arXiv:1509.03863v1 [math.CA] 13 Sep 2015 ftesequences the of and ( a ovre to converges h alc-tete rnfr nScin6 hr eprov we where functions. 6, step s Section some under in result transform concrete a Laplace-Stieltjes give the we (where Dirichlet for (1) imninmnflssc that such manifolds Riemannian l fwihcneg boueyi omnhl plane half common a in absolutely converge which of all npyist s h alc pcrma aaee pc nc in space parameter as spectrum Laplace the ma use Riemannian to to physics application in The spectra. their of vergence yaia aibe ncasclgaiy nprdb part by inspired geometry gravity, classical in variables dynamical ..Notation. 2.1. t on th metric of a convergence from introduce deduced is similarly topology to we using up equivalence, fields, manifolds number Riemannian and of spaces ifolds on metric a introduce ( ..Theorem. 2.2. λ λ strictly ν γ ν e od n phrases. and words Key Date n plcto st imningoer.If geometry. Riemannian to is application One ups htw aepitiecnegneo euneo g of a of convergence pointwise have we that Suppose ) ) µ ν ∞ nTerm43 ecnie h aewhere case the consider we 4.3, Theorem In . =1 ∗ coe 7 08(eso 1.0). (version 2018 17, October : log = Abstract. plctost oeseta eafntosta rs nR in arise that functions ind zeta their spectral of some convergence to applications the implies series Dirichlet other sasqec fpstv elnmes ednt by denote we numbers, real positive of sequence a is nraigsqec fra ubr.I hsppr estudy we paper, this In numbers. real of sequence increasing [9] ( , a [3] λ n ups that Suppose e soc n o l nrdc h olwn ovnetno convenient following the introduce all for and once us Let ∗ ) edsusti reyi eto ,weew hwhwt s s use to how show we where 3, Section in briefly this discuss We . n nTerm22 epoeta 1 seuvln othe to equivalent is (1) that prove we 2.2, Theorem in and , ( h otgnrlcs ssuidi eto rmtepito point the from 5 Section in studied is case general most The . egv oecniin ne hc uiom ovrec o convergence (uniform) which under conditions some give We a n,ν OVREC FFMLE FDRCLTSERIES DIRICHLET OF FAMILIES OF CONVERGENCE ) iihe eis eafnto,Reana aiod spe manifold, Riemannian , zeta series, Dirichlet and UTE ONLSE N RSIE KONTOGEORGIS ARISTIDES AND CORNELISSEN GUNTHER D ( µ n ( n,ν s d = ) D ) =spdim sup := is,w osdrtecs fcasclDrcltseries, Dirichlet classical of case the consider we First, . n .Cnegneo iihe series Dirichlet of Convergence 2. ( ν X s ≥ = ) 1 a n,ν ν X ≥ .Introduction 1. 1 e X − λ n,ν − sµ n { s X n,ν sfiie ovrec fterzt ucin mle con- implies functions zeta their of convergence finite, is µ and n,ν n 1 } → n ∞ eana emtyadphysics. and geometry iemannian vda ofcet n/repnns egv some give We exponents. and/or coefficients ividual = =1 Re D cemdl ope ogaiyi noncommutative in gravity to coupled models icle D ( µ ∪{ ( m yohss,adfr h on fve of view of point the form and hypotheses), ome s ( s ν iod em eeati h ih fproposals of light the in relevant seems nifolds i eeidzt functions. zeta Dedekind eir s h qiaec oLpcizcnegneof convergence Lipschitz to equivalence the e := ) ) = ) Λ sidpnetof independent is X sooia vrgn problems averaging osmological γ > esaeo ubrfilsu oarithmetic to up fields number of space he setaiy yteaaoybtenman- between analogy the By ospectrality. s } h sequence the ν X ν X sasqec fcnetdcoe smooth closed connected of sequence a is nrlDrcltseries Dirichlet eneral ≥ with , ≥ tu,convergence. ctrum, 1 1 htti mle bu “convergence” about implies this what aiyo iihe eist an- to series Dirichlet of family a f λ a ν − ν e s a − ∗ sµ ope ofcet,and coefficients, complex ν ℓ , ( iwo h ernformula Perron the of view f ain if tation: λ 1 n cnegneof -convergence ν s hn o every for then, ; ) eta eafntosto functions zeta pectral ν ∞ =1 . s ∈ where C [12] and ( n λ , n as and , a n,ν γ ( ∗ a µ = Λ 1 = n,ν ) ∗ to is ) 2 G. CORNELISSEN AND A. KONTOGEORGIS

∞ is a family of (generalized) for n = , 1, 2,... , where, for each n, Λn := (λn,ν ) forms ∅ ν=1 a sequence of increasing positive real numbers with finite multiplicities. Assume that all series Dn(s) are convergent in a common right half plane Re(s) >γ> 0. Then the following are equivalent:

(i) As n + , the functions Dn(s) converge to D(s), pointwise in s with Re(s) >γ; → ∞ ∞ (ii) For every fixed ν, any bounded subsequence of λn,ν n=1 converges to the same element λ Λ, and { } ∈ ∞ # (λn,ν )n=1 : lim λn,ν = λ = # λν : λν = λ . { n→∞ } { } −γ −γ 1 (iii) Λn converges to Λ in ℓ .

