arXiv:2009.04955v2 [math.NT] 13 Sep 2021 cecs(IS.Tersac n nig a o eettho reflect not may findings and research The (PIMS). Sciences Shimur functions, theta Partial polynomials, Jacobi bers, k Since nadto,let addition, In fmc oua om hi osrcini oiae ywor by motivated is functions construction the Their that form. modular mock of n the and ls ftitdadrsrce esoso lsia divis classical of versions restricted and twisted of class r ooopi weight holomorphic are etn,Oo n h hr uhrfcse ntecase the on focussed author third the and Ono, Mertens, n o-rva iihe hrce fmodulus of character Dirichlet non-trivial any rmtv iihe hrce fmodulus of character Dirichlet primitive let OA AMNCMA OM N OOOPI PROJECTION HOLOMORPHIC AND FORMS MAAß HARMONIC POLAR ≥ h eerhcnutdb h rtato o hsppri su is paper this for author first the by conducted research The eetpprb etn,Oo n h hr uhr[ author third the and Ono, Mertens, by paper recent A e od n phrases. and words Key 2020 τ ,te aldtee“okmdlrEsnti eis.Foll series”. Eisenstein modular “mock these called they 2, = σ ahmtc ujc Classification. Subject n rv htteegnrt h ooopi ato oa h polar of part holomorphic the generate these that prove and oncino u osrcint eoopi aoiforms normali Jacobi two meromorphic first to construction their our and of sums connection Appell-Lerch of combination k s o oesc dniis oevr eprove we Moreover, between identity identities. an such more obtain for we ask characters, t of on choice conditions growth certain struc suitable the and assuming invariance without project translation but holomorphic polynomials, the Jacobi c compute of terms a essentially we construct end, we this To Here, divi functions. small theta by given Shimura are classical coefficients by Their series. Eisenstein Abstract. when u ml iio function divisor small − 1 + ( n iv p saFuirceceto h lsia ooopi Eisens holomorphic classical the of coefficient Fourier a is ) sa d rm.If prime. odd an is 1 2 ∈ L H (1 eety etn,Oo n h hr uhrsuidmc mo mock studied author third the and Ono, Mertens, Recently, and − OHAMLS NRA OO N AR ROLEN LARRY AND MONO, ANDREAS MALES, JOSHUA ,φ k, pelLrhsm,Hroi aßfrs ooopi proje Holomorphic forms, Maaß Harmonic sums, Appell-Lerch 1. k D q + ) oua om nΓ on forms modular n : nrdcinadsaeeto results of statement and Introduction e = : = n X 2 ≥ χ  πiτ σ 1 d snntiilw ert h eeaigfnto of function generating the rewrite we non-trivial is 1 sm   ,ψ h notation The . | X d n λ ( | n n ψ 13,1F0 11F03. 11F30, 11F37, ) 1 : σ φ : = : k ( = − d N ≤ 1 ) 1 d d ( X ∈ d − n satisfying k D − ) ≤ M n ψ 1 0 2 : = ψ   ( ( ψ 1 ht ucin ml iio function. divisor Small function, theta a n d N − p

ai ogune forsaldvsrfunctions divisor small our of congruences -adic > X q d 1) L ihNebentypus with ) and n | n ( , ,dfietesto disbe“ml”divisors “small” admissible of set the define 1, n d ,φ s, pre ytePcfi nttt o h Mathematical the for Institute Pacific the by pported d eo h Institute. the of se { ∈  k φ rsums or 2 d − 4 ( eFuircecet.Seilzn oa to Specializing coefficients. Fourier he eest h Dirichlet the to refers ) − 0 k 1 − ≡ o ucin,adhv hdw given shadows have and functions, sor ueo uhfrs nta,w impose we Instead, forms. such of ture n , . 19 X o fmxdhroi aßfrsin forms Maaß harmonic mixed of ion σ findex of )=( = 1) ,adt hseddfie different a defined end this to and 2, = d 1 ≥ 2 sm n d aso ml iio functions divisor small of lass } 2 , endadivsiae e type new a investigated and defined ] 1 e eiaie.Lsl,w ffra offer we Lastly, derivatives. zed ½ roi aßfrso weight of forms Maaß armonic !   md2) (mod fHce[ Hecke of k n uwt ls ubr,and numbers, class Hurwitz and d. wn h etn f[ of setting the owing X d | − n − φ 1) n as ht functions. theta false and 1  k  eeadtruhu,we throughout, and Here . n d φ ,  if 14 d ua nlge of analogues dular enseries tein > k k to,Hrizcasnum- class Hurwitz ction, ,woerslsimply results whose ], − σ 1 2 sm   ,χ ,where 2, q salinear a as L n fnto of -function 19 E σ k ,let ], 2 sm ,χ 3 2 φ o even for . sany is ψ be φ . 2 JOSHUAMALES,ANDREASMONO,ANDLARRYROLEN

depending on the parity of ψ, let χ 4 be the unique odd character of modulus 4, and we recall Shimura’s

2 1 λψ n θψ(τ) := ψ(n)n q . 2 n Z X∈ Then the main result of [19] states that the function

1 +(τ) := α E (τ)+ σsm (n)qn θ (τ)  ψ 2 1,ψ  E ψ n 1 X≥  3  2 can be completed to a polar harmonic Maaß form of weight λψ on Γ (4M ) with Nebentypus 2 − 0 ψ λψ ψ χ 4, where αψ is an implicit constant to ensure a certain growth condition. That is, it has the transformation· − and analytic properties of a harmonic Maaß form, but it may have poles on the upper half plane arising from the θψ(τ) in the denominator. Similar ideas have been utilized for specific examples before by Andrews, Rhoades, Zwegers [2], and by Bringmann, Kane, Zwegers [4] for instance. Additionally, the authors of [19] presented some special choices of ψ where their polar harmonic Maaß forms happen to have no poles on H, and offer a p-adic property of + for primes p> 3. E A natural question is whether there are more classes of small divisor functions for which a similar phenomenon to the setting of [19] appears. Our two main results give such generalizations. We let χ be a second of modulus Mχ and define our small divisor function by n d n + d σsm (n) := χ d − ψ d d2. 2,χ 2 2 d Dn     X∈ sm We stipulate that ψ is odd and fixed throughout, and thus omit the dependency of σ2,χ on ψ. We moreover define (τ) := +(τ)+ (τ), where F F F − i + 1 sm n i ∞ θχ (w) (τ) := σ2,χ(n)q , −(τ) := 3 dw. F θ (τ) F π√ τ 2 ψ n 1 2 Z ( i (w + τ)) X≥ − − We obtain the following result. Theorem 1.1 If χ is even and non-trivial then the function is a polar harmonic Maaß form of 3 2 2 F 1 1 weight on Γ0 lcm 4Mχ, 4Mψ with Nebentypus χ (ψ χ 4)− . Its shadow is given by θχ. 2 · · − 2π   

Including the possibility of χ = ½, where a constant term arises, requires adjustments either in the holomorphic part + or in the non-holomorphic part . On one hand, we may subtract F F − the arising constant term from − again. The shadow of would be given by the partial theta function F F

1 n2 1 1

q = θ½ (τ) , π π π 2 n 1 2 − 4 X≥ and hence would not be modular anymore. In [6] it is proved that all one-dimensional partial theta functions are (strong) quantum modular forms, which were first introduced by Zagier [25]. Furthermore, such partial theta functions are related to Appell-Lerch sums of level 2, to meromor- phic Jacobi forms, and closely to false theta functions as well. An exposition on the former two connections was given by Bringmann, Zwegers, and the third author in [7]. To state their results, let ζ := e2πiz, and

2 n n 1 1 ∞ j+1 j 1 j+1 ϑ(z; τ) := q 2 ζ = iζ− 2 q 8 1 q 1 ζq 1 ζ− q n Z − j=0 − − − X∈ Y     1 be the standard Jacobi theta function of index and weight 2 . The second equality is the Jacobi triple product identity, from which we deduce that ϑ(z; τ) has zeros precisely in Zτ +Z as a function POLAR HARMONIC MAAß FORMS AND HOLOMORPHIC PROJECTION 3 of z and all zeros are simple. Moreover, letting ℓ N, the Appell-Lerch sum of level ℓ is defined by ∈ n n ℓn ℓ ( +1) 2πinz ( 1) q 2 e A (w, z; τ) := eπiℓw − . ℓ 2πiwqn n Z 1 e X∈ − The results of [7] can be specialized to our setting and read as follows. Proposition 1.2 ([7, Corollaries 1.2 and 1.5]) (i) We have 1 n2 2 2πi(2t) τ q = e A2 t , 0; τ dt. − 1 − 2 n 1 Z 2   X≥ − (ii) Let fn(τ) be defined by the expansion [7, eq. (2.5)] 1 =: f (τ)(2πiz)n. ϑ z τ 2 n ( ; ) n 2 X≥− Then it holds that 1 2πi(2t) n2 n2 2 e f 1(τ) q + 2f 2(τ) nq = dt, − − 1 τ 2 n 1 n 1 Z 2 ϑ t 2 ; τ X≥ X≥ − − and the functions f 1(τ), f 2(τ) both are known to be quasimodular forms (compare [12, Theorem 3.2], [3, Corollary 2.36]− , [7,− p. 8]).

