Home Generalized Ricci flow on nilpotent Lie groups
Article Open Access

Generalized Ricci flow on nilpotent Lie groups

  • Fabio Paradiso ORCID logo EMAIL logo
Published/Copyright: June 30, 2021

Abstract

We define solitons for the generalized Ricci flow on an exact Courant algebroid. We then define a family of flows for left-invariant Dorfman brackets on an exact Courant algebroid over a simply connected nilpotent Lie group, generalizing the bracket flows for nilpotent Lie brackets in a way that might make this new family of flows useful for the study of generalized geometric flows such as the generalized Ricci flow. We provide explicit examples of both constructions on the Heisenberg group. We also discuss solutions to the generalized Ricci flow on the Heisenberg group.

MSC 2010: 53D18; 53C44; 53C30

1 Introduction

Generalized geometry, building on the work of Hitchin [9] and Gualtieri [7] and the structure of Courant algebroids, constitutes a rich mathematical environment. The main idea behind it lies in the shift of point of view when studying structures on a differentiable manifold Mn, replacing the tangent bundle TM with the generalized tangent bundle

𝕋M=TMT*M.

More explicitly, in the language of G-structures, one studies reductions of GL(𝕋M), the GL2n-principal bundle of frames of 𝕋M.

A reduction to the orthogonal group O(n,n) always exists, thanks to the nondegenerate symmetric bilinear form of neutral signature

(1.1)X+ξ,Y+η=12(η(X)+ξ(Y)),

so that one usually only considers structures which are reductions of O(𝕋M), the O(n,n)-reduction of GL(𝕋M) determined by this natural pairing.

In this spirit, for example, a generalized almost complex structure on M2m, defined by an orthogonal automorphism 𝒥 of 𝕋M, i.e. 𝒥2=-Id𝕋M, determines a U(m,m)-reduction of GL(𝕋M). The integrability of such a structure is expressed through an involutivity condition with respect to a natural bracket operation, called the Dorfman bracket:

(1.2)[X+ξ,Y+η]=[X,Y]+Xη-ιYdξ.

On the other hand, a generalized Riemannian metric on Mn, defined by a symmetric (with respect to ,) and involutive automorphism 𝒢 of 𝕋M, determines an O(n)×O(n)-reduction of GL(𝕋M).

More generally, one can consider a Courant algebroidE over M, namely a smooth vector bundle over M endowed with a pairing , and a bracket [,] satisfying certain properties so that 𝕋M, endowed with (1.1) and (1.2), is a special case. On the generic Courant algebroid, one can then study reductions of GL(E) such as generalized almost complex structures and generalized (pseudo-)Riemannian metrics.

In [22, 5, 6], the classical Ricci flow of Hamilton [8] and the B-field renormalization group flow of Type II string theory (see [20]) were generalized to a flow of generalized (pseudo-)Riemannian metrics on a Courant algebroid E over a smooth manifold M. The generalized Ricci flow, as we shall refer to this flow from now on, is actually a flow for a pair of families of generalized (pseudo-)Riemannian metrics 𝒢Aut(E) and divergence operators div:Γ(E)C(M), the latter of which are required in order to “gauge-fix” curvature operators associated with a generalized (pseudo-)Riemannian metric.

The paper is organized as follows: Section 2 is devoted to a review of the setting of generalized geometry – including the notions of Courant algebroid, generalized curvature tensors and the definition of generalized Ricci flow – and of the algebraic framework of nilpotent Lie groups.

In Section 3, we introduce the notion of generalized Ricci soliton, which derives from the study of self-similar (in a suitable sense) solutions to the generalized Ricci flow on exact Courant algebroids. This condition generalizes the Ricci soliton condition Rcg=λg+Xg, where Rcg denotes the Ricci tensor of g, λ and Xg denotes the Lie derivative of g with respect to a vector field X. We show that, when working on a Lie group and considering left-invariant structures, this condition descends to an algebraic condition on the Lie algebra of the group.

Borrowing from the ideas of Lauret, in Section 4, we consider left-invariant Dorfman brackets on simply connected nilpotent Lie groups, describing them as elements of an algebraic subset of the vector space of skew-symmetric bilinear forms on n(n)* for the suitable n. We then define a family of flows of such structures, showing that they generalize the constructions known in literature as bracket flows, which have been extensively used to rephrase geometric flows on (nilpotent) Lie groups (see, for example, [15]). This justifies our definition of generalized bracket flows.

In Section 5, we perform explicit computations of generalized Ricci solitons and exhibit an example of generalized bracket flow on the three-dimensional Heisenberg group.

In Section 6, we study solutions of the generalized Ricci flow on the Heisenberg group, highlighting the differences with the classical Ricci flow.

2 Preliminaries

2.1 Courant algebroids

Let V be a real vector space of dimension n. We start by recalling a few facts about the algebra of the vector space VV*; for more details, see [7].

The vector space VV* can be endowed with a natural symmetric bilinear form of neutral signature

X+ξ,Y+η=12(η(X)+ξ(Y))

and with a canonical orientation provided by the preimage of 1 in the isomorphism

φ:Λ2n(VV*)=ΛnVΛnV*,

sending (X1Xn)(ξ1ξn) into det(ξi(Xj))ij.

Consider the Lie group SO(VV*)SO(n,n) of automorphisms of VV* preserving the pairing , and the canonical orientation. Its Lie algebra

𝔰𝔬(VV*)𝔰𝔬(n,n)

consists of endomorphisms T𝔤𝔩(VV*) which are skew-symmetric with respect to ,, namely

(2.1)Tz1,z2+z1,Tz2=0

for all z1,z2VV*. By seeing T as a block matrix, (2.1) dictates T to be of the form

T=(ϕβB-ϕ*)

for some ϕ𝔤𝔩(V), BΛ2V* and βΛ2V, recovering the fact that

𝔰𝔬(VV*)Λ2(VV*)*Λ2V*(V*V)Λ2V,

where the former isomorphism is given by TT,.

