Research Article the Laplacian on Riemannian Manifolds Peihe Wang

advertisement
Hindawi Publishing Corporation
Abstract and Applied Analysis
Volume 2013, Article ID 237418, 5 pages
http://dx.doi.org/10.1155/2013/237418
Research Article
A Global Curvature Pinching Result of the First Eigenvalue of
the Laplacian on Riemannian Manifolds
Peihe Wang1 and Ying Li2
1
2
School of Mathematical Sciences, Qufu Normal University, Shandong, Qufu 273165, China
College of Science, University of Shanghai for Science and Technology, Shanghai 200093, China
Correspondence should be addressed to Peihe Wang; peihewang@hotmail.com
Received 8 December 2012; Revised 19 February 2013; Accepted 9 March 2013
Academic Editor: Wenming Zou
Copyright © 2013 P. Wang and Y. Li. This is an open access article distributed under the Creative Commons Attribution License,
which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
The paper starts with a discussion involving the Sobolev constant on geodesic balls and then follows with a derivation of a lower
bound for the first eigenvalue of the Laplacian on manifolds with small negative curvature. The derivation involves Moser iteration.
1. Introduction
2. A Sobolev Constant on the Geodesic Ball
The Laplacian is one of the most important operator on
Riemannian manifolds, and the study of its first eigenvalue is
also an interesting subject in the field of geometric analysis.
In general, people would like to estimate the first eigenvalue
of Laplacian in terms of geometric quantities of the manifolds
such as curvature, volume, diameter, and injectivity radius. In
this sense, the first interesting result is that of Lichnerowicz
and Obata, which proved the following result in [1]: let 𝑀𝑛
be an 𝑛-dimensional compact Riemannian manifold without
boundary with Ric (𝑀) ≥ (𝑛 − 1), then the first eigenvalue
of Laplacian on 𝑀𝑛 will satisfy that πœ† 1 (𝑀) ≥ 𝑛, and the
inequality becomes equality if and only if 𝑀𝑛 ≅ 𝑆𝑛 .
The above result implies that the first eigenvalue of the
Laplacian will have a lower bound less than 𝑛 if the Ricci
curvature of manifolds involved has a lower bound 𝑛 − 1
except on a small part where the Ricci curvature satisfied
that Ric (𝑀) ≥ 0. Now a natural question arises: what is the
lower bound of the first eigenvalue of Laplacian on such a
manifold? In [2], Petersen and Sprouse gave a lower bound
under the assumption that the bad part of the manifolds is
small in the sense of 𝐿𝑝 -norm, where 𝑝 is a constant larger
than half of the dimension of the manifold. In this paper, we
are interested in the lower bound of the first eigenvalue under
the global pinching of the Ricci curvature and we obtain a
universal estimate of this lower bound on a certain class of
manifolds.
The Sobolev inequality is one of the most important tools
in geometric analysis, and the Sobolev constant plays an
important part in the study of this field. In this section, we
will obtain a general Sobolev constant only depending on the
dimension of the manifold on the geodesic ball with small
radius.
Definition 1. Let 𝐡𝑝 (𝑅) ⊆ 𝑀 be a geodesic ball with radius
𝑅; we define the Sobolev constant 𝐢𝑠 (𝑅) on it to be the
infimum among all the constant 𝐢 such that the inequality
||𝑓||22𝑛/(𝑛−2) ≤ 𝐢||∇𝑓||22 holds for all 𝑓 ∈ π‘Š01.2 (𝐡𝑝 (𝑅)).
Definition 2. Let 𝐡𝑝 (𝑅) ⊆ 𝑀 be a geodesic ball with radius
𝑅; we define the isoperimetric constant 𝐢0 (𝑅) on it to be the
supremum among all the constant 𝛼 such that the inequality
Area (πœ•Ω) ≥ 𝛼Vol (Ω)1−(1/𝑛) holds for all Ω ⊆ 𝐡𝑝 (𝑅) with
smooth boundary.
For any fixed point 𝑝 and radius 𝑅, Croke proves that the
equality 𝐢𝑠 (𝑅) = 4((𝑛 − 1)/(𝑛 − 2))2 𝐢0 (𝑅)−2 holds [3], but one
expects the constant 𝐢𝑠 (𝑅) to be independent on the location
of the point 𝑝, under some assumptions. In what follows, we
will give an upper bound to 𝐢𝑠 (𝑅) independent of the point
𝑝.
