YAlg 3
YAlg 3
LECTURES ON
ALGEBRA
/ Wen-Wei Li
Part 3
Wen-Wei Li
Chinese Academy of Sciences
Version: 2019-06-14
The cover page uses the fonts Bebas Neue and League Gothique, both licensed
under the SIL Open Font License.
Introduction 1
1 Warming up 3
1.1 Review on ring theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Localization of rings and modules . . . . . . . . . . . . . . . . . . . . . . 5
1.3 Radicals and Nakayama’s lemma . . . . . . . . . . . . . . . . . . . . . . . 9
1.4 Noetherian and Artinian rings . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.5 What is commutative algebra? . . . . . . . . . . . . . . . . . . . . . . . . . 12
2 Primary decompositions 15
2.1 The support of a module . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.2 Associated primes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.3 Primary and coprimary modules . . . . . . . . . . . . . . . . . . . . . . . 18
2.4 Primary decomposition: the main theorem . . . . . . . . . . . . . . . . . 19
2.5 Examples and remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
Bibliography 91
Index 93
Introduction
In the beginning, these lecture notes were prepared for the graduate course Algebra III
(ID: 011M4002Y) in Spring 2016, University of the Chinese Academy of Sciences. For
some reasons, it took place in the Yuquanlu campus instead of the Yanqi Lake campus
as it should be.
The course is a sequel to Algebra I (fields, modules and representations) and II (ho-
mological algebra). The topic of this Part III is commutative algebra, or more precisely
commutative ring theory. Each “Lecture” in these notes took roughly one week, say ap-
proximately four hours of lecture, but the materials were only partially covered. My
initial intention was to give a traditional course on commutative algebra as proposed
by the syllabus prescribed by UCAS. For various reasons, my plan failed. For exam-
ple, there are too few discussions on depth, regular sequences and Cohen–Macaulay
modules, too few applications of completions, and the computational aspects haven’t
been touched. Moreover, the homological aspect of commutative algebra is almost non-
existent in these notes, namely the Auslander–Buchsbaum Formula, the properties of
regular local rings, etc. Last but not least, the exercises herein are scarcely sufficient.
These notes were also used for the Enhanced Program for Graduate Study held at
the Beijing International Center of Mathematical Research, Peking University, during
Spring 2019 (course ID: 00102057).
As the title suggests, some backgrounds from the Part I are presumed, namely the
basic notions of rings, modules and their chain conditions, as well as familiarity with
tensor products and some Galois theory. We occasionally presume some basic knowl-
𝑖
edge of homological algebra, such as the functors Tor𝑖 and Ext .
Sometimes I made free use of the language of derived categories. This was indeed
covered in the preceding course Algebra II, and it should be common sense for the future
generations.
As the reader might have observed, these notes were prepared in a rush; certain
paragraphs have not been proofread yet and many proofs are silly. I am very grateful
to the students for various corrections and improvements, and I will try to polish these
notes in the future.
Conventions
Throughout these lectures, we consider only associative rings with unit 1, and the rings
and algebras are assumed to be commutative and nonzero unless otherwise specified.
The ideal generated by elements 𝑥1 , 𝑥2 , … in a ring 𝑅 is denoted by (𝑥1 , 𝑥2 , …) or some-
times ⟨𝑥1 , 𝑥2 , …⟩; the 𝑅-algebra of polynomials in variables 𝑋, 𝑌, … with coefficients in
⋅2⋅ Introduction
𝑅 is denoted by 𝑅[𝑋, 𝑌, …]. We write 𝑅× for the group of invertible elements in a ring 𝑅.
A ring without zero-divisors except 0 is called an integral domain, or simply a domain.
The localization of a ring 𝑅 with respect to a multiplicative subset 𝑆 will be written as
𝑅[𝑆−1 ].
For any sets 𝐸, 𝐹, let 𝐸 ∖ 𝐹 ∶= {𝑥 ∈ 𝐸 ∶ 𝑥 ∉ 𝐹}. The cardinality of 𝐸 is denoted by |𝐸|.
The usual logical connectives such as ∃, ∀, ∧, ∨ and so forth will occasionally be
used. Writing 𝐴 ∶= 𝐵 means that the expression 𝐴 is defined to be 𝐵.
We will use the standard notations ℤ, ℚ, ℝ, ℂ to denote the set of integers, of ra-
tional numbers, etc. Sans serif fonts are reserved for categories, such as Ab (abelian
groups) and Ring.
When denoting morphisms in a category by arrows, monomorphisms (resp. epi-
∼
morphisms, isomorphisms) will be indicated ↪ (resp. ↠, →).
Possible references
The reader is expected to have basic familiarity with groups, rings and modules, as
covered in my lecture notes on Algebra I. We will make use of some really elementary
homological algebra as our course proceeds — so keep calm.
Our main references will be [11] and [8]. The Bourbaki volumes [5, 3] serve as our
ultimate source. The readers are also encouraged to consult the relevant materials in
Stacks Project.
⋄ 𝐼 is prime if and only if 𝑅/𝐼 is an integral domain, i.e. has no zero divisors except
0;
⋄ 𝐼 is maximal if and only if 𝑅/𝐼 is a field; in particular, maximal ideals are prime;
Definition 1.1.1 (Local rings). The ring 𝑅 is called local if it has a unique maximal ideal,
semi-local if it has only finitely many maximal ideals.
Let 𝔪 be the maximal ideal of a local ring 𝑅. We call 𝑅/𝔪 the residue field of 𝑅. A
local homomorphism between local rings 𝜑 ∶ 𝑅1 → 𝑅2 is a ring homomorphism such
that 𝜑(𝔪1 ) ⊂ 𝔪2 . Consequently, local homomorphisms induce embeddings on the
level of residue fields.
They are called the spectrum and the maximal spectrum of 𝑅, respectively. The upshot is
that Spec(𝑅) comes with a natural topology.
Definition 1.1.3 (Zariski topology). For any ideal 𝔞 ⊂ 𝑅, set 𝑉(𝔞) ∶= {𝔭 ∈ Spec(𝑅) ∶
𝔭 ⊃ 𝔞}. Then there is a topology on Spec(𝑅), called the Zariski topology, whose closed
subset are precisely 𝑉(𝔞), for various ideals 𝔞.
Indeed, we only have to prove the family of subsets {𝑉(𝔞) ∶ 𝔞 ⊂ 𝑅} is closed un-
der finite union and arbitrary intersections. It boils down to the easy observation that
𝑉(𝔞) ∪ 𝑉(𝔟) = 𝑉(𝔞𝔟) (check this!) and ⋂𝔞∈𝒜 𝑉(𝔞) = 𝑉 (∑𝔞∈𝒜 𝔞), where 𝒜 is any
family of ideals.
Given a ring homomorphism 𝜑 ∶ 𝑅1 → 𝑅2 , if 𝐼 ⊂ 𝑅2 is an ideal, then 𝜑−1 (𝐼) ⊂ 𝑅1 is
also an ideal.
𝜑♯ ∶ Spec(𝑅2 ) ⟶ Spec(𝑅1 )
𝔭 ⟼ 𝜑−1 (𝔭)
⋄ Take 𝑅1 to be a subring of 𝑅2 and 𝜑 be the inclusion map, the map above becomes
𝔭 ↦ 𝔭 ∩ 𝑅1 .
Proof. If 𝑅 contains an infinite field 𝐹, the ideals are automatically 𝐹-vector subspaces
𝑟
of 𝑅. Since 𝐼 = ⋃𝑖=1 𝐼 ∩ 𝔭𝑖 whereas an 𝐹-vector space cannot be covered by finitely
many proper subspaces, there must exist some 𝑖 with 𝐼 ∩ 𝔭𝑖 = 𝐼.
Under the second assumption, let us argue by induction on 𝑛 that ∀𝑖 𝐼 ⊄ 𝔭𝑖 implies
𝑛
𝐼 ⊄ ⋃𝑖=1 𝔭𝑖 . The case 𝑛 = 1 is trivial. When 𝑛 ≥ 2, by induction we may choose, for
𝑛
each 𝑖, an element 𝑥𝑖 ∈ 𝐼 ∖ ⋃𝑗≠𝑖 𝔭𝑗 . Suppose on the contrary that 𝐼 ⊂ ⋃𝑗=1 𝔭𝑗 , then we
would have 𝑥𝑖 ∈ 𝔭𝑖 , for all 𝑖 = 1, … , 𝑛.
When 𝑛 = 2 we have 𝑥1 + 𝑥2 ∉ 𝔭1 ∪ 𝔭2 and 𝑥1 + 𝑥2 ∈ 𝐼, a contradiction. When 𝑛 > 2,
we may assume 𝔭1 is prime, therefore
𝑛 𝑛
𝑥1 + ∏ 𝑥𝑗 ∉ ⋃ 𝔭𝑖 ,
𝑗=2 𝑖=1
again a contradiction.
Exercise 1.1.6. The following construction from [8, Exercise 3.17] shows that the as-
sumptions of Proposition 1.1.5 cannot be weakened. Take 𝑅 = (ℤ/2ℤ)[𝑋, 𝑌]/(𝑋, 𝑌)2 ,
which has a basis {1, 𝑋, 𝑌} (modulo (𝑋, 𝑌)2 ) as a ℤ/2ℤ-vector space. Show that the
image 𝔪 of (𝑋, 𝑌) in 𝑅 is the unique prime ideal, and can be expressed as a union of
three ideals properly contained in 𝔪.
You should regard [𝑟, 𝑠] as a token for 𝑟/𝑠; the ring structure of 𝑅[𝑆−1 ] is therefore
evident. In brief, localization amounts to formally inverting the elements of 𝑆, whence
the notation 𝑅[𝑆−1 ]. Note that condition (c) guarantees 𝑅[𝑆−1 ] ≠ {0}.
Exercise 1.2.1. Given 𝑅 and 𝑆, show that 𝑟 ↦ 𝑟/1 yields a natural homomorphism
𝑅 → 𝑅[𝑆−1 ] and show that its kernel equals {𝑟 ∶ ∃𝑠 ∈ 𝑆, 𝑠𝑟 = 0}.
1∶1
Spec(𝑅[𝑆−1 ]) {𝔭 ∈ Spec(𝑅) ∶ 𝔭 ∩ 𝑆 = ∅} ⊂ Spec(𝑅)
𝔮 its preimage.
2. Take any 𝔭 ∈ Spec(𝑅) and 𝑆 ∶= 𝑅 ∖ 𝔭. From the definition of prime ideals, one
infers that 𝑆 is a multiplicative subset of 𝑅. The corresponding localization is
denoted by 𝑅 → 𝑅𝔭 ∶= 𝑅[𝑆−1 ]. We see from (1–1) that
MaxSpec(𝑅𝔭 ) = {𝔭𝑅𝔭 };
in particular, 𝑅𝔭 is a local ring with maximal ideal 𝔭𝑅𝔭 . This is the standard way
to produce local rings; we say that 𝑅𝔭 is the localization of 𝑅 at the prime 𝔭.
(a) 𝑅 = ℤ, and we localize at the prime ideal (𝑝) where 𝑝 is a prime number.
As in the case of rings, we shall write 𝑚/𝑠 instead of [𝑚, 𝑠]. It is an 𝑅[𝑆−1 ]-module,
equipped with a natural homomorphism 𝑀 → 𝑀[𝑆−1 ] of 𝑅-modules. This yields a
functor: for any homomorphism 𝑓 ∶ 𝑀 → 𝑁 we have a natural 𝑓 [𝑆−1 ] ∶ 𝑀[𝑆−1 ] →
𝑁[𝑆−1 ], mapping 𝑚/𝑠 to 𝑓 (𝑚)/𝑠; furthermore 𝑓 [𝑆−1 ] ∘ 𝑔[𝑆−1 ] = (𝑓 ∘ 𝑔)[𝑆−1 ] whenever
composition makes sense.
∼
A slicker interpretation is to use the natural isomorphism 𝑀[𝑆−1 ] → 𝑅[𝑆−1 ] ⊗ 𝑀
𝑅
which maps 𝑚/𝑠 to (1/𝑠) ⊗ 𝑚. Hereafter, we shall identify 𝑀[𝑆−1 ] and 𝑅[𝑆−1 ] ⊗ 𝑀
𝑅
without further comments.
In the same vein, we may define 𝑀[𝑓 −1 ] and 𝑀𝔭 , for non-nilpotent 𝑓 ∈ 𝑅 and
𝔭 ∈ Spec(𝑅) respectively. Localization “commutes” with several standard operation
on modules, which we sketch below. The details are left to the reader.
Same for arbitrary direct sums. This is easily seen by viewing 𝑀[𝑆−1 ] as 𝑀 ⊗𝑅
𝑅[𝑆−1 ].
⋄ Note that Hom𝑅 (𝑀, 𝑁) is also an 𝑅-module: simply set (𝑟𝑓 )(𝑚) = 𝑟 ⋅ 𝑓 (𝑚) for any
𝑓 ∈ Hom𝑅 (𝑀, 𝑁). There is a natural homomorphism
𝑅𝑎 → 𝑅𝑏 → 𝑀 → 0, 𝑎, 𝑏 ∈ ℤ≥0 .
In this case (1–2) is an isomorphism, as easily seen from the commutative dia-
gram with exact rows:
Here we used the fact that localization preserves exactness: see the Proposition
1.2.5 below.
have
∼
𝑅′ [(𝑆
⎵⎵ ⎵′⎵⎵
)−1 ] 𝑅′ [𝑆−1 ]
⎵⎵⎵ 𝑅[𝑆−1 ] ⊗ 𝑅′
𝑅
as ring as module
𝑟′ /𝜑(𝑠) (1/𝑠) ⊗ 𝑟′
𝑟′ 𝜑(𝑟)/𝜑(𝑠) (𝑟/𝑠) ⊗ 𝑟′
𝑅[𝑆−1 ]
(𝑅/𝐼)[(𝑆′ )−1 ] ≃ 𝑅[𝑆−1 ] ⊗ (𝑅/𝐼) =
𝑅 𝐼[𝑆−1 ]
We prove an easy yet fundamental property of localizations, namely they are exact
functors.
𝑓𝑖
⋯ → 𝑀𝑖 𝑀𝑖+1 → ⋯
𝑓𝑖 [𝑆−1 ]
⋯ → 𝑀𝑖 [𝑆−1 ] 𝑀𝑖+1 [𝑆−1 ] → ⋯
is also exact.
Proof. By homological common sense, we are reduced to the exact sequences (i) 0 →
𝑀′ → 𝑀 → 𝑀″ (i.e. left exactness), (ii) 𝑀′ → 𝑀 → 𝑀″ → 0 (i.e. right exactness). The
case (ii) is known for tensor products in general.
As for (i), note that for every homomorphism 𝑔 ∶ 𝑀 → 𝑀″ ,
𝑦 ∵𝑦/𝑠=𝑡𝑦/𝑡𝑠 𝑦
ker (𝑔[𝑆−1 ]) = { ∈ 𝑀[𝑆−1 ] ∶ ∃𝑡 ∈ 𝑆, 𝑡𝑔(𝑦) = 0} { ∶ 𝑦 ∈ ker(𝑔)}
𝑠 𝑠
= im [ker(𝑔)[𝑆−1 ] → 𝑀[𝑆−1 ]] .
Lemma 1.2.6. Let 𝑀 be an 𝑅-module. The localizations 𝑀 → 𝑀𝔪 for various maximal ideals
𝔪 assemble into an injection
𝑀↪ ∏ 𝑀𝔪 .
𝔪∈MaxSpec(𝑅)
§1.3 Radicals and Nakayama’s lemma ⋅9⋅
Proof. Let 𝑚 ∈ 𝑀 be such that 𝑚 ↦ 0 ∈ 𝑀𝔪 for all 𝔪. This means that for all 𝔪 there
exists 𝑠 ∈ 𝑅 ∖ 𝔪 such that 𝑠𝑚 = 0. Hence the annihilator ideal ann𝑅 (𝑚) ∶= {𝑟 ∈ 𝑅 ∶
𝑟𝑚 = 0} is not contained in any maximal ideal, thus ann𝑅 (𝑚) = 𝑅.
There is an analogue for rings. Observe that when 𝑅 is an integral domain, all the
localizations 𝑅[𝑆−1 ] can be regarded as subrings of the field of fractions Frac(𝑅).
Lemma 1.2.7. Let 𝑅 be an integral domain, then 𝑅 = ⋂𝔪∈MaxSpec(𝑅) 𝑅𝔪 as subrings of
Frac(𝑅).
Proof. Only the inclusion ⊃ requires proof. Let 𝑥 ∈ Frac(𝑅) and define 𝐷 ∶= {𝑟 ∈ 𝑅 ∶
𝑟𝑥 ∈ 𝑅} (the ideal of denominators). Suppose 𝑥 ∈ 𝑅𝔪 for all maximal 𝔪, then 𝐷 ⊄ 𝔪
for all maximal 𝔪. The same reasoning as above leads to 𝐷 = 𝑅.
It will be important to gain finer control on the ideals 𝔪 in the assertion above, say
by using some prime ideals “lower” than the maximal ones. We will return to this issue
later.
√𝐼 ∶= {𝑟 ∈ 𝑅 ∶ ∃𝑛, 𝑟𝑛 ∈ 𝐼}.
2𝑛
It is readily seen to be an ideal from the binomial identity (𝑎+𝑏)2𝑛 = ∑𝑘=0 (2𝑛 𝑘 2𝑛−𝑘 .
𝑘 )𝑎 𝑏
It should also be clear that √𝐼 ⊂ 𝑅 equals the preimage of √0 ⊂ 𝑅/𝐼.
√𝐼 = ⋂ 𝔭.
𝔭∈Spec(𝑅)
𝔭⊃𝐼
Proof. By replacing 𝑅 by 𝑅/𝐼, this is easily reduced to the case 𝐼 = {0}. If 𝑟 is nilpotent
and 𝔭 ∈ Spec(𝑅), then 𝑟𝑛 = 0 ∈ 𝔭 implies 𝑟 ∈ 𝔭. Conversely, suppose that 𝑟 is not
nilpotent. There exists a prime ideal in 𝑅[𝑟−1 ], which comes from some 𝔭 ∈ Spec(𝑅)
with 𝑟 ∈ 𝑅 ∖ 𝔭 by (1–1). Hence 𝑟 ∉ 𝔭.
Remark 1.3.3. A ring 𝑅 is called reduced if √0 = {0}. In any case, 𝑅red ∶= 𝑅/√0 is a
reduced ring. Furthermore, any reduced quotient of 𝑅 factors through 𝑅red .
The second radical is probably familiar to the readers. The Jacobson radical rad(𝑅)
of the ring 𝑅 is the intersection of all maximal ideals. The previous proposition implies
rad(𝑅) ⊃ √(0). Note that
𝑎 ∈ rad(𝑅) ⟹ (1 + 𝑎) ∈ 𝑅× .
⋅ 10 ⋅ Warming up
Indeed, 1+𝑎 cannot be contained in any maximal ideal 𝔪, for otherwise 1 = (1+𝑎)−𝑎 ∈
𝔪, which is absurd.
To prove the celebrated Nakayama’s Lemma, let us recall an easy variant of the
Cayley–Hamilton theorem from linear algebra.
Lemma 1.3.4. Suppose that 𝐼 ⊂ 𝑅 is an ideal, 𝑀 is an 𝑅-module with generators 𝑥1 , … , 𝑥𝑛 and
𝜑 ∈ End𝑅 (𝑀) satisfies 𝜑(𝑀) ⊂ 𝐼𝑀, then there exists a polynomial 𝑃(𝑋) = 𝑋 𝑛 +𝑎𝑛−1 𝑋 𝑛−1 +
⋯ + 𝑎0 ∈ 𝑅[𝑋] with 𝑎𝑖 ∈ 𝐼, such that 𝑃(𝜑) = 𝜑𝑛 + 𝑎𝑛−1 𝜑𝑛−1 + ⋯ + 𝑎0 = 0.
𝑛
Proof. Write 𝜑(𝑥𝑖 ) = ∑𝑗=1 𝑎𝑖𝑗 𝑥𝑗 where 𝑎𝑖𝑗 ∈ 𝐼. Set 𝐴 ∶= (𝑎𝑖𝑗 )1≤𝑖,𝑗≤𝑛 ∈ Mat𝑛 (𝑅). Regard
𝑀 as an 𝑅[𝑋]-module by letting 𝑋 act as 𝜑. Then we have the matrix equation over
𝑅[𝑋]
⎛
⎜ 𝑥1 ⎞
⎟ ⎛
⎜ 0⎞⎟
⎜
⎜ ⎟
⎟ ⎜
⎜ ⎟
⎟
⎜
(𝑋 ⋅ id𝑛×𝑛 − 𝐴) ⎜ ⋮ ⎟
⎟ = ⎜
⎜ ⋮ ⎟
⎟
⎜
⎜ ⎟ ⎜ ⎟
⎜ ⎟ ⎜ ⎟
⎟ ⎜ ⎟
⎝𝑥𝑛 ⎠ ⎝0⎠
Multiplying by the cofactor matrix (𝑋 ⋅ id𝑀 − 𝐴)∨ on the left, we see that 𝑃(𝑋) ∶=
det(𝑋 ⋅ id𝑛×𝑛 − 𝐴) ∈ 𝑅[𝑋] acts as 0 on each 𝑥𝑖 , thus on the whole 𝑀. This is the
required polynomial.
