We derive two types of sharp Moser-Trudinger inequalities on complete, simply connected, two-dimensional Riemannian manifolds whose sectional curvatures K satisfy the bounds −b2≤K≤−a2<0.
Citation: Carlo Morpurgo, Liuyu Qin. Moser-Trudinger inequalities on 2-dimensional Hadamard manifolds[J]. AIMS Mathematics, 2024, 9(7): 19670-19676. doi: 10.3934/math.2024959
[1] | Tao Zhang, Jie Liu . Anisotropic Moser-Trudinger type inequality in Lorentz space. AIMS Mathematics, 2024, 9(4): 9808-9821. doi: 10.3934/math.2024480 |
[2] | Yony Raúl Santaria Leuyacc . Supercritical Trudinger-Moser inequalities with logarithmic weights in dimension two. AIMS Mathematics, 2023, 8(8): 18354-18372. doi: 10.3934/math.2023933 |
[3] | Nagendra Singh, Sunil Kumar Sharma, Akhlad Iqbal, Shahid Ali . On relationships between vector variational inequalities and optimization problems using convexificators on the Hadamard manifold. AIMS Mathematics, 2025, 10(3): 5612-5630. doi: 10.3934/math.2025259 |
[4] | Ziqing Yuan, Jing Zhao . Solutions for gauged nonlinear Schrödinger equations on $ {\mathbb R}^2 $ involving sign-changing potentials. AIMS Mathematics, 2024, 9(8): 21337-21355. doi: 10.3934/math.20241036 |
[5] | Mohammad Dilshad, Aysha Khan, Mohammad Akram . Splitting type viscosity methods for inclusion and fixed point problems on Hadamard manifolds. AIMS Mathematics, 2021, 6(5): 5205-5221. doi: 10.3934/math.2021309 |
[6] | Kun Cheng, Wentao Huang, Li Wang . Least energy sign-changing solution for a fractional $ p $-Laplacian problem with exponential critical growth. AIMS Mathematics, 2022, 7(12): 20797-20822. doi: 10.3934/math.20221140 |
[7] | Yony Raúl Santaria Leuyacc . Hamiltonian elliptic system involving nonlinearities with supercritical exponential growth. AIMS Mathematics, 2023, 8(8): 19121-19141. doi: 10.3934/math.2023976 |
[8] | Xiuzhi Yang, G. Farid, Waqas Nazeer, Muhammad Yussouf, Yu-Ming Chu, Chunfa Dong . Fractional generalized Hadamard and Fejér-Hadamard inequalities for m-convex functions. AIMS Mathematics, 2020, 5(6): 6325-6340. doi: 10.3934/math.2020407 |
[9] | Yousaf Khurshid, Muhammad Adil Khan, Yu-Ming Chu . Conformable integral version of Hermite-Hadamard-Fejér inequalities via η-convex functions. AIMS Mathematics, 2020, 5(5): 5106-5120. doi: 10.3934/math.2020328 |
[10] | Jie Wu, Zheng Yang . Global existence and boundedness of chemotaxis-fluid equations to the coupled Solow-Swan model. AIMS Mathematics, 2023, 8(8): 17914-17942. doi: 10.3934/math.2023912 |
We derive two types of sharp Moser-Trudinger inequalities on complete, simply connected, two-dimensional Riemannian manifolds whose sectional curvatures K satisfy the bounds −b2≤K≤−a2<0.
A pinched Hadamard manifold is a complete, simply connected Riemannian manifold (M,g) whose Gaussian curvature K satisfies −b2≤K≤−a2 for some a,b>0. Let μ denote the Riemannian measure of M, induced by the metric g. Below, W1,2(M) denotes the usual Sobolev space on M.
In this note, we establish the following theorem:
Theorem 1. If (M,g) is a two-dimensional pinched Hadamard manifold, then there exists a constant C such that for all u∈W1,2(M) with ‖∇u‖2≤1, we have
∫M(e4πu2−1)dμ≤C, | (1.1) |
and
∫Me4πu2−11+|u|2dμ≤C‖u‖22. | (1.2) |
The exponential constant 4π is sharp in both inequalities.
