1. Introduction
Physics is a science devoted to the investigation of natural phenomena. The primary objective is to gather experimental data pertinent to a phenomenon and formulate a mathematical representation. The mathematical descriptions follow from assumptions taken as fundamental laws considered obvious. Among them, conservation laws are fundamental principles in physics. They describe the preservation of certain quantities in isolated systems over time by governing a wide range of phenomena and play a crucial role in our understanding of the universe. For example, the principle of energy conservation demands that energy cannot be generated or annihilated within an isolated system; linear and angular momentum conservation states that the total momentum and the angular momentum of an isolated system remain constant if no external forces act on it; and the conservation of the number of particles or of the electric charge must be conserved in a closed volume whenever annihilation/creation phenomena are absent inside the volume itself.
In a broader sense, by noting that a specific physical quantity must be conserved, we can formulate a mathematical expression for the related conservation law by examining a closed volume through the use of continuity equations. In fact, continuity equations play a fundamental role in several science and engineering contexts, describing the conservation of several quantities like particles, charges, or others within a system. These equations provide valuable insights into the behavior of fluids, electrical currents, and other physical phenomena. At its core, a continuity equation states the rate of change of a quantity within a given volume that must be equal to the net flow of the same quantity into or out of the volume. In fact, if the considered system can exchange a physical quantity with the external world, the rapidity of using the quantity contained in it has to be identical to the rapidity by means of which the same quantity is leaving the bulk across its limiting surface. This is the classical continuity equation in its integral form. Of course, to write it as a mathematical expression, we need to define the concept of rapidity of time variation of the given quantity
. A rather simple definition is to consider its time derivative
, which is a local-in-time property. In a more general fashion, the local expression of a continuity equation reads
where
represents the density of the physical quantity
, according to
while
the current density of particles.
In electromagnetism theory, continuity equations are crucial in describing the behavior of electrical currents. For instance, in the case of steady-state direct current (DC) circuits, the continuity equation states that the net flow of charge into or out of a closed system must be zero. This equation ensures that charge is conserved within the circuit, forming the basis for comprehending the operation of electrical devices. It is in this context that in 1860, James Clerk Maxwell, by noticing that Ampère law alone is insufficient to explain certain phenomena, particularly those involving time-varying electric fields, introduced the displacement current as an additional term charge particle current, given by
where
is the electric permittivity and
the local electric field.
The introduction of the displacement current in the framework of electromagnetism represents one of the most fundamental examples in physics in which the continuity equation played a fundamental role in the comprehension of the underlying physical law, and it was vital in the formulation of Maxwell’s equations, being a cornerstone in the development of the theory.
Often, continuity equations are strictly related to the famous Noether theorem, formulated in 1915, which plays an impressive role as a fundamental result in theoretical physics. It reveals a profound link between the symmetries inherent in a physical system and the conserved quantities that arise from the associated continuity equation. In fact, this theorem shows that if the dynamical evolution of a physical system can be obtained through the principle of minimal action, a variational problem involving the time-integral of the Lagrangian system, then there is a conserved quantity corresponding to each continuous symmetry of the action. Consequently, a corresponding continuity equation can be established. The power of the Noether theorem lies in its generality. It applies to a wide range of symmetries, including time symmetry as well as the translational and rotational symmetries that, according to this theorem, are at the base of the existence of the continuity equation for energy, linear, and angular momentum.
In this sense, the Noether theorem has profoundly impacted the development of modern physics, where symmetries and conservation laws are central concepts. In particular, it has become a powerful tool for understanding the hidden laws of nature, including the gauge symmetries, whose first example in physics arises exactly in the framework of the electromagnetic theory, these symmetries being the heart of local conservation of the electric charge.