Zwegers provided the non-holomorphic completion of Aℓ to a two-variable Jacobi form of weight ℓ 1 1 and matrix index −1 0 , see [27, Theorem 4]. Moreover, a third perspective arises from the close relation between partial and false theta functions. Bringmann and Nazaroglu described the  completion of false theta functions to functions with certain Jacobi transformation properties in a recent paper [5]. In particular, their result includes the completion for 2 n 2πinz sgn(n)q 2 e , n Z X∈ where sgn(0) := 0. On the other hand, we may compensate for the additional constant term by adjusting the holomorphic part +. Being more precise, we define (τ) := +(τ)+ (τ), where F G G G− 1 1 2 +(τ) := ψ(n)n2qn + σsm (n)qn , θ τ  2,½  G ψ( ) 2 n 1 n 1 X≥ X≥ i  i ∞ θ½ (w) −(τ) := 3 dw. G π√2 τ ( i (w + τ)) 2 Z− − It turns out that the strategy of the proof of Theorem 1.1 still applies, enabling us to complete the picture (regarding even χ) by the following result. Theorem 1.3 The function is a polar harmonic Maaß form of weight 3 on Γ 4M 2 with G 2 0 ψ 1 1   Nebentypus (ψ χ 4)− . Its shadow is given by θ½ . · − 2π Rouse and Webb showed in [22] that the only modular forms on Γ0(N) with integer Fourier co- efficients and no zeros on H are eta quotients. In addition, Mersmann and Lemke-Oliver completed the classification of theta functions which may be written as such an eta quotient. We cite their results in the formulation of [19, Theorem 1.2].

Theorem 1.4 ([17], [18]) The only nontrivial primitive characters ψ for which θψ is an eta quotient are contained in the set of Kronecker characters D Ψ := : D 8, 4, 3, 2, 12, 24 . ∈ {− − − }  ·   Combining Theorems 1.1, 1.3, and 1.4 immediately yields the following corollary. (Recall that there is no odd character of modulus 2.) 4 JOSHUAMALES,ANDREASMONO,ANDLARRYROLEN

Corollary 1.5 If ψ Ψ 2 is odd then + and + are mock theta functions. ∈ \ · F G n o The result of Theorem 1.4 motivates us to investigate the odd choices ψ Ψ 2 in greater ∈ \ · detail. To this end, let H(n) be the Hurwitz class number, counting the weighted numbern o of classes of positive definite binary quadratic forms of discriminant n. Phrased in todays terminology, Zagier discovered in [24] that − 1 +(τ) := + H(n)qn. 12 H − n 1 X≥ can be completed to a harmonic Maaß form (τ) := +(τ) 1 (τ) of weight 3 on Γ (4). To see H H − 4 G− 2 0 this, one may rewrite − as in Lemma 4.1 below, and compare [3, Theorem 6.3] for instance. The G 3 function is then often called Zagier’s weight 2 non-holomorphic Eisenstein series. A standard computationH using the Sturm bound yields the following.

Corollary 1.6 Let ψ = χ 4. Then we have − σsm n q8n θ τ H n q8n 1. 2,½ (8 ) = 4 ψ( ) (8 1) − n 1 − n 1 − X≥ X≥ Or in other words, by definition of ψ = χ 4 and θψ, −

σsm n j j H n j 2 . ½ (8 ) = 4 ( 1) (2 1) 8 (2 1) 2, − − − − j 1   (2j X1)≥2<8n − In addition, we note that both the coefficients of θ + and the values of σsm grow at most ψ 2,½ polynomially. Based on the observations from the preceedinH g discussion we inquire the following. Question For every ψ Ψ 2 , do there exist numbers 0 = C,t Q, such that ∈ \ · 6 ∈ n o C +(tτ) · G generates a linear combination of Hurwitz class numbers on some arithmetic progression? In the course of proving Theorem 1.1 and Theorem 1.3 we compute the holomorphic projection of a product similar to the structure of mixed harmonic Maaß forms. However, we just rely on translation invariance and impose suitable growth conditions on the coefficients to ensure conver- gence. The resulting expression can be written in terms of a particular class of Jacobi polynomials (a,b) r (sometimes also referred to as hypergeometric polynomials.). For r N0, these polynomials areP defined by ∈ r j (a,b) Γ(a + r + 1) r Γ(a + b + r + j + 1) z 1 r (z) := − P r! Γ(a + b + r + 1) j! Γ(a + j + 1) 2 jX=0   Γ(a + r + 1) 1 z = F r, a + b + r + 1, a + 1, − , r a 2 1 ! Γ( + 1) − 2  which can be found in [13, item 8.962] for example. Here, 2F1 denotes the usual Gauß hypergeo- metric function. Then we have the following result.

Proposition 1.7 Let kf R N, kg R ( N), such that κ := kf + kg Z 2. Let α(m), β(n) ∈ \ ∈ \ − ∈ ≥ be two complex sequences, and define1

kf 1 m n f(τ) := α(m)m − Γ(1 kf , 4πmv)q− , g(τ) := β(n)q . m 1 − n 1 X≥ X≥ Suppose that (1) the function (fg)(r + iv) grows at most polynomially as v 0, where r Q, and that (2) the function (fg)(iv) grows at most polynomially as v ց. ∈ ր∞ 1The function Γ(s,z) denotes the incomplete Gamma function, which will be introduced in Section 2. POLAR HARMONIC MAAß FORMS AND HOLOMORPHIC PROJECTION 5

Then the weight κ holomorphic projection of fg is given by m kf 1 (1 kf ,1 κ) kf 1 n m πκ (fg) (τ)= Γ(1 kf ) α(m)β(n) n − − − 1 2 m − q − . − − Pκ 2 − n − m 1 n m 1  −    X≥ −X≥ We provide two proofs of this result, which correspond to either definition of the Jacobi poly- nomials. The first one relies on identities of the Gauß hypergeometric function, while the second one relies on two Lemmas from [20]. Remarks (1) Assuming the framework of mixed harmonic Maaß forms, some variants of this result appear in [15, Theorem 3.5], and [20, Theorem 4.6]. Moreover, if kg = 2 kf , a,b − then 0 (z) = 1. Therefore, working with regularized holomorphic projection, our result includesP [19, Proposition 2.1]. m (2) Note that the summation conditions imply 1 < 1 2 n < 1. The asymptotic behavior of the Jacobi polynomials inside ( 1, 1) is well− known− and can be found in [13, item 8.965] for instance. − (3) One may choose various other special values of half integral kf , kg, which simplify the Jacobi polynomial and then the whole factor

k (1 kf ,1 κ) m k n f 1 − − 1 2 m f 1. − κ 2 n − P −  −  − n 2 This idea leads to other choices of polynomials P d , d Q[X,Y ] than d in the definition of σsm such that ∈ 2,χ  n d n + d n π χ d − ψ d P , d qn κ  2 2 d n 1 d Dn       X≥ X∈  2 2 2(k 1) 2 n2 m2 + α m β n m f − Γ(1 k , 4πm v)q − = 0. f  n 1 m 1 − X≥ X≥     We demonstrate during the proofs of Theorem 1.1 and 1.3 how to rewrite the corresponding generating function n d n + d n χ d − ψ d P , d 2 2 d n 1 d Dn       X≥ X∈ n to obtain a choice of P d , d Q[X,Y ], which matches the factor involving the Jacobi polynomial. ∈  A second application of Proposition 1.7 arises from p-adic properties of + and +. Mertens, Ono and the third author proved such a property for their mock modular EisensteinF G series +, c. f. [19, Theorem 1.4]. More precisely, the idea is to inspect the iterated action of the U-operatorE