Theorem 1.3.5 (Nakayama’s Lemma). Suppose that 𝑀 is a finitely generated 𝑅-module and
𝐼 is an ideal of 𝑅 such that 𝐼𝑀 = 𝑀. Then there exists 𝑎 ∈ 𝐼 such that (1 + 𝑎)𝑀 = 0. If
𝐼 ⊂ rad(𝑅), then we have 𝑀 = {0} under these assumptions.
Proof. Write 𝑀 = 𝑅𝑥1 + ⋯ + 𝑅𝑥𝑛 . Plug 𝜑 = id𝑀 into Lemma 1.3.4 to deduce that
𝑛−1 + ⋯ + 𝑎0 acts as 0 on 𝑀. This proves the first part. Assume further-
𝑃(id𝑀 ) = 1 + 𝑎⎵⎵⎵⎵⎵⎵⎵
=∶𝑎∈𝐼
more that 𝐼 ⊂ rad(𝑅), then 1 + 𝑎 ∈ 𝑅× so that 𝑀 = (1 + 𝑎)𝑀 = 0.
Corollary 1.3.6. Let 𝑀 be a finitely generated 𝑅-module, and let 𝐼 ⊂ rad(𝑅) be an ideal of 𝑅.
If the images of 𝑥1 , … , 𝑥𝑛 ∈ 𝑀 in 𝑀/𝐼𝑀 form a set of generators, then 𝑥1 , … , 𝑥𝑛 generate 𝑀.
§1.4 Noetherian and Artinian rings ⋅ 11 ⋅
𝔪1 ⊃ 𝔪1 𝔪2 ⊃ 𝔪1 𝔪2 𝔪3 ⊃ ⋯ .
𝑅 ⊃ 𝔪1 ⊃ 𝔪1 𝔪2 ⊃ ⋯ ⊃ 𝔪1 ⋯ 𝔪𝑛 = 𝔞
⊃ 𝔞𝔪1 ⊃ 𝔞𝔪1 𝔪2 ⊃ ⋯ ⊃ 𝔞𝔪1 ⋯ 𝔪𝑛 = 𝔞2
⊃ ⋯ ⊃ 𝔞𝑘 = {0}.
Corollary 1.4.3. A ring 𝑅 is Artinian if and only if it is Noetherian and every prime ideal of
𝑅 is maximal.
Proof. If 𝑅 is Artinian, then 𝑅 is of finite length, hence is Noetherian as well. For every
prime ideal 𝔭, we have 𝔭 ⊃ {0} = (𝔪1 ⋯ 𝔪𝑛 )𝑘 in the notations of the proof above,
therefore 𝔭 ⊃ 𝔪𝑖 for some 𝑖, so 𝔭 = 𝔪𝑖 is maximal.
Conversely, if 𝑅 is Noetherian but of infinite length, then the nonempty set of ideals
Rings whose prime ideals are all maximal are said to have dimension zero, in the
sense of Krull dimensions; we shall return to this point in §5.4.
One can deduce from this that the (closed) subvarieties of 𝔸𝑛 are in bijection with
ideals 𝔞 satisfying √𝔞 = 𝔞.
Furthermore, 𝕜[𝒳] ∶= 𝕜[𝑋1 , … , 𝑋𝑛 ]/𝐼(𝒳) may be regarded as the ring of “regular
functions” (i.e. functions definable by means of polynomials) on 𝒳, and MaxSpec(𝕜[𝒳])
is in bijection with the points of 𝒳: to 𝑥 = (𝑥1 , … , 𝑥𝑛 ) ∈ 𝒳 we attach
𝔪𝑥 = (𝑋1 − 𝑥1 , … , 𝑋𝑛 − 𝑥𝑛 ) ⊃ 𝐼(𝒳).
which is also an affine algebraic variety in 𝕜𝑛+1 , and one may verify that 𝕜[𝒳𝑓 ] =
𝕜[𝒳][𝑓 −1 ].
(B) Invariant theory. Let 𝐺 be a group acting on a finite-dimensional 𝕜-vector
space 𝑉 from the right, and let 𝕜[𝑉] be the 𝕜-algebra of polynomials on 𝑉. Thus 𝕜[𝑉]
carries a left 𝐺-action by 𝑔𝑓 (𝑣) = 𝑓 (𝑣𝑔). For “reasonable” groups 𝐺, say finite or 𝕜-
algebraic ones, the classical invariant theory seeks to describe the subalgebra 𝕜[𝑉]𝐺 of
invariants1 in terms of generators and relations.
In particular, one has to know when is the algebra 𝕜[𝑉]𝐺 finitely generated. This is
actually the source of many results in commutative algebra, such as the Basis Theorem
and Nullstellensatz of Hilbert. For example, let the symmetric group 𝐺 = 𝔖𝑛 act on
𝑉 = 𝕜𝑛 in the standard manner, then our question is completely answered by the
following classical result: 𝕜[𝑉]𝐺 equals the polynomial algebra 𝕜[𝑒1 , … , 𝑒𝑛 ], where 𝑒𝑖
stands for the 𝑖-th elementary symmetric function in 𝑛 variables.
The same questions may be posed for any affine algebraic variety 𝑉. From the geo-
metric point of view, if 𝕜[𝑉]𝐺 is finitely generated, it will consist of regular functions
of some kind of quotient variety 𝑉//𝐺. The study of quotients in this sense naturally
leads to geometric invariant theory, for which we refer to [16] for details.
1
More generally, we are also interested in the algebra of invariant differential operators with polyno-
mial coefficients.
⋅ 14 ⋅ Warming up
Lecture 2
Primary decompositions
We shall follow [8, §3] and [11, §8] closely.
Supp(𝑀) ∶= {𝔭 ∈ Spec(𝑅) ∶ 𝑀𝔭 ≠ 0} .
Proof. Again, we use the exactness of localization for the first assertion. For 𝔭 ∈ Spec(𝑅)
we have an exact 0 → 𝑀𝔭′ → 𝑀𝔭 → 𝑀𝔭″ → 0, hence 𝑀𝔭 ≠ 0 if and only if 𝔭 ∈
Supp(𝑀′ ) ∪ Supp(𝑀″ ). The second assertion is obvious.
⋅ 16 ⋅ Primary decompositions
From an 𝑅-module 𝑀, one can build a “field of modules” over Spec(𝑅) by assigning
to each 𝔭 the 𝑅𝔭 -module 𝑀𝔭 , and Supp(𝑀) is precisely the subset of 𝑀𝔭 over which the
field is non-vanishing. This is how the support arises in algebraic geometry. A more
precise description will require the notion of quasi-coherent sheaves on schemes.
(𝑀 ⊗ 𝕜) ⊗ (𝕜 ⊗ 𝑁) ≃ 𝑀 ⊗ (𝕜 ⊗ 𝕜) ⊗ 𝑁 ≃ (𝑀 ⊗ 𝑁) ⊗ 𝕜
𝑅 𝕜 𝑅 𝑅 𝕜 𝑅 𝑅 𝑅
is nonzero.
Example 2.2.2. For the ℤ-module 𝑀 ∶= ℤ/𝑛ℤ with 𝑛 ∈ ℤ>1 , one easily checks that
Ass(𝑀) is the set of prime factors of 𝑛.
Lemma 2.2.3. Consider the set 𝒮 ∶= {ann𝑅 (𝑥) ∶ 𝑥 ∈ 𝑀, 𝑥 ≠ 0} of ideals, partially ordered
by inclusion. Every maximal element in 𝒮 is prime.
(ii) The union of all 𝔭 ∈ Ass(𝑀) equals the set of zero divisors on 𝑀.
Proof. (i) Clearly Ass({0}) = ∅. If 𝑀 ≠ 0, the set 𝒮 in Lemma 2.2.3 is then nonempty,
hence contains a maximal element 𝔭 because 𝑅 is Noetherian; this yields 𝔭 ∈ Ass(𝑀).
(ii) Elements of any 𝔭 ∈ Ass(𝑀) are all zero divisors by the very definition of
associated primes. Conversely, if 𝑟 ∈ ann𝑅 (𝑥) for some 𝑥 ∈ 𝑀 ∖ {0}, there must exist
some maximal element 𝔭 of 𝒮 with 𝔭 ⊃ ann𝑅 (𝑥) as 𝑅 is Noetherian; so 𝔭 is the required
associated prime containing 𝑟.
(iii) If 𝔭 ∈ Spec(𝑅), 𝔭 ∩ 𝑆 = ∅ and there is some 𝑅/𝔭 ↪ 𝑀, then
by the exactness of localizations, hence 𝔭[𝑆−1 ] ∈ Ass(𝑀[𝑆−1 ]). Conversely, every ele-
ment of Ass(𝑀[𝑆−1 ]) has the form 𝔭[𝑆−1 ] for some 𝔭 ∈ Spec(𝑅) disjoint from 𝑆. Also
recall that 𝔭 equals the preimage of 𝔭[𝑆−1 ] under 𝑅 → 𝑅[𝑆−1 ]. There exist 𝑥 ∈ 𝑀 and
𝑠 ∈ 𝑆 such that 𝔭[𝑆−1 ] = ann𝑅[𝑆−1 ] (𝑥/𝑠) = ann𝑅[𝑆−1 ] (𝑥). Ideals in a Noetherian ring
being finitely generated, we infer that ∃𝑡 ∈ 𝑆 with 𝔭 ⊂ ann𝑅 (𝑡𝑥). It remains to show
𝔭 = ann𝑅 (𝑡𝑥). If 𝑟𝑡𝑥 = 0 for some 𝑟 ∈ 𝑅, then 𝑟 maps into 𝑟/1 ∈ 𝔭[𝑆−1 ] = ann𝑅[𝑆−1 ] (𝑥);
thus 𝑟 ∈ 𝔭.
(iv) It suffices to treat the second ⊂. Suppose that 𝔭 ∈ Ass(𝑀) and 𝑅/𝔭 ≃ 𝑁 for
some submodule 𝑁 ⊂ 𝑀. Identify 𝑀′ with ker(𝑀 → 𝑀″ ). If 𝑀′ ∩ 𝑁 = {0} then
𝑅/𝔭 ≃ 𝑁 ↪ 𝑀″ , so 𝔭 ∈ Ass(𝑀″ ). If there exists 𝑥 ∈ 𝑀′ ∩ 𝑁 ⊂ 𝑁 with 𝑥 ≠ 0, then we
have ann𝑅 (𝑥) = 𝔭 since 𝑁 ≃ 𝑅/𝔭 and 𝔭 is prime; in this case 𝔭 ∈ Ass(𝑀′ ).
Theorem 2.2.7. For every 𝑅-module 𝑀 we have Supp(𝑀) = ⋃𝔭∈Ass(𝑀) 𝑉(𝔭), in particu-
lar Ass(𝑀) ⊂ Supp(𝑀). Furthermore, every minimal element of Supp(𝑀) with respect to
inclusion is actually a minimal element of Ass(𝑀).
Proof. For any prime 𝔮 we have 𝑀𝔮 ≠ 0 ⟺ Ass(𝑀𝔮 ) ≠ ∅, and the latter condition
holds precisely when there exists 𝔭 ∈ Ass(𝑀) with 𝔭 ∩ (𝑅 ∖ 𝔮) = ∅, i.e. 𝔮 ⊃ 𝔭. This
proves the first assertion. The second assertion is a direct consequence.
(i) 𝑀 is coprimary;
(ii) for every zero divisor 𝑟 ∈ 𝑅 for 𝑀 and every 𝑥 ∈ 𝑀, there exists 𝑛 ≥ 1 such that 𝑟𝑛 𝑥 = 0.
𝔭 ∶= {𝑟 ∈ 𝑅 ∶ ∀𝑥 ∈ 𝑀 ∃𝑛 ≥ 1, 𝑟𝑛 𝑥 = 0}
is an ideal of 𝑅. For every 𝔮 ∈ Ass(𝑀) there exists 𝑥 ∈ 𝑀 with ann𝑅 (𝑥) = 𝔮. Every
𝑟 ∈ 𝔭 has some power falling in 𝔮, thus 𝔭 ⊂ 𝔮. Conversely, (ii) and Theorem 2.2.5 imply
𝔮 = ann𝑅 (𝑥) ⊂ 𝔭. From 𝔮 = 𝔭 we conclude 𝑀 is coprimary with the unique associated
prime 𝔭.
∀𝑎, 𝑏 ∈ 𝑅, (𝑎𝑏 ∈ 𝐼) ∧ (𝑎 ∉ 𝐼) ⟹ ∃𝑛 ≥ 1, 𝑏𝑛 ∈ 𝐼.
In this case we also say 𝐼 is a primary ideal of 𝑅. Show that {√𝐼} = Ass(𝑅/𝐼) if 𝐼 is a
primary ideal. Hint: apply Proposition 2.3.2.
§2.4 Primary decomposition: the main theorem ⋅ 19 ⋅
Exercise 2.3.4. Let 𝔪 be a maximal ideal of 𝑅. Show that every ideal 𝐼 ⊊ 𝑅 containing
some power of 𝔪 is primary, and Ass(𝑅/𝐼) = {𝔪}. Hint: show that 𝔪 is the only prime
ideal containing 𝐼 = ann𝑅 (𝑅/𝐼).
Lemma 2.3.5. Let 𝔭 ∈ Spec(𝑅) and 𝑁1 , 𝑁2 ⊂ 𝑀 are 𝔭-primary submodules. Then 𝑁1 ∩ 𝑁2
is a 𝔭-primary submodule of 𝑀.
Proof. We have 𝑀/𝑁1 ∩ 𝑁2 ↪ 𝑀/𝑁1 ⊕ 𝑀/𝑁2 . Since 𝑁1 ∩ 𝑁2 ≠ 𝑀, we have
by Theorem 2.2.5.
Granting this, 𝑄 ∶= ⋂𝔭∈Ass(𝑀) 𝑄(𝔭) yields the required decomposition since Ass(𝑀) is
finite and Ass(𝑄) = ∅.
Establish (2–2) as follows. Put 𝛹 ∶= {𝔭}. By Zorn’s Lemma we get a maximal
element 𝑄(𝔮) from the set
which is partially ordered by inclusion (details omitted, and you may also use the
Noetherian property of 𝑀). Since
𝑛
together with Theorem 2.2.5 yield Ass(𝑀/𝑁) ⊂ ⋃𝑖=1 Ass(𝑀/𝑀𝑖 ) = {𝔭1 , … , 𝔭𝑛 }.
Now assume the given primary decomposition is irredundant, we have
𝑀2 ∩ ⋯ ∩ 𝑀 𝑛 𝑀2 ∩ ⋯ ∩ 𝑀 𝑛
{0} ≠ =
𝑁 𝑀1 ∩ (𝑀2 ∩ ⋯ ∩ 𝑀𝑛 )
𝑀1 + 𝑀 2 ∩ ⋯ ∩ 𝑀 𝑛
≃ ↪ 𝑀/𝑀1 .
𝑀1
𝑁𝔭 = 𝑀1,𝔭 ⊂ 𝑀𝔭 .
It remains to show that the preimage of 𝑀1,𝔭 under 𝑀 → 𝑀𝔭 equals 𝑀1 , in other words
the injectivity of the natural map 𝑀/𝑀1 → (𝑀/𝑀1 )𝔭 = 𝑀𝔭 /𝑀1,𝔭 . Indeed, 𝑥 ̄ ∈ 𝑀/𝑀1
maps to 0 if and only if there exists 𝑠 ∉ 𝔭 with 𝑠𝑥 ̄ = 0, but Theorem 2.2.5 implies that
the zero divisors of 𝑀/𝑀1 must lie in 𝔭.
It follows that the non-uniqueness of minimal primary decompositions can only
arise from embedded primes in Ass(𝑀/𝑁).
Claim: this gives two minimal primary decompositions of 𝐼. The ideal (𝑋) is prime,
hence primary. In fact, (𝑋 2 , 𝑋𝑌, 𝑌 2 ) = (𝑋, 𝑌)2 and (𝑋 2 , 𝑌) are both primary ideals
associated to the maximal ideal (𝑋, 𝑌). This follows either by direct arguments or by
Exercise 2.3.4, noting that (𝑋, 𝑌)2 = (𝑋 2 , 𝑋𝑌, 𝑌 2 ) is contained in (𝑋 2 , 𝑌). The embed-
ded prime (𝑋, 𝑌) is seen to be responsible non-uniqueness of primary decompositions.
⋅ 22 ⋅ Primary decompositions
To see the geometry behind, recall that 𝑉(𝔞) ∪ 𝑉(𝔟) = 𝑉(𝔞𝔟) = 𝑉(𝔞 ∩ 𝔟) for any
ideals 𝔞, 𝔟, thus expressing 𝐼 as an intersection means breaking the corresponding ge-
ometric object into a union of simpler pieces. Also recall that for an ideal 𝐼 ⊂ 𝕜[𝑋, 𝑌],
the points in ⋂𝑓 ∈𝐼 {𝑓 = 0} are in bijection with the maximal ideals lying over 𝐼, at least
for 𝕜 algebraically closed (Nullstellensatz). Thus we may interpret these primary de-
compositions as equalities among “geometric objects” embedded in 𝕜2 :
⎧
{{𝑋 = 0} ∪ {𝑋 2 = 𝑋𝑌 = 𝑌 2 = 0}
{𝑋 2 = 0, 𝑋𝑌 = 0} = ⎨ 2
{
⎩{𝑋 = 0} ∪ {𝑋 = 𝑌 = 0} .
1. The geometric object defined by 𝑋 = 0 inside 𝕜2 is certainly the 𝑌-axis: the reg-
ular functions living on this space form the 𝕜-algebra 𝕜[𝑋, 𝑌]/(𝑋) = 𝕜[𝑌].
2. The geometric object defined by 𝑋 2 = 𝑋𝑌 = 𝑌 2 = 0 looks “physically” like
the origin (0, 0), but the 𝕜-algebra of “regular functions” (in an extended sense)
living on it equals 𝕜[𝑋, 𝑌]/(𝑋 2 , 𝑋𝑌, 𝑌 2 ): by restricting a polynomial function
𝜕𝑓
𝑓 (𝑋, 𝑌) to this “thickened point”, we see not only 𝑓 (0, 0) but also 𝜕𝑥 (0, 0) and
𝜕𝑓
𝜕𝑦
(0, 0). In other words, we shall view 𝑋 2 = 𝑋𝑌 = 𝑌 2 = 0 as the first-order
infinitesimal neighborhood of (0, 0) ∈ 𝕜2 .
3. In a similar vein, 𝑋 2 = 𝑌 = 0 physically defines (0, 0), but by restricting 𝑓 to that
𝜕𝑓
thickened point, we retrieve 𝑓 (0, 0) as well as 𝜕𝑥 (0, 0). Therefore we obtain the
first-order infinitesimal neighborhood of 0 inside the 𝑋-axis.
Both decomposition says that we obtain the 𝑌-axis together with first-order in-
finitesimal information at the origin (0, 0). This is also a nice illustration of the use
§2.5 Examples and remarks ⋅ 23 ⋅
Example 2.5.3 (Symbolic powers). Let 𝔭 be a prime ideal in a Noetherian ring 𝑅 and
fix 𝑛 ≥ 1. Observe that 𝔭 is the unique minimal element in Supp(𝑅/𝔭𝑛 ) = 𝑉(𝔭𝑛 ).
By Theorem 2.2.7 𝔭 ∈ Ass(𝑅/𝔭𝑛 ), so it makes sense to denote by 𝔭(𝑛) the 𝔭-primary
component (Corollary 2.4.2) of 𝔭𝑛 , called the 𝑛-th symbolic power of 𝔭. In general 𝔭(𝑛) ⊋
𝔭𝑛 . For a nice geometric interpretation of symbolic powers due to Nagata and Zariski,
we refer to [8, §3.9].
⎛
⎜ 𝑥1 ⎞
⎟
⎜
⎜ ⎟
⎟
⎜
(𝑋 ⋅ 1𝑚×𝑚 − 𝐴) ⎜ ⎟
⎜
⎜
⋮ ⎟ = 0,
⎟ 𝐴 = (𝑎𝑖𝑗 )1≤𝑖,𝑗≤𝑚 ∈ Mat𝑚×𝑚 (𝑅)
⎜ ⎟ ⎟
⎝𝑥𝑚 ⎠
over 𝑅[𝑋]. Now multiply both sides by the cofactor matrix (𝑋 ⋅ 1𝑚×𝑚 − 𝐴)∨ , we get
𝑃(𝑋)𝑥𝑖 = 0 for all 𝑖, where 𝑃 ∈ 𝑅[𝑋] is the characteristic polynomial of 𝐴, i.e. 𝑃(𝑥)𝑀 =
{0}. Since 𝑀 is faithful as an 𝑅[𝑥]-module, we get 𝑃(𝑥) = 0.
Corollary 3.1.3. The integral elements in an 𝑅-subalgebra 𝐴 form a subalgebra. In particular,
𝐴 is integral over 𝑅 if and only if it has a set of integral generators.