However, there are several anomalous phenomena for which the rapidity of variation of a given quantity at time
t depends on the previous story, i.e., these phenomena have a memory of the past and a nonlocal spatial dependence. In this case, identifying the rapidity of variation of a physical quantity
using
may not adequately describe the phenomenon under investigation. This is the case in many phenomena observed in complex fluids, predominantly liquid crystal materials. Because of molecular anisometry, these materials form anisotropic mesophases that present long-range orientational order and, in some cases, positional order. Furthermore, they are highly responsive to many stimuli, including applied field, confinement (and geometry), and several others. Therefore, it is not uncommon for reports of anomalous behavior of some sort to be received in liquid crystals. For example, anomalous diffusion [
1,
2] and transport in liquid crystals [
3], anomalous variation of elastic constants near interfaces [
4], anomalous behavior of ferroelectric pitch upon a change in dopant concentration [
5], anomalous phase behavior [
6], anomalous relaxation time [
7], and many others. Hence, liquid crystals are one of several examples of complex fluids in which physical quantities do not obey a simple
behavior. For this reason, a few generalizations of the derivative have been proposed from the beginning of differential calculus, which has become a powerful tool to investigate such
anomalous behavior in recent decades. Among the several mathematical generalizations proposed, the one based on fractional calculus has caught the interest of several people working in this field. Indeed, the foundation of fractional calculus lies in the concept of fractional operators, which accommodates various non-equivalent definitions. Among these, well-established expressions of fractional operators include Riemann–Liouville [
8,
9], Caputo [
10], Caputo–Fabrizio [
11], Riesz [
12], Grünwald–Letnikov [
13,
14], Weil [
15], Hadamard [
16], Marchaud [
17], and others [
18].
Fractional calculus can be seen as an extension of standard differentiation and integration to noninteger orders [
19]. The inquiry regarding the potential significance of a derivative of any order was posed by Guillaume François Antoine, Marquis de L’Hôpital, in a letter sent to Gottfried Wilhelm Leibniz, who gave a positive and correct answer to the instigating question. Since then fractional calculus has been the object of analysis by great mathematicians, including Jean-Baptiste Joseph Fourier, Georg Friedrich Bernhard Riemann, Pierre Simon Laplace, Leonhard Euler, Paul Mathieu Hermann Laurent, Niels Henrik Abel, Joseph Liouville, Nikolay Yakovlevich Sonin, and Aleksey Vasilievich Letnikov, among many others [
9,
20,
21,
22,
23,
24,
25,
26,
27,
28,
29]. A derivative of arbitrary order was first referenced in 1819 by Sylvestre François Lacroix (1765–1843), who provided an explicit formula for a fractional derivative [
20]. These fractional-order operators are formulated through integrals, making the fractional derivative reliant on the system’s past behavior, thereby enabling the description of memory effects. In a sense, it captures the function’s behavior over a fractional number of steps, allowing for a more nuanced understanding of complex systems.
One of the crucial applications of fractional calculus is the study of anomalous diffusion processes. In classical diffusion, particles follow a random walk described by the standard second-order differential equation. Nevertheless, diffusion manifests non-Gaussian characteristics and non-Markovian dynamics in numerous empirical contexts. Hence, fractional calculus provides a mathematical framework for modeling and analyzing these processes. Fractional calculus also finds applications in signal processing, where signals with fractional derivatives, known as fractional signals, have been shown to be helpful in analyzing and processing nonstationary and nonlinear responses. Fractional calculus can be used to characterize the scaling properties and long-term memory of signals, such as the electrical impedance of an electrical device, providing valuable tools for time-series analysis and filtering.
Over recent decades, fractional calculus has found applications in variational problems within physics and engineering. As demonstrated in [
30] (and further discussed in [
31]), the fractional Euler–Lagrange equations were formulated by employing the principle of minimal action on a Lagrangian functional that incorporates the field and its fractional derivatives. Remarkably, the resulting equations bear striking similarities to those obtained from ordinary variational problems involving integer-order derivatives. Consequently, it has been natural to explore the generalization of Noether theorem within the framework of fractional calculus [
32]. In this way, new conserved quantities associated with specific symmetries of the system and related to corresponding fractional continuity equations may be obtained, mirroring the remarkable result obtained within the realm of standard calculus. We aim to provide a general framework for analyzing systems where anomalous behavior may occur, which is the case in many complex fluids. Liquid crystals, for example, are especially prone to present anomalous behavior. However, it is surprising how only a few works employ fractional calculus to study liquid crystal materials. Hence, here we provide a general toolkit for this purpose, which may be extended as necessary and, being very general, always applicable, recovering the usual parameters (without fractional derivatives) if the system does not present anomalous behaviors.
