Next Article in Journal
Readout Methods to Enhance the Performance of Luminescence Thermometers
Previous Article in Journal
Simple Model for Tc and Pairing Symmetry Changes in Sr2RuO4 Under (100) Uniaxial Strain
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Magnetization Plateaus by the Field-Induced Partitioning of Spin Lattices

by
Myung-Hwan Whangbo
1,*,
Hyun-Joo Koo
2,
Reinhard K. Kremer
3 and
Alexander N. Vasiliev
4,*
1
Department of Chemistry, North Carolina State University, Raleigh, NC 27695-8204, USA
2
Department of Chemistry, Research Institute for Basic Sciences, Kyung Hee University, Seoul 02447, Republic of Korea
3
Max Planck Institute for Solid State Research, Heisenbergstrasse 1, D-70569 Stuttgart, Germany
4
Department of Low Temperature Physics and Superconductivity, Lomonosov Moscow State University, Moscow 119991, Russia
*
Authors to whom correspondence should be addressed.
Condens. Matter 2024, 9(4), 45; https://doi.org/10.3390/condmat9040045
Submission received: 26 September 2024 / Revised: 17 October 2024 / Accepted: 22 October 2024 / Published: 4 November 2024

Abstract

:
To search for a conceptual picture describing the magnetization plateau phenomenon, we surveyed the crystal structures and the spin lattices of those magnets exhibiting plateaus in their magnetization vs. magnetic field curves by probing the three questions: (a) why only certain magnets exhibit magnetization plateaus, (b) why there occur several different types of magnetization plateaus, and (c) what controls the widths of magnetization plateaus. We show that the answers to these questions lie in how the magnets under field absorb Zeeman energy, hence changing their magnetic structures. The magnetic structure of a magnet insulator is commonly described in terms of its spin lattice, which requires the determination of the spin exchanges’ nonnegligible strengths between the magnetic ions. Our work strongly suggests that a magnet under the magnetic field partitions its spin lattice into antiferromagnetic (AFM) or ferrimagnetic fragments by breaking its weak magnetic bonds. Our supposition of the field-induced partitioning of spin lattices into magnetic fragments is supported by the anisotropic magnetization plateaus of Ising magnets and by the highly anisotropic width of the 1/3-magnetization plateau in azurite. The answers to the three questions (a)–(c) emerge naturally by analyzing how these fragments are formed under the magnetic field.

1. Introduction

The properties of a magnet are primarily characterized by measuring thermodynamic quantities (e.g., magnetization M and/or magnetic specific heat Cmag) as a function of the temperature and external magnetic field. The values of M(H, T) and Cmag(H, T) for a given magnet depend on its magnetic energy spectrum and on how the individual states of this spectrum are thermally populated. The individual magnetic states differ in their magnetic properties, and their population at a given temperature is governed by the Boltzmann factor, so the thermodynamic quantity is a weighted average of the properties of various states with weights given by their Boltzmann factors at that particular temperature. As a function of temperature, the Boltzmann populations of the individual states changes. States with lower energy become exponentially more populated as the temperature is decreased. Thus, measuring the temperature dependence of the M or Cmag of a magnet is an indirect way of probing its magnetic energy spectrum. To confirm whether or not a magnet undergoes a long-range magnetic ordering as the temperature is lowered is often judged from the temperature dependence of its specific heat. The occurrence of such an ordering is signaled, e.g., by the presence of an anomaly in the Cmag vs. T curve, which reflects the loss of the magnetic entropy associated with the long-range magnetic ordering.
Often, to a very good approximation, the energy spectrum of a magnet can be described by the Heisenberg spin Hamiltonian H ^ s p i n (Equation (1)), which is written as the sum of the pairwise spin exchange interactions between spin operators S ^ i and S ^ j located at the magnetic ion sites i and j, respectively.
H ^ s p i n = i > j J i j S ^ i · S ^ j
Often, the spin operators S ^ i and S ^ j can be regarded as the classical spin vectors S i and S j , respectively, and hence the spin Hamiltonian has the classical expression,
H s p i n = i > j J i j S i · S j = i > j J i j S i S j cos θ
where θ is the angle between the two spin vectors S i   a n d S j .
Here, we consider the simplest case of the symmetric spin exchange and omit contributions of antisymmetric exchange to the Hamiltonian or anisotropies in the exchange. Being a dot product, the extrema of S i · S j occur when the two spins are parallel and antiparallel to each other, respectively. Thus, with the spin Hamiltonian defined as in Equations (1) and (2), the two spins prefer an AFM coupling if the spin exchange is positive ( J i j > 0 ), but a ferromagnetic (FM) coupling if it is negative ( J i j < 0 ). The “spin lattice” of a magnet refers to the repeat pattern of its spin exchange paths of nonnegligible strengths. If we consider such spin exchange paths as magnetic bonds, then the spin lattice of a magnet is the lattice of its magnetic bonds of various strengths. The spin lattice of a magnet is crucially important because it allows one to generate the energy spectrum relevant for the magnet using a model Hamiltonian H s p i n with a minimal number of spin exchange parameters.
Magnets including Heisenberg magnetic ions, with nonzero quantized spin moments in all directions, are described by the Heisenberg spin Hamiltonian, for which S i · S j   = S i x S j x + S i y S j y + S i z S j z . The magnets of uniaxial (i.e., Ising) magnetic ions, which possess nonzero spin moments only in one direction (by convention, the z-direction) so that S i · S j = S i z S j z , are described by the Ising spin Hamiltonian,
H I s i n g = i > j J i j S i z S j z
Transition-metal magnetic ions M in oxide magnets ions form MOn (typically, n = 3–6) polyhedra. The Ising magnetism is found for a magnetic ion M when the d-states of its MOn polyhedron has an unevenly occupied degenerate d-state (in the non-spin-polarized, one-electron picture of electronic structure description) [1,2]. Such a magnetic ion is susceptible to a Jahn–Teller distortion, which tends to lift, though weakly, the degeneracy responsible for the Ising magnetism [3]. Thus, pure Ising magnets are rather rare.
The dependence of magnetization M on the external magnetic field μ0H is usually measured at the lowest possible temperature to minimize the contributions of magnetic excited states lying close to the magnetic ground state through the Boltzmann averaging. As a function of the magnetic field μ0H, the magnetization of a paramagnet is well described by a Brillouin function, which increases steadily from zero to the magnetic saturation Msat. On increasing the magnetic field μ0H, the magnetization of an antiferromagnet exhibits spin flop and spin flip transitions (see below) while that of a ferromagnet rapidly reaches the saturation, depending on anisotropy and dipolar energies.
For some magnets among the wide variety of magnetic materials, their magnetization versus magnetic field (M vs. H) curves exhibit s plateaus at rational fractions f = m/n, where m and n are integers with m < n (most commonly, m = 1) of their saturation magnetization Msat. This phenomenon occurs not only in magnets undergoing a long-range magnetic order at low temperatures but also in low-dimensional or spin-frustrated magnets that do not undergo a magnetic order down to the lowest temperatures. In discussing the M vs. H curves observed for such magnets, it is convenient to distinguish magnets with and without uniaxial (i.e., Ising) anisotropy [1,2]. The idealized M vs. H curves observed for non-Ising magnets are illustrated in Figure 1a–c, and those for Ising magnets in Figure 1d–f. Consider first the M vs. H behaviors of non-Ising magnets. On increasing the magnetic field from zero, a gradual increase in M from zero to (m/n)Msat precedes before reaching the m/n-magnetization plateau at M = (m/n)Msat (Figure 1a), the (m/n)-magnetization plateau occurs as soon as the field increases from zero (Figure 1b), or the zero-magnetization plateau at M = 0 precedes until the field reaches a value from which a gradual increase in M to (m/n)Msat begins (Figure 1c). For an Ising magnet, the spin moment is nonzero only along one specific direction in space. An Ising magnet exhibits a highly anisotropic M vs. H behavior; its M vs. H curve exhibits a “step-like” feature when the applied field is parallel to the direction of the spin moment (Figure 1d,e), i.e., the easy axis direction, but the magnetization does not change with field showing no magnetization plateau when the applied field is perpendicular to the spin moment direction (Figure 1f). Experimentally, a very slight linear increase with field is observed, but this is often due to a minute misalignment of the crystal with respect to the external magnetic field.
Magnetic plateaus have been found in a large variety of magnets. Their spin lattices can be one-dimensional, two-dimensional (2D) or three-dimensional (3D), their ground state can be AFM or ferrimagnetic in the absence of the external magnetic field, their spin lattices may or may not be spin-frustrated, and their structures can be extended or discrete. For magnets of high symmetry, a number of theoretical studies examined their magnetization plateaus from the viewpoint of their magnetic energy spectra generated by model spin Hamiltonians [4,5]. So far, however, there has been no systematic study aimed at providing a conceptual picture for the magnetization plateau phenomenon. The primary objective of our survey is to come up with a conceptual framework useful for chemists, materials scientists and experimental physicists in organizing and thinking about magnetization plateaus. Therefore, we pursue the qualitative answers to the three questions (a)–(c) by analyzing not only the structural chemistry associated with the magnetic ions but also the relative magnitudes and the signs of the spin exchange interactions between them. Our study strongly suggests that the spin lattice of a magnet exhibiting a magnetic plateau is partitioned into ferrimagnetic or AFM fragments by breaking the weakest magnetic bonds one at a time by absorbing Zeeman energy provided by an external magnetic field. The M vs. H curve of a magnet is divided into two different regions: the regions where a magnet does not absorb Zeeman energy so nonzero (m/n > 0) magnetic plateaus occur and the regions where the magnet absorbs Zeeman energy so no magnetization plateau, except for the zero (m/n = 0) magnetization plateau, occurs.
Our survey is organized as follows: In Section 2, we analyze why the spin lattice of a certain magnet is partitioned into smaller magnetic fragments and what types of magnetization plateaus are possible. Section 3 describes the magnetization plateaus of magnets whose spin lattices are partitioned into AFM fragments (with an even number of spin sites) under field, and in Section 4 those of magnets whose spin lattices are partitioned into ferrimagnetic fragments (with an odd number of spin sites) under field. The magnetization plateaus of magnets possessing kagomé and trigonal layers are discussed in Section 5, and those of magnets with complex magnetic fragments in Section 6. Finally, our conclusions are summarized in Section 7.
We note that this work is not a comprehensive review on magnetization plateaus, but a survey on studies of magnetization plateaus that enabled us to put forward the concept that a magnet under field absorbs Zeeman energy by breaking its weak magnetic bonds. The associated partitioning of its spin lattice into magnetic fragments gives rise to magnetization plateaus. For magnets with low-symmetry crystal structures possessing a large number of magnetic ions per unit cell, describing their magnetic structures quantitatively using a model spin Hamiltonian is practically impossible. This difficulty has led us to search for a qualitative description of such magnets on the basis of their spin exchanges (i.e., magnetic bonds), because they can be readily determined by employing density functional theory (DFT) calculations. Our studies on numerous such magnets over the past two decades revealed that the spin exchanges obtained from DFT calculations are quite accurate in their relative magnitudes and are therefore reliable in finding which magnetic bonds are weak and hence will be broken preferentially under field. This realization led us to the concept of the field-induced partitioning of a spin lattice into magnetic fragments, initially from our own studies on magnets exhibiting magnetic plateaus. We then checked whether this concept is applicable to other magnets for which magnetization plateaus were reported. When the spin exchanges of these magnets were not available, we determined them by performing DFT calculations as summarized in the supporting information. This survey is the outcome of these efforts. Our choice of references is not as comprehensive as expected for a review article but is rather confined to those central to our supposition of the field-induced partitioning of a spin lattice into magnetic fragments.

2. Field-Induced Partitioning of Spin Lattices

2.1. Zeeman Energy and Magnetic Bonds

Two spins of an AFM exchange path tend to align antiparallel to each other, so it requires energy to force them to be ferromagnetically aligned. At very low temperatures where magnetization measurements are usually carried out, the energy needed for such a conversion in a spin exchange path of a magnet is supplied by Zeeman energy, EZ. For a magnetic ion with spin moment μ s = g μ B S under a magnetic field μ 0 H , the Zeeman energy is given by
E Z = g μ 0 μ B H · S ,
which is positive and negative if H and S are parallel and antiparallel, respectively. As the magnetic field is gradually increased from zero, the conversion from antiparallel to parallel spin alignment occurs initially in weak AFM exchange paths, i.e., those with small exchange J are converted first. For the convenience of our discussion, an AFM exchange path will be termed “a magnetic bond” if the two spins of the path are antiferromagnetically coupled. Likewise, an AFM exchange path may be termed “a broken magnetic bond” or “a magnetic antibond” if the two spins of the path are forced to be ferromagnetically coupled. Thus, an AFM magnetic bond can be broken by the Zeeman energy EZ (Figure 2a). Similarly, an FM exchange path may be termed “a magnetic bond” if the two spins of the path are ferromagnetically aligned, but “a broken magnetic bond or a “magnetic antibond” if the two spins of the path are antiferromagnetically aligned. Thus, an FM magnetic bond can be broken by Zeeman energy (Figure 2b). For the convenience of our discussion, we will represent the up-spin and down-spin at a magnetic ion site by unshaded and shaded circles, respectively (Figure 2). [Here, the up-spin and down-spin are parallel and antiparallel to the direction of an external magnetic field (taken to be the z-direction by convention), respectively. In the absence of an external field, what matters is that the up-spin and down-spin are antiparallel to each other, regardless of their absolute directions in space.] Breaking an AFM magnetic bond increases the spin moment (Figure 2a), while breaking an FM bond can either decrease or increase the spin moment (Figure 2b) because an FM coupling in an FM magnetic bond can be represented by the (↑↑) or (↓↓) spin arrangement. It is the FM bond breaking from (↓↓) to (↑↓), not from (↑↑) to (↑↓), that is relevant for our discussion of magnetization because the total magnetic moment should not decrease with field (see below for further discussion). Summarizing, magnetic bonds should be referred to as either AFM magnetic bonds or FM magnetic bonds. However, for most magnets showing magnetization plateaus, one deals with breaking AFM magnetic bonds, and it is very rare to find magnets whose magnetization requires the breaking of FM magnetic bonds. Thus, in the following, we use the term “magnetic bonds” to describe AFM bonds, unless stated otherwise.
To avoid a possible confusion in using the terminology, the broken or unbroken magnetic bond, it is necessary to distinguish between the eigenstates and the broken-symmetry states. For all practical evaluations of spin exchanges for any magnet, broken-symmetry states are used instead of the eigenstates simply because the latter are very difficult to determine [6]. For example, consider a spin dimer made up of two S = 1/2 magnetic ions representing, for example, the molecular Cu2Cl62− ion of edge-sharing CuCl4 square planes (see Section 3.2.2). This dimer can be described by the singlet and triplet states |S〉 and |T〉, which are the eigenstates of the dimer (Figure 3a). Then, the energy difference between the two states is the spin exchange J, i.e., ETES = J (Figure 3b). If the dimer is described by the broken-symmetry states ↑↑ and ↑↓, then the energy difference between the two states is given by E↑↑E↑↓ = J/2 (Figure 3c). In the following, by breaking an AFM J bond in an extended magnet, we mean the conversion from the AFM coupling ↑↓ to the FM coupling ↑↑. In the case of an isolated dimer, the breaking of the AFM J bond means the excitation from the singlet to the triplet state.

2.2. Causes for Magnetization Plateaus

In the following, we put forward the supposition that during the magnetization process the spin lattice of a magnet becomes partitioned into either AFM or ferrimagnetic fragments by breaking its weak magnetic bonds. As an example, consider how a 0-magnetization plateau arises by considering a chain in which AFM dimers described by spin exchange J1 are antiferromagnetically coupled in a tail-to-tail bridging pattern to make J2 bonds between adjacent dimers such that J1 > J2. An AFM chain made up of alternating J1 and J2 bonds is presented in Figure 4a. Since J2 is weaker than J1, the J2 bond will be broken “successively” (Figure 4b,c) as the magnetic field is increased from 0 until all J2 bonds are broken (Figure 4d). In Figure 4b–d, the red ellipses are used to indicate that the (↓↑) dimers, resulting from the (↑↓) dimers of Figure 4a, break the interdimer bonds J2. The energy needed for breaking a J2 bond is supplied by Zeeman energy, but the magnetization remains at zero while the field increases until all J2 bonds are broken because the spins stay paired in the J1 bonds. This leads to a 0-magnetization plateau (e.g., Figure 1c).
As the magnetic field increases further, the J1 bonds become broken one at a time, as depicted in Figure 4e–g, where the green ellipses are used to indicate the broken dimers, (↑↑), resulting from the (↓↑) dimers of Figure 4d. In principle, a broken dimer can be equally well represented by the configuration (↓↓). However, throughout our discussion, a broken dimer will be represented by (↑↑), because the spin moments of a magnet under the magnetic field should not decrease with field and because we use the convention that up-spin and down-spin have positive and negative moments, respectively. Since each J1 bond breaking creates unpaired up-spins, the magnetization M increases with the field: M = Msat/4, if one out of four J1 bonds is broken (Figure 4e), M = Msat/3 if one out of three J1 bonds is broken (Figure 4f), and M = Msat/2 if one out of two J1 bonds is broken (Figure 4g).
In general, when the spin lattice of a magnet becomes partitioned into identical AFM fragments by breaking its weak magnetic bonds interconnecting them, the magnet acquires a zero-magnetization given by the AFM fragment regardless of how many inter-fragment magnetic bonds there are. If there exist a large number of different arrangements between the AFM fragments, which differ only in the number of their inter-fragment magnetic bonds, then the magnetization of a magnet remains unchanged while the inter-fragment magnetic bonds are being broken by increasing the magnetic field. The 0-magnetization plateau of the AFM chain made up of AFM dimers discussed above is an example. To increase the magnetization beyond this level, a weak magnetic bond of each AFM fragment needs to be broken.
Suppose that the spin lattice of a magnet becomes partitioned into identical ferrimagnetic fragments when the field increases by breaking its weak inter-fragment magnetic bonds. Then, the magnetization increases gradually with field until all inter-fragment bonds are broken so that all ferrimagnetic fragments become ferromagnetically coupled, leading to a certain level of magnetization given by the ferrimagnetic fragments. To increase the magnetization beyond this level, it is necessary to break a weak magnetic bond of each ferrimagnetic fragment. If this magnetic bond is strong, the bond breaking will not happen unless the field reaches a high enough value, hence leading to a magnetization plateau.

2.3. Magnetic Bonding Pattern Affecting the Nature of Magnetization Plateaus

Consider now an AFM chain in which ferrimagnetic linear trimers with the (↑↓↑) configuration are antiferromagnetically coupled in a tail-to-tail pattern (Figure 5a) so that the (↑↓↑) and (↓↑↓) trimers alternate. We assume that the intra-trimer J1 bond is stronger than the inter-trimer bond J2. Then, as the field increases from 0, the J2 bonds become broken one at a time (Figure 5b) until all J2 bonds are broken (Figure 5c) by converting each (↓↑↓) trimer to a (↑↓↑) trimer, as indicated by the red ellipses. Since each trimer constitutes a ferrimagnetic unit, the magnetization increases with the number of broken J2 bonds until it reaches Msat/3, where all J2 bonds are broken. When the magnetic field is further raised, the bonds to break are the two J1 bonds in each (↑↓↑) trimer as indicated by the green ellipse in Figure 5d. The magnetic bonds J1 are strong, and two J1 bonds of a trimer should be broken simultaneously. Therefore, until the field reaches a high enough value, they are not broken hence leading to no increase in the magnetization. The magnetization plot of Figure 1a is a characteristic feature of an AFM chain in which ferrimagnetic fragments are antiferromagnetically coupled in a tail-to-tail manner.
When the ferrimagnetic linear trimers are combined in a head-to-tail bridging pattern, the resulting chain is a ferrimagnetic chain (Figure 6a), with magnetization M = Msat/3. Under a magnetic field, a J2 bond of this ferrimagnetic chain cannot be broken because, if broken, the resulting ferrimagnetic trimer will have the (↓↑↓) configuration (indicated by the red ellipse in Figure 6b) and hence will reduce the overall moment of the chain. Therefore, the only way to absorb the Zeeman energy is to break the two J1 bonds of a ferrimagnetic trimer successively (indicated by the green ellipses in Figure 6c,d) hence increasing the magnetization toward Msat. The J1 bond is strong and a simultaneous breaking of two J1 bonds requires high energy, so this will not occur until the applied field is high enough. Thus, the 1/3-magnetization plateau will be wide. The magnetization curve of the ferrimagnetic chain is well described by Figure 1b.

2.4. Field-Induced Ferrimagnetic Fragments in Spin-Frustrated Lattices

As described above, field-induced partitioning of a spin lattice into magnetic fragments lies at the heart of the magnetization plateau phenomenon. In understanding this field-induced partitioning, it is crucial to identify the weak magnetic bonds of a given spin lattice that can be readily broken by moderate magnetic fields. The absorption of Zeeman energy by a magnet is a consequence of the Le Chatelier’s principle. There are cases when it is not immediately obvious how a spin lattice under field will be partitioned into magnetic fragments when the spin lattice defined by a few spin exchanges of comparable magnitude is spin-frustrated, e.g., trigonal, kagomé and diamond chain spin lattices. For such cases as well, Le Chatelier’s principle enables us to put forward the supposition that a spin lattice is partitioned into small ferrimagnetic fragments of nonzero spin S such that these fragments fill the spin lattice without overlapping between them. This partitioning reduces the extent of spin frustration by absorbing Zeeman energy and hence breaking the inter-fragment bonds and making the ferrimagnetic fragment absorb Zeeman energy further when the field is raised. For example, consider a magnet consisting of diamond chains made up of two AFM spin exchange J1 and J2 with J1 > J2 (Figure 7a). The weaker magnetic bonds J2 form a continuous chain, but the 1/3-magnetization phenomenon observed for such a magnet can be readily understood by supposing that the diamond chain undergoes a field-induced partitioning into triangular ferrimagnetic fragments as depicted by shaded triangles in Figure 7b,c.
Each triangular fragment can have six possible spin arrangements (Figure 8). In the absence of an applied magnetic field, the (↑↑↑) and (↓↓↓) trimers are less stable than the other four ferrimagnetic trimers, and the (↑↓↓) and (↓↑↓) trimers with net negative moment are as stable as the (↓↑↑) and (↑↓↑) trimers with net positive moment. Applying a magnetic field stabilizes the latter but destabilizes the former. Likewise, an applied field stabilizes the (↑↑↑) trimer while destabilizing the (↓↓↓) trimer. Thus, for the discussion of the 1/3-magnetization plateau of the diamond chains, only the ferrimagnetic (↑↓↑) or (↓↑↑) trimers are relevant.
Let us consider theoretical and experimental justifications for our supposition that a spin-frustrated spin lattice will undergo a field-induced partitioning into small ferrimagnetic fragments of nonzero spin S . If a magnet can produce such ferrimagnetic fragments, the energy of each ferrimagnetic fragment is raised by Zeeman energy, E Z = g μ 0 μ B H S , which can be used for the magnet to break the magnetic bonds necessary for the partitioning. However, if a magnet cannot interact with the magnetic field, such fragmentation cannot occur so that the magnet cannot develop any magnetization plateau. An experimental test for these predictions is provided by magnetization studies on Ising magnets. The spins of an Ising magnet are nonzero in one direction in space (say, the ||z direction) but zero in all directions perpendicular to this direction (i.e., ⊥z). For field H | | along the ||z direction, the Zeeman energy is nonzero ( H | | · S ≠0). For field H along the ⊥z direction, however, the Zeeman energy is zero ( H · S = 0). (Here, we have assumed that the anisotropy energy that forced the spins to align either parallel or antiparallel to z is much larger than the Zeeman energy). Therefore, an Ising magnet can have a magnetization plateau when the field is along the ||z direction but cannot if the field is along the ⊥z direction (Figure 1d–f). The step-like feature of the magnetization curves found for Ising magnets under field H | | indicates that a large number of ferrimagnetic fragments develop simultaneously, because spin flipping is necessary for magnetic bond breaking.

2.5. Spin-Lattice Interactions

In our discussion of magnetic plateaus so far, the field-induced partitioning of a spin lattice into ferrimagnetic or antiferromagnetic fragments is discussed without considering spin-lattice interactions. The magnetic fragments broken off differ in their surroundings from the spin lattice (e.g., Figure 5a–c) and hence would require some relaxation in their atomic/electronic structures through magnetoelastic coupling. The concomitant change in the spin-lattice interaction can therefore affect the stability and nature of a magnetization plateau. A magnetization plateau generates a certain pattern of up-spin and down-spin arrangement and hence the associated spin-lattice interactions. When the spin-lattice interaction, associated with a certain magnetization plateau, leads to an energy-lowering relaxation, this magnetization plateau will arise. Otherwise, it will not be observed. It goes without saying that the spin-lattice interaction will be important for magnets of compact and high-symmetry atomic structure, because the pattern of up-spin and down-spin arrangement should be compatible with the symmetry of the underlying spin lattice.
Magnets of spinel structure such as CdCr2O4 and LiGaCr4O8 have a pyrochlore spin lattice (see Section 3.1.2). The ideal spinel structure is cubic. Studies on CdCr2O4 [7] and LiGaCr4O8 [8,9] reveal that their observed magnetization plateaus, namely, the 1/2-magnetization plateaus, are strongly stabilized by the magnetoelastic (i.e., spin-lattice) coupling. Another high-symmetry magnet showing the importance of the spin-lattice interaction is the quasi-2D tetragonal magnet SrCu2(BO3)2. Early magnetization measurements on single-crystal samples of SrCu2(BO3)2 identified 1/8-, 1/4- and 1/3-plateaus of Msat (see Section 3.1.1) [10]. More recent experiments with fields up to ~140 T [11,12] revealed that the transition into the regions of the 1/2-magnetization plateau is accompanied by strong magnetoelastic effects [13]. In stabilizing the magnetization plateaus of other magnets with less rigid and low-symmetry atomic structures, the magnetoelastic coupling would also be important, although elaborate studies carried out for the spinel magnets and SrCu2(BO3)2 are mostly not available yet. Thus, in what follows, we will discuss the field-induced fragmentation of a spin lattice based solely on the interaction of the external magnetic field with the spin arrangement in the spin lattice.

