452
Views
0
CrossRef citations to date
0
Altmetric
STATISTICS

Some compactness results by elliptic operators

& | (Reviewing editor)
Article: 1857577 | Received 05 Jul 2020, Accepted 21 Nov 2020, Published online: 04 Jan 2021

Abstract

In this paper, we get two compactness results for complete manifolds by applying a (sub-) elliptic second-order differential operator on distance functions. The first is an extension of a theorem of Galloway and gets an upper estimate for the diameter of the manifold and the second is an extension of a theorem of Ambrose.

Jel codes:

PUBLIC INTEREST STATEMENT

The compactness theorem by Myer’s and volume comparison theorem by Bishop-Gromov are essential tools in differential geometry and analysis on manifolds. In this paper, by using a elliptic second-order differential operator on distance functions, we give two compactness results for complete manifolds.

1. Introduction

One of the most important and celebrated results in Riemannain geometry is the Myer’s compactness theorem (Myers, Citation1941). This major theorem and its generalizations have many applications (Alvarez et al., Citation2015; Ambrose, Citation1957; Frankel & Galloway, Citation1981; Galloway, Citation1981). It states if M is a complete Riemannian manifold and its Ricci curvature is bounded bellow by (n1)a>0, then M is compact and its diameter satisfies diam(M)πa, also by the same argument for the universal covering space, one can conclude that M has finite first fundamental group. We recall two important generalizations of this theorem. The first is about Galloway’s theorem as follows.

Theorem 1.1. (Galloway, Citation1981)(Galloway)Let M be a complete Riemannian manifold and for any unit vector field X, one has

Ric(X,X)a+φ,X,

where a is a positive constant and φ is any smooth function satisfying φc. Then M is compact and its diameter is bounded from above by

diam(M)πac+c2+a(n1).

The second important generalization, is a theorem of Ambrose. Namely,

Theorem 1.2. (Ambrose, Citation1957)(Ambrose) Let M be a complete Riemannian manifold and p be a fixed point for which every geodesic γ(t) emanating p satisfies,

0+Ric(γ (t),γ (t))=.

Then M is compact.

Both theorems above have nice applications in relativistic cosmology (see (Alvarez et al., Citation2015; Ambrose, Citation1957; Frankel & Galloway, Citation1981; Galloway, Citation1981)). Myer’s theorem has been generalized in many ways, for example, Bakry and Qian applied the elliptic operator of the form Δ+X, (where X is some vector field) to the distance function and get some estimate of the diameter of the manifold under some curvature-dimension inequality (Bakry & Qian, Citation2005). Wei and Wylie extended the Myer’s theorem for weighted manifolds by the weighted Laplacian Δfu:=Δu+f,u as follows,

Theorem 1.3. (Wei & Willie, Citation2009) Let M be a complete Riemannian manifold and fC(M). If fk and

Ric+Hessf(n1)H>0,

them M is compact and its diameter satisfies,

diam(M)πH+4k(n1)H.

This result was extended to the operator ΔVu:=Δu+V,u, where VX(M) by J.Wu in (Wu, Citation2017). Also M. P. Cavalcante and et al. extended Theorems 1.1 and 1.2 for weighted manifolds as follows

Theorem 1.4. (Weighted Galloway theorem)(Cavalcante et al., Citation2015) Let M be a complete Riemannain weighted manifold, V be a smooth vector field, and for any unit vector field X on M one has,

RicVk=Ric(X,X)+12LVg(X,X)1kVV(n1)c+φ,X,

where k and c are positive constant and φ is a smooth function with φc. Then M is compact and its diameter satisfies,

diam(M)π(n1)cb(n1)c+b2(n1)c+(n1)+4k.

Theorem 1.5. (Weighted Ambrose theorem)(Cavalcante et al., Citation2015) Let M be a complete Riemannain weighted manifold, p a fixed point and for each geodesic γ(t) emanating from p we have,

0+Ricfk(γ (t),γ (t))dt=,

where Ricfk=Ric+Hessf1kdfdf and k(0,). Then M is compact.

In this paper, we generalize Theorems 1.1 and 1.2 by means of some kind of (sub-)elliptic operators and lower bound of the extended Ricci tensor Ric(X,AX)12LVgX,X as follows, the first is an extension of Galloway’s theorem.

