1. Introduction
In the context of population dynamics, the Gompertz curve represents one of the most adaptable models of population growth with variable rate. Precisely, Gompertz proposed the well-known curve to model population growth under the assumption that the mortality rate grows exponentially with the age (see [
1]). After this, the Gompertz curve has been shown to be a quite useful model for phenomena that exhibit an intrinsic ageing effect, as for instance tumour incidence (see [
2]). Gompertz-type models are not limited to phenomena involving human age. For instance, such kind of models have been widely used in the context of tumour growth (see [
3]), as the growth of the radius of a multicell spheroid is influenced by the inhibition effect of its necrotic core (see [
4]), which grows together with the tumour itself.
Different generalizations of the Gompertz curve have been presented in literature. An example is given by the generalized logistic curve (see [
5]), which covers not only the Gompertz curve, but a wide family of growth curves depending on the choice of the exponents of the logistic equation and the rate function. On the other hand, to include background noise effects, stochastic generalizations of the Gompertz curve, obtained via diffusion processes, have been introduced in [
6] and then further extended to non-homogeneous diffusions in [
7], while first passage time problems for them have been studied, e.g., in [
8,
9]. Such models have been widely used, for instance, to study the effects of therapy on tumours (see [
10] and references therein). The importance of the stochastic interpretation of the Gompertz curve is underlined in [
11], where the crucial role of the noise in growth phenomena is highlighted.
Among the generalizations of Gompertz-type models, we also find fractional order Gompertz curves. Fractional calculus has been applied to a lot of different fields (see [
12] and references therein) and several papers and books on fractional differential equations have been produced (see, for instance [
13,
14,
15] and references therein). Moreover, different papers on numerical algorithms to solve fractional differential equations have been published (see, for instance [
16] for a survey of numerical methods). Several works on the subject have been produced in the last two years, such as for instance [
17], in which a Crank–Nicolson scheme has been used to solve a fractional order PDE [
18,
19,
20,
21], in which different methods involving Chelyshkov polynomials have been studied, and [
22], in which Gegenbauer wavelets are used. Moreover, fractional stochastic differential equations, which are defined by means of the Lévy–Liouville fractional Brownian motion, have been recently studied in [
23]. Among the applications of fractional differential equations, in [
24] a fractional order Gompertz curve has been considered to study a tumour growth model whose rate exhibits a non-exponential trend. On the other hand, another fractional order Gompertz curve has been proposed in [
25] to study different biological phenomena, such as dark fermentation. The latter fractional-order curve is constructed by means of Caputo fractional derivatives with respect to increasing functions, that were introduced in [
26]. This type of derivative can be used to obtain chain rules for fractional derivatives of composite functions (with increasing ones), as obtained for Riemann–Liouville fractional derivatives in [
27]. A similar chain rule can be obtained for the change of variables
, by means of right-sided fractional derivatives that naturally arise in the integration by parts formula, as shown in [
28]. In [
29] both the approaches in [
24,
25] have been considered to introduce a Gompertz model with two, eventually different, fractional orders and its stochastic counterpart.
Lately, fractional calculus has been further generalized to include more complicated orders. This is the case, for instance, of the tempered fractional calculus [
30] and of the distributed order fractional calculus [
31]. These new operators fall into the wider setting of the generalized fractional calculus, introduced in [
32] via Stieltjes measures and in [
33] via Lévy measures. Both approaches are correlated with the good properties of Bernstein functions [
34]. As observed in the aforementioned literature, there is a strict link between the (generalized) fractional calculus and the theory of Lévy processes, which has been underlined, for instance, in [
35]. This link is revealed to be crucial to determining different properties of solutions of linear and nonlinear generalized fractional differential equations. In [
36], relaxation equations for generalized fractional derivatives are studied and different properties of the eigenfunctions of the aforementioned nonlocal derivatives are obtained, underlining the connection between such relaxation pattern and semi-Markov processes. The growth equation has been studied in [
37] in the case of complete Bernstein functions. In [
38], the existence and uniqueness of solutions for nonlinear generalized fractional differential equations have been proved by means of a fixed-point argument, while a generalized Grönwall inequality is proved with the same strategy as in [
39] for the fractional case. In [
36,
37,
38], eigenfunctions of the generalized fractional derivatives are recognized as Laplace transform or moment generating functions of particular stochastic processes. Explicit formulae for such functions are not known in the general case, while in the standard fractional one they are recognized as Mittag–Leffler functions in [
40]. In [
41], the existence, uniqueness and spectral decomposition of exact solutions of some time-nonlocal parabolic equations are obtained by means of stochastic representation results. Here, we will use both the representation of the eigenfunctions as obtained in [
36,
37,
38] and the regularity properties of solutions of generalized fractional linear differential equations proved in [
41].
