This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

More on chaos at weak coupling

Rohit R. Kalloor Department of Particle Physics and Astrophysics, Weizmann Institute of Science, Rehovot, Israel Adar Sharon Simons Center for Geometry and Physics, SUNY, Stony Brook, NY 11794, U.S.A.
Abstract

We discuss aspects of the quantum Lyapunov exponent λL\lambda_{L} in theories with an exactly marginal SYK-like random interaction, where λL\lambda_{L} can be computed as a continuous function of the interaction strength 𝒥\mathcal{J}. In 1d1d, we prove a conjecture from [1] which states that at small 𝒥\mathcal{J}, λL\lambda_{L} can be found by considering a specific limit of the four-point function in the decoupled theory. We then provide additional evidence for the 2d2d version of this conjecture by discussing new examples of Lyapunov exponents which can be computed at weak coupling.

1 Introduction

The quantum Lyapunov exponent λL\lambda_{L} is an intriguing and complicated observable in quantum field theories. Of particular interest are theories where λL\lambda_{L} saturates an upper bound [2], since this might indicate that they have semiclassical gravity duals. However, in this paper we will be interested in the opposite limit, and would like to study how the chaos exponent behaves as we turn on some small coupling 𝒥\mathcal{J}. The classical version of this question has been studied extensively, and many interesting behaviors have been found for chaos at weak coupling. For example, the KAM theorem schematically states that an integrable system deformed by a small (integrability-breaking) deformation remains non-chaotic even for a finite but small enough deformation. We will ask an analogous question in a quantum setting: starting with NN decoupled theories and slowly increasing the strength of a random interaction between them, how does λL\lambda_{L} behave?

Our setup is the following. The basic building block is a conformal field theory (CFT) 𝒞\mathcal{C}, called the “core CFT”, which contains a primary Φ\Phi. Next, we take NN copies of the core CFT, and deform it by a random interaction:

𝒞N+Ji1iqddxΦi1Φiq.\mathcal{C}^{N}+J_{i_{1}...i_{q}}\int d^{d}x\Phi_{i_{1}}...\Phi_{i_{q}}\;. (1.1)

The couplings Ji1iqJ_{i_{1}...i_{q}} are random with Gaussian measure and variance Ji1iq2=(q1)!𝒥2Nq1\langle J_{i_{1}...i_{q}}^{2}\rangle=(q-1)!\frac{\mathcal{J}^{2}}{N^{q-1}} (with no sum over repeated indices).111We emphasize that our disorder is spacetime-independent. The deformation should be understood in terms of conformal perturbation theory around NN copies of the CFT 𝒞\mathcal{C}, and we will be studying these theories in the large-NN limit. We will call such theories “disordered CFTs”.

Disordered CFTs are a generalization of ideas appearing in disordered free theories, which can be obtained by setting 𝒞\mathcal{C} to be a free field theory, and setting Φ\Phi to be the corresponding free field. A famous example of disordered free theories is the SYK model [3, 4], and some additional examples can be found in [5, 6, 7, 8, 9, 10, 11, 12, 13]. In general, the random interactions allow for some exact computations in the IR for disordered free fields, assuming the theory flows to an interacting CFT [3, 14, 15]. This was generalized to arbitrary core CFTs in [1, 16]. In these theories, λL\lambda_{L} can be read off from a certain out-of-time-ordered four-point function (or alternatively, the double-commutator) [4, 17, 18, 2]. This is a difficult computation in general, but it is aided by the large-NN limit and conformal invariance in the IR. In particular, we must also restrict ourselves to dimensions d2d\leq 2, since we will be interested in the theory at finite temperature, but in small dimensions a conformal transformation can map such theories to flat space.

An especially interesting subsector of disordered CFTs can be obtained by demanding that the interaction term in (1.1) is exactly marginal, such that 𝒥\mathcal{J} parametrizes a line of CFTs. This behavior is not generic, but can be obtained by using supersymmetry or by considering chiral theories, both of which we will discuss in this paper.222Similar behavior can be obtained in QM without these assumptions [11]. Having a line of CFTs allows us to follow interesting observables as we continuously move between different CFTs, and in particular will allow us to study λL\lambda_{L} as a function of 𝒥\mathcal{J}.

In [1], the chaos exponent was studied in the weak-coupling limit 𝒥0\mathcal{J}\to 0 in disordered CFTs where 𝒥\mathcal{J} is exactly marginal. In particular, in the examples that were discussed it was observed the chaos exponent in the limit 𝒥0\mathcal{J}\to 0 is also given by the leading exponential behavior of the double-commutator (DC) in a single core CFT. In other words, if the DC of a single core CFT behaves at large times as exp(λL0t)\exp(\lambda_{L}^{0}t), then the chaos exponent of the interacting theory in the limit 𝒥0\mathcal{J}\to 0 is given by

λL(J0)=λL0.\lambda_{L}(J\to 0)=\lambda_{L}^{0}\;. (1.2)

It was conjectured that this result is general; we will call it the continuity conjecture.

The equality (1.2) is surprising for two main reasons. First, λL0\lambda_{L}^{0} cannot be interpreted as a chaos exponent in a single core CFT (since some form of a large-NN limit is required for this interpretation), and yet in the interacting theory it dictates the behavior of the chaos exponent at weak coupling. Second, it is possible for λL0\lambda_{L}^{0} to be negative. A negative chaos exponent seems counter-intuitive, and indeed the result cannot be trusted in the standard method of computing λL\lambda_{L} due to some assumptions that are required. A more precise version of the continuity conjecture is then

λL(J0)=max(λL0,0).\lambda_{L}(J\to 0)=\max(\lambda_{L}^{0},0)\;. (1.3)

Despite this fact, if λL0\lambda_{L}^{0} is negative then one can still determine that the chaos exponent of the theory is at most zero for a finite range of values of 𝒥\mathcal{J}. As a result, it is enough to show that in a single core CFT λL0<0\lambda_{L}^{0}<0 in order to find a discontinuous transition into chaos, see figure 1. This is reminiscent of classical KAM theory.

Refer to caption
(a)
Refer to caption
(b)
Figure 1: Two types of behaviors for the dependence of λL\lambda_{L} on the exactly marginal interaction 𝒥\mathcal{J}: (a) continuous and (b) discontinuous.

This paper includes two main results. First we will prove the continuity conjecture (1.3) in quantum mechanics (QM). We then discuss two examples in 2d2d and show that they obey the continuity conjecture, providing further evidence for the 2d2d version of the conjecture. The first example is the chiral SYK model discussed in [12], where the chaos exponent was already found for all 𝒥\mathcal{J}. We compare the result at small 𝒥\mathcal{J} to λL0\lambda_{L}^{0} and find agreement. Next we discuss the disordered 𝒜3\mathcal{A}_{3} minimal model. Since this theory is dual to a free CFT, we can compute all nn-point functions for its primaries, and we use this result to compute the chaos exponent at small 𝒥\mathcal{J}. Comparing the result to λL0\lambda_{L}^{0} we again find agreement.

In the examples discussed in this paper, λL0\lambda_{L}^{0} always turns out to be non-negative, so that the physical picture is that of figure 1(a). We will discuss ideas for how to generate cases with a negative value and a discontinuous transition into chaos.

This paper is organized as follows. In section 2 we introduce the theories we consider in this paper and the methods for computing the chaos exponent. We then prove the continuity conjecture in QM in section 3. In section 4 we discuss our first 2d2d example, the chiral SYK model, and show that the continuity conjecture is obeyed. In section 5 we discuss our second example, the disordered 𝒜3\mathcal{A}_{3} minimal model. We compute the chaos exponent at small 𝒥\mathcal{J} and again show that the continuity conjecture is obeyed. We discuss some future directions in section 6.

2 Background

In this section we review the basic method discussed in [1] for obtaining the chaos exponent λL\lambda_{L} at weak coupling for a general disordered CFT.

2.1 Four-point function and double-commutator

In disordered free theories, it is possible to write down self-consistency equations for the two- and four-point functions of the fundamental field, which can then be solved using a conformal ansatz when the conformal symmetry is restored in the IR. In [1] it was shown that similar equations can be written down for general disordered CFT for the operator Φ\Phi in (1.1), which can in principle be solved for any value of 𝒥\mathcal{J} when it is an exactly marginal deformation. However, solving these equations requires knowledge of all nn-point functions of Φ\Phi in the core CFT, and so is very difficult in general. In this section we will review the derivation in [1], omitting some details.

In the following we will only use the four-point function, and so we only discuss the self-consistency equation for the four-point function and for the DC. The discussion can also be immediately generalized to SUSY theories, in which case the diagrams discussed should be understood as supergraphs.

We would first like to compute the connected four-point function

C=1N2i,j=1NΦiΦiΦjΦjconnC=\frac{1}{N^{2}}\sum_{i,j=1}^{N}\langle\Phi_{i}\Phi_{i}\Phi_{j}\Phi_{j}\rangle_{conn} (2.1)

in the deformed theory (1.1). The diagrams contributing to CC have an iterative ladder structure at large NN, so that the full result for the four-point function is given by

C=n=0KnF0=F01K,C=\sum_{n=0}^{\infty}K^{n}F_{0}=\frac{F_{0}}{1-K}\;, (2.2)

where F0,KF_{0},K are defined in figure 2.

Refer to caption
Figure 2: The kernel KK and initial diagram F0F_{0} for general disordered CFTs. Red lines denote full propagators GG, and black dots denote insertions of the disorder interaction, with q2q-2 red propagators between each pair.

The computation of KK and F0F_{0} requires knowing all “subtracted nn-point functions”, denoted by nsn_{s}^{\prime}. These are given by the standard nn-point functions but with some specific subtractions of lower-order correlators, and are explicitly defined in [1]. Examples of some subtracted nn-point functions appear in figure 3. From these it is in principle possible to compute the full four-point function CC. One can also perform perturbation theory in 𝒥\mathcal{J}; at leading order only the first diagram in the expression for KK in figure 2 contributes, which is determined by the CFT four-point function and two-point function, and takes the form

K(1,2;3,4)=(q1)J2Φ1Φ¯2Φ3Φ¯4sG(3,4)q2,K(1,2;3,4)=(q-1)J^{2}\langle\Phi_{1}\bar{\Phi}_{2}\Phi_{3}\bar{\Phi}_{4}\rangle_{s}^{\prime}G(3,4)^{q-2}\;, (2.3)

with G(3,4)G(3,4) the Φ\Phi propagator between the points x3,x4x_{3},x_{4}.

Next we discuss the computation of the double-commutator, defined as

WR(t1,t2)=1N2i,j=1N[Φi(β/2),Φj(β/2+it2)][Φi(0),Φj(it1)]=limε01N2i,j=1N(Φi(ε)Φi(ε))(Φi(β/2+ε)Φi(β/2ε))Φj(it1)Φj(β/2+it2).\begin{split}W_{R}(t_{1},t_{2})&=\frac{1}{N^{2}}\sum_{i,j=1}^{N}\left\langle[\Phi_{i}(\beta/2),\Phi_{j}(\beta/2+it_{2})][\Phi_{i}(0),\Phi_{j}(it_{1})]\right\rangle\\ &=\lim_{\varepsilon\rightarrow 0}\frac{1}{N^{2}}\sum_{i,j=1}^{N}\left\langle\left(\Phi_{i}\left(\varepsilon\right)-\Phi_{i}\left(-\varepsilon\right)\right)\left(\Phi_{i}\left(\beta/2+\varepsilon\right)-\Phi_{i}\left(\beta/2-\varepsilon\right)\right)\right.\\ &\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\cdot\left.\Phi_{j}\left(it_{1}\right)\Phi_{j}\left(\beta/2+it_{2}\right)\right\rangle\;.\end{split} (2.4)

We have suppressed the spatial coordinates, since they will be unimportant, and we have kept the (real and imaginary) time coordinates. By \langle...\rangle we mean the Euclidean time-ordered thermal trace. Note that (2.4) is just a combination of analytically-continued Euclidean four-point functions on the cylinder, which in a d=2d=2 CFT is an analytically-continued flat-space correlator (2.1).

Refer to caption
Figure 3: Examples of correlation functions nsn_{s}^{\prime}. Dashed lines corresponds to external points, while solid lines are connected via Σ\Sigma’s in figure 2.

