<<

Advances in Pure Mathematics, 2016, 6, 903-914 http://www.scirp.org/journal/apm ISSN Online: 2160-0384 ISSN Print: 2160-0368

On the Prime Theorem for Non-Compact Riemann Surfaces

Muharem Avdispahić, Dženan Gušić

Department of Mathematics, Faculty of Sciences and Mathematics, University of Sarajevo, Sarajevo, Bosnia and Herzegovina

How to cite this paper: Avdispahić, M. and Abstract Gušić, Dž. (2016) On the Prime Geodesic Theorem for Non-Compact Riemann Sur- We use B. Randol’s method to improve the error term in the prime geodesic theorem faces. Advances in Pure Mathematics, 6, 903- for a noncompact Riemann having at least one cusp. The case considered is a 914. general one, corresponding to a of the first kind and a multiplier http://dx.doi.org/10.4236/apm.2016.612068 system with a weight on it. Received: October 23, 2016 Accepted: November 19, 2016 Keywords Published: November 22, 2016 Selberg Trace Formula, Selberg Zeta Function, Prime Geodesic Theorem Copyright © 2016 by authors and Scientific Research Publishing Inc. This work is licensed under the Creative Commons Attribution International 1. Introduction License (CC BY 4.0). http://creativecommons.org/licenses/by/4.0/ The Selberg trace formula, introduced by A. Selberg in 1956, describes the spectrum of Open Access the hyperbolic Laplacian in terms of geometric data involving the lengths of on a . Motivated by analogy between this trace formula and the explicit formulas of relating the zeroes of the to prime numbers, Selberg [1] introduced a zeta function whose analytic properties are encoded in the Selberg trace formula. By focusing on the Selberg zeta function, H. Huber ([2], p. 386; [3], p. 464), proved an analogue of the prime number theorem for compact Rie- 3 1 − 4 mann surfaces with the error term Ox(log x) 2 that agrees with Selberg’s one.  Using basically the same method as in [4], D. Hejhal ([5], p. 475), established also the prime geodesic theorem for non-compact Riemann surfaces with the remainder 3 1 − 4 Ox(log x) 2 . However, in the compact case there exist several different proofs (see,  B. Randol [6], p. 245; P. Buser [7], p. 257, Th. 9.6.1; M. Avdispahić and L. Smajlović

DOI: 10.4236/apm.2016.612068 November 22, 2016 M. Avdispahić, Dž. Gušić

3 − 4 1 [8], Th. 3.1) that give the remainder Ox(log x) . Thanks to new integral repre-  sentations of the logarithmic derivative of the Selberg zeta function (cf. [9], p. 185; [10], p. 128), M. Avdispahić and L. Smajlović ([11], p. 13) were in position to improve 3 1 3 − − 4 4 1 Ox(log x) 2 error term in a non-compact, finite volume case up to Ox(log x) .   Whereas the authors in [8] and [11] approached the prime number theorem in various settings via explicit formulas for the Jorgenson-Lang fundamental class of functions, our main goal is to obtain this improvement for non-compact Riemann surfaces with cusps following a more direct method of B. Randol [6].

2. Preliminaries

Let X be a non-compact Riemann surface regarded as a quotient Γ \  of the upper half-plane  by a finitely-generated Fuchsian group Γ⊆PSL( 2, ) of the first

kind, containing n1 ≥ 1 cusps. Let ℑ denote the fundamental region of Γ . We shall assume that the fundamental region ℑ of Γ has a finite non-Euclidean area ℑ . We put ab az+ b Γ= ∈SL(2,) : ∈Γ cd cz+ d

and denote by v the multiplier system of the weight m ∈  for Γ . Let ψ be an ir- reducible rr× unitary representation on Γ and WT( ) =ψ ( T) vT( ) , T ∈Γ. For an r dimensional vector space V over  we consider an essentially self-adjoint oper- ator