Proof. Let 𝑎, 𝑏 ∈ 𝐴 be integral elements. One readily checks that
⋄ 𝑅[𝑎, 𝑏] is finitely generated as an 𝑅-module (say by certain monomials 𝑎𝑖 𝑏𝑗 );
3.2 Nullstellensatz
Our aim is to present a generalization of the celebrated Nullstellensatz, which is one of
the cornerstones of algebraic geometry. We shall also write the nilpotent radical of a
ring 𝑅 as
nil(𝑅) ∶= √0𝑅 .
For an ideal 𝔞 ⊂ 𝑅, we write 𝔞[𝑋] for the ideal of 𝑅[𝑋] formed by polynomials with all
coefficients lying in 𝔞.
Definition 3.2.1. A ring 𝑅 is called a Jacobson ring if every prime ideal 𝔭 satisfies
𝔭= ⋂ 𝔪. (3–1)
𝔪∶maximal ideal ⊃𝔭
Equivalently, we require that the Jacobson radical rad(𝑅/𝔭) = nil(𝑅/𝔭) = {0} for all 𝔭.
Note that ⊃ always holds.
Observations:
⋄ Quotients of Jacobson rings are still Jacobson.
⋄ Fields are trivially Jacobson.
Exercise 3.2.2. Prove that every principal ideal domain (a domain in which every ideal
is generated by one element) with infinitely many maximal ideals is Jacobson.
Theorem 3.2.3 (E. Snapper). For any 𝑅, the polynomial algebra 𝑅[𝑋] satisfies rad(𝑅[𝑋]) =
nil(𝑅[𝑋]).
Proof. To show that nil(𝑅[𝑋]) ⊃ rad(𝑅[𝑋]), let 𝑓 (𝑋) = ∑𝑖 𝑎𝑖 𝑋 𝑖 ∈ rad(𝑅[𝑋]), then
1 + 𝑋𝑓 (𝑋) = 1 + ∑𝑖 𝑎𝑖 𝑋 𝑖+1 ∈ 𝑅[𝑋]× . By looking at the reduction modulo 𝔭 of 𝑓 (𝑋)
for every prime ideal 𝔭 of 𝑅, we see that 𝑎𝑖 ∈ ⋂ 𝔭 = nil(𝑅) for all 𝑖. Hence 𝑓 (𝑋) ∈
nil(𝑅)[𝑋] ⊂ nil(𝑅[𝑋]).
Lemma 3.2.4. Let 𝑅 ⊂ 𝐴 be integral domains such that 𝐴 is a finitely generated 𝑅-algebra. If
rad(𝑅) = {0}, then rad(𝐴) = {0}.
Proof. We may assume that 𝐴 is generated by a single element 𝑎 ∈ 𝐴 over 𝑅. If 𝑎 is
transcendental over 𝐾 ∶= Frac(𝑅), Theorem 3.2.3 above can be applied as nil(𝑅[𝑋]) =
𝑛
{0}. Let us assume that 𝑎 satisfies 𝑓 (𝑎) = 0 for some 𝑓 (𝑋) = ∑𝑖=0 𝑟𝑖 𝑋 𝑖 ∈ 𝑅[𝑋] with
𝑟𝑛 ≠ 0. Let 𝑏 ∈ rad(𝐴) and suppose 𝑏 ≠ 0. Hereafter we embed everything into the
𝐾-algebra Frac(𝐴). Since 𝑎 is algebraic over 𝐾, so is every element from 𝑅[𝑎] or even
𝐾[𝑎] ⊂ Frac(𝐴). Hence 𝑏 is integral over 𝐾 as well. By cleaning denominators, we
§3.2 Nullstellensatz ⋅ 29 ⋅
𝑚
arrive at 𝑔(𝑏) = 0 for some 𝑔(𝑋) = ∑𝑖=0 𝑠𝑖 𝑋 𝑖 ∈ 𝑅[𝑋] with the smallest possible degree
𝑚. Since 𝐴 is a domain, we have 𝑠0 ≠ 0.
Using rad(𝑅) = {0}, there exists a maximal ideal 𝔪 of 𝑅 such that 𝑟𝑛 𝑠0 ∉ 𝔪. Taking
localization at 𝔪, we get the subring 𝐴′ ∶= 𝐴 ⊗𝑅 𝑅𝔪 ⊂ Frac(𝐴) containing 𝑅𝔪 , and 𝐴′
is also a finitely generated 𝑅𝔪 -module since we inverted 𝑟𝑛 . Nakayama’s Lemma for
𝑅𝔪 -modules implies 𝔪𝐴′ ⊊ 𝐴′ , thus 𝔪𝐴 ⊊ 𝐴. Now choose a maximal ideal 𝔪𝐴 of 𝐴
over 𝔪. We must have 𝔪𝐴 ∩ 𝑅 = 𝔪, which entails 𝑠0 ∉ 𝔪𝐴 . This is a contradiction since
𝑚
𝑠0 = − ∑𝑖=1 𝑠𝑖 𝑏𝑖 ∈ rad(𝐴).
Theorem 3.2.5 (Nullstellensatz). Let 𝐴 be a finitely generated 𝑅-algebra. Assume that 𝑅 is
a Jacobson ring, then the following statements hold.
(i) 𝐴 is a Jacobson ring.
(ii) Let 𝔫 ∈ Spec(𝐴) be maximal, then its image 𝔪 ∈ Spec(𝑅) is maximal as well, and 𝐴/𝔫
is a finite extension of the field 𝑅/𝔪.
Proof. We start with (i). One may assume 𝑅 ⊂ 𝐴 from the outset. Condition (3–1) for
𝐴 amounts to rad(𝐴/𝔭) = 0 for every 𝔭 ∈ Spec(𝐴). Apply the previous Lemma to the
integral domains 𝑅/𝑅 ∩ 𝔭 ⊂ 𝐴/𝔭 to prove (i).
Now turn to (ii). Using (i) and induction, we may assume that 𝐴 = 𝑅[𝑎] for some
𝑎 ∈ 𝐴. By considering the homomorphism 𝑅/𝔪 ↪ 𝐴/𝔫 between Jacobson rings, we
may further reduce to the case 𝔫 = {0} and 𝔪 = {0}, so that 𝑅 is a domain embedded in
the field 𝐴. In particular 𝑎 cannot be transcendental (as 𝑅[𝑋] is not a field) and must
𝑛
satisfy ∑𝑖=0 𝑐𝑖 𝑎𝑖 = 0 for some 𝑐0 , … , 𝑐𝑛 ∈ 𝑅 with 𝑐𝑛 ≠ 0. Let 𝔨 be any maximal ideal of
𝑅 not containing 𝑐𝑛 , which exists since rad(𝑅) = {0}.
As in the proof of the previous Lemma, 𝑎 becomes integral over 𝑅𝔨 and Nakayama’s
Lemma for 𝑅𝔨 -modules entails 𝔨𝐴 ≠ 𝐴, hence 𝔨 = 0 because 𝐴 is a field. This implies
that 𝑅 is a field and 𝐴 is a finite extension of 𝑅.
Corollary 3.2.6. Let 𝕜 be an algebraically closed field, and 𝐴 ∶= 𝕜[𝑋1 , … , 𝑋𝑛 ]. The maximal
ideals of 𝐴 are in bijection with 𝕜𝑛 by attaching to each 𝑥 ∶= (𝑥1 , … , 𝑥𝑛 ) ∈ 𝕜𝑛 the ideal
𝔪𝑥 = {𝑓 ∈ 𝐴 ∶ 𝑓 (𝑥) = 0} = (𝑋1 − 𝑥1 , … , 𝑋𝑛 − 𝑥𝑛 ).
∼
Proof. Since 𝐴/𝔪𝑥 → 𝕜 by evaluation at 𝑥, we see 𝔪𝑥 is indeed maximal. It is routine
to show that 𝑥 = 𝑦 ⟺ 𝔪𝑥 = 𝔪𝑦 . It remains to show that every maximal ideal
𝔫 contains some 𝔪𝑥 . Indeed, Theorem 3.2.5 implies the field 𝐴/𝔫 is algebraic over 𝕜,
∼
hence 𝐴/𝔫 ≃ 𝕜 as 𝕜-algebras. Let 𝑥𝑖 be the image of 𝑋𝑖 under 𝐴 ↠ 𝐴/𝔫 → 𝕜 and set
𝑥 ∶= (𝑥1 , … , 𝑥𝑛 ), then 𝔫 ⊃ 𝔪𝑥 as required.
Corollary 3.2.7. Keep the notations above and set
Proof. If 𝑓 ∈ 𝐴 and 𝑓 𝑛 ∈ 𝔞 for some 𝑛, the vanishing of 𝑓 𝑛 on 𝑍(𝔞) will entail that of
𝑓 , hence the inclusion ⊃ holds. Assume conversely that 𝑓 ∈ 𝐴 vanishes on 𝑍(𝔞). This
means: for every maximal ideal 𝔪𝑥 we have
𝔪𝑥 ⊃ 𝔞 ⟺ 𝑥 ∈ 𝑍(𝔞) ⟹ 𝑓 (𝑥) = 0 ⟺ 𝑓 ∈ 𝔪𝑥 .
Hence
𝑓 ∈ ⋂ 𝔪𝑥 = √𝔞,
𝔪𝑥 ⊃𝔞
the last equality being based on Definition 3.2.1 since 𝐴/𝔞 is a Jacobson ring. This
concludes the ⊂.
Remark 3.2.8. How about 𝑍𝐼(𝒳)? Unwinding definitions, it is seen to equal the set of
points that “satisfy the algebraic equations that 𝒳 satisfies.” The Zariski topology on
𝕜𝑛 is defined by stipulating the subsets {𝑥 ∈ 𝕜𝑛 ∶ 𝑓 (𝑥) = 0} to be closed, for all 𝑓 ∈ 𝐴,
so we obtain 𝑍𝐼(𝒳) = 𝒳,̄ the Zariski-closure of 𝒳. The reader is invited to verify that by
identifying 𝕜𝑛 with MaxSpec(𝐴), the foregoing topology is induced from the Zariski
topology on the prime spectrum Spec(𝐴).
An (algebraic, closed) subvariety of 𝕜𝑛 is the vanishing locus 𝑓1 = ⋯ = 𝑓𝑚 = 0
for some 𝑓1 , … , 𝑓𝑚 ∈ 𝕜[𝑋1 , … , 𝑋𝑛 ]; it is determined by the ideal 𝔞 = (𝑓1 , … , 𝑓𝑚 ), in fact
it depends only on √𝔞. An ideal 𝔞 is called radical if √𝔞 = 𝔞. To recap, we obtain a
dictionary:
If one allows arbitrary rings 𝐴 instead of just 𝕜[𝑋1 , … , 𝑋𝑛 ]/𝔞, and consider Spec(𝐴)
instead of MaxSpec(𝐴) (the latter is well-behaved only for Jacobson rings), the result
is the category of affine schemes. A proper treatment of these ideas should be left to the
Algebraic Geometry course, if it exists......
left and right exact is equivalent to that 𝐹 preserves all exact sequences; in this case we
say 𝐹 is an exact functor.
Definition 3.3.1. We say 𝑁 is a flat 𝐴-module if 𝑁 ⊗ − is exact. We say 𝑁 is faithfully flat
𝐴
if for every sequence 𝑀• = [⋯ → 𝑀𝑖 → 𝑀𝑖−1 → ⋯] of 𝑅-modules, we have 𝑀• ⊗ 𝑁 is
𝑅
exact if and only if 𝑀• is.
Remark 3.3.2. In view of the right-exactness of ⊗, to assure flatness of 𝑁 it suffices that
𝑁 ⊗ − preserves kernels.
𝐴
Now we consider a ring homomorphism 𝐴 → 𝐵, which makes 𝐵 into an 𝐴-algebra.
Tensor product now gives an additive functor, often called the base change:
𝐵 ⊗ − ∶ 𝐴-Mod → 𝐵-Mod.
𝐴
Thus we can also talk about flatness and faithful flatness of 𝐵 over 𝐴. Since 𝐵 is naturally
an 𝐴-module, this notion is compatible with the previous one.
Example 3.3.3. Let 𝑆 be a multiplicative subset of 𝐴, then 𝐴[𝑆−1 ] is flat over 𝐴. It is not
faithfully flat in general, however; see Theorem 3.5.6.
Example 3.3.4. A routine fact is that for any family (𝑀•(𝑖) ) of complexes of 𝐴-modules,
𝑖∈𝐼
we have
∀𝑖 ∈ 𝐼, 𝑀•(𝑖) is exact ⟺ ⨁ 𝑀•(𝑖) is exact.
𝑖∈𝐼
Recall that ⊗ preserves direct sums. It follows that a direct sum of modules is flat if
and only if each summand is flat. From this we deduce the flatness of free modules
since 𝐴 ⊗ 𝑀 ≃ 𝑀 functorially for each 𝑀. Furthermore, projective modules are flat as
𝐴
they are direct summands of free modules.
Exercise 3.3.5. Show that ℤ/𝑛ℤ is not flat over ℤ for 𝑛 > 1.
We list some basic properties below.
⊳ Tensor products Let 𝑀, 𝑁 be flat (resp. faithfully flat) 𝑅-modules, then so is 𝑀 ⊗
𝑅
𝑁. This follows the associativity constraint of tensor products: (− ⊗ 𝑀) ⊗ 𝑁 ≃
− ⊗ (𝑀 ⊗ 𝑁).
⊳ Transitivity Given ring homomorphisms 𝐴 → 𝐵 → 𝐶, if 𝐵 is flat (resp. faithfully
flat) over 𝐴 and 𝐶 is flat (resp. faithfully flat) over 𝐵, then 𝐶 is also flat (resp.
faithfully flat) over 𝐴. This follows from the transitivity of base change, namely
there is a isomorphism of functors 𝐴-Mod → 𝐶-Mod.
∼
(− ⊗ 𝐵) ⊗ 𝐶 → − ⊗ 𝐶.
𝐴 𝐵 𝐴
⊳ Base change Suppose 𝑁 is a flat (resp. faithfully flat) 𝐴-module and 𝐵 is any 𝐴-
algebra, then 𝑁 ⊗ 𝐵 is a flat (resp. faithfully flat) 𝐵-module. Again, any 𝐵-module
𝐴
𝑀 can be viewed as an 𝐴-module, and there is a functorial isomorphism
∼
(𝑁 ⊗ 𝐵) ⊗ 𝑀 → 𝑁 ⊗ 𝑀.
𝐴 𝐵 𝐴
⋅ 32 ⋅ Integral dependence, Nullstellensatz and flatness
𝑑𝑖
Remark 3.3.6. A sequence [⋯ → 𝑀𝑖 𝑀𝑖−1 → ⋯] of 𝑅-modules is a complex (resp.
exact) if and only if so is its localization at 𝔪, for every maximal ideal 𝔪. Indeed:
⋄ (𝑀• , 𝑑• ) is a complex if and only if im(𝑑𝑖−1 𝑑𝑖 ) = 0 for all 𝑖. Since localization is an
exact functor, it preserves images and we know a module 𝑁 is zero if and only if
𝑁𝔪 = 0 for all 𝔪.
⋄ a complex (𝑀• , 𝑑• ) is exact if and only if H𝑖 (𝑀• ) = 0 for all 𝑖. The same reasoning
applies since localization preserves 𝐻𝑖 .
Proposition 3.3.7. The following are equivalent for an 𝑅-module 𝑁. (i) 𝑁 is flat over 𝑅,
(ii) 𝑁𝔭 is flat over 𝑅𝔭 for all prime ideal 𝔭, (iii) 𝑁𝔪 is flat over 𝑅𝔪 for all maximal ideal 𝔪.
𝑅𝔭 → 𝑅 ′ [𝜑(𝑆)−1 ] → 𝑅′ .
⎵⎵⎵⎵⎵ 𝔭′
as a ring
Proof. (i) ⟹ (ii): Set 𝑆 ∶= 𝑅∖𝔭. Base change implies 𝑅′ [𝑆−1 ] is flat over 𝑅[𝑆−1 ] = 𝑅𝔭 .
Since 𝑅′𝔭′ is a localization of 𝑅[𝑆−1 ] (exercise), we conclude by transitivity.
(ii) ⟹ (iii): Trivial.
(iii) ⟹ (i): By Remark 3.3.6 applied to complexes of 𝑅′ -modules, it suffices to
show the exactness of the functor − ⊗ 𝑅′𝔭′ for all 𝔭′ ∈ MaxSpec(𝑅′ ). In view of Lemma
𝑅
3.3.8, the factorization of 𝑅 → 𝑅′𝔭′ into 𝑅 → 𝑅𝔭 → 𝑅′𝔭′ and the flatness of 𝑅 → 𝑅𝔭 show
that 𝑅′𝔭′ is indeed flat over 𝑅.
Lemma 3.3.10 (Equational criterion of flatness). An 𝑅-module 𝑁 is flat if and only if for
𝑟
all 𝑟 ≥ 1, 𝑎1 , … , 𝑎𝑟 ∈ 𝑅, 𝑥1 , … , 𝑥𝑟 ∈ 𝑁 verifying ∑𝑖=1 𝑎𝑖 𝑥𝑖 = 0, there exist 𝑠 ∈ ℤ≥1 , an
𝑅-valued matrix 𝐵 = (𝑏𝑖𝑗 )1≤𝑖≤𝑟 and 𝑦1 , … , 𝑦𝑠 ∈ 𝑁 such that
1≤𝑗≤𝑠
⎛
⎜ 𝑥1 ⎞⎟ ⎛
⎜ 𝑦1 ⎞⎟
⎜
⎜ ⎟
⎟ ⎜
⎜ ⎟
⎟
⎜ ⎟ = 𝐵 ⎜ ⎟
⎜
⎜
⎜
⋮ ⎟
⎟
⎟
⎜
⎜
⎜
⋮ ⎟,
⎟ (𝑎1 ⋯ 𝑎𝑟 ) 𝐵 = 0.
⎜ ⎟ ⎜ ⎟ ⎟
𝑥
⎝ ⎠ 𝑟 𝑦
⎝ ⎠ 𝑠
§3.4 Structure of flat modules ⋅ 33 ⋅
𝑓
Proof. Suppose 𝑀 is flat and consider the exact sequence 0 → ker(𝑓 ) → 𝑅⊕𝑟 𝑅 where
𝑓 (𝑡1 , … , 𝑡𝑟 ) = ∑𝑖 𝑎𝑖 𝑡𝑖 . We obtain an exact
(𝑥𝑖 )𝑖 ↦∑𝑖 𝑎𝑖 𝑥𝑖
0 → ker(𝑓 ) ⊗ 𝑀 → 𝑀⊕𝑟 𝑀.
𝐴
𝑠
Thus if (𝑥1 , … , 𝑥𝑟 ) ↦ 0 by the arrow above, we can express it as ∑𝑗=1 (𝑏1𝑗 , … , 𝑏𝑟𝑗 ) ⊗ 𝑦𝑗 ∈
ker(𝑓 ) ⊗ 𝑀 as required.
𝐴
To show the converse, we invoke the fact that flatness is equivalent to the injectivity
of 𝔞 ⊗ 𝑁 → 𝔞𝑁 for all finitely generated ideal 𝔞. See Proposition 3.4.1.
It will be useful to rephrase the condition in Lemma 3.3.10 as follows: for all
⋄ homomorphism 𝑥 ∶ 𝑅⊕𝑟 → 𝑀, where 𝑟 ∈ ℤ≥1 , and
⋄ 𝐾 ⊂ ker(𝑥): submodule generated by one element,
there exist some 𝑠 ∈ ℤ≥1 and a commutative diagram
𝐵
𝑅⊕𝑟 𝑅⊕𝑠
s.t. 𝐾 ⊂ ker(𝐵).
𝑥 𝑦
𝑀
(i) 𝑁 is flat,
𝑅
(ii) Tor𝑖 (𝑁, −) = 0 for all 𝑖 > 0,
𝑅
(iii) Tor1 (𝑁, −) = 0,
𝑅
(iv) Tor1 (𝑁, 𝑅/𝔞) = 0 for all finitely generated ideal 𝔞, or equivalently 𝔞 ⊗ 𝑁 → 𝑁 is
𝑅
injective.
⋅ 34 ⋅ Integral dependence, Nullstellensatz and flatness
Proof. First, the equivalence mentioned in (iv) is a consequence of the exact sequence
𝑅 𝑅
Tor1⎵(𝑁,
⎵⎵ ⎵⎵⎵ 𝑅) → Tor1 (𝑁, 𝑅/𝔞) → 𝑁 ⊗ 𝔞 → 𝑁 → 𝑁 ⊗ (𝑅/𝔞) → 0
𝑅 𝑅
=0
Theorem 3.4.6. Let 𝑅 be a local ring with maximal ideal 𝔪. Let 𝑀 be a finitely generated
𝑅-module. The following are equivalent:
⋄ 𝑀 is free,
⋄ 𝑀 is projective.
If we assume moreover that 𝑀 is finitely presented, both conditions are equivalent to the flatness
of 𝑀.
