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

institutetext: Kavli Institute for Theoretical Sciences (KITS),
University of Chinese Academy of Sciences (UCAS), Beijing 100190, China

Half-Wormholes and Ensemble Averages

Abstract

We study “half-wormhole-like” saddle point contributions to spectral correlators in a variety of ensemble average models, including various statistical models, generalized 0d SYK models, 1d Brownian SYK models and an extension of it. In statistical ensemble models, where more general distributions of the random variables could be studied in great details, we find the accuracy of the previously proposed approximation for the half-wormholes could be improved when the distribution of the random variables deviate significantly from Gaussian distributions. We propose a modified approximation scheme of the half-wormhole contributions that also work well in these more general theories. In various generalized 0d SYK models we identify new half-wormhole-like saddle point contributions. In the 0d SYK model and 1d Brownian SYK model, apart from the wormhole and half-wormhole saddles, we find new non-trivial saddles in the spectral correlators that would potentially give contributions of the same order as the trivial self-averaging saddles. However after a careful Lefschetz-thimble analysis we show that these non-trivial saddles should not be included. We also clarify the difference between “linked half-wormholes” and “unlinked half-wormholes” in some models.

1 Introduction

The AdS/CFT correspondence Maldacena:1997re ; Witten:1998qj ; Gubser:1998bc provides a non-perturbative definition of quantum gravity. An important lesson from the recently progress in understanding the black hole information paradox is that a summation of different configurations in the semi-classical gravitational path integral is crucial to probe some quantum mechanical properties of the system, such as the Page curve Penington:2019npb ; Almheiri:2019psf ; Almheiri:2019hni ; Penington:2019kki , the late-time behavior of the spectral form factor Saad:2019lba ; Saad:2018bqo , and correlation functions Saad:2019pqd ; Yan:2022nod , see also a recent review in Bousso:2022ntt . However, the inclusion of spacetime wormholes leads to an apparent factorization puzzle Maldacena:2004rf ; a holographic computation of the correlation functions of field theory partition functions living on different boundaries gives non-factorized results, i.e. ZLZRZL×ZR\langle Z_{L}Z_{R}\rangle\neq\langle Z_{L}\rangle\times\langle Z_{R}\rangle, which is in tension with the general expectation on the field theory side. This revitalizes the hypothetical connection between wormholes and ensemble averages Coleman:1988cy ; Giddings:1988wv ; Giddings:1988cx ; Polchinski:1994zs , and motivates an appealing conjectural duality between a bulk gravitational theory and (the average of) an ensemble of theories on the boundary Saad:2019lba ; Stanford:2019vob ; Iliesiu:2019lfc ; Kapec:2019ecr ; Maxfield:2020ale ; Witten:2020wvy ; Mefford:2020vde ; Altland:2020ccq ; Eberhardt:2021jvj ; Stanford:2021bhl ; Arefeva:2019buu ; Betzios:2020nry ; Anninos:2020ccj ; Berkooz:2020uly ; Mertens:2020hbs ; Turiaci:2020fjj ; Anninos:2020geh ; Gao:2021uro ; Godet:2021cdl ; Johnson:2021owr ; Blommaert:2021etf ; Okuyama:2019xbv ; Forste:2021roo ; Maloney:2020nni ; Afkhami-Jeddi:2020ezh ; Cotler:2020ugk ; Benjamin:2021wzr ; Perez:2020klz ; Cotler:2020hgz ; Ashwinkumar:2021kav ; Afkhami-Jeddi:2021qkf ; Collier:2021rsn ; Benjamin:2021ygh ; Dong:2021wot ; Dymarsky:2020pzc ; Meruliya:2021utr ; Bousso:2020kmy ; Janssen:2021stl ; Cotler:2021cqa ; Marolf:2020xie ; Balasubramanian:2020jhl ; Gardiner:2020vjp ; Belin:2020hea ; Belin:2020jxr ; Altland:2021rqn ; Belin:2021ibv ; Peng:2021vhs ; Banerjee:2022pmw ; Heckman:2021vzx ; Johnson:2022wsr ; Collier:2022emf ; Chandra:2022bqq ; Schlenker:2022dyo , whose prototype is the by-now well known duality between the two-dimensional Jackiw-Teitelboim (JT) gravity Jackiw:1984je ; Teitelboim:1983ux and the Schwarzian sector of the Sachdev-Ye-Kitaev (SYK) model Sachdev:1992fk ; KitaevTalk2 , or more directly the random matrix theories Saad:2019lba ; Stanford:2019vob . Alternatively, an interesting question is whether there exist other configurations whose inclusion into the gravitational path integral would capture properties of a single boundary theory that are washed out after averaging over the ensemble. This is closely related to the belief that solving the factorization problem will shed light on the microscopic structure of quantum gravity such as the microstates or the states behind the horizon of the black hole; these fine structures are not universal so they can not be captured by the ensemble averaged quantities Stanford:2020wkf ; Almheiri:2021jwq . In Saad:2021uzi , the factorization problem is carefully studied in a toy model introduced in Marolf:2020xie , where it is shown that the (approximate) factorization can be restored if other half-wormhole contributions are included. In the dual field theory analysis, these half-wormhole contributions are identified with non-self-averaging saddle points in the ensemble averaged theories. This idea is explicitly realized in a 0-dimensional “one-time” SYK model in Saad:2021rcu , followed by further analyses in different models Mukhametzhanov:2021nea ; Garcia-Garcia:2021squ ; Choudhury:2021nal ; Mukhametzhanov:2021hdi ; Okuyama:2021eju ; Goto:2021mbt ; Blommaert:2021fob ; Goto:2021wfs . An explicit connection between the gravity computation in Saad:2021uzi and the field theory computation in Saad:2021rcu is proposed in Peng:2021vhs .

The construction of half-wormhole in Saad:2021rcu is based on the G,ΣG,\Sigma effective action of the model that comes from the Gaussian statistics of the random coupling. Furthermore, a prescription to identify the half-wormhole contribution is proposed and verified for the 0-dimensional SYK model and GUE matrix model in Mukhametzhanov:2021hdi . This raised a question of whether half-wormhole contributions also exist in different ensemble theories, such as those with random variables from a Poisson distribution Peng:2020rno or a uniform distribution on the moduli space Maloney:2020nni ; Afkhami-Jeddi:2020ezh ; Cotler:2020ugk ; Perez:2020klz ; Benjamin:2021wzr ; Dong:2021wot ; Collier:2022emf ; Chandra:2022bqq , and whether these contributions share the same general properties as those discussed in Saad:2021rcu and Mukhametzhanov:2021hdi .

In this paper we study the half-wormhole-like contributions that characterize the distinct behaviors of each individual theory in an ensemble of theories, and test the approximation schemes of the half-wormholes in various models. Our main findings are summarized as follows.

1.1 Summary of our main results

  • To understand the nature of the half-wormhole contributions in the 1-time SYK model, an approximation scheme is proposed in Mukhametzhanov:2021hdi . Since the proposal does not rely on specific details of the SYK model, such as the collective GG and Σ\Sigma variables, it is interesting to understand if there is a similar approximation that applies to more general ensemble averaged theories. In this paper, we first consider various statistical models with a single or multiple random variables. We compute a variety of different quantities, such as simple observables, power-sum observables and product observables, before and after the statistical average. We propose an approximation formula for the half-wormhole like contributions in general statistical models, which generalizes the one in Mukhametzhanov:2021hdi , and show their validity explicitly. We find the validity of the “wormhole/half-wormhole” approximation crucially depend on the large-NN factorization property of the observables we consider. The large-NN constraints such as traces and determinants play crucial roles in the validity of this approximation.

  • We review the 0-dimensional SYK model introduced in Saad:2021rcu and fill in technical details of some calculations. In particular, in the saddle point analysis of various quantities, such as Φ(σ)2\langle\Phi(\sigma)^{2}\rangle and others, we find new non-trivial saddle points whose on-shell values, including the 1-loop corrections, are of the same order as the the trivial saddle that is accounted for the half-wormhole. We then carry out explicit Lefschetz-thimbles analyses to conclude that the contributions from these non-trivial saddle points should not be included in the path integral, which supports the previous results in Saad:2021rcu . We also extend some of the computations to two-loop order and again find our results support previous conclusions in Saad:2021rcu .

  • We generalize the 0-dimensional SYK model so that the random coupling Ji1iqJ_{i_{1}\dots i_{q}} can be drawn from more general distributions, with non-vanishing mean or higher order cumulants.

    When Ji1iqJ_{i_{1}\dots i_{q}} has a non-vanishing mean value, we find new half-wormhole saddle of zz in additional to the linked half-wormhole saddle of z2z^{2}. We introduce new collective variables G,ΣG,\Sigma to compute z\langle z\rangle and identify the contributions from the half-wormhole saddle. We further consider the half-wormhole proposal in this context. We find that depending on the relative ratio between the different cumulants, different “multiple-linked-wormholes” could be dominant. In particular, in very special limits approximate factorization could hold automatically and no other “half-wormholes” saddles are needed.

    In models with non-vanishing higher cumulants of the random coupling, e.g. Ji1,iq40\langle J_{i_{1},\dots i_{q}}^{4}\rangle\neq 0, we find a similar conclusion that the saddle point contributes. Equivalently, the bulk configurations that dominate the path integral depends crucially on the ratios of the various cumulants and the result is not universal.

    In addition, we do a preliminary analysis of models whose random couplings Ji1,iqJ_{i_{1},\dots i_{q}} are drawn from a discrete distribution, the Poisson distribution, where more complicated saddle points can be found.

  • We do a similar analysis explicitly to the Brownian SYK model, and identify the wormhole and half-wormhole saddles at late time. The results are computed from both an explicit integration and a saddle point analysis, and we find a perfect agreement between them. We test the approximation of the partition function by its mean value and the half wormhole saddle, and further show that this approximation is good by demonstrating that the error of this approximation is small. Interestingly, like in the 0-dimensional model we also find non-trivial saddles for Φ(σ)2\langle\Phi(\sigma)^{2}\rangle and they should be excluded by a similar Lefschetz thimble analysis.

  • We further investigate modified 0d and 1d SYK model whose random couplings have non-vanishing mean values that are written in terms of products of some background Majorana fermions Goto:2021wfs . We compute explicitly the wormhole and a new type of saddle point, the “unlinked half-wormholes”, that contribute to the partition function. We show these unlink half-wormholes are closely related to the disconnected saddles due to the non-vanishing mean value of the random coupling.

2 Statistical models

In this section we consider statistical models, which can be considered as toy models of the Random Matrix Theories, to test the idea of half-wormholes in ensemble theories with random variables drawn from different distributions.

2.1 Models of a single random variable

Let XX be a random variable with a PDF P(X)P(X) that satisfies the inequality

X2X2,\displaystyle\langle X^{2}\rangle\geq\langle X\rangle^{2}\,, (1)

that is valid for all conventional probability distributions. To identify the “half-wormhole contributions” in this model, we consider the unaveraged observable XX,X2X^{2} etc., and rewrite

Xn\displaystyle X^{n} =𝑑xδ(xX)xnP(x)P(X)=𝑑xdk2πeik(xX)xnP(x)P(X)=dk2πeikXP(X)xneikx,\displaystyle=\int dx\,\delta(x-X)\frac{x^{n}P(x)}{P(X)}=\int dx\int\frac{dk}{2\pi}\,e^{\text{i}k(x-X)}\frac{x^{n}P(x)}{P(X)}=\int\frac{dk}{2\pi}\frac{e^{-\text{i}kX}}{P(X)}\langle x^{n}e^{\text{i}kx}\rangle\,, (2)

where as usual the angle bracket denotes the average of xx with the probability distribution P(x)P(x)

𝒪eikx=𝑑x𝒪eikxP(x).\displaystyle\langle\mathcal{O}e^{\text{i}kx}\rangle=\int dx\,\mathcal{O}e^{\text{i}kx}P(x)\ . (3)

Such expectation values can further be decomposed into the connected and disconnected parts, for example

xeikx=xeikx+xeikxc,\displaystyle\langle xe^{\text{i}kx}\rangle=\langle x\rangle\langle e^{\text{i}kx}\rangle+\langle xe^{\text{i}kx}\rangle_{\text{c}}\,, (4)
x2eikx=x2eikx+2xxeikxc+x2eikxc,\displaystyle\langle x^{2}e^{\text{i}kx}\rangle=\langle x^{2}\rangle\langle e^{\text{i}kx}\rangle+2\langle x\rangle\langle xe^{\text{i}kx}\rangle_{c}+\langle x^{2}e^{\text{i}kx}\rangle_{\text{c}}, (5)
x3eikx=x3eikx+3x2xeikxc+3xx2eikxc+x3eikxc,\displaystyle\langle x^{3}e^{\text{i}kx}\rangle=\langle x^{3}\rangle\langle e^{\text{i}kx}\rangle+3\langle x^{2}\rangle\langle xe^{\text{i}kx}\rangle_{c}+3\langle x\rangle\langle x^{2}e^{\text{i}kx}\rangle_{c}+\langle x^{3}e^{\text{i}kx}\rangle_{\text{c}}\,, (6)
\displaystyle\dots

where the subscript cc denotes “connected” or “cumulant” which can be defined recursively as

xeikxc=xeikxxeikx,\displaystyle\langle xe^{\text{i}kx}\rangle_{\text{c}}=\langle xe^{\text{i}kx}\rangle-\langle x\rangle\langle e^{\text{i}kx}\rangle, (7)
x2eikxc=x2eikxx2eikx2xxeikxc,\displaystyle\langle x^{2}e^{\text{i}kx}\rangle_{\text{c}}=\langle x^{2}e^{\text{i}kx}\rangle-\langle x^{2}\rangle\langle e^{\text{i}kx}\rangle-2\langle x\rangle\langle xe^{\text{i}kx}\rangle_{\text{c}}, (8)
\displaystyle\dots

There is a diagrammatic way to understand this result that closely resembles the 2-dimensional topological gravity model which is introduced in Marolf:2020xie . Formally writing

𝒪eikx=𝒪|eikx,\displaystyle\langle\mathcal{O}e^{\text{i}kx}\rangle=\langle\mathcal{O}|e^{\text{i}kx}\rangle, (9)

we can interpret the state |eikx|e^{\text{i}kx}\rangle as a “spacetime” D-braneik{}_{\text{i}k} state that is similar to that introduced in Marolf:2020xie . Then the relation (5) can be understood as in Figure 1 where the meaning of the subscript cc is transparent.

Refer to caption
Refer to caption
Figure 1: Each xx denotes a circular boundary and the bracket \langle\cdot\rangle denotes a bulk amplitude. The first two diagrams denote x2eikx\langle x^{2}\rangle\langle e^{\text{i}kx}\rangle and the last two diagrams denote the “connected” parts of the correlation function 2xxeikxc+x2eikxc2\langle x\rangle\langle xe^{\text{i}kx}\rangle_{c}+\langle x^{2}e^{\text{i}kx}\rangle_{c}.

We would like to get an estimation of the difference between any quantity XnX^{n} and its ensemble average Xn\langle X^{n}\rangle, which requires a simple evaluation of xneikx\langle x^{n}e^{ikx}\rangle. Motivated by the diagrams in Figure 1 and a similar proposal in Mukhametzhanov:2021hdi , we propose the following approximation

x2eikxcxeikxc1eikxxeikxc,\displaystyle{\langle x^{2}e^{\text{i}kx}\rangle_{c}}\approx\langle xe^{\text{i}kx}\rangle_{\text{c}}\frac{1}{\langle e^{\text{i}kx}\rangle}\langle xe^{\text{i}kx}\rangle_{\text{c}}\,, (10)

which has a diagrammatic interpretation as a recursive computation of configurations with a higher number of contractions to the spacetime brane from gluing the fundamental building blocks xeikxc\langle xe^{\text{i}kx}\rangle_{\text{c}} with the “propagator” eikx1\langle e^{\text{i}kx}\rangle^{-1}.

Equivalently, this relation can be presented as

x2eikxeikx\displaystyle\frac{\langle x^{2}e^{\text{i}kx}\rangle}{\langle e^{\text{i}kx}\rangle} \displaystyle\approx x2x2+xeikxxeikxeikxeikx\displaystyle\langle x^{2}\rangle-\langle x\rangle^{2}+\frac{\langle xe^{\text{i}kx}\rangle\langle xe^{\text{i}kx}\rangle}{\langle e^{\text{i}kx}\rangle\langle e^{\text{i}kx}\rangle} (11)
=\displaystyle= x2+2xxeikxceikx+xeikxcxeikxceikxeikx.\displaystyle\langle x^{2}\rangle+2\langle x\rangle\frac{\langle xe^{\text{i}kx}\rangle_{\text{c}}}{\langle e^{\text{i}kx}\rangle}+\frac{\langle xe^{\text{i}kx}\rangle_{\text{c}}\langle xe^{\text{i}kx}\rangle_{\text{c}}}{\langle e^{\text{i}kx}\rangle\langle e^{\text{i}kx}\rangle}\ . (12)

Making use of the fact that the quantity eikxφ(k)\langle e^{\text{i}kx}\rangle\equiv\varphi(k) is the characteristic function of the probability distribution whose inverse Fourier transformation is the PDF

12πφ(k)eikX𝑑k=P(X),\displaystyle\frac{1}{2\pi}\int\varphi(k)e^{-\text{i}kX}dk=P(X)\,, (13)

the relation (10) is equivalent to

X2\displaystyle X^{2} \displaystyle\approx X2X2+Φ,Φ=12π𝑑keikXP(X)eikx(xeikxeikx)2.\displaystyle\langle X^{2}\rangle-\langle X\rangle^{2}+\Phi\,,\quad\Phi=\frac{1}{2\pi}\int dk\frac{e^{-\text{i}kX}}{P(X)}\langle e^{\text{i}kx}\rangle\left(\frac{\langle xe^{\text{i}kx}\rangle}{\langle e^{\text{i}kx}\rangle}\right)^{2}. (14)

A more instructive form of this approximation is

X2X2+Φ~,Φ~=12π𝑑keikXP(X)eikx(xeikxc2eikx2+2xxeikxceikx),\displaystyle X^{2}\approx\langle X^{2}\rangle+\tilde{\Phi}\,,\quad\tilde{\Phi}=\frac{1}{2\pi}\int dk\frac{e^{-\text{i}kX}}{P(X)}\langle e^{\text{i}kx}\rangle\left(\frac{\langle xe^{\text{i}kx}\rangle^{2}_{c}}{\langle e^{\text{i}kx}\rangle^{2}}+\frac{2\langle x\rangle\langle xe^{\text{i}kx}\rangle_{c}}{\langle e^{\text{i}kx}\rangle}\right)\,, (15)

where Φ~=0\langle\tilde{\Phi}\rangle=0. We will call the connected piece X2cX2X2\langle X^{2}\rangle_{c}\equiv\langle X^{2}\rangle-\langle X\rangle^{2} the “wormhole” contribution and Φ\Phi the “half-wormhole” contribution although it’s mean value is non-vanishing.

As a simple example, the Gaussian distribution 𝒩(μ,t2+μ2)\mathcal{N}(\mu,t^{2}+\mu^{2}) has the non-vanishing cumulants

c1=μ,c2=t2,\displaystyle c_{1}=\mu,\quad c_{2}=t^{2}, (16)

such that

xeikxeikx=μ+ikt2,(xeikxeikx)2=μ2k2t4+2ikμt2.\displaystyle\frac{\langle xe^{\text{i}kx}\rangle}{\langle e^{\text{i}kx}\rangle}=\mu+\text{i}kt^{2},\quad\left(\frac{\langle xe^{\text{i}kx}\rangle}{\langle e^{\text{i}kx}\rangle}\right)^{2}=\mu^{2}-k^{2}t^{4}+2\text{i}k\mu t^{2}. (17)

Substituting the above into (14) gives

Φ=X2t2,\displaystyle\Phi=X^{2}-t^{2}, (18)

which means that for Gaussian distribution the approximation (14) is actually exact. Clearly, this approximation cannot be exact for an arbitrarily general probability distribution. For example, for exponential distribution (λ)\mathcal{E}(\lambda) the half-wormhole part is given by

Φ=X22,x0,\displaystyle\Phi=\frac{X^{2}}{2}\ ,\qquad x\geq 0\,, (19)

and we quantify the error by its ratio to the variance of X2X^{2}

Error=X2X2+X2Φ,ρ=Error2X4=524.\displaystyle\text{Error}=X^{2}-\langle X^{2}\rangle+\langle X\rangle^{2}-\Phi\,,\qquad\rho=\frac{\langle\text{Error}^{2}\rangle}{\langle X^{4}\rangle}=\frac{5}{24}\ . (20)

In fact, the error of the approximation (10) or (14) can be derived explicitly for any general distribution. Denoting the cumulants of the probability distribution as cnc_{n}, namely

logeikxlogφ(k)=n=0cn(ik)nn!,\displaystyle\log\langle e^{\text{i}kx}\rangle\equiv\log\varphi(k)=\sum_{n=0}^{\infty}c_{n}\frac{(ik)^{n}}{n!}\,, (21)

we find111Notice that c\langle\cdot\rangle_{c} is not a linear functional, so we don’t expect similar relations for xneikx\langle x^{n}e^{ikx}\rangle.

(ik)logeikx=xeikxeikx=xeikxceikx+x=n=0cn+1(ik)nn!,\displaystyle(-\text{i}\partial_{k})\log\langle e^{\text{i}kx}\rangle=\frac{\langle xe^{\text{i}kx}\rangle}{\langle e^{\text{i}kx}\rangle}=\frac{\langle xe^{\text{i}kx}\rangle_{c}}{\langle e^{\text{i}kx}\rangle}+\langle x\rangle=\sum_{n=0}^{\infty}c_{n+1}\frac{(\text{i}k)^{n}}{n!}\,, (22)

which means

xeikxceikx=n=1cn+1(ik)nn!.\displaystyle\frac{\langle xe^{\text{i}kx}\rangle_{c}}{\langle e^{\text{i}kx}\rangle}=\sum_{n=1}^{\infty}c_{n+1}\frac{(\text{i}k)^{n}}{n!}\ . (23)

Similarly,

(ik)2logeikx=x2eikxeikxxeikxxeikxeikxeikx=n=0cn+2(ik)nn!,\displaystyle(-\text{i}\partial_{k})^{2}\log\langle e^{\text{i}kx}\rangle=\frac{\langle x^{2}e^{\text{i}kx}\rangle}{\langle e^{\text{i}kx}\rangle}-\frac{\langle xe^{\text{i}kx}\rangle\langle xe^{\text{i}kx}\rangle}{\langle e^{\text{i}kx}\rangle\langle e^{\text{i}kx}\rangle}=\sum_{n=0}^{\infty}c_{n+2}\frac{(\text{i}k)^{n}}{n!}\,, (24)

which means

x2eikxceikxxeikxcxeikxceikxeikx=n=1cn+2(ik)nn!.\displaystyle\frac{\langle x^{2}e^{\text{i}kx}\rangle_{c}}{\langle e^{\text{i}kx}\rangle}-\frac{\langle xe^{\text{i}kx}\rangle_{\text{c}}\langle xe^{\text{i}kx}\rangle_{\text{c}}}{\langle e^{\text{i}kx}\rangle\langle e^{\text{i}kx}\rangle}=\sum_{n=1}^{\infty}c_{n+2}\frac{(\text{i}k)^{n}}{n!}\ . (25)

The approximation (10) is thus originated from neglecting all higher ckc_{k} with k>2k>2.

This implies that indeed the approximation (10) or (14) is exact when the distribution is Gaussian, namely cn=0c_{n}=0 for n>2n>2.

Similarly we can consider the approximation of XnX^{n}. We first derive the approximation of the connected correlators in the presence of spacetime brane. Taking the higher order derivative of the cumulant generating functions, for example when n=3n=3, we get

(ik)3logeikx=x3eikxeikx3x2eikxxeikxeikxeikx+2(xeikxeikx)3.\displaystyle(-\text{i}\partial_{k})^{3}\log\langle e^{\text{i}kx}\rangle=\frac{\langle x^{3}e^{\text{i}kx}\rangle}{\langle e^{\text{i}kx}\rangle}-3\frac{\langle x^{2}e^{\text{i}kx}\rangle\langle xe^{\text{i}kx}\rangle}{\langle e^{\text{i}kx}\rangle\langle e^{\text{i}kx}\rangle}+2\left(\frac{\langle xe^{\text{i}kx}\rangle}{\langle e^{\text{i}kx}\rangle}\right)^{3}\ . (26)

Separating out connected and disconnected parts, we get

(ik)3logeikx=x3eikxceikx3x2eikxcxeikxceikxeikx+2(xeikxceikx)3+x3c,\displaystyle(-\text{i}\partial_{k})^{3}\log\langle e^{\text{i}kx}\rangle=\frac{\langle x^{3}e^{\text{i}kx}\rangle_{c}}{\langle e^{\text{i}kx}\rangle}-3\frac{\langle x^{2}e^{\text{i}kx}\rangle_{c}\langle xe^{\text{i}kx}\rangle_{c}}{\langle e^{\text{i}kx}\rangle\langle e^{\text{i}kx}\rangle}+2\left(\frac{\langle xe^{\text{i}kx}\rangle_{c}}{\langle e^{\text{i}kx}\rangle}\right)^{3}+\langle x^{3}\rangle_{c}\,, (27)

where

x3c=x33x2x+2x3,\displaystyle\langle x^{3}\rangle_{c}=\langle x^{3}\rangle-3\langle x^{2}\rangle\langle x\rangle+2\langle x\rangle^{3}\,, (28)

is the connected correlator that equals to c3c_{3}. Therefore we arrive at

x3eikxceikx3x2eikxcxeikxceikxeikx+2(xeikxceikx)3=n=1cn+3(ik)nn!.\displaystyle\frac{\langle x^{3}e^{\text{i}kx}\rangle_{c}}{\langle e^{\text{i}kx}\rangle}-3\frac{\langle x^{2}e^{\text{i}kx}\rangle_{c}\langle xe^{\text{i}kx}\rangle_{c}}{\langle e^{\text{i}kx}\rangle\langle e^{\text{i}kx}\rangle}+2\left(\frac{\langle xe^{\text{i}kx}\rangle_{c}}{\langle e^{\text{i}kx}\rangle}\right)^{3}=\sum_{n=1}^{\infty}c_{n+3}\frac{(\text{i}k)^{n}}{n!}\ . (29)

This means up to the third cumulant we have approximately

x3eikxceikx3x2eikxcxeikxceikxeikx2(xeikxceikx)3,\displaystyle\frac{\langle x^{3}e^{\text{i}kx}\rangle_{c}}{\langle e^{\text{i}kx}\rangle}\approx 3\frac{\langle x^{2}e^{\text{i}kx}\rangle_{c}\langle xe^{\text{i}kx}\rangle_{c}}{\langle e^{\text{i}kx}\rangle\langle e^{\text{i}kx}\rangle}-2\left(\frac{\langle xe^{\text{i}kx}\rangle_{c}}{\langle e^{\text{i}kx}\rangle}\right)^{3}\,, (30)

and the error of this approximation is due to neglecting all ckc_{k} with k>3k>3. It is clear from this computation that the error of this approximation can be determined by (14). If the accuracy requirement is only up to the second moment, it up to quadratic fluctuations, we can use the approximation (10) again to get

x3eikxceikx(xeikxceikx)3,\displaystyle\frac{\langle x^{3}e^{\text{i}kx}\rangle_{c}}{\langle e^{\text{i}kx}\rangle}\approx\left(\frac{\langle xe^{\text{i}kx}\rangle_{c}}{\langle e^{\text{i}kx}\rangle}\right)^{3}\ , (31)

which becomes exact when the distribution is Gaussian. In fact, we can derive similar relations by taking higher order derivatives in (26) to get relations among higher order xieikxc\langle x^{i}e^{\text{i}kx}\rangle_{c}’s. If again we need accuracy up to quadratic order one can prove by induction

xneikxceikx(xeikxceikx)n.\displaystyle\frac{\langle x^{n}e^{\text{i}kx}\rangle_{c}}{\langle e^{\text{i}kx}\rangle}\approx\left(\frac{\langle xe^{\text{i}kx}\rangle_{c}}{\langle e^{\text{i}kx}\rangle}\right)^{n}\ . (32)

We can then approximate the un-average X3X^{3} to a required accuracy. In practice, we rewrite the definition of XnX^{n} according to (2), then expand the xneikx\langle x^{n}e^{\text{i}kx}\rangle in (2) in terms of the connected correlators xieikxc\langle x^{i}e^{\text{i}kx}\rangle_{c} according to e.g. (4)-(6). Then depending on the accuracy requirement, we use relations analogous to either (30) or (61), (32), to write down the approximation and the error of the final approximation is the composition of the errors the different approximations of xneikx\langle x^{n}e^{ikx}\rangle. The general expression of the approximation of XnX^{n} and the corresponding errors are complicated. But we will present some general procedures that work for any distribution once an accuracy goal is given.

2.1.1 Recursion relations for approximations to arbitrary accuracy

Define Φn=12πeikXP(X)eikx1nxeikxn\Phi_{n}=\frac{1}{2\pi}\int\frac{e^{\text{i}kX}}{P(X)}\langle e^{\text{i}kx}\rangle^{1-n}\langle xe^{\text{i}kx}\rangle^{n}, we have

XmΦn\displaystyle X^{m}\Phi_{n} =12πXmeikXP(X)eikx1nxeikxn=12πeikXP(X)(ik)m(eikx1nxeikxn).\displaystyle=\frac{1}{2\pi}\int\frac{X^{m}e^{\text{i}kX}}{P(X)}\langle e^{\text{i}kx}\rangle^{1-n}\langle xe^{\text{i}kx}\rangle^{n}=\frac{1}{2\pi}\int\frac{e^{\text{i}kX}}{P(X)}\left(-i\partial_{k}\right)^{m}\left(\langle e^{\text{i}kx}\rangle^{1-n}\langle xe^{\text{i}kx}\rangle^{n}\right)\ . (33)

Evaluating the derivative gives a result involving xieikx\langle x^{i}e^{\text{i}kx}\rangle with 1im+11\leq i\leq m+1. Rewriting them in terms of xieikxc\langle x^{i}e^{\text{i}kx}\rangle_{c} with the help of e.g. (4)-(6). Then use the approximation either (30) or (61), (32) according to the required accuracy. Then rewrite the xieikxc\langle x^{i}e^{\text{i}kx}\rangle_{c} in the approximated results back in terms of xieikx\langle x^{i}e^{\text{i}kx}\rangle, and the result will be a relation among Φi\Phi_{i} with 1im+11\leq i\leq m+1. Making use of the fact that Φ1=X\Phi_{1}=X and recursively carrying out the above procedure to evaluate Xn1Φ1X^{n-1}\Phi_{1}, we get the approximation of XnX^{n} to the desired accuracy.

For example, if we require accuracy to the second order, we simply consider

XΦn\displaystyle X\Phi_{n} =12πeikXP(X)eikxn(nx2eikxxeikxn1eikx+(1n)xeikxn+1).\displaystyle=\frac{1}{2\pi}\int\frac{e^{\text{i}kX}}{P(X)}\langle e^{\text{i}kx}\rangle^{-n}\left(n\langle x^{2}e^{\text{i}kx}\rangle\langle xe^{\text{i}kx}\rangle^{n-1}\langle e^{\text{i}kx}\rangle+(1-n)\langle xe^{\text{i}kx}\rangle^{n+1}\right)\ . (34)

Following the above procedure to rewrite x2eikx\langle x^{2}e^{\text{i}kx}\rangle, we arrive at

XΦn=n(x2x2)Φn1+Φn+1.\displaystyle X\Phi_{n}=n\left(\langle x^{2}\rangle-\langle x\rangle^{2}\right)\Phi_{n-1}+\Phi_{n+1}\ . (35)

For example, we can evaluate

X3=X2Φ1=3(μ2μ12)Φ1+Φ3,\displaystyle X^{3}=X^{2}\Phi_{1}=3\left(\mu_{2}-\mu_{1}^{2}\right)\Phi_{1}+\Phi_{3}\,, (36)

where we keep only accuracy up to the quadratic order, so μ3\mu_{3} does not appear independently; it is simply replaced by

μ3=3μ1μ22μ13.\displaystyle\mu_{3}=3\mu_{1}\mu_{2}-2\mu_{1}^{3}\ . (37)

2.1.2 Explicit relations for Gaussian approximation

If we only want Gaussian approximations of XnX^{n}, we can get an explicit approximation formula. First let introduce some convenient notations

ϕn=xneikxeikx,ϕnc=xneikxceikx,\displaystyle\phi_{n}=\frac{\langle x^{n}e^{\text{i}kx}\rangle}{\langle e^{\text{i}kx}\rangle},\quad\phi^{c}_{n}=\frac{\langle x^{n}e^{\text{i}kx}\rangle_{c}}{\langle e^{\text{i}kx}\rangle}, (38)
xn=μn,xncumulant=cn.\displaystyle\langle x^{n}\rangle=\mu_{n},\quad\langle x^{n}\rangle_{\text{cumulant}}=c_{n}. (39)

The cumulant cmc_{m} can be expressed as a polynomial of moments

cm=Pm(μm,μm1,,μ1).\displaystyle c_{m}=P_{m}(\mu_{m},\mu_{m-1},\dots,\mu_{1}). (40)

Some examples are

c1=μ1,c2=μ2μ12,c3=μ33μ1μ2+2μ13,\displaystyle c_{1}=\mu_{1},\quad c_{2}=\mu_{2}-\mu_{1}^{2},\quad c_{3}=\mu_{3}-3\mu_{1}\mu_{2}+2\mu_{1}^{3},\dots (41)

Note that the coefficient of μm\mu_{m} is 1. Of course the relations can be inverted

μm=Qm(cm,cm1,c1).\displaystyle\mu_{m}=Q_{m}(c_{m},c_{m-1},\dots c_{1}). (42)

Similar to (4),(5) and (6), ϕn\phi_{n} can be decomposed as

ϕm=P~m(ϕmc,,ϕ0c),\displaystyle\phi_{m}=\tilde{P}_{m}(\phi_{m}^{c},\dots,\phi_{0}^{c}), (43)

for example

ϕ1=ϕ1c+μ1ϕ0c,ϕ2=ϕ2c+2μ1ϕ1c+μ2,.\displaystyle\phi_{1}=\phi_{1}^{c}+\mu_{1}\phi_{0}^{c},\quad\phi_{2}=\phi_{2}^{c}+2\mu_{1}\phi_{1}^{c}+\mu_{2},\dots. (44)

Since logeikx\log\langle e^{\text{i}kx}\rangle is the generating function of cnc_{n} we have 222The simplest way to see this is to set k=0k=0, then it reduces to (40) and to notice that the coefficients of the polynomial PmP_{m} do not depend on kk.

(ik)mlogeikx=Pm(ϕm,ϕm1,,ϕ1)=n=0cn+m(ik)nn!.\displaystyle(-\text{i}\partial_{k})^{m}\log\langle e^{\text{i}kx}\rangle=P_{m}(\phi_{m},\phi_{m-1},\dots,\phi_{1})=\sum_{n=0}c_{n+m}\frac{(\text{i}k)^{n}}{n!}. (45)

Using (43) and (42) the left-hand side can be expanded as a polynomial of cic_{i} with coefficients to be functions of ϕic\phi_{i}^{c}:

Pm(ϕm,ϕm1,,ϕ1)=Pm(P~m(ϕic),P~m1(ϕic),,P~1(ϕic)).\displaystyle P_{m}(\phi_{m},\phi_{m-1},\dots,\phi_{1})=P_{m}(\tilde{P}_{m}(\phi_{i}^{c}),\tilde{P}_{m-1}(\phi_{i}^{c}),\dots,\tilde{P}_{1}(\phi_{i}^{c}))\ . (46)

For example

P2\displaystyle P_{2} =\displaystyle= ϕ2ϕ12=P~22P~12\displaystyle\phi_{2}-\phi_{1}^{2}=\tilde{P}_{2}-2{\tilde{P}_{1}}^{2} (47)
=\displaystyle= ϕ2c+2μ1ϕ1c+μ2ϕ1c2μ12ϕ0c22μ1ϕ1cϕ0c\displaystyle\phi_{2}^{c}+2\mu_{1}\phi_{1}^{c}+\mu_{2}-{\phi_{1}^{c}}^{2}-\mu_{1}^{2}{\phi_{0}^{c}}^{2}-2\mu_{1}\phi_{1}^{c}\phi_{0}^{c} (48)
=\displaystyle= ϕ2cϕ1c2+c1(2ϕ1c2ϕ1cϕ0c)+c2.\displaystyle\phi_{2}^{c}-{\phi_{1}^{c}}^{2}+c_{1}(2\phi_{1}^{c}-2\phi_{1}^{c}\phi_{0}^{c})+c_{2}. (49)

Therefore we end up with

Pm=Mm+c1Mm1(1)+(c12Mm2(1)+c2Mm2(2))++cm=n=0cn+m(ik)nn!,\displaystyle P_{m}=M_{m}+c_{1}M_{m-1}^{(1)}+(c_{1}^{2}M_{m-2}^{(1)}+c_{2}M_{m-2}^{(2)})+\dots+c_{m}=\sum_{n=0}c_{n+m}\frac{(\text{i}k)^{n}}{n!}\,, (50)

where each Mi(k)M_{i}^{(k)} is a function of the ϕic\phi^{c}_{i}’s. Since the subscript ii of ϕic\phi^{c}_{i} and MiM_{i} both indicate the power of xx, it is clear that

aia=m,(aϕiac)Mm,\displaystyle\sum_{a}i_{a}=m\,,\qquad\forall\left(\prod_{a}\phi_{i_{a}}^{c}\right)\in M_{m}\,, (51)

where aϕiac\prod_{a}\phi_{i_{a}}^{c} is any term in MmM_{m}. Notice that these relations are true for arbitrary kk, mm and distributions, then the non-trivial solution is only

Mn(p)=0,Mm=Pm(ϕmc,ϕm1c,,ϕ1c)=n=1cn+m(ik)nn!.\displaystyle\quad M_{n}^{(p)}=0\,,\qquad M_{m}=P_{m}(\phi^{c}_{m},\phi^{c}_{m-1},\dots,\phi^{c}_{1})=\sum_{n=1}c_{n+m}\frac{(\text{i}k)^{n}}{n!}\ . (52)

The Gaussian approximation means cm=0c_{m}=0 for all m>2m>2. This requires

Pm(ϕmc,ϕm1c,,ϕ1c)0,m>1.\displaystyle P_{m}(\phi^{c}_{m},\phi^{c}_{m-1},\dots,\phi^{c}_{1})\approx 0\,,\quad\forall m>1\ . (53)

At m=2m=2 this relation means

P2(ϕ2c,ϕ1c)0,\displaystyle P_{2}(\phi^{c}_{2},\phi^{c}_{1})\approx 0\,, (54)

which combines with (51) means ϕ2c=α(ϕ1c)2\phi^{c}_{2}=\alpha\left(\phi^{c}_{1}\right)^{2} and

P2(αϕ2c,ϕ1c)=0.\displaystyle P_{2}(\alpha\phi^{c}_{2},\phi^{c}_{1})=0\ . (55)

To fix the normalization α\alpha, we notice that since the above relations (40) -(53), in particular the functional form of PP, are true for arbitrary distribution, we can choose the delta function distribution such that cn=0,n2c_{n}=0,\forall n\geq 2 and μm=μ1m\mu_{m}=\mu_{1}^{m}, we can get the identity

Pm(μm,μm1,μ)=0,m2,\displaystyle P_{m}(\mu^{m},\mu^{m-1},\dots\mu)=0,\quad m\geq 2, (56)

thus combining this with (55) we conclude α=1\alpha=1 and

ϕ2c(ϕ1c)2,\displaystyle\phi^{c}_{2}\approx\left(\phi^{c}_{1}\right)^{2}\,, (57)

where \approx is due to the Gaussian approximation. This is nothing but the approximation (10). Iterating this procedure successively for different mm, we reach to

ϕmc(ϕ1c)m,\displaystyle\phi_{m}^{c}\approx{(\phi_{1}^{c})}^{m}\,, (58)

in the Gaussian approximation. Then we can approximate XmX^{m} as

Xm\displaystyle X^{m} =\displaystyle= 12πeikXP(X)eikxϕm=12πeikXP(X)eikxP~m(ϕmc,,ϕ1c,1)\displaystyle\frac{1}{2\pi}\int\frac{e^{\text{i}kX}}{P(X)}\langle e^{\text{i}kx}\rangle\phi_{m}=\frac{1}{2\pi}\int\frac{e^{\text{i}kX}}{P(X)}\langle e^{\text{i}kx}\rangle\tilde{P}_{m}(\phi_{m}^{c},\dots,\phi_{1}^{c},1) (59)
\displaystyle\approx 12πeikXP(X)eikxP~m((ϕ1c)m,(ϕ1c)m1,,ϕ1c,1)\displaystyle\frac{1}{2\pi}\int\frac{e^{\text{i}kX}}{P(X)}\langle e^{\text{i}kx}\rangle\tilde{P}_{m}((\phi_{1}^{c})^{m},(\phi_{1}^{c})^{m-1},\dots,\phi_{1}^{c},1) (60)
=\displaystyle= i=0m(mi)μiΦmi,\displaystyle\sum_{i=0}^{m}{m\choose i}\mu_{i}\Phi_{m-i}\,, (61)

where Φi=12πdkeikXP(X)eikx(ϕ1c)i\Phi_{i}=\frac{1}{2\pi}\int\text{d}k\frac{e^{\text{i}kX}}{P(X)}\langle e^{\text{i}kx}\rangle(\phi_{1}^{c})^{i}, and it may be understood as generalized wormholes which we will report somewhere else.

It is easy to check that the result (61) agrees with (36) once the relation (37) is used.

2.2 Models with multiple independent identical random variables

In statistical models with a single random variable, the various moments are all observables that we can compute. On the other hand, we would like to consider other interesting observables. We therefore proceed to consider operators in statistical models with multiple independent identical random variables.

One class of operators in these models is the light operators that are simply linear combinations of the random variables XiX_{i}. We conjecture that if Y(Xi)Y(X_{i}) is some function of a large number NN independent random variables XiX_{i} such that YY is approximately Gaussian, then the approximation

Y2Y2Y2+Φ,\displaystyle Y^{2}\approx\langle Y^{2}\rangle-\langle Y\rangle^{2}+\Phi, (62)
Φ(X)=1(2π)Ni(dkieikiXiP(Xi))eiikixi(Y(x)eiikixieiikixi)2.\displaystyle\Phi(X)=\frac{1}{(2\pi)^{N}}\int\prod_{i}\left(dk_{i}\frac{e^{-\text{i}k_{i}X_{i}}}{P(X_{i})}\right)\langle e^{\text{i}\sum_{i}k_{i}x_{i}}\rangle\left(\frac{\langle Y(x)e^{\text{i}\sum_{i}k_{i}x_{i}}\rangle}{\langle e^{\text{i}\sum_{i}k_{i}x_{i}}\rangle}\right)^{2}. (63)

is good in the sense that

ρError2Y22,\displaystyle\rho\equiv\frac{\langle\text{Error}^{2}\rangle}{\langle Y^{2}\rangle^{2}}\,, (64)

is suppressed by 1/N1/N.

Like (15) we can rewrite it into

Y2Y2+Φ~,\displaystyle Y^{2}\approx\langle Y^{2}\rangle+\tilde{\Phi}, (65)
Φ~(X)=1(2π)Ni(dkieikiXiP(Xi))eiikixi(Y(x)eiikixic2eiikixi2+2YY(x)eiikixiceiikixi).\displaystyle\tilde{\Phi}(X)=\frac{1}{(2\pi)^{N}}\int\prod_{i}\left(dk_{i}\frac{e^{-\text{i}k_{i}X_{i}}}{P(X_{i})}\right)\langle e^{\text{i}\sum_{i}k_{i}x_{i}}\rangle\left(\frac{\langle Y(x)e^{\text{i}\sum_{i}k_{i}x_{i}}\rangle_{c}^{2}}{\langle e^{\text{i}\sum_{i}k_{i}x_{i}}\rangle^{2}}+\frac{2\langle Y\rangle\langle Y(x)e^{\text{i}\sum_{i}k_{i}x_{i}}\rangle_{c}}{\langle e^{\text{i}\sum_{i}k_{i}x_{i}}\rangle}\right). (66)

2.2.1 Simple observables

The fundamental logic in this section is that by the central limit theorem (CLT), summing over a large number of i.i.d random variables gives a random variable that approximately obey a Gaussian distribution. Explicitly, if XiX_{i} is from a normal distribution 𝒩(μ,σ2){\cal N}(\mu,\sigma^{2}), then the mean of NN such i.i.d’s

Y~=1Ni=1NXi,\displaystyle\tilde{Y}=\frac{1}{N}\sum_{i=1}^{N}X_{i}\,, (67)

is approximately a Gaussian random variable from 𝒩(μ,σ2/N){\cal N}(\mu,\sigma^{2}/N) when NN is large enough.

In this paper, it turns out that it is more convenient to define

Y=i=1NXi,\displaystyle Y=\sum_{i=1}^{N}X_{i}\,, (68)

so that the connection to the SYK model is more transparent. Then YY is a Gaussian random variable with probability distribution 𝒩(Nμ,Nσ2){\cal N}(N\mu,N\sigma^{2}) when NN is large. In particular, we expect

Y43Y222Y4,Y2N(X2X2)+N2X2,YNX.\displaystyle\langle Y^{4}\rangle\approx 3\langle Y^{2}\rangle^{2}-2\langle Y\rangle^{4}\,,\quad\langle Y^{2}\rangle\approx N\left(\langle X^{2}\rangle-\langle X\rangle^{2}\right)+N^{2}\langle X\rangle^{2}\,,\quad\langle Y\rangle\approx N\langle X\rangle\ . (69)

They can be checked by a direct calculation

Y2=NX2+N(N1)X2,\displaystyle\langle Y^{2}\rangle=N\langle X^{2}\rangle+N(N-1)\langle X\rangle^{2}, (70)
Y4=NX4+N(N1)(4X3X+3X22)\displaystyle\langle Y^{4}\rangle=N\langle X^{4}\rangle+N(N-1)\left(4\langle X^{3}\rangle\langle X\rangle+3\langle X^{2}\rangle^{2}\right)
+6N(N1)(N2)X2X2+N(N1)(N2)(N3)X4.\displaystyle\quad+6N(N-1)(N-2)\langle X^{2}\rangle\langle X\rangle^{2}+N(N-1)(N-2)(N-3)\langle X\rangle^{4}\ . (71)

Because all the XiX_{i} are independent so that it is straightforward to obtain

Yeikixieikixi=ixieikixieikixiiki[1].\displaystyle\frac{\langle Ye^{\text{i}k_{i}x_{i}}\rangle}{\langle e^{\text{i}k_{i}x_{i}}\rangle}=\sum_{i}\frac{\langle x_{i}e^{\text{i}k_{i}x_{i}}\rangle}{\langle e^{\text{i}k_{i}x_{i}}\rangle}\equiv\sum_{i}k_{i}[1]. (72)

Next we can rewrite the square of (72) into the diagonal terms and off-diagonal terms

(Yeikixieikixi)2=iki[1]2+ijki[1]kj[1].\displaystyle\left(\frac{\langle Ye^{\text{i}k_{i}x_{i}}\rangle}{\langle e^{\text{i}k_{i}x_{i}}\rangle}\right)^{2}=\sum_{i}k_{i}[1]^{2}+\sum_{i\neq j}k_{i}[1]k_{j}[1]. (73)

To compute the off-diagonal contributions to the half-wormhole, we observe that

1(2π)Ni(dkieikiXiP(Xi))eiikixiki[1]kj[1]\displaystyle\frac{1}{(2\pi)^{N}}\int\prod_{i}\left(dk_{i}\frac{e^{-\text{i}k_{i}X_{i}}}{P(X_{i})}\right)\langle e^{\text{i}\sum_{i}k_{i}x_{i}}\rangle k_{i}[1]k_{j}[1] (74)
=1(2π)2𝑑ki𝑑kjeikiXiikjXjP(Xi)P(Xj)xieikixixjeikjxj=XiXj.\displaystyle=\frac{1}{(2\pi)^{2}}\int dk_{i}dk_{j}\frac{e^{-\text{i}k_{i}X_{i}-\text{i}k_{j}X_{j}}}{P(X_{i})P(X_{j})}\langle x_{i}e^{\text{i}k_{i}x_{i}}\rangle\langle x_{j}e^{\text{i}k_{j}x_{j}}\rangle=X_{i}X_{j}\ . (75)

In terms of ki[n]m^\widehat{k_{i}[n]^{m}} which are defined in (585) the half-wormhole can be written as

Φ=iki[1]2^+ijXiXj,\displaystyle\Phi=\sum_{i}\widehat{k_{i}[1]^{2}}+\sum_{i\neq j}X_{i}X_{j}, (76)

and the error is given by

Error =i(Xi2ki[1]2^t2),t2=Xi2Xi2,\displaystyle=\sum_{i}\left(X_{i}^{2}-\widehat{k_{i}[1]^{2}}-t^{2}\right),\quad t^{2}=\langle X_{i}^{2}\rangle-\langle X_{i}\rangle^{2}\,, (77)
Error2\displaystyle\langle\text{Error}^{2}\rangle =i,j(Xi2ki[1]2^)(Xj2kj[1]2^)+N2t42Nt2i(Xi2ki[1]2^).\displaystyle=\sum_{i,j}\langle(X_{i}^{2}-\widehat{k_{i}[1]^{2}})(X_{j}^{2}-\widehat{k_{j}[1]^{2}})\rangle+N^{2}t^{4}-2Nt^{2}\sum_{i}(X_{i}^{2}-\widehat{k_{i}[1]^{2}})\ . (78)

Recalling that Y2Nt2\langle Y^{2}\rangle\sim Nt^{2} so to prove the conjecture (62) we need to show that the 𝒪(N2){\cal O}(N^{2}) term in (78) vanish. A direct calculation gives

ki[1]2^Xi\displaystyle\langle\widehat{k_{i}[1]^{2}}\rangle_{X_{i}} =\displaystyle= 𝑑XiP(Xi)ki[1]2eikixixieikiXiP(Xi)𝑑ki\displaystyle\int dX_{i}P(X_{i})k_{i}[1]^{2}\langle e^{\text{i}k_{i}x_{i}}\rangle_{x_{i}}\frac{e^{-\text{i}k_{i}X_{i}}}{P(X_{i})}dk_{i} (79)
=\displaystyle= 𝑑kiδ(ki)eikixixiki[1]2=Xi2.\displaystyle\int dk_{i}\delta(-k_{i})\langle e^{\text{i}k_{i}x_{i}}\rangle_{x_{i}}k_{i}[1]^{2}=\langle X_{i}\rangle^{2}\ . (80)

This means

(Xi2ki[1]2^=Xi2Xi2=t2.ki[1]2^XiXi2.\displaystyle\langle(X_{i}^{2}-\widehat{k_{i}[1]^{2}}\rangle=\langle X^{2}_{i}\rangle-\langle X_{i}\rangle^{2}=t^{2}\ .\quad\Leftrightarrow\quad\langle\widehat{k_{i}[1]^{2}}\rangle_{X_{i}}\approx\langle X_{i}\rangle^{2}\ . (81)

In particular, a consequence of this relation is that although all the 3 terms in (78) are of order 𝒪(N2){\cal O}(N^{2}), the sum of them cancelled exactly since (81) does not depend on ii. This then shows that Error2Y2\langle\text{Error}^{2}\rangle\ll\langle Y^{2}\rangle and hence the approximation (62) is valid.

We can derive this result in a more illuminating fashion. First using (23) ki[1]k_{i}[1] can be expressed as

ki[1]=n=0(ik)nn!cn+1.\displaystyle k_{i}[1]=\sum_{n=0}\frac{(-\text{i}k)^{n}}{n!}c_{n+1}. (82)

Then using the fact that the inverse Fourier transformation of the characteristic function is the PDF we find

ki[1]2^Xi=𝑑XiP(Xi)n,m=0cn+1cm+1n!m!(iki)n+meikixixieikiXiP(Xi)dki\displaystyle\langle\widehat{k_{i}[1]^{2}}\rangle_{X_{i}}=\int dX_{i}P(X_{i})\sum_{n,m=0}\frac{c_{n+1}c_{m+1}}{n!m!}(-\text{i}k_{i})^{n+m}\langle e^{\text{i}k_{i}x_{i}}\rangle_{x_{i}}\frac{e^{-\text{i}k_{i}X_{i}}}{P(X_{i})}dk_{i} (83)
=𝑑Xin,m=0cn+1cm+1n!m!(Xi)n+mP(Xi)=c12=Xi2.\displaystyle\quad=\int dX_{i}\sum_{n,m=0}\frac{c_{n+1}c_{m+1}}{n!m!}(\partial_{X_{i}})^{n+m}P(X_{i})=c_{1}^{2}=\langle X_{i}\rangle^{2}\ . (84)

2.2.2 Power-sum observables

In this section, we consider another class of more general observables

Y=if(Xi),Y2=i,jf(Xi)f(Xj),\displaystyle Y=\sum_{i}f(X_{i}),\quad Y^{2}=\sum_{i,j}f(X_{i})f(X_{j}), (85)

where XiX_{i} are still independent identical random variables with PDF PXiP_{X_{i}} and ff is some smooth function so that f(Xi)f(X_{i}) are also independent and identical random variables with a new PDF PfP_{f}:

𝑑XF[f(X)]PX=𝑑fF(f)Pf.\displaystyle\int dXF[f(X)]P_{X}=\int dfF(f)P_{f}. (86)

The CLT is still valid but the proposal may not because naively it depends on the function ff. By smooth function we mean f(Xi)f(X_{i}) is not singular anywhere such that it can be Taylor expanded

f(Xi)=nanXin,\displaystyle f(X_{i})=\sum_{n}a_{n}X_{i}^{n}\,, (87)

whose expansion coefficients satisfy

an0,n>n0,n0N.\displaystyle a_{n}\approx 0\,,\qquad\forall n>n_{0}\,,\quad n_{0}\ll N\ . (88)

Accordingly (72) and (73) become

Yeikixieikixi=inanki[n].\displaystyle\frac{\langle Ye^{\text{i}k_{i}x_{i}}\rangle}{\langle e^{\text{i}k_{i}x_{i}}\rangle}=\sum_{i}\sum_{n}a_{n}k_{i}[n]. (89)
(Yeikixieikixi)2=i,j(nanki[n]mamkj[m]).\displaystyle\left(\frac{\langle Ye^{\text{i}k_{i}x_{i}}\rangle}{\langle e^{\text{i}k_{i}x_{i}}\rangle}\right)^{2}=\sum_{i,j}\left(\sum_{n}a_{n}k_{i}[n]\sum_{m}a_{m}k_{j}[m]\right). (90)

So the error is given by

Error =i(f2(Xi)t2n,manamki[n]ki[m]^),\displaystyle=\sum_{i}\left(f^{2}(X_{i})-t^{2}-\sum_{n,m}a_{n}a_{m}\widehat{k_{i}[n]k_{i}[m]}\right)\,, (91)
Error2\displaystyle\langle\text{Error}^{2}\rangle =i,j(f2(Xi)n,manamki[n]ki[m]^)(f2(Xj)n,manamkj[n]kj[m]^)\displaystyle=\langle\sum_{i,j}(f^{2}(X_{i})-\sum_{n,m}a_{n}a_{m}\widehat{k_{i}[n]k_{i}[m]})(f^{2}(X_{j})-\sum_{n,m}a_{n}a_{m}\widehat{k_{j}[n]k_{j}[m]})\rangle
+N2t42Nt2i(f2(Xi)n,manamki[n]ki[m]^),\displaystyle\quad+N^{2}t^{4}-2Nt^{2}\sum_{i}(f^{2}(X_{i})-\sum_{n,m}a_{n}a_{m}\widehat{k_{i}[n]k_{i}[m]})\,, (92)

where t2=f2(Xi)f(Xi)2t^{2}=\langle f^{2}(X_{i})\rangle-\langle f(X_{i})\rangle^{2}. Similar to the calculation of (80), one can find

n,manamki[n]ki[m]^=f(Xi)2,\displaystyle\langle\sum_{n,m}a_{n}a_{m}\widehat{k_{i}[n]k_{i}[m]}\rangle=\langle f(X_{i})\rangle^{2}, (93)

which means the leading order terms, ie of order N2N^{2}, in (92) is

2((f2(Xj)n,manamkj[n]kj[m]^)2t4)N2=0,\displaystyle 2\left(\langle(f^{2}(X_{j})-\sum_{n,m}a_{n}a_{m}\widehat{k_{j}[n]k_{j}[m]})\rangle^{2}-t^{4}\right)N^{2}=0\,, (94)

As a result, the error is small and indeed the approximation (62) is reasonable in this case too. We also show some explicit examples in the Appendix (B). More generally, following the same procedure one can show that the half wormhole proposal is correct for the following family of functions

Yk=iN(f(Xi1,Xi2,,Xik)),\displaystyle Y_{k}=\sum_{i}^{N}\left(f(X_{i_{1}},X_{i_{2}},\dots,X_{i_{k}})\right), (95)

where XipX_{i_{p}} are independent and identical random variables.

2.2.3 Product observables

Previously the function YY we considered are a summation of (polynomials of) independent random variables. The proposal works very well for all the probability distributions. However in the original construction of half wormhole introduced in Saad:2021rcu , the function YY is a determinant observables which are “heavy” in the traditional field theory language

Y=PF(J)=A1<A2<<Apsgn(A)JA1JA2JAp,\displaystyle Y=\text{PF}(J)=\sum^{\prime}_{A_{1}<A_{2}<\dots<A_{p}}\text{sgn}(A){J}_{A_{1}}{J}_{A_{2}}\dots J_{A_{p}}, (96)

where the function PF(J)\text{PF}(J) is called the hyperpfaffian Barvinok which is a tensorial generalization of pfaffian and JAiJ_{A_{i}} are random variables. To mimic this construction let us consider a similar model:

Y=i1i2iqNXi1Xi2Xiq.\displaystyle Y=\sum_{i_{1}\neq i_{2}\neq\dots\neq i_{q}}^{N}X_{i_{1}}X_{i_{2}}\dots X_{i_{q}}. (97)

\bullet q=2q=2 Gaussian distribution
The simplest case is q=2q=2:

Y=ijXiXj,\displaystyle Y=\sum_{i\neq j}X_{i}X_{j}, (98)
Y2=ijpqXiXjXpXq+4ijpXi2XjXp+2ijXi2Xj2.\displaystyle Y^{2}=\sum_{i\neq j\neq p\neq q}X_{i}X_{j}X_{p}X_{q}+4\sum_{i\neq j\neq p}X_{i}^{2}X_{j}X_{p}+2\sum_{i\neq j}X_{i}^{2}X_{j}^{2}\ . (99)

It is straightforward to get

Y2=N(N1)(2t4+4(N1)μ2t2+N(N1)μ4),\displaystyle\langle Y^{2}\rangle=N(N-1)\left(2t^{4}+4(N-1)\mu^{2}t^{2}+N(N-1)\mu^{4}\right), (100)
Xi=μ,X2X2=t2.\displaystyle\langle X_{i}\rangle=\mu,\quad\langle X^{2}\rangle-\langle X\rangle^{2}=t^{2}. (101)

So in general Y2\langle Y^{2}\rangle will scale as N4N^{4} if μ0\mu\neq 0, while if μ=0\mu=0 it scales as N2N^{2}.