2 ∂∂ ∂ ∆=m y 22 + −imy ∂∂xy ∂x

on the space m of all twice continuously differentiable functions fV:  → , such

that f and ∆m ( f ) are square integrable on ℑ , and satisfy the equality

m (cz+ d ) ab f( Sz) = m W( S) f( z), for all z ∈ and  = ∈Γ. cz+ d cd

  The operator −∆m has the unique self-adjoint extension −∆m to the space m , a 2 dense subspace of L (Γ \ ) . Let Tj , jn= 1, , 1 be the set of parabolic transforma-

tions corresponding to n1 cusps of Γ . WT( j ) does not depend on the choice of a

representative of the parabolic class {Tj } and can be considered as a matrix from rr×  . By m j we will denote the multiplicity of 1 as an eigen-value of the matrix n1 * WT( j ) , and nm1 = ∑ j will be the degree of singularity of W. We mention that oper- j=1  * ator −∆m has both the discrete and continuous spectrum in the case n1 ≥ 1, and only * the discrete spectrum in the case n1 = 0 . The discrete spectrum will be denoted as λ =λλ < < < λ →∞ { n}n>0 ( 0 01 n ). The continuous spectrum is expressed through

904 M. Avdispahić, Dž. Gušić

zeros (or equivalently poles) of the hyperbolic scattering determinant (see, [12]).

3. Selberg Zeta Function

Let PΓh denotes the set of Γ -conjugacy classes of a primitive hyperbolic element P0 in Γ , and Γh denotes the set of Γ -conjugacy classes of a hyperbolic element P in Γ that satisfy property Tr(P) > 2 . Assume that m ≤ 1 . We define the Selberg zeta function associated to the pair ( Γ,W ) by ∞ − −−sk ZΓ,Wr( s) =∏∏ det (I WP( 00) NP( ) ). Pk0h∈=PΓ 0

ZsΓ,W ( ) is absolutely convergent for Re(s) > 1. Analytic considerations given in ([5], pp. 499-501) yield that the Selberg zeta function in this setting satisfies the func- tional equation

ZΓΓ,,WW( s) Ψ=( sZ) (1 − s) with the fudge factor  1 s η′ Ψ=(s) φη( s) ⋅∫ ( uu)d. (1) 2 1 η 2 Here, φ denotes the hyperbolic scattering determinant. It can be represented in the form

n* 1 1 πΓ(ss) Γ−∞ 2 an φ (s) = ∑ 2s , mm   n=1 g Γ+ss  Γ−  n 22   where the coefficients an and gn depend on the group Γ (see, [5], p. 437). Here, * n1 denotes the degree of singularity of W (see Section 2). An explicit expression for the fudge factor η in the Equation (1) is given in ([5], p. 501, Equation (5.10)).

The logarithmic derivative of the Selberg zeta function ZsΓ,W ( ) is given by Zs′ () Λ (P) Γ,W = ∑ s Tr(WP( )) , ZsΓ,W () P∈Γ h NP( )

log NP( 0 ) where NP( ) denotes the norm of the class P and Λ=(P) −1 for a primi- 1− NP( ) = n tive element P0 such that PP0 for some n ∈  . We will omit the indices in ZΓ,W in the sequel.

4. Counting Functions ψ n ( xW, )

Lemma 1. For Re(s) > 1,

Zs′′( ) −s Zs( +1) =∑ Λ+1 (P)Tr ( WP( )) NP( ) , Zs( ) P∈Γ h Zs( +1)

905 M. Avdispahić, Dž. Gušić

n where Λ=10(P) log NP( ) for a primitive element P0 such that PP= 0 for some n ∈  . Proof.

