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

Infinite-dimensional Ramsey theory for homogeneous structures with SDAP+

Natasha Dobrinen Department of Mathematics
University of Notre Dame
255 Hurley Bldg
Notre Dame, IN 46556 U.S.A.
[email protected]
Abstract.

We prove that for any homogeneous structure 𝐊\mathbf{K} in a language with finitely many relation symbols of arity at most two satisfying SDAP+ (or LSDAP+), there are spaces of subcopies of 𝐊\mathbf{K}, forming subspaces of the Baire space, in which all Borel sets are Ramsey. Structures satisfying SDAP+ include the rationals, the Rado graph and more generally, unrestricted structures, and generic kk-partite graphs, the latter three types with or without an additional dense linear order. As a corollary of the main theorem, we obtain an analogue of the Nash-Williams Theorem which recovers exact big Ramsey degrees for these structures, answering a question raised by Todorcevic at the 2019 Luminy Workshop on Set Theory. Moreover, for the rationals and similar homogeneous structures our methods produce topological Ramsey spaces, thus satisfying analogues of the Ellentuck theorem.

2020 Mathematics Subject Classification:
05D10, 05C55, 05C15, 05C05, 03C15, 03E75
This research was supported by National Science Foundation Grant DMS-1901753

1. Introduction

Ramsey theory was initiated by the following celebrated result.

Theorem 1.1 (Ramsey [13]).

Given a positive integer kk, suppose that [β„•]k[\mathbb{N}]^{k}, the collection of all kk-element subsets of the natural numbers, is partitioned into finitely many pieces. Then there is an infinite subset NβŠ†β„•N\subseteq\mathbb{N} such that [N]k[N]^{k} is contained in one piece of the partition.

Extensions of Ramsey’s Theorem to colorings of infinite subsets of β„•\mathbb{N} have been proved, subject to constraints necessitated by the Axiom of Choice. Considering the set of all infinite subsets of the natural numbers, denoted by [β„•]β„•[\mathbb{N}]^{\mathbb{N}}, as the Baire space with its metric topology, a set π’³βŠ†[β„•]β„•\mathcal{X}\subseteq[\mathbb{N}]^{\mathbb{N}} is called Ramsey if for each M∈[β„•]β„•M\in[\mathbb{N}]^{\mathbb{N}}, there is an N∈[M]β„•N\in[M]^{\mathbb{N}} such that either [N]β„•βŠ†π’³[N]^{\mathbb{N}}\subseteq\mathcal{X} or else [N]β„•βˆ©π’³=βˆ…[N]^{\mathbb{N}}\cap\mathcal{X}=\emptyset. From 1965 through 1974, a beautiful progression of results was obtained using topological properties to guarantee that certain subsets of the Baire space are Ramsey. Nash-Williams proved that clopen sets are Ramsey in [12]; Galvin and Prikry proved that Borel sets are Ramsey in [8]; and Silver extended this to analytic sets in [15]. This line of work culminated in the topological characterization of Ramsey sets found by Ellentuck in [7] in terms of a topology refining the metric topology on [β„•]β„•[\mathbb{N}]^{\mathbb{N}}, now referred to as the Ellentuck topology.

This paper is focused on developing analogues of the Galvin–Prikry and Ellentuck Theorems for topological spaces of subcopies of a given FraΓ―ssé structure. This line of inquiry was highlighted in Section 11 of [9], by Kechris, Pestov, and Todorcevic. Specifically, they asked for the development of infinite-dimensional Ramsey theory of the form πŠβ†’βˆ—(𝐊)β„“,t𝐊\mathbf{K}\rightarrow_{*}(\mathbf{K})^{\mathbf{K}}_{\ell,t}, where 𝐊\mathbf{K} is the FraΓ―ssé limit of some FraΓ―ssé class and β†’βˆ—\rightarrow_{*} means that the partitions of (𝐊𝐊){\mathbf{K}\choose\mathbf{K}} are required to be definable in some sense.

Identifying the universe of 𝐊\mathbf{K} with β„•\mathbb{N}, one can view the set of all subcopies of 𝐊\mathbf{K} as the subspace of the Baire space corresponding to the set of universes of subcopies of 𝐊\mathbf{K}. In [4], the author proved an infinite-dimensional Ramsey theorem for certain topological spaces of subcopies of the Rado graph. At the 2019 Luminy Workshop on Set Theory, Todorcevic asked the author whether the infinite-dimensional Ramsey theorem would directly recover the exact big Ramsey degrees of the Rado graph. The approach in [4] directly recovers exact big Ramsey degrees for vertex, edge, and non-edge colorings, but does not directly recover exact big Ramsey degrees for most graphs with three or more vertices. Thus, the first motivation for this paper was to develop infinite-dimensional Ramsey theory for the Rado graph which would directly recover known exact big Ramsey degrees from a Nash-Williams style corollary. This is done in Corollary 6.5.

The second motivation for this paper was to develop infinite-dimensional Ramsey theory for a large collection of Fraïssé structures for which exact big Ramsey degrees are already known, thus making progress on the question of Kechris, Pestov, and Todorcevic discussed above. The Main Theorem of this paper, Theorem 6.3, develops infinite-dimensional Ramsey theory for all Fraïssé structures with finitely many relations of arity at most two satisfying amalgamation properties called SDAP+ and LSDAP+, developed by Coulson, Dobrinen, and Patel in [2] and [3] to prove exact big Ramsey degrees with a simple characterization in terms of diagonal antichains in coding trees of 11-types (see Theorem 2.24). The class of homogeneous structures satisfying SDAP+ includes the rationals and the rationals with an equivalence relation with finitely many dense equivalence classes, as well as the Rado graph, generic nn-partite graphs, the generic tournament and digraph, more generally unrestricted structures with finitely many binary relations, as well as versions of these with an additional linear order forming a dense linear order on the Fraïssé limit. The class of LSDAP+ structures includes the rationals with a convex equivalence relation and a natural hiearchy of such structures with successively coarser convex equivalence relations. (See Section 5 of [3] for a catalogue of SDAP+ and LSDAP+ structures and their big Ramsey degree results.)

The infinite-dimensional Ramsey theorem in this paper recovers the big Ramsey degrees proved in [3] for SDAP+ and LSDAP+ structures in the following manner: For each diagonal antichain AA representing a finite substructure 𝐀\mathbf{A} of 𝐊\mathbf{K}, Corollary 6.5 shows that given any finite coloring of the copies of 𝐀\mathbf{A} in 𝐊\mathbf{K}, there is a subcopy of 𝐊\mathbf{K} in which all copies of 𝐀\mathbf{A} represented by the similarity type of AA have the same color. Note that the lower bound argument in [3] showing that each diagonal antichain representing 𝐀\mathbf{A} persists in each subcopy of 𝐊\mathbf{K} does not follow from the infinite-dimensional Ramsey theory in this paper. Rather, given that result, we can conclude that Corollary 6.5 recovers exact big Ramsey degrees.

We remark on the necessity (in one form or another) of diagonal coding antichains and similarity types for infinite dimensional Ramsey theory. In the context of Countable Choice, one can well-order the vertices of a countably infinite structure 𝐊\mathbf{K} in order type Ο‰\omega; that is, we may assume that the universe of 𝐊\mathbf{K} is 𝐍\mathbf{N}. Since our language is countable, we can linearly order its symbols. Taken together, these automatically induce a coding tree of 11-types representing 𝐊\mathbf{K}. Big Ramsey degrees of FraΓ―ssé structures present constraints for the development of infinite-dimensional structural Ramsey theory. A FraΓ―ssé limit 𝐊\mathbf{K} of a FraΓ―ssé class 𝒦\mathcal{K} is said to have finite big Ramsey degrees if for each π€βˆˆπ’¦\mathbf{A}\in\mathcal{K}, there is some positive integer tt such that for each β„“β‰₯2\ell\geq 2,

(1) πŠβ†’(𝐊)β„“,t𝐀.\mathbf{K}\rightarrow(\mathbf{K})^{\mathbf{A}}_{\ell,t}.

This is the structural analogue of the infinite Ramsey Theorem 1.1. When such a tt exists for a given 𝐀\mathbf{A}, using the conventions in [9], T​(𝐀,𝒦)T(\mathbf{A},\mathcal{K}) denotes the minimal such tt and is called the big Ramsey degree of 𝐀\mathbf{A} in 𝐊\mathbf{K}. In all known cases, the big Ramsey degree T​(𝐀,𝐊)T(\mathbf{A},\mathbf{K}) corresponds to a canonical partition of (πŠπ€){\mathbf{K}\choose\mathbf{A}} (the copies of 𝐀\mathbf{A} in 𝐊\mathbf{K}) into T​(𝐀,𝐊)T(\mathbf{A},\mathbf{K}) many pieces each of which is persistent, meaning that for any member 𝐌\mathbf{M} of (𝐊𝐊){\mathbf{K}\choose\mathbf{K}}, the set (πŒπ€){\mathbf{M}\choose\mathbf{A}} meets every piece in the partition. Through the view of coding trees of 11-types, canonical partitions for SDAP+ and LSDAP+ structures are characterized by finite diagonal coding antichains (see Theorem 2.24 and preceding definitions). It is useful to think of finite big Ramsey degrees as a structural Ramsey theorem where one finds an expanded structure which guarantees one color for all copies of 𝐀\mathbf{A} in that expansion. Big Ramsey degrees of size two or more present a fundamental constraint to the development of infinite-dimensional structural Ramsey theory: any infinite-dimensional theorem must restrict to a subspace of (𝐊𝐊){\mathbf{K}\choose\mathbf{K}} where all members have the same similarity type in the coding tree of 11-types.

Given any FraΓ―ssé structure 𝐊\mathbf{K} satisfying SDAP+ or LSDAP+ with universe β„•\mathbb{N}, and given a subcopy 𝐌\mathbf{M} of 𝐊\mathbf{K}, let πŠβ€‹(𝐌)\mathbf{K}(\mathbf{M}) denote the collection of all 𝐍∈(𝐌𝐊)\mathbf{N}\in{\mathbf{M}\choose\mathbf{K}} with the same (induced) similarity type as 𝐌\mathbf{M} has as an enumerated substructure of 𝐊\mathbf{K}. (This space will be precisely defined in Section 6.) Note that πŠβ€‹(𝐌)\mathbf{K}(\mathbf{M}) is a topological space, identified with the subspace of the Baire space consisting of the universes of all structures in πŠβ€‹(𝐌)\mathbf{K}(\mathbf{M}). The following is the main theorem of the paper.

Theorem 6.3.

Let 𝐊\mathbf{K} be an enumerated FraΓ―ssé structure satisfying SDAP+ (or LSDAP+) with finitely many relations of arity at most two, and let 𝐃\mathbf{D} be a subcopy of 𝐊\mathbf{K} such that the subtree 𝔻\mathbb{D} of the coding tree of 11-types over 𝐊\mathbf{K} induced by the vertices in 𝐃\mathbf{D} is a good diagonal antichain. Then each Borel subset of πŠβ€‹(𝐃)\mathbf{K}(\mathbf{D}) is completely Ramsey, and hence Ramsey.

From the methods used in this paper, we immediately obtain the following stronger Ellentuck analogue for certain structures.

Theorem 6.4.

Let 𝐊\mathbf{K} be any one of the following structures with universe β„•\mathbb{N}: The rationals, β„šn\mathbb{Q}_{n}, β„šβ„š\mathbb{Q}_{\mathbb{Q}}, and or any FraΓ―ssé  structure satisfying SDAP+ (or LSDAP+) for which the coding tree of 11-types π•Œβ€‹(𝐊)\mathbb{U}(\mathbf{K}) has the property that on any given level of π•Œβ€‹(𝐊)\mathbb{U}(\mathbf{K}), only the coding node splits. Then the spaces π’Ÿβ€‹(𝔻)\mathcal{D}(\mathbb{D}), where 𝔻\mathbb{D} is a diagonal coding antichain for 𝐊\mathbf{K}, are actually topological Ramsey spaces.

The paper is orginized as follows: Background from [2] and [3] is presented in Section 2. Section 3 defines the spaces of diagonal coding antichains representing subcopies of a given homogeneous structure 𝐊\mathbf{K}. The pretext for our notation and set-up is Chapter 5 of Todorcevic’s book [16] on topological Ramsey spaces. Theorem 4.5 proves an Extended Pigeonhole Principle, a strong version of Todorcevic’s Axiom A.4. Theorem 5.17 proves a Galvin–Prikry style theorem for spaces of diagonal coding antichains with a metric topology. Theorem 6.3 in Section 6 interprets this back into natural subspaces of the Baire space, proving the main theorem of this paper. Corollary 6.5 answers a question of Todorcevic, showing that the Nash-Williams style corollary of our main theorem recovers big Ramsey degrees. Theorem 6.4 proves Ellentuck analogues for structures which have a certain amount of rigidity.

2. Background

This section reviews Fraïssé theory, amalgamation properties, and coding tree notions from [2].

2.1. Fraïssé theory and substructure amalgamation properties

In this paper, all languages β„’\mathcal{L} will consist of finitely many relation symbols {Ri:i<n}\{R_{i}:i<n\}, with the arity kik_{i} of RiR_{i} being either 11 or 22. An β„’\mathcal{L}-structure is an object 𝐀=⟨A,R0𝐀,…,Rnβˆ’1π€βŸ©\mathbf{A}=\langle\mathrm{A},R^{\mathbf{A}}_{0},\dots,R^{\mathbf{A}}_{n-1}\rangle, where the universe of 𝐀\mathbf{A}, denoted by A\mathrm{A}, is non-empty and Riπ€βŠ†AkiR^{\mathbf{A}}_{i}\subseteq\mathrm{A}^{k_{i}}. Finite structures will be denoted by 𝐀,𝐁,𝐂,…\mathbf{A},\mathbf{B},\mathbf{C},\dots, and their universes by A,B,C,…\mathrm{A},\mathrm{B},\mathrm{C},\dots. Infinite structures will typically be denoted by 𝐊,𝐌,𝐍,…\mathbf{K},\mathbf{M},\mathbf{N},\dots, and their universes by K,M,N,…\mathrm{K},\mathrm{M},\mathrm{N},\dots. The elements of the universe of a structure will be called vertices.

With no loss of generality, we make the following assumptions: (a) β„’\mathcal{L} has at least one unary relation symbol. (b) Any structure π€βˆˆπ’¦\mathbf{A}\in\mathcal{K} has the property that each vertex v∈Av\in\mathrm{A} satisfies R𝐀​(v)R^{\mathbf{A}}(v) for exactly one unary relation symbol RR in β„’\mathcal{L}. (c) For each unary relation symbol Rβˆˆβ„’R\in\mathcal{L}, there is some π€βˆˆπ’¦\mathbf{A}\in\mathcal{K} and a vertex v∈Av\in\mathrm{A} such that R𝐀​(v)R^{\mathbf{A}}(v). We say that the unary relations are non-trivial exactly when β„’\mathcal{L} has two or more unary relation symbols.

For β„’\mathcal{L}-structures 𝐀\mathbf{A} and 𝐁\mathbf{B}, an embedding e:𝐀→𝐁e:\mathbf{A}\rightarrow\mathbf{B} is an injection on their universes e:Aβ†’Be:\mathrm{A}\rightarrow\mathrm{B} with the property that for all i<ni<n,

Ri𝐀​(a1,…,ani)⟷Ri𝐁​(e​(a1),…,e​(ani))R_{i}^{\mathbf{A}}(a_{1},\dots,a_{n_{i}})\longleftrightarrow R_{i}^{\mathbf{B}}(e(a_{1}),\dots,e(a_{n_{i}}))

The ee-image of 𝐀\mathbf{A} is called a copy of 𝐀\mathbf{A} in 𝐁\mathbf{B}. If ee is the identity map, then 𝐀\mathbf{A} is a substructure of 𝐁\mathbf{B}. If ee is onto 𝐁\mathbf{B} then ee is an isomorphism and we say that 𝐀\mathbf{A} and 𝐁\mathbf{B} are isomorphic. We write 𝐀≀𝐁\mathbf{A}\leq\mathbf{B} exactly when there is an embedding of 𝐀\mathbf{A} into 𝐁\mathbf{B}, and 𝐀≅𝐁\mathbf{A}\cong\mathbf{B} exactly when there is an isomorphism from 𝐀\mathbf{A} onto 𝐁\mathbf{B}.

A class 𝒦\mathcal{K} of finite structures is called a FraΓ―ssé class if it is nonempty, closed under isomorphisms, hereditary, and satisfies the joint embedding and amalgamation properties. 𝒦\mathcal{K} is hereditary if whenever πβˆˆπ’¦\mathbf{B}\in\mathcal{K} and 𝐀≀𝐁\mathbf{A}\leq\mathbf{B}, then also π€βˆˆπ’¦\mathbf{A}\in\mathcal{K}. 𝒦\mathcal{K} satisfies the joint embedding property if for any 𝐀,πβˆˆπ’¦\mathbf{A},\mathbf{B}\in\mathcal{K}, there is a π‚βˆˆπ’¦\mathbf{C}\in\mathcal{K} such that 𝐀≀𝐂\mathbf{A}\leq\mathbf{C} and 𝐁≀𝐂\mathbf{B}\leq\mathbf{C}. 𝒦\mathcal{K} satisfies the amalgamation property if for any embeddings f:𝐀→𝐁f:\mathbf{A}\rightarrow\mathbf{B} and g:𝐀→𝐂g:\mathbf{A}\rightarrow\mathbf{C}, with 𝐀,𝐁,π‚βˆˆπ’¦\mathbf{A},\mathbf{B},\mathbf{C}\in\mathcal{K}, there is a πƒβˆˆπ’¦\mathbf{D}\in\mathcal{K} and there are embeddings r:𝐁→𝐃r:\mathbf{B}\rightarrow\mathbf{D} and s:𝐂→𝐃s:\mathbf{C}\rightarrow\mathbf{D} such that r∘f=s∘gr\circ f=s\circ g. Note that in a finite relational language, there are only countably many finite structures up to isomorphism.

A FraΓ―ssé class 𝒦\mathcal{K} satisfies the disjoint amalgamation property (DAP) if given 𝐀,𝐁,π‚βˆˆπ’¦\mathbf{A},\mathbf{B},\mathbf{C}\in\mathcal{K} and embeddings e:𝐀→𝐁e:\mathbf{A}\rightarrow\mathbf{B} and f:𝐀→𝐂f:\mathbf{A}\rightarrow\mathbf{C}, there is some πƒβˆˆπ’¦\mathbf{D}\in\mathcal{K} and embeddings eβ€²:𝐁→𝐃e^{\prime}:\mathbf{B}\rightarrow\mathbf{D} and fβ€²:𝐂→𝐃f^{\prime}:\mathbf{C}\rightarrow\mathbf{D} such that eβ€²βˆ˜e=fβ€²βˆ˜fe^{\prime}\circ e=f^{\prime}\circ f, and e′​[B]∩f′​[C]=eβ€²βˆ˜e​[A]=fβ€²βˆ˜f​[A]e^{\prime}[B]\cap f^{\prime}[C]=e^{\prime}\circ e[A]=f^{\prime}\circ f[A]. The DAP is often called the strong amalgamation property and is equivalent to the strong embedding property, which says that for any π€βˆˆπ’¦\mathbf{A}\in\mathcal{K}, v∈Av\in\mathrm{A}, and embedding Ο†:(π€βˆ’v)β†’πŠ\varphi:(\mathbf{A}-v)\rightarrow\mathbf{K}, there are infinitely many different extensions of Ο†\varphi to embeddings of 𝐀\mathbf{A} into 𝐊\mathbf{K}. (See [1].) We say that 𝒦\mathcal{K} satisfies the free amalgamation property (FAP) if it satisfies the DAP and moreover, 𝐃\mathbf{D} can be chosen so that 𝐃\mathbf{D} has no additional relations other than those inherited from 𝐁\mathbf{B} and 𝐂\mathbf{C}.

The following amalgamation properties, SFAP and SDAP, were first formulated in [2].

Definition 2.1 ([2]).

A FraΓ―ssé class 𝒦\mathcal{K} has the Substructure Free Amalgamation PropertyΒ (SFAP) if 𝒦\mathcal{K} has free amalgamation and given 𝐀,𝐁,𝐂,πƒβˆˆπ’¦\mathbf{A},\mathbf{B},\mathbf{C},\mathbf{D}\in\mathcal{K}, the following holds: Suppose

  1. (1)

    𝐀\mathbf{A} is a substructure of 𝐂\mathbf{C}, where 𝐂\mathbf{C} extends 𝐀\mathbf{A} by two vertices, say Cβˆ–A={v,w}\mathrm{C}\setminus\mathrm{A}=\{v,w\};

  2. (2)

    𝐀\mathbf{A} is a substructure of 𝐁\mathbf{B} and Οƒ\sigma and Ο„\tau are 11-types over 𝐁\mathbf{B} with σ↾𝐀=tp⁑(v/𝐀)\sigma\restriction\mathbf{A}=\operatorname{tp}(v/\mathbf{A}) and τ↾𝐀=tp⁑(w/𝐀)\tau\restriction\mathbf{A}=\operatorname{tp}(w/\mathbf{A}); and

  3. (3)

    𝐁\mathbf{B} is a substructure of 𝐃\mathbf{D} which extends 𝐁\mathbf{B} by one vertex, say vβ€²v^{\prime}, such that tp⁑(vβ€²/𝐁)=Οƒ\operatorname{tp}(v^{\prime}/\mathbf{B})=\sigma.

Then there is an π„βˆˆπ’¦\mathbf{E}\in\mathcal{K} extending 𝐃\mathbf{D} by one vertex, say wβ€²w^{\prime}, such that tp⁑(wβ€²/𝐁)=Ο„\operatorname{tp}(w^{\prime}/\mathbf{B})=\tau, 𝐄↾(Aβˆͺ{vβ€²,wβ€²})≅𝐂\mathbf{E}\restriction(\mathrm{A}\cup\{v^{\prime},w^{\prime}\})\cong\mathbf{C}, and 𝐄\mathbf{E} adds no other relations over D\mathrm{D}.

SFAP can be stated in terms of embeddings, but its formulation via substructures and 11-types is more closely aligned with its uses in the forcing proof in Section 4. In [2] it was remarked that SFAP is equivalent to free amalgamation along with a model-theoretic property that may be termed free 3-amalgamation, which is a special case of the disjoint 3-amalgamation property defined in [10]. Kruckman showed in [10] that if the age of a Fraïssé limit 𝐊\mathbf{K} has disjoint amalgamation and disjoint 3-amalgamation, then 𝐊\mathbf{K} exhibits a model-theoretic tameness property called simplicity.

The next amalgamation property extends SFAPΒ to disjoint amalgamation classes.

Definition 2.2 ([2]).

A FraΓ―ssé class 𝒦\mathcal{K} has the Substructure Disjoint Amalgamation Property (SDAP) if 𝒦\mathcal{K} has disjoint amalgamation, and the following holds: Given 𝐀,π‚βˆˆπ’¦\mathbf{A},\mathbf{C}\in\mathcal{K}, suppose that 𝐀\mathbf{A} is a substructure of 𝐂\mathbf{C}, where 𝐂\mathbf{C} extends 𝐀\mathbf{A} by two vertices, say vv and ww. Then there exist 𝐀′,π‚β€²βˆˆπ’¦\mathbf{A}^{\prime},\mathbf{C}^{\prime}\in\mathcal{K}, where 𝐀′\mathbf{A}^{\prime} contains a copy of 𝐀\mathbf{A} as a substructure and 𝐂′\mathbf{C}^{\prime} is a disjoint amalgamation of 𝐀′\mathbf{A}^{\prime} and 𝐂\mathbf{C} over 𝐀\mathbf{A}, such that letting vβ€²,wβ€²v^{\prime},w^{\prime} denote the two vertices in Cβ€²βˆ–Aβ€²\mathrm{C}^{\prime}\setminus\mathrm{A}^{\prime} and assuming (1) and (2), the conclusion holds:

  1. (1)

    Suppose πβˆˆπ’¦\mathbf{B}\in\mathcal{K} is any structure containing 𝐀′\mathbf{A}^{\prime} as a substructure, and let Οƒ\sigma and Ο„\tau be 11-types over 𝐁\mathbf{B} satisfying σ↾𝐀′=tp⁑(vβ€²/𝐀′)\sigma\restriction\mathbf{A}^{\prime}=\operatorname{tp}(v^{\prime}/\mathbf{A}^{\prime}) and τ↾𝐀′=tp⁑(wβ€²/𝐀′)\tau\restriction\mathbf{A}^{\prime}=\operatorname{tp}(w^{\prime}/\mathbf{A}^{\prime}).

  2. (2)

    Suppose πƒβˆˆπ’¦\mathbf{D}\in\mathcal{K} extends 𝐁\mathbf{B} by one vertex, say vβ€²β€²v^{\prime\prime}, such that tp⁑(vβ€²β€²/𝐁)=Οƒ\operatorname{tp}(v^{\prime\prime}/\mathbf{B})=\sigma.

Then there is an π„βˆˆπ’¦\mathbf{E}\in\mathcal{K} extending 𝐃\mathbf{D} by one vertex, say wβ€²β€²w^{\prime\prime}, such that tp⁑(wβ€²β€²/𝐁)=Ο„\operatorname{tp}(w^{\prime\prime}/\mathbf{B})=\tau and 𝐄↾(Aβˆͺ{vβ€²β€²,wβ€²β€²})≅𝐂\mathbf{E}\restriction(\mathrm{A}\cup\{v^{\prime\prime},w^{\prime\prime}\})\cong\mathbf{C}.

It is straightforward to see that SDAPΒ implies SFAPΒ (let 𝐀′=𝐀\mathbf{A}^{\prime}=\mathbf{A} and 𝐂′=𝐂\mathbf{C}^{\prime}=\mathbf{C}), and that SFAPΒ and SDAPΒ are each preserved under free superposition. These two amalgamation properties were formulated by extracting properties of FraΓ―ssé classes for which the forcing partial order in Theorem 4.5 can just be extension, from which simple characterizations of their big Ramsey degrees in [3] follow.