This paper seeks to acquaint the nonspecialist reader with the topic of fractional calculus and to demonstrate several applications to physical problems involving time-fractional derivatives, especially in relation to anomalous diffusion and adsorption phenomena in an electrolytic cell, with a special emphasis on complex fluids like liquid crystals. Of course, spatial fractional derivatives could also be of some importance. In this respect, several papers are devoted to generalizing vectorial differential calculus concerning Stokes, Gauss, and Green formulae, and to the fractional form of Maxwell equations (see, for instance, [
33]). Our analysis mainly focuses on diffusion or adsorption problems involving just fractional time derivatives, and we propose a generalization of their phenomenological description.
The structure of our manuscript is delineated as follows. In
Section 2, we provide a comprehensive review of fractional calculus, with a primary focus on the Riemann–Liouville and Caputo methodologies. In
Section 3, we revisit briefly its application to the study of anomalous diffusion and its possible solution within the fractional calculus formalism.
Section 4 presents the diffusion and adsorption phenomena in the fractional kinetic approach by discussing the case of an electrolytic cell with adsorbing electrodes. The model is then detailed in
Section 5, where the fractional Poisson–Nernst–Planck model is proposed, including our new proposal for the fractional displacement current, and the impedance response is analyzed for the anomalous system proposed.
Section 6 contains our concluding comments.
2. Elements of Fractional Calculus
Nowadays, the Riemann–Liouville operator is a well-established fractional derivative, although it is the result of several contributions (mainly by Liouville), which is defined as
for
, where
is the gamma function [
34],
c is the left-hand point and
t is the terminal point, and
is the order of the derivative. We notice that
n is the smallest integer greater than
, with
being the usual derivative of
. In Equation (
5), we utilize a notation pioneered by the mathematician Harold Thayer Davis (1892–1974) [
35], which is particularly useful for making the left subscript
c explicit. This subscript is not immediately apparent in a derivative and has led to extensive discussions on the complementary function [
24]. We consider simple cases to illustrate the fractional time derivative for
and
. One of them is the function
which yields
Figure 1 shows
vs.
t for
for several values of
and
.
For the exponential function, i.e.,
, we have
where
is the generalized Mittag–Leffler function [
19]. Another function is
. In this case, the fractional time derivative gives
It is worth mentioning the if
in Equation (
8), we have that
.
One can reinterpret the Riemann–Liouville fractional derivative by using the concept of an integral of arbitrary order, as detailed in the subsequent method [
19]. The Riemann–Liouville integral of arbitrary order may be defined from the Cauchy integral formula:
In Equation (
9), the parameter
is allowed to be complex provided that its real part is strictly positive, i.e.,
. For the sake of simplicity, let us temporarily assume that
. The definition provided in Equation (
9) exhibits the following properties:
Now, let us define the fractional derivative as
But this identification inspired by the usual calculus has some problems. A problem arises because the gamma function is undefined for zero or negative integer values. Therefore, we should proceed with caution when using this approach. It is important to remember that for
, we are permitted to write
In other words, applying the derivative operator
n times to a function
after integrating it
n times results in the identity operator. Similarly, we anticipate discovering a fractional derivative operator such that
that is, according to the Riemann–Liouville framework, as indicated by Equation (
11), the operator
denotes the fractional derivative of order
, acting as the left-inverse of the Riemann–Liouville integral of order
, analogous to the relationship between the standard derivative and its integral.
where
is the identity operator.
One possible operator to accomplish this task could be constructed as
where
represents the ceiling function, which determines the smallest integer that is not less than
(i.e., the subsequent integer must be considered). Some illustrative examples are as follows:
Thus, the Riemann–Liouville fractional derivative [
29] can be more rigorously defined as
that coincides with the operator (
5) with the assumption
.
Now, it should be clear that for
, such that when
, the following holds:
In fact, by using this last relation and (
10), we have
Therefore, we can rewrite relation (
17) in
that coincides with (
5) and establishes the arbitrary-order Riemann–Liouville derivative operator.
If not differently specified, let us assume from now on the left subscript and denote, for the sake of simplicity, the corresponding operator in , that is, [risp. ].
