Skip to main content

PERSPECTIVE article

Front. Appl. Math. Stat., 15 June 2023
Sec. Dynamical Systems
Volume 9 - 2023 | https://doi.org/10.3389/fams.2023.1201043

Invariant forms and control dimensional parameters in complexity quantification

  • Department of Mathematics and Statistics, The University of Western Australia, Perth, WA, Australia

Non-equilibrium dynamics is omnipresent in nature and technology and can exhibit symmetries and order. In idealistic systems this universality is well-captured by traditional models of dynamical systems. Realistic processes are often more complex. This work considers two paradigmatic complexities—canonical Kolmogorov turbulence and interfacial Rayleigh-Taylor mixing. We employ symmetries and invariant forms to assess very different properties and characteristics of these processes. We inter-link, for the first time, to our knowledge, the scaling laws and spectral shapes of Kolmogorov turbulence and Rayleigh-Taylor mixing. We reveal the decisive role of the control dimensional parameters in their respective dynamics. We find that the invariant forms and the control parameters provide the key insights into the attributes of the non-equilibrium dynamics, thus expanding the range of applicability of dynamical systems well-beyond traditional frameworks.

1. Introduction

Non-equilibrium dynamics governs a broad range of processes in nature and technology and is a challenge to study in theory, experiments and simulations. An important aspect advanced our understanding of this complexity is symmetries of the dynamics. For instance, systems with pattern formations—a subject of active research in the field of dynamical systems—are well-described by universal theoretical models, such as the complex Ginzburg-Landau equation and the non-linear Schrödinger equation [18].

Realistic processes are often more complicated than idealistic systems studied within the traditional framework. Yet, they are observed to exhibit symmetries, universality and order. Their non-equilibrium dynamics is eligible to the first principle consideration, and can be investigated on the basis of group theory and representation theory. A critical aspect is the link of the theoretical attributes of the non-equilibrium dynamics—the symmetry groups and the invariant forms—to the dimensional parameters that can control the physical process, and to the observable quantities that can be diagnosed in experiments [1, 817].

In this work we consider two paradigmatic complexities—the classical fluid dynamics problems of canonical Kolmogorov turbulence and Rayleigh-Taylor interfacial mixing. These non-equilibrium processes have very different physical properties, symmetries and characteristics. We employ the invariant forms of these processes to inter-link their scaling laws and spectral shapes and to reveal the role of the control dimensional parameters in their respective dynamics [812, 1524].

Turbulence and Rayleigh-Taylor mixing are inherent to a broad range of phenomena having considerable scientific and technological importance. Examples include supernovae, solar flares, climate change, plasma fusion, nanofabrication, and purification of water [2539]. Turbulence is a state of a dissipative system and it decays unless it is driven by an external energy source. Canonical turbulence is self-similar, isotropic and homogeneous, with a non-dissipative energy transport between the scales. It is a stochastic process with strong fluctuations that may fully blackout deterministic conditions. For as much as turbulence is considered to be the last unsolved problem of the classical physics, Rayleigh-Taylor mixing is its more complex counterpart [812, 1719, 4047].

Rayleigh-Taylor instability develops at the interface between two fluids of different densities accelerated against their density gradient, and it is driven by the acceleration. The amplitude of the interface perturbation grows quickly, and the interface is transformed to a composition of small-scale shear driven vortical structures and a large-scale coherent structure. The scale interaction enhances with time, and the flow transitions to the final stage of intensive interfacial mixing of the fluids. Rayleigh-Taylor mixing is self-similar, like Kolmogorov turbulence, and it is anisotropic, heterogeneous, and sensitive to deterministic conditions, contrary to canonical turbulence [812, 2025, 4858].

Turbulence and Rayleigh-Taylor mixing are a subject of active research in contemporary science, mathematics and engineering. In-depth understanding of their fundamental properties is achieved over the recent decades. The following aspects are certain now: Turbulence is a super-diffusive stochastic process challenging to implement in practice. Realistic turbulent processes often exhibit anomalous scaling. Properties of self-similar interfacial Rayleigh-Taylor mixing depart from those of canonical turbulence, including scaling laws, spectral shapes, and sensitivity to deterministic conditions [812, 4258].

According to the classical approaches, in canonical turbulence, the velocity scales with length as a power-law with an exponent (1/3) and the wave-vector spectrum has the scaling exponent (−5/3). The group theory approach finds that in Rayleigh-Taylor mixing with constant acceleration the velocity scales with length as a power-law with an exponent (1/2) and the wave-vector spectrum has the scaling exponent (−2). When compared to canonical turbulence, Rayleigh-Taylor mixing has stronger correlations and steeper spectra, and can keep order and sense deterministic conditions even at high Reynolds numbers. The group theory results are consistent with, and explain, experiments on Rayleigh-Taylor mixing in fluids and plasmas. The order in Rayleigh-Taylor mixing is similar in spirit to laminarization of strongly accelerated turbulent flows, including flows in boundary layers and curved pipes [812, 1719, 23, 4860].

The canonical approaches for Kolmogorov turbulence and the group theory approach for Rayleigh-Taylor mixing are both based on the analysis of symmetries of these processes, including scaling transformations. Questions thus appear: (1) What is the influence of scaling symmetries and invariant forms on theoretical attributes of the non-equilibrium dynamics? (2) Can the properties of very different processes—canonical turbulence and Rayleigh-Taylor mixing—be linked to one another? (3) What is the role of the control dimensional parameters in their respective dynamics? [812, 1519, 23, 48].

The three questions motivate and frame our investigation. We handle mathematical challenges of Kolmogorov turbulence and Rayleigh-Taylor mixing by employing elegant physical concepts. We reveal that these paradigmatic complexities have lucid theoretical representations. We capture the decisive role of the control dimensional parameters in their non-equilibrium dynamics. Our results chart perspectives for future research and expand the range of non-equilibrium processes accessible for analysis, including group theory, representation theory and dynamical systems methodologies.

