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

aainstitutetext: Center for Gravitational Physics, Department of Space Science, Beihang University, Beijing 100191, Chinabbinstitutetext: Peng Huanwu Collaborative Center for Research and Education, Beihang University, Beijing 100191, China

Probing Krylov Complexity in Scalar Field Theory with General Temperatures

Peng-Zhang He a,b    , Hai-Qing Zhang [email protected] [email protected]
Abstract

Krylov complexity characterizes the operator growth in the quantum many-body systems or quantum field theories. The existing literatures have studied the Krylov complexity in the low temperature limit in the quantum field theories. In this paper, we extend and systematically study the Krylov complexity and Krylov entropy in a scalar field theory with general temperatures. To this end, we propose a new method to calculate the Wightman power spectrum which allows us to compute the Lanczos coefficients and subsequently to study the Krylov complexity (entropy) in general temperatures. We find that the Lanczos coefficients and Krylov complexity (entropy) in the high temperature limit will behave somewhat differently from those studies in the low temperature limit. We give an explanation of why the Krylov complexity does not oscillate in the high-temperature region. Moreover, we uncover the transition temperature that separates the oscillating and monotonic increasing behavior of Krylov complexity.

1 Introduction

In recent years, complexity Nielsen:2006cea ; Jefferson:2017sdb ; Parker:2018yvk has become a crucial notion to understand the chaotic properties of a quantum system or spacetime PhysRevLett.116.191301 ; Susskind:2014rva . It provides a new insight and an analytical approach to probe the chaotic behaviors of a physical system. The complexity is originated as a means to quantify the difficulty of a quantum system which can transit from one state to another. With the deepening of the research, complexity has received wide applications in the fields such as quantum chaos, quantum information, quantum field theory, and holographic theory, etc. Aaronson:2016vto .

Recent studies on complexity has led to the emergence of Krylov complexity as an important concept in quantum physics. Initially, it was developed to explore the growth of operator in the Heisenberg picture and to distinguish between chaotic and integrable systems Parker:2018yvk . Usually, this operator is defined in a local Hilbert space called Krylov space, which is the span of nested commutators (see the following section 2). The Lanczos coefficients can be constructed according to the basis in the Krylov space. Consequently, Krylov complexity can be computed from the Schrödinger-type equation of the Lanczos coefficients. The authors Parker:2018yvk found that the time evolution of the Krylov complexity satisfy an exponential growth as time goes by, which indicates a signature of chaos in quantum many-body systems, such as the spin chains and Sachdev-Ye-Kitaev (SYK) model Sachdev:1992fk ; kitaev . At finite temperature, the authors further found that there is an upper bound for this exponential growth rate of the Krylov complexity.

In nowadays, Krylov complexity has received increasingly widespread applications in many other quantum systems Bhattacharya:2022gbz ; Bhattacharjee:2022lzy ; Liu:2022god ; Hashimoto:2023swv ; Adhikari:2022whf ; Avdoshkin:2022xuw . Specifically, the study of Krylov complexity has unveiled some astonishing phenomena. In chaotic systems, its growth initially exhibits an exponential trend, gradually transits to linear, and eventually reaches a plateau, indicating a saturation, as the system size is finite Parker:2018yvk ; Rabinovici:2020ryf . However, even in integrable quantum systems, such as free scalar field theory Camargo:2022rnt or conformal field theory Dymarsky:2021bjq , the Lanczos coefficients display linear growth, while the Krylov complexity exhibits an exponential growth. The paper Bhattacharjee:2022vlt is the first to present non-trivial evidence that integrable systems, under specific conditions, can exhibit the linear behavior of the Lanczos coefficients. This implies that there are still some subtle aspects of Krylov complexity that we have not yet fully understood. So far, the study of Krylov complexity has permeated multiple fields such as free and interacting field theory, random matrix theory, and open quantum systems etc Bhattacharjee:2022qjw ; Balasubramanian:2023kwd ; Dymarsky:2019elm ; Dymarsky:2021bjq ; Camargo:2022rnt ; Vasli:2023syq ; PhysRevD.106.126022 ; Iizuka:2023pov ; Iizuka:2023fba ; Erdmenger:2023wjg ; Bhattacharjee:2022ave ; Bhattacharjee:2023uwx ; Bhattacharya:2023zqt ; Bhattacharyya:2023grv ; Malvimat:2024vhr ; Caputa:2024vrn ; Afrasiar:2022efk ; Tan:2024kqb ; Li:2024iji ; Li:2024kfm ; Vardian:2024fsp ; Camargo:2023eev ; Huh:2023jxt ; Camargo:2024deu . Moreover, the concept of Krylov complexity has been successfully applied to various specific problems, including the SYK model Rabinovici:2020ryf ; He:2022ryk , generalized coherent states Patramanis:2021lkx ; Caputa:2021sib , Ising and Heisenberg models Cao:2020zls ; Trigueros:2021rwj ; Heveling:2022hth , topological phases of matter Caputa:2022eye . Interested readers can refer to the recent seminal review Nandy:2024htc .

In this paper, we extend the study of Krylov complexity in five dimensional free field theory from Camargo:2022rnt to general temperatures. In Camargo:2022rnt the authors only studied the case in the low temperature limit which satisfies βm1\beta m\gg 1,111β\beta is the inverse of the temperature while mm is the mass of the field. Their definitions are defined in the following context. which can simplify the Wightman power spectrum in this limit. However, in our paper we find a new method to compute the Wightman power spectrum with general βm\beta m, i.e., with general temperatures. As a result, the Lanczos coefficients can be subsequently computed and the Krylov complexity and Krylov entropy can also be obtained accordingly. We found that the Lanczos coefficients bnb_{n} exhibit the ‘staggering’ behavior, which separates the odd and even nn into two families. For large nn one can linearly fit the Lanczos coefficients as bnαn+γb_{n}\sim\alpha n+\gamma. The linear coefficient α\alpha is approximately identical to π/β\pi/\beta. The differences Δγ\Delta\gamma between γ\gamma of odd nn and even nn is linearly proportional to the mass of the field as βm1\beta m\gg 1. These behaviors are consistent with previous studies in Camargo:2022rnt . However, in the high temperature limit βm1\beta m\ll 1, we observe that the linear behavior of the difference Δγ\Delta\gamma to the mass will break down, which may come from the effect of the high temperature. We also found that in the high temperature limit, the Krylov complexity and Krylov entropy behaves like those in conformal field theories Dymarsky:2021bjq . The exponential growth rate of the Krylov complexity in the high temperature limit is roughly 2π/β2\pi/\beta, however, in the low temperature limit the rate is smaller than 2π/β2\pi/\beta. Therefore, totally for the general temperatures, the exponential growth rate of the Krylov complexity has an upper bound 2π/β2\pi/\beta which is consistent with previous reports. For the Krylov entropy, we found that in the late time it is linearly proportional to time. This is consistent with the claims that Krylov entropy is proportional to logarithmic of the Krylov complexity. Interestingly, we found that the Krylov complexity in high-temperature limit does not oscillate, which has sharp comparison to those in low-temperature limit. We postulate that the different behaviors of Krylov complexity in high and low temperature limit are due to the different behavior of bnb_{n} as well as the auto-correlation function φ0(t)\varphi_{0}(t). Moreover, we numerically uncover the transition temperature at around βm=10\beta m=10 which separates the oscillations and monotonic increasing of the Krylov complexity.

This paper is arranged as follows: In section 2 we will give a brief review of Krylov space and Krylov complexity. In Section 2.3, we will briefly review some of the results from Camargo:2022rnt and provide a physical explanation for the oscillation of the Krylov complexity in the low-temperature region; In section 3, we will first introduce a new method to compute the Wightman power spectrum in general temperatures, and then compare the Krylov complexity and Krylov entropy for general temperatures in the free scalar field theory; Then we draw our conclusions in section 4. In the appendix A, we will discuss the Krylov complexity and Krylov entropy in the spontaneous symmetry breaking phase. Appendix B will provide some details about the spontaneous symmetry breaking in the real scalar field theory. These two appendices can be regarded as an application of Krylov complexity in high-temperature from section 3.3.

2 Overview on Krylov Space and Krylov Complexity

In this section, we will give a brief review on the basics of Krylov space and Krylov complexity. Interested readers can refer to Refs. Parker:2018yvk ; Camargo:2022rnt ; Nandy:2024htc for more details.

2.1 Lanczos algorithm

In quantum mechanics, time evolution of an operator 𝒪\mathcal{O} is determined by the Heisenberg equation

t𝒪(t)=i[H,𝒪(t)],\partial_{t}\mathcal{O}(t)=i[H,\mathcal{O}(t)], (1)

where HH is the Hamiltonian of the system. The solution of the above equation is

𝒪(t)=eiHt𝒪(0)eiHt.\mathcal{O}(t)=e^{iHt}\mathcal{O}(0)e^{-iHt}. (2)

where 𝒪(0)\mathcal{O}(0) is the value of the operator at time t=0t=0. From the well-known Baker-Campbell-Hausdorff formula wachter2011relativistic

eABeA=n=01n![A(n),B],e^{A}Be^{-A}=\sum_{n=0}^{\infty}\frac{1}{n!}[A^{(n)},B], (3)

where [A(n),B][A,,[A,[An,B]][A^{(n)},B]\equiv[\underbrace{A,\cdots,[A,[A}_{n},B]], the Eq.(2) can be rewritten as

𝒪(t)=n=0(it)nn![H(n),𝒪]n=0(it)nn!𝒪~n,\mathcal{O}(t)=\sum_{n=0}^{\infty}\frac{(it)^{n}}{n!}[H^{(n)},\mathcal{O}]\equiv\sum_{n=0}^{\infty}\frac{(it)^{n}}{n!}\tilde{\mathcal{O}}_{n}, (4)

where we have defined 𝒪~n[H(n),𝒪]\tilde{\mathcal{O}}_{n}\equiv[H^{(n)},\mathcal{O}] and set 𝒪𝒪(0)\mathcal{O}\equiv\mathcal{O}(0). This equation describes how a “simple” operator may become increasingly “complex” as time evolves. Then one can introduce a super-operator called Liouvillian

:=[H,],\mathcal{L}:=[H,\cdot], (5)

which is a linear map in the space of operators such that

𝒪=[H,𝒪].\mathcal{L}\mathcal{O}=[H,\mathcal{O}]. (6)

Obviously we have

n𝒪=𝒪~n,\mathcal{L}^{n}\mathcal{O}=\tilde{\mathcal{O}}_{n}, (7)

and the Eq.(2) can be reformulated as

𝒪(t)=n=0(it)nn!n𝒪=eit𝒪.\mathcal{O}(t)=\sum_{n=0}^{\infty}\frac{(it)^{n}}{n!}\mathcal{L}^{n}\mathcal{O}=e^{it\mathcal{L}}\mathcal{O}. (8)

Compared to the Schrödinger’s picture of ordinary wave function, the above equation can be explained as the operator’s “wave function” expanded in some local basis of “states” 𝒪~n\tilde{\mathcal{O}}_{n} belonging to a local Hilbert space 𝒪\mathcal{H}_{\mathcal{O}}, known as Krylov space. Formally, Eq.(8) resembles the wave function in the Schrödinger picture, with \mathcal{L} playing a role as the Hamiltonian. In the following, we will use the symbol |A)\left|A\right) to represent a state in the Krylov space, i.e. |A)𝒪|A)\in\mathcal{H}_{\mathcal{O}}.

