Geometric Theory of Weyl Structures

2 The bundle of Weyl structures

This section works in the setting of general parabolic geometries. We assume that the reader is familiar with the the basic concepts and only briefly collect what we need about parabolic geometries and Weyl structures. Then we define the bundle of Weyl structures and identify some of the geometric structures that are naturally induced on its total space. We then prove existence of a canonical connection and explain how these structures can be used as an equivalent encoding of the theory of Weyl structures.

2.1 Parabolic geometries

The basic ingredient needed to specify a type of parabolic geometry is a semisimple Lie algebra \(\mathfrak g\) that is endowed with a so-called \(|k|\)-grading. This is a decomposition \[\mathfrak g=\mathfrak g_{-k}\oplus\dots\oplus\mathfrak g_{-1}\oplus\mathfrak g_0\oplus\mathfrak g_1\oplus\dots\oplus\mathfrak g_k\] of \(\mathfrak g\) into a direct sum of linear subspaces such that

In particular, this implies that the Lie subalgebra \(\mathfrak g_0\) naturally acts on each of the spaces \(\mathfrak g_i\) via the restriction of the adjoint action. Moreover, \(\mathfrak p:=\mathfrak g_0\oplus\mathfrak p_+\) is a Lie subalgebra of \(\mathfrak g\), which turns out to be a parabolic subalgebra in the sense of representation theory.

Such \(|k|\)-gradings can be easily described in terms of the structure theory of semisimple Lie algebras, see Section 3.2 of [16]. In particular, it turns out that any parabolic subalgebra is obtained in this way and, essentially, the classification of gradings is equivalent to the classification of parabolic subalgebras. Further, the decomposition \(\mathfrak p=\mathfrak g_0\oplus\mathfrak p_+\) is the reductive Levi decomposition, so it is a semi-direct product, \(\mathfrak p_+\) is the nilradical of \(\mathfrak p\), and the subalgebra \(\mathfrak g_0\) is reductive. Of course, also \(\mathfrak g_-:=\mathfrak g_{-k}\oplus\dots\oplus\mathfrak g_{-1}\) is a Lie subalgebra of \(\mathfrak g\), which is nilpotent by the grading property. It turns out that \(\mathfrak g_-\) and \(\mathfrak p_+\) are isomorphic.

Next, one chooses a Lie group \(G\) with Lie algebra \(\mathfrak g\). Then the normalizer of \(\mathfrak p\) in \(G\) has Lie algebra \(\mathfrak p\), and one chooses a closed subgroup \(P\subset G\) lying between this normalizer and its connected component of the identity. The subgroup \(P\) naturally acts on \(\mathfrak g\) and \(\mathfrak p\) via the adjoint action. More generally, one puts \(\mathfrak g^i:=\oplus_{j\geq i}\mathfrak g_j\) to define a filtration of \(\mathfrak g\) by linear subspaces that is invariant under the adjoint action of \(P\). This makes \(\mathfrak g\) into a filtered Lie algebra in the sense that \([\mathfrak g^i,\mathfrak g^j]\subset\mathfrak g^{i+j}\).

Having made these choices, there is the concept of a parabolic geometry of type \((G,P)\) on a manifold \(M\) of dimension \(\dim(G/P)\). This is defined as a Cartan geometry \((p:\mathcal G\to M,\omega)\) of type \((G,P)\), which means that \(p:\mathcal G\to M\) is a principal \(P\)-bundle and that \(\omega\in\Omega^1(\mathcal G,\mathfrak g)\) is a Cartan connection. This in turn means that \(\omega\) is equivariant for the principal right action, so \((r^g)^*\omega=\operatorname{Ad}(g^{-1})\circ\omega\), reproduces the generators of fundamental vector fields, and that \(\omega(u):T_u\mathcal G\to\mathfrak g\) is a linear isomorphism for each \(u\in\mathcal G\). In addition, one requires two conditions on the curvature of \(\omega\), which are called regularity and normality, which we don’t describe in detail.

While Cartan geometries provide a nice uniform description of parabolic geometries, this should be viewed as the result of a theorem rather than a definition. To proceed towards more common descriptions of the geometries, one first observes that the Lie group \(P\) can be decomposed as a semi-direct product. On the one hand, the exponential map restricts to a diffeomorphism from \(\mathfrak p_+\) onto a closed normal subgroup \(P_+\subset P\). On the other hand, one defines a closed subgroup \(G_0\subset P\) as consisting of those elements, whose adjoint action preserves the grading of \(\mathfrak g\) and observes that this has Lie algebra \(\mathfrak g_0\). Then the inclusion of \(G_0\) into \(P\) induces an isomorphism \(G_0\to P/P_+\).

Using this, one can pass from the Cartan geometry \((p:\mathcal G\to M,\omega)\) to an underlying structure by first forming the quotient \(\mathcal G_0:=\mathcal G/P_+\), which is a principal \(G_0\)-bundle. Moreover, for each \(i=-k,\dots,k\), there is a smooth subbundle \(T^i\mathcal G\subset T\mathcal G\) consisting of those tangent vectors that are mapped to \(\mathfrak g^i\subset\mathfrak g\) by \(\omega\). Since \(T^1\mathcal G\) is the vertical bundle of \(\mathcal G\to\mathcal G_0\), these subbundles descend to a filtration \(\{T^i\mathcal G_0:i=-k,\dots,0\}\) of \(T\mathcal G_0\). Moreover, for each \(i<0\), the component of \(\omega\) in \(\mathfrak g_i\) descends define a smooth section of the bundle \(L(T^i\mathcal G_0,\mathfrak g_i)\) of linear maps, so this can be viewed as a partially defined \(\mathfrak g_i\)-valued differential form.

The simplest case here is \(k=1\), for which the geometries in question are often referred to as AHS structures. In this case, one obtains a \(\mathfrak g_{-1}\)-valued one-form \(\theta\) on \(\mathcal G_0\), which is \(G_0\)-equivariant and whose kernel in each point is the vertical subbundle. This means that \((p_0:\mathcal G_0\to M,\theta)\) in this case simply is a classical first order structure corresponding to the adjoint action of \(G_0\) on \(\mathfrak g_{-1}\) (which turns out to be infinitesimally effective). According to a result of Kobayashi and Nagano (see [24]), the resulting class of structures for simple \(\mathfrak g\) is very peculiar, since these are the only irreducible first order structures of finite type, for which the first prolongation is non-trivial. This class contains important examples, like conformal structures, almost quaternionic structures, and almost Grassmannian structures.

