The Stacks project

15.91 Derived Completion

Some references for the material in this section are [Dwyer-Greenlees], [Greenlees-May], [PSY], [dag12] (especially Chapter 4). Our exposition follows [BS]. The analogue (or “dual”) of this section for torsion modules is Dualizing Complexes, Section 47.9. The relationship between the derived category of complexes with torsion cohomology and derived complete complexes can be found in Dualizing Complexes, Section 47.12.

Let $K \in D(A)$. Let $f \in A$. We denote $T(K, f)$ a derived limit of the system

\[ \ldots \to K \xrightarrow {f} K \xrightarrow {f} K \]

in $D(A)$.

Lemma 15.91.1. Let $A$ be a ring. Let $f \in A$. Let $K \in D(A)$. The following are equivalent

  1. $\mathop{\mathrm{Ext}}\nolimits ^ n_ A(A_ f, K) = 0$ for all $n$,

  2. $\mathop{\mathrm{Hom}}\nolimits _{D(A)}(E, K) = 0$ for all $E$ in $D(A_ f)$,

  3. $T(K, f) = 0$,

  4. for every $p \in \mathbf{Z}$ we have $T(H^ p(K), f) = 0$,

  5. for every $p \in \mathbf{Z}$ we have $\mathop{\mathrm{Hom}}\nolimits _ A(A_ f, H^ p(K)) = 0$ and $\mathop{\mathrm{Ext}}\nolimits ^1_ A(A_ f, H^ p(K)) = 0$,

  6. $R\mathop{\mathrm{Hom}}\nolimits _ A(A_ f, K) = 0$,

  7. the map $\prod _{n \geq 0} K \to \prod _{n \geq 0} K$, $(x_0, x_1, \ldots ) \mapsto (x_0 - fx_1, x_1 - fx_2, \ldots )$ is an isomorphism in $D(A)$, and

  8. add more here.

Proof. It is clear that (2) implies (1) and that (1) is equivalent to (6). Assume (1). Let $I^\bullet $ be a K-injective complex of $A$-modules representing $K$. Condition (1) signifies that $\mathop{\mathrm{Hom}}\nolimits _ A(A_ f, I^\bullet )$ is acyclic. Let $M^\bullet $ be a complex of $A_ f$-modules representing $E$. Then

\[ \mathop{\mathrm{Hom}}\nolimits _{D(A)}(E, K) = \mathop{\mathrm{Hom}}\nolimits _{K(A)}(M^\bullet , I^\bullet ) = \mathop{\mathrm{Hom}}\nolimits _{K(A_ f)}(M^\bullet , \mathop{\mathrm{Hom}}\nolimits _ A(A_ f, I^\bullet )) \]

by Algebra, Lemma 10.14.4. As $\mathop{\mathrm{Hom}}\nolimits _ A(A_ f, I^\bullet )$ is a K-injective complex of $A_ f$-modules by Lemma 15.56.3 the fact that it is acyclic implies that it is homotopy equivalent to zero (Derived Categories, Lemma 13.31.2). Thus we get (2).

A free resolution of the $A$-module $A_ f$ is given by

\[ 0 \to \bigoplus \nolimits _{n \in \mathbf{N}} A \to \bigoplus \nolimits _{n \in \mathbf{N}} A \to A_ f \to 0 \]

where the first map sends the $(a_0, a_1, a_2, \ldots )$ to $(a_0, a_1 - fa_0, a_2 - fa_1, \ldots )$ and the second map sends $(a_0, a_1, a_2, \ldots )$ to $a_0 + a_1/f + a_2/f^2 + \ldots $. Applying $\mathop{\mathrm{Hom}}\nolimits _ A(-, I^\bullet )$ we get

\[ 0 \to \mathop{\mathrm{Hom}}\nolimits _ A(A_ f, I^\bullet ) \to \prod I^\bullet \to \prod I^\bullet \to 0 \]

Since $\prod I^\bullet $ represents $\prod _{n \geq 0} K$ this proves the equivalence of (1) and (7). On the other hand, by construction of derived limits in Derived Categories, Section 13.34 the displayed exact sequence shows the object $T(K, f)$ is a representative of $R\mathop{\mathrm{Hom}}\nolimits _ A(A_ f, K)$ in $D(A)$. Thus the equivalence of (1) and (3).

There is a spectral sequence

\[ E_2^{p, q} = \mathop{\mathrm{Ext}}\nolimits ^ p_ A(A_ f, H^ q(K)) \Rightarrow \mathop{\mathrm{Ext}}\nolimits ^{p + q}_ A(A_ f, K) \]

See Equation (15.67.0.1). This spectral sequence degenerates at $E_2$ because $A_ f$ has a length $1$ resolution by projective $A$-modules (see above) hence the $E_2$-page has only 2 nonzero columns. Thus we obtain short exact sequences

\[ 0 \to \mathop{\mathrm{Ext}}\nolimits ^1_ A(A_ f, H^{p - 1}(K)) \to \mathop{\mathrm{Ext}}\nolimits ^ p_ A(A_ f, K) \to \mathop{\mathrm{Hom}}\nolimits _ A(A_ f, H^ p(K)) \to 0 \]

This proves (4) and (5) are equivalent to (1). $\square$

Lemma 15.91.2. Let $A$ be a ring. Let $K \in D(A)$. The set $I$ of $f \in A$ such that $T(K, f) = 0$ is a radical ideal of $A$.

