-invariant Laplacian flow
Abstract.
The Laplacian flow is a geometric flow introduced by Bryant as a way for finding torsion free -structures starting from a closed one. If the flow is invariant under a free action then it descends to a flow of -structures on a -manifold. In this article we derive expressions for these evolution equations. In our search for examples we discover the first inhomogeneous shrinking solitons, which are also gradient. We also show that any compact non-torsion free soliton admits no infinitesimal symmetry.
Key words and phrases:
-structure, -structure, symmetry, Laplacian flow, Soliton2010 Mathematics Subject Classification:
53C10, 53C291. Introduction
1.1. Overview.
A symplectic -structure on a -manifold is given by a pair , where is a symplectic form and is a complex -form with real and imaginary parts given by and respectively. Our main goal in this article is to derive the evolution equations for such an -structure on , together with a Higgs fields and a connection -form on an bundle , such that
is a solution to the Laplacian flow (see below for the definition). Put differently we want to understand how the -structure on the quotient of an -invariant solution to the Laplacian flow evolves. This investigation is further motivated from the results of Lotay-Wei in [18] that assert that if the initial -form on a compact manifold is -invariant then so is the flow for as long as it exists. We study two simple cases and as a result we construct the first examples of inhomogeneous shrinking Laplacian solitons. It is known that shrinking solitons do not occur on compact manifolds and that the steady ones are actually torsion free [18]. Indeed the examples that we construct are non-compact. There are currently no known compact examples of expanding solitons. We prove that such expanders, if they exist at all, in fact do not admit any infinitesimal symmetry i.e. there are no Killing vector fields which preserve the associated -form . This gives a partial explanation why such solitons have not (yet) been found.
1.2. Motivation.
A closed -structure on a -manifold is given by a closed non-degenerate -form , which in turn determines a Riemannian metric and volume form . In particular, also defines a Hodge star operator . If is also closed then the holonomy of is a subgroup of and consequently the metric is Ricci-flat. The Laplacian flow, defined as the initial value problem
(1.1) | ||||
(1.2) | ||||
(1.3) |
where , was introduced by Bryant in [5] as a way of potentially deforming within its cohomology class to a torsion free one. The short time existence of the flow was proved by Bryant and Xu in [6], and long time existence, uniqueness and compactness results were proved by Lotay and Wei in [18].
In recent years there have been many works on the Laplacian flow cf. [10, 11, 12, 16, 17, 19] but the flow is nonetheless still very complicated to study in its full generality. A natural strategy to simplifying the equations is to impose symmetry. Aside from the homogeneous cases, three notable works in this direction were carried out by Fine and Yao in [10], where the authors, motivated by a question of Donaldson, study the flow on as a way of deforming a hypersymplectic triple on to a hyperKähler one, by Lotay and Lambert in [16], where the authors reinterpret the flow on as a spacelike mean curvature flow on , and by Fino and Raffero in [11], where the authors study the flow for warped -structures on . In each case the presence of a free abelian group action allows for a reduction of the Laplacian flow to a lower dimensional space. It is exactly this common feature that motivates the work undertaken here.
1.3. Outline.
In this article we consider the more general case when only admits a free action. In this case the Laplacian flow descends to a flow on a symplectic -structure on , also called almost Kähler, together with the data of a Higgs field and a connection -form. This leads us to try and understand the resulting flow. In section 3 we prove a Gibbons-Hawking type construction for -invariant closed -structures (see Theorem 3.3) and we express the intrinsic torsion of the -structure in terms of that of the -structure. In other words we encode that data of an -invariant closed -structure only in terms of data on the quotient. In section 4 we then derive the evolution equations for such data when is evolving by the Laplacian flow. We also prove that certain cohomology classes have to vanish on a compact manifold with an exact -structure (see Theorem 4.15) and as a consequence we deduce that compact expanding solitons cannot arise from our construction. Even if is torsion free the -structure on the quotient is generally not torsion free, though it is always symplectic. For a flow of -structures a natural question one can ask is if any class of -structure is preserved. For instance in [11], the authors show that under certain constraints if is initially symplectic half-flat then this is preserved by the flow. In section 5 we first consider the original example of the Laplacian flow studied by Bryant and show that the symplectic form on the quotient, for a given action, in fact remains constant under the flow, though the intrinsic torsion of the quotient -structure is generic. Motivated by the work of Apostolov and Salamon in [1] where the authors construct (incomplete) examples of metrics which admit Kähler reduction, we next search for solutions to the Laplacian flow on these spaces. We show that the flow indeed preserves the Kähler property in this setting and we find explicit (complete) shrinking gradient solitons. These are the only currently known examples of inhomogeneous shrinkers.
