natural structures in differential geometry lucas...

101
NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown A thesis submitted in partial fulfillment of the requirements for the degree of Master of Science in mathematics Trinity College Dublin March 2015

Upload: others

Post on 14-Mar-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

NATURAL STRUCTURES IN DIFFERENTIALGEOMETRY

Lucas Mason-Brown

A thesis submitted in partial fulfillment of therequirements for the degree of Master of Science in

mathematics

Trinity College Dublin

March 2015

Page 2: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

Declaration

I declare that the thesis contained herein is entirely my own work, ex-cept where explicitly stated otherwise, and has not been submitted for adegree at this or any other university. I grant Trinity College Dublin per-mission to lend or copy this thesis upon request, with the understandingthat this permission covers only single copies made for study purposes,subject to normal conditions of acknowledgement.

Page 3: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

Contents

1 Summary 3

2 Acknowledgment 6

I Natural Bundles and Operators 6

3 Jets 6

4 Natural Bundles 9

5 Natural Operators 11

6 Invariant-Theoretic Reduction 16

7 Classical Results 24

8 Natural Operators on Differential Forms 32

II Natural Operators on Alternating Multivector Fields 38

9 Twisted Algebras in VecZ 40

10 Ordered Multigraphs 48

11 Ordered Multigraphs and Natural Operators 52

12 Gerstenhaber Algebras 57

13 Pre-Gerstenhaber Algebras 58

14 A Lie-admissable Structure on Grairred

acyc 63

15 A Co-algebra Structure on Grairred

acyc 67

16 Loday’s Rigidity Theorem 75

17 A Useful Consequence of Kunneth’s Theorem 83

18 Chevalley-Eilenberg Cohomology 87

1 Summary

Many of the most important constructions in differential geometry are functorial. Take, for

example, the tangent bundle. The tangent bundle can be profitably viewed as a functor T :

Manm → Fib from the category of m-dimensional manifolds (with local diffeomorphisms) to

3

Page 4: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

the category of fiber bundles (with locally invertible bundle maps). This functor takes every

m-dimensional manifold M to its tangent bundle TM := tm∈MTmM and every local diffeomor-

phism f : M → N to its pushforward Tf := f∗ : TM → TN . The statement that T is a functor

is essentially the chain rule.

Functors of the general type Manm → Fib are ubiquitous in differential geometry. Other

examples include: the cotangent bundle T ∗, tensor bundles and their various sub-bundles (most

notably the exterior bundles ∧nT and ∧nT ∗), frame bundles, jet bundles, and others. Over

the years such functors have been variously termed “bundle functors,” “natural bundles,” and

“geometric objects.” In this thesis, we use the term “natural bundle” throughout.

If we view natural bundles as functors, certain constructions play the role of natural transfor-

mations. Other constructions play the role of natural operators, transforming smooth sections

of one bundle into smooth sections of another, through a rule or recipe which respects the local

action of diffeomorphisms. Classical examples include: the exterior derivative, the wedge prod-

uct, the contraction (i.e. the evaluation of a form at a vector field), the Lie bracket, and the

much-more-general Schouten-Nijenhuis bracket of alternating multivector fields.

The functorial approach to natural bundles and operators was first introduced by A. Nijenhuis

in his talk at the 1958 International Congress of Mathematicians. The foundations of a general

theory were later developed by C.L. Terng and R. Palais ([1], [2]) with important contributions

from D.B.A. Epstein, W.P. Thurston ([3]), and others. Following in the early footsteps of R.

Palais, who demonstrated in 1959 that the exterior derivative is, apart from the identity, the only

unary natural operator on differential forms ([4]), recent work in the field of “natural differential

geometry” has largely been devoted to the task of classifying natural bundles and operators

of a prescribed type. Much of this work is summarized in the foundational monograph of I.

Kolar, P. Michor, and J. Slovak ([5]), which was the primary reference for the first part of this

thesis. However, since the publication of this monograph, a few major holes have been filled.

Over the past decade, M. Markl has published a string of papers, culminating in an exhaustive

classification of natural n-ary operators on vector fields. He demonstrates, not surprisingly, that

over high-dimensional manifolds the only such operators are given by linear combinations of

various iterations of the Lie bracket. His work draws heavily on Operad theory, particularly on

the work of F. Chapoton, M. Livernet ([6]), and J.-L Loday ([7]), as well as a graphical approach

to tensor calculus originally formalized by S.A. Merkulov in [8].

We see in the foundational work of R. Palais and the more recent work of M. Markl the contours

of a much more general problem. Namely, what is the complete space of natural multilinear

operators on differential forms and alternating multivector fields? More explicitly, if ∧αiT ∗ is

the natural bundle of differential forms of degree αi and ∧βiT is the natural bundle of alternating

multivector fields of degree βi, can we classify the space of natural multilinear operators of the

4

Page 5: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

general type:

η : ∧α1T ∗ × ...× ∧αpT ∗ × ∧β1T × ...× ∧βqT ∧α′1T ∗ × ...× ∧α

′r × ∧β

′1T × ...× ∧β

′sT (1)

We conjecture that all such operators can be built up from 5 ‘elementary’ operations. On

differential forms, there is the exterior derivative and the wedge product of differential forms.

On alternating multivector fields, there is the Schouten-Nijenhuis bracket and the wedge product

of alternating multivector fields. Linking the two is the contraction. We believe that (over high-

dimensional manifolds) all natural operators of type (1) can be realized as linear combinations of

iterations of these 5 operations. A proof of this conjecture would have intrinsic value within the

field of natural differential geometry. Such a proof would also be valuable in a much broader sense.

Natural operators are precisely the kinds of operators that present themselves in mathematical

formulations of fundamental physics. Furthermore, differential forms and alternating multivector

fields are among the most pervasive mathematical objects in these fundamental laws. Thus, a

proof of this conjecture would radically limit the range of plausible natural laws that physicists

might be tempted to formulate. In this thesis, we make several steps towards proving this

conjecture.

In the first part, we develop the basic theory of natural differential operators, roughly following

chapters IV, V, and VI of [5]. The key concept in the first part is the method of invariant-

theoretic reduction, whereby the geometric problem of classifying natural differential operators is

reduced to the much more tractable algebraic problem of classifying module homomorphisms of

a particular source and target. We conclude part I with an original proof that all natural n-ary

operators on differential forms are exhausted by iterations of the wedge product and exterior

derivative (making no assumption of multilinearity).

In the second and final part, we shift our attention to the more obstreperous case of alternating

multivector fields. Generalizing the work of M. Markl, we introduce the notion of an ordered

multigraph as a general framework for the study of natural operators on alternating multivec-

tor fields. Inspired by Markl, we draw connections between the space of natural operators on

alternating multivector fields and the cohomology of a particular complex of multigraphs. We

then develop the notion of a pre-Gerstenhaber algebra, which combines the structure of a graded

commutative algebra with the structure of a graded Lie-admissable algebra, and prove that these

operads are compatible in a suitable sense. We immediately see that this is the correct algebraic

structure for the geometric problem at hand. Finally, we apply Loday’s rigidity theorem to

deduce certain facts about the structure of the relevant algebras.

5

Page 6: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

2 Acknowledgment

Special thanks go to Vladimir Dotsenko. Without his patient support and insight, none of this

would have been possible. Thanks also to my undergraduate mentors and advisors—Benedict

Gross, Mike Rosen, Dan Abramovic, and Tom Goodwillie, just to name a few—who enabled and

empowered me as a mathematician.

Part I

Natural Bundles and Operators

3 Jets

Suppose f : R→ R is a smooth function and x0 ∈ R. By Taylor’s theorem

f(x) = f(x0) + f ′(x0)(x− x0) + ...+f (r)(x0)

r!(x− x0)r + hr(x)(x− x0)r

where limx→x0hr(x) = 0.

Definition 1. The r-jet of f : R→ R is the rth order Taylor polynomial

jrx0(x) = f(x0) + f ′(x0)x+ ...+

f (r)(x0)

r!xr

How can we extend this notion to smooth maps between arbitrary smooth manifolds? One idea

is to pick chart maps near x and f(x) and consider multivariate Taylor polynomials in these

chosen local coordinates. We wish to formalize this approach in a coordinate-free manner. We

begin by defining jets of smooth curves. Let M be a manifold and x ∈ M . By a smooth curve

through x we mean a smooth map f : R → M with f(0) = x. Two smooth curves f and g

through x are equivalent to order r if for every smooth function ψ : M → R, we have

jr0(ψ ◦ f) = jr0(ψ ◦ g)

Note, these expressions make sense, since ψ ◦ f and ψ ◦ g are smooth functions from R to itself.

6

Page 7: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

Definition 2. The r-jet of a smooth curve f through x is its equivalence class under the equiv-

alence relation defined above.

At the risk of abusing notation, we will denote the r-jet of f by jr0(f). Two smooth maps

f, g : M → N are equivalent to order r at x if for every smooth curve φ : R→M through x, we

have

jr0(f ◦ φ) = jr0(g ◦ φ)

Definition 3. The r-jet of a smooth map f : M → N at x is its equivalence class under the

equivalence relation defined above.

We will denote the r-jet of f : M → N at x by jrx(f). The set of all r-jets is denoted by

Jr(M,N). The set of all r-jets at x is Jrx(M,N). Finally, the set of all r-jets at x with f(x) = y

is Jrx(M,N)y.

Next, we show that jet composition makes sense.

Proposition 1. Suppose f, f ′ : M → N and g, g′ : N → O such that f(x) = y = f ′(x). If

jrx(f) = jrx(f ′) and jry(g) = jry(g′), then jrx(g ◦ f) = jrx(g′ ◦ f ′).

Proof. First note that if two curves a, b : R→M are rth order equivalent at x, then jrx(f ◦ a) =

jrx(f ◦b) for any smooth map f : M → N . Indeed, for every function ψ : N → R, the composition

ψ ◦ f is a function on M .

Take any curve φ : R → M with φ(0) = x. jrx(f) = jrx(f ′) implies jrxf ◦ φ = jrxf′ ◦ φ. And thus

the remark above gives jrx(g′ ◦ f ◦ φ) = jrx(g′ ◦ f ′ ◦ φ). On the other hand, jry(g) = jry(g′) yields

jrx(g ◦ f ◦ φ) = jrx(g′ ◦ f ◦ φ). Thus, jrx(g ◦ f ◦ φ) = jrx(g′ ◦ f ′ ◦ φ) and so jrx(g ◦ f) = jrx(g′ ◦ f ′) by

definition.

Thus, r-jets are the morphisms of a category with smooth manifolds as objects. The set of

invertible r-jets between manifolds M and N is denoted invJr(M,N). Note that an r-jet is

invertible if and only if the underlying 1-jet is invertible (an immediate consequence of the

inverse function theorem). Such a jet exists between manifolds M and N if and only if dimM =

dimN .

One can also consider jets of infinite order. Jets of finite order form a projective system

7

Page 8: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

J1(M,N)π21← J2(M,N)

π32← J3(M,N)

π43← ...

We define J∞(M,N) as its inverse limit. Thus, a jet of infinite order is an infinite sequence of

finite order jets j1, j2, j3, ... such that ji = πi+1i (ji+1) for all i ∈ N.

Next we define a special group of invertible jets.

Definition 4. The rth jet group in dimension m is the set Grm := invJr0 (Rm,Rm)0. Grm forms

a group under jet composition.

In fact, Grm is a Lie group. Its manifold structure is determined by the inclusion

Grm ⊂ {m-tuples of degree r polynomials in m variables} = Rm(m+rm )

Note that G1m is canonically identified with GLm(R). The kernel of the jet projection Grm →

GLm(R) is a (connected, simply connected) normal subgroup which we will denote by NGLrm.

Thus, we have a short exact sequence

1→ NGLrm → Grm → GLm(R)→ 1

which splits on the right through the obvious inclusion GLm(R) ↪→ Grm.

Remark 1. It follows from these observations and the splitting lemma for (non-abelian) groups

that Grm is the semidirect product NGLrmoGLm(R) where GLm(R) acts on NGLrm by adjunc-

tion. This decomposition will prove particularly useful in part 2.

The Lie algebra of Grm is the space of r-jets of vector fields on Rm that vanish at the origin, or,

equivalently, vector fields on Rm with polynomial coefficients of degree not exceeding r and no

constant terms. As a vector space

grm = {a1 + a2 + ...+ ar | ai ∈ Sym(⊗iRm,Rm

)}

The Lie algebra of NGLrm is the (nilpotent) ideal:

nglrm = {a2 + a3...+ ar | ai ∈ Sym(⊗iRm,Rm

)}

For a thorough treatment of jet groups and their Lie algebras, see [5, chap. 4].

8

Page 9: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

Finally, we define rth order velocities on M .

Definition 5. T rk := Jr0 (Rk,M) is the set of k-dimensional velocities of order r on M .

Picking charts on M , T rk is seen to be a manifold (in fact, a fibered manifold over M).

4 Natural Bundles

A natural bundle, in the sense of [5], generalizes tangent and cotangent bundles in differential ge-

ometry. These objects provide recipes for constructing fiber bundles over arbitrarym-dimensional

manifolds—recipes which respect the action of local diffeomorphisms. For example, the tangent

bundle M 7→ TM comes equipped with the pushforward map f 7→ f∗ which satisfies the chain

rule (f ◦ g)∗ = f∗ ◦ g∗ on compositions.

Let Manm be the category of m-dimensional manifolds, where morphisms are local diffeomor-

phisms (i.e. globally defined maps with locally defined inverses) and let Fib be the category of

fiber bundles, where morphisms are locally invertible bundle maps.

Definition 6. A natural bundle over m-dimensional manifolds is a (covariant) functor F :

Manm → Fib satisfying

1. (Prolongation) B ◦ F = IdManm , where B : Fib → Manm is the base functor. In other

words, FM is a bundle over M for every manifold M , and Ff covers f for every local

diffeomorphism f : M → N .

2. (Locality) If i : U → M is an inclusion of an embedded submanifold, then FU = π−1M (U)

and Fi is the inclusion of π−1M (U) into FM .

A natural vector bundle is a natural bundle with values in the subcategory of vector bundles.

Example 1. Perhaps the most important and ubiquitous examples of natural bundles in dif-

ferential geometry are the tensor bundles (i.e. tangent bundles, cotangent bundles, and their

various tensor products). The tangent bundle acts on morphisms via the pushforward map, and

the cotangent bundle acts on morphisms by pulling back along their (local) inverses. Higher

tensor bundles act via tensor products of these actions.

Example 2. For every manifold F , there is a (trivial) natural bundle ×F : M 7→ M × F . On

morphisms, ×F : f 7→ f × id.

9

Page 10: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

Example 3. The bundle of rth order k-velocities is a natural bundle. On objects, T rk : M 7→T rkM := Jr0 (Rk,M) and on morphisms, T rk : f 7→ (jr0σ 7→ jr0f ◦ jr0σ). Dually, one can consider

the natural bundle T r∗k of rth order k-covelocities. On objects, T r∗k : M 7→ T r∗k M := Jr(M,Rk)0,

and on morphisms, T r∗k : f 7→ (jrσ 7→ jrσ ◦ jr0f−1), where f−1 is constructed locally.

Example 4. The frame bundles of natural vector bundles are, again, natural vector bundles.

Let E be a natural vector bundle with fiber dimension k and Fr(E) its frame bundle. On objects,

Fr(E) : M 7→ Fr(EM) := qx∈MLin(Rk ∼→ ExM). On morphisms, Fr(E) : f 7→ (σ 7→ Ef ◦ σ).

The following proposition follows trivially from the definition of a natural bundle.

Proposition 2. Let f : M → N be a local diffeomorphism of m-dimensional manifolds and let

x ∈M . For any natural bundle F , Ff |FxM : FxM → Ff(x)N depends only on the germ of f at

x.

Proof. Suppose g : M → N is a local diffeomorphism that agrees with f on an open submanifold

U ⊂M containing x. Let i denote the inclusion of U into M so that

f ◦ i = g ◦ i

and thus by functoriality Ff ◦Fi = Fg ◦Fi. By locality, Fi is the inclusion of π−1M (U) into FM .

Hence, Ff and Fg agree on π−1M (U) and, in particular, π−1

M (x).

A natural bundle has order r if its action on diffeomorphisms factors through r-jets. Formally,

Definition 7. A natural bundle F : Manm → Fib has order r, 0 ≤ r < ∞, if for every pair

of local diffeomorphisms f, g : M → N and every point x ∈ M , the equality jrxf = jrxg implies

Ff |FxM = Fg|FxM and if r is the minimal such integer.

Example 5. The tangent bundle has order 1. In coordinates, the pushforward map is given by

the Jacobian matrix, which depends exclusively on first-order derivatives. Similarly, the cotan-

gent and tensor bundles also have order 1.

Example 6. The trivial bundle ×F defined in Example 2 has order 0. Indeed, the behavior of

the map f × id at a point depends only on the value of f at that point.

Since sums, products, and tensor products are functorial constructions, we can form sums, prod-

10

Page 11: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

ucts, and tensor products of arbitrary natural bundles. These will, again, be natural bundles.

Furthermore, the order of a sum, product, or tensor product of two natural bundles of orders r1

and r2 will be the maximum of these two integers.

A foundational result in the theory of natural bundles is the finiteness theorem, proved by Palais

and Terng in [1].

Finiteness Theorem for Natural Bundles. If F is a natural bundle over m-dimensional

manifolds of fiber dimension k (i.e. dimF0Rm = k), then F has finite order less than 2k+1.

Note that this theorem is a significant strengthening of Proposition 2.

Another foundational result, which we state below without proof, is the regularity theorem for

natural bundles, originally proved by Epstein and Thurston in [3] and later generalized to natural

bundles of infinite-dimensional fiber by Mikulski in [9] .

Regularity Theorem for Natural Bundles. Natural bundles are regular in the following

sense: if F is a natural bundle and f : P ×M → N is a smoothly parameterized system of local

diffeomorphisms fp : M → N , then F f : P ×FM → FN , defined by (F f)p = Ffp, is a smoothly

parameterized system of local bundle isomorphisms.

5 Natural Operators

A natural operator η : F G between natural bundles F and G over m-dimensional manifolds

is a rule transforming smooth sections of F into smooth sections of G, in a way that respects the

functorial actions of F and G on diffeomorphisms. Formally,

Definition 8. Let F and G be natural bundles over m-dimensional manifolds. A natural op-

erator η : F G is a family of smooth maps ηM : C∞(FM) → C∞(GM), M ∈ Manm

satisfying

1. (Naturality) For every section s ∈ C∞(FM) and diffeomorphism f : M → N we have

ηN (Ff ◦ s ◦ f−1) = Gf ◦ ηM (s) ◦ f−1

2. (Locality) For every section s ∈ C∞(FM) and open submanifold U ⊂M we have

ηU (s|U ) = ηM (s)|U

11

Page 12: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

3. (Regularity) If ψ : P ×M → FM is a smoothly parameterized system of smooth sections

of F , the map ψ : P ×M → GM , defined by ψp = ηM (ψp), is a smoothly parameterized

system of smooth sections of G.

If F and G are, in particular, vector bundles, C∞(FM) and C∞(GM) are vector spaces. If, in

this case, the maps ηM : C∞(FM)→ C∞(GM) are linear (resp., polynomial), we say that η is

a linear (resp., polynomial) natural operator.

Put differently, a natural operator is a local differential operator, whose local coordinate descrip-

tion is invariant under arbitrary changes of coordinates.

Definition 9. Suppose F1, ..., Fk, G are natural vector bundles over m-dimensional manifolds.

By a multilinear natural operator η : F1 × F2 × ... × Fkmultilin G we mean a natural operator

η : F1 ×F2 × ...×Fk G for which the maps ηM : C∞(F1M)×C∞(F2M)× ...×C∞(FkM)→C∞(GM) are multilinear over R.

Remark 2. It is important to distinguish between multilinear natural operators F1 ×F2 × ...×Fk

multilin G, as defined above, and linear natural operators F1 ⊗ F2 ⊗ ... ⊗ Fk G. It is

well known that for any manifold M , C∞(F1M ⊗ F2M ⊗ ... ⊗ FkM) is canonically isomorphic

to C∞(F1M)⊗C∞(M) C∞(F2M)⊗C∞(M) ...⊗C∞(M) C

∞(FkM). In other words, linear natural

operators F1 ⊗F2 ⊗ ...⊗Fk G are multilinear over functions. Such operators correspond pre-

cisely to the punctual (order 0) multilinear operators F1×F2×...×Fkmultilin G. See Example 12.

We present a few simple examples of natural operators below.

Example 7. The exterior derivative d : ∧kT ∗ ∧k+1T ∗ is a linear natural operator transform-

ing k-forms into k+ 1-forms for every nonnegative integer k. Indeed, the exterior derivative has

a coordinate-free description given by:

dω(v0, v1, ..., vk) =∑i

(−1)iviω(v0, ..., vi, ..., vk) +∑i<j

(−1)i+jω([vi, vj ], v0, ..., vi, ..., vj , ...vk)

In fact, we will soon see that the exterior derivative is the only natural operator (up to scalar

multiplication) between k-forms and k + 1-forms.

Example 8. The wedge product ∧ : ∧aT ∗ × ∧bT ∗ ∧a+bT ∗ is a bilinear natural operator. It

has a coordinate-free description is given by:

12

Page 13: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

ω1 ∧ ω2(v1, v2, ..., va+b) =∑

(a,b)−shuffles

ω1(vσ(1), vσ(2), ..., vσ(a))ω2(vσ(a+1), vσ(a+2), ..., vσ(b))

The wedge product of alternating multivector fields ∧ : ∧aT × ∧bT ∧a+bT also defines a

bilinear natural operator with an analogous coordinate-free description.

Example 9. The Lie bracket [ , ] : T×T T is a bilinear natural operator. Indeed, if we regard

vector fields on M as derivations of the algebra C∞(M), the Lie bracket admits the following

coordinate-free description:

[X,Y ] = X ◦ Y − Y ◦X

In fact, for any pair of integers a, b ≥ 0, there is a bilinear natural operator [ , ] : ∧aT × ∧bT ∧a+b−1T called the Schouten-Nijenhuis bracket, which generalizes the Lie bracket (if a=b=0,

the bracket is identically zero). The Schouten-Nijenhuis bracket also admits a coordinate-free

expression. The Schouten-Nijenhuis bracket will receive quite a bit of attention in part 2.

Example 10. For integers a, b ≥ 0, the contraction i : ∧aT ∗ × ∧bT ∧a−bT ∗ is a bilinear

natural operator (where ∧−nT ∗ is shorthand for ∧nT ). Its coordinate-free description is given

by:

i(ω,X)(v1, ..., va−b) = ω(X ∧ v1 ∧ v2 ∧ ... ∧ va−b)

Definition 10. A natural operator η : F G has order r, 0 ≤ r <∞, if for every manifold M ,

point x ∈M , and section s ∈ C∞(FM), the value of ηM (s) at x depends only on jrx(s), and if r

is the minimal such integer.

Example 11. The exterior derivative d : ∧kT ∗ ∧k+1T ∗ has order 1. In coordinates,

d(fdx1 ∧ dx2 ∧ ... ∧ dxk) =∑i

∂f

∂xidx1 ∧ dx2 ∧ ... ∧ dxk ∧ dxi

which depends exclusively on first-order partials.

Example 12. The wedge product ∧ : ∧aT ∗×∧bT ∗ ∧a+bT ∗ has order 0. The value of ω1∧ω2

at a point depends entirely on the values of ω1 and ω2 at that point. Similarly for the wedge

13

Page 14: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

product of alternating multivector fields. Natural operators of order 0 are punctual.

Example 13. The Schouten-Nijenhuis bracket [ , ] : ∧aT × ∧bT ∧a+b−1T has order 1. In

the Lie bracket case, this can be seen in local coordinates:

[Xi∂i, Yi∂i] =

(Xj∂jY

i − Y j∂jXi)∂i

The general case can then be reduced to the Lie bracket and the wedge product via the following

formula:

[X1∧X2∧ ...∧Xa, Y1∧Y2∧ ...∧Yb] =∑i,j

(−1)i+j [Xi, Yj ]X1∧ ...∧Xi∧ ...∧Xa∧Y1∧ ...∧ Yj∧ ...∧Yb

Example 14. The contraction i : ∧aT ∗×∧bT ∧a−bT ∗ is of course order 0. It is given locally

by the evaluation of a form at a multivector field.

For every fiber bundle π : E → M and natural number r, there is a fiber bundle Jr(π) → M

consisting of r-jets of smooth sections of π.