α(n)qn U(p) := α(pn)qn n ! n ≫−∞X ≫−∞X 2a + on θψ p τ (τ) for every a N and p> 3 prime. (The notation n is explained in Lemma 2.7.) Then theyE show that this∈ is congruent to some meromorphic modular≫−∞ form of weight 2. In  P our case Theorems 1.1 and 1.3 imply that the products θψ(τ) (τ) and θψ(τ) (τ) are modular of weight 3 with Nebentypus χ or trivial Nebentypus respectively.F Therefore, weG find a different result. Theorem 1.8 Let a, b, p N and suppose that p is an odd prime. Then we have ∈ 2a + b min(a,b) θψ p τ (τ) U p 0 (mod p ), F ≡       and

2a + b min(a,b) θψ p τ (τ) U p 0 (mod p ). G ≡      

6 JOSHUAMALES,ANDREASMONO,ANDLARRYROLEN

sm The third remark on page 3 in [19] states that “the generating function of σ1,ψ can be given in terms of Appell-Lerch sums as studied by Zwegers” in [26], [27] (see also [21, Lemma 2]). This sm remark applies verbatim to the generating function of σ2,χ as well, and we present a strategy which 1 d applies to both generating functions. Let Dz := 2πi dz , giving n n ℓn ℓ ( +1) 2πinz ( 1) q 2 e (DjA )(w,z,τ) = eπiℓw nj − . z ℓ 2πiwqn n Z 1 e X∈ − for every integer j 0. Then we have the following result. ≥ Proposition 1.9 Suppose that χ is non-trivial and even. Additionally assume Mψ Mχ. Then we have that | sm n σ2,χ(n)q n 1 X≥ Mχ 1 Mχ 1 − − 1 c(c+2b Mχ) 2 2 = χ(b) ψ(b + c)q − (MχDz + c) A1 2Mχcτ, z, 2Mχτ . z=(2(b+c) M )M τ+ 1 2 c χ χ 2 bX=1 X=0   − sm Lastly, it is also likely that the function σ2,χ can be viewed as a Siegel theta lift in the following way. A recent paper of Bruinier and Schwagenscheidt [9] investigated the Siegel theta lift on Lorentzian lattices, and its connection to coefficients of mock theta functions. In an isotropic lattice of signature (1, 1), for example, they obtained a formula for the Siegel theta lift of a particular weakly holomorphic evaluated at a certain point in the Grassmanian in terms of the sum H(4m r2). r Z − r 1X (mod∈ 2) ≡ Furthermore, in [1], Alfes-Neumann, Bringmann, Schwagenscheidt, and the first author considered the same lift on a lattice of signature (1, 2), but with the inclusion of an iterated Maass raising operator acting on the weakly holomorphic modular form. There, we obtained an expression of the form (see [1, Example 1.2]) 4D 10n2 10m2 H D n2 m2 . − − − − n,m Z     n DX(mod∈ 2) ≡ The quadratic form in the Hurwitz class number is explained by the signature being (1, 2) in this case, and the addition of the polynomial is a consequence of the iterated Maass raising operators. In view of Corollary 1.6, our smallest divisor function seems to lie at the interface of these two situations, i.e. it is natural to expect that it can be realized as a theta lift in signature (1, 1) where the lift includes iterated Maass raising operators. Perhaps this phenomenon can also be extended to further classes of small divisor functions - the theta lift construction allows one to choose many examples of mock theta functions of appropriate weight, not just the generating function for Hurwitz class numbers.

The paper is organized as follows. The upcoming section briefly recaps the overall framework and collects some basic properties, which we require later. The aim of Section 3 is to present two proofs of Proposition 1.7. Combining the results of Section 2 and 3 enables us to prove our two main Theorems in Section 4. Some modifications of the computations during Section 4 prove Theorem 1.8, and we devote Section 5 to them. Finally, the purpose of Section 6 is to prove the remaining Propositions 1.2 and 1.9.

Acknowledgements The authors would like to thank Michael Mertens particularly, whose comments significantly improved this paper. Furthermore, we would like to thank Kathrin Bringmann as well as the anonymous referee for many valuable comments on an earlier version of this paper. POLAR HARMONIC MAAß FORMS AND HOLOMORPHIC PROJECTION 7

2. Preliminaries 2.1. Growth conditions and modular forms. To describe our various modular objects, we first require some terminology on growth conditions. Definition 2.1 Let n f(τ) := cf (n)q n Z X∈ with some complex coefficients cf (n). Then we say that

(i) the function f is holomorphic at i if cf (n) = 0 for every n< 0. (ii) the function f is of moderate growth∞ at i if f O (vm) as v for some m N. In other words f grows at most polynomially. ∞ ∈ →∞ ∈ 1 (iii) the function f is of governable growth at i if there exists Pf C q− such that for some δ > 0 we have ∞ ∈   δv f(τ) Pf (τ) O e− , v . − ∈ →∞   The polynomial Pf is called the principal part of f. Equivalently, f is permitted to have a pole at i . (iv)∞ f is of linear exponential growth if f(τ) O eδv as v for some δ > 0. ∈ →∞   These conditions can be phrased at all other cusps via suitable scaling matrices. To define the slash-operator, we take the principal branch of the holomorphic square root throughout. 1 a b Definition 2.2 Let k 2 Z, φ be a Dirichlet character, and γ = c d SL2(Z). Then the slash operator is defined as ∈ ∈  1 k φ(d)− (cτ + d)− f(γτ) if k Z, (f kγ) (τ) := 1 c 2k k ∈ 1 | (φ(d)− ε (cτ + d)− f(γτ) if k + Z, d d ∈ 2 c  where d denotes the extended Legendre symbol, and  1 if d 1 (mod 4), εd := ≡ (i if d 3 (mod 4). ≡ for odd integers d. For the sake of completeness, we define the variants of modular forms appearing in this paper.

Definition 2.3 Let f : H C be a function, Γ SL2(Z) be a subgroup, φ be a Dirichlet character, and k 1 Z. Then we say→ that ≤ ∈ 2 (i) The function f is a modular form of weight k on Γ with Nebentypus φ if (a) for every γ Γ and every τ H we have (f kγ) (τ)= f(τ), (b) f is holomorphic∈ on H, ∈ | (c) f is holomorphic at every cusp. We denote the vector space of functions satisfying these conditions by Mk(Γ, φ). (ii) If in addition f vanishes at every cusp, then we call f a . The subspace of cusp forms is denoted by Sk(Γ, φ). (iii) If f satisfies the conditions (a) and (b) from (i) and is allowed to have a pole at one or more cusps, then we call f a weakly holomorphic modular form of weight k on Γ with Nebentypus φ. ! The vector space of such functions is denoted by Mk(Γ, φ). Furthermore, we recall the following fact which we require throughout. Lemma 2.4 The following are true 2 M 1 (Γ0(4Mψ), ψ) if λψ = 0, θ 2 ψ  2 ∈ S 3 (Γ0(4Mψ), ψ χ 4) if λψ = 1.  2 · −  8 JOSHUAMALES,ANDREASMONO,ANDLARRYROLEN

A proof can be found in [16, Theorem 10.10] for instance. Finally, Corollary 1.6 follows directly from the following result.

Lemma 2.5 (Sturm bound) Let f, g Mk (Γ (N), φ), k > 1, and ∈ 0 1 m := [SL (Z): Γ (N)] = N 1+ . 2 0 p p N   Y| Define km B := , 12   and denote by cf (n), cg(n) the coefficients of the q-expansions of f and g respectively. If cf (n)= cg(n) for all n B then f = g. ≤ A proof can be found in [23, Corollary 9.20].

