- Research Article
- Open Access
- Published:
First-Order Twistor Lifts
Journal of Inequalities and Applications volume 2010, Article number: 594843 (2010)
Abstract
The use of twistor methods in the study of Jacobi fields has proved quite fruitful, leading to a series of results. L. Lemaire and J. C. Wood proved several properties of Jacobi fields along harmonic maps from the two-sphere to the complex projective plane and to the three- and four-dimensional spheres, by carefully relating the infinitesimal deformations of the harmonic maps to those of the holomorphic data describing them. In order to advance this programme, we prove a series of relations between infinitesimal properties of the map and those of its twistor lift. Namely, we prove that isotropy and harmonicity to first order of the map correspond to holomorphicity to first order of its lift into the twistor space, relatively to the standard almost complex structures and
. This is done by obtaining first-order analogues of classical twistorial constructions.
1. Introduction
Harmonic maps are mappings between Riemannian manifolds
and
which extremize the energy functional

Letting denote the tangent bundle of
, one can (locally) characterize harmonic maps as solutions of the nonlinear second-order differential equation

where denotes the tension field of
,

A bibliography can be found in [1] and for some useful summaries on this topic, see [2, 3].
The infinitesimal deformations of a harmonic map are called Jacobi fields. More precisely, let be a smooth map and denote by
the set of smooth sections of the pull-back bundle
. If
is harmonic and
is a vector field along it,
is said to be a Jacobi field (along
) if it satisfies the linear Jacobi equation
, where the Jacobi operator
is defined by

Here, is the Laplacian on
,

and, letting denote the curvature tensor of
,

Jacobi fields are characterized as lying in the kernel of the second variation of the energy functional. Indeed, if is a two-parameter variation of a harmonic map
, then, writing
and
, the Hessian
of
is the bilinear operator on
given by

so that a Jacobi field (along
) is characterized by the condition

A Jacobi field is called integrable if it is tangent to a deformation through harmonic maps. In [4, 5], the question of whether all Jacobi fields are integrable is answered for the case where the domain is the two-sphere and the codomain the two-dimensional complex projective space or the three- and four-dimensional sphere. This was done by relating the deformations of the map associated with the Jacobi field and those of the twistor lift of the map. More precisely, given an oriented even-dimensional manifold , we can construct its (positive) twistor space
. This manifold admits two natural almost complex structures
and
. Given a map
from a Riemann surface
, harmonicity is intimately related with the existence of a
-holomorphic lift
, whereas isotropy is related with the existence of a
-holomorphic lift
(see [6]). On the other hand, Jacobi vector fields induce families of maps which are harmonic to first-order and, in some cases, isotropic to first order. The translation of these first order properties in terms of twistor lifts plays an important role on the study of the Jacobi fields and we shall exhibit how this translation can be established in general.
This work is divided as follows: in the next two sections, we recall some well-known results concerning twistor lifts of harmonic and isotropic maps. In Section 4, we show how this constructions generalize to their parametric versions and examine more closely the construction when the codomain is a four-dimensional manifold. We leave to the last section some technical proofs.
2. The Setup
2.1. Twistor Spaces
Let be an oriented even-dimensional Euclidean space, equipped with a metric
. A Hermitian structure
on
is
with
and
for all
. We say that
is positive if there is a positive basis of
of the form
and negative otherwise. The set of all positive Hermitian structures (resp., all negative Hermitian structures) on
is denoted by
(resp.,
). The Lie group
acts transitively on
by the formula
and the isotropy subgroup at
is given by

Letting and
, we easily conclude that

In particular, the tangent space of at
is given by
. In this vector space, we can introduce a complex structure
defining

When equipped with , the manifold
is a complex manifold.
Given with a Hermitian structure
, we can consider on
its
and
-parts given as usual by

These are isotropic subspaces, in the sense that for all
in
(or in
). Associating an Hermitian structure
on
with its
-space
gives a 1–1 correspondence between Hermitian structures and maximal isotropic subspaces. We say that a maximal isotropic subspace is positive if the corresponding orthogonal complex structure is positive and we denote the set of all such subspaces by
.
Let be an oriented even-dimensional vector bundle over a manifold
equipped with a connection
and a parallel metric
. Then, we may take the bundle

whose fibre at is precisely
. If
is Riemann surface, the vector bundle
has vanishing
-part of its curvature tensor and therefore admits a unique structure as a holomorphic bundle over
, by a well-known theorem of Koszul and Malgrange [7]. This induces a holomorphic structure
on the bundle
for which a section
of
is holomorphic if and only if [6] (see [8]).

