1. Introduction
Let be a smooth map between Riemannian manifolds and .
A -harmonic map , proposed by Eells and Lemaire [8], is a critical point of the -energy functional
|
|
|
The Euler-Lagrange equation is given by , where is the -tension field.
The concept of -harmonic map is a natural generalization of harmonic map, and in particular for and , the critical points of or are biharmonic or triharmonic maps, respectively (c.f.[24], [25],[11], [12]).
In recent years, biharmonic maps and biharmonic submanifolds have been widely studied, see [10, 3, 19] and the references therein. There are also a lot of results on triharmonic maps and triharmonic submanifolds (c.f.[12, 11, 13, 14, 17, 16, 1, 2]). Concerning triharmonic hypersurfaces in a space form , Maeta [11] proved that any compact CMC triharmonic hypersurface in for is minimal. Recently, Montaldo-Oniciuc-Ratto [15] gave a systematic study of CMC triharmonic hypersurface in space forms. The authors proved that
Theorem 1.1.
([15])
Let be a CMC triharmonic hypersurface in with and assume that the squared norm of the second fundamental
form is constant. Then is minimal.
In particular, the assumption being constant was removed for in [15].
Theorem 1.2.
([15])
Let be a CMC triharmonic surface in .
Then is minimal if and is an open part of the small hypersphere if .
Very recently, Chen-Guan investigated triharmonic CMC hypersurfaces in a space form under some assumptions on the number of distinct principal curvatures in [6, 7].
Theorem 1.3.
([6])
Let () be a CMC proper triharmonic hypersurface with
at most three distinct principal curvatures in . Then has constant scalar curvature.
Theorem 1.4.
([7])
Let () be a CMC proper triharmonic hypersurface with
four distinct principal curvatures in . If zero is a principal curvature
with multiplicity at most one, then has constant scalar curvature.
We recall Chen’s conjecture in the literature of biharmonic submanifolds: any biharmonic submanifold in the Euclidean space is minimal.
There are some important progress in recent years to support the conjecture under some geometric restrictions (see, for instances [9, 19]), however the general case remains open. Taking into account Chen’s conjecture, Maeta [11] further proposed the generalized Chen’s conjecture on -harmonic submanifolds:
Conjecture : Any -harmonic submanifold in the Euclidean space is minimal.
In this paper, we are able to determine the geometry of CMC triharmonic hypersurfaces in a space form without the restrictions on the number of principal curvatures. We will prove the following statements:
Theorem 1.5.
Let be a CMC triharmonic
hypersurface in the hyperbolic space . Then is minimal.
Moreover, restricting ourselves on the case of , we get
Theorem 1.6.
Let be a CMC proper triharmonic
hypersurface in the sphere . Then has constant scalar curvature.
Combining our Theorem 1.6 with the results of Montaldo-Oniciuc-Ratto (Theorem 1.9, [15]) and Chen-Guan (Corollary 1.7, [7]), we provide a more general result for CMC triharmonic
hypersurfaces in .
Corollary 1.7.
Let be a CMC proper triharmonic
hypersurface in the sphere . Then either
(1) and is an open part of , or
(2)
and if and only is an open part of , where is given by
|
|
|
and is the unique real root belonging to of the polynomial
|
|
|
Let us recall the generalized Chern conjecture (c.f.[4, 5]), which says that:
any closed hypersurface in the unit sphere with constant mean curvature and constant scalar curvature is isoparametric. Since the
class of CMC proper triharmonic hypersurfaces in a sphere have constant scalar curvature, the next important problem is to study whether these hypersurfaces are isoparametric. The problem remains open in its full generality. The readers may refer to the recent important progress on the generalized Chern conjecture due to Tang and Yan et al. [22, 23].
Considering the case , we obtain a characterization under an assumption on the multiplicity of zero principal curvature.
Theorem 1.8.
Let be a CMC triharmonic
hypersurface in the Euclidean space . If zero is a principal curvature
with multiplicity at most one, then is minimal.
In particular, we can prove
Theorem 1.9.
Any CMC triharmonic
hypersurface in the Euclidean space is minimal.
