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

Regulating stochastic clocks

Zhe Fei Department of Finance, Boston University Questrom School of Business, Boston, MA, USA.Email: [email protected]    Weixuan Xia11footnotemark: 1 Corresponding author. Email: [email protected]
(Started 2021
This version: )
Abstract

Stochastic clocks represent a class of time change methods for incorporating trading activity into continuous-time financial models, with the ability to deal with typical asymmetrical and tail risks in financial returns. In this paper we propose a significant improvement of stochastic clocks for the same objective but without decreasing the number of trades or changing the trading intensity. Our methodology targets any Lévy subordinator, or more generally any process of nonnegative independent increments, and is based on various choices of regulating kernels motivated from repeated averaging. By way of a hyperparameter linked to the degree of regulation, arbitrarily large skewness and excess kurtosis of returns can be easily achieved. Generic-time Laplace transforms, characterizing triplets, and cumulants of the regulated clocks and subsequent mixed models are analyzed, serving purposes ranging from statistical estimation and option price calibration to simulation techniques. Under specified jump–diffusion processes and tempered stable processes, a robust moment-based estimation procedure with profile likelihood is developed and a comprehensive empirical study involving S&P500 and Bitcoin daily returns is conducted to demonstrate a series of desirable effects of the proposed methods.
MSC2020 Classifications: 60E07; 60G51; 60H30
JEL Classifications: C13; C65; G12
Keywords: Asymmetrical and tail risks; Lévy subordinators; regulating kernels; tempered stable processes; moment-based estimation; profile likelihood

1 Introduction

Financial returns have long been documented to have excessive skewness and kurtosis relative to those of an (un)conditional Gaussian distribution, leading to asymmetrical and tail risks; see, among others, [Bollerslev, 1987] [16], [Mittnik et al., 2000] [49], [Jondeau and Rockinger, 2003] [35], and [Ornthanalai, 2014] [51]. This phenomenon appears to be even more salient in the recently burgeoning cryptocurrency market which exhibits very high volatility due to its heavy reliance on supply and demand and market attention; see [Troster et al., 2019] [63], [Gkillas and Katsiampa, 2018] [30], [Chaim and Laurini, 2019] [20], and [Liu and Tsyvinski, 2021] [42], e.g., for the major cryptocurrencies including Bitcoin.

In continuous-time modeling, a well-known method for incorporating distributional asymmetry and heavy tails is to replace the usual, constantly forward-moving clock under which the underlying stochastic process is monitored with a generally independent nondecreasing stochastic process. The latter process is termed a “stochastic clock” and can be linked to trading volumes or the number of trades (see [Clark, 1973] [23], [Ané and Geman, 2000] [5], [Geman et al., 2001] [29], and [Geman, 2008] [28]). Stemming from the celebrated Dambis–Dubins–Schwarz theorem of time change, [Monroe, 1978] [50] first gave the possibility to represent any semimartingale as a time-changed Brownian motion. When the stochastic clock is modeled by a nonnegative process with independent and stationary increments, namely a Lévy subordinator, such a representation has had striking consequences in generating a variety of purely discontinuous Lévy processes as mixed models, which have flourished for the last two decades in time series econometrics, option pricing, and portfolio management, some recent contributions including [Madan et al., 2019] [47], [Aguilar et al., 2020] [2], and [Fallahgoul and Loeper, 2021] [27].

The key idea behind using stochastic clocks in the absence of a drift component in producing large skewness and kurtosis is rather comprehensible: Pure-jump stochastic clocks progress in a staircase-like manner and are less active – and less uniform – than the usual “sloped” clock of calendar time, and hence possess relatively slower local speed, representing trading activity that does not take place in a perfectly continuous manner, which enables them to adequately tell apart periods of intensive trading and other relatively calmer ones. Perhaps the two most predominant clocks of this type are the gamma process and the inverse Gaussian process, which were studied in [Madan et al., 1998] [46] and [Barndorff-Nielsen, 1997] [7], respectively, to time-change a drifted Brownian motion for building discontinuous models in dealing with skewed and leptokurtic financial returns.111The same processes also play an important role in many nonfinancial fields, especially in reliability engineering to model the degradation phenomena of structural components (see the overview by [Ye and Xie, 2015] [70]).

Noteworthily, when working with ordinary stochastic clocks under Lévy subordinators, the asymmetrical and tail risks of returns are largely controlled by the number of trades occurring for a significant period of time (i.e., the aggregate level of trading activity) through an infinite-divisibility (i.e., time-multiplicative) parameter (as mentioned in the overview in [Schoutens, 2003, Chap. 4] [59]), or the long-term trading intensity (equivalently, the variation of trading volumes) via an α\alpha-stability index (in the case of the stable family; see [Samorodnitsky and Taqqu, 1994] [57] e.g.). More specifically, the matching of a high empirical kurtosis level will require decreasing the value of the infinite-divisibility parameter which is a direct measurement of the amount of jumps,222The intuition is that as such a parameter tends to zero in value, the corresponding distribution becomes degenerate. or altering the stability index which is heavily tied to the microstructure or the local regularity of sample paths. Such an operation seems totally innocuous when return distributions are analyzed statically, after imposing assumptions of i.i.d. samples. However, in reality, the aggregate number of trades cannot be expected to be unlimitedly small, or close to zero, leading effectively to degeneracy, or trading halts; on the other hand, the long-term trading intensity is generally specified a priori without being optimized and may be estimated in a dynamical fashion using techniques of empirical power variations (see, e.g., [Aït-Sahalia and Jacod, 2009] [3], [Todorov and Tauchen, 2011] [62], [Jing et al., 2012] [34], and [Todorov, 2021] [61]), which of course varies with the frequencies at which data are observed. With restrictions on these two important dimensions, nearly all the noticed Lévy models will not function well in coping with diverse asymmetrical and tail risks of returns.

In connection with this, we are motivated to consider yet another possibility for the same purpose, by changing the size of local movements of a stochastic clock instead of decreasing their amount or adjusting the corresponding activity level. From the perspective of trading activity, this would mean that the spread of trade sizes may be enlarged at will without having to alter the level of trading intensity or stick to an implausibly small number of trades. In doing so, we prefer to operate in a continuous-time environment because discrete-time analogies are easy to develop subsequently, whereas it will be understandably difficult to go in the opposite direction, e.g., to ensure infinite divisibility in finite dimensions.

Besides, consideration of such a possibility will be beneficial in unraveling the microstructure of returns from low-frequency data via a connection to phenomena that trigger abnormal trading volumes, with ample empirical evidence. For instance, the goal to increase the sparsity of trade size distributions is in agreement with the presence of large speculators in traditional and cryptocurrency markets (mentioning, e.g., [Chang et al., 1997] [21] and [Blau, 2017] [14], respectively) or the prevalence of trade-size clustering (referring to [Alexander and Peterson, 2008] [4] and [Chen, 2019] [22]). In this regard, the linkage between skewed and leptokurtic returns and the theoretical moments of the driving stochastic clock is simply embodied by the famous price–volume relationship (see, e.g., [Karpoff, 1987] [37] and [Richardson and Smith, 1994] [55]).

With the foregoing aspects in mind, in this paper we aim to propose a unified approach towards regulating any existing stochastic clock without impairing important fundamental properties. In this context, the act of “regulating” can be understood in the ordinary sense – a stochastic clock is continually adjusted with the intent of achieving a desirable local speed, making it more reliable for capturing the differences in trading activity over time. As a starting point, however, our methodology is concentrated on the aforesaid Lévy subordinators but will be easily extendable to processes of independent but time-inhomogeneous increments, or Sato processes (see [Sato, 2006] [58] and [Eberlein and Madan, 2009] [26]), all of which can be subsequently applied to build structurally more complex stochastic clocks with mean-reverting acceleration and continuous movements that can deal with volatility clustering (see [Carr and Wu, 2004] [19]). The regulated stochastic clocks have essentially the same functionalities as the originals, and can be used to time-change diffusion processes to spawn multifarious semimartingales with discontinuities, albeit with much larger capacity in dealing with asymmetrical and tail risks.

That being said, our approach can also be exploited to generate a rich class of unexplored Lévy processes comfortable for financial modeling. As expected, this newfound dimension is encoded into a hyperparameter representing the degree of regulation, i.e., how much slower the regulated clock is compared with the unregulated one, and kept as independent as possible from the roles of other shape parameters. The resultant processes are shown to be analytically tractable whether it be statistical estimation or risk-neutral valuation of options that is of interest.

The remainder of this paper is structured as follows. In Section 2 we present our main methodology in three recipes, all of which serve to regulate a stochastic clock to any intended degrees. The effect of clock regulation on the jump behaviors depicting corresponding evolution of trading activity is analyzed (Theorem 1, Theorem 2, and Corollary 1) and the impact on asymmetry and tail heaviness of the original distribution is examined in detail (Corollary 2). Then, we present in Section 3 two simple ways for utilizing the (un)regulated stochastic clocks in order to construct real-valued mixed models. In Section 4, we demonstrate clock regulation on two important special cases featuring jump–diffusion models and purely discontinuous models and discuss their potential applications in finance as well as other fields. Section 5 outlines two possible candidates for optimizing the degree of regulation: one employing a robust moment-based estimation procedure combined with profile log-likelihood and the other using option price calibration with numerical Fourier inversion techniques, which underlie our empirical study in Section 6 concerning S&P500 and Bitcoin returns and Bitcoin options. Conclusions are drawn in Section 7 along with a summary of the properties of the three recipes and highlights on future research directions. All mathematical proofs are presented at the end of the paper, in Appendix A.

2 Regulated stochastic clocks

As already noted, we shall develop our methodology using Lévy processes exclusively. By having independent and stationary increments, such processes are easy to manage and sit at the core of statistical estimation with i.i.d. temporal data. To begin with, we outline some basic intuition.

Let X(Xt)t0X\equiv(X_{t})_{t\geq 0} be a Lévy subordinator (nondeterministic by default), acting as a stochastic clock. The objective of regulation lies in properly compressing the amplitude of its moderate jumps (for moderate values the jump component |ΔX||\Delta X|), and an attendant consequence is that the finite-dimensional distribution of XX will tend to have greater asymmetry and a heavier right tail, as measured by the skewness and kurtosis. Thus, suppose that ϕXt(u):=EeuXt\phi_{X_{t}}(u):=\textsf{E}e^{-uX_{t}} is the Laplace transform (LT) of the random variable XtX_{t} for generic t>0t>0, assumed to be definable in a neighborhood of the origin (in \mathbb{C}), and then with the canonical expansion

logϕXt(u)=m=0(u)mKt(m)m!,u\log\phi_{X_{t}}(u)=\sum^{\infty}_{m=0}\frac{(-u)^{m}K_{t}(m)}{m!},\quad u\in\mathbb{C} (2.1)

in terms of corresponding cumulants Kt(m)K_{t}(m), mm\in\mathbb{N} (with Kt(0)=0K_{t}(0)=0), the objective is converted to finding, for a fixed uu, a linear functional :{αu:α[0,1]}\mathcal{I}:\mathbb{C}^{\{\alpha u:\alpha\in[0,1]\}}\mapsto\mathbb{C} such that, from (2.1),

(logϕXt)(u)=m=0Cm(u)mKt(m)m!,(\mathcal{I}\log\phi_{X_{t}})(u)=\sum^{\infty}_{m=0}\frac{C_{m}(-u)^{m}K_{t}(m)}{m!},

where the time-invariant factors CmC_{m}’s satisfy |Cm|1|C_{m}|\leq 1 and Cm+1CmC_{m+1}\leq C_{m}, m\forall m. These relations will immediately render cumulants of larger (up to four) orders more reduced. Although there are obviously infinitely many possible ways to construct such a functional, \mathcal{I} should be conceptually matched with a recipe operating on the sample paths [0,t]sXs[0,t]\ni s\mapsto X_{s} to which the collection {ϕXtα:α[0,1]}\big{\{}\phi^{\alpha}_{X_{t}}:\alpha\in[0,1]\big{\}} of LTs correspond; otherwise temporal structures represented by path properties will be difficult to justify if possible at all. In what follows we are going to present three utterly simple and closely related recipes of this nature, by exploiting continuous temporal averaging.

2.1 Repeated averaging-induction: Three recipes

We design our first recipe as follows. Set X¯(0):=X(0)X\bar{X}^{(0)}:=X^{(0)}\equiv X and, for any n++{0}n\in\mathbb{N}_{++}\equiv\mathbb{N}\setminus\{0\}, repeatedly define the running averages

X¯0(n)=0;X¯t(n):=1t0tX¯s(n1)ds,t>0.\bar{X}^{(n)}_{0}=0;\quad\bar{X}^{(n)}_{t}:=\frac{1}{t}\int^{t}_{0}\bar{X}^{(n-1)}_{s}{\rm d}s,\quad t>0. (2.1.1)

For each n1n\geq 1, X¯(n)\bar{X}^{(n)} is a stochastic process having 𝖯\mathsf{P}-a.s. continuous sample paths, where continuity at 0 follows by the dominated convergence theorem. More specifically, the sample paths tX¯t(n)t\mapsto\bar{X}^{(n)}_{t} belong to the class 𝒞n1(+;+)\mathcal{C}^{n-1}(\mathbb{R}_{+};\mathbb{R}_{+}) of (n1n-1)th-order differentiable functions for n1n\geq 1, 𝖯\mathsf{P}-a.s., as a result of repeated integration.

Since (2.1.1) operates a linear functional, it is clear that for every n1n\geq 1 and fixed t>0t>0 the random variable X¯t(n)\bar{X}^{(n)}_{t} has an infinitely divisible distribution, which induces a unique Lévy process that is also nonnegative, which we denote as

Xt(n)=d.X¯t(n),t0,n++.X^{(n)}_{t}\overset{\text{d.}}{=}\bar{X}^{(n)}_{t},\quad t\geq 0,\;n\in\mathbb{N}_{++}. (2.1.2)

By nonnegativity every X(n)X^{(n)} can be used as a different stochastic clock, and will be referred to as the type-I regulated XX-clock of degree nn.

The intuition behind the first recipe is rather clear: Averaging in time has a natural effect of contracting moderate values of the sample paths and continuing to do so will tend to enlarge such an effect. To give some remarks, in the special case n=1n=1, the process X¯(1)=1/()0Xsds\bar{X}^{(1)}=1/(\cdot)\int^{\cdot}_{0}X_{s}{\rm d}s is precisely the running average of XX, which is heavily tied to linear integral functionals of Lévy processes; such integrals have been widely applied to trace the cumulative patterns of various stochastic systems; some notable examples include the total cost of random epidemics (see, e.g., [Downton, 1972] [25] and [Pollett, 2003] [54]), cumulation of count data generalized to non-integer values ([Orsingher and Polito, 2013] [52]) or count data with overdispersion ([Xia, 2019] [67]), and the modeling of structural degradation phenomena in the presence of memory effects (see [Tseng and Peng, 2004] [64] and [Xia, 2021, Sect. 4.1] [68]).

Note that the averaging–induction procedure also works for any integer n1n\geq 1, and hence it is intuitive to consider inducing a Lévy process for each nn using (2.1.2), before initiating a subsequent averaging step (2.1.1). In this way the integrand in the averaging step can be made a Lévy process even for n2n\geq 2, and each resultant averaged process will correspond to the linear functional of a certain Lévy process. Such an idea leads to our second recipe. Likewise, set X~(0):=Y(0)X\tilde{X}^{(0)}:=Y^{(0)}\equiv X, and then repeatedly utilize the first recipe for every n++n\in\mathbb{N}_{++} in the following fashion:

X~0(n)=0;X~t(n):=1t0tYs(n1)ds,t>0,\tilde{X}^{(n)}_{0}=0;\quad\tilde{X}^{(n)}_{t}:=\frac{1}{t}\int^{t}_{0}Y^{(n-1)}_{s}{\rm d}s,\quad t>0, (2.1.3)

where Y(n)Y^{(n)}’s are induced Lévy processes such that

Yt(n)=d.X~t(n),t0,n++.Y^{(n)}_{t}\overset{\text{d.}}{=}\tilde{X}^{(n)}_{t},\quad t\geq 0,\;n\in\mathbb{N}_{++}.

It is clear that the averaged processes X¯(1)\bar{X}^{(1)} and X~(1)\tilde{X}^{(1)} are indistinguishable, and so are the induced processes X(1)X^{(1)} and Y(1)Y^{(1)}, but for n2n\geq 2 they are different with dissimilar path properties – the sample paths tX~t(n)t\mapsto\tilde{X}^{(n)}_{t} will remain in the class (𝒞𝒞1)(+;+)(\mathcal{C}\setminus\mathcal{C}^{1})(\mathbb{R}_{+};\mathbb{R}_{+}) (𝖯\mathsf{P}-a.s.) for n2n\geq 2.

It can be shown (see the proof of Theorem 1 in Appendix A) that the second recipe can be essentially thought to average the log-LT, while in comparison the first recipe, by construction, has the averaging effect on the sample paths. We will refer to the process Y(n)Y^{(n)}, n1n\geq 1, as the type-II regulated XX-clock of degree nn.

In the averaging step (2.1.1), if we stack up the time scale (1/t1/t) outside of the integral operators for every n2n\geq 2, the averaging procedure becomes a folded cumulation with a one-time power scaling, and we shall have our third recipe. Set X˘(0):=Z(0)X\breve{X}^{(0)}:=Z^{(0)}\equiv X and define the quasi-averages

X˘0(n)=0;X˘t(n):=1tn0tnXsdsdsn,t>0,n++,\breve{X}^{(n)}_{0}=0;\quad\breve{X}^{(n)}_{t}:=\frac{1}{t^{n}}\underbrace{\int\cdots\int^{t}_{0}}_{n}X_{s}\underbrace{{\rm d}s\cdots{\rm d}s}_{n},\quad t>0,\;n\in\mathbb{N}_{++}, (2.1.4)

which has an infinitely divisible distribution for the same reason as in (2.1.1), and induce a unique nonnegative Lévy process,

Zt(n)=d.X˘t(n),t0,n++,Z^{(n)}_{t}\overset{\text{d.}}{=}\breve{X}^{(n)}_{t},\quad t\geq 0,\;n\in\mathbb{N}_{++},

referred to as the type-III regulated XX-clock of degree nn.

It is obvious that the processes X˘(n)\breve{X}^{(n)} and X¯(n)\bar{X}^{(n)} have the same degree of path smoothness but are guaranteed to coincide (up to indistinguishability) only if n=1n=1. Besides, the distribution of X˘t(n)\breve{X}^{(n)}_{t} is also infinitely divisible for every n1n\geq 1 and fixed tt.

We present some general formulae for the LTs of the three types of regulated stochastic clocks in Theorem 1, which are naturally extended to non-integer-valued n>0n>0 as well. This aspect hints at the extension of the three recipes to the sense of averaging of fractional degrees and will be considered in the optimization of the regulation degree nn.

Theorem 1.

For generic t>0t>0 and any n>0n>0, the LTs of X¯t(n)\bar{X}^{(n)}_{t}, X~t(n)\tilde{X}^{(n)}_{t}, and X˘t(n)\breve{X}^{(n)}_{t}, are given by

ϕX¯t(n)(u)\displaystyle\phi_{\bar{X}^{(n)}_{t}}(u) :=EeuX¯t(n)=exp(t01logϕX1((1Γ(n,logs)Γ(n))u)ds),\displaystyle:=\textsf{E}e^{-u\bar{X}^{(n)}_{t}}=\exp\bigg{(}t\int^{1}_{0}\log\phi_{X_{1}}\bigg{(}\bigg{(}1-\frac{\mathrm{\Gamma}(n,-\log s)}{\mathrm{\Gamma}(n)}\bigg{)}u\bigg{)}{\rm d}s\bigg{)},
ϕX~t(n)(u)\displaystyle\phi_{\tilde{X}^{(n)}_{t}}(u) :=EeuX~t(n)=exp(tΓ(n)01(logv)n1logϕX1(uv)dv),\displaystyle:=\textsf{E}e^{-u\tilde{X}^{(n)}_{t}}=\exp\bigg{(}\frac{t}{\mathrm{\Gamma}(n)}\int^{1}_{0}(-\log v)^{n-1}\log\phi_{X_{1}}(uv){\rm d}v\bigg{)},
ϕX˘t(n)(u)\displaystyle\phi_{\breve{X}^{(n)}_{t}}(u) :=EeuX˘t(n)=exp(t01logϕX1(snuΓ(n+1))ds),Reu0,\displaystyle:=\textsf{E}e^{-u\breve{X}^{(n)}_{t}}=\exp\bigg{(}t\int^{1}_{0}\log\phi_{X_{1}}\bigg{(}\frac{s^{n}u}{\mathrm{\Gamma}(n+1)}\bigg{)}{\rm d}s\bigg{)},\quad\mathrm{Re}u\geq 0, (2.1.5)

respectively, where Γ()\mathrm{\Gamma}(\cdot) and Γ(,)\mathrm{\Gamma}(\cdot,\cdot) symbolize, respectively, the usual gamma function and the (upper) incomplete gamma function.

In any case, the recipes have given rise to three families of stochastic clocks, controlled by the degree nn of regulation. The effect of operating the three recipes on the unregulated clock XX is twofold – on path properties and distributional properties, as we are bound to investigate next.

2.2 Characterizing triplets

From their primary motivation, the three recipes are all supposed to compress the jumps of XX into extreme values. To confirm such effects, it is necessary to derive the characterizing triplets of the regulated clocks X(n)X^{(n)}’s, Y(n)Y^{(n)}’s, and Z(n)Z^{(n)}’s governing their sample path properties. Doing so will also be helpful for developing simulation methods, the main reason being that the LTs that we have obtained in (1), having exponentiated integrals, are generally difficult to invert analytically, thus circumventing the implementation of inverse sampling methods.

First, there is no Brownian component in X(n)X^{(n)}’s, Y(n)Y^{(n)}’s or Z(n)Z^{(n)}’s and we assume without loss of generality that there is no drift as well, so that we can concentrate on the corresponding Lévy measure. More specifically, suppose that the LT ϕX1\phi_{X_{1}} takes the following general form due to the Lévy–Khintchine representation:

ϕX1(u)=exp0(euz1)ν(dz),Reu0,\phi_{X_{1}}(u)=\exp\int^{\infty}_{0}(e^{-uz}-1)\nu({\rm d}z),\quad\mathrm{Re}u\geq 0,

with the characterizing triplet (0,0,ν)(0,0,\nu), for some Lévy measure333The notation νX\nu_{X} will occasionally be used for the Lévy measure of XX in the mathematical proofs (Appendix A) to avoid ambiguity. ν\nu imposed on (+)\mathcal{B}(\mathbb{R}_{+}) such that ν({0})=0\nu(\{0\})=0 and 0(z1)ν(dz)<\int^{\infty}_{0}(z\wedge 1)\nu({\rm d}z)<\infty. Then, the characterizing triplets of X(n)X^{(n)}, Y(n)Y^{(n)}, and Z(n)Z^{(n)}, given n>0n>0, are denoted as (0,0,ν¯(n))(0,0,\bar{\nu}^{(n)}), (0,0,ν~(n))(0,0,\tilde{\nu}^{(n)}), and (0,0,ν˘(n))(0,0,\breve{\nu}^{(n)}) respectively, which make the representations

ϕX¯1(n)(u)\displaystyle\phi_{\bar{X}^{(n)}_{1}}(u) =exp0(euz1)ν¯(n)(dz),\displaystyle=\exp\int^{\infty}_{0}(e^{-uz}-1)\bar{\nu}^{(n)}({\rm d}z),
ϕX~1(n)(u)\displaystyle\phi_{\tilde{X}^{(n)}_{1}}(u) =exp0(euz1)ν~(n)(dz),\displaystyle=\exp\int^{\infty}_{0}(e^{-uz}-1)\tilde{\nu}^{(n)}({\rm d}z),
ϕX˘1(n)(u)\displaystyle\phi_{\breve{X}^{(n)}_{1}}(u) =exp0(euz1)ν˘(n)(dz).\displaystyle=\exp\int^{\infty}_{0}(e^{-uz}-1)\breve{\nu}^{(n)}({\rm d}z). (2.2.1)
Theorem 2.

For any n>0n>0, we have

ν¯(n)(dz)\displaystyle\bar{\nu}^{(n)}({\rm d}z) =Γ(n)1ν(d(yz))y2Qn1(n,11/y)dy,\displaystyle=\mathrm{\Gamma}(n)\int^{\infty}_{1}\frac{\nu({\rm d}(yz))}{y^{2}\mathrm{Q}^{n-1}(n,1-1/y)}{\rm d}y,
ν~(n)(dz)\displaystyle\tilde{\nu}^{(n)}({\rm d}z) =1Γ(n)1(logy)n1ν(d(yz))y2dy,\displaystyle=\frac{1}{\mathrm{\Gamma}(n)}\int^{\infty}_{1}\frac{(\log y)^{n-1}\nu({\rm d}(yz))}{y^{2}}{\rm d}y,
ν˘(n)(dz)\displaystyle\breve{\nu}^{(n)}({\rm d}z) =Γ(n+1)(Γ(n+1)y)1/nν(d(yz))nydy,z0,\displaystyle=\int^{\infty}_{\mathrm{\Gamma}(n+1)}\bigg{(}\frac{\mathrm{\Gamma}(n+1)}{y}\bigg{)}^{1/n}\frac{\nu({\rm d}(yz))}{ny}{\rm d}y,\quad z\geq 0, (2.2.2)

where Q(n,s)\mathrm{Q}(n,s) is the inverse regularized (upper) incomplete gamma function.444It is understood as the unique solution x0x\geq 0 to the transcendental equation Γ(n,x)/Γ(n)=s\mathrm{\Gamma}(n,x)/\mathrm{\Gamma}(n)=s for s[0,1)s\in[0,1); see [DiDonato and Morris, 1986] [24].

Furthermore, if ν\nu is absolutely continuous with respect to Lebesgue measure, then

¯(n)(z)\displaystyle\bar{\ell}^{(n)}(z) :=ν¯(n)(dz)dz=Γ(n)1(yz)yQn1(n,11/y)dy,\displaystyle:=\frac{\bar{\nu}^{(n)}({\rm d}z)}{{\rm d}z}=\mathrm{\Gamma}(n)\int^{\infty}_{1}\frac{\ell(yz)}{y\mathrm{Q}^{n-1}(n,1-1/y)}{\rm d}y,
~(n)(z)\displaystyle\tilde{\ell}^{(n)}(z) :=ν~(n)(dz)dz=1Γ(n)1(logy)n1(yz)ydy,\displaystyle:=\frac{\tilde{\nu}^{(n)}({\rm d}z)}{{\rm d}z}=\frac{1}{\mathrm{\Gamma}(n)}\int^{\infty}_{1}\frac{(\log y)^{n-1}\ell(yz)}{y}{\rm d}y,
˘(n)(z)\displaystyle\breve{\ell}^{(n)}(z) :=ν˘(n)(dz)dz=Γ(n+1)(Γ(n+1)y)1/n(yz)ndy,z0,\displaystyle:=\frac{\breve{\nu}^{(n)}({\rm d}z)}{{\rm d}z}=\int^{\infty}_{\mathrm{\Gamma}(n+1)}\bigg{(}\frac{\mathrm{\Gamma}(n+1)}{y}\bigg{)}^{1/n}\frac{\ell(yz)}{n}{\rm d}y,\quad z\geq 0, (2.2.3)

with ν(dz)=(z)dz\nu({\rm d}z)=\ell(z){\rm d}z.

