Partial Equilibration Scenario in 3D athermal martensites
quenched below first-order transition temperatures
Abstract
To test a Partial Equilibration Scenario (PES) of Ritort and colleagues, we do Monte Carlo simulations of discretized-strain spin models, for four 3D martensitic structural transitions under quenches to a bath temperature below a first-order transition. The ageing system faces entropy barriers, in searches for energy-lowering passages between quasi-microcanonical energy shells. We confirm the PES signature of an exponential-tail distribution of intermittent heat releases to the bath, scaled in an effective temperature, that in our case, depends on the quench. When its inverse vanishes below a ‘martensite start’ temperature of avalanche conversions, then entropy barriers vanish. When this search temperature vanishes, PES cooling is arrested, as entropy barriers diverge. We find a linear vanishing of , below a delay-divergence temperature in between, . Martensitic conversion delays thus have Vogel-Fulcher-Tammann like divergences. Post-quench delay data extracted from simulations and athermal martensitic alloys, are both consistent with predictions.
I Introduction
The re-equilibration of a system after a quench is a long-standing problem in non-equilibrium statistical mechanics, and a generic Partial Equilibration Scenario (PES) has been proposed by Ritort and colleagues R1 ; R2 ; R3 ; R4 After a quench, the system spreads rapidly over an energy shell in configuration space. The system lowers its energy by intermittent energy releases to the heat bath at , and then again spreads ergodically over the next energy shell . The iteration of the fast/ slow steps ratchets the system down to a canonical equilibrium at .The distribution of energy changes has a signature exponential tail R1 ; R2 ; R3 ; R4 in the heat release distribution , whose slope at the origin defines the (inverse) effective temperature , dependent on the post-quench waiting time . The tail for negative energies is part of a shifted gaussian that peaks at positive energies. The PES has been confirmed through analytic Monte Carlo (MC) calculations of relaxing independent harmonic oscillators R1 ; R2 ; by simulations of spin-glass models and Lennard-Jones binary mixturesR3 ; R4 , and through voltage noise intermittency R3 . The MC updates of ageing harmonic oscillators R2 have to hit the ever-shrinking target of as yet unrelaxed oscillators, and these rising entropy barriers induce slow decreases in energies and acceptances .
It is natural to test PES ideas in interacting systems with slow relaxations. The structural glass transition R5 is an attempted equilibration that is arrested at a glassy freezing temperature , pre-empting crystallization R5 ; R6 . The -dependent effective viscosity increases above a Vogel-Fulcher-Tammann (VFT) singularity R7 ; R8 , that has been studied for a century, but is not yet fully understood. Further, decays around the glass temperature have non-exponential time dependence; and non-Debye frequency responses R9 . We consider other systems with structural transitions and equilibration delays. Martensitic steel alloys, when quenched from high temperature parent austenite to low temperature ‘martensite’ R10 ; R11 , show strain domain-wallR12 patterns R13 ; R14 ; R15 .They can exhibit puzzling delays in conversion to martensite R16 ; R17 ; R18 ; R19 ; R20 , that increase rapidly with temperature: raising to nearer transition by a few percent, can raise delays from sec to seconds R18 .
We do Monte Carlo (MC) simulations of quenches in , of martensitic discretized-strainR21 ; R22 ; R23 models in 3D. The model hamiltonians describe the elastic Domain Walls (DW) or mobile twin boundaries, of four 3D structural transitions, each with characteristic anisotropic Compatibility interactions between order-parameter strainsR15 . The four transitions R21 can occur in martensite-related functional materials, where OP strains will be coupled to the functionality variables. The transitions are tetragonal-orthorhombic (YBCO, superconductivity); cubic-tetragonal (FePd, shape memory); cubic-orthorhombic (BaTiO, ferroelectrics); cubic trigonal (LaSrMnO, colossal magnetoresistance).
Our 3D simulations yield post-quench evolutions as in R24 ; R25 ; R26 2D, passing through three Domain Wall (DW) states. At first there is a majority-austenite DW Vapour state of a martensite droplet in an austenite background. This converts to majority-martensite DW Liquid, of randomly wandering walls. Finally the DW Liquid orders to a DW Crystal microstructure, with the walls along preferred directions R13 , as in parallel ‘twins’.
We focus on the conversion delays of the first evolution of DW Vapour DW Liquid, that corresponds to austenite to martensite conversions, or a rise of the martensite fraction from zero to unity. As in the earlier 2D case R24 ; R25 ; R26 , a phase diagram in material parameters is obtained. In the ‘athermal’ martensite regime, there are curious ‘incubations’, or no apparent changes after a quench, terminated by sudden avalanche conversions R16 ; R17 ; R18 ; R19 ; R20 . In this regime, we find three characteristic temperatures, with . Here is the meanfield transition to uniform ordering.
Avalanche conversions in a single MC time step ,occur for , identified as the martensite start temperatureR10 ; R11 . Quenches into show (postponed) avalanches R27 or incubation behaviour: the conversion fraction remains flat at zero, until there is a jump up to unity a time . The incubation delay time extends rapidly, on approaching a divergence temperature .The physical picture for delays is of Vapour-droplet Fourier profile attempting entry to a negative-energy region of effective Hamiltonian spectra . The profile has to pass through a zero-energy -space contour at a bottleneck, like a Golf Hole edge R28 ; R29 ; R30 ; R31 . This transit passage delay differs from the familiar critical slowing down from a divergent Order Parameter length. The -dependent, anisotropic bottleneck shrinks on warming, with a topological shape change at , that blocks entry, so entropy barriers diverge. The precursorR14 ; R15 region , where 2D results suggest dynamic tweed R27 , may be studied elsewhere.
We use generic equilibration scenario of Ritort and colleagues to analyze the statistics R1 ; R2 ; R3 ; R4 ; R31 , of the set of energy changes from each MC step (usually used but not retained R32 ). These heat releases are recorded only up to an aging time , so the effective temperature depends on the quench temperature, : non-stationary distributions become time independent. We confirm the signature PES exponential tail for all four transitions. We find the search temperature vanishes linearly , driving an entropy barrier divergence at . The martensite-conversion times are thus predicted to show glass-like VFT behaviour, here understood as an arrest of PES cooling. Such VFT behaviour, is extracted from simulation and experimental R16 ; R17 ; R18 data.