2.2. Harmonic Maaß forms and shadows. We define our main objects of interest.

1 Definition 2.6 Let k 2 Z, and choose N N such that 4 N whenever k Z. Let φ be a Dirichlet character of modulus∈ N. ∈ | 6∈

(i) A weight k harmonic Maaß form on a subgroup Γ0(N) with Nebentypus φ is any smooth function f : H C satisfying the following three properties: → (a) For all γ Γ0(N) and all τ H we have (f kγ) (τ)= f(τ). (b) The function∈ f is harmonic with∈ respect to the| weight k hyperbolic Laplacian on H, explicitly ∂2 ∂2 ∂ ∂ 0=∆ f := v2 + + ikv + i f. k ∂u2 ∂v2 ∂u ∂v − !  ! (c) The function f has at most linear exponential growth at all cusps. ! We denote the vector space of such functions by Hk(Γ0(N), φ). (ii) If we restrict the growth condition in (c) to governable growth then the vector space of such forms is denoted by Hk(Γ0(N), φ). (iii) A polar harmonic Maaß form is a harmonic Maaß form with isolated poles on the upper half plane. Remark If we restrict the growth condition (c) to moderate growth at all cusps, and allow arbitrary eigenvalues in (b), then f is a classical Maaß wave form. During the following summary, we may assume that φ is trivial for simplicity, since the gener- alization to a nontrivial Nebentypus is immediate. Bruinier and Funke observed in [8] that the Fourier expansion of a harmonic Maaß form2 natu- rally splits into two parts. One of them involves the incomplete Gamma function

∞ s 1 t Γ(s, z) := t − e− dt, Zz defined for Re(s) > 0 and z C. It can be analytically continued in s via the functional equation ∈ s z Γ(s + 1, z)= sΓ(s, z)+ z e− , and has the asymptotic behavior s 1 v Γ(s, v) v − e− , v ∼ | |→∞ for v R. We state their result in the formulation of [3, Lemma 4.3]. ∈

2Be aware that their terminology refers to our harmonic Maaß forms as “harmonic weak Maaß forms” instead. POLAR HARMONIC MAAß FORMS AND HOLOMORPHIC PROJECTION 9

Lemma 2.7 Let k 1 Z 1 and f H! (Γ (N)). Then f has a Fourier expansion of the shape ∈ 2 \ { } ∈ k 0 + n 1 k n f(τ)= cf (n)q + cf−(0)v − + cf−(n)Γ(1 k, 4πnv)q . n n − − ≫−∞X nX≪∞=0 6 In particular, if f Hk(Γ (N)) then f has a Fourier expansion of the shape ∈ 0 + n n f(τ)= cf (n)q + cf−(n)Γ(1 k, 4πnv)q . n n<0 − − ≫−∞X X The notation n abbreviates n m for some mf Z. The notation n is defined ≫−∞ ≥ f ∈ ≪∞ analogously, and similar expansions hold at the other cusps. P P P We follow the following terminology from [3, Definition 4.4]. Definition 2.8 We refer to the functions + + n 1 k n f (τ) := cf (n)q , f −(τ) := cf−(0)v − + cf−(n)Γ(1 k, 4πnv)q n n − − ≫−∞X nX≪∞=0 6 as the holomorphic part of f and to f − as its non-holomorphic part. In the same paper, Bruinier and Funke introduced the operator ∂ ∂ ∂ ξ := 2ivk = ivk + i . k ∂τ ∂u ∂v   We summarize its relevant properties. Lemma 2.9 Let f be a smooth function on H. Then the ξ-operator satisfies the following proper- ties.

(i) We have ξk(f) = 0 if and only if f is holomorphic. (ii) The slash operator intertwines with ξk, i. e. , we have

ξk (f kγ) = (ξkf) 2 kγ | | − 1 for every γ SL2(Z) if k Z or γ Γ0(4) if k 2 Z Z respectively. ∈ ∈ ∈ ∈ ! \ ! (iii) The kernel of ξk restricted to Hk(Γ0(N)) or Hk(Γ0(N)) is precisely the space Mk(Γ0(N)) in both cases. (iv) Let f H! (Γ (N)). Assuming the notation of Lemma 2.7 we have ∈ k 0 1 k 1 k n ! ξkf(τ)= ξkf −(τ) = (1 k)cf−(0) (4π) − cf−( n)n − q M2 k (Γ0 (N)) , − − n − ∈ − ≫−∞X and in particular if f Hk(Γ (N)) then ∈ 0 1 k 1 k n ξkf(τ)= (4π) − cf−( n)n − q S2 k (Γ0 (N)) . − − n 1 − ∈ X≥ ! ։ ! In addition, we have ξk : Hk(Γ0(N)) M2 k (Γ0 (N)). − The first item is simply a reformulation of the Cauchy-Riemann equations and the second one is induced by the corresponding well known property for the Maaß lowering operator ∂ 1 ∂ ∂ L := v2 = v2 + i . k ∂τ 2 ∂u ∂v   We fix some more terminology, following [3, Definition 5.16] and the second remark afterwards. Definition 2.10 (i) A function f is called a mock modular form if f is the holomorphic part of a harmonic Maaß form for which f − is nontrivial. ! + (ii) If f Hk(Γ0(N)) then we refer to the form (ξkf) as the shadow of f . ∈ 1 3 (iii) In particular, f is called a mock theta function if f is a mock modular form of weight 2 or 2 , whose shadow is a linear combination of unary theta functions. 10 JOSHUAMALES,ANDREASMONO,ANDLARRYROLEN

Moreover, we study the following objects, which were introduced first in [11, Section 7.3]. How- ever, we follow the definition given in [3, Section 13.2].

Definition 2.11 (i) A mixed harmonic Maaß form of weight (k1, k2) is a function h of the shape n h(τ)= fj(τ)gj(τ), jX=1 ! where fj Hk and gj M for every j. ∈ 1 ∈ k2 (ii) Analogously, a mixed mock modular form of weight (k1, k2) is a function h of the shape n h(τ)= fj(τ)gj(τ), jX=1 ! where each fj is a mock modular form of weight k1 and gj M for every j. ∈ k2 We extend the last result of Lemma 2.9 to mixed harmonic Maaß forms. It suffices to consider products involving the non-holomorphic part of a mixed harmonic Maaß form.

Lemma 2.12 Let kf , kg R, κ := kf + kg, and let α(m), β(n) be two complex sequences. Define ∈ kf 1 m n f(τ) := α(m)m − Γ(1 kf , 4πmv)q− , g(τ) := β(n)q . m 1 − n 1 X≥ X≥ Then

1 kf kg m n ξκ (fg) (τ)= (4π) − v α(m)q β(n)q . − m 1 n 1 X≥ X≥ Proof: We have

kf 1 n m (fg)(τ)= α(m)m − Γ(1 kf , 4πmv)β(n)q − m 1 n 1 − X≥ X≥ and ∂ a 1 v Γ(a, v)= v − e− . ∂v − n m We compute that ξκ (Γ(1 kf , 4πmv)q ) equals − − κ n m iv Γ(1 kf , 4πmv)2πi(n m)q − − − n kf 4πmv n m n m i (4πmv)− e− 4πmq + Γ(1 kf , 4πmv)( 2π(n m))q − − − − − − − hkg 1 kf 4πmv n m kg 1 kf m 2πinτ io = v (4πm) − e− q = v (4πm) − q e− , − − − and infer

1 kf kg m n (ξκ (fg)) (τ)= (4π) − v α(m)β(n)q q , − m 1 n 1 X≥ X≥ as claimed.  2.3. Holomorphic projection. We introduce the holomorphic projection operator. Its origin lies in the search for an operator which preserves the (regularized) Petersson inner product. Although this can be derived implicitly from the Riesz representation theorem, an explicit description comes in handy quite often.

Definition 2.13 Let f : H C be a translation invariant function and k Z 2. If f has at most moderate growth towards the→ cusps then the weight k holomorphic projection∈ ≥of f is defined by (k 1)(2i)k f (x + iy) yk dxdy (πkf) (τ) := − k , 4π H (τ x + iy) y2 Z − whenever the integral converges absolutely. Furthermore, we require the following result. POLAR HARMONIC MAAß FORMS AND HOLOMORPHIC PROJECTION 11

Lemma 2.14 (Lipschitz summation formula) For any r N we have that ∈ 1 ( 2πi)r = − jr 1e2πijw. w j r r − j Z ( + ) ( 1)! j 1 X∈ − X≥ A short proof is due to Zagier and can be found in [10, p. 16]. We summarize two further properties of the holomorphic projection operator, both of which are proven in [4, Section 3] for instance. Lemma 2.15 Let f : H C be a translation invariant function of moderate growth such that the → integral defining πkf converges absolutely, and k Z 2. Then πk enjoys the following properties. ∈ ≥ (i) If f is holomorphic then πkf = f. (ii) If f is modular with some Nebentypus φ (but not necessarily holomorphic), then

Mk (Γ0 (N) , φ) if k Z 3, πkf ∈ ≥ ∈ (M2 (Γ0 (N) , φ) M0 (Γ0 (N) , φ) E2 if k = 2. ⊕ · Therefore, the slash operator and πk commute if k 3. ≥

3. Two proofs of Proposition 1.7 During the first proof of Proposition 1.7, we appeal to the following technical results. Lemma 3.1 (i) If Re(b), Re(a + b) > 0 and Re(c + s) > 0 then a ∞ b 1 sz c Γ(a + b) s Γ(a, cz)z − e− dz = F 1, a + b, b + 1; . b(c + s)a+b 2 1 s + c Z0   (ii) The hypergeometric function 2F1 satisfies c a b F (a, b, c; z) = (1 z) − − F (c a, c b, c; z), 2 1 − 2 1 − − and Γ(c)Γ(c a b) F (a, b, c; z)= − − F (a, b, a + b c + 1; 1 z) 2 1 Γ(c a)Γ(c b) 2 1 − − − − c a b Γ(c)Γ(a + b c) + (1 z) − − − F (c a, c b, c a b + 1; 1 z). − Γ(a)Γ(b) 2 1 − − − − − The first identity is [13, item 6.455]. Both hypergeometric transformations can be found in [13, page 1008].