Zs′( ) −−−1 =Λ−1 s ∑ 1 (P) Tr( WP( )) ( 1 NP( ) ) NP( ) Zs( ) P∈Γh

−−ss−1 =Λ∑∑11(P) Tr ( WP( )) NP( ) +Λ( P) Tr ( WP( )) ( NP( ) −1) NP( ) PP∈Γhh∈Γ −1 =Λ −ss+Λ − −11−+( ) ∑∑11(P) Tr ( WP( )) NP( ) ( P) Tr( WP( )) ( 1 NP( ) ) NP( ) . PP∈Γhh∈Γ

We shall spend the rest of this section to derive a representation of ψ 2 ( xW, ) in the form (11) bellow. We choose not to write it in a separate statement because of the length of expressions involved. However, it will serve as a base for the proof of the prime geodesic theorem in Section 5. Let us recall the following theorem given in ([13], p. 51, Th. 40). −λn −s Theorem 1. If the Dirichlet’s series f( s) =Σ=Σ aen alnn is summable (lk, ) for s = β and c > 0 , c > β , then

ci+∞ − k 1 Γ(ks +Γ1) ( ) ωωks−= ω ∑ann( l) ∫ fs( ) d.s (2) ln <ω 2πici−∞ Γ( ks ++1 )

By Lemma 1,

Zs′′( ) Zs( +1) −s −=Λ∑ 1 (P)Tr ( WP( )) NP( ) . Zs( ) Zs( +1) P∈Γ h

We have,

ci+∞ 1 Zs′′( ) Zs( +1) −1 −−11s ∫ −s( s ++1d) ( s k) xs 2πici−∞ Zs( ) Zs( +1) −s 11ci+∞NP( ) Γ( k +Γ1) ( s) = Λ ∫ ∑ 1 (P)Tr ( WP( )) ds . ki!2π P∈Γ xΓ( k ++1 s) ci−∞h 

−s NP( ) = Λ Therefore, substituting ω = 1, fs( ) ∑ P∈Γ 1 ( P)Tr ( WP( )), and hence h x NP( ) a= Λ ( P)Tr ( WP( )) , ln = in (2), we get n 1 x

ci+∞ 1 Zs′′( ) Zs( +1) −1 −−11s ∫ −s( s ++1d) ( s k) xs 2πici−∞ Zs( ) Zs( +1) −s 1 NP( ) =Λ−∑ 1 (P)Tr( WP( )) 1 kx! NP( ) ≤1  x −s 1 NP( ) =Λ−∑ 1 (P)Tr( WP( )) 1 . kx! NP( )≤ x 

906 M. Avdispahić, Dž. Gušić

Then,

ci+∞ 1 Zs′( ) − −−11 ∫ s1 ( s++1d) ( s k) xss 2πici−∞ Zs( ) k 1 NP( ) =Λ−∑ 1 (P)Tr( WP( )) 1 (3) kx! NP( )≤ x  ci+∞ 1 Zs′( +1) −−11 +∫ s−1 ( s++1) ( s k) xssd. 2πici−∞ Zs( +1)

Now, put

ψ 01( xW, ) =∑ Λ ( P)Tr ( W( P)) NP( )≤ x and x ψψ= jj( xW,) ∫ −1 ( tW ,d) t 0 for j = 1, 2,. Using ([14], p. 12, Th. 1.3.5), it is easy to get that