2.3. Remark. The assumption that all series Dn(s) are convergent in a common right half plane Re(s) >γ is a minimal necessary assumption, since if this is not the case, the questions we ask are void.

2.4. Remark. The series Dn (when divergent at s = 0), converges for Re(s) > γn, where (see Chapter 1, Section 6 of [8]): log ν (2) γn = lim sup . ν→∞ log λn,ν

The hypothesis says that γ := sup γn is finite.

2.5. Remark. The spectral zeta function ζX of a closed smooth Riemannian manifold converges absolutely for Re(s) > d/2, where d is the dimension of X [11]. In the situation of Theorem 2.8, the assumption of a common half-plane of convergence hence follows from the fact that we assume that sup dim(Xn) is finite. Hence Theorem 2.8 follows from Theorem 2.2. Proof of Theorem 2.2. Since we later want to interchange some limits, we will first prove: 2.6. Lemma. The N 1 S (s) := n,N λs ν=1 n,ν X converge to Dn(s) for N + uniformly in n. → ∞ Proof of Lemma 2.6. The means that N 1 ǫ> 0, N0 1, N >N0, s Dn(s) ǫ ∀ ∃ ≥ ∀ λn,ν − ≤ ν=1 X where N0 does not depend on n. Now notice that by Equation (2) we have

log ν γn = lim sup ν→∞ log λn,ν and the series Dn(s) converge for Im(s) γn. The assumption that there is a common half plane of ≥ convergence for all Dn means that the sequence γn is bounded by γ = sup γn. The uniform convergence of Sn,N (s) follows by repeating the argument found in the proof on page 7 of [8]. The proof there uses γn for each n, but one may as well use (the same) γ for all n.  Next, we will show that unbounded subsequences do not contribute to the limit. For this, suppose that, for some fixed κ, (λnk ,κ)k∈N is an unbouded subsequence, with

lim λnk ,κ = k→∞ ∞ FAMILIES OF DIRICHLET SERIES 3 and, by enlarging the subsequence if necessary, such that the sequence

(λn,κ)n∈(N−{nk :k∈N}) is either bounded or the empty set. Now since λnk ,κ λnk ,µ for µ κ, all sequences (λnk ,µ)k∈N for µ κ tend to infinity as well. Observe now that the series≤ ≥ ≥ ∞ 1 D≥κ(s) := nk λs ν=κ nk,ν X tends to the zero function as nk tends to infinity: ∞ ≥κ 1 lim Dnk (s)= lim s =0, nk→∞ nk →∞ λ ν=κ nk ,ν X for s real positive (hence for all s by ). In the above equation we were allowed to interchange the order of the limits (in nk and the summation of Dnk ) since the series converge uniformly in nk. Since we assume that Dn is a convergent sequence of functions, it has the same limit as its the subsequence Dnk . Since we have now proven that unbounded subsequences do not contribute to the limit, we can assume that λn,ν is bounded in n, for all ν, i.e.,

n N λn,ν cν . ∀ ∈ ≤ Then we can select a subsequence so that for all ν, the limit

lim λnk ,ν = ℓν k→∞ exists. Not to overload notation, we will momentarily relabel the convergent subsequence λnk ,ν as λn,ν . In particular, λn,1 converges to ℓ1. We will prove that λn,1 converges to λ1. Let us rewrite ∞ s 1 λn,1 Dn(s)= s λ λn,ν n,1 ν=1 ! X   and ∞ s 1 λ1 D(s)= s . λ λν 1 ν=1 ! X   We now assume that s is an integer s>γ. Since Dn(s) D(s) we have that → ∞ s λ1 ∞ s λν s λ1 ν=1 λ1 (3) lim = ∞  s n→∞ λn,1 P λn,1 ≤ λν limn ν=1   →∞ λn,ν   ν=1 X P   For the last inequality, we have used the fact that λn,ν > 0 and that the denominator is 1. Set ≥ λ ℓ := lim 1 . n→∞ λn,1 We now consider the limit as s (along the integers) in Equation (3), to find → ∞ s lim ℓ # λi = λ1 . s→∞ ≤ { } 4 G. CORNELISSEN AND A. KONTOGEORGIS

Now 1 if ℓ =1 lim ℓs =  0 if ℓ< 1 . s→∞   if ℓ> 1 ∞ Hence we find ℓ 1.   We also have the≤ inequality ∞ s λ1 s λν s λ1 ν=1 1 (4) ℓ = lim = ∞  s ∞ s . n→∞ λ P λn,1 ≥ n,1 lim ℓ1   n→∞ λn,ν ℓν ν=1 ν=1     In the inequality, we have used that we can interchange limitP and summationP in the denominator, by uniform convergence. By taking the limit s (along the integers) we arrive at → ∞ 1 lim ℓs > 0. s→∞ ≥ # ℓn = ℓ { 1} We conclude from all the above that ℓ =1, and hence that

