<<

Notes on SL(2f , R) representations

Alexei Kitaev California Institute of Technology, Pasadena, CA 91125, U.S.A. 21 August 2018

Abstract

These notes describe representations of the universal of SL(2, R) with a view toward applications in . on the hyperbolic and the two-dimensional anti-de Sitter are also discussed.

Contents ∼ 1 The G = PSL(2, R) and its universal cover3 2 Irreducible unitary representations4

3 Non-unitary continuous and discrete series4 3.1 The individual representations ...... 4 3.2 and intertwiners ...... 6

4 on Ge and the Plancherel measure8 4.1 Fourier transform on a group ...... 8 4.2 The L- and R-actions on Ge in explicit coordinates ...... 10 4.3 Casimir eigenfunctions ...... 11 4.4 The elements and Plancherel for the irreps of Ge ...... 13

5 Spinors on the hyperbolic plane and anti-de Sitter space 15 arXiv:1711.08169v2 [hep-th] 21 Aug 2018 2 5.1 The H , AdS2, and their complex ...... 15 5.2 Definitions of spinors and two standard gauges ...... 17 5.3 Spinors on H2 ...... 19

6 products of unitary irreps 23 6.1 D+ ⊗ D+ ...... 24 λ1 λ2 6.2 D+ ⊗ D− ...... 25 λ1 λ2

1 Introduction

The representations of the group SL(2, R) were originally studied by Bargmann [1], and an ex- plicit Plancherel formula was obtained by Harish-Chandra [2]. The representation for the universal cover of that group is largely due to Puk´anszky[3]. These notes are intended as an accessible (though not very rigorous) exposition, with notation and equations that would be convenient for general use. My main motivation comes from the study of the Sachdev-Ye-Kitaev (SYK) model [4,5,6]. This model has zero spatial , so the only coordinate is t. At az+b low temperatures, all correlation functions are under M¨obiustransformations z 7→ cz+d of the z = exp(2πt/β). If t is real, then z is also real, and thus, the group is PSL(2, R) = SL(2, R)/{1, −1}. It is, actually, more common to consider the problem in Euclidean time, t = −iτ. In this case, the symmetry transformations preserve the (and also the orientation). The corresponding group G is isomorphic to PSL(2, R). In general, one is interested in the action of G on functions of two points on the circle. Fermionic wave functions (of one point on the circle) are transformed under the double cover of that group, i.e. SL(2, R). Furthermore, the study of the SYK model in Lorentzian time requires the universal cover of SL(2, R) (or G), which is denoted by Ge. Finite-dimensional representations of Ge are similar to those for SU(2): they are defined by an or half-integer spin S. However, these representations (except for the trivial one) are not unitary. All nontrivial unitary representations are infinitely dimensional. We will discuss three types of elementary representations: 1. Irreducible unitary representations, which include continuous and discrete .

2. Forms of an arbitrary degree on the . These are, essentially, functions on the circle, whereas the degree λ (which can be any ) enters the transformation law under diffeomorphisms.

3. Holomorphic λ-forms on the unit disk. The last two are also called “non-unitary” continuous and discrete series because they do not have a Hermitian inner product as part of their definition. However, they include the unitary irreps as special cases. Unitary representations of Ge are sufficiently well-behaved; for example, an arbitrary splits into irreps with multiplicities. (We do not prove this result because it involves a great deal of .) Some differences from the more familiar case of compact Lie groups are as follows: • Irreducible unitary representations can be infinitely-dimensional.

• Some irreps do not enter the decomposition of the left . The irreps α in that decomposition are those whose matrix elements hl|Uα(g)|mi are normalizable or δ-normalizable as functions of the group g. The normalization factor is known as the on the of irreps. With physical applications in mind, we give explicit formulas for the matrix element functions and more general Casimir eigenfunctions. The latter may be interpreted as spinors on the anti-de Sitter space AdS2, or as functions (or forms) of two points on a circle.

2 ∼ 1 The group G = PSL(2, R) and its universal cover

First, let us consider the the M¨obius group PSL(2, C) = SL(2, C)/{1, −1}. It consists of all linear fractional maps az + b z 7→ , where a, b, c, d ∈ , ad − bc = 1. (1) cz + d C The corresponding Lie is denoted by sl2(C). Its general element is a traceless matrix δa δb δW = , which generates an infinitesimal change of z by δz = δb + (δa − δd)z − (δc)z2. δc δd One can write δW in a similar form, namely, δW = (δb)L−1 + (δa − δd)L0 − (δc)L1, where δa + δd = 0 and 0 1 1 1 0   0 0 L = ,L = ,L = . (2) −1 0 0 0 2 0 −1 1 −1 0

The elements Ln of the may be interpreted as complex vector fields with the following z-components and commutation relations:1

z n+1 Ln(z) = z , [Ln,Lm] = (n − m)Ln+m. (3)

Now, the PSL(2, R) consists of the M¨obiustransformations with real parameters a, b, c, d. Such maps preserve the upper half-plane. However, it is more convenient to work with the subgroup G of those transformations that preserve the unit disk |z| < 1. It is conjugate to z−i PSL(2, R) by the Cayley z 7→ z+i , which takes the upper half-plane to the unit disk. A of the corresponding Lie algebra g is as follows:

1 i 0  Λ = = iL 0 2 0 −i 0

1 0 1 L − L Λ = = −1 1 (4) 1 2 1 0 2

1  0 i iL + iL Λ = = −1 1 2 2 −i 0 2

One can also define g abstractly, using the Lie bracket:

[Λ0, Λ1] = Λ2, [Λ0, Λ2] = −Λ1, [Λ1, Λ2] = −Λ0. (5) a constant factor, the Casimir operator is 1 Q = Λ2 − Λ2 − Λ2 = −L2 + (L L + L L ). (6) 0 1 2 0 2 −1 1 1 −1

The Lie algebra g canonically defines the simply connected Ge, the universal cover of G. Conversely, G is the of Ge by its , Z = {e2πnΛ0 : n ∈ Z}.

1 β We define the action of a vector field v on functions by the formula f 7→ −∂vf = −v ∂βf. Correspondingly, [u, v] = − Lu v, where L denotes the Lie .

3 2 Irreducible unitary representations

An irreducible representation of Ge is characterized by the eigenvalues of the Casimir operator Q and the central element e2πΛ0 . The latter has the form e−2πiµ, where µ ∈ R/Z. Representations −2πiµ of G (rather than its universal cover) are characterized by e = 1. By unitarity, Λ0,Λ1,Λ2 are represented by anti-Hermitian operators; hence the Casimir eigenvalue q is real. It and can be parametrized in a way similar to the expression S(S + 1) in the su(2) case:

1 q = λ(1 − λ), where λ ∈ R or λ = 2 + is, s ∈ R. (7)

The basis vectors |mi are eigenvectors of Λ0 with the eigenvalues −im, where m ∈ µ+Z. Recall that Λ0 = iL0 (see Eq. (4)); hence L0|mi = −m|mi. The commutation relations [L0,L−1] = L−1 and [L0,L1] = −L1 imply that L−1 lowers, and L1 raises, m by 1. Using the [L1,L−1] = 2 1 2L0 and the expression for the Casimir operator, Q = −L0 + 2 (L−1L1 + L1L−1), we find that

hm|L1L−1|mi = m(m − 1) + q = (m − λ)(m − 1 + λ), (8)

hm|L−1L1|mi = m(m + 1) + q = (m + λ)(m + 1 − λ). (9)

† Since Ln = L−n due to unitarity, the matrix elements in the above equations are nonnegative for all values of m that occur in a given representation. This condition is satisfied by three types of irreducible representations:

I : µ = q = 0, m = 0;

µ q > |µ|(1 − |µ|), • Continuous series Cq : m = µ + k (k ∈ Z); where |µ| 6 1/2, (10) m = λ, λ + 1, λ + 2,... or • Discrete series D+, D− : λ > 0, µ = ±λ, λ λ m = −λ, −λ − 1, −λ − 2,....

µ We will also use the generic notation Uλ that encompasses all three cases. The continuous series µ 1 1 Cq is further subdivided into the principal series, q > 4 (i.e. λ = 2 +is) and complementary series, 1 1 q < 4 (i.e. |µ| < λ < 2 ). The general solution to equations (8), (9) involves some arbitrary phase factors, which can be absorbed in the definition of the basis vectors. Thus,

p L−1|mi = − (m − λ)(m − 1 + λ) |m − 1i,

L0|mi = −m |mi, (11) p L1|mi = − (m + λ)(m + 1 − λ) |m + 1i.

3 Non-unitary continuous and discrete series

3.1 The individual representations µ Let λ ∈ C and µ ∈ C/Z. The non-unitary continuous series representation Fλ consists of “µ- twisted λ-forms” on the unit circle, which may be written as f = f˜(ϕ)(dϕ)λ. More formally, an

4 abstract form f is represented by a f˜ of a real variable that satisfies a twisted periodicity condition and transforms under diffeomorphisms in a particular way:

f˜(ϕ + 2π) = e2πiµf˜(ϕ) (12) ˜ −1 λ ˜ −1 (V f)(ϕ) = ∂ϕV (ϕ) f(V (ϕ)) (13)

1 Here V is an element of the group Diffg +(S ), that is, a smooth monotone map V : R → R such that V (ϕ + 2π) = V (ϕ) + 2π. Infinitesimal diffeomorphisms, i.e. vector fields, act as follows:

˜ ϕ ˜ ϕ ˜ (vf)(ϕ) = −v ∂ϕf(ϕ) − λ (∂ϕv ) f(ϕ). (14)

For example, if λ = −1, then f is a vector field, and vf is equal to − Lv f. One can also 1 characterize the diff(S ) action by applying Eq. (14) to the complex vector fields Ln (expressed ϕ −1 z inϕ as Ln = (∂z/∂ϕ) Ln = −ie ) and the Fourier basis of the space of forms:

˜ imϕ Lnfλ,m = −(m + nλ)fλ,m+n, where fλ,m(ϕ) = e for m ∈ µ + Z (15)