We remark that the Lévy measures ν¯(n)\bar{\nu}^{(n)}’s, ν~(n)\tilde{\nu}^{(n)}’s, and ν˘(n)\breve{\nu}^{(n)}’s from (2), for n1n\geq 1, are all absolutely continuous because of integration, and thus their associated Lévy densities on the left-hand side of (2) exist unconditionally. More importantly, we have the following equalities among their Blumenthal–Getoor indices ([Blumenthal and Getoor, 1961] [15]).

Corollary 1.

For any n>0n>0, it holds that

𝔅(X(n)):=inf{p>0:01|z|pν¯(n)(dz)<}=𝔅(X),\mathfrak{B}(X^{(n)}):=\inf\bigg{\{}p>0:\int^{1}_{0}|z|^{p}\bar{\nu}^{(n)}({\rm d}z)<\infty\bigg{\}}=\mathfrak{B}(X), (2.2.4)

and similarly 𝔅(Y(n))=𝔅(Z(n))=𝔅(X)\mathfrak{B}(Y^{(n)})=\mathfrak{B}(Z^{(n)})=\mathfrak{B}(X).

Corollary 1 means that none of the three recipes alters the path regularity of XX, which is beyond doubt a very desirable property as there is never an intention of changing the jump activity of the underlying models connected with trading intensities (recalling Section 1). On the other hand, all of the induced Lévy measures govern jumps of arbitrarily large amplitudes, whereas their actual weights (namely likelihood) will depend on the functional form of ν\nu (or that of the LT), subject to further adjustment of its parameters, and through the foregoing relations. This aspect will be clearer as we move to specific distribution families of interest in Section 4.

To gain additional insight into how the three recipes distort the Lévy measures ν¯(n)\bar{\nu}^{(n)}, ν~(n)\tilde{\nu}^{(n)}, and ν˘(n)\breve{\nu}^{(n)} from a comparative angle, it is first observed from (2) in Theorem 2 that these Lévy measures do coincide if and only if n=1n=1. Upon the substitution yzyyz\mapsto y, they can be viewed as fractional integrals of the Lévy measure ν\nu of XX under different kernels, which with y=1y=1 fixed boil down to the functions Γ(n)/Qn1(n,1z)\mathrm{\Gamma}(n)/\mathrm{Q}^{n-1}(n,1-z), (logz)n1/Γ(n)(-\log z)^{n-1}/\mathrm{\Gamma}(n), and z1/n1/nz^{1/n-1}/n, respectively, defined for z(0,1)z\in(0,1). Interestingly, we then observe that the first two functions are precisely the derivatives in magnitude of eQ(n,1z)e^{-\mathrm{Q}(n,1-z)} and 1Γ(n,logz)/Γ(n)1-\mathrm{\Gamma}(n,-\log z)/\mathrm{\Gamma}(n), respectively, which happen to be inverses of each other. Put differently, the effects on the (average-induced) Lévy measures from the first two recipes are perfectly complementary, in the sense that neither dominates the other. Differently, the third function integrates in magnitude to 1z1/n1-z^{1/n}, which is not in direct comparison with the others, for the third recipe, strictly speaking, does not embody a running average for n1n\neq 1. This discussion is further illustrated in Figure 1, with five integer degrees n[1,5]n\in\mathbb{N}\cap[1,5] of regulation. We see that the first recipe, which averages the sample paths, has a larger impact (distortion) when ν\nu is concentrated on small values of +\mathbb{R}_{+}, relative to the second, which at bottom averages the log-LT. In addition, the larger the regulation degree, the faster the impact differences are reflected.

Refer to caption
Figure 1: Regulation impact on induced Lévy measures

2.3 Asymmetry and tail heaviness

The static effect from operating the three recipes is to enlarge the asymmetry and (right) tail heaviness of the distribution of the unregulated clock XtX_{t} for any fixed t>0t>0. With the LTs in (1) at hand, computation of the statistics of the corresponding regulated clocks Xt(n)X^{(n)}_{t}’s and Yt(n)Y^{(n)}_{t}’s is straightforward. More particularly, we shall focus on the impact of regulation on skewness and excess kurtosis. Let us restate from (2.1) that the corresponding cumulants are defined to be the series coefficients {K¯t(m)}m=0\{\bar{K}_{t}(m)\}^{\infty}_{m=0} and {K~t(m)}m=0\{\tilde{K}_{t}(m)\}^{\infty}_{m=0} in

logϕX¯t(u)=m=0(u)mK¯t(m)m!,logϕX~t(u)=m=0(u)mK~t(m)m!,logϕX˘t(u)=m=0(u)mK˘t(m)m!,\log\phi_{\bar{X}_{t}}(u)=\sum^{\infty}_{m=0}\frac{(-u)^{m}\bar{K}_{t}(m)}{m!},\quad\log\phi_{\tilde{X}_{t}}(u)=\sum^{\infty}_{m=0}\frac{(-u)^{m}\tilde{K}_{t}(m)}{m!},\quad\log\phi_{\breve{X}_{t}}(u)=\sum^{\infty}_{m=0}\frac{(-u)^{m}\breve{K}_{t}(m)}{m!},

with K¯t(0)=K~t(0)=K˘t(0)=0\bar{K}_{t}(0)=\tilde{K}_{t}(0)=\breve{K}_{t}(0)=0. The next corollary explains how the cumulants vary with the regulation degree nn.

Corollary 2.

For any n>0n>0, we have the cumulant reduction relations

K¯t(n)(m)=Cm,nKt(m),K~t(n)(m)=Kt(m)(m+1)n,K˘t(n)(m)=Kt(m)(mn+1)Γm(n+1),\bar{K}^{(n)}_{t}(m)=C_{m,n}K_{t}(m),\quad\tilde{K}^{(n)}_{t}(m)=\frac{K_{t}(m)}{(m+1)^{n}},\quad\breve{K}^{(n)}_{t}(m)=\frac{K_{t}(m)}{(mn+1)\mathrm{\Gamma}^{m}(n+1)}, (2.3.1)

with the sequence

Cm,n=01(1Γ(n,logs)Γ(n))mds,m++,C_{m,n}=\int^{1}_{0}\bigg{(}1-\frac{\mathrm{\Gamma}(n,-\log s)}{\mathrm{\Gamma}(n)}\bigg{)}^{m}{\rm d}s,\quad m\in\mathbb{N}_{++}, (2.3.2)

which is rational for n++n\in\mathbb{N}_{++}.

Using Corollary 2, the relations for the mean, variance, skewness and excess kurtosis of the type-I nn-degree regulated XX-clocks are given by

EXt(n)=C1,nEXt,VarXt(n)=C2,nVarXt,\displaystyle\textsf{E}X^{(n)}_{t}=C_{1,n}\textsf{E}X_{t},\quad\textsf{Var}X^{(n)}_{t}=C_{2,n}\textsf{Var}X_{t},
SkewXt(n)=C3,nC2,n3/2SkewXt,EKurtXt(n)=C4,nC2,n2EKurtXt.\displaystyle\textsf{Skew}X^{(n)}_{t}=\frac{C_{3,n}}{C^{3/2}_{2,n}}\textsf{Skew}X_{t},\quad\textsf{EKurt}X^{(n)}_{t}=\frac{C_{4,n}}{C^{2}_{2,n}}\textsf{EKurt}X_{t}. (2.3.3)

Since the integrand in (2.3.2) admits the series representation ([Abramowitz and Stegun, 1972, Eq. 6.5.29] [1])

1Γ(n,logs)Γ(n)=s(logs)nΓ(n)k=0(logs)k(n)k+1=s(logs)nΓ(n+1)(1+O(n1)),as n,1-\frac{\mathrm{\Gamma}(n,-\log s)}{\mathrm{\Gamma}(n)}=\frac{s(-\log s)^{n}}{\mathrm{\Gamma}(n)}\sum^{\infty}_{k=0}\frac{(-\log s)^{k}}{(n)_{k+1}}=\frac{s(-\log s)^{n}}{\mathrm{\Gamma}(n+1)}(1+O(n^{-1})),\quad\text{as }n\rightarrow\infty,

carrying out the integral over s(0,1)s\in(0,1) gives

Cm,n=O(n(1m)/2(mm+1)mn),as n,C_{m,n}=O\bigg{(}n^{(1-m)/2}\bigg{(}\frac{m}{m+1}\bigg{)}^{mn}\bigg{)},\quad\text{as }n\rightarrow\infty, (2.3.4)

and hence the sequence decays exponentially as nn increases for m=1m=1 and (strictly) faster for m2m\geq 2. This indicates that the first recipe shrinks the cumulants of XX with acceleration – more aggressively so as regulation continues with nn. Note that for m=1m=1 the exponential decay is in fact an equality C1,n=2nC_{1,n}=2^{-n}. Looking at (2.3) and (2.3.4) we can also claim that the type-I regulated clocks have at least exponentially decaying mean and variance but faster-than-exponentially increasing skewness and excess kurtosis, with respect to the degree nn of regulation, other things (e.g., parameter values) held equal.

In the case of the type-II regulated clocks, the counterpart of (2.3) for Yt(n)Y^{(n)}_{t} is

EYt(n)=12nEXt,VarYt(n)=13nVarXt,SkewYt(n)=(334)nSkewXt,EKurtYt(n)=(95)nEKurtXt.\textsf{E}Y^{(n)}_{t}=\frac{1}{2^{n}}\textsf{E}X_{t},\quad\textsf{Var}Y^{(n)}_{t}=\frac{1}{3^{n}}\textsf{Var}X_{t},\quad\textsf{Skew}Y^{(n)}_{t}=\bigg{(}\frac{3\sqrt{3}}{4}\bigg{)}^{n}\textsf{Skew}X_{t},\quad\textsf{EKurt}Y^{(n)}_{t}=\bigg{(}\frac{9}{5}\bigg{)}^{n}\textsf{EKurt}X_{t}. (2.3.5)

Intuitively, this signifies that the cumulant-reduction effect of the second recipe is uniform, in the sense that for each additional round of regulation, the mean and variance are reduced by fixed proportions (1/21/2 and 1/31/3, respectively) while the skewness and excess kurtosis are simultaneously enlarged by fixed proportions (33/43\sqrt{3}/4 and 9/59/5, respectively). In consequence, for the type-II regulated XX-clocks the mean and variance decay exactly exponentially and so do the skewness and excess kurtosis increase, other things equal.

As for the type III, we have similarly from Corollary 2 the quotient K˘t(n)(m)/Kt(m)=1/((mn+1)Γm(n+1))\breve{K}^{(n)}_{t}(m)/K_{t}(m)=1/((mn+1)\mathrm{\Gamma}^{m}(n+1)), m++m\in\mathbb{N}_{++}, and subsequently the following moment relations for Zt(n)Z^{(n)}_{t}:

EZt(n)=1Γ(n+2)EXt,VarZt(n)=1(2n+1)Γ2(n+1)VarXt,\displaystyle\textsf{E}Z^{(n)}_{t}=\frac{1}{\mathrm{\Gamma}(n+2)}\textsf{E}X_{t},\quad\textsf{Var}Z^{(n)}_{t}=\frac{1}{(2n+1)\mathrm{\Gamma}^{2}(n+1)}\textsf{Var}X_{t},
SkewZt(n)=(2n+1)3/23n+1SkewXt,EKurtZt(n)=(2n+1)24n+1EKurtXt.\displaystyle\textsf{Skew}Z^{(n)}_{t}=\frac{(2n+1)^{3/2}}{3n+1}\textsf{Skew}X_{t},\quad\textsf{EKurt}Z^{(n)}_{t}=\frac{(2n+1)^{2}}{4n+1}\textsf{EKurt}X_{t}. (2.3.6)

Upon comparing the above quotient with (2.3.4), with Γ(n+1)=O(n(n/e)n)\mathrm{\Gamma}(n+1)=O(\sqrt{n}(n/e)^{n}) as nn\rightarrow\infty it is seen that the third recipe generates an even severer cumulant-reduction effect than the first recipe does. However, the skewness and excess kurtosis are enlarged with deceleration, noted that (2n+1)3/2/(3n+1)=O(n1/2)(2n+1)^{3/2}/(3n+1)=O(n^{1/2}) and (2n+1)2/(4n+1)=O(n)(2n+1)^{2}/(4n+1)=O(n), as nn\rightarrow\infty. Put together, the type-III regulated XX-clocks have mean and variance of faster-than-exponential decay whereas the skewness and excess kurtosis exhibit slower-than-linear growth, which properties are in stark contrast with those of the first two types.

On paper, if we do not alter the parameter (if any) values of XX, then all three recipes are able to effectively enlarge the asymmetry and tail heaviness of its distribution, with the first having the most significant effect, whereas the downside is that the mean and variance are reduced at the same time. The latter effect is, of course, undesirable for applications, and eluding it will require at least two independent parameters so that the mean and variance can be held constant across the regulation degree nn. More specifically, when a real-valued model is to be constructed from XX (see Section 3), with the aid of an additional location parameter the mean can be readily fixed by way of centralization, and so the other parameter will most likely (and ideally) be a scale parameter, to which the standard deviation is proportional and the skewness and excess kurtosis are immune. Such a requirement becomes rather standard in statistical inference, and in consequence the skewness and excess kurtosis relations stated in (2.3), (2.3.5), and (2.3) provide useful implications. We remark, however, that if the mean is to be kept invariant to regulation by way of a shape parameter of XX (oftentimes existent for one-sided distributions) then changing its value is likely to enlarge skewness and excess kurtosis at rates different from those in (2.3) and (2.3.5), despite exponential growth rates. Such a situation cannot be analyzed from a universal viewpoint and will again depend on the functional form of the LT ϕX1\phi_{X_{1}}, with respect to the shape parameter. More details are provided in Section 4.

Similarly, the results for the statistics stated in this section can be naturally extended for fractional degrees n>0n>0 of regulation. In Figure 2 we give a joint plot of the enlargement effect on asymmetry and tail heaviness for the two recipes with real regulation degrees n(0,5]n\in(0,5], by using (2.3), (2.3.5), and (2.3) in sequence.

Refer to caption
Figure 2: Enlargement effect on asymmetry and tail heaviness with respect to regulation degree

3 Time-changed processes

As aforementioned, the functionality of the regulated clocks X(n)X^{(n)}’s, Y(n)Y^{(n)}’s, and Z(n)Z^{(n)}’s is meant to be the same as that of XX – mathematically, to adjust the time underlying a (continuous) base process and generate a wide class of models for real-valued temporal data such as financial returns. In the following we shall consider two well-known base processes for the mixture, which are also shown to retain, leastways in part, the enlargement effect on asymmetry and tail heaviness through the clock regulation.

3.1 Constant mixtures

The simplest base process is beyond doubt a (deterministic) drift (t)(t) up to positive scaling, and time-changing it by the clock XX gives none but XX itself. Then, in order to create a real-valued random variable from XX it suffices to take the difference of two of its independent copies. Specifically, let XX^{\prime} be an independent copy of XX, and then define the difference process

ξtX=μt+κ1Xtκ2Xt,t0,\xi^{X}_{t}=\mu t+\kappa_{1}X_{t}-\kappa_{2}X^{\prime}_{t},\quad t\geq 0, (3.1.1)

where μ\mu\in\mathbb{R} is a location parameter and κ1,κ20\kappa_{1},\kappa_{2}\geq 0 are additional scale parameters. Then ξX\xi^{X} is a purely discontinuous real-valued Lévy process admitting the LT (fixing t>0t>0)

ϕξtX(u):=EeuξtX=eμtuϕXt(κ1u)ϕXt(κ2u),ui,\phi_{\xi^{X}_{t}}(u):=\textsf{E}e^{-u\xi^{X}_{t}}=e^{-\mu tu}\phi_{X_{t}}(\kappa_{1}u)\phi_{X_{t}}(-\kappa_{2}u),\quad u\in{\rm i}\mathbb{R}, (3.1.2)

where the domain of Reu\mathrm{Re}u can be extended to a neighborhood of 0 provided that (2.1) holds.

Note that since the integral operator 1/t0tds1/t\int^{t}_{0}{\rm d}s is for every t>0t>0 a linear isometry over the space 𝕃(+;+)\mathbb{L}^{\infty}(\mathbb{R}_{+};\mathbb{R}_{+}) of bounded functions, considering such independent differences under the nn-degree regulated clocks is no different from operating the three recipes on the initial difference process ξX\xi^{X}. As a result, we can define the corresponding difference processes under each X(n)X^{(n)}, Y(n)Y^{(n)}, and Z(n)Z^{(n)}, denoted ξ(n),X\xi^{(n),X}, ξ(n),Y\xi^{(n),Y}, and ξ(n),Z\xi^{(n),Z}, respectively, in the same way as in (3.1.1).

Corollary 3.

For generic t>0t>0 and any n>0n>0, the LTs of the constant mixtures subject to clock regulation are given by

ϕξt(n),X(u)\displaystyle\phi_{\xi^{(n),X}_{t}}(u) =eμtuϕX¯t(n)(κ1u)ϕX¯t(n)(κ2u),\displaystyle=e^{-\mu tu}\phi_{\bar{X}^{(n)}_{t}}(\kappa_{1}u)\phi_{\bar{X}^{(n)}_{t}}(-\kappa_{2}u),
ϕξt(n),Y(u)\displaystyle\phi_{\xi^{(n),Y}_{t}}(u) =eμtuϕX~t(n)(κ1u)ϕX~t(n)(κ2u),\displaystyle=e^{-\mu tu}\phi_{\tilde{X}^{(n)}_{t}}(\kappa_{1}u)\phi_{\tilde{X}^{(n)}_{t}}(-\kappa_{2}u),
ϕξt(n),Z(u)\displaystyle\phi_{\xi^{(n),Z}_{t}}(u) =eμtuϕX˘t(n)(κ1u)ϕX˘t(n)(κ2u),ui.\displaystyle=e^{-\mu tu}\phi_{\breve{X}^{(n)}_{t}}(\kappa_{1}u)\phi_{\breve{X}^{(n)}_{t}}(-\kappa_{2}u),\quad u\in{\rm i}\mathbb{R}. (3.1.3)

Besides, by independence the mmth cumulant of ξt(n),X\xi^{(n),X}_{t} for fixed t>0t>0 is nothing but μt𝟙{1}(m)+(κ1m+(κ2)m)K¯t(n)(m)\mu t\mathbb{1}_{\{1\}}(m)+(\kappa^{m}_{1}+(-\kappa_{2})^{m})\bar{K}^{(n)}_{t}(m) and similarly for ξt(n),Y\xi^{(n),Y}_{t} and ξt(n),Z\xi^{(n),Z}_{t}. If the mean and variance of ξt(n),X\xi^{(n),X}_{t} and ξt(n),Y\xi^{(n),Y}_{t} are kept constant by adjusting, respectively, the location parameter μ\mu and a scale parameter of XX, then the skewness and excess kurtosis increase at precisely the same rates as those of XX stated in (2.3), (2.3.5), and (2.3), respectively. Therefore, the constant mixture approach also preserves enlargement effects on asymmetry and tail heaviness of the constituent subordinator XX, regardless of the additional parameters μ\mu, κ1\kappa_{1}, and κ2\kappa_{2}.

Simplicity notwithstanding, a major drawback of this approach, due to linearity, is that it inevitably inherits the path regularity of XX into the processes ξ(n),X\xi^{(n),X}’s, ξ(n),Y\xi^{(n),Y}’s, and ξ(n),Z\xi^{(n),Z}’s, as the next corollary shows.

Corollary 4.

For any n>0n>0, we have 𝔅(ξ(n),X)=𝔅(ξ(n),Y)=𝔅(ξ(n),Z)=𝔅(ξX)=𝔅(X)\mathfrak{B}(\xi^{(n),X})=\mathfrak{B}(\xi^{(n),Y})=\mathfrak{B}(\xi^{(n),Z})=\mathfrak{B}(\xi^{X})=\mathfrak{B}(X).

However, in some situations, e.g., as mentioned in [Todorov and Tauchen, 2011] [62], a suitable model needs to have sample paths of infinite variation. This is clearly not possible via (3.1.1), because by nonnegativity 𝔅(X)[0,1)\mathfrak{B}(X)\in[0,1), thus leading us to consider a second approach.

3.2 Gaussian mixtures

Alternatively, one can construct a real-valued random variable by using it as a scale parameter for an independent Gaussian random variable – a technique called “Gaussian mixture,” which is implemented here by way of using XX to time-change a Brownian motion with drift (see, e.g., [Barndorff-Nielsen, 1997, Sect. 3.1] [7] and [Madan et al., 1998, Sect. 2] [46]). To be precise, let WW be a standard Brownian motion independent of XX, and define the mixed process

ΞtX=μt+θXt+WXt,t0,\varXi^{X}_{t}=\mu t+\theta X_{t}+W_{X_{t}},\quad t\geq 0, (3.2.1)

where μ\mu\in\mathbb{R} is a location parameter and θ\theta\in\mathbb{R} is a Brownian drift parameter representing signed random momenta. The dispersion of the Brownian motion can be fixed at 11 for simplicity as long as XX has a scale parameter, which is already able to capture jump volatility. Then, ΞX\varXi^{X} is also a purely discontinuous real-valued Lévy process. Upon applying the tower property of conditional expectations, the LT of ΞtX\varXi^{X}_{t} for fixed t>0t>0 is found to be

ϕΞtX(u):=EeuΞtX=eμtuϕXt(θuu22),ui.\phi_{\varXi^{X}_{t}}(u):=\textsf{E}e^{-u\varXi^{X}_{t}}=e^{-\mu tu}\phi_{X_{t}}\bigg{(}\theta u-\frac{u^{2}}{2}\bigg{)},\quad u\in{\rm i}\mathbb{R}. (3.2.2)

Again, a partial domain extension into (i)\mathbb{C}\setminus({\rm i}\mathbb{R}) is possible with (2.1).

In a similar fashion, we can time-change the Brownian motion with drift by the regulated clocks, though this would not be equivalent to operating the two recipes on ΞX\varXi^{X}. The resultant processes are denoted as, respectively, Ξ(n),X\varXi^{(n),X}, Ξ(n),Y\varXi^{(n),Y}, and Ξ(n),Z\varXi^{(n),Z}, with the LTs (t>0t>0) states as follows.

Corollary 5.

For generic t>0t>0 and any n>0n>0, the LTs of the Gaussian mixtures subject to clock regulation are given by

ϕΞt(n),X(u)\displaystyle\phi_{\varXi^{(n),X}_{t}}(u) =eμtuϕX¯t(n)(θuu22),\displaystyle=e^{-\mu tu}\phi_{\bar{X}^{(n)}_{t}}\bigg{(}\theta u-\frac{u^{2}}{2}\bigg{)},
ϕΞt(n),Y(u)\displaystyle\phi_{\varXi^{(n),Y}_{t}}(u) =eμtuϕX~t(n)(θuu22),\displaystyle=e^{-\mu tu}\phi_{\tilde{X}^{(n)}_{t}}\bigg{(}\theta u-\frac{u^{2}}{2}\bigg{)},
ϕΞt(n),Z(u)\displaystyle\phi_{\varXi^{(n),Z}_{t}}(u) =eμtuϕX˘t(n)(θuu22),ui.\displaystyle=e^{-\mu tu}\phi_{\breve{X}^{(n)}_{t}}\bigg{(}\theta u-\frac{u^{2}}{2}\bigg{)},\quad u\in{\rm i}\mathbb{R}. (3.2.3)

From Corollary 5, an application of Faà di Bruno’s formula yields the cumulants of Ξt(n),X\varXi^{(n),X}_{t} in terms the finite sums

(1)mdmdumlogϕΞt(n),X(u)|u=0=μt𝟙{1}(m)+k=1mBm,k(θ,1)K¯t(n)(k),m++,(-1)^{m}\frac{{\rm d}^{m}}{{\rm d}u^{m}}\log\phi_{\varXi^{(n),X}_{t}}(u)\bigg{|}_{u=0}=\mu t\mathbb{1}_{\{1\}}(m)+\sum^{m}_{k=1}\mathrm{B}_{m,k}(\theta,1)\bar{K}^{(n)}_{t}(k),\quad m\in\mathbb{N}_{++},

where Bm,k(θ,1)\mathrm{B}_{m,k}(\theta,1)’s are partial Bell polynomials. In particular, by the order of m{1,2,3,4}m\in\{1,2,3,4\} the first four cumulants are: μt+θK¯t(n)(1)\mu t+\theta\bar{K}^{(n)}_{t}(1), K¯t(n)(1)+θ2K¯t(n)(2)\bar{K}^{(n)}_{t}(1)+\theta^{2}\bar{K}^{(n)}_{t}(2), 3θK¯t(n)(2)+θ3K¯t(n)(3)3\theta\bar{K}^{(n)}_{t}(2)+\theta^{3}\bar{K}^{(n)}_{t}(3), and 3K¯t(n)(2)+6θ2K¯t(n)(3)+θ4K¯t(n)(4)3\bar{K}^{(n)}_{t}(2)+6\theta^{2}\bar{K}^{(n)}_{t}(3)+\theta^{4}\bar{K}^{(n)}_{t}(4), in order. Similar results hold for Ξ(n),Y\varXi^{(n),Y} and Ξ(n),Z\varXi^{(n),Z}. Notably, the elegance of the Gaussian mixture (3.2.1) lies in that the additional parameter θ\theta in its own right acts as a scaling factor on XX, so that scaling ΞX\varXi^{X} is fundamentally linked to doing XX, making it a feasible task to control the variance of Ξ(n),X\varXi^{(n),X} (and that of Ξ(n),Y\varXi^{(n),Y} and Ξ(n),Z\varXi^{(n),Z}) to be invariant to clock regulation, while the invariance of the mean is taken care of by the location parameter μ\mu. The consequence is that the skewness and kurtosis relations in (2.3), (2.3.5), and (2.3) set the upper bounds for those (lower-bounded by 1) of the Gaussian-mixed processes, whose exact rates of increase are subject to the value of θ\theta.

The main reason to consider Gaussian mixtures is that they allow for processes with infinite-variation sample paths. Indeed, it can be shown (see Appendix A) that the Blumenthal-Getoor indices of the Gaussian mixtures are precisely double that of XX.

Corollary 6.

