arXiv:1912.01121v4 [hep-th] 13 Jan 2021 enee nti ae.Ti eto snvrmatt earve fS of review a be w to results, the meant introduce never only is [ will we paper section So original This the one. to brief a paper. even not this theory, in needed Se be the modula study theory. e.g. the to theory, into Seibe number Theorem, insights the new from Picard’s provide tools called Great which the now and apply is derivative will work Schwarzian we beautiful paper, th Their this with In together singularities. theory the certain of near properties holomorphic the using by agMlster a locnutterve aes[ papers review the consult also can theory Yang-Mills the EBR-ITNTER N OUA ABAFUNCTION LAMBDA MODULAR AND THEORY SEIBERG-WITTEN is,ltu reyitouesm elkonrslso Seiberg-Wit of results well-known some introduce briefly us let First, ntefnaetlppr[ paper fundamental the In References prospects space further moduli and Acknowledgments the Conclusions of limit curvature non-perturbative scalar 8. the The period and Seiberg-Witten Theorem the Picard’s 7. of Great derivative periods Schwarzian its The 6. and functions oneform theta Seiberg-Witten and meromorphic 5. function The lambda modular The curves 4. elliptic of family Legendre 3. The 2. .Introduction 1. N uesmercYn-il hoywt ag ru U2 a be has SU(2) group gauge with theory Yang-Mills supersymmetric 2 = Abstract. smttclyfltna h etraielimit. perturbative the near comput the flat also of will asymptotically We space moduli functions. theta the the of by given is constant coupling otesuyo h ebr-itnter.W ilepesteholo the express will We theory. Seiberg-Witten the of study the to ilso h cwrindrvtv ftequotient the of derivative Schwarzian the show will h eid fteLgnr aiyo litccre.Tu ewl eab be will we Thus curves. elliptic of quotient the family of Legendre transformations the of periods the hc eeaetelattice the generate which nti ae,w ilapytetosfo ubrter n modular and theory number from tools the apply will we paper, this In 12 .Teraeswoaentfmla ihthe with familiar not are who readers The ]. 12 N Z ySiegadWte,telweeg ffcieato of action effective low-energy the Witten, and Seiberg by ] = uesmercYn-il hoy hc ssont be to shown is which theory, Yang-Mills supersymmetric 2 = a n D e 1. a /a + EZEYANG WENZHE ne h cino h oua ru PSL(2 group modular the of action the under Introduction n Contents m a D , 1 ( n e n , m a ) D ∈ /a 1 , Z ihrsett h complexified the to respect with 2 2 ]. fcnrlcags ntrsof terms in charges, central of opi functions morphic i oorm behaviors monodromy eir h clrcurvature scalar the e ielaealtedetails the all leave hile N et opt the compute to le br-itntheory, iberg-Witten supersymmetric 2 = e hoyta will that theory ten abafunction, lambda r gWte theory. rg-Witten ndetermined en , eiberg-Witten Z ,a a, .We ). forms D , 16 16 16 14 14 11 9 6 5 1 The N = 2 SUSY multiplet contains a massless gauge field Aµ, two Weyl fermions λ and ψ, and a scalar field φ, all of which live in the adjoint representation of the gauge group SU(2). The classical potential of the pure N = 2 theory is given by 1 V (φ)= Tr[φ,φ†]. (1.1) 2 Unbroken SUSY will require that we have V (φ) = 0 (1.2) in the vacuum. But this does not mean that φ itself is zero, since it is sufficient that if φ and φ† commute. Up to a gauge transformation, the scalar field φ can be taken to be 1 φ = aσ , (1.3) 2 3 where σ3 is the third Pauli matrix. For a given vacuum, a must be a constant. But a gauge transformation from the Weyl group of SU(2) can still transform a to a, so they are gauge − equivalent. Hence the gauge invariant quantity that parametrizes the inequivalent vacua is 1 u = a2 = Tr φ2. (1.4) 2 The space of gauge inequivalent vacua is also called the moduli space M of the N = 2 theory, which has u as its coordinate. Naively, one can consider M as the complex plane C. However, this moduli space M has special singularities, and the behaviors of the N =2 theory near these singularities will determine the low-energy effective action [12]. More explicitly, the low energy effective Lagrangian of the N = 2 supersymmetric Yang- Mills theory is completely determined by a prepotential (a), which is a multi-valued holo- F morphic function. The effective complexified coupling constant of the N = 2 theory is given by ∂2 τ (a)= F . (1.5) sw ∂a2 While a natural choice of metric for the moduli space M is 2 (ds) = Im τsw(a) dada, (1.6) which works at least in a small neighborhood of the perturbative limit u = . The dual ∞ variable aD is given by a = ∂ /∂a, (1.7) D F and the complexified coupling constant satisfies da τ (a)= D . (1.8) sw da In particular, the metric (ds)2 on the moduli space can also be expressed as i (ds)2 = Im da da = (da da dada ) . (1.9) D −2 D − D The key insight of Seiberg and Witten is that the multi-valued holomorphic functions a(u) and aD(u) can be realized as the period integrals of a meromorphic oneform defined for a 2 family of elliptic curves [12]. More explicitly, let be the family of elliptic curves defined Eu by the cubic equation : y2 =(x 1)(x + 1)(x u). (1.10) Eu − − On every , there is a meromorphic oneform Ωsw, called the Seiberg-Witten Eu u oneform, that is defined by √2 xdx udx Ωsw = . (1.11) u 2π  y − y  sw The residue of Ωu at its pole is zero, therefore it defines an element of the cohomology group H1( , C)[12]. The underlying differentiable manifold of is the torus. From the paper Eu Eu [12], there exists a basis γsw,γsw for the homology group H ( , Z) such that { 0 1 } 1 Eu sw sw a(u)= Ωu , aD(u)= Ωu . (1.12) sw sw Zγ0 Zγ1 In fact, the mass-lattice of the BPS states of the N = 2 theory is given by n a + n a :(n , n ) Z2 . (1.13) { e m D e m ∈ } Furthermore, the complexified coupling constant τsw(u) is just the period of the elliptic curve u that is given by E da /du τ (u)= D . (1.14) sw da/du