2. Conservation laws, symmetries and invariant forms

2.1. Governing equations

As in any physical process [17], a dynamics of an ideal fluid is governed by the conservation of mass, momentum and energy represented in continuous approximation in an inertial frame of reference as

ρt + ρvixi=0,ρvit + ρvivjxj + Pxi = 0,                       Et + (E + P)vixi =0    (1)

Here the spatial coordinates and time are (xi, t) = (x, y, z, t); the fields of density, velocity, pressure and energy density are (ρ, v, P, E), with E = ρ(e + v2/2) and the specific internal energy e. The closure equation of state related the internal energy and pressure, with constant (Pe). In the presence of kinematic viscosity ν the momentum equation is augmented with the term (-ρν2vi/xj2) and the energy equation is also modified [1012, 17].

For canonical Kolmogorov turbulence, the density field is uniform, ρ = ρ0, the dynamics is the density independent, and the process is driven by an external source supplying energy at a constant rate per unit mass E0. This specific power E0, with the dimension m2s-3, is the control parameter of the self-similar, isotropic and homogeneous turbulence [1719, 4046].

For Rayleigh-Taylor dynamics, the equations in the bulk are augmented with the boundary conditions at the interface and at the outside boundaries, so that the normal (tangential) component of velocity and pressure (enthalpy) are continuous (discontinuous) at the interface, and there are no external sources.

ρ(v · n +θ/|θ|) = 0, [v · n] = 0, [P] = 0,        [· τ] = any,[W] = any, v|z± = 0    (2)

Here the jump of a quantity at the interface is […]; the normal and tangential unit vectors of the interface are n, τ with n = ∇θ/|∇θ|, (n·τ) = 0, and the function θ(x, y, z, t) is θ = 0 at the interface and is θ > 0 (< 0) in the light (heavy) fluid sub-domain. The initial conditions prescribe the perturbations of the interface and the flow fields at some instance of time. The dynamics is specific and is driven by balance per unit mass, as follows from the independence of the boundary condition for the normal velocity from the fluid density.

Rayleigh-Taylor dynamics is driven by the acceleration g, g = (0, 0, −g), g = |g|. It is due to a body force, is directed from the heavy to the light fluid, and modifies the pressure field as PP − ρgz. For constant acceleration, g = g0, in the mixing regime the length scale in the acceleration direction (i.e., the amplitude of the interface perturbation) increases quadratic with time h~ g0t2. The acceleration strength g0, with the dimension ms-2, is the control parameter of the self-similar, anisotropic and heterogeneous Rayleigh-Taylor mixing [812, 23, 25, 6163].

2.2. Symmetries and invariant forms

Symmetries of isotropic homogeneous turbulence include Galilean transformations, translations in space and time and spatial rotations and reflections. Self-similar canonical turbulence is also invariant under the scaling transformation of the length, LLK, velocity vvKn, and time, ttK1−n, where n is an exponent and K > 0 is a constant. In the governing equations in the limit of vanishing viscosity, ν/vL → 0, conditional on ν → νK1+n, the exponent of the scaling transformation is n = 1/3. Its invariant form is the rate of energy dissipation, ε=ν(vi/xj),2, with ε~v3/L and ε → εK3n−1. The energy dissipation rate and the energy power are similar quantities, ε ~ ϵ0 [1719, 4047]:

n=13,ε ~ v3L,ε ~ ϵ0    (3)

Symmetries of non-inertial RT mixing include translations, rotations and reflections in the plane normal to the acceleration. Self-similar Rayleigh-Taylor mixing is also invariant with respect to the scaling transformation, LLK, vvKn, , conditional on g0g0K2n-1. In the governing equations in the limit of vanishing viscosity, ν/vL → 0, with ν → νK1+n, the exponent of the scaling transformation is n = 1/2. Its invariant form is (the component of) the rate of loss of specific momentum μ in the acceleration direction, with μ ~ v2/L and μ → μK2n−1, where the vector of the rate of momentum loss is μi=ν(2vi/xj2). The rate of momentum loss and the acceleration strength are similar quantities, μ ~ g0. In RT mixing the rate of energy dissipation is scale-dependent, with ε~ μ2t~ g02t at time t and ε~ μ3/2L1/2~ g03/2L1/2 at length L [812, 48, 63]:

n=12,μ ~  v2L,μ ~  g0    (4)

Distinct symmetries and invariant forms lead to substantial departures of properties of self-similar Rayleigh-Taylor mixing from those of Kolmogorov turbulence, including their scaling laws, spectral shapes, correlations and fluctuations [812, 48, 63].

The properties of Kolmogorov turbulence and Rayleigh-Taylor mixing are identified by the classical approaches and by the group theory approach, respectively, through analyzing symmetries and scaling transformations. We need to clarify whether the group theory approach and the classical approaches are consistent with one another, whether the distinctions in properties of Kolmogorov turbulence and Rayleigh-Taylor mixing are fully captured by their invariant forms, and whether the characteristics of these processes depend on their control dimensional parameters [812, 15, 1719, 48].

3. Interlink of Kolmogorov turbulence and Rayleigh-Taylor mixing

This section directly links the properties of Rayleigh-Taylor mixing and canonical turbulence, demonstrates the full consistency of their theoretical descriptions, and reveals the prominent role of the control dimensional parameters in physics of these processes.

3.1. Velocity scaling

Consider the velocity scaling law, with v (vl) being the velocity scale at the length scale L (l).

In canonical turbulence, the energy dissipation rate is ε ~ v3/L with υ ~ (εL)1/3 at the length scale L, and it is εl~ vl3/l with vl ~ (εll)1/3 at the length scale l. The invariance of the rate of energy dissipation, ε = εl with ε ~ ϵ0, leads to the velocity scaling law vl3/v3~ l/L [1719, 41, 46, 47].

In Rayleigh-Taylor mixing, the rate of energy dissipation is scale-dependent, with ε ~ v3/L ~ μ3/2L1/2 and with εl~ vl3/l~ μl3/2l1/2, where the rate of momentum loss is μ ~ v2/L at the length scales L and it is μl~ vl2/l at the length scale l. The rate of momentum loss is an invariant quantity, μ = μl with μ = g0, leading to the velocity scaling law vl2/v2~ (μl/μ)(l/L)~ l/L [1012, 48].