The plan of the paper is as follows. In Section II, we discuss the generic Partial Equilibration Scenario and our specific case of quenching across a phase transition. Section III describes for the four transitions in 3D, the discrete-strain clock-like spins and their -dependent Hamiltonians. In Section IV we describe the MC simulations, with delay results from the phase space bottlenecks in Section V. Section VI shows that PES signatures are seen in all four transitions. Section VII shows that both 3D simulations and metallic alloy experiments exhibit VFT behaviour. Finally Section VIII is a summary.
Appendix A illustrates how a continuum double-well Landau free energy induces a T-dependent, Ising effective Hamiltonian. Appendix B obtains the athermal phase diagram for four transitions. Movies of post-quench DW evolutions are in Supplementary Material Videos R27 .

II Scenario for post-quench equilibration:
How do systems re-equilibrate, after a temperature quench? Ritort and coworkers have suggested that if an equilibrium canonical ensemble in thermal contact with a heat bath suddenly has its bath quenched to a lower temperature , then the system goes into an ageing ensembleR1 ; R2 ; R3 ; R4 , that has a quasi- microcanonical description of states of the system. There are sequential passages through decreasing-energy configurational shells, and intervening entropy barriers in the system-search for the new equilibrium. While delays from energy barriers are from attempts through activated jumps, to cross mountains, delays from entropy barriers are from attempts through constant-energy searches, to find rare channels going through or around, the mountain R31 .
In this Section we i) outline (our understanding of) the generic Partial Equilibration Ensemble R1 ; R2 ; R3 ; R4 or PES, and ii) state how this ageing scenario is applied to our specific case, that has quenches across phase transitions, and order parameter emergence from zero.
i) Generic PES for ageing after quenches:
The PES for the equilibration process considers a system of energy in contact with a (larger) heat bath. A familiar textbook derivation R33 of the canonical ensemble, applies the microcanonical ensemble to the system plus heat bath, of constant total energy . The total number of states is the product of the states of the system, and of the states of the bath, summed over all allowed system energies :
(1) |
The system configurational entropy is , and the inverse effective temperature is . Similarly, the inverse bath temperature is . The change in total entropy depends on the system energy change as
(2) |
with the equality at equilibrium, when the system and bath temperatures are equal . The Second Law inequality must hold, for the irreversible cases. After a bath quench from initial to a final , the system is left hotter, , or . The energy changes are negative , with heat released by the system to the bath.
At equilibrium, the terms in the sum of Equ (1) are dominated by clusters of energy shells selected R33 by a sharp peak, arising from the product of a rising number of system states , and a falling bath factor . The peak width is the energy fluctuations from stochastic system-bath exchanges. These equilibrium ideas describe states, not processes.
The Partial Equilibration Scenario postulates a plausible post-quench non-equilibrium process for the system to evolve between the initial and the final equilibrium state. A sudden change in bath temperature or quench, will induce a shifted peak, around a different equilibrium state. The post-quench system, initially stranded in non-optimum states, is visualized as moving through the sequentially lower energy shells of Equ (1) in its search for the shifted peak, tracked by an ageing time .
The Scenario postulates that the system i) rapidly spreads ergodically through all states of a shell of energy E, and ii) slowly dribbles out energy to the ever-present energy bath. Since the system is partially equilibrated in the quasi-microcanonical shell, the equilibrium definitions can be retained, of the shell entropy and its energy derivative . The back-and-forth energy exchanges to rapidly surmount internal energy barriers and explore all shell configurations, are summoned by the system from the bath (‘stimulated’). The slow changes on passages to a lower-energy shell, are releases by the system to the bath (’spontaneous’).
Fig 1 is a schematic of the Partial Equilibration Scenario. The successive shells have lower energy and hence lower number of configurations and entropies . There is a generic entropy barrier to finding the rarer states. Key seeks lock: most attempts fail.
From an ageing Fluctuation Relation R3 the nonequilibrium energy-change probability R1 ; R2 ; R3 ; R4 is a peak at the origin, times an exponential tail for negative changes. This generic PES signature tail depends on the ageing time through the the effective temperature , that scales the heat releases:
(3) |
with an even prefactor .
In an important result, other effective temperatures, from the Fluctuation-Dissipation Theorem; and from non-equilibrium fluctuations of system variables, are shown to be equal to the PES effective temperature R1 : there is only one .


ii) Specific PES from quenching across a transition.
For our case of quenching to across a first-order transition, the Order Parameter (OP) has to rise from zero, and so the wait times to reach OP marker events will depend on the quench temperature, .The effective temperature and PES distribution will thus also depend on the quench temperature, and . The even prefactor can be exponentiated and expanded to quadratic order, Equ (3) is then a gaussian peaked at the origin, times an exponential falloff. Completing the square yields a PES signature of a shifted gaussian, peaked at positive mean changes , and scaled in :
(4) |
For small heat releases , the PES distribution takes a Boltzmann-like form . This gives a physical meaning to the effective temperature: it is a search range for accessible energy shells. If , entropy barriers collapse, and passages are immediate. If , then entropy barriers diverge, and passage-searches freeze.
We postulate that the OP-related bottlenecks can be of two types, depending on the depth of the quench. a) The DW Vapour Liquid delays are attributed to phase space bottlenecks R27 , suggested by concepts in protein folding R28 ; R29 ; R30 . Hamiltonian energy-spectrum contours in Fourier space of zero energy are like a -dependent Golf Hole (GH) edge, with a negative-energy Funnel region inside it leading to the final state R28 . The entropic delays are from finding and entering the bottleneck.
b) For deeper quenches, the DW Liquid Crystal delays could be conceptually related to spin facilitation models R27 ; R34 ; R35 , through the -dependent sparseness of austenitic-hotspot dynamical catalysts, or other facilitating fields R20 ; R27 .
III Domain-wall Hamiltonians for four structural transitions in 3D:
The transition-specific, derived effective Hamiltonians have been presented in detailR15 ; R21 ; R22 ; R23 ; R24 ; R25 ; R26 , and we just outline as conceptual background: A) Strains and Compatibility constraints. B) Reduction of continuum strains to discrete-strain ‘pseudo-spins’. C) Reduction of continuum strain free energies to effective ‘spin’ Hamiltonians.
It is useful to define the number of components of the OP strains; and the number of Landau ‘variant’ minima at nonzero OP strain values. In terms of discretized strains, is the number of vector spin components, that can point in variant directions. A double-well Landau free energy for a scalar magnetization maps onto (Appendix A) an Ising model with and . We consider four first-order transitions R21 with . The nonzero, unit-magnitude variant vectors point respectively to corners of symmetry-dictated polyhedra with corners, inscribed in a unit circle or unit sphere: a geometrically pleasing sequence of line, triangle, hexagon, and tetrahedron. These transitions are respectively, tetragonal-orthorhombic, cubic-tetragonal, cubic-orthorhombic, and cubic-trigonal.
III.1 Strains and Compatibility:
Strains are symmetric tensors , where the superscript is Transpose. In three spatial dimensions, there are 6 independent Cartesian strains R21 ; R23 . The physical strains are convenient linear combinations: one compressional ; two deviatoric or rectangular , and three shears .
The free energy R21 has a nonlinear Landau term that depends on a subset of these physical strains, as the Order Parameter(s). The remaining non-Order Parameter physical strains enter the free energy as harmonic springs, whose extensions cannot be simply be set equal to zero, as pointed out by Kartha R15 . This is because a local OP-strained unit cell will generate non-OP strains in surrounding unit cells. To maintain lattice integrity all strained unit cells must mutually adapt, to all fit together in a smoothly compatible way, without dislocations.
For electromagnetism, there is a no-monopole Maxwell condition of vanishing divergence of the magnetic induction vector, . For elasticity, there is a no-dislocation St Venant Compatibility condition of a vanishing double curl R15 , of the Cartesian strain tensor. In coordinate and Fourier space,
(5) |
Here for , and lattice constant . There are six differential-equation constraints, that are algebraic equations in Fourier space, of which only three are independent R21 ; R23 . Going to physical strains the three algebraic equations express the non-OP strains in terms of the OP strains. The uniform non-OP strains are not so constrained, and can be freely set to their minimum value of zero.
The harmonic non-OP terms can then be analytically minimized subject to the linear constraints, by direct substitution for non-OP strains or by Lagrange multipliers. This yields an OP-OP effective interaction, with a transition-specific Compatibility Fourier kernel R15 that depends on direction . The kernels all have a prefactor , that vanishes for . The Compatibility kernels are smallest (eg zero) for specific directions , explaining the observed DW orientation along preferred crystallographic directions. The Compatibility potential in coordinate space is an anisotropic powerlaw, with the spatial dimensionality as the fall-off exponent .

