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

A non-Archimedean Arens–Eells isometric embedding theorem on valued fields

Yoshito Ishiki Photonics Control Technology Team RIKEN Center for Advanced Photonics 2-1 Hirasawa, Wako, Saitama 351-0198, Japan [email protected]
Abstract.

In 1959, Arens and Eells proved that every metric space can be isometrically embedded into a real linear space as a closed subset. In later years, Michael pointed out that every metric space can be isometrically embedded into a real linear space as a linear independent subset and provided a short proof of Arens–Eells theorem as an application. In this paper, we prove a non-Archimedean analogue of the Arens–Eells isometric embedding theorem, which states that for every non-Archimedean valued field KK, every ultrametric space can be isometrically embedded into a non-Archimedean valued field that is a valued field extension of KK such that the image of the embedding is algebraically independent over KK. Using Levi-Civita fields, we also show that every Urysohn universal ultrametric sapce has a valued field structure.

Key words and phrases:
Ultrametrics, Isometric embeddings, and Non-Archimedean valued fields
2020 Mathematics Subject Classification:
Primary 54E35, Secondary 54E40, 51F99

1. Introduction

In 1956, Arens and Eells [1] established that for every metric space (X,d)(X,d) there exist a real normed linear space (V,)(V,\|*\|) and an isometric embedding I:XVI\colon X\to V such that I(X)I(X) is closed in VV. Michael [16] pointed out that for every metric space (X,d)(X,d), we can take a real normed linear space (V,)(V,\lVert*\rVert) and an isometric embedding I:XVI\colon X\to V such that I(X)I(X) is linearly independent over VV. Using this observation, Michael provided a short proof of the Arens–Eells theorem.

A metric dd on XX is said to be an ultrametric or a non-Archimedean metric if d(x,y)d(x,z)d(z,y)d(x,y)\leq d(x,z)\lor d(z,y) for all x,y,zXx,y,z\in X, where \lor stands for the maximal operator on \mathbb{R}. A set RR is a range set if 0R0\in R and R[0,)R\subseteq[0,\infty). An ultrametric dd on XX is said to be R-valued if d(x,y)Rd(x,y)\in R for all x,yXx,y\in X. Some authors try to investigate non-Archimedean analogue of theorems on metric spaces such as the Arens–Eells theorem. Megrelishvili and Shlossberg [13] proved a non-Archimedean Arens–Eells theorem, which embeds ultrametric spaces into linear spaces over 𝔽2\mathbb{F}_{2}. In [14, Theorem 4.3], as a improvement of their theorem, they prove a non-Archimedean Arens–Eells theorem on linear spaces over arbitrary non-Archimedean valued fields. In [7, Theorem 1.1], as a non-Archimedean analogue of the Arens–Eells theorem, the author showed that for every range set RR, for every integral domain AA with the trivial absolute value ||\lvert*\rvert (i.e., |x|=1\lvert x\rvert=1 for all x0x\neq 0), for every RR-valued ultrametric space (X,d)(X,d), there exist an RR-valued ultra-normed module (V,)(V,\lVert*\rVert) over (A,||)(A,\lvert*\rvert) and an isometric embedding I:XVI\colon X\to V such that I(X)I(X) is closed and linearly independent over (A,||)(A,\rvert*\rvert). Using this embedding theorem, the author proved a (RR-valued) non-Archimedean analogue of the Hausdorff extension theorem of metrics.

There are other attempts to construct an isometric embedding from an ultrametric spaces into a space with a non-Archimedean algebraic structure. Schikhof [21] established that every ultrametric spaces can be isometrically embedded into a non-Archimedean valued field using the Hahn fields, which is a generalization of a field of formal power series. In [2, Conjecture 5.34], Baroughan raise the following conjecture.

Conjecture 1.1.

Let pp be an odd prime and (X,d)(X,d) be an HpH_{p}-valued metric space, where Hp={0}{pnn}H_{p}=\{0\}\cup\{\,p^{n}\mid n\in\mathbb{Z}\,\}. Then there exist a non-Archimedean Banach algebra (B,)(B,\lVert*\rVert) over p\mathbb{Q}_{p} and an isometry i:XBi\colon X\to B.

In this paper, as a generalization of the Schikhof theorem ([21]) and known non-Archimedean analogues ([7, Theorem 1.1] and [14, Theorem 4.3]) of the Arens–Eells theorem, we prove that for every non-Archimedean valued field (K,K)(K,\rVert*\lVert_{K}), for every metric space (X,d)(X,d), there exist a valud field (L,L)(L,\rVert*\lVert_{L}) and an isometry embedding I:XKI\colon X\to K such that LL is a valued field extension of KK, the set I(X)I(X) is closed in LL, and I(X)I(X) is algebraically independent over KK (Theorem 4.6). A key point of the proof of Theorem 4.6 is to use the notion of pp-adic Hahn fields (pp-adic Mal’cev–Neumann fields), which was first introduced by Poonen [20] as pp-adic analogues of ordinary Hahn fields.

As an application, we also gives an affirmative solution of Conjecture 1.1 (see Theorem 4.8).

The theory of pp-adic Hahn fields have an application to Urysohn universal ultrametric spaces, which is defined as ultrametric spaces possessing high homogeneity. We introduce the concept of pp-adic Levi-Civita fields as subspaces of pp-adic Hahn fields, which is a pp-adic analogue of ordinary Levi-Civita fields (see, for instance, [3]). We also show that if pp is 0 or a prime, then every Urysohn universal ultrametric space has a field structure that is an extension of p\mathbb{Q}_{p}, where we consider that 0=\mathbb{Q}_{0}=\mathbb{Q} and it is equipped with the trivial valuation in the case of p=0p=0 (see Theorem 5.5).

The paper is organized as follows. Section 2 presents notions and notations of metric spaces and valued fields. We introduce Hahn fields and pp-adic Hahn fields, which plays an important role in the proofs of our main results. We also prepare some basic statements on valued fields and Hahn fields. In Section 3, we show statements on algebraic independence in (pp-adic) Hahn fields. Section 4 is devoted to proving Theorem 4.6. We also provide an affirmative solution of Conjecture 1.1. In Section 5, we show that (pp-adic) Levi-Civita fields become Urysohn universal ultrametric spaces. Our arguments in this section are baed on the author’s paper [8].

Acknowledgements.

The author would like to thank to Tomoki Yuji for helpful advices on algebraic arguments.

2. Preliminaries

2.1. Generalities

In this paper, we use the set-theoretic notations of ordinals. For example, for an ordinal α\alpha, we have β<α\beta<\alpha if and only if βα\beta\in\alpha.

2.1.1. Metric spaces

For a metric space (X,d)(X,d), and for a subset of AA of XX, and for xXx\in X, we define d(A,x)=inf{d(a,x)aA}d(A,x)=\inf\{\,d(a,x)\mid a\in A\,\}. For xXx\in X and r(0,)r\in(0,\infty), we denote by B(a,r;d)B(a,r;d)the closed ball centered at xx with radius rr. We often simply represent it as B(a,r)B(a,r) when no confusion can arise. Similarly, we define the open ball U(a,r)U(a,r).

The proofs of the next two lemmas are presented in Propositions 18.2, and 18.4, in [22], respectively.

Lemma 2.1.

Let XX be a set, and dd be a pseudo-metric on XX. Then dd satisfies the strong triangle inequality if and only if for all x,y,zXx,y,z\in X, the inequality d(x,z)<d(z,y)d(x,z)<d(z,y) implies d(z,y)=d(x,y)d(z,y)=d(x,y).

Lemma 2.2.

Let (X,d)(X,d) be a pseudo-ultrametric space, aXa\in X and r[0,)r\in[0,\infty). Then for every qB(p,r)q\in B(p,r), we have B(p,r)=B(q,r)B(p,r)=B(q,r).

2.1.2. Valued rings

Let AA be a commutative ring. We say that a function v:A{}v\colon A\to\mathbb{R}\sqcup\{\infty\} is a (additive) valuation if the following conditions are satisfied:

  1. (1)

    for every xAx\in A, we have v(x)=v(x)=\infty if and only if x=0x=0;

  2. (2)

    for every pair x,yAx,y\in A, we have v(xy)=v(x)+v(y)v(xy)=v(x)+v(y)

  3. (3)

    for every pair x,yAx,y\in A, we have v(x+y)v(x)v(y)v(x+y)\geq v(x)\land v(y), where \land stands for the minimum operator on \mathbb{R}.

If AA is a field, then it is called a valued field. Note that for every valued ring (A,v)(A,v), we can extend the valuation vv to the fractional field KK of AA. Namely, for x=b/aKx=b/a\in K, where a,bAa,b\in A and a0a\neq 0, we define v(x)=v(b)v(a)v(x)=v(b)-v(a). In this case, vv is well-defined and the pair (K,v)(K,v) naturally becomes a valued field. For example, for a prime pp, we define the pp-adic valuation vpv_{p} on \mathbb{Z} by declaring that vpv_{p} is the number of the factor pp in the prime factorization of xx. The completion p\mathbb{Z}_{p} of \mathbb{Z} with respect to vpv_{p} is called the ring of pp-adic integers. The fractional field of p\mathbb{Z}_{p} is called the field of pp-adic numbers. For more discussion on pp-adic numbers, we refer the readers to [19] and [22].

We say that a function :A[0,)\lVert*\rVert\colon A\to[0,\infty) is a (non-Archimedean) absolute value or multiplicative valuation if the following conditions are satisfied:

  1. (1)

    for every xAx\in A, we have x=0\lVert x\rVert=0 if and only if x=0x=0;

  2. (2)

    for every pair x,yAx,y\in A, we have xy=xy\lVert xy\rVert=\lVert x\rVert\cdot\lVert y\rVert

  3. (3)

    for every pair x,yAx,y\in A, we have x+yxy\lVert x+y\rVert\leq\lVert x\rVert\lor\lVert y\rVert, where \lor stands for the maximum operator on \mathbb{R}.

In what follows, for every η(1,)\eta\in(1,\infty), we consider that η=0\eta^{-\infty}=0 and logη(0)=-\log_{\eta}(0)=\infty. For a valuation vv on a ring AA, and for a real number η(1,)\eta\in(1,\infty), we define xv,η=ηv(x)\lVert x\rVert_{v,\eta}=\eta^{-v(x)} and v,η(x)=logη(x)v_{\rVert*\rVert,\eta}(x)=-\log_{\eta}(x).

Fix η(1,)\eta\in(1,\infty), valuations and absolute values on a ring AA are essentially equivalent. Namely, they are corresponding to each other as follows.

Proposition 2.3.

Let AA be a commutative ring. Then the following statements are true:

  1. (1)

    For every η(1,)\eta\in(1,\infty), and for every valuation vv on AA, the function v,η:A[0,)\lVert*\rVert_{v,\eta}\colon A\to[0,\infty) is an absolute value on AA;

  2. (2)

    For η(1,)\eta\in(1,\infty), and for every absolute value \lVert*\rVert on AA, the function v,η:Av_{\rVert*\rVert,\eta}\colon A\to\mathbb{R} is a valuation on AA.

In this paper, we use both of valuation and absolute values on a ring due to Proposition 2.3. Since we represent a valuation (resp. an absolute value) as a symbol like vv, or ww (resp. ||\rvert*\lvert or \rVert*\lVert), the readers will be able to distinguish them.

For a valued ring (K,v)(K,v), the set 𝔄(K,v)={xK0v(x)}\mathfrak{A}(K,v)=\{\,x\in K\mid 0\leq v(x)\,\} becomes a ring and 𝔬(K,v)={xK0<v(x)}\mathfrak{o}(K,v)=\{\,x\in K\mid 0<v(x)\,\} is a maximal ideal of 𝔄(K,v)\mathfrak{A}(K,v). We denote by 𝔎(K,v)\mathfrak{K}(K,v) the field 𝔄(K,v)/𝔬(K,v)\mathfrak{A}(K,v)/\mathfrak{o}(K,v), and we call it the residue class field of (K,v)(K,v). We also denote by ζK:𝔄(K,v)𝔎(K,v)\zeta_{K}\colon\mathfrak{A}(K,v)\to\mathfrak{K}(K,v) the canonical projection. We simply represent it as ζ\zeta when no confusions can arise. We say that a subset JJ of KK is a complete system of representative of the residue class field 𝔎(A,v)\mathfrak{K}(A,v) if J𝔄(A,v)J\subseteq\mathfrak{A}(A,v), 0J0\in J, and ζ|J:J𝔎(A,v)\zeta|_{J}\colon J\to\mathfrak{K}(A,v) is bijective.

2.2. Constructions of valued fields

In this section, we review some constructions of valued fields such as Hahn fields. For more discussion, we refer the readers to [3] and [4]. Most of the proofs in this section is refered to [20] and [23].

2.2.1. Hahn rings and fields

A non-empty subset SS of \mathbb{R} is said to be well-ordered if every non-empty subset of SS has a minimum.

We denote by 𝒢\mathcal{G} the set of all subgroups of \mathbb{R} containing 11\in\mathbb{R} (equivalently, G\mathbb{Z}\subseteq G). For the sake of convenience, we only consider the setting where G𝒢G\in\mathcal{G} in this paper.

Now we review the construction of the Hahn fields in [20]. Let G𝒢G\in\mathcal{G} and AA be a commutative ring. For a map a:GAa\colon G\to A, we define the support supp(a)\text{{\rm supp}}(a) of aa by the set {xGa(x)0}\{\,x\in G\mid a(x)\neq 0\,\}. We denote by (G,A)\mathbb{H}(G,A) the set of all a:GAa\colon G\to A such that supp(a)\text{{\rm supp}}(a) is well-ordered. We often symbolically represent a(G,A)a\in\mathbb{H}(G,A) as a=gGa(g)tga=\sum_{g\in G}a(g)t^{g}, where tt is an indeterminate. For every pair a,b(G,A)a,b\in\mathbb{H}(G,A), we define a+ba+b by

(a+b)(x)=a(x)+b(x).(a+b)(x)=a(x)+b(x).

We also define ab:GAab\colon G\to A by

ab=gG(i,jG,i+j=gaibj)tg.ab=\sum_{g\in G}\left(\sum_{i,j\in G,i+j=g}a_{i}b_{j}\right)t^{g}.

Define a valuation vG,Av_{G,A} on (G,A)\mathbb{H}(G,A) by vG,A(a)=minsupp(a)v_{G,A}(a)=\min\text{{\rm supp}}(a). Since supp(a)\text{{\rm supp}}(a) is well-ordered, the minimum minsupp(a)\min\text{{\rm supp}}(a) actually exists. Note that AA becomes a subring of (G,A)\mathbb{H}(G,A).

Proposition 2.4.

Let G𝒢G\in\mathcal{G}, and 𝐤\boldsymbol{k} be a field. The pair ((G,𝐤),vG,𝐤)(\mathbb{H}(G,\boldsymbol{k}),v_{G,\boldsymbol{k}}) becomes a valued field and it satisfies that 𝔎((G,𝐤),vG,𝐤)=𝐤\mathfrak{K}(\mathbb{H}(G,\boldsymbol{k}),v_{G,\boldsymbol{k}})=\boldsymbol{k}.

Proof.

See [20, Corollary 1]. ∎

We call ((G,A),vG,A)(\mathbb{H}(G,A),v_{G,A}) the Hahn ring associated with GG and AA and call it the Hahn field if AA is a field. Note that in general, we can define the Hahn fields even if GG is a linearly ordered Abelian group (see [20]).

2.2.2. The pp-adic Hahn fields

A pp-adic analogue of the Hahn fields was first introduced in [20]. Let us review a construction. A field 𝒌\boldsymbol{k} of characteristic pp is said to be perfect if p=0p=0, or p>0p>0 and every element of 𝒌\boldsymbol{k} has a pp-th root in 𝒌\boldsymbol{k}. The following proposition states the existence of rings of Witt vectors. The proof is presented in [23].

Proposition 2.5.

Let 𝐤\boldsymbol{k} be a perfect field of characteristic p>0p>0. The there exists a unique valued ring (A,v)(A,v) of characteristic 0 equipped with a valuation vv such that v(A)=0v(A)=\mathbb{Z}_{\geq 0}, vv is complete, and 𝔎(A,v)=𝐤\mathfrak{K}(A,v)=\boldsymbol{k}.

For each perfect field 𝒌\boldsymbol{k}, we denote by (𝕎(𝒌),w𝒌)(\mathbb{W}(\boldsymbol{k}),w_{\boldsymbol{k}}) the valuation field stated in Proposition 2.5 and denote by Fr𝕎(𝒌)\mathrm{Fr}\mathbb{W}(\boldsymbol{k}) the fractional field of 𝕎(𝒌)\mathbb{W}(\boldsymbol{k}). We use the same symbol w𝒌w_{\boldsymbol{k}} as the valuation on 𝕎(𝒌)\mathbb{W}(\boldsymbol{k}) induced by w𝒌w_{\boldsymbol{k}} in the canonical way. The ring (𝕎(𝒌),w𝒌)(\mathbb{W}(\boldsymbol{k}),w_{\boldsymbol{k}}) is called the ring of Witt vectors associated with 𝐤\boldsymbol{k}. Notice that for every prime pp, we have 𝕎(𝔽p)=p\mathbb{W}(\mathbb{F}_{p})=\mathbb{Z}_{p} and Fr𝕎(𝔽p)=p\mathrm{Fr}\mathbb{W}(\mathbb{F}_{p})=\mathbb{Q}_{p}, and the valuation w𝔽pw_{\mathbb{F}_{p}} coincides with the pp-adic valuation vpv_{p}.

The next proposition explains the concrete representation of an elements of a ring of Witt vectors.

Proposition 2.6.