Let be an oriented Riemannian manifold with dimension
. We call (positive) twistor space of
the bundle whose fibre at
is precisely
; that is,

The Riemannian connection on induces a splitting of the tangent space to
into vertical and horizontal parts,
. Namely, if
denotes the canonical projection defined by
, then

With respect to this decomposition, maps
isomorphically into
and allows to define an almost complex structure
on
as the pull-back of
on
. Together with (2.3), this allows to define two almost complex structures
and
on
by the formulae

When equipped with ,
is never a complex manifold; as for
, it is integrable if and only if
is conformally flat (
) or anti-self-dual (
) (for more details, see [6, 9, 10]; a discussion on this topic can also be found in [11] and references therein).
Notice that is a holomorphic map for any of these complex structures; that is, for
or
, one has that

Definition 2.1.
Let and
be two almost complex manifolds. Let
be a decomposition of
into
-stable subbundles; that is,
and
. We shall call such a decomposition a
-stable decomposition. Let
be a smooth map. We shall say that
is
-holomorphic if

Analogously we define -holomorphic maps.
A smooth map is holomorphic if and only if it is both
and
-holomorphic for some, and so any, stable decomposition
. Taking
, the decomposition
is clearly stable for both the almost complex structures
and
on
.
Remark 2.2.
We can easily introduce a metric on the twistor space : let
and consider the tangent space at this point,
. We know that we have the identifications
and
. To get a metric on
, transport the metric from that on
; that is,
, for all
,
, where
denotes the metric on
at
. For the vertical space
, we can consider the restriction of the metric on the space
. Finally, we declare
and
to be orthogonal under the metric
; that is,
, for all
,
. With this metric, the decomposition
is orthogonal and
-stable (
),
(
) are almost Hermitian manifolds and the projection map
is a Riemannian submersion.
2.2. Conformal and Isotropic Maps
Given a smooth map ,
is said to be weakly conformal at
if there is
with

If , then
is said to be a regular point (of
) and the map
is called conformal at
. Moreover, a map which is conformal (resp., weakly conformal) at all points
is said to be a conformal map (resp., a weakly conformal map).
If is an almost Hermitian manifold, any holomorphic (resp., antiholomorphic) map
is (weakly) conformal as it maps
to
(resp., to
). A stronger property than conformality is isotropy: if
is a smooth map from a Riemann surface,
is isotropic if [12]

where . Actually, the condition (2.13) can be weakened to

To check this, establish an induction on : if
the result is trivial. Assuming now that (2.14) implies (2.13) for all
and taking
with
, we may assume without loss of generality that
,
and we get

Since and
, both terms in the above expression vanish.
Moreover, letting in (2.13), it is easy to check that an isotropic map from a Riemann surface is a (weakly) conformal map.
Let be a smooth map from a Riemann surface
. We shall say that
is an umbilic point (of
) if
is a
-linearly dependent set. If
is such that all points
are umbilic, we shall say that
is totally umbilic (see [6]).
3. Nonparametric Twistorial Constructions
The following are well-known twistorial constructions [13] (see also [6]).
Theorem 3.1.
If is holomorphic, the projection map
is isotropic. Conversely, if
is a conformal totally umbilic immersion, there is (locally) a holomorphic map
such that
.
If is holomorphic, the projection map
is harmonic. Conversely, if
is a conformal harmonic map, there is (locally) a holomorphic map
such that
.
We shall sketch the proof of this result. We start by noticing that given a smooth map obtained as the projection of
,
, without requiring further conditions à priori on
, nothing guarantees that
is holomorphic relatively to the induced almost Hermitian structure
on
; if it is, we shall say that the structure
is strictly compatible with
(or that the map
is a strictly compatible twistor lift of
). Such a structure
can exist if and only if
is isotropic: in other words, if and only if
is (weakly) conformal. If
preserves
but does not necessarily render
holomorphic, we shall say that
(or the map
) is compatible with
.
If is given as the projection
of an
-holomorphic map
(
or
), then
is holomorphic with respect to the induced almost Hermitian structure
on
:

In particular, is (weakly) conformal. Moreover, the above equivalence shows that any strictly compatible lift of
is
-holomorphic. On the opposite direction, let
be a conformal map. Let
denote the normal bundle of
in
. We may decompose the connection on
into its tangent and normal parts,
. Hence, on
we have a metric
and connection
inherited from those on
. Moreover, we may take the bundle
over
which has the
-holomorphic structure. Since
is conformal, we may transfer the Hermitian structure
of
into
. Hence, we have a natural map

Taking any holomorphic section of and composing with
, we obtain a strictly compatible twistor lift
of
. Since any such lift is
-holomorphic, we have just proved the following result.
Proposition 3.2.
Given ,
is conformal if and only if
is (locally) the projection of an
-holomorphic map
.
To proceed, we need the following result [13].
Proposition 3.3.
Let and let
. Take
the section of
corresponding to
. Then, the map
is
-holomorphic if and only if
is holomorphic with respect to
and

The map is
-holomorphic if and only if
is holomorphic with respect to
and

Let be a
-holomorphic map and let
be the corresponding section of
. If
is the projection map, we have that
is
-holomorphic. In particular,
. Since
is
-holomorphic, we can write

and, inductively, it follows that for all
so that
is isotropic. Conversely, let
be a conformal totally umbilic map. In this case, we consider the manifold
equipped with the opposite holomorphic structure-
and again construct the holomorphic bundle
. For this structure, a section
is holomorphic if and only if
. Since
is conformal, we may define the map
as in (3.2). On the other hand, because
is totally umbilic, we know that
lies in the span of
; in other words, if
is any almost Hermitian structure on
strictly compatible with
,
lies in
. We may therefore conclude that
is
-holomorphic. As a matter of fact, given a holomorphic section
of
and considering
, we have that
. We may write

since and
lie in
. In particular,
and so

and therefore we conclude that is
-holomorphic.
With a similar argument, let be a
-holomorphic map and let
be the section of
corresponding to
. Then, taking the projection map
, we get

Since is
-holomorphic,
lies in
. From (iv) in Proposition 3.3, we conclude that
lies in
and therefore has vanishing
-part. From the reality of
, we have that
and so
is harmonic. Conversely, if
is conformal and harmonic, we may take the bundle
and the map
in (3.2). Since
is harmonic,
is holomorphic with respect to the
and
structures [6, 8]: letting
denote a holomorphic section of
and
the composition
, then,
and

Since is harmonic,
. On the other hand, since
is holomorphic,
. Finally,

shows that and so that
is holomorphic. Therefore, taking any holomorphic section of
and composing with
give a
-holomorphic lift of
.
Remark 3.4.
Notice that to guarantee the existence of the -holomorphic lift for
, the important fact was that
belongs to the
-part of
for any almost Hermitian structure strictly compatible with
. This is guaranteed if
is a totally umbilic map, but it is not strictly necessary. For instance, if
is an isotropic map, the vectors
,
span an isotropic subspace. If this vectors are linearly independent, taking this space as the
-space of
, then
is a
-holomorphic lift of
, although
may be a map into
; on the other hand, if
is totally umbilic, then we may take
either as the unique strictly compatible map into
or into
and both these maps are
-holomorphic.
4. First-Order Twistorial Constructions
4.1. Harmonicity and Isotropy to First Order
Let denote an interval of the real line containing
. Given a (family of) map(s)
,
, we say that
is harmonic to first order if

where and
.
Let be a harmonic map,
a vector field along
, and
a one-parameter variation of
. We say that
is tangent to
if
. The following result is a key ingredient in what follows [4]:
Proposition 4.1.
Let be a harmonic map between compact manifolds
and
. Let
be a vector field along
and
a one-parameter variation of
tangent to
. Then,