Definition 11. Let π : E → M be a fiber bundle and r a natural number. The rth order jet

bundle over π is the set {jrxs | x ∈ M, s ∈ C∞(E)} and is denoted by Jr(π) or, infrequently,

Jr(E) when the projection map is implied. The source projection jrxs 7→ x makes Jr(π) into a

fiber bundle over M .

Note that for every natural bundle F : Manm → Fib, the assignment M 7→ Jr(FM) defines a

covariant functor JrF : Manm → Fib. Indeed, for every local diffeomorphism f : M → N , there

is a smooth bundle map JrF (f) : Jr(FM)→ Jr(FN) given by JrF (f)(jrxs) = jrf(x)(Ff◦s◦f−1),

which respects compositions. It is easy to see that JrF satisfies both criteria in Definition 6 .

Thus, JrF is a natural bundle, which we call the rth order jet prolongation of F .

Proposition 3. Suppose F is a natural bundle of order k. Then JrF is a natural bundle of

order k + r.

Proof. For any local diffeomorphism f : M → N , we have

JrF (f)(jrxs) = jrf(x)(Ff ◦ s ◦ f−1)

14

Page 15: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

Since jrs(x)(Fjkxf) = Fjr+kx (f), it follows that JrF (f)(jrxs) depends only on the (k + r)-jet of f

at x.

Suppose F and G are natural bundles of order k and l, respectively. Every r-order natural

operator η : F G induces a natural transformation η : JrF → G defined by the formula

ηM (jrxs) = ηM (s)(x)

The smoothness of the components follows from the regularity of η. The naturality of η follows

from the naturality of η.

Proposition 4. The correspondence η 7→ η defines a bijection from natural r-order operators

F G onto natural transformations JrF → G.

Proof. For every natural transformation ξ : JrF → G define an operator ξ : F G by the

formula

ξM (s)(x) = ξM (jrxs)

Since the components of ξ are smooth (locally invertible) bundle maps, the maps x 7→ ξM (jrxs)

are smooth sections of GM . Since all categories in question are local, the induced operator is

local. Moreover, for every section s ∈ C∞(FM) and diffeomorphism f : M → N , we have, by

the naturality of ξ,

ξN (Ff ◦ s ◦ f−1)(f(x)) = ξN (jrf(x)(Ff ◦ s ◦ f−1))

= ξN (JrF (f)(jrxs))

= (ξN ◦ JrFf)(jrxs)

= (Gf ◦ ξM )(jrxs)

= Gf(ξM (s)(x))

= (Gf ◦ ξMs ◦ f−1)(f(x))

Hence, ξ is natural. To verify the regularity of ξ, consider a smoothly parameterized family of

smooth sections of F , ψ : P ×M → FM . ξ induces a map ψ : P ×M → GM given by

15

Page 16: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

ψ(p, x) = ξM (jrxψp)

Since ξ is a natural transformation, ξM is smooth. Hence, it suffices to show that the assignment

(p, x) 7→ jrxψp is smooth. In charts, jrxψp is just an rth order Taylor polynomial. In other words,

the map (p, x) 7→ jrxψp is, locally, a polynomial over derivatives of the smooth function ψ. Thus,

ψ is manifestly smooth.

We should, at this point, pause to remark that there is no general analogue of Palais-Terng

finiteness for natural operators. However, in many applications, the natural operators in question

are presumed to be linear. In these cases, finiteness often follows from the Invariant Tensor

Theorem (see section 7). In other situations, we will appeal to partial results, such as theorem

3 which we state without proof in section 6.

6 Invariant-Theoretic Reduction

The methods of invariant-theoretic reduction allow us to regard natural bundles over manifolds

as modules over jet groups and natural operators as module homomorphisms, thus reducing the

geometric problem of classifying natural bundles and operators to an essentially algebraic one.

We begin with a few preliminary observations.

Definition 12. The standard fiber of a natural bundle F is the fiber F0Rm over the origin.

Proposition 5. Let F be a natural bundle with standard fiber S. The fiber bundle FRm is

canonically isomorphic to the trivial bundle Rm × S.

Proof. For every x ∈ Rm, let tx : Rm → Rm denote the parallel shift tx(a) = x+ a. There is an

obvious bundle map ϕ : Rm × S → FRm given by

ϕ(x, s) = Ftx(s)

which is smooth by the regularity of F . In fact, ϕ is an isomorphism. Its inverse is given by

ϕ−1(v) = (π(v), F t−π(v)(v))

16

Page 17: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

Replacing the parallel shifts in the proof above with arbitrary chart maps, we get

Proposition 6. For every m-dimensional manifold M , the value FM is a locally trivial fiber

bundle with standard fiber S.

Proof. Let a ∈M and let φ : U → Rm be a chart map with φ(a) = 0. By the locality condition,

FU is π−1M U . Hence, the value Fφ is a diffeomorphism from π−1

M U onto FRm. Left composing

with the isomorphism in Proposition (5), we obtain a bundle isomorphism ϕ−1 ◦ Fφ : π−1M U →

Rm × S.

If F is a natural bundle of order less than or equal to r, there is a canonical (left) action of the

jet group Grm on the standard fiber S given by the formula

jr0(f)s = Ff(s)

for every jr0(f) ∈ Grm and s ∈ S. We call this the induced action of Grm. By the regularity of F ,

the induced action is smooth.

Theorem 1. Restricting to the standard fiber defines a bijective correspondence between (equiv-

alence classes of) natural bundles of order less than or equal to r and smooth left actions of Grm

on smooth manifolds.

Proof. We have already seen that for every natural bundle F of order not exceeding r, there is a

smooth left action of Grm on its standard fiber, S. We must exhibit a map in the other direction.

Let S be a manifold and ϕ : Grm × S → S a smooth left action. We begin by considering the

rth order frame bundle Grm(M) over an m-dimensional manifold M . Grm(M) is the set of r-jets

jrx(f) ⊂ Jrx(M,Rm)0 such that dfx maps TxM isomorphically onto Rm = T0Rm. As such, it is

an open submanifold of the bundle of rth order m-covelocities T r∗m (M) (see Example (3)). The

projection jrx(f) 7→ x turns Grm(M) into a fiber bundle over M . Moreover, there is a smooth

right action of Grm on Grm(M) given by

jrx(f)jr0(ϕ) = jrx(ϕ−1 ◦ f)

Clearly, this action is free and transitive on the fibers of Grm(M). Hence, Grm(M) is a principal

Grm-bundle over M . We then then take the product Grm(M)× S and let Grm act on the left via

jr0(ϕ)(jrx(f), s) = (jrx(f)jr0(ϕ−1), jr0(ϕ)s)

17

Page 18: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

denoting the orbit space of this action by S(M). We will quickly verify that S(M) is a natural

bundle over m-dimensional manifolds.

For any local diffeomorphism φ : M → N , there is an induced map Grmφ : Grm(M) → Grm(N),

defined by jrx(f) 7→ jrx(f◦φ−1), which is Grm-invariant by construction. Therefore, so is Grm×idS :

Grm(M) × S → Grm(N) × S, which induces a smooth bundle map Sφ : S(M) → S(N). The

naturality, locality, and regularity of S(M) follow trivially from the naturality, locality, and

regularity of Grm(M). Moreover, it is clear that the order of S(M) is less than or equal to r.

The standard fiber of S(M) is the set of equivalence classes {[id, s] | s ∈ S}, which we identify

canonically with S, and the canonical action of Grm on S0(Rm) is indeed jr0(ϕ)[id, s] = [id, jr0(ϕ)s].

On the other hand, if we start with a natural bundle F with standard fiber S, the functors

F : Manm → Fib and S : Manm → Fib are naturally equivalent.

The equivalence is defined as follows. For every manifold M and point x ∈ M , we pick a local

diffeomorphism ψ : Rm →M with ψ(0) = x. Then, for every y ∈ FxM , we put

ηM (y) = S0(ψ)[id, F0(ψ)−1(y)]

Note that these mappings are independent of ψ. Indeed, if φ is a another local diffeomorphism

φ : Rm →M with φ(0) = x, we have, for every s ∈ S,

[id, F0(φ−1 ◦ ψ)(s))] = S0(φ−1 ◦ ψ)[id, s]

and so by the functoriality of F and S,

S0(φ)[id, F0(φ−1)(s)] = S0(ψ)[id, F0(ψ−1)(s)]

Now, let g : M → N be a local diffeomorphism. For every x ∈ M , y ∈ FxM , and choice of

local diffeomorphism ψ : Rm → M with ψ(0) = x, the composition g ◦ ψ : Rm → N is a local

diffeomorphism with (g ◦ ψ)(0) = g(x) and

ηN (Fxg(y)) = S0(g ◦ ψ)[id, F0(g ◦ ψ)−1(Fxg(y))]

= S0(g)(S0(ψ)[id, F0(ψ)−1(y)])

= S0(g)(ηM (y))

Hence, η defines a natural isomorphism η : F∼→ S, which completes the proof.

18

Page 19: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

If F is, in particular, a natural vector bundle (i.e. F takes values in the subcategory of vector

bundles over m-dimensional manifolds with locally invertible, fiber-wise linear bundle maps),

then the standard fiber of F is a vector space and the action of Grm is linear. Conversely, every

Grm-module S gives rise to a natural vector bundle with standard fiber S and order not exceeding

r through the construction described above. Therefore,

Corollary 1. Restricting to the standard fiber defines a bijective correspondence between (equiv-

alence classes of) natural vector bundles of order less than or equal to r and Grm-modules.

Example 15. The tangent bundle T is a natural vector bundle of order 1 and standard fiber

Rm. G1m = GLm(R) acts on Rm canonically, i.e.

g • v = gv

The cotangent bundle T ∗ is a natural vector bundle of order 1 and standard fiber Rm∗. GLm(R)

acts on Rm∗ canonically, i.e.

g • v = vg−1

For every tensor bundle ⊗pT ⊗⊗qT ∗, the corresponding GLm(R)-module is the tensor product

⊗pRm ⊗⊗qRm∗, with the canonical linear action

g • (v1 ⊗ ...⊗ vp ⊗ w1 ⊗ ...⊗ wq) = gv1 ⊗ ...⊗ gvp ⊗ w1g−1 ⊗ ...⊗ wqg−1

Example 16. Suppose F is a natural bundle of order k and standard fiber S. Then the rth

order jet bundle Jr(F ) is a natural bundle of order k+r with standard fiber T rm(S) := Jr0 (Rm, S)

(see Example (3)). The action of the jet group Gr+km on T rm(S) is inherited directly from S, i.e.

g • jr0(f) = jr0(g • f)

In the special case where F is a vector bundle with, say, standard fiber V , T rm(V ) is canonically

identified as a vector space (and, indeed, a GLm(R)-module) with the direct sum

V ⊕ Rm∗ ⊗ V ⊕ S2Rm∗ ⊗ V ⊕ ...⊕ SrRm∗ ⊗ V

19

Page 20: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

To see this, select a basis {v1, ...vn} for V . Then, the set {∂|α|fi

∂xα }i,α forms a basis for T rm(V ).

The subspace Sj(Rm) ⊗ V coincides with the linear span of basis elements ∂|α|fi

∂xα with |α| = j

(all products are symmetric since mixed partials commute). This identification is independent

of our particular choice of basis.

Proposition 4 says that natural operators are essentially natural transformations between natural

bundles. Theorem (1) then says that natural bundles can be regarded as smooth Grm-spaces. It

is not surprising, therefore, that natural operators can be regarded as smooth, equivariant maps

between smooth Grm-spaces (or, in the vector bundle case, Grm-module homomorphisms).

Theorem 2. Let F andG be natural bundles of order k and l, respectively, and let r = max (k, l).

Then restriction defines a bijective correspondence between natural transformations F → G and

smooth, Grm-equivariant maps between their standard fibers.

Proof. Let S and Q be the standard fibers of F and G, respectively. Every natural transformation

η : F → G induces a smooth map ηRm |S : S → Q, which is Grm-equivariant, by the naturality of

η.

Conversely, given any Grm-equivariant map f : S → Q, we can define a natural transformation

η : F → G by pushing forward and backward along locally invertible chart maps. Specifically, for

every manifold M , point x ∈ M , and locally invertible chart map ψ : M → Rm with ψ(x) = 0,

we put

ηM |FxM := G(ψ−1)|Q ◦ f ◦ F (ψ)|FxM (2)

which is smooth, since f is smooth, and which is independent of ψ by the funtoriality of F and G

and the equivariance of f . The components ηM , formed by assembling these maps, are smooth,

since the chart maps {Fψ} and {Gψ} form smooth bundle atlases for FM and GM , respectively.

Moreover, for any local diffeomorphism g : M → N , the diagram

FM FN

GM GN

Fg

ηM ηN

Gg

is easily seen to commute. Indeed, for any y ∈ FxM , the composition ψ ◦ g−1 : N → Rm is a

locally invertible chart map with (ψ ◦ g−1)g(x) = 0 and

20

Page 21: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

ηN (Fg(y)) = G(g ◦ ψ−1)(f(F (ψ ◦ g−1)(Fg(y))))

= Gg(G(ψ−1)(f(F (ψ)(y)))

= Gg(ηM (y))

Thus, η : F → G is a natural transformation. Clearly, the restriction of η to S is simply f , the

equivariant map we started with. On the other hand, every natural transformation η : F → G

which restricts to f on S satisfies equation (2). Thus, if we start with a natural transformation

η : F → G, the natural transformation constructed via (2) from the restriction of η to S is

precisely η. Hence, the correspondence is a bijection, and we are done.

Combining this with the correspondence established in Proposition (4), we get

Corollary 2. Let F be a k-order natural bundle with standard fiber S, and let G be an l-

order natural bundle with standard fiber Q. Then restriction defines a bijective correspondence

between natural r-order operators F G and smooth Gpm-equivariant maps T rm(S)→ Q, where

p = max (k + r, l).

In the vector bundle case, we can apply the identification described in Example (16)

Corollary 3. Let F be a k-order natural vector bundle with standard fiber V , and let G be

an l-order natural vector bundle with standard fiber W . Then restriction defines a bijective

correspondence between natural r-order operators F G and Gpm-equivariant maps

V ⊕ Rm∗ ⊗ V ⊕ ...⊕ SrRm∗ ⊗ V →W

where p = max (k + r, l).

In summary, we have proved:

Natural r-order Operators

F G

Natural Transformations

JrF → G

Smooth Gpm-equivariant maps

T rm(S)→ Q

∼ ∼

Remark 3. Suppose F , G, and H are natural vector bundles (of order k, l, and n and with

standard fibers S, Q, and R) and F × G H is an r-order natural bilinear operator. Let p =

21

Page 22: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

max (r + k, r + l, n). The fibered product JrF×JrG has order max (r + k, r + l) and the induced

natural transformation JrF × JrG → H restricts to a bilinear Gpm-module homomorphism

T rm(S) × T rm(Q) → R. Conversely, a bilinear Gpm-module homomorphism T rm(S) × T rm(Q) → R

gives rise to a bilinear natural operator F × G H through the procedure described in the

proof of Theorem (2). In other words, there is a bijective correspondence between natural r-

order bilinear operators F×G H and linear Gpm-module homomorphisms T rm(S)⊗T rm(Q)→ R.

Similarly, natural multilinear operators F1 × F2 × ... × Fn G correspond bijectively to linear

module homomorphisms T rmS1 ⊗ T rmS2 ⊗ ...⊗ T rmSn → Q.

Let us reexamine some of our original examples of natural operators through an invariant-

theoretic lens.

Example 17. The first-order vector bundle ∧kT ∗ corresponds to the vector space ∧kRm∗, with

the canonical action of GLm(R). Thus, the jet bundle J1(∧kT ∗) corresponds to the G2m-module

T 1m(∧kRm∗) = ∧kRm∗ ⊕ Rm∗ ⊗ ∧kRm∗

The exterior derivative d : ∧kT ∗ ∧k+1T ∗ is then the linear G2m-module homomorphism

d : ∧kRm∗ ⊕ Rm∗ ⊗ ∧kRm∗ → ∧k+1Rm∗

dxi1 ∧ dxi2 ∧ ... ∧ dxik + dxj0 ⊗ dxj1 ∧ dxj2 ∧ ... ∧ dxjk 7→ dxj0 ∧ dxj1 ∧ ... ∧ dxjk

Example 18. The wedge product of differential forms ∧ : ∧aT ∗ ×∧bT ∗ ∧a+bT ∗ corresponds

to the GLm(R)-module homomorphism

∧ : ∧aRm∗ ⊗ ∧bRm∗ → ∧a+bRm∗

dxi1 ∧ dxi2 ∧ ... ∧ dxia ⊗ dxj1 ∧ dxj2 ∧ ... ∧ dxjb 7→ dxi1 ∧ ...dxia ∧ dxj1 ∧ ... ∧ dxjb

Similarly for the wedge product of alternating multivector fields.

Example 19. The Schouten-Nijenhuis bracket [ , ] : ∧aT ×∧bT ∧a+b−1T corresponds to the

G2m-module homomorphism

[ , ] : (∧aRm ⊕ Rm∗ ⊗ ∧aRm)⊗ (∧bRm ⊕ Rm∗ ⊗ ∧bRm)→ ∧a+b−1Rm

22

Page 23: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

given by

[ , ] : (∂i1 ∧ ... ∧ ∂ia + dxj0 ⊗ ∂j1 ∧ ... ∧ ∂ja)⊗ (∂k1 ∧ ... ∧ ∂kb + dxl0 ⊗ ∂l1 ∧ ... ∧ ∂lb)

7→a∑l=1

(−1)ldxl0(∂il)∂i1 ∧ ... ∧ ∂il ∧ ... ∧ ∂ia ∧ ∂l1 ∧ ... ∧ ∂lb

−(−1)(a−1)(b−1)b∑l=1

(−1)ldxj0(∂kl)∂j1 ∧ ... ∧ ∂ja ∧ ∂k1 ∧ ... ∧ ∂kl ∧ ... ∧ ∂kb−1

Example 20. The contraction i : ∧aT ∗ × ∧bT ∗ ∧a−bT ∗ (where ∧−nT ∗ = ∧nT ) corresponds

to the GLm(R)-homomorphism

i : ∧aRm∗ ⊗ ∧bRm → ∧a−bRm∗

dxi1 ∧ ... ∧ dxia ⊗ ∂j1 ∧ ... ∧ ∂jb 7→∑σ∈Sa

sgn(σ)dxσ(i1)(∂j1)...dxσ(ib)(∂jb)dxσ(ib+1) ∧ ... ∧ dxσ(ia)

as always, reinterpreting in the obvious way if a < b.

The following is a partial finiteness theorem for natural operators.

Theorem 3. Let F and G be natural bundles corresponding, via the identification described in

Theorem 1, to GLm(R)-submodules of the finite sums

⊕i

(⊗ai

Rm ⊗⊗bi

Rm∗)

and

⊕j

⊗a′j

Rm ⊗⊗b′j

Rm∗

respectively. Then every natural operator η : F G has finite order.

Proof. See Proposition 23.5 in [5].

23

Page 24: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

7 Classical Results

In this section, we state and prove a variety of classical results. Chief among these are the Ho-

mogenous Function Theorem and Weyl’s celebrated Invariant Tensor Theorem. These two results

are the most powerful tools in our arsenal for describing and classifying natural differential oper-

ators, and a large class of operators can be exhaustively classified using these two theorems alone.

Homogenous Function Theorem. Let V1, V2, ..., Vn,W be finite dimensional vector spaces

and f : V1 × ...× Vn →W a smooth function. Let a1, ..., an, b be real numbers with ai > 0 such

that

kbf(v1, ..., vn) = f(ka1v1, ..., kanvn) (3)

for every n-tuple (v1, ..., vn) and k > 0. Then f is a sum of polynomials of degree di in vi such

that

a1|d1|+ ...+ an|dn| = b

If no such non-negative integers d1, ..., dn exist, then f is identically 0.

Proof. First observe that if f satisfies (3) with b < 0, then f is identically 0. Indeed, if

f(x1, ..., xn) 6= 0 for some x1, ..., xn, then taking limits k → 0+, we get

limk→0+

kbf(x1, ..., xn) = limk→0+

f(ka1x1, ..., kanxn) = f(0, ..., 0)

But since b < 0, the limit on the left is improper.

Suppose now b ≥ 0. Assume, without loss of generality, that a1 = min(a1, ..., an), and set p

equal to the integer part of ba1

. Select a basis xji on each Vi. Differentiating (3) with respect to

xji , we get

kbfxji (v1, ..., vn) = kaifxji (ka1v1, ..., k

anvn)

Thus, the partial derivative fxji satisfies the homogeneity condition

kb−aifxji (v1, ..., vn) = fxji (ka1v1, ..., k

anvn)

24

Page 25: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

Iterating this argument, we find that every partial derivative of order p + 1 or greater satisfies

(3) with b < 0. Hence, by the argument above, all such partials are identically 0. It follows from

Taylor’s theorem that f is a polynomial of maximal order r. Suppose vt11 ...vtnn is a monomial of

degree |ti| in vi. Then,

(ka1v1)t1 ...(kanvn)tn = ka1|t1|+...+an|tn|vt11 ...vtnn

Since f satisfies (3) for all k > 0, it follows that a1|t1|+ ...+ an|tn| = b.

The Homogenous Function Theorem can often be used to demonstrate the linearity of certain

types of operators.

Example 21. Suppose η : ∧kT ∗ ∧k+1T ∗ is a natural operator and k > 1. By Theorem 3, η

is of finite order r. Thus, η corresponds to a Gr+1m -module homomorphism

f :

r⊕i=0

∧kRm∗ ⊗ Si(Rm∗)→ ∧k+1Rm∗

The jet group Gr+1m contains the subgroup GLm(R) = G1

m. The linear homotheties {t(Idm)|t ∈R∗}, in turn, form a subgroup of GLm(R). Naturality with respect to linear homotheties defines

a homogeneity condition

tk+1f(v0, v1, ..., vr) = f(tkv0, tk+1v1, ..., t

k+rvr) ∀t 6= 0

Therefore, by the Homogenous Function Theorem, f is a sum of polynomials of degree di in vi

such that

k|d0|+ (k + 1)|d1|+ ...+ (k + r)|dr| = k + 1

For k > 1, the only solution is given by |d1| = 1 and |di| = 0 ∀i 6= 1. Hence f is a linear function

in the second component ∧kRm∗ ⊗ Rm∗ and η is a linear first-order operator.

Every polynomial map f : V →W between finite-dimensional vector spaces decomposes uniquely

into homogenous components

f = f0 + f1 + ...+ fn

25

Page 26: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

Proposition 7. If V and W are finite-dimensional G-modules and the polynomial map f : V →W is G-equivariant, then each homogenous component is also G-equivariant.

Proof. For every g ∈ G, we have, by the linearity of the G-action

gf0 + ...+ gfn = f0g + ...+ fng

In a basis, multiplying by g corresponds to a linear change of coordinates. Thus, for each fi,

both gfi and fig are homogenous of degree i. Hence, gfi = fig for every g ∈ G.

Similarly, every polynomial map f : V1 × ... × Vp → W decomposes uniquely into multi-

homogenous components

f =∑

(d1,...,dp)

f(d1,...,dp)

where the multi-homogenous degree of a monomial xα11 ...xαnn in f is the n-tuple (|α1|, ..., |αn|).

Proposition 8. If V1, ..., Vp,W are finite-dimensional G-modules and the polynomial map f :

V1×...×Vp →W is G-equivariant, then each multi-homogenous component is also G-equivariant.

Proof. For every g ∈ G, we have, by the linearity of the G-action

∑(d1,...,dp)

gf(d1,...,dp) = gf = gf =∑

(d1,...,dp)

f(d1,...,dp)g

In a basis, multiplying by g corresponds to a linear change of coordinates. Thus, for each

f(d1,...,dp), both gf(d1,...,dp) and f(d1,...,dp)g are multi-homogenous of multi-homogenous degree

(d1, ..., dp). Hence, gf(d1,...,dp) = f(d1,...,dp)g for every (d1, ..., dp).

Let V be a finite-dimensional G-module. The induced action of G on the dual space V ∗ is given

by

(gw)(v) = w(g−1v) ∀w ∈ V ∗, v ∈ V, g ∈ G

The tensor product V ⊗W of G-modules is a G-module, where

g(v ⊗ w) = gv ⊗ gw

26

Page 27: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

In particular, every tensor product ⊗pV ⊗⊗qV ∗ is a canonical GL(V )-module.

Definition 13. A tensor x ∈ ⊗pV ⊗ ⊗qV ∗ is invariant if it is a fixed point of the canonical

GL(V ) action.