The quotient aD/a plays an important role in the Seiberg-Witten theory, especially in the study of the wall-crossing of BPS states [1, 2, 12]. The motivation of this paper is that since the transformation x 2x 1, y 2√2 y, u 2λ 1, (1.15) 7→ − 7→ 7→ − sends the family (1.10) to the Legendre family of elliptic curves E : y2 = x(x 1)(x λ), (1.16) λ − − could the tools in number theory provide any new insights into the Seiberg-Witten theory? We will see the answer is yes! The most significant result we have obtained through this approach is that the Schwarzian derivative of the quotient aD/a with respect to the complexified coupling constant τsw is given by the theta functions π2 a /a, τ = θ8(0, eπiτsw ). (1.17) { D sw} 2 4 Recall that the Schwarzian derivative of a holomorphic function f(z) with respect to the variable z is defined by f ′′(z) ′ 1 f ′′(z 2 f ′′′(z) 3 f ′′(z) 2 f, z = = . (1.18) { }  f ′(z)  − 2 f ′(z) f ′(z) − 2  f ′(z)  The Schwarzian derivative has many important properties [9, 11], e.g. (1) It is invariant under the M¨obius transformation pf + q p q , z = f, z , GL(2, C). (1.19) rf + s  { }  r s  ∈ 3 (2) If w is a function of z, then we have dw 2 f, z = f,w + w, z . (1.20) { } { }  dz  { } (3) If w is related to z by a M¨obius transformation pz + q p q w = , GL(2, C), (1.21) rz + s  r s  ∈ then we have (ps qr)2 f, z = f,w − . (1.22) { } { } (rz + s)4 (4) Given two holomorphic functions f(z) and g(z), we have pg(z)+ q p q f, z = g, z f(z)= for some GL(2, C). (1.23) { } { } ⇐⇒ rg(z)+ s  r s  ∈ These properties of the Schwarzian derivative immediately tell us that the equation (1.17) is an essential equation that characterize the central charge lattice (1.13), and more impor- tantly the Seiberg-Witten theory. This equation prompts us to ask whether there exists an automorphic representation that can be constructed from the the quotient aD/a which is related to the counting of the BPS states of the N = 2 supersymmetric Yang-Mills the- ory. It is also very interesting to find out whether the equation (1.17) admits any physical interpretations. In this paper, we will also apply the Great Picard’s Theorem to study the behaviors of the quotient a /a near the non-perturbative limit u = 1, based on which we prove that D ± the quotient aD/a indeed can take real values. Even though this is already a well-known property, but a mathematical proof could at least serve as an assurance. We will also compute the scalar curvature of the metric (1.6) on the moduli space M . We will show that near the perturbative limit u = , the scalar curvature approaches 0 exponentially, while near ∞ the non-perturbative limits u = 1, the scalar curvature approaches infinity exponentially. ± Therefore near the perturbative limit u = , the moduli space M is asymptotically flat. ∞ The outline of this paper is as follows. In Section 2, we will review some well-known results of the Legendre family of elliptic curves, e.g. its periods and the Picard-Fuchs equation. In Section 3, we will discuss the properties of the modular lambda function and the theta functions. In Section 4, we will look at the meromorphic Seiberg-Witten oneform and its periods. We will find out their relations to the periods of the Legendre family of elliptic curves. In Section 5, we will study the transformations of the periods of the meromorphic Seiberg-Witten oneform under the action of the modular group PSL(2, Z). We will show the Schwarzian derivative of the quotient aD/a with respect to the complexified coupling constant τsw is given by the theta functions. In Section 6, we will apply the Great Picard’s Theorem to study the behaviors of the quotient a /a near the non-perturbative limit u = 1. D ± In Section 7, we will compute the scalar curvature for the moduli space M of the N = 2 theory. In Section 8, we will summarize the results of this paper and raise several interesting open questions. 4 2. The Legendre family of elliptic curves In this section, we will review some well-known results about the Legendre family of elliptic curves, which are very elementary. They are included here solely to make this paper as self-contained as possible. The Legendre family of elliptic curves is defined by the cubic equation E : y2 = x(x 1)(x λ), (2.1) λ − − where λ is a free parameter. The singular fibers of this family are over the points λ =0, 1, . ∞ When λ is 0 or 1, this family degenerates to a nodal cubic. Let us now briefly recall how to construct a complex curve from the equation (2.1) using branch cuts [5]. First, cut the complex plane C along two lines: 0 to λ and 1 to . Second, take a second copy of the ∞ complex plane C and cut it along the same lines. Then, glue the two copies of the complex planes along the corresponding branch cuts. The result is a smooth torus, which has a complex structure induced by the cubic equation (2.1). On every smooth elliptic curve Eλ, there exists a nowhere vanishing oneform Ωλ which on the open locus where y = 0 is defined by 6 dx Ω = . (2.2) λ 2πy It is a straightforward exercise to check that dx/(2πy) actually extends to a nowhere vanish- ing oneform on Eλ [5]. We now construct a basis γ0,γ1 for the homology group H1(Eλ, Z), 2 { } which is isomorphic to Z . Let γ0 be a cycle on one copy of C that encircles the branch cut (1, ); while let γ be a circle that is the composite of two lines: the line from 1 to λ on ∞ 1 the first copy of C and the line from λ to 1 on the second copy of C. The integration of the oneform Ω over the basis γ ,γ gives us two periods ̟ (λ) and ̟ (λ) λ { 0 1} 0 1

