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

11institutetext: K.Sakuraba, S.Mori 22institutetext: Department of Mathematics and Physics, Graduate School of Science and Technology, Hirosaki University
Bunkyo-cho 3, Hirosaki, Aomori 036-8561, Japan
W. Kurebayashi
33institutetext: Department of Mechanical Science and Engineering, Graduate School of Science and Technology, Hirosaki University
Bunkyo-cho 3, Hirosaki, Aomori 036-8561, Japan
M. Hisakado
44institutetext: Nomura Holdings, Inc., Otemachi 2-2-2, Chiyoda-ku, Tokyo 100-8130, Japan

Self-exciting negative binomial distribution process and critical properties of intensity distribution

Kotaro Sakuraba    Wataru Kurebayashi    Masato Hisakado    Shintaro Mori
(Received: date / Accepted: date)
Abstract

We study the continuous time limit of a self-exciting negative binomial process and discuss the critical properties of its intensity distribution. In this limit, the process transforms into a marked Hawkes process. The probability mass function of the marks has a parameter ω\omega, and the process reduces to a ”pure” Hawkes process in the limit ω0\omega\to 0. We investigate the Lagrange–Charpit equations for the master equations of the marked Hawkes process in the Laplace representation close to its critical point and extend the previous findings on the power-law scaling of the probability density function (PDF) of intensities in the intermediate asymptotic regime to the case where the memory kernel is the superposition of an arbitrary finite number of exponentials. We develop an efficient sampling method for the marked Hawkes process based on the time-rescaling theorem and verify the power-law exponents.

1 Introduction

Recently, high-frequency financial data have become available, and many studies have been devoted to calibrating models of market micro-structure(Bacry et al. 2015). Initially, many studies adopted discrete time models that captured trade dynamics at regular time intervals or through a sequence of discrete time events, such as trades. Subsequently, the framework of the continuous time model—the point process— has been gradually applied to model financial data at the transaction level (Hasbrouck 1991, Engle and Russell 1998, Bowsher 2007, Hawkes 1971a, b, Hawkes and Oakes 1974, Hawkes 2018, Filimonov and Sornette 2012, 2015, Wheatley et al. 2019).

A point or counting process is characterized by its ”intensity” function, which represents the conditional probability of an event occurring in the immediate future. The Hawkes process, introduced by A. G. Hawkes (Hawkes 1971a, b, Hawkes and Oakes 1974), is a type of point process originally developed to model the occurrence of seismic events. Owing to its simplicity and flexibility, this model is being increasingly used in high-frequency finance (Bacry et al. 2015, Kirchner 2017, Blanc et al. 2017, Errais et al. 2010). It can easily capture the interactions between different types of events, incorporate the influence of intensive factors through marks, and accommodate non-stationarities.

The non-stationarity of the Hawkes process arise from the reproduction of an event induced by the interaction terms described by the kernel function that estimate the effect of the past events based on the elapsed time from them. The intensity of the process is determined by summing the effects of all past events through the kernel function. Additionally, by introducing a real number ”mark” that represents the strength of event effects, both the marks and the kernel values can be considered in the intensity estimation. If the average number of events triggered by one event, known as the branching ratio, exceeds one, the system becomes non-stationary. If the branching ratio approaches one from below, the system becomes critical, leading to power-law scaling in the PDF of the intensities in the intermediate asymptotic region. In the sub-critical case, the distribution decays exponentially beyond the characteristic intensity, which diverges as the branching ratio approaches one (Kanazawa and Sornette 2020b, a).

The discrete time self-exciting negative binomial distribution (DT-SE-NBD) was proposed for the modeling of time-series data related to defaults (Hisakado and Mori 2020). This model incorporates a parameter ω\omega that captures the correlation of events within the same term, and the process converges to a discrete-time self-exciting Poisson process in the limit ω0\omega\to 0. In the context of time-series default data, defaults are typically recorded on a yearly or quarterly basis, and defaults within the same period exhibit high correlation. Consequently, the probability distribution of the number of defaults displays overdispersion, where the variance is significantly larger than the mean value. To accurately capture this overdispersion, the negative binomial distribution (NBD), which has two parameters to control both the mean and variance, is more suitable than the Poisson distribution (Hisakado et al. 2022a, b). In the continuous time limit, the DT-SE-NBD process transforms into a point process with marks. As ω\omega tends to zero, the marked point process becomes a ”pure” Hawkes process as the value of the marks equals one. The PDF of the intensities also exhibits power-law scaling and exponentially decays beyond the characteristic intensity at the critical point and in the subcritical case, respectively.

This study extends the results of the SE-NBD process with a single exponential kernel to the case where the memory kernel is the sum of an arbitrary finite number of exponentials. Empirically, power-law memory kernels have been observed in studies on financial transaction data(Bacry et al. 2015). The theoretical analysis of Hawkes processes with power-law memory kernels requires decomposing the power-law memory kernel into a superposition of exponential kernels with weights following the inverse gamma distribution. By considering the case of multiple exponential kernels, we can estimate the power-law exponent of the intensity distribution for the marked Hawkes process with a power-law memory kernel. Additionally, we validate the theoretical predictions for the power-law behavior of the intensities through numerical simulations. The remaining sections of this paper are organized as follows: In Section 2, we introduce an SE-NBD process with an arbitrary finite number of exponentials and review related results on the process as well as on the Hawkes process. Section 3 focuses on studying the multivariate stochastic differential equation for the process, providing the solution to the two-variate master equations with a double exponential memory kernel, and deriving the PDF of the intensities. Furthermore, we discuss the PDF of the intensities for the case where the memory kernel is a sum of finite KK exponential functions. In Section 4, we verify the theoretical predictions for the power-law exponents of the PDF of the intensities at the critical point through numerical simulations. Finally, we present our conclusions in Section 5.

2 Model

We consider a DT-SE-NBD process {Xt},t=1,\{X_{t}\},t=1,\cdots. The variable Xt{0,1,}X_{t}\in\{0,1,\cdots\} represents the size of the event at time tt. In the context of modeling of time-series default data, XtX_{t} specifically represents the number of defaults that occur in the tt-th period(Hisakado et al. 2022a, b). XtX_{t} obeys NBD for the condition λ^t=λt\hat{\lambda}_{t}=\lambda_{t} as

Xt+1\displaystyle X_{t+1} \displaystyle\sim NBD(α=λ^tω,p=1ω+1),t0\displaystyle\mbox{NBD}\left(\alpha=\frac{\hat{\lambda}_{t}}{\omega},p=\frac{1}{\omega+1}\right),t\geq 0 (1)
λ^t\displaystyle\hat{\lambda}_{t} =\displaystyle= ν0+ns=1thtsXs,t1,λ^0=ν0\displaystyle\nu_{0}+n\sum_{s=1}^{t}h_{t-s}X_{s},t\geq 1\,\,,\,\,\hat{\lambda}_{0}=\nu_{0} (2)
ht\displaystyle h_{t} =\displaystyle= 1nk=1Knkhtk=1nk=1Knk(1e1/τk)et/τk,n=k=1Knk.\displaystyle\frac{1}{n}\sum_{k=1}^{K}n_{k}h^{k}_{t}=\frac{1}{n}\sum_{k=1}^{K}n_{k}(1-e^{-1/\tau_{k}})e^{-t/\tau_{k}},n=\sum_{k=1}^{K}n_{k}. (3)

Here, α,p\alpha,p are the parameters of the NBD. α>0\alpha>0 represents the number of successes until the experiment is terminated, and p(0,1]p\in(0,1] is the probability of success in each individual experiment. The sequence {ht}t=0,1,\{h_{t}\}_{t=0,1,\cdots} refers to the discount factor (memory kernel), which is assumed to be normal; specifically, t=0ht=1\sum_{t=0}^{\infty}h_{t}=1. The prefactor (1e/τk)(1-e^{-/\tau_{k}}) is introduced to ensure the normalization of the kk-th exponential term htk=(1e1/τk)et/τkh^{k}_{t}=(1-e^{-1/\tau_{k}})e^{-t/\tau_{k}}, where τk\tau_{k} represents the memory length associated with the kk-th term. Consequently, t=0(1e1/τk)et/τk=1\sum_{t=0}^{\infty}(1-e^{-1/\tau_{k}})e^{-t/\tau_{k}}=1. The set of coefficients {nk}k=1,,K\{n_{k}\}_{k=1,\cdots,K} quantifies the contribution of the kk-th exponential term to the memory kernel ht,t=0,1h_{t},t=0,1\cdots, with nn representing their sum.

By considering a double-scaling limit, the DT-SE-NBD process can be constructed as an extension of the multi-term Pólya urn process with a memory kernel as follows:

XtlimN,n0n0/N=1BBD(N,α=λ^tω,β=n0λ^tω)=NBD(α=λ^tω,p=1ω+1).X_{t}\sim\lim_{\begin{subarray}{c}N,n_{0}\to\infty\\ n_{0}/N=1\end{subarray}}\mbox{BBD}\left(N,\alpha=\frac{\hat{\lambda}_{t}}{\omega},\beta=\frac{n_{0}-\hat{\lambda}_{t}}{\omega}\right)=\mbox{NBD}\left(\alpha=\frac{\hat{\lambda}_{t}}{\omega},p=\frac{1}{\omega+1}\right).

Here, BBD represents the beta-binomial distribution, and the variables N,λ^t,n0λ^tN,\hat{\lambda}_{t},n_{0}-\hat{\lambda}_{t} represent the number of trials, the number of red balls, and the number of blue balls in an urn in the tt-th term, respectively. In each term, we sequentially remove a ball and return it along with ω\omega additional balls of the same color. Consequently, the total number of balls in the urn increases by ω\omega. The process is repeated NN times within one term. The term 1/(1+n0/ω)1/(1+n_{0}/\omega) represents Pearson’s correlation coefficient, , which reflects the correlation between the color choices of the balls in the process (Hisakado et al. 2006). The feedback term hth_{t} induces an intertemporal correlation of the ball choices. Furthermore, as ω\omega approaches zero, Xt+1X_{t+1} under the condition λ^t=λt\hat{\lambda}_{t}=\lambda_{t} follows a Poisson random variable with a mean of λt\lambda_{t}, thereby transforming the DT-SE-NBD process into a DT-SE Poisson process.