Via the exponential map

exp:𝔰𝔬(VV*)SO(VV*),

we obtain distinguished elements of SO(VV*):

  1. We have

    eB=(Id0BId):X+ξX+ξ+ιXB,

    called B-field transformations.

  2. We have

    eϕ=(eϕ00(e-ϕ)*),

    which extends to an embedding of the whole GL(V) into SO(VV*), sending AGL(V) into

    𝐀=(A00(A*)-1).

    In the case V=n, the image of this embedding will be denoted by GLn.

Let M be an oriented smooth manifold of positive dimension n.

Definition 2.1.

A Courant algebroid over M is a smooth vector bundle EM equipped with:

  1. a fiberwise nondegenerate bilinear form ,, which allows to identify E and its dual E*, viewing zE as z,E*,

  2. a bilinear operator [,] on Γ(E),

  3. a bundle homomorphism π:ETM, called the anchor,

which satisfy the following properties for all z,ziΓ(E), i=1,2,3, fC(M):

  1. [z1,[z2,z3]]=[[z1,z2],z3]+[z2,[z1,z3]] (Jacobi identity).

  2. π[z1,z2]=[π(z1),π(z2)].

  3. [z1,fz2]=f[z1,z2]+π(z1)(f)z2.

  4. [z,z]=12𝒟z,z, where 𝒟π*d:C(M)Γ(E).

  5. π(z1)z2,z3=[z1,z2],z3+z2,[z1,z3].

Definition 2.2.

A Courant algebroid E over M is exact if the short sequence

(2.2)0T*Mπ*E𝜋TM0

is exact, namely if the anchor map is surjective and its kernel is exactly the image of π*.

By the classification of Ševera [21], isomorphism classes of exact Courant algebroids over M are in bijection with the elements of the third de Rham cohomology group of M, i.e. H3(M): an exact Courant algebroid with Ševera class[H]H3(M) is isomorphic to the Courant algebroid EH=𝕋MTMT*M over M with pairing of neutral signature

(2.3)X+ξ,Y+η=12(η(X)+ξ(Y))

and (twisted) Dorfman bracket

(2.4)[X+ξ,Y+η]=[X,Y]+Xη-ιYdξ+ιYιXH

for any H[H]. Such isomorphisms are obtained explicitly via the choice of an isotropic splitting to (2.2), while B-field transformations, BΓ(Λ2T*M), provide explicit isomorphisms

eB:EHEH-dB.

In what follows, let E be a Courant algebroid over M, with rk(E)=2n and pairing , of neutral signature.

Definition 2.3.

A generalized Riemannian metric on E is an O(n)×O(n)-reduction of O(E), the O(n,n)-principal subbundle of orthonormal frames of E with respect to the pairing ,. Explicitly, it is equivalently determined by

  1. a subbundle E+ of E, rk(E+)=n, on which , is positive-definite;

  2. an automorphism 𝒢 of E which is involutive, namely 𝒢2=IdE, and such that 𝒢, is a positive-definite metric on E.

Given E+, denoting by E- its orthogonal complement with respect to ,, we define 𝒢 by 𝒢|E±=±IdE±. Then E± can be recovered as the ±1-eigenbundles of 𝒢. Given zE, we shall denote by z± its orthogonal projections along E±.

Example 2.4.

Every generalized Riemannian metric on the exact Courant algebroid EH is of the form

𝒢=eB(0g-1g0)e-B

for some Riemannian metric g and for some 2-form B on M (see [7, Section 6.2]). The corresponding E± are

E±=eB{X±g(X):XTM},

where by g(X) we mean g(X,). Notice that 𝒢 is of the form

(0g-1g0)

in the splitting EH+dB.

2.2 Generalized curvature

We now recall the definition of generalized connection on a Courant algebroid E, showing how these objects can be used to associate curvature operators with a generalized Riemannian metric 𝒢. Unlike the Riemannian case, where the uniqueness of the Levi-Civita connection allows to single out canonical curvature operators for a given Riemannian metric, in the generalized setting there are plenty of torsion-free generalized connections compatible with a generalized Riemannian metric 𝒢, and these may define different curvature operators. To gauge-fix them, one needs to additionally fix a divergence operator. For further details, we refer the reader to [5, 3].

Definition 2.5.

A generalized connection on a Courant algebroid E is a linear map

D:Γ(E)Γ(E*E)

which satisfies a Leibniz rule and a compatibility condition with ,:

D(fz)=f(Dz)+𝒟fz,
𝒟z1,z2,=Dz1,z2+z1,Dz2

for all z,z1,z2Γ(E) and fC(M), where Dz1z2Dz2(z1).

Given a generalized Riemannian metric 𝒢, a generalized connection D is compatible with 𝒢 if D𝒢=0, where D denotes the induced E-connection on the tensor bundle E*EEnd(E). Equivalently, D is compatible with 𝒢 if D(Γ(E±))Γ(E*E±).

The torsionTDΓ(Λ2E*E) of a generalized connection D on E is defined by

TD(z1,z2)=Dz1z2-Dz2z1-[z1,z2]+(Dz1)*z2.

If TD=0, the generalized connection D is said to be torsion-free.

Given a generalized connection D on E which is compatible with a generalized Riemannian metric 𝒢, one can define curvature operators

RD±Γ(E±*E*𝔬(E±)),

where 𝔬(E±)=,-1Λ2E±* denotes the Lie algebra of skew-symmetric endomorphisms of E± with respect to ,, by

RD±(z1±,z2)z3±=Dz1±Dz2z3±-Dz2Dz1±z3±-D[z1±,z2]z3±.

One then has associated Ricci tensors

RcD±Γ(E*E±*),
RicD±Γ(E*E±),

defined by

RcD±(z1,z2±)=tr(z±RD±(z±,z1)z2±),
RicD±=,-1RcD±.

Definition 2.6.

A divergence operator on E is a first-order differential operator div:Γ(E)C(M) satisfying the Leibniz rule

div(fz)=π(z)f+fdiv(z),

fC(M), zΓ(E). Given a generalized connection D on E, one may define the associated divergence operator

divD(z)=tr(Dz).

Remark 2.7.