There is a natural isomorphism

⊗pV ⊗⊗qV ∗ ∼= Lin (⊗qV,⊗pV )

given by

v1 ⊗ ...⊗ vp ⊗ w1 ⊗ ...⊗ vq 7→ (x1 ⊗ ...⊗ xq 7→ w1(x1)...wq(xq)v1 ⊗ ...⊗ vp)

If, in particular, p = q, there is an obvious collection of invariant tensors given by permuations of

indices. Explicity, for every σ ∈ Sp, consider the tensor Iσ identified with the linear map

Iσ : x1 ⊗ ...⊗ xp 7→ xσ(1) ⊗ ...⊗ xσ(p)

Clearly, Iσ is GL(V )-invariant. Invariant tensors defined in this way are called elementary in-

variant tensors.

Invariant Tensor Theorem. Let I denote the subspace of invariant tensors in ⊗pV ⊗⊗qV ∗.Then

1. If p 6= q, I = 0.

2. If p = q, I is generated by the elementary invariant tensors {Iσ}σ∈Sp .

3. If p = q and dimV ≥ p, the elementary invariant tensors {Iσ}σ∈Sp are linearly independent

and, thus, form a basis for I.

Proof. We closely follow the proof of proposition 24.4 in [5].

For 1, consider the linear homothety k(Id) for k 6= 0. For any x ∈ I, invariance with respect to

k(Id) implies kp−qx = x. Thus, if p 6= q, we have no choice but x = 0.

The only interesting case occurs when p = q. Let X = ei1...ipj1...jp

be an element of I and put

m = dimV . Invariance means

ai1...ipk1...kp

ek1...kpl1...lp

= ei1...ipj1...jp

aj1...jpl1...lp

27

Page 28: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

for every matrix aij ∈ GLm(R). Equivalently,

aj1...jpk1...kp

δ(i1, j1)...δ(ip, jp)ek1...kpl1...lp

= ei1...ipj1...jp

δ(k1, l1)...δ(kp, lp)aj1...jpk1...kp

Since this holds for every aij ∈ GLm(R), we can compare coefficients in each monomial in aij to

obtain

∑σ∈Sp

δ(i1, jσ(1))...δ(ip, jσ(p))ekσ(1)...kσ(p)l1...lp

=∑σ∈Sp

δ(kσ(1), l1)...δ(kσ(p), lp)ei1...ipjσ(1)jσ(p)

(4)

If m ≥ p, put cσ = e1...pσ(1)...σ(p) and i1 = 1 = j1, ..., ip = p = jp so that the only non-zero term on

the left-hand side of equation (4) is given by σ = 1. Hence

ek1...kpl1...lp

=∑σ∈Sp

cσδ(kσ(1), l1)...δ(kσ(p), lp) (5)

which establishes 1 in the special case m ≥ p. Furthermore, since the coefficients cσ in (4) are

uniquely determined by ei1...ipj1...jp

, we have also proved 2.

If, on the other hand, m < p, a subtler argument is needed. In this case, the coefficients cσ in

(5) are not uniquely determined, since for m < p, the system of m2p equations in p! variables dσ

∑σ∈Sp

δ(i1, jσ(1))...δ(ip, jσ(p))dσ = 0 (6)

has nontrivial solutions. For example, the alternation

∑σ∈Sp

sgn(σ)δ(i1, jσ(1))...δ(ip, jσ(p))

is 0, by the pigeonhole principle. Select a basis dασ , α = 1, ..., t for the space of solutions to (6)

and consider the equations in dσ

∑σ∈Sp

dασdσ = 0 α = 1, ..., t (7)

We claim the system (6) and (7) has rank p! (i.e. full column rank). To this end, it suffices to

show that the only solution to equations (6) and (7) is trivial. If dσ satisfy (6), there are kα such

that, for every σ ∈ Sp,

28

Page 29: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

dσ =

t∑α=1

kαdασ (8)

If dσ also satisfy (7), we have

t∑α=1

kα∑σ∈Sp

(dασdσ) = 0 (9)

Rewriting the equation above and applying equation (8), we obtain

0 =

t∑α=1

kα∑σ∈Sp

(dασdσ) =∑σ∈Sp

t∑α=1

kαdασ =

∑σ∈Sp

d2σ (10)

It follows that all dσ are 0. Thus, we have the following lemma:

Lemma 1. Suppose p! tensors Xσ ∈ ⊗pRm ⊗⊗qRm∗ satisfy the equations

∑σ∈Sp

δ(i1, jσ(1))...δ(ip, jσ(p))Xσ =∑σ∈Sp

ci1...ipjσ(1)...jσ(p)

Iσ (11)

with real coefficients ci1...ipjσ(1)...jσ(p)

and

∑σ∈Sp

dασXσ = 0 α = 1, ..., t (12)

with dασ as above. Then every Xσ is a linear combination of {Iσ}.

Indeed, since the system (6) and (7) has full rank p! and the equations in (7) are linearly

independent, there is a subsystem (7’) of (7) such that the system (6) and (7’) is invertible. Let

(11’) denote the subsystem of (11) corresponding to (7’) so that the system (11’) and (12) is also

invertible. Applying Cramer’s rule, we see that each Xσ must be a linear combination of the

right-hand sides of equations (11) and (12), and hence, a linear combination of {Iσ}.

With this lemma in hand, our proof quickly falls into place. Let X be an element of I, and for

every σ ∈ Sp, let Xσ be the result of applying σ to X. Then, equation (4) translates to

∑σ∈Sp

δ(i1, jσ(1))...δ(ip, jσ(p))Xσ =∑σ∈Sp

δ(kσ(1), l1)...δ(kσ(p), lp)Iσ (13)

29

Page 30: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

Since dασ are solutions to (6), the tensor

∑σ∈Sp

δ(i1, jσ(1))...δ(ip, jσ(p))dασ

is 0. Contracting with X we obtain

∑σ∈Sp

dασXσ = 0 (14)

Applying our lemma to equations (13) and (14), we conclude that each Xσ, and, therefore X in

particular, is a linear combination of {Iσ}.

We can apply the Homogenous Function Theorem and Invariant Tensor Theorem in tandem to

deduce all GL(V )-equivariant maps between symmetric and alternating powers of V .

Let Sym : ⊗pV → Sp(V ) and Alt : ⊗p → ∧p(V ) be the symmetrization and alternation maps,

respectively. Consider also the inclusions Sp(V ) ↪→ ⊗pV and ∧p(V ) ↪→ ⊗p(V ) which we denote

in both cases by i. Note that all four maps are linear and GL(V )-equivariant. Let f : Sp(V )→Sp(V ) be a smooth GL(V )-equivariant map and put f := i◦f ◦Sym. Since Sym : ⊗pV → Sp(V )

is a left inverse for i : Sp(V ) ↪→ ⊗pV , the following diagram commutes

Sp(V ) Sp(V )

⊗p(V ) ⊗p(V )

f

i

f

Sym

Since the maps Sym, i, and f are smooth and GL(V )-equivariant, so is f . Invariance with

respect to homotheties, as in example 21, yields an homogeneity condition

kp(f(x)) = f(kpx) ∀k 6= 0

which, by application of the Homogenous Function Theorem, implies that f is linear. Therefore,

via the obvious identification ⊗p(V )⊗⊗p(V ∗) ∼= Lin(⊗pV,⊗pV ), f corresponds to an invariant

tensor in the vector space ⊗p(V )⊗⊗p(V ∗). It follows from the Invariant Tensor Theorem that

f is a linear combination of elementary invariant tensors {Iσ}σ∈Sp . Take some Iσ and consider

the composition Sym ◦ Iσ ◦ i. We have, for every v1 ⊗ ...⊗ vp ∈ Sp(V ),

30

Page 31: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

Sym ◦ Iσ ◦ i(v1 ⊗ ...⊗ vp) = Sym ◦ Iσ(v1 ⊗ ...⊗ vp)

= Sym(v1 ⊗ ...⊗ vp)

= v1 ⊗ ...⊗ vp

Hence f is a constant multiple of the identity. Reproducing this reasoning with different types

of smooth, equivariant maps of variously alternating and symmetric domain and codomain, we

obtain the following classification:

Proposition 9. Every smooth, GL(V )-equivariant map

1. Sp(V )→ Sp(V ) is a constant multiple of the identity

2. ∧p(V )→ ∧p(V ) is a constant multiple of the identity

3. Sp(V )→ ⊗p(V ) is a constant multiple of the inclusion

4. ∧p(V )→ ⊗p(V ) is a constant multiple of the inclusion

5. ⊗p(V )→ Sp(V ) is a constant multiple of the symmetrization

6. ⊗p(V )→ ∧p(V ) is a constant multiple of the alternation

7. Sp(V )→ ∧p(V ) is identically 0

8. ∧p(V )→ Sp(V ) is identically 0

The final two claims can be strengthened considerably.

Proposition 10. Let M be a GL(V )-submodule of ⊗qV such that the inclusion M ↪→ ⊗qVadmits a linear GL(V )-equivariant left inverse (standard examples include ∧q(V ), Sq(V ) and

∧a(V )⊗ Sq−a(V )). Then, for any integer p > 1, every smooth, GL(V )-equivariant map

1. Sp(V )⊗M → ∧r(V ) is identically 0

2. ∧p(V )⊗M → Sr(V ) is identically 0

Proof. Let f : Sp(V ) ⊗M → ∧r(V ) be a smooth, GL(V )-equivariant map. Acting by homoth-

eties, we obtain the homogeneity condition

krf(x) = f(kp+qx) ∀k 6= 0

31

Page 32: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

By the Homogenous Function Theorem, f is identically 0 if (p+ q) - r. Otherwise, (p+ q)n = r

and f is a linear map

f : Sn (Sp(V )⊗M)→ ∧r(V )

Let φ : ⊗q(V )→M be a linear GL(V )-equivariant left inverse of the inclusion M ↪→ ⊗qV , which

is guaranteed by assumption, and consider the map

Φ := Sn(Sym⊗ φ) : ⊗r(V )→ Sn (Sp(V )⊗M)

Let f := i ◦ f ◦ Φ : ⊗r(V )→ ⊗r(V ) so that the following diagram commutes

Sn (Sp(V )⊗M) ∧r(V )

⊗r(V ) ⊗r(V )

f

i

f

Alt

As a composition of linear, GL(V )-equivariant maps, f is linear and GL(V )-equivariant. Hence,

by the Invariant Tensor Theorem, f is a linear combination of elementary invariant tensors. Take

one such tensor Iσ. Since p > 1 and n ≥ 1, for every v1 ⊗ ...⊗ vr ∈ Sn (Sp(V )⊗M), the tensor

Iσ(v1 ⊗ ... ⊗ vr) is symmetric in at least two components. Hence, Alt(Iσ(v1 ⊗ ... ⊗ vr) = 0, as

desired. The proof of 2 is analogous.

8 Natural Operators on Differential Forms

In this section, we apply invariant-theoretic reduction and the results of section 7 to classify all

natural operators of type

∧p1T ∗ × ...× ∧pnT ∗ ∧qT ∗

on m-dimensional manifolds. For the following lemma, let � denote the symmetric tensor prod-

uct.

Lemma 2. Let f :⊙n

i=1 ∧pi(V )⊗ Sqi(V )→ ∧r(V ) be a smooth, GL(V )-equivariant map with∑ni=1 pi+qi = r. Then f = 0 if any qi > 1. Otherwise, f is a constant multiple of the alternation.

32

Page 33: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

Proof. Our proof is quite similar to proposition 10. Acting by homotheties, we obtain the

homogeneity condition

krf(x) = f(krx) ∀k 6= 0

Thus, by the Homogenous Function Theorem, f is necessarily linear.

The inclusion⊙n

i=1 ∧pi(V )⊗ Sqi(V ) ↪→ ⊗rV admits an equivariant left inverse

Φ :=

n⊙i=1

Alt⊗ Sym : ⊗r(V )→n⊙i=1

∧pi(V )⊗ Sqi(V )

Put f := i ◦ f ◦ Φ so that the following diagram commutes

⊙ni=1 ∧pi(V )⊗ Sqi(V ) ∧r(V )

⊗r(V ) ⊗r(V )

f

i

f

Alt

As a composition of linear, GL(V )-equivariant maps, f is linear and GL(V )-equivariant. Hence,

by the Invariant Tensor Theorem, f is a linear combination of elementary invariant tensors.

Take one such tensor Iσ. If qi > 1, then every v1 ⊗ ...⊗ vr ∈⊙n

i=1 ∧pi(V )⊗ Sqi(V ), and hence

Iσ(v1 ⊗ ...⊗ vr), is symmetric in at least two components. Hence, Alt(Iσ(v1 ⊗ ...⊗ vr)) = 0, as

desired. Otherwise,

Alt(Iσ(v1 ⊗ ...⊗ vr)) = Alt(vσ(1) ⊗ ...⊗ vσ(r))

=∑τ∈Sr

sgn(τ)vτσ(1) ⊗ ...⊗ vτσ(r)

= sgn(σ)∑τσ∈Sr

sgn(τσ)vτσ(1) ⊗ ...⊗ vτσ(r)

= sgn(σ)Alt(v1 ⊗ ...⊗ vr)

It follows that f is a constant multiple of Alt.

Let Natformsm (p1, ...pn, q) denote the space of natural operators of type ∧p1T ∗×...×∧pnT ∗ ∧qT ∗

over m-dimensional manifolds. We prove that these spaces are generated (in an operadic sense)

by the wedge product and the exterior derivative.

33

Page 34: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

Theorem 4. For all nonnegative integers p1, ..., pn, q, the vector space Natformsm (p1, ..., pn, q) is

spanned by natural operators of form

η : (ω1, ..., ωn) 7→ (∧α1ω1) ∧(∧β1dω1

)∧ ... ∧ (∧αnωn) ∧

(∧βndωn

)where αi and βi are nonnegative integers satisfying

∑ni=1 αipi + βi(pi + 1) = q. If no such αi, βi

exist, f is identically 0.

Proof. Let η : ∧p1T ∗ × ... × ∧pnT ∗ ∧qT ∗ be a natural operator. By theorem 3, η has finite

order r. By theorem 2, η can therefore be identified with a smooth Gr+1m -equivariant map

f :⊕

0≤i≤r1≤j≤n

SiRm∗ ∧pj Rm∗ → ∧qRm∗

Acting by homotheties { 1k id} yields the homogeneity condition

kqf(xi,j) = f(ki+pjxi,j)

By the Homogenous Function Theorem, f decomposes into multi-homogenous components f{di,j}

of degree di,j in SiRm∗ ∧pj Rm∗ with degrees di,j satisfying

q =∑

0≤i≤r1≤j≤n

|di,j |(i+ pj)

If no such di,j exist, f is identically 0. By proposition 8, each multi-homogenous component

f{di,j} is Gr+1m -equivariant. Thus, each f{di,j} corresponds to a linear, Gr+1

m -equivariant map

f{di,j} :⊙

0≤i≤r1≤j≤n

Sdi,j(SiRm∗ ⊗ ∧pjRm∗

)→ ∧qRm∗

Applying lemma 2, we deduce that if di,j > 0 for any i > 1, then f{di,j} is identically 0. Hence,

the only non-zero components are given by linear, Gr+1m -equivariant maps

f{di,j} :

n⊙i=1

Sαi(∧piRm∗)� Sβi(Rm∗ ⊗ ∧piRm∗)→ ∧qRm∗ (15)

34

Page 35: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

Since GLm(R) = G1m ⊆ Gr+1

m , f{di,j} is in particular GLm(R)-equivariant. Moreover, both

domain and codomain of (15) are functorial over G-modules. The canonical identification Rm ∼=Rm∗ is an isomorphism of GLm(R)-modules and therefore induces GLm(R)-module isomorphisms

n⊙i=1

Sαi(∧piRm)� Sβi(Rm ⊗ ∧piRm) ∼=n⊙i=1

Sαi(∧piRm∗)� Sβi(Rm∗ ⊗ ∧piRm∗)

∧qRm ∼= ∧qRm∗

Applying these isomorphisms to (15), f{di,j} can be identified with a linear, GLm(R)-equivariant

map

˜f{di,j} :

n⊙i=1

Sαi(∧piRm)� Sβi(Rm ⊗ ∧piRm)→ ∧qRm

replacing Rm with its dual Rm∗. By lemma 2, ˜f{di,j} ,and hence f{di,j}, is a constant multiple of

the alternation. Let us examine how the alternation behaves on different factors of the domain.

Factors like ∧piRm∗ correspond (invariant-theoretically) to 0-jets of differential forms. Since

these factors are already alternating, the alternation behaves like the identity. Factors like

Rm∗⊗∧piRm∗ correspond to 1-jets of differential forms. The alternation maps the 1-jet xi0dxi1∧... ∧ dxip to dxi0 ∧ dxi1 ∧ ... ∧ dxip . Therefore, on these factors the alternation behaves like the

derivative. The multi-homogenous component f{di,j} therefore computes a (scalar multiple of a)

wedge product of pi-forms and their derivatives:

f{di,j}(ω1, ...ωn) = c (∧α1ω1) ∧(∧β1dω1

)∧ ... ∧ (∧αnωn) ∧

(∧βndωn

)c ∈ R

Thus, f is a linear combination of wedge products of forms and their derivatives. Conversely, all

such operators are natural, since both d and ∧ are natural (see examples 7 and 8). So we are

done.

In some cases, the operator

η : (ω1, ..., ωn) 7→ (∧α1ω1) ∧(∧β1dω1

)∧ ... ∧ (∧αnωn) ∧

(∧βndωn

)is identically 0. If |ω| is odd, the skew-symmetry of the wedge product forces ω ∧ ω = 0. If

|ω1| + |ω2| > m, ω1 ∧ ω2 = 0. However, by removing these degeneracies, we obtain a basis for

35

Page 36: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

Natm(p1, ..., pn, q).

Proposition 11. Consider the set S of non-zero operators

η : (ω1, ..., ωn) 7→ (∧α1ω1) ∧(∧β1dω1

)∧ ... ∧ (∧αnωn) ∧

(∧βndωn

)for nonnegative αi, βi. The elements of S are linearly independent.

Proof. Denote the operator η : (ω1, ..., ωn) 7→ (∧α1ω1)∧(∧β1dω1

)∧ ...∧ (∧αnωn)∧

(∧βndωn

)by

ηαi,βi . All operators are regarded invariant-theoretically (i.e. as maps∏i T

10 ∧piRm∗ → ∧qRm∗).

Consider a nontrivial linear dependence relation R on some collection of distinct non-zero ηαji ,βji:

R =

t∑j=1

cjηαji ,βji≡ 0

Since ηα1i ,β

1i

is non-zero, there are 1-jets ωi = dxji1 ∧ ... ∧ dxjipi + xki0dxki1 ∧ ... ∧ dxkipi such that

ηα1i ,β

1i(ω1, ..., ωn) 6= 0. For every i and r, s ∈ R, we put

ωi(r, s) = rdxji1 ∧ ... ∧ dxjipi + sxki0dxki1 ∧ ... ∧ dxkipi

Consider the function

f : R2n → ∧qRm∗

(r1, s1, ..., rn, sn) 7→ R(ω1(r1, s1), ..., ωn(rn, sn))

Since d is an R-linear function, f is a polynomial in ri, si. Since ηα1i ,β

1i(ω1, ..., ωn) is non-zero,

it can be completed to a basis. Let π : ∧qRm∗ → R denote the projection onto the linear

subspace Span(ηα1

i ,β1i(ω1, ..., ωn)

)induced by this basis, and let f1 = π ◦ f . By plugging the

forms ω1(r1, s1), ..., ωn(rn, sn) into R and applying the linearity of d, we see that the coefficient

of∏ni=1 r

αii s

βii in f1 is c1, which is non-zero by assumption. Hence, f is non-constant, and

therefore non-zero, a contradiction.

Consider the category ComDga of differential (cohomologically) graded associative supercom-

mutative algebras (commutative dga-algebras, for short) and the forgetful functor F : ComDga→VecZ mapping every commutative dga-algebra to the underlying graded vector space. Consider

36

Page 37: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

the functor P : VecZ → ComDga left adjoint to F . P exists and is unique up to natural

isomorphism. For any graded vector space V , we say that P (V ) is the free commutative dga-

algebra generated by V . As a vector space, P (V ) is the free graded associative supercommutative

algebra generated by the graded vector space V ⊕ dV . Note, this is distinct from the notion of

a quasi-free or semi-free dga-algebra, which is free as an associative graded algebra.

Consider the vector space Natformsm spanned by all natural operators of type ∧p1T ∗×...×∧pnT ∗

∧qT ∗ on m-dimensional manifolds, with no restrictions on n, p1, ..., pn, or q. This vector space

is Z-graded by q (as a matter of convention, Natm(p1, ..., pn, q) = 0 for any q < 0).The wedge

product and exterior derivative turn Natformsm into a commutative dga-algebra.

Corollary 4. For every q ∈ N, consider the identity operator idq : ∧qT ∗ ∧qT ∗. Let 〈idq〉 be

the 1-dimensional vector space spanned by idq, and let V be the graded vector space⊕

q∈N〈idq〉.The free commutative dga-algebra P (V ) contains an ideal I>m spanned by elements of degree

exceeding m, the dimension of the base manifolds in question. Natformsm is isomorphic to the

quotient P (V )/I>m.

Proof. Let V =⊕

q∈N〈idq〉. As a vector space, P (V ) is spanned by finite words constructed

from the formal symbols {idq}, a differential d, and an associative, supercommutative degree-0

product ∧, subject to the Leibniz rule:

d(a ∧ b) = da ∧ b+ (−1)|a|a ∧ db

By the universal property, the assignment idq 7→ idq extends to a map of commutative dgas

ϕ : P (V )→ Natformsm , which is surjective by theorem 4. Let x ∈ P (V ) be a homogenous element

of degree q > m. Since ϕ preserves degree, ϕ(x) is a natural operator mapping into smooth

sections of ∧qT ∗. However, over m-dimensional manifolds, the natural bundle ∧qT ∗ is the trivial

bundle M × {0}. Thus, ϕ(x) = 0. It follows that I>m ⊆ kerϕ.

By induction, we see that the non-zero elements

(∧α1idp1) ∧(∧β1d(idp1)

)∧ ... ∧ (∧αnidpn) ∧

(∧βnd(idpn)

)form a basis for P (V ). Suppose x ∈ kerϕ is a linear combination of elements x1, ..., xl as

above. The equation ϕ(c1x1 + ... + clxl) = 0 imposes a nontrivial linear dependence on the

non-zero elements of {ϕ(x1), ..., ϕ(xl)}. But, by proposition 11, these elements are linearly

independent. So, indeed kerϕ ⊆ I>m. Hence, kerϕ = I>m and ϕ descends to an isomorphism

ϕ : P (V )/I>m∼→ Natforms

m .

37

Page 38: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

Part II

Natural Operators on Alternating

Multivector Fields

There is a sense in which we got ‘lucky’ in our proof of theorem 4. To classify natural operators on

differential forms it was enough to check equivariance with respect to linear diffeomorphisms (i.e.

elements of the proper subgroup GLm(R) ⊂ Gr+1m ). We are not so lucky in the case of alternating

multivector fields. Recall, from section 3, the group isomorphism Gr+1m∼= NGLr+1

m o GLm(R).

Since the subgroups NGLr+1m and GLm(R) generate Gr+1

m , an operation is natural if and only

if the corresponding map between the standard fibers is equivariant with respect to GLm(R)

and NGLr+1M . In the case of differential forms, the action of NGLr+1

m on GLm(R)-equivariant

operators is manifestly trivial. This is not the case with alternating multivector fields, as demon-

strated below.

Example 22. Consider the bilinear ‘half-bracket’ of vector fields. In local coordinates,

[X,Y ] 12

:=∑i,j

(Xj∂jYi)∂i

For any g ∈ GLm(R), we have (gX)j =∑k g

jkX

k and ∂j(gY )i =∑r,s ∂sY

r(g−1)sjgir. Hence,

38

Page 39: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

[gX, gY ] 12

=∑i,j

((gX)j∂j(gY )i)∂i

=∑

i,j,k,r,s

gjkXk∂sY

r(g−1)sjgir∂i

=∑i,k,r,s

Xk∂sYrgir

∑j

(g−1)sjgjk

∂i

=∑i,k,r,s

Xk∂sYrgirδ(s, k)∂i

=∑i,k,r

Xk∂kYrgir∂i

= g[X,Y ] 12

Thus, [X,Y ] 12

is GLm(R)-equivariant. It is not, however, a natural operator. Consider the

invertible 2-jet ϕ : xi 7→ xi + x2i regarded as an element of NGL2

m. Let X =∑i ∂i so that

ϕX =∑i(1 + 2xi)∂i. Clearly, ϕ[X,X] 1

2= ϕ0 = 0 while on the other hand

[ϕX,ϕX] 12

=∑i,j

(ϕX)j∂j(ϕX)i∂i

=∑i

2∂i

= 2X

To classify natural operators on alternating multivector fields, we will adopt the following general

strategy, inspired by the work of Markl [10], [11]:

1. Apply the Invariant Tensor Theorem to model GLm(R)-equivariant operators as linear

combinations of a particular type of multigraph.