For general \(k\), there is an interpretation of \(\mathcal G_0\) and the partially defined forms as a filtered analogue of a first order structure. This involves a filtration of the tangent bundle \(TM\) by smooth subbundles \(T^iM\) for \(i=-k,\dots,-1\) with prescribed (non–)integrability properties together with a reduction of structure group of the associated graded vector bundle to the tangent bundle. This leads to examples like hypersurface-type CR structures, in which the filtration is equivalent to a contact structure, while the reduction of structure group is defined by an almost complex structure on the contact subbundle. Further important example of such structures are path geometries, quaternionic contact structures and various types of generic distributions.

Except for two cases, the Cartan geometry can be uniquely (up to isomorphism) recovered from the underlying structure (see Section 3.1 of [16]), and indeed this defines an equivalence of categories. So in this case, one has two equivalent descriptions of the structure. The two exceptional cases are projective structures and a contact analogue of those. In these cases, the underlying structure contains no information respectively describes only the contact structure, and one in addition has to choose an equivalence class of connections in order to describe the structure. Still, these fit into the general picture with respect to Weyl structures, which we discuss next.

2.2 Weyl structures

These provide the basic tool to explicitly translate between the description of a parabolic geometry as a Cartan geometry and the picture of the underlying structure. So let us suppose that \((p:\mathcal G\to M,\omega)\) is a Cartan geometry of type \((G,P)\) and that \(p_0:\mathcal G_0\to M\) is the underlying structure described in 2.1. The original definition of a Weyl structure used in [15] is as a \(G_0\)-equivariant section \(\sigma\) of the natural projection \(q:\mathcal G\to \mathcal G_0=\mathcal G/P_+\). One shows that such sections always exist globally and by definition, they provide reductions of the principal \(P\)-bundle \(p:\mathcal G\to M\) to the structure group \(G_0\subset P\). As a representation of \(G_0\), the Lie algebra \(\mathfrak g\) splits as \(\mathfrak g_-\oplus\mathfrak g_0\oplus\mathfrak p_+\) (and indeed further according to the \(|k|\)-grading). Thus, the pullback \(\sigma^*\omega\) splits accordingly into a sum of three \(G_0\)-equivariant one-forms with values in \(\mathfrak g_-\), \(\mathfrak g_0\) and \(\mathfrak p_+\), respectively, which then admit nice interpretations in terms of the underlying structure. The \(\mathfrak g_0\)-component defines a principal connection on \(\mathcal G_0\), which induces the Weyl connections on associated bundles. The component in \(\mathfrak p_+\) descends to a one-form on \(M\) with values in the associated graded to the cotangent bundle \(T^*M\), which is the Rho-tensor associated to the Weyl structure. The \(\mathfrak g_-\)-component also descends to \(M\) and provides an isomorphism between the tangent bundle \(TM\) and its associated graded bundle. For the structures we consider in this article, this component coincides with the soldering form that identifies \(\mathcal G_0\) as a reduction of structure group of \(TM\).

As observed in [21], any reduction of \(p:\mathcal G\to M\) to the structure group \(G_0\subset P\) comes from a Weyl structure. This is because the composition of \(q\) with the principal bundle morphism defining such a reduction clearly is an isomorphism of \(G_0\)-principal bundles. Thus one could equivalently define a Weyl structure as such a reduction of structure group and then observe that this defines a \(G_0\)-equivariant section of \(q:\mathcal G\to\mathcal G_0\). It is a classical result that reductions of \(\mathcal G\) to the structure group \(G_0\) can be equivalently described as smooth sections of the associated bundle with fiber \(P/G_0\). This motivates the following definition from [21].

Definition 2.1

The bundle of Weyl structures associated to the parabolic geometry \((p:\mathcal G\to M,\omega)\) is \(\pi:A:=\mathcal G\times_P(P/G_0)\to M\).

The correspondence between Weyl structures and smooth sections of \(\pi:A\to M\) can be easily made explicit. Given a \(G_0\)-equivariant section \(\sigma:\mathcal G_0\to\mathcal G\) one considers the map sending \(u_0\in \mathcal G_0\) to the class of \((\sigma(u_0),eG_0)\) in \(\mathcal G\times_P(P/G_0)\), where \(e\in P\) is the neutral element. By construction, the resulting smooth map \(\mathcal G_0\to A\) is constant on the fibers of \(p_0:\mathcal G_0\to M\) and thus descends to a smooth map \(s:M\to A\), which is a section of \(\pi\) by construction. Conversely, a section \(s\) of \(\pi\) corresponds to a smooth, \(G_0\)-equivariant map \(f:\mathcal G\to P/G_0\) characterized by the fact that \(s(x)\) is the class of \((u,f(u))\) for each \(u\) in the fiber of \(\mathcal G\) over \(x\). But then \(f^{-1}(eG_0)\) is a smooth submanifold of \(\mathcal G\) on which the projection \(q:\mathcal G\to\mathcal G_0\) restricts to a \(G_0\)-equivariant diffeomorphism. The inverse of this diffeomorphism gives the Weyl structure determined by \(s\).

From the definition, we can verify that the bundle of Weyl structures is similar to an affine bundle. This will also provide the well known affine structure on Weyl structures in our picture. To formulate this, recall first that the parabolic subgroup \(P\subset G\) is a semi-direct product of the subgroup \(G_0\subset P\) and the normal subgroup \(P_+\subset P\). In particular, any element \(g\in P\) can be uniquely written as \(g_0g_1\) with \(g_0\in G_0\) and \(g_1\in P_+\), compare with Theorem 3.1.3 of [16], and of course \(g_0g_1=(g_0g_1g_0^{-1})g_0\) provides the corresponding decomposition in the opposite order.

Proposition 2.2

Let \(\pi:A\to M\) be the bundle of Weyl structures associated to a parabolic geometry \((p:\mathcal G\to M,\omega)\). Then sections of \(\pi:A\to M\) can be naturally identified with smooth functions \(f:\mathcal G\to P_+\) such that \(f(u\cdot (g_0g_1))=g_1^{-1}g_0^{-1}f(u)g_0\) for each \(u\in\mathcal G\), \(g_0\in G_0\) and \(g_1\in P_+\).

