then a vector subspace EE of VV is called a cofinite-dimensional vector subspace of VV if dimV/E<\dim V/E<\infty. Now each VV defines naturally the filter basis J(V)J(V) of cofinite-dimensional closed vector subspaces WW such that VlimWJ(V)V/WV\cong\lim_{W\in J(V)}V/W where V/WV/W ranges through the finite-dimensional quotient spaces of VV so that the subcategory 𝒲fin{\cal W}_{\rm fin} of all finite-dimensional vector spaces is topologically dense in 𝒲{\cal W}.

The left adjoint functor L:𝒜L\colon{\cal A}\to{\cal B} preserves colimits. But it would be interesting to know whether it preserves also at least some of the significant limits in the contexts that interest us. For instance: If 𝒜=𝒲{\cal A}={\cal W}, the category of weakly complete 𝕂\mathbb{K}-vector spaces: does then LL under certain circumstances preserve projective limits in 𝒲{\cal W} such as limWJ(V)V/W\lim_{W\in J(V)}V/W? That is: Is the natural morphism φV:L(V)limWJ(V)L(V/W)\varphi_{V}\colon L(V)\to\lim_{W\in J(V)}L(V/W) an isomorphism for certain categories {\cal B}?


In the example of the category 𝒲{\cal W} of weakly complete vector spaces each object VV gave rise to the projective system

{fUW:V/WV/U;U,WJ(V),WU}\{f_{UW}\colon V/W\to V/U;U,W\in J(V),W\subseteq U\},

whose limit was naturally isomorphic to VV. Here the filter basis J(V)J(V) converges to 0 in the topological space underlying VV. If FF is any finite-dimensional 𝕂\mathbb{K}-vector space, then the filter base of vector subspaces {FW:WJ(F)}\{F\cap W:W\in J(F)\} of FF converges to zero in FF. Now, since dim𝕂F<\dim_{\mathbb{K}}F<\infty, there is some member WFJ(V)W_{F}\in J(V) such that WWFW\subseteq W_{F} implies FW={0}F\cap W=\{0\}. In terms of quotient morphisms of VV this can be expressed as follows:

Each object VV of 𝒲{\cal W} has canonical projective system of quotient maps qW:VV/Wq_{W}\colon V\to V/W, WJ(V)W\in J(V) and dimV/W<\dim V/W<\infty such that VlimWJ(V)V/WV\cong\lim_{W\in J(V)}V/W, and that for each morphism f:VFf\colon V\to F into a finite-dimensional vector space FF we have kerfJ(V)\ker f\in J(V) so that for every WJ(V)W\in J(V) with WkerfW\subseteq\ker f the morphism ff factors through qW:VV/Wq_{W}\colon V\to V/W. In other words, there is an index WfJ(V)W_{f}\in J(V) such that for every WJ(V)W\in J(V) such that WWfW\subseteq W_{f} there is a morphism pW:V/WFp_{W}\colon V/W\to F such that f=pWqWf=p_{W}\circ q_{W}, as in the following commutative diagram:

V=AqWfV/WpWF.\begin{matrix}V&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{=}}&A\\ \hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle q_{W}$}}$}\Big{\downarrow}&&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle f$}}$\hss}\\ V/W&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits_{p_{W}}}&F.\end{matrix}

Following this example we formulate the following definition:

{Definition}

For an object AA of 𝒜{\cal A}, a projective system

{qjk:AkAj:(j,k)Q(A)×Q(A),jk}\{q_{jk}\colon A_{k}\to A_{j}:(j,k)\in Q(A)\times Q(A),\quad j\leq k\} (PS)

in 𝒜{\cal A} will be called appropriate for AA if AlimjQ(A)AjA\cong\lim_{j\in Q(A)}A_{j} with limit morphisms qj:AAjq_{j}\colon A\to A_{j} such that the following conditions are satisfied at least for a cofinal set of indices jj in Q(A)Q(A):

  1. (i)

    For every 𝒜{\cal A}-morphism f:AFf\colon A\to F into an 𝒜d{\cal A}_{d}-object FF there is a j0Q(A)j_{0}\in Q(A) such that for all jj0j\geq j_{0} there is an 𝒜{\cal A}-morphism pj:AjFp_{j}\colon A_{j}\to F such that f=pjqjf=p_{j}\circ q_{j}.

  2. (ii)

    The limit morphisms qj:AAjq_{j}\colon A\to A_{j} are epic.

If for an object AA of 𝒜{\cal A} there is an appropriate projective system {qjk}\{q_{jk}\}, then we say that AA is appropriately representable.

{Definition}

A subcategory 𝒜d{\cal A}_{d} of a category 𝒜{\cal A} is called strictly dense in 𝒜{\cal A} if each object AA in 𝒜{\cal A} is appropriately representable by a projective system (PS)(PS) such that all AjA_{j} are objects from 𝒜d{\cal A}_{d}.

Note that in Condition 7.4(i) the factorisation f=pjqjf=p_{j}\circ q_{j} is depicted by the commuting diagram

A=AqjfAjpjF.\begin{matrix}A&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{=}}&A\\ \hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle q_{j}$}}$}\Big{\downarrow}&&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle f$}}$\hss}\\ A_{j}&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits_{p_{j}}}&F.\end{matrix}

Moreover, condition (ii) is certainly satified if the morphisms qj:AAjq_{j}\colon A\to A_{j} are quotient maps as in the case 𝒜=𝒲{\cal A}={\cal W} that we used as motivation above. Accordingly, we observe that in the cateory 𝒲{\cal W} of weakly complete 𝕂\mathbb{K}-vector spaces, the subcategory of finite-dimensional vector spaces is strictly dense.

A more sophisticated example is the category of pro-Lie groups, in which the subcategory of Lie groups is strictly dense.

However, the most relevant example for us is the category 𝒲𝒜{\cal W}\hskip-2.0pt{\cal A} of weakly complete associative unital algebras in which the full subcategory 𝒲𝒜fin{\cal W}\hskip-2.0pt{\cal A}_{\rm fin} of finite-dimensional 𝕂\mathbb{K}-algebras is strictly dense by Theorem 7.1.

For the applications of the next theorem it is useful to first recall the following general lemma:

{Lemma}

Left adjoint functors preserve epics. {Proof} By [11], Theorem A3.52 a left adjoint LL preserves colimits. A morphism e:A1A2e\colon A_{1}\to A_{2} is an epic if and only if

A1eA2eidA2A2idA2A2\begin{matrix}A_{1}&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{e}}&A_{2}\\ \hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle e$}}$}\Big{\downarrow}&&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle\mathop{\rm id}\nolimits_{A_{2}}$}}$\hss}\\ A_{2}&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits_{\mathop{\rm id}\nolimits_{A_{2}}}}&A_{2}\\ \end{matrix}

is a pushout. A pushout is a colimit (cf. [11] EA3.27), epimorphisms and pushouts are dual to monomorphisms and pullbacks; the latter are defined in  [11] Definition A3.9 and Definition A3.43(ii), respectively.

Now we apply Definition 7.4 to provide circumstances in which the morphism ()(\bullet) called φA\varphi_{A} and introduced in Lemma 7.4 is an epimorphism.

In order to simplify the language of our notation we introduce the following definition.

{Definition}

A pair of categories 𝒜{\cal A} and {\cal B} shall be called a suitable pair of categories if

  1. (i)

    𝒜{\cal A} posesses a strictly dense subcategory 𝒜d{\cal A}_{d},

  2. (ii)

    {\cal B} posseses a topologically dense subcategory d{\cal B}_{d},

  3. (iii)

    there is a pair of functors L:𝒜L\colon{\cal A}\to{\cal B} and R:𝒜R\colon{\cal B}\to{\cal A} such that LL is left adjoint to RR, and

  4. (iv)

    RR maps d{\cal B}_{d} into 𝒜d{\cal A}_{d}.

{Theorem}

Let 𝒜{\cal A} and {\cal B} be a suitable pair of categories. Assume that the object AA of 𝒜{\cal A} is appropriately representable in the form A=limjQ(A)AjA=\lim_{j\in Q(A)}A_{j} for an appropriate projective system

{qjk:AkAj:(j,k)Q(A)×Q(A),jk}.\{q_{jk}\colon A_{k}\to A_{j}:(j,k)\in Q(A)\times Q(A),j\leq k\}.

Then the following statements hold:

{L(qjk):L(Ak)L(Aj):(j,k)Q(A)×Q(A),jk}\{L(q_{jk})\colon L(A_{k})\to L(A_{j}):(j,k)\in Q(A)\times Q(A),\quad j\leq k\} (a)

is appropriate for L#(A):=limjJL(Aj)L^{\#}(A):=\lim_{j\in J}L(A_{j}) in {\cal B}.

(b) The morphism

φA:L(A)L#(A)\varphi_{A}\colon L(A)\to L^{\#}(A) ()

is an epimorphism. {Proof} For proving (b) let α,β:L#(A)B\alpha,\beta\colon L^{\#}(A)\to B be {\cal B}-morphisms such that αφA=βφA\alpha\circ\varphi_{A}=\beta\circ\varphi_{A}. We must show that α=β\alpha=\beta. We shall first argue that we may assume that BB is in d{\cal B}_{d}. Since d{\cal B}_{d} is topologically dense in {\cal B} by 7.4(ii), there is a projective system

{rmn:BnBn;(m,n)Q(B)×Q(B),mn}\{r_{mn}\colon B_{n}\to B_{n};\ (m,n)\in Q(B)\times Q(B),\ m\leq n\}

in d{\cal B}_{d} such that B=limmQ(B)BmB=\lim_{m\in Q(B)}B_{m} with limit morphisms rm:BBmr_{m}\colon B\to B_{m} such that rm=rmnrnr_{m}=r_{mn}\circ r_{n} for mnm\leq n. Then for each mQ(B)m\in Q(B) we have morphisms rmα,rmβ:L#(A)Bmr_{m}\circ\alpha,\ r_{m}\circ\beta\colon L^{\#}(A)\to B_{m} such that

rmαφA=rmβφA.r_{m}\circ\alpha\circ\varphi_{A}=r_{m}\circ\beta\circ\varphi_{A}. (8)

If we can show that for all mm we have rmα=rmβ:L#(A)Bmr_{m}\circ\alpha=r_{m}\circ\beta\colon L^{\#}(A)\to B_{m}, then by the uniquenes of the universal property of the limit this will show α=β\alpha=\beta, and we shall be done. So from now on we shall assume that BB is in d{\cal B}_{d}.

