Locally nilpotent derivation

In mathematics, a derivation {\displaystyle \partial } of a commutative ring A {\displaystyle A} is called a locally nilpotent derivation (LND) if every element of A {\displaystyle A} is annihilated by some power of {\displaystyle \partial } .

One motivation for the study of locally nilpotent derivations comes from the fact that some of the counterexamples to Hilbert's 14th problem are obtained as the kernels of a derivation on a polynomial ring.[1]

Over a field k {\displaystyle k} of characteristic zero, to give a locally nilpotent derivation on the integral domain A {\displaystyle A} , finitely generated over the field, is equivalent to giving an action of the additive group ( k , + ) {\displaystyle (k,+)} to the affine variety X = Spec ( A ) {\displaystyle X=\operatorname {Spec} (A)} . Roughly speaking, an affine variety admitting "plenty" of actions of the additive group is considered similar to an affine space.[2]

Definition

Let A {\displaystyle A} be a ring. Recall that a derivation of A {\displaystyle A} is a map : A A {\displaystyle \partial \colon \,A\to A} satisfying the Leibniz rule ( a b ) = ( a ) b + a ( b ) {\displaystyle \partial (ab)=(\partial a)b+a(\partial b)} for any a , b A {\displaystyle a,b\in A} . If A {\displaystyle A} is an algebra over a field k {\displaystyle k} , we additionally require {\displaystyle \partial } to be k {\displaystyle k} -linear, so k ker {\displaystyle k\subseteq \ker \partial } .

A derivation {\displaystyle \partial } is called a locally nilpotent derivation (LND) if for every a A {\displaystyle a\in A} , there exists a positive integer n {\displaystyle n} such that n ( a ) = 0 {\displaystyle \partial ^{n}(a)=0} .

If A {\displaystyle A} is graded, we say that a locally nilpotent derivation {\displaystyle \partial } is homogeneous (of degree d {\displaystyle d} ) if deg a = deg a + d {\displaystyle \deg \partial a=\deg a+d} for every a A {\displaystyle a\in A} .

The set of locally nilpotent derivations of a ring A {\displaystyle A} is denoted by LND ( A ) {\displaystyle \operatorname {LND} (A)} . Note that this set has no obvious structure: it is neither closed under addition (e.g. if 1 = y x {\displaystyle \partial _{1}=y{\tfrac {\partial }{\partial x}}} , 2 = x y {\displaystyle \partial _{2}=x{\tfrac {\partial }{\partial y}}} then 1 , 2 LND ( k [ x , y ] ) {\displaystyle \partial _{1},\partial _{2}\in \operatorname {LND} (k[x,y])} but ( 1 + 2 ) 2 ( x ) = x {\displaystyle (\partial _{1}+\partial _{2})^{2}(x)=x} , so 1 + 2 LND ( k [ x , y ] ) {\displaystyle \partial _{1}+\partial _{2}\not \in \operatorname {LND} (k[x,y])} ) nor under multiplication by elements of A {\displaystyle A} (e.g. x LND ( k [ x ] ) {\displaystyle {\tfrac {\partial }{\partial x}}\in \operatorname {LND} (k[x])} , but x x LND ( k [ x ] ) {\displaystyle x{\tfrac {\partial }{\partial x}}\not \in \operatorname {LND} (k[x])} ). However, if [ 1 , 2 ] = 0 {\displaystyle [\partial _{1},\partial _{2}]=0} then 1 , 2 LND ( A ) {\displaystyle \partial _{1},\partial _{2}\in \operatorname {LND} (A)} implies 1 + 2 LND ( A ) {\displaystyle \partial _{1}+\partial _{2}\in \operatorname {LND} (A)} [3] and if LND ( A ) {\displaystyle \partial \in \operatorname {LND} (A)} , h ker {\displaystyle h\in \ker \partial } then h LND ( A ) {\displaystyle h\partial \in \operatorname {LND} (A)} .