Let 𝐤\boldsymbol{k} be a perfect field of characteristic p>0p>0. Then there uniquely exists a map f𝐤:𝐤𝕎(𝐤)f_{\boldsymbol{k}}\colon\boldsymbol{k}\to\mathbb{W}(\boldsymbol{k}) such that {f𝐤(a)a𝐤}\{\,f_{\boldsymbol{k}}(a)\mid a\in\boldsymbol{k}\,\} is a complete system of representatives, and f𝐤(ab)=f𝐤(a)f𝐤(b)f_{\boldsymbol{k}}(ab)=f_{\boldsymbol{k}}(a)f_{\boldsymbol{k}}(b) for all a,b𝐤a,b\in\boldsymbol{k}. In this case, for every x𝕎(𝐤)x\in\mathbb{W}(\boldsymbol{k}), there uniquely exists a sequence {ai}i\{a_{i}\}_{i\in\mathbb{Z}} in 𝐤\boldsymbol{k} such that x=nf𝐤(an)pnx=\sum_{n\in\mathbb{Z}}f_{\boldsymbol{k}}(a_{n})p^{n} and for a sufficient large m0m\in\mathbb{Z}_{\geq 0}, we have ai=0a_{i}=0 for all i<mi<-m.

Proof.

See Proposition 8 in the page 35 and and the argument in the page 37 in [23]. ∎

Proposition 2.7.

Let 𝐤\boldsymbol{k} and 𝐥\boldsymbol{l} be perfect fields of characteristic p>0p>0. For every homomorphism ϕ:𝐤𝐥\phi\colon\boldsymbol{k}\to\boldsymbol{l}, there uniquely exists a homomorphism 𝕎(ϕ):𝕎(𝐤)𝕎(𝐥)\mathbb{W}(\phi)\colon\mathbb{W}(\boldsymbol{k})\to\mathbb{W}(\boldsymbol{l}) such that ζ𝕎(𝐥)𝕎(ϕ)=ϕζ𝕎(𝐤)\zeta_{\mathbb{W}(\boldsymbol{l})}\circ\mathbb{W}(\phi)=\phi\circ\zeta_{\mathbb{W}(\boldsymbol{k})}. Moreover, if x=nf𝐤(an)pn𝕎(ϕ)x=\sum_{n\in\mathbb{Z}}f_{\boldsymbol{k}}(a_{n})p^{n}\in\mathbb{W}(\phi), where an𝐤a_{n}\in\boldsymbol{k}, then 𝕎(ϕ)(x)=nf𝐥(ϕ(an))pn\mathbb{W}(\phi)(x)=\sum_{n\in\mathbb{Z}}f_{\boldsymbol{l}}(\phi(a_{n}))p^{n}. In particular, we have w𝐥(𝕎(𝐤)(x))=w𝐤(x)w_{\boldsymbol{l}}(\mathbb{W}(\boldsymbol{k})(x))=w_{\boldsymbol{k}}(x) for all x𝕎(𝐤)x\in\mathbb{W}(\boldsymbol{k}).

Proof.

See [23, Proposition 10 in the page 39]. ∎

Remark 2.1.

It is known that the construction of rings of Witt vectors is a functor (see [23, Page 39]).

Now we discuss a pp-adic analogue of Hahn fields, which is defined by a quotient field of a Hahn ring. For G𝒢G\in\mathcal{G}, and for a perfect field 𝒌\boldsymbol{k} of characteristic p>0p>0, we define a subset G,𝒌\mathbb{N}_{G,\boldsymbol{k}} of (G,𝕎(𝒌))\mathbb{H}(G,\mathbb{W}(\boldsymbol{k})) by the set of all α=gGαgtg(G,𝕎(𝒌))\alpha=\sum_{g\in G}\alpha_{g}t^{g}\in\mathbb{H}(G,\mathbb{W}(\boldsymbol{k})) such that nαg+npn=0\sum_{n\in\mathbb{Z}}\alpha_{g+n}p^{n}=0 in 𝕎(𝒌)\mathbb{W}(\boldsymbol{k}) for every gGg\in G.

Proposition 2.8.

For every G𝒢G\in\mathcal{G}, and for every perfect field 𝐤\boldsymbol{k} with characteristic p>0p>0, the set G,𝐤\mathbb{N}_{G,\boldsymbol{k}} is an ideal of the ring (G,𝕎(𝐤))\mathbb{H}(G,\mathbb{W}(\boldsymbol{k})), and (G,𝕎(𝐤))/G,𝐤\mathbb{H}(G,\mathbb{W}(\boldsymbol{k}))/\mathbb{N}_{G,\boldsymbol{k}} is a filed.

Proof.

See [20, Proposition 3 ] and [20, Corollary 3]. ∎

Lemma 2.9.

Let G𝒢G\in\mathcal{G}, pp be a prime, and 𝐤\boldsymbol{k} be a field of characteristic pp. Let J𝕎(𝐤)J\subseteq\mathbb{W}(\boldsymbol{k}) be a complete system of representatives of the residue class field of 𝐤\boldsymbol{k}. Then every element α=gGαgtg(G,𝕎(𝐤))\alpha=\sum_{g\in G}\alpha_{g}t^{g}\in\mathbb{H}(G,\mathbb{W}(\boldsymbol{k})) is equivalent to an element β=gGβgtg\beta=\sum_{g\in G}\beta_{g}t^{g} modulo G,K\mathbb{N}_{G,K}, where βg\beta_{g} is in JJ. In addition, for every x(G,𝕎(𝐤))x\in\mathbb{H}(G,\mathbb{W}(\boldsymbol{k})), the family {βg}gG\{\beta_{g}\}_{g\in G} is unique and supp(β)supp(α)+0\text{{\rm supp}}(\beta)\subseteq\text{{\rm supp}}(\alpha)+\mathbb{Z}_{\geq 0}.

Proof.

See [20, Proposition 4]. ∎

Definition 2.1.

Based on Lemma 2.9, for a fixed complete system JJ of representatives, and for every a(G,𝕎(𝒌))a\in\mathbb{H}(G,\mathbb{W}(\boldsymbol{k})), we denote by StG,𝒌,J(a)\mathrm{St}_{G,\boldsymbol{k},J}(a) the standard representation of aa with respect to JJ stated in Lemma 2.9. In this case, we have supp(StG,𝒌,J(a))supp(a)+0\text{{\rm supp}}(\mathrm{St}_{G,\boldsymbol{k},J}(a))\subseteq\text{{\rm supp}}(a)+\mathbb{Z}_{\geq 0}.

Let p(G,𝒌)\mathbb{P}_{p}(G,\boldsymbol{k}) denote the quotient field (G,𝕎(𝒌))/G,𝒌\mathbb{H}(G,\mathbb{W}(\boldsymbol{k}))/\mathbb{N}_{G,\boldsymbol{k}}, and let Pr:(G,𝕎(𝒌))p(G,𝒌)\mathrm{Pr}\colon\mathbb{H}(G,\mathbb{W}(\boldsymbol{k}))\to\mathbb{P}_{p}(G,\boldsymbol{k}) denote the canonical projection. We define VG,𝒌,p:p(G,𝒌)GV_{G,\boldsymbol{k},p}\colon\mathbb{P}_{p}(G,\boldsymbol{k})\to G by VG,𝒌,p(x)=minsupp(StG,𝒌,J(x))V_{G,\boldsymbol{k},p}(x)=\min\text{{\rm supp}}(\mathrm{St}_{G,\boldsymbol{k},J}(x)).

Proposition 2.10.

Let G𝒢G\in\mathcal{G}, and 𝐤\boldsymbol{k} be a perfect field of characteristic p>0p>0. Take a complete system J𝕎(𝐤)J\subseteq\mathbb{W}(\boldsymbol{k}) of representatives of the residue class field 𝐤\boldsymbol{k}. Then the following statements are true:

  1. (1)

    For every x(G,𝕎(𝒌))x\in\mathbb{H}(G,\mathbb{W}(\boldsymbol{k})), the value minsupp(StG,𝒌,J(x))\min\text{{\rm supp}}(\mathrm{St}_{G,\boldsymbol{k},J}(x)) is independent from the choice of JJ.

  2. (2)

    The map VG,𝒌,pV_{G,\boldsymbol{k},p} is a valuation on p(G,𝒌)\mathbb{P}_{p}(G,\boldsymbol{k}).

Proof.

See [20, Proposition 5]. ∎

We call the valued field (p(G,𝒌),VG,𝒌,p)(\mathbb{P}_{p}(G,\boldsymbol{k}),V_{G,\boldsymbol{k},p}) the pp-adic Mal’cev–Neumann field or pp-adic Hahn field. Notice that (p(,𝔽p),V,𝔽p,p)(\mathbb{P}_{p}(\mathbb{Z},\mathbb{F}_{p}),V_{\mathbb{Z},\mathbb{F}_{p},p}) is nothing but the (p,vp)(\mathbb{Q}_{p},v_{p}) field of pp-adic numbers.

To consider characteristics of a valued field and its residue class field, we supplementally define 𝒞\mathcal{CH} by the set of all pairs (q,p)(q,p) such that qq and pp are 0 or a prime satisfying the either of the following conditions:

  1. (Q1)

    q=pq=p,

  2. (Q2)

    q=0q=0 and 0<p0<p.

Note that (q,p)𝒞(q,p)\in\mathcal{CH} satisfies (Q2) if and only if qpq\neq p.

In order to discuss pp-adic and ordinary Hahn fields in unified manner, we make a notation as follows.

Definition 2.2.

Let G𝒢G\in\mathcal{G}, (q,p)𝒞(q,p)\in\mathcal{CH}, and let 𝒌\boldsymbol{k} be a perfect field of characteristic pp. We define a field 𝔸q,p(G,𝒌)\mathbb{A}_{q,p}(G,\boldsymbol{k}) by

𝔸q,p(G,𝒌)={(G,𝒌)if q=p;p(G,𝒌)if qp.\mathbb{A}_{q,p}(G,\boldsymbol{k})=\begin{cases}\mathbb{H}(G,\boldsymbol{k})&\text{if $q=p$;}\\ \mathbb{P}_{p}(G,\boldsymbol{k})&\text{if $q\neq p$.}\end{cases}

We also define a valuation UG,𝒌,q,pU_{G,\boldsymbol{k},q,p} on 𝔸q,p(G,𝒌)\mathbb{A}_{q,p}(G,\boldsymbol{k}) by

UG,𝒌,q,p={vG,𝒌if q=p;VG,𝒌,pif qp.U_{G,\boldsymbol{k},q,p}=\begin{cases}v_{G,\boldsymbol{k}}&\text{if $q=p$;}\\ V_{G,\boldsymbol{k},p}&\text{if $q\neq p$.}\end{cases}

A metric space (X,d)(X,d) is said to be spherically complete if for every sequence of (closed or open) balls {Bi}i0\{B_{i}\}_{i\in\mathbb{Z}_{\geq 0}} with Bi+1Bi+1B_{i+1}\subseteq B_{i+1} for all n0n\in\mathbb{Z}_{\geq 0}, we have i0Bi\bigcap_{i\in\mathbb{Z}_{\geq 0}}B_{i}\neq\emptyset.

Proposition 2.11.

Let G𝒢G\in\mathcal{G}, (q,p)𝒞(q,p)\in\mathcal{CH}, and 𝐤\boldsymbol{k} be a field of characteristic pp. Then the following statements are true:

  1. (1)

    UG,𝒌,q,p(𝔸q,p(G,𝒌))=GU_{G,\boldsymbol{k},q,p}(\mathbb{A}_{q,p}(G,\boldsymbol{k}))=G;

  2. (2)

    𝔎(𝔸q,p(G,𝒌),UG,𝒌,q,p)=𝒌\mathfrak{K}(\mathbb{A}_{q,p}(G,\boldsymbol{k}),U_{G,\boldsymbol{k},q,p})=\boldsymbol{k};

  3. (3)

    (𝔸q,p(G,𝒌),UG,𝒌,q,p)(\mathbb{A}_{q,p}(G,\boldsymbol{k}),U_{G,\boldsymbol{k},q,p}) is spherically complete. In particular, it is complete.

Proof.

The statements (1) and (2) follows from the construction of (𝔸q,p(G,𝒌),UG,𝒌,q,p)(\mathbb{A}_{q,p}(G,\boldsymbol{k}),U_{G,\boldsymbol{k},q,p}). The statement (3) is proven by [20, Theorem 1], and [12, Theorem 4] (see also [3, Theorem 6.11]). ∎

Next we consider homeomorphic embeddings between pp-adic or ordinary Hahn fields. We begin with ordinary ones. Let G,H𝒢G,H\in\mathcal{G} with GHG\subseteq H, and A,BA,B be commutative rings. We represent ι\iota as the inclusion map GHG\to H. Let ϕ:AB\phi\colon A\to B be a ring homomorphism. For x=gGxgtg(G,A)x=\sum_{g\in G}x_{g}t^{g}\in\mathbb{H}(G,A), we define (ι,ϕ)(x)(H,B)\mathbb{H}(\iota,\phi)(x)\in\mathbb{H}(H,B) by (ι,ϕ)(x)=hHyhth\mathbb{H}(\iota,\phi)(x)=\sum_{h\in H}y_{h}t^{h}, where

yh={ϕ(xh)if hH;0if hH.y_{h}=\begin{cases}\phi(x_{h})&\text{if $h\in H$;}\\ 0&\text{if $h\not\in H$.}\end{cases}

Then (ι,ϕ):(G,A)(H,B)\mathbb{H}(\iota,\phi)\colon\mathbb{H}(G,A)\to\mathbb{H}(H,B) becomes a map. If G=HG=H, we simply write it as (G,ϕ)\mathbb{H}(G,\phi). Let us observe properties of (ι,ϕ)\mathbb{H}(\iota,\phi).

Proposition 2.12.

Let A,BA,B be commutative rings, and ϕ:AB\phi\colon A\to B be a ring homomorphism. Let G,H𝒢G,H\in\mathcal{G} such that GHG\subseteq H, and denote by ι\iota the inclusion map GHG\to H. Then the map (ι,ϕ):(G,A)(H,B)\mathbb{H}(\iota,\phi)\colon\mathbb{H}(G,A)\to\mathbb{H}(H,B) is a ring homomorphism and satisfies ζ(H,B)(ι,ϕ)=ϕζ(G,A)\zeta_{\mathbb{H}(H,B)}\circ\mathbb{H}(\iota,\phi)=\phi\circ\zeta_{\mathbb{H}(G,A)} on 𝔄((G,A),vG,A)\mathfrak{A}(\mathbb{H}(G,A),v_{G,A}).

Proof.

The lemma follows from the definitions of (ι,ϕ)\mathbb{H}(\iota,\phi) and Hahn fields. ∎

Next we discuss pp-adic Hahn fields, which is defined by quotient fields of Hahn rings.

Proposition 2.13.

Let G,H𝒢G,H\in\mathcal{G}, 𝐤\boldsymbol{k} and 𝐥\boldsymbol{l} be perfect field with characteristic p>0p>0 and ϕ:𝐤𝐥\phi\colon\boldsymbol{k}\to\boldsymbol{l} is a homomorphism. Then the homomorphism (ι,𝕎(ϕ)):(G,𝕎(𝐤))(H,𝕎(𝐥))\mathbb{H}(\iota,\mathbb{W}(\phi))\colon\mathbb{H}(G,\mathbb{W}(\boldsymbol{k}))\to\mathbb{H}(H,\mathbb{W}(\boldsymbol{l})) satisfies

(ι,𝕎(ϕ))(G,𝒌)H,𝒍.\mathbb{H}(\iota,\mathbb{W}(\phi))(\mathbb{N}_{G,\boldsymbol{k}})\subseteq\mathbb{N}_{H,\boldsymbol{l}}.

in particular, the map (ι,𝕎(ϕ))\mathbb{H}(\iota,\mathbb{W}(\phi)) induces a homomorphism

p(ι,ϕ):p(G,𝒌)p(H,𝒍)\mathbb{P}_{p}(\iota,\phi)\colon\mathbb{P}_{p}(G,\boldsymbol{k})\to\mathbb{P}_{p}(H,\boldsymbol{l})

such that ζp(H,𝐥)p(ι,ϕ)=ϕζp(G,𝐤)\zeta_{\mathbb{P}_{p}(H,\boldsymbol{l})}\circ\mathbb{P}_{p}(\iota,\phi)=\phi\circ\zeta_{\mathbb{P}_{p}(G,\boldsymbol{k})} on 𝔄(p(G,𝐤),VG,𝐤,p)\mathfrak{A}(\mathbb{P}_{p}(G,\boldsymbol{k}),V_{G,\boldsymbol{k},p}).

Proof.

Take xG,𝒌x\in\mathbb{N}_{G,\boldsymbol{k}} and put x=gGx(g)pgx=\sum_{g\in G}x(g)p^{g}. Then for every gGg\in G we have nx(g+n)pn=0\sum_{n\in\mathbb{Z}}x(g+n)p^{n}=0 in 𝕎(𝒌)\mathbb{W}(\boldsymbol{k}). Note that for a fixed gGg\in G and for a sufficient large m0m\in\mathbb{Z}_{\geq 0}, we have x(g+n)=0x(g+n)=0 for all n<mn<-m since {gGx(g)0}\{\,g\in G\mid x(g)\neq 0\,\} is well-ordered. By the strong triangle inequality, nx(g+n)pn=0\sum_{n\in\mathbb{Z}}x(g+n)p^{n}=0 is equivalent to x(g+n)pn0x(g+n)p^{n}\to 0 in 𝕎(𝒌)\mathbb{W}(\boldsymbol{k}) as nn\to\infty (see [3, Theorem 2.24]). Since w𝒍(𝕎(ϕ)(x)=w𝒌(x)w_{\boldsymbol{l}}(\mathbb{W}(\phi)(x)=w_{\boldsymbol{k}}(x) for all x𝕎(𝒌)x\in\mathbb{W}(\boldsymbol{k}) (see Proposition 2.7), we also have 𝕎(ϕ)(x(g+n))pn0\mathbb{W}(\phi)(x(g+n))p^{n}\to 0 in 𝕎(𝒍)\mathbb{W}(\boldsymbol{l}) as nn\to\infty. Thus the infinite sum n𝕎(ϕ)(x(g+n))pn\sum_{n\in\mathbb{Z}}\mathbb{W}(\phi)(x(g+n))p^{n} is convergent and we have

n𝕎(ϕ)(x(g+n))pn=𝕎(ϕ)(nx(g+n)pn)=𝕎(ϕ)(0)=0.\sum_{n\in\mathbb{Z}}\mathbb{W}(\phi)(x(g+n))p^{n}=\mathbb{W}(\phi)\left(\sum_{n\in\mathbb{Z}}x(g+n)p^{n}\right)=\mathbb{W}(\phi)(0)=0.