Let 𝑀 be an 𝑛-dimensional Riemannian manifold, 𝑆𝑀
is the unit tangent bundle of 𝑀, and πœ‹ : 𝑆𝑀 → 𝑀 is
the canonical projective map. Ω ⊆ 𝑀, πœ‰ ∈ 𝑆Ω, π›Ύπœ‰ (𝑑) is the
2
Abstract and Applied Analysis
normalized geodesic from πœ‹(πœ‰) with the initial velocity πœ‰. We
define some notations as follows:
𝜏 (πœ‰) = sup {𝜏 > 0 | π›Ύπœ‰ (𝑑) ∈ Ω, ∀𝑑 ∈ (0, 𝜏)} .
(1)
𝐢(πœ‰) is the arc length from πœ‹(πœ‰) to the cut locus point
along π›Ύπœ‰ (𝑑). Consider
π‘ˆΩ = {πœ‰ ∈ 𝑆٠| 𝐢 (πœ‰) ≥ 𝜏 (πœ‰)} ,
π‘ˆπ‘₯ = (πœ‹ | π‘ˆΩ)−1 (π‘₯) ;
πœ”π‘₯ =
πœ‡π‘₯ (π‘ˆπ‘₯ )
;
𝑐𝑛−1
πœ” = inf πœ”π‘₯ ,
π‘₯∈Ω
(2)
where πœ‡π‘₯ is the standard surface measure of the unit sphere,
𝑐𝑛−1 is denoted to be the area of the unit sphere 𝑆𝑛−1 .
Definition 3. Using the Notation above, πœ” = inf π‘₯∈Ω πœ”π‘₯ is
called the visibility angle of Ω.
If the manifold has Inj (𝑀) ≥ 𝑖 which ensures that any
minimal geodesic starting from any point in 𝐡𝑝 (𝑖/2) will
reach the boundary πœ•π΅π‘ (𝑖/2) before it reaches its cut locus,
then the visibility angle of 𝐡𝑝 (𝑖/2) for any point 𝑝 which we
denote by πœ”(𝑖/2) satisfies πœ”(𝑖/2) = 1.
Lemma 4. Let 𝑀𝑛 be a closed Riemannian manifold with
Inj (𝑀) ≥ 𝑖, then for any 𝑝 ∈ 𝑀, the following Sobolev
inequality holds on 𝐡𝑝 (𝑖/2): ||𝑓||22𝑛/(𝑛−2) ≤ 𝐢||∇𝑓||22 , where
𝑓 ∈ π‘Š01.2 (𝐡𝑝 (𝑖/2)) and 𝐢 = 𝐢(𝑛).
Proof. Croke proved the following inequality [4]:
Area (πœ•Ω)
Vol (Ω)
1−(1/𝑛)
≥
𝑐𝑛−1
(𝑐𝑛 /2)
1−(1/𝑛)
πœ”1+(1/𝑛) ,
(3)
where Ω ⊆ 𝐡𝑝 (𝑖/2), πœ•Ω ∈ 𝐢∞ , and πœ” is just the visibility angle
of the domain Ω.
As discussed above, we will have πœ”(Ω) = 1 if Ω ⊆ 𝐡𝑝 (𝑖/2);
then according to Croke’s inequality, we obtain 𝐢0 (𝑖/2) ≥
𝑐𝑛−1 /(𝑐𝑛 /2)1−(1/𝑛) . The relation between 𝐢0 (𝑖/2) and 𝐢𝑠 (𝑖/2)
tells us that 𝐢𝑠 (𝑖/2) ≤ 𝐢(𝑛), where 𝐢(𝑛) is a constant only
depending on the dimension 𝑛.
Proposition 5. Let 𝑀𝑛 be a closed 𝑛-dimensional Riemannian
manifold with Inj (𝑀) ≥ 𝑖, then for all 𝑝 ∈ 𝑀, Vol (𝐡𝑝 (𝑖/2))
≥ 𝐢𝑛 𝑖𝑛 , where 𝐢𝑛 is a constant only depending on the dimension
𝑛.
Proof. Also take the inequality of Croke
Area (πœ•Ω)
Vol (Ω)
1−(1/𝑛)
≥
𝑐𝑛−1
(𝑐𝑛 /2)
1−(1/𝑛)
πœ”1+(1/𝑛) ,
(4)
then the result can easily be derived from the fact that
πœ”(𝑖/2) = 1 and Area (πœ•π΅π‘ (π‘Ÿ)) = 𝑑Vol (𝐡𝑝 (π‘Ÿ))/π‘‘π‘Ÿ after we
integrate both sides of the inequality.