Fixing one function \(f\) that satisfies this equivariancy condition, any other function which is equivariant in the same way can be written as \(\hat f(u)=f(u)h(u)\), where \(h:\mathcal G\to P_+\) is a smooth function such that \(h(u\cdot(g_0g_1))=g_0^{-1}h(u)g_0\).

Proof. The inclusion \(P_+\hookrightarrow P\) induces a smooth map \(P_+\to P/G_0\), and from the decomposition of elements of \(P\) described above, we readily see that this is surjective. On the other hand, writing the quotient projection \(P\to G_0\) as \(\alpha\), the map \(g\mapsto g\alpha(g)^{-1}\) induces a smooth inverse, so \(P/G_0\) is diffeomorphic to \(P_+\). Since \(A=\mathcal G\times_P(P/G_0)\), smooth sections of \(\pi:A\to M\) are in bijective correspondence with \(P\)-equivariant smooth functions \(\mathcal G\to P/G_0\), so these can be viewed as functions with values in \(P_+\). The equivariancy condition reads as \(f(u\cdot (g_0g_1))=g_1^{-1}g_0^{-1}\cdot f(u)\). But starting from \(\tilde g_1G_0\), we get \(g_1^{-1}g_0^{-1}\tilde g_1G_0=g_1^{-1}g_0^{-1}\tilde g_1g_0G_0\), and \(g_1^{-1}g_0^{-1}\tilde g_1g_0\in P_+\). This completes the proof of the first claim.

Given one function \(f:\mathcal G\to P_+\), of course any other such function can be uniquely written as \(\hat f=fh\) for a smooth function \(h:\mathcal G\to P_+\), so it remains to understand \(P\)-equivariance. What we assume is that \(f(u\cdot(g_0g_1))= g_1^{-1}g_0^{-1}f(u)g_0\) and we want \(\hat f\) to satisfy the analogous equivariancy condition. But this exactly requires that \(g_1^{-1}g_0^{-1}f(u)g_0h(u\cdot(g_0g_1))= g_1^{-1}g_0^{-1}f(u)h(u)g_0\), which is equivalent to the claimed equivariancy of \(h\).

To connect to the well-known affine structure on the set of Weyl structures, we observe two alternative ways to express things using the exponential map. On the one hand, we have observed above that \(\exp:\mathfrak p_+\to P_+\) is a diffeomorphism. Thus we can write \(h(u)=\exp(\Upsilon(u))\) and equivariancy of \(h\) is equivalent to \(\Upsilon(u\cdot(g_0g_1))=\operatorname{Ad}(g_0)^{-1}(\Upsilon(u))\). On the other hand, in the proof of Theorem 3.1.3 of [16] it shown that also \((Z_1,\dots,Z_k)\mapsto\exp(Z_1)\cdots\exp(Z_k)\) defines a diffeomorphism \(\mathfrak g_1\oplus\dots\oplus \mathfrak g_k\to P_+\). Correspondingly, we can write \(h(u)=\exp(\Upsilon_1(u))\cdots\exp(\Upsilon_k(u))\) where \(\Upsilon_i:\mathcal G\to\mathfrak g_i\) is a smooth map for each \(i=1,\dots,k\). Again, equivariancy of \(h\) translates to \(\Upsilon_i(u\cdot(g_0g_1))=\operatorname{Ad}(g_0)^{-1}(\Upsilon_i(u))\) for each \(i\).

There is also a nice global way to express the affine structure. The filtration of \(TM\) induced by a parabolic geometry dualizes to a filtration of the cotangent bundle \(T^*M\) and we can form the associated graded bundle \(\operatorname{gr}(T^*M)\). The general theory implies that this can be realized as \(\mathcal G\times_{P}\operatorname{gr}(\mathfrak p_+)\cong\mathcal G_0\times_{G_0}\mathfrak p_+\).

Proposition 2.3

Let \(\pi:A\to M\) be the bundle of Weyl structures associated to a parabolic geometry \((p:\mathcal G\to M,\omega)\). Then for any smooth section \(s\) of \(\pi\), there is an induced diffeomorphism \(\varphi_s:T^*M\to A\).

Proof. Let \(\sigma:\mathcal G_0\to\mathcal G\) be the \(G_0\)-equivariant section determined by \(s\). Since \(\exp:\mathfrak p_+\to P_+\) is a diffeomorphism, we conclude that \(\Phi_s(u_0,Z):=\sigma(u_0)\exp(Z)\) defines a diffeomorphism \(\Phi_s:\mathcal G_0\times\mathfrak p_+\to \mathcal G\). Given \(g_0\in G_0\), the definition readily implies that \(\Phi_s(u\cdot g_0,\operatorname{Ad}(g_0^{-1})(Z))=\Phi_s(u_0,Z)\cdot g_0\). Hence there is an induced diffeomorphisms between the orbit spaces \(\operatorname{gr}(T^*M)=\mathcal G_0\times_{G_0}\mathfrak p_+\) and \(A=\mathcal G_0/G_0\).

2.3 The basic geometric structures on \(A\)

It was shown in [21] that the parabolic geometry \((p:\mathcal G\to M,\omega)\) gives rise to a connection on the tangent bundle \(TA\) of \(A\). The argument used to obtain this connection is rather intricate: There is the opposite parabolic subgroup \(P^{op}\) to \(P\) which corresponds to the Lie subalgebra \(\mathfrak g_-\oplus\mathfrak g_0\subset\mathfrak g\) and one considers the homogeneous space \(G/P^{op}\). Restricting the \(G\)-action to \(P\) and forming the associated bundle \(\mathcal G\times_P(G/P^{op})\) the Cartan connection \(\omega\) induces a natural affine connection on the total space of this bundle. It is then easy to see that \(P\cap P^{op}=G_0\), so acting with \(P\) on \(eP^{op}\) defines a \(P\)-equivariant open embedding \(A\to\mathcal G\times_P(G/P^{op})\), thus providing a connection on \(TA\) as claimed. Our first main result provides a more conceptual description of this connection, which directly implies compatibility with several additional geometric structures on \(A\).

Proposition 2.4

The canonical projection \(\mathcal G\to A\) is a \(G_0\)-principal bundle and \(\omega\) defines a Cartan connection on that bundle, so \((\mathcal G\to A,\omega)\) is a Cartan geometry of type \((G,G_0)\). In particular, the \(\mathfrak g_0\)-component of \(\omega\) defines a canonical principal connection on \(\mathcal G\to A\) and \(TA\cong\mathcal G\times_{G_0}(\mathfrak g/\mathfrak g_0)\), so this inherits a canonical linear connection. Finally, there is a natural splitting \(TA=L^-\oplus L^+\) into a direct sum of two subbundles of rank \(\dim(M)\), which is parallel for the connection and such that \(L^+\) is the vertical bundle of \(\pi\).