Divergence operators on E form an affine space over the vector space Γ(E)Γ(E*). Fixing a divergence operator div0, any other divergence operator div is of the form

div=div0-z,

for some zΓ(E).

Proposition 2.8 ([5, Proposition 4.4]).

Let Di, i=1,2, be torsion-free generalized connections on E compatible with a given generalized Riemannian metric G. Suppose divD1=divD2. Then, RcD1±=RcD2±.

Moreover, for any divergence operator div and generalized Riemannian metric 𝒢 on E, the set of torsion-free generalized connections D on E which are compatible with 𝒢 and such that divD=div is nonempty (see [5, Section 3.2]). Thus, Ricci tensors Rc𝒢,div± are well-defined as equal to RcD± for any such generalized connection D.

Example 2.9.

On the exact Courant algebroid EH over M, let

𝒢=(0g-1g0)

and

divg,z(X+ξ)=dVg-1XdVg-z,X+ξ,

where g is a Riemannian metric, dVg its associated Riemannian volume form and zΓ(EH). Then, via the isomorphism π+=π|E+:E+TM, the Ricci tensor Rc+ of (𝒢,divg,θ) is given by

(2.5)Rc𝒢,divg,z+=Rcg-𝑔14HH-12dg*H+12g,H+θ,

where

  1. RcgΓ(S+2T*M) is the Ricci tensor associated with g,

  2. 𝑔HHΓ(S2T*M), with

    𝑔HH(X,Y)=g(ιXH,ιYH),
  3. dg*=-*𝑔d*𝑔:Γ(Λ3T*M)Γ(Λ2T*M) is the Hodge codifferential associated with the metric g and the fixed orientation, with *𝑔 being the Hodge star operator,

  4. g,H+=g+12g-1H is the Bismut connection with torsion H, with g denoting the Levi-Civita connection of g,

  5. θΓ(T*M) is given by θ=2g(πz+,)=g(X,)+ξ if z=X+ξ.

See [6, Proposition 3.30] for the proof of this fact (cf. also [13]).

2.3 Generalized Ricci flow

We now review the framework of the generalized Ricci flow first introduced in [22, 5] and later described and studied in [6] by Garcia-Fernandez and Streets. Consider a smooth family of generalized Riemannian metrics (𝒢(t))tI on E, I, with respective eigenbundles E±|t. Its variation 𝒢˙(t) exchanges the eigenbundles E±|t, so that 𝒢˙(t)=𝒢˙+(t)+𝒢˙-(t), with

𝒢˙±(t)Γ(E|t*E±|t).

Definition 2.10.

[5, Definition 5.1] A smooth pair of families (𝒢(t),div(t))tI of generalized Riemannian metrics and divergence operators on E is a solution to the generalized Ricci flow if it satisfies

𝒢˙+(t)=-2Rict+

for all t in the interior of I, where Rict+Ric𝒢(t),div(t)+.

On an exact Courant algebroid, the system may be written as follows.

Proposition 2.11 ([5, Example 5.4]).

Let E be an exact Courant algebroid on an oriented smooth manifold M, with Ševera class [H]H3(M). Fix an isotropic splitting EH=TM for E and consider the pair of smooth families (G(t),div(t))tI defined by

𝒢(t)=eB(t)(0g(t)-1g(t)0)e-B(t),
div(t)=divg(t),z(t),

where (g(t))Γ(S+2T*M), (B(t))Γ(Λ2T*M) and (z(t))Γ(E).

Then (G(t),div(t))tI is a solution of the generalized Ricci flow on E if and only if the families

(g(t),B(t),θ(t))tI,with θ(t)=2g(πz(t)+,)Γ(T*M),

solve the equation

(2.6)g˙(t)=-2(Rcg(t)-g(t)14H(t)H(t)-12dg(t)*H(t)+12g(t),H(t)+θ(t))+B˙(t),

where H(t)=H+dB(t).

Separating the symmetric and skew-symmetric part of (2.6), one gets (see [24])

{g˙(t)=-2Rcg(t)+g(t)12H(t)H(t)-12g(t)-1θ(t)g(t),B˙(t)=-dg(t)*H(t)+12dθ(t)-12ιg(t)-1θ(t)H(t),

where one has that

12g(t)-1θ(t)g(t)=S(g(t),H(t)+θ(t)),12dθ(t)-12ιg(t)-1θ(t)H(t)=A(g(t),H(t)+θ(t))

are respectively the symmetric and skew-symmetric parts of g(t),H(t)+θ(t).

The pair (g(t),H(t)) evolves as

(2.7){g˙(t)=-2Rcg(t)+g(t)12H(t)H(t)-12g(t)-1θ(t)g(t),H˙(t)=-Δg(t)H(t)-12g(t)-1θ(t)H(t),

where Δg=ddg*+dg*d denotes the Hodge–Laplacian operator associated with g and the fixed orientation. Notice how, up to scaling, the pluriclosed flow introduced in [23] is equivalent to a particular case of the generalized Ricci flow, as is proven in [24, Propositions 6.3 and 6.4]. By [24, Theorem 6.5], a solution to (2.7) can be pulled back to a solution of

(2.8){g˙(t)=-2Rcg(t)+g(t)12H(t)H(t),H˙(t)=-Δg(t)H(t),

via the one-parameter family of diffeomorphisms generated by 14g(t)-1θ(t).

2.4 Simply connected nilpotent Lie groups

We briefly recall the structure of simply connected nilpotent Lie groups, in the description of Lauret (see, for example, [15]).

Every simply connected nilpotent Lie group G is diffeomorphic to its Lie algebra of left-invariant fields 𝔤 via the exponential map. Identifying 𝔤 with n via the choice of a basis, denote by μΛ2(n)*n the induced Lie bracket. Now, exploiting the Campbell–Baker–Hausdorff formula, we get

exp(X)exp(Y)=exp(X+Y+pμ(X,Y)),

X,Y𝔤n, where pμ is an n-valued polynomial in the variables X,Y, and one can endow n with the operation μ, i.e.