III.2 Discrete-strain pseudo-spins:


The Landau free energy functionals for a first order transition can be scaled to be independent or weakly dependent, on material parameters R21 . With physical strains as a vector in OP space, the Landau free energy in the austenite phase always has a turning point at . In the martensite phase, it additionally develops variant minima at .
(6) |
All energies are scaled in the thermodynamic Landau temperature where austenite and martensite free energies cross, so the scaled . Here is the spinodal temperature where the austenite minimum vanishes, so uniform bulk austenite becomes unstable for .
The local vector OP can be written as a product of magnitude and direction . The directions of variant or ‘spin’ vectors identify the degenerate variants on either side of a Domain Wall (DW), with all having unit magnitude, . Since austenite is always a Landau turning point, and in any case austenite could be induced at any by local stresses, we always also include the austenite origin point as an allowed value R21 ; R22 ; R24 .
The strain magnitudes are flat, deep into domains on either side of narrow Domain Walls that are zeros of the OP. The local strain magnitude is set equal to the uniform Landau mean value R21 , , so components are approximated as
(7) |
Substituting into the variational free energy density R21 with Landau,Ginzburg, and Compatibility terms , generates a -dependent effective spin Hamiltonian , with the same three terms, inheriting material-specific parameters such as . Each of the discretized-strain clock-like Hamiltonians have been systematically derived R21 from continuous-strain free energies. They are bilinear in the spins, and encode the crystal symmetries, strain nonlinearities, and Compatibility constraints.
The DW Hamiltonian, with an energy scale (in units of ), is
(8) |
Notice has an inherent separation of time scales, with the magnitude responding immediately to quenches in a single time-step, while the more sluggish DW adjustments of can take hundreds or thousands of MC time-steps.
With , the Landau term is
(9) |
The Landau free energy density is
(10) |
defining a factor , that vanishes at the Landau transition temperature , and is negative below it.
For a uniform variant a constant vector, or in Fourier space , there is a vanishing of the Ginzburg term , and of the Compatibility kernel . The uniform (Landau) free energy then sets a lower bound to the energy, , where the martensite fraction is
(11) |
so or for uniform martensite or austenite.
As a 2D illustration R24 ; R25 ; R26 , the square-rectangle OP is a scalar, so . There are two variants (rectangles along either or axes), so . The Landau free energy is a triple well in the OP strains. For , the three well depths at are degenerate at zero. For the austenite well at is metastable, and goes unstable at , the spinodal temperature. The Hamiltonian is diagonal in space, , where . The energy spectrum for long wavelengths is . The square-rectangle transition kernel depends on direction , or the single polar angle , as
(12) |
where is the ratio of a non-OP (shear) elastic constant, and the non-OP compressional . The (positive) kernel has a maximum value at , and a minimum value at , driving a preferred DW orientation along both diagonals.
The energy spectrum for is a parabola pulled down to negative values by the Landau term, . A zero energy contour in space is a circle with a T-dependent radius , that shrinks to a point at . For , the bottleneck becomes angularly modulated, with a squared-radius , interpolating between a outer radius and a inner radius . The inner radius clearly vanishes at some temperature where . This characteristic temperature, from an interplay between Landau, Ginzburg, and Compatibility terms, is where the entropy barrier diverges.
Planes et al R36 consider a uniform-martensite model with a Landau variational term . Fast or slow behaviour is through first-passage-time jumps crossing an energy barrier, that collapses at , or is largest at . Our spatially non-uniform martensite model with Ginzburg, Compatibility and Landau variational terms, differs in detail, but is similar in spirit. Fast or slow behaviour is through MC searches crossing an entropy barrier, that collapses at and diverges at .
III.3 DW Hamiltonians for four transitions :
Clock models have discrete spins directed at points on a unit circle, and are denoted by , where the Ising model is . Here we generalize to include , and call these ‘clock-zero’ models, denoted by .
Drawing on Equ (7), the generic coordinate-space Hamiltonian is
(13) |
where the overall energy scale is . Here is the domain-wall thickness parameter, is the elastic constant for the non-OP compressional strain R24 ; R25 ; R26 . The kernel is an matrix potential, that carries the spatial dimensionality , and depends on ratios of other non-OP elastic constants to . Local meanfield treatmentsR22 yield even the complex strain textures of some real materials R13 ; R14 .
The generic space Hamiltonian is obtained from , and as is real, . The Hamiltonian and energy-spectrum are
(14) |