One example of the μ=0\mu=0 case is the Gaussian distribution 𝒩(μ=0,t2)\mathcal{N}(\mu=0,t^{2}). We then verifies

Y=0,Y2=2t4N(N1)\displaystyle\langle Y\rangle=0,\quad\langle Y^{2}\rangle=2t^{4}N(N-1) (102)

and

Φ=ijpqXiXjXpXq+4ijp(Xi2t2)XjXp+2ij(Xi2t2)(Xj2t2).\displaystyle\Phi=\sum_{i\neq j\neq p\neq q}X_{i}X_{j}X_{p}X_{q}+4\sum_{i\neq j\neq p}(X_{i}^{2}-t^{2})X_{j}X_{p}+2\sum_{i\neq j}(X_{i}^{2}-t^{2})(X_{j}^{2}-t^{2})\ . (103)

Therefore we obtain

Error=2t4N(N1)+4t2(N2)ijXiXj+4(N1)t2iXi22t4N(N1),\displaystyle\text{Error}=-2t^{4}N(N-1)+4t^{2}(N-2)\sum_{i\neq j}X_{i}X_{j}+4(N-1)t^{2}\sum_{i}X_{i}^{2}-2t^{4}N(N-1),
(Error/4)2=(2+1+12)N4t8+#N3+\displaystyle\langle(\text{Error}/4)^{2}\rangle=(2+1+1-2)N^{4}t^{8}+\#N^{3}+\dots (104)

the leading term does not vanish so the approximation

Y2Y2Y2+Φ,\displaystyle Y^{2}\approx\langle Y^{2}\rangle-\langle Y\rangle^{2}+\Phi\,, (105)

is not good.

However, for more general Gaussian distributions 𝒩(μ,t2)\mathcal{N}(\mu,t^{2}) similar calculation gives

Y=N(N1)μ2,Y2Y2t~2=2t2N(N1)(t2+2(N1)μ2),\displaystyle\langle Y\rangle=N(N-1)\mu^{2},\quad\langle Y^{2}\rangle-\langle Y\rangle^{2}\equiv\tilde{t}^{2}=2t^{2}N(N-1)(t^{2}+2(N-1)\mu^{2}), (106)

and

Error=t~2+4t2(N2)ijXiXj+4(N1)t2iXi22t4N(N1),\displaystyle\text{Error}=-\tilde{t}^{2}+4t^{2}(N-2)\sum_{i\neq j}X_{i}X_{j}+4(N-1)t^{2}\sum_{i}X_{i}^{2}-2t^{4}N(N-1), (107)

now we find that

Error2=32(3t2μ2+μ4)N5+32(t412t2μ24μ4)N4+\displaystyle\langle\text{Error}^{2}\rangle=32(3t^{2}\mu^{2}+\mu^{4})N^{5}+32(t^{4}-12t^{2}\mu^{2}-4\mu^{4})N^{4}+\dots (108)

and

Error2Y22=2(μ4+3t2μ23t4)(2t44t2μ2)2N+.\displaystyle\frac{\langle\text{Error}^{2}\rangle}{\langle Y^{2}\rangle^{2}}=\frac{2(\mu^{4}+3t^{2}\mu^{2}-3t^{4})}{(2t^{4}-4t^{2}\mu^{2})^{2}N}+\dots. (109)

Notice that the error is always small, even when μ0\mu\rightarrow 0, and the proposal is valid. This is because when μ0\mu\neq 0, the moments of YY behave as

YN2μ,Y2N4μ2,Y4N8μ4,\displaystyle\langle Y\rangle\approx N^{2}\mu,\quad\langle Y^{2}\rangle\approx N^{4}\mu^{2},\quad\langle Y^{4}\rangle\approx N^{8}\mu^{4}, (110)

as expected from (69). It is thus clear that the μ0\mu\to 0 limit is not smooth.

It seems μ0\mu\neq 0 is fundamentally better than the μ=0\mu=0 case in the sense that the approximation (62) is good. But as we will discuss shortly in section 2.3.1 this is not the case and the crucial point is that it is more appropriate to compare the error with the connected contributions and left out the disconnected contributions.

\bullet General qq
Next we consider general distributions. We show some details of the computation for exponential distribution and Poisson distribution in the Appendix (C). Here we only give a more abstract derivation. In terms of (585) the half wormhole (63) can be written as

Φ\displaystyle\Phi =\displaystyle= ijpqki[1]^kj[1]^kp[1]^kq[1]^+4ijpki[1]2^kj[1]^kp[1]^+2ijki[1]2^kj[1]2^\displaystyle\sum_{i\neq j\neq p\neq q}\widehat{k_{i}[1]}\widehat{k_{j}[1]}\widehat{k_{p}[1]}\widehat{k_{q}[1]}+4\sum_{i\neq j\neq p}\widehat{k_{i}[1]^{2}}\widehat{k_{j}[1]}\widehat{k_{p}[1]}+2\sum_{i\neq j}\widehat{k_{i}[1]^{2}}\widehat{k_{j}[1]^{2}} (111)
=\displaystyle= ijpqXiXjXpXq+4ijpki[1]2^XjXp+2ijki[1]2^kj[1]2^.\displaystyle\sum_{i\neq j\neq p\neq q}X_{i}X_{j}X_{p}X_{q}+4\sum_{i\neq j\neq p}\widehat{k_{i}[1]^{2}}X_{j}X_{p}+2\sum_{i\neq j}\widehat{k_{i}[1]^{2}}\widehat{k_{j}[1]^{2}}. (112)

Therefore the error of the proposal is

Error=4ijq(Xi2ki[1]2^)XjXp+2ij(Xi2Xj2ki[1]2^kj[1]2^)t~2.\displaystyle\text{Error}=4\sum_{i\neq j\neq q}(X_{i}^{2}-\widehat{k_{i}[1]^{2}})X_{j}X_{p}+2\sum_{i\neq j}(X_{i}^{2}X_{j}^{2}-\widehat{k_{i}[1]^{2}}\widehat{k_{j}[1]^{2}})-\tilde{t}^{2}. (113)

The maximal power of NN in Error2\langle\text{Error}^{2}\rangle will be 66.

When μ0\mu\neq 0, Y22N8\langle Y^{2}\rangle^{2}\sim N^{8}. So in this case the error is small and the approximation is good.

When μ=0\mu=0, Y22N4\langle Y^{2}\rangle^{2}\sim N^{4}. The terms of N4N^{4} in Error2\langle\text{Error}^{2}\rangle come from

Error2\displaystyle\langle\text{Error}^{2}\rangle =\displaystyle= ijpq{16×2(Xi2ki[1]2^)(Xj2kj[1]2^)Xp2Xq2\displaystyle\langle\sum_{i\neq j\neq p\neq q}\{16\times 2(X_{i}^{2}-\widehat{k_{i}[1]^{2}})(X_{j}^{2}-\widehat{k_{j}[1]^{2}})X_{p}^{2}X_{q}^{2} (114)
+\displaystyle+ 4(Xi2Xj2ki[1]2^kj[1]2^)(Xp2Xq2kp[1]2^kq[1]2^)}\displaystyle 4(X_{i}^{2}X_{j}^{2}-\widehat{k_{i}[1]^{2}}\widehat{k_{j}[1]^{2}})(X_{p}^{2}X_{q}^{2}-\widehat{k_{p}[1]^{2}}\widehat{k_{q}[1]^{2}})\}
+\displaystyle+ 4t16N48t4N2ij(Xi2Xj2ki[1]2^kj[1]2^)+\displaystyle 4t^{16}N^{4}-8t^{4}N^{2}\sum_{i\neq j}(X_{i}^{2}X_{j}^{2}-\widehat{k_{i}[1]^{2}}\widehat{k_{j}[1]^{2}})\rangle+\dots
=\displaystyle= N4t16(32+4+48)+#N3=32N4t16+#N3\displaystyle N^{4}t^{16}\left(32+4+4-8\right)+\#N^{3}\dots=32N^{4}t^{16}+\#N^{3}\dots (115)

which is not vanishing so the error is large and we cannot approximate Y2Y^{2} by Y2+Φ\langle Y^{2}\rangle+\Phi probably for the same reason as the q=2q=2 case. One could ask that when Xi2ki[1]2^=0\langle X_{i}^{2}-\widehat{k_{i}[1]^{2}}\rangle=0, the approximation might be fine, but it requires t2=0t^{2}=0 which we do not consider at the moment.

\bullet General distributions

Now we consider the general case (97):

Y=i1i2iqNXi1Xi2Xiq,\displaystyle Y=\sum_{i_{1}\neq i_{2}\neq\dots\neq i_{q}}^{N}X_{i_{1}}X_{i_{2}}\dots X_{i_{q}}, (116)
Y2=k=0q(q!/(qk)!)2k!j1j2jki1i2q2kXj12Xjk2Xi1Xi2q2k.\displaystyle Y^{2}=\sum_{k=0}^{q}\frac{(q!/(q-k)!)^{2}}{k!}\sum_{j_{1}\neq j_{2}\dots j_{k}\neq i_{1}\dots\neq i_{2q-2k}}X_{j_{1}}^{2}\dots X_{j_{k}}^{2}X_{i_{1}}\dots X_{i_{2q-2k}}. (117)

If NqN\gg q then the average Y2\langle Y^{2}\rangle will have the following scaling behavior in the large NN limit

Y2{N2qμ2qμ0Nqq!t2qμ=0\displaystyle\langle Y^{2}\rangle\sim\begin{cases}N^{2q}\mu^{2q}&\quad\mu\neq 0\\ N^{q}q!t^{2q}&\quad\mu=0\end{cases} (118)

Similar to (112), one can find that the half wormhole contribution Φ\Phi can be written as

Φ\displaystyle\Phi =\displaystyle= k=0q(q!/(qk)!)2k!j1j2jki1i2q2kkj1[1]2^kjk[1]2^Xi1Xi2q2k,\displaystyle\sum_{k=0}^{q}\frac{(q!/(q-k)!)^{2}}{k!}\sum_{j_{1}\neq j_{2}\dots j_{k}\neq i_{1}\dots\neq i_{2q-2k}}\widehat{k_{j_{1}}[1]^{2}}\dots\widehat{k_{j_{k}}[1]^{2}}X_{i_{1}}\dots X_{i_{2q-2k}}, (119)

so that the error is

Error =\displaystyle= k=1q(q!/(qk)!)2k!j1j2jki1i2q2k(Xj12Xjk2kj1[1]2^kjk[1]2^)Xi1Xi2q2k\displaystyle\sum_{k=1}^{q}\frac{(q!/(q-k)!)^{2}}{k!}\sum_{\begin{subarray}{c}j_{1}\neq j_{2}\dots j_{k}\neq\\ i_{1}\dots\neq i_{2q-2k}\end{subarray}}(X_{j_{1}}^{2}\dots X_{j_{k}}^{2}-\widehat{k_{j_{1}}[1]^{2}}\dots\widehat{k_{j_{k}}[1]^{2}})X_{i_{1}}\dots X_{i_{2q-2k}} (120)
\displaystyle- Y2+Y2.\displaystyle\langle Y^{2}\rangle+\langle Y\rangle^{2}.

When μ0\mu\neq 0, the leading contribution to Error2\langle\text{Error}^{2}\rangle scales as N2q2N^{2q-2} so the approximation (62) is correct.

However when μ=0\mu=0, the leading contributions to Error2\langle\text{Error}^{2}\rangle are

Error2\displaystyle\langle\text{Error}^{2}\rangle =E1+E2+#N2q1,\displaystyle=E_{1}+E_{2}+\#N^{2q-1}, (121)
E1\displaystyle E_{1} =k=1q((q!/(qk)!)2k!)2(2q2k)!\displaystyle=\langle\sum_{k=1}^{q}\left(\frac{(q!/(q-k)!)^{2}}{k!}\right)^{2}(2q-2k)! (122)
×j1j2j2ki1i2q2k(Xj12kj1[1]2^)(Xjk2kj2k[1]2^)Xi12Xi2q2k2\displaystyle\quad\times\sum_{\begin{subarray}{c}j_{1}\neq j_{2}\dots\neq j_{2k}\\ \neq i_{1}\neq\dots\neq i_{2q-2k}\end{subarray}}(X_{j_{1}}^{2}-\widehat{k_{j_{1}}[1]^{2}})\dots(X_{j_{k}}^{2}-\widehat{k_{j_{2k}}[1]^{2}})X_{i_{1}}^{2}\dots X_{i_{2q-2k}}^{2}\rangle
=N2qt4q(2q)!(F23(q,q,q;1,12q;14)1)0,\displaystyle=N^{2q}t^{4q}(2q)!\left(\,{}_{3}F_{2}\left(-q,-q,-q;1,\frac{1}{2}-q;\frac{1}{4}\right)-1\right)\neq 0, (123)
E2\displaystyle E_{2} =(q!i1i2iqN(Xi12Xiq2ki1[1]2^kiq[1]2^)q!Nq)2=0.\displaystyle=\langle\left(q!\sum_{i_{1}\neq i_{2}\neq\dots\neq i_{q}}^{N}(X_{i_{1}}^{2}\dots X_{i_{q}}^{2}-\widehat{k_{i_{1}}[1]^{2}}\dots\widehat{k_{i_{q}}[1]^{2}})-q!N^{q}\right)^{2}\rangle=0\ . (124)

So the error is large as in the previous case (115) and the approximation (62) is not good.

In our toy model (97) we did not include the “diagonal” terms while from our analysis above we have shown in the large NN limit it is the “off-diagonal” term that dominates. So our conclusions for (97) are also valid for the following general function

Y=i1,i2,,iq=1NXi1Xi2Xiq..\displaystyle Y=\sum_{i_{1},i_{2},\dots,i_{q}=1}^{N}X_{i_{1}}X_{i_{2}}\dots X_{i_{q}}.. (125)

As a simple demonstration, let us still consider the simplest case with q=2q=2:

Y=i,jXiXj,\displaystyle Y=\sum_{i,j}X_{i}X_{j}, (126)
Y2=iXi4+4ijXi3Xj+3ijXi2Xj2\displaystyle Y^{2}=\sum_{i}X_{i}^{4}+4\sum_{i\neq j}X_{i}^{3}X_{j}+3\sum_{i\neq j}X_{i}^{2}X_{j}^{2}
+6ijkXiXjXk2+ijmnXiXjXmXn.\displaystyle\qquad+6\sum_{i\neq j\neq k}X_{i}X_{j}X_{k}^{2}+\sum_{i\neq j\neq m\neq n}X_{i}X_{j}X_{m}X_{n}. (127)

Comparing

Y2=\displaystyle\langle Y^{2}\rangle= N4κ14+4N2κ3κ1+3N2κ22+6N3κ2κ12+Nκ4,\displaystyle N^{4}\kappa_{1}^{4}+4N^{2}\kappa_{3}\kappa_{1}+3N^{2}\kappa_{2}^{2}+6N^{3}\kappa_{2}\kappa_{1}^{2}+N\kappa_{4}, (131)
κ1=X=μ,κ2=X2X2=t2,\displaystyle\kappa_{1}=\langle X\rangle=\mu,\quad\kappa_{2}=\langle X^{2}\rangle-\langle X\rangle^{2}=t^{2},
κ3=X33XX2+2X3,\displaystyle\kappa_{3}=\langle X^{3}\rangle-3\langle X\rangle\langle X^{2}\rangle+2\langle X^{3}\rangle,
κ4=X43X224XX3+12X2X26X4,\displaystyle\kappa_{4}=\langle X^{4}\rangle-3\langle X^{2}\rangle^{2}-4\langle X\rangle\langle X^{3}\rangle+12\langle X\rangle^{2}\langle X^{2}\rangle-6\langle X\rangle^{4},

with (100) one find that if t0t\neq 0, the scaling behavior of Y2\langle Y^{2}\rangle is same as before. The half wormhole contribution Φ\Phi can be work out similarly:

Φ\displaystyle\Phi =\displaystyle= iki[2]2^+ijki[2]^kj[2]^+ijmnXiXjXmXn+4ijmki[1]2^XjXm\displaystyle\sum_{i}\widehat{k_{i}[2]^{2}}+\sum_{i\neq j}\widehat{k_{i}[2]}\widehat{k_{j}[2]}+\sum_{i\neq j\neq m\neq n}X_{i}X_{j}X_{m}X_{n}+4\sum_{i\neq j\neq m}\widehat{k_{i}[1]^{2}}X_{j}X_{m} (132)
+\displaystyle+ 2ijki[1]2^kj[1]2^+2ijkki[2]^XjXk+4ijki[2]ki[1]^Xj\displaystyle 2\sum_{i\neq j}\widehat{k_{i}[1]^{2}}\widehat{k_{j}[1]^{2}}+2\sum_{i\neq j\neq k}\widehat{k_{i}[2]}X_{j}X_{k}+4\sum_{i\neq j}\widehat{k_{i}[2]k_{i}[1]}X_{j}

Then the error is given by

Error =\displaystyle= 4ijk(Xi2ki[1]2^)XjXk+2ijk(Xi2ki[2]^)XjXk+2ij(Xi2Xj2ki[1]2^kj[1]2^)\displaystyle 4\sum_{i\neq j\neq k}(X_{i}^{2}-\widehat{k_{i}[1]^{2}})X_{j}X_{k}+2\sum_{i\neq j\neq k}(X_{i}^{2}-\widehat{k_{i}[2]})X_{j}X_{k}+2\sum_{i\neq j}(X_{i}^{2}X_{j}^{2}-\widehat{k_{i}[1]^{2}}\widehat{k_{j}[1]^{2}}) (133)
+\displaystyle+ ij(Xi2Xj2ki[2]^kj[2]^)+4ij(Xi3ki[2]ki[1]^)Xj+i(Xi4ki[2]2^)t~2.\displaystyle\sum_{i\neq j}(X_{i}^{2}X_{j}^{2}-\widehat{k_{i}[2]}\widehat{k_{j}[2]})+4\sum_{i\neq j}(X_{i}^{3}-\widehat{k_{i}[2]k_{i}[1]})X_{j}+\sum_{i}(X_{i}^{4}-\widehat{k_{i}[2]^{2}})\rangle-\tilde{t}^{2}.
=\displaystyle= 4ijk(Xi2ki[1]2^)XjXk+2ij(Xi2Xj2ki[1]2^kj[1]2^)\displaystyle 4\sum_{i\neq j\neq k}(X_{i}^{2}-\widehat{k_{i}[1]^{2}})X_{j}X_{k}+2\sum_{i\neq j}(X_{i}^{2}X_{j}^{2}-\widehat{k_{i}[1]^{2}}\widehat{k_{j}[1]^{2}})
+\displaystyle+ 4ij(Xi3ki[2]ki[1]^)Xj+i(Xi4ki[2]2^)t~2,\displaystyle 4\sum_{i\neq j}(X_{i}^{3}-\widehat{k_{i}[2]k_{i}[1]})X_{j}+\sum_{i}(X_{i}^{4}-\widehat{k_{i}[2]^{2}})\rangle-\tilde{t}^{2},

where we have used the identity

k[2]^=dkeikXP(X)x2eikx=X2.\displaystyle\widehat{k[2]}=\int\text{d}k\frac{e^{-\text{i}kX}}{P(X)}\langle x^{2}e^{\text{i}kx}\rangle=X^{2}. (134)

Comparing with (113), there are two extra terms in (133), but they will never contribute333If μ0\mu\neq 0 they maximally contribute to N5N^{5} and when μ=0\mu=0 they maximally contribute to N3N^{3}. to the leading power of NN when t0t\neq 0. So again it seems the approximation (62) is good when μ0\mu\neq 0 but not good when μ=0\mu=0. We will explain in the next section how to understand these results and modify the proposal (62).

2.3 Large-NN constraints and half-wormhole approximation

In the previous sections we consider a few different examples. To summarize, the half-wormhole conjecture (62) and (63) is valid for a large families of statistical models. However, for some examples discussed in section 2.2.3 this approximation is not good.

2.3.1 Why and how to modify the approximation proposal

The failed examples indicate that the proposed Φ\Phi does not capture all semi-classical components in the observable Y2Y^{2} to be approximated.

As discussed previously, the approximation (62) should come from the approximation (10). The relation (10) indeed fails for the case where the approximation (62) is not good in section (2.2.3). To see this explicitly, we consider the simplest example (98) where

Y2\displaystyle Y^{2} =ijpqXiXjXpXq+4ipqXi2XpXq+2ijXi2Xj2,\displaystyle=\sum_{i\neq j\neq p\neq q}X_{i}X_{j}X_{p}X_{q}+4\sum_{i\neq p\neq q}X_{i}^{2}X_{p}X_{q}+2\sum_{i\neq j}X_{i}^{2}X_{j}^{2}\,, (135)

which means we need to consider the following terms in the approximation Φ\Phi

XiXjXpXqeiakaXa,Xj2XpXqeiakaXa,,Xi2Xj2eiakaXa.\displaystyle\langle X_{i}X_{j}X_{p}X_{q}e^{i\sum_{a}k_{a}X_{a}}\rangle\,,\qquad\langle X_{j}^{2}X_{p}X_{q}e^{i\sum_{a}k_{a}X_{a}}\rangle,,\qquad\langle X_{i}^{2}X_{j}^{2}e^{i\sum_{a}k_{a}X_{a}}\rangle\ . (136)

However, in the proposal (62) the Φ\Phi term contains only Yeikaxa2\langle Ye^{ik_{a}x_{a}}\rangle^{2}, which means only terms like

XiXjeiakaXaXpXqeiakaXa,ij,pq,\displaystyle\langle X_{i}X_{j}e^{i\sum_{a}k_{a}X_{a}}\rangle\langle X_{p}X_{q}e^{i\sum_{a}k_{a}X_{a}}\rangle\,,\qquad i\neq j\,,p\neq q\,, (137)

contribute. Therefore to check why the proposal (62) that fails, we want to understand what is “missing” in (137) comparing with the correct answer involving (136).

Because the xix_{i}’s are identical independent random variables, the cumulant cnc_{n} for each xix_{i} are the same and the moment generating function is just a product of the moment generating functions of each xix_{i}. Therefore we can reduce the problem of finding a good approximation of the above product terms to each flavor of xix_{i} and find the approximation for each of them. This should give a good approximation for each term. 444Although this would obscure the interpretation of YY as an independent function, we still choose to proceed this way in order to check how the approximation (62) fails.

Recall the approximation is to replace Xineikxceikx\frac{\langle X_{i}^{n}e^{ikx}\rangle_{c}}{\langle e^{ikx}\rangle} by (Xieikxceikx)n\left(\frac{\langle X_{i}e^{ikx}\rangle_{c}}{\langle e^{ikx}\rangle}\right)^{n} for n>1n>1, ie (32), therefore only the last two terms in (136) are affected by the approximation. In particular, the first term in (136) gives the same contribution as the term (137) that leads to the inaccurate approximation (62). So the non-vanishing contributions from the last two terms in the leading order of 1/N1/N should then be responsible for the failure of the approximation (62) in this example. As discussed above, a good approximation to the xjx_{j} factor of Xj2XpXqeiakaXa\langle X_{j}^{2}X_{p}X_{q}e^{i\sum_{a}k_{a}X_{a}}\rangle should be

Xj2eikjXjeikjXjXj2+2XjXjeikjXjceikjXj+(XjeikjXjceikjXj)2.\displaystyle\frac{\langle X_{j}^{2}e^{ik_{j}X_{j}}\rangle}{\langle e^{ik_{j}X_{j}}\rangle}\approx\langle X_{j}^{2}\rangle+2\langle X_{j}\rangle\frac{\langle X_{j}e^{ik_{j}X_{j}}\rangle_{c}}{\langle e^{ik_{j}X_{j}}\rangle}+\left(\frac{\langle X_{j}e^{ik_{j}X_{j}}\rangle_{c}}{\langle e^{ik_{j}X_{j}}\rangle}\right)^{2}\ . (138)

The contribution to the half-wormhole Φ\Phi from this term Xj2XpXqeiakaXa\langle X_{j}^{2}X_{p}X_{q}e^{i\sum_{a}k_{a}X_{a}}\rangle is thus

(t2+μ2)XpXq+2μXjXpXq+Φ2jXpXq.\displaystyle(t^{2}+\mu^{2})X_{p}X_{q}+2\mu X_{j}X_{p}X_{q}+\Phi_{2}^{j}X_{p}X_{q}\ . (139)

Similarly, the Xp2Xq2eiakaXa\langle X_{p}^{2}X_{q}^{2}e^{i\sum_{a}k_{a}X_{a}}\rangle type terms gives a contribution

(t2+μ2)2+4μ2XpXq+Φ2pΦ2q+4μ(t2+μ2)(Xp+Xq)\displaystyle(t^{2}+\mu^{2})^{2}+4\mu^{2}X_{p}X_{q}+\Phi_{2}^{p}\Phi_{2}^{q}+4\mu(t^{2}+\mu^{2})\left(X_{p}+X_{q}\right)
+(t2+μ2)(Φ2p+Φ2q)+4μ(XpΦ2q+XqΦ2p).\displaystyle\quad+(t^{2}+\mu^{2})\left(\Phi_{2}^{p}+\Phi_{2}^{q}\right)+4\mu\left(X_{p}\Phi_{2}^{q}+X_{q}\Phi_{2}^{p}\right)\ . (140)

Now we should sum over j,p,qj,p,q to get all the contributions to the computation of Y2\langle Y^{2}\rangle and further to Error2.

To understand the structure of the contribution to Error2, we denote

Error=(Y2Φ+Y2)Y2=MY2,M=Y2,\displaystyle\text{Error}=(Y^{2}-\Phi+\langle Y\rangle^{2})-\langle Y^{2}\rangle=M-\langle Y^{2}\rangle\,,\qquad\langle M\rangle=\langle Y^{2}\rangle\,, (141)

then if we switch the notation of Error2\langle\text{Error}^{2}\rangle to a slightly more indicative one Error1Error2\langle\text{Error}_{1}\text{Error}_{2}\rangle, we have

Error1Error2=M1M2+Y222Y2M=M1M2Y22.\displaystyle\langle\text{Error}_{1}\text{Error}_{2}\rangle=\langle M_{1}M_{2}\rangle+\langle Y^{2}\rangle^{2}-2\langle Y^{2}\rangle\langle M\rangle=\langle M_{1}M_{2}\rangle-\langle Y^{2}\rangle^{2}\ . (142)

Therefore we find if M1M2M1M2\langle M_{1}M_{2}\rangle\approx\langle M_{1}\rangle\langle M_{2}\rangle to the leading order, then the error is small and the approximation (62) is good.

This is precisely how the previous proposal (62) failed. For example, in the Error (104), it is precisely the contraction among the two factors ijXiXj\sum_{i\neq j}X_{i}X_{j} that gives another factor of 2N4t82N^{4}t^{8} in (Error/4)2\langle(\text{Error}/4)^{2}\rangle and prevent it from vanishing. On the other hand, if we check the results (139) and (140) we find, to the leading order of NN, the term that have non-trivial contribution to Error2 is

4(N2)(t2+μ2)XpXq,\displaystyle 4(N-2)(t^{2}+\mu^{2})X_{p}X_{q}\,, (143)

that comes from summing the first terms in (139) over jj; the other terms are either suppressed by 1/N1/N or do not give nontrivial contraction between the two copies of Error as discussed above. Then we immediately notice that this is precisely the term, with μ=0\mu=0 in this case, that is missing in Φ\Phi to remove the “problematic” term in the Error that we just discussed. Therefore, once we use the correct approximation with all terms in (136), the error should be small and the approximation should be good. The other examples in section (2.2.3) could also be modified in a similar way so that the errors become small.

Further notice that one of the upshot of the approximation (62) is, as pointed out in Mukhametzhanov:2021hdi , that we can safely ignore the direct correlation between the two YY’s (or zz’s in the context of Mukhametzhanov:2021hdi ) and the two terms are “linked” through the correlation with eikxe^{ikx}. What we found in the previous section are however cases where these direct correlations cannot be ignored. The new ingredient of the approximation (145) we will present shortly is precisely a partial correlation between the YY’s directly, not just through the eikxe^{ikx} factors. In this sense the saddles in the general models discussed in section 2.2.3 are hyper-linked half-wormholes with extra partially direct connections.

With this we propose a modified approximation

Y2Y2+Φ~,\displaystyle Y^{2}\approx\langle Y^{2}\rangle+\tilde{\Phi}, (144)
Φ~(X)=1(2π)Ni(dkieikiXiP(Xi))eiikixi[Y(x)2eiikixi]eiikixi,\displaystyle\tilde{\Phi}(X)=\frac{1}{(2\pi)^{N}}\int\prod_{i}\left(dk_{i}\frac{e^{-\text{i}k_{i}X_{i}}}{P(X_{i})}\right)\langle e^{\text{i}\sum_{i}k_{i}x_{i}}\rangle\frac{\left[Y(x)^{2}e^{\text{i}\sum_{i}k_{i}x_{i}}\right]}{\langle e^{\text{i}\sum_{i}k_{i}x_{i}}\rangle}\,, (145)

where [Y(x)2eiikixi]\left[Y(x)^{2}e^{\text{i}\sum_{i}k_{i}x_{i}}\right] denotes all possible terms contains at least one contraction between Y2Y^{2} and the spacetime brane eikxe^{ikx}.

In the example (98), each term in YY contains two XiX_{i} legs, therefore we have

[Y(x)2eiikixi]eiikixi=2YY(x)eiikixiceiikixi+Y(x)eiikixic2eiikixi2+Y(x)2eiikixiceiikixi,\displaystyle\frac{\left[Y(x)^{2}e^{\text{i}\sum_{i}k_{i}x_{i}}\right]}{\langle e^{\text{i}\sum_{i}k_{i}x_{i}}\rangle}=\frac{2\langle Y\rangle\langle Y(x)e^{\text{i}\sum_{i}k_{i}x_{i}}\rangle_{c}}{\langle e^{\text{i}\sum_{i}k_{i}x_{i}}\rangle}+\frac{\langle Y(x)e^{\text{i}\sum_{i}k_{i}x_{i}}\rangle_{c}^{2}}{\langle e^{\text{i}\sum_{i}k_{i}x_{i}}\rangle^{2}}+\frac{\langle Y(x)^{2}e^{\text{i}\sum_{i}k_{i}x_{i}}\rangle_{c}}{\langle e^{\text{i}\sum_{i}k_{i}x_{i}}\rangle}\,, (146)

where the different terms correspond to one contraction to the brane, two separate contractions to the brane and a pair of connected contractions to the brane. The c here means the contribution cannot be made disconnected if we only cut on the brane. Among these terms the last one is precisely the one missed in the previous proposal (63). A demonstration of these terms are shown in Figure 2. We notice that this approximation is closely related to the relation between Z^\hat{Z} and W^\hat{W} discussed in Peng:2021vhs , see e.g. Figure. 9 there.

Refer to caption
Figure 2: A pictorial illustration of the 3 terms in (146) respectively. Each vertex on the left is a factor of Y(x)Y(x), brane on the right denotes eikxe^{ikx}, and each bracket eikxc\langle\cdot e^{ikx}\rangle_{c} corresponds to a component of the bulk amplitude that connects the brane with a set of vertices. Notice that the first diagrams should be considered as two diagrams each has a vertex connected to the brane.

From this analysis, it is more obvious to understand why the Errors are all small when μ0\mu\neq 0 in section 2.2.3. When disconnected contributions exist, the leading order contributions of the Error2 always come from the disconnected component and hence the Error is guaranteed to be small. However, this is not very meaningful since as in most of the large-NN theories studied in the literature, we isolate away the disconnected contributions and always focus on the connected contributions.

2.3.2 Why the proposal works for the Pfaffian in the SYK model

From the above discussion, it seems that for a generic operators with complicated product structure, the original proposal (62) almost surely fails. However, we know from explicit computations in Saad:2021rcu ; Mukhametzhanov:2021nea that the approximation works well for the hyperpfaffian of the random couplings which is also related to the partition function of the SYK model.

We believe the reason for this is the large-NN factorizations properties due to large-NN constraints. By this we mean when the operators are defined to have extra structures, for example as a trace or a determinant over the NN flavors, such extra structure remains to affect the computation of the Error. When this is true, which indeed is our case, then the contractions between the two copies of Error are necessarily suppressed by the large-NN factors; either 1/N1/N when the structure is trace as in (77) or higher powers of 1/N1/N when the structure is a determinant. Therefore all contractions between the two copies of Errors are suppressed and at the leading order the result factorizes and hence the original proposal (62) works. 555A related fact is that when the approximation is no longer good the relation between the 4th{}^{\text{th}} moment Y4\langle Y^{4}\rangle of the observable (98) and the second moment Y2\langle Y^{2}\rangle deviates significantly from the Gaussian distribution. In Gaussian distribution, this contribution is 3Y2Y43\langle Y^{2}\rangle\subset\langle Y^{4}\rangle, on the other hand, for the observable YY in (98) we get Y4\displaystyle\langle Y^{4}\rangle =\displaystyle= 8ijXi4Xj4+60ijpqXi2Xj2Xp2Xq2+48ijpXi4Xj2Xp2\displaystyle 8\sum_{i\neq j}\langle X_{i}^{4}X_{j}^{4}\rangle+60\sum_{i\neq j\neq p\neq q}\langle X_{i}^{2}X_{j}^{2}X_{p}^{2}X_{q}^{2}\rangle+48\sum_{i\neq j\neq p}\langle X_{i}^{4}X_{j}^{2}X_{p}^{2}\rangle (147) \displaystyle\approx 60N4t83Y222Y412N4t8.\displaystyle 60N^{4}t^{8}\neq 3\langle Y^{2}\rangle^{2}-2\langle Y\rangle^{4}\approx 12N^{4}t^{8}\ . (148) But at the moment we have not succeeded in making a causal relation between this fact and the fact that the Error is small. The explanation in the main text does better in doing so.

A somewhat ad hoc reason for the need of traces or determinant in the definition of the operator to make the discussion about (half-)wormhole meaningful is the following. There is no “spacetime” in our statistical models, so we cannot use any locality property to identify a function of the random variables as a single operator; the most we can do is to use a trace or determinant structure to identify a group of random variables as an operator. If there is no such trace/determinant constraints, it is equally legitimate to regard the result as computing correlations of a large number of the fundamental random variables and the (half-)wormhole interpretation is not necessarily relevant.

A different interpretation of the importance of the existence of such trace or determinant structure could be considered as some emergent global symmetry among the random variables (probably when appropriately analytically continued). By this we simply mean if we treat the random variables XiX_{i} as “fields”, then the action, ie the probability distribution, and the operators we considered in the computation all have SO(N)SO(N) symmetry among them. Then the invariant tensors of SO(N)SO(N) directly lead to the trace or determinant structures we just described. It is interesting to make this point more clear, and we plan to come back to this question somewhere else.

We did not find a general proof of the above assertion (145) or (146), but as a check we can, according to our assertion, modify the definition of the function YY and put in by hand some constraints, mimicking a trace structure. Then we find with this constraints the approximation (62) is indeed valid. For instance we could introduce a restriction in the sum

Y=i+j=MXiXj,N<M<2N,ij,\displaystyle Y=\sum_{i+j=M}X_{i}X_{j},\quad N<M<2N,\quad i\neq j\,, (149)

where NN is the total number of XX’s and MM is an integer. Without loss of generality we assume MM is even in the following, and the computation for odd MM is the same. Following the previous computations, we get

Y2\displaystyle Y^{2} =2i+j=MXi2Xj2+ijmnXiXjXmXn,\displaystyle=2\sum_{i+j=M}X_{i}^{2}X_{j}^{2}+\sum_{i\neq j\neq m\neq n}X_{i}X_{j}X_{m}X_{n}\,, (150)

and

Y=Kμ2,Y2=2K(t2+μ2)2+K(K2)μ4,K=2NM.\displaystyle\langle Y\rangle=K\mu^{2},\quad\langle Y^{2}\rangle=2K(t^{2}+\mu^{2})^{2}+K(K-2)\mu^{4},\quad K=2N-M\ . (151)

Taking XiX_{i} from the same Gaussian distribution in the previous cases we get the expression for the error

Error=4t2i+j=MXi24Kt44Kt2μ2.\displaystyle\text{Error}=4t^{2}\sum_{i+j=M}X_{i}^{2}-4Kt^{4}-4Kt^{2}\mu^{2}. (152)

It is straightforwardly to show that the expectation values

Error=0,(Error/4)2=2Kt8+4Kt6μ2.\displaystyle\langle\text{Error}\rangle=0\,,\qquad\langle(\text{Error}/4)^{2}\rangle=2Kt^{8}+4Kt^{6}\mu^{2}\ . (153)

Clearly in this case (Error/4)2\langle(\text{Error}/4)^{2}\rangle is 1/N1/N suppressed compared to Y22\langle Y^{2}\rangle^{2} independent on the value of μ\mu. Hence the approximation (62) is always valid in the presence of this extra constraint. Similar restrictions could be imposed to models with general qq. It turns out that again the computation is quite similar and we expect the approximation to be valid in these cases too.

3 SYK at one time point: Ja=0\langle J_{a}\rangle=0

In this section, we study the half-wormhole contributions in some 0d SYK model that can be considered as the usual 0+1d SYK model on a single instant of time. This section is largely a review of previous results in Saad:2021rcu ; Mukhametzhanov:2021nea ; Mukhametzhanov:2021hdi ; we provide more details of various saddle point results and carry out Lefschetz thimble analysis of some computations when needed.

3.1 SYK model with one time point

Let us first revisit the analysis of the 0-dimensional SYK model introduced in Saad:2021rcu . We are interested in the following Grassmann integral

z=dNψexp(iq/2Ji1iqψi1iq),\displaystyle z=\int d^{N}\psi\exp(\text{i}^{q/2}\sum J_{i_{1}\dots i_{q}}\psi_{i_{1}\dots i_{q}})\,, (154)

where ψi1iq=ψa1ψa2ψaq\psi_{i_{1}\dots i_{q}}=\psi_{a_{1}}\psi_{a_{2}}\dots\psi_{a_{q}} and ψi\psi_{i} are Grassmann numbers. The number zz can be understood as the partition function of 0+00+0 dimensional analogue of SYK model. The random couplings Ji1iqJ_{i_{1}\dots i_{q}} is drawn from a Gaussian distribution

Ji1iq=0,Ji1iqJj1jq=t2δi1j1δiqjq,t2=(q1)!Nq1.\displaystyle\langle J_{i_{1}\dots i_{q}}\rangle=0,\quad\langle J_{i_{1}\dots i_{q}}J_{j_{1}\dots j_{q}}\rangle=t^{2}\delta_{i_{1}j_{1}}\dots\delta_{i_{q}j_{q}},\quad t^{2}=\frac{(q-1)!}{N^{q-1}}\ . (155)

We sometimes use the collective indies A,BA,B to simplify the notation

A={a1<<aq},JAψAJa1aqψa1aq.\displaystyle A=\{a_{1}<\dots<a_{q}\}\,,\qquad J_{A}\psi_{A}\equiv J_{a_{1}\dots a_{q}}\psi_{a_{1}\dots a_{q}}\ . (156)

Integrating out the Grassmann numbers directly gives (96)666Here we choose the measure of Grassmann integral to be dNψψ1N=iN/2\int d^{N}\psi\psi_{1\dots N}=\text{i}^{-N/2}.:

z=dNψexp(iq/2JAψA)=A1<<Apsgn(A)JA1JAp,p=N/q,\displaystyle z=\int d^{N}\psi\exp(\text{i}^{q/2}J_{A}\psi_{A})=\sum^{\prime}_{A_{1}<\dots<A_{p}}\text{sgn}(A)J_{A_{1}}\dots J_{A_{p}}\,,\quad p=N/q\,, (157)

where the expression (157) is nothing but the hyperpfaffian Pf(J)\text{Pf}(J). Since z=0\langle z\rangle=0 due to (155), we focus on z2z^{2} and z2\langle z^{2}\rangle

z2=zLzR=dNψLdNψRexp{iq/2AJA(ψAL+ψAR)},\displaystyle z^{2}=z_{L}z_{R}=\int\text{d}^{N}\psi^{L}\text{d}^{N}\psi^{R}\exp\left\{\text{i}^{q/2}\sum_{A}J_{A}\left(\psi_{A}^{L}+\psi_{A}^{R}\right)\right\}\,, (158)
z2=d2Nψexp{Nq(1Ni=1NψiLψiR)q},\displaystyle\langle z^{2}\rangle=\int\text{d}^{2N}\psi\exp\left\{\frac{N}{q}\left(\frac{1}{N}\sum_{i=1}^{N}\psi_{i}^{L}\psi_{i}^{R}\right)^{q}\right\}\,, (159)

where we have assumed that qq and NN are even. The exact values of (159) can be computed by introducing the standard G,ΣG,\Sigma variables

z2\displaystyle\langle z^{2}\rangle =\displaystyle= d2NψdGδ(G1Ni=1NψiLψiR)exp(NqGq)\displaystyle\int\text{d}^{2N}\psi\int_{\mathbb{R}}\text{d}G\delta\left(G-\frac{1}{N}\sum_{i=1}^{N}\psi_{i}^{L}\psi_{i}^{R}\right)\exp\left(\frac{N}{q}G^{q}\right) (160)
=\displaystyle= dGidΣ2πi/Nexp{N(log(Σ)ΣG+1qGq)}\displaystyle\int_{\mathbb{R}}\text{d}G\int_{\text{i}\mathbb{R}}\frac{\text{d}\Sigma}{2\pi\text{i}/N}\exp\left\{N\left(\log(\Sigma)-\Sigma G+\frac{1}{q}G^{q}\right)\right\} (161)
=\displaystyle= NNdGexp(NqGq)(G)Nδ(G)\displaystyle N^{-N}\int_{\mathbb{R}}\text{d}G\exp\left(\frac{N}{q}G^{q}\right)(-\partial_{G})^{N}\delta(G) (162)
=\displaystyle= N!(N/q)N/qNN(N/q)!=e(11q)Nq(1+1q12N+𝒪(1N2)),\displaystyle\frac{N!(N/q)^{N/q}}{N^{N}(N/q)!}=e^{-(1-\frac{1}{q})N}\sqrt{q}\left(1+\frac{1-q}{12N}+\mathcal{O}(\frac{1}{N^{2}})\right)\,, (163)

where in the last step we expand around NN\to\infty to the next-to-leading order.

Next we consider the non-averaged quantity (158). Following Saad:2021rcu , we rewrite

z2=RdσΨ(σ)Φ(σ),Ψ(σ)=dg2π/Nexp[N(iσg1/qgq)],\displaystyle z^{2}=\int_{R}\text{d}\sigma\Psi(\sigma)\Phi(\sigma)\,,\quad\Psi(\sigma)=\int\frac{dg}{2\pi/N}\exp[N(-\text{i}\sigma g-1/qg^{q})]\,, (164)

where the coupling dependent piece Φ\Phi is

Φ(σ)=d2Nψexp{ieiπqσψiLψiR+iq/2JA(ψAL+ψAR)Nq(1NψiLψiR)q}.\displaystyle\Phi(\sigma)=\int\text{d}^{2N}\psi\exp\left\{\text{i}e^{-\frac{\text{i}\pi}{q}}\sigma\psi_{i}^{L}\psi_{i}^{R}+\text{i}^{q/2}J_{A}(\psi_{A}^{L}+\psi_{A}^{R})-\frac{N}{q}\left(\frac{1}{N}\psi_{i}^{L}\psi_{i}^{R}\right)^{q}\right\}\,. (165)

Its averaged value is

Φ(σ)=(ieiπqσ)N.\displaystyle\langle\Phi(\sigma)\rangle=(\text{i}e^{-\frac{\text{i}\pi}{q}}\sigma)^{N}\ . (166)

As suggested in Saad:2021rcu , to understand the relation between each individual result and the averaged result, we could figure out in what region of the σ\sigma-plane Φ\Phi is self-averaging. This is reflected in the quantity (Φ(σ)Φ(σ))2\langle\left(\Phi(\sigma)-\langle\Phi(\sigma)\rangle\right)^{2}\rangle. Therefore we compare Φ(σ)2\langle\Phi(\sigma)\rangle^{2} with Φ(σ)2\langle\Phi(\sigma)^{2}\rangle

Φ(σ)2=Rd4σABd4gAB(2π/N)4eN[log(e2iπq(σ2+σ14σ23σ13σ24))iσABgAB1qgABq],\displaystyle\langle\Phi(\sigma)^{2}\rangle=\int_{R}\frac{\text{d}^{4}\sigma_{AB}\text{d}^{4}g_{AB}}{(2\pi/N)^{4}}e^{N\left[\log(-e^{-\frac{2\text{i}\pi}{q}}(\sigma^{2}+\sigma_{14}\sigma_{23}-\sigma_{13}\sigma_{24}))-\text{i}\sigma_{AB}g_{AB}-\frac{1}{q}g_{AB}^{q}\right]}\,, (167)

where we relabel L=1,L=3,R=2,R=4L=1,L^{\prime}=3,R=2,R^{\prime}=4 and (AB)=(13),(14),(23),(24)(AB)=(13),(14),(23),(24). The integral can be done exactly Saad:2021rcu following a similar computation we used to get (163)

Φ(σ)2=(e2iπq)Nn1+n2+n3=Nq,ni0N!N2q(n2+n3)(Nq)2(n2+n3)σ2qn1(qn2)!(qn3)!(qn1)!(n2!)2(n3!)2,\displaystyle\langle\Phi(\sigma)^{2}\rangle=(-e^{-\frac{2\text{i}\pi}{q}})^{N}\sum_{n_{1}+n_{2}+n_{3}=\frac{N}{q},n_{i}\geq 0}\frac{N!}{N^{2q(n_{2}+n_{3})}}\left(\frac{N}{q}\right)^{2(n_{2}+n_{3})}\frac{\sigma^{2qn_{1}}(qn_{2})!(qn_{3})!}{(qn_{1})!(n_{2}!)^{2}(n_{3}!)^{2}}\,, (168)

which can be organized into a polynomial in σ\sigma

Φ(σ)2\displaystyle\langle\Phi(\sigma)^{2}\rangle =\displaystyle= (e2iπq)N(σ2N+2N!q!(Nq)!q2N2q2σ2N2q++e2N1qq2q)\displaystyle(-e^{-\frac{2\text{i}\pi}{q}})^{N}\left(\sigma^{2N}+\frac{2N!q!}{(N-q)!q^{2}N^{2q-2}}\sigma^{2N-2q}+\dots+e^{2N\frac{1-q}{q}}2q\right) (169)
\displaystyle\sim (e2iπq)N(σ2N+2(q1)!qNq2σ2N2q++e2N1qq2q),\displaystyle(-e^{-\frac{2\text{i}\pi}{q}})^{N}\left(\sigma^{2N}+\frac{2(q-1)!}{qN^{q-2}}\sigma^{2N-2q}+\dots+e^{2N\frac{1-q}{q}}2q\right)\,, (170)

where the phase factor is trivial whenever qq divides NN.

3.2 The saddle points analysis

The above results can be reproduced by saddle point approximation in large NN limit.

3.2.1 The averaged z2\langle z^{2}\rangle

To obtain the same result (163) from saddle point approximation, we first we rotate the contour

Σ=ieiπqσ,G=eiπqg,\displaystyle\Sigma=\text{i}e^{-\text{i}\frac{\pi}{q}}\sigma,\quad G=e^{\text{i}\frac{\pi}{q}}g\,, (171)

to get

z2=RdgRdσ2π/Nexp{N(log(ieiπqσ)iσg1qgq)}RdgRdσ2π/NeNS,\displaystyle\langle z^{2}\rangle=\int_{R}\text{d}g\int_{R}\frac{\text{d}\sigma}{2\pi/N}\exp\left\{N\left(\log(\text{i}e^{-\frac{\text{i}\pi}{q}}\sigma)-\text{i}\sigma g-\frac{1}{q}g^{q}\right)\right\}\equiv\int_{R}\text{d}g\int_{R}\frac{\text{d}\sigma}{2\pi/N}e^{NS}\,, (172)

so that the integral converges. The saddle point equations are

iσgq1=0,gq=1,g=e(2m+1)iπq,m=0,,q1.\displaystyle-\text{i}\sigma-g^{q-1}=0\,,\quad g^{q}=-1\,,\quad\rightarrow\quad g=e^{\frac{(2m+1)\text{i}\pi}{q}}\,,\quad m=0,\dots,q-1\ . (173)

All of them give the same on-shell action

z2s=N2πe(11q)N.\displaystyle\langle z^{2}\rangle_{s}=\frac{N}{2\pi}e^{-(1-\frac{1}{q})N}\ . (174)

To match with the exact result (163) we need to consider fluctuations around the saddle points. For simplicity let us take q=4q=4 and focus on one of the saddle points

σs=gs=(1)34,z2s=N2πe34N.\displaystyle\sigma_{s}=g_{s}=-(-1)^{\frac{3}{4}},\quad\langle z^{2}\rangle_{s}=\frac{N}{2\pi}e^{-\frac{3}{4}N}. (175)

Expanding the exponent around this saddle

σ=σs+x,g=gs+y\displaystyle\sigma=\sigma_{s}+x,\quad g=g_{s}+y (176)

to the second order

S234+3ix22ixyiy22+[(1)3/4x3+(1)3/43y3]+y4x44,\displaystyle S_{2}\sim-\frac{3}{4}+\frac{3\text{i}x^{2}}{2}-\text{i}xy-\frac{\text{i}y^{2}}{2}+[(-1)^{3/4}x^{3}+\frac{(-1)^{3/4}}{3}y^{3}]+\frac{y^{4}-x^{4}}{4}\,, (177)

and evaluating the integral directly gives the fluctuation that combines with the saddle contribution to

z2saddle+loop=e34N12(114N).\displaystyle\langle z^{2}\rangle_{\text{saddle}+\text{loop}}=e^{-\frac{3}{4}N}\frac{1}{2}\left(1-\frac{1}{4N}\right)\ . (178)

Adding contributions from all 4 saddles we arrive at

z2saddle+loop=2e34N(114N),\displaystyle\langle z^{2}\rangle_{\text{saddle}+\text{loop}}=2e^{-\frac{3}{4}N}\left(1-\frac{1}{4N}\right)\,, (179)

that agrees with (163) at the two-loop order.

3.2.2 The unaveraged z2z^{2}: the wormhole saddle

The result (170) can be reproduced from a saddle point analysis in the large-NN limit. The saddle point equations are

gABq1=iσAB,ig13=σ24f,ig14=σ23f,ig23=σ14f,ig24=σ13f,\displaystyle g_{AB}^{q-1}=-\text{i}\sigma_{AB}\,,\quad-\text{i}g_{13}=\frac{\sigma_{24}}{f},\quad\text{i}g_{14}=\frac{\sigma_{23}}{f},\quad\text{i}g_{23}=\frac{\sigma_{14}}{f},\quad-\text{i}g_{24}=\frac{\sigma_{13}}{f}\,, (180)

where fσ14σ23σ13σ24+σ2f\equiv\sigma_{14}\sigma_{23}-\sigma_{13}\sigma_{24}+\sigma^{2}. The trivial solution σAB=gAB=0\sigma_{AB}=g_{AB}=0 leads to

Φ(σ)2trivial+1loop=Φ(σ)2,\displaystyle\langle\Phi(\sigma)^{2}\rangle_{\text{trivial}+1\text{loop}}=\langle\Phi(\sigma)\rangle^{2}\,, (181)

which says the trivial saddle always agrees with the first term in (170).

Next let us consider non-trivial solutions with σAB0\sigma_{AB}\neq 0. From the equations of motion we obtain

xq2=yq2,(xq1yq1+σ2)2=xq2=yq2,\displaystyle x^{q-2}=y^{q-2},\quad(x^{q-1}-y^{q-1}+\sigma^{2})^{2}=x^{q-2}=y^{q-2}\,, (182)
g13q=g24q,g23q=g14q\displaystyle g_{13}^{q}=g_{24}^{q},\quad g_{23}^{q}=g_{14}^{q} (183)

where

x=g13g24,y=g14g23.\displaystyle x=g_{13}g_{24},\quad y=g_{14}g_{23}\ . (184)

It is easy to check that solutions of the above equation satisfies x=ye2mπiq2x=ye^{\frac{2m\pi\text{i}}{q-2}}, and for each choice of mm there are 2q22q^{2} solutions of gabg_{ab}. For simplicity let us again focus on the q=4q=4 case such that there are only two classes x=±yx=\pm y.

\bullet When x=yx=y we find another 32 non-trivial saddles. The on-shell action of all of them are the same

Φ(σ)2non-trivial+=N4Φ(σ)2=Φ(σ)2trivial,\displaystyle\langle\Phi(\sigma)^{2}\rangle_{\text{non-trivial}}^{+}=N^{4}\langle\Phi(\sigma)\rangle^{2}=\langle\Phi(\sigma)^{2}\rangle_{\text{trivial}}\,, (185)

where the factor N4N^{4} comes from the measure of (167). However the 1-loop fluctuations around them are different

trivial saddle:1N4,non-trivial saddles:18N4.\displaystyle\text{trivial saddle}:\frac{1}{N^{4}}\,,\quad\text{non-trivial saddles}:\frac{1}{8N^{4}}\ . (186)

We notice that including the 1-loop effect, the trivial saddle is larger and it reproduces the large NN behavior of the exact result. On the other hand, the non-trivial saddle contributions are also comparable; so it is possible that we should also take into account of their contributions as well. However, if we add all the trivial and non-trivial saddle-point values, the result will obviously exceed the exact value (170). In fact, by a simple Lefschetz-thimble analysis, see e.g. Witten:2010cx , which is reviewed In Appendix E, we conclude that these non-trivial saddles should not be included.

Refer to caption
Refer to caption
Figure 3: Anti-thimble on the σ13\sigma_{13} plane (left) and the σ24\sigma_{24} plane (right).

In particular, we choose a Morse function to be the real part of the action (167)

h(S)=\displaystyle h\equiv\Re(S)= abj(gabj44+3gab12gab222+gab1σab2+gab2σab1)\displaystyle\sum_{abj}\left(-\frac{g_{abj}^{4}}{4}+\frac{3g_{ab1}^{2}g_{ab2}^{2}}{2}+g_{ab1}\sigma_{ab2}+g_{ab2}\sigma_{ab1}\right)
+12log((σ142σ231+σ141σ232σ132σ241σ131σ242)2\displaystyle\quad+\frac{1}{2}\log\left((\sigma_{142}\sigma_{231}+\sigma_{141}\sigma_{232}-\sigma_{132}\sigma_{241}-\sigma_{131}\sigma_{242})^{2}\right.
+(1+σ141σ231σ142σ232σ131σ241+σ132σ242)2),\displaystyle\left.\quad+(1+\sigma_{141}\sigma_{231}-\sigma_{142}\sigma_{232}-\sigma_{131}\sigma_{241}+\sigma_{132}\sigma_{242})^{2}\right)\,, (187)

where we have chosen q=4q=4 for simplicity and σ=1\sigma=1 since we are interested in the case σ0\sigma\neq 0777The σ=0\sigma=0 case is analyzed in Saad:2021rcu . The gabig_{abi} and σabj\sigma_{abj} are the real and imaginary parts of the field gabg_{ab} and σab\sigma_{ab}

gab=gab1+igab2,σab=σab1+iσab2.\displaystyle g_{ab}=g_{ab1}+\text{i}g_{ab2},\quad\sigma_{ab}=\sigma_{ab1}+\text{i}\sigma_{ab2}\ . (188)

The downward flow equations of the Morse function are

dgabjdt=hgabj,dσabjdt=hσabj.\displaystyle\frac{dg_{abj}}{dt}=-\frac{\partial h}{\partial g_{abj}},\quad\frac{d\sigma_{abj}}{dt}=-\frac{\partial h}{\partial\sigma_{abj}}\ . (189)

The end point of each anti-thimble is one of the saddles at gabjcg_{abj}^{c} and gabjcg_{abj}^{c}, which leads to the following boundary conditions of the flow equation

limt+gabj=gabjc,limt+σabj=σabjc.\displaystyle\lim_{t\to+\infty}g_{abj}=g_{abj}^{c},\quad\lim_{t\to+\infty}\sigma_{abj}=\sigma_{abj}^{c}\ . (190)

We can then solve the flow equation and obtain the Lefschetz anti-thimbles going through each saddle point and if they intersect with the original integration contour the saddle point contributes to the integral.

For example in Figure 3 we illustrate examples of the anti-thimbles of the saddle point

g13=1,g24=1,g14=(1)3/4,g23=(1)1/4,\displaystyle g_{13}=1,\quad g_{24}=-1,\quad g_{14}=(-1)^{3/4},\quad g_{23}=(-1)^{1/4}, (191)
σ13=i,σ24=i,σ14=(1)3/4,σ23=(1)1/4,\displaystyle\sigma_{13}=\text{i},\quad\sigma_{24}=-\text{i},\quad\sigma_{14}=(-1)^{3/4},\quad\sigma_{23}=-(-1)^{1/4}\,, (192)

that do not intersect with the original integration contour, namely the real axis. This means the contribution of this saddle should not be included to the integral.

Examples of anti-thimbles of another saddle point

g13=(1)1/4,g24=(1)3/4,g14=1,g23=1,\displaystyle g_{13}=-(-1)^{1/4},\quad g_{24}=(-1)^{3/4},\quad g_{14}=-1,\quad g_{23}=-1, (193)
σ13=(1)1/4,σ23=(1)3/4,σ14=i,σ23=i,\displaystyle\sigma_{13}=(-1)^{1/4},\quad\sigma_{23}=(-1)^{3/4},\quad\sigma_{14}=-\text{i},\quad\sigma_{23}=-\text{i}\,, (194)

is shown in Figure 4. Again they do not intersect with the real axis so the contribution from this saddle should not be included either.