2.6. Different Magnetization Behaviors of Heisenberg and Ising Magnets

Heisenberg and Ising magnets change the direction of their magnetic moments as the external magnetic field increases and exhibit a slightly different dependence on the magnetic field. Though often described by an idealized Heisenberg spin Hamiltonian, such a magnet exhibits weak magnetic anisotropy if the orbital moment quenching of its magnetic ions is incomplete. In a similar manner, an ideal Ising magnet described by an Ising Hamiltonian is difficult to find because the associated Jahn–Teller distortion can lift, though weakly, the degeneracy of the d-state responsible for the Ising magnetism [3]. Therefore, by Heisenberg and Ising magnets, we mean those whose magnetic properties are well approximated by Heisenberg and Ising spin Hamiltonians, respectively.
The plateau formation at fractional values of the saturation magnetization Msat compete and coexist with spin-flop transitions in Heisenberg antiferromagnets and metamagnetic transitions in Ising antiferromagnets. In the absence of a magnetization plateau, the magnetization processes of Heisenberg antiferromagnets can be described as illustrated in Figure 9, and those of Ising antiferromagnets as illustrated in Figure 10. Most Heisenberg magnets possess weak magnetic anisotropy, so describing them with a Heisenberg spin Hamiltonian is an approximation. Similarly, the MOn polyhedra of the magnetic ions M in most oxide Ising magnets are weakly distorted due to the Jahn–Teller instability. Therefore it is an approximation to describe Ising magnets using an Ising spin Hamiltonian. The majority of magnetic measurements are conducted mainly on polycrystalline samples. For measurements on such samples, it is often difficult to distinguish whether a magnetic transition involved is a spin-flop or a metamagnetic type. However, it is generally observed that the M(H) curve is concave for spin-flop transitions, but convex for metamagnetic transitions.

2.6.1. Spin Flop and Spin Flip Processes of Antiferromagnets

From the viewpoint of phenomenological description, Heisenberg and Ising magnets differ in the strength of the single-ion magnetic anisotropy D with respect to that of the spin exchange J. The positions of the spin-flop and metamagnetic transitions in the magnetization curves are determined by J and D. General aspects of field-induced transitions were described by Néel [14], detailed discussions of spin-flop transitions by Morosov and Sigov [15], and those of metamagnetic transitions by Strujewski and Giordano [16]. Consider an antiferromagnet with single-ion anisotropy aligning the magnetic moments along a preferred crystal direction commonly called an “easy axis”, which is generally defined as the z-direction. For such a magnet, its behaviors in a field directed either along or perpendicular to the easy magnetization axis are important to analyze. First, let us consider a Heisenberg-type antiferromagnet with a small magnetic anisotropy D compared with the exchange J, which has two magnetic sublattices, M1 and M2 (namely, the up-spin and down-spin sublattices). Consider, for example, Fe2O(SeO3)2 [17], in which the Fe atoms Fe1, Fe2 and Fe3, lead to three spin exchange paths J1, J2 and J3 (for the nearest-neighbor Fe1-Fe2, Fe2-Fe3 and Fe3-Fe3 paths, respectively). These paths form 2D nets parallel to the ab-plane (Figure 11a) which are stacked along the c-direction. The Fe1, Fe2 and Fe3 atoms exist as Fe3+ (d5, S = 5/2) ions, so their magnetic anisotropy is weak (i.e., small D). The neutron diffraction studies [17] reveal that, in all these spin exchange paths, the spin moments parallel to the b-axis are antiferromagnetically coupled. The magnetic susceptibility of this magnet (Figure 11b) shows that, for an external magnetic field along the easy axis, μ0H||b, the transition to the AFM state at the Néel temperature TN manifests itself as a sharp decrease in the magnetic susceptibility χ||. For an external field perpendicular to the easy axis, μ0Hb, however, the magnetic susceptibility χ at T < TN remains almost unchanged. At low temperatures, χ|| < χ. Thus, upon reaching a certain critical field, the difference in the energy of the magnetic moments M1 and M2 oriented either parallel or perpendicular to an external magnetic field reaches a critical value
Δ E = 1 2 ( χ χ | | ) μ 0 H s p i n f l o p 2 ,
at which there is a 90° rotation of the magnetic moments M1 and M2 to the direction perpendicular to the magnetic field. Taking into account the fact that these two moments are related by the exchange interaction J, the field of the spin-flop transition is determined by the expression,
μ 0 H s p i n f l o p = ( 2 D J ) 1 / 2 .
At this field and at temperatures that are low compared with the Néel temperature TN, there is a sharp jump in the magnetization M||, after which the magnetization monotonically increases up to the saturation magnetization
M s a t u r a t i o n = | M 1 | + | M 2 | ,
which is reached in the field of the spin-flip transition,
μ 0 H s p i n f l i p H e i s e n b e r g = J .
In an external magnetic field μ0Hb, the magnetization M increases monotonically, reaching the saturation magnetization in the same field μ0Hspin-flip (we neglect here the anisotropy of the g factor). The field dependence of the magnetization of the easy-axis antiferromagnet Fe2O(SeO3)2 is shown in Figure 11c.
We now turn to an Ising-type antiferromagnet with a magnetic anisotropy D comparable in strength to or exceeding the spin exchange J. When the magnetic field is directed along the magnetic moments of the two sublattices M1 and M2 (i.e., along the b axis), the magnetic susceptibility χ|| of an Ising antiferromagnet is similar to that observed for a Heisenberg antiferromagnet, as shown in Figure 12a for the francisite-type compound [18]. When an external magnetic field μ0H||b reaches the critical value
μ 0 H metamagnetic = J ,
one of the sublattices (either M1 or M2) will reverse its moment direction by 180°, producing a sharp jump as shown in Figure 12b [19]. In this case, μ0Hmetamagnetic is equivalent to μ0Hspin-flip and corresponds to Msaturation. In a magnetic field μ0Hb, both M1 and M2 moments continuously rotate to the direction of the external magnetic field reaching the saturation magnetization Msaturation at
μ 0 H s p i n f l i p I s i n g = J + D .

2.6.2. Magnetization Plateaus

The magnetization behaviors of Heisenberg and Ising magnets differ in their M vs. H curves preceding a magnetization plateau (Figure 1); the magnetization exhibits a linear increase with field for a Heisenberg magnet but does not depend on field for an Ising magnet. We briefly comment on why this difference comes about. In a Heisenberg exchange coupled system, the exchange energy depends only on the relative orientation of the participating spin moments. To a first approximation, the effective spin Hamiltonian for an Ising magnet contains the z-components of the spins only. The only options of the crystal field anisotropy are to favor either a parallel or an antiparallel alignment of the spin moments along the easy axis. Therefore, an external magnetic field not only competes with the spin exchange but also with the anisotropy. Consequently, the magnetic response of an Ising magnet to an external magnetic field depends sensitively on the alignment of the magnetic field with respect to the easy axis as well as their relative magnitudes. Spin-flop transitions with sudden jumps of the magnetization from very low to large values are a phenomenon connected to the presence of a crystal field anisotropy. Thus, the magnetization curve of an Ising magnet deviates somewhat from the step-like features (Figure 1d,e) when the magnetic field is parallel to the easy axis, and from the flat line (Figure 1f) when the magnetic field is perpendicular to the easy axis.

2.7. Quantitative Evaluations of Spin Exchange Interactions

In understanding the magnetic properties of a magnet, it is essential to know the strengths of its magnetic bonds, i.e., the values of its spin exchanges Jij. The parameters Jij specify the spin Hamiltonian (Equation (1)) and determine the energy spectrum for a given magnet. These days, it is almost routine to evaluate the spin exchanges of a magnet composed of transition-metal magnetic ions by using the energy-mapping analysis [6,20,21] based on density functional theory (DFT) calculations. The quantitative values of the spin exchanges are determined by mapping the energy spectrum of a magnet generated by the spin Hamiltonian onto that obtained for a set of broken-symmetry states of the magnet by spin polarized DFT calculations.
The spin exchanges of some magnets discussed in our survey have not been determined before. To gain better insight into these magnets, we determined them by carrying out the energy-mapping analyses. For our DFT calculations, we employed the frozen core projector augmented plane wave (PAW) method [22] encoded in the Vienna ab Initio Simulation Packages (VASP) [23] and the PBE exchange-correlation functional [24]. To take into consideration the electron correlation associated with transition-metal magnetic ions, we performed DFT+U calculations [25] with an effective on-site repulsion Ueff = U − J = 3 and 4 eV on the magnetic ions to ensure that all broken-symmetry states employed for a magnet are magnetic insulating. For a certain magnet, the effect of spin–orbit coupling (SOC) was tested by performing DFT+U+SOC calculations [26]. Unless stated otherwise, the values of the calculated spin exchanges (in K), which are included in a figure describing each magnet, were those determined using DFT+U or DFT+U+SOC calculations with Ueff. Other details of our calculations are reported in the Supporting Information (SI).

3. Magnets of AFM Fragments

3.1. Spin Clusters with Even Number of Spin Sites

3.1.1. Orthogonal Spin Dimers in SrCu2(BO3)2

In the quasi-2D tetragonal compound SrCu2(BO3)2, the CuBO3 layers alternate with Sr layers along the c axis. In each CuBO3 layer, planar Cu2O6 dimers of two edge-sharing CuO4 units are interconnected by BO3 triangles (Figure 13a). The two dominant spin exchanges of this layer are the intradimer exchange J1 of the Cu-O-Cu type and the interdimer exchange J2 of the Cu-O…O-Cu type [6]. In each CuBO3 layer, the (Cu2+)2 dimer ions have an orthogonal arrangement such that the J1 and J2 exchange paths are interconnected by a “head-to-tail” bridging pattern (Figure 13b). The values of J1 and J2, estimated to be 84.7 and 53.4 K (J2/J1 = 0.63), respectively, by LDA+U calculations [27], are very close to the experimental values [28]. This orthogonal arrangement of the dimers represents a realization of the so-called Shastry–Sutherland spin lattice [29]. The magnetic susceptibility of SrCu2(BO3)2 has a sharp peak at about 20 K. Once the contribution of magnetic defects is removed, the susceptibility drops to zero indicating a nonzero spin gap Δ (Figure 14a) [30]. At low temperatures, early magnetization measurements on single-crystal samples of SrCu2(BO3)2 identified 1/8-, 1/4- and 1/3-plateaus of Msat (Figure 14b) [10]. More plateaus (in particular, 2/5 and 1/2) were found in recent experiments with fields up to ~140 T [11,12]. In the plateau regions, the triplet dimers [namely, the (Cu2+)2 dimers with broken J1 bonds] crystallize into magnetic superstructures, so the transitions into the plateau regions are accompanied by substantial magnetoelastic effects. The latter are detected by changes in the magnetostrictive length and volume and a drastic decrease in the sound velocity [13]. To theoretically model the 1/2-magnetization plateau, the interlayer spin exchange needs to be taken into account [13].
The essential aspect of the Shastry–Sutherland spin lattice is illustrated in Figure 13c. In the ground state, each J1 bond is surrounded by an equal number of unbroken and broken J2 bonds. Thus, the contribution of the J2 bonds to the total energy vanishes. The magnetization of SrCu2(BO3)2 is zero between 0 and ~20 T (Figure 14b), because it requires the breaking of J1 bonds to increase magnetization and because all different arrangements of the J2 and broken J2 bonds are identical in energy as long as the J1 bonds remain unbroken. The 1/n-magnetization plateau of SrCu2(BO3)2 requires that one out of every n dimers has a broken J1 bond, because the resulting spin configuration (↑↓)n−1(↑↑) has two net spins out of every 2n spins hence leading to the 1/n-plateau. The magnetic superstructure describing the 1/n-plateau is determined by how a (↑↑) dimer is arranged with respect to every (n − 1) (↑↓) dimer. Figure 14b reveals that the widths of the magnetization plateaus are not uniform, i.e., they decrease in the order, 1/3- > 1/2- > 2/5-, 1/4- > 1/8-plateau. This variation would be caused by their spin-lattice interactions and hence the associated energy lowering, because they will be different due to the differences in their magnetic superstructures. Several theoretical studies examined the magnetic textures and superstructures of the CuBO3 layer predicting many more plateaus [10], which were mostly detected in magnetization, magnetostriction, magnetocaloric effect and nuclear magnetic resonance measurements.

3.1.2. Spin Tetrahedra in Spinel CdCr2O4

CdCr2O4 is a spinel-type compound based on CrO6 octahedra containing Cr3+ (S = 3/2) ions. It is convenient to describe the structure of this compound in terms of the Cr4O16 cluster (Figure 15a), which is made up of four CrO6 octahedra by sharing their edges to form a Cr4O4 distorted cube containing a Cr4 tetrahedron. The spinel structure of CdCr2O4 is obtained by corner-sharing the Cr4O16 clusters, which is accompanied by the corner-sharing Cr4 tetrahedra such that each Cr4 tetrahedron shares a corner with four Cr4 tetrahedra in a tetrahedral arrangement (Figure 15b). Thus, the resulting arrangement of the Cr3+ ions is a pyrochlore spin lattice, which is highly spin-frustrated.
CdCr2O4 undergoes a long-range AFM ordering and a tetragonal lattice distortion at TN = 7.8 K [31]. Using neutron diffraction measurements, the ground state spin configuration was found to be a helical spin structure [32]. The magnetization curve determined at low temperatures reveals a gradual increase with field as the field increases from zero, which is followed by a sharp transition into a 1/2-plateau at Msat/2 = 1.5 μB/Cr3+ (Figure 16a). To account for the nature of the observed M vs. H curve, we treat the pyrochlore arrangement of Cr3+ ions as composed of isolated (Cr3+)4 tetrahedra by neglecting the interactions between them, as shown in Figure 15c. In the zero-magnetization state, each tetrahedron has the (2↑2↓) configuration (Figure 16b), which has four J1 and two broken J1 bonds. When a tetrahedron has the (3↑1↓) configuration (Figure 16b), which has three J1 and three broken J1 bonds, the magnetization increases. The 1/2-magnetization, M = Msat/2, is reached when all isolated tetrahedra have the (3↑1↓) configuration. The energy change required for each (Cr3+)4 tetrahedron to undergo the (2↑2↓) to (3↑1↓) transition is to break one J1 bond per tetrahedron. The sharp jump in the magnetization of CdCr2O4 takes place at about 29 T. At 1.8 K, M(H) increased linearly with H until ~28 T, where M ≈ 0.75 μB, namely, when half the (Cr3+)4 tetrahedra have the (3↑1↓) configuration. On increasing the field beyond this point, the J1 bond breaking occurs simultaneously everywhere such that all (Cr3+)4 tetrahedra have the (3↑1↓) configuration. As the number of broken J1 bonds increases, the J1 breaking at one tetrahedron becomes correlated with those at other places due to the inter-cluster tetrahedra, which were neglected in our discussion. Above 28 T, CdCr2O4 shows a flat magnetization. For the magnetization to increase beyond Msat/2, a tetrahedron must undergo the configuration change from (3↑1↓) to (4↑0↓). The (4↑0↓) configuration has no J1 bond (Figure 16b), while (3↑1↓) has three J1 bonds, so the (3↑1↓) to (4↑0↓) transition requires to break three J1 bonds per tetrahedron. That is, this transition requires more energy than does the (2↑2↓) to (3↑1↓) transition (i.e., one J1 bond per tetrahedron). This explains why the magnetization of CdCr2O4 is flat above 28 T.
A spinel magnet with the ideal pyrochlore spin lattice would not undergo a 3D magnetic long-range order due to the severe spin frustration. However, most spinel magnets undergo a 3D magnetic long-range order because the degeneracy of their ground state is lifted by a structural distortion. In a sense, this distortion can be considered a 3D spin-Peierls transition. A theoretical analysis of CdCr2O4 [7] showed that its 1/2-magnetization plateau is stabilized by magnetoelastic coupling and this plateau is robust. The pyrochlore spin lattice of the spinel magnet LiGaCr4O8 differs from that found for CdCr2O4 in that it consists of small and large tetrahedral (Cr3+)4 clusters which alternate by corner-sharing [8]. The magnetization and magnetostriction studies of LiGaCr4O8 under magnetic fields of up to 600 T [9] show that it exhibits a two-step coupled magnetic and structural phase transition between 150 T and 200 T, followed by a robust 1/2-magnetization plateau up to ~420 T, and that the intermediate-field phase is stabilized by the strong spin-lattice coupling. This phase can be considered a tetrahedron-based superstructure with a 3D periodic array of (3↑1↓) and canted (2↑2↓) configurations.

3.1.3. Spin Hexamers in Pyroxene CoGeO3 and Anisotropic Magnetization Plateau

CoGeO3 consists of two nonequivalent Co atoms, Co1 and Co2, forming Co1O6 and Co2O6 octahedra. By sharing their edges, these octahedra form zigzag ribbon chains parallel to the bc-plane, as shown in Figure 17a. The 3D structure of CoGeO3 is obtained from these zigzag ribbon chains when their oxygen atoms are shared with GeO4 tetrahedra [33]. The magnetic properties of CoGeO3 present a novel feature [34]. The 1/3-magnetization plateau of CoGeO3 is uniaxially anisotropic, that is, CoGeO3 exhibits a pronounced 1/3-plateau when measured with field applied along the c-direction but does not show any magnetization plateau when measured with field perpendicular to the c-direction (Figure 18a). This observation provides an experimental support for our supposition that field-induced partitioning of a spin lattice into ferrimagnetic fragments is essential for magnetization plateaus. To probe the cause for the anisotropic character of the 1/3-magnetization plateau in CoGeO3, mentioned above, we evaluate the four spin exchanges J1J4 defined in Figure 17b using DFT+U calculations. The intrachain exchanges J1J3 are of the Co-O-Co type exchange, while the interchain exchange J4 is of the Co-O…O-Co type. Results of these calculations are summarized in Figure 17c (see Section S1 of the Supplementary Materials).
The spin exchanges J1J4 are all AFM, and the interchain exchange J4 is considerably weaker than the intrachain exchanges J1J3. Since J1 is considerably weaker than J2 and J3, the magnetic ground state for the layer of the double chains has the AFM spin arrangement as depicted in Figure 18b, where each double chain is made up of J2 and J3 magnetic bonds as well as broken J1 bonds. The smallest fragment that can generate a ferrimagnetic fragment of 1/3-magnetization is the hexamer, composed of three J2 bonds with (3↑3↓) spin configuration (indicated by shading in Figure 18b). The ferrimagnetic fragment of (4↑2↓) configuration is generated when one of the three J2 bonds is broken, ultimately leading to the ferrimagnetic state (Figure 18c) when every hexamer has the (4↑2↓) spin configuration. The conversion from a (3↑3↓) to a (4↑2↓) is facilitated because the breaking a J2 bond is accompanied by the formation of two J1 bonds.
CoGeO3 exhibits uniaxial (i.e., Ising) magnetism with the spin moments oriented along the c-direction [33]. According to the selection rules governing the preferred spin orientations of magnetic ions [35], either Co12+ or Co22+ or both ions of CoGeO3 prefer to have their spins oriented along the c-direction. Consider an ideal axially compressed CoO6 octahedron containing Co2+ (d7, S = 3/2) ion with the short Co-O bonds oriented along the z-axis, as depicted in Figure 19a. The t2g-state of such an octahedron is split into the degenerate (xz, yz) state lying above the xy state, assuming that the axially compressed octahedron has an ideal shape with four-fold rotational symmetry. With two d-electrons to occupy the down-spin d-states, the split t2g states become occupied as depicted in Figure 19a. Thus, between the highest-occupied and the lowest-unoccupied d-states, the minimum difference in their magnetic quantum numbers, |ΔLz|, is zero so that the preferred spin orientation is parallel to the z-axis [35]. Of the Co1O6 and Co2O6 octahedra of CoGeO3, only the Co2O6 octahedra have a structure close to an axial-compression [namely, Co2-Oax = 1.994 (×2), and Co2-Oeq = 2.118 (×2), 2.278 (×2) Å]. Since the Co2O6 octahedra have no four-fold rotational symmetry, their xz and yz states are not degenerate, but they still lie above the xy state due to the strong axial compression. The latter guarantees |ΔLz| = 0, hence predicting that the spins of the Co2O6 octahedra are oriented along the short Co2-O bonds, i.e., along the c-direction. The Co12+ ions adopt the spin orientation of the Co22+ ions to maximize their spin exchanges (J2 and J4) with the Co22+ ions. This explains why CoGeO3 exhibits uniaxial magnetism with spin moment along the c-direction.
Based on the uniaxial magnetism of CoGeO3, one can understand why it shows a 1/3-magnetization plateau only when the field is along the c-direction. As already discussed in Section 2, the Zeeman energy EZ between the spin moment μ s = g μ B S and magnetic field μ 0 H is given by E Z = g μ 0 μ B H · S (Equation (4). Therefore, when the spin moment and magnetic field are parallel to each other (Figure 19b), E Z > 0 so that the formation of a ferrimagnetic fragment is energetically favored. However, E Z = 0 if the magnetic field is perpendicular to the spin moment. In such a case, the energy needed to break the magnetic bonds and hence form ferrimagnetic fragments is not available. This explains why the 1/3-magnetization plateau of CoGeO3 has a uniaxial character. In addition, this finding is in support of our suggestion that, for a magnet to exhibit magnetic plateaus, its spin lattice should undergo a field-induced partitioning into ferrimagnetic clusters. It should be noted that the magnetization curves of CoGeO3 (Figure 18a) have a “step-like” feature, because the uniaxial magnetism favors a spin-flip mechanism for magnetization.

3.2. Bose–Einstein Condensates

In a certain magnet composed of discrete units possessing two magnetic ions, such “dimers” have an S = 0 ground state, and the interactions between adjacent dimers are weak so that the first excited state of each dimer, which has S > 0, lies close to the S = 0 ground state. In such a case, the magnetic states of the magnet are well approximated by those of its dimer. The |S, Sz〉 = |S, −S〉 substate of the excited state is lowered in energy under magnetic field μ0H. When μ0H exceeds a certain value, μ0Hc, the |S, −S〉 substate becomes lower in energy than the ground state |0, 0〉, so that the magnetic ground state of each dimer becomes an S > 0 state. Magnets showing such a behavior, known as Bose–Einstein condensates, have been reviewed by Zapf et al. [36]. It should be noted that S = 0 dimers can be discrete molecular units such as Cu2Cl62− anions containing two magnetic ions or dimers composed of two monomers such as (MnO43−)2 (see below). In both cases, the interdimer spin exchange is weaker than the intradimer exchanges. In this section, we discuss the magnetization phenomena observed in two Bose–Einstein condensates.

3.2.1. 0- and 1/2-Plateaus of Ba3Mn2O8

The trigonal compound Ba3Mn2O8 [37] is composed of MnO4 tetrahedra containing Mn5+ (S = 1) ions. Every two tetrahedra combine to form a dimer unit (MnO4)2 such that one Mn-O bond of each MnO4 is parallel to the c-axis (hereafter the Mn-O|| bond). The two Mn-O|| bonds of each dimer are pointed in opposite directions (Figure 20a), and these dimers form trigonal layers. Adjacent layers are shifted from each other such that each dimer of one layer is pointed to the center of three dimers of the two adjacent layers (Figure 20a). Consequently, every (Mn5+)2 dimer ion of one layer is surrounded by six (Mn5+)2 dimer ions (Figure 20b). The spin exchanges of Ba3Mn2O8 are dominated by the intradimer exchange J0 and the interdimer exchange J1 (Figure 20c). (J0 = 15.2 K and J1 = 1.4 K according to our DFT+U calculations, see Section S2 of the Supplementary Materials).
The allowed spin states of each (Mn5+)2 dimer ion are singlet, triplet and quintuplet since Mn5+ is an S = 1 ion, apart from usually small zero-field splitting of these multiplets. The temperature dependence of the magnetic susceptibility χ measured for Ba3Mn2O8 is shown in Figure 21a [38], which evidences that Ba3Mn2O8 is in a singlet ground state with a spin gap Δ = 11.2 K, in which all (Mn5+)2 dimer ions are in the singlet state. At low temperatures, the magnetization plateaus are observed at M = 0 zero and Msat/2 = 2 μB per formula unit (Figure 21b) [38]. We now examine how these plateaus are related to the breaking of the J1 and J0 bonds. The most stable and least stable arrangements of J1 bonds around a J0 bond are shown in Figure 22a, and those around a broken J0 bond in Figure 22b. There are many other arrangements of the J1 and broken J1 bonds whose stabilities lie in between these two extremes. In general, the arrangement becomes more stable if it has more J1 bonds but becomes less stable if it has more broken J1 bonds. As the field increases from 0 to Hc1, each J1 bond begins to break without breaking the J0 bonds. Thus, M = 0 between 0 and Hc1. As the field increases from Hc1, the J0 bond breaking proceeds, hence increasing M. Two dimers with one J0 and one broken J0 bond have the (3↑1↓) configuration. When half the J0 bonds are broken, the M = Msat/2 point at Hc2 is reached. The 1/2-plateau between Hc2 and Hc3 means that there are more J1 bonds than broken J1 bonds at Hc2, while the opposite is the case at Hc3. That is, magnetic energy is absorbed without increasing magnetization from Msat/2. Since J1 is a weak magnetic bond, the width of the 1/2-magnetization plateau is narrow. When the field is stronger than Hc3, more J0 bonds begin to break, increasing the magnetization.

