Ring constructions on spectral spaces

A thesis submitted to the University of Manchester for the degree of Doctor of Philosophy in the Faculty of Science and Engineering

by

Christopher Francis Tedd School of Mathematics

2016 2 Contents

Abstract5

Declaration6

Copyright statement7

Acknowledgements9

Dedication 11

Epigraph 13

Introduction 15 Summary of notation ...... 21 Summary of references ...... 23 Literature on the Hochster construction ...... 23

1. Preliminaries 25 The prime ...... 25 Spectral spaces ...... 28 The specialisation order ...... 38 The inverse space of a spectral space ...... 41 Spec A as a spectral space ...... 44 Valuations and valuation rings ...... 49

2. Hochster’s construction 51 Representing a spectral space X within the spectrum of a ring of functions on X ...... 51

The example of X = Spec Z ...... 59

3 Extensions and indices ...... 60 Representing spectral maps via ring homomorphisms ...... 69

3. Ring constructions on finite posets 73 Summary of techniques ...... 73 Intersecting (discrete) valuation rings ...... 73 Fibre sums of spaces ...... 74 The method of Lewis ...... 78 ‘Gluing’ of domains over maximal ideals ...... 79 ‘Joining’ of maximal ideals ...... 83 The iterative procedure ...... 86 Application to 1-dimensional Noetherian spectral spaces .... 92 The method of Ershov ...... 97

4. Comparison of constructions 107 Comparison of finite constructions ...... 107 Gluing of 1-dimensional domains ...... 112 Joining of maximal points ...... 117 Comparison of constructions on a general spectral space ...... 120 The comparison of valuation rings ...... 122 The comparison of admissibility conditions ...... 126  The example of X = Spec Z inv ...... 132

Bibliography 137

Word count 29,506

4 Abstract

Ring constructions on spectral spaces; a thesis submitted to the University of Manchester for the degree of Doctor of Philosophy by Christopher Francis Tedd, September 27, 2016.

In the paper [14] Hochster gave a topological characterisation of those spaces X which arise as the prime spectrum of a commutative ring: they are the spectral spaces, defined as those topological spaces which are T0, quasi-compact and sober, whose quasi-compact and open subsets form a basis for the topology and are closed under finite intersections. It is well known that the prime spec- trum of a ring is always spectral; Hochster proved the converse by describing a construction which, starting from such a space X, builds a ring having the desired prime spectrum; however the construction given is (in Hochster’s own words) very intricate, and has not been further exploited in the literature (a passing exception perhaps being the use of [14] Theorem 4 in the example on page 272 of [24]). In the finite setting, alternative constructions of a ring having a given spectrum are provided by Lewis [17] and Ershov [7], which, particularly in light of the work of Fontana in [8], appear to be more tractable, and at least more readily understood. This insight into the methods of constructing rings with a given spectrum is used to prove a result about which spaces may arise as the prime spectrum of a Noetherian ring: it is shown that every 1-dimensional Noetherian spectral space may be realised as the prime spectrum of a Noethe- rian ring. A close analysis of the two finite constructions considered here re- veals considerable similarities between their underlying operation, despite their radically different presentations. Furthermore, we generalise the framework of Ershov’s construction beyond the finite setting, finding that the ring we thus define on a space X contains the ring defined by Hochster’s construction on X as a subring. We find that in certain examples these rings coincide, but that in general the containment is proper, and that the spectrum of the ring provided by our generalised construction is not necessarily homeomorphic to our origi- nal space. We then offer an additional condition on our ring which may (—and indeed in certain examples does) serve to repair this disparity. It is hoped that the analysis of the constructions presented herein, and the demonstration of the heretofore unrecognised connections between the disparate ring constructions proposed by various authors, will facilitate further investigation into the prime ideal structure of commutative rings.

5 Declaration

No portion of the work referred to in this thesis has been submitted in sup- port of an application for another degree or qualification of this or any other university or other institute of learning.

6 Copyright statement

The author of this thesis (including any appendices and/or schedules to this thesis) owns certain copyright or related rights in it (the “Copyright”) and they have given The University of Manchester certain rights to use such Copyright, including for administrative purposes.

Copies of this thesis, either in full or in extracts and whether in hard or elec- tronic copy, may be made only in accordance with the Copyright, Designs and Patents Act 1988 (as amended) and regulations issued under it or, where appro- priate, in accordance with licensing agreements which the University has from time to time. This page must form part of any such copies made.

The ownership of certain Copyright, patents, designs, trade marks and other in- tellectual property (the “Intellectual Property”) and any reproductions of copy- right works in the thesis, for example graphs and tables (“Reproductions”), which may be described in this thesis, may not be owned by the author and may be owned by third parties. Such Intellectual Property and Reproductions cannot and must not be made available for use without the prior written permis- sion of the owner(s) of the relevant Intellectual Property and/or Reproductions.

Further information on the conditions under which disclosure, publication and commercialisation of this thesis, the Copyright and any Intellectual Property and/or Reproductions described in it may take place is available in the Uni- versity IP Policy (see http://documents.manchester.ac.uk/display.aspx?DocID= 24420), in any relevant Thesis restriction declarations deposited in the Univer- sity Library, The University Library’s regulations (see http://www.manchester

7 .ac.uk/library/about/regulations/) and in The University’s policy on Presenta- tion of Theses.

8 Acknowledgements

Thanks must first of all go to my supervisor Dr. Marcus Tressl, whose knowl- edge, insight, confidence and equanimity throughout have been central to the success of this project.

Thanks also to my parents for their unfailing financial and logistical support, without which none of this would have been possible.

And thanks for constant friendship to Sophie and to all my fellow Maths stu- dents who have made Manchester such a supportive place: Alex, Adam, Lyn- dsey, Nic, Siân, Rosie, Alex, Jonas, Goran, Jamie, Laura, Joe, Inga, Matthew and David, my office-mates and all past and present attendees of the Pure Post- graduate seminar, Friday pub, Hallowe’en and Christmas parties.

I would also like to thank my examiners Dr. Vincent Astier and Dr. Hendrik Süss for their diligence, attention to detail, and constructive commentary on the thesis.

9 10 Dedicated to the memory of

Gerard McCavana 01/10/1958—31/12/2013

&

Robert Ernest Charles Tedd 29/7/1927—01/04/2014

11 12 “craving peace its slow so slow drop into our laps but as far from it and myself as ever”

— Tom Paulin, The Road to Inver (after Pessoa)

13 14 Introduction

A X is spectral if it satisfies the following properties:

(1) X is T0;

(2) X is quasi-compact;

(3) X has a basis of sets which are quasi-compact and open;

(4) The quasi-compact open sets of X are closed under finite intersections; and

(5) X is sober — every non-empty irreducible of X is the closure of a unique point.

A map f : X → Y of spectral spaces is a spectral map if the preimage under f of all quasi-compact open sets of Y are quasi-compact and open in X. In the paper [14], Hochster showed that the spectral spaces are exactly the spaces that arise as the prime spectrum of some (commutative, unital) ring (— throughout the thesis a ring will always be commutative with 1), and that every spectral map of spectral spaces arises from a ring homomorphism under the prime spectrum functor, which takes a ring homomorphism ϕ : A → B to the map which sends a prime ideal of B to its preimage under ϕ. Indeed, it is easy to show that the prime spectrum of a ring is spectral, and that homomorphisms of rings induce spectral maps; Hochster demonstrated the converse by describing a construction which, starting from such a space X, builds a ring having the desired prime spectrum, in such a way that a spectral map between any two spectral spaces may be realised by a homomorphism of the constructed rings. This construction is the central object of study of this thesis, which, due to its intricate nature, is relatively unexplored and unexploited in the subsequent

15 16 Introduction

literature. It is hoped that the elucidation of this construction, and of the heretofore unrecognised connections between this and other ring constructions proposed by various authors ([17], [7]) will facilitate further investigation into the prime ideal structure of commutative rings. The first chapter compiles relevant standard results from the literature for easy reference. In particular some detail is gone into regarding the theory of spectral spaces, due to the fairly idiosyncratic behaviour of such spaces and the relative unfamiliarity of their properties in the context of general topology. Of particular relevance to the remainder of this introduction is the specialisa-

tion order defined on a spectral space X (—indeed, on any T0 space). The

specialisation order on a space X, denoted , is defined by

x y ⇐⇒ y ∈ {x}.

The dimension of a spectral space X is then the dimension of (X, ) as a poset.

This ordering is trivial on any T1 space, in particular any Hausdorff space, and on the space Spec A for a ring A this ordering is exactly the subset inclusion ordering on prime ideals considered as subsets of A. The references for this chapter are [1], [4] and [3], with the paper [14] also containing many of the stated results. Chapter 2 gives an exposition of Hochster’s construction of a ring having a given spectral space as its spectrum, as in the paper [14]. In particular no- tation is fixed which it is hoped will facilitate comparison of aspects of this construction with those presented elsewhere in the thesis. To summarise the construction: one may obtain a mapping from any topological space X into the prime spectrum of a ring of functions A on X in the case that each function f of A takes values in a domain at each x ∈ X: a point x ∈ X is mapped to the prime ideal of A consisting of those functions f ∈ A that are zero at x. When X is a spectral space we obtain four criteria which are necessary and sufficient for 17

Spec A to be homeomorphic to X: 1) we require that the range of the functions of A under evaluation at each point x ∈ X is a domain; 2) if the preimage of the basic open sets D(a) of Spec A are an open basis for X then X homeomor- phically embeds into Spec A; 3) if furthermore the preimage of these sets are quasi-compact, the inclusion map is a spectral map and X is called a spectral subspace of Spec A; the properties of spectral subspaces are then used to pro- duce a rather technical fourth and final criterion, see proposition 2.13, which if satisfied along with conditions (1)–(3) guarantees that X ∼= Spec A. We find it not too difficult to construct such a ring of functions AX on a spectral space X which satisfies the first three criteria: we form a ring AX consisting of polyno- mial expressions of characteristic functions χU for a collection of quasi-compact open sets {U}U∈B which form a subbasis for the topology on X. However, the ring so constructed may yet be a long way from satisfying the fourth condition: our ring AX may have many more prime ideals than there are points in the space X. Our procedure is to extend the ring AX , or, to extend the domains which are the range of the functions in AX evaluated on X towards their field of fractions, thereby in some sense ‘inverting’ certain functions in AX and thus killing off unwanted prime ideals; however care must be taken that the three properties satisfied by AX are not lost under this process of extension. Thus arises a procedure for determining in exactly what way the ring may be extended without the required properties being lost; it is hoped to give as clear an expo- sition of this procedure, which Hochster himself describes as “very intricate” in

[14], as possible. Finally, we obtain a ring HX satisfying all four properties, so ∼ that Spec HX = X; however it is by this stage not especially easy to determine exactly what ring this HX actually is, being the union of an inductive process of ring extensions defined via the ‘intricate procedure’ mentioned above. On the other hand, it is comparatively simple to show that every spectral map between 18 Introduction spectral spaces arises as the map induced from a homomorphism of rings having the appropriate spectra. This observation closes the chapter. Chapter 3 gives details of the constructions of Lewis [17] and Ershov [7] of a ring having prime spectrum homeomorphic to a given finite spectral space. Again complete details are given in order to facilitate analysis and compari- son of the various procedures used. Lewis’s construction employs an inductive argument to iteratively build up a ring realising a given space via two tech- niques: the first technique adds a new layer of maximal ideals to a ring that is assumed to exist by induction, and the second technique “joins” points in the spectrum of the ring obtained which may have been inadvertently multi- plicated as a consequence of the first technique. Fontana, in [8], observed that the two techniques used in Lewis’s construction are both instances of taking the amalgamated or fibre sum of topological spaces, via its connection under certain hypotheses to the fibre product of rings having spectra homeomorphic to the spaces in question. That is, given spectral spaces and maps as in the following diagram:

X

f

? Y - Z, g one may form the topological fibre sum W = X tZ Y ; this is a “gluing” con- struction that in some precise way amalgamates the topological structure of the spaces X and Y , see page 75. Then, under certain conditions, if we already have rings and ring homomorphisms ϕ : A → C, ψ : B → C realising f : X → Z and g : Y → Z, one may construct a ring realising W by taking the fibre product

A ×C B over those homomorphisms. Phrasing the construction in the language of topological fibre sums not only simplifies the presentation of Lewis’s method, 19 but also allows us to extend the methods employed to demonstrate that every 1-dimensional Noetherian spectral space arises as the spectrum of a Noetherian ring; whilst it is expected that this fact is known, the author could not find an explicit reference to it in the literature, and it provides an example of how the fibre sum methodology may be used to control properties of rings having a certain given spectrum. It is worth noting that a finite space may arise as the prime spectrum of a Noetherian ring only in the case that the dimension of X is at most 1, see 1.41. The construction of Ershov proceeds initially along similar lines to that of Hochster. For a finite spectral space X, we may find a subbasis for the topology on X that is in bijection with the points of X. The value at a point z ∈ X of the characteristic function of a basis set corresponding to a point x ∈ X is then determined by whether z lies below x in the specialisation order on X. Ershov then corresponds to each point x ∈ X a ring Rx↑ , being the intersection of certain discrete valuation rings defined from the specialisation relations involving x in X; these are the domains in which our functions on X take their values at the point x. A ring RX is then defined as those functions on X satisfying a compatibility condition: the values taken by f at points x and y in X must agree under the quotient maps associated to the valuations ∼ rings involved. The ring RX then has the property that Spec RX = X. Chapter 4 then proceeds to compare the constructions we have set out. Similarities are immediately found between our two finite constructions: we find that the elementary building blocks of Lewis’s construction can also be viewed as intersections of discrete valuation rings arising in exactly the manner of Ershov’s construction, except under the choice of an alternative subbasis for the topology of a finite spectral space. Furthermore, a strong similarity is found between the compatibility conditions by which Ershov picks out a ring RX 20 Introduction

having spectrum homeomorphic to a finite spectral space X, and the conditions arising in the definition of the fibre product over the rings defined on the parts of the space X that would be “amalgamated” in Lewis’s methodology. However, whilst the topological consequences of taking either the ring RX , or taking the fibre product as described, are identical (—we still obtain a ring whose spectrum is homeomorphic to X), the Ershov ring RX is strictly smaller than the ring arising from the fibre product. The second half of the chapter then proceeds from the similarity noted be-

tween the ring RX defined by Ershov’s construction and the ring AX defined at the initial stage of Hochster’s construction. In the finite case, our ring RX was defined from a subbasis for X that was in bijection with the points of X; we find it is not at all difficult to generalise the definitions of our finite construction to start from any subbasis for X, thereby obtaining a ring RX corresponding to any spectral space X. We then find a remarkable correspondence between the objects and conditions present in Hochster’s construction and those present in this generalised Ershov construction. Firstly, Hochster defines a discrete val- uation v to each relation x y that holds in the specialisation order on x y X; the generalised Ershov construction defines a discrete valuation ring V x y corresponding to each such relation. We find 4.2 that V is the valuation ring x y of the discrete valuation v . Hochster then uses these valuations to impose x y conditions on which ring extensions are permitted in his construction; we find 4.3 that elements satisfying the condition of Hochster to be admitted to a ring extension always satisfy the condition of Ershov to be admitted to the general ring RX ; that is, where HX is the union of all ring extensions permitted by

Hochster’s construction, we have that HX ⊆ RX . Indeed we find in a simple example that the rings HX and RX in fact coincide, but we also provide an ex- ample demonstrating that this does not hold in general, and furthermore (whilst Summary of notation 21

∼ ∼ we always have that Spec HX = X) we may not have that Spec RX = X for RX the general Ershov ring defined on a general spectral space X. This prompts us to investigate in two directions: on one hand, to consider subclasses of spectral ∼ spaces for which indeed we have Spec RX = X, and on the other hand to con- sider conditions which may be applied to RX to guarantee isomorphism with the ring HX (and thus ensuring that for the ensuing ring RbX we indeed have ∼ Spec RbX = X for all spectral spaces X). Conjectural hypotheses are discussed; however due to the short time between the discovery of these lines of inquiry and the necessary completion date of this thesis, this concludes the extent of the concrete progress that has been able to be made.

Summary of notation

The symbol ⊆ is used for subset inclusion; we use the symbol ( to denote strict inclusion. Where 6 denotes an order relation the symbol < is reserved for strict ordering, i.e. x < y ⇐⇒ x 6 y & x 6= y. ↑ If 6 is an order relation on a set X, then given Y ⊆ X we use Y to denote the set

↑ Y = {x ∈ X | y 6 x for some y ∈ Y } and Y ↓ to denote the set

↓ Y = {x ∈ X | x 6 y for some y ∈ Y }.

In particular given some element y ∈ X we use y↑ for

↑ y = {x ∈ X | y 6 x}

and y  for

↑ y  = {x ∈ X | y < x} = y \{y}; 22 Introduction

likewise we have

↓ y = {x ∈ X | x 6 y}

and

↓ y  = {x ∈ X | x < y} = y \{y}.

Given a poset X, the height of an element x ∈ X is the maximum n such

that there is a chain x0 < x1 < ··· < xn−1 < xn = x in X, or ∞ where no such n exists (—note that minimal elements in X have height 0); we denote the height of x ∈ X by ht x. The dimension of X, denoted dim X, is the maximum height of an element of X, or equivalently is the maximum length of a chain of distinct elements in X (under the same convention that a chain consisting of n + 1 distinct elements has length n), again with dim X = ∞ if no such maximum exists. Given sets X and Y then X t Y denotes the disjoint union of X and Y . Given a subset Y ⊆ X then Y c = X \ Y denotes the complement of Y in X. Given a ring A, we use the notation A× for the set of units of A (in particular if A is a field then A× = A \{0}). Given a subset S ⊆ A, then hSi denotes the ideal generated by S; in particular, for an element a of A then hai denotes the ideal generated by a. We use the notation Q(A) to denote the total ring of quotients of A, the localisation of A at the set of non-zero-divisors of A; when A is an integral domain then Q(A) is the field of fractions or quotient field of A. Throughout a ring is commutative with 1 and ring homomorphisms preserve the identity. Given a topological space X and a subset Y ⊆ X, then Y denotes the closure of Y in X. All notation not specified herein is either defined in the body of the text, or is hoped to be sufficiently standard as to be intelligible without ambiguity. Summary of references 23

Summary of references

In addition to the textbooks [1], [3], [4] as mentioned above, the author relied on the excellent textbooks [15] and [10] as standard references on commutative algebra, with [19] providing alternative reference, and used [20] as a reference for basic general topology, with [11] providing a reference on the basic properties of Noetherian spaces.

Literature on the Hochster construction

Whilst frequent reference to Hochster’s result is found throughout the sub- sequent literature, very few authors have engaged directly with the techniques of its construction, and the insights such an analysis may provide; an exception, notable for its rarity, being the use of [14] Theorem 4 in the example on page 272 of [24]; most often, the result of Hochster is used to prove a result of the form “there exists a ring having spectrum with certain properties”, proved by producing a spectral space with such properties and then invoking Hochster’s Theorem 6 to assert the existence of such a ring, without addressing any ques- tions of what sort of a ring that provided by the construction which proves Theorem 6 might be. Of examples where the “inner workings” of Hochster’s method are actually referred to explicitly, the only examples this author is aware of are the appendix to the textbook [16], and the paper [2]. [16] is a re-presentation (in French) of the method of Hochster furnished with complete proofs and some small examples; it is in most regards identical to the presentation given in [14]. [2] uses the techniques of Hochster to prove constructive results concerning the class of lattices which may arise as the lattice of radical ideals of a ring; the nuances of constructivity aside (—which are of course the main concern of the paper in 24 Introduction question), the paper again essentially reproduces the method of [14] unaltered, and thereby doesn’t really furnish the reader with any greater insight as to why the construction works as it does. It is hoped the detailed analysis presented in this thesis may expose overlooked aspects of Hochster’s results that could find some use in future work. 1. Preliminaries

The prime spectrum of a ring

The set of prime ideals of a ring A can be considered as the points of a topological space, the prime spectrum of A, denoted

Spec A = {p | p is a prime ideal of A}, with topology (named the ) having closed sets defined by subsets S ⊆ A as follows:

V(S) = {p ∈ Spec A | S ⊆ p}, by analogy (— as will become evident in due course) with the “algebraic sets” defined as the common zero-locus of sets of polynomials in classical algebraic geometry. Given an element a ∈ A, we write V(a) for V({a}) = {p | {a} ⊆ p} = {p | a ∈ p}. It is common (— and equivalent) to see the topology defined as having closed sets V(I) = {p ∈ Spec A | I ⊆ p} for ideals I ⊆ A, a convention we will adopt from now on. We begin by reviewing (without proof) relevant facts about this collection of closed sets, starting with the verification that it does in fact define a topology. A standard reference for facts concerning the prime spectrum of a ring is the textbook [1].

1.1. For the subsets V(I) ⊆ Spec A of the prime spectrum of a ring A as defined above, we have

(1) V(A) = ∅;

25 26 1. Preliminaries

(2) V(h0i) = Spec A;

T P  (3) V(Iλ) = V Iλ ; λ∈Λ λ∈Λ (4) V(I) ∪ V(J) = V(IJ).

1.2. Given ideals I,J of A, we have I ⊆ J =⇒ I ⊆ rad J ⇐⇒ V(J) ⊆ V(I). In particular, if I and J are radical ideals, then I ⊆ J ⇐⇒ V(J) ⊆ V(I).

Then any closed set C ⊆ Spec A in the Zariski topology corresponds to a unique radical ideal I of A so that C = V(I); we use the notation I = I(C) to specify such a radical ideal. We may in a very natural way define Spec as a (contravariant) functor from the category of rings and ring homomorphisms to that of topological spaces and continuous functions.

1.3. The function Spec ϕ : Spec B → Spec A defined from a ring homomorphism ϕ : A → B by (Spec ϕ)(q) = ϕ−1(q) is a continuous function with respect to the Zariski topology on Spec B, Spec A.

ϕ ψ 1.4. Given homomorphisms of rings A - B - C then Spec(ψ ◦ ϕ) = Spec ϕ ◦ Spec ψ.

We observe an important property of the continuous functions between spec- tra which arise from surjective ring homomorphisms.

1.5. Given a surjective ring homomorphism ϕ : A  B then the induced map Spec ϕ : Spec B,→ Spec A is a closed embedding of topological spaces.

In fact Spec ϕ Spec B = V(ker ϕ). We will return to examine further topological properties of the prime spec- trum in due course. First, however, we introduce an important shift in perspec- tive (— providing the promised analogy with classical algebraic geometry) that The prime spectrum of a ring 27 will recur throughout this investigation: that is, to consider our original ring

A as a ring of functions on its spectrum. We define the value fp of an element f ∈ A at a prime ideal p ∈ Spec A as the residue class of f in the ring A/p; that is, fp = [f]A/p ∈ A/p.

Then, since [f]A/p = 0 ⇐⇒ f ∈ p, we see, for a subset S ⊆ A,

V(S) = {p ∈ Spec A | fp = 0 for all f ∈ S}.

Y We may always write a function on Spec A in terms of its graph, f = fp. p∈Spec A In such a way, we obtain a mapping of elements of A into the ring of all such functions taking values in the domain A/p at p ∈ Spec A, that is, we obtain a mapping Y ϕ : A → A/p, p∈Spec A Y Y defined by f 7→ fp = [f]A/p. It is natural to ask when do distinct p∈Spec A p∈Spec A elements of A actually define distinct functions on Spec A, that is, when does this “ring of functions” viewpoint faithfully represent the original ring A. In fact,

Y Y 1.6. The mapping ϕ : A → A/p given by f 7→ fp is injective if p∈Spec A p∈Spec A and only if A is a reduced ring, that is, nil A = {0}.

We see that given a space X = Spec A the prime spectrum of some ring A, there is in fact a reduced ring A0 such that Spec A0 ∼= X.

1.7. Given any ring A, then Spec A/ nil A ∼= Spec A.

We now proceed to investigate the prime spectrum of a ring as a topological space. Our topology was defined by stating which sets were to be considered closed; we now turn our attention to the open sets of the Zariski topology. 28 1. Preliminaries

Given a ring A, then the open sets of Spec A are sets of the form

D(I) = Spec A \V(I) = Spec A \{p | I ⊆ p}

= {p | I 6⊆ p} for I ⊆ A. We define a principal given by an element a ∈ A to be the set

D(a) = {p | hai 6⊆ p} = {p | a∈ / p}.

These sets turn out to be of central importance to the topological structure of Spec A.

1.8. The collection of principal open sets {D(a)}a∈A forms an open basis for the Zariski topology on Spec A.

1.9. The principal open sets D(a) are quasi-compact.

1.10. Spec A is quasi-compact, since D(1) = Spec A.

1.11. Given any a, b ∈ A, then D(a) ∩ D(b) = D(ab).

In summary, we have demonstrated that the prime spectrum of a ring is a quasi- with a basis of quasi-compact open sets which is closed under finite intersections. We continue, considering a class of topological spaces possessing similar properties.

Spectral spaces

We define that a topological space X is spectral if it satisfies the following properties: Spectral spaces 29

(1) X is T0;

(2) X is quasi-compact;

(3) X has a basis of sets which are quasi-compact and open;

(4) The quasi-compact open sets of X are closed under finite intersections; and

(5) X is sober — every non-empty irreducible closed set of X is the closure of a unique point.

