Skip to main content

MINI REVIEW article

Front. Phys., 05 February 2020
Sec. Soft Matter Physics
Volume 7 - 2019 | https://doi.org/10.3389/fphy.2019.00239

A Concise Review of Gradient Models in Mechanics and Physics

  • Laboratory of Mechanics and Materials, School of Engineering, Aristotle University of Thessaloniki, Thessaloniki, Greece

The various mathematical models developed over the years to interpret the behavior of materials and corresponding processes they undergo were based on observations and experiments made at that time. Classical laws for solids (Hooke) and fluids (Navier–Stokes) form the basis of current technology. The discovery of new phenomena with the aid of newly developed experimental probes have led to various modifications of these laws, especially at small scales. The emergence of nanotechnology is ultimately connected with the design of novel tools for observation and measurements, as well as with the development of new methods and approaches for quantification and understanding. This paper first reviews the author's previously developed weakly non-local or gradient models for elasticity, diffusion, and plasticity. It then proposes a similar extension for fluids and electrodynamics. Finally, it suggests a gradient modification of Newton's law of gravity, with a possible connection to the strong force of elementary particle physics.

Introduction

In a recent chapter in Advances of Applied Mechanics [1], an extensive review of the author's internal length gradient (ILG) mechanics framework is presented. It is based on the assignment of internal lengths/ILs (associated with the local geometry/topology of material nano/microstructures) as scalar multipliers of extra Laplacian terms that are introduced into classical constitutive equations to account for heterogeneity effects and weak non-locality. Related background work for this framework can be found in the references quoted therein, as well as in earlier published articles by the author and his coworkers on gradient elasticity/plasticity [2, 3] and higher order diffusion/heat conduction [4, 5].

The motivation for the ILG framework goes back to van der Waals work on liquid/vapor interfaces, as extended by Landau's school through the introduction of a gradient-dependent order parameter and, more recently, adopted in Cahn–Hilliard theory of spinodal decomposition [6, 7]. In this connection, it is also pointed out that the plethora of more recent phenomenological phase field models, as well as subatomic generalized functional density theories, are essentially based on the introduction of higher order gradients in the free energy of the system under consideration.

At very small scales, mechanical and chemical effects are often equipresent, and an extended chemomechanical ILG framework is necessary to consider higher order IL couplings, as suggested in Aifantis [1]. In view of the fact that mechanical and chemical ILs are introduced as scalar multipliers of corresponding Laplacian terms, it turns out that such coupled chemomechanical formulation is appealing and robust. Since in models of mathematical biology [8, 9], cells are commonly represented by scalar concentration fields (i.e., in the same way as chemical species), the formulation could be easily adapted for the description of higher order couplings between mechanical and biochemical ILs. Such an extended ILG mechanics framework, including synergistic effects between mechanical and chemical or biological ILs, can be employed to consider chemomechanical instabilities in lithium-ion battery/LiB anodes [10] and biomechanical instabilities in brain tumors [11]. In this connection, it should be pointed out that the Laplacian of the elastic Hookean stress in the author's model of gradient elasticity/GradEla [3, 5] enters in the same way as in Murray's treatment of morphogenesis in biology [8, 9]. Similarly, the Laplacians of the immotile/motile cancer cell densities in the go or grow/GoG model [11] were introduced in the same way as in the Walgraef-Aifantis/W-A model of immobile/mobile dislocation densities for persistent slip bands [12].

Ilg in Solids: Elasticity/Plasticity/Diffusion and Chemoelasticity

For elastic deformations, the term ε22[λεmmδij+2Gεij]—where ℓε denotes elastic IL, εij is the elastic strain, and (λ, G) are the Lamé constants—is incorporated into classical Hooke's law σij = λεmmδij + 2ij, which now reads [1, 3] σij=λεmmδij+2Gεij-c2[λεmmδij+2Gεij];c=ε2. Similarly, the term p22γp– where ℓp denotes plastic IL and γp=γ˙pdt;γ˙p=2ε˙ijpε˙ijp is the equivalent plastic strain with εijp denoting the plastic strain tensor and a superimposed “·” denotes time derivative—is introduced in the classical von Mises yield condition τ=κ(γp);τ12σijσij, σij=σij(1/3)σmmδij, which now reads [1, 2] τ=κ(γp)-c2γp;c= p2κ(γp).

For diffusion problems, the IL enters through the additional term d22ji—where ℓd is a diffusional internal length and ji denotes the diffusion flux—which generalizes the classical Fick's law ρ˙=D2ρ by the expression ρ˙=D2ρ-c4ρ;c=d2D, in a manner formally similar to the CahnHilliard theory [7] of spinodal decomposition. For collective dislocation phenomena, the IL enters through the extra Laplacian terms Di,m2ρi,m, where ρi, m denote (immobile, mobile) dislocation densities and Di, m “effective” diffusion-like transport coefficients [12], which are absent from conventional dislocation dynamics models.

For coupled elasto-diffusion processes, higher order IL couplings have not been adequately considered with the exception of some recent works on chemomechanical damage of LiB anodes and propagation of lithiation fronts. Colossal mechanical stresses develop during lithiation, and the following gradient enhanced stress-assisted diffusion equation may be used to model Li intercalation

ρt=(D+Nσh0)2[ρ-ρ22ρ]          -Mσh0·[ρ-ρ22ρ].    (1)

In the above expression, ρ is the Li concentration, σh0 denotes the hydrostatic component of mechanical stress, and (D, N, M) are phenomenological coefficients, whereas ℓρ denotes a diffusional internal length entering in an analogous (slightly different) way as ℓd above. It is noted that Equation (1) is an IL extension of the stress-assisted diffusion equation earlier used for hydrogen embrittlement [13]. It is an uncoupled equation, since σh0 is assumed to be determined by the solution of a separate elasticity problem.