2.2. Coding trees of 11-types

This subsection reproduces notions from [2] which will be used throughout this paper. Given a FraΓ―ssé class 𝒦\mathcal{K}, an enumerated FraΓ―ssé structure is a FraΓ―ssé limit 𝐊\mathbf{K} of 𝒦\mathcal{K} with universe β„•={0,1,2,…}\mathbb{N}=\{0,1,2,\dots\}. For the sake of clarity, we use vnv_{n} (rather than nn) to denote the nn-th vertex of 𝐊\mathbf{K}. We let 𝐊n\mathbf{K}_{n} denote πŠβ†Ύ{vi:i<n}\mathbf{K}\restriction\{v_{i}:i<n\}, the restriction of 𝐊\mathbf{K} to its first nn vertices. All types will be quantifier-free 1-types, with variable xx, over some finite initial segment of 𝐊\mathbf{K}; the notation β€œtp” denotes a complete quantifer-free 1-type. For nβ‰₯1n\geq 1, a type over 𝐊n\mathbf{K}_{n} must contain the formula Β¬(x=vi)\neg(x=v_{i}) for each i<ni<n. Given a type ss over 𝐊n\mathbf{K}_{n}, for any i<ni<n, sβ†ΎπŠis\restriction\mathbf{K}_{i} denotes the restriction of ss to parameters from 𝐊i\mathbf{K}_{i}.

Definition 2.3 (The Coding Tree of 11-Types, [2]).

The coding tree of 11-types π•Šβ€‹(𝐊)\mathbb{S}(\mathbf{K}) for an enumerated FraΓ―ssé structure 𝐊\mathbf{K} is the set of all complete 11-types over initial segments of 𝐊\mathbf{K} along with a function c:β„•β†’π•Šβ€‹(𝐊)c:\mathbb{N}\rightarrow\mathbb{S}(\mathbf{K}) such that c​(n)c(n) is the 11-type of vnv_{n} over 𝐊n\mathbf{K}_{n}. The tree-ordering is inclusion.

We will often write π•Š\mathbb{S} and cnc_{n} in place of π•Šβ€‹(𝐊)\mathbb{S}(\mathbf{K}) and c​(n)c(n), respectively. Let π•Šβ€‹(n)\mathbb{S}(n) denote the collection of all 11-types tp⁑(vi/𝐊n)\operatorname{tp}(v_{i}/\mathbf{K}_{n}), where iβ‰₯ni\geq n, and note that cnc_{n} is a node in π•Šβ€‹(n)\mathbb{S}(n). The set π•Šβ€‹(0)\mathbb{S}(0) consists of the 11-types over the empty structure 𝐊0\mathbf{K}_{0}. A level set is a subset XβŠ†π•Šβ€‹(n)X\subseteq\mathbb{S}(n) for some nn. For sβˆˆπ•Šβ€‹(n)s\in\mathbb{S}(n), the immediate successors of ss are exactly those tβˆˆπ•Šβ€‹(n+1)t\in\mathbb{S}(n+1) such that sβŠ†ts\subseteq t. Each set π•Šβ€‹(n)\mathbb{S}(n) is finite, since the language β„’\mathcal{L} has finitely many finitary relation symbols.

A node sβˆˆπ•Šβ€‹(n)s\in\mathbb{S}(n) has length n+1n+1, denoted by |s||s|, and uniquely induces the sequence ⟨s​(i):i<|s|⟩\langle s(i):i<|s|\rangle defined as follows: s​(0)s(0) denotes the set of formulas in ss involving no parameters, and for 1≀i<|s|1\leq i<|s|, s​(i)s(i) denotes the set of those formulas in sβ†ΎπŠis\restriction\mathbf{K}_{i} in which viβˆ’1v_{i-1} appears as the parameter. For j<|s|j<|s|, note that ⋃i≀js​(i)\bigcup_{i\leq j}s(i) is the predecessor of ss in π•Šβ€‹(j)\mathbb{S}(j). For ℓ≀|s|\ell\leq|s|, we let sβ†Ύβ„“s\restriction\ell denote ⋃i<β„“s​(i)\bigcup_{i<\ell}s(i). Given s,tβˆˆπ•Šs,t\in\mathbb{S}, s∧ts\wedge t denotes the meet of ss and tt, which is sβ†ΎπŠms\restriction\mathbf{K}_{m} where mm is maximal such that sβ†ΎπŠm=tβ†ΎπŠms\restriction\mathbf{K}_{m}=t\restriction\mathbf{K}_{m}.

Let Ξ“\Gamma denote π•Šβ€‹(0)\mathbb{S}(0), the set of complete 11-types over the empty set that are realized in 𝐊\mathbf{K}. For Ξ³βˆˆΞ“\gamma\in\Gamma, we write β€œΞ³β€‹(vn)\gamma(v_{n}) holds in 𝐊\mathbf{K}” when Ξ³\gamma is the 1-type of vnv_{n} over the empty set. The following modification of Definition 2.3 of π•Šβ€‹(𝐊)\mathbb{S}(\mathbf{K}) is useful for FraΓ―ssé classes which have both non-trivial unary relations and a linear order.

Definition 2.4 (The Unary-Colored Coding Tree of 11-Types, [2]).

Let 𝒦\mathcal{K} be a FraΓ―ssé class in language β„’\mathcal{L} and 𝐊\mathbf{K} be an enumerated FraΓ―ssé structure for 𝒦\mathcal{K}. For nβˆˆβ„•n\in\mathbb{N}, let cnc_{n} denote the 11-type of vnv_{n} over 𝐊n\mathbf{K}_{n} (exactly as in the definition of π•Šβ€‹(𝐊)\mathbb{S}(\mathbf{K})). Let β„’βˆ’\mathcal{L}^{-} denote the collection of all relation symbols in β„’\mathcal{L} of arity greater than one, and let πŠβˆ’\mathbf{K}^{-} denote the reduct of 𝐊\mathbf{K} to β„’βˆ’\mathcal{L}^{-} and 𝐊nβˆ’\mathbf{K}_{n}^{-} the reduct of 𝐊n\mathbf{K}_{n} to β„’βˆ’\mathcal{L}^{-}.

The nn-th level, denoted π•Œβ€‹(n)\mathbb{U}(n), is the collection of all 11-types ss over 𝐊nβˆ’\mathbf{K}^{-}_{n} in the language β„’βˆ’\mathcal{L}^{-} such that for some iβ‰₯ni\geq n, viv_{i} satisfies ss. Define π•Œ=π•Œβ€‹(𝐊)\mathbb{U}=\mathbb{U}(\mathbf{K}) to be ⋃n<Ο‰π•Œβ€‹(n)\bigcup_{n<\omega}\mathbb{U}(n). The tree-ordering on π•Œ\mathbb{U} is simply inclusion. The unary-colored coding tree of 11-types is the tree π•Œ\mathbb{U} along with the function c:Ο‰β†’π•Œc:\omega\rightarrow\mathbb{U} such that c​(n)=cnc(n)=c_{n}. Thus, cnc_{n} is the 11-type (in the language β„’βˆ’\mathcal{L}^{-}) of vnv_{n} in π•Œβ€‹(n)\mathbb{U}(n) along with the additional β€œunary color” Ξ³βˆˆΞ“\gamma\in\Gamma such that γ​(vn)\gamma(v_{n}) holds in 𝐊\mathbf{K}.

Remark 2.5.

If β„’\mathcal{L} has no non-trivial unary relation symbols then π•Œβ€‹(𝐊)=π•Šβ€‹(𝐊)\mathbb{U}(\mathbf{K})=\mathbb{S}(\mathbf{K}). If 𝒦\mathcal{K} satisfies SFAP, it suffices to work in π•Šβ€‹(𝐊)\mathbb{S}(\mathbf{K}). The purpose of π•Œβ€‹(𝐊)\mathbb{U}(\mathbf{K}) is to handle cases such as β„šn\mathbb{Q}_{n}, the rationals with an equivalence relation with finitely many equivalence classes each of which is dense in the rationals.

2.3. Passing types and similarity

This subsection reproduces definitions from [2], with simplified versions given due to the fact that all relation symbols in this article have arity at most two. Throughout, fix 𝐊\mathbf{K} and let π•Š\mathbb{S} denote π•Šβ€‹(𝐊)\mathbb{S}(\mathbf{K}). All of the instances of π•Š\mathbb{S} in this subsection may be substituted with π•Œ:=π•Œβ€‹(𝐊)\mathbb{U}:=\mathbb{U}(\mathbf{K}).

Definition 2.6 (Passing Type, [2]).

Given s,tβˆˆπ•Šs,t\in\mathbb{S} with |s|<|t||s|<|t|, we call t​(|s|)t(|s|) the passing type of tt at ss. We also call t​(|s|)t(|s|) the passing type of tt at cnc_{n}, where nn is the integer such that |cn|=|s||c_{n}|=|s|.

Note that passing types are partial 11-types which contain only binary relation symbols.

Definition 2.7 (Similarity of Passing Types, [2]).

Let m,nβˆˆβ„•m,n\in\mathbb{N}, and let f:{vm,x}β†’{vn,x}f:\{v_{m},x\}\to\{v_{n},x\} be given by f​(vm)=vnf(v_{m})=v_{n} and f​(x)=xf(x)=x. Suppose s,tβˆˆπ•Šs,t\in\mathbb{S} are such that |cm|<|s||c_{m}|<|s| and |cn|<|t||c_{n}|<|t|. We write

(2) s​(cm)∼t​(cn)s(c_{m})\sim t(c_{n})

when, given any relation symbol Rβˆˆβ„’R\in\mathcal{L} of arity two and ordered pair (z0,z1)(z_{0},z_{1}) such that {z0,z1}={vm,x}\{z_{0},z_{1}\}=\{v_{m},x\}, it follows that R​(z0,z1)∈s​(cm)R(z_{0},z_{1})\in s(c_{m}) if and only if R​(f​(z0),f​(z1))∈t​(cn)R(f(z_{0}),f(z_{1}))\in t(c_{n}). When s​(cm)∼t​(cn)s(c_{m})\sim t(c_{n}) holds, we say that the passing type of ss at cmc_{m} is similar to the passing type of tt at cnc_{n}.

It is clear that ∼\sim is an equivalence relation.

Definition 2.8.

Let 𝐀\mathbf{A}, 𝐁\mathbf{B} be finite substructures of 𝐊\mathbf{K} with universes ⟨vji:i<n⟩\langle v_{j_{i}}:i<n\rangle, ⟨vki:i<n⟩\langle v_{k_{i}}:i<n\rangle, respectively. Let sβˆˆπ•Šβ€‹(β„“)s\in\mathbb{S}(\ell) and tβˆˆπ•Šβ€‹(β„“β€²)t\in\mathbb{S}(\ell^{\prime}), where β„“β‰₯|cjnβˆ’1|+1\ell\geq|c_{j_{n-1}}|+1 and β„“β€²β‰₯|cknβˆ’1|+1\ell^{\prime}\geq|c_{k_{n-1}}|+1. We say that s↾𝐀s\restriction\mathbf{A} and t↾𝐁t\restriction\mathbf{B} are similar, and write sβ†Ύπ€βˆΌt↾𝐁s\restriction\mathbf{A}\sim t\restriction\mathbf{B}, if and only if for each i<ni<n, s​(cji)∼t​(cki)s(c_{j_{i}})\sim t(c_{k_{i}}).

Fact 2.9 ([2]).

Let A=⟨vji:i<n⟩A=\langle v_{j_{i}}:i<n\rangle and B=⟨vki:i<n⟩B=\langle v_{k_{i}}:i<n\rangle be sets of vertices in 𝐊\mathbf{K}, and let 𝐀:=πŠβ†ΎA\mathbf{A}:=\mathbf{K}\restriction A and 𝐁:=πŠβ†ΎB\mathbf{B}:=\mathbf{K}\restriction B. Then 𝐀\mathbf{A} and 𝐁\mathbf{B} are isomorphic as ordered substructures of 𝐊\mathbf{K}, if and only if

  1. (1)

    cjic_{j_{i}} and ckic_{k_{i}} contain the same parameter-free formulas, for each i<ni<n; and

  2. (2)

    cjiβ†Ύ(πŠβ†Ύ{vjm:m<i})∼ckiβ†Ύ(πŠβ†Ύ{vkm:m<i})c_{j_{i}}\restriction(\mathbf{K}\restriction\{v_{j_{m}}:m<i\})\sim c_{k_{i}}\restriction(\mathbf{K}\restriction\{v_{k_{m}}:m<i\}), for all i<ni<n.

A lexicographic order β‰Ί\prec on π•Š\mathbb{S} is induced by fixing a linear ordering on the relation symbols in β„’\mathcal{L} and their negations. We may assume that the negated relation symbols appear in the linear order before the relation symbols. Since any node of π•Š\mathbb{S} is completely determined by such atomic and negated atomic formulas, this lexicographic order gives rise to a linear order on π•Š\mathbb{S}, which we again denote by β‰Ί\prec, with the following properties: If s⊊ts\subsetneq t, then sβ‰Ίts\prec t. For any incomparable s,tβˆˆπ•Šs,t\in\mathbb{S}, if |s∧t|=n|s\wedge t|=n, then sβ‰Ίts\prec t if and only if sβ†Ύ(n+1)β‰Ίtβ†Ύ(n+1)s\restriction(n+1)\prec t\restriction(n+1). This order β‰Ί\prec generalizes the lexicographic order for the case of binary relational structures in [14], [11], [6], [5], and [17].

Given SβŠ†π•ŠS\subseteq\mathbb{S}, let ⟨ciS:i<n⟩\langle c^{S}_{i}:i<n\rangle enumerate the coding nodes in SS in increasing length, where nβˆˆβ„•βˆͺ{β„•}n\in\mathbb{N}\cup\{\mathbb{N}\}.

Definition 2.10 (Similarity Map, [2]).

Let SS and TT be meet-closed subsets of π•Š\mathbb{S}. A function f:Sβ†’Tf:S\rightarrow T is a similarity map of SS to TT if ff is a bijection and for all nodes s,tβˆˆπ•Šs,t\in\mathbb{S}, the following hold:

  1. (1)

    ff preserves β‰Ί\prec: sβ‰Ίts\prec t if and only if f​(s)β‰Ίf​(t)f(s)\prec f(t).

  2. (2)

    ff preserves meets, and hence splitting nodes: f​(s∧t)=f​(s)∧f​(t)f(s\wedge t)=f(s)\wedge f(t).

  3. (3)

    ff preserves relative lengths: |s|<|t||s|<|t| if and only if |f​(s)|<|f​(t)||f(s)|<|f(t)|.

  4. (4)

    ff preserves initial segments: sβŠ†ts\subseteq t if and only if f​(s)βŠ†f​(t)f(s)\subseteq f(t).

  5. (5)

    ff preserves coding nodes and their parameter-free formulas: Given ciS∈Sc_{i}^{S}\in S, then f​(ciS)=ciTf(c_{i}^{S})=c^{T}_{i}; moreover, for Ξ³βˆˆΞ“\gamma\in\Gamma, γ​(viS)\gamma(v^{S}_{i}) holds in 𝐊\mathbf{K} if and only if γ​(viT)\gamma(v^{T}_{i}) holds in 𝐊\mathbf{K}, where viSv^{S}_{i} and viTv^{T}_{i} are the vertices of 𝐊\mathbf{K} represented by coding nodes ciSc^{S}_{i} and ciTc^{T}_{i}, respectively.

  6. (6)

    ff preserves relative passing types: s​(ciS)∼f​(s)​(ciT)s(c^{S}_{i})\sim f(s)(c^{T}_{i}), for all coding nodes ciSc^{S}_{i} in SS.

When there is a similarity map between SS and TT, we say that SS and TT are similar and write S∼TS\sim T. Given a subtree SS of π•Š\mathbb{S}, let Sim⁑(S)\operatorname{Sim}(S) denote the collection of all subtrees TT of π•Š\mathbb{S} which are similar to SS. If Tβ€²βŠ†TT^{\prime}\subseteq T and ff is a similarity map of SS to Tβ€²T^{\prime}, then ff is called a similarity embedding of SS into TT.

2.4. The properties SDAP+ and LSDAP+

The property SDAP+ is a strengthening of SDAP, and LSDAP+ is a β€˜labeled’ version of SDAP+.

Definition 2.11 (Subtree).

Let TT be a subset of π•Š\mathbb{S}, and let LL be the set of lengths of coding nodes in TT and lengths of meets of two incomparable nodes (not necessarily coding nodes) in TT. Then TT is a subtree of π•Š\mathbb{S} if TT is closed under meets and closed under initial segments with lengths in LL.

Definition 2.12 (Diagonal tree).

A subtree TβŠ†π•ŠT\subseteq\mathbb{S} is diagonal if each level of TT has at most one splitting node, each splitting node in TT has degree two (exactly two immediate successors), and coding node levels in TT have no splitting nodes.

Subtrees of π•Š\mathbb{S} correspond to substructures of 𝐊\mathbf{K}, and vice versa for diagonal subtrees: Given a subtree TβŠ†π•ŠT\subseteq\mathbb{S}, let NTN^{T} denote the set of natural numbers nn such that cn∈Tc_{n}\in T; let πŠβ†ΎT\mathbf{K}\restriction T denote the substructure of 𝐊\mathbf{K} on universe {vn:n∈NT}\{v_{n}:n\in N^{T}\}. We call πŠβ†ΎT\mathbf{K}\restriction T the substructure of 𝐊\mathbf{K} represented by the coding nodes in TT, or simply the substructure represented by TT. In the reverse direction, given a substructure πŒβ‰€πŠ\mathbf{M}\leq\mathbf{K}, let π•Šβ†ΎπŒ\mathbb{S}\restriction\mathbf{M} denote the subtree of π•Š\mathbb{S} induced by the meet-closure of the coding nodes {cn:vn∈M}\{c_{n}:v_{n}\in\mathrm{M}\}, and call π•Šβ†ΎπŒ\mathbb{S}\restriction\mathbf{M} the subtree of π•Š\mathbb{S} induced by 𝐉\mathbf{J}. If TT is a diagonal subtree and 𝐌=πŠβ†ΎT\mathbf{M}=\mathbf{K}\restriction T, then π•Šβ†ΎπŒ=T\mathbb{S}\restriction\mathbf{M}=T, as TT being diagonal ensures that the coding nodes in π•Šβ†ΎπŒ\mathbb{S}\restriction\mathbf{M} are exactly those in TT.

Given two substructures 𝐀,𝐁\mathbf{A},\mathbf{B} of 𝐊\mathbf{K}, we write 𝐀≅ω𝐁\mathbf{A}\cong^{\omega}\mathbf{B} when there exists an β„’\mathcal{L}-isomorphism between 𝐀\mathbf{A} and 𝐁\mathbf{B} that preserves the linear order on their universes. It follows from Fact 2.9 that for any subtrees S,TβŠ†π•ŠS,T\subseteq\mathbb{S}, S∼TS\sim T implies that πŠβ†ΎSβ‰…Ο‰πŠβ†ΎT\mathbf{K}\restriction S\cong^{\omega}\mathbf{K}\restriction T.

Notation 2.13.

Given a diagonal subtree TT, let ⟨cnT:n<N⟩\langle c^{T}_{n}:n<N\rangle denote the enumeration of the coding nodes in TT in order of increasing length, and let β„“nT\ell^{T}_{n} denote |cnT||c^{T}_{n}|, the length of cnTc^{T}_{n}. For a finite subset AβŠ†TA\subseteq T, let

(3) β„“A=max⁑{|t|:t∈A}andmax⁑(A)={s∈A:|s|=β„“A}.\ell_{A}=\max\{|t|:t\in A\}\mathrm{\ \ and\ \ }\max(A)=\{s\in A:|s|=\ell_{A}\}.

For any subset AβŠ†TA\subseteq T, let

(4) Aβ†Ύβ„“={tβ†Ύβ„“:t∈A​and​|t|β‰₯β„“}A\restriction\ell=\{t\restriction\ell:t\in A\mathrm{\ and\ }|t|\geq\ell\}

and let

(5) Aβ†Ώβ„“={t∈A:|t|<β„“}βˆͺAβ†Ύβ„“.A\upharpoonleft\ell=\{t\in A:|t|<\ell\}\cup A\restriction\ell.

Thus, Aβ†Ύβ„“A\restriction\ell is a level set, while Aβ†Ώβ„“A\upharpoonleft\ell is the set of nodes in AA with length less than β„“\ell along with the truncation to β„“\ell of the nodes in AA of length at least β„“\ell. For A,BβŠ†TA,B\subseteq T, BB is an initial segment of AA if B=Aβ†Ώβ„“B=A\upharpoonleft\ell for some β„“\ell equal to the length of some node in AA. In this case, we also say that AA end-extends (or just extends) BB. If β„“\ell is not the length of any node in AA, then Aβ†Ώβ„“A\upharpoonleft\ell is not a subset of AA, but is a subset of A^\widehat{A}, where A^\widehat{A} denotes {tβ†Ύn:t∈A​and​n≀|t|}\{t\restriction n:t\in A\mathrm{\ and\ }n\leq|t|\}. Given t∈Tt\in T at the level of a coding node in TT, tt has exactly one immediate successor in T^\widehat{T}, which we denote by t+t^{+}.

Given a substructure 𝐌\mathbf{M} of 𝐊\mathbf{K}, let 𝐌n\mathbf{M}_{n} denote 𝐌\mathbf{M} restricted to its first nn vertices. A tree TT is perfect if each node in TT has at least two incomparable extensions in TT.

Definition 2.14 (Diagonal Coding Subtree).

A subtree TβŠ†π•ŠT\subseteq\mathbb{S} is called a diagonal coding subtree if TT is diagonal, perfect, 𝐌:=πŠβ†ΎTβ‰…πŠ\mathbf{M}:=\mathbf{K}\restriction T\cong\mathbf{K}, and the following holds:

  1. Suppose s∈Ts\in T with |s|=β„“iβˆ’1T+1|s|=\ell^{T}_{i-1}+1 for some iβ‰₯1i\geq 1, or else suppose ss is the stem of TT and let i=0i=0. Then for each n>in>i and each 11-type Ο„\tau over 𝐊n\mathbf{K}_{n} such that Ο„β†ΎπŠi∼s\tau\restriction\mathbf{K}_{i}\sim s, there is a t∈Tβ†Ύ(β„“nβˆ’1T+1)t\in T\restriction(\ell^{T}_{n-1}+1) extending ss such that tβ†ΎπŒnβˆΌΟ„t\restriction\mathbf{M}_{n}\sim\tau.

Definition 2.15 (Diagonal Coding Tree Property, [2]).

A FraΓ―ssé class 𝒦\mathcal{K} in language β„’\mathcal{L} satisfies the Diagonal Coding Tree Property (DCTP) if given any enumerated FraΓ―ssé structure 𝐊\mathbf{K} for 𝒦\mathcal{K}, there is a diagonal coding subtree.

Definition 2.16 (++-Similarity, [2]).

Let TT be a diagonal coding tree for the FraΓ―ssé limit 𝐊\mathbf{K} of a FraΓ―ssé class 𝒦\mathcal{K}, and suppose AA and BB are finite subtrees of TT. We write A∼+BA\stackrel{{\scriptstyle+}}{{\sim}}B and say that AA and BB are ++-similar if and only if A∼BA\sim B and one of the following two cases holds:

    1. Case 1.

      If max⁑(A)\max(A) has a splitting node in TT, then so does max⁑(B)\max(B), and the similarity map from AA to BB takes the splitting node in max⁑(A)\max(A) to the splitting node in max⁑(B)\max(B).

    1. Case 2.

      If max⁑(A)\max(A) has a coding node, say cnAc^{A}_{n}, and f:Aβ†’Bf:A\rightarrow B is the similarity map, then s+​(n)∼f​(s)+​(n)s^{+}(n)\sim f(s)^{+}(n) for each s∈max⁑(A)s\in\max(A).

Note that ∼+\stackrel{{\scriptstyle+}}{{\sim}} is an equivalence relation, and A∼+BA\stackrel{{\scriptstyle+}}{{\sim}}B implies A∼BA\sim B. When A∼BA\sim B (A∼+BA\stackrel{{\scriptstyle+}}{{\sim}}B), we say that they have the same similarity type (++-similarity type).

Definition 2.17 (Extension Property, [2]).

We say that 𝐊\mathbf{K} has the Extension Property when the following condition (EP) holds:

  1. (EP)

    Suppose AA is a finite or infinite subtree of some Tβˆˆπ’―T\in\mathcal{T}. Let kk be given and suppose max⁑(rk+1​(A))\max(r_{k+1}(A)) has a splitting node. Suppose that BB is a ++-similarity copy of rk​(A)r_{k}(A) in TT. Let uu denote the splitting node in max⁑(rk+1​(A))\max(r_{k+1}(A)), and let ss denote the node in max(B)+\max(B)^{+} which must be extended to a splitting node in order to obtain a ++-similarity copy of rk+1​(A)r_{k+1}(A). If sβˆ—s^{*} is a splitting node in TT extending ss, then there are extensions of the rest of the nodes in max(B)+\max(B)^{+} to the same length as sβˆ—s^{*} resulting in a ++-similarity copy of rk+1​(A)r_{k+1}(A) which can be extended to a copy of AA.

It was shown in Lemma 4.17 of [2] that SFAP implies the Extension Property, and that moreover, any SFAP class with an additional linear order also easily satisfy the Extension Property.

Definition 2.18 (SDAP+, [2]).