Theorem 1.6. Let M be a complete Riemannian manifold, A a (1,1)-self adjoint tensor field, VX(M) a smooth vector field and φ a smooth function with φK0. Fixed pM and define r(x)=dist(p,x). Assume H>0 be some constant and the following conditions are satisfied,

a) for some constant H>0 and any unit vector field XX(M) we have,

Ric(X,AX)12LVgX,Xn1δnHX2+φ,X,

b) secradMG(r(x)) and limt0+t2ln(f(t))=0, where G(t) and f(t) are defined in Lemma 3.1,

c) fAK2,

d) VK6,

e) divAK3,

f) K4tr(A)K5,

g) for some smooth function K1:RR, we have

X,TA(r,X)(x)1nK1(r(x))X2,XX(M),

such that 0+Kˉ1(t)f (t)f(t)dtK7, where

K1(t)=sin2HtK1(t);tπ2H,K1(t);tπ2H.

for some constants K2,K3,K5,K6,K7. Then M is compact, its fundamental group is finite and its diameter satisfies,

diam(M)πH+1δn(n1)HK6(n1)+K3+K7+2HK2+K5.

The second result is an extension of theorem of Ambrose.

Theorem 1.7. Let M be a complete Riemannian manifold, A a (1,1)-self adjoint tensor field, pM be a fixed point and define r(x)=dist(p,x), assume the following conditions are satisfied,

a) there exist some constant M such that fA+tr(A)divA<M,

b) secradMG(r(x)) and limt0+t2ln(f(t))=0, where G(t) and f(t) are defined in Lemma 3.1,

c) X,TA(r,X)(x)1nK1(r(x))X2,XX(M), where K1(r) is some function, such that 1+K1(t)f (t)f(t)dt<,

d) along every geodesic γ(t) emanating from p one has limt+1tRicγ (t),Aγ (t)dt=, then M is compact and its fundamental group is finite.

2. Preliminaries

In this section, we present the preliminaries. Throughout the paper M=(M,,) is a complete Riemannian manifold. First, we give some definitions.

Definition 2.1. A self-adjoint operator A on M is a 1,1-tensor field with the following property,

X,YX(M),AX,Y=X,AY.

Now we define bounded operator A as follows.

Definition 2.2. Let A be a self-adjoint positive definite operator on M, A is called bounded if there are constants δn>0 such that 0X,AX<δn for any unit vector field X,YX(M),AX,Y=X,AY..

By the following definition, we give some notations about second-order differential operator L on a manifold M with L1=0. In fact a second-order differential L with L1=0 can be written as Lu=ΔAu+X.u=div(Au)+Y.u, where X,Y are some suitable vector fields and X=divA+Y.

Definition 2.3. Let A be a self-adjoint operator on M, XX(M) and uC(M), We define

a) LA,X(u):=divAu+X,u,

b) ΔA(u):=ieiu,Aei,

c) ΔA,X(u):=ΔAu+u,X.

Definition 2.4. Let A be a (1,1)-tensor field on M. Define TA as,

TA(X,Y):=XAYYAX.

It is clear that TA is a (2,1) tensor field.

Example 2.5. If A is the shape operator of a hypersurface ΣnMn+1 then

TA(Y,X)=Rˉ(Y,X)NT,

where Rˉ is the curvature tensor of M and N is a unit normal vector field on ΣnMn+1.

We recall the following extended Bochner formula from (Alencar et al., Citation2015; Gomes & Miranda, Citation2018) to prove Theorem 2.10 which is the main tools to get the compactness results.

Proposition 2.6. Let A be a self-adjoint operator on M, then,

(1) 12ΔA(u2)=trAhess2u+u,(ΔAu)ΔuAu+RicA(u,u),(1)

where RicA is defined in (Fatemi & Azami, Citation2018).

The term ΔuAu in (2.1) is very complicated and depends on the algebraic and analytic properties of the tensor field A. So we try to simplify it to get the better estimates. First, by the following Lemmas, we show some relations about the second covariant derivative of the tensor field A.

Lemma 2.7. Let A be a (1,1)-self-adjoint tensor field on M and X,Y,ZX(M), then

a) 2AX,Y,Z=2AX,Z,Y+R(Z,Y)AXAR(Z,Y)X,

b) 2AX,Y,Z2AY,X,Z=ZTA(Y,X).

Proof. For part (a) we have,

2A(X,Y,Z)=A(X,Y,Z)=ZA(X,Y)=ZYAX+YAZXYA(ZX)ZYA(X)=ZYAXZYAX.