3.2.2. Gapped and Gapless Ground States of ACuCl3 (A = K, Tl, NH4)

A.
Singlet to triplet excitations under magnetic field
The molecular magnets ACuCl3 (A = K, Tl, NH4) [39,40,41] consist of planar Cu2Cl62− anions, which are made up of two CuCl4 square planes containing Cu2+ (S = 1/2) ions by edge-sharing (Figure 23a). Thus, each Cu2Cl62− anion contains a spin dimer (Cu2+)2. The ground spin state for such a dimer can be either singlet (ΔST > 0, “singlet dimer”, Figure 23b) or triplet (ΔST < 0, “triplet dimer” Figure 23c). When such a spin dimer is exposed to a magnetic field μ0H, the triplet state |S, Sz〉 (S = 1, Sz = −1, 0, 1) is split while the singlet state |S, Sz〉 (S = 0, Sz = 0) remains unaffected. For a singlet spin dimer (ΔST > 0) under magnetic field, the triplet state becomes more stable than the singlet state if the field is greater than a critical value μ0Hc (Figure 23d), so that every spin dimer occupies the Sz = −1 state, and the system undergoes a Bose–Einstein condensation. Likewise, for a triplet spin dimer (ΔST < 0) under the magnetic field, the singlet state becomes more stable than the triplet state if the field is higher than a critical value (Figure 23e). For a singlet dimer below μ0Hc, there are three possible spin-flip transitions from the singlet to the triplet under the magnetic field, namely, |0, 0〉 → |1, Sz〉 (Sz = −1, 0, 1), and the energy difference between the two states can be accessed, e.g., by inelastic neutron spectroscopy techniques. The energy difference immediately provides the magnitude of the spin exchange J in the spin dimer. For such transitions to be observed by inelastic neutron scattering measurements, the singlet state |0, 0〉 should be thermally populated and should be more populated than the triplet state(s) into which the transition occurs. This is the case for a singlet dimer because, for field lower than μ0Hc, the |0, 0〉 state is the lowest-lying in energy than any of the three triplet branches (Figure 23d). For a triplet dimer (ΔST < 0), the |0, 0〉 state can be thermally more populated than one branch of the triplet, i.e., the |1, +1〉 state, only when the field is substantially greater than μ0Hc (Figure 23e). Under this condition, the |0, 0〉 → |1, +1〉 transition can take place in a triplet dimer.
NH4CuCl3 is known to consist of three different spin dimers, termed A, B and C, in the 1:2:1 ratio. Results of inelastic neutron scattering experiments carried out for NH4CuCl3 at 0.13 and 1.80 K are summarized in Figure 23f [42], which shows the three branches of the triplet state for the dimers B and C at both temperatures (1.80 and 0.13 K). This finding proves that B and C are singlet dimers with ΔEST values of ~1.6 and ~3.0 meV, respectively. For dimer A, however, only one branch is found, which becomes visible only above ~8.5 T and only at 1.80 K, but apparently is not observed for the measurement at 0.13 K. A singlet–triplet splitting of 0.5 meV (~5.8 K) with the spin triplet lower than the singlet (ΔST < 0) derived by extrapolating the observed |1, +1〉 branch to zero field is consistent with this finding. The small energy difference between the triplet and singlet implies that any excitations with energies below ~0.5 meV could have been masked under the elastic peak or accidentally coincide with excitations for dimers B and C (see the inset in Figure 4 of [30]). These results suggest that the (Cu2+)2 dimer A is weakly coupled by ferromagnetic spin exchange, i.e., it is a triplet dimer. Further temperature dependent inelastic neutron scattering investigations are necessary to verify this conclusion. We found that the magnetization curve observed for NH4CuCl3 at 0.5 K using the H||a field is reasonably well reproduced by assuming that dimers A, B and C are all singlet dimers with the intradimer spin exchanges of 2.7 (1), 11.3 (1) and 19.3 (2) K, respectively (see Figure S1 in Section S3 of the Supplementary Materials). This suggests that dimer A is a singlet dimer but is inconsistent with the inelastic neutron scattering study described above. With dimer A as a triplet dimer, it is straightforward to understand the gapless excitation in the magnetization measurements of NH4CuCl3 (see below) because its triplet dimers A have a nonzero spin moment even in the absence of field. Though isostructural with NH4CuCl3 as far as the atom positions of the heavier atoms are concerned, KCuCl3 and TlCuCl3 consist of only one kind of singlet spin dimer. The excitation energy gaps measured for KCuCl3 and TlCuCl3 are 2.6 and 0.7 meV, respectively [43,44].
B.
Different magnetization behaviors of ACuCl3 (A = K, Tl, NH4)
Magnetization processes of KCuCl3, TlCuCl3 and NH4CuCl3 have been investigated up to 39 T at low temperatures (Figure 24). Both KCuCl3 and TlCuCl3 exhibit a 0-magnetization plateau, and this plateau has a much wider width for KCuCl3 (Figure 24a,b). The transition from a singlet ground state to a magnetic excited state occurs when the field is greater than ~6 and ~20 T for TlCuCl3 and KCuCl3, respectively. These critical fields μ0Hc are consistent with the excitation energy gaps of 0.7 and 2.6 meV observed for TlCuCl3 and KCuCl3, respectively [43,44]. Except for the 0-magnetization plateau, the magnetization of TlCuCl3 and KCuCl3 increases continuously with field showing no more plateau. The magnetization of NH4CuCl3 shows a very different behavior. As the field increases from zero, the magnetization reveals gapless excitations toward a 1/4-plateau, which was followed by a 3/4-plateau before reaching full saturation Msat (Figure 24c) [45].
The magnetization behaviors of KCuCl3 and TlCuCl3 can be readily understood by considering how the magnetic bonds of their spin dimers are broken under the magnetic field. When the field is zero, both KCuCl3 and TlCuCl3 have zero moment because each spin dimer has the singlet configuration, / 2 . The magnetization of KCuCl3 and TlCuCl3 can increase from zero only if dimers start to break their magnetic bonds, one at a time, to assume the triplet configuration (↑↑). The critical field μ0Hc needed to break each dimer magnetic bond is much higher for KCuCl3 than for TlCuCl3 (~20 vs. ~6 T) because the singlet-triplet energy difference ΔST is much larger for KCuCl3 than for TlCuCl3 (2.0 vs. 0.7 meV). This explains why the 0-magnetization plateau is much wider for KCuCl3. With increasing the field, the singlet to the triplet magnetic bond breaking will continue until all magnetic bonds are broken, namely, until the saturation magnetization is reached.
The magnetization behaviors of NH4CuCl3, though apparently more complex, can be similarly explained by noting that spin dimers A, B and C constitute 25%, 50% and 25% of all the dimers, and that dimers A are triplet dimers while dimers B and C are singlet dimers, and the singlet–triplet energy gap is greater for C than for B (3.0 vs.1.6 meV) [42]. When μ0H = 0, triplet dimers A should exist half in the (↑↑) configuration and half in the (↓↓) configuration. As μ0H increases from 0, dimers A with (↓↓) configuration will switch their configuration to (↑↑), successively, until all dimers A attain the (↑↑) configuration at μ0Hc1, where M = Msat/4 because all A dimers have the (↑↑) configuration while the dimers B and C are in the (↑↓) configuration and because 25% of the dimers are dimers of type A. The 1/4-plateau continues until μ0Hc2. When the field is greater than μ0Hc2, the magnetic bonds of dimers B start to break, one at a time, until all dimers B break their bonds at μ0Hc3, where M = 3Msat/4 because all dimers A and B have the (↑↑) configuration while the dimers C have the (↑↓) configuration and because dimers A and B with (↑↑) configuration represent 75% of the total dimers. The 3/4-plateau continues until μ0Hc4. When the field is greater than μ0Hc4, the magnetic bonds of dimers C start to break, successively, until all dimers C break their bonds at μ0Hc5, and the saturation magnetization is finally reached.
C.
Crystal structures of ACuCl3 (A = K, Tl, NH4)
As discussed above, the spin dimers of KCuCl3 are very different from those of TlCuCl3 in the singlet-to-triplet excitation energies. However, the Cu2Cl62− ions of KCuCl3 are very similar in crystal structure to those of TlCuCl3 [35,36]. Neutron scattering measurements reveal the existence of three different spin dimers in NH4CuCl3 [42], but the neutron diffraction studies to determine the crystal structure [46] carried out for ND4CuCl3 at various temperatures show that there is only one kind of Cu2Cl62− anion in ND4CuCl3. These apparently puzzling observations imply that the spin dimers used in interpreting the experimental magnetic data are the effective spin dimers which are affected by the interactions between dimers and by those with the cations A+ (A = K, Tl, NH4). Therefore, it is necessary to examine the crystal structures of ACuCl3 (A = K, Tl, NH4) in more detail with focus on why the magnetic behavior of NH4CuCl3 differs from those of KCuCl3 and TlCuCl3.
In ACuCl3, the Cu2Cl62− anions form stacks along the a-direction (Figure 25a). The spin exchanges describing the interactions within each stack are the intradimer exchange J1 and the two interdimer exchanges Ja and Ja. An important interdimer exchange between adjacent stacks of Cu2Cl62− anions is J2 (Figure 25b). The spin exchanges J1 and J2 are contained in a layer of Cu2Cl62− anions and A+ cations, which is parallel to the ad-plane, where the repeat vector d is defined as d = a + c/2 (Figure 25b). In this layer each Cu2Cl62− anion is surrounded by six A+ cations, and the adjacent J1-J2-J1-J2 alternating chains interact by the interdimer exchanges J3 and J4.
The crystal structure of NH4CuCl3 is slightly more complex than those of KCuCl3 and TlCuCl3 due to the orientation of each NH4+ cation. The crystal structures of ND4CuCl3 including the D atom positions were determined via neutron diffraction at various temperature [46]. In these structure determinations, the presence of three different dimers A, B and C were not taken into consideration (see below for further discussion).
D.
Interdimer exchanges of KCuCl3 and TlCuCl3
The values (in K) of the spin exchanges defined in Figure 25a–c, evaluated by using the energy-mapping analysis based on DFT+U calculations (see Sections S4 and S5 of the Supplementary Materials), are presented in Figure 25d. The intradimer exchange J1 is stronger than the interdimer exchanges in both KCuCl3 and TlCuCl3, and the interdimer spin exchanges are substantially stronger for TlCuCl3 than for KCuCl3. These findings are consistent with the results of the neutron scattering study of Matsumoto et al. [43]. (The intradimer exchange J1 is slightly smaller for KCuCl3 in their study, while the opposite is the case in our calculations). Thus, the excitation energy is substantially smaller for TlCuCl3 than for KCuCl3 essentially because the interdimer spin exchanges are substantially stronger for TlCuCl3 as found previously [43].
To find why the interdimer exchanges were stronger for TlCuCl3 than for KCuCl3, we consider the spin exchanges J2 and Ja as representative examples. The x2-y2 magnetic orbital of each Cu2+ ion lies in the CuCl4 square plane. Thus, the two Cu2+ ions of a spin exchange path are represented by two CuCl4 square planes, a (CuCl4)2 dimer, for short. The (CuCl4)2 dimers of the exchange paths J2 and Ja′ make short contacts with A+ cations, as depicted in Figure 26a and Figure 26b, respectively. In each (CuCl4)2 dimer, the x2-y2 magnetic orbitals of two Cu2+ ions form in-phase and out-of-phase combinations (see Figure S2, Section S3 of the Supplementary Materials), which we represent by the labels (+) and (−), respectively. The frontier orbitals of K+ and Tl+ that can interact with the (+) and (−) d-states are K 4s, Tl 6s and Tl 6p orbitals (Figure 26c). By symmetry, the K 4s orbital interacts with the (+) state, so the (+) level is lowered in energy (Figure 26c). The Tl 6s orbital interacts with the (+) state, which raises the (+) level, but the Tl 6p orbital also interacts with the (−) state, which lowers the (−) level (Figure 26c). Such interactions occur at every Cl…A+…Cl bridge each (CuCl4)2 dimer makes with the surrounding A+ cations. Consequently, the energy gap between the (+) and (−) d-states for the interdimer spin exchanges is larger for TlCuCl3 than for KCuCl3. The intra-stack exchange Ja′ is calculated to be slightly stronger than the strongest inter-stack exchange J2 in both KCuCl3 and TlCuCl3 (Figure 25d). This reflects that the Ja′ path has four A+ cations making the Cl…A+…Cl bridges (Figure 26b), while the J2 path has only two such bridges (Figure 26a).
E.
Intradimer exchange of NH4CuCl3
Let us now examine how the spin exchanges of NH4CuCl3 depend on the orientations of the NH4+ cations with respect to the Cu2Cl62− anions they surround (Figure 25c). Each N-H bond of a NH4+ cation has a σ*N-H orbital, which is highly anisotropic in shape because it is oriented along the N-H bond. Based on the crystal structure determined via X-ray diffraction [41], and assuming that the rotational mobility of the NH4+ cations ceases at low temperatures, we constructed three model orientations, termed YY, NY and NN, of the two NH4+ cations that bridge either side of the Cl…Cl contact in every J2 exchange path (Figure 27a). Under the constraint that two NH bonds of each NH4+ group are coplanar with the Cl…Cl contact and the other NH2 group bisects the Cl…Cl contact, only three different NH4+ arrangements were possible: both NH4+ cations make N-H…Cl hydrogen bonds with the Cl…Cl contact in the YY arrangement, only one NH4+ cation does in the NY arrangement, and no NH4+ cation does so in the NN arrangement (Figure 27b). The orientations of six NH4+ cations surround each Cu2Cl62− ion in the YY, NY and NN arrangements (see Figure S3, Section S3 of the Supplementary Materials).
The relative energies of these three structures and the values of their intradimer spin exchange J1 are summarized in Figure 27c, from which we note the following: (a) The YY, NY and NN arrangements of NH4CuCl3 have considerably different relative stabilities, with the stability increasing in the order, YY < NY < NN. (b) The intradimer exchange J1 of NH4CuCl3 depends strongly on the NH4+ orientations, with its value increasing in the order, NN < NY < YY. (c) The interdimer exchanges J2 and Ja′ of NH4CuCl3 are much weaker than those of KCuCl3 and TlCuCl3. This reflects the fact that the σ*N-H orbitals of NH4+ are strongly contracted compared with the K 4s and the Tl 6s/6p.
Since there are two equivalent ways of having the NY arrangement, the statistical probabilities for the YY, NY and NN arrangements are 1:2:1. The observations (a) and (b) are consistent with the experimental observation suggesting that NH4CuCl3 consists of three different Cu2Cl62− anions in the 1:2:1 ratio [42]. The Cu2Cl62− anions in the YY, NY and NN structures are surrounded by NH4+ cations with different orientations (Figure S3, Section S3 of the Supplementary Materials) and hence will undergo different local relaxations further changing the values of their spin exchanges. To test this hypothesis, we optimized the YY, NY and NN structures by relaxing only the Cu and Cl positions and then calculated the spin exchanges for the resulting structures (Figure 27c) to find a reduction of J1 by 0.05, 13 and 43% for the YY, NY and NN structures, respectively. These reductions reflect that the mid Cl atom of the Cu2Cl62− anion with (without) the short N-H…Cl contact moves away from (toward) the N atom thereby increasing (decreasing) the ∠Cu-Cl-Cu angle of the Cu2Cl62− anion [namely, 96.28° (×2) for the YY, 95.43° and 96.27° for the NY, and 95.35° (×2) for the NN structure]. The effect of the structure relaxation on other spin exchanges is weak (see Sections S6–S8 of the Supplementary Materials).
F.
Consequence of the interaction between NH4+ and Cu2Cl62− in NH4CuCl3
It is of interest to find why the intradimer exchange J1 of NH4CuCl3 depends so sensitively on the NH4+ orientations. The two d-states of a Cu2Cl62− anion, termed the [+] and [−] states in Figure 28a, are the in-phase and out-of-phase combinations of the two x2-y2 magnetic orbitals. With respect to the long axis of the Cu2Cl62− ion, the two p-orbitals at each bridging Cl atom (hereafter, the mid-Cl atom) in Figure 28a are hybridized to become a perpendicular p-orbital (p) in the [+] state, but a parallel p-orbital (p||) in the [−] state. The p orbital is spatially more extended out toward the surrounding cations NH4+ than is the p|| orbital and is hence more effective in the cation–anion interactions in the short NH4+…Cl contacts. The observation (b) reflects that the energy lowering by the (p-σ*N-H) interaction occurs in one and two places in the NY and YY structures, respectively (Figure 28b). Thus, the energy gap between the [+] and [−] states of a Cu2Cl62− anion interacting with the surrounding NH4+ cations increases in the order NN < NY < YY, as depicted in Figure 28c.
We now examine an important implication of the observation made in Figure 28c. In general, the spin exchange J of a spin dimer made up of two S = 1/2 ions is written as J = JF + JAF [6,47]. If the spin sites at i and j are represented by magnetic orbitals ϕi and ϕj, respectively, the FM component JF (<0) increases in magnitude with the overlap density ρij = ϕiϕj, and the AFM component JAF (>0) with the magnitude of the overlap integral Sij = 〈ϕij〉. The interaction between ϕi and ϕj leads to the energy split (Δe)ij between them, which is related to Sij as (Δe)ij ∝ (Sij)2. Therefore, the overall spin exchange J can be FM if (Δe)ij is small. Figure 28c shows that the energy gap between the [+] and [−] d-states of NH4CuCl3 decreases in the order, YY > NY > NN. If the energy split (Δe)ij becomes smaller, then the associated spin exchange can become FM, hence the associated dimer becomes a triplet dimer. It is most likely that the three spin dimers A, B and C of NH4CuCl3, as experimentally observed, might be assigned to the Cu2Cl62− anions surrounded with the NN, NY and YY orientations of the NH4+ cations, respectively. This is a consequence that a given NH4CuCl3 sample does not have a uniform orientation of the NH4+ cations. It rather consists of regions possessing mainly YY, NY and NN orientations of the NH4+ cations.

4. Magnets of Ferrimagnetic Fragments

4.1. Linear Trimers and Chains

4.1.1. Isolated Linear Trimers in Mn3(PO4)2

Manganese diphosphates Mn3(PO4)2 are found in several different phases, namely α, β’, and γ [48], which undergo a long-range AFM order at TN = 21.9, 12.3, and 13.3 K, respectively. The 3D crystal structures of these phases consist of corner- and edge-sharing MnO5 and MnO6 polyhedra, which are further bridged by PO4 tetrahedra. In γ-Mn3(PO4)2, each Mn2O6 octahedron corner-shares with two Mn1O5 trigonal bipyramids to form a Mn1-Mn2-Mn1 linear trimer (Figure 29a), and these trimers are edge-shared either in a head-to-tail (Figure 29b) or tail-to-tail (Figure 29c) fashion. The magnetization curves of these phases show spin-flop-like features at a low magnetic field, but a 1/3-magnetization plateau is found only for the α- and γ-Mn3(PO4)2 modifications. As shown in Figure 29d, the 1/3-plateau of the γ-phase is very wide.
A simplified view of the layer that linear Mn1-Mn2-Mn1 trimers form by a head-to-tail bridging is presented in Figure 30a. The spin lattice of this layer is defined by the intra-trimer exchange J3 and the inter-trimer exchange J2. Such layers make a 3D structure by a tail-to-tail bridging between the trimers lying in adjacent layers, which leads to the interlayer exchange J1 (Figure 30b). The spin exchanges J1, J2 and J3 of γ-Mn3(PO4)2 (Figure 29a) are all AFM and are estimated to be 1.7, 4.7 and 10.5 K, respectively [48]. Each layer defined by the exchanges J3 and J2 are ferrimagnetic because each linear trimer is ferrimagnetic due to the strong AFM exchange J3, and because the head-to-tail coupling between two ferrimagnetic trimers does not cancel their moments (Figure 30c). Such ferrimagnetic layers are coupled antiferromagnetically via the exchange J1 to form an AFM magnetic ground state responsible for the AFM ordering.
The gradual increase in the magnetization M of γ-Mn3(PO4)2 with increasing μ0H in the region of 0–7.5 T mirrors the breaking of the interlayer J1 bonds, leading to the ferrimagnetic layers. This field-induced ferrimagnetic state at 7.5 T explains the 1/3-plateau. For each ferrimagnetic layer to go beyond the 1/3-plateau, it is necessary to break two J3 bonds of a trimer, which is accompanied by the breaking of a J2 bond (Figure 30d). The wide plateau between 7.5 and 23.5 T reflects the difficulty of simultaneously breaking one J2 and two J3 bonds when a linear AFM trimer becomes FM.

4.1.2. Bent Trimers in Cu3(P2O6OH)2

The building blocks of Cu3(P2O6OH)2 are Cu2O6 octahedra and Cu1O5 trigonal bipyramids, as found in γ-Mn3(PO4)2. However, each Cu2O6 octahedron edge-shares with two Cu1O5 trigonal bipyramids in Cu3(P2O6OH)2 (Figure 31a) [49] to form linear Cu1-Cu2-Cu1 trimers, in contrast to the corner-sharing found in γ-Mn3(PO4)2 (Figure 29a). Cu3(P2O6OH)2 exhibits a 1/3-magnetization plateau above 12 T (Figure 32a) [50], which was initially interpreted by supposing that its spin lattice is a J1-J2-J2 chain made up of ferrimagnetic linear Cu1-Cu2-Cu1 trimers (Figure 31b). However, this model is not consistent with the spin exchanges of Cu3(P2O6OH)2 evaluated using DFT+U calculations [51]; the latter found that the exchange J2 is practically zero, and that Cu3(P2O6OH)2 has a 2D spin lattice made up of three spin exchanges J1, J3 and J6 (479, 69 and 90 K, respectively) as shown in Figure 31c.
The three AFM exchanges J1, J3 and J6 lead to an AFM spin arrangement in the 2D lattice (Figure 32b), where half the Cu2 sites have up-spins, and the remaining half down-spins (i.e., those in green circles). Since J3 and J6 are considerably weaker than J1, the increase in M with μ0H is achieved by breaking these magnetic bonds, i.e., by flipping the down-spin to up-spin at the Cu2 sites, one at a time. This spin flipping simultaneously breaks one J6 and two J3 bonds. When all down-spins at the Cu2 sites are flipped, a ferrimagnetic configuration with M = Msat/3 is reached (Figure 32c). Note that this spin arrangement is equivalent in energy to another ferrimagnetic spin arrangement shown in Figure 32d. Either ferrimagnetic arrangement can be decomposed into bent ferrimagnetic trimers, as illustrated in Figure 32d. The plateau above 12 T is wide because the J1 bond is strong and because a spin flip from (↑↓↑) to (↑↑↑) in each ferrimagnetic trimer, which must occur to increase the magnetization beyond M = Msat/3, simultaneously breaks one J1 and one J3 bond.

4.1.3. Heisenberg Chains in Volborthite Cu3V2O7(OH)2·2(H2O)

Volborthite, Cu3V2O7(OH)2·2H2O, has a layered crystal structure, in which the layers of composition Cu3O6(OH)2 parallel to the ab-plane are pillared by pyrovanadate V2O7 groups and crystal water molecules occupy the voids between the layers. The Cu2+ ions in each Cu3O6(OH)2 layer have a kagomé-like arrangement (Figure 33a). Below room temperature, volborthite undergoes two structural phase transitions, one at ~292 K from a C2/c phase to a I2/a phase, and the other at ~155 K from the I2/a phase to a P21/c phase [52]. The latter structural phase transition generates two kagomé layers slightly different in structure. Below 1.5 K, volborthite exhibits magnetic order, indicated by two anomalies in the magnetic specific heat [53].
Volborthite had been regarded as a kagomé spin lattice system [54]. However, according to a recent study [53], it is not a kagomé spin lattice but an S = 1/2 AFM uniform Heisenberg (AUH) chain that describes the magnetic properties of volborthite at low temperatures. This observation reflects the fact that the spin lattice of a magnet does not necessarily follow the geometrical pattern of its magnetic ion arrangement but is determined by that of strongly interacting spin exchange paths between the magnetic ions [55]. The CuO6 octahedra of volborthite accommodating the Cu2+ ions are axially elongated, so their x2-y2 magnetic orbitals lie in their CuO4 square planes perpendicular to the elongated Cu-O bonds. The arrangement of these CuO4 planes in volborthite, depicted in Figure 33b, is highly anisotropic forming the Cu2-Cu1-Cu2 linear trimers bridged by Cu2-O-Cu1 linkages. Within each Cu3O6(OH)2 layer, the Cu2-Cu1-Cu2 trimers are arranged as in Figure 33c. The spin exchanges of volborthite determined using DFT+U calculations show [53] that the strongest AFM spin exchange, J2 (550 and 582 K for the two different layers), makes each Cu2-Cu1-Cu2 linear trimer ferrimagnetic, and these ferrimagnetic trimers are coupled antiferromagnetically by the next strongest spin exchange J4 (78 K for both layers) to form two-leg spin ladders. All other spin exchanges are negligibly small, and adjacent spin ladders are entangled in their legs (Figure 33c). In essence, the kagomé-like arrangement of Cu2+ ions in Cu3V2O7(OH)2·2H2O gives rise to weakly interacting two-leg spin ladders with Cu2-Cu1-Cu2 trimers as rungs, which have an AFM spin arrangement as depicted in Figure 33d.
At low temperatures where thermal excitations within each trimer rung are absent, each rung acts as an effective S = 1/2 species due to a strong AFM coupling between adjacent Cu2+ sites, so that each two-leg spin ladder should behave as an effective S = 1/2 AUH chain (Figure 33e) [53]. Indeed, the magnetic susceptibility of volborthite at low temperatures (below 75 K) is very well described by an S = 1/2 AUH chain model to find the nearest-neighbor spin exchange JC = 27.8(5) K, as shown in Figure 34a [53]. On lowering the temperature, the susceptibility shows a broad maximum and converges to a nonzero value as the temperature approaches zero, a characteristic feature expected for an S = 1/2 AUH chain. Volborthite exhibits an extremely wide 1/3-magnetization plateau above 28 T continuing over 74 T at 1.4 K (Figure 34b) [56]. Before reaching the value of M = Msat/3, the magnetization increases with field because each linear (↓↑↓) trimer is converted to a linear (↑↓↑) trimer, breaking four J4 bonds, eventually reaching the ferrimagnetic state (Figure 33f) at ~28 T. A further increase in the magnetic field does not increase magnetization leading to the 1/3-plateau because it requires breaking two J2 bonds to convert a (↑↓↑) rung to a (↑↑↑) rung and because the J2 bond is very strong. There is a theoretical study on the magnetization plateau of a two-leg spin ladder [57]. In our analysis of the magnetization plateau of volborthite obtained at 1.4 K [55], we employed the S = 1/2 AUH chain model (Figure 34c) because each (↑↓↑) rung acted as an effective S = 1/2 entity at 1.4 K. As shown in Figure 34c, the experimental magnetization data were very well described by the S = 1/2 AUH chain model (Figure 34c) using the nearest-neighbor spin exchange JC of 27.5 K [53], just as were the magnetic susceptibility data below 75 K.