(a) For a proof of (a) we shall have to prove that the projective system

{L(qjk):L(Ak)L(Aj):(j,k)Q(A)×Q(A),jk}\{L(q_{jk})\colon L(A_{k})\to L(A_{j}):(j,k)\in Q(A)\times Q(A),\quad j\leq k\}

with the limit morphisms ρj:L#(A)L(Aj)\rho_{j}\colon L^{\#}(A)\to L(A_{j}) is appropriate, that is, for each morphism 𝐟:L#(A)B{\bf f}\colon L^{\#}(A)\to B for an object BB in d{\cal B}_{d}, and all sufficiently large jj there will be morphisms 𝐩j:L(Aj)B{\bf p}_{j}\colon L(A_{j})\to B such that 𝐩jρj=𝐟:L#(A)B{\bf p}_{j}\circ\rho_{j}={\bf f}:L^{\#}(A)\to B with the limit morphisms and ρj:L#(A)L(Aj)\rho_{j}\colon L^{\#}(A)\to L(A_{j}).

So let BB be a d{\cal B}_{d}-object and 𝐟:L#(A)B{\bf f}\colon L^{\#}(A)\to B a {\cal B}-morphism. We define f:=𝐟φAf:={\bf f}\circ\varphi_{A}. Then R(f):RL(A)R(B)R(f)\colon RL(A)\to R(B) is an 𝒜{\cal A}-morphism into an 𝒜d{\cal A}_{d} object R(B)R(B) since RR maps d{\cal B}_{d} into 𝒜d{\cal A}_{d}. By the hypothesis, that AA is appropriately represented in the form A=limjQ(A)AjA=\lim_{j\in Q(A)}A_{j}, the morphism

AηARL(A)R(f)R(B)A\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\eta_{A}}}RL(A)\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{R(f)}}R(B)

is a morphism from AA to an 𝒜d{\cal A}_{d} object for which we find a j0Q(A)j_{0}\in Q(A) such that for all jQ(A)j\in Q(A) with j0jj_{0}\leq j there is a pj:AjR(B)p_{j}\colon A_{j}\to R(B) such that R(f)ηA=pjqjR(f)\circ\eta_{A}=p_{j}\circ q_{j}. By the universal property of the left adjoint LL there is a unique {\cal B}-morphism pj:L(Aj)Bp_{j}^{\prime}\colon L(A_{j})\to B such that pj=R(pj)ηAjp_{j}=R(p_{j}^{\prime})\circ\eta_{A_{j}}. The following diagram illustrates the situation:

𝒜AηARL(A)L(A)φAL#(A)qjRL(qj)L(qj)ρjAjηAjRL(Aj)L(Aj)=L(Aj)pjR(pj)pjpjR(B)=R(B)B=B.R(pj)RL(qj)=R(f)pjL(qj)=f.\begin{matrix}&{\cal A}&&\hbox to19.91692pt{}&&{\cal B}\cr\vskip 3.0pt\cr\hrule\cr\cr\vskip 3.0pt\cr A&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\eta_{A}}}&RL(A)&\hbox to19.91692pt{}&L(A)&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\varphi_{A}}}&L^{\#}(A)\\ \hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle q_{j}$}}$}\Big{\downarrow}&&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle RL(q_{j})$}}$\hss}&\hbox to19.91692pt{}&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle L(q_{j})$}}$\hss}&&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle\rho_{j}$}}$\hss}\\ A_{j}&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\eta_{A_{j}}}}&RL(A_{j})&\hbox to19.91692pt{}&L(A_{j})&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{=}}&L(A_{j})\\ \hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle p_{j}$}}$}\Big{\downarrow}&&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle R(p_{j}^{\prime})$}}$\hss}&\hbox to19.91692pt{}&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle p_{j}^{\prime}$}}$\hss}&&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle p_{j}^{\prime}$}}$\hss}\\ R(B)&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits_{=}}&R(B)&\hbox to19.91692pt{}&B&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits_{=}}&B.\\ \vskip 12.0pt\cr&&\hskip-40.0ptR(p_{j}^{\prime})\circ RL(q_{j})=R(f)&\hbox to19.91692pt{}&p_{j}^{\prime}\circ L(q_{j})=f.&&\\ \end{matrix}

We claim that

pjρj=𝐟:L#(A)B{p_{j}}^{\prime}\circ\rho_{j}={\bf f}:L^{\#}(A)\to B

in the right half of the diagram. For a proof of this claim we invoke the functoriality of the limit in the following lemma, which we consider well understood: {Lemma} Let ξj:limkJXkXj\xi_{j}:\lim_{k\in J}X_{k}\to X_{j} and ωj:limkJYkYj\omega_{j}:\lim_{k\in J}Y_{k}\to Y_{j} the corresponding limit cones of two projective limits and assume that there is a compatible family φj:XjYj\varphi_{j}\colon X_{j}\to Y_{j} of morphism such that for all jkj\leq k the diagrams

YkωjkYjφkφjXkξjkXj\begin{matrix}Y_{k}&\smash{\mathop{\hbox to40.0pt{\leftarrowfill}}\limits^{\omega_{jk}}}&Y_{j}\\ \hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle\varphi_{k}$}}$}\Big{\downarrow}&&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle\varphi_{j}$}}$\hss}\\ X_{k}&\smash{\mathop{\hbox to40.0pt{\leftarrowfill}}\limits_{\xi_{jk}}}&X_{j}\end{matrix}

commute. Then there is a unique morphism φ:limkXklimkYk\varphi\colon\lim_{k}X_{k}\to\lim_{k}Y_{k} such that ωjφ=φjξj\omega_{j}\circ\varphi=\varphi_{j}\circ\xi_{j} for all jJj\in J, i.e., the following diagram commutes

limkXkφlimkYkξjωjXjφjYj.\begin{matrix}\lim_{k}X_{k}&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\varphi}}&\lim_{k}Y_{k}\\ \hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle\xi_{j}$}}$}\Big{\downarrow}&&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle\omega_{j}$}}$\hss}\\ X_{j}&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits_{\varphi_{j}}}&Y_{j}.\end{matrix}

Now we apply this lemma to the special case that the YjY_{j} arise as a a constant projective diagram with Yj=YY_{j}=Y and ωjk=idY\omega_{jk}=\mathop{\rm id}\nolimits_{Y} for all jkj\leq k in JJ, and limkYk=Y\lim_{k}Y_{k}=Y with ωk=idY\omega_{k}=\mathop{\rm id}\nolimits_{Y} for all kJk\in J. Then φ:limkXkY\varphi\colon\lim_{k}X_{k}\to Y agrees with φjξj\varphi_{j}\circ\xi_{j} for all jj, that is

limkXkφYξjidYXjφjY\begin{matrix}\lim_{k}X_{k}&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\varphi}}&Y\\ \Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle\xi_{j}$}}$\hss}&&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle\mathop{\rm id}\nolimits_{Y}$}}$\hss}\\ X_{j}&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits_{\varphi_{j}}}&Y\end{matrix} ()

commutes for all jJj\in J. This we apply with J={jQ(A):j0j}J=\{j\in Q(A):j_{0}\leq j\}, Xj=L(Aj)X_{j}=L(A_{j}), Y=Yj=BY=Y_{j}=B, ξj=ρj:L#(A)=limkXkXj=L(Aj)\xi_{j}=\rho_{j}\colon L^{\#}(A){=}\lim_{k}X_{k}{\to}X_{j}{=}L(A_{j}), φj=pj:L(Aj)=XjY=B\varphi_{j}{=}p_{j}^{\prime}\colon L(A_{j}){=}X_{j}{\to}Y=B, φ=𝐟:L#(A)B=Y\varphi={\bf f}\colon L^{\#}(A)\to B=Y. Then the commuting of ()({\dagger}) yields exactly pjρj=𝐟p_{j}^{\prime}\circ\rho_{j}={\bf f} for j0jj_{0}\leq j as asserted. So for all jQ(A)j\in Q(A), j0jj_{0}\leq j, the morphisms pj:L(Aj)Bp_{j}^{\prime}\colon L(A_{j})\to B are the required morphisms 𝐩j:L(Aj)B{\bf p}_{j}\colon L(A_{j})\to B.