# λi = λ 1= { 1}. # ℓn = ℓ { 1} Now recall that we have relabelled before, so that we have actually shown that every convergentsubsequence

(λnk ,1)k∈N of (λn,1)n∈N tend to some limit, and since ℓ = 1 all these subsequences converge to the same limit λ1. Therefore (λn)n∈N itself is convergent to λ1. We conclude that in general (viz., before erasing all unbounded subsequences), that every bounded subsequence of (λn)n∈N converges to λ1. We now use an inductive argument to treat the general term. Namely, consider the Dirichlet series ≥2 −s D (s) := Dn(s) λ n − n,1 which (by what we have proven) converges to D≥2(s) := D(s) λ−s. − 1 These are still sequences of Dirichlet series of the same form, but with first eigenvalues λn,2 and λ2. We can repeat the argument with this series, to conclude λn,2 λ2, etc. This finishes the proof that (i) implies (ii). → Since we assume that Re(s) >γ is a commonhalf plane of convergenceof all series Dn (n = , 1, 2,... ), the sums ∞ λ−γ converge, and hence the sequences Λ−γ (n = , 1, 2,... ) belong to the Banach∅ space ν=1 n,ν n ∅ ℓ1. We will now prove that Λ−γ Λ−γ as elements of ℓ1. P n In order to do so we have to prove→ that for every ǫ> 0 there is an n N such that n>n implies 0 ∈ 0 ∞ 1 1 γ γ ǫ λn,ν − λν ≤ ν=1 X ∞ It is known that if (aν ) is a sequence of positive real numbers so that ν=1 aν converges, then all of its “tails” tend to zero: ∞ P lim aν =0. N→∞ ν N X= FAMILIES OF DIRICHLET SERIES 5

So given an ǫ > 0 there is an n0, which does not depend on n (using the same γ for all n), such that for N n0 ≥ ∞ ∞ 1 1 γ + γ < ǫ/2. λn,ν λν ν=N ν=N X X Therefore,

∞ N ∞ ∞ 1 1 1 1 1 1 γ γ γ γ + γ + γ λn,ν − λν ≤ λn,ν − λν λn,ν λν ν=1 ν=1 ν=N ν=N X X X X N γ γ λν λn,ν ǫ | γ− γ | + ≤ λn,ν λν 2 ν=1 X N λγ λγ ǫ (5) | ν − n,ν | + , ≤ C 2 ν=1 X where γ γ γ γ 0 = C = inf (λ1,ν λν ) (λn,ν λν ). 6 1≤ν≤N ≤ γ γ Now the finite number (ν = 1,...,N) of sequences (λn,ν )n∈N can be made to uniformly converge to λν , that is, for every ǫ> 0 there is an n1 such that n>n1 implies ǫC λγ λγ | ν − n,ν |≤ 2N and inequality (5) gives us the desired result for all n max n0,n1 . −γ≥ { } −γ This proves that (ii) implies (iii). Finally, if Λn converges to Λ , we have for every s C with Re −s −s ∈ (s) > γ that Λn converges to Λ , and it follows easily that Dn(s) converges to D(s), pointwise in s. This proves that (iii) implies (i) and finishes the proof of the theorem. 

2.7. Application. Suppose X = (X,gX ) and Y = (Y,gY ) are two isospectral connected smooth closed Riemannian manifolds, i.e., suppose their Laplace-Beltrami operators ∆X and ∆Y have the same spectrum ∞ with multiplicities ΛX = ΛY [11]. The spectrum ΛX is considered as a sequence (λν ) with 0 λ ν=1 ≤ 1 ≤ λ2 ... , with finite repetitions. The≤ identity theorem for Dirichlet series [8] shows that such isospectrality can also be described as the manifolds having the same zeta function ζX = ζY , where 1 ζ (s) := tr(∆−s)= , X X λs λ 06=X∈ΛX since connectedness implies that the zero eigenvalue has multiplicity one. In this context, Theorem 2.2 says the following, which is a “convergent” version of the identity theorem: ∞ 2.8. Proposition. Suppose Xn is a sequence of connected closed smooth Riemannian manifolds such { }n=1 that d := sup dim Xn is finite, and suppose that X is another closed smooth Riemannian manifold. Then the following statements are equivalent Re (i) For (s) > d/2, the functions ζXn (s) converge pointwise to ζX ; (ii) For some γ C with Re(γ) > d/2, the sequence of eigenvalues Λ−γ converges to Λ−γ in ℓ1. ∈ Xn X 2.9. Remark. If the manifolds are closed and smooth and of odd dimension, but possibly disconnected, the equality ζX = ζY implies that also the multiplicity of the zero eigenvalue is equal for X and Y , namely, it is minus the value at 0 of the analytic continuation of ζX ([11], 5.2). 6 G. CORNELISSEN AND A. KONTOGEORGIS