Similarly,

2A(X,Z,Y)=YZAXYZAX.

Thus

2A(X,Y,Z)2A(X,Z,Y)=ZYAXYZAXZ,YAX=R(Z,Y)AX=R(Z,Y)AXAR(Z,Y)X.

For part (b), by definition of T, we have

2A(X,Y,Z)=ZAX,Y=ZAY,X+TAY,XAZX,YAX,ZY=ZAY,X+AZY,X+AY,ZX+ZTAY,XAZX,YAX,ZY=ZAY,X+ZTAY,X.

Lemma 2.8. Let A be a (1,1)self-adjoint tensor field on M, then

ΔAX,X=XdivA,XRicAX,X+RicX,AX+TA(X),X.

where is adjoint of and

TA(X)=ieiTA(ei,X).

Proof. For simplicity let ei be an orthonormal local frame field in a normal neighborhood of p such that with eiej=0 at p. At p Lemma 2.7 implies,

ΔAX,X=ieieiAX,X=i2A(X,ei,ei),X=i2A(ei,X,ei),X+TA(X),X.

So by Lemma 2.7, part (a) we have

ΔAX,X=i2A(ei,X,ei),X+TA(X),X
=i2A(ei,ei,X)+R(ei,X)AeiAR(ei,X)ei,X
+TA(X),X
=XdivA,XRicAX,X+RicX,AX+TA(X),X.

Now, we ready to simplify the term ΔuAu in (1).

Proposition 2.9. Let A be a (1,1)-self-adjoint tensor field on M and uC(M), then

ΔuAu=u.u.tr(A)udivA,u+iei,TA(u,eiu)
(2.2) +iTuA(ei,u),ei+RicAu,uRicu,Au.(2.2)

Proof. Let A be a (1,1)-tensor field, then

ΔuAu=ieiu,eiAu+ieiu,TA(u,ei)
=iei.u,eiAuiu,ei2Au
ieiu,eiAu+ieiu,TA(u,ei)
=iei.u,uAei+iei.u,TA(ei,u)u,ΔAu
ΔuAu+2ieiu,TA(u,ei).

Note

iei.u,TA(ei,u)+2ieiu,TA(u,ei)
=ieiu,TA(ei,u)+iu,eiTA(ei,u)
+iu,TA(ei,eiu)+2ieiu,TA(u,ei)
=iu,eiTA(ei,u)+iu,TA(ei,eiu)
+ieiu,TA(u,ei)
=u,TA(u)+iu,TA(ei,eiu)+ieiu,TA(u,ei)
=u,TA(u)+iei,TA(u,eiu).

In other words,

ΔuAu=u,divuAu,ΔAu+u,TA(u)
+iei,TA(u,eiu).

But,

u,divuA=iu,eiuAei=ieiuAu,ei
=iuuAei+TuA(ei,u),ei
=u.u.tr(A)+iTuA(ei,u),ei.

So,

ΔuAu=u.u.tr(A)u,ΔAu+u,TA(u)
+iei,TA(u,eiu)+iTuA(ei,u),ei.

Finally the result concludes by Lemma 2.8.

Here is another extension of Bochner- formula, which we use it as one of the mail tools to get the compactness results.

Theorem 2.10. (type-Bochner formula)Let X,Y,ZX(M) and A be a (1,1)-self-adjoint tensor field on M, then for any smooth function u we have,

(2.2) 12ΔA,V(u2)=trAhess2u+u,(ΔA,Vu)u.u.tr(A)+udivA,uiei,TA(u,eiu)iTuA(ei,u),ei+Ricu,Au12LVgu,u(2.2)

Proof. From (Alencar et al., Citation2015; Fatemi & Azami, Citation2018; Gomes & Miranda, Citation2018) we can write,

12ΔA(u2)=trAhess2u+u,(ΔAu)ΔuAu+RicA(u,u).

So Proposition 2.9 concludes,

12ΔA(u2)=trAhess2u+u,(ΔAu)u.u.tr(A)+udivA,u
iei,TA(u,eiu)iTuA(ei,u),ei+Ricu,Au.

Finally, the result follows, by considering

u.V,u12LVgu,u
=uV,u+V,uu12V.u,u+V,u,u
=uV,u+V,uu12V.u,u+VuuV,u
=uV,u+V,uu12V.u,u+Vu,uuV,u
=V,uu12V.u,u+Vu,u
=2HessuV,u12V.u,u=HessuV,u.