Relation to Ga-actions

Let A {\displaystyle A} be an algebra over a field k {\displaystyle k} of characteristic zero (e.g. k = C {\displaystyle k=\mathbb {C} } ). Then there is a one-to-one correspondence between the locally nilpotent k {\displaystyle k} -derivations on A {\displaystyle A} and the actions of the additive group G a {\displaystyle \mathbb {G} _{a}} of k {\displaystyle k} on the affine variety Spec A {\displaystyle \operatorname {Spec} A} , as follows.[3] A G a {\displaystyle \mathbb {G} _{a}} -action on Spec A {\displaystyle \operatorname {Spec} A} corresponds to a k {\displaystyle k} -algebra homomorphism ρ : A A [ t ] {\displaystyle \rho \colon A\to A[t]} . Any such ρ {\displaystyle \rho } determines a locally nilpotent derivation {\displaystyle \partial } of A {\displaystyle A} by taking its derivative at zero, namely = ϵ d d t ρ , {\displaystyle \partial =\epsilon \circ {\tfrac {d}{dt}}\circ \rho ,} where ϵ {\displaystyle \epsilon } denotes the evaluation at t = 0 {\displaystyle t=0} . Conversely, any locally nilpotent derivation {\displaystyle \partial } determines a homomorphism ρ : A A [ t ] {\displaystyle \rho \colon A\to A[t]} by ρ = exp ( t ) = n = 0 t n n ! n . {\displaystyle \rho =\exp(t\partial )=\sum _{n=0}^{\infty }{\frac {t^{n}}{n!}}\partial ^{n}.}

It is easy to see that the conjugate actions correspond to conjugate derivations, i.e. if α Aut A {\displaystyle \alpha \in \operatorname {Aut} A} and LND ( A ) {\displaystyle \partial \in \operatorname {LND} (A)} then α α 1 LND ( A ) {\displaystyle \alpha \circ \partial \circ \alpha ^{-1}\in \operatorname {LND} (A)} and exp ( t α α 1 ) = α exp ( t ) α 1 {\displaystyle \exp(t\cdot \alpha \circ \partial \circ \alpha ^{-1})=\alpha \circ \exp(t\partial )\circ \alpha ^{-1}}

The kernel algorithm

The algebra ker {\displaystyle \ker \partial } consists of the invariants of the corresponding G a {\displaystyle \mathbb {G} _{a}} -action. It is algebraically and factorially closed in A {\displaystyle A} .[3] A special case of Hilbert's 14th problem asks whether ker {\displaystyle \ker \partial } is finitely generated, or, if A = k [ X ] {\displaystyle A=k[X]} , whether the quotient X / / G a {\displaystyle X/\!/\mathbb {G} _{a}} is affine. By Zariski's finiteness theorem,[4] it is true if dim X 3 {\displaystyle \dim X\leq 3} . On the other hand, this question is highly nontrivial even for X = C n {\displaystyle X=\mathbb {C} ^{n}} , n 4 {\displaystyle n\geq 4} . For n 5 {\displaystyle n\geq 5} the answer, in general, is negative.[5] The case n = 4 {\displaystyle n=4} is open.[3]

However, in practice it often happens that ker {\displaystyle \ker \partial } is known to be finitely generated: notably, by the Maurer–Weitzenböck theorem,[6] it is the case for linear LND's of the polynomial algebra over a field of characteristic zero (by linear we mean homogeneous of degree zero with respect to the standard grading).