Proof. Mapping \(u\in\mathcal G\) to the class of \((u,eG_0)\) in \(\mathcal G\times_P(P/G_0)\) is immediately seen to be surjective and its fibers coincide with the orbits of \(G_0\) on \(\mathcal G\). Hence one obtains an identification of \(\mathcal G/G_0\) with \(A\), and it is well known that this makes the projection \(\mathcal G\to A\) into a \(G_0\)-principal bundle. The defining properties of \(\omega\) for the group \(P\) and the Lie algebra \(\mathfrak p\) then imply the corresponding properties for the group \(G_0\) and the Lie algebra \(\mathfrak g_0\), so \(\omega\) defines a Cartan connection on \(\mathcal G\to A\).

As a representation of \(G_0\), we get \(\mathfrak g=\mathfrak g_0\oplus(\mathfrak g_-\oplus\mathfrak p_+)\). This means that we have given a \(G_0\)-invariant complement to \(\mathfrak g_0\) in \(\mathfrak g\). Decomposing \(\omega\) accordingly, the component \(\omega_0\) in \(\mathfrak g_0\) is \(G_0\)-equivariant, thus defining a principal connection on \(\mathcal G\to A\), which induces linear connections on all associated vector bundles. Moreover, since \(\omega\) is a Cartan connection on \(\mathcal G\to A\), we can identify \(TA\) with the associated vector bundle \[\mathcal G\times_{G_0}(\mathfrak g/\mathfrak g_0)\cong\mathcal G\times_{G_0}(\mathfrak g_-\oplus\mathfrak p_+).\] This readily implies both the existence of a natural connection and of a compatible decomposition of \(TA\) with \(L^-=\mathcal G\times_{G_0}\mathfrak g_-\) and \(L^+=\mathcal G\times_{G_0}\mathfrak p_+\). The tangent map \(T\pi:TA\to TM\) is induced by the projection \(\mathfrak g/\mathfrak g_0\to\mathfrak g/\mathfrak p\). Identifying \(\mathfrak g/\mathfrak g_0\) with \(\mathfrak g_-\oplus\mathfrak p_+\) the kernel of this projection is \(\mathfrak p_+\), which shows that \(L^+\subset TA\) coincides with \(\ker(T\pi)\).

This result also gives us a basic supply of natural vector bundles on \(A\), namely the vector bundles associated to the principal bundle \(\mathcal G\to A\) via representations of \(G_0\). Moreover, the principal connection on that bundle coming from \(\omega_0\) gives rise to an induced linear connection on each of these associated bundles. We will denote all these induced connections by \(D\). Given an associated bundle \(E\to A\), we can view \(D\) as an operator \(D:\Gamma(E)\to\Gamma(T^*A\otimes E)\). Of course the splitting \(TA=L^-\oplus L^+\) from Proposition 2.4 induces an analogous splitting of \(T^*A\), which allows us to split \(D\) into two partial connections \(D=D^-\oplus D^+\). Here \(D^\pm:\Gamma(E)\to\Gamma((L^\pm)^*\otimes E)\). Viewing \(D\) as a covariant derivative, \(D^\pm\) is defined by differentiating only in directions of the corresponding subbundle of \(TA\).

2.4 Relations between natural vector bundles

Recall that the natural vector bundles for the parabolic geometry \((p:\mathcal G\to M,\omega)\) are the associated vector bundles of the form \(\mathcal VM=\mathcal G\times_P\mathbb V\) for representations \(\mathbb V\) of \(P\). Throughout this article, we will only consider the case that the center \(Z(G_0)\) of the subgroup \(G_0\subset P\) acts diagonalizably on \(\mathbb V\). Together with the fact that \(G_0\) is reductive, this implies that \(\mathbb V\) is completely reducible as a representation of \(G_0\). One important subclass of natural bundles is formed by completely reducible bundles that correspond to representations of \(P\) on which the subgroup \(P_+\subset P\) acts trivially, which is equivalent to complete reducibility as a representation of \(P\). On the other hand, there are tractor bundles which correspond to restrictions to \(P\) of representations of \(G\).

Any representation \(\mathbb V\) of \(P\) can naturally be endowed with a \(P\)-invariant filtration of the form \(\mathbb V=\mathbb V^0\supset\mathbb V^1\supset\dots\supset\mathbb V^N\) as follows (see Section 3.2.12 of [16]). The smallest component \(\mathbb V^N\) consists of those elements, on which \(\mathfrak p_+\) acts trivially under the infinitesimal action. The larger components are characterized iteratively by the fact that \(v\in\mathbb V^j\) if and only if it is sent to \(\mathbb V^{j+1}\) by the action of any element of \(\mathfrak p_+\). Then one defines the associated graded representation \(\operatorname{gr}(\mathbb V):=\oplus_{i=0}^N\operatorname{gr}_i(\mathbb V)\) with \(\operatorname{gr}_i(\mathbb V):=(\mathbb V^i/\mathbb V^{i+1})\) and \(\mathbb V^{N+1}=\{0\}\). By construction, this is a completely reducible representation of \(P\).

As an important special case, consider the restriction of the adjoint representation of \(G\) to \(P\). Then it turns out that, up to a shift in degree, the canonical \(P\)-invariant filtration is exactly the filtration \(\{\mathfrak g^i\}\) derived from the \(|k|\)-grading of \(\mathfrak g\) as in 2.1. In particular, this implies that \(\mathfrak g^2=[\mathfrak p_+,\mathfrak p_+]\) and similarly, the higher filtrations components form the lower central series of \(\mathfrak p_+\). Using this it is easy to see that the natural filtration on any representation \(\mathbb V\) of \(P\) has the property that \(\mathfrak g^i\cdot\mathbb V^j\subset\mathbb V^{i+j}\) for all \(i,j\geq 0\) under the infinitesimal representation of \(\mathfrak p=\mathfrak g^0\). This readily implies that there is a natural action of the associated graded \(\operatorname{gr}(\mathfrak p)\) on \(\operatorname{gr}(\mathbb{V} )\), which is compatible with the grading. Since the filtration of \(\mathfrak p\) is induced by the (non-negative part of the) grading on \(\mathfrak g\), we can identify \(\operatorname{gr}(\mathfrak p)\) with \(\mathfrak p\) via the inclusion of \(\mathfrak g_i\) into \(\mathfrak g^i\). Altogether, we get, for each \(i,j\geq 0\), bilinear maps \(\mathfrak g_i\times\operatorname{gr}_j(\mathbb V)\to\operatorname{gr}_{i+j}(\mathbb V)\), which are \(P\)-equivariant (with trivial action of \(P_+\)) by construction.