Our standard reference for facts about spectral spaces is the forthcoming book [4]; however many results described herein are contained in e.g. [14] and elsewhere. Of the four properties defining a spectral space, the least familiar is the latter; it in a certain sense guarantees that the topology exactly knows what set of points it is defined on. We remark that the requirement of uniqueness in the statement of property (5) actually implies that X is T0: since point-closures are always already irreducible, then the condition of uniqueness requires that distinct points have distinct closures, which is exactly equivalent to T0. Then in a distinct points are distinguished by the topology. The remainder of the condition — that every irreducible closed set has a generic point — in some sense stipulates that there are sufficient points to distinguish properties manifest in the topology. We note some simple examples of classes of spectral spaces.

Proposition 1.12. A Noetherian topological space X is spectral ⇐⇒ X is (T0 and) sober

as every subset of a Noetherian space X is quasi-compact, and so conditions (2)–(4) are automatically satisfied. 30 1. Preliminaries

Proposition 1.13. A finite T0 space X is spectral.

Again every subset of a finite space is quasi-compact (—indeed finite spaces are Noetherian) and so conditions (1)–(4) are satisfied. We observe that an irreducible closed set C of a finite space X is the closure of a point. Since S S C is finite, then c∈C {c} is a closed set, and so C = c∈C {c}. Then as C

is irreducible we must have that C = {c0} for some c0. Then if X is T0 this point-closure is necessarily unique, hence X is sober and so hence spectral as required.

The properties (2)–(4) of spectral spaces relate to the quasi-compact open sets of X. In fact, in the presence of the first four conditions, the final property is also dependent on the behaviour of the quasi-compact open sets of X, as we will see. We introduce the notation

◦ K(X) = {U ⊆ X | U is quasi-compact open}.

Observe that the collection of quasi-compact open sets of a topological space is closed under finite unions, since a union of open sets is open and a finite union of quasi-compact sets is quasi-compact. When X is a spectral space, we ◦ have that the set K(X) is also closed under finite intersections; that is, it is a sublattice of the lattice of open sets of X. We note at this stage,

Proposition 1.14. If U is a collection of quasi-compact open sets of a topolog- ical space X containing a basis for X, then if U is closed under finite unions ◦ then U = K(X),

◦ since given a quasi-compact open set U ∈ K(X) then as U contains a basis then U can be written as a union of sets from U; but as U is quasi-compact then it is a finite such union, and so is contained in U. Spectral spaces 31

Given the central role of quasi-compact open sets in the behaviour of spectral spaces, we introduce the notion of a spectral map: A function of spectral spaces ◦ ◦ f : X → Y is spectral if for every U ∈ K(Y ) then we have f −1(U) ∈ K(X). Note

Proposition 1.15. A spectral map f : X → Y of spectral spaces is continuous,

◦ ◦ since K(Y ), K(X) are an open basis for Y , X respectively.

We will generally restrict our attention to the subcategory of the category Top of topological spaces and continuous maps which consists of spectral spaces and spectral maps. It is important to note that this is not a full subcategory, in the sense that there are continuous maps in Top between spectral spaces which are not spectral maps. We then define a spectral subspace of a spectral space X: A subset Y ⊆ X is a spectral subspace of X if

(1) Y is a spectral space in the inherited from X; and

(2) The inclusion map Y,→ X is a spectral map.

Then spectral subspaces are those such that both the subspace Y and the inclusion map Y,→ X are present in our subcategory consisting of spectral spaces and spectral maps. The behaviour of spectral maps and spectral subspaces, along with the fifth property of the definition of a spectral space, that of being sober, can be stated in terms of a finer topology defined from the quasi-compact open sets of a spectral space X. We first introduce the notation

◦ K(X) = {X \ U | U ∈ K(X)} for the collection of complements of quasi-compact open sets of X. We then define the patch or constructible topology on X to be the topology having 32 1. Preliminaries

◦ K(X) ∪ K(X) as an open subbasis; that is, the topology generated by declaring the quasi-compact open sets of X to be closed and open. We denote X with

the constructible topology by Xcon. We see

Proposition 1.16. The constructible topology on a spectral space X refines the original topology,

◦ since it contains the basis K(X).

Proposition 1.17. The constructible topology on a spectral space X is Haus- dorff.

◦ The original topology on X is T0, and the sets K(X) are a basis for this ◦ topology. Then, given distinct points x 6= y, there is a basis set U ∈ K(X) with either x ∈ U & y∈ / U, or vice-versa. But then X \ U ∈ K(X), and so U and

X \ U are disjoint open sets in Xcon with x ∈ U and y ∈ X \ U (or vice-versa), so Xcon is Hausdorff.

◦ Proposition 1.18. The sets of the form U ∩ V for U ∈ K(X) and V ∈ K(X) form an open basis for the constructible topology on a spectral space X.

Each set U ∩ V is certainly open in Xcon, being a finite intersection of subbasic ◦ ◦ open sets. Further as X ∈ K(X) and (since ∅ ∈ K(X),) X ∈ K(X), then this ◦ collection contains every set U ∈ K(X) and V ∈ K(X) in our subbasis in the form U ∩X or X ∩V . Then as a collection of open sets containing the subbasis, it will suffice to show it is closed under finite intersections. As X is spectral we ◦ ◦ know K(X) is closed under finite intersections; furthermore as K(X) is closed under finite unions then K(X) is closed under finite intersections. Then, given ◦ Tn  a finite intersection i Ui ∩ Vi with Ui ∈ K(X) and Vi ∈ K(X), we have n n n \  \ \ Ui ∩ Vi = Ui ∩ Vi, i i i ◦ Tn Tn with i Ui ∈ K(X) and i Vi ∈ K(X) as required. Spectral spaces 33

◦ Corollary 1.19. The sets of the form U ∪ V for U ∈ K(X) and V ∈ K(X) form a closed basis for the constructible topology on a spectral space X,

◦ since given an open basic set U 0 ∩ V 0 with U 0 ∈ K(X) and V 0 ∈ K(X), then X \ (U 0 ∩ V 0) = (X \ U 0) ∪ (X \ V 0) = V ∪ U where V = (X \ U 0) ∈ K(X) and ◦ U = (X \ V 0) ∈ K(X).

The most significant property of the constructible topology on a spectral space X is that Xcon is itself quasi-compact. In fact,

Proposition 1.20. If X is a topological space satisfying properties (1)–(4) in the definition of a spectral space, then X is sober (that is, X is in fact a spectral space) ⇐⇒ Xcon is quasi-compact.

Let X be a topological space satisfying properties (1)–(4) in the definition of a ◦ spectral space; that is, X is T0, quasi-compact, and the collection K(X) contains a basis and is closed under finite intersections. Observe that the conclusions of the preceding three propositions and the corollary still hold for such a space

X, as their proofs (at most) use only the fact that X is T0 rather than the full sobriety property. We relate sets that are closed and irreducible in the original topology on X to collections of subsets closed in the constructible topology on X having the finite intersection property. Note that a closed subset C ⊆ X is irreducible iff any collection of open sets which intersect C has non-empty finite intersections within C. Indeed, it suffices to consider basic open sets; then, C is irreducible if, given any collection UΛ = ◦ {Uλ | Uλ ∈ K(X) and Uλ ∩ C 6= ∅}, we have that for any finite subcollection Tn  Uλ1 ,...,Uλn we have i Uλi ∩ C 6= ∅.

Suppose Xcon is quasi-compact, and take C a set closed and irreducible in the ◦ original topology on X. We consider the collection UC = {U ∈ K(X) | U ∩ C 6= ∅}. As C is irreducible, this has the finite intersection property. Furthermore, it is 34 1. Preliminaries

a collection of basic closed sets for the constructible topology. Then, if Xcon is T T quasi-compact, we must have UC 6= ∅. Take x ∈ UC . Then x is contained in every basic open set (for the original topology) that intersects C, that is,

C = {x}, and so (—since X is already T0, and so x is uniquely specified,) X is sober.

Alternatively let C be a collection of closed basic subsets of Xcon (—that ◦ is, sets of the form U ∪ V for U ∈ K(X) and V ∈ K(X)) having the finite intersection property. We see that chains of such collections (that is, of closed

basic subsets of Xcon having the finite intersection property) satisfy the condition of Zorn’s lemma, since the union of such a chain retains the finite intersection property, as any finite collection of sets from the union of the chain must be already present in some set in the chain. Let M be a maximal such collection containing C. Then define V = {V ∈ M | V ∈ K(X)} and let C = T V. We see that ◦ ◦ {U ∈ K(X) | U ∈ M} = {U ∈ K(X) | U ∩ C 6= ∅}.

Since, if U/∈ M then M ∪ {U} must fail the finite intersection property,

c that is, we have M1,...,Mn ∈ M with M1 ∩ · · · ∩ Mn ∩ U = ∅. Then U ∈ M,

c c since M1 ∩ · · · ∩ Mn ∩ U = M1 ∩ · · · ∩ Mn, so for any A1,...,Am ∈ M \ {U } we see that

m m n c \  c \  \  A1 ∩ · · · ∩ Am ∩ U ⊇ Ai ∩ M1 ∩ · · · ∩ Mn ∩ U = Ai ∩ Mj 6= ∅,

and so adding U c preserves the finite intersection property. Furthermore U c ∈ K(X), so U c ∈ V; then C = T V ⊆ U c and so U ∩ C = ∅. Conversely, if U ∈ M then we consider the collection {U ∩ V | V ∈ V}. This is a collection of relatively-closed (—per the original topology) subsets in the quasi-compact subset U; furthermore it has the finite intersection property, since U and each V are in M, and so a finite intersection from this collection Spectral spaces 35 is a finite intersection of sets from M. Then, by quasi-compactness of U this collection must have non-empty total intersection, that is, U ∩ T V 6= ∅, i.e. U ∩ C 6= ∅ as required. We show C is irreducible. We have seen that the set of basic open sets having non-empty intersection with C is exactly the set of basic open sets present in our collection M; then given a finite collection U1,...,Un of such basic open sets, we have that U1∩· · ·∩Un ∈ M, since adding this set preserves the finite intersection property of M since each constituent set is already present in M. Then as ◦  U1 ∩· · ·∩Un is in K(X), we have U1 ∩· · ·∩Un ∈ M =⇒ U1 ∩· · ·∩Un ∩C 6= ∅; so C is irreducible as required. Then if X is sober, we have that C = {x} for some point x ∈ X. We show that x ∈ T M. M is a collection of basic closed sets for the constructible topology on X, that is, it is a collection of sets of the ◦ form U ∪ V for U ∈ K(X) and V ∈ K(X). In fact we see, given such a set U ∪ V ∈ M, then we have either U ∈ M or V ∈ M. Since, if neither were Tm  in M, by maximality we must have finite intersections Ai ∩ U = ∅ and Tn  Tm  Tn  Bj ∩ V = ∅ for Ai,Bj ∈ M. But then Ai ∩ Bj ∩ (U ∪ V ) = ∅, contradicting U ∪ V ∈ M. Then x ∈ U ∪ V , since, if U ∈ M then U ∩ C 6= ∅, and so (since U is open) x ∈ U, and if V ∈ M then C ⊆ V by definition. Then x ∈ T M and so certainly x ∈ T C, and so we have shown that collections of closed basic subsets of Xcon having the finite intersection property have non- empty total intersection, that is, Xcon is quasi-compact.

We now see how the behaviour of spectral maps relates to the constructible topology on a spectral space.

Proposition 1.21. A map f : X → Y of spectral spaces X and Y is a spec- tral map ⇐⇒ f is continuous and is a continuous map with respect to the constructible topologies on X and Y .

We know a spectral map f : X → Y is continuous; furthermore since for a 36 1. Preliminaries

◦ ◦ spectral map we have f −1(K(Y )) ⊆ K(X) then likewise f −1(K(Y )) ⊆ K(X)

and so the preimage of a subbasic set for Ycon is in the subbasis for Xcon, so f is continuous as a map Xcon → Ycon. Conversely, to show f is a spectral map we need to show that the preimage of a quasi-compact open set U of Y is quasi-compact open in X. If f is continuous then the preimage of U is open. If furthermore f is a continuous map between the constructible topologies on X and Y , then as U is a closed set in Ycon then

−1 the preimage of U is closed in Xcon. Then, as Xcon is quasi-compact, f (U) is quasi-compact in Xcon, and as the original topology on X is a coarsening of the constructible topology then f −1(U) is quasi-compact in the original topology on X as required.

Subsets of a spectral space X that are closed in the constructible topology have some remarkable properties. Firstly,

Proposition 1.22. If Y ⊆ X is closed in the constructible topology of X, then we have x ∈ Y ⇐⇒ x ∈ {y} for some y ∈ Y .

Given some x ∈ Y we consider the collection of basic open neighbourhoods ◦ of x, Ux = {U ∈ K(X) | x ∈ U}. As x is in the closure of Y , then every

open neighbourhood of x intersects Y ; so the collection Ux ∪ {Y } has the finite

intersection property (since a finite intersection of sets from Ux is again an open neighbourhood of x, and so intersects Y ). Furthermore, each set in this collec-

tion is closed in the constructible topology. Then, as Xcon is quasi-compact, this collection has non-empty total intersection, that is there is an element y ∈ Y contained in every basic open neighbourhood of x; then x ∈ {y} as required.

In fact the spectral subspaces of a spectral space X are exactly the subsets closed in the constructible topology. Spectral spaces 37

Proposition 1.23. A subset Y of a spectral space X is a spectral subspace of X ⇐⇒ Y is a closed set in the constructible topology on X.

Suppose Y is a spectral subspace of X. Then Y is a spectral space in the inherited topology, and inclusion Y,→ X is a spectral map. As Y is a spectral space, we have that Ycon is quasi-compact. As inclusion is a spectral map, then it is continuous between Ycon → Xcon; then, being the image of a quasi-compact set under a continuous map, Y is quasi-compact in Xcon. Then finally as Xcon is Hausdorff, Y is closed in Xcon as required. ◦ ◦ Conversely suppose Y is closed in Xcon. We find K(Y ) = {U∩Y | U ∈ K(X)}. ◦ As Y and each U ∈ K(X) are closed in Xcon then U ∩ Y is closed, and hence quasi-compact, in Xcon (—since Xcon is quasi-compact), and so also in X (since the original topology on X is a coarsening of the constructible topology), and then also in Y (since quasi-compactness is inherited by subsets of Y in the subspace topology). Furthermore this collection is a basis for the inherited topology on Y and is closed under finite unions, so it accounts for all quasi- ◦ compact open sets in Y . Clearly K(Y ) is also closed under finite intersections; then Y will be a spectral space in the subspace topology iff it is sober. Since ◦ ◦ K(Y ) = {U ∩ Y | U ∈ K(X)}, and so also K(Y ) = {V ∩ Y | V ∈ K(X)}, it is clear that the constructible topology defined on Y as a spectral space (i.e. that ◦ generated by K(Y ) ∪ K(Y )) is exactly the same as the subspace topology inher- ited by Y as a subset of Xcon. As Y is closed in Xcon then it is quasi-compact in Xcon, so Ycon is quasi-compact (as quasi-compactness is inherited) and hence ◦ ◦ Y is sober as required. The fact that K(Y ) = {U ∩ Y | U ∈ K(X)} shows that preimages of quasi-compact open sets of X are quasi-compact open in Y , hence inclusion is a spectral map, and so Y is a spectral subspace.

We take this opportunity to investigate some further topological notions which will become relevant in later discussions. 38 1. Preliminaries

The specialisation order

We have seen in various places in the preceding section properties of spectral spaces that relate to point-closures, that is, to closures of singleton sets in the spectral topology. We introduce an order relation on a topological space that encapsulates this notion. Given a topological space X, we define the specialisation order on X, denoted

, by x y ⇐⇒ y ∈ {x}.

Where we have x y we say y is a specialisation of x and x is a generisation of y; x specialises to y and y generalises to x. We fix notation as follows:

↑ x = {y | x y}, the set of all specialisations of x;

↓ x = {z | z x}, the set of all generisations of x.

Clearly by the definition we have x↑ = {x}.

In general, is a pre-order, that is, it is reflexive and transitive, since closure in a topological space is reflexive and transitive, i.e.

•∀ x ∈ X, x ∈ {x}, so ∀x ∈ X, x x, and

• y ∈ {x} & z ∈ {y} =⇒ z ∈ {x}, so x y & y z =⇒ x z.

We see

Proposition 1.24. The specialisation order on a topological space X is a partial order (—i.e., it is anti-symmetric) ⇐⇒ X satisfies the T0 separation property.

Anti-symmetry of is the statement that x y & y x =⇒ x = y, which translates via the definition to y ∈ {x} & x ∈ {y} =⇒ x = y. Taking the Spectral spaces 39 contrapositive, we obtain x 6= y =⇒ y∈ / {x} or x∈ / {y}, i.e. distinct points are distinguished by some closed set; equivalently by some open set, which is the T0 property.

Relating to the specialisation order on a topological space is the notion of an order-compatible topology on a poset. Given a poset (X, 6), a topology τ on X is order-compatible if the specialisation order on (X, τ) is 6. Note that by stipulating that 6 is a partial order we have that any order-compatible topology will satisfy the T0 property, by the preceding observation. A comprehensive overview of order-compatible topology as it pertains to spectral spaces is given in the paper [22]; see also [23], [18], [5], [21]. It is easy to see that

Proposition 1.25. Given a set X, and a partial order 6 and a topology τ on X, we have τ is order-compatible (with 6) exactly when:

↑ (1) For each x ∈ X, the set x = {y | x 6 y} is a closed set in τ, and

(2) For each closed set C in τ, x↑ ⊆ C for each x ∈ C.

We note as an immediate consequence of this observation that if we are given a poset (X, 6), then the order-compatible topologies on X form an interval in the collection of all topologies on X: the coarsest possible order-compatible topology is that given by taking the collection {x↑ | x ∈ X} as a closed sub- basis; we call this topology the coarse lower topology or sometimes the closures- of-points or COP topology, and the finest topology is that given by allowing any set C containing x↑ for every x ∈ C (i.e. all upper sets C = C↑) to be closed; in which case we obtain exactly the lower sets U = U ↓ as the open sets, hence we can think of this as the topology having the collection {x↓ | x ∈ X} as an open basis; we call this topology the fine lower topology, or in some references 40 1. Preliminaries

l the left topology. Given a poset (X, 6), we write X for the space given by X with the coarse lower topology, and XL for X with the fine lower topology. Note that when X is finite then we have that all up-closed sets C ⊆ X (i.e. all closed sets of XL) are of the form S↑ for some finite set S ⊆ X (we can take the finite set S = C), that is, C is closed in Xl, hence XL = Xl and we may refer to this unique order-compatible topology as the lower topology on a finite poset X. We note that by 1.13, as an order-compatible topology on a poset is T0, then every finite poset (X, 6) arises as the specialisation order of a spectral space, namely the finite T0 space consisting of X equipped with the lower topology. We now see that the specialisation order exactly characterises the Noethe- rian spectral spaces; per 1.12 these are the sober Noetherian spaces. Our first observation in this direction is that Noetherian spectral spaces are uniquely determined (up to ) by their specialisation order; that is,

∼ Proposition 1.26. Given Noetherian spectral spaces X,Y where (X, X ) = ∼ (Y, Y ) as posets, then X = Y as topological spaces.

It is a well known property of Noetherian spaces that every closed set of a Noetherian space X is a finite union of irreducible closed sets. In the case that X is spectral, and so sober, then every irreducible closed set is the closure of a point, and so every closed set C of X is a finite union C = {x1} ∪ · · · ∪ {xn} =

↑ ↑ x1 ∪ · · · ∪ xn (and of course every such finite union gives a closed set of X). That is, the topology on X is entirely specified by the partial order relation

X , and so two isomorphic specialisation posets necessarily define the same topology.

Note as a consequence of this observation that the topology on a Noetherian

spectral space X is always equal to the coarse lower topology on (X, X ). Spectral spaces 41

In fact there is a purely order-theoretic characterisation of exactly those posets which arise as the specialisation poset of a Noetherian spectral space, first given in [23]:

Theorem 1.27 ([23], discussion following Proposition 1). A poset (X, 6) is 0 0 isomorphic to (X , X0 ) for X a Noetherian spectral space if and only if:

(1) X has the ascending chain condition;

↑ (2) X has finitely many minimal elements {x1, . . . , xn} and {x1, . . . , xn} = X; and

(3) Every pair of elements x, y ∈ X has finitely many minimal upper bounds in X.

We remark only on the necessity of these conditions, that is, that if X is a

Noetherian spectral space then the poset (X, ) satisfies these properties. An ascending chain ... xi xi+1 xi+2 ... corresponds to a descending chain of closed sets ... ⊇ {xi} ⊇ {xi+1} ⊇ {xi+2} ⊇ ... ; thus if X is Noethe- rian the chain must stabilise. As in 1.26, the closed set X has finitely many

↑ ↑ irreducible components X = {x1} ∪ · · · ∪ {xn} = x1 ∪ · · · ∪ xn which correspond to points minimal in the specialisation order; likewise the closed set x↑ ∩ y↑ has

↑ ↑ ↑ ↑ finitely many irreducible components x ∩ y = z1 ∪ · · · ∪ zm, corresponding to the minimal common upper bounds of x and y in (X, ).

The inverse space of a spectral space

Given a spectral space X we use the collection K(X) to define an alternative topology on the set X: We define the inverse space Xinv of X to be the set X with topology (—the inverse or opposite-order topology) defined by taking the ◦ set B = K(X) = {X \ U | U ∈ K(X)} as an open subbasis for X. In fact as 42 1. Preliminaries

we have previously observed, B = K(X) is closed under finite intersections, so

is in fact a basis for Xinv. Remarkably, the space so defined is again a spectral space. We see

Proposition 1.28. Xinv is quasi-compact,

since the topology of Xinv is a coarsening of the topology of Xcon, which is quasi-compact.

Furthermore,

Proposition 1.29. Xinv has a basis of quasi-compact open sets, closed under finite intersections,

since firstly B = K(X) is closed under finite intersections; furthermore each

set in K(X) is closed in Xcon, so is quasi-compact in Xcon and hence is quasi-

compact in Xinv (again since the topology of Xinv is a coarsening of the topology of Xcon).

In fact

◦ Proposition 1.30. K(Xinv) = B = K(X),

that is our basis B accounts for all quasi-compact open sets in Xinv: since, as ◦ K(X) is closed under finite intersections then B = K(X) is closed under finite unions. But then B is a basis of quasi-compact open sets which is closed under finite unions, therefore by 1.14 it accounts for all quasi-compact open sets in ◦ Xinv, that is, K(Xinv) = B.

◦ Corollary 1.31. K(Xinv) = K(X),

◦ since by definition, K(Xinv) is the set of complements {X \ U | U ∈ K(Xinv)} of quasi-compact open sets of Xinv. But then, by the preceding proposition we Spectral spaces 43 have

◦ K(Xinv) = {X \ U | U ∈ K(Xinv)}

= {X \ U | U ∈ K(X)} ◦ = K(X).

This allows us to observe,

Proposition 1.32.( Xinv)inv = X,

since (Xinv)inv is the space defined as having K(Xinv) as an open subbasis. But ◦ K(Xinv) = K(X) is a basis for the original topology on X, so (Xinv)inv = X.

Furthermore,

Proposition 1.33.( Xinv)con = Xcon,

◦ ◦ since K(Xinv) ∪ K(Xinv) = K(X) ∪ K(X).

And then finally, we have

Proposition 1.34. Xinv is a spectral space.

We have already noted that Xinv is a quasi-compact space with a basis of quasi- compact open sets closed under finite intersections. It is clearly T0, as we know ◦ distinct points in the set X are separated by some U ∈ K(X) (since the original topology on X is T0); and so they are likewise separated by X \ U ∈ K(X).

Furthermore (Xinv)con = Xcon is quasi-compact; so Xinv is sober, and so is hence a spectral space, as required.

We finally note how the inverse topology and the specialisation order on a spectral space X relate. 44 1. Preliminaries

Given an order relation 6, we may consider the opposite or dual order 6opp defined by

x 6opp y ⇐⇒ y 6 x.

Of course, 6opp is just the familiar order relation >.

Now, given a spectral space X, let X denote the specialisation order on

X, and consider the opposite order ( X )opp.

Then, considering the inverse space Xinv of X, and writing the specialisation

order on Xinv as Xinv , we find

Proposition 1.35. Xinv = ( X )opp,

that is, x Xinv y ⇐⇒ y X x.

Recall we have y X x ⇐⇒ x ∈ {y} ⇐⇒ every basic closed set of X T containing y contains x. Then y X x ⇐⇒ x ∈ {V ∈ K(X) | y ∈ V }. But considering this statement in Xinv, then as the sets V ∈ K(X) are an open basis for the inverse topology on X, this says that every basic open neighbourhood of y (in Xinv) contains x, which is exactly equivalent to y ∈ {x} in Xinv, i.e.

x Xinv y.

We now return to considering how the topological properties investigated here relate to the prime spectrum of a ring.

Spec A as a spectral space

As might be anticipated, we have

Proposition 1.36. The prime spectrum of a ring A is a spectral space.

