Spin-crossover (SC) compounds are fascinating materials that exhibit colorful phase transitions induced by temperature, pressure, photoirradiation, etc. From the microscopic point of view, the electronic (spin) state of a molecule changes between the low-spin (LS) and high-spin (HS) states by such stimuli, and a variation of the molecular size follows through the vibronic coupling between the electronic state and structure in the molecule. This causes an elastic distortion and then an elastic interaction. The elastic interaction is essential in cooperative properties of SC phenomena. In this paper, we present a tutorial study on elastic interaction models for SC phenomena, which are the recent trend of modeling of SC compounds. We focus on multistep transitions, which are a topic of SC phenomena. We analyze the phase diagrams including the metastable phases for several SC systems, in which antiferromagnetic-like and ferrimagnetic-like phases are associated in addition to the LS and HS phases. Making use of the phase diagrams, we show various patterns of thermally induced SC transition with steps. We also investigate SC transitions with steps in a core-shell SC nanocomposite composed of two different SC compounds. We focus on two cases: the core has a lower transition temperature than the shell, and the core has a higher transition temperature. We show characteristic features of difference in the two systems.

Spin-crossover (SC) compounds exhibit a variety of phase transitions induced by temperature, pressure, light irradiation, etc.1–21 Their switchable magnetic, optical, and structural properties have drawn much attention in their potential applications to memory devices and sensors. In SC phenomena, transitions between the low-spin (LS) and high-spin (HS) states occur, and a wide variety of transition patterns are realized.

With increasing and decreasing temperature, various SC compounds show one-step SC transitions between the HS and LS phases. Their temperature dependence of the HS fraction, f HS, shows a smooth single, discontinuous, or hysteresis curve. In the smooth single curve, the transition is a second-order transition or a mere change (not phase transition). In the discontinuous or hysteresis curve, a first-order transition occurs. In some SC compounds, besides the LS and HS phases, another phase with some pattern in alignment of LS and HS molecules is realized as a thermodynamics phase. An antiferromagnetic (AF)-like phase is a typical one, in which LS and HS molecules align alternately, and this is realized as an intermediate-temperature phase in two-step SC transitions.22–31 

In recent experiments, more complex SC transition patterns were observed in three-step transitions32–35 and more than three-step transitions.36–40 A variety of ordered phases such as a LS–LS–HS–HS layered phase41 and commensurate and incommensurate stripe-ordered phases14 were found. With the development of nanotechnologies, SC nanoparticles and nanocomposites have also attracted much attention because of controllable finite-size effects.42–47 Among them, recently, a core-shell nanocomposite made up of two different SC compounds48 has been synthesized, which shows SC transitions with steps. Their synergy or competing effects owing to the combination of different SC compounds with finite-size effects are fascinating. Multistep SC phenomena with complex orderings are a current topic, and metastability or multistability is a key to understand these phenomena. In this paper, we present a tutorial study for modeling and simulation of SC materials and show that various patterns of SC transitions with steps are realized.

From the viewpoint of theoretical studies of SC phenomena, Ising-like models have been intensively studied, and several important properties of SC phenomena have been derived.28–30,49–52 However, the Ising interaction was introduced phenomenologically for simplicity, and the origin of the cooperative interactions has been ignored. On the atomistic scale, the molecular volume (or size) varies between the LS and HS states (Fig. 1) through the vibronic coupling between the electronic state and the structure, and the difference in the molecular size induces lattice distortion, which causes an elastic interaction between SC molecules (Fig. 2). Thus, volume expansion or contraction of SC crystals is accompanied with SC transitions. This elastic interaction is an essential ingredient in the cooperative character of SC materials.

FIG. 1.

Intramolecular potential as a function of the molecular radius (metal–ligand distance). The HS molecule is larger than the LS molecule.

FIG. 1.

Intramolecular potential as a function of the molecular radius (metal–ligand distance). The HS molecule is larger than the LS molecule.

Close modal
FIG. 2.

Modeling of SC molecules. Red and blue molecules are HS and LS ones, respectively. Each molecule changes its size depending on the molecular state and varies its position by surrounding elastic interactions.

FIG. 2.

Modeling of SC molecules. Red and blue molecules are HS and LS ones, respectively. Each molecule changes its size depending on the molecular state and varies its position by surrounding elastic interactions.

Close modal

In recent years, modeling of the cooperative interaction from the microscopic viewpoint has been focused on, and a new class of SC models based on cooperative elastic interactions, in which variation of the molecular size is taken into account, has been developed.53–82 For the metal center whose spin switches between S = 0 and S = 2 in the case of Fe(II), for example, the molecular size (metal–ligand distance) changes depending on the electronic state as shown in Fig. 1. The cooperative nature of the elastic interaction has been intensively studied using this class of models, and important features of the cooperativity in SC phenomena have been found. Unlike the Ising-like models, in this modeling, the elastic interaction generates a long-range (LR) interaction. This LR interaction is important for the cooperativity as well as short-range (SR) interactions, and the interplay between the SR and LR interactions, which leads to frustration and synergy effect in SC ordering, is also significant especially for metastable and multistable states, leading to two-step and multi-step SC transitions.68,79,81,83–86 Interestingly, the importance of the LR interaction depends on the type of SC transitions. The LR interaction of elastic origin is relevant in first-order transitions79,81 and a one-step second-order (continuous) transition,55,65,68 while it is not relevant for second-order (continuous) transitions through the AF-like79 or ferrimagnetic (FR)-like phases.81 The dynamics of phase transitions in the elastic interaction models is also characteristic and different from that of Ising-like models. In one-step transitions between the LS and HS phases, clustering in the bulk is suppressed, while clustering occurs from the corners (less probability from the sides) and does not occur from the inside.61,66 On the other hand, in the Ising-like models, clustering occurs from the inside as well as the corners and sides. It should be noted that clustering from the inside is not always suppressed in the elastic interaction models. For example, in the second-order transitions through the AF-like or FR-like phases, clustering occurs also from the bulk,68,79 in which the elastic distortion energy is relatively small.

Lattice parameter misfit causes an elastic frustration leading to unexpected behavior in SC transitions with steps.76 SC models based on elastic interactions have been investigated for SC nanocomposites by modifying the surface environment,64,87,88 and they have shown hysteresis curves with size dependence, which are consistent with corresponding experiments. The effect of lattice parameter misfit is important especially around the interface. Very recently, elastic interaction models have also been applied in a SC core-SC shell nanocomposite, and the thermodynamic properties of the nanocomposite have been studied as a function of the lattice parameter misfit between the two components around the interface.89,90

The effect of the elastic interaction has been taken into account also in macroscopic approaches. SC transitions with steps have recently been studied in a Landau theory considering that non-symmetry-breaking and symmetry-breaking order parameters are coupled to the volume change.91,92

In the present paper, we study elastic interaction models for SC systems in which LS, HS, and AF-like phases are realized as thermodynamics phases (upper panels in Fig. 3) and also those in which the LS, HS, LS-rich FR-like, and HS-rich FR-like phases are realized as thermodynamics phases (lower panels in Fig. 3). We analyze several phase diagrams for those systems, and making use of the phase diagrams, we show SC transition patterns with one-step, two-step, and three-step SC transitions. We also investigate an elastic interaction model for core-shell SC nanocomposites and show the occurrence of several two-step SC transitions originating from the difference in the model parameters between the core and shell.

FIG. 3.

Molecular configurations of several thermodynamic phases in SC materials. LS (upper left), HS (upper middle), AF-like (upper right), LS-rich FR-like (lower left), and HS-rich FR-like (lower right) phases. Red and blue circles denote HS and LS molecules.

FIG. 3.

Molecular configurations of several thermodynamic phases in SC materials. LS (upper left), HS (upper middle), AF-like (upper right), LS-rich FR-like (lower left), and HS-rich FR-like (lower right) phases. Red and blue circles denote HS and LS molecules.

Close modal

The rest of the paper is organized as follows. In Sec. II, the model and method are presented. In Sec. III, we show the phase diagrams for square and triangular-lattice SC systems with metastable phases. Using the phase diagrams, we discuss thermodynamic properties of possible phases and SC transition patterns with steps. In Sec. IV, we model SC core-shell nanocomposites and give SC transition patterns with steps. We discuss characteristic features of the SC core-shell nanocomposites with an analysis of clustering and local pressure profiles. Section V is devoted to summary.

In simulation of elastic interaction models for SC transitions, Monte Carlo (MC) and molecular dynamics (MD) methods have often been used. In many cases, MC methods are easier to handle to simulate SC models than MD methods,53,57 and many studies have been performed by MC methods. A combination between MD and MC methods is effective to study dynamics with different time scales between spin and lattice motions.69 First, we give a brief instruction for the methods.

The Hamiltonian for the MD method consists of the kinetic energy term, K, and the potential energy term, V,
(1)
To treat the change of the molecular size accompanied by the change of the molecular state and the cooperativity of the elastic interaction, we take into account for K the translational motion of the center of mass (first term) of each molecule and its totally symmetric vibration mode (second term),
(2)
and for V, the intra- and intermolecular potentials,
(3)
Here, X i is the coordinate of the center of the ith molecule, and M is the mass of the molecule (Fig. 2). R i and P i are its molecular radius and momentum, respectively. p i represents the momentum of the totally symmetric (breathing) mode, and m is the reduced mass of this mode. N is the number of molecules. The intramolecular potential, V i intra, is a function of the radius, R i, shown in Fig. 1. The HS state has much larger density of states than the LS state. Thus, the entropy difference from other intramolecular vibration modes and also spins has to be included in the Hamiltonian, and this entropy difference is introduced by coupling oscillators whose frequencies depend on the molecular state.57 The temperature effect is studied by using a thermal reservoir such as the Nosé–Hoover thermostat.53,57,93,94
For the MC method, only the potential energy term is considered for the Hamiltonian,
(4)
The Ising spin formulation is applicable to the MC method (not to the MD); namely, the Ising up- and down-spins can be allocated to the HS and LS states, respectively. Thus, the model can be simpler, and simulations are generally easier than the MD method.