Proof. Free modules are known to be projective. Now let 𝑀 be finitely generated pro-
jective, and take a basis 𝑥1̄ , … 𝑥𝑛̄ of the 𝑅/𝔪-vector space 𝑀/𝔪𝑀, together with liftings
𝑀 ∋ 𝑥𝑖 ↦ 𝑥𝑖̄ . Nakayama’s Lemma then implies the surjectivity of
𝛷 ∶ 𝑅⊕𝑛 ⟶ 𝑀
(𝑎1 , … , 𝑎𝑛 ) ⟼ 𝑎1 𝑥1 + ⋯ + 𝑎𝑛 𝑥𝑛 .
and by comparing dimensions we see 𝑁/𝔪𝑁 = {0}, which in turn gives 𝑁 = {0} by
Nakayama’s Lemma (𝑁 is finitely generated since 𝑅⊕𝑛 ↠ 𝑁). Hence 𝑀 ≃ 𝑅⊕𝑛 is free.
Now turn to the second assertion. Projective modules are flat since they are direct
summands of free modules, and it remains to show that every flat 𝑀 with finite pre-
𝑥
sentation 𝑅⊕𝑞 → 𝑅⊕𝑟 𝑀 → 0 is a direct summand of a free module. Let’s plug
𝑥 ∶ 𝑅⊕𝑟 ↠ 𝑀 and 𝐾 ∶= ker(𝑥) into the equational criterion of flatness (Lemma 3.3.10),
rephrased as in Remark 3.3.11. Let 𝑁 be the image of 𝐵 ∶ 𝑅⊕𝑟 → 𝑅⊕𝑠 . One readily sees
∼
that 𝑦 induces 𝑁 → 𝑀. This furnishes a section 𝑠 ∶ 𝑀 → 𝑅⊕𝑠 for 𝑦.
We deduce the following result characterizing finitely presented projective mod-
ules: in geometric language, they correspond to vector bundles over the affine scheme
Spec(𝑅).
Corollary 3.4.7. Let 𝑀 be a finitely presented 𝑅-module. Then 𝑀 is projective if and only 𝑀𝔪
is free for every maximal ideal 𝔪.
Proof. In view of Theorem 3.4.6, it suffices to show 𝑀 is projective if and only if 𝑀𝔪
is for all maximal ideal 𝔪. One direction is easy: if 𝑀 is a direct summand of a free
module, then so is 𝑀𝔪 .
Conversely, the assumption on finite presentation entails an isomorphism between
additive functors 𝑅-Mod → 𝑅𝔪 -Mod
Hence 𝑀𝔪 is projective for all 𝔪 implies 𝑀 is projective, by Remark 3.3.6 and the exact-
ness of localizations.
⋅ 36 ⋅ Integral dependence, Nullstellensatz and flatness
Note that for finitely presented 𝑀, the equivalence between projectivity and flatness
holds for any ring 𝑅. The arguments are verbatim, and this can also be deduced from
the local case.
We close this section by a stronger result, whose proof is referred to [8, Theorem
A6.6].
Lemma 3.5.1. Let 𝐹 ∶ 𝒞 → 𝒞 ′ be an additive functor between abelian categories. The following
are equivalent.
Proof. (i) ⟹ (ii): If 𝐹𝑀 = 0 then 𝐹(id𝑀 ) = id𝐹𝑀 = 0, and the faithfulness implies
id𝑀 = 0 in End𝒞 (𝑀); this is possible only when 𝑀 = 0.
(ii) ⟹ (i): Suppose 𝑢 ∶ 𝑁 → 𝑀 is mapped to 0 under 𝐹. Then we have 𝐹(im(𝑢)) =
0, thereby im(𝑢) = 0 and 𝑢 = 0.
𝑢 𝑣 𝐹𝑢 𝐹𝑣
(i) ⟹ (iii): Suppose 𝑀′ 𝑀 𝑀″ induces an exact sequence 𝐹𝑀′ 𝐹𝑀
″
𝐹𝑀 . From 𝐹(𝑣𝑢) = 𝐹(𝑣)𝐹(𝑢) = 0 we get 𝑣𝑢 = 0. Thus it makes sense to define
𝐶 ∶= ker(𝑣)/ im(𝑢). One has an exact sequence
im(𝑢) → ker(𝑣) → 𝐶 → 0.
(iii) 𝑁 is flat and for every maximal ideal 𝔪 of 𝑅, we have 𝑁/𝔪𝑁 ≃ 𝑁 ⊗ 𝑅/𝔪 ≠ 0.
𝑅
Proof. The equivalence (i) ⟺ (ii) has just been established. An 𝑅-module 𝑀 is
nonzero if and only if there exist exact sequences
where 𝔞 is a proper ideal and 𝔪 is a maximal over-ideal of 𝔞. Next, let’s show (iii) ⟹
(i) or (ii): there are exact sequences
0 → 𝑇 → 𝑀 ⊗ 𝑁, 𝑇 → 𝑀 ⊗ (𝑅/𝔪) → 0
𝑅 𝑅
𝐵 𝐵𝔭 Spec(𝐵) Spec(𝐵𝔭 )
𝜑 𝜑𝔭 𝜑♯ 𝜑♯𝔭
𝐴 𝐴𝔭 Spec(𝐴) Spec(𝐴𝔭 )
(iii) 𝜑 is flat and for any maximal ideal 𝔭 ⊂ 𝐴 there exists a maximal ideal 𝔪 ⊂ 𝐵 such that
𝜑−1 (𝔪) = 𝔭.
Proof. (i) ⟹ (ii) is contained in Proposition 3.5.5. As for (ii) ⟹ (iii), take any
𝔮 ∈ Spec(𝐵) that pulls back to 𝔭 ∈ Spec(𝐴), then any maximal over-ideal 𝔪 of 𝔮 also
pulls back to 𝔭. To show (iii) ⟹ (i), apply the criterion of Proposition 3.5.2: for any
𝔭 ∈ MaxSpec(𝐴), the existence of 𝔪 ↦ 𝔭 implies 𝔪 ⊃ 𝜑(𝔭) ⋅ 𝐵 = 𝔭𝐵, hence 𝔭𝐵 ≠ 𝐵.
The notion of flatness was first introduced by J.-P. Serre in [15]. The surjections
𝜑♯ ∶ Spec(𝐵) → Spec(𝐴) (or rather their global avatars) for faithfully flat 𝜑 ∶ 𝐴 → 𝐵
are often employed as candidates of “coverings” in algebraic geometry, leading up to
the well-known fpqc (faithfully flat + quasi-compact) and fppf (faithfully flat + finitely
presented) topologies in the sense of Grothendieck. They have been indispensable
tools for contemporary geometers; [17] serves as a readable introduction to this circle
of ideas.
Lecture 4
Going-up, going-down,
gradings and filtrations
This lecture will be less self-contained than the other ones.
⋄ We say the going-up property holds for 𝜑 if for every 𝔭 ⊂ 𝔭′ in Spec(𝐴) and 𝔮 ∈
Spec(𝐵) with 𝜑♯ (𝔮) = 𝔭 (and we say 𝔮 lies over 𝔭...) there exists 𝔮′ ⊃ 𝔮 lying over
𝔭′ .
⋄ We say the going-down property holds for 𝜑 if for every 𝔭 ⊂ 𝔭′ in Spec(𝐴) and
𝔮′ ∈ Spec(𝐵) lying over 𝔭′ , there exists 𝔮 ⊂ 𝔮′ lying over 𝔭.
Pictorially:
𝔮′ 𝐵
𝔭′ 𝐴
Proof. One easily reduces to the case 𝔞 = {0}. We want to use Zorn’s Lemma to find a
minimal prime. It boils down to show that any chain (𝔭𝑖 )𝑖∈𝐼 of prime ideals (𝐼: totally
ordered set with 𝑗 > 𝑖 ⟹ 𝔭𝑖 ⊃ 𝔭𝑗 ) has a lower bound; we assume 𝔭𝑖 ⊂ 𝔓 when 𝔓
is prescribed. It suffices to show 𝔭 ∶= ⋂𝑖∈𝐼 𝔭𝑖 is prime: if 𝑥𝑦 ∈ 𝔭 but there exists 𝑖 with
𝑥 ∉ 𝔭𝑖 , then 𝑥 ∉ 𝔭𝑗 whenever 𝑗 ≥ 𝑖; in this case 𝑗 ≥ 𝑖 ⟹ 𝑦 ∈ 𝔭𝑗 . This entails 𝑦 ∈ 𝔭.
Lemma 4.1.2. The going-down property for 𝜑 is equivalent to the following: for every 𝔭 ∈
Spec(𝐴) with 𝜑(𝔭)𝐵 ≠ 𝐵 and any minimal over-prime 𝔮 of 𝜑(𝔭)𝐵, we have 𝜑♯ (𝔮) = 𝔭.
Proof. Assuming going-down for 𝜑, let 𝔭, 𝔮 be as above. Evidently 𝜑♯ (𝔮) ⊃ 𝜑−1 (𝜑(𝔭)𝐵) ⊃
𝔭. If we have ⊋, then going-down guarantees the existence of 𝔮♭ ⊊ 𝔮 lying over 𝔭. Thus
𝜑−1 (𝔮♭ ) = 𝔭 implies 𝔮♭ ⊃ 𝜑(𝔭)𝐵, contradicting the minimality of 𝔮.
To show the converse, consider 𝔭 ⊂ 𝔭′ with 𝔮′ lying over 𝔭′ . We have 𝜑(𝔭)𝐵 ⊂
𝜑(𝔭′ )𝐵 ⊂ 𝔮′ ≠ 𝐵. Take 𝔮 to be a minimal over-prime of 𝜑(𝔭)𝐵 (which exists by Lemma
4.1.1) to verify the going-down property.
Proof. Consider 𝔭 ⊂ 𝔭′ and 𝔮′ lying over 𝔭′ in the setting of going-down. First, 𝐵𝔮′ is
flat over 𝐴𝔭′ by Proposition 3.3.9. Secondly, 𝐴𝔭′ → 𝐵𝔮′ is faithfully flat since it is local
by Theorem 3.5.6 (iii), therefore induces a surjection on spectra. Take any prime of 𝐵𝔮′
mapping to 𝔭𝐴𝔭 ∈ Spec(𝐴𝔭′ ) and to 𝔮 ∈ Spec(𝐵). In view of the commutative diagrams
Theorem 4.1.4 (Krull–Cohen–Seidenberg). Suppose the ring 𝐵 is integral over its subring
𝐴. The following holds.
(v) Assume 𝐴, 𝐵 are domains and 𝐴 is normal. Then going-down holds for 𝐴 ↪ 𝐵.
(vi) Assume furthermore that 𝐵 is the integral closure of 𝐴 in a normal field extension 𝐿 ⊃
𝐾 ∶= Frac(𝐴), then 𝛤 ∶= Aut(𝐿/𝐾) acts transitively on each fiber of Spec(𝐵) →
Spec(𝐴).
The last assertion should be familiar to readers with a background in algebraic num-
ber theory.
§4.1 Going-up and going-down ⋅ 41 ⋅
Proof. (iv): Let 𝔮 ∈ MaxSpec(𝐵) and 𝔭0 ∶= 𝔮 ∩ 𝐴. We know 𝐵/𝔮 is a field, integral over
its subring 𝐴/𝔭0 . We claim that 𝐴/𝔭0 is also a field, therefore 𝔭0 = 𝔭. Let 𝑥 ∈ 𝐴/𝔭0 ∖{0}.
Its inverse in 𝐵/𝔮 satisfies an integral relation ( 1𝑥 )𝑛 +𝑎𝑛−1 ( 1𝑥 )𝑛−1 +⋯ +𝑎0 = 0 over 𝐴/𝔭0 .
Multiplying both sides by 𝑥𝑛−1 yields 1𝑥 ∈ (𝐴/𝔭0 )[𝑥].
Conversely, we have to show any 𝔮 ∈ Spec(𝐵) with 𝔮 ∩ 𝐴 = 𝔭 is maximal. Again,
there is an integral extension of domains 𝐴/𝔭 ↪ 𝐵/𝔮. Consider 𝑦 ∈ 𝐵/𝔮 satisfying
𝑦𝑛 + 𝑎𝑛−1 𝑦𝑛−1 + ⋯ + 𝑎0 = 0, with the smallest possible 𝑛. If 𝑦 ≠ 0 then 𝑎0 ≠ 0. Since
𝐴/𝔭 is a field, the recipe to produce 𝑦−1 ∈ (𝐴/𝔭)[𝑦] is well-known.
(i), (ii): Fix 𝔭 and consider the inclusion 𝐴𝔭 ↪ 𝐵𝔭 = 𝐵 ⊗ 𝐴𝔭 which is still integral
𝐴
(note that 𝐴 ∖ 𝔭 is a multiplicative subset of 𝐵, and 𝐵𝔭 is nonzero). We are reduced to
the case 𝐴 is local with maximal ideal 𝔭. By (iv) the fiber of 𝔭 in Spec(𝐵) is nothing but
MaxSpec(𝐵). This establishes (i) and (ii) since there are no inclusions among maximal
ideals.
(iii): Consider 𝔭 ⊂ 𝔭′ and 𝔮 over 𝔭 in the setting of going-up. Then (i) is applicable
to 𝐴/𝔭 ↪ 𝐵/𝔮 and yields the required 𝔮′ ∈ Spec(𝐵/𝔮) ↪ Spec(𝐵).
(vi): Observe that every 𝜎 ∈ 𝛤 induces an 𝐴-automorphism of 𝐵. Let 𝐾 ′ ∶= 𝐿𝛤 . By
(infinite) Galois theory we know 𝐿/𝐾 ′ is Galois and 𝐾 ′ /𝐾 is purely inseparable. Let 𝐴′
be the integral closure of 𝐴 in 𝐾 ′ . First observe that
Spec(𝐴′ ) ⟶ Spec(𝐴), 𝔭 ′ ↦ 𝔭 = 𝔭′ ∩ 𝐴
is a bijection. Indeed, 𝐾 ′ ≠ 𝐾 only when 𝑝 ∶= char(𝐾) > 0, in which case the inverse
𝑚
is given by 𝔭′ = {𝑡 ∈ 𝐴′ ∶ 𝑡𝑝 ∈ 𝔭, 𝑚 ≫ 0}. Thus we assume henceforth that 𝐿/𝐾 is
Galois.
Deal with the case [𝐿 ∶ 𝐾] < ∞ first. Consider 𝔮, 𝔮′ ∈ Spec(𝐵) in the fiber over 𝔭.
Suppose on the contrary that 𝛤𝔮 does not meet 𝔮′ , then by (ii) we have 𝔮′ ⊄ 𝜎(𝔮) for
each 𝜎 ∈ 𝛤. By the prime avoidance (Proposition 1.1.5), there exists 𝑥 ∈ 𝔮′ ∖ ⋃𝜎 𝜎(𝔮),
since 𝛤 is finite. Now define the norm 𝑦 ∶= 𝑁𝐿/𝐾 (𝑥) ∈ 𝐾, which is some positive power
(namely [𝐿 ∶ 𝐾]𝑖 ) of ∏𝜎∈𝛤 𝜎(𝑥), hence belongs to 𝐵. Notice that
⋄ 𝐴 normal implies 𝑦 ∈ 𝐴;
⋄ 𝑥 ∉ 𝜎 −1 (𝔮) for all 𝜎 ∈ 𝛤 implies 𝑦 ∉ 𝔭;
⋄ however 𝑦 ∈ 𝔮′ ∩ 𝐴 = 𝔭 since 𝑥 ∈ 𝔮′ . Contradiction.
Now suppose [𝐿 ∶ 𝐾] is infinite and 𝔮, 𝔮′ in the fiber over 𝔭. We need to use the Krull
topology on 𝛤. For every finite, Galois subextension 𝐸/𝐾 of 𝐿/𝐾, define the set
𝒯 (𝐸) ∶= {𝜎 ∈ 𝛤 ∶ 𝜎(𝔮 ∩ 𝐸) = 𝔮′ ∩ 𝐸} .
𝔯 𝔯1 𝔯2 𝐶
𝔮′ 𝐵
𝔭 𝔭′ 𝐴
we first construct 𝔯 and then 𝔯1 by going-up, then “tilt” it via some 𝜎 to match 𝔯1 with
some chosen 𝔯2 above 𝔮′ , so that 𝜎(𝔯) ∩ 𝐵 produces the required going-down:
Proof. Consider a closed subset 𝑉(𝔟) of Spec(𝐵). First, every 𝔮 ∈ Spec(𝐵) with 𝔮 ⊃ 𝔟
lies over a minimal over-prime of 𝔟, by Lemma 4.1.1. Secondly, 𝐵 is Noetherian im-
plies Ass(𝐵/𝔟) is finite; in particular there are only finitely many minimal over-primes
𝔮1 , … , 𝔮𝑛 of 𝔟.
Set 𝔭𝑖 ∶= 𝜑♯ (𝔮𝑖 ) for all 𝑖. By going-up, 𝑉(𝔭𝑖 ) is contained in 𝜑♯ (𝑉(𝔟)). On the other
hand, every 𝔭 = 𝜑♯ (𝔮) with 𝔮 ∈ 𝑉(𝔟) lies over some 𝔭𝑖 = 𝜑♯ (𝔮𝑖 ) by the foregoing
𝑛
discussion. This shows 𝜑♯ (𝑉(𝔟)) = ⋃𝑖=1 𝑉(𝔭𝑖 ) is closed.
Corollary 4.2.2. Suppose 𝐵 is Noetherian and integral over a subring 𝐴. Then Spec(𝐵) →
Spec(𝐴) is a closed surjection with finite fibers.
Proof. Apply Theorem 4.1.4 with Proposition 4.2.1 to show that Spec(𝐵) → Spec(𝐴) is
closed and surjective.
To show the finiteness of the fiber over 𝔭 ∈ Spec(𝐴), note that the preimage of
𝑉(𝔭) in Spec(𝐵) equals 𝑉(𝔭𝐵). Since there are no inclusions in the fiber over 𝔭, every
element in that fiber must be a minimal over-prime of 𝔭𝐵. We have seen in the proof of
Proposition 4.2.1 that there are only finitely many such minimal-over primes.
Note that the “closed surjection” part applies to any integral extension of rings. See
Exercise 4.1.5.
In order to obtain further results of this type, we have to introduce more notions.
Let 𝑅 be a ring.
§4.2 Subsets in the spectrum ⋅ 43 ⋅
To make geometric meaning from it, being “specialized” signifies that there are
“more equations” in 𝔭′ , therefore it corresponds a smaller embedded geometric object.
For example, in 𝑅 = ℂ[𝑋, 𝑌] the prime ideal (𝑋, 𝑌) is a specialization of (𝑋), as the
origin 𝑋 = 𝑌 = 0 belongs to the line 𝑋 = 0.
Lemma 4.2.4. With respect to the Zariski topology, 𝔭 is a generalization of 𝔭′ if and only if
𝔭′ ∈ {𝔭}.
Proof. The condition 𝔭′ ∈ {𝔭} means that for every ideal 𝔞, if 𝔭 ⊃ 𝔞 then 𝔭′ ⊃ 𝔞. Taking
𝔞 = 𝔭 yields 𝔭′ ⊃ 𝔭, and the converse is even easier.
A subset is called stable under specialization (resp. generalization) if the special-
ization (resp. generalization) of any member still belongs to that set. The following is
straightforward.
Lemma 4.2.5. Any closed subset is stable under specialization; any open subset is stable under
generalization.
⋄ The foregoing definition is standard only for 𝑅 Noetherian. The general definition
in EGA differs.
⋄ These notions can be applied to any topological space 𝑋. In practice one usually
suppose 𝑋 to be
– Noetherian: the closed subsets satisfy descending chain condition,
– sober: every irreducible has a generic point,
in order to get interesting results. This explains our Noetherian assumption.
(¬𝑋 = 0) ∨ (𝑋 = 0 ∧ 𝑌 = 0)
Exercise 4.2.7. Show that the set of constructible subsets is stable under finite ∪, finite
∩ and taking complements.
Exercise 4.2.8. Show that irreducible closed subsets 𝑍 admit the following topological
characterization: if 𝑋 = 𝐴 ∪ 𝐵 with 𝐴, 𝐵 closed, then either 𝑋 = 𝐴 or 𝑋 = 𝐵.
Lemma 4.2.9. Let 𝐸 be a subset of Spec(𝑅) where 𝑅 is a Noetherian ring. The following are
equivalent:
(i) 𝐸 is constructible;
(ii) for every irreducible subset 𝑍 of Spec(𝑅), either 𝑍 ∩ 𝐸 is not dense in 𝑍 or 𝑍 ∩ 𝐸 contains
a nonempty open subset of 𝑍.
Now we can give a partial converse to Lemma 4.2.5, albeit not in the strongest form.
Proof. Let 𝑈 = Spec(𝐵) ∖ 𝑉(𝔞) be an open subset. Going-down implies that 𝜑♯ (𝑈) is
stable under generalization. It suffices to show 𝜑♯ (𝑈) is constructible, and this is the
content of Chevalley’s Theorem 4.2.10.