First proof of Proposition 1.7: The q-expansion of (fg) (τ) is given by

kf 1 n m (fg)(τ)= α(m)β(n)m − Γ(1 kf , 4πmv)q − . m 1 n 1 − X≥ X≥ We see that fg is translation invariant, and recall that it has moderate growth towards all cusps by assumption, so πκ (fg) exists. Hence, we need to calculate (κ 1)(2i)κ (fg) (x + iy) yκ dxdy πκ (fg) (τ)= − κ . 4π H (τ x + iy) y2 Z − The integral converges since κ 2 0, and converges absolutely if κ > 2. Using the translation invariance of fg, we rewrite the− integral≥ over H as

κ 1 (fg) (x + iy) y dxdy ∞ κ 2 1 = (fg) (x + iy)y − dxdy, (τ x + iy)κ y2 (τ x + iy + j)κ ZH Z0 Z0 j Z − X∈ − 12 JOSHUAMALES,ANDREASMONO,ANDLARRYROLEN

and consequently (κ 1)(2i)κ π (fg) (τ)= − α(m)mkf 1β(n) κ π − 4 m 1 n 1 X≥ X≥ 1 ∞ κ 2 1 2πi(n m)(x+iy) Γ(1 kf , 4πmy)y − e − dxdy. × − (τ x + iy + j)κ Z0 Z0 j Z X∈ − By the Lipschitz summation formula and then Lemma 3.1 (i), we infer that πκ (fg) (τ) equals κ κ (κ 1)(2i) ( 2πi) k 1 − − α(m)m f − β(n) 4π (κ 1)! m 1 n 1 − X≥ X≥ 1 ∞ κ 2 κ 1 2πij(τ x+iy) 2πi(n m)(x+iy) Γ(1 kf , 4πmy)y − j − e − e − dxdy × − Z0 Z0 j 1 X≥ κ κ (κ 1)(2i) ( 2πi) k 1 κ 1 = − − α(m)m f − β(n)(n m) − 4π (κ 1)! m 1 n m 1 − − X≥ −X≥ ∞ κ 2 4π(n m)y n m Γ(1 kf , 4πmy)y − e− − dy q − × − Z0 κ 1 Γ(kg) (n m) − m n m = α(m)β(n) − 2F1 1, kg, κ; 1 q − . (κ 1)! nkg − n m 1 n m+1   − X≥ ≥X Finally, we apply the hypergeometric transformations from Lemma 3.1 (ii). Explicitly, 1 κ k 1 m (κ 1) m Γ(κ)Γ(1 kf ) m − m f − 2F1 1, kg, κ; 1 = − 2F1 1, kg, 2 kf ; + − 1 − n kf 1 − n Γ(kg) − n n   −       1 κ 1 κ k 1 (κ 1) m − m Γ(κ)Γ(1 kf ) m − m f − = − 1 2F1 1 kf , 2 κ, 2 kf ; + − 1 . kf 1 − n − − − n Γ(kg) − n n −         Thus, we arrive at κ 1 Γ(kg) (n m) − m − 2F1 1, kg, κ; 1 (κ 1)! nkg − n −   k κ m κ k Γ( g) kf 1 ( 1) Γ( )Γ(1 f ) kf 1 = n − − 2F1 1 kf , 2 κ, 2 kf ; + − m − (κ 1)! kf 1 − − − n Γ(kg) − −   ! k m kf 1 Γ( g) kf 1 = Γ(1 kf ) n − 2F1 2 κ, 1 kf , 2 kf , m − − − (κ 2)! Γ(2 kf ) − − − n − − −   ! k (1 kf ,1 κ) m k = Γ(1 k ) n f 1 − − 1 2 m f 1 , f − κ 2 n − − −  P −  −  −  and ultimately obtain m kf 1 (1 kf ,1 κ) kf 1 n m πκ (fg) (τ)= Γ(1 kf ) α(m)β(n) n − − − 1 2 m − q − , − − Pκ 2 − n − m 1 n m 1  −    X≥ −X≥ as desired.  Second proof of Proposition 1.7: One copies the first proof until the application of the Lipschitz summation formula, which produced the expression κ κ (κ 1)(2i) ( 2πi) k 1 κ 1 π (fg) (τ)= − − α(m)m f − β(n)(n m) − κ 4π (κ 1)! m 1 n m 1 − − X≥ −X≥ ∞ κ 2 4π(n m)y n m Γ(1 kf , 4πmy)y − e− − dy q − . × − Z0 Then, one shows the following two identities. The first relies on the fact that κ = kf + kg is an integer. POLAR HARMONIC MAAß FORMS AND HOLOMORPHIC PROJECTION 13

(i) Define the homogeneous polynomial a 2 − j + b 2 j a j 2 Pa,b(X,Y ) := − X (X + Y ) − − C[X,Y ] j ! ∈ jX=0 of degree a 2. Then we have − ∞ κ 2 4πry Γ(1 kf , 4πmy)y − e− dy − Z0 Γ(1 kf )(κ 2)! π 1 κm1 kf r m 1 kg P r, m mkf 1 . = (4 ) − − − κ 1 − ( + ) − κ,2 kf ( ) − − r − − −   A proof can be found in [20, Lemma 4.7]. (ii) If b = 1, 2 then the polynomial Pa,b(X,Y ) from the first item satisfies 6 a 2 − a + b 3 j + b 2 a 2 j j Pa,b(X,Y )= − − (X,Y ) − − ( Y ) . a 2 j! j ! − jX=0 − − A proof can be found in [20, Lemma 5.1]. Next, one proceeds by writing

1 kg kf 1 n m πκ (fg) (τ)= Γ(1 kf ) α(m)β(n) n − Pκ,2 k (n m,m) m − q − , − f − − m 1 n m 1 − − X≥ −X≥   according to the first identity, and then writing κ 2 − k j k 1 kg g 1 f kf 1 j j n − Pκ,2 kf (n m,m)= − − n − − ( m) − − κ 2 j! j ! − jX=0 − − k 1 κ 2 j k n f − κ m m Γ( g) − 2 1 kf 1 (1 kf ,1 κ) = − = n − κ −2 − 1 2 , (κ 2)! Γ(1 kf ) j !j + 1 kf − n P − − n − − jX=0 −     by virtue of the second identity. Summing up, one obtains m kf 1 (1 kf ,1 κ) kf 1 n m πκ (fg) (τ)= Γ(1 kf ) α(m)β(n) n − − − 1 2 m − q − , − − Pκ 2 − n − m 1 n m 1  −    X≥ −X≥ as claimed. 

4. Proof of Theorem 1.1 and Theorem 1.3 We collect the results needed to prove Theorem 1.1 and Theorem 1.3. We begin by rewriting the definitions of and . F − G− Lemma 4.1 We have

2 1 2 m2 −(τ)= χ(m)mΓ , 4πm v q− , F 1 −2 Γ m 1   − 2 X≥ and  

2 1 2 m2 1 −(τ)= mΓ , 4πm v q− 1 . G 1 −2 − 2 Γ m 1   2πv − 2 X≥   Proof: We compute i i 2πim2(z τ) 1 θχ (w) 1 e − 2 ∞ 2 ∞ (2π)− i 3 dw = (2π)− i χ(m) 3 dz − 2 − 2 Z τ ( i (w + τ)) m 1 Z2iv ( iz) − − X≥ − 1 ∞ 3 2πm2x m2 ∞ 1 1 t m2 = (2π)− 2 χ(m) x− 2 e− dx q− = χ(m)m t− 2 − e− dt q− , 2 m 1 Z2v  m 1 Z4πm v  X≥ X≥ 14 JOSHUAMALES,ANDREASMONO,ANDLARRYROLEN

and the first claim follows directly, since 2 = 1 . To prove the second claim, it remains to Γ( 1 ) √π − 2 −

separate the constant term of θ½ , and next to calculate

i 1 i ∞ 1 ∞ 3 1 2 2 3 dw = t− dt = 1 , π√2 τ ( i (w + τ)) 2 −2π√2 2v −2πv 2 Z− − Z as asserted. 