Assume ker {\displaystyle \ker \partial } is finitely generated. If A = k [ g 1 , , g n ] {\displaystyle A=k[g_{1},\dots ,g_{n}]} is a finitely generated algebra over a field of characteristic zero, then ker {\displaystyle \ker \partial } can be computed using van den Essen's algorithm,[7] as follows. Choose a local slice, i.e. an element r ker 2 ker {\displaystyle r\in \ker \partial ^{2}\setminus \ker \partial } and put f = r ker {\displaystyle f=\partial r\in \ker \partial } . Let π r : A ( ker ) f {\displaystyle \pi _{r}\colon \,A\to (\ker \partial )_{f}} be the Dixmier map given by π r ( a ) = n = 0 ( 1 ) n n ! n ( a ) r n f n {\displaystyle \pi _{r}(a)=\sum _{n=0}^{\infty }{\frac {(-1)^{n}}{n!}}\partial ^{n}(a){\frac {r^{n}}{f^{n}}}} . Now for every i = 1 , , n {\displaystyle i=1,\dots ,n} , chose a minimal integer m i {\displaystyle m_{i}} such that h i : = f m i π r ( g i ) ker {\displaystyle h_{i}\colon =f^{m_{i}}\pi _{r}(g_{i})\in \ker \partial } , put B 0 = k [ h 1 , , h n , f ] ker {\displaystyle B_{0}=k[h_{1},\dots ,h_{n},f]\subseteq \ker \partial } , and define inductively B i {\displaystyle B_{i}} to be the subring of A {\displaystyle A} generated by { h A : f h B i 1 } {\displaystyle \{h\in A:fh\in B_{i-1}\}} . By induction, one proves that B 0 B 1 ker {\displaystyle B_{0}\subset B_{1}\subset \dots \subset \ker \partial } are finitely generated and if B i = B i + 1 {\displaystyle B_{i}=B_{i+1}} then B i = ker {\displaystyle B_{i}=\ker \partial } , so B N = ker {\displaystyle B_{N}=\ker \partial } for some N {\displaystyle N} . Finding the generators of each B i {\displaystyle B_{i}} and checking whether B i = B i + 1 {\displaystyle B_{i}=B_{i+1}} is a standard computation using Gröbner bases.[7]

Slice theorem

Assume that LND ( A ) {\displaystyle \partial \in \operatorname {LND} (A)} admits a slice, i.e. s A {\displaystyle s\in A} such that s = 1 {\displaystyle \partial s=1} . The slice theorem[3] asserts that A {\displaystyle A} is a polynomial algebra ( ker ) [ s ] {\displaystyle (\ker \partial )[s]} and = d d s {\displaystyle \partial ={\tfrac {d}{ds}}} .

For any local slice r ker ker 2 {\displaystyle r\in \ker \partial \setminus \ker \partial ^{2}} we can apply the slice theorem to the localization A r {\displaystyle A_{\partial r}} , and thus obtain that A {\displaystyle A} is locally a polynomial algebra with a standard derivation. In geometric terms, if a geometric quotient π : X X / / G a {\displaystyle \pi \colon \,X\to X//\mathbb {G} _{a}} is affine (e.g. when dim X 3 {\displaystyle \dim X\leq 3} by the Zariski theorem), then it has a Zariski-open subset U {\displaystyle U} such that π 1 ( U ) {\displaystyle \pi ^{-1}(U)} is isomorphic over U {\displaystyle U} to U × A 1 {\displaystyle U\times \mathbb {A} ^{1}} , where G a {\displaystyle \mathbb {G} _{a}} acts by translation on the second factor.

However, in general it is not true that X X / / G a {\displaystyle X\to X//\mathbb {G} _{a}} is locally trivial. For example,[8] let = u x + v y + ( 1 + u y 2 ) z LND ( C [ x , y , z , u , v ] ) {\displaystyle \partial =u{\tfrac {\partial }{\partial x}}+v{\tfrac {\partial }{\partial y}}+(1+uy^{2}){\tfrac {\partial }{\partial z}}\in \operatorname {LND} (\mathbb {C} [x,y,z,u,v])} . Then ker {\displaystyle \ker \partial } is a coordinate ring of a singular variety, and the fibers of the quotient map over singular points are two-dimensional.

If dim X = 3 {\displaystyle \dim X=3} then Γ = X U {\displaystyle \Gamma =X\setminus U} is a curve. To describe the G a {\displaystyle \mathbb {G} _{a}} -action, it is important to understand the geometry Γ {\displaystyle \Gamma } . Assume further that k = C {\displaystyle k=\mathbb {C} } and that X {\displaystyle X} is smooth and contractible (in which case S {\displaystyle S} is smooth and contractible as well[9]) and choose Γ {\displaystyle \Gamma } to be minimal (with respect to inclusion). Then Kaliman proved[10] that each irreducible component of Γ {\displaystyle \Gamma } is a polynomial curve, i.e. its normalization is isomorphic to C 1 {\displaystyle \mathbb {C} ^{1}} . The curve Γ {\displaystyle \Gamma } for the action given by Freudenburg's (2,5)-derivation (see below) is a union of two lines in C 2 {\displaystyle \mathbb {C} ^{2}} , so Γ {\displaystyle \Gamma } may not be irreducible. However, it is conjectured that Γ {\displaystyle \Gamma } is always contractible.[11]