2. Define a linear map ∂ on multigraphs that corresponds to the tangent action of ngl∞m on

GLm(R)-equivariant operators.

3. Define a (Lie-admissable) product on multigraphs which behaves suitably with respect to

the ngl∞m -action.

4. Draw analogies to the shuffle coproduct on the tensor algebra to deduce the structure of

ker ∂.

39

Page 40: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

9 Twisted Algebras in VecZ

We begin by establishing some of the basic theory of symmetric collections and twisted algebras

over an operad. We deviate slightly from the standard formalism (see, for example, [12]) in that

all vector spaces are graded.

Definition 14. Consider the categories

• FinSet of finite sets with bijections

• VecZ of real, Z-graded vector spaces with linear degree-preserving maps.

Definition 15. A symmetric collection is a functor F : FinSet → VecZ. We write SymCol

for the category of symmetric collections with natural transformations as morphisms.

Thus, a symmetric collection is a family of graded vector spaces {F [S]}, one for each finite set

S, together with linear, degree-preserving maps {F (f) : F [S] → F [S′]}, one for each bijection

f : S → S′ such that F (idS) = idF (S) and F (f ◦ g) = F (f) ◦ F (g) for every pair (f, g) of

composable bijections. A morphism η : F → G of symmetric collections is a family of linear,

degree-preserving maps {ηS : F [S]→ G[S]}, one for each finite set S, such that for every bijection

f : S → S′, the following diagram commutes:

F [S] G[S]

F [S′] G[S′]

ηS

Ff Gf

ηS′

A subobject of F in the category of symmetric collections (a ‘subcollection,’ for lack of a better

term) is a symmetric collection G such that G[S] is contained (as a graded vector space) in F [S],

for every finite set S. Quotient objects are defined pointwise, i.e. (F/G)[S] = F [S]/G[S]. The

classical tensor product of representations has the following analogue in the category of symmet-

ric collections:

Definition 16. For symmetric collections F,G ∈ SymCol, let

(F ⊗G)[S] :=⊕

UtV=S

F [U ]⊗G[V ]

where the tensor product on the right is the ordinary tensor product of graded vector spaces. A

bijection f : S → S′ induces linear isomorphisms F [U ] ∼= F [U ′] and G[V ] ∼= G[V ′] for U ′ = f [U ]

and V ′ = f [V ], and therefore a linear isomorphism (F ⊗G)[S] ∼= (F ⊗G)[S′].

40

Page 41: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

Proposition 12. The tensor product of symmetric collections defines a monoidal structure on

SymCol.

Proof. The unit is given by the symmetric collection

1[S] =

{R if S = ∅0 otherwise

Associativity follows from the identity

(F ⊗ (G⊗H))[S] =⊕

UtV tW=S

F [U ]⊗G[V ]⊗H[W ] = ((F ⊗G)⊗H)[S]

we leave the details to the reader.

For every pair of symmetric collections F,G ∈ SymCol, we define the braiding map β(F,G) :

F ⊗G→ G⊗ F by

β(F,G)S : x⊗ y 7→ (−1)|x||y|y ⊗ x

This is the usual switch map in the category of ungraded vector spaces with the standard Koszul

signs. It is easy to see that the braiding maps turn SymCol into a symmetric monoidal category.

Definition 17. An algebra in the category SymCol is a symmetric collection F together with

a morphism × : F ⊗ F → F . Classically, an algebra in the symmetric monodical category of

symmetric collections is called a twisted algebra, a convention we will continue. The morphism

× : F ⊗ F → F is called multiplication on F .

Definition 18. A twisted algebra F with multiplication given by × : F ⊗ F → F is associative

if the following diagram commutes:

F ⊗ F ⊗ F F ⊗ F

F ⊗ F F

×⊗1

1⊗× ×

×

Furthermore, with braiding maps in hand, we can easily define a sensible notion of (super)commutativity.

41

Page 42: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

Definition 19. For any symmetric collection F , let S2(F ) denote the subcollection of F ⊗ Fgiven on each finite set U by

S2(F )[U ] = {x ∈ (F ⊗ F )[U ] | β(F, F )U (x) = x}

As in the ungraded case, there is a canonical epimorphism Sym : F ⊗ F → S2(F ) called the

symmetrization map, given by the formula

Sym(x⊗ y) =1

2(x⊗ y + β(F, F )(x⊗ y))

Naturally, we may also consider the nth symmetric power Sn(F ) of a symmetric collection F as

well as the symmetric algebra S(F ), defined as the pointwise sum S(F ) :=⊕

n≥0 Sn(F ).

Definition 20. A twisted supercommutative algebra is a twisted algebra F with multiplication

× : F ⊗ F → F which factors through the symmetric square S2(F ):

F ⊗ F F

S2(F )

×

Sym

One may consider the endomorphism operad of a symmetric collection F in the symmetric

monoidal category SymCol (for a thorough treatment of operads, see, for example, [13]).

Definition 21. For any symmetric collection F , define the endomorphism operad

EndF := {⊗nF → F}n≥0

with the obvious symmetric action and composition law.

This construction suggest an alternative way of defining associative and supercommutative

algebras—indeed, algebras of any desired type—in the context of symmetric collections.

Definition 22. Let O be an operad. A twisted O-algebra is a symmetric collection F together

with an operad map O → EndF .

A twisted associative algebra is therefore a symmetric collection F together with an operad map

Ass → EndF where Ass denotes the operad governing graded associative algebras. A twisted

associative, supercommutative algebra is a symmetric collection F together with an operad map

Com→ EndF where Com denotes the operad governing associative, supercommutative algebras.

42

Page 43: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

We conclude this section by presenting an alternative (equivalent) definition of the category of

symmetric collections in terms of sequences of (graded) representations:

Sequence Definition of Symmetric Collection. An object in SymCol is a sequence of

graded vector spaces V = {Vn}n≥0, where each Vn is a graded representation of the symmetric

group Sn. A morphism f : V → W is a sequence {fn : Vn → Wn}n≥0, where each fn is a

morphism of Sn-representations.

Notation. For each n, the subspace of homogenous degree-k elements in Vn is denoted by V kn

so that

Vn =⊕k∈Z

V kn

‘Graded representation of the symmetric group’ in the definition above means that this a direct

sum of Sn-representations.

The functor F 7→ {F [n]}n≥0 defines an obvious equivalence between our two definitions. The

tensor product of functors F : FinSet→ VecZ in definition 16 becomes:

(V ⊗W )n =⊕i+j=n

IndSnSi×Sj (Vi ⊗Wj)

where Ind denotes the induced representation. Concretely, each summand IndSnSi×Sj (Vi ⊗Wj) is

given by

IndSnSi×Sj (Vi ⊗Wj) =⊕

(i,j)−shuffles σ

σ Vi ⊗Wj

That is, a direct sum of(ni

)isomorphic copies of Vi⊗Vj , indexed by (i, j)-shuffles. Fix a particular

(i, j)-shuffle σ. Since the (i, j)-shuffles form a complete set of coset representatives for the normal

subgroup Si×Sj ⊂ Sn, for each permutation g ∈ Sn, there is a unique element xy ∈ Si×Sj and

(i, j)-shuffle τ such that gσ = τxy. The element g then acts via

g(σ v ⊗ w) = τ x(v)⊗ y(w)

The braiding map β(V,W ) : V ⊗W →W ⊗ V becomes:

43

Page 44: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

β(V,W )n(σ v ⊗ w) = (−1)|v||w|(σρi,j) w ⊗ v

where ρi,j is the (i, j)-shuffle switching the first i and last j elements of {1, ..., n}. We have the

following characterization of twisted associative algebras in the sequence model for symmetric

collections:

Proposition 13. A twisted associative algebra is a symmetric collection {Vn}n≥0 together with

a family of linear, degree-preserving, Si × Sj-equivariant maps ϕi,j : Vi ⊗ Vj → Vi+j such that

the following diagram commutes:

Vi ⊗ Vj ⊗ Vk Vi+j ⊗ Vk

Vi ⊗ Vj+k Vi+j+k

ϕi,j⊗1

1⊗ϕj,k ϕi+j,k

ϕi,j+k

for every i, j, k ∈ N.

Proof. Let V = {Vn}n≥0 be a twisted associative algebra with multiplication given by the mor-

phism f : V ⊗ V → V . For each n, we have a map of graded Sn-representations:

fn :⊕i+j=n

IndSnSi×Sj (Vi ⊗ Vj) = (V ⊗ V )n → Vn

By the discussion above, the tensor product Vi ⊗ Vj is contained as a Si × Sj-subrepresentation

in (V ⊗ V )n. Right-composing with the canonical inclusion i : Vi ⊗ Vj ↪→ (V ⊗ V )n, we obtain a

map of graded Si × Sj-representations

fn ◦ i : Vi ⊗ Vj → Vn

which we will denote by the symbol ϕi,j . Now, let i, j, and k be arbitrary natural numbers and

consider elements vi ∈ Vi, vj ∈ Vj , and vk ∈ Vk. Since V is twisted associative, we have

ϕi+j,k(ϕi,j(v1 ⊗ v2)⊗ v3) = f(f(vi ⊗ vj)⊗ vk)

= (f)(f ⊗ 1)(vi ⊗ vj ⊗ vk)

= (f)(1⊗ f)(vi ⊗ vj ⊗ vk)

= f(vi ⊗ f(vj ⊗ vk))

= ϕi,j+k(vi ⊗ ϕj,k(vj ⊗ vk))

44

Page 45: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

as desired. Conversely, suppose V is a symmetric collection with a family of linear, degree-

preserving, Si×Sj-equivariant maps ϕi,j : Vi⊗Vj → Vi+j such that the diagram in the proposition

commutes. For each n, define an Sn-equivariant map fn :⊕

i+j=n IndSnSi×Sj (Vi ⊗ Vj) = (V ⊗V )n → Vn by putting

fn(σ vi ⊗ vj) = σ ϕi,j(vi ⊗ vj)

and extending by linearity. The collection of maps f = {fn}n∈N defines a morphism V ⊗V → V .

We want to show that (V, f) is a twisted associative algebra. An element of the tensor cube

V ⊗ V ⊗ V is of form σ vi ⊗ vj ⊗ vk where σ ∈ Si+j+kSi×Sj×Sk is an arbitrary (i, j, k)-shuffle. The

groupSi+j+k

Si×Sj×Sk of (i, j, k) shuffles contains normal subgroupsSi+j+kSi+j×Sk and

Si+jSi×Sj with trivial

intersection. Hence, the commutator subgroup

[Si+j+kSi+j × Sk

,Si+jSi × Sj

] ⊆ Si+j+kSi+j × Sk

∩ Si+jSi × Sj

= 1

and these two subgroups commute. The multiplication map

Si+j+kSi+j × Sk

× Si+jSi × Sj

→ Si+j+kSi × Sj × Sk

is therefore a group homomorphism with trivial kernel. By simple counting, it follows that

the multiplication map is surjective and, hence, a group isomorphism. Similarly, we see that

multiplication defines a group isomorphism

Si+j+kSi × Sj+k

× Sj+kSj × Sk

∼=Si+j+k

Si × Sj × Sk

The upshot, of course, is that each (i, j, k)-shuffle σ decomposes uniquely as a commutative

product σi+j,kσi,j of an(i + j, k)-shuffle and an (i, j)-shuffle, as well as a commutative product

σj,kσi,j+k of an (i, j + k)-shuffle and a (j, k)-shuffle. The associativity of f follows trivially,

45

Page 46: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

(f)(f ⊗ 1)(σ vi ⊗ vj ⊗ vk) = fi+j+k(σi+j,k fi+j(σi,j vi ⊗ vj)⊗ vk)

= σi+j,k fi+j+k(σi,j fi+j(vi ⊗ vj)⊗ vk)

= σi+j,k ϕi+j,k(σi,j ϕi,j(vi ⊗ vj)⊗ vk)

= σi+j,kσi,j ϕi+j,k(ϕi,j(vi ⊗ vj)⊗ vk)

= σ ϕi,j+k(vi ⊗ ϕj,k(vj ⊗ vk))

= σi,j+kσj,k fi+j+k(vi ⊗ fj+k(vj ⊗ vk))

= fi+j+k(σi,j+k vi ⊗ σj,k fj+k(vj ⊗ vk))

= fi+j+k(σi,j+k vi ⊗ fj+k(σj,k vj ⊗ vk))

= (f)(1⊗ f)(σ vi ⊗ vj ⊗ vk)

Quite similarly, we obtain the following useful characterization of twisted supercommutative al-

gebras:

Proposition 14. A twisted supercommutative algebra is a symmetric collection {Vn}n≥0 to-

gether with a family of linear, degree-preserving, Si × Sj-equivariant maps ϕi,j : Vi ⊗ Vj → Vi+j

such that ϕi,j(v1 ⊗ v2) = (−1)|v1||v2|ρi,jϕj,i(v2 ⊗ v1).

Proof. Let V = {Vn}n≥0 be a twisted supercommutative algebra with multiplication given by

the morphism f : V ⊗ V → V . For each n, we have a map of graded Sn-representations:

fn :⊕i+j=n

IndSnSi×Sj (Vi ⊗ Vj) = (V ⊗ V )n → Vn

and the induced map of graded Si × Sj-representations

ϕi,j := fn ◦ i : Vi ⊗ Vj → Vn

Since V is twisted supercommutative, we have the following relation for every vi ∈ Vi and vj ∈ Vj :

46

Page 47: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

ϕi,j(vi ⊗ vj) = fn(vi ⊗ vj)

= fn (β(V, V )n(vi ⊗ vj))

= fn

((−1)|vi||vj |ρi,j vj ⊗ vi

)= (−1)|vi||vj |ρi,jfn(vj ⊗ vi)

= (−1)|vi||vj |ρi,jϕj,i(vj ⊗ vi)

as desired. Conversely, suppose V is a symmetric collection with linear, degree-preserving, Si×Sj-equivariant maps ϕi,j : Vi⊗Vj → Vi+j satisfying the condition ϕi,j(v1⊗v2) = (−1)|v1||v2|ρi,jϕj,i(v2⊗v1). For each n, define an Sn-equivariant map fn :

⊕i+j=n IndSnSi×Sj (Vi ⊗ Vj) = (V ⊗ V )n → Vn

by putting

fn(σ vi ⊗ vj) = σ ϕi,j(vi ⊗ vj)

and extending by linearity. The collection of maps f = {fn}n∈N defines a morphism V ⊗V → V .

We want to show that (V, f) is a twisted supercommutative algebra. Indeed, since ρi,jρj,i = 1,

f (β(V, V )n(σ vi ⊗ vj)) = fi+j (β(V, V )n(σ vi ⊗ vj))

= fi+j

((−1)|v1||v2|(σρi,j) vj ⊗ vi

)= (−1)|v1||v2|(σρi,j) fi+j(vj ⊗ vi)

= (−1)|v1||v2|(σρi,j) ϕj,i(vj ⊗ vi)

= (−1)|v1||v2|(−1)|v2||v1|(σρi,jρj,i) ϕi,j(vi ⊗ vj)

= σ ϕi,j(vi ⊗ vj)

= fi+j(σ vi ⊗ vj)

Let τ be a simple transposition on three letters with τ(l) = l + 1, and let τi,j,k ∈ Si+j+k be the

(i, j, k)-shuffle determined by τ . For example, if τ is (12), τi,j,k switches 1, ..., i with i+1, ..., i+j.

If τ is (23), τi,j,k switches i+1, ..., j with i+j+1, ..., i+j+k. Let σ v1⊗v2⊗v3 ∈ IndSi+j+kSi×Sj×Sk(Vi⊗

Vj ⊗ Vk) ⊂ ⊗3V be an arbitrary element of the tensor cube. τ acts on σ v1 ⊗ v2 ⊗ v3 via

τ (σ v1 ⊗ v2 ⊗ v3) = −(−1)(|vl|−1)(|vl+1|−1)στi,j,k vτ(1) ⊗ vτ(2) ⊗ vτ(3)

47

Page 48: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

This extends (uniquely) to a group action of S3 on the tensor cube ⊗3V . The fixed points form

a subcollection of ⊗3V , which we will denote by A3(V ) (notice the difference between A3(V )

and S3(V )). Consider the epimorphism ΦV : ⊗3V → A3(V ) defined by

ΦV (σ v1 ⊗ v2 ⊗ v3) =∑τ∈S3

1

6τ(σ v1 ⊗ v2 ⊗ v3)

Definition 23. A twisted Lie-admissable algebra is a symmetric collection V together with a

morphism f : V ⊗ V → V such that

(f(f ⊗ 1)− f(1⊗ f)) ΦV = 0

One easily sees that in arity 0, definition 23 coincides with the usual definition of Lie-admissability.

In many ways, the sequence definition is better suited for the particular context at hand. In

what follows a ‘symmetric collection’ will almost always refer to a sequence of graded represen-

tations.

10 Ordered Multigraphs

An ordered multigraph is a directed multigraph with total orderings on half-edges. Formally,

Definition 24. An ordered multigraph is a sextuple (V,E, s, t, <s, <t). V is a finite set of ver-

tices, E is a finite set of edges, s : E → V is the source function, t : E → V is the target function,

and <s and <t are functions which assign total orderings on s−1(v) and t−1(v), respectively, to

every vertex v ∈ V .

Definition 25. A morphism of ordered multigraphs is a pair of functions (f : V1 → V2, g : E1 →E2) such that the following diagrams commute:

E1 E2

V1 V2

g

s1 s2

f

E1 E2

V1 V2

g

t1 t2

f

and such that the edge function g : E1 → E2 is monotonic with respect to all relevant orderings.

When drawing these objects, we will adopt the (arbitrary) convention of ordering inputs and

outputs left to right on the page. When in doubt, all multigraphs in this and subsequent sec-

tions are ordered. Multigraphs may be labeled or unlabeled. A ‘rooted’ multigraph is a labeled,

48

Page 49: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

ordered multigraph with a special marked vertex (its root) with no outgoing edges. For example

•x

a b

is a simple rooted multigraph. The vertex set is {x,�}, with � denoting the root, the edge set

is {a, b}, the source function is identically x, and the target function is identically �. The inputs

to � are ordered by a <t b and the outputs from x are ordered by a <s b. In what follows, an

‘ordinary’ vertex in a rooted multigraph is any vertex distinct from the root.

Definition 26. For every natural number n, let Gra[n] denote the real vector space spanned by

rooted multigraphs with ordinary vertices labeled 1, ..., n, such that every ordinary vertex has at

least one outgoing edge. Each Gra[n] is a Z-graded vector space, with the degree of a multigraph

defined as the valence of its root.

For each n, Sn acts on the graded vector space Gra[n] by relabeling.

Definition 27. Let Gra denote the symmetric collection {Gra[n]}n≥0.

Many of the multigraphs we will consider are contained in one of the following two subcollections:

Definition 28. A rooted multigraph is irreducible if the multigraph formed by removing the

root is connected. Grairred is the subcollection of irreducible multigraphs (i.e. Grairred[n] is the

subspace of Gra[n] spanned by irreducible multigraphs, for every n ∈ N).

Definition 29. A rooted multigraph is internally acyclic if the multigraph formed by removing

the root is acyclic. Graacyc is the subcollection of internally acyclic multigraphs (i.e. Graacyc[n]

is the subspace of Gra[n] spanned by internally acyclic multigraphs, for every n ∈ N).

Remark 4. It is worth noting that every multigraph G ∈ Graacyc is necessarily connected.

Suppose G is disconnected and select a connected component G′ ⊂ G not containing the root.

Since G ∈ Graacyc, G′ is acyclic and therefore, by simple order theory, contains a maximal

element, i.e. a vertex with no outgoing edges. However, the only such vertex permitted in Gra

is the root, which, by assumption, is not contained in G′.

We will often work in the intersection Grairred ∩Graacyc, which we denote, somewhat unimagi-

49

Page 50: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

natively, by Grairredacyc . Finally,

Definition 30. We say that a rooted multigraph G ∈ Gra[n] is of type (α1, ..., αn) if for each

i ∈ [n], the vertex labeled i has precisely αi outgoing edges. Gra(α1, ..., αn) ⊂ Gra[n] is the

subspace spanned by type-(α1, ..., αn) multigraphs.

The symmetric collection Gra has the structure of a twisted associative algebra. The multipli-

cation maps ∧ : Gra[i]⊗Gra[j]→ Gra[n] are defined as follows:

Definition 31. Let G1 ∈ Gra[i] and G2 ∈ Gra[j]. Construct the multigraph G1∧G2 ∈ Gra[i+j]

by

1. Relabeling each vertex k ∈ G2 by the natural number i+ k

2. Removing the root vertex of G2

3. Grafting the free inputs from the root vertex of G2 onto the root vertex of G1

The inputs to the root vertex of the wedge product G1 ∧G2 are ordered by concatenation. All

other orderings on G1 ∧G2 are inherited from G1 or G2 in the obvious way. Finally, this map is

extended to the tensor product Gra[i]⊗Gra[j] by bilinearity.

Example 23.

•1

•1

=

•1

•2

We would like to introduce a certain collection of symmetries and anti-symmetries on the half-

edges of our multigraphs. This can be accomplished by modding out by a suitable subcollection.

Definition 32. Let Z ⊂ Gra be the subcollection spanned by elements of the following three

types

50

Page 51: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

· · · +

· · ·

Type 1

· · ·

· · ·

−•

· · ·

· · ·

Type 2

· · ·

· · ·

+

· · ·

· · ·

Type 3

Note that crossings in the pictures above represent arbitrary transpositions of the indicated type.

Definition 33. Let Gra denote the quotient Gra/Z and π : Gra � Gra the canonical epi-

morphism. Denote the images of the subcollections Grairred, Graacyc, and Grairredacyc under π by

Grairred

, Graacyc, and Grairred

acyc , respectively.

We regard elements of the spaces Gra[n] as rooted multigraphs labeled by [n] with

1. Symmetric inputs to all ordinary vertices

2. Anti-symmetric ouputs

3. Anti-symmetric inputs to the root

The wedge product ∧ : Gra ⊗ Gra → Gra induces a product on Gra, denoted by the same

symbol, in the obvious way:

51

Page 52: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

∧ : Gra[i]⊗Gra[j]→ Gra[n]

G1 ⊗G2 7→ G1 ∧G2

It is a simple exercise to verify that these maps are well-defined.

Proposition 15. The wedge product turns Gra into a twisted associative, super-commutative

algebra, freely generated over Grairred

.

Proof. Associativity is obvious. Let G1 ∈ Gra[i] and G2 ∈ Gra[j]. By proposition 14, we need

to show that G1 ∧ G2 = (−1)|G1||G2|ρi,j G2 ∧ G1. Reflecting on the definition of the wedge

product, it is clear that the multigraph G1 ∧G2 can be transformed into G2 ∧G1 by performing

|G1||G2| transpositions on inputs to the root and then applying ρi,j to the vertices of the resulting

multigraph. In Gra[i+ j], each transposition comes with a sign of −1. Hence,

G1 ∧G2 = G1 ∧G2 = (−1)|G1||G2|ρi,j G2 ∧G1 = (−1)|G1||G2|ρi,j G1 ∧G2

Moreover, each rooted multigraph G decomposes uniquely (up to re-ording) as a wedge product

of irreducible pieces. So indeed, Gra ∼= S(Gra

irred)

, as desired.

11 Ordered Multigraphs and Natural Operators

Rooted multigraphs correspond quite nicely to certain spaces of GLm(R)-invariant tensors, which

arise organically in the study of natural operators on alternating multivector fields. Let V = Rm.

For natural numbers α1, ..., αn, γ consider the GL(V )-module

T(α1, ..., αn)(γ) :=⊕

β1,...,βn∈N

n⊗i=1

(⊗βiV ⊗⊗αiV ∗

)⊗⊗γV

and the submodule

W(α1, ..., αn)(γ) :=⊕

β1,...,βn∈N

n⊗i=1

(SβiV ⊗ ∧αiV ∗

)⊗ ∧γV

Put

52

Page 53: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

T(α1, ..., αn) :=⊕γ∈N

T(α1, ..., αn)(γ) W(α1, ..., αn) :=⊕γ∈N

W(α1, ..., αn)(γ)

We regard each as a Z-graded vector space, with degree defined by γ.