This shows that (ι,𝕎(ϕ))(x)H,𝒍\mathbb{H}(\iota,\mathbb{W}(\phi))(x)\in\mathbb{N}_{H,\boldsymbol{l}}, and hence

(ι,𝕎(ϕ))(G,𝒌)H,𝒍.\mathbb{H}(\iota,\mathbb{W}(\phi))(\mathbb{N}_{G,\boldsymbol{k}})\subseteq\mathbb{N}_{H,\boldsymbol{l}}.

In particular the map (ι,𝕎(ϕ))\mathbb{H}(\iota,\mathbb{W}(\phi)) induces a map p(ι,ϕ):p(G,𝒌)p(G,𝒌)\mathbb{P}_{p}(\iota,\phi)\colon\mathbb{P}_{p}(G,\boldsymbol{k})\to\mathbb{P}_{p}(G,\boldsymbol{k}).

Proposition 2.12 implies that a map (ι,𝕎(ϕ)):(H,𝕎(𝒌))(G,𝕎(𝒍))\mathbb{H}(\iota,\mathbb{W}(\phi))\colon\mathbb{H}(H,\mathbb{W}(\boldsymbol{k}))\to\mathbb{H}(G,\mathbb{W}(\boldsymbol{l})) satisfies ζ(ι,ϕ)=ϕζ\zeta\circ\mathbb{H}(\iota,\phi)=\phi\circ\zeta. Then ζp(ι,ϕ)=ϕζ\zeta\circ\mathbb{P}_{p}(\iota,\phi)=\phi\circ\zeta. ∎

Let G,H𝒢G,H\in\mathcal{G} with GHG\subseteq H, (q,p)𝒞(q,p)\in\mathcal{CH}, 𝒌\boldsymbol{k} and 𝒍\boldsymbol{l} be fields of characteristic pp, and ϕ:𝒌𝒍\phi\colon\boldsymbol{k}\to\boldsymbol{l} be a homomorphism. Denote by ι:GH\iota\colon G\to H the inclusion map. We define

𝔸q,p(ι,ϕ)={(ι,ϕ)if q=p;p(ι,ϕ)if qp.\mathbb{A}_{q,p}(\iota,\phi)=\begin{cases}\mathbb{H}(\iota,\phi)&\text{if $q=p$;}\\ \mathbb{P}_{p}(\iota,\phi)&\text{if $q\neq p$.}\end{cases}

Let (K,v)(K,v) and (L,w)(L,w) be valued fields. We say that (L,w)(L,w) is a valued field extension of (K,v)(K,v) as a valued field if KLK\subseteq L and w|K=vw|_{K}=v.

Proposition 2.14.

Let 𝐤\boldsymbol{k} and 𝐥\boldsymbol{l} be perfect field with characteristic p>0p>0 and ϕ:𝐤𝐥\phi\colon\boldsymbol{k}\to\boldsymbol{l} be a homomorphism. The map 𝔸q,p(ι,ϕ):𝔸q,p(G,𝐤)𝔸q,p(H,𝐥)\mathbb{A}_{q,p}(\iota,\phi)\colon\mathbb{A}_{q,p}(G,\boldsymbol{k})\to\mathbb{A}_{q,p}(H,\boldsymbol{l}) is a homomorphism such that ζ𝔸q,p(H,𝐥)𝔸q,p(ι,ϕ)=ϕζ𝔸q,p(G,𝐤)\zeta_{\mathbb{A}_{q,p}(H,\boldsymbol{l})}\circ\mathbb{A}_{q,p}(\iota,\phi)=\phi\circ\zeta_{\mathbb{A}_{q,p}(G,\boldsymbol{k})} on 𝔄(𝔸q,p(G,𝐤),UG,𝐤,q,p)\mathfrak{A}(\mathbb{A}_{q,p}(G,\boldsymbol{k}),U_{G,\boldsymbol{k},q,p}). In addition, if qpq\neq p, then, the following are true:

  1. (1)

    (𝔸q,p(G,𝒌),UG,𝒌,q,p)(\mathbb{A}_{q,p}(G,\boldsymbol{k}),U_{G,\boldsymbol{k},q,p}) is a valued field extension of (Fr𝕎(𝒌),w𝒌)(\mathrm{Fr}\mathbb{W}(\boldsymbol{k}),w_{\boldsymbol{k}});

  2. (2)

    In particular, (𝔸q,p(G,𝒌),UG,𝒌,q,p)(\mathbb{A}_{q,p}(G,\boldsymbol{k}),U_{G,\boldsymbol{k},q,p}) is a valued field extension of (p,vp)(\mathbb{Q}_{p},v_{p}).

Proof.

Propositions 2.12 and 2.13 shows the former part of the statement. By the former part, since Fr𝕎(𝒌)=𝔸q,p(,𝒌)\mathrm{Fr}\mathbb{W}(\boldsymbol{k})=\mathbb{A}_{q,p}(\mathbb{Z},\boldsymbol{k}) and G\mathbb{Z}\subseteq G (see the definition of 𝒢\mathcal{G}), we can regard Fr𝕎(𝒌)\mathrm{Fr}\mathbb{W}(\boldsymbol{k}) as a subfield of 𝔸q,p(G,𝒌)\mathbb{A}_{q,p}(G,\boldsymbol{k}) Namely, (1) is true. Similarly, by G\mathbb{Z}\subseteq G, 𝔽p𝒌\mathbb{F}_{p}\subseteq\boldsymbol{k}, and since (p,vp)(\mathbb{Q}_{p},v_{p}) is equal to (𝔸q,p(,𝔽p),U,𝔽p,q,p)(\mathbb{A}_{q,p}(\mathbb{Z},\mathbb{F}_{p}),U_{\mathbb{Z},\mathbb{F}_{p},q,p}), the field 𝔸q,p(G,𝒌)\mathbb{A}_{q,p}(G,\boldsymbol{k}) is a valued field extension of (p,vp)(\mathbb{Q}_{p},v_{p}). This implies (2). ∎

Definition 2.3.

Let G𝒢G\in\mathcal{G}, (q,p)𝒞(q,p)\in\mathcal{CH}, and let 𝒌\boldsymbol{k} be a perfect field of characteristic pp. We make the following assumptions and definitions.

  1. (1)

    In the rest of the paper, whenever we take a complete system J𝔸q,p(G,𝒌)J\subseteq\mathbb{A}_{q,p}(G,\boldsymbol{k}) of representatives of the residue class field 𝒌\boldsymbol{k}, in the case of q=pq=p, we define J=𝒌J=\boldsymbol{k} using the fact that 𝒌(G,𝒌)\boldsymbol{k}\subseteq\mathbb{H}(G,\boldsymbol{k}). In the case of qpq\neq p, we take Jp(G,𝒌)J\subseteq\mathbb{P}_{p}(G,\boldsymbol{k}) such that J𝕎(𝒌)J\subseteq\mathbb{W}(\boldsymbol{k}) based on Proposition 2.14.

  2. (2)

    We define an element τ𝔸q,p(G,𝒌)\tau\in\mathbb{A}_{q,p}(G,\boldsymbol{k}) as follows. If q=pq=p, we define τ\tau by an indeterminate as in the definition of Hahn fields. If qpq\neq p, we define τ=pp(G,𝒌)\tau=p\in\mathbb{P}_{p}(G,\boldsymbol{k}). In this case, every element of 𝔸q,p(G,𝒌)\mathbb{A}_{q,p}(G,\boldsymbol{k}) can be represented as a power series of τ\tau with powers in GG.

  3. (3)

    For y𝔸q,p(G,𝒌)y\in\mathbb{A}_{q,p}(G,\boldsymbol{k}), and gGg\in G, if q=pq=p, then we define 𝑪(y,g)\boldsymbol{C}(y,g) the coefficient of τg\tau^{g} in the power series representation of yy. If qpq\neq p, then we fixed a complete system J𝕎(𝒌)J\subseteq\mathbb{W}(\boldsymbol{k}) of representatives of 𝒌\boldsymbol{k}, and we define 𝑪(y,g)J\boldsymbol{C}(y,g)\in J the coefficient of yy with respect to τg\tau^{g}. Of cause, in this case, the value 𝑪(y,g)\boldsymbol{C}(y,g) depends on a system JJ of representatives. Throughout this paper, we will not consider the situation where we change a system of representatives. Thus no confusion can arise even if JJ does not explicitly appear in the notation of 𝑪(y,g)\boldsymbol{C}(y,g). Notice that we can represent y=gG𝑪(y,g)τgy=\sum_{g\in G}\boldsymbol{C}(y,g)\tau^{g}.

Remark 2.2.

Related to Proposition 2.14, we make the next remarks.

  1. (1)

    𝔸q,p(ι,ϕ)\mathbb{A}_{q,p}(\iota,\phi) is injective since it is a homomorphism between fields.

  2. (2)

    The construction of 𝔸q,p(G,𝒌)\mathbb{A}_{q,p}(G,\boldsymbol{k}) is a functor.

  3. (3)

    In contrast to Proposition 2.7, the author does not know whether 𝔸q,p(ι,ϕ)\mathbb{A}_{q,p}(\iota,\phi) is a unique homeomorphism such that ζ𝔸q,p(ι,ϕ)=ϕζ\zeta\circ\mathbb{A}_{q,p}(\iota,\phi)=\phi\circ\zeta or not.

A group GG is said to be divisible if for every gGg\in G and for every n0n\in\mathbb{Z}_{\geq 0} there exists hGh\in G such that g=nhg=n\cdot h.

Proposition 2.15.

Let G𝒢G\in\mathcal{G} be divisible, (q,p)𝒞(q,p)\in\mathcal{CH}, 𝐤\boldsymbol{k} be an algebraically closed field with characteristic p>0p>0, and (K,v)(K,v) be a valued field such that v(K)G{}v(K)\subseteq G\sqcup\{\infty\} and 𝔎(K,v)𝐤\mathfrak{K}(K,v)\subseteq\boldsymbol{k}. Then there exists a homomorphic embedding ϕ:K𝔸q,p(G,𝐤)\phi\colon K\to\mathbb{A}_{q,p}(G,\boldsymbol{k}) such that v(x)=UG,𝐤,q,p(ϕ(x))v(x)=U_{G,\boldsymbol{k},q,p}(\phi(x)) fot all xKx\in K. Namely, the field (K,v)(K,v) can be regarded as a valued subfield of (𝔸q,p(G,𝐤),UG,𝐤,q,p)(\mathbb{A}_{q,p}(G,\boldsymbol{k}),U_{G,\boldsymbol{k},q,p}).

Proof.

See [20, Corollary 5]. ∎

2.2.3. Levi–Civita fields

We next discuss Levi–Civita fields and pp-adic Levi–Civita fields, which are will be used in Section 5.

Let G𝒢G\in\mathcal{G}, and 𝒌\boldsymbol{k} be a field. For f(G,𝒌)f\in\mathbb{H}(G,\boldsymbol{k}), in this subsection, we mainly consider the following condition.

  1. (Fin)

    For every nn\in\mathbb{Z}, the set supp(f)(,n]\text{{\rm supp}}(f)\cap(-\infty,n] is finite.

We denote by 𝕃[G,𝒌]\mathbb{L}[G,\boldsymbol{k}] the set of all f(G,𝒌)f\in\mathbb{H}(G,\boldsymbol{k}) satisfying the condition (Fin). For the next lemma, we refer the readers to [3, Theorem 3.18].

Lemma 2.16.

Let G𝒢G\in\mathcal{G}, 𝐤\boldsymbol{k} be a field. Then the set 𝕃[G,𝐤]\mathbb{L}[G,\boldsymbol{k}] is a subfield of (G,𝐤)\mathbb{H}(G,\boldsymbol{k}).

We call 𝕃[G,𝒌]\mathbb{L}[G,\boldsymbol{k}] the Levi–Civita field associated with GG and 𝐤\boldsymbol{k}

Fix (q,p)𝒞(q,p)\in\mathcal{CH} with qpq\neq p and assume that 𝒌\boldsymbol{k} has characteristic p>0p>0. Before defining a pp-adic analogue of Levi-Civita fields, we supplementally define a subset 𝔻[G,𝒌]\mathbb{D}[G,\boldsymbol{k}] of (G,𝕎(𝒌))\mathbb{H}(G,\mathbb{W}(\boldsymbol{k})) 𝕄p[G,𝒌]\mathbb{M}_{p}[G,\boldsymbol{k}] by the set of all members f𝕄p[G,𝒌]f\in\mathbb{M}_{p}[G,\boldsymbol{k}] satisfying the condition (Fin). We define 𝕄p[G,𝒌]\mathbb{M}_{p}[G,\boldsymbol{k}] by 𝕄p[G,𝒌]=Pr(𝔻[G,𝒌])\mathbb{M}_{p}[G,\boldsymbol{k}]=\mathrm{Pr}(\mathbb{D}[G,\boldsymbol{k}]), where Pr:(G,𝕎(𝒌))p(G,𝒌)\mathrm{Pr}\colon\mathbb{H}(G,\mathbb{W}(\boldsymbol{k}))\to\mathbb{P}_{p}(G,\boldsymbol{k}) is the canonical projection.

Lemma 2.17.

Let G𝒢G\in\mathcal{G}, pp be a prime, and 𝐤\boldsymbol{k} be a field of characteristic p>0p>0. Then the following statements are true:

  1. (1)

    For every complete system Jp(G,𝒌)J\subseteq\mathbb{P}_{p}(G,\boldsymbol{k}) of representatives of 𝒌\boldsymbol{k}, and for every a𝔻[G,𝒌]a\in\mathbb{D}[G,\boldsymbol{k}], the member f=StG,𝒌,J(G,𝕎(𝒌))(x)f=\mathrm{St}_{G,\boldsymbol{k},J}\in\mathbb{H}(G,\mathbb{W}(\boldsymbol{k}))(x) satisfies the condition (Fin) and f(g)Jf(g)\in J for all gGg\in G.

  2. (2)

    We have 𝔻[G,𝒌]\mathbb{D}[G,\boldsymbol{k}] is a subring of (G,𝕎(𝒌))\mathbb{H}(G,\mathbb{W}(\boldsymbol{k}));

  3. (3)

    If a𝔻[G,𝒌]a\in\mathbb{D}[G,\boldsymbol{k}] satisfies UG,𝒌,q,p(a)>0U_{G,\boldsymbol{k},q,p}(a)>0, then (1a)1𝔻[G,𝒌](1-a)^{-1}\in\mathbb{D}[G,\boldsymbol{k}].

  4. (4)

    For every a𝔻[G,𝒌]a\in\mathbb{D}[G,\boldsymbol{k}], there exists b𝔻[G,𝒌]b\in\mathbb{D}[G,\boldsymbol{k}] such that abab is equivalent to 11 modulo G,𝒌\mathbb{N}_{G,\boldsymbol{k}}.

Proof.

First we prove (1). By Lemma 2.9, we see that f(g)Jf(g)\in J for all gGg\in G. Put A=supp(a)A=\text{{\rm supp}}(a) and F=supp(f)F=\text{{\rm supp}}(f). Lemma 2.9 also shows that FA+0F\subseteq A+\mathbb{Z}_{\geq 0}. Due to this relation, since AA satisfies the condition (Fin), so does WW. Hence the statement (1) is true.

Next we prove (2). By the definitions of 𝔻[G,𝒌]\mathbb{D}[G,\boldsymbol{k}] and 𝕃[G,Fr𝕎(𝒌)]\mathbb{L}[G,\mathrm{Fr}\mathbb{W}(\boldsymbol{k})], we have 𝔻[G,𝒌]=(G,𝕎(𝒌))𝕃[G,Fr𝕎(𝒌)]\mathbb{D}[G,\boldsymbol{k}]=\mathbb{H}(G,\mathbb{W}(\boldsymbol{k}))\cap\mathbb{L}[G,\mathrm{Fr}\mathbb{W}(\boldsymbol{k})] (pay an attention to the difference between 𝕎(𝒌)\mathbb{W}(\boldsymbol{k}) and Fr𝕎(𝒌)\mathrm{Fr}\mathbb{W}(\boldsymbol{k}) appearing in (G,𝕎(𝒌))\mathbb{H}(G,\mathbb{W}(\boldsymbol{k})) and 𝕃[G,Fr𝕎(𝒌)]\mathbb{L}[G,\mathrm{Fr}\mathbb{W}(\boldsymbol{k})], respectively). Thus 𝔻[G,𝒌]\mathbb{D}[G,\boldsymbol{k}] is a subring of (G,𝕎(𝒌))\mathbb{H}(G,\mathbb{W}(\boldsymbol{k})).

Now we show (3). Put m=UG,𝒌,q,p(a)m=U_{G,\boldsymbol{k},q,p}(a). As in [20], we have (1a)1=1+a2+a3+(1-a)^{-1}=1+a^{2}+a^{3}+\cdots in (G,𝒌)\mathbb{H}(G,\boldsymbol{k}). For every i0i\in\mathbb{Z}_{\geq 0}, we have Ai=supp(StG,𝒌,J(ai))A_{i}=\text{{\rm supp}}(\mathrm{St}_{G,\boldsymbol{k},J}(a^{i})). In this case, we have imminAii\cdot m\leq\min A_{i} for all i1i\in\mathbb{Z}_{\geq 1}. Put B=i1AiB=\bigcup_{i\in\mathbb{Z}_{\geq 1}}A_{i}. Since AiA_{i} satisfies imminAii\cdot m\leq\min A_{i} and (,n]Ai(-\infty,n]\cap A_{i} is finite for all i,n0i,n\in\mathbb{Z}_{\geq 0}, we observe that BB satisfies that (,n]B(-\infty,n]\cap B is finite for all n0n\in\mathbb{Z}_{\geq 0}. Put E=supp(StG,𝒌,J(1a))E=\text{{\rm supp}}(\mathrm{St}_{G,\boldsymbol{k},J}(1-a)). Then EB+0E\subseteq B+\mathbb{Z}_{\geq 0}, and hence EE satisfies that (,n]E(-\infty,n]\cap E is finite for all n0n\in\mathbb{Z}_{\geq 0}. This implies that (1a)1𝔻[G,𝒌](1-a)^{-1}\in\mathbb{D}[G,\boldsymbol{k}].