We can use a combination between the MD and MC methods, in which the change of the spin state is performed by the MC method and the deformation of the lattice is done by the MD method. The merit of this method is to control the relative ratio of the time scales of the spin-state change and lattice relaxation, which was applied to the property of the roughness of the interface growth depending on the two time scales.69 

Throughout this article, we use a MC method. For H MC, we consider the following Hamiltonian:
(5)
which consists of the elastic interaction ( H Elastic), the SR (Ising) interaction ( H IS), and the effective field term ( H eff). Each SC molecule has the LS state ( σ i = 1) or the HS state ( σ i = 1). Its radius R i is a function of the state σ i because the LS molecule is smaller than the HS molecule; i.e., R LS < R HS, where R LS ( R HS) is the radius of the LS (HS) molecule.
On the triangular lattice, the nearest-neighbor (NN) interaction is considered for the elastic interaction, while in the square lattice, a small contribution of the next-nearest-neighbor (NNN) elastic interaction is also taken into account to maintain the lattice structure,
(6)
where
(7)
and
(8)
where k 2 = k 1 / 10 is used. Here, r i , j is the instantaneous distance between the ith and jth molecules; i.e., r i , j = | X j X i |.
H IS is the NN Ising SR interaction, which is, in general, not of a magnetic origin,
(9)
The energy difference between the HS and LS states is defined as D = E HS E LS. The entropy effect owing to the difference of the density of states ( ρ) between the LS and HS states is important for SC transitions. In the model, the difference is introduced with the Ising spin with the ρ LS-fold LS state and the ρ HS-fold HS state; i.e., σ i = 1 , , 1 ρ LS, 1 , , 1 ρ HS. This degenerated Ising spin is equivalent to σ i = 1 , 1 with an extra field described as 1 2 k B T ln g with g = ρ HS / ρ LS with respect to the partition function. Thus, the following effective field term is incorporated:
(10)
where
(11)

Here, we use periodic boundary conditions in the MC method to study bulk properties in the N P T ensemble (Sec. III) and open boundary conditions (OBCs) to investigate surface nucleation and clustering in finite size systems and properties of nanocomposite (Sec. IV) . The pressure is set to P = 0. In the N P T MC method, we choose a molecule (site i), update σ i, R i, and X i, and then update the volume of the total system under P = 0. In OBCs, the volume change is automatically realized. One MC step (MCS) is defined as N repetitions of the update of a molecule. We apply 10 4 10 6 MCS for measurement of physical quantities after the same number of MCS for equilibration.

As mentioned in Sec. I, the elastic interaction causes a LR interaction. Focusing on the infinite-range component of the LR interaction, we introduce a simple model for SC transitions with SR (Ising) and infinite-LR (Husimi-Temperley) interactions,73,83–86
(12)
The total Hamiltonian is
(13)
This form is derived from the elastic term [Eq. (6)] in a kind of mean-field treatment assuming uniform change of the intermolecular distance, r i , j r ¯, and uniform spin state, σ i 1 N i N σ i, under the condition that r ¯ takes the minimum energy.85,86 Here, N is the number of total molecules. This type of models has also been used for SC transitions with steps.83–86 

We found that in spite of drastic simplification, the model (13) with an AF-like coupling, J < 0, on the square lattice catches essential features of the phase diagram for the elastic interaction model (5) with J < 0 in both weak and strong elastic interaction cases. Owing to competing SR and LR interactions, unusual “horn structures” appear in the phase diagram of the model (13) with large A ( > 0 ) as realized in that of the model (5) in the strong elastic interaction case (see also in Sec. III B).79,83,84

For ferromagnetic (F)-like (LS or HS) order, we define magnetization
(14)
where N is the total number of the molecules and stands for the thermal average. Here, “magnetization” is used in analogy to magnetic systems, and it is the difference in the number between HS and LS molecules. It detects the F-like order. Then, the HS fraction is given as
(15)
In the same way, to detect AF-like order, a symmetry-breaking order parameter, i.e., staggered magnetization is defined as
(16)
where ( i x, i y) is the integer coordinate of the ith molecule, which numbers the two-dimensional lattice.
To study the LS-rich and HS-rich FR-like phases, we introduce a vector that quantifies the three-sublattice states shown in Fig. 4,
(17)
at the lth triangular plaquette. In a perfect FR-like phase, the vectors at the plaquettes align in one of six directions, and we define the following order parameter for the breaking of Z 6 symmetry:
(18)
in which N p is the number of the triangular plaquettes; i.e., N p = N / 3. To distinguish between HS-rich and LS-rich FR-like phases, we define the following order parameter using M :
(19)
This order parameter gives M 3 = 1 and M 3 = 1 for perfect HS-rich and LS-rich FR-like phases, respectively.
FIG. 4.

Six plaquette states of the FR-like state. Red and blue circles denote HS and LS molecules, respectively.

FIG. 4.

Six plaquette states of the FR-like state. Red and blue circles denote HS and LS molecules, respectively.

Close modal
Volume change is a characteristic in spin-crossover transitions, and we introduce the normalized volume defined as
(20)
Here, V is the (two-dimensional) volume of the system and V LS is that of the perfect LS phase.

We study one-step SC transitions between the LS and HS spin phases (upper left and middle configurations in Fig. 3). The transition is theoretically classified to three types: first-order transition, second-order transition, and gradual change. As we see later, the second-order transition is a special case, and practically, there are two types of transitions.

We can get a good outlook about the classification using the field–temperature ( h T) phase diagram of the model (5) with J = 0 or J > 0 (ferromagnetic-like case) illustrated in Fig. 5. The feature of the phase diagram is independent of the space dimension and the lattice symmetry of the system. The pink dashed line is the coexistence line of the LS and HS phases. The right edge of the coexistence line is the critical point ( T c). This point has the mean-field universality,55,65,68 which indicates a long-range interaction nature. The upper (lower) region is the HS (LS) phase, and the orange lines are spinodal lines. Therefore, the metastable LS (HS) phase exists between the upper (lower) spinodal line and the coexistence line. The region at high temperatures is a disordered phase, which connects the HS and LS phases continuously.

FIG. 5.

Phase diagram of the model (5) with J = 0 or J > 0. The pink dashed line shows the coexistence line of the LS and HS phases. The orange lines are spinodal lines. T c is the critical point. T SC is the center of the SC transition. MLS and MHS stand for metastable LS and metastable HS, respectively.

FIG. 5.

Phase diagram of the model (5) with J = 0 or J > 0. The pink dashed line shows the coexistence line of the LS and HS phases. The orange lines are spinodal lines. T c is the critical point. T SC is the center of the SC transition. MLS and MHS stand for metastable LS and metastable HS, respectively.

Close modal
Noting that the effective field (11) is a linear function of temperature T, and a variation of the temperature in the model (5) causes a shift of the position ( T, h) in the phase diagram along a straight line, e.g., the thin blue line in Fig. 5. The SC transition (change between m > 0 and m < 0) is realized when
(21)
and thus, the SC transition temperature is given by
(22)

If the elastic interaction is stronger, T c is higher. If T SC < T c, the SC transition is of first order and is accompanied with a thermal hysteresis whose width is T 2 T 1 in Fig. 5. If T SC = T c, the SC transition is a second-order (continuous) transition. If T SC > T c, the SC transition is a gradual change. Here, we give an example of the first-order transition in Fig. 6 for k 1 = 80 000 K / nm 2, D = 1000 K, J = 0, g = 200, and R HS / R LS = 1.1 ( R LS = 1 nm) on the square lattice for L = 50, where L = N 1 / 2.

FIG. 6.

First-order spin-crossover transition with a hysteresis. k 1 = 80 000 K / nm 2, D = 1000 K, J = 0, and g = 200.

FIG. 6.

First-order spin-crossover transition with a hysteresis. k 1 = 80 000 K / nm 2, D = 1000 K, J = 0, and g = 200.

Close modal

Next, we study two-step SC transitions through an intermediate AF-like phase (upper right configuration in Fig. 3) in which LS and HS molecules align alternately. Here, we consider the AF-like SR (Ising) interaction, i.e., J < 0, on the square lattice.

In the same manner as the one-step SC transition, we perform a phase diagram analysis to classify the type of the two-step SC transitions.79 There are basically three patterns for the two-step SC transition.

  • double second-order (continuous) transitions.

  • first-order and second-order (continuous) transitions.

  • double first-order transitions.

We depict the h T phase diagram in Fig. 7 when the elastic interaction is not so strong. The green line denotes a critical line. This line has the Ising universality, in which the SR interaction is relevant.68,79 The upper pink dashed line denotes the coexistence line of the HS and AF-like phases and the lower pink dashed line that of the AF-like and LS phases. The upper and lower orange lines are the spinodal lines of the AF-like phase, while the upper and lower violet lines are those of the HS and LS phases. The metastable AF phase exists between the upper orange solid and upper pink dashed lines and also between the lower pink dashed and lower orange solid lines. The metastable HS phase exists between the upper pink dashed and upper violet solid lines, while the metastable LS phase exists between the lower pink dashed and lower violet solid lines.

FIG. 7.

Phase diagram for the elastic interaction model (5) with J < 0 (AF-like) on the square lattice. The path (blue thin line) for a two-step SC transition with first-order and second-order transitions is given in (a), and the path for a two-step SC transition with double first-order transitions is given in (b). The upper and lower pink dashed lines denote the coexistence line of the HS and AF-like phases and that of the AF-like and LS phases, respectively. The green line denotes a critical line. The upper and lower orange lines are the spinodal lines of the AF-like phase. The upper and lower violet lines are those of the HS and LS phases, respectively. MAF, MHS, and MLS stand for metastable AF-like, metastable HS, and metastable LS phases, respectively.