Acknowledgements. The author is indebted to his PhD advisors Jason Lotay and Simon Salamon for their constant support and many helpful discussions that led to this article. The author would also like to thank Andrew Dancer and Lorenzo Foscolo for helpful comments on a version of this paper that figures in the author’s thesis. This work was supported by the Engineering and Physical Sciences Research Council [EP/L015234/1], The EPSRC Centre for Doctoral Training in Geometry and Number Theory (The London School of Geometry and Number Theory), University College London.
2. Preliminaries
The aim of this section is mainly to give a concise introduction to the basic objects of interest in this article and to set up notation.
2.1. -structures
Definition 2.1.
An -structure on a -manifold consists of an almost Hermitian structure determined by an almost complex structure , a non-degenerate -form and an almost Hermitian metric satisfying
(2.1) |
and a -form satisfying the two conditions
(2.2) | |||
(2.3) |
Note that condition (2.2) is just a consequence of the fact that is of type . Although an -structure consists of the data , one can in fact recover the whole -structure only from the pair , or . This observation is due to Hitchin in [13] where he shows that , or , determines . Abstractly this follows from the fact that the stabiliser of in is congruent to . The metric is then determined by (2.1) and , where is the Hodge star operator determined by and the volume form
(2.4) |
As modules the space of exterior differential forms splits as
(2.5) | ||||
where
(2.6) | ||||
(2.7) | ||||
(2.8) | ||||
(2.9) | ||||
and we get a corresponding decomposition for and via . The and notation refers to taking the corresponding real underlying vector space i.e.
cf. [7, 21]. As modules the spaces are all isomorphic. In computations we will often need to interchange between these spaces and to do so we use the following lemma which follows from a simple calculation.
Lemma 2.2.
Given a -form , let . Then the following hold:
-
(1)
-
(2)
-
(3)
We follow the convention that for a -form and vector , which might differ by a minus sign from other conventions in the literature. We also define the operator on a real function .
The intrinsic torsion of the -structure is determined by
(2.10) | ||||
(2.11) | ||||
(2.12) |
where are functions, , and , cf. [3]. Many well-known geometric structures can be recast using this formulation:
Definition 2.3.
The -structure is said to be
-
(1)
Calabi-Yau if all the torsion forms vanish,
-
(2)
almost-Kähler if ,
-
(3)
Kähler if all the torsion forms aside from vanish,
-
(4)
half-flat if .
The next notion that we will need is that of a -structure which we now define.
2.2. -structures
Definition 2.4.
A -structure on a -manifold is given by a -form that can be identified at each point with the standard one on :
where denote the coordinates on and is shorthand for .
The reason for this nomenclature stems from the fact that the subgroup of which stabilises is isomorphic to the Lie group . Since is a subgroup of [4, 21] it follows that determines (up to a choice of a constant factor) a Riemannian metric and volume form on . Explicitly we define these by
(2.13) |
where are vector fields on . In particular, defines a Hodge star operator The set of -forms defining a -structure is an open set in the space of sections of which we denote by . In this article we will only consider the situation when is closed. In this case we have
(2.14) |
for a unique -form where
(2.15) | ||||
The -form is called the intrinsic torsion of the closed -structure. On there is also a natural equivariant map given by
The kernel of is
and as modules we have
where is identified with the space of traceless symmetric -tensors via while . If is both closed and coclosed i.e. , then the holonomy of is contained in and . We refer the reader to the classical references [4, 21] for proofs of the aforementioned facts. The problem of constructing examples with holonomy equal to is very hard. The Laplacian flow was introduced as a means to tackle this problem.
2.3. The Laplacian flow
The Laplacian flow (1.1)-(1.3) preserves the closed condition i.e. for all and thus (1.1) is equivalent to
(2.16) |
In the compact setting, Hitchin gives the following interpretation of the flow. Consider the functional defined by
where denotes an open set in Then computing the Euler-Langrange equation Hitchin finds that the critical points of satisfy . The Laplacian flow is the gradient flow of with respect to an norm induced by on The Hessian of at a critical point is non-degenerate transverse to the action of the diffeomorphism group and in fact is negative definite. Thus can be interpreted as a Morse-Bott functional and the torsion free -structures correspond to the local maxima. In the non-compact setting this interpretation is not valid but the Laplacian flow is still well-defined and critical points are still torsion free -structures cf. [18].