XμY=X+Y+pμ(X,Y),

so that exp:(n,μ)G is an isomorphism of Lie groups. Therefore, the set of isomorphism classes of simply connected nilpotent Lie groups is parametrized by the set of nilpotent Lie brackets on n: these form an algebraic subset of the vector space of skew-symmetric bilinear forms on n, i.e.

𝒱nΛ2(n)*n,

which parametrizes all skew-symmetric algebra structures on n. Coordinates for 𝒱n can be obtained by fixing a basis {ei}i=1n for n: this allows to determine the so-called structure constants of any fixed μ𝒱n as the real numbers {μijk:i,j,k=1,,n} given by

μ(ei,ej)=μijkek.

One can then consider

n{μ𝒱n:μ satisfies the Jacobi identity},

the algebraic subset of 𝒱n consisting of Lie brackets on n, and

𝒩n{μn:μ is nilpotent},

which parametrizes all nilpotent Lie algebra structures on n. By the previous remarks, 𝒩n parametrizes all n-dimensional simply connected nilpotent Lie groups, up to isomorphism.

Let us consider the following family of Riemannian metrics on n:

(2.9){gμ,q:μ𝒩n,q is a positive definite bilinear form on n},

where gμ,q coincides with q at the origin and is left-invariant with respect to the nilpotent Lie group operation μ. The set (2.9) is actually the set of all Riemannian metrics on n which are invariant by some transitive action of a nilpotent Lie group. By [25, Theorem 3], the Riemannian manifolds (n,gμ,q) (varying n, μ and q) are, up to isometry, all the possible examples of simply connected homogeneous nilmanifolds, namely connected Riemannian manifolds admitting a transitive nilpotent Lie group of isometries.

The Riemannian metrics in (2.9) are not all distinct, up to isometry: it was shown again in [25, Theorem 3] that gμ,q is isometric to gμ,q if and only if there exists hGLn such that μ=h*μ and q=h*q. By convention we shall set gμgμ,,, where , denotes the standard scalar product.

Since the Riemannian metrics gμ,q are completely determined by their value at 0 and by the Lie bracket μ, so will be all curvature quantities related to gμ,q. In particular, we are interested in Riemannian metrics gμ and their Ricci tensor, which we shall encounter in two guises, which we denote by

RcμRcgμ(0)S2(n)*(n)*(n)*,
RicμRicgμ(0)(n)*n=𝔤𝔩n,

with Rcμ(X,Y)=Ricμ(X),Y, X,Yn.

For these, explicit formulas can be computed [14]. Let {ei}i=1n be the standard basis of n, which, in particular, is orthonormal with respect to ,: one has

(2.10)Rcμ(X,Y)=-12μ(X,ek),elμ(Y,ek),el+14μ(ek,el),Xμ(ek,el),Y,

so that, if Rcμ=(Rcμ)ijeiej and Ricμ=(Ricμ)ijeiej, one has

(2.11)(Rcμ)ij=(Ricμ)ij=-12μiklμjkl+14μkliμklj.

Notice that one can use formulas (2.10) and (2.11) to define RcμS2(n)* and Ricμ𝔤𝔩n for any μ𝒱n.

3 Generalized Ricci solitons

Just as Ricci soliton metrics arise from self-similar solutions of the Ricci flow, generalized Ricci solitons arise from self-similar solutions of the generalized Ricci flow. We focus on exact Courant algebroids, defining a family of generalized Riemannian metrics, whose initial one is determined by a Riemannian metric on the base manifold; imposing that this family (together with a family of divergence operators) is a solution of the generalized Ricci flow, we draw necessary conditions on said Riemannian metric: these conditions generalize the Ricci soliton condition, leading to the definition of generalized Ricci solitons.

Let E be a Courant algebroid over an oriented smooth manifold M with Ševera class [H0]H3(M). Fixing an isotropic splitting EH0=𝕋M, we consider a smooth self-similar pair of families (𝒢(t),div(t))tI, 0I, on E of the form

𝒢(t)=eB(t)(0(c(t)φt*g0)-1c(t)φt*g00)e-B(t),
div(t)=divg(t),θ(t),

where g0Γ(S+2(T*M)) is a Riemannian metric, c:I is smooth and positive, c(0)=1, (φt) is a one-parameter family of diffeomorphisms of M, (B(t))Γ(Λ2T*M), B(0)=0, (θ(t))Γ(T*M), θ(0)=θ0Γ(T*M), and g(t)=c(t)φt*g0.

By Proposition 2.11, such (𝒢(t),div(t))tI is a solution of the generalized Ricci flow if and only if

{c˙(t)φt*g0+c(t)φt*Ytg0=-2Rcg(t)+g(t)12H(t)H(t)+14g(t)-1θ(t)g(t),B˙(t)=-dg(t)*H(t)-14dθ(t)+14ιg(t)-1θ(t)H(t),

where H(t)=H0+dB(t) and (Yt)tIΓ(TM) is such that

ddtφt(x)=Yt(φt(x))

for all tI, xM.

Setting t=0 and rearranging the terms, we obtain

(3.1){Rcg0=λg0+Xg0+g014H0H0-14g0-1θ0g0,ω=-dg0*H0+12dθ0-12ιg0-1θ0H0,

where -2λ=c˙(0), -2X=Y0Γ(TM) and ω=B˙(0)Γ(Λ2T*M). Summing together the two equations of (3.1), which involve symmetric and skew-symmetric tensor fields, respectively, one has

(3.2)Rcg0=λg0+Xg0+g014H0H0-12g0,H0+θ0+12dg0*H0+12ω,

which is therefore equivalent to (3.1). We can now introduce the following definition, which generalizes the notion of Ricci soliton.

Definition 3.1.

A Riemannian metric g0 on M is called a generalized Ricci soliton if there exist λ, XΓ(TM), H0Γ(Λ3T*M) closed, θ0Γ(T*M), and ωΓ(Λ2T*M) such that (3.2), or equivalently (3.1), holds.

When working on a Lie group G, for simplicity one can assume all structures to be left-invariant, so that the generalized Ricci soliton condition reduces to an algebraic condition on structures on the Lie algebra of G, i.e. (𝔤,μ).