Xt+1Po(λ^t),t0X_{t+1}\sim\mbox{Po}(\hat{\lambda}_{t}),t\geq 0

Fig.1 illustrates the transformation from the multi-term Pòlya urn process to the DT-SE Negative binomial and DT-SE Poisson process.

Refer to caption
Figure 1: Transformation of multi-term Pòlya urn process (beta-binomial) to discrete-time Negative binomial and Poisson process.

We study the (unconditional) expected value of Xt+1X_{t+1}. As Xt+1NBD(λtω,1ω+1)X_{t+1}\sim\mbox{NBD}\left(\frac{\lambda_{t}}{\omega},\frac{1}{\omega+1}\right) for λ^t=λt\hat{\lambda}_{t}=\lambda_{t}, we have

E[Xt+1]=Eλt[E[Xt+1|λ^t=λt]]=Eλt[λ^t].E[X_{t+1}]=E_{\lambda_{t}}[E[X_{t+1}|\hat{\lambda}_{t}=\lambda_{t}]]=E_{\lambda_{t}}[\hat{\lambda}_{t}].

where Eλt[]E_{\lambda_{t}}[\,\,\,\,] is the ensemble average of the stochastic process {Xs,st}\{X_{s},s\leq t\} [i.e., Filtration FtF_{t}]. We are now interested in the steady state of the process and write the unconditional expected value of XtX_{t} as μ\mu. We have the following relationship:

μlimtEλt[λ^t].\mu\equiv\lim_{t\to\infty}E_{\lambda_{t}}[\hat{\lambda}_{t}].

With λt=ν0+ns=1tXshts\lambda_{t}=\nu_{0}+n\sum_{s=1}^{t}X_{s}h_{t-s},E[Xs]=μE[X_{s}]=\mu and the normalization of the discount factor hth_{t}, we have

μ=ν0+nμ.\mu=\nu_{0}+n\mu.

We then obtain,

μ=ν01n\mu=\frac{\nu_{0}}{1-n}

The phase transition between the steady and non-steady state occurs at n=nc=1n=n_{c}=1, which is the critical point. This is the same as that in the Hawkes process. In this sense, nn corresponds to the branching ratio of the Hawkes process, and the steady-state condition is n<1n<1.

We proceed to the continuous time SE-NBD process. We begin with the DT-SE-NBD process in (1), (2), and (3). The stochastic process {Xt},t=1,\{X_{t}\},t=1,\cdots is non-Markovian, so we focus on the time evolution of z^tλ^tν0\hat{z}_{t}\equiv\hat{\lambda}_{t}-\nu_{0}. We decompose z^t\hat{z}_{t} as follows:

z^t\displaystyle\hat{z}_{t} \displaystyle\equiv λ^tν0=kz^tk\displaystyle\hat{\lambda}_{t}-\nu_{0}=\sum_{k}\hat{z}^{k}_{t}
z^tk\displaystyle\hat{z}^{k}_{t} \displaystyle\equiv nks=1thtskXs,z^0k=0.\displaystyle n_{k}\sum_{s=1}^{t}h^{k}_{t-s}X_{s}\,\,,\,\,\ \hat{z}_{0}^{k}=0\color[rgb]{0,0,0}.

z^tk\hat{z}^{k}_{t} satisfies the following recursive equation:

z^t+1k=e1/τkz^tk+nkh0kXt+1.\hat{z}^{k}_{t+1}=e^{-1/\tau_{k}}\hat{z}^{k}_{t}+n_{k}h^{k}_{0}X_{t+1}.

Here, we use the relation s=1t+1Xsht+1sk=Xt+1h0k+e1/τks=1tXshtsk\sum_{s=1}^{t+1}X_{s}h^{k}_{t+1-s}=X_{t+1}h^{k}_{0}+e^{-1/\tau_{k}}\sum_{s=1}^{t}X_{s}h^{k}_{t-s}. The stochastic difference equation for z^tk\hat{z}^{k}_{t} is

Δz^tkz^t+1kz^tk=(e1/τk1)z^tk+nkh0kXt+1,k=1,,K.\Delta\hat{z}^{k}_{t}\equiv\hat{z}^{k}_{t+1}-\hat{z}^{k}_{t}=(e^{-1/\tau_{k}}-1)\hat{z}^{k}_{t}+n_{k}h^{k}_{0}X_{t+1},k=1,\cdots,K. (4)

The DT-SE-NBD process is now represented by a multivariate stochastic difference equation.

Now, we consider the continuous time limit. We divide the unit time interval by infinitesimal time intervals with width dtdt. The decreasing factor e1/τke^{-1/\tau_{k}} for unit time period is replaced by edt/τk1dt/τk+o(dt/τk)e^{-dt/\tau_{k}}\simeq 1-dt/\tau_{k}+o(dt/\tau_{k}) for dtdt. In addition, we adopt the notation hk(t)=1τket/τkh_{k}(t)=\frac{1}{\tau_{k}}e^{-t/\tau_{k}}. Here, we replace the normalization factor (1e1/τk)(1-e^{-1/\tau_{k}}) of htkh^{k}_{t} with 1/τk1/\tau_{k} in hk(t)h_{k}(t) to ensure that 0hk(t)𝑑t=1τk0et/τk=1\int_{0}^{\infty}h_{k}(t)dt=\frac{1}{\tau_{k}}\int_{0}^{\infty}e^{-t/\tau_{k}}=1. We introduce the continuous time memory kernel h(t)h(t) as

nh(t)=k=1Knkhk(t)=k=1Knk1τket/τk.nh(t)=\sum_{k=1}^{K}n_{k}h_{k}(t)=\sum_{k=1}^{K}n_{k}\frac{1}{\tau_{k}}e^{-t/\tau_{k}}. (5)

We denote the continuous time limits of z^tk\hat{z}^{k}_{t} and λ^t\hat{\lambda}_{t} as z^k(t)\hat{z}_{k}(t) and λ^(t)\hat{\lambda}(t), respectively. The timing, size of events, and counting process are denoted by t^i\hat{t}_{i}, mim_{i} and N^(t)\hat{N}(t), respectively. z^k(t)\hat{z}_{k}(t) and λ^(t)\hat{\lambda}(t) are defined as

z^k(t)\displaystyle\hat{z}_{k}(t) =\displaystyle= nki=1N^(t)hk(tt^i)mi\displaystyle n_{k}\sum_{i=1}^{\hat{N}(t)}h_{k}(t-\hat{t}_{i})\cdot m_{i}
λ^(t)\displaystyle\hat{\lambda}(t) =\displaystyle= ν0+k=1Kz^k(t)\displaystyle\nu_{0}+\sum_{k=1}^{K}\hat{z}_{k}(t)

XtX_{t} is the common noise for the time interval [t,t+1)[t,t+1) in the DT SE-NBD process. To take the continuous time limit, it is necessary to define the noise dξ(t)d\xi(t) for the infinitesimal interval [t,t+dt)[t,t+dt). There is freedom in defining dξ(t)d\xi(t), and we adopt the following, which is based on the reproduction properties of NBD. When X1NBD(α1,p),X2NBD(α2,p)X_{1}\sim NBD(\alpha_{1},p),X_{2}\sim\mbox{NBD}(\alpha_{2},p),X1+X2NBD(α1+α2,p)X_{1}+X_{2}\sim\mbox{NBD}(\alpha_{1}+\alpha_{2},p). We define dξ(α,p)NBD(t)d\xi^{NBD}_{(\alpha,p)}(t) for the interval [t,t+dt)[t,t+dt) as

dξ(α,p)NBD(t)NBD(αdt,p).d\xi^{NBD}_{(\alpha,p)}(t)\sim\mbox{NBD}(\alpha dt,p).

If we consider the above-mentioned noise, by the reproduction property, the noise for the interval [t,t+1)[t,t+1) becomes

tt+1𝑑ξ(α,p)NBD(s)NBD(α,p).\int_{t}^{t+1}d\xi^{NBD}_{(\alpha,p)}(s)\sim\mbox{NBD}(\alpha,p).

In the self-exciting process, we replace α\alpha and pp by λ^(t)ω\frac{\hat{\lambda}(t)}{\omega} and 1ω+1\frac{1}{\omega+1}, respectively. The conditional expected value and variance of dξ(α=λ^(t)ωdt,p=1ω+1)NBDd\xi^{NBD}_{(\alpha=\frac{\hat{\lambda}(t)}{\omega}dt,p=\frac{1}{\omega+1})} are

E[dξ(λ^(t)ωdt,1ω+1)NBD|λ^(t)=λ(t)]=λ(t)dt\displaystyle E\left[d\xi^{NBD}_{(\frac{\hat{\lambda}(t)}{\omega}dt,\frac{1}{\omega+1})}\right|\left.\hat{\lambda}(t)=\lambda(t)\right]=\lambda(t)dt
V[dξ(λ^(t)ω,1ω+1)NBD|λ^(t)=λ(t)]=λ(t)(ω+1)dt.\displaystyle V\left[d\xi^{NBD}_{(\frac{\hat{\lambda}(t)}{\omega},\frac{1}{\omega+1})}\right|\left.\hat{\lambda}(t)=\lambda(t)\right]=\lambda(t)(\omega+1)dt.