We directly link the velocity scaling laws in canonical turbulence and Rayleigh-Taylor mixing as:

v ~  (εL)1/3 ~  (μ3/2L1/2L)1/3 ~  (μL)1/2 ~  (g0L)1/2vl ~  (εll)1/3 ~  (μl3/2l1/2l)1/3 ~  (μll)1/2 ~  (g0l)1/2vlv ~  (εll)1/3(εL)1/3 ~  (μll)1/2(μL)1/2 ~  (g0l)1/2(g0L)1/2vlv ~  (lL)1/2    (5)

This reveals that the velocity scaling laws in Rayleigh-Taylor mixing and in Kolmogorov turbulence are consistent with each other. Due to their distinct invariant forms—μ and ε – the velocity correlations are stronger in Rayleigh-Taylor mixing than in Kolmogorov turbulence [1012, 23, 48].

3.2. Reynolds number scaling and viscous scale

Consider the Reynolds number scaling and the viscous scale [1719].

The Reynolds number is Re = vL/ν at the length scale L, and the Reynolds number is Rel = vll/ν at the length scale l. Since v ~ (εL)1/3 and vl~ (εll)1/3, we obtain Re ~ ε1/3L4/3/ν and Rel~ εl1/3l4/3/ν. For the viscous length scale l ~ lν, the local Reynolds number is Rel ~ 1 [1719, 47].

In canonical turbulence, the invariance of the energy dissipation rate, ε = εl with ε ~ ϵ0, leads to the scaling law for the Reynolds number Rel /Re~ (l/L)4/3 and determines the viscous scale lν~ (ν3/εl)1/4~ (ν3/ε)1/4~ (ν3/ϵ0)1/4 [1719, 47].

In Rayleigh-Taylor mixing, with account for the scale-dependence of the energy dissipation rates ε ~ μ3/2L1/2 and εl~ μl3/2l1/2, we obtain Re ~ μ1/2L3/2/v and Rel~ μl1/2l3/2/ν. The invariance of the rate of momentum loss, μ = μl with μ ~ g0, leads to the Reynolds number scaling Rel /Re~ (l/L)3/2 and the viscous scale lν~ (ν2/μl)1/3~ (ν2/μ)1/3~ (ν2/g0)1/3 [1012, 23, 48].

We directly link these quantities in canonical turbulence and in Rayleigh-Taylor mixing as:

Re ~  ε1/3L4/3ν ~  (μ3/2L1/2)1/3L4/3ν ~  μ1/2L3/2ν ~  g01/2L3/2νRel ~  εl1/3l4/3ν ~  (μl3/2l1/2)1/3l4/3ν ~  μl1/2l3/2ν ~  g01/2l3/2νRelRe ~  (lL)3/2,lν ~  (ν2μl)1/3 ~  (ν2μ)1/3 ~  (ν2g0)1/3    (6)

This reveals that the Reynolds number scaling and the viscous scale in Rayleigh-Taylor mixing are consistent with those in Kolmogorov turbulence. Due to their distinct invariant forms—μ and ε , respectively—the Reynolds number scaling is steeper in Rayleigh-Taylor mixing than in canonical turbulence, whereas the viscous scale is set by the acceleration g0 in Rayleigh-Taylor mixing and by the energy power ϵ0 in turbulence [1012, 23, 48].

3.3. Spectral shapes for velocity fluctuations

Consider the spectral shape for fluctuations of the velocity (the specific kinetic energy) [1719].

In canonical turbulence the spectral density of the velocity fluctuations is E(k). It is defined by the invariance of the energy dissipation rate ε and its independence of the wavevector k, leading to the exponent −5/3 of the k spectrum, with E(k)~ ε2/3k-5/3~ ϵ02/3k-5/3 [1719, 47].

In RT mixing, the energy dissipation rate depends on the wavevector, ε~ μ3/2k−1/2. We obtain:

E(k) ~  ε2/3k-5/3 ~  (μ3/2k-1/2)2/3k-5/3 ~  μk-2                     E(k) ~  μk-2 ~  g0k-2    (7)

This demonstrates that the spectral shapes in Rayleigh-Taylor mixing and in Kolmogorov turbulence are consistent with one another. In Rayleigh-Taylor mixing the velocity fluctuations spectra are steeper than in canonical turbulence, due to the distinct invariant forms of these processes, μ and ε respectively [1012, 23, 48].

In two-dimensional isotropic homogeneous turbulence, with account for invariance properties of the enstrophy Ω , with the dimension s-2, and the rate of enstrophy Ω~, with the dimension s-3, the spectral density for the kinetic energy fluctuations v2 has the form E(k)~ Ω~2/3k-3 [64, 65]. In Rayleigh-Taylor mixing the enstrophy Ω and the rate of enstrophy Ω~ depend on the wavevector as Ω~ μk and Ω~~ (μk)3/2. We derive

E(k)  ~  Ω~2/3k-3  ~  ((μk)3/2)2/3k-3  ~  μk-2                 E(k)  ~  μk-2  ~  g0k-2    (8)

and find that the spectral shapes in Rayleigh-Taylor mixing and in two dimensional turbulence are consistent with one another. In Rayleigh-Taylor mixing the velocity fluctuations spectra are more gradual than in two-dimensional turbulence, due to distinct invariant forms of these processes, μ and Ω~, respectively [1012, 23, 48, 6466].

3.4. Spectral shapes for density fluctuations

Consider the spectral shape for the density fluctuations [10, 17, 53].

In canonical turbulence, the spectral shape of the density fluctuations is E(k)~ ρ0ε0k-1, since the energy dissipation rate ε is independent of the fluid density ρ0. In Rayleigh-Taylor mixing, the independence of the rate of momentum loss on the fluid density ρ0 leads to the spectral shape E(k)~ ρ0μ0k-1 [10, 53].