2 2 2.10. Example. The circle of radius r has ∆ = r ∂θ (with θ [0, 2π) the angle coordinate), spectrum −2 2 2s− ∈ λr,ν = r ν/2 and zeta function ζr(s) = r ζ(2s), where ζ is the Riemann zeta funtion. For varying r r , the⌈ convergence⌉ in the theorem happens for γ > 1/2. → 0 2.11. Remark. Already in the case of families of Riemannian manifolds, it can happen that (λn,κ) has unbounded subsequences for some fixed κ; for example, a family of circles whose radius tends to zero. However, for fixed κ, we have bounds on the eigenvalues of the form ([2])

d 2 C2 d 2 C1 √κ λn,κ √κ , ≤ ≤ vol(Xn) where the constants Ci depend on the dimension d, the diameter D, and a lower bound R on the Ricci curvature of the manifolds under consideration. This implies that (at least if we fix the data d, D and R, so we are in the Gromov precompact moduli space [7]) in unbounded subsequences, the volume should shrink to zero.

3. Applications: metric theories derived from Dirichlet series Distances in cosmology. In connection with the averaging problem in cosmology and the question of topol- ogy change under evolution of the universe, Seriu [12] proposes to use eigenvalues of the Laplace-Beltrami operator on spatial sections of a cosmological model to construct a metric on the space of such Riemannian manifolds up to some notion of “large scale isospectrality”. More precisely, he raises two objections against the use of the plain difference of spectra as a measure: large energy contributions (corresponding to small scale geometry) should carry a lower weight, and the dominant weight should be put on the small spectrum (corresponding to large scale geometry); therefore, he introduces a cut-off N and only compares the first N eigenvalues. Secondly, the eigenvalue difference is not a dimensionless quantity, and because of this, he suggests comparing quotients of spectra. However, as N + , his distance diverges. From the above theorem, it also appears natural not to→ use∞ a cut-off function, but rather use a distance between the zeta functions (which, like partition functions, give more weight to low energy in their region of convergence), considered as complex functions; here, one may use classical notions of distance between complex functions [4] used in the study of limits of holomorphic or meromorphic functions. Also, the quo- tient of two zeta functions is a dimensionless function. Actually, a distance between Riemannian manifolds up to isometry was constructed by the first author and de Jong, who have furthermore given a spectral char- acterization of when a diffeomorphism of closed smooth Riemannian manifolds is an isometry, in terms of equality of more general zeta functions under pullback by the map [5]. Also this distance is based on the dimensionless object of quotients of zeta functions. In conclusion, we propose the following function as a distance on suitable spaces of Riemannian geome- tries up to isospectrality: 3.1. Proposition. Let denote a space of Riemannian manifolds up to isospectrality, with M sup dim(X): X < 2γ { ∈ M} finite. Then for any X ,X , the function 1 2 ∈M ζ (s) d(X ,X ) := sup log X1 1 2 ζ (s) γ γ X1 X2 metric on . M FAMILIES OF DIRICHLET SERIES 7

Proof. The function d is positive, and if d(X1,X2)=0, then ζX1 (s) = ζX2 (s) for all s in the interval ]γ,γ + 1[. Since this set has accumulation points, and since the| zeta function| | is positive| real for such values of s, we find that ζX1 = ζX2 as complex functions. Hence the main theorem (or the identity theorem for −1 Dirichlet series) implies that X1 and X2 are isospectral. The function d is symmetric, since log(x ) = log(x) . Finally, the triangle inequality follows from | |

ζ (s) ζ (s) ζ (s) X1 = X1 X2 ζX3 (s) ζX2 (s) · ζX3 (s) and the usual properties of the absolute value. 

This is a distance that weighs correctly the energy contributions, but does not depend on a cut-off, nor diverges if a cut-off tends to infinity. Also, convergence in the topology defined by this distance can be easily understood from Theorem 2.2. 3.2. Remark. Taking the supremum over γ

3.3. Example. If Sr denotes a circle of radius r, then

d(Sr ,Sr )=4 log(r /r ) . 1 2 | 1 2 | This example shows that the distance can be non-differentiable in the parameter space of a family. 3.4. Example. Let us compute the spectral distance between a sphere S and a real projective space RP2 with the same volume 4π. The zeta functions are ∞ 2ν +1 ∞ 4ν +1 ζ = and ζRP2 = . S νs(ν + 1)s νs(2ν + 1)s ν=1 ν=1 X X A numerical experiment suggests that the maximum in the distance formula is attained at s = 2, and there we get d(S, RP2) = log(4 π2/3) 0.342. − ≈ Eigenvalues as dynamical variables. Gravity coupled to matter can be given a spectral description using the framework of noncommutative geometry [3]. Even by ignoring the matter part, one arives at an inter- esting description of classical gravity (general relativity) in terms of spectral data. These spectra form a diffeomorphism invariant set of coordinates on the space of manifolds, up to isospectrality. Diffeomorphism invariant coordinates are an important prerequisite for certain programmes to quantize gravity. In this way, spectra were used as dynamical variables for classical gravity by Landi and Rovelli [9]. Our theorem shows that convergence in these spacetime variables is the same as convergence of classical Dirichlet series in .