Definition 34. Let T and W be the symmetric collections defined by

T[n] :=⊕

α1,...,αn∈NT(α1, ..., αn) W[n] :=

⊕α1,...,αn∈N

W(α1, ..., αn)

Sn acts on components by permuting factors.

The inclusion i : W ↪→ T admits a GL(V )-equivariant left-inverse p : T � W given on each

component by the formula:

pn :

n⊗i=1

(vi1⊗...⊗viβi⊗v∗i1⊗...⊗v

∗iαi

)⊗v1⊗...⊗vγ 7→n⊗i=1

Sym((vi1⊗...⊗viβi )⊗Alt(v∗i1⊗...⊗v

∗iαi

)⊗Alt(v1⊗...⊗vγ)

We denote the subcollections of GL(V )-invariants in the standard way by

TGL(V )[n] := T[n]GL(V ) WGL(V )[n] := W[n]GL(V )

Since p |TGL(V ) ◦ i |WGL(V )= id, p maps TGL(V ) surjectively onto WGL(V ).

Analogously, let Natfieldsm (α1, ..., αn)(γ) denote the vector space of natural multilinear operators

η : ∧α1T × ...× ∧αnT ∧γT over m-dimensional manifolds. Put

Natfieldsm (α1, ..., αn) :=

⊕γ∈N

Natfieldsm (α1, ..., αn)(γ)

grading, as before, by γ.

Definition 35. Let Natfieldsm be the symmetric collection defined by

Natfieldsm [n] =

⊕α1,...,αn∈N

Natfieldsm (α1, ..., αn)

Sn acts by permuting arguments.

53

Page 54: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

The connection between natural multilinear operators on alternating multivector fields and in-

variant tensors in W is expressed in the following simple proposition.

Proposition 16. For every α1, ..., αn ∈ N, restriction to the standard fiber defines a linear

isomorphism

Natfieldsm (α1, ..., αn) ∼=

(W(α1, ..., αn)GL(V )

)NGL∞mThese isomorphisms assemble into an isomorphism of symmetric collections:

Natfieldsm

∼=(WGL(V )

)NGL∞mProof. Consider the subspace Natfields

m,r (α1, ..., αn)(γ) ⊂ Natfieldsm (α1, ..., αn)(γ) consisting of nat-

ural multilinear operators η : ∧α1T × ...× ∧αnT ∧γT of order not exceeding r. By invariant-

theoretic reduction and remark 3, the restriction defines a linear isomorphism

Natfieldsm,r (α1, ..., αn)(γ) ∼=

((n⊗i=1

T rm ∧αi V ∗)⊗ ∧γV

)Gr+1m

We have the natural identifications

(n⊗i=1

T rm ∧αi V ∗)⊗ ∧γV =

n⊗i=1

r⊕j=0

SjV ⊗ ∧αiV ∗⊗ ∧γV

=⊕

(β1,...,βn)∈{0,...,r}n

n⊗i=1

((SβiV ⊗ ∧αiV ∗

)⊗ ∧γV

)

so that

54

Page 55: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

Natfieldsm (α1, ..., αn)(γ) ∼=

∞⋃r=0

⊕(β1,...,βn)∈{0,...,r}n

n⊗i=1

((SβiV ⊗ ∧αiV ∗

)⊗ ∧γV

)Gr+1m

=

∞⋃r=0

⊕(β1,...,βn)∈{0,...,r}n

n⊗i=1

((SβiV ⊗ ∧αiV ∗

)⊗ ∧γV

)G∞m

=

∞⋃r=0

⊕(β1,...,βn)∈{0,...,r}n

n⊗i=1

((SβiV ⊗ ∧αiV ∗

)⊗ ∧γV

)G∞m

=

⊕β1,...,βn∈N

n⊗i=1

((SβiV ⊗ ∧αiV ∗

)⊗ ∧γV

)G∞m

and

Natfieldsm (α1, ..., αn) =

⊕γ∈N

Natfieldsm (α1, ..., αn)(γ)

∼=⊕γ∈N

⊕β1,...,βn∈N

n⊗i=1

((SβiV ⊗ ∧αiV ∗

)⊗ ∧γV

)G∞m

=

⊕β1,...,βn,γ∈N

n⊗i=1

((SβiV ⊗ ∧αiV ∗

)⊗ ∧γV

)G∞m

= W(α1, ..., αn)G∞m

Recall, Gr+1m is isomorphic to a semidirect product NGLr+1

m oGLm(R). Taking the inverse limit,

we obtain, G∞m∼= NGLr+1

m o GLm(R). In particular, the subgroups NGL∞m , GLm(R) ⊆ G∞m

generate G∞m . Hence,

Natfieldsm (α1, ..., αn) ∼=

(W(α1, ..., αn)GL(V )

)NGL∞mwhich establishes the first part. Since Natfields

m [n] =⊕

α1,...,αn∈NNatfieldsm (α1, ..., αn) is a direct

sum of G∞m -modules, these isomorphisms assemble:

Natfieldsm [n] =

⊕α1,...,αn∈N

Natfieldsm (α1, ..., αn) ∼=

⊕α1,...,αn∈N

W(α1, ..., αn)G∞m = WG∞m =

(WGL(V )

)NGL∞m55

Page 56: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

Each isomorphism is (trivially) an isomorphism of Sn-representations, since permuting the ar-

guments of a natural operation permutes the factors in the standard fiber of the corresponding

jet bundle.

We wish to define a correspondence between elements of the symmetric collection Gra and

WGL(V ). Every multigraph G ∈ Gra[n] gives rise to an elementary invariant tensor fn(G) ∈TGL(V )[n] in the following way: the multigraph G has n ordinary vertices labeled 1, ..., n. For

each i ∈ [n], let αi, βi denote the number of outgoing and incoming edges, respectively, to the

ith vertex of G and let γ denote the number of incoming edges to the root. The structure of G

determines a bijection σ(g) from inputs onto outputs, which determines an elementary invariant

tensor Iσ(G) ∈⊗n

i=1

((⊗βiV ⊗⊗αiV ∗

)⊗⊗γV

). We put fn(G) = Iσ(g) and extend to Gra[n]

by linearity. The morphism f : Gra → TGL(V ) is surjective, by the Invariant Tensor Theo-

rem. Moreover, if m exceeds∑ni=1 αi, we have ker fn ∩ Gra(α1, ..., αn) = 0. Fix an inclusion

i : Gra ↪→ Gra right inverse to π.

Definition 36. Let f : Gra→WGL(V ) be the morphism of symmetric collections given by the

composition f =: p ◦ f ◦ i

The following diagram commutes:

Gra TGL(V )

Gra WGL(V )

f

π p

f

i i

We see immediately that

Proposition 17. f

1. is well-defined (i.e. independent of i)

2. is surjective

3. restricts to a linear isomorphism fn : Gra(α1, ..., αn)∼→W(α1, ..., α)GL(V ) for

∑ni=1 αi < m

Proof. All three claims follow from the trivial observation f(Z) = ker p.

1. If i1 and i2 are two right-inverses of π and g ∈ Gra, we have π(i1g − i2g) = g − g = 0.

Hence, i1g− i2g ∈ kerπ = Z and therefore, since f(Z) = ker p, (p ◦ f ◦ i1 − p ◦ f ◦ i2)(g) =

p(f((i1g − i2g))) = 0, as desired.

56

Page 57: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

2. Since f is surjective, it admits a right-inverse g. We claim π ◦ g ◦ i is a right-inverse for f .

Let x ∈ TGL(V ). Notice that (i ◦π ◦ g)(x)− g(x) ∈ kerπ since π((i ◦π ◦ g)(x)− g(x)) = (π ◦g)(x)−(π◦g)(x) = 0. Therefore, since f(Z) = ker p, we have (p◦f)((i◦π◦g)(x)−g(x)) = 0

and hence (p◦f ◦ i◦π ◦g)(x) = (p◦f ◦g)(x) = p(x). Now, letting x = i(y) for y ∈WGL(V ),

we see that (p ◦ f ◦ i ◦π ◦ g ◦ i)(y) = (p ◦ i)(y) = y. However, p ◦ f ◦ i ◦π ◦ g ◦ i = f ◦π ◦ g ◦ iso, indeed, f ◦ π ◦ g ◦ i = id, as desired.

3. Let α1, ..., αn be natural numbers with∑ni=1 αi < m. Let f ′, f

′, π′, and p′ denote

the restrictions of f , f , π, and p to Gra(α1, ..., αn), Gra(α1, ..., αn), Gra(α1, .., αn), and

T(α1, ..., αn)GL(V ), respectively. Note that the following diagram commutes

Gra(α1, ..., αn) T(α1, ..., αn)GL(V )

Gra(α1, ..., αn) W(α1, ..., αn)GL(V )

f ′

p′

f′

i

The top row is an isomorphism, by the Invariant Tensor Theorem. And fn is diagonal with

respect to the decomposition Gra[n] =⊕Gra(α1, ..., αn). Thus, f(Z) = ker p implies

f(kerπ′) = ker p′. Suppose x ∈ ker f′

so that (p′ ◦f ′ ◦ i)(x) = 0. It follows that (f ′ ◦ i)(x) ∈ker p′ = f ′(kerπ′). In other words, there is a y ∈ kerπ′ such that f ′(i(x)−y) = (f ′ ◦ i)(x)−f ′(y) = 0. Notice that π′(i(x) − y) = x − π′y = x and, clearly, i(x) − y ∈ ker f ′. Hence,

x ∈ π′ ker f ′. Thus, we have proved ker f′ ⊆ π′ ker f ′ = 0. f

′is surjective onto WGL(V ) by

part 2, so we are done.

In summary, we have proved

1. Natural multilinear operators on alternating multivector fields correspond to certain types

of invariant tensors

2. Invariant tensors correspond (in stable dimension) to certain spaces of rooted multigraphs

12 Gerstenhaber Algebras

A Gerstenhaber algebra is a graded associative, supercommutative algebra with a lie bracket of

degree −1 satisfying the poisson identity. More formally

Definition 37. A Gerstenhaber algebra is a graded vector space G =⊕∞

i=1Gi equipped with a

wedge product ∧ : G⊗G→ G and a lie bracket [ , ] : G⊗G→ G. The degree of a homogenous

element a is denoted by |a|, and the following identities hold:

57

Page 58: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

1. |a ∧ b| = |a|+ |b| (the wedge product has degree 0)

2. |[a, b]| = |a|+ |b| − 1 (the lie bracket has degree −1)

3. a ∧ (b ∧ c) = (a ∧ b) ∧ c (the wedge product is associative)

4. [a, [b, c]] = [[a, b], c] + (−1)(|a|−1)(|b|−1)[b, [a, c]] (the Jacobi Identity)

5. a ∧ b = (−1)|a||b|b ∧ a (the wedge product is supercommutative)

6. [a, b] = −(−1)(|a|−1)(|b|−1)[b, a] (the lie bracket is (graded) antisymmetric)

7. [a, b ∧ c] = [a, b] ∧ c+ (−1)(|a|−1)|b|b ∧ [a, c] (the Poisson identity)

Example 24. The Hochschild cohomology H∗(A,A) of an associative algebra A is a Gersten-

haber algebra under the cup product and a lie bracket defined by Gerstenhaber [14].

Example 25. The graded vector space of alternating multivector fields forms a Gerstenhaber

algebra under the wedge product of multivector fields and the Schouten-Nijenhuis bracket.

Example 26. The vector space of natural multilinear operations on alternating multivector

fields, graded by the degree of the output, forms a Gerstenhaber algebra under the wedge prod-

uct and Schouten-Nijenhuis bracket.

Example 27. The homology of the little-disks operad is the operad for Gerstenhaber algebras,

a result first proved in [15].

13 Pre-Gerstenhaber Algebras

We will now formalize the notion of a pre-Gerstenhaber algebra. Roughly speaking, pre-Gerstenhaber

algebras are to Gerstenhaber algebras as Lie-admissable algebras are to Lie Algebras. Formally,

Definition 38. A pre-Gerstenhaber algebra is a graded vector space P =⊕∞

i=1 Pi together with

two bilinear multiplications ∧, ∗ : P ⊗ P → P such that the following identities hold:

1. |a ∧ b| = |a|+ |b| (the wedge product has degree 0)

2. |a ∗ b| = |a|+ |b| − 1 (the asterisk has degree −1)

3. a ∧ (b ∧ c) = (a ∧ b) ∧ c (the wedge product is associative)

4. a ∧ b = (−1)|a||b|b ∧ a (the wedge product is supercommutative)

58

Page 59: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

5. a ∗ (b ∧ c) = (a ∗ b) ∧ c + (−1)(|a|−1)|b|b ∧ (a ∗ c) (the asterisk is a left-derivation over the

wedge product)

6. (a ∧ b) ∗ c = a ∧ (b ∗ c) + (−1)(|c|−1)|b|(a ∗ c) ∧ b (the asterisk is a right-derivation over the

wedge product)

Furthermore, if [a, b] = a ∗ b − (−1)(|a|−1)(|b|−1)b ∗ a, we have the following Lie-admissability

condition

7. [a, [b, c]] = [[a, b], c] + (−1)(|a|−1)(|b|−1)[b, [a, c]] (the asterisk is Lie-admissable)

Let A be a graded, Lie-admissable algebra, and consider the pre-Gerstenhaber algebra P (A)

generated freely over A (more precisely, consider the functor P : LieAdm → Pre-Gerst left

adjoint to the forgetful functor F : Pre-Gerst → LieAdm and apply P to A). What can we

say, if anything, about the structure of P (A)? As a vector space, P (A) is a quotient of the free

graded supercommutative associative algebra generated by the graded vector space A. Naively,

one might expect a rather unwieldy set of relations in the denominator of this quotient, imposed

by the interaction between the Poisson identities and the Lie-admissability condition. This is,

remarkably, not the case.

Theorem 5. Let A be a graded, Lie-admissable algebra and P (A) the free pre-Gerstenhaber

algebra generated by A. P (A) is isomorphic as a vector space (and, indeed, as a graded algebra)

to the free graded supercommutative associative algebra generated by the graded vector space

A.

Proof. We must demonstrate the consistency of pre-Gerstenhaber axioms 5, 6, and 7 on the

free graded supercommutative associative algebra generated by the graded vector space V . This

object can be constructed by

1. forming the tensor algebra over A (regarded as an ungraded vector space)

2. grading in accordance with the given grading on A and the formula |a ∧ b| = |a|+ |b|

3. modding out by supercommutativity

To check the consistency of axioms 5 and 6, we compute the product (a ∧ b) ∗ (c ∧ d) in two

different ways. Applying the left Poisson relation first, we obtain:

59

Page 60: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

(a ∧ b) ∗ (c ∧ d) = ((a ∧ b) ∗ c) ∧ d+ (−1)(|a|+|b|−1)|c|c ∧ ((a ∧ b) ∗ d)

= a ∧ (b ∗ c) ∧ d+ (−1)(|c|−1)|b|(a ∗ c) ∧ b ∧ d

+ (−1)(|a|+|b|−1)|c|(c ∧ a ∧ (b ∗ d) + (−1)(|d|−1)|b|c ∧ (a ∗ d) ∧ b)

= a ∧ (b ∗ c) ∧ d+ (−1)(|c|−1)|b|(a ∗ c) ∧ b ∧ d

+ (−1)(|a|+|b|−1)|c|c ∧ a ∧ (b ∗ d) + (−1)(|a|+|b|−1)|c|+(|d|−1)|b|c ∧ (a ∗ d) ∧ b

Applying the right Poisson relation first, we obtain:

(a ∧ b) ∗ (c ∧ d) = a ∧ (b ∗ (c ∧ d)) + (−1)(|c|+|d|−1)|b|(a ∗ (c ∧ d)) ∧ b

= a ∧ (b ∗ c) ∧ d+ (−1)(|c|−1)|d|a ∧ c ∧ (b ∗ d)

+ (−1)(|c|+|d|−1)|b|((a ∗ c) ∧ d ∧ b+ (−1)(|a|−1)|c|c ∧ (a ∗ d) ∧ b)

= a ∧ (b ∗ c) ∧ d+ (−1)(|c|−1)|d|a ∧ c ∧ (b ∗ d)

+ (−1)(|c|+|d|−1)|b|(a ∗ c) ∧ d ∧ b+ (−1)(|c|+|d|−1)|b|+(|a|−1)|c|c ∧ (a ∗ d) ∧ b

The first and fourth terms in both expansions are the same. The second term in the first

expansion is equal to the third term in the second (here, applying the supercommutativity of the

wedge product) and the third term in the first expansion is equal to the second in the second.

However, this part of the proof is hardly surprising. Clearly, the consistency of axioms 6 and

7 follows from the consistency of the right and left Poisson relations on the free Gerstenhaber

algebra over A.

The interesting observation is that if we assume the Poisson relations, we get the Jacobi identity

for free. We know that the Jacobi identity is satisfied by all triples in A. Thus, by induction, it

suffices to prove the following: if the triples (a, x, y) and (b, x, y) satisfy the Jacobi identity, so

does the triple (a ∧ b, x, y). For notational convenience, we will omit the absolute values when

writing degrees:

60

Page 61: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

[a ∧ b, [x, y]] = (a ∧ b) ∗ (x ∗ y − (−1)(x−1)(y−1)y ∗ x)− (−1)(a+b−1)(x+y)(x ∗ y − (−1)(x−1)(y−1)y ∗ x) ∗ (a ∧ b)

= a ∧ (b ∗ (x ∗ y)) + (−1)(x+y)b(a ∗ (x ∗ y)) ∧ b

− (−1)(x−1)(y−1)a ∧ (b ∗ (y ∗ x))− (−1)(x+y)b+(x−1)(y−1)(a ∗ (y ∗ x)) ∧ b

− (−1)(a+b−1)(x+y)((x ∗ y) ∗ a) ∧ b− (−1)(b−1)(x+y)a ∧ ((x ∗ y) ∗ b)

+ (−1)(a+b−1)(x+y)+(x−1)(y−1)((y ∗ x) ∗ a) ∧ b+ (−1)(b−1)(x+y)+(x−1)(y−1)a ∧ ((y ∗ x) ∗ b)

[[a ∧ b, x], y] = ((a ∧ b) ∗ x− (−1)(a+b−1)(x−1)x ∗ (a ∧ b)) ∗ y

− (−1)(a+b+x)(y−1)y ∗ ((a ∧ b) ∗ x− (−1)(a+b−1)(x−1)x ∗ (a ∧ b))

= a ∧ ((b ∗ x) ∗ y) + (−1)(y−1)(b+x−1)(a ∗ y) ∧ (b ∗ x)

+ (−1)(x−1)b(a ∗ x) ∧ (b ∗ y) + (−1)(x+y)b((a ∗ x) ∗ y) ∧ b

− (−1)(a+b−1)(x−1)(x ∗ a) ∧ (b ∗ y)− (−1)(a+b−1)(x−1)+(y−1)b((x ∗ a) ∗ y) ∧ b

− (−1)(b−1)(x−1)a ∧ ((x ∗ b) ∗ y)− (−1)(b−1)(x−1)+(y−1)(x+b−1)(a ∗ y) ∧ (x ∗ b)

− (−1)(a+b+x)(y−1)(y ∗ a) ∧ (b ∗ x)− (−1)(b+x)(y−1)a ∧ (y ∗ (b ∗ x))

− (−1)(a+b+x)(y−1)+(x−1)b(y ∗ (a ∗ x)) ∧ b− (−1)(b−1)(y−1)+(x−1)b(a ∗ x) ∧ (y ∗ b)

+ (−1)(a+b+x)(y−1)+(a+b−1)(x−1)(y ∗ (x ∗ a)) ∧ b+ (−1)(b−1)(y−1)+(a+b−1)(x−1)(x ∗ a) ∧ (y ∗ b)

+ (−1)(a+b+x)(y−1)+(b−1)(x−1)(y ∗ a) ∧ (x ∗ b) + (−1)(b+x)(y−1)+(b−1)(x−1)a ∧ (y ∗ (x ∗ b))

61

Page 62: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

[x, [a ∧ b, y]] = x ∗ ((a ∧ b) ∗ y − (−1)(a+b−1)(y−1)y ∗ (a ∧ b))

− (−1)(a+b+y)(x−1)((a ∧ b) ∗ y − (−1)(a+b−1)(y−1)y ∗ (a ∧ b)) ∗ x

= (x ∗ a) ∧ (b ∗ y) + (−1)(x−1)(a)a ∧ (x ∗ (b ∗ y))

+ (−1)(y−1)b(x ∗ (a ∗ y)) ∧ b+ (−1)(y−1)b+(x−1)(a+y−1)(a ∗ y) ∧ (x ∗ b)

− (−1)(a+b−1)(y−1)(x ∗ (y ∗ a)) ∧ b− (−1)(a+b−1)(y−1)+(x−1)(y+a−1)(y ∗ a) ∧ (x ∗ b)

− (−1)(b−1)(y−1)(x ∗ a) ∧ (y ∗ b)− (−1)(b−1)(y−1)+(x−1)aa ∧ (x ∗ (y ∗ b))

− (−1)(a+b+y)(x−1)a ∧ ((b ∗ y) ∗ x)− (−1)(a−1)(x−1)(a ∗ x) ∧ (b ∗ y)

− (−1)(a+b+y)(x−1)+(y−1)b(a ∗ y) ∧ (b ∗ x)− (−1)(a+y)(x−1)+(y−1)b((a ∗ y) ∗ x) ∧ b

+ (−1)(a+b+y)(x−1)+(a+b−1)(y−1)(y ∗ a) ∧ (b ∗ x) + (−1)(a+y)(x−1)+(a+b−1)(y−1)((y ∗ a) ∗ x) ∧ b

+ (−1)(a+b+y)(x−1)+(b−1)(y−1)a ∧ ((y ∗ b) ∗ x) + (−1)(a−1)(x−1)+(b−1)(y−1)(a ∗ x) ∧ (y ∗ b)

We can now compute the difference [[a ∧ b, x], y] + (−1)(a+b−1)(x−1)[x, [a ∧ b, y]] − [a ∧ b, [x, y]].

Inspecting the expansions above, one easily verifies that terms of shape (� ∗ �) ∧ (� ∗ �) cancel

each other out. What’s left is killed by the Jacobi identity

[[a ∧ b, x], y] + (−1)(a+b−1)(x−1)[x, [a ∧ b, y]]− [a ∧ b, [x, y]] =

a ∧(

(b ∗ x) ∗ y − (−1)(b−1)(x−1)(x ∗ b) ∗ y − (−1)(b+x)(y−1)y ∗ (b ∗ x) + (−1)(b+x)(y−1)+(b−1)(x−1)y ∗ (x ∗ b)

+ (−1)(x−1)(b−1)x ∗ (b ∗ y)− (−1)(x+y)(b−1)x ∗ (y ∗ b)− (−1)(x−1)(y−1)(b ∗ y) ∗ x+ (−1)(y−1)(x+b)(y ∗ b) ∗ x

− b ∗ (x ∗ y) + (−1)(x−1)(y−1)b ∗ (y ∗ x) + (−1)(b−1)(x+y)(x ∗ y) ∗ b− (−1)(b−1)(x+y)+(x−1)(y−1)(y ∗ x) ∗ b)

+ (−1)(x+y)b(

(a ∗ x) ∗ y − (−1)(a−1)(x−1)(x ∗ a) ∗ y − (−1)(a+x)(y−1)y ∗ (a ∗ x)

+ (−1)(a+x)(y−1)y ∗ (x ∗ a) + (−1)(a−1)(x−1)x ∗ (a ∗ y)− (−1)(a−1)(x+y)x ∗ (y ∗ a)

− (−1)(y−1)(x−1)(a ∗ y) ∗ x+ (−1)(a+y)(x−1)+(a−1)(x+y)(y ∗ a) ∗ x− a ∗ (x ∗ y)

+ (−1)(x−1)(y−1)a ∗ (y ∗ x) + (−1)(a−1)(x+y)(x ∗ y) ∗ a− (−1)(a−1)(x+y)+(x−1)(y−1)(y ∗ x) ∗ a)∧ b

= a ∧(

[[b, x], y] + (−1)(b−1)(x−1)[x, [b, y]]− [b, [x, y]])

+(

[[a, x], y] + (−1)(a−1)(x−1)[x, [a, y]]− [a, [x, y]])∧ b

= 0

62

Page 63: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

The compatibility of the operads Com and Lie in the Gerstenhaber case is not surprising, since

Gerstenhaber algebras come from something topological (see example 27). The compatibility of

the operads Com and LieAdm therefore begs the question: is there a topological construction—

perhaps related to the operad of little disks (little annuli, for example?)—linked to the operad

of pre-Gerstenhaber algebras? We conjecture that there is.