4.1.4. Head-to-Tail Coupling of Bent Trimers and Anisotropic 1/3-Plateau in Cs2Cu3(SeO3)4·2(H2O)

Cs2Cu3(SeO3)4·2H2O consists of two nonequivalent Cu atoms, Cu1 and Cu2 in the 1:2 ratio, each forming Cu1O4 and Cu2O4 square planes, respectively. The 3D framework of Cs2Cu3(SeO3)4·2H2O is formed by the corner-sharing of Cu1O4 and Cu2O4 square planes [58]. As depicted in Figure 35a,b, each Cu1O4 square plane is corner-shared with four Cu2O4 square planes such that the four Cu2 atoms around a Cu1 atom make a Cu1(Cu2)4 tetrahedron (Figure 35c). Condensing such Cu1(Cu2)4 tetrahedra by sharing their Cu2 corners leads to the 3D network of Cu2+ ions of Cs2Cu3(SeO3)4·2H2O, which can be described as resulting from the fusing of chair-shape hexagonal rings (Figure 35d).
Cs2Cu3(SeO3)4(H2O)2 is a ferrimagnet ordering at TC = 20 K with residual magnetization at about Msat/3 (Figure 36a). In general, ferrimagnetism occurs when ferrimagnetic fragments are combined antiferromagnetically in a head-to-tail bridging pattern so that the magnetic moment of each ferrimagnetic fragment is not quenched. Such ferrimagnetic units in Cs2Cu3(SeO3)4(H2O)2 should consist of one Cu1 and two Cu2 atoms, given that the Cu1 and Cu2 atoms occur in the 1:2 ratio. DFT+U calculations [58] show that the nearest-neighbor exchange J1 (Figure 35c) is strong (256 K) but other exchanges are negligibly weak. This makes all nearest-neighbor Cu1…Cu2 linkages antiferromagnetically coupled, so the ferrimagnetic fragments needed to explain the ferrimagnetism of Cs2Cu3(SeO3)4(H2O)2 are the bent Cu2-Cu1-Cu2 units with (↑↓↑) spin configuration.
When measured for a single crystal sample of Cs2Cu3(SeO3)4(H2O)2 parallel (||) and perpendicular (⊥) to the c-direction (Figure 36a,b), the values of the magnetization at μ0H = 7 T are quite different, namely, M = 1.18 μB, whereas M|| = 0.93 μB [58]. There are three factors contributing to this highly anisotropic magnetization plateau: the nearly orthogonal arrangements of the Cu2O4 square planes around each Cu1O4 square plane (Figure 35b), the strong nearest-neighbor antiferromagnetic exchange J1 and the anisotropic g-factor of Cu2+ ions in a square planar coordination site. The magnetic anisotropy of a magnetic ion arises from SOC. In the spin-only description, in which orbital information is suppressed, the effect of SOC on magnetic anisotropy is discussed by introducing g-factor g, different from 2 [2]. That is, the magnetic moment μ of a spin site with spin S is given by μ = gS, where g is the anisotropic g-factor of the magnetic ion. The g-factors of Cu2+ at a square planar coordination site along the c-axis (||c for short) and perpendicular to the c-axis (⊥c for short) are written as
g|| = 2 + Δg||
g = 2 + Δg
where Δg|| > Δg (approximately, 0.25 vs. 0.05) (Figure 37a). Δg|| and Δg are proportional to the amounts of unquenched orbital angular momenta [2], so the associated magnetic moments are also anisotropic, namely,
μ|| = g||S = (2 + Δg||)S
μ = gS = (2 + Δg)S
Thus, the magnetic moment of the Cu2+ ion is greater along the ||z direction than along the ⊥z direction.
To simplify our analysis of the observed magnetization anisotropy, we assume that each Cu2O4 unit is truly a square planar in shape, and the planes of the Cu1O4 units are truly orthogonal to the plane of the idealized Cu2O4 unit (Figure 37b,c). Then, all three CuO4 square planes of a bent Cu2-Cu1-Cu2 ferrimagnetic fragment are identical except for their spatial arrangement. Since J1 is AFM, the three Cu2+ spins of a bent Cu2-Cu1-Cu2 fragment have a (↑↓↑) spin arrangement. For the magnetic field H||c, the three spins are aligned along the crystallographic ||c direction (Figure 35b) so that the magnetic moments on the two up-spin sites are both μ, while that on the down-spin site is -μ|| (Figure 37b). For the magnetic field Hc, however, the three spins are aligned along the ⊥c direction so that the magnetic moments on the two up-spin sites are both μ||, while that on the down-spin site is -μ (Figure 37c). Therefore, the total moments of the ferrimagnetic fragment are given by
μtot (||c) = 2μμ|| = (2gg||)S ≈ 1 + (Δg − Δg||/2) ≈ 0.94
μtot (⊥c) = 2μ||μ = (2g||g)S ≈ 1 + (Δg|| − Δg/2) ≈ 1.195
This difference explains why the 1/3-magnetization plateau has a moment larger than 1 μB for Hc, but a moment smaller than 1 μB for H||c, and why the magnetization plateau deviates more from 1 μB for Hc than for H||c.

4.1.5. Haldane Chain of Cu6 Clusters and a 1/3-Magnetization Plateau in Fedotovite K2Cu3O(SO4)3

Fedotovite, K2Cu3O(SO4)3, consists of Cu6 clusters (Figure 38a), which are made up of three different Cu atoms, Cu1, Cu2 and Cu3, in strongly distorted square planar coordination. Two Cu3O4 planes are edge-shared to form a twisted Cu32O6 dimer, and one bridging O atom of this dimer is corner-shared with two Cu1O4 square planes while the other bridging O atom is corner-shared with two Cu2O4 square planes. Thus, the atoms of a Cu6 cluster have the shape of an edge-sharing tetrahedra (Figure 38b). Such Cu6 clusters form chains along the b-direction (Figure 38c) [59].
K2Cu3O(SO4)3 undergoes a 3D AFM ordering at TN = 3.1 K. Above this temperature, K2Cu3O(SO4)3 behaves as an S = 1 Haldane chain system (Figure 39a), with each Cu6 cluster acting as an S = 1 species [60]. This implies that the spin exchange coupling between six Cu2+ ions of the Cu6 cluster is very strong so that thermal excitations within each Cu6 cluster are weak. In addition, K2Cu3O(SO4)3 exhibits a 1/3-plateau above TN (Figure 39b) [60], implying that each Cu6 cluster forms a ferrimagnetic fragment with (4↑2↓) spin configuration. To confirm this interpretation, we examine the eight spin exchanges J1J8 defined in Figure 38. The values of these exchanges determined using DFT+U calculations are summarized in Figure 38 (see Section S9 of the Supplementary Materials). The exchange J1 between the Cu12+ ions is strongly AFM, and so is the exchange J2 between the Cu22+ ions. In contrast, the exchange J3 between the Cu32+ ions is strongly FM. There are four strong AFM exchanges between the Cu12+ and Cu32+ ions (namely, 2J4 + 2J5) and between the Cu22+ and Cu32+ ions (namely, 2J6 + 2J7). Since these AFM interactions dominate over J1 and J2, the energetically favorable spin arrangement for a Cu6 cluster is either a (2↑2↓2↑) or a (2↓2↑2↓) configuration (Figure 40a), which are both ferrimagnetic. Due to the AFM inter-cluster exchange J8, the ferrimagnetic Cu6 clusters prefer to couple antiferromagnetically (Figure 40b). The gradual increase in the magnetization with the magnetic field from 0 to about 20 T is explained by the field-induced breaking of the inter-cluster magnetic bonds J8, one at a time, eventually reaching the ferrimagnetic state (Figure 40c), in which all J8 bonds are broken with M = Msat/3.
It should be noted that the magnetic susceptibility of K2Cu3O(SO4)3 is rather weak (Figure 39a). This is a direct consequence of the fact that the spin exchanges J1J7 leading to the ferrimagnetic fragment Cu6 are rather strong. The latter is necessary for the effective S = 1 behavior of the Cu6 clusters. The magnetic susceptibility of this Haldane chain system is weak despite the presence of six Cu2+ cations in each cluster due to the (↑↓↑) arrangement of three FM dimers.

4.1.6. Trigonal Arrangement of Ferromagnetic Chains in Ca3Co2O6

Ca3Co2O6 consists of Co2O6 chains in which Co2O6 trigonal prisms alternate with Co1O6 octahedra by sharing their triangular faces (Figure 41a) [61]. These chains running along the c-direction have a trigonal arrangement (Figure 41b), with Ca2+ cations occupying the positions in between these chains. Each Co2O6 chain is FM [62] so that the spin lattice of Ca3Co2O6 can be described as a trigonal lattice by treating each FM chain a pseudo-magnetic ion with a giant spin moment. Both Co1 and Co2 atoms of Ca3Co2O6 are in the oxidation state of +3 [3,63], indicating that each Co2O6 trigonal prism has six electrons to occupy its d-states, and so does each Co1O6 octahedron. This made it difficult to explain why Ca3Co2O6 exhibited a uniaxial magnetism [1,3,63] because the configuration (3z2 − r2)1(xy, x2 − y2)0 predicted for a Co2O6 trigonal prism does not lead to uniaxial magnetism (Figure 42a, left). A systematic study [3] of Ca3Co2O6 based on DFT+U and DFT+U+SOC calculations, including geometry relaxations allowing for Jahn–Teller distortions, showed that the uniaxial magnetism of Ca3Co2O6 is a consequence of three effects: (a) the FM spin arrangement between the Co3+ ions of adjacent Co2O6 and Co1O6 polyhedra, (b) a direct metal–metal interaction between adjacent Co3+ ions mediated by their 3z2 − r2 orbitals (Figure 41c), and (c) the SOC of the Co3+ ion at the trigonal prism site (Figure 42a, right).
A single crystal sample of Ca3Co2O6 exhibits a 1/3-magnetization plateau when the field is parallel to the chain direction, with the magnetization curve showing a step-like feature. When the field is perpendicular to the chain direction, there occurs no magnetization plateau [62]. (Experimentally, it is very difficult to align a single crystal sample of a uniaxial magnet precisely perpendicular to the field. A very slight misalignment can easily give rise to nonzero magnetization). As discussed for CoGeO3 in the previous section, these observations are a direct consequence of the fact that Ca3Co2O6 is a uniaxial magnet with a spin moment along the chain direction. Ca3Co2O6 can be described in terms of a regular trigonal spin lattice of simple magnetic ions once each FM Co2O6 chain is treated as a single magnetic ion (see Section 5). It is noteworthy that Ca3Co2O6 reaches a full saturation magnetization at a rather low field (namely, at about 3.5 T). This reflects that the interchain magnetic bonds are weak.

4.2. Distorted Triangular Fragments

4.2.1. Diamond Chains of NaFe3(HPO3)2(H2PO3)6

NaFe3(HPO3)2(H2PO3)6 has two nonequivalent Fe atoms, Fe1 and Fe2, forming Fe1O6 and Fe2O6 octahedra [64]. The HPO3 unit occurs in two different forms, i.e., H-PO3 and PO2(OH), but the H2PO3 unit only in the form H-PO2(OH). Consequently, both Fe1 and Fe2 atoms are present as Fe3+ (S = 5/2) ions. These Fe3+ ions are bridged by H-PO3, PO2(OH) or H-PO2(OH), as illustrated by Figure 43a. DFT+U calculations [65] showed that four AFM spin exchanges (i.e., J2, J3, J4 and J6 depicted in Figure 43a) are relevant and comparable in magnitude (~2 K). The three spin exchanges J2, J3 and J6 couple the Fe3+ cations to diamond chains, which are interlinked by J4 to form 2D layers (Figure 43b). Such layers are stacked to form the 3D spin lattice of NaFe3(HPO3)2(H2PO3)6. In addition, there are weak interlayer AFM spin exchanges (see below for further discussion).
As shown in Figure 44a (inset), NaFe3(HPO3)2(H2PO3)6 undergoes a ferrimagnetic ordering below TC = 9.5 K and exhibits a 1/3-magnetization plateau in the magnetization [65]. The plateau extends to ∼8 T. Above this field, the magnetization increases linearly with field until the saturation is reached at ∼27 T. As discussed in Section 2.4, we suppose that the spin lattice of NaFe3(HPO3)2(H2PO3)6 undergoes field-induced partitioning into ferrimagnetic triangular clusters (Figure 44b). Then, the spin arrangement, (↑↓↑), (↑↑↓) or (↓↑↑), of each cluster leads to one positive moment per cluster. Among these three, the (↑↓↑) arrangement at each triangular fragment is energetically most favorable because of the inter-diamond spin exchange J4, thereby leading to the ferrimagnetic state with M = Msat/3 (Figure 44c). To increase the magnetization beyond Msat/3, each ferrimagnetic triangle must break two magnetic bonds (Figure 44d) within a cluster, which is accompanied by the breaking of two inter-diamond J4 bonds. This needs a high enough magnetic field, hence explaining the 1/3-plateau extending to ~8 T. When the field increases beyond 8 T toward the saturation magnetization, each ferrimagnetic triangle begins to break two magnetic bonds (Figure 44d) within a cluster.
In general, the ground state of a magnet composed of ferrimagnetic layers is AFM because the weak high-spin orbital interactions between adjacent ferrimagnetic layers favor an AFM coupling rather than an FM coupling [66]. Indeed, DFT+U calculations found [65] that the interlayer spin exchanges J1 and J5, which are weakly AFM (~0.6 and ~0.4 K, respectively), and form spin-frustrated (J1, J5, J4) triangles between adjacent ferrimagnetic layers as depicted in Figure 45a. The interlayer FM coupling (Figure 45b) leads to the (J1, J5, J4) triangles, which have J5 magnetic bonds and J1 broken magnetic bonds. In contrast, the interlayer AFM coupling (Figure 45c) leads to the (J1, J5, J4) triangles, which have J5 broken magnetic bonds and J1 magnetic bonds. Since J1 is slightly stronger than J5, the interlayer AFM coupling is energetically favored over the interlayer FM coupling. This is consistent with the general observation that the magnetic ground state of a magnet composed of ferrimagnetic layers is AFM rather than ferrimagnetic. Furthermore, we note from Figure 44a that, below ~2 T, the magnetization rises sharply with field toward Msat/3. This observation can be related to the breaking of the weak interlayer magnetic bonds (i.e., J5 and J1) in the AFM ground state. It will be interesting to examine whether the magnetic ground state of NaFe3(HPO3)2(H2PO3)6 is AFM or ferrimagnetic.

4.2.2. Three-Dimensional Spin Lattice and Anisotropic Plateau Width in Azurite Cu3(CO3)2(OH)2

The important structural building blocks of azurite Cu3(CO3)2(OH)2 [67] are the Cu1O4 and Cu2O4 square planes containing their x2-y2 magnetic orbitals. Each Cu12+ ion is surrounded by four Cu22+ ions to form a Cu5 ribbon (Figure 46a) where the four Cu2O4 square planes of a Cu5 ribbon are nearly perpendicular to the central Cu1O4 square plane. This structural feature implies that the orientations of the x2-y2 magnetic orbitals are the keys to understanding the magnetic properties and especially the magnetization plateau observed for azurite. By sharing their edges, such Cu5 ribbons form “diamond chains” along the b-direction (Figure 46b). Each Cu5 ribbon is described by three spin exchanges J1J3, as used early on by Rule et al. to discuss the temperature dependence of the magnetic susceptibility [68]. Kang et al. carried out DFT+U calculations [69] to find that adjacent diamond chains interact through the spin exchanges J4, which takes place through the bridging of CO32− ions (Figure 46c) to form a layer of interacting diamond chains, and that the dimer exchange J2 (=363 K) dominates with J1/J2J3/J2 = 0.24 and J4/J2 = 0.13. Jeschke et al. proposed a modified diamond chain model by including a spin exchange between the Cu1 cations [70]. Topologically, the 2D spin lattice of azurite is identical with that for NaFe3(HPO3)2(H2PO3)6 discussed in the previous section. So, one might expect that each layer of azurite is ferrimagnetic as shown in Figure 44c, and such ferrimagnetic layers are antiferromagnetically coupled to form an ordered 3D AFM state, and that azurite exhibits a 1/3-plateau as NaFe3(HPO3)2(H2PO3)6 does. Indeed, azurite orders antiferromagnetically at TN ≈ 1.9 K [71] and exhibits a 1/3-magnetization plateau below this temperature, as shown in Figure 47.
The 1/3-plateau of azurite presents two features remarkably different from that of NaFe3(HPO3)2(H2PO3)6 (Figure 44a): (1) The field Hc1 where the M = Msat/3 point starts on increasing the field from zero is much greater for azurite (over 10 T) than for NaFe3(HPO3)2(H2PO3)6 (~1 T). In NaFe3(HPO3)2(H2PO3)6, the gradual increase in M with μ0H in the region of 0–Hc1 is ascribed to the breaking of the interlayer magnetic bonds. Since Hc1 is much higher for azurite, the interlayer AFM spin exchange of azurite must be substantial. (2) The width of the 1/3-plateau is much wider when the field is perpendicular to the b-axis (⊥b) than parallel to the b-axis (||b); the Hc1 is greater for H||b than for Hb (16 vs. 11 T), whereas the Hc2 is smaller for H||b than for Hb (26 vs. 30 T). Thus, the plateau width is significantly larger for Hb. However, in contrast to these differences in the plateau widths, the saturation fields Hc3 in both orientations are identical (32.5 T) [71]. In the following, we examine why these observations occur.
  • Interlayer spin exchange in azurite
Two-dimensional layers of interlinked diamond chains are stacked as depicted in Figure 48a. There occur two Cu-O…O-Cu type spin exchange paths, J5 and J6, between adjacent layers (Figure 48b). The J5 paths take place between Cu12+ and Cu22+ ions (Figure 48c), and the J6 paths between two Cu22+ ions (Figure 48d). The values of J5 and J6, determined using energy-mapping analysis (see Section S10 of the Supplementary Materials) are not negligible compared with the inter-diamond exchange J4 within a layer; J5 is AFM while J6 is FM, and J5/J4 = 0.7 and J6/J4 = −0.5. The presence of the AFM interlayer exchange J5, which is only slightly weaker than J4, confirms our suggestion that the increase in magnetization with a field in the 0–Hc1 region is related to the breaking of the interlayer magnetic bonds, and azurite reaches the state consisting of (↑↓↑) ferrimagnetic triangular fragments at Hc1.
B.
Magnetic anisotropy affecting Dzyaloshinskii–Moriya (DM) interactions
The different widths of the 1/3-plateaus were explained by Kikuchi et al. [71] in terms of DM interactions by assuming a DM vector perpendicular to both the J2 bond and the b-axis. So far, however, it is unknown why such a DM vector should exist in azurite, or why the 1/3-plateau starts at a higher field for H||b than for Hb. To resolve these questions, we examine how the Zeeman energies of the Cu2+ ions in azurite are affected by the field direction based on the following three observations:
(1)
The g-factor for the Cu2+ ion of a CuO4 square plane is anisotropic; the g-factor along the four-fold rotational axis, g|| = 2 + Δg|| ≈ 2.25, is substantially greater than that perpendicular to this axis, g = 2 + Δg ≈ 2.05 [72].
(2)
In general, the g-factor of a magnetic ion measured with magnetic field H in a certain direction can be written as g = 2 + Δg, where Δg is related to the unquenched orbital moment δL on the magnetic ion along that direction as [2]
g = λ δ L μ B H δ L ,
where λ is the SOC constant of the magnetic ion, i.e., Δg is a measure of δL.
(3)
In a DM interaction D a b · ( S a × S b ) between two spins located at the sites a and b and coupled by spin exchange Jab, the DM vector D a b is related to the unquenched orbital moments δ L a and δ L b of the magnetic ions at the sites a and b, respectively, as [2,73]
D a b = λ J a b δ L a δ L b ( g a g b )
The essential key to understanding the observation of the different widths of the plateaus in azurite is that each Cu1O4 square plane is nearly perpendicular to the four Cu2O4 square planes within each diamond chain, and also nearly perpendicular to the two Cu2O4 square planes between two adjacent diamond chains (Figure 49a). To simplify our analysis, we assume that the Cu1O4 and Cu2O4 units have an ideal planar square shape, and that the arrangement of these square planes are ideally orthogonal such that the two edges of the Cu1O4 plane are aligned along the y- and z-axes, but those of the Cu2O4 planes along the x- and y-axes (Figure 49b). Then, the four-fold rotational axis of the Cu1O4 plane is parallel to the x-axis (||x), but that of each Cu2O4 plane is parallel to the z-axis (||z). With this choice of the Cartesian coordinate system, the y-direction is approximately aligned along the b-direction, i.e., the diamond chain direction. Then, for the Cu1O4, the g-factor of the Cu2+ cations is g|| for H||x, but g for Hx (Figure 49c). For the Cu2O4 planes, however, the g-factor of the Cu2+ is g|| for H||z, but g for Hz (Figure 49c).
Using the results summarized in Figure 49c, the Zeeman energy for the three Cu2+ ions of each ferrimagnetic triangle (namely, one Cu12+ and two Cu22+ ions, Figure 46d) is calculated as follows:
For H||x: E Z = ( g | | + 2 g ) μ 0 μ B H S
For H||y: E Z = 3 g μ 0 μ B H S
For H||z: E Z = ( 2 g | | + g ) μ 0 μ B H S
For the Cu2+ ion, g < g | | (i.e., ~2.05 vs. ~2.25). Thus, at a given magnetic field strength μ0H, the Zeeman energy is lower for H||y than either for H||x or for H||z. This implies that in reaching the energy required for breaking a certain magnetic bond, a higher magnetic field is necessary when the field is aligned along the y-direction. This explains why the Hc1 is higher for H||b than for Hb (16 vs. 11 T) since the y-direction is approximately aligned along the b-direction of azurite. Interestingly, this identifies the magnetic bonds to break in this process are the interlayer magnetic bonds, i.e., the interlayer exchange coupling is a crucial factor leading to the different widths of the 1/3-plateaus in azurite.
We now examine why Hc1 is lower for H||b than for Hb (i.e., 26 vs. 30 T) by noting that the Hc2 marks the point where each (2↑1↓) ferrimagnetic triangle of Figure 46d begins to change into a (3↑0↓) ferromagnetic triangle. This change breaks six magnetic bonds (namely, 2J1 + 2J3 +2J4) around a Cu12+ ion. As pointed out earlier, J1/J2J3/J2 ≈ 0.24 and J4/J2 ≈ 0.13, so (2J1 + 2J3 +2J4) ≈ 0.74J2. The DM interactions of the six magnetic bonds are identical except for the magnetic bond strengths. Therefore, we treat the six DM interactions involving one Cu12+ ion as one DM interaction of the Cu12+ ion with a hypothetical Cu22+ ion with effective bond Jeff = 0.74J2. Then, by considering the Cu12+ and the hypothetical Cu22+ ions at sites a and b, respectively, we obtain the following results,
For μ0H||x: D a b λ J e f f ( g | | g )     0.2 λ J e f f < 0
For μ0H||y: D a b λ J e f f g g     0
For μ0H||z: D a b λ J e f f ( g g | | )   0.2 λ J e f f > 0 ,
where we used the fact that λ < 0 for Cu2+ with more than a half-filled d-shell. The above results show that the DM interaction vanishes for H||y. The DM vector for H||x is opposite in sign to that for H||z. For H||z, the DM interaction raises the Zeeman energy, so the magnetic bond breaking occurs at a lower field (compared with the H||y case). For H||x, however, the DM interaction lowers Zeeman energy, forcing the magnetic bond breaking to a higher field. What is observed for azurite can be understood if the ⊥b direction is close to the ||x direction. The ||z direction is also approximately the ⊥b direction, but the DM interaction for H||z raises the Zeeman energy while that for H||x lowers it. This leads to the prediction that the plateau widths increase in the order,
H||z < H||y < H||x.
It would be interesting to verify this prediction experimentally.

5. Trigonal vs. Kagomé Magnets

The magnetization plateaus of magnets with triangular [74] and kagomé [75,76,77,78,79,80] spin lattices have received more attention in theoretical studies than in experimental studies. These plateaus are less prominent compared with those found for other magnets of lower symmetry.

5.1. Cause for the Presence or Absence of a Clear-Cut 1/3-Magnetization Plateau

Magnets of triangular and kagomé spin lattices show contrasting behaviors in their magnetization, especially, in the development of magnetization plateaus. The 1/3-magnetization plateaus observed for trigonal spin-lattice magnets are generally narrow in their widths [81,82]. In the case of kagomé spin lattice magnets, it is often difficult to detect magnetization plateaus in terms of their M vs. H curves. Therefore, sometimes dM/dH vs. H plots have been employed to discuss the plateaus [83,84,85]. However, 1/3-plateaus are clearly observed in their M vs. H plots for trigonal spin lattice magnets. In the following, we examine the probable cause of this difference by regarding the kagomé and trigonal spin lattices as made up of non-overlapping ferrimagnetic fragments, namely, ferrimagnetic triangles indicated by shading in Figure 50a and Figure 50b, respectively. As discussed in Section 2.4, each ferrimagnetic triangle can assume three different spin arrangements (Figure 50c). Then, all possible ordered and disordered spin configurations representing the 1/3-magnetization plateau are generated by how each ferrimagnetic triangle adopts one of the three spin arrangements. For example, Figure 51 shows three ordered spin arrangements creating the 1/3-plateau state for a kagomé spin lattice, and Figure 52 shows those for a trigonal spin lattice. To probe the question of whether kagomé and trigonal spin lattices have a 1/3-magnetization plateau, we note that the magnetization of the whole spin lattice remains at Msat/3 regardless of whether there are more or fewer inter-fragment magnetic bonds. Thus, in the following, we examine the most and least stable arrangements that a given (↑↓↑) ferrimagnetic fragment can have with the surrounding ferrimagnetic fragments.
Let us first examine possible spin arrangements around one ferrimagnetic fragment in a kagomé spin lattice. As depicted in Figure 53a, each shaded triangle, representing a ferrimagnetic fragment, is corner-shared with three unshaded triangles. The two sites on each edge of the unshaded triangle belong to two different ferrimagnetic fragments (see Figure 51) so that each ferrimagnetic fragment interacts with six different ferrimagnetic neighboring fragments. A chosen ferrimagnetic fragment makes the most stable inter-fragment spin arrangement by making six inter-fragment magnetic bonds (Figure 53b), and the least stable spin arrangement by making six inter-fragment broken bonds (Figure 53c). Obviously, it is not possible for every ferrimagnetic fragment to make six magnetic bonds with the six adjacent ferrimagnetic fragments. For, in making six bonds (broken bonds) with a chosen fragment, the six surrounding ferrimagnetic fragments should possess specific spin arrangements. These arrangements cannot be altered to make six bonds (broken bonds) for another ferrimagnetic fragment next to the chosen fragment. This means that there is a variation in the number of inter-fragment magnetic bonds each ferrimagnetic fragment can make, from six bonds to six broken bonds. The kagomé spin lattice as a whole is more (less) stable if it has more inter-fragment bonds (broken bonds) on average, implying that a good indicator for the width of the 1/3-plateau is the energy difference between the most and the least stable inter-fragment magnetic bonding. This energy difference corresponds to effectively 12 magnetic bonds, i.e., from six bonds (Figure 53b) to six broken bonds (Figure 53c) per ferrimagnetic fragment.
In a trigonal spin lattice, each ferrimagnetic fragment of a trigonal lattice is surrounded by 12 unshaded triangles and interacts with six adjacent ferrimagnetic fragments (Figure 54a). Three of these six make interactions through a corner, and the remaining three through an edge (indicated by red rectangles in Figure 54a). That is, in a trigonal spin lattice as well, a given ferrimagnetic fragment is surrounded by six ferrimagnetic fragments. In the interactions through a corner, the corner spin site can be either up-spin or down-spin. In the interactions through an edge, the two spins on the edge can be both up-spins or a combination of one up-spin and one down-spin, because this edge is a part of a (↑↓↑) triangle. Consequently, with six adjacent ferrimagnetic fragments, a ferrimagnetic fragment makes nine bonds and three broken bonds (i.e., effectively six bonds) in the most stable spin arrangement (Figure 54b), but two bonds and 10 broken bonds (i.e., effectively, eight broken bonds) in the least stable arrangement (Figure 54c). Thus, the energy difference between the most stable and the least stable arrangements is effectively 14 bonds.
The above analysis indicates that, between the most stable and the least stable arrangements, the trigonal spin lattice has only a slightly greater energy difference than the kagomé spin lattice (i.e., 14 vs. 12 magnetic bonds). From this, one might be led to speculate if trigonal and kagomé spin lattices have similar 1/3-plateau properties. However, we note that the end point of the 1/3-plateau occurs when a ferrimagnetic fragment starts to have a configuration change from (↑↓↑) to (↑↑↑). In a trigonal spin lattice, the (↑↓↑) to (↑↑↑) spin flip requires the breaking of six bonds in the most stable inter-fragment arrangement (Figure 54b), and that of two bonds in the least stable inter-fragment arrangement (Figure 54c). Namely, the spin flip requires energy in the most and least stable inter-fragment arrangements. This is not the case for a kagomé spin lattice. There, the (↑↓↑) to (↑↑↑) spin flip requires the breaking of four magnetic bonds at the site of the most stable inter-fragment arrangement (Figure 53b), but no energy at the site of the least stable inter-fragment arrangement (Figure 53c) because breaking two bonds within a ferrimagnetic fragment generates two bonds between the fragments. This implies that, during the field sweep from the most stable to the least stable distribution of the inter-fragment magnetic bonding, Zeeman energy causes (↑↓↑) to (↑↑↑) spin flips at certain down-spin sites with less favorable bonding connections with its neighboring fragments (e.g., Figure 53c,d) because, in such a case, the spin flip requires less energy than does the breaking of the inter-fragment bonds. This reasoning predicts that a kagomé spin lattice has a narrower 1/3-magnetization plateau than does a trigonal spin lattice, and this might be the reason why the M vs. H curves of kagomé spin lattices show a steady increase in magnetization with the field through the Msat/3 point.