Proof. We will use the results of Lemma 15.91.1 without further mention. If $f \in I$, and $g \in A$, then $A_{gf}$ is an $A_ f$-module hence $\mathop{\mathrm{Ext}}\nolimits ^ n_ A(A_{gf}, K) = 0$ for all $n$, hence $gf \in I$. Suppose $f, g \in I$. Then there is a short exact sequence

\[ 0 \to A_{f + g} \to A_{f(f + g)} \oplus A_{g(f + g)} \to A_{gf(f + g)} \to 0 \]

because $f, g$ generate the unit ideal in $A_{f + g}$. This follows from Algebra, Lemma 10.24.2 and the easy fact that the last arrow is surjective. From the long exact sequence of $\mathop{\mathrm{Ext}}\nolimits $ and the vanishing of $\mathop{\mathrm{Ext}}\nolimits ^ n_ A(A_{f(f + g)}, K)$, $\mathop{\mathrm{Ext}}\nolimits ^ n_ A(A_{g(f + g)}, K)$, and $\mathop{\mathrm{Ext}}\nolimits ^ n_ A(A_{gf(f + g)}, K)$ for all $n$ we deduce the vanishing of $\mathop{\mathrm{Ext}}\nolimits ^ n_ A(A_{f + g}, K)$ for all $n$. Finally, if $f^ n \in I$ for some $n > 0$, then $f \in I$ because $T(K, f) = T(K, f^ n)$ or because $A_ f \cong A_{f^ n}$. $\square$

Lemma 15.91.3. Let $A$ be a ring. Let $I \subset A$ be an ideal. Let $M$ be an $A$-module.

  1. If $M$ is $I$-adically complete, then $T(M, f) = 0$ for all $f \in I$.

  2. Conversely, if $T(M, f) = 0$ for all $f \in I$ and $I$ is finitely generated, then $M \to \mathop{\mathrm{lim}}\nolimits M/I^ nM$ is surjective.

Proof. Proof of (1). Assume $M$ is $I$-adically complete. By Lemma 15.91.1 it suffices to prove $\mathop{\mathrm{Ext}}\nolimits ^1_ A(A_ f, M) = 0$ and $\mathop{\mathrm{Hom}}\nolimits _ A(A_ f, M) = 0$. Since $M = \mathop{\mathrm{lim}}\nolimits M/I^ nM$ and since $\mathop{\mathrm{Hom}}\nolimits _ A(A_ f, M/I^ nM) = 0$ it follows that $\mathop{\mathrm{Hom}}\nolimits _ A(A_ f, M) = 0$. Suppose we have an extension

\[ 0 \to M \to E \to A_ f \to 0 \]

For $n \geq 0$ pick $e_ n \in E$ mapping to $1/f^ n$. Set $\delta _ n = fe_{n + 1} - e_ n \in M$ for $n \geq 0$. Replace $e_ n$ by

\[ e'_ n = e_ n + \delta _ n + f\delta _{n + 1} + f^2 \delta _{n + 2} + \ldots \]

The infinite sum exists as $M$ is complete with respect to $I$ and $f \in I$. A simple calculation shows that $fe'_{n + 1} = e'_ n$. Thus we get a splitting of the extension by mapping $1/f^ n$ to $e'_ n$.

Proof of (2). Assume that $I = (f_1, \ldots , f_ r)$ and that $T(M, f_ i) = 0$ for $i = 1, \ldots , r$. By Algebra, Lemma 10.96.7 we may assume $I = (f)$ and $T(M, f) = 0$. Let $x_ n \in M$ for $n \geq 0$. Consider the extension

\[ 0 \to M \to E \to A_ f \to 0 \]

given by

\[ E = M \oplus \bigoplus Ae_ n\Big/\langle x_ n - fe_{n + 1} + e_ n\rangle \]

mapping $e_ n$ to $1/f^ n$ in $A_ f$ (see above). By assumption and Lemma 15.91.1 this extension is split, hence we obtain an element $x + e_0$ which generates a copy of $A_ f$ in $E$. Then

\[ x + e_0 = x - x_0 + fe_1 = x - x_0 - f x_1 + f^2 e_2 = \ldots \]

Since $M/f^ nM = E/f^ nE$ by the snake lemma, we see that $x = x_0 + fx_1 + \ldots + f^{n - 1}x_{n - 1}$ modulo $f^ nM$. In other words, the map $M \to \mathop{\mathrm{lim}}\nolimits M/f^ nM$ is surjective as desired. $\square$

Motivated by the results above we make the following definition.

Definition 15.91.4. Let $A$ be a ring. Let $K \in D(A)$. Let $I \subset A$ be an ideal. We say $K$ is derived complete with respect to $I$ if for every $f \in I$ we have $T(K, f) = 0$. If $M$ is an $A$-module, then we say $M$ is derived complete with respect to $I$ if $M[0] \in D(A)$ is derived complete with respect to $I$.

The full subcategory $D_{comp}(A) = D_{comp}(A, I) \subset D(A)$ consisting of derived complete objects is a strictly full, saturated triangulated subcategory, see Derived Categories, Definitions 13.3.4 and 13.6.1. By Lemma 15.91.2 the subcategory $D_{comp}(A, I)$ depends only on the radical $\sqrt{I}$ of $I$, in other words it depends only on the closed subset $Z = V(I)$ of $\mathop{\mathrm{Spec}}(A)$. The subcategory $D_{comp}(A, I)$ is preserved under products and homotopy limits in $D(A)$. But it is not preserved under countable direct sums in general. We will often simply say $M$ is a derived complete module if the choice of the ideal $I$ is clear from the context.