§4.3 Graded rings and modules ⋅ 45 ⋅
Definition 4.3.1. A 𝛤-graded ring is a ring 𝑅 whose underlying additive group is en-
dowed with a decomposition 𝑅 = ⨁𝛾∈𝛤 𝑅𝛾 , such that 𝑅𝛾 𝑅𝜂 ⊂ 𝑅𝛾+𝜂 for all 𝛾, 𝜂 ∈ 𝛤.
For 𝑅 as above, a 𝛤-graded 𝑅-module is an 𝑅-module 𝑀 whose underlying additive
group decomposes as 𝑀 = ⨁𝛾∈𝛤 𝑀𝛾 , such that 𝑅𝛾 𝑀𝜂 ⊂ 𝑀𝛾+𝜂 for all 𝛾, 𝜂 ∈ 𝛤; in
particular, 𝑅 itself is a 𝛤-graded 𝑅-module. If 𝑥 ∈ 𝑀𝛾 ∖ {0}, we say 𝑥 is homogeneous
of degree 𝛾.
We will often omit 𝛤 when there is no worry of confusion. Note that if 0 is allowed
to be homogeneous, as people sometimes do, it will be homogeneous of any degree.
Exercise 4.3.2. Show that in a graded ring 𝑅 we always have 1 ∈ 𝑅0 , provides that (𝛤, +)
satisfies the cancellation law: 𝛾 + 𝜂 = 𝜂 ⟺ 𝛾 = 0. Hint: let 𝑒0 be the component of 1𝑅
in degree 0, argue that 𝑥𝑒0 = 𝑥 = 𝑒0 𝑥 for all homogeneous 𝑥 ∈ 𝑅. The condition 1 ∈ 𝑅0
is sometimes built into the definition of graded rings.
(i) 𝑁 ⊂ 𝑀 is graded;
Example 4.3.4. Let 𝐴 be a ring and 𝑅 ∶= 𝐴[𝑋1 , … , 𝑋𝑛 ]. Then 𝑅 is naturally ℤ≥0 -graded
by degrees: for each 𝑑 ∈ ℤ≥0 , let 𝑅𝑑 be the set of homogeneous polynomials of total de-
gree 𝑑. Ideals generated by homogeneous polynomials are precisely the graded ideals.
The importance of this grading comes from projective algebraic geometry.
On the other hand, 𝑅 can also be graded by monomials by taking 𝛤 = ℤ≥0 ×⋯×ℤ≥0
𝑑 𝑑
(𝑛 copies), and we set 𝑅(𝑑1 ,…,𝑑𝑛 ) = 𝐴 ⋅ 𝑋1 1 ⋯ 𝑋𝑛𝑛
Lemma 4.3.5. Let 𝑀 be a ℤ-graded module over a ℤ-graded ring 𝑅. If 𝑥 ∈ 𝑀 and 𝔭 ∶= ann(𝑥)
is a prime ideal, then
Proof. We begin with (i). Let 𝑡 ∈ 𝔭 and 𝑥 ∈ 𝑀 be such that 𝔭 = ann(𝑀). Write
𝑡 = ∑ 𝑡𝛾 , 𝑥 = ∑ 𝑥𝜂 ,
𝛾∈𝒜 𝜂∈ℬ
where 𝑡𝛾 ∈ 𝑅𝛾 ∖ {0} and 𝑥𝜂 ∈ 𝑀𝜂 ∖ {0} for all 𝛾, 𝜂. Denote by 𝛾0 and 𝜂0 the minimal
elements in 𝒜 and ℬ, respectively. Homogenity amounts to 𝑡𝛾 ∈ 𝔭 for each 𝛾 ∈ 𝒜,
and this will be done by induction on |ℬ|. By a further induction on |𝒜|, for fixed 𝑥 and
𝔭, this can be reduced to showing 𝑡𝛾0 ∈ 𝔭.
First of all, considerations of degrees and 𝑡𝑥 = 0 lead to 𝑡𝛾0 𝑥𝜂0 = 0. If 𝑥 = 𝑥𝜂0 (i.e.
|ℬ| = 1), we obtain 𝑡𝛾0 ∈ ann(𝑥) = 𝔭 as required. In general:
𝑡𝛾0 𝑥 = ∑ 𝑡𝛾0 𝑥𝜂
𝜂∈ℬ
𝜂≠𝜂0
⋄ Suppose there exists 𝑠 ∈ 𝑅 ∖ 𝔭 such that 𝑠(𝑡𝛾0 𝑥) = 0, then 𝑠𝑡𝛾0 ∈ 𝔭, hence 𝑡𝛾0 ∈ 𝔭.
This concludes the homogeneity (i).
From the homogeneity 𝔭 we infer that 𝔭 ⊂ ann(𝑥𝜂 ) for each 𝜂 ∈ ℬ. Now that
𝔭 = ann(𝑥) ⊃ ∏ ann(𝑥𝜂 ),
𝜂∈ℬ
4.4 Filtrations
Now turn to filtrations. We only deal with decreasing filtrations indexed by ℤ≥0 .
𝑅 = 𝐹0 𝑅 ⊃ 𝐹1 𝑅 ⊃ 𝐹2 𝑅 ⊃ ⋯
of ideals such that 𝐹𝑖 𝑅 ⋅ 𝐹𝑗 𝑅 ⊂ 𝐹𝑖+𝑗 𝑅. Define the associated ℤ≥0 -graded ring
Proposition 4.4.5. Let 𝔞 be a proper ideal of 𝑅 and 𝑀 a finitely generated 𝑅-module. Suppose
𝑀 is endowed with an 𝔞-stable filtration such that 𝐹𝑖 𝑀 is finitely generated for each 𝑖, and
𝐹≤0 𝑀 = 𝑀. Then gr(𝑀) is a finitely generated gr(𝑅)-module.
Proof. Take 𝑛 such that 𝔞 ⋅ 𝐹𝑖 𝑀 = 𝐹𝑖+1 𝑀 for all 𝑖 ≥ 𝑛. Then in gr(𝑀) = ⨁𝑖 gr𝑖 𝑀 we
have
(𝔞/𝔞2 ) ⋅ gr𝑖 𝑀 = gr𝑖+1 𝑀, 𝑖 ≥ 𝑛.
Therefore it suffices to take generators from gr0 𝑀, … , gr𝑛 𝑀, each of whom is finitely
generated over 𝑅/𝔞 = gr0 𝑅.
1
In view of later applications, the filtration on a module is indexed by ℤ instead of ℤ≥0 .
⋅ 48 ⋅ Going-up, going-down, gradings and filtrations
Definition 4.5.1 (Morphisms between filtered objects). Let (𝐴, 𝐹• 𝐴) and (𝐵, 𝐹• 𝐵) be
filtered rings. A morphism between them means a ring homomorphism 𝜑 ∶ 𝐴 → 𝐵
satisfying 𝜑(𝐹𝑖 𝐴) ⊂ 𝐹𝑖 𝐵 for all 𝑖. Similarly, suppose 𝐴 is filtered and let 𝑀, 𝑁 be filtered
𝐴-modules. A morphism 𝑀 → 𝑁 means a homomorphism 𝜓 ∶ 𝑀 → 𝑁 of 𝑅-modules
satisfying 𝜓(𝐹𝑖 𝑀) ⊂ 𝐹𝑖 𝑁 for all 𝑖.
This makes the filtered rings and the filtered modules over a filtered ring into cat-
egories. Obviously, morphisms 𝜑 between filtered objects induce graded morphisms
gr 𝜑 between the associated graded objects. Therefore we obtain a functor from the
category of filtered rings or modules into their graded avatars.
Remark 4.5.2. Suppose 𝜑 ∶ 𝑀 → 𝑁 is a morphism between filtered 𝐴-modules. The
quotient 𝑀/ ker(𝜑) inherits a filtration from 𝑀, whereas the submodule im(𝜑) inherits
one from 𝑁. When the natural isomorphism 𝑀/ ker(𝜑) → im(𝜑) is an isomorphism
between filtered modules, or equivalently
∀𝑖 ∈ ℤ, 𝜑(𝐹𝑖 𝑀) = 𝜑(𝑀) ∩ 𝐹𝑖 𝑁,
Bl𝔞 𝑅 ∶= ⨁ 𝔞𝑛 𝑋 𝑛 ⊂ 𝑅[𝑋].
𝑛≥0
̃ −1 ] = ⨁ 𝑅 ⋅ 𝑇 −𝑛 = 𝑅[𝑇 ±1 ].
𝐵[𝑇
𝑛∈ℤ
Lemma 4.5.6. Consider a ring 𝑅 with proper ideal 𝔞, together with a filtered 𝑅-module 𝑀,
assume furthermore that each 𝐹𝑖 𝑀 is finitely generated over 𝑅. The following are equivalent:
Theorem 4.5.7 (Artin–Rees). Let 𝑅 be a Noetherian ring endowed with 𝔞-adic filtration. Let
𝑀 be a finitely generated 𝑅-module and 𝑁 ⊂ 𝑀 an 𝑅-submodule. Then the filtration on 𝑁
induced by the 𝔞-adic filtration of 𝑀, namely 𝐹𝑖 𝑁 ∶= 𝔞𝑖 𝑀 ∩ 𝑁, is 𝔞-stable.
Proof. Since the induced filtration on 𝑁 is 𝔞-stable by Theorem 4.5.7, for 𝑛 ≫ 0 we have
𝑁 = 𝔞𝑛 𝑀 ∩ 𝑁 = 𝔞 ⋅ (𝔞𝑛−1 𝑀 ∩ 𝑁) = 𝔞𝑁.
Corollary 4.5.9 (Krull). If 𝔞 ⊂ rad(𝑅), then ⋂𝑛≥0 𝔞𝑛 𝑀 = {0} for any finitely generated
𝑅-module 𝑀. In particular ⋂𝑛≥0 𝔞𝑛 = {0} whenever 𝔞 ⊂ rad(𝑅).
From completions to
dimensions
The main references are [11, 8].
5.1 Completions
Consider a ring 𝑅 together with a family of ideals ℐ ≠ ∅, such that for any 𝐼, 𝐽 ∈ ℐ
there exists 𝐾 ∈ ℐ with 𝐾 ⊂ 𝐼 ∩ 𝐽. This turns 𝑅 into a topological ring, characterized by
the property that ℐ forms a local base of open neighborhoods of 0. Recall that being a
topological ring means that addition, multiplication and 𝑥 ↦ −𝑥 are all continuous. By
standard arguments, 𝑅 is Hausdorff if and only if {0} is closed, if and only if ⋂𝐼∈ℐ 𝐼 =
{0}.
To simplify matters, we assume that
⋄ the family ℐ is countable, so that the topological properties (accumulation points,
etc.) are detected by convergence of sequences as in the case of metric spaces;
⋄ furthermore, we may arrange that ℐ = {𝐼 ⊃ 𝐽 ⊃ 𝐾 ⊃ ⋯}, in other words our
topology comes from filtrations.
Without the countability assumption, the sequences will have to be replaced by filters.
It makes sense to talk about topological 𝑅-modules for a topological ring 𝑅. By re-
placing filtration by ideals by filtration by 𝑅-submodules subject to the usual compati-
bility relation 𝐹𝑖 𝑅 ⋅ 𝐹𝑗 𝑀 ⊂ 𝐹𝑖+𝑗 𝑀, the recipe above applies to 𝑅-modules as well. Given
𝑁 ⊂ 𝑀, the topology so obtained on 𝑀 passes to 𝑀/𝑁 by taking the quotient topol-
ogy, or equivalently the quotient filtration (𝐹• 𝑀 + 𝑁)/𝑁. If the filtration in question is
𝐼-adic, where 𝐼 ⊊ 𝑅 is an ideal, we obtain the 𝐼-adic topology on rings and modules.
An 𝑅-module 𝑀 equipped with a topology as above is complete if every Cauchy
sequence (𝑥𝑛 )𝑛≥1 has a limit; a Cauchy sequence (𝑥𝑛 )𝑛≥1 is a sequence satisfying
∀𝐼 ∈ ℐ, ∃𝑁 𝑖, 𝑗 ≥ 𝑁 ⟹ 𝑥𝑖 − 𝑥𝑗 ∈ 𝐼.
As in the familiar case of metric spaces, one has the completion of 𝑀. It is actually a mor-
phism 𝑀 → 𝑀̂ with 𝑀̂ complete Hausdorff, characterized by the following universal
⋅ 52 ⋅ From completions to dimensions
property:
𝜑∶ cont. homo.
𝑀 𝑀̂
𝑀 𝐿
⇝ ∃!𝜑̂
complete Hausdorff 𝜑
𝐿
The uniqueness results immediately, and the formation of 𝑀 ↦ 𝑀̂ is seen to be
functorial in 𝑀. If 𝑀 is already complete Hausdorff, one may take 𝑀̂ = 𝑀. Certainly,
the same applies to the ring 𝑅.
Exercise 5.1.1. Suppose that 𝑅 is complete Hausdorff with respect to the 𝐼-adic topol-
ogy, where 𝐼 is a proper ideal. Show that every element of the form 𝑢 + 𝑥, 𝑢 ∈ 𝑅× and
𝑥 ∈ 𝐼, is invertible.
From the algebraic perspective, the completion of a filtered 𝑅-module 𝑀 = 𝐹0 𝑀 ⊃
𝐹1 𝑀 ⊃ ⋯ can be constructed as the projective limit
𝑀̂ ∶= lim 𝑀/𝐹𝑖 𝑀
←−−
𝑖≥1
The morphism 𝑀 → 𝑀̂ is the diagonal map. The topology on 𝑀̂ arises from the filtra-
tion
𝐹𝑖 𝑀̂ ∶= ker [𝑝𝑖 ∶ 𝑀̂ → 𝑀/𝐹𝑖 𝑀] = {(𝑥𝑛 )𝑛 ∈ 𝑀̂ ∶ 𝑖 ≤ 𝑘 ⟹ 𝑥𝑖 = 0} ,
so that the preimage of 𝐹𝑖 𝑀̂ in 𝑀 is precisely 𝐹𝑖 𝑀. In the case where 𝑀 = 𝑅 and 𝐹𝑖 𝑅
are ideals, we obtain the complete Hausdorff ring 𝑅,̂ which is a subring of ∏𝑖≥1 𝑅/𝐹𝑖 𝑅.
Since the filtrations on 𝑅 and 𝑀 are assumed compatible, 𝑀̂ is an 𝑅-module.
̂
Example 5.1.2. Fix a prime number 𝑝. The completion of ℤ with respect to the ideal
𝑝ℤ is nothing but the ring ℤ𝑝 of 𝑝-adic integers. Similarly, the completion of 𝕜[𝑋] with
respect to (𝑋) is isomorphic to the 𝕜-algebra 𝕜J𝑋K.
Exercise 5.1.3. Describe the kernel of 𝑀 → 𝑀̂ and show 𝑀 ↪ 𝑀̂ if and only if 𝑀 is
Hausdorff.
Exercise 5.1.4. Show that the topology of 𝑀̂ is the restriction of the product topology of
∏𝑖 𝑀/𝐹𝑖 𝑀, provided that each 𝑀/𝐹𝑖 𝑀 is endowed with the discrete topology. Show
that 𝑀̂ is a closed subspace of ∏𝑖 𝑀/𝐹𝑖 𝑀
Lemma 5.1.5. Let 𝑀 be a complete 𝑅-module with respect to some filtration 𝐹• 𝑀. For any
submodule 𝑁, the quotient 𝑀/𝑁 is also complete with respect to the quotient topology, or equiv-
alently with respect to the quotient filtration (𝐹• 𝑀 + 𝑁)/𝑁.
Proof. Let 𝑥𝑛̄ be a Cauchy sequence in 𝑀/𝑁. Choose preimages 𝑀 ∋ 𝑥𝑛 ↦ 𝑥𝑛̄ for all
𝑛. We have 𝑥𝑛+1̄ − 𝑥𝑛̄ ∈ 𝐹𝑖(𝑛) 𝑀 + 𝑁 where lim𝑛→∞ 𝑖(𝑛) = ∞, therefore we can write
𝑥𝑛+1 − 𝑥𝑛 = 𝑦𝑛 + 𝛿𝑛 where 𝑦𝑛 ∈ 𝐹𝑖(𝑛) 𝑀 and 𝛿𝑛 ∈ 𝑁. We contend that 𝑥𝑛′ ∶= 𝑥1 + ∑𝑖<𝑛 𝑦𝑖
is a Cauchy sequence in 𝑀. Indeed, for any 𝑖 > 𝑗 we have 𝑥𝑖′ − 𝑥𝑗′ = ∑𝑗≤𝑘<𝑖 𝑦𝑘 , which
lies in 𝐹inf𝑘 𝑖(𝑘) 𝑀. This implies (𝑥𝑛′ )𝑛 is a Cauchy sequence, hence has a limit 𝑥 ∈ 𝑀. It
is also clear that 𝑥𝑛′ ↦ 𝑥𝑛̄ . Hence 𝑥𝑛′̄ has a limit 𝑥 ̄ = 𝑥 mod 𝑁 in 𝑀/𝑁.
§5.1 Completions ⋅ 53 ⋅
Let us turn to the exactness of completion. This should be understood in the broader
framework of lim of arbitrary projective systems. For simplicity, we only consider 𝐼-adic
←−−
topologies on finitely generated modules over a Noetherian ring.
Observe that any homomorphism 𝜑 ∶ 𝑀 → 𝑁 between 𝑅-modules is automatically
𝔞-adically continuous, for that 𝜑(𝔞𝑛 𝑀) ⊂ 𝔞𝑛 𝑁.
For any 𝑅-module 𝑀 endowed with 𝐼-adic topology, there is a canonical homomor-
phism 𝑀 ⊗ 𝑅̂ → 𝑀.̂ Indeed, 𝑀̂ = lim 𝑀/𝐼 𝑛 𝑀 is a 𝑅̂ = lim 𝑅/𝐼 𝑛 -module by
𝑅 ←−−𝑛 ←−−𝑛
(𝑟𝑛 )𝑛≥1 ⋅ (𝑥𝑛 )𝑛≥1 = (𝑟𝑛 𝑥𝑛 )𝑛≥1 , (𝑟𝑛 )𝑛 ∈ 𝑅,̂ (𝑥𝑛 )𝑛 ∈ 𝑀,̂
𝑅⊕𝑎 ⊗ 𝑅̂ 𝑅⊕𝑎 ⊗ 𝑅̂ 𝑀 ⊗ 𝑅̂ 0
𝑅 𝑅 𝑅
̂
𝑅⊕𝑎 ̂
𝑅⊕𝑏 𝑀̂ 0
∼
Remark 5.1.8. In fact 𝑀 ⊗ 𝑅̂ → 𝑀̂ is also a homeomorphism. It suffices to observe that
𝑅
in the rows of the commutative diagram above, 𝑀 ⊗ 𝑅̂ and 𝑀̂ are both realized as
𝑅
̂ ∼
quotient topological 𝑅-modules, and that 𝑅⊕⋆ ⊗ 𝑅̂ → 𝑅̂ ⊕⋆ is a homeomorphism. The
𝑅
second point has been observed in the proof of Proposition 5.1.6, and the first follows
̂
from the fact completed modules carry the 𝐼-adic topology. See Proposition 5.2.2. We
do not need this result.
In the following statements, 𝑅 is Noetherian and an ideal 𝐼 ⊊ 𝑅 is chosen.
Proof. Flatness can be tested on short exact sequences of the form 0 → 𝔞 → 𝑅 → 𝑅/𝔞 →
0 where 𝔞 is a finitely generated ideal of 𝑅. Its base-change to 𝑅̂ is the same as comple-
tion, and completion is an exact functor by Proposition 5.1.6.
Corollary 5.1.10. Assume 𝑅 is 𝐼-adically complete Hausdorff. Then every finitely generated
𝑅-module 𝑀 is 𝐼-adically complete Hausdorff, and any submodule 𝑁 ⊂ 𝑀 is closed.
Proof. For the first assertion: the completion of 𝑀 can be identified with the composite
𝑀 = 𝑀 ⊗ 𝑅 → 𝑀 ⊗ 𝑅̂ → 𝑀,̂ which is bijective. Therefore every Cauchy sequence in 𝑀
𝑅 𝑅
has a limit in 𝑀.
As to the second assertion, recall that the 𝐼-adic topology on 𝑁 is the same as the
one restricted from 𝑀 by (5–1). It remains to notice that complete subspaces must be
closed.
§5.2 Further properties of completion ⋅ 55 ⋅
𝔞𝑀 ⊗ 𝑅̂ 𝑀̂
𝑅
The upper horizontal arrow is an isomorphism by Theorem 5.1.7, therefore the diago-
nal arrow has image equal to 𝔞𝑀.̂ The lower horizontal arrow is just the completion of
𝔞𝑀 ↪ 𝑀, thus injective with image (𝔞𝑀)∧ by Proposition 5.1.6. A comparison yields
(𝔞𝑀)∧ = 𝔞𝑀.̂ Also note that 𝔞𝑀 ̂ ̂ = 𝔞𝑅̂ 𝑀̂ = 𝔞𝑀.̂ The final assertion results from the
exactness of completion.