5.2. Variation in the 1/3-Plateau Widths in RbFe(MoO4)2, Ba3CoSb2O9 and Ba2LaNiTe2O12

The 2D antiferromagnets RbFe(MoO4)2 [86], Ba3CoSb2O9 [87] and Ba2LaNiTe2O12 [82] consist of trigonal layers made up of MO6 (M = Fe, Co, Ni) octahedra (Figure 55a). In RbFe(MoO4)2, the upper and lower surfaces of such a layer are condensed by corner-sharing with MoO4 tetrahedra (Figure 55b,c) in RbFe(MoO4)2, with Sb2O9 double octahedra in Ba3CoSb2O9, and with TeO6 octahedra in Ba2LaNiTe2O12 (Figure 55d). RbFe(MoO4)2 consists of trigonal layers of Fe3+ (d5, S = 5/2) ions, Ba3CoSb2O9—those of Co2+ (d7, S = 3/2) ions—and Ba2LaNiTe2O12—those of Ni2+ (d8, S = 1) ions. RbFe(MoO4)2 undergoes a phase transition at TN = 3.8 K into a 120° spin structure with all the spins confined in the basal plane. Application of an in-plane magnetic field induces a collinear spin state between 4.7 and 7.1 T, producing a 1/3-magnetization plateau (Figure 56a) [88]. Ba3CoSb2O9 exhibits an AFM transition at TN = 3.8 K and the powder neutron diffraction measurements show that it adopts a 120° spin structure in the ab-plane [87]. Under the magnetic field applied in the ab-plane, Ba3CoSb2O9 exhibits a 1/3-magnetization plateau between 10 and 15 T (Figure 56b) [89]. Ba2LaNiTe2O12 undergoes successive magnetic phase transitions at TN1 = 9.8 K and TN2 = 8.9 K [90]. The ground state is accompanied by a weak ferromagnetic moment, suggesting that it adopts a slightly canted 120° spin structure. The magnetization curve exhibits a 1/3-magnetization plateau between 35 and 45 T (Figure 56c) [82].
The observed widths Δ(μ0H) of the 1/3-plateaus found for RbFe(MoO4)2, Ba3CoSb2O9 and Ba2LaNiTe2O12 are 2.4, 5.0 and ~10 T, respectively. In the previous section, we argued that the energy difference between the most stable and the least stable arrangements involving a (↑↓↑) ferrimagnetic triangle amounts to 14 nearest-neighbor magnetic bonds J. Thus, the Δ(μ0H) values of these magnets should be related to their nearest-neighbor spin exchanges J as Δ(μ0H) ∝ J. Precisely speaking, the spin exchange between two magnetic ions of spin S coupled by exchange constant J generates the energy JS2 (Equation (1)). Since we compare the relative strengths of the magnetic bonds involving the ions of different spins, it is necessary to use the relationship Δ(μ0H) ∝ JS2. Then, according to the observed experimental Δ(μ0H) values, the JS2 values should increase in the order, RbFe(MoO4)2 < Ba3CoSb2O9 < Ba2LaNiTe2O12.
As depicted in Figure 57a, the nearest-neighbor exchange J in the three magnets is of the M-O…O-M type exchange. In general, the strength of such an exchange becomes stronger as the O…O contact distance decreases [2,91]. As summarized in Figure 57b, the O…O contact distances of the J exchange paths decrease in the order RbFe(MoO4)2 > Ba3CoSb2O9 > Ba2LaNiTe2O12. The standard deviation of the O…O distance in Ba2LaNiTe2O12 is rather large. However, even if the largest O…O distance of 2.74 Å allowed by the standard deviation is considered, the trend RbFe(MoO4)2 > Ba3CoSb2O9 > Ba2LaNiTe2O12 still remains valid. We carried out an energy mapping analysis based on DFT+U calculations (see Sections S11–S13 of the Supplementary Materials) to find that J = 1.5 K for RbFe(MoO4)2, J = 6.2 K for Ba3CoSb2O9 and J = 56 K for Ba2LaNiTe2O12. (As expected, the J value of Ba2LaNiTe2O12 is very large due to the unusually short O…O distance of 2.67 Å reported in the structure determination. A more accurate crystal structure would reduce the J value). As summarized in Figure 57b, the JS2 values of the three antiferromagnets increase in the order RbFe(MoO4)2 < Ba3CoSb2O9 < Ba2LaNiTe2O12. This provides experimental and theoretical support for our arguments presented in the previous section.

6. Complex Clusters

6.1. Trimer–Dimer Zigzag Chains for the 3/5-Plateau in Na2Cu5(Si2O7)2

Sodium copper pyrosilicate, Na2Cu5(Si2O7)2, consists of ferrimagnetic zigzag chains in which trimer units alternate with dimer units (Figure 58a) [92]. Each trimer becomes ferrimagnetic due to the nearest-neighbor AFM exchange (J1); the exchange (J2) between adjacent trimer and dimer units is AFM, and the dimer exchange (J3) is FM [92]. Thus, the ground state of the zigzag chain is ferrimagnetic (Figure 58b). Since Na2Cu5(Si2O7)2 undergoes a 3D AFM ordering below TN = 8K, the interchain interaction is weakly AFM. The fitting analysis of the magnetic susceptibility data using the ferrimagnetic chain model led to J1 = 236 K, J2 = 8 K and J3 = −40 K [92]. The repeat unit of this ferrimagnetic state has the (3↑2↓) configuration with M = Msat/5. Under a magnetic field, Na2Cu5(Si2O7)2 exhibits a 3/5-magnetization plateau as shown in Figure 58c [93]. It occurs because the weak J2 bond is broken under field leading to the higher-energy ferrimagnetic state (Figure 58d) with the (4↑1↓) configuration and hence M = 3Msat/5. To increase the magnetization beyond 3Msat/5, the two J1 bonds in each trimer should be broken. Since J1 is a strong bond, this does not occur unless the magnetic field is strong enough. Thus, the 3/5-magnetization plateau arises.

6.2. Linear Heptamer of One Trimer and Two Dimers for the 3/7-Plateau in Y2Cu7(TeO3)6Cl6(OH)2

Viewed solely geometrically, the Cu2+ ions of Y2Cu7(TeO3)6Cl6(OH)2 make chains of diamond-like tetramers which are interconnected by linear trimers [94]. In each trimer, a CuO2Cl2 plane corner-shares its oxygen atoms with two CuO3Cl planes such that the adjacent CuO2Cl2 and CuO3Cl planes are nearly perpendicular (Figure 59a). In each diamond-like tetramer, the planes of the two Cu2O3Cl dimers are separated and are nearly parallel to each other (Figure 59b). When viewed from the point of the magnetic orbitals contained in square planar units containing Cu2+ ions, a somewhat different picture emerges. In each diamond-like tetramer, the spin exchange between two Cu2O3Cl dimer units cannot be strong because their magnetic orbital planes are nearly parallel to each other. However, each Cu2O3Cl dimer can interact with an adjacent trimer through the Cu-O…O-Cu type spin exchange because the O…O distance is short (2.659 Å) and the Cu-O bonds are nearly directed toward each other (∠Cu-O…O = 158.8, 165.7°). Thus, each trimer is connected to two adjacent Cu2O3Cl dimers forming a linear heptamer (Figure 59c), and such heptamers are expected to be important for Y2Cu7(TeO3)6Cl6(OH)2.
The magnetization curve presents a field-induced metamagnetic transition at 0.2 T, which is followed by a magnetization plateau within a wide magnetic field range from 7 T to at least 55 T (Figure 59d). To account for this observation, an energy-mapping analysis based on DFT+U calculations was carried out to find the spin exchanges (in K), summarized in Figure 60a (see Section S14 of the Supplementary Materials). The latter shows that the dominating AFM spin exchanges are the inter-trimer–dimer exchange and the intradimer exchange J3. The intra-trimer exchange J2 is FM, and so are the exchanges between the Cu2O3Cl units within each diamond-like tetramer, but their magnitudes are weaker. Therefore, these considerations lead to the (5↑2↓) spin arrangement for each linear heptamer (Figure 60b). One heptamer is coupled to two other heptamers through the AFM exchanges J5, leading to an AFM chain of heptamers (Figure 60c,d). (Since Y2Cu7(TeO3)6Cl6(OH)2 undergoes an AFM ordering below TN = 4.1 K, the chains of heptamers have a very weak AFM interchain coupling). Therefore, the breaking of all J5 bonds is necessary to reach the M = 3Msat/7 point. To increase the magnetization beyond 3Msat/7, it is necessary to break the intradimer J3 bonds. These bonds are strong so that their breaking does not take place unless the magnetic field is strong enough. This explains the occurrence of the 3/7-plateau.

6.3. Zigzag Pentamer as an Effective S = 1/2 Unit in Cu5(VO4)2(OH)4+

Turanite, Cu5(VO4)2(OH)4, has layers made up of three nonequivalent CuO6 octahedra, which are interconnected by VO4 groups. Within each layer, the CuO4 square planes containing the x2-y2 orbitals are arranged as presented in Figure 61a [95], so the pattern of the Cu2+ ion arrangement has interconnected chains of edge-sharing hexagons composed of six triangles (hereafter, the hexagon chains, for short) as depicted in Figure 61b. This magnet undergoes a ferrimagnetic ordering at TC = 4.5 K, and its magnetization evidences a rapid increase below about 0.01 T. The latter is followed by a much slower increase, eventually reaching a 1/5-magnetization plateau at 8 T (Figure 61c) [96]. There are two puzzling observations to note: the M vs. H curve is smooth and resembles that observed for a paramagnet of S = 1/2 ions, and only ~23.6% of the spins participate in the ferrimagnetic ordering, which led to the suggestion that the remaining spins are still fluctuating [96].
To probe the cause for these observations, it is necessary to know the spin exchanges in each layer of interlinked hexagon chains (Figure 61b). What matters for spin exchanges is not the geometrical arrangement of the magnetic ions but that of their magnetic orbitals. To see if the spin lattice of Cu5(VO4)2(OH)4 was spin-frustrated, we examined the seven spin exchanges defined in Figure 62a. Note that due to the absence of a vertical mirror plane of symmetry in each hexagon chain, the exchanges J5 and J6 were treated as different (see Figure 62b), and so were the spin exchanges J3 and J4. The spin exchanges of J1J7 determined using DFT+U calculations are summarized in Figure 62c (see Section S15 of the Supplementary Materials), from which we observed the following:
(1)
Within each hexagon chain, the AFM exchanges J1 and J5 dominated over the FM exchanges J2, J4 and J6 so that there was effectively no spin frustration in all spin triangles of the hexagon chains.
(2)
The exchanges J1 and J5 formed zigzag pentamer ferrimagnetic fragments of (3↑2↓) spin configuration with M = Msat/5 (Figure 62d).
(3)
Since J6 is more strongly FM than J4, each hexagon chain preferred to have adjacent (3↑2↓) ferrimagnetic fragments to have an AFM coupling rather than an FM coupling within each hexagon chain (Figure 63a,b).
(4)
Since the interchain exchange J7 is AFM, an AFM coupling was preferred to an FM coupling between adjacent (3↑2↓) ferrimagnetic fragments between hexagon chains.
(5)
Thus, the most stable arrangement between adjacent (3↑2↓) ferrimagnetic fragments was AFM both within and between hexagon chains (Figure 63c), and the least stable arrangement was an FM both within and between hexagon chains (Figure 63d).
Though seemingly paradoxical, the above theoretical analysis is entirely consistent with the experimental observations for Cu5(VO4)2(OH)4. The key point to understand is that the AFM exchanges J5 and J1 leading to a (3↑2↓) ferrimagnetic fragment are very strong, so that each (3↑2↓) ferrimagnetic fragment acts as an effective S = 1/2 unit. Indeed, the observed magnetization curve is similar to the one found for a paramagnet of S = 1/2 ions (see the red curve in Figure 61c). This realization explains the low-temperature magnetic properties of Cu5(VO4)2(OH)4, namely, why only one out of five spins appears to participate in the magnetic ordering below TC = 4.5 K and why the magnetization behavior resembles that expected for a paramagnet of S = 1/2 ions. The magnetic susceptibility of Cu5(VO4)2(OH)4 reveals the presence of five spins per formula unit at high temperature, because thermal agitation would break the AFM coupling leading to the (3↑2↓) ferrimagnetic fragment. As already discussed in Section 4.1.3, such a phenomenon of reduced spin moments due to strong AFM coupling was also found for volborthite, which consists of two-leg spin ladders with rungs made up of linear trimers of Cu2+ ions. In this case, the AFM coupling between adjacent Cu2+ ions was so strong that each linear trimer acted as an effective S = 1/2 unit.

6.4. Cu7 Cluster of Corner-Sharing Tetrahedra for the 3/7-Plateau in Pb2Cu10O4(SeO3)4Cl7 and Na2Cu7(SeO3)4O2Cl4

Pb2Cu10O4(SeO3)4Cl7 has a complex crystal structure consisting of one nonmagnetic Cu+ (S = 0) ion and nine Cu2+ (S = 1/2) ions per formula unit [97]. Of the nine Cu2+ ions, two were found in a dimer unit and seven in a heptamer unit made up of two corner-sharing (Cu2+)4 tetramers (Figure 64a). Pb2Cu10O4(SeO3)4Cl7 orders antiferromagnetically at TN = 10.2 K and below this temperature exhibited a sequence of spin-flop transition at 1.3 T and a 1/3-plateau at 4.4 T, which persisted at least up to 53.5 T (Figure 64b). The spin exchanges of Pb2Cu10O4(SeO3)4Cl7 evaluated using DFT+U calculations showed that the magnetic properties of Pb2Cu10O4(SeO3)4Cl7 were governed by the ferrimagnetic heptamer (Cu2+)7 with (5↑2↓) spin configuration (Figure 64c). The heptamers formed chains (Figure 64d) with interchain AFM coupling, so the magnetic ground state of the chain was an AFM state (Figure 64e). Under the magnetic field, the inter-cluster bonds became broken, eventually reaching the ferrimagnetic state in which the ferrimagnetic clusters were ferromagnetically coupled (Figure 64f). A further increase in magnetization required a high magnetic field because it was necessary to break the magnetic bonds within a ferrimagnetic cluster, hence leading to the 3/7-plateau.
Note that this discussion is based solely on the seven Cu2+ ions of a heptamer (Cu2+)7. The two Cu2+ ions, strongly coupled antiferromagnetically in a dimer, were magnetically “silent”. If we include these two magnetic ions in our analysis, the 3/7-plateau discussed above becomes equivalent to a 1/3-plateau. A similar 3/7-plateau was found for Na2Cu7(SeO3)4O2Cl4 (Figure 65a) [98], which also consisted of (Cu2+)7 heptamers made up of two corner-sharing (Cu2+)4 tetramers (Figure 65b). Na2Cu7(SeO3)4O2Cl4 differs from Pb2Cu10O4(SeO3)4Cl7 in the bridging mode between adjacent heptamers (Figure 65c), but the composition of the heptamers is identical, namely, it is composed of five square planar and two trigonal bipyramid units (Figure 66a). Our DFT+U calculations summarized in Figure 65d (see Section S16 of the Supplementary Materials) show that each (Cu2+)7 heptamer had a (5↑2↓) spin configuration (Figure 64e), thereby explaining why Na2Cu7(SeO3)4O2Cl4 exhibits a 3/7-plateau as does Pb2Cu10O4(SeO3)4Cl7.
Finally, we comment on why the (Cu2+)7 heptamer adopts the (5↑2↓) spin configuration. From one trigonal bipyramid (TBP) to the central square plane (SP) to another trigonal bipyramid (TBP) in a heptamer, the three Cu2+ ions form a linear path (Figure 66a), and the two nearest-neighbor spin exchanges are of the Cu-O-Cu type. These two spin exchanges are strongly AFM in this linear path because the atoms associated with the Cu-O-Cu-O-Cu exchange paths are coplanar, so that the a1 magnetic orbitals of the two TBPs [99] and the x2-y2 magnetic orbital of the central SP are coplanar (Figure 66b). This makes the in-plane 2p orbitals of the two bridging O atoms intact efficiently with the a1 and x2-y2 magnetic orbitals, leading to a strong ↑↓↑ coupling of the three Cu2+ spins along the TBP-SP-TBP linear path. Note that the central SP is nearly orthogonal to every SP at both ends. This makes their x2-y2 magnetic orbitals nearly orthogonal to each other, so the associated spin exchange becomes FM. Consequently, the heptamer adopts the (5↑2↓) spin configuration shown in Figure 66c.

7. Concluding Remarks

In an effort to find a conceptual picture describing the magnetization plateau phenomenon, we surveyed the crystal structures, the spin exchanges and the spin lattices of numerous magnets exhibiting magnetic plateaus. Our analyses show that an important key to understanding this phenomenon is the realization that a magnet under field absorbs Zeeman energy in accordance with Le Chartlier’s principle, which occurs by breaking its magnetic bonds. For a magnet with spin lattice defined by several spin exchanges of different strengths, its weakest bonds are broken preferentially to partition the spin lattice into either antiferromagnetic or ferrimagnetic fragments, which fill the whole spin lattice without overlapping each other. For a magnet with a spin-frustrated spin lattice defined by a few spin exchanges of comparable strengths, the weaker magnetic bonds are broken to partition the spin lattice into small ferrimagnetic fragments filling the whole spin lattice without overlapping each other. Such field-induced fragmentation is influenced by the spin-lattice interactions brought about by the fragmentation.
As illustrated in this survey, the conceptual aspects of the magnetization plateau phenomenon in any magnet can be readily explained or predicted once its crystal structure and its spin exchanges are known. Given the crystal structure of a magnet, it is straightforward to evaluate the spin exchanges for various exchange paths between its magnetic ions, and hence the relative strengths of its magnetic bonds, using the energy-mapping analysis based on DFT calculations. It goes without saying that this approach is not designed to provide quantitative descriptions. The latter lie in the realm of quantitative calculations using model Hamiltonians with a minimal number of adjustable parameters (e.g., spin exchanges) to generate the magnetic energy spectrum of a magnet under investigation. Even with powerful computers currently available, such quantitative analyses cannot be carried out for most magnets because their spin lattices are complex and low in symmetry. The conceptual picture of the magnetization plateau phenomenon, based on the supposition of field-induced partitioning of a spin lattice into magnetic fragments, is valid for all magnets regardless of whether their spin lattices are complex or not.
The magnetization plateau phenomenon can be highly anisotropic, as found for the Ising magnets Ca3Co2O6 and CoGeO3, in that their 1/3-magnetization plateaus observed with the field along the easy axis do not occur if the field is perpendicular to the easy axis. A strong plateau anisotropy, though weaker than those found for the Ising magnets, is also observed for Cs2Cu3(SeO3)4·2H2O and azurite Cu3(CO3)2(OH)2. In Cs2Cu3(SeO3)4·2H2O, the value of M = Msat/3 depends on the field direction (i.e., Hc vs. H||c) because the Cu2+ ion has a higher spin moment when the magnetic field is perpendicular than parallel to the CuO4 square plane. In azurite, the width of the 1/3-magnetization plateau depends on the field direction (H||b vs. Hb) for three reasons: (1) the Dzyaloshinskii–Moriya interactions between the Cu2+ ions depend on the relative orientations of their CuO4 square planes, (2) the spin moment of a Cu2+ ion depends on the field direction with respect to the CuO4 square plane, and (3) the spin lattice of azurite is three-dimensional with nonnegligible interlayer spin exchange. In both Cs2Cu3(SeO3)4·2H2O and azurite, the strong anisotropy of their magnetization plateaus stems from the presence of near orthogonal arrangements of their CuO4 square planes. This emphasizes once more the importance of analyzing the structural chemistry associated with magnetic ion arrangements.
Our supposition that the spin lattice of a magnet exhibiting one or more magnetic plateaus is partitioned into magnetic fragments is supported by the experimental observation that an Ising magnet can exhibit a magnetization plateau when the applied field is parallel to the easy axis of the magnet, but this magnetization plateau disappears when the field is perpendicular to the easy axis. This reflects that Zeeman energy, being a dot product between the magnetic field and the spin moment, is nonzero for the parallel field but zero for the perpendicular field. More support comes from the observation that the highly anisotropic width of the 1/3-magnetization plateau (H||b vs. Hb) in azurite Cu3(CO3)2(OH)2 arises from the field-dependent Zeeman energy available for the magnet.
This survey reflects our efforts to comprehend the magnetization plateau phenomenon based on the relative strengths of magnetic bonds. Thus, our discussion focused on the arrangement of magnetic ions and their spin exchanges leading to the spin lattices responsible for magnetization plateaus. It is our hope that the conceptual picture of the magnetization plateau phenomenon presented in this survey will promote further developments in this and related research fields.
Near the completion stage of this survey, new magnets were reported to exhibit magnetization plateaus at low temperatures. They include CsCo2Br4 [100], YCu3(OD)6+xBr3−x (x ≈ 0.5) [79], Ni2V2O7 [101], TbTi3Bi4 [102], TbRh6Ge4 [103], GdInO3 [104], Cu5(PO4)2(OH)4 [105], Ba2Cu3(SeO3)4F2 [106], Sr2CoTeO6 [107], Cu3Bi(TeO3)2O2Cl [108] and Na3Ni2BiO6 [109]. We expect that the magnetic plateau phenomena of these new magnets can be readily explained using the concept of the field-induced partitioning of their spin lattices once their spin exchanges are determined.
In this survey, we focused on the “classical” magnetization plateaus that are readily described by the field-induced partitioning of spin lattices into spin superstructures. It should be pointed out that there may be cases that require a more sophisticated description beyond this classical picture [110,111,112,113,114,115,116,117,118,119,120,121,122,123,124,125]. For instance, as recently reviewed by Yoshida [125], quantum plateau phases may emerge from quantum spin liquids.
Note added in proof: The essential results of our paper were recently summarized and commented in arXivpp [126].

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/condmat9040045/s1, DFT calculations for the spin exchanges of CoGeO3 in S1, Ba3Mn2O8 in S2, KCuCl3 in S4, TlCuCl3 in S5, the YY structures of NH4CuCl3 in S6, the NY structures of NH4CuCl3 in S7, the NN structures of NH4CuCl3 in S8, K2Cu3O(SO4)3 in S9, Cu3(CO3)2(OH)2 in S10, RbFe(MoO4)2 in S11, Ba3CoSb2O9 in S12, Ba2LaNiTe2O12 in S13, Y2Cu7(TeO3)6Cl6(OH)2 in S14, Cu5(VO4)2(OH)4 in S15, Na2Cu7(SeO3)4O2Cl4 in S16. Supplementary Figures S1–S3 in S3.