As we already stated, fractional calculus was introduced in population dynamics models to consider phenomena with underlying memory effects, as done in [
24,
25]. However, fractional calculus relies only on one particular type of memory kernel, that is, the Riesz kernel. To consider more complicated memory effects, other more general kernels have to be considered. This can be achieved by using the tools of the generalized fractional calculus, as done, for instance, in [
37] in the case of the (Malthusian) growth equation and in [
42] in the case of the logistic equation. On the other hand, the introduction of noise in growth phenomena is also mandatory to describe possible random fluctuations, as done in [
6,
7] for the classical Gompertz curve and in [
29] for the fractional Gompertz curves.
The problem we address in this paper is twofold: on one hand, we want to extend the definition of fractional Gompert curves to cover a wider range of memory kernels; on the other hand, we want to introduce noise in the aforementioned models in a coherent way. Thus, here we consider a further generalization of the construction presented in [
29] to the case of fractional orders expressed in terms of complete Bernstein functions. In particular, we first construct some deterministic generalized fractional Gompertz curves and then we introduce their stochastic counterparts. Differently from [
29], we consider a generalization of lognormal processes to guarantee that the stochastic Gompertz curves remain positive and, at the same time, their medians represent the deterministic ones. The paper is structured as follows:
In
Section 2 we give some preliminaries on Bernstein functions;
In
Section 3 we introduce generalized fractional integrals and derivatives. In particular we give some chain rules involving generalized fractional derivatives with respect to other functions and we study the properties of the eigenfunctions of the defined operators;
In
Section 4, we study the properties of the solutions of some linear stochastic equations involving generalized fractional integrals. In particular, we show that such equations admit a unique Gaussian solution and that, under some suitable assumptions on the noise process, its expectation solves a linear generalized fractional differential equation;
In
Section 5, we use the operators introduced in
Section 3 to construct generalized fractional Gompertz curves and the Gaussian processes given in
Section 4 to determine their stochastic counterparts;
In
Section 6 we present a short summary of the main results of the paper and we compare them with some pre-existing literature. Finally, some highlights for future works are given.
2. Preliminaries and Notation
In this section we provide some preliminary definitions and lemmas.
Definition 1. We say that a function (where ) is a Bernstein function (see [34]) if and only if and, for any , it holds The convex cone (see [34] [Corollary ]) of Bernstein functions will be denoted as . A Bernstein function is said to be special if and only if the conjugate functionis still a Bernstein function. The class of special Bernstein functions will be denoted as . To characterize Bernstein functions, let us recall the following theorem, known as Lévy-Khinchine representation theorem (see [
34] [Theorem
]).
Theorem 1. A function belongs to if and only if there exist two constants and a non-negative measure on such thatand The measure is called the Lévy measure of . Vice versa, any triplet , where and is a non-negative measure on satisfying condition (1), defines a unique via Equation (2). The constants and are called respectively the killing and the drift coefficient of . We will denote . However, we need Bernstein functions whose Lévy measure is more regular. To do this, we refer to the following definition.
Definition 2. A Bernstein function is said to be complete if its Lévy measure admits a completely monotone density, that is, for some function such that We denote the convex cone (see [34] [Corollary ]) of complete Bernstein functions as . Bernstein functions can be recognized as Laplace exponents of particular Lévy processes. Indeed, let us recall the following definition.