Refer to caption
Refer to caption
Figure 4: Anti-thimble on the g13g_{13} plane (left) and the g24g_{24} plane (right).

We can run this analysis over all the nontrivial saddles and find none of them contribute to the integral. As a result, the path integral can be approximated entirely by the trivial saddle.

Refer to caption
Figure 5: The shaded region is where a non-trivial saddle in (195) dominates over the trivial saddle. The plot for the other two non-trivial saddles can be obtained from this plot by simple rotations.

\bullet When x=yx=-y, there are also nontrivial saddle points and a similar analysis of Lefschetz thimbles demonstrate that they do not contribute to the integral.

Actually, there is a quicker way to arrive at the same conclusion. We find that the on-shell actions corresponding to these saddle points are

(σ22)N3eN±32213Ne2imπ3σ43,m=0,±1,σ.\displaystyle\left(\frac{\sigma^{2}}{2}\right)^{\frac{N}{3}}e^{-N\pm\frac{3}{2}2^{\frac{1}{3}}Ne^{\frac{2\text{i}m\pi}{3}}\sigma^{\frac{4}{3}}},\quad m=0,\pm 1,\quad\sigma\rightarrow\infty\ . (195)

However these saddle points should be saddle points of the entire multi-dimensional integral including the integral over σ\sigma. As a result this saddle should also satisfy the fall-off condition of the σ\sigma integral, otherwise they will not contribute to the σ\sigma integral. Therefore we should only consider the decaying saddle points namely

(σ22)N3eN+32213Ne±2iπ3σ43,(σ22)N3eN32213Nσ43.\displaystyle\left(\frac{\sigma^{2}}{2}\right)^{\frac{N}{3}}e^{-N+\frac{3}{2}2^{\frac{1}{3}}Ne^{\pm\frac{2\text{i}\pi}{3}}\sigma^{\frac{4}{3}}},\quad\left(\frac{\sigma^{2}}{2}\right)^{\frac{N}{3}}e^{-N-\frac{3}{2}2^{\frac{1}{3}}N\sigma^{\frac{4}{3}}}\ . (196)

We plot the region where these non-trivial saddle dominates over the trivial saddle in Figure 5, and it is easy to observe from the figure that the wormhole saddle (317) of z2\langle z^{2}\rangle, located at |σ|=1|\sigma|=1, is in the region where the trivial saddle dominates.

Another family of solutions to the equation of motion (180) has x=0x=0 or y=0y=0. On shell actions on these saddles behave as

σ2N3eN+32Ne±2iπ3σ43,σ2N3eN32Nσ43,\displaystyle\sigma^{\frac{2N}{3}}e^{-N+\frac{3}{2}Ne^{\pm\frac{2\text{i}\pi}{3}}\sigma^{\frac{4}{3}}},\quad\sigma^{\frac{2N}{3}}e^{-N-\frac{3}{2}N\sigma^{\frac{4}{3}}}\,, (197)

whose dominant regions are similar to Figure 5 and they are sub-leading comparing with the trivial saddle.

Putting all the result together we confirm that the trivial saddle point dominate in the gabg_{ab} and σab\sigma_{ab} integral and the wormhole saddle (317) is self-averaging.

3.2.3 The unaveraged z2z^{2}: the linked half-wormhole saddles

The trivial saddle point discussed in the previous section gives vanishing contribution at σ0\sigma\sim 0, so we expect other saddle points dominate the path integral here. In Saad:2021rcu they are referred to as the (linked) half-wormhole saddles. Here we provide some further details of the saddle contribute at σ0\sigma\sim 0 and show that it agrees with the exact result in (170), ie

Φ(0)2ext2qe32N.\displaystyle\langle\Phi(0)^{2}\rangle_{\text{ext}}\sim 2qe^{-\frac{3}{2}N}\ . (198)

We can apply the same analysis, except that now we evaluate at σ0\sigma\sim 0, as in the previous section. As expected, the trivial saddle gives

eNlog(σ)0.\displaystyle e^{N\log(\sigma)}\sim 0\ . (199)

The subleading non-trivial saddles (196) and (197) discussed in the previous section has on-shell values

e32N2N/2,e32N,\displaystyle\frac{e^{-\frac{3}{2}N}}{2^{N/2}},\quad e^{-\frac{3}{2}N}\ , (200)

respectively when σ=0\sigma=0. So (197) dominates. Adding them up precisely gives the exact solution (198)

2qe32N,\displaystyle 2qe^{-\frac{3}{2}N}\,, (201)

The general lesson is that the linked half wormhole saddle points are always in the integral, and furthermore they are also always saddles. It’s only that they are, for most of the time, hidden behind the leading saddles. They can only be exposed in regions where the leading saddle decreases faster, namely the σ0\sigma\sim 0 region in this case.

4 SYK at one time point: Ja0\langle J_{a}\rangle\neq 0

In the following, we will generalize the study of half-wormhole along several directions. The main question we want to address is how the distribution of the random coupling affects the wormhole and half-wormhole saddles.

First let us consider the case where the random coupling is drawn from a general Gaussian distribution 𝒩(u,t2){\cal N}(u,t^{2})888When we write JAJ_{A}, we have in mind that the index set AA is automatically sorted, and all JJ’s with other permutations of AA picks up signs accordingly.

JA=JA0=u,JA2JA2=τ2(q1)!Nq1t2,\displaystyle\langle J_{A}\rangle=J_{A}^{0}=u,\quad\langle J_{A}^{2}\rangle-\langle J_{A}\rangle^{2}=\tau^{2}\frac{(q-1)!}{N^{q-1}}\equiv t^{2}\,, (202)

in particular, the mean value of the random coupling could be non-vanishing.

The ensemble averaged quantities can be computed directly by first averaging over the couplings and then integrating out the fermions

z\displaystyle\langle z\rangle =\displaystyle= PF(J0),\displaystyle\text{PF}(J^{0})\,, (203)
z2\displaystyle\langle z^{2}\rangle =\displaystyle= d2Nψexp(iqt2AψALψAR+iq/2JA0(ψAL+ψAR))\displaystyle\int d^{2N}\psi\exp\left(\text{i}^{q}t^{2}\sum_{A}\psi_{A}^{L}\psi_{A}^{R}+\text{i}^{q/2}J_{A}^{0}(\psi_{A}^{L}+\psi_{A}^{R})\right) (204)
=\displaystyle= A,Bsgn(A)sgn(B)(JA10JB10+δA1B1t2))(JAp0JBp0+δApBpt2)).\displaystyle\sum^{\prime}_{A,B}\text{sgn}(A)\text{sgn}(B)\left(J_{A_{1}}^{0}J_{B_{1}}^{0}+\delta_{A_{1}B_{1}}t^{2})\right)\dots\left(J_{A_{p}}^{0}J_{B_{p}}^{0}+\delta_{A_{p}B_{p}}t^{2})\right)\ . (205)

4.1 Half-wormhole saddle in zz

Since z0\langle z\rangle\neq 0, we expect a disk saddle point in the path integral presentation of zz that gives the contribution of z\langle z\rangle. Moreover, like linked half-wormhole contribution to z2z^{2} in the model with u=0u=0, it is possible that there are also single half-wormhole saddles contributing to zz, 999This single half-wormhole saddle is related to the half-wormhole saddle of JT gravity introduced in Blommaert:2021fob . as shown in Figure. 6. We will show in the following that such saddles indeed exist and together with their contribute Θ1\Theta_{1} the following approximation is good

zz+Θ1.\displaystyle z\approx\langle z\rangle+\Theta_{1}\ . (206)

Let us clarify the notation we use in this paper, we call the non-self-averaged component in zz as “single half-wormhole” or simply “half-wormhole”, and we refer to the non-self-averaged saddle in z2z^{2} as “linked half-wormhole”.

Refer to caption
Figure 6: The single half-wormhole saddle of zz.

To demonstrate (206) explicitly, recall that the partition function is given by

z=dNψexp(iq/2Ji1iqψi1iq).\displaystyle z=\int\text{d}^{N}\psi\exp\left(\text{i}^{q/2}\sum J_{i_{1}\dots i_{q}}\psi_{i_{1}\dots i_{q}}\right)\ . (207)

The ensemble averaged quantity z\langle z\rangle does not vanish

z=dNψexp(iq/2Ji1iq(0)ψi1iq)=up(pq/2)!p!((q/2)!)pmpup,pq=N.\displaystyle\langle z\rangle=\int\text{d}^{N}\psi\exp(\text{i}^{q/2}\sum J^{(0)}_{i_{1}\dots i_{q}}\psi_{i_{1}\dots i_{q}})=u^{p}\frac{(pq/2)!}{p!((q/2)!)^{p}}\equiv m_{p}u^{p}\,,\quad pq=N\ . (208)

In the following we present a heuristic but simple proof of this result. A more rigorous but technical proof is presented in Appendix G. For simplicity let us first consider the q=4q=4 case

z=dNψeuAψA,A={a1<<a4}.\displaystyle\langle z\rangle=\int d^{N}\psi\,e^{-u\sum_{A}\psi_{A}},\quad A=\{a_{1}<\dots<a_{4}\}\ . (209)

We introduce the collective variable GG

G=1N1i<jNψiψj,G2=2!N2AψA,\displaystyle G=\frac{1}{N}\sum_{1\leq i<j\leq N}\psi_{i}\psi_{j},\quad G^{2}=\frac{2!}{N^{2}}\sum_{A}\psi_{A}\,, (210)

then z\langle z\rangle can be rewritten as

z=dGidΣ2πi/NdNψeu2N2G2eΣ(NGi<jψiψj).\displaystyle\langle z\rangle=\int_{\mathbb{R}}\text{d}G\int_{\text{i}\mathbb{R}}\frac{d\Sigma}{2\pi\text{i}/N}\text{d}^{N}\psi\,e^{-\frac{u}{2}N^{2}G^{2}}e^{-\Sigma(NG-\sum_{i<j}\psi_{i}\psi_{j})}\ . (211)

Now we can integrate the out the fermions to get

dNψeΣi<jψiψj=(Σ)N/2mp|(q=2)=ΣN/2.\displaystyle\int d^{N}\psi\,e^{\Sigma\sum_{i<j}\psi_{i}\psi_{j}}=(\Sigma)^{N/2}m_{p}\,|_{(q=2)}=\Sigma^{N/2}\ . (212)

Then (211) becomes

zq=4\displaystyle\langle z\rangle_{q=4} =\displaystyle= dGidΣ2πi/NΣN/2euN2G22eNΣG\displaystyle\int_{\mathbb{R}}\text{d}G\int_{\text{i}\mathbb{R}}\frac{d\Sigma}{2\pi\text{i}/N}\Sigma^{N/2}e^{-\frac{uN^{2}G^{2}}{2}}e^{-N\Sigma G}\, (213)
=\displaystyle= NN/2(G)N/2euN2G22|G=0=(u2)N/4(N/2)!(N/4)!=mpup|q=4.\displaystyle N^{-N/2}(\partial_{G})^{N/2}e^{-\frac{uN^{2}G^{2}}{2}}\,|_{G=0}\,=\left(\frac{u}{2}\right)^{N/4}\frac{(N/2)!}{(N/4)!}=m_{p}u^{p}|_{q=4}\,.

For general qq, the proof is similar with the modification

AψA=Nq/2(q/2)!Gq/2.\displaystyle\sum_{A}\psi_{A}=\frac{N^{q/2}}{(q/2)!}G^{q/2}\,. (214)

In summary, we have generalized the G,ΣG,\Sigma trick and derived an effective action to compute z\langle z\rangle:

z=dGidΣ2πi/NΣN/2euiq/2Nq/2(q/2)!Gq/2eNΣG.\displaystyle\langle z\rangle=\int_{\mathbb{R}}\text{d}G\int_{\text{i}\mathbb{R}}\frac{d\Sigma}{2\pi\text{i}/N}\Sigma^{N/2}e^{u\text{i}^{q/2}\frac{N^{q/2}}{(q/2)!}G^{q/2}}e^{-N\Sigma G}\,. (215)

It would be convenient to rotate the integral contour as

Σiei2πqσ,Gei2πqg\displaystyle\Sigma\rightarrow\text{i}e^{-\text{i}\frac{2\pi}{q}}\sigma,\quad G\rightarrow e^{\text{i}\frac{2\pi}{q}}g (216)

such that we obtain a “standard” action:

z=dgdσ2π/Nexp{N2(log(ie2πiqσ)2iσg2μqgq/2)},\displaystyle\langle z\rangle=\int_{\mathbb{R}}\frac{dgd\sigma}{2\pi/N}\exp\left\{\frac{N}{2}\left(\log(\text{i}e^{-\frac{2\pi\text{i}}{q}}\sigma)-2\text{i}\sigma g-\frac{2\mu}{q}g^{q/2}\right)\right\}, (217)

where we define

μiq/2u2Nq/21(q/21)!,u=(i)q/2μ(q/21)!2Nq/21.\displaystyle\mu\equiv\text{i}^{q/2}u\frac{2{N}^{q/2-1}}{(q/2-1)!},\quad\leftrightarrow\quad u=(-\text{i})^{q/2}\mu\frac{(q/2-1)!}{2N^{q/2-1}}. (218)

Rescaling μ\mu to 1, the saddle point equations are then

1σ2ig=0,2iσμgq/21=0,μgq/2=1.\displaystyle\frac{1}{\sigma}-2\text{i}g=0,\quad-2\text{i}\sigma-\mu g^{q/2-1}=0,\quad\rightarrow\quad\mu g^{q/2}=-1\ . (219)

Comparing (217) with (172) it is easy to find that to reproduce the exact result (208) we have to added the contributions from all the q/2q/2 saddles.

Having found the suitable saddle contributions to the averaged partition function z\langle z\rangle, we proceed to analyze the difference between the non-averaged quantity and the mean value zzz-\langle z\rangle. We start with inserting the identity

1=𝑑GhiiNdΣh2πieΣh(NGhi<jψiψj)+Nμq(Ghq/2(1Ni<jψiψj)q/2),\displaystyle 1=\int_{-\infty}^{\infty}dG_{h}\int_{-\text{i}\infty}^{\text{i}\infty}\frac{Nd\Sigma_{h}}{2\pi\text{i}}e^{-\Sigma_{h}(NG_{h}-\sum_{i<j}\psi_{i}\psi_{j})+\frac{N\mu}{q}\left(G_{h}^{q/2}-\left(\frac{1}{N}\sum_{i<j}\psi_{i}\psi_{j}\right)^{q/2}\right)}\,,

into the non-averaged partition function zz. To make the integral well defined, we again rotate the contour by Σh=ie2iπ/qσh,Gh=e2iπ/qgh\Sigma_{h}=\text{i}e^{-2\text{i}\pi/q}\sigma_{h},G_{h}=e^{2\text{i}\pi/q}g_{h}, then zz can be cast into the form

z=Ndσh2πΨ(σh)Θ^(σh),\displaystyle z=\int_{-\infty}^{\infty}\frac{N\text{d}\sigma_{h}}{2\pi}\Psi(\sigma_{h})\hat{\Theta}(\sigma_{h})\,, (221)

where the first factor is similar to (164)

Ψ(σh)=dgh2π/Nexp[N(iσhghμqghq/2)],\displaystyle\Psi(\sigma_{h})=\int_{\mathbb{R}}\frac{\text{d}g_{h}}{2\pi/N}\exp[N(-\text{i}\sigma_{h}g_{h}-\frac{\mu}{q}g_{h}^{q/2})]\,, (222)

and the second factor is

Θ^(σh)=dNψexp[ie2iπqσhi<jψiψj+iq/2JAψAiq/2uAψA].\displaystyle\hat{\Theta}(\sigma_{h})=\int\text{d}^{N}\psi\exp[\text{i}e^{-\frac{2\text{i}\pi}{q}}\sigma_{h}\sum_{i<j}\psi_{i}\psi_{j}+\text{i}^{q/2}J_{A}\psi_{A}-\text{i}^{q/2}u\sum_{A}\psi_{A}]\ . (223)

Averaging over the coupling, we get back to the computation in (217) where σh=12i(μ2/qe4πi(n+12)/q)\sigma_{h}=\frac{1}{2i}\left(\mu^{-2/q}e^{4\pi i(n+\frac{1}{2})/q}\right). We expect a separate saddle point to appear in this integral which leads to the difference zzz-\langle z\rangle. The Ψ(σh)\Psi(\sigma_{h}) is peaked at σh=0\sigma_{h}=0, so we look for dominant contributions around σh0\sigma_{h}\approx 0, which is

Θ1=Θ^(0)=Pf(JJ0)=Asgn(A)(JA1JA10)(JApJAp0).\displaystyle\Theta_{1}=\hat{\Theta}(0)=\text{Pf}(J-J^{0})=\sum^{\prime}_{A}\text{sgn}(A)(J_{A_{1}}-J_{A_{1}}^{0})\dots(J_{A_{p}}-J_{A_{p}}^{0})\ . (224)

It is clear that its average vanishes Θ1=0\langle\Theta_{1}\rangle=0. Then we propose the approximation

zz+Θ1.\displaystyle z\approx\langle z\rangle+\Theta_{1}\ . (225)

which is (206). According to the power of JA0=uJ^{0}_{A}=u, we can further expand

Θ1\displaystyle\Theta_{1} =k=0pΘ1(k)uk.\displaystyle=\sum_{k=0}^{p}\Theta_{1}^{(k)}u^{k}\ . (226)

To verify this approximation, we define the error function

Error=zzΘ1.\displaystyle\text{Error}=z-\langle z\rangle-\Theta_{1}\ . (227)

A direct calculation gives

Error2=z2z2+Θ22zΘ\displaystyle\langle\text{Error}^{2}\rangle=\langle z^{2}\rangle-\langle z\rangle^{2}+\langle\Theta^{2}\rangle-2\langle z\Theta\rangle (228)

The quantities z2,Θ2,zΘ\langle z^{2}\rangle,\langle\Theta^{2}\rangle,\langle z\Theta\rangle can be computed with the Feynman diagrams as shown in Fig. 7.

Refer to caption
Figure 7: Feynman diagrams for z2,Θ12,zΘ1\langle z^{2}\rangle,\langle\Theta_{1}^{2}\rangle,\langle z\Theta_{1}\rangle. Each black dot represents a zz or Θ1\Theta_{1}, each red dot and the attached line represents a contraction with the JA0J_{A}^{0} source, and each blue line is a contraction of a pair of JAJ_{A}.

Recall that value of z\langle z\rangle is given by the star diagram that is one connected component of the last term in Fig. 7

z=(pq/2)!p!((q/2)!)pμpmpμp,\displaystyle\langle z\rangle=\frac{(pq/2)!}{p!((q/2)!)^{p}}\mu^{p}\equiv m_{p}\mu^{p}\,, (229)

The value of z2\langle z^{2}\rangle can be computed either from summing over the diagrams,

z2=k=0pckmpk2t2ku2p2kkz2(k),\displaystyle\langle z^{2}\rangle=\sum_{k=0}^{p}c_{k}m_{p-k}^{2}t^{2k}u^{2p-2k}\equiv\sum_{k}z_{2}^{(k)}\,, (230)

where

ck=1k!(Nq)(Nqq)(N(k1)qq)=N!k!(q!)k(Nkq)!,\displaystyle c_{k}=\frac{1}{k!}{N\choose q}{N-q\choose q}\dots N-(k-1)q\choose q=\frac{N!}{k!(q!)^{k}(N-kq)!}\,, (231)

or by introducing the collective variables

GLR=1NiψiLψiR,GL=1Ni<jψiLψjL,GR=1Ni<jψiRψjR,\displaystyle G_{LR}=\frac{1}{N}\sum_{i}\psi_{i}^{L}\psi_{i}^{R},\quad G_{L}=\frac{1}{N}\sum_{i<j}\psi_{i}^{L}\psi_{j}^{L},\quad G_{R}=\frac{1}{N}\sum_{i<j}\psi_{i}^{R}\psi_{j}^{R}\,, (232)

and doing the path integral

z2\displaystyle\langle z^{2}\rangle =\displaystyle= Rd3Giid3ΣieNq(τ2GLRq+μGLq/2+μGRq/2)N(ΣiGi)d2Nψe12ΨMΨ,\displaystyle\int_{R}\text{d}^{3}G_{i}\int_{\text{i}\mathbb{R}}d^{3}\Sigma_{i}\,e^{\frac{N}{q}(\tau^{2}G_{LR}^{q}+\mu G_{L}^{q/2}+\mu G_{R}^{q/2})-N(\Sigma_{i}G_{i})}\int\text{d}^{2N}\psi e^{\frac{1}{2}{\Psi}M{\Psi}},
=\displaystyle= Rd3Giid3ΣieNq(τ2GLRq+μGLq/2+μGRq/2)N(ΣiGi)det[ΣLΣRA2+ΣLR2]\displaystyle\int_{R}\text{d}^{3}G_{i}\int_{\text{i}\mathbb{R}}d^{3}\Sigma_{i}\,e^{\frac{N}{q}(\tau^{2}G_{LR}^{q}+\mu G_{L}^{q/2}+\mu G_{R}^{q/2})-N(\Sigma_{i}G_{i})}\sqrt{\text{det}[\Sigma_{L}\Sigma_{R}A^{2}+\Sigma_{LR}^{2}]}
=\displaystyle= Rd3Giid3ΣieNq(τ2GLRq+μGLq/2+μGRq/2)N(ΣiGi)det[iΣLΣRA+ΣLR]\displaystyle\int_{R}\text{d}^{3}G_{i}\int_{\text{i}\mathbb{R}}d^{3}\Sigma_{i}\,e^{\frac{N}{q}(\tau^{2}G_{LR}^{q}+\mu G_{L}^{q/2}+\mu G_{R}^{q/2})-N(\Sigma_{i}G_{i})}\text{det}[\text{i}\sqrt{\Sigma_{L}\Sigma_{R}}A+\Sigma_{LR}]
=\displaystyle= Rd3Giid3ΣieNq(τ2GLRq+μGLq/2+μGRq/2)N(ΣiGi)12((ΣLR+iΣLΣR)N+(ΣLRiΣLΣR)N)\displaystyle\int_{R}\text{d}^{3}G_{i}\int_{\text{i}\mathbb{R}}d^{3}\Sigma_{i}\,e^{\frac{N}{q}(\tau^{2}G_{LR}^{q}+\mu G_{L}^{q/2}+\mu G_{R}^{q/2})-N(\Sigma_{i}G_{i})}\frac{1}{2}\left((\Sigma_{LR}+\text{i}\sqrt{\Sigma_{L}\Sigma_{R}})^{N}+(\Sigma_{LR}-\text{i}\sqrt{\Sigma_{L}\Sigma_{R}})^{N}\right)
=\displaystyle= Rd3Giid3Σim=0N/2(N2m)(ΣLR)2m(i2ΣLΣR)N2meNq(τ2GLRq+μGLq/2+μGRq/2)eN(ΣiGi),\displaystyle\int_{R}\text{d}^{3}G_{i}\int_{\text{i}\mathbb{R}}d^{3}\Sigma_{i}\,\sum_{m=0}^{N/2}{N\choose 2m}(\Sigma_{LR})^{2m}(\text{i}^{2}\Sigma_{L}\Sigma_{R})^{\frac{N}{2}-m}e^{\frac{N}{q}(\tau^{2}G_{LR}^{q}+\mu G_{L}^{q/2}+\mu G_{R}^{q/2})}e^{-N(\Sigma_{i}G_{i})}\,,

where we have defined

Ψ=(ψ1L,,ψNL,ψ1R,,ψNR),M=(ΣLAΣLRINΣLRINΣRA),\displaystyle\Psi=\left(\psi_{1}^{L},\dots,\psi_{N}^{L},\psi_{1}^{R},\dots,\psi_{N}^{R}\right),\quad M=\begin{pmatrix}\Sigma_{L}A&\Sigma_{LR}I_{N}\\ -\Sigma_{LR}I_{N}&\Sigma_{R}A\\ \end{pmatrix}, (234)
A=AT,Aij=1,i<j.\displaystyle A=-A^{T},\quad A_{ij}=1,\quad\forall i<j. (235)

Using the same tricks as (213), (4.1) can be evaluated exactly as

z2\displaystyle\langle z^{2}\rangle =\displaystyle= NNk=0p(Nkq)(GLR)kq(i2GLGR)Nkq2eNq(τ2GLRq+μGLq/2+μGRq/2)|Gi=0\displaystyle N^{-N}\sum_{k=0}^{p}{N\choose kq}(\partial_{G_{LR}})^{kq}(\text{i}^{2}\partial_{G_{L}}\partial_{G_{R}})^{\frac{N-kq}{2}}e^{\frac{N}{q}(\tau^{2}G_{LR}^{q}+\mu G_{L}^{q/2}+\mu G_{R}^{q/2})}|_{G_{i}=0} (236)
=\displaystyle= NNk=0piNkq(Nkq)(kq)!k!(Nτ2q)k[(q(pk)2)!(pk)!]2(Nμq)2p2k\displaystyle N^{-N}\sum_{k=0}^{p}\text{i}^{N-kq}{N\choose kq}\frac{(kq)!}{k!}\left(\frac{N\tau^{2}}{q}\right)^{k}\left[\frac{(\frac{q(p-k)}{2})!}{(p-k)!}\right]^{2}\left(\frac{N\mu}{q}\right)^{2p-2k} (237)
=\displaystyle= k=0pckmpk2t2ku2p2k,\displaystyle\sum_{k=0}^{p}c_{k}m_{p-k}^{2}t^{2k}u^{2p-2k}, (238)

which agrees with (230) as it should be.

Furthermore, from this result we find z2(0)=z2z_{2}^{(0)}=\langle z\rangle^{2} which is given by the last diagram in Fig. 7 and z2(p)=z2μ=0z_{2}^{(p)}=\langle z^{2}\rangle_{\mu=0} which is given by the first diagram in Fig. 7. The expression of Θ1\Theta_{1} (224) implies that Θ12=Θ1z=z2(p)\langle\Theta_{1}^{2}\rangle=\langle\Theta_{1}z\rangle=z_{2}^{(p)}, therefore we find

Error2=k=1p1ckmpk2t2ku2p2kk=1p1z2(k),\displaystyle\langle\text{Error}^{2}\rangle=\sum_{k=1}^{p-1}c_{k}m_{p-k}^{2}t^{2k}u^{2p-2k}\equiv\sum_{k=1}^{p-1}z_{2}^{(k)}\,, (239)

where mpm_{p} is defined in (208). In the large-NN limit, some of the terms in the summation (230) dominate. If z2(p)z_{2}^{(p)} or z2(0)z_{2}^{(0)} dominates then the error is small.

However the dominant term is not always given by a fixed z2(k)z_{2}^{(k)}. A simple argument is the following. To find the dominant term we can compute the ratio101010Recall that p=N/qp=N/q.

rk=z2(k)z2(k1)=t2(k+p+1)(4k+4p+1)(4k+4p+3)3u2(2k(pk)+k),\displaystyle r_{k}=\frac{z_{2}^{(k)}}{z_{2}^{(k-1)}}=\frac{t^{2}(-k+p+1)(-4k+4p+1)(-4k+4p+3)}{3u^{2}(2k(p-k)+k)}\,, (240)
rp=t2pu2,r1p2t2u2,\displaystyle r_{p}=\frac{t^{2}}{pu^{2}},\quad r_{1}\sim\frac{p^{2}t^{2}}{u^{2}}\,, (241)

here for simplicity we have chosen q=4q=4. First we notice that rkr_{k} decreases with respect to kk. Therefore if r11r_{1}\leq 1 i.e.

utp,\displaystyle\frac{u}{t}\geq{p}\,, (242)

then the dominant term will be z2(0)z_{2}^{(0)}. It means that all the wormhole saddles are suppressed. However if rp1r_{p}\geq 1 i.e.

ut1p\displaystyle\frac{u}{t}\leq\frac{1}{\sqrt{p}} (243)

then the dominant term will be z2(p)z_{2}^{(p)}, in other words the effect of μ\mu can be neglected. For other cases with

1p<ut<p,\displaystyle\frac{1}{\sqrt{p}}<\frac{u}{t}<p, (244)

by fine tuning the value of u/tu/t, every diagram in Fig. (7) is possible to be dominant. For the choices (202) and (218) which lead to reasonable large NN behavior we have

utμτ(q/21)!(q1)!N12p,\displaystyle\frac{u}{t}\sim\frac{\mu}{\tau}\frac{(q/2-1)!}{\sqrt{(q-1)!}}N^{\frac{1}{2}}\sim\sqrt{p}, (245)

which exactly lies in the (244). It also implies there should be other saddles contributing to (223).

On the other hand, the can derive the saddle point equations

GL(R)1+q2=2μΣL(R),GLR1+q=1τ2ΣLR,\displaystyle G_{L(R)}^{-1+\frac{q}{2}}=\frac{2}{\mu}\Sigma_{L(R)},\quad G_{LR}^{-1+q}=\frac{1}{\tau^{2}}\Sigma_{LR}, (246)
GL(R)=iΣR(L)2ΣLΣRf+n1fn1f+n+fn,GLR=f+n1+fn1f+n+fn,\displaystyle G_{L(R)}=\frac{\text{i}\Sigma_{R(L)}}{2\sqrt{\Sigma_{L}\Sigma_{R}}}\frac{f_{+}^{n-1}-f_{-}^{n-1}}{f_{+}^{n}+f_{-}^{n}},\quad G_{LR}=\frac{f_{+}^{n-1}+f_{-}^{n-1}}{f_{+}^{n}+f_{-}^{n}}\,, (247)

where f±=ΣLR±iΣLΣRf_{\pm}=\Sigma_{LR}\pm\text{i}\sqrt{\Sigma_{L}\Sigma_{R}}. Again for simplicity we will choose τ2=μ=1\tau^{2}=\mu=1. There are always two types of trivial solutions

wormhole solution:GL=GR=0,GLR=e2imπq,\displaystyle\text{wormhole solution}:\quad G_{L}=G_{R}=0,\quad G_{LR}=e^{\frac{2\text{i}m\pi}{q}}, (248)
disconnect solution:GLR=0,GL=e4imLπq,GR=e4imRπq\displaystyle\text{disconnect solution}:\quad G_{LR}=0,\quad G_{L}=e^{\frac{4\text{i}m_{L}\pi}{q}},\quad G_{R}=e^{\frac{4\text{i}m_{R}\pi}{q}} (249)

with on-shell action

wormhole solution:z2wh=eN(11q)e2imπNq\displaystyle\text{wormhole solution}:\quad\langle z^{2}\rangle_{\text{wh}}=e^{-N(1-\frac{1}{q})}e^{\frac{2\text{i}m\pi N}{q}} (250)
disconnect solution:z2dis=2NeN(12q)e4imπNq.\displaystyle\text{disconnect solution}:\quad\langle z^{2}\rangle_{\text{dis}}={2^{-N}}e^{-N(1-\frac{2}{q})}{e^{\frac{4\text{i}m\pi N}{q}}}. (251)

Note that the ratio of these two contribution is

z2whz2dis=(2e1/q)N,\displaystyle\frac{\langle z^{2}\rangle_{\text{wh}}}{\langle z^{2}\rangle_{\text{dis}}}=\left(2e^{-1/q}\right)^{N}, (252)

so when q2q\geq 2 it is the wormhole saddle dominates. The general analytic solution is hard to obtain. However in the large NN limit we expect that only f+f_{+} or ff_{-} will survive. Assuming fN0,Nf^{N}_{-}\rightarrow 0,N\rightarrow\infty, (247) get dramatically simplified

GL(R)=ΣR(L)2iΣRΣL1ΣLR+iΣLΣR,GLR=1ΣLR+iΣLΣR,\displaystyle G_{L(R)}=\frac{\Sigma_{R(L)}}{-2\text{i}\sqrt{\Sigma_{R}\Sigma_{L}}}\frac{1}{\Sigma_{LR}+\text{i}\sqrt{\Sigma_{L}\Sigma_{R}}},\quad G_{LR}=\frac{1}{\Sigma_{LR}+\text{i}\sqrt{\Sigma_{L}\Sigma_{R}}}, (253)

from which we obtain

GLRq+GRq/2+GLq/2=1,GRq/2=GLq/2.\displaystyle G_{LR}^{q}+G_{R}^{q/2}+G_{L}^{q/2}=1,\quad G_{R}^{q/2}=G_{L}^{q/2}. (254)

For the case of q=4q=4, (246) and (253) can be solved explicitly and it contributes the on-shell action

z2non-trivial+e0.63Ne2miπN4>z2wh=e0.75Ne2miπN4.\displaystyle\langle z^{2}\rangle_{\text{non-trivial}+}\approx e^{-0.63N}e^{\frac{2m\text{i}\pi N}{4}}>\langle z^{2}\rangle_{\text{wh}}=e^{-0.75N}e^{\frac{2m\text{i}\pi N}{4}}. (255)

We also checked that for these solutions limNfN=0\lim_{N\rightarrow\infty}f_{-}^{N}=0. Similar saddles can also be found for the case of f+N=0f_{+}^{N}=0. Therefore we conclude that in the large NN limit the dominate saddles are the non-trivial ones.

In the regime of (244), the ansatz (224) of half-wormhole saddle is not adequate. We have to consider the contribution from the σh\sigma_{h} fluctuation to Θ\Theta. This can be done by expanding Θ^(σh)\hat{\Theta}(\sigma_{h}) with respect σh\sigma_{h}, substituting into zz and integrating over σh\sigma_{h}. Equivalently this can be done by expanding the exact value of zz

z\displaystyle z =\displaystyle= PF(JA)=PF(u+JAJA0)\displaystyle\text{PF}(J_{A})=\text{PF}(u+J_{A}-J_{A}^{0}) (256)
=\displaystyle= Asgn(A)(u+JA1JA10)(u+JApJAp0)n=0pΘ(n),\displaystyle\sum^{\prime}_{A}\text{sgn}(A)(u+J_{A_{1}}-J_{A_{1}}^{0})\dots(u+J_{A_{p}}-J_{A_{p}}^{0})\equiv\sum_{n=0}^{p}\Theta^{(n)}\,,

with respect to uu. For examples

Θ(p1)=Asgn(A)(JA1JA10)JAi0(JApJAp0),\displaystyle\Theta^{(p-1)}=\sum_{A}^{\prime}\text{sgn}(A)(J_{A_{1}}-J_{A_{1}}^{0})\dots J_{A_{i}}^{0}\dots(J_{A_{p}}-J_{A_{p}}^{0})\,, (257)
Θ(0)=z,Θ(p)=Θ.\displaystyle\Theta^{(0)}=\langle z\rangle\,,\quad\Theta^{(p)}=\Theta. (258)

Then from the Feynman diagrams it is not hard to find in Fig. 7 that

Θ(k)Θ(k)=Θ(k)z=z2(k).\displaystyle\langle{\Theta^{(k)}}{\Theta^{(k)}}\rangle=\langle\Theta^{(k)}z\rangle=z_{2}^{(k)}. (259)

So if z2(k)z_{2}^{(k)} is the dominant term, we can choose the half-wormhole saddle to be Θ(k)\Theta^{(k)}. Or we can think of that for each wormhole saddle z2(k)z_{2}^{(k)} there is a corresponding half-wormhole saddle Θ(k)\Theta^{(k)} such that

zz+Θ(k).\displaystyle z\approx\langle z\rangle+\Theta^{(k)}. (260)

We will present a further analysis on this model somewhere else.

4.2 Linked half-wormhole saddles in z2z^{2}

In this section we study the linked half-wormhole contribution to z2z^{2}, and, in particular, we would like to understand the relation with the single half-wormhole saddles in zz,

To get a general picture, we first compute z4\langle z^{4}\rangle from the Feynman diagrams shown in Fig.8. In general it is a cumbersome combinatorial problem but in the large NN limit we know that it should be factorized into disconnected diagrams as

z43z2(k)2,z2z2(k),\displaystyle\langle z^{4}\rangle\approx 3{z_{2}^{(k)}}^{2}\,,\qquad\langle z^{2}\rangle\approx z_{2}^{(k)}\,, (261)

which is shown in Fig.9 and here we have assumed that z2(k)z_{2}^{(k)} is the dominant wormhole saddles.

This means there are more refined structures of the nontrivial saddles in z2z^{2}, comparing with the general discussion in Saad:2021rcu . Inspired by our analysis of the single half-wormhole for zz, we insert another two copies of identities (4.1) in z2z^{2}

z2=dσwdσhLdσhRΨ(σw,σhL,σhL)Λ^(σw,σhL,σhL),\displaystyle z^{2}=\int\text{d}\sigma_{w}\text{d}\sigma_{h_{L}}\text{d}\sigma_{h_{R}}\Psi(\sigma_{w},\sigma_{h_{L}},\sigma_{h_{L}})\hat{\Lambda}(\sigma_{w},\sigma_{h_{L}},\sigma_{h_{L}})\,, (262)
Ψ(σw,σhL,σhL)=Ψ(σw)Ψ(σhL)Ψ(σhR),\displaystyle\Psi(\sigma_{w},\sigma_{h_{L}},\sigma_{h_{L}})=\Psi(\sigma_{w})\Psi(\sigma_{h_{L}})\Psi(\sigma_{h_{R}})\,, (263)
Λ^(σw,σhL,σhL)=d2Nψexp[ie2iπqσhLi<jψijL+ie2iπqσhRi<jψijR+ieiπqσwψiLψiR\displaystyle\hat{\Lambda}(\sigma_{w},\sigma_{h_{L}},\sigma_{h_{L}})=\int\text{d}^{2N}\psi\exp[\text{i}e^{-\frac{2\text{i}\pi}{q}}\sigma_{h_{L}}\sum_{i<j}\psi^{L}_{ij}+\text{i}e^{-\frac{2\text{i}\pi}{q}}\sigma_{h_{R}}\sum_{i<j}\psi^{R}_{ij}+\text{i}e^{\frac{\text{i}\pi}{q}}\sigma_{w}\psi_{i}^{L}\psi_{i}^{R}
+iq/2JA(ψAL+ψAR)iq/2uA(ψAL+ψAR)iqt2ψALψAR],\displaystyle\qquad\qquad\qquad\qquad+\text{i}^{q/2}J_{A}(\psi_{A}^{L}+\psi_{A}^{R})-\text{i}^{q/2}u\sum_{A}(\psi_{A}^{L}+\psi_{A}^{R})-\text{i}^{q}t^{2}\psi_{A}^{L}\psi_{A}^{R}], (264)

where we have introduced three pairs of G,ΣG,\Sigma variables

Gw=1NψiLψiR,GhL=1Ni<jψijL,GhR=1Ni<jψijR,\displaystyle G_{w}=\frac{1}{N}\psi_{i}^{L}\psi_{i}^{R},\quad G_{h_{L}}=\frac{1}{N}\sum_{i<j}\psi^{L}_{ij},\quad G_{h_{R}}=\frac{1}{N}\sum_{i<j}\psi^{R}_{ij}, (265)

and rotated the contour as before. As before, the function Ψ\Psi is highly peaked around Ψ(0,0,0)\Psi(0,0,0) so we expect that there is a half-wormhole saddle point

Λ=Λ^(0,0,0)\displaystyle\Lambda=\hat{\Lambda}(0,0,0) =A,Bsgn(A)sgn(B)k=1p((JAkJAk0)(JBkJBk0)δAkBkt2),\displaystyle=\sum^{\prime}_{A,B}\text{sgn}(A)\text{sgn}(B)\prod_{k=1}^{p}\left((J_{A_{k}}-J_{A_{k}}^{0})(J_{B_{k}}-J_{B_{k}}^{0})-\delta_{A_{k}B_{k}}t^{2}\right)\,, (266)

whose average manifestly vanishes Λ=0\langle\Lambda\rangle=0 and it further satisfies Λ2=2z2(p)2\langle\Lambda^{2}\rangle=2{z_{2}^{(p)}}^{2}.

Refer to caption
Figure 8: Feynman diagrams for z4\langle z^{4}\rangle
Refer to caption
Figure 9: z43z2(k)2\langle z^{4}\rangle\approx 3{z_{2}^{(k)}}^{2}

However because of the large NN behavior (261), again we have to consider the fluctuations of σh\sigma_{h}. It is achieved by expand Λ^(0,σhL,σhR)\hat{\Lambda}(0,\sigma_{h_{L}},\sigma_{h_{R}}) with respect to σhL(R)\sigma_{h_{L(R)}} or equivalently by expanding

A,Bsgn(A)sgn(B)k=1p((u+JAkJAk0)(u+JBkJBk0)δAkBkt2)n=0pΛ(k).\displaystyle\sum^{\prime}_{A,B}\text{sgn}(A)\text{sgn}(B)\prod_{k=1}^{p}\left((u+J_{A_{k}}-J_{A_{k}}^{0})(u+J_{B_{k}}-J_{B_{k}}^{0})-\delta_{A_{k}B_{k}}t^{2}\right)\equiv\sum_{n=0}^{p}\Lambda^{(k)}. (267)

Some examples are

Λ(p1)=iA,Bsgn(A)sgn(B)((JA1JA10)(JB1JB10)δA1B1t2)\displaystyle\Lambda^{(p-1)}=\sum_{i}\sum^{\prime}_{A,B}\text{sgn}(A)\text{sgn}(B)\left((J_{A_{1}}-J_{A_{1}}^{0})(J_{B_{1}}-J_{B_{1}}^{0})-\delta_{A_{1}B_{1}}t^{2}\right)\dots
JAi0JBi0((JApJAp0)(JBpJBp0)δApBpt2),Λ(0)=z2,Λ(p)=Λ.\displaystyle J_{A_{i}}^{0}J_{B_{i}}^{0}\dots\left((J_{A_{p}}-J_{A_{p}}^{0})(J_{B_{p}}-J_{B_{p}}^{0})-\delta_{A_{p}B_{p}}t^{2}\right),\quad\Lambda^{(0)}=\langle z\rangle^{2},\quad\Lambda^{(p)}=\Lambda\ .

Then similarly one can find that

Λ(k)Λ(k)=zΛ(k)=2z2(k)2\displaystyle\langle\Lambda^{(k)}\Lambda^{(k)}\rangle=\langle z\Lambda^{(k)}\rangle=2{z_{2}^{(k)}}^{2} (268)

so that when z2(k)z_{2}^{(k)} is the dominant wormhole saddle in the large NN limit the

z2z2+Λ(k)z2(k)+Λ(k),\displaystyle z^{2}\approx\langle z^{2}\rangle+\Lambda^{(k)}\approx z_{2}^{(k)}+\Lambda^{(k)}\,, (269)

is a good approximation.

5 SYK at one time point: Ja=0,Ja4c0\langle J_{a}\rangle=0,\quad\langle J_{a}^{4}\rangle_{c}\neq 0

Another class of interesting distributions of the random coupling is non-Gaussian. In this section we consider a special subset of them that have vanishing mean values, namely

JA=0,JA2=t2,JA4=v4+3JA22.\displaystyle\langle J_{A}\rangle=0\,,\qquad\langle J_{A}^{2}\rangle=t^{2}\,,\qquad\langle J_{A}^{4}\rangle=v^{4}+3\langle J_{A}^{2}\rangle^{2}\ . (270)

It is easy to compute that the partition function of the 0d SYK model with such random couplings are

z=0,z2=N!p!(q!)pt2,.\displaystyle\langle z\rangle=0,\quad\langle z^{2}\rangle=\frac{N!}{p!(q!)^{p}}t^{2},\ . (271)

The higher moments of JAJ_{A} in (6) contributes nontrivially to z4\langle z^{4}\rangle

z4\displaystyle\langle z^{4}\rangle =\displaystyle= A,B,C,Dsgn(A)sgn(B)sgn(C)sgn(D)JA1JB1JC1JD1JApJBpJCpJDp,\displaystyle\sum_{A,B,C,D}^{\prime}\text{sgn}(A)\text{sgn}(B)\text{sgn}(C)\text{sgn}(D)\langle J_{A_{1}}J_{B_{1}}J_{C_{1}}J_{D_{1}}\dots J_{A_{p}}J_{B_{p}}J_{C_{p}}J_{D_{p}}\rangle\,, (272)

which can be expanded

z4=k=0pcknNqkv4kt4(pk)kz4(k),\displaystyle\langle z^{4}\rangle=\sum_{k=0}^{p}c_{k}n_{N-qk}v^{4k}t^{4(p-k)}\equiv\sum_{k}z_{4}^{(k)},
nN=N!(q!)2N/qn1+n2+n3=N/qni0(qn1)!(qn2)!(qn3)!(n1!n2!n3!)2,\displaystyle n_{N}=\frac{N!}{(q!)^{2N/q}}\sum_{\begin{subarray}{c}n_{1}+n_{2}+n_{3}=N/q\\ n_{i}\geq 0\end{subarray}}\frac{(qn_{1})!(qn_{2})!(qn_{3})!}{(n_{1}!n_{2}!n_{3}!)^{2}},
cknNqk=N!k!(q!)2pkn1+n2+n3=N/qkni0(qn1)!(qn2)!(qn3)!(n1!n2!n3!)2\displaystyle c_{k}n_{N-qk}=\frac{N!}{k!(q!)^{2p-k}}\sum_{\begin{subarray}{c}n_{1}+n_{2}+n_{3}=N/q-k\\ n_{i}\geq 0\end{subarray}}\frac{(qn_{1})!(qn_{2})!(qn_{3})!}{(n_{1}!n_{2}!n_{3}!)^{2}} (273)

where ckc_{k} is the number of ways to choose kk qq-subsets out of NN and nNn_{N} is the multiplicities coming from the different Wick contractions, i.e.

z4v=0=nNt4p.\displaystyle\langle z^{4}\rangle_{v=0}=n_{N}t^{4p}. (274)

To find the dominant term in the large NN limit let us define the ratio

r~k=z4(k)z4(k1)v4t41k+pk4!(4pkp)!(4p4k+4)!,\displaystyle\tilde{r}_{k}=\frac{z_{4}^{(k)}}{z_{4}^{(k-1)}}\sim\frac{v^{4}}{t^{4}}\frac{1-k+p}{k}\frac{4!(4p-kp)!}{(4p-4k+4)!}, (275)
r~1v4t41p2,r~pυ4t41p,\displaystyle\tilde{r}_{1}\sim\frac{v^{4}}{t^{4}}\frac{1}{p^{2}},\quad\tilde{r}_{p}\sim\frac{\upsilon^{4}}{t^{4}}\frac{1}{p}\,, (276)

where we have taken q=4q=4 for simplicity. By taking the derivative with respect to kk we find that r~k\tilde{r}_{k} will initially decrease and then increase with increasing kk so r~p\tilde{r}_{p} is the maximal value. If r~p1\tilde{r}_{p}\leq 1 i.e.

v4t4p,\displaystyle\frac{v^{4}}{t^{4}}\leq p\,, (277)

then the dominant term will be z4(0)z_{4}^{(0)} therefore the contributions of higher moments can be ignored in this limit. Recall that the half-wormhole saddle of z2z^{2} when JA=0\langle J_{A}\rangle=0 can be written as

Φ=A,Bsgn(A)sgn(B)(JA1JB1δA1B1t2)(JApJBpδApBpt2),\displaystyle\Phi=\sum^{\prime}_{A,B}\text{sgn}(A)\text{sgn}(B)\left(J_{A_{1}}J_{B_{1}}-\delta_{A_{1}B_{1}}t^{2}\right)\dots\left(J_{A_{p}}J_{B_{p}}-\delta_{A_{p}B_{p}}t^{2}\right)\,, (278)

such that

Φ2Φz22z22,\displaystyle\langle\Phi^{2}\rangle\approx\langle\Phi z^{2}\rangle\approx 2\langle z^{2}\rangle^{2}, (279)

and

Error2\displaystyle\langle\text{Error}^{2}\rangle =\displaystyle= z4z22+Φ22z2Φ2\displaystyle\langle z^{4}\rangle-\langle z^{2}\rangle^{2}+\langle\Phi^{2}\rangle-2\langle z^{2}\Phi^{2}\rangle (280)
\displaystyle\approx 3z22z22+2z224z22=0,\displaystyle 3\langle z^{2}\rangle^{2}-\langle z^{2}\rangle^{2}+2\langle z^{2}\rangle^{2}-4\langle z^{2}\rangle^{2}=0,

in the leading order of NN as before. However if r~p>1\tilde{r}_{p}>1, then it will be possible that z4(p)z_{4}^{(p)} is the leading term whose corresponding Feynman diagram is shown in Fig.10.

Refer to caption
Figure 10: z4(p)z_{4}^{(p)}

Therefore there will be no half-wormhole saddle anymore since the (two-mouth) wormhole saddles are not dominant.

One can consider more general distribution with all the cumulants to be non-vanishing. The analysis and the results will be similar. If vv is very large then it is the four-way wormhole saddle that dominate. It is therefore possible to introduce a new ”four-linked-wormhole” saddle as we show in next section. However, if vv is relatively small it is still the two-mouth wormhole (with some legs as shown in Fig.7) that dominates. We will present a more thorough analysis of these points separately.

6 SYK at one time point: Ja=Ja2=Ja3=0\langle J_{a}\rangle=\langle J_{a}^{2}\rangle=\langle J_{a}^{3}\rangle=0

In this section, we consider a special model where we could focus on the “multi-linked” wormhole saddle points. In this model the random coupling only have non-vanishing 4th4^{\text{th}} cumulant

Ja=Ja2=Ja3=0,Ja4=v4.\displaystyle\langle J_{a}\rangle=\langle J_{a}^{2}\rangle=\langle J_{a}^{3}\rangle=0,\quad\langle J_{a}^{4}\rangle=v^{4}\ . (281)

Such a distribution could also be considered as an extremal limit of other distributions.

6.1 Averaged quantities: z4\langle z^{4}\rangle and z8\langle z^{8}\rangle

Due to our special choice (281) the first non-vanishing averaged quantity is

z4\displaystyle\langle z^{4}\rangle =\displaystyle= d4Nψexp(v4A1<<AqψA11ψA12ψA13ψA14ψAq1ψAq2ψAq3ψAq4)\displaystyle\int\text{d}^{4N}\psi\exp\left(v^{4}\sum_{A_{1}<\dots<A_{q}}\psi_{A_{1}}^{1}\psi_{A_{1}}^{2}\psi_{A_{1}}^{3}\psi_{A_{1}}^{4}\dots\psi_{A_{q}}^{1}\psi_{A_{q}}^{2}\psi_{A_{q}}^{3}\psi_{A_{q}}^{4}\right) (282)
=\displaystyle= d4Nψexp(v4q!(iNψi1ψi2ψi3ψi4)q).\displaystyle\int\text{d}^{4N}\psi\exp\left(\frac{v^{4}}{q!}(\sum_{i}^{N}\psi_{i}^{1}\psi_{i}^{2}\psi_{i}^{3}\psi_{i}^{4})^{q}\right)\,.

Then we can introduce the G,ΣG,\Sigma trick

z4\displaystyle\langle z^{4}\rangle =\displaystyle= d4NψdGδ(G4iNψi1ψi2ψi3ψi4)exp(v4q!G4q)\displaystyle\int\text{d}^{4N}\psi\int\text{d}G\,\delta(G_{4}-\sum_{i}^{N}\psi_{i}^{1}\psi_{i}^{2}\psi_{i}^{3}\psi_{i}^{4})\exp\left(\frac{v^{4}}{q!}G_{4}^{q}\right) (283)
=\displaystyle= d4NψdGdΣ2πiexp(Σ(G4iNψi1ψi2ψi3ψi4))exp(v4q!G4q)\displaystyle\int\text{d}^{4N}\psi\int\text{d}G\frac{\text{d}\Sigma}{2\pi\text{i}}\exp\left(-\Sigma(G_{4}-\sum_{i}^{N}\psi_{i}^{1}\psi_{i}^{2}\psi_{i}^{3}\psi_{i}^{4})\right)\exp\left(\frac{v^{4}}{q!}G_{4}^{q}\right)
=\displaystyle= dGdΣ2πiexp(NlogΣΣG4+v4q!Gq)\displaystyle\int\text{d}G\int\frac{\text{d}\Sigma}{2\pi\text{i}}\exp\left(N\log\Sigma-\Sigma G_{4}+\frac{v^{4}}{q!}G^{q}\right)
=\displaystyle= (G4)Nexp(υ4q!G4q)|G4=0=(v4q!)N/qN!(N/q)!=v4pN!p!(q!)p.\displaystyle(\partial_{G_{4}})^{N}\exp\left(\frac{\upsilon^{4}}{q!}G_{4}^{q}\right)\,|_{G_{4}=0}=\left(\frac{v^{4}}{q!}\right)^{N/q}\frac{N!}{(N/q)!}=v^{4p}\frac{N!}{p!(q!)^{p}}\,.

Alternatively, we can obtain this result by integrating out the fermions first to get the hyperpfaffin, taking the 4th4^{\text{th}} power, and then do the average

z4=ABCDsgn(A,B,C,D)JA1JB1JC1JD1JApJBpJCpJDp=v4pA 1=v4pN!p!(q!)p.\displaystyle\langle z^{4}\rangle=\sum_{ABCD}\text{sgn}(A,B,C,D)\langle J_{A_{1}}J_{B_{1}}J_{C_{1}}J_{D_{1}}\dots J_{A_{p}}J_{B_{p}}J_{C_{p}}J_{D_{p}}\rangle=v^{4p}\sum_{A}\,1=v^{4p}\frac{N!}{p!(q!)^{p}}\,. (284)

The computation of z8\langle z^{8}\rangle is more involved

z8=d8Nψexp(v4q!(iNψiaψibψicψid)q),\displaystyle\langle z^{8}\rangle=\int\text{d}^{8N}\psi\exp\left(\frac{v^{4}}{q!}(\sum_{i}^{N}\psi_{i}^{a}\psi_{i}^{b}\psi_{i}^{c}\psi_{i}^{d})^{q}\right)\,, (285)

where

(a,b,c,d){1a<b<c<d8}.\displaystyle(a,b,c,d)\in\{1\leq a<b<c<d\leq 8\}\ . (286)

In the following we use the collective index AA^{\prime} to label the 44-element subset. Then we introduce antisymmetric tensors Gabcd=GAG_{abcd}=G_{A^{\prime}} and Σabcd=ΣA\Sigma_{abcd}=\Sigma_{A^{\prime}} as the collective field variables such that (284) can be expressed as

z8\displaystyle\langle z^{8}\rangle =\displaystyle= dGAdΣA(2πi)70(PF(ΣA))Nexp(A[ΣAGA+v4q!GAq])\displaystyle\int\frac{\text{d}G_{A^{\prime}}\text{d}\Sigma_{A^{\prime}}}{(2\pi\text{i})^{70}}(\text{PF}(\Sigma_{A^{\prime}}))^{N}\exp\left(-\sum_{A^{\prime}}[\Sigma_{A^{\prime}}G_{A^{\prime}}+\frac{v^{4}}{q!}G_{A^{\prime}}^{q}]\right) (287)
=\displaystyle= (A1<A2sgn(A)GA1GA2)Nexp(v4q!GAq)|GA=0\displaystyle\left(\sum^{\prime}_{A^{\prime}_{1}<A^{\prime}_{2}}\text{sgn}(A^{\prime})\partial_{G_{A^{\prime}_{1}}}\partial_{G_{A^{\prime}_{2}}}\right)^{N}\exp\left(\frac{v^{4}}{q!}G_{A^{\prime}}^{q}\right)|_{G_{A^{\prime}}=0}
\displaystyle\approx (v4q!)2NqN!2p!212(84)=35(v4q!)2NqN!2p!2,\displaystyle\left(\frac{v^{4}}{q!}\right)^{\frac{2N}{q}}\frac{N!^{2}}{p!^{2}}\frac{1}{2}{8\choose 4}=35\left(\frac{v^{4}}{q!}\right)^{\frac{2N}{q}}\frac{N!^{2}}{p!^{2}}\,,

where in the last line we have taken the large NN limit. In this limit we have

z835z42.\displaystyle\langle z^{8}\rangle\approx 35\langle z^{4}\rangle^{2}\ . (288)

6.2 The un-averaged z4z^{4}

Following similar ideas as in the previous sections, we insert a suitable identity to the expression of z4z^{4}

z4\displaystyle z^{4} =\displaystyle= d4Nψexp(iq/2A,iJAψAi)dG4δ(G4iNa=14ψia)exp(v4q![G4q(iNa=14ψia)q])\displaystyle\int\text{d}^{4N}\psi\exp\left(\text{i}^{q/2}\sum_{A,i}J_{A}\psi_{A}^{i}\right)\int\text{d}G_{4}\delta(G_{4}-\sum_{i}^{N}\prod_{a=1}^{4}\psi_{i}^{a})\exp\left(\frac{v^{4}}{q!}[G_{4}^{q}-(\sum_{i}^{N}\prod_{a=1}^{4}\psi_{i}^{a})^{q}]\right)\,

Rotating the contour as before we can rewrite z4z^{4} as

z4=dσΨ(σ)Γ^(σ),\displaystyle z^{4}=\int\text{d}\sigma\Psi(\sigma)\hat{\Gamma}(\sigma)\,, (290)

where Ψ(σ)\Psi(\sigma) is same as (164) and the second factor is

Γ^(σ)=d4Nψexp(ieiπqσaψia+iq/2A,aJAψAav4AaψAa).\displaystyle\hat{\Gamma}(\sigma)=\int\text{d}^{4N}\psi\exp\left(\text{i}e^{-\frac{\text{i}\pi}{q}}\sigma\prod_{a}\psi^{a}_{i}+\text{i}^{q/2}\sum_{A,a}J_{A}\psi^{a}_{A}-v^{4}\sum_{A}\prod_{a}\psi_{A}^{a}\right). (291)

Therefore we expect the half-wormhole saddle is given by

Γ=Γ^(0)\displaystyle\Gamma=\hat{\Gamma}(0) =\displaystyle= ABCDsgn(A,B,C,D)k=1p(JAkJBkJCkJDkδAkBkδCkBkδCkDkv4),\displaystyle\sum_{ABCD}\text{sgn}(A,B,C,D)\prod_{k=1}^{p}(J_{A_{k}}J_{B_{k}}J_{C_{k}}J_{D_{k}}-\delta_{A_{k}}^{B_{k}}\delta_{C_{k}}^{B_{k}}\delta_{C_{k}}^{D_{k}}v^{4})\,, (292)

which satisfies

Γ=0,Γ2=Γz434z42,\displaystyle\langle\Gamma\rangle=0\,,\qquad\langle\Gamma^{2}\rangle=\langle\Gamma z^{4}\rangle\approx 34\langle z^{4}\rangle^{2}\,, (293)
(z4z4Γ)2=z8z42+Γ22Γz40.\displaystyle\langle(z^{4}-\langle z^{4}\rangle-\Gamma)^{2}\rangle=\langle z^{8}\rangle-\langle z^{4}\rangle^{2}+\langle\Gamma^{2}\rangle-2\langle\Gamma z^{4}\rangle\approx 0\,. (294)

We find clearly that the contribution from this four-linked-wormhole saddle is not equal to the square of (two-linked) half-wormhole saddle. Even though we derive it in the 0-SYK toy model, it should exist in other SYK-like theory as long as the G,ΣG,\Sigma trick can be applied. We will present some more details about these more general discussions somewhere else.