(b) We finally prove that φA\varphi_{A} is an epic. By (b) there is a j0j_{0} such that for all jj0j\geq j_{0} there exist {\cal B}-morphisms αj:L(Aj)B\alpha_{j}\colon L(A_{j})\to B and βj:L(Aj)B\beta_{j}\colon L(A_{j})\to B such that α=αjρj\alpha=\alpha_{j}\rho_{j} and β=βjρj\beta=\beta_{j}\rho_{j}:

L#(A)ρjL(Aj)αβαjβjBidBB.\begin{matrix}L^{\#}(A)&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\rho_{j}}}&L(A_{j})\\ \hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle\alpha$}}$}\Big{\downarrow}\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle\beta$}}$\hss}&&\hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle\alpha_{j}$}}$}\Big{\downarrow}\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle\beta_{j}$}}$\hss}\\ B&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits_{\mathop{\rm id}\nolimits_{B}}}&B.\end{matrix} (9)

Then we must show that

(jj0)αj=βj.(\forall j\geq j_{0})\,\alpha_{j}=\beta_{j}. ()

Now by the Definition of φA\varphi_{A} we have

(jJ)ρjφA=L(qj).(\forall j\in J)\,\rho_{j}\varphi_{A}=L(q_{j}). (10)

We consider the following diagram

L(A)idL(A)L(A)φAL(qj)L#(A)ρjL(Aj)αβαjβjBidBB.\begin{matrix}L(A)&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\mathop{\rm id}\nolimits_{L(A)}}}&L(A)\\ \hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle\varphi_{A}$}}$}\Big{\downarrow}&&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle L(q_{j})$}}$\hss}\\ L^{\#}(A)&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\rho_{j}}}&L(A_{j})\\ \hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle\alpha$}}$}\Big{\downarrow}\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle\beta$}}$\hss}&&\hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle\alpha_{j}$}}$}\Big{\downarrow}\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle\beta_{j}$}}$\hss}\\ B&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits_{\mathop{\rm id}\nolimits_{B}}}&B.\end{matrix} (11)

The top square commutes by (10) for all jJj\in J. The two bottom squares commute for each α\alpha and β\beta and all jj0j\geq j_{0} by (9). Accordingly, the outside rectangles commute for both α\alpha and β\beta. The left vertical edges αφA=βφA\alpha\varphi_{A}=\beta\varphi_{A} agree by assumption on α\alpha and β\beta. So for each jj0j\geq j_{0} we compute

αjL(qj)=idBαφAidL(A)1=idBβφaidL(A)1=βjL(qj).\alpha_{j}L(q_{j})=\mathop{\rm id}\nolimits_{B}\alpha\varphi_{A}\mathop{\rm id}\nolimits_{L(A)}^{-1}=\mathop{\rm id}\nolimits_{B}\beta\varphi_{a}\mathop{\rm id}\nolimits_{L(A)}^{-1}=\beta_{j}L(q_{j}). (12)

The morphisms qj:AAjq_{j}\colon A\to A_{j} are epic by Definition 7.8(ii). Then Lemma 7.10 shows that the morphisms L(qj):L(A)L(Aj)L(q_{j})\colon L(A)\to L(A_{j}) are all epic. Now (11) implies that αj=βj\alpha_{j}=\beta_{j} for all jj0j\geq j_{0}. So ()(*) is proved and this is what we had to show.

Notice that Theorem 7.4 does not assert that the objects L(Aj)L(A_{j}) are (even cofinally) objects of the topologically dense subcategory d{\cal B}_{d}. In fact, in the applications, which we aim for, this is not the case. It is nevertheless assumed by Definition 7.4(ii) that every object BB of {\cal B} is a projective limit of objects from d{\cal B}_{d}.

{Lemma}

Let 𝒜{\cal A} and {\cal B} be a suitable pair of categories. Assume that the object AA of 𝒜{\cal A} is appropriately representable as A=limjQ(A)AjA=\lim_{j\in Q(A)}A_{j}. Abbreviating limjQ(A)L(Aj)\lim_{j\in Q(A)}L(A_{j}) by L#(A)L^{\#}(A), define an 𝒜{\cal A}-morphism ηA#:AR(L#(A))\eta_{A}^{\#}\colon A\to R(L^{\#}(A)) by ηA#=R(φA)ηA\eta_{A}^{\#}=R(\varphi_{A})\circ\eta_{A}.

Then for each object BdB\in{\cal B}_{d} and each 𝒜{\cal A}-morphism f:ARBf\colon A\to RB there is a unique {\cal B}-morphism f#:L#(A)Bf^{\#}\colon L^{\#}(A)\to B such that f=R(f#)ηA#f=R(f^{\#})\circ\eta^{\#}_{A}.

{Proof}

By Definition 7.4(iv), the functor RR maps d{\cal B}_{d} into 𝒜d{\cal A}_{d}. So RBRB is in 𝒜d{\cal A}_{d}. By Definition 7.4(i), the subcategory 𝒜d{\cal A}_{d} is strictly dense in 𝒜{\cal A}. Since AA is appropriately representable, there indeed exists a projective system

{qjk:AkAj:(j,k)Q(A)×Q(a),jk}\{q_{jk}\colon A_{k}\to A_{j}:(j,k)\in Q(A)\times Q(a),\quad j\leq k\}

which is appropriate for AA. So there is a j0Q(A)j_{0}\in Q(A) such that for all jj with j0jj_{0}\leq j there are 𝒜d{\cal A}_{d} morphisms pj:AjR(B)p_{j}\colon A_{j}\to R(B) such that f=pjqjf=p_{j}\circ q_{j}. Since LL is left adjoint to RR, there are unique {\cal B}-morphisms f:LABf^{\prime}\colon LA\to B and (pj):LAjB(p_{j})^{\prime}\colon LA_{j}\to B such that f=RfηAf=Rf^{\prime}\circ\eta_{A} and pj=R(pj)ηAp_{j}=R(p_{j})^{\prime}\circ\eta_{A}. Now from f=pjqjf=p_{j}\circ q_{j}, by [11], Proposition A3.33 we deduce

f=(pj)Lqj.f^{\prime}=(p_{j})^{\prime}\circ Lq_{j}. (13)

The fill-in morphism φA:LAL#A\varphi_{A}\colon LA\to L^{\#}A of Lemma 7.4 satisfies Lqj=ρjφALq_{j}=\rho_{j}\circ\varphi_{A} with the limit morphism ρj:L#ALAj\rho_{j}\colon L^{\#}A\to LA_{j}. We set fj#=(pj)ρjf^{\#}_{j}=(p_{j})^{\prime}\circ\rho_{j}. If kQ(A)k\in Q(A) satisfies jkj\leq k, then we have a commutative diagram

L#A=L#AρkρjL(Ak)L(qjk)L(Aj)(pk)(pj)B=B\begin{matrix}L^{\#}A&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits_{=}}&L^{\#}A\\ \hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle\rho_{k}$}}$}\Big{\downarrow}&&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle\rho_{j}$}}$\hss}\\ L(A_{k})&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{L(q_{jk})}}&L(A_{j})\\ \hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle(p_{k})^{\prime}$}}$}\Big{\downarrow}&&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle(p_{j})^{\prime}$}}$\hss}\\ B&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{{=}}}&B\\ \end{matrix}

with fj#=(pj)ρj=(pk)ρk=fk#f^{\#}_{j}=(p_{j})^{\prime}\circ\rho_{j}=(p_{k})^{\prime}\circ\rho_{k}=f^{\#}_{k}. That is, for a cofinal subset Qc(A)Q(A)Q_{c}(A)\subseteq Q(A) of Q(A)Q(A) the function kfk#:Qc(A)(L#(A),B)k\mapsto f^{\#}_{k}\colon Q_{c}(A)\to{\cal B}(L^{\#}(A),B) is constant. Hence we have a unique morphism f#:L#ABf^{\#}\colon L^{\#}A\to B such that f#=fj#f^{\#}=f^{\#}_{j} for all sufficiently large jj such that (7) and (13) imply

f#φA=(pj)ρjφA=(pj)Lqj=ff^{\#}\circ\varphi_{A}=(p_{j})^{\prime}\circ\rho_{j}\circ\varphi_{A}=(p_{j})^{\prime}\circ Lq_{j}=f^{\prime} (14)

for all sufficiently large jj, and thus

R(f#)ηA#=R(f#)R(φA)ηA=R(f)ηA=f.R(f^{\#})\circ\eta_{A}^{\#}=R(f^{\#})\circ R(\varphi_{A})\circ\eta_{A}=R(f^{\prime})\circ\eta_{A}=f.

If f:L#ABf^{*}\colon L^{\#}A\to B is a {\cal B}-morphism such that f=R(f)ηA#=R(f)R(φA)ηAf=R(f^{*})\circ\eta_{A}^{\#}=R(f^{*})\circ R(\varphi_{A})\circ\eta_{A}, then fφA=ff^{*}\circ\varphi_{A}=f^{\prime} by the uniqueness in determining ff^{\prime}. Since also f#φA=ff^{\#}\circ\varphi_{A}=f^{\prime} by (14) above, we may conclude f=f#f^{*}=f^{\#}, since φA\varphi_{A} is an epimorphism by Theorem 7.4. This completes the proof of Lemma 7.4. As a corollary of the epimorphism Theorem 7.4 we now have the following main result, in which it happens that a left adjoint functor LL preserves, in addition to all colimits, also certain limits. In its formulation we retain the notation of Lemma 7.4 and Definition 7.4.

{Theorem}

Let 𝒜{\cal A} and {\cal B} be a suitable pair of categories and assume that the projective system

{qjk:AkAj:(j,k)Q(A)×Q(A),jk}\{q_{jk}\colon A_{k}\to A_{j}:(j,k)\in Q(A)\times Q(A),\quad j\leq k\}

is appropriate for AA. Then the morphism

φA:L(limjQ(A)Aj)limjQ(A)L(Aj)\varphi_{A}\colon L(\lim_{j\in Q(A)}A_{j})\to\lim_{j\in Q(A)}L(A_{j}) ()

is an isomorphism.

Proof.