,
The 3D transitions are given below in the sequence of , plus cases respectively.
a) Tetragonal-Orthorhombic case ():
The scalar OP is the first deviatoric-strain OP , with having values: at the origin, and pointing to the two endpoints of a unit-circle diameter. The Hamiltonian is like a 3D Spin-1 Blume Capel model R24 , but with anisotropic powerlaw interactions, and with the quadratic term where .
The Hamiltonian is a clock-zero model R21 .The scalar compatibility kernel for the tetragonal-orthorhombic transition is given in Equ. (A26) of Ref 21.
With , the Tetragonal-orthorhombic (and also 2D square-rectangle) case has the squared-mean OP and factor as R21
(15) |
b) Cubic-tetragonal case ():
This 3D transition has been considered earlierR23 ; R31 .
The OP strains are the two deviatoric strains .
The spin values vectors are in a plane in OP space,
with having values: at the origin and pointing to the three corners of a triangle inscribed in a unit circle.
The Hamiltonian is a clock-zero model R21 .The compatibility kernel is a matrix, , with , as in Equ (A23) of Ref 21.
The mean OP and of the cubic-tetragonal transition are
(16) |
c) Cubic-orthorhombic case ():
The OP strains are again the two deviatoric strains as above.
The nonzero spin vectors are in a plane in OP space,
with having values: at the origin and pointing to the six corners of a hexagon inscribed in a unit circle.
The Hamiltonian is a clock-zero model R21 . The Compatibility kernel is a matrix , again with and is the same as the cubic-tetragonal case, given in Equ (A23) of Ref 21.
The squared-mean OP and for the Cubic-orthorhombic case, are the same as the tetragonal-orthorhombic in Equ (15).
d) Cubic-trigonal case ():
The three OP for the cubic-trigonal transition are the three shears , and the variant vector has
three vector components,
with having values: at the origin and pointing to the four corners of a tetrahedron inscribed in a unit sphere. The Compatibility kernel is now a matrix , with , or six components as in Equ. (A20) of Ref 21.
The mean OP and for the cubic-trigonal case are the same as the cubic-tetragonal case in Equ (16).
IV Monte Carlo Simulations:
Simulations were done on models for four transitions in 3D. The initial state is randomly and dilutely seeded martensite cells, in an austenite sea of . Evolutions proceed at quench temperatures . Typical parameters are ; ; to ; to ; ; ; ; and holding times MC sweeps, each over sites.
The Compatibility kernels arise from the non-OP harmonic terms, with elastic constants. For all transitions, we specify the fixed ratio of other non-OP elastic constants, to . For the tetragonal-orthorhombic transition, with and as the OP spin, the non-OP elastic constants are the other deviatoric constant , and the three shear constants , set to be . Similarly, for the cubic-tetragonal and cubic-orthorhombic cases with and as the OP spins, the non-OP constants are set to be . Finally, for the cubic-trigonal case with and the shears as the OP, the constants are set as .
The standard MC procedure R32 is followed, with an extra data retention R1 ; R2 ; R3 ; R4 of energy changes.
0. Take lattice sites, each with a vector spin of components, in one of possible values (including zero) at MC time . Each set is a ‘configuration’. With or for uniform austenite or martensite, the average martensite fraction is . The conversion time is defined as when R24 .
1. Randomly pick one of sites, and randomly flip the spin on it to a new direction/value, and find the (positive or negative) energy change .
2. If , then accept the flip. If , then accept flip with probability . Record this spin-flip , that is usually not retained after use.
3. Repeat steps 1 and 2. Stop after such spin-flips. This is the configuration with .
4. We collect all recorded over each MC sweep of every run while tracking . The collection is only up to a waiting time equal to the martensite conversion time.
We do six quenches, from upwards to .
A single-variant athermal martensite droplet or ‘embryo’ R11 can form anywhere, and after a local conversion at a waiting time , can propagate rapidly to the rest of the system R11 ; R19 . Hence it is the mean rates (or inverse times), that are averaged over runs, analogous to total resistors in parallel determined by the smallest resistance. The mean times are defined as the inverse mean ratesR24 .
(17) |
For an un-successful -th run that does not convert in a holding time , the rate could be assigned a value of either (conversion right after cutoff) or (conversion never occurs). We choose the cutoff, that shows up as a flattening of the mean rate near . An extrapolation of the linear part of to the temperature axis yields a valueR24 for .


V Delays and bottlenecks:
For the four transitions, we consider i) Delays from sluggish Domain Walls; ii) Bottlenecks in Fourier space.
V.1 Delays from sluggish Domain Walls
Figure 2 shows the twinned microstructures in all four transitions, colour coded through the variant-label , where is always austenite. For the single-OP tetragonal-orthorhombic transition, has two variant colours. For the two-component OP, the cubic-tetragonal and cubic-orthorhombic transitions have respectively, and variant colours. For the three-component OP, the cubic trigonal has four variant colours. Fig 2 shows that all allowed degenerate variants are present, for all four transitions.
The tetragonal-orthorhombic, and cubic-tetragonal twins can have Domain Walls decorated with austenite, as also found in the 2D case R25 . Such observed austenite retentionsR11 can be understood: they are energetically favoured, when the lower Ginzburg costs of austenite-martensite walls compensate for the absence of the negative condensation energy of martensite unit cells. As is lowered, the energy accounting is reversed, and the austenite inclusions are expelled, replaced by martensite, so DW are only between martensite variants R25 .
We define a fractional deviation from a characteristic temperature , as
(18) |
Figure 3 shows the martensite conversion-fraction versus MC time in a single run, for different . For quenches to below , there is an immediate avalanche conversion in a single sweep, characteristic of athermal martensite. For temperatures there is a strange ‘incubation’ behaviour, or a postponement of these avalanches. The fraction remains virtually unchanged, up till when rises sharply through , to unity. The cubic-trigonal transition has order parameter components, variants, and can show unusual behaviour. Here, there is an initial jump to followed by incubation, and then a jump to unity.
Fig 4 shows that for there is an immediate spike in the MC acceptance fraction at , and drop in energy to negative values. For , during incubations they both remain zero, up to , when the acceptance spikes and the energy drops. Again, the cubic-trigonal case is unusual.
What goes on microscopically, during incubation ? In 2D, Video R27 A shows random initial seeds of both variants (red,blue) can quickly form an almost zero energy single-variant martensitic droplet or embryo (red) in an austenite background (green), in the DW Vapour state. The small droplet extends and retracts amoeba-like arms, searching for energy-lowering pathways. After a long conversion delay of hundreds of time steps, when , the single-variant droplet (red) suddenly expands rapidly and generates the opposite variant (blue). This is the wandering-wall or DW Liquid state. After a shorter orientation delay, the walls of the DW Liquid orient to a DW crystal.
Figure 5 shows for all four transitions, the mean fraction of successful conversions during a holding time over 100 runs, versus the temperature deviation . For temperatures , every run converts, and . However, for , falls through 1/2 at , and then to zero. The success fraction is not exponentially sensitive to overall energy scales , so the probability of conversion is not activated over an energy barrier.
Just above there are unsuccessful runs (not shown), when the martensite seeds can dissolve back into austenite. The seeds will not be regenerated, even if the holding time is increased, or if the temperature is loweredR20 .
Figure 6 shows a linear-linear plot of the mean conversion times versus for various . The absence of exponential sensitivity to implies the delays are not activated over energy barriers, but are due to entropy barriers. For fluctuations are small, while for standard-deviation error bars over the runs, are larger, on approaching .
If the probability of entropy-barrier crossings coming from a product of sequential random steps, a logarithm of rates would be an additive random variable, suggesting a log-normal distribution of rates R25 . Fig 7 shows optimized-bin histograms R37 of data, with (asymmetric) log-normal lines as guides to the eye. Although data are too sparse to decide distributions, clearly fast conversions are narrow, and slow conversions are broad, similar to protein folding R29 ; R30 .
Fig 8 shows an Angell-type plot R8 of log delay time versus . Arrhenius-type activations over energy barriers would be linear. The delays have curvatures, and so must be from entropy barriers.The solid curve is VFT behaviour of Equ (29) below, with .
V.2 Bottlenecks in Fourier-space:
A purely random search of protein configurations could take astronomical times (Leventhal paradox). Rapid protein folding is attributed to a configuration-space Golf Hole (GH) opening into a Funnel of negative-energy configurations, leading rapidly to the folded protein state R28 ; R29 . Bicout and Szabo R30 consider a random walk of a Brownian particle in a space of eigen-labels of protein folding modes (that could be analogous to propagative martensitic twinning modesR38 ). The Brownian particle has to locate and enter spherical zero-energy GH contour of marginal modes. Unusual delays can occur, at the GH edge.