1 j ψ j ( xW, ) =Λ−∑ 1 (P)Tr ( W( P))( x N( P)) . (4) j! NP( )≤ x

1 11 1 For 0 <<λ , let s=−=− ir i λ −, nK= 1, 2, , , be the zeros of Zs( ) n 4 nn22 n 4 1 in ,1 . Let ρk , kM= 0,1, , e denote all zeros of the hyperbolic scattering de- 2 1 terminant in  ,1 . 2  1 1 Assume T > 2 , Tlr±≠n , l ∈  , where snn= + ir and snr= − ir for 2 2 1 1 rinn=−−λ , λn > . Following ([5], p. 468), we may also assume Tl±≠γ , l ∈  , 4 4 where, 1− ρ , 1− ρ are the zeros of the Selberg zeta function Zs( ) for each zero 1 3 ρ=++ ηγi , η ≥ 0 , γ > 0 , of the hyperbolic scattering determinant φ. Let A ≥ 2 0 2 1 m be a large constant such that A0 ∉  , A0 +∉ , A0 ±∉ . We put AA=0 +1 . 2 2 Without loss of generality we may assume that c =1 + ε , ε > 0 , x ≥ 2 . Let R( AT,) =−[ Ac ,,] ×−[ T T] . By the Cauchy residue theorem one has c+ iT 1 Zs′( ) −−11 ∫ s−1 ( s++1d) ( s k) xss 2πic− iT Zs( ) 11 11 −+εε ++ +−εε −− −+ iT iT iT iT −− 1 A00 iT 22 c+ iT 22 A iT −−A iT =∫∫∫∫∫∫∫∫ +++++++ (5) 2πi −+ −+ 11− 11−− A iT A00 iT −+εεiT ++ iT c iT +−εεiT −− iT A iT 22 22 −+A iT 1 Zs′( ) −−11 ++ −1 ++s ∫ ∑ Ressz= s (s1) ( sk) x 2πi−−A iT z∈ R( AT, ) Zs( )

907 M. Avdispahić, Dž. Gušić

and

c+ iT 1 Zs′( +1) −−11 ∫ s−1 ( s++1d) ( sk) xss 2πic− iT Zs( +1)

11 11 −+εε ++ +−εε −− −+ iT iT iT iT −− 1 A00 iT 22 c+ iT 22 A iT −−A iT =∫∫∫∫∫∫∫∫ +++++++ (6) 2πi −+A iT −+ A iT 11 c− iT 11−−A iT 00−+εεiT ++ iT +−εεiT −− iT 22 22

−+A iT 1 Z′(s +1) −−11 ++ −1 ++s ∫ ∑ Ressz= s( s1.) ( sk) x 2πi −−A iT z∈ R( AT, ) Zs( +1)

Arguing as in [5] (p. 474) and [4] (pp. 105-108), we easily find that the sum of the x1+ε first eight integrals on the right hand side of (5) is O k . Similarly, taking into ac- εT Zs′( ) count that is bounded for Re(s) > 1, we obtain that the sum of the first eight Zs( ) x1+ε integrals on the right hand side of (6) is O k . Following [5] (p. 474) and [4] (p. 85, εT Prop. 5.7), we obtain that the ninth resp. the third integral on the right hand side of (5) resp. (6) are Ox( − A ) . Now, if we take k = 2 , (5) and (6) will give us

1 c+ iT Zs′( ) xs ∫ ds 2πic− iT Zs( ) ss( ++12)( s ) (7) xx1+ε Zs′( ) s =++Ox− A O Res ( ) 2 ∑ sz=  εT z∈ R( AT, ) Zs( ) ss( ++12)( s )

and

1 c+ iT Zs′( +1) xs ∫ ds 2πic− iT Zs( +1) ss( ++ 12)( s ) (8) xx1+ε Zs′( +1) s =++Ox− A O Res . ( ) 2 ∑ sz=  εT z∈ R( AT, ) Zs( +1) ss( ++ 12)( s )

Zs′( ) Bearing in mind location of the poles of given in ([5], p. 439, Th. 2.16; or [5], Zs( ) p. 498, Th. 5.3) and the fact that m ≤ 1 , we may assume without loss of generality that