Since we are interested in the effects of finite temperature, we can define the inner product in Krylov space as the Wightman inner product

(A|B):=eβH/2AeβH/2Bβ1𝒵βtr(eβH/2AeβH/2B),\left(A|B\right):=\expectationvalue{e^{\beta H/2}A^{\dagger}e^{-\beta H/2}B}_{\beta}\equiv\frac{1}{\mathcal{Z}_{\beta}}\tr(e^{-\beta H/2}A^{\dagger}e^{-\beta H/2}B), (9)

where β\beta is the inverse of the temperature T=β1T=\beta^{-1} and 𝒵β=tr(eβH)\mathcal{Z}_{\beta}=\tr(e^{-\beta H}) is the thermal partition function. With the definition of the Wightman inner product, the Liouvillian \mathcal{L} is Hermitian, i.e, (A|B)=(A|B)(A|\mathcal{L}B)=(\mathcal{L}A|B). Note that Eq.(8) can be considered as the expansion of |𝒪(t))\left|\mathcal{O}(t)\right) in the basis {n|𝒪)}\{\mathcal{L}^{n}\left|\mathcal{O}\right)\}, which are not necessarily orthonormal. However, we can use the Gram-Schmidt orthogonalization procedure gram1883ueber ; schmidt1907theorie to obtain a set of orthonormal basis {|𝒪n)}\{\left|\mathcal{O}_{n}\right)\} called the Krylov basis Parker:2018yvk . Suppose the initially given operator 𝒪\mathcal{O} itself is a normalized state in the Krylov space. We can define

|𝒪0):=|𝒪).\left|\mathcal{O}_{0}\right):=\left|\mathcal{O}\right). (10)

Afterwards, we can continuously construct vectors that are orthonormal to the existing basis vectors to obtain the Krylov basis. 222Besides, we assume 𝒪\mathcal{O} is a Hermitian operator 𝒪=𝒪\mathcal{O}^{\dagger}=\mathcal{O}, so that it is an observable. Thus, |𝒪1)\left|\mathcal{O}_{1}\right) can be constructed as follows 333From the inner product Eq.(9) and Eq.(6), it is readily to get that (𝒪0||𝒪0)=0\left(\mathcal{O}_{0}\right|\mathcal{L}\left|\mathcal{O}_{0}\right)=0.

b1|𝒪1)=|𝒪0)|𝒪0)(𝒪0||𝒪0)=|𝒪0),b_{1}\left|\mathcal{O}_{1}\right)=\mathcal{L}\left|\mathcal{O}_{0}\right)-\left|\mathcal{O}_{0}\right)\left(\mathcal{O}_{0}\right|\mathcal{L}\left|\mathcal{O}_{0}\right)=\mathcal{L}\left|\mathcal{O}_{0}\right), (11)

where b1:=(𝒪0|𝒪0)1/2b_{1}:=(\mathcal{L}{\mathcal{O}_{0}}|\mathcal{L}{\mathcal{O}_{0}})^{1/2}. For n>1n>1, |𝒪n)\left|\mathcal{O}_{n}\right) can be constructed as

bn|𝒪n):=|An)=|𝒪n1)bn1|𝒪n2),b_{n}\left|\mathcal{O}_{n}\right):=\left|A_{n}\right)=\mathcal{L}\left|\mathcal{O}_{n-1}\right)-b_{n-1}\left|\mathcal{O}_{n-2}\right), (12)

where bn=(An|An)1/2b_{n}=(A_{n}|A_{n})^{1/2}. The sequence {bn}\{b_{n}\} is called the Lanczos coefficients. The above procedure is also known as Lanczos algorithm.

2.2 Krylov complexity

The information about the growth of the operator 𝒪(t)\mathcal{O}(t) is contained in the sequence Eq.(12). We can expand |𝒪(t))\left|\mathcal{O}(t)\right) in terms of Krylov basis {|𝒪n)}\{\left|\mathcal{O}_{n}\right)\} as

|𝒪(t))=n=0|𝒪n)(𝒪n|𝒪(t))n=0inφn(t)|𝒪n),\left|\mathcal{O}(t)\right)=\sum_{n=0}^{\infty}\left|\mathcal{O}_{n}\right)\left(\mathcal{O}_{n}\right|\left.\mathcal{O}(t)\right)\equiv\sum_{n=0}^{\infty}i^{n}\varphi_{n}(t)\left|\mathcal{O}_{n}\right), (13)

where φn(t)\varphi_{n}(t) are the probability amplitudes and satisfy

n=0|φn(t)|2=1.\sum_{n=0}^{\infty}\absolutevalue{\varphi_{n}(t)}^{2}=1. (14)

Combined with Eq.(1), we can obtain a discrete “Schrödinger” equation

tφn(t)=bnφn1(t)bn+1φn+1(t).\partial_{t}\varphi_{n}(t)=b_{n}\varphi_{n-1}(t)-b_{n+1}\varphi_{n+1}(t). (15)

According to Eq.(13), one can find that φn(0)=δn,0\varphi_{n}(0)=\delta_{n,0} and from the recursion Eq.(15) we can define φ1(t)0\varphi_{-1}(t)\equiv 0. The Krylov complexity of an operator 𝒪\mathcal{O} is defined as

K(t):=(𝒪(t)|n|𝒪(t))=n=0n|φn(t)|2.K(t):=\left(\mathcal{O}(t)|n|\mathcal{O}(t)\right)=\sum_{n=0}^{\infty}n\absolutevalue{\varphi_{n}(t)}^{2}. (16)

For convenience, the definition of Krylov complexity we are going to use will differ slightly from Eq.(16). In order to be consistent with the Krylov complexity in Dymarsky:2021bjq , we redefine it as

K(t):=1+n=0n|φn(t)|2.K(t):=1+\sum_{n=0}^{\infty}n\absolutevalue{\varphi_{n}(t)}^{2}. (17)

It is worth noting that Eq.(15) links the growth of the operator with the hopping problem on a one-dimensional chain, where the Lanczos coefficients bnb_{n} can be considered as the hopping amplitudes Parker:2018yvk . Therefore, the definition of the Krylov complexity indicates that K(t)K(t) represents the average position of the wave function on the chain. Hence, the growth of the operator implies an increase in the number of nn that contributes to K(t)K(t).

In the study of Krylov complexity, there exists a very important quantity, i.e., the autocorrelation function or thermal Wightman 2-point function,

C(t)\displaystyle C(t) :=φ0(t)(𝒪(t)𝒪(0))\displaystyle:=\varphi_{0}(t)\equiv(\mathcal{O}(t)\mid\mathcal{O}(0)) (18)
ei(tiβ/2)H𝒪(0)ei(tiβ/2)H𝒪(0)β\displaystyle\equiv\left\langle e^{i(t-i\beta/2)H}\mathcal{O}^{\dagger}(0)e^{-i(t-i\beta/2)H}\mathcal{O}(0)\right\rangle_{\beta}
=𝒪(tiβ/2)𝒪(0)β:=ΠW(t).\displaystyle=\left\langle\mathcal{O}^{\dagger}(t-i\beta/2)\mathcal{O}(0)\right\rangle_{\beta}:=\Pi^{W}(t).

It is not difficult to find that as long as the derivatives of the autocorrelation function at all orders are known, the Krylov complexity can also be determined. To this end, we can expand the autocorrelation function around t=0t=0,

ΠW(t)\displaystyle\Pi^{W}(t) =n=0(𝒪|(it)nn!n|𝒪)\displaystyle=\sum_{n=0}^{\infty}\left(\mathcal{O}\right|\frac{(-it)^{n}}{n!}\mathcal{L}^{n}\left|\mathcal{O}\right) (19)
=n=0μ2n(it)2n(2n)!,\displaystyle=\sum_{n=0}^{\infty}\mu_{2n}\frac{(it)^{2n}}{(2n)!},

in which we have used the facts that (𝒪|p|𝒪)=0(\mathcal{O}|\mathcal{L}^{p}|\mathcal{O})=0 with pp an odd number. Therefore, only terms with even powers of \mathcal{L} are left. In the above Eq.(19), {μ2n}\{\mu_{2n}\} is called the moments with the expression

μ2n:=(𝒪|2n|𝒪)=1i2nd2ndt2nΠW(t)|t=0.\mu_{2n}:=\left(\mathcal{O}\right|\mathcal{L}^{2n}\left|\mathcal{O}\right)=\left.\frac{1}{i^{2n}}\frac{d^{2n}}{dt^{2n}}\Pi^{W}(t)\right|_{t=0}. (20)

These moments can also be obtained from the Wightman power spectrum fW(ω)f^{W}(\omega), which is related to the Wightman 2-point function via a Fourier transformation

fW(ω)=𝑑teiωtΠW(t).f^{W}(\omega)=\int_{-\infty}^{\infty}dte^{i\omega t}\Pi^{W}(t). (21)

From Eqs.(20) and (21), the moments are related to fW(ω)f^{W}(\omega) via

μ2n=12π𝑑ωω2nfW(ω).\mu_{2n}=\frac{1}{2\pi}\int_{-\infty}^{\infty}d\omega\omega^{2n}f^{W}(\omega). (22)

In particular, there is a nonlinear recursive relationship between moments and Lanczos coefficients, b12nbn2=det(μi+j)0i,jnb_{1}^{2n}\cdots b_{n}^{2}=\det\left(\mu_{i+j}\right)_{0\leq i,j\leq n}, where μi+j\mu_{i+j} is a Hankel matrix built from the moments viswanath1994recursion . Alternatively, this expression can be reformulated in the following recursive relations,

bn=M2n(n),\displaystyle b_{n}=\sqrt{M_{2n}^{(n)}}, (23)

where,

M2l(j)=M2l(j1)bj12M2l2(j2)bj22,l=j,,n,\displaystyle M^{(j)}_{2l}=\frac{M^{(j-1)}_{2l}}{b_{j-1}^{2}}-\frac{M^{(j-2)}_{2l-2}}{b^{2}_{j-2}},\qquad l=j,\dots,n, (24)

with M2l(0)=μ2l,b1b0:=1,M2l(1)=0M^{(0)}_{2l}=\mu_{2l},b_{-1}\equiv b_{0}:=1,M^{(-1)}_{2l}=0.

2.3 Krylov complexity in scalar field theory with low temperatures βm1\beta m\gg 1

In this subsection, we will give a brief review on the Krylov complexity in low-temperature scalar field theory based on Camargo:2022rnt . Consider a real massive scalar field ϕ\phi in five-dimensional spacetime, we can write its Lagrangian in the following form

Lϕ=12(μϕμϕm2ϕ2),L_{\phi}=\frac{1}{2}(\partial_{\mu}\phi\partial^{\mu}\phi-m^{2}\phi^{2}), (25)

where μ=0,1,2,3,4\mu=0,1,2,3,4 and x0=iτx^{0}=-i\tau with τ\tau the Euclidean time. Then the Wightman power spectrum fW(ω)f^{W}(\omega) can be obtained at the finite temperature β1=T\beta^{-1}=T as,

fW(ω)=N~(m,β)(ω2m2)Θ(|ω|m)/sinh(β|ω|2),f^{W}(\omega)=\tilde{N}(m,\beta)(\omega^{2}-m^{2})\Theta(\absolutevalue{\omega}-m)\bigg{/}\sinh\left(\frac{\beta|\omega|}{2}\right), (26)

where the symbol Θ\Theta represents the Heaviside step function and N~(m,β)\tilde{N}(m,\beta) is the normalization factor satisfying

12π𝑑ωfW(ω)=1.\frac{1}{2\pi}\int_{-\infty}^{\infty}d\omega{f^{W}(\omega)}=1. (27)

In order to get the Krylov complexity, we need to calculate the Lanczos coefficients bnb_{n} by solving the discrete Schrödinger equation (15). In general, this problem is not easy to do. However, it is relatively simple at low temperatures Camargo:2022rnt . It is worth noting that in finite temperature field theories, low temperature means βm1\beta m\gg 1, i.e., temperature is much less than the mass of the field, rather than simply β1\beta\gg 1 Kapusta:2006pm . Therefore, even if β\beta is not very large, it may still be considered as low temperature if mm is sufficiently large. In the low temperature limit βm1\beta m\gg 1, we can rewrite the Wightman power spectrum Eq.(26) as

fW(ω)N(m,β)eβ|ω|/2(ω2m2)Θ(|ω|m),βm1,f^{W}(\omega)\approx N(m,\beta)e^{-\beta\absolutevalue{\omega}/2}(\omega^{2}-m^{2})\Theta(\absolutevalue{\omega}-m),\qquad\beta m\gg 1, (28)

in which we have absorbed the extra term 22 into the normalization factor N(m,β)N(m,\beta), i.e., N(m,β)=2N~(m,β)N(m,\beta)=2\tilde{N}(m,\beta). Using Eq.(27), N(m,β)N(m,\beta) can be determined analytically,

N(m,β)=πβ3eβm/216+8βm,βm1.N(m,\beta)=\frac{\pi\beta^{3}e^{\beta m/2}}{16+8\beta m},\qquad~{}\beta m\gg 1. (29)

Therefore, the moments μ2n\mu_{2n} in Eq.(22) can be directly computed as

μ2n=22eβm22+βm(2β)2n[4Γ~(3+2n,βm2)β2m2Γ~(1+2n,βm2)],\mu_{2n}=\frac{2^{-2}e^{\frac{\beta m}{2}}}{2+\beta m}\left(\frac{2}{\beta}\right)^{2n}\left[4\tilde{\Gamma}\left(3+2n,\frac{\beta m}{2}\right)-\beta^{2}m^{2}\tilde{\Gamma}\left(1+2n,\frac{\beta m}{2}\right)\right], (30)

where Γ~(n,z)\tilde{\Gamma}(n,z) is the incomplete Gamma function.

Following the non-linear recursive relation in Eqs.(23)-(24), the Lanczos coefficients {bn}\{b_{n}\} can be obtained from the Eq.(30). For example, we show the behavior of the Lanczos coefficients in Figure 1 with βm=50\beta m=50. We see that bnb_{n} exhibits staggering behavior, meaning that it is divided into two groups: one is with odd nn (blue dots) while the other is with even nn (brown dots). Each group is a monotonically increasing function with respect to nn. In Camargo:2022rnt the authors also observed that the separation Δbn:=|bnoddbneven|\Delta b_{n}:=|b_{n}^{\rm odd}-b_{n}^{\rm even}| is of the order of the mass mm.