Higher order elasto-diffusion couplings were considered more recently in Tsagrakis and Aifantis [14] to model the propagation of lithiation fronts. The corresponding constitutive equations for the stress (σ) and the chemical potential (μ) are given by

σ=2Gε+λ(tr ε)1-ε22[2Gε+λ(tr ε)1        -(2G+3λ)Mρ1,μ=μ0+RT[ln(ρ1-ρ)+α(1-2ρ)]        -κ2ρ-Ωσh.    (2)

Here again, ρ is a normalized Li concentration, ε is the strain, (λ, G) denote the Lamé constants, and (M, Ω) are chemomechanical phenomenological coefficients. The quantities μ0(reference value of μ) and α denote chemical parameters of ideal solution theory, whereas ℓε and κ denote the mechanical and diffusional internal lengths, respectively.

Ilg in Rheology: Newtonian and Complex Fluids

In this section, we suggest various possibilities for a gradient enhancement of constitutive equations used in the fluid mechanics and rheology communities. In this connection, it is pointed out that (following the author's work on gradient theory), a number of such generalizations have already been proposed in these communities [1519]. In the spirit of the ILG formulation, such type of generalizations can readily be deduced by formally replacing the local fields for the fluid density ρ, stretching tensor D= 12[v+(v)T], and vorticity tensor W=12[v-(v)T], with their gradient-dependent counterparts ρ-ρ22ρ,D-D22D, W-W22W, respectively.

Another possibility is to include the Laplacian of the viscoelastic stress Σ, as proposed in the diffusive Johnson–Segalman (DJS) model, i.e.,

T=-p1+2μD+Σ;div T=0,    (3)

with p denoting pressure, μ being the solvent's shear viscosity [not to be confused with the same symbol used for the chemical potential in Equation (2)], and the second equation standing for quasi-static equilibrium. The newly introduced viscoelastic stress Σ obeys the following evolution equation

Σo+1τΣ=2μ*τ+D2Σ,    (4)

where τ denotes relaxation time, μ* is the micelle (polymer-like) viscosity, D is a diffusion-like coefficient, and a superimposed “o” denotes corotational time derivative.

On returning to the topic of an appropriate generalization of the Navier–Stokes (N–S) equations for incompressible fluids, i.e., of the constitutive equation T = −p1 + 2μD, we can propose the following gradient model

T-T22T=-p1+2μ(D-D22D),    (5)

where ℓT and ℓD denote internal lengths associated with stress and strain rate inhomogeneities. On assuming that ℓT can be neglected and introducing Equation (5) in the equation of momentum balance ρv˙=divT, we obtain the following gradient generalization of the N–S equations

ρv˙=-p+μ(Δv-D2Δ2v),    (6)

where Δ = ∇2 and Δ2 = ∇4 denote the Laplacian and biharmonic operators, respectively [16]. A slightly generalized model to consider turbulence, reads

ρv˙=-p+μ(1-α2Δ)Δv+2ρα2divD,    (7)

where the α parameter [not to be connected with the same symbol used in Equation (2)] denotes a statistical correlation length and D=D +DW-WD denotes the usual Jaumann rate.

Steady-state solutions of Equation (6) may be determined by employing the operator split method (or the use of Ru–Aifantis theorem [20]) utilized to eliminate singularities from dislocation lines and crack tips in the theory of gradient elasticity (see also [1]). This same procedure leads to the cancelation of singularities in typical fluid flow calculations involving immersed objects. It turns out, for example, that the resulting gradient Oseen tensor OijG, which generalizes its classical counterpart Oij

Oij=18πμ r(δij+rirjr2),    (8)

where ri denotes the position vector and r its magnitude, reads

OijG=18πμ r{[1-2e-r/-2re-r/+22r2(1-e-r/)]δij           +[1+2e-r+6re-r-62r2(1-e-r/)]rirjr2}.    (9)

This expression is essentially the same with the exponential regularization of the Green's tensor emerged in gradient elasticity/GradEla and used to rederive non-singular expressions for the stresses and strains in dislocation lines and crack tips. More details on such gradient fluids aspects can be found in Giusteri and Fried [21], the authors of which it seems not to have been aware of analogous developments in gradient elasticity.

Ilg in Other Disciplines and Scales

In this section, we summarize the applicability of the ILG framework to other disciplines and scales ranging from earth scales to quantum scales.

ILG in Geology

Some initial work on introducing internal lengths and Laplacians of strain in geomodels has been published by the author and coworkers to model shear banding and related instability phenomena in soils/rocks and snow/ice (see, for example [2233]). Various types of gradient-dependent constitutive equations for such classes of geomaterials have also been proposed and elaborated upon in detail by many other authors. This was due to the fact that the Laplacian regularizes unstable behavior in the geomaterial's softening regime and allows for the determination of shear band thickness/spacing, eliminates mesh-size dependence, and establishes convergence of corresponding finite element calculations [2]. The popularization of the approach in the geomechanics community is mainly due to the follow-up works by Vardoulakis and collaborators for soils, as well as de Borst and collaborators for concrete. These are too many to mention here, and most of them can be found in the web.

In connection with the above, it is worth noting that the Walgraef-Aifantis/W-A model for dislocation patterning has recently been used by Ord and Hobbs [34] to interpret fracture patterns in frictional, cohesive, granular materials. Their article was one contribution of 17 to a Theme Issue “Patterns in our planet: Applications of multi-scale non-equilibrium thermodynamics to Earth-system science”.