Examples

Example 1

The standard coordinate derivations x i {\displaystyle {\tfrac {\partial }{\partial x_{i}}}} of a polynomial algebra k [ x 1 , , x n ] {\displaystyle k[x_{1},\dots ,x_{n}]} are locally nilpotent. The corresponding G a {\displaystyle \mathbb {G} _{a}} -actions are translations: t x i = x i + t {\displaystyle t\cdot x_{i}=x_{i}+t} , t x j = x j {\displaystyle t\cdot x_{j}=x_{j}} for j i {\displaystyle j\neq i} .

Example 2 (Freudenburg's (2,5)-homogeneous derivation[12])

Let f 1 = x 1 x 3 x 2 2 {\displaystyle f_{1}=x_{1}x_{3}-x_{2}^{2}} , f 2 = x 3 f 1 2 + 2 x 1 2 x 2 f 1 + x 5 {\displaystyle f_{2}=x_{3}f_{1}^{2}+2x_{1}^{2}x_{2}f_{1}+x^{5}} , and let {\displaystyle \partial } be the Jacobian derivation ( f 3 ) = det [ f i x j ] i , j = 1 , 2 , 3 {\textstyle \partial (f_{3})=\det \left[{\tfrac {\partial f_{i}}{\partial x_{j}}}\right]_{i,j=1,2,3}} . Then LND ( k [ x 1 , x 2 , x 3 ] ) {\displaystyle \partial \in \operatorname {LND} (k[x_{1},x_{2},x_{3}])} and rank = 3 {\displaystyle \operatorname {rank} \partial =3} (see below); that is, {\displaystyle \partial } annihilates no variable. The fixed point set of the corresponding G a {\displaystyle \mathbb {G} _{a}} -action equals { x 1 = x 2 = 0 } {\displaystyle \{x_{1}=x_{2}=0\}} .

Example 3

Consider S l 2 ( k ) = { a d b c = 1 } k 4 {\displaystyle Sl_{2}(k)=\{ad-bc=1\}\subseteq k^{4}} . The locally nilpotent derivation a b + c d {\displaystyle a{\tfrac {\partial }{\partial b}}+c{\tfrac {\partial }{\partial d}}} of its coordinate ring corresponds to a natural action of G a {\displaystyle \mathbb {G} _{a}} on S l 2 ( k ) {\displaystyle Sl_{2}(k)} via right multiplication of upper triangular matrices. This action gives a nontrivial G a {\displaystyle \mathbb {G} _{a}} -bundle over A 2 { ( 0 , 0 ) } {\displaystyle \mathbb {A} ^{2}\setminus \{(0,0)\}} . However, if k = C {\displaystyle k=\mathbb {C} } then this bundle is trivial in the smooth category[13]

LND's of the polynomial algebra

Let k {\displaystyle k} be a field of characteristic zero (using Kambayashi's theorem one can reduce most results to the case k = C {\displaystyle k=\mathbb {C} } [14]) and let A = k [ x 1 , , x n ] {\displaystyle A=k[x_{1},\dots ,x_{n}]} be a polynomial algebra.

n = 2 (Ga-actions on an affine plane)

Rentschler's theorem — Every LND of k [ x 1 , x 2 ] {\displaystyle k[x_{1},x_{2}]} can be conjugated to f ( x 1 ) x 2 {\displaystyle f(x_{1}){\tfrac {\partial }{\partial x_{2}}}} for some f ( x 1 ) k [ x 1 ] {\displaystyle f(x_{1})\in k[x_{1}]} . This result is closely related to the fact that every automorphism of an affine plane is tame, and does not hold in higher dimensions.[15]

n = 3 (Ga-actions on an affine 3-space)