From Lemma 7.4 and Theorem 7.2 it now follows that L#L^{\#} extends to a functor L#:𝒜L^{\#}\colon{\cal A}\to{\cal B} which is left adjoint to RR. Thus LL and L#L^{\#} are naturally isomorphic functors. Then there is a commutative diagram of natural functions

(L#(A),B)βAB(L(A),B)αAB#αAB𝒜(A,R(B))=𝒜(A,R(B)),\begin{matrix}{\cal B}(L^{\#}(A),B)&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits_{\beta_{AB}}}&{\cal B}(L(A),B)\\ \hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle\alpha^{\#}_{AB}$}}$}\Big{\downarrow}&&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle\alpha_{AB}$}}$\hss}\\ {\cal A}(A,R(B))&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{=}}&{\cal A}(A,R(B)),\\ \end{matrix} (15)
  1. (a)

    βAB(h)=hφA\beta_{AB}(h)=h\circ\varphi_{A} for h:L#(A)Bh\colon L^{\#}(A)\to B,

  2. (b)

    αAB#(h)=R(hφA)ηA\alpha^{\#}_{AB}(h)=R(h\circ\varphi_{A})\circ\eta_{A}, for h:LABh\colon LA\to B,

  3. (c)

    αAB(h)=R(h)ηA\alpha_{AB}(h)=R(h)\circ\eta_{A}, for h:L(A)Bh\colon L(A)\to B.

The bijectivity of αAB\alpha_{AB} expresses the fact that LL is left adjoint to RR, and likewise the bijectivity of αAB#\alpha^{\#}_{AB} is now secured since we proved that L#L^{\#} is left adjoint to RR. the commutativity of the diagram (15) then shows the bijectivity of βAB\beta_{AB} which in turn proves that φA\varphi_{A} is an isomorphism. This completes the proof. ∎

Since the right adjoint R:𝒜R\colon{\cal B}\to{\cal A} preserve limits, the following corollary is immediate:

{Corollary}

Under the hypotheses of   Theorem 7.4, for each A=limj(A)AjA{=}\lim_{j\in\mathbb{Q}(A)}A_{j}, the 𝒜{\cal A}-morphism R(φA):RL(A)=RL(limjQ(A)Aj)limjQ(A)RL(Aj)R(\varphi_{A})\colon RL(A)=RL(\lim_{j\in Q(A)}A_{j})\to\lim_{j\in Q(A)}RL(A_{j}) is an isomorphism.

If rj:L#(A):=limkL(Ak)L(Aj)r_{j}\colon L^{\#}(A):=\lim_{k}L(A_{k})\to L(A_{j}), jQ(A)j\in Q(A) denotes the limit morphisms, the situation is illustrated by the following diagram:

𝒜AηARL(A)R(φA)limkRL(Ak)L(A)qjRL(qj)R(rj)!qjAj=AjηAjRL(Aj)L(Aj)\begin{matrix}&&{\cal A}&&&\hbox to71.13188pt{}&{\cal B}\cr\vskip 3.0pt\cr\hrule\cr\cr\vskip 3.0pt\cr A&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\eta_{A}}}&RL(A)&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{R(\varphi_{A})}}&\lim_{k}RL(A_{k})&\hbox to71.13188pt{}&L(A)\\ \hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle\forall q_{j}$}}$}\Big{\downarrow}&&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle RL(q_{j}^{\prime})$}}$\hss}&&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle R(r_{j})$}}$\hss}&\hbox to71.13188pt{}&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle\exists!q_{j}^{\prime}$}}$\hss}&&\\ A_{j}&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits_{=}}&A_{j}&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits_{\eta_{A_{j}}}}&RL(A_{j})&\hbox to71.13188pt{}&L(A_{j})\\ \end{matrix}
{Corollary}

Assume the hypotheses of Theorem 7.4, and, in addition, that for all objects Aob(𝒜d)A\in\mathop{\rm ob}\nolimits({\cal A}_{d}) the front adjunction ηA:ARL(A)\eta_{A}\colon A\to RL(A) is monic. Then it is monic for all objects Aob(𝒜)A\in\mathop{\rm ob}\nolimits({\cal A}) in 𝒜{\cal A}. {Proof} Let α,β:XA\alpha,\beta\colon X\to A be morphisms such that ηAα=ηAβ\eta_{A}\alpha=\eta_{A}\beta. Then for jQ(A)j\in Q(A) we have ηAjqjα=ηAjRL(qj)ηAα=ηAjRL(qj)ηAβ=ηAjqjβ\eta_{A_{j}}q_{j}\alpha=\eta_{A_{j}}RL(q_{j}^{\prime})\eta_{A}\alpha=\eta_{A_{j}}RL(q_{j}^{\prime})\eta_{A}\beta=\eta_{A_{j}}q_{j}\beta. Since ηAj\eta_{A_{j}} is monic, we have

(jQ(A))qjα=qjβ.(\forall j\in Q(A))\,q_{j}\alpha=q_{j}\beta.

Since A=limjQ(A)AjA=\lim_{j\in Q(A)}A_{j}, the uniqueness of the morphism in the universal property of the limit (see. e.g. [11], Definition A3.41) implies α=β\alpha=\beta.


7.5 Application: =𝒲𝒜{\cal B}={\cal W}\hskip-2.0pt{\cal A}.

Our main target category for various left adjoint functors is the category 𝒲𝒜{\cal W}\hskip-2.0pt{\cal A} of weakly complete unital algebras over 𝕂=,\mathbb{K}=\mathbb{R},\mathbb{C}. {Proposition} The subcategory 𝒲𝒜d{\cal W}\hskip-2.0pt{\cal A}_{d} of finite-dimensional associative unital 𝕂\mathbb{K}-algebras is strictly dense in 𝒲𝒜{\cal W}\hskip-2.0pt{\cal A}. {Proof} Each weakly complete algebra AA has a filter base Q(A)Q(A) of closed two sided ideals II such tht A/IA/I is a finite-dimensional 𝕂\mathbb{K} algebra, and the natural morphims qA:AlimIQ(A)A/Iq_{A}\colon A\to\lim_{I\in Q(A)}A/I is an isomorphism according to Theorem 7.1. Thus by Definition 7.1, 𝒲𝒜d{\cal W}\hskip-2.0pt{\cal A}_{d} is topologically dense in 𝒲𝒜{\cal W}\hskip-2.0pt{\cal A}.

We need to verify the conditions of Definition 7.4. Let f:AFf\colon A\to F be an 𝒲𝒜{\cal W}\hskip-2.0pt{\cal A}-morphism for a finite-dimensional algebra FF and let I=kerfI=\ker f. Then A/IimfA/I\cong\mathop{\rm im}\nolimits f (see [11], Theorem A7.12(b)) and so, since dimimfdimF<\dim\mathop{\rm im}\nolimits f\leq\dim F<\infty, IQ(A)I\in Q(A). Let qI:AA/Iq_{I}\colon A\to A/I denote the quotient morphism and pI:A/IFp_{I}\colon A/I\to F the injective morphism induced by ff. Then f=pIqIf=p_{I}\circ q_{I}. Hence condition (i) of Definition 7.4 is satisfied. Since the quotient morphisms qI:AA/Iq_{I}\colon A\to A/I are surjective and therefore epimorphisms, condition (ii) is satisfied as well. This completes the proof.

For observing first applications, we let 𝒜{\cal A} be a category 𝒯𝒜{\cal T}\hskip-3.0pt{\cal A} of topological algebraic structures and R:𝒲𝒜𝒯𝒜R\colon{\cal W}\hskip-2.0pt{\cal A}\to{\cal T}\hskip-3.0pt{\cal A} a limit preserving functor. We assume that RR satisfies the Solution Set Condition (see [11] A3.58), as is the case in the examples we discuss below (cf. [2, 7]). Hence a left adjoint functor L:𝒯𝒜𝒲𝒜L\colon{\cal T}\hskip-3.0pt{\cal A}\to{\cal W}\hskip-2.0pt{\cal A} exists (see [11], Theorem A3.60.)

Assume that L:𝒯𝒜𝒲𝒜L\colon{\cal T}\hskip-3.0pt{\cal A}\to{\cal W}\hskip-2.0pt{\cal A} is left adjoint to RR and that ηX:XRL(X)\eta_{X}\colon X\to RL(X) is the front adjunction (see [11], Definition A3.37). Now L(X)L(X) is a weakly complete unital algebra. In practically all examples of interest to us, for a weakly complete unital algebra AA, the 𝒯𝒜{\cal T}\hskip-3.0pt{\cal A}-object R(A)R(A) is a subset of AA such as for instance A×A^{\times} (if 𝒯𝒜{\cal T}\hskip-3.0pt{\cal A} is a category of topological groups), or ALieA_{\rm Lie} (if 𝒯𝒜{\cal T}\hskip-3.0pt{\cal A} is a category of topological Lie algebras), or the underlying topological space |A||A| (if 𝒯𝒜{\cal T}\hskip-3.0pt{\cal A} is a category of topological spaces).

In such a situation ηX:XRL(X)\eta_{X}\colon X\to RL(X) is a function and we can consider its image ηX(X)\eta_{X}(X) as a subset of RL(X)RL(X). Then ηX(X)\langle\eta_{X}(X)\rangle denotes the smallest unital subalgebra containing ηX(X)\eta_{X}(X) and ηX(X)¯\overline{\langle\eta_{X}(X)\rangle} the smallest 𝒲𝒜{\cal W}\hskip-2.0pt{\cal A} subobject of L(X)L(X). Under these circumstances we define a function

Lo:ob(𝒯𝒜)ob(𝒲𝒜) by Lo(X)=ηX(X)¯,L_{o}\colon\mathop{\rm ob}\nolimits({\cal T}\hskip-3.0pt{\cal A})\to\mathop{\rm ob}\nolimits({\cal W}\hskip-2.0pt{\cal A})\mbox{ by }L_{o}(X)=\overline{\langle\eta_{X}(X)\rangle},

the smallest 𝒲𝒜{\cal W}\hskip-2.0pt{\cal A}-subobject for which the morphism ηX:XLR(X)\eta_{X}\colon X\to LR(X) factors through the inclusion morphism R(Lo(X))RL(X)R(L_{o}(X))\to RL(X). Then one observes immediately that LoL_{o} is conditionally left adjoint to RR with respect to 𝒲𝒜{\cal W}\hskip-2.0pt{\cal A}. (Cf. Definition 7.2.) However, here Remark 7.2 applies and shows that the containment Lo(X)L(X)L_{o}(X)\subseteq L(X) is equality in all cases. Thus we have

{Proposition}

Assume that R:𝒲𝒜𝒯𝒜R\colon{\cal W}\hskip-2.0pt{\cal A}\to{\cal T}\hskip-3.0pt{\cal A} and L:𝒯𝒜𝒲𝒜L\colon{\cal T}\hskip-3.0pt{\cal A}\to{\cal W}\hskip-2.0pt{\cal A} is a pair of adjoint functors where 𝒯𝒜{\cal T}\hskip-3.0pt{\cal A} is a category of topological algebraic structures for which the front adjunctions ηX:XRL(X)\eta_{X}\colon X\to RL(X) are functions whose image ηX(X)\eta_{X}(X) is a subset of the weakly complete unital algebra L(X)L(X). Then for each Xob(𝒯𝒜)X\in\mathop{\rm ob}\nolimits({\cal T}\hskip-3.0pt{\cal A}), the abstract unital algebra ηX(X)\langle\eta_{X}(X)\rangle generated by the image of ηX\eta_{X} is dense in the weakly complete unital algebra L(X)L(X).