In [5], Bryant computes the evolution equations for the following geometric quantities under the Laplacian flow
(2.17) | |||
(2.18) | |||
(2.19) |
We see immediately that up to lower order terms (2.18) coincides with the Ricci flow of . In [5, (4.30)] Bryant also derives an expression for the Ricci tensor in terms of the torsion form only and thus, one can express (2.18) only in terms of the torsion form as
(2.20) |
The simplest solutions to (1.1) are those that evolve by the symmetry of the flow. If satisfies
(2.21) |
for a vector field and constant then
where is the diffeomorphism group generated by , is a solution to flow and is called a Laplacian soliton [18]. Depending on whether is positive, zero or negative the soliton is called expanding, steady or shrinking respectively. If is a gradient vector field then we call them gradient solitons.
We refer the reader to [18] for foundational theory for the Laplacian flow.
Notations. In what follows we shall write for the space of differential forms, as well as their sections, omitting reference to the underlying space or unless there is any possible ambiguity. For a -form we will denote by its projection in . We shall also omit pullback signs and identify forms on with their pullbacks on .
3. reduction of closed -structure
Our aim in this section is to characterise -invariant closed -structures purely in terms of the data on the quotient space. In particular, we derive a Gibbons-Hawking Ansatz type construction for closed -structures, see Theorem 3.3.
Let be a closed -structure on which is invariant under a free action generated by a vector field . We define a connection -form on by
(3.1) |
where . The hypothesis that this circle action is free i.e. is nowhere vanishing ensures that is well-defined and moreover, by definition is also -invariant. The quotient space inherits an -structure given by
(3.2) | |||
(3.3) | |||
(3.4) |
The curvature form is invariant under and horizontal i.e.
(3.5) |
and as such it descends to . Since and is closed, we have that
(3.6) | |||
(3.7) |
From (2.2) and the fact that is closed it follows that
(3.8) |
We can also express the torsion form as
(3.9) |
for a -form and a -form which are both basic i.e. they are horizontal and -invariant (since ). Thus, and are really pullback of forms on . Our goal is to encode the intrinsic torsion of the -invariant closed -structure (and its derivatives) only in terms of the data .
As it follows from (2.15) that
(3.10) | |||
(3.11) |
From (2.5), we note that (3.8) and (3.10) imply that and have no -component i.e.
(3.12) |
The latter is not surprising since . In terms of the -structure we can express the condition as
(3.13) | |||
(3.14) |
where we recall Observe that the forms , and are related by (3.11), (3.14) and the fact that appears in both (2.11) and (2.12). Thus, these forms are all essentially equivalent, to be more precise we have:
Lemma 3.1.
The torsion forms and are determined by the curvature component via
where the -form is defined by .
Proof.
Let for a -form (note that the existence of -forms and follows from (2.7)), then using Lemma 2.2 we can express the torsion forms from (3.7) and (3.13) in irreducible summands as
(3.15) | |||
(3.16) |
and hence comparing the torsion -form from (2.11) and (2.12) we have that
From (3.11) and using of Lemma 2.2 we find that
Another application of Lemma 2.2 now shows that
i.e. and this completes the proof. ∎
In particular Lemma 3.1 asserts that the torsion form completely determines . We can now express the torsion forms as:
(3.17) | |||
(3.18) |
Since , it follows, using Lemma 2.2, that
(3.19) |
where denotes the codifferential on and hence the torsion form can be expressed as:
(3.20) |
In particular we have which is equivalent to
(3.21) |
i.e. . One can also extract the -component of from the next Proposition.
Proposition 3.2.
The following holds
(3.22) |
or equivalently . In particular, if is compact then
Proof.
Having now expressed only in terms of the data , we need to show that we can recover from , which is achieved by the next Theorem.
Theorem 3.3 (Gibbons-Hawking Ansatz for closed -structures).
Given a symplectic manifold admitting an -structure and a positive function satisfying
(3.23) |
then defines a closed -structure on the total space of the bundle determined by (3.23) and the curvature of the connection form is given by
(3.24) |
Proof.
Remark 3.4.