The diagrams contributing to the double-commutator also obey an iterative ladder structure. The corresponding kernel KRK_{R} and initial diagram F0,RF_{0,R} have the same diagrammatics as the four-point function (appearing in figure 2), but the correlators are analytically-continued versions of the ones appearing above. The details appear in [1], and we will write down explicitly only the leading contribution in 𝒥\mathcal{J}, which comes from the four-point function:

KR(t1,t2,t3,t4)=𝒥2Δ𝒪(it1)Δ𝒪(β2+it2)𝒪(it3)𝒪(β2+it4)sGlr,Δq2(3,4)+O(𝒥4).K_{R}(t_{1},t_{2},t_{3},t_{4})=\mathcal{J}^{2}\left\langle\Delta\mathcal{O}\left(it_{1}\right)\Delta\mathcal{O}\left(\frac{\beta}{2}+it_{2}\right)\mathcal{O}\left(it_{3}\right)\mathcal{O}\left(\frac{\beta}{2}+it_{4}\right)\right\rangle^{\prime}_{s}G^{q-2}_{lr,\Delta}(3,4)+O(\mathcal{J}^{4})\;. (2.5)

Here we have denoted Δ𝒪(z)=𝒪(z+ε)𝒪(zε)\Delta\mathcal{O}(z)=\mathcal{O}(z+\varepsilon)-\mathcal{O}(z-\varepsilon) where we eventually take the limit ϵ0\epsilon\to 0. The integration range in principle for all points is over the complex time contour appearing in figure 4, although various cancellations as we take ϵ0\epsilon\to 0 lead to the result above. In particular, we have used the expression for the thermal two-point function for a scalar operator of dimension Δ\Delta between points from different “rails” (see figure 4):

Glr,Δ(1,2)=1(4cosh(t12x122)cosh(t12+x122))Δ.G_{lr,\Delta}(1,2)=\frac{1}{\left(4\cosh(\frac{t_{12}-x_{12}}{2})\cosh(\frac{t_{12}+x_{12}}{2})\right)^{\Delta}}\;. (2.6)
Refer to caption
Figure 4: The complex time contour chosen for the computation of the DC. Red dots denote insertion points of operators. We call the two excursions from the real axis the “left rail” and the “right rail”.

2.2 Chaos

Having written down the ladder diagrams which contribute to the DC, we can now compute the chaos exponent. In principle, this is done by computing the DC explicitly, and then taking the large-time limit t1,t2=tt_{1},t_{2}=t\to\infty, which should lead to exponentially growing behavior,

WR(t1,t2)exp(λL2(t1+t2))f(t1t2).W_{R}(t_{1},t_{2})\sim\exp(\frac{\lambda_{L}}{2}(t_{1}+t_{2}))f(t_{1}-t_{2})\;. (2.7)

where from now on we set β=2π\beta=2\pi. However, computing the full DC is difficult, and instead the existence of an iterative ladder structure offers us a shortcut. The ladder structure leads to the self-consistency equation

WR=F0,R+KRWR.W_{R}=F_{0,R}+K_{R}W_{R}\;. (2.8)

At large times, we assume that the F0,RF_{0,R} term is negligible, and so WRW_{R} obeys the equation

WR=KRWR,W_{R}=K_{R}W_{R}\;, (2.9)

which is just an eigenvalue equation for KRK_{R}. As a result, the exponentially-growing solution for WRW_{R} must be an eigenfunction of the retarded kernel KRK_{R} with eigenvalue 11. Thus, the chaos exponent is found by guessing solutions of the form (2.7) and finding their eigenvalue kR(λL)k_{R}(\lambda_{L}) under KRK_{R}. The largest λL\lambda_{L} for which kR(λL)=1k_{R}(\lambda_{L})=1 is the chaos exponent.

The precise form of the eigenvalue is constrained due to conformal invariance, and takes the form of a two-point function on the cylinder. In 2d2d we have the ansatz

Wλ(1,2)=exp(h+h~2(t1+t2)+hh~2(x1+x2))(2cosh(t12x122))Δh(2cosh(t12+x122))Δh~.W_{\lambda}(1,2)=\frac{\exp(-\frac{h+\tilde{h}}{2}(t_{1}+t_{2})+\frac{h-\tilde{h}}{2}(x_{1}+x_{2}))}{(2\cosh(\frac{t_{12}-x_{12}}{2}))^{\Delta-h}(2\cosh(\frac{t_{12}+x_{12}}{2}))^{\Delta-\tilde{h}}}\;. (2.10)

In general we require h=λ2+ip,h~=λ2iph=-\frac{\lambda}{2}+ip,\tilde{h}=-\frac{\lambda}{2}-ip for real λ,p\lambda,p, but in practice in the examples appearing here the maximal chaos exponent will have p=0p=0.

Finding the chaos exponent now amounts to computing the eigenvalue of the eigenfunction WλW_{\lambda} under KRK_{R}. Since the theories we consider are conformally invariant for any value of 𝒥\mathcal{J}, we can perform this computation in a perturbative expansion in 𝒥\mathcal{J}. Generally the eigenvalue is given by

kR(λ,𝒥)=d2x3d2x4KRWW,k_{R}(\lambda,\mathcal{J})=\frac{\int d^{2}x_{3}d^{2}x_{4}K_{R}\cdot W}{W}\;, (2.11)

and plugging in the leading-order result for KRK_{R} we find that in 2d2d and at leading order in 𝒥\mathcal{J},

kR(λ,𝒥)=𝒥2d2x3d2x4Δ𝒪(it1)Δ𝒪(β/2+it2)𝒪(it3)𝒪(β/2+it4)sGlr,Δ+λ2(3,4)Glr,Δ+λ2(1,2)exp(λ2(t3+t4t1t2))Glr,2Δ(q2)(3,4)+O(J4)=𝒥24exp(λ2(t1+t2))Glr,λ/2(1,2)u1𝑑u3u2𝑑u4u342+λ2v1𝑑v3v2𝑑v4v342+λ2𝒢R(χ,χ¯)+O(𝒥4).\begin{split}k_{R}(\lambda,\mathcal{J})=&\mathcal{J}^{2}\int d^{2}x_{3}d^{2}x_{4}\left\langle\Delta\mathcal{O}\left(it_{1}\right)\Delta\mathcal{O}\left(\beta/2+it_{2}\right)\mathcal{O}\left(it_{3}\right)\mathcal{O}\left(\beta/2+it_{4}\right)\right\rangle^{\prime}_{s}\\ &\cdot\frac{G_{lr,\Delta+\frac{\lambda}{2}}(3,4)}{G_{lr,\Delta+\frac{\lambda}{2}}(1,2)}\exp\left(\frac{\lambda}{2}(t_{3}+t_{4}-t_{1}-t_{2})\right)\cdot G_{lr,2\Delta(q-2)}(3,4)+O(J^{4})\\ =&\frac{\mathcal{J}^{2}}{4}\frac{\exp(-\frac{\lambda}{2}(t_{1}+t_{2}))}{G_{lr,\lambda/2}(1,2)}\frac{\int^{\infty}_{u_{1}}du_{3}\int_{-\infty}^{u_{2}}du_{4}}{u_{34}^{2+\frac{\lambda}{2}}}\frac{\int_{v_{1}}^{\infty}dv_{3}\int_{-\infty}^{v_{2}}dv_{4}}{v_{34}^{2+\frac{\lambda}{2}}}\mathcal{G}_{R}(\chi,\bar{\chi})+O(\mathcal{J}^{4})\;.\end{split} (2.12)

With χ,χ¯\chi,\bar{\chi} the conformal cross-ratios. Here we have performed the change of variables

u3=ex3t3,v3=ex3t3,u4=ex4t4,v4=ex4t4.\begin{split}u_{3}=e^{x_{3}-t_{3}}\;,&\quad v_{3}=e^{-x_{3}-t_{3}}\;,\\ u_{4}=-e^{x_{4}-t_{4}}\;,&\quad v_{4}=-e^{-x_{4}-t_{4}}\;.\end{split} (2.13)

𝒢R\mathcal{G}_{R} is the retarded normalized 4-point of the undeformed CFT at 𝒥=0\mathcal{J}=0,

𝒢R(χ,χ¯)=[O(β/2+it2),O(β/2+it4)][O(it1),O(it3)]sGlr,Δ(1,2)Glr,Δ(3,4)=limε1,ε20𝒢++𝒢+𝒢++𝒢\begin{split}\mathcal{G}_{R}(\chi,\bar{\chi})&=\frac{\left\langle[O(\beta/2+it_{2}),O(\beta/2+it_{4})][O(it_{1}),O(it_{3})]\right\rangle_{s}^{\prime}}{G_{lr,\Delta}(1,2)G_{lr,\Delta}(3,4)}\\ &=\lim_{\varepsilon_{1},\varepsilon_{2}\rightarrow 0}\mathcal{G}_{++}-\mathcal{G}_{+-}-\mathcal{G}_{-+}+\mathcal{G}_{--}\end{split} (2.14)

with the normalized four-point 𝒢±1,±2=𝒢(u1e±iε1,v1e±iε1,u2e±iε2,v1e±iε2,u3,v3,u4,v4)\mathcal{G}_{\pm_{1},\pm_{2}}=\mathcal{G}(u_{1}e^{\pm i\varepsilon_{1}},v_{1}e^{\pm i\varepsilon_{1}},u_{2}e^{\pm i\varepsilon_{2}},v_{1}e^{\pm i\varepsilon_{2}},u_{3},v_{3},u_{4},v_{4}).

2.2.1 Consistency of the perturbative expansion

We now discuss perturbative corrections to the chaos exponent which are subleading in the 𝒥0\mathcal{J}\to 0 limit. The computation of the chaos exponent in perturbation theory in 𝒥\mathcal{J} requires several assumptions. Expanding the eigenvalues of the retarded kernel, we expect to see

kR(λ,𝒥)=𝒥2f2(λ)+𝒥4f4(λ)+,k_{R}(\lambda,\mathcal{J})=\mathcal{J}^{2}f_{2}(\lambda)+\mathcal{J}^{4}f_{4}(\lambda)+...\;, (2.15)

First, we would like to understand when it is enough to find leading term f2(λ)f_{2}(\lambda) discussed above in order to find the leading value of the chaos exponent in the limit 𝒥0\mathcal{J}\to 0. The chaos exponent is found by setting kR=1k_{R}=1, and so if we keep only the 𝒥2\mathcal{J}^{2} term, λL\lambda_{L} is found by analyzing at which values of λ\lambda the function f2(λ)f_{2}(\lambda) diverges as 1/𝒥21/\mathcal{J}^{2}. The largest such λ\lambda can then be identified with the chaos exponent. In order for this procedure to be consistent it is enough to require two conditions on the higher-order terms:

  1. 1.

    If the largest value of λ\lambda at which f2f_{2} diverges is λ0\lambda_{0}, then all other fnf_{n} diverge at values λλ0\lambda\leq\lambda_{0}.

  2. 2.

    If f2f_{2} diverges as 1(λλ0)α\frac{1}{(\lambda-\lambda_{0})^{\alpha}}, then other fnf_{n} diverge as 1(λλ0)βn\frac{1}{(\lambda-\lambda_{0})^{\beta_{n}}} for βnnα/2\beta_{n}\leq n\alpha/2.

In this case the leading value of λL\lambda_{L} is indeed λ0\lambda_{0}. All examples discussed in [1] were shown to obey these conditions, and we will show that the examples discussed here also obey them).

Next, it is natural to ask when it is consistent to perform a perturbative expansion around λ0\lambda_{0} in the limit of small 𝒥\mathcal{J} in order to obtain the chaos exponent in a series in 𝒥\mathcal{J} of the form

λL=λ0+𝒥2λ1+\lambda_{L}=\lambda_{0}+\mathcal{J}^{2}\lambda_{1}+... (2.16)

This requires a strict inequality in the second condition above, i.e. we require

βn<nα/2.\beta_{n}<n\alpha/2\;. (2.17)

To see why this is the case, assume an expansion of the form

kR=an𝒥2nλ2n,k_{R}=\sum a_{n}\frac{\mathcal{J}^{2n}}{\lambda^{2n}}\;, (2.18)

so that α=2\alpha=2 and βn=n=nα/2\beta_{n}=n=n\alpha/2 and the inequality is exactly saturated. As a result, the expansion is not in terms of 𝒥\mathcal{J}, but in terms of 𝒥/λ\mathcal{J}/\lambda. We can now try to compute λL\lambda_{L} in perturbation theory. Write

λL=λ0+𝒥2λ1+\lambda_{L}=\lambda_{0}+\mathcal{J}^{2}\lambda_{1}+... (2.19)

Plugging in this expansion, we immediately learn that the leading order is given by λ0=0\lambda_{0}=0, as discussed above. However, in order to find the value of the subleading term λ1\lambda_{1}, we must take into account all of the terms in the expansion, since all terms they are all of the same order in 𝒥\mathcal{J}. As a result, while we can still compute λ0\lambda_{0} in such theories, we cannot perform perturbation theory around it without knowing kRk_{R} completely (or at least to all orders in some double-scaling limit). So a perturbative calculation of λL\lambda_{L} beyond the leading order using this method fails. We will see this happen in the chiral SYK example discussed in section 4.