Our immediate examples for the category 𝒯𝒜{\cal T}\hskip-3.0pt{\cal A} are as follows:

(A)  𝒯𝒜=𝒫ROGR{\cal T}\hskip-3.0pt{\cal A}=\mathcal{P}\hskip-2.0ptROG\hskip-1.0ptR, the category of pro-Lie groups, RA=A×RA=A^{\times} the group of units of the weakly complete algebra AA. For the fact that A×A^{\times} is a pro-Lie group see [2] or [11], Proposition A7.37. The left adjoint L is the weakly complete group algebra G𝕂[G]G\mapsto\mathbb{K}[G] over 𝕂\mathbb{K}. It was discussed in [2], [7], and [11] (mostly for 𝕂=\mathbb{K}=\mathbb{R} and compact groups GG).

A prominent subcategory of 𝒫ROGR\mathcal{P}\hskip-2.0ptROG\hskip-1.0ptR is the full subcategory 𝒞OMPGR\mathcal{C}O\hskip-2.0ptM\hskip-2.0ptPG\hskip-1.0ptR of compact groups for which the real weakly complete group algebra [G]\mathbb{R}[G] is particularly effective. See [2].

(B)  𝒯𝒜=𝒫ROLIE{\cal T}\hskip-3.0pt{\cal A}=\mathcal{P}\hskip-2.0ptRO\hskip-1.0ptLI\hskip-1.0ptE, the category of profinite-dimensional Lie algebras over 𝕂\mathbb{K} (cf. [7], [8]). The functor R:𝒲𝒜𝒫ROLIER\colon{\cal W}\hskip-2.0pt{\cal A}\to\mathcal{P}\hskip-2.0ptRO\hskip-1.0ptLI\hskip-1.0ptE associates with a weakly complete unital algebra AA the profinite-dimensional Lie algebra ALieA_{\rm Lie} defined on the weakly complete underlying weakly complete 𝕂\mathbb{K}-vector space endowed with the Lie algebra multiplication [x,y]=xyyx[x,y]=xy-yx. Then the left adjoint L:𝒫ROLIE𝒲𝒜L\colon\mathcal{P}\hskip-2.0ptRO\hskip-1.0ptLI\hskip-1.0ptE\to{\cal W}\hskip-2.0pt{\cal A} is the weakly complete universal enveloping algebra 𝔤𝐔𝕂(𝔤)\mathfrak{g}\mapsto\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{K}}(\mathfrak{g}) over 𝕂\mathbb{K} [7, 8] which we shall address again below.

(C) 𝒯𝒜=𝒲{\cal T}\hskip-3.0pt{\cal A}={\cal W}, the category of weakly complete 𝕂\mathbb{K}-vector spaces. The functor R:𝒲𝒜𝒲R\colon{\cal W}\hskip-2.0pt{\cal A}\to{\cal W} associates with a weakly complete unital algebra AA the underlying weakly complete topological 𝕂\mathbb{K}-vector space |A||A|. The left adjoint L:𝒲𝒲𝒜L\colon{\cal W}\to{\cal W}\hskip-2.0pt{\cal A} of RR is, as we shall discuss in the subsequent section, the functor which associates with any weakly complete vector space WW the weakly complete tensor algebra 𝐓(W)\mathop{\bf T\hphantom{}}\nolimits(W) of WW over 𝕂\mathbb{K}.

{Proposition}

(i) In each of the categories 𝒯𝒜=𝒞OMPGR{\cal T}\hskip-3.0pt{\cal A}=\mathcal{C}O\hskip-2.0ptM\hskip-2.0ptPG\hskip-1.0ptR, 𝒫ROLIE\mathcal{P}\hskip-2.0ptRO\hskip-1.0ptLI\hskip-1.0ptE, and 𝒲{\cal W}, a monomorphism f:XYf\colon X\to Y induces an isomorphism Xf(X)X\to f(X) onto the image, that is, an embedding in the respective category 𝒯𝒜{\cal T}\hskip-3.0pt{\cal A}.

(ii) The front adjunction ηX:XRL(X)\eta_{X}\colon X\to RL(X), namely,

G𝕂[G]×G\to\mathbb{K}[G]^{\times}, 𝔤𝐔K(𝔤)Lie\mathfrak{g}\to\mathop{\bf U\hphantom{}}\nolimits_{K}(\mathfrak{g})_{\rm Lie}, and W|𝐓(W)|W\to|\mathop{\bf T\hphantom{}}\nolimits(W)|,

is an embedding in the respective category. That is, XX may be considered as a 𝒯𝒜{\cal T}\hskip-3.0pt{\cal A}-subobject of RL(X)RL(X) and a subset of L(X)L(X).

(iii) If Xob(𝒯𝒜)X\in\mathop{\rm ob}\nolimits({\cal T}\hskip-3.0pt{\cal A}) and XL(X)X\subset L(X) as in (ii) above, then X\langle X\rangle, the abstract unital algebra generated by XX in L(X)L(X) is dense in L(X)L(X).

{Proof}

Part (i) may be safely considered as an exercise; for the two categories 𝒫ROLIE\mathcal{P}\hskip-2.0ptRO\hskip-1.0ptLI\hskip-1.0ptE and 𝒲{\cal W} see also [11], Theorem A7.12.

Part (ii) is then a consequence of Part (i), Corollary 7.4 and the following facts which secure that the front adjunction ηX:XRL(X)\eta_{X}\colon X\to RL(X) is injective, hence monic for Xob(𝒯𝒜)X\in\mathop{\rm ob}\nolimits({\cal T}\hskip-3.0pt{\cal A}):

Part(iii) follows from Proposition 7.5.

(a) Every compact Lie group has a faithful linear representation (see e.g. [11], Corollary 2.40).

(b) Every finite-dimensional Lie algebra over a field of characteristic 0 has a faithful linear representation (Ado’s Theorem, see [1], Chap. 1, Paragraph 7, no 3, Théorème 3).

(c) It suffices to observe that the one-dimensional vector space 𝕂\mathbb{K} has a faithful linear representation, e.g.

c(1c01).c\mapsto\begin{pmatrix}1&c\\ 0&1\end{pmatrix}.

We now secure the validity of the hypotheses of Theorem 7.4 for the examples (A), (B), and (C).

{Proposition}

For the functors L=𝕂[]L=\mathbb{K}[-], 𝐔𝕂{\bf U}_{\mathbb{K}}, and 𝐓\bf T the morphism φA:AlimjQ(A)L(Aj)\varphi_{A}\colon A\to\lim_{j\in Q(A)}L(A_{j}) is an isomorphism.

{Proof}

We show that the hypothesis of Theorem 7.4 is satisfied in each of the three examples (A), (B), and (C).

(A) G𝕂[G]:𝒫ROGR𝒲𝒜G\mapsto\mathbb{K}[G]:\mathcal{P}\hskip-2.0ptROG\hskip-1.0ptR\to{\cal W}\hskip-2.0pt{\cal A}, is left adjoint to AA×:𝒲𝒜𝒫ROGRA\mapsto A^{\times}:{\cal W}\hskip-2.0pt{\cal A}\to\mathcal{P}\hskip-2.0ptROG\hskip-1.0ptR. We note that the subcategory {\cal LIE} of Lie groups is strictly dense in 𝒫ROGR\mathcal{P}\hskip-2.0ptROG\hskip-1.0ptR:

Each pro-Lie group GG is the projective limit of its Lie group quotients G/NG/N, N𝒩(G)N\in{\cal N}(G), where 𝒩(G){\cal N}(G) denotes the filter basis of normal subgroups NN of GG such that G/NG/N is a Lie group. If f:GLf\colon G\to L is a morphism of GG into a Lie group, let NN be the kernel of ff. Then ff factors through the quotient morphism q:GG/Nq\colon G\to G/N followed by an injection of Lie groups G/NLG/N\to L. Each quotient morphism GG/NG\to G/N is an epimorphism. So

{qMN:G/NG/M:(M,N)𝒩(G)×𝒩(G),NM}\{q_{MN}:G/N\to G/M:(M,N)\in{\cal N}(G)\times{\cal N}(G),N\subseteq M\}

is appropriate for GG. The functor AA×:𝒲𝒜𝒫ROGRA\mapsto A^{\times}:{\cal W}\hskip-2.0pt{\cal A}\to\mathcal{P}\hskip-2.0ptROG\hskip-1.0ptR maps finite-dimensional algebras to Lie groups. So the hypothesis of Theorem 7.4 is satisfied and so

φG:GlimN𝒩(G)𝕂[G/N]×\varphi_{G}\colon G\to\lim_{N\in{\cal N}(G)}\mathbb{K}[G/N]^{\times}

is an isomorphism. The cases (B) and (C) are equally simple and are left as an exercise.