The conditional probability of occurrence of an event with size mm during the time interval [t,t+dt)[t,t+dt) is given as

P(dξ(λ^(t)ω,1ω+1)NBD=m|λ^(t)=λ(t))={1λ(t)ωlog(ω+1)dtm=01m(λ(t)ω)(ωω+1)mdtm1.P\left(d\xi^{NBD}_{(\frac{\hat{\lambda}(t)}{\omega},\frac{1}{\omega+1})}=m\right|\left.\hat{\lambda}(t)=\lambda(t)\right)=\left\{\begin{array}[]{cc}1-\frac{\lambda(t)}{\omega}\log(\omega+1)dt&m=0\\ \frac{1}{m}\left(\frac{\lambda(t)}{\omega}\right)\left(\frac{\color[rgb]{1,0,0}\omega\color[rgb]{0,0,0}}{\omega+1}\right)^{m}dt&m\geq 1.\end{array}\right.

In the limit ω0\omega\to 0, the probabilities converge to

limω0P(dξ(λ^(t)ω,1ω+1)NBD=m|λ^(t)=λ(t))={1λ(t)dtm=0λ(t)dtm=1o(dt)m2,\lim_{\omega\to 0}P\left(d\xi^{NBD}_{(\frac{\hat{\lambda}(t)}{\omega},\frac{1}{\omega+1})}=m\right|\left.\hat{\lambda}(t)=\lambda(t)\right)=\left\{\begin{array}[]{cc}1-\lambda(t)dt&m=0\\ \lambda(t)dt&m=1\\ o(dt)&m\geq 2,\end{array}\right.

and the event size mm is restricted to one.

The NBD noise ξNBD(t)\xi^{NBD}(t) is then written as

ξ(λ^(t)ω,1ω+1)NBD(t)=i=1N^(t)miδ(tt^i).\xi^{NBD}_{\left(\frac{\hat{\lambda}(t)}{\omega},\frac{1}{\omega+1}\right)}(t)=\sum_{i=1}^{\hat{N}(t)}m_{i}\delta(t-\hat{t}_{i}).

The probability mass function (PMF) for event size m=1,m=1,\cdots is given by

ρ(m)=P(dξ(λ^(t)ω,1ω+1)NBD=m)P(dξNBD>0)=1mlog(ω+1)(ωω+1)m.\rho(m)=\frac{P\left(d\xi^{NBD}_{(\frac{\hat{\lambda}(t)}{\omega},\frac{1}{\omega+1})}=m\right)}{P(d\xi^{NBD}>0)}=\frac{1}{m\log(\omega+1)}\left(\frac{\color[rgb]{1,0,0}\omega\color[rgb]{0,0,0}}{\omega+1}\right)^{m}. (6)

The intensity of the process, P(dξNBD>0)/dtP(d\xi^{NBD}>0)/dt, is then given as

ν^(t)log(ω+1)ωλ^(t),λ^(t)=ν0+i=1N^(t)k=1Knkhk(tt^i)mi.\hat{\nu}(t)\equiv\frac{\log(\omega+1)}{\omega}\hat{\lambda}(t)\,\,,\,\,\hat{\lambda}(t)=\nu_{0}+\sum_{i=1}^{\hat{N}(t)}\sum_{k=1}^{K}n_{k}h_{k}(t-\hat{t}_{i})\cdot m_{i}. (7)

This process is known as the marked Hawkes process. Each event has a distinct influential power that is encoded in the marks {mi}i=1,,N^(t)\{m_{i}\}_{i=1,\cdots,\hat{N}(t)}. {mi}i=1,,N^(t)\{m_{i}\}_{i=1,\cdots,\hat{N}(t)} are IID random numbers and obey the PMF of (6). When the parameter ω\omega tends to zero (ω0\omega\to 0), the PMF ρ(m)\rho(m) approaches the delta function δ(m1)\delta(m-1), and the marked Hawkes process reduces to the “pure” Hawkes process. In the following analysis, we will focus on studying the PDF of the intensities in the marked Hawkes process.

3 Solution

The multivariate difference equation (4) can be transformed into a multivariate SDE as follows:

z^k(t)\displaystyle\hat{z}_{k}(t) =\displaystyle= 1τkz^k(t)dt+nkτkdξ(λ^(t)ω,1ω+1)NBD,k=1,,K.\displaystyle-\frac{1}{\tau_{k}}\hat{z}_{k}(t)dt+\frac{n_{k}}{\tau_{k}}d\xi^{NBD}_{(\frac{\hat{\lambda}(t)}{\omega},\frac{1}{\omega+1})},k=1,\cdots,K. (8)
λ^(t)\displaystyle\hat{\lambda}(t) =\displaystyle= ν0+k=1Kz^k(t)\displaystyle\nu_{0}+\sum_{k=1}^{K}\hat{z}_{k}(t) (9)
ξ(λ^(t)ω,1ω+1)NBD(t)\displaystyle\xi^{NBD}_{\left(\frac{\hat{\lambda}(t)}{\omega},\frac{1}{\omega+1}\right)}(t) =\displaystyle= i=1N^(t)miδ(tt^i).\displaystyle\sum_{i=1}^{\hat{N}(t)}m_{i}\delta(t-\hat{t}_{i}). (10)

Note that the same state-dependent NBD noise ξ(λ^ω,1ω+1)NBD\xi^{\mathrm{NBD}}_{(\frac{\hat{\lambda}}{\omega},\frac{1}{\omega+1})} affects every component of the multivariate SDE {z^k}k=1,K\{\hat{z}_{k}\}_{k=1,...K}. In other words, each shock event simultaneously affects the trajectories for all excess intensities {z^k}k=1,K\{\hat{z}_{k}\}_{k=1,...K}.

The formal solution of the SDE is

z^k(t)=nkτk0t𝑑te(tt)/τkξ(λ^ω,1ω+1)NBD(t).\hat{z}_{k}(t)=\frac{n_{k}}{\tau_{k}}\int^{t}_{0}dt^{\prime}e^{-(t-t^{\prime})/\tau_{k}}\xi^{\mathrm{NBD}}_{(\frac{\hat{\lambda}}{\omega},\frac{1}{\omega+1})}(t^{\prime}).

The SDE(8) can be interpreted as

z^k(t+dt)z^k(t)={z^k(t)τkdtProb.=1λ^(t)ωlog(ω+1)dtnkmτkProb.=1m(λ(t)ω)(ωω+1)mdt,m1.\hat{z}_{k}(t+dt)-\hat{z}_{k}(t)=\left\{\begin{aligned} -\frac{\hat{z}_{k}(t)}{\tau_{k}}dt\qquad&\mathrm{Prob.}=1-\frac{\hat{\lambda}(t)}{\omega}\log(\omega+1)dt\\ \frac{n_{k}m}{\tau_{k}}\hskip 25.0pt&\mathrm{Prob.}=\frac{1}{m}\left(\frac{\lambda(t)}{\omega}\right)\left(\frac{\color[rgb]{1,0,0}\omega\color[rgb]{0,0,0}}{\omega+1}\right)^{m}dt,m\geq 1.\end{aligned}\right.

We adopted the same procedure to derive the master equation for the PDF of z^k\hat{z}_{k} in Ref.(Kanazawa and Sornette 2020b, a). As the SDEs for 𝒛:=(z1,,zK)\boldsymbol{z}:=(z_{1},...,z_{K}) are standard Markovian stochastic processes, we obtain the corresponding master equation

tPt\displaystyle\frac{\partial}{\partial t}P_{t} (𝒛)=k=1KzkzkτkPt(𝒛)+m=11mω(ωω+1)m\displaystyle(\boldsymbol{z})=\sum^{K}_{k=1}\frac{\partial}{\partial z_{k}}\frac{z_{k}}{\tau_{k}}P_{t}(\boldsymbol{z})+\sum_{m=1}^{\infty}\frac{1}{m\omega}\left(\frac{\omega}{\omega+1}\right)^{m}
×{[ν0+k=1K(zknkmτk)]Pt(𝒛𝒉)[ν0+k=1Kzk]Pt(𝒛)}\displaystyle\times\left\{\left[\nu_{0}+\sum^{K}_{k=1}\left(z_{k}-\frac{n_{k}m}{\tau_{k}}\right)\right]P_{t}(\boldsymbol{z-h})-\left[\nu_{0}+\sum^{K}_{k=1}z_{k}\right]P_{t}(\boldsymbol{z})\right\} (11)

The jump-size vector is given by 𝒉:=(n1m/τ1,,nKm/τK)\boldsymbol{h}:=(n_{1}m/\tau_{1},...,n_{K}m/\tau_{K}).

The master equation (3) takes a simplified form under the Laplace representation:

P~t(𝒔):=K[Pt(𝒛);𝒔],\tilde{P}_{t}(\boldsymbol{s}):=\mathcal{L}_{\color[rgb]{1,0,0}K\color[rgb]{0,0,0}}[P_{t}(\boldsymbol{z});\boldsymbol{s}], (12)

where the Laplace transformation in the KK-dimensional space is defined as

K[Pt(𝒛);𝒔]:=0𝑑𝒛e𝒔𝒛Pt(𝒛)\mathcal{L}_{K}[P_{t}(\boldsymbol{z});\boldsymbol{s}]:=\int^{\infty}_{0}d\boldsymbol{z}e^{-\boldsymbol{s\cdot z}}P_{t}(\boldsymbol{z}) (13)

with the volume element d𝒛:=k=1Kdzkd\boldsymbol{z}:=\prod^{K}_{k=1}dz_{k}. The wave vector 𝒔:=(s1,,sK)\boldsymbol{s}:=(s_{1},...,s_{K}) is the conjugate of the excess intensity vector 𝒛:=(z1,,zK)\boldsymbol{z}:=(z_{1},...,z_{K}).