In particular, is Jacobi if and only if any tangent one-parameter variation is harmonic to first order.
We have seen in Theorem 3.1 that harmonicity was not enough to establish a relation with possible twistor lifts of a map conformality and was also a key ingredient, as maps obtained as projections of twistorial maps must be holomorphic with respect to some almost Hermitian structure along the map. On the other hand, when the domain is the -sphere, harmonicity implies (weak) conformality or even isotropy, the last case occurring if the target manifold is itself also a sphere or the complex projective space [12, 14].
Let be a Riemann surface and
a smooth map. The map
is said to be conformal to first order if [15]

Analogously, is said to be isotropic to first order,

As for the nonparametric case, one can prove [16] that condition (4.4) can be weakened to the following

As in the nonparametric case, harmonicity to first order implies conformality to first order for maps defined on the two-sphere and even isotropy when the codomain is itself a real or complex space form [15].
4.2. Twistorial Constructions
As we have seen, Jacobi fields induce variations that are harmonic (and, in some cases, conformal or even isotropic) to first order. On the other hand, in Section 2 we have seen that conformality, harmonicity, and isotropy of the map correspond to
,
and
-holomorphicity of the twistor lift
. As we shall see, these results do have a first-order version as follows. We start with a definition.
Definition 4.2.
Let be an almost complex manifold and
an almost Hermitian manifold. Given a smooth map
, we say that
is holomorphic to first order if
is holomorphic and

where is the Levi-Civita connection on
induced by the metric
. Moreover, if
is a
-stable decomposition of
, orthogonal with respect to
, we shall say that
is
-holomorphic to first order if
is
-holomorphic and

where is the restriction of
to
. Changing
to
gives the definition of
-holomorphicity to first order.
In contrast with the nonparametric case, it is not obvious that -holomorphicity to first order implies
-holomorphicity to first order. As a matter of fact, from (4.6), it only follows that
. However, we do have the following.
Lemma 4.3.
Let be a smooth map and let
be a
-stable decomposition of
, orthogonal with respect to
. Then,
is holomorphic to first order if and only if
is both
and
-holomorphic to first order.
Proof.
Assume that is holomorphic to first order. Then,
is
-holomorphic. As for (4.7), letting
denote an arbitrary section of
and
its projection into
, we have

which is true since is
-holomorphic to first order. Hence,
is
-holomorphic to first order. Changing
to
shows that
is
-holomorphic to first order. The converse also follows using analogous arguments.
Remark 4.4.
The importance of choosing the Levi-Civita connection on is illusory. In particular, we can define the concept of holomorphicity to first order (or
,
-holomorphicity to first order) for maps defined between almost complex manifolds, not necessarily equipped with any metric.
Indeed, let be a holomorphic to first-order map with respect to
, so that (4.6) holds. Let
denote a (local) frame for
. Then,

Since is holomorphic to first order,
is holomorphic, the above equation is equivalent to

Now, since holomorphicity does not depend on the chosen connection, we can deduce that holomorphicity with respect to
reduces to the same condition (4.10). Thus,
being holomorphic to first order does not depend on the chosen connection. For
(resp.,
) holomorphicity to first order, we use similar arguments, replacing
for a horizontal (resp., vertical) frame.
4.3. The
-Holomorphic Case
In the nonparametric case, given a conformal map , we can always find a lift
such that
is holomorphic with respect to
. In other words, (locally defined) strictly compatible lifts always exist. In general, this lift may not be
or
-holomorphic but it is
-holomorphic. Let
be a variation of
, conformal to first order. Then, if a lift
to the twistor space that makes
holomorphic for all small
exists,
is necessarily conformal for all small
, which may not be the case. So, even if conformality is preserved to first order, there might be no strictly compatible twistor lift for all
; hence, we should relax the condition on conformality. We shall say that a twistor lift
of a conformal to first order map
is compatible to first order (with
) if

We start with a technical lemma, whose proof the reader can find in Section 5.
Lemma 4.5.
Let be a conformal to first-order map. Let
be a twistor lift compatible to first order with
. Then for all
there is a function
and a vector field
with
,
and
,
such that

In particular, is
-holomorphic to first order in the sense that

Lemma 4.6.
Let be
-holomorphic to first order. Then,
is
-holomorphic to first order.
Proof.
Since is
-holomorphic to first order,
. Therefore, for all
, since
is
-holomorphic,