mm51 mm 5 −−1 ,1 −+ ∈− , − ,3 −− ,3 −+ ∉−A , − . 2 2 42  2 2  2

Calculating residues and passing to the limit TA→ +∞, → +∞ in (7) and (8) we get

908 M. Avdispahić, Dž. Gušić

+∞ −ρ 11ci Zs′( ) xs  KK xxsnns Me x1 k =+ ++ ∫ dsx ∑∑∑ 2πici−∞ Zs( ) ss( ++12) ( s ) 6 nnk=1 ssnn( ++12) ( s n ) =1 ssnn( ++ 12) ( s  n ) =0 ( 1 −ρρρk) ( 2 − k) ( 3 − k) m mm − −−1 x2 xx 22 ++AA + B 01mm  m   mmm    0 mmm    − −+1  −+ 2   −− 1  −  −+1   + 12  + 22  2   222    222    m −+1 11 2 − x 8 1822* *1−− *1 +B1 +Tr Ir −Φ x − nx1 − gnx 11 − nx 1 log x mmm    15 2 3 −+11   +  222     ** ssnn 3 nn11   xx +− +hx1 +log  +∑∑+ (9)   ++ ++ 2 22   rrnn>0 ssnn( 12) ( s n ) >0 ssnn( 12) ( s  n ) 

1 1  −−ηγi −+ηγi  x 2 x 2 + ∑ + ∑  γ >0 1 35  γ >0  135     −−ηγi −−ηγii  −− ηγ   −+ ηγ iii  −+ ηγ  −+ ηγ  2 22    222   

mm −− −+ 2235 * * 22−− nn xx88**22 31 −21 −2 +A2 +B2 +nx11 − nx ++ f 1 x +xxlog m  m  m   m  mm  3 15 2 2 2 −−21  −−  −   −+ 21  −+  2  2  2   2  22  and

 m −− +∞  − − 1 1 ci Zs′( +1) xs KK xxsnn11s x2 = ++ ∫ dsA∑∑ 0 2πi−∞ Zs( +1) ss( ++ 12) ( s )  nn=11(s−+11) ss( ) = ( s −+ 11) ss ( ) m  mm   ci  n nn n nn −−11  −  −+   2  22   m −+ 1 1 −ρ xx2 81 − Me k 2 ′ *1−− *1 +B0 −Tr Ixr −Φ +∑ −−gnx11 nx 1 log x mmm  32 k =0 (−−ρρ) ( +12) ( −+ρ ) −+11  +  k kk 222  

  1 −−−ηγi   ssnn−−11 2 3′ 11  xx x +− +hx1 +log  +∑∑∑+ + 2 22   rr>0(s−+11) ss( ) >> 00( s −+ 11) ss ( ) γ 113    (10)   nnn nn n nn −−−ηγiii  −− ηγ  −− ηγ    222   

1  m −−+ηγi −−2 2  2 x  x + ∑ + A1 γ >0 11  3  m  mm   −−+ηγii −+ ηγ −+ηγi    −−−−−21    22 2  2  22  

m  −+ 2 ** 35 x 2 3nn 88−− ′ 11−−22 *22 * +B1 ++f1 x + xlog x + nx11 − nx . m  mm  2 2 2 3 15  −+21  −+   2  22   The implied constants on the right sides of (9) and (10) depend solely on Γ , m and W. With k = 2 in (3), j = 2 in (4), Equations (4), (3), (9) and (10) yield

909 M. Avdispahić, Dž. Gušić

1 2 ψ 21( xW, ) =Λ−∑ (P) Tr ( W( P)) ( x N( P)) 2! NP( )≤ x 1 ci+∞Zs′′( ) xxss++22ci+∞Zs( +1) = ∫∫ddss− 2πici−∞ Zs( ) ss( ++12) ( s) ci−∞ Zs( + 1) ss( ++ 12) ( s ) ++ −ρ 1  KKxxssnn22 Me x 3 k =3 + ++ x  ∑∑∑ 6  nnk=110ssnn( ++12) ( s n ) = ssnn( ++ 12) ( s  n ) = ( 1 −ρρρk) ( 2 − k) ( 3 − k) m mm −+2 12−+ x 2 xx22 + A ++AB 0 mmm 10  mmm    mmm    − −+1 −+2   −− 1  −  −+1   + 12  +  222   222    222    m 1+ 2 53 * x 8 1822** 3 n1 2 +B1 +Tr Ir −Φ x − nx1 − gnx 11 + − + h1  x mmm    15 2 3 2  2 −+11   +  222    m −+1 * K sn +11K sn + 2 n1 2 x xx +−xxlog ∑ −−∑ A0 2 n=1 (s−+11) ss( ) n=1 ( s −+ 11) ss ( ) m  mm   n nn n nn −−11  −  −+  2  22   m + 1 3 −+ρ 2  Me k 2 xx8 112 2 −B0 +Tr Ir −Φ x − xxlog −∑ mmm  3 22 k =0 (−ρρ) ( −+12) ( −+ ρ) −+11  +  kk k 222   (11)   ssnn++22 ′′*231 xx +gnx11 −− + h1 x  +∑∑+  ++ ++ 22  rn >>00ssnn( 12) ( s n ) rn ssnn( 12) ( s  n )  55 −−ηγii−+ηγ xx22 ++∑∑ γγ>>00135     135    −−ηγηγηγiii  −−  −−   −+ ηγηγηγ iii  −+  −+  222     222    3 −−ηγi xxssnn++11 x 2 −−−∑∑∑ rr>>00(s−+11) ss( ) ( s −+ 11) ss ( ) γ >0 113    nnn nn n nn −−−ηγiii  −− ηγ  −− ηγ  222   