̂ 𝑛 𝑀̂ = 𝑀/
𝑀/𝐼 𝑛 𝑀 = 𝑀/𝐼 ̂ 𝐼 𝑛̂ 𝑀,̂ ∀𝑛 ≥ 0,
̂ = gr (̂ 𝑅),
gr (𝑅) = gr (𝑅) ̂
𝐼 𝐼 𝐼
̂ = gr (̂ 𝑀).
gr𝐼 (𝑀) = gr𝐼 (𝑀) ̂
𝐼
Proposition 5.2.2. For any finitely generated 𝑅-module 𝑀, the topology on 𝑀̂ coincides with
̂
the 𝐼-adic one.
Proof. Consider the closure of the image of 𝐼 𝑛 𝑀 in 𝑀.̂ It is readily seen to be {(𝑥𝑘 )𝑘 ∶ 𝑖 ≤
𝑛 ⟹ 𝑥𝑖 = 0} = 𝐹𝑛 𝑀.̂ On the other hand, we have seen that this closure is 𝐼̂ 𝑛 𝑀 ⊂ 𝑀. ̂
By virtue of Proposition 5.2.1, we have 𝐼̂
𝑛 𝑀 = 𝐼̂𝑀̂ and 𝐼̂ = 𝐼 𝑅̂ = (𝐼 𝑅)
𝑛 𝑛 𝑛 ̂ = 𝐼̂ .
𝑛 𝑛
Proof. Let 𝑦 ∈ 𝐹𝑑 𝑁, we may take 𝑥 ∈ 𝐹𝑑 𝐿 such that 𝑦′ ∶= 𝑦 − 𝜑(𝑥) ∈ 𝐹𝑑+1 𝑁. Next, take
𝑥′ ∈ 𝐹𝑑+1 𝐿 with 𝑦″ ∶= 𝑦′ − 𝜑(𝑥′ ) ∈ 𝐹𝑑+2 𝑁, and so forth. Use the completeness of 𝐿 to
define 𝑥∞ ∶= 𝑥 + 𝑥′ + 𝑥″ + ⋯, which maps to 𝑦 since 𝑁 is Hausdorff.
Proof. Let 𝔄 be any ideal of 𝑅,̂ equipped with the filtration 𝐹𝑛 𝔄 ∶= 𝐼 𝑛̂ ∩ 𝔄. We have
to show 𝔄 is finitely generated. Since gr𝐼 (𝑅) = gr𝐼 (̂ 𝑅) ̂ is Noetherian, so is gr (𝔄).
𝐹
𝑑
Take 𝑡1 , … , 𝑡𝑛 ∈ 𝔄, 𝑡𝑖 ∈ 𝐹𝑑𝑖 𝔄, whose images 𝑡𝑖̄ in gr𝐹𝑖 (𝔄) generates gr𝐹 (𝔄). Using
an appropriately shifted filtration on 𝐿 ∶= 𝑅̂ ⊕𝑛 , we obtain a filtered homomorphism
𝜑 ∶ 𝐿 → 𝔄 with image (𝑡1 , … , 𝑡𝑛 ), such that gr(𝜑) is surjective. Now apply the previous
Lemma to obtain the first assertion.
One of the characterizations of Jacobson radical says that 𝐼 ̂ ⊂ rad(𝑅) ̂ if and only if
̂ ̂ × −1 2
1 − 𝐼 ⊂ 𝑅 . This is verified by noting that (1 − 𝑡) = 1 + 𝑡 + 𝑡 + ⋯ converges 𝐼-adically.
This proves the second assertion.
̂ satisfying
Proposition 5.2.5. The map 𝔭 ↦ 𝔭̂ furnishes an injection from 𝑉(𝐼) to Spec(𝑅)
1∶1
̂ 𝔭̂ (as rings). It restricts to a bijection MaxSpec(𝑅) ∩ 𝑉(𝐼)
𝑅/𝔭 ≃ 𝑅/ ̂
MaxSpec(𝑅).
Consequently, if 𝑅 is local (resp. semi-local), so is 𝑅.̂
𝑅(𝜂)𝛾 ∶= 𝑅𝛾+𝜂 .
𝑛
⨁𝑖=1 𝑅(−𝜂𝑖 ) 𝑀
(5–2)
(… , 0, 1⎵ , 0, …) 𝑥𝑖 .
𝑖−th slot
Lemma 5.3.1. For 𝑅 as above and 𝑀 a finitely generated graded 𝑅-module, each graded piece
𝑀𝛾 is an 𝑅0 -module of finite length.
Proof. Using (5–2) this is readily reduced to the case 𝑀 = 𝑅(−𝜂), and then to 𝑀 = 𝑅.
Write 𝑅 = 𝑅0 [𝑥1 , … , 𝑥𝑛 ] where each 𝑥𝑖 is homogeneous of degree 𝑑𝑖 . Given 𝛾, the 𝑅0 -
𝑎 𝑎
module 𝑀𝛾 is generated by monomials 𝑥11 ⋯ 𝑥𝑛𝑛 with ∑𝑖 𝑎𝑖 𝑑𝑖 = 𝛾 and 𝑎1 , … , 𝑎𝑛 ∈ ℤ≥0 ;
this admits only finitely many solutions (𝑎1 , … , 𝑎𝑛 ). We conclude that 𝑀𝛾 has finite
length since 𝑅0 /𝑅0 ∩ ann(𝑀) is an Artinian ring.
Recall that saying a module 𝑁 over a ring 𝐴 has finite length means that there exists
a composition series
⋅𝑥𝑖 ⋅𝑥𝑖
𝑍 ∶= ker (𝑀 𝑀(𝜂𝑖 )) , 𝑌 ∶= coker (𝑀(−𝜂𝑖 ) 𝑀)
which are again finitely generated, so that we have the exact sequence
0 → 𝑍𝛾 → 𝑀𝛾 → 𝑀𝛾+𝜂𝑖 → 𝑌𝛾+𝜂𝑖 → 0, 𝛾 ∈ 𝛤.
the right-hand side being quasi-polynomials of period lcm(… , 𝜂̂𝑖 , …) for large |𝛾| and
of degrees ≤ 𝑛−2, since 𝑥𝑖 acts trivially on 𝑍 and 𝑌. Doing this for all 𝑖 yields difference
equations that witness the polynomiality of 𝜒(𝑀, 𝛾) for |𝛾| ≫ 0 in every congruence
class modulo 𝑒.
In particular, if 𝑅 is generated by 𝑅1 over 𝑅0 , the period 𝑒 = 1 and we have the
notion of Hilbert–Samuel polynomials.
Definition 5.4.1 (Height and dimension). For any prime ideal 𝔭 of 𝑅, define its height
ht(𝔭) as the supremum of the lengths of prime chains
𝔭 = 𝔭 0 ⊋ 𝔭 1 ⊋ ⋯ ⊋ 𝔭𝑛 , length ∶= 𝑛.
⋄ Fields have dimension zero. In fact, a ring has dimension zero if and only if every
prime ideal is maximal.
Exercise 5.4.3. Show that every principal ideal domain which is not a field has dimen-
sion one.
Lemma 5.4.4. Suppose 𝑅 is Noetherian. The following are equivalent for a finitely generated
𝑅-module 𝑀 ≠ {0}.
(i) dim 𝑀 = 0.
Proof. (i) ⟺ (ii) is already known: recall that a Noetherian ring is Artinian if and
only if its prime ideals are all maximal (Corollary 1.4.3). Let us show (i) or (ii) ⟹
(iii). By writing 𝑀 = 𝑀1 + ⋯ + 𝑀𝑛 where each 𝑀𝑖 is generated by one element, we
may assume 𝑀 ≃ 𝑅/𝔞 for some ideal 𝔞 = ann(𝑀). It has been shown that 𝑅/𝔞 has
finite length as a module since it is an Artinian ring.
(iii) ⟹ (i). Upon modulo ann(𝑀) we may assume ann(𝑀) = {0}. Take any
minimal prime 𝔭 in 𝑅. As ann(𝑀) = {0} we have 𝑀𝔭 ≠ {0}. Therefore 𝔭 is a minimal
element of Supp(𝑀), hence belongs to Ass(𝑀). We may embed 𝑅/𝔭 into 𝑀. The 𝑅-
module 𝑅/𝔭 has finite length since 𝑀 does, therefore 𝑅/𝔭 is an Artinian ring. This
implies 𝔭 is a maximal ideal, therefore dim 𝑅 = 0 since every prime in 𝑅 lies over a
minimal prime.
Our strategy is to study the Krull dimension via completions and Hilbert polyno-
mials. As a preparation, we begin with the local, or more generally the semi-local rings.
Definition 5.4.5. Let 𝑅 be a Noetherian semi-local ring (i.e. there are finitely many
maximal ideals 𝔪1 , … , 𝔪𝑛 ). Let 𝑀 ≠ {0} be a finitely generated 𝑅-module. We say an
ideal 𝐼 is a parameter ideal for 𝑀 if 𝐼 ⊂ rad(𝑅) and 𝑀/𝐼𝑀 has finite length.
Parameter ideals are often called ideals of definition. Here we follow the terminolo-
gies of [8].
Exercise 5.4.6. Show that 𝐼 is a parameter ideal for 𝑅 if and only if there exists 𝑘 with
rad(𝑅)𝑘 ⊂ 𝐼 ⊂ rad(𝑅).
Show that such an ideal is a parameter ideal for every 𝑀. Hint: If 𝐼 ⊃ rad(𝑅)𝑘 , every
prime ideal 𝔭 ⊃ 𝐼 must contain (𝔪1 ⋯ 𝔪𝑛 )𝑘 , hence 𝔭 = 𝔪𝑖 for some 𝑖. Conversely, show
that in an Artinian ring we have rad(𝑅)𝑘 = 0 for 𝑘 ≫ 0, using Corollary 1.4.3. Hint: for
Artinian rings, rad(𝑅) equals the nilpotent radical, and is finitely generated.
⋅ 60 ⋅ From completions to dimensions
Dimension theory for modules can be built solely on the parameter ideals for 𝑅, but
we opt to introduce the general notion here.
Hereafter we fix a Noetherian semi-local ring 𝑅 and a finitely generated 𝑅-module
𝑀 ≠ {0}.
Lemma 5.4.7. An ideal 𝐼 ⊂ rad(𝑅) is a parameter ideal for 𝑀 if and only if there exists 𝑘 with
rad(𝑅)𝑘 ⊂ 𝐼 + ann(𝑀). In this case 𝑅/(𝐼 + ann(𝑀)) is an Artinian ring.
In particular, rad(𝑅) is a parameter ideal for any 𝑀.
Proof. First we claim that 𝑉(ann(𝑀/𝐼𝑀)) = Supp(𝑀/𝐼𝑀) equals 𝑉(𝐼 + ann(𝑀)). By
the exactness of localizations together with Nakayama’s Lemma, we have
Supp(𝑀/𝐼𝑀) = Supp(𝑀) ∩ {𝔭 ∶ 𝐼𝑅𝔭 ⊊ 𝑅𝔭 } ;
the last term equals Supp(𝑀) ∩ 𝑉(𝐼) = 𝑉(ann(𝑀) + 𝐼), thereby proving our claim.
By applying to 𝑀/𝐼𝑀 the Lemma 5.4.4, 𝐼 ⊂ rad(𝑅) being a parameter ideal for 𝑀 is
equivalent to any one of the following
𝑅
is Artinian ⟺ 𝑉(ann(𝑀/𝐼𝑀)) ⊂ MaxSpec(𝑅)
ann(𝑀/𝐼𝑀)
⟺ 𝑉(𝐼 + ann(𝑀)) ⊂ MaxSpec(𝑅)
𝑅
⟺ 𝑅̄ ∶= is Artinian.
𝐼 + ann(𝑀)
If 𝑅̄ is Artinian, then the image of rad(𝑅) in 𝑅̄ is contained in rad(𝑅),
̄ and we know
rad(𝑅) ̄ = 0 for large 𝑘.
𝑘
Recall that
⋄ gr𝐼 (𝑅) is finitely generated over gr0𝐼 (𝑅) = 𝑅/𝐼 as an algebra and is Noetherian
(Proposition 4.4.4);
⋄ more precisely, gr𝐼 (𝑅) is generated by gr1𝐼 (𝑅) over 𝑅/𝐼.
⋄ gr𝐼 (𝑀) is a finitely generated gr𝐼 (𝑅)-module (Proposition 4.4.5);
⋄ the ring gr0𝐼 (𝑅) = 𝑅/𝐼 becomes Artinian after modulo gr0𝐼 (𝑅) ∩ ann(gr𝐼 (𝑀)),
which contains (ann(𝑀) + 𝐼)/𝐼 (use Lemma 5.4.7).
Upon recalling Lemma 5.3.1, it are justified to define
𝑛−1
𝜒(𝑀, 𝐼; 𝑛) ∶= ℓ𝑅/𝐼 𝑛 (𝑀/𝐼 𝑛 𝑀) = ∑ ℓ𝑅/𝐼 (𝐼 𝑗 𝑀/𝐼 𝑗+1 𝑀), 𝑛 ∈ ℤ≥0 .
𝑗=0
Lemma 5.4.8. Let 𝑀 ≠ {0} be a finitely generated 𝑅-module with parameter ideal 𝐼.
(i) The degree 𝑑(𝑀) of 𝐻𝐼 (𝑀, ⋅) is independent of the choice of the parameter ideal 𝐼.
(ii) In a short exact sequence 0 → 𝑀′ → 𝑀 → 𝑀″ → 0 of finitely generated 𝑀-modules, we
have
deg 𝐻𝐼 (𝑀′ , ⋅), deg 𝐻𝐼 (𝑀″ , ⋅) ≤ deg 𝐻𝐼 (𝑀, ⋅),
and 𝐻𝐼 (𝑀, ⋅) − 𝐻𝐼 (𝑀′ , ⋅) − 𝐻𝐼 (𝑀″ , ⋅) has degree < 𝑑(𝑀).
Proof. (i): To compare the graded objects associated to two parameter ideals 𝐼, 𝐽, we
apply the characterization in Lemma 5.4.7: it suffices to take 𝐽 = rad(𝑅), so that
for some 𝑚 ≥ 1. This implies 𝜒(𝑀, 𝐽; 𝑛) ≤ 𝜒(𝑀, 𝐼; 𝑛) and 𝜒(𝑀, 𝐼; 𝑛) ≤ 𝜒(𝑀, 𝐽; 𝑚𝑛) for
all 𝑛 ≥ 0. Whence (i).
(ii): For the first part, note that 𝑀/𝐼 𝑛 𝑀 → 𝑀″ /𝐼 𝑛 𝑀″ is surjective, so 𝜒(𝑀″ , 𝐼; 𝑛) ≤
′ 𝑛
𝜒(𝑀, 𝐼; 𝑛). On the other hand, 𝑀′ /𝐼 𝑛 𝑀′ → 𝑀/𝐼 𝑛 𝑀 has kernel 𝑀𝐼 𝑛∩𝐼𝑀′𝑀 . For 𝑛 ≥ 𝑛0 ≫ 0,
Artin–Rees (Theorem 4.5.7) gives
Theorem 5.5.1 (Krull). Suppose 𝔞 = (𝑡1 , … , 𝑡𝑟 ) is a proper ideal of 𝑅, then for every minimal
over-prime ideal 𝔭 of 𝔞, we have ht(𝔭) ≤ 𝑟.
From Definition 5.4.1 we infer that ht(𝔞) ≤ 𝑟. The special case 𝑟 = 1 says that every
principal ideal (𝑡) ≠ 𝑅 has height at most one (exactly one if 𝑡 is not a zero divisor —
use Theorem 2.2.5 (ii)); this is called the Hauptidealsatz.
Now assume 𝑅 is Noetherian and local with maximal ideal 𝔪. The parameter ideals
of 𝑅 are precisely those squeezed between 𝔪 and 𝔪𝑘 for some 𝑘 ≥ 1, by Lemma 5.4.7.
Set 𝑑 ∶= dim 𝑅, which is finite by Theorem 5.4.9. The same theorem tells us that we can
generate some parameter ideal 𝐼 ⊂ 𝔪 (namely 𝑅/𝐼 Artinian) by elements 𝑡1 , … , 𝑡𝑑 ∈ 𝔪.
These elements form a system of parameters of 𝑅.
Proof. Consider the sequence 𝑅, 𝑅/(𝑡1 ), 𝑅/(𝑡1 , 𝑡2 ), … , 𝑅/(𝑡1 , … , 𝑡𝑑 ). Recall the formal-
ism in Theorem 5.4.9: at each stage 𝑑(𝑅/ ⋯) drops at most by one, by (5–3). After 𝑑
steps we arrive at 𝑅/𝐼 with dim(⋅) = 𝑑(⋅) = 𝑠(⋅) = 0, since it has finite length. Hence 𝑑
drops exactly by one at each stage. The remaining assertions are immediate.
Definition 5.5.3. We say a Noetherian local ring 𝑅 is a regular local ring if 𝔪 can be
generated by 𝑑 = dim 𝑅 elements 𝑡1 , … , 𝑡𝑑 . In this case we say 𝑡1 , … , 𝑡𝑟 form a regular
system of parameters.
Exercise 5.5.4. Let 𝕜 be a field. The 𝕜-algebra of formal power series 𝕜J𝑋1 , … , 𝑋𝑑 K is
a regular local ring. Indeed, it is Noetherian with maximal ideal 𝔪 = (𝑋1 , … , 𝑋𝑑 ), and
𝑋1 , … , 𝑋𝑑 form a regular system of parameters. On the other hand, 𝔪/𝔪2 has a 𝕜-basis
formed by the images of 𝑋1 , … , 𝑋𝑑 . One way to determine its dimension and prove its
regularity is to calculate the functions 𝑛 ↦ dim𝕜 (𝔪𝑛 /𝔪𝑛+1 ) explicitly, i.e. count the
monomials in 𝑑 variables with total degree 𝑛. You should get a polynomial in 𝑛 with
degree 𝑑 − 1, cf. Exercise 5.3.4.
Theorem 5.5.5. For any Noetherian local ring 𝑅 with maximal ideal 𝔪 and residue field 𝕜, we
have dim 𝑅 ≤ dim𝕜 𝔪/𝔪2 . Equality holds if and only if 𝑅 is a regular local ring.
⋅ 64 ⋅ From completions to dimensions
Proof. By Nakayama’s Lemma (more precisely, Corollary 1.3.6), 𝔪/𝔪2 can be generated
over 𝕜 by 𝑠 elements if and only if 𝔪 can be generated over 𝑅 by 𝑠 elements, for any
𝑠 ∈ ℤ≥0 . Hence Theorem 5.4.9 imposes the bound 𝑠 ≥ dim 𝑅, and equality holds if and
only if 𝑅 admits a regular system of parameters.
Due to time constraints, we cannot say too much about regular local rings. Below
is one of their wonderful properties.
dim 𝑅′ = dim 𝑅 − 1,
dim𝕜 𝔪′ /(𝔪′ )2 = dim𝕜 𝔪/𝔪2 − 1 = dim 𝑅 − 1.
Hence 𝑅′ is still regular local, and by induction it is a domain. This implies (𝑡) is prime.
Take any minimal prime 𝔭 below (𝑡); note that 𝑡 ∉ 𝔭 by construction. To show 𝑅 is a
domain it suffices to prove 𝔭 = {0}. Indeed, every 𝑠 ∈ 𝔭 can be written as 𝑠 = 𝑎𝑡, 𝑎 ∈ 𝑅.
Since 𝑡 ∉ 𝔭, we must have 𝑎 ∈ 𝔭. Hence 𝔭 = 𝑡𝔭 ⊂ 𝔪𝔭. Nakayama’s Lemma (Theorem
1.3.5) implies 𝔭 = {0}.
Lecture 6
Dimension of finitely
generated algebras
The main reference is [11, §13].
The fiber (𝜑♯ )−1 (𝔭) is then identified with the spectrum of
which is empty if and only if 𝐵 ⊗ 𝜅(𝔭) is zero. This also equips (𝜑♯ )−1 (𝔭) with an extra
𝐴
structure: it is the spectrum of an explicit quotient ring of 𝐵𝔭 .
Observe that for all 𝔮 ∈ (𝜑♯ )−1 (𝔭), the localization of 𝐵𝔭 ⊗ 𝜅(𝔭) at the image of 𝔮 is
𝐴𝔭
canonically isomorphic to 𝐵𝔮 ⊗ 𝜅(𝔭).
𝐴𝔭
(iii) if going-down holds and 𝜑♯ is surjective, then dim(𝐵) ≥ dim(𝐴), and for all ideal 𝔞 ⊊ 𝐴
we have 𝜑(𝔞)𝐵 ≠ 𝐵 and ht(𝔞) = ht(𝜑(𝔞)𝐵).
It is crucial to notice that dim 𝐵𝔮 ⊗ 𝜅(𝔭) = ht(𝔮𝐵𝔮 /𝜑(𝔭)𝐵𝔮 ) = ht(𝔮/𝜑(𝔭)𝐵).
𝐴𝔭
This bounds the source dimension of a morphism by the target dimension plus the
fiber dimension, localized both at 𝔭 and 𝔮 ∈ (𝜑♯ )−1 (𝔭). In order to have an equality, a
certain submersion-like condition on 𝜑♯ is evidently required; this explains the going-
down condition. Cf. [11, (6.H)].