We obtain

E(k) ~  ρ0ε0k-1 ~  ρ0(μ3/2k-1/2)0k-1 ~  ρ0μ0k-1              E(k) ~  ρ0μ0k-1 ~  ρ0g00k-1    (9)

For the density fluctuations, the exponent −1 of the k spectrum is the same in the anisotropic and heterogeneous Rayleigh-Taylor mixing and in the isotropic and homogeneous Kolmogorov turbulence. In either case the dynamics is specific and is balanced per unit mass (rather than per unit volume), as displayed in the independence of the invariant forms of these processes—μ or ε —on the fluid density ρ0 [10, 53].

3.5. Link to other modeling approaches

We further illustrate in step-by-step derivations that our results on Rayleigh-Taylor mixing are consistent with other models of realistic turbulent processes and with some empirical models [53, 6775].

In modeling realistic turbulent processes, the spectral density E(k) is often related to the energy dissipation rate ε , the wavevector k, and the process time scale τ as ε ~ τk4E2 [73].

In canonical turbulence the time-scale is τ ~ (k3E)−1/2, leading to ε ~ (k3E)−1/2k4E2 ~ k5/2E3/2 and, due to the invariance of the energy dissipation rate ε , to the spectral density E(k) ~ ε2/3k−5/3. In RT mixing, the time-scale is τ~ (g0k)-1/2, leading to ε~ (g0k)-1/2k4E2~ g0-1/2k7/2E2 and the spectral density E~ ε1/2g01/4k-7/4. By further accounting for the scale-dependence of the energy dissipation rate ε~ g03/2k-1/2~ μ 3/2k-1/2, we obtain the spectral density in Rayleigh-Taylor mixing with constant acceleration as E~ g0k-2~ μk-2, similarly to the foregoing [73, 74]:

ε ~  τ k4E2,τ ~  (g0k)-1/2,ε ~  g03/2k-1/2 ~  μ 3/2k-1/2                                E ~  g0k-2 ~  μk-2    (10)

By considering the Rayleigh-Taylor time scale τ~ (g0k)-1/2 and by formally treating the energy dissipation rate ε , the empirical model [74] identifies the spectral density as E ~ k−7/4. We derive this result from the spectral density E ~ μk−2 in Rayleigh-Taylor mixing, by accounting for the scale dependence of the energy dissipation rate ε ~ μ3/2k−1/2, the invariant form of the rate of momentum loss μ ~ g0 and the time-scale τ~ (g0k)-1/2 as:

E ~  μk-2 ~  g0k-2 ~  (μ 3/2k-1/2)1/2g01/4k-7/4~  (g03/2k-1/2)1/2g01/4k-7/4E ~  ε1/2g01/4k-7/4E ~  ε1/2(g0k)1/4k-2 ~  ε1/2 τ-1/2 k-2 ~  μk-2    (11)

The phenomenological model [75] postulates that in Rayleigh-Taylor mixing the spectral density is the same as in canonical turbulence E ~ k−5/3. We reproduce this prospect from the spectral density defined by the invariant form of Rayleigh-Taylor mixing E ~ μk−2, with μ ~ g0, and with relations of the rates of momentum loss and energy dissipation as μ ~ ε2/3k1/3:

E ~  μk-2 ~  (ε2/3k1/3) k-2 ~  ε2/3k-5/3    (9)

The model further states that in Rayleigh-Taylor mixing the viscous scale vanishes with time [75]. For testing this statement, we consider the local Reynolds number set by the invariant form of Rayleigh-Taylor mixing, Rel~ μl1/2l3/2/ν, with μl ~ μ ~ g0, relate the rates of momentum loss and energy dissipation, μl~ εl2/3l-1/3, and derive:

Rel ~  μl1/2l3/2νl ~  Rel2/3ν2/3μl1/3 ~  Rel2/3ν2/3(εl2/3l-1/3)1/3                                        l ~  Rel3/4ν3/4εl1/4lν ~  ν3/4εl1/4    (12)

The model's result can be further reproduced with a formal replacement εl ~ ε and substitution ε~ g02t.

The results of empirical models of Rayleigh-Taylor mixing can be obtained within group theory approach by formally treating the energy dissipation rate and by masking its scale-dependence [53, 74, 75].

3.6. Velocity structure function

To conclude this section, we consider the velocity structure function, Sn, of the order n, nN, with the dimension (ms-1)n [1719]. It scales with the length scale l as Sn~ (εll)n/3 in Kolmogorov turbulence, and as Sn~ (μll)n/2 in Rayleigh-Taylor mixing [1012, 1719, 23, 48]. Since in Rayleigh-Taylor mixing the energy dissipation rate is scale-dependent, εl~ μl3/2l1/2, we obtain

Sn ~  (εll)n/3 ~  (μl3/2l1/2l)n/3 ~  (μll)n/2                           Sn ~  (μll)n/2    (13)

The structure functions in Rayleigh-Taylor mixing and in Kolmogorov turbulence are consistent with one another and are set by their respective invariant forms. In Rayleigh-Taylor mixing the structure function has a steeper dependence on the order number when compared to Kolmogorov turbulence [1012, 1719, 23, 48].

In isotropic homogeneous turbulence in realistic environments the structure function is known to depart from the Kolmogorov scenario: It exhibits intermittency and multi-fractality mathematically, is influenced by the flow structures physically, and has remarkable statistics [17, 66, 7679]. We believe that the approaches developed for canonical turbulence [66, 76, 78] can be generalized to the case of Rayleigh-Taylor mixing with variable accelerations [811], to be done in the future.

4. Invariant forms and control dimensional parameters

Symmetries and their associated invariant forms are common in physical processes. They relate the process insights to the control dimensional parameter and enable the problem solution. We give some examples to accentuate the role of the control dimensional parameters and the associated invariant forms in understanding complex processes [37, 1012, 1517].

For gravitational process, the invariance of the gravitational constant G with the dimension kg−1m3s−2 is compatible with the Kepler's third law, L3 ~ t2 [17]. In standard diffusion the invariance of diffusion coefficient D with the dimension m2s−1 leads to the diffusion scaling law L ~ t1/2 and the Gaussian distribution [17].