ILG in Electrodynamics

The inclusion of higher order gradients in deforming materials under the action of electromagnetic fields has also become very popular in recent years due to emerging applications and novel design of piezoelectric (induction of electricity due to applied stress/strain) and flexoelectric (induction of electricity due to strain gradients) components. The number of written articles is prohibitive for mentioning them here, and we only refer to recent publications by the author and coworkers [3537], as well as to the bibliography listed there for recent literature related to size effects in micro-/nanoelectromechanical systems.

In relation to the issue of eliminating singularities and introducing screening effects (e.g., Debye screening) in the electric field, the following gradient modification of Coulomb's law of electrostatics has been used [see, for example, [38] where, in addition, a fractional generalization of Debye screening is also discussed], as follows:

2Φ(r)-1rD2Φ(r)=-1ε0ρ(r).    (10)

In this equation, Φ is the electrostatic potential [E(r) = −∇Φ(r); E(r) is the electric field], ρ(r) denotes the charge density, ε0 is the vacuum permittivity, rD is the Debye screening distance, and r denotes the spatial radial coordinate. The classical Coulomb's potential for a point charge of strength Q has the form Φ(r) = Q/4π ε0r, while its Debye-screened counterpart obtained from Equation (10)—which is identical in form to the reduced equation for gradient elasticity [20]—reads

Φ(r)=Q4πε0re-r/rD.    (11)

In concluding this discussion on gradient electrodynamics, an interesting possibility is outlined below for a gradient extension of the MacCullagh–Fitzgerald formal analogy between elastic deformation and light transmission/reflection/refraction in transparent media (also discussed briefly by Sommerfeld in his lectures), as pointed out in Truesdell and Toupin [39]. On assuming first that the ether behaves as an elastic medium with its stress T depending linearly on rotations ω(instead of strains), we can write

T=2kω,ω=1/2[u-(u)T];divT=ρu¨,    (12)

where u denotes displacement, ρ denotes density, and k is a constant. These lead to the equation k curl curl u+ρu¨=0 and by formally setting k curl ü → α E and ρu˙αβ (with E/B denoting electric/magnetic fields and α being an arbitrary proportionality constant), we arrive at Maxwell's equations [39]

Bt+curl E=0;divB=0 &Et-1μ0ε0curl B=0;             divE=0,    (13)

where the identities divcurlu = 0 and curl(∂u/∂t) − ∂(curlu)/∂t = 0 were used, along with the following identification of the various coefficients (β = 0, μ0 = ρ/β, with β being an arbitrary constant). The symbols (ε0, μ0) are the fundamental vacuum constants of electromagnetism with ε0μ0=1/c2 and c denoting the speed of light in vacuum. By repeating the same procedure as above but replacing Equation (12)1 for the ether's elastic stress with its gradient counterpart

T=2k(ω-22ω),    (14)

we arrive at the following generalization of Maxwell's equations

Bt+curl[(1-22)E]=0; divB=0,               Et-1μ0ε0curlB=0; divE=0.    (15)

It is noted that for electrostatics (with the usual assumption that the electric field E is proportional to a potential gradient ∇Φ), Podolsky's non-quantum equation for the electric potential, ∇2(1 − ℓ22)Φ = 0, is obtained [40, 41]1.

ILG in Atomistics and Quantum Mechanics

We conclude this section on the applicability of the ILG framework to various disciplines and scales by focusing on two specific topics: a possible gradient generalization of the molecular dynamics stress, and an analogous generalization of the quantum mechanical stress. In this connection, it is noted that the following expressions were proposed for these stresses:

σ=1V[12ifij(ri-rj)-imiυiυi],    (16)

in Zimmerman et al. [42], and

σαβ=-1Vipiαpiβmi                -12Vi,j(j  i)(ri-rj)α(ri-rj)β|ri-rj|Uij(|ri-rj|),    (17)

in Maranganti and Sharma [43], with the various symbols having their usual meaning [42, 43]. The formal similarity between these two expressions and their resemblance to the virial stress and other statistical stress measures is striking. However, the problem to connect such discrete “microscopic” stress measures with the continuum “macroscopic” measure of Cauchy stress in a “seamless” way is a challenging issue. A gradient generalization of the force fields fij in Equation (16) and the interaction potential Uij in Equation (17) may turn out to be quite useful in this respect. Among other things, it could also naturally introduce screening distances and eliminate associated singularities.

The effect of strain ε on the electronic structure has been described [44] through the equations

(Ec-22m*2)ψ(r)+actr(ε)ψ(r)=Eψ(r) and ε=ε0;σ=[C]ε;divσ=0,    (18)

where ψ(r) denotes the wavefunction, [C] is the Hookean (generally anisotropic) elasticity matrix, ac is the so-called deformation potential constant, and the rest of the symbols have their usual quantum mechanical meaning [44, 45]. This is an uncoupled framework where strain can affect the electronic state but not vice versa. A generalization to account for the inverse effect on strain due to changes in the quantum field through the wavefunction ψ(r) has already been proposed as follows [45]:

(Ec-22m*2)ψ(r)+actr(ε)ψ(r)=Eψ(r) and           ε=ε0-ac3K|ψ(r)|21;σ=[C]ε;divσ=0,    (19)

where K is the isotropic bulk elastic modulus. A possible gradient modification is then to replace ε with its gradient counterpart ε-ε22ε, and this formal generalization may be of interest to further explore.

Ilg Modification of Newton's Gravitation Law

In this final section, we venture a gradient generalization of Newton's Law, which allows for the corresponding gravitational force to attain values larger than the electromagnetic force and even reach the levels of the nuclear/strong force which keeps matter together2.