̟0(λ)= Ωλ, ̟1(λ)= Ωλ, (2.3) Zγ0 Zγ1 which are multi-valued holomorphic functions. The computations of ̟0(λ) and ̟1(λ) are certainly very well-known and have a very long history. Up to a sign, they are given by the integrals 1 ∞ dx ̟0(λ)= , π Z1 x(x 1)(x λ) − − (2.4) 1 λ p dx ̟1(λ)= . π Z1 x(x 1)(x λ) − − After a change of variable by x =1/z, the firstp integral in formula (2.4) becomes 1 1 dz ̟0(λ)= . (2.5) π Z0 z(1 z)(1 λz) − − When λ is in a small neighborhood of 0,p the factor (1 λz)−1/2 admits a power series − expansion in λz. Then the integration (2.5) becomes ∞ 1/2 2 ̟ (λ)= − λn, (2.6) 0  n  Xn=0 5 which is the hypergeometric function 2F1(1/2, 1/2; 1; λ)[6]. The period ̟1(λ) can be com- puted similarly, and it admits an expansion with leading term given by 1 1 dx 1 ̟1(λ)= = (4 log 2 log λ)+ , (2.7) −πi Zλ x(1 x)(x λ) −πi − ··· − − where the limit of the terms inp are zero when λ 0. Thus, we deduce that the period ··· → ̟1(λ) must be of the form 1 log 16 ̟ (λ)= (̟ (λ) log λ + h(λ)) ̟ (λ), (2.8) 1 πi 0 − πi 0 where h(λ) is a holomorphic function in a small neighborhood of λ = 0 and it admits a power series expansion of the form [5] 1 21 185 h(λ)= λ + λ2 + λ3 + . (2.9) 2 64 768 ··· The derivative of the oneform Ωλ with respect to λ is given by d Ω 1 dx λ = , (2.10) dλ 4π x(x 1)(x λ)3 − − p which is a meromorphic oneform on Eλ with a pole at the point (x = λ,y = 0). However, 1 its residue at this pole is zero, hence it defines an element of H (Eλ, C) through integration. ′ 1 E C Moreover, Ωλ, Ωλ form a basis for the two dimensional vector space H ( λ, ). The second { } 2 2 1 derivative of Ωλ, i.e. d Ω/dλ , also defines an element of H (Eλ, C), hence it must be a linear ′ combination of Ωλ and Ωλ. From the Chapter 1 of [5], the oneform Ωλ satisfies a second order ordinary differential equation (ODE) d2Ω d Ω 1 λ(1 λ) + (1 2λ) Ω=0, (2.11) − dλ2 − dλ − 4 which is called the Picard-Fuchs equation of Ωλ. Therefore, the periods ̟i(λ), i =0, 1 also satisfy this ODE 1 λ(1 λ) ̟′′(λ)+(1 2λ) ̟′(λ) ̟ (λ)=0, i =0, 1. (2.12) − i − i − 4 i In fact, ̟ (λ),̟ (λ) form a basis for the solution space of the Picard-Fuchs equation { 0 1 } (2.11).