In our case, the bottleneck is fixed by the energy spectrum from Equ (14) in 3D Fourier space (taken as diagonal in , for discussion):
(19) |
Energy spectra set to a constant or , define contours in -space. The Ginzburg term at long wavelengths , forms a -space spherical bowl in 3D, with zero-energy minimum at the origin. The Compatibility term angularly modulates its 2D surface to produce an anisotropic zero-energy contour R24 .
The phase-space boundary separates an outside region of positive (austenite) energies, from a -region inside, of (martensitic) negative energies. Video B shows R27 a 2D circular droplet Fourier profile in as it distorts, to enter the phase space bottleneck.
We consider spectra with for the first three transitions, and for the cubic-trigonal case. Consistent with the twin orientations in Fig 1, the plane intersecting the 3D bottleneck to yield a 2D cross-section is taken as . The plane through the Brillouin Zone origin is , and Fig 9 shows for all four transitions, the -dependent contours of constant versus ( for temperature range to .As mentioned, the cubic-tetragonal and cubic-orthorhombic have the same kernel but different Landau factor , so one would expect the second and third panels to show the same overall shapes, but slightly different energy contours for a given : this is indeed the case.
The bottleneck sizes are large at low and small at high . The contours are angularly modulated between a smaller inner-radius wave-vector and larger outer-radius wave-vector . From the spectrum Equ (19)
(20) |
For , the outer square-radius vanishes at the Landau temperature, . Close to transition,
(21) |
where the Taylor expansion coefficient .
With a positive there is a temperature where the inner radius can pinch off to zero, , while the outer radius is still nonzero. Near ,
(22) |
Figure 10 shows the inner and outer squared-radii, both almost linear, and vanishing respectively at and .
The conversion-delay divergence comes from a pinch-off of the inner radius of the bottleneck. The topology of a 2D slice of the 3D negative energy states, goes from an open butterfly to a segmented four-petalled flower R25 . It is impossible for the broad Fourier profile of a small droplet to distort at zero total energy, into four separated petal-like segments. The lower-energy states for are thus available, but not accessible. The intershell configurational pathway closes; the success-fraction vanishes; and the entropy barrier diverges.
VI PES signatures in all four transitions
We now exhibit PES signatures in the four athermal martensite transitions. Entropy barriers are insensitive to energy scales, so we consider only .


As noted the MC procedures of Section IV retain the set of single spin-flip energy changes from all N spin-flips, in each of runs, up to t MC times . The histograms can be dense since the data set size can be large: has up to data points.
An ageing-state fluctuation relation is postulated R4 . The probability to hit configurations from , is proportional to the target size, or the number of accessible states: , with , where . The reverse path has . The ratio of forward and backward probabilities is related to the entropy change and entropy barrier . Thus
(23) |
Fig 11 shows four-panel log-linear plots of the fluctuation ratio . Since , the entropy change is odd, , and a solution is
(24) |
where the (even) prefactor is the geometric mean, . For small energy changes, , and hence .
Fig 12 is just a check that the prefactor versus has no linear contribution near the origin, that might modify the exponential tail. For temperatures near it is a single-peak gaussian, while near it can go bimodal.
Fig 13 shows that the variance of the weight versus a scaled is linear and nonsingular, for all four transitions.
Fig 14 shows four-panel linear-linear plots of (normalized) versus for four transitions, each at one of six temperatures in the Legend. The peak is at positive energy as in the oscillator case R2 , and moves left as decreases towards . The exponential tails near the origin are barely visible.
Fig 15 shows the same four plots but now in log-linear form, and zoomed in. The curves all show the PES signature of linearity around the origin . The cubic tetragonal panel has been shown earlier R31 . As is lowered to , the slopes all flatten.


Fig 16 shows a central result, namely the search temperature and its inverse, versus . The left-axis search temperature intersects the temperature axis at an extrapolated , defining a quench temperature . The vanishing appears to be linear, .
The mean conversion rate involves an integral over the heat releases of the distribution R39 . The mean time is then a singular exponential . The Vogel Fulcher-Tammann form R5 ; R7 ; R8 thus emerges naturally from a search temperature freezing inducing a rapid arrest of PES cooling, and an entropy-barrier divergence.
Similarly, the inverse effective temperature on the right-axis of Fig 16 seems to go to zero linearly at a search explosion where entropy barriers vanish.
Once again, the cubic-trigonal last panel is unusual, with showing a smaller slope near . If the linear slope actually vanishes as then that would yield ‘super-VFT’ behaviour R5 , .
Fig 17 shows the scaled ratio of occurrence probability to its value at the origin, versus . The dashed white lines have slopes close to the predicted universal slope of unity, as given in the Figure caption.
The normalized probability of a Monte Carlo spin-flip is the product of the occurrence probability and an MC acceptance factor with step functions,
(25) |
with a normalization constant. The ratio of the MC probability and its value at the origin defines . With ,
(26) |
Fig 18 shows a log-linear plot of along the positive axis . The closeness of data to theoretical lines with slope is further evidence for PES.