Miyanishi's theorem — The kernel of every nontrivial LND of A = k [ x 1 , x 2 , x 3 ] {\displaystyle A=k[x_{1},x_{2},x_{3}]} is isomorphic to a polynomial ring in two variables; that is, a fixed point set of every nontrivial G a {\displaystyle \mathbb {G} _{a}} -action on A 3 {\displaystyle \mathbb {A} ^{3}} is isomorphic to A 2 {\displaystyle \mathbb {A} ^{2}} .[16][17]

In other words, for every 0 LND ( A ) {\displaystyle 0\neq \partial \in \operatorname {LND} (A)} there exist f 1 , f 2 A {\displaystyle f_{1},f_{2}\in A} such that ker = k [ f 1 , f 2 ] {\displaystyle \ker \partial =k[f_{1},f_{2}]} (but, in contrast to the case n = 2 {\displaystyle n=2} , A {\displaystyle A} is not necessarily a polynomial ring over ker {\displaystyle \ker \partial } ). In this case, {\displaystyle \partial } is a Jacobian derivation: ( f 3 ) = det [ f i x j ] i , j = 1 , 2 , 3 {\textstyle \partial (f_{3})=\det \left[{\tfrac {\partial f_{i}}{\partial x_{j}}}\right]_{i,j=1,2,3}} .[18]

Zurkowski's theorem — Assume that n = 3 {\displaystyle n=3} and LND ( A ) {\displaystyle \partial \in \operatorname {LND} (A)} is homogeneous relative to some positive grading of A {\displaystyle A} such that x 1 , x 2 , x 3 {\displaystyle x_{1},x_{2},x_{3}} are homogeneous. Then ker = k [ f , g ] {\displaystyle \ker \partial =k[f,g]} for some homogeneous f , g {\displaystyle f,g} . Moreover,[18] if deg x 1 , deg x 2 , deg x 3 {\displaystyle \deg x_{1},\deg x_{2},\deg x_{3}} are relatively prime, then deg f , deg g {\displaystyle \deg f,\deg g} are relatively prime as well.[19][3]

Bonnet's theorem — A quotient morphism A 3 A 2 {\displaystyle \mathbb {A} ^{3}\to \mathbb {A} ^{2}} of a G a {\displaystyle \mathbb {G} _{a}} -action is surjective. In other words, for every 0 LND ( A ) {\displaystyle 0\neq \partial \in \operatorname {LND} (A)} , the embedding ker A {\displaystyle \ker \partial \subseteq A} induces a surjective morphism Spec A Spec ker {\displaystyle \operatorname {Spec} A\to \operatorname {Spec} \ker \partial } .[20][10]

This is no longer true for n 4 {\displaystyle n\geqslant 4} , e.g. the image of a quotient map A 4 A 3 {\displaystyle \mathbb {A} ^{4}\to \mathbb {A} ^{3}} by a G a {\displaystyle \mathbb {G} _{a}} -action t ( x 1 , x 2 , x 3 , x 4 ) = ( x 1 , x 2 , x 3 t x 2 , x 4 + t x 1 ) {\displaystyle t\cdot (x_{1},x_{2},x_{3},x_{4})=(x_{1},x_{2},x_{3}-tx_{2},x_{4}+tx_{1})} (which corresponds to a LND given by x 1 x 4 x 2 x 3 ) {\displaystyle x_{1}{\tfrac {\partial }{\partial x_{4}}}-x_{2}{\tfrac {\partial }{\partial x_{3}}})} equals A 3 { ( x 1 , x 2 , x 3 ) : x 1 = x 2 = 0 , x 3 0 } {\displaystyle \mathbb {A} ^{3}\setminus \{(x_{1},x_{2},x_{3}):x_{1}=x_{2}=0,x_{3}\neq 0\}} .