For any n>0n>0 we have 𝔅(Ξ(n),X)=𝔅(Ξ(n),Y)=𝔅(Ξ(n),Z)=𝔅(ΞX)=2𝔅(X)\mathfrak{B}\big{(}\varXi^{(n),X}\big{)}=\mathfrak{B}\big{(}\varXi^{(n),Y}\big{)}=\mathfrak{B}\big{(}\varXi^{(n),Z}\big{)}=\mathfrak{B}\big{(}\varXi^{X}\big{)}=2\mathfrak{B}(X).

Therefore, as long as 𝔅(X)1/2\mathfrak{B}(X)\geq 1/2, ΞX\varXi^{X}, and hence Ξ(n),X\varXi^{(n),X}, Ξ(n),Y\varXi^{(n),Y}, and Ξ(n),Z\varXi^{(n),Z}, n>0n>0, on the regulated clocks, will all have jumps of infinite variation. The same idea goes for the Gaussian mixtures on the regulated clocks. This property overcomes the limitation of the constant mixtures and is particularly attractive when flexibility to include highly active jumps in asset returns is required for model formulation.

4 Some important special cases

We now consider two specific distributions of the unregulated clock XX for practical interests. By infinite divisibility we fix time at t=1t=1 unless otherwise specified. Explicit formulae will be presented for the distributions of all three types of regulated clocks, as well as their time-changed processes.

4.1 Poisson process

Let the LT of X1X_{1} be given by ϕX1(u)=exp(λ(eu1))\phi_{X_{1}}(u)=\exp(\lambda(e^{-u}-1)) with shape parameter λ>0\lambda>0 (Poisson intensity). Then we have the following specialized formulae.

Proposition 1.

If XX is a Poisson process with intensity λ\lambda, then for any n>0n>0,

ϕX¯1(n)(u)\displaystyle\phi_{\bar{X}^{(n)}_{1}}(u) =exp(λ(Γ(n)01euzQn1(n,1z)dz1)),\displaystyle=\exp\bigg{(}\lambda\bigg{(}\mathrm{\Gamma}(n)\int^{1}_{0}\frac{e^{-uz}}{\mathrm{Q}^{n-1}(n,1-z)}{\rm d}z-1\bigg{)}\bigg{)},
ϕX~1(n)(u)\displaystyle\phi_{\tilde{X}^{(n)}_{1}}(u) =exp(1Γ(n)01(logv)n1(euv1)dv)=n++exp(λ(nFn(1,,1;2,,2;u)1)),\displaystyle=\exp\bigg{(}\frac{1}{\mathrm{\Gamma}(n)}\int^{1}_{0}(-\log v)^{n-1}(e^{-uv}-1){\rm d}v\bigg{)}\overset{n\in\mathbb{N}_{++}}{=}\exp(\lambda(\;_{n}\mathrm{F}_{n}(1,\dots,1;2,\dots,2;-u)-1)),
ϕX˘1(n)(u)\displaystyle\phi_{\breve{X}^{(n)}_{1}}(u) =exp(λ((Γ(n+1)u)1/nΓ(1/n)Γ(1/n,u/Γ(n+1))n1)),u,\displaystyle=\exp\bigg{(}\lambda\bigg{(}\bigg{(}\frac{\mathrm{\Gamma}(n+1)}{u}\bigg{)}^{1/n}\frac{\mathrm{\Gamma}(1/n)-\mathrm{\Gamma}(1/n,u/\mathrm{\Gamma}(n+1))}{n}-1\bigg{)}\bigg{)},\quad u\in\mathbb{C}, (4.1.1)

where F(;;)\;{}_{\cdot}\mathrm{F}_{\cdot}(\cdots;\cdots;\cdot) denotes the generalized hypergeometric function (see [Slater, 1966] [60]); also,

¯(n)(z)\displaystyle\bar{\ell}^{(n)}(z) =λΓ(n)𝟙(0,1](z)Qn1(n,1z),\displaystyle=\frac{\lambda\mathrm{\Gamma}(n)\mathbb{1}_{(0,1]}(z)}{\mathrm{Q}^{n-1}(n,1-z)},
~(n)(z)\displaystyle\tilde{\ell}^{(n)}(z) =λ(logz)n1𝟙(0,1](z)Γ(n),\displaystyle=\frac{\lambda(-\log z)^{n-1}\mathbb{1}_{(0,1]}(z)}{\mathrm{\Gamma}(n)},
˘(n)(z)\displaystyle\breve{\ell}^{(n)}(z) =λ(Γ(n+1)z)1/n𝟙(0,1/Γ(n+1)](z)nz,z>0.\displaystyle=\frac{\lambda(\mathrm{\Gamma}(n+1)z)^{1/n}\mathbb{1}_{(0,1/\mathrm{\Gamma}(n+1)]}(z)}{nz},\quad z>0. (4.1.2)

It is an easy implication from Proposition 1 that the type-I regulated clock X(n)X^{(n)} is of compound Poisson type: In light of the discussion in Subsection 2.2 and with inverse sampling in mind,

Xt(n)=d.k=1Xt(1Γ(n,logUk)Γ(n)),t0,X^{(n)}_{t}\overset{\text{d.}}{=}\sum^{X_{t}}_{k=1}\bigg{(}1-\frac{\mathrm{\Gamma}(n,-\log U_{k})}{\mathrm{\Gamma}(n)}\bigg{)},\quad t\geq 0, (4.1.3)

where {Uk}k=1\{U_{k}\}^{\infty}_{k=1} is a sequence of i.i.d. standard uniform random variables. The transformed variable representing the jump amplitudes reduces to the standard uniform for n=1n=1 and takes values in [0,1][0,1] for all values of nn. In particular, X(1)X^{(1)} is a compound Poisson-uniform process, which result has been discovered as a special case of fractionally integrated Poisson processes in [Orsingher and Polito, 2013] [52].

The first LT (1) cannot be evaluated explicitly555Apparently, this is caused by the lack of an explicit antiderivative of the function essse^{-s}s^{s}, s[0,1)s\in[0,1). Even so, one can always derive alternative series representations for the log-LT. except for n=1n=1, in which case it is ϕX¯1(1)(u)=exp(λ((1eu)/u1))\phi_{\bar{X}^{(1)}_{1}}(u)=\exp(\lambda((1-e^{-u})/u-1)). Nonetheless, the first integral in (1) is very amenable to numerical computations using a standard Gauss–Kronrod quadrature rule because the integrand is an increasing function (of s[0,1)s\in[0,1)) with values in [eu,1][e^{-u},1] for any u0u\geq 0.

Likewise, under the second recipe, the following representation exists for Y(n)Y^{(n)}:

Yt(n)=d.k=1XteQ(n,1Uk),t0,Y^{(n)}_{t}\overset{\text{d.}}{=}\sum^{X_{t}}_{k=1}e^{-\mathrm{Q}(n,1-U_{k})},\quad t\geq 0, (4.1.4)

which is also a compound Poisson process. Similar to the type I, the jump amplitudes are all valued in [0,1][0,1], and are standard uniformly distributed for n=1n=1.

For the same reason as the first, the second LT in (1) can be efficiently carried out as a numerical integral thanks to the monotonicity and boundedness of the integrand, while an explicit expression has been given for integer regulation degrees. Only for two degree values can it be expressed in terms of more elementary functions. Of course, if n=1n=1, we have the compound Poisson-uniform LT ϕX~1(1)(u)=exp(λ((1eu)/u1))=ϕX¯1(1)(u)\phi_{\tilde{X}^{(1)}_{1}}(u)=\exp(\lambda((1-e^{-u})/u-1))=\phi_{\bar{X}^{(1)}_{1}}(u). If n=2n=2, one can consult [Bateman, 1954, Eq. 4.6.1 and Eq. 4.6.2] [10] to obtain

ϕX~1(2)(u)=exp(λ(logu+Γ(0,u)Γ(1)u1)),\phi_{\tilde{X}^{(2)}_{1}}(u)=\exp\bigg{(}\lambda\bigg{(}\frac{\log u+\mathrm{\Gamma}(0,u)-\mathrm{\Gamma}^{\prime}(1)}{u}-1\bigg{)}\bigg{)},

where Γ(1)0.577216-\mathrm{\Gamma}^{\prime}(1)\approx 0.577216 is known as the Euler–Mascheroni constant.

Due to simplicity, the third LT in (1) (under the third recipe) permits explicit evaluation for any real regulation degree, from which we have the representation

Zt(n)=d.1Γ(n+1)k=1Xt(1Uk)n=d.1Γ(n+1)k=1XtBk,(1/n,1),t0,Z^{(n)}_{t}\overset{\text{d.}}{=}\frac{1}{\mathrm{\Gamma}(n+1)}\sum^{X_{t}}_{k=1}(1-U_{k})^{n}\overset{\text{d.}}{=}\frac{1}{\mathrm{\Gamma}(n+1)}\sum^{X_{t}}_{k=1}B_{k,(1/n,1)},\quad t\geq 0, (4.1.5)

showing that Z(n)Z^{(n)} is a compound Poisson process with 1/Γ(n+1)1/\mathrm{\Gamma}(n+1)-scaled beta(1/n,11/n,1)-distributed jumps. Unlike the other two types, due to such scaling the support of the jump amplitudes of Z(n)Z^{(n)} shrinks with nn, and converges to the origin in its lower limit.

According to (4.1.3), (4.1.4), and (4.1.5), we conclude that when XX is a Poisson process, all three types of regulated clocks follow compound Poisson processes with bounded jumps. These representations provide easy ways for conducting simulations. Also, the two corresponding transformations under the first two recipes form inverse functions of each other, in line with the demonstration by Figure 1.

In financial applications, compound Poisson processes with finite jump amplitudes, coupled with an independent Brownian motion, form adequate jump–diffusion models for log-price dynamics when sudden movements in returns are believed not to be arbitrarily large in reality; see [Yan and Hanson, 2006] [69] and also [Baustian et al., 2017] [12]. These models are built from uniform distributions and are deemed to enlarge the tail heaviness of Gaussian distributions to an extent comparable to normally or double-exponentially distributed jumps (e.g., [Kou, 2002] [40]). Noted that X(1)X^{(1)} is a compound Poisson-uniform process, they can be simply expressed in terms of

ξt(1),X/b+σWt=μt+σWt+k=1XtUkb,t0\xi^{(1),X/b}_{t}+\sigma W_{t}=\mu t+\sigma W_{t}+\sum^{X_{t}}_{k=1}\frac{U_{k}}{b},\quad t\geq 0 (4.1.6)

for a rate (reciprocal scale) parameter b>0b>0 and a Brownian dispersion parameter σ>0\sigma>0. Since the continuous part (μt+σWt)(\mu t+\sigma W_{t}) is immune to regulation by the stability of the Gaussian distribution, clock regulation can still be understood on the entire jump–diffusion process ξt(1),X/b+σWt\xi^{(1),X/b}_{t}+\sigma W_{t}. Viewing from the representations (4.1.3), (4.1.4), and (4.1.5), the effect of regulation is then to assign greater (probabilistic) weights to smaller price jumps, and the consequent jump amplitude distribution, which possesses a strictly decreasing density over a finite domain, is a “give-and-take” between the uniform and the exponential family.

To give an example, let us consider the scaled clock X/bnX/b_{n}, where bnb_{n} is for every n>0n>0 a rate parameter (assuming b0=1b_{0}=1 for simplicity), and derive from it the regulated clocks X(n)X^{(n)}, Y(n)Y^{(n)}, and Z(n)Z^{(n)} accordingly, whose constant mixtures are constructed following Subsection 3.1, with μ=0\mu=0 and κ1=κ2=1\kappa_{1}=\kappa_{2}=1. Then the processes ξ(n),X\xi^{(n),X}, ξ(n),Y\xi^{(n),Y}, and ξ(n),Z\xi^{(n),Z} are centered symmetric compound Poisson processes with intensity λ\lambda and positive and negative jumps equally distributed in magnitude. To fix the variance of ξ(n),X\xi^{(n),X} (at generic time) over nn, it suffices to set bn=C2,nb_{n}=\sqrt{C_{2,n}}, for ξ(n),Y\xi^{(n),Y}, bn=3n/2b_{n}=3^{-n/2}, and for ξ(n),Z\xi^{(n),Z}, bn=1/(2n+1Γ(n+1))b_{n}=1/(\sqrt{2n+1}\mathrm{\Gamma}(n+1)). Then, as nn increases, the domain of jump amplitudes spreads out, albeit bounded, and the probability of smaller jumps also rises, as Figure 3 shows. Therefore, ξ(n),X\xi^{(n),X}, ξ(n),Y\xi^{(n),Y}, and ξ(n),Z\xi^{(n),Z} all constitute a rich class of desirable models, when return jumps are known to be bounded in size but unevenly distributed.

Refer to caption
Refer to caption
Refer to caption
Figure 3: Distorted uniform distributions of (positive) jumps for fixed variance

Moreover, if XX and XX^{\prime} are two independent Poisson processes with possibly different intensities, say λ,λ>0\lambda,\lambda^{\prime}>0, the modified constant mixture666The superscript is expanded to include the process XX^{\prime} because here it is not an exact copy of XX. ξX,X/b\xi^{X,X^{\prime}}/b with parameters b>0b>0, μ=0\mu=0, and κ1=κ2=1\kappa_{1}=\kappa_{2}=1 is seen to be a scaled Skellam process which has jumps of fixed amplitude (equal to 1/b1/b). This type of discrete-valued processes have gained popularity in modeling tick-by-tick price changes with high-frequency data, as pioneered in [Barndorff-Nielsen et al., 2011] [9]. Recent developments have dealt with stochastic-volatility extensions for memory effects; see, e.g., [Kerss et al., 2014] [38], and [Gupta et al., 2020] [33], the latter also having studied associated running average processes beside discussing natural applications to scoring in ball sports. In connection with this, the process X¯(n)X¯(n)\bar{X}^{(n)}-\bar{X}^{\prime(n)} (n++n\in\mathbb{N}_{++}), up to scaling by a positive number, defines exactly the nn-degree running average of ξX\xi^{X} (by linearity) and enables analysis of arbitrarily long-term behaviors of such phenomena (tick price changes versus scoring gaps). On the other hand, the scaled Lévy processes ξ(n),X,X/b\xi^{(n),X,X^{\prime}}/b, ξ(n),Y,Y/b\xi^{(n),Y,Y^{\prime}}/b, and ξ(n),Z,Z/b\xi^{(n),Z,Z^{\prime}}/b of possibly fractional degrees n>0n>0, when applied for the same modeling purposes, allow for imprecision (e.g., ambiguity between change versus no change) in possible outcomes, by assigning positive probability to jumps of amplitudes strictly less than 1/b1/b. In consequence, these possible outcomes are extended to real values to balance out false judgments, but still cannot fall out of their constructional range (tick sizes or score scales). In particular, the extent of imprecision is directly related to nn – if n(0,1)n\in(0,1), ambiguity in favor of no change stays minor. Obviously, in the limit as n0n\searrow 0 the original discrete-valued processes are restored.

4.2 Tempered stable subordinators

Tempered stable distributions are an infinitely divisible family that was initially introduced in [Koponen, 1995] [39] from a procedure incorporating the finite moments property of Gaussian processes into Lévy stable processes, where it bore the name “truncated Lévy flights.” It is most commonly known in its one-sided form, characterized by up to three parameters a>0a>0 (shape), b>0b>0 (rate) and c[0,1)c\in[0,1) (family), with which the corresponding Lévy processes are often called tempered stable subordinators and so are readily usable as stochastic clocks. We refer to [Schoutens, 2003, Sect. 5.3.6] [59], [Rosiński, 2007] [56], and [Küchler and Tappe, 2013] [41] for a comprehensive treatment of these processes; see also [Grabchak, 2021] [31] for their discrete-valued analogy. The tempered stable family includes several commonly used subordinators as special cases, such as the gamma process (c=0c=0) and the inverse Gaussian process (c=1/2c=1/2). If XX is a tempered stable subordinator with parametrization {a,b,c}\{a,b,c\}, its LT (at t=1t=1) is given by

ϕX1(u)={exp(aΓ(c)((u+b)cbc))if c(0,1)(ub+1)aif c=0,u(,b],\phi_{X_{1}}(u)=\begin{cases}\exp(a\mathrm{\Gamma}(-c)((u+b)^{c}-b^{c}))\quad&\text{if }c\in(0,1)\\ \displaystyle\bigg{(}\frac{u}{b}+1\bigg{)}^{-a}\quad&\text{if }c=0,\end{cases}\quad u\in\mathbb{C}\setminus(-\infty,-b],

which is right-continuous at c=0c=0. The Lévy density \ell of XX exists and equals (z)=aebz/zc+1\ell(z)=ae^{-bz}/z^{c+1}, z>0z>0.

For n=1n=1, the indistinguishable order-1 regulated clocks X(1)X^{(1)}, Y(1)Y^{(1)}, and Z(1)Z^{(1)} have been studied in depth in [Xia, 2021] [68], by the name of “average-tempered stable subordinators.” For all other values of n>0n>0, in the presence of the gamma-related functions, for the type I it does not seem possible to further write the corresponding formulae in (1) and (2) explicitly, whereas both in the present form can be computed numerically with high efficiency thanks to the positivity and boundedness of the integrands.

In contrast, the formulae in (1) and (2) for the type-II clocks are reducible to some closed-form expressions provided n++n\in\mathbb{N}_{++}, by exploiting series representations of the integrands; further, due to the constructional simplicity of the type III, its corresponding formulae readily permit explicit evaluation, even for fractional degrees, which results are given in Proposition 2.

Proposition 2.

If XX is a tempered stable subordinator with with parametrization {a,b,c}\{a,b,c\}, then for any n>0n>0,

ϕX~1(n)(u)\displaystyle\phi_{\tilde{X}^{(n)}_{1}}(u) ={exp(abcΓ(c)(n+1Fn(1,,1,c;2,,2;ub)1))if c>0ean(ub+1)a(b/u+1)exp(abuj=2nLij(ub))if c=0,\displaystyle=\begin{cases}\displaystyle\exp\bigg{(}ab^{c}\mathrm{\Gamma}(-c)\bigg{(}\;_{n+1}\mathrm{F}_{n}\bigg{(}1,\dots,1,-c;2,\dots,2;-\frac{u}{b}\bigg{)}-1\bigg{)}\bigg{)}\quad&\text{if }c>0\\ \displaystyle e^{an}\bigg{(}\frac{u}{b}+1\bigg{)}^{-a(b/u+1)}\exp\Bigg{(}\frac{ab}{u}\sum^{n}_{j=2}\mathrm{Li}_{j}\bigg{(}-\frac{u}{b}\bigg{)}\Bigg{)}\quad&\text{if }c=0,\end{cases}
u(,b],n++,\displaystyle\quad u\in\mathbb{C}\setminus(-\infty,-b],\;n\in\mathbb{N}_{++},
ϕX˘t(n)|c>0(u)\displaystyle\phi_{\breve{X}^{(n)}_{t}|c>0}(u) ={exp(atbcΓ(c)(2F1(c,1n;1n+1;ubΓ(n+1))1))if c>0(1+ubΓ(n+1))atexp(atnubΓ(n+2)2F1(1,1n+1;1n+2;ubΓ(n+1)))if c=0,\displaystyle=\begin{cases}\displaystyle\exp\bigg{(}atb^{c}\mathrm{\Gamma}(-c)\bigg{(}\;_{2}\mathrm{F}_{1}\bigg{(}-c,\frac{1}{n};\frac{1}{n}+1;-\frac{u}{b\mathrm{\Gamma}(n+1)}\bigg{)}-1\bigg{)}\bigg{)}\quad&\text{if }c>0\\ \displaystyle\bigg{(}1+\frac{u}{b\mathrm{\Gamma}(n+1)}\bigg{)}^{-at}\exp\bigg{(}\frac{atnu}{b\mathrm{\Gamma}(n+2)}\;_{2}\mathrm{F}_{1}\bigg{(}1,\frac{1}{n}+1;\frac{1}{n}+2;-\frac{u}{b\mathrm{\Gamma}(n+1)}\bigg{)}\bigg{)}\quad&\text{if }c=0,\\ \end{cases}
u(,bΓ(n+1)],n>0,\displaystyle\quad u\in\mathbb{C}\setminus(-\infty,-b\mathrm{\Gamma}(n+1)],\;n>0, (4.2.1)

where Lij()=k=1()kkj\mathrm{Li}_{j}(\cdot)=\sum^{\infty}_{k=1}(\cdot)^{k}k^{-j} the poly-logarithm, and where analytic continuation is understood in each slit region (D0(b))(,b](D_{0}(b))^{\complement}\setminus(-\infty,-b] and (D0(bΓ(n+1)))(,bΓ(n+1)](D_{0}(b\mathrm{\Gamma}(n+1)))^{\complement}\setminus(-\infty,-b\mathrm{\Gamma}(n+1)], n>0n>0, respectively, and

~(n)(z)\displaystyle\tilde{\ell}^{(n)}(z) =abc+1Gn,n+1n+1,0(1,,10,,0,c1|bz),z>0,n++,\displaystyle=ab^{c+1}\mathrm{G}^{n+1,0}_{n,n+1}\bigg{(}\begin{array}[]{cc}1,\dots,1\\ 0,\dots,0,-c-1\end{array}\bigg{|}bz\bigg{)},\quad z>0,\;n\in\mathbb{N}_{++}, (4.2.4)
˘(n)(z)\displaystyle\breve{\ell}^{(n)}(z) =abc(bΓ(n+1)z)1/nΓ(c1/n,bΓ(n+1)z)nz,z>0,n>0,\displaystyle=\frac{ab^{c}(b\mathrm{\Gamma}(n+1)z)^{1/n}\mathrm{\Gamma}(-c-1/n,b\mathrm{\Gamma}(n+1)z)}{nz},\quad z>0,\;n>0, (4.2.5)

where G,,(|)\mathrm{G}^{\cdot,\cdot}_{\cdot,\cdot}(\cdots|\cdot) denotes the Meijer G-function (see the proof in Appendix A).

Note that if n=1n=1, then with F12(1,c;2;u/b)=(bc(u+b)c+1b)/((c+1)u)\;{}_{2}\mathrm{F}_{1}(1,-c;2;-u/b)=(b^{-c}(u+b)^{c+1}-b)/((c+1)u) and j=210\sum^{1}_{j=2}\equiv 0 the elementary formulae follow:

ϕX~1(1)|c>0(u)=expaΓ(c)((u+b)c+1bc((c+1)u+b))(c+1)u,ϕX~1(1)|c=0(u)=ea(ub+1)a(b/u+1),\phi_{\tilde{X}^{(1)}_{1}|c>0}(u)=\exp\frac{a\mathrm{\Gamma}(-c)((u+b)^{c+1}-b^{c}((c+1)u+b))}{(c+1)u},\quad\phi_{\tilde{X}^{(1)}_{1}|c=0}(u)=e^{a}\bigg{(}\frac{u}{b}+1\bigg{)}^{-a(b/u+1)},

as were initially obtained in [Xia, 2021, Thm. 1 and Corol. 2] [68]. For rational values of the family parameter c(0,1)c\in(0,1), it is possible to rewrite (2) in terms of multiply nested sums of poly-logarithms; see, e.g., [Kalmykov and Kniehl, 2009] [36]. However, it is our preference not to straggle into those representations which are extremely lengthy in general, as interest is more for computational efficiency. For example, the following elementary formula applies to n=2n=2 in the special case c=1/2c=1/2, corresponding to the inverse Gaussian process:

ϕX~1(2)|c=12(u)=(12(ub+1+1))8ab3π/(3u)exp(2abπ(4b9u(4(ub+4)ub+1)+1)).\phi_{\tilde{X}^{(2)}_{1}|c=\frac{1}{2}}(u)=\bigg{(}\frac{1}{2}\bigg{(}\sqrt{\frac{u}{b}+1}+1\bigg{)}\bigg{)}^{8a\sqrt{b^{3}\pi}/(3u)}\exp\bigg{(}2a\sqrt{b\pi}\bigg{(}\frac{4b}{9u}\bigg{(}4-\bigg{(}\frac{u}{b}+4\bigg{)}\sqrt{\frac{u}{b}+1}\bigg{)}+1\bigg{)}\bigg{)}.

It can be easily verified using the antiderivative of (logv)uv+b(\log v)\sqrt{uv+b} in v(0,1)v\in(0,1); see [Gradshteyn and Ryzhik, 2007, Eq. 2.725] [32].

Also, we note that none of the Lévy densities given in (4.2.4) is integrable on ++\mathbb{R}_{++}, meaning that the processes Y(n)Y^{(n)}’s all have essentially infinite jump activity, as discussed earlier; in particular, for the tempered stable subordinator 𝔅(X(n))=𝔅(Y(n))=𝔅(Z(n))=𝔅(X)=c[0,1)\mathfrak{B}(X^{(n)})=\mathfrak{B}(Y^{(n)})=\mathfrak{B}(Z^{(n)})=\mathfrak{B}(X)=c\in[0,1). In the special case n=1n=1, the first formula in (4.2.4) reduces to

~(1)(z)=abc+1Γ(c1,bz),z>0,\tilde{\ell}^{(1)}(z)=ab^{c+1}\mathrm{\Gamma}(-c-1,bz),\quad z>0,

corresponding to the Lévy density derived in [Xia, 2021, Corol. 1] [68], while for n2n\geq 2, the formula (4.2.4) stands and cannot be simplified to more elementary functions.777The first formula (4.2.4) is not really useful beyond deciphering the functional form of the induced Lévy densities; more specifically, because the G-function is defined through complex line integration by most computer algebra systems, its implementation is understandably much less computationally efficient relative to its numerical counterpart (via (2) by the quadrature rule), which also accommodates fractional values of nn.

Unlike the other two, the third recipe benefits by having no effect of increasing the computational complexity of the LT when operating on a tempered stable subordinator,888Numerical computations through the formulae (2) and (4.2.4) still intensify with the regulation degree n++n\in\mathbb{N}_{++} which controls the dimensionality of the hypergeometric functions. and by this advantage the formulae (2) can be confidently applied to characteristic function-based optimization problems such as calibrating the parameters of a market model on strings of derivatives prices using numerical Fourier transforms.