A FraΓ―ssé class 𝒦\mathcal{K} satisfies SDAP+ if and only if 𝒦\mathcal{K} satisfies SDAP and any FraΓ―ssé limit 𝐊\mathbf{K} of 𝒦\mathcal{K} with universe β„•\mathbb{N} satisfies the Diagonal Coding Tree Property and the Extension Property.

Remark 2.19.

If there exists an enumerated FraΓ―ssé limit 𝐊\mathbf{K} of 𝒦\mathcal{K} satisfying DCTP and EP, then every enumerated FraΓ―ssé limit of 𝒦\mathcal{K} also satisfies DCTP and EP. Thus, the property SDAP+ is truly a property of the FraΓ―ssé class 𝒦\mathcal{K}. This being the case, we will refer to both a FraΓ―ssé class and its FraΓ―ssé limit as satisfying SDAP+.

The Labeled Substructure Disjoint Amalgamation Property is a labeled version applicable to structures such as β„šβ„š\mathbb{Q}_{\mathbb{Q}}, β„šβ„šβ„š\mathbb{Q}_{{\mathbb{Q}}_{\mathbb{Q}}}, and so forth (see Subsection 4.4 in [2]). Since the proofs for SDAP+ and LSDAP+ structures are almost identical, for the sake of space, we will provide proofs for SDAP+ structures.

Note that any SDAP+ and LSDAP+ structures can be encoded inside π•Œ\mathbb{U} according to the following convention.

Convention 2.20.

There is some kβ‰₯1k\geq 1, a partition PiP_{i} (i<ki<k) of the unary relation symbols in β„’\mathcal{L} and subtrees TiβŠ†π•ŒT_{i}\subseteq\mathbb{U} so that the following hold:

  1. (1)

    T:=⋃i<kTiT:=\bigcup_{i<k}T_{i} forms a kk-rooted diagonal coding tree;

  2. (2)

    For each i<ki<k, the coding nodes in TiT_{i} have only unary relations from PiP_{i}, and all the unary relation symbols from PiP_{i} occur densely in TiT_{i};

  3. (3)

    Property (2) persists in every coding subtree of TT.

In this way, the partition PiP_{i}, i<ki<k, is optimal and persistent.

For simplicity, though, we will assume the following convention.

Convention 2.21.

Let 𝒦\mathcal{K} be a FraΓ―ssé class in a language β„’\mathcal{L} and 𝐊\mathbf{K} a FraΓ―ssé limit of 𝒦\mathcal{K}. If (a) 𝒦\mathcal{K} satisfies SFAP, or (b) 𝐊\mathbf{K} satisfies SDAP+ and either has no unary relations or has no transitive relations, then we work inside a diagonal coding subtree 𝕋\mathbb{T} of π•Š\mathbb{S}. Otherwise, we work inside a diagonal coding subtree 𝕋\mathbb{T} of π•Œ\mathbb{U}.

Convention 2.21 is a special case of Convention 2.20. We will prove the main theorems assuming Convention 2.21, noting that it is straightforward to recover the general results under Convention 2.20.

We conclude this section with the result on big Ramsey degrees from [3]. A set of incomparable coding nodes AβŠ†π•ŒA\subseteq\mathbb{U} is called an antichain. An antichain of coding nodes in π•Œ\mathbb{U} is called a diagonal antichain if the tree induced by the meet closure of the antichain is diagonal.

Definition 2.22 (Diagonal Coding Antichain, [2]).

A diagonal antichain AβŠ†π•ŒA\subseteq\mathbb{U} is called a diagonal coding antichain (DCA) if πŠβ†ΎAβ‰…πŠ\mathbf{K}\restriction A\cong\mathbf{K} and (the tree induced by) AA is a Diagonal Coding Subtree.

Lemma 2.23 ([2]).

Suppose 𝒦\mathcal{K} is a FraΓ―ssé class satisfying SDAP+ or LSDAP+. Then there is an infinite diagonal antichain of coding nodes π”»βŠ†π•Œ\mathbb{D}\subseteq\mathbb{U} so that πŠβ†Ύπ”»β‰…Ο‰πŠ\mathbf{K}\restriction\mathbb{D}\cong^{\omega}\mathbf{K}.

Theorem 2.24 (Coulson, Dobrinen, Patel, [3]).

Let 𝒦\mathcal{K} be a FraΓ―ssé class satisfying SDAP+ or LSDAP+ in a finite relational language with relation symbols of arity at most two. Then for each finite structure π€βˆˆπ’¦\mathbf{A}\in\mathcal{K}, the big Ramsey degree of 𝐀\mathbf{A} in 𝐊\mathbf{K} is exactly the number of similarity types of diagonal antichains representing a copy of 𝐀\mathbf{A}.

3. Baire spaces of diagonal coding antichains

We now set up the subspaces of the Baire space for which we will prove analogues of the Galvin–Prikry and Ellentuck Theorems. Given a FraΓ―ssé structure 𝐊\mathbf{K} with universe β„•\mathbb{N}, each subcopy of 𝐊\mathbf{K} can be identified with its universe, an infinite subset of β„•\mathbb{N}. Thus, the collection (𝐊𝐊){\mathbf{K}\choose\mathbf{K}} of all subcopies of 𝐊\mathbf{K} is naturally identified with a subspace of the Baire space [β„•]β„•[\mathbb{N}]^{\mathbb{N}}.

The existence of big Ramsey degrees greater than one precludes any simplistic approach to infinite-dimensional Ramsey theorems in terms only of definable sets on the full space (𝐊𝐊){\mathbf{K}\choose\mathbf{K}} of subcopies of 𝐊\mathbf{K}. At the same time, Theorem 2.24 shows us where to look for viable infinite-dimensional Ramsey theorems: precisely on subspaces of the Baire space determined by collections of coding subtrees which are all similar to each other. While infinite-dimensional Ramsey theorems for all such spaces follow from the methods in this paper, we will concentrate on spaces of antichains similar to some fixed diagonal coding antichain, as such spaces will additionally recover exact big Ramsey degrees.

Definition 3.1 (Spaces of Diagonal Coding Antichains).

Let 𝒦\mathcal{K} be a FraΓ―ssé class satisfying SDAP+, and let 𝐊\mathbf{K} be a FraΓ―ssé limit of 𝒦\mathcal{K} with universe β„•\mathbb{N}. Let 𝔻\mathbb{D} be any diagonal coding antichain; that is, 𝔻\mathbb{D} is a diagonal antichain of coding nodes in π•Šβ€‹(𝐊)\mathbb{S}(\mathbf{K}) or π•Œβ€‹(𝐊)\mathbb{U}(\mathbf{K}) representing a subcopy of 𝐊\mathbf{K}. (Recall Convention 2.21.)

Let π’Ÿβ€‹(𝔻)\mathcal{D}(\mathbb{D}) denote the collection of all subsets MβŠ†π”»M\subseteq\mathbb{D} such that MβˆΌπ”»M\sim\mathbb{D}. The partial ordering ≀\leq on π’Ÿβ€‹(𝔻)\mathcal{D}(\mathbb{D}) is simply inclusion. For Mβˆˆπ’Ÿβ€‹(𝔻)M\in\mathcal{D}(\mathbb{D}), when we write N≀MN\leq M, it is implied that Nβˆˆπ’Ÿβ€‹(𝔻)N\in\mathcal{D}(\mathbb{D}). When 𝔻\mathbb{D} is understood, we simply write π’Ÿ\mathcal{D}.

Each diagonal coding antichain Mβˆˆπ’ŸM\in\mathcal{D} uniquely determines the substructure 𝐌:=πŠβ†Ύ{vi:ci∈M}\mathbf{M}:=\mathbf{K}\restriction\{v_{i}:c_{i}\in M\}. Since MβˆΌπ”»M\sim\mathbb{D}, it follows that πŒβ‰…Ο‰πŠ\mathbf{M}\cong^{\omega}\mathbf{K}. Let 𝐃\mathbf{D} denote πŠβ†Ύπ”»\mathbf{K}\restriction\mathbb{D}, and let

(6) πŠβ€‹(𝐃)={𝐌:Mβˆˆπ’Ÿ}.\mathbf{K}(\mathbf{D})=\{\mathbf{M}:M\in\mathcal{D}\}.

Then πŠβ€‹(𝐃)\mathbf{K}(\mathbf{D}) is identified with the subspace {{iβˆˆβ„•:vi∈𝐌}:πŒβˆˆπŠβ€‹(𝐃)}\{\{i\in\mathbb{N}:v_{i}\in\mathbf{M}\}:\mathbf{M}\in\mathbf{K}(\mathbf{D})\} of the Baire space. These are the spaces for which we will prove infinite-dimensional Ramsey theorems in Section 6.

Each diagonal coding antichain Mβˆˆπ’ŸM\in\mathcal{D} can also be identified with the tree induced by its meet-closure. The set of coding and splitting nodes in (the tree induced by) MM are called the critical nodes of MM, and we let ⟨dnM:nβˆˆβ„•βŸ©\langle d^{M}_{n}:n\in\mathbb{N}\rangle denote their enumeration in order of increasing length. For nβˆˆβ„•n\in\mathbb{N}, M​(n)M(n) denotes the set of nodes in MM of length |dnM||d^{M}_{n}|. Given kβˆˆβ„•k\in\mathbb{N}, rk​(M)r_{k}(M) denotes the finite subset of MM consisting of all nodes in MM with length less than |dkM||d^{M}_{k}|. Thus, r0​(M)r_{0}(M) is the empty set and

(7) rk​(M)=⋃n<kM​(n).r_{k}(M)=\bigcup_{n<k}M(n).

We define the following notation in line with topological Ramsey space theory from [16]. For kβˆˆβ„•k\in\mathbb{N}, define

(8) π’œβ€‹π’Ÿk={rk​(M):Mβˆˆπ’Ÿ},\mathcal{AD}_{k}=\{r_{k}(M):M\in\mathcal{D}\},

the set of all kk-th restrictions of members of π’Ÿ\mathcal{D}. Let

(9) π’œβ€‹π’Ÿ=⋃k=1βˆžπ’œβ€‹π’Ÿk,\mathcal{AD}=\bigcup_{k=1}^{\infty}\mathcal{AD}_{k},

the set of all finite approximations to members of π’Ÿ\mathcal{D}. Note that having identified MM with the tree it induces, members of π’œβ€‹π’Ÿ\mathcal{AD} are finite diagonal trees which are initial segments of the diagonal tree MM.

For A,Bβˆˆπ’œβ€‹π’ŸA,B\in\mathcal{AD} we write AβŠ‘BA\sqsubseteq B if and only if there is some Mβˆˆπ’ŸM\in\mathcal{D} and some j≀kj\leq k such that A=rj​(M)A=r_{j}(M) and B=rk​(M)B=r_{k}(M). In this case, AA is called an initial segment of BB; we also say that BB extends AA. If AβŠ‘BA\sqsubseteq B and Aβ‰ BA\neq B, then we say that AA is a proper initial segment of BB and write A⊏BA\sqsubset B. When A=rj​(M)A=r_{j}(M) for some jj, we also write A⊏MA\sqsubset M and call AA an initial segment of MM.

The metric topology on π’Ÿ\mathcal{D} is the topology induced by basic open cones of the form

(10) [A,𝔻]={Mβˆˆπ’Ÿ:βˆƒk​(rk​(M)=A)},[A,\mathbb{D}]=\{M\in\mathcal{D}:\exists k\,(r_{k}(M)=A)\},

for Aβˆˆπ’œβ€‹π’ŸA\in\mathcal{AD}. The Ellentuck topology on π’Ÿ\mathcal{D} is induced by basic open sets of the form

(11) [A,M]={Nβˆˆπ’Ÿ:βˆƒk​(rk​(N)=A)​and​N≀M},[A,M]=\{N\in\mathcal{D}:\exists k\,(r_{k}(N)=A)\mathrm{\ and\ }N\leq M\},

where Aβˆˆπ’œβ€‹π’ŸA\in\mathcal{AD} and Mβˆˆπ’ŸM\in\mathcal{D}. Thus, the Ellentuck topology refines the metric topology.

Given Aβˆˆπ’œβ€‹π’ŸA\in\mathcal{AD}, let β„“A\ell_{A} denote the maximum of the lengths of nodes in AA, and let

(12) max⁑(A)={s∈A:|s|=β„“A}.\max(A)=\{s\in A:|s|=\ell_{A}\}.

The partial ordering ≀fin\leq_{\mathrm{fin}} on π’œβ€‹π’Ÿ\mathcal{AD} is defined as follows: For A,Bβˆˆπ’œβ€‹π’ŸA,B\in\mathcal{AD}, write A≀finBA\leq_{\mathrm{fin}}B if and only if AA is a subtree of BB. Define depthM⁑(A)\operatorname{depth}_{M}(A) to be the least integer kk such that A≀finrk​(M)A\leq_{\mathrm{fin}}r_{k}(M), if it exists; otherwise, define depthM⁑(A)=∞\operatorname{depth}_{M}(A)=\infty. Lastly, given j<kj<k, Aβˆˆπ’œβ€‹π’ŸjA\in\mathcal{AD}_{j} and Mβˆˆπ’ŸM\in\mathcal{D}, define

(13) rk​[A,M]={rk​(N):N∈[A,M]}.r_{k}[A,M]=\{r_{k}(N):N\in[A,M]\}.

4. Forcing the Extended Pigeonhole Principle

This section proves an enhanced pigeonhole principle for Baire spaces of diagonal coding antichains which preserves the width in some finite initial segment of the ambient antichain. This is done to prove the pigeonhole principle (axiom A.4 of Todorcevic) while simultaneously overcoming the fact that the amalgamation axiom A.3(2) of Todorcevic does not hold for many spaces of the form π’Ÿβ€‹(𝔻)\mathcal{D}(\mathbb{D}). The proof will use forcing techniques to do infinitely many unbounded searches for finite objects with some homogeneity properties. Since the objects are finite, they exist in the ground model; no generic extension is needed for the main theorem of this section.

Definition 4.1 (Good Diagonal Coding Antichains).

Fix an enumerated Fraïssé structure 𝐊\mathbf{K} satisfying SDAP+. We call a diagonal coding antichain MM good if it satisfies the following:

  1. (1)

    For each nβˆˆβ„•n\in\mathbb{N}, the longest splitting node in MM with length less than |cnM||c^{M}_{n}| extends β‰Ί\prec-right to cnMc^{M}_{n}. We call this splitting node the splitting predecessor of cnMc^{M}_{n} and denote it by spM​(cnM)\mathrm{sp}_{M}(c^{M}_{n}).

  2. (2)

    For any m<nm<n, letting ss be the β‰Ί\prec-left extension of spM​(cmM)\mathrm{sp}_{M}(c^{M}_{m}) in Mβ†Ύ(|cmM|+1)M\restriction(|c^{M}_{m}|+1) and tt be the β‰Ί\prec-left extension of spM​(cnM)\mathrm{sp}_{M}(c^{M}_{n}) in Mβ†Ύ(|cnM|+1)M\restriction(|c^{M}_{n}|+1), we have s​(cmM)∼t​(cnM)s(c^{M}_{m})\sim t(c^{M}_{n}).

  3. (3)

    There is a kβˆˆβ„•k\in\mathbb{N} such that for all nβ‰₯kn\geq k, to each 11-type Οƒ\sigma over 𝐊n+1\mathbf{K}_{n+1} there corresponds a unique node s∈Mβ†Ύ(|cnM|+1)s\in M\restriction(|c^{M}_{n}|+1) such that tp⁑(s/𝐌n+1)βˆΌΟƒ\operatorname{tp}(s/\mathbf{M}_{n+1})\sim\sigma.

Fix throughout this section a good diagonal coding antichain 𝔻\mathbb{D} for 𝐊\mathbf{K}, and let π’Ÿ\mathcal{D} denote π’Ÿβ€‹(𝔻)\mathcal{D}(\mathbb{D}). For any subset UβŠ†π”»U\subseteq\mathbb{D}, finite or infinite, let LUL_{U} denote {|t|:t∈U}\{|t|:t\in U\}, the set of lengths of nodes in UU, and let U∧U^{\wedge} denote the meet-closure of UU. Define

(14) U^={tβ†Ύβ„“:t∈U​and​ℓ≀|t|},\widehat{U}=\{t\restriction\ell:t\in U\mathrm{\ and\ }\ell\leq|t|\},

the tree of all initial segments of members of UU, and

(15) tree⁑(U)={t∈U^:|t|∈LU∧},\operatorname{tree}({U})=\{t\in\widehat{U}:|t|\in L_{U^{\wedge}}\},

the tree induced by the meet-closure of UU. We will abuse notation an identify UU with tree(U)(U).

Let

(16) π’œβ€‹π’Ÿ^={Aβ†Ώβ„“:Aβˆˆπ’œβ€‹π’Ÿβ€‹and​ℓ≀ℓA}.\widehat{\mathcal{AD}}=\{A\upharpoonleft\ell:A\in\mathcal{AD}\mathrm{\ and\ }\ell\leq\ell_{A}\}.

Given Mβˆˆπ’ŸM\in\mathcal{D}, let π’œβ€‹π’Ÿβ€‹(M)\mathcal{AD}(M) denote the members of π’œβ€‹π’Ÿ\mathcal{AD} which are contained in MM. For kβˆˆβ„•k\in\mathbb{N}, let π’œβ€‹π’Ÿk​(M)\mathcal{AD}_{k}(M) denote the set of those Aβˆˆπ’œβ€‹π’ŸkA\in\mathcal{AD}_{k} such that AA is a subtree of MM. Define

(17) π’œβ€‹π’Ÿ^​(M)={Aβ†Ώβ„“:Aβˆˆπ’œβ€‹π’Ÿβ€‹(M)​andβ€‹β„“βˆˆLM}.\widehat{\mathcal{AD}}(M)=\{A\upharpoonleft\ell:A\in\mathcal{AD}(M)\mathrm{\ and\ }\ell\in L_{M}\}.

Note that for any Mβˆˆπ’ŸM\in\mathcal{D}, there are members of π’œβ€‹π’Ÿ^​(M)\widehat{\mathcal{AD}}(M) which are not similar to rn​(𝔻)r_{n}(\mathbb{D}) for any nn, and hence are not members of π’œβ€‹π’Ÿβ€‹(M)\mathcal{AD}(M).

Definition 4.2.

Given Mβˆˆπ’ŸM\in\mathcal{D} and Bβˆˆπ’œβ€‹π’Ÿ^​(M)B\in\widehat{\mathcal{AD}}(M), letting mm be the least integer for which there exists Bβ€²βˆˆπ’œβ€‹π’ŸmB^{\prime}\in\mathcal{AD}_{m} such that max⁑(B)βŠ‘max⁑(Bβ€²)\max(B)\sqsubseteq\max(B^{\prime}), define

(18) [B,M]βˆ—={Nβˆˆπ’Ÿ:max⁑(B)βŠ‘max⁑(rm​(N))​and​N≀M}.[B,M]^{*}=\{N\in\mathcal{D}:\max(B)\sqsubseteq\max(r_{m}(N))\mathrm{\ and\ }N\leq M\}.

For nβ‰₯mn\geq m, define

(19) rn​[B,M]βˆ—={rn​(N):N∈[B,M]βˆ—},r_{n}[B,M]^{*}=\{r_{n}(N):N\in[B,M]^{*}\},

and let

(20) r​[B,M]βˆ—=⋃m≀nrn​[B,M]βˆ—.r[B,M]^{*}=\bigcup_{m\leq n}r_{n}[B,M]^{*}.

Given Bβˆˆπ’œβ€‹π’Ÿ^B\in\widehat{\mathcal{AD}} and Mβˆˆπ’ŸM\in\mathcal{D}, notice that the set [B,M]βˆ—[B,M]^{*} from Definition 4.2 is open in the Ellentuck topology on π’Ÿ\mathcal{D}: If BB is in π’œβ€‹π’Ÿk\mathcal{AD}_{k} for some kk, then the set [B,M]βˆ—[B,M]^{*} is the union of [B,M][B,M] along with all [C,M][C,M], where Cβˆˆπ’œβ€‹π’ŸkC\in\mathcal{AD}_{k} and max⁑(C)\max(C) end-extends max⁑(B)\max(B). If BB is in π’œβ€‹π’Ÿ^\widehat{\mathcal{AD}} but not in π’œβ€‹π’Ÿ\mathcal{AD}, then letting kk be the least integer for which there is some Cβˆˆπ’œβ€‹π’ŸkC\in\mathcal{AD}_{k} with max⁑(C)⊐max⁑(B)\max(C)\sqsupset\max(B), we see that [B,M]βˆ—[B,M]^{*} equals the union of all [C,M][C,M], where Cβˆˆπ’œβ€‹π’ŸkC\in\mathcal{AD}_{k} and BβŠ‘CB\sqsubseteq C. For the same reasons, the set [B,𝔻]βˆ—[B,\mathbb{D}]^{*} is open in the metric topology on π’Ÿ\mathcal{D}. Notice also that rn​[B,M]βˆ—r_{n}[B,M]^{*} defined in equation (19) is equal to {Cβˆˆπ’œβ€‹π’Ÿn​(M):max⁑(B)βŠ‘max⁑(C)}\{C\in\mathcal{AD}_{n}(M):\max(B)\sqsubseteq\max(C)\}.

Given Mβˆˆπ’ŸM\in\mathcal{D}, a splitting node s∈Ms\in M is called a splitting predecessor of a coding node in MM (or just splitting predecessor if MM is understood) if and only if there is a coding node c∈Mc\in M such that sβŠ‚cs\subset c and |c||c| is minimal in LML_{M} above |s||s|. Given a coding node cc in MM, we write spM⁑(c)\operatorname{sp}_{M}(c) to denote the splitting predecessor of cc in MM. Note that s=spM⁑(c)s=\operatorname{sp}_{M}(c) if and only if the minimal node in MM extending the β‰Ί\prec-right extension of ss is a coding node. When M=𝔻M=\mathbb{D}, we will usually write sp⁑(c)\operatorname{sp}(c) in place of sp𝔻⁑(c)\operatorname{sp}_{\mathbb{D}}(c).

Given Aβˆˆπ’œβ€‹π’Ÿ^​(M)A\in\widehat{\mathcal{AD}}(M), let A+A^{+} denote the union of AA with the set of immediate successors in M^\widehat{M} of the members of max⁑(A)\max(A); thus,

(21) A+=Aβˆͺ{t∈Mβ†Ύ(β„“A+1):tβ†Ύβ„“A∈A}A^{+}=A\cup\{t\in M\restriction(\ell_{A}+1):t\restriction\ell_{A}\in A\}

and max⁑(A+)\max(A^{+}) is the set of all nodes in A+A^{+} with length β„“A+1\ell_{A}+1.

A set of nodes is called a level set if all nodes in the set have the same length. For level sets X,YβŠ†π”»X,Y\subseteq\mathbb{D}, we say that YY end-extends XX and write X⊏YX\sqsubset Y if and only if XX and YY have the same cardinality, β„“X<β„“Y\ell_{X}<\ell_{Y}, and Yβ†Ύβ„“X=XY\restriction\ell_{X}=X. More generally, for A,Bβˆˆπ’œβ€‹π’Ÿ^A,B\in\widehat{\mathcal{AD}}, write AβŠ‘BA\sqsubseteq B if and only if A=Bβ†Ώβ„“AA=B\upharpoonleft\ell_{A}; in this case, write A⊏BA\sqsubset B if also β„“A<β„“B\ell_{A}<\ell_{B}.

Assumption 4.3.

Fix a good diagonal coding antichain 𝔻\mathbb{D} and let π’Ÿ:=π’Ÿβ€‹(𝔻)\mathcal{D}:=\mathcal{D}(\mathbb{D}). We will be working with triples (A,B,k)(A,B,k), where Aβˆˆπ’œβ€‹π’Ÿ^A\in\widehat{\mathcal{AD}}, BβŠ†π”»^B\subseteq\widehat{\mathbb{D}}, and A⊏BβŠ†A+A\sqsubset B\subseteq A^{+}. Assume that all splitting nodes in AA, BB, and DD are not splitting predecessors in 𝔻\mathbb{D}. We consider the following combinations of one of Cases (a) or (b) with one of Cases (i) or (ii).

    1. Case (a).

      max⁑(rk+1​(𝔻))\max(r_{k+1}(\mathbb{D})) has a splitting node.

    1. Case (b).

      max⁑(rk+1​(𝔻))\max(r_{k+1}(\mathbb{D})) has a coding node.

    1. Case (i).

      kβ‰₯1k\geq 1, Aβˆˆπ’œβ€‹π’ŸkA\in\mathcal{AD}_{k}, and B=A+B=A^{+}.

    1. Case (ii).

      kβ‰₯0k\geq 0, AA has at least one node, each member of max⁑(A)\max(A) has exactly one extension in max⁑(B)\max(B) (that is, max⁑(A)⊏max⁑(B)\max(A)\sqsubset\max(B)), and A=Cβ†Ώβ„“A=C\upharpoonleft\ell for some Cβˆˆπ’œβ€‹π’Ÿk+1C\in\mathcal{AD}_{k+1} and β„“<β„“C\ell<\ell_{C} such that rk​(C)βŠ‘Ar_{k}(C)\sqsubseteq A and BβŠ‘CB\sqsubseteq C.

We point out that in Case (ii), AA may or may not be a member of π’œβ€‹π’Ÿ\mathcal{AD}.

The following theorem of ErdΕ‘sΒ and Rado will be used in the proof of Theorem 4.5.

Theorem 4.4 (ErdΕ‘s-Rado).

For r<Ο‰r<\omega and ΞΌ\mu an infinite cardinal,

β„Άr​(ΞΌ)+β†’(ΞΌ+)ΞΌr+1.\beth_{r}(\mu)^{+}\rightarrow(\mu^{+})_{\mu}^{r+1}.

We are now set up to prove the theorem which will form the basis of the result that all Borel sets in ℬ​(𝔻)\mathcal{B}(\mathbb{D}) are Ramsey.

Theorem 4.5 (Extended Pigeonhole Principle).