3. The First Eigenfunction and Eigenvalue
Let 𝑀𝑛 be a closed 𝑛-dimensional Riemannian manifold;
suppose that πœ† 1 (𝑀) is the first eigenvalue of the Laplacian and
𝑒 is the first eigenfunction. In other words, they will satisfy
that Δ𝑒 + πœ† 1 (𝑀)𝑒 = 0. By linearity, we can assume that
−1 ≤ 𝑒 ≤ 1 and inf π‘₯∈𝑀 𝑒 = −1 for the linearity. For the
convenience, we call it the normalized eigenfunction. Next we
will study some properties of the normalized eigenfunction
and the eigenvalue.
Lemma 6. Let 𝑀𝑛 be a closed 𝑛-dimensional Riemannian
manifold with Ric (𝑀) ≥ 0 and Inj (𝑀) ≥ 𝑖. Then, a constant
𝐢1 (𝑛, 𝑖) > 0 can be found such that πœ† 1 (𝑀) ≤ 𝐢1 (𝑛, 𝑖).
Proof. One of the theorems of Yau and Schoen [1] shows that
πœ† 1 (𝑀) ≤ 𝐸𝑛 /𝑑2 ≤ 𝐸𝑛 /𝑖2 β‰œ 𝐢1 (𝑛, 𝑖) if Ric (𝑀) ≥ 0, where 𝑑 is
the diameter of the manifold and 𝐸𝑛 is a constant depending
only on 𝑛.
We will now introduce some notation. Let Ric− (π‘₯) denote
the lowest eigenvalue of the Ricci curvature tensor at π‘₯. For
a function 𝑓(π‘₯) on 𝑀𝑛 , we denote 𝑓+ (π‘₯) = max{𝑓(π‘₯), 0}.
Notice that a Riemannian manifold satisfies Ric ≥ 𝑛 − 1 if
and only if ((𝑛 − 1) − Ric− )+ ≡ 0.
The well-known Myers theorem shows that a closed
manifold with Ric ≥ 𝑛 − 1 would have a bounded diameter
𝑑 ≤ πœ‹. In other words, one can deduce that 𝑑 ≤ πœ‹ if one has
(1/Vol(𝑀)) ∫𝑀((𝑛 − 1) − Ric− )+ 𝑑vol = 0. We will show next
a result analogous to the one in [5] which we will use in our
estimation of the eigenvalue. The proof follows identically; so
it will be omitted (the reader can refer to the aforementioned
article).
Lemma 7. Let 𝑀𝑛 be a closed 𝑛-dimensional Riemannian
manifold with Ric (𝑀) ≥ 0, then for any 𝛿 > 0, there exists
πœ–0 = πœ–0 (𝑛, 𝛿) > 0 such that if
1
∫ ((𝑛 − 1) − Ric − )+ 𝑑vol ≤ πœ–0 (𝑛, 𝛿) ,
Vol (𝑀) 𝑀
(5)
then the diameter will satisfy 𝑑 < πœ‹ + 𝛿. In particular, there
exists πœ–1 = πœ–1 (𝑛) such that if
1
∫ ((𝑛 − 1) − Ric − )+ 𝑑vol ≤ πœ–1 (𝑛) ,
Vol (𝑀) 𝑀
(6)
then the diameter will satisfy 𝑑 < 2πœ‹. This fact, together with
the volume comparison theorem, implies that Vol (𝑀) ≤
𝐢2 (𝑛), where 𝐢2 (𝑛) is also a constant only dependent of 𝑛.
Now, we can get a rough lower bound for the first eigenvalue.
Lemma 8. For 𝑛 ∈ N, let πœ–1 = πœ–1 (𝑛) > 0 as above and suppose
that 𝑀𝑛 is a closed manifold with
Ric (𝑀) ≥ 0,
1
∫ ((𝑛 − 1) − Ric − )+ 𝑑vol ≤ πœ–1 (𝑛) ;
Vol (𝑀) 𝑀
(7)
Abstract and Applied Analysis
3
then there exists a constant 𝐢3 (𝑛) > 0 such that πœ† 1 (𝑀) ≥
𝐢3 (𝑛).