Since is a
-holomorphic Riemannian submersion, the above equation can be written as

Hence, for all , using the fact that
is
holomorphic,

showing that and concluding our proof.
Proposition 4.7.
Let be
-holomorphic to first order map. Then, the projected map
is conformal to first order. Conversely, let
be a conformal to first order map. Then there is a (local)
-holomorphic to first order map
which is compatible to first order with
.
Proof.
Take an
-holomorphic to first order map and let
. We know that
is conformal (Proposition 3.2). As for the first-order variation, using the preceding Lemma 4.6,

Using similar arguments, we can show that

concluding the first part of the proof.
Conversely, let be any twistor lift of
compatible to first order. Then, there is a function
and a vector field
verifying the conditions on Lemma 4.5 with

Now, is
-holomorphic to first order if and only if
is
-holomorphic (which follows from Proposition 3.2) and (4.6) holds. Using the same argument as in Lemma 4.3, (4.6) is equivalent to

But

from the given conditions on and
and thus concluding the proof.
4.4. The
-Holomorphic Case
Next, we give a useful characterization for maps to be -holomorphic to first order (
or
), whose proof is in Section 5 (compare with Proposition 3.3).
Lemma 4.8.
Let be a smooth map and let
be its projection. Then,
is
-holomorphic to first order (
or
) if and only if


or

From the preceding lemma, we can also deduce the following.
Lemma 4.9.
Let be a map
-holomorphic to first order and consider the projected map
. Then for all
there is
with

Proof.
For , we have
so that using (4.22) we have

Taking , we obtain the result. To establish an induction, assume now that the result is valid for
; that is, there is
such that

Taking and noticing that
,

as is antisymmetric on the first two arguments. Now, since
is
-holomorphic, (4.23) holds so that there is
such that

But the second condition gives , whereas the first holds precisely that
, as we wanted to show.
Proposition 4.10 (projections of maps -holomorphic to first order).
Let be a map
-holomorphic to first order, where
is any Riemann surface. Then, the projection map
is isotropic to first order.
Notice that we could replace with
, as real isotropy (to first order) does not depend on the fixed orientation on
.
Proof.
That is isotropic follows from the nonparametric case. Therefore, we are left with proving that

Using Lemma 4.9, for fixed , choose
with
and
. Then, the conclusion follows from

We now turn our attention to the existence of lifts -holomorphic to first order for a given isotropic to first-order map
. Recall that in the nonparametric case such lift exists (see Remark 3.4).
Theorem 4.11.
Let be an isotropic to first order with
and
linearly independent. Then, there is either a (local) map
or a map
which is
-holomorphic to first order and compatible to first order with
.
Before proving Theorem 4.11, we give the following lemma, which we prove in Section 5
Lemma 4.12.
Let be as in the preceding Theorem 4.11.
(i)Suppose that the -holomorphic lift of
is
(resp.,
). Take
the unique positive (resp., negative) almost Hermitian structure on
compatible with
. Then, taking

we have that

(ii)There are ,
such that

We are now ready to prove Theorem 4.11.
Proof of Theorem 4.11.
As before, take or
the
-holomorphic lift of
. Assume, without loss of generality, that it is
. Then, at each
take
the unique positive almost Hermitian structure compatible with
and let us prove that this map
is
-holomorphic to first order. Using Lemma 4.5,
is
-holomorphic to first order and we are left with proving that (4.23) holds. It is enough to prove that there is a basis
of
for which (4.23) holds. Now, take
and
where
is as in (4.32). Then,

Analogously,

where we have used ,
,
, and
. Hence,
and
satisfy equation (4.23), concluding our proof.
4.5. The
Holomorphic Case
We prove the following.
Theorem 4.13.
Let be a map
-holomorphic to first order. Then,
is harmonic to first order (and conformal to first order from Lemma 4.3 and Proposition 4.7).
We first give the following characterization of -holomorphic to first order maps.
Lemma 4.14.
Let (
or
) be a smooth map. Then,
is a
-holomorphic to first order map if and only if
is
-holomorphic and