A distance on Number Fields. Consider the Dedekind-zeta function for a number field K as ζK (s) := −s N(I) , where the sum runs over all non-zero ideals I of K , the of K, and N(I) is the norm of the I. For any fixed a> 0, we can define a distanceO P ζ (s) d(K ,K ) := sup log K1 . 1 2 ζ (s) 1

8 G. CORNELISSEN AND A. KONTOGEORGIS

For a prime number p and a positive integer f, let IK (p,f) denote the number of ideals of K with norm f O p . Since ζK (s) admits an 1 ζ (s)= , K 1 N(P )−s P ∈SpecY OK − we find

−fs d(K1,K2) = sup (IK1 (p,f) IK2 (p,f)) log 1 p . 1

3.5. Example. Let us estimate the distance between the field of rational numbers Q and a real quadratic

field Q(√D) for D> 0. Let I denote the primes inert in Q(√D), S the set of split primes and R the set of ramified primes, i.e., the divisors of the squarefree part of D. Using the Euler products, we find

d(Q, Q(√D) = sup log L(χD,s) 1

Q As D 1, the distance goes to zero. If Di denotes the product of the first i prime numbers, then R → increases to the set of all primes, so then lim d(Q, Q(√Di))=0, too. We verified numerically in SAGE i→+∞ that L(χD , 1) 1 as i + ; The figure is a plot of L(χD , 1) as a function of i. i → → ∞ | i |

1.2

1.15

1.1

1.05

1

0.95

0.9

0.85

0 50 100 150 200

Figure 1. L(χD , 1) a a function of i | i |

4. Series with general coefficients In this section, we study what happens if we have pointwise convergence of general Dirichlet series

−sµn,ν −sµν Dn(s)= an,ν e D(s)= aν e , → ν ν X≥1 X≥1 FAMILIES OF DIRICHLET SERIES 9 all of which converge absolutely in a common half plane Re(s) > γ, with a∗ complex coefficients, and µ∗ is a strictly increasing sequence of real numbers. The previous case occurs when µ = log λ and a { ∗} { ∗} ∗ counts the multiplicities in (λ∗). In this paper, we will not discuss subtleties that arise from such series that have a different region of convergence and . We start by discussing two special cases.

−s 4.1 (Taylor series). The first is when µn,ν = log ν for all n = , 1, 2,... . In this case, we set z = e and we get a (pointwise) convergence of Taylor series ∅

ν ν Dn(z)= an,ν z D(z)= aν z . → ν ν X≥0 X≥0 In this case, the individual series D converge in Ω := z>e−γ to a holomorphic function (by assump- ∗ { } tion). Evaluation at zero gives limn→+∞ an,0 = a0, and we can proceed by induction to conclude that

lim an,ν = aν n→+∞ for all ν. Alternatively, one can use the representation of the coefficients by a complex contour integral to deduce the result “in a more complicated way”. Namely, fix ǫ > 0, and let n satisfy that Dn(z) D(z) < ǫ 0 | − | for n>n0, uniformly in z K Ω, where K is a compact set. For a contour C K around z = 0 (independent of n), we have ∈ ⊂ ⊂

1 −n−1 an,ν aν Dn(z) D(z) z dz ε. | − |≤ 2π | − || | ≤ ZC 4.2 (Constant exponents). The reason for providing this second proof is that it leads us to the next special case, in which we use the analogue of the integral representation for the coefficients for general Dirichlet series, also called Perron’s formula. This formula gives a representation of the terms of a general Dirichlet series by integration over a vertical line in the , and since this integration domain, unlike the contour in the Taylor series proof, is not compact, we will need to work more to establish the result (or assume uniform convergence on an entire half-line, which seems too strong an assumption). This second special case occurs if µn,ν is constant in n. Then we have the following result:

4.3. Theorem. Assume that Dn (n = , 1, 2,... ) is a set of Dirichlet series that converge absolutely in ∅ a common half plane Re(s) > γ, and such that Dn(s) D(s) converges pointwise there. Assume that → µn,ν = µν is independent of n. Then for every n, we have

lim an,ν = aν ; n→+∞ actually, for σ1 >γ, we have a convergence of sequences

−σ1µν ∞ −σ1µν ∞ ∞ (an,ν e ) (aν e ) in ℓ . ν=1 → ν=1 Proof. Consider the difference