Let 𝔻\mathbb{D}, (A,B,k)(A,B,k), DD, dd be as in Assumption 4.3, where (A,B,k)(A,B,k) satisfies one of Cases (i) or (ii) and one of Cases (a) or (b). Let h:rk+1​[D,𝔻]βˆ—β†’2h:r_{k+1}[D,\mathbb{D}]^{*}\rightarrow 2 be a coloring. Then there is an N∈[D,𝔻]βˆ—N\in[D,\mathbb{D}]^{*} such that hh is monochromatic on rk+1​[B,N]βˆ—r_{k+1}[B,N]^{*}.

Proof.

Assume the hypotheses. Let 𝐒+1\mathbf{i}+1 be the number of nodes in max⁑(B)\max(B), and fix an enumeration s0,…,s𝐒s_{0},\dots,s_{\mathbf{i}} of the nodes in max⁑(B)\max(B) with the property that for any C∈rk+1​[B,𝔻]βˆ—C\in r_{k+1}[B,\mathbb{D}]^{*}, the critical node in max⁑(C)\max(C) extends s𝐒s_{\mathbf{i}}. Note that in Case (b), 𝐒\mathbf{i} must be at least two. Let dd denote the integer such that Dβˆˆπ’œβ€‹π’ŸdD\in\mathcal{AD}_{d}, and let II denote the set of all n>dn>d such that for some (equivalently, all) M∈[D,𝔻]M\in[D,\mathbb{D}] there is a member C∈rk+1​[B,M]βˆ—C\in r_{k+1}[B,M]^{*} with depthM⁑(C)=n\operatorname{depth}_{M}(C)=n. Let LL denote the set {β„“rn​(M):M∈[D,𝔻]\{\ell_{r_{n}(M)}:M\in[D,\mathbb{D}] and n∈I}n\in I\}. In Case (b), for β„“βˆˆL\ell\in L we let β„“β€²\ell^{\prime} denote the length of the splitting predecessor in 𝔻\mathbb{D} of the coding node in 𝔻\mathbb{D} of length β„“\ell, and let Lβ€²={β„“β€²:β„“βˆˆL}L^{\prime}=\{\ell^{\prime}:\ell\in L\}.

Given Uβˆˆπ’œβ€‹π’Ÿβˆͺπ’ŸU\in\mathcal{AD}\cup\mathcal{D} with DβŠ‘UD\sqsubseteq U, define the set ExtU⁑(B)\operatorname{Ext}_{U}(B) as follows: In Case (a), let ExtU⁑(B)\operatorname{Ext}_{U}(B) consist of those level sets XβŠ†UX\subseteq U such that X=max⁑(C)X=\max(C) for some C∈rk+1​[B,𝔻]βˆ—C\in r_{k+1}[B,\mathbb{D}]^{*}, where the splitting node in XX is not a splitting predecessor in 𝔻\mathbb{D}. In Case (b), let ExtU⁑(B)\operatorname{Ext}_{U}(B) consist of those sets XβŠ†UX\subseteq U such that XX consists of the non-coding nodes in max⁑(C)\max(C) along with the splitting predecessor in 𝔻\mathbb{D} of the coding node in max⁑(C)\max(C), for some C∈rk+1​[B,𝔻]βˆ—C\in r_{k+1}[B,\mathbb{D}]^{*}. We will simply write Ext⁑(B)\operatorname{Ext}(B) to mean Ext𝔻⁑(B)\operatorname{Ext}_{\mathbb{D}}(B).

The coloring hh induces a coloring hβ€²:Ext⁑(B)β†’2h^{\prime}:\operatorname{Ext}(B)\rightarrow 2 as follows: For X∈Ext⁑(B)X\in\operatorname{Ext}(B), in Case (a) define h′​(X)=h​(C)h^{\prime}(X)=h(C), where CC is the member of rk+1​[B,𝔻]βˆ—r_{k+1}[B,\mathbb{D}]^{*} such that X=max⁑(C)X=\max(C). In Case (b), define h′​(X)=h​(C)h^{\prime}(X)=h(C), where CC is the member of rk+1​[B,𝔻]βˆ—r_{k+1}[B,\mathbb{D}]^{*} such that X=(max⁑(C)βˆ–{c})βˆͺ{sp⁑(c)}X=(\max(C)\setminus\{c\})\cup\{\operatorname{sp}(c)\}, where cc denotes the coding node in max⁑(C)\max(C).

For i≀𝐒i\leq\mathbf{i}, let Ti={tβˆˆπ”»^:tβŠ‡si}T_{i}=\{t\in\widehat{\mathbb{D}}:t\supseteq s_{i}\}. Let ΞΊ=β„Ά2​𝐒\kappa=\beth_{2\mathbf{i}}, so that the partition relation ΞΊβ†’(β„΅1)β„΅02​𝐒\kappa\rightarrow(\aleph_{1})^{2\mathbf{i}}_{\aleph_{0}} holds by the ErdΕ‘s-Rado Theorem 4.4. The following forcing notion adds ΞΊ\kappa many paths through each TiT_{i}, i<𝐒i<\mathbf{i}, and one path through T𝐒T_{\mathbf{i}}.

In both Cases (a) and (b), define β„™\mathbb{P} to be the set of finite partial functions pp such that

  1. (1)

    dom⁑(p)={𝐒}Γ—Ξ΄β†’p\operatorname{dom}(p)=\{\mathbf{i}\}\times\vec{\delta}_{p}, where Ξ΄β†’p\vec{\delta}_{p} is a finite subset of ΞΊ\kappa;

  2. (2)

    p​(𝐒)∈T𝐒p(\mathbf{i})\in T_{\mathbf{i}} and {p​(i,Ξ΄):Ξ΄βˆˆΞ΄β†’p}βŠ†Ti\{p(i,\delta):\delta\in\vec{\delta}_{p}\}\subseteq T_{i} for each i<𝐒i<\mathbf{i};

  3. (3)

    For any choices of Ξ΄iβˆˆΞ΄β†’p\delta_{i}\in\vec{\delta}_{p}, i<𝐒i<\mathbf{i}, the set {p​(i,Ξ΄i):i<𝐒}βˆͺ{p​(𝐒)}\{p(i,\delta_{i}):i<\mathbf{i}\}\cup\{p(\mathbf{i})\} is a member of Ext⁑(B)\operatorname{Ext}(B).

Let β„“p\ell_{p} denote the maximal length of nodes in ran⁑(p)\operatorname{ran}(p). All nodes in ran⁑(p)\operatorname{ran}(p) will have length β„“p\ell_{p} except in Case (b), where the splitting predecessor p​(𝐒)p(\mathbf{i}) will have length β„“pβ€²\ell_{p}^{\prime}.

The partial ordering on β„™\mathbb{P} is just reverse inclusion: q≀pq\leq p if and only if β„“qβ‰₯β„“p\ell_{q}\geq\ell_{p}, Ξ΄β†’qβŠ‡Ξ΄β†’p\vec{\delta}_{q}\supseteq\vec{\delta}_{p}, q​(𝐒)βŠ‡p​(𝐒)q(\mathbf{i})\supseteq p(\mathbf{i}), and q​(i,Ξ΄)βŠ‡p​(i,Ξ΄)q(i,\delta)\supseteq p(i,\delta) for each (i,Ξ΄)βˆˆπ’Γ—Ξ΄β†’p(i,\delta)\in\mathbf{i}\times\vec{\delta}_{p}. It is routine to check that (β„™,≀)(\mathbb{P},\leq) is a separative, atomless partial order.

We point out that condition (3) in the definition of β„™\mathbb{P} is easy to satisfy since 𝐊\mathbf{K} has SDAP+: Let CC be any member of rk+1​[B,𝔻]βˆ—r_{k+1}[B,\mathbb{D}]^{*} and let ⟨ti:iβ‰€π’βŸ©\langle t_{i}:i\leq\mathbf{i}\rangle enumerate max⁑(C)\max(C) so that so that each tit_{i} extends sis_{i}. In Case (a), each p​(i,Ξ΄)p(i,\delta) only need be a node in TiT_{i} of length β„“p\ell_{p}. If (1) of the Extension Property holds for 𝐊\mathbf{K}, then p​(𝐒)p(\mathbf{i}) just needs to be a splitting node in T𝐒T_{\mathbf{i}} which is not a splitting predecessor; if (2) of the Extension Property holds, it suffices for p​(𝐒)p(\mathbf{i}) to additionally satisfy Οˆβ€‹(p​(𝐒))=Οˆβ€‹(t𝐒)\psi(p(\mathbf{i}))=\psi(t_{\mathbf{i}}). In Case (b), p​(𝐒)p(\mathbf{i}) need only be the splitting predecessor of some coding node cc in T𝐒T_{\mathbf{i}}, and each p​(i,Ξ΄)p(i,\delta) need only be a node in TiT_{i} of length β„“p\ell_{p} such that p​(i,Ξ΄)+​(c)∼ti+​(t𝐒)p(i,\delta)^{+}(c)\sim t_{i}^{+}(t_{\mathbf{i}}).

Given pβˆˆβ„™p\in\mathbb{P}, the range of pp is the set

ran⁑(p)={p​(i,Ξ΄):(i,Ξ΄)βˆˆπ’Γ—Ξ΄β†’p}βˆͺ{p​(𝐒)}.\operatorname{ran}(p)=\{p(i,\delta):(i,\delta)\in\mathbf{i}\times\vec{\delta}_{p}\}\cup\{p(\mathbf{i})\}.

If also qβˆˆβ„™q\in\mathbb{P} and q≀pq\leq p, then we let

(22) ran⁑(q)β†Ύdom⁑(p)={q​(i,Ξ΄):(i,Ξ΄)βˆˆπ’Γ—Ξ΄β†’p}βˆͺ{q​(𝐒)}.\operatorname{ran}(q)\restriction\operatorname{dom}(p)=\{q(i,\delta):(i,\delta)\in\mathbf{i}\times\vec{\delta}_{p}\}\cup\{q(\mathbf{i})\}.

For (i,Ξ±)βˆˆπ’Γ—ΞΊ(i,\alpha)\in\mathbf{i}\times\kappa, let

(23) bΛ™i,Ξ±={⟨p​(i,Ξ±),p⟩:pβˆˆβ„™β€‹andβ€‹Ξ±βˆˆΞ΄β†’p},\dot{b}_{i,\alpha}=\{\langle p(i,\alpha),p\rangle:p\in\mathbb{P}\mathrm{\ and\ }\alpha\in\vec{\delta}_{p}\},

a β„™\mathbb{P}-name for the Ξ±\alpha-th generic branch through TiT_{i}. Let

(24) b˙𝐒={⟨p​(𝐒),p⟩:pβˆˆβ„™},\dot{b}_{\mathbf{i}}=\{\langle p(\mathbf{i}),p\rangle:p\in\mathbb{P}\},

a β„™\mathbb{P}-name for the generic branch through M𝐒M_{\mathbf{i}}. Given a generic filter GβŠ†β„™G\subseteq\mathbb{P}, notice that b˙𝐒G={p​(𝐒):p∈G}\dot{b}_{\mathbf{i}}^{G}=\{p(\mathbf{i}):p\in G\}, which is a cofinal path in T𝐒T_{\mathbf{i}}. We point out that given pβˆˆβ„™p\in\mathbb{P},

(25) pβŠ©βˆ€(i,Ξ±)βˆˆπ’Γ—Ξ΄β†’p​(bΛ™i,Ξ±β†Ύβ„“p=p​(i,Ξ±)).p\Vdash\forall(i,\alpha)\in\mathbf{i}\times\vec{\delta}_{p}\,(\dot{b}_{i,\alpha}\restriction\ell_{p}=p(i,\alpha)).

Furthermore, in Case (a), p⊩(b˙𝐒↾ℓp=p​(𝐒))p\Vdash(\dot{b}_{\mathbf{i}}\restriction\ell_{p}=p(\mathbf{i})), while in Case (b), p⊩(b˙𝐒↾ℓpβ€²=p​(𝐒))p\Vdash(\dot{b}_{\mathbf{i}}\restriction\ell^{\prime}_{p}=p(\mathbf{i})).

Let GΛ™\dot{G} be the β„™\mathbb{P}-name for a generic filter, and let LΛ™G\dot{L}_{G} be a β„™\mathbb{P}-name for the set of lengths {β„“p:p∈GΛ™}\{\ell_{p}:p\in\dot{G}\}. Note that β„™\mathbb{P} forces that LΛ™GβŠ†L\dot{L}_{G}\subseteq L. Let 𝒰˙\dot{\mathcal{U}} be a β„™\mathbb{P}-name for a non-principal ultrafilter on LΛ™G\dot{L}_{G}.

We will write sets {Ξ±i:i<𝐒}\{\alpha_{i}:i<\mathbf{i}\} in [ΞΊ]𝐒[\kappa]^{\mathbf{i}} as vectors Ξ±β†’=⟨α0,…,Ξ±π’βˆ’1⟩\vec{\alpha}=\langle\alpha_{0},\dots,\alpha_{\mathbf{i}-1}\rangle in strictly increasing order. For Ξ±β†’βˆˆ[ΞΊ]𝐒\vec{\alpha}\in[\kappa]^{\mathbf{i}}, let

(26) bΛ™Ξ±β†’=⟨bΛ™0,Ξ±0,…,bΛ™π’βˆ’1,Ξ±π’βˆ’1,bΛ™π’βŸ©.\dot{b}_{\vec{\alpha}}=\langle\dot{b}_{0,\alpha_{0}},\dots,\dot{b}_{\mathbf{i}-1,\alpha_{\mathbf{i}-1}},\dot{b}_{\mathbf{i}}\rangle.

For β„“βˆˆL\ell\in L, in Case (a) let

(27) bΛ™Ξ±β†’β†Ύβ„“=⟨bΛ™0,Ξ±0β†Ύβ„“,…,bΛ™π’βˆ’1,Ξ±π’βˆ’1β†Ύβ„“,bΛ™π’β†Ύβ„“βŸ©;\dot{b}_{\vec{\alpha}}\restriction\ell=\langle\dot{b}_{0,\alpha_{0}}\restriction\ell,\dots,\dot{b}_{\mathbf{i}-1,\alpha_{\mathbf{i}-1}}\restriction\ell,\dot{b}_{\mathbf{i}}\restriction\ell\rangle;

and in Case (b), let

(28) bΛ™Ξ±β†’β†Ύβ„“=⟨bΛ™0,Ξ±0β†Ύβ„“,…,bΛ™π’βˆ’1,Ξ±π’βˆ’1β†Ύβ„“,bΛ™π’β†Ύβ„“β€²βŸ©.\dot{b}_{\vec{\alpha}}\restriction\ell=\langle\dot{b}_{0,\alpha_{0}}\restriction\ell,\dots,\dot{b}_{\mathbf{i}-1,\alpha_{\mathbf{i}-1}}\restriction\ell,\dot{b}_{\mathbf{i}}\restriction\ell^{\prime}\rangle.

Note that hβ€²h^{\prime} is a coloring on bΛ™Ξ±β†’β†Ύβ„“\dot{b}_{\vec{\alpha}}\restriction\ell whenever this is forced to be a member of ExtM⁑(B)\operatorname{Ext}_{M}(B). Given Ξ±β†’βˆˆ[ΞΊ]𝐒\vec{\alpha}\in[\kappa]^{\mathbf{i}} and pβˆˆβ„™p\in\mathbb{P} with Ξ±β†’βŠ†Ξ΄β†’p\vec{\alpha}\subseteq\vec{\delta}_{p}, let

(29) X​(p,Ξ±β†’)={p​(i,Ξ±i):i<𝐒}βˆͺ{p​(𝐒)}.X(p,\vec{\alpha})=\{p(i,\alpha_{i}):i<\mathbf{i}\}\cup\{p(\mathbf{i})\}.

For each Ξ±β†’βˆˆ[ΞΊ]𝐒\vec{\alpha}\in[\kappa]^{\mathbf{i}}, choose a condition pΞ±β†’βˆˆβ„™p_{\vec{\alpha}}\in\mathbb{P} satisfying the following:

  1. (1)

    Ξ±β†’βŠ†Ξ΄β†’pΞ±β†’\vec{\alpha}\subseteq\vec{\delta}_{p_{\vec{\alpha}}}.

  2. (2)

    There is an Ξ΅Ξ±β†’βˆˆ2\varepsilon_{\vec{\alpha}}\in 2 such that pΞ±β†’βŠ©p_{\vec{\alpha}}\Vdash β€œh′​(bΛ™Ξ±β†’β†Ύβ„“)=Ρα→h^{\prime}(\dot{b}_{\vec{\alpha}}\restriction\ell)=\varepsilon_{\vec{\alpha}} for 𝒰˙\dot{\mathcal{U}} many β„“\ell in LΛ™G\dot{L}_{G}”.

  3. (3)

    h′​(X​(pΞ±β†’,Ξ±β†’))=Ρα→h^{\prime}(X(p_{\vec{\alpha}},\vec{\alpha}))=\varepsilon_{\vec{\alpha}}.

Such conditions can be found as follows: Fix some X∈Ext⁑(B)X\in\operatorname{Ext}(B) and let xix_{i} denote the node in XX extending sis_{i}, for each i≀𝐒i\leq\mathbf{i}. For Ξ±β†’βˆˆ[ΞΊ]𝐒\vec{\alpha}\in[\kappa]^{\mathbf{i}}, define

pΞ±β†’0={⟨(i,Ξ΄),xi⟩:i<𝐒,Ξ΄βˆˆΞ±β†’}βˆͺ{⟨𝐒,x𝐒⟩}.p^{0}_{\vec{\alpha}}=\{\langle(i,\delta),x_{i}\rangle:i<\mathbf{i},\ \delta\in\vec{\alpha}\}\cup\{\langle\mathbf{i},x_{\mathbf{i}}\rangle\}.

Then (1) will hold for all p≀pΞ±β†’0p\leq p^{0}_{\vec{\alpha}}, since Ξ΄β†’pΞ±β†’0=Ξ±β†’\vec{\delta}_{p_{\vec{\alpha}}^{0}}=\vec{\alpha}. Next, let pΞ±β†’1p^{1}_{\vec{\alpha}} be a condition below pΞ±β†’0p^{0}_{\vec{\alpha}} which forces h′​(bΛ™Ξ±β†’β†Ύβ„“)h^{\prime}(\dot{b}_{\vec{\alpha}}\restriction\ell) to be the same value for 𝒰˙\dot{\mathcal{U}} many β„“βˆˆLΛ™G\ell\in\dot{L}_{G}. Extend this to some condition pΞ±β†’2≀pΞ±β†’1p^{2}_{\vec{\alpha}}\leq p_{\vec{\alpha}}^{1} which decides a value Ξ΅Ξ±β†’βˆˆ2\varepsilon_{\vec{\alpha}}\in 2 so that pΞ±β†’2p^{2}_{\vec{\alpha}} forces h′​(bΛ™Ξ±β†’β†Ύβ„“)=Ρα→h^{\prime}(\dot{b}_{\vec{\alpha}}\restriction\ell)=\varepsilon_{\vec{\alpha}} for 𝒰˙\dot{\mathcal{U}} many β„“\ell in LΛ™G\dot{L}_{G}. Then (2) holds for all p≀pΞ±β†’2p\leq p_{\vec{\alpha}}^{2}. If pΞ±β†’2p_{\vec{\alpha}}^{2} satisfies (3), then let pΞ±β†’=pΞ±β†’2p_{\vec{\alpha}}=p_{\vec{\alpha}}^{2}. Otherwise, take some pΞ±β†’3≀pΞ±β†’2p^{3}_{\vec{\alpha}}\leq p^{2}_{\vec{\alpha}} which forces bΛ™Ξ±β†’β†Ύβ„“βˆˆExt⁑(B)\dot{b}_{\vec{\alpha}}\restriction\ell\in\operatorname{Ext}(B) and h′​(bΛ™Ξ±β†’β†Ύβ„“)=Ρα→h^{\prime}(\dot{b}_{\vec{\alpha}}\restriction\ell)=\varepsilon_{\vec{\alpha}} for some β„“βˆˆLΛ™G\ell\in\dot{L}_{G} with β„“pΞ±β†’2<ℓ≀ℓpΞ±β†’3\ell_{p^{2}_{\vec{\alpha}}}<\ell\leq\ell_{p^{3}_{\vec{\alpha}}}; then h′​(X​(pΞ±β†’3β†Ύβ„“,Ξ±β†’))=Ρα→h^{\prime}(X(p^{3}_{\vec{\alpha}}\restriction\ell,\vec{\alpha}))=\varepsilon_{\vec{\alpha}}. Thus, letting pΞ±β†’p_{\vec{\alpha}} be pΞ±β†’3β†Ύβ„“p^{3}_{\vec{\alpha}}\restriction\ell, we see that pΞ±β†’p_{\vec{\alpha}} satisfies (1)–(3).

Let ℐ\mathcal{I} denote the collection of all functions ΞΉ:2​𝐒→2​𝐒\iota:2\mathbf{i}\rightarrow 2\mathbf{i} such that for each i<𝐒i<\mathbf{i}, {ι​(2​i),ι​(2​i+1)}βŠ†{2​i,2​i+1}\{\iota(2i),\iota(2i+1)\}\subseteq\{2i,2i+1\}. For ΞΈβ†’=⟨θ0,…,ΞΈ2β€‹π’βˆ’1⟩∈[ΞΊ]2​𝐒\vec{\theta}=\langle\theta_{0},\dots,\theta_{2\mathbf{i}-1}\rangle\in[\kappa]^{2\mathbf{i}}, ι​(ΞΈβ†’)\iota(\vec{\theta}\,) determines the pair of sequences of ordinals ⟨ιe​(ΞΈβ†’),ΞΉo​(ΞΈβ†’)⟩\langle\iota_{e}(\vec{\theta}\,),\iota_{o}(\vec{\theta}\,)\rangle, where

(30) ΞΉe​(ΞΈβ†’)\displaystyle\iota_{e}(\vec{\theta}\,) =βŸ¨ΞΈΞΉβ€‹(0),θι​(2),…,ΞΈΞΉ(2π’βˆ’2))⟩\displaystyle=\langle\theta_{\iota(0)},\theta_{\iota(2)},\dots,\theta_{\iota(2\mathbf{i}-2))}\rangle
(31) ΞΉo​(ΞΈβ†’)\displaystyle\iota_{o}(\vec{\theta}\,) =βŸ¨ΞΈΞΉβ€‹(1),θι​(3),…,θι​(2β€‹π’βˆ’1)⟩.\displaystyle=\langle\theta_{\iota(1)},\theta_{\iota(3)},\dots,\theta_{\iota(2\mathbf{i}-1)}\rangle.

We now proceed to define a coloring ff on [ΞΊ]2​𝐒[\kappa]^{2\mathbf{i}} into countably many colors. Let Ξ΄β†’Ξ±β†’\vec{\delta}_{\vec{\alpha}} denote Ξ΄β†’pΞ±β†’\vec{\delta}_{p_{\vec{\alpha}}}, kΞ±β†’k_{\vec{\alpha}} denote |Ξ΄β†’Ξ±β†’||\vec{\delta}_{\vec{\alpha}}|, β„“Ξ±β†’\ell_{\vec{\alpha}} denote β„“pΞ±β†’\ell_{p_{\vec{\alpha}}}, and let βŸ¨Ξ΄Ξ±β†’(j):j<kΞ±β†’βŸ©\langle\delta_{\vec{\alpha}}(j):j<k_{\vec{\alpha}}\rangle denote the enumeration of Ξ΄β†’Ξ±β†’\vec{\delta}_{\vec{\alpha}} in increasing order. Given ΞΈβ†’βˆˆ[ΞΊ]2​𝐒\vec{\theta}\in[\kappa]^{2\mathbf{i}} and ΞΉβˆˆβ„\iota\in\mathcal{I}, to reduce subscripts let Ξ±β†’\vec{\alpha} denote ΞΉe​(ΞΈβ†’)\iota_{e}(\vec{\theta}\,) and Ξ²β†’\vec{\beta} denote ΞΉo​(ΞΈβ†’)\iota_{o}(\vec{\theta}\,), and define

(32) f​(ΞΉ,ΞΈβ†’)=\displaystyle f(\iota,\vec{\theta}\,)=\, ⟨ι,Ρα→,kΞ±β†’,pΞ±β†’(𝐒),⟨⟨pΞ±β†’(i,δα→(j)):j<kΞ±β†’βŸ©:i<𝐒⟩,\displaystyle\langle\iota,\varepsilon_{\vec{\alpha}},k_{\vec{\alpha}},p_{\vec{\alpha}}(\mathbf{i}),\langle\langle p_{\vec{\alpha}}(i,\delta_{\vec{\alpha}}(j)):j<k_{\vec{\alpha}}\rangle:i<\mathbf{i}\rangle,
(33) ⟨⟨i,j⟩:i<𝐒,j<kΞ±β†’,and​δα→​(j)=Ξ±i⟩,\displaystyle\langle\langle i,j\rangle:i<\mathbf{i},\ j<k_{\vec{\alpha}},\ \mathrm{and\ }\delta_{\vec{\alpha}}(j)=\alpha_{i}\rangle,
(34) ⟨⟨j,k⟩:j<kΞ±β†’,k<kΞ²β†’,δα→(j)=δβ→(k)⟩⟩.\displaystyle\langle\langle j,k\rangle:j<k_{\vec{\alpha}},\ k<k_{\vec{\beta}},\ \delta_{\vec{\alpha}}(j)=\delta_{\vec{\beta}}(k)\rangle\rangle.

Fix some ordering of ℐ\mathcal{I} and define

(35) f(ΞΈβ†’)=⟨f(ΞΉ,ΞΈβ†’):ΞΉβˆˆβ„βŸ©.f(\vec{\theta}\,)=\langle f(\iota,\vec{\theta}\,):\iota\in\mathcal{I}\rangle.