Definition 3. A subordinator (see [43] [Chapter ]) is a non-decreasing Lévy process. Given a subordinator and a positive constant , the processwhere is an exponentially distributed random variable, with rate a, independent of , is called a subordinator killed at rate a. Concerning the link between subordinators and Bernstein functions, we have the following Theorem (see [
34] [Theorem
]).
Theorem 2. For any there exists a unique (possibly killed) subordinator such that Vice versa, for any (possibly killed) subordinator there exists a Bernstein function such that Equation (3) holds. Since we focus on Bernstein functions, we will usually denote a subordinator by , referring to the fact that its Laplace exponent is given by .
For any subordinator, the following occupation measure can be defined.
Definition 4. Let be a subordinator. The potential measure of on is defined aswhere is the Borel σ-algebra of and is the indicator function of the Borel set A. We will denote the distribution function of the potential measure and we will usually refer to it directly as a potential measure. Moreover, we can define a right-continuous inverse for the process .
Definition 5. Let be a subordinator. For any we definecalled inverse subordinator. Remark 1. As shown in [35] [Theorem ], if , then is an absolutely continuous random variable for any . Let us denote by its probability density function. Once we have defined the inverse subordinator, by using the fact that
and
are increasing, we have
Concerning special Bernstein functions, the associated potential measure is
almost absolutely continuous, except at most for a jump in 0, as stated in the following theorem (see [
34] [Theorem
]).
Theorem 3. Let . Then if and only if there exists a non-negative and non-increasing function such that andwhere is Dirac’s δ measure centered in 0 and In particular, if or and , then is absolutely continuous with density given by , called the potential density of .
Remark 2. Let us emphasize that, for fixed , denoting by , , respectively the killing coefficient, the drift coefficient and the Lévy measure of , it holds (see [34] [Equation (10.9)]): Hence, if and only if , that is to say if and only if or Let us recall that the following inclusion holds:
We will work with a specific subset of complete Bernstein functions. Hence, let us introduce the following notation:
where
is the identity map. Hereafter, we consider the following Assumption.
Assumption A1. and there exist , and such that Remark 3. The previous Assumption guarantees that, for any , there exists a constant C such that 3. Generalized Fractional Integrals and Derivatives
Let us first introduce some generalized fractional integrals.
Definition 6. Set a Banach space . For any , given a function , we say that if f is Bochner-integrable, that, by Bochner’s Theorem (see [44] [Theorem ]) is equivalent to asking that f is measurable and is Lebesgue-integrable, i.e., . Given a function we say that if and only if for any . For any function we define the generalized fractional integral of the first kind of f induced by aswhere is the potential density of and the integral is a Bochner integral. We define the generalized fractional integral of the second kind of f induced by aswhere is the tail of the Lévy measure of and the integral is a Bochner integral. Remark 4. First, let us observe that, by [44] [Proposition ], the quantites defined in Equations (4) and (5) are well-defined. Let us also underscore that, by Remark 2, as , the operators and coincide if and only if or . In general, it holds, for any , With the help of the previously introduced operators, we can define the following generalizations of both Riemann–Liouville and Caputo derivatives, introduced first in [
32] for complete Bernstein functions and then in [
33] in the general case.
Definition 7. Set a Banach space . For any function we define the generalized Riemann–Liouville derivative induced by , with , aswhere the integral is a Bochner integral, provided that the involved quantities exist. Given , we say that a function is absolutely continuous if there exists a function such that andand we denote it by . We denote by the set of functions such that for any it holds . When , we do not specify X. For any we define the generalized Caputo derivative of f induced by , with , aswhere the integral is a Bochner integral. Remark 5. Let us remark that, if , then one can show that . Thus, the quantity defined in Equation (7) is well-defined for any . However, if , there exist some functions such that . For instance, if , this is the case of . Indeed, by Equation (6) and [34] [Theorem ] we getthus . However, since , we have , hence . Thus, in general, the domain of properly contains . On the other hand, the quantities defined in Equation (8) are well-defined if and only if by [44] [Proposition ]. Such operators are generalizations of the well-known Caputo and Riemann–Liouville fractional derivatives, which are achieved in the case with . Indeed, in this case, and .