14 A Lie-admissable Structure on Grairredacyc

We now define a Lie-admissable structure on the symmetric collection Grairred

acyc , which general-

izes the pre-Lie product of rooted trees described by Chapoton and Livernet in [6]. We start by

considering a much smaller set of multigraphs.

Definition 39. An elementary multigraph is a multigraph contained in the graded vector space

Graacyc[1]

Elementary multigraphs look like:

•1

· · ·

Suppose G1 ∈ Gra[n] is irreducible and G2 is elementary. For every vertex v ∈ G1, construct

the multigraph G1 ∗v G2 ∈ Gra[n+ 1] by

1. Relabeling the singular ordinary vertex in G2 by the natural number n+ 1

2. Removing the root of G2

3. Grafting the leftmost free input in G2 onto the vertex v of G1, to the right of all existing

inputs to v

4. Grafting all remaining free inputs in G2 onto the root of G1, to the right of all existing

inputs, preserving the original ordering on grafted inputs from G2

We define the product G1 ∗G2 by the formula

G1 ∗G2 :=∑v

G1 ∗v G2 (16)

This operation extends to a product on the symmetric collection Grairredacyc . The multiplication

maps Grairredacyc [i]⊗Grairredacyc [j]→ Grairredacyc [i+ j] are defined through the following general proce-

63

Page 64: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

dure. For multigraphs, G1 ∈ Grairredacyc [i] and G2 ∈ Grairredacyc [j], the multigraph G1∗G2 ∈ Gra[i+j]

is constructed by:

1. Relabeling the ordinary vertices 1, ..., j in G2 by i+ 1, ..., i+ j

2. Erasing all of the edges and vertices in G2 not directly connected to the root, creating a

multigraph G′2 which we regard as a wedge product of elementary rooted multigraphs

3. Applying the the left Poisson relation and formula (16) to obtain G1 ∗G′2

4. Re-inserting all of the edges and vertices erased to obtain G1 ∗G2

We illustrate this procedure with a few simple examples.

Example 28.

•1

· · · ∗

•1

· · · =

•1

•2

· · · · · ·

Example 29.

•1

•2

• 1

· · · =

•1

•2

•3

· ··

+

•1

•2

• 3

· ··

Example 30.

•1

•2

•1 • 2

=

•1

•2

• 4

•3

+

•1

•2

• 4

• 3

•1

•2

• 4

•3

•1

•2

• 4

• 3

Finally, ∗ is extended to Grairredacyc [i] ⊗ Grairredacyc [j] by bilinearity. ∗ induces a multiplication on

64

Page 65: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

the quotient space Gra, which we will also denote by ∗, in the obvious way.

Definition 40. For every G1, G2 ∈ Gra, let G1 ∗G2 = G1 ∗G2.

It is a simple exercise to verify that G1 ∗G2 is well-defined.

Theorem 6. Grairred

acyc forms a twisted Lie-admissable algebra under the ∗-product defined above.

Proof. Let G1, G2, and G3 be rooted multigraphs in Grairred

acyc , of degree p, q, and r, respectively.

We will use the following shorthand to denote the product G1 ∗G2

1 2

This diagram actually represents a sum over all possible graftings of G2 onto G1, with signs as

specified above. Hence

(G1 ∗G2) ∗G3 =

1 2

3+

1 2

3

And

G1 ∗ (G2 ∗G3) =

1 2

3+

1 2

3

Hence,

(G1 ∗G2) ∗G3 −G1 ∗ (G2 ∗G3) =

1 2

3−

1 2

3

The Jacobi identity is the twelve term sum

[[G1, G2], G3] + (−1)(p−1)(q−1)[G2, [G1, G3]]− [G1, [G2, G3]] = 6 (∗(∗ ⊗ 1)− ∗(1⊗ ∗)) ΦGra

irredacyc

(G1 ⊗G2 ⊗G3)

=∑τ∈S3

τ

1 2

3−

1 2

3

65

Page 66: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

with S3 acting on Grairred

acyc as described in section 9.To demonstrate Lie-admissability, it therefore

suffices to show that

1 2

3− (−1)(q−1)(r−1)

1 3

2= 0 (17)

And

1 2

3− (−1)(p−1)(q−1)

2 1

3= 0 (18)

Equation (17) is more or less trivial. Since all ordinary inputs are symmetric, the only sign

incurred by swapping G2 and G3 comes from the antisymmetry of the root vertex of G1. One

root-input from both G2 and G3 is diverted to an ordinary vertex of G1, leaving q − 1 and

r − 1 inputs, respectively, attached to the root. Swapping these inputs results in a sign of

(−1)(q−1)(r−1), from which equation (17) immediately follows.

Equation (18) is a bit more subtle. There are two cases to consider. The edges connecting

G3 to G1 and G3 to G2 may have either the same or different source vertex. Suppose, first,

that these connecting edges have different source vertices. These edges were originally inputs to

the root vertex of G3 in positions, say, a and b in the total ordering on these inputs. Without

loss of generality, suppose a < b. Thus, for any vertices v1 ∈ G1 and v2 ∈ G2 (it doesn’t

matter which ones), the term on the left of equation (18) coming from G1 ∗v1 (G2 ∗v2 G3)

has sign (−1)b−1(−1)q+a−1. The term on the right coming from G2 ∗v2 (G1 ∗v1 G3) has sign

(−1)a−1(−1)p+b. The term on the right has an additional sign of (−1)pq incurred from swapping

inputs to the root. Therefore, we have

(−1)b−1(−1)q+a−1G1 ∗v1 (G2 ∗v2 G3)− (−1)(p−1)(q−1)(−1)a−1(−1)p+b(−1)pqG1 ∗v1 (G2 ∗v2 G3)

= (−1)a+b+qG1 ∗v1 (G2 ∗v2 G3)(

1− (−1)(p−1)(q−1)+p+q+qp−1)

= (−1)a+b+qG1 ∗v1 (G2 ∗v2 G3) (1− 1)

= 0

Next, suppose the connecting edges share a source. These edges were originally inputs to the

root vertex of G3. Reflecting on the construction of G1 ∗ (G2 ∗G3), it is clear that these inputs

66

Page 67: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

were consecutive, say a and a + 1 in the total ordering on the root-inputs in G3. The term on

the left of equation (18) coming from G1 ∗v1 (G2 ∗v2 G3) has sign (−1)a−1(−1)q+a−1. The term

on the right coming from G2 ∗v2 (G1 ∗v1 G3) has sign (−1)a−1(−1)p+a−1 and then an additional

sign of (−1)pq+1 incurred from swapping inputs to the root and swapping the connecting edges

from G3. Therefore, we have

(−1)a−1(−1)q+a−1G1 ∗v1 (G2 ∗v2 G3)− (−1)(p−1)(q−1)(−1)a−1(−1)p+a−1(−1)pq+1G1 ∗v1 (G2 ∗v2 G3)

= (−1)qG1 ∗v1 (G2 ∗v2 G3)(

1− (−1)(p−1)(q−1)+p+q+qp−1)

= (−1)a+b+qg1 ∗v1 (G2 ∗v2 G3) (1− 1)

= 0

which completes the proof.

Corollary 5. The wedge product and the ∗-product turn the symmetric collection Graacyc into

a twisted pre-Gerstenhaber algebra.

Proof. This follows immediately from theorem 5 and theorem 6.

Notice that:

1. The restriction of ∗ to the space of rooted trees (i.e., the subcollection of Grairred

acyc spanned

in each component by rooted multigraphs containing no vertex with more than one outgoing

edge) is precisely the pre-Lie concatenation product defined by Chapoton and Livernet in

[6].

2. The Lie algebra product [a, b] = a ∗ b− (−1)(|a|−1)(|b|−1)b ∗ a corresponds to the usual Lie

bracket of differential operators.

15 A Co-algebra Structure on Grairredacyc

Let V = {Vn}n≥0 be a symmetric collection in the category of graded vector spaces. The dual

of V , denoted V ∗, is defined by (V ∗)n = (Vn)∗, where (Vn)∗ is the dual graded vector space

associated to Vn. Duality defines a functor SymCol→ SymColop with f∗ : W ∗ → V ∗ defined

by (f∗)n = (fn)∗ for every morphism f : V →W . Recall the following notation: for every n ∈ Nand k ∈ Z, the subspace of Vn consisting of homogenous degree-k elements is denoted by V kn .

67

Page 68: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

Definition 41. A symmetric collection V is locally finite-dimensional if for every n ∈ N and

k ∈ Z, V kn is a finite-dimensional vector space.

Proposition 18. If V is locally-finite dimensional, the natural map ϕ : V → V ∗∗ is an isomor-

phism of symmetric collections.

Proof. For every n ∈ N, ϕn : Vn → (V ∗∗)n = (Vn)∗∗ is a map of graded vector spaces ϕn =

⊕k∈Zϕkn, where ϕkn : V kn → V k∗∗n is given by the evaluation:

ϕkn : vk 7−→(wk 7→ wk(vk)

)for every vk ∈ V kn and wk ∈ V k∗n . Each ϕkn is injective by the standard argument involving

the axiom of choice, and therefore surjective by the finiteness of dimV kn and the rank-nullity

theorem

Thus, duality induces an equivalence of categories SymCollocally finite∼= SymColoplocally finite.

Duality behaves nicely with respect to tensor products. For every pair of symmetric collections

V,W ∈ SymCol, there is a natural morphism

ψ : V ∗ ⊗W ∗ → (V ⊗W )∗

which satisfies

ψn(v∗ ⊗ w∗)(v ⊗ w) = v∗(v)w∗(w)

for v ∈ Vi, v∗ ∈ V ∗i , w ∈Wj , and w∗ ∈W ∗j and i+ j = n.

Proposition 19. For V,W locally finite-dimensional, the natural morphism ψ : V ∗ ⊗W ∗ →(V ⊗W )∗ is an isomorphism of symmetric collections.

Proof. Recall,

(V ∗ ⊗W ∗)n =⊕i+j=n

IndSnSi×SjV∗i ⊗W ∗j

and

68

Page 69: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

(V ⊗W )∗n =

⊕i+j=n

IndSnSi×SjVi ⊗Wj

∗ =⊕i+j=n

IndSnSi×Sj (Vi ⊗Wj)∗

Each V ∗i ⊗W ∗j , being a tensor product of graded vector spaces, is a direct sum

Vi ⊗Wj =⊕

k1+k2=k

V k1∗i ⊗W k2∗j (19)

And each (Vi ⊗Wj)∗ is given by

(Vi ⊗Wj)∗ =

( ⊕k1+k2=k

V k1i ⊗Wk2j

)∗=

⊕k1+k2=k

(V k1i ⊗Wk2j )∗ (20)

Moreover, since the direct sums in decompositions (19) and (20) are direct sums of Si × Sj-

representations, these sums commute with induction, i.e.

(V ∗ ⊗W ∗)n =⊕i+j=nk1+k2=k

IndSnSi×SjVk1∗i ⊗W k2∗

j =⊕i+j=nk1+k2=k

⊕(i,j)−shuffles σ

σ V k1∗i ⊗W k2∗j

and

(V ⊗W )∗n =⊕i+j=nk1+k2=k

IndSnSi×Sj (Vk1i ⊗W

k2j )∗ =

⊕i+j=nk1+k2=k

⊕(i,j)−shuffles σ

σ (V k1i ⊗Wk2j )∗

The nth component of the canonical map ψ : V ∗ ⊗W ∗ → (V ⊗W )∗ is a direct sum

ψn =⊕i+j=nk1+k2=k

⊕(i,j)−shuffles σ

σ ψk1,k2i,j

where σ ψk1,k2i,j : σ V k1∗i ⊗W k2∗j → σ (V k1i ⊗W

k2j )∗ is the linear map given by

σ ψk1,k2i,j (σ vk1∗i ⊗ wk2∗j )(σ vk1i ⊗ wk2j ) = vk1∗i (vk1i )wk2∗j (wk2j )

for every vk1i ∈ Vk1i , vk1∗i ∈ V k1∗i , wk2j ∈ W

k2j , and wk2∗j ∈ W k2∗

j . The important point is that

each σ ψk1,k2i,j is a linear map between finite-dimensional vector spaces and hence an isomorphism

by the standard argument involving the axiom of choice.

69

Page 70: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

Recall from section 9 the construction of the epimorphism ΦV : ⊗3V → A3(V ): for every

symmetric collection V , there is an action of S3 on ⊗3V given on simple transpositions τ(l) = l+1

and elements σ v1 ⊗ v2 ⊗ v3 ∈ IndSi+j+kSi×Sj×Sk (Vi ⊗ Vj ⊗ Vk) ⊂ ⊗3V by

τ (σ v1 ⊗ v2 ⊗ v3) = −(−1)(|vl|−1)(|vl+1|−1)τi,j,kσ vτ(1) ⊗ vτ(2) ⊗ vτ(3)

and extended uniquely to S3. The fixed points form a subcollection A3(V ) ⊂ ⊗3V and ΦV :

⊗3V → A3(V ) is the morphism defined by

ΦV (σ v1 ⊗ v2 ⊗ v3) =∑τ∈S3

1

6τ(σ v1 ⊗ v2 ⊗ v3)

We have the following proposition:

Proposition 20. Suppose V and W are locally finite-dimensional and f : V → W an isomor-

phism. Then Φ satisfies

1. ΦV = ⊗3f−1 ◦ ΦW ◦ ⊗3f (naturality)

2. ΦV ∗ = (ΦV )∗ (self-duality)

Proof. It is easy to see that the isomorphism ⊗3f commutes with the action of S3. Indeed, for

any σ v1 ⊗ v2 ⊗ v3 ∈ ⊗3V , and simple transposition τ ∈ S3, τ(l) = l + 1 we have:

⊗3f (τ (σ v1 ⊗ v2 ⊗ v3)) = ⊗3f(−(−1)(|vl|−1)(|vl+1|−1)τi,j,kσ vτ(1) ⊗ vτ(2) ⊗ vσ(3)

)= −(−1)(|vl|−1)(|vl+1|−1)τi,j,kσ ⊗3 f

(vτ(1) ⊗ vτ(2) ⊗ vτ(3)

)= −(−1)(|vi|−1)(|vi+1|−1)τi,j,kσ f(vσ(1))⊗ f(vσ(2))⊗ f(vσ(3))

= τ(⊗3f(σ v1 ⊗ v2 ⊗ v3)

)And, therefore,

70

Page 71: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

⊗3f−1 ◦ ΦW ◦ ⊗3f(σ v1 ⊗ v2 ⊗ v3) = ⊗3f−1 ◦ ΦW (σ f(v1)⊗ f(v2)⊗ f(v3))

= ⊗3f−1

(∑τ∈S3

1

6τ (σ f(v1)⊗ f(v2)⊗ f(v3))

)

=∑τ∈S3

1

6⊗3 f−1 (τ (σ f(v1)⊗ f(v2)⊗ f(v3)))

=∑τ∈S3

1

6τ (⊗3f−1(σ f(v1)⊗ f(v2)⊗ f(v3)))

=∑τ∈S3

1

6τ (σ v1 ⊗ v2 ⊗ v3)

= ΦV (σ v1 ⊗ v2 ⊗ v3)

which establishes 1. Now, select a basis {vi}i∈N for V (i.e. select a basis for each Vn and take

the disjoint union). For each vi, let vi denote the associated covector vi(vj) = δ(i, j). Since V

is locally finite-dimensional, {vi}i∈N forms a dual basis for V ∗ and {σ vi1 ⊗ vi2 ⊗ vi3} forms a

basis for ⊗3V ∗. By definition

(ΦV )∗(σ vi1 ⊗ vi2 ⊗ vi3)(σ′ vj1 ⊗ vj2 ⊗ vj3) = (σ vi1 ⊗ vi2 ⊗ vi3) (ΦV (σ′ vj1 ⊗ vj2 ⊗ vj3))

This is non-zero if and only if σ = σ′ and ∃τ ∈ S3 such that ik = jτ(k) for k = 1, 2, 3. Assume

that τ is a simple transposition τ(t) = t+ 1 (the general case follows since the symmetric group

is generated by simple transpositions). Hence,

(σ vi1 ⊗ vi2 ⊗ vi3) ((ΦV )∗(σ vj1 ⊗ vj2 ⊗ vj3)) = −(−1)(|vjt |−1)(|vjt+1|−1)

= −(−1)(|vjτ(t+1)

|−1)(|vjτ(t) |−1)

= −(−1)(|vit+1 |−1)(|vit |−1)

= ΦV ∗(σ vi1 ⊗ vi2 ⊗ vi3)(σ vj1 ⊗ vj2 ⊗ vj3)

which establishes 2.

Recall from section 9, a twisted algebra (A,×) is graded Lie-admissable if, as a map ⊗3A →A,

71

Page 72: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

(×(×⊗ 1)−×(1⊗×)) ΦA = 0

This motivates the following obvious definition:

Definition 42. A twisted co-algebra (A, ∂) is co-lie-admissable if, as a map A→ ⊗3A,

ΦA ((∂ ⊗ 1)∂ − (1⊗ ∂)∂) = 0

The ∗-product on rooted multigraphs defines a morphism

∗ : Grairred

acyc ⊗Grairred

acyc → Grairred

acyc

Since Grairred

acyc is locally finite-dimensional, dualizing yields a morphism

∗∗ : Grairred∗acyc → Gra

irred∗acyc ⊗Gra

irred∗acyc

Let B denote the standard basis of multigraphs for Grairredacyc . Applying the projection map

π : Grairredacyc → Grairred

acyc to B yields a spanning set for Grairred

acyc . We can obtain a basis

B for Grairred

acyc by selecting an element from each equivalence class in the quotient π(B)/±.

Since Grairred

acyc is locally finite-dimensional, a basis B for Grairred

acyc induces an isomorphism

ψB : Grairred

acyc∼→ Gra

irred∗acyc .

Proposition 21. The isomorphism ψB : Grairred

acyc∼→ Gra

irred∗acyc is well-defined.

Proof. Let B and B′ be two bases obtained from the standard basis B. Every element x ∈ B has

an associated covector x∗ ∈ Grairredacyc∗ defined by

x∗(b) =

{1 if b = x

0 otherwise

for every b ∈ B and extended by linearity. Since Grairred

acyc is locally finite-dimensional, the set

B∗ := {x∗ | x ∈ B} forms a dual basis for Grairred∗acyc . The assignment x 7→ x∗ is therefore an

isomorphism ψB : Grairred

acyc → Grairred∗acyc which corresponds to the identity matrix with respect

to the bases B and B∗. The important fact in this case is that the change of basis matrix M

72

Page 73: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

transforming B into B′ is a diagonal matrix with 1’s and −1’s in the diagonal. In particular,

M is orthogonal and satisfies (M−1)T = MT = M . The change of basis x′ = Mx in Grairred

acyc

corresponds to a change of basis x∗′

= (M−1)Tx∗ in the dual space Grairred∗acyc . Thus, the

isomorphism ψB′ : Grairred

acyc → Grairred∗acyc induced by B′ is represented by the matrix

(M−1)TM−1 = MM−1 = 1

It follows that the isomorphism ψB : Grairred

acyc → Grairred∗acyc is independent of our particular choice

of B.

We can apply the isomorphism ψB betweenGrairred

acyc and its dual to the morphism ∗∗ : Grairred∗acyc →

Grairred∗acyc ⊗Gra

irred∗acyc to obtain a coproduct on Gra

irred

acyc .

Definition 43. The co-product

∆ : Grairred

acyc → Grairred

acyc ⊗Grairred

acyc

is defined as the composition

∆ = ψ−1B ⊗ ψ

−1B ◦ ∗

∗ ◦ ψB

Proposition 22. The twisted co-algebra (Grairred

acyc ,∆) is co-Lie-admissable.

Proof. For notational convenience, put G = Grairred

acyc . In Theorem 6 we proved that the algebra

(G, ∗) is graded Lie-admissable. Hence, by definition,

(∗(∗ ⊗ 1)− ∗(1⊗ ∗)) ΦG = 0

Dualizing, we get

(ΦG)∗ ((∗∗ ⊗ 1) ∗∗ − (1⊗ ∗∗)∗∗) = 0

Thus,

73

Page 74: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

0 = ⊗3ψ−1B ◦ (ΦG)∗ ((∗∗ ⊗ 1) ∗∗ − (1⊗ ∗∗)∗∗) ◦ ψB

= (⊗3ψ−1B ◦ (ΦG)∗ ◦ ⊗3ψB)

((⊗2ψ−1

B ◦ ∗∗ ◦ ψB)⊗ 1)(⊗2ψ−1

B ◦ ∗∗ ◦ ψB)− (1⊗ (⊗2ψ−1

B ◦ ∗∗ ◦ ψB))(⊗2ψ−1

B ◦ ∗∗ ◦ ψB)

)= (⊗3ψ−1

B ◦ (ΦG)∗ ◦ ⊗3ψB) ((∆⊗ 1)∆− (1⊗∆)∆)

However, by parts 1 and 2 of Proposition 20, we have

⊗3ψ−1B ◦ (ΦG)∗ ◦ ⊗3ψB = ΦG

So,

ΦG ((∆⊗ 1)∆− (1⊗∆)∆)

as desired.

For a particular choice of B, ∆ is given on basis elements x ∈ Grairredacyc [n] by the formula

∆(x) :=∑a,b∈B

ε(a,b)x∈σa∗b

σ ε(a, b)a⊗b ∈⊕i+j=n

IndSnSi×SjGrairred

acyc [i]⊗Grairredacyc [j] =(Gra

irred

acyc ⊗Grairred

acyc

)n

(21)

where ε(a, b) is ±1, σ ∈ Sn is a permutation shuffling the vertices of a and b, and the notation

ε(a, b)x ∈ σa ∗ b indicates that the term ε(a, b)x is represented in the unique B-decomposition of

σ a ∗ b.

Intuitively, ∆ is the opposite of ∗: whereas G1 ∗G2 is the sum of all possible graftings of G2 onto

G1, ∆(G) is the sum of all possible ‘splitting aparts’ of G.

Example 31.

•2

•1

� =

•1

�⊗

•1

74

Page 75: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

Example 32.

•2 • 3

• 1

� =

•2

•1

⊗•1

�+ (23)

•2

•1

⊗•1

with ‘(23)’ denoting, in cycle notation, the (2, 1)-shuffle σ ∈ S3 given by 1 7→ 1, 2 7→ 3, 3 7→ 2.

Example 33.

•2

•3

• 1

•4

= (234) •2

• 1

•3

⊗•1

�− (123)

•1

•2

•2 • 1

+

•2

•3

• 1 ⊗•1

16 Loday’s Rigidity Theorem

In [7], Loday formalizes the notion of a generalized bialgebra, generalizing the classical notion of

a Hopf algebra, among other things. A type of generalized bialgebra is determined by an algebra

structure, a co-algebra structure, and a compatibility relation between them. The algebra and

co-algebra structures are encoded in the data of a finitely generated operad A and finitely gen-

erated co-operad Cc (see [7] for relevant definitions). For example, if A = As and Cc = Asc with

the standard Hopf compatibility relation, we recover the familiar Hopf algebra. Loday’s Rigidity

Theorem, a special case of the much more general Structure Theorem for generalized bialgebras,

is a powerful tool for deducing the freeness of certain types of bialgebras. We assume all operads

and co-operads are graded vector spaces over R.

Definition 44. A compatibility relation is an equation involving compositions of operations and

co-operations. A compatibility relation between an operation µ and co-operation δ is distributive

if it has the form

75

Page 76: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

δ ◦ µ =∑i

(µi1 ⊗ ...⊗ µim) ◦ ωi ◦ (δi1 ⊗ ...⊗ δin)

where

µ ∈ A(n), µij ∈ A(kj),

δ ∈ C(m), δij ∈ C(lj),m∑i=1

ki = r =

n∑i=1

li,

ωi ∈ R[Sr]

Definition 45. Let H be a generalized A− Cc bialgebra. The primitive part of H, denoted by

Prim(H), is the subset

PrimH := {x ∈ H | δ(x) = 0 ∀δ ∈ Cc(n), n ≥ 2}

Rigidity Theorem for Generalized Bialgebras. LetA−Cc be a type of generalized bialgebra

such that the following hypotheses hold:

H0 The operad C is finitely generated and for every generating operation µ and generating

co-operation δ, there is a distributive compatibility relation.

H1 The free A-algebra A(V ) is naturally endowed with an A−Cc bialgebra structure, inducing

a natural co-algebra map ϕ : A(V )→ Cc(V ).