7 SYK at one time point: Poisson distribution

Up to now we have only considered random couplings with continuous probability distributions. It is also interesting to consider random couplings that take discrete values such as the Poisson distribution.

In fact the Poisson distribution, whose PDF and moments are given by (596) and (597), can be regarded as an opposite extremum to what we have considered above in the sense that all the cumulants are equal Jnc=Nλ\langle J^{n}\rangle_{c}=N\lambda, n\forall n. From the gravity point of view, it means that all the wormholes with different number of boundaries have the same amplitude. Ensemble theory or theories with random coupling with Poisson distribution have been studied in Marolf:2020xie ; Peng:2020rno ; Peng:2021vhs . If we view the index ii of ψi\psi^{i} as the label of different time points, then the effect of ensemble average is to introduce (“non-local”) interaction between different time points. In particular, starting with action (154) we can compute the first few moments111111Here we have rescaled q2qq\rightarrow 2q, N2NN\rightarrow 2N.

z\displaystyle\langle z\rangle =d2NψeNiqλAψA1,\displaystyle=\int\text{d}^{2N}\psi\,e^{N\text{i}^{q}\lambda\sum_{A}\psi^{1}_{A}}, (295)
z2\displaystyle\langle z^{2}\rangle =d4NψeNiqλA(ψA1+ψA2)eNi2qλAψA1ψA2,\displaystyle=\int\text{d}^{4N}\psi\,e^{N\text{i}^{q}\lambda\sum_{A}(\psi_{A}^{1}+\psi_{A}^{2})}e^{N\text{i}^{2q}\lambda\sum_{A}\psi_{A}^{1}\psi_{A}^{2}}, (296)
z3\displaystyle\langle z^{3}\rangle =d6NψeNiqλA(ψA1+ψA2+ψA3)eNi2qλA(ψA1ψA2+ψA1ψA3+ψA2ψA3)eNi3qλAψA1ψA2ψA3.\displaystyle=\int\text{d}^{6N}\psi\,e^{N\text{i}^{q}\lambda\sum_{A}(\psi_{A}^{1}+\psi_{A}^{2}+\psi_{A}^{3})}e^{N\text{i}^{2q}\lambda\sum_{A}(\psi_{A}^{1}\psi_{A}^{2}+\psi_{A}^{1}\psi_{A}^{3}+\psi_{A}^{2}\psi_{A}^{3})}e^{N\text{i}^{3q}\lambda\sum_{A}\psi_{A}^{1}\psi_{A}^{2}\psi_{A}^{3}}\ . (297)

For a generic kk, we find

zk=d2kNψeλAn=1k1n!(iqi=1k(ψAi))n.\displaystyle\langle z^{k}\rangle=\int\text{d}^{2kN}\psi e^{\lambda\sum_{A}\sum_{n=1}^{k}\frac{1}{n!}(\text{i}^{q}\sum^{k}_{i=1}(\psi_{A}^{i}))^{n}}\ . (298)

Formally we can define

𝒵(λ)z=dψexp{NλA(eiqi=1ψAi1)}.\displaystyle{\cal Z}(\lambda)\equiv\langle z^{\infty}\rangle=\int\text{d}\psi\exp\left\{N\lambda\sum_{A}(e^{\text{i}^{q}\sum_{i=1}\psi_{A}^{i}}-1)\right\}\,. (299)

We can compute these moments by integrating out the fermions directly

zn=Pf(JA)n.\displaystyle\langle z^{n}\rangle=\langle\text{Pf}(J_{A})^{n}\rangle\ . (300)

However the ensemble average of PF(JA)n\text{PF}(J_{A})^{n} is very complicated. Alternatively, if we only care about the large NN behavior we can use the G,ΣG,\Sigma trick and do a saddle point approximation. For example, the G,ΣG,\Sigma expression of z\langle z\rangle is similar to (215)

z=𝑑Σ𝑑G(i)NΣNeNiqλGqq!eiNΣG.\displaystyle\langle z\rangle=\int d\Sigma dG(-\text{i})^{N}\Sigma^{N}e^{N\text{i}^{{q}}\lambda\frac{G^{{q}}}{{q}!}}e^{\text{i}N\Sigma G}. (301)

The saddle point equations are

ΣG=i,λ(q1)!(iG)q=1,\displaystyle\Sigma G=\text{i},\quad\frac{\lambda}{(q-1)!}(\text{i}G)^{q}=1\,, (302)

whose solutions are

iG=((q1)!λ)1/qe2mπiq,m=1,,q.\displaystyle\text{i}G=\left(\frac{(q-1)!}{\lambda}\right)^{1/q}e^{\frac{2m\pi\text{i}}{q}},\quad m=1,\dots,q\,. (303)

It has been argued in Saad:2021rcu these qq saddle points should be added together to reproduce the correct large NN behavior in a very similar calculation. We expect the same to apply in the current situation121212Here we have dropped the normalization factor iN\text{i}^{N}.

zDisk=eN(11q)(Nqλ(q1)!)pme2mπiq=qeN(11q)(Nqλ(q1)!)p,\displaystyle\langle z\rangle_{\text{Disk}}=e^{-N(1-\frac{1}{q})}\left(\frac{N^{q}\lambda}{(q-1)!}\right)^{p}\sum_{m}e^{\frac{2m\pi\text{i}}{q}}=qe^{-N(1-\frac{1}{q})}\left(\frac{N^{q}\lambda}{(q-1)!}\right)^{p}, (304)

where p=N/qp=N/q as before. Adding the 1-loop factor 1/q1/\sqrt{q} we end up with the correct large-NN behavior

zDisk+1 loop=1qeN(11q)(Nqλ(q1)!)p.\displaystyle\langle z\rangle_{\text{Disk}+1\text{ loop}}=\frac{1}{\sqrt{q}}e^{-N(1-\frac{1}{q})}\left(\frac{N^{q}\lambda}{(q-1)!}\right)^{p}. (305)

Other moments can be computed similarly. For example, to compute z2\langle z^{2}\rangle, we need to introduce three collective variables

G1=i<jiψi1ψj1,G2=i<jiψi2ψj2,G12=iψi1ψi2\displaystyle G_{1}=\sum_{i<j}\text{i}\psi_{i}^{1}\psi_{j}^{1},\quad G_{2}=\sum_{i<j}\text{i}\psi_{i}^{2}\psi_{j}^{2},\quad G_{12}=\sum_{i}\psi_{i}^{1}\psi_{i}^{2} (306)

such that

iqAψA1=G1qq!,iqAψA2=G2qq!,i2qAψA1ψA2=G122q(2q)!.\displaystyle\text{i}^{q}\sum_{A}\psi_{A}^{1}=\frac{G_{1}^{q}}{q!},\quad\text{i}^{q}\sum_{A}\psi_{A}^{2}=\frac{G_{2}^{q}}{q!},\quad\text{i}^{2q}\sum_{A}\psi_{A}^{1}\psi_{A}^{2}=\frac{G_{12}^{2q}}{(2q)!}. (307)

Imposing these relations with the help of a set of Lagrangian multiplier fields Σ1\Sigma_{1}, Σ2\Sigma_{2} and Σ12\Sigma_{12}, the z2\langle z^{2}\rangle can be expressed as

z2\displaystyle\langle z^{2}\rangle =\displaystyle= [d3Gid3Σi]eNλq!(G1q+G2q+q!(2q)!G122q)eiNi(ΣiGi)d2Nψe12ΨMΨ,\displaystyle\int[\text{d}^{3}G_{i}d^{3}\Sigma_{i}]e^{N\frac{\lambda}{q!}(G_{1}^{q}+G_{2}^{q}+\frac{q!}{(2q)!}G_{12}^{2q})}e^{\text{i}N\sum_{i}(\Sigma_{i}G_{i})}\int\text{d}^{2N}\psi e^{\frac{1}{2}{\Psi}M{\Psi}}, (308)
=\displaystyle= [d3Gid3Σi]det[Σ1Σ2A2Σ122I2N]eNλq!(G1q+G2q+q!(2q)!G122q)eiNi(ΣiGi)\displaystyle\int[\text{d}^{3}G_{i}d^{3}\Sigma_{i}]\sqrt{\text{det}[\Sigma_{1}\Sigma_{2}A^{2}-\Sigma_{12}^{2}I_{2N}]}e^{\frac{N\lambda}{q!}(G_{1}^{q}+G_{2}^{q}+\frac{q!}{(2q)!}G_{12}^{2q})}e^{\text{i}N\sum_{i}(\Sigma_{i}G_{i})} (309)
=\displaystyle= [d3Gid3Σi]i2Ndet[Σ1Σ2A+Σ12IN]eNλq!(G1q+G2q+q!(2q)!G122q)eiNi(ΣiGi)\displaystyle\int[\text{d}^{3}G_{i}d^{3}\Sigma_{i}]\text{i}^{2N}\text{det}[\sqrt{\Sigma_{1}\Sigma_{2}}A+\Sigma_{12}I_{N}]e^{N\frac{\lambda}{q!}(G_{1}^{q}+G_{2}^{q}+\frac{q!}{(2q)!}G_{12}^{2q})}e^{\text{i}N\sum_{i}(\Sigma_{i}G_{i})}
=\displaystyle= [d3Gid3Σi]i2N2((Σ12+Σ1Σ2)2N+(Σ12Σ1Σ2)2N)eNNλq!(G1q+G2q+q!(2q)!G122q)eNii(ΣiGi)\displaystyle\int[\text{d}^{3}G_{i}d^{3}\Sigma_{i}]\frac{\text{i}^{2N}}{2}\left((\Sigma_{12}+\sqrt{\Sigma_{1}\Sigma_{2}})^{2N}+(\Sigma_{12}-\sqrt{\Sigma_{1}\Sigma_{2}})^{2N}\right)e^{N\frac{N\lambda}{q!}(G_{1}^{q}+G_{2}^{q}+\frac{q!}{(2q)!}G_{12}^{2q})}e^{N\text{i}\sum_{i}(\Sigma_{i}G_{i})}
=\displaystyle= [d3Gid3Σi]i2Nk=1N(2N2k)Σ122N2k(Σ1Σ2)keNλq!(G1q+G2q+q!(2q)!G122q)eNii(ΣiGi)\displaystyle\int[\text{d}^{3}G_{i}d^{3}\Sigma_{i}]\text{i}^{2N}\sum_{k=1}^{N}{2N\choose 2k}\Sigma_{12}^{2N-2k}(\Sigma_{1}\Sigma_{2})^{k}e^{N\frac{\lambda}{q!}(G_{1}^{q}+G_{2}^{q}+\frac{q!}{(2q)!}G_{12}^{2q})}e^{N\text{i}\sum_{i}(\Sigma_{i}G_{i})} (311)

where we have defined

Ψ=(ψ11,,ψ2N1,ψ12,,ψ2N2),M=(Σ1AiΣ12I2NiΣ12I2NΣ2A),\displaystyle\Psi=\left(\psi_{1}^{1},\dots,\psi_{2N}^{1},\psi_{1}^{2},\dots,\psi_{2N}^{2}\right),\quad M=\begin{pmatrix}\Sigma_{1}A&-\text{i}\Sigma_{12}I_{2N}\\ \text{i}\Sigma_{12}I_{2N}&\Sigma_{2}A\\ \end{pmatrix}, (312)
A=AT,Aij=1,i<j.\displaystyle A=-A^{T},\quad A_{ij}=1,\quad\forall i<j. (313)

The saddle point equations lead to

iΣi+λ(q1)!Giq1=0,i=1,2,\displaystyle\text{i}\Sigma_{i}+\frac{\lambda}{(q-1)!}G_{i}^{q-1}=0,\quad i=1,2, (314)
iΣ12+λ(2q1)!G122q1=0,iΣiGi=2i.\displaystyle\text{i}\Sigma_{12}+\frac{\lambda}{(2q-1)!}G_{12}^{2q-1}=0\,,\quad\sum_{i}\Sigma_{i}G_{i}=2\text{i}\ . (315)

This set of equations have multiple solutions. For example, the wormhole saddle is

G1=G2=Σ1=Σ2=0,G12=(2(2q1)!λ)1/2qe2mπi2q,\displaystyle G_{1}=G_{2}=\Sigma_{1}=\Sigma_{2}=0,\quad G_{12}=\left(\frac{2(2q-1)!}{\lambda}\right)^{1/2q}e^{\frac{2m\pi\text{i}}{2q}}, (316)
z2WH+1loop=12qe2N(112q)((2N)2qλ2(2q1)!)p\displaystyle\langle z^{2}\rangle_{WH+1\text{loop}}=\frac{1}{\sqrt{2q}}e^{-2N(1-\frac{1}{2q})}\left(\frac{(2N)^{2q}\lambda}{2(2q-1)!}\right)^{p} (317)

and the disconnected saddle is

G12=Σ12=0,G1=G2=((q1)!λ)1/q,\displaystyle G_{12}=\Sigma_{12}=0,\quad G_{1}=G_{2}=\left(\frac{(q-1)!}{\lambda}\right)^{1/q}, (318)
z2disc+1loop=1qe2N(11q)(Nqλ(q1)!)2p=zDisk+1loop2.\displaystyle\langle z^{2}\rangle_{disc+1\text{loop}}=\frac{1}{q}e^{-2N(1-\frac{1}{q})}\left(\frac{N^{q}\lambda}{(q-1)!}\right)^{2p}=\langle z\rangle_{\text{Disk}+1\text{loop}}^{2}. (319)

The ratio of these two saddles is

z2WH+1loopz2disc+1loop=q2(q!222qeλq(2q)!)p.\displaystyle\frac{\langle z^{2}\rangle_{WH+1\text{loop}}}{\langle z^{2}\rangle_{disc+1\text{loop}}}=\sqrt{\frac{q}{2}}\left(\frac{q!^{2}2^{2q}}{e\lambda q(2q)!}\right)^{p}\,. (320)

In the large NN or p=N/qp=N/q limit, the wormhole saddle can dominate only when λ<q!222qeλq(2q)!(q2)12p\lambda<\frac{q!^{2}2^{2q}}{e\lambda q(2q)!}\left(\frac{q}{2}\right)^{\frac{1}{2p}} which is consistent with our previous results.

Then a natural question is that in this limit how about other n-boundary wormhole saddles? In the following let us focus on a particular nn-linked-wormhole saddles. When n=2kn=2k is even, the situation is similar to the one in section 6:

z2kconnected\displaystyle\langle z^{2k}\rangle_{\text{connected}} =\displaystyle= d4kNψdGdΣ2πexp(iNΣ(Gi2Na=12kψia))exp(Nλ(2q)!G2q)\displaystyle\int d^{4kN}\psi\text{d}G\frac{\text{d}\Sigma}{2\pi}\exp\left(\text{i}N\Sigma\left(G-\sum_{i}^{2N}\prod_{a=1}^{2k}\psi_{i}^{a}\right)\right)\exp\left(N\frac{\lambda}{(2q)!}G^{2q}\right) (321)
=\displaystyle= dGdΣ2π(iΣ)2Nexp(Nλ(2q)!G2q+iNΣG),\displaystyle\int\text{d}G\frac{\text{d}\Sigma}{2\pi}(\text{i}\Sigma)^{2N}\exp\left(\frac{N\lambda}{{(2q)}!}G^{2q}+\text{i}N\Sigma G\right)\,, (322)

where the collective variable GG is

G=i2Na=12kψia.\displaystyle G=\sum_{i}^{2N}\prod_{a=1}^{2k}\psi_{i}^{a}\ . (323)

The expression (322) is of the same form as (301) so the saddle point approximation is

z2k2kWH+1loop=z22WH+1loop=12qe2N(112q)((2N)2qλ2(2q1)!)p.\displaystyle\langle z^{2k}\rangle_{2k-WH+1\text{loop}}=\langle z^{2}\rangle_{2-WH+1\text{loop}}=\frac{1}{\sqrt{2q}}e^{-2N(1-\frac{1}{2q})}\left(\frac{(2N)^{2q}\lambda}{2(2q-1)!}\right)^{p}. (324)

When n=2k+1n=2k+1 is odd, the situation is similar to the one of n=1n=1:

z2k+1connected\displaystyle\langle z^{2k+1}\rangle_{\text{connected}} =\displaystyle= d(4k+2)NψdGdΣ2πexp(iNΣ(Gi<j2Na=12k+1ψiaa=12k+1ψja)exp(Nλq!Gq)\displaystyle\int d^{(4k+2)N}\psi\text{d}G\frac{\text{d}\Sigma}{2\pi}\exp\left(\text{i}N\Sigma(G-\sum_{i<j}^{2N}\prod_{a=1}^{2k+1}\psi_{i}^{a}\prod_{a=1}^{2k+1}\psi_{j}^{a}\right)\exp\left(\frac{N\lambda}{q!}G^{q}\right) (325)
=\displaystyle= dGdΣ2π(iΣ)2Nexp(Nλq!Gq+iNΣG),\displaystyle\int\text{d}G\frac{\text{d}\Sigma}{2\pi}(\text{i}\Sigma)^{2N}\exp\left(\frac{N\lambda}{{q}!}G^{q}+\text{i}N\Sigma G\right),

where the collective variable GG is obviously defined as

G=i<j2Na=12k+1ψiaa=12k+1ψja,\displaystyle G=\sum_{i<j}^{2N}\prod_{a=1}^{2k+1}\psi_{i}^{a}\prod_{a=1}^{2k+1}\psi_{j}^{a}, (326)

therefore the saddle point approximation is

z2k+12k+1HW+1loop=zDisk+1loop=1qeN(11q)(Nqλ(q1)!)p.\displaystyle\langle z^{2k+1}\rangle_{2k+1-HW+1\text{loop}}=\langle z\rangle_{\text{Disk}+1\text{loop}}=\frac{1}{\sqrt{q}}e^{-N(1-\frac{1}{q})}\left(\frac{N^{q}\lambda}{(q-1)!}\right)^{p}\ . (327)

These higher nn-linked-wormholes should be compared with the corresponding powers of the disk solution, and furthermore since z22WH+1loop1\langle z^{2}\rangle_{2-WH+1\text{loop}}\gg 1, we conclude that all these multiple-linked-wormholes are suppressed. In other words, the ensemble of zz can be approximated by a Gaussian when the ratio (320) is of order 1.

8 The Brownian SYK model

In this section, we study the wormhole and half-wormholes saddles in the Brownian SYK model Saad:2018bqo . In the Brownian SYK model, the couplings are only correlated at the same instant of time so that after integrating over the coupling we end up with a local effective action131313See Appendix (F) for general discussion on averaged model.. The quantity that is analogous to the partition function but with some information of real time evolution is

U(T)=𝐓ei0T𝑑tH(t).\displaystyle U(T)=\mathbf{T}e^{-\text{i}\int_{0}^{T}dtH(t)}\ . (328)

To check the nature of its fluctuations that is not caused by the phase factor, we consider the norm square of its trace

|TrU(T)|2.\displaystyle\left|{\rm Tr}\,U(T)\right|^{2}\ . (329)

This quantity is manifest real in the sense the complex conjugate maps TrU(T){\rm Tr}\,U(T) to TrU(T){\rm Tr}\,U(T)^{*}. The trace is over the Hilbert space, which has a path integral interpretation

TrU(T)\displaystyle{\rm Tr}\,U(T) =𝒟ψaexp{i0T𝑑t[i2ψatψa+Ja1aq(t)iq2ψa1aq]},\displaystyle=\int\mathcal{D}\psi_{a}\exp\left\{-\text{i}\int_{0}^{T}dt\left[-\frac{\text{i}}{2}\psi_{a}\partial_{t}\psi_{a}+J_{a_{1}\ldots a_{q}}(t)\text{i}^{\frac{q}{2}}\psi_{a_{1}\ldots a_{q}}\right]\right\}\,, (330)

where the Lagrangian density is manifestly real.

To compute (329), we introduce two replicas of fermions; ψ(L)\psi^{(L)} constitute the fermions in HH of UU and ψ(R)\psi^{(R)} in UU^{*}. Therefore the complex conjugate should map between ψ(L)\psi^{(L)} and ψ(R)\psi^{(R)}. One conventional way to define ψ(R)\psi^{(R)} from ψ(L)\psi^{(L)} is

ψa(R)=(ψa(L)).\displaystyle\psi^{(R)}_{a}=\left(\psi^{(L)}_{a}\right)^{*}\ . (331)

Then the complex conjugation of (330) is

TrU(T)\displaystyle{\rm Tr}\,U(T)^{*} =𝒟ψa(R)exp{i0T𝑑t[i2ψa(R)tψa(R)Ja1aq(t)iq2ψa1aq(R)]},\displaystyle=\int\mathcal{D}\psi_{a}^{(R)}\exp\left\{-\text{i}\int_{0}^{T}dt\left[\frac{\text{i}}{2}\psi^{(R)}_{a}\partial_{t}\psi^{(R)}_{a}-J_{a_{1}\ldots a_{q}}(t)\text{i}^{\frac{q}{2}}\psi^{(R)}_{a_{1}\ldots a_{q}}\right]\right\}\,, (332)

We can further do a field redefinition ψ(R)iψ(R)\psi^{(R)}\to\text{i}\psi^{(R)} so that the kinetic term has the “right” sign141414Here we choose to absorb an extra iNi^{N} phase factor into the definition of the path integral measure. There might be Nmod4N\bmod 4 effects that we will discuss separately.

TrU(T)\displaystyle{\rm Tr}\,U(T)^{*} =𝒟ψa(R)ei0T𝑑t[i2ψa(R)tψa(R)Ja1aq(t)(i)q2ψa1aq(R)],\displaystyle=\int\mathcal{D}\psi_{a}^{(R)}e^{-\text{i}\int_{0}^{T}dt\left[-\frac{\text{i}}{2}\psi^{(R)}_{a}\partial_{t}\psi^{(R)}_{a}-J_{a_{1}\ldots a_{q}}(t)(-\text{i})^{\frac{q}{2}}\psi^{(R)}_{a_{1}\ldots a_{q}}\right]}\,, (333)

Combining (330), with ψa\psi_{a} replaced by ψa(L)\psi_{a}^{(L)}, and (333), the quantity we would like to compute is

|TrU(T)|2=𝒟ψa(L)𝒟ψa(R)ei0T𝑑t[i2ψa(j)tψa(j)Ja1aq(t)(iq2ψa1aq(L)(i)q2ψa1aq(R))].\displaystyle|\operatorname{Tr}U(T)|^{2}=\int\mathcal{D}\psi_{a}^{(L)}\mathcal{D}\psi_{a}^{(R)}e^{\text{i}\int_{0}^{T}dt\left[\frac{\text{i}}{2}\psi_{a}^{(j)}\partial_{t}\psi_{a}^{(j)}-J_{a_{1}\ldots a_{q}}(t)\left(\text{i}^{\frac{q}{2}}\psi_{a_{1}\ldots a_{q}}^{(L)}-(-\text{i})^{\frac{q}{2}}\psi_{a_{1}\ldots a_{q}}^{(R)}\right)\right]}\ . (334)

A side remark is that the complex conjugation is closely related to time reversal symmetry 𝒯\mathcal{T}, and also because [𝒯,H]=0[\mathcal{T},H]=0, we expect TrU(T)=TrU(T){\rm Tr}\,U(T)^{*}={\rm Tr}\,U(-T). Indeed, we find

TrU(T)=𝒟ψa(R)exp{i0T𝑑t[i2ψa(R)tψa(R)+Ja1aq(t)iq2ψa1aq(R)]}\displaystyle{\rm Tr}\,U(-T)=\int\mathcal{D}\psi_{a}^{(R)}\exp\left\{-\text{i}\int_{0}^{-T}dt\left[-\frac{\text{i}}{2}\psi^{(R)}_{a}\partial_{t}\psi^{(R)}_{a}+J_{a_{1}\ldots a_{q}}(t)\text{i}^{\frac{q}{2}}\psi^{(R)}_{a_{1}\ldots a_{q}}\right]\right\} (335)
=𝒟ψa(R)exp{iT0𝑑t[i2ψa(R)tψa(R)+Ja1aq(t)iq2ψa1aq(R)]}\displaystyle=\int\mathcal{D}\psi_{a}^{(R)}\exp\left\{\text{i}\int_{-T}^{0}dt\left[-\frac{\text{i}}{2}\psi^{(R)}_{a}\partial_{t}\psi^{(R)}_{a}+J_{a_{1}\ldots a_{q}}(t)\text{i}^{\frac{q}{2}}\psi^{(R)}_{a_{1}\ldots a_{q}}\right]\right\} (336)
=𝒟ψa(R)exp{i0T𝑑t[i2ψa(R)tψa(R)+Ja1aq(t)iq2ψa1aq(R)]}=TrU(T),\displaystyle=\int\mathcal{D}\psi_{a}^{(R)}\exp\left\{\text{i}\int_{0}^{T}dt\left[-\frac{\text{i}}{2}\psi^{(R)}_{a}\partial_{t}\psi^{(R)}_{a}+J_{a_{1}\ldots a_{q}}(t)\text{i}^{\frac{q}{2}}\psi^{(R)}_{a_{1}\ldots a_{q}}\right]\right\}={\rm Tr}\,U(T)^{*}\,, (337)

where we simply use ψ(R)\psi^{(R)} to represent a different set of fermions that will be integrated over in the path integral; in particular, we do not think of them as the complex conjugate of the ψ(L)\psi^{(L)}. In the last line we assume the system to be invariant under time translation, and the last equality is clear from (332). Therefore the quantity we are interested can also be written as TrU(T)TrU(T){\rm Tr}\,U(T){\rm Tr}\,U(-T).

Note that the random couplings satisfy

JA=0,JA(t)JB(t)=δ(tt)δAB𝒥2,𝒥2=2J(q1)!Nq1,\displaystyle\langle J_{A}\rangle=0,\quad\langle J_{A}(t)J_{B}(t^{\prime})\rangle=\delta(t-t^{\prime})\delta_{AB}{\cal J}^{2},\quad{\cal J}^{2}=2J\frac{(q-1)!}{N^{q-1}}, (338)

and the our normalization of one-dimensional Majorana fermions is

{ψi,ψj}=δij.\displaystyle\{\psi_{i},\psi_{j}\}=\delta_{ij}\ . (339)

To simplify our notation, we simply denote |TrU(T)|2\left|{\rm Tr}\,U(T)\right|^{2} by |z|2\langle|z|^{2}\rangle in the rest computation.

8.1 |z|2\langle|z|^{2}\rangle in the Brownian SYK model: accurate evaluation

As argued in Saad:2018bqo , we focus on the time independent configurations. Therefore we can directly integrate out the fermions and averaging over the random coupling according to (338). In the large NN limit and for even qq this leads to

|z|2\displaystyle\langle|z|^{2}\rangle =𝒟GLRi𝒟ΣLR2πi/(TN)eN[log(2cosTΣLR2)2JTq2q+iq2JTqGLRq2T2ΣLRGLR].\displaystyle=\int_{\mathbb{R}}\mathcal{D}G_{LR}\int_{i\mathbb{R}}\frac{\mathcal{D}\Sigma_{LR}}{2\pi i/(TN)}e^{N\left[\log\left(2\cos\frac{T\Sigma_{LR}}{2}\right)-\frac{2JT}{q2^{q}}+i^{q}\frac{2JT}{q}G^{q}_{LR}-2\frac{T}{2}\Sigma_{LR}G_{LR}\right]}\ . (340)

The integration measure is normalized such that if we first to the GG integral then the Σ\Sigma integral, we get the result of free fermions |z|2|J=0=2N\langle|z|^{2}\rangle|_{J=0}=2^{N}. Notice that the GijG_{ij} function defined above is real under the complex conjugation (331). Making use of the identity

idΣ2πi/(TN)eNTΣ(Gx)=δ(Gx)\displaystyle\int_{\text{i}\mathbb{R}}\frac{d\Sigma}{2\pi i/(TN)}e^{-NT\Sigma(G-x)}=\delta(G-x) (341)

we get

|z|2=eJNTq2q1k=0N(Nk)eJTNq2q1(2kN1)qce.\displaystyle\langle|z|^{2}\rangle=e^{-\frac{JNT}{q2^{q-1}}}\sum_{k=0}^{N}{N\choose k}e^{\frac{JTN}{q2^{q-1}}(\frac{2k}{N}-1)^{q}}\equiv c_{e}\,. (342)

In the large-NN ( and also large-NTNT ) limit, the dominant contribution are determined from two factors: the combinatoric factor and the exponential. At early time T0T\ll 0, contributions from the different exponential factors are roughly the same, so the dominant term is determined from the largest term in the combinatoric factor

(NN/2)=N!((N/2)!)22N2πN,\displaystyle{N\choose N/2}=\frac{N!}{\left((N/2)!\right)^{2}}\sim 2^{N}\sqrt{\frac{2}{\pi N}}\,, (343)

which leads to the contribution

cs=2πN2NeJTNq2q1.\displaystyle c_{s}=\sqrt{\frac{2}{\pi N}}2^{N}e^{-\frac{JTN}{q2^{q-1}}}\ . (344)

At late time, the different exponential factors dominant over the combinatoric factors, so the dominant contribution is from the maximal exponential factor, which is at k=0,Nk=0,N with contributions to the sum being

cl=2.\displaystyle c_{l}=2\ . (345)

The behavior of |z|2\langle|z|^{2}\rangle is shown in Figure. 11 where the early time exponential decay and the late time constant behavior is manifest.

Refer to caption
Figure 11: The behavior of |z|2\langle|z|^{2}\rangle, where the horizontal axis is log10T\log_{10}T and the vertical axis is log10|z|2\log_{10}\langle|z|^{2}\rangle. The red dots are computed using (342) with q=4q=4, J=1J=1 and a cutoff at n=100n=100. The blue curve is the contribution from the trivial saddle (367), and the cyan curve is the contribution from the wormhole saddle (378).

8.2 |z|2\langle|z|^{2}\rangle in the Brownian SYK model: large-NN saddle point evaluation

In the following, we perform a saddle point analysis to reproduce these distinct behaviors. We deform the integration contour together with a change of variable

G=ieiπqg,Σ=eiπqσ.\displaystyle G=\text{i}e^{\frac{\text{i}\pi}{q}}g,\quad\Sigma=e^{-\frac{\text{i}\pi}{q}}\sigma\ . (346)

The action then reduces to

|z|2\displaystyle\langle|z|^{2}\rangle =𝒟g𝒟σLR2πi/(TN)eN[log(2cosTeiπqσ2)2JTq2q2JTqgqTiσg].\displaystyle=\int\mathcal{D}g\int\frac{\mathcal{D}\sigma_{LR}}{2\pi i/(TN)}e^{N\left[\log\left(2\cos\frac{Te^{-i\frac{\pi}{q}}\sigma}{2}\right)-\frac{2JT}{q2^{q}}-\frac{2JT}{q}g^{q}-Ti\sigma g\right]}\ . (347)

The equation of motion of the σ\sigma field leads to

g=i2eiπqtan(Tσ2eiπq),\displaystyle g=\frac{\text{i}}{2}e^{-\frac{\text{i}\pi}{q}}\tan(\frac{T\sigma}{2}e^{-\frac{\text{i}\pi}{q}})\,, (348)

while the equation of motion of gg is

igσ+2Jgq=0.\displaystyle\text{i}g\sigma+2Jg^{q}=0. (349)

The two equations indicate a condition that gg should satisfy

g+12eiπqtanh(JTgq1eiπq)=0.\displaystyle g+\frac{1}{2}e^{-\frac{\text{i}\pi}{q}}\tanh({JT}g^{q-1}e^{-\frac{\text{i}\pi}{q}})=0\ . (350)

Solutions to this equation are in general irrational. In the following we solve it with different approximations.

8.2.1 Saddle point solution: the q=1q=1 case

Formally, we can consider the q=1q=1 case where the saddle point solution can be found explicitly. In particular, when q=1q=1, the saddle point equation (350) reduces to

g=12tanh(TJ),σ=iJ.\displaystyle g=-\frac{1}{2}\tanh\left({TJ}\right)\,,\qquad\sigma=\text{i}J\ . (351)

The on-shell action (with the 1-loop correction) is then

|z|2\displaystyle\langle|z|^{2}\rangle =e2JTN2(eJT+eJT)N.\displaystyle=e^{{-\frac{2JTN}{2}}}\left(e^{-{JT}}+e^{{JT}}\right)^{N}\,. (352)

On the other hand, when q=1q=1, the summation expression (342) can be evaluated explicitly to

|z|2=e2JTN(eJT+eJT)N.\displaystyle\langle|z|^{2}\rangle=e^{{-{2JTN}{}}}\left(e^{-{JT}}+e^{{JT}}\right)^{N}\ . (353)

The exact result agrees with the above saddle point result.

8.2.2 Saddle point solution: q>1q>1 at short time Tσ1T\sigma\ll 1

In the saddle point approach, the effective action in the short time limit can be expanded into

S2=N(2T(Jgqqigσ2)18e2iπqσ2T2e4iπqσ4T4192+𝒪(T5)),\displaystyle S_{\text{2}}=N\left(2T\left(-\frac{Jg^{q}}{q}-\frac{ig\sigma}{2}\right)-\frac{1}{8}e^{-\frac{2i\pi}{q}}\sigma^{2}T^{2}-\frac{e^{-\frac{4i\pi}{q}}\sigma^{4}T^{4}}{192}+\mathcal{O}(T^{5})\right)\,, (354)

where

S2=SNlog(2)+NT2Jq2q,\displaystyle S_{\text{2}}=S-N\log(2)+NT\frac{2J}{q2^{q}}\,, (355)

is the part of the action that depends on the dynamical fields, in other words, the constant piece in SS has been factored out to define S2S_{2}. Notice that although we have T1T\ll 1 in this limit, we still want the saddle point approximation to be good, this means we want NT1NT\gg 1.

Before going to the details, we first discuss the region where this is a valid perturbative analysis. In the above expansion, the only gg dependence is in the (iσg2Jqgq)\left(-\frac{\text{i}\sigma g}{2}-\frac{J}{q}g^{q}\right) term, which means the set of saddle point equations always contain the following equation

2iJgq1=σ.\displaystyle 2\text{i}Jg^{q-1}=\sigma\ . (356)

This relation means the saddle point contribution to the on shell action has the general form

S2=N(2(i2)qq1(q1)TJ11qσ~qq1q18e2iπqσ2T2e4iπqσ4T4192+𝒪(T3)).\displaystyle S_{\text{2}}=N\left(\frac{2\left(-\frac{i}{2}\right)^{\frac{q}{q-1}}(q-1)TJ^{\frac{1}{1-q}}\tilde{\sigma}^{\frac{q}{q-1}}}{q}-\frac{1}{8}e^{-\frac{2i\pi}{q}}\sigma^{2}T^{2}-\frac{e^{-\frac{4i\pi}{q}}\sigma^{4}T^{4}}{192}+\mathcal{O}(T^{3})\right)\ . (357)

Next, we would like to make sure our expansion of the log\log term is valid, this requires

Tσ<1.\displaystyle T\sigma<1\ . (358)

The remaining terms could switch dominance depending on the value of σ\sigma in the saddle point solution.

short time: Tσqq1>T2σ2>T4σ4,0<σ<Tq12q,\displaystyle T\sigma^{\frac{q}{q-1}}>T^{2}\sigma^{2}>T^{4}\sigma^{4}\,,\quad\Leftrightarrow\quad 0<\sigma<T^{\frac{q-1}{2-q}}\,, (359)
intermediate time: T2σ2>Tσqq1>T4σ4,Tq12q<σ<T3(1+q)43q.\displaystyle T^{2}\sigma^{2}>T\sigma^{\frac{q}{q-1}}>T^{4}\sigma^{4}\,,\quad\Leftrightarrow\quad T^{\frac{q-1}{2-q}}<\sigma<T^{\frac{3(-1+q)}{4-3q}}\ . (360)

In all these cases the saddle point equation (350) reduces to the approximate form

g=ie2iπqσT,σ=igq1J.\displaystyle g=\text{i}e^{-\frac{2\text{i}\pi}{q}}\sigma T\,,\quad\sigma=\text{i}g^{q-1}J\,. (361)

\bullet Short time

In this case, the dominant term in the action is

S2=2NT(Jgqqgσ2)2N(((1)1q1(q1))TJ11qσqq1q2qq1).\displaystyle S_{\text{2}}=2NT\left(-\frac{Jg^{q}}{q}-\frac{g\sigma}{2}\right)\approx 2N\left(-\frac{\left((-1)^{\frac{1}{q-1}}(q-1)\right)TJ^{\frac{1}{1-q}}\sigma^{\frac{q}{q-1}}}{q2^{\frac{q}{q-1}}}\right)\ . (362)

The σT\sigma T term in the saddle point equation (361) can be dropped and the only solution is

g=0,σ=0,\displaystyle g=0\,,\qquad\sigma=0\,, (363)

This gives the following saddle contribution to the on-shell action

NTexp{N[log(2)2JTq2q]}=2NeJTNq2q1NT.\displaystyle{NT}\exp\left\{N\left[\log\left(2\right)-\frac{2JT}{q2^{q}}\right]\right\}=2^{N}e^{-\frac{JTN}{q2^{q-1}}}NT\ . (364)

Next we need to consider the quadratic fluctuations around this saddle

gg+δg,σσ+δσ.\displaystyle g\to g+\delta g\,,\qquad\sigma\to\sigma+\delta\sigma\ . (365)

This gives the 1-loop factor

1NT,\displaystyle\frac{1}{NT}\,, (366)

Therefore this saddle point approximation gives

|z|2s+1loop\displaystyle\langle|z|^{2}\rangle_{s+1\text{loop}} =exp{N[log(2)JTq2q]}=2NeJTNq2q1.\displaystyle=\exp\left\{N\left[\log\left(2\right)-\frac{JT}{q2^{q}}\right]\right\}=2^{N}e^{-\frac{JTN}{q2^{q-1}}}\ . (367)

We next want to compare this saddle point approximation with the exact result (342). It is clear that in the small TT region the dominant saddle should be (344). However, it is also clear that the kN2k\sim\frac{N}{2} terms in the sum should also give comparable contributions. Indeed, if we compare the result (344) with (367), we find

|z|2s+1loop=πN2cs.\displaystyle\langle|z|^{2}\rangle_{s+1\text{loop}}=\sqrt{\frac{\pi N}{2}}c_{s}\ . (368)

On the other hand, as we can check numerically,

ce=πN2cs,whenJ1,T1.\displaystyle c_{e}=\sqrt{\frac{\pi N}{2}}c_{s}\ ,\quad\text{when}\quad J\sim 1,\quad T\ll 1\ . (369)

An example of this numerical check is shown in Figure. 12. Or we can understand this approximation as the following. When JJ is of order 1 the exponent eTN(12kN)qeTNe^{TN(1-\frac{2k}{N})^{q}}\leq e^{TN}. So if T1NT\sim\frac{1}{N} this exponent is always of order 11 so that (342) can be approximated by eJNTq2q1k(Nk)=eJNTq2q12Ne^{-\frac{JNT}{q2^{q-1}}}\sum_{k}{N\choose k}=e^{-\frac{JNT}{q2^{q-1}}}2^{N}. Actually T1/NT\sim 1/N is only a sufficient condition for (369) to hold; as can be observed from the numerical data TT can be much larger than 1/N1/N.

Refer to caption
Figure 12: N=105N=10^{5}, q=4q=4, J=1J=1, T=0.1T=0.1. The vertical axis is ce/(πN2cs)c_{e}/\left(\sqrt{\frac{\pi N}{2}}c_{s}\right) with a cutoff 2πN2\sqrt{\pi N} in the summation (342). The horizontal axis is the number of terms kept in the summation.

Therefore, indeed we find our saddle point approximation agrees very well with the exact result.

\bullet Intermediate time

In this region both the σ2\sigma^{2} and σqq1\sigma^{\frac{q}{q-1}} terms in the action are roughly of the same order, so we need to solve the approximated saddle point equation (361). There are two solutions

σ=0,g=0,\displaystyle\sigma=0\,,\qquad g=0\,, (370)

and

g=(e2πim2iπqJT2)12q,σ=2iJgq1.\displaystyle g=\left(-e^{2\pi\text{i}m-\frac{2\text{i}\pi}{q}}\frac{JT}{2}\right)^{\frac{1}{2-q}}\,,\quad\sigma=2\text{i}Jg^{q-1}\ . (371)

and both saddles should in principle be taken into account. Notice that the expression seems to blow up at T0T\to 0, but we have fixed the range to be T0\sim T_{0} so the expression remains finite. The saddle point contribution from the non-trivial solution (371) is proportional to

(NT)2Ne2JNTq2qexp((121q)gq),\displaystyle(NT)2^{N}e^{-\frac{2JNT}{q2^{q}}}\exp\left(\left(\frac{1}{2}-\frac{1}{q}\right)g^{q}\right)\,, (372)

The loop correction around each of the saddle is

1NTq2.\displaystyle\frac{1}{NT\sqrt{q-2}}\ . (373)

The full contribution is

2Ne2JNTq2q1q2exp(N2q2qe2qmiπ2q(JT2)22q)\displaystyle 2^{N}e^{-\frac{2JNT}{q2^{q}}}\frac{1}{\sqrt{q-2}}\exp\left(N\frac{2-q}{2q}e^{\frac{2qm\text{i}\pi}{2-q}}\left(\frac{JT}{2}\right)^{\frac{2}{2-q}}\right) (374)

However, it is easy to check numerically that the contributions to the on-shell action from these saddles (374) never dominate when q>2q>2. Therefore the trivial saddle point always has larger contribution and dominate the path integral in this range of time.

8.2.3 Long time Tσ1T\sigma\gg 1

At long time, we can replace the cosh\cosh function by an exponential function. There are two choices, which leads to two different solutions

log(2cos(Tσ2eiπq))\displaystyle\log\left(2\cos(\frac{T\sigma}{2}e^{-\frac{\text{i}\pi}{q}})\right) i2Teiπqσ,(ieiπqσ)>0\displaystyle\sim\frac{\text{i}}{2}Te^{-\frac{\text{i}\pi}{q}}\sigma\,,\qquad\Re(\text{i}e^{-\text{i}\frac{\pi}{q}}\sigma)>0 (375)
log(2cos(Tσ2eiπq))\displaystyle\log\left(2\cos(\frac{T\sigma}{2}e^{-\frac{\text{i}\pi}{q}})\right) i2Teiπqσ,(ieiπqσ)<0.\displaystyle\sim-\frac{\text{i}}{2}Te^{-\frac{\text{i}\pi}{q}}\sigma\,,\qquad\Re(\text{i}e^{-\text{i}\frac{\pi}{q}}\sigma)<0\ . (376)

The solution of the saddle point equation in this case is time independent,

g=±12eiπq,σ=2iJgq1.\displaystyle g=\pm\frac{1}{2}e^{-\frac{\text{i}\pi}{q}}\,,\sigma=2\text{i}Jg^{q-1}\ . (377)

For even qq the on-shell actions of these saddles, including the 1-loop corrections, are also time independent

|z|2WH+11oop=2×1=2,\displaystyle\langle|z|^{2}\rangle_{\text{WH}+1\text{1oop}}=2\times 1=2\,, (378)

where the factor of 2 comes from adding up the contributions from the two saddles (377) and the result reproduces (345). In addition, the contribution from the trivial saddle vanishes at late time, so the non-trivial saddles (375) and (376) dominate. Since g0g\neq 0, these saddle points are identified with wormhole saddles.

8.3 |z|4\langle|z|^{4}\rangle in the Brownian SYK model

One of our goal in this section is to find possible half-wormhole saddles and study their relation to the wormhole saddle. To achieve this, it is helpful to first consider |z|4z1z2z3z4|z|^{4}\equiv z_{1}z_{2}z_{3}z_{4}:

𝒟4Nψei0T𝑑t[i2ψa(i)tψa(i)Ja1aq(t)(iq2ψa1aq(1)(i)q2ψa1aq(2)+iq2ψa1aq(3)(i)q2ψa1aq(4))].\displaystyle\int\mathcal{D}^{4N}\psi e^{\text{i}\int_{0}^{T}dt\left[\frac{\text{i}}{2}\psi_{a}^{(i)}\partial_{t}\psi_{a}^{(i)}-J_{a_{1}\ldots a_{q}}(t)\left(\text{i}^{\frac{q}{2}}\psi_{a_{1}\ldots a_{q}}^{(1)}-(-\text{i})^{\frac{q}{2}}\psi_{a_{1}\ldots a_{q}}^{(2)}+\text{i}^{\frac{q}{2}}\psi_{a_{1}\ldots a_{q}}^{(3)}-(-\text{i})^{\frac{q}{2}}\psi_{a_{1}\ldots a_{q}}^{(4)}\right)\right]}\,. (379)

We first compute the ensemble averaged version z1z2z3z4\langle z_{1}z_{2}z_{3}z_{4}\rangle:

|z|4\displaystyle\langle|z|^{4}\rangle =\displaystyle= 𝒟4Nψe0T𝑑t[12ψa(i)tψa(i)𝒥22(iq2ψa1aq(1)(i)q2ψa1aq(2)+iq2ψa1aq(3)(i)q2ψa1aq(4))2]\displaystyle\int{\cal D}^{4N}\psi\,e^{\int_{0}^{T}dt\left[-\frac{1}{2}\psi_{a}^{(i)}\partial_{t}\psi_{a}^{(i)}-\frac{{\cal J}^{2}}{2}\left(\text{i}^{\frac{q}{2}}\psi_{a_{1}\ldots a_{q}}^{(1)}-(-\text{i})^{\frac{q}{2}}\psi_{a_{1}\ldots a_{q}}^{(2)}+\text{i}^{\frac{q}{2}}\psi_{a_{1}\ldots a_{q}}^{(3)}-(-\text{i})^{\frac{q}{2}}\psi_{a_{1}\ldots a_{q}}^{(4)}\right)^{2}\right]} (380)
=\displaystyle= 𝒟4Nψ[a<b𝒟Gab𝒟Σab2πi/N]exp{0Tdt[12ψa(i)tψa(i)4JNTq2q+\displaystyle\int{\cal D}^{4N}\psi[\prod_{a<b}{\cal D}G_{ab}\frac{{\cal D}\Sigma_{ab}}{2\pi\text{i}/N}]\exp\left\{\int_{0}^{T}\text{d}t\left[-\frac{1}{2}\psi_{a}^{(i)}\partial_{t}\psi_{a}^{(i)}-\frac{4JNT}{q2^{q}}+\right.\right.
2a<b(JNqsabGabqNΣab2Gab+iΣab2ψiaψib)]}\displaystyle\qquad\left.\left.2\sum_{a<b}\left(\frac{JN}{q}s_{ab}G_{ab}^{q}-N\frac{\Sigma_{ab}}{2}G_{ab}+\sum_{i}\frac{\Sigma_{ab}}{2}\psi_{i}^{a}\psi_{i}^{b}\right)\right]\right\}

where s12=s14=s23=s34=iq,s13=s24=1s_{12}=s_{14}=s_{23}=s_{34}=\text{i}^{q},\quad s_{13}=s_{24}=-1 and the other orders of (a,b)(a,b) have been absorbed into the factor of 2. We again focus on the time-independent saddle points, and the integration over fermions gives

d4Nψexp(a<bΣabψiaψib)\displaystyle\int\text{d}^{4N}\psi\exp\left(\sum_{a<b}\Sigma_{ab}\psi_{i}^{a}\psi_{i}^{b}\right) (381)
=2N[cos(12(Σ14Σ23)2+(Σ13+Σ24)2+(Σ12Σ34)2)\displaystyle\quad=2^{N}\left[\cos\left(\frac{1}{2}\sqrt{(\Sigma_{14}-\Sigma_{23})^{2}+(\Sigma_{13}+\Sigma_{24})^{2}+(\Sigma_{12}-\Sigma_{34})^{2}}\right)\right.
+cos(12(Σ14+Σ23)2+(Σ13Σ24)2+(Σ12+Σ34)2)]N,\displaystyle\qquad\left.+\cos\left(\frac{1}{2}\sqrt{(\Sigma_{14}+\Sigma_{23})^{2}+(\Sigma_{13}-\Sigma_{24})^{2}+(\Sigma_{12}+\Sigma_{34})^{2}}\right)\right]^{N}\,, (382)

thus

|z|4\displaystyle\langle|z|^{4}\rangle =[a<b𝒟Gab𝒟Σab2πi/(NT)]e4JNTq2qexp(Seff),\displaystyle=\int[\prod_{a<b}{\cal D}G_{ab}\frac{{\cal D}\Sigma_{ab}}{2\pi\text{i}/(NT)}]e^{-\frac{4JNT}{q2^{q}}}\exp\left(S_{\text{eff}}\right)\,, (383)
Seff\displaystyle S_{\text{eff}} =2NTa<b(JqsabGabqΣab2Gab)+Nlog(eiTf++eiTf++eiTf+eiTf),\displaystyle=2NT\sum_{a<b}\left(\frac{J}{q}s_{ab}G_{ab}^{q}-\frac{\Sigma_{ab}}{2}G_{ab}\right)+N\log(e^{\text{i}Tf_{+}}+e^{-\text{i}Tf_{+}}+e^{\text{i}Tf_{-}}+e^{-\text{i}Tf_{-}})\,, (384)

with

f±=12(Σ14±Σ23)2+(Σ13Σ24)2+(Σ12±Σ34)2.\displaystyle f_{\pm}=\frac{1}{2}\sqrt{(\Sigma_{14}\pm\Sigma_{23})^{2}+(\Sigma_{13}\mp\Sigma_{24})^{2}+(\Sigma_{12}\pm\Sigma_{34})^{2}}\ . (385)

8.3.1 Exact evaluation

Similar to the exact calculation of |z|2\langle|z|^{2}\rangle we can integrate ΣAB\Sigma_{AB} first to obtain

|z|4\displaystyle\langle|z|^{4}\rangle =\displaystyle= e4JNTq2q(eif^+N+eif^+N+eif^N+eif^N)Ne2NTJqa<bsabGabq|Gab=0\displaystyle e^{-\frac{4JNT}{q2^{q}}}\left(e^{\frac{\text{i}\hat{f}_{+}}{N}}+e^{-\frac{\text{i}\hat{f}_{+}}{N}}+e^{\frac{\text{i}\hat{f}_{-}}{N}}+e^{-\frac{\text{i}\hat{f}_{-}}{N}}\right)^{N}e^{\frac{2NTJ}{q}\sum_{a<b}s_{ab}G_{ab}^{q}}\,\Big{|}_{G_{ab}=0}\, (386)
=\displaystyle= e4JNTq2q0niN,ini=Nei(n1n2)Nf^+ei(n3n4)Nf^e2NTJqa<bsabGabq|Gab=0,\displaystyle e^{-\frac{4JNT}{q2^{q}}}\sum_{0\leq n_{i}\leq N,\sum_{i}n_{i}=N}e^{\frac{\text{i}(n_{1}-n_{2})}{N}\hat{f}_{+}}e^{\frac{\text{i}(n_{3}-n_{4})}{N}\hat{f}_{-}}e^{\frac{2NTJ}{q}\sum_{a<b}s_{ab}G_{ab}^{q}}\,\Big{|}_{G_{ab}=0}\,, (387)

where we have introduced the differential operators

f^±=12(G14±G23)2+(G13G24)2+(G12±G34)2.\displaystyle\hat{f}_{\pm}=\frac{1}{2}\sqrt{(\partial_{G_{14}}\pm\partial_{G_{23}})^{2}+(\partial_{G_{13}}\mp\partial_{G_{24}})^{2}+(\partial_{G_{12}}\pm\partial_{G_{34}})^{2}}. (388)

Expanding the exponentials into Taylor series and keeping only the non-vanishing terms we get

ei(n1n2)Nf^+ei(n3n4)Nf^e2NTJqa<bsabGabq|Gab=0,\displaystyle e^{\frac{\text{i}(n_{1}-n_{2})}{N}\hat{f}_{+}}e^{\frac{\text{i}(n_{3}-n_{4})}{N}\hat{f}_{-}}e^{\frac{2NTJ}{q}\sum_{a<b}s_{ab}G_{ab}^{q}}\,\Big{|}_{G_{ab}=0},
=n=0m=0(i(n1n2)N)mq(i(n3n4)N)nqf^+mqf^nq(mq)!(nq)!e2NTJqa<bsabGabq|Gab=0,\displaystyle=\sum_{n=0}^{\infty}\sum_{m=0}^{\infty}\left(\frac{\text{i}(n_{1}-n_{2})}{N}\right)^{mq}\left(\frac{\text{i}(n_{3}-n_{4})}{N}\right)^{nq}\frac{\hat{f}_{+}^{mq}\hat{f}_{-}^{nq}}{(mq)!(nq)!}e^{\frac{2NTJ}{q}\sum_{a<b}s_{ab}G_{ab}^{q}}\,\Big{|}_{G_{ab}=0}\,, (389)

with

f^+mqf^nq\displaystyle\hat{f}_{+}^{mq}\hat{f}_{-}^{nq} =12(n+m)q0ki+m,ki+=m0kin,ki=n((G14+G23)k1+q(G14G23)k1q\displaystyle=\frac{1}{2^{(n+m)q}}\sum_{\begin{subarray}{c}0\leq k^{+}_{i}\leq m,\sum k_{i}^{+}=m\\ 0\leq k^{-}_{i}\leq n,\sum k^{-}_{i}=n\end{subarray}}\left((\partial_{G_{14}}+\partial_{G_{23}})^{k_{1}^{+}q}(\partial_{G_{14}}-\partial_{G_{23}})^{k_{1}^{-}q}\right.
×(G13G24)k2+q(G13+G24)k2q(G12+G34)k3+q(G12G34)k3q).\displaystyle\times\left.(\partial_{G_{13}}-\partial_{G_{24}})^{k_{2}^{+}q}(\partial_{G_{13}}+\partial_{G_{24}})^{k_{2}^{-}q}(\partial_{G_{12}}+\partial_{G_{34}})^{k_{3}^{+}q}(\partial_{G_{12}}-\partial_{G_{34}})^{k_{3}^{-}q}\right)\,. (390)

For each pair of differential operators in (390) the contribution can be obtained for example as

(G14+G23)k1+q(G14G23)k1qe2NTJq(G14q+G23q)|Gab=0\displaystyle(\partial_{G_{14}}+\partial_{G_{23}})^{k_{1}^{+}q}(\partial_{G_{14}}-\partial_{G_{23}})^{k_{1}^{-}q}e^{-\frac{2NTJ}{q}(G_{14}^{q}+G_{23}^{q})}\Big{|}_{G_{ab}=0}
=l1+=0,l1=0k1+,k1(k1+ql1+q)(k1ql1q)G14(l1++l1)qG23(k1++k1l1+l1)qeiq2NTJq(G14q+G23q)|Gab=0\displaystyle=\sum_{l^{+}_{1}=0,l^{-}_{1}=0}^{k_{1}^{+},k_{1}^{-}}{k_{1}^{+}q\choose l_{1}^{+}q}{k_{1}^{-}q\choose l_{1}^{-}q}\partial_{G_{14}}^{(l_{1}^{+}+l_{1}^{-})q}\partial_{G_{23}}^{(k_{1}^{+}+k_{1}^{-}-l_{1}^{+}-l_{1}^{-})q}e^{\text{i}^{q}\frac{2NTJ}{q}(G_{14}^{q}+G_{23}^{q})}\Big{|}_{G_{ab}=0} (391)
=(iq2NJTq)(k1++k1)ql1+=0,l1=0k1+,k1(k1+ql1+q)(k1ql1q)[(l1++l1)q]![(k1++k1l1+l1)q]!(l1++l1)!(k1++k1l1+l1)!\displaystyle=\left(\text{i}^{q}\frac{2NJT}{q}\right)^{(k_{1}^{+}+k_{1}^{-})q}\sum_{l^{+}_{1}=0,l^{-}_{1}=0}^{k_{1}^{+},k_{1}^{-}}{k_{1}^{+}q\choose l_{1}^{+}q}{k_{1}^{-}q\choose l_{1}^{-}q}\frac{[(l_{1}^{+}+l_{1}^{-})q]![(k_{1}^{+}+k_{1}^{-}-l_{1}^{+}-l_{1}^{-})q]!}{(l_{1}^{+}+l_{1}^{-})!(k_{1}^{+}+k_{1}^{-}-l_{1}^{+}-l_{1}^{-})!}
(iq2NJTq)(k1++k1)qΔ(k1+,k1).\displaystyle\equiv\left(\text{i}^{q}\frac{2NJT}{q}\right)^{(k_{1}^{+}+k_{1}^{-})q}\Delta(k_{1}^{+},k_{1}^{-}). (392)

Thus the full expression of (386) is

|z|4=e4JNTq2q0niN,ini=Nn=0m=00ki+m,ki+=m0kin,ki=nli+=0,li=0ki+,ki(in12N)mq(mq)!2mq(in34N)nq(nq)!2nq\displaystyle\langle|z|^{4}\rangle=e^{-\frac{4JNT}{q2^{q}}}\sum_{\begin{subarray}{c}0\leq n_{i}\leq N,\\ \sum_{i}n_{i}=N\end{subarray}}\sum_{n=0}^{\infty}\sum_{m=0}^{\infty}\sum_{\begin{subarray}{c}0\leq k^{+}_{i}\leq m,\sum k_{i}^{+}=m\\ 0\leq k^{-}_{i}\leq n,\sum k^{-}_{i}=n\end{subarray}}\sum_{l^{+}_{i}=0,l^{-}_{i}=0}^{k_{i}^{+},k_{i}^{-}}\frac{(\frac{\text{i}n_{12}}{N})^{mq}}{(mq)!2^{mq}}\frac{(\frac{\text{i}n_{34}}{N})^{nq}}{(nq)!2^{nq}}
(iq2NJTq)(k1++k1+k3++k3)q(2NJTq)(k2++k2)qi=13Δ(ki+,ki),\displaystyle\quad\left(\text{i}^{q}\frac{2NJT}{q}\right)^{(k_{1}^{+}+k_{1}^{-}+k_{3}^{+}+k_{3}^{-})q}\left(-\frac{2NJT}{q}\right)^{(k_{2}^{+}+k_{2}^{-})q}\prod_{i=1}^{3}\Delta(k_{i}^{+},k_{i}^{-})\,, (393)

where n12=n1n2n_{12}=n_{1}-n_{2} and n34=n3n4n_{34}=n_{3}-n_{4}. This is very complicated expression but at large T and large NN, the leading contributions come from the cases when ni=Nn_{i}=N and nj=0,jin_{j}=0,j\neq i. For each of these cases, we show in Appendix (D) that it contributes 2 when q=4mq=4m and 3 when q=4m+2q=4m+2. So in total, z4T\langle z_{4}\rangle_{T\rightarrow\infty} approaches to 8 when q=4mq=4m and 12 when q=4m+2q=4m+2.