Refer to caption
Figure 1: Lanczos coefficients bnb_{n} for βm=50\beta m=50. The blue dots are for odd nn while the brown dots are for even nn.

The Krylov complexity can be computed from the auto-correlation function φ0(t)\varphi_{0}(t) in Eq.(18) and the Lanczos coefficients by means of Eq.(15). Performing the Fourier transformation of fW(ω)f^{W}(\omega) one can yield the auto-correlation function as

φ0(t)\displaystyle\varphi_{0}(t) =12π𝑑ωeiωtfW(ω)\displaystyle=\frac{1}{2\pi}\int_{-\infty}^{\infty}d\omega e^{-i\omega t}f^{W}(\omega) (31)
=1π0𝑑ωcos(ωt)fW(ω)\displaystyle=\frac{1}{\pi}\int_{0}^{\infty}d\omega\cos(\omega t)f^{W}(\omega)
=β3((β3(2+βm)24βt216mt4)cos(mt)4t(β2(3+βm)+4(βm1)t2)sin(mt))(2+βm)(β2+4t2)3.\displaystyle=\frac{\beta^{3}\left((\beta^{3}(2+\beta m)-24\beta t^{2}-16mt^{4})\cos(mt)-4t(\beta^{2}(3+\beta m)+4(\beta m-1)t^{2})\sin(mt)\right)}{(2+\beta m)(\beta^{2}+4t^{2})^{3}}.

in which we have used the fact that φ0(t)\varphi_{0}(t) is a real function and fW(ω)f^{W}(\omega) is an even function of ω\omega. Afterwards, we can use the fourth Runge-Kutta method to numerically solve Eq.(15) and consequently obtain the Krylov complexity K(t)K(t) and the Krylov entropy SK(t)S_{K}(t), which is defined by Barbon:2019wsy

SK(t):=n=0|φn(t)|2log|φn(t)|2.S_{K}(t):=-\sum_{n=0}^{\infty}\absolutevalue{\varphi_{n}(t)}^{2}\log\absolutevalue{\varphi_{n}(t)}^{2}. (32)
Refer to caption
Refer to caption
Figure 2: Time evolution of Krylov complexity (left panel) and Krylov entropy (right panel) of the free scalar field theories for βm=50\beta m=50 in five dimensional spacetime. The vertical axis is in a logarithmic scale for K(t)K(t) in the left panel.

We show the time evolution of K(t)K(t) and SK(t)S_{K}(t) in Figure 2 as an example with βm=50\beta m=50. In the left panel of Figure 2, it is found that the Krylov complexity oscillates when time is short. As was explained in Camargo:2022rnt , these oscillations are due to the trigonometric functions in the auto-correlation function (31). The oscillatory behavior of φ0(t)\varphi_{0}(t) will subsequently pass on to φn(t)\varphi_{n}(t) according to the discrete Schrödinger equation (15). In the early time the period of the oscillations is roughly π/m\pi/m, which is consistent with the plots in the left panel of Figure 2 where the period is roughly 0.060.06. In the late time, the amplitudes of the oscillations becomes smaller, which is due to the cancellation of φn(t)\varphi_{n}(t) coming from various nn. However, in the early time only small number of nn will contribute to φn(t)\varphi_{n}(t). Thus, the oscillation in the early time is more significant than in the late time. Intuitively, it might seem difficult to comprehend this oscillatory behavior, as a ‘simple’ operator will evolve towards a more ‘complex’ one Parker:2018yvk . However, in fact, the operator becoming more ‘complex’ is directly related to K(t)K(t) which receives contributions from more number of nn. Therefore, K(t)K(t) as the position of the wave function on the one-dimensional chain, is not necessarily monotonically increasing. In the following subsection 3.3 we will explain in detail whether the position of the wave function increases monotonically or not is related to the behavior of the hopping amplitude bnb_{n} as well as the wave function φ0(t)\varphi_{0}(t).

The above calculations only applies to the low temperature case of βm1\beta m\gg 1 since Eq.(28) merely works for βm1\beta m\gg 1. Therefore, we cannot use it for more general temperatures. In the next section, we will provide detailed studies of Krylov complexity and Krylov entropy for more general situations of βm\beta m.

3 Krylov complexity and Krylov entropy with General Temperatures

In this section, we will relax the constraint on βm\beta m to study the Krylov complexity and Krylov entropy in a more general situation. The key idea is that we can rewrite the Wightman power spectrum fW(ω)f^{W}(\omega) (26) in a series of summation with general βm\beta m, which subsequently makes the computation of Krylov complexity and Krylov entropy feasible.

3.1 Truncations of the Wightman power spectrum

For general values of βm\beta m, the Wightman power spectrum (26) can be rewritten as an infinite series. For convenience, we consider only the part for ω>m\omega>m and the other part for ω>m-\omega>m can be obtained from the fact that Eq.(26) is an even function with respect to ω\omega. Therefore, for ω>m\omega>m we have

fW(ω)\displaystyle f^{W}(\omega) =N~(β,m)(ω2m2)/sinh(βω2)\displaystyle=\tilde{N}(\beta,m)(\omega^{2}-m^{2})/\sinh(\frac{\beta\omega}{2}) (33)
=2N~(β,m)(ω2m2)eβω21eβω\displaystyle=2\tilde{N}(\beta,m)(\omega^{2}-m^{2})\frac{e^{-\frac{\beta\omega}{2}}}{1-e^{-\beta\omega}}
=2N~(β,m)(ω2m2)eβω2k=0eβωk\displaystyle=2\tilde{N}(\beta,m)(\omega^{2}-m^{2})e^{-\frac{\beta\omega}{2}}\sum_{k=0}^{\infty}e^{-\beta\omega k}
=2k=0N~(β,m)(ω2m2)eβω(k+1/2),\displaystyle=2\sum_{k=0}^{\infty}\tilde{N}(\beta,m)(\omega^{2}-m^{2})e^{-\beta\omega(k+1/2)},

in which we have used the series expansion 11x=k=0xk\frac{1}{1-x}=\sum_{k=0}^{\infty}x^{k}. Let βkβ(2k+1)\beta_{k}\equiv\beta(2k+1) and absorb the factor 22 into N(β,m)N(\beta,m) we obtain

fW(ω)=N(β,m)k=0(ω2m2)eβkω/2,f^{W}(\omega)=N(\beta,m)\sum_{k=0}^{\infty}(\omega^{2}-m^{2})e^{-\beta_{k}\omega/2}, (34)

where N(β,m)N(\beta,m) can be calculated from the normalization conditions (27) and its exact form is

N(β,m)=πβ3e3βm22βmΦ(eβm,2,3/2)+2eβm(4βm+Φ(eβm,3,1/2)),N(\beta,m)=\frac{\pi\beta^{3}e^{\frac{3\beta m}{2}}}{2\beta m\Phi(e^{-\beta m},2,3/2)+2e^{\beta m}(4\beta m+\Phi(e^{-\beta m},3,1/2))}, (35)

in which Φ(z,s,a)k=0zk/(k+a)s\Phi(z,s,a)\equiv\sum_{k=0}^{\infty}z^{k}/(k+a)^{s} is the Lerch transcendent function lerch . Therefore, the series of the Wightman power spectrum in Eq.(34) can in principle work for general βm\beta m’s. It has much broader applications than that in Eq.(28) from Camargo:2022rnt , which can only apply in the low temperature limit.

In order to compute the moments μ2n\mu_{2n} (22) we first define

I(n)mωneβkω2𝑑ωI(n)\equiv\int_{m}^{\infty}\omega^{n}e^{-\frac{\beta_{k}\omega}{2}}d\omega (36)

which can be integrated by parts as,

I(n)=2βkeβkm2mn+2nβkI(n1).\displaystyle I(n)=\frac{2}{\beta_{k}}e^{-\frac{\beta_{k}m}{2}}m^{n}+\frac{2n}{\beta_{k}}I(n-1). (37)

In particular, for n=0n=0 we can get it directly from the integration Eq.(36),

I(0)=meβkω2𝑑ω=2βkeβkm2.\displaystyle I(0)=\int_{m}^{\infty}e^{-\frac{\beta_{k}\omega}{2}}d\omega=\frac{2}{\beta_{k}}e^{-\frac{\beta_{k}m}{2}}. (38)

These equations imply that

I(n)\displaystyle I(n) =(2βk)n+1Γ~(n+1,βkm2).\displaystyle=\left(\frac{2}{\beta_{k}}\right)^{n+1}\tilde{\Gamma}\left(n+1,\frac{\beta_{k}m}{2}\right). (39)

Therefore, from the definition of moments Eq.(22), we have

μ2n\displaystyle\mu_{2n} =ω2n2π[N(βm)k=0(ω2m2)eβkω/2]\displaystyle=\int_{-\infty}^{\infty}\frac{\omega^{2n}}{2\pi}\left[N(\beta m)\sum_{k=0}^{\infty}(\omega^{2}-m^{2})e^{-\beta_{k}\omega/2}\right] (40)
=N(β,m)πk=0[I(2(n+1))m2I(2n)]\displaystyle=\frac{N(\beta,m)}{\pi}\sum_{k=0}^{\infty}\left[I(2(n+1))-m^{2}I(2n)\right]
=N(β,m)k=022n+1βk(2n+3)π[4Γ~(2n+3,βkm2)βk2m2Γ~(2n+1,βkm2)].\displaystyle=N(\beta,m)\sum_{k=0}^{\infty}\frac{2^{2n+1}\beta_{k}^{-(2n+3)}}{\pi}\left[4\tilde{\Gamma}\left(2n+3,\frac{\beta_{k}m}{2}\right)-\beta_{k}^{2}m^{2}\tilde{\Gamma}\left(2n+1,\frac{\beta_{k}m}{2}\right)\right].

Unfortunately, this formula is too complex to be computed analytically. Therefore, we resort to numerical computations. To this end, we can truncate the summation of μ2n\mu_{2n} in Eq.(40) at a suitable kk to obtain approximate results. In order to ensure sufficient computational speed and accuracy, the truncation position of kk should not be too large or too small. We find that a suitable choice is to make the normalization factor of the truncated Wightman power spectrum close to the exact form in Eq.(35). In the remainder of this paper, we always take

fWN(β,m)k=0kmax(ω2m2)eβkω/2,f^{W}\approx N(\beta,m)\sum_{k=0}^{k_{\text{max}}}(\omega^{2}-m^{2})e^{-\beta_{k}\omega/2}, (41)

in which kmaxk_{\rm max} is a large cut-off of kk. We can see that the exact fW(ω)f^{W}(\omega) in Eq.(34) satisfies kmax=k_{\text{max}}=\infty, while the approximated fW(ω)f^{W}(\omega) in Eq.(28) in the low temperature limit satisfies kmax=0k_{\text{max}}=0. In our case we choose kmax=200k_{\text{max}}=200 in Eq.(41), because in most cases it is sufficient to obtain accurate numerical results. In Figure 3 we compare the exact Wightman power spectrum fWf^{W} to the approximated functions with β=1\beta=1 and m=0.01m=0.01, i.e. βm1\beta m\ll 1 in the limit of high temperature. In the left panel of Figure 3, we see that fW(ω)f^{W}(\omega) for kmax=0k_{\rm max}=0 (blue dash-dotted line) is only consistent with the exact form (black line with kmax=k_{\rm max}=\infty) in the regime of large frequency; However, they will deviate from each other as ω12\omega\lesssim 12, which means the formula in Eq.(28) is not accurate in the low frequency regime with high temperatures. On the contrary, still in the left panel of Figure 3, we can see that there is no distinguishable differences for the Wightman power spectrum fW(ω)f^{W}(\omega) between kmax=k_{\text{max}}=\infty (black line) and kmax=200k_{\text{max}}=200 (red dashed line) in the whole regime of the frequency, which verifies that our choice of kmax=200k_{\text{max}}=200 is sufficient for the accuracy.

Refer to caption
Refer to caption
Figure 3: Relation between the Wightman power spectrum fW(ω)f^{W}(\omega) against the frequency ω\omega with β=1\beta=1 and m=0.01m=0.01. (Left) The black line (kmax=k_{\text{max}}=\infty) corresponds to Eq.(34), while the blue dash-dotted line (kmax=0k_{\text{max}}=0) corresponds to Eq.(28). The two lines deviate from each other as ω12\omega\lesssim 12. The red dashed line (kmax=200k_{\text{max}}=200) corresponds to Eq.(41). It matches with the black line in the whole regime of the frequency, reflecting that taking kmax=200k_{\text{max}}=200 is accurate enough for the numerical calculations; (Right) The logarithmic plot of fW(ω)f^{W}(\omega) against ω\omega in the limit of large frequency. The three lines (black, red dashed and blue dash-dotted) for different kmaxk_{\rm max} collapse together and the green dashed line is the best fit of them. They share the same scaling law as fW(ω)1611.14×e0.4919ωf^{W}(\omega)\sim 1611.14\times e^{-0.4919\omega}.