Because the LTs ϕX¯t(n)\phi_{\bar{X}^{(n)}_{t}}, ϕX~t(n)\phi_{\tilde{X}^{(n)}_{t}}, and ϕX˘t(n)\phi_{\breve{X}^{(n)}_{t}} (fixing t>0t>0) are all analytic in the complex plane except for a branch point on the negative real axis, the corresponding inverses (i.e., (a posteriori existent) density functions) can be found via a standard deformation argument employing a counterclockwise keyhole contour avoiding the cut extending to -\infty. Consequently, we obtain the following formulae:

fX¯t(n)(x)\displaystyle f_{\bar{X}^{(n)}_{t}}(x) :=𝖯{X¯t(n)dx}dx=1πbduexp(RelogϕX¯t(n)(eiπu)ux)sin(ImlogϕX¯t(n)(eiπu)),\displaystyle:=\frac{\mathsf{P}\big{\{}\bar{X}^{(n)}_{t}\in{\rm d}x\big{\}}}{{\rm d}x}=\frac{1}{\pi}\int^{\infty}_{b}{\rm d}u\;\exp\big{(}\mathrm{Re}\log\phi_{\bar{X}^{(n)}_{t}}(e^{{\rm i}\pi}u)-ux\big{)}\sin\big{(}-\mathrm{Im}\log\phi_{\bar{X}^{(n)}_{t}}(e^{{\rm i}\pi}u)\big{)},
fX~t(n)(x)\displaystyle f_{\tilde{X}^{(n)}_{t}}(x) :=𝖯{X~t(n)dx}dx=1πbduexp(RelogϕX~t(n)(eiπu)ux)sin(ImlogϕX~t(n)(eiπu)),\displaystyle:=\frac{\mathsf{P}\big{\{}\tilde{X}^{(n)}_{t}\in{\rm d}x\big{\}}}{{\rm d}x}=\frac{1}{\pi}\int^{\infty}_{b}{\rm d}u\;\exp\big{(}\mathrm{Re}\log\phi_{\tilde{X}^{(n)}_{t}}(e^{{\rm i}\pi}u)-ux\big{)}\sin\big{(}-\mathrm{Im}\log\phi_{\tilde{X}^{(n)}_{t}}(e^{{\rm i}\pi}u)\big{)},
fX˘t(n)(x)\displaystyle f_{\breve{X}^{(n)}_{t}}(x) :=𝖯{X˘t(n)dx}dx=1πbΓ(n+1)duexp(RelogϕX˘t(n)(eiπu)ux)sin(ImlogϕX˘t(n)(eiπu))\displaystyle:=\frac{\mathsf{P}\big{\{}\breve{X}^{(n)}_{t}\in{\rm d}x\big{\}}}{{\rm d}x}=\frac{1}{\pi}\int^{\infty}_{b\mathrm{\Gamma}(n+1)}{\rm d}u\;\exp\big{(}\mathrm{Re}\log\phi_{\breve{X}^{(n)}_{t}}(e^{{\rm i}\pi}u)-ux\big{)}\sin\big{(}-\mathrm{Im}\log\phi_{\breve{X}^{(n)}_{t}}(e^{{\rm i}\pi}u)\big{)} (4.2.6)

for x,t>0x,t>0, where the real and imaginary parts are efficiently computable according to Theorem 1.999From a computational viewpoint, the formulae in (4.2) are stabler than performing a numerical inverse Fourier transform and allow the density functions to evaluated (or plotted) on a fine grid. In particular, for n++n\in\mathbb{N}_{++} we have, after the substitution u/b1/yu/b\mapsto 1/y

fX~t(n)|c>0(x)\displaystyle f_{\tilde{X}^{(n)}_{t}|c>0}(x) =beatbcΓ(c)π01dyy2exp(atbcΓ(c)Ren+1Fn(1,,1,c;2,,2;1y)bxy)\displaystyle=\frac{be^{-atb^{c}\mathrm{\Gamma}(-c)}}{\pi}\int^{1}_{0}\frac{{\rm d}y}{y^{2}}\exp\bigg{(}atb^{c}\mathrm{\Gamma}(-c)\mathrm{Re}\;_{n+1}\mathrm{F}_{n}\bigg{(}1,\dots,1,-c;2,\dots,2;\frac{1}{y}\bigg{)}-\frac{bx}{y}\bigg{)}
×sin(atbcΓ(c)Imn+1Fn(1,,1,c;2,,2;1y)),x>0\displaystyle\qquad\times\sin\bigg{(}-atb^{c}\mathrm{\Gamma}(-c)\mathrm{Im}\;_{n+1}\mathrm{F}_{n}\bigg{(}1,\dots,1,-c;2,\dots,2;\frac{1}{y}\bigg{)}\bigg{)},\quad x>0 (4.2.7)

and for any n>0n>0, with u/(bΓ(n+1))1/yu/(b\mathrm{\Gamma}(n+1))\mapsto 1/y,

fX˘t(n)|c>0(x)\displaystyle f_{\breve{X}^{(n)}_{t}|c>0}(x) =bΓ(n+1)eatbcΓ(c)π01dyy2exp(atbcΓ(c)Re2F1(c,1n;1n+1;1y)bΓ(n+1)xy)\displaystyle=\frac{b\mathrm{\Gamma}(n+1)e^{-atb^{c}\mathrm{\Gamma}(-c)}}{\pi}\int^{1}_{0}\frac{{\rm d}y}{y^{2}}\exp\bigg{(}atb^{c}\mathrm{\Gamma}(-c)\mathrm{Re}\;_{2}\mathrm{F}_{1}\bigg{(}-c,\frac{1}{n};\frac{1}{n}+1;\frac{1}{y}\bigg{)}-\frac{b\mathrm{\Gamma}(n+1)x}{y}\bigg{)}
×sin(atbcΓ(c)Im2F1(c,1n;1n+1;1y)),x>0.\displaystyle\qquad\times\sin\bigg{(}-atb^{c}\mathrm{\Gamma}(-c)\mathrm{Im}\;_{2}\mathrm{F}_{1}\bigg{(}-c,\frac{1}{n};\frac{1}{n}+1;\frac{1}{y}\bigg{)}\bigg{)},\quad x>0. (4.2.8)

Despite stability, implementation of (4.2) will require some effort for n2n\geq 2 since the hypergeometric functions need to be computed by analytic continuation of the definitional series which are only convergent within D0(1)D_{0}(1), while there is no such issue for (4.2) because of fixed dimensionality of the hypergeometric functions.

With its overall flexibility and explicitness, processes with marginal tempered stable distributions are of great applicability in operations research, especially in the fields of reliability engineering and mathematical finance; we mention [van Noortwijk, 2009] [65], [Wang and Xu, 2015] [66], and [Ye and Xie, 2015] [70] for résumés on related structural degradation models and [Boyarchenko and Levendorskii, 2000] [17] and [Carr et al., 2002] [18] for derivatives pricing theory. For Gaussian mixtures, the path regularity is directly captured by the family parameter c[0,1)c\in[0,1). Recalling that 𝔅(ΞX)=𝔅(Ξ(n),X)=𝔅(Ξ(n),Y)=𝔅(Ξ(n),Z)=2c\mathfrak{B}(\varXi^{X})=\mathfrak{B}(\varXi^{(n),X})=\mathfrak{B}(\varXi^{(n),Y})=\mathfrak{B}(\varXi^{(n),Z})=2c, for any n>0n>0 (by Corollary 6), if c[0,1/2)c\in[0,1/2), then the mixed processes Ξ(n),X\varXi^{(n),X}, Ξ(n),Y\varXi^{(n),Y}, and Ξ(n),Z\varXi^{(n),Z} all have sample paths of finite variation, while the paths are of infinite variation for c[1/2,1)c\in[1/2,1). In the special cases c=0c=0 and c=1/2c=1/2, ΞX\varXi^{X} becomes the variance gamma process and the normal inverse Gaussian process, respectively.101010In [Carr et al., 2002] [18], two-sided tempered stable processes were studied under the name “CGMY,” which include the variance gamma process as a special case. However, for c0c\neq 0 such processes cannot be interpreted as Gaussian mixtures of tempered stable subordinators. With 𝔅(X)=0\mathfrak{B}(X)=0 the sample paths of the variance gamma process are of finite variation while with 𝔅(X)=1\mathfrak{B}(X)=1 those of the normal inverse Gaussian process have infinite variation. In such cases, the processes Ξ(n),X\varXi^{(n),X}, Ξ(n),Y\varXi^{(n),Y}, and Ξ(n),Z\varXi^{(n),Z} derived from stochastic clock regulation are natural extensions of the foregoing models with arbitrarily large skewness and kurtosis, the mean and variance fixed and path regularity unchanged.

Subject to the addition parameters μ,θ\mu,\theta\in\mathbb{R}, the density function of Ξt(n),X\varXi^{(n),X}_{t} can be written either with (3.2.2) as the Fourier inverse

fΞt(n),X(x):=𝖯{Ξt(n),Xdx}dx\displaystyle f_{\varXi^{(n),X}_{t}}(x):=\frac{\mathsf{P}\big{\{}\varXi^{(n),X}_{t}\in{\rm d}x\big{\}}}{{\rm d}x} =1π0Re(eiuxϕΞt(n),X(iu))du\displaystyle=\frac{1}{\pi}\int^{\infty}_{0}\mathrm{Re}\big{(}e^{-{\rm i}ux}\phi_{\varXi^{(n),X}_{t}}(-{\rm i}u)\big{)}{\rm d}u
=1π0Re(ei(μtx)uϕX¯t(n)(iθu+u22))du,x,t>0\displaystyle=\frac{1}{\pi}\int^{\infty}_{0}\mathrm{Re}\bigg{(}e^{{\rm i}(\mu t-x)u}\phi_{\bar{X}^{(n)}_{t}}\bigg{(}-{\rm i}\theta u+\frac{u^{2}}{2}\bigg{)}\bigg{)}{\rm d}u,\quad x\in\mathbb{R},\;t>0 (4.2.9)

or, after applying the Bayes theorem to (4.2), as the marginalization

fΞt(n),X(x)\displaystyle f_{\varXi^{(n),X}_{t}}(x) =0dw2πwe(xθw)2/(2w)fX¯t(n)(w)\displaystyle=\int^{\infty}_{0}\frac{{\rm d}w}{\sqrt{2\pi w}}e^{-(x-\theta w)^{2}/(2w)}f_{\bar{X}^{(n)}_{t}}(w)
=1πbdu2u+θ2exp(RelogϕX¯t(n)(eiπu)+θ(xμt)|xμt|2u+θ2)\displaystyle=\frac{1}{\pi}\int^{\infty}_{b}\frac{{\rm d}u}{\sqrt{2u+\theta^{2}}}\exp\big{(}\mathrm{Re}\log\phi_{\bar{X}^{(n)}_{t}}(e^{{\rm i}\pi}u)+\theta(x-\mu t)-|x-\mu t|\sqrt{2u+\theta^{2}}\big{)}
×sin(ImlogϕX¯t(n)(eiπu)),\displaystyle\qquad\times\sin\big{(}-\mathrm{Im}\log\phi_{\bar{X}^{(n)}_{t}}(e^{{\rm i}\pi}u)\big{)}, (4.2.10)

where the use of Fubini’s theorem in the second equality is under the condition that c<1/2c<1/2;111111It can be shown that the polarity of RelogϕXt(eiπu)\mathrm{Re}\log\phi_{X_{t}}(e^{{\rm i}\pi}u) with XX being a tempered stable subordinator exactly equals that of 2c12c-1, which feature is also unaffected by regulation (nn). similar results hold for the densities fΞt(n),Yf_{\varXi^{(n),Y}_{t}} and fΞt(n),Zf_{\varXi^{(n),Z}_{t}} in terms of ϕX~t(n)\phi_{\tilde{X}^{(n)}_{t}} and ϕX˘t(n)\phi_{\breve{X}^{(n)}_{t}}, respectively.

It is straightforward to keep the mean and variance of the Gaussian mixture Ξ(n),Y\varXi^{(n),Y} invariant to nn with the parametrization {a,b,μ,θ}\{a,b,\mu,\theta\}. One can for example adjust the location and Brownian drift parameters μ\mu and θ\theta according to the rules μn+θnEX1/2n=[target mean]\mu_{n}+\theta_{n}\textsf{E}X_{1}/2^{n}=[\text{target mean}] and EX1/2n+θn2(VarX1)/3n=[target variance]\textsf{E}X_{1}/2^{n}+\theta^{2}_{n}(\textsf{Var}X_{1})/3^{n}=[\text{target variance}];121212Since we know from Subsection 3.2 that the value of θ\theta directly controls the skewness, this approach can very quickly result in abnormally large asymmetry as nn increases. to maintain the moderate enlargement effects in (2.3.5), one can as well adjust the location and (tempered stable) rate parameters μ\mu and bb by the rules μn+θEX1/(2nbn)=[target mean]\mu_{n}+\theta\textsf{E}X_{1}/(2^{n}b_{n})=[\text{target mean}] and EX1/(2nbn)+θ2(VarX1)/(3nbn2)=[target variance]\textsf{E}X_{1}/(2^{n}b_{n})+\theta^{2}(\textsf{Var}X_{1})/(3^{n}b^{2}_{n})=[\text{target variance}], where X1X_{1} has unit scale; alternatively, disregarding the additional parameters μ\mu and θ\theta one can change the shape and rate parameters aa and bb of the unregulated clock XX, as discussed in Subsection 3.2.131313In contrast to the first, this one tends to give milder-than-moderate enlargement effects and so can fit in with a wide range of values of nn, hence also more manageable; see Appendix B for details. Along these lines one can come up with many hybrid methods for fixing the mean and variance, e.g., by changing μ\mu, aa and powers of bb simultaneously. Similar arguments go for Ξ(n),X\varXi^{(n),X} and Ξ(n),Z\varXi^{(n),Z}.

5 Optimal degrees of regulation

It is time to consider the problem of finding an optimal degree nn of regulation with available data. Optimization may be accomplished by many different rules, depending on various aspects of the resultant distributions and data types. Instead of a comprehensive analysis, the aim of this section is to discuss two popular and comfortable methods that can justify values significantly deviating from zero (unregulated case) with due contrast control. The methods by themselves accommodate both the constant mixture ξX\xi^{X} and the Gaussian mixture ΞX\varXi^{X}.

5.1 Moment-based estimation

In financial modeling, it is a standard assumption that a probability distribution of asset returns can be defined by the knowledge of their first several (usually four) moments; see, e.g., [Ané and Geman, 2000] [5]. This suggests that one can use the first sample moments – in particular the sample mean, variance, skewness, and excess kurtosis – to retrieve certain model parameters, and underlies our moment-based estimation procedure, which goes as follows.

We start by fixing a truncated and discretized domain 𝔑+\mathfrak{N}\subsetneq\mathbb{R}_{+} of degrees of regulation, and conduct moment-based estimation of the model parameters for each n𝔑n\in\mathfrak{N}, with n=0n=0 corresponding to the unregulated case. Recall that regulation leads to simple cumulant reduction relations (Corollary 2), with regulation type-specific coefficients

ρ(m,n):={Cm,ntype I1(m+1)ntype II1(mn+1)Γm(n+1)type III,m++,n0,\rho(m,n):=\begin{cases}C_{m,n}\quad&\text{type I}\\ \displaystyle\frac{1}{(m+1)^{n}}\quad&\text{type II}\\ \displaystyle\frac{1}{(mn+1)\mathrm{\Gamma}^{m}(n+1)}\quad&\text{type III},\end{cases}\quad m\in\mathbb{N}_{++},\;n\geq 0, (5.1.1)

so it is most convenient to construct corresponding moment conditions from the model cumulants; for similar estimation procedures in the literature we refer to [Beckers, 1981] [13] and [Bandi and Nguyen, 2003] [6]. After experimenting with every member in 𝔑\mathfrak{N}, the optimal value of nn is identified (a posteriori) based on the maximal profile likelihood as a criterion for model selection. The reason behind choosing moment-based estimation over maximum likelihood estimation is twofold: It is much more computationally efficient because the density function of Lévy models are generally difficult to obtain in closed form (e.g., see [Schoutens, 2003, Chap. 4] [59]), with or without clock regulation; it is also more robust by making fewer assumptions for the underlying data-generating mechanism as only a finite number of moments are utilized.

Suppose that we have obtained an i.i.d. sample of some log-returns {xˇk}k=1N\{\check{x}_{k}\}^{N}_{k=1}, with N1N\gg 1 being the sample size, at a uniform frequency Δ>0\varDelta>0 in fractions of a year; from now on we shall write as [M][M], [V][V], [SK][SK], and [EK][EK] the sample mean, variance, skewness, and excess kurtosis, respectively, and KΔ(n)(m)K^{(n)}_{\varDelta}(m), for m{1,2,3,4}m\in\{1,2,3,4\}, the mixed model cumulants with regulation degrees n0n\geq 0. Then, the four moment conditions can be written

KΔ(n)(1)=[M],KΔ(n)(2)=[V],KΔ(n)(3)=[SK][V]3/2,KΔ(n)(3)KΔ(n)(4)=[SK][EK][V]1/2.K^{(n)}_{\varDelta}(1)=[M],\quad K^{(n)}_{\varDelta}(2)=[V],\quad K^{(n)}_{\varDelta}(3)=[SK][V]^{3/2},\quad\frac{K^{(n)}_{\varDelta}(3)}{K^{(n)}_{\varDelta}(4)}=\frac{[SK]}{[EK][V]^{1/2}}. (5.1.2)

First, we take XX to be a Poisson process and its constant mixture ξX\xi^{X} plus an independent Brownian motion, as discussed in Subsection 4.1. From (4.1.3), (4.1.4), and (4.1.5) we know that in this case the model is of jump–diffusion type with bounded (distorted uniform) jumps. For convenience, we shall focus on downside jumps, in view of excessive tail risk, by imposing the auxiliary parameter values κ1=0\kappa_{1}=0 and κ2=1\kappa_{2}=1 in (3.1.1). Note that the mixed model with clock regulation has a total of four parameters to be estimated, including an infinite-divisibility parameter λn>0\lambda_{n}>0, a rate parameter bn>0b_{n}>0, a location parameter μ\mu\in\mathbb{R}, and a Brownian dispersion parameter σn>0\sigma_{n}>0, all of which are subscripted by the corresponding degree of regulation, n0n\geq 0.

Proposition 3.

Given any regulation degree n0n\geq 0, under the Poisson-based jump–diffusion model (see (4.1.6)) with κ1=0\kappa_{1}=0 and κ2=1\kappa_{2}=1 and parametrization {λn,bn,μn,σn}\{\lambda_{n},b_{n},\mu_{n},\sigma_{n}\}, the moment-based estimators are uniquely determined as

b^n=|[SK]|[EK][V]1/2ρ(4,n)ρ(3,n),λ^n=|[SK]|[V]3/2b^n3Δρ(3,n),σ^n=[V]Δρ(2,n)λ^nb^n2,μ^n=[M]Δ+ρ(1,n)λ^nb^n,\hat{b}_{n}=\frac{|[SK]|}{[EK][V]^{1/2}}\frac{\rho(4,n)}{\rho(3,n)},\quad\hat{\lambda}_{n}=\frac{|[SK]|[V]^{3/2}\hat{b}^{3}_{n}}{\varDelta\rho(3,n)},\quad\hat{\sigma}_{n}=\frac{[V]}{\varDelta}-\rho(2,n)\frac{\hat{\lambda}_{n}}{\hat{b}^{2}_{n}},\quad\hat{\mu}_{n}=\frac{[M]}{\varDelta}+\rho(1,n)\frac{\hat{\lambda}_{n}}{\hat{b}_{n}}, (5.1.3)

where ρ(m,n)\rho(m,n)’s are specified in (5.1.1).

In terms of the above estimates, the profile likelihood function under the mixed model (with an nn-degree regulated clock) can be directly computed using the Bernoulli approximation technique in [Kou, 2002, Eq. (5)] [40] as

g^Δ(n)\displaystyle\hat{g}_{\varDelta}(n) =k=1N(Δb^n2πσ^n2exp((xˇkzμ^nΔ)22σ^n2Δ)n(b^nz)dz\displaystyle=\prod^{N}_{k=1}\bigg{(}\varDelta\int_{\mathbb{R}}\frac{\hat{b}_{n}}{\sqrt{2\pi\hat{\sigma}^{2}_{n}}}\exp\bigg{(}-\frac{(\check{x}_{k}-z-\hat{\mu}_{n}\varDelta)^{2}}{2\hat{\sigma}^{2}_{n}\varDelta}\bigg{)}\ell_{n}(-\hat{b}_{n}z){\rm d}z
+(1λ^nΔ)12πσ^n2exp((xˇkμ^nΔ)22σ^n2Δ)),\displaystyle\quad+(1-\hat{\lambda}_{n}\varDelta)\frac{1}{\sqrt{2\pi\hat{\sigma}^{2}_{n}}}\exp\bigg{(}-\frac{(\check{x}_{k}-\hat{\mu}_{n}\varDelta)^{2}}{2\hat{\sigma}^{2}_{n}\varDelta}\bigg{)}\bigg{)},

provided that Δ\varDelta is small – e.g., as in the case of daily data, where the Lévy density n\ell_{n} represents any one of ¯n\bar{\ell}_{n}, ~n\tilde{\ell}_{n}, and ˘n\breve{\ell}_{n} (with scaling factor λ^n\hat{\lambda}_{n}) in Proposition 1 according to the type of regulation. Whichever model yields the largest value of g^Δ(n)\hat{g}_{\varDelta}(n) is taken as the best.

Second, we let XX be a tempered stable subordinator and construct its Gaussian mixture ΞX\varXi^{X}. The family parameter c[0,1)c\in[0,1) is fixed a priori for its special connection to local path regularity, which leaves us with also four parameters to be estimated – (with regulation degree subscripts) the infinite-divisibility parameter an>0a_{n}>0, rate parameter bn>0b_{n}>0, location parameter μ\mu\in\mathbb{R}, and the Brownian drift parameter θn\theta_{n}\in\mathbb{R}. In this case, the moment-based estimators can be shown to be generally unique under the conditions in (5.1.2).

Proposition 4.

Given any regulation degree n0n\geq 0, under the Gaussian-mixed tempered stable model with c[0,1)c\in[0,1) and parametrization {an,bn,μn,θn}\{a_{n},b_{n},\mu_{n},\theta_{n}\}, the moment-based estimators are given by the conditions

b^n=2(2c)ρ(3,n)θ^n2ρ(2,n)P(|θ^n|),a^n=[V]b^n1cΔΓ(1c)(ρ(1,n)+(1c)ρ(2,n)θ^n2/b^n),\displaystyle\hat{b}_{n}=\frac{2(2-c)\rho(3,n)\hat{\theta}^{2}_{n}}{\rho(2,n)P(|\hat{\theta}_{n}|)},\quad\hat{a}_{n}=\frac{[V]\hat{b}^{1-c}_{n}}{\varDelta\mathrm{\Gamma}(1-c)\big{(}\rho(1,n)+(1-c)\rho(2,n)\hat{\theta}^{2}_{n}/\hat{b}_{n}\big{)}},
μ^n=[M]Δρ(1,n)θ^n[V]Δ(ρ(1,n)+(1c)ρ(2,n)θ^n2/b^n),\displaystyle\hat{\mu}_{n}=\frac{[M]}{\varDelta}-\frac{\rho(1,n)\hat{\theta}_{n}[V]}{\varDelta\big{(}\rho(1,n)+(1-c)\rho(2,n)\hat{\theta}^{2}_{n}/\hat{b}_{n}\big{)}}, (5.1.4)

where θ^n=sgn([SK])|θ^n|\hat{\theta}_{n}=\mathrm{sgn}([SK])|\hat{\theta}_{n}| and |θ^n||\hat{\theta}_{n}| solves the rational equation

(3c)ρ(2,n)ρ(4,n)4(2c)ρ(3,n)2P2(|θn|)+(3[EK]|θn|2|[SK]|)P(|θn|)+3(1[EK]|θn||[SK]|)=0,\frac{(3-c)\rho(2,n)\rho(4,n)}{4(2-c)\rho(3,n)^{2}}P^{2}(|\theta_{n}|)+\bigg{(}3-\frac{[EK]|\theta_{n}|}{2|[SK]|}\bigg{)}P(|\theta_{n}|)+3\bigg{(}1-\frac{[EK]|\theta_{n}|}{|[SK]|}\bigg{)}=0, (5.1.5)

with

P(|θn|)\displaystyle P(|\theta_{n}|) =(|[SK]||θn|3)+D1/2(|θn|),\displaystyle=(|[SK]||\theta_{n}|-3)+D^{1/2}(|\theta_{n}|),
D(|θn|)\displaystyle D(|\theta_{n}|) =(|[SK]||θn|3)2+4(2c)ρ(1,n)ρ(3,n)(1c)ρ(2,n)2|[SK]||θn|.\displaystyle=(|[SK]||\theta_{n}|-3)^{2}+\frac{4(2-c)\rho(1,n)\rho(3,n)}{(1-c)\rho(2,n)^{2}}|[SK]||\theta_{n}|.

With these estimates, the corresponding profile likelihood function is then simply g^Δ(n)=k=1NfΞΔ(n),#(xˇk)\hat{g}_{\varDelta}(n)=\prod^{N}_{k=1}f_{\varXi^{(n),\#}_{\varDelta}}(\check{x}_{k}) (see (4.2) and (4.2)), where #\# is a placeholder for XX, YY, or ZZ depending on the regulation type. Again, the best model is taken to be the one with the largest value of g^Δ(n)\hat{g}_{\varDelta}(n).

5.2 Model calibration on option prices

Alternatively, instead of estimating parameters from sample moments, one can recover them by model calibration on option price data. With the assumption of no-arbitrage, such an exercise will require that the LT be evaluated under some risk-neutral probability measure 𝖰\mathsf{Q} (other than 𝖯\mathsf{P}) under which the asset price is discounted at some (supposedly) constant risk-free rate r0r\geq 0. Allowing for calibration efficiency, here we concentrate on the type-II and type-III regulated clocks which permit some explicit formulae when acting on Gaussian mixtures of tempered stable subordinators (recall Subsection 4.2). By way of the “mean-correcting measure” outlined in [Schoutens, 2003, Sect. 6.2.2] [59], we can then directly assume that the LT of the log asset price at time t>0t>0 under the measure 𝖰\mathsf{Q} is given by

φt(u;n)=(S0ertϕΞt(n),Y(1))uϕΞt(n),Y(u)orφt(u;n)=(S0ertϕΞt(n),Z(1))uϕΞt(n),Z(u),n𝔑,\varphi_{t}(u;n)=\bigg{(}\frac{S_{0}e^{rt}}{\phi_{\varXi^{(n),Y}_{t}}(-1)}\bigg{)}^{-u}\phi_{\varXi^{(n),Y}_{t}}(u)\quad\text{or}\quad\varphi_{t}(u;n)=\bigg{(}\frac{S_{0}e^{rt}}{\phi_{\varXi^{(n),Z}_{t}}(-1)}\bigg{)}^{-u}\phi_{\varXi^{(n),Z}_{t}}(u),\quad n\in\mathfrak{N},

provided that the denominator in the power is finite, where r0r\geq 0 is the constant risk-free rate in the associated money market and 𝔑\mathfrak{N} is the same select domain of the regulation degree nn for experiments.