3  m −+ηγi − 22 xx −+∑ A2 γ >0 113    m  mm   −−+ηγiii  −+ ηγ  −+ ηγ    −−21  −−  −  222    222    m m − 2 * * 2 x 3 n1 n1 x +Bf21+++−log xA1 m  mm  222 m  mm   −+21  −+  −−−−−21    2  22  2  22  

m  **  2 nn x 3 ′ 11  −B11−+fx −log m  mm  2 22  −+21  −+   2  22   1 =x3 +++∑∑∑, 6 I II III

910 M. Avdispahić, Dž. Gušić

where the first sum ranges over the finite set of poles s of Zs′′( ) xxssZs( +1) , Zs( ) ss( ++12)( s) Zs( + 1) ss( ++ 12)( s ) 5 with Re(s) >− , Im(s) = 0 , the second sum ranges over the set of poles s of the 4 same functions with Im(s) > 0 , and the third sum ranges over the finite set of their 5 poles s with Re(s) <− . 4

5. Prime Geodesic Theorem

In our setting, the prime geodesic counting function is defined by

π 00( xW,) = ∑ Tr( W( P)) , x≥ 1, NP( 0 )≤ x where the sum on the right is taken over all primitive hyperbolic classes P0h∈ΓP with respect to Γ (see, [5], p. 473, [11], p. 13). Theorem 2. For x ≥ 2 , the formula

K 3 −1 π =sn + 4 0 ( x, W) ∑ li( x) O x(log x) n=0 

11 1 holds true, where, s =+−λ for 0 ≤<λ , and the implied constant de- nn24 n 4 pends solely on Γ , m and W. Proof. Following [6] (p. 245) and [15] (p. 11), for a positive number d > 0 , we de- + fine the second difference operator ∆2 by xd+ yd+ ∆=+ ′′ 2 f( x) ∫∫ f( t)dd ty . (12) xy

Here, d is a constant which will be fixed later. By the mean value theorem, we have

+−θθ22 ∆=2 xdθθ( −1) x (13) for some x ∈+[ xx,2 d] . It is easy to verify that + ∆2 fx( ) = fx( +22 d) − fx( ++ d) fx( ) . (14)

Reasoning as in [5] (p. 475), we may assume without loss of generality that ψ 0 ( xW, ) is non-decreasing. Hence, (12) implies

−+2 ψ0( xW, ) ≤∆ d22 ψψ( xW,) ≤0( x + 2, dW) . (15)

+++ Since (14) holds true, one can easily deduce that ∆=∆22CfxC( ) fx( ) , ∆=2C 0 , 13   ∆=+ −+22∆= −+22∆=44 = −+2∆= 2 x 0 , d2 xO(1) , d2 xlog x Ox log x  Ox , d2 log xO( 1) ,   

−+231 d∆=+2  x x Od( ) . Thus, (13) and finiteness of the sums contained in ∑ on 6 I the right hand side (11) yield