In addition, we have the following immediate corollary of Proposition 1.7.

+ Corollary 4.2 Let f(τ) := f (τ)+ f −(τ) be the splitting of f into its holomorphic and non- holomorphic part. Assume the notation and hypotheses as in Proposition 1.7. Then

πκ (fg) (τ) m + kf 1 (1 kf ,1 κ) kf 1 n m = f g (τ) Γ(1 kf ) α(m)β(n) n − − − 1 2 m − q − . − − Pκ 2 − n − m 1 n m 1  −      X≥ −X≥ Moreover, we combine the properties of the slash-operator, the shadow operator, and of the holomorphic projection operator. This yields a third preparatory result. Proposition 4.3 ([19, Proposition 2.3]) Let f : H C be a translation invariant function such that f(τ) vδ is bounded on H for some δ > 0. If the→ weight k holomorphic projection of f vanishes | | identically for some k > δ + 1 and ξkf is modular of weight 2 k for some subgroup Γ < SL2(Z), then f is modular of weight k for Γ. − Proof: This is a straightforward adaption of [4, Proposition 3.5]. Let γ Γ. Then the modularity ∈ of ξkf implies that

ξk ((f kγ) f) = (ξkf) 2 kγ ξkf = 0. | − | − − Hence, (f kγ) f is holomorphic. This yields | − (f kγ) f = πκ ((f kγ) f)= πκ (f) kγ πκ (f) , | − | − | − and by assumption the right hand side vanishes. This proves the claim. 

Remark The subtle growth conditions are required to include the case π2, and are clearly satisfied if we deal with higher weight holomorphic projections, in which case the integral defining πk converges absolutely. Proof of Theorem 1.1: We need to check the three conditions required by the definition of a har- monic Maaß form.

(a) Growth conditions. Recall that θψ is a cusp form, namely it decays exponentially towards all cusps. In turn, the function + admits at most linear exponential growth towards all cusps. Note F + that in particular i is a removable singularity of , since i is a simple zero of both θψ and + ∞ F ∞ θψ . To inspect the non-holomorphic part, we have F 3 1 2 m2 2 − 2 2πm2v Γ , 4πm y q− 4πm v e− , v , −2 ∼ →∞     and hence the function decays exponentially towards the cusp i . By the transformation F − ∞ properties of θχ under the full modular group SL2(Z), we deduce that − is of moderate growth towards all cusps. This establishes the growth condition required by DefinitionF 2.6. + As pointed out in the introduction after Corollary 1.6, the Fourier coefficients of θψ at i + F ∞ are of moderate growth, wherefore the growth of the function θψ towards any cusp is moderate. One may see this by choosing suitable scaling matrices, whose actionF yields additional polynomial factors in τ. Consequently, the growth of θψ is moderate. This justifies the existence of π3 (θψ ) F + F as well as the application of Proposition 4.3 to θψ during the upcoming item. F POLAR HARMONIC MAAß FORMS AND HOLOMORPHIC PROJECTION 15

(b) The transformation law. First, we compute

1 1 2 2 2 1 ( 1 , 2) m 1 1 1 m 1 m n n 2 − 2 − 1 2 m 2 = n 2 + m 2 = − . 1 n 2 2n  1  P  −  −   − 2n 2 Comparing our initial setting with Proposition 1.7 and Lemma 2.12, we need to switch to squares above. By virtue of Lemma 4.1 and the definition of θψ, we have the coefficients 2 α m2 = χ(m), β n2 = ψ(n)n. Γ 1   − 2    sm We rewrite the generating function of σ2,χ as n n + sm n d d d + d 2 n θψ (τ)= σ (n)q = χ − ψ d q F 2,χ 2 2 n 1 n 1 d Dn       X≥ X≥ X∈ j d j + d = χ − ψ d2qdj 2 2 d 1 j d     X≥ j d X(mod≥ 2) ≡ 2 n2 m2 = χ(m)ψ(n) (n m) q − , m 1 n m 1 − X≥ −X≥ and apply Corollary 4.2 to obtain

+ 2 n2 m2 π3 ( θψ) (τ)= θψ (τ) χ(m)ψ(n) (m n) q − = 0. F F − m 1 n m 1 −   X≥ −X≥ Furthermore, we apply Lemma 2.12, obtaining 2 1 2 3 m2 n2 1 3 θψ(τ) ξ3 ( θψ) (τ)= (4π)− 2 v 2 χ(m)q ψ(n)nq = v 2 θχ(τ)| | . 1 2π θ (τ) F − Γ m 1 n 1 ψ − 2 X≥ X≥   2 2 2 2 We observe that ξ ( θψ) (τ) is modular of weight 1forΓ (4M ) Γ (4M )=Γ lcm 4M , 4M 3 F − 0 χ ∩ 0 ψ 0 χ ψ 1 1 a b 2 2    with Nebentypus χ ψ− χ−4. Indeed, for any γ = c d Γ0 lcm 4Mχ, 4Mψ we have · · − ∈     3 2 3 2 1 v 2 1 ψ(d)χ 4(d)(cτ + d) θψ(τ) ξ θ γτ χ d cτ d 2 θ τ − 3 ( ψ) ( )= 3 ( )( + ) χ( ) 3 F 2π cτ + d ψ(d)χ 4(d)(cτ + d) 2 θψ(τ) | | − 1 1 1 = χ(d)ψ(d)− χ−4(d)(cτ + d)− ξ3 ( θψ) (τ). − F Finally, Proposition 4.3 applies directly, because the growth conditions are met thanks to absolute convergence. We deduce that θψ is modular of weight 3 with respect to the same data, as desired. +F (c) Harmonicity. Clearly, is holomorphic away from the zeros of θψ. Hence, the Cauchy- F + Riemann equations imply ∆ 3 = 0 directly. The computation for − is standard, so we only 2 F F sketch its results. It holds that 2πm2(iu+v) ∂ ∂ 1 2 i e + i Γ , 4πm2v q m = − , ∂u ∂v − 1 3   −2  −2 π 2 mv 2 2 2 2πm2(iu+v) ∂ ∂ 1 2 m2 3 e− + Γ , 4πm v q− = , ∂u2 ∂v2 2 4 1 5 ! −  π 2 mv 2 2πm2(iu+v) 2πm2(iu+v) 2 3 e− 3i i e− ∆ 3 − (τ)= v 1 5 + v 1 3 = 0, 2 F − 4 π 2 mv 2 ! 2 −2 π 2 mv 2 !   and thus ∆ 3 = 0 away from the zeros of θψ. 2 F Altogether, this completes the proof, since the shadow is a byproduct of the second item.  16 JOSHUAMALES,ANDREASMONO,ANDLARRYROLEN

Proof of Theorem 1.3: The proof of Theorem 1.3 uses the same ideas as the proof of Theorem 1.1, so we just emphasize the differences. Recall from Lemma 4.1 that

2 1 2 m2 1 −(τ)= mΓ , 4πm v q− 1 . G 1 −2 − 2 Γ m 1   2πv − 2 X≥   Therefore, to compute π (θψ ), it suffices to deal with the second term. We see that 3 G− θψ(τ) 1 − 2πv 2 is translation invariant, vanishes at i , and has a removable singularity at all other cusps inspecting the order of vanishing as v 0. Hence∞ the integral defining its weight 3 holomorphic projection exists and converges absolutely.ց The computation begins exactly as in the proof of Proposition 1.7 and we employ the Lipschitz summation formula. This yields

3 1 1 2πin2(x+iy) θψ(τ) 2(2i) ∞ y 2 e π3 1 = 2 ψ(n)n 3 dxdy − 2 − 8π (τ x + iy + j)  2πv  n 1 j Z Z0 Z0 X≥ X∈ − 1 2 ∞ 1 2πi(n2(x+iy)+j(τ x+iy)) = 8π ψ(n)n j y 2 e − dxdy − n 1 j 1 Z0 Z0 X≥ X≥ 5 ∞ 1 4πn2y n2 1 2 n2 = 8π ψ(n)n y 2 e− dy q = ψ(n)n q , − −2 n 1 Z0  n 1 X≥ X≥ which also appeared in [20, Lemma 4.4] in the framework of mixed harmonic Maaß forms. Fur- thermore, it follows by Corollary 4.2 (with identical parameters of the Jacobi polynomial as in the proof of Theorem 1.1) that