Proof. Set π‘Ž = 2 in the (13) from above. Then applying
Lemma 6, one obtains
Proof. The proof mainly belongs to Li and Yau [6]. Let 𝑒 be
the normalized eigenfunction of 𝑀, set V = log (π‘Ž + 𝑒) where
π‘Ž > 1. Then, we can easily get that
1
|∇𝑒|2 ≤ |∇V|2 (π‘₯) ≤ 2𝐢1 (𝑛, 𝑖) (𝑛 + 2) ;
9
(16)
|∇𝑒|2 ≤ 𝐢4 (𝑛, 𝑖) .
(17)
ΔV =
−πœ† 1 (𝑀) 𝑒
− |∇V|2 .
π‘Ž+𝑒
(8)
Denote that 𝑄(π‘₯) = |∇V|2 (π‘₯), and we then have by the
Ricci identity on manifolds with Ric (𝑀) ≥ 0:
Δ𝑄 = 2V𝑖𝑗2 + 2V𝑗 V𝑗𝑖𝑖 ≥ 2V𝑖𝑗2 + 2 ⟨∇V, ∇ΔV⟩ .
(9)
Proposition 10. Let 𝑀𝑛 be a closed 𝑛-dimensional Riemannian manifold, 𝑒 the first eigenfunction of the Laplacian, and
πœ† 1 (𝑀) the corresponding eigenvalue, then Δ|𝑒|+πœ† 1 (𝑀)|𝑒| ≥ 0
holds in the sense of distribution. Moreover, if 𝑀𝑛 is compact
with boundary, then the same conclusion holds for its Neumann
boundary value problem.
(10)
Proof. From the definition, we know that Δ𝑒 + πœ† 1 (𝑀)𝑒 = 0
holds on 𝑀. Denote
For the term V𝑖𝑗2 , we have
∑V𝑖𝑗2 ≥
𝑖,𝑗
2πœ† (𝑀) 𝑒
(ΔV)2 1
),
≥ (𝑄2 + 1
𝑛
𝑛
π‘Ž+𝑒
therefore,
𝑀+ = {π‘₯ ∈ 𝑀 | 𝑒 (π‘₯) > 0} ,
and for the term ⟨∇V, ∇ΔV⟩, we have
⟨∇V, ∇ΔV⟩ = −
π‘Žπœ† 1 (𝑀)
𝑄 − ⟨∇V, ∇π‘„βŸ© .
π‘Ž+𝑒
Therefore, assume π‘₯0 ∈ 𝑀 to be the maximum of 𝑄; then at
π‘₯0 , we have
0≥
4πœ† (𝑀) 2 (𝑛 + 2) π‘Žπœ† 1 (𝑀)
2
).
𝑄 (π‘₯0 ) + ( 1
−
𝑛
𝑛
𝑛 (π‘Ž − 1)
(12)
Therefore,
𝑄 (π‘₯) ≤ 𝑄 (π‘₯0 ) ≤
(𝑛 + 2) π‘Žπœ† 1 (𝑀)
.
π‘Ž−1
(13)
(𝑛 + 2) π‘Žπœ† 1 (𝑀) 1/2
π‘Ž
π‘Ž + max 𝑒
log (
) ≤ log (
) ≤ 𝑑(
) .
π‘Ž−1
π‘Ž−1
π‘Ž−1
(14)
Let 𝑑 = (π‘Ž − 1)/π‘Ž; then for any 𝑑 ∈ (0 1), we have (𝑛 +
2)πœ† 1 (𝑀) ≥ 𝑑(𝑑−2 (log (1/𝑑))2 ).
Considering the maximum of the right hand and the
upper bound of the diameter derived in Lemma 7, we can
deduce that a positive constant 𝐢3 (𝑛) can be found such that
4𝑒−2
≥ 𝐢3 (𝑛) ,
(𝑛 + 2) 𝑑2
(18)
𝑀0 = {π‘₯ ∈ 𝑀 | 𝑒 (π‘₯) = 0} .
According to the maximum principle of elliptic equation and
the discussion about nodal set and nodal regions in [1], we
can conclude that πœ•π‘€+ = πœ•π‘€− = 𝑀0 is a smooth manifold
with dimension 𝑛 − 1.