911 M. Avdispahić, Dž. Gušić

1 K xsn 3 −+23∆ +=++ + 4 d2  x∑∑ x Od( ) O x . (16) 6 I1n= sn  Similarly,

3 − −+2∆= 2 d2 ∑ Ox. (17) III  −+2∆ In order to estimate d 2 ∑ II , we will first consider

sn +2 −+2 x d ∆2 ∑ . rn >0 ssnn( ++12)( s n )

By (14) it is evident that

sn +2 5 x −3 −+22∆=− 2 d 2 Od sn x . (18) ssnn( ++12)( s n )  On the other hand, the mean value theorem (13) gives us

sn +2 1 x −1 −+2∆=2 d 2 Osn x . (19) ssnn( ++12)( s n )  1 Let Nt( ) be the number of roots of Zs( ) on the critical line + ix in the inter- 2 r ℑ val 0 <≤xt. It is known ([5], p. 477, Th. 3.8) that Nt( ) ~ t2 . Taking M > 1 4π

and following ([3], pp. 463-464; [6], p. 246), we use (19) resp. (18) in the sums over sn , 1 < (below) to get 2 n n n n

sn +2 −+2 x d ∆2 ∑ rn >0 ssnn( ++12)( s n )

ssnn++22 −+22xx−+ ≤∆∑∑dd22 + ∆ 1 ss++12 s ≤< ss++12 s <

sn +2 −+2 x +∆∑ d 2 sMn > ssnn( ++12)( s n )

11 5 −− − 2211−2 2 3 ≤+Cx11∑∑ snn Cx s + Cd 2 x ∑ s n 1 ≤< > <

Thus,

sn +2 11 5 −+2 x −21 −  d∆ =++OxOMxOdxM22 2 . (20) 2 ∑ ++   rn >0 ssnn( 12)( s n )  

912 M. Avdispahić, Dž. Gušić

Similarly,

sn +2 11 5 −+2 x −21 −  d∆ =++OxOMxOdxM22 2 . (21) 2 ∑ ++   rn >0 ssnn( 12)( s  n )   = 2 Observe that ∑ γ ≤t1 Ot( ) (see, [5], p. 437, Prop. 2.13). Thus, application of −+2∆ d 2 to the third and the fourth sum in ∑ II gives us 11 5  22−−21 2 OxOMxOdxM++ .   Let us write ∑∑∑= − , (22) II 1 2 where ∑1 denotes the sum of the first four sums in ∑ II and ∑ 2 denotes the sum of the last four sums in ∑ II . Now, Equations (11), (16), (17), (20), (21) and (22) give us

K sn 31 −+ x    2∆ψ =+++42 + d22( x, W) x O( d) ∑ O x O  Mx  n=1 sn    (23) 5 +−2122 − −∆ −+ O d xM d 2 ∑. 2

1 3 Putting Mx= 4 , dx= 4 , the Equation (23) becomes

K xsn 3 −+22∆ψ =+ +4 −∆−+ d22( xW,.) x∑∑ O x d 2 (24) n=12sn  3 4 Since the left sides of Equations (20), (21) are Ox for such choice of M and d,  3 1  − d−+2∆= Ox4 d−+2∆= Ox4 we get 2 ∑1 . Now, it is obvious that 2 ∑1 . Finally, Equ-   ation (24) gives us

K sn 3 −+ x  2∆ψ =++4 d22( xW,.) x∑ O x n=1 sn  Returning to (15), we conclude that inequality K xsn 3 ψ ≤+ + 4 0 ( xW, ) x∑ O x n=1 sn  holds true. Following ([15], p. 11), we analogously obtain that K xsn 3 ++4 ≤ψ x∑ O x0 ( xW,.) n=1 sn  Hence, K xsn 3 ψ =++4 0 ( xW,.) x∑ O x (25) n=1 sn 

913 M. Avdispahić, Dž. Gušić

Arguing as in [5] (p. 475) and [4] (p. 113), one immediately sees that equality (25) proves the theorem.