Proposition 15.91.5. Let $I \subset A$ be a finitely generated ideal of a ring $A$. Let $M$ be an $A$-module. The following are equivalent

  1. $M$ is $I$-adically complete, and

  2. $M$ is derived complete with respect to $I$ and $\bigcap I^ nM = 0$.

Proof. This is clear from the results of Lemma 15.91.3. $\square$

The next lemma shows that the category $\mathcal{C}$ of derived complete modules is abelian. It turns out that $\mathcal{C}$ is not a Grothendieck abelian category, see Examples, Section 110.11.

Lemma 15.91.6. Let $I$ be an ideal of a ring $A$.

  1. The derived complete $A$-modules form a weak Serre subcategory $\mathcal{C}$ of $\text{Mod}_ A$.

  2. $D_\mathcal {C}(A) \subset D(A)$ is the full subcategory of derived complete objects.

Proof. Part (2) is immediate from Lemma 15.91.1 and the definitions. For part (1), suppose that $M \to N$ is a map of derived complete modules. Denote $K = (M \to N)$ the corresponding object of $D(A)$. Pick $f \in I$. Then $\mathop{\mathrm{Ext}}\nolimits _ A^ n(A_ f, K)$ is zero for all $n$ because $\mathop{\mathrm{Ext}}\nolimits _ A^ n(A_ f, M)$ and $\mathop{\mathrm{Ext}}\nolimits _ A^ n(A_ f, N)$ are zero for all $n$. Hence $K$ is derived complete. By (2) we see that $\mathop{\mathrm{Ker}}(M \to N)$ and $\mathop{\mathrm{Coker}}(M \to N)$ are objects of $\mathcal{C}$. Finally, suppose that $0 \to M_1 \to M_2 \to M_3 \to 0$ is a short exact sequence of $A$-modules and $M_1$, $M_3$ are derived complete. Then it follows from the long exact sequence of $\mathop{\mathrm{Ext}}\nolimits $'s that $M_2$ is derived complete. Thus $\mathcal{C}$ is a weak Serre subcategory by Homology, Lemma 12.10.3. $\square$

We will generalize the following lemma in Lemma 15.91.19.

Lemma 15.91.7. Let $I$ be a finitely generated ideal of a ring $A$. Let $M$ be a derived complete $A$-module. If $M/IM = 0$, then $M = 0$.

Proof. Assume that $M/IM$ is zero. Let $I = (f_1, \ldots , f_ r)$. Let $i < r$ be the largest integer such that $N = M/(f_1, \ldots , f_ i)M$ is nonzero. If $i$ does not exist, then $M = 0$ which is what we want to show. Then $N$ is derived complete as a cokernel of a map between derived complete modules, see Lemma 15.91.6. By our choice of $i$ we have that $f_{i + 1} : N \to N$ is surjective. Hence

\[ \mathop{\mathrm{lim}}\nolimits (\ldots \to N \xrightarrow {f_{i + 1}} N \xrightarrow {f_{i + 1}} N) \]

is nonzero, contradicting the derived completeness of $N$. $\square$

If the ring is $I$-adically complete, then one obtains an ample supply of derived complete complexes.

Lemma 15.91.8. Let $A$ be a ring and $I \subset A$ an ideal. If $A$ is derived complete (eg. $I$-adically complete) then any pseudo-coherent object of $D(A)$ is derived complete.

Proof. (Lemma 15.91.3 explains the parenthetical statement of the lemma.) Let $K$ be a pseudo-coherent object of $D(A)$. By definition this means $K$ is represented by a bounded above complex $K^\bullet $ of finite free $A$-modules. Since $A$ is derived complete it follows that $H^ n(K)$ is derived complete for all $n$, by part (1) of Lemma 15.91.6. This in turn implies that $K$ is derived complete by part (2) of the same lemma. $\square$

Lemma 15.91.9. Let $A$ be a ring. Let $f, g \in A$. Then for $K \in D(A)$ we have $R\mathop{\mathrm{Hom}}\nolimits _ A(A_ f, R\mathop{\mathrm{Hom}}\nolimits _ A(A_ g, K)) = R\mathop{\mathrm{Hom}}\nolimits _ A(A_{fg}, K)$.

Proof. This follows from Lemma 15.73.1. $\square$

slogan

Lemma 15.91.10. Let $I$ be a finitely generated ideal of a ring $A$. The inclusion functor $D_{comp}(A, I) \to D(A)$ has a left adjoint, i.e., given any object $K$ of $D(A)$ there exists a map $K \to K^\wedge $ of $K$ into a derived complete object of $D(A)$ such that the map

\[ \mathop{\mathrm{Hom}}\nolimits _{D(A)}(K^\wedge , E) \longrightarrow \mathop{\mathrm{Hom}}\nolimits _{D(A)}(K, E) \]

is bijective whenever $E$ is a derived complete object of $D(A)$. In fact, if $I$ is generated by $f_1, \ldots , f_ r \in A$, then we have

\[ K^\wedge = R\mathop{\mathrm{Hom}}\nolimits \left((A \to \prod \nolimits _{i_0} A_{f_{i_0}} \to \prod \nolimits _{i_0 < i_1} A_{f_{i_0}f_{i_1}} \to \ldots \to A_{f_1\ldots f_ r}), K\right) \]

functorially in $K$.