+ 2 n2 m2 1 2 n2 π3 ( θψ) (τ)= θψ (τ) ψ(n) (m n) q − ψ(n)n q . G G − m 1 n m 1 − − 2 n 1   X≥ −X≥ X≥ In addition, we note that

j + d 2 2 σsm (n)qn = ψ d2qjd = ψ(n)(n m)2qn m , ½ − 2, 2 − n 1 d 1 j d   m 1 n m+1 X≥ X≥ j d X(mod≥ 2) X≥ ≥X ≡ where we substituted d = n m, j = n + m in the last equation. Collecting these observations and inserting the definition of− +, we obtain G π ( θψ) (τ) = 0. 3 G Moreover, 2 1 3 m2 1 3 1 3 θψ(τ) m2 ξ ( θ ) (τ)= v 2 θ (τ) q + v 2 θ (τ)= v 2 | | q , 3 ψ 2π ψ 4π ψ 4π θ (τ) G m 1 ψ m Z X≥ X∈ 2 2 2 which is modular of weight 1onΓ0 4Mψ Γ0(4) = Γ0 lcm 4, 4Mψ =Γ0(4Mψ) with Neben- 1 − ∩ typus (ψ χ 4)− by the same argument as in the proof of Theorem 1.3. This establishes weight · − 3 modularity of θψ via Proposition 4.3 again. Additionally, weG clearly have 1 ∆ 3 1 = 0, 2 −2πv 2  and hence harmonicity is preserved. Finally, the growth properties of + towards all cusps agree verbatim with +. Summing up, this establishes the Theorem. G  F POLAR HARMONIC MAAß FORMS AND HOLOMORPHIC PROJECTION 17

5. Proof of Theorem 1.8 Proof of Theorem 1.8: We prove the first claim. On one hand, by the same computation as during the proof of Theorem 1.1 we infer

2a 2a + a 2 r π θψ p τ (τ) = θψ p τ (τ)+  χ(m)ψ(n) (m p n)  q =: g(τ). 3 F F −       rX1  m,nX 1  ≥ (pan)2 ≥m2=r   −    Invoking Theorem 1.1 and the properties of holomorphic projection we deduce that g is a modular form of weight 3 on Γ lcm 4M 2p2a, 4M 2 with Nebentypus χ. However, clearly I Γ (N) 0 ψ χ − ∈ 0 for every level N and hence gvanishes identically (recall that χ is assumed to be even throughout). On the other hand, the inner sum can be rewritten as a sum over small divisors of r. To this end, the set of admissible small divisors is given by r r r D (p) := d r : 1 d , d (mod 2), d + 0 (mod 2pa) , r d d d  | ≤ ≤ ≡ ≡  and exactly as in the proof of Theorem 1.1 we see that n d n + d χ(m)ψ(n) (m pan)2 = χ d − ψ d d2. − 2 2pa m,n 1 d Dr(p)     (pan)X2 ≥m2=r ∈X − If we apply the operator U(pb) to g then we need to replace r by pbr everywhere above. This produces the condition pbr d + 0 (mod 2pa) d ≡ in the set of admissible small divisors, which eventually forces d 0 (mod pmin(a,b)), ≡ since d is a divisor of r. Combining, we arrive at

b 2a + b min(a,b) 0= g(τ)= g(τ) U(p ) θψ p τ (τ) U(p ) (mod p ), ≡ F     as claimed. The proof of the second claim is completely analogous. The character χ is trivial, Mχ = 1, and one r can remove the condition d d (mod 2) from the definition of the set of admissible small divisors. However, this does not affect≡ the rest of the proof and we provided the necessary computations during the proof of Theorem 1.3 essentially. 

6. Proof of Proposition 1.2 and of Proposition 1.9 Proof of Proposition 1.9: The first step is to apply geometric expansion. We compute

sm n 2 n2 m2 2 s2+2ms σ2,χ(n)q = χ(m)ψ(n) (n m) q − = χ(m)ψ(m + s)s q n 1 m 1 n m 1 − m 1 s 1 X≥ X≥ −X≥ X≥ X≥ Mχ 1 − 2 2 = χ(b)ψ(s + b)s2qs +2(aMχ+b)s χ(0)ψ(s + b)s2qs s 1 a 0 b=0 − X≥ X≥ X Mχ 1 2 − qs +2bs = χ(b) ψ(s + b)s2 , 1 q2Mχs b=1 s 1 X X≥ − where we have used the assumption Mψ Mχ after the substitution m = aMχ + b and the assumption that χ is non-trivial in the last| equation. The second step is to convert the sum in s 18 JOSHUAMALES,ANDREASMONO,ANDLARRYROLEN

to a sum over Z instead of N. Note that qs2+2bs qs2+2bs qs2 2bs 2 ψ(s + b)s2 = ψ(s + b)s2 + ψ( s + b)s2 − 1 q2Mχs 1 q2Mχs 1 q 2Mχs s 1 s 1 s 1 − − X≥ − X≥ − X≤− − 2 2 qs +2bs qs +2Mχs 2bs = ψ(s + b)s2 + ψ(s b)s2 − , q2Mχs q2Mχs s 1 1 s 1 − 1 X≥ − X≤− − using that ψ is odd. The key observation is

M 2 M 2 χ 1 s +2Mχs 2bs χ 1 s +2bs − 2 q − − 2 q χ(b) ψ(s b)s = χ(Mχ b) ψ(s (Mχ b))s − 1 q2Mχs − − − 1 q2Mχs b=1 s 1 b=1 s 1 X X≤− − X X≤− − Mχ 1 2 − qs +2bs = χ(b) ψ(s + b)s2 . 1 q2Mχs b=1 s 1 X X≤− − using that χ is even. Thus, we have

Mχ 1 s2+2bs sm n 1 − 2 q σ ,χ(n)q = χ(b) ψ(s + b)s , 2 2 1 q2Mχs n 1 b=1 s Z X≥ X X∈ − since the constant term in s vanishes as well. The third step is to substitute s = nMχ + c to isolate ψ from its s-dependency (recall Mψ Mχ), getting | M M 2 χ 1 χ 1 (nMχ+c) +2b(nMχ+c) sm n 1 − − 2 q σ2,χ(n)q = χ(b) ψ(b + c) (nMχ + c) , 2 q2Mχ(nMχ+c) n 1 b=1 c=0 n Z 1 X≥ X X X∈ − from which we read off the claim. 

Proof of Proposition 1.2 (i): We provide a purely computational proof. To this end, let

1 2 2πi(2t) τ J(τ) := e A2 t , 0; τ dt. 1 − 2 Z− 2   We split the sum defining A2 into its positive, constant, and negative summands (with respect to 2πit n 1 the summation variable n). Next, we note that e± q ∓ 2 < 1 for every n 1 and t R, which enables us to apply geometric expansion, which| yields | ≥ ∈ n(n+1) q τ n(n+1)+jn 2πij(t 2 ) τ = q e − , 2πi(t ) n n 1 1 e − 2 q n 1 j 0 X≥ − X≥ X≥ n(n+1) n2 2πi(t τ ) q q e− − 2 2 τ n +jn 2πi(1+j)(t 2 ) τ = τ = q e− − . 2πi(t ) n 2πi(t ) n n 1 1 e − 2 q − n 1 1 e− − 2 q − n 1 j 0 X≤− − X≥ − X≥ X≥ We collect all terms, getting

1 1 2 2πi(2t) 2πi(t τ ) n(n+1)+jn 2πij(t τ ) 2 2πi(2t) 2πi(t τ ) 1 J(τ)= e e − 2 q e − 2 dt + e e − 2 dt 1 1 2πi(t τ ) n 1 j 0 Z 2 Z 2 1 e − 2 X≥ X≥ − − − 1 2 2πi(2t) 2πi(t τ ) n2+jn 2πi(1+j)(t τ ) e e − 2 q e− − 2 dt. − 1 n 1 j 0 Z 2 X≥ X≥ − The first term vanishes and the third term reduces to the case j = 2. Moreover, the second term simplifies to

1 2πi(2t) 2πi(t τ ) 1 3πiτ 2 e e − 2 2 2πit+πiτ e 2πi(2t) τ dt = e q e dt = q, 1 2πi(t ) 1 − − e2πit eπiτ − − − 2 1 e − 2 2 Z− − Z− − POLAR HARMONIC MAAß FORMS AND HOLOMORPHIC PROJECTION 19 and we arrive at

2 2 J(τ)= q qn +2n+1 = qn , − − n 1 − n 1 X≥ X≥ as claimed. 