3. Extended Laplacian comparison theorem

In this section we shall extend the mean curvature comparison theorem by some (sub-) elliptic operators, so we apply Theorem 2.10 to the distance function r(x)=dist(p,x), where p is a fixed point. First, we need some estimate for terms iei,TA(r,eir) and iTrA(r,ei),ei.

Lemma 3.1. Let M be a complete Reimannian manifold and p be a fixed point, r(x)=dist(p,x). Assume the radial sectional curvature of M satisfies secradMG(r(x)) and there is some smooth function K1:RR, which satisfies

X,TA(r,X)(x)1nK1(r(x))X2,X(M),

then we have the following estimates,

iei,TA(r,eir)(x)K1(r(x))f (r(x))f(r(x)),

where

(3.1) f  Gf=0,f(0)=0,f (0)=1,(3.1)

where G:RR be a smooth function.

Proof. By secradMG(r(x)) we have the following estimate for Hessr(X,X):=Xr,X, (see (Pigola et al., Citation2008))

Hessr(X,X)f (r)f(r)X2r,X2.

So for the local orthonormal vector field ei which diagonalize hessr(X):=Xr we have,(hii:=Hessr(ei,ei))

iei,TA(r,eir)(x)=ihiiei,TA(r,ei)nmaxihiimaxiei,TA(r,ei)K1(r(x))f (r(x))f(r(x)).

To approximate iTrA(r,ei),ei, we give the following definition.

Definition 3.2. Let A be a (1,1) tensor field on a complete Riemannian manifold Mn, pM be a fixed point and r(x):=dist(p,x) be the distance function, we define fA:MR with the following property,

iTrA(r,ei),eiHessfA(r,r).

Remark 3.3. Here, we give some description and example a bout fA. Note that, when 2A=0, then fA=0, In fact fA is note unique and it depends on the Algebraic and Analytic properties of the tensor field A. In general, let,

K(r):=maxXTxM,X=1,xB(p,r)iTXA(ei,X),ei(x),

where the radial sectional curvature satisfies secradG and f be the solution of differential Equationequation (3.1), then fA can be the solution of the following differential inequality,

K(r)f ′′+f ffA,fA>0.

Here we get an extension of mean curvature comparison theorem.

Theorem 3.4 (Extended mean curvature comparison) Let M be a complete Riemannian manifold, A a (1,1)-self-adjoint tensor field, VX(M) a smooth vector field and φ a smooth function with φK0. Fixed pM and define r(x)=dist(p,x). Assume H>0 be some constant and the following conditions are satisfied,

a) for any unit vector field XX(M) we have,

Ric(X,AX)12LVgX,Xn1δnHX2+φ,X,

b) K4tr(A)K5,

c) fAK2 for some constant K2

d) secradMG(r(x)) and limt0+t2ln(f(t))=0, where G(t) and f(t) are defined in Lemma 3.1.

e) divAK3,

f) VK6.

Then along any minimal geodesic segment from x0 we have,

a) for rπ4H,

LA,VfA+tr(A)rδn1+4K2+2K52K4δnn1ΔHr+K6(n1)+K3+2K0+1snH2(r)0rsnH2(t)K1(t)f (t)f(t)dt,
b) for π4Hrπ2H,
LA,V(fA+tr(A))rδn1+8K2+4K5K4(n1)sin(2Hr)ΔHr+1(n1)K6+K3+2K0+1snH2(r)0rsnH2(t)K1(t)f (t)f(t)dt.

Proof of Theorem 3.4. We are inspired by the proof of (Fatemi & Azami, Citation2018). By Lemma 3.1 and Theorem 2.10, we get the following differential inequality,

0ΔAr2(n1)δn+r.ΔA,Vrr.r.tr(A)+r.divA,rK1(r)f (r)f(r)r.r.fA
+Ricr,Ar12LVgr,r.