Then, the no-arbitrage price of a European-style call option written on SS at time t=0t=0 with strike price KK and maturity T>0T>0 fixed is given by

Π0call(n)=S0eqTQ1(n)KerTQ2(n),\varPi^{\text{call}}_{0}(n)=S_{0}e^{-qT}Q_{1}(n)-Ke^{-rT}Q_{2}(n), (5.2.1)

where q>0q>0 is the (constant) dividend yield and

Q1(n)=12+1π0ReKiuφT(iu1;n)iuφT(1;n)du,Q2(n)=12+1π0ReKiuφT(iu;n)iuduQ_{1}(n)=\frac{1}{2}+\frac{1}{\pi}\int^{\infty}_{0}\mathrm{Re}\frac{K^{-{\rm i}u}\varphi_{T}(-{\rm i}u-1;n)}{{\rm i}u\varphi_{T}(-1;n)}{\rm d}u,\quad Q_{2}(n)=\frac{1}{2}+\frac{1}{\pi}\int^{\infty}_{0}\mathrm{Re}\frac{K^{-{\rm i}u}\varphi_{T}(-{\rm i}u;n)}{{\rm i}u}{\rm d}u

are associated risk-neutralized in-the-money probabilities.

Suppose that we have obtained contemporaneous market quotes of MM call options on SS for various strike prices and maturities, denoted as Πˇcall\check{\varPi}^{\text{call}}. Focusing on the Gaussian mixed model ΞX\varXi^{X} where XX is the tempered stable subordinator with a fixed family parameter c[0,1)c\in[0,1), then after mean correction there will be only three parameters, ana_{n}, bnb_{n}, and θn\theta_{n} (with degree subscripts as before), whose (locally) optimal values can be found by solving the following minimization problem:

{a^n,b^n,θ^n}=argmina>0,b>0,θ1MK,T|Πˇ0callΠ0call(n)|Πˇ0call,\{\hat{a}_{n},\hat{b}_{n},\hat{\theta}_{n}\}=\underset{a>0,b>0,\theta\in\mathbb{R}}{\arg\min}\frac{1}{M}\sum_{K,T}\frac{\big{|}\check{\varPi}^{\text{call}}_{0}-\varPi^{\text{call}}_{0}(n)\big{|}}{\check{\varPi}^{\text{call}}_{0}}, (5.2.2)

where the objective function is the mean absolute percentage error (MAPE) measuring average pricing errors and the summation runs over all available strikes and maturities. The best model is then characterized by imposing the criterion minn𝔑\min_{n\in\mathfrak{N}} on the right side of (5.2.2). A similar exercise using put options is immediate by applying a standard parity argument to (5.2.1). Nevertheless, compared with moment-based estimation, calibration is expected to be less robust because it generally constitutes a highly non-convex optimization problem which must be solved numerically.

6 Empirical modeling

In this section we demonstrate the effect of clock regulation on real financial returns. We focus on the equity market and the cryptocurrency market; in particular, we have collected S&P500 index and Bitcoin price data on the daily basis over the 2020 calendar year (data source: Yahoo Finance) and have computed corresponding log-returns. The two data sets hence have sample sizes (NN) of 252 and 366, respectively, with Δ=1/252,1/366\varDelta=1/252,1/366.

In accordance with the moment-based estimation procedures discussed in Subsection 5.1, our empirical study will feature two exercises, one using the constant mixture of a Poisson process augmented by an independent Brownian motion and the other using the Gaussian mixture of a tempered stable subordinator, respectively. We highlight that the first exercise is interesting for modeling (presumably) bounded jumps in asset returns ([Yan and Hanson, 2006] [69] and [Baustian et al., 2017] [12]) and aims to compare the distorted jump distributions with commonly used uniform distributions; the second, on the other hand, is especially significant for possible improvement of tempered stable processes ([Rosiński, 2007] [56] and [Küchler and Tappe, 2013] [41]) in terms of flexibility, by investigating how the degree nn of regulation affects the model fit subject to moment matching for different choices of the family parameter cc.

First, we report the four sample statistics of the collected return data in Table 1. Relative to the S&P500, Bitcoin returns have clearly exhibited a much larger (absolute) skewness and excess kurtosis, speaking to high asymmetrical and tail risk.

Table 1: Descriptive statistics of daily returns (\simsix significant digits)
Data set mean variance skewness excess kurtosis
S&P500 0.000564705 0.00047912 0.861012-0.861012 8.46843
Bitcoin 0.00384156 0.00160601 4.07498-4.07498 50.8679

In both exercises, the choice of the degree domain is 𝔑[0,10]\mathfrak{N}\subsetneq[0,10] with fixed modulus |𝔑|=11|\mathfrak{N}|=11 to ease comparison. In the first exercise, the exponential jump–diffusion model ([Kou, 2002] [40]) is also included as a contrast model because the distorted uniform distributions due to regulation have exponential shapes. In the second exercise, we experiment with four values of cc: 0 (variance gamma), 0.25, 0.5 (normal inverse Gaussian), and 0.75. Estimation results are presented in ten (sub)tables – Table 2, Table 3, Table 4, Table 5, Table 6, Table 7, Table 8, Table 9, Table 10, and Table 11, and are comprised of parameter estimates and the profile log-likelihood (PLL), logg^Δ(n)\log\hat{g}_{\varDelta}(n), for every n𝔑n\in\mathfrak{N}. In each model setting (including each chosen value of cc), the model with the highest PLL value for a certain regulation type is marked “\ast” and that with the highest PLL value taking into account all three types is marked “\star.” To enhance visual impact, changes in the PLL are plotted separately in Figure 4, Figure 6, and Figure 7.

Furthermore, we provide Figure 5, Figure 8, and Figure 9 to compare the estimated density functions of the best fit model (namely the one with the highest PLL value across all three types) and the original model (n=0n=0) against the empirical kernel density function which is estimated using the Gaussian kernel, i.e.,

fˇΔ(x)=1N𝔟k=1N12πe(xxˇk)2/(2𝔟2),x;\check{f}_{\varDelta}(x)=\frac{1}{N\mathfrak{b}}\sum^{N}_{k=1}\frac{1}{\sqrt{2\pi}}e^{-(x-\check{x}_{k})^{2}/(2\mathfrak{b}^{2})},\quad x\in\mathbb{R};

the optimal bandwidth 𝔟\mathfrak{b} is determined by the Silverman rule of thumb:

𝔟=0.9N1/5min([V]1/2,IQR(xˇ)1.349),\mathfrak{b}=\frac{0.9}{N^{1/5}}\min\bigg{(}[V]^{1/2},\frac{\mathrm{IQR}(\check{x})}{1.349}\bigg{)},

where IQR\mathrm{IQR} is the interquartile range (the difference between the third and the first quartiles).

*

Table 2: Jump diffusion estimation results on S&P500 daily returns (\simsix significant digits)
type I nn λ^n\hat{\lambda}_{n} b^n\hat{b}_{n} μ^n\hat{\mu}_{n} σ^n\hat{\sigma}_{n} PLL
0 0.228049 4.64498 0.191402 0.331917 605.28
1 0.467045 3.71599 0.205148 0.330853 620.36
2 1.03524 3.37102 0.219081 0.330044 621.566
3 2.27627 3.11901 0.233531 0.329334 622.22
4 4.98085 2.91283 0.249179 0.328666 622.669
5 10.8658 2.73578 0.266423 0.328018 623.007
6 23.6553 2.5796 0.285589 0.327377 623.276
7 51.4231 2.43945 0.306992 0.326739 623.496
8 111.664 2.31218 0.330954 0.326099 623.68
9 242.275 2.19558 0.357827 0.325455 623.837
10 525.314 2.08804 0.387992 0.324803 623.972\ast
type II nn λ^n\hat{\lambda}_{n} b^n\hat{b}_{n} μ^n\hat{\mu}_{n} σ^n\hat{\sigma}_{n} PLL
0 0.228049 4.64498 0.191402 0.331917 605.28
1 0.467045 3.71599 0.205148 0.330853 620.36
2 0.956508 2.97279 0.222744 0.329716 621.973
3 1.95893 2.37823 0.245267 0.328497 622.917
4 4.01189 1.90259 0.274096 0.327193 623.557
5 8.21634 1.52207 0.310998 0.325796 624.018
6 16.8271 1.21765 0.358232 0.3243 624.359
7 34.4618 0.974124 0.418691 0.322695 624.612
8 70.5779 0.779299 0.496079 0.320976 624.798
9 144.543 0.623439 0.595135 0.319131 624.935
10 296.025 0.498751 0.721927 0.317151 625.046\star
type III nn λ^n\hat{\lambda}_{n} b^n\hat{b}_{n} μ^n\hat{\mu}_{n} σ^n\hat{\sigma}_{n} PLL
0 0.228049 4.64498 0.191402 0.331917 605.28
1 0.467045 3.71599 0.205148 0.330853 620.36
2 0.751093 1.80638 0.211606 0.330498 620.954
3 1.038 0.595511 0.214933 0.330338 621.161
4 1.32573 0.148002 0.216952 0.330247 621.267
5 1.6138 0.029492 0.218306 0.330189 621.33
6 1.90206 0.00490304 0.219277 0.330148 621.373
7 2.19041 0.000699163 0.220007 0.330118 621.404
8 2.47883 0.000087275 0.220576 0.330095 621.427
9 2.7673 9.68674×1069.68674\times 10^{-6} 0.221031 0.330077 621.445
10 3.05579 9.6783×1079.6783\times 10^{-7} 0.221404 0.330062 621.459 \ast
exponential - 2.43253 18.5799 0.273228 0.326566 624.048

PLL: profile log-likelihood

Table 3: Jump diffusion estimation results on Bitcoin daily returns (\simsix significant digits)
type I nn λ^n\hat{\lambda}_{n} b^n\hat{b}_{n} μ^n\hat{\mu}_{n} σ^n\hat{\sigma}_{n} PLL
0 0.766746 1.99897 1.78958 0.62922 641.506
1 1.5703 1.59918 1.89698 0.618971 736.065
2 3.48068 1.45072 2.00583 0.611076 736.11
3 7.65327 1.34227 2.11873 0.604087 736.174
4 16.7466 1.25354 2.24098 0.597444 736.2\star
5 36.533 1.17735 2.3757 0.590935 736.171
6 79.5338 1.11014 2.52544 0.584454 736.08
7 172.894 1.04982 2.69265 0.577936 735.924
8 375.437 0.995048 2.87986 0.571337 735.702
9 814.575 0.944869 3.08981 0.564624 735.414
10 1766.21 0.89859 3.32547 0.557771 735.063
type II nn λ^n\hat{\lambda}_{n} b^n\hat{b}_{n} μ^n\hat{\mu}_{n} σ^n\hat{\sigma}_{n} PLL
0 0.766746 1.99897 1.78958 0.62922 641.506
1 1.5703 1.59918 1.89698 0.618971 736.065
2 3.21597 1.27934 2.03445 0.607849 736.076
3 6.5863 1.02347 2.21041 0.595756 736.101\ast
4 13.4887 0.81878 2.43565 0.582581 735.963
5 27.6249 0.655024 2.72395 0.568191 735.511
6 56.5759 0.524019 3.09297 0.552428 734.623
7 115.867 0.419215 3.56532 0.535103 733.21
8 237.297 0.335372 4.16992 0.515982 731.201
9 485.983 0.268298 4.94382 0.494772 728.473
10 995.294 0.214638 5.9344 0.471097 724.378
type III nn λ^n\hat{\lambda}_{n} b^n\hat{b}_{n} μ^n\hat{\mu}_{n} σ^n\hat{\sigma}_{n} PLL
0 0.766746 1.99897 1.78958 0.62922 641.506
1 1.5703 1.59918 1.89698 0.618971 736.065
2 2.52532 0.777379 1.94743 0.615517 736.092
3 3.48997 0.256279 1.97342 0.613954 736.116
4 4.45737 0.0636928 1.9892 0.613067 736.131
5 5.42592 0.0126919 1.99977 0.612496 736.141
6 6.39508 0.00211003 2.00736 0.612097 736.148
7 7.36459 0.000300886 2.01306 0.611803 736.153
8 8.33432 0.0000375589 2.01751 0.611578 736.157
9 9.30419 4.1687×1064.1687\times 10^{-6} 2.02107 0.611399 736.16
10 10.2742 4.16507×1074.16507\times 10^{-7} 2.02398 0.611254 736.162\ast
exponential - 8.17863 7.9959 2.42886 0.576157 735.574

PLL: profile log-likelihood

Refer to caption
Refer to caption
Figure 4: Jump diffusion log-likelihood with regulation degrees for S&P500 (left) and Bitcoin (right) daily returns
Refer to caption
Refer to caption
Figure 5: Jump diffusion estimated densities with regulation degrees for S&P500 (top) and Bitcoin (bottom) daily returns

*

Table 4: Tempered stable estimation results on S&P500 daily returns (\simsix significant digits)
c=0c=0 (VG)
type I nn a^n\hat{a}_{n} b^n\hat{b}_{n} μ^n\hat{\mu}_{n} θ^n\hat{\theta}_{n} PLL
0 94.8367 811.158 0.744442 5.1502-5.1502 654.012
1 128.456 549.545 0.752422 5.22027-5.22027 662.425
2 203.096 434.589 0.760763 5.29356-5.29356 666.79
3 334.145 357.634 0.7691 5.36685-5.36685 668.065\ast
4 561.015 300.338 0.777794 5.44332-5.44332 667.986
5 954.266 255.534 0.787051 5.52481-5.52481 667.252
6 1638.64 219.491 0.797029 5.6127-5.6127 666.333
7 2835.01 189.959 0.807875 5.70831-5.70831 665.136
8 4935.86 165.446 0.81974 5.81298-5.81298 662.129
9 8641.45 144.906 0.832792 5.92824-5.92824 662.009
10 15206.5 127.574 0.847223 6.05579-6.05579 660.34
type II nn a^n\hat{a}_{n} b^n\hat{b}_{n} μ^n\hat{\mu}_{n} θ^n\hat{\theta}_{n} PLL
0 94.8367 811.158 0.744442 5.1502-5.1502 654.012
1 128.456 549.545 0.752422 5.22027-5.22027 662.425
2 174.38 373.157 0.761651 5.30136-5.30136 666.218
3 237.339 254.059 0.772375 5.39565-5.39565 667.676\ast
4 324.016 173.515 0.784904 5.50592-5.50592 667.625
5 443.949 118.945 0.799644 5.63577-5.63577 666.73
6 610.907 81.8998 0.817127 5.78997-5.78997 665.336
7 845.057 56.6956 0.838074 5.97499-5.97499 663.626
8 1176.48 39.5076 0.863484 6.19982-6.19982 661.708
9 1651.07 27.7586 0.894794 6.47743-6.47743 659.645
10 2340.97 19.7109 0.93414 6.82722-6.82722 657.469
type III nn a^n\hat{a}_{n} b^n\hat{b}_{n} μ^n\hat{\mu}_{n} θ^n\hat{\theta}_{n} PLL
0 94.8367 811.158 0.744442 5.1502-5.1502 654.012
1 128.456 549.545 0.752422 5.22027-5.22027 662.425
2 174.86 249.402 0.756665 5.25754-5.25754 665.585
3 223.022 79.5317 0.758921 5.27736-5.27736 666.776
4 271.766 19.384 0.760309 5.28957-5.28957 667.32
5 320.774 3.81341 0.761249 5.29782-5.29782 667.615
6 369.923 0.628261 0.761926 5.30377-5.30377 667.795
7 419.157 0.0889868 0.762437 5.30826-5.30826 667.915
8 468.446 0.0110503 0.762836 5.31177-5.31177 667.999
9 517.773 0.0012214 0.763157 5.31459-5.31459 668.061
10 567.126 0.000121622 0.76342 5.31691-5.31691 668.108\star

VG: variance gamma — PLL: profile log-likelihood

Table 5: Tempered stable estimation results on S&P500 daily returns (\simsix significant digits)
c=0.25c=0.25
type I nn a^n\hat{a}_{n} b^n\hat{b}_{n} μ^n\hat{\mu}_{n} θ^n\hat{\theta}_{n} PLL
0 11.8708 621.521 0.755058 5.24342-5.24342 668.251
1 17.7616 422.34 0.764691 5.32808-5.32808 668.672\star
2 29.8475 335.05 0.774806 5.41703-5.41703 668.277
3 51.6775 276.597 0.784966 5.50645-5.50645 667.305
4 90.8543 233.057 0.795618 5.60026-5.60026 666.07
5 161.326 198.997 0.807023 5.70079-5.70079 664.739
6 288.564 171.591 0.819393 5.80991-5.80991 661.032
7 519.198 149.132 0.832928 5.92943-5.92943 652.583
8 938.875 130.495 0.847844 6.06127-6.06127 660.391
9 1705.49 114.887 0.864387 6.20765-6.20765 659.088
10 3111.25 101.731 0.882844 6.37118-6.37118 657.695
type II nn a^n\hat{a}_{n} b^n\hat{b}_{n} μ^n\hat{\mu}_{n} θ^n\hat{\theta}_{n} PLL
0 11.8708 621.521 0.755058 5.24342-5.24342 668.251
1 17.7616 422.34 0.764691 5.32808-5.32808 668.672\star
2 26.6293 287.789 0.775895 5.42662-5.42662 668.241
3 40.0192 196.745 0.789001 5.54199-5.54199 667.224
4 60.311 135.026 0.804439 5.67803-5.67803 665.812
5 91.1977 93.0987 0.82278 5.83985-5.83985 664.09
6 138.465 64.553 0.844796 6.03439-6.03439 662.253
7 211.286 45.0713 0.871567 6.27138-6.27138 660.317
8 324.434 31.744 0.904647 6.56486-6.56486 657.45
9 502.195 22.6087 0.946369 6.93602-6.93602 655.531
10 785.628 16.3417 1.0004 7.41838-7.41838 655.78
type III nn a^n\hat{a}_{n} b^n\hat{b}_{n} μ^n\hat{\mu}_{n} θ^n\hat{\theta}_{n} PLL
0 11.8708 621.521 0.755058 5.24342-5.24342 668.251
1 17.7616 422.34 0.764691 5.32808-5.32808 668.672\star
2 29.4917 191.979 0.769827 5.37324-5.37324 668.582
3 50.086 61.2722 0.772562 5.39729-5.39729 668.423
4 86.8967 14.9415 0.774248 5.41212-5.41212 668.29
5 154.046 2.94048 0.775389 5.42216-5.42216 668.186
6 278.893 0.484571 0.776211 5.42939-5.42939 668.103
7 515.191 0.0686477 0.776833 5.43486-5.43486 668.038
8 970.033 0.00852589 0.777318 5.43913-5.43913 667.985
9 1859.66 0.000942492 0.777708 5.44256-5.44256 667.941
10 3626.32 0.0000938585 0.778028 5.44538-5.44538 667.904

PLL: profile log-likelihood

Table 6: Tempered stable estimation results on S&P500 daily returns (\simsix significant digits)
c=0.5c=0.5 (NIG)
type I nn a^n\hat{a}_{n} b^n\hat{b}_{n} μ^n\hat{\mu}_{n} θ^n\hat{\theta}_{n} PLL
0 1.3704 432.849 0.777245 5.43847-5.43847 667.041\star
0.5 1.72512 343.139 0.783233 5.49118-5.49118 666.37
1 2.26522 296.003 0.790516 5.5553-5.5553 665.447
1.5 3.01643 262.598 0.797581 5.61754-5.61754 664.544
2 4.04629 236.4 0.804581 5.67924-5.67924 663.674
2.5 5.45309 214.807 0.811644 5.74153-5.74153 662.832
3 7.37333 196.483 0.818859 5.80518-5.80518 662.016
3.5 9.99496 180.635 0.826291 5.87079-5.87079 661.22
4 13.5762 166.743 0.833993 5.93881-5.93881 660.442
4.5 18.472 154.447 0.842012 6.00967-6.00967 659.68
5 25.17 143.48 0.850391 -6.08376 658.933
type II nn a^n\hat{a}_{n} b^n\hat{b}_{n} μ^n\hat{\mu}_{n} θ^n\hat{\theta}_{n} PLL
0 1.3704 432.849 0.777245 5.43847-5.43847 667.041\star
0.5 1.76152 357.786 0.783616 5.49455-5.49455 666.261
1 2.26522 296.003 0.790516 5.5553-5.5553 665.447
1.5 2.9143 245.126 0.797999 5.62123-5.62123 664.604
2 3.75125 203.209 0.806132 5.69292-5.69292 663.738
2.5 4.83121 168.655 0.814989 5.77105-5.77105 662.853
3 6.22584 140.155 0.824658 5.85639-5.85639 661.951
3.5 8.02838 116.634 0.835242 5.94987-5.94987 661.036
4 10.3604 97.2105 0.846859 6.05256-6.05256 660.108
4.5 13.3806 81.1608 0.859651 6.16575-6.16575 659.17
5 17.2969 67.8906 0.87379 6.29097-6.29097 658.222
type III nn a^n\hat{a}_{n} b^n\hat{b}_{n} μ^n\hat{\mu}_{n} θ^n\hat{\theta}_{n} PLL
0 1.3704 432.849 0.777245 5.43847-5.43847 667.041\star
0.5 1.68675 371.306 0.784277 5.50037-5.50037 666.26
1 2.26522 296.003 0.790516 5.5553-5.5553 665.447
1.5 3.173 210.414 0.7947 5.59216-5.59216 664.868
2 4.58808 135.007 0.797632 5.61799-5.61799 664.453
2.5 6.81669 79.3114 0.799788 5.63699-5.63699 664.144
3 10.3757 43.1666 0.801435 5.65151-5.65151 663.907
3.5 16.1436 21.9721 0.802734 5.66296-5.66296 663.719
4 25.6301 10.5381 0.803784 5.67221-5.67221 663.567
4.5 41.459 4.7913 0.804649 5.67984-5.67984 663.442
5 68.2414 2.07545 0.805375 5.68624-5.68624 663.337

NIG: normal inverse Gaussian — PLL: profile log-likelihood

Table 7: Tempered stable estimation results on S&P500 daily returns (\simsix significant digits)
c=0.75c=0.75
type I nn a^n\hat{a}_{n} b^n\hat{b}_{n} μ^n\hat{\mu}_{n} θ^n\hat{\theta}_{n} PLL
0 0.127463 248.604 0.852286 6.10039-6.10039 654.027\star
0.1 0.134541 233.914 0.853433 6.11053-6.11053 653.976
0.2 0.142421 222.685 0.855778 6.13128-6.13128 653.863
0.3 0.151034 213.54 0.858568 6.15597-6.15597 653.721
0.4 0.160364 205.785 0.861531 6.18219-6.18219 653.566
0.5 0.170423 199.021 0.864556 6.20896-6.20896 653.405
0.6 0.181235 193.002 0.867597 6.23588-6.23588 653.242
0.7 0.192836 187.563 0.870632 6.26276-6.26276 653.079
0.8 0.205268 182.593 0.873656 6.28954-6.28954 652.917
0.9 0.218579 178.007 0.876666 6.3162-6.3162 652.757
1 0.232824 173.745 0.879663 6.34276-6.34276 652.599
type II nn a^n\hat{a}_{n} b^n\hat{b}_{n} μ^n\hat{\mu}_{n} θ^n\hat{\theta}_{n} PLL
0 0.127463 248.604 0.852286 6.10039-6.10039 654.027\star
0.1 0.135368 239.776 0.854794 6.12257-6.12257 653.885
0.2 0.143766 231.278 0.857349 6.14518-6.14518 653.743
0.3 0.152687 223.097 0.859953 6.16822-6.16822 653.601
0.4 0.162164 215.22 0.862607 6.19171-6.19171 653.458
0.5 0.172233 207.637 0.865313 6.21566-6.21566 653.316
0.6 0.182929 200.335 0.868072 6.24008-6.24008 653.173
0.7 0.194293 193.306 0.870884 6.26499-6.26499 653.03
0.8 0.206367 186.537 0.873753 6.29039-6.29039 652.887
0.9 0.219194 180.02 0.876679 6.31631-6.31631 652.743
1 0.232824 173.745 0.879663 6.34276-6.34276 652.599
type III nn a^n\hat{a}_{n} b^n\hat{b}_{n} μ^n\hat{\mu}_{n} θ^n\hat{\theta}_{n} PLL
0 0.127463 248.604 0.852286 6.10039-6.10039 654.027\star
0.1 0.132234 240.172 0.853793 6.11372-6.11372 653.96
0.2 0.13829 233.905 0.856772 6.14007-6.14007 653.818
0.3 0.14551 228.168 0.86014 6.16988-6.16988 653.65
0.4 0.153881 222.222 0.863506 6.19967-6.19967 653.475
0.5 0.163442 215.729 0.86672 6.22812-6.22812 653.305
0.6 0.174271 208.561 0.869727 6.25475-6.25475 653.143
0.7 0.186477 200.706 0.872516 6.27944-6.27944 652.991
0.8 0.200194 192.219 0.875092 6.30226-6.30226 652.85
0.9 0.21558 183.194 0.877469 6.32332-6.32332 652.72
1 0.232824 173.745 0.879663 6.34276-6.34276 652.599

PLL: profile log-likelihood

*

Table 8: Tempered stable estimation results on Bitcoin daily returns (\simsix significant digits)
c=0c=0 (VG)
type I nn a^n\hat{a}_{n} b^n\hat{b}_{n} μ^n\hat{\mu}_{n} θ^n\hat{\theta}_{n} PLL
0 27.7898 56.0074 3.00165 3.21585-3.21585 455.411
1 39.9891 40.7099 3.10861 3.46658-3.46658 524.494
2 67.5402 34.7769 3.23098 3.75877-3.75877 601.924
3 119.117 31.0558 3.36652 4.08912-4.08912 655.587
4 215.901 28.5552 3.52426 4.48257-4.48257 691.131
5 400.41 26.9375 3.71367 4.96792-4.96792 715.837
6 759.406 26.0819 3.94698 5.58529-5.58529 729.963
7 1474.97 25.9914 4.24121 6.39498-6.39498 737.741
8 2941.94 26.7808 4.62072 7.49154-7.49154 741.75
9 6047.17 28.7032 5.12027 9.02652-9.02652 742.155\star
10 12850.6 32.2046 5.78719 11.2431-11.2431 740.616
type II nn a^n\hat{a}_{n} b^n\hat{b}_{n} μ^n\hat{\mu}_{n} θ^n\hat{\theta}_{n} PLL
0 27.7898 56.0074 3.00165 3.21585-3.21585 455.411
1 39.9891 40.7099 3.10861 3.46658-3.46658 524.494
2 58.4843 30.162 3.24722 3.79825-3.79825 591.811
3 87.4167 22.9374 3.43303 4.25497-4.25497 645.474
4 134.707 18.1054 3.69371 4.91968-4.91968 691.284
5 217.137 15.123 4.08334 5.96699-5.96699 717.942
6 375.668 13.859 4.72003 7.82461-7.82461 730.62
7 730.375 14.9641 5.89275 11.7664-11.7664 730.654\ast
8 1689.11 21.4225 8.28461 22.3332-22.3332 716.365
9 4447.82 42.7097 12.7363 55.7046-55.7046 675.673
10 11431.1 114.716 19.1751 182.6-182.6 558.882
type III nn a^n\hat{a}_{n} b^n\hat{b}_{n} μ^n\hat{\mu}_{n} θ^n\hat{\theta}_{n} PLL
0 27.7898 56.0074 3.00165 3.21585-3.21585 455.411
1 39.9891 40.7099 3.10861 3.46658-3.46658 524.494
2 56.2539 19.1976 3.16875 3.60939-3.60939 573.498
3 73.0303 6.24998 3.20178 3.6884-3.6884 598.04
4 89.9752 1.543 3.22249 3.73814-3.73814 613.948
5 106.996 0.306219 3.23666 3.77226-3.77226 623.336
6 124.059 0.05077 3.24696 3.79711-3.79711 630.002
7 141.146 0.00722557 3.25478 3.816-3.816 635.146
8 158.249 0.000900634 3.26092 3.83085-3.83085 639.394
9 175.363 0.0000998491 3.26587 3.84282-3.84282 642.235
10 192.485 9.9672×1069.9672\times 10^{-6} 3.26994 3.85268-3.85268 644.439\ast