Theorem 7.4 now has immediately the following corollaries:

{Theorem}

Each pro-Lie group GG has an appropriate projective limit representation G=limjJGjG=\lim_{j\in J}G_{j} in terms of Lie groups. Therefore

𝕂[G]limjJ𝕂[Gj].\mathbb{K}[G]\cong\lim_{j\in J}\mathbb{K}[G_{j}].
{Theorem}

Each profinite-dimensional weakly complete Lie algebra has an appropriate projective limit representation limjJ𝔤j\lim_{j\in J}\mathfrak{g}_{j} in terms of finite-dimensional Lie algebras. Therefore

𝐔𝕂(𝔤)limjJ𝐔𝕂(𝔤j).\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{K}}(\mathfrak{g})\cong\lim_{j\in J}\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{K}}(\mathfrak{g}_{j}).
{Theorem}

If a weakly complete 𝕂\mathbb{K}-vector space WW is represented in terms of an appropriate projective limit representation W=limjJWjW=\lim_{j\in J}W_{j} in terms of finite-dimensional vector spaces. Therefore

𝐓(W)limjJ𝐓(Wj).\mathop{\bf T\hphantom{}}\nolimits(W)\cong\lim_{j\in J}\mathop{\bf T\hphantom{}}\nolimits(W_{j}).

8 Appendix: The Definition of the Tensor Algebra

In Paragraph (C) above we already introduced the tensor algebra of a weakly complete vector space. Let us now review this concept more systematically. So we let 𝕂\mathbb{K} again denote one of the topological fields \mathbb{R} or \mathbb{C}, and 𝒲𝒜{\cal W}\hskip-2.0pt{\cal A} the category of weakly complete associative unital algebras over 𝕂\mathbb{K}.

Here is the definition of the tensor algebra via its universal property:

{Theorem}

(The Existence Theorem of 𝐓\mathop{\bf T\hphantom{}}\nolimits) The underlying weakly complete vector space functor A|A|A\mapsto|A| from 𝒲𝒜{\cal W}\hskip-2.0pt{\cal A} to 𝒲{\cal W} has a left adjoint 𝐓:𝒲𝒲𝒜\mathop{\bf T\hphantom{}}\nolimits\colon{\cal W}\to{\cal W}\hskip-2.0pt{\cal A}.

The front adjunction ωV:V|𝐓(V)|\omega_{V}\colon V\to|\mathop{\bf T\hphantom{}}\nolimits(V)| is an embedding of topological vector spaces.

{Proof}

The category 𝒲{\cal W} is complete. (Exercise. Cf. Theorem A3.48 of [11], p. 819.) The “Solution Set Condition” (of Definition A3.58 in [11], p. 824) holds. (Exercise: Cf. the proof Lemma 3.58 of [11].) Hence 𝐓\mathop{\bf T\hphantom{}}\nolimits exists by the Adjoint Functor Existence Theorem (i.e., Theorem A3.60 of [11], p. 825).

The assertion about ωV\omega_{V} follows from Proposition 7.5 (iii).

In other words, each weakly complete vector space VV may be considered as a weakly complete vector subspace of the weakly complete tensor algebra 𝐓(V)\mathop{\bf T\hphantom{}}\nolimits(V) with the property that each continuous linar map f:V|A|f\colon V\to|A| with some weakly complete associative unital algebra AA and its underlying weakly complete vector space |A||A| extends uniquely to a 𝒲𝒜{\cal W}\hskip-2.0pt{\cal A}-morphism f:𝐓(V)Af^{\prime}\colon\mathop{\bf T\hphantom{}}\nolimits(V)\to A.

𝒲𝒲𝒜V|𝐓(V)|𝐓(V)f|f|!f|A|id|A|A.\begin{matrix}&{\cal W}&&\hbox to19.91692pt{}&{\cal W}\hskip-2.0pt{\cal A}\cr\vskip 3.0pt\cr\hrule\cr\cr\vskip 3.0pt\cr V&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\subseteq}}&|\mathop{\bf T\hphantom{}}\nolimits(V)|&\hbox to19.91692pt{}&\mathop{\bf T\hphantom{}}\nolimits(V)\\ \hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle\forall f$}}$}\Big{\downarrow}&&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle|f^{\prime}|$}}$\hss}&\hbox to19.91692pt{}&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle\exists!f^{\prime}$}}$\hss}\\ |A|&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits_{\mathop{\rm id}\nolimits}}&|A|&\hbox to19.91692pt{}&A.\end{matrix}

If necessary we shall write 𝐓𝕂\mathop{\bf T\hphantom{}}\nolimits_{\mathbb{K}} instead of 𝐓\mathop{\bf T\hphantom{}}\nolimits whenever the ground field should be emphasized.

{Definition}

For each weakly complete 𝕂\mathbb{K}-vector space VV we shall call 𝐓𝕂(V)\mathop{\bf T\hphantom{}}\nolimits_{\mathbb{K}}(V) the weakly complete tensor algebra of VV (over 𝕂\mathbb{K}).

We record what we already saw in Section 1 in Theorem 7.5:

{Corollary}

If VV is represented as a projective limit limjJVj\lim_{j\in J}V_{j} of finite-dimensional vector spaces, then 𝐓(V)limjJ𝐓(Vj)\mathop{\bf T\hphantom{}}\nolimits(V)\cong\lim_{j\in J}\mathop{\bf T\hphantom{}}\nolimits(V_{j}).

Every unital associative algebra AA has injective morphism ιA:𝕂A\iota_{A}\colon\mathbb{K}\to A given by ιA(t)=t1\iota_{A}(t)=t{\cdot}1. In some circumstances, ι\iota is a coretraction:

{Remark}

For every weakly complete vector space VV, the morphism v0:V𝕂v\mapsto 0:V\to\mathbb{K}, according to the definition of 𝕋(V)\mathbb{T}(V), induces a natural 𝒲{\cal W}-morphism αV:𝐓(V)𝕂\alpha_{V}\colon{\bf T}(V)\to\mathbb{K} such that αV(V)={0}\alpha_{V}(V)=\{0\} and that αVι𝐓(V)=id𝕂\alpha_{V}\circ\iota_{{\bf T}(V)}=\mathop{\rm id}\nolimits_{\mathbb{K}}.

The retraction αV\alpha_{V} is frequently called the augmentation. of 𝐓(V){\bf T}(V).

Let us compare the weakly complete tensor algebra with the abstract tensor algebra T(E)T(E) of a plain 𝕂\mathbb{K}-vector space EE. By the universal property of the abstract tensor product T(|V|)T(|V|), the linear inclusion map ξ|V|:|V|T(|V|)\xi_{|V|}\colon|V|\to T(|V|) extends to a unique morphism of unital algebras j|V|:T(|V|)|𝐓(V)|j_{|V|}\colon T(|V|)\to|\mathop{\bf T\hphantom{}}\nolimits(V)| such that

|V|ξ|V|T(|V|)|ωV|jV|𝐓(V)|id|𝐓(V)|\begin{matrix}|V|&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\xi_{|V|}}}&T(|V|)\\ \hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle|\omega_{V}|$}}$}\Big{\downarrow}&&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle j_{V}$}}$\hss}\\ |\mathop{\bf T\hphantom{}}\nolimits(V)|&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits_{\mathop{\rm id}\nolimits}}&|\mathop{\bf T\hphantom{}}\nolimits(V)|\\ \end{matrix} (1)

commutes.

For a natural number mm and a weakly complete vector space VV, set Am=𝒲mVA_{m}=\bigotimes^{m}_{\cal W}V. Then AmA_{m} is a weakly complete vector space. If mm, and nn are natural numbers, then there is a canonical continuous bilinear map

(am,an)aman:=am𝒲an:Am×AnAm+n,(a_{m},a_{n})\mapsto a_{m}a_{n}:=a_{m}\otimes_{\cal W}a_{n}:A_{m}\times A_{n}\to A_{m+n},

where we have identified the naturally isomorphic weakly complete vector spaces Am𝒲AnA_{m}\otimes_{\cal W}A_{n} and Am+nA_{m+n}. Set A0=𝕂A_{0}=\mathbb{K}. Then m=0Am\bigoplus_{m=0}^{\infty}A_{m} is a graded unital algebra DD with the multiplication

(am)m0(bm)m0=(j+k=majbk)m0(a_{m})_{m\in\mathbb{N}_{0}}(b_{m})_{m\in\mathbb{N}_{0}}=(\sum_{j+k=m}a_{j}b_{k})_{m\in\mathbb{N}_{0}},

dense in the weakly complete vector space A(V):=m=0AmA(V):=\prod_{m=0}^{\infty}A_{m}. By the definition of the unital algebra DD, there is a unique injective morphism of unital algebrs iV:T(|V|)=m=0𝒱mV|A(V)|i_{V}\colon T(|V|)=\bigoplus_{m=0}^{\infty}\bigotimes^{m}_{\cal V}V\to|A(V)| so that we have a commutative diagram:

|V|ξ|V|T(|V|)idiVVinclA(V).\begin{matrix}|V|&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\xi_{|V|}}}&T(|V|)\\ \hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle\mathop{\rm id}\nolimits$}}$}\Big{\downarrow}&&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle i_{V}$}}$\hss}\\ V&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits_{\mathop{\rm incl}\nolimits}}&A(V).\\ \end{matrix} (2)