−sµν Bn(s) := Dn(s) D(s)= bn,ν e , − ν X≥1 where

bn,ν := an,ν aν . − 10 G. CORNELISSEN AND A. KONTOGEORGIS

Now according to Theorem I.3.1 in [10] we have the following integral representation for every n and every fixed ν: T −σ1µν 1 µν it bn,νe = lim Bn(σ1 + it)e dt T →∞ T 0 Z T 1 lim Bn(σ1 + it) dt ≤ T →∞ T | | Z0 We could finish the proof here by assuming that Dn converges uniformly to D on the entire line Re(s)= σ1. However, we can avoid this (strong) hypothesis by proving the following lemma: 4.4. Lemma. For every ǫ> 0 there is a t R such that for t R and all n, 0 ∈ ∈ Bn(σ + it) ǫ + Bn(σ + it ) . | 1 |≤ | 1 0 | We are then finished with the proof of Theorem 4.3, since now, given any ǫ > 0, the pointwise conver- gence at t0 implies that there exists n0 such that for all n>n0,

Bn(σ + it) ǫ + Bn(t ) 2ǫ | 1 |≤ | 0 |≤ and then the above inequality becomes −σ1µν bn,ν e 2ǫ, ≤ −σ1µν Since σ1 and ν are fixed, e is a non-zero constant, and this proves that bn,ν 0 as n + . Since the ǫ-bound holds uniformly in ν, we do find the ℓ∞ convergence as stated. → → ∞

Proof of Lemma 4.4. Since the series Dn are absolutely convergenton a common half plane, their sequences of tails tend to zero uniformly in n, that is, for every ǫ> 0 there is an N that is independent of n such that ∞ ∞ −sµν −sµν an,ν e + aν e < ǫ. | | | | ν N ν N =X+1 =X+1 Hence ≤N >N >N ≤N (6) Bn(s) B (s)+ D (s) D (s) B (s) + ǫ | |≤ n n − ≤ n Re We will now estimate the sum of the first N terms on a vertical line (s) = σ1. Consider the function N f : R S1 given by → t (eitµ1 ,...,eitµN )  7→ and the function N F : S1 C → sending  N −σ1µν (P ,...,PN ) e Pν . 1 7→ ν=1 X The function F is continuous on a compact set therefore it attains a maximal value M at a point A0 := 0 0 (P1 ,...,PN ). ∞ 4.5. Lemma. There exists t R such that all the numbers t µν are irrational. 0 ∈ { 0 }ν=1 Proof. The set of multiples bρ of a given ρ R such that bρ Q is just 1 Q and this set { } ∈ ∈ ρ is denumerable. A denumerable union of denumerable sets cannot exhaust the set of reals and the result follows.  FAMILIES OF DIRICHLET SERIES 11

N This proves that the set f(R) is dense in S1 . Therefore, for every δ > 0 there exists t R such that 0 ∈ f(t ) A0 δ, and hence, since F is continuous, | 0 − |≤  (7) B≤N (t ) M = F (f(t )) F (A0) < ǫ. | n 0 − | | 0 − | Since M is the maximum, for all t R, we have ∈ (8) B≤N (σ + it) M B≤N (σ + it ) + B≤N (σ + it ) M . | n 1 |≤ ≤ | n 1 0 | | n 1 0 − | By equations (6), (7) and (8) we now have ≤N ≤N ≤N ≤N Bn(σ +it) ǫ+ B (σ +it) ǫ+ B (σ +it ) + B (σ +it ) M 2ǫ+ B (σ +it ) , | 1 |≤ | n 1 |≤ | n 1 0 | | n 1 0 − |≤ | n 1 0 | and this finishes the proof of lemma 4.4. 

The proof of the lemma also completes the proof of the theorem.  4.6. Application. Let X denote a closed smooth Riemannian manifold and let a C∞(X) denote a smooth function. Define a generalized Dirichlet series by ∈ −s ζX,a := tr(a∆X ), cf. [5]. Then 1 ζX,a = aσX,λ, λs · λ X 06=X∈ΛX Z where 2 σX,λ := Ψ | | λ X⊣Ψ is the sum of the elements Ψ of an orthonormal basis of eigenfunctions that belong to the eigenvalue λ. In [5], it was proven that a diffeomorphism ϕ: Y X between closed Riemannian manifolds with simple → ∞ spectrum is an isometry precisely if ζX,a = ζY,ϕ∗(a) for all a C (X) (and there is also a version if the spectrum is not simple). ∈ Now assume that we have a compact manifold X and a family gr (r R) of isospectral metrics with { } ∈ simple eigenvalues on X (cf. Gordon and Wilson [6] for the existence of such families). Denote by Ψr,λ the normalized real eigenfunction for the metric gr corresponding to the eigenvalue λ. If all zeta functions converge in the sense that

(9) ζX,g ,a ζX,g ,a for all a r → s then we find from the above result that

2 2 aΨ dµr aΨ dµs r,λ → s,λ Z Z ∞ for all functions a C (X), where µr is the measure belonging to the metric gr. ∈ Taking residues at dim(X)/2 in (9) for a = 1, we find that the volume of X in gr is constant, and then taking residues for general a, we find that for all a C∞(X), ∈

adµr adµs. → Z Z Changing variables, we get that dµr a 1 dµs 0, − dµ → Z  s  12 G. CORNELISSEN AND A. KONTOGEORGIS and hence that dµr 1. From the above theorem, we conclude that Ψ2 dµr Ψ2 , and hence that dµs → r,λ dµs → s,λ Ψ2 Ψ2 , r,λ → s,λ a convergence of squared eigenfunctions.