In canonical turbulence, the invariance of the energy dissipation rate ε is associated with the power E0 of the external source supplying energy to the system at a constant rate, both having the dimension m2s−3 [1719]. This leads to the scaling laws for the length L ~ t3/2 and the velocity v ~ t1/2 and displays the stochastic nature of canonical turbulence having normal distribution of velocity fluctuations.

In Rayleigh-Taylor mixing, the invariance of the rate of loss of specific momentum μ is associated with the acceleration strength g0, both having the dimension ms−2. This leads to the scaling laws for the length L ~ t2 and the velocity L ~ t and exhibits the deterministic nature of Rayleigh-Taylor dynamics having ballistic velocity fluctuations [1012].

We thus find that the scaling laws, spectral shapes, properties of correlations and fluctuations in canonical turbulence and in Rayleigh-Taylor mixing are set by their invariant forms and the associated control dimensional parameters, ε ~ ϵ0 and μ ~ g0, respectively [1012, 1719, 47, 48, 53]. These theoretical insights call for experimental investigations. As noted by the 1923 Nobel Laureate Robert A. Millikan, “The fact that Science walks forward on two feet, namely theory and experiment… Sometimes it is one foot which is put forward first, sometimes the other, but continuous progress is only made by the use of both—by theorizing and then testing, or by finding new relations in the process of experimenting and then bringing the theoretical foot up and pushing it on beyond, and so on in unending alternations.”1

5. Discussion

We considered two paradigmatic complexities of non-equilibrium dynamics—canonical turbulence and Rayleigh-Taylor mixing. These processes are a long-standing theoretical challenge, requiring one to solve the system of conservation laws—non-linear partial differential equations, augmented also in the Rayleigh-Taylor case with the singular boundary value problem at the unstable interface and the ill-posed initial value problem. We handle the mathematical challenges on the basis of the physical concept of symmetries and reveal the effect of invariant forms on attributes and characteristics of these processes. We assess that in Rayleigh-Taylor mixing the correlations are stronger, the velocity spectra are steeper and the deterministic conditions are more authoritative, than in Kolmogorov turbulence. For the first time, to our knowledge, we interlink the scaling laws and spectral shapes of these processes, and identify the decisive role of the control dimensional parameters in their non-equilibrium dynamics.

The concept of symmetries advanced our understanding of dynamical systems and enabled development of universal theoretical models of pattern formation in idealistic systems. Our work finds that symmetries and invariant forms encapsulate information on characteristics of non-equilibrium dynamics and are associated with dimensional parameters controlling physical processes. They can provide an important insight on properties of realistic complex systems. This expands the range of applicability of dynamical systems beyond traditional frameworks, and allow us to systematically investigate a broad range of phenomena, having considerable scientific and technological importance, and including supernovae, climate change, plasma fusion, and purification of water [163] (see text footnote 1).

Data availability statement

The original contributions presented in the study are included in the article/supplementary material, further inquiries can be directed to the corresponding author.

Author contributions

SIA contributed to conceptualization, formal analysis, investigation, methodology, resources, and writing—original draft.

Funding

The author thanks the National Science Foundation (USA) (Award No. 1404449) and the Australian Research Council (AUS) (Award No. LE220100132).

Conflict of interest

The author declares that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher's note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

Footnotes

1. ^The Nobel Prize. Available online at: https://www.nobelprize.org/.

References

1. Abarzhi SI, Goddard WA. Interfaces and 3mixing: non-equilibrium transport across the scales. Proc Natl Acad Sci USA. (2019) 116:18171. doi: 10.1073/pnas.1818855116

PubMed Abstract | CrossRef Full Text | Google Scholar

2. Abarzhi SI, Sreenivasan KR. Turbulent mixing and beyond. Introduction. Philos Trans Roy Soc A. (2010) 368:1539. doi: 10.1098/rsta.2010.0021

PubMed Abstract | CrossRef Full Text | Google Scholar

3. Kadanoff LP. Statistical physics: statistics, dynamics and renormalization. World Sci. (2000). doi: 10.1142/4016

CrossRef Full Text | Google Scholar

4. Shubnikov AV, Koptsik VA. Symmetry in Science and Art. Plenum Press (1974).

Google Scholar

5. Cross MC, Hohenberg PC. Pattern formation outside of equilibrium. Rev Mod Phys. (1993) 65:851. doi: 10.1103/RevModPhys.65.851

CrossRef Full Text | Google Scholar

6. Aranson IS, Kramer L. The world of the complex Ginzburg-Landau equation. Rev Mod Phys. (2002) 74:99. doi: 10.1103/RevModPhys.74.99

PubMed Abstract | CrossRef Full Text | Google Scholar

7. Pitaevskii L, Stringari S. Bose-Einstein Condensation. Clarendon: Oxford (2003).

Google Scholar

8. Abarzhi SI, Hill DL, Williams KC Li JT, Remington BA, Arnett WD. Fluid dynamics mathematical aspects of supernova remnants. Phys Fluids. (2023) 35:034106. doi: 10.1063/5.0123930

PubMed Abstract | CrossRef Full Text | Google Scholar

9. Abarzhi SI, Bhowmick AK, Naveh A, Pandian A, Swisher NC, Stellingwerf RF, et al. Supernova, nuclear synthesis, fluid instabilities and mixing. Proc Natl Acad Sci USA. (2019) 116:18184. doi: 10.1073/pnas.1714502115

PubMed Abstract | CrossRef Full Text | Google Scholar

10. Abarzhi SI, Sreenivasan KR. Self-similar Rayleigh-Taylor mixing with accelerations varying in time and space. Proc Natl Acad Sci USA. (2022) 119:e2118589119. doi: 10.1073/pnas.2118589119

PubMed Abstract | CrossRef Full Text | Google Scholar

11. Abarzhi SI. Self-similar interfacial mixing with variable acceleration. Phys Fluids. (2021) 33:122110. doi: 10.1063/5.0064120

CrossRef Full Text | Google Scholar

12. Abarzhi SI. Review of theoretical modeling approaches of Rayleigh-Taylor instabilities and turbulent mixing. Phil Trans R Soc A. (2010) 368:1809. doi: 10.1098/rsta.2010.0020