Proof.
If is any smooth map, then [17]
. Hence, from (2.9),

Thus, we can rephrase equation (4.37) as

which is the condition for -holomorphicity.
Proof of Theorem 4.13.
That is harmonic follows from Theorem 3.1. Hence, we are left with proving that
. Since
is
-holomorphic to first order, we deduce that
is both
and
-holomorphic to first order (Lemma 4.3). From
-holomorphicity (4.37), we have

Using Lemma 4.6 and (4.13) together with symmetry of the second fundamental form of , the right-hand side of the above identity becomes

so that , concluding the proof.
Theorem 4.15.
Let be a harmonic and conformal to first order map. Then, there is (locally) a map
which is
-holomorphic to first order and with
.
Since harmonicity (to first order) does not depend on the orientation on , we could have replaced
by
in both Theorems 4.13 and 4.15
Proof.
For each consider
, bundle over
. Since
is a Riemann surface,
and we can conclude that for each
there is a Koszul-Malgrange holomorphic structure on
. Hence ([16, Theorem
.]), there is a smooth section
with
a Koszul-Malgrange holomorphic section of
:
. So,

equivalently,

Take where
is the
-part on
determined by
= rotation by +
on
(
). (Notice that as
is not conformal we might not get a Hermitian structure by setting
; on the other hand, positive rotation by
comes from the natural orientation on
imported from
via
.) Then
defines a compatible twistor lift of
. Let us check that
is
-holomorphic to first order. That
is holomorphic is immediate. From the proof of Proposition 4.7, we deduce that
is
-holomorphic to first order as it is compatible to first order with
and the latter is conformal to first order. Hence, we are left with proving that (4.37)

holds. We shall establish this equation by showing that both sides agree when applied to any vector . For that, we consider, in turn, the three cases
,
, and
. The first two have similar arguments so that we prove only the first.
(i). From
holomorphicity, we have
. On the other hand, as
is
-holomorphic to first order, (4.13) is satisfied. Finally, for all
,

Since is harmonic to first order, our condition follows from

(ii). We now have

Now, the first condition follows from (4.43) since is Koszul-Malgrange holomorphic for each
. As for the second, letting
denote its left-hand side, we shall prove that
for all
. We do this by establishing the three cases
,
and
(since the first two cases have similar arguments, we prove only the first).
( )When we have

( )Let . Then,

The first term on the right side of the above equation vanishes as lies in
, whereas the second is zero from
-holomorphicity, concluding our proof.
4.6. The 4-Dimensional Case
Theorem 4.16.
Let be harmonic and isotropic to first-order map and with
and
being linearly independent. Then, (locally) there is either a map
or a map
which is simultaneously
and
-holomorphic to first order and with
. Conversely, if
(or
) is
and
-holomorphic to first order, the projected map
is harmonic and isotropic to first order.
Proof.
The converse is obvious from Proposition 4.10 and Theorem 4.13. As for the first part, in Theorem 4.11 we saw that we can lift the map to a map
-holomorphic to first order. Moreover, this lift could be defined as the unique positive or negative almost complex structure compatible with
. On the other hand, in Theorem 4.15 we have seen that there is a map
-holomorphic to first order with
and for which
is compatible. From the comment after Theorem 4.15, there is also a twistor lift of
into
. Therefore, from the dimension of
, we conclude that the lifts constructed in both cited results are the same and, therefore, simultaneously
and
-holomorphic to first order.
We would now like to guarantee the uniqueness to first order of our twistor lift. Before stating such a result, we start with a lemma, proved in Section 5.
Lemma 4.17.
Let be a map
-holomorphic to first order. Consider the twistor projection
and the vectors
and
defined by

so that

Then, for all for which
and
are linearly independent

Notice that in Lemma 4.12, we were given and defined the twistor lift as the unique lift compatible with
. Now, we are given the twistor map
but nothing guarantees that projecting the map to
makes
compatible; that is,
may not preserve
.
Proposition 4.18.
Let be two
-holomorphic to first-order maps such that
and the variational vector fields induced on
are the same; that is, writing
,
, we have
. Then, at all points
for which
and
are linearly independent, writing
,
, we have
.
Proof.
Let (
) denote the projection maps. From our hypothesis, it follows that
. Hence, the only thing left is to prove that the vertical parts coincide. Now, from the proof of Lemma 4.14,
so that our result follows if
. We prove this identity showing that
for all
, where