Kaliman's theorem — Every fixed-point free action of G a {\displaystyle \mathbb {G} _{a}} on A 3 {\displaystyle \mathbb {A} ^{3}} is conjugate to a translation. In other words, every LND ( A ) {\displaystyle \partial \in \operatorname {LND} (A)} such that the image of {\displaystyle \partial } generates the unit ideal (or, equivalently, {\displaystyle \partial } defines a nowhere vanishing vector field), admits a slice. This results answers one of the conjectures from Kraft's list.[10]

Again, this result is not true for n 4 {\displaystyle n\geqslant 4} :[21] e.g. consider the x 1 x 2 + x 2 x 3 + ( x 2 2 2 x 1 x 3 1 ) x 4 LND ( C [ x 1 , x 2 , x 3 , x 4 ] ) {\displaystyle x_{1}{\tfrac {\partial }{\partial x_{2}}}+x_{2}{\tfrac {\partial }{\partial x_{3}}}+(x_{2}^{2}-2x_{1}x_{3}-1){\tfrac {\partial }{\partial x_{4}}}\in \operatorname {LND} (\mathbb {C} [x_{1},x_{2},x_{3},x_{4}])} . The points ( x 1 , 1 , 0 , 0 ) {\displaystyle (x_{1},1,0,0)} and ( x 1 , 1 , 0 , 0 ) {\displaystyle (x_{1},-1,0,0)} are in the same orbit of the corresponding G a {\displaystyle \mathbb {G} _{a}} -action if and only if x 1 0 {\displaystyle x_{1}\neq 0} ; hence the (topological) quotient is not even Hausdorff, let alone homeomorphic to C 3 {\displaystyle \mathbb {C} ^{3}} .

Principal ideal theorem — Let LND ( A ) {\displaystyle \partial \in \operatorname {LND} (A)} . Then A {\displaystyle A} is faithfully flat over ker {\displaystyle \ker \partial } . Moreover, the ideal ker im {\displaystyle \ker \partial \cap \operatorname {im} \partial } is principal in A {\displaystyle A} .[14]

Triangular derivations

Let f 1 , , f n {\displaystyle f_{1},\dots ,f_{n}} be any system of variables of A {\displaystyle A} ; that is, A = k [ f 1 , , f n ] {\displaystyle A=k[f_{1},\dots ,f_{n}]} . A derivation of A {\displaystyle A} is called triangular with respect to this system of variables, if f 1 k {\displaystyle \partial f_{1}\in k} and f i k [ f 1 , , f i 1 ] {\displaystyle \partial f_{i}\in k[f_{1},\dots ,f_{i-1}]} for i = 2 , , n {\displaystyle i=2,\dots ,n} . A derivation is called triangulable if it is conjugate to a triangular one, or, equivalently, if it is triangular with respect to some system of variables. Every triangular derivation is locally nilpotent. The converse is true for 2 {\displaystyle \leq 2} by Rentschler's theorem above, but it is not true for n 3 {\displaystyle n\geq 3} .

Bass's example

The derivation of k [ x 1 , x 2 , x 3 ] {\displaystyle k[x_{1},x_{2},x_{3}]} given by x 1 x 2 + 2 x 2 x 1 x 3 {\displaystyle x_{1}{\tfrac {\partial }{\partial x_{2}}}+2x_{2}x_{1}{\tfrac {\partial }{\partial x_{3}}}} is not triangulable.[22] Indeed, the fixed-point set of the corresponding G a {\displaystyle \mathbb {G} _{a}} -action is a quadric cone x 2 x 3 = x 2 2 {\displaystyle x_{2}x_{3}=x_{2}^{2}} , while by the result of Popov,[23] a fixed point set of a triangulable G a {\displaystyle \mathbb {G} _{a}} -action is isomorphic to Z × A 1 {\displaystyle Z\times \mathbb {A} ^{1}} for some affine variety Z {\displaystyle Z} ; and thus cannot have an isolated singularity.

Freudenburg's theorem — The above necessary geometrical condition was later generalized by Freudenburg.[24] To state his result, we need the following definition:

A corank of LND ( A ) {\displaystyle \partial \in \operatorname {LND} (A)} is a maximal number j {\displaystyle j} such that there exists a system of variables f 1 , , f n {\displaystyle f_{1},\dots ,f_{n}} such that f 1 , , f j ker {\displaystyle f_{1},\dots ,f_{j}\in \ker \partial } . Define rank {\displaystyle \operatorname {rank} \partial } as n {\displaystyle n} minus the corank of {\displaystyle \partial } .