3. The modular lambda function and theta functions In this section, we will review the properties of the modular lambda function, the periods ̟i(λ), i = 0, 1 and the theta functions. All the results in this section have been known since the 19th century, and the readers who are familiar with them can skip this section completely. From the computations in Section 2, the monodromy of the periods ̟ (λ),̟ (λ) near { 0 1 } the singular point λ = 0 is ̟ (λ) ̟ (λ), ̟ (λ) ̟ (λ)+2 ̟ (λ). (3.1) 0 → 0 1 → 1 0 Let the column vector ̟ be ⊤ ̟ =(̟0,̟1) , (3.2) 6 then the monodromy matrix of ̟ near λ = 0 is 1 0 T = . (3.3) 0 2 1 The monodromy matrices of the period vector ̟ near the three singular points 0, 1, { ∞} generate the modular group Γ(2). The readers are referred to the book [5] for more details. The period τ of the elliptic curve Eλ is given by the quotient ̟ (λ) τ = 1 . (3.4) ̟0(λ) In a small neighborhood of λ = 0, τ has an explicit expansion of the form 1 h(λ) log 16 τ(λ)= log λ + . (3.5) πi  ̟0(λ) − πi The elliptic curve E is isomorphic to the quotient of C by the lattice generated by 1, τ λ { } [5]. The inverse of τ(λ), i.e. λ(τ), is the famous modular lambda function. Moreover, λ(τ) generates the function field of the X(2), i.e. it is a Hauptmodul for X(2) [6]. The index of Γ(2) in PSL(2, Z) is 6 and the quotient group PSL(2, Z)/Γ(2) is generated by 1 T : τ τ +1, S : τ . (3.6) 7→ 7→ −τ The modular curve X(2) = Γ(2) H is a sixfold cover of PSL(2, Z) H [7]. \ \ Let the variable q be defined by q := exp πiτ. (3.7) From formula (3.5), we immediately obtain 1 q = λ exp (h(λ)/̟ (λ)) . (3.8) 16 0 This equation can be inverted order by order, and the result is the power series expansion of λ with respective to q [5] λ(q)=16q 128q2 + 704q3 3072q4 + . (3.9) − − ··· The modular lambda function can also be expressed as a product of the theta functions [6] 4 4 θ2(0, q) θ4(0, q) λ(q)= 4 , 1 λ(q)= 4 ; (3.10) θ3(0, q) − θ3(0, q) where we have used the Jacobi identity 4 4 4 θ3(0, q)= θ2(0, q)+ θ4(0, q). (3.11) Following the mirror symmetry of Calabi-Yau threefolds [8, 10], let us define the normalized Yukawa coupling by Y 1 dΩ := 2 Ω dλ, (3.12) Y ̟0(λ)  ZEλ ∧ dλ  which is a oneform defined on C 0, 1 . In fact, the Yukawa coupling is just dτ −{ } Y 1 d̟1(λ) d̟0(λ) d(̟1(λ)/̟0(λ)) = 2 ̟0(λ) ̟1(λ) dλ = dλ = dτ. (3.13) Y ̟0(λ)  dλ − dλ  dλ 7 Now define Wk to be dkΩ Wk = Ω k , k =0, 1, 2, . (3.14) ZEλ ∧ dλ ··· From the definition, we immediately have dW W =0, W = 1 . (3.15) 0 2 dλ The Picard-Fuchs equation (2.11) implies dW λ(λ 1) 1 + (2λ 1) W =0, (3.16) − dλ − 1 a general solution of which is of the form C W = , (3.17) 1 λ(1 λ) − Here C is a nonzero constant, which can be shown to be 1/πi by an explicit computation using the series expansion of ̟ (λ), i = 0, 1. Hence the normalized Yukawa coupling is i Y also equal to 1 dλ = . (3.18) Y πi ̟2(λ) λ(1 λ) 0 − Then formula (3.13) immediately implies dτ 1 πi = , (3.19) dλ ̟2(λ) λ(1 λ) 0 − which is equivalent to 1 dλ = ̟2(λ) λ(1 λ). (3.20) πi dτ 0 − Let us now prove a well-know result [4, 13]. Lemma 3.1. 2 ̟0(λ(q)) = θ3(0, q). (3.21) Proof. From formula (3.20), we have 1 dλ 1 ̟2(λ)= . (3.22) 0 πi dτ λ(1 λ) − The actions of the two generators T and S of PSL(2, Z) on λ are given by λ T : λ , 7→ λ 1 (3.23) S : λ 1 −λ. 7→ − 2 From formula (3.22), the actions of T and S on ̟0(λ) are given by T : ̟2(λ) (1 λ) ̟2(λ), 0 7→ − 0 2 2 2 (3.24) S : ̟0(λ) τ ̟0(λ). 7→ −8 4 But the actions of T and S on the θ3(0, q) are [6] 4 4 4 T : θ3(0, q) θ4(0, q)=(1 λ) θ3(0, q), 7→ − (3.25) S : θ4(0, q) τ 2 θ4(0, q); 3 7→ − 3 where we have used formula (3.10). Since ̟0(λ(q)) does not vanish on the upper half plane, 4 2 H the quotient θ3(0, q)/̟0(λ(q)) is a holomorphic function on that is invariant under the actions of PSL(2, Z). Near the cusp i , the q-expansion of θ2(0, q)/̟ (λ(q)) is given by ∞ 3 0 2 4 θ3(0, q)/̟0(λ(q))=1+ O(q ), (3.26) 2  hence we deduce θ3(0, q)/̟0(λ(q)) must be 1. The upshot is that we have the following identities θ4(0, q)= λ̟2(λ(q)), θ2(0, q)= ̟ (λ(q)), θ4(0, q)=(1 λ) ̟2(λ(q)), (3.27) 2 0 3 0 4 − 0 which will be important in this paper.