Proof. Define $K^\wedge $ by the last displayed formula of the lemma. There is a map of complexes

\[ (A \to \prod \nolimits _{i_0} A_{f_{i_0}} \to \prod \nolimits _{i_0 < i_1} A_{f_{i_0}f_{i_1}} \to \ldots \to A_{f_1\ldots f_ r}) \longrightarrow A \]

which induces a map $K \to K^\wedge $. It suffices to prove that $K^\wedge $ is derived complete and that $K \to K^\wedge $ is an isomorphism if $K$ is derived complete.

Let $f \in A$. By Lemma 15.91.9 the object $R\mathop{\mathrm{Hom}}\nolimits _ A(A_ f, K^\wedge )$ is equal to

\[ R\mathop{\mathrm{Hom}}\nolimits \left((A_ f \to \prod \nolimits _{i_0} A_{ff_{i_0}} \to \prod \nolimits _{i_0 < i_1} A_{ff_{i_0}f_{i_1}} \to \ldots \to A_{ff_1\ldots f_ r}), K\right) \]

If $f \in I$, then $f_1, \ldots , f_ r$ generate the unit ideal in $A_ f$, hence the extended alternating Čech complex

\[ A_ f \to \prod \nolimits _{i_0} A_{ff_{i_0}} \to \prod \nolimits _{i_0 < i_1} A_{ff_{i_0}f_{i_1}} \to \ldots \to A_{ff_1\ldots f_ r} \]

is zero in $D(A)$ by Lemma 15.29.5. (In fact, if $f = f_ i$ for some $i$, then this complex is homotopic to zero by Lemma 15.29.4; this is the only case we need.) Hence $R\mathop{\mathrm{Hom}}\nolimits _ A(A_ f, K^\wedge ) = 0$ and we conclude that $K^\wedge $ is derived complete by Lemma 15.91.1.

Conversely, if $K$ is derived complete, then $R\mathop{\mathrm{Hom}}\nolimits _ A(A_ f, K)$ is zero for all $f = f_{i_0} \ldots f_{i_ p}$, $p \geq 0$. Thus $K \to K^\wedge $ is an isomorphism in $D(A)$. $\square$

Remark 15.91.11. Let $A$ be a ring and let $I \subset A$ be a finitely generated ideal. The left adjoint to the inclusion functor $D_{comp}(A, I) \to D(A)$ which exists by Lemma 15.91.10 is called the derived completion. To indicate this we will say “let $K^\wedge $ be the derived completion of $K$”. Please keep in mind that the unit of the adjunction is a functorial map $K \to K^\wedge $.

Lemma 15.91.12. Let $A$ be a ring and let $I \subset A$ be a finitely generated ideal. Let $K^\bullet $ be a complex of $A$-modules such that $f : K^\bullet \to K^\bullet $ is an isomorphism for some $f \in I$, i.e., $K^\bullet $ is a complex of $A_ f$-modules. Then the derived completion of $K^\bullet $ is zero.

Proof. Indeed, in this case the $R\mathop{\mathrm{Hom}}\nolimits _ A(K, L)$ is zero for any derived complete complex $L$, see Lemma 15.91.1. Hence $K^\wedge $ is zero by the universal property in Lemma 15.91.10. $\square$

Lemma 15.91.13. Let $A$ be a ring and let $I \subset A$ be a finitely generated ideal. Let $K, L \in D(A)$. Then

\[ R\mathop{\mathrm{Hom}}\nolimits _ A(K, L)^\wedge = R\mathop{\mathrm{Hom}}\nolimits _ A(K, L^\wedge ) = R\mathop{\mathrm{Hom}}\nolimits _ A(K^\wedge , L^\wedge ) \]

Proof. By Lemma 15.91.10 we know that derived completion is given by $R\mathop{\mathrm{Hom}}\nolimits _ A(C, -)$ for some $C \in D(A)$. Then

\begin{align*} R\mathop{\mathrm{Hom}}\nolimits _ A(C, R\mathop{\mathrm{Hom}}\nolimits _ A(K, L)) & = R\mathop{\mathrm{Hom}}\nolimits _ A(C \otimes _ A^\mathbf {L} K, L) \\ & = R\mathop{\mathrm{Hom}}\nolimits _ A(K, R\mathop{\mathrm{Hom}}\nolimits _ A(C, L)) \end{align*}

by Lemma 15.73.1. This proves the first equation. The map $K \to K^\wedge $ induces a map

\[ R\mathop{\mathrm{Hom}}\nolimits _ A(K^\wedge , L^\wedge ) \to R\mathop{\mathrm{Hom}}\nolimits _ A(K, L^\wedge ) \]

which is an isomorphism in $D(A)$ by definition of the derived completion as the left adjoint to the inclusion functor. $\square$

Lemma 15.91.14. Let $A$ be a ring and let $I \subset A$ be an ideal. Let $(K_ n)$ be an inverse system of objects of $D(A)$ such that for all $f \in I$ and $n$ there exists an $e = e(n, f)$ such that $f^ e$ is zero on $K_ n$. Then for $K \in D(A)$ the object $K' = R\mathop{\mathrm{lim}}\nolimits (K \otimes _ A^\mathbf {L} K_ n)$ is derived complete with respect to $I$.