At last, we point out that for a CMC proper triharmonic
hypersurface in , the two equations in (2.7) are quite similar to the equations of a proper biharmonic hypersurface in a space form, see for instance [9]. This is reasonable because the geometry property of triharmonicity is much weaker than biharmonicity. Hence, it is expected that more geometric features of triharmonic hypersurfaces could be found. Interestingly, we can achieve a complete classification of CMC triharmonic hypersurfaces in with . This will benefit us in studying biharmonic hypersurfaces in .
The paper is organized as follows. In Section 2, we recall some
background on the theory of triharmonic hypersurfaces in space forms and derive some useful lemmas, which are very important for us to study the geometric properties of triharmonic hypersurfaces. In Section 3, we give the proofs of Theorems 1.5 and 1.6. In Section 4, we finish the proofs of Theorems 1.8 and 1.9.
2. Preliminaries
Let be an -dimensional Riemannian space form
with constant sectional curvature . For an isometric immersion
, we denote by the
Levi-Civita connection of and the Levi-Civita
connection of . The Riemannian curvature tensors of
are respectively given by
|
|
|
|
|
|
The Gauss and Weingarten formulae are stated, respectively, as
|
|
|
|
|
|
|
|
Here are tangent vector fields on , is the unit normal vector field on , is the second fundamental form of , and is the shape operator.
Let us choose an orthonormal frame of . With this frame, define , where are the connection coefficients.
Denote by
|
|
|
|
|
|
|
|
|
|
|
|
From the definition of the Gauss curvature tensor we obtain
|
|
|
|
(2.1) |
Moreover, the Gauss and Codazzi equations are given, respectively, by
|
|
|
|
(2.2) |
|
|
|
|
(2.3) |
The mean curvature function and the squared norm of the second fundamental form are written respectively as
|
|
|
(2.4) |
From the Gauss equation, the scalar curvature is given by
|
|
|
(2.5) |
We recall a fundamental characterization result on CMC triharmonic hypersurfaces in .
Proposition 2.1.
(c.f.[15])
A CMC hypersurface is triharmonic
if the mean curvature and the squared norm of the second fundamental form on
satisfy
|
|
|
(2.6) |
According to (2.6), it is clear that minimal hypersurfaces are automatically triharmonic in
. A triharmonic hypersurfaces in is called proper if it is not minimal.
In the following, we will consider a CMC proper hypersurface
in a space form . Then (2.6)
becomes
|
|
|
(2.7) |
For a hypersurface in , we denote by for its principal curvatures. The number of distinct principal curvatures is locally constant and the set of all points here is an open and dense
subset of . Denote by this set. On a non-empty connected component
of , which is open, the number of distinct principal curvatures is
constant. On that connected component, the multiplicities of the distinct
principal curvatures are constant and hence are always smooth and the shape operator is
locally diagonalizable (see[18, 20, 21]).
Denote by and . If is constant, then is an empty set. From
now on, we assume that is not constant, that is . We will work in
.
Observing from the second equation of (2.7),
it is known that is a principal direction with the corresponding principal curvature . Hence, we may choose an orthonormal frame such that is parallel to and the shape operator is diagonalizable with respect to , i.e., , where is the principal curvature and .
Suppose that has distinct principal curvatures with , that is
|
|
|
where
|
|
|
with and satisfying , namely, is the multiplicity of .
For convenience, we will use the range of indices except special declaration.
We collect a lemma for later use.
Lemma 2.2.
(c.f.[7])
The connection coefficients satisfy:
(1) .
(2) for and .
(3) if the indices satisfy one of the following conditions:
(3a) but ;
(3b) and .
(4) if the indices satisfy one of the following conditions:
(4a) ;
(4b) and ;
(4c) and ;
(4d) with and .
(5)
for , and are mutually different.
(6)
for , and are mutually different.
We first derive some crucial lemmas for studying CMC proper triharmonic hypersurfaces in a space form.
Lemma 2.3.
Denoting by for , we have
|
|
|
(2.8) |
|
|
|
(2.9) |
Proof.
For and , it follows from (2.1) and the terms (1), (2) and (4) in Lemma 2.2 that
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
On the other hand, from the Gauss equation (2.2) we get . Therefore, we obtain (2.8).