Let be an open set of endowed with the Euclidean metric and volume form, and hence Hodge star operator . The Gibbons-Hawking Ansatz states that given a positive harmonic function such that
(3.25) |
then
defines a hyperKähler metric on the total space of the bundle determined by (3.25), where is a connection -form satisfying In Theorem 3.3 condition (3.23) is the higher dimensional analogue of the ‘integrality’ condition (3.25) that figures in the Gibbons-Hawking Ansatz and the (linear) harmonic condition on is replaced by the (non-linear) condition
(3.26) |
on the pair . So the data is sufficient to recover the -structure since the curvature of is already determined by (3.24).
We can also characterise the torsion free -structures in terms of the data on the base by the following Proposition.
Proposition 3.5.
Assuming we are in the situation of Theorem 3.3, then
If this holds, then is a Calabi-Yau -fold if and only if is constant.
Proof.
The first part follows from the equivalence between the conditions and and the fact that implies (from Lemma 3.1) that . If furthermore is constant then from Lemma 3.1 we have that i.e. . Differentiating the relation
from (2.9) and using (3.7) then shows that i.e. . To complete the proof we need to show that if is Calabi-Yau then is constant which follows immediately from
∎
Having encoded the data of a closed -invariant -structure in terms of the data on the quotient we derive the evolution equations for the data under the Laplacian flow in the next section.
4. The -invariant Laplacian flow
4.1. -invariant flow equations
Consider the Laplacian flow starting from an -invariant closed -structure. Then by the existence and uniqueness of the flow, at least in the compact case, it follows that this symmetry persists i.e the solution to the flow can be expressed as (3.2) for as long as it exists cf. [18, Corollary ]. Note also that since and it follows that is a basic -form i.e. it corresponds to a -form on . Thus, in the -invariant setting, from (3.20) we see that (2.16) becomes equivalent to the following evolution equations on :
(4.1) | |||
(4.2) |
We omit writing the dependence on to ease the notation.
Remark 4.1.
Observe that (4.1) agrees with the fact that since remains in its cohomology class so does .
The main result of this section can be summed up as follows:
Theorem 4.2.
Given a compact symplectic -structure together with the data of an bundle with connection -form and positive function as in Theorem 3.3, then the coupled flow defined by
(4.3) | |||
(4.4) | |||
(4.5) | |||
(4.6) |
where and with defined by , admits short time existence and uniqueness for the initial data . Moreover, stays symplectic and holds for as long as the flow exists.
Proof.
Note that in view of Theorem 3.3 we already know that the evolution equation for is a consequence of (4.3)-(4.5). This can also be seen by inspecting expression (4.6). Before deriving the evolution equations for the data on we first give expressions for quantities that will appear in the evolution equations.
Lemma 4.3.
The norms of the torsion forms can be expressed as
-
(1)
-
(2)
-
(3)
-
(4)
Proof.
The proof is a direct calculation using the expressions from the previous section. We prove () as an example:
where the first equality follows from () of Lemma 2.2 and the definition of . The second equality is again just by the definition of . The proofs for the rest follow by similar computations. ∎
Proposition 4.4.
The evolution equation for the connection form is given by
(4.7) |
Proof.
Proposition 4.5.
(4.9) | ||||
(4.10) |
Proof.
This follows directly from (4.1) and the fact that ∎
Proposition 4.6.
(4.11) | ||||
We can also expressed the above more compactly as
(4.12) |
Proof.
Since we can use (2.20) to write down its evolution equation:
(4.13) |
From Lemma 4.3 we can express in terms of the torsion of the -structure as
(4.14) |
Thus, we only need to simplify the term
(4.15) |
which is straightforward to do, except for the term involving . Using Lemma 2.2 and (3.15) we compute
We also have that
and
Using the last three expressions and the fact that , one can rewrite (4.15) in terms of data on . Substituting all this in (4.13) completes the proof for the first equation. Rather than unwinding the various relations between the torsion forms given in the previous section, one can prove the second expression more directly using (2.19) together with the fact that and the evolution equation for . ∎
Remark 4.7.
Observe that even if is initially constant, i.e. the orbits have constant size, this is not generally preserved in time. Indeed from (4.11) we see that if then , so the size of the orbit is expected to shrink initially.
Proposition 4.8.
The evolution equation for is given by
(4.16) | ||||
Proof.
Proposition 4.9.
The evolution equation for is given by
(4.17) |
Proof.
Proposition 4.10.
(4.18) | ||||
Proof.