The Laplace representation of the master equation (3) is given by

P~t(𝒔)t=k=1KskτkP~t(𝒔)sk+m=11mω(ωω+1)m(ν0k=1Ksk)(e𝒉𝒔1)P~t(𝒔)\frac{\partial\tilde{P}_{t}(\boldsymbol{s})}{\partial t}=-\sum^{K}_{k=1}\frac{s_{k}}{\tau_{k}}\frac{\partial\tilde{P}_{t}(\boldsymbol{s})}{\partial s_{k}}+\sum^{\infty}_{m=1}\frac{1}{m\omega}\left(\frac{\omega}{\omega+1}\right)^{m}\left(\nu_{0}-\sum^{K}_{k=1}\frac{\partial}{\partial s_{k}}\right)(e^{-\boldsymbol{h\cdot s}}-1)\tilde{P}_{t}(\boldsymbol{s}) (14)

Then, the Laplace representation (12) of Pt(𝒛)P_{t}(\boldsymbol{z}), which is the solution of (14), enables us to obtain the (one-dimensional) Laplace representation Q~t(s)\tilde{Q}_{t}(s) of the intensity PDFPt(ν)\mathrm{PDF}\ P_{t}(\nu) according to

Q~t(s):=[Pt(ν);s]=eν0sP~t(𝒔=(s,s,,)).\tilde{Q}_{t}(s):=\mathcal{L}[P_{t}(\nu);s]=e^{-\nu_{0}s}\tilde{P}_{t}(\boldsymbol{s}=(s,s,...,)). (15)

A. Single exponential kernel

We now consider the case in which the memory function (5) consists of K=1K=1 exponential functions(Hisakado et al. 2022a). The Laplace representation of the master equation (3) is given by

dP~t(s)dt=sτdP~t(s)ds+m=11mω(ωω+1)m(λ0dds)(enkτs1)P~t(s).\frac{d\tilde{P}_{t}(s)}{dt}=-\frac{s}{\tau}\frac{d\tilde{P}_{t}(s)}{ds}+\sum^{\infty}_{m=1}\frac{1}{m\omega}\left(\frac{\omega}{\omega+1}\right)^{m}\left(\lambda_{0}-\frac{d}{ds}\right)\left(e^{-\frac{nk}{\tau}s}-1\right)\tilde{P}_{t}(s).

The steady-state PDF of the intensity λ^(t)\hat{\lambda}(t) is

PSS(λ)λ1+2ν0τω+1e2τϵω+1λ,ϵ=1n.P_{SS}(\lambda)\propto\lambda^{-1+2\frac{\nu_{0}\tau}{\omega+1}}e^{-\frac{2\tau\epsilon}{\omega+1}\lambda}\,\,,\,\,\epsilon=1-n. (16)

The power-law exponent of the PDF of the excess intensity is 12ν0τω+11-\frac{2\nu_{0}\tau}{\omega+1} and depends on ω\omega. In the limit ω0\omega\rightarrow 0, the result coincides with that in Ref.(Kanazawa and Sornette 2020b, a). The power-law exponent increases with ω\omega and converges to 11 in the limit ω\omega\rightarrow\infty. In addition, the length scale beyond which the intensity shows exponential decay for the off-critical case is (ω+1)/(2τϵ)(\omega+1)/(2\tau\epsilon), and it diverges in the limit ϵ=1n0\epsilon=1-n\to 0. This is also an increasing function of ω\omega.

B. Double exponential kernel

We consider the case where the memory function (5) consists of K=2K=2 exponential functions. To derive the solution for the Laplace representation of the master equation (14), we apply the method of characteristics as in Ref.(Kanazawa and Sornette 2020a). One can find a brief explanation in Refs.(Gardiner 2009) and (Kanazawa and Sornette 2020a). In appendix A, we provide a brief review.

We start from the Lagrange–Charpit equations, which are given by

ds1dl\displaystyle\frac{ds_{1}}{dl} =\displaystyle= m=11mω(ωω+1)m(e𝒉𝒔1)s1τ1,\displaystyle-\sum^{\infty}_{m=1}\frac{1}{m\omega}\left(\frac{\omega}{\omega+1}\right)^{m}\left(e^{-\boldsymbol{h\cdot s}}-1\right)-\frac{s_{1}}{\tau_{1}}, (17)
ds2dl\displaystyle\frac{ds_{2}}{dl} =\displaystyle= m=11mω(ωω+1)m(e𝒉𝒔1)s2τ2,\displaystyle-\sum^{\infty}_{m=1}\frac{1}{m\omega}\left(\frac{\omega}{\omega+1}\right)^{m}\left(e^{-\boldsymbol{h\cdot s}}-1\right)-\frac{s_{2}}{\tau_{2}}, (18)
ddllogP~ss\displaystyle\frac{d}{dl}\log\tilde{P}_{ss} =\displaystyle= m=11mω(ωω+1)mν0(e𝒉𝒔1)\displaystyle-\sum^{\infty}_{m=1}\frac{1}{m\omega}\left(\frac{\omega}{\omega+1}\right)^{m}\nu_{0}\left(e^{-\boldsymbol{h\cdot s}}-1\right) (19)

and ll is the auxiliary “time” parameterizing the position on the characteristic curve. Let us develop the stability analysis around s=0s=0 (i.e., for large λs\lambda^{\prime}s) for this pseudo-dynamical system.

a. Sub-critical case n<1n<1

Assuming n:=n1+n2<1n:=n_{1}+n_{2}<1, we first expand e𝒉𝒔e^{-\boldsymbol{h\cdot s}} to compute the sum of mm:

e(n1s1τ1+n2s2τ2)m1(n1s1τ1+n2s2τ2)m+12(n1s1τ1+n2s2τ2)2m2+e^{-\left(\frac{n_{1}s_{1}}{\tau_{1}}+\frac{n_{2}s_{2}}{\tau_{2}}\right)m}\simeq 1-\left(\frac{n_{1}s_{1}}{\tau_{1}}+\frac{n_{2}s_{2}}{\tau_{2}}\right)m+\frac{1}{2}\left(\frac{n_{1}s_{1}}{\tau_{1}}+\frac{n_{2}s_{2}}{\tau_{2}}\right)^{2}m^{2}+\cdots (21)

We obtain the linearized dynamics of the system (17),(18),(19) as follows:

d𝒔dl𝑯𝒔,ddllogP~ssν0𝑲𝒔\frac{d\boldsymbol{s}}{dl}\simeq-\boldsymbol{H}\boldsymbol{s},\ \frac{d}{dl}\log\tilde{P}_{ss}\simeq\nu_{0}\boldsymbol{K}\boldsymbol{s} (22)

with

𝑯:=(1n1τ1n2τ2n1τ11n2τ2),𝑲:=(n1τ1,n2τ2).\boldsymbol{H}:=\begin{pmatrix}\frac{1-n_{1}}{\tau_{1}}&\frac{-n_{2}}{\tau_{2}}\\ \frac{-n_{1}}{\tau_{1}}&\frac{1-n_{2}}{\tau_{2}}\end{pmatrix},\ \boldsymbol{K}:=\left(\frac{n_{1}}{\tau_{1}},\frac{n_{2}}{\tau_{2}}\right).

We introduce the eigenvalues β1,β2\beta_{1},\beta_{2} and eigenvectors 𝒆1,𝒆2\boldsymbol{e}_{1},\boldsymbol{e}_{2} of 𝑯\boldsymbol{H} such that

𝑷:=(e1,e2),𝑷1𝑯𝑷=(β100β2).\boldsymbol{P}:=(e_{1},e_{2}),\ \boldsymbol{P}^{-1}\boldsymbol{H}\boldsymbol{P}=\begin{pmatrix}\beta_{1}&0\\ 0&\beta_{2}\end{pmatrix}.

The matrix 𝑯\boldsymbol{H} is the same with the results of Ref.(Kanazawa and Sornette 2020b, a) and all eigenvalues are real. We denote them as β1β2\beta_{1}\geq\beta_{2}. The determinant of 𝑯\boldsymbol{H} is given by

det𝑯=1nτ1τ2.\mathrm{det}\boldsymbol{H}=\frac{1-n}{\tau_{1}\tau_{2}}.

This implies that the zero eigenvalue β1=0\beta_{1}=0 appears at the critical point n=1n=1. Below the critical point n<1n<1, all eigenvalues are positive (β1,β2>0\beta_{1},\beta_{2}>0). For n<1n<1, the dynamics can be rewritten as

ddl𝑷1𝒔=(β100β2)𝑷1𝒔𝒔(l)=𝑷(eβ1(ll0)eβ2(ll0)/C1).\frac{d}{dl}\boldsymbol{P}^{-1}\boldsymbol{s}=-\begin{pmatrix}\beta_{1}&0\\ 0&\beta_{2}\end{pmatrix}\boldsymbol{P}^{-1}\boldsymbol{s}\Longrightarrow\boldsymbol{s}(l)=\boldsymbol{P}\begin{pmatrix}e^{-\beta_{1}(l-l_{0})}\\ e^{-\beta_{2}(l-l_{0})}/C_{1}\end{pmatrix}.

Here, l0l_{0} and C1C_{1} are integration constants. We can assume l0=0l_{0}=0 as the initial point of the characteristic curve without loss of generality. Integrating the second equation in (22), we obtain

logP~ss\displaystyle\log\tilde{P}_{ss} =\displaystyle= ν0𝑲𝒔(l)𝑑l+C2=ν0𝑲𝑷(1/β1001/β2)𝑷1𝒔+C2\displaystyle\nu_{0}\boldsymbol{K}\int\boldsymbol{s}(l)dl+C_{2}=-\nu_{0}\boldsymbol{KP}\begin{pmatrix}1/\beta_{1}&0\\ 0&1/\beta_{2}\end{pmatrix}\boldsymbol{P}^{-1}\boldsymbol{s}+C_{2}
=\displaystyle= ν𝑲𝑯1𝒔+C2.\displaystyle-\nu\boldsymbol{KH}^{-1}\boldsymbol{s}+C_{2}.