5. General case Finally, in the most general case of varying coefficients and varying exponents, we prove a theorem about accumulation points. First, we do some preparation. 5.1. Definition. For a fixed strictly positive real function g, define for a real function f, the g-sup norm as f(x) f ∞,g := sup , || || R g(x) x∈ when it is defined. We say that a sequence of functions fn converges multiplicatively to a real function f if there exists a strictly positive real function g that is{ integrable } with respect to the multiplicative Haar measure on R∗ (i.e., such that g(x) dx < + ), such that R |x| ∞ R fn(x) f(x) ∞,g 0 || − || → for n + . → ∞ 5.2. Definition. For f a complex function defined for Re(s)= c, and x R, denote by ∈ c xs ds Ix(f) := f(s)e . Re s Z (s)=c The relevance of this integral for the theory of Dirichlet series lies in the following formula of Perron: if −sµν D(s)= aν e is convergent for s = β + iγ and c> 0, c>β, x R, x β, then ν≥1 ∈ ≥ 1 P a = Ic(D), n 2πi x λ x Xn≤ with the convention that the last summand on the left hand side is multiplied by 1/2 if x equals some λν . c Since Ix(D) does not depend on c once it satisfies the conditions for Perron’s formula, we will now write c Ix(D) for Ix(D) with any suitable c.

5.3. Lemma. If fn(c + it) converges multiplicatively to f(c + it) in t, then for all x R, { } ∈ lim Ix(fn)= Ix(lim fn)= Ix(f). n n Proof. We have

fn(c + it) f(c + it) x(c+it) Ix(fn) Ix(f) − e dt | − |≤ R c + it Z

fn(c + it) f(c + it) x(c+it) − e dt ≤ R c + it Z

cx g(t) e dt fn(c + it ) f(c + it) ∞,g ≤ R √c2 + t2 · || − || Z  cx g(t) e dt fn(c + it) f(c + it) ∞,g ≤ R t · || − || Z | |  Cε, ≤ FAMILIES OF DIRICHLET SERIES 13

cx g(t)  with C = e R |t| dt finite constant, for n sufficiently large. This proves the desired result.   Before statingR the main result of this section, we need to introduce some notation:

∞ (j) 5.4. Notation. Assume that all sequences (λn,j )n=1 are bounded. Let ℓi , i Ij be the accumulation ∞ ∈ points of sequence (λn,j )n=1. We consider a subsequence such that for all (j) for a selection . Notice nk j lim λnk ,j = ℓij ij Ij nk→∞ ∈ ∞ ∞ that the sequences (λnk ,j )k=1 and (λnk ,j+1)k=1 satisfy λnk,j < λnk ,j+1 but they can tend to the same accumulation point. For the infinite vector of convergentsequences (λ )∞ converging to the infinite vector ℓ(j) nk ,j k=1 j≥1 ij j≥1 we consider the sequence m ,m ,..., such that 1 2   (1) (2) (m1) (m1+1) (m1+2) (m2) ℓi = ℓi = = ℓi ,ℓi = ℓi = = ℓi , etc. 1 2 ··· m1 m1+1 m1+2 ··· m2

(λnk ,1) (λnk ,m1 ) (λnk ,m1+1) (λnk ,m2 ) ✉✉ rr ··· ✉✉ rr ✉✉ rrr ✉✉ rr  z✉✉  yr ℓ(1) ℓ(m1+1) i1 im1+1

−s log λj 5.5. Theorem. We use the notation of 5.4. Assume that Dn converges multiplicatively to D(s)= j≥1 aj e . ∞ Then, λ are accumulation points for some sequence (λ ′ ) . j n,j n=1 P Consider the set of subsequences (λ )∞ converging to the infinite vector ℓ(j) . Suppose n,j n=1 j≥1 ij j≥1 that the sequences (λ )∞ for j = m +1,...,m converge to ℓ. Set nk ,j n=1 µ  µ+1  mµ+1 (µ) A := an ,j , for µ 0. nk k ≥ j=m +1 Xµ Then,

(µ) ai if ℓ = λi, (10) lim An = k  0 otherwise.  Proof. Assume that the set of subsequences ∞ converges to the set of accumulation points (λnk,j )k=1 j≥1 (j) (ℓi ).  (j) Consider the first eigenvalue λ1 of D. If ℓ is the first element in the set ℓi that is smaller than λ1 then by m1 (0) choosing x such that ℓ

lim ank,j =0, nk→∞ j=m +1 X1 and so the desired result is proved for all ℓ < λ1. 14 G. CORNELISSEN AND A. KONTOGEORGIS

We will prove now that λ1 is an accumulation point. Indeed, for sufficiently small ǫ> 0 the quantity

Iλ ǫ(D) Iλ ǫ(D)= a =0. 1− − 1+ 1 6 Using the above equation and lemma 5.3 we obtain

lim Iλ1−ǫ(Dnk ) Iλ1+ǫ(Dnk ) = a1 = lim ank,j . nk→∞ − λ1−ǫ<λ <λ1+ǫ  Xnk,j So by taking small ǫ we can find a subsequence tending to λ1 so λ1 is one of the accumulation points of the sequence ∞ . Notice also that desired result of eq. (10) is also proved. (λnk ,j )k=1 j≥1 We continue the proof by induction by taking ℓ to be inside λ and λ so the corresponding sum tends to  1 2 zero, then we take ℓ to be λ2, then inside λ2 and λ3 etc. 