Funding

The work at KHU was supported by the National Research Foundation of Korea of the Ministry of Education through the grant number 2020R1A6A1A03048004, and the work at MSU by Russian Ministry of Science and Higher Education through grant No. 075-15-2021-1353.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dai, D.; Whangbo, M.-H. Analysis of the uniaxial magnetic properties of high-spin d6 ions at trigonal prism and linear two-coordinate sites: Uniaxial magnetic properties of Ca3Co2O6 and Fe[C(SiMe3)3]2. Inorg. Chem. 2005, 44, 4407–4414. [Google Scholar] [CrossRef] [PubMed]
  2. Dai, D.; Xiang, H.J.; Whangbo, M.-H. Effects of spin-orbit coupling on magnetic properties of discrete and extended magnetic systems. J. Comput. Chem. 2008, 29, 2187–2209. [Google Scholar] [CrossRef] [PubMed]
  3. Zhang, Y.; Kan, E.J.; Xiang, H.J.; Villesuzanne, A.; Whangbo, M.-H. Density Functional Theory Analysis of the Interplay between Jahn-Teller Instability, Uniaxial Magnetism, Spin Arrangement, Metal-Metal Interaction, and Spin-Orbit Coupling in Ca3CoMO6 (M = Co, Rh, Ir). Inorg. Chem. 2011, 50, 1758–1766. [Google Scholar] [CrossRef] [PubMed]
  4. Honecker, A.; Schulenburg, J.; Richter, J. Magnetization plateaus in frustrated antiferromagnetic quantum spin models. J. Phys. Condens. Matter 2004, 16, S749. [Google Scholar] [CrossRef]
  5. Takigawa, M.; Mila, F. Magnetization plateaus. In Introduction to Frustrated Magnetism; Lacroix, C., Mendels, P., Mila, F., Eds.; Springer: New York, NY, USA, 2010; pp. 241–268. [Google Scholar]
  6. Xiang, H.J.; Lee, C.; Koo, H.-J.; Gong, X.G.; Whangbo, M.-H. Magnetic properties and energy-mapping analysis. Dalton Trans. 2013, 42, 823–853. [Google Scholar] [CrossRef] [PubMed]
  7. Penc, K.; Shannon, N.; Shiba, H. Half-magnetization plateau stabilized by structural distortion in the antiferromagnetic Heisenberg model on a pyrochlore lattice. Phys. Rev. Lett. 2004, 93, 197202. [Google Scholar] [CrossRef]
  8. Aoyama, K.; Gen, M.; Kawamura, H. Effects of spin-lattice coupling and a magnetic field in classical Heisenberg antiferromagnets on the breathing pyrochlore lattice. Phys. Rev. B 2021, 104, 184411. [Google Scholar] [CrossRef]
  9. Gen, M.; Ikeda, A.; Aoyama, K.; Jeschke, H.O.; Ishii, Y.; Ishikawa, H.; Yajima, T.; Okamoto, Y.; Zhou, X.; Nakamura, D.; et al. Signatures of a magnetic superstructure phase induced by ultrahigh magnetic fields in a breathing pyrochlore antiferromagnet. Proc. Nat. Acad. Sci. USA 2023, 120, e2302756120. [Google Scholar] [CrossRef]
  10. Onizuka, K.; Kageyama, H.; Narumi, Y.; Kindo, K.; Ueda, Y.; Goto, T. 1/3 magnetization plateau in SrCu2(BO3)2—Stripe order of excited triplets. J. Phys. Soc. Jpn. 2000, 69, 1016–1018. [Google Scholar] [CrossRef]
  11. Jaime, M.; Daou, R.; Crooker, S.A.; Gaulin, B.D. Magnetostriction and magnetic texture to 100.75 Tesla in frustrated SrCu2(BO3)2. Proc. Natl. Acad. Sci. USA 2012, 109, 12404. [Google Scholar] [CrossRef]
  12. Matsuda, Y.H.; Abe, N.; Takeyama, S.; Kageyama, H.; Corboz, P.; Honecker, A.; Manmana, S.R.; Foltin, G.R.; Schmidt, K.P.; Mila, F. Magnetization of SrCu2(BO3)2 in Ultrahigh Magnetic Fields up to 118 T. Phys. Rev. Lett. 2013, 111, 137204. [Google Scholar] [CrossRef] [PubMed]
  13. Nomura, T.; Corboz, P.; Miyata, A.; Zherlitsyn, S.; Ishii, Y.; Kohama, Y.; Matsuda, Y.H.; Ikeda, A.; Zhong, C.; Kageyama, H.; et al. Unveiling new quantum phases in the Shastry-Sutherland compound SrCu2(BO3)2 up to the saturation magnetic field. Nat. Commun. 2023, 14, 3769. [Google Scholar] [CrossRef] [PubMed]
  14. Néel, L. Les métamagnétiques ou substances antiferromagnétiques à champ seuil. II Nuovo C. 1957, 6, 942–960. [Google Scholar] [CrossRef]
  15. Morosov, A.I.; Sigov, A.S. Surface spin-flop transition in an antiferromagnet. Physics-Uspekhi 2010, 53, 677–689. [Google Scholar] [CrossRef]
  16. Strujewski, E.; Giordano, N. Metamagnetism. Adv. Phys. 1977, 26, 487–650. [Google Scholar] [CrossRef]
  17. Nawa, K.; Avdeev, M.; Perdonosov, P.; Sobolev, A.; Presniakov, I.; Aslandukova, A.; Kozlyakova, E.; Vasiliev, A.; Shchetinin, I.; Sato, T. Magnetic structure study of the sawtooth chain antiferromagnet Fe2Se2O7. Sci. Rep. 2021, 11, 24049. [Google Scholar] [CrossRef]
  18. Zakharov, K.V.; Zvereva, E.A.; Berdonosov, P.S.; Kuznetsova, E.S.; Dolgikh, V.A.; Clark, L.; Black, C.; Lightfoot, P.; Kockelmann, W.; Pchelkina, Z.V.; et al. Thermodynamic properties, electron spin resonance, and underlying spin model in Cu3Y(SeO3)2O2Cl. Phys. Rev. B 2014, 90, 214417. [Google Scholar] [CrossRef]
  19. Zakharov, K.V.; Zvereva, E.A.; Kuznetsova, E.S.; Kuznetsova, E.S.; Berdonosov, P.S.; Dolgikh, V.A.; Markina, M.M.; Olenev, A.V.; Shakin, A.A.; Volkova, O.S.; et al. Two new lanthanide members of francisite family Cu3Ln(SeO3)2O2Cl (Ln = Eu, Lu). J. Alloys Compd. 2016, 685, 442–447. [Google Scholar] [CrossRef]
  20. Whangbo, M.-H.; Xiang, H.J. Magnetic Properties from the Perspectives of Electronic Hamiltonian: Spin Exchange Parameters, Spin Orientation and Spin-Half Misconception. In Handbook of Solid State Chemistry, Volume 5: Theoretical Descriptions; Dronskowski, R., Kikkawa, S., Stein, A., Eds.; Wiley: Hoboken, NJ, USA, 2017; pp. 285–343. [Google Scholar]
  21. Whangbo, M.-H.; Koo, H.-J.; Kremer, R.K. Spin Exchanges between Transition Metal Ions Governed by the Ligand p-Orbitals in Their Magnetic Orbitals. Molecules 2021, 26, 531. [Google Scholar] [CrossRef]
  22. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  23. Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  24. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [PubMed]
  25. Dudarev, S.L.; Botton, G.A.; Savrasov, S.Y.; Humphreys, C.J.; Sutton, A.P. Electron-energy-loss spectra and the structural stability of nickel oxide: An LSDA+U study. Phys. Rev. B 1998, 57, 1505–1509. [Google Scholar] [CrossRef]
  26. Kuneš, K.; Novák, P.; Schmid, R.; Blaha, P.; Schwarz, K. Electronic structure of fcc Th: Spin-orbit calculation with 6p1/2 local orbital extension. Phys. Rev. Lett. 2001, 64, 153102. [Google Scholar]
  27. Mazurenko, V.V.; Skornyakov, S.L.; Anisimov, V.I.; Mila, F. First-principles investigation of symmetric and antisymmetric exchange interactions of SrCu2(BO3)2. Phys. Rev. B 2008, 78, 195110. [Google Scholar] [CrossRef]
  28. Miyahara, S.; Ueda, K. Theory of the orthogonal dimer Heisenberg spin model for SrCu2(BO3)2. J. Phys. Condens. Matter 2003, 15, R237–R366. [Google Scholar] [CrossRef]
  29. Shastry, B.S.; Sutherland, B. Exact ground state of a quantum mechanical antiferromagnet. Phys. B&C 1981, 108, 1069–1070. [Google Scholar]
  30. Kageyama, H.; Yoshimura, K.; Stern, R.; Mushnikov, N.; Onizuka, K.; Kato, M.; Kosuge, K.; Slichter, C.P.; Goto, T.; Ueda, Y. Exact dimer ground state and quantized magnetization plateaus in the two-dimensional spin system SrCu2(BO3)2. Phys. Rev. Lett. 1999, 82, 3168–3171. [Google Scholar] [CrossRef]
  31. Ueda, H.; Katori, H.A.; Mitamura, H.; Goto, H.; Takagi, H. Magnetic-field induced transition to the 1/2 magnetization plateau state in the geometrically frustrated magnet CdCr2O4. Phys. Rev. Lett. 2005, 94, 047202. [Google Scholar] [CrossRef]
  32. Chung, J.H.; Matsuda, M.; Lee, S.-H.; Kakurai, K.; Ueda, H.; Sato, T.J.; Takagi, H.; Hong, K.-P.; Park, S. Statics and dynamics of incommensurate spin order in a geometrically frustrated antiferromagnet CdCr2O4. Phys. Rev. Lett. 2005, 95, 247204. [Google Scholar] [CrossRef]
  33. Peakor, D.R. The crystal structure of CoGeO3. Z. Krist. 1968, 126, 299–306. [Google Scholar] [CrossRef]
  34. Guo, H.; Zhao, M.; Baenitz, X.; Gukasov, A.; Melendez Sans, A.; Khomskii, D.I.; Tjeng, L.H.; Komareck, A.C. Emergent 1/3 magnetization plateaus in pyroxene CoGeO3. Phys. Rev. Res. 2021, 3, L032037. [Google Scholar] [CrossRef]
  35. Whangbo, M.-H.; Gordon, E.E.; Xiang, H.J.; Koo, H.-J.; Lee, C. Prediction of spin orientations in terms of HOMO-LUMO interactions using spin-orbit coupling as perturbation. Acc. Chem. Res. 2015, 48, 3080–3087. [Google Scholar] [CrossRef]
  36. Zapf, V.; Jaime, M.; Batista, C.D. Bose-Einstein condensation in quantum magnets. Rev. Mod. Phys. 2014, 86, 563, Erratum in Rev. Mod. Phys. 2014, 86, 1453. [Google Scholar] [CrossRef]
  37. Weller, M.T.; Skinner, S.J. Ba3Mn2O8 determined from neutron powder diffraction. Acta Crystallogr. C 1999, 55, 154–156. [Google Scholar] [CrossRef]
  38. Uchida, M.; Tanaka, H.; Mitamura, H.; Ishkawa, F.; Goto, T. High-field magnetization process in the S = 1 quantum spin system Ba3Mn2O8. Phys. Rev. B 2002, 66, 054429. [Google Scholar] [CrossRef]
  39. Willett, R.D.; Dwiggins, C., Jr.; Kruh, R.F.; Rundle, R.E. Crystal structures of KCuCl3 and NH4CuCl3. J. Chem. Phys. 1963, 38, 2429–2436. [Google Scholar] [CrossRef]
  40. Ryu, G.; Son, K. Surface defect free growth of a spin dimer TlCuCl3 compound crystals and investigations on its optical and magnetic properties. J. Solid State Chem. 2016, 237, 358–363. [Google Scholar] [CrossRef]
  41. O’Bannon, G.; Willet, R.D. A redetermination of the crystal structure of NH4CuCl3 and a magnetic study of NH4CuX3 (X = Cl, Br). Inorg. Chim. Acta 1981, 53, L131–L132. [Google Scholar] [CrossRef]
  42. Rüegg, C.; Schefer, J.; Zaharko, O.; Furrer, A.; Tanaka, H.; Krämer, K.W.; Güdel, H.-U.; Vorderwisch, P.; Habicht, K.; Polinski, T.; et al. Neutron scattering study of the field-dependent ground state and the spin dynamics in spin-one-half NH4CuCl3. Phys. Rev. Lett. 2004, 93, 037207. [Google Scholar] [CrossRef]
  43. Matsumoto, M.; Normand, B.; Rice, T.M.; Sigrist, M. Field- and pressure-induced magnetic quantum phase transitions in TlCuCl3. Phys. Rev. B 2004, 69, 054423. [Google Scholar] [CrossRef]
  44. Oosawa, A.; Kato, T.; Tanaka, H.; Kakurai, K.; Müller, M.; Mikeska, H.-J. Magnetic excitations in the spin gap system TlCuCl3. Phys. Rev. B 2002, 65, 094426. [Google Scholar] [CrossRef]
  45. Tanaka, H.; Shiramura, W.; Takatsu, T.; Kurniawan, B.; Takahashi, M.; Kamishima, K.; Takizawa, K.; Mitamura, H.; Goto, T. High-field magnetization processes of quantum double spin chain systems KCuCl3, TlCuCl3 and NH4CuCl3. Phys. B 1998, 246–247, 230–233. [Google Scholar] [CrossRef]
  46. Rüegg, C. Neutron Scattering Investigations of the S = ½ Quantum Spin Systems TlCuCl3 and NH4CuCl3. Ph.D. Thesis, Laboratorium für Neutronenstreuung, ETH Zürich, Zürich, Switzerland, 2001. CH-5232 Villigen PSI. [Google Scholar]
  47. Hay, P.J.; Thibeault, J.C.; Hoffmann, R. Orbital interactions in metal dimer complexes. J. Am. Chem. Soc. 1975, 97, 4884–4899. [Google Scholar] [CrossRef]
  48. Volkova, O.S.; Shvanskaya, L.V.; Ovchenkov, E.A.; Zvereva, E.A.; Volkov, A.S.; Chareev, D.A.; Molla, K.; Rahaman, B.; Saha-Dasgupta, T.; Vasiliev, A.N. Structure-property relationships in α-, β’- and ℽ-modifications of Mn3(PO4)2. Inorg. Chem. 2016, 55, 10692–10700. [Google Scholar] [CrossRef]
  49. Baies, R.; Caignaert, V.; Pralong, V.; Raveau, B. Copper hydroxydiphosphate with a one-dimensional arrangement of copper polyhedra: Cu3[P2O6OH]2. Inorg. Chem. 2005, 44, 2376–2380. [Google Scholar] [CrossRef]
  50. Hase, M.; Kohno, M.; Kitazawa, H.; Tsujii, N.; Suzuki, O.; Ozawa, K.; Kido, G.; Imai, M.; Hu, X. 1/3 magnetization plateau observed in the spin-1/2 trimer chain compound Cu3(P2O6OH)2. Phys. Rev. B 2006, 73, 104419. [Google Scholar] [CrossRef]
  51. Koo, H.-J.; Whangbo, M.-H. Finite magnetization plateau from a two-dimensional antiferromagnet: Density functional analysis of the magnetic structure of Cu3(P2O6OH)2. Inorg. Chem. 2010, 49, 9253–9256. [Google Scholar] [CrossRef]
  52. Ishikawa, H.; Yamaura, J.; Okamoto, Y.; Yoshida, H.; Nilsen, G.J.; Hiroi, Z. A novel crystal polymorph of volborthite, Cu3V2O7(OH)2⋅2H2O. Acta Cryst. C 2012, 68, i41–i44. [Google Scholar] [CrossRef]
  53. Whangbo, M.-H.; Koo, H.-J.; Brücher, E.; Puphal, P.; Kremer, R.K. Absence of spin frustration in the kagomé layers of Cu2+ ions in volborthite Cu3V2O7(OH)2·2H2O and observation of the suppression and re-entrance of specific heat anomalies in volborthite under an external magnetic field. Condens. Matter 2022, 7, 24. [Google Scholar] [CrossRef]
  54. Yoshida, M.; Takigawa, M.; Yoshida, H.; Okamoto, Y.; Hiroi, Z. Phase diagram and spin dynamics in volborthite with a distorted kagomé lattice. Phys. Rev. Lett. 2009, 103, 077207. [Google Scholar] [CrossRef] [PubMed]
  55. Ishikawa, H.; Yoshida, M.; Nawa, K.; Jeong, M.; Krämer, S.; Horvatić, M.; Berthier, C.; Takigawa, M.; Akaki, M.; Miyake, A.; et al. One-third magnetization plateau with a preceding novel phase in volborthite. Phys. Rev. Lett. 2015, 114, 227202. [Google Scholar] [CrossRef] [PubMed]
  56. Janson, O.; Furukawa, S.; Momoi, T.; Sindzingre, P.; Richter, J.; Held, K. Magnetic behavior of volborthite Cu3V2O7(OH)2⋅2H2O determined by coupled trimers rather than frustrated chains. Phys. Rev. Lett. 2016, 117, 037206. [Google Scholar] [CrossRef] [PubMed]
  57. Wiessner, R.M.; Fledderjohann, A.; Mütter, K.-H.; Karbach, M. Spontaneous magnetization in spin-ladder systems with competing interactions. Phys. Rev. B 1999, 60, 6545. [Google Scholar] [CrossRef]
  58. Moskin, A.V.; Kozlyakova, E.S.; Shvanskaya, L.V.; Chareev, D.A.; Koo, H.-J.; Whangbo, M.-H.; Vasiliev, A.N. Highly anisotropic 1/3-magnetization plateau in a ferrimagnet Cs2Cu3(SeO3)4·2H2O: Topology of magnetic bonding necessary for magnetization plateau. Dalton Trans. 2023, 52, 118–127. [Google Scholar] [CrossRef]
  59. Starova, L.P.; Filatov, S.K.; Fundamenskii, V.S.; Vergasova, L.P. The crystal structure of fedotovite, K2Cu3O(SO4)3. Mineral. Mag. 1991, 55, 613–616. [Google Scholar] [CrossRef]
  60. Hase, M.; Rule, R.C.; Hester, J.R.; Fernandez-Baca, J.A.; Masuda, T.; Matsuo, Y. A possible magnetic structure of the cluster-based Haldane compound fedotovite K2Cu3O(SO4)3. J. Phys. Soc. Jpn. 2019, 88, 094708. [Google Scholar] [CrossRef]
  61. Miyazaki, Y.; Yang, H.X.; Kajitani, T. Compounds and subsolidus phase relations in the CaO-Co3O4-CuO system. J. Solid State Chem. 2005, 178, 2973–2979. [Google Scholar] [CrossRef]
  62. Kageyama, H.; Yoshimura, K.; Kosuge, K.; Azuma, M.; Takano, M.; Mitamura, H.; Goto, T. Magnetic anisotropy of Cas3Co2O6 with ferromagnetic Ising chains. J. Phys. Soc. Jpn. 1997, 66, 3996–4000. [Google Scholar] [CrossRef]
  63. Whangbo, M.-H.; Dai, D.; Koo, H.-J.; Jobic, S. Investigations of the oxidation states and spin distributions in Ca3Co2O6 and Ca3CoRhO6 by spin-polarized electronic band structure calculations. Solid State Commun. 2003, 125, 413–417. [Google Scholar] [CrossRef]
  64. Munaò, I.; Zvereva, E.A.; Volkova, O.A.; Vasiliev, A.N.; Armstrong, A.R.; Lightfoot, P. NaFe3(HPO3)2((H,F)PO2OH)6: A Potential Cathode Material and a Novel Ferrimagnet. Inorg. Chem. 2016, 55, 2558–2564. [Google Scholar] [CrossRef] [PubMed]
  65. Vasiliev, A.N.; Volkova, O.S.; Zvereva, E.A.; Ovchenkov, E.A.; Munao, I.; Clark, L.; Lightfoot, P.; Vavilova, E.L.; Kamusella, S.; Klauss, H.-H.; et al. 1/3 magnetization plateau and frustrated ferrimagnetism in a sodium iron phosphite. Phys. Rev. B 2016, 93, 134401. [Google Scholar] [CrossRef]
  66. Koo, H.-J.; Kremer, R.K.; Whangbo, M.-H. High-Spin Orbital Interactions Across van der Waals Gaps Controlling the Interlayer Ferromagnetism in van der Waals Ferromagnets. J. Am. Chem. Soc. 2022, 144, 16272–16275. [Google Scholar] [CrossRef]
  67. Zigan, F.; Schuster, H.D. Verfeinerung der Struktur von Azurit, Cu3(OH)2(CO3)2, durch Neutronenbeugung. Z. Kristallogr. 1972, 135, 416–436. [Google Scholar] [CrossRef]
  68. Rule, K.C.; Reehuis, M.; Gibson, M.C.R.; Ouladdiaf, B.; Gutmann, M.J.; Hoffmann, J.-U.; Gerischer, S.; Tennant, D.A.; Süllow, S.; Lang, M. Magnetic and crystal structure of azurite Cu3(CO3)2(OH)2 as determined by neutron diffraction. Phys. Rev. B 2011, 83, 104401. [Google Scholar] [CrossRef]
  69. Kang, J.; Lee, C.; Kremer, R.K.; Whangbo, M.-H. Consequences of the intrachain dimer–monomer spin frustration and the interchain dimer–monomer spin exchange in the diamond-chain compound azurite Cu3(CO3)2(OH)2. J. Phys. Condens. Matter 2009, 21, 392201. [Google Scholar] [CrossRef]
  70. Jeschke, H.; Opahle, I.; Kandpal, H.; Valentí, R.; Das, H.; Saha-Dasgupta, T.; Janson, O.; Rosner, H.; Brühl, A.; Wolf, B.; et al. Multistep Approach to Microscopic Models for Frustrated Quantum Magnets: The Case of the Natural Mineral Azurite. Phys. Rev. Lett. 2011, 106, 217201. [Google Scholar] [CrossRef]
  71. Kikuchi, H.; Fujii, Y.; Chiba, M.; Mitsudo, S.; Idehara, T.; Tonegawa, T.; Okamoto, K.; Sakai, T.; Kuwai, T.; Ohta, H. Experimental Observation of the 1/3 Magnetization Plateau in the Diamond-Chain Compound Cu3(CO3)2(OH)2. Phys. Rev. Lett. 2005, 94, 227201. [Google Scholar] [CrossRef] [PubMed]
  72. Ohta, H.; Okubo, S.; Kamikawa, T.; Kunimoto, T.; Inagaki, Y.; Kikuchi, H.; Saito, T.; Azuma, M.; Takano, M. High Field ESR Study of the S = 1/2 Diamond-Chain Substance Cu3(CO3)2(OH)2 up to the Magnetization Plateau Region. J. Phys. Soc. Jpn. 2003, 72, 2464. [Google Scholar] [CrossRef]
  73. Lee, C.; Kan, E.J.; Xiang, H.J.; Kremer, R.K.; Lee, S.-H.; Hiroi, Z.; Whangbo, M.-H. Spin Reorientation in the Square-Lattice Antiferromagnets RMnAsO (R = Ce, Nd): Density Functional Analysis of the Spin-Exchange Interactions between the Rare-Earth and Transition-Metal Ions. Inorg. Chem. 2012, 51, 6890–6897. [Google Scholar] [CrossRef]
  74. Chubukov, A.V.; Golosov, D.I. Quantum theory of an antiferromagnet on a triangular lattice in a magnetic field. J. Phys. Condens. Matter 1991, 3, 69. [Google Scholar] [CrossRef]
  75. Nishimoto, S.; Shibata, N.; Hotta, C. Controlling frustrated liquids and solids with an applied field in a kagome Heisenberg antiferromagnet. Nat. Commun. 2013, 4, 2287. [Google Scholar] [CrossRef]
  76. Capponi, S.; Derzhko, O.; Honecker, A.; Läuchli, A.M.; Richter, J. Numerical study of magnetization plateaus in the spin-1/2 kagome Heisenberg antiferromagnet. Phys. Rev. B 2013, 88, 144416. [Google Scholar] [CrossRef]
  77. Schnack, J.; Schulenburg, J.; Honecker, A.; Richter, J. Magnon Crystallization in the Kagome Lattice Antiferromagnet. Phys. Rev. Lett. 2020, 125, 117207. [Google Scholar] [CrossRef] [PubMed]
  78. Suetsugu, S.; Asaba, T.; Kasahara, Y.; Kohsaka, Y.; Totsuka, K.; Li, B.; Zhao, Y.; Li, Y.; Tokunaga, M.; Matsuda, Y. Emergent spin-gapped magnetization plateaus in a spin-1/2 perfect kagome antiferromagnet. arXiv 2023, arXiv:2310.10069. [Google Scholar] [CrossRef]
  79. Jeon, S.; Wulferding, D.; Choi, Y.; Lee, S.; Nam, K.; Kim, K.H.; Lee, M.; Jang, T.-H.; Park, J.-H.; Lee, S.; et al. One-ninth magnetization plateau stabilized by spin entanglement in a kagome antiferromagnet. Nat. Phys. 2024, 20, 435–441. [Google Scholar] [CrossRef]
  80. Cabra, D.C.; Grynberg, M.D.; Holdsworth, P.C.W.; Honecker, A.; Pujol, P.; Richter, J.; Schmalfuss, D.; Schulenburg, J. Quantum kagome antiferromagnet in a magnetic field: Low-lying non-magnetic excitations versus valence-bond crystal order. Phys. Rev. B 2005, 71, 144420. [Google Scholar] [CrossRef]
  81. Petrenko, O.A.; Lees, M.R.; Balakrishnan, G.; de Brion, S.; Chouteau, G. Revised magnetic properties of CuFeO2—A case of mistaken identity. J. Phys. Condens. Matter 2005, 17, 2741–2747. [Google Scholar] [CrossRef]
  82. Saito, M.; Watanabe, M.; Kurita, N.; Matsuo, A.; Kindo, K.; Avdeev, M.; Jeschke, H.O.; Tanaka, H. Successive phase transitions and magnetization plateau in the spin-1 triangular-lattice antiferromagnet Ba2LaNiTe2O12 with small easy-axis anisotropy. Phys. Rev. B 2019, 100, 064417. [Google Scholar] [CrossRef]
  83. Goto, M.; Ueda, H.; Michioka, C.; Matsuo, A.; Kindo, K.; Yoshimura, K. Various disordered ground states and 1/3 magnetization-plateau-like behavior in the S=1/2 Ti3+ kagomé lattice antiferromagnets Rb2NaTi3F12, Cs2NaTi3F12, and Cs2KTi3F12. Phys. Rev. B 2016, 94, 104432. [Google Scholar] [CrossRef]
  84. Okuta K, Hara S, Sato H, Narumi Y and Kindo K Observation of 1/3 magnetization-plateau-like anomaly in S=3/2 perfect kagomé lattice antiferromagnet KCr3(OH)6(SO4)2 (Cr-jarosite). J. Phys. Soc. Jpn. 2011, 80, 063703. [CrossRef]
  85. Okuma, R.; Nakamura, D.; Takeyama, S. Magnetization plateau observed by ultrahigh-field Faraday rotation in the kagomé antiferromagnet herbertsmithite. Phys. Rev. B 2020, 102, 104429. [Google Scholar] [CrossRef]
  86. Smirnov, A.I.; Soldatov, T.A.; Petrenko, O.A.; Takata, A.; Kida, T.; Hagiwara, M.; Shapiro, A.Y.; Zhitomirsky, M.E. Order by quenched disorder in the model triangular antiferromagnet RbFe(MoO4)2. Phys. Rev. Lett. 2017, 119, 047204. [Google Scholar] [CrossRef] [PubMed]
  87. Svistov, L.E.; Smirnov, A.I.; Prozorova, L.A.; Petrenko, O.A.; Demianets, L.N.; Shapiro, A.Y. Quasi-two-dimensional antiferromagnet on a triangular lattice RbFe(MoO4)2. Phys. Rev. B 2003, 67, 094434. [Google Scholar] [CrossRef]
  88. Doi, Y.; Hinatsu, Y.; Ohoyama, K. Structural and magnetic properties of pseudo-two-dimensional triangular antiferromagnets Ba3MSb2O9 (M = Mn, Co, and Ni). J. Phys. Condens. Matter 2004, 16, 8923–8935. [Google Scholar] [CrossRef]
  89. Kamiya, Y.; Ge, L.; Hong, T.; Qiu, Y.; Quintero-Castro, D.L.; Lu, Z.; Cao, H.B.; Matsuda, M.; Choi, E.S.; Batista, C.D.; et al. The nature of spin excitations in the one-third magnetization plateau phase of Ba3CoSb2O9. Nat. Commun. 2018, 9, 2666. [Google Scholar] [CrossRef]
  90. Tanaka, H.; Ono, T.; Katori, H. Magnetic phase transition and magnetization plateau in Cs2CuBr4. Progr. Theor. Phys. Suppl. 2022, 145, 101–106. [Google Scholar] [CrossRef]
  91. Whangbo, M.-H.; Koo, H.-J.; Dai, D. Spin exchange interactions and magnetic structures of extended magnetic solids with localized spins: Theoretical descriptions on formal, quantitative and qualitative level. J. Solid State Chem. 2005, 176, 417–481. [Google Scholar] [CrossRef]
  92. dos Santos, A.M.; Brandao, P.; Fitch, A.; Reis, M.S.; Amaral, V.S.; Rocha, J. Synthesis, crystal structure and magnetic characterization of Na2Cu5(Si2O7)2: An inorganic ferrimagnetic chain. J. Solid State Chem. 2007, 180, 16–21. [Google Scholar] [CrossRef]
  93. Vasiliev, A.N. Private communication.
  94. Liu, X.C.; Ouyang, Z.W.; Yue, X.Y.; Jiang, X.; Wang, Z.X.; Wang, J.F.; Xia, Z.C. Structure and 3/7-like magnetization plateau of layered Y2Cu7(TeO3)6Cl6(OH)2 containing diamond chains and trimers. Inorg. Chem. 2019, 58, 10680–10685. [Google Scholar] [CrossRef]
  95. Volkova, L.M.; Marinin, D.V. Antiferromagnetic spin-frustrated layers of corner-sharing Cu4 tetrahedra on the kagomé lattice in volcanic minerals Cu5O2(VO4)2(CuCl), NaCu5O2(SeO3)2Cl3, and K2Cu5Cl8(OH)4·2H2O. J. Phys. Condens. Matter 2018, 30, 425801. [Google Scholar] [CrossRef] [PubMed]
  96. Zhang, S.-Y.; Guo, W.-B.; Yang, M.; Tang, Y.-Y.; Cui, M.-Y.; Wang, N.-N.; He, Z.-Z. A frustrated ferrimagnet Cu5(VO4)2(OH)4 with a 1/5 magnetization plateau on a new spin-lattice of alternating triangular and honeycomb strips. Dalton Trans. 2015, 44, 20562–20567. [Google Scholar] [CrossRef] [PubMed]
  97. Vasiliev, A.N.; Berdonosov, P.S.; Kozlyakova, E.S.; Maximova, O.V.; Murtazoev, A.F.; Dolgikh, V.A.; Lyssenko, K.A.; Pchelkina, Z.V.; Gorbunov, D.I.; Chung, S.H.; et al. Observation of a 1/3 magnetization plateau in Pb2Cu10O4(SeO3)4Cl7 arising from (Cu2+)7 clusters of corner-sharing (Cu2+)4 tetrahedra. Dalton Trans. 2022, 51, 15017–15021. [Google Scholar] [CrossRef] [PubMed]