We have 1 rank n {\displaystyle 1\leq \operatorname {rank} \partial \leq n} and rank ( ) = 1 {\displaystyle \operatorname {rank} (\partial )=1} if and only if in some coordinates, = h x n {\displaystyle \partial =h{\tfrac {\partial }{\partial x_{n}}}} for some h k [ x 1 , , x n 1 ] {\displaystyle h\in k[x_{1},\dots ,x_{n-1}]} .[24]

Theorem: If LND ( A ) {\displaystyle \partial \in \operatorname {LND} (A)} is triangulable, then any hypersurface contained in the fixed-point set of the corresponding G a {\displaystyle \mathbb {G} _{a}} -action is isomorphic to Z × A rank {\displaystyle Z\times \mathbb {A} ^{\operatorname {rank} \partial }} .[24]

In particular, LND's of maximal rank n {\displaystyle n} cannot be triangulable. Such derivations do exist for n 3 {\displaystyle n\geq 3} : the first example is the (2,5)-homogeneous derivation (see above), and it can be easily generalized to any n 3 {\displaystyle n\geq 3} .[12]

Makar-Limanov invariant

The intersection of the kernels of all locally nilpotent derivations of the coordinate ring, or, equivalently, the ring of invariants of all G a {\displaystyle \mathbb {G} _{a}} -actions, is called "Makar-Limanov invariant" and is an important algebraic invariant of an affine variety. For example, it is trivial for an affine space; but for the Koras–Russell cubic threefold, which is diffeomorphic to C 3 {\displaystyle \mathbb {C} ^{3}} , it is not.[25]