We saw at the end of the first section that Spec A satisfies the second and Spec A as a spectral space 45 third properties of a spectral space via the basis D(A) of quasi-compact ‘prin- cipal open’ sets, that is that Spec A is quasi-compact, and has a basis of quasi- compact open sets. We also observed that the principal open sets were closed under finite intersection; note the definition of a spectral space requires that the collection of all quasi-compact open sets be closed under finite intersections. We can see that, as D(A) is a basis, then a general quasi-compact open set U in Spec A is a finite union of sets from D(A),

U = D(a1) ∪ · · · ∪ D(an)

for some a1, . . . , an ∈ A. Then given two such quasi-compact open sets U, V , we see that

  U ∩ V = D(a1) ∪ · · · ∪ D(an) ∩ D(b1) ∪ · · · ∪ D(bm) [  = D(ai) ∩ D(bj) i,j [  = D(aibj) , i,j i.e. it is a finite union of principal quasi-compact open sets, and so is quasi- compact open. To show Spec A is a spectral space, then, it remains to show that every non-empty irreducible closed set of Spec A is the closure of a unique point (—recall that uniqueness will then imply that Spec A is T0). Let ∅ 6= C ⊆ Spec A be an irreducible closed set. We consider I = I(C) the radical ideal of A such that C = V(I). We show that C being irreducible implies that I is in fact a prime ideal. Since, suppose we have ideals J, K of A with JK ⊆ I. Then (by 1.2 and 1.14) we have V(I) ⊆ V(JK) = V(J) ∪ V(K). As V(I) is irreducible we must have one of V(I) ⊆ V(J) or V(I) ⊆ V(K); then, as I is a radical ideal we have either J ⊆ I or K ⊆ I; this shows that I is a prime ideal of A. Then the set C = V(I) is the closure in Spec A of the point 46 1. Preliminaries

I, since for any closed set V(S) defined by some subset S ⊆ A, we have

I ∈ V(S) =⇒ S ⊆ I =⇒ V(I) ⊆ V(S), that is, V(I) is the smallest closed set of Spec A containing I. Then we have shown that every irreducible closed set of Spec A is the closure of a point; the required uniqueness is equivalent to distinct points having distinct closures, but given distinct prime ideals p and q in Spec A then (— since prime ideals are radical ideals) the sets {p} = V(p) and {q} = V(q) are distinct.

It is much harder to show—though no less true—that every spectral space arises as the prime spectrum of some ring, which shall be the topic of the following chapter. In the meantime, we investigate some further properties of the prime spec- trum of a ring, from the point of view of spectral spaces. Again the textbooks [4],[1] and in some instances the paper [14] provide standard references for such facts.

1.37. Every ring homomorphism ϕ : A → B induces a spectral map Spec ϕ : Spec B → Spec A.

We have already that Spec ϕ is continuous, from 1.3; we must show the preim- age of a quasi-compact open set in Spec A is quasi-compact open in Spec B. But this is true for the basis sets

Spec ϕ−1(D(a)) = {q ∈ Spec B | Spec ϕ(q) ∈ D(a)} = {q | a∈ / ϕ−1(q)}

= {q | ϕ(a) ∈/ q}

= D(ϕ(a)) and so is true for finite unions of such, which account for all quasi-compact open sets in Spec A. Spec A as a spectral space 47

1.38. The specialisation order on Spec A is the subset-inclusion relation between prime ideals as subsets of A, that is, p q ⇐⇒ p ⊆ q as subsets of A.

We have

0 0 p q ⇐⇒ q ∈ {p} = V(p) ⇐⇒ q ∈ {q ∈ Spec A | p ⊆ q } ⇐⇒ p ⊆ q.

We have observed that the quasi-compact open sets of Spec A are the finite unions of basic open sets D(a). We use this to note the following:

Tn  1.39. Given X = Spec A, sets of the form D(a)∩ i V(bi) for a, b1, . . . , bn ∈ A  form a basis for the space Xcon = Spec A con.

◦ From 1.18 we have that sets of the form U ∩ V for U ∈ K(X) and V ∈ K(X) form a basis for the constructible topology on a spectral space X. Given X = ◦ Sm Spec A, a general set in K(X) is of the form j D(aj) for some finite subset a1, . . . , am ∈ A, and a general set in K(X) is the complement of such a set, Sn c Tn c Tn i.e. is of the form i D(bi) = i D(bi) = i V(bi) for some finite subset b1, . . . , bn ∈ A. Then we have that sets of the form

m n [ \ D(aj) ∩ V(bi) j i for finite subsets a1, . . . , am, b1, . . . , bn ∈ A give a basis for the constructible topology on Spec A. But, as any set Y open in Xcon is a union of such sets, then T Y is also a union of the sets D(ai) ∩ b∈B V(b) for each ai; thus the collection of sets of this form also forms a basis for Xcon.

Sn 1.40. Given X = Spec A, sets of the form V(a) ∪ i D(bi) for a, b1, . . . , bn ∈ A  form a closed basis for the space Xcon = Spec A con

by taking complements in the preceding result. 48 1. Preliminaries

We remarked in the preceding section on the characterisation of Noetherian spectral spaces by their specialisation posets. Whilst we have that if A is a Noetherian ring, then Spec A is indeed a Noetherian spectral space, unfortu- nately it is not the case that every specialisation poset of a Noetherian spectral space can arise as the inclusion ordering of prime ideals of a Noetherian ring; that is, there are Noetherian spectral spaces X for which there is no Noetherian ring A such that Spec A ∼= X. Since, observe that any finite chain, for example

X = {x0 < x1 < x2}, satisfies the conditions of 1.27, such that the coarse lower topology on X is a Noetherian spectral space. However it is not hard to see

1.41. If A is a Noetherian ring of dimension at least 2, then Spec A is infinite.

Let p ⊆ A be a prime ideal of height 2. We write

[ p = hai. a∈p

The Krull principal ideal theorem (see, for example, [15]) states that a prime ideal qa in a Noetherian ring that is minimal with respect to the inclusion of a principal ideal hai ⊆ qa has height at most 1. For each a ∈ p let qa be a prime ideal minimal over a; then [ p ⊆ qa. a∈p If Spec A were finite then p would necessarily be contained in a finite union of height-1 primes; then by prime avoidance it must be contained in one of them, giving p ⊆ qa0 for some qa0 of height less than or equal to 1. But this contradicts that the height of p is at least 2; and so A must have infinitely many prime ideals.

Thus the poset X defined above cannot arise as the inclusion ordering of any Noetherian ring. Valuations and valuation rings 49

Valuations and valuation rings

The theory of (discrete) valuations on a field is centrally involved in many of the discussions in this thesis; we summarise without proof the basic properties and results of this theory. The standard reference for this section is [3]. A valuation on a field k is a map v : k → Γ ∪ {∞} for Γ an ordered group and some symbol ∞ ∈/ Γ defined as greater than every element of Γ (with the obvious extension of the group operation: ∞ + g = g + ∞ = ∞ for any g ∈ Γ), satisfying the following properties:

(1) v(α) = ∞ ⇐⇒ α = 0;

(2) v(αβ) = v(α) + v(β) for all α, β ∈ k;

(3) v(α + β) > min(v(a), v(b)) for all α, β ∈ k.

A valuation v : k → Γ ∪ {∞} on a field k is improper if v(α) = 0 for all α ∈ k×, v being proper otherwise; some authors exclude improper valuations from their definition. ∼ A valuation v : k → Γ∪ {∞} is a discrete valuation on k if Γ = Z; in practice all valuations considered in the remainder of this thesis will be discrete, however unless otherwise specified the results summarised in this section are valid for all valuations. We obtain the following identities:

1.42. Given a field k and v : k → Γ ∪ {∞} a valuation on k, then:

(1) v(1) = 0;

(2) v(α) = v(−α) for all α ∈ k;

(3) If v(α) 6= v(β) then v(α + β) = min(v(α), v(β)); 50 1. Preliminaries

Given a valuation v : k → Γ ∪ {∞} on a field k the valuation ring of v is the

ring V = {α ∈ k | v(α) > 0}. Generally, a ring R (—necessarily a domain) is a valuation ring if it is the valuation ring of some valuation v on Q(R). We note

1.43. A domain R is a valuation ring if for every α ∈ Q(R)×, we have either α ∈ R or α−1 ∈ R.

1.44. Given a valuation v : k → Γ ∪ {∞} on a field k, the valuation ring V is a local ring, having unique maximal ideal m = {α ∈ k | v(α) > 0}.

In fact,

1.45. The set of ideals of a valuation ring are totally-ordered by subset inclusion.

Then every valuation ring comes with a canonical quotient map π : R  R/m, and we refer to the field K = R/m as the quotient field of R. A valuation ring V of a valuation v is a discrete valuation ring if the valuation v is discrete.

1.46. A ring R is a discrete valuation ring if and only if R is a local Noetherian domain whose unique maximal ideal m is a principal ideal; equivalently iff R is a local principal ideal domain. In particular (at least in the case of a proper discrete valuation ring, so that m 6= 0), m is generated by any element r of R for which v(r) = 1. 2. Hochster’s construction

Representing a spectral space X within the spectrum of a ring of functions on X

We have seen that given a reduced ring A, we may represent the ring A as a ring of functions on Spec A taking values in the domain A/p at each p ∈ Spec A. We aim to replicate that situation starting from a general spectral space X. Given X a topological space and A a ring of functions X → R for some ring

R, then considering the “evaluation at x” map evx : A → R for each x ∈ X that takes a function f ∈ A to the value f(x) ∈ R, this map defines

• a subring of R, Rx = evx(A) = {f(x) | f ∈ A}; and

• an ideal of A, Ix = ker(evx) = {f | f(x) = 0R}.

We may then, as we did earlier, write elements of A ‘graph-wise’ as elements Q Q Q of x∈X Rx, with f 7→ x∈X evx(f) = x∈X f(x).

Proposition 2.1. If Rx is a domain (for every x ∈ X) then Ix is a prime ideal (for every x ∈ X),

since Ix is the kernel of the surjective homomorphism evx : A  Rx.

Thus in the case that Rx is a domain for every x ∈ X, we obtain a mapping of X into the prime spectrum of A, which we will write as Φ : X → Spec A given by Φ(x) = ker(evx) = Ix. We compare the (induced) topology on Φ(X) with the topology on X, par- ticularly in the case that X is a spectral space, which will be assumed from now

51 52 2. Hochster’s construction on. We have the collection D(A) as a basis for the topology on Spec A; we look at the preimage of principal open sets D(f) under Φ. To each f ∈ A we define the subset d(f) ⊆ X by

d(f) = Φ−1D(f) = {x | Φ(x) ∈ D(f)}

= {x | ker(evx) ∈ D(f)}

= {x | f∈ / ker(evx)}

= {x | f(x) 6= 0}.[1]

Then the following propositions, per Section 3 of [14], give successive condi- tions on the sets d(f) ⊆ X which relate the topology of X to that of Spec A.

=∼ Proposition 2.2. If {d(f)}f∈A is a basis for X then X ,→ Spec A.

Since X is a spectral space, then it is T0, and so if {d(f)}f∈A is a basis for X then Φ is injective, since distinct points x 6= y must be distinguished by some basic set d(f), in which case we cannot have Φ(x) = Φ(y). Then identifying X with its image Φ(X), clearly we have d(f) = D(f) ∩ Φ(X), and so, the topology defined by {d(f)}f∈A is exactly the subspace topology inherited from Spec A.

Proposition 2.3. If furthermore each d(f) is quasi-compact in X then X is a spectral subspace of Spec A, since in that case given a quasi-compact open set of Spec A, being a finite union D(a1)∪· · ·∪D(an), then its preimage under Φ, the set d(a1)∪· · ·∪d(an), is quasi-compact open in X, and so inclusion is a spectral map.

[1]It is worth noting for posterity that convincing myself of a suspected typo at exactly this point in Hochster’s exposition was the first non-trivial bit of maths I undertook in this project; his paper [14] gives the definition of d(f) as the set of those x such that f(x) = 0. The spectrum of a ring of functions 53

We shall in fact not find it too difficult, given a spectral space X, to define rings A and R as above so that X is a spectral subspace of Spec A. We will however find it significantly more difficult to obtain a ring for which X ∼= Spec A. ◦ Given a spectral space X take BX ⊆ K(X) an open subbasis for X. We fix

a field k. Take TX = {tU }U∈BX a collection of indeterminates indexed by BX and define corresponding “characteristic” functions χU : X → k[TX ] by   tU if x ∈ U χU (x) =  0 if x∈ / U.

Let XX = {χU }U∈BX ; we consider the ring k[XX ], that is, the ring of poly- nomial expressions in χ’s with coefficients in k. We then consider this as a ring of functions X → k[TX ] in the obvious way, with the value of a polynomial ex- pression P α χβ1 · ... · χβm in k[X ] at a point x ∈ X defined via (fg) = f g , i U1 Um X x x x

(f + g)x = fx + gx, and the definitions of the functions χU ; we then define the ring AX to be this ring of functions. An important caveat at this stage is that distinct expressions in AX may not truly be distinct functions X → k[TX ].

For example, suppose that our subbasis BX contains non-empty sets U, V such that U ∩ V = ∅ (—note in this case that the space X is reducible). Then the expression χU χV ∈ k[XX ] is identically zero on X. We examine this set-up in light of our earlier discussion. Given x ∈ X, we see that the range of evaluation at x, that is, the ring Rx as given above, is the subring of k[TX ] generated by those indeterminates tU for which x ∈ U. That is, given x ∈ X, define Tx = {tU ∈ TX | x ∈ U}. Then Rx = k[Tx]. Evidently, this is a domain for each x ∈ X. Then, we consider AX as a subring of the product of Q these domains, that is, we identify an element f ∈ AX with its graph x∈X fx. Note, as per the preceding paragraph, the action of identifying expressions in k[XX ] with the function they induce in AX may not be injective; however this identification is (of course) onto, and so we will frequently consider elements of 54 2. Hochster’s construction

AX as expressions in k[XX ], when strictly we mean the result of evaluating this

expression at each x ∈ X, that is, the function in AX induced by the expression.

Thus we consider AX as the subring

Y Y Y AX = { evx(f) | f ∈ k[XX ]} ⊆ evx(k[XX ]) = k[Tx], x∈X x∈X x∈X

and so obtain a mapping Φ : X → Spec AX by Φ(x) = ker(evx). We find

Proposition 2.4. X is a spectral subspace of Spec AX , that is, the sets d(f) =

−1 Φ (D(f)) for f ∈ AX are quasi-compact open and form a basis for X.

P × A general f ∈ AX is of the form f = i λimi for coefficients λi ∈ k and mi distinct monomials from k[X ] (that is, m = χα1 χα2 . . . χαN for some U ∈ B X i U1 U2 UN j X and αj ∈ N, where by convention we include 1 as the ‘empty’ monomial). Then at a particular x ∈ X, we have that

  fx = λ1m1 x + ··· + λnmn x,

where   α1 α2 αN λitU tU . . . tU in the case x ∈ Uj for all j λ m  = λ χα1 χα2 . . . χαN  = 1 2 N i i x i U1 U2 UN x  0 otherwise.

Then \ d(mi) = {x | x ∈ Uj for all j} = Uj, j and

d(f) = d(m1) ∪ d(m2) ∪ · · · ∪ d(mn)

(note that as the distinct monomials mi evaluate either to distinct monomials

(mi)x or to 0, we cannot have terms cancelling each other to zero, so if mi is non-zero at x then f is indeed non-zero at x, that is x ∈ d(mi) =⇒ x ∈ d(f); the reverse inclusion is obvious). The spectrum of a ring of functions 55

That is, we have that d(mi) is a finite intersection of quasi-compact open sets of X, so is itself quasi-compact open; and then d(f) is a finite union of quasi-compact open sets, and so is quasi-compact open. We need finally

that our collection {d(f)}f∈AX is a basis for X. But for U ∈ BX we have

χU ∈ AX , and d(χU ) = U, so BX ⊆ {d(f)}f∈AX ; furthermore we observed that T d(χU1 χU2 . . . χUn ) = i Ui, so {d(f)}f∈AX contains our subbasis and all finite intersections thereof, and therefore is a basis for X as required.

∼ We may at this stage be a long way from having Spec AX = X (which we will have exactly when Φ(X) = Spec AX ). We observe a necessary condition for ∼ Spec AX = X, which is in fact also sufficient in the case that X is a spectral subspace of Spec AX (in which case we make the identification between X and

Φ(X) as a subspace of Spec AX ). We have X ( Spec AX if and only if there is a prime ideal p ∈ Spec AX such that p ∈/ X. One situation where we may be able to detect this is where such a p lies in some V(I) thereby distinguishing it from some other set V(J) with p ∈/ V(J), so that V(I) 6⊆ V(J); but such that all primes in V(I)\V(J) fall outside the range of Φ (that is, outside the image of X), so that in the subspace X we do in fact have (V(I) ∩ X) ⊆ (V(J) ∩ X) (we take this opportunity to define z(I) = V(I) ∩ X = {x ∈ X | fx = 0 for all f ∈ I};

−1 equivalently z(I) = Φ (V(I)), and note for f ∈ AX then z(f) = X \ d(f)). In fact for X a spectral subspace of Spec AX , it is sufficient to check such situations do not arise for I,J such that J is principal and I is finitely generated, in order to guarantee X = Spec AX . That is,

Proposition 2.5 ([14], Theorem 2). Given a spectral space X and a ring of functions B from X into some ring R satisfying the conditions of propositions 2.1, 2.2 and 2.3, so that X is a spectral subspace of Spec B (and hence we iden- tify X with Φ(X) as a subspace of Spec B), then X = Spec B if and only if for all f ∈ B and finite subsets G ⊆ B, we have that z(G) ⊆ z(f) =⇒ f ∈ rad G. 56 2. Hochster’s construction

From the preceding discussion, since V(G) ⊆ V(f) ⇐⇒ f ∈ rad G we clearly have necessity. Conversely, given that X is a spectral subspace of Spec B, then

from 1.23 we have that X is closed in the patch space (Spec B)con. Then, if

X ( Spec B then there is some closed basis subset C of (Spec B)con such that Sn X ⊆ C ( Spec B. From 1.40 we have that the sets of the form V(f) ∪ i D(gi)

for f, g1, . . . , gn ∈ B form a closed basis for (Spec B)con. Then if X ( Spec B

there is some f ∈ B and some finite G = {g1, . . . , gn} ⊆ B such that n [ X ⊆ V(f) ∪ D(gi) ( Spec B, i or, taking complements, n n [ c \ c c X ⊆ V(f) ∪ V(gi) = V(f) ∪ V(gi) = V(f) ∪ V(hGi) ( Spec B i i Tn (the equality i V(gi) = V(hGi) = V(G)) coming from property3 in 1.1). Then

X ⊆ V(f) ∪ V(G)c ⇐⇒ (V(f) ∩ X) ∪ (V(G)c ∩ X) = X

⇐⇒ V(G) ∩ X ⊆ (V(f) ∩ X)

⇐⇒ z(G) ⊆ z(f).

But if for all f ∈ B and finite G ⊆ B we have that z(G) ⊆ z(f) =⇒ f ∈ rad G, then z(G) ⊆ z(f) =⇒ f ∈ rad G =⇒ V(G) ⊆ V(f) =⇒ c Sn V(f) ∪ V(G) = Spec B, i.e. V(f) ∪ i D(gi) = Spec B for all basic closed Sn subsets of (Spec B)con such that X ⊆ V(f) ∪ i D(gi); in which case we must have that X = Spec B as required.

In fact in the example of our ring AX as defined above (and for all subsequent rings considered in this chapter), this condition simplifies, due to the following sequence of observations:

Proposition 2.6 ([14], Theorem 4). Given a ring of functions B on X as in proposition 2.5 (i.e. so that X is a spectral subspace of Spec B), if for ev- ery function f ∈ B we have that the image of f in R is finite (i.e. the set The spectrum of a ring of functions 57

{evx(f)}x∈X = {fx}x∈X is a finite subset of R), then for all finite collections g1, . . . , gn ∈ B there is some g ∈ B such that z(g1) ∩ · · · ∩ z(gn) = z(g) (equiv- alently, such that d(g1) ∪ · · · ∪ d(gn) = d(g)).

By an elementary induction it suffices to consider the situation for two elements

f, g ∈ B. Let fx1 6= fx2 6= · · ·= 6 fxm , gz1 6= gz2 6= · · ·= 6 gzn enumerate all possible non-zero values for f and g (Note that if either f or g takes no non- zero values, i.e. if z(f) = X or z(g) = X, then either z(f) ∩ z(g) = z(f) or z(f) ∩ z(g) = z(g) and so we are done). Furthermore consider the sets

Wij = {y ∈ X | fy = fxi & gy = gzj } for some xi and zj, and choose an element yij ∈ Wij whenever Wij is non-empty. Let

Y = {x1, . . . , xm, z1, . . . , zn} ∪ {yij where defined}.

We consider the set S = {P f kgl} of all combinations of f and g (i.e. the

k1 l1 kn ln set of all sums f g + ··· + f g for all possible k1, l1, . . . , kn, ln, n ∈ N).

Since for each y ∈ Y we have that one of fy or gy is non-zero, then S 6⊆ Φ(y) = ker(evy) = {h ∈ B | hy = 0} for each y ∈ Y . Then as each Φ(y) is a prime ideal of B and S is an additively and multiplicatively closed subset of B, by prime S avoidance we have that S 6⊆ y∈Y Φ(y) = {h ∈ B | hy = 0 for some y ∈ Y }. Then, there is some r ∈ S non-zero on the whole of Y , in which case r will be non-zero at all points x ∈ X where f or g take non-zero values, since the values of fx and gx at such a point (and hence the value of rx) must be the same as the value taken at some y ∈ Y . Then, d(r) = d(f) ∪ d(g) as required.

Proposition 2.7. Given a ring of functions B on X as in proposition 2.5, such that for all finite collections g1, . . . , gn ∈ B there is some g ∈ B such that z(g1) ∩ · · · ∩ z(gn) = z(g), then we will have X = Spec B if and only if for all f, g ∈ B we have that z(g) ⊆ z(f) =⇒ f ∈ rad g, 58 2. Hochster’s construction

Tn as is obvious by replacing the set z(G) = i z(gi) in 2.5 by the set z(g) guaranteed by our hypothesis.

Corollary 2.8. Given a ring of functions B on X as in proposition 2.5, if for every function f ∈ B we have that the image of f in R is finite, we have X = Spec B if and only if for all f, g ∈ B we have that z(g) ⊆ z(f) =⇒ f ∈ rad g.

Then we note

Proposition 2.9. The ring AX as defined on page 54 has the property that for

each f ∈ AX the image of f in R = k[TX ] is finite.

Given that for a general f = P λ m ∈ A the monomials m = χα1 . . . χαN i i i X i U1 UN take only the values λ tα1 . . . tαN or 0 according as whether x ∈ T U or not, i U1 UN j j then clearly such an f only takes finitely many distinct values in k[TX ] under evaluation at points x ∈ X.

Indeed, looking forward to our eventual goal, the ring HX we ultimately ∼ construct such that Spec HX = X also shares this property, so that for every

g1, . . . , gn ∈ HX there is some g ∈ HX such that z(g1) ∩ · · · ∩ z(gn) = z(g). But ∼ as we will have Spec HX = X then we have z(gi) = V(gi) and d(gi) = D(gi), so

that, as noted in [24], the ring HX has the property that for every g1, . . . , gn ∈

HX there is some g ∈ HX such that

V(g1) ∩ · · · ∩ V(gn) = V(g)

and

D(g1) ∪ · · · ∪ D(gn) = D(g);

that is, the collections V(HX ) = {V(a)}a∈HX and D(HX ) = {D(a)}a∈HX are lat- tices under the operations ∩, ∪ (note that as V(f)∪V(g) = V(fg); D(f) ∩ D(g) = D(fg) they are in general semi-lattices under ∪, ∩ respectively); a property The spectrum of a ring of functions 59 that is shared, for example, by all unique factorisation domains, as well as by all B´ezoutrings, but which is not satisfied in general in an arbitrary ring B.

The semi-lattice D(A) is strongly linked to divisibility in the ring A: we have a | b ⇐⇒ b ∈ hai ⇐⇒ hbi ⊆ hai =⇒ D(b) ⊆ D(a) (—the reverse implication proceeds D(b) ⊆ D(a) =⇒ b ∈ rad a =⇒ a | bn for some n ∈ N); it would be interesting to see what further insights may follow from the fact that for any ring A, there is a ring having the same prime spectrum as A (namely, the ring HX for X = Spec A) with the property that D(HX ) is a lattice. We now give an illustration of the first step of our construction, and of how we may fall short of having X = Spec AX at this first stage, by calculating an example.

The example of X = Spec Z

Let X = Spec Z. The first step of our construction of a ring AX from X is to pick a subbasis for X; we may choose the sets D(p) for (positive) prime numbers p. We then take a set of corresponding indeterminates TX = {tp | p ∈ N is a prime} and a set XX of characteristic functions χp : X → k[TX ] defined by   tp if q ∈ D(p) χp(q) =  0 if q ∈/ D(p) for points of X, i.e. prime ideals q ∈ Spec Z. Then the range of evaluation at x = q, that is, the ring Rx as given above, is the subring of k[TX ] generated by those indeterminates tp for which q ∈ D(p). But we have q ∈ D(p) ⇐⇒ p∈ / q ⇐⇒ q 6= hpi. Then for a prime ideal q = hqi we have that Tq = TX \{tq}, Q and where q = h0i we have Th0i = TX . We then define AX = { evq(f) | q∈Spec Z Q f ∈ k[X ]} ⊆ k[Tq]. q∈Spec Z

In fact clearly expressions f ∈ k[XX ] evaluate “injectively” at the point h0i 60 2. Hochster’s construction

under the mapping χp 7→ tp, and so AX is isomorphic to the ring k[TX ] = k[t2, t3, t5, t7,... ] = Rh0i together with projections Rh0i → Rp given by fh0i 7→ fh0i mod tp for p the prime number such that p = hpi. We observe that for q 6= h0i we have

∼ Φ(q) = ker(evq) = {f ∈ AX | fq = 0} = htqi, and ∼ Φ(h0i) = {f ∈ AX | fh0i = 0} = {0}, thus seeing how X embeds into Spec AX . But we can without difficulty find suitable f, g ∈ AX so that z(g) ⊆ z(f) but f∈ / rad g: Let f = χ2 and g =

χ2 + χ3, that is, f is the function taking value fp = 0 at the point p = h2i and fq = t2 elsewhere, and g takes values gh2i = t3; gh3i = t2 and gp = t2 + t3 elsewhere. Then z(g) = ∅ and z(f) = {h2i} so that z(g) ⊆ z(f) as required, but f ∈ rad g ⇐⇒ ∃r ∈ A, n ∈ N such that f n = r · g, in which case we must n have, for example, that rh0i · gh0i = fh0i for some rh0i ∈ Rh0i = k[TX ], that is that n rh0i · (t2 + t3) = t2 . But clearly there is no rh0i ∈ k[TX ] for which this is possible, n hence there is no r ∈ AX such that f = r · g and so f∈ / rad g and so hence

X 6= Spec AX . Indeed the ideal ht2 + t3i is prime in Rh0i = k[t2, t3, t5, t7,... ], and so the ideal of AX consisting of those functions induced by some expression in the ideal hχ2 + χ3i in k[XX ] is a prime ideal of AX outside the range of Φ.