We consider the four possible cases for ; namely, when
is equal to
,
,
or
, where
and
are as in the preceding lemma (notice that, since
, then
and
).
(i)When (
uses similar arguments), we have

where we have used Lemma 4.17, as well as the fact that and
.
(ii)Taking (
uses similar arguments), we have

But

As , we deduce
; analogously,
so that
.
Hence, the twistor lifts constructed in Theorem 4.16 are unique to first order, in the sense that the vector field induced on
(or
) by the map
,
depends only on the initial projected map
and on the Jacobi field
along
. Moreover, taking
the
-sphere or the complex projective plane, letting
be a harmonic map, and
a Jacobi field, isotropy to first order is immediately guaranteed. Hence, the previous construction allows a (local) unified proof of the twistor correspondence between Jacobi fields and twistor vector fields that are tangent to variations on
which are simultaneously
and
-holomorphic (infinitesimal horizontal holomorphic deformations in [5]). We can also conclude which different properties (namely, conformality, real isotropy or harmonicity) are related with those of the twistor lift (resp.,
,
or
-holomorphicity).
5. Additional Proofs
Proof of Lemma 4.5.
Since is compatible to first order,
preserves
for all
. Hence, there are
and
such that

Take . Since, at
,
we deduce
and
. Now, since
and
form an orthogonal basis for
, we have

Differentiating with respect to at the point
, the above identity yields

Computing the inner product of (5.1) with and using the fact that
vanishes for all
, we get
. Hence,

and we deduce , as
and (5.3) hold. Using (5.1) again, we can now write

which vanishes since for all
, the second and last terms cancelling as
is
-holomorphic and
is conformal to first order. Analogously,
vanishes so that
. For the orthogonal part, taking
,

showing that and concluding the proof.
Proof of Lemma 4.8.
We shall do the proof only for the case, with the
case being similar. Assume that
is
-holomorphic to first order. Then, using Lemmas 4.3 and 4.6,
is
-holomorphic to first order. On the other hand,
satisfies equation (4.37), which implies that for all

Take in
. Let
be such that
and write
. Then,

Moreover, since is holomorphic,
and

finishing the first part of our proof.
Conversely, suppose that (4.22) and (4.23) hold. Take and
with
and
. Then

and we can now easily conclude that (5.7) holds. Together with the fact that is
-holomorphic (Proposition 3.3), we can conclude that (4.37) is verified. As for the horizontal part, we have that

as is holomorphic. Since (4.22) holds, the last condition is trivially satisfied and we can conclude that our map is
-holomorphic to first order, as desired.
Proof of Lemma 4.12.
-
(i)
Since
is isotropic to first order,
, equivalently,
. Thus, we have
(5.12)
Similarly, is equivalent to
and implies

As is compatible with
, Lemma 4.5 guarantees that
is
-holomorphic to first order. On the other hand, since
and
are linearly independent, we deduce that
,
and
form a basis for
. Hence, (4.33) will be satisfied if and only on evaluating the inner product of
and
with which one of these four vectors one obtains the same result. We shall only prove for the first and fourth vectors, the other two cases being similar.
(a)Since ,

(b)Using (5.13),

-
(ii)
We know that
,
,
,
span
. Hence, there are
, and
with
(5.16)
where since
. Therefore,

Now, using the fact that is isotropic to first order, we have

Together with

we deduce

Similarly, from

we have

Equations (5.20) and (5.22) imply that either (which is absurd as
does not lie in
) or
and consequently (4.34) holds, as wanted.
Proof of Lemma 4.17.
That follows from the proof of Proposition 4.7. Since
, and
are linearly independent vectors, we can deduce that
,
,
and
form a basis for
for
is a neighbourhood of
. On the other hand, as
is the projection of a map
-holomorphic to first order, we know that it must be isotropic to first order from Proposition 4.10. Hence,