Next we turn to the saddle point analysis and try to match the results.

8.3.2 Saddle point analysis

To make the integral (384) convergent, we do the following change of variables and deform the original integral contour so that the integral over gabg_{ab} and σab\sigma_{ab} are along the real lines

G12=ieiπqg12,G23=ieiπqg23,G14=ieiπqg14,G34=ieiπqg34,\displaystyle G_{12}=\text{i}e^{\frac{\text{i}\pi}{q}}g_{12},\quad G_{23}=\text{i}e^{\frac{\text{i}\pi}{q}}g_{23},\quad G_{14}=\text{i}e^{\frac{\text{i}\pi}{q}}g_{14},\quad G_{34}=\text{i}e^{\frac{\text{i}\pi}{q}}g_{34}, (394)
Σ12=eiπqσ12,Σ23=eiπqσ23,Σ14=eiπqσ14,Σ34=eiπqσ34,\displaystyle\Sigma_{12}=e^{-\frac{\text{i}\pi}{q}}\sigma_{12},\quad\Sigma_{23}=e^{-\frac{\text{i}\pi}{q}}\sigma_{23},\quad\Sigma_{14}=e^{-\frac{\text{i}\pi}{q}}\sigma_{14},\quad\Sigma_{34}=e^{-\frac{\text{i}\pi}{q}}\sigma_{34}, (395)
G13=g13,G24=g24,Σ13=iσ13,Σ24=iσ24.\displaystyle G_{13}=g_{13},\quad G_{24}=g_{24},\quad\Sigma_{13}=\text{i}\sigma_{13},\quad\Sigma_{24}=\text{i}\sigma_{24}\ . (396)

The effective action then becomes

S=2NTab(Jqgabqiσab2gab)+Nlog(eiTf++eiTf++eiTf+eiTf).\displaystyle S=2NT\sum_{ab}\left(-\frac{J}{q}g_{ab}^{q}-\text{i}\frac{\sigma_{ab}}{2}g_{ab}\right)+N\log(e^{\text{i}Tf_{+}}+e^{-\text{i}Tf_{+}}+e^{\text{i}Tf_{-}}+e^{-\text{i}Tf_{-}})\ . (397)

Let us again focus on the large TT limit. In this limit, we expect only one of the four exponentials e±iTf±e^{\pm\text{i}Tf_{\pm}} dominates the integral. To be explicit, let us assume the dominant one to be

eiaT2(Σ14+bΣ23)2+(Σ13bΣ24)2+(Σ12+bΣ34)2=eiaT2e2iπq(σ14+bσ23)2(σ13bσ24)2+e2iπq(σ12+bσ34)2,\displaystyle e^{\text{i}a\frac{T}{2}\sqrt{(\Sigma_{14}+b\Sigma_{23})^{2}+(\Sigma_{13}-b\Sigma_{24})^{2}+(\Sigma_{12}+b\Sigma_{34})^{2}}}=e^{\text{i}a\frac{T}{2}\sqrt{e^{-\frac{2\text{i}\pi}{q}}(\sigma_{14}+b\sigma_{23})^{2}-(\sigma_{13}-b\sigma_{24})^{2}+e^{-\frac{2\text{i}\pi}{q}}(\sigma_{12}+b\sigma_{34})^{2}}}\,, (398)

where a,ba,b can be ±1\pm 1. The saddle point equations leads to

g34=bg12,g24=bg13,g23=bg14,σab=2iJgabq1,g122e2iπq+g142e2iπqg132=1.\displaystyle g_{34}=bg_{12},\quad g_{24}=-bg_{13},\quad g_{23}=bg_{14},\quad\sigma_{ab}=2\text{i}Jg^{q-1}_{ab},\quad g_{12}^{2}e^{\frac{2\text{i}\pi}{q}}+g_{14}^{2}e^{\frac{2\text{i}\pi}{q}}-g_{13}^{2}=1\ . (399)

There is always a trivial saddle solution

gab=σab=0,|z|4trivial=22NeJNTq2q2,\displaystyle g_{ab}=\sigma_{ab}=0,\quad\langle|z|^{4}\rangle_{\text{trivial}}=2^{2N}e^{-\frac{JNT}{q2^{q-2}}}\,, (400)

which corresponds to the disconnected topology.

There is in addition a large number of non-trivial saddle solutions. ones with largest contributions to the on-shell actions, including the 1-loop corrections, are

g12=bg34=a12eiπq,g13=g14=g23=g24=0,|z|412=1,\displaystyle g_{12}=bg_{34}=a\frac{1}{2}e^{-\frac{\text{i}\pi}{q}},\quad g_{13}=g_{14}=g_{23}=g_{24}=0,\quad\langle|z|^{4}\rangle_{12}=1, (401)
g13=bg24=ai2,g12=g14=g23=g34=0,|z|413=e2(1+(i)q)NJTq,\displaystyle g_{13}=-bg_{24}=a\frac{\text{i}}{2},\quad g_{12}=g_{14}=g_{23}=g_{34}=0,\quad\langle|z|^{4}\rangle_{13}=e^{-\frac{2(1+(\text{i})^{q})NJT}{q}}, (402)
g14=bg23=a12eiπq,g13=g12=g34=g24=0,|z|414=1,\displaystyle g_{14}=bg_{23}=a\frac{1}{2}e^{-\frac{\text{i}\pi}{q}},\quad g_{13}=g_{12}=g_{34}=g_{24}=0,\quad\langle|z|^{4}\rangle_{14}=1\,, (403)

where the last equation in each line is the on-shell action of the corresponding solution. Apparently (401) and (403) correspond to the wormhole saddles appearing in |z|2\langle|z|^{2}\rangle and (402) correspond to the possible wormhole saddle appearing in z2\langle z^{2}\rangle. Therefore we find that in the late time

z4WH={3z2WH2q=4k+2,2z2WH2q=4k.\displaystyle\langle z^{4}\rangle_{WH}=\begin{cases}3\langle z^{2}\rangle^{2}_{WH}&q=4k+2,\\ 2\langle z^{2}\rangle^{2}_{WH}&q=4k.\end{cases} (404)

Notice that for a=±1a=\pm 1 the real parts of eiTfpe^{iTf_{p}} and eiTfpe^{-iTf_{p}} are the same, so it is not possible that only the a=1a=1 term dominate; what happens is that when the eiaTfpe^{iaTf_{p}} dominates, the eiaTfpe^{-iaTf_{p}} term also dominates and the resulting path integral result is just twice of the above results (401)-(403). Further taking into account that bb can be ±1\pm 1 we find that total saddle point contributions are 8 when q=4kq=4k and 12 when q=4k+2q=4k+2 as we found in the exact evaluation.

The interesting qmod4q\bmod 4 relation is consistent with the time reversal symmetry.

8.4 z2z^{2} at fixed coupling in the Brownian SYK model

In the following we consider the non-average expression (334), which we recall here

|TrU(T)|2=𝒟ψa(L)𝒟ψa(R)ei0T𝑑t[i2ψa(j)tψa(j)Ja1aq(t)(iq2ψa1aq(L)(i)q2ψa1aq(R))].\displaystyle|\operatorname{Tr}U(T)|^{2}=\int\mathcal{D}\psi_{a}^{(L)}\mathcal{D}\psi_{a}^{(R)}e^{\text{i}\int_{0}^{T}dt\left[\frac{\text{i}}{2}\psi_{a}^{(j)}\partial_{t}\psi_{a}^{(j)}-J_{a_{1}\ldots a_{q}}(t)\left(\text{i}^{\frac{q}{2}}\psi_{a_{1}\ldots a_{q}}^{(L)}-(-\text{i})^{\frac{q}{2}}\psi_{a_{1}\ldots a_{q}}^{(R)}\right)\right]}\ . (405)

We can again introduce

1=𝒟GLRi𝒟ΣLR2×2πi/Nedt𝑑tΣLR(t,t)2(NGLR(t,t)aψaL(t)ψaR(t))efLR(NGLR)fLR(aψaLψaR).\displaystyle 1=\int_{\mathbb{R}}\mathcal{D}G_{LR}\int_{\text{i}\mathbb{R}}\frac{\mathcal{D}\Sigma_{LR}}{2\times 2\pi\text{i}/N}e^{-\int\text{d}tdt^{\prime}{\frac{\Sigma_{LR}(t,t^{\prime})}{2}}\left(NG_{LR}(t,t^{\prime})-\sum_{a}\psi^{L}_{a}(t)\psi^{R}_{a}(t^{\prime})\right)}e^{f_{LR}\left(NG_{LR}\right)-f_{LR}\left(\sum_{a}\psi^{L}_{a}\psi^{R}_{a}\right)}\ . (406)

The quantity we would like to compute is

|TrU(T)|2\displaystyle|\operatorname{Tr}U(T)|^{2} =𝒟GLRi𝒟ΣLR2×2πi/N𝒟ψa(L)𝒟ψa(R)\displaystyle=\int_{\mathbb{R}}\mathcal{D}G_{LR}\int_{\text{i}\mathbb{R}}\frac{\mathcal{D}\Sigma_{LR}}{2\times 2\pi\text{i}/N}\int\mathcal{D}\psi_{a}^{(L)}\mathcal{D}\psi_{a}^{(R)} (407)
×exp{i0T𝑑t[i2ψa(j)tψa(j)Ja1aq(t)(iq2ψa1aq(L)(i)q2ψa1aq(R))]}\displaystyle\times\exp\left\{\text{i}\int_{0}^{T}dt\left[\frac{\text{i}}{2}\psi_{a}^{(j)}\partial_{t}\psi_{a}^{(j)}-J_{a_{1}\ldots a_{q}}(t)\left(\text{i}^{\frac{q}{2}}\psi_{a_{1}\ldots a_{q}}^{(L)}-(-\text{i})^{\frac{q}{2}}\psi_{a_{1}\ldots a_{q}}^{(R)}\right)\right]\right\} (408)
×edtdtΣLR(t,t)2(NGLR(t,t)aψaL(t)ψaR(t))efLR(NGLR)fLR(aψaLψaR).\displaystyle\times e^{-\int\text{d}t\text{d}t^{\prime}{\frac{\Sigma_{LR}(t,t^{\prime})}{2}}\left(NG_{LR}(t,t^{\prime})-\sum_{a}\psi^{L}_{a}(t)\psi^{R}_{a}(t^{\prime})\right)}e^{f_{LR}\left(NG_{LR}\right)-f_{LR}\left(\sum_{a}\psi^{L}_{a}\psi^{R}_{a}\right)}\ . (409)

We further rewrite

|TrU(T)|2=𝒟ΣLR2π/NΦ(ΣLR)Ψ(ΣLR).\displaystyle|\operatorname{Tr}U(T)|^{2}=\int_{\mathbb{R}}\frac{\mathcal{D}\Sigma_{LR}}{2\pi/N}\Phi(\Sigma_{LR})\Psi(\Sigma_{LR})\ . (410)

The ψa\psi_{a} independent part reads

Ψ(ΣLR)=𝒟GLRedtdtNΣLR(t,t)2GLR(t,t)efLR(NGLR),\displaystyle\Psi(\Sigma_{LR})=\int_{\mathbb{R}}\mathcal{D}G_{LR}e^{-\int\text{d}t\text{d}t^{\prime}N{\frac{\Sigma_{LR}(t,t^{\prime})}{2}}G_{LR}(t,t^{\prime})}e^{f_{LR}\left(NG_{LR}\right)}\,, (411)

where

fLR(NGLR)=2JNq1q(1)q2(NGLR,ii)q.\displaystyle f_{LR}(NG_{LR})=\frac{2J}{N^{q-1}q}(-1)^{\frac{q}{2}}\left(NG_{LR,ii}\right)^{q}\ . (412)

The ψa\psi_{a} dependent part is

Φ(ΣLR)=\displaystyle\Phi(\Sigma_{LR})= 𝒟ψa(L)𝒟ψa(R)e12dtdtΣLR(t,t)(ψaL(t)ψaR(t))fLR(aψaLψaR)\displaystyle\int\mathcal{D}\psi_{a}^{(L)}\mathcal{D}\psi_{a}^{(R)}e^{\frac{1}{2}\int\text{d}t\text{d}t^{\prime}\Sigma_{LR}(t,t^{\prime})\left(\psi^{L}_{a}(t)\psi^{R}_{a}(t^{\prime})\right)-f_{LR}(\sum_{a}\psi^{L}_{a}\psi^{R}_{a})} (413)
×exp{i0T𝑑t[i2ψa(j)tψa(j)Ja1aq(t)(iq2ψa1aq(L)(i)q2ψa1aq(R))]}.\displaystyle\times\exp\left\{\text{i}\int_{0}^{T}dt\left[\frac{\text{i}}{2}\psi_{a}^{(j)}\partial_{t}\psi_{a}^{(j)}-J_{a_{1}\ldots a_{q}}(t)\left(\text{i}^{\frac{q}{2}}\psi_{a_{1}\ldots a_{q}}^{(L)}-(-\text{i})^{\frac{q}{2}}\psi_{a_{1}\ldots a_{q}}^{(R)}\right)\right]\right\}\ . (414)

For the integral over GLRG_{LR} to converge, we rotate the contour so that

ΣLR=eiπqσ,GLR=ieiπqg.\displaystyle\Sigma_{LR}=e^{-\text{i}\frac{\pi}{q}}\sigma\,,\quad G_{LR}=\text{i}e^{\text{i}\frac{\pi}{q}}g\ . (415)

If we now compute the average of Φ(σ)\Phi(\sigma), we get

Φ(σ)\displaystyle\langle\Phi(\sigma)\rangle =𝒟ψa(L)𝒟ψa(R)e12dtdtΣLR(t,t)(ψaL(t)ψaR(t))fLR(aψaLψaR)\displaystyle=\int\mathcal{D}\psi_{a}^{(L)}\mathcal{D}\psi_{a}^{(R)}e^{\frac{1}{2}\int\text{d}t\text{d}t^{\prime}\Sigma_{LR}(t,t^{\prime})\left(\psi^{L}_{a}(t)\psi^{R}_{a}(t^{\prime})\right)-f_{LR}(\sum_{a}\psi^{L}_{a}\psi^{R}_{a})} (416)
×exp{0Tdt[12ψa(j)tψa(j)𝒥22(iq2ψa1aq(L)(i)q2ψa1aq(R))2]}\displaystyle\times\exp\left\{\int_{0}^{T}\text{d}t\left[\frac{-1}{2}\psi_{a}^{(j)}\partial_{t}\psi_{a}^{(j)}-\frac{{\cal J}^{2}}{2}\left(\text{i}^{\frac{q}{2}}\psi_{a_{1}\ldots a_{q}}^{(L)}-(-\text{i})^{\frac{q}{2}}\psi_{a_{1}\ldots a_{q}}^{(R)}\right)^{2}\right]\right\} (417)
=𝒟ψa(L)𝒟ψa(R)e12𝑑t𝑑tΣLR(t,t)(ψaL(t)ψaR(t))e0T𝑑t[12ψa(j)tψa(j)+𝒥22q(Nq)].\displaystyle=\int\mathcal{D}\psi_{a}^{(L)}\mathcal{D}\psi_{a}^{(R)}e^{\frac{1}{2}\int dtdt^{\prime}\Sigma_{LR}(t,t^{\prime})\left(\psi^{L}_{a}(t)\psi^{R}_{a}(t^{\prime})\right)}e^{-\int_{0}^{T}dt\left[\frac{1}{2}\psi_{a}^{(j)}\partial_{t}\psi_{a}^{(j)}+\frac{{\cal J}^{2}}{2^{q}}{N\choose q}\right]}\ . (418)

Integrating over the fermions, we get

Φ(σ)\displaystyle\langle\Phi(\sigma)\rangle =exp{𝒥22qT(Nq)+Nlog[2cos(Tσeiπq4)]}\displaystyle=\exp\left\{-\frac{{\cal J}^{2}}{2^{q}}T{N\choose q}+N\log\left[2\cos({\frac{T\sigma e^{-\frac{\text{i}\pi}{q}}}{4}})\right]\right\} (419)
exp{2JTNq2q+Nlog[2cos(Tσeiπq4)]},\displaystyle\sim\exp\left\{-\frac{2JTN}{q2^{q}}+N\log\left[2\cos({\frac{T\sigma e^{-\frac{\text{i}\pi}{q}}}{4}})\right]\right\}\,, (420)

where in the last line we substitute (338) and adopt the leading large-NN approximation. For example, at Σ=0\Sigma=0,

Φ(0)\displaystyle\langle\Phi(0)\rangle 2Ne2JTNq2q,\displaystyle\sim 2^{N}e^{-\frac{2JTN}{q2^{q}}}\,, (421)

and at T=0T=0,

Φ(0)\displaystyle\langle\Phi(0)\rangle 2N,\displaystyle\sim 2^{N}\,, (422)

which is independent of JJ and qq.

We still want to find the region in the σ\sigma plane where Φ\Phi is self-averaging. So next we compute the square

Φ(σ)2=\displaystyle\Phi(\sigma)^{2}= 𝒟ψa(L,1)𝒟ψa(R,1)𝒟ψa(L,2)𝒟ψa(R,2)exp{12dtΣLR(t,t)(ψaL,1ψaR,1+ψaL,2ψaR,2)\displaystyle\int\mathcal{D}\psi_{a}^{(L,1)}\mathcal{D}\psi_{a}^{(R,1)}\mathcal{D}\psi_{a}^{(L,2)}\mathcal{D}\psi_{a}^{(R,2)}\exp\left\{\frac{1}{2}\int\text{d}t\Sigma_{LR}(t,t^{\prime})\left(\psi^{L,1}_{a}\psi^{R,1}_{a}+\psi^{L,2}_{a}\psi^{R,2}_{a}\right)\right.
fLR(aψaL,1ψaR,1)fLR(aψaL,2ψaR,2)0Tdt(12ψa(j,1)tψa(j,1)+12ψa(j,2)tψa(j,2))\displaystyle-f_{LR}(\sum_{a}\psi^{L,1}_{a}\psi^{R,1}_{a})-f_{LR}(\sum_{a}\psi^{L,2}_{a}\psi^{R,2}_{a})-\int_{0}^{T}\text{d}t\left(\frac{1}{2}\psi_{a}^{(j,1)}\partial_{t}\psi_{a}^{(j,1)}+\frac{1}{2}\psi_{a}^{(j,2)}\partial_{t}\psi_{a}^{(j,2)}\right)
i0TdtdtJa1aq(t)(iq2ψa1aq(L,1)(i)q2ψa1aq(R,1)+iq2ψa1aq(L,2)(i)q2ψa1aq(R,2))}.\displaystyle\left.-\text{i}\int_{0}^{T}\text{d}t\text{d}t^{\prime}J_{a_{1}\ldots a_{q}}(t)\left(\text{i}^{\frac{q}{2}}\psi_{a_{1}\ldots a_{q}}^{(L,1)}-(-\text{i})^{\frac{q}{2}}\psi_{a_{1}\ldots a_{q}}^{(R,1)}+\text{i}^{\frac{q}{2}}\psi_{a_{1}\ldots a_{q}}^{(L,2)}-(-\text{i})^{\frac{q}{2}}\psi_{a_{1}\ldots a_{q}}^{(R,2)}\right)\right\}\ . (423)

Its average over the random coupling is

Φ(σ)2\displaystyle\langle\Phi(\sigma)^{2}\rangle =𝒟ψa(L,1)𝒟ψa(R,1)𝒟ψa(L,2)𝒟ψa(R,2)exp{12dtΣLR(t,t)(ψaL,1ψaR,1+ψaL,2ψaR,2)\displaystyle=\int\mathcal{D}\psi_{a}^{(L,1)}\mathcal{D}\psi_{a}^{(R,1)}\mathcal{D}\psi_{a}^{(L,2)}\mathcal{D}\psi_{a}^{(R,2)}\exp\left\{\frac{1}{2}\int\text{d}t\Sigma_{LR}(t,t^{\prime})\left(\psi^{L,1}_{a}\psi^{R,1}_{a}+\psi^{L,2}_{a}\psi^{R,2}_{a}\right)\right.
fLR(aψaL,1ψaR,1)fLR(aψaL,2ψaR,2)0Tdt(12ψa(j,1)tψa(j,1)+12ψa(j,2)tψa(j,2))\displaystyle-f_{LR}(\sum_{a}\psi^{L,1}_{a}\psi^{R,1}_{a})-f_{LR}(\sum_{a}\psi^{L,2}_{a}\psi^{R,2}_{a})-\int_{0}^{T}\text{d}t\left(\frac{1}{2}\psi_{a}^{(j,1)}\partial_{t}\psi_{a}^{(j,1)}+\frac{1}{2}\psi_{a}^{(j,2)}\partial_{t}\psi_{a}^{(j,2)}\right)
𝒥22(iq2ψa1aq(L,1)(i)q2ψa1aq(R,1)+iq2ψa1aq(L,2)(i)q2ψa1aq(R,2))2}.\displaystyle\left.-\frac{{\cal J}^{2}}{2}\left(\text{i}^{\frac{q}{2}}\psi_{a_{1}\ldots a_{q}}^{(L,1)}-(-\text{i})^{\frac{q}{2}}\psi_{a_{1}\ldots a_{q}}^{(R,1)}+\text{i}^{\frac{q}{2}}\psi_{a_{1}\ldots a_{q}}^{(L,2)}-(-\text{i})^{\frac{q}{2}}\psi_{a_{1}\ldots a_{q}}^{(R,2)}\right)^{2}\right\}\ . (424)

Expanding out the square, and introducing the extra GG, Σ\Sigma variables for the quantities between the two copies, we get

Φ(σ)2\displaystyle\langle\Phi(\sigma)^{2}\rangle =𝒟Gab𝒟Σab2πi/N𝒟ψa(L,1)𝒟ψa(R,1)𝒟ψa(L,2)𝒟ψa(R,2)exp{12dtΣLR(t,t)(ψaL,1ψaR,1+ψaL,2ψaR,2)\displaystyle=\int{\cal D}G_{ab}\frac{{\cal D}\Sigma_{ab}}{2\pi\text{i}/N}\mathcal{D}\psi_{a}^{(L,1)}\mathcal{D}\psi_{a}^{(R,1)}\mathcal{D}\psi_{a}^{(L,2)}\mathcal{D}\psi_{a}^{(R,2)}\exp\left\{\frac{1}{2}\int\text{d}t\Sigma_{LR}(t,t^{\prime})\left(\psi^{L,1}_{a}\psi^{R,1}_{a}+\psi^{L,2}_{a}\psi^{R,2}_{a}\right)\right.
fLR(aψaL,1ψaR,1)fLR(aψaL,2ψaR,2)0Tdt(12ψa(j,1)tψa(j,1)+12ψa(j,2)tψa(j,2))\displaystyle-f_{LR}(\sum_{a}\psi^{L,1}_{a}\psi^{R,1}_{a})-f_{LR}(\sum_{a}\psi^{L,2}_{a}\psi^{R,2}_{a})-\int_{0}^{T}\text{d}t\left(\frac{1}{2}\psi_{a}^{(j,1)}\partial_{t}\psi_{a}^{(j,1)}+\frac{1}{2}\psi_{a}^{(j,2)}\partial_{t}\psi_{a}^{(j,2)}\right)
+dt(Σ13ψA(L,1)ψA(L,2)+Σ23ψA(R,1)ψA(L,2)+Σ14ψA(L,1)ψA(R,2)+Σ24ψA(R,1)ψA(R,2))\displaystyle+\int\text{d}t\left(\Sigma_{13}\psi^{(L,1)}_{A}\psi^{(L,2)}_{A}+\Sigma_{23}\psi^{(R,1)}_{A}\psi^{(L,2)}_{A}+\Sigma_{14}\psi^{(L,1)}_{A}\psi^{(R,2)}_{A}+\Sigma_{24}\psi^{(R,1)}_{A}\psi^{(R,2)}_{A}\right)
Ndt(Σ13G13+Σ14G14+Σ23G23+Σ24G24)\displaystyle-N\int\text{d}t\left(\Sigma_{13}G_{13}+\Sigma_{14}G_{14}+\Sigma_{23}G_{23}+\Sigma_{24}G_{24}\right)
+0Tdt[2𝒥22q(Nq)2JNq(GLL,12q+GRR,12q)\displaystyle\left.+\int_{0}^{T}\text{d}t\left[-\frac{2{\cal J}^{2}}{2^{q}}{N\choose q}-\frac{2JN}{q}\left(G_{LL,12}^{q}+G_{RR,12}^{q}\right)\right.\right.
+2JNqiq(GLR,11q+GLR,22q+GLR,12q+GLR,21q)]}.\displaystyle+\left.\left.\frac{2JN}{q}\text{i}^{q}\left(G_{LR,11}^{q}+G_{LR,22}^{q}+G_{LR,12}^{q}+G_{LR,21}^{q}\right)\right]\right\}\ . (425)

Notice that in the first line the fermion bilinears are all in the same copy; these terms come from the Φ\Phi itself. In the second line, we added in a few other terms that couple the fermions between the two copies. Then we choose the fLRf_{LR} to cancel the GLR,11qG_{LR,11}^{q} and GLR,22qG_{LR,22}^{q} terms in the square, namely

fLR(aψaLψaR)=2JNq(1)q2GLR,iiq,\displaystyle f_{LR}(\sum_{a}\psi^{L}_{a}\psi^{R}_{a})=\frac{2JN}{q}(-1)^{\frac{q}{2}}G_{LR,ii}^{q}\,, (426)

so that the above result simplifies to

Φ(σ)2\displaystyle\langle\Phi(\sigma)^{2}\rangle =𝒟Gab𝒟Σab4πi/N𝒟ψa(L,1)𝒟ψa(R,1)𝒟ψa(L,2)𝒟ψa(R,2)exp{12dtΣLR(t,t)(ψaL,1ψaR,1+ψaL,2ψaR,2)\displaystyle=\int\int{\cal D}G_{ab}\frac{{\cal D}\Sigma_{ab}}{4\pi\text{i}/N}\mathcal{D}\psi_{a}^{(L,1)}\mathcal{D}\psi_{a}^{(R,1)}\mathcal{D}\psi_{a}^{(L,2)}\mathcal{D}\psi_{a}^{(R,2)}\exp\left\{\frac{1}{2}\int\text{d}t\Sigma_{LR}(t,t^{\prime})\left(\psi^{L,1}_{a}\psi^{R,1}_{a}+\psi^{L,2}_{a}\psi^{R,2}_{a}\right)\right.
0Tdt(12ψa(j,1)tψa(j,1)+12ψa(j,2)tψa(j,2))\displaystyle-\int_{0}^{T}\text{d}t\left(\frac{1}{2}\psi_{a}^{(j,1)}\partial_{t}\psi_{a}^{(j,1)}+\frac{1}{2}\psi_{a}^{(j,2)}\partial_{t}\psi_{a}^{(j,2)}\right)
+12dt(Σ13ψA(L,1)ψA(L,2)+Σ23ψA(R,1)ψA(L,2)+Σ14ψA(L,1)ψA(R,2)+Σ24ψA(R,1)ψA(R,2))\displaystyle+\frac{1}{2}\int\text{d}t\left(\Sigma_{13}\psi^{(L,1)}_{A}\psi^{(L,2)}_{A}+\Sigma_{23}\psi^{(R,1)}_{A}\psi^{(L,2)}_{A}+\Sigma_{14}\psi^{(L,1)}_{A}\psi^{(R,2)}_{A}+\Sigma_{24}\psi^{(R,1)}_{A}\psi^{(R,2)}_{A}\right)
N2dt(Σ13G13+Σ14G14+Σ23G23+Σ24G24)\displaystyle-\frac{N}{2}\int\text{d}t\left(\Sigma_{13}G_{13}+\Sigma_{14}G_{14}+\Sigma_{23}G_{23}+\Sigma_{24}G_{24}\right)
+0Tdt[2𝒥22q(Nq)2JNq(GLL,12q+GRR,12q)+2JNqiq(GLR,12q+GRL,12q)].\displaystyle+\int_{0}^{T}\text{d}t\left[-\frac{2{\cal J}^{2}}{2^{q}}{N\choose q}-\frac{2JN}{q}\left(G_{LL,12}^{q}+G_{RR,12}^{q}\right)+\frac{2JN}{q}\text{i}^{q}\left(G_{LR,12}^{q}+G_{RL,12}^{q}\right)\right]\ . (427)

Integrating out the fermions with the help of the relation (382), shortening the labels according to (L,1)1(L,1)\to 1, (R,1)2(R,1)\to 2, (L,2)3(L,2)\to 3, (R,2)4(R,2)\to 4, and using the fact that Σ12=Σ34=ΣLR\Sigma_{12}=\Sigma_{34}=\Sigma_{LR} by construction, we get

Φ(σ)2\displaystyle\langle\Phi(\sigma)^{2}\rangle =dGABdΣAB4πi/(NT)exp{N[log(2cos(T4(Σ14Σ23)2+(Σ13+Σ24)2))\displaystyle=\int\frac{\text{d}G_{AB}\text{d}\Sigma_{AB}}{4\pi\text{i}/(NT)}\exp\left\{N\left[\log\left(2\cos\left(\frac{T}{4}\sqrt{(\Sigma_{14}-\Sigma_{23})^{2}+(\Sigma_{13}+\Sigma_{24})^{2}}\right)\right)\right.\right.
+2cos(T4(Σ14+Σ23)2+(Σ13Σ24)2+4Σ2)]NT2ABΣABGAB\displaystyle\left.+2\cos\left(\frac{T}{4}\sqrt{(\Sigma_{14}+\Sigma_{23})^{2}+(\Sigma_{13}-\Sigma_{24})^{2}+4\Sigma^{2}}\right)\right]-\frac{NT}{2}\sum_{AB}\Sigma_{AB}G_{AB}
2𝒥22qT(Nq)2JNTq(G13q+G24q)+2JNTqiq(G14q+G23q)},\displaystyle\left.-\frac{2{\cal J}^{2}}{2^{q}}T{N\choose q}-\frac{2JNT}{q}\left(G_{13}^{q}+G_{24}^{q}\right)+\frac{2JNT}{q}\text{i}^{q}\left(G_{14}^{q}+G_{23}^{q}\right)\right\}\,, (428)

where again we have focused on the time-independent saddles.

8.4.1 Exact computation

We can first evaluate the integral explicitly. The calculation is similar to the one of |z|4\langle|z|^{4}\rangle

Φ(σ)2\displaystyle\langle\Phi(\sigma)^{2}\rangle =e4JNTq2q(eif^+N+eif^+N+eif^N+eif^N)Ne2NTJqsABGABq|GAB=0\displaystyle=e^{-\frac{4JNT}{q2^{q}}}\left(e^{\frac{\text{i}\hat{f^{\prime}}_{+}}{N}}+e^{-\frac{\text{i}\hat{f^{\prime}}_{+}}{N}}+e^{\frac{\text{i}\hat{f^{\prime}}_{-}}{N}}+e^{-\frac{\text{i}\hat{f^{\prime}}_{-}}{N}}\right)^{N}e^{\frac{2NTJ}{q}\sum s_{AB}G_{AB}^{q}}\,\Big{|}_{G_{AB}=0}\, (429)
=e4JNTq2q0niN,ini=Nei(n1n2)Nf^+ei(n3n4)Nf^e2NTJqsABGABq|GAB=0,\displaystyle=e^{-\frac{4JNT}{q2^{q}}}\sum_{0\leq n_{i}\leq N,\sum_{i}n_{i}=N}e^{\frac{\text{i}(n_{1}-n_{2})}{N}\hat{f^{\prime}}_{+}}e^{\frac{\text{i}(n_{3}-n_{4})}{N}\hat{f^{\prime}}_{-}}e^{\frac{2NTJ}{q}\sum s_{AB}G_{AB}^{q}}\,\Big{|}_{G_{AB}=0}\,, (430)

where s14=s23=iq,s13=s24=1s_{14}=s_{23}=\text{i}^{q},s_{13}=s_{24}=-1 and we have introduced the differential operators

f^+=12(G14+G23)2+(G13G24)2+4T2Σ2,\displaystyle\hat{f^{\prime}}_{+}=\frac{1}{2}\sqrt{(\partial_{G_{14}}+\partial_{G_{23}})^{2}+(\partial_{G_{13}}-\partial_{G_{24}})^{2}+4T^{2}\Sigma^{2}}, (431)
f^=12(G14G23)2+(G13+G24)2,\displaystyle\hat{f^{\prime}}_{-}=\frac{1}{2}\sqrt{(\partial_{G_{14}}-\partial_{G_{23}})^{2}+(\partial_{G_{13}}+\partial_{G_{24}})^{2}}, (432)

Following a similar calculation as in section (386), we get

Φ(σ)2=e4JNTq2q0niN,ini=Nn=0m=00k1+,k2+,k3m,0k1,k2nk1++k2++k3=mki=nli+=0,li=0ki+,ki(in12N)mq(mq)!2mq(in34N)nq(nq)!2nq\displaystyle\langle\Phi(\sigma)^{2}\rangle=e^{-\frac{4JNT}{q2^{q}}}\sum_{\begin{subarray}{c}0\leq n_{i}\leq N,\\ \sum_{i}n_{i}=N\end{subarray}}\sum_{n=0}^{\infty}\sum_{m=0}^{\infty}\sum_{\begin{subarray}{c}0\leq k^{+}_{1},k^{+}_{2},k_{3}\leq m,\\ 0\leq k^{-}_{1},k^{-}_{2}\leq n\\ \sum k_{1}^{+}+k_{2}^{+}+k_{3}=m\\ \sum k^{-}_{i}=n\end{subarray}}\sum_{l^{+}_{i}=0,l^{-}_{i}=0}^{k_{i}^{+},k_{i}^{-}}\frac{(\frac{\text{i}n_{12}}{N})^{mq}}{(mq)!2^{mq}}\frac{(\frac{\text{i}n_{34}}{N})^{nq}}{(nq)!2^{nq}}
(iq2NJTq)(k1++k1)q(2NJTq)(k2++k2)q(2Teπqσ)k3i=12Δ(ki+,ki),\displaystyle\quad\left(\text{i}^{q}\frac{2NJT}{q}\right)^{(k_{1}^{+}+k_{1}^{-})q}\left(-\frac{2NJT}{q}\right)^{(k_{2}^{+}+k_{2}^{-})q}(2Te^{-\frac{\pi}{q}}\sigma)^{k_{3}}\prod_{i=1}^{2}\Delta(k_{i}^{+},k_{i}^{-})\,, (433)

where n12=n1n2n_{12}=n_{1}-n_{2} and n34=n3n4n_{34}=n_{3}-n_{4}. In the large TT limit, as we show in the exact computation of z4\langle z^{4}\rangle, the leading contributions are the summation of the contribution obtained by keeping only one exponential differential operator in (429). The operator e±if^Ne^{\pm\frac{\text{i}\hat{f^{\prime}}_{-}}{N}} contributes 1 when q=4mq=4m and 2 when q=4m+2q=4m+2. The result of operator e±if^+Ne^{\pm\frac{\text{i}\hat{f^{\prime}}_{+}}{N}} will be a monomial of σ\sigma while its expression is not very illuminating so we omit here.

8.4.2 Saddle point computation

We deform the contour so that the integral converge. In the current case, the contours are rotated as

G13=g13,G24=g24,Σ13=iσ13,Σ24=iσ24\displaystyle G_{13}=g_{13}\,,\quad G_{24}=g_{24}\,,\quad\Sigma_{13}=i\sigma_{13}\,,\quad\Sigma_{24}=i\sigma_{24} (434)

and

G14=ieiπqg14,G23=ieiπqg23,Σ14=eiπqσ14,Σ23=eiπqσ23,\displaystyle G_{14}=\text{i}e^{\text{i}\frac{\pi}{q}}g_{14}\,,\quad G_{23}=\text{i}e^{\text{i}\frac{\pi}{q}}g_{23}\,,\quad\Sigma_{14}=e^{-\text{i}\frac{\pi}{q}}\sigma_{14}\,,\quad\Sigma_{23}=e^{-\text{i}\frac{\pi}{q}}\sigma_{23}\,, (435)

so that the effective action for computing Φ(σ)2\langle\Phi(\sigma)^{2}\rangle is 151515To get rid of the factor 1/21/2 we have scaled the variables as σAB2σAB\sigma_{AB}\rightarrow 2\sigma_{AB} .

S=4NTJq2q2NTab(Jqgabq+i2σabgab)+Nlog(2cos(Tf+)+2cos(Tf)),\displaystyle S=-\frac{4NTJ}{q2^{q}}-2NT\sum_{ab}\left(\frac{J}{q}g_{ab}^{q}+\frac{\text{i}}{2}\sigma_{ab}g_{ab}\right)+N\log(2\cos(Tf^{\prime}_{+})+2\cos(Tf^{\prime}_{-})),

where we have defined

f+=12e2iπq(σ14+σ23)2(σ13σ24)2+e2iπq4σ2,\displaystyle f^{\prime}_{+}=\frac{1}{2}\sqrt{e^{-\frac{2\text{i}\pi}{q}}(\sigma_{14}+\sigma_{23})^{2}-(\sigma_{13}-\sigma_{24})^{2}+e^{-\frac{2\text{i}\pi}{q}}4\sigma^{2}}, (437)
f=12e2iπq(σ14σ23)2(σ13+σ24)2.\displaystyle f^{\prime}_{-}=\frac{1}{2}\sqrt{e^{-\frac{2\text{i}\pi}{q}}(\sigma_{14}-\sigma_{23})^{2}-(\sigma_{13}+\sigma_{24})^{2}}. (438)

The equation of motion of gABg_{AB} gives universally

iσab=2Jgabq1,(ab)=(13),(24),(14),(23).\displaystyle-\text{i}\sigma_{ab}=2Jg_{ab}^{q-1},\quad(ab)=(13),(24),(14),(23)\ . (439)

As discussed in the previous |z|2\langle|z|^{2}\rangle computation, the equation of motion of σab\sigma_{ab} depends on the value of time TT.

Very short time

When time TT is very short, the cosine function can be approximated by a constant. Then the only saddle point is

gab=0,σab=0.\displaystyle g_{ab}=0\,,\qquad\sigma_{ab}=0\ . (440)

The on-shell action, including a 1-loop determinant 1(TN)4\frac{1}{(TN)^{4}}, around this saddle point is

Φ(σ)2=(2cos(Teiπqσ))Ne4JTNq2q=(2cos(TΣ))Ne4JTNq2q.\displaystyle\langle\Phi(\sigma)^{2}\rangle=\left(2\cos\left(Te^{-\frac{\text{i}\pi}{q}}\sigma\right)\right)^{N}e^{-\frac{4JTN}{q2^{q}}}=\left(2\cos\left(T\Sigma\right)\right)^{N}e^{-\frac{4JTN}{q2^{q}}}\ . (441)

The results agree with Φ2\langle\Phi\rangle^{2} as can be seen from (420). Notice that the trivial saddle (440) remains a saddle point for a large range of TT, and the on-shell action around this saddle point, ie (441), is true within this large range.

In summary, we have shown that at very short time the trivial saddles dominate the Φ2\langle\Phi^{2}\rangle which approximately equals to Φ2\langle\Phi\rangle^{2}, we conclude that at short time the trivial saddle dominates and Φ(σ)\Phi(\sigma) is self-averaging.

Short time

When the time is larger, the determinant term in the action cannot be approximated by a constant, and thus we expand it to the second order of TT. The saddle point equation for σab\sigma_{ab} is now

σ13T=4ig13,σ24T=4ig24\displaystyle\sigma_{13}T=4\text{i}g_{13}\,,\qquad\sigma_{24}T=4\text{i}g_{24} (442)
σ14T=4ie2πiqg14,σ23T=4ie2πiqg23.\displaystyle\sigma_{14}T=-4\text{i}e^{\frac{2\pi\text{i}}{q}}g_{14}\,,\qquad\sigma_{23}T=-4\text{i}e^{\frac{2\pi\text{i}}{q}}g_{23}\ . (443)

Combining with (439), we get the non-trivial saddle solutions

2JT=g13q2=g24q2=e2iπqg23q2=e2iπqg14q2\displaystyle\frac{2}{JT}=g_{13}^{q-2}=g_{24}^{q-2}=-e^{-\frac{2\text{i}\pi}{q}}g_{23}^{q-2}=-e^{-\frac{2\text{i}\pi}{q}}g_{14}^{q-2} (444)

Each solution gives a contribution to Φ(σ)2\langle\Phi(\sigma)^{2}\rangle as

22Ne2JTNq2q1Ne2iπqσ2T2/4e4Nq22q(JT/2)22q(e4m1πi2q+e4m2πi2q)e4Nq22q(JT/2)22q(e4m3πi2q+e4m4πi2q),\displaystyle 2^{2N}e^{-\frac{2JTN}{q2^{q-1}}-Ne^{-\frac{2\text{i}\pi}{q}}\sigma^{2}T^{2}/4}e^{4N\frac{q-2}{2q}(JT/2)^{\frac{2}{2-q}}(e^{\frac{4m_{1}\pi\text{i}}{2-q}}+e^{\frac{4m_{2}\pi\text{i}}{2-q}})}e^{-4N\frac{q-2}{2q}(JT/2)^{\frac{2}{2-q}}(e^{\frac{4m_{3}\pi\text{i}}{2-q}}+e^{\frac{4m_{4}\pi\text{i}}{2-q}})}\,, (445)

where mim_{i} are integers. It is cumbersome in the following to discuss the most generic case with arbitrary qq, we therefore focus on the q=4q=4 case. Then the above contribution uces to

22Ne2JTNq2q1+iNσ2T2/4.\displaystyle 2^{2N}e^{-\frac{2JTN}{q2^{q-1}}+\text{i}N\sigma^{2}T^{2}/4}\ . (446)

Recall that in the end we need to perform the integral (410), as we argued in last section we have to deform the contour to a steepest contour such that the (446) vanishes at infinity. It means that this saddle point behaves like

22Ne2JTNq2q1N|Im(σ2)T2/4.\displaystyle 2^{2N}e^{-\frac{2JTN}{q2^{q-1}}-N|\text{Im}(\sigma^{2})T^{2}/4}\ . (447)

so it is sub-dominant comparing with the trivial saddle. Therefore in the regime of time, Φ\Phi is still self-averaging.

Long time

At very long time we rewrite the log\log term as

log(eiTf++eiTf++eiTf+eiTf).\displaystyle\log\left(e^{\text{i}Tf^{\prime}_{+}}+e^{-\text{i}Tf^{\prime}_{+}}+e^{\text{i}Tf^{\prime}_{-}}+e^{\text{i}Tf^{\prime}_{-}}\right)\,. (448)

When TT is sufficiently large, only one term in the above expression is dominant in the log\log function, there are thus two different cases to be discussed; either the term that is independent of σ\sigma or the term that depend on σ\sigma. When the σ\sigma independent term dominates, the saddle point contribution is independent on σ\sigma, which is not related to either the wormhole or half-wormhole contribution that we are interested in. Therefore it is tentative to consider the case where the σ\sigma dependent term dominates

log(eiTf++eiTf++eiTf+eiTf)iaT2e2iπq(σ14+σ23)2(σ13σ24)2+e2iπq4σ2,\displaystyle\log\left(e^{\text{i}Tf^{\prime}_{+}}+e^{-\text{i}Tf^{\prime}_{+}}+e^{\text{i}Tf^{\prime}_{-}}+e^{\text{i}Tf^{\prime}_{-}}\right)\approx{\frac{\text{i}aT}{2}\sqrt{e^{-\frac{2\text{i}\pi}{q}}(\sigma_{14}+\sigma_{23})^{2}-(\sigma_{13}-\sigma_{24})^{2}+e^{-\frac{2\text{i}\pi}{q}}4\sigma^{2}}},

where a=±1a=\pm 1. The saddle point equations are

2g13+iag13q1g132q2e2iπqg142q2+e2iπqσ~2=0,\displaystyle 2g_{13}+\frac{\text{i}ag_{13}^{q-1}}{\sqrt{g_{13}^{2q-2}-e^{-\frac{2\text{i}\pi}{q}}g_{14}^{2q-2}+e^{-\frac{2\text{i}\pi}{q}}\tilde{\sigma}^{2}}}=0, (449)
2g14iae2iπqg14q1g132q2e2iπqg142q2+e2iπqσ~2=0,\displaystyle 2g_{14}-\frac{\text{i}ae^{-\frac{2\text{i}\pi}{q}}g_{14}^{q-1}}{\sqrt{g_{13}^{2q-2}-e^{-\frac{2\text{i}\pi}{q}}g_{14}^{2q-2}+e^{-\frac{2\text{i}\pi}{q}}\tilde{\sigma}^{2}}}=0, (450)
g14=g23,g24=g13,σ14=σ23,σ24=σ13.\displaystyle g_{14}=g_{23}\,,\quad g_{24}=-g_{13}\,,\quad\sigma_{14}=\sigma_{23}\,,\quad\sigma_{24}=-\sigma_{13}\ . (451)

where σ~2=σ2/J2\tilde{\sigma}^{2}=\sigma^{2}/J^{2} and the equation of motion (439) has been used. Solutions of this set of equations are complicated in general, here we only provide the solutions for q=4q=4. First let us consider the non-trivial solutions gab0g_{ab}\neq 0. There is a set of 8 solutions

g13=±eiπ8σ~a,g14=±e7iπ8σ~a,g13=±eiπ8σ~a,g14=e7iπ8σ~a,\displaystyle g_{13}=\pm e^{\frac{\text{i}\pi}{8}}\sqrt{\frac{\tilde{\sigma}}{a}}\,,\quad g_{14}=\pm e^{\frac{7\text{i}\pi}{8}}\sqrt{\frac{\tilde{\sigma}}{a}}\,,\quad g_{13}=\pm e^{\frac{\text{i}\pi}{8}}\sqrt{\frac{\tilde{\sigma}}{a}}\,,\quad g_{14}=\mp e^{\frac{7\text{i}\pi}{8}}\sqrt{\frac{\tilde{\sigma}}{a}}, (452)
g13=±e5iπ8σ~a,g14=±e3iπ8σ~a,g13=±e5iπ8σ~a,g14=e3iπ8σ~a,\displaystyle g_{13}=\pm e^{\frac{5\text{i}\pi}{8}}\sqrt{\frac{\tilde{\sigma}}{a}}\,,\quad g_{14}=\pm e^{\frac{3\text{i}\pi}{8}}\sqrt{\frac{\tilde{\sigma}}{a}}\,,\quad g_{13}=\pm e^{\frac{5\text{i}\pi}{8}}\sqrt{\frac{\tilde{\sigma}}{a}}\,,\quad g_{14}=\mp e^{\frac{3\text{i}\pi}{8}}\sqrt{\frac{\tilde{\sigma}}{a}}\,, (453)

where all the unlisted variables are given by (451). The first four solutions lead to the same on-shell action

12eiaTNe2iπqσ2e4JTNq2q,\displaystyle\frac{1}{2}e^{\text{i}aTN\sqrt{e^{-\frac{2\text{i}\pi}{q}}\sigma^{2}}}e^{-\frac{4JTN}{q2^{q}}}\,, (454)

while the last four saddles lead to another on-shell action

14eiaTNe2iπqσ2e4JTNq2q.\displaystyle\frac{1}{4}e^{\text{i}aTN\sqrt{e^{-\frac{2\text{i}\pi}{q}}\sigma^{2}}}e^{-\frac{4JTN}{q2^{q}}}\ . (455)

Comparing with the contribution from the trivial saddles

Φ2=(2cos(Teiπqσ))Ne2JTNqe±iTNe2iπqσ2e4JTN2qq,\displaystyle\langle\Phi^{2}\rangle=\left(2\cos\left(Te^{-\frac{\text{i}\pi}{q}}\sigma\right)\right)^{N}e^{-\frac{2JTN}{q}}\sim e^{\pm\text{i}TN\sqrt{e^{-\frac{2\text{i}\pi}{q}}\sigma^{2}}}e^{-\frac{4JTN}{2^{q}q}}\,, (456)

we find that all of them are comparable with the trivial saddle. We have seen this phenomenon in the 0-dimensional SYK model. We expect these non-trivial saddles to not contribute to the path integral, which could be checked in the Lefschetz thimble analysis.

In the end let us consider the special non-trivial solution with g23=0=g14g_{23}=0=g_{14}. Focusing on the case of a=1a=1, the saddle point which has the proper fall-off behavior at infinite of σ\sigma is

exp(32NTσ43),σ.\displaystyle\exp\left(-\frac{3}{2}NT\sigma^{\frac{4}{3}}\right),\quad\sigma\rightarrow\infty. (457)

Then we plot the region where this non-trivial saddle dominates over the trivial saddle in Fig. 13.

Refer to caption
Figure 13: The shaded region is where non-trivial saddle dominates. We have set J=1/2J=1/2.

It turns out that wormhole saddle σwh=±12ieiπ4\sigma_{wh}=\pm\frac{1}{2}\text{i}e^{\frac{\text{i}\pi}{4}} is not in this region. For example at J=1/2J=1/2, Re(Φ2(σwh)non-trivialΦ2(σwh)trivial)=0.23\text{Re}(\langle\Phi^{2}(\sigma_{wh})\rangle_{\text{non-trivial}}-\langle\Phi^{2}(\sigma_{wh})\rangle_{\text{trivial}})=-0.23. It suggests that like the 0-dimensional model, the wormhole saddle of |z|2\langle|z|^{2}\rangle are within the self-averaging region of the σ\sigma plane.

8.5 Half-wormholes

With all the results in the previous sections, we expect that the integral (410) can be approximated by

|z|2\displaystyle|z|^{2} \displaystyle\approx 0dσ2πΨ(σ)Φ(σ)+( half-wormholes at σ=0),\displaystyle\int_{\mathbb{R}-0}\frac{\text{d}\sigma}{2\pi}\Psi(\sigma)\Phi(\sigma)+\left(\text{ half-wormholes at }\sigma=0\right), (458)
=\displaystyle= 0dσ2πΨ(σ)Φ(σ)+( half-wormholes at σ=0),\displaystyle\int_{\mathbb{R}-0}\frac{\text{d}\sigma}{2\pi}\Psi(\sigma)\langle\Phi(\sigma)\rangle+\left(\text{ half-wormholes at }\sigma=0\right), (459)
=\displaystyle= |z|2+Φ(0),\displaystyle\langle|z|^{2}\rangle+\Phi(0)\,, (460)

at late time. First we notice that in the late time

Φ(0)=2Ne4JTNq2q0\displaystyle\langle\Phi(0)\rangle=2^{N}e^{-\frac{4JTN}{q2^{q}}}\rightarrow 0 (461)

therefore at least (460) is consistent. To confirm (460) we need to compute Φ(0)2\langle\Phi(0)^{2}\rangle, Φ(0)|z|\langle\Phi(0)|z|\rangle and the error:

Error=(|z|2|z|2Φ(0))2=|z|4|z|22+Φ(0)22|z|2Φ(0).\displaystyle\text{Error}=\langle(|z|^{2}-\langle|z|^{2}\rangle-\Phi(0))^{2}\rangle=\langle|z|^{4}\rangle-\langle|z|^{2}\rangle^{2}+\langle\Phi(0)^{2}\rangle-2\langle|z|^{2}\Phi(0)\rangle. (462)

In the late time, each term in (462) is given by the non-trivial saddle points. It is clear that Φ(0)2\langle\Phi(0)^{2}\rangle, Φ(0)|z|\langle\Phi(0)|z|\rangle can be obtain from |z|4\langle|z|^{4}\rangle by setting g12=0g_{12}=0. Therefore we have

Error={(31+24)z2WH2=0q=4k+2,(21+12)z2WH2=0q=4k,\displaystyle\text{Error}=\begin{cases}(3-1+2-4)\langle z^{2}\rangle^{2}_{WH}=0&q=4k+2,\\ (2-1+1-2)\langle z^{2}\rangle^{2}_{WH}=0&q=4k,\end{cases} (463)

therefore (460) indeed is good in the large TT limit.

9 Modified Brownian SYK model

In this section, we study the wormhole and half-wormholes contributions in some modified (Brownian) SYK model.

9.1 Brownian SYK with non-vanishing mean value

Let us first consider to turn on the mean value of the random couplings:

JA=JA(0)=μ,JA(t)JB(t)=δ(tt)(δABτ2+μ2),\displaystyle\langle J_{A}\rangle=J_{A}^{(0)}=\mu,\quad\langle J_{A}(t)J_{B}(t^{\prime})\rangle=\delta(t-t^{\prime})(\delta_{AB}\tau^{2}+\mu^{2})\,, (464)

and in this section we use the convention {ψi,ψj}=2hδi,j\{\psi_{i},\psi_{j}\}=2h\delta_{i,j}.