In conclusion all four transitions show PES signatures, lending support to the Partial Equilibration Scenario.
VII PES delays in simulations and experiment
We show that delay data both in simulations and in experiment, are consistent with a picture of diverging entropy barriers from linearly vanishing PES effective temperatures.
VII.1 VFT delays in 3D simulations

Vogel-Fulcher-Tamman temperature dependences can be written near as
(27) |
where are the time and energy scales, for DW shifts of a lattice spacing.
The logarithms of VFT times can be written in two useful forms, to extract constants from simulations and experiments. Thus
(28) |
and
(29) |
For simulations, and can be extracted from data using Equ (28) and Equ (29). Fig 19 shows for all four transitions, the data in a scaled form of
(30) |
There is data clustering around the Vogel-Fulcher straight line showing universality over 3 orders of magnitude near . There is also a peel-off toward shorter times, near (downward arrow).
VII.2 VFT delays in martensitic alloys
Delayed athermal martensitic transformations in metallic alloys have been tracked by conversion diagnostics suited to the scale of the waiting timesR16 ; R17 ; R18 such as electrical resistivity drops; or surface optical or X-ray reflectivity. In pioneering experiments, Kakeshita et al R17 used resistivity drops to detect the austenite to martensite delayed conversions for alloys . The alloying percentage is , with start temperatures of K.The delays discovered were of macroscopically long times. Other work such as the Klemradt groupR16 ; R18 on alloy delays, also rose rapidly: for temperature increments above the values of K, K and , the delay times went from R18 several seconds, to seconds, to forever.
The data analysis of simulations is used again for experiment, as now described in more detail. To extract divergence temperatures from simulation or experimental data, we plot versus , and extrapolateR24 straight line segments to the x-axis (not shown). For FeNi dataR17 this yields K, well above , with large fractional delay windows . Similarly extrapolation of data for the NiAl alloyR18 with K yields but with a smaller window .
Another group considered NiTi alloys R20 , and argued that if the largest delay is at , then for a long enough annealing time, conversions should be seen at any in a wide window . However no conversion was detected, for holding at above , for days. The absence of conversion was attributed to a sparseness of (atomic) catalyst fields in facilitation-type delay models R20 . In our picture this absence could also be due to being outside the narrower conversion window ,
Fig 20 shows measurement data of conversion times , in seconds, versus temperature in degrees Kelvin. The left column shows a linear-linear plot like Equ (28) of versus to extract slope from Kakeshita data R17 for three alloys of FeNi (top panel); and from Klemradt dataR18 for a NiAl alloy (bottom panel). The right column shows a linear-linear plot Equ (29) of versus , using lines of the extracted slope , to determine the intercept for FeNi (top panel) and for NiAl (bottom panel). Downward arrow marks for NiAl data. The ‘fragility’ R8 parameter , and a basic time scale for DW hops is second.


Fig 21 shows that combined experimental data R16 ; R17 ; R18 cluster around the Vogel-Fulcher straight line of Equ (30), with universality over 3 orders of magnitude near . For NiAl dataR18 , there is a linear falloff on approaching (downward arrow), consistent with Fig 16.
It would be interesting to get more data for these and other martensitic alloys, through systematic quenches in steps of , over the entire delay range between barrier divergence and collapse. It would also be interesting to include a -like onset of sluggishness, in fitting analyses of glass-former viscosity data R7
VIII Summary and further work
In this paper, we present Monte Carlo (MC) simulations, on discrete-strain Hamiltonians for four 3D structural transitions, under systematic temperature quenches from seeded austenite, to study austenite-to-martensite conversion times. The results and scenario are as follows.
For athermal martensites, there are explosive conversions below a martensite start temperature so there are no barriers. Above this start temperature, entropy barriers emerge , and incubation time delays rise sharply towards a divergence temperature . The entropy barrier collapse/ divergence, is understood through temperature-controlled phase-space bottlenecks.
Partial equilibration ideas provide an understanding of fast/ slow times, based on effective temperatures for energy-lowering searches. The inverse search temperature vanishes linearly at or . The search temperature vanishes linearly at , or . This rapid search arrest explains the singular Vogel-Fulcher-Tammann form, extracted from martensitic experimental data.
Further simulations of crystallisation models R6 could try to record heat releases. Further experiments could record strain signals and intermittencyR40 over the delay region ; and over the tweed precursor R14 ; R25 ; R27 region above it . Non-stationary distributions of energy changes in martensites could be measured through concurrent acoustic, photonic or strain probesR40 . The VFT temperature regime in glasses shows non-Debye frequency responses R9 : this might be more general. Finally, quenches of complex oxides R41 near their structural / functional transitions, might yield interesting PES signatures in functional variables, induced by their coupling to ageing strain domains.
Acknowledgements:
It is a pleasure to thank Turab Lookman for helpful early conversations; and Smarajit Karmakar for valuable discussions on the glass transition.
Appendix A: PES signatures in the model:
As a toy model illustration of the Section III procedure we consider a 2D magnetization free energy with double-well Landau term and a Ginzburg term, , Here is the single component OP, of both signs, so . The Landau term is
(31) |
where , with all energies/ temperatures scaled in the physical transition temperature, so the scaled . The Ginzburg term is .

.
Domain-walls are solitonic solutions with a profile, interpolating between flat OP variants of opposite sign, within a DW thickness . The OP can be written as a magnitude times a variant ‘spin’ of unit length. The nearly flat magnitude in the domains is approximated by the mean field OP,
(32) |
where . Compare Equ (7).
The mean-field Landau free energy is
(33) |
where . Compare Equ (10).
Substituting Equ (32) in Equ (31) yields the DW coordinate space Hamiltonian, with ,
(34) |
although in this case, at all sites. With , the Hamiltonian in Fourier space is
(35) |
Compare Equ (13) and Equ (14).
We can do MC simulations with this T-dependent, Ising-variant effective Hamiltonian. The parameters are , with holding times . The spin-flips are . Fig 22a shows the magnetization fraction versus waiting time analogous to Fig 3, but here with a gradual rise, and no incubation behaviour.
We record for every spin flip up to an OP marker event time when . For nearest-neighbour couplings on a square lattice, the energy changes will be discrete. Fig 22b shows the log-linear versus discrete energy changes . The spike heights decrease linearly with energy changes, consistent with PES. Compare Fig 15.