µ To add some extra rigor, the elements of Fν are required to have finite Sobolev of degree 1 P pP 2 d 2 d = 2 − Re λ. By definition, the norm of f = m amfλ,m is kfk = m(1 + m ) |am| . If λ < 0, then f is continuous; the larger the λ, the more singular can such functions be. While the norm remains finite under smooth diffeomorphisms, it is not preserved. There is, however, a 1 nondegenerate Diffg +(S )-invariant Z 2π ˜ dϕ −µ µ (g, f) = g˜(ϕ) f(ϕ) , where g ∈ F1−λ, f ∈ Fλ . (16) 0 2π

1 ∗ If µ is real and λ = 2 + is for s ∈ R, then the formula hg|fi = (g , f) defines a Hermitian inner µ product on Fλ . µ 1 We now regard Fλ as a representation of Ge ⊆ Diffg +(S ), i.e. specialize Eq. (15) to n = −1, 0, 1. As expected, the central element e2πiL0 and the Casimir operator Q are represented by e−2πiµ and λ(1 − λ), respectively. Thus, the current use of µ and λ is consistent with that for unitary irreps. µ 1/2 The representation Fλ of the group Ge can be reducible. For example F1/2 splits into the − + subspaces F1/2 and F1/2 that are spanned by the basis vectors f1/2,m with negative and positive µ m, respectively. In some other cases, the representation Fλ has an but does 1 not split. For example, let µ = λ 6= 2 . Then L1fλ,λ−1 = afλ,λ, where a = 1 − 2λ 6= 0. However, L−1fλ,λ = 0. This situation is indicated by the arrow in the diagram

where the small represent the basis vectors fλ,m labeled by m. A between m − 1 and m without an arrow means that both L1fλ,m−1 and L−1fλ,m are nonzero. The vectors fλ,λ, fλ,λ+1,... + (represented by the full circles) form a Ge-invariant subspace Fλ , which has no invariant com- + ⊥ plement. Indeed, suppose that such a complement (Fλ ) exists. Any invariant subspace has + ⊥ a homogeneous basis, that is, the basis vectors are eigenvectors of L0. Thus, (Fλ ) should be

5 λ Fλ :

1/2 F1/2 : λ F1−λ :

3 λ = 1, 2 , 2,... µ Figure 1: The structure of representations Fλ for some λ and µ. The full circles indicate those basis vectors fλ,m that span an invariant subspace, whereas the empty circles correspond to .

equal to the linear span of fλ,λ−1, fλ,λ−2,... (the empty circles). But this subspace is not, actually, invariant because L1fλ,λ−1 does not belong to it. By definition, the “non-unitary discrete series” representations are the invariant subspaces + λ − −λ Fλ ⊆ Fλ and Fλ ⊆ Fλ that are spanned by basis vectors fλ,m with m = λ, λ + 1,... in the first + − case and m = −λ, −λ − 1,... in the second case. Elements of Fλ and Fλ may be written as f = f˚(z)(−i dz)λ, f˚(eiϕ) = e−iλϕf˜(ϕ) or f = f˚(z)(i dz−1)λ, f˚(eiϕ) = eiλϕf˜(ϕ), (17)

respectively, where f˚(z) is holomorphic in the disk |z| < 1 or its complementary domain on the Riemann sphere, |z−1| < 1. In this notation, the transformation rule (13) becomes

 dz −λ  dz−1 −λ (V f˚)(z) = f˚(w) if f ∈ F +, (V f˚)(z) = f˚(w) if f ∈ F − (18) dw λ dw−1 λ where z = V (w) and V ∈ Ge.

3.2 Isomorphisms and intertwiners

µ ± The structure of representations Fλ , Fλ , and their relations can be described as follows (see also Figure1):

µ µ 1. If λ, −λ∈ / µ + Z, then Fλ is irreducible and isomorphic to F1−λ. 2. Let λ ∈ µ + Z or −λ ∈ µ + Z. (Without loss of generality, we may assume that µ = ±λ.)  1 3 ± µ If λ∈ / 0, − 2 , −1, − 2 ,... , then the representation Fλ ⊆ Fλ is irreducible and isomorphic µ ∓ to F1−λ/F1−λ. µ ± 3. If the unitary representation Cλ(1−λ) or Dλ exists for given λ and µ, it is isomorphic (except µ ∼ µ ± ∼ ±λ ∓ for the inner product) to Fλ = F1−λ or Fλ = F1−λ/F1−λ, respectively. The statements about irreducibility follow directly from Eq. (15). The isomorphisms in 1 and 2 can be obtained in a unified fashion using the intertwiners

µ± µ µ µ± ± Ξλ : F1−λ → Fλ , Ξλ f1−λ,m = bλ,mfλ,m for m ∈ µ + Z, (19) where Γ(λ + m) Γ(λ − m) b+ = if λ + m∈ / , b− = if λ − m∈ / ; (20) λ,m Γ(1 − λ + m) Z λ,m Γ(1 − λ − m) Z

6 Γ(2λ + k) + −  for k = 0, 1, 2,...  1 bλ,λ+k = bλ,−(λ+k) = k! if λ∈ / 0, − 2 , −1,... . (21) 0 for k = −1, −2....

µ,± The above definitions agree when they overlap. One can check that Ξλ is an intertwiner and that sin π(λ + µ) Ξµ,− = Ξµ,+, Ξµ,± = Ξµ,±−1 if ±λ∈ / µ + ; (22) λ sin π(λ − µ) λ 1−λ λ Z

±λ,± ∓ ±λ,± ± (Ξλ ) = F1−λ, (Ξλ ) = Fλ . (23) µ,± Thus, Ξλ defines the required (either between the full spaces or between the sub- µ,± −µ,∓ space and the quotient). Furthermore, Ξλ and Ξλ are conjugate to each other: µ,±  −µ,∓  g, Ξλ f = Ξλ g, f . (24) µ± µ+ The maps Ξλ are bounded with respect to the Sobolev norm. In fact, Ξλ f1−λ,m ≈ f1−λ,m µ− for large positive m and Ξλ f1−λ,m ≈ f1−λ,m for large negative m. µ µ ± µ Now, let Uλ be a nontrivial unitary irrep, i.e. either Cλ(1−λ) or Dλ . Its isomorphism with Fλ ± µ± or Fλ is obtained by factoring Ξλ as follows:

Ξµ± Ξµ± µ λ↓ µ λ↑ µ µ± ± µ± ± F1−λ −−→Uλ −−−→Fλ , Ξλ↓ f1−λ,m = cλ,m|mi, Ξλ↑ |mi = cλ,mfλ,m (25)

± such that the first map is onto and the second is injective. The coefficients cλ,m are given by these formulas: s s Γ(λ + m) Γ(λ − m) c+ = , c− = (−1)m−µ . (26) λ,m Γ(1 − λ + m) λ,m Γ(1 − λ − m) For the continuous series, they are both well-defined and their ratio is a function of λ and µ, + − p + − + namely cλ,m/cλ,m = sin π(λ − µ)/ sin π(λ + µ). For the discrete series Dλ (or Dλ ), only cλ,m − (resp. cλ,m) exists. The signs of the roots in Eq. (26) require some care. Let us fix them on a case-by-case ± ± basis. For the continuous series, we set c1/2,0 = 1 and analytically continue the functions cλ,m to 1 λ = 2 + is while keeping m equal to 0. If s 6= 0, we further continue to all real m. If s = 0 but 1 ± ± |µ| < 2 , then we can take the limit c1/2,m = lims→0 c1/2+is,m to obtain the following expressions: ( + − 1 if k > 0, c = c = γm−µ, where γk = (27) 1/2,m 1/2,m (−1)k if k < 0. For the complementary series, the correct signs are these: s s Γ(λ + m) Γ(λ − m) (using positive c+ = γ , c− = γ (28) λ,m m−µ Γ(1 − λ + m) λ,m m−µ Γ(1 − λ − m) square roots).

+ − For the representation Dλ or Dλ , one of the above equations is applicable. More explicitly, r r Γ(2λ + k) Γ(2λ + k) c+ = or c− = (−1)k (k = 0, 1, 2,...). (29) λ, λ+k k! λ, −(λ+k) k! ± For other values of m, the numbers cλ,m vanish.

7 4 Fourier transform on Ge and the Plancherel measure We first remind the reader of some algebraic terminology. An action of a group G on a set X is a function A : G × X → X such that A(1, x) = x and A(gh , x) = Ag, A(h, x) for all g, h, and x. Instead of A(g, x), it is more common to write A(g) x or A(g)·x. This construction is also called a “left ”. A “right” action is a similar function B that satisfies the equation B(hg, x) = B(g, B(h, x)). Any right action B can be turned into the left action A(g, x) = B(g−1, x). Thus, the concept of a right action is redundant when working with groups (as opposed to ). Invertible maps V : X → X act on functions as follows: (V f)(x) = f(V −1(x)). (30) Any group G acts on itself in two different ways: L(h) · g = hg, R(h) · g = gh−1. (31) We call the (left) actions L, R the L-action and R-action, respectively. They can also be applied to functions on G. If f is such a function, then L(h) · f(g) = f(h−1g), R(h) · f(g) = f(gh). (32)

4.1 Fourier transform on a group Let G be a group. We denote its unitary irreps by α, β, etc., the corresponding Hilbert spaces by Uα, and the group actions on them by Uα. Given some orthonormal basis of Uα, the matrix j j element functions Uα,k and Uα,k are defined as follows:

j j j −1 ∗ Uα,k(g) = j Uα(g) k , Uα,k(g) = Uα,k(g ) = k Uα(g) j (33)

The latter are slightly more convenient as a basis of the space of functions on G. Let us first write some elementary things, namely, the composition law and the group action on ket- and bra-vectors in the matrix notation: j X j n Uα,k(gh) = Uα,n(g) Uα,k(h), (34) n

X n −1 X k Uα(h) |ji = Uα,j(h) |ni, hk| Uα(h ) = Uα,n(h) hn|. (35) n n (In the case of bra-vectors, we use h−1 to conform to the definition of the group action.) Now, one k can easily check that the functions Uα,j are transformed under the L- and R-actions as |jihk| ∈ ∗ Uα ⊗ Uα: k X n k k X k n L(h) · Uα,j = Uα,j(h) Uα,n, R(h) · Uα,j = Uα,n(h) Uα,j. (36) n n k If G is compact, the matrix element functions Uα,j satisfy the Schur orthogonality relation, where the inner product is defined by the (arbitrarily normalized) . Let dα be the of α; then the orthogonality relation is: Z k l −1 R −1 Uα,j Uβ,m = j Uα(g) k l Uβ(g ) m dg = dα/ G dg δαβ δkl δjm. (37) G