On the other hand, for any theory where there is a strict inequality βn<nα/2\beta_{n}<n\alpha/2, a perturbative expansion in 𝒥\mathcal{J} should be possible. This is the result in e.g. the generalized free fields examples in [1].

2.3 The continuity conjecture

We now discuss the behavior of the chaos exponent as 𝒥0\mathcal{J}\to 0. Note that in equation (2.12), the prefactor 𝒥2\mathcal{J}^{2} vanishes in this limit, and so in order for us to be able to solve for kR(λL,𝒥)=1k_{R}(\lambda_{L},\mathcal{J})=1, we must find values of λ\lambda for which the integral in (2.12) diverges as 1/𝒥21/\mathcal{J}^{2}. The chaos exponent in this limit is then given by the largest value of λ\lambda for which this divergence occurs. Assuming that the retarded four-point function of the theory at a single core CFT behaves at large times t1,t2t_{1},t_{2} as333In d2d\leq 2 this is related to the Regge limit.

𝒢R(t1,t2)exp(λL0(t1+t2)/2),\mathcal{G}_{R}(t_{1},t_{2})\sim\exp(\lambda_{L}^{0}(t_{1}+t_{2})/2)\;, (2.20)

it is easily seen that the divergence occurs precisely at λ=λL0\lambda=\lambda_{L}^{0}. As a result, it was conjectured in [1] that the chaos exponent in the limit 𝒥0\mathcal{J}\to 0 is given precisely by the leading exponential behavior of the double-commutator in the free theory at 𝒥=0\mathcal{J}=0. Thus the computation of the leading behavior of λL\lambda_{L} is particularly simple, and is a property of a single core CFT.444Note that λL0\lambda_{L}^{0} does not have any interpretation in terms of a chaos exponent in a single core CFT, since this requires some form of a large-NN limit or another weak-coupling expansion. Several examples were discussed in [1] which were shown to obey this conjecture.

3 Proving the continuity conjecture in QM

The continuity conjecture discussed in 2.3 relates the chaos exponent in disordered CFTs at small 𝒥\mathcal{J} to the late-time behavior of the DC in a single core CFT. It states that assuming that 𝒢R\mathcal{G}_{R} defined in (2.14) (or equivalently, the double-commutator) behaves at large times as exp(λL0t)\exp(\lambda_{L}^{0}t), then the chaos exponent as 𝒥0\mathcal{J}\to 0 approaches λL0\lambda_{L}^{0}. We will now prove this statement under the assumption that perturbation theory in 𝒥\mathcal{J} is valid for the eigenvalues of the retarded kernel, so that the leading order is obtained by considering just the contribution of the four-point function to the retarded kernel, see section 2.2.

Consider the leading contribution to kRk_{R}, which comes from the four-point function diagram in figure 2. In 1d1d this contribution simplifies to (see Appendix B for a derivation):

kR=1|z12|2Δz1𝑑z3z2𝑑z4𝒢R(χ)|z34|2+λ.k_{R}=\frac{1}{|z_{12}|^{2\Delta}}\int^{\infty}_{z_{1}}dz_{3}\int^{z_{2}}_{-\infty}dz_{4}\frac{\mathcal{G}_{R}(\chi)}{|z_{34}|^{2+\lambda}}\;. (3.1)

Here, we performed the change of variables z=etz=e^{-t} on the “left rail” (points 1,31,3) and z=etz=-e^{-t} on the “right rail” (points 2,42,4), see e.g. [5] for details. The conformal cross-ratio is

χ=z12z34z13z24.\chi=\frac{z_{12}z_{34}}{z_{13}z_{24}}\;. (3.2)

Next we perform another change of coordinates to the coordinates κ,χ\kappa,\chi, where κ\kappa is defined as z3=κ/χz_{3}=\kappa/\chi. The integral becomes is

kR(λ)=0𝑑χ𝒢R(χ)χ1λχ1dκκ1+λ(κχ)1+λ(κ(χ1))λ.k_{R}(\lambda)=\int_{-\infty}^{0}d\chi\frac{\mathcal{G}_{R}(\chi)}{\chi^{1-\lambda}}\int_{-\infty}^{\chi-1}\frac{d\kappa}{\kappa^{1+\lambda}(\kappa-\chi)^{1+\lambda}(\kappa-(\chi-1))^{-\lambda}}\;. (3.3)

The chaos exponent λL\lambda_{L} is now given by the largest value of λ\lambda for which the integral diverges in the limit t3,t4t_{3},t_{4}\to-\infty, which in these coordinates corresponds to χ0\chi\rightarrow 0. Note that the κ\kappa integral is finite as long as λ>1\lambda>-1 (corresponding to λL0>1\lambda_{L}^{0}>-1). At χ=0\chi=0 the κ\kappa integral is smooth; it does not vanish or diverge. Therefore we can expand 𝒢R(χ)\mathcal{G}_{R}(\chi) around χ=0\chi=0 inside the integral. Let us denote the leading order term by 𝒢R(χ)=c0χλL0+\mathcal{G}_{R}(\chi)=c_{0}\chi^{-\lambda_{L}^{0}}+.... Note that since at large times we have χet\chi\sim e^{-t}, we can identify λL0\lambda_{L}^{0} with the rate of growth of the DC of a single core CFT as in (2.20). Plugging in this expansion, we see that the integral converges only for λ>λL0\lambda>\lambda_{L}^{0}, which means that the chaos exponent at small 𝒥\mathcal{J} is λL0\lambda_{L}^{0}. We have thus proved the continuity conjecture in 1d1d.

4 The chiral SYK model

The chiral SYK model was introduced in [12] (see also [19]). The theory is a disordered free theory in 2d2d, where the core CFT is a free chiral Majorana fermion. The theory was shown to have a line of fixed points characterized by 𝒥\mathcal{J}, so that the chaos exponent can be found as a function of 𝒥\mathcal{J}. Since the theory is chiral, instead of the standard chaos exponent it is more interesting to consider the velocity-dependent chaos exponent, given by considering a large time and large distance limit with the ratio v=x/tv=x/t kept constant. The corresponding chaos exponent is denoted λv\lambda_{v}. It was found that the chaos exponent always starts at zero as 𝒥0\mathcal{J}\to 0, and rises as we increase 𝒥\mathcal{J}. In particular, choosing vv such that λv\lambda_{v} is maximal, one finds that at infinite 𝒥\mathcal{J} the chaos exponent λv\lambda_{v} saturates the bound on chaos [2]. We will be interested in the weak-coupling limit, where 𝒥\mathcal{J} is close to zero. We will show that the continuity conjecture is obeyed for the velocity-dependent chaos exponent, and discuss corrections in 𝒥\mathcal{J}.

The velocity-dependent chaos exponent at weak coupling can be easily extracted from Appendix B of [12] (see equation (B.6)), and in the limit 𝒥0\mathcal{J}\to 0 one finds

λv(𝒥)={2πβη(v1)+O(𝒥),u<v<u+0,else\lambda_{v}(\mathcal{J})=\begin{cases}\frac{2\pi}{\beta}\eta(v-1)+O(\mathcal{J}),&u_{-}<v<u_{+}\\ 0,&\text{else}\end{cases} (4.1)

where u±=1±𝒥2πu_{\pm}=1\pm\frac{\mathcal{J}}{2\pi} and

η=31𝒥21v𝒥2(1v)2.\eta=\frac{\sqrt{3}}{\sqrt{1-\mathcal{J}^{2}}}\frac{1-v}{\sqrt{\mathcal{J}^{2}-(1-v)^{2}}}\;. (4.2)

Note that since u<v<u+u_{-}<v<u_{+}, η\eta is finite in the limit 𝒥0\mathcal{J}\to 0.

Since the theory is chiral, it is not surprising that the chaos exponent vanishes outside of a cone around the speed of light. Taking the strict 𝒥0\mathcal{J}\to 0 limit, we find that λv(𝒥0)\lambda_{v}(\mathcal{J}\to 0) vanishes trivially unless v=1v=1 (since u=u+=1u_{-}=u_{+}=1), but at v=1v=1 it turns out to vanish as well. Thus we find λv(𝒥0)=0\lambda_{v}(\mathcal{J}\to 0)=0.

We would like to compare this to λL0\lambda_{L}^{0}, which describes the large-time behavior of the DC in a single copy of the free core CFT. The DC in the core CFT is given by GR(13)GR(24)G_{R}(13)G_{R}(24), where GR(ij)G_{R}(ij) is the retarded propagator between spacetime points i,ji,j and is given by

GR(t,x)=1βu+uΘ(tu+1x)Θ(u1xt)sinh[πβ(tu+1x)]sinh[πβ(u1xt)].G_{R}(t,x)=\frac{1}{\beta\sqrt{u_{+}u_{-}}}\frac{\Theta\left(t-u_{+}^{-1}x\right)\Theta\left(u_{-}^{-1}x-t\right)}{\sqrt{\sinh\left[\frac{\pi}{\beta}\left(t-u_{+}^{-1}x\right)\right]\sinh\left[\frac{\pi}{\beta}\left(u_{-}^{-1}x-t\right)\right]}}\;. (4.3)

Next we take the large-time limit where t=t1=t2t=t_{1}=t_{2} and x=x1=x2x=x_{1}=x_{2} are large while v=x/tv=x/t is kept constant. We find (up to unimportant constants)

DC{exp(πβ(u1u+1)vt),u<v<u+0,elseDC\propto\begin{cases}\exp\left(-\frac{\pi}{\beta}(u_{-}^{-1}-u_{+}^{-1})vt\right),&u_{-}<v<u_{+}\\ 0,&\text{else}\end{cases} (4.4)

Plugging in 𝒥=0\mathcal{J}=0 we find as a result that λv0=0\lambda_{v}^{0}=0 always. Thus we have found

λv0=λv(𝒥0)=0,\lambda_{v}^{0}=\lambda_{v}(\mathcal{J}\to 0)=0\;, (4.5)

and so the continuity conjecture is obeyed.

We can also try to understand subleading corrections to the chaos exponent as discussed in section 2.2.1. The eigenvalues of the chiral SYK model were computed in [12], and we can expand the result in 𝒥\mathcal{J}:

kR=3𝒥2λ26𝒥3λ3+O(𝒥4).k_{R}=\frac{3\mathcal{J}^{2}}{\lambda^{2}}-\frac{6\mathcal{J}^{3}}{\lambda^{3}}+O\left(\mathcal{J}^{4}\right)\;. (4.6)

Expanding to higher orders one finds α=2\alpha=2 and βn=n=nα/2\beta_{n}=n=n\alpha/2 in the notation of section 2.2.1. As a result, the expansion is not in terms of 𝒥\mathcal{J}, but in terms of 𝒥/λ\mathcal{J}/\lambda, and so a perturbative computation of λL\lambda_{L} will fail at higher orders in 𝒥\mathcal{J}.

5 The disordered 𝒩=2\mathcal{N}=2 𝒜3\mathcal{A}_{3} minimal model

We now discuss the case where the core CFT is the 𝒩=2\mathcal{N}=2 supersymmetric 𝒜3\mathcal{A}_{3} minimal model. This model can be constructed using a single chiral superfield XX with superpotential

W=X4.W=X^{4}\;. (5.1)

The model has central charge c=3/2c=3/2 and its spectrum includes a chiral primary of dimension 1/41/4 which we will call Φ\Phi. This theory can be identified with the theory of a free boson HH an a free fermion χ\chi, which combine into a single free 𝒩=1\mathcal{N}=1 chiral multiplet [20]. The core CFT thus has a free field representation.