FIG. 7.

Phase diagram for the elastic interaction model (5) with J < 0 (AF-like) on the square lattice. The path (blue thin line) for a two-step SC transition with first-order and second-order transitions is given in (a), and the path for a two-step SC transition with double first-order transitions is given in (b). The upper and lower pink dashed lines denote the coexistence line of the HS and AF-like phases and that of the AF-like and LS phases, respectively. The green line denotes a critical line. The upper and lower orange lines are the spinodal lines of the AF-like phase. The upper and lower violet lines are those of the HS and LS phases, respectively. MAF, MHS, and MLS stand for metastable AF-like, metastable HS, and metastable LS phases, respectively.

Close modal

In Fig. 7(a), we show a path (blue thin line) for a temperature change (case II), and it crosses the critical line (green) ( T = T 3), the spinodal line of the LS phase ( T = T 2), and the spinodal line of the AF phase ( T = T 1). In the heating process, the LS phase changes to the AF-like phase at T = T 2 and the AF-like phase changes to the HS phase at T = T 3, while in the cooling process, the HS phase changes to the AF-like phase at T = T 3 and the AF-like phase changes to the LS phase at T = T 1. We plot the temperature dependences of f HS and V n in Fig. 8(a) for case II, where V n = 1.21 in the perfect HS phase. Here, we used J = 100 K, k 1 = 40 000 K / nm 2, D = 1030 K, g = 1000, and R HS / R LS = 1.1 ( R LS = 1 nm) for L = 40. The corresponding m sg is given in Fig. 8(b). We find that the feature of the volume change is similar to that of f HS and observe volume jumps in the hysteresis loop.

FIG. 8.

Temperature dependences of (a) f HS, V n, and (b) m sg in a two-step SC transition with first-order and second-order transitions. J = 100 K, k 1 = 40 000 K / nm 2, D = 1030 K, and g = 1000.

FIG. 8.

Temperature dependences of (a) f HS, V n, and (b) m sg in a two-step SC transition with first-order and second-order transitions. J = 100 K, k 1 = 40 000 K / nm 2, D = 1030 K, and g = 1000.

Close modal

In Fig. 7(b), we give a path for a two-step SC transition with double first-order transitions (case III), in which in the heating process, the LS phase changes to the AF-like phase at T 2 and the AF-like phase changes to the HS phase at T 4, while in the cooling process, the HS phase changes to the AF-like phase at T 3 and the AF-like phase changes to the LS phase at T 1. In addition to the two-step transition, the one-step transition between the AF-like and HS phases can be realized with another parameter set.79 

If the elastic interaction is strong, the feature of the phase diagram changes. Their unusual “horn structures,” each of which is surrounded by F-like spinodal lines, disorder (D) spinodal lines, and a critical line, appear at high temperatures. These structures cause a different kind of a two-step transition, in which a transition from the LS to AF to D to HS phase in the heating process and from the HS to D to AF to LS phase in the cooling process occurs.79 

We investigate spin-crossover transitions on triangular-lattice systems.73,81 If a SR AF-like interaction is introduced into a triangular-lattice (nonbipartite) SC system such as [ Fe ( bbtr ) 3 ] ( ClO 4 ) 2,95 a frustration of the spin-state configuration, which accompanies an elastic frustration, is induced. Here, in addition to the NN SR (Ising) interaction, a small NNN SR (Ising) interaction, J 2, is also considered. If J 1 0 (F-like) and/or J 2 0 (F-like), the phase diagram is similar to that of Fig. 5. Here, we take into account J 1 < 0 (AF-like) and J 2 0, which causes FR-like thermodynamic phases (lower left and right configurations in Fig. 3).

For the case of relatively weak elastic interactions, a typical phase diagram is given in Fig. 9. The two black solid lines in the phase diagram are second-order phase-transition lines with the three-state Potts universality.81,96,97 The pink dashed line at H = 0 is the coexistence line of the HS-rich and LS-rich FR-like phases. The yellow line at H = 0 is a Berezinskii–Kosterlitz–Thouless (BKT) critical line, in which the right end point (right black circle) is a BKT transition point,98–103 while the left end point is a second-order phase-transition point (left black circle).73 The upper (lower) pink dashed line is the coexistence line of HS (LS) and HS-rich (LS-rich) FR-like phases. The violet lines are spinodal lines for F-like (HS and LS) phases. The upper (lower) orange line is the spinodal line for the HS-rich (LS-rich) FR-like phase against the stable F-like phase. The upper (lower) green line is the spinodal line for the LS-rich (HS-rich) FR-like phase against the stable HS-rich (LS-rich) FR-like phase.

FIG. 9.

Phase diagram for the weak elastic interaction case ( R HS / R LS = 1.03) with the J 1 < 0 and J 2 > 0 short-range interactions on the triangular lattice. The two black lines denote critical lines. The upper (lower) orange line is the spinodal line for the HS (LS) rich FR-like phase against the stable F-like phase. The violet lines are the F-like spinodal (HS and LS spinodal) lines. The upper (lower) green line is the spinodal line for the LS (HS) rich FR-like phase against the HS (LS) rich FR-like phase. The upper, middle, and lower dashed pink lines denote coexistence lines of the HS and HS-rich FR-like phases, the HS-rich and LS-rich FR-like phases, and the LS-rich FR-like and LS phases, respectively. MHS, MLS, and MFerri stand for metastable HS, metastable LS, and metastable FR-like phases, respectively. Noted that there is no phase boundary between the HS (LS) and disorder (D) phases. The blue thin line is a path for a three-step SC transition.

FIG. 9.

Phase diagram for the weak elastic interaction case ( R HS / R LS = 1.03) with the J 1 < 0 and J 2 > 0 short-range interactions on the triangular lattice. The two black lines denote critical lines. The upper (lower) orange line is the spinodal line for the HS (LS) rich FR-like phase against the stable F-like phase. The violet lines are the F-like spinodal (HS and LS spinodal) lines. The upper (lower) green line is the spinodal line for the LS (HS) rich FR-like phase against the HS (LS) rich FR-like phase. The upper, middle, and lower dashed pink lines denote coexistence lines of the HS and HS-rich FR-like phases, the HS-rich and LS-rich FR-like phases, and the LS-rich FR-like and LS phases, respectively. MHS, MLS, and MFerri stand for metastable HS, metastable LS, and metastable FR-like phases, respectively. Noted that there is no phase boundary between the HS (LS) and disorder (D) phases. The blue thin line is a path for a three-step SC transition.

Close modal

We give a path in temperature variation depicted by a thin blue line in Fig. 9. This line crosses one of the critical lines at T 4, the BKT line at T 3, the spinodal line for the LS phase at T 2, and the spinodal line for the LS-rich FR-like phase at T 1. We set the parameter values as k 1 = 40 000 K / nm 2, R HS / R LS = 1.03 ( R LS = 1 nm), J 1 = 100 K, and J 2 = 20 K. This parameter set gives a phase diagram like Fig. 9. For D = 1520 K and ln g = 11.692, the path of the temperature change is like the blue line. The temperature dependences of f HS and V n for L = 36 for this path are given in Fig. 10(a) and that of M 3 is shown in Fig. 10(b). The feature of V n is similar to that of f HS, where V n = 1.06 in the perfect HS phase. We find a three-step SC transition with first-order, BKT, and second-order phase transitions at low, middle, and high temperatures, respectively, and observe that the LS-rich and HS-rich FR-like phases appear in the middle temperature region. In the relatively weak elastic interaction case, five typical SC transition patterns for one-step to four-step SC transitions can be realized.81 

FIG. 10.

Temperature dependences of (a) f HS, V n, and (b) M 3 for the weak elastic interaction case ( k 1 = 40 000 K / nm 2, R HS / R LS = 1.03) with J 1 = 100 K and J 2 = 20 K.

FIG. 10.

Temperature dependences of (a) f HS, V n, and (b) M 3 for the weak elastic interaction case ( k 1 = 40 000 K / nm 2, R HS / R LS = 1.03) with J 1 = 100 K and J 2 = 20 K.

Close modal

In the relatively strong elastic interaction with J 1 < 0 and J 2 > 0, the feature of the phase diagram changes.81 The BKT phase disappears, and alternatively, a first-order transition point appears as the end point of the coexistence line between the HS-rich and LS-rich FR-like phases, at which the two critical lines terminate. The three coexistence lines and six spinodal lines are extended to the higher-temperature side. When J 1 < 0 and J 2 = 0, i.e., only NN SR interaction, the topology of that is similar to the phase diagram with J 1 < 0 and J 2 > 0 for the relatively strong elastic interaction, but the regions of the coexistence lines are very small.81 Theoretically, there can be a variety of transition patterns in the triangular SC system: three, four, three, and two typical patterns for one-step, two-step, three-step, and four-step SC transitions, respectively.81 

In the properties of SC nanocomposites, finite size effects are important. Before investigating SC nanocomposites, we study clustering features in a SC nanosystem. Here, we use the elastic interaction model (5) with open boundary conditions for L = 40. The parameters are set as k 1 = 90 000 K / nm 2, D = 1200 K, J = 0, and g = 200.

The temperature dependence of f HS is plotted in Fig. 11. Corresponding snapshots of the configuration are given in Fig. 12. It is a characteristic that clustering of LS (HS) domains in the HS (LS) phase occurs from the corners during the transition from the HS (LS) to LS (HS) phase, and the clustering of LS molecules is sharper than that of HS molecules because of lower temperature. Clustering from the sides has a much lower probability owing to lower stability of the clusters, and clustering from the inner part is unrealistic owing to high distortion energy. This clustering feature is different from that of the Ising-like model, in which clustering occurs not only from the corners but also from the sides and the inner part. J > 0 enforces the cluster formation, and the clustering feature is similar to the case of J = 0.