6. Relation with Laplace-Stieltjes Transform The notion of Dirichlet series and Laplace transforms can be unified in terms of the Riemann-Stieltjes integrals (Widder [13], compare [1]). 6.1. Definition. Suppose ω 0 is a real number. ≥ (1) The space Lip is defined as the set of functions F : R R with bounded norm ω ≥0 → F (t) F (s) F Lip,ω := sup | − ωt | < . || || sω k! ds ∞ k∈N

The main result is now that the so-called Laplace-Stieltjes transform ∞ F e−stdF (t) 7→ Z0 induces an isometric isomorphism Lipω Widω ([1], Thm. 2.4.1). → −sµν Widder ([13], Theorems 11.2 and 12.4) proved that a Dirichlet series of the form D(s) = ν aν e convergent for Re(s) >ω is in the space Widω. Also, such D is the Laplace-Stieltjes transform of P ∞ F (t)= aν H(t µν ), − ν=0 X where H is the Heaviside step function. Thus, we immediately conclude the following:

−sµn,ν 6.2. Theorem. Suppose Dn(s) = ν≥1 an,ν e is a sequence of Dirichlet series each converging absolutely in a common half plane Re(s) > γ; then for any ω>γ, Dn converges to a Dirichlet series −sµν P D(s)= ν≥1 aν e in Widω-norm if and only if P ∞ an,ν H(t µn,ν ) aν H(t µν ) 0 − − − → ν=0 X   in Lipω-norm. It would be interesting to deduce Theorem 4.3 and 5.5 from Theorem 6.2. FAMILIES OF DIRICHLET SERIES 15

References [1] Wolfgang Arendt, Charles J. K. Batty, Matthias Hieber, and Frank Neubrander. Vector-valued Laplace transforms and Cauchy problems, volume 96 of Monographs in Mathematics. Birkhäuser Verlag, Basel, 2001. [2] Pierre Bérard, Gérard Besson, and Sylvain Gallot. Embedding Riemannian manifolds by their heat kernel. Geom. Funct. Anal., 4(4):373–398, 1994. [3] Alain Connes. Gravity coupled with matter and the foundation of non-commutative geometry. Comm. Math. Phys., 182(1):155– 176, 1996. [4] John B. Conway. Functions of one complex variable, volume 11 of Graduate Texts in Mathematics. Springer-Verlag, New York, second edition, 1978. [5] Gunther Cornelissen and Jan Willem de Jong. The spectral length of a map between Riemannian manifolds. J. Noncommut. Geom., 6(4):721–748, 2012. [6] Carolyn S. Gordon and Edward N. Wilson. Continuous families of isospectral Riemannian metrics which are not locally isometric. J. Differential Geom., 47(3):504–529, 1997. [7] Mikhael Gromov. Metric structures for Riemannian and non-Riemannian spaces. Modern Birkhäuser Classics. Birkhäuser Boston Inc., Boston, MA, 2007. [8] Godfrey H. Hardy and Marcel Riesz. The general theory of Dirichlet’s series. Cambridge Tracts in Mathematics and Mathematical Physics, No. 18. Stechert-Hafner, Inc., New York, 1964. [9] Giovanni Landi and Carlo Rovelli. General relativity in terms of Dirac eigenvalues. Phys. Rev. Lett., 78(16):3051–3054, 1997. [10] Szolem Mandelbrojt. Séries de Dirichlet. Principes et méthodes, volume 11 of Monographies Internationales de Mathématiques Modernes. Gauthier-Villars, Paris, 1969. [11] Steven Rosenberg. The Laplacian on a Riemannian manifold, volume 31 of London Mathematical Society Student Texts. Cam- bridge University Press, Cambridge, 1997. [12] Masafumi Seriu. Spectral representation of the spacetime structure: the “distance” between universes with different topologies. Phys. Rev. D (3), 53(12):6902–6920, 1996. [13] David V. Widder. A generalization of Dirichlet’s series and of Laplace’s integrals by means of a Stieltjes integral. Trans. Amer. Math. Soc., 31(4):694–743, 1929.

Mathematisch Instituut, Universiteit Utrecht, Postbus 80.010, 3508 TA Utrecht, Nederland E-mail address: [email protected]

Department of Mathematics, National and Kapodistrian University of Athens, Panepistimioupolis, GR-157 84, Athens, Greece E-mail address: [email protected]