By the ErdΕ‘s-Rado Theorem 4.4, there is a subset KβŠ†ΞΊK\subseteq\kappa of cardinality β„΅1\aleph_{1} which is homogeneous for ff. Take Kβ€²βŠ†KK^{\prime}\subseteq K so that between each two members of Kβ€²K^{\prime} there is a member of KK. Then take KiβŠ†Kβ€²K_{i}\subseteq K^{\prime} satisfying K0<β‹―<Kπ’βˆ’1K_{0}<\dots<K_{\mathbf{i}-1}, where Ki<Ki+1K_{i}<K_{i+1} means that each ordinal in KiK_{i} is less than each ordinal in Ki+1K_{i+1}. Let Kβ†’\vec{K} denote ∏i<𝐒Ki\prod_{i<\mathbf{i}}K_{i}.

Fix some Ξ³β†’βˆˆKβ†’\vec{\gamma}\in\vec{K}, and define

(36) Ξ΅βˆ—=Ργ→,kβˆ—=kΞ³β†’,t𝐒=pγ→​(𝐒),\displaystyle\varepsilon_{*}=\varepsilon_{\vec{\gamma}},\ \ k_{*}=k_{\vec{\gamma}},\ \ t_{\mathbf{i}}=p_{\vec{\gamma}}(\mathbf{i}),
(37) ti,j\displaystyle t_{i,j} =pγ→​(i,δγ→​(j))​for​i<𝐒,j<kβˆ—.\displaystyle=p_{\vec{\gamma}}(i,\delta_{\vec{\gamma}}(j))\mathrm{\ for\ }i<\mathbf{i},\ j<k_{*}.

The next three lemmas show that the values in equation (36) are the same for any choice of γ→\vec{\gamma} in K→\vec{K}.

Lemma 4.6.

For all Ξ±β†’βˆˆKβ†’\vec{\alpha}\in\vec{K}, Ρα→=Ξ΅βˆ—\varepsilon_{\vec{\alpha}}=\varepsilon_{*}, kΞ±β†’=kβˆ—k_{\vec{\alpha}}=k_{*}, pα→​(𝐒)=t𝐒p_{\vec{\alpha}}(\mathbf{i})=t_{\mathbf{i}}, and ⟨pΞ±β†’(i,δα→(j)):j<kΞ±β†’βŸ©=⟨ti,j:j<kβˆ—βŸ©\langle p_{\vec{\alpha}}(i,\delta_{\vec{\alpha}}(j)):j<k_{\vec{\alpha}}\rangle=\langle t_{i,j}:j<k_{*}\rangle for each i<𝐒i<\mathbf{i}.

Proof.

Let Ξ±β†’\vec{\alpha} be any member of Kβ†’\vec{K}, and let Ξ³β†’\vec{\gamma} be the set of ordinals fixed above. Take ΞΉβˆˆβ„\iota\in\mathcal{I} to be the identity function on 2​𝐒2\mathbf{i}. Then there are ΞΈβ†’,ΞΈβ†’β€²βˆˆ[K]2​𝐒\vec{\theta},\vec{\theta}^{\prime}\in[K]^{2\mathbf{i}} such that Ξ±β†’=ΞΉe​(ΞΈβ†’)\vec{\alpha}=\iota_{e}(\vec{\theta}\,) and Ξ³β†’=ΞΉe​(ΞΈβ†’β€²)\vec{\gamma}=\iota_{e}(\vec{\theta}^{\prime}\,). Since f​(ΞΉ,ΞΈβ†’)=f​(ΞΉ,ΞΈβ†’β€²)f(\iota,\vec{\theta}\,)=f(\iota,\vec{\theta}^{\prime}\,), it follows that Ρα→=Ργ→\varepsilon_{\vec{\alpha}}=\varepsilon_{\vec{\gamma}}, kΞ±β†’=kΞ³β†’k_{\vec{\alpha}}=k_{\vec{\gamma}}, pα→​(𝐒)=pγ→​(𝐒)p_{\vec{\alpha}}(\mathbf{i})=p_{\vec{\gamma}}(\mathbf{i}), and ⟨⟨pΞ±β†’(i,δα→(j)):j<kΞ±β†’βŸ©:i<𝐒⟩=⟨⟨pΞ³β†’(i,δγ→(j)):j<kΞ³β†’βŸ©:i<𝐒⟩\langle\langle p_{\vec{\alpha}}(i,\delta_{\vec{\alpha}}(j)):j<k_{\vec{\alpha}}\rangle:i<\mathbf{i}\rangle=\langle\langle p_{\vec{\gamma}}(i,\delta_{\vec{\gamma}}(j)):j<k_{\vec{\gamma}}\rangle:i<\mathbf{i}\rangle. ∎

Let β„“βˆ—\ell_{*} denote the length of the nodes ti,jt_{i,j}, (i,j)∈dΓ—kβˆ—(i,j)\in d\times k_{*}. In Case (a), |t𝐒||t_{\mathbf{i}}| also equals β„“βˆ—\ell_{*}; in Case (b), let β„“βˆ—β€²\ell_{*}^{\prime} denote |t𝐒||t_{\mathbf{i}}|.

Lemma 4.7.

Given any Ξ±β†’,Ξ²β†’βˆˆKβ†’\vec{\alpha},\vec{\beta}\in\vec{K}, if j,k<kβˆ—j,k<k_{*} and δα→​(j)=δβ→​(k)\delta_{\vec{\alpha}}(j)=\delta_{\vec{\beta}}(k), then j=kj=k.

Proof.

Let Ξ±β†’,Ξ²β†’\vec{\alpha},\vec{\beta} be members of Kβ†’\vec{K} and suppose that δα→​(j)=δβ→​(k)\delta_{\vec{\alpha}}(j)=\delta_{\vec{\beta}}(k) for some j,k<kβˆ—j,k<k_{*}. For i<𝐒i<\mathbf{i}, let ρi\rho_{i} be the relation from among {<,=,>}\{<,=,>\} such that Ξ±i​ρi​βi\alpha_{i}\,\rho_{i}\,\beta_{i}. Let ΞΉ\iota be the member of ℐ\mathcal{I} such that for each ΞΈβ†’βˆˆ[K]2​𝐒\vec{\theta}\in[K]^{2\mathbf{i}} and each i<𝐒i<\mathbf{i}, θι​(2​i)​ρi​θι​(2​i+1)\theta_{\iota(2i)}\ \rho_{i}\ \theta_{\iota(2i+1)}. Fix some ΞΈβ†’βˆˆ[Kβ€²]2​𝐒\vec{\theta}\in[K^{\prime}]^{2\mathbf{i}} such that ΞΉe​(ΞΈβ†’)=Ξ±β†’\iota_{e}(\vec{\theta})=\vec{\alpha} and ΞΉo​(ΞΈβ†’)=Ξ²β†’\iota_{o}(\vec{\theta})=\vec{\beta}. Since between any two members of Kβ€²K^{\prime} there is a member of KK, there is a ΞΆβ†’βˆˆ[K]𝐒\vec{\zeta}\in[K]^{\mathbf{i}} such that for each i<𝐒i<\mathbf{i}, Ξ±i​ρi​΢i\alpha_{i}\,\rho_{i}\,\zeta_{i} and ΞΆi​ρi​βi\zeta_{i}\,\rho_{i}\,\beta_{i}. Let ΞΌβ†’,Ξ½β†’\vec{\mu},\vec{\nu} be members of [K]2​𝐒[K]^{2\mathbf{i}} such that ΞΉe​(ΞΌβ†’)=Ξ±β†’\iota_{e}(\vec{\mu})=\vec{\alpha}, ΞΉo​(ΞΌβ†’)=ΞΆβ†’\iota_{o}(\vec{\mu})=\vec{\zeta}, ΞΉe​(Ξ½β†’)=ΞΆβ†’\iota_{e}(\vec{\nu})=\vec{\zeta}, and ΞΉo​(Ξ½β†’)=Ξ²β†’\iota_{o}(\vec{\nu})=\vec{\beta}. Since δα→​(j)=δβ→​(k)\delta_{\vec{\alpha}}(j)=\delta_{\vec{\beta}}(k), the pair ⟨j,k⟩\langle j,k\rangle is in the last sequence in f​(ΞΉ,ΞΈβ†’)f(\iota,\vec{\theta}). Since f​(ΞΉ,ΞΌβ†’)=f​(ΞΉ,Ξ½β†’)=f​(ΞΉ,ΞΈβ†’)f(\iota,\vec{\mu})=f(\iota,\vec{\nu})=f(\iota,\vec{\theta}), also ⟨j,k⟩\langle j,k\rangle is in the last sequence in f​(ΞΉ,ΞΌβ†’)f(\iota,\vec{\mu}) and f​(ΞΉ,Ξ½β†’)f(\iota,\vec{\nu}). It follows that δα→​(j)=δ΢→​(k)\delta_{\vec{\alpha}}(j)=\delta_{\vec{\zeta}}(k) and δ΢→​(j)=δβ→​(k)\delta_{\vec{\zeta}}(j)=\delta_{\vec{\beta}}(k). Hence, δ΢→​(j)=δ΢→​(k)\delta_{\vec{\zeta}}(j)=\delta_{\vec{\zeta}}(k), and therefore jj must equal kk. ∎

For each Ξ±β†’βˆˆKβ†’\vec{\alpha}\in\vec{K}, given any ΞΉβˆˆβ„\iota\in\mathcal{I}, there is a ΞΈβ†’βˆˆ[K]2​𝐒\vec{\theta}\in[K]^{2\mathbf{i}} such that Ξ±β†’=ΞΉo​(Ξ±β†’)\vec{\alpha}=\iota_{o}(\vec{\alpha}). By the second line of equation (32), there is a strictly increasing sequence ⟨ji:i<𝐒⟩\langle j_{i}:i<\mathbf{i}\rangle of members of kβˆ—k_{*} such that δγ→​(ji)=Ξ±i\delta_{\vec{\gamma}}(j_{i})=\alpha_{i}. By homogeneity of ff, this sequence ⟨ji:i<𝐒⟩\langle j_{i}:i<\mathbf{i}\rangle is the same for all members of Kβ†’\vec{K}. Then letting tiβˆ—t^{*}_{i} denote ti,jit_{i,j_{i}}, one sees that

(38) pα→​(i,Ξ±i)=pα→​(i,δα→​(ji))=ti,ji=tiβˆ—.p_{\vec{\alpha}}(i,\alpha_{i})=p_{\vec{\alpha}}(i,\delta_{\vec{\alpha}}(j_{i}))=t_{i,j_{i}}=t^{*}_{i}.

Let tπ’βˆ—t_{\mathbf{i}}^{*} denote t𝐒t_{\mathbf{i}}.

Lemma 4.8 (Homogeneity of {pΞ±β†’:Ξ±β†’βˆˆKβ†’}\{p_{\vec{\alpha}}:\vec{\alpha}\in\vec{K}\,\}).

For any finite subset Jβ†’βŠ†Kβ†’\vec{J}\subseteq\vec{K}, pJβ†’:=⋃{pΞ±β†’:Ξ±β†’βˆˆJβ†’}p_{\vec{J}}:=\bigcup\{p_{\vec{\alpha}}:\vec{\alpha}\in\vec{J}\,\} is a member of β„™\mathbb{P} which is below each pΞ±β†’p_{\vec{\alpha}}, Ξ±β†’βˆˆJβ†’\vec{\alpha}\in\vec{J}.

Proof.

Given Ξ±β†’,Ξ²β†’βˆˆJβ†’\vec{\alpha},\vec{\beta}\in\vec{J}, if j,k<kβˆ—j,k<k_{*} and δα→​(j)=δβ→​(k)\delta_{\vec{\alpha}}(j)=\delta_{\vec{\beta}}(k), then jj and kk must be equal, by Lemma 4.7. Then Lemma 4.6 implies that for each i<𝐒i<\mathbf{i},

(39) pα→​(i,δα→​(j))=ti,j=pβ→​(i,δβ→​(j))=pβ→​(i,δβ→​(k)).p_{\vec{\alpha}}(i,\delta_{\vec{\alpha}}(j))=t_{i,j}=p_{\vec{\beta}}(i,\delta_{\vec{\beta}}(j))=p_{\vec{\beta}}(i,\delta_{\vec{\beta}}(k)).

Hence, for all Ξ΄βˆˆΞ΄β†’Ξ±β†’βˆ©Ξ΄β†’Ξ²β†’\delta\in\vec{\delta}_{\vec{\alpha}}\cap\vec{\delta}_{\vec{\beta}} and i<𝐒i<\mathbf{i}, pα→​(i,Ξ΄)=pβ→​(i,Ξ΄)p_{\vec{\alpha}}(i,\delta)=p_{\vec{\beta}}(i,\delta). Thus, pJβ†’:=⋃{pΞ±β†’:Ξ±β†’βˆˆJβ†’}p_{\vec{J}}:=\bigcup\{p_{\vec{\alpha}}:\vec{\alpha}\in\vec{J}\} is a function with Ξ΄β†’pJβ†’=⋃{Ξ΄β†’Ξ±β†’:Ξ±β†’βˆˆJβ†’}\vec{\delta}_{p_{\vec{J}}}=\bigcup\{\vec{\delta}_{\vec{\alpha}}:\vec{\alpha}\in\vec{J}\,\}; hence, pJβ†’p_{\vec{J}} is a member of β„™\mathbb{P}. Since for each Ξ±β†’βˆˆJβ†’\vec{\alpha}\in\vec{J}, ran⁑(pJβ†’β†ΎΞ΄β†’Ξ±β†’)=ran⁑(pΞ±β†’)\operatorname{ran}(p_{\vec{J}}\restriction\vec{\delta}_{\vec{\alpha}})=\operatorname{ran}(p_{\vec{\alpha}}), it follows that pJ→≀pΞ±β†’p_{\vec{J}}\leq p_{\vec{\alpha}} for each Ξ±β†’βˆˆJβ†’\vec{\alpha}\in\vec{J}. ∎

We now proceed to build an N∈[D,𝔻]N\in[D,\mathbb{D}] so that the coloring hβ€²h^{\prime} will be monochromatic on ExtN⁑(B)\operatorname{Ext}_{N}(B), from which it will follow that hh is monochromatic in rk+1​[B,N]βˆ—r_{k+1}[B,N]^{*}. Set

(40) Uβˆ—={tiβˆ—:i≀𝐒}βˆͺ{uβˆ—:u∈max(D)+βˆ–B},U^{*}=\{t^{*}_{i}:i\leq\mathbf{i}\}\cup\{u^{*}:u\in\max(D)^{+}\setminus B\},

where for each uu in max(D)+βˆ–B\max(D)^{+}\setminus B, uβˆ—u^{*} is some extension of uβˆ—u^{*} in π”»β†Ύβ„“βˆ—\mathbb{D}\restriction\ell_{*}. Then Uβˆ—U^{*} end-extends max(D)+\max(D)^{+}. One may take each uβˆ—u^{*} to be the β‰Ί\prec-leftmost extension of uu to be deterministic, but SDAPΒ implies that any extensions will suffice. (Since tπ’βˆ—t^{*}_{\mathbf{i}} is either a splitting node or a splitting predecessor, all possible choices of uβˆ—u^{*} for u∈max(D)+βˆ–Bu\in\max(D)^{+}\setminus B are are automatically never splitting nodes nor splitting predecessors nor coding nodes in 𝔻\mathbb{D}.)

Let {nj:jβˆˆβ„•}\{n_{j}:j\in\mathbb{N}\} be the strictly increasing enumeration of II, and note that n0>dn_{0}>d. From now on, Cases (a) and (b) are different enough to warrant separate treatment.

Case (a). If n0=d+1n_{0}=d+1, then DβˆͺUβˆ—D\cup U^{*} is a member of rn0​[D,𝔻]r_{n_{0}}[D,\mathbb{D}]. In this case, we let Un0=DβˆͺUβˆ—U_{n_{0}}=D\cup U^{*}, and let Un1βˆ’1U_{n_{1}-1} be any member of rn1βˆ’1​[Un0,𝔻]r_{n_{1}-1}[U_{n_{0}},\mathbb{D}], noting that Uβˆ—U^{*} is the only member of ExtUn0⁑(B)\operatorname{Ext}_{U_{n_{0}}}(B) and that h′​(Uβˆ—)=Ξ΅βˆ—h^{\prime}(U^{*})=\varepsilon_{*}. Otherwise, n0>d+1n_{0}>d+1. In this case, take some Un0βˆ’1∈rn0βˆ’1​[D,𝔻]U_{n_{0}-1}\in r_{n_{0}-1}[D,\mathbb{D}] such that max⁑(Un0βˆ’1)\max(U_{n_{0}-1}) end-extends Uβˆ—U^{*}, and notice that ExtUn0βˆ’1⁑(B)\operatorname{Ext}_{U_{n_{0}-1}}(B) is empty.

Now assume that jβ‰₯0j\geq 0 and we have constructed Unjβˆ’1∈rnjβˆ’1​[D,𝔻]U_{n_{j}-1}\in r_{n_{j}-1}[D,\mathbb{D}] so that every member of ExtUnjβˆ’1⁑(B)\operatorname{Ext}_{U_{n_{j}-1}}(B) has hβ€²h^{\prime}-color Ξ΅βˆ—\varepsilon_{*}. Fix some E∈rnj​[Unjβˆ’1,𝔻]E\in r_{n_{j}}[U_{n_{j}-1},\mathbb{D}] and let Y=max⁑(E)Y=\max(E). We will extend the nodes in YY to construct Unj∈rnj​[Unjβˆ’1,𝔻]U_{n_{j}}\in r_{n_{j}}[U_{n_{j}-1},\mathbb{D}] with the property that all members of ExtUnj⁑(B)\operatorname{Ext}_{U_{n_{j}}}(B) have the same hh-value Ξ΅βˆ—\varepsilon_{*}. This will be achieved by constructing the condition qβˆˆβ„™q\in\mathbb{P}, below, and then extending it to some condition r≀qr\leq q which decides that all members of Ext⁑(B)\operatorname{Ext}(B) coming from the nodes in ran⁑(r)\operatorname{ran}(r) have hh-color Ξ΅βˆ—\varepsilon_{*}.

Let q​(𝐒)q(\mathbf{i}) denote the splitting node in YY and let β„“q=β„“Y\ell_{q}=\ell_{Y}. For each i<𝐒i<\mathbf{i}, let YiY_{i} denote Y∩TiY\cap T_{i}, and let JiβŠ†KiJ_{i}\subseteq K_{i} be a set of the same cardinality as YiY_{i} and label the members of YiY_{i} as {zΞ±:α∈Ji}\{z_{\alpha}:\alpha\in J_{i}\}. Let Jβ†’\vec{J} denote ∏i<𝐒Ji\prod_{i<\mathbf{i}}J_{i}, and note that for each Ξ±β†’βˆˆJβ†’\vec{\alpha}\in\vec{J}, the set {zΞ±i:i<𝐒}βˆͺ{q​(𝐒)}\{z_{\alpha_{i}}:i<\mathbf{i}\}\cup\{q(\mathbf{i})\} is a member of Ext⁑(B)\operatorname{Ext}(B). By Lemma 4.8, the set {pΞ±β†’:Ξ±β†’βˆˆJβ†’}\{p_{\vec{\alpha}}:\vec{\alpha}\in\vec{J}\} is compatible, and pJβ†’:=⋃{pΞ±β†’:Ξ±β†’βˆˆJβ†’}p_{\vec{J}}:=\bigcup\{p_{\vec{\alpha}}:\vec{\alpha}\in\vec{J}\} is a condition in β„™\mathbb{P}.

Let Ξ΄β†’q=⋃{Ξ΄β†’Ξ±β†’:Ξ±β†’βˆˆJβ†’}\vec{\delta}_{q}=\bigcup\{\vec{\delta}_{\vec{\alpha}}:\vec{\alpha}\in\vec{J}\}. For i<𝐒i<\mathbf{i} and α∈Ji\alpha\in J_{i}, define q​(i,Ξ±)=zΞ±q(i,\alpha)=z_{\alpha}. It follows that for each Ξ±β†’βˆˆJβ†’\vec{\alpha}\in\vec{J} and i<𝐒i<\mathbf{i},

(41) q​(i,Ξ±i)βŠ‡tiβˆ—=pα→​(i,Ξ±i)=pJ→​(i,Ξ±i),q(i,\alpha_{i})\supseteq t^{*}_{i}=p_{\vec{\alpha}}(i,\alpha_{i})=p_{\vec{J}}(i,\alpha_{i}),

and

(42) q​(𝐒)βŠ‡tπ’βˆ—=pα→​(𝐒)=pJ→​(𝐒).q(\mathbf{i})\supseteq t^{*}_{\mathbf{i}}=p_{\vec{\alpha}}(\mathbf{i})=p_{\vec{J}}(\mathbf{i}).

For i<𝐒i<\mathbf{i} and Ξ΄βˆˆΞ΄β†’qβˆ–Ji\delta\in\vec{\delta}_{q}\setminus J_{i}, let q​(i,Ξ΄)q(i,\delta) be any node in 𝔻↾ℓq\mathbb{D}\restriction\ell_{q} extending pJ→​(i,Ξ΄)p_{\vec{J}}(i,\delta). Define

(43) q={q​(𝐒)}βˆͺ{⟨(i,Ξ΄),q​(i,Ξ΄)⟩:i<𝐒,Ξ΄βˆˆΞ΄β†’q}.q=\{q(\mathbf{i})\}\cup\{\langle(i,\delta),q(i,\delta)\rangle:i<\mathbf{i},\ \delta\in\vec{\delta}_{q}\}.

This qq is a condition in β„™\mathbb{P}, and q≀pJβ†’q\leq p_{\vec{J}}.

Take an r≀qr\leq q in β„™\mathbb{P} which decides some β„“\ell in LΛ™G\dot{L}_{G} for which h′​(bΛ™Ξ±β†’β†Ύβ„“)=Ξ΅βˆ—h^{\prime}(\dot{b}_{\vec{\alpha}}\restriction\ell)=\varepsilon_{*}, for all Ξ±β†’βˆˆJβ†’\vec{\alpha}\in\vec{J}. Without loss of generality, we may assume that β„“r=β„“\ell_{r}=\ell. Since rr forces bΛ™Ξ±β†’β†Ύβ„“=X​(r,Ξ±β†’)\dot{b}_{\vec{\alpha}}\restriction\ell=X(r,\vec{\alpha}) for each Ξ±β†’βˆˆJβ†’\vec{\alpha}\in\vec{J}, and since the coloring hβ€²h^{\prime} is defined in the ground model, it follows that h′​(X​(r,Ξ±β†’))=Ξ΅βˆ—h^{\prime}(X(r,\vec{\alpha}))=\varepsilon_{*} for each Ξ±β†’βˆˆJβ†’\vec{\alpha}\in\vec{J}. Let

(44) Yβ€²={q​(𝐒)}βˆͺ{q​(i,Ξ±):i<𝐒,α∈Ji},Y^{\prime}=\{q(\mathbf{i})\}\cup\{q(i,\alpha):i<\mathbf{i},\ \alpha\in J_{i}\},

and let

(45) Zβ€²={r​(𝐒)}βˆͺ{r​(i,Ξ±):i<𝐒,α∈Ji}.Z^{\prime}=\{r(\mathbf{i})\}\cup\{r(i,\alpha):i<\mathbf{i},\ \alpha\in J_{i}\}.

Let ZZ be the level set consisting of the nodes in Zβ€²Z^{\prime} along with a node zyz_{y} in 𝔻↾ℓ\mathbb{D}\restriction\ell extending yy, for each y∈Yβˆ–Yβ€²y\in Y\setminus Y^{\prime}. Then ZZ end-extends YY, and moreover, letting Unj=Unjβˆ’1βˆͺZU_{n_{j}}=U_{n_{j}-1}\cup Z, we see that UnjU_{n_{j}} is a member of rnj​[Unjβˆ’1,𝔻]r_{n_{j}}[U_{n_{j}-1},\mathbb{D}] such that hβ€²h^{\prime} has value Ξ΅βˆ—\varepsilon_{*} on ExtUnj⁑(B)\operatorname{Ext}_{U_{n_{j}}}(B).

Case (b). Notice that in this case, n0n_{0} must be at least d+2d+2 and that tπ’βˆ—t^{*}_{\mathbf{i}} is the splitting predecessor of the coding node in π”»β†Ύβ„“βˆ—\mathbb{D}\restriction\ell_{*}, which we shall denote by cβˆ—c^{*}. Let Uβˆ—U^{*} be as in equation (40). In Case (b), all nodes in Uβˆ—U^{*} have length β„“βˆ—\ell_{*} except for tπ’βˆ—t^{*}_{\mathbf{i}}, which has length β„“βˆ—β€²\ell^{\prime}_{*}. There is exactly one non-terminal (i.e.Β non-coding) node in π”»β†Ύβ„“βˆ—\mathbb{D}\restriction\ell_{*} extending tπ’βˆ—t^{*}_{\mathbf{i}}; denote this node by uπ’βˆ—u^{*}_{\mathbf{i}}. If n0=d+2n_{0}=d+2, let Un0U_{n_{0}} be the tree induced by DβˆͺUβˆ—βˆͺ{uπ’βˆ—,cβˆ—}D\cup U^{*}\cup\{u_{\mathbf{i}}^{*},c^{*}\}. Then let Un1βˆ’2U_{n_{1}-2} be any member of rn1βˆ’2​[Un0,𝔻]r_{n_{1}-2}[U_{n_{0}},\mathbb{D}].