H2 The natural co-algebra map ϕ : A(V )→ Cc(V ) is an isomorphism.

Then any A− Cc bialgebra H is both free and co-free over its primitive elements, i.e.

A(PrimH) ∼= H ∼= Cc(PrimH)

Note that Loday’s Rigidity Theorem generalizes trivially to twisted bialgebras.

Grairred

acyc contains two important subalgebras.

Definition 46. A1 ⊆ Grairred

acyc is the subcollection spanned by irreducible, internally acyclic

multigraphs containing no vertices with more than one outgoing edge.

76

Page 77: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

Definition 47. A≥2 ⊂ Grairred

acyc is the subcollection spanned by irreducible, internally acyclic

multigraphs containing no ordinary vertices with fewer than two outgoing edge.

It is clear from definitions that A1 and A≥2 are both subalgebras and subcoalgebras of Grairred

acyc .

Definition 48. For every pair of multigraphs a, b ∈ Grairred

acyc , let n(a, b) denote the product

(#V ert(a)) (#Out(b)) where, as usual, V ert(a) denotes the set of vertices in a and Out(b)

denotes the set of vertices in b directly connected to the root.

Note that if b ∈ A1, n(a, b) = #V ert(a).

Theorem 7. For every a, b ∈ A1, ∆ and ∗ satisfy the following compatibility relation:

∆(a ∗ b) = ∆(a) ∗ (1⊗ b) + ∆(a) ∗ (b⊗ 1) + (a⊗ 1) ∗∆(b) + n(a, b)a⊗ b (22)

Proof. Since ∆, ∗, and ⊗ are bilinear, it suffices to prove equation (22) on a basis. Let B be any

basis for Grairred

acyc obtained from the standard basis of multigraphs for Grairredacyc and let a, b ∈ B.

By definition

∆(a ∗ b) = ∆

(∑v∈a

a ∗v b

)=∑v∈a

∆(a ∗v b)

We have ∆(a) =∑i ai(1) ⊗ a

i(2) where ai(1) and ai(2) are full subgraphs of a (no signs are present

since a ∈ A1). In Sweedler’s notation, all indices are suppressed: ∆(a) =∑

(a) a(1)⊗a(2). Hence,

∆(a∗v b) =∑

(a∗vb)(a∗v b)(1)⊗ (a∗v b)(2). The edge connecting b to the vertex v in a is contained

in (a ∗v b)(1), (a ∗v b)(2), or neither. Thus,

∆(a ∗v b) =∑a(2)3v

a(1) ⊗ (a(2) ∗v b) +∑a(1)3v

(a(1) ∗v b)⊗ a(2) +∑(b)

(a ∗v b(1))⊗ b(2) + a⊗ b

Hence,

77

Page 78: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

∆(a ∗ b) =∑v∈a

∑a(2)3v

a(1) ⊗ (a(2) ∗v b) +∑a(1)3v

(a(1) ∗v b)⊗ a(2) +∑(b)

(a ∗v b(1))⊗ b(2) + a⊗ b

=∑(a)

∑v∈a(2)

a(1) ⊗ (a(2) ∗v b) +∑(a)

∑v∈a(1)

(a(1) ∗v b)⊗ a(2) +∑(b)

∑v∈a

(a ∗v b(1))⊗ b(2) +∑v∈a

a⊗ b

=∑(a)

a(1) ⊗ (a(2) ∗ b) +∑(a)

(a(1) ∗ b)⊗ a(2) +∑(b)

(a ∗ b(1))⊗ b(2) + n(a, b)a⊗ b

= ∆(a) ∗ (1⊗ b) + ∆(a) ∗ (b⊗ 1) + (a⊗ 1) ∗∆(b) + n(a, b)a⊗ b

Theorem 8. Bialgebras of type PreLie–PreLiec with compatibility relation defined by equation

(22) satisfy the hypotheses of Loday’s Rigidity Theorem.

Sketch of Proof. Clearly, equation (22) js distributive. We must exhibit a natural co-pre-Lie

structure on the free pre-Lie algebra PreLie(V ) over a symmetric collection V . Let I ⊂PreLie(V ) denote the ideal spanned by elements of form (∗(∗ ⊗ 1)− ∗(1⊗ ∗)) (a⊗b⊗c−a⊗c⊗b).The assignment ∆(v) = 0,∀v ∈ V extends to a coproduct on PreLie(V ) if and only if ∆(I) = 0.

And indeed, for every a, b, c ∈ PreLie(V ),

∆((ab)c− a(bc)) = ∆(ab)(1⊗ c) + ∆(ab)(c⊗ 1) + (ab⊗ 1)∆(c) + n(ab)ab⊗ c

− (∆(a)(1⊗ bc) + ∆(a)(bc⊗ 1) + (a⊗ 1)∆(bc) + n(a)a⊗ bc)

= ∆(a) (1⊗ bc+ b⊗ c+ c⊗ b+ bc⊗ 1) + (a⊗ 1)∆(b)(1⊗ c+ c⊗ 1)

+ n(a)(a⊗ bc) + n(a)(ac⊗ b) + (ab⊗ 1)∆(c) + n(ab)(ab⊗ c)

− (∆(a)(1⊗ bc+ bc⊗ 1) + (a⊗ 1)∆(b)(1⊗ c+ c⊗ 1)

+ (ab⊗ 1)∆(c) + n(b)(ab⊗ c) + n(a)(a⊗ bc))

= ∆(a)(b⊗ c+ c⊗ b) + n(a)(ac⊗ b) + n(ab)(ab⊗ c)− n(b)(ab⊗ c)

where, for notational convenience, all products have been written as concatenations and n(a) =

#V ert(a). Since n(ab) = n(a) + n(b), the last three terms simplify to

n(a)(ac⊗ b) + n(ab)(ab⊗ c)− n(b)(ab⊗ c) = n(a)(ac⊗ b+ ab⊗ c)

Thus,

78

Page 79: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

∆((ab)c− a(bc)) = ∆(a)(b⊗ c+ c⊗ b) + n(a)(ac⊗ b+ ab⊗ c)

which is symmetric in the second two factors. So, indeed, ∆(v) = 0 extends to a coproduct on

PreLie(V ). We must show that this coproduct is co-pre-Lie. We proceed by induction, noting

that the base case is trivial, since (∆⊗ 1)∆(v)− (1⊗∆)∆(v) = 0 for every v ∈ V . Suppose that

the co-associators of a and b are co-pre-Lie and consider the co-associator of their product ab.

We compute, in Sweedler’s notation,

(∆⊗ 1)∆(ab)− (1⊗∆)∆(ab) = (∆⊗ 1− 1⊗∆)(∆(a)(1⊗ b+ b⊗ 1) + (a⊗ 1)∆(b) + n(a)a⊗ b)

= (∆⊗ 1− 1⊗∆)(a(1) ⊗ a(2)b+ a(1)b⊗ a(2) + ab(1) ⊗ b(2) + n(a)a⊗ b)

= ∆(a(1))⊗ a(2)b+ ∆(a(1)b)⊗ a(2) + ∆(ab(1))⊗ b(2) + n(a)∆(a)⊗ b

−(a(1) ⊗∆(a(2)b) + a(1)b⊗∆(a(2)) + ab(1) ⊗∆(b(2)) + n(a)a⊗∆(b)

)= ∆(a(1))⊗ a(2)b+ ∆(a(1))(1⊗ b+ b⊗ 1)⊗ a(2) + a(1)b(1) ⊗ b(2) ⊗ a(2)

+ n(a(1))a(1) ⊗ b⊗ a(2) + a(1) ⊗ a(2)b(1) ⊗ b(2) + a(1)b(1) ⊗ a(2) ⊗ b(2)

+ (a⊗ 1)∆(b(1))⊗ b(2) + n(a)a⊗ b(1) ⊗ b(2) + n(a)a(1) ⊗ a(2) ⊗ b

−(a(1) ⊗∆(a(2))(1⊗ b+ b⊗ 1) + a(1) ⊗ a(2)b(1) ⊗ b(2) + n(a(2))a(1) ⊗ a(2) ⊗ b

+a(1)b⊗∆(a(2)) + ab(1) ⊗∆(b(2)) + n(a)a⊗ b(1) ⊗ b(2)

)= ∆(a(1))⊗ a(2)b+ ∆(a(1))(1⊗ b+ b⊗ 1)⊗ a(2) + a(1)b(1) ⊗ b(2) ⊗ a(2)

+ n(a(1))a(1) ⊗ b⊗ a(2) + a(1)b(1) ⊗ a(2) ⊗ b(2)

+ (a⊗ 1)∆(b(1))⊗ b(2) + n(a)a(1) ⊗ a(2) ⊗ b

−(a(1) ⊗∆(a(2))(1⊗ b+ b⊗ 1) + n(a(2))a(1) ⊗ a(2) ⊗ b

+a(1)b⊗∆(a(2)) + ab(1) ⊗∆(b(2)))

The three terms with non-trivial coefficients simplify, via the distributivity of n(a):

n(a(1))a(1)⊗b⊗a(2)+n(a)a(1)⊗a(2)⊗b−n(a(2))a(1)⊗a(2)⊗b = n(a(1))(a(1) ⊗ a(2) ⊗ b+ a(1) ⊗ b⊗ a(2)

)which is symmetric in the second two factors. This leaves

79

Page 80: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

∆(a(1))⊗ a(2)b+ ∆(a(1))(1⊗ b+ b⊗ 1)⊗ a(2) + a(1)b(1) ⊗ b(2) ⊗ a(2)

+ a(1)b(1) ⊗ a(2) ⊗ b(2) + (a⊗ 1)∆(b(1))⊗ b(2)

−(a(1) ⊗∆(a(2))(1⊗ b+ b⊗ 1) + a(1)b⊗∆(a(2)) + ab(1) ⊗∆(b(2))

)= (a⊗ 1⊗ 1)

(∆(b(1))⊗ b(2) − b(1) ⊗∆(b(2))

)+ a(1)b(1) ⊗ b(2) ⊗ a(2) + a(1)b(1) ⊗ a(2) ⊗ b(2)

+ ∆(a(1))⊗ a(2)b+ ∆(a(1))(1⊗ b+ b⊗ 1)⊗ a(2)

−(a(1) ⊗∆(a(2))(1⊗ b+ b⊗ 1) + a(1)b⊗∆(a(2))

)The first term is co-pre-Lie by induction and the following two terms (considered as a sum)

a(1)b(1) ⊗ b(2) ⊗ a(2) + a(1)b(1) ⊗ a(2) ⊗ b(2) are clearly symmetric in the second two factors. We

are left with the expression

∆(a(1))⊗ a(2)b+ ∆(a(1))(1⊗ b+ b⊗ 1)⊗ a(2) − a(1) ⊗∆(a(2))(1⊗ b+ b⊗ 1)− a(1)b⊗∆(a(2))

This factors as(∆(a(1))⊗ a(2) − a(1) ⊗∆(a(2))

)(b⊗ 1⊗ 1 + 1⊗ b⊗ 1 + 1⊗ 1⊗ b) which is, again,

co-pre-Lie by induction.

Thus, PreLie(V ) is canonically endowed with the structure of a twisted co-pre-Lie co-algebra.

To complete the proof, we must show that the canonical surjection p : PreLie(V )→ PreLiec(V )

is an isomorphism. For every operation µ ∈ PreLie(n) and co-operation β ∈ PreLiec(n), we

have

(β ◦ µ)(x1 ⊗ ...⊗ xn) =∑σ∈Sn

cσxσ(1) ⊗ ...⊗ xσ(n) (23)

The assignment 〈µ, β〉 = cid defines a pairing 〈 , 〉 : PreLie(n) ⊗ PreLiec(n) → R. Suppose

0 6= u = µ(x1 ⊗ ...⊗ xn) and p(u) = 0. Take any non-trivial co-operation β of the same arity as

µ. Since p is a morphism of co-algebras,

(p⊗ ...⊗ p)((β ◦ µ)(x1 ⊗ ...⊗ xn)) = β(p(µ(x1 ⊗ ...⊗ xn))) = β(p(u)) = 0

Hence, by equation (23),

∑σ∈Sn

cσxσ(1) ⊗ ...⊗ xσ(n) = (p⊗ ...⊗ p)∑σ∈Sn

cσxσ(1) ⊗ ...⊗ xσ(n) = 0

80

Page 81: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

And in particular, 〈µ, β〉 = 0. Thus, it suffices to show that this pairing is non-degenerate.

This can be done, for example, by selecting bases for PreLie(n) and PreLiec(n), computing the

pairing on these basis elements, and then calculating the determinant. We have not yet been

able to produce a succinct formula for this pairing, so we leave this step incomplete.

Corollary 6. A1 is a freely generated as a twisted pre-Lie algebra.

Proof. This follows immediately from theorems 7, 8, and Loday’s Rigidity Theorem.

On A≥2, the relationship between ∗ and ∆ is a bit more subtle. To describe this relationship,

we must introduce a bit of additional notation. Fix a basis B of multigraphs for Grairred

acyc . For

multigraphs a, b ∈ B,

a ∗ b =∑v∈a

w∈Out(b)

sb(w) a ∗v,w b

where sb(w) = ±1 and a∗v,w b is the multigraph formed by decapitating b, grafting the left-most

free output from w onto v, and grafting all other free outputs onto the root vertex of a. The

sign sb(w) can be computed explicitly by counting the number x of free outputs to the left of

the left-most free output from w and then computing (−1)x. Thus,

∆(a ∗ b) = ∆

∑v∈a

w∈Out(b)

sb(w) a ∗v,w b

=∑v∈a

w∈Out(b)

sb(w)∆(a ∗v,w b) (24)

On the other hand,

∆(a) =∑(a)

εa(a(1), a(2))a(1) ⊗ a(2)

where a(1) and a(2) are full subgraphs of a such that, in its unique B-decomposition, εa(a(1), a(2))a(1)∗a(2) contains the term a with coefficient 1. For notational convenience, we will suppress the sub-

script a in εa(a(1), a(2)) when there is no potential for confusion.

Theorem 9. For every pair of multigraphs a, b ∈ A≥2 with the property that every ordinary

vertex is directly connected to the root, ∆ and ∗ satisfy the following compatibility relation:

81

Page 82: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

∆(a ∗ b) =∑(a)

ε(a(1), a(2))(a(1) ⊗ (a(2) ∗ b) + (−1)(|b|−1)(|a(2)|−1)(a(1) ∗ b)⊗ a(2)

)+∑(b)

ε(b(1), b(2))(

(a ∗ b(1))⊗ b(2) + (−1)(|a|−1)(|b(1)|−1)b(1) ⊗ (a ∗ b(2)))

+ n(a, b)a⊗ b

(25)

In other words, the five-term sum ∆(a ∗ b) = ∆(a) ∗ (1 ⊗ b) + ∆(a) ∗ (b ⊗ 1) + (a ⊗ 1) ∗∆(b) +

(1⊗ a) ∗∆(b) + n(a, b)a⊗ b with the appropriate Koszul signs.

Proof. The multigraph a∗v,wb can be broken up into two pieces in five different ways, as expressed

in the following sum

∆(a ∗v,w b) =∑a(2)3v

ε(a(1), (a(2) ∗v,w b))a(1) ⊗ (a(2) ∗v,w b) +∑a(1)3v

ε((a(1) ∗v,w b), a(2))(a(1) ∗v,w b)⊗ a(2)

+∑b(1)3w

ε((a ∗v,w b(1)), b(2))(a ∗v,w b(1))⊗ b(2) +∑b(2)3w

ε(b(1), (a ∗v,w b(2)))b(1) ⊗ (a ∗v,w b(2))

+ ε(a, b)a⊗ b(26)

with an implied subscript of a ∗v,w b for each ε above. Unraveling definitions, we see that

ε(a(1), (a(2) ∗v,w b)) = εa(a(1), a(2))

ε((a(1) ∗v,w b), a(2)) = (−1)(|b|−1)(|a(2)|−1)εa(a(1), a(2))

ε((a ∗v,w b(1)), b(2)) = εb(b(1), b(2))

ε(b(1), (a ∗v,w b(2))) = (−1)(−1)|a|(−1)|a||b(1)|εb(b(1), b(2))

Substituting these identities into equation (26) and equation (26) into equation (24), we obtain

by commuting sums

82

Page 83: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

∆(a ∗ b) =∑v∈a

w∈Out(b)

s(w)∆(a ∗v,w b)

=∑(a)

ε(a(1), a(2))∑v∈a(2)

w∈Out(b)

sb(w)a(1) ⊗ (a(2) ∗v,w b)

+∑(a)

(−1)(|b|−1)(|a(2)|−1)ε(a(1), a(2))∑v∈a(1)

w∈Out(b)

sb(w)(a(1) ∗v,w b)⊗ a(2)

+∑(b)

εb(b(1), b(2))∑v∈a

w∈Out(b)∩b(1)

sb(w)(a ∗v,w b(1))⊗ b(2)

+∑(b)

(−1)|a||b(1)|+|a|+1ε(b(1), b(2))∑v∈a

w∈Out(b)∩b(2)

sb(w)b(1) ⊗ (a ∗v,w b(2))

+∑v∈a

w∈Out(b)

sb(w)εa∗v,wb(a, b)a⊗ b

Applying the fact that Out(b)∩b(1) = Out(b(1)) and Out(b)∩b(2) = Out(b(2)), this can be further

simplified:

∆(a ∗ b) =∑(a)

ε(a(1), a(2))a(1) ⊗ (a(2) ∗ b) +∑(a)

ε(a(1), a(2))(−1)(|b|−1)(|a(2)|−1)(a(1) ∗ b)⊗ a(2)

+∑(b)

ε(b(1), a(2))(a ∗ b(1))⊗ b(2) + ε(b(1), a(2))(−1)(|a|−1)(|b(1)|−1)b(1) ⊗ (a ∗ b(2))

+∑v∈a

w∈Out(b)

sb(w)εa∗v,wb(a, b)a⊗ b

Since εa∗v,wb(a, b) = sb(w), the final sum reduces to n(a, b)a⊗ b, which completes the proof.

We conjecture that A≥2 is a free twisted Lie-admissable algebra, generated by elementary multi-

graphs of valence ≥ 2.

17 A Useful Consequence of Kunneth’s Theorem

We pause here to prove a simple consequence of Kunneth’s theorem for cohomology. This corol-

lary will prove useful in section 18.

83

Page 84: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

Suppose V and W are cochain complexes of right and left R-modules, respectively. The tensor

product V ⊗RW is the cochain complex

(V ⊗RW )n =⊕i+j=n

V i ⊗RW j

with coboundary ∂ : (V ⊗R W )n → (V ⊗R W )n+1 given on pure tensors v ⊗ w ∈ V i ⊗R W j by

the formula:

∂(v ⊗ w) = (∂(v)⊗ w) + (−1)i(v ⊗ ∂(w)) ∈ (V ⊗RW )n+1

If R is commutative, V is a cochain complex of R-modules. In this case, tensor products such

as T kV can be formed by tensoring V with itself repeatedly. The tensor algebra T (V ) is the

pointwise sum of all T kV . For each k, there is an action of the symmetric group Sk on the

components of T kV , which commutes with the differential. We can therefore consider the sub-

complex SkV ⊂ T kV of symmetric invariants. The pointwise sum S(V ) :=⊕

k Sk(V ) is the

symmetric algebra of V .

Note that there is a graded version of this construction. Suppose V is a cochain complex of graded

R-modules. Then the tensor product T kV is also a complex of graded modules, with gradings

inherited from V . Sk then acts on the complex T kV via the Koszul sign rule. Explicitly, for

every simple transposition σα : α↔ α+ 1, we put

σα(v1 ⊗ v2 ⊗ ...⊗ vk) = (−1)|vα||vα+1|v1 ⊗ v2 ⊗ ...vα+1 ⊗ vα ⊗ ...⊗ vk

It is an easy verification that the formula above respects all of the relations among simple

transpositions. Indeed, ∀α < k:

σα ◦ σα(v1 ⊗ v2 ⊗ ...⊗ vk) = (−1)2|vα||vα+1|v1 ⊗ v2 ⊗ ...⊗ vk = v1 ⊗ v2 ⊗ ...⊗ vk

∀α < β < k, β 6= α+ 1:

σα ◦ σβ(v1 ⊗ v2 ⊗ ...⊗ vk) = (−1)|vα||vα+1|+|vβ ||vβ+1|v1 ⊗ ...vα+1 ⊗ vα ⊗ ...vβ+1 ⊗ vβ ⊗ ...vk

= σβ ◦ σα(v1 ⊗ v2 ⊗ ...⊗ vk)

84

Page 85: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

and ∀α < k:

σα ◦ σα+1 ◦ σα(v1 ⊗ v2 ⊗ ...⊗ vk) = (−1)|vα||vα+1|+|vα||vα+2|+|vα+1||vα+2|v1 ⊗ ...⊗ vα+2 ⊗ vα+1 ⊗ vα ⊗ ...⊗ vk

= σα+1 ◦ σα ◦ σα+1(v1 ⊗ v2 ⊗ ...⊗ vk)

Hence, the action defined above extends uniquely to an action of the symmetric group. This

action, again, commutes with the differential. We can therefore consider the subcomplex SkV ⊂T kV of invariants. The pointwise sum S(V ) :=

⊕k S

k(V ) is the free associative, supercommu-

tative algebra over V .

Proposition 23. Let V be a cochain complex of graded vector spaces. Then, there is a natural

isomorphism S(H•(V )) ∼= H•(S(V )).

Proof. By Kunneth’s theorem, we have T k(H•(V )) ∼= H•(T k(V )), for ever k ∈ N, by a natural

isomorphism

φ : [v1]⊗ [v2]⊗ ...⊗ [vk] 7→ [v1 ⊗ v2 ⊗ ...⊗ vk]

Clearly, φ commutes with the (graded) action of the symmetric group. Hence, φ descends to

a natural isomorphism Sk(H•(V )) ∼= H•(Sk(V )). Assembling these maps for every k ∈ N, we

obtain a natural isomorphism S(H•(V )) ∼= H•(S(V )), as desired.

Corollary 7. Suppose A is an associative, supercommutative algebra, B is a graded A-module,

and ∂ : A → B is a linear degree-preserving map. Suppose, in addition, that the following

conditions hold:

1. A is freely generated by a subspace V ⊂ A as an algebra over Com

2. B is freely generated over A by a subspace W ⊂ B as a module over an algebra over Com

3. ∂(V ) ⊂W

4. ∂(a1a2) = a1∂(a2) + ∂(a1)a2, ∀a1, a2 ∈ A

Then ker ∂ is the free associative, supercommutative algebra generated by ker ∂ ∩ V .

Proof. Let ∂ = ∂|V , and let C• be the following cochain complex of graded vector spaces

V∂→W → 0→ 0→ ...

85

Page 86: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

S(C•) is the cochain complex

S(V )S(∂)−→

∞⊕n=1

(n−1⊕k=0

⊗kV ⊗W ⊗⊗n−k−1V

)Sn−→ ...

By 1, there is an algebra isomorphism p0 : S(V )∼→ A which restricts to the identity on V , and

by 2, there is a A-module isomorphism p1 :⊕∞

n=1

(⊕n−1k=0 ⊗kV ⊗W ⊗⊗n−k−1V

)Sn ∼→ B which

restricts to the identity on W .The diagram

S(V )⊕∞

n=1

(⊕n−1k=0 ⊗kV ⊗W ⊗⊗n−k−1V

)Sn

A B

S(∂)

p0 p1

commutes trivially on V . Commutativity on the full symmetric algebra is then established by

induction. Indeed, if the diagram commutes for a, b ∈ S(v) then

p1(S(∂)(a� b)) = p1(S(∂)(a)� b+ a� S(∂)(b))

= p1(S(∂)(a))p0(b) + p0(a)p1(S(∂)(b))

= ∂(p0(a))p0(b) + p0(a)∂(p0(b))

= ∂(p0(a� b))

Applying Proposition (23) in homological degree 0 to the complex S(C•), we get kerS(∂) =

S(ker ∂). Therefore, ker ∂ = p0(kerS(∂)) = p0(S(ker ∂)), as desired.

We have the following obvious analogues of proposition 23 and corollary 7 in the twisted context:

Proposition 24. Let V be a cochain complex of symmetric collections. Then, there is a natural

isomorphism S(H•(V )) ∼= H•(S(V )).