By recalling the Laplace transform definition of a function
applied to the Riemann–Liouville fractional derivative of
gives
Furthermore,
represents the Riemann–Liouville integral operator of order
. It is well-known that the Laplace transform of the
n-th derivative of a function
, for
, is expressed as
where
denotes the
n-th derivative of
.
Then, from Equation (
21) and using Equation (
17), we obtain the Laplace transform of the Riemann–Liouville fractional derivative of
:
An important remark is in order here if we want to apply this formalism to solve initial and boundary-value problems in mathematical physics. We observe that the initial conditions will be expressed using fractional derivatives, specifically,
This is not a usual way to establish initial conditions and, in some sense, may be an additional problem in practical applications of this powerful formalism.
In 1967, Caputo presented a novel idea for a fractional derivative. This idea is linked to the fractional Riemann–Liouville integral and differential operators. It is now referred to as the Caputo fractional derivative [
10]. In terms of this operator, the initial conditions can be formulated in the usual way, i.e., by stating them in terms of integer-order operators. The Caputo operator is defined as
allowing us to view it as equivalent to the Riemann–Liouville integral of the
derivative of the function
. Observe that, similar to how the Riemann–Liouville derivative operator is defined in Equation (
17), the fractional derivative of order
in the Caputo sense can be defined as
where, with respect to Equation (
17), the roles of operators
and
have been interchanged. This relation clarifies the difference between the Riemann–Liouville and the Caputo derivative operators [
36] and, in a very direct way, gives the definition above:
which matches (
25) and characterizes the Caputo derivative operator of any order. For noninteger orders, fractional derivatives generally differ because usual derivatives of order
k do not commute with integral fractional operators. The Caputo differential operator uses regular conditions on
compared to the Riemann–Liouville differential operator. According to the Cauchy integral formula, we can express
Thus, as already stated [Cfr. Equation (
14)], we expect that if
is the left derivative, then
Nevertheless, if
and its first
derivatives do not vanish at
, generally, we obtain
By performing integration by parts, we have that
Alternatively, it can be demonstrated that the Caputo operator satisfies
The substitution of Equation (
32) into Equation (
33) yields
Thus, a relationship between the previous differential operators can be expressed as follows [
37]:
which, when
, i.e.,
, yields
Based on the equation above, we can infer that these operators are equivalent when .
The paramount advantage of employing the Caputo approach lies in its initial conditions for fractional differential equations. Unlike other methods, the Caputo derivative ensures that these initial conditions mirror those of integer-order differential equations. This means they encapsulate the limit values of integer-order derivatives of the unknown functions precisely when
, namely:
meaning conditions (initial or boundary) can be represented as integer-order derivatives. This idea becomes more understandable when we look at the Laplace transform of the Caputo operator:
Upon deriving the Laplace transform of the convolution and noting that
, one can determine the Caputo derivative of order
in the Laplace domain as follows:
Notice that the initial condition is expressed using the integer derivatives.
For physical applications, it is useful to put
in the Riemann–Liouville derivative; then, we obtain that
In this scenario, both definitions coincide if
and its derivatives exhibit appropriate behavior as
, meaning
for
, where
. This property is important from a physical perspective because it allows for the consideration of stationary processes. For example, it can be applicable to fractional-order dynamical systems with periodic signals, which is necessary for impedance problems [
38], wave propagation in continuous media, and other similar scenarios.
One way to generalize these operators is by introducing the distributed-order fractional operator [
39]. This operator is essentially a linear combination of fractional operators with various arbitrary orders. A simple example is obtained if we consider the Caputo operator in the following way:
for
and
. Some special regularity and boundary behavior conditions are required for the weight function
, also called order density [
39]. For instance, when
, the operator (
40) reduces to
As we discuss in
Section 3, this formalism is shown to be very useful in handling diffusion problems involving time-fractional derivatives.