The idea is to again use the evolution equation for . Since only evolves on the base we can ignore terms involving in (2.20). Thus, we have that
(4.19) |
As modules we have the following decomposition
By abuse of notation we are identifying the cotangent spaces of and with and in the above. It follows that the only terms in that contribute to the evolution of belong to the last summands. Since we have that the only terms that can arise in the evolution of are the components of which we write as
(4.20) |
and the components of which we can write as
(4.21) |
A direct computation (in a coframe) shows that
Since the orthonormal symmetric tensors and span the rank module it follows that as modules we have
We now compute
Substituting the latter and (4.12) in (4.19) gives the result. ∎
The reader might find the presence of the map in (4.18) rather unusual as the latter is strictly speaking a -equivariant map but one can replace it by the corresponding -equivariant map
defined in [3, Sect. 2.3]. To conclude this section we derive the evolution equations for certain types of differential forms on .
Lemma 4.11.
-
(1)
Let then
-
(2)
Let then
-
(3)
Let then
-
(4)
Let then
Proof.
To prove (1) we simply differentiate the relation and use (4.1). The proofs for the rest are completely analogous. For (2) we differentiate , for (3) we differentiate and for (4) we differentiate . ∎
These expressions can be quite useful for extracting the evolution equations for specific components of a given quantity. For instance we can apply (1) and (2) to to find the evolution equation for the component of in and respectively. From (2.17) one can also compute the evolution equation for :
(4.22) |
From this one can deduce the evolution equations for and . The resulting expressions are rather involved so we won’t write them down here.
Remark 4.12.
The evolution equations derived in this section generalise those derived in [11] in the special case that is a warped product. Note however that their choice of -structure differs from ours by a conformal factor so that . In particular, is not symplectic but on the other hand with respect to , instead of , equation (4.11) becomes parabolic. Since the induced flow on the data is still generally quite complicated we shall only study it in a couple of simple examples in the last section, which exclude their case i.e. not warped products.
4.2. The -invariant soliton equation
Having derived the general -invariant flow equations the natural next step is to work out the soliton equation. But before doing so we first prove a non-existence result in the compact case. By compact we always mean without boundary.
4.2.1. Non-existence of compact solitons with continuous symmetry
Proposition 4.13.
Let be an exact -structure on a compact -manifold so that for . Then admits no non-trivial infinitesimal symmetry i.e.
Proof.
Suppose that then
The first equality is from Stokes’ Theorem, the second follows from the fact that and the last is a consequence of (2.13). Hence is identically zero i.e. . ∎
Corollary 4.14.
A non-torsion free Laplacian soliton on a compact manifold admits no non-trivial infinitesimal symmetry.
Proof.
Recall that the soliton equation is
for some , or equivalently
On a compact manifold, from [18, Proposition 9.5] we know that with equality if and only if is torsion free. So we only need to consider the case when and the result follows from the previous Proposition. ∎
This makes the construction of expanding solitons on compact manifolds quite a hard problem as one cannot use continuous symmetries to simplify the PDEs, so this suggests that one might have to use hard analysis to find examples (if any exist at all).
More generally, let be a closed -form on compact and . Then
(4.23) |
Observe that only the cohomology class of is relevant here. In Proposition 4.13 we used the fact that if then we can set . Note also that for any closed -structure (not necessarily exact) we have the following one-to-one correspondences:
Here we are using the fact that for , we have that and hence any closed is also coclosed. Likewise the analogous argument applies for closed since .
Although for torsion free -structures it is known that
where denotes the space of harmonic -forms [14], this is not generally true for strictly closed -structures. One way of seeing this is suppose is a harmonic -form on then we have
Thus, there could exist harmonic -forms not in In any case, we know that these spaces are finite dimensional vector spaces since from Hodge theory is the second Betti number of .
So (4.23) says that in fact there does not exist any closed -form strictly of type or on and hence we deduce the following:
Theorem 4.15.
If a compact manifold admits an exact -structure then has no non-trivial infinitesimal symmetry and it also does not have any closed (equivalently harmonic) -form of pure type i.e. .
In particular, this Theorem applies to compact expanding solitons. Another immediate consequence is that any bundle on cannot admit a connection whose curvature -form belongs to or i.e. it cannot be a instanton or anti-instanton respectively. Note that it is still an important open problem whether a compact manifold can even admit an exact -structure.
Remark 4.16.