Proof. Since the category of derived complete objects is preserved under $R\mathop{\mathrm{lim}}\nolimits $ it suffices to show that each $K \otimes _ A^\mathbf {L} K_ n$ is derived complete. By assumption for all $f \in I$ there is an $e$ such that $f^ e$ is zero on $K \otimes _ A^\mathbf {L} K_ n$. Of course this implies that $T(K \otimes _ A^\mathbf {L} K_ n, f) = 0$ and we win. $\square$

Situation 15.91.15. Let $A$ be a ring. Let $I = (f_1, \ldots , f_ r) \subset A$. Let $K_ n^\bullet = K_\bullet (A, f_1^ n, \ldots , f_ r^ n)$ be the Koszul complex on $f_1^ n, \ldots , f_ r^ n$ viewed as a cochain complex in degrees $-r, -r + 1, \ldots , 0$. Using the functoriality of Lemma 15.28.3 we obtain an inverse system

\[ \ldots \to K_3^\bullet \to K_2^\bullet \to K_1^\bullet \]

compatible with the inverse system $H^0(K_ n^\bullet ) = A/(f_1^ n, \ldots , f_ r^ n)$ and compatible with the maps $A \to K_ n^\bullet $.

A key feature of the discussion below will use that for $m > n$ the map

\[ K_ m^{-p} = \wedge ^ p(A^{\oplus r}) \to \wedge ^ p(A^{\oplus r}) = K_ n^{-p} \]

is given by multiplication by $f_{i_1}^{m - n} \ldots f_{i_ p}^{m - n}$ on the basis element $e_{i_1} \wedge \ldots \wedge e_{i_ p}$.

Lemma 15.91.16. In Situation 15.91.15. For $K \in D(A)$ the object $K' = R\mathop{\mathrm{lim}}\nolimits (K \otimes _ A^\mathbf {L} K_ n^\bullet )$ is derived complete with respect to $I$.

Proof. This is a special case of Lemma 15.91.14 because $f_ i^ n$ acts by an endomorphism of $K_ n^\bullet $ which is homotopic to zero by Lemma 15.28.6. $\square$

Lemma 15.91.17. In Situation 15.91.15. Let $K \in D(A)$. The following are equivalent

  1. $K$ is derived complete with respect to $I$, and

  2. the canonical map $K \to R\mathop{\mathrm{lim}}\nolimits (K \otimes _ A^\mathbf {L} K_ n^\bullet )$ is an isomorphism of $D(A)$.

Proof. If (2) holds, then $K$ is derived complete with respect to $I$ by Lemma 15.91.16. Conversely, assume that $K$ is derived complete with respect to $I$. Consider the filtrations

\[ K_ n^\bullet \supset \sigma _{\geq -r + 1}K_ n^\bullet \supset \sigma _{\geq -r + 2}K_ n^\bullet \supset \ldots \supset \sigma _{\geq -1}K_ n^\bullet \supset \sigma _{\geq 0}K_ n^\bullet = A \]

by stupid truncations (Homology, Section 12.15). Because the construction $R\mathop{\mathrm{lim}}\nolimits (K \otimes E)$ is exact in the second variable (Lemma 15.87.11) we see that it suffices to show

\[ R\mathop{\mathrm{lim}}\nolimits \left( K \otimes _ A^\mathbf {L} (\sigma _{\geq p}K_ n^\bullet / \sigma _{\geq p + 1}K_ n^\bullet ) \right) = 0 \]

for $p < 0$. The explicit description of the Koszul complexes above shows that

\[ R\mathop{\mathrm{lim}}\nolimits \left( K \otimes _ A^\mathbf {L} (\sigma _{\geq p}K_ n^\bullet / \sigma _{\geq p + 1}K_ n^\bullet ) \right) = \bigoplus \nolimits _{i_1, \ldots , i_{-p}} T(K, f_{i_1}\ldots f_{i_{-p}}) \]

which is zero for $p < 0$ by assumption on $K$. $\square$

Lemma 15.91.18. In Situation 15.91.15. The functor which sends $K \in D(A)$ to the derived limit $K' = R\mathop{\mathrm{lim}}\nolimits ( K \otimes _ A^\mathbf {L} K_ n^\bullet )$ is the left adjoint to the inclusion functor $D_{comp}(A) \to D(A)$ constructed in Lemma 15.91.10.

First proof. The assignment $K \leadsto K'$ is a functor and $K'$ is derived complete with respect to $I$ by Lemma 15.91.16. By a formal argument (omitted) we see that it suffices to show $K \to K'$ is an isomorphism if $K$ is derived complete with respect to $I$. This is Lemma 15.91.17. $\square$

Second proof. Denote $K \mapsto K^\wedge $ the adjoint constructed in Lemma 15.91.10. By that lemma we have

\[ K^\wedge = R\mathop{\mathrm{Hom}}\nolimits \left((A \to \prod \nolimits _{i_0} A_{f_{i_0}} \to \prod \nolimits _{i_0 < i_1} A_{f_{i_0}f_{i_1}} \to \ldots \to A_{f_1\ldots f_ r}), K\right) \]