Let γ(t) be a minimal geodesic through the point x0. Then,

0ΔAr2(n1)δn+(ΔA,Vr) tr(A)+fA(t) ′′t+divA,γ (t)K1(t)f (t)f(t)
(3.2) +n1δnH+φ (r).(3.2)

On the space form MHn with constant sectional curvature H, we have

ΔA,VrδnΔHr(ΔAr)2n1δn2(ΔHr)2n1+1δnfA+tr(A) ′′(t)divA,γ (t)+φ (t)
(3.3) +1δnK1(t)f (t)f(t).(3.3)

Formula (3.3) gives,

snH2(r)ΔA,VrδnΔHr=2sn H(r)snH(r)ΔA,VrδnΔHr+snH2(r)ΔA,VrδnΔHr
snH2(r)2ΔHr(n1)ΔA,VrδnΔHr(ΔAr)2n1δn2(ΔHr)2n1
+snH2(r)δnfA+tr(A) ′′(t)divA,γ (t)+φ (t)
+snH2(r)δnK1(t)f (t)f(t)
=snH2(r)n1ΔArδnΔHr2+snH2(t)δnn1V,γ (t)
+snH2(r)δnfA+tr(A) ′′(t)divA,γ (t)+φ (t)
+snH2(r)δnK1(t)f (t)f(t)
snH2(t)δnn1V,γ (t)+snH2(r)δnfA+tr(A) ′′(t)
+snH2(r)δndivA,γ (t)+φ (t)+K1(t)f (t)f(t).

Note limr0snH2(r)ΔArδnΔHr=0. So integration with respect to r concludes,

1δnsnH2(r)ΔA,VrsnH2(r)ΔHr+1δnn10rsnH2(t)V,γ (t)dt+1δn0rsnH2(t)fA(t) ′′dt
+1δn0rsnH2(t)tr(A)t ′′divA,γ (t)+φ (t)+K1(t)f (t)f(t)dt
=snH2(r)ΔHr+1δnn10rsnH2(t)V,γ (t)dt
+1δnsnH2(r)fA(r)+tr(A)r+1δnsnH2(r)φ(r)divA,r
+1δn0rsnH2(t)K1(t)f (t)f(t)dt1δn0rsnH2(t)fA(t)+tr(A)tdt
1δn0rsnH2(t)φ(t)divA,γ (t)dt.

By definition one has,

1δnsnH2(r)LA,VfA+tr(A)rsnH2(r)ΔHr+1δnn10rsnH2(t)V,γ (t)dt
+1δnsnH2(r)φ(r)+1δn0rsnH2(t)K1(t)f (t)f(t)dt
1δn0rsnH2(t)fA(t)+tr(A)tdt
1δn0rsnH2(t)φ(t)divA,γ (t)dt,

note snH2(t)0, hence

0rsnH2(t)divA,γ (t)dtK30rsnH2(t)dt=K3snH2(r),

and

0rsnH2(t)V,γ (t)dtK60rsnH2(t)dt=K6snH2(r).

Therefore the integration by parts implies,

1δnsnH2(r)LA,VfA+tr(A)rsnH2(r)ΔHr+1δn1(n1)K6+K3+2K0snH2(r)
+1δn0rsnH2(t)K1(t)f (t)f(t)dt
(3.4) 1δnfA(r)+tr(A)rsnH2(r)(3.4)
+1δn0rsnH2(t) ′′fA(t)+tr(A)tdt.

For proof (a) we have inequality rπ4H, then (3.4) concludes,

1δnsnH2(r)LA,VfA+tr(A)rsnH2(r)ΔHr+1δnK6(n1)+K3+2K0snH2(r)
+1δn0rsnH2(t)K1(t)f (t)f(t)dt+2K2+K5K4δnsnH2(r).

We know snH2(r)=2n1ΔHrsnH2(r) so

LA,VfA+tr(A)rδn1+4K2+2K52K4δnn1ΔHr+K6(n1)+K3+2K0
+1snH2(r)0rsnH2(t)K1(t)f (t)f(t)dt.

For proof (b), we have

0rsnH2(t) ′′fA(t)+tr(A)dtK2+K50π4HsnH2(t) ′′dtK2π4HrsnH2(t) ′′dt
+K4π4HrsnH2(t) ′′dt
=4K2+2K5K4H+K4K2snH(2r).

Notice that

1snH2(r)4K2+2K5K4H+K4K2snH(2r)=8K2+4K5K4(n1)sin(2Hr)+2K4K2n1ΔHr.

Therefore,

(3.5) LA,VfA+tr(A)rδn1+8K2+4K5K4(n1)sin(2Hr)ΔHr+1(n1)K6+K3+2K0+1snH2(r)0rsnH2(t)K1(t)f (t)f(t)dt(3.5)

Remark 3.5. Note LA,VfA+tr(A)=ΔA,X, for X:=divA+VfA+tr(A). So by Lemma 3.8 in (Fatemi & Azami, Citation2018) the inequality (3.5) is valid in barrier sense.