Extensions and indices

The construction continues by defining a discrete valuation v on Q(R ) = k(T ) x y x x for every specialisation x y in X (this collection of valuations Hochster refers to as an index). We note that where we have x y, then Ty ⊆ Tx. Since, if x y then y ∈ {x}, so that every open neighbourhood of y contains x. Then given U ∈ B, if y ∈ U (so that tU ∈ Ty) then x ∈ U, so tU ∈ Tx. We define Extensions and indices 61 the valuation v as follows: given t ∈ T (—that is, where x ∈ U), then x y U x v (t ) is either 1 or 0 according as whether t vanishes at y or not. That is, x y U U   1 y∈ / U, v (t ) = x y U  0 y ∈ U. We then define v on R = k[T ] by x y x x

• v (tα1 . . . tαn ) = P α for those i such that v (t ) = 1; and x y 1 n i x y i

• v (P λ m ) = min({v (m )}) for m distinct monomials in T and x y i i x y i i x × λi ∈ k .

It then has a uniquely defined extension to a discrete valuation v : x y k(Tx) → Z ∪ {∞}. Furthermore it has the following properties, which will be important not only in the ensuing discussions but also in light of later chapters:

Proposition 2.10. For all f ∈ AX , then at any Rx, and for all specialisations x y, we have that v (f ) 0. x y x >

Evidently, if f ∈ A then f ∈ k[T ] and so v (f ) 0. X x x x y x >

Proposition 2.11. For all f ∈ AX , then at any Rx, and for all specialisations x y, we have that v (f ) > 0 ⇐⇒ f = 0. x y x y

Since, given f ∈ AX then fx ∈ k[Tx]. Then by the definition of our valuation, we have v (f ) = 0 ⇐⇒ there is some term λm of f for some λ ∈ k× x y x and some monomial m such that v (λm ) = 0, which occurs if and only if x y x m consists solely of χU ’s such that y ∈ U (which we consider to be vacuously satisfied in the case that the term in question is a constant). But in such a case then λmy 6= 0 and so fy 6= 0; and conversely if fy 6= 0 then there must be some term λm of f which does not vanish at y, in which case we must have that m consists wholly of χ ’s such that y ∈ U, and so then v (f ) = 0 as U x y x required. 62 2. Hochster’s construction

Furthermore the next property will be crucial.

Proposition 2.12. For all f ∈ AX there is an integer Nf such that, at any Rx, and for all specialisations x y, if f 6= 0 then v (f ) N . x x y x 6 f

As we have observed several times before, monomials in AX evaluate either to

the corresponding monomial in k[Tx] or to zero; therefore the degree of some f ∈ k[T ], and so hence any possible value v (f ) (— provided f 6= 0), is x x x y x x

bounded by the degree of f as a polynomial in k[XX ].

Our construction then proceeds by considering f, g ∈ AX where at all R , for all x y, that firstly v (f ) v (g ); and furthermore that if x x y x > x y x v (f ) = v (g ) then either v (g ) = ∞ or v (g ) = 0. Once again x y x x y x x y x x y x these conditions will be not only significant for the progress of the current con- struction (—in some sense they ensure that the two properties discussed in the preceding propositions are maintained when we extend our ring), but also for when we reconsider this set-up in light of later chapters. Q The na¨ıve idea is then that we extend AX to the subring of x∈X Q(Rx) generated by A and the element Q fx . Of course we may have that at a X x∈X gx

certain x ∈ X we have that both fx and gx are zero (— the condition on our

valuations guarantees that whenever gx = 0 we have already that fx = 0); then, Q technically speaking, we extend to the subring of x∈X Q(Rx) generated by AX f Q and the element k ∈ Q(Rx) defined as g x∈X   fx f  gx 6= 0;   gx k x = g  0 gx = 0.

Hochster observes that, when the valuation conditions are satisfied for a cer- f 0 tain f and g, then we maintain X as a spectral subspace of the ring A = AX [ k ]. g As we have seen, this will be the case if Extensions and indices 63

0 0 0 0 0 • The range Rx = evx(A ) = {f (x) | f ∈ A } is a domain for all x ∈ X;

0 • The collection {d(g )}g0∈A0 is an open basis for X, and

• For every h0 ∈ A0 we have that d(h0) is quasi-compact open in X.

0 We have the first condition since Rx is a subring of Q(Rx); and so we 0 0 have a map Φ : X → Spec A given by x 7→ ker(evx). Furthermore clearly

0 the last condition implies the second condition, as the collection {d(h )}h0∈A0

contains the basis {d(f)}f∈AX for X —note that for f ∈ AX the set d(f) =

{x | ker(evx) ∈ D(f)} = {x | fx 6= 0} still defines the same subset of X when

0 0 we extend AX to A . Then furthermore we have that Φ is injective (as the sets d(f) for f ∈ AX still separate points in X) and so identifying X with its image Φ0(X) we have for each h0 ∈ A0 that d(h0) = D(h0) ∩ Φ0(X). We need to check that each d(h0) is quasi-compact open in X.

Proposition 2.13 ([14], Theorem 3). Given f, g ∈ AX such that for every spe- cialisation x y in X we have that v (f ) v (g ), with v (f ) = v (g ) x y x > x y x x y x x y x f f 0 only if either vx y(gx) = ∞ or vx y(gx) = 0, and defining A = AX [ k ] for k g g as given above, then for every h0 ∈ A0 we have that d(h0) is quasi-compact open in X.

f 0 0 We consider a general element h in A = AX [ k ], g

0 f n f  h = fn k + ··· + f1 k + f0 g g

0 for f0, . . . , fn ∈ AX . It will be not too difficult to show that d(h ) is quasi- compact in X. We will then show it is open by showing its complement, the

0 0 set z(h ) = {x ∈ X | hx = 0}, is closed in X. The behaviour of h0 at a point x ∈ X falls into two possible cases according f f   0 as whether k = 0 or not. We can see that if k = 0 then hx = 0 ⇐⇒ g x g x f  (f0)x = 0, and we know that we will have k = 0 exactly when fx = 0. g x 64 2. Hochster’s construction

f 

In the alternative case, that is, where k 6= 0, we consider the function g x f n 0 0 

g h in A . We see in fact (— regardless of which case in the definition of k g x applies at x) that

n 0 n n−1 n gx hx = (fn)xfx + ··· + (f1)xfxgx + (f0)xgx

n 0 n n−1 n n 0 at all x ∈ X, that is g h = fnf + ··· + f1fg + f0g and so g h is an f  element of A. Then in the case when k 6= 0 then certainly gx 6= 0, and so g x 0 n 0 hx = 0 ⇐⇒ gx hx = 0 (—since Rx is a domain). Then, to sum up, if fx = 0 0 0 n 0 then hx = 0 ⇐⇒ (f0)x = 0 and if fx 6= 0 then hx = 0 ⇐⇒ gx hx = 0; that is,

0 n 0 d(h ) = (z(f) ∩ d(f0)) ∪ (d(f) ∩ d(g h )).

Observe (as for every a ∈ AX we have d(a) is quasi-compact open in X,

n 0 and we have that f, f0, g h are all in AX ) that each set in the above expression

0 0 0 for d(h ) is a closed set in Xcon, and so d(h ) is closed in Xcon; then d(h ) is

quasi-compact in Xcon, and so is quasi-compact in X. We now consider the complement of d(h0). From the above discussion, we can see that

0 n 0 z(h ) = (z(f) ∩ z(f0)) ∪ (d(f) ∩ z(g h )),

0 and so we can likewise conclude that z(h ) is closed in Xcon. Then by proposition 1.22 it is closed in X if it contains the closure of each of its points; that is, if

0 0 0 we have for each x ∈ z(h ) and x y, that y ∈ z(h ). Then, given x ∈ z(h ) we n 0 have either x ∈ z(f) ∩ z(f0) or x ∈ d(f) ∩ z(g h ). If x ∈ z(f) ∩ z(f0) then as

f, f0 ∈ AX we have z(f) ∩ z(f0) is closed in X (since their complements d(f),

d(f0) are open), so certainly for any x y we have y ∈ z(f) ∩ z(f0) and so 0 n 0 y ∈ z(h ). Then, suppose x∈ / z(f), that is, fx 6= 0. Then x ∈ d(f) ∩ z(g h )

n 0 n 0 and so given x y we have (as z(g h ) is closed in X) that y ∈ z(g h ). If y ∈ d(f) then y ∈ d(f) ∩ z(gnh0) and so we are done. Suppose not, that is, Extensions and indices 65

suppose fy = 0. Then we need to show that y ∈ z(f) ∩ z(f0); we have y ∈ z(f)

0 and so we will have y ∈ z(h ) if and only if y ∈ z(f0). f  fx 0 We have fx 6= 0 so that k = , and we have hx = 0. Then g x gx

n h0 = (f ) fx  + ··· + (f ) fx + (f ) = 0, so x n x gx 1 x gx 0 x n (f ) fx  + ··· + (f ) fx = −(f ) and so n x gx 1 x gx 0 x n X fx i vx y (fi)x = vx y(−(f0)x) = vx y((f0)x). gx i=1

By proposition 2.11 then as f = 0 we have v (f ) 6= 0, and so by as- y x y x

fx sumption vx y(fx) > vx y(gx); then vx y( ) > 0 and so gx

n X fx i fx i vx y((f0)x) = vx y (fi)x min({vx y (fi)x }) gx > i 1 gx i=1 >

 fx = min({vx y (fi)x + ivx y( )}) gx i>1 by the standard properties of valuations; then as f ∈ A then v (f )  0 i X x y i x >

 fx for each i, and so vx y (fi)x +ivx y( ) > 0 for all i 1 and hence vx y((f0)x) > 0 gx > likewise. But again by proposition 2.11 we have v ((f ) ) > 0 =⇒ (f ) = 0, x y 0 x 0 y that is, y ∈ z(f0) as required.

We observe

Proposition 2.14. For all h0 ∈ A0 we have that h0 has finite image as a function

X → k(TX ),

since this is the case for f, g and all possible f0, . . . , fn ∈ AX .

As mentioned above, for the continuation of our construction we need the following properties for the ring A0:

0 0 Proposition 2.15. For all h ∈ A , then for all specialisations x y we have that v (h0 ) 0. x y x > 66 2. Hochster’s construction

0 0 Proposition 2.16. For all h ∈ A , then for all specialisations x y we have that v (h0 ) = 0 ⇐⇒ h0 6= 0. x y x y

0 0 Proposition 2.17. For all h ∈ A , there is an integer Nh0 such that for all

0 0 specialisations x y where h 6= 0 we have that v (h ) N 0 . x x y x 6 h

f  The first result follows since vx y( k ) 0 for all specialisations x y by g x > our condition on f and g, and so for

0 f n f  h = fn k + ··· + f1 k + f0 g g

f 0  i we have that the value vx y(hx) min(vx y (fi)x( k ) ) 0. > g x > 0 Furthermore if fx = 0 then hx = (f0)x and the second result will hold since f ∈ A ; furthermore those values v (h0 ) for which f = 0 but h0 6= 0 will 0 X x y x x x

be bounded by Nf0 .

Then to complete the result we need only look at those x for which fx 6= 0, in which case

n h0 = (f ) fx  + ··· + (f ) fx + (f ) . x n x gx 1 x gx 0 x

Note then that

n n−1 n−1 n 0 (fn)xfx + gx(fn−1)xfx ··· + gx (f1)xfx + gx (f0)x hx = n , gx

0 n n where the numerator is the image at x of the element f = fnf + ··· + g f0 of

0 AX , therefore at those x where fx 6= 0 and hx 6= 0 the value of any valuation 0 0 n v (h ) = v (f ) − v (g ) is bounded by N 0 (since g ∈ A and so x y x x y x x y x f X hence v (g ) 0); therefore in general the value of v (h0 ) is bounded by x y x > x y x

0 max(Nf0 ,Nf ). It remains therefore to address the second required property: that v (h0 ) = 0 ⇐⇒ x y x 0 hy 6= 0. Extensions and indices 67

fx If fy = 0 then vx y(fx) > 0 and so vx y( ) > 0; and furthermore if fy = 0 gx f  0 then k = 0 so that hy = (f0)y. We have that g y

0  fx vx y(h ) min({vx y (fi)x + ivx y( )}), x > gx i>0

 fx 0 and for all i 1 we have vx y (fi)x + ivx y( ) > 0 so that vx y(h ) = 0 =⇒ > gx x v (f )  = 0; but furthermore if v (f )  = 0 then by the property 1.423 x y 0 x x y 0 x of valuations (as every other term in the range on min is strictly bigger than zero) we in fact have equality, i.e. if v (f )  = 0 then x y 0 x

0  fx  vx y(h ) = min({vx y (fi)x + ivx y( )}) = vx y (f0)x = 0. x gx i>0

Then we have v (h0 ) = 0 ⇐⇒ v (f )  = 0 ⇐⇒ (f ) 6= 0 ⇐⇒ x y x x y 0 x 0 y 0 hy 6= 0. Otherwise if f 6= 0 then furthermore g 6= 0, and so firstly v (g ) = 0, so y y x y x that v ((gnh0) ) = nv (g ) + v (h0 ) = v (h0 ); x y x x y x x y x x y x and also as in the previous proposition we have that

0 n 0 hy = 0 ⇐⇒ (g h )y = 0; and so v (h0 ) = 0 ⇐⇒ v ((gnh0) ) = 0 ⇐⇒ (gnh0) 6= 0 ⇐⇒ h0 6= 0 as x y x x y x y y n 0 required (the second equivalence holding since g h is an element of AX ).

This allows us to extend our original ring AX in the following way: call a pair f, g admissible if they satisfy the condition of proposition 2.13. Note that if

0 0 0 0 we have f, g and f , g admissible in AX , then the pair f , g is still admissible in f 0 the ring A = AX [ k ] (since admissibility depends only on the valuations vx y, g Q which are already defined on the whole of x∈X Q(Rx), and therefore unchanged by whatever extensions we may carry out within this ring), and furthermore, by propositions 2.15 and 2.16, every part of the proof of 2.13 is still valid when 68 2. Hochster’s construction

f 0 0

extending the ring A by k (Hochster phrases this by stating that the collection g0 of valuations v is also an index for the ring A0). Then in fact the proofs of x y f 0 0

2.15, 2.16 and 2.17 are also valid for the extension A [ k ] (that is, our valuations g0 f 0 00 0 00 are also an index for the ring A = A [ k ] and each h ∈ A has a bound Nh on its g0 values at these valuations), and each h ∈ A00 again has finite image as a function

X → k(TX ), since it is a finite combination of functions with this property. Then we may extend in any order by finitely many admissible pairs and still have that the ensuing ring A is a ring of functions, each of finite image, with X as a spectral subspace of Spec A (i.e. the conclusion of 2.13 holds for A), that our collection of valuations are still ‘an index’ for A — that is that we may again extend by

any admissible pair in A, and that each h ∈ A has a bound Nh on its values.

This allows the following inductive argument: Set A0 = AX and let B = B0 = f

{ k | f, g ∈ AX satisfy the condition of proposition 2.13}. Then given any h ∈ g

A1 = AX [B], h arises from some finite extension of AX by admissible elements.

Therefore for all h ∈ A1 we have that the conclusions of propositions 2.13—

2.17 hold for h, so that X is a spectral subspace of Spec A1, with A1 a ring of functions of finite image in their codomain, and our collection of valuations v are an index for A bounded on each h ∈ A , so that we may extend A x y 1 1 1

by any admissible pair in A1. We then apply identical reasoning, inductively f

forming the set Bi = { k | f, g ∈ Ai satisfy the condition of proposition 2.13} g

and extending to the ring Ai+1 = Ai[Bi], where at all stages n we maintain X S as a spectral subspace of the ring An. We then set Aω = An. We find n∈N

Theorem 2.18 ([14], Theorem 6). Defining HX = Aω as given above, we have ∼ X = Spec HX .

We summarise some vital facts about the ring of functions HX . These are

(1) X is a spectral subspace of Spec HX ; Representing spectral maps via ring homomorphisms 69

(2) Each element f ∈ HX has finite image in the ring k(TX );

(3) For each element f ∈ HX there is an integer Nf such that for all special- isations x y in X where f 6= 0, we have that v (f ) N . x x y x 6 f

We show that it has the property that for all f, g ∈ HX we have that z(g) ⊆ z(f) =⇒ f ∈ rad g.

Suppose we have f, g ∈ HX such that z(g) ⊆ z(f). We need to show that

n there is some r ∈ HX and some n ∈ N such that f = r · g, so that f ∈ rad g.

We consider a ring Ai in our inductive procedure for i sufficiently large that both f and g are contained in Ai. There is an integer Ng such that for all specialisations x y where g 6= 0 we have that v (g ) N . Then at all x x y x 6 g specialisations x y we have v f Ng+1 = (N + 1)v (f ) > v (g ) x y x g x y x x y x except if v (g ) = ∞ or v (g ) = 0, where v f Ng+1 = v (g ). Since x y x x y x x y x x y x firstly g = 0 =⇒ f = 0 since z(g) ⊆ z(f); and secondly v (f ) = 0 x x x y x only if fy 6= 0, in which case by the same reasoning necessarily gy 6= 0 and so f Ng+1 f Ng+1 vx y(gx) = 0. Then k is admissible in Ai, and so k ∈ Ai+1 and hence g g f Ng+1 f Ng+1 Ng+1 k ∈ HX ; finally for r = k ∈ HX we have that f = r · g, as either g g Ng+1 Ng+1 fx f = 0 in which case r = 0 and fx = 0 = 0 · g , or r = and so x x x x gx Ng+1 fx Ng+1 r g = g = fx ; thus f ∈ rad g as required. x x gx x

Then by2 .8 we have that X = Spec HX .

Representing spectral maps via ring homomorphisms

As we saw in the preliminaries, every ring homomorphism gives rise to a spectral map between the corresponding spectral spaces. Our construction given above allows us, without too much difficulty, to show that every spectral map

∗ f : X → Y of spectral spaces arises from a ring homomorphism f : HY → HX ∼ ∼ ∗ ∼ such that Spec HX = X, Spec HY = Y and Spec(f ) = f. 70 2. Hochster’s construction

◦ ◦ ◦ A spectral map f : X → Y induces a mapping K(f): K(Y ) → K(X) of the quasi-compact open sets of X and Y by U 7→ f −1(U); when we consider our construction of a ring from the spaces X and Y we see that this map restricts ◦ to a map r : BY → K(X). We then require that we choose BX so that r(BY ) ⊆

BX ; furthermore we wish that the map r be injective, as this will simplify the situation greatly. Obviously in general this may not be possible, when the ◦ map K(f) is not itself injective; thus we extend the definition of BX , and so the corresponding definitions of the sets TX and XX , to allow us to specify an injective mapping r : BY ,→ BX , as follows: we let BX be any set, equipped with ◦ an identification function i : BX → K(X) so that the range i(BX ) is a subbasis for X; thus we think of the characteristic function χu corresponding to some ◦ element u ∈ BX as being defined as taking the value tu on the set i(u) ∈ K(X) and 0 elsewhere. We can then define an injective function r : BY ,→ BX simply ◦ −1 by taking BX to be the set BY × K(X) with r defined as r(U) = (U, f (U)) ◦ and the identification function i : BX → K(X) to be projection onto the second ◦ co-ordinate (of course instead of K(X) in the above definition we could take any ◦ −1 subbasis B ⊆ K(X) such that f (BY ) ⊆ B, with maps defined in an identical way). We can see by examination of the argument in 2.4 that this definition does not materially alter the behaviour of our ring of functions AX defined as on page 54 from XX , and so all the properties claimed for the ring AX and ∼ all subsequent constructions thereon are maintained; hence Spec HX = X as required. Furthermore, injectivity of r allows us to guarantee, when we define a ring

∗ homomorphism f : HY → HX from the map r, that distinct monomials in HY evaluate to distinct monomials in HX , greatly simplifying the argument.

Recall the observation that not every distinct expression in k[XX ] corre- sponds to a unique function. In the case of a single characteristic function χu, Representing spectral maps via ring homomorphisms 71

this is only a problem if i(u) = ∅ so that χu is identically zero on X; as otherwise it will take the value tu somewhere on X, which is taken by no other character-

0 istic function χu0 for u 6= u . Then the homomorphism ϕ : AY → AX defined by ϕ(χU ) = χr(U) is injective on individual characteristic functions χU except

−1 where i(r(U)) = f (U) = ∅, where ϕ(χU ) = 0. Then for distinct monomials

0 0 0 m, m in AY , we have ϕ(m) 6= ϕ(m ) except where ϕ(m) = 0 = ϕ(m ). This P allows us to use the argument as in 2.4 that given h = j λjmj ∈ AY we have that d(ϕ(h)) = d(ϕ(m1)) ∪ · · · ∪ d(ϕ(mn)).

−1 Then for all h ∈ AY we have that f (d(h)) = d(ϕ(h)): since, per the pre- ceding discussion, this reduces to checking f −1(d(m)) = d(ϕ(m)) on monomials m = χα1 . . . χαN , in which case we have d(m) = T U . But U1 UN j j

ϕ(m) = ϕ(χα1 ) · ... · ϕ(χαN ) = χα1 · ... · χαN U1 UN r(U1) r(UN ) so that \ \ d(ϕ(m)) = dχα1 · ... · χαN  = i(r(U )) = f −1(U ) = f −1(d(m)) r(U1) r(UN ) j j j j as required.

The homomorphism ϕ is defined (via the identification of AY with the ring Q { y∈Y evy(g) | g ∈ k[XY ]}) on all of Ry for each y ∈ Y and so obviously has Q a uniquely defined extension to the ring y∈Y Q(Ry); and so remains defined throughout our extension process.

−1 Then if we verify that ϕ (ΦX (x)) = ΦY (f(x)) for all x ∈ X, that is, that

−1 ϕ (ker(evx)) = ker(evf(x)) for all x, we will have that the natural extension of Q ∗ ∗ ϕ to HY ⊆ y∈Y Q(Ry) is the ring homomorphism f such that f = Spec ϕ. We see

−1 g ∈ ϕ (ker(evx)) ⇐⇒ ϕ(g)x = 0 ⇐⇒ x∈ / d(ϕ(g)) ⇐⇒ f(x) ∈/ d(g)

⇐⇒ gf(x) = 0 ⇐⇒ g ∈ ker(evf(x)) as required. 72 2. Hochster’s construction 3. Ring constructions on finite posets

We now present two constructions, per Lewis [17], and Ershov [7], of a ring having prime spectrum homeomorphic to any finite spectral space. Note that in the finite case, from 1.26, that two spectral spaces are homeomorphic if and only if their specialisation posets are isomorphic; note furthermore that every

finite poset (X, 6) arises as the specialisation poset of some finite spectral space, namely the spectral space consisting of X in the lower topology; therefore we may consider finite spectral spaces and finite posets interchangeably, and to achieve the desired result it suffices to find for every finite poset (X, 6) a ring ∼ R such that (Spec R, ⊆) = (X, 6). As a consequence, the topological details of a finite spectral space X are reduced to the details of a partial order relation on X, which discards many of the intricacies present in the general topological situation, thereby greatly simplifying the presentation and discussion of the results.

Summary of techniques

We introduce two techniques that are used repeatedly in both the construc- tions under consideration, to obtain rings with certain (finite) prime ideal struc- ture.

Intersecting (discrete) valuation rings

We have per 1.45 that the ideals, in particular the prime ideals, of a valuation ring V are totally-ordered by subset inclusion; that is, the prime spectrum of

73 74 3. Ring constructions on finite posets

V is a total order in the specialisation order. In particular, a (proper) discrete valuation ring is a local principal ideal domain, so that it has exactly two prime

ideals h0i ( m. We see how this may be used, via the following propositions, to obtain certain finite posets as the prime spectrum of a ring.

Proposition 3.1 ([3], §7, Proposition 2). If A1,...,An are inclusion-incomparable T valuation rings of a field k, then for R = i Ai we have that the maximal ideals

of R are in bijection with the maximal ideals mi of the valuation rings Ai, with each maximal ideal ni of R being of the form ni = (mi ∩ R).

Proposition 3.2 (op. cit, §7, Proposition 1). Given R,A1,...,An as above, the prime ideals of Ai are in order-preserving bijection with the prime ideals of

R contained in ni,

since we have that Rni = Ai.

∼ Corollary 3.3 (loc. cit.). R/ni = Ai/mi,

∼ ∼ since R/ni = Rni /niRni = Ai/mi.