For all πœ™ ∈ 𝐢∞ (𝑀), πœ™ ≥ 0, integrating by parts we then
have
∫ |𝑒| Δπœ™ + πœ† 1 (𝑀) πœ™ |𝑒|
𝑀
Denote 𝛾 to be the minimizing unit speed geodesic joining the maximum and minimum points of 𝑒; then integrating
𝑄1/2 along 𝛾, one will get:
πœ† 1 (𝑀) ≥
𝑀− = {π‘₯ ∈ 𝑀 | 𝑒 (π‘₯) < 0} ,
(11)
(15)
where 𝑑 is the diameter of the manifold.
Corollary 9. If the manifold one discussed satisfies all the
conditions in Lemma 8 and its injectivity radius satisfies
Inj (𝑀) ≥ 𝑖 and if one let 𝑒 to be the normalized eigenfunction,
then there exists a constant 𝐢4 (𝑛, 𝑖) > 0 such that |∇𝑒|2 ≤
𝐢4 (𝑛, 𝑖).
=∫
𝑀+
|𝑒| Δπœ™ + πœ† 1 (𝑀) πœ™ |𝑒| + ∫
=∫
πœ•π‘€+
𝑀−
(𝑒
−∫
πœ•π‘€−
=∫
πœ•π‘€−
πœ™
|𝑒| Δπœ™ + πœ† 1 (𝑀) πœ™ |𝑒|
πœ•πœ™
πœ•π‘’
− πœ™ + ) + ∫ πœ™ (Δ𝑒 + πœ† 1 (𝑀) 𝑒)
πœ•π‘›+
πœ•π‘›
𝑀+
(𝑒
πœ•πœ™
πœ•π‘’
− πœ™ − ) − ∫ πœ™ (Δ𝑒 + πœ† 1 (𝑀) 𝑒)
πœ•π‘›−
πœ•π‘›
𝑀−
πœ•π‘’
πœ•π‘’
− ∫ πœ™ + ≥ 0,
−
+
πœ•π‘›
πœ•π‘›
πœ•π‘€
(19)
where 𝑛+ and 𝑛− denote the outward normal direction with
respect to the boundaries of 𝑀+ and 𝑀− , respectively. Note
that πœ•π‘’/πœ•π‘›− ≥ 0 on πœ•π‘€− and πœ•π‘’/πœ•π‘›+ ≤ 0 on πœ•π‘€+ for the
definition of 𝑀− and 𝑀+ . This completes the proof.
When 𝑀 has boundary, we can apply the same reasoning,
except that the test function will require πœ™ ∈ 𝐢0∞ (𝑀). This
gives the proof.
As long as the given manifold is compact, one knows
that the first normalized eigenfunction is then determined.
This indicates that the first normalized eigenfunction of
the Laplacian has a close relation with the geometry of
4
Abstract and Applied Analysis
the manifold. In particular, one would hope to bound the
𝐿2 -norm of first normalized eigenvalue of Laplacian from
below by the geometric quantities. In this sense, we have the
following result.
𝑛
Theorem 11. Let 𝑀 be a closed 𝑛-dimensional Riemannian
manifold with Ric (𝑀) ≥ 0 and Inj (𝑀) ≥ 𝑖. If 𝑒 is the
normalized eigenfunction of the Laplacian, then there exists a
constant 𝐢5 (𝑛, 𝑖) > 0 such that ∫𝑀 𝑒2 ≥ 𝐢5 (𝑛, 𝑖).
Proof. We use Moser iteration to get the result. From Proposition 10, we know that Δ|𝑒| + πœ† 1 (𝑀)|𝑒| ≥ 0 holds on 𝑀 in
the sense of distribution. Set V = |𝑒| and take the point 𝑝 ∈ 𝑀
such that 𝑒(𝑝) = −1.
For π‘Ž ≥ 1, denote 𝑅 = 𝑖/2; πœ™ is a cut-off function on
𝐡𝑝 (𝑅), then we have by integrating by parts:
πœ† 1 (𝑀) ∫
𝐡𝑝 (𝑅)
πœ™2 V2π‘Ž = − ∫
𝐡𝑝 (𝑅)
πœ™2 V2π‘Ž−1 ΔV
= 2∫
πœ™V2π‘Ž−1 + (2π‘Ž − 1) ∫
≥ 2∫
πœ™V2π‘Ž−1 + π‘Ž ∫
𝐡𝑝 (𝑅)
𝐡𝑝 (𝑅)
𝐡𝑝 (𝑅)
𝐡𝑝 (𝑅)
𝐡𝑝
󡄨󡄨
π‘Ž 󡄨2
󡄨󡄨∇ (πœ™V )󡄨󡄨󡄨 = ∫
(𝑅)
𝐡
𝑝
𝐡𝑝 (𝑅)
where we denote ||𝑓||𝑝;Ω = (∫٠𝑓𝑝 )1/𝑝 only for emphasizing
the integral domain.