FIG. 11.

Temperature dependence of f HS in the nano SC system. k 1 = 90 000 K / nm 2, D = 1200 K, J = 0, and g = 200.

FIG. 11.

Temperature dependence of f HS in the nano SC system. k 1 = 90 000 K / nm 2, D = 1200 K, J = 0, and g = 200.

Close modal
FIG. 12.

Snapshots of the configuration in the nano SC system.

FIG. 12.

Snapshots of the configuration in the nano SC system.

Close modal

Here, the elastic interaction model (5) is applied to core-shell nanosystems. We use a core-shell nanostructure illustrated in Fig. 13.89,90 The core has a square shape composed of L C × L C molecules, and the shell has a shape of a thick frame of width L S surrounding the core. The numbers of molecules in the core and the shell are L C 2 and 4 L S ( L S + L C ), respectively, and the total number is ( L C + 2 L S ) 2.

FIG. 13.

Core-shell spin-crossover nanocomposite model. Pink and green circles denote molecules belonging to the core and the shell, respectively.

FIG. 13.

Core-shell spin-crossover nanocomposite model. Pink and green circles denote molecules belonging to the core and the shell, respectively.

Close modal
The Hamiltonian for the core-shell system is expressed as
(23)
Here, H core ( H shell) denotes Eq. (5) of the core (shell) part, and H inter indicates the interface bonds [Eq. (6)] between the core and shell. H core and H shell contain parameters ( k 1 C, D C, J C, g C, R HS C, R LS C) and ( k 1 S, D S, J S, g S, R HS S, R LS S), respectively, and H inter contains k 1 I and k 2 I.

We use the same elastic constant in the core, shell, and interface, i.e., k 1 C = k 1 S = k 1 I, and the same HS/LS degeneracy ratio g, i.e., g C = g S = 200 for simplicity. The Ising interaction is not considered: J C = J S = 0. We set L C = 40 and L S = 5.

Variations of R HS C and R HS S cause changes of the strength of the elastic interactions in the core and shell, respectively, which influences the interaction between the core and shell. These variations lead to different transition features. In the present paper, we consider the following two cases.

  • the shell SC transition temperature is higher: T SC C < T SC S.

  • the core SC transition temperature is higher: T SC S < T SC C.

Here, to study the HS (or LS) ordering in the core-shell system (finite size system), we use the expression of the SC transition temperature in the bulk system (22). The SC transition temperatures in the core and shell are
(24)
and
(25)
respectively. Since g = 200 is common in the core and shell, D C < D S and D C > D S are set in cases I and II, respectively. In addition to the high-spin fraction, f HS, for the whole system, we define the core high-spin fraction, f HS C, and the shell high-spin fraction, f HS S, which are estimated within the core and shell, respectively.
Firstly, we investigate case I, and D C = 200 K and D S = 800 K are used, which yields
(26)
and
(27)
In both the core and shell, k 1 = 30 000 K / nm 2 is commonly used. Here, we fix R HS S = 1.1 nm, R LS C = R LS S = 1.0 nm, and change the value of R HS C.

As reference data for the core-shell system, we depict the temperature dependences of f HS of the uncoated core with R HS C = 1.02, 1.06, and 1.1 nm and of the hollow shell in Fig. 14. For larger R HS C, the first-order transition nature is stronger in the uncoated core.

FIG. 14.

Temperature dependences of f HS of the uncoated core with R HS C = 1.02, 1.06, and 1.1 nm and of the hollow shell in case I.

FIG. 14.

Temperature dependences of f HS of the uncoated core with R HS C = 1.02, 1.06, and 1.1 nm and of the hollow shell in case I.

Close modal

In Fig. 15, we show the temperature dependences of f HS, f HS C, and f HS S for R HS C = 1.02 nm in the core-shell system. We find a smooth two-step transition with the first step at T T SC C and the second step at T T SC S. Both f HS C and f HS S show smooth curves without hysteresis. The size of the core HS molecule ( R HS C = 1.02 nm) is close to the core LS molecule, and the elastic interaction in the core is small. Above the lower transition temperature, the shell LS molecules around the interface are stabilized around T T SC S since R HS C is close to R LS S, and the transition point in the shell slightly shifts to the high-temperature side. This causes a smooth transition in the shell, in which the small hysteresis observed in the hollow shell in Fig. 14 disappears. The size of the shell HS molecule ( R HS S = 1.1) is not close to that of the shell LS molecule, and the elastic interaction in the shell is larger than that in the core, but the shell has a large surface (open boundary) and the elastic interaction around the surface is suppressed.

FIG. 15.

Temperature dependences of the total high-spin fraction f HS, the core high-spin fraction f HS C, and the shell high-spin fraction f HS S for k 1 = 30 000 K / nm 2, D C = 200 K, D S = 800 K, and g = 200. R HS C = 1.02 nm is used.

FIG. 15.

Temperature dependences of the total high-spin fraction f HS, the core high-spin fraction f HS C, and the shell high-spin fraction f HS S for k 1 = 30 000 K / nm 2, D C = 200 K, D S = 800 K, and g = 200. R HS C = 1.02 nm is used.

Close modal

Next, we consider a case of a larger size of the core HS molecule. In Fig. 16, we show the temperature dependences of f HS, f HS C, and f HS S for R HS C = 1.06. We find a hysteresis loop in f HS, which reflects a hysteresis loop of f HS C. The SC transition point in the core is located at a little higher temperature than T = T SC C. Compared to the case R HS C = 1.02 nm (Fig. 15), the shell HS fraction, f HS S, changes at a lower temperature, and the transition becomes shaper. A large elastic interaction caused by a large difference in the molecular size between the LS and HS states in the core is the origin of the hysteresis. The core HS molecule is bigger than the shell LS molecule. Therefore, the core HS molecules around the interface have a positive pressure before the transition to the HS state in the core, which shifts the transition point in the core to a higher temperature, while the shell LS molecules around the interface have a negative pressure before the transition to the HS state in the shell, which accelerates a change to the HS state in the shell.

FIG. 16.

Temperature dependences of f HS, f HS C, and f HS S for k 1 = 30 000 K / nm 2, D C = 200 K, D S = 800 K, and g = 200. R HS C = 1.06 nm is used.

FIG. 16.

Temperature dependences of f HS, f HS C, and f HS S for k 1 = 30 000 K / nm 2, D C = 200 K, D S = 800 K, and g = 200. R HS C = 1.06 nm is used.

Close modal

We study a case of a further large size of the core HS molecule. In Fig. 17, we show the temperature dependences of f HS, f HS C, and f HS S for R HS C = 1.1. We find that f HS shows a large hysteresis at low temperatures. It should be noted that in this hysteresis loop, there are two contributions, namely, hysteresis loops of f HS C and f HS S. In the heating process, some of the shell LS molecules become the HS state before the transition to the HS state in the core is completed, and in the cooling process, some of the core HS molecules become the LS state before the transition to the LS state in the shell is completed. Specifically, the LS–HS transition in the core and that in the shell occur concertedly. Here, the change of f HS cannot be expected from the values of T SC C and T SC S.

FIG. 17.

Temperature dependences of f HS, f HS C, and f HS S for k 1 = 30 000 K / nm 2, D C = 200 K, D S = 800 K, and g = 200. R HS C = 1.1 nm is used.

FIG. 17.

Temperature dependences of f HS, f HS C, and f HS S for k 1 = 30 000 K / nm 2, D C = 200 K, D S = 800 K, and g = 200. R HS C = 1.1 nm is used.

Close modal

The larger size of the core HS molecule causes a strong elastic interaction and a large hysteresis of f HS C. In the heating process, the metastability in the core is very strong because the transition to the HS state in the core occurs against a positive pressure from the shell LS molecules around the interface, and thus, the right branch of the hysteresis loop shifts to the high-temperature side. On the other hand, in the cooling process, the metastability in the core is not so strong because the transition to the LS state in the core is driven by a positive pressure caused by the shell LS molecules around the interface.

We study microscopic features of the SC transition of the core-shell system for R HS C = 1.1 by analyzing snapshots of the molecular configuration in the heating and cooling processes and of the spatial profile of the local pressure defined as
(28)
(29)
where NN and NNN stand for the nearest and next-nearest-neighbor molecules, respectively. Snapshots of the configuration and the local pressure for R HS C = 1.1 (Fig. 17) are shown in Figs. 18 and 19, respectively.
FIG. 18.

Snapshots of the core-shell system in the heating and cooling processes for Fig. 17. Red and blue molecules denote those in high-spin and low-spin states, respectively.

FIG. 18.

Snapshots of the core-shell system in the heating and cooling processes for Fig. 17. Red and blue molecules denote those in high-spin and low-spin states, respectively.

Close modal
FIG. 19.

Snapshots of the local pressure in the core-shell system in the heating and cooling processes for Fig. 17. Each color is allocated as yellow ( p i > 5000 K / nm), light green ( 5000 K / nm p i > 3000 K / nm), spring green ( 3000 K / nm p i > 1000 K / nm), lime ( 1000 K / nm p i > 1000 K / nm), green ( 1000 K / nm p i > 3000 K / nm), dark green ( 3000 K / nm p i > 5000 K / nm), and black ( 5000 K / nm p i).

FIG. 19.

Snapshots of the local pressure in the core-shell system in the heating and cooling processes for Fig. 17. Each color is allocated as yellow ( p i > 5000 K / nm), light green ( 5000 K / nm p i > 3000 K / nm), spring green ( 3000 K / nm p i > 1000 K / nm), lime ( 1000 K / nm p i > 1000 K / nm), green ( 1000 K / nm p i > 3000 K / nm), dark green ( 3000 K / nm p i > 5000 K / nm), and black ( 5000 K / nm p i).