In particular if A1,...,An are inclusion-incomparable discrete valuation T rings of a field k, then R = i Ai has exactly n non-zero prime ideals p1,..., pn

of height 1 corresponding to the maximal ideals mi of each Ai; furthermore

the quotient map R  R/pi is isomorphic to the restriction of the canonical

quotient map πi : Ai  Ai/mi to R.

Fibre sums of spaces

The fibre sum of topological spaces is a construction that gives the pushout of a diagram in the category of topological spaces and continuous maps. That Summary of techniques 75 is, given a diagram X 6 f

g Y  Z of topological spaces and continuous maps, the fibre sum provides a space XtZ Y and maps f 0, g0 having the universal property in the diagram 0 g X tZ Y X 6 6 f 0 f

g Y  Z. Most details in the following discussion may be found in [8] and the references therein; a detailed exploration of fibre sum constructions in relation to spectral spaces may be found in the textbook [4].

The fibre sum X tZ Y as a set is the disjoint union X t Y modulo the equivalence relation ∼ given by f(z) ∼ g(z) for each z ∈ Z, topologised with the quotient topology defined from the surjection X t Y  X t Y/∼. We will in fact only need to consider the circumstance where one of f or g is a closed embedding, which simplifies the description of the fibre sum. In the case that g is a closed embedding (and therefore an injective function) we identify

Z with its image under g as a subset of Y , and the underlying set of X tZ Y is then the disjoint union (Y \ Z) t X, where the maps g0, f 0 are the identity on X, (Y \ Z) respectively, and where f 0 maps z ∈ Z to f(z) ∈ X. Then the topology of X tZ Y has as closed sets C ⊆ (Y \ Z) t X such that C ∩ X is closed in X and (C ∩ (Y \ Z)) ∪ f −1(C ∩ f(Z)) is closed in Y . We note that X homeomorphically embeds into X tZ Y as a closed set and (Y \ Z) (considered as a subspace of Y ) homeomorphically embeds into X tZ Y as an open set. A particularly important observation for the purposes of our construction is that 76 3. Ring constructions on finite posets

we have a specialisation x y in X tZ Y if and only if one of the following hold:

(1) We have a specialisation x y in (Y \ Z);

(2) We have a specialisation x y in X, or

(3) There is some z ∈ Z and specialisations x z in Y and f(z) y in X.

It is worth noting (per [4]) that in general the topological fibre sum X tZ Y of two spectral spaces X,Y over a third spectral space Z may not itself be a spectral space; however in the case that one of the maps f or g is a closed

embedding then the space X tZ Y is indeed spectral. On the other side of our considerations, we have that Spec is a contravariant functor from the category of rings and ring homomorphisms to the category of topological spaces and continuous maps, so that the relevant corresponding notion in the category of rings is the pullback; it is well known that this is given by the fibre product of rings. That is, given rings and ring homomorphisms A

ϕ

ψ ? B - C

then the fibre product A ×C B = {(a, b) ∈ A × B | ϕ(a) = ψ(b)} has the universal property in the diagram 0 ψ - A ×C B A

ϕ0 ϕ

? ψ ? B - C (with the maps ϕ0, ψ0 being the obvious projections). We consider the situation where we have spectral spaces X = Spec A, Y = Spec B, Z = Spec C, and corresponding maps f = Spec ϕ, g = Spec ψ, as in the following diagram. Summary of techniques 77

AX 6 Spec ϕ =⇒ f

ψ ? g B - C Y  Z

As noted above, it may be the case that the fibre sum X tZ Y is not a spectral space; in which case we cannot hope to associate the fibre product of rings with the corresponding fibre sum of spaces. However we know from 1.5 that in the case the ring homomorphism ψ (for example) is surjective, we have that the corresponding spectral map g = Spec ψ is a closed embedding, in which case the fibre sum is a spectral space; in fact in this case we have

Theorem 3.4 ([8], 1.4). Given rings A, B, C and corresponding spectra X =

Spec A, Y = Spec B, Z = Spec C as in the above diagram, then if ψ : B  C ∼ is a surjective ring homomorphism, then Spec(A ×C B) = X tZ Y .

This then gives us a general method to use the properties of the fibre sum discussed above to construct rings with particular spectra. An important addition to theorem 3.4 concerns the properties of the fibre product in relation to Noetherian rings. Unfortunately, it is not sufficient that the rings A, B, C be Noetherian to guarantee their fibre product D = A ×C B is Noetherian. Recalling that a ring homomorphism ϕ : A → B is said to be finite in the case that B can be finitely generated as a module by the image of A under ϕ, we find

Proposition 3.5 ([8], 1.8). Given A, B, C and ϕ : A → C, ψ : B  C as above, then the fibre product D = A ×C B is Noetherian if and only if A and B are Noetherian and ϕ is a finite homomorphism.

This observation will be critical when considering which spectral spaces can arise as the prime spectrum of a Noetherian ring. 78 3. Ring constructions on finite posets

The method of Lewis

The construction in [17] uses iteratively two techniques to build up a given finite poset as the spectrum of a ring: the first being “gluing” (the spectrum of) a domain onto a maximal ideal in (the spectrum of) some other ring; the second being “joining” maximal ideals of some ring so that they form a single point in the spectrum. In fact it was observed by Fontana [8] that both techniques are instances of forming the fibre sum of spectral spaces, via its connection with the fibre product of the corresponding rings. This insight greatly simplifies the analysis of Lewis’s construction, and it is within this framework that it will be presented here. The basic unit of Lewis’s construction is the spectrum of a polynomial ring in finitely many indeterminates, localised at the height-1 prime ideals corre- sponding to the principal ideals generated by each indeterminate. This gives a ring having n (—where n is the number of indeterminates) non-zero prime ide- als each of height 1. In detail: we fix a field k, and consider the polynomial ring k[t1, . . . , tn] for some n. For each ti, we have that the ideal htii is a prime ideal c of height 1. Forming the set S = ht1i ∪ · · · ∪ htni we take A = k[t1, . . . , tn]S. By the well-known properties of localisation, we have that the prime ideals of

A are in order-preserving bijection with the prime ideals of k[t1, . . . , tn] disjoint

from S; these are exactly the height-1 prime ideals htii for i = 1, . . . , n, and the zero ideal h0i. Then the space Spec A has specialisation structure as in the following diagram:

n z }| {

••• ... •,

• The method of Lewis 79

•y where specialisation relations proceed “upwards”, that is, a diagram denotes

•x the existence of a specialisation x y. We summarise the two instances of the fibre sum of spectral spaces utilised by Lewis and describe the iterative process by which we may obtain a ring with spectrum homeomorphic to a given finite poset.

‘Gluing’ of domains over maximal ideals

Working over some field K, let A be some domain contained in K (—in practice A will be the ring k[t1, . . . , tn]S as described above, and we will take  K = k t1, . . . , tn ), and let B be a K-algebra having some maximal ideal m. Then setting C = B/m we have an inclusion ϕ : A,→ C, and a surjective quotient map ψ : B  C, as in the following diagram:

A ∩

ϕ

ψ ? B  C

Considering the image of this diagram under the functor Spec, we have that Spec ϕ : Spec C → Spec A maps the single point h0i of the spectrum of the field C to ϕ−1({0}) = ker ϕ = {0} in Spec A, and the map Spec ψ : Spec C → Spec B is a closed embedding taking the point h0i of Spec C to ψ−1({0}) = ker ψ = m in Spec B (hence the image of Spec C is the closed subset {m}). The following picture describes the situation: 80 3. Ring constructions on finite posets

. . . .

•• ... • A ...... ∩

• 6 ... ••m ...

......  • • • ... •

. . ? . . B  B/m = C

Then, as the map ψ is surjective, forming the fibre product D = A ×C B, we will have that Spec D is homeomorphic to the fibre sum of X = Spec A and Y = Spec B over Z = Spec C; that is, we have the situation represented by the following diagram:

ψ0 g0  ⊃ D = A ×C B AW = X tZ Y X  ∩ ∩ 6 6 Spec ϕ0 ϕ =⇒ f 0 f

? ψ ? g  ⊃ B  C Y Z

Then the spectrum of D, being the space W = X tZ Y , has the topological structure of Y = Spec B with the closed set {m} replaced by the closed set X = Spec A; that is, it is the result of gluing the space Spec A onto the maximal point m of Spec B, as in the following picture: The method of Lewis 81

. . Spec D . .

•• ... •

... ••m ...

......

• • ... •

. . . .

Examining D = A ×C B as a ring, we have

D = {(a, b) ∈ A × B | ϕ(a) = ψ(b)} = {(a, b) ∈ A × B | b mod m = a}

∼= {b ∈ B | b mod m ∈ A}

= A + m ⊆ B.

We may perform this operation on finitely many maximal points of Spec B simultaneously. That is, suppose m1,..., mn are among the maximal ideals of a

K-algebra B, and we have finitely many domains A1,...,An each contained in

K. Then the product A = A1 ×· · ·×An is contained in the product C = B/m1 ×

· · ·×B/mn, giving an inclusion homomorphism ϕ : A,→ C, and by the Chinese remainder theorem we again obtain a surjective quotient map ψ : B  C. We examine the image of these maps under the functor Spec. Write Ci for the field

B/mi, so that C = C1 ×· · ·×Cn. Then C, being a finite product of fields, has n prime ideals each of the form pi = (C1,...,Ci−1, 0,Ci+1,...,Cn) for i = 1, . . . , n.

Then the image of the point pi ∈ Spec C under the induced map Spec ϕ, that is, the preimage of pi under ϕ, is the prime ideal (A1,...,Ai−1, 0,Ai+1,...,An) of A. 82 3. Ring constructions on finite posets

•• ... • •• •• ... • Spec A ...

• • •

6 6 6 Spec ϕ

Spec C •• ... •

The image of such a point pi under Spec ψ is the kernel of the quotient

map B  B/mi; that is, it is the maximal ideal mi of B. Then the image of Z = Spec C under the closed embedding g = Spec ψ is the closed set of maximal

points {m1,..., mn} ⊆ Spec B

Then taking the fibre product D = A×C B, again Spec D is homeomorphic to

the fibre sum W = XtZ Y of X = Spec A and Y = Spec B over Z = Spec C. We retain the labelling of the diagram on page 80. Then Spec D has the structure of Spec B with the closed set {m1,..., mn} replaced by the closed set Spec A. ∼ In particular, recall that we have a specialisation x y in W = Spec D if and only if either there is already such a specialisation in X = Spec A or in

Y \ Z = Spec B \{m1,..., mn}, or else that for x ∈ Y and y ∈ X there is a

point mi ∈ {m1,..., mn}, such that there is a specialisation x mi in Y and

a specialisation f(mi) y in X. As can be seen from the above description of

f = Spec ϕ, this has the result of attaching the spectrum of each Ai directly on

to the maximal point mi of Spec B, with no further relations other than those guaranteed by transitivity ensuing. The method of Lewis 83

Then examining D = A ×C B as a ring, we have

D = {(a, b) ∈ A × B | ϕ(a) = ψ(b)}

= {((a1, a2, . . . , an), b) ∈ (A1 × · · · × An) × B |

(b mod m1, . . . , b mod mn) = (a1, . . . , an)} ∼ = {b ∈ B | b mod mi ∈ Ai for all i} \  = Ai + mi ⊆ B. i

‘Joining’ of maximal ideals

Suppose that m1,..., mn are among the maximal ideals of a K-algebra B and again let C = B/m1 × · · · × B/mn = C1 × · · · × Cn, with ψ : B  C the surjective quotient homomorphism guaranteed by the Chinese remainder theorem. Then let A = K, and let ϕ : A,→ C = C1 × · · · × Cn be the diagonal map K,→ K × · · · × K, i.e the map that sends a 7→ (a, . . . , a) for a ∈ K.

A = K ∩

ϕ

ψ ? B  C = C1 × · · · × Cn

Again, applying the functor Spec, the image of the point pi = (C1,..., 0,...,Cn) ∈

Spec C under the induced map Spec ψ is the kernel of the quotient map B  B/mi; that is, it is the maximal ideal mi of B. Then the image of Z = Spec C under the closed embedding g = Spec ψ is the closed set of maximal points

{m1,..., mn} ⊆ Y = Spec B.

The preimage of pi = (C1,..., 0,...,Cn) under ϕ is {0}, so that for every

−1 pi ∈ Z = Spec C we have that f(pi) = ϕ (pi) = h0i, the sole point in X = Spec K. 84 3. Ring constructions on finite posets

• A = K 6 ∩

 •• ... • ... •m1 •m2 •m3 ... •mn

. ..

• • ... •

. . ? . . B  B/m1 × · · · × B/mn = C

Then once again, as ψ is surjective, taking D = A×C B we have that Spec D is homeomorphic to the fibre sum W = X tZ Y .

ψ0 g0  ⊃ D = A ×C B AW = X tZ Y X  ∩ ∩ 6 6 Spec ϕ0 ϕ =⇒ f 0 f

? ψ ? g  ⊃ B  C Y Z

Then Spec D has the structure of Spec B with the closed set Z = {m1,..., mn} replaced by the single point X = Spec A = Spec K; that is, it is the result of

0 amalgamating the points m1,..., mn into a single point p :

Spec D... •p0

. ..

• • ... •

. . . .

0 ∼ (note we have a specialisation x p in W = Spec D if and only if we have a specialisation x mi in Y = Spec B for some i). The method of Lewis 85

We examine the ensuing ring:

D = {(a, b) ∈ A × B | ϕ(a) = ψ(b)}

= {(a, b) | (a, . . . , a) = (b mod m1, . . . , b mod mn)}

= {(a, b) | a = b mod m1 = ··· = b mod mn} ∼ = {b ∈ B | ∃m1 ∈ m1, m2 ∈ m2, . . . , mn ∈ mn, a ∈ K:

b = a + m1 = a + m2 = ··· = a + mn} \ = {b | ∃m ∈ mi, a ∈ K: b = a + m} i \ = K + mi = K + (m1m2 ... mn) ⊆ B. i Once again, we may apply this procedure on finitely many finite sets of maximal ideals at once, joining the maximal ideals of each set to a (distinct) single point in the ensuing spectrum of the fibre product: We take A to be a product of copies of K, one for each set of maximal ideals, we take C to again be the product of the quotient rings of B modulo each maximal ideal, and we map A diagonally component-wise into C, that is, the component of A corresponding to a particular set of maximal ideals maps diagonally into all quotient rings which are the quotient of B by a member of this set of maximal

ideals. Explicitly, let Mj = {mij }, j = 1, . . . , m be disjoint finite subsets of

m Y Y maximal ideals of B. Let A = K and C = B/mij . Then we take

j ij ϕ : A,→ B to be the map sending

(a1, a2, . . . , am) 7→ ((a1, a1, . . . , a1), (a2, a2, . . . , a2),..., (am, am, . . . , am)),

Y that is, where each component aj maps diagonally into B/mij . Then a

ij point mij ∈ Mj maps under Spec ϕ to the point (K,...,Kj−1, 0,Kj+1,...,K) in Spec A; and so, as in the comparable situation for ‘gluing’ finitely many

spaces at once, we have that each point mij of Mj is joined to a single point 86 3. Ring constructions on finite posets

mj in Spec D, and no other relations are created between points from Spec B in the ensuing space Spec D. Finally,

D = {(a, b) ∈ A × B |ϕ(a) = ψ(b)}

= {((a1, . . . , am), b) |(a1, . . . , a1) = (b mod m11 , . . . , b mod mn1 ) & . .

0 (am, . . . , am) = (b mod m1m , . . . , b mod mnm )} ∼  \  = {b ∈ B | ∃d1 ∈ mi1 , a1 ∈ K: b = a1 + d1 & i . .  \  ∃dm ∈ mim , am ∈ K: b = am + dm } i \ \ = {b | b ∈K + mi1 & ... & b ∈ K + mim } i i \ \  = K+ mij ⊆ B. j i

The iterative procedure

We now describe how we build a ring LX having spectrum isomorphic to a given finite poset X. We proceed by induction on the dimension of X; that is, we build up X ‘layer-by-layer’, increasing the dimension of the poset obtained by 1 at each stage. The idea is that by removing the points of maximum height of X, we ob- tain a poset X0 of dimension n − 1, to which an inductive hypothesis can be

∼ 0 applied to give a ring LX0 with Spec LX0 = X . We then “reconstruct” the poset X first by gluing a layer of (spectra of) 1-dimensional domains above appropriate maximal ideals of LX0 using the ‘gluing of domains’ technique de- scribed from page 79 onwards, and then, where we have in X that some point The method of Lewis 87 of maximum height lies above several different predecessors, we join together maximal points in the ring ensuing from our gluing of domains to obtain a ring having the structure of X, using the ‘joining of maximal ideals’ technique as described from page 83 onwards. More precisely, let Xht n = {x1, . . . , xm} be the points of maximum height of X. Let X

•y1 •y2

• • •y3

• • •

Then we can see that the point y3 is an immediate predecessor of the point x1 of maximum height in X, but that y3 is not maximal in the ensuing poset

X

• • •y3

• • •

We would therefore not be able to create the relation between y3 and x1 by gluing a point above y3, as our technique only applies to maximal points. 88 3. Ring constructions on finite posets

However, we know that, given a ring having a certain prime ideal structure, we may use the “joining of maximal ideals” technique to combine several maximal points into one. We are then able to avoid the difficulties of non-maximal immediate predecessors of points of maximum height by considering a poset where we have “split” the points of maximum height into multiple copies, so that each copy has a unique immediate predecessor, as shall be seen. We know then know that, given a ring representing this poset, we may obtain a ring representing our original poset by joining the multiplicated maximal points.

Then, labelling elements of Y as y1, . . . , yM , we define a new set of points

+ X = {xi,j | xi ∈ Xht n and yj ∈ Y is an immediate predecessor of xi in X}.

+ We then let X1 = X z ⇐⇒ yj > z

(with relations in X

x x x •1,1 •2,1 •2,1

x •1,3

•y1 •y2

• • •y3

• • •

The poset X1 has two key properties. Firstly, every point of maximum height

in X1 has a unique immediate predecessor: note that the height n points in X1

will be the points xi,j such that ht yj = n − 1 (—in general, ht xi,j = ht yj + 1);

guaranteeing that in X1 every immediate predecessor of a point of maximum

0 height is maximal in the poset X = X1 \ (X1)ht n. Secondly, we can obtain our

original poset X from X1 by joining maximal points of X1: for each original The method of Lewis 89

+ point of maximum height xi ∈ Xht n we wish to join all duplicates xi,j ∈ X to a single point. We now describe how we set up an induction to provide rings with the required spectra.

Theorem 3.6 ([17], Theorem (2.10)). Given X a finite poset, equivalently a ∼ finite spectral space, there exists a ring LX such that Spec LX = X.

We build a ring LX having spectrum isomorphic to a given finite poset X by induction on the dimension of X. More precisely, our induction hypothesis will be that for any field k and

0 0 for any finite poset X with dim X < n, there is a k-algebra LX0 such that ∼ 0 Spec LX0 = X ; an important point is that although at the induction step we ∼ will be trying to show that a k-algebra LX such that Spec LX = X exists for a given field k, our hypothesis allows us to assume the existence of a k0-algebra

∼ 0 0 such that LX0 such that Spec LX0 = X for any field k . In the base case where dim X = 0 we have that X consists of finitely many (m, say) points, each of

m height 0. Then setting LX = k clearly we have that LX is a k-algebra with m ∼ Spec LX = Spec k = X. Now suppose X is a finite poset with dim X = n. We assume that for any

0 0 0 0 finite X with dim X < n and any field k there is a k -algebra LX0 such that ∼ 0 Spec LX0 = X . We must prove that for a given field k, there is a k-algebra LX ∼ such that Spec LX = X. We fix k. We first show

Lemma 3.7. Under the present inductive assumptions, if X1 is the poset ob- tained from X by the above-described procedure of “splitting” the points of max-

imum height of X into multiple copies, then there is a k-algebra LX1 such that ∼ Spec LX1 = X1. 90 3. Ring constructions on finite posets

Let y1, . . . , yM0 be the immediate predecessors of the points of maximum height G  of X1; then X1 = (X1)

minimum point corresponding to the point yj itself) for each j = 1,...,M0.

Take T = {txi | xi ∈ Xht n} a set of indeterminates corresponding to each point

x ∈ X , and let T = {t | x ∈ y  }. Then for each j = 1,...,M , working i ht n  xi i j 0 yj over the ring k[T ], let S = Sht ic for t ∈ T . Then set A = k[T ] .  j xi xi  j  Sj yj yj yj Per the discussion at the beginning of this section, this is a 1-dimensional do-

main having height-1 prime ideals corresponding to the indeterminates txi ; that ↑ is, as a poset, it has prime ideal structure isomorphic to the sub-poset yj .

Let K = k(T ) so that Aj ,→ K for all j = 1,...,M0, let A = A1 × · · · × AM0 and let B be the K-algebra guaranteed by our induction hypothesis such that

∼ 0 0 Spec B = X . Then for each yj ∈ X there is a maximal ideal mj of B corre-

sponding to yj. Let C = B/m1 × · · · × B/mM0 and take the obvious inclusion

map ϕ : A,→ C and the canonical surjective quotient map ψ : B  C. We are in exactly the situation for applying the “gluing of domains over maximal ide- T  als” construction as per page 79 onwards. Then for D = A×C B = j Aj +mj we have that Spec D has the structure of Spec B, which is to say of X0, with ∼ ↑ each point mj replaced by Spec Aj = yj ; that is, Spec D has exactly the struc-

ture of the poset X1. Furthermore, since k ,→ Aj for every j = 1,...,M0, we

have that D is a k-algebra. We set LX1 = D. The method of Lewis 91

x x x •1,1 •2,1 •2,1

Spec A1 Spec A2

•y1 •y2

? ? •m1 ••m2

Spec B • • •

• • •

∼ Lemma 3.8. Given a k-algebra LX1 such that Spec LX1 = X1, where X1 is the poset obtained from X from the procedure described above, then there is a ∼ k-algebra LX such that Spec LX = X.

We have that to each xi,j ∈ X1 there corresponds a unique maximal ideal ni,j of LX1 . We wish to join those ideals ni,j as j ranges over the immediate prede- cessors yj of xi into a single point corresponding to xi, for each i = 1, . . . , m. m 0 m 0 Y Y Let A = k and let C = LX1 /ni,j over all such j as yj is an imme- i j 0 0 0 diate predecessor of xi in X; then take ϕ : A ,→ C the “component-wise” diagonal map as described in our “joining of maximal ideals” section that takes

0 Y 0 the i-th component of A diagonally into LX1 /ni,j, and let ψ : LX1  C j be the canonical surjective quotient map. Then we are in exactly the situation for applying the “joining of maximal ideals” construction per page 83 onwards.

0  0 T T Thus setting LX = LX1 ×C A = i k + j ni,j we have that Spec LX has the structure of Spec LX1 , which is to say of X1, with the exception that for each xi, 92 3. Ring constructions on finite posets

the points xi,j corresponding to the maximal ideals ni,j of LX1 have been joined ∼ into a single point; thus Spec LX = X, and LX is a k-algebra as required.

x1,1 x2,1 x2,1 • • • •x1 •x2

x •1,3

•y1 •y2 •y1 •y2

• - •

• • •y3 • • •y3

• • • • • •

So, given our n-dimensional poset X, we form the poset X1. From lemma

0 ∼ 3.7, we are provided a k-algebra LX such that Spec LX1 = X1. Therefore from ∼ 3.8 we are provided a k-algebra LX such that Spec LX = X, and the induction is complete.

Application to 1-dimensional Noetherian spectral spaces

In this subsection we temporarily broaden our scope beyond finite spectral spaces, to investigate an aspect of the long-standing problem of characterising which posets can arise as the inclusion-ordering of prime ideals of a Noetherian ring; equivalently, as the specialisation order on the prime spectrum of a Noethe- rian ring. We noted in the first chapter that, whilst this problem has been solved completely in the case of Noetherian spectral spaces, it is also known that this characterisation is insufficient for spectra of Noetherian rings: whilst the spec- trum of a Noetherian ring always gives rise to a specialisation poset satisfying the conditions of 1.27, we saw an example of such a poset which can never arise from the spectrum of a Noetherian ring. Note that in view of 1.26 the problem The method of Lewis 93 of finding a ring with spectrum homeomorphic to a given Noetherian spectral space X remains equivalent to finding one with isomorphic specialisation order; we wish to investigate to what extent can we represent such a space, or poset, by the spectrum of a Noetherian ring. We use the tools we have just developed, to show,

Theorem 3.9. If X is a Noetherian spectral space of dimension 1, then there ∼ is a Noetherian ring NX such that Spec NX = X.

We recall that if X is a Noetherian space, then under , we have

(1) X has the ascending chain condition;

↑ ↑ (2) X has finitely many minimal points y1, . . . , yn with y1 ∪ · · · ∪ yn = X; and

(3) Every pair x, y ∈ X has finitely many minimal upper bounds in X.

We see that this means we can take a product of finitely many (Noetherian)

↑ domains whose spectra correspond to yj for each minimal point yj, and that we need perform only finitely many ‘joining’ operations between height-1 points of these spaces to obtain the desired space X. However, the crucial concern will be to ensure that our operations preserve Noetherianeity for the rings constructed.