Acknowledgements

We thank the Editor and the referee for their comments.

References [1] Selberg, A. (1956) Harmonic Analysis and Discontinuous Groups in Weakly Symmetric Riemannian Spaces with Applications to Dirichlet Series. Journal of the Indian Mathemati- cal Society, 20, 47-87. [2] Huber, H. (1961) Zur analytischen Theorie hyperbolischer Raumformen und Bewe- gungsgrupen II. Mathematische Annalen, 142, 385-398. https:/doi.org/10.1007/BF01451031 [3] Huber, H. (1961) Nachtrag zu. Mathematische Annalen, 143, 463-464. https:/doi.org/10.1007/BF01470758

[4] Hejhal, D. (1973) The Selberg Trace Formula for PSL( 2, ) , Vol. I. Lecture Notes in Ma- thematics, Volume 548. Springer-Verlag, Berlin-Heidelberg.

[5] Hejhal, D. (1983) The Selberg Trace Formula for PSL( 2, ) , Vol. II. Lecture Notes in Mathematics, Volume 1001. Springer-Verlag, Berlin-Heidelberg. [6] Randol, B. (1977) On the Asymptotic Distribution of Closed Geodesics on Compact Rie- mann Surfaces. Transactions of the American Mathematical Society, 233, 241-247. https:/doi.org/10.1090/S0002-9947-1977-0482582-9 [7] Buser, P. (1992) Geometry and Spectra of Compact Riemann Surfaces, Progress in Mathe- matics, Vol. 106. Birkhäuser, Boston-Basel-Berlin. [8] Avdispahić, M. and Smajlović, L. (2009) On the Prime Number Theorem for a Compact Riemmann Surface. Rocky Mountain Journal of Mathematics, 39, 1837-1845. https:/doi.org/10.1216/RMJ-2009-39-6-1837 [9] Avdispahić, M. and Smajlović, L. (2006) An explicit Formula and Its Application to the Selberg Trace Formula. Monatshefte für Mathematik, 147, 183-198. https:/doi.org/10.1007/s00605-005-0317-0 [10] Avdispahić, M. and Smajlović, L. (2008) Euler Constants for a Fuchsian Group of the First Kind. Acta Arithmetica, 131, 125-143. https:/doi.org/10.4064/aa131-2-2 [11] Avdispahić, M. and Smajlović, L. (2016) Selberg Trace Formula as an Explicit Formula and the Prime Geodesic Theorem. (Submitted) [12] Fischer, J. (19760 An Approach to the Selberg Trace Formula via Selberg Zeta-Function. Lecture Notes in Mathematics, Volume 1253. Springer-Verlag, Berlin-Heidelberg. [13] Hardy, G.H. and Riesz. M. (1915) The General Theory of Dirichlet’s Series. Cambridge University Press, Cambridge. [14] Jameson, G.J. (2003) The Prime Number Theorem. Cambridge University Press, Cambridge. https:/doi.org/10.1017/CBO9781139164986 [15] Park, J. (2010) Ruelle Zeta Function and Prime Geodesic Theorem for Hyperbolic Mani- folds with Cusps. In: van Dijk, G. and Wakayama, M., Eds., Casimir Force, Casimir Opera- tors and Riemann Hypothesis, de Gruyter, Berlin, 89-104.

914

Submit or recommend next manuscript to SCIRP and we will provide best service for you: Accepting pre-submission inquiries through Email, Facebook, LinkedIn, Twitter, etc. A wide selection of journals (inclusive of 9 subjects, more than 200 journals) Providing 24-hour high-quality service User-friendly online submission system Fair and swift peer-review system Efficient typesetting and proofreading procedure Display of the result of downloads and visits, as well as the number of cited articles Maximum dissemination of your research work Submit your manuscript at: http://papersubmission.scirp.org/ Or contact [email protected]