We begin with the following nonlocal generalization of the gravitational force f in its component form (fi)

fi(r)=Gij(r-r)Fj(r)d3r    (20)

where Gij(r-r) is a non-local interaction kernel, and Fj is the classical Newton's force. By Fourier transforming Equation [26], Taylor series expanding up to the second-order term and inverting, the following differential equation is obtained for the above listed expression of the generalized force f

(1-22)f=F;2δij=12|d2G~ij(0)dk2|    (21)

where k = |k| denotes wave vector, G~ij is the Fourier transform of Gij, and ℓ is an internal length, with δij appearing due to the assumed isotropy/spherical symmetry [In general the sign in front of the Laplacian term in Equation (21) may be positive or negative depending on the sign of d2G~ij(0)/dk2 of the second order term in the Taylor expansion without assuming its absolute due to stability considerations]. Such a formal derivation can also be established by viewing the two point masses M0 and M in the classical Newton's Law, as being distributed and bounded by spheres of finite radii. By considering, for example, the reference mass M0 (M0 = ∑i mi) as being distributed within a sphere of radius R0, summing up the interactions of each point mass mi (located at distanceri from the center of the sphere where ri = 0) with the other distant mass M (viewed as a point mass), and expanding in Taylor series the density ρ(ri) around ρ0 = ρ(0) by keeping terms up to the second order, we obtain the following relationship

f=GMVR2(ρ0+22ρ0)eR;2=R0210,    (22)

where eR = R/R denotes the unit vector along the line connecting the center of the reference mass M0 with the second distant point mass M. On setting Vρ0dV=M0, we then have f = (1 + ℓ22)F, which by inversion leads to (1 − ℓ22)f = F.

On assuming a radial dependence of f and F[f=f(r)er;F = F er,F=A/r2, with A denoting Newton's classical gravitational proportionality parameter and setting eR = er], we can readily solve the scalar radial counterpart of Equation (21) in spherical coordinates, i.e.

f-2(2fr2+2rfr-2fr2)=Ar2,    (23)

by also requiring that F → 0 as r → ∞. The result is

f=Ar2[1+Be-r/(1+r)],    (24)

where B is a new parameter to evaluate in connection with experiments. It is noted that the above expression of Equation [30] reduces to Newton's classical force FN=A/r2 as r → ∞ and to the expression FSF=AB/r2 as r → 0. By adjusting the value of the new parameter B (B >> 1), we can attain values of the nuclear and strong force for FSF, as outlined below.

First, we note that the internal length parameter (ℓ) can be identified with de Broglie relativistic length, the Compton length, the Planck length, or the Schwarzschild distance, according to the configuration at hand. If, for example, ℓ is identified with de Broglie relativistic length used in the Vayenas [46] rotating neutrino model (RNM), it follows that =/γm0c=6.31×10-16 m, where ℏ denotes the Planck constant, c is the speed of light, and m0 denotes the rest mass of neutrino, whereas G is the classical Newton's gravitational constant (not to be confused with the same symbol earlier used for the shear modulus), and γ is the Lorentz factor (γ=1/1-(v/c)2; with v denoting particle speed).

To proceed further, we adopt next the Vayenas and coworkers RNM method [46, 47] as applied to the nucleus (proton or neutron), but we use the gradient enhanced Newton's Law as given by Equation (24) with B ≠ 0 instead of the classical Newton's Law expression (B = 0). To this end, we shall equate f in Equation (24) with the centrifugal force FC=γm0c2/r. Here, r denotes the radius of the nucleus modeled by the three rotating neutrinos whose total relativistic mass is mN = 3γm0. For a proton nucleus (mNmp), an estimate of γ can be obtained by equating the proton energy mpc2 with the one corresponding to the relativistic neutrino mass. This gives the value of γ=mp3m0, which, according to experimental measurements for mp (= 9.38 × 108eV/c2 = 1.67 × 10−27kg) and m0 (= 0.04 eV/c2 = 7.13 × 10−38 kg) turns out to be equal to 7.818 × 109.

Having such a value of γ available, we can make effective use of the equality between gravitational and centrifugal forces, as proposed in the RNM approach, to deduce the relationship

FR=A3r2[1+Be-3r/(1+3r)]=FC=γm0c2r,    (25)

where the factor 3 rises by considering the resultant gravitational force FR=f/3. One possibility for the constantA is to assume that it is given by the expression A=Gm02γ2 to account for relativistic effects (during the interaction of each pair of the rotating neutrinos in the assumed RNM configuration) in the same manner as in the expression for the centrifugal force FC (where the square is now introduced due to the fact that two neutrinos are involved). The above relationship (with ℓ identified with de Broglie's relativistic length, as discussed in the previous paragraph, i.e., ℓ = 6.31 × 10−16m) gives

A=2.07×10-65 N ·m2 and B=3e3c(1+3)Gm02γ2=5.47×1039.    (26)

The resulting value of FR is

FR=7.92×104N,    (27)

i.e., the value of the strong force obtained for the RNM configuration [46, 47] using an entirely different approach. In that approach, Equation (25) with B = 0 was used with A=Gm02γ6, giving a value for γ=31/12mpl1/3m0-1/3=7.167×109, where mpl is the Planck mass (mpl=c/G), and the value of m0 was taken as m0=0.0436 eV/c2. Since γ = mp/3m0, this gives mp=9.38×108 eV/c2, i.e. the same value as the one used in the previous paragraph by properly adjusting the parameters (A, B), as well as by identifying the internal length parameter ℓ with de Broglie relativistic length. Other choices of (A, B, ℓ) are possible not only for the RNM configuration at hand but also other more complex geometric models for elementary particles represented by several neutrinos where a potential is convenient to use. This topic will be pursued independently in a future article.