On pinched Hadamard manifolds of dimension n≥3, a sharp version of estimate (1.1) was derived by Bertrand-Sandeep [1] for the operators Dα=(−Δ)α2 if α is even, and Dα=∇(−Δ)α−12 if α is odd (1≤α<n). A sharp version of (1.2) on the same manifolds and for the same operators was obtained by Morpurgo-Qin [14], also for n≥3. Estimate (1.2) is known as "Moser-Trudinger inequality with exact growth condition", and was obtained first on R2 by Ibrahim-Masmoudi-Nakanishi [6], later extended in higher dimensions and for higher order operators in Masmoudi-Sani [11,12,13], and also in Lu-Tang-Zhu [8].
The first step of the strategy used in [1] was to write u=T(Dαu), where T is an integral operator with a kernel Kα, given in terms of the Green function G(x,y) of the Laplace-Beltrami operator on M; in particular, K2=G and K1=∇yG. The second step was to derive sharp asymptotic estimates and critical integrability estimates on Kα, which allowed the authors to apply general results in [5] regarding sharp Adams type inequalities on measure spaces.
The same ideas were then used in [14] where further estimates on Kα were obtained, allowing the authors to apply general results in [14] on Adams inequalities with exact growth conditions on metric measure spaces.
The case n=2 was somehow left out in the above works, due to small technical reasons. The purpose of this note is to fill in the gap and to establish the required estimates on K1=∇yG even when n=2.
It is enough to prove the inequalities of the theorem for u∈C∞c(M). Let G(x,y) be the minimal positive Green function on M. Then, for each x∈M we can write
u(x)=∫Mλ∇yG(x,y), ∇yu(y)⟩dμ(y), | (2.1) |
where λZ,W⟩=g(Z,W) for Z,W∈TyM, and ∇yG is the gradient of G(x,y) with respect to the y variable. Also, below we will denote the open ball centered at x and with radius r as B(x,r)={y∈M:d(x,y)<r} and its volume as Vx(r)=μ(B(x,r)).
For a measurable function f on M, we define its nonincreasing rearrangement as
f∗(t)=inf{s>0:μ{x:|f(x)|>s})≤t},t>0. | (2.2) |
In view of (2.1) and Theorem 1 in [5] (see also (10.8), (10.10), and related remarks in [14]), to prove (1.1) it is enough to show that
|∇yG(x,⋅)|∗(t)≤12√πt−1/2+C,0<t≤1,x∈M | (2.3) |
|∇yG(⋅,y)|∗(t)≤Ct−1/2,t>0,y∈M | (2.4) |
and
∫∞1(|∇yG(x,⋅)|∗(t))2dt≤C,x∈M. | (2.5) |
On the other hand, to prove (1.2), it is enough to verify the following additional conditions:
∫r1≤d(x,y)≤r2|∇yG(x,y)|2dμ(y)≤14πlogVx(r2)Vx(r1)+C,0<Vx(r1)<Vx(r2)≤1, | (2.6) |
∫d(x,y)≥r|∇yG(x,y)|2dμ(y)≤C,Vx(r)≥1, | (2.7) |
|∇yG(x,⋅)|∗(t)≤Ct−1/2,t>0, | (2.8) |
|∇yG(x,y)|≤CVx(d(x,y))−1/2,Vx(d(x,y))≤1, | (2.9) |
and for each δ>0, there is Bδ>0 such that
∫d(x,y)>R|∇yG(x′,y)−∇yG(x,y)|2dμ(y)≤Bδ, Vx(R)≥(1+δ)Vx(r), Vx(r)≤1, x′∈B(x,r). | (2.10) |
We first show that for any given R>0, there exists C such that
|∇yG(x,y)|≤C{d(x,y)−1if d(x,y)≤R1if d(x,y)≥R | (2.11) |
with C independent of x,y (but depending on R). First, let us recall this consequence of the well-known Li-Yau gradient estimate ((3.14) in [1], Thm. 6.1 in [9], Lemma 2.1 in [10]):
|∇yG(x,y)|≤G(b+Cd(x,y)−1),x,y∈M. | (2.12) |
Observe that such an estimate is valid in any dimension n, and that it was used in [1] when n≥3 to derive the bound |∇yG(x,y)|≤Cd(x,y)1−n, given the bound G(x,y)≤Cd(x,y)2−n. However, in dimension 2 this does not work for small distances, since G(x,y) behaves like −logd(x,y).