Remark 6. If , then we define . In general, for such that the operators are defined asand Let us consider any
. By a simple application of Fubini’s theorem we get
where we also used [
45] [Chapter 6, Theorem 11]. Differentiating (almost everywhere) in both sides of the previous relation we get
With this relation in mind we can extend the definition of generalized Caputo derivative to a (possibly) larger class of functions.
Definition 8. Set a Banach space . For any function we define the regularized generalized Caputo derivative induced by aswhenever the right-hand side is well-defined. Remark 7. Equation (10) justifies the fact that we are using the same symbol of the generalized Caputo derivative. Concerning the inversion of such operators, let us observe that, as shown in [
36] [Section
], if
f is a function such that
is well defined, then
thus we can see the operator
as the inverse of the generalized Caputo derivative
.
In the case
, it holds
, where
is the fractional integral of order
(see [
13] [Chapter 2]),
is the Riemann–Liouville fractional derivative of order
and
is the Caputo fractional derivative of order
.
Here, we also need other generalized fractional operators, that is, the integral and the derivative of a function with respect to an increasing function. The definitions we give are analogous to the ones given in [
27] for the Riemann–Liouville fractional derivative and [
26] for the Caputo one.
Definition 9. Set a Banach space and consider a strictly increasing function . For any measurable function , we define the generalized fractional integral of the second kind of f induced by with respect to the function aswhere is the tail of the Lévy measure of and the integral is a Bochner integral, provided that the involved quantities exist. Suppose now some . For any measurable function we define the generalized Riemann–Liouville derivative induced by , , with respect to aswhere the integral is a Bochner integral, provided that the involved quantities exist. Moreover, for any , we define the generalized Caputo derivative of f induced by , , aswhere the integral is a Bochner integral, provided that the involved quantities exist. Remark 8. Let us underline that we do not really need for all , but only on the points in which we want to define . Moreover, we can formally define the Caputo derivative even if , since does not play any role in the first equality of Formula (13). Finally, if , we define, for any , Let us stress that if
the operators coincide with the ones introduced in [
26,
27]. Now we want to prove a chain rule, analogous to the one given in [
29] [Proposition 1].
Proposition 1. Fix strictly increasing with and . Let and for . Then, the following properties are true:
- 1.
- 2.
If it holdsprovided one of the involved quantities exists; - 3.
If and is locally Lipschitz in , then and it holds - 4.
If , is locally Lipschitz in and , it holds
Proof. Let us argue for , since the case is trivial.
By the definitions of
in Equation (
12) and
in Equation (
5) we have
where we used the change of variables
.
Concerning claim
, it follows from
by considering
in place of
, differentiating both sides of (
14) and multiplying by
.
Let us prove claim
. First of all
since it is composition of an absolutely continuous function with a locally Lipschitz one. In particular, it holds
and then
where we again used the change of variables
.
Finally, concerning claim
, we have, by claims
and
and the fact that
(and thus
),
□
Remark 9. Let us observe that claims and also hold if Ψ is not locally Lipschitz, but . Indeed, both claims directly follow from the fact that in such case .
Moreover, the last Proposition tells us that the quantity (12) is well defined for any measurable function with the property that there exists such that (equivalently ). This is the case, for instance, of , since, by Continuous Inverse Theorem [46] [Theorem ], . Finally, the quantity in Equation (13) is well defined if is locally Lipschitz in whenever with the property that there exists such that (this is, for instance, the case in which and is bi-Lipschitz). If is not locally Lipschitz, then the quantity in Equation (13) is still well-defined if the function g defined as above belongs to . Eigenfunctions of the Generalized Fractional Derivatives of Caputo Type
In the following we need to characterize the eigenfunctions of the generalized fractional derivatives of Caputo type that have been previously introduced. First of all, let us recall that the function,
is well-defined for any
(see [
38] [Lemma
]). Let us also recall the following Proposition (see [
38] [Proposition
]).