In the context of semi-algebraic Ricci solitons, it was proven in [12, Theorem 1.5] that, if g0 is a left-invariant Riemannian metric on G, the Lie derivative of g0 with respect to a left-invariant vector field X can be written as

Xg0=g0(12(D+Dt))=g0(12(D+Dt),)

for some D=DXDer(𝔤), where Der(𝔤) denotes the algebra of derivations of 𝔤. It was then shown in [11, Theorem 1] (generalizing the already known fact for the simply connected nilpotent case in [14, Proposition 1.1]) that D can be chosen to be symmetric with respect to g0, so that one always has

Xg0=g0(D)=g0(D,)

for some D=DXDer(𝔤)Sym(𝔤,g0). Then (3.2) becomes

(3.3)Rcg0=λg0+g0(D)+g014H0H0+12dg0*H0-12g0,H0+θ0+12ωS2𝔤*

for g0S+2𝔤*, λ, DDer(𝔤)Sym(𝔤,g0), H0Λ3𝔤* (with dμH0=0, and dμ:Λ3𝔤*Λ4𝔤* denoting the Chevalley–Eilenberg differential of the Lie algebra (𝔤,μ)), θ0𝔤*, and ωΛ2𝔤*, or equivalently

(3.4){Rcg0=λg0+g0(D)+g014H0H0-14g0-1θ0g0,ω=-dg0*H0+12dθ0-12ιg0-1θ0H0.

Notice that dg0*H0 is still a left-invariant form since the Hodge star operator commutes with pull-backs via orientation-preserving isometries of g0, such as left translations Lg, gG, by left-invariance of g0.

4 Generalized bracket flows

Bracket flows have proven to be a powerful tool in the study of geometric flows on homogeneous spaces. This technique was first fully formalized by Lauret to study the Ricci flow on nilpotent Lie groups [15]. In particular, Lauret proved that the Ricci flow on an n-dimensional simply connected nilpotent Lie group G starting from a left-invariant Riemannian metric g0 is equivalent to an ODE system defined on the variety of nilpotent Lie algebras 𝒩n:

{μ˙(t)=-π(Ricμ(t))μ(t),μ(0)=μ0,

where μ0 is the nilpotent Lie bracket associated with a fixed g0-orthonormal left-invariant frame and π:𝔤𝔩n𝔤𝔩(𝒱n), given by

(π(ϕ)μ)(X,Y)=ϕμ(X,Y)-μ(ϕX,Y)-μ(X,ϕY),ϕ𝔤𝔩n,μ𝒱n,X,Yn,

is the differential of the standard GLn-action on 𝒱n:

(Aμ)(X,Y)=Aμ(A-1X,A-1Y),AGLn,μ𝒱n,X,Yn.

More generally, in literature many other bracket flows have been considered (see, for example, [1, 4, 16, 19, 17, 18, 2]): these can be written in the form

(4.1){μ˙(t)=-π(ϕ(μ(t)))μ(t),μ(0)=μ0

for some smooth function ϕ:𝒱n𝔤𝔩n.

4.1 Left-invariant Dorfman brackets

Let E be an exact Courant algebroid over a real Lie group G. We shall be interested in the case when G is simply connected and nilpotent, so that we know that G is isomorphic to (n,μ) for some Lie bracket μ𝒩n.

As we have recalled, there exists a unique cohomology class [H]H3(n) such that, for any H[H], E is isomorphic to EH=TnT*n, endowed with the inner product , in (2.3) and Dorfman bracket [,]H in (2.4).

The whole structure descends to a structure on left-invariant sections, viewed as elements of n(n)*, if and only if the 3-form H is left-invariant. Explicitly, when X+ξ,Y+ηn(n)*, the Dorfman bracket [,]H reduces to the operator

MH(X+ξ,Y+η)=μ(X,Y)-ηadμ(X)+ξadμ(Y)+ιYιXH
=μ(X,Y)-ημ(X,)+ξμ(Y,)+H(X,Y,).

We call such a bilinear operator a (nilpotent) left-invariant Dorfman bracket.

As one can check directly and also deduce from the axioms of Courant algebroids, a left-invariant Dorfman bracket is totally skew-symmetric, namely

MH(,),Λ3(n(n)*)*.

By a little abuse, we can say MHΛ3(n(n)*)*, by identifying Λ3(n(n)*)* with a subset of

𝓥nΛ2(n(n)*)*(n(n)*).

We shall denote the set of left-invariant Dorfman brackets on n by 𝓒n. By definition, it is clear that

𝓒n1:1{(μ,H)n×Λ3(n)*,dμH=0}.

Equivalently, a quick analysis using the axioms of Courant algebroids and the previous remarks shows that 𝓒n can be identified with the algebraic subset of 𝓥n consisting of all brackets M𝓥n such that:

  1. MΛ3(n(n)*)*,

  2. M((n)*,(n)*)=0,

  3. M satisfies the Jacobi identity.

Given any M𝓥n, one can define the structure constants with respect to the standard basis of n as the (2n)3=8n3 real numbers Mi¯¯j¯¯k¯¯, i,j,k=1,,n, given by

M(ei,ej)=Mi¯j¯k¯ek+Mi¯j¯k¯ek,
M(ei,ej)=Mi¯j¯k¯ek+Mi¯j¯k¯ek,
M(ei,ej)=Mi¯j¯k¯ek+Mi¯j¯k¯ek,
M(ei,ej)=Mi¯j¯k¯ek+Mi¯j¯k¯ek.

Taking MH𝓒n, the structure constants are skew-symmetric in all three indices and vanish when two or more indices are overlined. The remaining structure constants are determined by μ and H. More precisely,

(MH)i¯j¯k¯=μijk,(MH)i¯j¯k¯=Hijk.

The set of nilpotent left-invariant Dorfman brackets on n, denoted by 𝓝n, is an algebraic subset of 𝓥n contained in 𝓒n. It is easy to see that its elements are exactly those Dorfman brackets MH for which μ𝒩n.