The disordered 𝒜3\mathcal{A}_{3} minimal model with q=4q=4 can then be constructed as

(𝒜3)N+i1i4Ji1i4d2xd2θΦi1Φi4.(\mathcal{A}_{3})^{N}+\sum_{i_{1}\neq...\neq i_{4}}J_{i_{1}...i_{4}}\int d^{2}xd^{2}\theta\Phi_{i_{1}}...\Phi_{i_{4}}\;. (5.2)

In particular, this deformation is marginal. In fact, as discussed in [1], it can be shown using standard arguments [21, 22, 23] that any realization of this theory is a CFT, so that the interaction is exactly marginal (even without averaging).

In order to compute λL\lambda_{L} we need to know nn-point functions of Φ\Phi. Since the 𝒜3\mathcal{A}_{3} minimal model has a free field realization in terms of 𝒩=1\mathcal{N}=1 superfields, we should be able to identify the components of Φ\Phi with products of operators from the free boson and free fermion CFT. In practice, the various components will be mapped to products of vertex operators and fermionic twist fields, whose nn-point functions are known. Plugging the results into the retarded kernel, we will be able to read off λL\lambda_{L}. Following the discussion above, we will focus on the four-point function which gives the leading contribution to λL\lambda_{L} at small 𝒥\mathcal{J}.

Our notations appear in appendix A. In particular, we use lightcone coordinates x±x^{\pm} so that a two-point function takes the form (x+x)Δ|x|2Δ(x^{+}x^{-})^{\Delta}\equiv|x|^{2\Delta}.

5.1 Details of duality to free fields

The free field representation of the 𝒜3\mathcal{A}_{3} minimal model consists of a free Majorana fermion χ\chi and a free compact boson HH at the self-dual radius R=1/2R=1/\sqrt{2} (see for instance [20] or [24]). The SUSY algebra is generated by the operators

G±\displaystyle G_{\pm} =χ±exp(i2H±),\displaystyle=\chi_{\pm}\exp\left(i\sqrt{2}H_{\pm}\right)\;,
G¯±\displaystyle\bar{G}_{\pm} =χ±exp(i2H±),\displaystyle=\chi_{\pm}\exp\left(-i\sqrt{2}H_{\pm}\right)\;,
j±(R)\displaystyle j_{\pm}^{(R)} =i2±H,\displaystyle=\frac{i}{\sqrt{2}}\partial_{\pm}H\;, (5.3)

where ±\pm stand for the left/right moving parts of the fields and operators.555The theory actually has 𝒩=3\mathcal{N}=3 SUSY, but we will only need the 𝒩=2\mathcal{N}=2 subalgebra above for our purposes.

The fundamental superfield Φ\Phi has scaling dimensions (18,18)\left(\tfrac{1}{8},\tfrac{1}{8}\right), and there are five superprimaries in the NS sector: Φ\Phi, Φ2\Phi^{2}, Φ¯\bar{\Phi}, Φ¯2\bar{\Phi}^{2}, Φ¯Φ\bar{\Phi}\Phi. Their bottom components map to the following free theory operators:

ϕ=σexp(iH22),\displaystyle\phi=\sigma\exp\left(i\frac{H}{2\sqrt{2}}\right)\;, ϕ¯=σexp(iH22),\displaystyle\qquad\bar{\phi}=\sigma\exp\left(-i\frac{H}{2\sqrt{2}}\right)\;,
ϕ2=exp(iH2),\displaystyle\phi^{2}=\exp\left(i\frac{H}{\sqrt{2}}\right)\;, ϕ¯2=exp(iH2),\displaystyle\qquad\bar{\phi}^{2}=\exp\left(-i\frac{H}{\sqrt{2}}\right)\;,
ϕ¯ϕ\displaystyle\bar{\phi}\phi =ϵ=χ+χ,\displaystyle=\epsilon=\chi_{+}\chi_{-}\;, (5.4)

where σ\sigma and μ\mu (which will be of use later), with (h,h¯)=(1/16,1/16)(h,\bar{h})=(1/16,1/16), are the twist fields of the fermion theory. The other components of these superfields may be worked out via free field OPEs (see Appendix A). For convenience, we list in Table 1 the components of the basic superfield:

Φ(y+,y)=ϕ(y+,y)+θ+ψ+(y+,y)+θψ(y+,y)+θ+θF(y+,y)\displaystyle\Phi(y^{+},y^{-})=\phi(y^{+},y^{-})+\theta^{+}\psi_{+}(y^{+},y^{-})+\theta^{-}\psi_{-}(y^{+},y^{-})+\theta^{+}\theta^{-}F(y^{+},y^{-}) (5.5)

along with their dimensions and R-charges.

Table 1: The operators in the multiplet of Φ\Phi in the 𝒜3\mathcal{A}_{3} minimal model.
𝒪\mathcal{O} 𝒪free\mathcal{O}_{\text{free}} (h,h¯)(h,\bar{h}) qRq_{R}
ϕ\phi σeiH22\sigma\ e^{i\frac{H}{2\sqrt{2}}} (18,18)\left(\frac{1}{8},\frac{1}{8}\right) (14,14)\left(\frac{1}{4},\frac{1}{4}\right)
ψ+\psi_{+} μei3H++H22\mu\ e^{i\frac{-3H_{+}+H_{-}}{2\sqrt{2}}} (58,18)\left(\frac{5}{8},\frac{1}{8}\right) (34,14)\left(-\frac{3}{4},\frac{1}{4}\right)
ψ\psi_{-} μeiH+3H22\mu\ e^{i\frac{H_{+}-3H_{-}}{2\sqrt{2}}} (18,58)\left(\frac{1}{8},\frac{5}{8}\right) (14,34)\left(\frac{1}{4},-\frac{3}{4}\right)
FF σei3H22\sigma\ e^{-i\frac{3H}{2\sqrt{2}}} (58,58)\left(\frac{5}{8},\frac{5}{8}\right) (34,34)\left(-\frac{3}{4},-\frac{3}{4}\right)

5.2 Correlation functions

Having identified the components of the superfield Φ\Phi with operators from free field theories, we can now compute all nn-point functions of the field Φ\Phi.

5.2.1 Two-point function

Superconformal symmetry fixes the form of the two-point function to be

ΦΦ¯=1|12|2Δ\langle\Phi\bar{\Phi}\rangle=\frac{1}{|\langle 12\rangle|^{2\Delta}} (5.6)

where the relevant value for our theory is Δ=1/4\Delta=1/4. By expanding the result in the superspace coordinates on both sides, we can identify two-point functions of the various components of Φ\Phi, which allows us to set their normalization. We find

ϕϕ¯=1|x12|2Δ,FF¯=4Δ2|x12|2(Δ+1),ψ+ψ¯+=2Δ|x12|2Δx12+,ψψ¯=2Δ|x12|2Δx12.\begin{split}\langle\phi\bar{\phi}\rangle=\frac{1}{|x_{12}|^{2\Delta}}\;,&\qquad\langle F\bar{F}\rangle=\frac{4\Delta^{2}}{|x_{12}|^{2(\Delta+1)}}\;,\\ \langle\psi^{+}\bar{\psi}^{+}\rangle=-\frac{2\Delta}{|x_{12}|^{2\Delta}x_{12}^{+}}\;,&\qquad\langle\psi^{-}\bar{\psi}^{-}\rangle=-\frac{2\Delta}{|x_{12}|^{2\Delta}x_{12}^{-}}\;.\end{split} (5.7)

5.2.2 Four-point function

We move on to the computation of the four-point function. Superconformal symmetry fixes its form to be

Φ(x1)Φ¯(x2)Φ(x3)Φ¯(x4)=1|1234|12f(χs+,χs),\displaystyle\langle\Phi(x_{1})\bar{\Phi}(x_{2})\Phi(x_{3})\bar{\Phi}(x_{4})\rangle=\frac{1}{|\langle 12\rangle\langle 34\rangle|^{\frac{1}{2}}}f(\chi_{s}^{+},\chi_{s}^{-})\;, (5.8)

where ff is an arbitrary function of the superconformal cross ratios

χs±=12±34±14±32±.\chi_{s}^{\pm}=\frac{\langle 12\rangle^{\pm}\langle 34\rangle^{\pm}}{\langle 14\rangle^{\pm}\langle 32\rangle^{\pm}}\;. (5.9)

Note that the bottom component of χs±\chi_{s}^{\pm} is the usual non-supersymmetric cross-ratio χ±=x12±x34±x14±x32±\chi^{\pm}=\frac{x_{12}^{\pm}x_{34}^{\pm}}{x_{14}^{\pm}x_{32}^{\pm}}.

We start by calculating the bottom component of this four point function. Using the identification (5.1), this is equivalent to computing the product of a four-point function of vertex operators and of twist operators σ\sigma. The results are well known, and we find

ϕ(x1)ϕ¯(x2)ϕ(x3)ϕ¯(x4)=|1x12x34|12121+|χ|+|χ1|.\langle\phi(x_{1})\bar{\phi}(x_{2})\phi(x_{3})\bar{\phi}(x_{4})\rangle=\left|\frac{1}{x_{12}x_{34}}\right|^{\frac{1}{2}}\frac{1}{\sqrt{2}}\sqrt{1+|\chi|+|\chi-1|}\;. (5.10)

Since there is a single superconformal cross-ratio of each chirality, it is simple to uplift this result to the full supermultiplet; it suffices to replace χ±χS±\chi^{\pm}\to\chi_{S}^{\pm} and the prefactor of |1x12x34|12\left|\frac{1}{x_{12}x_{34}}\right|^{\frac{1}{2}} with its supersymmetric analog 1|1234|12\frac{1}{|\langle 12\rangle\langle 34\rangle|^{\frac{1}{2}}}. The result is thus

Φ(x1)Φ¯(x2)Φ(x3)Φ¯(x4)=1|1234|12121+|χs|+|χs1|.\displaystyle\langle\Phi(x_{1})\bar{\Phi}(x_{2})\Phi(x_{3})\bar{\Phi}(x_{4})\rangle=\frac{1}{|\langle 12\rangle\langle 34\rangle|^{\frac{1}{2}}}\frac{1}{\sqrt{2}}\sqrt{1+|\chi_{s}|+|\chi_{s}-1|}\;. (5.11)

As a consistency check, we have checked that expanding both sides in the superspace coordinates gives the expected result for four-point functions of other components of Φ\Phi.

5.3 Chaos

We can now compute the chaos exponent using the procedure outlined in section 2.2. This requires diagonalizing the retarded kernel. We perform this diagonalization by first writing down the kernel for the standard four-point function, and then performing the analytic continuation required to obtain the retarded kernel. In the following we will focus on the bottom component of all four-point functions, assuming that they produce that leading behavior at long times, as was the case in similar models [5, 6, 1].

5.3.1 The chaos exponent at weak coupling

We start with the eigenvalues of the standard kernel. The eigenvalues of the bosonic kernel are given by

k(h,𝒥)=d2X3d2X¯4KWW,k(h,\mathcal{J})=\frac{\int d^{2}X_{3}d^{2}\bar{X}_{4}K\cdot W}{W}\;, (5.12)

where d2X=d2xd2θd^{2}X=d^{2}xd^{2}\theta and d2X¯=d2xd2θ¯d^{2}\bar{X}=d^{2}xd^{2}\bar{\theta}. At leading order in 𝒥\mathcal{J} the kernel KK is given by the leading (super-)diagram in figure 2 and WW is given by (2.10). Plugging in the form of the four-point function (5.11) we find666We are ignoring the subtractions here. They are necessary to compute the eigenvalues kk, but will not contribute to the eigenvalues of the retarded kernel which are our main interest.