Close modal

At low temperatures, e.g., 30 K, the local pressure is almost uniform with relatively small positive or negative values. It should be noted that at around T = 80 K in the heating process, only small HS domains are generated in the core to reduce elastic distortion. The LS and HS molecules in the core have relatively low and high pressures, respectively. With further heating, the HS molecules stay in the HS, but many LS molecules switch to the HS state because of entropy gain. At 90 K in the heating process, the shell LS molecules at the sides (not corners) change to HS ones, and the transition to the HS state in the shell around the corners is delayed. There, the local pressure of the side molecules in the shell is low, and the HS domains are stable. At high temperatures ( 150 K), the state is close to the saturated HS state, but the fluctuation of the local pressure is found. At T = 80 K in the cooling process, clustering of the LS molecules occurs from the corners in the shell. With further cooling to T = 30 K, the rest shell HS molecules change to LS molecules, and simultaneously, clustering of the core LS molecule occurs from the corners of the core, around which the local pressure of the core HS molecules is high and this accelerates the transition to the LS state. Accompanying the clustering, small domains of the LS molecule appear inside the core to minimize the elastic distortion.

Next, we study case II. The parameter values D C = 800 K and D S = 200 K with g = 200 are used for the effective field term, which yields
(30)
and
(31)
In both the core and shell, k 1 = 60 000 K / nm 2 is commonly used. We fix R HS C = 1.1 nm, R LS C = R LS S = 1.0 nm and change the value of R HS S.

As reference data for the core-shell system, we illustrate the temperature dependences of f HS of the hollow shell with R HS S = 1.01, 1.08, and 1.1 nm and of the uncoated core in Fig. 20. For larger R HS S, the first-order transition nature in f HS is stronger in the hollow shell.

FIG. 20.

Temperature dependences of f HS of the hollow shell with R HS S = 1.01, 1.08, and 1.1 nm and of the uncoated core in case II.

FIG. 20.

Temperature dependences of f HS of the hollow shell with R HS S = 1.01, 1.08, and 1.1 nm and of the uncoated core in case II.

Close modal

In Fig. 21, we show the temperature dependences of f HS, f HS C, and f HS S for R HS S = 1.01 nm. f HS shows a two-step transition with a low-temperature smooth transition and a high-temperature transition with a hysteresis loop. In the heating (cooling) process, after the transition in the shell (core) is completed, the transition in the core (shell) occurs. The size of the HS molecule in the shell is small; thus, the elastic interaction in the shell is small, and a continuous transition between the LS and HS states in the shell takes place at T T SC S.

FIG. 21.

Temperature dependences of f HS, f HS C, and f HS S for k 1 = 60 000 K / nm 2, D C = 800 K, and D S = 200 K, and g = 200. R HS S = 1.01 nm is used.

FIG. 21.

Temperature dependences of f HS, f HS C, and f HS S for k 1 = 60 000 K / nm 2, D C = 800 K, and D S = 200 K, and g = 200. R HS S = 1.01 nm is used.

Close modal

It should be noted that a hysteresis does not appear at the lower transition temperature but appears at the higher transition temperature (175–190 K), although the hysteresis width is smaller than that of the uncoated core in Fig. 20. In the case of the two-step transition through the AF-like phase studied in Sec. III B, a hysteresis appears at the lower transition temperature and a stronger metastable state is realized at low temperatures. In this core-shell system, however, a hysteresis at the higher transition temperature is easier to realize. The shell HS molecular size, 1.01 nm, is small, and the molecules in the core change to the HS state under a large positive pressure. This effect reinforces the stability of the LS state in the core and causes a higher transition temperature, T 185 K, in the core than T = T SC C. This also makes the location of the left branch of the hysteresis loop at a higher-temperature side, and as a result, the hysteresis width becomes smaller than that of the uncoated core in Fig. 20.

We consider a larger size of the HS molecule in the shell, R HS S = 1.08. In Fig. 22, we show the temperature dependences of f HS, f HS C, and f HS S. f HS shows a two-step transition with a small hysteresis loop at a little higher temperature than T SC S and a large hysteresis loop at a high temperature, which is lower than T = T SC C. It should be noted that in the heating process, before the transition to the HS states is completed in the shell, the transition to the HS state starts in the core, and in the cooling process, before the transition to the LS state is completed in the core, the transition to the LS state starts in the shell.

FIG. 22.

Temperature dependences of f HS, f HS C, and f HS S for k 1 = 60 000 K / nm 2, D C = 800 K, D S = 200 K, and g = 200. R HS S = 1.08 nm is used.

FIG. 22.

Temperature dependences of f HS, f HS C, and f HS S for k 1 = 60 000 K / nm 2, D C = 800 K, D S = 200 K, and g = 200. R HS S = 1.08 nm is used.

Close modal

We study this case (Fig. 22) microscopically by analyzing snapshots of molecular configuration in Fig. 23 and of the local pressure in Fig. 24. In the heating process, clustering of HS molecules in the shell occurs from the corners around T = 50 K. On further heating to T = 140 K, the HS domains expand in the shell and simultaneously clustering of HS molecules in the core occurs from the corners, in which the core LS molecules have a low pressure and this helps them to easily change to the HS state. However, this clustering is not so clear, and small HS domains appear inside the core.

FIG. 23.

Snapshots of the core-shell system in heating and cooling for Fig. 22. Red and blue molecules denotes those in high-spin and low-spin states, respectively.

FIG. 23.

Snapshots of the core-shell system in heating and cooling for Fig. 22. Red and blue molecules denotes those in high-spin and low-spin states, respectively.

Close modal
FIG. 24.

Snapshots of the local pressure in the core-shell system in heating and cooling for Fig. 22.

FIG. 24.

Snapshots of the local pressure in the core-shell system in heating and cooling for Fig. 22.

Close modal

In the cooling process, small LS domains appear inside the core at T 110 K to reduce the distortion due to the negative pressure appearing in the bulk. On further cooling to 50 K, the rest core HS molecules, which have high pressure, change to the LS state, and clustering of shell LS molecules from the sides of the shell occurs. The corner shell molecules change to the LS state in the end.

For a further large size of the shell HS molecule, R HS C = 1.1, we give in Fig. 25 the temperature dependences of f HS, f HS C, and f HS S. f HS shows a large hysteresis with a neck. In the heating process, a part of the shell LS molecules changes fast to the HS ones at T 60 K. The rest of the shell LS molecules gradually change to the HS ones up to T 150 K, and simultaneously, the core LS molecules change to the HS ones. In the cooling process, the shell HS molecules convert to the LS ones following the change of the core HS molecules to the LS ones, and in the ending of the transition to the LS state, f HS and f HS S change slowly.

FIG. 25.

Temperature dependences of f HS, f HS C, and f HS S for k 1 = 60 000 K / nm 2, D C = 800 K, D S = 200 K, and g = 200. R HS S = 1.1 nm is used.

FIG. 25.

Temperature dependences of f HS, f HS C, and f HS S for k 1 = 60 000 K / nm 2, D C = 800 K, D S = 200 K, and g = 200. R HS S = 1.1 nm is used.

Close modal

The elastic interaction caused by lattice distortion owing to molecular size difference between the low-spin (LS) and high-spin (HS) molecules plays an essential role in cooperativity of SC phenomena. In this paper, we presented a tutorial study on elastic interaction models for spin-crossover (SC) compounds. Using the models, we studied two- and multi-step behavior in SC transitions in several systems. The competition and/or synergy effects between short-range interactions and long-range interactions of elastic origin cause a variety of thermodynamic and metastable phases. We analyzed several typical phase diagrams with metastable phases. Considering the path corresponding to temperature variation in the phase diagrams, we can classify what kind of transitions and what transition patterns can be realized with increasing and decreasing temperature.

Firstly, we showed basic one-step SC transitions between the LS and HS phases. Secondly, we presented two-step SC transitions between the LS, antiferromagnetic (AF)-like, HS phases on the square lattice, and possible patterns of two-step transitions. The qualitative feature of the phase diagram for the one-step SC transition and that for the two-step transition with relatively small elastic interactions do not change in three dimensions. Next, we investigated three-step SC transitions between the LS, LS-rich ferrimagnetic (FR)-like, HS-rich FR-like, and HS phases on the triangular lattice, and showed a three-step SC transition. In the systems introduced in this study, many patterns of SC transitions with one to four steps are theoretically plausible. Applying this analysis to recently discovered multi-step SC materials in experiments, details in transition mechanisms would be found.

We also investigated SC transitions with steps in a SC core-SC shell nanocomposite model. We analyzed the two cases: (I) the core has a lower transition temperature than the shell ( T SC C < T SC S), and (II) the core has a higher transition temperature ( T SC C > T SC S). We presented and discussed the temperature dependences of the HS fractions of the total system, core, and shell with an analysis of snapshots of the molecular configuration and the local pressure. We focus on lattice parameter misfit leading to an elastic frustration, which is an important ingredient for complex SC transitions in core-shell systems. The molecular size difference between the LS and HS state in the core and that in the shell affect the interaction between the core and shell, as well as the elastic interaction at the interface. Tuning the differences, a variety of SC transitions with steps are realized. For stronger interaction between the core and shell, the LS–HS transition in the core and that in the shell occur concertedly. In two-step transitions through the AF-like phase, the lower-temperature transition has a larger hysteresis. In two-step transitions in case II, however, the higher-temperature transition with a larger hysteresis is possible.

We hope that our study contributes to fully understanding of multistep SC phenomena and further development of SC studies.

This work was supported by Grants-in-Aid for Scientific Research C (Nos. 18K03444 and 20K03809) from the MEXT of Japan. Y.S. and K.B. acknowledge the ANR (Agence Nationale de la Recherche) for the funding received from project Mol-CoSM (No. ANR-20-CE07-0028-02), and all the authors acknowledge the LIA (International Associate French Japan Laboratory) for the financial support.