4.2 Generalized bracket flows

To introduce classical bracket flows, one uses the differential of the GLn-action on 𝒱n. In the same spirit, one can consider the natural GL(n(n)*) on 𝓥n:

(FM)(z1,z2)=FM(F-1z1,F-1z2),FGL(n(n)*),M𝓥n,z1,z2n(n)*,

which induces an action of GLnGLnSO(n(n)*) on 𝓥𝒏, preserving both 𝓒n and 𝓝n.

Now, identifying M𝓒n with (μ,H)𝒱n×Λ3(n)*, it is evident that this action distributes as

A(μ,H)=(Aμ,AH),

where AGLn and AH(A-1)*H.

We denote the differential of this action again by π:𝔤𝔩n𝔤𝔩(𝓥n): for M𝓥n and ϕ𝔤𝔩n, one has

π(ϕ)M=dds|s=0(esϕM)TM𝓥n𝓥n.

Since the curve sesϕM is contained in the orbit GLnM, in this interpretation one has

(4.2)π(ϕ)MTM(GLnM).

Following the ideas in the work of Lauret (see [15]), these remarks suggest the idea of defining a flow, which we shall refer to as generalized bracket flow, on the vector space 𝓥n, of the form

(4.3){M˙(t)=-π(ϕ(M(t)))M(t),M(0)=M0

for some smooth function ϕ:𝓥n𝔤𝔩n and some M0𝓝n. By (4.2), a solution M(t) to (4.3) satisfies

M˙(t)TM(t)(GLnM(t))TM(t)𝓝nfor all t,

so that the curve M(t) is entirely contained in 𝓝n. For this reason, the function ϕ may also be defined on 𝓝n only.

System (4.3) may be rewritten as the following ODE system on 𝒩n×Λ3(n)*:

(4.4){μ˙(t)=-π(ϕ(μ(t),H(t)))μ(t),H˙(t)=-π(ϕ(μ(t),H(t)))H(t),μ(0)=μ0𝒩n,H(0)=H0Λ3(n)*,dμ0H0=0,

where π denotes the differential of the GLn-action on 𝒱n or Λ3(n)*.

In what follows, we shall omit the time dependencies of the quantities involved. Fixing the standard basis {ei}i=1n for n, we shall denote by ϕij, i,j=1,,n, the entries of the generic ϕGLn with respect to it, such that ϕ(ei)=ϕijej for all i=1,,n. One can then compute the coordinate expression for the evolution equations (4.4), obtaining

(4.5)μ˙ijk=ϕilμljk+ϕjlμilk-ϕlkμijl,
(4.6)H˙ijk=ϕilHljk+ϕjlHilk+ϕklHijl

for i,j,k=1,,n.

Special generalized bracket flows are obtained when the 𝔤𝔩n-valued smooth function ϕ only depends on μ, i.e. ϕ=ϕ(μ): when this happens, the first equation of (4.4) is independent from the second one and corresponds to a usual bracket flow (4.1) on 𝒩n.

Classical bracket flows have proved to be a powerful tool in the study of geometric flows on (nilpotent) Lie groups. We thus expect the generalized bracket flows we have defined to be useful in the context of geometric flows in generalized geometry.

5 Examples on the Heisenberg group

In this section, we perform explicit computations for the constructions introduced in the previous sections. We focus in particular on the Heisenberg group.

The Heisenberg group H3 is a three-dimensional simply connected Lie group, which can be defined as a closed subgroup of GL3:

H3={(1ac01b001)GL3,a,b,c}.

Via the exponential map, H3 is diffeomorphic to its Lie algebra

𝔥3={(0ac00b000)𝔤𝔩3,a,b,c}.

By fixing the basis

(5.1)e1=(010000000),e2=(000001000),e3=(001000000)

for 𝔥3, the induced bracket μ𝒩3 is μ=e1e2e3 since [e1,e2]=e3 and [e1,e3]=[e2,e3]=0.

5.1 Generalized Ricci solitons on the Heisenberg group

Let H3 be the Heisenberg group, and fix the basis (5.1) for its Lie algebra 𝔥3.

In order to find generalized Ricci solitons on H3, we first notice that the codifferential dg0* is the null map for every g0S+2𝔥3* since *g0 sends Λ3𝔥3* to and d:𝔥3* is the null map. With respect to the basis {e1,e2,e3} in (5.1), the generic derivation D of 𝔥3 can be written in matrix form as

D=(a1a20a3a40a5a6a1+a4),

with ai, i=1,,6.

Let g0 be the standard metric

g0=e1e1+e2e2+e3e3

such that {e1,e2,e3} is an orthonormal basis. Now, symmetric derivations with respect to g0 are simply represented by symmetric matrices with respect to this basis:

(5.2)D=(a1a20a2a3000a1+a3),

ai, i=1,2,3. In what follows, assume

H0=ae123=a6εijkeijk,θ0=θiei,ω=12ωijeij,

where a,θi,ωij, ωij=-ωji, i,j=1,2,3, ei1ikei1eik, and εijk is equal to the sign of the permutation sending (1,2,3) into (i,j,k) whenever i, j and k are all different, and equal to 0 otherwise, by definition.

We are now ready to compute the coordinate expression for all the terms involved in (3.3):

  1. Rcg0: from (2.11), since the basis {e1,e2,e3} is orthonormal, by a direct computation, we get

    Rcg0=(-12000-1200012)

    in the fixed basis.

  2. g0(D): it is simply represented by the matrix (5.2) with respect to the orthonormal basis.

  3. g0H0H0: one has

    g0H0H0(ei,ej)=g(ιeiH0,ιejH0)=g0rlg0st(H0)irs(H0)jlt=a2εistεjst,

    so that, in matrix form, we get

    g0H0H0=(2a20002a20002a2).
  4. g0,H0+θ0: writing + instead of g0,H0+ and by the left-invariance of the quantities involved, one has

    +θ0(ei,ej)=-θ0(ei+ej).

    Now, +=g0+12g0-1H0 and, letting eig0ej=Γijkek and recalling the Koszul formula, one computes

    Γijk=-12(μjki+μikj+μjik),12g0-1H0(ei,ej)=12aεijkek,

    so that

    +θ0(ei,ej)=12θk(μjki+μikj+μjik-aεijk).