Instead, we use the fact that
G(x,y)=∫∞0H(t,x,y)dt | (2.13) |
where H(t,x,y) is the heat kernel for the Laplace-Beltrami operator on M, and give some good (even if rough) estimates on |∇yH|. First, we recall that on any n−dimensional Hadamard manifold satisfying K≤−a2, (a≥0), we have the comparison theorem
H(t,x,y)≤Ha(t,d(x,y)),t>0,x,y∈M, | (2.14) |
where Ha(t,da(˜x,˜y)) is the heat kernel on the space form of constant curvature −a2, and where da(˜x,˜y) denotes the distance of two points ˜x,˜y in such space (see Théorème 1 in [4]). When a>0, the heat kernel Ha is well-known and somewhat explicit. In particular, we have that Ha(t,r)=anH1(a2t,ar), for all t,r>0, where H1(t,r) yields the heat kernel on the hyperbolic space Hn, and satisfies the estimate
H1(t,r)≤Cnt−n/2(1+r)(1+r+t)n−32e−(n−1)24t−n−12r−r24t≤Cn{(1+r)n−12e−n−12rt−n/2e−r24tif 0<t≤1(1+r)n2e−n−12rt−3/2e−(n−1)24t−r24tif t≥1 | (2.15) |
for some Cn depending on n (see [3], Thm. 3.1]). Hence, there are some c1,c2,c3>0 depending on n,a such that
H(t,x,y)≤c1e−c2d(x,y){t−n/2e−d(x,y)24tif 0<t≤1e−c3t−d(x,y)24tif t≥1. | (2.16) |
Now, we can appeal to a result by E. B. Davies [2], Theorem 6, which, under the additional condition Ric≥−(n−1)b2, gives
|∇yH(t,x,y)|≤c4{t−n/2−1e−d(x,y)28tif 0<t≤1e−c3t if t≥1, | (2.17) |
for some c4 depending on a,b,n. From (2.13) we then easily get
|∇yG(x,y)|≤C{d(x,y)1−nif d(x,y)≤R1if d(x,y)≥R | (2.18) |
in any dimension n≥2 (and, hence, (2.11)).
We remark that (2.17) can be refined somewhat by replacing c4 with c4e−c2d(x,y), which also gives |∇yG(x,y)|≤Ce−c5d(x,y) for d(x,y)≥R, some c5>0. This can be done using the same method as in [2], using the the gradient estimate for the heat kernel (see, e.g., [9], Thm. 12.2)
|∇yH|2≤αH∂tH+H2(nα22t+Cnα2α−1b2), | (2.19) |
for any α>1, combined with (2.16) and the time derivative estimate
|∂H∂t(t,x,y)|≤c6e−c2d(x,y){t−n/2−1e−d(x,y)28tif 0<t≤1e−c3t−d(x,y)28t if t≥1. | (2.20) |
The latter estimate can be obtained using the method in [2], Theorem 4.
Using (2.11) and following the same argument in [1], Theorem 3.2, with some minor changes, we can now obtain
|∇yG(x,y)|≤12πd(x,y)−1+C, d(x,y)≤1, | (2.21) |
which implies (2.3), (2.9), and (2.6) by the volume comparison theorem.
For the convenience of the reader, we outline the proof of (2.21), keeping a part of the notation used in [1], Thm. 3.2, so the changes are slightly more evident. For any fixed x∈M, an n-dimensional Hadamard manifold, we have a unique chart given by the exponential map expx(rξ). The volume element in geodesic polar coordinate is given by dμ=rn−1√|g|drdξ, where g=(gij) is the metric tensor, evaluated at expx(rξ), |g|=det(gij), and where dξ is the measure on the unit sphere Sn−1.