4. The meromorphic Seiberg-Witten oneform and its periods

sw In this section, we will study the meromorphic Seiberg-Witten oneform Ωu (1.11) and its periods. Under the transformation (1.15) which sends the family u (1.10) to the Legendre E E sw family λ (2.1) of elliptic curves, the meromorphic Seiberg-Witten oneform Ωu (1.11) is sent to 1 xdx λdx Ωsw = . (4.1) λ π  y − y  sw Notice that Ωλ can also be expressed as Ωsw =2(x Ω λ Ω ), (4.2) λ λ − λ where Ωλ is the nowhere vanishing oneform (2.2) on the elliptic curve Eλ. sw Let us now show that up to a total differential, Ωλ is equal to 2 Ωsw =4 λ(λ 1) Ω′ + df, (4.3) λ − λ π where f is a meromorphic function on the elliptic curve Eλ f := x1/2(x 1)1/2(x λ)−1/2. (4.4) − − The total differential of f is 1 1 df = x−1/2(x 1)1/2(x λ)−1/2dx + x1/2(x 1)−1/2(x λ)−1/2dx 2 − − 2 − − (4.5) 1 x1/2(x 1)1/2(x λ)−3/2dx, − 2 − − ′ whose residues at the singularities are all 0. The oneform Ωλ and its derivative Ωλ are

1 −1/2 −1/2 −1/2 Ωλ = x (x 1) (x λ) dx, 2π − − (4.6) 1 Ω′ = x−1/2(x 1)−1/2(x λ)−3/2 dx, λ 4π − − 9 from which we deduce 1 1 1 df = (x 1) Ω + x Ω x (x 1) Ω′ . (4.7) 2π 2 − λ 2 λ − − λ To simplify this formula, we will need the identity 1 (x λ) Ω′ = Ω , (4.8) − λ 2 λ which tells us 1 1 x (x 1) Ω′ = x Ω +(λ 1)( Ω + λ Ω′ ). (4.9) − λ 2 λ − 2 λ λ Hence formula (4.7) becomes 1 1 1 df = x Ω λ Ω λ(λ 1) Ω′ , (4.10) 2π 2 λ − 2 λ − − λ from which we immediately obtain the formula (4.3). The integrals of Ωsw with respect to the homology basis γ ,γ of E constructed in λ { 0 1} λ Section 2 are given by

′ ′ π0(λ)= 4 λ(λ 1) Ω =4 λ(λ 1) ̟ (λ), Z − λ − 0 γ0 (4.11) ′ ′ π1(λ)= 4 λ(λ 1) Ωλ =4 λ(λ 1) ̟1(λ), Zγ1 − − where we have used formula (2.3). From the Picard-Fuchs equation (2.11), we deduce (π )′ = ̟ , (π )′ = ̟ . (4.12) 0 − 0 1 − 1 The periods πi(λ), i = 0, 1 will be called the BPS periods in this paper. Since the periods a , a (1.12) are given by the integrals of the Seiberg-Witten oneform Ωsw (1.11) over an { D } u integral basis of H ( , Z), hence they are related to π , π by an integral linear transfor- 1 Eu { 0 1} mation. In particular, the central charge lattice of the BPS states of the N = 2 theory is also given by m π + n π :(m, n) Z2 . (4.13) { 0 1 ∈ } After studying the behaviors of the limits of π , π near the singularities λ =0, 1, and { 0 1} ∞ compare them with that of a, a near u = 1, , which have been computed in [12], we { D} ± ∞ find that [12] a = π π , a = π . (4.14) − 1 − 0 D − 1 Hence from formula (1.14), the complexified coupling constant τsw of the N = 2 theory is also given by ′ π1 ̟1 τ τsw = ′ ′ = = , (4.15) π0 + π1 ̟0 + ̟1 τ +1 where we have used formula (4.12). ′ The derivatives ̟i(λ), i = 0, 1 are two linearly independent solutions for the hypergeo- metric differential equation d2(̟′) d(̟′) 9 λ(1 λ) i + (2 4λ) i ̟′ =0. (4.16) − dλ2 − dλ − 4 i 10 ′ In fact, ̟0(λ) is the hypergeometric function 1 3 3 ̟′ (λ)= F ( , ; 2; λ). (4.17) 0 4 2 1 2 2

On the other hand, the BPS periods πi(λ), i =0, 1 satisfy the following ODE d2π i + Q(λ)π =0, i =0, 1; (4.18) dλ2 i where Q(λ) is the rational expression 1 Q(λ)= . (4.19) −4λ(1 λ) − The ODE 4.18 is the Q-form of the hypergeometric differential equation 4.16. This Q-form is closely related to Schwarzian derivative [9]. More explicitly, we will further define the Seiberg-Witten period ϕ to be the quotient

ϕ(λ) := π1(λ)/π0(λ), (4.20) which is also equal to ′ ′ ϕ(λ)= ̟1(λ)/̟0(λ). (4.21)

While the quotient aD/a can be expressed as a π ϕ D = 1 = . (4.22) a π0 + π1 ϕ +1 There is a fundamental relation between the Schwarzian derivative and ODE in the complex plane, which immediately tells us that [9] ϕ,λ =2Q(λ). (4.23) { } However, we will see that it is the Schwarzian derivative of ϕ with respect to the variable τ of the upper half-plane H that is going to be more interesting and important in this paper [11].