In the right panel of Figure 3 we show the logarithmic relations between fW(ω)f^{W}(\omega) and ω\omega in the large frequency limit. We can find that in the large frequency limit, the Wightman power spectrum with various kmaxk_{\rm max}’s will overlap together and decay exponentially with the frequency,

fW(ω)eω/ω0,ω.\displaystyle f^{W}(\omega)\sim e^{-\omega/\omega_{0}},\qquad~{}~{}\omega\to\infty. (42)

The fitted line (green dashed line) in the right panel of Figure 3 shows that this decay rate is 1/ω00.49191/\omega_{0}\approx 0.4919 as ω\omega\to\infty. Therefore, we get ω02.0329\omega_{0}\approx 2.0329, which is consistent with discussions in Camargo:2022rnt that the leading exponential decay of the power spectrum is ω0=2/β\omega_{0}=2/\beta where we have set β=1\beta=1.

3.2 Lanczos coefficients with truncations

Before studying the Krylov complexity, let’s first look at the behaviors of the Lanczos coefficients bnb_{n} with the truncations kmax=200k_{\rm max}=200.

Refer to caption
Figure 4: Lanczos coefficients bnb_{n} for βm=10\beta m=10 (dots) and βm=50\beta m=50 (squares) with kmax=200k_{\text{max}}=200. They are grouped in two families with odd nn (blue) and even nn (brown).

In Figure 4, we present the behaviors of bnb_{n} for βm=10\beta m=10 (dots) and βm=50\beta m=50 (squares). These coefficients also exhibit clear behavior of “staggering”, which are grouped in two families with odd (in blue) and even nn (in brown). As nn is relatively large, bnb_{n} becomes linearly proportional to nn. Intuitively, we can find that the separation Δbn\Delta b_{n} between bnb_{n} for the odd nn and even nn with βm=50\beta m=50 are greater than those with βm=10\beta m=10. This phenomenon is consistent with the discussions in the preceding section that Δbn\Delta b_{n} is of the order of the mass mm Camargo:2022rnt . In order to analyze them quantitatively, we fit bnb_{n} linearly for odd nn and even nn separately similar to Camargo:2022rnt :

bnαoddn+γodd,odd n,\displaystyle b_{n}\sim\alpha_{\text{odd}}n+\gamma_{\text{odd}},\qquad\text{odd $n$}, (43)
bnαevenn+γeven,even n,\displaystyle b_{n}\sim\alpha_{\text{even}}n+\gamma_{\text{even}},\qquad\text{even $n$}, (44)

where αodd,αeven,γodd\alpha_{\text{odd}},\alpha_{\text{even}},\gamma_{\text{odd}} and γeven\gamma_{\text{even}} are constants which are independent of nn. Specifically, we perform the linear fitting for bnb_{n} in the range of n[100,450]n\in[100,450].

Refer to caption
Refer to caption
Figure 5: (Left) Mass-dependence of the slope αodd\alpha_{\text{odd}} and αeven\alpha_{\text{even}} in Eqs.(43)-(44); (Right) Mass-dependence of the difference γoddγeven\gamma_{\text{odd}}-\gamma_{\text{even}} in Eqs.(43)-(44). The inset plot exhibits the difference in high-temperatures regime.

In the left panel of Figure 5, we show the relation between βα\beta\alpha (α\alpha means αodd\alpha_{\rm odd} (blue) and αeven\alpha_{\rm even} (brown)) against βm\beta m. It is not difficult to observe that the two coefficients αodd\alpha_{\text{odd}} and αeven\alpha_{\text{even}} will overlap exactly, meaning that the slope in the fitting relations Eq.(43) and Eq.(44) are exactly identical. Moreover, we see that as βm0\beta m\to 0 the value of αodd\alpha_{\text{odd}} and αeven\alpha_{\text{even}} are approaching π/β\pi/\beta. However, as βm\beta m increases, αodd\alpha_{\text{odd}} and αeven\alpha_{\text{even}} slightly decrease from π/β\pi/\beta. This may come from the selected range of nn when fitting bnb_{n} as argued in Camargo:2022rnt . It is expected that larger range of nn will improve the value of α\alpha, and it will finally approach π/β\pi/\beta more closely Camargo:2022rnt . Therefore, in our case we can speculate that αodd=αevenπ/β\alpha_{\text{odd}}=\alpha_{\text{even}}\approx\pi/\beta. This analysis is also consistent with the universal operator growth hypothesis in Parker:2018yvk that the Lanczos coefficients bnb_{n} in a generic chaotic system will grow as fast as bnαn+γb_{n}\sim\alpha n+\gamma with α=πω0/2\alpha=\pi\omega_{0}/2 as nn\to\infty. In the fitting of the Wightman power spectrum Eq.(42) we already get ω02/β\omega_{0}\approx 2/\beta, therefore, we can get απ/β\alpha\approx\pi/\beta which is consistent with the left panel of Figure 5.

In the right panel of Figure 5, we show the relation between the difference β(γoddγeven)\beta(\gamma_{\text{odd}}-\gamma_{\text{even}}) with respect to βm\beta m. For larger mm, we can see that β(γoddγeven)\beta(\gamma_{\text{odd}}-\gamma_{\text{even}}) is linearly proportional to βm\beta m with the ratio β(γoddγeven)/(βm)1\beta(\gamma_{\text{odd}}-\gamma_{\text{even}})/(\beta m)\approx 1, which is consistent with the separations of bnb_{n} in Figure 4 that Δbn\Delta b_{n} is of the order of the mass mm. However, we find that this proportionality will break down for small βm[0,0.1]\beta m\in[0,0.1], as shown in the inset plot in the right panel of Figure 5. In the inset plot, the ratio in the range of small βm\beta m is roughly β(γoddγeven)/(βm)0.2\beta(\gamma_{\text{odd}}-\gamma_{\text{even}})/(\beta m)\approx 0.2 which is much smaller than the ratio 11 in the large βm\beta m regime. In the region of small βm1\beta m\ll 1, the system is in the high temperature limit, therefore, we speculate that the linear proportionality between γoddγeven\gamma_{\text{odd}}-\gamma_{\text{even}} and mm may not apply due to the high temperatures. As far as we know, this phenomenon is not explored previously.

Therefore, we see that high temperature limit will bring out different phenomena from the studies in Camargo:2022rnt with low temperatures. Next, we will first study the Krylov complexity and Krylov entropy in high-temperature limit, i.e., βm1\beta m\ll 1 and then discuss the general cases with general βm\beta m. After that, we apply the results to the finite-temperature Higgs scalar field theory.

3.3 High-temperature limit with βm1\beta m\ll 1

For convenience, in the numerics we always keep β1\beta\equiv 1 and study the variations of the Krylov complexity and Krylov entropy with respect to βm\beta m by adjusting the values of mm. The high temperature limit implies βm1\beta m\ll 1 Kapusta:2006pm , therefore, the approximation of Wightman power spectrum in Eq.(28) from Camargo:2022rnt is in invalid in this case. We need to use Eq.(41) to numerically study the Wightman power spectrum and then to study other quantities such as Krylov complexity and Krylov entropy.

Refer to caption
Refer to caption
Figure 6: (Left) Time evolution of Krylov complexity for βm1\beta m\ll 1 in the logarithmic scale. They overlap with each other and scale as e6.368t/βe^{6.368t/\beta} (as the black reference line indicates) at late time; (Right) The corresponding time evolution of Krylov entropy and the black line is the reference line with scalings ΔSK(t)/(Δt/β)=6.310\Delta S_{K}(t)/(\Delta t/\beta)=6.310.

As examples, we will take the cases of βm=0,0.001\beta m=0,0.001, 0.010.01 and 0.10.1. We have plotted the time evolution of Krylov complexity K(t)K(t) (in logarithmic scale) and Krylov entropy SK(t)S_{K}(t) in the Figure 6. In both panels, the red lines for βm=0\beta m=0 are taken from the conformal field theory in Dymarsky:2021bjq . From the figure, we can see that these curves (i.e., βm=0.001\beta m=0.001, 0.010.01, 0.10.1 and 0) will overlap together and cannot be distinguished from each other. This is consistent in numerics itself since in this case βm\beta m are very small and close to zero. Figure 6 indicates that at high temperatures βm1\beta m\ll 1, the behaviors of the Krylov complexity and Krylov entropy for a free scalar field theory are quite similar to those of the conformal field theory.

From the discussions in Parker:2018yvk , Krylov complexity will grow exponentially at late time as

K(t)eλKt,\displaystyle K(t)\propto e^{\lambda_{K}t}, (45)

in which λK=πω0=2π/β\lambda_{K}=\pi\omega_{0}=2\pi/\beta. In the left panel of Figure 6, the black line is the reference line for the Krylov complexity at late time. Therefore, we see that in the high temperature limit, the exponential growth rate for Krylov complexity is roughly λK6.368\lambda_{K}\approx 6.368 which is a little bit greater than 2π/β2\pi/\beta. This is because we are fitting the line in a finite time region. For larger time, the slope of the exponential growth will tend to 2π/β2\pi/\beta which is consistent with the discussions in Camargo:2022rnt ; Dymarsky:2021bjq .

In the right panel of Figure 6 the black line is the reference line and scale as ΔSK(t)/(Δt/β)=6.310\Delta S_{K}(t)/(\Delta t/\beta)=6.310, indicating the linear growth of the Krylov entropy in late time. The linear slope is also close to λK\lambda_{K}, which is consistent with the results in Fan:2022xaa ; Camargo:2022rnt that SK(t)logK(t)S_{K}(t)\sim\log K(t) at late time.

3.4 Origins of the absence of oscillations in the high-temperature limit

In the high-temperature limit, the behavior of the Krylov complexity and Krylov entropy is significantly different from that observed at low temperatures. Specifically, the oscillations that are present at low temperatures are absent at high temperatures. From the discrete Schrödinger equation, we can speculate that the behavior of φn(t)\varphi_{n}(t) will depend on the hopping amplitudes bnb_{n} and the value of the wave function φ0(t)\varphi_{0}(t). The hopping amplitudes bnb_{n} describes how the wave function can ‘jump’ from one position to the next position. Therefore, if bnb_{n} is staggering between even and odd positions, we can intuitively speculate that this staggering behavior will certainly affect the monotonic increase of φn(t)\varphi_{n}(t), and further affects the behavior the Krylov complexity. Moreover, the recursive relation of φn(t)\varphi_{n}(t) in the discrete Schrödinger equation indicates that the behavior of φn(t)\varphi_{n}(t) will also depend on φ0(t)\varphi_{0}(t). Therefore, the behavior of φ0(t)\varphi_{0}(t) will also affect the behavior of φn(t)\varphi_{n}(t) as well as the Krylov complexity. In the following we will compare the behaviors of bnb_{n} and φ0(t)\varphi_{0}(t) in the high-temperature and in the low-temperature regions, and we will see that the numerical results are consistent with our speculations.

  • Dependence on φ0(t)\varphi_{0}(t): In the Figure 7, we have plotted the profiles of the φ0(t)\varphi_{0}(t) for various values of βm\beta m, i.e., βm=0.01\beta m=0.01, βm=10\beta m=10 and βm=50\beta m=50. We can observe that in the low-temperature (βm=50\beta m=50) regime, φ0(t)\varphi_{0}(t) (green line) has significant oscillations. However, on the contrary, in the high-temperature (βm=0.01\beta m=0.01) regime, φ0(t)\varphi_{0}(t) (blue line) does not oscillate! Therefore, from the behavior of φ0(t)\varphi_{0}(t), we can conclude that the lower the temperature is, the more oscillations of φ0(t)\varphi_{0}(t) will become. Therefore, in the high-temperature regime, the absence of oscillations of Krylov complexity can be attributed to the lack of oscillations in φ0(t)\varphi_{0}(t).