Letting Φ(r)=−12πlogr and r=d(x,y), viewed as a function of y for fixed x, we have, for all r>0,
ΔΦ(r)=Φ′′(r)|∇r|2+Φ′(r)Δr=−12πr∂rlog√|g| | (2.22) |
where we used |∇r|=1 and Δr=∂rlog(rn−1√|g|) in any dimensions ([7], Lemma 11.13). Letting
Hx(y)=12πr∂rlog√|g|,y=expx(rξ) | (2.23) |
it is then easy to check that, in the sense of distributions,
ΔΦ(d(x,⋅))=−δx−Hx. | (2.24) |
From the Laplacian comparison ([7], Thm. 11.5), we have, in any dimension n,
(n−1)acoth(ar)≤Δr=n−1r+∂rlog√|g|≤(n−1)bcoth(br) | (2.25) |
from which we deduce, when n=2,
|∂rlog√|g||≤Cr,r≤4 | (2.26) |
where C is independent of x. Hence, we get
|Hx(y)|≤C,d(x,y)≤4. | (2.27) |
Define now
Ux(y)=∫MG(y,z)ψ(d(x,z))Hx(z)dμ(z) | (2.28) |
where ψ:[0,∞)→R, is smooth and such that ψ=1 on [0,2] and ψ=0 on [4,∞). One then has that the function hx(y):=G(x,y)−Φ(d(x,y))−Ux(y) is harmonic in the ball B(x,2), as a function of y, and also uniformly bounded on ∂B(0,2). This last fact follows from a uniform estimate on Ux, which can be verified as in [1] with some small modifications. First, we have the Green function comparison
G(x,y)≤Ga(d(x,y))=−12πlogtanh(ad(x,y)2),x,y∈M | (2.29) |
where Ga(da(˜x,˜y))) gives the Green function on the two-dimensional space form of constant curvature −a2<0. Since for r>0 we have |Ga(r)|≤max{log2r,1}+Ca for some Ca>0, then using the volume and the Rauch comparison theorems, we get for y=expx(¯y),z=expx(¯z), with ¯y,¯z∈R2,
|Ux(y)|≤C∫B(x,4)G(y,z)dμ(z)≤C∫|¯y−¯z|≤2log2|¯y−¯z|d¯z+C∫|¯z|≤4(1+Ca)d¯z≤C. | (2.30) |
(We note that in [1], the authors used the bound G(x,y)≤G0(d(x,y))=cnd(x,y)2−n, but this cannot be done when n=2.)
Using the gradient bound for harmonic functions, we can conclude that |∇hx| is bounded on B(x,1), uniformly w.r. to x. Moreover, using (2.11), the fact that G is symmetric, and arguing as in (2.30), if d(x,y)≤1, we obtain
|∇yUx(y)|≤C∫B(x,4)|∇yG(y,z)|dμ(z)≤C∫d(x,z)≤4d(y,z)−1dμ(z)≤C. | (2.31) |
Finally, we have
|∇yG(x,y)−∇yΦ(d(x,y))|≤|∇yUx(y)+∇yhx(y)|≤C,d(x,y)≤1, | (2.32) |
which gives (2.21).
Next, note that estimate (3.25) in[1] still holds when n=2, i.e.,
μ({y:G(x,y)>s})≤4a2s, s>0. | (2.33) |
Using (2.33) and the gradient estimate (2.12), for s<1 we get
μ({y:|∇yG|>s})≤μ({y:G(b+Cd−1)>s})≤μ({y:d≤1})+μ({y:d>1, G(b+Cd−1)>s})≤C+μ({y:(C+b)G>s})≤C+Cs−1≤Cs−1. | (2.34) |
Hence,
|∇yG(x,⋅)|∗(t)≤Ct−1, t>1 | (2.35) |
and (2.5) holds. Note also that |∇yG|≤C for large distances therefore, by Remark 17 in [14], (2.7) follows. Furthermore, (2.35) and (2.3) imply (2.8).
By (2.33), (2.12), and the symmetry of G(x,y) in x,y, we have
|∇yG(⋅,y)|∗(t)≤Ct−1, t>1, | (2.36) |
which combined with (2.21) gives (2.4). To prove (2.10), we follow the same argument as in [14], which still works when n=2, given (2.11).
The sharpness of the exponential constant 4π in (1.1) and (1.2) is proved in the usual way, by considering the "Moser family" of functions vϵ∈W1,n0(B(x0,1)), and some fixed x0∈M, defined as
vϵ(y)={log1ϵif d(x0,y)≤ϵlog1d(x0,y)if ϵ<d(x0,y)≤10if d(x0,y)>1. | (2.37) |
We derived two types of sharp Moser-Trudinger inequalities on complete, simply connected, two-dimensional Riemannian manifolds whose sectional curvatures K satisfy the bounds −b2≤K≤−a2<0. The results fill gaps that were left in [1] and [14], where such inequalities were proved in dimensions n≥3.