VG: variance gamma — PLL: profile log-likelihood

Table 9: Tempered stable estimation results on Bitcoin daily returns (\simsix significant digits)
c=0.25c=0.25
type I nn a^n\hat{a}_{n} b^n\hat{b}_{n} μ^n\hat{\mu}_{n} θ^n\hat{\theta}_{n} PLL
0 7.17612 47.0281 3.1418 3.54481-3.54481 641.25
1 11.3467 35.0184 3.28923 3.8994-3.8994 674.554
2 20.2654 30.7629 3.46274 4.32733-4.32733 703.597
3 37.443 28.3513 3.66136 4.83148-4.83148 722.383
4 70.7801 27.0578 3.90086 5.45996-5.45996 734.063
5 136.543 26.7025 4.19975 6.27619-6.27619 740.223
6 268.857 27.3237 4.5831 7.3758-7.3758 743.733
7 541.024 29.136 5.08604 8.91013-8.91013 744.02\star
8 1113.88 32.5598 5.756 11.1203-11.1203 741.869
9 2344.96 38.2747 6.65067 14.3797-14.3797 737.46
10 5028.11 47.2474 7.82714 19.2312-19.2312 730.984
type II nn a^n\hat{a}_{n} b^n\hat{b}_{n} μ^n\hat{\mu}_{n} θ^n\hat{\theta}_{n} PLL
0 7.17612 47.0281 3.1418 3.54481-3.54481 641.25
1 11.3467 35.0184 3.28923 3.8994-3.8994 674.554
2 18.2308 26.8011 3.48754 4.39002-4.39002 701.452
3 29.9485 21.3312 3.76679 5.10797-5.10797 721.23
4 50.8037 18.0118 4.18535 6.24526-6.24526 733.655
5 90.4781 16.731 4.86847 8.26702-8.26702 737.997\ast
6 173.503 18.2688 6.10689 12.5041-12.5041 734.271
7 364.579 25.7088 8.51289 23.2474-23.2474 718.548
8 795.808 47.4225 12.7543 53.8354-53.8354 682.923
9 1616.39 110.052 18.8531 153.237-153.237 606.713
10 2565.25 544.578 26.6876 928.402-928.402 320.315
type III nn a^n\hat{a}_{n} b^n\hat{b}_{n} μ^n\hat{\mu}_{n} θ^n\hat{\theta}_{n} PLL
0 7.17612 47.0281 3.1418 3.54481-3.54481 641.25
1 11.3467 35.0184 3.28923 3.8994-3.8994 674.554
2 19.4167 16.7381 3.37351 4.10569-4.10569 692.703
3 33.5171 5.48975 3.42025 4.2212-4.2212 701.522
4 58.7421 1.36161 3.44971 4.29443-4.29443 706.549
5 104.855 0.271079 3.46994 4.34489-4.34489 709.757
6 190.782 0.0450476 3.48468 4.38175-4.38175 711.971
7 353.758 0.00642237 3.49589 4.40984-4.40984 713.587
8 668.049 0.000801617 3.5047 4.43195-4.43195 714.816
9 1283.77 0.0000889697 3.51181 4.4498-4.4498 715.782
10 2508.23 8.88926×1068.88926\times 10^{-6} 3.51766 4.46452-4.46452 716.56\ast

PLL: profile log-likelihood

Table 10: Tempered stable estimation results on Bitcoin daily returns (\simsix significant digits)
c=0.5c=0.5 (NIG)
type I nn a^n\hat{a}_{n} b^n\hat{b}_{n} μ^n\hat{\mu}_{n} θ^n\hat{\theta}_{n} PLL
0 1.69903 40.2719 3.4866 4.3844-4.3844 741.57
0.5 2.18658 34.0606 3.60284 4.6784-4.6784 744.017
1 2.95041 31.8517 3.75372 5.06738-5.06738 746.459
1.5 4.037 30.6647 3.91344 5.48857-5.48857 748.356
2 5.56725 30.0464 4.08646 5.95593-5.95593 749.25
2.5 7.72111 29.8558 4.27762 6.48574-6.48574 749.903
3 10.7581 30.0452 4.49173 7.09616-7.09616 747.875
3.5 15.0512 30.61 4.73399 7.80857-7.80857 750.611\star
4 21.1362 31.5722 5.01006 8.64893-8.64893 749.197
4.5 29.7841 32.9755 5.32626 9.64906-9.64906 747.685
5 42.1034 34.8824 5.68939 10.8479-10.8479 745.767
type II nn a^n\hat{a}_{n} b^n\hat{b}_{n} μ^n\hat{\mu}_{n} θ^n\hat{\theta}_{n} PLL
0 1.69903 40.2719 3.4866 4.3844-4.3844 741.57
0.5 2.23523 35.6267 3.60867 4.69302-4.69302 744.468
1 2.95041 31.8517 3.75372 5.06738-5.06738 746.459
1.5 3.90955 28.8422 3.92845 5.5294-5.5294 748.107
2 5.20409 26.5277 4.14219 6.11131-6.11131 748.78\ast
2.5 6.96431 24.8739 4.40824 6.86179-6.86179 748.142
3 9.37804 23.891 4.74579 7.85667-7.85667 746.481
3.5 12.7184 23.648 5.18271 9.21738-9.21738 747.08
4 17.3826 24.2981 5.75891 11.1428-11.1428 743.584
4.5 23.9386 26.1153 6.52903 13.9616-13.9616 739.578
5 33.1604 29.5378 7.56059 18.2114-18.2114 731.809
type III nn a^n\hat{a}_{n} b^n\hat{b}_{n} μ^n\hat{\mu}_{n} θ^n\hat{\theta}_{n} PLL
0 1.69903 40.2719 3.4866 4.3844-4.3844 741.57
0.5 2.14646 37.2922 3.62438 4.73348-4.73348 743.909
1 2.95041 31.8517 3.75372 5.06738-5.06738 746.459
1.5 4.19843 23.7357 3.84507 5.30685-5.30685 748.233
2 6.13874 15.7484 3.9114 5.48266-5.48266 748.608
2.5 9.19582 9.4846 3.96141 5.61627-5.61627 749.061
3 14.0854 5.26192 4.00036 5.72095-5.72095 749.327
3.5 22.0249 2.71928 4.03151 5.80506-5.80506 749.486
4 35.1087 1.32035 4.05697 5.87407-5.87407 749.581
4.5 56.9807 0.606458 4.07815 5.93167-5.93167 749.634
5 94.0523 0.264957 4.09606 5.98048-5.98048 749.657\ast

NIG: normal inverse Gaussian — PLL: profile log-likelihood

Table 11: Tempered stable estimation results on Bitcoin daily returns (\simsix significant digits)
c=0.75c=0.75
type I nn a^n\hat{a}_{n} b^n\hat{b}_{n} μ^n\hat{\mu}_{n} θ^n\hat{\theta}_{n} PLL
0 0.304774 50.5501 5.21772 9.19796-9.19796 752.733\star
0.1 0.321949 48.3428 5.26152 9.33471-9.33471 752.507
0.2 0.341336 47.5163 5.34882 9.6086-9.6086 752.06
0.3 0.362624 47.2887 5.4522 9.9359-9.9359 751.546
0.4 0.385726 47.3793 5.56273 10.2896-10.2896 750.975
0.5 0.410649 47.664 5.67699 10.6595-10.6595 750.398
0.6 0.437445 48.0803 5.79359 11.0415-11.0415 749.811
0.7 0.466199 48.594 5.91194 11.434-11.434 749.224
0.8 0.497011 49.1848 6.03186 11.8366-11.8366 748.614
0.9 0.530001 49.8404 6.1533 12.2493-12.2493 748.005
1 0.565299 50.5528 6.27632 12.6725-12.6725 747.39
type II nn a^n\hat{a}_{n} b^n\hat{b}_{n} μ^n\hat{\mu}_{n} θ^n\hat{\theta}_{n} PLL
0 0.304774 50.5501 5.21772 9.19796-9.19796 752.733\star
0.1 0.324226 50.3274 5.30404 9.46608-9.46608 752.302
0.2 0.344917 50.1521 5.39414 9.74882-9.74882 751.851
0.3 0.366924 50.0249 5.48823 10.0472-10.0472 751.379
0.4 0.39033 49.9464 5.5865 10.3622-10.3622 750.885
0.5 0.41522 49.9176 5.68916 10.695-10.695 750.367
0.6 0.441685 49.9393 5.79643 11.0469-11.0469 749.825
0.7 0.469821 50.0124 5.90855 11.4192-11.4192 749.257
0.8 0.499731 50.1382 6.02574 11.8131-11.8131 748.663
0.9 0.531519 50.3179 6.14824 12.2304-12.2304 748.041
1 0.565299 50.5528 6.27632 12.6725-12.6725 747.39
type III nn a^n\hat{a}_{n} b^n\hat{b}_{n} μ^n\hat{\mu}_{n} θ^n\hat{\theta}_{n} PLL
0 0.304774 50.5501 5.21772 9.19796-9.19796 752.733\star
0.1 0.316506 49.8956 5.27563 9.37895-9.37895 752.433
0.2 0.331647 50.6222 5.38789 9.73239-9.73239 751.857
0.3 0.349699 51.6632 5.51481 10.1368-10.1368 751.21
0.4 0.370556 52.6039 5.64265 10.5493-10.5493 750.562
0.5 0.394296 53.2529 5.76595 10.9521-10.9521 749.939
0.6 0.421103 53.5212 5.88255 11.3377-11.3377 749.352
0.7 0.451243 53.3751 5.99177 11.7029-11.7029 748.805
0.8 0.485048 52.815 6.09357 12.0468-12.0468 748.297
0.9 0.522912 51.8624 6.18827 12.3697-12.3697 747.826
1 0.565299 50.5528 6.27632 12.6725-12.6725 747.39

PLL: profile log-likelihood

* Refer to caption Refer to caption Refer to caption Refer to caption

Figure 6: Tempered stable log-likelihood with regulation degrees for S&P500 daily returns
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 7: Tempered stable log-likelihood with regulation degrees for Bitcoin daily returns

* Refer to caption Refer to caption Refer to caption Refer to caption

Figure 8: Tempered stable estimated densities with regulation degrees for S&P500 daily returns
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 9: Tempered stable estimated densities with regulation degrees for Bitcoin daily returns

For the first exercise involving jump–diffusion models, from Table 2 and Table 3, the steady changes in the parameter values are clear for both data sets, which show robustness of the moment-based estimation. Specifically, while the estimate of the drift parameter, μ^n\hat{\mu}_{n}, has only marginally increased and that of the Brownian dispersion parameter, σ^n\hat{\sigma}_{n}, remains fairly stable, the infinite-divisibility parameter estimate λ^n\hat{\lambda}_{n} has increased geometrically with the degree nn regardless of the regulation type. This is in keeping with our original motivation to not reduce the measure of the number of trades. While decreasing, the rate of change of the rate parameter estimate b^n\hat{b}_{n}, on the other hand, does depend on the regulation type, and is the lowest under the type I, moderate under the type II, and the highest (actually in orders of reciprocal factorials) under the type III. This is because the rate parameter plays the role of adjusting the local movement sizes of stochastic clocks in consonance with the decreasing effect from regulation. Speaking of the profile log-likelihood, we see that any type of regulation has resulted in a notable improvement for some n1n\geq 1. While the first two types tend to achieve a maximal value relatively quickly for n10n\leq 10, the type III seems to be increasing the likelihood much more slowly. A simple explanation is that the third type has the mildest enlargement effect on distributional asymmetry and tail heaviness among the three. Besides, for both data sets the best models (with a “\star”) have all outperformed their counterparts with exponentially distributed jumps, also implying that jump–diffusion models with distorted uniformly distributed jump amplitudes can be used as an ideal replacement of the widely accepted exponential jump–diffusion models in most applications. The density plots in Figure 5 mainly show that despite closeness of the three estimated model densities, the peak of the one of the best fit model tends to lie between those of the models with uniform or exponential jumps, confirming its balancing characteristic mentioned in Subsection 4.1.

For the second exercise with Gaussian-mixed tempered stable models, it is seen from Table 4, Table 5, Table 6, Table 7, Table 8, Table 9, Table 10, and Table 11 that the parameter estimates also change steadily with the regulation degree, for all regulation types and both data sets. For any chosen value of the family parameter cc and any regulation type, apart from marginal increases in the location parameter estimate μ^n\hat{\mu}_{n} and the absolute Brownian drift estimate |θ^n||\hat{\theta}_{n}|, the conspicuous value increase of the estimate a^n\hat{a}_{n} of the infinite-divisibility parameter shows that clock regulation is indeed able to cater to a large number of trades while maintaining a significant level of asymmetry and tail heaviness. In the meantime, the interpretation of the value change of the rate parameter estimate b^n\hat{b}_{n} is the same as in the jump–diffusion case in the first exercise. Viewing from the profile log-likelihood, the effect of regulation deteriorates rapidly as cc increases, with the variance gamma case (c=0c=0) undergoing the greatest improvement. This phenomenon is within expectation, because in the limit as c1c\nearrow 1 the tempered stable subordinator will approach the usual drift clock of calendar time, which is by no means affected by regulation (or path averaging); equivalently, its Gaussian mixture will tend to a drifted Brownian motion, by the central limit theorem.141414When minute-level data are employed, the value of cc will probably be near 1 for liquid assets (see [Todorov and Tauchen, 2011, Table 5] [62]), in which case the underlying stochastic clock is close to the calendar time. This in turn shows that clock regulation is particularly useful for daily or lower-frequency data, making up their deficiency relative to the high-frequency counterparts which inevitably come at additional costs. Such a trend is more salient for the S&P500 returns than for the Bitcoin returns, depending on the relative extent of asymmetrical and tail risks. To be fair, an intriguing observation is that by increasing the degree nn of regulation the same level of log-likelihood (distributional information) has been achieved, whichever cc-value is being used. Since cc has an important dynamical meaning in terms of the local regularity of sample paths, and hence long-term trading intensity (recalling Section 1), clock regulation has therefore introduced a useful dimension for parameter estimation of a tempered stable model for any fixed value of the family parameter cc. The fundamental difference, as we shall restate, is that the regulation degree nn only changes the sizes of local movements and does not affect the Blumenthal–Getoor index (Corollary 1). The density plots in Figure 8 and Figure 9 further confirm the benefits of increasing the value of aa (a large number of trades), as the original model density without clock regulation seems overly “peaked” (almost degenerate, or near trading halts) when its moments are matched to the empirical, while proper clock regulation helps get rid of such degeneracy tendency.

To demonstrate the method outlined in Subsection 5.2, we also do a calibration exercise on Bitcoin options using the Gaussian-mixed tempered stable models. The data set is directly taken from [Xia, 2021, Sect. 4.2] [68] and consists of M=40M=40 Bitcoin call option prices quoted as of July 11th, 2020. The strike prices KK range from $3,000 to $32,000 and there are four maturities: T=19,47,166,257T=19,47,166,257 days. At the time of quoting, the Bitcoin price stood at S0=$9,232.98S_{0}=\$9,232.98. We have q=r=0q=r=0 for digital assets. To keep the demonstration concise, we also only consider four values of the family parameter c{0,0.25,0.5,0.75}c\in\{0,0.25,0.5,0.75\} and choose the domain of regulation degrees 𝔑={0,1,2,3,4,5}\mathfrak{N}=\{0,1,2,3,4,5\} (with |𝔑|=6|\mathfrak{N}|=6), which already give rise to 48 independent models. Each model is calibrated according to (5.2.2).

*

Table 12: Calibration results on Bitcoin options (\simsix significant digits)
type II
c=0c=0 (VG) nn a^n\hat{a}_{n} b^n\hat{b}_{n} θ^n\hat{\theta}_{n} MAPE (%) CPU time (sec)
0 52.8148 97.315 1.00541 24.7965 213.984
1 69.312 63.5709 0.835575 24.423 221.781
2 79.9247 34.5063 0.0841175-0.0841175 24.368 1208.5
3 99.9403 21.5024 0.271352-0.271352 24.2491 1011.41
4 99.8584 10.7058 0.806309-0.806309 23.7375 2647.44
5 99.9986 5.30486 1.21115-1.21115 23.3089\ast 3961.56
c=0.25c=0.25 nn a^n\hat{a}_{n} b^n\hat{b}_{n} θ^n\hat{\theta}_{n} MAPE (%) CPU time (sec)
0 5.29036 24.9255 1.4834-1.4834 22.5461 347.219
1 5.55635 9.93689 1.70271-1.70271 20.484 501.266
2 6.89235 5.1679 1.64386-1.64386 21.5499 3029.7
3 2.81083 0.476929 1.48728-1.48728 19.065\ast 13629.9
4 - - - - -
5 - - - - -
c=0.5c=0.5 (NIG) nn a^n\hat{a}_{n} b^n\hat{b}_{n} θ^n\hat{\theta}_{n} MAPE (%) CPU time (sec)
0 0.485557 1.37746 1.59762-1.59762 19.2856 135.625
1 0.664932 0.572124 1.46531-1.46531 19.2146 378.313
2 1.03635 0.343634 1.44776-1.44776 19.1939\ast 618.609
3 - - - - -
4 - - - - -
5 - - - - -
c=0.75c=0.75 nn a^n\hat{a}_{n} b^n\hat{b}_{n} θ^n\hat{\theta}_{n} MAPE (%) CPU time (sec)
0 0.288703 9.88632 1.18918-1.18918 23.3422 164.516
1 0.331241 0.608326 1.32828-1.32828 20.6779\ast 304.75
2 - - - - -
3 - - - - -
4 - - - - -
5 - - - - -

MAPE: mean absolute percentage error — VG: variance gamma — NIG: normal inverse Gaussian

Table 13: Calibration results on Bitcoin options (\simsix significant digits)
type III
c=0c=0 (VG) nn a^n\hat{a}_{n} b^n\hat{b}_{n} θ^n\hat{\theta}_{n} MAPE (%) CPU time (sec)
0 52.8148 97.315 1.00541 24.7965 213.984
1 69.312 63.5709 0.835575 24.423 221.781
2 25.7893 7.01917 1.82847-1.82847 21.9319\ast 1386.55
3 80.8854 5.83995 0.551056-0.551056 23.9882 1071.81
4 90.3674 1.28759 0.786714-0.786714 23.8168 968.281
5 97.9971 0.232557 0.926549-0.926549 23.6565 764.641
c=0.25c=0.25 nn a^n\hat{a}_{n} b^n\hat{b}_{n} θ^n\hat{\theta}_{n} MAPE (%) CPU time (sec)
0 5.29036 24.9255 1.4834-1.4834 22.5461 347.219
1 5.55635 9.93689 1.70271-1.70271 20.484 501.266
2 4.40503 1.49887 1.62404-1.62404 18.5857\ast 1442.34
3 7.26858 0.477655 1.8991-1.8991 19.6144 673.453
4 9.61409 0.0816453 1.64024-1.64024 18.9021 681.328
5 9.94952 0.0053706 2.12896-2.12896 24.8041 1469.98
c=0.5c=0.5 (NIG) nn a^n\hat{a}_{n} b^n\hat{b}_{n} θ^n\hat{\theta}_{n} MAPE (%) CPU time (sec)
0 0.485557 1.37746 1.59762-1.59762 19.2856 135.625
1 0.664932 0.572124 1.46531-1.46531 19.2146 378.313
2 1.30094 0.242459 1.45114-1.45114 19.1893\ast 492.594
3 3.02094 0.0752527 1.54985-1.54985 19.2294 1282.11
4 9.89913 0.0457464 1.59855-1.59855 19.9204 815.641
5 9.95359 0.0000857145 1.25003-1.25003 23.4282 1602.34
c=0.75c=0.75 nn a^n\hat{a}_{n} b^n\hat{b}_{n} θ^n\hat{\theta}_{n} MAPE (%) CPU time (sec)
0 0.288703 9.88632 1.18918-1.18918 23.3422 164.516
1 0.331241 0.608326 1.32828-1.32828 20.6779 304.75
2 0.978496 0.728833 1.24462-1.24462 22.117 762.734
3 1.82492 0.00378187 1.30428-1.30428 19.4049\ast 750.078
4 6.33734 0.000522405 1.31269-1.31269 19.43 708.563
5 27.7777 0.000615278 1.37646-1.37646 19.8326 903.156

MAPE: mean absolute percentage error — VG: variance gamma — NIG: normal inverse Gaussian

As discussed in Subsection 5.2, the optimal parameter values are subject to minor instability with joint calibration. For this reason, the main goal of this calibration exercise is to demonstrate that tempered stable models with clock regulation can be used with high efficiency and well expected to outperform the originals with unregulated clocks for financial derivatives valuation. The calibration results are summarized in Table 12 and Table 13, along with (locally) minimal pricing error and CPU time (with only a quad-core processor). Notably, under type-II regulated clocks, since the computational complexity of the formula (2) can increase rapidly with the degree of regulation, some models for c>0c>0 and n2n\geq 2 have not been successfully calibrated within a reasonable time limit (24 hours), and hence their results have been deliberately left out for comparability, while the remaining results are sufficient to serve our conclusions. Likewise, for each chosen value of cc, the model with the best fit is marked “\ast.”

From Table 12 and Table 13 we can see that adopting regulated clocks has significantly improved the calibration results, regardless of the value of cc. More specifically, for type-II regulation, better performance is generally produced by increasing nn, though the computation time grows drastically simultaneously. The latter feature is especially notable if c>0c>0, when one is forced to apply (2) outside the unit disk, whereas for the variance gamma model with c=0c=0, the more elementary formula (4.2.4) is used instead. In contrast, under type-III regulation, computation time does not scale with the degree of regulation, but there is oftentimes an optimal value n<5n<5 that gives a best fit.

Speaking of the optimized parameter values, we observe that, for each fixed value of cc, a^n\hat{a}_{n} usually increases with nn while b^\hat{b} decreases, which is similar to the results of statistical estimation. Besides, θ^n\hat{\theta}_{n} seems to be very stable in general, implying, yet again, that the Brownian drift parameter is largely uninfluenced by clock regulation. From Table 12 and Table 13 we identify the best fit model as the one with c=0.25c=0.25 and n=2n=2 under type-III regulation, whose implied prices are further visually compared with the market prices in Figure 10.

Refer to caption
Figure 10: Bitcoin option prices (market vs best fit model)

Furthermore, we compare in Figure 11 the risk-neutral densities (namely the 𝖰\mathsf{Q}-densities) of annualized Bitcoin log-returns (driven by Ξ1(n),Z\varXi^{(n),Z}_{1} for n[0,5]n\in\mathbb{N}\cap[0,5]) to further illustrate the static effect of clock regulation, though for succinctness the illustration is only done for the subcase “type III, c=0.25c=0.25” of which the best fit model is a special case. The formula (4.2) is applied because c<1/2c<1/2, along with the choice μn=logϕΞ1(n),Z(1)\mu_{n}=-\log\phi_{\varXi^{(n),Z}_{1}}(-1).

Refer to caption
Figure 11: Option-implied risk-neutral densities of Bitcoin log-returns (type III, c=0.25c=0.25)

7 Concluding remarks

In this paper we have studied the regulation of stochastic clocks built from Lévy subordinators for the purpose of tackling asymmetrical and tail risks typically observed in financial returns without making undesirable changes on parameters linked to important aspects of the trading activity. These parameters are those measuring infinite divisibility or α\alpha-stability and so making such an improvement has potential not only in enhancing capacity of the original processes but in balancing various modeling emphases involving the microstructure of data as well. From the viewpoint of static distributions, a direct effect is to arbitrarily enlarge skewness and (excess) kurtosis, while being able to fix the mean and variance and not decrease the jump intensity (or approach degeneracy).

The methodology comes in three concrete recipes. Despite a common motivation from repeated continuous-time averaging, the recipes vary in their impact on the original distributional asymmetry and tail heaviness as well as the computational complexity of resultant distribution formulae, which properties are summarized and compared in Table 14 for the convenience of the reader.

Table 14: Summary of properties of regulated stochastic clocks
Category type I type II type III
induction nature average of sample paths average of log-LT
quasi-average of
sample paths
regulating kernel incomplete gamma inverse gamma Riemann-Liouville
Lévy measure distortion severe moderate mild
Skew/EKurt enlargement faster than exponential exponential
slower than linear
or bounded
explicit formulae
(Poisson)
LT: n=1n=1 only
LM: any n>0n>0
LT: n++n\in\mathbb{N}_{++} only
LM: any n>0n>0
LT and LM: any n>0n>0
explicit formulae
(tempered stable)
LT and LM : n=1n=1 only LT and LM: n++n\in\mathbb{N}_{++} LT and LM: any n>0n>0

LT: Laplace transform — LM: Lévy measure

Generally speaking, stochastic clock regulation works well on any Lévy subordinator as long as the prerequisite (2.1) of generic-time log-LT analyticity is satisfied and can be utilized to obtain a good number of newfangled Lévy models for financial modeling and other uses. It is only futile for processes whose generic-time log-LT is linear or undefined in the neighborhood of zero.151515Note that the log-LT cannot be a finite-order polynomial of degree greater than two; see [Lukacs, 1970, Thm. 7.3.5] [45]. Clearly, in the former case the subordinator is degenerate and effectively identical to the clock of calendar time, and in the latter it does not have a finite third-order cumulant. The resultant mixed models then have skewness and excess kurtosis either equal to zero or undefined (including infinity) – obviously, both zero and infinity cannot be further “enlarged” by multiplication.