Appendix B: Athermal phase diagram:

Fig 23 shows data for the martensitic tetragonal-orthorhombic transition, for several , showing both athermal and non-athermal or mixed behaviour R11 ; R23 ; R24 ; R25 . For the athermal regime, Temperature-Time-Transformation (TTT) curves have flat lines along the temperature axis where there are immediate, explosive conversions for . The embryo or droplet is small in coordinate space, and so is broad and flat in Fourier space. With approximated by a constant, the Hamiltonian energy of Equ (18) is then involves averages of terms over the Brillouin Zone, denoted by square brackets. The droplet energy is
(36) |
The energy vanishing defines . For small values in a Taylor expansion, and with coefficient as defined in the text,
(37) |
We define ‘athermal’ behaviour as a nonzero ; explosive conversions for ; and incubation delays in the window . The behaviour not precisely athermal, is called ‘mixed’, with conversions occurring gradually, without flat incubations. Vanishing of the athermal case start temperature then determines a phase boundary.
The four-panel Fig 24 plot of versus , shows the phase boundary. Above the phase boundary the system is purely athermal, while below the phase boundary where , is a mixed regime. Data from Fig 23 and other TTT diagrams are seen to be consistent with the theoretical phase boundary. Fig 23 shows that for some parameters, curves on the upper left, have a fall and then a rise with temperature, like a ‘U shape, or downward ‘nose’ . For the large used, the shape is distorted, but for smaller the U shape is more well defined R26 . The shape is from an Arrhenius activation over a temperature-dependent energy barrierR42 .
References
-
(1)
F. Ritort in Unifying Concepts in Granular Media and Glasses, (eds.) A. Coniglio, A. Fierro, H.J. Herrmann and M. Nicodemi, Elsevier B.V. (2004);
arXiv/ condmat/ 03113710, v1. -
(2)
L. L. Bonilla, F.G. Padilla and F. Ritort, Physica A 250, 315 (1998);
F. Ritort, J. Phys. Chem. B 108, 6893 (2004);
L. Garriga and F. Ritort, Phys. Rev. E 72, 031505 (2005). -
(3)
A. Crisanti and F. Ritort, Europhys. Lett. 66,253 (2004);
L. Buisson, L. Bellon and S. Ciliberto, J. Phys. C, 15, S1163 (2003). - (4) A. Crisanti, M. Picco and F. Ritort, Phys. Rev. Lett. 110, 080601 (2013).
- (5) K. Binder, W. Kob, Glassy Materials and Disordered Solids: An Introduction to Their Statistical Mechanics, World Scientific, Singapore (2005).
-
(6)
H. Fang, M.F. Hagan, W.B. Rogers, Proc. Nat. Acad. Sci., 117, 27927 (2020);
X- Yang, H. Tong, W-H. Wang, and K. Chen, Phys Rev E, 99, 062610 (2019);
H- Shintani and H. Tanaka, Nature Physics 2, 200 (2006). - (7) P. Lunkenheimer, S. Kastner, M. Koehler and A. Loidl, Phys. Rev. E 81,051504 (2011).
- (8) Fragility of glass-forming liquids ed. A.L. Greer, K.F. Kelton, S. Sastry, Hindustan Book Agency, New Delhi (2014).
- (9) R. Boehmer, K.L. Ngai, C.A. Angell, D.J. Plazek, J. Chem. Phys., 99, 4201 (1993); T.V. Ramakrishnan and M. Rajlalakshmi, Eds. Non-Debye Relaxation in Condensed Matter, World Scientific, Singapore (1987).
-
(10)
K. Bhattacharya, Microstructure of Martensite, Oxford University Press (2003),
EKH Salje, Phase Transitions in Ferroelastic and co-Elastic Crystals Cambridge (1993). -
(11)
J.R.C. Guimaraes and P.R. Rios, J. Mater. Sci.,43, 5206 (2008);
M. Rao and S. Sengupta, Current Science (Bangalore) 77, 382 (1999);
Physical properties of martensite and bainite: Proceedings of the joint conference, the British Iron and Steel Research Association and the Iron and Steel Institute, Special report 93, London (1965);
A. R. Entwistle, Metall. Trans. 2, 2395 (1971);
SR Pati and M Cohen, Acta Metall., 17, 189 (1960);
G.V. Kurdjumov and O.P. Maximova, Doklad. Akad. Nauk USSR 61, 83 (1948); 73, 95 (1950). - (12) G.R. Barsch and J.A. Krumhansl, Phys. Rev. Lett. 53, 1069 (1984).
-
(13)
J.W. Seo and D. Schryvers, Acta Mater., 46, 1165 (1998);
C. Manolikas and S. Amelinckx, Phys. Stat. Sol. A 61,179 (1980). - (14) L.E. Tanner, Phil. Mag 14,111 (1966).
-
(15)
S. Kartha, J. A. Krumhansl, J. P. Sethna, and L. K. Wickham, Phys. Rev. B 52, 803 (1995);
M. Baus and R. Lovett, Phys. Rev. A 44, 1211 (1991);
S.R. Shenoy, T. Lookman, A. Saxena and A.R. Bishop, Phys. Rev. B 60, R12537 (1999). - (16) L. Müller, U. Klemradt, T.R. Finlayson, Mat. Sci. and Eng. A 438, 122 (2006).
-
(17)
T. Kakeshita, T. Fukuda and T. Saburi, Scripta Mat. 34, 1 (1996).
T. Kakeshita, K. Kuroiwa, K. Shimizu, T. Ikeda, A. Yamagishi, and M. Date, Materials Transactions, JIM, 34, 423 (1993). -
(18)
M. Aspelmeyer, U Klemradt, L.T. Wood and S.C. Moss, Phys. Stat. Sol. (a) 174, R9 (1999);
U. Klemradt, M.Aspelmeyer, H. Abe, L.T. Wood, S.C. Moss and E. Dimasi, J.Peisl, Mat. Res. Soc. Symp. Proc. 580, 293 (2000);
L. Mueller, M. Waldorf, C. Gutt, G. Gruebel, A. Madsen, T.R. Finlayson, and U. Klemradt, Phys. Rev. Lett. 107, 105701 (2011). -
(19)
P.R. Rios and J.R.C. Guimaraes, Scripta Mat., 57, 1105 (2007);
P.R. Rios, F.G. Cardoso, T.A. Neves, J.R.C. Guimaraes, in The Minerals, Metals and Materials Society (eds) TMS 2015 144th Annual Meeting and Exhibition, Springer, Cham (2015),
https://doi.org/10.1007/978-3-319-48127-2_84. -
(20)
K. Otsuka, X. Ren and T. Takeda, Scripta Mat. 45,145 (2001).