Now, multiplication in DD is continuous w.r.t. the topology induced from A(V)A(V) and therefore extends continuously to a multiplication on A(V)A(V), making A(V)A(V) a weakly complete unital algebra. There is an injective continuous linear map ιV:VA(V)\iota_{V}\colon V\to A(V) given by ιV(v)=(0,v,0,0,)𝕂×A1×A2×=A(V)\iota_{V}(v)=(0,v,0,0,\dots)\in\mathbb{K}\times A_{1}\times A_{2}\times\cdots=A(V) which by Theorem 8 yields a unique morphism of weakly complete unital algebras ιV:T(V)A(V){\iota_{V}}^{\prime}\colon T(V)\to A(V) such that ιV(v)=ιV(ωV(v))\iota_{V}(v)={\iota_{V}}^{\prime}(\omega_{V}(v)) for all vVv\in V, i.e. such that the following diagram commutes:

VωV|𝐓(V)|𝐓(V)ιV|ιV|ιV|A(V)|id|A(V)|A(V).\begin{matrix}V&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\omega_{V}}}&|\mathop{\bf T\hphantom{}}\nolimits(V)|&\hbox to19.91692pt{}&\mathop{\bf T\hphantom{}}\nolimits(V)\\ \hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle\iota_{V}$}}$}\Big{\downarrow}&&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle|{\iota_{V}}^{\prime}|$}}$\hss}&\hbox to19.91692pt{}&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle{\iota_{V}}^{\prime}$}}$\hss}\\ |A(V)|&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits_{\mathop{\rm id}\nolimits}}&|A(V)|&\hbox to19.91692pt{}&A(V).\end{matrix} (3)

We now have jVξ|V|=|ωV|j_{V}\circ\xi_{|V|}=|\omega_{V}| by (1), iVξ|V|=|ιV|i_{V}\circ\xi_{|V|}=|\iota_{V}| by (2), and |ιV|=|ιV||ωV||\iota_{V}|=|{\iota_{V}}^{\prime}|\circ|\omega_{V}| by (3). Therefore iVξ|V|=|ιV|jVξ|V|i_{V}\circ\xi_{|V|}=|{\iota_{V}}^{\prime}|\circ j_{V}\circ\xi_{|V|}, and so the uniqueness in the universal property of T(|V|)T(|V|) allows us to conclude

iV=|ιV|jV.i_{V}=|{\iota_{V}}^{\prime}|\circ j_{V}. (4)

But iVi_{V} is injective, and so jV:T(|V|)|𝐓(V)|j_{V}\colon T(|V|)\to|\mathop{\bf T\hphantom{}}\nolimits(V)| is injective.

Collecting the information we have collected we now arrive at the following insight:

{Lemma}

For each weakly complete vector space VV, the weakly complete unital algebra 𝐓(V)\mathop{\bf T\hphantom{}}\nolimits(V) contains a copy of the algebraic tensor algebra T(|V|)=m=0m|V|T(|V|)=\bigoplus_{m=0}^{\infty}\bigotimes^{m}|V| algebraically generated by V𝐓(V)V\subseteq\mathop{\bf T\hphantom{}}\nolimits(V). {Proof} The completion of the proof is now an exercise.

{Theorem}

(i) For any weakly complete vector space VV, the unital associative subalgebra V\langle V\rangle generated algebraically in 𝐓(V)\mathop{\bf T\hphantom{}}\nolimits(V) by VV is dense in 𝐓(V)\mathop{\bf T\hphantom{}}\nolimits(V).

(ii) Moreover, V\langle V\rangle is algebraically isomorphic to the algebraic tensor algebra T(|V|)T(|V|) generated by |V||V|.

{Proof}

(i) The assertion was proved in Proposition 7.5(iii).

(ii) By the universal property of the algebraic tensor algebra T(|V|)T(|V|) generated by |V||V| there is a morphism jV:T(|V|)|𝐓(V)|j_{V}\colon T(|V|)\to|\mathop{\bf T\hphantom{}}\nolimits(V)| (see (1) above) whose corestriction to its image is a morphism of unital algebras from T(|V|)T(|V|) to V\langle V\rangle which is is injective by Lemma 8 and therefore is an isomorphism of unital algebras.


Let us now use the weakly complete tensor algebra to construct 𝐔𝕂(𝔤)\mathop{\bf U\hphantom{}}\nolimits_{\mathbb{K}}(\mathfrak{g}) as a quotient of T(|𝔤|)T(|\mathfrak{g}|). In the classical theory of universal enveloping algebras of Lie algebras, the construction usually does proceed from the tensor algebra as an origin and progresses to the enveloping algebra as a quotient. In the world 𝒲{\cal W} of weakly complete vector spaces we proceeded systematically via universal properties using category theoretical standard methods. In this fashion we have developed the weakly complete tensor algebra and the weakly complete universal enveloping algebra separately albeit with unified methods. Now let us pause and bring the two together again using the principle of the universal property.


Let 𝔤\mathfrak{g} be a profinite-dimensional Lie algebra and |𝔤||\mathfrak{g}| the weakly complete vector space on which it is based. Let 𝔤\|\mathfrak{g}\| denote the underlying vector space and 𝔤¯\underline{\mathfrak{g}} the underlying abstract Lie algebra. There is a quotient morphism of unital algebras p|𝔤|:T(𝔤)U(𝔤¯)p_{|\mathfrak{g}|}\colon T(\|\mathfrak{g}\|)\to U(\underline{\mathfrak{g}}) well known form the apparatus of the Poincaré-Birkhoff-Witt-Theorem where we may consider 𝔤\|\mathfrak{g}\| as a vector subspace of T(𝔤)T(\|\mathfrak{g}\|). Now we elevate this quotient to the level of the weakly complete unital algebras. From Theorem 1 we know that 𝔤𝐔(𝔤)\mathfrak{g}\subseteq\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}). This give us an embedding of weakly complete vector spaces |𝔤||𝐔(𝔤)||\mathfrak{g}|\to|\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})| where 𝐔(𝔤)\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}) is a weakly complete unital algebra. Then Theorem 8 provides us with a unique morphism q𝔤:𝐓(|𝔤|)𝐔(𝔤)q_{\mathfrak{g}}\colon\mathop{\bf T\hphantom{}}\nolimits(|\mathfrak{g}|)\to\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}) extending the identity function |𝔤|𝔤|\mathfrak{g}|\to\mathfrak{g}. As a morphism of weakly complete algebras, q𝔤q_{\mathfrak{g}} has a closed image (see e.g.  [11], Theorem A7.12), and by Theorem 8 (i) has a dense image. Thus q𝔤q_{\mathfrak{g}} is surjective and thus a quotient map (againby [11], A7.12). We summarize this in the following Theorem whose proof is clear from what we know:

{Theorem}

There is a canonical quotient morphism of weakly complete algebras q𝔤:𝐓(|𝔤|)𝐔(𝔤)q_{\mathfrak{g}}\colon\mathop{\bf T\hphantom{}}\nolimits(|\mathfrak{g}|)\to\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g}) such that

𝔤inclT(𝔤)=𝔤j𝔤𝐓(|𝔤|)idp𝔤¯q𝔤𝔤¯inclU(𝔤¯)=𝔤¯incl𝐔(𝔤)\begin{matrix}\|\mathfrak{g}\|&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{\mathop{\rm incl}\nolimits}}&T(\|\mathfrak{g}\|)&{=}&\langle\|\mathfrak{g}\|\rangle&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits^{j_{\|\mathfrak{g}\|}}}&\mathop{\bf T\hphantom{}}\nolimits(|\mathfrak{g}|)\\ \hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle\mathop{\rm id}\nolimits$}}$}\Big{\downarrow}&&\hbox to0.0pt{\hss$\vbox{\hbox{$\scriptstyle p_{\underline{\mathfrak{g}}}$}}$}\Big{\downarrow}&&&&\Big{\downarrow}\hbox to0.0pt{$\vbox{\hbox{$\scriptstyle q_{\mathfrak{g}}$}}$\hss}\\ \underline{\mathfrak{g}}&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits_{\mathop{\rm incl}\nolimits}}&U(\underline{\mathfrak{g}})&{=}&\langle\underline{\mathfrak{g}}\rangle&\smash{\mathop{\hbox to40.0pt{\rightarrowfill}}\limits_{\mathop{\rm incl}\nolimits}}&\mathop{\bf U\hphantom{}}\nolimits(\mathfrak{g})\\ \end{matrix}

is commutative.

{Remark}

The quotient morphism q𝔤q_{\mathfrak{g}} respects augmentations in the sense that α𝔤q𝔤=α|𝔤|\alpha_{\mathfrak{g}}\circ q_{\mathfrak{g}}=\alpha_{|\mathfrak{g}|}.


9 Appendix: Some facts on weakly complete symmetric Hopf algebras

{Definition}

Let AA be a weakly complete symmetric Hopf algebra, i.e. a group object in the monoidal category (𝒲,W)({\cal W},\otimes_{W}) of weakly complete vector spaces (see [11], Appendix 7 and Definition A3.62), with comultiplication c:AAAc\colon A\to A\otimes A and coidentity k:A𝕂k\colon A\to\mathbb{K}.

(i) An element aAa\in A is called grouplike if c(a)=aac(a)=a\otimes a and k(a)=1k(a)=1. The subgroup of grouplike elements in the group of units A×A^{\times} will be denoted 𝔾(A)\mathbb{G}(A).

(ii) An element aAa\in A is called primitive, if c(a)=a1+1ac(a)=a\otimes 1+1\otimes a. The Lie algebra of primitive elements of ALieA_{\rm Lie} will be denoted (A)\mathbb{P}(A).

Any weakly complete unital algebra AA has an everywhere defined exponential function exp:ALieA×\exp\colon A_{\rm Lie}\to A^{\times} into the pro-Lie group A×A^{\times} of invertible elements defined as expx=1+x+12!x2+13!x3+\exp x=1+x+\frac{1}{2!}{\cdot}x^{2}+\frac{1}{3!}{\cdot}x^{3}+\cdots. As a function exp:ALieA×\exp\colon A_{\rm Lie}\to A^{\times} it is the exponential function of the pro-Lie group A×A^{\times} in the sense of pro-Lie groups.