PubMed Abstract | CrossRef Full Text | Google Scholar

13. Abarzhi SI. Review of nonlinear dynamics of the unstable fluid interface: conservation laws and group theory. Phys Scr. (2008) 2008:014012. doi: 10.1088/0031-8949/2008/T132/014012

CrossRef Full Text | Google Scholar

14. Kargopolov MI, Merzlyakov YI. Fundamentals of Group Theory (in Russian). Moscow: Nauka (1982).

Google Scholar

15. Bridgman PW. Dimensional Analysis. New Haven, CT: Yale University Press (1931).

Google Scholar

16. Migdal AB. Qualitative Methods in Quantum Theory. Benjamin WA, Translation from: Migdal AB 1973 Qualitative methods in quantum theory [Kachestvennye metody v kvantovoi teorii] (In Russian). Moscow: Nauka (1977).

Google Scholar

17. Landau LD, Lifshitz EM. Theory Course I-X. New York, NY: Pergamon Press (1987).

Google Scholar

18. Kolmogorov AN. Local structure of turbulence in an incompressible fluid for very large Reynolds numbers. Dokl Akad Nauk SSSR. (1941) 30:299.

Google Scholar

19. Kolmogorov AN. Energy dissipation in locally isotropic turbulence. Dokl Akad Nauk SSSR. (1941) 32:19.

Google Scholar

20. Rayleigh L. Investigations of the character of the equilibrium of an incompressible heavy fluid of variable density. Proc London Math Soc. (1883) 14:170. doi: 10.1112/plms/s1-14.1.170

CrossRef Full Text | Google Scholar

21. Taylor GI. The instability of liquid surfaces when accelerated in a direction perpendicular to their planes. Proc R Soc London A. (1950) 201:192–6. doi: 10.1098/rspa.1950.0052

CrossRef Full Text | Google Scholar

22. Davies RM, Taylor GI. The mechanics of large bubbles rising through extended liquids and through liquids in tubes. Proc R Soc A. (1950) 200:375. doi: 10.1098/rspa.1950.0023

CrossRef Full Text | Google Scholar

23. Meshkov EE, Abarzhi SI. On Rayleigh-Taylor interfacial mixing. Fluid Dyn Res. (2019) 51:065502. doi: 10.1088/1873-7005/ab3e83

CrossRef Full Text | Google Scholar

24. Meshkov EE. Some peculiar features of hydrodynamic instability development. Phil Trans R Soc A. (2013) 371:20120288. doi: 10.1098/rsta.2012.0288

PubMed Abstract | CrossRef Full Text | Google Scholar

25. Anisimov SI, Drake RP, Gauthier S, Meshkov EE, Abarzhi SI. What is certain and what is not so certain in our knowledge of Rayleigh–Taylor mixing? Phil Trans R Soc A. (2013) 371:20130266. doi: 10.1098/rsta.2013.0266

PubMed Abstract | CrossRef Full Text | Google Scholar

26. Arnett D. Supernovae and Nucleosynthesis. Princeton University Press (1996).

Google Scholar

27. Zeldovich YB, Raizer YP. Physics of Shock Waves and High-Temperature Hydrodynamic Phenomena. New York, NY: Dover (2002).

Google Scholar

28. Drake RP. Perspectives on high-energy-density physics. Phys Plasmas. (2009) 16:055501. doi: 10.1063/1.3078101

CrossRef Full Text | Google Scholar

29. Remington BA, Park H-S, Casey DT, Cavallo RM, Clark DS, Huntington CM, et al. Rayleigh–Taylor instabilities in high-energy density settings on the National Ignition Facility. Proc Natl Acad Sci USA. (2019) 116:18233. doi: 10.1073/pnas.1717236115

PubMed Abstract | CrossRef Full Text | Google Scholar

30. Betti R, Hurricane OA. Inertial-confinement fusion with lasers. Nat Phys. (2006) 12:435. doi: 10.1038/nphys3736

CrossRef Full Text | Google Scholar

31. Popov NA, Shcherbakov VA, Mineev VN, Zaydel' PM, Funtikov AI. Thermonuclear fusion in the explosion of a spherical charge - the problem of a gas-dynamic thermonuclear fusion. Physics. (2008) 51:1047. doi: 10.1070/PU2008v051n10ABEH006688

CrossRef Full Text | Google Scholar

32. Liang Z, Bu W, Schweighofer KJ, Walwark DJ Jr, Harvey JS, Hanlon GR, et al. Nanoscale view of assisted ion transport across the liquid–liquid interface. Proc Natl Acad Sci USA. (2019) 116:18227. doi: 10.1073/pnas.1701389115

PubMed Abstract | CrossRef Full Text | Google Scholar

33. Zhakhovsky VV, Kryukov AP, Levashov VY, Shishkov IN, Anisimov SI. Mass and heat transfer between evaporation and condensation surfaces: Atomistic simulation and solution of Boltzmann kinetic equation. Proc Natl Acad Sci USA. (2019) 116:18209. doi: 10.1073/pnas.1714503115

PubMed Abstract | CrossRef Full Text | Google Scholar

34. Grigoryev SY, Dyachkov SA, Parshikov AN, Zhakhovsky VV. Limited and unlimited spike growth from grooved free surface of shocked solid. J Appl Phy. (2022) 131:065104. doi: 10.1063/5.0078138

CrossRef Full Text | Google Scholar

35. Ilyin DV, Goddard WA III, Oppenheim JJ, Cheng T. First principles–based reaction kinetics from reactive molecular dynamics simulations: application to hydrogen peroxide decomposition. Proc Natl Acad Sci USA. (2019) 116:18202. doi: 10.1073/pnas.1701383115

PubMed Abstract | CrossRef Full Text | Google Scholar

36. Mayer F, Richter S, Westhauser J, Blasco E, Barner-Kowollik C, Wegener M. Multi-material 3D laser micro-printing using an integrated microfluidic system. Sci Adv. (2019) 5:eaau916. doi: 10.1126/sciadv.aau9160