Taking the disorder averaging of the coupling we obtain the averaged theory

z(T)J\displaystyle\langle z(T)\rangle_{J} =dNψeSa,\displaystyle=\int d^{N}\psi\,e^{-S_{a}}\,, (465)
Sa\displaystyle S_{a} =120T𝑑tiNψitψiiq/2𝑑tAJA(0)ψAτ22𝑑t(AψA2)\displaystyle=\frac{1}{2}\int_{0}^{T}dt\sum_{i}^{N}\psi_{i}\partial_{t}\psi_{i}-\text{i}^{q/2}\int dt\sum_{A}J_{A}^{(0)}\psi_{A}-\frac{\tau^{2}}{2}\int dt(\sum_{A}\psi_{A}^{2}) (466)

We can convert the effective Hamiltonian of the averaged theory as a spin system

zJ=Tr(eT),=iq/2AJA(0)ψAt22Ahq=iq/2AJA(0)ψAτ22(Nq)hq.\displaystyle\langle z\rangle_{J}=\text{Tr}(e^{-T{\cal H}})\,,\quad{\cal H}=-\text{i}^{q/2}\sum_{A}J_{A}^{(0)}\psi_{A}-\frac{t^{2}}{2}\sum_{A}h^{q}=-\text{i}^{q/2}\sum_{A}J_{A}^{(0)}\psi_{A}-\frac{\tau^{2}}{2}{N\choose q}h^{q}\ . (467)

When μ=0\mu=0, the averaged partition function is given by

zJ=eTτ22(Nq)hq2NeTE0,E0=τ22(Nq)hqτ22Nqhq.\displaystyle\langle z\rangle_{J}=e^{T\frac{\tau^{2}}{2}{N\choose q}h^{q}}\equiv 2^{N}e^{TE_{0}},\quad E_{0}=\frac{\tau^{2}}{2}{N\choose q}h^{q}\sim\frac{\tau^{2}}{2}N^{q}h^{q}\,. (468)

When μ0\mu\neq 0, we have to evaluate the trace

zJ=eTE0Tr(eTiq/2μAψA)=eTE0dNψiexp(Tiq/2iψi).\displaystyle\langle z\rangle_{J}=e^{TE_{0}}\text{Tr}(e^{T\text{i}^{q/2}\mu\sum_{A}\psi_{A}})=e^{TE_{0}}\int\text{d}^{N}\psi_{i}\exp(T\text{i}^{q/2}\sum_{i}\psi_{i})\ . (469)

However there is no simple expression for z\langle z\rangle. We first consider the simplest case with q=1q=1

If=dNψiexp(aiψi).\displaystyle I_{f}=\int\text{d}^{N}\psi_{i}\exp(a\sum_{i}\psi_{i})\ . (470)

The idea is to transfer the Majorana fermions to Dirac fermions which have a well-defined rules of integrals. Assuming the total number of fermions is even N=2KN=2K then we introduce KK Dirac fermions as

ci=12h(ψ2i1iψ2i),ci=12h(ψ2i1+iψ2i),i=1,,K,\displaystyle c_{i}=\frac{1}{2\sqrt{h}}(\psi_{2i-1}-\text{i}\psi_{2i}),\quad c_{i}^{\dagger}=\frac{1}{2\sqrt{h}}(\psi_{2i-1}+\text{i}\psi_{2i}),\quad i=1,\dots,K\,, (471)
ψ2i1=h(ci+ci),ψ2i=ih(cici),\displaystyle\psi_{2i-1}=\sqrt{h}(c_{i}+c_{i}^{\dagger}),\quad\psi_{2i}=\text{i}\sqrt{h}(c_{i}-c_{i}^{\dagger})\,, (472)

which obey

{ci,cj}\displaystyle\{c_{i},c_{j}\} =\displaystyle= {ci,cj}=0,{ci,cj}=δij\displaystyle\{c_{i}^{\dagger},c_{j}^{\dagger}\}=0,\quad\{c_{i},c_{j}^{\dagger}\}=\delta_{ij}\, (473)

The integration measure changes as

𝒟ψ2i𝒟ψ2i1=2h𝒟ci𝒟ci.\displaystyle{\cal D}\psi_{2i}{\cal D}\psi_{2i-1}=2h{\cal D}c_{i}{\cal D}c_{i}^{\dagger}\,. (475)

Thus the integral can be evaluated as

I1\displaystyle I_{1} =\displaystyle= (2h)Ki𝒟ci𝒟ciexp(aiKh[(1+i)ci+(1i)ci])\displaystyle(2h)^{K}\int\prod_{i}{\cal D}c_{i}{\cal D}c_{i}^{\dagger}\exp\left(a\sum_{i}^{K}\sqrt{h}[(1+\text{i})c_{i}+(1-\text{i})c_{i}^{\dagger}]\right) (476)
=\displaystyle= (2h)K(2cosh(2ah))K=(2h)N[cosh(2ah)]N/2.\displaystyle(2h)^{K}(2\cosh(\sqrt{2}ah))^{K}=(2h)^{N}[\cosh(\sqrt{2}ah)]^{N/2}\,. (477)

Now we let us consider the case of q=2q=2

I2(a)=dNψiexp(a2ijψiAijψj),with (Aij=Aij=a,i<j),\displaystyle I_{2}(a)=\int\text{d}^{N}\psi_{i}\exp(\frac{a}{2}\sum_{i\neq j}\psi_{i}A_{ij}\psi_{j}),\quad\text{with }\quad\left(A_{ij}=-A_{ij}=a,\quad i<j\right), (478)

which looks like a Gaussian but we need to replace ψi\psi_{i} with cic_{i}:

I2=(2h)Ni𝒟ci𝒟cie\displaystyle I_{2}=(\sqrt{2h})^{N}\int\prod_{i}{\cal D}c_{i}{\cal D}c_{i}^{\dagger}\,e^{{\cal H}} (479)
=(iah(iK[cicicici]+2i<j[cicjcicj])+2ahi<j[cicj+cicj])\displaystyle\quad{\cal H}=\left(\text{i}ah(\sum_{i}^{K}[c_{i}^{\dagger}c_{i}-c_{i}c_{i}^{\dagger}]+2\sum_{i<j}[c_{i}c_{j}-c_{i}^{\dagger}c_{j}^{\dagger}])+2ah\sum_{i<j}[c_{i}^{\dagger}c_{j}+c_{i}c_{j}^{\dagger}]\right) (480)

To get an idea how to compute this integral let us consider a simple case of N=4N=4:

ψ1=h(c1+c1),ψ2=ih(c1c1),ψ3=h(c2+c2),ψ4=ih(c2c2),\displaystyle\psi_{1}=\sqrt{h}(c_{1}+c_{1}^{\dagger}),\psi_{2}=\text{i}\sqrt{h}(c_{1}-c_{1}^{\dagger}),\psi_{3}=\sqrt{h}(c_{2}+c_{2}^{\dagger}),\psi_{4}=\text{i}\sqrt{h}(c_{2}-c_{2}^{\dagger}), (481)
i<jψiψj=ih(c1c1c1c1+c2c2c2c2+2c1c22c1c2)+2h(c1c2+c1c2).\displaystyle\sum_{i<j}\psi_{i}\psi_{j}=\text{i}h(c_{1}^{\dagger}c_{1}-c_{1}c_{1}^{\dagger}+c_{2}^{\dagger}c_{2}-c_{2}c_{2}^{\dagger}+2c_{1}c_{2}-2c_{1}^{\dagger}c_{2}^{\dagger})+2h(c_{1}^{\dagger}c_{2}+c_{1}c_{2}^{\dagger}). (482)

We have four different states |Ψi|\Psi_{i}\rangle:

|00,c1|00,c2|00,c1c2|00.\displaystyle|00\rangle,\quad c_{1}^{\dagger}|00\rangle,\quad c_{2}^{\dagger}|00\rangle,\quad c_{1}^{\dagger}c_{2}^{\dagger}|00\rangle. (483)

So the operator i<jψiψj\sum_{i<j}\psi_{i}\psi_{j} can be written as a 4×44\times 4 matrix :

i<jψiψj=(2ih002ih002h002h002ih002ih)\displaystyle\sum_{i<j}\psi_{i}\psi_{j}=\left(\begin{array}[]{cccc}-2\text{i}h&0&0&-2\text{i}h\\ 0&0&-2h&0\\ 0&2h&0&0\\ -2\text{i}h&0&0&2\text{i}h\\ \end{array}\right) (488)

with 4 eigenvalues {±i2h,±i22h}\{\pm\text{i}2h,\pm\text{i}2\sqrt{2}h\} so path integral over cic_{i} and cic_{i}^{\dagger} can be computed as

iΨi|eaiψiψj|Ψi=2(cos(2ah)+cos(22ah)).\displaystyle\sum_{i}\langle\Psi_{i}|e^{a\sum_{i}\psi_{i}\psi_{j}}|\Psi_{i}\rangle=2\left(\cos(2ah)+\cos(2\sqrt{2}ah)\right)\,. (489)

For example of N=6N=6, the corresponding matrix is

i<jψiψj=(3ih0002ih2ih2ih00ih2h2h0002ih02hih2h0002ih02h2hih0002ih2ih000ih2h2h02ih0002hih2h02ih0002h2hih002ih2ih2ih0003ih)\displaystyle\sum_{i<j}\psi_{i}\psi_{j}=\left(\begin{array}[]{cccccccc}-3\text{i}h&0&0&0&-2\text{i}h&-2\text{i}h&-2\text{i}h&0\\ 0&-\text{i}h&2h&2h&0&0&0&-2\text{i}h\\ 0&-2h&-\text{i}h&2h&0&0&0&2\text{i}h\\ 0&-2h&-2h&-\text{i}h&0&0&0&-2\text{i}h\\ -2\text{i}h&0&0&0&\text{i}h&2h&-2h&0\\ -2\text{i}h&0&0&0&-2h&\text{i}h&2h&0\\ -2\text{i}h&0&0&0&2h&-2h&\text{i}h&0\\ 0&-2\text{i}h&2\text{i}h&-2\text{i}h&0&0&0&3\text{i}h\\ \end{array}\right) (498)

which can be divided into two blocks. We get the eigenvalues by directly diagonalizing the matrix:

±5ih,±(23+1)ih,±3ih,±(231)ih.\displaystyle\pm 5\text{i}h,\quad\pm(2\sqrt{3}+1)\text{i}h,\quad\pm 3\text{i}h,\quad\pm(2\sqrt{3}-1)\text{i}h. (499)

Similarly for general NN, we can write effective Hamiltonian defined in (479) as

=ij=1k(αijcicj+βijcicj+γijcicj+θijcicj),\displaystyle{\cal H}=\sum_{i\leq j=1}^{k}\left(\alpha_{ij}c_{i}^{\dagger}c_{j}+\beta_{ij}c_{i}c_{j}^{\dagger}+\gamma_{ij}c_{i}^{\dagger}c_{j}^{\dagger}+\theta_{ij}c_{i}c_{j}\right), (500)

with

αii=ih,βii=ih,αij=2h,βij=2h,\displaystyle\alpha_{ii}=\text{i}h,\quad\beta_{ii}=-\text{i}h,\quad\alpha_{ij}=2h,\quad\beta_{ij}=2h, (501)
γij=2ih,θij=2ih,γii=0,θii=0.\displaystyle\gamma_{ij}=-2\text{i}h,\quad\theta_{ij}=2\text{i}h,\quad\gamma_{ii}=0,\quad\theta_{ii}=0. (502)

This Hamiltonian is quadratic and famously can be diagonalized by the Bogoliubov and Valatin’s method Bogolyubov:1947zz ; Valatin:1958ja . Explicitly we can do the transformation by taking an operator basis for the Hamiltonian

H=cMc\displaystyle H=c^{\dagger}Mc (503)

where we have

c=(c1,c2,,c1,c2,).\displaystyle c^{\dagger}=\left(c_{1}^{\dagger},c_{2}^{\dagger},\dots,c_{1},c_{2},\dots\right). (504)

In the simple case with N=4N=4 the matrix can be expressed as

M=(ihh0ihhihih00ihihhih0hih),\displaystyle M=\left(\begin{array}[]{cccc}\text{i}h&h&0&-\text{i}h\\ -h&\text{i}h&\text{i}h&0\\ 0&\text{i}h&-\text{i}h&h\\ -\text{i}h&0&-h&-\text{i}h\\ \end{array}\right), (509)

we can directly take the diagonalization and get the eigenvalues

i(1+2)h,i(1+2)h,i(12)h,i(1+2)h.\displaystyle\text{i}(1+\sqrt{2})h,\quad-\text{i}(1+\sqrt{2})h,\quad-\text{i}(1-\sqrt{2})h,\quad-\text{i}(-1+\sqrt{2})h. (510)

For simplicity we take the notation as

λ1=i(2+1)h,λ2=i(21)h,\displaystyle\lambda_{1}=\text{i}(\sqrt{2}+1)h,\quad\lambda_{2}=\text{i}(\sqrt{2}-1)h, (511)

then the resulting effective Hamiltonian becomes

H=λ1(d1d1d1d1)+λ2(d2d2d2d2).\displaystyle H=\lambda_{1}\left(d_{1}^{\dagger}d_{1}-d_{1}d_{1}^{\dagger}\right)+\lambda_{2}\left(d_{2}^{\dagger}d_{2}-d_{2}d_{2}^{\dagger}\right). (512)

To evaluate the trace we still take the states as (483) therefore we have

Tr(eH)=ea(λ1+λ2)+ea(λ1λ2)+ea(λ1+λ2)+ea(λ1+λ2),\displaystyle\text{Tr}(e^{H})=e^{-a(\lambda_{1}+\lambda_{2})}+e^{a(\lambda_{1}-\lambda_{2})}+e^{a(-\lambda_{1}+\lambda_{2})}+e^{a(\lambda_{1}+\lambda_{2})}, (513)

so we can recover the result (489). For general NN the operator (500) can be expressed as a block matrix

M=(A+ihINiAiAAihIN),\displaystyle M=\left(\begin{array}[]{cc}A+\text{i}hI_{N}&-\text{i}A\\ \text{i}A&A-\text{i}hI_{N}\end{array}\right), (516)

with

A=(0hhh0hhh0)\displaystyle A=\left(\begin{array}[]{cccc}0&h&h&\cdots\\ -h&0&h&\cdots\\ -h&-h&0&\cdots\\ \vdots&\vdots&\vdots&\ddots\end{array}\right) (521)

The characteristic equation is

det(A+(ihλ))(A(ih+λ)H2)\displaystyle\det\left(A+(\text{i}h-\lambda))(A-(\text{i}h+\lambda)-H^{2}\right) =\displaystyle= det((h2+λ2)IN2λA)\displaystyle\det\left((h^{2}+\lambda^{2})I_{N}-2\lambda A\right) (522)
=\displaystyle= (λ+h)N+(λh)N=0.\displaystyle(\lambda+h)^{N}+(\lambda-h)^{N}=0\,. (523)

So the eigenvalues are

λm=ihtan(mπ2N),m=1,3,,N1.\displaystyle\lambda_{m}=\text{i}h\tan(\frac{m\pi}{2N}),\quad m=1,3,\dots,N-1\,. (524)

then the Hamiltonian becomes

H=i=1Nλi(didididi).\displaystyle H=\sum_{i=1}^{N}\lambda_{i}(d_{i}^{\dagger}d_{i}-d_{i}d_{i}^{\dagger}). (525)

and the trace will have the form

Tr(eH)=σ=±1eai=1kσiλi=σi=1keaσiλi=i=1kσeaσiλi=2ki=1kcosh(aλi),\displaystyle{\rm Tr}(e^{H})=\sum_{\sigma=\pm 1}e^{a\sum_{i=1}^{k}\sigma_{i}\lambda_{i}}=\sum_{\sigma}\prod_{i=1}^{k}e^{a\sigma_{i}\lambda_{i}}=\prod_{i=1}^{k}\sum_{\sigma}e^{a\sigma_{i}\lambda_{i}}=2^{k}\prod_{i=1}^{k}\cosh\left(a\lambda_{i}\right)\,, (526)

Now let us consider the function

Xn=1i1<inNψi1ψiN.\displaystyle X_{n}=\sum_{1\leq i_{1}<\dots i_{n}\leq N}\psi_{i_{1}}\dots\psi_{i_{N}}. (527)

We would like to argue that in the large NN limit, we have the approximation

n!X2n(X2)n,\displaystyle n!X_{2n}\approx(X_{2})^{n}, (528)

as we find for the 0-dimensional theory. Note that unlike the situation of the 0-dimensional theory, {Xn}\{X_{n}\} do not form a basis for X2nX_{2}^{n}. For example, let us take N=6N=6, there is indeed the identity

X22=15+2!X4\displaystyle X_{2}^{2}=-15+2!X_{4} (529)

but we find that

X23\displaystyle X_{2}^{3} =\displaystyle= 3!X6+15X2+12(ψ1ψ2+ψ1ψ6+ψ3ψ4+ψ4ψ5+ψ5ψ6)\displaystyle 3!X_{6}+15X_{2}+12(\psi_{1}\psi_{2}+\psi_{1}\psi_{6}+\psi_{3}\psi_{4}+\psi_{4}\psi_{5}+\psi_{5}\psi_{6}) (530)
4(ψ1ψ4+ψ2ψ4+ψ3ψ6).\displaystyle\qquad-4(\psi_{1}\psi_{4}+\psi_{2}\psi_{4}+\psi_{3}\psi_{6}).

Let us focus on the second last term in X2nX_{2}^{n}

X2nc1X2n4+n!X2n,c1=(n2)!(n2)(N2),\displaystyle X_{2}^{n}\approx\dots c_{1}X_{2n-4}+n!X_{2n},\quad c_{1}=(n-2)!{n\choose 2}{N\choose 2}\,, (531)

where c1c_{1} is computed as follows. We need to pick 2 X2X_{2} out of nn and contract them, and the (n2)(n-2) X2X_{2}’s remain not contracted and gives (n1)!(n-1)! X2n4X_{2n-4}. Notice that the subleading term is X2n4X_{2n-4} instead of X2n2X_{2n-2}, since if we contract one fermion in X2X_{2} to get

ψ1ψ2ψ1ψ3ψ3ψ2,\displaystyle\psi_{1}\psi_{2}\psi_{1}\psi_{3}\mapsto\psi_{3}\psi_{2}\,, (532)

there is going to be another contraction that gives

ψ1ψ3ψ1ψ2ψ2ψ3.\displaystyle\psi_{1}\psi_{3}\psi_{1}\psi_{2}\mapsto\psi_{2}\psi_{3}\ . (533)

The two outcomes simply cancel with each other. The main conclusion of this computation is, given that X2nN2nX_{2n}\sim N^{2n}, the subleading terms can be safely neglected and approximate X2nX_{2n} by X2nX_{2}^{n}. So in the large NN limit, we can use the G,ΣG,\Sigma trick to compute the fermionic integral

Iq(a)\displaystyle I_{q}(a) =\displaystyle= dNψiexp(aAψA),A={1a1<<aqN},\displaystyle\int\text{d}^{N}\psi_{i}\exp(a\sum_{A}\psi_{A}),\quad A=\{1\leq a_{1}<\dots<a_{q}\leq N\}, (534)
\displaystyle\approx dNψieaGq2q2!eiσ(Gi<jψiψj)dGdσ\displaystyle\int\text{d}^{N}\psi_{i}\,e^{a\frac{G^{\frac{q}{2}}}{\frac{q}{2}!}}e^{\text{i}\sigma(G-\sum_{i<j}\psi_{i}\psi_{j})}\text{d}G\text{d}\sigma (535)
=\displaystyle= dGdσI2(iσ)eaGq2q2!eiσG=I2(iG)eaGq2q2!|G=0.\displaystyle\int\text{d}G\text{d}\sigma I_{2}(-\text{i}\sigma)e^{a\frac{G^{\frac{q}{2}}}{\frac{q}{2}!}}e^{\text{i}\sigma G}=I_{2}(\text{i}\partial_{G})e^{a\frac{G^{\frac{q}{2}}}{\frac{q}{2}!}}|_{G=0}\,. (536)

where the function I2I_{2} is defined in (478). We can evaluate this expression and we expect the half-wormhole contributions to be similar as the 0-SYK model

zz+Θ,Θ=dNψe0T𝑑t12iNψitψi+iq/20T𝑑tA(JAμ)ψA.\displaystyle z\approx\langle z\rangle+\Theta\,,\qquad\Theta=\int d^{N}\psi\,e^{-\int_{0}^{T}dt\,\frac{1}{2}\sum_{i}^{N}\psi_{i}\partial_{t}\psi_{i}+\text{i}^{q/2}\int_{0}^{T}dt\sum_{A}(J_{A}-\mu)\psi_{A}}\ . (537)

The detailed analysis is similar to the Brownian SYK model as we have shown above, but it is not particularly illuminating, so we omit them here.

In the next section, we instead consider a modified SYK-like model where half-wormhole saddle can be verified explicitly.

9.2 Random coupling from product of Grassmann variables JA(0)=JiθAiJ_{A}^{(0)}=J\prod_{i}\theta_{A_{i}}

A modified SYK-like model dubbed as partially disorder-averaged SYK model is proposed in Goto:2021wfs . In this model, the random coupling J~A\widetilde{J}_{A} consists of two pieces

J~A=JA+JA(0)\displaystyle\widetilde{J}_{A}=J_{A}+J_{A}^{(0)} (538)

where JAJ_{A} is the standard random coupling of the SYK model while JA(0)J_{A}^{(0)} is specially chosen as

Jiiiq(0)=iq/2q!μθi1θiq,with{θi,θj}=2δij\displaystyle J_{i_{i}\dots i_{q}}^{(0)}=\text{i}^{q/2}q!\mu\,\theta_{i_{1}}\dots\theta_{i_{q}},\quad\text{with}\quad\{\theta_{i},\theta_{j}\}={2}\delta_{ij} (539)

so we can think of it as coupling the fermions ψi\psi_{i} in the original model with some background Majorana fermions θi\theta_{i} (or non-dynamical fermions living in another universe Goto:2021wfs ). Note that JA(0)J_{A}^{(0)} is not a c-number which is different from our models studied in the previous section.

9.2.1 0d model

Let us first consider the 0-dimensional model to see the difference explicitly. In this case the integral (154) can be written as

z\displaystyle z =\displaystyle= dNψexp(iq/2J~i1iqψi1iq)\displaystyle\int\text{d}^{N}\psi\exp(\text{i}^{q/2}\sum\widetilde{J}_{i_{1}\dots i_{q}}\psi_{i_{1}\dots i_{q}})\, (540)
=\displaystyle= dNψexp(iq/2AJAψA+μ(iθiψi)q).\displaystyle\int\text{d}^{N}\psi\exp\left(\text{i}^{q/2}\sum_{A}J_{A}\psi_{A}+\mu(\sum_{i}\theta_{i}\psi_{i})^{q}\right).

The averaged quantity z\langle z\rangle

z=dNψexp(μ(iθiψi)q),\displaystyle\langle z\rangle=\int\text{d}^{N}\psi\exp\left(\mu(\sum_{i}\theta_{i}\psi_{i})^{q}\right)\,, (541)

can be computed in two ways. One can integrate out the fermions ψi\psi_{i} directly. The result is

z=μN/qN!(N/q)!dNψ(θ1ψ1)(θNψN)=[iθi]μN/qN!(N/q)![iθi]𝔪p.\displaystyle\langle z\rangle=\frac{\mu^{N/q}N!}{(N/q)!}\int\text{d}^{N}\psi(\theta_{1}\psi_{1})\dots(\theta_{N}\psi_{N})=\left[\prod_{i}\theta_{i}\right]\frac{\mu^{N/q}N!}{(N/q)!}\equiv\left[\prod_{i}\theta_{i}\right]\mathfrak{m}_{p}\,. (542)

Note that zz is not a c-number and depends on the background fermions living in other universe. Here we will not think of this as a problem but a feature since the model is not exactly the original SYK model. Alternatively we can compute this average quantity by the G,ΣG,\Sigma trick:

Gσ=iθiψi,\displaystyle G_{\sigma}=\sum_{i}\theta_{i}\psi_{i}\,,
z=dNψdGσdΣσ2πei[Σσ(Gσiθiψi)]eμGσq\displaystyle\langle z\rangle=\int\text{d}^{N}\psi\int\text{d}G_{\sigma}\frac{\text{d}\Sigma_{\sigma}}{2\pi}e^{\text{i}[\Sigma_{\sigma}(G_{\sigma}-\sum_{i}\theta_{i}\psi_{i})]}e^{\mu G_{\sigma}^{q}}
=[iθi]dGσdΣσ2πΣσNeiΣσGσ+μGσq\displaystyle\qquad=\left[\prod_{i}\theta_{i}\right]\int\text{d}G_{\sigma}\frac{\text{d}\Sigma_{\sigma}}{2\pi}\Sigma_{\sigma}^{N}e^{\text{i}\Sigma_{\sigma}G_{\sigma}+\mu G_{\sigma}^{q}} (543)
=[iθi](Gσ)NeμGσq|Gσ=0=[iθi]μN/qN!(N/q)!.\displaystyle\qquad=\left[\prod_{i}\theta_{i}\right](\partial_{G_{\sigma}})^{N}e^{\mu G_{\sigma}^{q}}|_{G_{\sigma}=0}=\left[\prod_{i}\theta_{i}\right]\frac{\mu^{N/q}N!}{(N/q)!}\,. (544)

One can also use the effective action (543) to derive the large NN result of (541) as shown in Goto:2021wfs . We will not repeat that analysis here. Instead, we would like to consider the half-wormhole saddle of zz

zz+Θ\displaystyle z\approx\langle z\rangle+\Theta (545)

as we did in last section. The subtlety is that as we stressed zz is not a c-number so the approximation (545) is in the sense

[z(z+Θ)]20,\displaystyle\langle[z-(\langle z\rangle+\Theta)]^{2}\rangle\approx 0\,, (546)

which is a c-number due to (539) is small. Let us proceed by computing the averaged quantity z2\langle z^{2}\rangle

z2\displaystyle\langle z^{2}\rangle =\displaystyle= d2Nψexp(τ2q!(iψiLψiR)q+μ(iθiψiL)q+μ(iθiψiR)q)\displaystyle\int\text{d}^{2N}\psi\,\exp\left({\frac{\tau^{2}}{q!}}(\sum_{i}\psi_{i}^{L}\psi_{i}^{R})^{q}+\mu(\sum_{i}\theta_{i}\psi_{i}^{L})^{q}+\mu(\sum_{i}\theta_{i}\psi_{i}^{R})^{q}\right)
=\displaystyle= d2Nψk(τ2q!)ki1<<ikq(ψi1LRψiqkLR)(qk)!k!((Nkq)!(N/qk)!)2\displaystyle\int\text{d}^{2N}\psi\sum_{k}\left(\frac{\tau^{2}}{q!}\right)^{k}\sum_{{i_{1}<\dots<i_{kq}}}(\psi^{LR}_{i_{1}}\dots\psi^{LR}_{i_{qk}})\frac{(qk)!}{k!}\left(\frac{(N-kq)!}{(N/q-k)!}\right)^{2}
μ2p2kj1<<jNqk{i}(θj1ψj1L)(θjNqkψjNkqL)(θj1ψj1R)(θjNkψjNkqR)\displaystyle\mu^{2p-2k}\sum_{j_{1}<\dots<j_{N-qk}\neq\{i\}}(\theta_{j_{1}}\psi_{j_{1}}^{L})\dots(\theta_{j_{N-qk}}\psi^{L}_{j_{N-kq}})(\theta_{j_{1}}\psi_{j_{1}}^{R})\dots(\theta_{j_{N-k}}\psi^{R}_{j_{N-kq}})
=\displaystyle= d2Nψk(τ2q!)kμ2p2k(qk)!k!((Nkq)!(N/qk)!)2(Nkq)ψ1LRψNLR\displaystyle\int\text{d}^{2N}\psi\sum_{k}\left(\frac{\tau^{2}}{q!}\right)^{k}\mu^{2p-2k}\frac{(qk)!}{k!}\left(\frac{(N-kq)!}{(N/q-k)!}\right)^{2}{N\choose kq}\psi^{LR}_{1}\dots\psi_{N}^{LR}
=\displaystyle= k(τ2q!)kμ2p2k(qk)!k!((Nkq)!(N/qk)!)2(Nkq)\displaystyle\sum_{k}\left(\frac{\tau^{2}}{q!}\right)^{k}\mu^{2p-2k}\frac{(qk)!}{k!}\left(\frac{(N-kq)!}{(N/q-k)!}\right)^{2}{N\choose kq}
=\displaystyle= kpτ2kμ2(pk)ck𝔪pk2k𝔷2(k),\displaystyle\sum_{k}^{p}\tau^{2k}\mu^{2(p-k)}c_{k}\mathfrak{m}_{p-k}^{2}\equiv\sum_{k}\mathfrak{z}_{2}^{(k)}\,, (548)

where ckc_{k} of defined in (231) and 𝔪p\mathfrak{m}_{p} is defined in (542). The result (9.2.1) is in the same form of (230). So the analysis of the half-wormhole saddle will be similar; we insert the a suitable identity to (540)

z\displaystyle z =\displaystyle= dNψexp(iq/2J~i1iqψi1iq)dGσδ(Gσiθiψi)exp(μq!(Gσq(iθiψi)q))\displaystyle\int\text{d}^{N}\psi\exp(\text{i}^{q/2}\sum\widetilde{J}_{i_{1}\dots i_{q}}\psi_{i_{1}\dots i_{q}})\int\text{d}G_{\sigma}\delta(G_{\sigma}-\sum_{i}\theta_{i}\psi_{i})\exp(\frac{\mu}{q!}(G_{\sigma}^{q}-(\sum_{i}\theta_{i}\psi_{i})^{q})) (549)
=\displaystyle= dNψdΣσdGσ2πiexp(iq/2AJAψA+Σσiθiψi)exp(ΣσGσ+μq!Gσq).\displaystyle\int\text{d}^{N}\psi\frac{\text{d}\Sigma_{\sigma}\text{d}G_{\sigma}}{2\pi\text{i}}\exp(\text{i}^{q/2}\sum_{A}J_{A}\psi_{A}+\Sigma_{\sigma}\sum_{i}\theta_{i}\psi_{i})\exp(-\Sigma_{\sigma}G_{\sigma}+\frac{\mu}{q!}G_{\sigma}^{q})\,.

Following the arguments below (4.1) one can obtain the half-wormhole saddle161616Here the factor iθi\prod_{i}\theta_{i} should be present.

Θ=[iθi]dNψexp(iq/2AJAψA).\displaystyle\Theta=\left[\prod_{i}\theta_{i}\right]\int\text{d}^{N}\psi\exp(\text{i}^{q/2}\sum_{A}J_{A}\psi_{A})\,. (550)

Then it is easy to find that the half-wormhole saddle satisfies

Θ=0,Θ2=Θz=𝔷2(p),\displaystyle\langle\Theta\rangle=0,\quad\langle\Theta^{2}\rangle=\langle\Theta z\rangle=\mathfrak{z}_{2}^{(p)}\,, (551)

so the approximation (545) will be sufficient if 𝔷2(p)\mathfrak{z}_{2}^{(p)} is the dominant term in (548) as we have shown in last section. When 𝔷2(p)\mathfrak{z}_{2}^{(p)} is not the dominant term we have to consider the contribution of fluctuation of Σσ\Sigma_{\sigma}. To finish our analysis of the half-wormhole saddle for zz, let us redo the computation of z2\langle z^{2}\rangle with the G,ΣG,\Sigma trick. We need introduce three GG variables

GLR=1NiψiLψiR,GL=1NiθiψiL,GR=1NiθiψiR\displaystyle G_{LR}=\frac{1}{N}\sum_{i}\psi_{i}^{L}\psi_{i}^{R},\quad G_{L}=\frac{1}{{N}}\sum_{i}\theta_{i}\psi_{i}^{L},\quad G_{R}=\frac{1}{{N}}\sum_{i}\theta_{i}\psi_{i}^{R} (552)

then z2\langle z^{2}\rangle can be written as

z2\displaystyle\langle z^{2}\rangle =\displaystyle= d2NψadGaexp(Nq(tGLRq+uGLq+uGRq))adΣa2πi/Nexp(ΣLR(NGLRiψiLψiR))\displaystyle\int\text{d}^{2N}\psi\prod_{a}\text{d}G_{a}\exp\left(\frac{N}{q}(tG_{LR}^{q}+uG_{L}^{q}+uG_{R}^{q})\right)\int\prod_{a}\frac{d\Sigma_{a}}{2\pi\text{i}/N}\exp(-\Sigma_{LR}(NG_{LR}-\sum_{i}\psi_{i}^{L}\psi_{i}^{R})) (553)
exp(ΣL(NGLiθiψiL))exp(ΣR(NGRiθiψiR))\displaystyle\exp(-\Sigma_{L}(NG_{L}-\sum_{i}\theta_{i}\psi_{i}^{L}))\exp(-\Sigma_{R}(NG_{R}-\sum_{i}\theta_{i}\psi_{i}^{R}))
=\displaystyle= [adGadΣa2πi/N]exp(N(tqGLRq+uqGLq+uqGRqaΣaGa))(ΣLR+ΣLΣR)N,\displaystyle\int[\prod_{a}\text{d}G_{a}\frac{d\Sigma_{a}}{2\pi\text{i}/N}]\exp\left(N(\frac{t}{q}G_{LR}^{q}+\frac{u}{q}G_{L}^{q}+\frac{u}{q}G_{R}^{q}-\sum_{a}\Sigma_{a}G_{a})\right)(\Sigma_{LR}+\Sigma_{L}\Sigma_{R})^{N}\,,

where in order to have a well-defined large NN scaling we have introduced

t=τ2(q1)!Nq1,u=qμNq1.\displaystyle t=\frac{\tau^{2}}{(q-1)!}N^{q-1},\quad u=q\mu N^{q-1}. (554)

The saddle point equations are

tGLRq1=ΣLR,uGLq1=ΣL,uGRq1=ΣR,\displaystyle tG_{LR}^{q-1}=\Sigma_{LR},\quad uG_{L}^{q-1}=\Sigma_{L},\quad uG_{R}^{q-1}=\Sigma_{R}\,, (555)
GLR=1ΣLR+ΣLΣR,GL=ΣRΣLR+ΣLΣR,GR=ΣLΣLR+ΣLΣR.\displaystyle G_{LR}=\frac{1}{\Sigma_{LR}+\Sigma_{L}\Sigma_{R}},\quad G_{L}=-\frac{\Sigma_{R}}{\Sigma_{LR}+\Sigma_{L}\Sigma_{R}},\quad G_{R}=-\frac{\Sigma_{L}}{\Sigma_{LR}+\Sigma_{L}\Sigma_{R}}\,. (556)

The obvious solutions are the “wormhole” saddles with

GL=GR=ΣL=ΣR=0,\displaystyle G_{L}=G_{R}=\Sigma_{L}=\Sigma_{R}=0, (557)

which corresponds to 𝔷2p\mathfrak{z}_{2}^{p}. There are also other saddles corresponding to other 𝔷2k\mathfrak{z}_{2}^{k}. For the simplest case q=2q=2, these solutions can be written explicitly. The “wormhole” saddles are

GL=GR=ΣL=ΣR=0,GLR=±1t,ΣLR=±t,\displaystyle G_{L}=G_{R}=\Sigma_{L}=\Sigma_{R}=0,G_{LR}=\pm\frac{1}{\sqrt{t}},\quad\Sigma_{LR}=\pm\sqrt{t}, (558)
z2WH=eN2tN/2,\displaystyle\langle z^{2}\rangle_{\text{WH}}=e^{-\frac{N}{2}}t^{N/2}\,, (559)

which do not depend on μ\mu and the other four solutions are

ΣL=ΣR=uGL=uGR=±u2tu,ΣLR=tu,GLR=1u\displaystyle\Sigma_{L}=\Sigma_{R}=uG_{L}=uG_{R}=\pm\sqrt{\frac{u^{2}-t}{u}},\quad\Sigma_{LR}=\frac{t}{u},\quad G_{LR}=\frac{1}{u} (560)
ΣL=uGL=ΣR=uGR=±u2tu,ΣLR=tu,GLR=1u,\displaystyle\Sigma_{L}=uG_{L}=-\Sigma_{R}=-uG_{R}=\pm\sqrt{\frac{u^{2}-t}{u}},\quad\Sigma_{LR}=-\frac{{t}}{u},\quad G_{LR}=-\frac{1}{u}\,, (561)
z2new=eN2(2tu2)uN\displaystyle\langle z^{2}\rangle_{\text{new}}=e^{-\frac{N}{2}(2-\frac{t}{u^{2}})}u^{N} (562)

Apparently when uu\rightarrow\infty, ΣLR,GLR0,\Sigma_{LR},G_{LR}\rightarrow 0, then we expect that in this limit the dominant saddle will correspond to 𝔷20\mathfrak{z}_{2}^{0} since in this limit saddle point value does not depend on tt. Comparing these two saddle values we find

z2WHz2new=exp(N2(1x+logx))1,x=tu2.\displaystyle\frac{\langle z^{2}\rangle_{\text{WH}}}{\langle z^{2}\rangle_{\text{new}}}=\exp\left(\frac{N}{2}(1-x+\log x)\right)\leq 1,\quad x=\frac{t}{u^{2}}\,. (563)

Note that when x=1x=1 such that z2WH=z2new\langle z^{2}\rangle_{\text{WH}}=\langle z^{2}\rangle_{\text{new}} the new saddle just reduces to the wormhole saddle. Therefore it implies that the new saddle always dominates.

This new saddle is named as “unlinked half-wormhole” in Goto:2021wfs to distinguish it from the half-wormhole saddle which was found in Saad:2021rcu . One interpretation of this new saddle is that it is the analogue of the disconnected saddle in this model; indeed, we do not find other disconnected saddle with GLR=0G_{LR}=0, ΣLR=0\Sigma_{LR}=0 and GL/R,ΣL/R0G_{L/R},\Sigma_{L/R}\neq 0, in addition, this saddle is present only when u0u\neq 0, and this saddle is more and more important as uu increases.

The analysis of the half-wormhole saddle for z2z^{2} will be similar to one we did in last section so we will not repeat here.

9.2.2 1d model

Now we come back to the 0+1d model that is a variant of the Brownian SYK model. Let us begin by deriving the wormhole saddle of z2\langle z^{2}\rangle171717Here we have assumed the large NN limit, the exact treatment can be found in Saad:2021rcu

zLzR=d2Nψexp{0Tdt12i(ψiLtψiL+ψiRtψiR)+iq/20TdtAJ~A(ψAL+ψAR)}\displaystyle z_{L}z_{R}=\int\text{d}^{2N}\psi\exp\left\{-\int_{0}^{T}\text{d}t\frac{1}{2}\sum_{i}(\psi_{i}^{L}\partial_{t}\psi_{i}^{L}+\psi_{i}^{R}\partial_{t}\psi_{i}^{R})+\text{i}^{q/2}\int_{0}^{T}\text{d}t\sum_{A}\tilde{J}_{A}(\psi_{A}^{L}+\psi_{A}^{R})\right\}
zLzR=d2Nψexp{0Tdt12i(ψiLtψiL+ψiRtψiR)+\displaystyle\langle z_{L}z_{R}\rangle=\int\text{d}^{2N}\psi\exp\left\{-\int_{0}^{T}\text{d}t\frac{1}{2}\sum_{i}(\psi_{i}^{L}\partial_{t}\psi_{i}^{L}+\psi_{i}^{R}\partial_{t}\psi_{i}^{R})+\right.
0Tdt(τ2q!(iψiLψiR)q+μ(iθiψiL)q+μ(iθiψiR)q)+τ2E0T}.\displaystyle\qquad\qquad\qquad\left.\int_{0}^{T}\text{d}t\left({\frac{\tau^{2}}{q!}}(\sum_{i}\psi_{i}^{L}\psi_{i}^{R})^{q}+\mu(\sum_{i}\theta_{i}\psi_{i}^{L})^{q}+\mu(\sum_{i}\theta_{i}\psi_{i}^{R})^{q}\right)+\tau^{2}E_{0}T\right\}\,. (564)
=d2Nψ[adGadΣa2πi]exp{0Tdt12i(ψiLtψiL+ψiRtψiR)+τ2TE0+\displaystyle\qquad\qquad=\int\text{d}^{2N}\psi[\prod_{a}\text{d}G_{a}\frac{d\Sigma_{a}}{2\pi\text{i}}]\exp\left\{-\int_{0}^{T}\text{d}t\frac{1}{2}\sum_{i}(\psi_{i}^{L}\partial_{t}\psi_{i}^{L}+\psi_{i}^{R}\partial_{t}\psi_{i}^{R})+\tau^{2}TE_{0}+\right.
0Tdt(τ2q!GLRq+μGLq+μGRqaΣaGa+i[ΣLRψiLψiR+ΣLθiψiL+ΣRθiψiR])},\displaystyle\qquad\left.\int_{0}^{T}\text{d}t\left(\frac{\tau^{2}}{q!}G_{LR}^{q}+\mu G_{L}^{q}+\mu G_{R}^{q}-\sum_{a}\Sigma_{a}G_{a}+\sum_{i}[\Sigma_{LR}\psi_{i}^{L}\psi_{i}^{R}+\Sigma_{L}\theta_{i}\psi_{i}^{L}+\Sigma_{R}\theta_{i}\psi_{i}^{R}]\right)\right\}\,,

where E0=(Nq)E_{0}={N\choose q} is the constant term coming from ψAL(R)ψAL(R)=(1)q2\psi_{A}^{L(R)}\psi_{A}^{L(R)}=(-1)^{\frac{q}{2}}. As explained in Saad:2021rcu , we can focus on the time-independent saddles then the fermions can be simply integrated out. The result is 181818This is result is different from the one derived in Goto:2021wfs . It seems that they used a wrong formula for the fermion integral.

zLzR\displaystyle\langle z_{L}z_{R}\rangle =\displaystyle= eTτ2E0[adGadΣa2πi]eTN(tqGLRq+uqGLq+uqGRqaΣaGa)[cosh(TΣL2+ΣR2ΣLR2)]N.\displaystyle e^{T\tau^{2}E_{0}}\int[\prod_{a}\text{d}G_{a}\frac{d\Sigma_{a}}{2\pi\text{i}}]e^{TN(\frac{t}{q}G_{LR}^{q}+\frac{u}{q}G_{L}^{q}+\frac{u}{q}G_{R}^{q}-\sum_{a}\Sigma_{a}G_{a})}[\cosh(T\sqrt{\Sigma_{L}^{2}+\Sigma_{R}^{2}-\Sigma_{LR}^{2}})]^{N}. (566)
=\displaystyle= [adGadΣa2πi]eτ2TE0eSeff\displaystyle\int[\prod_{a}\text{d}G_{a}\frac{d\Sigma_{a}}{2\pi\text{i}}]e^{\tau^{2}TE_{0}}e^{S_{eff}}

For general TT, the saddle equation is very hard to solve due to the complicity of cosh\cosh function. However the equations simplify in the large TT limit because of the following approximations

log(cosh(TΣL2+ΣR2ΣLR2))±iTΣLR2ΣL2ΣR2.\displaystyle\log(\cosh(T\sqrt{\Sigma_{L}^{2}+\Sigma_{R}^{2}-\Sigma_{LR}^{2}}))\approx\pm\text{i}T\sqrt{\Sigma_{LR}^{2}-\Sigma_{L}^{2}-\Sigma_{R}^{2}}\,. (567)

Then in this limit the effective action becomes

Seff=TN(tqGLRq+uqGLq+uqGRqaΣaGa)±iNTΣLR2ΣL2ΣR2,\displaystyle S_{eff}=TN(\frac{t}{q}G_{LR}^{q}+\frac{u}{q}G_{L}^{q}+\frac{u}{q}G_{R}^{q}-\sum_{a}\Sigma_{a}G_{a})\pm\text{i}NT\sqrt{\Sigma_{LR}^{2}-\Sigma_{L}^{2}-\Sigma_{R}^{2}}\,, (568)

and corresponding saddle point equations are

tGLRq1=ΣLR,uGLq1=ΣL,uGRq1=ΣR,\displaystyle tG_{LR}^{q-1}=\Sigma_{LR},\quad uG_{L}^{q-1}=\Sigma_{L},\quad uG_{R}^{q-1}=\Sigma_{R}, (569)
GLR=±iΣLRΣLR2ΣL2ΣR2,\displaystyle G_{LR}=\pm\frac{\text{i}\Sigma_{LR}}{\sqrt{\Sigma_{LR}^{2}-\Sigma_{L}^{2}-\Sigma_{R}^{2}}}\,, (570)
GL=iΣLΣLR2ΣL2ΣR2,GR=iΣRΣLR2ΣL2ΣR2.\displaystyle G_{L}=\mp\frac{\text{i}\Sigma_{L}}{\sqrt{\Sigma_{LR}^{2}-\Sigma_{L}^{2}-\Sigma_{R}^{2}}},\quad G_{R}=\mp\frac{\text{i}\Sigma_{R}}{\sqrt{\Sigma_{LR}^{2}-\Sigma_{L}^{2}-\Sigma_{R}^{2}}}\,. (571)

So the wormhole saddle still presents Saad:2018bqo

GL=GR=ΣL=ΣR=0,GLR=±i,\displaystyle G_{L}=G_{R}=\Sigma_{L}=\Sigma_{R}=0,\quad G_{LR}=\pm\text{i}, (572)
eSeff|WH=eiqTNtq.\displaystyle e^{S_{eff}}\Big{|}_{\text{WH}}=e^{\text{i}^{q}TN\frac{t}{q}}. (573)

The unlinked half-wormhole saddle is:

GLR=ΣLR=0,GL=sinα,GR=cosα,\displaystyle G_{LR}=\Sigma_{LR}=0,\quad G_{L}=\sin\alpha,\quad G_{R}=\cos\alpha, (574)
eSeff|unlink=eTNuq(cosqα+sinqα)eSeff|unlink,α=0,π/2=eTNuq\displaystyle e^{S_{eff}}\Big{|}_{\text{unlink}}=e^{TN\frac{u}{q}(\cos^{q}\alpha+\sin^{q}\alpha)}\leq e^{S_{eff}}\Big{|}_{\text{unlink},\alpha=0,\pi/2}=e^{TN\frac{u}{q}}\, (575)

where the relation

GL2+GR2GLR2=1,\displaystyle G_{L}^{2}+G_{R}^{2}-G_{LR}^{2}=1\,, (576)

is fulfilled and α\alpha satisfies

cosα=±cosq1αcos2q2α+sin2q2α.\displaystyle\cos\alpha=\pm\frac{\cos^{q-1}\alpha}{\sqrt{\cos^{2q-2}\alpha+\sin^{2q-2}\alpha}}. (577)

In the late time (T)(T\rightarrow\infty), there is indeed a wormhole saddle so it possible to include a linked half-wormhole saddle for zz. We also assume that the half-wormhole saddle is time independent since the wormhole saddle is time independent. Then the analysis is completely same as the one for the 0-dimensional model. So the half-wormhole saddle will be given by

Θ=[iθi]dNψexp(Tiq/2AJAψA),\displaystyle\Theta=\left[\prod_{i}\theta_{i}\right]\int\text{d}^{N}\psi\exp(T\text{i}^{q/2}\sum_{A}J_{A}\psi_{A})\,, (578)
Θ2ΘzzLzR|Wormhole saddle.\displaystyle\langle\Theta^{2}\rangle\approx\langle\Theta z\rangle\approx\langle z_{L}z_{R}\rangle|_{\text{Wormhole saddle}}\,. (579)

10 Discussion

In this paper we consider the half-wormhole proposal in some statistical models and simple SYK-like models. We showed that in all statistical models we have consider the half-wormhole conjecture (63) is valid almost for all the distributions except for some special cases where the mean value of the random variable vanish. In the 0-dimensional SYK model which is introduced in Saad:2021rcu we have shown that the half-wormhole construction depends on the distribution of the couplings. When the mean value of the coupling is very large then only the disconnected saddles dominate therefore the correlation functions automatically factorize. If the mean value is very small such that only the wormhole saddles dominate then factorization can be restored by adding half-wormhole saddles. When the disconnected saddles and wormhole saddles are comparable, we have to modified the half-wormhole saddle to restore the factorization. We also generalized G,ΣG,\Sigma trick to compute z\langle z\rangle. As a by-product, we can construct a a new saddle, the single half-wormhole saddle, for zz. Moreover we argued if the random couplings satisfy a general distribution, new half-wormhole saddles can be constructed 191919Interestingly, irrelevant deformation of 0-SYK model is studied in Das:2022uhj where they show after deformation half-wormhole saddle survives. It is very possible that our new half-wormholes will also survive under the same irrelevant deformation.. We also generalize the construction of half-wormhole saddles to (modified) Brownian SYK model.

Partially averaged models and spacetime branes

The meaning of higher cumulants of the random coupling can be understood from the idea of Coleman’s Coleman:1988cy and Giddings’s and Strominger’s Giddings:1988wv ; Giddings:1988cx . Just as we showed they are related to the non-local interaction induced by the spacetime wormholes. However the first cumulant or the mean value seems to be puzzling. In Blommaert:2021gha ; Goto:2021wfs , ensemble theories with non-vanishing mean value random couplings are also consider where they call such models partially averaged models. In these models, the mean values of random couplings can be understood as external sources or spacetime branes which describe the non-perturbative corrections. It is shown in Blommaert:2021fob by fine tuning these non-perturbative corrections the JT gravity can factorize for all orders. But it seems that the original half-wormhole Saad:2021rcu constructed in 0-SYK model is not related to the branes but just a result of applying G,ΣG,\Sigma trick in the non-averaged theory. However we can understand this construction in an opposite way: the half-wormhole is constructed by adding eigenbranes Blommaert:2019wfy ; Goel:2020yxl in the averaged theory. This opposite point of view can also be viewed as an explicit realization of the idea Saad:2021uzi about factorization.

Standard SYK model

There are already proposals Saad:2021uzi ; Mukhametzhanov:2021hdi ; Mukhametzhanov:2021nea ; Goto:2021wfs of the half-wormhole saddle for z2z^{2} in the standard SYK model. But due to technical difficulty it has not been confirmed. It would be interesting to generalize our single half-wormhole saddle to the SYK model since zz is much simpler than z2z^{2}. It would be also interesting to generalize the new half-wormholes we found in section 3 to the standard SYK model and understand their possible relations to saddles in JT gravity.

Non-trivial saddles and Null states

In the simple 0-dimensional SYK model and Brownian SYK model, we find some non-trivial saddles whose on-shell values are comparable with one of the trivial self-averaging saddles. It would be very interesting if such non-trivial saddle also exists in the standard SYK model. It implies that there are also some non-trivial solutions in the dual (deformed) JT gravity. In the semiclassical physics, this coexistence of bulk description can be understood as the consequence of null states. In Blommaert:2022ucs , the null states of (deformed) JT gravity are proposed. However these null states do not show up in the dual matrix model. It seems to be promising to identify these null states in the SYK model.

Acknowledgements.
We thank many of the members of KITS for interesting related discussions. We also want to thank Kenta Suzuki for comments on a draft of this paper. CP is supported by the Fundamental Research Funds for the Central Universities, by funds from the University of Chinese Academy of Science (UCAS), and funds from the Kavli Institute for Theoretical Science (KITS). JT is supported by the National Youth Fund No.12105289 and funds from the UCAS program of special research associate.

Appendix A Explicit examples: simple observables

Thanks to the central limit theorem (CLT), the simplest choice of Y(X)Y(X) is just the summation of NN independent and identical random samples. We will first check proposal with three explicit distributions: the Gaussian distribution, the exponential distribution and the Poisson distribution and then give a general proof for general cases. Readers who are bored with these examples can jump into the general proof directly.

Let YY to be a summation of NN independent and identical random variables, i.e.

Y=i=1NXi,Y2=i,jNXiXj.\displaystyle Y=\sum_{i=1}^{N}X_{i},\quad Y^{2}=\sum_{i,j}^{N}X_{i}X_{j}. (580)

\bullet Gaussian distribution

The PDF of Gaussian distribution 𝒩(μ,t2)\mathcal{N}(\mu,t^{2}) is

P(x)=1t2πe12(xμt)2.\displaystyle P(x)=\frac{1}{t\sqrt{2\pi}}e^{-\frac{1}{2}(\frac{x-\mu}{t})^{2}}. (581)

Given (581) one can straightforwardly compute the averaged quantities

Xi=μ,eikiXi=eki2t22+ikiμ,Y2=Nt2+N2μ2,\displaystyle\langle X_{i}\rangle=\mu,\quad\langle e^{\text{i}k_{i}X_{i}}\rangle=e^{-\frac{k_{i}^{2}t^{2}}{2}+\text{i}k_{i}\mu},\quad\langle Y^{2}\rangle=Nt^{2}+N^{2}\mu^{2}, (582)
YeiikiXieiikiXi=ixieikixieikixiiki[1],ki[1]=(u+ikit2).\displaystyle\frac{\langle Ye^{\sum_{i}\text{i}k_{i}X_{i}}\rangle}{\langle e^{\text{i}\sum_{i}k_{i}X_{i}}\rangle}=\sum_{i}\frac{\langle x_{i}e^{\text{i}k_{i}x_{i}}\rangle}{\langle e^{\text{i}k_{i}x_{i}}\rangle}\equiv\sum_{i}k_{i}[1],\quad k_{i}[1]=(u+\text{i}k_{i}t^{2}). (583)

where we have defined

ki[n]=xineikixieikixi.\displaystyle k_{i}[n]=\frac{\langle x^{n}_{i}e^{\text{i}k_{i}x_{i}}\rangle}{\langle e^{\text{i}k_{i}x_{i}}\rangle}. (584)

Let us introduce another convenient quantity

ki[n]m^=12π𝑑kieikiXiP(Xi)eikixiki[n]m,\displaystyle\widehat{k_{i}[n]^{m}}=\frac{1}{2\pi}\int dk_{i}\frac{e^{-\text{i}k_{i}X_{i}}}{P(X_{i})}\langle e^{\text{i}k_{i}x_{i}}\rangle k_{i}[n]^{m}, (585)
ki[1]^=Xi,ki[1]2^=Xi2t2,\displaystyle\widehat{k_{i}[1]}=X_{i},\quad\widehat{k_{i}[1]^{2}}=X_{i}^{2}-t^{2},\quad\dots (586)

then the half-wormhole can be written as

Φ\displaystyle\Phi =\displaystyle= iki[1]2^+ijki[1]^kj[1]^=i(Xi2t2)+ijXiXj,\displaystyle\sum_{i}\widehat{k_{i}[1]^{2}}+\sum_{i\neq j}\widehat{k_{i}[1]}\widehat{k_{j}[1]}=\sum_{i}(X_{i}^{2}-t^{2})+\sum_{i\neq j}X_{i}X_{j}, (587)
=\displaystyle= Y2Nt2\displaystyle Y^{2}-Nt^{2} (588)

Substituting into (20) one can computed the error and the ration ρ\rho directly

Error =\displaystyle= Y2(Nt2+N2μ2N2μ2+Y2Nt2)\displaystyle Y^{2}-\left(Nt^{2}+N^{2}\mu^{2}-N^{2}\mu^{2}+Y^{2}-Nt^{2}\right) (589)
=\displaystyle= 0.\displaystyle 0. (590)

The proposal is exact as expected.

\bullet Exponential distribution

The PDF of exponential distribution is given by

Pλ(x)={λeλx,x0,0,x<0,\displaystyle P_{\lambda}(x)=\begin{cases}\lambda e^{-\lambda x},\quad&x\geq 0,\\ 0,&x<0,\end{cases} (591)

and the moments are given by

xn=n!λn.\displaystyle\langle x^{n}\rangle=\frac{n!}{\lambda^{n}}. (592)

The relevant averaged quantities are

Xi=1λ,eikXi=λλik,Y2=N(N+1)λ2.\displaystyle\langle X_{i}\rangle=\frac{1}{\lambda},\quad\langle e^{\text{i}kX_{i}}\rangle=\frac{\lambda}{\lambda-\text{i}k},\quad\langle Y^{2}\rangle=\frac{N(N+1)}{\lambda^{2}}. (593)

From the example of Gaussian distribution we have shown that to compute Φ\Phi (588) and the error (589) we only need to compute

ki[1]^=Xi,ki[1]2^=Xi22,\displaystyle\widehat{k_{i}[1]}=X_{i},\quad\widehat{k_{i}[1]^{2}}=\frac{X_{i}^{2}}{2}, (594)

which lead to

Error=i(Xi221λ2),ρError2Y221N3.\displaystyle\text{Error}=\sum_{i}\left(\frac{X_{i}^{2}}{2}-\frac{1}{\lambda^{2}}\right),\quad\rho\approx\frac{\langle\text{Error}^{2}\rangle}{\langle Y^{2}\rangle^{2}}\sim\frac{1}{N^{3}}. (595)

So the proposal is correct.

\bullet Poisson distribution

Next let us examine the proposal for a discrete probability distribution: the Poisson distribution. The PDF is

Pλ(k)=eλλkk!,k=0,1,2,\displaystyle P_{\lambda}(k)=\frac{e^{-\lambda}\lambda^{k}}{k!},\quad k=0,1,2,\dots (596)

and the moments are given by

xn=Bn(λ),\displaystyle\langle x^{n}\rangle=B_{n}(\lambda), (597)

where Bn(λ)B_{n}(\lambda) is the Bell polynomial. The relevant averaged quantities can be easily computed

Xi=λ,eikXi=eλ+eikλ,Y2=Nλ+N2λ2,\displaystyle\langle X_{i}\rangle={\lambda},\quad\langle e^{\text{i}kX_{i}}\rangle=e^{-\lambda+e^{\text{i}k}\lambda},\quad\langle Y^{2}\rangle=N\lambda+N^{2}\lambda^{2}, (598)
ki[1]=eikiλ.\displaystyle k_{i}[1]=e^{\text{i}k_{i}}\lambda. (599)

The computation of ki[1]^\widehat{k_{i}[1]} and ki[1]2^\widehat{k_{i}[1]^{2}} is a little subtle and needs some explanation. According to the definition (585), we have

ki[1]^\displaystyle\widehat{k_{i}[1]} =\displaystyle= λ2πP(Xi)𝑑kieikiXieikixieiki,\displaystyle\frac{\lambda}{2\pi P(X_{i})}\int dk_{i}{e^{-\text{i}k_{i}X_{i}}}{}\langle e^{\text{i}k_{i}x_{i}}\rangle e^{\text{i}k_{i}}, (600)
=\displaystyle= λP(Xi)P(Xi1)=Xi,\displaystyle\frac{\lambda}{P(X_{i})}P(X_{i}-1)=X_{i}, (601)

where in the second line we have used the fact that the inverse Fourier transformation of the characteristic function is the PDF

12π𝑑kieikiXieikixi=P(Xi).\displaystyle\frac{1}{2\pi}\int dk_{i}{e^{-\text{i}k_{i}X_{i}}}{}\langle e^{\text{i}k_{i}x_{i}}\rangle=P(X_{i}). (602)

Similarly one can derive

ki[1]2^\displaystyle\widehat{k_{i}[1]^{2}} =\displaystyle= λ2P(Xi2)P(Xi)=Xi(Xi1).\displaystyle\lambda^{2}\frac{P(X_{i}-2)}{P(X_{i})}=X_{i}(X_{i}-1). (603)

Then using (589) we can obtain the error

Error=i(Xiλ),ρError2Y221N3,\displaystyle\text{Error}=\sum_{i}(X_{i}-\lambda),\quad\rho\approx\frac{\langle\text{Error}^{2}\rangle}{\langle Y^{2}\rangle^{2}}\sim\frac{1}{N^{3}}, (604)

so the proposal is correct while the effective parameter is λN\lambda N instead of NN.

Appendix B Explicite Examples: composite observables

Gaussian distribution

We still start from the simple model

Y=i=1NXi2,Xi2=t2,Y2=i,jXi2Xj2.\displaystyle Y=\sum_{i=1}^{N}X_{i}^{2},\quad\langle X_{i}^{2}\rangle=t^{2},\quad Y^{2}=\sum_{i,j}X_{i}^{2}X_{j}^{2}. (605)

It is easy to evaluate

eikiXi=exp(iki2t22),\displaystyle\langle e^{\text{i}k_{i}X_{i}}\rangle=\exp(-\sum_{i}\frac{k_{i}^{2}t^{2}}{2}), (606)
YeiikiXi/eiikiXi=i(ki2t4+t2).\displaystyle\langle Ye^{\text{i}\sum_{i}k_{i}X_{i}}\rangle/\langle e^{\text{i}\sum_{i}k_{i}X_{i}}\rangle=\sum_{i}(-k_{i}^{2}t^{4}+t^{2}). (607)

Then we find

Φ=i(Xi44t2Xi2+2t4)+ijXi2Xj2,\displaystyle\Phi=\sum_{i}\left(X_{i}^{4}-4t^{2}X_{i}^{2}+2t^{4}\right)+\sum_{i\neq j}X_{i}^{2}X_{j}^{2}, (608)

and the error is given by

Error=4Nt44t2iXi2.\displaystyle\text{Error}=4Nt^{4}-4t^{2}\sum_{i}X_{i}^{2}. (609)

From this we can easily find that the proposal is correct.

Let us consider

Y=iXi3,Xi3=0,\displaystyle Y=\sum_{i}X_{i}^{3},\quad\langle X_{i}^{3}\rangle=0, (610)
Y2=i,jXi3Xj3,Y2=15Nt6.\displaystyle Y^{2}=\sum_{i,j}X_{i}^{3}X_{j}^{3},\quad\langle Y^{2}\rangle=15Nt^{6}. (611)

Then we find can evaluate

YeikiXi/eikiXi=iikit4(3ki2t2),\displaystyle\langle Ye^{\text{i}k_{i}X_{i}}\rangle/\langle e^{\text{i}k_{i}X_{i}}\rangle=\sum_{i}\text{i}k_{i}t^{4}(3-k_{i}^{2}t^{2}), (612)

and

Φ=i(Xi69t2Xi4+18t4Xi26t6)+ijXi3Xj3.\displaystyle\Phi=\sum_{i}(X_{i}^{6}-9t^{2}X_{i}^{4}+18t^{4}X_{i}^{2}-6t^{6})+\sum_{i\neq j}X_{i}^{3}X_{j}^{3}. (613)

So that the error is given by

Error=i(9t2Xi418t4Xi29t6),\displaystyle\text{Error}=\sum_{i}(9t^{2}X_{i}^{4}-18t^{4}X_{i}^{2}-9t^{6}), (614)

the leading order of Error2\langle\text{Error}^{2}\rangle is

0N2.\displaystyle 0N^{2}. (615)

Let us consider

Y=iXi4,Xi4=3t4,Y2=i,jXi4Xj4\displaystyle Y=\sum_{i}X_{i}^{4},\quad\langle X_{i}^{4}\rangle=3t^{4},\quad Y^{2}=\sum_{i,j}X_{i}^{4}X_{j}^{4} (616)

then we find can evaluate

YeikiXi/eikiXi=it4(ki4t46ki2t2+3)\displaystyle\langle Ye^{\text{i}k_{i}X_{i}}\rangle/\langle e^{\text{i}k_{i}X_{i}}\rangle=\sum_{i}t^{4}(k_{i}^{4}t^{4}-6k_{i}^{2}t^{2}+3) (617)

and

Φ=8i(3t812t6Xi2+9t3Xi42t2Xi6)+Y2\displaystyle\Phi=8\sum_{i}(3t^{8}-12t^{6}X_{i}^{2}+9t^{3}X_{i}^{4}-2t^{2}X_{i}^{6})+Y^{2} (618)

such that the error is

Error=8i(15t812t6Xi2+9t3Xi42t2Xi6).\displaystyle\text{Error}=-8\sum_{i}(15t^{8}-12t^{6}X_{i}^{2}+9t^{3}X_{i}^{4}-2t^{2}X_{i}^{6}). (619)

We can also find the leading order of Error2\langle\text{Error}^{2}\rangle is zero.