We shall prove (4) Put f=StG,𝒌,J(a)f=\mathrm{St}_{G,\boldsymbol{k},J}(a). We only need to show that ff is invertible in 𝔻[G,𝒌]\mathbb{D}[G,\boldsymbol{k}]. We represent ff as f=gGfgtgf=\sum_{g\in G}f_{g}t^{g}. Take m=minsupp(f)m=\min\text{{\rm supp}}(f) and put y=m<gfgtgmy=\sum_{m<g}f_{g}t^{g-m}. Since w𝒌(fm)=0w_{\boldsymbol{k}}(f_{m})=0, we see that fmf_{m} is invertible in 𝕎(𝒌)\mathbb{W}(\boldsymbol{k}). Then f=fmtm(1(fm1y))f=f_{m}t^{m}(1-(-f_{m}^{-1}y)). UG,𝒌,q,p(fm1y)>0U_{G,\boldsymbol{k},q,p}(f_{m}^{-1}y)>0. Thus, by f=pmfm(1(fm1y))f=p^{m}f_{m}(1-(-f_{m}^{-1}y)), and by (3), we see that xx is invertible in 𝔻[G,𝒌]\mathbb{D}[G,\boldsymbol{k}]. ∎

As a consequence of Lemma 2.17, we obtain the next corollary.

Corollary 2.18.

Let G𝒢G\in\mathcal{G}, pp be a prime, and 𝐤\boldsymbol{k} be a field of characteristic p>0p>0. Then the following statements are true:

  1. (1)

    For every complete system Jp(G,𝒌)J\subseteq\mathbb{P}_{p}(G,\boldsymbol{k}) of representatives of 𝒌\boldsymbol{k}, the set 𝕄p[G,𝒌]\mathbb{M}_{p}[G,\boldsymbol{k}] is equal to the set Pr(StG,𝒌,J(𝔻[G,𝒌]))\mathrm{Pr}(\mathrm{St}_{G,\boldsymbol{k},J}(\mathbb{D}[G,\boldsymbol{k}])). Moreover, for every a𝕄p[G,𝒌]a\in\mathbb{M}_{p}[G,\boldsymbol{k}], there uniquely exists fStG,𝒌,J(𝔻[G,𝒌])f\in\mathrm{St}_{G,\boldsymbol{k},J}(\mathbb{D}[G,\boldsymbol{k}]) such that a=Pr(f)a=\mathrm{Pr}(f).

  2. (2)

    The set 𝕄p[G,𝒌]\mathbb{M}_{p}[G,\boldsymbol{k}] is a subfield of p(G,𝒌)\mathbb{P}_{p}(G,\boldsymbol{k}).

We call 𝕄p[G,𝒌]\mathbb{M}_{p}[G,\boldsymbol{k}] the pp-adic Levi–Civita field associated with GG and 𝐤\boldsymbol{k}. We simply represent the restriction UG,𝒌,q,p|𝔹q,p(G,𝒌)U_{G,\boldsymbol{k},q,p}|_{\mathbb{B}_{q,p}(G,\boldsymbol{k})} as the same symbol UG,𝒌,q,pU_{G,\boldsymbol{k},q,p}. In this setting, the field (𝔹q,p(G,𝒌),UG,𝒌,q,p)(\mathbb{B}_{q,p}(G,\boldsymbol{k}),U_{G,\boldsymbol{k},q,p}) becomes a valued subfield of (𝔸q,p(G,𝒌),UG,𝒌,q,p)(\mathbb{A}_{q,p}(G,\boldsymbol{k}),U_{G,\boldsymbol{k},q,p}).

To use pp-adic and ordinary Levi-Civita fields in a unified way, we make the next definition.

Definition 2.4.

Let G𝒢G\in\mathcal{G}, (q,p)𝒞(q,p)\in\mathcal{CH}, and 𝒌\boldsymbol{k} be a field of characteristic pp. We define 𝔹q,p(G,𝒌)\mathbb{B}_{q,p}(G,\boldsymbol{k}) by

𝔹q,p(G,𝒌)={𝕃[G,𝒌]if q=p;𝕄p[G,𝒌]if qp.\mathbb{B}_{q,p}(G,\boldsymbol{k})=\begin{cases}\mathbb{L}[G,\boldsymbol{k}]&\text{if $q=p$;}\\ \mathbb{M}_{p}[G,\boldsymbol{k}]&\text{if $q\neq p$.}\end{cases}
Proposition 2.19.

Let G𝒢G\in\mathcal{G}, (q,p)𝒞(q,p)\in\mathcal{CH}, and 𝐤\boldsymbol{k} be a field of characteristic pp. Then the set 𝔹q,p(G,𝐤)\mathbb{B}_{q,p}(G,\boldsymbol{k}) is a subfield of 𝔸q,p(G,𝐤)\mathbb{A}_{q,p}(G,\boldsymbol{k}).

Proof.

The case of q=pq=p is presented in Lemma 2.16. The case of qpq\neq p is proven in Corollary 2.18. ∎

3. Algebraic independence over valued fields

First we remark that, for valued fields (K,v)(K,v) and (L,w)(L,w) such that (L,w)(L,w) is a valued field extension of (K,v)(K,v), there exists a canonical injective embedding from 𝔎(K,v)\mathfrak{K}(K,v) into 𝔎(L,w)\mathfrak{K}(L,w). Namely, we can regard 𝔎(K,v)\mathfrak{K}(K,v) as a subset of 𝔎(L,w)\mathfrak{K}(L,w) since the inclusion map ι:KL\iota\colon K\to L satisfies ι(𝔄(K,v))𝔄(L,w))\iota(\mathfrak{A}(K,v))\subseteq\mathfrak{A}(L,w)) and ι(𝔬(K,v))𝔬(L,w)\iota(\mathfrak{o}(K,v))\subseteq\mathfrak{o}(L,w). Thus it naturally induce an homomorphic embedding 𝔎(K,v)𝔎(L,w)\mathfrak{K}(K,v)\to\mathfrak{K}(L,w).

Let KK and LL be fields with KLK\subseteq L. A member xx of LL is said to be transcendental over KK if xx is not a root of any non-trivial polynomial with coefficients in KK. A subset SS of LL is said to be algebraically independent if any finite collection x1,,xnx_{1},\dots,x_{n} in SS does not satisfy any non-trivial polynomial equation with coefficients in KK. Note that a singleton {x}\{x\} of LL is algebraically independent over KK if and only if xx is transcendental over KK.

Lemma 3.1.

Let (K0,v0)(K_{0},v_{0}) and (K1,v1)(K_{1},v_{1}) be valued fields. Assume that (K1,v1)(K_{1},v_{1}) is a valued field extension of (K0,v0)(K_{0},v_{0}). If x𝔄(K1,v1)x\in\mathfrak{A}(K_{1},v_{1}) such that ζ(x)𝔎(K1,v1)\zeta(x)\in\mathfrak{K}(K_{1},v_{1}) is transcendental over 𝔎(K0,v0)\mathfrak{K}(K_{0},v_{0}), then xx is transcendental over K0K_{0}.

Proof.

The lemma follows from [5, Theorem 3.4.2]. ∎

Lemma 3.2.

Let (K0,v0)(K_{0},v_{0}) and (K1,v1)(K_{1},v_{1}) be valued fields such that (K1,v1)(K_{1},v_{1}) is a valued field extension of (K0,v0)(K_{0},v_{0}). If x1,,xn,yK1x_{1},\dots,x_{n},y\in K_{1} satisfy that:

  1. (1)

    the set {x1,,xn}\{x_{1},\dots,x_{n}\} is algebraically independent over K0K_{0};

  2. (2)

    there exists a filed LL containing x1,,xnx_{1},\dots,x_{n} and satisfying K0LK1K_{0}\subseteq L\subseteq K_{1} for which there exist zLz\in L and cK0c\in K_{0} satisfying that c(yz)𝔄(K1,v1)c(y-z)\in\mathfrak{A}(K_{1},v_{1}) and ζ(c(yz))𝔎(K1,v1)\zeta(c(y-z))\in\mathfrak{K}(K_{1},v_{1}) is transcendental over 𝔎(L,v1|L)\mathfrak{K}(L,v_{1}|_{L}),

then the set {x1,,xn,y}\{x_{1},\dots,x_{n},y\} is algebraically independent over K0K_{0}.

Proof.

For the sake of contradiction, suppose that the set {x1,,xn,y}\{x_{1},\dots,x_{n},y\} is not algebraically independent over K0K_{0}. From (1) and the fact that LL contains x1,,xnx_{1},\dots,x_{n}, it follows that yy is algebraic over LL, and hence so is c(yz)c(y-z). Using Lemma 3.1 together with (2), we see that c(yz)c(y-z) is transcendental over LL. This is a contradiction. Therefore, the set {x1,,xn,y}\{x_{1},\dots,x_{n},y\} is algebraically independent over K0K_{0}. ∎

Lemma 3.3.

Let η(1,)\eta\in(1,\infty), G𝒢G\in\mathcal{G}, (q,p)𝒞(q,p)\in\mathcal{CH}, and let 𝐤\boldsymbol{k} be a perfect field with characteristic pp. Fix a cardinal θ\theta and a complete system J𝔸q,p(G,𝐤)J\subseteq\mathbb{A}_{q,p}(G,\boldsymbol{k}) of representatives of the residue class field 𝐤\boldsymbol{k}. If a set SS of non-zero elements of 𝔸q,p(G,𝐤)\mathbb{A}_{q,p}(G,\boldsymbol{k}) satisfies that:

  1. (N1)

    for every pair x,ySx,y\in S with xyx\neq y, and for every gGg\in G satisfying that g[v(xy),)g\in[v(x-y),\infty), if either of 𝑪(x,g)\boldsymbol{C}(x,g) and 𝑪(y,g)\boldsymbol{C}(y,g) is non-zero, then 𝑪(x,g)𝑪(y,g)\boldsymbol{C}(x,g)\neq\boldsymbol{C}(y,g),

then for every finite subset A={z1,,zn}A=\{z_{1},\dots,z_{n}\} of SS, there exist i{1,,n}i\in\{1,\dots,n\} and uGu\in G such that 𝐂(zi,u)0\boldsymbol{C}(z_{i},u)\neq 0 and 𝐂(zi,u)𝐂(zj,u)\boldsymbol{C}(z_{i},u)\neq\boldsymbol{C}(z_{j},u) for all j{1,,n}j\in\{1,\dots,n\}.

Proof.

Put u=min{UG,𝒌,q,p(zizj)ij}u=\min\{\,U_{G,\boldsymbol{k},q,p}(z_{i}-z_{j})\mid i\neq j\,\} and take a pair {i,n}\{i,n\} such that u=UG,𝒌,q,p(zizn)u=U_{G,\boldsymbol{k},q,p}(z_{i}-z_{n}). Then either 𝑪(zi,u)\boldsymbol{C}(z_{i},u) or 𝑪(zn,u)\boldsymbol{C}(z_{n},u) is non-zero. We my assume that 𝑪(zi,u)0\boldsymbol{C}(z_{i},u)\neq 0. Using the condition (N1) and the minimality of uu, we have 𝑪(yi,u)𝑪(yj,u)\boldsymbol{C}(y_{i},u)\neq\boldsymbol{C}(y_{j},u) for all j{1,,n}j\in\{1,\dots,n\}. ∎

Let G𝒢G\in\mathcal{G}, (q,p)𝒞(q,p)\in\mathcal{CH}, 𝒌\boldsymbol{k} be a field of characteristic pp, and KK be a subfield of 𝔸q,p(G,𝒌)\mathbb{A}_{q,p}(G,\boldsymbol{k}). Fix a complete system J𝔸q,p(G,𝒌)J\subseteq\mathbb{A}_{q,p}(G,\boldsymbol{k}) of representatives if qpq\neq p. We denote by 𝐀𝐇(K)\mathbf{AH}(K) the subfield of 𝒌\boldsymbol{k} generated by {ζ(𝑪(x,g))xK,gG}\{\,\zeta(\boldsymbol{C}(x,g))\mid x\in K,g\in G\,\} over 𝔎(K,v)\mathfrak{K}(K,v). The definition of 𝐀𝐇(K)\mathbf{AH}(K) is “ad-hoc”, which means that it depends on not only information of KK, but also the inclusion map K𝔸q,p(G,𝒌)K\to\mathbb{A}_{q,p}(G,\boldsymbol{k}). Namely, even if K,L𝔸q,p(G,𝒌)K,L\subseteq\mathbb{A}_{q,p}(G,\boldsymbol{k}) are isomorphic each other as fields, it can happen that 𝐀𝐇(K)𝐀𝐇(L)\mathbf{AH}(K)\neq\mathbf{AH}(L).

Proposition 3.4.

Let G𝒢G\in\mathcal{G}, (q,p)𝒞(q,p)\in\mathcal{CH}, and 𝐤,𝐥\boldsymbol{k},\boldsymbol{l} be perfect fields with characteristic pp such that 𝐤𝐥\boldsymbol{k}\subseteq\boldsymbol{l}, Take a subfield KK of 𝔸q,p(G,𝐥)\mathbb{A}_{q,p}(G,\boldsymbol{l}) such that 𝔎(K,v)𝐤\mathfrak{K}(K,v)\subseteq\boldsymbol{k}. Fix a system J𝔸q,p(G,𝐥)J\subseteq\mathbb{A}_{q,p}(G,\boldsymbol{l}) of representatives of 𝐥\boldsymbol{l}. If a set of SS of non-zero elements of 𝔸q,p(G,𝐥)\mathbb{A}_{q,p}(G,\boldsymbol{l}) satisfies the condition (N1) in Lemma 3.3 and the following:

  1. (T1)

    for every pair x,ySx,y\in S, and for every distinct pair g,gGg,g^{\prime}\in G, if 𝑪(x,g)0\boldsymbol{C}(x,g)\neq 0, then we have 𝑪(x,g)𝑪(y,g)\boldsymbol{C}(x,g)\neq\boldsymbol{C}(y,g^{\prime});

  2. (T2)

    the set {ζ(𝑪(x,g))xS,gG,𝑪(x,g)0}\{\,\zeta(\boldsymbol{C}(x,g))\mid x\in S,g\in G,\boldsymbol{C}(x,g)\neq 0\,\} is algebraically independent over 𝐀𝐇(K)\mathbf{AH}(K),

then the set SS is algebraically independent over KK.

Proof.

Let τ\tau be the same element of 𝔸q,p(G,𝒌)\mathbb{A}_{q,p}(G,\boldsymbol{k}) as in Definition 2.4. Take nn-many distinct members z1,,znz_{1},\dots,z_{n} in SS. Now we prove that {z1,,zn}\{z_{1},\dots,z_{n}\} is algebraically independent over KK by induction on nn.

In the case of n=1n=1, put z1=gG𝑪(z1,g)τgz_{1}=\sum_{g\in G}\boldsymbol{C}(z_{1},g)\tau^{g}. Take uGu\in G such that 𝑪(z1,u)0\boldsymbol{C}(z_{1},u)\neq 0. Put A={𝑪(z1,g)gG,gu}A=\{\,\boldsymbol{C}(z_{1},g)\mid g\in G,g\neq u\,\}. Then, due to the condition (T1), we have 𝑪(z1,u)A\boldsymbol{C}(z_{1},u)\not\in A. Note that the set ζ(A)\zeta(A) is algebraically independent over 𝒌\boldsymbol{k}. Let 𝒎\boldsymbol{m} be the perfect subfield of 𝒍\boldsymbol{l} generated by 𝐀𝐇(K)ζ(A)\mathbf{AH}(K)\cup\zeta(A), and put L=𝔸q,p(G,𝒎)L=\mathbb{A}_{q,p}(G,\boldsymbol{m}). Notice that 𝔸q,p(G,𝒎)\mathbb{A}_{q,p}(G,\boldsymbol{m}) is a subfield of 𝔸q,p(G,𝒍)\mathbb{A}_{q,p}(G,\boldsymbol{l}) (see Proposition 2.14). The fact that 𝐀𝐇(K)𝒎\mathbf{AH}(K)\subseteq\boldsymbol{m} implies that KLK\subseteq L. By assumption (T2) and 𝑪(z1,u)A\boldsymbol{C}(z_{1},u)\not\in A, we see that 𝑪(z1,u)\boldsymbol{C}(z_{1},u) is transcendental over 𝒎\boldsymbol{m}. Thus, by Lemma 3.1, we conclude that z1z_{1} is transcendental over LL. In particular, z1z_{1} is transcendental over KK.

Next, we fix k0k\in\mathbb{Z}_{\geq 0} and assume that the case of n=kn=k is true. We consider the case of n=k+1n=k+1. Since SS satisfies the condition (N1), we can take i{1,,n}i\in\{1,\dots,n\} and uGu\in G stated in Lemma 3.3. We may assume that i=k+1i=k+1. Put

A={𝑪(zi,g)gG,i=1,,k}{𝑪(zk+1,g)gG,gu}.A=\{\,\boldsymbol{C}(z_{i},g)\mid g\in G,i=1,\dots,k\,\}\cup\{\,\boldsymbol{C}(z_{k+1},g)\mid g\in G,g\neq u\,\}.

According to (T1) and the conclusion of Lemma 3.3, we see that 𝑪(zk+1,u)A\boldsymbol{C}(z_{k+1},u)\not\in A. Let 𝒎\boldsymbol{m} be a perfect subfield of 𝒍\boldsymbol{l} generated by 𝐀𝐇(K)ζ(A)\mathbf{AH}(K)\cup\zeta(A), and put L=𝔸q,p(G,𝒎)L=\mathbb{A}_{q,p}(G,\boldsymbol{m}). Similarly to the case of n=1n=1, we observe that KLK\subseteq L. Since A𝒎A\subseteq\boldsymbol{m}, we have ziLz_{i}\in L for all i{1,,k}i\in\{1,\dots,k\}. Define f=zk+1𝑪(zk+1,u)τuf=z_{k+1}-\boldsymbol{C}(z_{k+1},u)\tau^{u}. Due to the condition (T1), we have fLf\in L. By the definition, we also have τu(zk+1f)=𝑪(zk+1,u)\tau^{-u}(z_{k+1}-f)=\boldsymbol{C}(z_{k+1},u). Thus the condition (T2) shows that ζ(τu(zk+1f))\zeta(\tau^{-u}(z_{k+1}-f)) is transcendental over 𝒎\boldsymbol{m}. Hence Lemma 3.2 shows that the set {z1,,zk,zk+1}\{z_{1},\dots,z_{k},z_{k+1}\} is algebraically independent over KK. This finishes the proof. ∎

Remark 3.1.