Carlo Morpurgo and Liuyu Qin: Conceptualization, Methodology, Software, Validation, Formal analysis, Investigation, Resources, Data curation, Writing-original draft, Writing-review & editing, Visualization, Supervision. All authors of this article have been contributed equally. All authors have read and approved the final version of the manuscript for publication.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
The second author was supported by the National Natural Science Foundation of China (12201197).
The authors declare no conflicts of interest.
[1] |
J. Bertrand, K. Sandeep, Sharp Green's function estimations on Hadamard manifolds and Adams inequality, Int. Math. Res. Not. IMRN, 6 (2021), 4729–4767. https://doi.org/10.1093/imrn/rnaa216 doi: 10.1093/imrn/rnaa216
![]() |
[2] | E. B. Davies, Pointwise bounds on the space and time derivatives of heat kernels, J. Oper. Theory, 21 (1989), 367–378. |
[3] |
E. B. Davies, N. Mandouvalos, Heat kernel bounds on hyperbolic space and Kleinian groups, Proc. London Math. Soc., 57 (1988), 182–208. http://dx.doi.org/10.1112/plms/s3-57.1.182 doi: 10.1112/plms/s3-57.1.182
![]() |
[4] | A. Debiard, B. Gaveau, E. Mazet, Théorèmes de comparaison en géométrie riemannienne, Publ. Res. Inst. Math. Sci., 12 (1976/77), 391–425. |
[5] |
L. Fontana, C. Morpurgo, Adams inequalities for Riesz subcritical potentials, Nonlinear Anal., 192 (2020), 111662. http://dx.doi.org/10.1016/j.na.2019.111662 doi: 10.1016/j.na.2019.111662
![]() |
[6] |
S. Ibrahim, N. Masmoudi, K. Nakanishi, Trudinger-Moser inequality on the whole plane with the exact growth condition, J. Eur. Math. Soc. (JEMS), 17 (2015), 819–835. http://dx.doi.org/10.4171/jems/519 doi: 10.4171/jems/519
![]() |
[7] | J. Lee, Introduction to Riemannian Manifolds, 2 Eds., Berlin: Springer, 2018. http://dx.doi.org/10.1007/978-3-319-91755-9 |
[8] |
G. Lu, H. Tang, M. Zhu, Best constants for Adams' inequalities with the exact growth condition in Rn, Adv. Nonlinear Stud., 15 (2015), 763–788. http://dx.doi.org/10.1515/ans-2015-0402 doi: 10.1515/ans-2015-0402
![]() |
[9] | P. Li, Geometric Analysis, Cambridge: Cambridge University Press, 2012. http://dx.doi.org/10.1017/CBO9781139105798 |
[10] |
P. Li, J. Wang, Complete manifolds with positive spectrum. Ⅱ, J. Diff. Geom., 62 (2002), 143–162. http://dx.doi.org/10.4310/jdg/1090425532 doi: 10.4310/jdg/1090425532
![]() |
[11] |
N. Masmoudi, F. Sani, Adams' inequality with the exact growth condition in R4, Commun. Pure Appl. Math., 67 (2014), 1307–1335. http://dx.doi.org/10.1002/cpa.21473 doi: 10.1002/cpa.21473
![]() |
[12] |
N. Masmoudi, F. Sani, Trudinger-Moser inequalities with the exact growth condition in Rn and applications, Commun. Partial Diff. Equ., 40 (2015), 1408–1440. https://doi.org/10.1080/03605302.2015.1026775 doi: 10.1080/03605302.2015.1026775
![]() |
[13] |
N. Masmoudi, F. Sani, Higher order Adams' inequality with the exact growth condition, Commun. Contemp. Math., 20 (2018), 1750072. http://dx.doi.org/10.1142/S0219199717500729 doi: 10.1142/S0219199717500729
![]() |
[14] | C. Morpurgo, L. Qin, Sharp Adams inequalities with exact growth conditions on metric measure spaces and applications, Math. Ann., 2023. https://doi.org/10.1007/s00208-023-02771-y |