3. Fractional Diffusion Equations
Fractional diffusion equations can be derived, for instance, from continuous-time random walk (CTRW) formalisms by appropriately selecting the waiting-time and jump probability distributions [
40,
41] and from the master Equation [
42] or comb models [
43]; they can also be expressed using time-fractional derivatives of distributed order [
44]. In this latter case, the main novelty is the presence of different diffusion regimes in the theoretical approach [
45]. To investigate the capabilities of fractional calculus in modeling the diffusion process, we revisit, for demonstration purposes, the task of finding the fundamental solution to the Cauchy problem for a time- and space-fractional diffusion equation, closely following the methodology of Mainardi et al. [
46]. The fractional diffusion equation is
for
. To comprehend the fractional indices, we can employ the continuous-time random walk method to link these indices with the waiting-time and jumping distributions. For simplicity, we limit our discussion to the diffusive scenario, where
, to explain the significance of
and
. When
, we need to use the approach presented in Refs. [
47,
48]. In the first case, the process is governed by the probability density function
from which it is possible to obtain the waiting-time distribution and the jumping distributions, respectively, as follows:
By using
, the distribution related to the diffusion process
is found by combining the equations
and
where
denotes the survival probability at the initial location. By applying the Fourier–Laplace transforms (
=
and
), it is feasible to derive
By considering
and
with a long-tailed behavior, i.e., in the limit of
and
, we have
with
, we obtain
with a direct connection with Equation (
42) for
and
. This feature implies that the fractional derivative in time and space appear connected with the behavior of the particles in the bulk, i.e., the waiting-time and jumping distributions, which determine the system’s behavior. Notice that Equation (
46) is formally equivalent to the generalized master equation:
with
Let us assume
to work out the fractional diffusion in space and time with
where
is a specified function. In Equation (
42), the fractional operator with respect to the spatial variable is known as the Riesz–Feller operator [
40], which is defined here such that its Fourier transform is straightforward:
where
represents the Fourier transform of
. For
, the initial condition on the first derivative must also be taken into account:
The form of the solution to Equation (
42) will be
where
represents the Green’s function of the given problem, meaning it solves Equation (
42) when
. By utilizing the Fourier transform on the Riesz–Feller operator and the Laplace transform on the Caputo time-fractional derivative, Equation (
42) can be reformulated as
where
represents the Green’s function of the problem in the Fourier–Laplace domain, i.e.,
=
and
. We need to invert both transforms to find the desired solution
. This is not straightforward, but it can be achieved using the Mellin transform along with some scaling properties of the Fourier and Laplace transforms (refer to refs. [
19,
46] for more details). The fundamental solution of the problem, specifically for
, is given by
where the function
is the Fourier transform of the Mittag–Leffler function:
Notably, an important result can be derived without the exact form of
. By setting
, we can construct the quantity
where
Based on Equation (
59), with the assumption that
if
exists, it can be generally inferred that subdiffusion occurs when
because
; superdiffusion occurs when
, i.e.,
; and normal diffusion occurs when
, i.e.,
. A similar analysis for a broad range of diffusion-like equations can be found in ref. [
49].
In what follows, we consider some particular cases. For
,
, the solution becomes
where
is the Lévy distribution given in
For
and
, the solution is
i.e., it is expressed in terms of the
-Wright function,
, often called the Mainardi function, which is significant in the analysis of stochastic processes [
50].
Furthermore, for
and
, it yields the classical diffusion equation, and the solution to the Cauchy problem is given by
which is the expected result for normal diffusion.
In the broader context, it is necessary to address the space-time fractional diffusion equation formulated as [
46,
51,
52]
with the real parameters constrained by the following conditions:
Here,
denotes the most general Riesz–Feller fractional derivative such that for
, it reduces to the case discussed in Equation (
42). The Green’s function can be represented using the
-function of Fox, beginning with its general Mellin–Barnes integral form for
[
51]:
Table 1 shows some cases of the aforementioned solution.
An approach similar to the one outlined above can be modified to address time-fractional diffusion equations of distributed order. Here, the fundamental solution can be represented as an integral involving a Mellin–Barnes integral representation with two parameters that are functionals of the order density,
(refer to ref. [
46] for more details). In the following sections, we apply the formalism presented above to two general challenging physical problems, dealing with adsorption phenomena and impedance spectroscopy response of materials to the external fields.