The general solution is given by

(C1)=C2\mathcal{H}(C_{1})=C_{2} (23)

with function \mathcal{H} determined by the initial condition of the characteristic curve. Let us introduce

𝒔¯:=𝑷1𝒔=(s¯1s¯2)C1=(s¯1)β2/β1(s¯2)1.\bar{\boldsymbol{s}}:=\boldsymbol{P}^{-1}\boldsymbol{s}=\begin{pmatrix}\bar{s}_{1}\\ \bar{s}_{2}\end{pmatrix}\Longrightarrow C_{1}=(\bar{s}_{1})^{\beta_{2}/\beta_{1}}(\bar{s}_{2})^{-1}.

This implies that the solution is given in the following form:

logP~ss(𝒔)=ν𝑲𝑯1𝒔+((s¯1)β2/β1(s¯2)1).\log\tilde{P}_{ss}(\boldsymbol{s})=-\nu\boldsymbol{KH}^{-1}\boldsymbol{s}+\mathcal{H}((\bar{s}_{1})^{\beta_{2}/\beta_{1}}(\bar{s}_{2})^{-1}).

Owing to the renormalization of the PDF, the relation

lim𝒔𝟎logP~ss(𝒔)=0\lim_{\boldsymbol{s\rightarrow 0}}\log\tilde{P}_{ss}(\boldsymbol{s})=0

must hold for any path in the (s1,s2s_{1},s_{2}) space ending at the origin (limit𝒔𝟎\mathrm{limit}\ \boldsymbol{s\rightarrow 0}). Let us consider a specific limit such that s¯2=x1(s¯1)β2/β1\bar{s}_{2}=x^{-1}(\bar{s}_{1})^{\beta_{2}/\beta_{1}} and s¯10\bar{s}_{1}\rightarrow 0 for an arbitrary positive xx.

lims¯10logP~ss(𝒔)=(x)\lim_{\bar{s}_{1}\rightarrow 0}\log\tilde{P}_{ss}(\boldsymbol{s})=\mathcal{H}(x)

Because the left-hand side is zero for any xx, the function ()\mathcal{H}(\cdot) must be exactly zero. Thus, this leads to

logP~ss(𝒔)=ν𝑲𝑯1𝒔.\log\tilde{P}_{ss}(\boldsymbol{s})=-\nu\boldsymbol{KH}^{-1}\boldsymbol{s}.

By substituting 𝒔=(s1=s,s2=s)\boldsymbol{s}=(s_{1}=s,s_{2}=s), we obtain

logQ~ss(𝒔)=ν0s+logP~t=(s(1,1))ν01ns.\log\tilde{Q}_{ss}(\boldsymbol{s})=-\nu_{0}s+\log\tilde{P}_{t=\infty}(s(1,1))\simeq-\frac{\nu_{0}}{1-n}s.

This is consistent with the asymptotic mean intensity in the steady state (Kanazawa and Sornette 2020b, a).

b. Critical case n=1n=1

In this case, the eigenvalues and eigenvectors of 𝑯\boldsymbol{H} are given by

β1=0,β2=n1τ1+n2τ2τ1τ2,𝒆1=(τ1τ2),𝒆2=(n2n1).\beta_{1}=0,\ \beta_{2}=\frac{n_{1}\tau_{1}+n_{2}\tau_{2}}{\tau_{1}\tau_{2}},\ \boldsymbol{e}_{1}=\begin{pmatrix}\tau_{1}\\ \tau_{2}\end{pmatrix},\ \boldsymbol{e}_{2}=\begin{pmatrix}-n_{2}\\ n_{1}\end{pmatrix}.

This means that the eigenvalue matrix and its inverse matrix are given by

𝑷=(τ1n2τ2n1),𝑷1=1α(n1n2τ2τ1),α:=det𝑷=τ1n1+τ2n2.\boldsymbol{P}=\begin{pmatrix}\tau_{1}&-n_{2}\\ \tau_{2}&n_{1}\end{pmatrix},\ \boldsymbol{P}^{-1}=\frac{1}{\alpha}\begin{pmatrix}n_{1}&n_{2}\\ -\tau_{2}&\tau_{1}\end{pmatrix},\ \alpha:=\det\boldsymbol{P}=\tau_{1}n_{1}+\tau_{2}n_{2}.

Here, let us introduce

𝑿=(XY)=𝑷1𝒔,X=n1s1+n2s2α,Y=τ2s1+τ1s2α.\boldsymbol{X}=\begin{pmatrix}X\\ Y\end{pmatrix}=\boldsymbol{P}^{-1}\boldsymbol{s},\Longleftrightarrow X=\frac{n_{1}s_{1}+n_{2}s_{2}}{\alpha},\ Y=\frac{-\tau_{2}s_{1}+\tau_{1}s_{2}}{\alpha}. (24)

We then obtain

dXdl=0,dYdl=β2Y\frac{dX}{dl}=0,\frac{dY}{dl}=-\beta_{2}Y

at the leading linear order in the expansions of the powers of XX and YY. Because the first linear term is zero in the dynamics of XX, corresponding to a transcritical bifurcation for the Lagrange–Charpit equations (16), we need to consider the second-order term in XX, namely,

e(n1τ1s1+n2τ2s2)m1Xm+X2m22+n1n2(1τ11τ2)mY+𝒪(XY,X2Y,Y2)e^{-\left(\frac{n_{1}}{\tau_{1}}s_{1}+\frac{n_{2}}{\tau_{2}}s_{2}\right)m}\simeq 1-Xm+\frac{X^{2}m^{2}}{2}+n_{1}n_{2}\left(\frac{1}{\tau_{1}}-\frac{1}{\tau_{2}}\right)mY+\mathcal{O}(XY,X^{2}Y,Y^{2})

where we dropped the terms of the order Y2Y^{2}, XYXY, and X2YX^{2}Y. We then obtain the dynamic equations at the transcritical bifurcation to the leading order:

dYdlβ2Y,dXdlω+12αX2\frac{dY}{dl}\simeq-\beta_{2}Y,\ \frac{dX}{dl}\simeq-\frac{\omega+1}{2\alpha}X^{2}

whose solutions are given by

X(l)=2αω+11ll0,Y(l)=C1eβ2(ll0)X(l)=\frac{2\alpha}{\omega+1}\frac{1}{l-l_{0}},\ Y(l)=C_{1}e^{-\beta_{2}(l-l_{0})}

with constants of integration l0l_{0} and C1C_{1}. We can assume l0=0l_{0}=0 is the initial point on the characteristic curve. Remarkably, only the contribution along the XX axis is dominant for the large ll limit (i.e., |X||Y||X|\gg|Y| for ll\rightarrow\infty), which corresponds to the asymptotic 𝒔0\boldsymbol{s}\rightarrow 0. We then obtain

logP~ss\displaystyle\log\tilde{P}_{ss} \displaystyle\simeq ν0𝑑l(n1s1(l)τ1+n2s2(l)τ2)\displaystyle\nu_{0}\int dl\left(\frac{n_{1}s_{1}(l)}{\tau_{1}}+\frac{n_{2}s_{2}(l)}{\tau_{2}}\right)
\displaystyle\simeq 2ν0αω+1logX+ν0n1n2β2(1τ11τ2)Y+C2\displaystyle-\frac{2\nu_{0}\alpha}{\omega+1}\log X+\frac{\nu_{0}n_{1}n_{2}}{\beta_{2}}\left(\frac{1}{\tau_{1}}-\frac{1}{\tau_{2}}\right)Y+C_{2}

with integration constant C2C_{2}. C2C_{2} is a divergent constant because it has to compensate the diverging logarithm logX\log X to ensure that logPss(s=0)=0\log P_{ss}(s=0)=0. This divergent constant appears as a result of neglecting the ultraviolet (UV) cutoff for small ss (which corresponds to neglecting the exponential tail of the PDF of intensities) (Kanazawa and Sornette 2020a).

Therefore, we obtain the steady solution

logP~ss(𝒔)=2ν0αω+1logX+ν0n1n2β2(1τ11τ2)Y\log\tilde{P}_{ss}(\boldsymbol{s})=-\frac{2\nu_{0}\alpha}{\omega+1}\log X+\frac{\nu_{0}n_{1}n_{2}}{\beta_{2}}\left(\frac{1}{\tau_{1}}-\frac{1}{\tau_{2}}\right)Y

for small XX and YY by ignoring the UV cutoff and constant contribution. This recovers the power-law formula of the intermediate asymptotic of the PDF of the Hawkes intensities:

logQ~ss(s):\displaystyle\log\tilde{Q}_{ss}(s): =\displaystyle= ν0s+logP~ss(s,s)2ν0αω+1logs(s0)\displaystyle-\nu_{0}s+\log\tilde{P}_{ss}(s,s)\simeq-\frac{2\nu_{0}\alpha}{\omega+1}\log s\quad(s\sim 0) (25)
\displaystyle\Longleftrightarrow P(λ)λ1+2ν0αω+1(λ),\displaystyle P(\lambda)\sim\lambda^{-1+\frac{2\nu_{0}\alpha}{\omega+1}}\quad(\lambda\rightarrow\infty),

where α=τ1n1+τ2n2\alpha=\tau_{1}n_{1}+\tau_{2}n_{2}, as defined in (24). The power-law exponent of the PDF of the excess intensity is 12ν0αω+11-\frac{2\nu_{0}\alpha}{\omega+1} and depends on ω\omega. In the limit ω0\omega\rightarrow 0, the result coincides with that in Ref.(Kanazawa and Sornette 2020b, a).