In Lemma 15.29.6 we have seen that the extended alternating Čech complex

\[ A \to \prod \nolimits _{i_0} A_{f_{i_0}} \to \prod \nolimits _{i_0 < i_1} A_{f_{i_0}f_{i_1}} \to \ldots \to A_{f_1\ldots f_ r} \]

is a colimit of the Koszul complexes $K^ n = K(A, f_1^ n, \ldots , f_ r^ n)$ sitting in degrees $0, \ldots , r$. Note that $K^ n$ is a finite chain complex of finite free $A$-modules with dual (as in Lemma 15.74.15) $R\mathop{\mathrm{Hom}}\nolimits _ A(K^ n, A) = K_ n$ where $K_ n$ is the Koszul cochain complex sitting in degrees $-r, \ldots , 0$ (as usual). Thus it suffices to show that

\[ R\mathop{\mathrm{Hom}}\nolimits _ A(\text{hocolim} K^ n, K) = R\mathop{\mathrm{lim}}\nolimits (K \otimes _ A^\mathbf {L} K_ n) \]

This follows from Lemma 15.74.16. $\square$

sloganreference

Lemma 15.91.19. Let $I$ be a finitely generated ideal of a ring $A$. Let $K$ be a derived complete object of $D(A)$. If $K \otimes _ A^\mathbf {L} A/I = 0$, then $K = 0$.

Proof. Choose generators $f_1, \ldots , f_ r$ of $I$. Denote $K_ n$ the Koszul complex on $f_1^ n, \ldots , f_ r^ n$ over $A$. Recall that $K_ n$ is bounded and that the cohomology modules of $K_ n$ are annihilated by $f_1^ n, \ldots , f_ r^ n$ and hence by $I^{nr}$. By Lemma 15.88.7 we see that $K \otimes _ A^\mathbf {L} K_ n = 0$. Since $K$ is derived complete by Lemma 15.91.18 we have $K = R\mathop{\mathrm{lim}}\nolimits K \otimes _ A^\mathbf {L} K_ n = 0$ as desired. $\square$

As an application of the relationship with the Koszul complex we obtain that derived completion has finite cohomological dimension.

Lemma 15.91.20. Let $A$ be a ring and let $I \subset A$ be an ideal which can be generated by $r$ elements. Then derived completion has finite cohomological dimension:

  1. Let $K \to L$ be a morphism in $D(A)$ such that $H^ i(K) \to H^ i(L)$ is an isomorphism for $i \geq 1$ and surjective for $i = 0$. Then $H^ i(K^\wedge ) \to H^ i(L^\wedge )$ is an isomorphism for $i \geq 1$ and surjective for $i = 0$.

  2. Let $K \to L$ be a morphism of $D(A)$ such that $H^ i(K) \to H^ i(L)$ is an isomorphism for $i \leq -1$ and injective for $i = 0$. Then $H^ i(K^\wedge ) \to H^ i(L^\wedge )$ is an isomorphism for $i \leq -r - 1$ and injective for $i = -r$.

Proof. Say $I$ is generated by $f_1, \ldots , f_ r$. For any $K \in D(A)$ by Lemma 15.91.18 we have $K^\wedge = R\mathop{\mathrm{lim}}\nolimits K \otimes _ A^\mathbf {L} K_ n$ where $K_ n$ is the Koszul complex on $f_1^ n, \ldots , f_ r^ n$ and hence we obtain a short exact sequence

\[ 0 \to R^1\mathop{\mathrm{lim}}\nolimits H^{i - 1}(K \otimes _ A^\mathbf {L} K_ n) \to H^ i(K^\wedge ) \to \mathop{\mathrm{lim}}\nolimits H^ i(K \otimes _ A^\mathbf {L} K_ n) \to 0 \]

by Lemma 15.87.4.

Proof of (1). Pick a distinguished triangle $K \to L \to C \to K[1]$. Then $H^ i(C) = 0$ for $i \geq 0$. Since $K_ n$ is sitting in degrees $\leq 0$ we see that $H^ i(C \otimes _ A^\mathbf {L} K_ n) = 0$ for $i \geq 0$ and that $H^{-1}(C \otimes _ A^\mathbf {L} K_ n) = H^{-1}(C) \otimes _ A A/(f_1^ n, \ldots , f_ r^ n)$ is a system with surjective transition maps. The displayed equation above shows that $H^ i(C^\wedge ) = 0$ for $i \geq 0$. Applying the distinguished triangle $K^\wedge \to L^\wedge \to C^\wedge \to K^\wedge [1]$ we get (1).

Proof of (2). Pick a distinguished triangle $K \to L \to C \to K[1]$. Then $H^ i(C) = 0$ for $i < 0$. Since $K_ n$ is sitting in degrees $-r, \ldots , 0$ we see that $H^ i(C \otimes _ A^\mathbf {L} K_ n) = 0$ for $i < -r$. The displayed equation above shows that $H^ i(C^\wedge ) = 0$ for $i < r$. Applying the distinguished triangle $K^\wedge \to L^\wedge \to C^\wedge \to K^\wedge [1]$ we get (2). $\square$

Lemma 15.91.21. Let $A$ be a ring and let $I \subset A$ be a finitely generated ideal. Let $K^\bullet $ be a filtered complex of $A$-modules. There exists a canonical spectral sequence $(E_ r, \text{d}_ r)_{r \geq 1}$ of bigraded derived complete $A$-modules with $d_ r$ of bidegree $(r, -r + 1)$ and with

\[ E_1^{p, q} = H^{p + q}((\text{gr}^ pK^\bullet )^\wedge ) \]

If the filtration on each $K^ n$ is finite, then the spectral sequence is bounded and converges to $H^*((K^\bullet )^\wedge )$.