The argument to establish the second and third identities in (4.52), will now be similar to the one in Lemma 4.12(i).
References
Burstall FE, Lemaire L, Rawnsley JH: Harmonic maps bibliography. http://people.bath.ac.uk/masfeb/harmonic.html
Eells J, Lemaire L: A report on harmonic maps. The Bulletin of the London Mathematical Society 1978, 10(1):1–68. 10.1112/blms/10.1.1
Eells J, Lemaire L: Another report on harmonic maps. The Bulletin of the London Mathematical Society 1988, 20(5):385–524. 10.1112/blms/20.5.385
Lemaire L, Wood JC: Jacobi fields along harmonic 2-spheres in are integrable. Journal of the London Mathematical Society. Second Series 2002, 66(2):468–486. 10.1112/S0024610702003496
Lemaire L, Wood JC: Jacobi fields along harmonic 2-spheres in 3- and 4-spheres are not all integrable. The Tohoku Mathematical Journal. Second Series 2009, 61(2):165–204.
Salamon S: Harmonic and holomorphic maps. In Geometry Seminar, Lecture Notes in Mathematics. Volume 1164. Springer, Berlin, Germany; 1985:161–224. 10.1007/BFb0081912
Koszul J-L, Malgrange B: Sur certaines structures fibrées complexes. Archiv der Mathematik 1958, 9: 102–109. 10.1007/BF02287068
Simões BA, Svensson M: Twistor spaces, pluriharmonic maps and harmonic morphisms. The Quarterly Journal of Mathematics 2009, 60(3):367–385. 10.1093/qmath/han019
Atiyah MF, Hitchin NJ, Singer IM: Self-duality in four-dimensional Riemannian geometry. Proceedings of the Royal Society A 1978, 362(1711):425–461. 10.1098/rspa.1978.0143
O'Brian NR, Rawnsley JH: Twistor spaces. Annals of Global Analysis and Geometry 1985, 3(1):29–58. 10.1007/BF00054490
Davidov J, Sergeev AG: Twistor spaces and harmonic maps. Uspekhi Matematicheskikh Nauk 1993, 48(3):3–96. English translation in Russian Mathematical Surveys, vol. 48, no. 3, pp. 1–91, 1993
Calabi E: Quelques applications de l'analyse complexe aux surfaces d'aire minima. In Topics in Complex Manifolds. Université de Montréal, Montreal, Canada; 1967:59–81.
Rawnsley JH: -structures, -twistor spaces and harmonic maps. In Geometry Seminar, Lecture Notes in Mathematics. Volume 1164. Springer, Berlin, Germany; 1985:85–159. 10.1007/BFb0081911
Chern SS: On the minimal immersions of the two-sphere in a space of constant curvature. In Problems in Analysis. Edited by: Gunning RC. Princeton University Press, Princeton, NJ, USA; 1970:27–40.
Wood JC: Jacobi fields along harmonic maps. In Differential Geometry and Integrable Systems (Proceedings of the 9th MSJ-IRI, Tokyo, 2000), Contemporary Mathematics. Volume 308. American Mathematical Society, Providence, RI, USA; 2002:329–340.
Simões BA: Twistorial constructions of harmonic morphisms and Jacobi fields, Ph.D. thesis. University of Leeds, Leeds, UK; 2007.
Eells J, Salamon S: Twistorial construction of harmonic maps of surfaces into four-manifolds. Annali della Scuola Normale Superiore di Pisa. Classe di Scienze. Serie IV 1985, 12(4):589–640.
Acknowledgments
The author is grateful to Professor J. C. Wood for helpful comments and stimulating discussions during the preparation of this work. The author would also like to thank the referee for valuable comments.
Author information
Authors and Affiliations
Corresponding author
Rights and permissions
Open Access This article is distributed under the terms of the Creative Commons Attribution 2.0 International License (https://creativecommons.org/licenses/by/2.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
About this article
Cite this article
Simões, B.A. First-Order Twistor Lifts. J Inequal Appl 2010, 594843 (2010). https://doi.org/10.1155/2010/594843
Received:
Accepted:
Published:
DOI: https://doi.org/10.1155/2010/594843
Keywords
- Riemann Surface
- Complex Manifold
- Twistor Space
- Holomorphic Section
- Hermitian Structure