Proof. We have an induced local homomorphism 𝐴𝔭 → 𝐵𝔮 since 𝔮 ↦ 𝔭. Since (i) and
(ii) depend only on this induced homomorphism, we may assume from the outset that
𝐴, 𝐵 are local with maximal ideals 𝔭, 𝔮, and 𝜑 is a local homomorphism. Let 𝑑 ∶= dim 𝐴
and take a parameter ideal 𝐼 = (𝑡1 , … , 𝑡𝑑 ) of 𝐴, so that 𝔭𝑘 ⊂ 𝐼 ⊂ 𝔭 for some 𝑘. It follows
that √𝜑(𝔭)𝐵 = √𝜑(𝐼)𝐵, therefore dim(𝐵 ⊗ 𝜅(𝔭)) = dim(𝐵/𝜑(𝔭)𝐵) = dim(𝐵/𝜑(𝐼)𝐵);
𝐴
denote this number as 𝑒. Take a 𝑠1 , … , 𝑠𝑒 ∈ 𝔮 whose images generate a parameter ideal
for 𝐵/𝜑(𝐼)𝐵, and put 𝐽 ∶= (𝜑(𝑡1 ), … , 𝜑(𝑡𝑑 ), 𝑠1 , … , 𝑠𝑒 ). Then 𝐵/𝐽 is Artinian, therefore
dim 𝐵 ≤ 𝑑 + 𝑒 establishes (i).
As for (ii), we conserve the same hypotheses and take a prime chain 𝔮 = 𝔮0 ⊋ ⋯ ⊋
𝔮𝑒 with 𝔮𝑒 ⊃ 𝜑(𝔭)𝐵 in 𝐵, as well as a prime chain 𝔭 = 𝔭0 ⊋ ⋯ ⊋ 𝔭𝑑 in 𝐴. Note that
𝜑−1 (𝔮𝑒 ) = 𝔭 since 𝑒 = dim 𝐵/𝜑(𝔭)𝐵. By applying going-down repeatedly to the chain
𝔭𝑖 , we obtain a prime chain in 𝐵
To obtain ≤, choose 𝔭 ⊃ 𝔞 with ht(𝔭) = ht(𝔞) and take 𝔮 ∈ Spec(𝐵) with 𝔭 = 𝜑−1 (𝔮);
this implies 𝔮 ⊃ 𝜑(𝔭)𝐵 ⊃ 𝜑(𝔞)𝐵. Upon shrinking 𝔮, we may even assume 𝔮 is minimal
over 𝜑(𝔭)𝐵, i.e. ht(𝔮/𝜑(𝔭)𝐵) = 0. Using (ii), this entails ht(𝔞) = ht(𝔭) = ht(𝔮) ≥
ht(𝜑(𝔞)𝐵).
Going-down holds for flat 𝜑 by Theorem 4.1.3, therefore the dimension equality
dim 𝐵𝔮 = dim 𝐴𝔭 + 1
⋅ 68 ⋅ Dimension of finitely generated algebras
As 𝜅[𝑋] is a principal ideal domain which is not a field, every maximal ideal thereof
has height one. Hence dim 𝜅[𝑋]𝔮′ = ht(𝔮′ ) = 1.
For the initial case 𝑒 = 𝑛, simply take 𝑥𝑖′ ∶= 𝑌𝑖 . Our aim is 𝑒 = 0. Let us explain the
induction step from 𝑒 ≥ 1 to 𝑒 − 1. The prior argument based on transcendence degrees
implies the algebraic independence among
If 𝑒 ≤ 𝑑𝑗 for all 𝑗, we are done. Otherwise set 𝑗 ∶= min{𝑗′ ∶ 𝑒 > 𝑑𝑗′ }, we contend that
For a general Noetherian domain 𝐵, we call a prime chain maximal if it is not properly
contained in any prime chain. A priori, a maximal prime chain does not necessarily
have length equal to dim 𝐵. If
we say 𝐵 is a catenary domain. This means that ht(𝔮′ /𝔮) = dim 𝐵𝔮′ /𝔮𝐵𝔮′ is the common
length of all maximal prime chains between 𝔮′ and 𝔮. In particular, maximal prime
chains in 𝐵𝔮′ are automatically longest. As shown above, finitely generated domains
over a field are catenary. For a finer analysis of catenary and universally catenary rings,
we refer to [11, §14].
Corollary 6.3.5. Suppose 𝐵 is a domain finitely generated over a field 𝕜. Set 𝐿 ∶= Frac(𝐵).
Then dim 𝐵 = tr.deg𝕜 (𝐿) and it is the common length of maximal prime chains.
which equals 𝑛.
Remark 6.3.6. Recall that finitely generated domains over an algebraically closed field
𝕜 are objects “opposite” to the irreducible affine 𝕜-varieties. It is instructive to make a
comparison with the analytic theory when 𝕜 = ℂ. Let 𝒳 be a compact connected com-
plex manifold of (complex) dimension 𝑛. Denote by ℳ(𝒳) the field of meromorphic
functions on 𝒳. Siegel proved that tr.degℂ (ℳ(𝒳)) ≤ 𝑛. When 𝒳 is a projective alge-
braic ℂ-variety, equality holds and we have ℳ(𝒳) = Frac(𝐴) if Spec(𝐴) is any open
dense affine subscheme in 𝒳. In general, the abundance of meromorphic functions
on 𝒳 is a subtle issue, cf. the case of Riemann surfaces (𝑛 = 1). Compact connected
complex manifolds with tr.degℂ (ℳ(𝒳)) = dimℂ 𝒳 are called Moishezon manifolds.
Non-algebraic Moishezon manifolds do exist, and Moishezon proved that a Moishe-
zon manifold is projective, hence algebraic, if and only if it is Kähler. Following M.
Artin and D. Knutson, one can enlarge the category of ℂ-schemes into that of alge-
braic spaces over ℂ, and there is an analytification functor 𝒳 ↦ 𝒳an that sends algebraic
spaces of finite type over ℂ to complex analytic varieties. M. Artin [2, §7] showed that
the analytification establishes an equivalence between the category of smooth proper
algebraic spaces of finite type over ℂ and that of Moishezon manifolds. Therefore,
such manifolds still retain an algebraic flavor: they are quotients of certain ℂ-schemes
by étale equivalence relations.
⋅ 72 ⋅ Dimension of finitely generated algebras
Lecture 7
Exercise 7.1.4. Show that 𝑡 is a uniformizer if and only if it generates the maximal ideal
of 𝑅.
Recall that a regular local ring 𝑅 with dim 𝑅 = 1 is a Noetherian local ring whose
maximal ideal 𝔪 is principal and nonzero; elements generating 𝔪 are called the regular
parameters for 𝑅.
Proposition 7.1.6. Suppose 𝑡 is a regular parameter in a regular local ring 𝑅 of dimension one,
then 𝑅 is a domain, and every element 𝑥 ∈ 𝑅 ∖ {0} can be uniquely written as 𝑥 = 𝑡𝑟 𝑢 with
𝑟 ≥ 0 and 𝑢 ∈ 𝑅× . This makes 𝑅 into a discrete valuation ring by setting 𝑣(𝑥) = 𝑟, for which
𝑡 is a uniformizer.
Therefore 𝑅 is a discrete valuation ring. Conversely, every discrete valuation ring is regular
local of dimension 1.
Proof. From Theorem 5.5.6 we know regular local rings are Noetherian domains. By
applying Krull’s Intersection Theorem (Corollary 4.5.10) to the powers of (𝑡), we see
that 𝑟 ∶= sup{𝑘 ≥ 0 ∶ 𝑥 ∈ (𝑡)𝑘 } is finite. Write 𝑥 = 𝑡𝑟 𝑢. Since 𝑅× = 𝑅 ∖ 𝔪, we see 𝑢 ∈ 𝑅× .
As to uniqueness, suppose 𝑡𝑟 𝑢 = 𝑡𝑠 𝑤 with 𝑟 ≥ 𝑠, then 𝑡𝑟−𝑠 = 𝑢−1 𝑤 ∈ 𝑅× implies 𝑟 = 𝑠,
hence 𝑢 = 𝑤 as 𝑅 is a domain. As every element of Frac(𝑅)× is uniquely expressed
as 𝑡𝑟 𝑢 with 𝑟 ∈ ℤ, one readily checks that 𝑣(𝑡𝑟 𝑢) = 𝑟 satisfies all the requirements of
discrete valuation.
The converse direction has been addressed in Lemma 7.1.3.
To recap, in dimension one we have
Exercise 7.1.7. Explain that the regular local rings of dimension 0 are just fields.
Lemma 7.2.1. Let 𝑀 be a finitely generated 𝑅-module. An element 𝑥 ∈ 𝑀 is zero if and only
if its image in 𝑀𝔭 is zero for every maximal element 𝔭 in Ass(𝑀).
Proof. Suppose 𝑥 ≠ 0. Since 𝑀 is Noetherian, among ideals of the form ann(𝑦) there is
a maximal one containing ann(𝑥), and we have seen in Lemma 2.2.3 that such an ideal
𝔭 belongs to Ass(𝑀). Since ann(𝑥) ⊂ 𝔭, we have 𝑥/1 ∈ 𝑀𝔭 ∖ {0}.
Call a ring reduced if it has no nilpotent element except zero.
§7.2 Auxiliary results on the total fraction ring ⋅ 75 ⋅
For the next result, we denote by 𝑇 the set of non zero-divisors of 𝑅. Recall that the
total fraction ring 𝐾(𝑅) is 𝑅[𝑇 −1 ]; this is the largest localization such that 𝑅 → 𝐾(𝑅) is
injective, and 𝐾(𝑅) = Frac(𝑅) when 𝑅 is a domain. The map 𝔭 ↦ 𝔭𝐾(𝑅) sets up an
∼
order-preserving bijection Ass(𝑅) → Ass(𝐾(𝑅)): indeed, if 𝔭 ∋ 𝑡 for some 𝑡 ∈ 𝑇, then
𝔭 cannot belong to Ass(𝑅) because the union of Ass(𝑅) equals 𝑅 ∖ 𝑇.
∼
Lemma 7.2.3. Let 𝑅 be reduced. Then 𝐾(𝑅) → ∏𝔭 𝐾(𝑅/𝔭) as 𝑅-algebras, where 𝔭 ranges
over the minimal prime ideals of 𝑅. For any multiplicative subset 𝑆 ⊂ 𝑅 there is a canonical
isomorphism of 𝑅[𝑆−1 ]-algebras
∼
𝐾(𝑅[𝑆−1 ]) → 𝐾(𝑅)[𝑆−1 ].
In other words, the formation of total fraction ring commutes with localizations.
Proof. Each element in 𝐾(𝑅) = 𝑅[𝑇 −1 ] is either a zero-divisor or invertible. The set
of zero-divisors of 𝐾(𝑅) is the union of minimal prime ideals 𝔭𝑖 𝐾(𝑅) of 𝐾(𝑅) (where
Ass(𝑅) = {𝔭1 , … , 𝔭𝑚 } by an earlier discussion), therefore each prime ideal of 𝐾(𝑅) must
equal some 𝔭𝑖 𝐾(𝑅), by prime avoidance (Proposition 1.1.5). Hence 𝔭1 𝐾(𝑅), … , 𝔭𝑚 𝐾(𝑅)
are also the maximal ideals in 𝐾(𝑅), with zero intersection. Chinese Remainder The-
𝑚
orem entails that 𝐾(𝑅) ≃ ∏𝑖=1 𝐾(𝑅)/𝔭𝑖 𝐾(𝑅). To conclude the first part, notice that
𝐾(𝑅)/𝔭𝑖 𝐾(𝑅) = (𝑅/𝔭𝑖 )[𝑇 −1 ]; this is a field in generated by an isomorphic copy of 𝑅/𝔭𝑖
since 𝔭𝑖 ∩ 𝑇 = ∅, hence equals Frac(𝑅/𝔭𝑖 ).
As for the second part, one decomposes 𝐾(𝑅[𝑆−1 ]) and 𝐾(𝑅)[𝑆−1 ] by the previous
step, noting that
⋄ 𝑅[𝑆−1 ] is reduced;
⋄ Ass(𝑅[𝑆−1 ]) = {𝔭𝑖 𝑅[𝑆−1 ] ∶ 1 ≤ 𝑖 ≤ 𝑚, 𝔭𝑖 ∩ 𝑆 = ∅} consists of minimal primes;
⋄ 𝐾 (𝑅[𝑆−1 ]/𝔭𝑖 𝑅[𝑆−1 ]) ≃ 𝐾(𝑅/𝔭𝑖 ) = 𝐾(𝑅/𝔭𝑖 )[𝑆−1 ] when 𝔭𝑖 ∩ 𝑆 = ∅, by the argu-
ments above;
⋄ 𝐾(𝑅/𝔭𝑖 )[𝑆−1 ] = {0} when 𝔭𝑖 ∩ 𝑆 ≠ ∅.
A term-by-term comparison finishes the proof.
Lemma 7.2.4. Suppose 𝑅 is reduced. Then 𝑥 ∈ 𝐾(𝑅) belongs to 𝑅 if and only if its image in
𝐾(𝑅)𝔭 = 𝐾(𝑅𝔭 ) belongs to 𝑅𝔭 for every prime 𝔭 associated to a non zero-divisor.
Proof. Only the “if” direction requires a proof. Write 𝑥 = 𝑎/𝑡 with 𝑡 not a zero-divisor.
Suppose that 𝑎 ∉ (𝑡), i.e. 𝑎 does not map to zero in 𝑅/(𝑡). Lemma 7.2.1 asserts there
exists 𝔭 ∈ Ass(𝑅/(𝑡)) such that 𝑎 does not map to 0 ∈ (𝑅/(𝑡))𝔭 = 𝑅𝔭 /𝑡𝑅𝔭 . It follows
that the image of 𝑎/𝑡 in 𝐾(𝑅𝔭 ) does not lie in 𝑅𝔭 .
⋅ 76 ⋅ Serre’s criterion for normality and depth
7.3 On normality
Fix a Noetherian ring 𝑅.
Proposition 7.3.2. Let 𝑅 be a Noetherian domain. Then 𝑅 is normal if and only if for every
principal ideal (𝑡) ⊂ 𝑅 and every 𝔭 ∈ Ass(𝑅/(𝑡)), the ideal 𝔭𝑅𝔭 is principal.
Proof. Assume the conditions above. To prove the normality of 𝑅, it suffices to use
𝑅 = ⋂ 𝑅𝔭 where 𝔭 ranges over the primes associated to nonzero principal ideals (con-
sequence of Lemma 7.2.4). Indeed, each 𝑅𝔭 is regular, hence normal by Proposition
7.1.6, therefore so is their intersection by the previous exercise.
Conversely, assume 𝑅 is normal and let 𝔭 ∈ Ass(𝑅/(𝑡)) with 𝑡 ≠ 0, we have to show
𝔭𝑅𝔭 is principal. Upon replacing 𝑅 by 𝑅𝔭 and recalling how associated primes behave
under localization, we may even assume 𝑅 is local with maximal ideal 𝔭. Express 𝔭 as
the annihilator of some 𝑥 ̄ ∈ 𝑅/(𝑡) with 𝑥 ∈ 𝑅. Define the fractional ideal
Proof. (iii) ⟹ (i). We have seen that discrete valuation rings are principal ideal rings
of dimension 1, therefore also normal by unique factorization property.
(i) ⟹ (ii). Under the normality assumption, choose any 𝑡 ∈ 𝑅 ∖ {0}. Since
dim 𝑅 = 1 and {0} is a prime ideal, the associated prime of (𝑡) can only be the maximal
ideal 𝔪, which is principal by Proposition 7.3.2. This shows that 𝑅 is regular local.
(ii) ⟹ (iii) is included in Proposition 7.1.6.
𝑅= ⋂ 𝑅𝔭
ht(𝔭)=1
inside Frac(𝑅).
§7.4 Serre’s criterion ⋅ 77 ⋅
Proof. Evidently ⊂ holds. By Lemma 7.2.4 together with Proposition 7.3.2, 𝑅 can be
written as an intersection of 𝑅𝔭 where 𝔭 is associated to some non zero-divisor, such
that 𝔭𝑅𝔭 is principal; it suffices to show ht(𝔭) = 1. From 𝔭 ≠ {0} we see ht(𝔭) ≥ 1; on
the other hand, by Hauptidealsatz or by the discussion on regular local rings, we see
ht(𝔭) = ht(𝔭𝑅𝔭 ) ≤ 1.
Theorem 7.4.2 (J.-P. Serre). A Noetherian ring 𝑅 is a finite direct product of normal domains
if and only if the following two conditions hold.
⊳ S2 For every non zero-divisor 𝑡 of 𝑅, the primes in Ass(𝑅/(𝑡)) are all of height 1; the
primes in Ass(𝑅) are all of height 0.
Assume conversely R1 and S2. They imply the conditions of Lemma 7.4.1, hence 𝑅
is reduced. Now for every prime 𝔭 associated to a non zero-divisor, we have ht(𝔭) = 1
and 𝑅𝔭 is a normal domain by R1 ∧ S2. By Lemma 7.2.4 (as 𝑅 is reduced), 𝑅 is integrally
closed in 𝐾(𝑅): indeed, if 𝑥 ∈ 𝐾(𝑅) is integral over 𝑅, so is its image in 𝐾(𝑅𝔭 ) =
𝑚
𝐾(𝑅)𝔭 for every 𝔭 as above, therefore lies in 𝑅𝔭 . Decompose 𝐾(𝑅) = ∏𝑖=1 𝐾(𝑅/𝔭𝑖 ) as in
Lemma 7.2.3. The idempotents 𝑒𝑖 ∈ 𝐾(𝑅) associated to this decomposition are trivially
integral over 𝑅: 𝑒𝑖2 − 𝑒𝑖 = 0, hence 𝑒𝑖 ∈ 𝑅 for all 𝑖. It follows that 𝑅 = 𝑅𝑒1 + ⋯ +
𝑚
𝑅𝑒𝑚 ∏𝑖=1 𝑅𝑒𝑖 ⊂ 𝐾(𝑅) and one easily checks that 𝑅𝑒𝑖 = 𝑅/𝔭𝑖 .
Finally, since 𝑅 is integrally closed in 𝐾(𝑅), the decomposition above implies 𝑅/𝔭𝑖 is
integrally closed in 𝐾(𝑅/𝔭𝑖 ). All in all, we have written 𝑅 as a direct product of normal
domains.
Exercise 7.4.3. Recall that for an 𝑅-module 𝑀, a prime ideal 𝔭 ∈ Ass(𝑀) is called embed-
ded if 𝔭 is not a minimal element in Ass(𝑀). Show that for 𝑀 = 𝑅, embedded primes
are primes in Ass(𝑅) with height > 0. For 𝑀 = 𝑅/(𝑡) where 𝑡 is not a zero-divisor,
embedded primes are primes in Ass(𝑅/(𝑡)) with height > 1. Use this to rephrase S2
as follows: there are no embedded primes in Ass(𝑅/(𝑡)) (𝑡 not a zero-divisor) or in
Ass(𝑅).
Exercise 7.4.4. Suppose a ring 𝑅 is isomorphic to a direct product ∏𝑖∈𝐼 𝑅𝑖 . Show that
𝑅 is a domain if and only if |𝐼| = 1 (say 𝐼 = {𝑖0 }), and 𝑅𝑖0 is a domain.
Corollary 7.4.5. A Noetherian domain 𝑅 is normal if and only if R1 and S2 hold for 𝑅.
Definition 7.5.1 (Depth of a module). Let 𝐼 be a proper ideal of 𝑅. We define the depth
of 𝑀 relative to 𝐼 as
𝑛
depth𝐼 (𝑀) ∶= inf {𝑛 ≥ 0 ∶ Ext𝑅 (𝑅/𝐼, 𝑀) ≠ 0}
𝑥
(ii) for all 𝑥 ∈ 𝐼, the homomorphism 𝑀 𝑀 is not injective;
𝑥
Proof. In each case we have 𝑀 ≠ {0}. If (i) holds, then 𝑀 𝑀 vanishes on the image
of some nonzero 𝑅/𝐼 → 𝑀, hence (ii). If (ii) holds, the union of Ass(𝑀) will cover 𝐼,
and (iii) follows by prime avoidance. Finally, suppose 𝔭 ∈ Ass(𝑀) ∩ 𝑉(𝐼), there is an
embedding 𝑅/𝔭 ↪ 𝑀, which yields a non-zero 𝑅/𝐼 → 𝑀.
𝑥𝑘
0 → 𝑀/(𝑥1 , … , 𝑥𝑘−1 ) 𝑀/(𝑥1 , … , 𝑥𝑘−1 )
Proof. The case 𝑟 = 1 follows by staring at the long exact sequence attached to 0 →
𝑥1
𝑀 𝑀 → 𝑀/𝑥1 𝑀 → 0. The general case follows by induction on 𝑟.
(i) depth𝐼 (𝑀) is the supremum of the lengths of 𝑀-regular sequences with elements in 𝐼.
(ii) Suppose depth𝐼 (𝑀) < +∞. Every 𝑀-regular sequence with elements in 𝐼 can be ex-
tended to one of length depth𝐼 (𝑀).