As a representation of the subgroup \(G_0\subset P\), the associated graded \(\operatorname{gr}(\mathbb V)\) is isomorphic to \(\mathbb V\). Indeed, we have observed above that \(\mathbb V\) is completely reducible as a representation of \(G_0\), so the same holds for each of the subrepresentations \(\mathbb V^j\subset\mathbb V\). In particular, there always is a \(G_0\)-invariant complement \(\mathbb V_j\) to the invariant subspace \(\mathbb V^{j+1}\subset\mathbb V^j\) and we put \(\mathbb V_N=\mathbb V^N\). By construction, we on the one hand get \(\mathbb V\cong\oplus\mathbb V_j\) and on the other hand \(\mathbb V_j\cong\mathbb V^j/\mathbb V^{j+1}\) which implies that claimed statement. Otherwise put, one can interpret the passage from \(\mathbb V\) to \(\operatorname{gr}(\mathbb V)\) as keeping the restriction to \(G_0\) of the \(P\)-action on \(\mathbb V\) and extending this by the trivial action of \(P_+\) to a new action of \(P\).

The construction of the associated graded has a direct counterpart on the level of associated bundles. Putting \(\mathcal VM:=\mathcal G\times_P\mathbb V\to M\), any of the filtration components \(\mathbb V^i\) defines a smooth subbundle \(\mathcal V^iM:=\mathcal G\times_P\mathbb V^i\to M\). Thus \(\mathcal VM\) is filtered by the smooth subbundles \(\mathcal V^iM\) and we can form the associated graded vector bundle \(\operatorname{gr}(\mathcal VM)=\oplus (\mathcal V^iM/\mathcal V^{i+1}M)\). It is easy to see that this can be identified with the associated bundle \(\mathcal G\times_P\operatorname{gr}(\mathbb V)\). However, the fact that \(\mathbb V\) and \(\operatorname{gr}(\mathbb V)\) are isomorphic as representations of \(G_0\) does not have a geometric interpretation without making additional choices. Hence on the level of associated bundles, it is very important to carefully distinguish between a filtered vector bundle and its associated graded.

Any representation \(\mathbb V\) of \(P\) defines a representation of \(G_0\) by restriction. Hence denoting by \(\pi:A\to M\) the bundle of Weyl structures, \(\mathbb V\) also gives rise to a natural vector bundle over \(A\) that we denote by \(\mathcal VA:=\mathcal G\times_{G_0}\mathbb V\to A\). Some information is lost in that process, however, for example \(\mathcal G\times_{G_0}\mathbb V\cong\mathcal G\times_{G_0}\operatorname{gr}(\mathbb V)\) for any representation \(\mathbb V\) of \(P\). Next, sections of \(\mathcal VA\to A\) can be naturally identified with smooth functions \(\mathcal G\to\mathbb V\) that are \(G_0\)-equivariant. Similarly, sections of \(\mathcal VM\to M\) are in bijective correspondence with smooth functions \(\mathcal G\to\mathbb V\), which are \(P\)-equivariant. Thus we see that there is a natural inclusion of \(\Gamma(\mathcal VM\to M)\) as a linear subspace of \(\Gamma(\mathcal VA\to A)\). We will denote this by putting a tilde over the name of a section of \(\mathcal VM\to M\) in order to indicate the corresponding section of \(\mathcal VA\to A\). So both the sections of \(\mathcal VM\to M\) and of its associated graded vector bundle can be interpreted as (different) subspaces of the space of sections of \(\mathcal VA\to A\).

Now we can describe the relations of bundles and sections explicitly.

Theorem 2.5

Let \((p:\mathcal G\to M,\omega)\) be a parabolic geometry of type \((G,P)\) and let \(\pi:A\to M\) be the corresponding bundle of Weyl structures. Fix a representation \(\mathbb V\) of \(P\) and consider the corresponding natural bundles \(\mathcal VM=\mathcal G\times_P\mathbb V\to M\) and \(\mathcal VA=\mathcal G\times_{G_0}\mathbb V\to A\). Then we have:

(1) \(\mathcal VA\) can be naturally identified with the pullback bundle \(\pi^*\mathcal VM\). In particular, \(L^-\cong\pi^*TM\) and \(L^+\cong\pi^*T^*M\).

(2) The operation \(\sigma\mapsto\tilde\sigma\) identifies \(\Gamma(\mathcal VM\to M)\) with the subspace of \(\Gamma(\mathcal VA\to A)\) consisting of those sections \(\tau\) for which \(D^+_{\varphi}\tau=-\varphi\bullet\tau\) for all \(\varphi\in\Gamma(L^+)\). Here \(\bullet:L^+\times\mathcal VA\to\mathcal VA\) is induced by the infinitesimal representation \(\mathfrak p_+\times\mathbb V\to\mathbb V\).

In particular, for a completely reducible bundle \(\mathcal VM\), \(\tilde\sigma=\pi^*\sigma\) and the image consists of all sections that are parallel for \(D^+\).

(3) Any section \(s:M\to A\) of \(\pi\) determines a natural pullback operator \(s^*:\Gamma(\mathcal VA\to A)\to\Gamma(\operatorname{gr}(\mathcal VM)\to M)\). In particular, choosing \(s\), \(\sigma\mapsto s^*\tilde\sigma\) defines a map \(\Gamma(\mathcal VM)\to\Gamma(\operatorname{gr}(\mathcal VM))\). This map is induced by a vector bundle isomorphism \(\mathcal VM\to\operatorname{gr}(\mathcal VM)\) that coincides with the isomorphism determined by the Weyl structure corresponding to \(s\) as in Section 5.1.3 of [16].