Proof. By Lemma 15.91.10 we know that derived completion is given by $R\mathop{\mathrm{Hom}}\nolimits _ A(C, -)$ for some $C \in D^ b(A)$. By Lemmas 15.91.20 and 15.68.2 we see that $C$ has finite projective dimension. Thus we may choose a bounded complex of projective modules $P^\bullet $ representing $C$. Then

\[ M^\bullet = \mathop{\mathrm{Hom}}\nolimits ^\bullet (P^\bullet , K^\bullet ) \]

is a complex of $A$-modules representing $(K^\bullet )^\wedge $. It comes with a filtration given by $F^ pM^\bullet = \mathop{\mathrm{Hom}}\nolimits ^\bullet (P^\bullet , F^ pK^\bullet )$. We see that $F^ pM^\bullet $ represents $(F^ pK^\bullet )^\wedge $ and hence $\text{gr}^ pM^\bullet $ represents $(\text{gr}K^\bullet )^\wedge $. Thus we find our spectral sequence by taking the spectral sequence of the filtered complex $M^\bullet $, see Homology, Section 12.24. If the filtration on each $K^ n$ is finite, then the filtration on each $M^ n$ is finite because $P^\bullet $ is a bounded complex. Hence the final statement follows from Homology, Lemma 12.24.11. $\square$

Example 15.91.22. Let $A$ be a ring and let $I \subset A$ be a finitely generated ideal. Let $K^\bullet $ be a complex of $A$-modules. We can apply Lemma 15.91.21 with $F^ pK^\bullet = \tau _{\leq -p}K^\bullet $. Then we get a bounded spectral sequence

\[ E_1^{p, q} = H^{p + q}(H^{-p}(K^\bullet )^\wedge [p]) = H^{2p + q}(H^{-p}(K^\bullet )^\wedge ) \]

converging to $H^{p + q}((K^\bullet )^\wedge )$. After renumbering $p = -j$ and $q = i + 2j$ we find that for any $K \in D(A)$ there is a bounded spectral sequence $(E'_ r, d'_ r)_{r \geq 2}$ of bigraded derived complete modules with $d'_ r$ of bidegree $(r, -r + 1)$, with

\[ (E'_2)^{i, j} = H^ i(H^ j(K)^\wedge ) \]

and converging to $H^{i + j}(K^\wedge )$.

Lemma 15.91.23. Let $A \to B$ be a ring map. Let $I \subset A$ be an ideal. The inverse image of $D_{comp}(A, I)$ under the restriction functor $D(B) \to D(A)$ is $D_{comp}(B, IB)$.

Proof. Using Lemma 15.91.2 we see that $L \in D(B)$ is in $D_{comp}(B, IB)$ if and only if $T(L, f)$ is zero for every local section $f \in I$. Observe that the cohomology of $T(L, f)$ is computed in the category of abelian groups, so it doesn't matter whether we think of $f$ as an element of $A$ or take the image of $f$ in $B$. The lemma follows immediately from this and the definition of derived complete objects. $\square$

Lemma 15.91.24. Let $A \to B$ be a ring map. Let $I \subset A$ be a finitely generated ideal. If $A \to B$ is flat and $A/I \cong B/IB$, then the restriction functor $D(B) \to D(A)$ induces an equivalence $D_{comp}(B, IB) \to D_{comp}(A, I)$.

Proof. Choose generators $f_1, \ldots , f_ r$ of $I$. Denote $\check{\mathcal{C}}^\bullet _ A \to \check{\mathcal{C}}^\bullet _ B$ the quasi-isomorphism of extended alternating Čech complexes of Lemma 15.89.4. Let $K \in D_{comp}(A, I)$. Let $I^\bullet $ be a K-injective complex of $A$-modules representing $K$. Since $\mathop{\mathrm{Ext}}\nolimits ^ n_ A(A_ f, K)$ and $\mathop{\mathrm{Ext}}\nolimits ^ n_ A(B_ f, K)$ are zero for all $f \in I$ and $n \in \mathbf{Z}$ (Lemma 15.91.1) we conclude that $\check{\mathcal{C}}^\bullet _ A \to A$ and $\check{\mathcal{C}}^\bullet _ B \to B$ induce quasi-isomorphisms

\[ I^\bullet = \mathop{\mathrm{Hom}}\nolimits _ A(A, I^\bullet ) \longrightarrow \text{Tot}(\mathop{\mathrm{Hom}}\nolimits _ A(\check{\mathcal{C}}^\bullet _ A, I^\bullet )) \]

and

\[ \mathop{\mathrm{Hom}}\nolimits _ A(B, I^\bullet ) \longrightarrow \text{Tot}(\mathop{\mathrm{Hom}}\nolimits _ A(\check{\mathcal{C}}^\bullet _ B, I^\bullet )) \]