KW=d2X3d2X¯4Φ1Φ¯2Φ3Φ¯41|34|3/22h=12|12|1/2d2X3d2X¯41+|χs|+|χs1||34|22h.\int KW=\int d^{2}X_{3}d^{2}\bar{X}_{4}\langle\Phi_{1}\bar{\Phi}_{2}\Phi_{3}\bar{\Phi}_{4}\rangle\frac{1}{|\langle 34\rangle|^{3/2-2h}}=\frac{1}{\sqrt{2}|\langle 12\rangle|^{1/2}}\int d^{2}X_{3}d^{2}\bar{X}_{4}\frac{\sqrt{1+|\chi_{s}|+|\chi_{s}-1|}}{|\langle 34\rangle|^{2-2h}}\;. (5.13)

We will focus on the bosonic part of this integral. Using

12±=x12±,34±=x34±2θ3±θ¯4±,χs±=x12±(x34±2θ3±θ¯4±)x14±x32±=χ±(12θ3±θ¯4±x34±).\begin{split}\langle 12\rangle^{\pm}=x_{12}^{\pm}\;,\qquad&\langle 34\rangle^{\pm}=x_{34}^{\pm}-2\theta_{3}^{\pm}\bar{\theta}^{\pm}_{4}\;,\\ \chi_{s}^{\pm}=\frac{x_{12}^{\pm}(x_{34}^{\pm}-2\theta_{3}^{\pm}\bar{\theta}^{\pm}_{4})}{x_{14}^{\pm}x_{32}^{\pm}}&=\chi^{\pm}(1-\frac{2\theta_{3}^{\pm}\bar{\theta}^{\pm}_{4}}{x_{34}^{\pm}})\;.\end{split} (5.14)

we can perform the Grassman integrals to obtain

k=1|z12|1/2d2x3d2x4(x34)h2(x34+)h242(χ±)k=\frac{1}{|z_{12}|^{1/2}}\int d^{2}x_{3}d^{2}x_{4}\frac{(x_{34}^{-})^{h-2}({x}_{34}^{+})^{h-2}}{4\sqrt{2}}\mathcal{I}(\chi^{\pm}) (5.15)

where

=1(|χ1|+|χ|+1)3/2[|χ|(4h(|χ1|χ++1)2|χ1|+3χ+2)+χ(7|χ|2χ++4)|χ1|+χ(4h(2χ+|χ1|+2χ+|χ||χ1|2|χ|+χ+1)6χ+(|χ1|+|χ|)+4|χ1|)|χ1|4(h1)(|χ1|+|χ|+1)(|χ|χ+(|χ1|+|χ|)χ+14(h1)(|χ1|+|χ|+1))].\begin{split}\mathcal{I}=&\frac{1}{\left(|\chi-1|+|\chi|+1\right)^{3/2}}\left[\frac{|\chi|\left(4h\left(|\chi-1|-\chi^{+}+1\right)-2|\chi-1|+3\chi^{+}-2\right)+\chi^{-}\left(7|\chi|-2\chi^{+}+4\right)}{|\chi-1|}\right.\\ &+\frac{\chi^{-}\left(4h\left(2\chi^{+}|\chi-1|+2\chi^{+}|\chi|-|\chi-1|-2|\chi|+\chi^{+}-1\right)-6\chi^{+}\left(|\chi-1|+|\chi|\right)+4|\chi-1|\right)}{|\chi-1|}\\ &\left.-4(h-1)\left(|\chi-1|+|\chi|+1\right)\left(\frac{|\chi|-\chi^{+}\left(|\chi-1|+|\chi|\right)}{\chi^{+}-1}-4(h-1)\left(|\chi-1|+|\chi|+1\right)\right)\right]\;.\end{split} (5.16)

We can now use analytic continuation to find the eigenvalues of the retarded kernel at finite temperature. More precisely, first we need to go to the cylinder, the do the analytic continuation. The mapping to the cylinder is given by [5]

u=ext,v=ext, (left rail) u=ext,v=ext, (right rail) \begin{array}[]{lll}u=e^{x-t},&v=e^{-x-t},&\text{ (left rail) }\\ u=-e^{x-t},&v=-e^{-x-t},&\text{ (right rail) }\end{array} (5.17)

with the rails identified in figure 4. The analytic continuation is done by taking the following times:

τ1=β/2±ϵ1+it1,τ2=±ϵ2+it2,τ3=β/2+it3,τ4=it4.\tau_{1}=\beta/2\pm\epsilon_{1}+it_{1},\quad\tau_{2}=\pm\epsilon_{2}+it_{2},\quad\tau_{3}=\beta/2+it_{3},\quad\tau_{4}=it_{4}\;. (5.18)

The procedure is then to compute the four-point function for each of the four choices of signs for the ϵ\epsilon’s, and then add them up where each term gets a sign which is positive if both epsilons have the same sign and negative otherwise. This gives the double-commutator

[Φi(β/2+it1),Φj(β/2+it3)][Φ¯i(it2),Φ¯j(it4)]\left\langle[\Phi_{i}\left(\beta/2+it_{1}\right),\Phi_{j}\left(\beta/2+it_{3}\right)][\overline{\Phi}_{i}\left(it_{2}\right),\overline{\Phi}_{j}\left(it_{4}\right)]\right\rangle (5.19)

which can be written as

(Φi(β/2+ϵ1+it1)Φi(β/2ϵ1+it1))(Φ¯i(ϵ2+it2)Φ¯i(ϵ2+it2))Φj(β/2+it3)Φ¯j(it4),\left\langle\left(\Phi_{i}\left(\beta/2+\epsilon_{1}+it_{1}\right)-\Phi_{i}\left(\beta/2-\epsilon_{1}+it_{1}\right)\right)\left(\overline{\Phi}_{i}\left(\epsilon_{2}+it_{2}\right)-\overline{\Phi}_{i}\left(-\epsilon_{2}+it_{2}\right)\right)\Phi_{j}\left(\beta/2+it_{3}\right)\overline{\Phi}_{j}\left(it_{4}\right)\right\rangle\;, (5.20)

where we eventually take the limit ϵi0\epsilon_{i}\to 0. Note that for most combinations of τ\tau’s, the ϵ\epsilon-dependence drops out in this limit. For example, in the combination τ1τ4\tau_{1}-\tau_{4} the real part is always dominated by β\beta and not by the ϵ\epsilon term, and so the sign of ϵ\epsilon does not affect the result when we take ϵ0\epsilon\to 0. As a result, χ±(ϵ)=χ±\chi^{\pm}(\epsilon)=\chi^{\pm} does not depend on ϵ\epsilon:

χ±=sinhx12±iτ122sinhx34±iτ342sinhx14±iτ142sinhx32±iτ322.\chi^{\pm}=\frac{\sinh\frac{x_{12}\pm i\tau_{12}}{2}\sinh\frac{x_{34}\pm i\tau_{34}}{2}}{\sinh\frac{x_{14}\pm i\tau_{14}}{2}\sinh\frac{x_{32}\pm i\tau_{32}}{2}}\;. (5.21)

However, |χ1||\chi-1| does depend on ϵ\epsilon. Following this procedure for |χ1||\chi-1| we find that the analytic continuation takes

|χ1|=(χ1)(χ+1)4|(χ1)(χ+1)|,|\chi-1|=\sqrt{(\chi^{-}-1)(\chi^{+}-1)}\to-4\sqrt{|(\chi^{-}-1)(\chi^{+}-1)|}\;, (5.22)

assuming t13>|x13|t_{13}>\left|x_{13}\right| and t24>|x24|t_{24}>\left|x_{24}\right|, and the expression vanishes otherwise.

Up to overall factors which will not be important, the integral then becomes

kR=(KW)R=𝑑u3𝑑v3𝑑u4𝑑v41u342hv342hR\begin{split}k_{R}=\int(KW)_{R}=\int du_{3}dv_{3}du_{4}dv_{4}\frac{1}{u_{34}^{2-h}v_{34}^{2-h}}\mathcal{I}_{R}\end{split} (5.23)

where R\mathcal{I}_{R} is given by taking \mathcal{I} and subtracting \mathcal{I} but where each |χ1||\chi-1| is replaced by |χ1|-|\chi-1|. In these new variables, we have

χ=u12u34u14u32,χ¯=v12v34v14v32,\chi=\frac{u_{12}u_{34}}{u_{14}u_{32}},\qquad\bar{\chi}=\frac{v_{12}v_{34}}{v_{14}v_{32}}\;, (5.24)

and the integration region is u3>u1,u2>u4,v3>v1,v2>v4u_{3}>u_{1},\;u_{2}>u_{4},\;v_{3}>v_{1},\;v_{2}>v_{4}.

As discussed in 2.2, the chaos exponent as 𝒥0\mathcal{J}\to 0 is found by looking for divergences, which should appear at large |ui|,|vi||u_{i}|,|v_{i}|; the chaos exponent is λ0=2h\lambda_{0}=-2h_{*} where hh_{*} is the value for which the integral diverges. We can find this analytically, since we expect the divergence to come from taking large u3,v3,u4,v4u_{3},v_{3},u_{4},v_{4} (and they are all of the same order of magnitude), which means we are taking χ,χ¯0\chi,\bar{\chi}\to 0. In this limit the integrand behaves as

1u342hv342h\frac{1}{u_{34}^{2-h}v_{34}^{2-h}} (5.25)

which diverges for h0h\geq 0. So the chaos exponent at weak coupling is

λL(𝒥0)=0.\lambda_{L}(\mathcal{J}\to 0)=0\;. (5.26)

5.3.2 Contributions from higher nn-point functions

As discussed in section 2.2, we must make sure that contributions to the kernel from higher nn-point functions don’t diverge at lower values of λ\lambda than λL(𝒥0)=0\lambda_{L}(\mathcal{J}\to 0)=0, otherwise the approximations made above are invalid. We know all of the higher nn-point functions since we have mapped the theory to a free theory, and so it is a matter of plugging the results into the kernel.777As in [1], we can ignore the contribution of subtractions in this calculation.

As an example, we consider the contribution from the bottom component of the 2n2n-point function for any nn to the kernel in figure 2. Using the free field realization, we find that any 2n2n-point function of the bottom component ϕ\phi of the superfield Φ\Phi takes the form

ϕ(x1)ϕ(xn)ϕ¯(y1)ϕ¯(yn)=12n/2ϵix=±1,ϵiy=±1,ϵix+ϵiy=0i<j|xij|(ϵixϵjx+1)/2|yij|(ϵiyϵjy+1)/2i,j|xiyj|(1ϵixϵjy)/2.\langle\phi(x_{1})...\phi(x_{n})\bar{\phi}(y_{1})...\bar{\phi}(y_{n})\rangle=\frac{1}{2^{n/2}}\sqrt{\sum_{\begin{subarray}{c}\epsilon_{i}^{x}=\pm 1,\;\;\epsilon_{i}^{y}=\pm 1,\\ \sum\epsilon_{i}^{x}+\epsilon_{i}^{y}=0\end{subarray}}\frac{\prod_{i<j}|x_{ij}|^{\left(\epsilon_{i}^{x}\epsilon_{j}^{x}+1\right)/2}|y_{ij}|^{\left(\epsilon_{i}^{y}\epsilon_{j}^{y}+1\right)/2}}{\prod_{i,j}|x_{i}-y_{j}|^{\left(1-\epsilon_{i}^{x}\epsilon_{j}^{y}\right)/2}}}\;. (5.27)

Combined with the propagators (5.7), we can find the contributions to the kernel at any order and look for divergences as x3+,x4+x^{+}_{3}\to\infty,x^{+}_{4}\to\-\infty and similarly for x3,4x_{3,4}^{-}. In this limit, the contribution from the six point function behaves as 1|x34|42h\frac{1}{|x_{34}|^{4-2h}}, which matches the behavior found in the contribution from the four-point function. The divergence is thus also at h=0h=0, and so our approximation is consistent at this level. One can repeat the analysis for other components of the 2n-point function which contribute to kRk_{R}, and as a result the approximations discussed above are consistent.

5.3.3 Continuity conjecture

We now compute λL0\lambda_{L}^{0} and compare to λL(𝒥0)\lambda_{L}(\mathcal{J}\to 0) in order to show that the continuity conjecture is obeyed. This amounts to taking the bosonic part of the analytically-continued four-point function of a single core CFT (5.11) and studying its behavior in the large-time limit t3t4=tt_{3}\sim t_{4}=t\to\infty.