If n0>d+2n_{0}>d+2, the same argument will handle the base case and the induction step. For the base case, let EE be a member of rn0​[D,𝔻]r_{n_{0}}[D,\mathbb{D}] such that Eβ†Ύβ„“βˆ—E\restriction\ell_{*} equals (Uβˆ—βˆ–{tπ’βˆ—})βˆͺ{uπ’βˆ—}(U^{*}\setminus\{t^{*}_{\mathbf{i}}\})\cup\{u^{*}_{\mathbf{i}}\}. (In particular, β„“βˆ—<β„“rd+1​(Un0)\ell_{*}<\ell_{r_{d+1}(U_{n_{0}})}.) For jβ‰₯1j\geq 1, supposing we have constructed Unjβˆ’2∈rnjβˆ’2​[Unjβˆ’1,𝔻]U_{n_{j}-2}\in r_{n_{j}-2}[U_{n_{j-1}},\mathbb{D}] so that every member of ExtUnjβˆ’2⁑(B)\operatorname{Ext}_{U_{n_{j}-2}}(B) has hβ€²h^{\prime}-color Ξ΅βˆ—\varepsilon_{*}, let EE be any member of rnj​[Unjβˆ’2,D]r_{n_{j}}[U_{n_{j}-2},D]. In each of these two cases, let cEc^{E} denote the coding node in max⁑(E)\max(E), and let YY denote the set max⁑(E)\max(E) but with the two extensions of sp⁑(cE)\operatorname{sp}(c^{E}) in max⁑(E)\max(E) deleted and replaced by sp⁑(cE)\operatorname{sp}(c^{E}).

Let β„“q=β„“E\ell_{q}=\ell_{E}, and let q​(𝐒)q(\mathbf{i}) denote sp⁑(cE)\operatorname{sp}(c^{E}). For each i<𝐒i<\mathbf{i}, let YiY_{i} denote the set of nodes y∈Y∩Tiy\in Y\cap T_{i} such that yy is a member of some X∈ExtE⁑(B)X\in\operatorname{Ext}_{E}(B). Equivalently, YiY_{i} is the set of those y∈Y∩Tiy\in Y\cap T_{i} such that y+​(cE)∼(tiβˆ—)+​(cβˆ—)y^{+}(c^{E})\sim(t^{*}_{i})^{+}(c^{*}). For each i<𝐒i<\mathbf{i}, take a set JiβŠ†KiJ_{i}\subseteq K_{i} of the same cardinality as YiY_{i} and label the members of YiY_{i} as {zΞ±:α∈Ji}\{z_{\alpha}:\alpha\in J_{i}\}. Let Jβ†’\vec{J} denote ∏i<𝐒Ji\prod_{i<\mathbf{i}}J_{i}, noting that for each Ξ±β†’βˆˆJβ†’\vec{\alpha}\in\vec{J}, {zΞ±i:i<𝐒}βˆͺ{q​(𝐒)}\{z_{\alpha_{i}}:i<\mathbf{i}\}\cup\{q(\mathbf{i})\} is a member of Ext⁑(B)\operatorname{Ext}(B). By Lemma 4.8, the set {pΞ±β†’:Ξ±β†’βˆˆJβ†’}\{p_{\vec{\alpha}}:\vec{\alpha}\in\vec{J}\} is compatible, and pJβ†’:=⋃{pΞ±β†’:Ξ±β†’βˆˆJβ†’}p_{\vec{J}}:=\bigcup\{p_{\vec{\alpha}}:\vec{\alpha}\in\vec{J}\} is a condition in β„™\mathbb{P}.

Let Ξ΄β†’q=⋃{Ξ΄β†’Ξ±β†’:Ξ±β†’βˆˆJβ†’}\vec{\delta}_{q}=\bigcup\{\vec{\delta}_{\vec{\alpha}}:\vec{\alpha}\in\vec{J}\}. For i<𝐒i<\mathbf{i} and α∈Ji\alpha\in J_{i}, define q​(i,Ξ±)=zΞ±q(i,\alpha)=z_{\alpha}. It follows that for each Ξ±β†’βˆˆJβ†’\vec{\alpha}\in\vec{J} and i<𝐒i<\mathbf{i},

(46) q​(i,Ξ±i)βŠ‡tiβˆ—=pα→​(i,Ξ±i)=pJ→​(i,Ξ±i),q(i,\alpha_{i})\supseteq t^{*}_{i}=p_{\vec{\alpha}}(i,\alpha_{i})=p_{\vec{J}}(i,\alpha_{i}),

and

(47) q​(𝐒)βŠ‡tπ’βˆ—=pα→​(𝐒)=pJ→​(𝐒).q(\mathbf{i})\supseteq t^{*}_{\mathbf{i}}=p_{\vec{\alpha}}(\mathbf{i})=p_{\vec{J}}(\mathbf{i}).

For i<𝐒i<\mathbf{i} and Ξ΄βˆˆΞ΄β†’qβˆ–Ji\delta\in\vec{\delta}_{q}\setminus J_{i}, let q​(i,Ξ΄)q(i,\delta) be an extension of pJ→​(i,Ξ΄)p_{\vec{J}}(i,\delta) in TiT_{i} of length β„“q\ell_{q} satisfying

(48) q​(i,Ξ΄)+​(cq)∼pJ→​(i,Ξ΄)+​(cβˆ—),q(i,\delta)^{+}(c_{q})\sim p_{\vec{J}}(i,\delta)^{+}(c^{*}),

where cqc_{q} denotes the coding node in 𝔻↾ℓq\mathbb{D}\restriction\ell_{q}. Such nodes q​(i,Ξ΄)q(i,\delta) exist by SDAP. Define

(49) q={q​(𝐒)}βˆͺ{⟨(i,Ξ΄),q​(i,Ξ΄)⟩:i<𝐒,Ξ΄βˆˆΞ΄β†’q}.q=\{q(\mathbf{i})\}\cup\{\langle(i,\delta),q(i,\delta)\rangle:i<\mathbf{i},\ \delta\in\vec{\delta}_{q}\}.

This qq is a condition in β„™\mathbb{P}, and q≀pJβ†’q\leq p_{\vec{J}}.

Now take an r≀qr\leq q in β„™\mathbb{P} which decides some β„“\ell in LΛ™G\dot{L}_{G} for which h′​(bΛ™Ξ±β†’β†Ύβ„“)=Ξ΅βˆ—h^{\prime}(\dot{b}_{\vec{\alpha}}\restriction\ell)=\varepsilon_{*}, for all Ξ±β†’βˆˆJβ†’\vec{\alpha}\in\vec{J}. Without loss of generality, we may assume that β„“r=β„“\ell_{r}=\ell. Since rr forces bΛ™Ξ±β†’β†Ύβ„“=X​(r,Ξ±β†’)\dot{b}_{\vec{\alpha}}\restriction\ell=X(r,\vec{\alpha}) for each Ξ±β†’βˆˆJβ†’\vec{\alpha}\in\vec{J}, and since the coloring hβ€²h^{\prime} is defined in the ground model, it follows that h′​(X​(r,Ξ±β†’))=Ξ΅βˆ—h^{\prime}(X(r,\vec{\alpha}))=\varepsilon_{*} for each Ξ±β†’βˆˆJβ†’\vec{\alpha}\in\vec{J}. Let

(50) Z0={r​(𝐒)}βˆͺ{r​(i,Ξ±):i<𝐒,α∈Ji}.Z_{0}=\{r(\mathbf{i})\}\cup\{r(i,\alpha):i<\mathbf{i},\ \alpha\in J_{i}\}.

Recall that ran⁑(q)βŠ†Y\operatorname{ran}(q)\subseteq Y, and note that Z0Z_{0} end-extends ran⁑(q)\operatorname{ran}(q).

Let crc_{r} denote the coding node in 𝔻\mathbb{D} of length β„“r\ell_{r}. For each y∈Yβˆ–ran⁑(q)y\in Y\setminus\operatorname{ran}(q), choose a member zyβŠƒyz_{y}\supset y in 𝔻↾ℓr\mathbb{D}\restriction\ell_{r} so that

(51) zy+​(cr)∼y+​(cq).z_{y}^{+}(c_{r})\sim y^{+}(c_{q}).

Again, SDAPΒ ensures the existence of such zyz_{y}. Let ZZ be the level set consisting of the nodes zyz_{y} for y∈Yβˆ–ran⁑(q)y\in Y\setminus\operatorname{ran}(q), the nodes in Z0βˆ–{r​(𝐒)}Z_{0}\setminus\{r(\mathbf{i})\}, and the two nodes in 𝔻↾ℓr\mathbb{D}\restriction\ell_{r} extending r​(𝐒)r(\mathbf{i}). Let

(52) Unj=Unjβˆ’2βˆͺZβˆͺ(Zβ†Ύβ„“rβ€²).U_{n_{j}}=U_{n_{j}-2}\cup Z\cup(Z\restriction\ell^{\prime}_{r}).

Then UnjU_{n_{j}} is a member of rnj​[Unjβˆ’2,𝔻]r_{n_{j}}[U_{n_{j}-2},\mathbb{D}].

Now that we have constructed UnjU_{n_{j}}, let Unj+1βˆ’2U_{n_{j+1}-2} be any member of rnj+1βˆ’2​[Unj,𝔻]r_{n_{j+1}-2}[U_{n_{j}},\mathbb{D}]. This completes the inductive construction. Let N=⋃j<Ο‰UnjN=\bigcup_{j<\omega}U_{n_{j}}. Then NN is a member of [D,𝔻]βˆ—[D,\mathbb{D}]^{*} and for each X∈ExtN⁑(B)X\in\operatorname{Ext}_{N}(B), h′​(X)=Ξ΅βˆ—h^{\prime}(X)=\varepsilon_{*}. Thus, NN satisfies the theorem. ∎

Remark 4.9.

A simple modification of the proof yields the same theorem for structures with LSDAP+. (See proof of Theorem 5.4 in [2].)

5. Borel sets of π’Ÿβ€‹(𝔻)\mathcal{D}(\mathbb{D}) are completely Ramsey

The main result of this section is Theorem 5.17: For any enumerated FraΓ―ssé  structure 𝐊\mathbf{K} satisfying SDAP+, for each good diagonal coding antichain 𝔻\mathbb{D} representing 𝐊\mathbf{K}, the space π’Ÿβ€‹(𝔻)\mathcal{D}(\mathbb{D}) of all diagonal antichains MβŠ†π”»M\subseteq\mathbb{D} similar to 𝔻\mathbb{D} has the property that all Borel subsets are Ramsey. The proof generally follows the outline of the Galvin–Prikry Theorem in [8] with the following exceptions: The proof of the Nash-Williams-style Theorem 5.5 uses an asymmetric version of combinatorial forcing as well as applications of the Extended Pigeonhole Principle. This Principle is also needed to show that countable unions of completely Ramsey sets are completely Ramsey. Finally, working with diagonal coding antichains requires extra care.

Fix a FraΓ―ssé  structure 𝐊\mathbf{K} with universe β„•\mathbb{N} satisfying SDAP+. Fix a good diagonal coding antichain 𝔻\mathbb{D} representing 𝐊\mathbf{K}, and let π’Ÿ\mathcal{D} denote π’Ÿβ€‹(𝔻)\mathcal{D}(\mathbb{D}). We hold to the following convention:

Convention 5.1.

Given Mβˆˆπ’ŸM\in\mathcal{D} and Aβˆˆπ’œβ€‹π’Ÿβ€‹(M)A\in\mathcal{AD}(M), when we write [A,M][A,M], it is assumed that max⁑(A)\max(A) does not contain a splitting predecessor in MM. When we write B∈r​[A,M]βˆ—B\in r[A,M]^{*}, if max⁑(B)\max(B) has a splitting node then it is assumed that that splitting node is not a splitting predecessor in MM.

Definition 5.2.

A family β„±βŠ†π’œβ€‹π’Ÿ\mathcal{F}\subseteq\mathcal{AD} has the Nash-Williams property if for any two distinct members in β„±\mathcal{F}, neither is an initial segment of the other.

Nash-Williams families correspond to metrically open sets in π’Ÿ\mathcal{D}.

Definition 5.3.

Suppose Mβˆˆπ’ŸM\in\mathcal{D} and Bβˆˆπ’œβ€‹π’Ÿ^​(M)B\in\widehat{\mathcal{AD}}(M). A Nash-Williams family β„±βŠ†r​[B,M]βˆ—\mathcal{F}\subseteq r[B,M]^{*} is a front on [B,M]βˆ—[B,M]^{*} if for each N∈[B,M]βˆ—N\in[B,M]^{*}, there is some Cβˆˆβ„±C\in\mathcal{F} such that C⊏NC\sqsubset N.

A front β„±\mathcal{F} on [B,M]βˆ—[B,M]^{*} determines a collection of disjoint (Ellentuck) basic open sets [C,M][C,M], Cβˆˆβ„±C\in\mathcal{F}, whose union is exactly [B,M]βˆ—[B,M]^{*}.

Recall that for level sets X,YβŠ†π”»X,Y\subseteq\mathbb{D}, we write X⊏YX\sqsubset Y exactly when XX and YY have the same cardinality, β„“X<β„“Y\ell_{X}<\ell_{Y}, and Yβ†Ύβ„“X=XY\restriction\ell_{X}=X. More generally, for B,Fβˆˆπ’œβ€‹π’Ÿ^B,F\in\widehat{\mathcal{AD}}, write BβŠ‘FB\sqsubseteq F exactly when B=Fβ†Ώβ„“BB=F\upharpoonleft\ell_{B}; in this case, write B⊏FB\sqsubset F if also β„“B<β„“F\ell_{B}<\ell_{F}. Given β„±βŠ†π’œβ€‹π’Ÿ\mathcal{F}\subseteq\mathcal{AD} and Bβˆˆπ’œβ€‹π’Ÿ^B\in\widehat{\mathcal{AD}}, define

(53) β„±B={Fβˆˆβ„±:B⊏F}.\mathcal{F}_{B}=\{F\in\mathcal{F}:B\sqsubset F\}.

In particular, if β„±βŠ†r​[B,𝔻]βˆ—\mathcal{F}\subseteq r[B,\mathbb{D}]^{*}, then β„±B=β„±\mathcal{F}_{B}=\mathcal{F}. If β„±\mathcal{F} is a Nash-Williams family, then Bβˆˆβ„±B\in\mathcal{F} if and only if β„±B={B}\mathcal{F}_{B}=\{B\}.

Given Mβˆˆπ’ŸM\in\mathcal{D}, let

(54) β„±|M={Fβˆˆβ„±:Fβˆˆπ’œβ€‹π’Ÿβ€‹(M)}.\mathcal{F}|M=\{F\in\mathcal{F}:F\in\mathcal{AD}(M)\}.

With this notation, β„±B|M=β„±βˆ©r​[B,M]βˆ—\mathcal{F}_{B}|M=\mathcal{F}\cap r[B,M]^{*}, for any Bβˆˆπ’œβ€‹π’Ÿ^B\in\widehat{\mathcal{AD}}. For Fβˆˆπ’œβ€‹π’ŸF\in\mathcal{AD}, let |F||F| denote the kk for which Fβˆˆπ’œβ€‹π’ŸkF\in\mathcal{AD}_{k}. Given a set β„±βŠ†π’œβ€‹π’Ÿ\mathcal{F}\subseteq\mathcal{AD}, let

(55) β„±~={rk​(F):Fβˆˆβ„±β€‹and​k≀|F|}.\widetilde{\mathcal{F}}=\{r_{k}(F):F\in\mathcal{F}\mathrm{\ and\ }k\leq|F|\}.

If β„±\mathcal{F} is a Nash-Williams family, then β„±\mathcal{F} consists of the βŠ‘\sqsubseteq-maximal members of β„±~\widetilde{\mathcal{F}}.

We now prove an analogue of the Nash-Williams Theorem for our spaces of good diagonal coding antichains.

Assumption 5.4.

Recall that we are under Convention 5.1. Given Mβˆˆπ’ŸM\in\mathcal{D} and Aβˆˆπ’œβ€‹π’Ÿ^​(M)A\in\widehat{\mathcal{AD}}(M), let d=depthM⁑(A)d=\operatorname{depth}_{M}(A) and D=rd​(M)D=r_{d}(M). Recall that A+A^{+} denotes the union of AA with the set of immediate extensions in M^\widehat{M} of the members of max⁑(A)\max(A). Let BB be a member of π’œβ€‹π’Ÿ^​(M)\widehat{\mathcal{AD}}(M) such that A⊏BβŠ†A+A\sqsubset B\subseteq A^{+}. In what follows we consider simultaneously the two pairs of cases for triples (A,B,k)(A,B,k), from Section 4:

    1. Case (a).

      max⁑(rk+1​(𝔻))\max(r_{k+1}(\mathbb{D})) has a splitting node.

    1. Case (b).

      max⁑(rk+1​(𝔻))\max(r_{k+1}(\mathbb{D})) has a coding node.

    1. Case (i).

      kβ‰₯1k\geq 1, Aβˆˆπ’œβ€‹π’ŸkA\in\mathcal{AD}_{k}, and B=A+B=A^{+}.

    1. Case (ii).

      kβ‰₯0k\geq 0, AA has at least one node, max⁑(A)⊏max⁑(B)\max(A)\sqsubset\max(B), and A=Cβ†Ώβ„“A=C\upharpoonleft\ell for some Cβˆˆπ’œβ€‹π’Ÿk+1C\in\mathcal{AD}_{k+1} and β„“<β„“C\ell<\ell_{C} such that rk​(C)βŠ‘Ar_{k}(C)\sqsubseteq A and BβŠ‘CB\sqsubseteq C.

Theorem 5.5.

Given Mβˆˆπ’ŸM\in\mathcal{D}, (A,B,k)(A,B,k), d=depthM⁑(A)d=\operatorname{depth}_{M}(A), and D=rd​(M)D=r_{d}(M) as in Assumption 5.4, let β„±βŠ†r​[B,M]βˆ—\mathcal{F}\subseteq r[B,M]^{*} be a Nash-Williams family. Then there is an N∈[D,M]N\in[D,M] such that either β„±|N\mathcal{F}|N is a front on [B,N]βˆ—[B,N]^{*} or else β„±|N=βˆ…\mathcal{F}|N=\emptyset.

Proof.

Since AA and BB are fixed, we shall use lower case a,b,c,…a,b,c,\dots to denote members of π’œβ€‹π’Ÿ\mathcal{AD} in this proof. Recall that the notation N≀MN\leq M means that N∈[βˆ…,M]N\in[\emptyset,M]. We say that N≀MN\leq M accepts a∈r​[B,N]βˆ—a\in r[B,N]^{*} if β„±a|N\mathcal{F}_{a}|N is a front on [a,N][a,N]. We say that NN widely-rejects (w-rejects) aa if either

  1. (a)

    aβˆ‰r​[B,N]βˆ—a\not\in r[B,N]^{*}; or

  2. (b)

    a∈r​[B,N]βˆ—a\in r[B,N]^{*} and βˆ€P∈[depthN⁑(a),N]β€‹βˆƒQ∈[a,P]β€‹βˆ€n​(rn​(Q)βˆ‰β„±)\forall P\in[\operatorname{depth}_{N}(a),N]\ \exists Q\in[a,P]\ \forall n(r_{n}(Q)\not\in\mathcal{F}).

We say that NN decides aa if either NN accepts aa or else NN w-rejects aa.111After the author had developed this proof, it was pointed out that a similar asymmetric version of combinatorial forcing was developed by Todorcevic in notes for a graduate course in Ramsey theory. However, those notes do not directly apply to sets of the form [B,M]βˆ—[B,M]^{*}, nor do they include the concluding argument in our proof after Lemma 5.10. For nβˆˆΟ‰n\in\omega, let [n,N][n,N] denote [rn​(N),N][r_{n}(N),N].

Fact 5.6.

If NN accepts aa, then so does each P≀NP\leq N with aβˆˆπ’œβ€‹π’Ÿβ€‹(P)a\in\mathcal{AD}(P). If NN w-rejects aa, then either aβˆ‰r​[B,N]βˆ—a\not\in r[B,N]^{*} and every P≀NP\leq N also rejects aa, or else a∈r​[B,N]βˆ—a\in r[B,N]^{*} and every P∈[depthN⁑(a),N]P\in[\operatorname{depth}_{N}(a),N] w-rejects aa.

Proof.

Suppose NN accepts aa and P≀NP\leq N with aβˆˆπ’œβ€‹π’Ÿβ€‹(P)a\in\mathcal{AD}(P). Since β„±a|N\mathcal{F}_{a}|N is a front on [a,N][a,N], it follows that β„±a|P\mathcal{F}_{a}|P is a front on [a,P][a,P]. Hence PP accepts ss.

Suppose NN w-rejects aa. If aβˆ‰π’œβ€‹π’Ÿβ€‹(N)a\not\in\mathcal{AD}(N), then also for each P≀NP\leq N, aβˆ‰π’œβ€‹π’Ÿβ€‹(P)a\not\in\mathcal{AD}(P) and hence PP w-rejects aa. Otherwise, aβˆˆπ’œβ€‹π’Ÿβ€‹(N)a\in\mathcal{AD}(N). Let n=depthN⁑(a)n=\operatorname{depth}_{N}(a) and suppose P∈[n,N]P\in[n,N]. Since NN w-rejects aa, for each Q∈[n,N]Q\in[n,N] there is an R∈[n,Q]R\in[n,Q] such that for all mm, rm​(R)βˆ‰β„±r_{m}(R)\not\in\mathcal{F}. Note that P∈[n,N]P\in[n,N] implies [n,P]βŠ†[n,N][n,P]\subseteq[n,N]; so for each Q∈[n,P]Q\in[n,P] there is an R∈[a,Q]R\in[a,Q] such that for all mm, rm​(Q)βˆ‰β„±r_{m}(Q)\not\in\mathcal{F}. Therefore, PP w-rejects aa. ∎

Lemma 5.7.

Given a∈r​[B,N]βˆ—a\in r[B,N]^{*} and n=depthN⁑(a)n=\operatorname{depth}_{N}(a), either βˆƒP∈[n,N]\exists P\in[n,N] which w-rejects aa, or else βˆ€P∈[n,N]β€‹βˆƒQ∈[n,P]\forall P\in[n,N]\ \exists Q\in[n,P] which accepts aa.

Proof.

Suppose there is no P∈[n,N]P\in[n,N] which w-rejects aa. Then βˆ€P∈[n,N]\forall P\in[n,N],

(56) βˆƒQ∈[n,P]β€‹βˆ€X∈[a,Q]β€‹βˆƒm​(rm​(X)βˆˆβ„±).\exists Q\in[n,P]\ \forall X\in[a,Q]\ \exists m(r_{m}(X)\in\mathcal{F}).

Thus, for all P∈[n,N]P\in[n,N] there is a Q∈[n,P]Q\in[n,P] such that β„±a|Q\mathcal{F}_{a}|Q is a front on [a,Q][a,Q]; that is, QQ accepts aa. ∎

Fact 5.8.
  1. (a)

    For each a∈r​[B,M]βˆ—a\in r[B,M]^{*}, there is an N∈[depthM⁑(a),M]N\in[\operatorname{depth}_{M}(a),M] which decides aa.

  2. (b)

    If a∈r​[B,M]βˆ—a\in r[B,M]^{*}, then N∈[B,M]βˆ—N\in[B,M]^{*} with aβˆˆπ’œβ€‹π’Ÿβ€‹(N)a\in\mathcal{AD}(N) accepts aa if and only if NN accepts each b∈r|a|+1​[a,N]b\in r_{|a|+1}[a,N].

Proof.

For (a), let n=depthM⁑(a)n=\operatorname{depth}_{M}(a). By Lemma 5.7, either there is an N∈[n,M]N\in[n,M] which w-rejects aa, or else there is an N∈[n,M]N\in[n,M] which accepts aa.

For (b), given the hypotheses, NN accepts aa iff β„±a|N\mathcal{F}_{a}|N is a front on [a,N][a,N] iff for each b∈r|a|+1​[a,N]b\in r_{|a|+1}[a,N], β„±a|N\mathcal{F}_{a}|N is a front on [b,N][b,N] iff NN accepts each b∈r|a|+1​[a,N]b\in r_{|a|+1}[a,N]. ∎

Recall that Bβˆˆπ’œβ€‹π’Ÿ^B\in\widehat{\mathcal{AD}}, but is not necessarily a member of π’œβ€‹π’Ÿ\mathcal{AD}. We shall say that NN accepts BB if NN accepts aa for all a∈rk+1​[B,N]βˆ—a\in r_{k+1}[B,N]^{*}.

Fact 5.9.

If N∈[B,M]βˆ—N\in[B,M]^{*} accepts BB, then β„±B|N\mathcal{F}_{B}|N is a front on [B,N]βˆ—[B,N]^{*}.

Proof.

For each a∈rk+1​[B,N]βˆ—a\in r_{k+1}[B,N]^{*}, NN accepts aa implies that β„±a|N\mathcal{F}_{a}|N is a front on [a,N][a,N]. Since [B,N]βˆ—=⋃{[a,N]:a∈rk+1​[B,N]βˆ—}[B,N]^{*}=\bigcup\{[a,N]:a\in r_{k+1}[B,N]^{*}\}, it follows that β„±|N=⋃{β„±a|N:a∈rk+1​[B,N]βˆ—}\mathcal{F}|N=\bigcup\{\mathcal{F}_{a}|N:a\in r_{k+1}[B,N]^{*}\}, which is a front on [B,N]βˆ—[B,N]^{*}. ∎

Lemma 5.10.

There is an N∈[d,M]N\in[d,M] which decides each aa in r​[B,N]βˆ—r[B,N]^{*}.

Proof.