4. Proofs of theorems 1.6 and 1.7

Now we prove the Theorem 1.6 by using the so called excess functions. We recall for the point p,qM the excess function is defined by ep,q(x)=d(p,x)+d(q,x)d(p,q). For the proof, we use the idea in (Wei & Willie, Citation2009; Wu, Citation2017). By adapting their approach we obtain the compactness result using the extended mean curvature Theorem 3.4 for the elliptic differential operator LA,VfA+tr(A) to the excess function.

Proof of Theorem 1.6. Let p,q are two points in M with distp,qπH. Define B:=distp,qπH, r1(x):=distp,x and r2(x):=distq,x. Let ep,q(x) be the excess function associated to the points p,q. By triangle inequality, we have ep,q(x)0 and ep,qγ(t)=0, where γ is the minimal geodesic joining p,q. Hence by Remark 3.4

LA,VfA+tr(A)eγ(t)0,

in the barrier sense. Let y1=γπ2H and y2=γB+π2H. So riyi=π2H,i=1,2. Theorem 3.4 (a) concludes that

LA,VfA+tr(A)riyiK6(n1)+K3+2K0+Hδn4K2+2(K5K4)
(4.1) +0π2Hsin2(Ht)K1(t)f (t)f(t)dt.(4.1)

Also, By integration with (3.2), we get

(4.2) LA,VfA+tr(A)r1(y2)LA,VfA+tr(A)r1(y1)B(n1)δnH+K0+π2HB+π2HK1(t)f (t)f(t)dt.(4.2)

So by (4.1) and (4.2), we have

0LA,VfA+tr(A)ep,q(y2)=LA,VfA+tr(A)r1(y2)+r2(y2)
K6(n1)+K3+3K0+Hδn4K2+2(K5K4)+K7B(n1)δnH,

thus

B1(n1)δnHK6(n1)+3K0+K3+K7+2Hδn2K2+K5K4,

and

dist(p,q)πH+1δn(n1)HK6(n1)+3K0+K3+K7+2Hδn2K2+K5K4.

The finiteness of its fundamental group can be proved by the similar argument in (Fatemi & Azami, Citation2018).

Finally, we prove Theorem 1.2, the proof is an adaptations of (Cavalcante et al., Citation2015; Wraith, Citation2006).

Proof of Theorem 1.7. Assume the contrary, i.e. the manifold M is not compact. So there is a ray γ(t) emanating from the fixed point x0. The geodesic γ(t) is minimal, so it has no conjugate point, thus ΔAr is smooth along γ(t). Along the geodesic γ(t) we know

Ricγ (t),Aγ (t)ΔAr2(n1)δn(ΔAr) +tr(A) ′′t+fA(t) ′′divA,γ (t)
+K1(t)f (t)f(t).

By integration along the geodesic γ(t), we have

limt+1tRicγ (t),Aγ (t)dtlimt+1t1(n1)δnΔAt2+(ΔAt) dt
+limt+1tfA(t) ′′+tr(A)divA,γ (t)dt
+limt+1tK1(t)f (t)f(t)dt
limt+1t1(n1)δnΔAt2+(ΔAt) dt+M.

For simplicity Let ΔAr=f(r), by smoothness of ΔAr on the geodesic γ, f(r) is smooth for r>0. From the assumption, we have

limt+1tf (t)1(n1)δnf2(t)dt=.

So, limr+f(r)= and

(4.3) limr+f(r)1rf2(t)dt=.(4.3)

By (4.3), there exists some r1>1 such that f(r)1rf2(t)dt>10 for all r>r1. Inductively, define rn+1=rn+101n. Note that when f(r)α, for rrn1, then for rrn+1, f(r)rnrn1α2. By induction, we have f(r)10n,rrn, so limn+f(rn)=. But limn+rn=rr1+10/9, which is a contradiction with the smoothness of f(r) on (0,+).

Additional information

Funding

The authors received no direct funding for this research.

Notes on contributors

Shahroud Azami

The author received his PhD from the Amirkabir University of Technology. He is working as professor at department of mathematics, faculty of sciences, Imam Khomeini international university, Qazvin, Iran. He published a number of research articles in international journals. He guided many postgraduate students. His research area is differential geometry.

References