8 k Furthermore, the functions U form a basis of the H of functions on G. The R α,j number dα/ G dg is called the Plancherel measure of the irrep α. It appears as a coefficient in the decomposition of identity into the projectors onto the basis functions. For example, if G is finite, the becomes a sum, and the Plancherel measure is dα/|G|. The decomposition into matrix element functions is known as the Fourier transform on G. The generalization of these results to non-compact Lie groups, such as Ge, is not straightforward. To understand the problems that can and do arise, let us prove the orthogonality and completeness α,jl for finite groups. To show the orthogonality, let Mβ,km denote the middle expression in (37). If α, β, k, l are fixed, then the expression Z X α,jl −1 Mβ,km|jihm| = Uα(g) k l Uβ(g ) dg (38) j,m G

defines a from Uβ to Uα, which is easily seen to be an intertwiner. By Schur’s lemma, it α,jl is proportional to the identity map if α = β and vanishes otherwise. Hence, Mβ,km is proportional α j α l α,jl α j l to δβ δm and, by a similar , to δβ δk. These two statements imply that Mβ,km ∝ δβ δmδk. The proportionality coefficient can be found by contracting the indices k and l. To show the completeness, we consider the Hilbert space H as the left regular representation of G (that is, the representation given by the L-action) and decompose it into unitary irreps. Any of these sub-representations has a basis of functions fj that transform as the basis vectors |ji of some standard representation α. Hence, −1  X n fj(g) = L(g ) · fj (1) = Uα,j(g) fn(1), (39) n n which means that fj is the linear combination of the functions Uα,j with the coefficients fn(1). Therefore, any f ∈ H is a combination of matrix element functions. What changes if G is a general locally compact Lie group? The use of Schur’s lemma is valid if the Haar measures are the same (which is true for G = Ge). However the integral in iαg Eq. (37) may diverge. For example, if G = R, then Uα(g) = e (for α, g ∈ R). In this case, the functions Uα are δ-normalizable, namely hUα|Uβi = 2πδ(α − β). This gives the decomposition of −1 R −1 identity 1 = (2π) |UαihUα| dα; thus, the Plancherel measure on the set of irreps is (2π) dα. The trivial representation of Ge is a different case. The corresponding function U0 is identically equal to 1, and its norm is clearly infinite. One might hope to remedy the situation by regarding 0 the trivial representation as a limiting case of the complementary series Cq for q → 0 or the discrete ± series Dλ for λ → 0. However, it turns out that all complementary series representations and the 1 discrete series representations with λ < 2 have matrix elements that are not even δ-normalizable. Therefore, these representations do not appear in the Fourier transform. The main assumption in the completeness proof was the existence of an irreducible decompo- sition of the left regular representation. It is true that any unitary representation of Ge splits into irreducible pieces, which are isomorphic to the standard unitary irreps. However, this fact should not be taken for granted. For some other groups (e.g. SL(2, Z)), a general unitary representation does not split into irreducible representations (because the process of splitting into progressively smaller pieces may not converge). A more general decomposition into isotypical components ex- ists, but it involves type II and type III von Neumann factors. We will not prove or use the existence of an irreducible decomposition for an arbitrary unitary representation of Ge but show the completeness of the matrix element functions directly.

9 4.2 The L- and R-actions on Ge in explicit coordinates It is convenient to parametrize Ge by three variables similar to the Euler angles:

ϕΛ ξΛ −ϑΛ g(ξ, ϕ, ϑ) = e 0 e 1 e 0 , ξ > 0 (40) Note that g(ξ, ϕ+2π, ϑ+2π) = g(ξ, ϕ, ϑ). For a nonsingular, one-to-one parametrization, one can use z = eiϕ tanh(ξ/2) (the image of 0 under the action of g(ξ, ϕ, ϑ) on the unit disk) and ϑ − ϕ. Let us find the L- and R-actions in the infinitesimal form, i.e. calculate (1+δh)g and g(1−δh), where δh is a Lie algebra element and g ∈ Ge. When δh is fixed, the expressions (δh)g and g(δh) define some vector fields on Ge. In general, (δh)g = XL(g)(δh), g(δh) = XR(g)(δh), (41) where XL(g) and XR(g) are linear maps from the Lie algebra to the of Ge at point g. They are represented by matrices if we write vector fields in components and express δh as j (δh )Λj. Thus, the L- and R-actions of Λj on functions are given by these formulas: ∂ ∂ ΛL = −XL(g)α , ΛR = XR(g)α , (42) j j ∂gα j j ∂gα where j = 0, 1, 2 and the index α refers to ξ, ϕ, or ϑ. It is easier to calculate the inverse matrices, Y L(g) = (XL(g))−1 and Y R(g) = (XR(g))−1, which represent two variants of the Maurer-Cartan form: −1  L j α −1  R j α (dg)g = Λj Y (g) α dg , g (dg) = Λj Y (g) α dg . (43) To obtain, for example, the first column of Y L = Y L(g) for g = g(ξ, ϕ, ϑ), we consider

∂g −1 g−1 = eϕΛ0 Λ eξΛ1 e−ϑΛ0 eϕΛ0 eξΛ1 e−ϑΛ0  = eϕΛ0 Λ e−ϕΛ0 = (cos ϕ)Λ + (sin ϕ)Λ , ∂ξ 1 1 1 2 L T and collect the coefficients in front of Λj. Thus, the first column of Y is (0, cos ϕ, sin ϕ) . Continuing in this manner, we find that  0 1 − cosh ξ   0 cosh ξ −1 L R Y = cos ϕ 0 − sin ϕ sinh ξ ,Y = cos ϑ − sin ϑ sinh ξ 0  , (44) sin ϕ 0 cos ϕ sinh ξ sin ϑ cos ϑ sinh ξ 0

0 cos ϕ sin ϕ   0 cos ϑ sin ϑ   cosh ξ cosh ξ   1 1  L 1 − sin ϕ cos ϕ  R  0 − sin ϑ cos ϑ  X =  sinh ξ sinh ξ  ,X =  sinh ξ sinh ξ  . (45)      1 1   cosh ξ cosh ξ  0 − sin ϕ cos ϕ −1 − sin ϑ cos ϑ sinh ξ sinh ξ sinh ξ sinh ξ We can now plug the expressions for XL, XR into Eq. (42) to obtain the L- and R-actions of the operators Λj on functions. It is convenient to have the result for L0 = −iΛ0 and L±1 = ∓Λ1 −iΛ2:  cosh ξ 1  LL = i∂ ,LL = e±iϕ ±∂ + (i∂ ) + (i∂ ) (46) 0 ϕ ±1 ξ sinh ξ ϕ sinh ξ ϑ  1 cosh ξ  LR = i∂ ,LR = e±iϑ ∓∂ − (i∂ ) − (i∂ ) (47) 0 ϑ ±1 ξ sinh ξ ϕ sinh ξ ϑ 10 4.3 Casimir eigenfunctions

The space of square-integrable functions on Ge can be decomposed into common eigenfunctions of L R three commuting Hermitian operators, L0 , L0 , and Q, where 1 1 Q = −(LL)2 + LL LL + LLLL  = −(LR)2 + LR LR + LRLR . (48) 0 2 −1 1 1 −1 0 2 −1 1 1 −1 We now find the common eigenfunctions without asking if they are normalizable. That question will be addressed later. L R Let us first impose the conditions L0 Ψ = −lΨ and L0 Ψ = −rΨ, where l and r are arbitrary complex numbers.2 Thus, ξ ΨeϕΛ0 eξΛ1 e−ϑΛ0  = ei(lϕ+rϑ)f(u), where u = tanh2 . (49) 2 The use of the variable u instead of ξ will help to simplify some subsequent equations. In what follows, Ψ is treated as a function of (ξ, ϕ, ϑ) rather than g = eϕΛ0 eξΛ1 e−ϑΛ0 . We will later require that Ψ(g) only depend on g and be regular at g = 1. L R The action of Ln, Ln on f depends on the parameters l and r, l + r l − r LL(l, r) = −l, LL (l, r) = ±(1 − u)u1/2∂ − u−1/2 − u1/2, (50) 0 ±1 u 2 2 R n L Ln (l, r) = (−1) Ln(r, l), (51) and the parameters also change:

L R Ln :(l, r) 7→ (l + n, r),Ln :(l, r) 7→ (l, r + n). (52) It is now easy to write the Casimir operator explicitly: 1 − u  Q = −(1 − u)2 u∂2 + ∂  + (l + r)2 − (l − r)2u . (53) u u 4u The eigenvalue equation, Qf = λ(1−λ)f is equivalent to the hypergeometric differential equation (u∂u + c)∂uh = (u∂u + a)(u∂u + b)h for a closely related function h. Indeed, both differential equations have regular singular points at u = 0, 1, ∞ and no other singularities. To find the exact relation, it is sufficient to compare the exponents that define the asymptotics of the fundamental solutions, namely, f(u) ∼ uαv for u → v (v = 0, 1) and f(u) ∼ u−α∞ for u → ∞. The Casimir eigenvalue equation and the hypergeometric equation have the following exponents:   Equation for f α0 = ±(l + r)/2, Equation for h α0 = 0, 1 − c, with parameters α1 = λ, 1 − λ, with parameters α1 = 0, c − a − b, (54) λ, l, r : α∞ = ±(l − r)/2; a, b, c : α∞ = a, b. Since each exponent has two different values, there are several ways to match them. For example,

f(u) = u(l+r)/2(1 − u)λh(u), a = λ + l, b = λ + r, c = 1 + l + r. (55)

2 R The parameter ν = −r may be called “spin” because a function Ψ satisfying the condition L0 Ψ = νΨ has the as a ν- on the hyperbolic plane, see section 5.2.