First we must do the analytic continuation discussed above. The bosonic part of the 4-pt function is

121+|χ|+|χ1||z12z34|12.\frac{1}{\sqrt{2}}\frac{\sqrt{1+|\chi|+|\chi-1|}}{|z_{12}z_{34}|^{\frac{1}{2}}}\;. (5.28)

Mapping to the cylinder and performing the analytic continuation discussed above, we find

1+|χ|+|χ1||z12z34|122θ(t13|x13|)θ(t24|x24|)1+|χ|+|χ1|1+|χ||χ1||z12z34|12.\frac{\sqrt{1+|\chi|+|\chi-1|}}{|z_{12}z_{34}|^{\frac{1}{2}}}\to 2\theta(t_{13}-|x_{13}|)\theta(t_{24}-|x_{24}|)\frac{\sqrt{1+|\chi|+|\chi-1|}-\sqrt{1+|\chi|-|\chi-1|}}{|z_{12}z_{34}|^{\frac{1}{2}}}\;. (5.29)

Now we can take the large-time limit, which amounts to taking χ,χ¯et\chi,\bar{\chi}\sim e^{-t} for tt large. We find

1+|χ|+|χ1|1+|χ||χ1|22et/2+\sqrt{1+|\chi|+|\chi-1|}-\sqrt{1+|\chi|-|\chi-1|}\approx\sqrt{2}-\sqrt{2}e^{-t/2}+... (5.30)

and so the analytically-continued four-point function is finite in the large-tt limit, since it is not exponentially vanishing in tt. As a result, λL0=0\lambda_{L}^{0}=0 since there is no exponential behavior. We thus find that λL0=λ(𝒥0)=0\lambda_{L}^{0}=\lambda(\mathcal{J}\to 0)=0 and so the continuity conjecture is obeyed.

6 Conclusions

In this paper we have proven the continuity conjecture in QM and given additional evidence for it in 2d2d. We also studied the disordered 𝒜3\mathcal{A}_{3} minimal model at leading order in the random coupling 𝒥\mathcal{J}.

There are many interesting questions left for future work. For example, in both of the 2d2d examples given here (the chiral SYK model and the disordered 𝒜3\mathcal{A}_{3} model), the chaos exponent at small coupling was λL0=0\lambda_{L}^{0}=0. In [1], the example of the disordered 𝒜2\mathcal{A}_{2} minimal model was also shown to have λL0=0\lambda_{L}^{0}=0. These examples should be compared to the case of disordered generalized free fields where λL0<0\lambda_{L}^{0}<0, leading to a discontinuous transition into chaos as 𝒥\mathcal{J} increases [1].888This behavior is not related to the non-locality of the generalized free field theory; for example, a negative λL0\lambda_{L}^{0} was found in [25], although this corresponds to the chaos exponent on thermal Rindler space and not flat space. We can ask why this interesting behavior did not occur here, and an obvious conjecture is that λL0\lambda_{L}^{0} can only be non-zero when the core CFT does not have a free theory realization. Unfortunately, this means that performing computations in theories with λL0<0\lambda_{L}^{0}<0 will be difficult, since the computations require knowledge of exact nn-point functions of the CFT. One possible direction would be to perform these computations in 𝒩=2\mathcal{N}=2 minimal models with higher central charges, where the four-point functions are known exactly, but performing the necessary integrals over them is a very complicated technical task which we leave to future work.

Another future direction would be to prove the continuity conjecture in 2d2d. This becomes more complicated due to the appearance of the butterfly velocity. The divergence which leads to the chaos exponent at weak coupling appears at large times, but can appear at any value of the ratio v=t/xv=t/x, and an additional integral must be performed over these values. The final result depends nontrivially on vv, and the result must be studied carefully in order for the chaos exponent to be read off. However, the chiral example discussed above is evidence that even with a nontrivial butterfly velocity, the continuity conjecture is obeyed. It is yet to be seen whether it is obeyed only for the velocity which leads to the maximal chaos exponent, or whether it is obeyed for any velocity.

It would also be interesting to extend calculations of λL(𝒥)\lambda_{L}(\mathcal{J}) for a continuous range of 𝒥\mathcal{J} to higher dimensions, building on [9]. For example, considering NN free bosons in 3d3d, the deformation ((φi)2)3((\varphi_{i})^{2})^{3} is exactly marginal at leading order in 1/N1/N [26]. Similarly for NN free matter multiplets in a 3d3d 𝒩=1\mathcal{N}=1 supersymmetric theory, the superpotential deformation (|Φi|2)2(|\Phi_{i}|^{2})^{2} is also exactly marginal at leading order in 1/N1/N [27]. One could perform a computation of the chaos exponent analogous to that of the O(N)O(N) model [28] as a function of the exactly marginal deformation.999A better understanding of whether it is enough to be exactly marginal at leading order in 1/N1/N is required in order to perform these computations. The chaos exponent is expected to be non-positive since the theory is close to being free, and it would be interesting to see if it is exactly zero.

Acknowledgements

The authors would like to thank E. Y. Urbach and N. Silberstein for collaborating on this work in its early stages, and B. Lian for collaboration on related topics. The authors would also like to thank Micha Berkooz and Doron Gepner for enlightening discussions. The work of RRK was supported in part by an Israel Science Foundation (ISF) center for excellence grant (grant number 2289/18), by ISF grant no. 2159/22, by Simons Foundation grant 994296 (Simons Collaboration on Confinement and QCD Strings), by grant no. 2018068 from the United States-Israel Binational Science Foundation (BSF), by the Minerva foundation with funding from the Federal German Ministry for Education and Research, by the German Research Foundation through a German-Israeli Project Cooperation (DIP) grant “Holography and the Swampland”, and by a research grant from Martin Eisenstein.

Appendix A Superappendix

A.1 The supersymmetry algebra

𝒩=(2,2)\mathcal{N}=(2,2) SUSY in two dimensions is closely related to 𝒩=1\mathcal{N}=1 in four dimensions. There are two real spinors’ worth of supercharges: Qα(a)Q^{(a)}_{\alpha}, where α\alpha is a Lorentz (spinor) index and a=1,,𝒩a=1,...,\mathcal{N} is a separate (R-symmetry) index. The SUSY algebra is

{Qαa,Qβb}\displaystyle\{Q^{a}_{\alpha},Q^{b}_{\beta}\} =2γαβμPμδab,\displaystyle=2\gamma^{\mu}_{\alpha\beta}P_{\mu}\delta^{ab}\;, (A.1)

and the gamma matrices are:

γαβ0=[1001],γαβ1=[1001].\displaystyle\gamma^{0}_{\alpha\beta}=\begin{bmatrix}1&0\\ 0&-1\end{bmatrix}\;,\qquad\gamma^{1}_{\alpha\beta}=\begin{bmatrix}1&0\\ 0&1\end{bmatrix}\;.

We can break the spinor representation into the left/right-moving sectors, which are one-dimensional and transform by a phase under the Lorentz group:

Q±a\displaystyle Q^{a}_{\pm} e±12ηQ±a\displaystyle\rightarrow e^{\pm\frac{1}{2}\eta}Q^{a}_{\pm} (A.2)

In Lorentzian signature, the supercharges are real:

Q±a¯\displaystyle\overline{Q^{a}_{\pm}} =(Q±a)=Q±a\displaystyle=\left(Q^{a}_{\pm}\right)^{*}=Q^{a}_{\pm} (A.3)

A.1.1 𝒩=(1,1)\mathcal{N}=(1,1)

A real 𝒩=(1,1)\mathcal{N}=(1,1) scalar superfield 𝒜(x,θ)\mathcal{A}(x,\theta) consists of a real scalar ϕ(x)\phi(x) and its superpartners: a Majorana spinor ψα(x)\psi_{\alpha}(x), and another scalar F(x)F(x).

𝒜(x,θ)\displaystyle\mathcal{A}(x,\theta) =eθαQαϕ(x)\displaystyle=e^{\theta^{\alpha}Q_{\alpha}}\phi(x) (A.4)
=ϕ(x)+θαψα(x)+θαθβϵαβF(x)+\displaystyle=\phi(x)+\theta^{\alpha}\psi_{\alpha}(x)+\theta^{\alpha}\theta^{\beta}\epsilon_{\alpha\beta}F(x)+\dots
=ϕ(x)+θ+ψ+(x)+θψ(x)+θ+θF(x)+\displaystyle=\phi(x)+\theta^{+}\psi_{+}(x)+\theta^{-}\psi_{-}(x)+\theta^{+}\theta^{-}F(x)+\dots (A.5)

where θα\theta^{\alpha} is a real Grassman spinor,101010(θ±)2=0\left(\theta^{\pm}\right)^{2}=0. and is given a Lorentz transformation so that the superfield is a scalar. We’ve broken up ψα(x)\psi_{\alpha}(x) into Lorentz irreps, and the equations of motion will tell us that ψ±\psi_{\pm} are left/right moving (holomorphic/antiholomorphic in Euclidean signature). Higher components are descendants since the supercharge squares to a derivative. There is a discrete R-symmetry

Qα\displaystyle Q_{\alpha} Qα,\displaystyle\rightarrow-Q_{\alpha}\;, (A.6)

that leaves the SUSY algebra invariant. This gives the following transformations for the fields:

ϕ(x)\displaystyle\phi(x) (1)γϕ(x),\displaystyle\rightarrow(-1)^{\gamma}\phi(x)\;,
ψα(x)\displaystyle\psi_{\alpha}(x) (1)γ+1ψα(x),\displaystyle\rightarrow(-1)^{\gamma+1}\psi_{\alpha}(x)\;,
F(x)\displaystyle F(x) (1)γ+2F(x).\displaystyle\rightarrow(-1)^{\gamma+2}F(x)\;. (A.7)

With θαθα\theta^{\alpha}\rightarrow-\theta^{\alpha}, the superfield 𝒜(x,θ)\mathcal{A}(x,\theta) has the same R-charge as its bottom component (here, the scalar; see (A.4)). The Lagrangian (free-field plus superpotential) in terms of superfields may be written as

\displaystyle\mathcal{L} =d2θ(12𝒟𝒜𝒟¯𝒜+W(𝒜)).\displaystyle=\int d^{2}\theta(\frac{1}{2}\mathcal{D}\mathcal{A}\bar{\mathcal{D}}\mathcal{A}+W(\mathcal{A}))\;. (A.8)

A.1.2 𝒩=(2,2)\mathcal{N}=(2,2)

The 𝒩=(2,2)\mathcal{N}=(2,2) algebra has two Majorana supercharges, and we need two real Grassman spinors θaα\theta^{\alpha}_{a}. A complex superfield 𝒳\mathcal{X} takes the form

𝒳(x,θa)\displaystyle\mathcal{X}(x,\theta^{a}) =eθaαQαaϕ(x)=ϕ+θaαψαa+\displaystyle=e^{\theta_{a}^{\alpha}Q^{a}_{\alpha}}\phi(x)=\phi+\theta_{a}^{\alpha}\psi^{a}_{\alpha}+...
=ϕ(x)+θa+ψ+a(x)+θaψa(x)+\displaystyle=\phi(x)+\theta^{+}_{a}\psi^{a}_{+}(x)+\theta^{-}_{a}\psi^{a}_{-}(x)+...
𝒳¯(x,θa)\displaystyle\bar{\mathcal{X}}(x,\theta^{a}) =ϕ¯+θaαψ¯αa+\displaystyle=\bar{\phi}+\theta_{a}^{\alpha}\bar{\psi}^{a}_{\alpha}+...
=ϕ¯(x)+θa+ψ¯+a(x)+θaψ¯a(x)+\displaystyle=\bar{\phi}(x)+\theta^{+}_{a}\bar{\psi}^{a}_{+}(x)+\theta^{-}_{a}\bar{\psi}^{a}_{-}(x)+... (A.9)

The bar stands for complex conjugation (see (A.3)). It is convenient to use complex combinations of the supercharges:

Q±=Q±1+iQ±2,Q¯±=Q±1iQ±2,(Q±)¯=Q¯±.\displaystyle Q_{\pm}=Q^{1}_{\pm}+iQ^{2}_{\pm}\;,\qquad\overline{Q}_{\pm}=Q^{1}_{\pm}-iQ^{2}_{\pm}\;,\qquad\overline{(Q_{\pm})}=\overline{Q}_{\pm}\;. (A.10)

The 𝒩=(2,2)\mathcal{N}=(2,2) SUSY algebra may be written as:

{Q±,Q¯±}\displaystyle\{Q_{\pm},\overline{Q}_{\pm}\} =2P±=2(±P0+P1).\displaystyle=2P_{\pm}=2(\pm P_{0}+P_{1})\;. (A.11)

The Grassman variables are now:

θ±=θ1±+iθ2±,θ¯±=θ1±iθ2±.\displaystyle\theta^{\pm}=\theta_{1}^{\pm}+i\theta_{2}^{\pm}\;,\qquad\overline{\theta}^{\pm}=\theta_{1}^{\pm}-i\theta_{2}^{\pm}\;. (A.12)