K. Otsuka and X. Ren, Scripta Mater., 50, 207 (20004);
T. Okuzuno, Y. Yamazaki and T. Ohta, Phys. Rev. B, 67, 654106 (2005). - (21) S.R. Shenoy, T. Lookman and A. Saxena, Phys. Rev. B 82, 144103 (2010).
- (22) R. Vasseur , T. Lookman and S. R. Shenoy, Phys. Rev. B 82, 094118 (2010).
-
(23)
K.O. Rasmussen, T. Lookman, A. Saxena, A.R. Bishop, R.C. Albers and S.R. Shenoy, Phys. Rev. Lett. 87 (2001);
SR Shenoy, T Lookman and A Saxena in Magnetism and structure in functional materials,Eds. A Planes and A Saxena Springer-Verlag, Berlin (2005). - (24) N. Shankaraiah, K. P. N. Murthy, T. Lookman and S. R. Shenoy, Europhys. Lett. 92, 36002 (2010).
-
(25)
N. Shankaraiah, K.P.N. Murthy, T. Lookman and S.R. Shenoy, J. of Alloys Compds 577, S66 (2013);
Phys. Rev. B 84, 064119 (2011);
Phys Rev. B 91, 214108 (2015). - (26) N. Shankaraiah, Phys. Rev. E 95, 063003 (2017); J. Alloys Compds. 675, 211 (2016).
-
(27)
Videos for the 2D square-rectangle transition, with austenite (green) and the two martensite variants (red/ blue), showR24 ; R25 sequentially evolving Domain Wall phases, from DW Vapour to DW Liquid to DW Crystal or oriented ‘twins’. The 3D evolution is similar, but we present the earlierR24 ; R25 2D system as illustrations.
Video A. The droplet in -space: This shows the coordinate-space evolution for , for the Golf Hole type constrained passage of Fig 1. Random initial martensite seeds rapidly form a single-variant DW Vapour droplet. The droplet fluctuates while slowly searching for a conversion passage to a bidiagonal DW Liquid. There is then a quick symmetry-breaking to single-diagonal DW crystal.
Video B. The droplet in -space: This shows for , the evolution as in Video A but now in Fourier space.
Videos C,D are in different quench regimes than A,B but are also given for completeness.
Video C. The -space dynamic catalysts: This shows the evolution for deep quenches , of the facilitation type constrained passage. The DW Vapour now converts rapidly to a DW Liquid of sluggish walls. Sparse transient hotspots of austenite in martensite act as dynamic catalysts, that non-locally delete even far-off minority-diagonal DW segments, leaving a single-diagonal DW crystal.
Video D. The -space tweed precursor: This shows dynamical tweed in the region , as an array of vibrating martensitic islands in an austenite sea, possibly induced by a space profile, vibrating on top of a topologically blocked bottleneck. - (28) P.G. Wolynes, Proc. of the Am. Phil. Society 145, 4 (2001).
- (29) M. Cieplak and I. Sulkowska, J. Chem. Phy. 123, 194908 (2005).
- (30) D. J. Bicout and A. Szabo, Protein Science 9, 452 (2000).
- (31) N. Shankaraiah, K.P.N. Murthy and S.R. Shenoy, https://arxiv.org/pdf/2111.04290.pdf
- (32) KPN Murthy, Monte Carlo Methods in Statistical Physics, Universities Press, Hyderabad (2004).
- (33) K Huang, Statistical Mechanics 2nd edition Wiley, NY (1987).
- (34) P. Sollich and F. Ritort, Adv Phys., 52, 219 (2003).
- (35) S.R. Shenoy and T. Lookman, Phys. Rev. B., 78, 144103 (2008). See Fig 7.
- (36) A. Planes, F.-J. Perez-Roche, E. Vives and L. Manosa, Scr. Mater., 50 181 (2004).
- (37) D.W. Scott, Biometrika 66, 605 (1979).
- (38) G.S. Bales and R.J. Gooding, Phys. Rev. Lett., 67, 34412 (1991).
- (39) The mean equilibration rate is the average over an inter-shell rate over all heat releases. With a normalized PES probability, . For slow intershell variation , the mean rate is proportional to the acceptance fraction over energy decrements. Taking the PES probability, as a shifted gaussian with peak mean energy change , and variance , the acceptance fraction is a complementary error function, whose asymptotic behaviour yields . Dropping constants, the martensitic mean conversion time near is written simply as .
-
(40)
U. Chandni, S Kar-Narayan, A. Ghosh, H.S. Vijaya and S. Mohan, Acta Mater., 57,6113 (2009);
A. Planes, Ll. Manosa and E. Vives, J. Alloys and Comp., 577, S699 (2013);
Z. Eranson, B. Ruta, S. Hechler, M. Stolpe, E. Pineda, I. Gallimo, and R. Busch, Phys. Rev. Lett., 115, 175701 (2015).See also Physics 8, S5121 (2015);
X. Balandraud, N. Barrara, P. Biscarin, M. Grediac, G. Zanzotto, Phys. Rev. B, 91,174111 (2015);
B. Blaysat, X. Balandraud, M. Grediac, E. Vives, N. Barrera and G. Zanzotto, Nature Commun. Mater., 1,3 (2020). -
(41)
W.I.F. David,C.C. Wilson, P. P. Edwards, R. Jones, M.R. Harrison, Nature 331, 245 (1988);
M. Uehara and S-W. Cheong. Europhys. Lett., 52, 674 (2000);
V. Podzorov, B.G. Kim, V. Kiryukin, M.E. Gershenson and S.W. Cheong, Phys. Rev. B 64, R140406 (2001);
V. Podzorov, C.H. Chen, M.E. Gershenson and S.W. Cheong, Europhys. Lett., 55, 411 (2001). V. Hardy, A. Maignan, S. Hebert, C. Yaicle, C. Martin, M. Hervieu, M.R. Lees, G. Rowlands, D. McK. Paul and B. Raveau, Phys. Rev. B 68, R220402 (2003);
V. Hardy, S. Majumdar, S.J. Crowe, M¿R. Lee, D. McK. Paul, L. Herve’, A. Maignan, S. Hebert, C. Martin, C. Yaicle, M. Hervieu and B. Raveau, Phys. Rev. B 69, R020407 (2004). - (42) Activated times over rising energy barriers diverge at . The times fall for increasing from zero, have a minimum, and then rise exponentially at large . The minimum in is at . Close to , the TTT curve has an ‘isothermal’ U-shape or nose: .