τ 1 ∗ Proof of Proposition 1.2 (ii): We follow the original proof. Let z∗ = 2 2 and Pz be the − − 2 fundamental parallelogram around 0. Then Pz∗ contains no other zeros of ϑ(z; τ) . In the notation of [7] we set

+ n2 2n ϑ 2,0,1(z; τ) := q ζ , − n 1 X≥ and the idea is to evaluate the integral + ϑ 2,0,1(z; τ) − I(τ) := 2 dz ∗ ϑ(z; τ) Z∂Pz 1 in two different ways. On one hand, the Laurent expansion of ϑ(z;τ)2 with respect to z around 0 has the shape

1 f 2(τ) f 1(τ) = − + − + O(1), ϑ(z; τ)2 (2πiz)2 2πiz and thus + + ϑ 2,0,1( ; τ) d 2 ϑ 2,0,1(z; τ) I(τ) = 2πiRes 0, − · = 2πi z − ϑ( ; τ)2 ! dz ϑ(z; τ)2 ! · z=0

+ 1 d + n2 n2 = f 1(τ)ϑ 2,0,1(0; τ)+ f 2(τ) ϑ 2,0,1(z; τ) = f 1(τ) q + 2f 2(τ) nq . − − − 2πi dz − − −   z=0 n 1 n 1 X≥ X≥

On the other hand, the integrand is one-periodic in z, and hence the integrals along the vertical edges cancel each other. Moreover, we have the elliptic transformation property

λ+µ πi(λ2τ+2λz) ϑ(z + λτ + µ; τ) = ( 1) e− ϑ(z; τ) − for any λ, µ Z. Combining yields ∈ z∗+1 + z∗+τ+1 + ϑ 2,0,1(z; τ) ϑ 2,0,1(z; τ) I(τ)= − dz − dz ∗ ϑ(z; τ)2 − ∗ ϑ(z; τ)2 Zz Zz +τ z∗+1 + + ϑ 2,0,1(z; τ) ϑ 2,0,1(z + τ; τ) = − − dz ∗ ϑ(z; τ)2 − ϑ(z + τ; τ)2 Zz ! z∗+1 + + ϑ 2,0,1(z; τ) 2πi(τ+2z) ϑ 2,0,1(z + τ; τ) = − e − dz. ∗ ϑ(z; τ)2 − ϑ(z; τ)2 Zz ! The last step is to compute

+ 2 + n2 2n (n+1)2 2(n+1) 2 ϑ 2,0,1(z; τ) qζ ϑ 2,0,1(z + τ; τ)= q ζ q ζ = qζ . − − − n 1 − n 1 X≥ X≥ Hence,

z∗+1 2πi(2z) 1 2πi(2t) e 2 e I(τ)= q 2 dz = 2 dt. z∗ ϑ(z; τ) 1 ϑ t τ ; τ Z Z− 2 − 2 This completes the proof.   20 JOSHUAMALES,ANDREASMONO,ANDLARRYROLEN

References [1] C. Alfes-Neumann, Kathrin Bringmann, J. Males, and M. Schwagenscheidt, Cycle integrals of meromorphic modular forms and coefficients of harmonic Maass forms, J. Math. Anal. Appl. 497 (2021), no. 2, 124898, 15. [2] G. E. Andrews, R. C. Rhoades, and S. Zwegers, Modularity of the concave composition generating function, Algebra Number Theory 7 (2013), no. 9, 2103–2139. [3] K. Bringmann, A. Folsom, K. Ono, and L. Rolen, Harmonic Maass forms and mock modular forms: theory and applications, American Mathematical Society Colloquium Publications, vol. 64, American Mathematical Society, Providence, RI, 2017. [4] K. Bringmann, B. Kane, and S. Zwegers, On a completed generating function of locally harmonic Maass forms, Compos. Math. 150 (2014), no. 5, 749–762. [5] K. Bringmann and C. Nazaroglu, A framework for modular properties of false theta functions, Res. Math. Sci. 6 (2019), no. 3, Paper No. 30, 23. [6] K. Bringmann and L. Rolen, Half-integral weight Eichler integrals and quantum modular forms, J. Number Theory 161 (2016), 240–254. [7] K. Bringmann, L. Rolen, and S. Zwegers, On the Fourier coefficients of negative index meromorphic Jacobi forms, Res. Math. Sci. 3 (2016), Paper No. 5, 9. [8] J. H. Bruinier and J. Funke, On two geometric theta lifts, Duke Math. J. 125 (2004), no. 1, 45–90. [9] J. H. Bruinier and M. Schwagenscheidt, Theta lifts for Lorentzian lattices and coefficients of mock theta functions, Math. Z. (2020), 1–25. [10] J. H. Bruinier, G. van der Geer, G. Harder, and D. Zagier, The 1-2-3 of modular forms, Universitext, Springer- Verlag, Berlin, 2008. Lectures from the Summer School on Modular Forms and their Applications held in Nordfjordeid, June 2004; Edited by Kristian Ranestad. [11] A. Dabholkar, S. Murthy, and D. Zagier, Quantum black holes, wall crossing, and mock modular forms (2012), https://arxiv.org/abs/1208.4074. to appear in Cambridge Monographs in Mathematical Physics. [12] M. Eichler and D. Zagier, The theory of Jacobi forms, Progress in Mathematics, vol. 55, Birkhäuser Boston, Inc., Boston, MA, 1985. [13] I. S. Gradshteyn and I. M. Ryzhik, Table of integrals, series, and products, Seventh, Elsevier/Academic Press, Amsterdam, 2007. Translated from the Russian, Translation edited and with a preface by Alan Jeffrey and Daniel Zwillinger, With one CD-ROM (Windows, Macintosh and UNIX). [14] E. Hecke, Theorie der Eisensteinschen Reihen höherer Stufe und ihre Anwendung auf Funktionentheorie und Arithmetik, Abh. Math. Sem. Univ. Hamburg 5 (1927), no. 1, 199–224 (German). [15] Ö. Imamoğlu, M. Raum, and O. K. Richter, Holomorphic projections and Ramanujan’s mock theta functions, Proc. Natl. Acad. Sci. USA 111 (2014), no. 11, 3961–3967. [16] H. Iwaniec, Topics in classical automorphic forms, Graduate Studies in Mathematics, vol. 17, American Math- ematical Society, Providence, RI, 1997. [17] R. J. Lemke-Oliver, Eta-quotients and theta functions, Adv. Math. 241 (2013), 1–17. [18] G. Mersmann, Holomorphe η-produkte und nichtverschwindende ganze modulformen für Γ0(N), Diploma Thesis, Universität Bonn, 1991. [19] M. H. Mertens, K. Ono, and L. Rolen, Mock modular Eisenstein series with Nebentypus, Int. J. Number Theory (special issue for Bruce Berndt’s 80th birthday conference). to appear. [20] M. H. Mertens, Eichler-Selberg type identities for mixed mock modular forms, Adv. Math. 301 (2016), 359–382. [21] , Mock modular forms and class number relations, Res. Math. Sci. 1 (2014), Art. 6, 16. [22] J. Rouse and J. J. Webb, On spaces of modular forms spanned by eta-quotients, Adv. Math. 272 (2015), 200–224. [23] William Stein, Modular forms, a computational approach, Graduate Studies in Mathematics, vol. 79, American Mathematical Society, Providence, RI, 2007. With an appendix by Paul E. Gunnells. [24] D. Zagier, Nombres de classes et formes modulaires de poids 3/2, C. R. Acad. Sci. Paris Sér. A-B 281 (1975), no. 21, Ai, A883–A886 (French, with English summary). [25] , Quantum modular forms, Quanta of maths, Clay Math. Proc., vol. 11, Amer. Math. Soc., Providence, RI, 2010, pp. 659–675. [26] S. Zwegers, Mock theta functions, Ph.D. Thesis, Universiteit Utrecht, 2002. [27] , Multivariable Appell functions and non-holomorphic Jacobi forms, Res. Math. Sci. 6 (2019), no. 1, Paper No. 16, 15.

450 Machray Hall, Department of Mathematics, University of Manitoba, Winnipeg, Canada Email address: [email protected]

Department of Mathematics and Computer Science, Division of Mathematics, University of Cologne, Weyertal 86-90, 50931 Cologne, Germany Email address: [email protected]

Department of Mathematics, 1420 Stevenson Center, Vanderbilt University, Nashville, TN 37240 Email address: [email protected]