In general let us consider

Y=iemXi,emXi=em2t22,Y2=i,jemXiemXj,\displaystyle Y=\sum_{i}e^{mX_{i}},\quad\langle e^{mX_{i}}\rangle=e^{\frac{m^{2}t^{2}}{2}},\quad Y^{2}=\sum_{i,j}e^{mX_{i}}e^{mX_{j}}, (620)

and

Φ=i(em(mt2+2Xi)e2mXi)+Y2,\displaystyle\Phi=\sum_{i}(e^{m(-mt^{2}+2X_{i})}-e^{2mX_{i}})+Y^{2}, (621)

such that

Error=i(em2t21)(e2m2t2e2mXi),\displaystyle\text{Error}=\sum_{i}(e^{-m^{2}t^{2}}-1)(e^{2m^{2}t^{2}}-e^{2mX_{i}}), (622)
Error2=Ne4m2t2(e4m2t21)(em2t21)2,\displaystyle\langle\text{Error}^{2}\rangle=Ne^{4m^{2}t^{2}}(e^{4m^{2}t^{2}}-1)(e^{-m^{2}t^{2}}-1)^{2}, (623)

which is sub-dominant comparing to Y22\langle Y^{2}\rangle^{2}.

Exponential distribution

First let us choose

Y=iXi2,Xi2=2λ2,Y2=i,jXi2Xj2.\displaystyle Y=\sum_{i}X_{i}^{2},\quad\langle X_{i}^{2}\rangle=\frac{2}{\lambda^{2}},\quad Y^{2}=\sum_{i,j}X_{i}^{2}X_{j}^{2}. (624)

Correspondingly we find

YeikXieikXi=2(k+iλ)2\displaystyle\frac{\langle Ye^{\text{i}kX_{i}}\rangle}{\langle e^{\text{i}kX_{i}}\rangle}=-\frac{2}{(k+\text{i}\lambda)^{2}} (625)

and

Φ\displaystyle\Phi =\displaystyle= 56iXi4+i,jXi2Xj2.\displaystyle-\frac{5}{6}\sum_{i}X_{i}^{4}+\sum_{i,j}X_{i}^{2}X_{j}^{2}. (626)

So the error is given by

Error=i(5Xi4620λ4),\displaystyle\text{Error}=\sum_{i}\left(\frac{5X_{i}^{4}}{6}-\frac{20}{\lambda^{4}}\right), (627)
Error2=25N36λ8(8!4!2).\displaystyle\langle\text{Error}^{2}\rangle=\frac{25N}{36\lambda^{8}}(8!-4!^{2}). (628)

In general, let us consider

Y=ieβXi,eβXi=λλβ,Y2=i,jeβXieβXj.\displaystyle Y=\sum_{i}e^{\beta X_{i}},\quad\langle e^{\beta X_{i}}\rangle=\frac{\lambda}{\lambda-\beta},\quad Y^{2}=\sum_{i,j}e^{\beta X_{i}}e^{\beta X_{j}}. (629)

Then we can obtain

Φ\displaystyle\Phi =\displaystyle= i(eβXi(1eβXi+βXi))+i,jeβXieβXj.\displaystyle\sum_{i}\left(e^{\beta X_{i}}(1-e^{\beta X_{i}}+\beta X_{i})\right)+\sum_{i,j}e^{\beta X_{i}}e^{\beta X_{j}}. (630)

So the error is

Error=i(e2βXieβXi(1+βXi))Nβ2λ(βλ)2(λ2β),\displaystyle\text{Error}=\sum_{i}\left(e^{2\beta X_{i}}-e^{\beta X_{i}}(1+\beta X_{i})\right)-\frac{N\beta^{2}\lambda}{(\beta-\lambda)^{2}(\lambda-2\beta)}, (631)
Error2=N(2β4(3λ8β)(2βλ)3(4βλ)(λ3β)λ2β4(λβ)4(λ2β)2),\displaystyle\langle\text{Error}^{2}\rangle=N\left(\frac{2\beta^{4}(3\lambda-8\beta)}{(2\beta-\lambda)^{3}(4\beta-\lambda)(\lambda-3\beta)}-\frac{\lambda^{2}\beta^{4}}{(\lambda-\beta)^{4}(\lambda-2\beta)^{2}}\right), (632)

given that the condition

β<λ4,\displaystyle\beta<\frac{\lambda}{4}, (633)

to ensure that all the integrals to be convergent.

Poisson distribution

Let us choose

Y=iXi2,Xi2=λ2+λ,Y2=i,jXi2Xj2\displaystyle Y=\sum_{i}X_{i}^{2},\quad\langle X_{i}^{2}\rangle=\lambda^{2}+\lambda,\quad Y^{2}=\sum_{i,j}X_{i}^{2}X_{j}^{2} (634)

Then we find

YeikiXieikiXi=iλeiki(1+λeiki)\displaystyle\frac{\langle Ye^{\text{i}k_{i}X_{i}}\rangle}{\langle e^{\text{i}k_{i}X_{i}}\rangle}=\sum_{i}\lambda e^{\text{i}k_{i}}(1+\lambda e^{\text{i}k_{i}}) (635)

and

Φ=i(Xi(3+6Xi4Xi2))+i,jXi2Xj2.\displaystyle\Phi=\sum_{i}\left(X_{i}(-3+6X_{i}-4X_{i}^{2})\right)+\sum_{i,j}X_{i}^{2}X_{j}^{2}. (636)

Therefore the error is

Error=i(Xi(36Xi+4Xi2)λ(1+6λ+4λ2))\displaystyle\text{Error}=\sum_{i}\left(X_{i}(3-6X_{i}+4X_{i}^{2})-\lambda(1+6\lambda+4\lambda^{2})\right) (637)
Error2=Nλ(24λ(λ(6λ(λ+4)+23)+4)+1)\displaystyle\langle\text{Error}^{2}\rangle=N\lambda(24\lambda(\lambda(6\lambda(\lambda+4)+23)+4)+1) (638)
Error2Y221N3\displaystyle\frac{\langle\text{Error}^{2}\rangle}{\langle Y^{2}\rangle^{2}}\sim\frac{1}{N^{3}} (639)

In general let us consider

Y=ieβXi,eβXi=expλ(eβ1),Y2=i,jeβXieβXj.\displaystyle Y=\sum_{i}e^{\beta X_{i}},\quad\langle e^{\beta X_{i}}\rangle=\exp\lambda(e^{\beta}-1),\quad Y^{2}=\sum_{i,j}e^{\beta X_{i}}e^{\beta X_{j}}. (640)

Then we find

YeikiXieikiXi=e(eβ1)eikiλ\displaystyle\frac{\langle Ye^{\text{i}k_{i}X_{i}}\rangle}{\langle e^{\text{i}k_{i}X_{i}}\rangle}=e^{\left(e^{\beta}-1\right)e^{\text{i}k_{i}}\lambda} (641)

and

Φ\displaystyle\Phi =\displaystyle= i((2eβ1)Xie2βXi)+Y2.\displaystyle\sum_{i}(\left(2e^{\beta}-1\right)^{X_{i}}-e^{2\beta X_{i}})+Y^{2}. (642)

Therefore the error is

Error=i((2eβ1)Xi+e2βXie2λ(ee2βλ+λe2eβλ))\displaystyle\text{Error}=\sum_{i}\left(-\left(2e^{\beta}-1\right)^{X_{i}}+e^{2\beta X_{i}}-e^{-2\lambda}\left(e^{e^{2\beta}\lambda+\lambda}-e^{2e^{\beta}\lambda}\right)\right) (643)
Error2=e4λN(e4eβλ2e(eβ+1)2λ+e2(e2β+1)λe(e4β+3)λe4(eβ(eβ1)+1)λ+2e(e2β(2eβ1)+3)λ)\displaystyle\langle\text{Error}^{2}\rangle=-e^{-4\lambda}N\left(e^{4e^{\beta}\lambda}-2e^{\left(e^{\beta}+1\right)^{2}\lambda}+e^{2\left(e^{2\beta}+1\right)\lambda}-e^{\left(e^{4\beta}+3\right)\lambda}-e^{4\left(e^{\beta}\left(e^{\beta}-1\right)+1\right)\lambda}+2e^{\left(e^{2\beta}\left(2e^{\beta}-1\right)+3\right)\lambda}\right)

and

Error2Y221N3.\displaystyle\frac{\langle\text{Error}^{2}\rangle}{\langle Y^{2}\rangle^{2}}\sim\frac{1}{N^{3}}. (645)

Appendix C Explicte Examples: generalized statistical models

The exponential distribution

First let us consider the exponential distribution. It is straightforward to derive

Y=ijXiXj=N(N1)1λ2,\displaystyle\langle Y\rangle=\langle\sum_{i\neq j}X_{i}X_{j}\rangle=N(N-1)\frac{1}{\lambda^{2}}, (646)
Y2=2N(N1)(2N1)+N2(N1)2λ4\displaystyle\langle Y^{2}\rangle=\frac{2N(N-1)(2N-1)+N^{2}(N-1)^{2}}{\lambda^{4}} (647)

and

Φ\displaystyle\Phi =\displaystyle= ijpqXiXjXpXq+4ijpXi22XjXp+2ijXi22Xj22\displaystyle\sum_{i\neq j\neq p\neq q}X_{i}X_{j}X_{p}X_{q}+4\sum_{i\neq j\neq p}\frac{X_{i}^{2}}{2}X_{j}X_{p}+2\sum_{i\neq j}\frac{X_{i}^{2}}{2}\frac{X_{j}^{2}}{2} (648)

in particular as a consistency check

Φ=N2(N1)21λ4.\displaystyle\langle\Phi\rangle=N^{2}(N-1)^{2}\frac{1}{\lambda^{4}}. (649)

Therefore the error is

Error=2ijpXi2XjXp+32ijXi2Xj2Y2,\displaystyle\text{Error}=2\sum_{i\neq j\neq p}X_{i}^{2}X_{j}X_{p}+\frac{3}{2}\sum_{i\neq j}X_{i}^{2}X_{j}^{2}-\langle Y^{2}\rangle, (650)
Error2(422+16424)N6+#N5=#N5/λ8.\displaystyle\langle\text{Error}^{2}\rangle\sim(4*2*2+16-4*2*4)N^{6}+\#N^{5}=\#N^{5}/\lambda^{8}. (651)

Poisson distribution

First we compute

ijXiXj=N(N1)λ,\displaystyle\langle\sum_{i\neq j}X_{i}X_{j}\rangle=N(N-1)\lambda, (652)
Y2=2N(N1)(1+2(N1)λ)λ2+N2(N1)2λ2,\displaystyle\langle Y^{2}\rangle=2N(N-1)(1+2(N-1)\lambda)\lambda^{2}+N^{2}(N-1)^{2}\lambda^{2}, (653)

and

Φ\displaystyle\Phi =\displaystyle= ijpqXiXjXpXq+4ijp(Xi2Xi)XjXp+2ij(Xi2Xi)(Xj2Xj).\displaystyle\sum_{i\neq j\neq p\neq q}X_{i}X_{j}X_{p}X_{q}+4\sum_{i\neq j\neq p}(X_{i}^{2}-X_{i})X_{j}X_{p}+2\sum_{i\neq j}(X_{i}^{2}-X_{i})(X_{j}^{2}-X_{j}). (654)

One can check that

Φ=N2(N1)2λ2.\displaystyle\langle\Phi\rangle=N^{2}(N-1)^{2}\lambda^{2}. (655)

Therefore the error is

Error=Y2+4ijpXiXjXp+4ijXi2Xj2ijXiXj\displaystyle\text{Error}=-\langle Y^{2}\rangle+4\sum_{i\neq j\neq p}X_{i}X_{j}X_{p}+4\sum_{i\neq j}X_{i}^{2}X_{j}-2\sum_{i\neq j}X_{i}X_{j} (656)

and it is easy to check (note that only the first two term in (656) will contribute)

Error20N6+#N5.\displaystyle\langle\text{Error}^{2}\rangle\sim 0N^{6}+\#N^{5}. (657)

Appendix D Exact evaluation of z4\langle z^{4}\rangle for Brownian SYK at large T

We can only consider one ff term in the logarithm, we expect it gives the non-trivial contribution in the large TT limit. For instance we have

ei2(G14+G23)2+(G13G24)2+(G12+G34)2e2NTJ/qa,bsabGabq\displaystyle e^{\frac{\text{i}}{2}\sqrt{(\partial_{G_{14}}+\partial_{G_{23}})^{2}+(\partial_{G_{13}}-\partial_{G_{24}})^{2}+(\partial_{G_{12}}+\partial_{G_{34}})^{2}}}e^{2NTJ/q\sum_{a,b}s_{ab}G_{ab}^{q}} (658)
=m(i2)mm![(G14+G23)2+(G13G24)2+(G12+G34)2]m/2e2NTJ/qa,bsabGabq\displaystyle=\sum_{m}^{\infty}\frac{\left(\frac{\text{i}}{2}\right)^{m}}{m!}\left[(\partial_{G_{14}}+\partial_{G_{23}})^{2}+(\partial_{G_{13}}-\partial_{G_{24}})^{2}+(\partial_{G_{12}}+\partial_{G_{34}})^{2}\right]^{m/2}e^{2NTJ/q\sum_{a,b}s_{ab}G_{ab}^{q}} (659)
=m=0(i2)mm!k1k2k3m2!k12!k22!k32!(G14+G23)k1(G13G24)k2(G12+G34)k3e2NTJqa,bsabGabq,\displaystyle=\sum_{m=0}^{\infty}\frac{\left(\frac{\text{i}}{2}\right)^{m}}{m!}\sum_{k_{1}k_{2}k_{3}}\frac{\frac{m}{2}!}{\frac{k_{1}}{2}!\frac{k_{2}}{2}!\frac{k_{3}}{2}!}(\partial_{G_{14}}+\partial_{G_{23}})^{k_{1}}(\partial_{G_{13}}-\partial_{G_{24}})^{k_{2}}(\partial_{G_{12}}+\partial_{G_{34}})^{k_{3}}e^{\frac{2NTJ}{q}\sum_{a,b}s_{ab}G_{ab}^{q}}, (660)

and

(G14+G23)k1e2NTJ/q(G14q+G23q)=l1+l2=k1k1!l1!l2!G14l1G23l2e(2NTJ/q)G14qe(2NTJ/q)G23q\displaystyle(\partial_{G_{14}}+\partial_{G_{23}})^{k_{1}}e^{-2NTJ/q\left(G_{14}^{q}+G_{23}^{q}\right)}=\sum_{l_{1}+l_{2}=k_{1}}\frac{k_{1}!}{l_{1}!l_{2}!}\partial_{G_{14}}^{l_{1}}\partial_{G_{23}}^{l_{2}}e^{-(2NTJ/q)G_{14}^{q}}e^{-(2NTJ/q)G_{23}^{q}} (661)
=l1+l2=k1(2NTJq)k1/qk1!(l1/q)!(l2/q)!,\displaystyle=\sum_{l_{1}+l_{2}=k_{1}}\left(-\frac{2NTJ}{q}\right)^{k_{1}/q}\frac{k_{1}!}{(l_{1}/q)!(l_{2}/q)!}, (662)

where now k1k_{1} is a multiple of qq. Then (658) becomes

m=0(i/2)mm!k1k2k3(m/2)!(k1/2)!(k2/2)!(k3/2)!l1+l2=k1(2NTJq)k1/qk1!(l1/q)!(l2/q)!\displaystyle\sum_{m=0}^{\infty}\frac{(\text{i}/2)^{m}}{m!}\sum_{k_{1}k_{2}k_{3}}\frac{(m/2)!}{(k_{1}/2)!(k_{2}/2)!(k_{3}/2)!}\sum_{l_{1}+l_{2}=k_{1}}\left(-\frac{2NTJ}{q}\right)^{k_{1}/q}\frac{k_{1}!}{(l_{1}/q)!(l_{2}/q)!} (663)
×r1+r2=k2(iq2NTJq)k2/qk2!(r1/q)!(r2/q)!s1+s2=k3(iq2NTJq)k3/qk3!(s1/q)!(s2/q)!,\displaystyle\times\sum_{r_{1}+r_{2}=k_{2}}\left(\text{i}^{q}\frac{2NTJ}{q}\right)^{k_{2}/q}\frac{k_{2}!}{(r_{1}/q)!(r_{2}/q)!}\sum_{s_{1}+s_{2}=k_{3}}\left(\text{i}^{q}\frac{2NTJ}{q}\right)^{k_{3}/q}\frac{k_{3}!}{(s_{1}/q)!(s_{2}/q)!}, (664)
=m=0(i2)mm!k1+k2+k3=mm2!k12!k22!k32!(4NTJq)k1+k2+k3qk1!k2!k3!k1q!k2q!k3q!(iq)k2+k3q\displaystyle=\sum_{m=0}^{\infty}\frac{(\frac{\text{i}}{2})^{m}}{m!}\sum_{k_{1}+k_{2}+k_{3}=m}\frac{\frac{m}{2}!}{\frac{k_{1}}{2}!\frac{k_{2}}{2}!\frac{k_{3}}{2}!}\left(-\frac{4NTJ}{q}\right)^{\frac{k_{1}+k_{2}+k_{3}}{q}}\frac{k_{1}!k_{2}!k_{3}!}{\frac{k_{1}}{q}!\frac{k_{2}}{q}!\frac{k_{3}}{q}!}(-\text{i}^{q})^{\frac{k_{2}+k_{3}}{q}} (665)

where mm is also a multiple of qq. To evaluate it we can take an approximation when m>qm>q

k1+k2+k3=m(iq)k2+k3qk1!k2!k3!m!m2!k12!k22!k32!mq!k1q!k2q!k3q!{3,q=4k+23,q=4k,m/q is even1,q=4k,m/q is odd\displaystyle\sum_{k_{1}+k_{2}+k_{3}=m}(-\text{i}^{q})^{\frac{k_{2}+k_{3}}{q}}\frac{k_{1}!k_{2}!k_{3}!}{m!}\frac{\frac{m}{2}!}{\frac{k_{1}}{2}!\frac{k_{2}}{2}!\frac{k_{3}}{2}!}\frac{\frac{m}{q}!}{\frac{k_{1}}{q}!\frac{k_{2}}{q}!\frac{k_{3}}{q}!}\cong\begin{cases}3\,,\qquad q=4k+2\\ 3\,,\qquad q=4k,\quad m/q\text{ is even}\\ -1\,,~{}\quad q=4k,\quad m/q\text{ is odd}\end{cases}\ (666)

then up to a constant (658) becomes

{3e4NTJq2q,q=4k+23cosh(4NTJq2q)+sinh(4NTJq2q),q=4k.\displaystyle\begin{cases}3e^{\frac{4NTJ}{q2^{q}}},\hskip 128.0374ptq=4k+2\\ 3\cosh\left(\frac{4NTJ}{q2^{q}}\right)+\sinh\left(\frac{4NTJ}{q2^{q}}\right),\quad q=4k.\end{cases} (667)

Appendix E Lefschetz Thimbles

In this appendix, we review the method of Lefschetz thimble Witten:2010cx . Suppose we would like to evaluate the integral

Z=𝑑xieS,\displaystyle Z=\int_{\mathcal{M}_{\mathbb{R}}}dx^{i}e^{S}\,, (668)

where the integration contour is \mathcal{M}_{\mathbb{R}}. Then we complexify the manifold on which the integration is done to \mathcal{M}_{\mathbb{C}}. If we choose (S)\Re(S) to be the Morse function. The saddle points of the integral are the critical points of the Morse function. Around each critical point on \mathcal{M}_{\mathbb{C}} we introduce a set of local coordinates {zi}\{z_{i}\}. The Morse flow is determined by the flow equations

dzidt=gij¯S¯z¯j,dz¯idt=gij¯Szj.\displaystyle\frac{dz^{i}}{dt}=-g^{i\bar{j}}\frac{\partial\bar{S}}{\partial\bar{z}^{j}},\quad\frac{d\bar{z}^{i}}{dt}=-g^{i\bar{j}}\frac{\partial S}{\partial z^{j}}\ . (669)

We find

d(SS¯)dt=SzidzidtS¯z¯idz¯idt=0,\displaystyle\frac{d(S-\bar{S})}{dt}=\frac{\partial S}{\partial z^{i}}\frac{dz^{i}}{dt}-\frac{\partial\bar{S}}{\partial\bar{z}^{i}}\frac{d\bar{z}^{i}}{dt}=0\,, (670)

which implies that the imaginary part of SS is a constant along the flow. Each of the critical points is associated with a pair of flows, the thimble and the anti-thimble. The thimble is the “stable” direction such that the Morse function \mathcal{M}_{\mathbb{R}} decays along the thimble and the integral of exp(S)\exp(S) along the thimble converges. On the contrary, the anti-thimble is the “unstable” direction. Explicitly the boundary conditions for a particular critical point pσp_{\sigma} are

limtz(t)=pσ,for thimbles,\displaystyle\lim_{t\to-\infty}z(t)=p_{\sigma},\quad\text{for thimbles}, (671)
limt+z(t)=pσ,for anti-thimbles.\displaystyle\lim_{t\to+\infty}z(t)=p_{\sigma},\quad\text{for anti-thimbles}. (672)

The main statement in Witten:2010cx that we will use repeatedly is that the integral can be approximated by a weighted sum over integrals along the thimbles of each critical point

Z=ini𝒥i𝑑teS[t],\displaystyle Z=\sum_{i}n_{i}\int_{\mathcal{J}_{i}}dt\,e^{S[t]}\,, (673)

where ii runs over all the critical points, 𝒥i\mathcal{J}_{i} is the Lefschetz thimble attaching to the ithi^{\text{th}} critical point, and the weight nin_{i} is given by the intersection number between the anti-thimble and the original integration contour \mathcal{M}_{{\mathbb{R}}}.

E.1 some examples

To illustrate how this works, we first go through some simple examples.

E.1.1 The Gaussian function

Let us consider a simple example with

S=x2/2+σx.\displaystyle S=-x^{2}/2+\sigma x\ . (674)

The integral can be regarded as a zero-dimension theory with quadratic interaction and a complex source σ\sigma. The only critical point is at x=σ=a+ibx=\sigma=a+\text{i}b. The flow equation is

dxdt=x¯(aib).\displaystyle\frac{\text{d}x}{\text{d}t}=\bar{x}-(a-\text{i}b)\ . (675)

Expressing x=x1+ix2x=x_{1}+\text{i}x_{2}, we get the following equations

dx1dt=x1a,dx2dt=bx2.\displaystyle\frac{\text{d}x_{1}}{\text{d}t}=x_{1}-a,\quad\frac{\text{d}x_{2}}{\text{d}t}=b-x_{2}\ . (676)

The general solution can be easily solved

x1=a+c1et,x2=b+c2et,\displaystyle x_{1}=a+c_{1}e^{t},\quad x_{2}=b+c_{2}e^{-t}, (677)

where c1c_{1},c2c_{2} are two undetermined constants. The boundary conditions for the thimble is

(x1,x2)(a,b),t,\displaystyle(x_{1},x_{2})\to(a,b),\quad t\to-\infty, (678)

while for the anti-thimble we have

(x1,x2)(a,b),t+,\displaystyle(x_{1},x_{2})\to(a,b),\quad t\to+\infty, (679)

where (a,b)(a,b) is the critical point. Then with these boundary conditions we can get the equations for the thimble and the anti-thimble respectively

x2=b,\displaystyle x_{2}=b, (680)
x1=a.\displaystyle x_{1}=a. (681)

We plot the thimble and the anti-thimble in this case in Figure 14, where for simplicity we let σ=1+i\sigma=1+\text{i}.

We can also compare the saddle point solution with the exact result. The integral can evaluated as

dxex2/2+σx=2πeσ2/2.\displaystyle\int\text{d}xe^{-x^{2}/2+\sigma x}=\sqrt{2\pi}e^{\sigma^{2}/2}. (682)

While the saddle point solution gives

eσ2/2,\displaystyle e^{\sigma^{2}/2}, (683)

with the one-loop correction 2π\sqrt{2\pi} the saddle point analysis recovers the exact result.

Refer to caption
Figure 14: The red line denotes the thimble and the blue line denotes the anti-thimble. The anti-thimble intersects with the real line, so this saddle point contributes

E.1.2 The Airy function

A slightly less trivial example is the Airy action

Zλ=dxeS,S=iλ(x33x).\displaystyle Z_{\lambda}=\int_{-\infty}^{\infty}\text{d}xe^{S}\,,\quad S=\text{i}\lambda\left(\frac{x^{3}}{3}-x\right)\ . (684)

It is not hard to find that for real λ\lambda there are three “convergent” regions, namely (S)<\Re(S)<\infty, on the complex xx-plane:

x=reiθ,2nπ3θπ3+2nπ3,n=0,1,2.\displaystyle x=re^{\text{i}\theta},\quad\frac{2n\pi}{3}\leq\theta\leq\frac{\pi}{3}+\frac{2n\pi}{3},\quad n=0,1,2. (685)

In each convergent region, the Airy integrand is exponentially small. As we vary λ\lambda to complex values, we should deform the integration contour of xx accordingly so that the integral remains converge. This gives an analytic continuation of ZλZ_{\lambda}. The two critical points are located at x=±1x=\pm 1. The values of saddle points are

S±=2iλ3.\displaystyle S_{\pm}=\mp\frac{2\text{i}\lambda}{3}. (686)

Since along the (anti-)thimbles, the imaginary part of SS is a constant and

Im(S±)=2Re(λ)3.\displaystyle\text{Im}(S_{\pm})=\mp\frac{2\text{Re}(\lambda)}{3}. (687)

Therefore the two (anti-)thimbles associated with the two critical points will not intersect except for the case of Re(λ)=0\text{Re}(\lambda)=0. The thimble which connects critical points is called the Stoke ray. Using the Lefschetz thimbles 𝒥±\mathcal{J}_{\pm}, we can rewrite the integral as

Zλ=n+𝒥+expS+n𝒥expS.\displaystyle Z_{\lambda}=n_{+}\int_{\mathcal{J}_{+}}\exp S+n_{-}\int_{\mathcal{J}_{-}}\exp S. (688)

To solve the thimbles, let us take λ=1\lambda=1, then the flow equations are

dxdt=i(x¯21).\displaystyle\frac{\text{d}x}{\text{d}t}=\text{i}(\bar{x}^{2}-1). (689)

Expressing x=x1+ix2x=x_{1}+\text{i}x_{2}, we obtain

dx1dt=2x1x2,dx2dt=x12x221.\displaystyle\frac{\text{d}x_{1}}{\text{d}t}=2x_{1}x_{2},\quad\frac{\text{d}x_{2}}{\text{d}t}=x_{1}^{2}-x_{2}^{2}-1. (690)

We plot the anti-thimbles in Fig. 15

Refer to caption
Figure 15: The red line denotes the anti-thimble of x=1x=-1 and the Green line denotes the anti-thimble of x=1x=1.The grey regions are the convergent regions.

Therefore for λ=1\lambda=1 both of the saddle points contribute. This result is expected since that the two critical points are already located on the real line.

The problem we met in the main text is better illustrated by the following toy model

Z~=dxexpS,S=i(x33+x).\displaystyle\tilde{Z}=\int_{-\infty}^{\infty}\text{d}x\exp S,\quad S=\text{i}\left(\frac{x^{3}}{3}+x\right)\ . (691)

The integral is convergent and can be expressed by the Airy function

Z~=2πAi(133)33=0.83.\displaystyle\tilde{Z}=\frac{2\pi\text{Ai}\left(\frac{1}{\sqrt[3]{3}}\right)}{\sqrt[3]{3}}=0.83\ . (692)

We now try to compute the integral with saddle point approximation, where the saddle points are located at x=±ix=\pm\text{i}. The saddle point value, plus the 1-loop correction, of the integral at these two saddle points, Z~±\tilde{Z}_{\pm} are the same, and the sum of them is larger than the exact evaluation of the integral

Z~+=Z~=0.733,Z~++Z~>Z~.\displaystyle\tilde{Z}_{+}=\tilde{Z}_{-}=0.733,\quad\tilde{Z}_{+}+\tilde{Z}_{-}>\tilde{Z}. (693)

This is exactly the situation we are encountering. In this toy model, it is easy to show that the anti-thimble associated with the saddle point x=ix=-\text{i} does not intersect with the real axis, Figure. 16, so the saddle point x=ix=-\text{i} does not contribute to the integral.

Refer to caption
Figure 16: The red line is the anti-thimble of x=ix=\text{i} which intersects with the real line while the Green line is the anti-thimble of x=ix=-\text{i} which does not intersect with the real line. The grey regions are the convergent regions.

E.2 Multi-variable cases

Let us consider another example with two variables

Z=dσdg2πeS,S=logσiσg12g2.\displaystyle Z=\int\text{d}\sigma\frac{\text{d}g}{2\pi}e^{S}\,,\quad S=\log\sigma-\text{i}\sigma g-\frac{1}{2}g^{2}\ . (694)

The integral can be done directly to get

Z=0.\displaystyle Z=0\ . (695)

There are two saddle points

g±=±i,σ±=1.\displaystyle g_{\pm}=\pm\text{i},\quad\sigma_{\pm}=\mp 1\ . (696)

with saddle point contributions to the integral (on-shell action)

Z±=1e,Z+Z+=0,\displaystyle Z_{\pm}=\mp\frac{1}{\sqrt{e}},\quad Z_{-}+Z_{+}=0\ , (697)

Matching this with the exact solution suggests that n±=1n_{\pm}=1. Note that σ±=1\sigma_{\pm}=\mp 1 are already on the real line so corresponding anti-thimbles always intersect with the original contour. The flow equations are

dσdt=1σ¯ig¯,dgdt=iσ¯+g¯.\displaystyle\frac{d\sigma}{dt}=-\frac{1}{\bar{\sigma}}-\text{i}\bar{g},\quad\frac{\text{d}g}{dt}=-\text{i}\bar{\sigma}+\bar{g}. (698)

Expressing σ=σ1+iσ2\sigma=\sigma_{1}+\text{i}\sigma_{2} and g=g1+ig2g=g_{1}+\text{i}g_{2} we obtain the following differential equations

dσ1dt+g2+σ1σ12+σ22=0,dσ2dt+g1+σ2σ12+σ22=0,\displaystyle\frac{\text{d}\sigma_{1}}{\text{d}t}+g_{2}+\frac{\sigma_{1}}{\sigma_{1}^{2}+\sigma_{2}^{2}}=0,\quad\frac{\text{d}\sigma_{2}}{\text{d}t}+g_{1}+\frac{\sigma_{2}}{\sigma_{1}^{2}+\sigma_{2}^{2}}=0, (699)
dg1dt+σ2g1=0,dg2dt+σ1+g2=0.\displaystyle\frac{\text{d}g_{1}}{\text{d}t}+\sigma_{2}-g_{1}=0,\quad\frac{\text{d}g_{2}}{\text{d}t}+\sigma_{1}+g_{2}=0. (700)

We find that indeed these two saddles should both be included. We plot the gg-plane of the anti-thimbles in Fig. 17.

Refer to caption
Figure 17: Since σ±\sigma_{\pm} are already on the real line here we only plot the gg-plane of the anti-thimbles. Clearly both of these two anti-thimbles intersect with the real axis so these two saddle points both contribute to the integral.

Note that this example is special case of (172) with q=2q=2.

Flow equations in real coordinates

Sometimes it more convenient to use real form of the flow equations (669). We start with the relations

Sz=12Sx+12iSy,\displaystyle\frac{\partial S}{\partial z}=\frac{1}{2}\frac{\partial S}{\partial x}+\frac{1}{2\text{i}}\frac{\partial S}{\partial y}, (701)
S¯z¯=12S¯x12iS¯y,\displaystyle\frac{\partial\bar{S}}{\partial\bar{z}}=\frac{1}{2}\frac{\partial\bar{S}}{\partial x}-\frac{1}{2\text{i}}\frac{\partial\bar{S}}{\partial y}, (702)

where

z=x+iy.\displaystyle z=x+\text{i}y. (703)

Then we evaluate the equation as

dzdt+dz¯dt\displaystyle\frac{\text{d}z}{\text{d}t}+\frac{\text{d}\bar{z}}{\text{d}t} =SzS¯z¯=Re(S)xIm(S)y,\displaystyle=-\frac{\partial S}{\partial z}-\frac{\partial\bar{S}}{\partial\bar{z}}=-\frac{\partial\text{Re}(S)}{\partial x}-\frac{\partial\text{Im}(S)}{\partial y}, (704)
dzdtdz¯dt\displaystyle\frac{\text{d}z}{\text{d}t}-\frac{\text{d}\bar{z}}{\text{d}t} =SzS¯z¯=iRe(S)y+iIm(S)x,\displaystyle=\frac{\partial S}{\partial z}-\frac{\partial\bar{S}}{\partial\bar{z}}=-\text{i}\frac{\partial\text{Re}(S)}{\partial y}+\text{i}\frac{\partial\text{Im}(S)}{\partial x}, (705)

where we work in the flat space. Recall the Cauchy-Riemann equation we get

dxdt=Re(S)x,dydt=Re(S)y.\displaystyle\frac{\text{d}x}{\text{d}t}=-\frac{\partial\text{Re}(S)}{\partial x},\quad\frac{\text{d}y}{\text{d}t}=-\frac{\partial\text{Re}(S)}{\partial y}. (706)

To illustrate it we consider a special case in the Airy function

S=i(x33+x).\displaystyle S=\text{i}\left(\frac{x^{3}}{3}+x\right). (707)

In the complex plane its conjugate is

S¯=i(x¯33+x¯),\displaystyle\bar{S}=-\text{i}\left(\frac{\bar{x}^{3}}{3}+\bar{x}\right), (708)

and we can define the components

x=x1+ix2,x¯=x1ix2.\displaystyle x=x_{1}+\text{i}x_{2},\quad\bar{x}=x_{1}-\text{i}x_{2}. (709)

The flow equation in complex coordinates becomes

dxdt=S¯x¯=i(x¯2+1),\displaystyle\frac{\text{d}x}{\text{d}t}=-\frac{\partial\bar{S}}{\partial\bar{x}}=\text{i}(\bar{x}^{2}+1), (710)

which leads to the equations in real coordinates

dx1dt=2x1x2,dx2dt=x12x22+1.\displaystyle\frac{\text{d}x_{1}}{\text{d}t}=2x_{1}x_{2},\quad\frac{\text{d}x_{2}}{\text{d}t}=x_{1}^{2}-x_{2}^{2}+1. (711)

On the other hand we can get the equations with the real part of SS:

Re(S)=x2x12x2+x233.\displaystyle\text{Re}(S)=-x_{2}-x_{1}^{2}x_{2}+\frac{x_{2}^{3}}{3}. (712)

From the equations (706) we can recover the two flow equations (711).

Appendix F Averaged models

In this section, before talking about higher dimensional model we detour the main topic a little bit and consider the averaged theory in general. Let us consider a real scalar field with a chemical potential. The partition function of the theory is

Z=𝒟Φexp(𝑑xμϕ(x)μϕ(x)+J(x)ϕ(x))\displaystyle Z=\int{\cal D}\Phi\exp\left(-\int dx\,\partial_{\mu}\phi(x)\partial^{\mu}\phi(x)+J(x)\phi(x)\right)\, (713)

where J(x)J(x) is a random source drawn from some probability distribution. If the random coupling JJ does not depends on the spacetime i.e.

Jnconnected=κn\displaystyle\langle J^{n}\rangle_{\text{connected}}=\kappa_{n}\,\quad (714)

where κn\kappa_{n} is the nn-th cumulants of the probability distribution. This situation is like the regular SYK theory. To take the ensemble average we can expand the exponential as

ZJ\displaystyle\langle Z\rangle_{J} =\displaystyle= 𝒟Φexp(𝑑x0)nJnn!(𝑑xϕ(x))n\displaystyle\int{\cal D}\Phi\exp\left(-\int dx\,{\cal L}_{0}\right)\sum_{n}\frac{\langle J^{n}\rangle}{n!}\left(\int dx\,\phi(x)\right)^{n} (715)
=\displaystyle= 𝒟Φexp(𝑑x0)MGF(𝑑xϕ(x))\displaystyle\int{\cal D}\Phi\exp\left(-\int dx\,{\cal L}_{0}\right)\text{MGF}\left(\int dx\phi(x)\right) (716)

where MGF is the moment generating function of JJ. If JJ is Gaussian 𝒩(0,t2){\cal N}(0,t^{2}) then

MGF(𝑑xϕ(x))=exp(t22𝑑x𝑑yϕ(x)ϕ(y)),\displaystyle\text{MGF}\left(\int dx\phi(x)\right)=\exp\left(\frac{t^{2}}{2}\int dx\,dy\,\phi(x)\phi(y)\right), (717)
ZJ=𝒟Φexp(𝑑x0+t22𝑑x𝑑yϕ(x)ϕ(y))\displaystyle\langle Z\rangle_{J}=\int{\cal D}\Phi\exp\left(-\int dx\,{\cal L}_{0}+\frac{t^{2}}{2}\int dx\,dy\,\phi(x)\phi(y)\right)\, (718)

which is similar to SYK model that after the Gaussian average, bi-local interaction is generated. For the general distribution (714), the resulting averaged action is

ZJ=𝒟Φexp(𝑑x0+nκnn![𝑑xϕ(x)]n),\displaystyle\langle Z\rangle_{J}=\int{\cal D}\Phi\exp\left(-\int dx\,{\cal L}_{0}+\sum_{n}\frac{\kappa_{n}}{n!}[\int dx\phi(x)]^{n}\right)\,, (719)

so multi-local interactions will be generated. In particular, if the distribution is Poisson we have

ZJ=𝒟Φexp(𝑑x0+eλ𝑑xϕ(x)1),\displaystyle\langle Z\rangle_{J}=\int{\cal D}\Phi\exp\left(-\int dx\,{\cal L}_{0}+e^{\lambda\int dx\phi(x)}-1\right)\,, (720)

which is a highly non-local theory.

Alternatively we can require the random coupling JJ to depend on the spacetime i.e.

inJ(xi)=i=1,j=inδ(xixj)κn.\displaystyle\langle\prod^{n}_{i}J(x_{i})\rangle=\prod_{i=1,j=i}^{n}\delta(x_{i}-x_{j})\kappa_{n}\,. (721)

This situation is like the Brownian SYK theory Saad:2018bqo . Now we can take the ensemble average to get

ZJ\displaystyle\langle Z\rangle_{J} =\displaystyle= 𝒟Φexp(𝑑x0)nJnn!(𝑑xϕ(x))n\displaystyle\int{\cal D}\Phi\exp\left(-\int dx\,{\cal L}_{0}\right)\sum_{n}\frac{\langle J^{n}\rangle}{n!}\left(\int dx\,\phi(x)\right)^{n}
=\displaystyle= 𝒟Φexp(𝑑x0)nκnn!(𝑑xϕn(x))\displaystyle\int{\cal D}\Phi\exp\left(-\int dx\,{\cal L}_{0}\right)\sum_{n}\frac{\kappa_{n}}{n!}\left(\int dx\,\phi^{n}(x)\right)
=\displaystyle= 𝒟Φexp(𝑑x0)𝑑xMGF(ϕ(x)).\displaystyle\int{\cal D}\Phi\exp\left(-\int dx\,{\cal L}_{0}\right)\int dx\,\text{MGF}(\phi(x)). (723)

If JJ is Gaussian 𝒩(0,t2){\cal N}(0,t^{2}) then we have

ZJ=𝒟Φexp(𝑑x(0+t22ϕ2(x))).\displaystyle\langle Z\rangle_{J}=\int{\cal D}\Phi\exp\left(-\int dx\,({\cal L}_{0}+\frac{t^{2}}{2}\phi^{2}(x))\right). (724)

and if JJ is Poisson then we simply get

ZJ=𝒟Φexp(𝑑x[0+λ(eϕ(x)1)]),\displaystyle\langle Z\rangle_{J}=\int{\cal D}\Phi\exp\left(-\int dx\,[{\cal L}_{0}+\lambda(e^{\phi(x)}-1)]\right), (725)

which also coincides with results in Peng:2020rno .

Appendix G Computations in large N

CLT with μ=0\mu=0

In the appendix, we will consider the half-wormhole conjecture for statistical model in a more systematic way. The error functions we care about are

Error =Y2Y2+Y2Φ,\displaystyle=Y^{2}-\langle Y^{2}\rangle+\langle Y\rangle^{2}-\Phi, (726)
Error2\displaystyle\langle\text{Error}^{2}\rangle =(Y2Y2+Y2Φ)2,\displaystyle=\langle\left(Y^{2}-\langle Y^{2}\rangle+\langle Y\rangle^{2}-\Phi\right)^{2}\rangle, (727)

where Φ\Phi is defined in (63). And if the approximation is proper we must have

Error=0,Error2/Y41.\displaystyle\langle\text{Error}\rangle=0,\quad\langle\text{Error}^{2}\rangle/\langle Y^{4}\rangle\ll 1. (728)

The first requirement leads to

Φ=Y2\displaystyle\langle\Phi\rangle=\langle Y\rangle^{2} (729)

which is proved around (63), then we have

Error2=(Y2+Y2Φ)2Y22.\displaystyle\langle\text{Error}^{2}\rangle=\langle\left(Y^{2}+\langle Y\rangle^{2}-\Phi\right)^{2}\rangle-\langle Y^{2}\rangle^{2}. (730)

In the following we’ll give one type of function and try to explore the feasibility of the proposal, i.e. we compute the two error functions and check the relations (728) respectively.

First we consider the function YY consisting of NN identical and independent variables XiX_{i} as

Y=ijXiXj,\displaystyle Y=\sum_{i\neq j}X_{i}X_{j}, (731)

and for simplicity each combination only appears once since XiX_{i}’s commute with each other. The properties for XiX_{i} in this section are set as

X=0,X2=t2,\displaystyle\langle X\rangle=0,\quad\langle X^{2}\rangle=t^{2}, (732)

in the next section we’ll consider the distribution with a non-zero mean. But actually for any distribution the property (732) can always be satisfied since we can always take the subtraction

X~=XX.\displaystyle\tilde{X}=X-\langle X\rangle. (733)

therefore our discussion may be applied into the case with any distribution.

Obviously the first requirement in (728) is satisfied then we consider the computation for the second requirement

Y2=ijXi2Xj2+2ijkXi2XjXk+6ijklXiXjXkXl,\displaystyle Y^{2}=\sum_{i\neq j}X_{i}^{2}X_{j}^{2}+2\sum_{i\neq j\neq k}X^{2}_{i}X_{j}X_{k}+6\sum_{i\neq j\neq k\neq l}X_{i}X_{j}X_{k}X_{l}, (734)

where the first sum contains pp different terms

p=(N2),\displaystyle p={N\choose 2}, (735)

while we have 2p(N2)/22p(N-2)/2 for the second sum and p(p2(N2)1)/6p(p-2(N-2)-1)/6 for the third one. For the Φ\Phi in the approximation we have

Φ\displaystyle\Phi =1(2π)N𝑑keikXP(X)eikx(Yeikxeikx)2,\displaystyle=\frac{1}{(2\pi)^{N}}\int d\vec{k}\frac{e^{-\text{i}\vec{k}\vec{X}}}{P(\vec{X})}\langle e^{\text{i}\vec{k}\vec{x}}\rangle\left(\frac{\langle Ye^{\text{i}\vec{k}\vec{x}}\rangle}{\langle e^{\text{i}\vec{k}\vec{x}}\rangle}\right)^{2}, (736)
=1(2π)N𝑑keikXP(X)eikx(ijxixjeikxeikx)2,\displaystyle=\frac{1}{(2\pi)^{N}}\int d\vec{k}\frac{e^{-\text{i}\vec{k}\vec{X}}}{P(\vec{X})}\langle e^{\text{i}\vec{k}\vec{x}}\rangle\left(\sum_{i\neq j}\frac{\langle x_{i}x_{j}e^{\text{i}\vec{k}\vec{x}}\rangle}{\langle e^{\text{i}\vec{k}\vec{x}}\rangle}\right)^{2}, (737)

the computation is similar to that in Y2Y^{2} so that we can split the Φ\Phi function into three parts according to the three sums in (734)

Φ0\displaystyle\Phi_{0} =2(2π)N𝑑keikXP(X)eikxijxixjeikxeikxxixjeikxeikx,\displaystyle=\frac{2}{(2\pi)^{N}}\int d\vec{k}\frac{e^{-\text{i}\vec{k}\vec{X}}}{P(\vec{X})}\langle e^{\text{i}\vec{k}\vec{x}}\rangle\sum_{i\neq j}\frac{\langle x_{i}x_{j}e^{\text{i}\vec{k}\vec{x}}\rangle}{\langle e^{\text{i}\vec{k}\vec{x}}\rangle}\frac{\langle x_{i}x_{j}e^{\text{i}\vec{k}\vec{x}}\rangle}{\langle e^{\text{i}\vec{k}\vec{x}}\rangle}, (738)
=ij1(2π)2𝑑ki𝑑kjeikiXieikjXjP(Xi)P(Xj)eikixieikjxj(xieikixieikixi)2(xjeikjxjeikjxj)2,\displaystyle=\sum_{i\neq j}\frac{1}{\left(2\pi\right)^{2}}\int dk_{i}dk_{j}\frac{e^{-\text{i}k_{i}X_{i}}e^{-\text{i}k_{j}X_{j}}}{P(X_{i})P(X_{j})}\langle e^{\text{i}k_{i}x_{i}}\rangle\langle e^{\text{i}k_{j}x_{j}}\rangle\left(\frac{\langle x_{i}e^{\text{i}k_{i}x_{i}}\rangle}{\langle e^{\text{i}k_{i}x_{i}}\rangle}\right)^{2}\left(\frac{\langle x_{j}e^{\text{i}k_{j}x_{j}}\rangle}{\langle e^{\text{i}k_{j}x_{j}}\rangle}\right)^{2}, (739)
Φ1\displaystyle\Phi_{1} =2(2π)N𝑑keikXP(X)eikxijkxixjeikxeikxxixkeikxeikx,\displaystyle=\frac{2}{(2\pi)^{N}}\int d\vec{k}\frac{e^{-\text{i}\vec{k}\vec{X}}}{P(\vec{X})}\langle e^{\text{i}\vec{k}\vec{x}}\rangle\sum_{i\neq j\neq k}\frac{\langle x_{i}x_{j}e^{\text{i}\vec{k}\vec{x}}\rangle}{\langle e^{\text{i}\vec{k}\vec{x}}\rangle}\frac{\langle x_{i}x_{k}e^{\text{i}\vec{k}\vec{x}}\rangle}{\langle e^{\text{i}\vec{k}\vec{x}}\rangle}, (740)
=2ijkXjXk12π𝑑kieikiXiP(Xi)eikixi(xieikixieikixi)2,\displaystyle=2\sum_{i\neq j\neq k}X_{j}X_{k}\frac{1}{2\pi}\int dk_{i}\frac{e^{-\text{i}k_{i}X_{i}}}{P(X_{i})}\langle e^{\text{i}k_{i}x_{i}}\rangle\left(\frac{\langle x_{i}e^{\text{i}k_{i}x_{i}}\rangle}{\langle e^{\text{i}k_{i}x_{i}}\rangle}\right)^{2}, (741)
Φ2\displaystyle\Phi_{2} =6(2π)N𝑑keikXP(X)eikxijklxixjeikxeikxxkxleikxeikx,\displaystyle=\frac{6}{(2\pi)^{N}}\int d\vec{k}\frac{e^{-\text{i}\vec{k}\vec{X}}}{P(\vec{X})}\langle e^{\text{i}\vec{k}\vec{x}}\rangle\sum_{i\neq j\neq k\neq l}\frac{\langle x_{i}x_{j}e^{\text{i}\vec{k}\vec{x}}\rangle}{\langle e^{\text{i}\vec{k}\vec{x}}\rangle}\frac{\langle x_{k}x_{l}e^{\text{i}\vec{k}\vec{x}}\rangle}{\langle e^{\text{i}\vec{k}\vec{x}}\rangle}, (742)
=6ijklXiXjXkXl.\displaystyle=6\sum_{i\neq j\neq k\neq l}X_{i}X_{j}X_{k}X_{l}. (743)

For simplicity we can define the function below

ϕi=12π𝑑keikXiP(Xi)eikx(xeikxeikx)2,\displaystyle\phi_{i}=\frac{1}{2\pi}\int dk\frac{e^{-\text{i}kX_{i}}}{P(X_{i})}\langle e^{\text{i}kx}\rangle\left(\frac{\langle xe^{\text{i}kx}\rangle}{\langle e^{\text{i}kx}\rangle}\right)^{2}, (744)

so that the Φ\Phi functions can be expressed as

Φ0=ijϕiϕj,\displaystyle\Phi_{0}=\sum_{i\neq j}\phi_{i}\phi_{j}, (745)
Φ1=2ijkϕiXjXk.\displaystyle\Phi_{1}=2\sum_{i\neq j\neq k}\phi_{i}X_{j}X_{k}. (746)

And note that the ϕ\phi function has the property

ϕi=Xi2,\displaystyle\langle\phi_{i}\rangle=\langle X_{i}\rangle^{2}, (747)

which is useful in the later computation, where XiX_{i}’s are identical.

Then following the procedure we have

Error2\displaystyle\langle\text{Error}^{2}\rangle =(Y2+Y2Φ)2Y22,\displaystyle=\langle\left(Y^{2}+\langle Y\rangle^{2}-\Phi\right)^{2}\rangle-\langle Y^{2}\rangle^{2}, (748)
=(ij(Xi2Xj2ϕiϕj)+2ijk(Xi2ϕi)XjXk)2p2X24,\displaystyle=\left\langle\left(\sum_{i\neq j}\left(X_{i}^{2}X_{j}^{2}-\phi_{i}\phi_{j}\right)+2\sum_{i\neq j\neq k}\left(X^{2}_{i}-\phi_{i}\right)X_{j}X_{k}\right)^{2}\right\rangle-p^{2}\langle X^{2}\rangle^{4}, (749)

where the last sum cancels out and pp is defined in (735). The cross terms between different sums are zero due to the zero mean of XiX_{i}, therefore we can only consider the squares of each sum

Error2=(ij(Xi2Xj2ϕiϕj))2+(2ijk(Xi2ϕi)XjXk)2p2X24.\displaystyle\langle\text{Error}^{2}\rangle=\left\langle\left(\sum_{i\neq j}\left(X_{i}^{2}X_{j}^{2}-\phi_{i}\phi_{j}\right)\right)^{2}+\left(2\sum_{i\neq j\neq k}\left(X^{2}_{i}-\phi_{i}\right)X_{j}X_{k}\right)^{2}\right\rangle-p^{2}\langle X^{2}\rangle^{4}. (750)

The numbers of the diagonal terms in the two sums in the first average are

p,4p(N2),\displaystyle p,\quad 4p(N-2), (751)

while the non-diagonal terms in the first sum have

p(p2(N2)1) for Xi2Xj2Xk2Xl2\displaystyle p(p-2(N-2)-1)\quad\text{ for }\quad\langle X_{i}^{2}X_{j}^{2}X_{k}^{2}X_{l}^{2}\rangle (752)
2p(N2) for Xi4Xj2Xk2,\displaystyle 2p(N-2)\quad\text{ for }\quad\langle X_{i}^{4}X_{j}^{2}X_{k}^{2}\rangle, (753)

and the non-diagonal terms in the second sum vanish due to the zero mean of XiX_{i}.

We want to talk about (728) in this case, and the ϕ\phi function has its role in the discussion. If the function ϕ\phi satisfies the relation

Xi2ϕi=f(Xi),f(Xi)=0\displaystyle X_{i}^{2}-\phi_{i}=f(X_{i}),\quad\langle f(X_{i})\rangle=0 (754)

where f(Xi)f(X_{i}) is an unknown function of XiX_{i} . Actually this relation leads to

t2=Xi2ϕi=0,\displaystyle t^{2}=\langle X_{i}^{2}-\phi_{i}\rangle=0, (755)

therefore in this case the distribution should be non-trivial with μ=0\mu=0 and t=0t=0. Usually we do not consider such cases, but we can still show how it works following this computation. Then in this case the dominant terms in Error2\langle\text{Error}^{2}\rangle will be (752) as it contains the largest number of terms. The ϕ\phi terms are not included since the averages of the non-diagonal ϕ\phi terms are zero because of (732) and (747). Then we can find that the dominant terms of Error2\langle\text{Error}^{2}\rangle which locate at the order p2p^{2} or N4N^{4} cancel out in (750). Therefore the relation (728) is satisfied since the dominant terms in Y2Y^{2} are also at the order N4N^{4}, so that the proposal holds in this case when NN is sufficiently large.

While if ϕ\phi has the relation below like what we have in the Gaussian case

Xi2ϕi=g(Xi),g(Xi)0\displaystyle X_{i}^{2}-\phi_{i}=g(X_{i}),\quad\langle g(X_{i})\rangle\neq 0 (756)

where g(Xi)g(X_{i}) is also unknown and we denote the average with a nonzero constant CC. Then the diagonal terms in the second sum can contribute to the dominant terms. We can have the explicit computation as the following,

2ijk(Xi2ϕi)XjXk2(N2)CjkXjXk,\displaystyle 2\sum_{i\neq j\neq k}\left(X^{2}_{i}-\phi_{i}\right)X_{j}X_{k}\sim 2(N-2)C\sum_{j\neq k}X_{j}X_{k}, (757)

where the approximation is valid as we only care about the maximal numbers. In the square this term is similar to (734) except we have an extra parameter 4(N2)24(N-2)^{2}. The non-vanishing contribution after the average will be the diagonal terms in the square as XiX_{i} has a zero mean, which means the number of the terms is

4(N2)2p=2N(N1)(N2)2.\displaystyle 4(N-2)^{2}p=2N(N-1)(N-2)^{2}. (758)

This contribution also locate at the order N4N^{4} which is additional comparing to the previous case, so that the dominant terms can not cancel out in (750). Therefore the proposal fails in this case.

After having discussed a simple example We can try to consider the general one

Y=i1iqXi1Xiq,\displaystyle Y=\sum_{i_{1}\dots i_{q}}X_{i_{1}}\dots X_{i_{q}}, (759)

where the sum contains pp terms

p=(Nq)\displaystyle p={N\choose q} (760)

and XiX_{i}’s are still identical and independent variables. Following the procedure above we have

Y2=i1iqXi12Xiq2+i1iq+1Xi12Xiq12XiqXiq+1++i1i2qXi1Xi2q,\displaystyle Y^{2}=\sum_{i_{1}\dots i_{q}}X_{i_{1}}^{2}\dots X_{i_{q}}^{2}+\sum_{i_{1}\dots i_{q+1}}X_{i_{1}}^{2}\dots X_{i_{q-1}}^{2}X_{i_{q}}X_{i_{q+1}}+\dots+\sum_{i_{1}\dots i_{2q}}X_{i_{1}}\dots X_{i_{2q}}, (761)

the numbers of the total terms for each sum are respectively

(Nq),(Nq)(q1)(Nq1),,(Nq)(Nqq),\displaystyle{N\choose q},\quad{N\choose q}{q\choose 1}{N-q\choose 1},\dots,{N\choose q}{N-q\choose q}, (762)

and the numbers for each single term in each sum or the repetitions are

1,(21),(42),,(2qq).\displaystyle 1,\quad{2\choose 1},\quad{4\choose 2},\dots,{2q\choose q}. (763)

When q=2q=2 we can find that the computations in the two cases match. For the function Φ\Phi similarly we can define the functions

Φ0,Φ1,,Φq,\displaystyle\Phi_{0},\quad\Phi_{1},\dots,\Phi_{q}, (764)

which correspond to the terms in (761) respectively. And we can also introduce the ϕ\phi functions to express each Φs\Phi_{s}, substituting the Xi2X^{2}_{i} by ϕi\phi_{i} in (761) we can get all the Φs\Phi_{s}’s just like (745),(746).

Following the previous procedure the square of the error function can be evaluated as

Error2\displaystyle\langle\text{Error}^{2}\rangle =(Y2+Y2Φ)2Y22,\displaystyle=\langle\left(Y^{2}+\langle Y\rangle^{2}-\Phi\right)^{2}\rangle-\langle Y^{2}\rangle^{2}, (765)
=(s=0qi1iq+s(Ys2Φs))2p2X22q,\displaystyle=\left\langle\left(\sum_{s=0}^{q}\sum_{i_{1}\dots i_{q+s}}\left(Y^{2}_{s}-\Phi_{s}\right)\right)^{2}\right\rangle-p^{2}\langle X^{2}\rangle^{2q}, (766)
=s=0q(i1iq+s(Ys2Φs))2p2X22q,\displaystyle=\left\langle\sum_{s=0}^{q}\left(\sum_{i_{1}\dots i_{q+s}}\left(Y^{2}_{s}-\Phi_{s}\right)\right)^{2}\right\rangle-p^{2}\langle X^{2}\rangle^{2q}, (767)

where the index ss labels the components in both Y2Y^{2},Φ\Phi and the cross terms among the different sums vanish due to zero mean of XiX_{i}.

Then we’ll encounter the previous problem again that there are two different conditions (754),(756) for ϕ\phi, and we also first consider the former one. The total numbers of the diagonal terms in the above each sum ss are respectively

(Nq),(21)(Nq)(q1)(Nq1),,(2(q1)q1)(Nq)(qq1)(Nqq1),\displaystyle{N\choose q},\quad{2\choose 1}{N\choose q}{q\choose 1}{N-q\choose 1},\dots,{2(q-1)\choose q-1}{N\choose q}{q\choose q-1}{N-q\choose q-1}, (768)

which are the combinations of (762) and (763) and the last one in (761) cancels out. While about the non-diagonal terms as the averages of XiX_{i} is zero, the nonzero non-diagonal terms only appears in the sum s=0s=0. And as the average of Φi\Phi_{i} is also zero, the nonzero non-diagonal terms in the sum s=0s=0 will only come from Y02Y^{2}_{0}. The computation of the square of this term is similar to (761) except the replacement of XiX_{i} by Xi2X_{i}^{2},

Ynn4=i1iq+1Xi14Xiq14Xiq2Xiq+12++i1i2qXi12Xi2q2,\displaystyle Y^{4}_{nn}=\sum_{i_{1}\dots i_{q+1}}X_{i_{1}}^{4}\dots X_{i_{q-1}}^{4}X_{i_{q}}^{2}X_{i_{q+1}}^{2}+\dots+\sum_{i_{1}\dots i_{2q}}X_{i_{1}}^{2}\dots X_{i_{2q}}^{2}, (769)

where nnnn means nonzero non-diagonal. The non-zero contributions in (767) come from the square of the first sum s=0s=0 and the diagonal terms in the square of the other sums s>0s>0. And note that the last one labelled by s=qs=q cancels out, and in (768) the last one is subordinate to the square of the first term when NN,qq,N/qN/q are sufficiently large. Then under the condition (754) the dominant terms will be the first sum with square according to the equation (768). The square of the first sum in (767) gives the dominant terms which are the last one in (769), it contains the number

(Nq)(Nqq).\displaystyle{N\choose q}{N-q\choose q}. (770)

Then after the average the dominant terms in the square of the error becomes

Error2=((Nq)(Nqq)(Nq)(Nq))X22q,\displaystyle\langle\text{Error}^{2}\rangle=\left({N\choose q}{N-q\choose q}-{N\choose q}{N\choose q}\right)\langle X^{2}\rangle^{2q}, (771)

which is subordinate comparing to Y4\langle Y^{4}\rangle.