C. Discrete superposition of exponential kernels

We now study the case in which the memory kernel is the sum of KK exponentials for an arbitrary finite number KK. We obtain the steady-state PDF of the intensity λ^(t)\hat{\lambda}(t) as

logQ~ss(s)2ν0αω+1logsP(λ)λ1+2ν0αω+1,withα:=k=1Knkτk.\log\tilde{Q}_{ss}(s)\simeq-\frac{2\nu_{0}\alpha}{\omega+1}\log s\Longleftrightarrow P(\lambda)\sim\lambda^{-1+\frac{2\nu_{0}\alpha}{\omega+1}},\quad\mathrm{with}\ \alpha:=\sum^{K}_{k=1}n_{k}\tau_{k}. (26)

The power-law exponent of the PDF of the excess intensity is 12ν0αω+11-\frac{2\nu_{0}\alpha}{\omega+1} and depends on ω\omega. In the limit ω0\omega\rightarrow 0, the result coincides with that in Ref.(Kanazawa and Sornette 2020b, a).

D. General case

We study the case where the memory kernel is a continuous superposition of exponential functions. We decompose the kernel as

h(t)=1n0n(τ)1τet/τ𝑑τ,n=0n(τ)𝑑τ.h(t)=\frac{1}{n}\int^{\infty}_{0}n(\tau)\frac{1}{\tau}e^{-t/\tau}d\tau\ ,\ n=\int^{\infty}_{0}n(\tau)d\tau.

This decomposition satisfies the normalization condition 0h(t)=1\int^{\infty}_{0}h(t)=1. The function n(τ)n(\tau) quantifies the contribution of the exponential kernel 1τet/τ\frac{1}{\tau}e^{-t/\tau} to the branching ratio. The steady-state PDF of the intensity λ^t\hat{\lambda}_{t} is

logQ~ss(s)2ν0αω+1logsP(λ)λ1+2ν0αω+1,withα:=0𝑑τn(τ)τ.\log\tilde{Q}_{ss}(s)\simeq-\frac{2\nu_{0}\alpha}{\omega+1}\log s\Longleftrightarrow P(\lambda)\sim\lambda^{-1+\frac{2\nu_{0}\alpha}{\omega+1}},\quad\mathrm{with}\ \alpha:=\int^{\infty}_{0}d\tau n(\tau)\tau. (27)

The power-law exponent of the PDF of the excess intensity is 12ν0αω+11-\frac{2\nu_{0}\alpha}{\omega+1} and depends on ω\omega. In the limit ω0\omega\rightarrow 0, the result coincides with that in Ref.(Kanazawa and Sornette 2020b, a).

We now consider the case in which the memory kernel exhibits power-law decay. We express h(t)h(t) as the superposition of erte^{-rt} with the weight of the gamma distribution fγ(r)=rγ1er/Γ(γ)f_{\gamma}(r)=r^{\gamma-1}e^{-r}/\Gamma(\gamma) as

h(t)=γ(1+t)(γ+1)=0rertfγ(r)𝑑r,γ>0.h(t)=\gamma(1+t)^{-(\gamma+1)}=\int^{\infty}_{0}re^{-rt}\cdot f_{\gamma}(r)dr,\ \gamma>0.

By the change of variable r1/τr\rightarrow 1/\tau, we rewrite h(t)h(t) as the superposition of et/τe^{-t/\tau} with the weight of the inverse gamma distribution fγ(τ)=τγ1e1/τ/Γ(γ)f^{\prime}_{\gamma}(\tau)=\tau^{-\gamma-1}e^{-1/\tau}/\Gamma(\gamma) as

h(t)=01τet/τfγ(1/τ)1τ2𝑑τ=0n(τ)1τet/τ𝑑τ,n(τ)=fγ(τ).h(t)=\int^{\infty}_{0}\frac{1}{\tau}e^{-t/\tau}\cdot f_{\gamma}(1/\tau)\cdot\frac{1}{\tau^{2}}d\tau=\int^{\infty}_{0}n(\tau)\frac{1}{\tau}e^{-t/\tau}d\tau\ ,n(\tau)=f^{\prime}_{\gamma}(\tau). (28)

α\alpha of Eq.(27) is obtained as the expected value of the inverse gamma distribution,

α=0τfγ(τ)𝑑τ=1γ1,γ>1.\alpha=\int^{\infty}_{0}\tau f^{\prime}_{\gamma}(\tau)d\tau=\frac{1}{\gamma-1},\ \gamma>1. (29)

The power-law exponent is 12ν0(ω+1)(γ1)1-\frac{2\nu_{0}}{(\omega+1)(\gamma-1)}.

4 Numerical verification

We conducted numerical simulations to study the steady-state PDF of the intensity in the marked Hawkes process. In particular, our goal was to verify the theoretical predictions of Eqs. (16), (25), (26), and (29). We employed the time-rescaling theorem, which enables us to sample data more efficiently than the rejection method. The time-rescaling theorem states that any point process with an integrable conditional intensity function can be transformed into a Poisson process with a unit rate (Brown et al. 2002). That is, it is possible to convert a marked Hawkes process into a marked point process with a constant intensity of 1 by performing time-dependent rescaling. A point process with intensity 1 can easily generate the time of the next event occurrence because the event interval follows an exponential distribution. The time-rescaling theorem introduces the following time transformations:

Λ(t)0tν^(s)𝑑s.\Lambda(t)\equiv\int^{t}_{0}\hat{\nu}(s)ds. (30)

When event 𝒕n={t1,t2,,tn}\boldsymbol{t}_{n}=\{t_{1},t_{2},...,t_{n}\} follows a marked Hawkes process with intensity function ν^(t)\hat{\nu}(t) in the observation interval [0,T][0,T], event 𝒕={Λ(t1),Λ(t2),,Λ(tn)}\boldsymbol{t}^{\prime}=\{\Lambda(t_{1}),\Lambda(t_{2}),...,\Lambda(t_{n})\} obtained by the time transformation (30) follows a marked point process with intensity 11 in the observation interval [0,Λ(T)][0,\Lambda(T)]. The event interval Λ(ti)Λ(ti1)\Lambda(t_{i})-\Lambda(t_{i-1}) is given by

ΛΛ(ti)Λ(ti1)=ti1tiν^(s)𝑑sEx(1).\Lambda^{\prime}\equiv\Lambda(t_{i})-\Lambda(t_{i-1})=\int^{t_{i}}_{t_{i-1}}\hat{\nu}(s)ds\sim\mbox{Ex}(1).

Hence, to numerically realize the marked Hawkes process, it is sufficient to generate a random number Λ\Lambda^{\prime} following an exponential distribution with mean 1 and successively find tit_{i} that satisfies the above relation.

We rewrite the relation more concretely. For ti1stit_{i-1}\leq s\leq t_{i}, zk(s)z_{k}(s) is written as

zk(s)=nkj=1i11τke(stj)/τkmj=zk(ti1)e(sti1/τk.z_{k}(s)=n_{k}\sum_{j=1}^{i-1}\frac{1}{\tau_{k}}e^{-(s-t_{j})/\tau_{k}}m_{j}=z_{k}(t_{i-1})e^{-(s-t_{i-1}/\tau_{k}}.

The integral of ν(s)\nu(s) is

ti1tiν(s)𝑑s=ln(ω+1)ω(ν0(titi1)+k=1Kzk(ti1)ti1tie(sti1)/τk𝑑s).\int_{t_{i-1}}^{t_{i}}\nu(s)ds=\frac{\ln(\omega+1)}{\omega}\left(\nu_{0}(t_{i}-t_{i-1})+\sum_{k=1}^{K}z_{k}(t_{i-1})\int_{t_{i-1}}^{t_{i}}e^{-(s-t_{i-1})/\tau_{k}}ds\right).

We have to solve the next relation,

Λ\displaystyle\Lambda^{\prime} =\displaystyle= ln(ω+1)ω(ν0(titi1)+k=1Kzk(ti1)τk(1e(titi1)/τk)\displaystyle\frac{\ln(\omega+1)}{\omega}\left(\nu_{0}(t_{i}-t_{i-1})+\sum_{k=1}^{K}z_{k}(t_{i-1})\tau_{k}(1-e^{-(t_{i}-t_{i-1})/\tau_{k}}\right) (31)
\displaystyle\Longrightarrow ν0(titi1)+k=1Kzk(ti1)τk(1exp(titi1τk))ωΛln(ω+1)=0\displaystyle\nu_{0}(t_{i}-t_{i-1})+\sum_{k=1}^{K}z_{k}(t_{i-1})\tau_{k}\left(1-\mbox{exp}\left(-\frac{t_{i}-t_{i-1}}{\tau_{k}}\right)\right)-\frac{\omega\Lambda^{\prime}}{\ln(\omega+1)}=0

Note that the jump size of this process is determined according to the distribution of the mark ρ(m)\rho(m) in Eq.(6) when the event occurrence time tit_{i} is determined. The numerical procedures are given below.

Algorithm Sampling process of marked Hawkes process
We denote zk,i=zk(ti)z_{k,i}=z_{k}(t_{i}).
  Step 1    i1i\leftarrow 1.
  Step 2    Get a random number ΛEx(1)\Lambda^{\prime}\sim\mbox{Ex}(1).
  Step 3    t1Λν0t_{1}\leftarrow\frac{\Lambda^{\prime}}{\nu_{0}}.
  Step 4    Get a random number miρ(m)m_{i}\sim\rho(m) in Eq.(6).
  Step 5    z^k,1nkm1τk,k=1,,K\hat{z}_{k,1}\leftarrow\frac{n_{k}m_{1}}{\tau_{k}},k=1,\cdots,K.
  Step 6    The following steps are repeated.
     Step 6.1    ii+1i\leftarrow i+1.
     Step 6.2    Get a random number ΛEx(1)\Lambda^{\prime}\sim\mbox{Ex}(1).
     Step 6.3    Solve Eq.(31) to get tit_{i}.
     Step 6.4    Get a random number miρ(m)m_{i}\sim\rho(m) in Eq.(6).
     Step 6.5    z^k,iz^k,i1etiti1τk+nkmiτk,k=1,,K\hat{z}_{k,i}\leftarrow\hat{z}_{k,i-1}e^{-\frac{t_{i}-t_{i-1}}{\tau_{k}}}+\frac{n_{k}m_{i}}{\tau_{k}},k=1,\cdots,K.