4. Diffusion and Adsorption Phenomena
Particles diffusing in confined spaces can undergo adsorption (desorption) onto solid substrates. The kinetics of this process are crucial for a range of problems from fundamental science to industrial separation processes [
53,
54,
55,
56,
57,
58,
59]. The most straightforward method to model adsorption–desorption phenomena is the Langmuir adsorption model [
60], commonly known as Langmuir kinetics [
61,
62]. Regarding particles in the bulk, their diffusion process is typically described by Fick’s law [
63], as adsorption–desorption is expected to happen in practical scenarios involving transport [
64,
65]. The adsorption behavior at this boundary is governed by a phenomenological kinetic equation that includes appropriate adsorption–desorption rates [
66].
Adsorption–desorption phenomena are particularly important when dealing with complex fluids. Because liquids generally need some sort of confinement, the solid substrate often plays an important role because of adsorption. This process may be due to the adsorption of molecules in the fluid itself, of added dopants, or of any impurities from molecular dissociation or the synthesis process, since removing all impurities in a liquid sample is almost impossible [
67]. For example, polymers in solution are often adsorbed by confining solid surfaces, having tremendous consequences that are important in pharmaceutical applications [
57], biophysics [
58], and nanocomposite materials [
68]. Liquid crystals, known for their high sensitivity to surface modifications, have demonstrated significant effects due to adsorption phenomena. These include alterations in anchoring energy [
55], surface transitions triggered by light-controlled adsorbed dyes [
69], and degradation of display performance linked to ion adsorption [
56]. In fact, liquid crystals, one of the best representative materials of complex fluids, are mainly applied in electro-optical devices such as displays. In most devices, the liquid crystal material is in direct contact with an electrode (which may be coated with an alignment layer), and thus, the confining walls present adsorption–desorption phenomena. As will be discussed later, the dielectric response of many liquid crystal cells requires the use of constant phase elements (CPEs) when the data are analyzed by equivalent circuits, and the need for CPEs is directly related to interfacial heterogeneity [
70], which in turn may be associated with the fractal nature of the electrodes. We thus now describe how fractional calculus can be introduced to formulate anomalous processes in the adsorption–desorption process, aiming to model the phenomena in heterogeneous interfaces with complex fluid materials.
4.1. Fractional Kinetic Equation
Within the Langmuir framework for adsorption–desorption processes occurring at the boundary between two distinct media (consider, for instance, an isotropic fluid with suspended particles interacting with a solid surface), the kinetic equation states that time variation in the density of adsorbed particles,
, is influenced by the bulk density of particles in the vicinity of the surface,
. Additionally, it is affected by the surface density of particles already adsorbed. We can thus use the following straightforward equation:
where
denotes the adsorption rate constant and
signifies the characteristic desorption time, both of which are phenomenological parameters.
The solution of Equation (
68) in the Laplace domain is given by
where
and
, assuming
. Equation (
69) provides the density of adsorbed particles. The time-domain solution can be found via the inverse Laplace transform of Equation (
69) or by directly integrating Equation (
68).
By performing integration by parts, Equation (
70) can be reformulated as
highlighting that two terms contribute to the effective density of particles near the surface. The first term represents particle adsorption in the first surface layer; the second term describes desorption with exponential relaxation times. Note that the second term is essentially the Caputo–Fabrizio fractional time derivative [
71], which has a nonsingular kernel.
An alternative method to tackle the intricacy of these relaxation phenomena, which considers potential frequency-dependent changes in effective adsorption parameters, involves extending the kinetic equation through fractional calculus [
66]. This approach assumes the adsorption–desorption process is governed by a generalized kinetic equation formulated using a fractional derivative of order
, as outlined below:
where
represents an intrinsic timescale associated with internal dynamics. In the case where
, Equation (
72) clearly simplifies to Equation (
68).