Put w=𝑪(x,g)w=\boldsymbol{C}(x,g). The condition (T1) means that 𝑪(x,g)\boldsymbol{C}(x,g) is zero or gg is a unique member in GG such that 𝑪(y,g)=w\boldsymbol{C}(y,g)=w for some yXy\in X.

4. Isometric embeddings of ultrametric spaces

In this section, we prove our non-Archimedean analogue of the Arens–Eells theorem. As a consequence, we give an affirmative solution of Conjecture 1.1.

4.1. A non-Archimedean Arens–Eells theorem

4.1.1. Preparations

This subsection is devoted to proving the following technical theorem, which plays a central role of our first main theorem. Our proof of the next theorem can be considered as a sophisticated version of the proof of the main theorem of [21].

Theorem 4.1.

Let η(1,)\eta\in(1,\infty), (q,p)𝒞(q,p)\in\mathcal{CH}, G𝒢G\in\mathcal{G}, 𝐤\boldsymbol{k} be a field, and 𝐥\boldsymbol{l} be a perfect field of characteristic pp Fix a cardinal θ\theta and a complete system J𝔸q,p(G,𝐥)J\subseteq\mathbb{A}_{q,p}(G,\boldsymbol{l}) of representatives of the residue class field 𝐥\boldsymbol{l}. Let CC be a subset of JJ. Put R={0}{ηggG}R=\{0\}\sqcup\{\,\eta^{-g}\mid g\in G\,\}. If the following condition are satisfied:

  1. (A1)

    𝒍\boldsymbol{l} is a field extension of 𝒌\boldsymbol{k};

  2. (A2)

    the subset ζ(C)\zeta(C) of 𝒍\boldsymbol{l} is algebraically independent over 𝒌\boldsymbol{k};

  3. (A3)

    Card(C)=θ\operatorname{Card}(C)=\theta,

then for every RR-valued ultrametric space (X,d)(X,d) with Card(X)θ\operatorname{Card}(X)\leq\theta, there exists a map I:X𝔸q,p(G,𝐥)I\colon X\to\mathbb{A}_{q,p}(G,\boldsymbol{l}) such that:

  1. (B1)

    each I(x)I(x) is non-zero;

  2. (B2)

    the map II is an isometric embedding from (X,d)(X,d) into the ultrametric space (𝔸q,p(G,𝒌),UG,𝒌,q,p,η)(\mathbb{A}_{q,p}(G,\boldsymbol{k}),\lVert*\rVert_{U_{G,\boldsymbol{k},q,p},\eta});

  3. (B3)

    for every pair x,yXx,y\in X, and for every r(0,d(x,y)]r\in(0,d(x,y)], if either of 𝑪(I(x),logη(r))\boldsymbol{C}(I(x),-\log_{\eta}(r)) or 𝑪(I(y),logη(r))\boldsymbol{C}(I(y),-\log_{\eta}(r)) is non-zero, then we have 𝑪(I(x),logη(r))𝑪(I(y),logη(r))\boldsymbol{C}(I(x),-\log_{\eta}(r))\neq\boldsymbol{C}(I(y),-\log_{\eta}(r));

  4. (B4)

    for every pair x,yXx,y\in X, and for every distinct pair g,gGg,g^{\prime}\in G, if 𝑪(I(x),g)0\boldsymbol{C}(I(x),g)\neq 0, then 𝑪(I(x),g)𝑪(I(y),g)\boldsymbol{C}(I(x),g)\neq\boldsymbol{C}(I(y),g^{\prime}).

  5. (B5)

    for every α<θ\alpha<\theta, the set {𝑪(I(ξα),g)gG}\{\,\boldsymbol{C}(I(\xi_{\alpha}),g)\mid g\in G\,\} is contained in CC.

In this subsection, in what follows, we fix objects in the assumption of Theorem 4.1. We divide the proof of Theorem 4.1 into some lemmas.

First we prepare notations. Take ϖX\varpi\not\in X, and put E=X{ϖ}E=X\sqcup\{\varpi\}. Fix r0R{0}r_{0}\in R\setminus\{0\} and x0Xx_{0}\in X, and define an RR-valued ultrametric hh on EE by h|X×X=dh|_{X\times X}=d and h(x,ϖ)=d(x,x0)d(x0,ϖ)h(x,\varpi)=d(x,x_{0})\lor d(x_{0},\varpi). Then hh is actually an ultrametric (see, for example, [7, Lemma 5.1]). A one-point extension of a metric space is a traditional method to prove analogues of the Arens–Eells theorem.

Put C={bαα<θ}C=\{\,b_{\alpha}\mid\alpha<\theta\,\} and E={ξαα<θ}E=\{\,\xi_{\alpha}\mid\alpha<\theta\,\} with ξ0=ϖ\xi_{0}=\varpi. For every β<κ\beta<\kappa, we also put Eβ={ξαα<β}E_{\beta}=\{\,\xi_{\alpha}\mid\alpha<\beta\,\} and Cβ={bαα<β}C_{\beta}=\{\,b_{\alpha}\mid\alpha<\beta\,\}.

Fix λ<θ\lambda<\theta. We say that a map f:Eλ𝔸q,p(G,𝒌)f\colon E_{\lambda}\to\mathbb{A}_{q,p}(G,\boldsymbol{k}) is well-behaved if the following conditions are true:

  1. (C1)

    if λ=0\lambda=0, then H0H_{0} is the empty map and if λ>0\lambda>0, then Hλ(ϖ)=0H_{\lambda}(\varpi)=0, where 0 is the zero element of 𝔸q,p(G,𝒌)\mathbb{A}_{q,p}(G,\boldsymbol{k});

  2. (C2)

    the map ff is an isometric embedding from EλE_{\lambda} into 𝔸q,p(G,𝒌)\mathbb{A}_{q,p}(G,\boldsymbol{k});

  3. (C3)

    for every pair x,yEλx,y\in E_{\lambda}, and for every r(0,d(x,y)]r\in(0,d(x,y)], if either of 𝑪(f(x),logη(r))\boldsymbol{C}(f(x),-\log_{\eta}(r)) or 𝑪(f(y),logη(r))\boldsymbol{C}(f(y),-\log_{\eta}(r)) is non-zero, then we have 𝑪(f(x),logη(r))𝑪(f(y),logη(r))\boldsymbol{C}(f(x),-\log_{\eta}(r))\neq\boldsymbol{C}(f(y),-\log_{\eta}(r));

  4. (C4)

    for every pair x,yEλx,y\in E_{\lambda}, and for every distinct pair g,rGg,r^{\prime}\in G, if 𝑪(f(x),g)0\boldsymbol{C}(f(x),g)\neq 0, then 𝑪(f(x),g)𝑪(f(y),g)\boldsymbol{C}(f(x),g)\neq\boldsymbol{C}(f(y),g^{\prime});

  5. (C5)

    for every α<λ\alpha<\lambda, the set {𝑪(f(ξα),g)gG}\{\,\boldsymbol{C}(f(\xi_{\alpha}),g)\mid g\in G\,\} is contained in Cα+1C_{\alpha+1}.

For an ordinal λ<θ\lambda<\theta, a family {Hα:Eα𝔸q,p(G,𝒌)}α<λ\{H_{\alpha}\colon E_{\alpha}\to\mathbb{A}_{q,p}(G,\boldsymbol{k})\}_{\alpha<\lambda} is said to be coherent if the following condition is true:

  1. (Coh)

    for every β<θ\beta<\theta and for every α<β\alpha<\beta, we have Hβ|Eα=HαH_{\beta}|_{E_{\alpha}}=H_{\alpha}.

For the sake of simplicity, we define an ultrametric ee on 𝔸q,p(G,𝒌)\mathbb{A}_{q,p}(G,\boldsymbol{k}) by e(x,y)=xyUG,𝒌,q,p,ηe(x,y)=\lVert x-y\rVert_{U_{G,\boldsymbol{k},q,p},\eta}.

We shall construct a coherent family {Hα}α<θ\{H_{\alpha}\}_{\alpha<\theta} of well-behaved maps using transfinite recursion. We begin with the following convenient criterion.

Lemma 4.2.

Fix λ<θ\lambda<\theta with λ0\lambda\neq 0. Put u=d(E,ξλ)u=d(E,\xi_{\lambda}) and m=logη(u)m=-\log_{\eta}(u). Let Hλ:Eλ𝔸q,p(G,𝐤)H_{\lambda}\colon E_{\lambda}\to\mathbb{A}_{q,p}(G,\boldsymbol{k}) be a well-behaved map. Assume that an isometric embedding Hλ+1:Eλ+1𝔸q,p(G,𝐤)H_{\lambda+1}\colon E_{\lambda+1}\to\mathbb{A}_{q,p}(G,\boldsymbol{k}) satisfies Hλ+1|Eλ=HλH_{\lambda+1}|_{E_{\lambda}}=H_{\lambda} and the following property:

  1. (P1)

    For every gG(m,)g\in G\cap(m,\infty), we have 𝑪(Hλ+1(ξλ),g)=0\boldsymbol{C}(H_{\lambda+1}(\xi_{\lambda}),g)=0;

  2. (P2)

    If m<m<\infty and mGm\in G, then 𝑪(Hλ+1(ξλ),m){0,bλ}\boldsymbol{C}(H_{\lambda+1}(\xi_{\lambda}),m)\in\{0,b_{\lambda}\};

Then tha map HλH_{\lambda} satisfies the conditions (C3)(C5).

Proof.

First we note that the following claim is true:

  1. (CL)

    For every a(u,)a\in(u,\infty), there exists zEλz\in E_{\lambda} such that

    e(Hλ+1(z),Hλ(ξλ))<a.e(H_{\lambda+1}(z),H_{\lambda}(\xi_{\lambda}))<a.

    On the other words, for every g(,m)g\in(-\infty,m), there exists zEλz\in E_{\lambda} such that

    𝑪(Hλ+1(z),n)=𝑪(Hλ(ξλ),n)\boldsymbol{C}(H_{\lambda+1}(z),n)=\boldsymbol{C}(H_{\lambda}(\xi_{\lambda}),n)

    for all n<gn<g

Due to (C3) for HλH_{\lambda}, it suffices to consider the case of y=ξλy=\xi_{\lambda}. Take xEλ+1x\in E_{\lambda+1} and r(0,d(x,ξλ)]r\in(0,d(x,\xi_{\lambda})]. Assume that either of 𝑪(f(x),logη(r))\boldsymbol{C}(f(x),-\log_{\eta}(r)) or 𝑪(f(ξλ),logη(r))\boldsymbol{C}(f(\xi_{\lambda}),-\log_{\eta}(r)) is non-zero. In this case, we have rd(x,ξλ)ur\leq d(x,\xi_{\lambda})\leq u. Thus mlogη(r)m\leq-\log_{\eta}(r). If m<logη(r)m<-\log_{\eta}(r), then the property (P1) shows that 𝑪(Hλ+1(ξλ),logη(r))=0\boldsymbol{C}(H_{\lambda+1}(\xi_{\lambda}),-\log_{\eta}(r))=0. Thus the condition (C3) is valid. If m=logη(r)m=-\log_{\eta}(r), then the property (P2) implies that 𝑪(Hλ+1(ξλ),m)Cλ\boldsymbol{C}(H_{\lambda+1}(\xi_{\lambda}),m)\not\in C_{\lambda}. Thus, using (C5) for HλH_{\lambda}, the condition (C3) is satisfied. In any case, we conclude that the condition (C3) is true.

Owing to (C4) for HλH_{\lambda}, it is enough to confirm the case where x=ξλx=\xi_{\lambda} and yEλy\in E_{\lambda}, or xEλx\in E_{\lambda} and y=ξλy=\xi_{\lambda}. Take arbitrary distinct pair g,gGg,g^{\prime}\in G. First assume that x=ξλx=\xi_{\lambda}, i.e., 𝑪(Hλ+1(ξλ),g)0\boldsymbol{C}(H_{\lambda+1}(\xi_{\lambda}),g)\neq 0. By (P1), we have gmg\leq m. In the case of g=mg=m, using (P2) and (C5) for HλH_{\lambda}, we have 𝑪(f(x),g)𝑪(f(y),g)\boldsymbol{C}(f(x),g)\neq\boldsymbol{C}(f(y),g^{\prime}). In the case of g<mg<m. Then ηg(u,)\eta^{-g}\in(u,\infty). Thus the property (CL) enables us to take zEλz\in E_{\lambda} such that e(yi,ξλ)<ηge(y_{i},\xi_{\lambda})<\eta^{-g}. Thus we also have 𝑪(Hλ+1(z),a)=𝑪(Hλ+1(ξλ+1),a)\boldsymbol{C}(H_{\lambda+1}(z),a)=\boldsymbol{C}(H_{\lambda+1}(\xi_{\lambda+1}),a) for all aga\leq g. Applying (C4) for HλH_{\lambda} to yy and zz, we obtain that 𝑪(Hλ+1(z),g)𝑪(Hλ+1(y),g)\boldsymbol{C}(H_{\lambda+1}(z),g)\neq\boldsymbol{C}(H_{\lambda+1}(y),g^{\prime}). Hence, we have 𝑪(Hλ+1(ξλ+1),g)𝑪(Hλ+1(y),g)\boldsymbol{C}(H_{\lambda+1}(\xi_{\lambda+1}),g)\neq\boldsymbol{C}(H_{\lambda+1}(y),g^{\prime}). Second assume that y=ξλy=\xi_{\lambda}, i.e., 𝑪(Hλ+1(x),g)0\boldsymbol{C}(H_{\lambda+1}(x),g)\neq 0. By (P1), it is sufficient consider that gmg\leq m. By (P2), if g=mg^{\prime}=m, then we have 𝑪(Hλ+1(x),g)𝑪(Hλ+1(y),g)\boldsymbol{C}(H_{\lambda+1}(x),g)\neq\boldsymbol{C}(H_{\lambda+1}(y),g^{\prime}). If g<mg^{\prime}<m, then the property (CL) enables us to take zEλz\in E_{\lambda} such that e(yi,ξλ)<ηge(y_{i},\xi_{\lambda})<\eta^{-g}. The remaining proof is similar to that of the first case.

By (CL), if g<mg<m we have 𝑪(Hλ(ξλ),g)Cλ\boldsymbol{C}(H_{\lambda}(\xi_{\lambda}),g)\in C_{\lambda}. If m<gm<g, (P1) implies that 𝑪(Hλ(ξλ),g)=0\boldsymbol{C}(H_{\lambda}(\xi_{\lambda}),g)=0. Thus By (P2) we have the set {𝑪(f(ξα),g)gG}\{\,\boldsymbol{C}(f(\xi_{\alpha}),g)\mid g\in G\,\} is contained in Cλ{bλ}=Cλ+1C_{\lambda}\sqcup\{b_{\lambda}\}=C_{\lambda+1}. Namely, the condition (C5) is true. ∎

We next see the elementary lemma.

Lemma 4.3.

Let τ\tau be the same element of 𝔸q,p(G,𝐤)\mathbb{A}_{q,p}(G,\boldsymbol{k}) as in Definition 2.4. Then, for every a𝔸q,p(G,𝐤)a\in\mathbb{A}_{q,p}(G,\boldsymbol{k}) and for every r(0,)r\in(0,\infty), we can take γB(a,r;e)\gamma\in B(a,r;e) such that γ=gG(,logη(r))sgτg\gamma=\sum_{g\in G\cap(-\infty,-\log_{\eta}(r))}s_{g}\tau^{g}, where sgJs_{g}\in J. In this case, we have B(a,r;e)=B(γ,r;e)B(a,r;e)=B(\gamma,r;e).

Proof.

We put a=gG(,logη(r))sgτga=\sum_{g\in G\cap(-\infty,-\log_{\eta}(r))}s_{g}\tau^{g}, where sgJs_{g}\in J and we define γ=gG(,logη(r))sgτg\gamma=\sum_{g\in G\cap(-\infty,-\log_{\eta}(r))}s_{g}\tau^{g}. By the definition of ee, or UG,𝒌,q,pU_{G,\boldsymbol{k},q,p}, we have γB(a,r;e)\gamma\in B(a,r;e). Thus Lemma 2.2 implies that B(a,r;e)=B(γ,r;e)B(a,r;e)=B(\gamma,r;e). ∎

Next we show lemmas corresponding to steps of isolated ordinals in transfinite induction.

Lemma 4.4.

Fix λ<θ\lambda<\theta with λ0\lambda\neq 0, and let Hλ:Eλ𝔸q,p(G,𝐤)H_{\lambda}\colon E_{\lambda}\to\mathbb{A}_{q,p}(G,\boldsymbol{k}) be a well-behaved map. If h(Eλ,ξλ)>0h(E_{\lambda},\xi_{\lambda})>0, then we can obtain a well-behaved isometric embedding Hλ+1:Eλ+1𝔸q,p(G,𝐤)H_{\lambda+1}\colon E_{\lambda+1}\to\mathbb{A}_{q,p}(G,\boldsymbol{k}) such that Hλ+1|Eλ=HλH_{\lambda+1}|_{E_{\lambda}}=H_{\lambda}.

Proof.

Put Yλ=Hλ(Eλ)Y_{\lambda}=H_{\lambda}(E_{\lambda}). Put u=h(Eλ,ξλ)>0u=h(E_{\lambda},\xi_{\lambda})>0 and m=logη(u)m=-\log_{\eta}(u). Let τ\tau be the same element in 𝔸q,p(G,𝒌)\mathbb{A}_{q,p}(G,\boldsymbol{k}) as in Lemma 4.3.