Some details omitted. Since $\check{\mathcal{C}}^\bullet _ A \to \check{\mathcal{C}}^\bullet _ B$ is a quasi-isomorphism and $I^\bullet $ is K-injective we conclude that $\mathop{\mathrm{Hom}}\nolimits _ A(B, I^\bullet ) \to I^\bullet $ is a quasi-isomorphism. As the complex $\mathop{\mathrm{Hom}}\nolimits _ A(B, I^\bullet )$ is a complex of $B$-modules we conclude that $K$ is in the image of the restriction map, i.e., the functor is essentially surjective

In fact, the argument shows that $F : D_{comp}(A, I) \to D_{comp}(B, IB)$, $K \mapsto \mathop{\mathrm{Hom}}\nolimits _ A(B, I^\bullet )$ is a left inverse to restriction. Finally, suppose that $L \in D_{comp}(B, IB)$. Represent $L$ by a K-injective complex $J^\bullet $ of $B$-modules. Then $J^\bullet $ is also K-injective as a complex of $A$-modules (Lemma 15.56.1) hence $F(\text{restriction of }L) = \mathop{\mathrm{Hom}}\nolimits _ A(B, J^\bullet )$. There is a map $J^\bullet \to \mathop{\mathrm{Hom}}\nolimits _ A(B, J^\bullet )$ of complexes of $B$-modules, whose composition with $\mathop{\mathrm{Hom}}\nolimits _ A(B, J^\bullet ) \to J^\bullet $ is the identity. We conclude that $F$ is also a right inverse to restriction and the proof is finished. $\square$


Comments (14)

Comment #5568 by Pavel Čoupek on

I would suggest a remark that mentions that derived -completion is not always the left derived functor of -adic completion, despite the suggestive terminology. (The difference is discussed e.g. in papers of Yekutieli (https://arxiv.org/abs/2002.04901 ) and Positselski (https://arxiv.org/abs/2002.12331 ).)

Comment #5754 by Pavel Čoupek on

I perhaps meant some different derived functor - namely the left derived functor of the -adic completion treated as a functor . That functor is quite close to the derived -completion (for example, it always lands in ) and agrees with it in some notable cases (e.g. Noetherian, generated by a reular sequence) but not always. Since itself does not land in -adically complete modules, the existence of derived complete modules that are not -adically complete does not justify why these functors are different in general. (But if this interpretation is not something that occurs to people, perhaps the warning is not needed after all...)

Comment #5763 by on

In the stacks project there is a well defined notion of taking the left or right derived functor of an additive functor between abelian categories which I believe to be entirely standard and must surely agree with what you mean by left derived functor as well (when you talk to your students for example). The example shows that doing this doesn't give derived completion (as defined in the Stacks project) even if the ring is Noetherian (even a complete dvr). I'm guessing what you mean is that there is a difference between the functor called derived completion (in the Stacks project) and the functor sending an objec in to which is indeed an object of . The agreement/difference between these two functors is discussed in depth in the following two sections so it does seem very unlikely that any confusion will arise (as you point out yourself in #5754). Thanks!

Comment #5769 by Pavel Čoupek on

If you do mean the standard definition of left derived functor for additive functor between abelian categories, then we do mean the same thing. The cat. of -adically complete modules is not abelian, and the functor of -adic completion is not even right exact (not even for a complete dvr). The point I was trying to make in #5754 is that the left derived functor of the -adic completion is thus not expected to land in complexes with -adically complete homologies or anything of that sort. Not even for the zeroth homology (as the -th left derived functor is not equivalent to -adic completion). So mere existence of modules that are derived -complete and not -adically complete does not say much about the derived functor.

Comment #5770 by on

Sorry, yes, of course. My bad; don't know what I was thinking. I have fixed this by removing the erroneous statement here. Thanks for insisting!

Comment #5974 by Bjorn Poonen on

At the top, the link to [dag12] goes to a bibliographic entry that has a bad URL. The URL for that entry should be http://people.math.harvard.edu/~lurie/papers/DAG-XII.pdf I think it may be a problem with the encoding of the tilde.

Comment #6152 by on

Dear Bjorn, thanks for reporting and we've listed this as a bug for the website.

Comment #6371 by on

@Pavel Čoupek. If we take and then we get an example where the derived completion of has nonzero cohomology in degrees and . Thus derived completion does not agree with the left derived functor of any additive functor as the value of on is always just .

Amusingly, the same example shows that derived completion of an object of is not the same as . See introduction to Section 15.94.

Anyway, we could add this discussion of "derived completion isn't a left derived functor" to the text. Question to all readers: should we do this? What do you think?

Comment #6490 by Daichi Takeuchi on

The E_2 terms of the spectral sequence in Lemma 091P vanish except for q=0,1, not for p=0,1, which I think merely shows that (4), (5) are stronger than the others. But they are equivalent for a two-term complex, so Lemma 091U (1) is OK.

Comment #6491 by Daichi Takeuchi on

I'm Sorry. I misunderstood. p, q are misplaced?


Post a comment

Your email address will not be published. Required fields are marked.

In your comment you can use Markdown and LaTeX style mathematics (enclose it like $\pi$). A preview option is available if you wish to see how it works out (just click on the eye in the toolbar).

Unfortunately JavaScript is disabled in your browser, so the comment preview function will not work.

All contributions are licensed under the GNU Free Documentation License.




In order to prevent bots from posting comments, we would like you to prove that you are human. You can do this by filling in the name of the current tag in the following input field. As a reminder, this is tag 091N. Beware of the difference between the letter 'O' and the digit '0'.