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
Spin Crossover in Transition Metal Compounds I, II, III, edited by P. Gütlich and H. A. Goodwin (Springer, Berlin, 2004).
2.
Spin-Crossover Materials—Properties and Applications, edited by M. A. Halcrow (John Wiley & Sons, Chichester, UK, 2013).
3.
O.
Kahn
and
C. J.
Martinez
, “
Spin-transition polymers: From molecular materials toward memory devices
,”
Science
279
,
44
48
(
1998
).
4.
Y.
Garcia
,
O.
Kahn
,
J.-P.
Ader
,
A.
Buzdin
,
Y.
Meurdesoif
, and
M.
Guillot
, “
The effect of a magnetic field on the inversion temperature of a spin crossover compound revisited
,”
Phys. Lett. A
271
,
145
154
(
2000
).
5.
A.
Bousseksou
,
K.
Boukheddaden
,
M.
Goiran
,
C.
Consejo
,
M.-L.
Boillot
, and
J.-P.
Tuchagues
, “
Dynamic response of the spin-crossover solid Co ( H 2 ( fsa ) 2 en ) ( py ) 2 to a pulsed magnetic field
,”
Phys. Rev. B
65
,
172412
(
2002
).
6.
G.
Molnár
,
V.
Niel
,
A. B.
Gaspar
,
J.-A.
Real
,
A.
Zwick
,
A.
Bousseksou
, and
J. J.
McGarvey
, “
Vibrational spectroscopy of cyanide-bridged, iron(II) spin-crossover coordination polymers estimation of vibrational contributions to the entropy change associated with the spin transition
,”
J. Phys. Chem. B
106
,
9701
9707
(
2002
).
7.
S.
Pillet
,
J.
Hubsch
, and
C.
Lecomte
, “
Single crystal diffraction analysis of the thermal spin conversion in [ Fe ( btr ) 2 ( NCS ) 2 ] ( H 2 O ): Evidence for spin-like domain formation
,”
Eur. Phys. J. B
38
,
541
552
(
2004
).
8.
R.
Tanasa
,
C.
Enachescu
,
A.
Stancu
,
J.
Linares
,
E.
Codjovi
,
F.
Varret
, and
J.
Haasnoot
, “
First-order reversal curve analysis of spin-transition thermal hysteresis in terms of physical-parameter distributions and their correlations
,”
Phys. Rev. B
71
,
014431
(
2005
).
9.
S.
Pillet
,
V.
Legrand
,
M.
Souhassou
, and
C.
Lecomte
, “
Kinetics of light-induced first-order phase transformation in molecular solids: Fe ( btr ) 2 ( NCS ) 2 H 2 O
,”
Phys. Rev. B
74
,
140101
(
2006
).
10.
N.
Bréfuel
,
H.
Watanabe
,
L.
Toupet
,
J.
Come
,
N.
Matsumoto
,
E.
Collet
,
K.
Tanaka
, and
J.-P.
Tuchagues
, “
Concerted spin crossover and symmetry breaking yield three thermally and one light-induced crystallographic phases of a molecular material
,”
Angew. Chem. Int. Ed.
48
,
9304
9307
(
2009
).
11.
C.
Chong
,
H.
Mishra
,
K.
Boukheddaden
,
S.
Denise
,
G.
Bouchez
,
E.
Collet
,
J.-C.
Ameline
,
A. D.
Naik
,
Y.
Garcia
, and
F.
Varret
, “
Electronic and structural aspects of spin transitions observed by optical microscopy. The case of [Fe(ptz) 6](BF 4) 2
,”
J. Phys. Chem. B
114
,
1975
1984
(
2010
).
12.
A.
Bousseksou
,
G.
Molnár
,
L.
Salmon
, and
W.
Nicolazzi
, “
Molecular spin crossover phenomenon: Recent achievements and prospects
,”
Chem. Soc. Rev.
40
,
3313
3335
(
2011
).
13.
A.
Slimani
,
F.
Varret
,
K.
Boukheddaden
,
C.
Chong
,
H.
Mishra
,
J.
Haasnoot
, and
S.
Pillet
, “
Visualization and quantitative analysis of spatiotemporal behavior in a first-order thermal spin transition: A stress-driven multiscale process
,”
Phys. Rev. B
84
,
094442
(
2011
).
14.
E.
Collet
,
H.
Watanabe
,
N.
Bréfuel
,
L.
Palatinus
,
L.
Roudaut
,
L.
Toupet
,
K.
Tanaka
,
J.-P.
Tuchagues
,
P.
Fertey
,
S.
Ravy
,
B.
Toudic
, and
H.
Cailleau
, “
Aperiodic spin state ordering of bistable molecules and its photoinduced erasing
,”
Phys. Rev. Lett.
109
,
257206
(
2012
).
15.
A.
Slimani
,
F.
Varret
,
K.
Boukheddaden
,
D.
Garrot
,
H.
Oubouchou
, and
S.
Kaizaki
, “
Velocity of the high-spin low-spin interface inside the thermal hysteresis loop of a spin-crossover crystal, via photothermal control of the interface motion
,”
Phys. Rev. Lett.
110
,
087208
(
2013
).
16.
S.
Brooker
, “
Spin crossover with thermal hysteresis: Practicalities and lessons learnt
,”
Chem. Soc. Rev.
44
,
2880
2892
(
2015
).
17.
R.
Bertoni
,
M.
Lorenc
,
H.
Cailleau
,
A.
Tissot
,
J.
Laisney
,
M.-L.
Boillot
,
L.
Stoleriu
,
A.
Stancu
,
C.
Enachescu
, and
E.
Collet
, “
Elastically driven cooperative response of a molecular material impacted by a laser pulse
,”
Nat. Mater.
15
,
606
610
(
2016
).
18.
K.
Ridier
,
G. A.
Craig
,
F.
Damay
,
T.
Fennell
,
M.
Murrie
, and
G.
Chaboussant
, “
Probing photoinduced spin states in spin-crossover molecules with neutron scattering
,”
Phys. Rev. B
95
,
094403
(
2017
).
19.
B. R.
Mullaney
,
L.
Goux-Capes
,
D. J.
Price
,
G.
Chastanet
,
J.-F.
Létard
, and
C. J.
Kepert
, “
Spin crossover-induced colossal positive and negative thermal expansion in a nanoporous coordination framework material
,”
Nat. Commun.
8
,
1053
(
2017
).
20.
M.
Sy
,
R.
Traiche
,
H.
Fourati
,
Y.
Singh
,
F.
Varret
, and
K.
Boukheddaden
, “
Spatiotemporal investigations on light-driven high-spin-low-spin interface dynamics in the thermal hysteresis region of a spin-crossover single crystal
,”
J. Phys. Chem. C
122
,
20952
20962
(
2018
).
21.
K.
Boukheddaden
,
H.
Fourati
,
Y.
Singh
, and
G.
Chastanet
, “
Evidence of photo-thermal effects on the first-order thermo-induced spin transition of [ Fe ( NCSe ) ( py ) 2 ] 2(m-bpypz)] spin-crossover material
,”
Magnetochemistry
5
,
21
(
2019
).
22.
H.
Köppen
,
E.
Müller
,
C.
Köhler
,
H.
Spiering
,
E.
Meissner
, and
P.
Gütlich
, “
Unusual spin-transition anomaly in the crossover system [ Fe ( 2 pic ) 3 ] Cl 2 EtOH
,”
Chem. Phys. Lett.
91
,
348
352
(
1982
).
23.
V.
Petrouleas
and
J.-P.
Tuchagues
, “
Fe [ 5 NO 2 sal N ( 1 , 4 , 7 , 10 ) ]: A new iron(II) complex exhibiting an unusual two-step spin conversion afforded by a hexadentate ligand with a N 4 O 2, donor set
,”
Chem. Phys. Lett.
137
,
21
25
(
1987
).
24.
R.
Jakobi
,
H.
Spiering
, and
P.
Gütlich
, “
Thermodynamics of the spin transition in [ Fe x Zn 1 x ( 2 pic ) 3 ] Cl 2 EtOH
,”
J. Phys. Chem. Solids
53
,
267
275
(
1992
).
25.
J. A.
Real
,
H.
Bolvin
,
A.
Bousseksou
,
A.
Dworkin
,
O.
Kahn
,
F.
Varret
, and
J.
Zarembowitch
, “
Two-step spin crossover in the new dinuclear compound [Fe(bt)(NCS)2]2bpym, with bt = 2, 2’-bi-2-thiazoline and bpym = 2, 2’-bipyrimidine: Experimental investigation and theoretical approach
,”
J. Am. Chem. Soc.
114
,
4650
4658
(
1992
).
26.
D.
Boinnard
,
A.
Bousseksou
,
A.
Dworkin
,
J. M.
Savariault
,
F.
Varret
, and
J. P.
Tuchagues
, “
Two-step spin conversion of [FeII(5-NO 2-sal-N(1, 4, 7, 10))]: 292, 153, and 103 K x-ray crystal and molecular structure, infrared, magnetic, moessbauer, calorimetric, and theoretical studies
,”
Inorg. Chem.
33
,
271
281
(
1994
).
27.
M.
Buron-Le Cointe
,
N.
Ould Moussa
,
E.
Trzop
,
A.
Moréac
,
G.
Molnar
,
L.
Toupet
,
A.
Bousseksou
,
J. F.
Létard
, and
G. S.
Matouzenko
, “
Symmetry breaking and light-induced spin-state trapping in a mononuclear FeII complex with the two-step thermal conversion
,”
Phys. Rev. B
82
,
214106
(
2010
).
28.
A.
Bousseksou
,
J.
Nasser
,
J.
Linares
,
K.
Boukheddaden
, and
F.
Varret
, “
Ising-like model for the two-step spin-crossover
,”
J. Phys. I
2
,
1381
1403
(
1992
).
29.
M.
Nishino
,
K.
Boukheddaden
,
S.
Miyashita
, and
F.
Varret
, “
Arrhenius Monte Carlo study of two-step spin crossover: Equilibrium and relaxation paths
,”
Phys. Rev. B
68
,
224402