Case 1. [There is no aEλa\in E_{\lambda} such that h(a,ξλ)=uh(a,\xi_{\lambda})=u]: Take a sequence {yi}i0\{y_{i}\}_{i\in\mathbb{Z}_{\geq 0}} in EλE_{\lambda} such that h(yi+1,ξλ)<h(yi,ξλ)h(y_{i+1},\xi_{\lambda})<h(y_{i},\xi_{\lambda}) for all i0i\in\mathbb{Z}_{\geq 0} and h(yi,ξλ)uh(y_{i},\xi_{\lambda})\to u as ii\to\infty. Put ri=h(yi,ξλ)r_{i}=h(y_{i},\xi_{\lambda}). Since ri+1<rir_{i+1}<r_{i} and B(H(yi+1),ri+1;e)B(H(yi),ri;e)B(H(y_{i+1}),r_{i+1};e)\subseteq B(H(y_{i}),r_{i};e) for all i0i\in\mathbb{Z}_{\geq 0} (see Lemma 2.2), using the spherical completeness of 𝔸q,p(G,𝒍)\mathbb{A}_{q,p}(G,\boldsymbol{l}) ((3) in Proposition 2.11), we obtain i0B(Hλ(yi),ri;e)\bigcap_{i\in\mathbb{Z}_{\geq 0}}B(H_{\lambda}(y_{i}),r_{i};e)\neq\emptyset. In this case, the set i0B(Hλ(yi),ri;e)\bigcap_{i\in\mathbb{Z}_{\geq 0}}B(H_{\lambda}(y_{i}),r_{i};e) is a closed ball of radius rr centered at some point in p(G,𝒍)\mathbb{P}_{p}(G,\boldsymbol{l}). Lemma 4.3 implies that there exists γ=gG(,m)srτg\gamma=\sum_{g\in G\cap(-\infty,m)}s_{r}\tau^{g} in 𝔸q,p(G,𝒌)\mathbb{A}_{q,p}(G,\boldsymbol{k}) such that B(γ,r;e)=i0B(Hλ(yi),ri;e)B(\gamma,r;e)=\bigcap_{i\in\mathbb{Z}_{\geq 0}}B(H_{\lambda}(y_{i}),r_{i};e). We define a map Hλ+1:Eλ+1𝔸q,p(G,𝒌)H_{\lambda+1}\colon E_{\lambda+1}\to\mathbb{A}_{q,p}(G,\boldsymbol{k}) by Hλ+1|Eλ=HλH_{\lambda+1}|_{E_{\lambda}}=H_{\lambda} and Hλ+1(ξλ)=γH_{\lambda+1}(\xi_{\lambda})=\gamma. This definition implies the condition (C1). Next we verify the condition (C2). It suffices to show that d(x,ξλ)=e(Hλ+1(x),Hλ+1(ξλ))d(x,\xi_{\lambda})=e(H_{\lambda+1}(x),H_{\lambda+1}(\xi_{\lambda})). Take n0n\in\mathbb{Z}_{\geq 0} such that d(yn,ξλ)<d(x,ξλ)d(y_{n},\xi_{\lambda})<d(x,\xi_{\lambda}). Then Lemma 2.1 implies h(x,ξλ)=h(yn,x)h(x,\xi_{\lambda})=h(y_{n},x), and hence h(yn,x)=e(Hλ+1(yn),Hλ+1(x))h(y_{n},x)=e(H_{\lambda+1}(y_{n}),H_{\lambda+1}(x)). By γB(Hλ+1(yn),rn;e)\gamma\in B(H_{\lambda+1}(y_{n}),r_{n};e), we have

e(Hλ+1(ξλ),Hλ+1(yn))h(ξλ,yn)<h(ξλ,x)=e(Hλ+1(yn),Hλ+1(x)).e(H_{\lambda+1}(\xi_{\lambda}),H_{\lambda+1}(y_{n}))\leq h(\xi_{\lambda},y_{n})<h(\xi_{\lambda},x)=e(H_{\lambda+1}(y_{n}),H_{\lambda+1}(x)).

Thus, using Lemma 2.1 again, we have

e(Hλ+1(yn),Hλ+1(x))=e(Hλ+1(ξλ),Hλ+1(x)),e(H_{\lambda+1}(y_{n}),H_{\lambda+1}(x))=e(H_{\lambda+1}(\xi_{\lambda}),H_{\lambda+1}(x)),

and hence e(Hλ+1(ξλ),Hλ+1(x))=h(ξλ,x)e(H_{\lambda+1}(\xi_{\lambda}),H_{\lambda+1}(x))=h(\xi_{\lambda},x). By the construction, the mao Hλ+1H_{\lambda+1} satisfies the properties (P1)(CL). Thus, we see that Hλ+1H_{\lambda+1} satisfies the conditions (C3)(C5).

Case 2. [There exists aXλa\in X_{\lambda} such that d(a,ξλ)=ud(a,\xi_{\lambda})=u]: By Lemma 4.3, there exists γ=gG(,m)sgτg𝔸q,p(G,𝒌)\gamma=\sum_{g\in G\cap(-\infty,m)}s_{g}\tau^{g}\in\mathbb{A}_{q,p}(G,\boldsymbol{k}) such that B(γ,u;e)=B(Hλ(a),u;e)B(\gamma,u;e)=B(H_{\lambda}(a),u;e). We put

δ=bλτm+γ=bλτm+gG(,m)sgτg.\delta=b_{\lambda}\cdot\tau^{m}+\gamma=b_{\lambda}\cdot\tau^{m}+\sum_{g\in G\cap(-\infty,m)}s_{g}\tau^{g}.

Then Lemma 2.2 states that B(δ,u;e)=B(Hλ(a),u;e)B(\delta,u;e)=B(H_{\lambda}(a),u;e). We define a map Hλ+1:Eλ+1𝔸q,p(G,𝒌)H_{\lambda+1}\colon E_{\lambda+1}\to\mathbb{A}_{q,p}(G,\boldsymbol{k}) by Hλ+1|Eλ=HλH_{\lambda+1}|_{E_{\lambda}}=H_{\lambda} and Hλ+1(ξλ)=δH_{\lambda+1}(\xi_{\lambda})=\delta. To verify the condition (C2), we show that e(Hλ+1(ξλ),Hλ+1(x))=e(ξλ,x)e(H_{\lambda+1}(\xi_{\lambda}),H_{\lambda+1}(x))=e(\xi_{\lambda},x) for all xXλx\in X_{\lambda}. If xB(a,u;h)x\not\in B(a,u;h), then we have h(x,a)=h(x,ξλ)h(x,a)=h(x,\xi_{\lambda}). Similarly, we have e(Hλ+1(x),Hλ+1(a))=e(Hλ+1(x),Hλ+1(ξλ))e(H_{\lambda+1}(x),H_{\lambda+1}(a))=e(H_{\lambda+1}(x),H_{\lambda+1}(\xi_{\lambda})). Since e(Hλ(x),Hλ(a))=e(Hλ(x),Hλ(a))=d(x,a)e(H_{\lambda}(x),H_{\lambda}(a))=e(H_{\lambda}(x),H_{\lambda}(a))=d(x,a), we have

e(Hλ+1(x),Hλ+1(ξλ))=h(x,ξλ).e(H_{\lambda+1}(x),H_{\lambda+1}(\xi_{\lambda}))=h(x,\xi_{\lambda}).

If xB(a,u;h)x\in B(a,u;h), then h(x,ξλ)h(x,a)h(a,ξλ)uh(x,\xi_{\lambda})\leq h(x,a)\lor h(a,\xi_{\lambda})\leq u. By the definition of uu, we have h(ξλ,x)=uh(\xi_{\lambda},x)=u. We also see that Hλ+1(x)B(Hλ+1(a),u;e)H_{\lambda+1}(x)\in B(H_{\lambda+1}(a),u;e). Then we can represent Hλ+1(x)=γ+gG[m,)cgτgH_{\lambda+1}(x)=\gamma+\sum_{g\in G\cap[m,\infty)}c_{g}\tau^{g}, where cgCλc_{g}\in C_{\lambda}. Take α<λ\alpha<\lambda with ξα=x\xi_{\alpha}=x. Then the condition (C2) for HλH_{\lambda} implies that cgCα+1c_{g}\in C_{\alpha+1} for all gG(,m]g\in G\cap(-\infty,m]. From the definition of UG,𝒌,q,pU_{G,\boldsymbol{k},q,p} and bλCα+1b_{\lambda}\not\in C_{\alpha+1}, it follows that e(Hλ+1(x),Hλ+1(ξλ))=ue(H_{\lambda+1}(x),H_{\lambda+1}(\xi_{\lambda}))=u. Hence

e(Hλ+1(x),Hλ+1(ξλ))=h(x,ξλ).e(H_{\lambda+1}(x),H_{\lambda+1}(\xi_{\lambda}))=h(x,\xi_{\lambda}).

Similarly to Case 1, by the construction, we see that Hλ+1H_{\lambda+1} satisfies the properties (P1)(CL). Thus, we see that Hλ+1H_{\lambda+1} satisfies the conditions (C3)(C5). ∎

Lemma 4.5.

Fix λ<θ\lambda<\theta with λ0\lambda\neq 0, and let Hλ:Eλ𝔸q,p(G,𝐤)H_{\lambda}\colon E_{\lambda}\to\mathbb{A}_{q,p}(G,\boldsymbol{k}) be a well-behaved map. If h(Eλ,ξλ)=0h(E_{\lambda},\xi_{\lambda})=0, then there exists a well-behaved map Hλ+1:Eλ+1𝔸q,p(G,𝐤)H_{\lambda+1}\colon E_{\lambda+1}\to\mathbb{A}_{q,p}(G,\boldsymbol{k}) such that Hλ+1|Xλ=HλH_{\lambda+1}|_{X_{\lambda}}=H_{\lambda}.

Proof.

We now define Hλ+1(ξλ)H_{\lambda+1}(\xi_{\lambda}) as follows. Take a sequence {xi}\{x_{i}\} in EλE_{\lambda} such that xiξλx_{i}\to\xi_{\lambda} as ii\to\infty, and define Hλ+1(ξλ)=limiHλ(xi)H_{\lambda+1}(\xi_{\lambda})=\lim_{i\to\infty}H_{\lambda}(x_{i}). The value Hλ+1(ξλ)H_{\lambda+1}(\xi_{\lambda}) is independent of the choice of a sequence {xi}i0\{x_{i}\}_{i\in\mathbb{Z}_{\geq 0}}.

First we confirm that Hλ+1H_{\lambda+1} is well-behaved. Since HλH_{\lambda} satisfies the condition (C1), so does Hλ+1H_{\lambda+1}.

To prove (C2), it is enough to show d(x,ξξλ)=e(Hλ+1(x),Hλ+1(ξλ))d(x,\xi_{\xi_{\lambda}})=e(H_{\lambda+1}(x),H_{\lambda+1}(\xi_{\lambda})) for all xEλ+1x\in E_{\lambda+1}. Take a sufficient large n0n\in\mathbb{Z}_{\geq 0} so that d(xn,ξλ)<d(x,ξλ)d(x_{n},\xi_{\lambda})<d(x,\xi_{\lambda}) and e(Hλ+1(xn),Hλ+1(ξλ))<e(Hλ+1(x),Hλ+1(ξλ))e(H_{\lambda+1}(x_{n}),H_{\lambda+1}(\xi_{\lambda}))<e(H_{\lambda+1}(x),H_{\lambda+1}(\xi_{\lambda})). Lemma 2.1 implies that

d(x,ξλ)=d(xn,x)d(x,\xi_{\lambda})=d(x_{n},x)

and

e(Hλ+1(x),Hλ+1(ξλ))=e(Hλ+1(xn),Hλ+1(x)).e(H_{\lambda+1}(x),H_{\lambda+1}(\xi_{\lambda}))=e(H_{\lambda+1}(x_{n}),H_{\lambda+1}(x)).

Since HλH_{\lambda} is an isometry and Hλ+1|Eλ=HλH_{\lambda+1}|_{E_{\lambda}}=H_{\lambda}, we have

d(xn,x)=e(Hλ+1(xn),Hλ+1(x)).d(x_{n},x)=e(H_{\lambda+1}(x_{n}),H_{\lambda+1}(x)).

Therefore we conclude that the condition (C2) is ture.

By the construction, we see that Hλ+1H_{\lambda+1} satisfies the properties (P1)(CL). Thus, we see that Hλ+1H_{\lambda+1} satisfies the conditions (C3)(C5). ∎

Let us prove Theorem 4.1.

Proof of Theorem 4.1.

The proof is based on [21]. Using transfinite recursion, we first construct a coherent family {Hμ:Eμ𝔸q,p(G,𝒌)}μ<θ\{H_{\mu}\colon E_{\mu}\to\mathbb{A}_{q,p}(G,\boldsymbol{k})\}_{\mu<\theta} of well-behaved maps. Fix μθ\mu\leq\theta and assume that we have already obtained a coherent family {Hα}α<μ\{H_{\alpha}\}_{\alpha<\mu} of well-behaved maps. Now we construct HμH_{\mu} as follows. If λ=0\lambda=0, then we define H0H_{0} as the empty map. If λ=1\lambda=1, then we have E1={b0}E_{1}=\{b_{0}\} and we define H1:E1𝔸q,p(G,𝒌)H_{1}\colon E_{1}\to\mathbb{A}_{q,p}(G,\boldsymbol{k}) by H1(ξ0)=0H_{1}(\xi_{0})=0. If μ=λ+1\mu=\lambda+1 for some λ<μ\lambda<\mu with λ0\lambda\neq 0, then using Lemmas 4.4 and 4.5, we obtain a well-behaved map Hλ+1:Eλ+1𝔸q,p(G,𝒌)H_{\lambda+1}\colon E_{\lambda+1}\to\mathbb{A}_{q,p}(G,\boldsymbol{k}) such that Hλ+1|Eλ=HλH_{\lambda+1}|_{E_{\lambda}}=H_{\lambda}. If μ\mu is a limit ordinal, then we define Hμ:Eμ𝔸q,p(G,𝒌)H_{\mu}\colon E_{\mu}\to\mathbb{A}_{q,p}(G,\boldsymbol{k}) by Hμ(x)=Hα(x)H_{\mu}(x)=H_{\alpha}(x), where xEαx\in E_{\alpha}. In this case, HμH_{\mu} is well-defined since the family {Hα}α<θ\{H_{\alpha}\}_{\alpha<\theta} is coherent. Therefore, according to transfinite recursion, we obtain a well-behaved isometric embedding Hθ:Eθ𝔸q,p(G,𝒌)H_{\theta}\colon E_{\theta}\to\mathbb{A}_{q,p}(G,\boldsymbol{k}). In this case, note that Eθ=EE_{\theta}=E. Now we define I:X𝔸q,p(G,𝒌)I\colon X\to\mathbb{A}_{q,p}(G,\boldsymbol{k}) by I=Hθ|XI=H_{\theta}|_{X}. Due to (C1), we have Hθ(ϖ)=0H_{\theta}(\varpi)=0 and ϖX\varpi\not\in X. Thus the condition (B1) is true. Since HθH_{\theta} satisfies (C2)(C5), the map II satisfies the conditions (B2)(B5). This finishes the proof of Theorem 4.1. ∎

4.1.2. The proof of the first main result

The following is our first main result.

Theorem 4.6.

Let η(1,)\eta\in(1,\infty), (q,p)𝒞(q,p)\in\mathcal{CH}, θ\theta be a cardinal, and G𝒢G\in\mathcal{G} be dividable. Put R={ηggG{}}R=\{\,\eta^{-g}\mid g\in G\sqcup\{\infty\}\,\}. Take an arbitrary valued field (K,v)(K,v) of characteristic qq such that v(K)G{}v(K)\subseteq G\sqcup\{\infty\} and 𝔎(K,v)\mathfrak{K}(K,v) has characteristic pp. Then there exists a valued field (L,w)(L,w) such that:

  1. (L1)

    the field (L,w)(L,w) is a valued field extension of (K,v)(K,v);

  2. (L2)

    w(L)=G{}w(L)=G\sqcup\{\infty\};

  3. (L3)

    for each RR-valued ultrametric (X,d)(X,d) with Card(X)θ\operatorname{Card}(X)\leq\theta, there exists an isometric embedding I:(X,d)(L,w,η)I\colon(X,d)\to(L,\lVert*\rVert_{w,\eta}) such that the set I(X)I(X) is algebraic independent over KK.

Moreover, for every RR-valued ultrametric space (X,d)(X,d), there exist a valued field (F,u)(F,u) and a map I:XFI\colon X\to F such that:

  1. (F1)

    the map I:(X,d)(F,u,η)I\colon(X,d)\to(F,\lVert*\rVert_{u,\eta}) is an isometric embedding;

  2. (F2)

    the field (L,u)(L,u) is a valued field extension of (K,v)(K,v);

  3. (F3)

    u(F)G{}u(F)\subseteq G\sqcup\{\infty\};

  4. (F4)

    the set I(X)I(X) is closed in FF;

  5. (F5)

    the set I(X)I(X) is algebraically independent over KK;

  6. (F6)

    if (X,d)(X,d) is complete, then (F,u)(F,u) can be chosen to be complete.

Proof.