    Refer to caption
    Figure 7: Behaviors of φ0(t)\varphi_{0}(t) for various temperatures βm\beta m. In the low-temperature regime (green line), φ0(t)\varphi_{0}(t) has significant oscillation; However, in the high-temperature regime (blue line), there are no oscillations of φ0(t)\varphi_{0}(t).
  • Dependence on bnb_{n}: It is known that if bnb_{n} is linearly related to nn, then the Krylov complexity grows exponentially Parker:2018yvk . However, in our case bnb_{n} separates into two groups of linear dependence on nn, i.e., there is staggering behavior of bnb_{n}. Therefore, we can speculate that the staggering behavior of bnb_{n} may cause the Krylov complexity to oscillate as well. Let γ1β(γodd+γeven)\gamma_{1}\equiv\beta(\gamma_{\text{odd}}+\gamma_{\text{even}}), γ2β(γoddγeven)\gamma_{2}\equiv\beta(\gamma_{\text{odd}}-\gamma_{\text{even}}), then we can reformulate bnb_{n} in Eqs.(43)-(44) as,

    βbn=βαn+γ12+(1)n+1γ22,nZ.\beta b_{n}=\beta\alpha n+\dfrac{\gamma_{1}}{2}+(-1)^{n+1}\frac{\gamma_{2}}{2},~{}~{}~{}~{}n\in Z. (46)

    When nn is sufficiently large, the linear term becomes significant. Then, the third term on the right side of Eq.(46) can be neglected. Therefore, at this point, the Krylov complexity should grow exponentially. To clarify why there is no oscillation at high temperatures and why the oscillation disappears over time at low temperatures, we can make the function φn(t)\varphi_{n}(t) continuous such that

    φ(x,t):=φn(t),x=ϵn,\varphi(x,t):=\varphi_{n}(t),\qquad x=\epsilon n, (47)

    where ϵ\epsilon is a small lattice cutoff. Assuming that when n>n^n>\hat{n}, the third term of βbn\beta b_{n} in Eq.(46) can be neglected, and let x^=ϵn^\hat{x}=\epsilon\hat{n}. Then, the Krylov complexity can also be rewritten in a continuous form

    K(t)\displaystyle K(t) =\displaystyle= 1+n=0n|φn(t)|2\displaystyle 1+\sum_{n=0}^{\infty}n\absolutevalue{\varphi_{n}(t)}^{2} (48)
    =\displaystyle= 1+1ϵ20𝑑xx|φ(x,t)|2\displaystyle 1+\frac{1}{\epsilon^{2}}\int_{0}^{\infty}dx~{}x\absolutevalue{\varphi(x,t)}^{2}
    =\displaystyle= 1+1ϵ20x^𝑑xx|φ(x,t)|2+1ϵ2x^𝑑xx|φ(x,t)|2.\displaystyle 1+\frac{1}{\epsilon^{2}}\int_{0}^{\hat{x}}dx~{}x\absolutevalue{\varphi(x,t)}^{2}+\frac{1}{\epsilon^{2}}\int_{\hat{x}}^{\infty}dx~{}x\absolutevalue{\varphi(x,t)}^{2}.

    The Krylov complexity represents the position of the wave function φ(x,t)\varphi(x,t) on the Krylov chain. When tt is small, its position is relatively close to the origin (x,t)=(0,0)(x,t)=(0,0), and the Krylov complexity is dominated by the second term in (48), which can disrupt the exponential growth and potentially lead to oscillations. When tt is sufficiently large, the position of the wave function is far from the origin, and the Krylov complexity is dominated by the third term in equation (48), thus exhibiting exponential growth behavior. This is why oscillations gradually fade away over time in low-temperature conditions. For a detailed discussion on this topic, please refer to reference Barbon:2019wsy .

    In the high-temperature limit, the Krylov complexity exhibits exponential growth from the very beginning. From our speculations, then the third term of βbn\beta b_{n} should be negligible when nn is very small. Therefore, in the high-temperature limit and when nn is very small, it should satisfy γ2/2βαn+γ1/2\gamma_{2}/2\ll\beta\alpha n+\gamma_{1}/2. This corresponds exactly to the high-temperature region as shown in the right panel of Figure 5 and in the Figure 8 below. In the high-temperature region, γ1\gamma_{1} is approximately around 6.3 (as βm\beta m is small in the Figure 8), and γ2\gamma_{2} is less than 0.6 (see the inset plot in the right panel of Figure 5), i.e., γ2/2βαn+γ1/2\gamma_{2}/2\ll\beta\alpha n+\gamma_{1}/2 for nn is small.

    Refer to caption
    Figure 8: Relation between βγ1\beta\gamma_{1} and the value of βm\beta m. We can see that in the high-temperature regime (βm\beta m is small), βγ16.3\beta\gamma_{1}\approx 6.3.

    Therefore, in high-temperature regime, Krylov complexity does not have oscillations. However, in the low-temperature regime (as βm\beta m is big enough) and when nn is small, γ2\gamma_{2} cannot be neglected. As we can see from the right panel of Figure 5 and the Figure 8, when βm=100\beta m=100, βγ1\beta\gamma_{1} is roughly 3535 and βγ2\beta\gamma_{2} is roughly 100100 which cannot be neglected. This means the staggering behavior of bnb_{n} cannot be neglected in low-temperature regime. Therefore, the Krylov complexity will exhibit oscillations in the low-temperature as nn is small.

Refer to caption
Figure 9: The Krylov complexity for βm=9,10,11\beta m=9,10,11. As βm=10\beta m=10 (black line), there exists a plateau at round K(t)2.4K(t)\approx 2.4 (red dashed line), which implies that βm=10\beta m=10 is the transition temperature that the Krylov complexity will change its behavior from monotonic increase to oscillations.

It will be interesting to see whether theres exists a transition temperature that Krylov complexity will change its behavior from the oscillations to monotonic increase. After careful seeking, we numerically found that at around βm=10\beta m=10 there indeed exists such kind of transition temperature. We plot the profile of K(t)K(t) as βm=9,10,11\beta m=9,10,11 in the Figure 9. We observe that at this transition temperature there exists a plateau at K(t)2.4K(t)\approx 2.4 (red dashed line). The existence of plateau indicates that K(t)K(t) at βm=10\beta m=10 is a critical line, since as βm<10\beta m<10 the Krylov complexity will increase monotonically and while βm>10\beta m>10 the Krylov complexity will have oscillation behaviors. Therefore, βm=10\beta m=10 is such kind of transition temperature.

3.5 Comparison with general temperatures

For the case of βm1\beta m\sim 1, the Eq.(28) for the low temperature limit in Camargo:2022rnt is also not applicable. Only the approximation in Eq.(41) is feasible in this case.

Refer to caption
(a) Krylov complexity for βm=1\beta m=1.
Refer to caption
(b) Krylov entropy for βm=1\beta m=1.
Refer to caption
(c) Krylov complexity for βm=0.1\beta m=0.1.
Refer to caption
(d) Krylov entropy for βm=0.1\beta m=0.1.
Refer to caption
(e) Krylov complexity for βm=50\beta m=50.
Refer to caption
(f) Krylov entropy for βm=50\beta m=50.
Figure 10: Time evolutions of Krylov complexity K(t)K(t) and Krylov entropy SK(t)S_{K}(t) for various values of βm\beta m. The panels (a), (c) and (e) for Krylov complexity are plotted in logarithmic scales. For all of the plots, the black lines are for kmax=200k_{\text{max}}=200 corresponding to Eq.(41), while the red (dashed) lines are for kmax=0k_{\text{max}}=0 corresponding to Eq.(28). The green lines are the reference lines for the scalings at late time.

In Figure 10, we show the time evolution of the Krylov complexity K(t)K(t) (in left column) and Krylov entropy SK(t)S_{K}(t) (in right column) for different values of βm\beta m. In these plots, the black curves are for kmax=200k_{\rm max}=200 corresponding to Eq.(41) while the red (dashed) curves are for kmax=0k_{\rm max}=0 corresponding to Eq.(28). From the panels in the first row, i.e. βm=1\beta m=1, we can see that the red lines will deviate from the black lines at late time. This reflects that the approximations in Eq.(28) is not accurate as the approximations in Eq.(41) at late time, which confirms our assertion that the approximation in Eq.(28) is only applicable in low temperature limit, i.e., βm1\beta m\gg 1. The green lines in the first row are the reference lines for K(t)K(t) and SK(t)S_{K}(t), respectively. In the panel (a) of Figure 10 we can see that the exponential growth rate of the Krylov complexity is roughly λK6.216\lambda_{K}\approx 6.216 which is close to 2π/β2\pi/\beta. This exponential scaling is consistent with our discussions in the preceding subsection. In the panel (b) of Figure 10, the linear scaling of the Krylov entropy is roughly ΔSK(t)/(Δt/β)6.189\Delta S_{K}(t)/(\Delta t/\beta)\approx 6.189, which is consistent with the claims that SK(t)logK(t)S_{K}(t)\sim\log K(t) at late time as we discussed in the preceding subsection. The small differences between the two scalings, i.e., 6.2166.216 and 6.1896.189 are due to the fittings of the two quantities at a finite time. It is expected that much longer time fitting will reduce this difference.

In the second row of Figure 10 we compare the Krylov complexity and Krylov entropy with distinct kmaxk_{\rm max} (i.e., kmax=200k_{\rm max}=200 and kmax=0k_{\rm max}=0) in high temperature limit (βm=0.11\beta m=0.1\ll 1). Again, the deviations between the red lines (kmax=0k_{\rm max}=0) and black lines (kmax=200k_{\rm max}=200) tell us that in the high temperature limit, the approximation in Eq.(28) is not valid. The late time scalings in K(t)K(t) and SK(t)S_{K}(t) (blue reference lines) are also close to 2π/β2\pi/\beta and consistent with the preceding discussions.

However, for βm1\beta m\gg 1, both the approximations in Eqs.(28) and (41) are valid, see for instance the last row in Figure 10, where we have set βm=50\beta m=50 for the computation of Krylov complexity and Krylov entropy. In the time region we have considered, both of kmax=200k_{\rm max}=200 and kmax=0k_{\rm max}=0 collapse together and cannot be distinguished from each other. This is different from the cases with βm=1\beta m=1 and βm=0.1\beta m=0.1 in the first two rows of Figure 10. Interestingly, the late time scalings for both Krylov quantities are much smaller than 2π/β2\pi/\beta, i.e., λK<2π/β\lambda_{K}<2\pi/\beta in the low temperature limit βm1\beta m\gg 1. Compared to high temperature limit, we may conjecture that the exponential growth rate of the Krylov complexity has a bound λK2π/β\lambda_{K}\leq 2\pi/\beta, which is consistent with the discussions in Camargo:2022rnt .

Another interesting thing is that in the low temperature limit βm1\beta m\gg 1, there are oscillations in K(t)K(t) in the short time region. The period is roughly π/m\pi/m. However, in the opposite limit, i.e., βm1\beta m\lesssim 1, there are no such oscillations in K(t)K(t). These phenomena are consistent with the observations in Camargo:2022rnt ; Dymarsky:2021bjq .

4 Conclusions

In this paper, we have performed a detailed study of the Krylov complexity and Krylov entropy for a free massive scalar field in five-dimensional spacetime under various temperature conditions. Previously, studies on the Krylov complexity of free scalar fields only considers the low-temperature limit, i.e., βm1\beta m\gg 1, since in that case the Wightman power spectrum can be easily approximated as in Eq.(28). In the general temperatures, i.e., general values of βm\beta m, we found a new expression for fW(ω)f^{W}(\omega) in the form of an infinite series in Eq.(34). We further obtained a good approximation of fW(ω)f^{W}(\omega) by truncating the infinite summations to a finite summation, which allowed us to study the Krylov complexity and Krylov entropy in the regimes where βm\beta m does not necessarily subject to βm1\beta m\gg 1.

Then we examined the performance of the Lanczos coefficients bnb_{n}, which exhibited the ‘staggering’ behavior by grouping bnb_{n} into odd and even nn families. As usual, bnb_{n} can be linearly fitted separately for odd and even nn for large nn. The linear coefficients α\alpha satisfies the relation απ/β\alpha\approx\pi/\beta which is consistent with the previous findings in the universal operator growth hypothesis. We further observed that the difference γoddγeven\gamma_{\text{odd}}-\gamma_{\text{even}} is linearly proportional to the mass mm for larger mm, however, this proportionality breaks down for small mm. This phenomenon was not explored in the existing literatures and we argued that the breakdown of the linear proportionality may come from the high temperature limit.

Next, we investigated the cases for general temperatures with the general values of βm\beta m. In the late time, we found that the Krylov complexity exhibited an exponential growth K(t)eλKtK(t)\sim e^{\lambda_{K}t} where the exponent satisfies the bound λK2π/β\lambda_{K}\leq 2\pi/\beta. The Krylov entropy is linearly proportional to tt in the late time and satisfying the relation S(t)logK(t)S(t)\sim\log K(t). In the high temperature limit βm1\beta m\ll 1, the Krylov complexity exhibits behaviors remarkably similar to those of the conformal field theory where βm=0\beta m=0. Moreover, it was observed that in the high-temperature limit the Krylov complexity would monotonically increase, while in the low-temperature limit it would oscillate. We speculated that this difference was due to the staggering of bnb_{n} as well as the function φ0(t)\varphi_{0}(t). Our speculation was consistent with the numerical results. Subsequently, we numerically found that there existed a plateau in the Krylov complexity, which indicated that the temperature at this point was the transition temperature. Lower or higher temperatures would destroy this plateau, leading to oscillations or monotonic increasing of the Krylov complexity. Therefore, the temperature at which this plateau appears can be considered a ‘critical temperature’.

We also checked that the finite truncations kmax=200k_{\rm max}=200 in our paper is accurate enough to get the correct results by comparing it to the previous studies with kmax=0k_{\rm max}=0. Therefore, the schemes in our paper to study the Krylov complexity and Krylov entropy with general temperatures is reliable.