When applied to a Poisson process, the regulated clocks have led to compound Poisson processes with bounded jumps having a truncated exponential-like distribution and hence corresponding jump–diffusion models inheriting the advantages of both the uniform distribution ([Yan and Hanson, 2006] [69]) and the (usual) exponential distribution ([Kou, 2002] [40]). As our empirical study on S&P500 and Bitcoin returns has shown, when moment-based estimation is employed to match the empirical skewness and kurtosis, the profile likelihood for the new models with regulation degree n>0n>0 is indeed improved relative to the uniform and the pure-exponential, which signifies that the distorted uniform distributions are suitable for modeling downside risks in a jump–diffusion framework.

Secondly, operating the regulated clocks on the Gaussian-mixed tempered stable models generates a series of modified processes with unchanged path regularity but enhanced distributional asymmetry and tail heaviness. The empirical results imply that the effect of regulation is most conspicuous when the family parameter cc is small, close to the case of variance gamma processes, while it gradually diminishes as cc approaches the upper limit 1, when the mixed model tends to Gaussianity. More importantly, by varying the degree nn of regulation, one can expect to easily take account of empirical skewness and kurtosis of returns without losing essential information on the actual distribution, regardless of which value of cc is in use. Put differently, in the presence of the hyperparameter nn, the value of cc can be confidently fixed for a certain data frequency or taken from other sources of estimation, since nn, benignly, has no effect on the local regularity of the stochastic process.

The methodology that the present paper has put forward is a jump-off point and so can be advanced in various directions. For instance, as we have mentioned since Section 1, it can be extended to subordinators of the Sato type (with non-stationary increments) with reasonable effort as well as the establishment of mean-reverting time changes of Ornstein–Uhlenbeck type ([Carr and Wu, 2004] [19]) for stochastic volatility. Also, an ongoing research is concerned with developments into a multidimensional setting, especially for the Gaussian mixed models, which will be conducive to portfolio analysis and derived optimization problems. We note that multivariate subordination of Brownian motion via tempered stable subordinators has received much attention in the literature; see, e.g., [Luciano and Semeraro, 2010a] [43], [Luciano and Semeraro, 2010b] [44], [Pérez-Abreu and Stelzer, 2014] [53], and [Mandoza-Arriaga and Linetsky, 2016] [48]. Of course, the two special cases discussed in Section 4 are also far from exhaustive, and it is possible to consider operating the regulating kernels on other choices of Lévy subordinators, e.g., compound Poisson-exponential processes or generalized inverse Gaussian processes ([Barndorff-Nielsen and Halgreen, 1977] [8]), and derive the moment-based estimation procedures in the same spirit as in Subsection 5.1. Last but not least, the empirical aspect may be strengthened by studying other kinds of financial time series, such as individual stocks, exchange rates, volatility indices, etc., in order to explore how the degree of clock regulation is fundamentally reflected in trading volumes or relative illiquidity and its variation with different levels of asymmetrical and tail risks.

References

  • [1] Abramowitz, M., Stegun, I.A. Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables, 10th Printing, U.S. National Bureau of Standards, Washington, D.C., 1972.
  • [2] Aguilar, J-P., Kirkby, J.L., Korbel, J. (2020). Pricing, risk and volatility in subordinated market models. Risks, 8: No. 124.
  • [3] Aït-Sahalia, Y., Jacod, J. (2009). Testing for jumps in a discretely observed process. Annals of Statistics, 37: 184–222.
  • [4] Alexander, G.J., Peterson, M.A. (2007). An analysis of trade-size clustering and its relation to stealth trading. Journal of Financial Economics, 84: 435–471.
  • [5] Ané, T., Geman, H. (2000). Order flow, transaction clock, and normality of asset returns. Journal of Finance, 55: 2259–2284.
  • [6] Bandi, F.M., Nguyen, T.H. (2003). On the functional estimation of jump–diffusion models. Journal of Econometrics, 116: 293–328.
  • [7] Barndorff-Nielsen, O.E. (1997). Normal inverse Gaussian distributions and stochastic volatility models. Scandinavian Journal of Statistics, 24: 1–13.
  • [8] Barndorff-Nielsen, O.E., Halgreen, C. (1977). Infinitely divisibility of the hyperbolic and generalized inverse Gaussian distributions. Zeitschrift für Wahrscheinlichkeitstheorie und verwandte Gebiete, 38: 309–311.
  • [9] Barndorff-Nielsen, O.E., Pollard, D.G., Shephard, N. (2011). Integer-valued Lévy processes and low latency financial econometrics. Quantitative Finance, 12: 587–605.
  • [10] Bateman, H. Tables of Integral Transforms, 1st Vol., McGraw-Hill, New York–Toronto–London, 1954.
  • [11] Bateman, H., Erdélyi, A. Higher Transcendental Functions, 2nd Vol., McGraw-Hill, New York, 1954.
  • [12] Baustian, F., Mrázek, M., Pospíšil, J., Sobotka, T. (2017). Unifying pricing formula for several stochastic volatility models with jumps. Applied Stochastic Models in Business and Industry, 33: 422–442.
  • [13] Beckers, S. (1981). A note on estimating the parameters of the diffusion–jump model of stock returns. Journal of Financial and Quantitative Analysis, 16: 127–140.
  • [14] Blau, B.M. (2017). Price dynamics and speculative trading in bitcoin. Research in International Business and Finance, 41: 493–499.
  • [15] Blumenthal, R.M., Getoor, R.K. (1961). Sample functions of stochastic processes with stationary independent increments. Journal of Mathematics and Mechanics, 10: 493–516.
  • [16] Bollerslev, T. (1987). A conditionally heteroskedastic time series model for speculative prices and rates of return. Review of Economics and Statistics, 69: 542–547.
  • [17] Boyarchenko, S.I., Levendorskii, S.Z. (2000). Option pricing for truncated Lévy processes. International Journal of Theoretical and Applied Finance, 3: 549–552.
  • [18] Carr, P., Gemen, H., Madan, D.B., Yor, M. (2002). The fine structure of asset returns: An empirical investigation. The Journal of Business, 75: 305–332.
  • [19] Carr, P., Wu, L. (2004). Time-changed Lévy processes and option pricing. Journal of Financial Economics, 71: 113–141.
  • [20] Chaim, P., Laurini, M.P. (2019). Nonlinear dependence in cryptocurrency markets. North American Journal of Economics and Finance, 48: 32–47.
  • [21] Chang, E.C., Pinegar, J.M., Schachter, B. (1997). Interday variations in volume, variance and participation of large speculators. Journal of Banking and Finance, 21: 797–810.
  • [22] Chen, T. (2019). The price impact of trade-size clustering: Evidence from an intraday analysis. Journal of Business Research, 101: 300–314.
  • [23] Clark, P.K. (1973). A subordinated stochastic process model with finite variance for speculative prices. Econometrica, 41: 135–155.
  • [24] DiDonato, A.R., Morris, A.H. (1986). Computation of the incomplete gamma function ratios and their inverse. ACM Transactions on Mathematical Software, 12: 377–393.
  • [25] Downton, F. (1972). The area under the infectives trajectory of the general stochastic epidemic. Journal of Applied Probability, 9: 414–417.
  • [26] Eberlein, E., Madan, D.B. (2009). Sato processes and the valuation of structured products. Quantitative Finance, 9: 27–42.
  • [27] Fallahgoul, H., Loeper, G. (2021). Modelling tail risk with tempered stable distributions: An overview. Annals of Operations Research, 299: 1253–1280.
  • [28] Geman, H. Stochastic clock and financial markets. In: Aspects of Mathematical Finance, Springer, Berlin, Heidelberg, 37–52, 2008.
  • [29] Geman, H., Madan, D.B., Yor, M. (2001). Time changes for Lévy processes. Mathematical Finance, 11: 79–96.
  • [30] Gkillas, K., Katsiampa, P. (2018). An application of extreme value theory to cryptocurrencies. Economics Letters, 164: 109–111.
  • [31] Grabchak, M. (2021). Discrete tempered stable distributions. Methodology and Computing in Applied Probability, Online First, 1–14.
  • [32] Gradshteyn, I.S., Ryzhik, I.M. Table of Integrals, Series, and Products, 7th Ed. Academic Press, Elsevier, Burlington, 2007.
  • [33] Gupta, N., Kumar, A., Leonenko, N.N. (2020). Skellam type processes of order kk and beyond. Entropy, 22: 1193.
  • [34] Jing, B-Y., Kong, X-B., Liu, Z., Mykland, P. (2012). On the jump activity index for semimartingales. Journal of Econometrics, 166: 213–223.
  • [35] Jondeau, E., Rockinger, M. (2003). Conditional volatility, skewness, and kurtosis: Existence, persistence, and comovements. Journal of Economic Dynamics and Control, 27: 1699–1737.
  • [36] Kalmykov, M.Y., Kniehl, B.A. (2009). Towards all-order Laurent expansion of generalised hypergeometric functions about rational values of parameters. Nuclear Physics B, 809 [PM]: 365–405.
  • [37] Karpoff, J.M. (1987). The relation between price changes and trading volume: A survey. Journal of Financial and Quantitative Analysis, 22: 109–126.
  • [38] Kerss, A., Leonenko, N.N., Sikorskii, A. (2014). Fractional Skellam processes with applications to finance. Fractional Calculus & Applied Analysis, 17: 532-551.
  • [39] Koponen, I. (1995). Analytic approach to the problem of convergence of truncated Lévy flights towards the Gaussian stochastic process. Physical Review E, 52: 1197–1199.
  • [40] Kou, S. (2002). A jump-diffusion model for option pricing. Management Science, 48: 1086–1101.
  • [41] Küchler, U., Tappe, S. (2013). Tempered stable distributions and processes. Stochastic Processes and their Applications, 123: 4256–4293.
  • [42] Liu, Y., Tsyvinski, A. (2021). Risks and returns of cryptocurrency. The Review of Financial Studies, 34: 2689–2727.
  • [43] Luciano, E., Semeraro, P. (2010a). A generalized normal mean–variance mixture for return processes in finance. International Journal of Theoretical and Applied Finance, 13: 415–440.
  • [44] Luciano, E., Semeraro, P. (2010b). Multivariate time changes for Lévy asset models: Characterization and calibration. Journal of Computational and Applied Mathematics, 233: 1937–1953.
  • [45] Lukacs, E. Characteristic Functions, 2nd Ed., Griffin, London, 1970.
  • [46] Madan, D.B., Carr, P., Chang, E.C. (1998). The variance gamma process and option pricing. European Finance Review, 2: 79–105.
  • [47] Madan, D.B., Reyners, S., Schoutens, W. (2019). Advanced model calibration on bitcoin options. Digital Finance, 1: 117–137.
  • [48] Mandoza-Arriaga, R., Linetsky, V. (2016). Multivariate subordination of Markov processes with financial applications. Mathematical Finance, 26: 699–747.
  • [49] Mittnik, S., Paolella, M.S., Rachev, S.T. (2000). Diagnosing and treating the fat tails in financial returns data. Journal of Empirical Finance, 7: 389–416.
  • [50] Monroe, I. (1978). Processes that can be embedded in Brownian motion. Annals of Probability, 6: 42–56.
  • [51] Ornthanalai, C. (2014). Lévy jump risk: Evidence from options and returns. Journal of Financial Economics, 112: 69–90.
  • [52] Orsingher, E., Polito, F. (2013). On the integral of fractional Poisson processes. Statistics and Probability Letters, 83: 1006–1017.
  • [53] Pérez-Abreu, V., Stelzer, R. (2014). Infinitely divisible multivariate and matrix gamma distributions. Journal of Multivariate Analysis, 130: 155–175.
  • [54] Pollett, P.K. (2003). Integrals for continuous-time Markov chains. Mathematical Biosciences, 182: 213–225.
  • [55] Richardson, M., Smith, T. (1994). A direct test of mixture distribution hypothesis: Measuring the daily flow of information. Journal of Financial and Quantitative Analysis, 29: 101–116.
  • [56] Ronsiński, J. (2007). Tempering stable processes. Stochastic Processes and their Applications, 117: 677–707.
  • [57] Samorodnitsky, G., Taqqu, M.S. Stable Non-Gaussian Random Processes: Stochastic Models with Infinite Variance. Chapman & Hall / CRC, New York, 1994.
  • [58] Sato, K-I. (2006). Additive processes and stochastic integrals. Illinois Journal of Mathematics, 50: 825–851.
  • [59] Schoutens, W. Lévy Processes in Finance: Pricing Financial Derivatives. The Atrium, Southern Gate, Chichestor: John Wiley & Sons Ltd, 2003.
  • [60] Slater, L.J. Generalized Hypergeometric Functions, Cambridge University Press, Cambridge, UK, 1966.
  • [61] Todorov, V. (2021). Nonparametric jump variation measures from options. Journal of Econometrics, Online First.
  • [62] Todorov, V., Tauchen, G. (2011). Volatility jumps. Journal of Business and Economic Statistics, 29: 356–371.
  • [63] Troster, V., Tiwari, A.K., Shahbaz, M., Macedo, D.N. (2019). Bitcoin returns and risk: A general GARCH and GAS analysis. Finance Research Letters, 30: 187–193.
  • [64] Tseng, S-P., Peng, C-Y. (2004). Optimal burn-in policy by using an integrated Wiener process. IIE Transactions, 36: 1161–1170.
  • [65] Van Noortwijk, J.M. (2009). A survey of the application of gamma processes in maintenance. Reliability Engineering & System Safety, 94: 2–21.
  • [66] Wang, X., Xu, D. (2010). An inverse Gaussian process model for degradation data. Technometrics, 52: 188–197.
  • [67] Xia, W. (2019). The average of a negative-binomial Lévy process and a class of Lerch distributions. Communications in Statistics - Theory and Methods, 49: 1008–1024.
  • [68] Xia, W. (2021). Average-tempered stable subordinators with applications. Applied Stochastic Models in Business and Industry, 37: 1097–1122.
  • [69] Yan, G., Hanson, F.B. Option pricing for a stochastic-volatility jump–diffusion model with log-uniform jump-amplitude. In: Proceedings of American Control Conference, Minneapolis, MI, USA, June, 14–16, 2006.
  • [70] Ye, Z-S., Xie, M. (2015). Stochastic modelling and analysis of degradation for highly reliable products. Applied Stochastic Models in Business and Industry, 31: 16–32.

Appendix Appendix A Mathematical proofs

Theorem 1

Proof..

Let t>0t>0 be fixed. By applying integration-by-parts to (2.1.1) nn times we have the fractional integral representation

X¯t(n)=1tΓ(n)0tXslogn1tsds=0t(1Γ(n,log(t/s))Γ(n))dXs,\bar{X}^{(n)}_{t}=\frac{1}{t\mathrm{\Gamma}(n)}\int^{t}_{0}X_{s}\log^{n-1}\frac{t}{s}{\rm d}s=\int^{t}_{0}\bigg{(}1-\frac{\mathrm{\Gamma}(n,\log(t/s))}{\mathrm{\Gamma}(n)}\bigg{)}{\rm d}X_{s}, (A.1)

where the second equality follows from the substitution log(t/s)s\log(t/s)\mapsto s.161616The first recipe generalizes the fractional integration of nonnegative processes to logarithmic kernels, as for every n1n\geq 1 and t>0t>0 the integrand (1/(tΓ(n))Xslogn1(t/s))s(0,t](1/(t\mathrm{\Gamma}(n))X_{s}\log^{n-1}(t/s))_{s\in(0,t]} in (A.1) constitutes a nonnegative stochastic process. The representation (A.1) together with the fact that XX is a Lévy process allows to write

EeuX¯t(n)=0tE(exp(uX1(1Γ(n,log(t/s))Γ(n))))ds=exp(0tlogϕX1((1Γ(n,log(t/s))Γ(n))u)ds),\textsf{E}e^{-u\bar{X}^{(n)}_{t}}=\prod^{t}_{0}\textsf{E}\bigg{(}\exp\bigg{(}-uX_{1}\bigg{(}1-\frac{\mathrm{\Gamma}(n,\log(t/s))}{\mathrm{\Gamma}(n)}\bigg{)}\bigg{)}\bigg{)}^{{\rm d}s}=\exp\bigg{(}\int^{t}_{0}\log\phi_{X_{1}}\bigg{(}\bigg{(}1-\frac{\mathrm{\Gamma}(n,\log(t/s))}{\mathrm{\Gamma}(n)}\bigg{)}u\bigg{)}{\rm d}s\bigg{)},

where \prod^{\cdot}_{\cdot} is the geometric integral sign and which leads to the first formula in (1) with the substitution s/tss/t\mapsto s.

For the other two formulae it suffices to identify the corresponding fractional integral representations. For (2.1.3), observe that for each n1n\geq 1, X~(n)\tilde{X}^{(n)} is the running average of the given Lévy process Y(n1)Y^{(n-1)}, so that the LT of X~t(n)\tilde{X}^{(n)}_{t} at generic time satisfies the recurrence relation

ϕX~t(n)(u)=exp(tu0ulogϕX~1(n1)(v)dv),Reu0,\phi_{\tilde{X}^{(n)}_{t}}(u)=\exp\bigg{(}\frac{t}{u}\int^{u}_{0}\log\phi_{\tilde{X}^{(n-1)}_{1}}(v){\rm d}v\bigg{)},\quad\mathrm{Re}u\geq 0, (A.2)

due to the independent increments property of Y(n1)Y^{(n-1)}. The recurrence can be resolved easily via integration-by-parts, given the initial condition X~(0)X\tilde{X}^{(0)}\equiv X, in the same vein as (A.1), yielding the second formula in (1).

Also, since the integrals in (2.1.4) are simply iterated, an application of Cauchy’s repeated integration formula immediately gives

X˘t(n)=1tnΓ(n+1)0t(ts)ndXs,\breve{X}^{(n)}_{t}=\frac{1}{t^{n}\mathrm{\Gamma}(n+1)}\int^{t}_{0}(t-s)^{n}{\rm d}X_{s},

which result shows that each X˘(n)\breve{X}^{(n)} can be treated as a (times-scaled) usual fractional Lévy process under the Riemann–Liouville kernel. ∎

Theorem 2

Proof..

Comparing the first formula in (1) with the first formula in (2.2), we obtain the following Fredholm integral equation:

0(euz1)ν¯(n)(dz)=010(e(1Γ(n,logs)/Γ(n))uz1)ν(dz)ds,\int^{\infty}_{0}(e^{-uz}-1)\bar{\nu}^{(n)}({\rm d}z)=\int^{1}_{0}\int^{\infty}_{0}\big{(}e^{-(1-\mathrm{\Gamma}(n,-\log s)/\mathrm{\Gamma}(n))uz}-1\big{)}\nu({\rm d}z){\rm d}s,

but as 1Γ(n,logs)/Γ(n)01-\mathrm{\Gamma}(n,-\log s)/\mathrm{\Gamma}(n)\geq 0 for all n1n\geq 1 and s[0,1]s\in[0,1], the substitution (1Γ(n,logs)/Γ(n))zz(1-\mathrm{\Gamma}(n,-\log s)/\mathrm{\Gamma}(n))z\mapsto z together with the Fubini–Tonelli theorem yields

ν¯(n)(dz)=01ν(dΓ(n)zΓ(n)Γ(n,logs))ds,z0.\bar{\nu}^{(n)}({\rm d}z)=\int^{1}_{0}\nu\bigg{(}{\rm d}\frac{\mathrm{\Gamma}(n)z}{\mathrm{\Gamma}(n)-\mathrm{\Gamma}(n,-\log s)}\bigg{)}{\rm d}s,\quad z\geq 0. (A.3)

Since [0,1]sΓ(n)/(Γ(n)Γ(n,logs))[1,)[0,1]\ni s\mapsto\mathrm{\Gamma}(n)/(\mathrm{\Gamma}(n)-\mathrm{\Gamma}(n,-\log s))\in[1,\infty) forms a strictly increasing function for each n1n\geq 1, we can apply the further substitution Γ(n)/(Γ(n)Γ(n,logs))y\mathrm{\Gamma}(n)/(\mathrm{\Gamma}(n)-\mathrm{\Gamma}(n,-\log s))\mapsto y to obtain the first formula in (2), after some rearrangement.

Then, for each fixed n1n\geq 1, comparing the second formula in (1) with the second formula in (2.2) and interchanging integrals gives the following integral equation:

0(euz1)ν~(n)(dz)=001(euvz1)dvν~(n1)(dz),\int^{\infty}_{0}(e^{-uz}-1)\tilde{\nu}^{(n)}({\rm d}z)=\int^{\infty}_{0}\int^{1}_{0}(e^{-uvz}-1){\rm d}v\tilde{\nu}^{(n-1)}({\rm d}z),

which is equivalent, on rearranging the right-hand side, to

0(euz1)ν~(n)(dz)=0(euz1)1ν~(n1)(d(yz))y2dy.\int^{\infty}_{0}(e^{-uz}-1)\tilde{\nu}^{(n)}({\rm d}z)=\int^{\infty}_{0}(e^{-uz}-1)\int^{\infty}_{1}\frac{\tilde{\nu}^{(n-1)}({\rm d}(yz))}{y^{2}}{\rm d}y.

It leaves us with no choice but

ν~(n)(dz)=1ν~(n1)(d(yz))y2dy.\tilde{\nu}^{(n)}({\rm d}z)=\int^{\infty}_{1}\frac{\tilde{\nu}^{(n-1)}({\rm d}(yz))}{y^{2}}{\rm d}y.

This is yet another recurrence equation similar to the ones solved before, subject to the initial condition ν~(0)ν\tilde{\nu}^{(0)}\equiv\nu. Upon integration-by-parts we obtain the second formula in (2).

It is straightforward to obtain the third formula in (2) by mimicking the steps taken for the first.

With absolute continuity with respect to Lebesgue measure, the corresponding formulae for Lévy densities in (2) follow from direct calculations. ∎

Corollary 1

Proof..

According to the definition of the Blumenthal–Getoor index, it suffices to observe that the relations in (2) (or (2)) entail integration over domains bounded away from zero. Then the sigma-finiteness of the Lévy measures along with the Fubini–Tonelli theorem implies that the finiteness of the integrals 01|z|pν¯(n)(dz)\int^{1}_{0}|z|^{p}\bar{\nu}^{(n)}({\rm d}z), 01|z|pν~(n)(dz)\int^{1}_{0}|z|^{p}\tilde{\nu}^{(n)}({\rm d}z), and 01|z|pν˘(n)(dz)\int^{1}_{0}|z|^{p}\breve{\nu}^{(n)}({\rm d}z), for any n>0n>0, are all equivalent to that of 01|z|pν(dz)\int^{1}_{0}|z|^{p}\nu({\rm d}z), for any p>0p>0. ∎

Corollary 2

Proof..

Fix n>0n>0. The construction of the coefficients Cm,nC_{m,n} and 1/((mn+1)Γm(n+1))1/((mn+1)\mathrm{\Gamma}^{m}(n+1)), m++m\in\mathbb{N}_{++}, for the type I and type III, respectively, is a direct result of Theorem 1, or the fractional integral representations (A.1) and (A.2). For the type II, since induction takes place for each nn, the coefficients 1/(m+1)n1/(m+1)^{n}, m++m\in\mathbb{N}_{++}, are obtained by noting that Cm,1=01(1s)mds=1/(m+1)C_{m,1}=\int^{1}_{0}(1-s)^{m}{\rm d}s=1/(m+1).

The rationality of Cm,nC_{m,n}’s is seen from the fact that, for every n++n\in\mathbb{N}_{++}, Γ(n)(n1)!\mathrm{\Gamma}(n)\equiv(n-1)! and the finite series representation Γ(n,logs)=(n1)!sk=0n1(logs)k/k!\mathrm{\Gamma}(n,-\log s)=(n-1)!s\sum^{n-1}_{k=0}(-\log s)^{k}/k!, and 01sj(logs)kds=k!/(j+1)k+1\int^{1}_{0}s^{j}(-\log s)^{k}{\rm d}s=k!/(j+1)^{k+1} for any j,kj,k\in\mathbb{N}. ∎

Corollary 3

Proof..

The formulae follow immediately by replacing XX with its regulated counterparts in (3.1.2). ∎

Corollary 4

Proof..

According to Corollary 1, it is enough to consider the constant mixture ξX\xi^{X} based on the Lévy subordinator XX. Suppose κ1,κ2>0\kappa_{1},\kappa_{2}>0. Following (3.1.2) and the Lévy–Khintchine representation, the Lévy measure of ξX\xi^{X} is given by

νξX(dz)=νX(dzκ1)+νX(dzκ2),z{0}.\nu_{\xi^{X}}({\rm d}z)=\nu_{X}\bigg{(}{\rm d}\frac{z}{\kappa_{1}}\bigg{)}+\nu_{-X}\bigg{(}{\rm d}\frac{z}{\kappa_{2}}\bigg{)},\quad z\in\mathbb{R}\setminus\{0\}.

Thus, it follows from the inequalities

01zpνX(dzκ1)11|z|pνξX(dz)2max{01zpνX(dzκ1),01zpνX(dzκ2)},p>0\int^{1}_{0}z^{p}\nu_{X}\bigg{(}{\rm d}\frac{z}{\kappa_{1}}\bigg{)}\leq\int^{1}_{-1}|z|^{p}\nu_{\xi^{X}}({\rm d}z)\leq 2\max\bigg{\{}\int^{1}_{0}z^{p}\nu_{X}\bigg{(}{\rm d}\frac{z}{\kappa_{1}}\bigg{)},\int^{1}_{0}z^{p}\nu_{X}\bigg{(}{\rm d}\frac{z}{\kappa_{2}}\bigg{)}\bigg{\}},\quad p>0

that 𝔅(ξX)=𝔅(X)\mathfrak{B}(\xi^{X})=\mathfrak{B}(X), as required. If either κ1\kappa_{1} or κ2\kappa_{2} is zero the result is immediate. ∎

Corollary 5

Proof..

The formulae follow immediately by replacing XX with its regulated counterparts in (3.2.2). ∎

Corollary 6

Proof..