11 General solutions: These are two solutions of the equation Qf = λ(1 − λ)f on the interval 0 < u < 1: (l+r)/2 λ  Aλ,l,r(u) = u (1 − u) F λ + l, λ + r, 1 + l + r; u (56)

(l+r)/2 λ  Bλ,l,r(u) = u (1 − u) F λ + l, λ + r, 2λ; 1 − u (57)

−1 where F(a, b, c; x) = Γ(c) 2F 1(a, b, c; x) is the scaled hypergeometric function. (It is well-defined for all values of a, b, c but vanishes if a, c − a ∈ {0, −1, −2,...} or b, c − b ∈ {0, −1, −2,...}.) Let us mention some useful identities:

A1−λ,l,r = Aλ,l,r,Bλ,l,r = Bλ,−l,−r; (58) sin(2πλ) B B A = λ,l,r − 1−λ,l,r . (59) π λ,l,r Γ(1 − λ + l) Γ(1 − λ + r) Γ(λ + l) Γ(λ + r)

For a more complete picture, Aλ,l,r(u), Aλ,−l,−r(u) make a pair of fundamental solutions near u = 0, −1 the functions Bλ,l,r(u), B1−λ,l,r(u) are the fundamental solutions near u = 1, and Aλ,l,−r(u ), −1 Aλ,−l,r(u ) near u = ∞. (The first four functions are defined for u ∈ (0, 1) and the last two for L u ∈ (1, ∞).) The operators L±1(l, r) act on Aλ,l,r and Bλ,l,r as follows:

L L L−1(l, r) Aλ,l,r = −Aλ,l−1,r ,L1 (l, r) Aλ,l,r = −(l + λ)(l + 1 − λ)Aλ,l+1,r ; (60) L L L−1(l, r) Bλ,l,r = −(l − λ)Bλ,l,r ,L1 (l, r) Bλ,l,r = −(l + λ)Bλ,l,r . (61)

L L To find the action on the other fundamental solutions, we note that Ln(l, r) = −L−n(−l, −r) and L −1 L that Ln(l, r) acts on functions of u as Ln(l, −r) on functions of u. The R-action is obtained from Eq. (51). i(ϕ1−ϕ2) In some applications (e.g. spinors on AdSg 2, where u = e ), u lies on the unit circle. One can analytically continue functions from the interval (0, 1) to the simply connected D = C − [0, ∞) (containing the unit circle without the point 1) through the upper half-plane or through the lower half-plane. The first option is preferred when the circle is parametrized as u = eiθ with 0 < θ < 2π. However, let us give both definitions and some related identities:

± ) ( ) Aλ,l,r Aλ,l,r ± = analytic cont. of through the upper (+) or lower (−) half-plane; (62) Bλ,l,r Bλ,l,r

l+r l+r −iπ 2 + iπ 2 − e Aλ,l,r(u) = e Aλ,l,r(u), (63)

π π π π i 2 λ + i 2 λ + −i 2 λ − −1 −i 2 λ − −1 e Bλ,l,r(u) = e Bλ,−l,−r(u) = e Bλ,l,−r(u ) = e Bλ,−l,r(u ). (64) We now write all 6 fundamental solutions that are continued from their original definition domains to D through the half-plane Im u > 0:

+ + − −1 Aλ,l,r(u),Bλ,l,r(u),Aλ,l,−r(u ), (65) + + − −1 Aλ,−l,−r(u),B1−λ,l,r(u),Aλ,−l,r(u ). They span the two-dimensional solution space.

12 Nonsingular solutions: Having studied the Casimir eigenfunctions of the form ei(lϕ+rϑ)f(u) in full generality, we select those that depend only on g = g(ξ, ϕ, ϑ) ∈ Ge and are regular at g = 1. The first condition means the invariance under (ξ, ϕ, ϑ) 7→ (ξ, ϕ + 2π, ϑ + 2π); hence, l + r is an integer. To check if a function is regular, we examine its u → 0 asymptotics using the variables iϕ 1/2 −iϕ 1/2 z = e u ,z ¯ = e u , and ϑ − ϕ. The space of regular solutions is spanned by Aλ,l,r and Aλ,−l,−r. These functions are linearly dependent because

Γ(λ + l) Γ(λ + r) Aλ,l,r = Γ(λ − l) Γ(λ − r) Aλ,−l,−r if l + r ∈ Z. (66)

One of them may vanish, but Aλ,l,r 6= 0 if l + r > 0 and Aλ,−l,−r 6= 0 if l + r 6 0.

Normalizability: Finally, we select the nonsingular solutions that are normalizable or δ-nor- malizable. The inner product is given by the Haar measure on Ge. When group elements are rep- iϕΛ0 iξΛ1 −iϑΛ0 i(lαϕ+rαϑ) 2  resented as e e e , the measure is (sinh ξ) dξ dϕ dϑ. If Ψα(g) = e fα tanh (ξ/2) with lα + rα ∈ Z and lα, rα ∈ R (where α = 1, 2), then Z 1 2 ∗ 2 du hΨ1|Ψ2i = 4π δl1+r1, l2+r2 δ(r1 − r2) hf1|f2i, where hf1|f2i = f1(u) f2(u) 2 . (67) 0 (1 − u) 1 Since λ and 1 − λ define the same Casimir eigenspace, we may assume that Re λ > 2 , or 1 1 λ = 2 + is with s > 0, or λ = 2 . Let us consider one of the linearly dependent candidate solutions Aλ,l,r and Aλ,−l,−r: λ 1−λ Aλ,l,r(u) ≈ aλ,l,r(1 − u) + a1−λ,l,r(1 − u) for u → 1, (68) where Γ(1 − 2λ) a = . (69) λ,l,r Γ(1 − λ + l) Γ(1 − λ + r) 1 λ If Re λ > 2 , a function with the asymptotic behavior f(u) ∼ (1 − u) for u → 1 is normalizable, 1−λ 1 and (1 − u) is not even δ-normalizable. In the marginal case of λ = 2 + is, s > 0, we have: 1/2+is ∗ 1/2−is 2 0 if fs(u) ≈ a(1 − u) + a (1 − u) for u → 1, then hfs|fs0 i = 4π|a| δ(s − s ). (70)

1 If λ = 2 , one has to take the in Eq. (68); the result is that the function is not normalizable. It is now easy to find all cases where both Aλ,l,r and Aλ,−l,−r are normalizable or δ-normalizable: 1   • λ > and λ + l, λ − r ∈ {0, −1, −2,...} or λ + r, λ − l ∈ {0, −1, −2,...} ; (71) 2 1 • λ = + is, s, l, r ∈ , s > 0, l + r ∈ . (72) 2 R Z

4.4 The matrix elements and Plancherel measure for the irreps of Ge µ µ ± Let Uλ be a nontrivial unitary irrep, i.e. Cλ(1−λ) or Dλ . The corresponding matrix element functions ν Uλ,m (for m, ν ∈ Z+µ) transform as |mihν| under the L- and R-actions of the group. In particular,

L ν ν R ν ν ν ν L0 Uλ,m = −mUλ,m,L0 Uλ,m = νUλ,m,Q Uλ,m = λ(1 − λ)Uλ,m, (73)

L ν p ν L±1 Uλ,m = − (m ± λ)(m ± (1 − λ)) Uλ,m±1. (74)

13 ν ϕΛ0 ξΛ1 −ϑΛ0  i(mϕ−νϑ) 2  The first set of equations implies that Uλ,m e e e = e f tanh (ξ/2) , where f is L proportional to the fundamental solution Aλ,m,−ν or Aλ,−m,ν. Comparing the action of L±1 on ν the functions Uλ,m with the corresponding action (60) on the fundamental solutions and using the ν ν identity Uλ,m(1) = δm for normalization, we find that

s ν   Γ(λ + m) Γ(1 − λ + m)  ξ  U eϕΛ0 eξΛ1 e−ϑΛ0 = ei(mϕ−νϑ)A tanh2 λ,m Γ(λ + ν) Γ(1 − λ + ν) λ,m,−ν 2 (75) s Γ(λ − m) Γ(1 − λ − m)  ξ  = (−1)ν−m ei(mϕ−νϑ)A tanh2 Γ(λ − ν) Γ(1 − λ − ν) λ,−m,ν 2

µ + µ − The first formula is applicable to the irreps Cλ(1−λ), Dλ , and the second to Cλ(1−λ), Dλ . Note that + − the range of m and ν is restricted to {λ, λ + 1,...} for Dλ and to {−λ, −λ − 1,...} for Dλ . Thus, the normalizable and δ-normalizable Casimir eigenfunctions (see equations (71), (72)) are exactly ± 1 µ 1 the matrix element functions for Dλ with λ > 2 and for Cλ(1−λ) with λ = 2 + is. This shows the completeness of the matrix element functions. The analogue of orthogonality relation (37) for discrete series representations is as follows:

0 0 2 D ±(λ+k) ±(λ +k ) E 8π 0 U U 0 0 0 = δ(λ − λ ) δ 0 δ 0 λ, ±(λ+j) λ , ±(λ +j ) 2λ − 1 jj kk (76) 1 for λ, λ0 > , j, k, j0, k0 ∈ {0, 1, 2,...}. 2 The overall factor is found by setting j, k, j0, k0 to 0 so that the functions in questions are e±iλ(ϕ−ϑ)(1 − u)λ and e±iλ0(ϕ−ϑ)(1 − u)λ0 , where u = tanh2(ξ/2). The inner product between these functions is obtained using Eq. (67). For the principal series, the orthogonality relation is:

0 0 D µ+k µ +k E 2 cosh(2πs) + cos(2πµ) 0 0 U U 0 0 0 = 4π δ(s − s ) δ(µ − µ ) δ 0 δ 0 1/2+is, µ+j 1/2+is , µ +j s sinh(2πs) jj kk (77) 1 1 for s, s0 > 0, − < µ, µ0 , j, k, j0, k0 ∈ . 2 6 2 Z (The range of s and µ has been restricted to avoid redundancy.) To derive equation (77), we again 0 0 µ consider the case j = k = j = k = 0. The function U1/2+is, µ is proportional to A1/2+is, µ,−µ, which has the asymptotics (68) with

2 2 cosh(2πs) + cos(2πµ) a1/2+is, µ,−µ = a1/2−is, µ,−µ = . (78) 4πs sinh(2πs)

The Plancherel measure is found by inverting the coefficients in the orthogonality relations:

 (2π)−2λ − 1  dλ, λ > 1 for D± Plancherel measure on  2 2 λ = (79) the irreps of G −2 s sinh(2πs) µ e (2π) ds dµ, s > 0 for C 2  cosh(2πs) + cos(2πµ) 1/4+s

14 ∼ These formulas can be specialized to the irreps of G = PSL(2, R), which are characterized by ± µ = 0. In particular, the discrete series representations are Dn with n = 1, 2,... Thus,

( −2 1  ± Plancherel measure on (2π) n − for Dn = 2 (80) the irreps of PSL(2, ) −2 0 R (2π) tanh(πs) s ds for C1/4+s2

5 Spinors on the hyperbolic plane and anti-de Sitter space

2 5.1 The spaces H , AdS2, and their complex embeddings 2 The hyperbolic plane H is the quotient of G by the subgroup K generated by Λ0; it is also equal to the quotient of the corresponding universal covers, Ge/Ke. Conversely, Ge is the total space of a principal Ke- over H2. Among the three “Euler angle” coordinates, ξ and ϕ parametrize the base and ϑ the fiber. The on H2 can be obtained from the L- and R-invariant metric on Ge, which is in turn determined by the η = diag(−1, 1, 1). Using the notation of Section 4.2, 2 L j L k α β R j R k α β d` = ηjk(Y )α(Y )β dg dg = ηjk(Y )α(Y )β dg dg (81) = dξ2 − dϕ2 − dϑ2 + 2 cosh ξ dϕ dϑ. The distance between infinitesimally close fibers is found by taking the extremum of d` over dϑ with dξ and dϕ fixed. The result is: 4 dz dz¯ ξ d`2 = dξ2 + (sinh ξ)2dϕ2 = , where z = eiϕ tanh . (82) (1 − zz¯)2 2 Thus, we have recovered the well-known Poincar´edisk model on the hyperbolic plane. The space H2 = Ge/Ke inherits the L-action of Ge, while the R-action has been used up in the quotient construction. The anti-de Sitter space AdS2 is the quotient of G by the subgroup generated by Λ2. Recall az+b that a general element of G is a linear fractional map g : z 7→ cz+d preserving the unit disk. The subgroup generated by Λ2 consists of those g’s that preserve i and −i. Thus, AdS2 is the of (i, −i) under the simultaneous action of G on pairs of points. This orbit, actually, includes iϕ1 iϕ2 all pairs of distinct points on the unit circle, z1 = e and z2 = e . The standard G → AdS2 takes g to (z1, z2) = (g(i), g(−i)). To describe the metric on AdS2 and its universal cover AdSg 2, we consider ϕ1, ϕ2 as real numbers subject to the constraint 0 < ϕ1 − ϕ2 < 2π. Then

−t − dϕ dϕ e tanh(ξ/2) = tan(π/4 − ϕ1/2), d`2 = dξ2 − (sinh ξ)2dt2 = 1 2 , (83) 2 ϕ1−ϕ2  t sin 2 e tanh(ξ/2) = tan(π/4 + ϕ2/2).

π π (The first expression gives the metric in the region ϕ1 < 2 , ϕ2 > − 2 , which we call the “Schwarzschild patch”, see Figure2.) Another, more standard way to write the AdSg 2 metric 2 −2 2 2 is d` = (cos θ) (−dϕ + dθ ), where ϕ = (ϕ1 + ϕ2)/2 and θ = (π − ϕ1 + ϕ2)/2. It is often useful to analytically continue functions between the hyperbolic plane and the anti-de Sitter space. From the physical perspective, there are two ways to define the analytic continuation:

15 −−−−−−→

Figure 2: Schwarzschild patch of the anti-de Sitter space. one is natural for the study of the SYK model in imaginary time (by interpreting a pair of points on the time circle as a point of AdS2 [7]) and the other corresponds to the transition to real time. We will obtain those continuations using a standard of H2 and two different embeddings of AdS2 in some complex M. The latter consists of pairs of distinct points on the C = C ∪ {∞} and comes with a complex metric:

 2 −4 dz1 dz2 M = (z1, z2): z1, z2 ∈ C, z1 6= z2 , d` = 2 . (84) (z1 − z2)

This metric is invariant under the simultaneous action of PSL(2, C) on z1 and z2. One can also describe M as the quotient of PSL(2, C) by the stabilizer of the point (0, ∞), that is, the complex subgroup generated by Λ0. We now construct the three embeddings together with some related structure. Each embedding is extended to a “map of principal bundles”, which consists of compatible maps between their bases, total spaces, and structure groups. Let us begin with the map from the 2 Ge → H to PSL(2, C) → M and denote its constituent parts by ζ, J, and ωH. The next equation includes the condition for compatibility between ζ and J (expressed as a ) as well as the definitions of these maps; Z denotes the subgroup of Ge generated by e2πΛ0 .

J Ge −−−→ PSL(2, C)   J(g) = g (reduced Z),   (85) y y ζ(z) = z, z¯−1. ζ H2 −−−→ M

θΛ0 The map ωH (from the group of elements h = e , θ ∈ R to such elements with complex θ) should be a group and satisfy the compatibility condition J ◦ R(h) = R(ωH(h)) ◦ J, i.e. −1 −1 J(gh ) = J(g) ωH(h ) for all h. Since J is, essentially, trivial, such is ωH, namely, ωH(h) = h (modulo Z). On the other hand, both principal bundle maps from Ge → AdSg 2 to PSL(2, C) → M involve this homomorphism of structure groups:

θΛ2  iθΛ0 −1 ωAdS e = e for all θ ∈ R, i.e. ωAdS(h) = W hW, (86) where 1 1 i z + i W = ei(π/2)Λ1 = √ ,W (z) = . (87) 2 i 1 iz + 1

16 The other parts are defined below. Although ζ˜ and ˚ζ are not injective, they factor as the projection onto AdS2 followed by an embedding.

˜ G J e −−−→ PSL(2, C) ˜   J(g) = gW,   (88) y y ˜ iϕ1 iϕ2  ζ(ϕ1, ϕ2) = e , e ; ζ˜ AdSg 2 −−−→ M

J˚ ˚ −1 Ge −−−→ PSL(2, C) J(g) = W gW,     ˚ −1 iϕ1 −1 iϕ2  (89) y y ζ(ϕ1, ϕ2) = W (e ),W (e ) ˚  ζ = tan(π/4 − ϕ1/2), tan(π/4 − ϕ2/2) . AdSg 2 −−−→ M 2 ˜ When the map ζ :H → M is used together with ζ : AdSg 2 → M, the coordinates z,z ¯ on the iϕ1 −1 −iϕ2 hyperbolic plane correspond to the functions z1 = e and z2 = e on the anti-de Sitter space by the analytic continuation through M. If, on the other hand, AdSg 2 is mapped to M using ˚ 2 ζ, then the coordinate ξ is consistent between H and the Schwarzschild patch of AdSg 2, and ϕ analytically continues to it.

5.2 Definitions of spinors and two standard gauges Spinors on H2, or any Riemannian , are associated with representations of the universal cover of SO(2), that is, the group Ke generated by Λ0. Let us consider the one-dimensional repre- sentation such that Λ0 acts as the by −iν. Sections of the vector bundle associated with this representation and some principal Ke-bundle are called “ν-spinors”. More explicitly, a ν-spinor is a function Ψ from the total space of the principal bundle to the representation space (or simply the complex numbers) such that

R (Λ0 − iν)Ψ = 0. (90) Here the superscript “R” refers to the action of the structure group on the total space. In the hyperbolic plane case, the principal Ke-bundle is given by the quotient map Ge → H2, and the action in question is the R-action considered previously. For calculational purposes, it is convenient to represent spinors by functions on the base space. 2 This requires fixing a gauge, i.e. a cross section of the principal Ke-bundle. Let sH :H → Ge 2 be such a cross section, and let ψ(x) = Ψ(sH(x)) Any point of the fiber over x ∈ H can be −θΛ0 represented as sH(x) e ; hence,

−θΛ0  −iνθ Ψ sH(x) e = e ψ(x). (91)

−τ(x)Λ0 −iντ(x) If sH(x) is replaced with sH(x) e , then ψ(x) changes to e ψ(x). Spinors on AdSg 2 are defined by the condition that Λ2 acts as the multiplication by ν. This definition is motivated by the relation between the structure group maps ωAdS and ωH; indeed, the Λ2 in the anti-de Sitter case corresponds to iΛ0 in the hyperbolic plane case. Thus,

−θΛ2  νθ Ψ sAdS(x) e = e ψ(x). (92)

17 2 Now, let us define cross sectionss ˜H,˚sH :H → Ge whose analytic continuations to M are consistent with the respective embeddings of AdS2. We will give explicit formulas as well as some pictures. Cross sections of the principal bundle Ge → H2 can be visualized as local frames on the unit disk. Indeed, consider the vector fieldsv ˜0, v˜1, v˜2 on Ge that correspond to the R-action of ϕΛ0 ξΛ1 −ϑΛ0 R Λ0, Λ1, Λ2. At each point g = e e e , they are given by the columns of the matrix X (g), see Eq. (45). The first of them lies in the fiber and the other two are orthogonal to it. Projecting v˜1 andv ˜2 on the base, we get these vectors v1, v2:

ξ !  cos ϑ  ξ !  sin ϑ  v1 v2 =  sin ϑ  , =  cos ϑ  . (93) vϕ − vϕ 1 sinh ξ 2 sinh ξ 2 Setting g = sH(x) gives an orthonormal frame at x ∈ H . The two particular cross sections are:

ϕΛ0 ξΛ1 ϕΛ0 ξΛ1 −ϕΛ0 s˜H(z) = e e ˚sH(z) = e e e (94)

where z = eiϕ tanh(ξ/2). The first one is multivalued because (ξ, ϕ) and (ξ, ϕ + 2π) correspond to the same z but eϕΛ0 eξΛ1 6= e(ϕ+2π)Λ0 eξΛ1 . As previously alluded to,s ˜H is related to some cross sections ˜AdS of the principal bundle Ge → AdSg 2. The correspondence betweens ˜H,s ˜AdS, and their common analytic continuations ˜ can be expressed by a commutative diagram and then translated to explicit equations:

J J˜ Ge −−−→ PSL(2, C) ←−−− Ge −1 x x x s˜(z, z¯ ) =s ˜H(z), s˜H s˜ s˜AdS (95)    iϕ1 iϕ2 s˜(e , e ) =s ˜AdS(ϕ1, ϕ2) W. ˜ 2 ζ ζ H −−−→ M ←−−− AdSg 2

(Althoughs ˜H is multivalued ands ˜ double-valued,s ˜AdS is well-defined.) The solution to the above equations is:

ϕ1 + ϕ2 ϕ1 − ϕ2 s˜ (ϕ , ϕ ) = eϕΛ0 eγΛ1 , where ϕ = , γ = − ln tan . (96) AdS 1 2 2 4 It is important that all the functions involved (namely, ϕ and γ) are real; that would not be the case if we began with ˚sH. The cross section ˚sH is consistent with the other embedding of AdS2. Specifically,

J J˚ Ge −−−→ PSL(2, C) ←−−− Ge −1 x x x ˚s(z, z¯ ) = ˚sH(z), ˚sH ˚s ˚sAdS (97)    −1 iϕ1 −1 iϕ2  −1 ˚s W (e ),W (e ) = W ˚sAdS(ϕ1, ϕ2) W. ˚ 2 ζ ζ H −−−→ M ←−−− AdSg 2 Solving these equations requires slightly more work. The result is this: sin(π/4 + ϕ /2) −τΛ2 1 ˚sAdS(ϕ1, ϕ2) =s ˜AdS(ϕ1, ϕ2) e , where τ = ln . (98) sin(π/4 − ϕ2/2)

18 Spinors written in the “tilde gauge” (i.e. usings ˜H ors ˜AdS) and in the “disk gauge” (using ˚sH or ˚sAdS) are related as follows:

 ν ˚H −ν/2 ˜H ˚AdS sin(π/4 + ϕ1/2) ˜AdS ψ (z) = z/z¯ ψ (z), ψ (ϕ1, ϕ2) = ψ (ϕ1, ϕ2) (99) sin(π/4 − ϕ2/2)

In either gauge, the analytic continuation of spinors is similar to that of ordinary functions. It still depends on the embedding of AdS2 in M, therefore the explicit expressions are different:

˜H ˜ −1 ˜AdS ˜ iϕ1 iϕ2  ψ (z) = ψ z, z¯ , ψ (ϕ1, ϕ2) = ψ e , e , (100) ˚H ˚ −1 ˚AdS ˚  ψ (z) = ψ z, z¯ , ψ (ϕ1, ϕ2) = ψ tan(π/4 − ϕ1/2), tan(π/4 − ϕ2/2) . (101)

5.3 Spinors on H2 Analytic spinors: Many applications involve ν-spinors that can be expressed by a convergent −1 Taylor series in z1 = z and z2 =z ¯ for |z|, |z¯| < 1. Such spinors transform under maps V ∈ Ge as follows:

 −ν/2 −1 ν/2 ˚ dz1 dz2 ˚ (V ψ)(z1, z2) = −1 ψ(w1, w2) for z1 = V (w1), z2 = V (w2). (102) dw1 dw2

This is a special case of the transformation rule for holomorphic (λ1, λ2)-forms, i.e. functions of z , z that transform as elements of F + in the first variable and F − in the second variable, cf. 1 2 λ1 λ2 Eq. (18). Such forms are written symbolically as

˚ λ1 −1 λ2 f = f(z1, z2)(−i dz1) (i dz2 ) . (103)

−2λ λ −1 λ For example, (1 − z1/z2) (dz1) (dz2 ) is a (λ, λ)-form that is invariant under Ge (and more ν ν  general linear fractional maps in a suitable domain). Thus, ν-spinors are 2 , − 2 -forms with respect to maps V ∈ Ge. Conversely, any (λ1, λ2)-form f can be written as

˚ −(λ1+λ2) ˚ f(z1, z2) = 1 − z1/z2 ψ(z1, z2), (104) where ψ is a (λ1 − λ2)-spinor.

Casimir eigenfunctions: Among ν-spinors, let us consider the common eigenfunctions of the Casimir operator with the eigenvalue λ(1 − λ) and the operator L0 with the eigenvalue −m. Since spinors on H2 are a certain type of functions on Ge on which the group operates by the L-action, we can simply use the equations from Section 4.3 with l = m, r = −ν, u = zz¯ = z1/z2, and iϕ p √ i(lϕ+rϑ) i(lϕ+rϑ) e = z/z¯ = z1z2. In particular, the eigenfunctions e Aλ,l,r(u) and e Aλ,−l,−r(u) in the disk gauge (i.e. with ϑ = ϕ) become

˚ν,+ m−ν λ  ψλ,m(z1, z2) = z1 (1 − z1/z2) F λ + m, λ − ν, 1 + m − ν; z1/z2 (105) ˚ν,− m−ν λ  ψλ,m(z1, z2) = z2 (1 − z1/z2) F λ − m, λ + ν, 1 − m + ν; z1/z2

19 ν 6 −λ −λ < ν < λ ν > λ

3 Figure 3: The action of L−1, L1 on Casimir eigenfunctions for λ = 1, 2 , 2,... and ν ∈ λ + Z.A ν,+ ν,− circle with label m ∈ ν + Z represents the basis function ψλ,m if m > ν and ψλ,m if m 6 ν.

These functions are nonsingular if and only if m ∈ ν+Z, in which case they are linearly dependent. ν,+ ν,− Choosing one of them that is nonzero for each given m, e.g. ψλ,m for m > ν and ψλ,m for m 6 ν, we obtain a basis of some representation of Ge. The group action on the basis functions is characterized by the equations below and illustrated by Figure3.

ν,+ ν,+ ν,+ ν,+ L−1ψλ,m = −ψλ,m−1,L1ψλ,m = −(m + λ)(m + 1 − λ) ψλ,m+1, (106) ν,− ν,− ν,− ν,− L−1ψλ,m = (m − λ)(m − 1 + λ) ψλ,m−1,L1ψλ,m = ψλ,m+1.

ν Intertwiner from the space F1−λ to ν-spinors on the hyperbolic plane: A nontrivial ν 3 intertwiner Eλ of this type exists and is unique up to an overall factor. Its action on the basis ν vectors f1−λ,m ∈ F1−λ is given by the equation Γ(λ + m) Γ(λ − m) Eνf = ψ ν,+ = ψ ν,− for m ∈ ν + . (107) λ 1−λ,m Γ(λ + ν) λ,m Γ(λ − ν) λ,m Z

ν If we regard the ν-spinor Eλf1−λ,m as a function of g ∈ Ge, then

ν   ν −1  Eλf1−λ,m (g) = fλ,−ν,F1−λ(g )f1−λ,m . (108)

ν −1 −1 ν where F1−λ(g ) is the action of the group element g in the representation space F1−λ and the dϕ big parentheses denote the integral of the product of two functions with 2π . While this integral can be calculated directly, we note that the right-hand side of the above equation is a non-unitary version of the matrix element functions considered in Section 4.1. Such functions are transformed ν as the basis vectors f1−λ,m ∈ F1−λ, which is exactly the intertwiner property. ν ν Let us also write the ν-spinor Eλf for an arbitrary f ∈ F1−λ in the disk gauge:

Z 2π λ ν ◦ (1 − zz¯) −iνϕ ˜ dϕ Eλf (z) = iϕ λ−ν −iϕ λ+ν e f(ϕ) (109) 0 (1 − ze¯ ) (1 − ze ) 2π

This equation is proved by expanding the integrand in z andz ¯. Once again, there is an independent argument showing that the integral operator on the right-hand side defines an intertwiner. Indeed, its kernel function corresponds to the Ge-invariant form

λ (1 − z1/z2) ν/2 −1 −ν/2 λ −1 iϕ λ−ν iϕ λ+ν (−i dz1) (i dz2 ) (dϕ) , z1 = z, z2 =z ¯ . (110) (1 − e /z2) (1 − z1/e ) 3In quantum holography, such an intertwiner is interpreted as a bulk-boundary propagator, where λ is the scaling dimension of the field from the boundary point of view.

20 1 2 Figure 4: The spectra of the operators Q and 2 (Q + ν ) as functions of the spin value ν.

ν Square-integrable spinors: A basis in the Hilbert space HH of square-integrable ν-spinors ν consists of the matrix element functions Uλ,m(g), see equation (75). They can also be written in ν,± terms of the variables (z1, z2) using the new notation ψλ,m: s s Γ(λ + m) Γ(1 − λ + m) Γ(λ − m) Γ(1 − λ − m) ψ ν = ψ ν,+ = (−1)m−ν ψ ν,−. (111) λ,m Γ(λ + ν) Γ(1 − λ + ν) λ,m Γ(λ − ν) Γ(1 − λ − ν) λ,m

ν These functions transform as the basis vectors |mi of the principal series representation C1/4+s2 1 + 1 with λ = 2 + is and m ∈ ν + Z, or the discrete series representation Dλ with λ > 2 and − 1 ν, m ∈ {λ, λ + 1,...}, or the representation Dλ with λ > 2 and ν, m ∈ {−λ, −λ − 1,...}. Thus, ν HH splits into the indicated irreps, each of which enters with multiplicity 1. Symbolically, Z ∞ ν ∼ ν M sgn ν 1 H = C 2 ds ⊕ D , where λ = |ν| − p > , p ∈ {0, 1, 2,...}. (112) H 1/4+s λ 2 0 λ

1 2 The spectrum of the Casimir operator Q, i.e. the set of numbers q = 4 + s and q = λ(1 − λ) in the above equation, is plotted in Figure4. As an aside, the qualitative form of the spectrum has an interesting physical interpretation. 1 2 1 2 An operator closely related to Q, namely, − 2 ∇ = 2 (Q + ν ) describes a quantum particle with spin ν on the hyperbolic plane. It may also be viewed as the Hamiltonian of a nonrelativistic spinless particle with unit mass and electric in a magnetic field of strength ν. In the flat geometry, such a Hamiltonian has a purely discrete spectrum, the Landau levels. This is because all classical trajectories are closed. However, the trajectory of a charged particle on the hyperbolic plane can be either closed or open, depending on the ratio between the velocity and magnetic field:

v v < 1 : > 1 : (113) ν ν

Therefore, for velocities v > ν, i.e. energies greater than ν2/2, the spectrum becomes continuous. Of course, this argument is very rough and does not give the exact spectrum.