PubMed Abstract | CrossRef Full Text | Google Scholar

37. Underwood TC, Loebner KT, Miller VA, Cappelli MA. Dynamic formation of stable current-driven plasma jets. Sci Rep. (2019) 9:2588. doi: 10.1038/s41598-019-39827-6

PubMed Abstract | CrossRef Full Text | Google Scholar

38. Gorokhovski M, Herrmann M. Modeling primary atomization. Ann Rev Fluid Mech. (2008) 40:343–66. doi: 10.1146/annurev.fluid.40.111406.102200

CrossRef Full Text | Google Scholar

39. Mahalov A. Multiscale modeling and nested simulations of three-dimensional ionospheric plasmas: Rayleigh–Taylor turbulence and nonequilibrium layer dynamics at fine scales. Phys Scripta. (2014) 89:098001. doi: 10.1088/0031-8949/89/9/098001

CrossRef Full Text | Google Scholar

40. Barenblatt GI. Scaling Self-Similarity and Intermediate Asymptotics. Cambridge: Cambridge University Press (1996).

Google Scholar

41. Zakharov VE, L'vov VS, Falkovich G. Kolmogorov Spectra of Turbulence. Springer (1992).

Google Scholar

42. Shraiman BI, Siggia ED. Scalar turbulence. Nature. (2000) 405:639. doi: 10.1038/35015000

PubMed Abstract | CrossRef Full Text | Google Scholar

43. Pouquet A, Mininni PD. The interplay between helicity and rotation in turbulence: implications for scaling laws and small-scale dynamics. Phil Trans R Soc A. (2010) 368:1635. doi: 10.1098/rsta.2009.0284

PubMed Abstract | CrossRef Full Text | Google Scholar

44. Yakhot V, Donzis D. Emergence of multi-scaling in a random-force stirred fluid. Phys Rev Lett. (2017) 119:044501. doi: 10.1103/PhysRevLett.119.044501

PubMed Abstract | CrossRef Full Text | Google Scholar

45. Schumacher J, Sreenivasan KR. Colloquium: unusual dynamics of convection in the Sun. Rev Modern Phys. (2020) 92:041001. doi: 10.1103/RevModPhys.92.041001

CrossRef Full Text | Google Scholar

46. Sreenivasan KR. Turbulent mixing: a perspective. Proc Natl Acad Sci USA. (2019) 116:18175. doi: 10.1073/pnas.1800463115

PubMed Abstract | CrossRef Full Text | Google Scholar

47. Sreenivasan KR. Fluid turbulence. Rev Mod Phys. (1999) 71:S383. doi: 10.1103/RevModPhys.71.S383

CrossRef Full Text | Google Scholar

48. Abarzhi SI. On fundamentals of Rayleigh-Taylor turbulent mixing. Europhys Lett. (2010) 91:12867. doi: 10.1209/0295-5075/91/35001

CrossRef Full Text | Google Scholar

49. Kadau K, Barber JL, Germann TC, Holian BL, Alder BJ. Atomistic methods in fluid simulation. Phil Trans R Soc A. (2010) 368:1547. doi: 10.1098/rsta.2009.0218

PubMed Abstract | CrossRef Full Text | Google Scholar

50. Robey HF, Zhou Y, Buckingham AC, Keiter P, Remington BA, Drake RP. The time scale for the transition to turbulence in a high Reynolds number, accelerated flow. Phys Plasmas. (2003) 10:614. doi: 10.1063/1.1534584

CrossRef Full Text | Google Scholar

51. Pandian A, Stellingwerf RF, Abarzhi SI. Effect of wave interference on nonlinear dynamics of Richtmyer-Meshkov flows. Phys Rev Fluids. (2017) 2:073903. doi: 10.1103/PhysRevFluids.2.073903

CrossRef Full Text | Google Scholar

52. Swisher NC, Kuranz C, Arnettt WD, Hurricane O, Robey H, Remington BA, et al. Rayleigh-Taylor mixing in supernova experiments. Phys Plasmas. (2015) 22:102707. doi: 10.1063/1.4931927

CrossRef Full Text | Google Scholar

53. Williams KC, Abarzhi SI. Fluctuations spectra of specific kinetic energy, density and mass flux in Rayleigh-Taylor mixing. Phys Fluids. (2022) 34:12211. doi: 10.1063/5.0120521

CrossRef Full Text | Google Scholar

54. Meshkov EE, Nikiforov VV, Tolshmyakov AI. On the structure of turbulent mixing zone at the interface between two gases accelerated by shock wave. Combust Explos Shock Waves. (1990) 26:315. doi: 10.1007/BF00751371

CrossRef Full Text | Google Scholar

55. Volchenko OI, Zhidov IG, Meshkov EE, Rogachev VG. Development of localized perturbations at unstable interface of accelerated liquid layer. ZhTF Lett. (1989) 15:47 (in Russian).

Google Scholar

56. Akula B, Suchandra P, Mikhaeil M, Ranjan D. Dynamics of unstably stratified free shear flows: an experimental investigation of coupled Kelvin–Helmholtz and Rayleigh–Taylor instability. J Fluid Mech. (2017) 816:619. doi: 10.1017/jfm.2017.95

CrossRef Full Text | Google Scholar

57. Lugomer, S. Laser generated Richtmyer–Meshkov instability and nonlinear wave paradigm in turbulent mixing. I. Central region of Gaussian spot. Laser Part. Beams 34, 687; 2017. Laser generated Richtmyer–Meshkov instability and nonlinear wave paradigm in turbulent mixing. II. Near-central region of Gaussian spot. Laser Part. Beams 35, 210. Lugomer S. Laser-generated Richtmyer–Meshkov and Rayleigh–Taylor instabilities. III. Near-peripheral region of Gaussian spot. Laser Part Beams. (2016) 35:597. doi: 10.1017/S026303461700009X