Set
{πœŽπ‘— } : 𝜎0 =
𝑅
𝑅
𝑅
, 𝜎 = , . . . , πœŽπ‘— = −(2+𝑗) , . . . ,
4 1 8
2
(20)
And putting {π‘Žπ‘— }, {πœŽπ‘— }, {πœŒπ‘— } into (25), we can derive after
iteration that
β€–Vβ€–2π‘Žπ‘—+1 ; 𝐡𝑝 (πœŒπ‘— )
1
≤ [𝐢(πœ† 1 (𝑀)π‘Žπ‘— + 2 ]
πœŽπ‘—
𝑗
πœ™V2π‘Ž−1 ⟨∇πœ™, ∇V⟩
𝐡𝑝 (𝑅)
𝐡𝑝
󡄨󡄨
π‘Ž 󡄨2
󡄨󡄨∇ (πœ™V )󡄨󡄨󡄨 − ∫
(𝑅)
𝐡
𝑝 (𝑅)
󡄨 󡄨2
V2π‘Ž 󡄨󡄨󡄨∇πœ™σ΅„¨σ΅„¨σ΅„¨ ;
therefore, using the Sobolev inequality in Lemma 4,
𝐡𝑝 (𝑅)
V2π‘Ž πœ™2 + ∫
𝐡𝑝 (𝑅)
β€–Vβ€–2;𝐡𝑝 (𝑅) .
Let 𝑗 → +∞, then
β€–Vβ€–∞;𝐡𝑝 (𝑅/2) ≤ ∏[𝐢(πœ† 1 (𝑀)π‘Žπ‘— +
𝑗=0
𝑗=0
1 σ΅„©σ΅„© π‘Ž σ΅„©σ΅„©2
.
σ΅„©πœ™V σ΅„©
𝐢 σ΅„© σ΅„©2𝑛/(𝑛−2)
16 1/2π‘Žπ‘— 𝑗/2π‘Žπ‘—
≤ ∏[𝐢 (πœ† 1 (𝑀) + 2 )]
4
.
𝑅
𝑗=0
(𝑛−2)/𝑛
(23)
Let
in
𝐡𝑝 (𝜌 + 𝜎)
𝐡𝑝 (𝜌)
𝐡𝑝 (𝑅)
𝐡𝑝 (𝜌 + 𝜎)
(28)
(29)
The right hand will converge to a fixed number by using
πœ‡π‘—
πœ‡/(πœ‡−1)
−𝑗
and the fact ∑∞
the fact that ∏∞
𝑗=0 𝐡 = 𝐡
𝑗=0 π‘—πœ‡ is finite
for some 𝐡 ∈ 𝑅, πœ‡ > 1. From πœ† 1 (𝑀) ≤ 𝐢1 (𝑛, 𝑖), we can find a
positive constant 𝐢5 (𝑛, 𝑖) > 0 such that
β€–Vβ€–∞;𝐡𝑝 (𝑅/2) ≤ 𝐢5−1/2 (𝑛, 𝑖) β€–Vβ€–2;𝐡𝑝 (𝑅) ≤ 𝐢5−1/2 (𝑛, 𝑖) β€–Vβ€–2;𝑀.
(30)
in 𝐡𝑝 (𝜌) ,
in
β€–Vβ€–2;𝐡𝑝 (𝑅) .