  98. Tang, Y.; Guo, W.; Zhang, S.; Xiang, H.; Cui, M.; He, Z. Na2Cu7(SeO3)4O2Cl4: A selenite chloride compound with Cu7 units showing spin-frustration and a magnetization plateau. Dalton Trans. 2016, 45, 8324–8326. [Google Scholar] [CrossRef]
  99. Albright, T.A.; Burdett, J.K.; Whangbo, M.-H. Orbital Interactions in Chemistry; Wiley: New York, NY, USA, 1985; Chapter 17. [Google Scholar]
  100. Facheris, L.; Povarov, K.Y.; Nabi, S.  D.; Mazzone, D. G.; Lass, J.; Roessli, B.; Ressouche, E.; Yan, Z.; Gvasaliya, S.; Zheludev, A. Spin Density Wave versus Fractional Magnetization Plateau in a Triangular Antiferromagnet. Phys. Rev. Lett. 2022, 129, 087201. [Google Scholar] [CrossRef]
  101. Cao, J.J.; Ouyang, Z.W.; Liu, X.C.; Xiao, T.T.; Song, Y.R.; Wang, J.F.; Ishii, Y.; Zhou, X.G.; Matsuda, Y.H. Unusual dimerization and magnetization plateaus in S = 1 skew chain Ni2V2O7 observed at 120 T. Phys. Rev. B 2022, 106, 184409. [Google Scholar] [CrossRef]
  102. Guo, K.; Ma, Z.; Liu, H.; Wu, Z.; Wang, J.; Shi, Y.; Li, Y.; Jia, S. 1/3 and other magnetization plateaus in the quasi-one-dimensional Ising magnet TbTi3Bi4 with zigzag spin chain. Phys. Rev. B 2024, 110, 064416. [Google Scholar] [CrossRef]
  103. Chen, Y.; Zhang, Y.; Li, R.; Su, H.; Shan, Z.; Smidman, M.; Yuan, H. Multiple magnetic phases and magnetization plateaus in TbRh6Ge4. Phys. Rev. B 2023, 107, 094414. [Google Scholar] [CrossRef]
  104. Yuan, N.; Elghandour, A.; Hergett, W.; Ohlendorf, R.; Gries, L.; Klingeler, R. 1/3 plateau and 3/5 discontinuity in the magnetization and the magnetic phase diagram of hexagonal GdInO3. Phys. Rev. B 2023, 108, 224403. [Google Scholar] [CrossRef]
  105. Kikuchi, H.; Fujii, Y.; Ishikawa, Y.; Matsuo, A.; Kindo, K.; Widyaiswari, U.; Watanabe, I. Magnetic Plateaux of the Frustrated Magnet, Pseudomalachite. JPS Conf. Proc. 2023, 38, 011127. [Google Scholar]
  106. Yalikun, A.; Wang, Y.; Shi, N.; Chen, Y.; Huang, H.; Koo, H.-J.; Ouyang, Z.; Xia, Z.; Kremer, R.K.; Whangbo, M.-H.; et al. Polar Layered Magnet Ba2Cu3(SeO3)4F2 Composed of Bitriangular Chains: Observation of 1/3-Magnetization Plateau. Chem. Mater. 2024, 36, 7887–7896. [Google Scholar] [CrossRef]
  107. Tanaka, H.; Matsuo, A.; Kindo, K. One-half magnetization plateau in the spin-1/2 fcc lattice antiferromagnet Sr2CoTeO6. Phys. Rev. B 2024, 109, 094414. [Google Scholar] [CrossRef]
  108. Liu, X.; Ouyang, Z.; Jiang, D.; Cao, J.; Xiao, T.; Wang, Z.; Xia, Z.; Tong, W. Syntheses, structure and magnetization plateau of staircase kagome-lattice antiferromagnet Cu3Bi(TeO3)2O2Cl. J. Magn. Magn. Mater. 2023, 565, 170228. [Google Scholar] [CrossRef]
  109. Shangguan, Y.; Bao, S.; Dong, Z.-Y.; Xi, N.; Gao, Y.-P.; Ma, Z.; Wang, W.; Qi, Z.; Zhang, S.; Huang, Z.; et al. A one-third magnetization plateau phase as evidence for the Kitaev interaction in a honeycomb-lattice antiferromagnet. Nat. Phys. 2023, 19, 1883–1889. [Google Scholar] [CrossRef]
  110. Hida, K.; Affleck, I. Quantum vs classical magnetization plateaus of S=1/2 frustrated Heisenberg chains. J. Phys. Soc. Jpn. 2005, 74, 1849–1857. [Google Scholar] [CrossRef]
  111. Sakai, T.; Okamoto, K. Quantum magnetization plateaux of an anisotropic ferrimagnetic spin chain. Phys. Rev. B 2002, 64, 214403. [Google Scholar] [CrossRef]
  112. Momoi, T.; Totsuka, K. Magnetization plateaus as insulator-superfluid transitions in quantum spin systems. Phys. Rev. B 2000, 61, 3231–3234. [Google Scholar] [CrossRef]
  113. Heidrich-Meissner, F.; Sergienko, I.A.; Feiguin, A.E.; Dagotto, E.R. Universal emergence of the one-third plateau in the magnetization process of frustrated quantum spin chains. Phys. Rev. B 2007, 75, 064413. [Google Scholar] [CrossRef]
  114. Takayoshi, S.; Totsuka, K.; Tanaka, A. Symmetry-protected topological order in magnetization plateau states of quantum spin chains. Phys. Rev. B 2015, 91, 155136. [Google Scholar] [CrossRef]
  115. Narumi, Y.; Katsumata, K.; Honda, Z.; Domenge, J.-C.; Sindzingre, P.; Lhuillier, C.; Shimaoka, Y.; Kobayashi, T.C.; Kindo, K. Observation of a transient magnetization plateau in a quantum antiferromagnet on the kagomé lattice. Europhys. Lett. 2004, 65, 705–711. [Google Scholar] [CrossRef]
  116. Hu, H.; Cheng, C.; Xu, Z.; Luo, H.-G.; Chen, S. Topological nature of magnetization plateaus in periodically modulated quantum spin chains. Phys. Rev. B 2014, 90, 035150. [Google Scholar] [CrossRef]
  117. Coletta, T.; Zhitomirsky, M.E.; Mila, F. Quantum stabilization of classically unstable plateau structures. Phys. Rev. B 2013, 87, 060407. [Google Scholar] [CrossRef]
  118. Tandon, K.; Lal, S.; Pati, S.K.; Ramasesha, S.; Sen, D. Magnetization properties of some quantum spin ladders. Phys. Rev. B 1999, 59, 396–410. [Google Scholar] [CrossRef]
  119. Adhikary, M.; Ralko, A.; Kumar, B. Quantum paramagnetism and magnetization plateaus in a kagomé-honeycomb Heisenberg antiferromagnet. Phys. Rev. B 2021, 104, 094416. [Google Scholar] [CrossRef]
  120. Ohanyan, V.; Rojas, O.; Strečka, J.; Bellucci, S. Absence of actual plateaus in zero-temperature magnetization curves of quantum spin clusters and chains. Phys. Rev. B 2015, 92, 214423. [Google Scholar] [CrossRef]
  121. Ananikian, N.S.; Strečka, J.; Hovhannisyan, V. Magnetization plateaus of an exactly solvable spin-1 Ising–Heisenberg diamond chain. Sol. St. Comm. 2014, 194, 48–53. [Google Scholar] [CrossRef]
  122. Totsuka, K. Magnetization process of random quantum spin chains. Phys. Rev. B 2001, 64, 134420. [Google Scholar] [CrossRef]
  123. Pal, S.; Lal, S. Magnetization plateaus of the quantum pyrochlore Heisenberg antiferromagnet. Phys. Rev. B 2019, 100, 104421. [Google Scholar] [CrossRef]
  124. Plat, X.; Alet, F.; Capponi, S.; Totsuka, K. Magnetization plateaus of an easy-axis kagomé antiferromagnet with extended interactions. Phys. Rev. B 2015, 92, 174402. [Google Scholar] [CrossRef]
  125. Yoshida, H.K. Frustrated kagomé antiferromagnets under high magnetic fields. J. Phys. Soc. Jpn. 2022, 91, 101003. [Google Scholar] [CrossRef]
  126. Whangbo, M.-H.; Koo, H.-J.; Kremer, R.K.; Vasilieve, A.N. Magnetization Plateaus by the Field-Induced Partitioning of Spin Lattices. arXiv 2024, arXiv:2409.19529. [Google Scholar]
Figure 1. Idealized magnetization versus magnetic field (M vs. H) curves for two types of magnets exhibiting magnetization plateaus, where the magnetization is given as the fraction f of the saturation magnetization Msat. (ac) M vs. H curves expected for magnets with isotropic magnetism, and (df) those expected for magnets with Ising magnetism. On increasing the magnetic field from 0, a gradual increase in M from 0 to (m/n)Msat precedes the m/n-magnetization plateau in (a); the magnetization plateau occurs immediately at (m/n)Msat in (b), and a 0-magnetization plateau occurs at M = 0 until the field reaches a value from which a gradual increase in M to (m/n)Msat starts in (c). M vs. H curves for Ising magnets exhibit step-like features when the field parallel to the direction of the spin moment is illustrated as in (d,e), but the magnetization does not change with field when the field is perpendicular to the direction of the spin moment as depicted in (f).
Figure 1. Idealized magnetization versus magnetic field (M vs. H) curves for two types of magnets exhibiting magnetization plateaus, where the magnetization is given as the fraction f of the saturation magnetization Msat. (ac) M vs. H curves expected for magnets with isotropic magnetism, and (df) those expected for magnets with Ising magnetism. On increasing the magnetic field from 0, a gradual increase in M from 0 to (m/n)Msat precedes the m/n-magnetization plateau in (a); the magnetization plateau occurs immediately at (m/n)Msat in (b), and a 0-magnetization plateau occurs at M = 0 until the field reaches a value from which a gradual increase in M to (m/n)Msat starts in (c). M vs. H curves for Ising magnets exhibit step-like features when the field parallel to the direction of the spin moment is illustrated as in (d,e), but the magnetization does not change with field when the field is perpendicular to the direction of the spin moment as depicted in (f).
Condensedmatter 09 00045 g001
Figure 2. Conventions and terminologies employed in discussing the magnetization behaviors of various magnets.
Figure 2. Conventions and terminologies employed in discussing the magnetization behaviors of various magnets.
Condensedmatter 09 00045 g002
Figure 3. (a) Expressions of the singlet and triplet states of an isolated spin dimer made up of two S = 1/2 magnetic ions. (b) Single and triplet states of a spin dimer in which the singlet state is lower in energy than the triplet state. (c) Broken-symmetry states of a spin dimer in which the AFM coupling is more stable than the FM coupling.
Figure 3. (a) Expressions of the singlet and triplet states of an isolated spin dimer made up of two S = 1/2 magnetic ions. (b) Single and triplet states of a spin dimer in which the singlet state is lower in energy than the triplet state. (c) Broken-symmetry states of a spin dimer in which the AFM coupling is more stable than the FM coupling.
Condensedmatter 09 00045 g003
Figure 4. Effect of the external magnetic field on the magnetic structure of an AFM chain made up of AFM dimers, where labels 1 and 2 refer to the spin exchanges J1 and J2, respectively. It is assumed that the intradimer exchange J1 is stronger than the interdimer exchange J2. (a) Ground state in the absence of the external magnetic field. (b,c) Breaking of the J2 bonds one at a time with increasing field. (d) State in which all J2 bonds are broken. (e) M = Msat/4 state that results when one out of four J1 bonds is broken. (f) M = Msat/3 state the results when one out of three J1 bonds is broken. (g) M = Msat/2 state that results when one out of two J1 bonds is broken.
Figure 4. Effect of the external magnetic field on the magnetic structure of an AFM chain made up of AFM dimers, where labels 1 and 2 refer to the spin exchanges J1 and J2, respectively. It is assumed that the intradimer exchange J1 is stronger than the interdimer exchange J2. (a) Ground state in the absence of the external magnetic field. (b,c) Breaking of the J2 bonds one at a time with increasing field. (d) State in which all J2 bonds are broken. (e) M = Msat/4 state that results when one out of four J1 bonds is broken. (f) M = Msat/3 state the results when one out of three J1 bonds is broken. (g) M = Msat/2 state that results when one out of two J1 bonds is broken.
Condensedmatter 09 00045 g004
Figure 5. Effect of the external magnetic field on the magnetic structure of an AFM chain made up of linear AFM trimers in a tail-to-tail bridging pattern, where labels 1 and 2 refer to the spin exchanges J1 and J2, respectively. It is assumed that the intra-trimer exchange J1 is stronger than the inter-trimer exchange J2. (a) Ground state in the absence of the external magnetic field. (b) Breaking of the J2 bonds one at a time with increasing field. (c) State in which all J2 bonds are broken, leading to the ferrimagnetic state with M = Msat/3. (d) Breaking of two J1 bonds of a linear trimer, enhancing the magnetization toward M = Msat.
Figure 5. Effect of the external magnetic field on the magnetic structure of an AFM chain made up of linear AFM trimers in a tail-to-tail bridging pattern, where labels 1 and 2 refer to the spin exchanges J1 and J2, respectively. It is assumed that the intra-trimer exchange J1 is stronger than the inter-trimer exchange J2. (a) Ground state in the absence of the external magnetic field. (b) Breaking of the J2 bonds one at a time with increasing field. (c) State in which all J2 bonds are broken, leading to the ferrimagnetic state with M = Msat/3. (d) Breaking of two J1 bonds of a linear trimer, enhancing the magnetization toward M = Msat.
Condensedmatter 09 00045 g005
Figure 6. Effect of the external magnetic field on the magnetic structure of a ferrimagnetic chain made up of linear AFM trimers in a head-to-tail bridging pattern, where labels 1 and 2 refer to the spin exchanges J1 and J2, respectively. It is assumed that the intra-trimer exchange J1 is stronger than the inter-trimer exchange J2. (a) Ferrimagnetic ground state in the absence of the external magnetic field. (b) Breaking one J2 bond leads to one trimer in the (↓↑↓) configuration, which reduces the overall moment of the chain. Hence, breaking a J2 bond will not take place since the moment of a magnet cannot decrease under field. (c,d) Breaking of two J1 bonds of a linear trimer, enhancing the magnetization toward M = Msat.
Figure 6. Effect of the external magnetic field on the magnetic structure of a ferrimagnetic chain made up of linear AFM trimers in a head-to-tail bridging pattern, where labels 1 and 2 refer to the spin exchanges J1 and J2, respectively. It is assumed that the intra-trimer exchange J1 is stronger than the inter-trimer exchange J2. (a) Ferrimagnetic ground state in the absence of the external magnetic field. (b) Breaking one J2 bond leads to one trimer in the (↓↑↓) configuration, which reduces the overall moment of the chain. Hence, breaking a J2 bond will not take place since the moment of a magnet cannot decrease under field. (c,d) Breaking of two J1 bonds of a linear trimer, enhancing the magnetization toward M = Msat.
Condensedmatter 09 00045 g006
Figure 7. (a) Diamond chain defined by two magnetic bonds J1 and J2, where J1 > J2. (b,c) Two possible ways of fragmenting a diamond chain into non-overlapping magnetic triangles. The labels 1 and 2 refer to J1 and J2, respectively.
Figure 7. (a) Diamond chain defined by two magnetic bonds J1 and J2, where J1 > J2. (b,c) Two possible ways of fragmenting a diamond chain into non-overlapping magnetic triangles. The labels 1 and 2 refer to J1 and J2, respectively.
Condensedmatter 09 00045 g007
Figure 8. Spin arrangements possible for a triangular fragment. In each case, the arrows from the left to right indicate the spins at the magnetic sites 1, 2 and 3, respectively.
Figure 8. Spin arrangements possible for a triangular fragment. In each case, the arrows from the left to right indicate the spins at the magnetic sites 1, 2 and 3, respectively.
Condensedmatter 09 00045 g008
Figure 9. Change in the spin moment orientations in Heisenberg antiferromagnets as a function of the magnetic field strength when the magnetic field is (a) parallel and (b) perpendicular to the direction of the spin moment. The thick white arrows represent the magnetic field direction, while thin black arrows represent the directions of the spin moments.
Figure 9. Change in the spin moment orientations in Heisenberg antiferromagnets as a function of the magnetic field strength when the magnetic field is (a) parallel and (b) perpendicular to the direction of the spin moment. The thick white arrows represent the magnetic field direction, while thin black arrows represent the directions of the spin moments.
Condensedmatter 09 00045 g009
Figure 10. Change in the spin moment orientations in Ising antiferromagnets as a function of the magnetic field strength when the magnetic field is (a) parallel and (b) perpendicular to the direction of the spin moments. The thick white arrows represent the magnetic field direction, while thin black arrows represent the directions of the spin moments.
Figure 10. Change in the spin moment orientations in Ising antiferromagnets as a function of the magnetic field strength when the magnetic field is (a) parallel and (b) perpendicular to the direction of the spin moments. The thick white arrows represent the magnetic field direction, while thin black arrows represent the directions of the spin moments.
Condensedmatter 09 00045 g010
Figure 11. (a) Arrangement of the three spin exchange paths J1, J2 and J3 in Fe2O(SeO3)2 forming 2D nets parallel to the ab-plane [17], where the labels 1, 2 and 3 refer to J1, J2 and J3, respectively. (b) Temperature dependence of the magnetic susceptibility measured under the magnetic field of 0.1 T along the a, b, and c axes in Fe2O(SeO3)2 [17]. The inset shows a zoomed-in view around the magnetic transition. (c) Magnetization curve at 2 K. The inset shows a zoomed-in view for the magnetization at low fields [17].
Figure 11. (a) Arrangement of the three spin exchange paths J1, J2 and J3 in Fe2O(SeO3)2 forming 2D nets parallel to the ab-plane [17], where the labels 1, 2 and 3 refer to J1, J2 and J3, respectively. (b) Temperature dependence of the magnetic susceptibility measured under the magnetic field of 0.1 T along the a, b, and c axes in Fe2O(SeO3)2 [17]. The inset shows a zoomed-in view around the magnetic transition. (c) Magnetization curve at 2 K. The inset shows a zoomed-in view for the magnetization at low fields [17].
Condensedmatter 09 00045 g011
Figure 12. (a) Magnetic susceptibility of Cu3Y(SeO3)2O2Cl at various probe magnetic fields between 1 and 9 T [18] measured with field μ0H||b. The inset shows the susceptibility measured at 0.1 T and the Curie constant C refers to (χχ0)(TΘ). The dashed line (see the inset) represents the Curie–Weiss law. (Reproduced with permission from [18]). (b) Metamagnetic phase transition in Cu3Eu(SeO3)2O2Cl and Cu3Lu(SeO3)2O2Cl under field μ0H||b. Inset: Schematic representations of the Cu2+ spin moments in weak and strong magnetic fields [19]. (Reproduced with permission from [19]).
Figure 12. (a) Magnetic susceptibility of Cu3Y(SeO3)2O2Cl at various probe magnetic fields between 1 and 9 T [18] measured with field μ0H||b. The inset shows the susceptibility measured at 0.1 T and the Curie constant C refers to (χχ0)(TΘ). The dashed line (see the inset) represents the Curie–Weiss law. (Reproduced with permission from [18]). (b) Metamagnetic phase transition in Cu3Eu(SeO3)2O2Cl and Cu3Lu(SeO3)2O2Cl under field μ0H||b. Inset: Schematic representations of the Cu2+ spin moments in weak and strong magnetic fields [19]. (Reproduced with permission from [19]).
Condensedmatter 09 00045 g012
Figure 13. (a) A projection view of one CuBO3 layer of SrCu2(BO3)2 along the c direction, where the blue, green and red circles represent the Cu, B and O atoms, respectively. (b) The spin lattice of a CuBO3 layer showing an orthogonal arrangement of (Cu2+)2 dimers, where the labels 1 and 2 refer to the spin exchanges J1 and J2, respectively. (c) Arrangement of the intradimer bonds J1 in the Shastry–Sutherland spin lattice leading to a J2 bond and a broken J2 bond between adjacent dimers.
Figure 13. (a) A projection view of one CuBO3 layer of SrCu2(BO3)2 along the c direction, where the blue, green and red circles represent the Cu, B and O atoms, respectively. (b) The spin lattice of a CuBO3 layer showing an orthogonal arrangement of (Cu2+)2 dimers, where the labels 1 and 2 refer to the spin exchanges J1 and J2, respectively. (c) Arrangement of the intradimer bonds J1 in the Shastry–Sutherland spin lattice leading to a J2 bond and a broken J2 bond between adjacent dimers.
Condensedmatter 09 00045 g013
Figure 14. (a) Temperature dependence of the magnetic susceptibility measured for SrCu2(BO3)2 powder. The solid and dashed lines show the theoretical approximations. The inset enlarges low-temperature data [30]. (Reproduced with permission from [30]). (b) Field dependence of the magnetization measured for SrCu2(BO3)2 single crystal. Adapted from Matsuda et al. [12]. (Adapted with permission from [12]).
Figure 14. (a) Temperature dependence of the magnetic susceptibility measured for SrCu2(BO3)2 powder. The solid and dashed lines show the theoretical approximations. The inset enlarges low-temperature data [30]. (Reproduced with permission from [30]). (b) Field dependence of the magnetization measured for SrCu2(BO3)2 single crystal. Adapted from Matsuda et al. [12]. (Adapted with permission from [12]).
Condensedmatter 09 00045 g014
Figure 15. (a) Cr4O16 cluster in CdCr2O4 made up of three CrO6 octahedra by edge-sharing, where the dark blue and red spheres represent the Cr and O atoms, respectively. (b) Pyrochlore spin lattice of Cr3+ ions in which each (Cr3+)4 tetrahedron is corner-shared with four (Cr3+)4 tetrahedra in a tetrahedral manner. (c) Generating isolated (Cr3+)4 tetrahedra that fill the pyrochlore lattice without overlapping between them. In (c), the shaded (Cr3+)4 tetrahedra represent those that become isolated when the inter-tetrahedra interactions are neglected.
Figure 15. (a) Cr4O16 cluster in CdCr2O4 made up of three CrO6 octahedra by edge-sharing, where the dark blue and red spheres represent the Cr and O atoms, respectively. (b) Pyrochlore spin lattice of Cr3+ ions in which each (Cr3+)4 tetrahedron is corner-shared with four (Cr3+)4 tetrahedra in a tetrahedral manner. (c) Generating isolated (Cr3+)4 tetrahedra that fill the pyrochlore lattice without overlapping between them. In (c), the shaded (Cr3+)4 tetrahedra represent those that become isolated when the inter-tetrahedra interactions are neglected.
Condensedmatter 09 00045 g015
Figure 16. (a) Field dependence of the magnetization M(H) observed for a CdCr2O4 single crystal at various temperatures [31]. (Reproduced with permission from [31]). (b) Three spin configurations of a (Cr3+)4 tetramer, where the tetramer is shown in terms of two dimers.
Figure 16. (a) Field dependence of the magnetization M(H) observed for a CdCr2O4 single crystal at various temperatures [31]. (Reproduced with permission from [31]). (b) Three spin configurations of a (Cr3+)4 tetramer, where the tetramer is shown in terms of two dimers.
Condensedmatter 09 00045 g016
Figure 17. (a) Zigzag ribbon chains of CoGeO3 parallel to the bc-plane. (b) Arrangement of Co12+ and Co22+ ions in the ribbon chains with the four spin exchange paths J1J4 represented by the labels 1–4, respectively. (c) Values of J1J4 (in K) determined using DFT+U calculations.
Figure 17. (a) Zigzag ribbon chains of CoGeO3 parallel to the bc-plane. (b) Arrangement of Co12+ and Co22+ ions in the ribbon chains with the four spin exchange paths J1J4 represented by the labels 1–4, respectively. (c) Values of J1J4 (in K) determined using DFT+U calculations.
Condensedmatter 09 00045 g017
Figure 18. (a) Magnetization M(H) of CoGeO3 for H||c. (Data obtained by ramping up to 9 T and down to 0 T after virgin zero-field cooling from 200 K down). Pronounced 1/3-magnetization plateaus can be seen. The magnetization for Hc obtained at 5 K (inset) is linear in H and unsaturated up to 9 T [34]. (b) Spin arrangement of the AFM ground state, where the intrachain spin arrangement is dictated by two strong AFM spin exchanges J2 and J4. (c) Spin arrangement of the ferrimagnetic state representing the 1/3-magnetization plateau. In (b,c), the “six-spin” units are shaded.
Figure 18. (a) Magnetization M(H) of CoGeO3 for H||c. (Data obtained by ramping up to 9 T and down to 0 T after virgin zero-field cooling from 200 K down). Pronounced 1/3-magnetization plateaus can be seen. The magnetization for Hc obtained at 5 K (inset) is linear in H and unsaturated up to 9 T [34]. (b) Spin arrangement of the AFM ground state, where the intrachain spin arrangement is dictated by two strong AFM spin exchanges J2 and J4. (c) Spin arrangement of the ferrimagnetic state representing the 1/3-magnetization plateau. In (b,c), the “six-spin” units are shaded.