Conclusions

A concise review of gradient models (across scales, materials, and processes) was provided based on the author's ILG approach. As a result, earlier references on generalized continuum mechanics and recent contributions on gradient and non-local theories were not discussed due to space limitation. For solids, one should single out the contributions of Eringen [48], Fleck and Hutchinson [49, 50], Gurtin and Anand [51], Gao et al. [52], Nix and Gao [53], de Borst et al. [54, 55], Geers et al. [56], Peerlings et al. [57], Willis et al. [58], Aifantis and Willis [59], and Polizzotto [60, 61]. Many more are included in a most recent and detailed article by Voyiadjis and Song [62] focusing on gradient plasticity. Gradients in fluid and granular flows were considered most recently by Goddard [63, 64]. For additional recent developments on granular flow, one may also consult references [6569], while for internal length interpretations based on kinetic theory, one may consult [70]. After writing this article, it came to the attention of the author that an expression similar to that derived herein and given by Equation (24) was also proposed on rather intuitive grounds by Fischbach et al. [71] in an effort to re-interpret existing measurements on earth's gravity (see also [72]). The values of their constants were entirely different than ours, as they used it for a reanalysis of the Eötvös experiment on Earth's gravitational field. There has been a vast literature on this expression, subsequently referred to as the “fifth force,” which we will discuss in a forthcoming publication, as this is beyond the scope of the present review.

Author Contributions

The author confirms being the sole contributor of this work and has approved it for publication.

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.

Acknowledgments

Useful discussions with my daughter Katerina Aifantis and my student Kostas Parisis along with my former university classmate Constantinos Vayenas, as well as the invitation of Joshua Dijksman to contribute to this volume are gratefully acknowledged. The work was greatly benefited from interactions within the RISE/FRAMED project no. 734485 (https://cordis.europa.eu/project/rcn/207050/factsheet/en) for which Aristotle University of Thessaloniki (AUTh) acts as coordinator. In this connection, thanks are extended to all beneficiary nodes and international partners of FRAMED. The gradient fluids section is a topic of the RISE/ATM2BT project no. 824022 (https://cordis.europa.eu/project/rcn/219192/factsheet/en) for which EU support is also acknowledged. This section was included in anticipation of a follow-up joint work between AUTh, Akita Univeristy/Japan, and Aston University/UK (which acts as project's ATM2BT coordinator). More details on this minireview can be found in a forthcoming arXiv preprint.

Footnotes

1. ^Podolsky has derived a generalization of Maxwell's equations through a variational principle, leading also to the appearance of ∇2B in addition to∇2E. This is also possible through the aforementioned analogy by replacing u with u−ℓ22u.

2. ^This question first emerged in discussions with my daughter during my visit in February 2019 to Florida and follow-up discussions with my university classmate Vayenas during his visit in June 2019 to Thessaloniki.

References

1. Aifantis EC. Internal length gradient (ILG) material mechanics across scales and disciplines. Adv Appl Mech. (2016) 49:1–110. doi: 10.1016/bs.aams.2016.08.001

CrossRef Full Text | Google Scholar

2. Aifantis EC. The physics of plastic deformation. Int J Plast. (1987) 3:211–47. doi: 10.1016/0749-6419(87)90021-0

CrossRef Full Text | Google Scholar

3. Askes H, Aifantis EC. Gradient elasticity in statics and dynamics: an overview of formulations, length scale identification procedures, finite element implementations and new results. Int J Solids Struct. (2011) 48:1962–90. doi: 10.1016/j.ijsolstr.2011.03.006

CrossRef Full Text | Google Scholar

4. Aifantis EC. Gradient nanomechanics: applications to deformation, fracture, and diffusion in nanopolycrystals. Metall Mater Trans A. (2011) 42:2985–98. doi: 10.1007/s11661-011-0725-9

CrossRef Full Text | Google Scholar

5. Forest S, Aifantis EC. Some links between recent gradient thermo-elasto-plasticity theories and the thermomechanics of generalized continua. Int J Solids Struct. (2010) 47:3367–76. doi: 10.1016/j.ijsolstr.2010.07.009

CrossRef Full Text | Google Scholar

6. Aifantis EC, Serrin JB. The mechanical theory of fluid interfaces and Maxwell's rule. J Coll Inter Sci. (1983) 96:517–29. doi: 10.1016/0021-9797(83)90053-X. [see also: EC Aifantis. Mechanics of Phase Transformations in: Phase Transformations. Aifantis EC, Gittus J, editors. London: Elsevier Appl. Sci. Pub. 1986].

CrossRef Full Text | Google Scholar

7. Cahn JW, Hilliard JE. Free energy of a nonuniform system. I. Interfacial free energy. J Chem Phys. (1958) 28:258–67. doi: 10.1063/1.1744102

CrossRef Full Text | Google Scholar

8. Murray JD. Mathematical Biology; Vol I – An Introduction. New York, NY: Springer (2002).

Google Scholar

9. Murray JD. Mathematical Biology, Vol II – Spatial Models and Biomedical Applications. New York, NY: Springer (2003) [corrected Second Printing, 2004].