5. The Schwarzian derivative of the Seiberg-Witten period In this section, we will first apply the results in Section 3 to obtain the transformations of the BPS periods π , π under the action of PSL(2, Z). Then we will study the Schwarzian { 0 1} derivative of the Seiberg-Witten period ϕ (4.20) with respect to τ, which is given by the .

5.1. The transformations of the BPS periods. From formula (3.4), the derivative of ̟1 with respect to λ can also be written as dτ 1 1 ̟′ =(τ̟ )′ = ̟ + τ̟′ = + τ̟′ , (5.1) 1 0 dλ 0 0 πi λ(1 λ)̟ 0 − 0 where we have used formula (3.19). Under the action of S : τ 1/τ, we have 7→ − 1 S : λ 1 λ; dλ dλ; ̟ iτ̟ = i̟ ; ̟ = τ̟ iτ̟ = i̟ ; (5.2) 7→ − 7→ − 0 7→ 0 1 1 0 7→ −τ 0 − 0 11 hence the transformations of π , π under S are given by { 0 1} S : π0 i π1, 7→ − (5.3) S : π i π . 1 7→ 0 Therefore the Seiberg-Witten period ϕ (4.20) transforms according to 1 S : ϕ . (5.4) 7→ −ϕ The transformation of π , π under the action of T is more complicated. Under the { 0 1} action of the transformation T : τ τ + 1, we have 7→ λ 1 T : λ , dλ dλ, ̟ √1 λ̟ ; (5.5) 7→ λ 1 7→ −(λ 1)2 0 7→ − 0 − − from which we obtain

2 λ ′ T : π0 2(λ 1)̟ + ̟0 , 7→ √1 λ − 0 −  (5.6) 2 λ dτ ′ T : π1 2(λ 1) + τ +1 ̟0 + 2(λ 1)(τ + 1)̟ . 7→ √1 λ  − dλ  − 0 − However under the action of T 2 : τ τ + 2, we have 7→ T 2 : λ λ, dλ dλ, ̟ ̟ , (5.7) 7→ 7→ 0 7→ 0 hence the BPS periods π , π transform in the way { 0 1} 2 T : π0 π0, 7→ (5.8) T 2 : π π +2 π . 1 7→ 1 0 Thus the Seiberg-Witten period ϕ (4.20) transforms in the way T 2 : ϕ ϕ +2. (5.9) 7→ The upshot is that ϕ is equivariant with respect to the actions of S and T 2.

5.2. The fixed point of the S-transformation. The point λ = 1/2, i.e. u = 0, has an interesting property. Recall that the period τ of Eλ can also be expressed as [6] F (1/2, 1/2;1;1 λ) τ = i 2 1 − , (5.10) 2F1(1/2, 1/2; 1; λ) hence the value of τ at λ =1/2 is τ = i. (5.11) |λ=1/2 The elliptic curve E1/2 E : y2 = x(x 1)(x 1/2) (5.12) 1/2 − − admits CM (complex multiplication), and its j-invariant is 1728. From the formula (5.1), the Seiberg-Witten period ϕ can also be expressed as 1 1 ϕ = τ + ′ . (5.13) πi λ(1 λ)̟0̟0 12 − From the following special values of hypergeometric functions Γ(1/4) F (1/2, 1/2; 1; λ) = , 2 1 |λ=1/2 (2π)1/2Γ(3/4) (5.14) Γ( 1/4) F (3/2, 3/2; 2; λ) = − , 2 1 |λ=1/2 −(2π)1/2Γ(5/4) we immediately obtain that the value of ϕ at λ =1/2 is i. Thus the central charge lattice of − the BPS states (4.13) has an additional symmetry. It is very interesting to find out whether this symmetry has any physical significance.

5.3. The Schwarzian derivative. Inspired by the paper [11], let us now compute the Schwarzian derivative of the Seiberg-Witten period ϕ with respect to the variable τ. From the results in Section 5.1, ϕ(τ) is equivariant with respect to the modular group Γ(2). First, ϕ can be expressed as

′ ′ ̟1 1 1 ̟0 h ϕ = ′ = log λ + ′ + ′ log 16 . (5.15) ̟0 πi  λ ̟0 ̟0 −  An explicit computation of the first several terms of the q-expansion of ϕ, τ , the transfor- { } mation properties in the formula 1.22 and results in the paper [11] show that π2 ϕ, τ = θ8(0, q), q = exp(πiτ). (5.16) { } 2 3 We have also proved this equation by using the property (1.20) of the Schwarzian derivative, formula (4.23) and the results in Section 3. Furthermore, using the properties (1.19) and (1.22), we obtain π2 a /a, τ = θ8(0, eπiτsw ), (5.17) { D sw} 2 4 where we have used the equation