    The corresponding matrix with respect to the orthonormal basis is thus

    +θ0=(0-12θ3(1+a)12θ2(1+a)12θ3(1+a)0-12θ1(1+a)12θ2(1-a)-12θ1(1-a)0),

    so that its symmetric and skew-symmetric parts are

    S(+θ0)=(0012θ200-12θ112θ2-12θ10),
    A(+θ0)=(0-12θ3(1+a)12aθ212θ3(1+a)0-12aθ1-12aθ212aθ10).

The first equation of (3.4) gives now rise to a system of six equations in the unknowns λ, a1, a2, a3, θ1, θ2, θ3:

{12+λ+a1+12a2=0,a2=0,θ2=0,12+λ+a3+12a2=0,θ1=0,-12+λ+a1+a3+12a2=0,

which is equivalent to

{λ=-12(3+a2),a1=a3=1,a2=θ1=θ2=0.

The second equation of (3.4) now implies

ω12=-ω21=-12θ3(1+a),

while all the other ωij’s vanish.

We thus obtain generalized Ricci solitons with the data

{g0=e1e1+e2e2+e3e3,λ=-12(3+a2),D=e1e1+e2e2+2e3e3,H0=ae123,θ0=θ3e3,ω=-12θ3(1+a)e12.

Remark 5.1.

The metric g0 above is actually also a Ricci soliton in the classical sense since, by setting a=θ3=0, H0, θ0 and ω vanish, leaving g0 satisfying Rcg0=λg0+g0(D), or equivalently, applying g0-1, Ricg0=λId+D for λ=-32 and D as above. By [14, Theorem 3.5], g0 is the only left-invariant Ricci soliton on H3, up to isometry and rescaling.

5.2 A generalized bracket flow on the Heisenberg group

The definition of the gauge-corrected generalized Ricci flow (2.8) suggest the generalized bracket flow

(5.3){M˙(t)=-π(Ricμ(t)-14H(t)2)M(t),M(0)=M0𝓝n.

Here, for every HΛ3(n)*, we set H2,-1(HH), where is meant with respect to ,. Recalling (2.11), the whole endomorphism ϕ(M)=ϕ(μ,H)=Ricμ-14H2 can then be written in coordinates as

ϕij=-12μiklμjkl+14μkliμklj-14HiklHjkl,

with respect to the standard basis of n.

Now, let n=3 and let μ0 be the Heisenberg Lie bracket μ0=e1e2e3. Let H0 be the generic (trivially dμ0-closed) 3-form H0=ce123, c. Then, using (2.11), (4.5) and (4.6), it is easy to compute that the solution to (5.3) is of the form M(t)=(x(t)μ0,y(t)e123), with x(t) and y(t) satisfying the ODE system

(5.4){x˙=-32x3-12xy2,y˙=-32y3-12x2y,x(0)=1,y(0)=a.

It is easy to see that the solution (x(t),y(t)) is defined for all positive times and converges to (0,0) since

ddt(x(t)2+y(t)2)=2(x(t)x˙(t)+y(t)y˙(t))-(x(t)2+y(t)2)2,

so that, by comparison, we get

x(t)2+y(t)21+a21+(1+a2)t

for all t0. For a=1, the explicit solution to (5.4) is given by

x(t)=y(t)=(1+4t)-12.

defined for t(-14,).

6 Generalized Ricci flow on the Heisenberg group

Let us consider the gauge-corrected generalized Ricci flow (2.8) on the three-dimensional Heisenberg group H3, with initial data

g0=e1e1+e2e2+e3e3,H0=ae123,a.

Denoting by (g(t),H(t)) the solution at time t, we adopt the ansatz

g(t)=g1(t)e1e1+g2(t)e2e2+g3(t)e3e3,gi(0)=1,i=1,2,3,

while H(t)=H0 is necessarily constant since dg* and Δg are null maps for every left-invariant Riemannian metric g, as remarked in Section 5.1.

An explicit computation yields

Rcg(t)=(-12g3g2000-12g3g100012g32g1g2),H0H0=(2a2g2g30002a2g1g30002a2g1g2),

so that (2.8) reduces to the ODE system

{g˙1=g3g2+a2g2g3,g˙2=g3g1+a2g1g3,g˙3=-g32g1g2+a2g1g2,gi(0)=1,i=1,2,3.

By uniqueness, we thus have g1(t)=g2(t) for all t and we obtain

(6.1){g˙1=a2+g32g1g3,g˙3=a2-g32g12,g1(0)=g3(0)=1.

Special cases are given by

  1. a=0: the generalized Ricci flow reduces to the classical Ricci flow and an explicit solution to (6.1) is given by

    g1(t)=(1+3t)13,g3(t)=(1+3t)-13,

    defined on the maximal definition interval I=(-13,) (cf. [10]).

  2. a=±1: the system reduces to

    {g˙1=2g1,g˙3=0,g1(0)=g3(0)=1,

    with solution

    g1(t)=(1+4t)12,g3(t)=1

    for tI=(-14,).

A quick qualitative analysis of (6.1) shows that, for all a, the solution to (6.1) exists for all positive times, with

limtg1(t)=,limtg3(t)=|a|.

The maximal definition interval is always of the form Ia=(Tmin(a),), where Tmin:<0 is an even function, with Tmin(0)=-13 and monotonically converging to 0 as a goes to infinity (see Figure 1). We also have

limtTmin(a)+g1(t)=0,limtTmin(a)+g3(t)={,|a|<1,1,a=±1,0,|a|>1.

In Figure 2, we show some solutions of (6.1), sampled for a=k4, k=0,,9, and viewed as curves in the phase plane (g1,g3). The red and blue curves correspond to a=0 and a=1, respectively.

Figure 1 Behavior of the map Tmin{T_{\min}}.
Figure 1

Behavior of the map Tmin.

Figure 2 Examples of solutions to (6.1) viewed in the phase plane (g1,g3){(g_{1},g_{3})}.
Figure 2

Examples of solutions to (6.1) viewed in the phase plane (g1,g3).