Google Scholar

10. Aifantis KE, Hackney SA, Kumar VR (Editors.) High Energy Density Lithium Batteries: Materials, Engineering, Applications. Weinheim: Wiley-VCH (2010). [also translated in Chinese: China Machine Press, ISBN:9787111371786, 2011] doi: 10.1002/9783527630011

CrossRef Full Text | Google Scholar

11. Hatzikirou H, Basanta D, Simon M, Schaller K, Deutsch A. ‘Go or Grow’: the key to the emergence of invasion in tumour progression? Math Med Biol. (2010) 29:49–65. doi: 10.1093/imammb/dqq011

PubMed Abstract | CrossRef Full Text | Google Scholar

12. Walgraef D, Aifantis EC. Dislocation patterning in fatigued metals as a result of dynamical instabilities. J Appl Phys. (1985) 58:688–91. doi: 10.1063/1.336183

CrossRef Full Text | Google Scholar

13. Unger DJ, Gerberich WW, Aifantis EC. Further remarks on the implications of steady state stress assisted diffusion on environmental cracking. Scripta Metall. (1982) 16:1059–64. doi: 10.1016/0036-9748(82)90456-2

CrossRef Full Text | Google Scholar

14. Tsagrakis I, Aifantis EC. Gradient elasticity effects on the two-phase lithiation of LiB anodes. In: Altenbach H, et al., editors. Generalized Models and Non-classical Approaches in Complex Materials 2. Cham: Springer (2018). p. 221–35. doi: 10.1007/978-3-319-77504-3_11

CrossRef Full Text | Google Scholar

15. Silber G, Trostel R, Alizadeh M, Benderoth G. A continuum mechanical gradient theory with applications to fluid mechanics. J Phys IV France. (1998) 8:365–73. doi: 10.1051/jp4:1998845

CrossRef Full Text | Google Scholar

16. Fried E, Gurtin ME. Tractions, balances, and boundary conditions for nonsimple materials with application to liquid flow at small-length scales. Arch Rat Mech Anal. (2006) 182:513–54. doi: 10.1007/s00205-006-0015-7

CrossRef Full Text | Google Scholar

17. Adams JM, Fielding SM, Olmsted PD. The interplay between boundary conditions and flow geometries in shear banding: hysteresis, band configurations, and surface transitions. J Nonnewton Fluid Mech. (2008) 151:101–18. doi: 10.1016/j.jnnfm.2008.01.008

CrossRef Full Text | Google Scholar

18. Cates ME, Fielding SM. Rheology of giant micelles. Adv Phys. (2006) 55:799–879. doi: 10.1080/00018730601082029

CrossRef Full Text | Google Scholar

19. Dhont JKG, Briels WJ. Gradient and vorticity banding. Rheol Acta. (2008) 47:257–81. doi: 10.1007/s00397-007-0245-0

CrossRef Full Text | Google Scholar

20. Ru CQ, Aifantis EC. A simple approach to solve boundary-value problems in gradient elasticity. Acta Mech. (1993) 101:59–68. doi: 10.1007/BF01175597

CrossRef Full Text | Google Scholar

21. Giusteri GG, Fried E. Slender-body theory for viscous flow via dimensional reduction and hyperviscous regularization. Meccanica. (2014) 49:2153–67. doi: 10.1007/s11012-014-9890-4

CrossRef Full Text | Google Scholar

22. Vardoulakis I, Aifantis EC. Gradient dependent dilatancy and its implications in shear banding and liquefaction. Ingenieur-Arch. (1989) 59:197–208. doi: 10.1007/BF00532250

CrossRef Full Text | Google Scholar

23. Oka F, Yashima A, Sawada K, Aifantis EC. Instability of gradient-dependent elastoviscoplastic model for clay and strain localization. Comp Meth Appl Mech Eng. (2000) 183:67–86. doi: 10.1016/S0045-7825(99)00212-1

CrossRef Full Text | Google Scholar

24. di Prisco C, Imposimato S, Aifantis EC. A visco-plastic constitutive model for granular soils modified according to non-local and gradient approaches. Int J Num Anal Meth Geomech. (2002) 26:121–38. doi: 10.1002/nag.195

CrossRef Full Text | Google Scholar

25. Fyffe B, Schwerdtfeger J, Blackford JR, Zaiser M, Konstantinidis A, Aifantis EC. Fracture toughness of snow: the influence of layered microstructure. J Mech Behav Mater. (2007) 18:195–215. doi: 10.1515/JMBM.2007.18.3.195

CrossRef Full Text | Google Scholar

26. Konstantinidis A, Cornetti P, Pugno N, Aifantis EC. Application of gradient theory and quantized fracture mechanics in snow avalanches. J Mech Behav Mater. (2009) 19:39–48. doi: 10.1515/JMBM.2009.19.1.39

CrossRef Full Text | Google Scholar

27. Haoxiang C, Qi C, Peng L, Kairui L, Aifantis EC. Modeling the zonal disintegration of rocks near deep level tunnels by gradient internal variable continuous phase transition theory. J Mech Behav Mater. (2015) 24:161–71. doi: 10.1515/jmbm-2015-0025

CrossRef Full Text | Google Scholar

28. Qi C, Wei X, Hongsen W, Aifantis EC. On temporal-structural dynamic failure criteria for rocks. J Mech Behav Mater. (2015) 24:173–81. doi: 10.1515/jmbm-2015-0026

CrossRef Full Text | Google Scholar

29. Efremidis G, Avlonitis M, Konstantinidis A, Aifantis EC. A statistical study of precursor activity in earthquake-induced landslides. Comput Geotech. (2017) 81:137–42. doi: 10.1016/j.compgeo.2016.08.010

CrossRef Full Text | Google Scholar

30. Chen H, Qi C, Efremidis G, Dorogov M, Aifantis EC. Gradient elasticity and size effect for the borehole problem. Acta Mech. (2018) 229:3305–18. doi: 10.1007/s00707-018-2109-3

CrossRef Full Text | Google Scholar

31. Vardoulakis I, Muhlhaus HB, Aifantis EC. Continuum models for localized deformations in pressure sensitive materials. In: Beer G, et al., editors. Computer Methods and Advances in Geomechanics. Rotterdam: Balkema Publications. (1991). p. 441–8.