In conclusion, our study provides a comprehensive analysis of the Krylov complexity and Krylov entropy in a five-dimensional scalar field theory under varying temperature conditions. These findings enhance our understanding of operator growth and complexity in quantum field theories, particularly in relation to thermal effects. Future work may extend these results to other types of field theories and different spacetime dimensions, by exploring the rich structures of Krylov complexity in diverse physical settings.

Acknowledgements.
This work was partially supported by the National Natural Science Foundation of China (Grants No.12175008).

Appendix A Krylov complexity and Krylov entropy in symmetry breaking phase

In this appendix, we will study the Krylov complexity and Krylov entropy in the symmetry breaking phase by adopting a model of a five-dimensional real Higgs scalar field theory. The steps will be similar to those in the main context except that the mass mm will be changed to M(φ)M(\varphi) which will be shown later.

Assume ϕ\phi is a Higgs scalar field which can be described by the following Lagrangian,

LHiggs=12μϕμϕ12m2ϕ2λϕ4,L_{\rm Higgs}=\frac{1}{2}\partial_{\mu}\phi\partial^{\mu}\phi-\frac{1}{2}m^{2}\phi^{2}-\lambda\phi^{4}, (49)

in which m2<0m^{2}<0 and λ>0\lambda>0 in order to render the potential

U(ϕ)=12m2ϕ2+λϕ4\displaystyle U(\phi)=\frac{1}{2}m^{2}\phi^{2}+\lambda\phi^{4} (50)

behaves like a double well potential. According to the Appendix B, we can write ϕ(X)\phi(X) as (61)

ϕ(X)=φ+σ(X),\phi(X)=\varphi+\sigma(X), (51)

where φ\varphi is the classical value corresponding to the minimal point of the potential while σ(X)\sigma(X) is a new dynamical field. XX is the Euclidean coordinates of the spacetime. In the thermal field theory, the propagator of the σ\sigma field is (73)444For more details, see Appendix B.

𝒟σ(K)=1ωn2+𝐤2+M2(φ),\mathcal{D}_{\sigma}(K)=\frac{1}{\omega_{n}^{2}+\mathbf{k}^{2}+M^{2}(\varphi)}, (52)

where

M2(φ)=m2+12λφ2.\displaystyle M^{2}(\varphi)=m^{2}+12\lambda\varphi^{2}. (53)

After quantum corrections in d=5d=5 Higgs field theory, φ\varphi here will take the value as (97)

φQ.C.=φ¯=±12πm2π23λT3ζ(3)λ.\varphi_{\rm Q.C.}=\bar{\varphi}=\pm\frac{1}{2\pi}\sqrt{\frac{-m^{2}\pi^{2}-3\lambda T^{3}\zeta(3)}{\lambda}}. (54)

In computing the Krylov complexity the key point is to get the Wightman power spectrum fW(ω)f^{W}(\omega) in Eq.(26). However, fW(ω)f^{W}(\omega) is calculated from the propagators in the field theory Camargo:2022rnt . Compared to the free field theory in Camargo:2022rnt , we find that the difference of the propagators between Higgs field theory and free field theory is the mass. Therefore, when calculating the Krylov complexity and Krylov entropy in the symmetry breaking phase, the only thing we need to do is to substitute mm in free field theory in the main context with the M(φ)M(\varphi) in Eq.(53). From this respect, we will not show the figures of Krylov complexity and Krylov entropy further in the symmetry breaking phase. They are expected to behave similarly to those in the main context by just regarding mm as M(φ)M(\varphi).

Appendix B Recapitulation of Higgs Scalar Field Theory

In this appendix, we will compute the propagators and mass shift in the symmetry breaking phase of a real Higgs scalar field theory. In the formula of the Lagrangian Eq.(49), we adopt the metric signature as (+,,,)(+,-,-,\cdots). One can check that the Lagrangian is invariant under Z2Z_{2} symmetry. The Higgs potential Eq.(50) has two global minima located at

ϕ=±v=±m24λ.\phi=\pm v=\pm\sqrt{\frac{-m^{2}}{4\lambda}}. (55)

If we choose the vacuum expectation value ϕ=v\expectationvalue{\phi}=v, we can write ϕ(X)\phi(X)555ϕ(X):=ϕ(τ,𝐱)\phi(X):=\phi(\tau,\bf{x}) and we have used the imaginary-time formalism. In this formalism tt is replaced by iτ-i\tau and 𝐱\bf{x} represents the spatial coordinates. as

ϕ(X)=v+σ(X).\phi(X)=v+\sigma(X). (56)

The Lagrangian (49) can be rewritten as

LHiggs=12μσμσ12(m2+12λv2)σ24λvσ3λσ4U(v),L_{\rm Higgs}=\frac{1}{2}\partial_{\mu}\sigma\partial^{\mu}\sigma-\frac{1}{2}(m^{2}+12\lambda v^{2})\sigma^{2}-4\lambda v\sigma^{3}-\lambda\sigma^{4}-U(v), (57)

where the tree-level potential reads

U(v)=12m2v2+λv4.U(v)=\frac{1}{2}m^{2}v^{2}+\lambda v^{4}. (58)

Therefore, Eq.(57) describes a scalar field σ\sigma with mass m2+12λv2\sqrt{m^{2}+12\lambda v^{2}} and it is not symmetric under the transformation σσ\sigma\leftrightarrow-\sigma. The previous Z2Z_{2} symmetry is spontaneously broken.

On the other hand, we can perform Fourier transformation for ϕ(X)\phi(X)

ϕ(X)=βVKeiKXϕK,Kn=𝐤,KXωnτ+𝐤𝐱,\phi(X)=\sqrt{\frac{\beta}{V}}\sum_{K}e^{-iK\cdot X}\phi_{K},\qquad\sum_{K}\equiv\sum_{n=-\infty}^{\infty}\sum_{\mathbf{k}},\qquad-K\cdot X\equiv\omega_{n}\tau+\mathbf{k}\cdot\mathbf{x}, (59)

where ωn2πn/β(n)\omega_{n}\equiv 2\pi n/\beta~{}(n\in\mathbb{Z}) are called Matsubara frequency and Kμ=(iωn,𝐤),Xμ=(iτ,𝐱)K^{\mu}=(-i\omega_{n},\mathbf{k}),X^{\mu}=(-i\tau,\mathbf{x}). We can write down the partition function following the conventions in Kapusta:2006pm that

𝒵=𝑑ϕ0K0dReϕKdImϕKexp(S[ϕ]),\mathcal{Z}=\int_{-\infty}^{\infty}d\phi_{0}\prod_{K\neq 0}\int_{-\infty}^{\infty}d\real\phi_{K}\int_{-\infty}^{\infty}d\imaginary\phi_{K}\exp(S[\phi]), (60)

in which S[ϕ]=0β𝑑τd3𝐱LHiggsS[\phi]=\int_{0}^{\beta}d\tau d^{3}\mathbf{x}L_{\rm Higgs}. By extracting out the zero mode of ϕ(X)\phi(X), we get

ϕ(X)=βVKϕKeiKX=βVϕ0+βVK0ϕKeiKX=φ+σ(X),\phi(X)=\sqrt{\frac{\beta}{V}}\sum_{K}\phi_{K}e^{-iK\cdot X}=\sqrt{\frac{\beta}{V}}\phi_{0}+\sqrt{\frac{\beta}{V}}\sum_{K\neq 0}\phi_{K}e^{-iK\cdot X}=\varphi+\sigma(X), (61)

where φ=v=βVϕ0\varphi=v=\sqrt{\frac{\beta}{V}}\phi_{0} is independent of XX and σ(X)K0eiKXϕK=K0eiKXσK\sigma(X)\equiv\sum_{K\neq 0}e^{-iK\cdot X}\phi_{K}=\sum_{K\neq 0}e^{-iK\cdot X}\sigma_{K}.

Then we can define the effective potential as

𝒱eff(φ)=TVln(K0dReσKdImσKexp(S[φ+σ]))=TVln(𝒟σeS[φ+σ]),\mathcal{V}_{\text{eff}}(\varphi)=-\frac{T}{V}\ln(\prod_{K\neq 0}\int_{-\infty}^{\infty}d\real\sigma_{K}\int_{-\infty}^{\infty}d\imaginary\sigma_{K}\exp(S[\varphi+\sigma]))=-\frac{T}{V}\ln(\int\mathcal{D}\sigma e^{S[\varphi+\sigma]}), (62)

where 𝒟σK0dReσKdImσK\int\mathcal{D}\sigma\equiv\prod_{K\neq 0}\int_{-\infty}^{\infty}d\real\sigma_{K}\int_{-\infty}^{\infty}d\imaginary\sigma_{K} and S[φ+σ]=S[ϕ]S[\varphi+\sigma]=S[\phi]. Then we have

𝒵=VT𝑑φexp(VT𝒱eff(φ)).\mathcal{Z}=\sqrt{VT}\int_{-\infty}^{\infty}d\varphi\exp(-\frac{V}{T}\mathcal{V}_{\text{eff}}(\varphi)). (63)

Without loss of generality, we assume that the minimum of effective potential is located at φ=φ¯>0\varphi=\bar{\varphi}>0. In the thermodynamic limit VV\rightarrow\infty, the partition function can be written nearby φ¯\bar{\varphi} as

𝒵=exp(VT𝒱eff(φ¯))2πT2𝒱eff′′(φ¯).\mathcal{Z}=\exp(-\frac{V}{T}\mathcal{V}_{\text{eff}}(\bar{\varphi}))\sqrt{\frac{2\pi T^{2}}{\mathcal{V}^{\prime\prime}_{\text{eff}}(\bar{\varphi})}}. (64)

Next, we need to compute φ¯\bar{\varphi}. To this end, we need to calculate the effective potential (62) first. Note that S[φ+σ]S[\varphi+\sigma] is actually the spacetime integral of the Lagrangian (57) with v=φv=\varphi. For convenience, we rewrite the Lagrangian as,

[σ,φ]=12μσμσ12M2(φ)σ24λφσ3λσ4U(φ),\mathcal{L}[\sigma,\varphi]=\frac{1}{2}\partial_{\mu}\sigma\partial^{\mu}\sigma-\frac{1}{2}M^{2}(\varphi)\sigma^{2}-4\lambda\varphi\sigma^{3}-\lambda\sigma^{4}-U(\varphi), (65)

where

M2(φ)=m2+12λφ2,\displaystyle M^{2}(\varphi)=m^{2}+12\lambda\varphi^{2}, (66)
U(φ)=12m2φ2+λφ4.\displaystyle U(\varphi)=\frac{1}{2}m^{2}\varphi^{2}+\lambda\varphi^{4}. (67)

Treating σ\sigma as a dynamical field, the Lagrangian can be divided into a free part 0\mathcal{L}_{0} and an interacting part I\mathcal{L}_{I} as,

[σ,φ]=0+IU(φ),\mathcal{L}[\sigma,\varphi]=\mathcal{L}_{0}+\mathcal{L}_{I}-U(\varphi), (68)

where

0=12μσμσ12M2(φ)σ2,I=4λφσ3λσ4.\mathcal{L}_{0}=\frac{1}{2}\partial_{\mu}\sigma\partial^{\mu}\sigma-\frac{1}{2}M^{2}(\varphi)\sigma^{2},\qquad\mathcal{L}_{I}=-4\lambda\varphi\sigma^{3}-\lambda\sigma^{4}. (69)

Since φ\varphi is independent of XX and σ\sigma, the effective potential (62) can be decomposed into (notice that T=β1T=\beta^{-1})

𝒱eff(φ)=U(φ)TVln(𝒟σeS0[σ]+SI[σ]),\mathcal{V}_{\text{eff}}(\varphi)=U(\varphi)-\frac{T}{V}\ln\left(\int\mathcal{D}\sigma e^{S_{0}[\sigma]+S_{I}[\sigma]}\right), (70)

in which S0[σ]=0β𝑑τd3𝐱0S_{0}[\sigma]=\int_{0}^{\beta}d\tau d^{3}\mathbf{x}\mathcal{L}_{0} and SI[σ]=0β𝑑τd3𝐱IS_{I}[\sigma]=\int_{0}^{\beta}d\tau d^{3}\mathbf{x}\mathcal{L}_{I}.