Proof. (1) follows directly from the construction: Mapping a \(G_0\)-orbit in \(\mathcal G\times\mathbb V\) to the \(P\)-orbit it generates, defines a bundle map \(\mathcal VA\to\mathcal VM\) with base map \(\pi:A\to M\). This evidently restricts to a linear isomorphism in each fiber and hence defines an isomorphism \(\mathcal VA\to\pi^*\mathcal VM\). The second statement follows from the well known facts that \(TM\cong\mathcal G\times_P(\mathfrak g/\mathfrak p)\) and \(T^*M\cong\mathcal G\times_P\mathfrak p_+\) and the fact that \(\mathfrak g/\mathfrak p\cong\mathfrak g_-\) as a representation of \(G_0\).

(2) Since \(P\) is a semi-direct product, \(P\)-equivariancy of a function is equivalent to equivariancy under \(G_0\) and \(P_+\) and equivariancy under \(P_+\) is equivalent to equivariancy for the infinitesimal action of \(\mathfrak p_+\). Hence for a \(G_0\)-equivariant function \(f:\mathcal G\to\mathbb V\), \(P\)-equivariancy is equivalent to the fact that for each \(u\in\mathcal G\) and \(Z\in\mathfrak p_+\) with fundamental vector field \(\zeta_Z\), we get \(\zeta_Z(u)\cdot f=Z\cdot f(u)\). Here in the left hand side the vector field differentiates the function, while in the right hand side we use the infinitesimal representation of \(\mathfrak p_+\) on \(\mathbb V\). Suppose that \(u\) projects to \(y\in A\). Then by definition, \(\zeta_Z(u)\) is the horizontal lift with respect to \(D\) of a tangent vector \(\varphi\in L^+_y\subset T_yA\). Hence \(\zeta_Z(u)\cdot f\) represents \(D_\varphi\tau(y)=D^+_\varphi\tau(y)\) in the, while \(Z\cdot f(u)\) of course represents \(\varphi\bullet\tau(y)\).

In the case of a completely reducible bundle, \(\bullet\) is the zero map, so we see that our subspace coincides with the \(D^+\)-parallel sections. On the other hand, for any section \(\sigma\in\Gamma(\mathcal VM)\), the pullback \(\pi^*\sigma\) is constant along the fibers of \(\pi\). Since we know from Proposition 2.4 that \(L^+\) is the vertical subbundle of \(\pi\), this implies that \(\pi^*\sigma=\tilde\sigma\).

(3) As we have noted already, for any representation \(\mathbb V\) of \(P\), the associated graded \(\operatorname{gr}(\mathbb V)\) is isomorphic to \(\mathbb V\) as a representation of \(G_0\). Thus we conclude from (1) that we can not only identify \(\mathcal VA\) with the pullback of \(\mathcal VM\) but also with the pullback of the associated graded vector bundle \(\operatorname{gr}(\mathcal VM)\). Hence for a smooth section \(s:M\to A\) and a point \(x\in M\), we can naturally identify the fiber \(\mathcal V_{s(x)}A\) with the fiber over \(x\) of \(\operatorname{gr}(\mathcal VM)\). This provides a pullback operator \(s^*:\Gamma(\mathcal VA)\to\Gamma(\operatorname{gr}(\mathcal VM))\), so \(\sigma\mapsto s^*\tilde\sigma\) defines an operator \(\Gamma(\mathcal VM)\to\Gamma(\operatorname{gr}(\mathcal VM))\). This operator is evidently linear over \(C^\infty(M,\mathbb R)\) and thus induced by a vector bundle homomorphism \(\mathcal VM\to\operatorname{gr}(\mathcal VM)\) with base map \(\operatorname{id}_M\). Suppose that for \(\sigma\in\Gamma(\mathcal VM)\) and \(x\in M\) we have \((s^*\tilde\sigma)(x)=0\). Then the function \(f:\mathcal G\to\mathbb V\) which corresponds to both \(\sigma\) and \(\tilde\sigma\) has to vanish along the fiber of \(\mathcal G\to A\) over \(s(x)\). But \(P\)-equivariancy then implies that \(f\) vanishes along the fiber of \(\mathcal G\to M\) over \(x\), so \(\sigma(x)=0\). This implies that our bundle map is injective in each fiber and since both bundles have the same rank it is an isomorphism of vector bundles.

The standard description of the isomorphism \(\mathcal VM\to\operatorname{gr}(\mathcal VM)\) induced by a Weyl structure is actually also phrased in the language of sections; Given the \(P\)-equivariant function \(f:\mathcal G\to\mathbb V\) corresponding to \(\sigma\) and the \(G_0\)-equivariant section \(\overline{s}:\mathcal G_0\to\mathcal G\) determined by \(s\), one considers the \(G_0\)-equivariant function \(f\circ\overline{s}\). This describes a section of \(\mathcal G_0\times_{G_0}\mathbb V\cong\mathcal G_0\times_{G_0}\operatorname{gr}(\mathbb V)\). Going through the identifications, it is clear that this coincides with the isomorphism described above.

Remark 2.6

(1) In principle, the pullback operation defined in part (3) of Theorem 2.5 could also be interpreted as having values in \(\Gamma(\mathcal VM\to M)\). Since \(\mathcal VA\) does not contain any information about the \(\mathfrak p_+\)-action on \(\mathbb V\), the interpretation with values in \(\Gamma(\operatorname{gr}(\mathcal VM)\to M)\) seems much more natural to us.

(2) The comparison to the standard description of Weyl structures in part (3) of the theorem also implies how the isomorphisms \(\mathcal VM\to\operatorname{gr}(\mathcal VM)\) induced by sections \(s\) of \(A\to M\) are compatible with the affine structure on the space of these sections from Proposition 2.2, compare with Proposition 5.1.5 of [16]. It is also easy to give a direct proof of this result in our picture. One just has to interpret the affine structure in terms of sections of \(L^+\to A\) and then use the obvious solution of the differential equation \(D^+_\varphi\tau=\varphi\bullet\tau\) for appropriate sections \(\varphi\).