1
)]
πœŽπ‘—2
∞
1
2𝑛/(𝑛−2)
󡄨󡄨
π‘Ž 󡄨2
(πœ™Vπ‘Ž )
)
󡄨󡄨∇ (πœ™V )󡄨󡄨󡄨 ≥ (∫
𝐢 𝐡𝑝 (𝑅)
𝐡𝑝 (𝑅)
1/2π‘Žπ‘—
1/2π‘Žπ‘—
∞
∏[𝐢 (πœ† 1 (𝑀) π‘Žπ‘— +
󡄨 󡄨2
V2π‘Ž 󡄨󡄨󡄨∇πœ™σ΅„¨σ΅„¨σ΅„¨
1
]
πœŽπ‘—2
The product can be estimated as follows:
≥∫
1,
{
{
{
{
{
{
𝜌+𝜎−π‘Ÿ
{
{
{
,
𝜎
πœ™={
{
{
{
{
{
{
{
{0,
{
1/2π‘Žπ‘ 
(27)
(22)
πœ† 1 (𝑀) π‘Ž ∫
β€–Vβ€–2π‘Žπ‘— ;𝐡𝑝 (πœŒπ‘—−1 )
πœ™2 V2π‘Ž−2 |∇V|2 ,
∞
V2π‘Ž πœ™2 ≥ ∫
1/2π‘Žπ‘—
1
)]
πœŽπ‘ 2
≤ ∏[𝐢 (πœ† 1 (𝑀) π‘Žπ‘  +
𝑠=0
we have
πœ† 1 (𝑀) π‘Ž ∫
(26)
𝑠=0
(21)
=
𝑛
𝑛 𝑗
, . . . , π‘Žπ‘— = (
) ,...,
𝑛−2
𝑛−2
{π‘Žπ‘— } : π‘Ž0 = 1, π‘Ž1 =
πœ™2 V2π‘Ž−2 |∇V|2 .
𝑝 (𝑅)
1 1/2π‘Ž
)] β€–Vβ€–2π‘Ž;𝐡𝑝 (𝜌+𝜎) ,
𝜎2
(25)
𝑗
πœ™2 V2π‘Ž−2 |∇V|2
󡄨󡄨 󡄨󡄨2 2π‘Ž
󡄨󡄨∇πœ™σ΅„¨σ΅„¨ V + 2π‘Ž ∫
(𝑅)
𝐡
+ π‘Ž2 ∫
β€–Vβ€–2π‘Žπ‘›/(𝑛−2);𝐡𝑝 (𝜌) ≤ [𝐢 (πœ† 1 (𝑀) π‘Ž +
{πœŒπ‘— } : 𝜌−1 = 𝑅, . . . , πœŒπ‘— = 𝑅 − ∑πœŽπ‘™ , . . . .
However, using the identity
∫
Putting πœ™ into the inequality above and we then have by
splitting the integral into three parts and using the values of
πœ™ on each of them:
,
(24)
.
Therefore,
∫ 𝑒2 ≥ 𝐢5 (𝑛, 𝑖) .
𝑀
(31)
Abstract and Applied Analysis
5
4. The Lower Bound of the First Eigenvalue
Acknowledgments
Using the same notation as above, we can state the following
result.
The authors owe great thanks to the referees for their
careful efforts to make the paper clearer. Research of the
first author was supported by STPF of University (no.
J11LA05), NSFC (no. ZR2012AM010), the Postdoctoral Fund
(no. 201203030) of Shandong Province, and Postdoctoral
Fund (no. 2012M521302) of China. Part of this work was done
while the first author was staying at his postdoctoral mobile
research station of QFNU.
Theorem 12. For 𝑛 ∈ N, 𝑖, 𝛿 ∈ R+ , there is an πœ– = πœ–(𝑛, 𝑖, 𝛿) >
0 such that any closed manifold 𝑀𝑛 with Ric (𝑀) ≥ 0,
Inj (𝑀) ≥ 𝑖 and
1
∫ ((𝑛 − 1) − Ric − )+ 𝑑vol ≤ πœ–
Vol (𝑀) 𝑀
(32)
will satisfy that πœ† 1 (𝑀) ≥ 𝑛 − 𝛿.
References
Proof. Assume that 𝑒 is the normalized eigenfunction of
Laplacian on 𝑀𝑛 , let 𝑄(π‘₯) = |∇𝑒|2 + (πœ† 1 (𝑀)/𝑛)𝑒2 , direct
computation shows that
[1] S. T. Yau and R. Schoen, Lectures in Differential Geometry,
Scientific Press, 1988.
[2] P. Petersen and C. Sprouse, “Integral curvature bounds, distance
estimates and applications,” Journal of Differential Geometry,
vol. 50, no. 2, pp. 269–298, 1998.
[3] D. Yang, “Convergence of Riemannian manifolds with integral
bounds on curvature. I,” Annales Scientifiques de l’École Normale
Supérieure. QuatrieΜ€me Série, vol. 25, no. 1, pp. 77–105, 1992.
[4] I. Chavel, Riemannian Geometry: A Mordern Introduction,
Cambridge University Press, Cambridge, UK, 2000.
[5] C. Sprouse, “Integral curvature bounds and bounded diameter,”
Communications in Analysis and Geometry, vol. 8, no. 3, pp. 531–
543, 2000.