(
2003
).
30.
K.
Boukheddaden
,
J.
Linares
,
E.
Codjovi
,
F.
Varret
,
V.
Niel
, and
J. A.
Real
, “
Dynamical Ising-like model for the two-step spin-crossover systems
,”
J. Appl. Phys.
93
,
7103
7105
(
2003
).
31.
V.
Niel
,
A. L.
Thompson
,
A. E.
Goeta
,
C.
Enachescu
,
A.
Hauser
,
A.
Galet
,
M. C.
Muñoz
, and
J. A.
Real
, “
Thermal- and photoinduced spin-state switching in an unprecedented three-dimensional bimetallic coordination polymer
,”
Chem. Eur. J.
11
,
2047
2060
(
2005
).
32.
N. F.
Sciortino
,
K. R.
Scherl-Gruenwald
,
G.
Chastanet
,
G. J.
Halder
,
K. W.
Chapman
,
J.-F.
Létard
, and
C. J.
Kepert
, “
Hysteretic three-step spin crossover in a thermo- and photochromic 3D pillared Hofmann-type metal-organic framework
,”
Angew. Chem. Int. Ed.
51
,
10154
10158
(
2012
).
33.
Z.-Y.
Li
,
H.
Ohtsu
,
T.
Kojima
,
J.-W.
Dai
,
T.
Yoshida
,
B. K.
Breedlove
,
W.-X.
Zhang
,
H.
Iguchi
,
O.
Sato
,
M.
Kawano
, and
M.
Yamashita
, “
Direct observation of ordered high-spin-low-spin intermediate states of an iron(III) three-step spin-crossover complex
,”
Angew. Chem. Int. Ed.
55
,
5184
5189
(
2016
).
34.
M.
Nihei
,
Y.
Yanai
,
I.-J.
Hsu
,
Y.
Sekine
, and
H.
Oshio
, “
A hydrogen-bonded cyanide-bridged [ Co 2 Fe 2 ] square complex exhibiting a three-step spin transition
,”
Angew. Chem. Int. Ed.
56
,
591
594
(
2017
).
35.
M. J.
Murphy
,
K. A.
Zenere
,
F.
Ragon
,
P. D.
Southon
,
C. J.
Kepert
, and
S. M.
Neville
, “
Guest programmable multistep spin crossover in a porous 2D Hofmann-type material
,”
J. Am. Chem. Soc.
139
,
1330
1335
(
2017
).
36.
Z.-Y.
Li
,
J.-W.
Dai
,
Y.
Shiota
,
K.
Yoshizawa
,
S.
Kanegawa
, and
O.
Sato
, “
Multi-step spin crossover accompanied by symmetry breaking in an FeIII complex: Crystallographic evidence and DFT studies
,”
Chem. Eur. J.
19
,
12948
12952
(
2013
).
37.
J. E.
Clements
,
J. R.
Price
,
S. M.
Neville
, and
C. J.
Kepert
, “
Hysteretic four-step spin crossover within a three-dimensional porous Hofmann-like material
,”
Angew. Chem. Int. Ed.
55
,
15105
15109
(
2016
).
38.
N. F.
Sciortino
,
K. A.
Zenere
,
M. E.
Corrigan
,
G. J.
Halder
,
G.
Chastanet
,
J.-F.
Létard
,
C. J.
Kepert
, and
S. M.
Neville
, “
Four-step iron(II) spin state cascade driven by antagonistic solid state interactions
,”
Chem. Sci.
8
,
701
707
(
2017
).
39.
C.-J.
Zhang
,
K.-T.
Lian
,
G.-Z.
Huang
,
S.
Bala
,
Z.-P.
Ni
, and
M.-L.
Tong
, “
Hysteretic four-step spin-crossover in a 3D Hofmann-type metal-organic framework with aromatic guest
,”
Chem. Commun.
55
,
11033
11036
(
2019
).
40.
C.-J.
Zhang
,
K.-T.
Lian
,
S.-G.
Wu
,
Y.
Liu
,
G.-Z.
Huang
,
Z.-P.
Ni
, and
M.-L.
Tong
, “
The substituent guest effect on four-step spin-crossover behavior
,”
Inorg. Chem. Front.
7
,
911
917
(
2020
).
41.
H.
Watanabe
,
K.
Tanaka
,
N.
Bréfuel
,
H.
Cailleau
,
J.-F. M. C.
Létard
,
S.
Ravy
,
P.
Fertey
,
M.
Nishino
,
S.
Miyashita
, and
E.
Collet
, “
Ordering phenomena of high-spin/low-spin states in stepwise spin-crossover materials described by the ANNNI model
,”
Phys. Rev. B
93
,
014419
(
2016
).
42.
F.
Volatron
,
L.
Catala
,
E.
Rivière
,
A.
Gloter
,
O.
Stéphan
, and
T.
Mallah
, “
Spin-crossover coordination nanoparticles
,”
Inorg. Chem.
47
,
6584
6586
(
2008
).
43.
I.
Boldog
,
A.
Gaspar
,
V.
Martínez
,
P.
Pardo-Ibañez
,
V.
Ksenofontov
,
A.
Bhattacharjee
,
P.
Gütlich
, and
J.
Real
, “
Spin-crossover nanocrystals with magnetic, optical, and structural bistability near room temperature
,”
Angew. Chem. Int. Ed.
47
,
6433
6437
(
2008
).
44.
A.
Tokarev
,
J.
Long
,
Y.
Guari
,
J.
Larionova
,
F.
Quignard
,
P.
Agulhon
,
M.
Robitzer
,
G.
Molnár
,
L.
Salmon
, and
A.
Bousseksou
, “
Spin crossover polysaccharide nanocomposites
,”
New J. Chem.
37
,
3420
3432
(
2013
).
45.
G.
Félix
,
W.
Nicolazzi
,
L.
Salmon
,
G.
Molnár
,
M.
Perrier
,
G.
Maurin
,
J.
Larionova
,
J.
Long
,
Y.
Guari
, and
A.
Bousseksou
, “
Enhanced cooperative interactions at the nanoscale in spin-crossover materials with a first-order phase transition
,”
Phys. Rev. Lett.
110
,
235701
(
2013
).
46.
M.
Mikolasek
,
G.
Félix
,
H.
Peng
,
S.
Rat
,
F.
Terki
,
A. I.
Chumakov
,
L.
Salmon
,
G.
Molnár
,
W.
Nicolazzi
, and
A.
Bousseksou
, “
Finite-size effects on the lattice dynamics in spin crossover nanomaterials. I. Nuclear inelastic scattering investigation
,”
Phys. Rev. B
96
,
035426
(
2017
).
47.
A. C.
Felts
,
A.
Slimani
,
J. M.
Cain
,
M. J.
Andrus
,
A. R.
Ahir
,
K. A.
Abboud
,
M. W.
Meisel
,
K.
Boukheddaden
, and
D. R.
Talham
, “
Control of the speed of a light-induced spin transition through mesoscale core-shell architecture
,”
J. Am. Chem. Soc.
140
,
5814
5824
(
2018
).
48.
Y.-X.
Wang
,
D.
Qiu
,
S.-F.
Xi
,
Z.-D.
Ding
,
Z.
Li
,
Y.
Li
,
X.
Ren
, and
Z.-G.
Gu
, “
Iron(II)-triazole core-hell nanocomposites: Toward multistep spin crossover materials
,”
Chem. Commun.
52
,
8034
8037
(
2016
).
49.
K.
Boukheddaden
,
I.
Shteto
,
B.
Hôo
, and
F.
Varret
, “
Dynamical model for spin-crossover solids. I. Relaxation effects in the mean-field approach
,”
Phys. Rev. B
62
,
14796
14805
(
2000
).
50.
K.
Boukheddaden
,
I.
Shteto
,
B.
Hôo
, and
F.
Varret
, “
Dynamical model for spin-crossover solids. II. Static and dynamic effects of light in the mean-field approach
,”
Phys. Rev. B
62
,
14806
14817
(
2000
).
51.
M.
Nishino
,
S.
Miyashita
, and
K.
Boukheddaden
, “
Effective interaction range in the spin crossover phenomenon: Wajnflasz and domain models
,”
J. Chem. Phys.
118
,
4594
4597
(
2003
).
52.
S.
Miyashita
,
Y.
Konishi
,
H.
Tokoro
,
M.
Nishino
,
K.
Boukheddaden
, and
F.
Varret
, “
Structures of metastable states in phase transitions with a high-spin low-spin degree of freedom
,”
Prog. Theor. Phys.
114
,
719
735
(
2005
).
53.
M.
Nishino
,
K.
Boukheddaden
,
Y.
Konishi
, and
S.
Miyashita
, “
Simple two-dimensional model for the elastic origin of cooperativity among spin states of spin-crossover complexes
,”
Phys. Rev. Lett.
98
,
247203
(
2007
).
54.
Y.
Konishi
,
H.
Tokoro
,
M.
Nishino
, and
S.
Miyashita
, “
Monte Carlo simulation of pressure-induced phase transitions in spin-crossover materials
,”
Phys. Rev. Lett.
100
,
067206
(
2008
).
55.
S.
Miyashita
,
Y.
Konishi
,
M.
Nishino
,
H.
Tokoro
, and
P. A.
Rikvold
, “
Realization of the mean-field universality class in spin-crossover materials
,”
Phys. Rev. B
77
,
014105
(
2008
).
56.
W.
Nicolazzi
,
S.
Pillet
, and
C.
Lecomte
, “
Two-variable anharmonic model for spin-crossover solids: A like-spin domains interpretation
,”
Phys. Rev. B
78
,
174401
(
2008
).
57.
M.
Nishino
,
K.
Boukheddaden
, and
S.
Miyashita
, “
Molecular dynamics study of thermal expansion and compression in spin-crossover solids using a microscopic model of elastic interactions
,”
Phys. Rev. B
79
,
012409
(
2009
).
58.
C.
Enachescu
,
L.
Stoleriu
,
A.
Stancu
, and
A.
Hauser
, “
Model for elastic relaxation phenomena in finite 2D hexagonal molecular lattices
,”
Phys. Rev. Lett.
102
,
257204
(
2009
).
59.
S.
Miyashita
,
P. A.
Rikvold
,
T.
Mori
,
Y.
Konishi
,
M.
Nishino
, and
H.
Tokoro
, “
Threshold phenomena under photoexcitation of spin-crossover materials with cooperativity due to elastic interactions