In perturbation theory, the effective potential can be expanded in terms of \hbar as (we have set 1\hbar\equiv 1)

𝒱eff(φ)=𝒱eff(0)(φ)+𝒱eff(1)(φ)+𝒱eff(2)+,\mathcal{V}_{\text{eff}}(\varphi)=\mathcal{V}^{(0)}_{\text{eff}}(\varphi)+\mathcal{V}^{(1)}_{\text{eff}}(\varphi)+\mathcal{V}_{\text{eff}}^{(2)}+\cdots, (71)

where the superscript index ii in 𝒱eff(i)\mathcal{V}^{(i)}_{\text{eff}} indicates the perturbations in the ii-th power in \hbar. Obviously, 𝒱eff(0)=U(φ)\mathcal{V}_{\text{eff}}^{(0)}=U(\varphi) and

𝒱eff(1)(φ)=TVln(𝒟σeS0[σ])=12TVKln(𝒟σ1(K)T2),\displaystyle\mathcal{V}_{\text{eff}}^{(1)}(\varphi)=-\frac{T}{V}\ln\left(\int\mathcal{D}\sigma e^{S_{0}[\sigma]}\right)=\frac{1}{2}\frac{T}{V}\sum_{K}\ln(\frac{\mathcal{D}_{\sigma}^{-1}(K)}{T^{2}}), (72)
𝒟σ(K)=1ωn2+𝐤2+M2(φ).\displaystyle\mathcal{D}_{\sigma}(K)=\frac{1}{\omega_{n}^{2}+\mathbf{k}^{2}+M^{2}(\varphi)}. (73)

The value of φ\varphi needs to be determined by the minimum point of the effective potential. We will only consider zeroth order and first order terms of the effective potential, i.e., 𝒱eff(0)(φ)\mathcal{V}_{\text{eff}}^{(0)}(\varphi) and 𝒱eff(1)(φ)\mathcal{V}_{\text{eff}}^{(1)}(\varphi). Assuming the spacetime is dd-dimensional, then following the steps in thermal field theory Laine:2016hma , we get

𝒱eff(0)(φ)=12m2φ2+λφ4,\displaystyle\mathcal{V}^{(0)}_{\text{eff}}(\varphi)=\frac{1}{2}m^{2}\varphi^{2}+\lambda\varphi^{4}, (74)
𝒱eff(1)=dd1𝐤(2π)d1[E𝐤2+Tln(1eE𝐤/T)],\displaystyle\mathcal{V}^{(1)}_{\text{eff}}=\int\frac{d^{d-1}\mathbf{k}}{(2\pi)^{d-1}}\left[\frac{E_{\mathbf{k}}}{2}+T\ln(1-e^{-E_{\mathbf{k}}/T})\right], (75)

where E𝐤=𝐤2+M2(φ)E_{\mathbf{k}}=\sqrt{\mathbf{k}^{2}+M^{2}(\varphi)}. We can recombine these two equations into a TT-dependent part and a TT-independent part

𝒱0(φ)=12m2φ2+λφ4+dd1𝐤(2π)d1E𝐤2,\displaystyle\mathcal{V}_{0}(\varphi)=\frac{1}{2}m^{2}\varphi^{2}+\lambda\varphi^{4}+\int\frac{d^{d-1}\mathbf{k}}{(2\pi)^{d-1}}\frac{E_{\mathbf{k}}}{2}, (76)
𝒱T(φ)=Tdd1𝐤(2π)d1ln(1eE𝐤/T).\displaystyle\mathcal{V}_{T}(\varphi)=T\int\frac{d^{d-1}\mathbf{k}}{(2\pi)^{d-1}}\ln(1-e^{-E_{\mathbf{k}}/T}). (77)

Using the identity Dolan:1973qd

E𝐤=dk02πiln(k02+E𝐤2iϵ),E_{\mathbf{k}}=\int_{-\infty}^{\infty}\frac{dk_{0}}{2\pi i}\ln(-k_{0}^{2}+E_{\mathbf{k}}^{2}-i\epsilon), (78)

the integration in (76) can be rewritten as

J(M)dd1𝐤(2π)d1E𝐤2=i2ddK(2π)dln(K2+M2iϵ),J(M)\equiv\int\frac{d^{d-1}\mathbf{k}}{(2\pi)^{d-1}}\frac{E_{\mathbf{k}}}{2}=-\frac{i}{2}\int\frac{d^{d}K}{(2\pi)^{d}}\ln(-K^{2}+M^{2}-i\epsilon), (79)

where we have set Kμ=(k0,𝐤)K^{\mu}=(k^{0},\mathbf{k}) and M=M(φ)M=M(\varphi). We can rotate the k0k_{0}-integral onto the imaginary axis,let k0=ik4k_{0}=ik_{4}, without crossing the branching cuts of the logarithm. With Euclidean dd-momentum variable KE(kd,𝐤)K_{E}\equiv(k_{d},\mathbf{k}) we have

J(M)12ddKE(2π)dln(KE2+M2).J(M)\equiv\frac{1}{2}\int\frac{d^{d}K_{E}}{(2\pi)^{d}}\ln(K^{2}_{E}+M^{2}). (80)

Using the following identities Peskin:1995ev

ddKE(2π)d=dΩd(2π)d0𝑑KEKEd1,𝑑Ωd=2πd/2Γ(d/2),\int\frac{d^{d}K_{E}}{(2\pi)^{d}}=\int\frac{d\Omega_{d}}{(2\pi)^{d}}\int_{0}^{\infty}dK_{E}K^{d-1}_{E},\qquad\int d\Omega_{d}=\frac{2\pi^{d/2}}{\Gamma(d/2)}, (81)

we have

J(M)=122πd/2Γ(d/2)(2π)d0𝑑xxd1ln(x2+M2).J(M)=\frac{1}{2}\frac{2\pi^{d/2}}{\Gamma(d/2)(2\pi)^{d}}\int_{0}^{\infty}dxx^{d-1}\ln(x^{2}+M^{2}). (82)

However, this integral is divergent. Therefore, we need to introduce a truncation Λ\Lambda such that

J(M)\displaystyle J(M) =122πd/2Γ(d/2)(2π)d0Λ𝑑xxd1ln(x2+M2)\displaystyle=\frac{1}{2}\frac{2\pi^{d/2}}{\Gamma(d/2)(2\pi)^{d}}\int_{0}^{\Lambda}dxx^{d-1}\ln(x^{2}+M^{2}) (83)
=12d2πd/2Γ(d/2)(2π)d[Λdln(Λ2+M2)0Λ2xxdx2+M2𝑑x]\displaystyle=\frac{1}{2d}\frac{2\pi^{d/2}}{\Gamma(d/2)(2\pi)^{d}}\left[\Lambda^{d}\ln(\Lambda^{2}+M^{2})-\int_{0}^{\Lambda}\frac{2x\cdot x^{d}}{x^{2}+M^{2}}dx\right]
=12d2πd/2Γ(d/2)(2π)d{Λdln(Λ2+M2)2Λ2+d(2+d)M2F12[1,2+d2,4+d2,Λ2M2]},\displaystyle=\frac{1}{2d}\frac{2\pi^{d/2}}{\Gamma(d/2)(2\pi)^{d}}\left\{\Lambda^{d}\ln(\Lambda^{2}+M^{2})-\frac{2\Lambda^{2+d}}{(2+d)M^{2}}{}_{2}F_{1}\left[1,\frac{2+d}{2},\frac{4+d}{2},-\frac{\Lambda^{2}}{M^{2}}\right]\right\},

where F12{}_{2}F_{1} is the hypergeometric function. In the following, we will compute J(M)J(M) with the dimensions with d=4d=4 and d=5d=5.

B.1 d=4d=4

If d=4d=4, we can expand J(M)J(M) in terms of 1/Λ1/\Lambda

J(M)={132π2[M2Λ2M42(lnΛ2M2+12)]}+Λ464π2[lnΛ212]+O(1/Λ2),J(M)=\left\{\frac{1}{32\pi^{2}}\left[M^{2}\Lambda^{2}-\frac{M^{4}}{2}\left(\ln\frac{\Lambda^{2}}{M^{2}}+\frac{1}{2}\right)\right]\right\}+\frac{\Lambda^{4}}{64\pi^{2}}\left[\ln\Lambda^{2}-\frac{1}{2}\right]+O(1/\Lambda^{2}), (84)

in which the second term is divergence and independent of φ\varphi, therefore, we can directly drop it. Let

lnΛ2M2=lnΛ2μ2+lnμ2M2,withM2=m2+12λφ2,\ln\frac{\Lambda^{2}}{M^{2}}=\ln\frac{\Lambda^{2}}{\mu^{2}}+\ln\frac{\mu^{2}}{M^{2}},\qquad{\rm with}~{}~{}M^{2}=m^{2}+12\lambda\varphi^{2}, (85)

where μ\mu is an arbitrary energy scale. Then we have

J(M)=\displaystyle J(M)= [Λ2m232π2m464π2lnΛ2μ2+φ2(3Λ28π2λ3m28π2λlnΛ2μ2)φ49λ24π2lnΛ2μ2]\displaystyle\left[\frac{\Lambda^{2}m^{2}}{32\pi^{2}}-\frac{m^{4}}{64\pi^{2}}\ln\frac{\Lambda^{2}}{\mu^{2}}+\varphi^{2}\left(\frac{3\Lambda^{2}}{8\pi^{2}}\lambda-\frac{3m^{2}}{8\pi^{2}}\lambda\ln\frac{\Lambda^{2}}{\mu^{2}}\right)-\varphi^{4}\frac{9\lambda^{2}}{4\pi^{2}}\ln\frac{\Lambda^{2}}{\mu^{2}}\right] (86)
m4128π2m464π2lnμ2m2+12φ2λφ2(3m216π2λ+3m2λ8π2lnμ2m2+12φ2λ)\displaystyle-\frac{m^{4}}{128\pi^{2}}-\frac{m^{4}}{64\pi^{2}}\ln\frac{\mu^{2}}{m^{2}+12\varphi^{2}\lambda}-\varphi^{2}\left(\frac{3m^{2}}{16\pi^{2}}\lambda+\frac{3m^{2}\lambda}{8\pi^{2}}\ln\frac{\mu^{2}}{m^{2}+12\varphi^{2}\lambda}\right)
φ4(9λ28π2+9λ24π2lnμ2m2+12φ2λ).\displaystyle-\varphi^{4}\left(\frac{9\lambda^{2}}{8\pi^{2}}+\frac{9\lambda^{2}}{4\pi^{2}}\ln\frac{\mu^{2}}{m^{2}+12\varphi^{2}\lambda}\right).

The part in the square brackets [][\dots] is divergent and can be directly discarded. In addition, m4128π2-\frac{m^{4}}{128\pi^{2}} is independent of φ\varphi, thus can also be discarded without affecting our results. Therefore, the part of the effective potential (76) that does not depend on the temperature is

𝒱0(φ)=(13λ8π2)m22φ2+(19λ8π2)λφ4+164π2(m2+12λφ2)2ln(m2+12λφ2μ2).\mathcal{V}_{0}(\varphi)=\left(1-\frac{3\lambda}{8\pi^{2}}\right)\frac{m^{2}}{2}\varphi^{2}+\left(1-\frac{9\lambda}{8\pi^{2}}\right)\lambda\varphi^{4}+\frac{1}{64\pi^{2}}\left(m^{2}+12\lambda\varphi^{2}\right)^{2}\ln\left(\frac{m^{2}+12\lambda\varphi^{2}}{\mu^{2}}\right). (87)

Assume that the coupling constant λ1\lambda\ll 1, then Laine:2016hma

𝒱0(φ)=m22φ2+λφ4,\mathcal{V}_{0}(\varphi)=\frac{m^{2}}{2}\varphi^{2}+\lambda\varphi^{4}, (88)

where we have discarded all terms that do not depend on φ\varphi, as such terms will not affect our results. Now consider the part of the effective potential that depends on temperature (77)

𝒱T(φ)\displaystyle\mathcal{V}_{T}(\varphi) =Td3𝐤(2π)3ln(1eE𝐤/T)\displaystyle=T\int\frac{d^{3}\mathbf{k}}{(2\pi)^{3}}\ln(1-e^{-E_{\mathbf{k}}/T}) (89)
=4πT40dx(2π)3x2ln(1ex2+α2),\displaystyle=4\pi T^{4}\int_{0}^{\infty}\frac{dx}{(2\pi)^{3}}x^{2}\ln(1-e^{-\sqrt{x^{2}+\alpha^{2}}}),

where α=M/T\alpha=M/T. In the high temperature limit α1\alpha\ll 1,

ln(1ex2+α2)ln(1ex)+1ex1α22x.\ln(1-e^{-\sqrt{x^{2}+\alpha^{2}}})\approx\ln(1-e^{-x})+\frac{1}{e^{x}-1}\frac{\alpha^{2}}{2x}. (90)

Then we get

𝒱T(φ)=π290T4+T224(m2+12φ2λ),\mathcal{V}_{T}(\varphi)=-\frac{\pi^{2}}{90}T^{4}+\frac{T^{2}}{24}(m^{2}+12\varphi^{2}\lambda), (91)

in which the terms independent of φ\varphi can be discarded again. Therefore, combining Eqs.(88) and (91) the effective potential under weak coupling λ1\lambda\ll 1 and high temperature α1\alpha\ll 1 is Laine:2016hma

𝒱eff(φ)=12(m2+λT2)φ2+λφ4.\mathcal{V}_{\text{eff}}(\varphi)=\frac{1}{2}(m^{2}+\lambda T^{2})\varphi^{2}+\lambda\varphi^{4}. (92)

The minimum point of the effective potential is located at

φ¯=±12m2λT2λ.\bar{\varphi}=\pm\frac{1}{2}\sqrt{\frac{-m^{2}-\lambda T^{2}}{\lambda}}. (93)

B.2 d=5d=5

Following the previous operations, we can obtain the effective potential with d=5d=5. The specific process will not be repeated here. The part of the effective potential that does not depend on temperature is

𝒱0(5)(φ)=\displaystyle\mathcal{V}_{0}^{(5)}(\varphi)= (1+2λm2+12λφ25π2)m22φ2+(1+6λm2+12λφ25π2)λφ4\displaystyle\left(1+\frac{2\lambda\sqrt{m^{2}+12\lambda\varphi^{2}}}{5\pi^{2}}\right)\frac{m^{2}}{2}\varphi^{2}+\left(1+\frac{6\lambda\sqrt{m^{2}+12\lambda\varphi^{2}}}{5\pi^{2}}\right)\lambda\varphi^{4} (94)
+m4m2+12λφ2120π2.\displaystyle+\frac{m^{4}\sqrt{m^{2}+12\lambda\varphi^{2}}}{120\pi^{2}}.