Communicated by Jan Frahm


Funding statement: The author was supported by GNSAGA of INdAM.

Acknowledgements

This paper is an adaptation of the author’s master’s thesis, written under the supervision of Anna Fino. To her the author wishes to express his most sincere gratitude. The author also wishes to thank Mario Garcia-Fernandez for useful comments and Jeffrey Streets for pointing out reference [22]. He also thanks David Krusche for noting an imprecision in formula (2.5), and an anonymous referee for useful comments which helped improve the presentation of the paper.

References

[1] R. M. Arroyo, The Ricci flow in a class of solvmanifolds, Differential Geom. Appl. 31 (2013), no. 4, 472–485. 10.1016/j.difgeo.2013.04.002Search in Google Scholar

[2] R. M. Arroyo and R. A. Lafuente, The long-time behavior of the homogeneous pluriclosed flow, Proc. Lond. Math. Soc. (3) 119 (2019), no. 1, 266–289. 10.1112/plms.12228Search in Google Scholar

[3] V. Cortés and L. David, Generalized connections, spinors, and integrability of generalized structures on Courant algebroids, preprint (2019), https://arxiv.org/abs/1905.01977. 10.17323/1609-4514-2021-21-4-695-736Search in Google Scholar

[4] N. Enrietti, A. Fino and L. Vezzoni, The pluriclosed flow on nilmanifolds and tamed symplectic forms, J. Geom. Anal. 25 (2015), no. 2, 883–909. 10.1007/s12220-013-9449-ySearch in Google Scholar

[5] M. Garcia-Fernandez, Ricci flow, Killing spinors, and T-duality in generalized geometry, Adv. Math. 350 (2019), 1059–1108. 10.1016/j.aim.2019.04.038Search in Google Scholar

[6] M. Garcia-Fernandez and J. Streets, Generalized Ricci Flow, AMS University Lect. Ser. 76, American Mathematical Society, Providence, 2021. 10.1090/ulect/076Search in Google Scholar

[7] M. Gualtieri, Generalized complex geometry, preprint (2004), https://arxiv.org/abs/math/0401221; PhD Thesis, University of Oxford. 10.4007/annals.2011.174.1.3Search in Google Scholar

[8] R. S. Hamilton, Three-manifolds with positive Ricci curvature, J. Differential Geom. 17 (1982), no. 2, 255–306. 10.4310/jdg/1214436922Search in Google Scholar

[9] N. Hitchin, Generalized Calabi–Yau manifolds, Q. J. Math. 54 (2003), no. 3, 281–308. 10.1093/qmath/hag025Search in Google Scholar

[10] J. Isenberg and M. Jackson, Ricci flow of locally homogeneous geometries on closed manifolds, J. Differential Geom. 35 (1992), no. 3, 723–741. 10.4310/jdg/1214448265Search in Google Scholar

[11] M. Jablonski, Homogeneous Ricci solitons are algebraic, Geom. Topol. 18 (2014), no. 4, 2477–2486. 10.2140/gt.2014.18.2477Search in Google Scholar

[12] M. Jablonski, Homogeneous Ricci solitons, J. Reine Angew. Math. 699 (2015), 159–182. 10.1515/crelle-2013-0044Search in Google Scholar

[13] D. Krusche, private communication. Search in Google Scholar

[14] J. Lauret, Ricci soliton homogeneous nilmanifolds, Math. Ann. 319 (2001), no. 4, 715–733. 10.1007/PL00004456Search in Google Scholar

[15] J. Lauret, The Ricci flow for simply connected nilmanifolds, Comm. Anal. Geom. 19 (2011), no. 5, 831–854. 10.4310/CAG.2011.v19.n5.a1Search in Google Scholar

[16] J. Lauret, Curvature flows for almost-hermitian Lie groups, Trans. Amer. Math. Soc. 367 (2015), no. 10, 7453–7480. 10.1090/S0002-9947-2014-06476-3Search in Google Scholar

[17] J. Lauret, Geometric flows and their solitons on homogeneous spaces, Rend. Semin. Mat. Univ. Politec. Torino 74 (2016), no. 1, 55–93. Search in Google Scholar

[18] J. Lauret, Laplacian flow of homogeneous G2-structures and its solitons, Proc. Lond. Math. Soc. (3) 114 (2017), no. 3, 527–560. 10.1112/plms.12014Search in Google Scholar

[19] J. Lauret and E. A. Rodríguez Valencia, On the Chern–Ricci flow and its solitons for Lie groups, Math. Nachr. 288 (2015), no. 13, 1512–1526. 10.1002/mana.201300333Search in Google Scholar

[20] J. Polchinski, String Theory. Vol. I: An Introduction to the Bosonic String, Cambridge Monogr. Math. Phys., Cambridge University, Cambridge, 1998. Search in Google Scholar

[21] P. Ševera, Letters to Alan Weinstein about Courant algebroids, preprint (2017), https://arxiv.org/abs/1707.00265. Search in Google Scholar

[22] J. Streets, Generalized geometry, T-duality, and renormalization group flow, J. Geom. Phys. 114 (2017), 506–522. 10.1016/j.geomphys.2016.12.017Search in Google Scholar

[23] J. Streets and G. Tian, A parabolic flow of pluriclosed metrics, Int. Math. Res. Not. IMRN 2010 (2010), no. 16, 3101–3133. 10.1093/imrn/rnp237Search in Google Scholar

[24] J. Streets and G. Tian, Regularity results for pluriclosed flow, Geom. Topol. 17 (2013), no. 4, 2389–2429. 10.2140/gt.2013.17.2389Search in Google Scholar

[25] E. N. Wilson, Isometry groups on homogeneous nilmanifolds, Geom. Dedicata 12 (1982), no. 3, 337–346. 10.1007/BF00147318Search in Google Scholar

Received: 2020-06-26
Revised: 2021-05-11
Published Online: 2021-06-30
Published in Print: 2021-07-01

© 2021 Walter de Gruyter GmbH, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 28.10.2025 from https://www.degruyterbrill.com/document/doi/10.1515/forum-2020-0171/html
Scroll to top button