References

  1. ^ Daigle, Daniel. "Hilbert's Fourteenth Problem and Locally Nilpotent Derivations" (PDF). University of Ottawa. Retrieved 11 September 2018.
  2. ^ Arzhantsev, I.; Flenner, H.; Kaliman, S.; Kutzschebauch, F.; Zaidenberg, M. (2013). "Flexible varieties and automorphism groups". Duke Math. J. 162 (4): 767–823. arXiv:1011.5375. doi:10.1215/00127094-2080132. S2CID 53412676.
  3. ^ a b c d e f Freudenburg, G. (2006). Algebraic theory of locally nilpotent derivations. Berlin: Springer-Verlag. CiteSeerX 10.1.1.470.10. ISBN 978-3-540-29521-1.
  4. ^ Zariski, O. (1954). "Interprétations algébrico-géométriques du quatorzième problème de Hilbert". Bull. Sci. Math. (2). 78: 155–168.
  5. ^ Derksen, H. G. J. (1993). "The kernel of a derivation". J. Pure Appl. Algebra. 84 (1): 13–16. doi:10.1016/0022-4049(93)90159-Q.
  6. ^ Seshadri, C.S. (1962). "On a theorem of Weitzenböck in invariant theory". J. Math. Kyoto Univ. 1 (3): 403–409. doi:10.1215/kjm/1250525012.
  7. ^ a b van den Essen, A. (2000). Polynomial automorphisms and the Jacobian conjecture. Basel: Birkhäuser Verlag. doi:10.1007/978-3-0348-8440-2. ISBN 978-3-7643-6350-5. S2CID 252433637.
  8. ^ Deveney, J.; Finston, D. (1995). "A proper G a {\displaystyle \mathbb {G} _{a}} -action on C 5 {\displaystyle \mathbb {C} ^{5}} which is not locally trivial". Proc. Amer. Math. Soc. 123 (3): 651–655. doi:10.1090/S0002-9939-1995-1273487-0. JSTOR 2160782.
  9. ^ Kaliman, S; Saveliev, N. (2004). " C + {\displaystyle \mathbb {C} _{+}} -Actions on contractible threefolds". Michigan Math. J. 52 (3): 619–625. arXiv:math/0209306. doi:10.1307/mmj/1100623416. S2CID 15020160.
  10. ^ a b c Kaliman, S. (2004). "Free C + {\displaystyle \mathbb {C} _{+}} -actions on C 3 {\displaystyle \mathbb {C} ^{3}} are translations" (PDF). Invent. Math. 156 (1): 163–173. arXiv:math/0207156. doi:10.1007/s00222-003-0336-1. S2CID 15769378.
  11. ^ Kaliman, S. (2009). "Actions of C {\displaystyle \mathbb {C} ^{*}} and C + {\displaystyle \mathbb {C} _{+}} on affine algebraic varieties" (PDF). Algebraic geometry-Seattle 2005. Part 2. Proceedings of Symposia in Pure Mathematics. Vol. 80. pp. 629–654. doi:10.1090/pspum/080.2/2483949. ISBN 9780821847039.
  12. ^ a b Freudenburg, G. (1998). "Actions of G a {\displaystyle \mathbb {G} _{a}} on A 3 {\displaystyle \mathbb {A} ^{3}} defined by homogeneous derivations". Journal of Pure and Applied Algebra. 126 (1): 169–181. doi:10.1016/S0022-4049(96)00143-0.
  13. ^ Dubouloz, A.; Finston, D. (2014). "On exotic affine 3-spheres". J. Algebraic Geom. 23 (3): 445–469. arXiv:1106.2900. doi:10.1090/S1056-3911-2014-00612-3. S2CID 119651964.
  14. ^ a b Daigle, D.; Kaliman, S. (2009). "A note on locally nilpotent derivations and variables of k [ X , Y , Z ] {\displaystyle k[X,Y,Z]} " (PDF). Canad. Math. Bull. 52 (4): 535–543. doi:10.4153/CMB-2009-054-5.
  15. ^ Rentschler, R. (1968). "Opérations du groupe additif sur le plan affine". Comptes Rendus de l'Académie des Sciences, Série A-B. 267: A384–A387.
  16. ^ Miyanishi, M. (1986). "Normal affine subalgebras of a polynomial ring". Algebraic and Topological Theories (Kinosaki, 1984). pp. 37–51.
  17. ^ Sugie, T. (1989). "Algebraic Characterization of the Affine Plane and the Affine 3-Space". Topological Methods in Algebraic Transformation Groups (New Brunswick, NJ, 1988). Progress in Mathematics. Vol. 80. Birkhäuser Boston. pp. 177–190. doi:10.1007/978-1-4612-3702-0_12. ISBN 978-1-4612-8219-8.
  18. ^ a b D., Daigle (2000). "On kernels of homogeneous locally nilpotent derivations of k [ X , Y , Z ] {\displaystyle k[X,Y,Z]} ". Osaka J. Math. 37 (3): 689–699.
  19. ^ Zurkowski, V.D. "Locally finite derivations" (PDF).
  20. ^ Bonnet, P. (2002). "Surjectivity of quotient maps for algebraic ( C , + ) {\displaystyle (\mathbb {C} ,+)} -actions and polynomial maps with contractible fibers". Transform. Groups. 7 (1): 3–14. arXiv:math/0602227. doi:10.1007/s00031-002-0001-6.
  21. ^ Winkelmann, J. (1990). "On free holomorphic C {\displaystyle \mathbb {C} } -actions on C n {\displaystyle \mathbb {C} ^{n}} and homogeneous Stein manifolds" (PDF). Math. Ann. 286 (1–3): 593–612. doi:10.1007/BF01453590.
  22. ^ Bass, H. (1984). "A non-triangular action of G a {\displaystyle \mathbb {G} _{a}} on A 3 {\displaystyle \mathbb {A} ^{3}} ". Journal of Pure and Applied Algebra. 33 (1): 1–5. doi:10.1016/0022-4049(84)90019-7.
  23. ^ Popov, V. L. (1987). "On actions of G a {\displaystyle \mathbb {G} _{a}} on A n {\displaystyle \mathbb {A} ^{n}} ". Algebraic Groups Utrecht 1986. Lecture Notes in Mathematics. Vol. 1271. pp. 237–242. doi:10.1007/BFb0079241. ISBN 978-3-540-18234-4.
  24. ^ a b c Freudenburg, G. (1995). "Triangulability criteria for additive group actions on affine space". J. Pure Appl. Algebra. 105 (3): 267–275. doi:10.1016/0022-4049(96)87756-5.
  25. ^ Kaliman, S.; Makar-Limanov, L. (1997). "On the Russell-Koras contractible threefolds". J. Algebraic Geom. 6 (2): 247–268.

Further reading

  • A Nowicki, the fourteenth problem of hilbert for polynomial derivations