CrossRef Full Text | Google Scholar

58. Kuranz CC, Park H-S, Huntington CM, Miles RA, Remington BA, Plewa T, et al. How high energy fluxes may affect Rayleigh–Taylor instability growth in young supernova remnants. Nat Commun. (2018) 9:1564. doi: 10.1038/s41467-018-03548-7

PubMed Abstract | CrossRef Full Text | Google Scholar

59. Narasimha R, Sreenivasan KR. Relaminarization in highly accelerated turbulent boundary layers. J Fluid Mech. (1973) 61:417. doi: 10.1017/S0022112073000790

CrossRef Full Text | Google Scholar

60. Taylor GI. The criterion for turbulence in curved pipes. Proc R Soc A. (1929) 124:243. doi: 10.1098/rspa.1929.0111

CrossRef Full Text | Google Scholar

61. Abarzhi SI, Williams KC. Scale-dependent Rayleigh-Taylor dynamics with variable acceleration by group theory approach. Phys Plasmas. (2020) 27:072107. doi: 10.1063/5.0012035

CrossRef Full Text | Google Scholar

62. Abarzhi SI, Hill DL, Williams KC, Wright CE. Buoyancy and drag in Rayleigh-Taylor and Richtmyer-Meshkov linear, nonlinear and mixing dynamics. Appl Math Lett. (2022) 31:108036. doi: 10.1016/j.aml.2022.108036

CrossRef Full Text | Google Scholar

63. Abarzhi SI, Gorobets A, Sreenivasan KR. Turbulent mixing in immiscible, miscible and stratified media. Phys Fluids. (2005) 17:081705. doi: 10.1063/1.2009027

CrossRef Full Text | Google Scholar

64. Kraichnan RH. Inertial ranges in two-dimensional turbulence. Phys Fluids. (1967) 10:1417–23. doi: 10.1063/1.1762301

CrossRef Full Text | Google Scholar

65. Batchelor GK. Computation of the energy spectrum in homogeneous two-dimensional turbulence. Phys Fluids. (1969) 12: II-233–9. doi: 10.1063/1.1692443

CrossRef Full Text | Google Scholar

66. Frisch U. Turbulence the Legacy of Kolmogorov. Cambridge: Cambridge University Press (1995).

Google Scholar

67. Neuvazhaev VE. Theory of turbulent mixing. Sov Phys Dokl. (1975) 20:398.

Google Scholar

68. Cabot WH, Cook AW. Reynolds number effects on Rayleigh–Taylor instability with possible implications for type Ia supernovae. Nat Phys. (2006) 2:562. doi: 10.1038/nphys361

CrossRef Full Text | Google Scholar

69. Dalziel SB, Linden PF, Youngs DL. Self-similarity and internal structure of turbulence induced by Rayleigh–Taylor instability. J Fluid Mech. (1999) 399:1. doi: 10.1017/S002211209900614X

CrossRef Full Text | Google Scholar

70. Glimm J, Sharp DH, Kaman T, Lim H. New directions for Rayleigh–Taylor mixing. Phil Trans R Soc A. (2013) 371:20120183. doi: 10.1098/rsta.2012.0183

PubMed Abstract | CrossRef Full Text | Google Scholar

71. Youngs DL. The density ratio dependence of self-similar Rayleigh–Taylor mixing. Phil Trans R Soc A. (2013) 371:20120173. doi: 10.1098/rsta.2012.0173

PubMed Abstract | CrossRef Full Text | Google Scholar

72. Schilling O. Self-similar Reynolds-averaged mechanical-calar turbulence models for Rayleigh-Taylor, Richtmyer-Meshkov, and Kelvin-Helmholtz instability-induced mixing in the small Atwood number limit. Phys Fluids. (2021) 33:085129. doi: 10.1063/5.0055193

CrossRef Full Text | Google Scholar

73. Pouquet A, Frsich U, Leorad J. Strong MHD helical turbulence and the nonlinear dynamo effect. J Fluid Mech. (1976) 77:321. doi: 10.1017/S0022112076002140

CrossRef Full Text | Google Scholar

74. Zhou Y. A scaling analysis of turbulent flows driven by Rayleigh-Taylor and Richtmyer-Meshkov instabilities. Physics of Fluids. (2001) 13:538. doi: 10.1063/1.1336151

CrossRef Full Text | Google Scholar

75. Chertkov M. Phenomenology of Rayleigh-Taylor turbulence. Phys Rev Lett. (2003) 91:115001. doi: 10.1103/PhysRevLett.91.115001

PubMed Abstract | CrossRef Full Text | Google Scholar

76. She Z-S, Leveque E. Universal scaling laws in fully developed turbulence. Phys Rev Lett. (1994) 72:336–9. doi: 10.1103/PhysRevLett.72.336

PubMed Abstract | CrossRef Full Text | Google Scholar

77. Birnir B. The Kolmogorov-Obukhov-She-Leveque scaling in turbulence. Commun Pure Appl Anal. (2014) 13:1737–57. doi: 10.3934/cpaa.2014.13.1737

PubMed Abstract | CrossRef Full Text | Google Scholar

78. Sreenivasan KR, Yakhot V. Dynamics of three-dimensional turbulence from Navier-Stokes equations. Phys Rev Fluids. (2021) 6:104604. doi: 10.1103/PhysRevFluids.6.104604

CrossRef Full Text | Google Scholar

79. Hsu A, Kaufman R, Glimm J. Scaling laws for partially developed turbulence. Front Appl Math Stat. (2022) 7:812330. doi: 10.3389/fams.2021.812330

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: fluid instabilities, interfacial mixing, self-similarity, symmetry, invariant forms, dynamical systems

Citation: Abarzhi SI (2023) Invariant forms and control dimensional parameters in complexity quantification. Front. Appl. Math. Stat. 9:1201043. doi: 10.3389/fams.2023.1201043

Received: 13 April 2023; Accepted: 29 May 2023;
Published: 15 June 2023.

Edited by:

Axel Hutt, Inria Nancy - Grand-Est Research Centre, France

Reviewed by:

James Glimm, Stony Brook University, United States

Copyright © 2023 Abarzhi. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Snezhana I. Abarzhi, snezhana.abarzhi@gmail.com

Download