Let 𝒌\boldsymbol{k} be the algebraic closure of 𝔎(K,v)\mathfrak{K}(K,v). Notice that 𝒌\boldsymbol{k} is perfect. Take a perfect filed 𝒍\boldsymbol{l} such that the transcendental degree of 𝒍\boldsymbol{l} over 𝒌\boldsymbol{k} is θ\theta, and take a transcendental basis BB of 𝒍\boldsymbol{l} over 𝒌\boldsymbol{k}. In this case, we have Card(B)=θ\operatorname{Card}(B)=\theta. Take a system J𝔸q,p(G,𝒍)J\subseteq\mathbb{A}_{q,p}(G,\boldsymbol{l}) of representatives of 𝒍\boldsymbol{l}. Notice that Bζ(J)B\subseteq\zeta(J). Put C=ζ1(B)={bαα<θ}C=\zeta^{-1}(B)=\{\,b_{\alpha}\mid\alpha<\theta\,\}. Applying Theorem 4.1 to (X,d)(X,d) and (K,v,η)(K,\lVert*\rVert_{v,\eta}), we can take an isometric embedding ψ:K𝔸q,p(G,𝒌)\psi\colon K\to\mathbb{A}_{q,p}(G,\boldsymbol{k}) satisfying the conditions (B1)(B5). Let ϕ:𝒌𝒍\phi\colon\boldsymbol{k}\to\boldsymbol{l} be the inclusion map. Using a homomorphic embedding 𝔸q,p(G,ϕ)ψ:K𝔸q,p(G,𝒍)\mathbb{A}_{q,p}(G,\phi)\circ\psi\colon K\to\mathbb{A}_{q,p}(G,\boldsymbol{l}), in what follows, we consider that KK is a subfield of 𝔸q,p(G,𝒍)\mathbb{A}_{q,p}(G,\boldsymbol{l}). In this case, we see that 𝐀𝐇(K)𝒌\mathbf{AH}(K)\subseteq\boldsymbol{k}. Put L=𝔸q,p(G,𝒍)L=\mathbb{A}_{q,p}(G,\boldsymbol{l}) and w=UG,𝒍,q,pw=U_{G,\boldsymbol{l},q,p}. Then (L,w)(L,w) satisfies the conditions (L1) and (L2) (see Proposition 2.11). Now we prove (L3). Take an RR-valued ultrametric space (X,d)(X,d). Applying Theorem 4.1 to CC, JJ, 𝒍\boldsymbol{l}, 𝒌\boldsymbol{k}, and (X,d)(X,d) of Card(X)θ\operatorname{Card}(X)\leq\theta, we can take I:XLI\colon X\to L satisfying the conditions (B1)(B5). The condition (B2) shows that II is isometric. Due to the conditions (B1), (B3), and (B4), and due to the fact that 𝐀𝐇(K)𝒌\mathbf{AH}(K)\subseteq\boldsymbol{k}, the set {I(x)xX}\{\,I(x)\mid x\in X\,\} satisfies the assumptions in Proposition 3.4. Then, according to Proposition 3.4, we see that I(X)I(X) is algebraic over KK. This means that the condition (L3).

We now prove the latter part. Let (Y,e)(Y,e) be the completion of (X,d)(X,d). Then (Y,e)(Y,e) is RR-valued (see [3, (12) in Theorem 1.6]). Take an isometric embedding I:YLI\colon Y\to L stated in the former part of Theorem 4.6. Since (Y,e)(Y,e) is complete and II is isometric, the set I(Y)I(Y) is closed in LL. Let FF be the subfield of LL generated by {I(x)xX}\{\,I(x)\mid x\in X\,\} over KK and put w=UG,𝒍,q,p|Kw=U_{G,\boldsymbol{l},q,p}|_{K}. Since I(Y)I(Y) is algebraically independent over KK, we have I(Y)L=I(X)I(Y)\cap L=I(X). Thus I(X)I(X) is closed in FF. The condition (F6) follows from the construction. This finishes the proof. ∎

Letting K=Fr𝕎(𝒌)K=\mathrm{Fr}\mathbb{W}(\boldsymbol{k}), and using Proposition 2.14, we obtain the next corollary.

Corollary 4.7.

Let η(1,)\eta\in(1,\infty), (q,p)𝒞(q,p)\in\mathcal{CH} with qpq\neq p, θ\theta be a cardinal, and G𝒢G\in\mathcal{G} be dividable. Put R={ηggG{}}R=\{\,\eta^{-g}\mid g\in G\sqcup\{\infty\}\,\}. Then there exists a valued field (L,v)(L,v) such that:

  1. (1)

    the field (L,v)(L,v) is a valued field extension of (Fr𝕎(𝒌),w𝒌)(\mathrm{Fr}\mathbb{W}(\boldsymbol{k}),w_{\boldsymbol{k}});

  2. (2)

    the absolute value v,η\lVert*\rVert_{v,\eta} is RR-valued;

  3. (3)

    for each RR-valued ultrametric (X,d)(X,d) with Card(X)θ\operatorname{Card}(X)\leq\theta, there exists an isometric embedding I:(X,d)(L,w,η)I\colon(X,d)\to(L,\lVert*\rVert_{w,\eta}) such that the set I(X)I(X) is algebraic independent over Fr𝕎(𝒌)\mathrm{Fr}\mathbb{W}(\boldsymbol{k}).

Moreover, for every RR-valued ultrametric space (X,d)(X,d), there exist a valued field (F,u)(F,u) and a map I:XFI\colon X\to F such that

  1. (1)

    the map I:(X,d)(F,u,η)I\colon(X,d)\to(F,\lVert*\rVert_{u,\eta}) is an isometric embedding;

  2. (2)

    the field (L,u)(L,u) is a valued field extension of (Fr𝕎(𝒌),w𝒌)(\mathrm{Fr}\mathbb{W}(\boldsymbol{k}),w_{\boldsymbol{k}});

  3. (3)

    u(F)G{}u(F)\subseteq G\sqcup\{\infty\};

  4. (4)

    I(X)I(X) is closed in FF;

  5. (5)

    I(X)I(X) is algebraically independent over Fr𝕎(𝒌)\mathrm{Fr}\mathbb{W}(\boldsymbol{k});

  6. (6)

    if (X,d)(X,d) is complete, then FF can be chosen to be complete.

Proof.

The proof is the same to that of Theorem 4.6. Remark that, since Fr𝕎(𝒌)\mathrm{Fr}\mathbb{W}(\boldsymbol{k}) is naturally a subset of 𝔸0,p(G,𝒌)\mathbb{A}_{0,p}(G,\boldsymbol{k}) (see Proposition 2.14), we do not need to use Proposition 2.15. Thus, the assumption of the corollary does not require that GG be divisible. ∎

4.2. Broughan’s conjecture

Next we give an affirmative solution of Conjecture 1.1. We begin with the definition of non-Archimedean Banach algebras. Let (K,||)(K,\lvert*\rvert) be a field equipped with a non-Archimedean absolute value, and (B,)(B,\lVert*\rVert) be a normed linear space over a field KK. We say that (B,)(B,\lVert*\rVert) is a non-Archimedean Banach algebra over KK (see [26]) if the following conditions are fullfiled:

  1. (1)

    BB is a ring (not necessarily commutative and unitary);

  2. (2)

    (B,)(B,\lVert*\rVert) is a normed linear space over (K,||)(K,\lvert*\rvert);

  3. (3)

    for every pair x,yBx,y\in B, we have x+yxy\lVert x+y\rVert\leq\lVert x\rVert\lor\lVert y\rVert and xyxy\lVert xy\lVert\leq\lVert x\rVert\cdot\lVert y\rVert;

  4. (4)

    BB is complete with respect to \lVert*\rVert;

  5. (5)

    if BB has a unit ee, then e=1\lVert e\rVert=1.

Some authors assume commutativity and the existence of a unit in the definition of a Banach algebra (see, for example, [24] and [17]). For instance, a valued field extension (L,||L)(L,\lvert*\rvert_{L}) of (K,||K)(K,\lvert*\rvert_{K}) is a (unitary) Banach algebra over (K,||K)(K,\lvert*\rvert_{K}).

As an application of Corollary 4.7, we obtain the next theorem.

Theorem 4.8.

Let pp be a prime, θ\theta be a cardinal, and 𝐤\boldsymbol{k} be a perfect field of characteristic pp such that the transcendental degree of 𝐤\boldsymbol{k} over 𝔽p\mathbb{F}_{p} is equal to θ\theta. Then every HpH_{p}-valued ultrametric (X,d)(X,d) with Card(X)θ\operatorname{Card}(X)\leq\theta can be isometrically embedded into Fr𝕎(𝐤)\mathrm{Fr}\mathbb{W}(\boldsymbol{k}). In particular, Conjecture 1.1 is true.

Proof.

The theorem follows from Corollary 4.7. Theorem 4.8 is actually an affirmative solution of Conjecture 1.1 since for every cardinal θ\theta, there exists a perfect filed 𝒌\boldsymbol{k} of characteristic pp such that the transcendental degree of 𝒌\boldsymbol{k} over 𝔽p\mathbb{F}_{p} is equal to θ\theta, and since Fr𝕎(𝒌)\mathrm{Fr}\mathbb{W}(\boldsymbol{k}) is a non-Archimedean (commutative and unitary) Banach algebra over p\mathbb{Q}_{p} (see Proposition 2.14). ∎

Remark 4.1.

An solution of Conjecture 1.1 is given as the ring 𝕎(𝒌)\mathbb{W}(\boldsymbol{k}) of Witt vectors.

5. Algebraic structures on Urysohn universal ultrametric spaces

In this section, we show that every Urysohn universal ultrametric space has a valued field structure that are extensions of a given prime valued field. Such an algebraic structure is realized as a pp-adic or ordinary Levi–Civita field.

For a class 𝒞\mathscr{C} of metric spaces, a metric space (X,d)(X,d) is said to be 𝒞\mathscr{C}-injective if for every pair (A,a)(A,a) and (B,b)(B,b) in 𝒞\mathscr{C} and for every pair of isometric embeddings ϕ:(A,a)(B,b)\phi\colon(A,a)\to(B,b) and ψ:(A,a)(X,d)\psi\colon(A,a)\to(X,d), there exists an isometric embedding θ:(B,b)(X,d)\theta\colon(B,b)\to(X,d) such that θϕ=ψ\theta\circ\phi=\psi. We denote by \mathscr{F} (resp. 𝒩(R)\mathscr{N}(R) for a range set RR) the class of all finite metric spaces (resp. all finite RR-valed ultrametric spaces). There exists a complete \mathscr{F}-injetive separable complete metric space (𝕌,ρ)(\mathbb{U},\rho) unique up to isometry, and this space is called the Urysohn universal metric space (see [25] and [15]).

Similarly, if RR is countable, there exists a separable complete RR-valued 𝒩(R)\mathscr{N}(R)-injective ultrametric space up to isometry, and it is called the RR-Urysohn universal ultrametric space, which is a non-Archimedean analogue of (𝕌,ρ)(\mathbb{U},\rho) (see [6] and [18]). Remark that if RR is uncountable, then every RR-injective ultrametric space is not separable (see [6]). Due to this phenomenon, mathematicians often consider only the case where RR is countable for separability.

In [11], the author discovered the concept of petaloid spaces, which can be considered to as RR-Urysohn universal ultrametric spaces in the case where RR is uncountable.

We begin with preparation of notations. A subset EE of [0,)[0,\infty) is said to be semi-sporadic if there exists a strictly decreasing sequence {ai}i0\{a_{i}\}_{i\in\mathbb{Z}_{\geq 0}} in (0,)(0,\infty) such that limiai=0\lim_{i\to\infty}a_{i}=0 and E={0}{aii0}E=\{0\}\cup\{\,a_{i}\mid i\in\mathbb{Z}_{\geq 0}\,\}. A subset of [0,)[0,\infty) is said to be tenuous if it is finite or semi-sporadic (see [10]). For a range set RR, we denote by 𝐓𝐄𝐍(R)\mathbf{TEN}(R) the set of all tenuous range subsets of RR. Let us recall the definition of petaloid spaces ([11]).

Definition 5.1.

Let RR be an uncountable range set. We say that a metric space (X,d)(X,d) is RR-petaloid if it is an RR-valued ultrametric space and there exists a family {Π(X,S)}S𝐓𝐄𝐍(R)\{\Pi(X,S)\}_{S\in\mathbf{TEN}(R)} of subspaces of XX satisfying the following properties:

  1. (P1)

    For every S𝐓𝐄𝐍(R)S\in\mathbf{TEN}(R), the subspace (Π(X,S),d)(\Pi(X,S),d) is isometric to the SS-Urysohn universal ultrametric space.

  2. (P2)

    We have S𝐓𝐄𝐍(R)Π(X,S)=X\bigcup_{S\in\mathbf{TEN}(R)}\Pi(X,S)=X.

  3. (P3)

    If S,T𝐓𝐄𝐍(R)S,T\in\mathbf{TEN}(R), then Π(X,S)Π(X,T)=Π(X,ST)\Pi(X,S)\cap\Pi(X,T)=\Pi(X,S\cap T).

  4. (P4)

    If S,T𝐓𝐄𝐍(R)S,T\in\mathbf{TEN}(R) and xΠ(X,T)x\in\Pi(X,T), then d(x,Π(X,S))d(x,\Pi(X,S)) belongs to (TS){0}(T\setminus S)\cup\{0\}.

We call the family {Π(X,S)}S𝐓𝐄𝐍(R)\{\Pi(X,S)\}_{S\in\mathbf{TEN}(R)} an RR-petal of XX, and call Π(X,S)\Pi(X,S) the SS-piece of the RR-petal {Π(X,S)}S𝐓𝐄𝐍(R)\{\Pi(X,S)\}_{S\in\mathbf{TEN}(R)}.

Notice that even if RR is countable, the RR-Urysohn universal ultrametric space has a petal structure satisfying the conditions (P1)(P4) (see [11]). This means that a petal space is a natural generalization of Urysohn universal ultrametric spaces.

The following is [9, Theorem 2.3] (see the property (P1) and [22, Propositions 20.2 and 21.1]).

Theorem 5.1.

Let RR be an uncountable range set. The following statements hold:

  1. (1)

    There exists an RR-petaloid ultrametric space and it is unique up to isometry.

  2. (2)

    The RR-petaloid ultrametric space is complete, non-separable, and (R)\mathscr{F}(R)-injective.

  3. (3)

    Every separable RR-valued ultrametric space can be isometrically embedded into the RR-petaloid ultrametric space.

Based on Theorem 5.1, for a range set RR, we denote by (𝕍R,σR)(\mathbb{V}_{R},\sigma_{R}) the RR-Urysohn universal ultrametric space if RR is countable; otherwise, the RR-petaloid ultrametric space. In what follows, by abuse of notation, we call (𝕍R,σR)(\mathbb{V}_{R},\sigma_{R}) the RR-Urysohn universal ultrametric space even if RR is uncountable.

We explain an example of a petaloid space. Let NN be a countable set. Let RR be a range set. We also denote by G(R,N)\mathrm{G}(R,N) the set of all function f:RNf\colon R\to N such that f(0)=0f(0)=0 and the set {0}{xRf(x)0}\{0\}\cup\{\,x\in R\mid f(x)\neq 0\,\} is tenuous. For f,gG(R,N)f,g\in\mathrm{G}(R,N), we define an RR-ultrametric \mathord{\vartriangle} on G(R,N)\mathrm{G}(R,N) by (f,g)=max{rRf(r)g(r)}\mathord{\vartriangle}(f,g)=\max\{\,r\in R\mid f(r)\neq g(r)\,\} if fgf\neq g; otherwise, (f,g)=0\mathord{\vartriangle}(f,g)=0. For more information of this construction, we refer the readers to [6] and [18].

Lemma 5.2.

For every countable set NN and every range set RR, the ultrametric space (G(R,N),)(\mathrm{G}(R,N),\mathord{\vartriangle}) is isometric to (𝕍R,σR)(\mathbb{V}_{R},\sigma_{R}).

Proof.

The lemma follows from [11, Theorem 1.3]. Notice that whereas in [11], it is only shown that (G(R,0),)(\mathrm{G}(R,\mathbb{Z}_{\geq 0}),\mathord{\vartriangle}) is isometric to (𝕍R,σR)(\mathbb{V}_{R},\sigma_{R}), the proof in [11] is still valid even if NN is a countable set in general. ∎

Lemma 5.3.

Let AA be a subset of \mathbb{R}, and η(0,)\eta\in(0,\infty). Put S={0}{ηggA}S=\{0\}\sqcup\{\,\eta^{-g}\mid g\in A\,\}. Then the following statements are equivalent:

  1. (1)

    for every n0n\in\mathbb{Z}_{\geq 0}, the set A(,n]A\cap(-\infty,n] is finite;

  2. (2)

    the set RR is tenuous.

Proof.

The lemma follows from [10, Lemma 2.12] and the fact that a map f:f\colon\mathbb{R}\to\mathbb{R} defined by f(x)=ηxf(x)=\eta^{-x} reverses the order on \mathbb{R}. ∎

In the next lemma, in order to treat the cases of q=pq=p and qpq\neq p, even if qpq\neq p, we define StG,𝒌,J\mathrm{St}_{G,\boldsymbol{k},J} by StG,𝒌,J(x)=x\mathrm{St}_{G,\boldsymbol{k},J}(x)=x.

Lemma 5.4.

Let η(1,)\eta\in(1,\infty), G𝒢G\in\mathcal{G}, (q,p)𝒞(q,p)\in\mathcal{CH}, 𝐤\boldsymbol{k} be a field of Card(𝐤)=0\operatorname{Card}(\boldsymbol{k})=\aleph_{0} and characteristic pp. Put R={0}{ηggG}R=\{0\}\cup\{\,\eta^{-g}\mid g\in G\,\}. Fix a complete system J𝔹q,p(G,𝐤)J\subseteq\mathbb{B}_{q,p}(G,\boldsymbol{k}) of representatives of the residue class field 𝐤\boldsymbol{k}. Take S𝐓𝐄𝐍(R)S\in\mathbf{TEN}(R). Define a subset Π(𝔹q,p(G,𝐤),S)\Pi(\mathbb{B}_{q,p}(G,\boldsymbol{k}),S) of 𝔹q,p(G,𝐤)\mathbb{B}_{q,p}(G,\boldsymbol{k}) by the set of all f𝔹q,p(G,𝐤)f\in\mathbb{B}_{q,p}(G,\boldsymbol{k}) such that {ηggsupp(StG,𝐤,J(f))}S\{\,\eta^{-g}\mid g\in\text{{\rm supp}}(\mathrm{St}_{G,\boldsymbol{k},J}(f))\,\}\subseteq S, and define an ultrametric dd on 𝔹q,p(G,𝐤)\mathbb{B}_{q,p}(G,\boldsymbol{k}) by d(x,y)=xyUG,𝐤,q,p,ηd(x,y)=\lVert x-y\rVert_{U_{G,\boldsymbol{k},q,p},\eta}. Then (Π(𝔹q,p(G,𝐤),S),d)(\Pi(\mathbb{B}_{q,p}(G,\boldsymbol{k}),S),d) is isometric to (𝕍S,σS)(\mathbb{V}_{S},\sigma_{S}).