In [20] Podestà and Raffero used similar arguments to show that for closed -structures on compact manifolds the Lie algebra of infinitesimal symmetry is in fact abelian and of dimension at most . They also exhibit an example showing that this bound is sharp. Theorem 4.15 shows that the exact case is very different from the closed one.
4.2.2. The -invariant soliton equation
We now derive the equations for -invariant Laplacian solitons in terms of the data on . Note that in view of Corollary 4.14 this only applies to the non-compact case. We now consider the soliton equation
with i.e. we assume that is a horizontal vector field on . Then under the free action generated by the vector field as before, this reduces to the pair
(4.24) | |||
(4.25) |
with and as defined in section 3. Observe that if then the symplectic form is necessarily exact (which implies that is non-compact and hence as we already noted), as it is for . If is the gradient vector field for some function on then we can rewrite the soliton equation using Lemma 3.1 and (3.19) as
(4.26) | ||||
(4.27) | ||||
We shall not attempt to solve this system here but we will give an example of a solution in the next section.
5. Examples of -invariant Laplacian flow
5.1. The Bryant-Fernández example
The compact nilmanifold associated to the -step nilpotent Lie algebra111Here we are using Salamon’s notation [22] to mean that the Lie algebra admits a coframing with , where denotes the th entry.
admits a closed -structure given by
This example was discovered by Fernández in [8] and Bryant worked out the Laplacian flow on this example in [5]. The solution to the Laplacian flow is given by
where cf. [9]. This solution is immortal and the volume grows as in time. Bryant also showed that cannot admit a torsion free -structure for topological reasons and hence one cannot expect the flow to converge. Nonetheless we have that converges to zero as and [9, Theorem 4.2] shows that converges in suitable sense to a flat metric.
Following the construction described in section 3 we choose the vector field generating an action preserving to be so that the connection form The solution to the induced flow on the quotient nilmanifold is then given by
We see that the symplectic form, and hence the volume, stays constant while the metric (equivalently the complex structure) degenerates as . As in [9] one can verify that the curvature decays to zero as Note that neither nor is zero in this case, so this example can be viewed as a generic case with regards to the type of the -structure.
5.2. New examples from the Apostolov-Salamon Ansatz
As a flow on -structures one can ask if the flow preserves any interesting geometric quantity. For instance we already saw that since stays closed under the flow stays symplectic. A natural question to ask is: if the almost complex structure is initially integrable, does this persists under the flow?
If is torsion free and is Kähler then Apostolov and Salamon proved that is in fact a bundle over a -manifold , which in special cases turns out to be hyperKähler cf. [1, Theorem 1]. This motivates us to search for solutions to the flow preserving the Kähler condition on these spaces.
Consider the manifold where is a compact nilmanifold associated to the Lie algebra
The -structure determined by
(5.1) |
defines a coframing on given by , , , , , and , where are (nowhere vanishing) functions of only and denote the standard self-dual -forms in i.e.
Lemma 5.1.
The intrinsic torsion of -structure defined by (5.1) is given by
-
(1)
if and only if
-
(2)
if and only if and
Proof.
A direct calculation shows
(5.2) |
and
(5.3) |
∎
The explicit torsion free -structure given by setting , and corresponds to Example 1 in [1].
Let us now impose that is closed, so that is determined by condition (1) of Lemma 5.1, and consider the action generated by the vector field . Then applying the construction of section 3 one can compute that
Since , from (3.15) and (3.16) we see that the only non-zero component of the torsion is and hence it follows that these closed -structures all admit Kähler reductions (see Definition 2.3).
Lemma 5.2.
The torsion form of closed is given by
We now search for solutions to the Laplacian flow (2.16) of the form (5.1), where only the functions depend on . Computing the Laplacian flow gives the system
(5.4) | |||
(5.5) | |||
(5.6) |
Equation (5.6) is a consequence of (5.5) by differentiating with respect to and using the closed condition from Lemma 5.1. Thus, the Laplacian flow is reduced to the pair
(5.7) | |||
(5.8) |
Since we are on a non-compact manifold the existence and uniqueness of a solution, given initial data, to (5.7) and (5.8) is not always guaranteed. We do however know that there exists at least one solution namely the (incomplete) torsion free one of Apostolov-Salamon [1].
Rather than addressing the general existence problem, we shall instead find another explicit solution as follows.
A shrinking gradient soliton.