21 ν Finally, let us consider the decomposition of the unit operator of the Hilbert space HH into the projectors onto its irreducible components:

Z ∞ s sinh(2πs) X  1 = (2π)−1 ds Πν + λ − 1  Πν (114) cosh(2πs) + cos(2πν) 1/2+is 2 λ 0 λ=|ν|−p>1/2 p=0,1,2,...

ν X ν ν Πλ = |ψλ,mihψλ,m|, m ∈ ν + Z (restricted for discrete series) (115) m

Here we have used the Plancherel measure (79) together with the constraint ν − µ ∈ Z. This is a ν ν slightly more detailed explanation. The basis functions |ψλ,mi are, essentially, the same as |Uλ,mi, which are related to the Plancherel measure. The constraints on λ, ν, m in the above equations follow from those in Eqs. (76), (77). However, in the first case the inner product is defined as an integral over H2, and in the second over Ge. Therefore,

0 ν ν 0 ν ν Uλ,m Uλ0,m0 = 2π δ(ν − ν ) ψλ,m ψλ0,m0 . (116) The relation between the decomposition measures is the inverse one, that is, equation (114) uses the Plancherel measure multiplied by 2π P δ(ν − µ − n). n∈Z ν 2 The projectors Πλ are integral operators with the kernel functions that depend on z, w ∈ H −2 −1 and the integration measure 4(1 − ww¯) dw dw¯. In terms of the variables z1 = z, z2 =z ¯ , −1 w1 = w, w2 =w ¯ , the kernel function is ˚ν X m−ν ˚ν ˚−ν Πλ(z1, z2; w1, w2) = (−1) ψλ,m(z1, z2) ψλ,−m(w1, w2), (117) m

If (w1, w2) = (0, ∞), then only the m = ν term in the sum is nonzero: ˚ν ˚ν λ  Πλ(z1, z2; 0, ∞) = ψλ,ν(z1, z2) = (1 − z1/z2) F λ + ν, λ − ν, 1; z1/z2 . (118) The general case is reduced to this one using a symmetry argument. Indeed, the kernel function ν ν  ν ν  defines a form of degree 2 , − 2 with respect to (z1, z2) and − 2 , 2 with respect to (w1, w2) when both pairs of variables are close to (0, ∞). This form is invariant under linear fractional maps because the projector commutes with the sl2 action. The map z − w V : z 7→ 1 (119) 1 − z/w2 ˚ν sends w1 to 0 and w2 to ∞. Therefore, Πλ(z1, z2; w1, w2) can be written as an arbitrary invariant form of the same type multiplied by some ordinary function of V (z1) and V (z2). More concretely, ˚ν −ν ν ˚ν  Πλ(z1, z2; w1, w2) = (1 − z1/w2) (1 − w1/z2) Πλ V (z1),V (z2); 0, ∞ . (120) Thus,  ν ˚ν 1 − w1/z2 λ  Πλ(z1, z2; w1, w2) = (1 − u) F λ + ν, λ − ν, 1; u , (121) 1 − z1/w2 (z − w )(z − w ) where u = 1 1 2 2 (122) (z1 − w2)(z2 − w1)

22 6 Tensor products of unitary irreps

A standard problem in is to decompose the product of two unitary irreps into irreps with multiplicities. For the group SL(2, R), this task was accomplished by Repka [8,9]. We will not attempt to give a complete solution for Ge. Rather, we will sketch a general recipe and work out the discrete series cases, D± ⊗ D± , which are relatively simple and relevant to the λ1 λ2 SYK model. The problem of splitting the representation U µ1 ⊗ U µ2 is equivalent to finding all intertwiners λ1 λ2 Υ: U µ → U µ1 ⊗ U µ2 , µ = µ + µ (123) λ λ1 λ2 1 2 and selecting those that are normalizable or δ-normalizable with respect to λ. The matrix elements hm1, m2|Υ|mi will be called “Clebsch-Gordan coefficients”. For now, let us not worry about normalizability and discuss a related task: find all Ge-invariant forms

˜ iϕ1 iϕ2 iϕ3  λ1 λ2 λ3 Y e , e , e (dϕ1) (dϕ2) (dϕ3) , (124)

iϕ1 iϕ2 iϕ3 where, e , e , e represent three points on the unit circle. Choosing some numbers µ1, µ2, µ3 with zero sum and using the twisted periodicity conditions, we can extend Y˜ from the 2π + ϕ3 > ϕ1, ϕ2 > ϕ3 to a function of real variables ϕ1, ϕ2, ϕ3. It is understood as a generalized function and may be defined by the Fourier expansion

˜ X m1 m2 m3 Y (z1, z2, z3) = Cm1,m2,m3 z1 z2 z3 (125) m1∈µ1+Z, m2∈µ2+Z m1+m2+m3=0

with the coefficients Cm1,m2,m3 growing at most polynomially. To express the Clebsch-Gordan ˜ coefficients, let λ3 = λ, µ3 = −µ, and let us consider Y as the integral kernel of the intertwiner

Y : F µ → F µ1 ⊗ F µ2 ,Y = Ξµ1± ⊗ Ξµ2± ΥΞµ±. (126) 1−λ λ1 λ2 λ1↑ λ2↑ λ↓

µ± µ± Here Ξλ↓ ,Ξλ↑ are defined by equation (25) and each of the three signs is individually chosen. (This choice only matters when the corresponding irrep belongs to a discrete series.) Thus,

C = c± c± c± hm , m |Υ|mi. (127) m1,m2,−m λ1,m1 λ2,m2 λ,m 1 2 ˜ The invariance of the generating function Y under L1 and L−1 is expressed as linear relations

between Cm1−1,m2,m3 , Cm1,m2−1,m3 , Cm1,m2,m3−1 and between Cm1+1,m2,m3 , Cm1,m2+1,m3 , Cm1,m2,m3+1. In general, this system of equations has multiple linearly independent solutions. But we are considering only those values of λj, mj that correspond to unitary irreps. With this restriction, the linear relations in the allowed region of (m1, m2, m3) are nondegenerate and can be turned into recurrences, which are solved beginning with just two Fourier coefficients. Thus, the solution space is at most two-dimensional. Its general form is easy to guess:

ϕj − ϕk Y˜ eiϕ1 , eiϕ2 , eiϕ3  = a |ϕ |−λ1−λ2+λ3 |ϕ |−λ1+λ2−λ3 |ϕ |λ1−λ2−λ3 , ϕ = 2 sin , (128) 12 13 23 jk 2

where a takes on two different values depending on the cyclic of ϕ1, ϕ2, ϕ3. However, this expression might require regularization when two or three points coincide. The problem

23 arises if any of the singularities is non-integrable, that is, if one of the exponents −λ1 − λ2 + λ3, −λ1 + λ2 − λ3, λ1 − λ2 − λ3 has real part less than or equal to −1 or if the real part of the sum of all three exponents is less than or equal to −2. This can only happen if a discrete series representation is involved. But if, say, U µ1 = D+ , then the generating form (124) is holomorphic λ1 λ1 iϕ1 in z1 = e for |z1| < 1. In this case, the regularization is achieved by analytic continuation. Since both cyclic orders are just limiting cases of z1 being inside the circle, the intertwiner space is one-dimensional.

6.1 D+ ⊗ D+ λ1 λ2 By analogy with SU(2) representations, it is clear that ∞ M D+ ⊗ D+ ∼ D+ . (129) λ1 λ2 = λ1+λ2+n n=0 For each given n, there is a unique (up to an overall factor) intertwiner Υ++ : D+ → D+ ⊗ D+ , λ = λ + λ + n. (130) λ1,λ2;λ λ λ1 λ2 1 2 Its generating function (125) with all three signs in (127) set to “+” is Y˜ ++ + (z , z , w) = zλ1 zλ2 w−λ (z − z )n(1 − z /w)−2λ1−n(1 − z /w)−2λ2−n (131) λ1,λ2,λ 1 2 1 2 2 1 1 2 The rest of this subsection is concerned with the decomposition of identity for D+ ⊗D+ . First, λ1 λ2 we define partial generating functions as the Taylor coefficients with respect to w−1, excluding the + cλ,m factor: X Y˜ ++ (z , z ) = c+ c+ m , m Υ++ m zm1 zm2 . (132) λ1,λ2;λ,m 1 2 λ1,m1 λ2,m1 1 2 λ1,λ2;λ 1 2 m1,m2 + p Recall that cα,α+k = Γ(2α + k)/k!. A straightforward calculation (where we use the Pochham- mer , (α)r = α ··· (α + r − 1)) shows that

s k k! X (2λ1 + n)k (2λ2 + n)k Y˜ ++ (z , z ) = zλ1 zλ2 (z − z )n 1 2 zk1 zk2 λ1,λ2; λ, λ+k 1 2 1 2 2 1 1 2 Γ(2λ + k) k1! k2! k1+k2=k (133) r Γ(2λ + k) = zλ1 zλ2+n+k (1 − z /z )n F−k, 2λ + n, 2λ; 1 − z /z . k! 1 2 1 2 1 1 2 For example, Y˜ ++ (z , z ) = Γ(2λ)−1/2zλ1 zλ2 (z −z )n. Summing up the squares of the Clebsch- λ1,λ2;λ,λ 1 2 1 2 2 1 Gordan coefficients in this special case, we find the norm of the intertwiner: n! (Υ++ )† Υ++ = 1+. (134) λ1,λ2;λ λ1,λ2;λ λ (2λ − 1) Γ(2λ1 + 2λ2 + n − 1) Γ(2λ1 + n) Γ(2λ2 + n) Thus,

X (2λ − 1) Γ(λ − 1 + λ1 + λ2) Γ(λ + λ1 − λ2) Γ(λ − λ1 + λ2) 1+ ⊗ 1+ = Π++ (135) λ1 λ2 n! λ1,λ2; λ λ=λ1+λ2+n n=0,1,2,... Π++ = Υ++ (Υ++ )† (136) λ1,λ2;λ λ1,λ2;λ λ1,λ2;λ