0 Let Xmin = {y1, . . . , yn} consist of the minimal elements of X and let X =

{x ∈ X | ht x = 1} = X \ Xmin. If X is infinite take k an algebraically-closed field with |k| = |X0| and identify elements of X0 with elements of k under the implied bijection; otherwise take k to be any algebraically-closed field of

0 characteristic 0 and identify each x1, x2, x3, . . . , xi,... in X with 1, 2, 3, . . . , i, . . . in k. We consider localisations of k[T ]. Observe that as the maximal ideals of k[T ] are in bijection with k, to each point x ∈ X0 there corresponds a maximal ideal mx = hT − xi of k[T ]. 94 3. Ring constructions on finite posets

↑ Then considering yj for yj ∈ Xmin, let [ Sj = k[T ] \ hT − xi ↑ x∈yj ∼ ↑ and set Bj = k[T ]Sj ; this is a Noetherian ring such that Spec Bj = yj ; that is, ↑ the points x ∈ yj correspond bijectively to the maximal ideals nx = mxk[T ]Sj of Spec Bj Then we take

0 [ ↑ ↑ ↑ ↑ X1 = yi ∩ yj = {x | x ∈ yi ∩ yj for some i 6= j} i6=j to be the set of all elements of X lying above more than one minimal element.

0 0 0 By property3, this is finite; let X1 = {x1, . . . , xm}.

We set B = B1 × · · · × Bn; this is again a Noetherian ring and we have that

↑ 0 0 Spec B has the structure of the disjoint union of the yj ; note that the xi ∈ X1 are exactly those points of X appearing more than once as maximal ideals in

0 Spec B; we label each instance of a x , equivalently, a n 0 , as n 0 in the case i xi xi,j 0 ↑ that xi ∈ yj . m Y Y 0 Then we let C = B/n 0 for such i, j as appear; note as X is finite xi,j 1 i j then C is a Noetherian ring and (—provided it is not empty) we get a surjective

homomorphism ψ : B  C. Let Ci,j denote the component of C corresponding

to B/n 0 for each i, j. We see that the image under Spec ψ of the point corre- xi,j

sponding to the zero ideal of Ci,j (that is, the ideal consisting of all elements in

Y 0 C 0 0 that are zero in the C component) is the point n 0 (—if X is empty i ,j i,j xi,j 1 i0,j0 then Spec B ∼= X and we are done already. We suppose not).

0 0 We now take A = k[T ]/m 0 for each x ∈ X , where m 0 is the “original” i xi i 1 xi 0 maximal ideal of k[T ] corresponding to x ; we have A ∼ C = B/n 0 for i i = i,j xi,j 0 ↑ every j such that xi ∈ yj . Y Then we take A = Ai so that again A is Noetherian, and we define ϕ : i The method of Lewis 95

A,→ B as in the “joining of maximal ideals” section to be the ‘component-wise Y diagonal’ map where each component ai maps diagonally into B/mi,j. Then j each point corresponding to the zero ideal of some Ci,j0 maps under Spec ϕ to the point (A1,...,Ai−1, 0,Ai+1,...,Am) in Spec A. Thus we have the following set-up.

A

ϕ

ψ ? B  C Importantly at this point we note that the map ϕ as so defined is a finite m Y Y ring homomorphism; that is, we observe that the ring C = Ci,j is finitely i j generated by the image of A and the standard basis (1, 0,..., 0), (0, 1, 0,..., 0), . . . , (0, 0,..., 1).

Then, taking the fibre product D = A×C B of A and B over C, we have that in Spec D each maximal ideal ni,j of Spec B corresponding to a multiplicated

0 point xi has been joined into a single point, with exactly the relevant speciali- sations and no others being preserved; thus Spec D ∼= X; and furthermore, as A and B are Noetherian rings and ϕ is a finite homomorphism, we have from

3.5 that D = A ×C B is a Noetherian ring, and the result is proved.

Note that we only used one of our two techniques: that of ‘joining’ maximal ideals. In fact, certainly in the set-up we have considered here, it is not possible to use the other technique, that of ‘gluing’ the spectrum of some domain above a maximal ideal, and still preserve Noetherianeity of the ensuing constructed ring. Observe, to set up the maps required for the application of our topological fibre sum construction, given a domain A we took k = Q(A) and a k-algebra

B, so that we would obtain a surjective map ψ : B  B/m and an inclusion 96 3. Ring constructions on finite posets

ϕ : A,→ k ,→ B/m for m the maximal ideal in question. But in such a case the map ϕ cannot be finite, since ([1](5.3)) a homomorphism is finite if and only if it is integral and of finite type; but ([1](5.7)) if k is a field and A,→ k is integral then A is a field. Thus any non-trivial gluing operation that utilises the inclusion of a domain into its field of fractions necessarily occasions a non- Noetherian ring; indeed, this method, christened the “D + m” construction (viz. the description of the ensuing ring, page 81), has been exploited [9] as a technique for providing examples of non-Noetherian rings! In the context of Noetherian ring constructions, the paper [6] offers an in- teresting alternative approach to exploring what partial-order structures can arise within the spectrum of a Noetherian ring: there it is shown, by a similar methodology to the above, that any finite poset X can be order-preservingly embedded into the spectrum of a Noetherian ring A in a particularly strong way: the embedding ϕ : X,→ Spec A is bijective on the sets of minimal points and on the sets of maximal points of the respective posets, and there exists a saturated chain of a given length d between ϕ(x1) and ϕ(x2) in Spec A if and only if a saturated chain of length d exists between x1 and x2 in X. The construction employed utilises exactly the “joining of maximal ideals” method as one of two elements of its procedure; the other being “adding a layer” to (the spectrum of) a ring B by extending to B[X], then localising at those new maximal ideals that are to be preserved. It would be of great interest to explore further how much this combination of techniques allows one to control what possibilities arise as the prime spectrum of a Noetherian ring. In the literature addressing the question of which orderings can arise from the spectrum a Noetherian ring, the survey paper [23] provides a comprehensive overview of the state of the art in dimension 2; also relevant in this context are the papers [12], [24], [13], [26]. The method of Ershov 97

The method of Ershov

In the paper [7] Ershov gives his own construction of a ring RX from a finite poset X such that the prime ideals of RX have the structure of X under subset inclusion; equivalently, such that the (spectral) space given by X in the lower topology is homeomorphic to Spec RX . We examine this construction here, in light of the considerations of the previous chapter. We saw that the construction of Hochster begins with taking a subbasis of

open sets BX , a collection of indeterminates TX = {tU }U∈BX indexed by BX , and rings Rx for x ∈ X given by considering Tx = {tU ∈ TX | x ∈ U}, then setting Rx = k[Tx] over our chosen field k. Given a finite spectral space X, we know that the sets x↓ give an open subbasis for the (lower) topology on X. Then, following Hochster, we may take one indeterminate tx for each point x ∈ X, and characteristic functions χx defined by   ↓ tx if z ∈ x χx(z) =  ↓ 0 if z∈ / x .

We then see that   tx if z 6 x χx(z) =  0 if z 6 x, and so the ring Rz corresponding to a point z is that generated by those indeter-

↑ minates tx for which x > z, i.e. Rz = k[Tz] for Tz = {tx | x > z} = {tx | x ∈ z }.

We will adopt the shorthand of writing Tz↑ for this collection of indeterminates.

Note as before that if x y then Ty↑ ⊆ Tx↑ . Thinking of the poset X as a spectral space, we define a discrete valuation ring corresponding to each specialisation y z in X, which we will write as 98 3. Ring constructions on finite posets

V . y z

We utilise the sets of indeterminates Tz↑ etc. as defined above. Given a specialisation y z, firstly, in the case that z = y we define the ring V to y y be the improper valuation ring

V = k(T ↑ ) = Q(R ). y y y y

In the case that z 6= y, so that y ∈ y↑ \ z↑, then, writing in the obvious

T \Tz↑ y  tx0 way T = T ↑ \{t }, we define a set = { | t 0 ∈ T \ T ↑ } = y  y y ty ty x y  z

tx0 0 ↑ { | x ∈ y  \ z }. Note that this set may well be empty. We then define the ty discrete valuation ring V to be y z

T \ T ↑ y  z V = k(T ↑ , )[t ] . y z z y htyi ty

T \Tz↑ y  That is, writing Ky z for the field k(T ↑ , ), then Vy z is the localisa- z ty tion of the polynomial ring K [t ] at the ideal generated by t . y z y y This ring has unique maximal ideal m = t V ; thus we have verified y z y y z (per 1.46) it is a discrete valuation ring. Corresponding to each V we then y z have the canonical quotient map π : V → V /m = K (in the y z y z y z y z y z case of the improper valuation rings V we take the map π to be the y y y y identity). We then define rings corresponding to y↑ for each y ∈ X, as follows:

\ R ↑ = V y y z z y (it is worth noting in passing, though this will not become relevant until later

in our discussions, that the ring Ry↑ is an intersection of finitely many discrete valuation rings). We need to observe some important facts about how the rings of the form

R ↑ and those of the form V relate to each other. Given some x ∈ X we x x y The method of Ershov 99

have that R ↑ ⊆ V for each y x, so that given r ∈ R ↑ , then π (r) x x y > x x y 0 is defined for every y > x. But furthermore, fixing some y > x, we find that 0 V 0 ⊆ K 0 for any z y so that R 0↑ ⊆ K 0 : since y z x y > y x y

T 0 \ T ↑ 0↑  z T \ T y x  y V 0 = k(T ↑ , )[t 0 ] ⊆ k(T 0↑ ) ⊆ k(T 0↑ , ) = K 0 . y z z y hty0 i y y x y ty0 tx

Then for r ∈ Rx↑ it makes sense to compare elements s ∈ Ry0↑ with the element

π 0 (r) ∈ K 0 , as indeed we will do in the next definition. x y x y Q Then finally we define the ‘Ershov ring’ RX ⊆ x∈X Rx↑ to be the sub-ring of functions on X taking values at x in the ring Rx↑ defined by

Y R = {f ∈ R ↑ | ∀x ∈ X, y x then f = π (f )}. X x > y x y x x∈X Remark. In Ershov’s paper, his construction is phrased very generally in terms of families of subsets of a finite poset X. In practice, however, the subsets considered are exactly those of the form x↑ for elements x ∈ X. For comparison purposes, we now relate the notions and notation of Ershov’s paper with the objects we have defined above.

↑ Let W be the collection {x }x∈X of subsets of X. Ershov’s construction

Y begins by defining discrete valuation rings RZ corresponding to subsets Z ⊆ Y in W; that is, to subsets z↑ ⊆ y↑ for z, y ∈ X; equally, to elements y, z such that y z in X; Ershov’s ring RY is the ring we denote by R . Z y z Given a set Y ∈ W (that is, where Y = y↑ for some y ∈ X), Ershov defines ↓ Y to be the subfamily {Z ∈ W | Z ⊆ Y }; we see this is the collection

↑ ↑ ↑ ↑ ↑ ↑ {z ∈ W | z ⊆ y } = {z | z > y} = {z | z ∈ y }. The ring Ershov denotes as Y R , defined in reference to the notion ↓ Y , is our ring Ry↑ . Given a subcollection W ⊆ W, he defines ↓ W to be S ↓ Y . We have a 0 0 Y ∈W0 ↑ ↑ correspondence Yi = yi for each Yi ∈ W0; let us write X0 = {yi | yi = Yi ∈ W0}. ↑ ↑ Then ↓ W0 is the collection of z where z is contained in some ↓ Yi, that is,

↑ where z > yi, for some i. Then ↓ W0 = {z | z > yi for some yi ∈ X0} = 100 3. Ring constructions on finite posets

↑ ↑ {z | z ∈ X0 }. A subcollection W0 ⊆ W is called lower if ↓ W0 = W0; we ↑ see that the collection W0 is lower exactly when X0 = X0 . Then, on lower

subfamilies W0 ⊆ W, equivalently, on up-closed subsets X0 ⊆ X, Ershov defines the ring RW0 ; in the case that W0 = W then RW is exactly our “Ershov ring”

RX . Given an up-closed subset X0 ⊆ X, then the ring RX0 , being the Ershov ring defined on the sub-poset X0 with the inherited order, corresponds to the ring RW0 in Ershov’s notation.

As RX is a ring of functions on X taking values in domains Rx↑ ⊆ Q(Rx), we obtain a mapping X → Spec RX by x 7→ ker(evx) as before. In fact ([7], ∼ Theorem 1), when X is finite, we have Spec RX = X. We proceed towards a proof of this fact, noting other important properties of this construction along the way. We note,

3.10. If the poset X has a unique minimal element x0, then the ring RX can be identified with a certain subring of the ring R ↑ : x0 since we have for all f ∈ R that f = π (f ) then R ∼ {f | f ∈ R } = X y x0 y x0 X = x0 X

0 0 {r ∈ R ↑ | πx0 y(r) ∈ Ry↑ for all y ∈ X & πy z(πx0 y(r)) = πy z(πx0 y (r)) x0 0 for all y z y in X}. In general, for x1, . . . , xn the minimal elements

of X then RX can be identified with a subring of R ↑ × · · · × R ↑ , namely x1 xn ∼ RX = {(fx1 , . . . , fxn ) | f ∈ RX }, which can be characterised by a similar condition as above, but further requiring that π (r ) = π (r ) whenever xi y i xj y j

we have xi y xj.

We now observe some relationships between the ring RX we have constructed and certain rings RY corresponding to subsets Y of X. Note that given any subset Y ⊆ X, we may define a ring of functions on Y The method of Ershov 101

Q Q from RX simply by projecting from x∈X Rx↑ to x∈Y Rx↑ , that is, let Y RX Y = { fx | for f ∈ RX }. x∈Y In fact

Theorem 3.11 ([7], Proposition 6). If Y is up-closed in X then we have

RX Y = RY for RY the Ershov ring defined on the sub-poset Y with the in- herited order.

Let us assume for the purpose of an induction that |X| = n and all posets X0 with |X0| < n have the property that for every upper-closed subset Y 0 ⊆ X0 then RX0 Y 0 = RY 0 . In the base case where |X| = 1 we have the only upper- closed subset of X is X itself, and the restriction map RX → RX X is identity. Continuing the induction, we consider two possible situations: Where X has a unique minimal element x0, and where X has more than one minimal element.

0 Suppose X has a unique minimal element x0, and let X = X\{x0}. We note

(—provided that there is some y 6= x0 ∈ X) that R ↑ is the intersection of the x0

finitely many discrete valuation rings V for y ∈ x  = X0. We claim these x0 y 0 are inclusion-incomparable; that is, for y 6= z then we have that V 6⊆ V x0 y x0 z and vice-versa. Since, if y 6= z suppose without loss of generality that y 6 z.

Then t ∈/ T ↑ ; and so t ∈ T \ T ↑ . We compare

z y z  y x0

T \ T ↑

 y x0 V = k(T ↑ , )[t ] ; x0 y y x0 htx0 i tx0 T \ T ↑

 z x0 V = k(T ↑ , )[t ] , x0 z z x0 htx0 i tx0 1 1 t t seeing that ∈ V , ∈/ V ; z ∈ V , z ∈/ V . x0 z x0 y x0 y x0 z tz tz tx0 tx0 Then from the discussion at the start of this chapter, each maximal ideal m of such valuation rings corresponds to a unique maximal ideal n = x0 y y

(mx0 y ∩ R ↑ ) of R ↑ , and furthermore the map πx0 y R : R ↑ Kx0 y x0 x0  ↑ x0  x0 102 3. Ring constructions on finite posets

∼ is surjective, since it is isomorphic to the quotient map R ↑ R ↑ /ny = x0  x0 V /m = K . Then furthermore by the Chinese remainder theorem x0 y x0 y x0 y Q we get a surjection π0 : R ↑ 0 Kx0 y. Note that as we have Ry↑ ⊆ Kx0 y x0  y∈X 0 0 Q 0 for all y ∈ X then RX ⊆ 0 Kx0 y. Let R0 = {r ∈ R ↑ | π0(r) ∈ RX } = y∈X x0 −1 π0 (RX0 ).

∼ 0 Then RX = R0, as given f ∈ RX then fX0 ∈ RX and π0(fx0 ) = fX0 (as π (f ) = π (f ) = f ) so that f ∈ R . Conversely, given r ∈ R then 0 x0 y x0 y x0 y x0 0 0 0 Q mapping r to the element fr of x∈X Rx↑ defined by   r y = x0, (fr)y = π (r) y ∈ X0  x0 y we have that f is in R , since π (r) is in R 0 for all y 6= x ; therefore the r X x0 y X 0 condition of R that π ((f ) ) = (f ) is satisfied either trivially in the case X y z r y r z 0 0 y = x0 or by the condition on X in the case y, z ∈ X ; furthermore clearly ∼ (fr)x0 = r, so that RX = R0 as claimed. Then identifying RX with R0 we have that RX X0 = π0(RX ) = RX0 as required.

Alternatively, suppose X has more than one minimal element; let x1 be ↑ some minimal element of X and set X1 = x1, considered as a sub-poset of X.

Since there is some minimal element x2 6= x1 in X we have X1 ( X so that ↑ |X1| < n. Now let X2 = (X \ x1) . As x1 is minimal then x1 ∈/ X2, so that

∗ again X2 ( X and |X2| < n; note that X = X1 ∪X2. Finally let X = X1 ∩X2. 0 Let R = RX1 ×RX∗ RX2 be the fibre product of RX1 and RX2 over RX∗ as in the following 0 R = RX1 ×RX∗ RX2  RX1

↓ ↓

∗ RX2  RX ,

∗ ∗ where the maps RX1  RX , RX2  RX are surjective by hypothesis (since The method of Ershov 103

∗ X is an up-closed subset of both X1 and X2 of size < n), thereby inducing surjectivity in the projections RX1 ×RX∗ RX2  RX1 , RX1 ×RX∗ RX2  RX2 , ∼ 0 We show that RX = R .

0 A general element of R is of the form r = (f1, f2) for f1 ∈ RX1 , f2 ∈ RX2 such that f1X∗ = f2X∗ . As X1 ∪ X2 = X, we may define an element f = Q ϕ(r) ∈ x∈X Rx↑ by   (f1)x x ∈ X1 fx =  (f2)x x ∈ X2; note this is well defined since f1 and f2 agree on their intersection. In fact we have f ∈ R , since every condition π (f ) = f appears in one of either X x y x y X or X by their upwards-closedness, and so the condition π (f ) = f is 1 2 x y x y satisfied for fx as so defined as it is satisfied by whichever of (f1)x ∈ RX1 or

(f2)x ∈ RX2 applies at x. Evidently the map ϕ is injective. Conversely given f 0 ∈ R then we have (f 0 , f 0 ) ∈ R0 since obviously these restrictions X X1 X2 agree on their further restriction to X∗; and clearly ϕ(f 0 , f 0 ) = f 0, hence X1 X2 ∼ 0 RX = R as required; and so the restriction maps RX  RX1 , RX  RX2 are surjective as in the above diagram. Finally given any (proper) up-closed subset

Y of X clearly either Y ⊆ X1 or Y ⊆ X2 (according as whether x1 ∈ Y or not), and so we get a surjective restriction map RX  RY via RX  RXi  RY by the result just proved and the induction hypothesis applied to the appropriate

RXi for i = 1 or 2.

Thus far we have examined relationships between rings defined on subsets of a finite poset X without any comment as to their prime ideal structure. We rectify this forthwith.

Theorem 3.12 ([7], Theorem 1). Given X a finite poset, equivalently, a finite ∼ spectral space, then Spec RX = X. 104 3. Ring constructions on finite posets

We note that the ring RX is a ring of functions on a finite spectral space X

taking values in a field k(TX ); then the range evx(RX ) of a point x ∈ X, being Q just projection onto the x-coordinate of RX ⊆ x∈X Rx↑ is a domain and so we

have a mapping x 7→ ker(evx) of X into Spec RX ; it will be seen that this gives the required homeomorphism. We use an induction identical in structure to that of the preceding proposi- tion.

In the base case where X consists of a single element X = {x0} we have the following rings

V = k(t ); x0 x0 x0 \ R ↑ = Vx0 z = Vx0 x0 = k(tx0 ); x0 z x0 Y Y RX = {f ∈ Rx↑ | ∀x ∈ X, y x then fy = πx y(fx)} = Rx↑ = R ↑ = k(tx0 ) > x0 x∈X x∈X

as the condition on RX is trivial, so that ker(evx0 ) = {0} = Spec RX and so ∼ Spec RX = X as required. In the inductive step we consider two possibilities: that X has a unique minimal element, or that X has more than one minimal element.

0 In the case that X has a unique minimal element x0 let X = X \{x0} as before; by induction we have that x 7→ ker(evx) gives a homeomorphism ∼ 0 Spec RX0 = X , and we have a surjective map π0 : RX  RX0 .

We claim that for every non-zero prime p of RX we have ker π0 ⊆ p. Let a ∈ ker π and take a non-zero b ∈ p. Let v be the discrete valuation 0 x0 y associated to the valuation ring V , and let N = max({v (b)} 0 );. We x0 y b x0 y y∈X see that a ∈ ker π =⇒ a ∈ m for all y, so that v (a) > 0 for all 0 x0 y x0 y y. Then v ( aNb+1 ) > 0 for all y so that aNb+1 ∈ ker π ⊆ R . But then x0 y b b 0 X N +1 Nb+1 a b  a = b · b ∈ p, so that a ∈ p as required. Then the non-zero prime

ideals of RX are in order-preserving bijection with the prime ideals of RX0 , The method of Ershov 105

0 which by assumption are isomorphic to the sub-poset X = X \{x0} (and

0 0 clearly for all f ∈ RX and x ∈ X we have fx0 = 0 ⇐⇒ (π0(f))x0 = 0 so that ker(evx0 ) in RX corresponds exactly with ker(evx0 ) in RX0 ); thus Spec RX

0 has the structure of Spec RX plus the ideal h0i. Verifying that ker(evx0 ) = {0},

0 since f = 0 =⇒ f 0 = 0 for all x ∈ X (as f 0 = π 0 (f )), then we have x0 x x x0 x x0 ∼ that x 7→ ker(evx) gives a homeomorphism Spec RX = X as required.

Alternatively suppose X has more than one minimal element. Again, let x1

↑ ↑ ∗ be some minimal element of X and take X1 = x1, X2 = (X \x1) , X = X1 ∩X2 ∗ as before. We have |X1|, |X2|, |X | < n and so an induction applies; we have that

∗ ∗ ∼ ∼ ∗ ∼ Spec RX1 = X1, Spec RX2 = X2 and Spec RX = X . Note that as X1,X2,X are up-closed, they are closed sets in the lower topology on X; furthermore X∗ is a closed subspace of both X1 and X2. We may then easily observe, in light of ∼ the discussion at the start of this chapter, that X = X1 tX∗ X2 (perhaps this is most easily observed by considering that any specialisation x y in X must lie within X1 or X2, recalling that the topology on X is uniquely determined by its specialisations). But then we are in the set-up of theorem 3.4 as follows:

RX1 X1 6 Spec =⇒ ↓ ↓ ∪ ∗ ∗  ⊃ RX2  RX X2 X ,

∼ ∗ ∼ so that Spec(RX1 ×RX∗ RX2 ) = X1 tX X2 = X. But we have from the proof of ∼ ∼ proposition 3.11 that RX = RX1 ×RX∗ RX2 , so that Spec RX = X as required. 0 Again f 0 = 0 ⇐⇒ (f ) 0 = 0 for the appropriate X containing x ; so that x Xi x i x 7→ ker(evx) again gives the homeomorphism. 106 3. Ring constructions on finite posets 4. Comparison of constructions

Comparison of finite constructions

We now proceed to compare aspects of the constructions considered in the preceding chapter, and of the rings LX and RX we defined from a finite spectral space X.

The fundamental “building blocks” of the construction of LX were the do- mains Aj defined as the localisation of a polynomial ring in finitely many inde- terminates at the (union of the) principal ideals generated by each indetermi- nate, thus giving a ring with prime spectrum having the following structure n z }| {

••• ... •

• for n the number of indeterminates of our polynomial ring. That is, to obtain a ring A having prime ideal structure as in the above diagram for a certain n, we begin by working over the ring k[t1, . . . , tn]. We then take the set S = c ht1i ∪ · · · ∪ htni and define A = k[t1, . . . , tn]S.

On the other hand, the basic objects in the construction of RX were the rings Rx↑ , being an intersection of the finitely many discrete valuation rings V where y ∈ x↑. We did not have cause to examine the prime ideal structure x y of these rings in our exposition, utilising only the properties of the associated quotient maps π . However, we know from the observation following 3.3 x y

that (—at least in the case that there is some y 6= x in x  ) such a ring is

107 108 4. Comparison of constructions

a 1-dimensional domain with maximal ideals corresponding bijectively to the

maximal ideals m of each V for y ∈ x  . x y x y Conversely, the rings of the form of A as above can be viewed as an in- tersection of discrete valuation rings. Observe if we wished to represent the poset

by a ring of such form, the ring obtained would A0 = k[t]hti. As A0 is a local Noetherian domain with principal maximal ideal tA0, we have that it is a discrete valuation ring of the field Q(A0) = k(t). Suppose we have n such rings chosen so that they are each a valuation ring of the same field; that is, we choose

A1 = k(t2, . . . , tn)[t1]ht1i,

A2 = k(t1, t3, . . . , tn)[t2]ht2i, . .

Ai = k(t1, . . . , ti−1, ti+1, . . . , tn)[ti]htii, . .