8 πiτ/(τ+1) 4 8 πiτ θ4 0, e =(τ + 1) θ3(0, e ). (5.18)  8 The function θ4(0, q) has an expansion of the form θ8(0, q)=1 16q + 112q2 448q3 + 1136q4 2016q5 + 3136q6 + . (5.19) 4 − − − ···

On the other hand, the modular group Γ0(2) has a unique weight-4 normalized entire modular form E0,4(τ) given by [7, 14] 16 8 E0,4(τ)= η (τ)/η (2τ). (5.20)

8 The modular form E0,4(τ) is related to θ4(0, q) by

8 θ4(0, q)= E0,4(τ/2). (5.21)

The readers are referred to the paper [3] for some applications of the modular form E0,4(τ). 13 6. Great Picard’s Theorem and the non-perturbative limit

In this section, we will prove that the quotient aD/a does take real values in a small neighborhood of the nonperturbative limits u = 1 using the Great Picard’s Theorem. In ± fact, we will prove an equivalent result: the Seiberg-Witten period ϕ (4.20) takes real values near λ =0, 1. First, let us look at the case where λ = 0. Define qsw by qsw := exp πi ϕ. (6.1) Thus ϕ is real if and only if the modulus of qsw is 1. Formula (5.15) implies

sw ′ ′ 1 ̟0 16 q = λ exp(h /̟0) exp ′ . (6.2) λ ̟0  ′ The power series expansion of ̟0/̟0 in a small neighborhood of λ = 0 is of the form 7 ̟ /̟′ =4 λ + O(λ2), (6.3) 0 0 − 2 from which we deduce that

sw ′ ′ 1 ̟0 16 q = F (λ) exp(4/λ), with F (λ)= λ exp(h /̟0) exp ′ 4 . (6.4) λ ̟0 −  The function F (λ) is holomorphic in a neighborhood of λ = 0 and it admits a series expansion with the first term given by F (λ)= λ + O(λ2). (6.5) The function F (λ) exp(4/λ) is holomorhpic in a small punctured neighborhood of 0, while it has an essential singularity at λ = 0. Now recall the following important theorem of Picard. Great Picard’s Theorem: If G(z) is a holomorphic function and z0 is an essential singularity of G(z), then on any punctured neighborhood of z0, G takes on all possible complex values, with at most a single exception, infinitely many times. In particular, the Great Picard’s Theorem implies that in every punctured neighbor- hood of λ = 0, which is assumed to be small enough to avoid other singularities, the function F (λ) exp(4/λ) takes on all possible values in exp πi t : t R , with at most a single excep- { ∈ } tion, infinitely many times. Therefore the behavior of ϕ in a small punctured neighborhood of λ = 0 is of a very complicated nature. The behavior of the Seiberg-Witten period ϕ at λ = 1 can be analyzed similarly, which is related to the λ = 0 case by an S-transformation.

7. The scalar curvature of the moduli space In this section, we will compute the scalar curvature of the moduli space M for the N =2 supersymmetric Yang-Mills theory. We will look at the behaviors of the scalar curvature near the singularities u = 1, . ± ∞ From the formulas (4.12) and (4.14), the metric (ds)2 (1.9) on the moduli space M can also be expressed as i (ds)2 = (̟ ̟ ̟ ̟ )dλdλ. (7.1) −2 1 0 − 0 1 14 We immediately recognize that this metric can be written down in a geometrically invariant way 2 i (ds) =( Ωλ Ωλ)dλdλ. (7.2) 2 ZEλ ∧ The pull-back of the metric (7.1) to the upper half plane H becomes dλ 2 (ds)2 = ̟ (Im τ)dτdτ, (7.3) 0 dτ

which is of a rather simple and explicit form. From the formula (3.20), we deduce that (ds)2 = π2 ̟3λ(1 λ) 2 (Im τ)dτdτ. (7.4) 0 −