The superfield takes the form

𝒳(x,θ,θ¯)\displaystyle\mathcal{X}(x,\theta,\bar{\theta}) =eθ¯αQα+θαQ¯αϕ(x)\displaystyle=e^{\bar{\theta}^{\alpha}Q_{\alpha}+\theta^{\alpha}\bar{Q}_{\alpha}}\phi(x) (A.13)
=ϕ(x)+θ¯+ψ~+(x)+θ+ψ+(x)+θ¯ψ~(x)+θψ(x)+\displaystyle=\phi(x)+\bar{\theta}^{+}\tilde{\psi}_{+}(x)+\theta^{+}\psi_{+}(x)+\bar{\theta}^{-}\tilde{\psi}_{-}(x)+\theta^{-}\psi_{-}(x)+...
𝒳¯(x,θ,θ¯)\displaystyle\bar{\mathcal{X}}(x,\theta,\bar{\theta}) =ϕ¯(x)+θ¯+ψ¯+(x)+θ+ψ~¯+(x)+θ¯ψ¯(x)+θψ~¯(x)+\displaystyle=\bar{\phi}(x)+\bar{\theta}^{+}\bar{\psi}_{+}(x)+\theta^{+}\bar{\tilde{\psi}}_{+}(x)+\bar{\theta}^{-}\bar{\psi}_{-}(x)+\theta^{-}\bar{\tilde{\psi}}(x)+... (A.14)

with

ψ±=ψ±1iψ±2,ψ~±=ψ±1+iψ±2.\displaystyle\psi_{\pm}=\psi_{\pm}^{1}-i\psi_{\pm}^{2}\;,\qquad\tilde{\psi}_{\pm}=\psi_{\pm}^{1}+i\psi_{\pm}^{2}\;. (A.15)

The 𝔲(1)R\mathfrak{u}(1)_{R} symmetry rotates (Q,Q¯)(Q,\bar{Q}) in opposite directions:

Q±eiαQ±,Q¯±eiαQ¯±,\displaystyle Q_{\pm}\rightarrow e^{i\alpha}Q_{\pm}\;,\qquad\bar{Q}_{\pm}\rightarrow e^{-i\alpha}\bar{Q}_{\pm}\;,
θ±eiαθ±,θ¯±eiαθ¯±.\displaystyle\theta^{\pm}\rightarrow e^{i\alpha}\theta^{\pm}\;,\qquad\bar{\theta}^{\pm}\rightarrow e^{-i\alpha}\bar{\theta}^{\pm}\;. (A.16)

There is also an axial 𝔲(1)R~\mathfrak{u}(1)_{\tilde{R}} which rotates the ±\pm components with opposite phases.

Chirality

The superfield (A.13) is in a reducible representation. Imposing

Q(ϕ(x))\displaystyle Q(\phi(x)) =Q¯(ϕ¯(x))=0,\displaystyle=\bar{Q}(\bar{\phi}(x))=0\;,

we have

Φ(x,θ,θ¯)\displaystyle\Phi(x,\theta,\bar{\theta}) =e(θ¯αQα+θαQ¯α)ϕ(x)=e(θαQ¯α+iθαγμθ¯Pμ)eθ¯αQαϕ(x)\displaystyle=e^{(\bar{\theta}^{\alpha}Q_{\alpha}+\theta^{\alpha}\bar{Q}_{\alpha})}\phi(x)=e^{(\theta^{\alpha}\bar{Q}_{\alpha}+i\theta^{\alpha}\gamma^{\mu}\bar{\theta}P_{\mu})}\cancel{e^{\bar{\theta}^{\alpha}Q_{\alpha}}}\phi(x)
=ϕ(x)+θ+ψ+(x)+θψ(x)+θ+θF(x)+\displaystyle=\phi(x)+\theta^{+}\psi_{+}(x)+\theta^{-}\psi_{-}(x)+\theta^{+}\theta^{-}F(x)+...
Φ¯(x,θ,θ¯)\displaystyle\bar{\Phi}(x,\theta,\bar{\theta}) =ϕ¯(x)+θ¯+ψ¯+(x)+θ¯ψ¯(x)+θ¯+θ¯F¯(x)+\displaystyle=\bar{\phi}(x)+\bar{\theta}^{+}\bar{\psi}_{+}(x)+\bar{\theta}^{-}\bar{\psi}_{-}(x)+\bar{\theta}^{+}\bar{\theta}^{-}\bar{F}(x)+... (A.17)

This makes Φ(x,θ,θ¯)\Phi(x,\theta,\bar{\theta}) a chiral superfield. An 𝒩=(2,2)\mathcal{N}=(2,2) chiral (scalar) multiplet contains only one complex fermion. It is easy to write down interactions for such superfields:

\displaystyle\mathcal{L} =d2θd2θ¯Φ¯Φ+d2θW(Φ)+d2θ¯W¯(θ¯).\displaystyle=\int d^{2}\theta d^{2}\bar{\theta}\bar{\Phi}\Phi+\int d^{2}\theta W(\Phi)+\int d^{2}\bar{\theta}\bar{W}(\bar{\theta})\;. (A.18)

Of course, these are not the only possible interactions.

A.2 The 𝒜3\mathcal{A}_{3} minimal model

We are interested in the following 𝒩=(2,2)\mathcal{N}=(2,2) theory:

\displaystyle\mathcal{L} =d2θd2θ¯X¯X+d2θX4+d2θ¯X¯4,\displaystyle=\int d^{2}\theta d^{2}\bar{\theta}\bar{X}X+\int d^{2}\theta X^{4}+\int d^{2}\bar{\theta}\bar{X}^{4}\;, (A.19)

with XX a chiral superfield.

Table 2: Operators from the Ising model. The last entry χ\chi is the fermion. σ\sigma and μ\mu are the spin and disorder operators; these aren’t mutually local.
𝒪\mathcal{O} Δ\Delta \ell
σ\sigma 18\frac{1}{8} 0
μ\mu 18\frac{1}{8} 0
ϵ\epsilon 11 0
χ±\chi_{\pm} 12\tfrac{1}{2} 12\tfrac{1}{2}

It’s easily seen that the interactions preserve 𝔲(1)R\mathfrak{u}(1)_{R} with qX=1/2q_{X}=1/2; this blocks the superpotential from picking up other terms under RG flow. The theory in the IR is a Landau-Ginsburg minimal model with central charge c=32c=\tfrac{3}{2} that may also be obtained by putting together a free scalar and the Ising model. In the following, HH is a free boson, and Table 2 are the operators we need from the Ising model.

Firstly, we identify the SUSY generators111111Recall that the SUSY currents are operators with dimensions (32,0)(\tfrac{3}{2},0) or (0,32)(0,\tfrac{3}{2}) (G±G_{\pm} respectively). and R symmetry current:

G±\displaystyle G_{\pm} =χ±exp(i2H±),\displaystyle=\chi_{\pm}\exp\left(i\sqrt{2}H_{\pm}\right)\;,
G¯±\displaystyle\bar{G}_{\pm} =χ±exp(i2H±),\displaystyle=\chi_{\pm}\exp\left(-i\sqrt{2}H_{\pm}\right)\;,
j±(R)\displaystyle j_{\pm}^{(R)} =i2±H.\displaystyle=\frac{i}{\sqrt{2}}\partial_{\pm}H\;. (A.20)

H±H_{\pm} are the left/right moving parts of the scalar field HH. The bottom components of the superconformal primaries are given by:

ϕ=σexp(iH22),\displaystyle\phi=\sigma\exp\left(i\frac{H}{2\sqrt{2}}\right)\;, ϕ¯=σexp(iH22),\displaystyle\qquad\bar{\phi}=\sigma\exp\left(-i\frac{H}{2\sqrt{2}}\right)\;,
ϕ2=exp(iH2),\displaystyle\phi^{2}=\exp\left(i\frac{H}{\sqrt{2}}\right)\;, ϕ¯2=exp(iH2),\displaystyle\qquad\bar{\phi}^{2}=\exp\left(-i\frac{H}{\sqrt{2}}\right)\;,
ϕ¯ϕ\displaystyle\bar{\phi}\phi =ϵ.\displaystyle=\epsilon\;. (A.21)

We can then find the free field expressions for their SUSY descendants using (A.20) and the OPE’s:

χ±(x±)×σ(x)1x±μ,χ±(x±)×μ(x)1x±σ.\displaystyle\chi_{\pm}(x^{\pm})\times\sigma(x)\sim\frac{1}{\sqrt{x^{\pm}}}\mu\;,\qquad\chi_{\pm}(x^{\pm})\times\mu(x)\sim\frac{1}{\sqrt{x^{\pm}}}\sigma\;. (A.22)

This leads to:

G¯+(x+)×ϕ\displaystyle\bar{G}_{+}(x^{+})\times\phi 1x+μexp(i22(3H++H))=1x+ψ+,\displaystyle\sim\frac{1}{x^{+}}\ \mu\exp\left(\frac{i}{2\sqrt{2}}\left(-3H_{+}+H_{-}\right)\right)=\frac{1}{x^{+}}\psi_{+}\;,
G¯(x)×ϕ\displaystyle\bar{G}_{-}(x^{-})\times\phi 1xμexp(i22(H+3H))=1xψ,\displaystyle\sim\frac{1}{x^{-}}\ \mu\exp\left(\frac{i}{2\sqrt{2}}\left(H_{+}-3H_{-}\right)\right)=\frac{1}{x^{-}}\psi_{-}\;,
G¯+(x+)×ψ\displaystyle\bar{G}_{+}(x^{+})\times\psi_{-} 1x+σexp(i3H22)=1x+F,\displaystyle\sim\frac{1}{x^{+}}\ \sigma\exp\left(-i\frac{3H}{2\sqrt{2}}\right)=\frac{1}{x^{+}}F\;,
G¯(x)×ψ+\displaystyle\bar{G}_{-}(x^{-})\times\psi_{+} 1xσexp(i3H22)=1xF.\displaystyle\sim\frac{1}{x^{-}}\ \sigma\exp\left(-i\frac{3H}{2\sqrt{2}}\right)=\frac{1}{x^{-}}F\;. (A.23)

We summarise these results in Table 3.

Table 3: The operators in the multiplet of XX in the 𝒜3\mathcal{A}_{3} minimal model.
𝒪\mathcal{O} 𝒪free\mathcal{O}_{\text{free}} (h,h¯)(h,\bar{h}) qRq_{R}
ϕ\phi σeiH22\sigma\ e^{i\frac{H}{2\sqrt{2}}} (18,18)\left(\frac{1}{8},\frac{1}{8}\right) (14,14)\left(\frac{1}{4},\frac{1}{4}\right)
ψ+\psi_{+} μei3H++H22\mu\ e^{i\frac{-3H_{+}+H_{-}}{2\sqrt{2}}} (58,18)\left(\frac{5}{8},\frac{1}{8}\right) (34,14)\left(-\frac{3}{4},\frac{1}{4}\right)
ψ\psi_{-} μeiH+3H22\mu\ e^{i\frac{H_{+}-3H_{-}}{2\sqrt{2}}} (18,58)\left(\frac{1}{8},\frac{5}{8}\right) (14,34)\left(\frac{1}{4},-\frac{3}{4}\right)
FF σei3H22\sigma\ e^{-i\frac{3H}{2\sqrt{2}}} (58,58)\left(\frac{5}{8},\frac{5}{8}\right) (34,34)\left(-\frac{3}{4},-\frac{3}{4}\right)
Mutual locality of the operators

Firstly, GG, G¯\bar{G}, and jRj^{R} are mutually local (of course, the fermionic operators are mutually local only in the fermionic sense – they pick up a 1-1) since the bosonic pieces are the conserved currents of the 𝔰𝔲(2)1\mathfrak{su}(2)_{1} model. Now, all the operators within the XX and X¯\bar{X} multiplet are obtained by operator products with the currents. So to show that these are mutually local, we just need to show that the bottom components don’t see branch cuts with respect to each other or the currents.