The part of the effective potential that depends on temperature is

𝒱T(5)(φ)=3λT3φ2ζ(3)2π2,\mathcal{V}_{T}^{(5)}(\varphi)=\frac{3\lambda T^{3}\varphi^{2}\zeta(3)}{2\pi^{2}}, (95)

where ζ\zeta is the Riemann-zeta function. In the limit of high temperature and weak coupling, the effective potential is

𝒱eff(5)(φ)=m22φ2+λφ4+3λT3φ2ζ(3)2π2=12(m2+3λT3ζ(3)π2)φ2+λφ4.\mathcal{V}^{(5)}_{\text{eff}}(\varphi)=\frac{m^{2}}{2}\varphi^{2}+\lambda\varphi^{4}+\frac{3\lambda T^{3}\varphi^{2}\zeta(3)}{2\pi^{2}}=\frac{1}{2}\left(m^{2}+\frac{3\lambda T^{3}\zeta(3)}{\pi^{2}}\right)\varphi^{2}+\lambda\varphi^{4}. (96)

The minimum point of the effective potential is located at

φ¯=±12πm2π23λT3ζ(3)λ.\bar{\varphi}=\pm\frac{1}{2\pi}\sqrt{\frac{-m^{2}\pi^{2}-3\lambda T^{3}\zeta(3)}{\lambda}}. (97)

References

  • (1) M.A. Nielsen, M.R. Dowling, M. Gu and A.C. Doherty, Quantum Computation as Geometry, Science 311 (2006) 1133 [quant-ph/0603161].
  • (2) R. Jefferson and R.C. Myers, Circuit complexity in quantum field theory, JHEP 10 (2017) 107 [1707.08570].
  • (3) D.E. Parker, X. Cao, A. Avdoshkin, T. Scaffidi and E. Altman, A Universal Operator Growth Hypothesis, Phys. Rev. X 9 (2019) 041017 [1812.08657].
  • (4) A.R. Brown, D.A. Roberts, L. Susskind, B. Swingle and Y. Zhao, Holographic complexity equals bulk action?, Phys. Rev. Lett. 116 (2016) 191301.
  • (5) L. Susskind, Computational Complexity and Black Hole Horizons, Fortsch. Phys. 64 (2016) 24 [1403.5695].
  • (6) S. Aaronson, The Complexity of Quantum States and Transformations: From Quantum Money to Black Holes, 7, 2016 [1607.05256].
  • (7) S. Sachdev and J. Ye, Gapless spin fluid ground state in a random, quantum Heisenberg magnet, Phys. Rev. Lett. 70 (1993) 3339 [cond-mat/9212030].
  • (8) A. Kitaev, A simple model of quantum holography, Talks at KITP (2015) .
  • (9) A. Bhattacharya, P. Nandy, P.P. Nath and H. Sahu, Operator growth and Krylov construction in dissipative open quantum systems, JHEP 12 (2022) 081 [2207.05347].
  • (10) B. Bhattacharjee, X. Cao, P. Nandy and T. Pathak, Operator growth in open quantum systems: lessons from the dissipative SYK, JHEP 03 (2023) 054 [2212.06180].
  • (11) C. Liu, H. Tang and H. Zhai, Krylov complexity in open quantum systems, Phys. Rev. Res. 5 (2023) 033085 [2207.13603].
  • (12) K. Hashimoto, K. Murata, N. Tanahashi and R. Watanabe, Krylov complexity and chaos in quantum mechanics, JHEP 11 (2023) 040 [2305.16669].
  • (13) K. Adhikari, S. Choudhury and A. Roy, Krylov Complexity in Quantum Field Theory, Nucl. Phys. B 993 (2023) 116263 [2204.02250].
  • (14) A. Avdoshkin, A. Dymarsky and M. Smolkin, Krylov complexity in quantum field theory, and beyond, 2212.14429.
  • (15) E. Rabinovici, A. Sánchez-Garrido, R. Shir and J. Sonner, Operator complexity: a journey to the edge of Krylov space, JHEP 06 (2021) 062 [2009.01862].
  • (16) H.A. Camargo, V. Jahnke, K.-Y. Kim and M. Nishida, Krylov complexity in free and interacting scalar field theories with bounded power spectrum, JHEP 05 (2023) 226 [2212.14702].
  • (17) A. Dymarsky and M. Smolkin, Krylov complexity in conformal field theory, Phys. Rev. D 104 (2021) L081702 [2104.09514].
  • (18) B. Bhattacharjee, X. Cao, P. Nandy and T. Pathak, Krylov complexity in saddle-dominated scrambling, JHEP 05 (2022) 174 [2203.03534].
  • (19) B. Bhattacharjee, S. Sur and P. Nandy, Probing quantum scars and weak ergodicity breaking through quantum complexity, Phys. Rev. B 106 (2022) 205150 [2208.05503].
  • (20) V. Balasubramanian, J.M. Magan and Q. Wu, Quantum chaos, integrability, and late times in the Krylov basis, 2312.03848.
  • (21) A. Dymarsky and A. Gorsky, Quantum chaos as delocalization in Krylov space, Phys. Rev. B 102 (2020) 085137 [1912.12227].
  • (22) M.J. Vasli, K. Babaei Velni, M.R. Mohammadi Mozaffar, A. Mollabashi and M. Alishahiha, Krylov complexity in Lifshitz-type scalar field theories, Eur. Phys. J. C 84 (2024) 235 [2307.08307].
  • (23) A. Banerjee, A. Bhattacharyya, P. Drashni and S. Pawar, From cfts to theories with bondi-metzner-sachs symmetries: Complexity and out-of-time-ordered correlators, Phys. Rev. D 106 (2022) 126022.
  • (24) N. Iizuka and M. Nishida, Krylov complexity in the IP matrix model, JHEP 11 (2023) 065 [2306.04805].
  • (25) N. Iizuka and M. Nishida, Krylov complexity in the IP matrix model. Part II, JHEP 11 (2023) 096 [2308.07567].
  • (26) J. Erdmenger, S.-K. Jian and Z.-Y. Xian, Universal chaotic dynamics from Krylov space, JHEP 08 (2023) 176 [2303.12151].
  • (27) B. Bhattacharjee, P. Nandy and T. Pathak, Krylov complexity in large q and double-scaled SYK model, JHEP 08 (2023) 099 [2210.02474].
  • (28) B. Bhattacharjee, P. Nandy and T. Pathak, Operator dynamics in Lindbladian SYK: a Krylov complexity perspective, JHEP 01 (2024) 094 [2311.00753].
  • (29) A. Bhattacharya, P. Nandy, P.P. Nath and H. Sahu, On Krylov complexity in open systems: an approach via bi-Lanczos algorithm, JHEP 12 (2023) 066 [2303.04175].
  • (30) A. Bhattacharyya, S.S. Haque, G. Jafari, J. Murugan and D. Rapotu, Krylov complexity and spectral form factor for noisy random matrix models, JHEP 10 (2023) 157 [2307.15495].
  • (31) V. Malvimat, S. Porey and B. Roy, Krylov Complexity in 2d2d CFTs with SL(2,)(2,\mathbb{R}) deformed Hamiltonians, 2402.15835.
  • (32) P. Caputa, H.-S. Jeong, S. Liu, J.F. Pedraza and L.-C. Qu, Krylov complexity of density matrix operators, JHEP 05 (2024) 337 [2402.09522].
  • (33) M. Afrasiar, J. Kumar Basak, B. Dey, K. Pal and K. Pal, Time evolution of spread complexity in quenched Lipkin–Meshkov–Glick model, J. Stat. Mech. 2310 (2023) 103101 [2208.10520].
  • (34) C. Tan, Z. Wei and R. Zhang, Scaling Relations of Spectrum Form Factor and Krylov Complexity at Finite Temperature, 2401.10499.
  • (35) T. Li and L.-H. Liu, Inflationary complexity of thermal state, 2405.01433.
  • (36) T. Li and L.-H. Liu, Inflationary Krylov complexity, JHEP 04 (2024) 123 [2401.09307].
  • (37) N. Vardian, Krylov complexity for 1-matric quantum mechanics, 2407.00155.
  • (38) H.A. Camargo, V. Jahnke, H.-S. Jeong, K.-Y. Kim and M. Nishida, Spectral and Krylov complexity in billiard systems, Phys. Rev. D 109 (2024) 046017 [2306.11632].
  • (39) K.-B. Huh, H.-S. Jeong and J.F. Pedraza, Spread complexity in saddle-dominated scrambling, JHEP 05 (2024) 137 [2312.12593].
  • (40) H.A. Camargo, K.-B. Huh, V. Jahnke, H.-S. Jeong, K.-Y. Kim and M. Nishida, Spread and Spectral Complexity in Quantum Spin Chains: from Integrability to Chaos, 2405.11254.
  • (41) S. He, P.H.C. Lau, Z.-Y. Xian and L. Zhao, Quantum chaos, scrambling and operator growth in TT¯T\overline{T} deformed SYK models, JHEP 12 (2022) 070 [2209.14936].
  • (42) D. Patramanis, Probing the entanglement of operator growth, PTEP 2022 (2022) 063A01 [2111.03424].
  • (43) P. Caputa, J.M. Magan and D. Patramanis, Geometry of Krylov complexity, Phys. Rev. Res. 4 (2022) 013041 [2109.03824].
  • (44) X. Cao, A statistical mechanism for operator growth, J. Phys. A 54 (2021) 144001 [2012.06544].
  • (45) F.B. Trigueros and C.-J. Lin, Krylov complexity of many-body localization: Operator localization in Krylov basis, SciPost Phys. 13 (2022) 037 [2112.04722].
  • (46) R. Heveling, J. Wang and J. Gemmer, Numerically probing the universal operator growth hypothesis, Phys. Rev. E 106 (2022) 014152 [2203.00533].
  • (47) P. Caputa and S. Liu, Quantum complexity and topological phases of matter, Phys. Rev. B 106 (2022) 195125 [2205.05688].
  • (48) P. Nandy, A.S. Matsoukas-Roubeas, P. Martínez-Azcona, A. Dymarsky and A. del Campo, Quantum Dynamics in Krylov Space: Methods and Applications, 2405.09628.
  • (49) A. Wachter, Relativistic quantum mechanics, Springer (2011).
  • (50) J. Gram, Ueber die entwickelung reeller functionen in reihen mittelst der methode der kleinsten quadrate., Journal für die reine und angewandte Mathematik 95 (1883) 41.
  • (51) E. Schmidt, Zur theorie der linearen und nichtlinearen integralgleichungen: I. teil: Entwicklung willkürlicher funktionen nach systemen vorgeschriebener, Mathematische Annalen 63 (1907) 433.
  • (52) V. Viswanath and G. Müller, The recursion method: application to many body dynamics, vol. 23, Springer Science & Business Media (1994).
  • (53) J.I. Kapusta and C. Gale, Finite-temperature field theory: Principles and applications, Cambridge Monographs on Mathematical Physics, Cambridge University Press (2011), 10.1017/CBO9780511535130.
  • (54) J.L.F. Barbón, E. Rabinovici, R. Shir and R. Sinha, On The Evolution Of Operator Complexity Beyond Scrambling, JHEP 10 (2019) 264 [1907.05393].
  • (55) E.W. Weisstein, “Lerch transcendent. from mathworld–a wolfram web resource.”
  • (56) Z.-Y. Fan, Universal relation for operator complexity, Phys. Rev. A 105 (2022) 062210 [2202.07220].
  • (57) M. Laine and A. Vuorinen, Basics of Thermal Field Theory, vol. 925, Springer (2016), 10.1007/978-3-319-31933-9, [1701.01554].
  • (58) L. Dolan and R. Jackiw, Symmetry Behavior at Finite Temperature, Phys. Rev. D 9 (1974) 3320.
  • (59) M.E. Peskin and D.V. Schroeder, An Introduction to quantum field theory, Addison-Wesley, Reading, USA (1995).