Again, with Corollary 1 we only consider the Gaussian mixture ΞX\varXi^{X} based on XX. By construction, the Lévy measure of ΞX\varXi^{X} satisfies

νΞX(dy)=0νX(ds)γs(dy),y{0},\nu_{\varXi^{X}}({\rm d}y)=\int^{\infty}_{0}\nu_{X}({\rm d}s)\gamma_{s}({\rm d}y),\quad y\in\mathbb{R}\setminus\{0\},

where γt\gamma_{t}, for t>0t>0, represents the Gaussian measure with mean θt\theta t and variance tt, that is, the probability measure of the drifted Brownian motion θt+Wt\theta t+W_{t}. For every p>0p>0, by the Fubini–Tonelli theorem we have

11|y|p0νX(ds)γs(dy)=0νX(ds)E[|θs+sζ|p𝟙As],\int_{-1}^{1}|y|^{p}\int^{\infty}_{0}\nu_{X}({\rm d}s)\gamma_{s}({\rm d}y)=\int^{\infty}_{0}\nu_{X}({\rm d}s)\textsf{E}\big{[}|\theta s+\sqrt{s}\zeta|^{p}\mathbb{1}_{A_{s}}\big{]},

ζ\zeta being a standard normal random variable and As:={|θs+sζ|1}A_{s}:=\{|\theta s+\sqrt{s}\zeta|\leq 1\}. From the fact that

1νX(ds)E[|θs+sζ|p𝟙As]νX([1,))<,\int^{\infty}_{1}\nu_{X}({\rm d}s)\textsf{E}[|\theta s+\sqrt{s}\zeta|^{p}\mathbb{1}_{A_{s}}]\leq\nu_{X}([1,\infty))<\infty,

the finiteness of 0νX(ds)E[|θs+sζ|p𝟙As]\int^{\infty}_{0}\nu_{X}({\rm d}s)\textsf{E}\big{[}|\theta s+\sqrt{s}\zeta|^{p}\mathbb{1}_{A_{s}}\big{]} is the same as that of 01νX(ds)E[|θs+sζ|p𝟙As]\int^{1}_{0}\nu_{X}({\rm d}s)\textsf{E}\big{[}|\theta s+\sqrt{s}\zeta|^{p}\mathbb{1}_{A_{s}}\big{]}. Towards that end, we first apply the CpC^{p}-inequality to obtain for every p>2𝔅(X)p>2\mathfrak{B}(X)

01E[|θs+sζ|p𝟙As]νX(ds)\displaystyle\int^{1}_{0}\textsf{E}\big{[}|\theta s+\sqrt{s}\zeta|^{p}\mathbb{1}_{A_{s}}\big{]}\nu_{X}({\rm d}s) 2(p1)+(θp01spE[𝟙As]νX(ds)+01sp/2E[|ζ|p𝟙As]νX(ds))\displaystyle\leq 2^{(p-1)^{+}}\bigg{(}\theta^{p}\int^{1}_{0}s^{p}\textsf{E}[\mathbb{1}_{A_{s}}]\nu_{X}({\rm d}s)+\int^{1}_{0}s^{p/2}\textsf{E}\big{[}|\zeta|^{p}\mathbb{1}_{A_{s}}\big{]}\nu_{X}({\rm d}s)\bigg{)}
2(p1)+(θp01spνX(ds)+E[|ζ|p]01sp/2νX(ds))\displaystyle\leq 2^{(p-1)^{+}}\bigg{(}\theta^{p}\int^{1}_{0}s^{p}\nu_{X}({\rm d}s)+\textsf{E}[|\zeta|^{p}]\int^{1}_{0}s^{p/2}\nu_{X}({\rm d}s)\bigg{)}
<,\displaystyle<\infty,

so that 𝔅(ΞX)2𝔅(X)\mathfrak{B}(\varXi^{X})\leq 2\mathfrak{B}(X). Second, for every p<2𝔅(X)p<2\mathfrak{B}(X), observe that

01E[|θs+sζ|p𝟙As]νX(ds)\displaystyle\int^{1}_{0}\textsf{E}\big{[}|\theta s+\sqrt{s}\zeta|^{p}\mathbb{1}_{A_{s}}\big{]}\nu_{X}({\rm d}s) 01sp/22πθsθs+1/s(xθs)pez2/2dxνX(ds)\displaystyle\geq\int^{1}_{0}\frac{s^{p/2}}{\sqrt{2\pi}}\int^{\theta\sqrt{s}+1/\sqrt{s}}_{\theta\sqrt{s}}(x-\theta\sqrt{s})^{p}e^{-z^{2}/2}{\rm d}x\;\nu_{X}({\rm d}s)
=01sp/22π01/sxpexp((x+θs)22)dxνX(ds)\displaystyle=\int^{1}_{0}\frac{s^{p/2}}{\sqrt{2\pi}}\int_{0}^{1/\sqrt{s}}x^{p}\exp\left(-\frac{(x+\theta\sqrt{s})^{2}}{2}\right){\rm d}x\;\nu_{X}({\rm d}s)
01sp/22π01xpexp((x+θ)22)dxνX(ds)\displaystyle\geq\int^{1}_{0}\frac{s^{p/2}}{\sqrt{2\pi}}\int^{1}_{0}x^{p}\exp\left(-\frac{(x+\theta)^{2}}{2}\right){\rm d}x\;\nu_{X}({\rm d}s)
=E[(ζθ)p𝟙[θ,θ+1](ζ)]01sp/2νX(ds)\displaystyle=\textsf{E}\big{[}(\zeta-\theta)^{p}\mathbb{1}_{[\theta,\theta+1]}(\zeta)\big{]}\int^{1}_{0}s^{p/2}\nu_{X}({\rm d}s)
=,\displaystyle=\infty,

where the second equality uses the substitution xθsxx-\theta\sqrt{s}\mapsto x and which implies that 𝔅(ΞX)2𝔅(X)\mathfrak{B}(\varXi^{X})\geq 2\mathfrak{B}(X). It is therefore concluded that 𝔅(Z)=2𝔅(X)\mathfrak{B}(Z)=2\mathfrak{B}(X), as required. ∎

Proposition 1

Proof..

For the type I, plugging the stated LT of X1X_{1} into the first formula in (1) and applying substitution as in (A.3) the LT of X¯1(n)\bar{X}^{(n)}_{1} reads for n>0n>0

ϕX¯1(n)(u)\displaystyle\phi_{\bar{X}^{(n)}_{1}}(u) =exp(λ(01e(1Γ(n,logs)/Γ(n))uds1))\displaystyle=\exp\bigg{(}\lambda\bigg{(}\int^{1}_{0}e^{-(1-\mathrm{\Gamma}(n,-\log s)/\mathrm{\Gamma}(n))u}{\rm d}s-1\bigg{)}\bigg{)}
=exp(λ(Γ(n)01euzQn1(n,1z)dz1)),\displaystyle=\exp\bigg{(}\lambda\bigg{(}\mathrm{\Gamma}(n)\int^{1}_{0}\frac{e^{-uz}}{\mathrm{Q}^{n-1}(n,1-z)}{\rm d}z-1\bigg{)}\bigg{)}, (A.4)

which is clearly an entire function.

The general formula for the type II is immediate from the second formula in (1). If n++n\in\mathbb{N}_{++}, then with 01(logv)n1/Γ(n)dv=1\int^{1}_{0}(-\log v)^{n-1}/\mathrm{\Gamma}(n){\rm d}v=1 we have that

1Γ(n)01euv(logv)n1dv\displaystyle\frac{1}{\mathrm{\Gamma}(n)}\int^{1}_{0}e^{-uv}(-\log v)^{n-1}{\rm d}v =k=001(logv)n1(uv)kΓ(n)k!dv\displaystyle=\sum^{\infty}_{k=0}\int^{1}_{0}\frac{(-\log v)^{n-1}(-uv)^{k}}{\mathrm{\Gamma}(n)k!}{\rm d}v
=k=0(u)k(k+1)nk!\displaystyle=\sum^{\infty}_{k=0}\frac{(-u)^{k}}{(k+1)^{n}k!}
=k=0(u)kk!(j=1kjj=1k(j+1))n\displaystyle=\sum^{\infty}_{k=0}\frac{(-u)^{k}}{k!}\Bigg{(}\frac{\prod^{k}_{j=1}j}{\prod^{k}_{j=1}(j+1)}\Bigg{)}^{n}
=nFn(1,,1;2,2;u),\displaystyle=\;_{n}\mathrm{F}_{n}(1,\dots,1;2\dots,2;-u),

where the second equality follows from the proof of Corollary 2 and which is also entire.

For the type III, note that

ϕX˘1(n)(u)=exp(λ(01esnu/Γ(n+1)ds1)),u\phi_{\breve{X}^{(n)}_{1}}(u)=\exp\bigg{(}\lambda\bigg{(}\int^{1}_{0}e^{-s^{n}u/\mathrm{\Gamma}(n+1)}{\rm d}s-1\bigg{)}\bigg{)},\quad u\in\mathbb{C}

applies to non-integer values of nn as well. The integral can be directly evaluated with the substitution snss^{n}\mapsto s, leading to the third formula in (1).

The formulae (1) for the Lévy densities are direct consequences of (2) or (2) after taking ν(dz)=λδ{1}(dz)\nu({\rm d}z)=\lambda\delta_{\{1\}}({\rm d}z) with the dirac measure. ∎

Proposition 2

Proof..

Given n++n\in\mathbb{N}_{++}, starting with the LT of Y1(n)Y^{(n)}_{1}, we have using binomial expansion and the proof of Corollary 2 that for c>0c>0,

1Γ(n)01(logv)n1(uv+b)cdv\displaystyle\frac{1}{\mathrm{\Gamma}(n)}\int^{1}_{0}(-\log v)^{n-1}(uv+b)^{c}{\rm d}v =k=001(ck)bc(logv)n1Γ(n)(uvb)kdv\displaystyle=\sum^{\infty}_{k=0}\int^{1}_{0}\binom{c}{k}\frac{b^{c}(-\log v)^{n-1}}{\mathrm{\Gamma}(n)}\bigg{(}\frac{uv}{b}\bigg{)}^{k}{\rm d}v
=bck=0(ck)(u/b)k(k+1)n\displaystyle=b^{c}\sum^{\infty}_{k=0}\binom{c}{k}\frac{(u/b)^{k}}{(k+1)^{n}}
=bck=01k!(ub)k(1)k(k!)nj=1k(jc1)((k+1)!)n\displaystyle=b^{c}\sum^{\infty}_{k=0}\frac{1}{k!}\bigg{(}\frac{u}{b}\bigg{)}^{k}\frac{(-1)^{k}(k!)^{n}\prod^{k}_{j=1}(j-c-1)}{((k+1)!)^{n}}
=bn+1cFn(1,,1,c;2,,2;ub),\displaystyle=b^{c}\;_{n+1}\mathrm{F}_{n}\bigg{(}1,\dots,1,-c;2,\dots,2;-\frac{u}{b}\bigg{)},

where in the first equality the interchange of integration and summation for u(D0(b))(,b]u\in(D_{0}(b))^{\complement}\setminus(-\infty,-b] is justified by analytic continuation.

The limiting case c=0c=0 (corresponding to the gamma process) cannot be immediately gleaned from the last result. Similarly, by performing a separate computation we have

1Γ(n)01(logv)n1logbuv+bdv\displaystyle\frac{1}{\mathrm{\Gamma}(n)}\int^{1}_{0}(-\log v)^{n-1}\log\frac{b}{uv+b}{\rm d}v =k=101(logv)n1kΓ(n)(uvb)kdv\displaystyle=\sum^{\infty}_{k=1}\int^{1}_{0}\frac{(-\log v)^{n-1}}{k\mathrm{\Gamma}(n)}\bigg{(}-\frac{uv}{b}\bigg{)}^{k}{\rm d}v
=k=1(u/b)kk(k+1)n\displaystyle=\sum^{\infty}_{k=1}\frac{(-u/b)^{k}}{k(k+1)^{n}}
=k=1(ub)k(1kj=1n(k+1)j)\displaystyle=\sum^{\infty}_{k=1}\bigg{(}-\frac{u}{b}\bigg{)}^{k}\Bigg{(}\frac{1}{k}-\sum^{n}_{j=1}(k+1)^{-j}\Bigg{)}
=n(1+bu)logu+bb+buj=2nLij(ub).\displaystyle=n-\bigg{(}1+\frac{b}{u}\bigg{)}\log\frac{u+b}{b}+\frac{b}{u}\sum^{n}_{j=2}\mathrm{Li}_{j}\bigg{(}-\frac{u}{b}\bigg{)}.

Thus, after some rearrangement we arrive at the first formula in (2).

Explicit computation of the Lévy density ~(n)\tilde{\ell}^{(n)} (n++n\in\mathbb{N}_{++}) is subtler because the integral inside (2) is concentrated on the tail behavior of ν\nu, which obstructs direct expansion. Alternatively, observe that the recurrence relation leading to it, i.e.,

~(n)(z)=1~(n1)(yz)ydy,z>0,\tilde{\ell}^{(n)}(z)=\int^{\infty}_{1}\frac{\tilde{\ell}^{(n-1)}(yz)}{y}{\rm d}y,\quad z>0, (A.5)

is nothing but an Euler integral equation for the G-function (documented in [Bateman and Erdélyi, 1954, Eq. 20.5.2] [11]), or

01yGp,ql,m(α1,,αpβ1,,βq|wy)dy=Gp+1,q+1l+1,m(α1,,αp,10,β1,,βq|w)\int^{\infty}_{0}\frac{1}{y}\mathrm{G}^{l,m}_{p,q}\bigg{(}\begin{array}[]{cc}\alpha_{1},\dots,\alpha_{p}\\ \beta_{1},\dots,\beta_{q}\end{array}\bigg{|}wy\bigg{)}{\rm d}y=\mathrm{G}^{l+1,m}_{p+1,q+1}\bigg{(}\begin{array}[]{cc}\alpha_{1},\dots,\alpha_{p},1\\ 0,\beta_{1},\dots,\beta_{q}\end{array}\bigg{|}w\bigg{)}

to be precise, where 0mp0\leq m\leq p and 0lq0\leq l\leq q are all integers and ww(z)>0w\equiv w(z)>0 is linear. By employing an induction argument the general solution to (A.5) reads

~(n)(z)=dGp+n,q+nl+n,m(α1,,αp,1,,1n0,,0n,β1,,βq|w),n,\tilde{\ell}^{(n)}(z)=d\mathrm{G}^{l+n,m}_{p+n,q+n}\bigg{(}\begin{array}[]{cc}\alpha_{1},\dots,\alpha_{p},\overbrace{1,\dots,1}^{n}\\ \underbrace{0,\dots,0}_{n},\beta_{1},\dots,\beta_{q}\end{array}\bigg{|}w\bigg{)},\quad n\in\mathbb{N},

for some yet-to-be-determined numbers m,l,p,qm,l,p,q\in\mathbb{N} and d>0d>0. Given the initial condition that ~(0)(z)=aebz/zc+1\tilde{\ell}^{(0)}(z)=ae^{-bz}/z^{c+1}, z>0z>0, we are forced to choose m=p=0m=p=0 and l=q=1l=q=1, with β1=c1\beta_{1}=-c-1, in which case the G-function specializes for n=0n=0 to

G0,11,0(c1|w)=ewwc+1,\mathrm{G}^{1,0}_{0,1}\bigg{(}\begin{array}[]{cc}\\ -c-1\end{array}\bigg{|}w\bigg{)}=\frac{e^{-w}}{w^{c+1}},

from where it remains to set w=bzw=bz and d=abc+1d=ab^{c+1}.

The formulae for the type III are easily obtained by adapting the binomial expansion argument used for the type II. ∎

Proposition 3

Proof..

Let XX be a Poisson process and consider its constant mixture plus an independent Brownian motion. From Subsection 3.1 and Subsection 4.1 we know that the model cumulants (with regulation degrees) over the time period of Δ\varDelta are given by

KΔ(n)(1)=Δ(μnρ(1,n)λnbn),KΔ(n)(2)=Δ(σn2+ρ(2,n)λnbn2),\displaystyle K^{(n)}_{\varDelta}(1)=\varDelta\bigg{(}\mu_{n}-\rho(1,n)\frac{\lambda_{n}}{b_{n}}\bigg{)},\quad K^{(n)}_{\varDelta}(2)=\varDelta\bigg{(}\sigma^{2}_{n}+\rho(2,n)\frac{\lambda_{n}}{b^{2}_{n}}\bigg{)},
KΔ(n)(3)=Δρ(3,n)λnbn3,KΔ(n)(4)=Δρ(4,n)λnbn4,n𝔑,\displaystyle K^{(n)}_{\varDelta}(3)=-\varDelta\rho(3,n)\frac{\lambda_{n}}{b^{3}_{n}},\quad K^{(n)}_{\varDelta}(4)=\varDelta\rho(4,n)\frac{\lambda_{n}}{b^{4}_{n}},\quad n\in\mathfrak{N},

where recall that 𝔑+\mathfrak{N}\subsetneq\mathbb{R}_{+} is the truncated and discretized nn-domain for experimentation, and ρ(m,n)\rho(m,n)’s are specified in (5.1.1). We then immediately have the following quotient relation:

KΔ(n)(4)KΔ(n)(3)=ρ(4,n)ρ(3,n)1bn,\frac{K^{(n)}_{\varDelta}(4)}{K^{(n)}_{\varDelta}(3)}=-\frac{\rho(4,n)}{\rho(3,n)}\frac{1}{b_{n}},

which combined with the four moment conditions in (5.1.2) pins down the rate parameter estimate b^n\hat{b}_{n}, and subsequently the other estimates in (5.1.3). ∎

Proposition 4

Proof..

Consider the case of the Gaussian mixture of XX being a tempered stable subordinator. Using the results obtained in Subsection 3.2 and Subsection 4.2 we have the following model cumulants:

KΔ(n)(1)\displaystyle K_{\varDelta}^{(n)}(1) =Δ(μn+θnρ(1,n)Kn;Δ(1)),\displaystyle=\Delta(\mu_{n}+\theta_{n}\rho(1,n)K_{n;\varDelta}(1)),
KΔ(n)(2)\displaystyle K_{\varDelta}^{(n)}(2) =Δ(ρ(1,n)Kn;Δ(1)+θn2ρ(2,n)Kn;Δ(2)),\displaystyle=\Delta(\rho(1,n)K_{n;\varDelta}(1)+\theta^{2}_{n}\rho(2,n)K_{n;\varDelta}(2)),
KΔ(n)(3)\displaystyle K_{\varDelta}^{(n)}(3) =Δ(3θnρ(2,n)Kn;Δ(2)+θn3ρ(3,n)Kn;Δ(3)),\displaystyle=\Delta(3\theta_{n}\rho(2,n)K_{n;\varDelta}(2)+\theta^{3}_{n}\rho(3,n)K_{n;\varDelta}(3)),
KΔ(n)(4)\displaystyle K_{\varDelta}^{(n)}(4) =Δ(3ρ(2,n)Kn;Δ(2)+6θn2ρ(3,n)Kn;Δ(3)+θn4ρ(4,n)Kn;Δ(4)),\displaystyle=\Delta(3\rho(2,n)K_{n;\varDelta}(2)+6\theta^{2}_{n}\rho(3,n)K_{n;\varDelta}(3)+\theta^{4}_{n}\rho(4,n)K_{n;\varDelta}(4)),

where

Kn;Δ(1)=Γ(1c)anbn1c,Kn;Δ(2)=1cbnKn;Δ(1),\displaystyle K_{n;\varDelta}(1)=\Gamma(1-c)\frac{a_{n}}{b_{n}^{1-c}},\quad K_{n;\varDelta}(2)=\frac{1-c}{b_{n}}K_{n;\varDelta}(1),
Kn;Δ(3)=(2c)(1c)bn2Kn;Δ(1),Kn;Δ(4)=(3c)(2c)(1c)bn3Kn;Δ(1)\displaystyle K_{n;\varDelta}(3)=\frac{(2-c)(1-c)}{b_{n}^{2}}K_{n;\varDelta}(1),\quad K_{n;\varDelta}(4)=\frac{(3-c)(2-c)(1-c)}{b_{n}^{3}}K_{n;\varDelta}(1)

are the first four cumulants of XΔX_{\varDelta} with regulation degree-subscripted parameters. From here one can deduce the following cumulant quotient relations:

KΔ(n)(3)KΔ(n)(2)\displaystyle\frac{K_{\varDelta}^{(n)}(3)}{K_{\varDelta}^{(n)}(2)} =θn(1c)(3ρ(2,n)ρ(1,n)+θn2ρ(3,n)ρ(1,n)(2c)bn)(bn+θn2ρ(2,n)ρ(1,n)(1c))1,\displaystyle=\theta_{n}(1-c)\bigg{(}3\frac{\rho(2,n)}{\rho(1,n)}+\theta_{n}^{2}\frac{\rho(3,n)}{\rho(1,n)}\frac{(2-c)}{b_{n}}\bigg{)}\bigg{(}b_{n}+\theta_{n}^{2}\frac{\rho(2,n)}{\rho(1,n)}(1-c)\bigg{)}^{-1},
KΔ(n)(4)KΔ(n)(2)\displaystyle\frac{K_{\varDelta}^{(n)}(4)}{K_{\varDelta}^{(n)}(2)} =(1c)(3ρ(2,n)ρ(1,n)+6θn2ρ(3,n)ρ(1,n)2cbn+θn4ρ(4,n)ρ(1,n)(3c)(2c)bn2)(bn+θn2ρ(2,n)ρ(1,n)(1c))1,\displaystyle=(1-c)\bigg{(}3\frac{\rho(2,n)}{\rho(1,n)}+6\theta_{n}^{2}\frac{\rho(3,n)}{\rho(1,n)}\frac{2-c}{b_{n}}+\theta_{n}^{4}\frac{\rho(4,n)}{\rho(1,n)}\frac{(3-c)(2-c)}{b_{n}^{2}}\bigg{)}\bigg{(}b_{n}+\theta_{n}^{2}\frac{\rho(2,n)}{\rho(1,n)}(1-c)\bigg{)}^{-1},

which define a rational system for bnb_{n} and θn\theta_{n}. Note that the latter two conditions in (5.1.2) can be equivalently written as KΔ(n)(3)/KΔ(n)(2)=[SK][V]1/2K_{\varDelta}^{(n)}(3)\big{/}K_{\varDelta}^{(n)}(2)=[SK][V]^{1/2} and KΔ(n)(4)/KΔ(n)(2)=[EK][V]K_{\varDelta}^{(n)}(4)\big{/}K_{\varDelta}^{(n)}(2)=[EK][V]; from the first condition the sign of θn\theta_{n} is seen to be determined by that of [SK][SK], and then using the second the equation (5.1.5) governing its absolute value |θn||\theta_{n}| can be easily derived. Notice that, as long as the condition

3c2cρ(4,n)ρ(2,n)ρ(3,n)2<[EK][SK]2\frac{3-c}{2-c}\frac{\rho(4,n)\rho(2,n)}{\rho(3,n)^{2}}<\frac{[EK]}{[SK]^{2}}

is fulfilled, a real solution of (5.1.5) always exists and can be easily located numerically; otherwise, the equation may not admit a real solution. Upon rearranging terms we obtain the estimates in (4) as desired. ∎

Appendix Appendix B Alternative skewness and excess kurtosis enlargement for tempered stable subordinators

For each nn\in\mathbb{N} (including zero), let XnX_{n} be a tempered stable subordinator with parametrization {an>0,bn>0,c[0,1)}\{a_{n}>0,b_{n}>0,c\in[0,1)\}, and X0XX_{0}\equiv X (a0=aa_{0}=a and b0=bb_{0}=b). We know from [Küchler and Tappe, 2013, Eq. 2.19] [41] that the mean and variance at fixed t>0t>0 are respectively

EXn,t=Γ(1c)antbn1c,VarXn,t=Γ(2c)antbn2c.\textsf{E}X_{n,t}=\frac{\mathrm{\Gamma}(1-c)a_{n}t}{b_{n}^{1-c}},\quad\textsf{Var}X_{n,t}=\frac{\mathrm{\Gamma}(2-c)a_{n}t}{b_{n}^{2-c}}.

We stick to writing #{X,Y,Z}\#\in\{X,Y,Z\} as a placeholder for any type of regulated stochastic clocks, whose mmth cumulant is ρ(m,n)(0,1]\rho(m,n)\in(0,1] times that of XtX_{t} (see (5.1.1)). Then, from the quotient relation VarXn,t/EXn,t=(1c)/bn\textsf{Var}X_{n,t}/\textsf{E}X_{n,t}=(1-c)/b_{n} and the first two formulae in (2.3), (2.3.5), and (2.3) we have

Var#t(n)E#t(n)=(1c)ρ(2,n)bnρ(1,n).\frac{\textsf{Var}\#^{(n)}_{t}}{\textsf{E}\#^{(n)}_{t}}=\frac{(1-c)\rho(2,n)}{b_{n}\rho(1,n)}.

Following #(n)\#^{(n)} is understood to be XnX_{n}-clocks with parameters ana_{n}, bnb_{n}, and cc, as opposed to XX-clocks. If we keep the mean and variance of #t(n)\#^{(n)}_{t} to be invariant across n1n\geq 1, we need

an=ρ1c(2,n)ρ2c(1,n)a,bn=ρ(2,n)ρ(1,n)b.\displaystyle a_{n}=\frac{\rho^{1-c}(2,n)}{\rho^{2-c}(1,n)}a,\quad b_{n}=\frac{\rho(2,n)}{\rho(1,n)}b.

It then follows that

SkewXn,t\displaystyle\textsf{Skew}X_{n,t} =Γ(3c)an/bn3c(Γ(2c)an/bn2c)3/2=2cΓ(2c)1/2an1/2bnc/2=ρ(1,n)ρ1/2(2,n)SkewXt,\displaystyle=\frac{\mathrm{\Gamma}(3-c)a_{n}/b_{n}^{3-c}}{(\mathrm{\Gamma}(2-c)a_{n}/b_{n}^{2-c})^{3/2}}=\frac{2-c}{\mathrm{\Gamma}(2-c)^{1/2}a_{n}^{1/2}b_{n}^{c/2}}=\frac{\rho(1,n)}{\rho^{1/2}(2,n)}\textsf{Skew}X_{t},
EKurtXn,t\displaystyle\textsf{EKurt}X_{n,t} =Γ(4c)an/bn4c(Γ(2c)an/bn2c)2=(3c)(2c)Γ(2c)anbnc=ρ2(1,n)ρ(2,n)EKurtXt.\displaystyle=\frac{\mathrm{\Gamma}(4-c)a_{n}/b_{n}^{4-c}}{(\mathrm{\Gamma}(2-c)a_{n}/b_{n}^{2-c})^{2}}=\frac{(3-c)(2-c)}{\mathrm{\Gamma}(2-c)a_{n}b_{n}^{c}}=\frac{\rho^{2}(1,n)}{\rho(2,n)}\textsf{EKurt}X_{t}.

Therefore, according to the last two formulae in (2.3), (2.3.5), and (2.3) we obtain for any nn\in\mathbb{N}

Skew#t(n)=ρ(1,n)ρ(3,n)ρ2(2,n)SkewXt,EKurt#t(n)=ρ2(1,n)ρ(4,n)ρ3(2,n)EKurtXt,t>0,\textsf{Skew}\#^{(n)}_{t}=\frac{\rho(1,n)\rho(3,n)}{\rho^{2}(2,n)}\textsf{Skew}X_{t},\quad\textsf{EKurt}\#^{(n)}_{t}=\frac{\rho^{2}(1,n)\rho(4,n)}{\rho^{3}(2,n)}\textsf{EKurt}X_{t},\quad t>0,

which as well apply to any non-integer values of n>0n>0. Specifying the last results for ρ(m,n)\rho(m,n) according to (5.1.1) delivers the results.