The OPE of two vertex operators (analytically continued to Euclidean space) take the form:

eik1.H(z)×eik2.H(0)\displaystyle e^{ik_{1}.H}(z)\times e^{ik_{2}.H}(0) zk1Rk2Rz¯k1Lk2Lei(k1+k2).H(0)\displaystyle\sim z^{k_{1}^{R}k_{2}^{R}}\bar{z}^{k_{1}^{L}k_{2}^{L}}e^{i(k_{1}+k_{2}).H}(0)
=(zz¯)12k1.k2(zz¯)12(k1Lk2Lk1Rk2R)ei(k1+k2).H(0)\displaystyle=(z\bar{z})^{\tfrac{1}{2}k_{1}.k_{2}}\left(\frac{z}{\bar{z}}\right)^{-\tfrac{1}{2}(k_{1}^{L}k_{2}^{L}-k_{1}^{R}k_{2}^{R})}e^{i(k_{1}+k_{2}).H}(0) (A.24)

where k1,2.H=k1,2RH++k1,2LHk_{1,2}.H=k_{1,2}^{R}H_{+}+k_{1,2}^{L}H_{-} and k1.k2=k1Rk2R+k1Lk2Lk_{1}.k_{2}=k_{1}^{R}k_{2}^{R}+k_{1}^{L}k_{2}^{L}. When one operator goes around the other, the right-hand side picks up e2πi(k1Lk2Lk1Rk2R)e^{2\pi i(k_{1}^{L}k_{2}^{L}-k_{1}^{R}k_{2}^{R})}. Setting this phase to 1 gives the Narain condition:

k1k2=k1Lk2Lk1Rk2R\displaystyle k_{1}\odot k_{2}=k_{1}^{L}k_{2}^{L}-k_{1}^{R}k_{2}^{R}\in\mathbbm{Z} (A.25)

This will be used to check locality between vertex operators. The R-current plays well with all vertex operators, so we won’t have to bother with it.

Coming back to our operators, the Ising parts of ϕ\phi and ϕ¯\bar{\phi} are mutually local. So we just need to examine the vertex operators, which are easily seen to satisfy (A.25). Actually it quite easy to see that all the bottom components listed in (A.21) are mutually local. The first nontrivial check is between (G,G¯)(G,\bar{G}) and ϕ\phi. But we can easily verify from (A.23) that there are no branch cuts; the same conclusion holds for (G,G¯)(G,\bar{G}) and ϕ¯\bar{\phi} via the conjugate of (A.23). ϕ¯ϕ\bar{\phi}\phi and the currents are also easily kosher. To work out ϕ2\phi^{2} case, we need only focus on the vertex operators. The only non-zero \odot products are (±2,0)(12,0)=±1(\pm\sqrt{2},0)\odot(\tfrac{1}{\sqrt{2}},0)=\pm 1.

Appendix B Simplification of 1d1d integral

Using the iterative ladder structure, we can write the full retarded kernel as

KR(1,2)=K0(1,2;3,4)Glrq2(3,4).K_{R}(1,2)=K_{0}(1,2;3,4)G_{lr}^{q-2}(3,4)\;. (B.1)

At leading order, K0K_{0} is just the subtracted 4-pt function. It is more convenient to work with the rescaled expression

KR=eΔ(t1+t2t3t4)KR.K_{R}^{\prime}=e^{\Delta\left(t_{1}+t_{2}-t_{3}-t_{4}\right)}K_{R}\;. (B.2)

Next, we perform the change of variables z=etz=e^{-t} on the left rail (points 1,3) and z=etz=-e^{-t} on the right (points 2,4), see figure 4, leading to

KRdt3dt4=KRdz3z3dz4z4.K_{R}^{\prime}dt_{3}dt_{4}=K_{R}^{\prime}\frac{dz_{3}}{z_{3}}\frac{dz_{4}}{z_{4}}\;. (B.3)

The two-points are fixed up to a constant that we will ignore:

Glr=1(2cosht3t42)2Δ=z3Δz4Δ(z34)2Δ,GR(1,3)=θ(t1t3)2cos(πΔ)b(2sinht1t32)2Δ=θ(z3z1)2cos(πΔ)bz1Δz3Δ(z13)2Δ.\begin{split}G_{lr}&=\frac{1}{\left(2\cosh\frac{t_{3}-t_{4}}{2}\right)^{2\Delta}}=\frac{z_{3}^{\Delta}z_{4}^{\Delta}}{(z_{34})^{2\Delta}}\;,\\ G_{R}\left(1,3\right)&=\theta\left(t_{1}-t_{3}\right)\frac{2\cos(\pi\Delta)b}{\left(2\sinh\frac{t_{1}-t_{3}}{2}\right)^{2\Delta}}=\theta(z_{3}-z_{1})\frac{2\cos(\pi\Delta)bz_{1}^{\Delta}z_{3}^{\Delta}}{(z_{13})^{2\Delta}}\;.\end{split} (B.4)

Now we use the fact that K0K_{0} is a (sum of) four-point functions, so we can write it as

K0=eΔ(t1+t2t3t4)Glr(1,2)Glr(3,4)𝒢(χ)K_{0}^{\prime}=e^{\Delta\left(t_{1}+t_{2}-t_{3}-t_{4}\right)}G_{lr}(1,2)G_{lr}(3,4)\mathcal{G}(\chi) (B.5)

with χ\chi the conformal cross-ratio and GG some function. Putting this together and setting Δ=1/q\Delta=1/q we find

KRdt3dt4=𝒢R(χ)dz3dz4(z12)2Δ(z34)2Δ(q1)K_{R}^{\prime}dt_{3}dt_{4}=\frac{\mathcal{G}_{R}(\chi)dz_{3}dz_{4}}{(z_{12})^{2\Delta}(z_{34})^{2\Delta\left(q-1\right)}} (B.6)

Finally, applying KRK_{R}^{\prime} to the eigenfunction W+=1|z34|2Δ+λW_{+}^{\prime}=\frac{1}{|z_{34}|^{2\Delta+\lambda}} we find

KRW+dt3dt4=𝒢R(χ)dz3dz4|z12|2Δ|z34|2+λ.K_{R}^{\prime}W_{+}^{\prime}dt_{3}dt_{4}=\frac{\mathcal{G}_{R}(\chi)dz_{3}dz_{4}}{|z_{12}|^{2\Delta}|z_{34}|^{2+\lambda}}\;. (B.7)

References

  • [1] Micha Berkooz, Adar Sharon, Navot Silberstein and Erez Y. Urbach “Onset of quantum chaos in disordered CFTs” In Phys. Rev. D 106.4, 2022, pp. 045007 DOI: 10.1103/PhysRevD.106.045007
  • [2] Juan Maldacena, Stephen H. Shenker and Douglas Stanford “A bound on chaos” In JHEP 08, 2016, pp. 106 DOI: 10.1007/JHEP08(2016)106
  • [3] Subir Sachdev and Jinwu Ye “Gapless spin-fluid ground state in a random quantum Heisenberg magnet” In Phys. Rev. Lett. 70 American Physical Society, 1993, pp. 3339–3342 DOI: 10.1103/PhysRevLett.70.3339
  • [4] A Kitaev “A simple model of quantum holography” Talks at KITP, 2015 URL: https://online.kitp.ucsb.edu/online/entangled15/kitaev/
  • [5] Jeff Murugan, Douglas Stanford and Edward Witten “More on Supersymmetric and 2d Analogs of the SYK Model” In JHEP 08, 2017, pp. 146 DOI: 10.1007/JHEP08(2017)146
  • [6] Ksenia Bulycheva 𝒩=2\mathcal{N}=2 SYK model in the superspace formalism” In JHEP 04, 2018, pp. 036 DOI: 10.1007/JHEP04(2018)036
  • [7] Wenbo Fu, Davide Gaiotto, Juan Maldacena and Subir Sachdev “Supersymmetric Sachdev-Ye-Kitaev models” [Addendum: Phys.Rev.D 95, 069904 (2017)] In Phys. Rev. D 95.2, 2017, pp. 026009 DOI: 10.1103/PhysRevD.95.026009
  • [8] Cheng Peng 𝒩=(0,2)\mathcal{N}=(0,2) SYK, Chaos and Higher-Spins” In JHEP 12, 2018, pp. 065 DOI: 10.1007/JHEP12(2018)065
  • [9] Chi-Ming Chang, Sean Colin-Ellerin, Cheng Peng and Mukund Rangamani “A 3d disordered superconformal fixed point” In JHEP 11, 2021, pp. 211 DOI: 10.1007/JHEP11(2021)211
  • [10] Chi-Ming Chang, Sean Colin-Ellerin, Cheng Peng and Mukund Rangamani “Disordered Vector Models: From Higher Spins to Incipient Strings” In Phys. Rev. Lett. 129.1, 2022, pp. 011603 DOI: 10.1103/PhysRevLett.129.011603
  • [11] David J. Gross and Vladimir Rosenhaus “A line of CFTs: from generalized free fields to SYK” In JHEP 07, 2017, pp. 086 DOI: 10.1007/JHEP07(2017)086
  • [12] Biao Lian, S. L. Sondhi and Zhenbin Yang “The chiral SYK model” In JHEP 09, 2019, pp. 067 DOI: 10.1007/JHEP09(2019)067
  • [13] Fedor K. Popov “Supersymmetric tensor model at large NN and small ϵ\epsilon In Phys. Rev. D 101.2, 2020, pp. 026020 DOI: 10.1103/PhysRevD.101.026020
  • [14] Juan Maldacena and Douglas Stanford “Remarks on the Sachdev-Ye-Kitaev model” In Phys. Rev. D 94.10, 2016, pp. 106002 DOI: 10.1103/PhysRevD.94.106002
  • [15] Alexei Kitaev and S. Josephine Suh “The soft mode in the Sachdev-Ye-Kitaev model and its gravity dual” In JHEP 05, 2018, pp. 183 DOI: 10.1007/JHEP05(2018)183
  • [16] Micha Berkooz, Adar Sharon, Navot Silberstein and Erez Y. Urbach “Onset of Quantum Chaos in Random Field Theories” In Phys. Rev. Lett. 129.7, 2022, pp. 071601 DOI: 10.1103/PhysRevLett.129.071601
  • [17] A Kitaev “Hidden correlations in the Hawking radiation and thermal noise” Talks at KITP, 2015 URL: https://online.kitp.ucsb.edu/online/joint98/kitaev/
  • [18] Anatoly I. Larkin and Yu. N. Ovchinnikov “Quasiclassical Method in the Theory of Superconductivity” In Journal of Experimental and Theoretical Physics, 1969
  • [19] Yichen Hu and Biao Lian “The Chiral Sachdev-Ye Model: Integrability and Chaos of Anyons in 1+1d”, 2021 arXiv:2109.13263 [cond-mat.str-el]
  • [20] G. Mussardo, G. Sotkov and M. Stanishkov “N=2 superconformal minimal models” In Int. J. Mod. Phys. A 4, 1989, pp. 1135 DOI: 10.1142/S0217751X89000522
  • [21] Daniel Green et al. “Exactly Marginal Deformations and Global Symmetries” In JHEP 06, 2010, pp. 106 DOI: 10.1007/JHEP06(2010)106
  • [22] Barak Kol “On conformal deformations” In JHEP 09, 2002, pp. 046 DOI: 10.1088/1126-6708/2002/09/046
  • [23] Barak Kol “On Conformal Deformations II”, 2010 arXiv:1005.4408 [hep-th]
  • [24] J. Polchinski “String theory. Vol. 2: Superstring theory and beyond”, Cambridge Monographs on Mathematical Physics Cambridge University Press, 2007 DOI: 10.1017/CBO9780511618123
  • [25] Simon Caron-Huot “Analyticity in Spin in Conformal Theories” In JHEP 09, 2017, pp. 078 DOI: 10.1007/JHEP09(2017)078
  • [26] William A. Bardeen, Moshe Moshe and Myron Bander “Spontaneous Breaking of Scale Invariance and the Ultraviolet Fixed Point in O(N)\mathrm{O}(N)-Symmetric (φ36{\varphi}_{3}^{6}) Theory” In Phys. Rev. Lett. 52 American Physical Society, 1984, pp. 1188–1191 DOI: 10.1103/PhysRevLett.52.1188
  • [27] Ofer Aharony and Adar Sharon “Large N renormalization group flows in 3d 𝒩\mathcal{N} = 1 Chern-Simons-Matter theories” In JHEP 07, 2019, pp. 160 DOI: 10.1007/JHEP07(2019)160
  • [28] Debanjan Chowdhury and Brian Swingle “Onset of many-body chaos in the O(N)O(N) model” In Phys. Rev. D 96.6, 2017, pp. 065005 DOI: 10.1103/PhysRevD.96.065005