Google Scholar

32. Vardoulakis I, Aifantis EC. A gradient flow theory of plasticity for granular materials. Acta Mech. (1991) 87:197–217. doi: 10.1007/BF01299795

CrossRef Full Text | Google Scholar

33. Vardoulakis I, Aifantis EC. On the role of microstructure in the behavior of soils: effects of higher order gradients and internal inertia. Mech Mater. (1994) 18:151–8. doi: 10.1016/0167-6636(94)00002-6

CrossRef Full Text | Google Scholar

34. Ord A, Hobbs BE. Fracture pattern formation in frictional, cohesive, granular material. Philos Trans R Soc A Math Phys Eng Sci. (2010) 368:95–118. doi: 10.1098/rsta.2009.0199

PubMed Abstract | CrossRef Full Text | Google Scholar

35. Yue YM, Xu KY, Aifantis EC. Microscale size effects on the electromechanical coupling in piezoelectric material for anti-plane problem. Smart Mater Struct. (2014) 23:125043. doi: 10.1088/0964-1726/23/12/125043

CrossRef Full Text | Google Scholar

36. Yue YM, Xu KY, Chen T, Aifantis EC. Size effects on magnetoelectric response of multiferroic composite with inhomogeneities. Phys B. (2015) 478:36–42. doi: 10.1016/j.physb.2015.08.056

CrossRef Full Text | Google Scholar

37. Yue Y, Xu KY, Aifantis EC. Strain gradient and electric field gradient effects in piezoelectric cantilever beams. J Mech Behav Mater. (2015) 24:121–7. doi: 10.1515/jmbm-2015-0014

CrossRef Full Text | Google Scholar

38. Tarasov VE, Trujillo JJ. Fractional power-law spatial dispersion in electrodynamics. Ann Phys. (2013) 334:1–23. doi: 10.1016/j.aop.2013.03.014

CrossRef Full Text | Google Scholar

39. Truesdell C, Toupin R. The classical field theories. In: Flügge S, editor. Principles of Classical Mechanics and Field Theory/Prinzipien der Klassischen Mechanik und Feldtheorie. Encyclopedia of Physics/Handbuch der Physik. Vol 2/3/1. Berlin; Heidelberg: Springer. (1960). p. 226–858. doi: 10.1007/978-3-642-45943-6_2

CrossRef Full Text | Google Scholar

40. Podolsky B. A generalized electrodynamics Part I—Non-quantum. Phys Rev. (1942) 62:68–71.

Google Scholar

41. Podolsky B, Schwed P. Review of a generalized electrodynamics. Rev Mod Phys. (1948) 20:40–50.

Google Scholar

42. Zimmerman JA, Webb EB III, Hoyt JJ, Jones RE, Klein PA, Bammann DJ. Calculation of stress in atomistic simulation. Model Simul Mater Sci Eng. (2004) 12:S319–32. doi: 10.1088/0965-0393/12/4/S03

CrossRef Full Text | Google Scholar

43. Maranganti R, Sharma P. Revisiting quantum notions of stress. Proc R Soc A. (2010) 466:2097–116. doi: 10.1098/rspa.2009.0636

CrossRef Full Text | Google Scholar

44. Davies H. The Physics of Low-Dimensional Semiconductors. Cambridge: Cambridge University Press (2000).

Google Scholar

45. Zhang X, Gharbi M, Sharma P, Johnson HT. Quantum field induced strains in nanostructures and prospects for optical actuation. Int J Solids Struct. (2009) 46:3810–24. doi: 10.1016/j.ijsolstr.2009.07.006

CrossRef Full Text | Google Scholar

46. Vayenas CG, Souentie SNA. Gravity, Special Relativity, and the Strong Force. Boston, MA: Springer US (2012). doi: 10.1007/978-1-4614-3936-3

CrossRef Full Text | Google Scholar

47. Vayenas CG, Souentie S, Fokas A. A Bohr-type model of a composite particle using gravity as the attractive force. Phys A Stat Mech Appl. (2014) 405:360–79. doi: 10.1016/j.physa.2014.03.045

CrossRef Full Text | Google Scholar

48. Eringen AC. Microcontinuum Field Theories I: Foundations and Solids. New York, NY: Springer (1999). doi: 10.1007/978-1-4612-0555-5

CrossRef Full Text | Google Scholar

49. Fleck NA, Hutchinson JW. Strain gradient plasticity. Adv Appl Mech. (1997) 33:295–361. doi: 10.1016/S0065-2156(08)70388-0

CrossRef Full Text | Google Scholar

50. Fleck NA, Hutchinson JW. A reformulation of strain gradient plasticity. J Mech Phys Solids. (2001) 49:2245–71. doi: 10.1016/S0022-5096(01)00049-7

CrossRef Full Text | Google Scholar

51. Gurtin ME, Anand L. Thermodynamics applied to gradient theories involving the accumulated plastic strain: the theories of Aifantis and Fleck and Hutchinson and their generalization. J Mech Phys Solids. (2009) 57:405–21. doi: 10.1016/j.jmps.2008.12.002

CrossRef Full Text | Google Scholar

52. Gao H, Huang Y, Nix WD, Hutchinson JW. Mechanism-based strain gradient plasticity - I. Theory. J Mech Phys Solids. (1999) 47:1239–63. doi: 10.1016/S0022-5096(98)00103-3

CrossRef Full Text | Google Scholar

53. Nix WD, Gao HJ. Indentation size effects in crystalline materials: a law for strain gradient plasticity. J Mech Phys Solids. (1998) 46:411–25. doi: 10.1016/S0022-5096(97)00086-0

CrossRef Full Text | Google Scholar

54. de Borst R, Muhlhaus HB. Gradient-dependent plasticity - formulation and algorithmic aspects. Int J Numer Methods Eng. (1992) 35:521–39. doi: 10.1002/nme.1620350307

CrossRef Full Text | Google Scholar

55. de Borst R, Pamin J, Sluys LJ. Computational issues in gradient plasticity. In: Mühlhaus HB, editor. Continuum Models for Materials With Microstructure. Chichester: Wiley (1995). p. 159–200.