We performed numerical simulations using Julia 1.7.3. The simulation time for t[0,T=107]t\in[0,T=10^{7}] is approximately 0.5 h for background intensity ν0=0.01\nu_{0}=0.01 and K=1K=1. The execution environment is as follows: OS: Ubuntu 20.045 LTS, Memory: 64 GB, CPU: 4.7 GHz. The code is available on github(Sakuraba 2023). We sampled the point process {ti},i=1,,N(T)\{t_{i}\},i=1,...,N(T) with T=107T=10^{7}. The common settings of the sampling processes are ν0{0.01,0.2,1.0},ω{0.01,1.0,10.0},τ=1.0\nu_{0}\in\{0.01,0.2,1.0\},\omega\in\{0.01,1.0,10.0\},\tau=1.0 and n{0.999,0.99,0.9}n\in\{0.999,0.99,0.9\}.

a. Single exponential kernel

Here, we test the theoretical prediction of Eq. (16),

PSS(λ)λ1+2ν0τω+1e2τϵω+1λ.P_{SS}(\lambda)\propto\lambda^{-1+2\frac{\nu_{0}\tau}{\omega+1}}e^{-\frac{2\tau\epsilon}{\omega+1}\lambda}.

Fig.2 presents the PDFs of λ^(t)\hat{\lambda}(t) for the cases.

Refer to caption Refer to caption Refer to caption
Refer to caption Refer to caption Refer to caption
Refer to caption Refer to caption Refer to caption
Figure 2: Plot of the steady state PDF Pss(λ)P_{ss}(\lambda) for the single exponential kernel case K=1K=1 near the critical point n=nc=1n=n_{c}=1. We set n=0.9()n=0.9(\bullet), n=0.99)()n=0.99)(\circ), and n=0.999)()n=0.999)(\bullet) for τ=1.0,ω{0.01,1.0,10.0}\tau=1.0,\omega\in\{0.01,1.0,10.0\} and ν0{0.01,0.2,1.0}\nu_{0}\in\{0.01,0.2,1.0\}. (a-c) Background intensity ν0=0.01\nu_{0}=0.01 and power-law exponents are 0.98020.9802, 0.990.99, and 0.99820.9982 for ω=0.01\omega=0.01, 1.01.0, and 1010, respectively. (d-f) Background intensity ν0=0.2\nu_{0}=0.2 and power-law exponents are 0.60390.6039, 0.80.8, and 0.96360.9636 for ω=0.01\omega=0.01, 1.01.0, and 1010, respectively. (g-i) Background intensity ν0=1.0\nu_{0}=1.0 and power-law exponents are 0.9802-0.9802, 0.00.0, and 0.81820.8182 for ω=0.01\omega=0.01, 1.01.0, and 1010, respectively.

For a small background intensity ν0=0.011\nu_{0}=0.01\ll 1, the power-law exponents are 0.98020.9802 and 0.99820.9982 for ω=0.01\omega=0.01 and ω=10.0\omega=10.0, respectively. These exponents are close to 1 and show little dependence on ω\omega. For large ν0\nu_{0}, the power-law exponent becomes more dependent on ω\omega. When ν0=1.0\nu_{0}=1.0, the power-law exponents are 0.9802-0.9802 and 0.81820.8182 for ω=0.01\omega=0.01 and ω=10.0\omega=10.0, respectively. When ω2ν0τ\omega\gg 2\nu_{0}\tau, The power-law exponent becomes universally 1, as the expression 12ν0τ/(ω+1)11-2\nu_{0}\tau/(\omega+1)\simeq 1 holds. Furthermore, the length scale of the power-law region, denoted as (ω+1)/(2τϵ)(\omega+1)/(2\tau\epsilon), increases as ω\omega increases. The numerical results confirm these findings by demonstrating agreement between the slopes of the PDFs and the theoretical values, and the exponents of the power-law were verified. In addition, by observing the xx-axis, it can be seen that the range of the straight power-law region becomes wider as ω\omega increases.

b. Double exponential kernel

We verified the theoretical predictions of Eq. (25).

PSS(λ)λ1+2ν0(τ1n1+τ2n2)ω+1.P_{SS}(\lambda)\propto\lambda^{-1+2\frac{\nu_{0}(\tau_{1}n_{1}+\tau_{2}n_{2})}{\omega+1}}.

To realize n=n1+n2{0.9,0.99,0.999}n=n_{1}+n_{2}\in\{0.9,0.99,0.999\}, we set (n1,n2)=(0.5,0.4),(0.5,0.49)(n_{1},n_{2})=(0.5,0.4),(0.5,0.49) and (0.5,0.499)(0.5,0.499). Fig.3 presents the double logarithmic plot of PDFs of λ^t\hat{\lambda}_{t}. One can observe the agreement between the slopes of the PDFs and the theoretical values and the theoretical prediction is verified.

Refer to caption Refer to caption Refer to caption
Refer to caption Refer to caption Refer to caption
Refer to caption Refer to caption Refer to caption
Figure 3: Plot of the steady-state PDF Pss(λ)P_{ss}(\lambda) for the double exponential kernel case K=2K=2 near the critical point n=nc=1n=n_{c}=1. We set n=0.0()n=0.0(\bullet), n=0.99()n=0.99(\circ), and n=0.999()n=0.999(\bullet) for (τ1,τ2)=(1.0,3.0),ω{0.01,1.0,10.0}(\tau_{1},\tau_{2})=(1.0,3.0),\omega\in\{0.01,1.0,10.0\}, and ν0{0.01,0.2,1.0}\nu_{0}\in\{0.01,0.2,1.0\}. (a-c) Background intensity ν0=0.01\nu_{0}=0.01 and power-law exponents are 0.96040.9604, 0.980.98, and 0.99640.9964 for ω=0.01\omega=0.01, 1.01.0, and 1010, respectively. (d-f) Background intensity ν0=0.2\nu_{0}=0.2 and power-law exponents are 0.20790.2079, 0.60.6, and 0.92730.9273 for ω=0.01\omega=0.01, 1.01.0, and 1010, respectively. (g-i) Background intensity ν0=1.0\nu_{0}=1.0 and power-law exponents are 2.96-2.96, 1.0-1.0, and 0.63630.6363 for ω=0.01\omega=0.01, 1.01.0, and 1010, respectively.

c. Triple exponential kernel

We verified the theoretical predictions of Eq. (26) for K=3K=3.

PSS(λ)λ1+2ν0(τ1n1+τ2n2+τ3n3)ω+1.P_{SS}(\lambda)\propto\lambda^{-1+2\frac{\nu_{0}(\tau_{1}n_{1}+\tau_{2}n_{2}+\tau_{3}n_{3})}{\omega+1}}.

To achieve n=n1+n2+n3{0.9,0.99,0.999}n=n_{1}+n_{2}+n_{3}\in\{0.9,0.99,0.999\}, we set (n1,n2,n3)=(0.3,0.2,0.4),(0.3,0.2,0.49)(n_{1},n_{2},n_{3})=(0.3,0.2,0.4),(0.3,0.2,0.49), and (0.3,0.2,0.499)(0.3,0.2,0.499). Fig.4 shows the resulting PDFs of λ^(t)\hat{\lambda}(t). One can observe the agreement between the slopes of the PDFs and the theoretical values and the theoretical prediction is verified.

Refer to caption Refer to caption Refer to caption
Refer to caption Refer to caption Refer to caption
Refer to caption Refer to caption Refer to caption
Figure 4: Plot of the numerical results for the steady-state PDF Pss(λ)P_{ss}(\lambda) for the double exponential case K=3K=3. λ\lambda near the critical point n=nc=1n=n_{c}=1. We set n=0.9()n=0.9(\bullet), n=0.99()n=0.99(\circ), and n=0.999()n=0.999(\bullet) for (τ1,τ2,τ3)=(1.0,2.0,3.0),ω{0.01,1.0,10.0}(\tau_{1},\tau_{2},\tau_{3})=(1.0,2.0,3.0),\omega\in\{0.01,1.0,10.0\}, and ν0{0.01,0.2,1.0}\nu_{0}\in\{0.01,0.2,1.0\}. (a-c) Background intensity ν0=0.01\nu_{0}=0.01 and power-law exponents are 0.9560.956, 0.9780.978, and 0.9960.996 for ω=0.01\omega=0.01, 1.01.0, and 1010, respectively. (d-f) Background intensity ν0=0.2\nu_{0}=0.2 and power-law exponents are 0.1290.129, 0.560.56, and 0.920.92 for ω=0.01\omega=0.01, 1.01.0, and 1010, respectively. (g-i) Background intensity ν0=1.0\nu_{0}=1.0 and power-law exponents are 3.356-3.356, 1.2-1.2, and 0.60.6 for ω=0.01\omega=0.01, 1.01.0, and 1010, respectively.

d. Power-law memory kernel

We verified the theoretical predictions of Eq.(29) for γ=11\gamma=11.

P(λ)λ1+2ν010(ω+1).P(\lambda)\sim\lambda^{-1+\frac{2\nu_{0}}{10(\omega+1)}}.

As τInvGamma(γ,1)\tau\sim\mathrm{InvGamma}(\gamma,1) in (28), we set K=100K=100 and {τi}i=1,,K\{\tau_{i}\}_{i=1,\cdots,K} as the (i1/2)(i-1/2)% point of the distribution. To achieve n=n1++nK{0.9,0.99,0.999}n=n_{1}+...+n_{K}\in\{0.9,0.99,0.999\}, we set ni=n/K,i=1,,Kn_{i}=n/K,i=1,\cdots,K. Fig.5 illustrates the resulting PDFs of λ^(t)\hat{\lambda}(t), confirming the validity of the theoretical prediction for γ=11\gamma=11.