[6] P. Li and S. T. Yau, “Estimates of eigenvalues of a compact
Riemannian manifold,” in AMS Proceedings of Symposia in Pure
Mathematics, pp. 205–239, American Mathematical Society,
Providence, RI, USA, 1980.
πœ† (𝑀)
πœ† (𝑀)
1
Δ𝑄 = 𝑒𝑖𝑗2 + 𝑒𝑗 𝑒𝑖𝑖𝑗 + 𝑅𝑖𝑗 𝑒𝑖 𝑒𝑗 + 1
𝑒Δ𝑒
|∇𝑒|2 + 1
2
𝑛
𝑛
≥
1−𝑛
πœ† (𝑀) |∇𝑒|2 + 𝑅𝑖𝑗 𝑒𝑖 𝑒𝑗 .
𝑛 1
(33)
Integrating both sides on 𝑀𝑛 , we have
0≥
≥
1−𝑛
πœ† (𝑀) ∫ |∇𝑒|2 𝑑vol + ∫ 𝑅𝑖𝑗 𝑒𝑖 𝑒𝑗 𝑑vol
𝑛 1
𝑀
𝑀
1−𝑛
πœ† (𝑀) ∫ |∇𝑒|2 𝑑vol + ∫ Ric− |∇𝑒|2 𝑑vol
𝑛 1
𝑀
𝑀
1−𝑛
=
πœ† (𝑀) ∫ |∇𝑒|2 𝑑vol + (𝑛 − 1) ∫ |∇𝑒|2 𝑑vol
𝑛 1
𝑀
𝑀
(34)
− ∫ [(𝑛 − 1) − Ric− ] |∇𝑒|2 𝑑vol;
𝑀
therefore,
πœ† 1 (𝑀) ≥ 𝑛 −
Vol (𝑀)
1
πœ† 1 (𝑀) ∫𝑀 |𝑒|2 𝑑vol Vol (𝑀)
(35)
2
× ∫ ((𝑛 − 1) − Ric− )+ |∇𝑒| 𝑑vol
𝑀
if we suppose that
1
∫ ((𝑛 − 1) − Ric− )+ 𝑑vol ≤ πœ–1 (𝑛) .
Vol (𝑀) 𝑀
(36)
If πœ–1 is the one obtained in Lemma 7, then one has:
πœ† 1 (𝑀) ≥ 𝑛 −
𝐢2 𝐢4
1
∫ ((𝑛 − 1) − Ric− )+ 𝑑vol.
𝐢3 𝐢5 Vol (𝑀) 𝑀
(37)
Finally, if one chooses
0 < πœ– = πœ– (𝑛, 𝑖, 𝛿) ≤ min {πœ–1 , 𝛿
𝐢3 𝐢5
},
𝐢2 𝐢4
(38)
then πœ† 1 (𝑀) ≥ 𝑛 − 𝛿 as long as (1/Vol (𝑀)) ∫𝑀((𝑛 − 1) −
Ric− )+ 𝑑vol ≤ πœ–, and this proves the theorem.
Advances in
Operations Research
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Advances in
Decision Sciences
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Mathematical Problems
in Engineering
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Journal of
Algebra
Hindawi Publishing Corporation
http://www.hindawi.com
Probability and Statistics
Volume 2014
The Scientific
World Journal
Hindawi Publishing Corporation
http://www.hindawi.com
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
International Journal of
Differential Equations
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Volume 2014
Submit your manuscripts at
http://www.hindawi.com
International Journal of
Advances in
Combinatorics
Hindawi Publishing Corporation
http://www.hindawi.com
Mathematical Physics
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Journal of
Complex Analysis
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
International
Journal of
Mathematics and
Mathematical
Sciences
Journal of
Hindawi Publishing Corporation
http://www.hindawi.com
Stochastic Analysis
Abstract and
Applied Analysis
Hindawi Publishing Corporation
http://www.hindawi.com
Hindawi Publishing Corporation
http://www.hindawi.com
International Journal of
Mathematics
Volume 2014
Volume 2014
Discrete Dynamics in
Nature and Society
Volume 2014
Volume 2014
Journal of
Journal of
Discrete Mathematics
Journal of
Volume 2014
Hindawi Publishing Corporation
http://www.hindawi.com
Applied Mathematics
Journal of
Function Spaces
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Optimization
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Download