Proof.

According to Card(𝒌)=0\operatorname{Card}(\boldsymbol{k})=\aleph_{0}, we see that Card(J)=0\operatorname{Card}(J)=\aleph_{0}. Then, using Lemma 2.9 or (1) in Corollary 2.18, for every xΠ(𝔹q,p(G,𝒌),S)x\in\Pi(\mathbb{B}_{q,p}(G,\boldsymbol{k}),S), there uniquely exists s(G,𝒌)s\in\mathbb{H}(G,\boldsymbol{k}) such that StG,𝒌,J(x)=s\mathrm{St}_{G,\boldsymbol{k},J}(x)=s and s(g)Js(g)\in J. Put s=gGsx(g)tgs=\sum_{g\in G}s_{x}(g)t^{g}, where sx(g)Js_{x}(g)\in J. We define a map T(x):RJT(x)\colon R\to J by T(x)(0)=0T(x)(0)=0 and T(x)(r)=sx(logη(r))T(x)(r)=s_{x}(-\log_{\eta}(r)) if r0r\neq 0. Then {rRsx(r)0}S\{\,r\in R\mid s_{x}(r)\neq 0\,\}\subseteq S. Thus T(x)G(S,J)T(x)\in\mathrm{G}(S,J) for all x𝔹q,p(G,𝒌)x\in\mathbb{B}_{q,p}(G,\boldsymbol{k}). By the definition of TT, we see that the map T:𝔹q,p(G,𝒌)G(S,J)T\colon\mathbb{B}_{q,p}(G,\boldsymbol{k})\to\mathrm{G}(S,J) defined by xT(x)x\mapsto T(x) is bijective. Using the definitions of UG,𝒌,q,pU_{G,\boldsymbol{k},q,p} and \mathord{\vartriangle}, the map T:𝔹q,p(G,𝒌)G(S,J)T\colon\mathbb{B}_{q,p}(G,\boldsymbol{k})\to\mathrm{G}(S,J) becomes an isometric bijection. Thus (G(S,J),)(\mathrm{G}(S,J),\mathord{\vartriangle}) is isometric to (Π(𝔹q,p(G,𝒌),S),d)(\Pi(\mathbb{B}_{q,p}(G,\boldsymbol{k}),S),d). Due to Card(J)=0\operatorname{Card}(J)=\aleph_{0}, Lemma 5.2 finishes the proof. ∎

Theorem 5.5.

Let η(1,)\eta\in(1,\infty), G𝒢G\in\mathcal{G}, (q,p)𝒞(q,p)\in\mathcal{CH}, 𝐤\boldsymbol{k} be a field of Card(𝐤)=0\operatorname{Card}(\boldsymbol{k})=\aleph_{0} and characteristic pp. Define an ultrametric dd on 𝔹q,p(G,𝐤)\mathbb{B}_{q,p}(G,\boldsymbol{k}) by d(x,y)=xyUG,𝐤,q,p,ηd(x,y)=\lVert x-y\rVert_{U_{G,\boldsymbol{k},q,p},\eta}. Then the Levi-Civita field (𝔹q,p(G,𝐤),d)(\mathbb{B}_{q,p}(G,\boldsymbol{k}),d) is isometric to (𝕍R,σR)(\mathbb{V}_{R},\sigma_{R}).

Proof.

We define Π(𝔹q,p(G,𝒌),S)\Pi(\mathbb{B}_{q,p}(G,\boldsymbol{k}),S) as in Lemma 5.4. Let us prove that {Π(𝔹q,p(G,𝒌),S)}S𝐓𝐄𝐍(R)\{\Pi(\mathbb{B}_{q,p}(G,\boldsymbol{k}),S)\}_{S\in\mathbf{TEN}(R)} satisfies the conditions (P1)(P4).

Lemma 5.4 implies the conditions (P1). From the definition of {Π(𝔹q,p(G,𝒌),S)}S𝐓𝐄𝐍(R)\{\Pi(\mathbb{B}_{q,p}(G,\boldsymbol{k}),S)\}_{S\in\mathbf{TEN}(R)}, the conditions (P2) is true. Next we show the condition (P3). It is sufficient to verify

(5.1) 𝔹q,p(G,𝒌)S𝐓𝐄𝐍(R)Π(𝔹q,p(G,𝒌),S).\displaystyle\mathbb{B}_{q,p}(G,\boldsymbol{k})\subseteq\bigcup_{S\in\mathbf{TEN}(R)}\Pi(\mathbb{B}_{q,p}(G,\boldsymbol{k}),S).

Take f𝔹q,p(G,𝒌)f\in\mathbb{B}_{q,p}(G,\boldsymbol{k}) and put A=supp(StG,𝒌,J(f))A=\text{{\rm supp}}(\mathrm{St}_{G,\boldsymbol{k},J}(f)). We also define S={ηggA}S=\{\,\eta^{-g}\mid g\in A\,\}. Then S𝐓𝐄𝐍(R)S\in\mathbf{TEN}(R) by Lemma 5.3. Hence fΠ(𝔹q,p(G,𝒌),S)f\in\Pi(\mathbb{B}_{q,p}(G,\boldsymbol{k}),S). Thus, the inclusion (5.1) is true.

Now we show (P4). We may assume that xΠ(𝔹q,p(G,𝒌),S)x\not\in\Pi(\mathbb{B}_{q,p}(G,\boldsymbol{k}),S). Put U=supp(StG,𝒌,J(x))U=\text{{\rm supp}}(\mathrm{St}_{G,\boldsymbol{k},J}(x)). Since xΠ(𝔹q,p(G,𝒌),S)x\not\in\Pi(\mathbb{B}_{q,p}(G,\boldsymbol{k}),S), we have USU\setminus S\neq\emptyset. Put u=max(US)u=\max(U\setminus S) and m=logη(u)m=-\log_{\eta}(u). Notice that 𝑪(x,m)0\boldsymbol{C}(x,m)\neq 0. Define a point y𝔹q,p(G,𝒌)y\in\mathbb{B}_{q,p}(G,\boldsymbol{k}) by y=gG(,m)x(g)τgy=\sum_{g\in G\cap(-\infty,m)}x(g)\tau^{g}. Then yΠ(𝔹q,p(G,𝒌),S)y\in\Pi(\mathbb{B}_{q,p}(G,\boldsymbol{k}),S) and d(x,y)=ud(x,y)=u. Then d(x,Π(𝔹q,p(G,𝒌),S))ud(x,\Pi(\mathbb{B}_{q,p}(G,\boldsymbol{k}),S))\leq u. For the sake of contradiction, suppose that d(x,Π(𝔹q,p(G,𝒌),S))<ud(x,\Pi(\mathbb{B}_{q,p}(G,\boldsymbol{k}),S))<u. In this setting, we can take zΠ(𝔹q,p(G,𝒌),S)z\in\Pi(\mathbb{B}_{q,p}(G,\boldsymbol{k}),S) such that m<v(xz)m<v(x-z). Then 𝑪(z,m)=𝑪(x,m)\boldsymbol{C}(z,m)=\boldsymbol{C}(x,m). In particular, 𝑪(z,m)0\boldsymbol{C}(z,m)\neq 0. Namely, we have supp(StG,𝒌,J(z))S\text{{\rm supp}}(\mathrm{St}_{G,\boldsymbol{k},J}(z))\not\subseteq S. This is a contradiction. Thus we have d(x,Π(𝔹q,p(G,𝒌),S))=uTSd(x,\Pi(\mathbb{B}_{q,p}(G,\boldsymbol{k}),S))=u\in T\setminus S, which means that the condition (P4) holds.

Therefore, we conclude that the Levi-Civita field 𝔹q,p(G,𝒌)\mathbb{B}_{q,p}(G,\boldsymbol{k}) is isometric to (𝕍R,σR)(\mathbb{V}_{R},\sigma_{R}). ∎

We find another application of the theory of Urysohn universal ultrametric spaces to that of valued field. For a class 𝒞\mathscr{C} of metric spaces, we say that a metric space (X,d)(X,d) is 𝒞\mathscr{C}-universal or universal for 𝒞\mathscr{C} if for every (A,e)𝒞(A,e)\in\mathscr{C} there exists an isometric embedding f:AXf\colon A\to X. An ultrametric space (X,d)(X,d) is said to be (R,0)(R,\aleph_{0})-haloed if for every aXa\in X and for every rR{0}r\in R\setminus\{0\}, there exists a subset AA of B(a,r)\mathrm{B}(a,r) such that κCard(A)\kappa\leq\operatorname{Card}(A) and d(x,y)=rd(x,y)=r for all distinct x,yAx,y\in A (see [9]).

Theorem 5.6.

Let η(1,)\eta\in(1,\infty), G𝒢G\in\mathcal{G}, pp be a prime, and let (K,v)(K,v) be a valued field such that 𝔎(K,v)\mathfrak{K}(K,v) is an infinite set and has characteristic pp. Put R={ηggG{}}R=\{\,\eta^{-g}\mid g\in G\sqcup\{\infty\}\,\}. Then (K,v,η)(K,\lVert*\rVert_{v,\eta}) is universal for all separable RR-valued ultrametric spaces.

Proof.

According to 0Card(𝒌)\aleph_{0}\leq\operatorname{Card}(\boldsymbol{k}), we observe that (K,v,η)(K,\lVert*\rVert_{v,\eta}) is (R,ω0)(R,\omega_{0})-haloed, and hence it is 𝒩(R,ω0)\mathscr{N}(R,\omega_{0})-injective (see [9, Theorem 1.1]). In [9, Theorem 1.5], it is stated that every complete 𝒩(R,ω0)\mathscr{N}(R,\omega_{0})-injective ultrametric space contains an isometric copy of (𝕍R,σR)(\mathbb{V}_{R},\sigma_{R}). Thus (K,v,η)(K,\lVert*\rVert_{v,\eta}) contains a metric subspace isometric to (𝕍R,σR)(\mathbb{V}_{R},\sigma_{R}). Hence (K,v,η)(K,\lVert*\rVert_{v,\eta}) is universal for all separable RR-valued ultrametric spaces. ∎

For a prime pp, the field p\mathbb{C}_{p} of pp-adic complex numbers is defined as the completion of the algebraic closure of p\mathbb{Q}_{p}. The pp-adic valuation vpv_{p} can be extended on p\mathbb{C}_{p}. In this case, we have vp(p)=v_{p}(\mathbb{C}_{p})=\mathbb{Q} (see [22]).

Corollary 5.7.

Let η(0,)\eta\in(0,\infty), and pp be a prime. Put R={ηgg}R=\{\,\eta^{-g}\mid g\in\mathbb{Q}\,\}. The field (p,vp)(\mathbb{C}_{p},v_{p}) of pp-adic complex numbers is universal for all separable RR-valued ultrametric spaces.

6. Questions

As a sophisticated version of a non-Archimedean Arens–Eells theorem for linear spaces, we ask the next question.

Question 6.1.

Let R,TR,T be range sets, and SS be a range set such that S{0}S\setminus\{0\} is a multiplicative subgroup of 0\mathbb{R}_{\geq 0}. Take a non-Archimedean valued field (K,||K)(K,\rvert*\lvert_{K}) and assume that

  1. (1)

    RS={rsrR,sS}TR\cdot S=\{\,r\cdot s\mid r\in R,s\in S\,\}\subseteq T;

  2. (2)

    ||K\rvert*\lvert_{K} is SS-valued.

For every RR-valued ultrametric space (X,d)(X,d), do there exist a TT-valued ultra-normed linear space (V,L)(V,\rVert*\lVert_{L}) over (K,||K)(K,\rvert*\lvert_{K}), and an isometric map I:XFI\colon X\to F?

Similarly to Corollary 5.7, as a counterpart of Theorem 4.6, we make the following question.

Question 6.2.

Let pp be 0 or a prime, 𝒌\boldsymbol{k} be an infinite field of characteristic pp, and RR be a rang set such that R{0}R\setminus\{0\} is a multiplicative subgroup of >0\mathbb{R}_{>0}. Take an arbitrary complete valued field (K,||)(K,\rvert*\lvert_{*}) with the residue class field 𝒌\boldsymbol{k} such that {|x|KxK}=R\{\,\rvert x\lvert_{K}\,\mid x\in K\,\}=R and (K,v)(K,v) is a valued field extension of (p,vp)(\mathbb{Q}_{p},v_{p}). For every RR-valued ultrametric space (X,d)(X,d) with Card(X)Card(𝒌)\operatorname{Card}(X)\leq\operatorname{Card}(\boldsymbol{k}), does there exist an isometric embedding I:XKI\colon X\to K?

References

  • [1] R. F. Arens and J. Eells, Jr., On embedding uniform and topological spaces, Pacific J. Math. 6 (1956), 397–403. MR 81458
  • [2] K. A. Broughan, Invariants for Real-Generated Uniform Topological and Algebraic Categories, Lecture Notes in Mathematics, Vol. 491, Springer-Verlag, Berlin-New York, 1975. MR 0425916
  • [3] A. B. Comicheo and K. Shamseddine, Summary on non-Archimedean valued fields, Advances in ultrametric analysis (A. Escassut, C. Perez-Garcia, and K. Shamseddine, eds.), Contemp. Math., vol. 704, American Mathematical Society, Providence, RI, 2018, pp. 1–36, DOI:10.1090/conm/704/14158. MR 3782290
  • [4] by same author, On non-Archimedean valued fields: a survey of algebraic, topological and metric structures, analysis and applications, Advances in non-Archimedean analysis and applications—the p-adic methodology in STEAM-H, STEAM-H: Sci. Technol. Eng. Agric. Math. Health, Springer, Cham, [2021] ©2021, pp. 209–254, DOI:10.1007/978-3-030-81976-7_6. MR 4367483
  • [5] A. J. Engler and A. Prestel, Valued Fields, Springer Monographs in Mathematics, Springer-Verlag, Berlin, 2005. MR 2183496
  • [6] S. Gao and C. Shao, Polish ultrametric Urysohn spaces and their isometry groups, Topology Appl. 158 (2011), no. 3, 492–508, DOI:10.1016/j.topol.2010.12.003. MR 2754373
  • [7] Y. Ishiki, An embedding, an extension, and an interpolation of ultrametrics, pp-Adic Numbers Ultrametric Anal. Appl. 13 (2021), no. 2, 117–147, DOI:10.1134/S2070046621020023. MR 4265905
  • [8] by same author, Simultaneous extensions of metrics and ultrametrics of high power, Topology Appl. 336 (2022), no. 15, Paper No. 108624, 37 pages.
  • [9] by same author, Characterizations of urysohn universal ultrametric spaces, (2023), arXiv:2303.17471.
  • [10] by same author, Constructions of Urysohn universal ultrametric spaces, (2023), arXiv:2302.00305, to appear in p-Adic Numbers Ultrametric Anal. Appl.
  • [11] by same author, Uniqueness and homogeneity of non-separable Urysohn universal ultrametric spaces, (2023), arXiv:2302.00306.
  • [12] I. Kaplansky, Maximal fields with valuations, Duke Math. J. 9 (1942), 303–321. MR 6161
  • [13] M. Megrelishvili and M. Shlossberg, Free non-Archimedean topological groups, Comment. Math. Univ. Carolin. 54 (2013), no. 2, 273–312, DOI:10.1016/j.ijar.2012.08.009. MR 3067710
  • [14] by same author, Non-Archimedean transportation problems and Kantorovich ultra-norms, p-Adic Numbers Ultrametric Anal. Appl. 8 (2016), no. 2, 125–148, DOI:10.1134/S2070046616020035. MR 3503300
  • [15] J. Melleray, Some geometric and dynamical properties of the Urysohn space, Topology Appl. 155 (2008), no. 14, 1531–1560, DOI:10.1016/j.topol.2007.04.029. MR 2435148
  • [16] E. Michael, A short proof of the Arens-Eells embedding theorem, Proc. Amer. Math. Soc. 15 (1964), 415–416, DOI:10.2307/2034516. MR 162222
  • [17] Lawrence Narici, Edward Beckenstein, and George Bachman, Functional analysis and valuation theory, Pure and Applied Mathematics, vol. Vol. 5, Marcel Dekker, Inc., New York, 1971. MR 361697
  • [18] L. Nguyen Van Thé, Structural Ramsey theory of metric spaces and topological dynamics of isometry groups, Mem. Amer. Math. Soc. 206 (2010), no. 968, 155 pages, DOI:10.1090/S0065-9266-10-00586-7. MR 2667917
  • [19] H. Ochsenius and W. H. Schikhof, Banach spaces over fields with an infinite rank valuation, pp-adic functional analysis (Poznań, 1998), Lecture Notes in Pure and Appl. Math., vol. 207, Dekker, New York, 1999, pp. 233–293. MR 1703500
  • [20] B. Poonen, Maximally complete fields, Enseign. Math. (2) 39 (1993), no. 1-2, 87–106. MR 1225257
  • [21] W. H. Schikhof, Isometrical embeddings of ultrametric spaces into non-Archimedean valued fields, Nederl. Akad. Wetensch. Indag. Math. 46 (1984), no. 1, 51–53. MR 748978
  • [22] by same author, Ultrametric Calculus, Cambridge Studies in Advanced Mathematics, vol. 4, Cambridge University Press, Cambridge, 2006, Reprint of the 1984 original [MR0791759]. MR 2444734
  • [23] J.-P. Serre, Local fields, GTM, vol. 67., Springer-Verlag, New York-Berlin, 1979, Translated from the French by Marvin Jay Greenberg. MR 554237
  • [24] N. Shilkret, Non-Archimedean Banach algebras, Duke Math. J. 37 (1970), 315–322. MR 261361
  • [25] P. Urysohn, Sur un espace métrique universel, Bull. Sci. Math. 51 (1927), 43–64 and 74–90.
  • [26] A. C. M. van Rooij, Non-Archimedean functional analysis, Monographs and Textbooks in Pure and Applied Mathematics, vol. 51, Marcel Dekker, Inc., New York, 1978. MR 512894