As mentioned in
Section 3, the generalized kinetic equation (Equation (
72)) can be solved using either the Fox
-function or the two-parameter Mittag–Leffler function [
72,
73]. For demonstration purposes, we rewrite Equation (
72) with
as follows:
It is noted that Saxena et al. [
72,
73] examine an alternative fractional
integral kinetic equation of order
:
with
,
, and
being an integrable function over the interval
, where
. In Equation (
74), the Riemann–Liouville fractional integral of order
u is specified as in Equation (
9), specifically,
Equation (
74) has as solution:
specifically, it is expressed using the H-function of Fox [
19]. For the scenario we are examining,
. Consequently, the integral
remains finite if the initial conditions ensure that
. Consequently, the solution to the fractional
differential kinetic Equation (
72) can be derived from Equation (
76), specifically
Moreover, this specific Fox H-function can be related to the two- or three-parameter Mittag–Leffler function, thereby providing a formal solution to the problem. It is important to mention that other different fractional operators may be used to extend the kinetic equation (or diffusion equation), such as the ones characterized by nonsingular kernels [
74], implying different relaxation processes. One of them is the Atangana–Baleanu fractional operator given by
where
is a normalization constant. Equation (
72) can be written, in terms of the Atangana–Baleanu fractional time operator, as follows:
The solution to this equation can be determined and is expressed as
with
,
, and
. It is worth mentioning that the relaxation process present in Equation (
81) is different from the previous one as a consequence of the kernel expressed in terms of a Mittag–Leffler function.
Figure 2 illustrates the adsorption process for the different kinetic equations considered above. It shows that the differential operator is important in the relaxation process.
Thus, we observe that different fractional operators can extend the standard approach and be powerful tools for capturing experimental behavior when the standard formalism is unsuitable. In the next sections, we consider the Riemann–Liouville and Caputo fractional operators, which can capture the experimental behavior of the scenarios discussed here.
4.2. The Poisson–Nernst–Planck Model
We provide a concise review of the main equations of the Poisson–Nernst–Planck (PNP) model in three dimensions as a preliminary step towards exploring its fractional extension [
75]. We examine a system consisting of neutral particles that can, through dissociation, become positive and negative charges with varying mobilities [
76,
77]. When an external electric field
is applied, these charges move, leading to currents of neutral, positive, and negative charges within the sample. In three dimensions, the bulk density of these particles will be denoted by
,
, and
and, likewise, the current density as
,
, and
. The following continuity equations may assure the conservation of particles:
where
is connected to the processes (dissociation and recombination) in bulk. For total dissociation, neutral particle terms vanish. The Poisson equation completes the PNP model.
where
denotes the dielectric permittivity of the ion-free insulating medium, assumed to be nondispersive over the relevant frequency spectrum. This equation connects the bulk densities of positive and negative ions, with an absolute charge of
q, to the resulting electric field distribution within the sample.
The total electric current is composed of the conduction and the displacement currents as follows:
If we now combine the last two formulae of Equation (
82) with Equation (
84), we obtain
and accounting for the Poisson Equation (
83), we conclude that
. Thus, the model requires that the total current has to be solenoidal, i.e.,
In a one-dimensional scenario, the current density
J does not vary with position. This requirement is essential for the concept of electrical impedance, which is the ratio of the voltage difference to the overall current, to be valid [
78].
4.3. Bulk Diffusion and Surface Adsorption
Before moving on to the fractional generalization of the PNP model discussed here, let us consider the question of determining the equilibrium distribution of charges and neutral particles in an insulating medium containing ionic impurities that are diffusing in a bulk in contact with adsorbing surfaces. We consider a system confined within a slab of thickness
d, with adsorbing surfaces (which are electrodes) located at
. Here,
x is the coordinate perpendicular to the surfaces in a Cartesian reference frame, making the mathematical problem one-dimensional. The system consists of neutral and charged particles, and the generation–recombination process is taken into account. The goal is to incorporate a general kinetic equation similar to Equation (
68), which includes a time-fractional derivative and a memory kernel for the adsorption–desorption phenomena [
79].
Thus, we rewrite the continuity Equation (
82) as
in which the bulk current densities of particles are given by
and the source term reads
In Equations (
87) and (
88),
is the electrical potential profile across the sample,
is the thermal energy, and
,
, and
are the diffusion coefficients for positive, negative, and neutral particles, respectively. In addition, the coefficients
and
are the constants of association and dissociation, respectively, for a first-order chemical reaction assumed in the bulk.