While under the condition (756) the diagonal terms in the squares of the sums with s>0s>0 will contribute. The computation is similar to the previous case with q=2q=2, the nonzero average increases the number of the diagonal or nonzero terms. As an example we calculate the last non-zero one in (767)

ji1i2q2(Xj2ϕj)Xi1Xi2q2=αi1i2q2Xi1Xi2q2\displaystyle\sum_{ji_{1}\dots i_{2q-2}}(X_{j}^{2}-\phi_{j})X_{i_{1}}\dots X_{i_{2q-2}}=\alpha\sum_{i_{1}\dots i_{2q-2}}X_{i_{1}}\dots X_{i_{2q-2}} (772)

where

α=C(Nq)(qq1)(Nqq1)/(N2q2),\displaystyle\alpha=C{N\choose q}{q\choose q-1}{N-q\choose q-1}/{N\choose 2q-2}, (773)

and for simplicity we denote the ϕ\phi term with CC. Since we only care about the number of the terms, the denotation does not matter for the result. The constant CC is defined in (756) and the numerator is the number of the terms in the sum labelled by s=q1s=q-1, while the denominator is the number of the sum of the right hand side of (772). In the square of this term only the diagonal ones survive in the average, we have the number

(Nq)2(qq1)2(Nqq1)2/(N2q2)N2qp2.\displaystyle{N\choose q}^{2}{q\choose q-1}^{2}{N-q\choose q-1}^{2}/{N\choose 2q-2}\sim N^{2q}\sim p^{2}. (774)

Actually we can find that except the last sum all the sums in (767) will contribute to the dominant terms. So that the proposal will fail as the first sum cancels out with the last one in (767), while we have other additional contributions to the dominant terms.

CLT with μ0\mu\neq 0

In the above section we have discussed the case with μ=0\mu=0, after that we can move to consider the case with μ0\mu\neq 0. Remember that for the distribution with a non-zero mean we deal it with the subtraction (733), now we keep this non-trivial mean. The difference between them may be illustrated by a simple example

Y=ij(XiXi)(XjXj),\displaystyle Y=\sum_{i\neq j}\left(X_{i}-\langle X_{i}\rangle\right)\left(X_{j}-\langle X_{j}\rangle\right), (775)
Y=ijXiXjijXiXj.\displaystyle Y=\sum_{i\neq j}X_{i}X_{j}-\left\langle\sum_{i\neq j}X_{i}X_{j}\right\rangle. (776)

The two definitions are not equivalent which is manifest in the higher powers of YY. Actually the former one is equivalent to the case with μ=0\mu=0 in the previous section, the latter without the subtraction is what we’ll consider in this section.

We still start with the simple case with q=2q=2

Y=ijXiXj,\displaystyle Y=\sum_{i\neq j}X_{i}X_{j}, (777)
Y2=ijXi2Xj2+2ijkXi2XjXk+6ijklXiXjXkXl.\displaystyle Y^{2}=\sum_{i\neq j}X_{i}^{2}X_{j}^{2}+2\sum_{i\neq j\neq k}X^{2}_{i}X_{j}X_{k}+6\sum_{i\neq j\neq k\neq l}X_{i}X_{j}X_{k}X_{l}. (778)

About Φ\Phi we can have similar definition to (736), and the definition and the requirement for the error are the same as before. We also separate each sum in both Y2Y^{2} and Φ\Phi so that the square of the error can be expressed as

Error2\displaystyle\langle\text{Error}^{2}\rangle =(Y2+Y2Φ)2Y22,\displaystyle=\langle\left(Y^{2}+\langle Y\rangle^{2}-\Phi\right)^{2}\rangle-\langle Y^{2}\rangle^{2}, (779)
=(Y02+Y02Φ0+Y12+Y12Φ1)2Y22.\displaystyle=\left\langle\left(Y_{0}^{2}+\langle Y\rangle^{2}_{0}-\Phi_{0}+Y_{1}^{2}+\langle Y\rangle^{2}_{1}-\Phi_{1}\right)^{2}\right\rangle-\langle Y^{2}\rangle^{2}. (780)

And note that here we can not separate the square of the sum into the sum of the square of each sum, since now the cross terms are not zero due to the nonzero mean of XiX_{i}. When NN\to\infty we can only consider the dominant terms, but as the case is different to before the terms we want will also be different. Now the dominant terms in the two parts in (780) are both in the second sum in (778) or Y12Y_{1}^{2} (780), since this sum contains the largest number of terms and is non-zero under the averagre because of the non-zero mean. Which means we can only consider the cross terms in the square of this sum which contains the largest number of terms. By the direct calculation we have

Φ1=2ijkϕiXjXk,\displaystyle\Phi_{1}=2\sum_{i\neq j\neq k}\phi_{i}X_{j}X_{k}, (781)

where ϕ\phi is defined in (744). Then we have

Y12+Y12Φ1=2ijk((Xi2ϕi)XjXk+Xi2XjXk),\displaystyle Y_{1}^{2}+\langle Y\rangle^{2}_{1}-\Phi_{1}=2\sum_{i\neq j\neq k}\left(\left(X_{i}^{2}-\phi_{i}\right)X_{j}X_{k}+\langle X_{i}\rangle^{2}\langle X_{j}\rangle\langle X_{k}\rangle\right), (782)

while the dominant terms can be written as

Error2d=(2ijk((Xi2ϕi)XjXk+Xi2XjXk))22ijkXi2XjXk2.\displaystyle\langle\text{Error}^{2}\rangle_{d}=\left\langle\left(2\sum_{i\neq j\neq k}\left(\left(X_{i}^{2}-\phi_{i}\right)X_{j}X_{k}+\langle X_{i}\rangle^{2}\langle X_{j}\rangle\langle X_{k}\rangle\right)\right)^{2}\right\rangle-\left\langle 2\sum_{i\neq j\neq k}X_{i}^{2}X_{j}X_{k}\right\rangle^{2}. (783)

The dominant terms of the first part are the cross terms, whose number is at the same order of NN as the second part. So to evaluate the approximation we can only consider the two terms below

(Xi2ϕi)XjXk+Xi2XjXk,Xi2XjXk,\displaystyle\langle\left(X_{i}^{2}-\phi_{i}\right)X_{j}X_{k}+\langle X_{i}\rangle^{2}\langle X_{j}\rangle\langle X_{k}\rangle\rangle,\quad\langle X_{i}^{2}X_{j}X_{k}\rangle, (784)

where the variables are identical and independent. Recall that we have the equation (747) therefore the two expressions are equal under the average. Explicitly the cross terms also contain many parts, we can also write down the one with the largest number of terms,

(Xi2ϕi)XjXk(Xl2ϕl)XmXn,4(N2)(N2)(N32)(N5),\displaystyle\left(X_{i}^{2}-\phi_{i}\right)X_{j}X_{k}\left(X_{l}^{2}-\phi_{l}\right)X_{m}X_{n},\quad 4{N\choose 2}(N-2){N-3\choose 2}(N-5), (785)

while the second part in (783) contains

4(N2)2(N2)2.\displaystyle 4{N\choose 2}^{2}(N-2)^{2}. (786)

Therefore the dominant contribution in (783) will cancel out, which makes the proposal hold.

We can also consider the problem that occurs in the case with μ=0\mu=0 when the average of X2ϕX^{2}-\phi is not zero, and we can find that there’s no such problem here. As the dominant term (783) contains the largest number of XX terms, therefore even when X2ϕX^{2}-\phi is a constant the total number of XX terms in (783) will not change.

Next we consider the case with general qq. Following the previous procedure for the error we can get an expression similar to (780)

Error2=(s=0q1(Ys2+Ys2Φs))2Y22,\displaystyle\langle\text{Error}^{2}\rangle=\left\langle\left(\sum_{s=0}^{q-1}(Y_{s}^{2}+\langle Y\rangle^{2}_{s}-\Phi_{s})\right)^{2}\right\rangle-\langle Y^{2}\rangle^{2}, (787)

and the dominant sum will also be the last non-zero one, i.e. the sum labelled with s=q1s=q-1. To find out the behavior of the dominant terms the relation of the two functions below is important,

(Xi12ϕi1)Xi2Xi2q1+Xi12Xi2Xi2q1,Xi12Xi2Xi2q1,\displaystyle\langle\left(X_{i_{1}}^{2}-\phi_{i_{1}}\right)X_{i_{2}}\dots X_{i_{2q-1}}+\langle X_{i_{1}}\rangle^{2}\langle X_{i_{2}}\rangle\dots\langle X_{i_{2q-1}}\rangle\rangle,\quad\langle X_{i_{1}}^{2}X_{i_{2}}\dots X_{i_{2q-1}}\rangle, (788)

we want to determine whether they are equal. And we can find that the relation (747) still holds thus the approximation is also proper in general qq.

SYK with one time point with μ=0\mu=0

Like in the previous section given any distribution we can always construct a variable with zero mean, such as

X~=XX,\displaystyle\tilde{X}=X-\langle X\rangle, (789)

so that we can consider any distribution we want.

We first consider a simple case with q=2q=2, the Hamiltonian has the form

Y=ijsgn(ij)XiXj,\displaystyle Y=\sum_{i\neq j}\text{sgn}(ij)X_{i}X_{j}, (790)

where XiX_{i}’s are identical and independent variables and the sign term mimics the sign function in the SYK model. Note that here we have no need to compute the explicit form of the sign function, since it has no effect yet. It is very similar to the previous case, but the difference is that here all XiX_{i}’s only appear once in the sum. Therefore for the square we have

Y2=ijXi2Xj2+2ijklsgn(ijkl)XiXjXkXl,\displaystyle Y^{2}=\sum_{i\neq j}X_{i}^{2}X_{j}^{2}+2\sum_{i\neq j\neq k\neq l}\text{sgn}(ijkl)X_{i}X_{j}X_{k}X_{l}, (791)

comparing to (734) it lacks the middle term. Our previous computation shows that the approximation fails in the Gaussian case with μ=0\mu=0 due to the middle term, thus the approximation is proper in this case no matter what the distribution is.

We can have an explicit computation for the proposal, but first the convention needs to be consistent with the SYK model. Now the total number NN counts the whole Majorana fermions and the number qq is for the number of the fermions in the interaction. Then for N/q=2N/q=2 the expression (790) is still valid for the partition function, except that the NN and qq have different meanings. Then the number of XiX_{i} in (790) becomes

N!((N/2)!)2,\displaystyle\frac{N!}{\left(\left(N/2\right)!\right)^{2}}, (792)

and the sum contains the number of the terms

N!2!((N/2)!)2.\displaystyle\frac{N!}{2!\left(\left(N/2\right)!\right)^{2}}. (793)

Following the previous process we have

Y2Φ=ij(Xi2Xj2ϕiϕj)\displaystyle Y^{2}-\Phi=\sum_{i\neq j}\left(X_{i}^{2}X_{j}^{2}-\phi_{i}\phi_{j}\right) (794)

where ϕ\phi is defined in (744) and for the error

Error2=(ij(Xi2Xj2ϕiϕj))2ijXi2Xj22.\displaystyle\langle\text{Error}^{2}\rangle=\left\langle\left(\sum_{i\neq j}\left(X_{i}^{2}X_{j}^{2}-\phi_{i}\phi_{j}\right)\right)^{2}\right\rangle-\left\langle\sum_{i\neq j}X_{i}^{2}X_{j}^{2}\right\rangle^{2}. (795)

Since there are only two parts in the above function and the sums are the same, we can only compare the two expressions in the brackets. The average of ϕ\phi is zero due to the zero mean of XiX_{i}, so the proposal holds in the case. Explicitly the dominant terms in (795) come from the cross term in the first part and the whole second part, the number becomes

N!2!((N/2)!)2(N!2!((N/2)!)21)(N!2!((N/2)!)2)2.\displaystyle\frac{N!}{2!\left(\left(N/2\right)!\right)^{2}}\left(\frac{N!}{2!\left(\left(N/2\right)!\right)^{2}}-1\right)-\left(\frac{N!}{2!\left(\left(N/2\right)!\right)^{2}}\right)^{2}. (796)

We can find the dominant terms cancel out so that the proposal is valid in this simple case.

Then we can consider a more general Hamiltonian with arbitrary NN and qq. Here we take the more familiar convention in SYK and still define p=N/qp=N/q, then the partition function becomes

z=Asgn(A)JA1JAp.\displaystyle z=\sum_{A}\text{sgn}(A)J_{A_{1}}\dots J_{A_{p}}. (797)

Before we check the approximation we can first have some computation about the numbers of the different terms in this partition function. The total number of the JAJ_{A}’s is

(Nq),\displaystyle{N\choose q}, (798)

and the number of the combinations of JAJ_{A}’s is

N!p!(q!)p,\displaystyle\frac{N!}{p!(q!)^{p}}, (799)

while the times for a single JAJ_{A} appearing in the Hamiltonian is

(Nq)!(p1)!(q!)p1.\displaystyle\frac{(N-q)!}{(p-1)!(q!)^{p-1}}. (800)

The square of the partition function can be given as

z2\displaystyle z^{2} =Asgn(A)2JA12JAp2+A,Bsgn(A)sgn(B)JA12JAp22JAp1JApJBp1JBp\displaystyle=\sum_{A}\text{sgn}(A)^{2}J_{A_{1}}^{2}\dots J_{A_{p}}^{2}+\sum_{A,B}\text{sgn}(A)\text{sgn}(B)J_{A_{1}}^{2}\dots J_{A_{p-2}}^{2}J_{A_{p-1}}J_{A_{p}}J_{B_{p-1}}J_{B_{p}}
++A,Bsgn(A)sgn(B)JA1JApJB1JBp\displaystyle\quad+\dots+\sum_{A,B}\text{sgn}(A)\text{sgn}(B)J_{A_{1}}\dots J_{A_{p}}J_{B_{1}}\dots J_{B_{p}} (801)

the first non-diagonal term appears with four different JAJ_{A}’s as the combination AA should be the permutation of 1,,N1,\dots,N. And in the sums there may be many identical terms, here we may not give the explicit forms. To figure out the explicit expressions for the numbers in each sum we can count the numbers of different combinations of rr different JAJ_{A}’s in the product of the combinations of AA and BB. Obviously there’s only one combination in the case r=1r=1, actually r=1r=1 is identical to r=0r=0 since the last JAiJ_{A_{i}} in a combination AA can be determined by the other p1p-1 JAiJ_{A_{i}}’s. For the larger rr we have to subtract the combinations in rir-i (i<r)(i<r) to get the correct ones, as it can give rir-i cases in the product of two combinations. Therefore we have

r=2,(2q)!2!(q!)21,\displaystyle r=2,\quad\frac{(2q)!}{2!(q!)^{2}}-1, (802)
r=3,(3q)!3!(q!)3(32)((2q)!2!(q!)21)1,\displaystyle r=3,\quad\frac{(3q)!}{3!(q!)^{3}}-{3\choose 2}\left(\frac{(2q)!}{2!(q!)^{2}}-1\right)-1, (803)

from the above we can introduce the function nrn_{r} for this counting

nr=(rq)!r!(q!)rs=2r1(rs)ns1.\displaystyle n_{r}=\frac{(rq)!}{r!(q!)^{r}}-\sum_{s=2}^{r-1}{r\choose s}n_{s}-1. (804)

From the calculation we can take mathematical induction to derive a simpler form

nr=s=1r(sq)!s!(q!)s(rs)(1)rs+(1)r,r>1,\displaystyle n_{r}=\sum_{s=1}^{r}\frac{(sq)!}{s!(q!)^{s}}{r\choose s}(-1)^{r-s}+(-1)^{r},\quad r>1, (805)

and

n0=1,\displaystyle n_{0}=1, (806)

where n1n_{1} is absent according to the equation (801).

Then we calculate the numbers of the terms in the each sum in (801),

N!p!(q!)pn0,N!p!(q!)p(p2)n2,N!p!(q!)p(p3)n3,,N!p!(q!)pnp,\displaystyle\frac{N!}{p!(q!)^{p}}n_{0},\frac{N!}{p!(q!)^{p}}{p\choose 2}n_{2},\frac{N!}{p!(q!)^{p}}{p\choose 3}n_{3},\dots,\frac{N!}{p!(q!)^{p}}n_{p}, (807)

and note that the first one is different to the others. And about the last one the dominant term of the subtrahend should be at the order (Nq)!(N-q)!, therefore the subtrahend is almost at the order N(Nq)!N(N-q)!. Actually this number is far larger than the accurate one, since in (805) there’s a factor (1)rs(-1)^{r-s} then the accurate number is approximately at (Nq)!(N-q)!.

Here we also have two conditions similar to (754),(756), and first we consider the former one. As in this section we assume the mean of the JAiJ_{A_{i}} is zero, therefore we have

z2=Asgn(A)2JA12JAp2=N!p!(q!)pJ¯2p\displaystyle\langle z^{2}\rangle=\left\langle\sum_{A}\text{sgn}(A)^{2}J_{A_{1}}^{2}\dots J_{A_{p}}^{2}\right\rangle=\frac{N!}{p!(q!)^{p}}\bar{J}^{2p} (808)

where

JAi2=J¯2.\displaystyle\langle J_{A_{i}}^{2}\rangle=\bar{J}^{2}. (809)

For the error we evaluate it as

Error2\displaystyle\langle\text{Error}^{2}\rangle =(z2Φ)2z22,\displaystyle=\langle\left(z^{2}-\Phi\right)^{2}\rangle-\langle z^{2}\rangle^{2}, (810)
=s=0,2p(i1iq+s(zs2Φs))2z22,\displaystyle=\left\langle\sum_{s=0,2}^{p}\left(\sum_{i_{1}\dots i_{q+s}}\left(z^{2}_{s}-\Phi_{s}\right)\right)^{2}\right\rangle-\langle z^{2}\rangle^{2}, (811)

where ss does not take value 11 for clarity and the cross terms of different sums are zero due to the zero mean of JAiJ_{A_{i}}. Then we find that the computation is similar to before that for s>0s>0 the nonzero contribution under the average is the diagonal terms in the square, while for s=0s=0 the whole terms contribute to the computation of the error. Which means to find out the dominant sum ss we should compare the square of the first term with the rest terms in (807), and note that the last one cancels out in (811) so that the non-trivial largest number should be the one labeled by p1p-1. And we give an upper bound for it

N!p!(q!)p(pp1)np1<p2N!p!(q!)p(Nq)!(p1)!(q!)p1.\displaystyle\frac{N!}{p!(q!)^{p}}{p\choose p-1}n_{p-1}<p^{2}\frac{N!}{p!(q!)^{p}}\frac{(N-q)!}{(p-1)!(q!)^{p-1}}. (812)

Note that the number of the identical terms in this sum may have an effect but we expect the upper bound is still valid. Then we can find that when NN,qq are sufficiently large it is subordinate to the square of the first term in (807). Therefore we write the dominant terms in (811)

N!p!(q!)pnp(N!p!(q!)p)2\displaystyle\frac{N!}{p!(q!)^{p}}n_{p}-\left(\frac{N!}{p!(q!)^{p}}\right)^{2} (813)

where the cross terms of the first part and the whole second part contribute. Recall the expression (805) the dominant terms in the above equation cancel out, which implies the proposal holds.

Then we consider the second condition (756) that for a single variable we can view JAi2ΦiJ_{A_{i}}^{2}-\Phi_{i} as a constant. The expression (811) is still valid so that the thing we need to consider is to find the correction to (813). As an example we consider the behavior of the second sum labeled by s=2s=2, the diagonal contribution after the subtraction should be

αJAp1JApJBp1JBp,\displaystyle\alpha\sum J_{A_{p-1}}J_{A_{p}}J_{B_{p-1}}J_{B_{p}}, (814)

where α\alpha is determined by the quotient of the two numbers of the two sum

α=N!p!(q!)p(p2)n2/((N2q)(2q)!2!(q!)2n2).\displaystyle\alpha=\frac{N!}{p!(q!)^{p}}{p\choose 2}n_{2}/\left({N\choose 2q}\frac{(2q)!}{2!(q!)^{2}}n_{2}\right). (815)

Therefore after the square the total number becomes

(N!p!(q!)p(p2)n2)2/((N2q)(2q)!2!(q!)2n2),\displaystyle\left(\frac{N!}{p!(q!)^{p}}{p\choose 2}n_{2}\right)^{2}/\left({N\choose 2q}\frac{(2q)!}{2!(q!)^{2}}n_{2}\right), (816)

comparing to the two terms in (813) the above term is subordinate. Similarly we can find that after the subtraction the largest number in the sums s>0s>0 occurs in the case s=p1s=p-1, which gives

(N!p!(q!)p(pp1)np1)2/((NNq)(Nq)!(p1)!(q!)p1np1).\displaystyle\left(\frac{N!}{p!(q!)^{p}}{p\choose p-1}n_{p-1}\right)^{2}/\left({N\choose N-q}\frac{(N-q)!}{(p-1)!(q!)^{p-1}}n_{p-1}\right). (817)

The above contribution is still subordinate to the two terms in (813), which means the proposal is proper in this case.

Thus for the SYK model we can conclude that when the mean is zero the proposal works well no matter what the distribution is.

SYK with one time point with μ0\mu\neq 0

Like what we have discussed before the case with non-zero mean is different to last section. The whole computation is similar to the case with μ=0\mu=0 except the mean terms, we can take relevant expressions from the previous sections. We consider the case with general N/qN/q,

z=Asgn(A)JA1JAp,\displaystyle z=\sum_{A}\text{sgn}(A)J_{A_{1}}\dots J_{A_{p}}, (818)

the main task here is find the dominant terms. The sign functions may have some effect which makes s=q1s=q-1 not dominant, such as when N/qN/q=2 we have

Asgn(A)JA1JA2=Asgn(A)JAi2=(N/21[N/4])JAi2,\displaystyle\left\langle\sum_{A}\text{sgn}(A)J_{A_{1}}J_{A_{2}}\right\rangle=\sum_{A}\text{sgn}(A)\langle J_{A_{i}}\rangle^{2}={N/2-1\choose[N/4]}\langle J_{A_{i}}\rangle^{2}, (819)

where the square bracket means the integer part. Actually here N/4N/4 is q/2q/2, in the later computation we always have even qq therefore we’ll omit the square bracket.

To derive this parameter explicitly we have two different ways, one is to find the sequence functions over different NN which can be solved by Mathematica. Another is to consider different combinations, we can define the positive and the negative combinations then their difference gives the parameter. Given NN different numbers we can have a permutation group which can be divided into the even and odd parts, and the two parts contain the same number of elements. If the indices of JJ have no order then the numbers of the even and odd parts are equal, so that the parameter is zero. But in the SYK model the indices are listed in a particular order, it makes the numbers of the two parts different.

To proceed the computation we define the sign of a qq-sequence,

sgn(JAi)=(1)ai1++aiq,q=0mod(4)\displaystyle\text{sgn}(J_{A_{i}})=(-1)^{a_{i_{1}}+\dots+a_{i_{q}}},\quad q=0\text{mod}(4) (820)

where when q=2mod(4)q=2\text{mod}(4) the sign will be opposite. Note that here we take the aia_{i} from the indices of JAiJ_{A_{i}}, but what in the superscript should be the initial sites of the indices. When the indices take the values 1,,N1,\dots,N, the values themselves naturally label their initial sites. But when the indices take arbitrary numbers or symbols, we can define an initial order and the sites with any permutation. After the assignment complete we can take the definition (820) to define the sign of any given qq-sequence.

Then for a combination AA containing rr negative q-sequences we have the argument

sgn(A)=1,if i=1rsi=0mod(2),\displaystyle\text{sgn}(A)=1,\quad\text{if }\sum_{i=1}^{r}s_{i}=0\text{mod}(2), (821)

where sis_{i} is the site of the ii-th negative qq-sequence. We give the condition above as the negative qq-sequences seems appearing in pairs. For N/q=2N/q=2 the positive combination AA contains two positive q-sequences, so our mission is to find the first sequence with positive sign. As the combination has an order the first index in JA1J_{A_{1}} is 1, we only need to consider the others. The positive qq-sequence has the number

(N/211)(N/2q2)+(N/213)(N/2q4)++(N/21q1)(N/20),\displaystyle{N/2-1\choose 1}{N/2\choose q-2}+{N/2-1\choose 3}{N/2\choose q-4}+\dots+{N/2-1\choose q-1}{N/2\choose 0}, (822)

where q=N/2q=N/2 the first chooses odd number of odd integers while the second chooses even number of even integers. Then the negative qq-sequence has

(N/210)(N/2q1)+(N/212)(N/2q3)++(N/21q2)(N/21),\displaystyle{N/2-1\choose 0}{N/2\choose q-1}+{N/2-1\choose 2}{N/2\choose q-3}+\dots+{N/2-1\choose q-2}{N/2\choose 1}, (823)

the difference of the two numbers will give the parameter

d(p,q)=i=0q1(1)i+1(pq/21i)(pq/2q1i),\displaystyle d(p,q)=\sum_{i=0}^{q-1}(-1)^{i+1}{pq/2-1\choose i}{pq/2\choose q-1-i}, (824)

which coincides with (819) when N/q=2N/q=2. For general p=N/qp=N/q we can also find an expression for finding the first positive qq-sequence

d(p,q)=1p1(pq/21q/2),\displaystyle d(p,q)=\frac{1}{p-1}{pq/2-1\choose q/2}, (825)

where qq is even and it can be verified by numerics.

And note that for different qq’s there may be a difference with the minus sign, which is explained in (820).

For general N/qN/q we can try to compute γ\gamma in the expression below by the recursion

Asgn(A)JA1JAp=Asgn(A)JAip=γJAip.\displaystyle\left\langle\sum_{A}\text{sgn}(A)J_{A_{1}}\dots J_{A_{p}}\right\rangle=\sum_{A}\text{sgn}(A)\langle J_{A_{i}}\rangle^{p}=\gamma\langle J_{A_{i}}\rangle^{p}. (826)

In the previous computation we have calculated the case with p=2p=2, we only need to derive the p+1p+1 case from the pp one. Given NN numbers we can give the difference between the positive combinations and the negative ones which is denoted as D(p,q)D(p,q) and we have

D(2,q)=d(2,q),D(p+1)=f(D(p,q)).\displaystyle D(2,q)=d(2,q),\quad D(p+1)=f(D(p,q)). (827)

where ff is an unknown function. When we add extra qq numbers to the NN case which is already known, in the new permutation we can fix one number in the extra qq ones at the p+1p+1 site in case of repetition. For simplicity we let the additional qq numbers be 1,,q1,\dots,q and put it at the first site, and we always fix the number 11 in this site.

We determine the first site through the procedure in the case of p=2p=2, which can be divided into the positive and negative parts. After choosing qq numbers from the N+qN+q ones, about the left NN numbers we assume we already know the difference between the positive combinations and the negative ones. To illustrate the product of the sequences, we define the positive (negative) sequences as P+P_{+} (PP_{-}) so that we have

P+P+=P+,PP=P+,P+P=P.\displaystyle P_{+}P_{+}=P_{+},\quad P_{-}P_{-}=P_{+},\quad P_{+}P_{-}=P_{-}. (828)

The product of two positive or two negative sequences gives a positive one while the product of one positive and one negative sequences gives a negative one. Then we can derive the sequences in the case N+qN+q as

P+N+q=P+qP+N+PqPN,PN+q=P+qPN+P+qPN,\displaystyle P_{+}^{N+q}=P^{q}_{+}P^{N}_{+}+P^{q}_{-}P^{N}_{-},\quad P_{-}^{N+q}=P^{q}_{+}P^{N}_{-}+P^{q}_{+}P^{N}_{-}, (829)

therefore the difference of the positive and negative sequences can be illustrated as

P+N+qPN+q=(P+qPq)(P+NPN).\displaystyle P_{+}^{N+q}-P_{-}^{N+q}=(P_{+}^{q}-P_{-}^{q})(P_{+}^{N}-P_{-}^{N}). (830)

Which means the recursion can be expressed as

D(p,q)=d(p,q)D(p1,q)=s=2pd(s,q),\displaystyle D(p,q)=d(p,q)D(p-1,q)=\prod_{s=2}^{p}d(s,q), (831)

inserting the equation (825) it becomes

D(p,q)=(pq/21)!(p1)!(q/21)!((q/2)!)p1=(pq/2)!p!((q/2)!)p.\displaystyle D(p,q)=\frac{(pq/2-1)!}{(p-1)!(q/2-1)!((q/2)!)^{p-1}}=\frac{(pq/2)!}{p!((q/2)!)^{p}}. (832)

Then we can try to evaluate the cross terms in the square of the partition function (801). Consider the terms with psp-s squares in the sequence,

A,Bsgn(A)sgn(B)JA12JAps2JAps+1JApJBps+1JBp,\displaystyle\sum_{A,B}\text{sgn}(A)\text{sgn}(B)J_{A_{1}}^{2}\dots J_{A_{p-s}}^{2}J_{A_{p-s+1}}\dots J_{A_{p}}J_{B_{p-s+1}}\dots J_{B_{p}}, (833)

after the average over JJ the number in the sum becomes

ds=r=2s(prsr)(Npqrq)(pqrq)!(pr)!(q!)prD(r,q)2(1)sr+(ps)(s1)(1)s1(pq)!p!(q!)p,\displaystyle d_{s}=\sum_{r=2}^{s}{p-r\choose s-r}{N\choose pq-rq}\frac{(pq-rq)!}{(p-r)!(q!)^{p-r}}D(r,q)^{2}(-1)^{s-r}+{p\choose s}(s-1)(-1)^{s-1}\frac{(pq)!}{p!(q!)^{p}}, (834)

where the derivation is similar to the equation (805).

We can find an upper bound for the number (834)

(Npqsq)(pqsq)!(ps)!(q!)psD(s,q)2,\displaystyle{N\choose pq-sq}\frac{(pq-sq)!}{(p-s)!(q!)^{p-s}}D(s,q)^{2}, (835)

Then for s=2,,p1s=2,\dots,p-1 we can compare the above number to the first one in (807), from this we may find out the dominant one.

By a few numerical comparisons it seems that the dominant terms do not locate at a particular ss as the two parameters pp,qq vary. In this computation we can directly take the equation (834) rather than the one with approximation.

Appendix H Some details about the fermion integral

The relevant integral is

IfN\displaystyle I_{f}^{N} =\displaystyle= d2Nψexp(T[i[ΣLRψiLψiR+ΣLθiψiL+ΣRθiψiR])\displaystyle\int\text{d}^{2N}\psi\exp\left(T[\sum_{i}[\Sigma_{LR}\psi_{i}^{L}\psi_{i}^{R}+\Sigma_{L}\theta_{i}\psi_{i}^{L}+\Sigma_{R}\theta_{i}\psi_{i}^{R}]\right)\, (836)
=\displaystyle= (dψLdψRexp(TΣLRψLψR)exp(TΣLθψL)exp(TΣRθψR))N\displaystyle\left(\int\text{d}\psi^{L}\text{d}\psi^{R}\exp(T\Sigma_{LR}\psi^{L}\psi^{R})\exp(T\Sigma_{L}\theta\psi^{L})\exp(T\Sigma_{R}\theta\psi^{R})\right)^{N} (837)

Before evaluating this integral we need some specifications. Here ψL(R)\psi^{L(R)} are not Grassmann numbers but can be transferred into Dirac fermions as

c=ψL+iψR2,c=ψLiψR2,ψL=c+c,ψR=i(cc)\displaystyle c=\frac{\psi^{L}+\text{i}\psi^{R}}{2},\quad c^{\dagger}=\frac{\psi^{L}-\text{i}\psi^{R}}{2},\quad\psi_{L}=c+c^{\dagger},\quad\psi_{R}=\text{i}(c^{\dagger}-c)\, (838)
{c,c}=1,{c,c}={c,c}=0.\displaystyle\{c,c^{\dagger}\}=1,\quad\{c,c\}=\{c^{\dagger},c^{\dagger}\}=0\,. (839)

Then the integral IfI_{f} becomes

If\displaystyle I_{f} =\displaystyle= 2dcdcea1(cccc)+ac+a+c,a1=iTΣLR,a±=T(ΣL±iΣR)θ,\displaystyle 2\int\text{d}c\text{d}c^{\dagger}e^{a_{1}(cc^{\dagger}-c^{\dagger}c)+a_{-}c+a_{+}c^{\dagger}},\quad a_{1}=\text{i}T\Sigma_{LR},\quad a_{\pm}=T(\Sigma_{L}\pm\text{i}\Sigma_{R})\theta, (840)
=\displaystyle= 4cosh(a12+a+a)=4cosh(TΣL2+ΣR2ΣLR2)\displaystyle 4\cosh(\sqrt{a_{1}^{2}+a_{+}a_{-}})=4\cosh(T\sqrt{\Sigma_{L}^{2}+\Sigma_{R}^{2}-\Sigma_{LR}^{2}}) (841)

Next let us use the same method to compute (382):

I4N\displaystyle I_{4}^{N} =\displaystyle= d4Nψexp(Σabψiaψib)\displaystyle\int\text{d}^{4N}\psi\exp\left(\Sigma_{ab}\psi_{i}^{a}\psi_{i}^{b}\right) (843)
=\displaystyle= {d4ψexp(12Σ12ψ1ψ2+Σ13ψ1ψ3+Σ14ψ1ψ4\displaystyle\{\int\text{d}^{4}\psi\exp\left(\frac{1}{2}\Sigma_{12}\psi^{1}\psi^{2}+\Sigma_{13}\psi^{1}\psi^{3}+\Sigma_{14}\psi^{1}\psi^{4}\right.
Σ23ψ2ψ3+Σ24ψ2ψ4+Σ34ψ3ψ4)}N.\displaystyle\quad\left.\Sigma_{23}\psi^{2}\psi^{3}+\Sigma_{24}\psi^{2}\psi^{4}+\Sigma_{34}\psi^{3}\psi^{4}\right)\}^{N}.

Generally to compute I4I_{4} let us introduce two Dirac fermions

ψ1=h(c1+c1),ψ2=ih(c1c1),ψ3=h(c2+c2),ψ4=ih(c2c2),\displaystyle\psi_{1}=\sqrt{h}(c_{1}+c_{1}^{\dagger}),\quad\psi_{2}=\text{i}\sqrt{h}(c_{1}-c_{1}^{\dagger}),\quad\psi_{3}=\sqrt{h}(c_{2}+c_{2}^{\dagger}),\quad\psi_{4}=\text{i}\sqrt{h}(c_{2}-c_{2}^{\dagger}), (844)

such that Σabψiaψib\Sigma_{ab}\psi_{i}^{a}\psi_{i}^{b} can be written as a 4×44\times 4 matrix HH with non-vanishing elements

H11=ih(Σ12+Σ34),H14=h(Σ13Σ24)ih(Σ14+Σ23)\displaystyle H_{11}=-\text{i}h(\Sigma_{12}+\Sigma_{34}),\quad H_{14}=-h(\Sigma_{13}-\Sigma_{24})-\text{i}h(\Sigma_{14}+\Sigma_{23}) (845)
H22=ih(Σ34Σ12),H23=h(Σ13+Σ24)+ih(Σ14Σ23)\displaystyle H_{22}=\text{i}h(\Sigma_{34}-\Sigma_{12}),\quad H_{23}=-h(\Sigma_{13}+\Sigma_{24})+\text{i}h(\Sigma_{14}-\Sigma_{23}) (846)
H32=h(Σ13+Σ24)+ih(Σ14Σ23),H33=ih(Σ34Σ12)\displaystyle H_{32}=h(\Sigma_{13}+\Sigma_{24})+\text{i}h(\Sigma_{14}-\Sigma_{23}),\quad H_{33}=-\text{i}h(\Sigma_{34}-\Sigma_{12}) (847)
H41=h(Σ13Σ24)ih(Σ14+Σ23),H44=ih(Σ12+Σ34).\displaystyle H_{41}=h(\Sigma_{13}-\Sigma_{24})-\text{i}h(\Sigma_{14}+\Sigma_{23}),\quad H_{44}=\text{i}h(\Sigma_{12}+\Sigma_{34}). (848)

By diagonalizing this matrix we can compute I4I_{4} as the trace of the exponential of the matrix

I4=2(cos(h(Σ14Σ23)2+(Σ13+Σ24)2+(Σ12Σ34)2)\displaystyle I_{4}=2\left(\cos\left(h\sqrt{(\Sigma_{14}-\Sigma_{23})^{2}+(\Sigma_{13}+\Sigma_{24})^{2}+(\Sigma_{12}-\Sigma_{34})^{2}}\right)\right.
+cos(h(Σ14+Σ23)2+(Σ13Σ24)2+(Σ12+Σ34)2)).\displaystyle\qquad\left.+\cos\left(h\sqrt{(\Sigma_{14}+\Sigma_{23})^{2}+(\Sigma_{13}-\Sigma_{24})^{2}+(\Sigma_{12}+\Sigma_{34})^{2}}\right)\right). (849)

And note that the change of the variable from ψ\psi to cc,cc^{\dagger} will introduce an additional coefficient in the integral

(2h)2,\displaystyle(2h)^{2}, (850)

the final result is the product of the two terms.

References

  • (1) J. M. Maldacena, “The Large N limit of superconformal field theories and supergravity,” Adv. Theor. Math. Phys. 2, 231-252 (1998) doi:10.1023/A:1026654312961 [arXiv:hep-th/9711200 [hep-th]].
  • (2) E. Witten, “Anti-de Sitter space and holography,” Adv. Theor. Math. Phys. 2, 253-291 (1998) doi:10.4310/ATMP.1998.v2.n2.a2 [arXiv:hep-th/9802150 [hep-th]].
  • (3) S. S. Gubser, I. R. Klebanov and A. M. Polyakov, “Gauge theory correlators from noncritical string theory,” Phys. Lett. B 428, 105-114 (1998) doi:10.1016/S0370-2693(98)00377-3 [arXiv:hep-th/9802109 [hep-th]].
  • (4) G. Penington, “Entanglement Wedge Reconstruction and the Information Paradox,” JHEP 09, 002 (2020) doi:10.1007/JHEP09(2020)002 [arXiv:1905.08255 [hep-th]].
  • (5) A. Almheiri, N. Engelhardt, D. Marolf and H. Maxfield, “The entropy of bulk quantum fields and the entanglement wedge of an evaporating black hole,” JHEP 12, 063 (2019) doi:10.1007/JHEP12(2019)063 [arXiv:1905.08762 [hep-th]].
  • (6) A. Almheiri, R. Mahajan, J. Maldacena and Y. Zhao, “The Page curve of Hawking radiation from semiclassical geometry,” JHEP 03, 149 (2020) doi:10.1007/JHEP03(2020)149 [arXiv:1908.10996 [hep-th]].
  • (7) G. Penington, S. H. Shenker, D. Stanford and Z. Yang, “Replica wormholes and the black hole interior,” [arXiv:1911.11977 [hep-th]].
  • (8) S. R. Coleman, “Black Holes as Red Herrings: Topological Fluctuations and the Loss of Quantum Coherence,” Nucl. Phys. B 307, 867-882 (1988) doi:10.1016/0550-3213(88)90110-1
  • (9) S. B. Giddings and A. Strominger, “Baby Universes, Third Quantization and the Cosmological Constant,” Nucl. Phys. B 321, 481-508 (1989) doi:10.1016/0550-3213(89)90353-2
  • (10) S. B. Giddings and A. Strominger, “Loss of Incoherence and Determination of Coupling Constants in Quantum Gravity,” Nucl. Phys. B 307, 854-866 (1988) doi:10.1016/0550-3213(88)90109-5
  • (11) J. Polchinski and A. Strominger, “A Possible resolution of the black hole information puzzle,” Phys. Rev. D 50, 7403-7409 (1994) doi:10.1103/PhysRevD.50.7403 [arXiv:hep-th/9407008 [hep-th]].
  • (12) P. Saad, S. H. Shenker and D. Stanford, “JT gravity as a matrix integral,” [arXiv:1903.11115 [hep-th]].
  • (13) P. Saad, S. H. Shenker and D. Stanford, “A semiclassical ramp in SYK and in gravity,” [arXiv:1806.06840 [hep-th]].
  • (14) P. Saad, “Late Time Correlation Functions, Baby Universes, and ETH in JT Gravity,” [arXiv:1910.10311 [hep-th]].
  • (15) C. Yan, “Crosscap Contribution to Late-Time Two-Point Correlators,” [arXiv:2203.14436 [hep-th]].
  • (16) R. Bousso, X. Dong, N. Engelhardt, T. Faulkner, T. Hartman, S. H. Shenker and D. Stanford, “Snowmass White Paper: Quantum Aspects of Black Holes and the Emergence of Spacetime,” [arXiv:2201.03096 [hep-th]].
  • (17) J. M. Maldacena and L. Maoz, “Wormholes in AdS,” JHEP 02, 053 (2004) doi:10.1088/1126-6708/2004/02/053 [arXiv:hep-th/0401024 [hep-th]].
  • (18) D. Stanford and E. Witten, “JT gravity and the ensembles of random matrix theory,” Adv. Theor. Math. Phys. 24, no.6, 1475-1680 (2020) doi:10.4310/ATMP.2020.v24.n6.a4 [arXiv:1907.03363 [hep-th]].
  • (19) L. V. Iliesiu, “On 2D gauge theories in Jackiw-Teitelboim gravity,” [arXiv:1909.05253 [hep-th]].
  • (20) D. Kapec, R. Mahajan and D. Stanford, “Matrix ensembles with global symmetries and ’t Hooft anomalies from 2d gauge theory,” JHEP 04, 186 (2020) doi:10.1007/JHEP04(2020)186 [arXiv:1912.12285 [hep-th]].
  • (21) H. Maxfield and G. J. Turiaci, “The path integral of 3D gravity near extremality; or, JT gravity with defects as a matrix integral,” JHEP 01, 118 (2021) doi:10.1007/JHEP01(2021)118 [arXiv:2006.11317 [hep-th]].
  • (22) E. Witten, “Matrix Models and Deformations of JT Gravity,” Proc. Roy. Soc. Lond. A 476, no.2244, 20200582 (2020) doi:10.1098/rspa.2020.0582 [arXiv:2006.13414 [hep-th]].
  • (23) E. Mefford and K. Suzuki, “Jackiw-Teitelboim quantum gravity with defects and the Aharonov-Bohm effect,” JHEP 05, 026 (2021) doi:10.1007/JHEP05(2021)026 [arXiv:2011.04695 [hep-th]].
  • (24) A. Altland and J. Sonner, “Late time physics of holographic quantum chaos,” SciPost Phys. 11, 034 (2021) doi:10.21468/SciPostPhys.11.2.034 [arXiv:2008.02271 [hep-th]].
  • (25) L. Eberhardt, “Summing over Geometries in String Theory,” JHEP 05, 233 (2021) doi:10.1007/JHEP05(2021)233 [arXiv:2102.12355 [hep-th]].
  • (26) D. Stanford, Z. Yang and S. Yao, “Subleading Weingartens,” JHEP 02, 200 (2022) doi:10.1007/JHEP02(2022)200 [arXiv:2107.10252 [hep-th]].
  • (27) I. Aref’eva and I. Volovich, “Gas of Baby Universes in JT Gravity and Matrix Models,” Symmetry 12, no.6, 975 (2020) doi:10.3390/sym12060975 [arXiv:1905.08207 [hep-th]].
  • (28) P. Betzios and O. Papadoulaki, “Liouville theory and Matrix models: A Wheeler DeWitt perspective,” JHEP 09, 125 (2020) doi:10.1007/JHEP09(2020)125 [arXiv:2004.00002 [hep-th]].
  • (29) D. Anninos and B. Mühlmann, “Notes on matrix models (matrix musings),” J. Stat. Mech. 2008, 083109 (2020) doi:10.1088/1742-5468/aba499 [arXiv:2004.01171 [hep-th]].
  • (30) M. Berkooz, V. Narovlansky and H. Raj, “Complex Sachdev-Ye-Kitaev model in the double scaling limit,” JHEP 02, 113 (2021) doi:10.1007/JHEP02(2021)113 [arXiv:2006.13983 [hep-th]].
  • (31) T. G. Mertens and G. J. Turiaci, “Liouville quantum gravity – holography, JT and matrices,” JHEP 01, 073 (2021) doi:10.1007/JHEP01(2021)073 [arXiv:2006.07072 [hep-th]].
  • (32) G. J. Turiaci, M. Usatyuk and W. W. Weng, “Dilaton-gravity, deformations of the minimal string, and matrix models,” doi:10.1088/1361-6382/ac25df [arXiv:2011.06038 [hep-th]].
  • (33) D. Anninos and B. Mühlmann, “Matrix integrals & finite holography,” JHEP 06, 120 (2021) doi:10.1007/JHEP06(2021)120 [arXiv:2012.05224 [hep-th]].
  • (34) P. Gao, D. L. Jafferis and D. K. Kolchmeyer, “An effective matrix model for dynamical end of the world branes in Jackiw-Teitelboim gravity,” [arXiv:2104.01184 [hep-th]].
  • (35) V. Godet and C. Marteau, “From black holes to baby universes in CGHS gravity,” JHEP 07, 138 (2021) doi:10.1007/JHEP07(2021)138 [arXiv:2103.13422 [hep-th]].
  • (36) C. V. Johnson, F. Rosso and A. Svesko, “Jackiw-Teitelboim supergravity as a double-cut matrix model,” Phys. Rev. D 104, no.8, 086019 (2021) doi:10.1103/PhysRevD.104.086019 [arXiv:2102.02227 [hep-th]].
  • (37) A. Blommaert and M. Usatyuk, “Microstructure in matrix elements,” [arXiv:2108.02210 [hep-th]].
  • (38) K. Okuyama and K. Sakai, “JT gravity, KdV equations and macroscopic loop operators,” JHEP 01, 156 (2020) doi:10.1007/JHEP01(2020)156 [arXiv:1911.01659 [hep-th]].
  • (39) S. Forste, H. Jockers, J. Kames-King and A. Kanargias, “Deformations of JT Gravity via Topological Gravity and Applications,” [arXiv:2107.02773 [hep-th]].
  • (40) A. Maloney and E. Witten, “Averaging over Narain moduli space,” JHEP 10, 187 (2020) doi:10.1007/JHEP10(2020)187 [arXiv:2006.04855 [hep-th]].
  • (41) N. Afkhami-Jeddi, H. Cohn, T. Hartman and A. Tajdini, “Free partition functions and an averaged holographic duality,” JHEP 01, 130 (2021) doi:10.1007/JHEP01(2021)130 [arXiv:2006.04839 [hep-th]].
  • (42) J. Cotler and K. Jensen, “AdS3 gravity and random CFT,” JHEP 04, 033 (2021) doi:10.1007/JHEP04(2021)033 [arXiv:2006.08648 [hep-th]].
  • (43) A. Pérez and R. Troncoso, “Gravitational dual of averaged free CFT’s over the Narain lattice,” JHEP 11, 015 (2020) doi:10.1007/JHEP11(2020)015 [arXiv:2006.08216 [hep-th]].
  • (44) N. Benjamin, C. A. Keller, H. Ooguri and I. G. Zadeh, “Narain to Narnia,” Commun. Math. Phys. 390, no.1, 425-470 (2022) doi:10.1007/s00220-021-04211-x [arXiv:2103.15826 [hep-th]].
  • (45) J. Cotler and K. Jensen, “AdS3 wormholes from a modular bootstrap,” JHEP 11, 058 (2020) doi:10.1007/JHEP11(2020)058 [arXiv:2007.15653 [hep-th]].
  • (46) M. Ashwinkumar, M. Dodelson, A. Kidambi, J. M. Leedom and M. Yamazaki, “Chern-Simons Invariants from Ensemble Averages,” doi:10.1007/JHEP08(2021)044 [arXiv:2104.14710 [hep-th]].
  • (47) N. Afkhami-Jeddi, A. Ashmore and C. Cordova, “Calabi-Yau CFTs and Random Matrices,” [arXiv:2107.11461 [hep-th]].
  • (48) S. Collier and A. Maloney, “Wormholes and Spectral Statistics in the Narain Ensemble,” [arXiv:2106.12760 [hep-th]].
  • (49) N. Benjamin, S. Collier, A. L. Fitzpatrick, A. Maloney and E. Perlmutter, “Harmonic analysis of 2d CFT partition functions,” doi:10.1007/JHEP09(2021)174 [arXiv:2107.10744 [hep-th]].
  • (50) J. Dong, T. Hartman and Y. Jiang, “Averaging over moduli in deformed WZW models,” doi:10.1007/JHEP09(2021)185 [arXiv:2105.12594 [hep-th]].
  • (51) A. Dymarsky and A. Shapere, “Comments on the holographic description of Narain theories,” JHEP 10, 197 (2021) doi:10.1007/JHEP10(2021)197 [arXiv:2012.15830 [hep-th]].
  • (52) V. Meruliya, S. Mukhi and P. Singh, “Poincaré Series, 3d Gravity and Averages of Rational CFT,” JHEP 04, 267 (2021) doi:10.1007/JHEP04(2021)267 [arXiv:2102.03136 [hep-th]].
  • (53) R. Bousso and E. Wildenhain, “Gravity/ensemble duality,” Phys. Rev. D 102, no.6, 066005 (2020) doi:10.1103/PhysRevD.102.066005 [arXiv:2006.16289 [hep-th]].
  • (54) O. Janssen, M. Mirbabayi and P. Zograf, “Gravity as an ensemble and the moment problem,” JHEP 06, 184 (2021) doi:10.1007/JHEP06(2021)184 [arXiv:2103.12078 [hep-th]].
  • (55) J. Cotler and K. Jensen, “Wormholes and black hole microstates in AdS/CFT,” JHEP 09, 001 (2021) doi:10.1007/JHEP09(2021)001 [arXiv:2104.00601 [hep-th]].
  • (56) D. Marolf and H. Maxfield, “Transcending the ensemble: baby universes, spacetime wormholes, and the order and disorder of black hole information,” JHEP 08, 044 (2020) doi:10.1007/JHEP08(2020)044 [arXiv:2002.08950 [hep-th]].
  • (57) V. Balasubramanian, A. Kar, S. F. Ross and T. Ugajin, “Spin structures and baby universes,” JHEP 09, 192 (2020) doi:10.1007/JHEP09(2020)192 [arXiv:2007.04333 [hep-th]].
  • (58) J. G. Gardiner and S. Megas, “2d TQFTs and baby universes,” JHEP 10, 052 (2021) doi:10.1007/JHEP10(2021)052 [arXiv:2011.06137 [hep-th]].
  • (59) A. Belin and J. de Boer, “Random statistics of OPE coefficients and Euclidean wormholes,” Class. Quant. Grav. 38, no.16, 164001 (2021) doi:10.1088/1361-6382/ac1082 [arXiv:2006.05499 [hep-th]].
  • (60) A. Belin, J. De Boer, P. Nayak and J. Sonner, “Charged Eigenstate Thermalization, Euclidean Wormholes and Global Symmetries in Quantum Gravity,” [arXiv:2012.07875 [hep-th]].
  • (61) A. Altland, D. Bagrets, P. Nayak, J. Sonner and M. Vielma, “From operator statistics to wormholes,” Phys. Rev. Res. 3, no.3, 033259 (2021) doi:10.1103/PhysRevResearch.3.033259 [arXiv:2105.12129 [hep-th]].
  • (62) A. Belin, J. de Boer, P. Nayak and J. Sonner, “Generalized Spectral Form Factors and the Statistics of Heavy Operators,” [arXiv:2111.06373 [hep-th]].
  • (63) C. Peng, J. Tian and J. Yu, “Baby universes, ensemble averages and factorizations with matters,” [arXiv:2111.14856 [hep-th]].
  • (64) A. Banerjee and G. W. Moore, “Comments on Summing over bordisms in TQFT,” [arXiv:2201.00903 [hep-th]].
  • (65) J. J. Heckman, A. P. Turner and X. Yu, “Disorder averaging and its UV discontents,” Phys. Rev. D 105, no.8, 086021 (2022) doi:10.1103/PhysRevD.105.086021 [arXiv:2111.06404 [hep-th]].
  • (66) C. V. Johnson, “The Microstate Physics of JT Gravity and Supergravity,” [arXiv:2201.11942 [hep-th]].
  • (67) S. Collier and E. Perlmutter, “Harnessing S-Duality in 𝒩=4\mathcal{N}=4 SYM & Supergravity as SL(2,)SL(2,\mathbb{Z})-Averaged Strings,” [arXiv:2201.05093 [hep-th]].
  • (68) J. Chandra, S. Collier, T. Hartman and A. Maloney, “Semiclassical 3D gravity as an average of large-c CFTs,” [arXiv:2203.06511 [hep-th]].
  • (69) J. M. Schlenker and E. Witten, “No Ensemble Averaging Below the Black Hole Threshold,” [arXiv:2202.01372 [hep-th]].
  • (70) R. Jackiw, “Lower Dimensional Gravity,” Nucl. Phys. B 252, 343-356 (1985) doi:10.1016/0550-3213(85)90448-1
  • (71) C. Teitelboim, “Gravitation and Hamiltonian Structure in Two Space-Time Dimensions,” Phys. Lett. B 126, 41-45 (1983) doi:10.1016/0370-2693(83)90012-6
  • (72) S. Sachdev and J. Ye, “Gapless spin fluid ground state in a random, quantum Heisenberg magnet,” Phys. Rev. Lett. 70, 3339 (1993) doi:10.1103/PhysRevLett.70.3339 [arXiv:cond-mat/9212030 [cond-mat]].
  • (73) A. Kitaev, ”A simple model of quantum holography.” Talks at KITP http://online.kitp.ucsb.edu/online/entangled15/kitaev/ and http://online.kitp.ucsb.edu/online/entangled15/kitaev2
    A. Kitaev, ”Hidden correlations in the Hawking radiation and thermal noise.” Talk at KITP http://online.kitp.ucsb.edu/online/joint98/kitaev/
  • (74) D. Stanford, “More quantum noise from wormholes,” [arXiv:2008.08570 [hep-th]].
  • (75) A. Almheiri and H. W. Lin, “The Entanglement Wedge of Unknown Couplings,” [arXiv:2111.06298 [hep-th]].
  • (76) P. Saad, S. Shenker and S. Yao, “Comments on wormholes and factorization,” [arXiv:2107.13130 [hep-th]].
  • (77) P. Saad, S. H. Shenker, D. Stanford and S. Yao, “Wormholes without averaging,” [arXiv:2103.16754 [hep-th]].
  • (78) B. Mukhametzhanov, “Half-wormholes in SYK with one time point,” [arXiv:2105.08207 [hep-th]].
  • (79) A. M. García-García and V. Godet, “Half-wormholes in nearly AdS2 holography,” [arXiv:2107.07720 [hep-th]].
  • (80) S. Choudhury and K. Shirish, “Wormhole calculus without averaging from O(N)q1O(N)^{q-1} tensor model,” [arXiv:2106.14886 [hep-th]].
  • (81) B. Mukhametzhanov, “Factorization and complex couplings in SYK and in Matrix Models,” [arXiv:2110.06221 [hep-th]].
  • (82) K. Okuyama and K. Sakai, “FZZT branes in JT gravity and topological gravity,” JHEP 09, 191 (2021) doi:10.1007/JHEP09(2021)191 [arXiv:2108.03876 [hep-th]].
  • (83) K. Goto, Y. Kusuki, K. Tamaoka and T. Ugajin, “Product of random states and spatial (half-)wormholes,” JHEP 10, 205 (2021) doi:10.1007/JHEP10(2021)205 [arXiv:2108.08308 [hep-th]].
  • (84) A. Blommaert, L. V. Iliesiu and J. Kruthoff, “Gravity factorized,” [arXiv:2111.07863 [hep-th]].
  • (85) K. Goto, K. Suzuki and T. Ugajin, “Factorizing Wormholes in a Partially Disorder-Averaged SYK Model,” [arXiv:2111.11705 [hep-th]].
  • (86) A.I. Barvinok, New algorithms for linear k-matroid intersection and matroid k-parity problems, Mathematical Programming 69 (1995) 449A470.
  • (87) C. Peng, “Ensemble averages, Poisson processes, and microstates,” Phys. Rev. D 103, no.6, L061901 (2021) doi:10.1103/PhysRevD.103.L061901 [arXiv:2010.11192 [hep-th]].
  • (88) N. N. Bogolyubov, “On the theory of superfluidity,” J. Phys. (USSR) 11, 23-32 (1947) Valatin:1958ja
  • (89) J. G. Valatin, “Comments on the theory of superconductivity,” Nuovo Cim. 7, 843-857 (1958) doi:10.1007/BF02745589
  • (90) E. Witten, “Analytic Continuation Of Chern-Simons Theory,” AMS/IP Stud. Adv. Math. 50, 347-446 (2011) [arXiv:1001.2933 [hep-th]].
  • (91) A. Blommaert and J. Kruthoff, “Gravity without averaging,” [arXiv:2107.02178 [hep-th]].
  • (92) A. Blommaert, T. G. Mertens and H. Verschelde, “Eigenbranes in Jackiw-Teitelboim gravity,” JHEP 02, 168 (2021) doi:10.1007/JHEP02(2021)168 [arXiv:1911.11603 [hep-th]].
  • (93) A. Goel, L. V. Iliesiu, J. Kruthoff and Z. Yang, “Classifying boundary conditions in JT gravity: from energy-branes to α\alpha-branes,” JHEP 04, 069 (2021) doi:10.1007/JHEP04(2021)069 [arXiv:2010.12592 [hep-th]].
  • (94) A. Blommaert, L. V. Iliesiu and J. Kruthoff, “Alpha states demystified: Towards microscopic models of AdS2 holography,” [arXiv:2203.07384 [hep-th]].
  • (95) D. Das, S. Pal and A. Sarkar, “(Half) Wormholes under Irrelevant Deformation,” [arXiv:2203.14988 [hep-th]].