By finitely many applications of Fact 5.8, we obtain an M1∈[d+1,M]M_{1}\in[d+1,M] such that M1M_{1} decides each a∈r​[B,M]βˆ—a\in r[B,M]^{*} with a≀finrd+1​(M)a\leq_{\mathrm{fin}}r_{d+1}(M). Given MiM_{i}, by finitely many applications of Fact 5.8, we obtain an Mi+1∈[d+i+1,Mi]M_{i+1}\in[d+i+1,M_{i}] such that Mi+1M_{i+1} decides each a∈r​[B,Mi]βˆ—a\in r[B,M_{i}]^{*} with a≀finrd+i+1​(Mi)a\leq_{\mathrm{fin}}r_{d+i+1}(M_{i}). Let N=⋃i=1∞rd+i​(Mi)N=\bigcup_{i=1}^{\infty}r_{d+i}(M_{i}), which is the same as ⋃i=1∞rd+i+1​(Mi)\bigcup_{i=1}^{\infty}r_{d+i+1}(M_{i}). Then N∈[d,M]N\in[d,M] (in fact, N∈[d+1,M]N\in[d+1,M]) and for a∈r​[B,N]βˆ—a\in r[B,N]^{*}, MiM_{i} decides aa, where ii is the index satisfying a≀finrd+i​(Mi)a\leq_{\mathrm{fin}}r_{d+i}(M_{i}). Since N∈[d+i,Mi]N\in[d+i,M_{i}], it follows that NN decides aa in the same way that MiM_{i} does. Thus, NN decides all a∈r​[B,N]βˆ—a\in r[B,N]^{*}. ∎

Now we finish the proof of the theorem. Take NN as in Lemma 5.10 and define a coloring f:r​[B,N]βˆ—β†’2f:r[B,N]^{*}\rightarrow 2 by f​(a)=0f(a)=0 if NN accepts aa and f​(a)=1f(a)=1 if NN w-rejects aa. By the Extended Pigeonhole Principle, Theorem 4.5, there is a P∈[d,N]P\in[d,N] for which ff is monochromatic on rk+1​[B,P]βˆ—r_{k+1}[B,P]^{*}. Now if ff has color 0 on this set, then PP accepts BB and by Fact 5.9, β„±|P\mathcal{F}|P is a front on [B,P]βˆ—[B,P]^{*}.

Otherwise, ff has color 11 on rk+1​[B,P]βˆ—r_{k+1}[B,P]^{*} so PP w-rejects each member of rk+1​[B,P]βˆ—r_{k+1}[B,P]^{*}. Let P0=PP_{0}=P. Apply Theorem 4.5 finitely many (possibly 0) times, to obtain some P1∈[d+1,P0]P_{1}\in[d+1,P_{0}] such that for each a∈r​[B,P1]βˆ—a\in r[B,P_{1}]^{*} with aβŠ†rd+1​(P0)a\subseteq r_{d+1}(P_{0}), all members of r|a|+1​[B,P1]βˆ—r_{|a|+1}[B,P_{1}]^{*} have the same ff-color. Since such an aa is necessarily in rk+1​[B,P0]βˆ—r_{k+1}[B,P_{0}]^{*} and P0P_{0} w-rejects aa, Fact 5.8 implies that this ff-color must be 11.

For iβ‰₯1i\geq 1, we have the following the induction hypothesis: Pi∈[d+i,Piβˆ’1]P_{i}\in[d+i,P_{i-1}] and for each a∈r​[B,Piβˆ’1]βˆ—a\in r[B,P_{i-1}]^{*} with aβŠ†rd+i​(Piβˆ’1)a\subseteq r_{d+i}(P_{i-1}), PiP_{i} w-rejects all members of r|a|+1​[a,Pi]r_{|a|+1}[a,P_{i}]. Apply Theorem 4.5 finitely many times to obtain a Pi+1∈[d+i+1,Pi]P_{i+1}\in[d+i+1,P_{i}] such that ff is monochromatic on r|a|+1​[a,Pi+1]r_{|a|+1}[a,P_{i+1}] for each a∈r​[B,Pi]βˆ—a\in r[B,P_{i}]^{*} with aβŠ†rd+i+1​(Pi)a\subseteq r_{d+i+1}(P_{i}). Fix an a∈r​[B,Pi]βˆ—a\in r[B,P_{i}]^{*} with aβŠ†rd+i+1​(Pi)a\subseteq r_{d+i+1}(P_{i}). If |a|=k+1|a|=k+1 then Pi+1P_{i+1} w-rejects aa, since a∈rk+1​[B,P]βˆ—a\in r_{k+1}[B,P]^{*} and Pi+1∈[B,P]βˆ—P_{i+1}\in[B,P]^{*}. Suppose now that |a|>k+1|a|>k+1. By the induction hypothesis, PiP_{i} w-rejects aa since a∈r|b|+1​[b,Pi]a\in r_{|b|+1}[b,P_{i}], where b=r|a|βˆ’1​(a)βŠ†rd+i​(Piβˆ’1)b=r_{|a|-1}(a)\subseteq r_{d+i}(P_{i-1}). Now if the ff-color on r|a|+1​[a,Pi+1]r_{|a|+1}[a,P_{i+1}] is 0, then Pi+1P_{i+1} accepts aa by Fact 5.8, a contradiction. Hence, ff has color 11 on r|a|+1​[a,Pi+1]r_{|a|+1}[a,P_{i+1}]; in particular, Pi+1P_{i+1} w-rejects each member of r|a|+1​[a,Pi+1]r_{|a|+1}[a,P_{i+1}].

Let Q=⋃i=1∞rd+i​(Pi)Q=\bigcup_{i=1}^{\infty}r_{d+i}(P_{i}). Then QQ w-rejects each member of r​[B,Q]βˆ—r[B,Q]^{*}. By definition of w-rejects, for each a∈r​[B,Q]βˆ—a\in r[B,Q]^{*},

(57) βˆ€R∈[depthQ⁑(a),Q]β€‹βˆƒX∈[a,R]β€‹βˆ€n​(rn​(X)βˆ‰β„±)\forall R\in[\operatorname{depth}_{Q}(a),Q]\ \exists X\in[a,R]\ \forall n(r_{n}(X)\not\in\mathcal{F})

Suppose toward a contradiction that there is an aβˆˆβ„±|Qa\in\mathcal{F}|Q. Then for all X∈[a,Q]X\in[a,Q], r|a|​(X)=aβˆˆβ„±r_{|a|}(X)=a\in\mathcal{F}. So Q∈[depthQ⁑(a),Q]Q\in[\operatorname{depth}_{Q}(a),Q] and for all X∈[a,Q]X\in[a,Q], βˆƒn​(rn​(X)βˆˆβ„±)\exists n(r_{n}(X)\in\mathcal{F}). But this contradicts (57). Thus β„±|Q\mathcal{F}|Q must be empty. ∎

Definition 5.11.

Let 𝒳\mathcal{X} be a subset of π’Ÿ\mathcal{D}. We say that 𝒳\mathcal{X} is Ramsey if for each Mβˆˆπ’ŸM\in\mathcal{D} there is an N≀MN\leq M such that either π’³βŠ†[βˆ…,N]\mathcal{X}\subseteq[\emptyset,N] or else π’³βˆ©[βˆ…,N]=βˆ…\mathcal{X}\cap[\emptyset,N]=\emptyset. 𝒳\mathcal{X} is said to be completely Ramsey (CR) if for each Cβˆˆπ’œβ€‹π’ŸC\in\mathcal{AD} and each Mβˆˆπ’ŸM\in\mathcal{D}, there is an N∈[C,M]N\in[C,M] such that either [C,N]βŠ†π’³[C,N]\subseteq\mathcal{X} or else [C,N]βˆ©π’³=βˆ…[C,N]\cap\mathcal{X}=\emptyset. We shall say that 𝒳\mathcal{X} is CRβˆ— if given Mβˆˆπ’ŸM\in\mathcal{D} and (A,B)(A,B) as in Assumption 5.4, there is an N∈[D,M]N\in[D,M] such that either [B,N]βˆ—βŠ†π’³[B,N]^{*}\subseteq\mathcal{X} or else [B,N]βˆ—βˆ©π’³=βˆ…[B,N]^{*}\cap\mathcal{X}=\emptyset.

Remark 5.12.

Since metrically open sets correspond to Nash-Williams families, Theorem 5.5 implies that metrically open sets are not only completely Ramsey but moreover CRβˆ—, even when relativized below some Mβˆˆπ’ŸM\in\mathcal{D}.

Lemma 5.13.

Complements of CRβˆ— sets are CRβˆ—.

Proof.

Suppose π’³βŠ†π’Ÿ\mathcal{X}\subseteq\mathcal{D} is CRβˆ—, and Mβˆˆπ’ŸM\in\mathcal{D}, (A,B,k)(A,B,k), d=depthM⁑(A)d=\operatorname{depth}_{M}(A), and D=rd​(M)D=r_{d}(M) are as in Assumption 5.4. By definition of CRβˆ—, there is an N∈[D,M]N\in[D,M] such that either [B,N]βˆ—βŠ†π’³[B,N]^{*}\subseteq\mathcal{X} or else [B,N]βˆ—βˆ©π’³=βˆ…[B,N]^{*}\cap\mathcal{X}=\emptyset. Letting 𝒴=π’Ÿβˆ–π’³\mathcal{Y}=\mathcal{D}\setminus\mathcal{X}, we see that either [B,N]βˆ—βˆ©π’΄=βˆ…[B,N]^{*}\cap\mathcal{Y}=\emptyset or else [B,N]βˆ—βŠ†π’΄[B,N]^{*}\subseteq\mathcal{Y}. ∎

In the rest of this section, given Mβˆˆπ’ŸM\in\mathcal{D}, endow [βˆ…,M][\emptyset,M] with the subspace topology inherited from π’Ÿ\mathcal{D} with the metric topology. The next two lemmas build up to Lemma 5.16, which will show that countable unions of CRβˆ— sets are CRβˆ—.

Lemma 5.14.

Suppose π’³βŠ†π’Ÿ\mathcal{X}\subseteq\mathcal{D} is CRβˆ—. Then for each Mβˆˆπ’ŸM\in\mathcal{D} and each Cβˆˆπ’œβ€‹π’Ÿβ€‹(M)C\in\mathcal{AD}(M), there is an N∈[C,M]N\in[C,M] such that π’³βˆ©[βˆ…,N]\mathcal{X}\cap[\emptyset,N] is metrically open in [βˆ…,N][\emptyset,N].

Proof.

Fix Mβˆˆπ’ŸM\in\mathcal{D} and Cβˆˆπ’œβ€‹π’Ÿβ€‹(M)C\in\mathcal{AD}(M). Notice that [βˆ…,M]=⋃j<j~[Bj,M]βˆ—[\emptyset,M]=\bigcup_{j<\tilde{j}}[B_{j},M]^{*}, where ⟨(Aj,Bj):j<j~⟩\langle(A_{j},B_{j}):j<\tilde{j}\rangle enumerates all pairs (A,B)(A,B) with depthM⁑(A)=depthM⁑(C)\operatorname{depth}_{M}(A)=\operatorname{depth}_{M}(C) satisfying Assumption 5.4.

Let Mβˆ’1=MM_{-1}=M. Given Mjβˆ’1M_{j-1} for j<j~j<\tilde{j}, 𝒳\mathcal{X} being CRβˆ— implies there is an Mj∈[C,Mjβˆ’1]M_{j}\in[C,M_{j-1}] such that either [Bj,Mj]βˆ—βŠ†π’³[B_{j},M_{j}]^{*}\subseteq\mathcal{X} or else π’³βˆ©[Bj,Mj]βˆ—=βˆ…\mathcal{X}\cap[B_{j},M_{j}]^{*}=\emptyset. Let N=Mj~βˆ’1N=M_{\tilde{j}-1}. Then N∈[C,M]N\in[C,M] and for each j<j~j<\tilde{j}, [Bj,N]βˆ—βŠ†[Bj,Mj]βˆ—[B_{j},N]^{*}\subseteq[B_{j},M_{j}]^{*}. Since [βˆ…,N]=⋃j<j~[Bj,N]βˆ—[\emptyset,N]=\bigcup_{j<\tilde{j}}[B_{j},N]^{*}, it follows that

(58) π’³βˆ©[βˆ…,N]=⋃j<j~(π’³βˆ©[Bj,N]βˆ—).\mathcal{X}\cap[\emptyset,N]=\bigcup_{j<\tilde{j}}(\mathcal{X}\cap[B_{j},N]^{*}).

For j<j~j<\tilde{j}, if [Bj,Mj]βˆ—βŠ†π’³[B_{j},M_{j}]^{*}\subseteq\mathcal{X} then π’³βˆ©[Bj,N]βˆ—=[Bj,N]βˆ—\mathcal{X}\cap[B_{j},N]^{*}=[B_{j},N]^{*}; and if π’³βˆ©[Bj,Mj]βˆ—=βˆ…\mathcal{X}\cap[B_{j},M_{j}]^{*}=\emptyset then π’³βˆ©[Bj,N]βˆ—=βˆ…\mathcal{X}\cap[B_{j},N]^{*}=\emptyset. Thus,

(59) π’³βˆ©[βˆ…,N]=⋃j∈J[Bj,N]βˆ—,\mathcal{X}\cap[\emptyset,N]=\bigcup_{j\in J}[B_{j},N]^{*},

where J={j<j~:[Bj,Mj]βˆ—βŠ†π’³}J=\{j<\tilde{j}:[B_{j},M_{j}]^{*}\subseteq\mathcal{X}\}. As each [Bj,N]βˆ—[B_{j},N]^{*} is metrically open in the subspace [βˆ…,N][\emptyset,N], π’³βˆ©[βˆ…,N]\mathcal{X}\cap[\emptyset,N] is also metrically open in the subspace [0,N][0,N]. ∎

Lemma 5.15.

Suppose 𝒳n\mathcal{X}_{n}, nβˆˆβ„•n\in\mathbb{N}, are CRβˆ— sets. Then for each Mβˆˆπ’ŸM\in\mathcal{D} and each Cβˆˆπ’œβ€‹π’Ÿβ€‹(M)C\in\mathcal{AD}(M), there is an N∈[C,M]N\in[C,M] such that for each nβˆˆβ„•n\in\mathbb{N}, 𝒳n∩[βˆ…,N]\mathcal{X}_{n}\cap[\emptyset,N] is metrically open in [βˆ…,N][\emptyset,N].

Proof.

Assume the hypotheses and let d=depthM⁑(C)d=\operatorname{depth}_{M}(C) and D=rd​(M)D=r_{d}(M). Since 𝒳0\mathcal{X}_{0} is CRβˆ—, Lemma 5.14 implies there is an M0∈[D,M]M_{0}\in[D,M] and a metrically open set π’ͺ0βŠ†π’Ÿ\mathcal{O}_{0}\subseteq\mathcal{D} satisfying 𝒳0∩[βˆ…,M0]=π’ͺ0∩[βˆ…,M0]\mathcal{X}_{0}\cap[\emptyset,M_{0}]=\mathcal{O}_{0}\cap[\emptyset,M_{0}]. In general, given MiM_{i}, by Lemma 5.14 there is some Mi+1∈[rd+i+1​(Mi),Mi]M_{i+1}\in[r_{d+i+1}(M_{i}),M_{i}] and some metrically open π’ͺiβŠ†π’Ÿ\mathcal{O}_{i}\subseteq\mathcal{D} satisfying 𝒳i∩[βˆ…,Mi]=π’ͺi∩[βˆ…,Mi]\mathcal{X}_{i}\cap[\emptyset,M_{i}]=\mathcal{O}_{i}\cap[\emptyset,M_{i}]. Let N=⋃i=0∞rd+i​(Mi)N=\bigcup_{i=0}^{\infty}r_{d+i}(M_{i}). Then NN is a member of [D,M][D,M].

Letting Mβˆ’1=MM_{-1}=M, note that N∈[rd+i​(Mi),Miβˆ’1]N\in[r_{d+i}(M_{i}),M_{i-1}] for each iβˆˆβ„•i\in\mathbb{N}. It follows that for each iβˆˆβ„•i\in\mathbb{N}, 𝒳i∩[βˆ…,N]=π’ͺi∩[βˆ…,N]\mathcal{X}_{i}\cap[\emptyset,N]=\mathcal{O}_{i}\cap[\emptyset,N]. Hence 𝒳i∩[βˆ…,N]\mathcal{X}_{i}\cap[\emptyset,N] is metrically open in [βˆ…,N][\emptyset,N]. ∎

Lemma 5.16.

Countable unions of CRβˆ— sets are CRβˆ—.

Proof.

Suppose 𝒳n\mathcal{X}_{n}, nβˆˆβ„•n\in\mathbb{N}, are CRβˆ— subsets of π’Ÿ\mathcal{D}, and let 𝒳=⋃n=0βˆžπ’³n\mathcal{X}=\bigcup_{n=0}^{\infty}\mathcal{X}_{n}. Let (M,A,B,k)(M,A,B,k) be as in Assumption 5.4, and let d=depthM⁑(A)d=\operatorname{depth}_{M}(A) and D=rd​(M)D=r_{d}(M). By Lemma 5.15, there is a Mβ€²βˆˆ[D,M]M^{\prime}\in[D,M] such that for each nn, 𝒳n∩[βˆ…,Mβ€²]\mathcal{X}_{n}\cap[\emptyset,M^{\prime}] is metrically open in [βˆ…,Mβ€²][\emptyset,M^{\prime}]. Thus, π’³βˆ©[βˆ…,Mβ€²]\mathcal{X}\cap[\emptyset,M^{\prime}] is metrically open in [βˆ…,Mβ€²][\emptyset,M^{\prime}], so π’³βˆ©[βˆ…,Mβ€²]=π’ͺ∩[βˆ…,Mβ€²]\mathcal{X}\cap[\emptyset,M^{\prime}]=\mathcal{O}\cap[\emptyset,M^{\prime}] for some metrically open set π’ͺβŠ†π’Ÿ\mathcal{O}\subseteq\mathcal{D}.

Theorem 5.5 implies that π’ͺ\mathcal{O} is CRβˆ— in [βˆ…,Mβ€²][\emptyset,M^{\prime}]. Hence, there is some N∈[D,Mβ€²]N\in[D,M^{\prime}] such that either [B,N]βˆ—βŠ†π’ͺ[B,N]^{*}\subseteq\mathcal{O} or else [B,N]βˆ—βˆ©π’ͺ=βˆ…[B,N]^{*}\cap\mathcal{O}=\emptyset. Therefore, either

(60) [B,N]βˆ—=[B,N]βˆ—βˆ©[βˆ…,Mβ€²]βŠ†π’ͺ∩[βˆ…,Mβ€²]=π’³βˆ©[βˆ…,Mβ€²],[B,N]^{*}=[B,N]^{*}\cap[\emptyset,M^{\prime}]\subseteq\mathcal{O}\cap[\emptyset,M^{\prime}]=\mathcal{X}\cap[\emptyset,M^{\prime}],

or else

(61) [B,N]βˆ—βˆ©π’³\displaystyle[B,N]^{*}\cap\mathcal{X} =[B,N]βˆ—βˆ©[βˆ…,Mβ€²]βˆ©π’³\displaystyle=[B,N]^{*}\cap[\emptyset,M^{\prime}]\cap\mathcal{X}
(62) βŠ†[B,N]βˆ—βˆ©[βˆ…,Mβ€²]∩π’ͺ\displaystyle\subseteq[B,N]^{*}\cap[\emptyset,M^{\prime}]\cap\mathcal{O}
(63) βŠ†[B,N]βˆ—βˆ©π’ͺ=βˆ….\displaystyle\subseteq[B,N]^{*}\cap\mathcal{O}=\emptyset.

Thus, 𝒳\mathcal{X} is CRβˆ—. ∎

Theorem 5.17.

Let 𝐊\mathbf{K} be an enumerated FraΓ―ssé structure satisfying SDAP+, with finitely many relations of arity at most two. Let 𝔻\mathbb{D} be a good diagonal coding antichain representing 𝐊\mathbf{K}. Then the collection of CRβˆ— subsets of π’Ÿβ€‹(𝔻)\mathcal{D}(\mathbb{D}) contains all Borel subsets of π’Ÿβ€‹(𝔻)\mathcal{D}(\mathbb{D}). In particular, Borel subsets of the space π’Ÿβ€‹(𝔻)\mathcal{D}(\mathbb{D}) are completely Ramsey, and hence Ramsey.

Proof.

This follows from Theorem 5.5 and Lemmas 5.13 and 5.16. ∎

Remark 5.18.

A simple modification of the proof yields the same result for LSDAP+ structures.

6. Main Theorems

This section contains the main theorem that Borel sets in our spaces of subcopies of a given structure 𝐊\mathbf{K} are Ramsey, conditions under which analogues of the Ellentuck theorem hold, and a Nash-Williams-style corollary recovering exact big Ramsey degrees.

6.1. Borel sets are Ramsey

We now prove the Main Theorem. Fix an enumerated FraΓ―ssé structure 𝐊\mathbf{K} satisfying SDAP+ and a good diagonal coding antichain π”»βŠ†π•Œβ€‹(𝐊)\mathbb{D}\subseteq\mathbb{U}(\mathbf{K}) representing a subcopy of 𝐊\mathbf{K}. Recall that the universe of 𝐊\mathbf{K} is β„•\mathbb{N}. Each substructure 𝐌\mathbf{M} of 𝐊\mathbf{K} is uniquely identified with its universe MβŠ†β„•\mathrm{M}\subseteq\mathbb{N}, which in turn, is uniquely identified with the set of coding nodes {cn:n∈M}\{c_{n}:n\in\mathrm{M}\}. To avoid any ambiguity, we will use T𝐌T_{\mathbf{M}} (rather than MM) to denote the subtree of 𝔻\mathbb{D} induced by the set of coding nodes {cn:n∈M}\{c_{n}:n\in\mathrm{M}\}. Define

(64) ℬ​(𝔻)={M∈[β„•]β„•:TπŒβˆˆπ’Ÿβ€‹(𝔻)}.\mathcal{B}(\mathbb{D})=\{\mathrm{M}\in[\mathbb{N}]^{\mathbb{N}}:T_{\mathbf{M}}\in\mathcal{D}(\mathbb{D})\}.

That is, MβŠ†β„•\mathrm{M}\subseteq\mathbb{N} is a member of ℬ​(𝔻)\mathcal{B}(\mathbb{D}) if and only if {cn:n∈M}βŠ†π”»\{c_{n}:n\in\mathrm{M}\}\subseteq\mathbb{D} and the tree induced by {cn:n∈M}\{c_{n}:n\in\mathrm{M}\} is similar to the tree induced by 𝔻\mathbb{D}. Note that ℬ​(𝔻)\mathcal{B}(\mathbb{D}) is a subspace of the Baire space.

Let 𝐃\mathbf{D} denote the substructure πŠβ†Ύπ”»\mathbf{K}\restriction\mathbb{D}, and let ⟨dn:nβˆˆβ„•βŸ©\langle d_{n}:n\in\mathbb{N}\rangle be the increasing enumeration of the universe D\mathrm{D} of 𝐃\mathbf{D}. Notice that ⟨cdn:n∈𝐍⟩\langle c_{d_{n}}:n\in\mathbf{N}\rangle enumerates the coding nodes in 𝔻\mathbb{D}. Define

(65) πŠβ€‹(𝐃)={πŒβ‰€πƒ:Mβˆˆβ„¬β€‹(𝔻)}.\mathbf{K}(\mathbf{D})=\{\mathbf{M}\leq\mathbf{D}:\mathrm{M}\in\mathcal{B}(\mathbb{D})\}.

That is, πŠβ€‹(𝐃)\mathbf{K}(\mathbf{D}) is the subspace of (𝐊𝐊){\mathbf{K}\choose\mathbf{K}} consisting of all substructures 𝐌\mathbf{M} of 𝐃\mathbf{D} with universe Mβˆˆβ„¬β€‹(𝔻)\mathrm{M}\in\mathcal{B}(\mathbb{D}). Notice that πŠβ€‹(𝐃)\mathbf{K}(\mathbf{D}) is identified with a subspace of the Baire space via its identification with ℬ​(𝔻)\mathcal{B}(\mathbb{D}). For πŒβˆˆπŠβ€‹(𝐃)\mathbf{M}\in\mathbf{K}(\mathbf{D}), we will let πŠβ€‹(𝐌)\mathbf{K}(\mathbf{M}) denote the cube of all substructures of 𝐌\mathbf{M} in πŠβ€‹(𝐃)\mathbf{K}(\mathbf{D}).

For πŒβˆˆπŠβ€‹(𝐃)\mathbf{M}\in\mathbf{K}(\mathbf{D}), let ⟨mi:iβˆˆβ„•βŸ©\langle m_{i}:i\in\mathbb{N}\rangle be the increasing enumeration of M\mathrm{M}. Then increasing bijection mi↦dim_{i}\mapsto d_{i} induces an isomorphism from 𝐌\mathbf{M} to 𝐃\mathbf{D}, and cmi↦cdic_{m_{i}}\mapsto c_{d_{i}} induces a similarity map from T𝐌T_{\mathbf{M}} to 𝔻\mathbb{D}. Given nβˆˆβ„•n\in\mathbb{N}, define 𝐌n=πŒβ†Ύ{mi:i<n}\mathbf{M}_{n}=\mathbf{M}\restriction\{m_{i}:i<n\}. Let

(66) π’œβ€‹πŠβ€‹(𝐃)={𝐌n:πŒβˆˆπŠβ€‹(𝐃)​and​nβˆˆβ„•}.\mathcal{A}\mathbf{K}(\mathbf{D})=\{\mathbf{M}_{n}:\mathbf{M}\in\mathbf{K}(\mathbf{D})\mathrm{\ and\ }n\in\mathbb{N}\}.

For π€βˆˆπ’œβ€‹πŠβ€‹(𝐃)\mathbf{A}\in\mathcal{A}\mathbf{K}(\mathbf{D}) and πŒβˆˆπŠβ€‹(𝐃)\mathbf{M}\in\mathbf{K}(\mathbf{D}), write π€βŠπŒ\mathbf{A}\sqsubset\mathbf{M} if and only if 𝐀=𝐌n\mathbf{A}=\mathbf{M}_{n} for some nn. Define

(67) [𝐀,𝐌]={πβˆˆπŠβ€‹(𝐃):π€βŠπ}.[\mathbf{A},\mathbf{M}]=\{\mathbf{N}\in\mathbf{K}(\mathbf{D}):\mathbf{A}\sqsubset\mathbf{N}\}.