An = k(t1, . . . , tn−1)[tn]htni, so that Q(A1) = ··· = Q(An) = k(t1, . . . , tn). Then for each i we have that 1 1 ∈/ Ai, ∈ Aj for all j 6= i, so that the rings are inclusion-incomparable, and ti ti it is easy to see that the intersection

\ Ai = k(t2, . . . , tn)[t1]ht1i ∩ · · · ∩ k(t1, . . . , tn−1)[tn]htni i

f is the set of g ∈ k(t1, . . . , tn) such that g is not in any of the ideals ht1i,...,

T c htni; that is, i Ai is the ring k[t1, . . . , tn](ht1i∪···∪htni) = A. Comparison of finite constructions 109

We find however that the ring associated by the construction of Ershov to a finite irreducible 1-dimensional spectral space X (that is, a space X having specialisation structure as in the diagram at the start of this section) appears significantly different to the ring just described. Let us label the points of X as follows:

•x1 •x2 •x3 ... •xn

•x0 Then for all x ∈ X the ring V = k(t ), and for i = 1, . . . , n, we have i xi xi xi T V = k(t , j6=i )[t ] x0 xi xi x0 htx0 i tx0

T t where j6=i = { xj | j 1, j 6= i}. tx0 > tx0

Then for each i 1 we have that R ↑ = Vxi xi = k(txi ), and > xi

\ \ Tj6=i  Rx↑ = Vx0 xi = k(txi , )[tx0 ]htx i ; 0 t 0 i i x0 then

Y R = {f ∈ R ↑ | ∀x ∈ X, y x then f = π (f )} X x > y x y x x∈X Y = {f ∈ R ↑ | f = π (f ) for all i 1} x xi x0 xi x0 > x∈X ∼ = {r ∈ R ↑ | πx0 xi (r) ∈ R ↑ = k(txi ) for all i 1}. x0 xi >

\ Tj6=i  Note that R ↑ is the ring k(txi , )[tx0 ]htx i and we have the quotient x0 tx0 0 i Tj6=i Tj6=i maps πx0 xi : k(txi , )[tx0 ]htx i Kx0 xi = k(txi , ); it is not especially tx0 0  tx0 clear what restriction the condition on RX imposes on members of R ↑ but it x0 seems plausible that it may be non-trivial. It is even not at all clear to this author what the set R ↑ itself actually is. x0 110 4. Comparison of constructions

However, recall that our initial step in the definition of our ring RX was

choosing a suitable subbasis BX for the topology on X; as the topology is the

↓ lower topology, we chose the basis BX = {x | x ∈ X}, being, in general, a basis for the fine lower topology on a poset. Alternatively we could have chosen

↑ c BX = {(x ) | x ∈ X}, a subbasis for the coarse lower topology, as the fine and

coarse lower topologies on a finite poset coincide. Let T X = {tx | x ∈ X} be the corresponding set of indeterminates. We would then obtain the characteristic functions   ↑ c tx if z ∈ (x ) χx(z) =  ↑ 0 if z ∈ x . We then see that   tx if x 6 z χx(z) =  0 if x 6 z,

and so in this new set-up the ring Rz corresponding to a point z is that

generated by those indeterminates tx for which x 6 z, i.e. Rz = k[T z] for ↓ c T z = {tx | x 6 z} = {tx | x ∈ (z ) }.

Note that we still maintain the property that if x y then T y ⊆ T x. Now let us follow the remainder of the Ershov construction on the poset X

as given above, in this different setting. We first find that for each xi, the set

T xi = {txj | xj 6 xi} = {txj | for j such that i 6= j 6= 0}; we will label this set

T xi = T j6=i. Then, associated to each trivial specialisation x x is the ring V = k(T ), i i xi xi j6=i

and for i = 1, . . . , n, considering the specialisation x0 xi, we pick some uni- formising element from the set T x0 \ T j6=i and divide the remaining elements by

it. But we see that T x0 \ T j6=i = {txi }, so there is nothing to do. Then, we have that

V x x = k(T j6=i)[tx ] . 0 i i htxi i Comparison of finite constructions 111

The situation thus begins to look a lot more like that of the ring A in Lewis’s construction. We first observe that the ring V has unique maximal ideal x0 xi

m = t V , x0 xi xi x0 xi with corresponding quotient map

π : V V /m = k(T ). x0 xi x0 xi  x0 xi x0 xi j6=i

We form

\ R ↑ = k(T j6=i)[txi ]ht i x0 xi i

= k[tx ,..., tx ] c ; 1 n (htx1 i∪···∪htxn i)

finally,

Y R = {f ∈ R ↑ | ∀x ∈ X, y x then f = π (f )} X x > y x y x x∈X Y = {f ∈ R ↑ | f = π (f ) for all i 1} x xi x0 xi x0 > x∈X ∼ = {r ∈ R ↑ | r mod mx0 xi ∈ R ↑ = k(T j6=i) for all i 1} x0 xi >

= R ↑ x0

= k[tx ,..., tx ] c . 1 n (htx1 i∪···∪htxn i)

Ultimately though, as we will see, it will not be especially important which ring we choose to represent a finite 1-dimensional irreducible spectral space (or, that is, which subbasis we choose for the topology on a finite spectral space); note that for the purposes of Lewis’s construction, we may always take the field K on which we apply our induction hypothesis to be the field of fractions of whichever 1-dimensional domain we are using as a “building block”. Note that whenever we apply either of the techniques of Lewis’s construction, if B is the pre-existing ring having certain spectral structure created or assumed 112 4. Comparison of constructions

to exist from an earlier stage of our procedure, that the ring D that we obtain after having applied the technique in question can always be taken to be a subring of B. At the very first step of the construction the ring B is a finite product of fields corresponding to the minimal elements of X. Compare this to the observation 3.10 that the ring RX given by Ershov’s construction can always be taken to be a subring of the product R ↑ × · · · × R ↑ ⊆ k(T ↑ ) × · · · × k(T ↑ ) x1 xn x1 xn

where x1, . . . , xn are the minimal elements of X. Indeed it is not difficult to

convince oneself that, given a fixed poset X, the construction of LX takes place

entirely within the ring k(T x1 ) × · · · × k(T xn ) for T xi as given above.

Gluing of 1-dimensional domains

Encouraged by this apparent similarity between the respective constituent building blocks of each of our constructions, we proceed to investigate how far the similarity is maintained when we apply the operations by which the ring

LX is built up from a finite spectral space X. We first investigate the process by which we glue the spectrum of a 1- dimensional domain onto a maximal ideal, as described from page 79 onwards. We take X to be a finite spectral space of dimension n occurring at the n-th stage of our iterative procedure 3.6, such that all the height-n points of X have

a unique immediate predecessor in X. Let Xht n = {z1, . . . , zm} be the points of maximum height of X. For simplicity, let us suppose that there is in fact

a common immediate predecessor y for the zi, that is, that Xht n = y  . Let

0 X = X \ Xht n. We firstly choose a domain A having m maximal ideals, by either of the methods discussed in the preceding section, so that Spec A ∼= y↑. We then set K = Q(A) and take a K-algebra B with Spec B ∼= X0 as guaran-

teed by the inductive hypothesis of 3.6, and pick out my the maximal ideal of

0 B corresponding to the point y of X . Finally, we take the ring LX to be the Comparison of finite constructions 113

fibre product of A and B as in the following diagram, where all the maps are the obvious ones,

ψ0 LX =A ×C B  A ∩ ∩

ϕ0 ϕ

? ψ ? B  B/my = C;

0 we then have that Spec LX has the structure of X with the point y replaced by ↑ ∼ the space y ; that is, we have Spec LX = X. Note that as the method is essentially topological in nature it doesn’t matter which representation of a 1-dimensional irreducible space we use as our ring A, provided we may still obtain the relevant maps ϕ and ψ (—for example, we require that Spec ϕ maps the single point of Spec C to the minimal point of Spec A). We then compare this situation with the rings defined by the Ershov con- struction. Writing Y = y↑ to avoid notational difficulties, we are provided with rings RX , RX0 and RY whose respective prime spectra are homeomorphic to X,

0 X and Y . We will then compare the Ershov ring RX with the fibre product of the rings RY and RX0 over the appropriate maps: we want to take the quotient

0 of RX0 by the maximal ideal my corresponding to the point y of X ; and we want an inclusion of RY into this quotient ring. To this end, we define RX and

0 RY over some chosen field k, and then define RX0 to be the ring defined on X over the base field K = k(tz1 , . . . , tzn ); note this is similar to how we chose B above to be a K-algebra for K the field of fractions of our ring A.

RY ∩

? RX0  RX0 /my 114 4. Comparison of constructions

We also choose to take the subbasis BX , BX0 etc. as defined in the previous section to be the subbasis used in the definition of all the components of our constructions, that is, we take the subbasis for the coarse lower topology on

↑ c a finite poset that consists of the collection BX = {(x ) | x ∈ X}. This is to facilitate the comparison of the rings we obtain. We thus re-label the rings

0 corresponding to the spaces X, X and Y as RX , RX0 and RY respectively.

Every constituent component of the construction of RX0 will be marked

0 0 0 with a prime mark, e.g. as R ↑ , V , π etc., to distinguish them from x x y x y

those defined in the construction of R , which will be unmarked, as R ↑ , V , X x x y π . It will be important to distinguish these objects as they will often be x y defined on the same points and specialisation relations, since X0 is a sub-poset of X; however the “ambient” poset they are defined within will significantly affect the definitions, as we will see. Our first observation is that we have:

Proposition 4.1. For RX the Ershov ring defined on X from the subbasis BX

0 over some chosen field k, and for RX0 the Ershov ring defined on X from the

0 0 subbasis BX over the field k(tz1 ,..., tzn ), we have that RX X0 ⊆ RX .

We remind ourselves of the definition of RX X0 as immediately preceding theo- 0 rem 3.11: it is simply the projection of RX onto X . Then, our claim is that we

0 0 0 have R ↑ ⊆ R ↑ for every x ∈ X , and all conditions of the form π (fx) = fw x x x w 0 0 for x w in X are satisfied by f ∈ RX ; note that as π is not in general the x w same as the map π , some justification is required. We unpick the definitions x w 0 0 as follows. Given x ∈ X , the definitions of Rx↑ , Rx↑ respectively are as follows:

\ R ↑ = V , x x w w x in X 0 \ 0 R ↑ = V . x x w w x in X0 Comparison of finite constructions 115

\ \ Note that V ⊆ V ; so that we may satisfy our x w x w w x in X w x in X0 0 claim if it holds when we compare just those rings V x w and V where x x w and w are in X0. We see the definitions:

T ↓c \ T c x w  V = k(T ↓c , )[t ] ; x w w x htxi tx 0 0 c ↓c 0 0 T x \ T w  V = K(T ↓c , )[tx] , x w w htxi tx

↓ c where the hopefully suggestive notation is defined as T w↓c = {tv | v ∈ (w ) }; T c \T c x↓ tv w  ↓ c c = { | v ∈ (x ) \ (w  ) , v 6= x} where down-sets and complements tx tx are taken in X; the prime-marked versions have identical definitions except with down-sets and complements taken in X0. We then see that for all w ∈ X such

↓ c that w 6= zi we have that {z1, . . . , zn} ⊆ (w ) , so that tz1 ,..., tzm are units in V for all x and w in X0; thus justifying our particular choice of subbasis. But x w 0 then we in fact see that V x w and V define identical rings for all x and w x w 0 \ \ 0 0 0 in X , so V x w = V = R ↑ for all x ∈ X . Furthermore x w x w x in X0 w x in X0 0 0 note that V x w = V means that the maps πx w, π are likewise identical x w x w 0 0 for all x, w ∈ X , and so the condition on RX0 that π (fx) = fw for all f ∈ RX0 x w is satisfied for all f ∈ R 0 by the analogous condition π (f ) = f on R . X X x w x w X

Hence we have RX X0 ⊆ RX0 as required.

↑ Now recall from 3.11 that, since Y = y is up-closed, we have that RX Y =

RY . Then, as we may identify RY with the subring {fy | f ∈ RY }, or equiv- alently, from the above equality, the subring {fy | f ∈ RX } of Ry↑ , then since

0 0 Ry↑ ⊆ Ry↑ , we may identify RY with a subring of R y↑ . This suggests the following comparison with the construction of Lewis: ob- serve that projection onto the y-component is the (surjective, since y is maximal

0 0 in X ) homomorphism evy : RX0  Ry↑ . Suppose we wished to consider only those elements f of RX0 for which fy ∈ RY (via the identification discussed in 116 4. Comparison of constructions

the previous paragraph). That is, we wish to consider a ring

R0 = {f ∈ RX0 | fy ∈ RY }

= {f ∈ RX0 | evy(f) ∈ RY }

= {f ∈ RX0 | f mod ker(evy) ∈ RY },

where we notice that ker(evy) is the maximal ideal my of RX0 corresponding to

0 0 0 the point y in X . That is, R0 = {f ∈ RX0 | fy ∈ RY } is exactly the fibre

product of RX0 and RY over my.

We may reverse the identification of RY with a subring of Ry↑ by sending f to Q π (f ); thus we may define R (isomorphically) on the whole of y x∈Y y x y 0 X by defining for each f ∈ R that f = π (f ). Since for every f ∈ R we 0 zi y zi y X

have that fy ∈ RY , we have that RX ⊆ R0. We find in fact that the inclusion is strict. To show this, we examine in more detail the distinction between the rings 0 Rx↑ and Rx↑ . In the case that x 6 y, we have that

\ 0 \ R ↑ = V = R ↑ ∩ V . x x w x x zi w x in X i ↓ c To analyse the rings V we first consider the set W = (y ) \ y  ; the x zi indeterminates corresponding to such points are always units in the definition of a ring V ; we label the corresponding set T . Then we can see from x zi W the definition and from comparison with the similar situation in the preceding section that

T x↓c \ T y↓c V x z = k(T W , T j6=i, )[tz ] , i i htzi i tzi

for T j6=i = {tzj | j 6= i}, where, as in the earlier section, we allow ourselves a

different choice of uniformising element for the ring, choosing tzi as opposed to

tw ↓ c ↓ c tx. Then, we can see that no element of the form , say, for any w ∈ (x ) \(y ) t2 zj (e.g. for w = y) can appear in V , and so cannot appear in R ↑ . x zj x Comparison of finite constructions 117

For f ∈ R we have that f ∈ R ↑ , and that we have f = π (f ), where 0 y y y x y x

R ↑ = k[tz ,..., tz ] c , y 1 m (htz1 i∪···∪htzm i) and π : V → K , x y x y x y that is,

T ↓c \ T c T ↓c \ T c x y  x y  π : k(T ↓c , )[t ] → k(T ↓c , ), x y y y htyi y ty ty so that f = f + g where g ∈ m = t V . But note that in R , that is, x y x y y x y 0 0 ty in Rx↑ such that f ∈ RX0 and fy ∈ Ry↑ , there is nothing to prevent fx = 2 , tz1 ty say (and then fy = πx y( 2 ) = 0 ∈ R ↑ ), in which case fx ∈/ V x z and so t y 1 z1 fx ∈/ Rx↑ ; hence RX ( R0.

It might be contended that by choosing the ring RX0 to be constructed over a different field than the rings RX and RY , that we are examining a somewhat artificial situation. However I believe the similarity drawn out between the conditions by which the ring RX is defined; the role of maximal ideals which correspond to evaluating functions at maximal points of a space; and the fibre product method of Lewis’s construction, is sufficiently evident to be of interest.

Joining of maximal points

We likewise consider the joining of maximal points. Suppose as in the expo- sition of Lewis’s iterative construction that from the poset X we have obtained a poset X0 having the following properties: firstly, that we may obtain X from X0 by joining certain maximal points of X0 to a single point; and secondly, each of these maximal points of X0 under consideration has a unique immedi- ate predecessor in X0. In Lewis’s construction, we then apply the joining of maximal points procedure to a ring B (—in fact, a K-algebra for any choice of 118 4. Comparison of constructions

0 K) having spectrum homeomorphic to X , in order to obtain a ring LX with spectrum homeomorphic to X. We compare the rings B, LX with the rings

RX0 , RX provided by the Ershov construction. For simplicity suppose we are joining a single set of maximal points in X0

0 to a single point in X. Then, if z1,1, . . . , z1,m are the maximal points of X corresponding to the maximal point z1 of X, each with immediate predecessor

0 y1, . . . , ym in X respectively, and if n1,1,..., n1,m are the maximal ideals of B corresponding to the maximal points in question, then we map the field K diagonally into, and the ring B surjectively onto, the product C = B/n1,1 ×

· · · × B/n1,m. To carry out the joining procedure we then take the fibre product over these ∼ maps, giving the ring LX = K×C B = {b ∈ B | b mod n1,1 = ··· = b mod n1,m = ∼ a ∈ K}, such that Spec LX = X.

ψ0 LX =K ×C B  K ∩ ∩

ϕ0 ϕ

? ψ ? B  B/n1,1 × · · · × B/n1,m = C

0 We now compare the rings RX0 , RX defined from X , X by the Ershov construction. The definitions differ in two ways: firstly RX0 is defined as a product over points in X0, that is,

Y 0 Y 0  0 0 RX0 ⊆ Rx↑ = Rx↑ × R ↑ × · · · × R ↑ , z1,1 z1,m x∈X0  x∈z1 whereas Y  RX ⊆ Rx↑ × R ↑ ; z1  x∈z1 Comparison of finite constructions 119

and where the definition of RX0 has compatibility criteria

0 π (fy ) = fz y1 z1,1 1 1,1 0 π (fy ) = fz y2 z1,2 2 1,2 . .

0 π (fy ) = fz , ym z1,m m 1,m these correspond to the criteria

π (f ) = π (f ) = ··· = π (f ) = f y1 z1 y1 y2 z1 y2 ym z1 ym z1 in RX . The situation then looks somewhat reminiscent of the description of our

fibre product just above. We may think one way to proceed is, defining by ϕi

0 the isomorphism of fields ϕi : R ↑ = k(tz1,i ) → k(tz1 ) = Rz↑ given by tz1,i 7→ tz1 , z1,i 1 then taking

0 R0 = {f ∈ RX | ϕ1(fz1,1 ) = ··· = ϕm(fz1,m )}, we can identify R0 with a product

Y 0 R0 ⊆ R ↑ × R ↑ x z1  x∈z1

by defining fz1 = ϕi(fz1,i ) ∈ R ↑ z1 0 The difficulty is that elements in Rx↑ may refer to any of the indeterminates tz1,1 , . . . , tz1,m where those in Rx↑ only refer to the indeterminate tz1 , and the stipulation that ϕ1(fz1,1 ) = ··· = ϕm(fz1,m ) is not a strong enough condition on f to allow us to define a map on the indeterminates tz1,i that would allow us

0 to compare an element f ∈ R0 with an element f ∈ RX . However, allowing ourselves to be slightly more “impressionistic” for a moment (—and allowing ourselves to notice, with reference to the analogous argument of the previous

0 section, that R0 is indeed the fibre product {f ∈ RX | f mod ker(evz1,1 ) = ··· = 120 4. Comparison of constructions

f mod ker(evz1,m )}, at least under the identification of each field R ↑ = k(tz1,i ) z1,i with k(tz1 )), we see that there are in some sense a great deal more elements in

R0 than in RX , in exactly the sense that R0 has indeterminates tz1,1 , . . . , tz1,m where RX has only tz1 . We now leave impressionism strictly behind us for the remainder of this thesis.

Comparison of constructions on a general spectral space

We now proceed to explore how the ‘Ershov ring’ RX associated to a finite spectral space X may be defined in the case that X is no longer finite. Recall we began our construction of RX in an identical way to how we began the construction of the Hochster ring HX (—initially, the ring AX ) associated to a general spectral space X: we identified a suitable subbasis BX for the topology of X and formed corresponding sets of indeterminates TX and characteristic functions χ taking values in the polynomial ring k[TX ] over some chosen field k. Our finite construction began to proceed along distinct lines to that of Hochster when, instead of extending to elements defined by having some prop- erty with respect to an ‘index’ of valuations, we specified valuation rings V x y corresponding to each specialisation x y in our space X and took the condi- tion on our final ring RX to be “compatibility” of values taken by functions at points x and y via the canonical quotient maps corresponding to these valuation rings. We recall that the definition of V in our original exposition (on page x y

98) made reference to the order relation on X via the sets T ↑ , T and so on. y x  However, as we will now see, we can without difficulty extend the definition of V beyond the finite setting, in fact using concepts we have already defined: x y Comparison of constructions on a general spectral space 121

◦ Let X be any spectral space, and take BX ⊆ K(X) an open subbasis for X,

TX = {tU }U∈BX a collection of indeterminates indexed by BX , and characteristic functions χU : X → k[TX ] defined by   tU if x ∈ U χU (x) =  0 if x∈ / U.

Recall that given a point x ∈ X we defined the set of indeterminates Tx ⊆ TX to be the set Tx = {tU ∈ TX | x ∈ U}. We note that in the case that X is finite

↓ and BX is the basis BX = {y | y ∈ X} as chosen at the start of our exposition of Ershov’s construction then Tx is exactly the set Tx↑ ; furthermore as noted at this stage in both the Hochster and the Ershov procedure, given a specialisation x y (—in any spectral space X) we have that Ty ⊆ Tx. Thus we make a general definition of the valuation ring V corresponding to the specialisation x y x y in a general spectral space X (—as our definitions are consistent with those previously given in the finite case we make no change to the notation of the objects defined): Firstly, in the case that y = x we define the ring V to be the improper x x valuation ring V = k(T ). x x x

In the case that y 6= x we have that Ty ( Tx (since distinct points are topologically distinguishable, and so separated by some subbasis set); we then

take some U0 such that x ∈ U0, y∈ / U0 (i.e. such that tU0 ∈ Tx\Ty), then, writing ∧ Tx \Ty tU t0 for the indeterminate tU we define the set = { | tU ∈ Tx \ Ty, U 6= 0 t0 t0 U }. We then define the discrete valuation ring V to be 0 x y T ∧ \ T V = k(T , x y )[t ] , x y y 0 ht0i t0 in a manner practically identical to the finite case. Note that the choice of U0 in fact in no way affects the definition of the ring V : for any choice of U 0 x y 122 4. Comparison of constructions

0 0 such that x ∈ U , y∈ / U , the indeterminate tU 0 ∈ Tx \ Ty has the same value as

tU0 t0 (since is a unit in Vx y so has value 0), and so tU 0 in fact has value 1, and t0 0 is hence a uniformiser for the valuation ring; that is, we could replace U0 by U in all our definitions and we would obtain an identical valuation ring V . x y We again have a unique maximal ideal m = t V and canonical quo- x y 0 x y tient map π : V → V /m = K . x y x y x y x y x y

The comparison of valuation rings

Recall that in chapter 2 we defined a discrete valuation v : k(T ) → ∪ {∞} x y x Z corresponding to each specialisation x y in a spectral space X. We have just seen how in analogy with Ershov’s construction described in the preceding chap- ter, we may define a valuation ring V corresponding to each specialisation x y x y in X, for a general spectral space X. We see in fact that V as so x y defined is the valuation ring of the valuation v . x y Recall the valuation v was defined first on k[T ] by x y x   1 y∈ / U, that is, tU ∈/ Ty; • v (t ) = x y U  0 y ∈ U, or, tU ∈ Ty;

• v (tα1 . . . tαn ) = P α for those i such that v (t ) = 1; and x y 1 n i x y i

• v (P λ m ) = min({v (m )}) for m distinct monomials in T and x y i i x y i i x × λi ∈ k ;

It then has a uniquely defined extension to a discrete valuation v : k(T ) → ∪ {∞}. x y x Z Then we see

Proposition 4.2. Given v : k(T ) → ∪ {∞}, V ⊆ k(T ) as defined x y x Z x y x above, we have that V is the valuation ring of the discrete valuation v , x y x y that is, V = {r ∈ k(T ) | v (r) 0}. x y x x y > Comparison of constructions on a general spectral space 123

Let V = {r ∈ k(T ) | v (r) 0} be the valuation ring of v . For t ∈ T x x y > x y U y we have that v (t ) = 0, so that t is a unit in the valuation ring; then to x y U U ∧ 0 0 Tx \Ty simplify our argument let k = k(Ty), then Vx y = k ( )[t0] and we need t0 ht0i ∧ 0 Tx \Ty to show that k ( )[t0] = V . This modification allows us to treat tU ∈ Ty t0 ht0i as coefficients in our base field. Since v is a valuation on k(T ) = k(T )(T \ T ) = k0(T \ T ), every x y x y x y x y f 0 r ∈ V is of the form g for f, g ∈ k [Tx \ Ty], g 6= 0. f Take some g ∈ V . We pick a term of least degree of g, which we will write

α1 αn 0 × as g0 = γ0t1 . . . tn for γ0 ∈ (k ) and ti ∈ Tx \ Ty (where of course we may have n = 0 so that g0 = γ0). Let g g0 = . g0

∧ ∧ 0 0 Tx \Ty 0 Tx \Ty 0 We claim g ∈ k ( )[t0]\ht0i. For convenience we write k ( )[t0] = R . t0 t0 Writing g termwise, we have

γ m γ m g g0 = d d + ··· + 1 1 + 0 g0 g0 g0

0 × for monomials mi and coefficients γi ∈ (k ) , where each mi has degree greater or equal to the degree of g0 (and where in describing g0 as a term of g we implicitly mean that each monomial mi is distinct from the monomial of g0). Then we write

0 00 mi = mimi

00 where mi has degree equal to the degree of g0, so that

00 00 0 γdmd 0 γ1m1 0 g = md0 + ··· + m1 + 1. g0 g0

Then

00 0 β1 0 βn0 γ m γ t . . . t 0 i i = i 1 n , α1 αn g0 γ0t1 . . . tn 124 4. Comparison of constructions

P P tU 0 t0 0 where j0 βj0 = j αj = N. Note for all tU 6= t0 we have ∈ R and ∈ R ; t0 tU t furthermore 0 = 1 ∈ R0. t0 Then

00 N 0 β1 0 βn0 γ m t γ t . . . t 0 i i = 0 i 1 n N α1 αn g0 t0 γ0t1 . . . tn 0 0 γ t t 0 t t = i 1 β1 · ... · n βn0 · 0 α1 · ... · 0 αn ∈ R0 γ0 t0 t0 t1 tn

0 0 0 0 for each i, and mi ∈ k[Tx \ Ty] ⊆ R , so that g ∈ R . Furthermore as each

monomial mi is distinct from the monomial of g0, we do not obtain any new constant terms; so g0 has constant term equal to 1, and so lies outside the ideal

generated by t0. Writing f termwise, we have

f f λd0 Md0 λ1M1 λ0M0 = 0 = 0 + ··· + 0 + 0 . g g0g g0g g0g g0g

× for monomials Mj and coefficients λj ∈ k . Note as v (f) v (g) then every monomial M has degree greater than x y > x y j or equal to the degree of g0. Then write

0 00 Mj = MjMj

00 where Mj has degree equal to the degree of g0, so that

00 0 00 0 00 0 f λd0 Md0 Md0 λ1M1 M1 λ0M0 M0 = 0 + ··· + 0 + 0 . g g0 g g0 g g0 g

Then to verify our first claim, we use exactly the procedure we have just

00 demonstrated to observe that, as g0 and each Mj have the same degree, then 00 00 0 λj Mj 0 0 0 0 λj Mj Mj 0 lies in R . Furthermore, M ∈ k[Tx \Ty] ⊆ R . Then let f = ∈ R . g0 j j g0 Then 0 0 0 f f 0 + ··· + f + f = d 1 0 , g g0 0 0 0 0 0 0 f 0 with f 0 + ··· + f + f ∈ R and g ∈ R \ ht i; so ∈ R = V . d 1 0 0 g ht0i x y Comparison of constructions on a general spectral space 125

∧ 0 Tx \Ty Conversely note that the elements with non-zero value in k ( )[t0] are t0 exactly those in the ideal generated by t0; therefore we may divide by any element outside this ideal without changing the value assigned, so that every

∧ 0 Tx \Ty element in k ( )[t0] = Vx y has value 0 and so V = Vx y as claimed. t0 ht0i >

The next stage of Ershov’s construction defines a ring Rx↑ for each x ∈ X, being the intersection of the valuation rings

\ R ↑ = V . x x y y x We maintain this definition (and notation) in the general case, noting that in general the ring Rx↑ is now no longer a finite intersection of discrete valuation rings.