,”
Phys. Rev. B
80
,
064414
(
2009
).
60.
W.
Nicolazzi
,
S.
Pillet
, and
C.
Lecomte
, “
Photoinduced phase separation in spin-crossover materials: Numerical simulation of a dynamic photocrystallographic experiment
,”
Phys. Rev. B
80
,
132102
(
2009
).
61.
M.
Nishino
,
C.
Enachescu
,
S.
Miyashita
,
K.
Boukheddaden
, and
F.
Varret
, “
Intrinsic effects of the boundary condition on switching processes in effective long-range interactions originating from local structural change
,”
Phys. Rev. B
82
,
020409
(
2010
).
62.
C.
Enachescu
,
M.
Nishino
,
S.
Miyashita
,
A.
Hauser
,
A.
Stancu
, and
L.
Stoleriu
, “
Cluster evolution in spin crossover systems observed in the frame of a mechano-elastic model
,”
Europhys. Lett.
91
,
27003
(
2010
).
63.
M.
Nishino
,
C.
Enachescu
,
S.
Miyashita
,
P. A.
Rikvold
,
K.
Boukheddaden
, and
F.
Varret
, “
Macroscopic nucleation phenomena in continuum media with long-range interactions
,”
Sci. Rep.
1
,
162
(
2011
).
64.
L.
Stoleriu
,
P.
Chakraborty
,
A.
Hauser
,
A.
Stancu
, and
C.
Enachescu
, “
Thermal hysteresis in spin-crossover compounds studied within the mechanoelastic model and its potential application to nanoparticles
,”
Phys. Rev. B
84
,
134102
(
2011
).
65.
T.
Nakada
,
T.
Mori
,
S.
Miyashita
,
M.
Nishino
,
S.
Todo
,
W.
Nicolazzi
, and
P. A.
Rikvold
, “
Critical temperature and correlation length of an elastic interaction model for spin-crossover materials
,”
Phys. Rev. B
85
,
054408
(
2012
).
66.
C.
Enachescu
,
M.
Nishino
,
S.
Miyashita
,
L.
Stoleriu
, and
A.
Stancu
, “
Monte Carlo metropolis study of cluster evolution in spin-crossover solids within the framework of a mechanoelastic model
,”
Phys. Rev. B
86
,
054114
(
2012
).
67.
W.
Nicolazzi
and
S.
Pillet
, “
Structural aspects of the relaxation process in spin crossover solids: Phase separation, mapping of lattice strain, and domain wall structure
,”
Phys. Rev. B
85
,
094101
(
2012
).
68.
M.
Nishino
and
S.
Miyashita
, “
Effect of the short-range interaction on critical phenomena in elastic interaction systems
,”
Phys. Rev. B
88
,
014108
(
2013
).
69.
M.
Nishino
,
T.
Nakada
,
C.
Enachescu
,
K.
Boukheddaden
, and
S.
Miyashita
, “
Crossover of the roughness exponent for interface growth in systems with long-range interactions due to lattice distortion
,”
Phys. Rev. B
88
,
094303
(
2013
).
70.
A.
Slimani
,
K.
Boukheddaden
,
F.
Varret
,
H.
Oubouchou
,
M.
Nishino
, and
S.
Miyashita
, “
Microscopic spin-distortion model for switchable molecular solids: Spatiotemporal study of the deformation field and local stress at the thermal spin transition
,”
Phys. Rev. B
87
,
014111
(
2013
).
71.
K.
Boukheddaden
, “
Monte Carlo investigations on surface elastic energy of spin-crossover solids: Direct access to image pressure and the Eshelby constant
,”
Phys. Rev. B
88
,
134105
(
2013
).
72.
A.
Slimani
,
K.
Boukheddaden
,
F.
Varret
,
M.
Nishino
, and
S.
Miyashita
, “
Properties of the low-spin high-spin interface during the relaxation of spin-crossover materials, investigated through an electro-elastic model
,”
J. Chem. Phys.
139
,
194706
(
2013
).
73.
M.
Nishino
and
S.
Miyashita
, “
Termination of the Berezinskii-Kosterlitz-Thouless phase with a new critical universality in spin-crossover systems
,”
Phys. Rev. B
92
,
184404
(
2015
).
74.
C.
Enachescu
,
M.
Nishino
,
S.
Miyashita
,
K.
Boukheddaden
,
F.
Varret
, and
P. A.
Rikvold
, “
Shape effects on the cluster spreading process of spin-crossover compounds analyzed within an elastic model with Eden and Kawasaki dynamics
,”
Phys. Rev. B
91
,
104102
(
2015
).
75.
H.-Z.
Ye
,
C.
Sun
, and
H.
Jiang
, “
Monte-Carlo simulations of spin-crossover phenomena based on a vibronic Ising-like model with realistic parameters
,”
Phys. Chem. Chem. Phys.
17
,
6801
6808
(
2015
).
76.
M.
Paez-Espejo
,
M.
Sy
, and
K.
Boukheddaden
, “
Elastic frustration causing two-step and multistep transitions in spin-crossover solids: Emergence of complex antiferroelastic structures
,”
J. Am. Chem. Soc.
138
,
3202
3210
(
2016
).
77.
C.
Enachescu
,
L.
Stoleriu
,
M.
Nishino
,
S.
Miyashita
,
A.
Stancu
,
M.
Lorenc
,
R.
Bertoni
,
H.
Cailleau
, and
E.
Collet
, “
Theoretical approach for elastically driven cooperative switching of spin-crossover compounds impacted by an ultrashort laser pulse
,”
Phys. Rev. B
95
,
224107
(
2017
).
78.
L.
Stoleriu
,
M.
Nishino
,
S.
Miyashita
,
A.
Stancu
,
A.
Hauser
, and
C.
Enachescu
, “
Cluster evolution in molecular three-dimensional spin-crossover systems
,”
Phys. Rev. B
96
,
064115
(
2017
).
79.
M.
Nishino
,
S.
Miyashita
, and
P. A.
Rikvold
, “
Nontrivial phase diagram for an elastic interaction model of spin crossover materials with antiferromagnetic-like short-range interactions
,”
Phys. Rev. B
96
,
144425
(
2017
).
80.
C.
Enachescu
and
W.
Nicolazzi
, “
Elastic models, lattice dynamics and finite size effects in molecular spin crossover systems
,”
C. R. Chim.
21
,
1179
1195
(
2018
).
81.
M.
Nishino
,
C.
Enachescu
, and
S.
Miyashita
, “
Multistep spin-crossover transitions induced by the interplay between short- and long-range interactions with frustration on a triangular lattice
,”
Phys. Rev. B
100
,
134414
(
2019
).
82.
M.
Ndiaye
and
K.
Boukheddaden
, “
Electro-elastic modelling of the two-step high-spin to low-spin relaxation with transient self-organized spin states in 2D spin crossover solids
,”
J. Phys. Soc. Jpn.
89
,
014004
(
2020
).
83.
P. A.
Rikvold
,
G.
Brown
,
S.
Miyashita
,
C.
Omand
, and
M.
Nishino
, “
Equilibrium, metastability, and hysteresis in a model spin-crossover material with nearest-neighbor antiferromagnetic-like and long-range ferromagnetic-like interactions
,”
Phys. Rev. B
93
,
064109
(
2016
).
84.
M.
Nishino
,
P. A.
Rikvold
,
C.
Omand
, and
S.
Miyashita
, “
Multistability in an unusual phase diagram induced by the competition between antiferromagnetic-like short-range and ferromagnetic-like long-range interactions
,”
Phys. Rev. B
98
,
144402
(
2018
).
85.
J.
Cruddas
and
B. J.
Powell
, “
Spin-state ice in elastically frustrated spin-crossover materials
,”
J. Am. Chem. Soc.
141
,
19790
19799
(
2019
).
86.
J.
Cruddas
and
B. J.
Powell
, “
Structure-property relationships and the mechanisms of multistep transitions in spin crossover materials and frameworks
,”
Inorg. Chem. Front.
7
,
4424
4437
(
2020
).
87.
H.
Oubouchou
,
A.
Slimani
, and
K.
Boukheddaden
, “
Interplay between elastic interactions in a core-shell model for spin-crossover nanoparticles
,”
Phys. Rev. B
87
,
104104
(
2013
).
88.
A.
Slimani
,
H.
Khemakhem
, and
K.
Boukheddaden
, “
Structural synergy in a core-shell spin crossover nanoparticle investigated by an electroelastic model
,”
Phys. Rev. B
95
,
174104
(
2017
).
89.
H.
Oubouchou
,
Y.
Singh
, and
K.
Boukheddaden
, “
Magnetoelastic modeling of core-shell spin-crossover nanocomposites
,”
Phys. Rev. B
98
,
014106
(
2018
).
90.
Y.
Singh
,
H.
Oubouchou
,
M.
Nishino
,
S.
Miyashita
, and
K.
Boukheddaden
, “
Elastic-frustration-driven unusual magnetoelastic properties in a switchable core-shell spin-crossover nanostructure
,”
Phys. Rev. B
101
,