{Theorem}

(Weakly Complete Symmetric Hopf Algebras) If AA is a weakly complete symmetric Hopf algebra, then the set 𝔾(A)\mathbb{G}(A) of grouplike elements is a closed pro-Lie subgroup of the pro-Lie group A×A^{\times}, and the set (A)\mathbb{P}(A) of primitive elements is a closed Lie subalgebra of the profinite-dimensional Lie algebra ALieA_{\rm Lie} and exp((A))𝔾(A)\exp(\mathbb{P}(A))\subseteq\mathbb{G}(A) in such a fashion that the restriction and corestriction of exp\exp is the exponential function exp𝔾(A):(A)𝔾(A)\exp_{\mathbb{G}(A)}:\mathbb{P}(A)\to\mathbb{G}(A) of the pro-Lie group 𝔾(A)\mathbb{G}(A).

(See e.g. [2], [11], [10].)


A simple observation tells us something about the geometry of the set expA\exp A. Indeed, for t𝕂t\in\mathbb{K} we have exp(t1+x)=etexpx\exp(t{\cdot}1+x)=e^{t}{\cdot}\exp x. Thus

{Remark}

In any weakly complete unital algebra AA, we have

expA={<expA, if 𝕂=,(×)expA, if 𝕂=.\exp A=\begin{cases}\mathbb{R}_{<}{\cdot}\exp A,\mbox{ if $\mathbb{K}=\mathbb{R}$,}\\ (\mathbb{C}^{\times}){\cdot}\exp A,\mbox{ if $\mathbb{K}=\mathbb{C}$.}\end{cases} ()

Now we assume that AA has a coidentity α:A𝕂\alpha\colon A\to\mathbb{K} which is a morphism of unital algebras, and we call AA an augmented algebra. We set 𝕀=kerα={aA:α(a)=0}{\mathbb{I}}=\ker\alpha=\{a\in A:\alpha(a)=0\}. Then 𝕀{\mathbb{I}} is a maximal ideal of AA and A/𝕀𝕂A/{\mathbb{I}}\cong\mathbb{K} and since 𝕂1\mathbb{K}{\cdot}1 is central we have

A=𝕂1𝕀A=\mathbb{K}{\cdot}1\oplus{\mathbb{I}} (1)

as a direct sum of closed subalgebras.

{Lemma}

Let AA be an augmented weakly complete unital algebra. Then 1+𝕀expA1+{\mathbb{I}}\subseteq\exp A.

{Proof}

Let x𝕀x\in{\mathbb{I}} and set a=1x1+𝕀a=1-x\in 1+{\mathbb{I}}. By Lemma 5.2 (ii) we find a morphism φ:𝕂XA\varphi\colon\mathbb{K}\langle X\rangle\to A of weakly complete unital algebras such that φ(X)=x\varphi(X)=x. By Theorem 5.2 (ii), in 𝕂X\mathbb{K}\langle X\rangle the element Y=log(1X)=(log(1Xf))f𝔓𝕂Y=\log(1-X)=(\log(1-X_{f}))_{f\in\mathfrak{P}_{\mathbb{K}}} is well defined. Then exp𝕂X(Y)=1X\exp_{\mathbb{K}\langle X\rangle}(Y)=1-X and so a=1φ(X)=φ(1X)=φ(exp𝕂X(Y))=expA(φ(Y))expA(A)a=1-\varphi(X)=\varphi(1-X)=\varphi(\exp_{\mathbb{K}\langle X\rangle}(Y))=\exp_{A}(\varphi(Y))\in\exp_{A}(A).

{Lemma}

If 𝕂=\mathbb{K}=\mathbb{R} then

<(1+𝕀)=(<1)𝕀1𝕀=A,\mathbb{R}_{<}{\cdot}(1+{\mathbb{I}})=(\mathbb{R}_{<}{\cdot}1)\oplus{\mathbb{I}}\subseteq\mathbb{R}{\cdot}1\oplus{\mathbb{I}}=A,

and if 𝕂=\mathbb{K}=\mathbb{C}, then

(×)(1+𝕀)=(×)(1𝕀)=A𝕀.(\mathbb{C}^{\times}){\cdot}(1+{\mathbb{I}})=(\mathbb{C}^{\times})\,(1\oplus{\mathbb{I}})=A\setminus{\mathbb{I}}.
{Proof}

In view of (1) above, the proof is elementary.

{Theorem}

In any weakly complete algebra AA with augmentation α:A𝕂\alpha\colon A\to\mathbb{K} let 𝕀:=α1(0){\mathbb{I}}:=\alpha^{-1}(0).

  1. (i)

    𝕂=\mathbb{K}=\mathbb{R}: Then exp(A)=α1(<)=(<1)𝕀\exp(A)=\alpha^{-1}(\mathbb{R}_{<})=(\mathbb{R}_{<}{\cdot}1)\oplus{\mathbb{I}}.

  2. (ii)

    𝕂=\mathbb{K}=\mathbb{C}: Then expA=α1(×)=A𝕀=A×\exp A=\alpha^{-1}(\mathbb{C}^{\times})=A\setminus{\mathbb{I}}=A^{\times}

{Proof}

By Lemma 9 we have 1+𝕀exp(A)1+{\mathbb{I}}\subseteq\exp(A). By Lemma 9 e𝕂(1+𝕀)expAe^{\mathbb{K}}{\cdot}(1+{\mathbb{I}})\subseteq\exp A. Since e𝕂=<e^{\mathbb{K}}=\mathbb{R}_{<} for 𝕂=\mathbb{K}=\mathbb{R} and e𝕂=×e^{\mathbb{K}}=\mathbb{C}^{\times} for 𝕂=\mathbb{K}=\mathbb{C}, Lemma 9 completes the proof.

These simple facts complement Theorem 9.


Acknowledgments. The authors are deeply grateful to the referee who has contributed sustantially to the final form of this text in its orthography, typography, and, notably in the context of Theorems 2 and 5.2, in its mathematics.

An essential part of this text was written while the authors were partners in the program Research in Pairs at the Mathematisches Forschungsinstitut Oberwolfach MFO in the Black Forest from February 2 through 22, 2020. The authors are grateful for the environment and infrastructure of MFO which made this research possible.

References

  • [1] Bourbaki, N., «Groupes et algèbres de Lie», Hermann, Paris, 1971.
  • [2] Dahmen, R., and K.H. Hofmann, The Pro-Lie Group Aspect of Weakly Complete Algebras and Weakly Complete Group Hopf Algebras, J. of Lie Theory 29 (2019), 413–455.
  • [3] Dixmier, J., “Enveloping Algebras”, North-Holland, 1977, xvi+375pp.
  • [4] Hamilton, A., A Poincaré-Birkhoff-Witt Theorem for Profinite Pronilpotent Lie Algebras, J. of Lie Theory 29 (2019), 611–618.
  • [5] Hilgert, J., K.H. Hofmann, and J.D. Lawson, Lie Groups, Convex Cones, and Semigroups, Oxford Mathematical Monographs 18, Clarendon Press Oxford, 1989, xxxviii+645pp.
  • [6] Hilgert, J., and K.-H. Neeb, “Structure and Geometry of Lie Groups”, Springer Monographs in Mathematics, Springer NewYork etc., 2012, x+744pp.
  • [7] Hofmann, K.H., and L. Kramer, On Weakly Complete Group Algebras of Compact Groups, J. of Lie Theory 30 (2020), 407–424.
  • [8] Hofmann, K.H., and L. Kramer, On Weakly Complete Enveloping Algebras of Profinite-Dimensional Lie Algebras, Preprint, Mathematisches Forschungsinstitut Oberwolfach, 2020, 22pp.
  • [9] Hofmann, K.H., and S.A. Morris, Sophus Lie’s Third Fundamental Theorem and the Adjoint Functor Theorem, J. of Group Theory 8 (2005), 115–133.
  • [10] Hofmann, K.H., and S.A. Morris, “The Lie Theory of Connected Pro-Lie Groups,–A Structure Theory for Pro-Lie Algebras, Pro-Lie Groups, and Connected Locally Compact Groups”, European Mathematical Society Publishing House, Zürich, 2006, xii+663pp.
  • [11] Hofmann, K.H., and S.A. Morris, “The Structure Theory of Compact Groups–A Primer for the Student–a Handbook for the Expert”, 4th Edition, De Gruyter Studies in Mathematics 25, Berlin/Boston, 2020, xxvi+1006pp.
  • [12] Hofmann, K.H., and S.A. Morris, Advances in the Theory of Compact Groups and Pro-Lie Groups in the last Quarter Century, Axioms 10, 2021, 190 (13pp.).
  • [13] Hofmann, K.H., and W.A.F. Ruppert, Lie Groups and Subsemigroups with Surjective Exponential Function, Memoirs of the Amer.Math.Soc. 130, 1997, viii+174pp.
  • [14] Kelley, J.L., “General Topology”, D. van Nostrand, Princeton, New Jersey, 1966.
  • [15] MacLane, S., “Categories for the Working Mathematician”, Graduate Texts 5, Springer-Verlag, New York etc., 1971, ix+262pp.
  • [16] Montgomery, D., and L. Zippin, “Topological Transformation Groups”, Interscience Publishers, New York, 1955.
  • [17] Mostert, P.S., and A. Shields, Semigroups with identity on a manifold, Trans. Amer. Math. Soc. 91 (1959), 380-389.
  • [18] Serre, J.-P., “Lie Algebras and Lie Groups”, W.A. Benjamin, Inc., New York-Amsterdam, 1965.
  • [19] Troshkin, M., Group-like elements in a universal enveloping algebra, Discussion in mathoverflow, Feb. 4, 2021.