2.5 The Weyl connections

We next describe the interpretation of Weyl connections in our picture. At the same time, we obtain a nice description of the Rho-corrected derivative associated to a Weyl structure, that was first introduced in [10], see Section 5.1.9 of [16] for a discussion. The Rho-corrected derivative comes from a principal connection on \(\mathcal G\) determined by an equivariant section \(\sigma:\mathcal G_0\to\mathcal G\). One takes the component of \(\omega\) in \(\mathfrak p\) along the image of \(\sigma\) and extends it equivariantly to a principal connection. The name "Rho-corrected derivative" comes from the explicit formula of this derivative in terms of Weyl connection an the Rho-tensor. To obtain our description, we first observe that the pullback operation from part (3) of Theorem 2.5 clearly extends to differential forms with values in a natural vector bundle. Let \(\mathbb V\) be a representation of \(P\) with corresponding natural bundles \(\mathcal VM\to M\) and \(\mathcal VA\to A\). Then one can pull back a \(\mathcal VA\)-valued \(k\) form \(\varphi\) on \(A\) along a section \(s:M\to A\) to a \(\operatorname{gr}(\mathcal VM)\)-valued \(k\)-form \(s^*\varphi\) on \(M\) in an obvious way.

Theorem 2.7

Let \(\mathbb V\) be a representation of \(P\) and let \(\mathcal VM\to M\) and \(\mathcal VA\to A\) be the corresponding natural bundles. For \(\sigma\in\Gamma(\mathcal VM)\) consider the natural lift \(\tilde\sigma\in\Gamma(\mathcal VA)\). For a smooth section \(s:M\to A\) let \(\nabla^s\) be the Weyl connection of the Weyl structure determined by \(s\). Let \(\xi\in\mathfrak X(M)\) be a vector field with natural lift \(\tilde\xi\in\Gamma(L^-)\).

(1) The pullback \(s^*D\tilde\sigma\in\Omega^1(M,\operatorname{gr}(\mathcal VM))\) of \(D\tilde\sigma\in\Omega^1(A,\mathcal VA)\) coincides with the image of \(\nabla^s\sigma\in\Omega^1(M,\mathcal VM)\) under the isomorphism \(\mathcal VM\to\operatorname{gr}(\mathcal VM)\) induced by \(s\) as in Theorem 2.5.

(2) The pullback \(s^*(D^-_{\tilde\xi}\tilde\sigma)\in\Gamma(\operatorname{gr}(\mathcal VM))\) coincides with the image of the Rho-corrected derivative \(\nabla^{\mbox{\textsf{P}}}_\xi\sigma\in\Gamma(\mathcal VM)\) under the isomorphism induced by \(s\) as in Theorem 2.5.

Proof. Let \(\bar s:\mathcal G_0\to\mathcal G\) be the \(G_0\)-equivariant section corresponding to \(s\). For a point \(x\in M\), an element \(u\in\mathcal G\) with \(p(u)=x\) lies in the image of \(\bar s\) if and only if \(u\) projects to \(s(x)\in A\). Assuming this, put \(u_0=q(u)\) where \(q:\mathcal G\to\mathcal G_0\) is the projection, so \(u=\bar s(u_0)\). To compute \(\nabla^s\), we need the horizontal lift \(\hat\xi\in\mathfrak X(\mathcal G_0)\) of \(\xi\) for the principal connection \(\bar s^*\omega_0\). This is characterized by the fact that \(\hat\xi(u_0)\) projects onto \(\xi(x)\) and that \(\omega(u)(T_{u_0}\bar s\cdot\hat\xi(u_0))\) has vanishing \(\mathfrak g_0\)-component. But by construction \(T_{u_0}\bar s\cdot\hat\xi(u_0)\) projects onto \(T_xs\cdot\xi(x)\) and so vanishing of the \(\mathfrak g_0\)-component implies that this is the horizontal lift of \(T_xs\cdot\xi(x)\) in \(u\) corresponding to the principal connection \(\omega_0\) that induces \(D\). From this, (1) follows immediately.

The argument for (2) is closely similar. By definition, \(\tilde\xi(s(x))\) is the unique tangent vector that lies in \(L^-\) and projects onto \(\xi(x)\). The \(D\)-horizontal lift of this tangent vector in \(u\), by construction, is mapped to \(\mathfrak g_-\) by \(\omega\) and projects onto \(\xi(x)\in T_xM\). But this is exactly the characterizing property of the horizontal lift with respect to the principal connection \(\gamma^{\bar s}\) used in Section 5.1.9 of [16] to define the Rho-corrected derivative. Thus the restriction of the \(G_0\)-equivariant function \(\mathcal G\to\mathbb V\) representing \(D^-_{\tilde\xi}\tilde\sigma\) to \(\bar s(\mathcal G_0)\) coincides with the restriction of the \(P\)-equivariant function representing \(\nabla^{\mbox{\textsf{P}}}_\xi\sigma\) and the claim follows from Theorem 2.5.

2.6 The universal Rho-tensor

Using the pullback of bundle valued forms, we can also describe the Rho tensor in our picture. Recall that we use the convention of [15] and [16] for Rho tensors in the setting of general parabolic geometries, which differ by sign from the standard conventions for projective and conformal structures.

Proposition 2.8

Let us view the projection \(TA\to L^+\) as \(\mbox{\textsf{P}}\in\Omega^1(A,L^+)\). Then for each smooth section \(s:M\to A\), the pullback \(s^*\mbox{\textsf{P}}\in\Omega^1(M,\operatorname{gr}(T^*M))\) coincides with the Rho-tensor of the Weyl structure determined by \(s\) as defined in Section 5.1.2 of [16].

Proof. Take a point \(x\in M\), a tangent vector \(\xi\in T_xM\) and consider \(T_xs\cdot\xi\in T_{s(x)}A\). Choose a point \(u\in\mathcal G\) over \(s(x)\) and consider its image \(u_0=q(u)\in\mathcal G_0\). Since \(u\) projects to \(s(x)\), it lies in the image of the \(G_0\)-equivariant section \(\bar s:\mathcal G_0\to\mathcal G\) determined by \(s\), so \(u=\bar s(u_0)\). Taking a tangent vector \(\hat\xi\in T_{u_0}\mathcal G_0\), the tangent vector \(T_{u_0}\bar s\cdot\hat\xi\in T_u\mathcal G\), by construction, projects onto \(T_xs\cdot\xi\in T_{s(x)}A\). But then, by definition, the \(L^+\) component of \(T_xs\cdot\xi\) is obtained by projecting \(\omega(u)^{-1}(\omega_+(T_{u_0}\bar s\cdot\hat\xi))\) to \(T_{s(x)}A\), where \(\omega_+\) denotes the \(\mathfrak p_+\)-component of the Cartan connection \(\omega\). But the Rho-tensor of \(\bar s\) is defined as the \(\operatorname{gr}(T^*M)\)-valued form induced by the \(G_0\)-equivariant form \(\bar s^*\omega_+\), which completes the argument.