With and we have
Taking and , one verifies directly that the soliton equation (2.21) is satisfied. Thus, it defines a gradient shrinking soliton with the induced metric
which is clearly complete. To the best of our knowledge this is the first example of an inhomogeneous shrinker.
To derive the general soliton equation we first observe that an invariant vector field is of the form , for functions , and . Comparing with the expressions for and it is easy to see that we get a consistent system only if . By reparametrising the -coordinate we can set , and defining and the closed condition becomes equivalent to . We compute the soliton equation for the unknowns as
(5.9) | |||
(5.10) |
With the ansatz , we find the solution and . The scalar curvature is
Observe that this construction applies to any hyperKähler -manifold such that . In this case can be taken to be the total space of the bundle determined by these integral cohomology classes and are the connection -forms with curvature respectively [1].
Remark 5.3.
We should also point out that recently Ball found the first examples of inhomogeneous steady solitons on the same spaces [2, Example 2], also arising from what we referred to as the Apostolov-Salamon Ansatz.
References
- [1] Vestislav Apostolov and Simon Salamon. Kähler reduction of metrics with holonomy . Communications in Mathematical Physics, 246(1):43–61, 2004.
- [2] Gavin Ball. Quadratic closed -structures. arXiv:2006.14155, 2020.
- [3] Lucio Bedulli and Luigi Vezzoni. The Ricci tensor of SU(3)-manifolds. Journal of Geometry and Physics, 57(4):1125–1146, 2007.
- [4] Robert L. Bryant. Metrics with exceptional holonomy. Annals of Mathematics, 126(3):525–576, 1987.
- [5] Robert L. Bryant. Some remarks on -structures. In Proceedings of Gökova Geometry-Topology Conference 2005. International Press, 2006.
- [6] Robert L. Bryant and Feng Xu. Laplacian flow for closed -structures: Short time behavior. arXiv:1101.2004, 2011.
- [7] Simon Chiossi and Simon Salamon. The intrinsic torsion of and structures. In Differential Geometry, Valencia 2001, pages 115–133. World Scientific, 2002.
- [8] Marisa Fernández. An example of a compact calibrated manifold associated with the exceptional Lie group . Journal of Differential Geometry, 26(2):367–370, 1987.
- [9] Marisa Fernández, Anna Fino, and Víctor Manero. Laplacian flow of closed -structures inducing nilsolitons. The Journal of Geometric Analysis, 26(3):1808–1837, 2016.
- [10] Joel Fine and Chengjian Yao. Hypersymplectic 4-manifolds, the -Laplacian flow, and extension assuming bounded scalar curvature. Duke Mathematical Journal, 167(18):3533–3589, 2018.
- [11] Anna Fino and Alberto Raffero. Closed warped -structures evolving under the Laplacian flow. Annali Scuola Normale Superiore - Classe Di Scienze, 2017.
- [12] Anna Fino and Alberto Raffero. A class of eternal solutions to the Laplacian flow. The Journal of Geometric Analysis, 2020.
- [13] Nigel Hitchin. The geometry of three-forms in six and seven dimensions. Journal Differential Geometry, 55(3):547–576, 2000.
- [14] Dominic Joyce. Riemannian holonomy groups and calibrated geometry, volume 12. Oxford University Press, 2007.
- [15] Shoshichi Kobayashi. Principal fibre bundles with the -dimensional toroidal group. Tohoku Mathematical Journal, 8(1):29–45, 1956.
- [16] Ben Lambert and Jason D. Lotay. Spacelike mean curvature flow. The Journal of Geometric Analysis, 2019.
- [17] Jorge Lauret. Laplacian flow of homogeneous -structures and its solitons. Proceedings of the London Mathematical Society, 114(3):527–560, 2017.
- [18] Jason D. Lotay and Yong Wei. Laplacian flow for closed structures: Shi-type estimates, uniqueness and compactness. Geometric and Functional Analysis, 27(1):165–233, 2017.
- [19] Marina Nicolini. New Examples Of Shrinking Laplacian Solitons. The Quarterly Journal of Mathematics, 2021. haab029.
- [20] Fabio Podestà and Alberto Raffero. On the automorphism group of a closed -structure. The Quarterly Journal of Mathematics, 70(1):195–200, 2018.
- [21] Simon Salamon. Riemannian geometry and holonomy groups. Longman Scientific and Technical, 1989.
- [22] Simon Salamon. Complex structures on nilpotent Lie algebras. Journal of Pure and Applied Algebra, 157(2):311–333, 2001.