The next step to construct the continuous PNP model is to impose the Poisson equation relating the net bulk density of charge to the actual electric potential, which may be written as
If we now combine Equations (
87) and (
88), we obtain the following diffusion equations:
The equations of the model are thus Equations (
90) and (
91), which have to be solved for chosen boundary conditions. To proceed, we assume that the limiting surfaces are blocking for neutral particles, i.e.,
and adsorbing for charged particles:
Finally, the electric potential has to be such that, at the boundaries,
where
is the amplitude of the potential applied to the system by means of an external power supply. The mathematical problem is thus well posed in the sense that the number of coupled equations to be solved are the four partial differential Equations (
90) and (
91), necessitating eight integration constants, which are to be found using the eight boundary conditions given by Equations (
92)–(
94).
Before proceeding, let us focus our attention, for a moment, on a possible meaning of Equation (
93). In the present approach, we assume that these equations are stated for the first and last bulk layers of an insulating liquid sample containing ions. They are stated for the last bulk layer of the sample at
, i.e., on the left of the right surface (placed at
), and are also stated on the first layer of the bulk of the sample at
, i.e., on the right of the left surface (placed at
). These are the bulk positions at which the rates of change in the density of charges in the surface layer (first and last) are represented by
and are equal to the bulk density of current arriving at the same layers of the liquid sample.
When solid surfaces (electrodes or solid phase) are put in contact with the liquid sample (liquid phase) at these points, we have in practice a contact between two different systems, forming a liquid–solid interface. Eventually, due to electrochemical forces, the charged particles present in the bulk can be adsorbed and, after a while, desorbed
by the solid surfaces. This usually happens because the liquid phase delivers charges to the solid phase at a certain rate, which combines the particles arriving from the bulk, as a result of the diffusion process, and the particles coming back to that position, as a result of the desorption process occurring at the surface. This process yields a net density
(
and −) of charges per unit surface which become again available to the solid surface. Once these charges are supplied again to the solid surface, usually they are assumed to undergo a dynamical adsorption–desorption process governed by an equation of the type of Equation (
68).
Here, to implement a possible extension of the Langmuir approximation to the field of fractional calculus, we assume that the adsorption–desorption phenomenon occurring for charged particles, which have been delivered to the surface by the bulk and received back from the surface, is governed by the following general kinetic equation:
To generalize further the approach, a memory kernel
has been inserted in the desorption term to account in a phenomenological way for the presence of roughness on the surface and its effect on the process. The roughness on the surface (electrodes in the case of impedance spectroscopy focused here) plays an important role in determining the anomalous character of the impedance [
80]. Indeed, this effect may lead to a nonusual, i.e., non-Debye, relaxation and, consistently, to an anomalous diffusion behavior.
From the perspective we adopt here, we solve Equations (
90) and (
91) to obtain
,
, and
. Then, we use
to obtain
to be inserted into Equation (
95), accordingly. Once this is performed, Equation (
95) is solved to obtain the profiles of
. Finally, we impose the boundary conditions, Equation (
93), together with the other ones, Equations (
92) and (
95), to fix the values of the integration constants. At this point, the problem is again formally solved, and an analysis of the electrical response of the material to an external difference of potential, in the presence of the adsorption–desorption phenomenon, can be carried out.
Proceeding in this way is equivalent to assuming that the sources (the reservoir) of the charged particles, which are supplied to the surface, are represented by the densities
, which are bulk quantities (liquid phase), if only adsorption is concerned. However, when desorption from the surface is also invoked, a net density of charged particles is present in the last (or the first) layer of the bulk. These net quantities are in turn supplied to the surface (solid phase) at the rate
, and once they are adsorbed on the surface, these charged particles will be subjected to the dynamics imposed by that specific surface on the particles, which is now governed by Equation (
95). Thus, imposing the boundary conditions Equation (
93) is a way to put the liquid phase (the bulk) and solid phase (the surface) in physical contact. The charged particles of the system will follow dynamical processes that are different when they are in the bulk (drift diffusion) or when they are on the surface (fractional adsorption kinetics in the present case). In this way as well, the extension of the kinetic equation to the fractional calculus does not require any change in the continuity equations that guarantee the conservation of the number of charges in the whole system. The details of a partial analysis of this complex phenomenon in the linear approximation, i.e., in the small AC voltage limit, can be found in Ref. [
79].