24 The generating function for the (unnormalized) projector Π++ , λ1,λ2;λ X Π˜ ++ (z , z ; w , w ) = Y˜ ++ (z , z ) Y˜ ++ (w−1, w−1), (137) λ1,λ2;λ 1 2 1 2 λ1,λ2;λ,m 1 2 λ1,λ2;λ,m 1 2 m

is calculated by analogy with spinors. We first assume that z2 → 0 and w1 → ∞, so that only the (k , k ) = (k, 0) terms in the expression for Y˜ ++ (z , z ) and the (k , k ) = (0, k) terms 1 2 λ1,λ2;λ,λ+k 1 2 1 2 in Y˜ ++ (w−1, w−1) are present (see Eq. (133)). The general case is reduced to this one using λ1,λ2;λ,λ+k 1 2 symmetry. The result is as follows, where u is defined by equation (122):

(z /w )λ1 (z /w )λ2 Π˜ ++ (z , z ; w , w ) = 1 1 2 2 (−χ)n F2λ + n, 2λ + n, 2λ; χ (138) λ1,λ2;λ 1 2 1 2 2λ 2λ 1 2 (1 − z1/w1) 1 (1 − z2/w2) 2

−1 (z1 − z2)(w1 − w2) λ = λ1 + λ2 + n, χ = 1 − u = (139) (z1 − w1)(z2 − w2)

6.2 D+ ⊗ D− λ1 λ2 The space D+ ⊗ D− maps onto F + ⊗ F − . The latter consists of holomorphic ν-spinors with λ1 λ2 λ1 λ2 ν = λ1 − λ2. However, the norm on the original space differs from the spinor norm defined by the integral over H2. One could use some functional to characterize the relation between the Hilbert spaces D+ ⊗ D− and Hν , cf. Proposition 7.2 in [9]. We instead directly construct the λ1 λ2 H irreducible decomposition Z ∞   + − ∼ ν M sgn ν ν 1 D ⊗ D C 2 ds ⊕ D ⊕ C for λ = λ + λ < (140) λ1 λ2 = 1/4+s λ λ(1−λ) 1 2 2 0 λ=|ν|−p>1/2 p=0,1,2,...

by analogy with the derivation of Eq. (112). In doing so, we reuse the Casimir eigenfunctions as partial generating functions, but calculate their norms using the inner product on D+ ⊗ D− . The λ1 λ2 key observation is that the asymptotics of the spinors and the Clebsch-Gordan coefficients are closely related, and the corresponding integral and sum converge for the same values of λ (with one exception that results in the extra term in Eq. (140)). According to the general , we consider an intertwiner

Υ+− : U ν → D+ ⊗ D− , ν = λ − λ (141) λ1,λ2;λ λ λ1 λ2 1 2 with the generating function

s λ λ Γ(λ ± ν) z  1  w  2 (1 − z /z )λ−λ1−λ2 Y˜ +− ± (z , z , w) = 1 1 2 (142) λ1,λ2,λ 1 2 λ−ν λ+ν Γ(1 − λ ± ν) w z2 (1 − w/z2) (1 − z1/w)

± where the ± sign is linked to cλ,m in Eq. (127). Independent of that sign, the partial generating functions are Y˜ +− (z , z ) = zλ1 z−λ2 (1 − z /z )−λ1−λ2 ψ˚ν (z , z ). (143) λ1,λ2; λ, ν+k 1 2 1 2 1 2 λ,ν+k 1 2 ˚ν where ψλ,m is defined by equation (111) with further reference to Eq. (105).

25 The normalizability of the intertwiner Υ+− is related to the asymptotics of the partial λ1,λ2;λ generating function. We have Y˜ +− (z , z ) = zλ1+kz−λ2 f(z /z ) with λ1,λ2; λ, ν+k 1 2 1 2 1 2

α+ α− f(z) ≈ a+(1−z) +a−(1−z) for z → 1, α+ = λ−λ1 −λ2, α− = 1−λ−λ1 −λ2. (144)

The function f is analytic in the with a branch cut from 1 to +∞. Its n-th Taylor coefficient fn can be expressed as a Cauchy integral over a circle of radius r > 1 and around the branch cut section [1, r], resulting in this asymptotic formula:

−1−α+ −1−α− a+ n a− n fn ≈ + for n → ∞. (145) Γ(−α+) Γ(−α−)

λ −1/2 On the other hand, c+ ≈ m 1 for m → +∞; similarly, c− ≈ (−1)m2+λ2 |m |λ2−1/2 λ1,m1 1 1 λ2,m2 2 for m2 → −∞. Thus, equation (144) translates to the following n → ∞ asymptotics of the Clebsch-Gordan coefficients:  a n−λ a n−1+λ  λ + n + k, −λ − n Υ+− ν + k ≈ (−1)n + + − . (146) 1 2 λ1,λ2;λ Γ(λ1 + λ2 − λ) Γ(λ1 + λ2 − 1 + λ)

It is now easy to see that the vector Υ+− |ν + ki is normalizable or δ-normalizable if the spinor λ1,λ2;λ ˚ν 1 ψλ,ν+k is normalizable or δ-normalizable. In , the said vector is normalizable if λ1 +λ2 < 2 and λ (or, equivalently, 1 − λ) is equal to λ1 + λ2. This case corresponds to the complementary ν series representation Cλ(1−λ). 1 To calculate the norm of the intertwiner for λ = 2 + is, we set k = 0 so that the coefficients a+ = aλ,ν,−ν and a− = a1−λ,ν,−ν are given by Eq. (69). Thus,

+− † +− cosh(2πs) + cos(2πν) 0 ν Υ Υ 0 = δ(s − s ) 1 . (147) λ1,λ2; 1/2+is λ1,λ2; 1/2+is 2 1/2+is 2s sinh(2πs) |Γ(λ1 + λ2 − 1/2 − is)|

+ 1 In the discrete series case Dλ (with λ = λ1 − λ2 − p > 2 and p ∈ {0, 1, 2,...}), it is convenient to consider the Clebsch-Gordan coefficients with m = λ. We proceed with the calculation of the norm: s Γ(2λ + p) ˜ +− p λ1 −λ2−p −2λ2−p Y (z1, z2) = (−1) z z (1 − z1/z2) , (148) λ1,λ2;λ,λ Γ(2λ) p! 1 2 s Γ(2λ + p) Γ(2λ + p + r)(p + r)! λ + r, −λ − p − r Υ+− λ − λ − p = (−1)r 2 , (149) 1 2 λ1,λ2;λ 1 2 2 Γ(2λ) Γ(2λ2 + p) Γ(2λ1 + r) p! r!

1 (Υ+− )†Υ+− = 1+. (150) λ1,λ2;λ λ1,λ2;λ λ (2λ − 1) Γ(λ1 + λ2 − λ) Γ(λ1 + λ2 − 1 + λ)

1 In the special case λ = λ1 + λ2 < 2 , the partial generating function for m = ν is quite simple, namely Y˜ +− (z , z ) = zλ1 zλ2 F(λ , λ , 1; z /z ). Summing up the squares of the Clebsch- λ1,λ2;λ,ν 1 2 1 2 1 2 1 2 Gordan coefficients gives this result: sin(2πλ ) sin(2πλ ) Γ(1 − 2λ) (Υ+− )†Υ+− = 1 2 1ν . (151) λ1,λ2;λ λ1,λ2;λ π2 λ 26 In conclusion, the decomposition of identity is as follows:

Z ∞ 2s sinh(2πs) |Γ(λ + λ − 1/2 − is)|2 1+ ⊗ 1− = ds 1 2 Π+− λ1 λ2 λ1,λ2; 1/2+is 0 cosh(2πs) + cos(2πν) X  +− + 2λ − 1 Γ(λ1 + λ2 − λ) Γ(λ1 + λ2 − 1 + λ)Π λ1,λ2;λ (152) λ=|ν|−p>1/2 p=0,1,2,...  π2 1 + Π+− for λ = λ + λ < λ1,λ2; λ 1 2 sin(2πλ1) sin(2πλ2) Γ(1 − 2λ) 2 where ν = λ1 − λ2. The expression for the projector is similar to that for spinors:

(z /w )λ1 (w /z )λ2 (1 − v)λ−λ1−λ2 Π˜ +− (z , z ; w , w ) = 1 1 2 2 Fλ + ν, λ − ν, 1; v (153) λ1,λ2;λ 1 2 1 2 2λ 2λ (1 − z1/w1) 1 (1 − w2/z2) 2 (z − w )(z − w ) v = u−1 = 1 2 2 1 (154) (z1 − w1)(z2 − w2)

Acknowledgments

I thank Josephine Suh for helpful comments as well as catching a number of errors in the paper draft. I gratefully acknowledge the by the Simons Foundation under grant 376205 and through the “It from Qubit” program, and from the Institute of Quantum Information and Matter, a NSF Frontier center funded in part by the Gordon and Betty Moore Foundation.

References

[1] V. Bargmann, “Irreducible unitary representations of the ”, Annals of Mathe- matics 48 (3), 568–640 (1947). [2] Harish-Chandra, “Plancherel formula for the 2 × 2 real unimodular group”, Proc. Natl. Acad. Sci. U.S.A., 38 (4), 337–342, (1952).

[3] L. Puk´anszky, “The Plancherel formula for the universal of SL(R, 2)”, Math. Annalen 156, 96–143 (1964). [4] S. Sachdev and J. Ye, “Gapless spin-fluid ground state in a random quantum Heisenberg magnet”, Phys. Rev. Lett. 70, 3339 (1993), arXiv:cond-mat/9212030. [5] A. Kitaev, “A simple model of quantum holography”, KITP talks, April 7 and May 27, 2015. [6] J. Maldacena and D. Stanford, “Remarks on the Sachdev-Ye-Kitaev model”, Phys. Rev. D 94, 106002 (2016), arXiv:1604.07818. [7] A. Jevicki, K. Suzuki, J. Yoon, “Bi-local holography in the SYK model”, JHEP 2016 (7), 1–25 (2016), arXiv:1603.06246.

27 [8] J. Repka, “Tensor products of unitary representations of SL2(R)”, Bull. Am. Math. Soc. 82 (6), 930–932 (1976).

[9] J. Repka, “Tensor products of unitary representations of SL2(R)”, Am. J. Math. 100 (4), 747–774 (1978).

28