We note that for all elements h of Hochster’s ring HX , by proposition 2.15 we have that v (h ) 0 for all specialisations x y in X; that is, for x y x > every h ∈ H we have that h lies within the valuation ring of v for all X x x y specialisations x y, which is the ring V by the preceding proposition. But x y then h is in the intersection of the valuation rings V for y x, so that x x y hx ∈ Rx↑ at every x ∈ X for all h ∈ HX . Q Then HX ⊆ x∈X Rx↑ . The Ershov ring RX is also originally defined as a Q subring of x∈X Rx↑ , being the ring

Y R = {f ∈ R ↑ | ∀x ∈ X, y x then f = π (f )}. X x > y x y x x∈X

We see that in our updated general setting we still have

∧ ∧ Ty \ Tz Tx \ Ty V = k(T ↑ , )[t ] ⊆ k(T ) ⊆ k(T , ) = K y z z 0y ht0y i y y x y t0y t0x

(where t0y ∈ Ty \ Tz, t0x ∈ Tx \ Ty are the indeterminates corresponding to our chosen basis sets U0y such that y ∈ U0y , z∈ / U0y and U0x such that x ∈ U0x , 126 4. Comparison of constructions

y∈ / U0x respectively), for any x y and y z in X, so that for all x ∈ X and 0 0 for all y with x y we have R 0↑ ⊆ K 0 and it again makes sense to compare y x y

π 0 (f ) ∈ K 0 with f 0 ∈ R 0↑ ; therefore we may essentially maintain our x y x x y y y definition for RX as above in the case that X is any spectral space, giving the general Ershov ring RX corresponding to a spectral space X, defined as

Y R = {f ∈ R ↑ | ∀x ∈ X, y x then f = π (f )}. X x y x y x x∈X

The comparison of admissibility conditions

We now consider how the condition of admissibility to RX as given above Q applies to elements h ∈ HX ⊆ x∈X Rx↑ . In fact we have for all h ∈ HX that π (h ) = h for all x ∈ X and x y; that is, x y x y

Theorem 4.3. Given any spectral space X and the rings HX as defined on page

68 and RX the general Ershov ring as defined above, we have HX ⊆ RX .

Q We have that HX ⊆ x∈X Rx↑ ; we must show that for all h ∈ HX we have π (h ) = h for all x ∈ X and x y. We prove this by an induction on the x y x y construction of HX from the rings Ai as defined on page 68.

Given h ∈ HX we have that for some n ∈ N, that h is contained in the ring

An obtained at the n-th stage of our inductive definition.

We apply an induction to the construction of the ring Ai+1 from the ring A as follows: if A is such that for all f ∈ A we have that π (f ) = f , we i i i x y x y prove that for all h0 ∈ A we have that π (h0 ) = h0 . i+1 x y x y

Our base case is the ring A0 = AX defined on page 54; we show that for all Q f ∈ A ⊆ R ↑ we have that π (f ) = f . X x∈X x x y x y P × Given a general f = i λimi in AX for λi ∈ k and mi monomials in k[XX ] then   fx = λ1m1 x + ··· + λdmd x, Comparison of constructions on a general spectral space 127 where   α1 α2 αN λitU tU . . . tU x ∈ Uj for all j λ m  = λ χα1 χα2 . . . χαN  = 1 2 N i i x i U1 U2 UN x  0 otherwise and   fy = λ1m1 y + ··· + λdmd y, for   α1 α2 αN λitU tU . . . tU y ∈ Uj for all j λ m  = λ χα1 χα2 . . . χαN  = 1 2 N i i y i U1 U2 UN y  0 otherwise.

Note that as x y then we have y ∈ Uj for all Uj in some particular term of f only if we have x ∈ Uj for all those Uj, so that for any term λimi for which   λimi y 6= 0 we necessarily have λimi x 6= 0. We examine π (f ) = π (λ m  ) + ··· + π (λ m  ) termwise, x y x x y 1 1 x x y d d x  considering those terms for which λimi x 6= 0. Observe that for an indeter- minate tU ∈ Tx, then y ∈ U ⇐⇒ tU ∈ Ty ⇐⇒ tU ∈ k(Ty), so that π (t ) = t ; conversely y∈ / U ⇐⇒ v (t ) > 0 ⇐⇒ t ∈ m ⇐⇒ x y U U x y U U x y π (t ) = 0. Then we have x y U

π (λ m  ) = π (λ tα1 tα2 . . . tαN ) = λ π (tα1 ) · ... · π (tαN ) x y i i x x y i U1 U2 UN i x y U1 x y UN   α1 αN λitU · ... · tU if y ∈ Uj for all j = 1 N  0 otherwise  = λimi y, thus π (f ) = f as required. x y x y

We now make the assumption for the purposes of an induction that Ai is such that for all f ∈ A we have that π (f ) = f . i x y x y 128 4. Comparison of constructions

Every element of Ai+1 arises from some finite extension of Ai by elements f

k for f, g ∈ Ai satisfying the condition of 2.13. Therefore per the discussion g preceding theorem 2.18, it suffices to show that if we have π (f ) = f for x y x y f 0 0 0 every f ∈ Ai then we have that πx y(hx) = hy for all h ∈ Ai+1[ k ] for some g

f, g ∈ Ai satisfying the condition of 2.13. f 0 We consider a general element h in Ai[ k ], g

0 f n f  h = fn k + ··· + f1 k + f0 g g

for f0, . . . , fn ∈ Ai. We have

0 f n f   πx y(hx) = πx y (fn)x k + ··· + (f1)x k + (f0)x g x g x f  n f   = (fn)yπx y k + ··· + (f1)yπx y k + (f0)y, g x g x

by our inductive assumption and the properties of ring homomorphisms. f f 0 0    Then πx y(hx) = hy ⇐⇒ πx y k = k ; hence we reduce to g x g y f f

considering h = k ∈ Ai[ k ]. g g If h = 0 then f = 0 and so f = π (f ) = 0 so that h = 0, in which x x y x y x y case we have π (h ) = π (0) = 0 = h as required. x y x x y y Otherwise assume h = fx 6= 0. x gx In the case π (h ) = 0 we have equality: x y x

π (h ) = 0 ⇐⇒ h ∈ m ⇐⇒ v (h ) > 0 x y x x x y x y x ⇐⇒ v (f ) > v (g ) x y x x y x ⇐⇒ v (f ) > 0 x y x

⇐⇒ fy = 0

⇐⇒ hy = 0. Comparison of constructions on a general spectral space 129

Alternatively,

π (h ) 6= 0 ⇐⇒ h ∈/ m ⇐⇒ v (h ) = 0 x y x x x y x y x ⇐⇒ v (f ) = v (g ). x y x x y x

Then, since hx 6= 0 so that fx, gx 6= 0, then by the condition of 2.13 we must

1 1 1 have vx y(fx) = vx y(gx) = 0, so that ∈ Vx y. Then πx y( ) = = gx gx πx y(gx) 1 fx 1 fy , so that πx y(hx) = πx y( ) = πx y(fx) · πx y( ) = = hy as required. gy gx gx gy f 0 0 0 Then for every h ∈ Ai[ k ] we have that πx y(hx) = hy; hence by induction g every finite extension of Ai by admissible pairs f, g has this property. But then as every element in Ai+1 arises from some such finite extension of Ai by admissible pairs, we have π (h0 ) = h0 for all h0 ∈ A , which completes the x y x y i+1 induction.

Thus we have that for all n ∈ N, the ring An has the property that for all h ∈ A we have that π (h ) = h ; thus this property holds for all h ∈ H n x y x y X and the theorem is proved.

Two obvious questions strike us at this point. We have generalised the ring

RX from the finite setting to be defined on any spectral space X. When X is ∼ finite, we have seen that Spec RX = X. However firstly this result was proved by an induction on |X|, and furthermore one part of the proof relied on the fact that the ring R ↑ was a finite intersection of discrete valuations rings for x0 x0 a minimum element of X, in order to argue about the structure of the non- zero primes of RX . Since both these arguments are no longer relevant in the general setting, we may not expect that we still have that Spec X ∼= X for X any spectral space; however whether this does in fact hold remains a question to be decided one way or the other.

We have further seen that for any spectral space X, the Hochster ring HX is contained in the general Ershov ring RX . An obvious question to consider is 130 4. Comparison of constructions

whether we in fact have HX = RX ; note that if this is the case then indeed we ∼ ∼ must have Spec RX = X since we know Spec HX = X for all X. We revisit a basic example from earlier.

Given X = Spec Z then we saw (page 59) that the ring AX initially assigned to X by the Hochster construction was the ring

Y AX = { evq(f) | f ∈ k[XX ]} q∈Spec Z

where k[XX ]} consists of polynomial expressions over k of functions   tp q 6= hpi χp(q) =  0 q = hpi for positive primes p; we observed that AX can just be viewed as the ring of all polynomials in k[TX ] = k[t2, t3, t5, t7,... ] = Rh0i together with its projections

Rh0i 7→ Rp given by f 7→ f mod tp for p the prime number such that p = hpi (—a notational convention we will assume from now on). We consider when f, g ∈ A are such that v (f) v (g) for all p X h0i p > h0i p and v (f) = v (g) =⇒ v (g) = ∞ or 0. If g has non-zero constant h0i p h0i p h0i p term then v (g) = 0 for all p ∈ X so the inequality holds. Alternatively h0i p

if g has constant term equal to zero, we have gp = 0 ⇐⇒ g ∈ htpi; then if v (f) v (g) for all p then for any p such that g ∈ ht i we must have h0i p > h0i p p f ∈ ht i; furthermore, as v (g) is the greatest power of t dividing g and as p h0i p p v (g) > 0 then we must have v (f) > v (g) and so f is divisible by h0i p h0i p h0i p Q α some higher power of tp. Then we can take mg = tp for α the greatest power 0 0 of tp dividing g, over all p such that g ∈ htpi. Then we have elements f , g ∈ AX f f 0 0 g

where g = has non-zero constant term, such that k = k (since, as the 0 mg g g power of tp dividing f was always strictly greater than that dividing g, we have f 0 f 0 for f = that fp = 0 ⇐⇒ fp = 0). Thus the admissible elements k of mg g Comparison of constructions on a general spectral space 131

AX are exactly those for which g has non-zero constant term, and so hence the ring A1 obtained at the first stage of our extension is the ring k[t2, t3, t5, t7,... ] localised at the set of all g with non-zero constant term, that is, the set of those g not lying in any principal ideal htpi.

We now examine the general Ershov ring RX associated to X = Spec Z. The set Tq for each q ∈ X is Tq = {tp | p ∈ D(p)} = {tp | q 6= hpi} = TX \{tq} where q = hqi; we have Th0i = TX . For q 6= 0 let us write Tq = T6=q. Then

V = k(T ) and for p 6= h0i we have R ↑ = V = k(T ); we see p p 6=p p p p 6=p

V = k(T )[t ] , h0i p 6=p p htpi and then \ R ↑ = V = k[T ] S c . h0i h0i p X ( htpi) p∈X

Then we have

Y R = {f ∈ R ↑ | ∀p ∈ X, q p then f = π (f )} X p q p q p p∈X Y = {f ∈ R ↑ | f = π (f ) for all p ∈ X} p q h0i q h0i p∈X ∼ = {r ∈ Rh0i↑ | (r mod tpk(T6=p)[tp]htpi) ∈ k(T6=p) for all p}

= Rh0i↑

S c = k[TX ]( htpi) .

But this is exactly the ring A1 obtained at the first stage of our extension procedure, and so (since HX ⊆ RX ) we have A1 = HX = RX , and furthermore ∼ we must have that Spec RX = X. We now proceed to give an example showing that the general conclusion to both of our questions is negative. 132 4. Comparison of constructions

 The example of X = Spec Z inv

 We examine the ring RX where X = Spec Z inv. We first pick a subbasis for X. We may take the complements of a subbasis for the space Spec Z. Let

DZ(n) denote the basic open set of Spec Z given by DZ(n) = {p | n∈ / p}.

Write Dinv(n) for its complement, Spec Z \ DZ(n) = VZ(n) = {p | n ∈ p}.  The sets Dinv(p) for p ∈ N prime form an open subbasis for X = Spec Z inv; observe Dinv(p) = VZ(p) = {q ∈ X | p ∈ q} = {hpi}. We take a set of

indeterminates TX = {tDinv(p) | p ∈ N prime} where we will write tp for tDinv(p); the set of indeterminates Tq corresponding to a point q ∈ X is the set Tq =

{tp | q ∈ Dinv(p)} = {tp | p ∈ q} = {tq} where q is the positive prime number such that q = hqi; except where q = h0i, in which case Th0i = {tp | p ∈ h0i} = ∅.

In X our non-trivial specialisations are p h0i for p 6= h0i; we examine the corresponding valuation ring V . This involves picking some element in p h0i

Tp \ Th0i and dividing any remaining indeterminates in Tp \ Th0i by this element; but Tp \ Th0i = Tp \ ∅ = {tp} for p the positive prime such that p = hpi, so there is nothing to do. Then

V = k[t ] , p h0i p htpi and we see \ R ↑ = V = V = k[t ] ; p p q p h0i p htpi q p and

Rh0i↑ = k(Th0i) = k(∅) = k.

Q Then RX = {f ∈ p∈X Rp↑ | fp mod htpi = fh0i}.

We show that there exist f, g ∈ RX with z(g) ⊆ z(f) but where f∈ / rad g ∼ (note this condition is necessary in all cases to have Spec RX = X; the fact that it was also sufficient in the case of those rings built up in chapter 2 does not Comparison of constructions on a general spectral space 133 concern us at present). Define   fp = tp ∈ Rp↑ for p 6= h0i f =  fh0i = 0 ∈ Rh0i, and   p gp = (tp) ∈ Rp↑ for p 6= h0i g =  gh0i = 0 ∈ Rh0i.

Clearly f, g ∈ RX , and z(g) = z(f) = {h0i}. Now f ∈ rad g ⇐⇒ ∃r ∈ RX ,

n n n f = rg for some power n ∈ N. But f = rg ⇐⇒ fp = rpgp at all p ∈ X, n p n−p in which case then we would have tp = rptp at all p, so that rp = tp . But to have r ∈ R we must have r ∈ V for all p, that is we must have that X p p h0i v (r ) = v (tn−p) = (n − p)v (t ) = n − p 0 for all p. But p is p h0i p p h0i p p h0i p > n unbounded, and so there is no n so that f = rg for any r ∈ RX .

∼ Thus Spec RX 6= X, and a fortiori HX ( RX . We could in fact see directly this latter fact, since we know that all h ∈ HX have finite image as functions

X → k(TX ) (that is, the set {hx}x∈X is a finite subset of k(TX )), and for both our functions f, g considered in our example we have that the sets {fx}x∈X and

{gx}x∈X are infinite, so that f, g ∈ RX \ HX . This suggests a possible modification to our definition of the general Ershov ring RX : let

RbX = {f ∈ RX | f has finite image as a function X → k(TX )}

= {f ∈ RX | the set {fx}x∈X is a finite subset of k(TX )}.

Note that in the case that X is finite this modification does not affect the ring defined, that is for finite X we have RbX = RX .

Clearly, as remarked, since every element of the ring HX has finite image as a function X → k(TX ), we still have HX ⊆ RbX . We claim that, in the case of 134 4. Comparison of constructions

 the example just given, at least—that is, in the case for X = Spec Z inv—the

ring RbX is in fact the Hochster ring HX constructed from X.

 Proposition 4.4. For X = Spec Z inv and for HX as defined in chapter 2,

RbX as defined above, then HX = RbX .

Maintaining the subbasis and corresponding indeterminates etc. as in the pre- ceding example, at the first stage of Hochster’s construction we define the ring Q Q AX = { p∈X evp(f) | f ∈ k[XX ]} ⊆ p∈X k[Tp], being the evaluation at each point p ∈ X of polynomial expressions formed over k from the characteristic

functions χp taking value tp at p = hpi and zero elsewhere. It is easy to see that

AX = {f ∈(k × k[t2] × k[t3] × k[t5] × ... ) | f contains finitely many distinct

indeterminates and the constant term at each co-ordinate is equal}.

We consider for which f, g ∈ A do we have v (f ) v (g ) for all X p h0i p > p h0i p p ∈ X and v (f ) = v (g ) =⇒ v (g ) = ∞ or 0. p h0i p p h0i p p h0i p In the case that g has non-zero constant term then v (g ) = 0 at all p h0i p p ∈ X, so that the inequality holds; note that g is only non-constant in finitely many places. Alternatively if g has constant term equal to zero, then for our

inequality to hold we must have that fp = 0 whenever gp = 0 (in particular f

has constant term equal to zero since gh0i = 0), furthermore wherever gp 6= 0

d 2 (that is, where gp = αdtp + ··· + α2tp + α1tp), where necessarily we have that v (g ) = min({i | α 6= 0}) > 0, we must have either f = 0, so that p h0i p i p

fp D 2 = 0, or, for fp = βDt + ··· + β2t + β1tp, as we must have v (fp) = gp p p p h0i min({j | β 6= 0}) > v (g ), we may cancel out a factor of α ti from both f j p h0i p i p p 0 fp fp 0 1 and gp (for i the least degree of a term of gp) so that = 0 where g = gp gp gp p i αitp 0 has constant term 1 (but where the constant term of fp is still zero, so that 0 fp 0 0 0 mod htpi = 0). Then clearly we can define functions f , g in AX so that gp Comparison of constructions on a general spectral space 135

f 0 f 0 0 k = k and g has constant term 1 (—we simply need to define gq in the case g0 g 0 0 that fq = 0; the natural choice is to define gq = 1). f f

That is the admissible functions k for which k is not already in AX are g g those for any f ∈ AX (—that is, a member of the product of polynomials rings which has equal constant term at each co-ordinate and only finitely many non-constant values) and for g with non-zero constant term (and again only f

finitely many non-constant values). Taking B = { k | for those admissible g Q f and g just described}, we set A1 = AX [B] ⊆ p∈X k(Tp).

The ring RbX consists of all h ∈ (k × k[t2]ht2i × k[t3]ht3i × k[t5]ht5i × ... ) such

0 0 f f0+α f f0+γ 0 0 that for hp = = and hq = 0 = 0 where f0, g0, f0, g0 have constant g g0+β g g0+δ α γ term equal to zero, that β = δ , and that f0, g0 etc. are non-zero for only finitely many p ∈ X (that is, there are only finitely many non-zero f0’s, g0’s); another

α way of stating this latter property is that hp is different to β only at finitely many p. Clearly every such h is obtained at the first stage of our extension of the ring AX to A1, and so (since HX ⊆ RbX ) we have A1 = HX = RbX as claimed.

The two examples we have seen lead us to four interrelated conjectures.

∼ (1) Spec RbX = X;

(2) HX = RbX ;

(3) If X has finitely many minimal elements then RX = RbX ;

(4) If X has finitely many minimal elements then RX = HX .

Cautionary note must however be taken that the examples thus far consid- ered are fairly elementary, in two different respects. Firstly the specialisation structure of the respective spaces Spec Z and (Spec Z)inv is almost as simple as it can be, outside of a finite setting. Secondly, the respective topologies on each 136 4. Comparison of constructions

space are at the “extremes” of the range of possible spectral topologies hav-

ing such a specialisation structure: the topology on Spec Z is the coarse lower topology on its specialisation poset, and the topology on (Spec Z)inv is the fine lower topology. This may obscure nuances which arise in either the definition of HX or of RX and RbX on a spectral space with more sophisticated topological structure. It is with some considerable regret that due to the short time be- tween the discovery of these lines of inquiry and the necessary completion date of this thesis, no concrete further progress has been able to be made on these questions.

Christopher F. Tedd Manchester, 2016. Bibliography

[1] M. F. Atiyah & I. G. Macdonald, Introduction to commutative algebra, Addison-Wesley, Reading, 1969.

[2] B. Banaschewski, Radical ideals and coherent frames, Comment. Math. Univ. Carolin., 37:349–370, 1996.

[3] N. Bourbaki, Alg`ebre commutative, Chapitre 6, Springer-Verlag, Berlin, 2006.

[4] M. Dickman, N. Schwartz & M. Tressl, Spectral spaces, book in preparation.

[5] D. E. Dobbs, M. Fontana & I. J. Papick, On certain distinguished spectral sets, Ann. Mat. Pura Appl. (4), 128:227–240, 1981.

[6] A. M. Doering & Y. Lequain, The gluing of maximal ideals—spectrum of a Noetherian ring—going up and going down in polynomial rings, Trans. Amer. Math. Soc., 260:583–593, 1980.

[7] Yu. Ershov, Spectra of rings and lattices, Siberian Math. J., 46:283–292, 2005.

[8] M. Fontana, Topologically defined classes of commutative rings, Ann. Mat. Pura Appl. (4), 123:331–355, 1980.

[9] R. Gilmer, An existence theorem for non-Noetherian rings, Amer. Math. Monthly 77:621–623, 1970.

[10] R. Gilmer, Multiplicative ideal theory, Dekker, New York, 1972.

[11] R. Hartshorne, Algebraic geometry, Springer-Verlag, New York, 1977.

137 138 Bibliography

[12] R. C. Heitmann, Prime ideal posets in Noetherian rings, Rocky Mountain J. Math., 7:667–673, 1977.

[13] R. C. Heitmann, Examples of noncatenary rings, Trans. Amer. Math. Soc., 247:125–136, 1979.

[14] M. Hochster, Prime ideal structure in commutative rings, Trans. Amer. Math. Soc., 142:43–60, 1969.

[15] I. Kaplansky, Commutative rings, Allyn and Bacon, Boston, 1970.

[16] J. P. Lafon, Alg`ebre commutative. Langages g´eom´etriqueet alg´ebrique, Her- mann, Paris, 1977.

[17] W. J. Lewis, The spectrum of a ring as a partially ordered set, J. Algebra., 25:419–434, 1973.

[18] W. J. Lewis & J. Ohm, The ordering of Spec R, Canad. J. Math., 28:820– 835, 1976.

[19] H. Matsumura, Commutative ring theory, Cambridge University Press, 1989.

[20] S. Lipschutz, General Topology, Schaum, New York, 1965.

[21] H. A. Priestley, Spectral sets, J. Pure and Appl. Alg., 94:101–114, 1994.

[22] H. A. Priestley, Intrinsic spectral topologies, in Papers on general topology and applications, Ann. New York Acad. Sci. 728 pp. 78–95, 1994.

[23] R. Wiegand & S. Wiegand, The maximal ideal space of a Noetherian ring, J. Pure and Appl. Alg., 8:129–141, 1976. Bibliography 139

[24] R. Wiegand, Prime ideal structure in Noetherian rings, in Ring theory II, Proceedings of the Second Oklahoma Conference, Lecture Notes in Pure and Appl. Math. 26 pp. 267–279, Dekker, New York, 1977.

[25] R. Wiegand & S. Wiegand, Prime ideals in Noetherian rings: a survey, in Ring and module theory, Trends Math. pp. 175–193, Springer, Basel, 2010.

[26] S. Wiegand, Intersections of prime ideals in Noetherian rings, Comm. Alg., 11:1853–1876, 1983.