For and , and , it follows from (2.1) and Lemma 2.2 that
|
|
|
|
|
|
|
|
|
|
|
|
(2.10) |
which together with gives (2.9). Note that follows from the Gauss equation (2.2) directly.
We thus complete the proof.
∎
Lemma 2.4.
For any , we have
|
|
|
(2.11) |
Proof.
Since the case has been obtained in [7], we only need to prove it for .
Taking into account the definition of from the first expression of (2.4), we have
|
|
|
(2.12) |
Since is constant, differentiating (2.12) with respect to , we obtain
|
|
|
which is equivalent to
|
|
|
(2.13) |
Differentiating (2.12) with respect to for and , we obtain
|
|
|
(2.14) |
Differentiating (2.13) with respect to , from (2.8), (2.12), and (2.14) we have
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
which is equivalent to
|
|
|
(2.15) |
Equations (2.13) and (2.15) imply that (2.11) holds for . Next we will prove that it holds for general by induction.
Differentiating (2.11) with respect to yields
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
(2.16) |
On the other hand, since (2.11) holds for , we differentiate with respect to for and .
It follows from (2.9) that
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Hence the following relation holds for any and
|
|
|
(2.17) |
Since (2.17) holds for any , letting , (2.17) reduces to
|
|
|
which together with (2.14) yields
|
|
|
(2.18) |
Letting and using (2.18), (2.17) reduces to
|
|
|
Similarly, we can gradually show
|
|
|
(2.19) |
Hence, combing (2) with (2.19) gives
|
|
|
(2.20) |
When is even, both of and are odd. From (2.20), can yield as well.
When is odd, both of and are even. We conclude from (2.20) that
|
|
|
|
|
|
|
|
|
|
|
|
(2.21) |
which completes the proof of Lemma 2.4.
∎
Lemma 2.6.
The equation is equivalent to
|
|
|
|
|
|
Proof.
Since for , it follows from Lemma 2.2 that
|
|
|
|
|
|
|
|
|
|
|
|
(2.22) |
Noting , it follows from (2.8) that
|
|
|
|
(2.23) |
|
|
|
|
|
|
|
|
(2.24) |
Differentiating with respect to for and , we get
|
|
|
(2.25) |
Combining (2.24) with (2.25) gives
|
|
|
(2.26) |
Substituting (2.23) and (2.26) into (2), we have
|
|
|
|
|
|
|
|
Hence, the proof has been completed.
∎
3. Proof of Theorems 1.5 and 1.6
The proof of Theorems 1.5 and 1.6:
We will prove Theorems 1.5 and 1.6 by deriving a contradiction from the assumption that
.
Taking in Lemma 2.4, we have
|
|
|
(3.1) |
which is a -th order equation system with a non-zero solution. Hence on we have
|
|
|
(3.2) |
Taking in Lemma 2.4, we obtain
|
|
|
(3.3) |
Next we consider all possible cases.
Case 1. at some . Then from (3.2) we have that
at .
Without loss of generality, we assume , i.e., .
Now the former equations in the system of equations (3.1) determine a new system of equations as follows:
|
|
|
(3.4) |
which has a non-zero solution. Then we have
|
|
|
Since , we have that .
Without loss of generality,
we assume that .
Proceeding in this way, we obtain that
at .
Now (3.3) becomes
|
|
|
Since , the first two equations of the above system imply that
|
|
|
and hence . It is a contradiction, so this case is ruled out.
Case 2. for all at some . In this case, we have
|
|
|
This contradicts at and this case is also ruled out.
Case 3. For any given point , some terms of are zero and the others are not zero. In this case, without loss of generality,
assume for and for . Then the first equations in (3.1) form a new system of equations
|
|
|
(3.5) |
which is a -th order equation system with non-zero solutions. So the coefficient determinant is zero, that is . Without loss of generality,
we assume that . Proceeding in this way, we can show that
. Denote by . Then (3.3) becomes
|
|
|
(3.6) |
The above system of equations means that since and . The first two equations of (3.6) force that , and the second and the third equation of (3.6) force that . Hence we have , a contradiction.
In conclusion, we have that is empty and hence has to be a constant.