These are the basic open sets for the Ellentuck topology on πŠβ€‹(𝐃)\mathbf{K}(\mathbf{D}) corresponding to the basic Ellentuck open sets [A,M][\mathrm{A},\mathrm{M}] in the Baire space, where A\mathrm{A} and M\mathrm{M} are the universes of 𝐀\mathbf{A} and 𝐌\mathbf{M}, respectively. The basic open sets for the metric topology on πŠβ€‹(𝐃)\mathbf{K}(\mathbf{D}) are those of the form [𝐀,𝐃][\mathbf{A},\mathbf{D}], where π€βˆˆπ’œβ€‹πŠβ€‹(𝐃)\mathbf{A}\in\mathcal{A}\mathbf{K}(\mathbf{D}).

Let ΞΈ:πŠβ€‹(𝐃)β†’π’Ÿβ€‹(𝔻)\theta:\mathbf{K}(\mathbf{D})\rightarrow\mathcal{D}(\mathbb{D}) denote the map which sends each πŒβˆˆπŠβ€‹(𝐃)\mathbf{M}\in\mathbf{K}(\mathbf{D}) to the tree T𝐌T_{\mathbf{M}} in π’Ÿβ€‹(𝔻)\mathcal{D}(\mathbb{D}). This map is certainly a bijection. We will show that ΞΈ\theta is in fact a homeomorphism between these two spaces with their metric topologies.

For nβˆˆβ„•n\in\mathbb{N}, let knk_{n} denote the least integer such that cnβˆ’1π”»βˆˆrkn​(𝔻)c^{\mathbb{D}}_{n-1}\in r_{k_{n}}(\mathbb{D}). Since each Tβˆˆπ’Ÿβ€‹(𝔻)T\in\mathcal{D}(\mathbb{D}) is similar to 𝔻\mathbb{D}, it follows that knk_{n} is the least integer such that the (nβˆ’1)(n-1)-st coding node of TT is in rkn​(T)r_{k_{n}}(T). In particular, knk_{n} is least such that 𝐃n=πŠβ†Ύrkn​(𝔻)\mathbf{D}_{n}=\mathbf{K}\restriction r_{k_{n}}(\mathbb{D}). For the following lemma, recall that since 𝔻\mathbb{D} is a good diagonal coding antichain, there is some n𝔻n_{\mathbb{D}} such that for each nβ‰₯n𝔻n\geq n_{\mathbb{D}}, there is a one-to-one correspondence between the nodes in max(rkn(𝔻))+\max(r_{k_{n}}(\mathbb{D}))^{+} and the 11-types over 𝐃n\mathbf{D}_{n}.

Lemma 6.1.

Suppose πŒβˆˆπŠβ€‹(𝐃)\mathbf{M}\in\mathbf{K}(\mathbf{D}) and 𝐀=𝐌n\mathbf{A}=\mathbf{M}_{n}, where nβ‰₯n𝔻n\geq n_{\mathbb{D}}. Then θ​([𝐀,𝐌])=[rkn​(T𝐌),T𝐌]\theta([\mathbf{A},\mathbf{M}])=[r_{k_{n}}(T_{\mathbf{M}}),T_{\mathbf{M}}].

Proof.

Since nβ‰₯n𝔻n\geq n_{\mathbb{D}}, there is a one-to-one correspondence between the nodes in max(rkn(T𝐌))+\max(r_{k_{n}}(T_{\mathbf{M}}))^{+} and the 11-types over 𝐀\mathbf{A}. For 𝐍∈[𝐀,𝐌]\mathbf{N}\in[\mathbf{A},\mathbf{M}], 𝐍\mathbf{N} extends 𝐀\mathbf{A} to some isomorphic subcopy of 𝐌\mathbf{M}, and T𝐍T_{\mathbf{N}} is a subtree of T𝐌T_{\mathbf{M}}. In order for 𝐍\mathbf{N} to be isomorphic to 𝐌\mathbf{M}, each 11-type over 𝐀\mathbf{A} must be represented by a node in max(rkn(T𝐍))+\max(r_{k_{n}}(T_{\mathbf{N}}))^{+}. The only way this is possible is if rkn​(T𝐍)=rkn​(T𝐌)r_{k_{n}}(T_{\mathbf{N}})=r_{k_{n}}(T_{\mathbf{M}}). Thus, letting A=rkn​(T𝐌)A=r_{k_{n}}(T_{\mathbf{M}}),

(68) θ​([𝐀,𝐌])\displaystyle\theta([\mathbf{A},\mathbf{M}]) ={T𝐍:𝐍∈[𝐀,𝐌]}\displaystyle=\{T_{\mathbf{N}}:\mathbf{N}\in[\mathbf{A},\mathbf{M}]\}
(69) ={T𝐍:𝐍∈[𝐀,𝐌]​and​rkn​(T𝐍)=rkn​(T𝐌)}\displaystyle=\{T_{\mathbf{N}}:\mathbf{N}\in[\mathbf{A},\mathbf{M}]\mathrm{\ and\ }r_{k_{n}}(T_{\mathbf{N}})=r_{k_{n}}(T_{\mathbf{M}})\}
(70) ={T𝐍:A⊏T𝐍​and​T𝐍≀T𝐌}\displaystyle=\{T_{\mathbf{N}}:A\sqsubset T_{\mathbf{N}}\mathrm{\ and\ }T_{\mathbf{N}}\leq T_{\mathbf{M}}\}
(71) =[A,T𝐌].\displaystyle=[A,T_{\mathbf{M}}].

∎

Thus, ΞΈ\theta takes the basic Ellentuck open set [𝐌n,𝐌][\mathbf{M}_{n},\mathbf{M}] to the basic Ellentuck open set [rkn​(T𝐌),T𝐌][r_{k_{n}}(T_{\mathbf{M}}),T_{\mathbf{M}}] whenever nβ‰₯n𝔻n\geq n_{\mathbb{D}}. Furthermore, ΞΈ\theta is a homeomorphism from πŠβ€‹(𝐃)\mathbf{K}(\mathbf{D}) with its metric topology to π’Ÿβ€‹(𝔻)\mathcal{D}(\mathbb{D}) with its metric topology, as follows from the next lemma.

Lemma 6.2.

The map ΞΈ\theta takes each basic metrically open set in πŠβ€‹(𝐃)\mathbf{K}(\mathbf{D}) to a metrically open set in ℬ​(𝔻)\mathcal{B}(\mathbb{D}), and ΞΈβˆ’1\theta^{-1} takes each basic metrically open set in ℬ​(𝔻)\mathcal{B}(\mathbb{D}) to a metrically open set in πŠβ€‹(𝐃)\mathbf{K}(\mathbf{D}).

Proof.

Let [𝐀,𝐃][\mathbf{A},\mathbf{D}] be a basic open set in the metric topology on πŠβ€‹(𝐃)\mathbf{K}(\mathbf{D}), and let nn be the number of vertices in 𝐀\mathbf{A}. Then

(72) θ​([𝐀,𝐃])\displaystyle\theta([\mathbf{A},\mathbf{D}]) =⋃{[rkn​(T𝐌),𝔻]:π€βŠπŒ}\displaystyle=\bigcup\{[r_{k_{n}}(T_{\mathbf{M}}),\mathbb{D}]:\mathbf{A}\sqsubset\mathbf{M}\}
(73) =⋃{[B,𝔻]:Bβˆˆπ’œβ€‹π’Ÿkn​and​𝐃↾B=𝐀}\displaystyle=\bigcup\{[B,\mathbb{D}]:B\in\mathcal{AD}_{k_{n}}\mathrm{\ and\ }\mathbf{D}\restriction B=\mathbf{A}\}

which is a countable union of metrically open sets in π’Ÿβ€‹(𝔻)\mathcal{D}(\mathbb{D}). Conversely, given a basic open set [A,𝔻][A,\mathbb{D}] in the metric topology on π’Ÿβ€‹(𝔻)\mathcal{D}(\mathbb{D}), we may without loss of generality assume that Aβˆˆπ’œβ€‹π’ŸknA\in\mathcal{AD}_{k_{n}} for some nn. Let nβ€²n^{\prime} denote the least integer such that for each πŒβˆˆπŠβ€‹(𝐃)\mathbf{M}\in\mathbf{K}(\mathbf{D}),

(74) rkn​(T𝐌nβ€²)=rkn​(T𝐌).r_{k_{n}}(T_{\mathbf{M}_{n^{\prime}}})=r_{k_{n}}(T_{\mathbf{M}}).

Then

(75) ΞΈβˆ’1​([A,𝔻])\displaystyle\theta^{-1}([A,\mathbb{D}]) ={πŒβˆˆπŠβ€‹(𝐃):T𝐌∈[A,𝔻]}\displaystyle=\{\mathbf{M}\in\mathbf{K}(\mathbf{D}):T_{\mathbf{M}}\in[A,\mathbb{D}]\}
(76) =⋃{[𝐁,𝐃]:πβˆˆπ’œβ€‹πŠβ€‹(𝐃)n′​and​rkn​(T𝐁)=A},\displaystyle=\bigcup\{[\mathbf{B},\mathbf{D}]:\mathbf{B}\in\mathcal{A}\mathbf{K}(\mathbf{D})_{n^{\prime}}\mathrm{\ and\ }r_{k_{n}}(T_{\mathbf{B}})=A\},

which is a countable union of basic metrically open sets in πŠβ€‹(𝐃)\mathbf{K}(\mathbf{D}). ∎

A set π’³βŠ†πŠβ€‹(𝐃)\mathcal{X}\subseteq\mathbf{K}(\mathbf{D}) is Ramsey if for any πŒβˆˆπŠβ€‹(𝐃)\mathbf{M}\in\mathbf{K}(\mathbf{D}), there is some πβ‰€πŒ\mathbf{N}\leq\mathbf{M} in πŠβ€‹(𝐃)\mathbf{K}(\mathbf{D}) such that either πŠβ€‹(𝐍)βŠ†π’³\mathbf{K}(\mathbf{N})\subseteq\mathcal{X} or else πŠβ€‹(𝐍)βˆ©π’³=βˆ…\mathbf{K}(\mathbf{N})\cap\mathcal{X}=\emptyset. A set π’³βŠ†πŠβ€‹(𝐃)\mathcal{X}\subseteq\mathbf{K}(\mathbf{D}) is completely Ramsey if for any π€βˆˆπ’œβ€‹πŠβ€‹(𝐃)\mathbf{A}\in\mathcal{A}\mathbf{K}(\mathbf{D}) and πŒβˆˆπŠβ€‹(𝐃)\mathbf{M}\in\mathbf{K}(\mathbf{D}), there is some 𝐍∈[𝐀,𝐌]\mathbf{N}\in[\mathbf{A},\mathbf{M}] such that either [𝐀,𝐍]βŠ†π’³[\mathbf{A},\mathbf{N}]\subseteq\mathcal{X} or else [𝐀,𝐍]βˆ©π’³=βˆ…[\mathbf{A},\mathbf{N}]\cap\mathcal{X}=\emptyset.

Theorem 6.3.

Let 𝐊\mathbf{K} be an enumerated FraΓ―ssé structure satisfying SDAP+ (or LSDAP+) with finitely many relations of arity at most two, let 𝔻\mathbb{D} be a good diagonal coding antichain, and let 𝐃=πŠβ†Ύπ”»\mathbf{D}=\mathbf{K}\restriction\mathbb{D}. Then every Borel subset π’³βŠ†πŠβ€‹(𝐃)\mathcal{X}\subseteq\mathbf{K}(\mathbf{D}) is completely Ramsey, and hence Ramsey.

Proof.

Let 𝒳\mathcal{X} be a Borel subset of πŠβ€‹(𝐃)\mathbf{K}(\mathbf{D}), and suppose π€βˆˆπ’œβ€‹πŠβ€‹(𝐃)\mathbf{A}\in\mathcal{A}\mathbf{K}(\mathbf{D}) and πŒβˆˆπŠβ€‹(𝐃)\mathbf{M}\in\mathbf{K}(\mathbf{D}). If [𝐀,𝐌]=βˆ…[\mathbf{A},\mathbf{M}]=\emptyset then we are done, so assume that [𝐀,𝐌][\mathbf{A},\mathbf{M}] is non-empty. By shrinking 𝐌\mathbf{M} if necessary, we may assume that 𝐀\mathbf{A} is an initial segment of 𝐌\mathbf{M}. Let nn be the integer such that 𝐀=𝐌n\mathbf{A}=\mathbf{M}_{n}. By Lemma 6.1, θ​([𝐀,𝐌])=[rkn​(T𝐌),T𝐌]\theta([\mathbf{A},\mathbf{M}])=[r_{k_{n}}(T_{\mathbf{M}}),T_{\mathbf{M}}]. Let AA denote rkn​(T𝐌)r_{k_{n}}(T_{\mathbf{M}}).

Let 𝒴\mathcal{Y} be the ΞΈ\theta-image of 𝒳\mathcal{X}, noting that 𝒴\mathcal{Y} is Borel in π’Ÿβ€‹(𝔻)\mathcal{D}(\mathbb{D}) with the metric topology by Lemma 6.2. Apply Theorem 5.17 to obtain an N∈[A,T𝐌]N\in[A,T_{\mathbf{M}}] such that either [A,N]βŠ†π’΄[A,N]\subseteq\mathcal{Y} or else [A,N]βˆ©π’΄=βˆ…[A,N]\cap\mathcal{Y}=\emptyset. Let 𝐍=𝐃↾N\mathbf{N}=\mathbf{D}\restriction N. Then T𝐍=NT_{\mathbf{N}}=N, A=rkn​(T𝐍)A=r_{k_{n}}(T_{\mathbf{N}}), and ΞΈβˆ’1​([A,N])=ΞΈβˆ’1​([rkn​(T𝐍),T𝐍])=[𝐀,𝐍]\theta^{-1}([A,N])=\theta^{-1}([r_{k_{n}}(T_{\mathbf{N}}),T_{\mathbf{N}}])=[\mathbf{A},\mathbf{N}], by Lemma 6.1. Thus, either [𝐀,𝐍]βŠ†π’³[\mathbf{A},\mathbf{N}]\subseteq\mathcal{X} or else [𝐀,𝐍]βˆ©π’³=βˆ…[\mathbf{A},\mathbf{N}]\cap\mathcal{X}=\emptyset.

Minor modifications of the proofs yield the same result for structures with LSDAP+. ∎

6.2. Topological Ramsey spaces of homogeneous structures

Theorem 6.4.

Let 𝐊\mathbf{K} be any one of the following structures with universe β„•\mathbb{N}: The rationals, β„šn\mathbb{Q}_{n}, β„šβ„š\mathbb{Q}_{\mathbb{Q}}, and or any FraΓ―ssé  structure satisfying SDAP+ or LSDAP+ for which the coding tree of 11-types π•Œβ€‹(𝐊)\mathbb{U}(\mathbf{K}) has the property that on any given level of π•Œβ€‹(𝐊)\mathbb{U}(\mathbf{K}), only the coding node splits. Then the spaces π’Ÿβ€‹(𝔻)\mathcal{D}(\mathbb{D}), where 𝔻\mathbb{D} is a diagonal coding antichain for 𝐊\mathbf{K}, are actually topological Ramsey spaces.

Proof.

For structures as in the theorem statement, it is straightforward to check that Todorcevic’s Axiom A.3(2) holds. (This is the axiom which fails for the Rado graph and similar structures if one works with good diagonal antichains.) It is simple to check that Axioms A.1, A.2, and A.3(1) hold, and Axiom A.4 is a special case of Theorem 4.5 (these in fact hold for all structures considered in this paper). Then by Todorcevic’s Abstract Ellentuck Theorem in [16], the spaces π’Ÿβ€‹(𝔻)\mathcal{D}(\mathbb{D}) satisfy analogues of Ellentuck’s Theorem. ∎

6.3. Exact big Ramsey degrees from infinite-dimensional Ramsey theory

Let 𝔻\mathbb{D} be a good diagonal coding antichain for 𝐊\mathbf{K}, and let Mβˆˆπ’Ÿβ€‹(𝔻)M\in\mathcal{D}(\mathbb{D}). Given a finite antichain of coding nodes AβŠ†MA\subseteq M, let ⟨cjA:j<n⟩\langle c^{A}_{j}:j<n\rangle enumerate the coding nodes in AA and let 𝐀\mathbf{A} denote the structure πŠβ†ΎA\mathbf{K}\restriction A. Recall that we identify AA with the tree which it induces, and that 𝐀j\mathbf{A}_{j} denotes 𝐀↾{ciA:i<j}\mathbf{A}\restriction\{c^{A}_{i}:i<j\}. Let kk be least such that AβŠ†rk​(M)A\subseteq r_{k}(M). An envelope E​(A)E(A) of AA in MM is a minimal set of nodes in rk+1​(M)r_{k+1}(M) containing AA such that for each j<nj<n, the splitting predecessor of cjAc^{A}_{j} in MM is in E​(A)E(A), and each 11-type over 𝐀\mathbf{A} is represented by exactly one maximal node in E​(A)E(A).

Envelopes can be made canonically as follows: First, add all the splitting predecessors of coding nodes in AA and extend them β‰Ί\prec-leftmost in MM to length β„“nβˆ’1A+1\ell^{A}_{n-1}+1; let Aβ€²A^{\prime} denote this extension of AA. Then proceed by induction on j<nj<n: For each 11-type Ο„\tau over 𝐀1\mathbf{A}_{1} not already represented by a node in Aβ€²A^{\prime}, add one node tt in MM of length β„“0A+1\ell^{A}_{0}+1 such that t/𝐀0βˆΌΟ„t/\mathbf{A}_{0}\sim\tau; let E0E_{0} denote the set of these nodes of length β„“0A+1\ell^{A}_{0}+1. Whenever there is a choice of more than one node tt, add the β‰Ί\prec-leftmost such node. Given Ejβˆ’1E_{j-1} for 1≀j<n1\leq j<n, for each 11-type Ο„\tau over 𝐀j+1\mathbf{A}_{j+1} which is not represented by any node in Aβ€²β†Ύ(β„“jA+1)A^{\prime}\restriction(\ell^{A}_{j}+1), take the β‰Ί\prec-leftmost node ss in Ejβˆ’1βˆͺAβ€²β†Ύ(β„“jβˆ’1A+1)E_{j-1}\cup A^{\prime}\restriction(\ell^{A}_{j-1}+1) such that s/𝐀jβˆΌΟ„/𝐀js/\mathbf{A}_{j}\sim\tau/\mathbf{A}_{j}, and extend ss β‰Ί\prec-leftmost to a node tt in MM of length β„“jA+1\ell^{A}_{j}+1 such that t/𝐀j+1βˆΌΟ„t/\mathbf{A}_{j+1}\sim\tau. Let EjE_{j} denote the set of these nodes of length β„“jA+1\ell^{A}_{j}+1. Then, let E​(A)=Aβ€²βˆͺ⋃j<nEjE(A)=A^{\prime}\cup\bigcup_{j<n}E_{j}.

Notice that for each Mβˆˆπ’Ÿβ€‹(𝔻)M\in\mathcal{D}(\mathbb{D}), every finite antichain AA of coding nodes in MM has such an envelope in MM. Moreover, for any A,BβŠ†MA,B\subseteq M such that A∼BA\sim B, the canonical construction of envelopes produces envelopes E​(A)E(A) and E​(B)E(B) such that E​(A)∼E​(B)E(A)\sim E(B).

Now, given a good diagonal coding antichain 𝔻\mathbb{D} and a finite antichain AβŠ†π”»A\subseteq\mathbb{D} with nn coding nodes, let E​(A)E(A) be the canonical envelope of AA in 𝔻\mathbb{D}. Define 𝔼\mathbb{E} to be a good diagonal coding antichain contained in 𝔻\mathbb{D} such that 𝔼↾(β„“nβˆ’1A+1)=E​(A)\mathbb{E}\restriction(\ell^{A}_{n-1}+1)=E(A), and above E​(A)E(A), each 11-type over an initial structure of 𝔼\mathbb{E} is represented by exactly one node in 𝔼\mathbb{E}.

The following theorem of Coulson–Dobrinen–Patel in [3] is recovered as a Nash-Williams style corollary from the Main Theorem in this paper.

Corollary 6.5.

Let 𝐊\mathbf{K} be an enumerated FraΓ―ssé structure satisfying SDAP+ (or LSDAP+) with finitely many relations of arity at most two, and let 𝔻\mathbb{D} be a good diagonal coding antichain representing a copy of 𝐊\mathbf{K}. Let AβŠ†π”»A\subseteq\mathbb{D} be a finite diagonal antichain, and let ff color all similarity copies of AA in 𝔻\mathbb{D} into finitely many colors. Then there is a good diagonal coding antichain π”ΌβŠ†π”»\mathbb{E}\subseteq\mathbb{D} representing 𝐊\mathbf{K} in which all copes of AA have the same color.

Proof.

Let 𝔼\mathbb{E} be an end-extension of the envelope E​(A)E(A) in 𝔻\mathbb{D} to a good diagonal coding antichain, and let ff color all similarity copies of AA in 𝔼\mathbb{E} into finitely many colors. Let kk be the least integer such that rk​(𝔼)r_{k}(\mathbb{E}) contains AA. Notice that for each Mβˆˆπ’Ÿβ€‹(𝔼)M\in\mathcal{D}(\mathbb{E}), rk​(M)∼E​(A)r_{k}(M)\sim E(A), so the coding nodes in any Cβˆˆπ’œβ€‹π’Ÿk​(𝔼)C\in\mathcal{AD}_{k}(\mathbb{E}) induce a tree similar to AA; denote this tree by CAC_{A}. Moreover, for each similarity copy BB of AA in 𝐄\mathbf{E}, the canonical envelope E​(B)E(B) in 𝔼\mathbb{E} is in π’œβ€‹π’Ÿk​(𝔼)\mathcal{AD}_{k}(\mathbb{E}). Thus, ff induces a coloring gg on π’œβ€‹π’Ÿk​(𝔼)\mathcal{AD}_{k}(\mathbb{E}) by g​(C)=f​(CA)g(C)=f(C_{A}). This in turn induces an open, hence Borel, coloring hh on π’Ÿβ€‹(𝔼)\mathcal{D}(\mathbb{E}) via h​(M)=g​(rk​(M))h(M)=g(r_{k}(M)). By Theorem 6.3, there is an Nβˆˆπ’Ÿβ€‹(𝔼)N\in\mathcal{D}(\mathbb{E}) on which hh is constant. Thus, ff is constant on the similarity copies of AA in NN. ∎

Remark 6.6.

As pointed out in the introduction, the fact that the number of similarity types of diagonal antichains yields the exact big Ramsey degrees is a theorem of Coulson–Dobrinen-Patel in [3].

References

  • [1] Peter Cameron, Oligomorphic Permutation Groups, Cambridge University Press, 1990.
  • [2] Rebecca Coulson, Natasha Dobrinen, and Rehana Patel, FraΓ―ssΓ© classes with SDAP+, Part I: Indivisibility, (2021), 56 pp, Submitted. arXiv:2207.06393.
  • [3] by same author, FraΓ―ssΓ© structures with SDAP+, Part II: simply characterized big Ramsey structures, (2021), 59 pp, Submitted. arXiv:2207.06505.
  • [4] Natasha Dobrinen, Borel sets of Rado graphs and Ramsey’s theorem, European Journal of Combinatorics, Proceedings of the 2016 Prague DocCourse on Ramsey Theory, 29 pp, To appear. arXiv:1904.00266v1.
  • [5] by same author, Ramsey theory of the universal homogeneous k-clique-free graph, Journal of Mathematical Logic, 75 pp, To appear. arXiv:1901.06660.
  • [6] by same author, The Ramsey theory of the universal homogeneous triangle-free graph, Journal of Mathematical Logic 20 (2020), no.Β 2, 2050012, 75 pp.
  • [7] Erik Ellentuck, A new proof that analytic sets are Ramsey, Journal of Symbolic Logic 39 (1974), no.Β 1, 163–165.
  • [8] Fred Galvin and Karel Prikry, Borel sets and Ramsey’s Theorem, Journal of Symbolic Logic 38 (1973), no.Β 2, 193–198.
  • [9] Alexander Kechris, Vladimir Pestov, and Stevo Todorcevic, FraΓ―ssΓ© limits, Ramsey theory, and topological dynamics of automorphism groups, Geometric and Functional Analysis 15 (2005), no.Β 1, 106–189.
  • [10] Alex Kruckman, Disjoint nn-amalgamation and pseudofinite countably categorical theories, Notre Dame Journal of Formal Logic 60 (2019), no.Β 1, 139–160.
  • [11] Claude Laflamme, Norbert Sauer, and Vojkan Vuksanovic, Canonical partitions of universal structures, Combinatorica 26 (2006), no.Β 2, 183–205.
  • [12] C.Β St. J.Β A. Nash-Williams, On well-quasi-ordering transfinite sequences, Proceedings of the Cambridge Philosophical Society 61 (1965), 33–39.
  • [13] FrankΒ P. Ramsey, On a problem of formal logic, Proceedings of the London Mathematical Society 30 (1929), 264–296.
  • [14] Norbert Sauer, Coloring subgraphs of the Rado graph, Combinatorica 26 (2006), no.Β 2, 231–253.
  • [15] Jack Silver, Some applications of model theory in set theory, Annals of Mathematical Logic 3 (1971), no.Β 1, 45–110.
  • [16] Stevo Todorcevic, Introduction to Ramsey Spaces, Princeton University Press, 2010.
  • [17] Andy Zucker, A Note on Big Ramsey degrees, (2020), 21pp, Submitted. arXiv:2004.13162.