Definition 2.9

The form \(\mbox{\textsf{P}}\in\Omega^1(A,L^+)\) defined by the projection \(TA\to L^+\) is called the universal Rho-tensor of the parabolic geometry \((p:\mathcal G\to M,\omega)\).

2.7 Curvature and torsion quantities

The curvature \(K\in\Omega^2(\mathcal G,\mathfrak g)\) of the Cartan connection \(\omega\) is defined by \(K(\xi,\eta)=d\omega(\xi,\eta)+[\omega(\xi),\omega(\eta)]\) for \(\xi,\eta\in\mathfrak X(\mathcal G)\). Since \(K\) is horizontal and \(P\)-equivariant, it can be interpreted as \(\kappa\in\Omega^2(M,\mathcal AM)\), where \(\mathcal AM=\mathcal G\times_P\mathfrak g\) is the adjoint tractor bundle. In the same way, we can interpret it as a two-form on \(A\) with values in the associated bundle \(\mathcal G\times_{G_0}\mathfrak g\). Since \(K\) is horizontal over \(M\), it follows that this two form vanishes upon insertion of one tangent vector from \(L^+\subset TA\). In view of the \(G_0\)-invariant decomposition \(\mathfrak g=\mathfrak g_-\oplus\mathfrak g_0\oplus\mathfrak p_+\) we can decompose that two-form further. To do this, we denote by \(\operatorname{End}_0(TA)\) the associated bundle \(\mathcal G\times_{G_0}\mathfrak g_0\). Via the adjoint action, this can naturally be viewed as a subbundle of \(L(TA,TA)\).

Definition 2.10

The components \(T\in \Omega^2(A,L^-)\), \(W\in\Omega^2(A,\operatorname{End}_0(TA))\) and \(Y\in \Omega^2(A,L^+)\) of the two form on \(A\) induced by \(K\) are called the universal torsion, the universal Weyl curvature and the universal Cotton-York tensor of the parabolic geometry \((p:\mathcal G\to M,\omega)\).

The following result follows directly from the definitions.

Proposition 2.11

For any smooth section \(s:M\to A\), the pullbacks \(s^*T\in\Omega^2(M,\operatorname{gr}(TM))\), \(s^*W\in\Omega^2(M,\operatorname{End}_0(TM))\) and \(s^*Y\in\Omega^2(M,\operatorname{gr}(T^*M))\) correspond to the components of the Cartan curvature \(\kappa\in\Omega^2(M,\mathcal AM)\) under the isomorphism \(\mathcal AM\cong\operatorname{gr}(\mathcal AM)\cong \operatorname{gr}(TM)\oplus\operatorname{End}_0(TM)\oplus\operatorname{gr}(T^*M)\) induced by the Weyl structure determined by \(s\).

These quantities are related to data associated to the Weyl structure determined by \(s\) in Section 5.2.9 of [16] and these results can be easily recovered in the current context.

On the level of \(A\), the best way to interpret the components of the Cartan curvature is via the torsion and curvature of the canonical connection \(D\). This interpretation will also be crucial for the analysis of the intrinsic geometric structure on \(A\) in 3 below. To formulate the result, we need a bit more notation. The Lie bracket is a \(G\)-equivariant, skew symmetric bilinear map \(\mathfrak g\times\mathfrak g\to\mathfrak g\). Now we can restrict this to entries from \(\mathfrak g_-\oplus\mathfrak p_+\) and then decompose the values according to \(\mathfrak g=(\mathfrak g_-\oplus\mathfrak p_+)\oplus\mathfrak g_0\), and the result will still be \(G_0\)-equivariant. The first component induces a two-form on \(A\) with values in \(TA\) which we denote by \(\{\ ,\ \}\). Similarly, the \(\mathfrak g_0\)-component of the bracket defines a two-form \(\{\ ,\ \}_0\) on \(A\) with values in \(\operatorname{End}_0(TA)\). Using this, we formulate

Theorem 2.12

Let \(A\to M\) be the bundle of Weyl structures associated to a parabolic geometry, and let \(D\) be the canonical connection on \(TA\). Let \(\tau\in\Omega^2(A,TA)\) be the torsion and \(\rho\in\Omega^2(A,L(TA,TA))\) be the curvature of \(D\). Then we have:

(1) The \(TA\)-valued two form \(\tau+\{\ ,\ \}\) vanishes upon insertion of one section of \(L^+\). On \(\Lambda^2L^-\), its components in \(L^-\) and \(L^+\) are the tensors \(T\) and \(Y\) from Definition 2.10, respectively.

(2) The curvature \(\rho\) has values in \(\operatorname{End}_0(TA)\subset L(TA,TA)\). Moreover, \(\rho+\{\ ,\ \}_0\) vanishes upon insertion of one section of \(L^+\) and coincides with the tensor \(W\) from Definition 2.10 on \(\Lambda^2L^-\).

Proof. This follows from well known results on the curvature and torsion of the affine connection induced by a reductive Cartan geometry. For \(\xi\in\mathfrak X(A)\), let \(\xi^h\in\mathfrak X(\mathcal G)\) be the horizontal lift. Then \(\omega(\xi^h):\mathcal G\to (\mathfrak g_-\oplus\mathfrak p_+)\) is the equivariant function corresponding to \(\xi\). Taking a second field \(\eta\), the bracket \([\xi^h,\eta^h]\) lifts \([\xi,\eta]\) so \(\omega_{\pm}([\xi^h,\eta^h])\) is the equivariant function representing \([\xi,\eta]\).

(1) From these considerations and the definition of the exterior derivative, it follows readily that \(d\omega_{\pm}(\xi^h,\eta^h)\) is the equivariant function representing \(\tau(\xi,\eta)\). On the other hand, the component of \([\omega(\xi^h),\omega(\eta^h)]\) in \(\mathfrak g_-\oplus\mathfrak p_+\) of course represents \(\{\xi,\eta\}\), so the claim follows from the definition of the curvature of a Cartan connection.

(2) It is also well known that \(-\omega_0([\xi^h,\eta^h])\) is the function representing \(\rho(\xi,\eta)\). Since the \(\mathfrak g_0\)-component of \([\omega(\xi^h),\omega(\eta^h)]\) clearly represents \(\{\xi,\eta\}_0\), the result again follows from the definition of the Cartan curvature.