Google Scholar

56. Geers MGD, Peerlings RHJ, Brekelmans WAM, de Borst R. Phenomenological nonlocal approaches based on implicit gradient-enhanced damage. Acta Mech. (2000) 144:1–15. doi: 10.1007/BF01181824

CrossRef Full Text | Google Scholar

57. Peerlings RHJ, Poh LH, Geers MGD. An implicit gradient plasticity-damage theory for predicting size effects in hardening and softening. Eng Fract Mech. (2012) 95:2–12. doi: 10.1016/j.engfracmech.2011.12.016

CrossRef Full Text | Google Scholar

58. Willis JR. Some forms and properties of models of strain-gradient plasticity. J Mech Phys Solids. (2019) 123:348–56. doi: 10.1016/j.jmps.2018.09.011

CrossRef Full Text | Google Scholar

59. Aifantis KE, Willis JR. The role of interfaces in enhancing the yield strength of composites and polycrystals. J Mech Phys Solids. (2005) 53:1047–70. doi: 10.1016/j.jmps.2004.12.003

CrossRef Full Text | Google Scholar

60. Polizzotto C. Unified thermodynamic framework-for nonlocal/gradient continuum theories. Eur J Mech A Solid. (2003) 22:651–68. doi: 10.1016/S0997-7538(03)00075-5

CrossRef Full Text | Google Scholar

61. Polizzotto C. Interfacial energy effects within the framework of strain gradient plasticity. Int J Solids Struct. (2009) 46:1685–94. doi: 10.1016/j.ijsolstr.2008.12.010

CrossRef Full Text | Google Scholar

62. Voyiadjis GZ, Song Y. Strain gradient continuum plasticity theories: theoretical, numerical and experimental investigations. Int J Plast. (2019) 121:21–75. doi: 10.1016/j.ijplas.2019.03.002

CrossRef Full Text | Google Scholar

63. Goddard JD. On linear non-local thermo-viscoelastic waves in fluids. Mat Mech Compl Syst. (2018) 6:321–38. doi: 10.2140/memocs.2018.6.321

CrossRef Full Text | Google Scholar

64. Goddard JD. On the stability of the μ(I) rheology for granular flow. J Fluid Mech. (2017) 833:302–31. doi: 10.1017/jfm.2017.651

CrossRef Full Text | Google Scholar

65. Kamrin K, Koval G. Nonlocal constitutive relation for steady granular flow. Phys Rev Lett. (2012) 108:178301. doi: 10.1103/PhysRevLett.108.199904

PubMed Abstract | CrossRef Full Text | Google Scholar

66. Forterre Y, Pouliquen O. Flows of dense granular media. Annu Rev Fluid Mech. (2008) 40:1–24. doi: 10.1146/annurev.fluid.40.111406.102142

CrossRef Full Text | Google Scholar

67. Fenistein D, van Hecke M. Wide shear zones in granular bulk flow. Nature. (2003) 425:256. doi: 10.1038/425256a

PubMed Abstract | CrossRef Full Text | Google Scholar

68. Dijksman JA, Wortel GH, van Dellen LTH, Dauchot O, van Hecke M. Jamming, yielding, and rheology of weakly vibrated granular media. Phys Rev Lett. (2011) 107:108303. doi: 10.1103/PhysRevLett.107.108303

PubMed Abstract | CrossRef Full Text | Google Scholar

69. Henann DL, Kamrin K. A predictive, size-dependent continuum model for dense granular flows. Proc Natl Acad Sci USA. (2013) 110:6730–5. doi: 10.1073/pnas.1219153110

PubMed Abstract | CrossRef Full Text | Google Scholar

70. Bocquet L, Colin A, Ajdari A. Kinetic theory of plastic flow in soft glassy materials. Phys Rev Lett. (2009) 103:036001. doi: 10.1103/PhysRevLett.103.036001

PubMed Abstract | CrossRef Full Text | Google Scholar

71. Fischbach E, Sudarsky D, Szafer A, Talmadge C, Aronson SH. Reanalysis of the Eötvös experiment. Phys Rev Lett. (1986) 56:3–6. doi: 10.1103/PhysRevLett.56.3

CrossRef Full Text | Google Scholar

72. Fischbach E. The fifth force: a personal history. Eur Phys J. (2015) 40:385–467. doi: 10.1140/epjh/e2015-60044-5

CrossRef Full Text | Google Scholar

Keywords: gradient elasticity, plasticity, fluidity, gradient electrodynamics, gradient gravity

Citation: Aifantis EC (2020) A Concise Review of Gradient Models in Mechanics and Physics. Front. Phys. 7:239. doi: 10.3389/fphy.2019.00239

Received: 12 September 2019; Accepted: 16 December 2019;
Published: 05 February 2020.

Edited by:

Joshua Albert Dijksman, Wageningen University & Research, Netherlands

Reviewed by:

Masoud Hoore, Helmholtz Center for Infection Research, Helmholtz Association of German Research Centers (HZ), Germany
Prateek Kumar Jha, Indian Institute of Technology Roorkee, India

Copyright © 2020 Aifantis. 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: Elias C. Aifantis, mom@mom.gen.auth.gr

Download