The scalar curvature for the metric (ds )2 (7.4) is 1 R = . (7.5) π2 ̟3λ(1 λ) 2 (Im τ)3 | 0 − | The scalar curvature R is invariant under the action of S. While under the action of T , R transforms according to T : R 1 λ 3 R. (7.6) →| − | Moreover, R is invariant under the action of T 2. The scalar curvature R has three singular- ities at λ = 0, 1, . Let us now analyze the the behaviors of R near them. When τ = i , ∞ ∞ we have λ = 0. On the imaginary axis, τ = it, t R+, (7.7) ∈ when t goes to infinity, the leading term of R is 1 R exp(2πt), (7.8) ∼ 256π2t3 which approaches infinity exponentially. The behavior of R near λ =1(τ = 0), is the same as that of R near λ = 0, which can be deduced from the property that R is invariant under the S-transformation. Notice that the S-transformation maps a neighborhood of λ =0toa neighborhood of λ = 1. Now let us look at the behavior R in a small neighborhood of λ = . First do a T - ∞ transformation, which sends λ = to λ = 1. Next do an S-transformation, which sends ∞ λ =1 to λ = 0. Under the composition of these two transformations, R transforms to λ R λ 3R = | | . (7.9) →| | π2 ̟3(1 λ) 2 (Im τ)3 | 0 − | On the imaginary axis (7.7), when t is large, the leading part of λ 3R is | | 16 λ 3R exp( πt), (7.10) | | ∼ π2t3 − which approaches 0 exponentially. Hence we deduce that R approaches 0 exponentially when λ goes to infinity. So the moduli space M is asymptotically flat near λ = , which reflects ∞ the property that λ = (u = ) is a perturbative limit of the N = 2 supersymmetric ∞ ∞ Yang-Mills theory. 15 8. Conclusions and further prospects In this paper, we have applied the tools from number theory to the study of Seiberg-Witten theory. We have reviewed the properties of the periods of the Legendre family of elliptic curves, and their relations to the modular lambda function and the theta functions, which are included here in order to make this paper as self-contained as possible. Based on these results, we have studied the transformations of the Siberg-Witten period ϕ (4.20) under the action of the modular group PSL(2, Z). We have shown that the Schwarzian derivative of ϕ with respect to the period τ is given by the theta functions. This allows us to obtain the equation (1.17), i.e. the Scharzian derivative of the quotient aD/a with respect to the complexified coupling constant τsw of the N = 2 theory is also given by the theta functions. We have also studied the behaviors of the quotient aD/a near the non-perturbative limit u = 1 by using the Great Picard’s Theorem. In particular, we have shown that a /a ± D can take on any real values, with at most a single exception, infinitely many times. The scalar curvature for the moduli space of the N = 2 theory has also been computed. It has been shown that the scalar curvature approaches 0 exponentially near the perturbative limit u = , so the moduli space is asymptotically flat near u = . While the scalar curvature of ∞ ∞ the moduli space approaches infinity exponentially near the the two non-perturbative limit u = 1. ± There are several open questions left unanswered in this paper. Perhaps the most im- portant open question is could the equation (1.17) bridge any new relations between the Seiberg-Witten theory and the theory of automorphic representations. For example, could we construct an automorphic representation from the quotient aD/a which is related to the counting of the BPS states of the N = 2 supersymmetric Yang-Mills theory. It is also very interesting to see whether the equation (1.17) admits any physical interpretations. As has been shown in Section 5.2, the point u =0(λ =1/2) of the moduli space M is a special point. The lattice for the elliptic curve E1/2, which has j-invariant 1728, is m + ni :(m, n) Z2 . (8.1) { ∈ } While the central charge lattice of the N = 2 theory for the vacuum with coordinate u = 0 is isomorphic to this lattice. It is very interesting to see whether this property has any physical significance in the N = 2 supersymmetric Yang-Mills theory.

Acknowledgments The author is grateful to Shamit Kachru and Arnav Tripathy for many helpful discussions and a reading of the draft.

References [1] L. Alvarez-Gaum´eand S. F. Hassan, Introduction to S-Duality in N = 2 Supersymmetric Gauge Theories, arXiv:hep-th/9701069. [2] A. Bilal, DUALITY IN N = 2 SUSY SU(2) YANG-MILLS THEORY: A pedagogical introduction to the work of Seiberg and Witten. arXiv:hep-th/9601007. [3] R. E. Borcherds, Automorphic forms with singularities on Grassmannians, Invent math 132, 491–562 (1998). https://doi.org/10.1007/s002220050232. 16 [4] J. H. Bruinier, G. van der Geer, G. Harder and D. Zagier, The 1–2–3 of Modular Forms, Lectures at a Summer School in Nordfjordeid, Norway. Springer. [5] J. Carlson, S. M¨uller-Stach, and C. Peters. Period Mappings and Period Domains. Cambridge University Press. [6] K. Chandrasekharan, Elliptic Functions, Grundlehren der mathematischen Wissenschaften, 281, Springer-Verlag. [7] F. Diamond and J. Shurman, A First Course in Modular Forms, GTM, vol. 228, Springer (2005). [8] M. Gross, D. Huybrechts and D. Joyce, Calabi-Yau Manifolds and Related Geometries. Springer. [9] E. Hille, Ordinary differential equations in the complex domain, Dover Publications (1997). [10] M. Kim and W. Yang, Mirror symmetry, mixed motives and ζ(3). arXiv:1710.02344. [11] A. Sebbar and H. Saber, Automorphic Schwarzian equations. Forum Mathematicum — Volume 32: Issue 6. DOI: https://doi.org/10.1515/forum-2020-0025. arXiv: 2002.00493. [12] N. Seiberg, E. Witten, Electric-magnetic duality, monopole condensation, and confinement in N=2 supersymmetric Yang-Mills theory. Nucl.Phys. B426 (1994), 19-52. arXiv: hep-th/9407087. [13] D. Zagier, Arithmetic and topology of differential equations, in Proceedings of the Seventh European Congress of Mathematics (Berlin, July 18–22, 2016), Editors V. Mehrmann, M. Skutella, European Mathematical Society, Berlin, 2018, 717–776. [14] A035016, The On-Line Encyclopedia of Integer Sequences. https://oeis.org/A035016.

SITP, Physics Department, Stanford University, Stanford, CA, 94305 Email address: [email protected]

17