(iii) The depth of 𝑀 relative to 𝐼 is finite if and only if 𝑉(𝐼) ∩ Supp(𝑀) ≠ ∅, or equivalently
𝐼𝑀 ≠ 𝑀.
Proof. To prove (i) and (ii), by the previous Lemma we are reduced to show that
depth𝐼 (𝑀) > 0 implies the existence of 𝑥 ∈ 𝐼 which is not a zero-divisor of 𝑀; this
follows from Proposition 7.5.2.
Now pass to the word “equivalently” in (iii). We have 𝑀 ≠ 𝐼𝑀 if and only if
(𝑀/𝐼𝑀)𝔭 = 𝑀𝔭 /𝐼𝔭 𝑀𝔭 ≠ 0 for some prime ideal 𝔭. That quotient always vanishes when
𝔭 ⊅ 𝐼, in which case 𝐼𝔭 = 𝑅𝔭 . On the other hand, when 𝔭 ∈ 𝑉(𝐼) we have 𝐼𝔭 ⊂ rad(𝑅𝔭 ),
thus the non-vanishing is equivalent to 𝑀𝔭 ≠ {0} by Nakayama’s Lemma 1.3.5.
The “if” direction of (iii) is based on the following fact
which will be proved in the next lecture (Theorem 8.3.2) using Koszul complexes. As for
the “only if” direction, 𝑉(𝐼) ∩ Supp(𝑀) = ∅ implies 𝐼 + ann(𝑀) = 𝑅, but the elements
𝑛
in 𝐼 + ann(𝑀) annihilate each Ext (𝑅/𝐼, 𝑀), hence depth𝐼 (𝑀) = +∞.
Corollary 7.5.6. With the same assumptions, let (𝑥1 , … , 𝑥𝑟 ) be an 𝑀-regular sequence with
𝑥𝑖 ∈ 𝐼. It is of length depth𝐼 (𝑀) if and only if Ass(𝑀/(𝑥1 , … , 𝑥𝑟 )𝑀) ∩ 𝑉(𝐼) ≠ ∅.
Proof. The sequence has length depth𝐼 (𝑀) if and only if 𝑅-module 𝑀/(𝑥1 , … , 𝑥𝑟 )𝑀 has
depth zero, so it remains to apply Proposition 7.5.2.
⋅ 80 ⋅ Serre’s criterion for normality and depth
Corollary 7.5.7. Let 𝑅 be a Noetherian local ring with maximal ideal 𝔪, and 𝑀 ≠ {0} a finitely
generated 𝑅-module, then depth𝔪 (𝑀) ≤ dim 𝑀.
Example 7.5.9. Regular local rings are Cohen–Macaulay, although we do not prove this
here. Another important class of Cohen–Macaulay rings is the algebra of invariants
𝐴𝐺 where 𝐴 is the algebra of regular functions on an affine 𝕜-variety 𝒳 with rational
singularities (eg. 𝐴 = 𝕜[𝑋1 , … , 𝑋𝑛 ]) with an action by a reductive 𝕜-group 𝐺 (finite
§7.5 Introduction to depth ⋅ 81 ⋅
groups allowed), and we assume char(𝕜) = 0. Here is the reason: Boutot [6] proved
the GIT quotient 𝒳//𝐺 has rational singularities as well, hence is Cohen–Macaulay;
in characteristic zero this strengthens an earlier theorem of Hochster–Roberts. These
algebras are interesting objects from the geometric, algebraic or even combinatorial
perspectives.
Exercise 7.5.10. Show by using Proposition 7.5.2 that depth𝔪 (𝑅) = 0 if 𝑅 is local with
maximal ideal 𝔪 and has dimension zero.
Proposition 7.5.11. The condition S2 in Theorem 7.4.2 is equivalent to
It is the quotient of the tensor algebra 𝑇(𝐿) = ⨁𝑛≥0 (𝑇 𝑛 (𝐿) ∶= 𝐿⊗𝑛 ) by the graded ideal
generated by the pure tensors
⋯ ⊗ 𝑥 ⊗ 𝑥 ⊗ ⋯, 𝑥 ∈ 𝐿.
0
The multiplication operation in ⋀ 𝐿 is written as ∧. Note that ⋀ 𝐿 = 𝑇 0 (𝐿) = 𝑅
by convention. The traditional notion of exterior algebras encountered in differential
geometry is recovered when ℚ ⊂ 𝑅.
𝑛+1
Given 𝑢 ∈ Hom𝑅 (𝐿, 𝑅), one can define the corresponding contractions 𝑖𝑢 ∶ ⋀ 𝐿→
𝑛
⋀ 𝐿, given concretely as
𝑛
𝑖𝑢 (𝑥0 ∧ ⋯ ∧ 𝑥𝑛 ) = ∑(−1)𝑖 𝑢(𝑥𝑖 ) ⋅ 𝑥0 ∧ ⋯ 𝑥̂𝑖 ⋯ ∧ 𝑥𝑛
𝑖=0
where 𝑥̂𝑖 means 𝑥𝑖 is omitted. It is routine to check that 𝑖𝑢 satisfies 𝑖𝑢 ∘ 𝑖𝑢 = 0, thereby
giving rise to a chain complex.
Definition 8.1.1. Let 𝐿 and 𝑢 be as above. Define the corresponding Koszul complex as
•
𝐾• (𝑢) ∶= (⋀ 𝐿, 𝑖𝑢 ). For any 𝑅-module 𝑀, put
𝐾• (𝑢; 𝑀) ∶= 𝑀 ⊗ 𝐾• (𝑢),
𝑅
𝐾 • (𝑢; 𝑀) ∶= Hom𝐴 (𝐾• (𝑢), 𝑀)
⋅ 84 ⋅ Some aspects of Koszul complexes
which is naturally a chain (resp. cochain) complex in positive degrees; here one re-
gards 𝑀 as a complex in degree zero. These definitions generalize to the case of any
complex 𝑀, and are functorial in 𝑀.
The reader might have encountered the following result in differential geometry.
𝑛
Proposition 8.1.2 (Homotopy formula). For any 𝑥 ∈ 𝐿 and 𝜔 ∈ ⋀ 𝐿, we have
The terms with a 𝑥̂𝑖 (non-existant — hopefully this won’t generate metaphysical issues)
where 𝑖 > 0 cancel out. We are left with 𝑢(𝑥0 )𝑥1 ∧ ⋯ ∧ 𝑥𝑛 = 𝑢(𝑥)𝜔.
Obviously, the same formula extends to all 𝜔 ∈ ⋀ 𝐿 by linearity.
Proposition 8.1.3. Set 𝔮 ∶= 𝑢(𝐿), which is an ideal of 𝑅. Then 𝔮 annihilates each homology
(resp. cohomology) of 𝐾• (𝑢; 𝑀) (resp. 𝐾 • (𝑢; 𝑀)). Again, this generalizes to general complexes
𝑀.
Proof. Given 𝑡 ∈ 𝔮, the homotopy formula implies that the endomorphism 𝜔 ↦ 𝑡𝜔 of
𝐾• (𝑢) is homotopic to zero, hence so are the induced endomorphisms of 𝐾• (𝑢; 𝑀) and
𝐾 • (𝑢; 𝑀) by standard homological algebra.
Proposition 8.1.4. Suppose 𝐿 is projective over 𝑅 and 0 → 𝑀′ → 𝑀 → 𝑀″ → 0 is exact.
Then there is a natural short exact sequence of complexes
which gives rise to a long exact sequence of cohomologies of the Koszul complexes in question.
𝑛
Proof. Standard. It suffices to note that 𝐿 is projective implies each graded piece ⋀ 𝐿
of ⋀ 𝐿 is projective as well.
𝑛
Exercise 8.1.5. Justify the assertion above concerning the projectivity of ⋀ 𝐿. Hint:
𝑛
suppose 𝐿 is a direct summand of a free module 𝐹, show that ⋀ 𝐿 is a direct summand
𝑛 𝑛
of ⋀ 𝐹 and ⋀ 𝐹 is free.
Similar properties hold for the homological version when 𝐿 is flat over 𝑅. One needs
𝑛
the property that ⋀ 𝐿 is flat if 𝐿 is.
𝑛
Exercise 8.1.6. Prove the assertion above concerning flatness of ⋀ 𝐿. Consult the proof
in [4, p.15] if necessary.
§8.2 Auxiliary results on depth ⋅ 85 ⋅
⎛ ⎞
depth𝐼 ⎜
⎜∏ 𝑀𝛽 ⎟
⎟ = inf depth𝐼 (𝑀𝛽 ).
𝛽
⎝ 𝛽 ⎠
𝑛 𝑛
Proof. This follows from Ext𝑅 (𝑅/𝐼, ∏𝑖 𝑀𝑖 ) = ∏𝑖∈𝐼 Ext𝑅 (𝑅/𝐼, 𝑀𝑖 ), as is easily seen by
taking a projective resolution of 𝑅/𝐼 and using the fact the Hom𝑅 preserves direct prod-
ucts in the second variable.
Remark 8.2.2. By stipulation, the empty product is 0, the zero object of the category
𝑅-Mod. In parallel we define inf ∅ ∶= ∞, so that Proposition 8.2.1 remains true in this
case, since the zero module has infinite depth.
0 → 𝐾 → (𝑅/𝐼)⊕𝐼 → 𝑀 → 0,
<0
together with induction on 𝑖 (note that Ext𝑅 (𝑁, 𝑀) = 0 for trivial reasons). The case
of general 𝑚 follows by a standard dévissage using
0 → 𝐽𝑁 → 𝑁 → 𝑁/𝐽𝑁 → 0
Lemma 8.2.4. Suppose 𝐽 is an ideal satisfying 𝐽 ⊃ 𝐼 𝑚 for some 𝑚 ≥ 1. Then depth𝐼 (𝑀) ≤
depth𝐽 (𝑀).
𝑖
Proof. From the previous Proposition, we have Ext𝑅 (𝑅/𝐽, 𝑀) = 0 for 𝑖 < depth𝐼 (𝑀)
since 𝐼 𝑚 annihilates 𝑅/𝐽. The assertion follows upon recalling the definition of depth.
The following technical results will be invoked in the proof of Theorem 8.3.2.
In what follows, we write 𝐻 𝑛 = 𝐻 𝑛 (𝐶• ) = 𝑍𝑛 /𝐵𝑛 in the usual notation for homo-
logical algebra.
⋅ 86 ⋅ Some aspects of Koszul complexes
Proof. Assume on the contrary that there exists 𝑘 ≤ ℎ with 𝐻 <𝑘 = 0 whereas 𝐻 𝑘 ≠ 0.
Write 𝐽 = 𝐽𝑘 . As 𝐽 annihilates the nonzero 𝑅-module 𝐻 𝑘 , the criterion of depth-zero
modules (Proposition 7.5.2) implies that depth𝐽 (𝐻 𝑘 ) = 0. By assumption depth𝐽 (𝐶𝑘 ) >
𝑘 −𝑘 = 0, it follows that depth𝐽 (𝑍𝑘 ) > 0 since 𝑍𝑘 ⊂ 𝐶𝑘 , by applying the aforementioned
criterion of depth-zero. From the short exact sequence
0 → 𝐵𝑘 → 𝑍 𝑘 → 𝐻 𝑘 → 0
+1
ℋom(𝑅/𝐽, 𝐵𝑘 ) → ℋom(𝑅/𝐽, 𝑍𝑘 ) → ℋom(𝑅/𝐽, 𝐻 𝑘 ) ,
+1
ℋom(𝑅/𝐽, 𝐻 𝑘 )[−1] → ℋom(𝑅/𝐽, 𝐵𝑘 ) → ℋom(𝑅/𝐽, 𝑍𝑘 ) .
As the leftmost and rightmost terms of the last line are in 𝐷≥1 (𝑅-Mod), we see
depth𝐽 (𝐵𝑘 ) ≥ 1;
0 0 1
moreover, the piece Ext𝑅 (𝑅/𝐽, 𝑍𝑘 ) → Ext𝑅 (𝑅/𝐽, 𝐻 𝑘 ) → Ext𝑅 (𝑅/𝐽, 𝐵𝑘 ) from the long ex-
act sequence shows that depth𝐽 (𝐵𝑘 ) = 1. Now for 𝑛 < 𝑘 we have short exact sequences
0→ 𝐵
⎵ 𝑛 → 𝐶𝑛 → 𝐵𝑛+1 → 0.
=𝑍𝑛
+1
ℋom(𝑅/𝐽, 𝐵𝑛+1 )[−1] → ℋom(𝑅/𝐽, 𝐵𝑛 ) → ℋom(𝑅/𝐽, 𝐶𝑛 ) ,
the assumption depth𝐽 (𝐶𝑛 ) > 𝑘 − 𝑛 and descending induction on 𝑛 that 𝑛 < 𝑘 ⟹
depth𝐽 (𝐵𝑛 ) = 𝑘 − 𝑛 + 1. This is impossible since 𝐵≪0 = 0 has infinite depth.
𝑥
𝐾(𝑥) ∶= [𝑅 𝑅] .
§8.3 Koszul complexes and depth ⋅ 87 ⋅
More generally, for any 𝑅-module 𝑀, viewed as a complex concentrated in degree zero,
and a family x = (𝑥𝛼 )𝛼∈𝒜 of element of 𝑅, we define the associated Koszul complex as
with the well-known sign convention. Also note that 𝑅⊕𝒜 is projective, hence the
Proposition 8.1.4 can always be applied to Koszul cohomologies.
Let 𝐼 denote the ideal generated by {𝑥𝛼 ∶ 𝛼 ∈ 𝒜}. The reader is invited to verify that
⋄ 𝐾 • (x, ∏𝛽∈ℬ 𝑀𝛽 ) = ∏𝛽∈ℬ 𝐾 • (x; 𝑀𝛽 ) for any family {𝑀𝛽 }𝛽∈ℬ of 𝑅-modules,
and same for their cohomologies;
Theorem 8.3.2. Let 𝑀 be an 𝑅-module and x = {𝑥𝛼 }𝛼∈𝒜 be a family of elements of 𝑅, which
generate an ideal 𝐼 of 𝐴. Then depth𝐼 (𝑀) equals
depth𝐼 (𝑀) ≤ 𝑛.
⋅ 88 ⋅ Some aspects of Koszul complexes
Proof. Let 𝑑 ∶= depth𝐼 (𝑀). Since 𝐾 ℎ (x; 𝑀) is a direct product of copies of 𝑀 (possibly
the empty product = 0), by Proposition 8.2.1 its depth equals either 𝑑 or ∞. Combining
Lemma 8.3.1 and Corollary 8.2.6 (with ℎ = 𝑑 − 1), we see that 𝐻 <𝑑 (𝐾 • (x; 𝑀)) = 0. It
remains to show 𝐻 𝑑 (𝐾 • (x; 𝑀)) ≠ 0 provided that 𝑑 < ∞, which we assume from now
onwards.
The case 𝑑 = 0 is clear since there exists 𝔭 ∈ Ass(𝑀) ∩ 𝑉(𝐼), therefore 𝑅/𝔭 ↪ 𝑀
and ∃𝑥 ∈ 𝑀 with 𝔭𝑥 ⊃ 𝐼𝑥 = {0}, whence 𝐻 0 (𝐾 • (x; 𝑀)) ≠ 0. Now suppose 𝑑 ∈ ℤ≥1
and assume 𝐻 𝑑 (𝐾 • (x; 𝑀)) = 0. Take a free resolution 𝐹• → 𝑅/𝐼 → 0 and put 𝐶• ∶=
Hom𝑅 (𝐹• , 𝑀), so that
𝑖
𝐻 𝑖 (𝐶• ) ≃ Ext𝑅 (𝑅/𝐼, 𝑀), 𝑖 ≥ 0.
Hence for 𝑖 < 𝑑 we have short exact sequences
0 → 𝐵𝑖 → 𝐶𝑖 → 𝐵𝑖+1 → 0
𝐻 0 (𝐾 • (x; 𝑀)) ≠ 0.
However, this contradicts the earlier result that 𝐻 <𝑑 (𝐾 • (x; 𝑀)) = 0 since 𝑑 ≥ 1.
This yields an alternative characterization of depth. It also completes the proof of
Theorem 7.5.5 as promised in the previous lecture.
§8.3 Koszul complexes and depth ⋅ 89 ⋅
[10] Robin Hartshorne. Algebraic geometry. Graduate Texts in Mathematics, No. 52.
Springer-Verlag, New York-Heidelberg, 1977, pp. xvi+496. isbn: 0-387-90244-9
(cit. on p. 44).
[11] Hideyuki Matsumura. Commutative algebra. Second Edition. Vol. 56. Mathematics
Lecture Note Series. Benjamin/Cummings Publishing Co., Inc., Reading, Mass.,
1980, pp. xv+313. isbn: 0-8053-7026-9 (cit. on pp. 2, 15, 19, 25, 44, 51, 65–67, 71).
[12] John Milnor. Singular points of complex hypersurfaces. Annals of Mathematics Stud-
ies, No. 61. Princeton University Press, Princeton, N.J.; University of Tokyo Press,
Tokyo, 1968, pp. iii+122 (cit. on p. 27).
[13] David Mumford. The red book of varieties and schemes. expanded. Vol. 1358. Lecture
Notes in Mathematics. Includes the Michigan lectures (1974) on curves and their
Jacobians, With contributions by Enrico Arbarello. Springer-Verlag, Berlin, 1999,
pp. x+306. isbn: 3-540-63293-X. doi: 10.1007/b62130. url: http://dx.doi.org/
10.1007/b62130 (cit. on p. 28).
[14] “Obituary: Tadasi Nakayama”. In: Nagoya Math. J. 27.1 (1966), pp. i–vii. url:
http://projecteuclid.org/euclid.nmj/1118801606 (cit. on p. 10).
[15] Jean-Pierre Serre. “Géométrie algébrique et géométrie analytique”. In: Ann. Inst.
Fourier, Grenoble 6 (1955–1956), pp. 1–42. issn: 0373-0956 (cit. on p. 38).
[16] I. R. Shafarevich, ed. Algebraic geometry. IV. Vol. 55. Encyclopaedia of Mathe-
matical Sciences. Linear algebraic groups. Invariant theory, A translation of ıt
Algebraic geometry. 4 (Russian), Akad. Nauk SSSR Vsesoyuz. Inst. Nauchn. i
Tekhn. Inform., Moscow, 1989 [ MR1100483 (91k:14001)], Translation edited by
A. N. Parshin and I. R. Shafarevich. Springer-Verlag, Berlin, 1994, pp. vi+284.
isbn: 3-540-54682-0. doi: 10.1007/978- 3- 662- 03073- 8. url: http://dx.doi.
org/10.1007/978-3-662-03073-8 (cit. on p. 13).
[17] Angelo Vistoli. “Grothendieck topologies, fibered categories and descent the-
ory”. In: Fundamental algebraic geometry. Vol. 123. Math. Surveys Monogr. Amer.
Math. Soc., Providence, RI, 2005, pp. 1–104 (cit. on p. 38).
Index
A going-up, 39
ann𝑅 (𝑀), ann𝑅 (𝑥), 15 gr𝐹 (𝑀), 46
Artinian, 11 graded, 45
Ass(𝑀), 16
associated prime, 16 H
Hauptidealsatz, 63
B height, 58
blow-up algebra, 48 Hilbert’s basis theorem, 11
C Hilbert–Samuel polynomial, 58
catenary, 71 homogeneous, 45
Cayley–Hamilton theorem, 10
Chevalley’s theorem, 44 I
Cohen–Macaulay module, 80 integral closure, 26
completion, 52 integrality, 25
constructible subset, 43 invariant theory, 13
coprimary module, 18
J
D Jacobson radical, 10
depth, 78, 87 Jacobson ring, 28
dimension, 58
dimension formula, 70 K
discrete valuation ring, 73 𝐾 • (𝑢; 𝑀), 84
𝐾 • (x; 𝑀), 87
E Koszul complex, 83, 87
embedded prime, 17 Krull’s intersection theorem, 50
exterios algebra, 83
F L
faithful functor, 36 length, 11, 57
faithfully flat, 31, 36 local ring, 3
field of fractions, 6 localization, 5
filtration, 46
M
𝔞-adic, 47
𝑀[𝑆−1 ], 5
finitely presented, 7
MaxSpec(𝑅), 4
flat, 31
minimal prime ideal, 39
G Moishezon manifolds, 71
going-down, 39 multiplicative subset, 5
⋅ 94 ⋅ INDEX
N
Nakayama’s lemma, 10
nilpotent radical, 9
Noether normalization, 69
Noetherian, 11
normal, 26
Nullstellensatz, 13, 29
P
parameter ideal, 59
primary decomposition, 19
primary module, 18
prime avoidance, 4
R
reduced ring, 9, 74
Rees algebra, 49
regular local ring, 63
regular sequence, 79
S
SageMath, 23
Serre’s criterion, 77
Spec(𝑅), 4
Supp(𝑀), 15
support, 15
T
topological ring, 51
𝐼-adic, 51
total fraction ring, 6
U
uniformizer, 73
V
𝑉(𝔞), 4
variety, 12
Z
Zariski topology, 4
zero divisor, 16
Zorn’s Lemma, 3, 20, 40