Refer to caption Refer to caption Refer to caption
Refer to caption Refer to caption Refer to caption
Refer to caption Refer to caption Refer to caption
Figure 5: Plot of the numerical results for the steady-state PDF Pss(λ)P_{ss}(\lambda) for the general case K=100K=100 and γ=11\gamma=11. We set n=0.9()n=0.9(\bullet), n=0.99()n=0.99(\circ), and n=0.999()n=0.999(\bullet) for (τ1,,τ100)InvGamma(11,1),ω{0.01,1.0,10.0}(\tau_{1},...,\tau_{100})\sim\mathrm{InvGamma}(11,1),\omega\in\{0.01,1.0,10.0\}, and ν0{0.01,0.2,1.0}\nu_{0}\in\{0.01,0.2,1.0\}. (a-c) Background intensity ν0=0.01\nu_{0}=0.01 and power-law exponents are 0.998020.99802, 0.9990.999, and 0.999820.99982 for ω=0.01\omega=0.01, 1.01.0, and 1010, respectively. (d-f) Background intensity ν0=0.2\nu_{0}=0.2 and power-law exponents are 0.96040.9604, 0.980.98, and 0.99640.9964 for ω=0.01\omega=0.01, 1.01.0, and 1010, respectively. (g-i) Background intensity ν0=1.0\nu_{0}=1.0 and power-law exponents are 0.8020.802, 0.90.9, and 0.9820.982 for ω=0.01\omega=0.01, 1.01.0, and 1010, respectively.

5 Conclusion

Here, we investigated the continuous time self-exciting negative binomial process with a memory kernel consisting of a sum of finite KK exponential functions. This process is equivalent to the marked Hawkes process. Our aim was to extend the previous findings on the intermediate power-law behavior of the intensities near the critical point for the steady and non-steady state phase transition to the case of K>1K>1. To achieve this, we developed an efficient sampling method for the marked Hawkes process based on the time-rescaling theorem. We conducted extensive simulations, focusing specifically on the scenario where K=100K=100 to capture the process with a power-law memory kernel. Through these simulations, we were able to verify the theoretical predictions and confirm their accuracy.

Appendix A Method of Characteristics

The method of characteristics is a standard method to solve first-order partial differential equations (Gardiner 2009, Kanazawa and Sornette 2020b, a). The equation for the characteristic function

ϕ(s)=eisxp(x,t|x0,0)𝑑x\phi(s)=\int^{\infty}_{-\infty}\mathrm{e^{isx}}p(x,t|x_{0},0)dx

is

tϕ+kssϕ=12Ds2ϕ.\partial_{t}\phi+ks\partial_{s}\phi=-\frac{1}{2}Ds^{2}\phi. (A.1)

We consider the corresponding Lagrange–Charpit equations

dtdl=1,dsdl=ks,dϕdl=12Ds2ϕ\frac{dt}{dl}=-1,\quad\frac{ds}{dl}=-ks,\quad\frac{d\phi}{dl}=\frac{1}{2}Ds^{2}\phi

with the parameter ll encoding the position along the characteristic curves. These equations are equivalent to an invariant form in terms of ll

dt1=dsks=dϕ12Ds2ϕ.\frac{dt}{1}=\frac{ds}{ks}=-\frac{d\phi}{\frac{1}{2}Ds^{2}\phi}.

The method of characteristics can be used to solve this equation. Namely, if

u(s,t,ϕ)=a,andv(s,t,ϕ)=bu(s,t,\phi)=a,\qquad\mathrm{and}\qquad v(s,t,\phi)=b

are two integrals of the subsidiary equation (with aa and bb arbitrary constants), then a general solution of (A.1) is given by

f(u,v)=0.f(u,v)=0.

Appendix B Addendum

In a recent publication by K.Kanazawa and D.Sornette (Kanazawa and Sornette 2023), the power-law exponent of the intensity distribution of general marked Hawkes process was given. We explain the correspondence for the reader’s convenience.

In (Kanazawa and Sornette 2023), the power-law exponent is given using the second moment of the mark’s PDF E[m2][m^{2}] as

PSS(ν)ν1a,a=2τν0E[m2].P_{SS}(\nu)\propto\nu^{-1-a},a=\frac{2\tau\nu_{0}}{\mbox{E}[m^{2}]}.

The normalization of ρ(m)\rho(m) with E[m]=1[m]=1 is adopted. In our model, aa is given as

a=2τν0E[m2]/E[m].a=\frac{2\tau\nu_{0}}{\mbox{E}[m^{2}]/\mbox{E}[m]}.

We use ρ(m)\rho(m) in Eq.(6) to estimate the power-law exponent. As E[m]=ω/ln(ω+1)[m]=\omega/\ln(\omega+1), E[m2]=ω(ω+1)/ln(ω+1)[m^{2}]=\omega(\omega+1)/\ln(\omega+1), we obtain

PSS(ν)ν1a,a=2τν0E[m2]/E[m]=2τν0/(ω+1).P_{SS}(\nu)\propto\nu^{-1-a},a=\frac{2\tau\nu_{0}}{\mbox{E}[m^{2}]/\mbox{E}[m]}=2\tau\nu_{0}/(\omega+1).

The result is consistent with ours in (Hisakado et al. 2022a).

Acknowledgements.
This work was supported by JPSJ KAKENHI [Grant No. 22K03445]. We would like to thank Editage (www.editage.com) for English language editing.

References

  • Bacry et al. (2015) Bacry E, Mastromatteo I, Muzy J (2015) Hawkes processes in finance. market microstructure and liquidity, 1, 1550005
  • Blanc et al. (2017) Blanc P, Donier J, Bouchaud JP (2017) Quadratic hawkes processes for financial prices. Quantitative Finance 17(2):171–188
  • Bowsher (2007) Bowsher CG (2007) Modelling security market events in continuous time: Intensity based, multivariate point process models. Journal of Econometrics 141(2):876–912
  • Brown et al. (2002) Brown EN, Barbieri R, Ventura V, Kass RE, Frank LM (2002) The time-rescaling theorem and its application to neural spike train data analysis. Neural computation 14(2):325–346
  • Engle and Russell (1998) Engle RF, Russell JR (1998) Autoregressive conditional duration: a new model for irregularly spaced transaction data. Econometrica pp 1127–1162
  • Errais et al. (2010) Errais E, Giesecke K, Goldberg LR (2010) Affine point processes and portfolio credit risk. SIAM Journal on Financial Mathematics 1(1):642–665
  • Filimonov and Sornette (2012) Filimonov V, Sornette D (2012) Quantifying reflexivity in financial markets: Toward a prediction of flash crashes. Physical Review E 85(5):056108
  • Filimonov and Sornette (2015) Filimonov V, Sornette D (2015) Apparent criticality and calibration issues in the hawkes self-excited point process model: application to high-frequency financial data. Quantitative Finance 15(8):1293–1314
  • Gardiner (2009) Gardiner C (2009) Stochastic Methods A Handbook for the Natural and Social Sciences. Springer, Berlin
  • Hasbrouck (1991) Hasbrouck J (1991) Measuring the information content of stock trades. The Journal of Finance 46(1):179–207
  • Hawkes (1971a) Hawkes A (1971a) Point spectra of some mutually exciting point processes. Journal of the Royal Statistical Society Series B 33, DOI 10.1111/j.2517-6161.1971.tb01530.x
  • Hawkes (1971b) Hawkes AG (1971b) Spectra of some self-exciting and mutually exciting point processes. Biometrika 58(1):83–90
  • Hawkes (2018) Hawkes AG (2018) Hawkes processes and their applications to finance: a review. Quantitative Finance 18(2):193–198
  • Hawkes and Oakes (1974) Hawkes AG, Oakes D (1974) A cluster process representation of a self-exciting process. Journal of Applied Probability 11(3):493–503
  • Hisakado and Mori (2020) Hisakado M, Mori S (2020) Phase transition in the bayesian estimation of the default portfolio. Physica A: Statistical Mechanics and its Applications 544:123480
  • Hisakado et al. (2006) Hisakado M, Kitsukawa K, Mori S (2006) Correlated binomial models and correlation structures. Journal of Physics A: Mathematical and General 39(50):15365
  • Hisakado et al. (2022a) Hisakado M, Hattori K, Mori S (2022a) From the multiterm urn model to the self-exciting negative binomial distribution and hawkes processes. Physical Review E 106(3):034106
  • Hisakado et al. (2022b) Hisakado M, Hattori K, Mori S (2022b) Multi-dimensional self-exciting nbd process and default portfolios. The Review of Socionetwork Strategies 16(2):493–512
  • Kanazawa and Sornette (2020a) Kanazawa K, Sornette D (2020a) Field master equation theory of the self-excited hawkes process. Physical Review Research 2(3):033442
  • Kanazawa and Sornette (2020b) Kanazawa K, Sornette D (2020b) Nonuniversal power law distribution of intensities of the self-excited hawkes process: A field-theoretical approach. Physical Review Letters 125(13):138301
  • Kanazawa and Sornette (2023) Kanazawa K, Sornette D (2023) Asymptotic solutions to nonlinear hawkes processes: A systematic classification of the steady-state solutions. Physical Review Research 5(1):013067
  • Kirchner (2017) Kirchner M (2017) An estimation procedure for the hawkes process. Quantitative Finance 17(4):571–595
  • Sakuraba (2023) Sakuraba K (2023) Self-exciting negative binomial distribution process and critical properties of intensity distribution. URL https://github.com/Kotaro-Sakuraba/SE-NBD_Process.git
  • Wheatley et al. (2019) Wheatley S, Wehrli A, Sornette D (2019) The endo–exo problem in high frequency financial price fluctuations and rejecting criticality. Quantitative Finance 19(7):1165–1178