Proposition 2. Consider , with and , and . Then is the unique solution of Remark 10. If then .
Remark 11. As a consequence of [41] [Theorem ], if then not only belongs to , but also admits an analytic extension on a sector for some . In the following we need to extend the definition of
for fixed
to negative values of
t. To do this we set
that is a continuous monotone function.
Remark 12. Observe that .
can be recognized as the eigenfunction of a particular non-local operator. Indeed, let us define the following operator.
Definition 10. Set a Banach space . We say that a measurable function belongs to , where , if and only if the function , defined as for any , belongs to .
For any function we define the right generalized Riemann–Liouville derivative induced by , with , as We say that belongs to if and only if the function belongs to .
Moreover, for any , we define the right generalized Caputo derivative induced by , with , as Since for any it holdswe can extend the definition of right generalized Caputo derivative to non absolutely continuous functions via Equation (15), supposed that the function admits a generalized Riemann–Liouville derivative. If , we define .
We say that a measurable function belongs to if . We say that a measurable function belongs to if .
For any function we define the bilateral generalized Riemann–Liouville derivative induced by asand the bilateral generalized Caputo derivative induced by asprovided the involved quantities exist. Remark 13. The definition of right derivative is given by taking in consideration the integration by parts formula (see, for instance [28]). However, with we want to mimic the behaviour of the derivative on the whole real line, thus we need to introduce another − sign on the right derivative. Recall that if we get . Obviously, the bilateral derivatives are well-defined on .
Proposition 3. Fix and let and such that for any . The following properties are true:
- 1.
It holdsprovided one of the involved quantities exists; - 2.
It holdsprovided one of the involved quantities exists.
Proof. Let us first observe that
where we used the change of variables
. Differentiating on both sides we get
By using (
15), we conclude the proof. □
As a direct consequence of Propositions 1–3, we obtain the following result.
Proposition 4. Fix and with . The following statements hold:
The function is the unique solution of The function is the unique solution of
We already know, by definition, that is nondecreasing. We want to prove that is strictly increasing. We actually have a stronger result.
Proposition 5. Let satisfy Assumption A1. Then, for any , it holds and Proof. By the fact that
, we know, by Remark 11, that
. Moreover, let us recall, from [
38] [Lemma
],
where
First of all, since
, then
with derivative
and
. Moreover, let us observe that, by definition,
and then
for any
. Observe also that
is well defined as
.
Now let us show that if
with derivative
, then also
with derivative
. Indeed, we have that
is absolutely continuous with
and then
Now let us observe that , being the convolution product of and that are both in . Thus, is sum of two functions in . By induction, we know that, for any , it holds with derivative .
Furthermore, let us observe that
for
and that, if
, then Equation (
17) implies that
, so that we can conclude that
Being
, then, by [
34] [Proposition
], also
. By Remark 2 we know that
and
If
, then
and
for any
. If
, then
. Hence, by [
34] [Proposition
], being
, we know that
cannot have bounded support. As it is also decreasing, it holds
for any
. Thus, in general, we conclude that
By [
47] [Lemma
] we can differentiate both sides of (
16) to achieve, by also using Equations (
18) and (
19),
concluding the proof. □
Corollary 1. Let . Then, for any , the function is continuous and strictly increasing.