From (2.5), we conclude that the scalar curvature of is constant as well. But for , the first equation of (2.7) means that , a contradiction. This completes the proof of Theorems 1.5 and 1.6.
4. Proofs of Theorems 1.8 and 1.9
In this section, we mainly concern CMC triharmonic hypersurfaces in the Euclidean space . For any dimension , we need another assumption that the multiplicity of the zero principal curvature is at most one as discussed in [7] for four distinct principal curvatures.
The proof of Theorem 1.8:
Assume that . We will prove Theorems 1.8 by deriving a contradiction.
Taking in Lemma 2.4, we have
|
|
|
(4.1) |
which is a -th order equation system with a non-zero solution. Hence on we have
|
|
|
(4.2) |
Next we consider three possible cases.
Case 1. at some . Then from (4.2) we have that
at .
Without loss of generality, we assume .
Now the former equations in the system of equations (4.1) determine a new system of equations as follows:
|
|
|
(4.3) |
which has a non-zero solution. Then we have
|
|
|
Since , we have that .
Without loss of generality,
we assume that .
Similarly, we obtain that
at .
Since , the first equation of the above system (4.3) implies that
|
|
|
and hence . It is a contradiction.
Case 2. for all at some . Then
|
|
|
which contradicts at .
Case 3. For any given point , some terms of are zero and the others are not zero. In this case, without loss of generality,
assume for and for . Then the first equations in (4.1) form a new system of equations
|
|
|
(4.4) |
which is a -th order equation system with non-zero solutions. Thus the coefficient determinant is zero, that is . Without loss of generality,
we assume that . Similar discussion as the above yields
. Denote by .
It follows from (4.4) that , because of . In this case, we have .
Therefore we deduce from Lemma 2.6 that
|
|
|
that is
|
|
|
|
|
|
|
|
(4.5) |
where we have used for .
Because we have assumed that the multiplicity of the zero principal curvature is one, i.e. , (4) becomes
|
|
|
(4.6) |
Differentiating with respect to on both sides of equation (4.6), it follows from (2.8) that
|
|
|
(4.7) |
Dividing (4.7) by , and then subtracting (4.6), we have
|
|
|
(4.8) |
Since , it follows from (4.8) that
, and hence for .
This is a contradiction since .
The proof of Theorem 1.9:
Let us restrict the case of a CMC triharmonic hypersurface in . We will still work on and prove Theorem 1.9 with a contradiction.
When the multiplicity of zero principal curvature is one, we can derive a contradiction from Theorem 1.8.
By Theorem 1.3, we only need to deal with the case that the multiplicity of zero principal curvature is two, i.e. . In this case, the principle curvatures on are respectively with .
According to the proof of Theorem 1.8, we can deduce a contradiction when or for all . Let us consider the remaining case that , . It follows from (4.4) that and . For simplicity, we denote and . Then the squared norm of the second fundamental form is given by . Since , we have and hence . In this case, we deduce from (2.9) that . Moreover,
and hence . Since , for and , we find
|
|
|
which means that . Then (2.8) turns into
|
|
|
(4.9) |
One the other hand, taking into account the Gauss equation and (2.2), from Lemma 2.2 we obtain
|
|
|
|
|
|
|
|
(4.10) |
Based on the above discussion, (4) becomes
|
|
|
(4.11) |
and hence
|
|
|
(4.12) |
Noting and differentiating (4.11) with respect to , from (4.9) and (4.10) we have
|
|
|
(4.13) |
Substituting (4.12) into (4.13) and eliminating the terms of one gets
|
|
|
|
|
|
that is
|
|
|
(4.14) |
Because , (4.14) becomes
|
|
|
(4.15) |
Differentiating with respect to on (4.15), from (4.9) and (4.10) we have
|
|
|
(4.16) |
Differentiating with respect to on (4.16) yields
|
|
|
(4.17) |
Eliminating the terms of between (4.16) and (4.17) gives
|
|
|
that is
|
|
|
(4.18) |
which implies that is constant and is constant as well. This contradicts .
In summary, we conclude that and according to (2.6) we know that is minimal. This completes the proof of Theorem 1.9.
Acknowledgement: The authors are supported by the NSFC (No.11801246) and Liaoning
Provincial Education Department Project (No.LJKMZ20221561)