Corollary 8. Suppose A is an associative, supercommutative twisted algebra, B is a twisted

A-module, and ∂ : A → B is a morphism of symmetric collections. Suppose, in addition, that

the following conditions hold:

1. A is freely generated by a subcollection V ⊂ A as a twisted algebra over Com

86

Page 87: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

2. B is freely generated over A by a subcollection W ⊂ B as a twisted module over an algebra

over Com

3. ∂(V ) ⊂W

4. ∂(a1a2) = a1∂(a2) + ∂(a1)a2, ∀a1, a2 ∈ A

Then ker ∂ is the free twisted associative, supercommutative algebra generated by ker ∂ ∩ V .

18 Chevalley-Eilenberg Cohomology

Consider a Lie algebra h and a h-module W .

Definition 49. The Chevalley-Eilenberg cohomology H•(h,W ) of h with coefficients in W is

the cohomology of the cochain complex (C•(h,W ), δCE) where

Cn(h,W ) := Lin (∧nh,W )

and

(δf)(h1 ∧ ... ∧ hn) :=∑

1≤i≤n+1

(−1)i+1hif(h1 ∧ ... ∧ hi ∧ ... ∧ hn+1)

+∑

1≤i≤j≤n+1

(−1)i+jf([hi, hj ] ∧ h1 ∧ ... ∧ hi ∧ ... ∧ hj ∧ ... ∧ hn+1)

Suppose H o G is a Lie group and W is a H o G-module. The action of H ⊂ H o G on W

induces an infinitesimal action of the Lie algebra h on W . We can therefore consider the cochain

complex (C•(h,W ), δCE). Since G acts on H by group homomorphisms, the identity element

1 ∈ H is fixed by G. Thus, by differentiating, we obtain a linear action of G on h. On the

other hand, G acts on W via the inclusion G ⊂ H o G. Let G act on the nth component

Cn(h,W ) = Lin (∧nh,W ) by “conjugation”:

(g · f)(h1 ∧ ... ∧ hn) = gf(g−1h1 ∧ ... ∧ g−1hn)

Proposition 25. For every component Cn(h,W ), let CnG(h,W ) denote the subspace of G-

invariants. The restrictions δCE : Cn(h,W )→ Cn+1(h,W ) are well-defined for every n.

87

Page 88: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

Proof. We want to show that the differential δCE : Cn(h,W )→ Cn+1(h,W ) commutes with the

action of G for every integer n. Let g ∈ G, f ∈ Lin (∧nh,W ) = Cn(h,W ), and h1, ..., hn+1 ∈ h.

We have

(δCE(g · f))(h1 ∧ ... ∧ hn+1)

=∑

1≤i≤n+1

(−1)i+1hi

(gf(g−1h1 ∧ ... ∧ g−1hi ∧ ... ∧ g−1hn+1)

)+

∑1≤i≤j≤n+1

(−1)i+jgf(g−1[hi, hj ] ∧ g−1h1 ∧ ... ∧ g−1hi ∧ ... ∧ g−1hj ∧ ...g−1 ∧ hn+1)

Let exp : h→ H denote the exponential map. Notice for the first summand that

hi

(gf(g−1h1 ∧ ... ∧ g−1hi ∧ ... ∧ g−1hn+1)

)=

d

dt|0 exp(thi)

(gf(g−1h1 ∧ ... ∧ g−1hi ∧ ... ∧ g−1hn+1)

)=

d

dt|0 g

((g−1exp(thi)g)f(g−1h1 ∧ ... ∧ g−1hi ∧ ... ∧ g−1hn+1)

)= g

((g−1hi)f(g−1h1 ∧ ... ∧ g−1hi ∧ ... ∧ g−1hn+1)

)For the second summand, notice that g−1[hi, hj ] = [g−1hi, g

−1hj ], since the derivative at the

identity of a Lie group homomorphism is a morphism of Lie algebras. Hence,

(δCE(g · f))(h1 ∧ ... ∧ hn+1)

=∑

1≤i≤n+1

(−1)i+1g(

(g−1hi)f(g−1h1 ∧ ... ∧ g−1hi ∧ ... ∧ g−1hn+1))

+∑

1≤i≤j≤n+1

(−1)i+jgf([g−1hi, g−1hj ] ∧ g−1h1 ∧ ... ∧ g−1hi ∧ ... ∧ g−1hj ∧ ...g−1 ∧ hn+1)

= g · (δf)(h1 ∧ ... ∧ hn+1)

Thus, δCE commutes with the action of G and, therefore, takes G-invariants to G-invariants.

It is therefore sensible to consider the subcomplex C•G(h,W ) ⊂ C•(h,W ) of G-invariants. We

denote the cohomology of this subcomplex by H•G(h,W ).

88

Page 89: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

If W is a graded vector space and the h-action is degree-preserving, C•(h,W ) is a cochain complex

of graded vector spaces and the obvious analogue of proposition 25 holds. In our case, W is in fact

a symmetric collection {W [k]}k∈N. h acts on each W [k] by an action which commutes with the

action of Sk. The definitions of C•(h,W ) and δCE in the twisted context are virtually identical:

Cn(h,W ) is the symmetric collection defined component-wise by

Cn(h,W )[k] = Lin (∧nh,W [k])

Likewise, δnCE is a morphism of symmetric collections with kth component equal to the usual

Chevalley-Eilenberg differential Cn(h,W [k]) → Cn+1(h,W [k]). Clearly, proposition 25 holds in

the twisted case as well.

Now, consider the Lie group G∞m = NGL∞m oGL(V ) of infinity jets of origin-preserving local dif-

feomorphisms. The symmetric collection W of differential operators on alternating multi-vector

fields is a module over G∞m . In light of proposition 25, we may consider the cochain complex

C•GL(V )(ngl∞m ,W) of symmetric collections and the subcomplex C•GL(V )(ngl

∞m ,W(α1, ..., αn)) for

natural numbers α1, ..., αn.

Proposition 26. Restriction to the standard fiber induces isomorphisms

Natfieldsm (α1, ..., αn) ∼= H0

GL(V ) (ngl∞m ,W(α1, ..., αn))

for natural numbers α1, ..., αn and

Natfieldsm

∼= H0GL(V ) (ngl∞m ,W)

Proof. This is simply a matter of parsing definitions. The first two components of the cochain

complex C•GL(V ) (ngl∞m ,W) are the vector spaces

C0GL(V ) (ngl∞m ,W) = WGL(V )

C1GL(V ) (ngl∞m ,W) = LinGL(V ) (ngl∞m ,W)

Reflecting on the definition of the Chevalley-Eilenberg differential, we see that the map δCE :

WGL(V ) → LinGL(V ) (ngl∞m ,W) is given by (δCEf)(g) = gf for f ∈ WGL(V ) and g ∈ ngl∞m .

Hence,

89

Page 90: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

H0GL(V ) (ngl∞m ,W) = ker δ : WGL(V ) → LinGL(V ) (ngl∞m ,W)

= {f ∈WGL(V ) | gf = 0 ∀g ∈ ngl∞m}

Since the Lie group NGL∞m is both connected and simply connected,

{f ∈WGL(V ) | gf = 0 ∀g ∈ ngl∞m} =(WGL(V )

)NGL∞mBy proposition 16, restriction to the standard fiber defines a linear isomorphism Natfields

m∼=(

WGL(V ))NGL∞m . Thus, Natfields

m∼= H0

GL(V ) (ngl∞m ,W), as desired. The proof of the first iso-

morphism is absolutely analogous, replacing W above with W(α1, ..., αn) and Natfieldsm with

Natfieldsm (α1, ..., αn).

The map f defined in section 11 gives a correspondence between a particular space of rooted

multigraphs and C0GL(V ) (ngl∞m ,W) = WGL(V ). We wish to define maps analogous to f for other

components of the cochain complex C•GL(V ) (ngl∞m ,W). As a GL(V )-module

ngl∞m∼=∞⊕i=1

SiV ∗ ⊗ V

Hence,

CnGL(V ) (ngl∞m ,W) = LinGL(V ) (∧nngl∞m ,W)

∼=⊕

i1,...,in∈NLinGL(V )

n∧j=1

SijV ∗ ⊗ V

,W

∼=

⊕i1,...,in∈N

n∧j=1

SijV ⊗ V ∗⊗W

GL(V )

We define a symmetric collection of rooted multigraphs which, for stable values of m, corre-

sponds isomorphically to CnGL(V ) (ngl∞m ,W). Note: a bit of extra care is required for n > 1 to

handle the wedge product in the vector space above. This is managed by making suitable identifi-

cations in the corresponding space of rooted multigraphs, as described in the following definition.

90

Page 91: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

Definition 50. For each n ≥ 0 and k ≥ 1, Let Gran[k] denote the vector space spanned by

finite ordered multigraphs containing

1. k ordinary vertices • labeled 1, ..., k with symmetric inputs and alternating outputs

2. n white vertices ◦ labeled 1, ..., n, each with exactly 1 output and at least 2 symmetric

inputs

3. 1 root vertex � with symmetric inputs and no outputs

Each Gran[k] is graded by the valence of the root and Sk acts by relabeling ordinary vertices.

Further, we make the following identifications

1. For multigraphs G and G′ in Gran[k] that differ only by an even permutation of the white

vertex labelings, we put G = G′

2. For multigraphs G and G′ in Gran[k] that differ only by an odd permutation of the white

vertex labelings, we put G = −G′

Gran

is the symmetric collection {Gran[k]}k∈N. The subcollections Gran,irred

, Gran

acyc and

Gran,irred

acyc are defined analogously to their degree-0 counterparts.

Through a procedure absolutely analogous to the construction of f in section 11, we define an

epimorphism fn

: Gran → CnGL(V ) (ngl∞m ,W) for each natural number n. In lieu of explicit defi-

nitions, we present several illustrative examples. For explicit definitions (in the slightly simpler,

but entirely analogous vector field case), see [16].

Example 34. Gra1(1, 1) is a four-dimensional vector space spanned by the rooted multigraphs

•1

•2

◦ 1

,•1

•2

◦ 1�

,•2

•1

◦ 1�

,

•1

•2

◦ 1

The first three basis elements map under f1

to elementary invariant tensors in the vector space

LinGL(V )

(S2V ⊗ V ∗, V ∗ ⊗ V ∗ ⊗ V

)and the fourth basis element maps to an elementary in-

variant tensor in LinGL(V )

(S3V ⊗ V ∗, V ∗ ⊗ V ∗

). Both of these vector spaces are canonically

regarded as subspaces of C1GL(V ) (ngl∞m ,W).

Example 35. Gra1(2) is a one-dimensional vector space spanned by the multigraph

91

Page 92: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

•1

◦ 1 �

This multigraph is mapped by f1

to an elementary invariant tensor in LinGL(V )

(S2V ⊗ V ∗,∧2V ∗ ⊗ V

),

again, regarded as a subspace of C1GL(V ) (ngl∞m ,W).

Example 36. Gra2(1, 1) is a three-dimensional vector space spanned by the multigraphs

•1

•2

◦1

◦2

,

•1

•2

◦1

◦2

,

•1

•2

◦1

◦2

Note that swapping the two white vertices in any of these three basis elements produces a minus

sign. As does swapping the two ordinary vertices in the first two multigraphs. Transposing the

two ordinary vertices in the third multigraph has no effect. All three multigraphs are mapped

by f2

to elementary invariant tensors in LinGL(V )

(∧2(S2V ⊗ V ∗), V ∗ ⊗ V ∗ ⊗ V

), regarded as a

subspace of C2GL(V ) (ngl∞m ,W).

Remark 5. A key fact about fn

is that it restricts to an isomorphism fn

: Gran(α1, ..., αk)

∼→CnGL(V ) (ngl∞m ,W(α1, ..., αk)) for any collection of α1, ..., αk with the property that n+

∑ki=1 αi <

m (note: this is analogous to part 3 of proposition 17, with the additional n on the left coming

from the n outgoing edges from the n white vertices). The proof of this fact is a straightforward

application of the Invariant Tensor Theorem.

Next, we define a differential δ on the graph complex Gra•

corresponding to the Chevalley-

Eilenberg differential on the cochain complex C•GL(V ) (ngl∞m ,W). We will define δ as a sum

δ1 + δ2 and define each piece separately.

Definition 51. Let G be a rooted multigraph in Gran[k]. Given any i, 1 ≤ i ≤ k, construct the

multigraph δ1,i(G) via the local replacement rule

92

Page 93: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

δ1,i : •i

· · ·

· · ·

v

w

7−→∑

s+u=v+1s≥2

•i

· · ·

· · ·· · ·

( )( ) alt

sym

u

s

w

−◦

• i

· · ·

· · ·

· · ·

( )sym

s

u

w

where “sym” denotes the symmetrization, obtained by summation over all (u, s)-unshuffles, and

“alt” denote the alternation, obtained by summation over all (w, 1) unshuffles, with signs deter-

mined by the signs of the permutations. The new white vertex introduced in the replacement

rule above is labeled 1 in δ1,i(G) and all other white vertices are labeled so as to retain the

original ordering on white vertices in G (i.e. white vertex j is relabeled j+1 for all j, 1 ≤ j ≤ n).

We then define δ1 as a sum over all ordinary vertices:

δ1(G) :=

k∑i=1

δ1,i(G)

It is clear that δ1 defines a morphism from Gran

to Gran+1

for every n.

Definition 52. Let G be a rooted multigraph in the symmetric collection Gran

and 1 ≤ i ≤ n.

Construct the multigraph δ2,i(G) via the local replacement rule

δ2,i :◦i

· · ·

v

7−→∑

u+s=v+1s,u≥2

◦i

◦ i+ 1

· · ·

· · ·( )sym

s

u

In the multigraph δ2,i(G), relabel each white vertex j > i as j+ 1 and retain the labelings on all

white vertices j < i. We then define δ2 as a sum over all white vertices:

93

Page 94: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

δ2(G) :=

n∑i=1

(−1)i+1δ2,i(G)

It is similarly clear that δ2 defines a morphism from Gran

to Gran+1

.

Definition 53. δ := δ1 + δ2

We give some simple examples in low cohomological degree.

Example 37.

δ

•1

= 0

Example 38.

δ

•1

•2

� =

•1

◦1

• 2

Example 39.

δ

•2

◦1 • 4

•1

•3

=

•2

◦2 ◦1

•1

•3

•4

•2

◦2 ◦1

•1

•3

•4

•2

◦2 ◦ 1

•1

•3

�• 4

δ1

+

•2

◦1 • 4

•1

•3

◦2

+

•2

◦1 • 4

•1

•3

◦ 2

+

•2

◦1

• 4

•1

•3

◦2

δ2

It is clear from the definition of δ that the following diagram commutes:

94

Page 95: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

Gra Gra1

Gra2 · · ·

C0GL(V ) (ngl∞m ,W) C1

GL(V ) (ngl∞m ,W) C2GL(V ) (ngl∞m ,W) · · ·

δ

f

δ

f1

δ

f2

δCE δCE δCE

Indeed, δ1 is precisely the graphical representation of the first part of the Chevalley-Eilenberg

differential and δ2 is precisely the graphical representation of the second. In particular, for natural

numbers α1, ..., αn satisfying 1 +∑ni=1 αi < m, f restricts to an isomorphism Gra(α1, ..., αn)

∼→C0GL(V ) (ngl∞m ,W(α1, ..., αn)) (see part 3 of proposition 17) and f

1restricts to an isomorphism

Gra1(α1, ..., αn)

∼→ C1GL(V ) (ngl∞m ,W(α1, ..., αn)) (see remark 5). In this case, the first square in

the diagram above restricts to:

Gra(α1, ..., αn) Gra1(α1, ..., αn)

C0GL(V ) (ngl∞m ,W(α1, ..., αn)) C1

GL(V ) (ngl∞m ,W(α1, ..., αn))

∼ ∼

δ

δCE

f f1

Hence, H0GL(V ) (ngl∞m ,W(α1, ..., αn)) = f

(ker δ : Gra(α1, ..., αn)→ Gra

1(α1, ..., αn)

). But by

proposition 26, H0GL(V ) (ngl∞m ,W(α1, ..., αn)) ∼= Natfields

m (α1, ..., αn). Thus, we have shown:

Theorem 10. For natural numbers α1, ..., αn satisfying 1+∑ni=1 αi < m, invariant-theoretic re-

duction and the correspondence outlined in section 11 between invariant tensors and multigraphs

induce a linear isomorphism

Natfieldsm (α1, ..., αn) ∼= ker δ : Gra(α1, ..., αn)→ Gra

1(α1, ..., αn)

Proposition 27. The morphism δ : Gra→ Gra1

satisfies the hypotheses of corollary 8.

Proof. From proposition 15, we have

Gra ∼= S(Gra

irred)

It is equally clear that the wedge product turns Gra1

into a twisted Gra-module which is freely

generated (as a twisted module over an algebra over Com) by the subcollection Gra1,irred

. In-

deed, every multigraph G ∈ Gra1decomposes as a wedge product of irreducible pieces, precisely

one of which is contained in Gra1,irred

, and this decomposition is unique up to re-ordering

95

Page 96: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

of pieces. Thus, hypotheses 1 and 2 are satisfied. For 3, consider an irreducible multigraph

G ∈ Gra1,irred. Hence, G−� is connected. Since the replacement rules defining δ are connected,

δ(G) − � = δ (G−�) is connected, and therefore, δ(G) is irreducible, which establishes 3. For

4, notice that δ2 is 0 on Gra. Hence,

δ(G1 ∧G2) = δ1(G1 ∧G2) + δ2(G1 ∧G2)

=∑

v∈G1tG2

δ1,v(G1 ∧G2)

=∑v∈G1

δ1,v(G1 ∧G2) +∑v∈G2

δ1,v(G1 ∧G2)

=∑v∈G1

δ1,v(G1) ∧G2 +∑v∈G2

G1 ∧ δ1,v(G2)

=

(∑v∈G1

δ1,v(G1)

)∧G2 +G1 ∧

(∑v∈G2

δ1,v(G2)

)= δ(G1) ∧G2 +G1 ∧ δ(G2)

which establishes 4.

It follows from corollary 8 that every element of ker δ : Gra→ Gra1

is a linear combination of

wedge products of irreducible kernel elements. Symbolically,

ker δ : Gra→ Gra1 ∼= S

(ker δ : Gra

irred → Gra1)

Furthermore, by a simple graph-theoretic argument, we see that all kernel elements are internally

acyclic:

Theorem 11.(

ker δ : Gra→ Gra1)⊂ Graacyc

We begin by establishing a bit of extra notation. Let Gracyc ⊂ Gra denote the subcollec-

tion spanned by internally cyclic multigraphs (i.e. multigraphs G with G − � cyclic) so that

Gra ∼= Graacyc ⊕ Gracyc. For each n, k ∈ N, let Gracyc[n, k] ⊂ Gracyc[n] denote the subspace

spanned by multigraphs with n ordinary vertices and k ordinary vertices contained in an internal

cycle. Similarly, we can define Gra1

cyc and Gra1

cyc[n, k], with Gra1

cyc[n, k] ⊂ Gra1

cyc[n] being the

subspace spanned by multigraphs with n ordinary vertices, 1 white vertex, and k vertices (of

either type) contained in an internal cycle.

We want to show that the restriction

96

Page 97: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

δ : Gracyc → Gra1

cyc

is injective, i.e that

δ[n] : Gracyc[n]→ Gra1

cyc[n]

is injective for every n ∈ N. Fix a particular n ∈ N. It is clear from the definition of δ that

δ[n] = δ+0[n] + δ+1[n]

where δ+0[n] is the component of δ[n] : Gracyc[n]→ Gra1

cyc[n] which fixes the number of vertices

contained in an internal cycle, and δ+1[n] is the component which increases this number by one.

In this notation, we have the following simple criterion.

Lemma 3. δ[n] : Gracyc[n] → Gra1

cyc[n] is injective if δ+0[n] : Gracyc[n, k] → Gra1

cyc[n, k] is

injective for every k ∈ N.

Proof. Suppose δ[n] : Gracyc[n]→ Gra1

cyc[n] has nontrivial kernel. Hence, there is a multigraph

x ∈ Gracyc[n] not equal to 0 such that δ[n](x) = 0. x can be decomposed as a finite sum

x = xk + xk+1 + ...+ xk+a

where xi ∈ Gracyc[n, i] for i = k, k+ 1, ..., k+a and xk 6= 0. Fix a basis of multigraphs for Gra1.

Our particular choice of basis induces a projection

π : Gra1

cyc[n]� Gra1

cyc[n, k]

Since δ[n] cannot reduce the number of cycle vertices in a multigraph, we have

0 = π ◦ δ[n](x) = δ+0[n](xk)

Yet, xk 6= 0 be assumption, so δ+0[n] : Gracyc[n, k]→ Gra1

cyc[n, k] cannot be injective.

Proof of theorem 11. Consider the subspace Gra1

cyc[n, k]′ ⊂ Gra1

cyc[n, k] spanned by multigraphs

with one white vertex contained in an internal cycle such that precisely one input to this vertex

does not lie on an internal cycle. Fixing a basis of multigraphs for Gra1, we obtain a projection

97

Page 98: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

π : Gra1

cyc[n, k]� Gra1

cyc[n, k]′

and contracting along the unique input not contained in a cycle, we obtain a linear map

r : Gra1

cyc[n, k]′ → Gracyc[n, k]

We claim r ◦ π ◦ δ+0[n] = w id where w is some nonzero integer. From this, the injectivity of

δ+0[n] follows trivially, and by lemma 3, we are done. Select a multigraph G ∈ Gracyc[n, k] and

consider a particular (internal) cycle C ⊂ G. Vertices in C are of three possible types:

1. v has two inputs contained in C

2. v has both an input and an output contained C

3. v has two outputs contained in C

Clearly, C contains at least one vertex of type 1 or 2. Denote the number of such vertices by wC

and let w be the sum∑wC over all internal cycles in G. Notice:

98

Page 99: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

δ

= ◦•

+ ...

δ

= ◦•

+ ◦

+ ◦•

− •

+ ...

δ

= ◦•

+ ◦•

+ ...

Figure 1: Replacement Rules for Type 1 Vertices

δ

= ◦• + ◦

+ ◦•

− ◦• + ...

Figure 2: Replacement Rules for Type 2 Vertices

99

Page 100: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

δ

= 0 + ...

δ

= ◦

+ ◦•

+ ...

Figure 3: Replacement Rules for Type 3 Vertices

From these illustrations it is clear that r ◦ π ◦ δ+0[n] = w id, as desired.

References

[1] R. S. Palais and C.-L. Terng, “Natural Bundles Have Finite Order,” Topology, 1976.

[2] C.-L. Terng, “Natural Vector Bundles and Natural Differential Operators,” American Jour-

nal of Mathematics, 1978.

[3] D. B. A. Epstein and W. P. Thurston, “Transformation Groups and Natural Bundles,” Proc.

London Math. Soc., vol. 3, no. 38, 1979.

[4] R. S. Palais, “Natural Operations on Differential Forms,” Transactions of the American

Mathematical Society, 1959.

[5] P. W. M. Ivan Kolar and J. Slovak, Natural Operations in Differential Geometry. Springer-

Verlag, 1993.

[6] F. Chapoton and M. Livernet, “Pre-Lie Algebras and the Rooted Trees Operad,” Interna-

tional Mathematics Research Notices, 2001.

[7] J.-L. Loday, Generalized Bialgebras and Triples of Operads. Astrisque, 2008.

[8] S. Merkulov, Frobenius Manifolds: Quantum Cohomology and Singularities, vol. E36 of

Aspects Math. Vieweg, 2004.

100

Page 101: NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas …math.mit.edu/~lmbrown/pdfs/mastersthesis.pdf · NATURAL STRUCTURES IN DIFFERENTIAL GEOMETRY Lucas Mason-Brown ... 15 A Co-algebra

[9] W. M. Mikulski, “Locally Determined Associated Spaces,” J. London Math. Soc., vol. 32,

1985.

[10] M. Markl, “Lie Elements in Pre-Lie Algebras, Trees, and Cohomology Operations,” Journal

of Lie Theory, 2006.

[11] M. Markl, “Natural Differential Operators and Graph Complexes,” Differential Geometry

and its Applications, 2009.

[12] A. Joyal and R. Street, “Braided tensor categories,” Advances in Mathematics, 1993.

[13] J.-L. Loday and B. Vallette, Algebraic Operads. Springer-Verlag, 2012.

[14] M. Gerstenhaber, “The Cohomology Structure of an Associative Ring,” Annals of Mathe-

matics, vol. 78, no. 2, 1963.

[15] T. J. L. Frederick R. Cohen and J. P. May, The Homology of Iterated Loop Spaces. Springer-

Verlag, 1976.

[16] M. Markl, “Invariant Tensors and Graphs.” Preprint arXiv:0801.0418 [math.CA].

101