Proof. Let us prove the statement for
, since the proof is analogous as
. Let us first recall that if
, then
is completely monotone (see [
36] [Theorem
]). Thus, we have that
is strictly increasing on
. On the other hand, Proposition 5 implies that
is strictly increasing on
. Finally, the fact that
concludes the proof. □
Remark 14. The function could be non (right-)differentiable in 0. For instance, consider with . Then it is known (see [40]) that , where is the Mittag–Leffler function defined as Thus, we haveand then . However, since is monotone, by [47] [Theorem ] we know that . 6. Conclusions
In this paper, we used the theory of complete Bernstein functions and the tools from generalized fractional calculus to extend the fractional Gompertz curves introduced in [
24,
25,
29]. Precisely, the classical Gompertz equation is decomposed in the rate equation and the curve equation and then a fractionalization (or, in this case, nonlocalization) procedure is applied to both, obtaining the generalized fractional Gompertz curves. For their stochastic counterparts, we first studied in
Section 4 a linear integral equation that plays the role of the equation of the rate process and then, in
Section 5, we introduced a generalization of lognormal processes to obtain the desired generalized fractional stochastic Gompertz curves.
A further extension to other growth curves with different rate functions could rely on nonlinear generalized fractional differential equations. While, on one hand, a theory of nonlinear generalized fractional differential equations is currently being developed (see, for instance [
38]), to obtain stochastic growth curves of such type one could need some sort of nonlinear generalized fractional stochastic differential equations. In future works we will focus on generalizations of the Lévy–Liouville fractional Brownian motion (which is different from the well-known Mandelbrot-Van Ness fractional Brownian motion) to consider stochastic differential equations of the aforementioned type (see, for instance [
23] for the fractional case). Let us also underline that the approach adopted in
Section 4 can be easily extended not only to the multivariate setting, but also to stochastic processes defined on Banach spaces, the latter by using cylindrical noise processes, such as the cylindrical (possibly fractional) Brownian motion (see, for instance [
51]). One can also define multivariate
-normal distributions by a simple
vectorization argument. However, once one moves from the one-dimensional to the
n-dimensional setting, the discussion on the median cannot be reproduced as it is, but, instead, one should consider quantile contours (see, for instance [
52]).
The models presented in this paper represent a natural step forward from [
24,
25]. Indeed, while in [
24,
25] the authors consider only Riesz kernels, here we propose a wide family of memory kernels (of which Riesz kernels are particular cases) that can be used. Moreover, analogously as done in [
29] for the deterministic models, here one can consider two different memory kernels acting on the rate function and/or on the curve itself. Concerning the introduction of the noise, let us observe that stochastic models are shown to be useful, for instance, to study random fluctuations in mathematical oncology models (see, e.g., [
8,
9,
10]). In [
24] a fractional-order model has been used to describe tumour growth. Both the models presented in [
29] and this paper provides some methods to introduce the noise in such fractional-order growth curves. There are three main differences between the present paper and [
29]:
In [
29] we considered only couples of Riesz kernels, while here we can consider any couple of suitable memory kernels;
Independently of the fact that memory effects are introduced in the rate function and/or in the curve, the generalized fractional stochastic Gompertz curve presented here is non-negative (as shown in Proposition 9), while this is not true in [
29] in the case in which the curve function is defined itself via a fractional differential equation;
In [
29], when the curve is defined via a fractional differential equation, the deterministic model is re-obtained by considering the mean of the stochastic curve. Here, in any case, the deterministic model is provided by the median of the stochastic curve, as in the classic case.
For these reasons, we think that the generalized fractional stochastic Gompertz curves defined in this paper are more realistic and more general models to describe Gompertz-type growth phenomena with memory and noise, thus they represent a step forward with respect to [
29]. However, restricting the view to a specific (parametrized) family of Bernstein functions permits a better
calibration of the used model. Hence, before using the models we discussed here, it is advisable to consider an ansatz on the couple of Bernstein functions to consider.
In both deterministic and stochastic models, the tools to obtain such generalizations are provided by fractional and generalized fractional calculus, with particular attention to inversion formulae given in [
36] and Grönwall-type inequalities (see [
38,
39]).
Finally, let us remark that the aim of the paper is to present a family of models that can be used for population growth phenomena with memory and/or noise. As evidenced by the cited literature, these kinds of phenomena naturally arise in physics, engineering, biology and social sciences and then more general models are useful tools for improving the knowledge about them.