Condensedmatter 09 00045 g018
Figure 19. (a) Split down-spin t2g states of an ideally axially compressed CoO6 octahedron. With one electron to fill the degenerate (xz, yz) state, the smallest difference in the Lz values of the highest-occupied and the lowest-unoccupied d-states of such an octahedron is zero, namely, |ΔLz| = 0. (b,c) Magnetic fields H||c and Hc acting on the (4↑2↓) ferrimagnetic fragment with spins oriented along the c-direction.
Figure 19. (a) Split down-spin t2g states of an ideally axially compressed CoO6 octahedron. With one electron to fill the degenerate (xz, yz) state, the smallest difference in the Lz values of the highest-occupied and the lowest-unoccupied d-states of such an octahedron is zero, namely, |ΔLz| = 0. (b,c) Magnetic fields H||c and Hc acting on the (4↑2↓) ferrimagnetic fragment with spins oriented along the c-direction.
Condensedmatter 09 00045 g019
Figure 20. (a) Arrangement of the MnO4 tetrahedra in Ba3Mn2O8 around one dimer unit (MnO4)2, indicated by a green ellipse. (b) Simplified view of the arrangement of six (Mn5+)2 dimer ions surrounding one (Mn5+)2 dimer ion indicated by a green ellipse. (c) Definitions of the intradimer exchange J0 and the interdimer exchange J1.
Figure 20. (a) Arrangement of the MnO4 tetrahedra in Ba3Mn2O8 around one dimer unit (MnO4)2, indicated by a green ellipse. (b) Simplified view of the arrangement of six (Mn5+)2 dimer ions surrounding one (Mn5+)2 dimer ion indicated by a green ellipse. (c) Definitions of the intradimer exchange J0 and the interdimer exchange J1.
Condensedmatter 09 00045 g020
Figure 21. (a) Temperature dependence of magnetic susceptibility in Ba3Mn2O8 powder at 0.1 T. (b) Field dependence of magnetization at 1.4 K [38]. (Reproduced with permission from [38]).
Figure 21. (a) Temperature dependence of magnetic susceptibility in Ba3Mn2O8 powder at 0.1 T. (b) Field dependence of magnetization at 1.4 K [38]. (Reproduced with permission from [38]).
Condensedmatter 09 00045 g021
Figure 22. (a) Two arrangements of J1 bonds around a J0 bond in Ba3Mn2O8. (b) Two arrangements of J1 bonds around a broken J0 bond in Ba3Mn2O8.
Figure 22. (a) Two arrangements of J1 bonds around a J0 bond in Ba3Mn2O8. (b) Two arrangements of J1 bonds around a broken J0 bond in Ba3Mn2O8.
Condensedmatter 09 00045 g022
Figure 23. (a) Planar Cu2Cl62− anion of ACuCl3. (b) Singlet dimer with ΔE > 0. (c) Triplet dimer with ΔE < 0. (d) Splitting of a singlet spin dimer as a function of μ0Hc. (e) Splitting of a triplet spin dimer as a function of μ0Hc. (f) Singlet to triplet excitation energies of NH4CuCl3 measured by neutron scattering experiments [42]. (Reproduced with permission from [42]).
Figure 23. (a) Planar Cu2Cl62− anion of ACuCl3. (b) Singlet dimer with ΔE > 0. (c) Triplet dimer with ΔE < 0. (d) Splitting of a singlet spin dimer as a function of μ0Hc. (e) Splitting of a triplet spin dimer as a function of μ0Hc. (f) Singlet to triplet excitation energies of NH4CuCl3 measured by neutron scattering experiments [42]. (Reproduced with permission from [42]).
Condensedmatter 09 00045 g023
Figure 24. Field dependence of the magnetization M in (a) KCuCl3, (b) TlCuCl3 and (c) NH4CuCl3 [45]. (Reproduced with permission from [45]).
Figure 24. Field dependence of the magnetization M in (a) KCuCl3, (b) TlCuCl3 and (c) NH4CuCl3 [45]. (Reproduced with permission from [45]).
Condensedmatter 09 00045 g024
Figure 25. (a) A stack of Cu2Cl62− anions along the a-direction. (b) Two adjacent stacks of Cu2Cl62− anions viewed along the b-direction. (c) Arrangement of the Cu2Cl62− anions and the A+ cations in a layer parallel to the ad-plane. (d) Spin exchanges calculated for KCuCl3 and TlCuCl3 using DFT+U calculations (see text).
Figure 25. (a) A stack of Cu2Cl62− anions along the a-direction. (b) Two adjacent stacks of Cu2Cl62− anions viewed along the b-direction. (c) Arrangement of the Cu2Cl62− anions and the A+ cations in a layer parallel to the ad-plane. (d) Spin exchanges calculated for KCuCl3 and TlCuCl3 using DFT+U calculations (see text).
Condensedmatter 09 00045 g025
Figure 26. (a,b) Arrangements of the cations A+ around the (CuCl4)2 dimers constituting the interdimer spin exchange paths J2 and Ja′. (c) The orbitals of the cations K+ and Tl+ that can interact with the d-states, (+) and (−), states of the (CuCl4)2 dimers. (d) A schematic diagram showing how the d-states, (+) and (−), of the (CuCl4)2 dimer are affected by the K 4s orbital in KCuCl3 (left), and by the Tl 6s and Tl 6p orbitals in TlCuCl3 (right).
Figure 26. (a,b) Arrangements of the cations A+ around the (CuCl4)2 dimers constituting the interdimer spin exchange paths J2 and Ja′. (c) The orbitals of the cations K+ and Tl+ that can interact with the d-states, (+) and (−), states of the (CuCl4)2 dimers. (d) A schematic diagram showing how the d-states, (+) and (−), of the (CuCl4)2 dimer are affected by the K 4s orbital in KCuCl3 (left), and by the Tl 6s and Tl 6p orbitals in TlCuCl3 (right).
Condensedmatter 09 00045 g026
Figure 27. (a) The YY, NY and NN arrangements of two NH4+ cations around the Cl…Cl contact of each J2 exchange path in NH4CuCl3. (b) Short N-H…Cl contacts to the mid Cl atoms of the Cu2Cl62− anion in the YY, NY and NN arrangements of two NH4+ cations. (c) The relative energies ΔE (in meV/f.u). of NH4CuCl3 with the YY, NY and NN arrangements, and values of their intradimer exchanges (in K), where the labels 1–4, a and a′ refer to the spin exchanges J1J4, Ja and Ja′, respectively. The J1 values in the parentheses were obtained using the optimized NH4CuCl3 structures.
Figure 27. (a) The YY, NY and NN arrangements of two NH4+ cations around the Cl…Cl contact of each J2 exchange path in NH4CuCl3. (b) Short N-H…Cl contacts to the mid Cl atoms of the Cu2Cl62− anion in the YY, NY and NN arrangements of two NH4+ cations. (c) The relative energies ΔE (in meV/f.u). of NH4CuCl3 with the YY, NY and NN arrangements, and values of their intradimer exchanges (in K), where the labels 1–4, a and a′ refer to the spin exchanges J1J4, Ja and Ja′, respectively. The J1 values in the parentheses were obtained using the optimized NH4CuCl3 structures.
Condensedmatter 09 00045 g027
Figure 28. (a) The in-phase, [+], and out-of-phase, [−], combinations of the two x2-y2 magnetic orbitals describing the intradimer exchange path J1 of a Cu2Cl62− anion. The two magnetic orbitals are given in red and blue colors for the ease of distinction. The two p-orbitals at each bridging Cl atoms (encircled by a dashed green circle) in the [+] state become a p orbital perpendicular to the Cu…Cu axis, and those in the [−] state become a p|| orbital parallel to the Cu…Cu axis. (b) Sigma bonding (p-σ*N-H) interaction(s) in the YY and NY structures of NH4CuCl3 that the σ*NH orbital of NH4+ makes with the p orbital(s) in the [+] d-state of Cu2Cl62− ion. (c) Lowering of the [+] level by the (p-σ*N-H) interaction(s) in the YY and NY structures of NH4CuCl3.
Figure 28. (a) The in-phase, [+], and out-of-phase, [−], combinations of the two x2-y2 magnetic orbitals describing the intradimer exchange path J1 of a Cu2Cl62− anion. The two magnetic orbitals are given in red and blue colors for the ease of distinction. The two p-orbitals at each bridging Cl atoms (encircled by a dashed green circle) in the [+] state become a p orbital perpendicular to the Cu…Cu axis, and those in the [−] state become a p|| orbital parallel to the Cu…Cu axis. (b) Sigma bonding (p-σ*N-H) interaction(s) in the YY and NY structures of NH4CuCl3 that the σ*NH orbital of NH4+ makes with the p orbital(s) in the [+] d-state of Cu2Cl62− ion. (c) Lowering of the [+] level by the (p-σ*N-H) interaction(s) in the YY and NY structures of NH4CuCl3.
Condensedmatter 09 00045 g028
Figure 29. (a) Linear Mn1-Mn2-Mn1 trimer in γ-Mn3(PO4)2, which results when a Mn2O6 octahedron corner-shares with two Mn1O5 trigonal bipyramids. (b) Two Mn1-Mn2-Mn1 trimers edge-sharing in a head-to-tail fashion. (c) Two Mn1-Mn2-Mn1 trimers edge-sharing in a tail-to-tail fashion. The labels 1, 2 and 3 refer to the spin exchange paths J1, J2 and J3, respectively. (d) Field dependence of the magnetization in α-, β’- and γ-phases of Mn3(PO4)2 at 2 K [48]. (Reproduced with permission from [48]).
Figure 29. (a) Linear Mn1-Mn2-Mn1 trimer in γ-Mn3(PO4)2, which results when a Mn2O6 octahedron corner-shares with two Mn1O5 trigonal bipyramids. (b) Two Mn1-Mn2-Mn1 trimers edge-sharing in a head-to-tail fashion. (c) Two Mn1-Mn2-Mn1 trimers edge-sharing in a tail-to-tail fashion. The labels 1, 2 and 3 refer to the spin exchange paths J1, J2 and J3, respectively. (d) Field dependence of the magnetization in α-, β’- and γ-phases of Mn3(PO4)2 at 2 K [48]. (Reproduced with permission from [48]).
Condensedmatter 09 00045 g029
Figure 30. (a) Layer of Mn1-Mn2-Mn1 linear trimers in γ-Mn3(PO4), parallel to the bc-plane, formed by a head-to-tail bridge. (b) Tail-to-tail bridging between the trimers lying in adjacent layers. (c) Ferrimagnetic state of a layer defined by the spin exchange J3 and J2 due to the head-to-tail coupling between the linear trimers. A green ellipse indicates a trimer that will undergo a field-induced J3 bond breaking. In (ac), the labels 1–3 refer to the spin exchanges J1J3, respectively. (d) Breaking of an inter-trimer bond J2 as a consequence of breaking the two J3 bonds of a linear trimer.
Figure 30. (a) Layer of Mn1-Mn2-Mn1 linear trimers in γ-Mn3(PO4), parallel to the bc-plane, formed by a head-to-tail bridge. (b) Tail-to-tail bridging between the trimers lying in adjacent layers. (c) Ferrimagnetic state of a layer defined by the spin exchange J3 and J2 due to the head-to-tail coupling between the linear trimers. A green ellipse indicates a trimer that will undergo a field-induced J3 bond breaking. In (ac), the labels 1–3 refer to the spin exchanges J1J3, respectively. (d) Breaking of an inter-trimer bond J2 as a consequence of breaking the two J3 bonds of a linear trimer.
Condensedmatter 09 00045 g030
Figure 31. (a) Chain of edge-sharing Cu1O5 trigonal bipyramids and Cu2O6 octahedra in Cu3(P2O6OH)2. (b) A spin lattice composed of ferrimagnetic linear Cu1-Cu2-Cu1 trimers making chains by a tail-to-tail coupling. (c) A spin lattice of ferrimagnetic bent Cu1-Cu1-Cu2 trimers (e.g., those enclosed in red rectangles) making chains by a tail-to-tail coupling, and such chains make a 2D net by a tail-to-tail coupling. The labels 1, 2, 3 and 6 refer to the spin exchanges J1, J2, J3 and J6, respectively.
Figure 31. (a) Chain of edge-sharing Cu1O5 trigonal bipyramids and Cu2O6 octahedra in Cu3(P2O6OH)2. (b) A spin lattice composed of ferrimagnetic linear Cu1-Cu2-Cu1 trimers making chains by a tail-to-tail coupling. (c) A spin lattice of ferrimagnetic bent Cu1-Cu1-Cu2 trimers (e.g., those enclosed in red rectangles) making chains by a tail-to-tail coupling, and such chains make a 2D net by a tail-to-tail coupling. The labels 1, 2, 3 and 6 refer to the spin exchanges J1, J2, J3 and J6, respectively.
Condensedmatter 09 00045 g031
Figure 32. (a) Field dependence of magnetization of Cu3(P2O6OH)2 at 1.6 K and its quantum Monte Carlo (QMC) simulation [50]. (Reproduced with permission from [50]). (b) AFM arrangement of ferrimagnetic bent trimers via the tail-to-tail coupling along the J3 and J6 exchange paths. The down-spin at each site encircled with a green circle becomes up-spin to increase the moment under magnetization. (c) A ferrimagnetic state resulting from the down-spin to up-spin conversion at each circled down-spin site in (b). (d) A ferrimagnetic state, equivalent to the one shown in (c), composed of ferrimagnetic bent trimers.
Figure 32. (a) Field dependence of magnetization of Cu3(P2O6OH)2 at 1.6 K and its quantum Monte Carlo (QMC) simulation [50]. (Reproduced with permission from [50]). (b) AFM arrangement of ferrimagnetic bent trimers via the tail-to-tail coupling along the J3 and J6 exchange paths. The down-spin at each site encircled with a green circle becomes up-spin to increase the moment under magnetization. (c) A ferrimagnetic state resulting from the down-spin to up-spin conversion at each circled down-spin site in (b). (d) A ferrimagnetic state, equivalent to the one shown in (c), composed of ferrimagnetic bent trimers.
Condensedmatter 09 00045 g032
Figure 33. (a) Arrangement of the Cu2+ ions in a Cu3O6(OH)2 layer of volborthite. (b) Arrangement of the CuO4 square planes containing the x2-y2 magnetic orbitals in a Cu3O6(OH)2 layer of volborthite. (c) Arrangement of Cu2-Cu1-Cu2 linear trimers in a Cu3O6(OH)2 layer of volborthite. (d) AFM state of a two-leg spin ladder with rungs of ferrimagnetic linear trimers defined by J2 and legs defined by J4. In (c,d), the labels 2 and 4 refer to J2 and J4, respectively. (e) Effective S = 1/2 AUH chain representing the two-leg spin ladder of (d) at low temperature, where thermal excitations within each rung are absent. (f) Ferrimagnetic state of a two-leg spin ladder with rungs of ferrimagnetic linear trimers.
Figure 33. (a) Arrangement of the Cu2+ ions in a Cu3O6(OH)2 layer of volborthite. (b) Arrangement of the CuO4 square planes containing the x2-y2 magnetic orbitals in a Cu3O6(OH)2 layer of volborthite. (c) Arrangement of Cu2-Cu1-Cu2 linear trimers in a Cu3O6(OH)2 layer of volborthite. (d) AFM state of a two-leg spin ladder with rungs of ferrimagnetic linear trimers defined by J2 and legs defined by J4. In (c,d), the labels 2 and 4 refer to J2 and J4, respectively. (e) Effective S = 1/2 AUH chain representing the two-leg spin ladder of (d) at low temperature, where thermal excitations within each rung are absent. (f) Ferrimagnetic state of a two-leg spin ladder with rungs of ferrimagnetic linear trimers.
Condensedmatter 09 00045 g033
Figure 34. (a) Magnetic susceptibility of volborthite for one formula unit (i.e., comprising three Cu atoms) of volborthite (black circles) with probe field applied along the crystallographic b axis fitted (for T < 75 K) to the theoretical prediction for an S = 1/2 AUH chain (see the solid blue curve) [53]. The difference between the two is displayed as a solid green line. The experimental susceptibilities for 75 K ≤ T ≤ 320 K are well fitted by the susceptibility of a linear spin S = 1/2 trimer with a spin exchange of 197 K (red dotted curve). (b) Field dependence of the magnetization (per Cu) measured for single crystal and polycrystalline samples of Cu3V2O7(OH)2·2H2O at 1.4 K [55]. (Reproduced with permission from [55]). (c) Field dependence of the magnetization (per three Cu) measured for volborthite at 1.4 K (taken from Ishikawa et al. [55]) compared with quantum Monte Carlo calculations for an S = 1/2 AUH chain with JC = 27.5 K (solid red line) [53].
Figure 34. (a) Magnetic susceptibility of volborthite for one formula unit (i.e., comprising three Cu atoms) of volborthite (black circles) with probe field applied along the crystallographic b axis fitted (for T < 75 K) to the theoretical prediction for an S = 1/2 AUH chain (see the solid blue curve) [53]. The difference between the two is displayed as a solid green line. The experimental susceptibilities for 75 K ≤ T ≤ 320 K are well fitted by the susceptibility of a linear spin S = 1/2 trimer with a spin exchange of 197 K (red dotted curve). (b) Field dependence of the magnetization (per Cu) measured for single crystal and polycrystalline samples of Cu3V2O7(OH)2·2H2O at 1.4 K [55]. (Reproduced with permission from [55]). (c) Field dependence of the magnetization (per three Cu) measured for volborthite at 1.4 K (taken from Ishikawa et al. [55]) compared with quantum Monte Carlo calculations for an S = 1/2 AUH chain with JC = 27.5 K (solid red line) [53].
Condensedmatter 09 00045 g034
Figure 35. (a,b) Four Cu2O4 square planes sharing their oxygen corners with a Cu1O4 square plane in Cs2Cu3(SeO3)4(H2O)2 viewed approximately along the b-direction in (a) and along the c-direction in (b). (c) Cu1(Cu2)4 tetrahedron associated with the five CuO4 planes in (a). (d) Chair-form hexagonal ring made up of Cu1(Cu2)4 tetrahedra by sharing their Cu2 corners. (e) Head-to-tail coupling of the bent ferrimagnetic Cu2-Cu1-Cu2 units with (↑↓↑) spin configuration leading to the ferrimagnetic state of Cs2Cu3(SeO3)4(H2O)2.
Figure 35. (a,b) Four Cu2O4 square planes sharing their oxygen corners with a Cu1O4 square plane in Cs2Cu3(SeO3)4(H2O)2 viewed approximately along the b-direction in (a) and along the c-direction in (b). (c) Cu1(Cu2)4 tetrahedron associated with the five CuO4 planes in (a). (d) Chair-form hexagonal ring made up of Cu1(Cu2)4 tetrahedra by sharing their Cu2 corners. (e) Head-to-tail coupling of the bent ferrimagnetic Cu2-Cu1-Cu2 units with (↑↓↑) spin configuration leading to the ferrimagnetic state of Cs2Cu3(SeO3)4(H2O)2.
Condensedmatter 09 00045 g035
Figure 36. (a) Temperature dependence of the magnetic susceptibility χ = M/H in Cs2Cu3(SeO3)4·2H2O for both H||c and Hc taken at μ0H = 0.1 T. Left inset: Temperature dependence of magnetic susceptibility corrected for demagnetization effects. Right inset: Temperature dependence of the inverse magnetic susceptibility for H||c, where the solid line represents the Néel law. (b) Anisotropic 1/3-magnetization plateau in Cs2Cu3(SeO3)4(H2O)2 [58]. (Reproduced with permission from [58]).
Figure 36. (a) Temperature dependence of the magnetic susceptibility χ = M/H in Cs2Cu3(SeO3)4·2H2O for both H||c and Hc taken at μ0H = 0.1 T. Left inset: Temperature dependence of magnetic susceptibility corrected for demagnetization effects. Right inset: Temperature dependence of the inverse magnetic susceptibility for H||c, where the solid line represents the Néel law. (b) Anisotropic 1/3-magnetization plateau in Cs2Cu3(SeO3)4(H2O)2 [58]. (Reproduced with permission from [58]).
Condensedmatter 09 00045 g036
Figure 37. (a) Anisotropic g-factors of the Cu2+ ion at a square planar coordination site. (b,c) The moments associated with the (↑↓↑) spin arrangement of a bent ferrimagnetic Cu2-Cu1-Cu2 fragment. The magnetic field is applied along the ||c direction in (b), and along the ⊥c direction in (c). The filled and unfilled circles in (b,c) represent the Cu12+ and Cu22+ ions, respectively [58].
Figure 37. (a) Anisotropic g-factors of the Cu2+ ion at a square planar coordination site. (b,c) The moments associated with the (↑↓↑) spin arrangement of a bent ferrimagnetic Cu2-Cu1-Cu2 fragment. The magnetic field is applied along the ||c direction in (b), and along the ⊥c direction in (c). The filled and unfilled circles in (b,c) represent the Cu12+ and Cu22+ ions, respectively [58].
Condensedmatter 09 00045 g037
Figure 38. (a) The structure of a Cu6 cluster present in K2Cu3O(SO4)3, which is constructed from distorted Cu1O4, Cu2O4 and Cu3O4 square planes. (b) View of a Cu6 cluster resulting from two Cu4 tetrahedra by edge-sharing. (c) Definitions of the eight spin exchanges J1J8. The labels 1–8 refer to J1J8, respectively. (d) Values of the J1J8 determined via DFT+U calculations.
Figure 38. (a) The structure of a Cu6 cluster present in K2Cu3O(SO4)3, which is constructed from distorted Cu1O4, Cu2O4 and Cu3O4 square planes. (b) View of a Cu6 cluster resulting from two Cu4 tetrahedra by edge-sharing. (c) Definitions of the eight spin exchanges J1J8. The labels 1–8 refer to J1J8, respectively. (d) Values of the J1J8 determined via DFT+U calculations.
Condensedmatter 09 00045 g038
Figure 39. (a) Temperature dependence of the magnetic susceptibility χbulk (filled red circles) of K2Cu3O(SO4)3 measured at 0.1 T, obtained by subtracting Pascal’s diamagnetic contribution χdia and an estimated contribution of impurity χimp (gray solid line) from the experimental data χobs (filled green circles). (b) High-field magnetization at 4.2 K (pink solid line) and 20 K (black solid line). The blue dashed line denotes a theoretical magnetization curve [60]. (Reproduced with permission from [60]).
Figure 39. (a) Temperature dependence of the magnetic susceptibility χbulk (filled red circles) of K2Cu3O(SO4)3 measured at 0.1 T, obtained by subtracting Pascal’s diamagnetic contribution χdia and an estimated contribution of impurity χimp (gray solid line) from the experimental data χobs (filled green circles). (b) High-field magnetization at 4.2 K (pink solid line) and 20 K (black solid line). The blue dashed line denotes a theoretical magnetization curve [60]. (Reproduced with permission from [60]).
Condensedmatter 09 00045 g039
Figure 40. (a) Ferrimagnetic state of a Cu6 cluster in K2Cu3O(SO4)3. (b) AFM arrangement of ferrimagnetic Cu6 clusters. (c) Ferrimagnetic arrangement of ferrimagnetic Cu6 clusters.
Figure 40. (a) Ferrimagnetic state of a Cu6 cluster in K2Cu3O(SO4)3. (b) AFM arrangement of ferrimagnetic Cu6 clusters. (c) Ferrimagnetic arrangement of ferrimagnetic Cu6 clusters.
Condensedmatter 09 00045 g040
Figure 41. (a) An isolated Co2O6 chain of Ca3Co2O6, in which Co1O6 octahedra alternate with Co2O6 trigonal prisms by sharing their triangular faces. (b) Trigonal arrangement of the Co2O6 chains in Ca3Co2O6, where each chain is represented by showing only the Co atoms. (c) The 3z2 − r2 orbitals of the Co1 and Co2 atoms in each Co2O6 chain.
Figure 41. (a) An isolated Co2O6 chain of Ca3Co2O6, in which Co1O6 octahedra alternate with Co2O6 trigonal prisms by sharing their triangular faces. (b) Trigonal arrangement of the Co2O6 chains in Ca3Co2O6, where each chain is represented by showing only the Co atoms. (c) The 3z2 − r2 orbitals of the Co1 and Co2 atoms in each Co2O6 chain.
Condensedmatter 09 00045 g041
Figure 42. (a) Left: Down-spin split d-states of a Co2+ ion at the Co2O6 trigonal prism with (3z2 − r2)1(xy, x2 − y2)0 configuration in the absence of SOC in Ca3Co2O6. Right: Effect of SOC on the down-spin split d-states. In terms of the spherical harmonics, the angular parts of the xy and x2 − y2 states are given as linear combinations of |2, 2〉 and |2, −2〉, and that of 3z2 − r2 as |2, 0〉 [3]. (b) Field dependence of the magnetization measured for a single crystal sample (M|| and M) and a powder sample (MΓ) of Ca3Co2O6 at 12 and 35 K [62]. (Reproduced with permission from [62]).
Figure 42. (a) Left: Down-spin split d-states of a Co2+ ion at the Co2O6 trigonal prism with (3z2 − r2)1(xy, x2 − y2)0 configuration in the absence of SOC in Ca3Co2O6. Right: Effect of SOC on the down-spin split d-states. In terms of the spherical harmonics, the angular parts of the xy and x2 − y2 states are given as linear combinations of |2, 2〉 and |2, −2〉, and that of 3z2 − r2 as |2, 0〉 [3]. (b) Field dependence of the magnetization measured for a single crystal sample (M|| and M) and a powder sample (MΓ) of Ca3Co2O6 at 12 and 35 K [62]. (Reproduced with permission from [62]).
Condensedmatter 09 00045 g042
Figure 43. (a) Geometrical arrangements associated with the spin exchange paths, J2, J3, J6 and J4 in NaFe3(HPO3)2(H2PO3)6. All these exchanges are of the Fe-O…O-Fe type with the O…O contact making a O…P5+…O bridge. (b) 2D spin lattice of NaFe3(HPO3)2(H2PO3)6 made up of the spin exchanges J2, J3, J6 and J4, which are indicated by the labels 2, 3, 6 and 4, respectively.
Figure 43. (a) Geometrical arrangements associated with the spin exchange paths, J2, J3, J6 and J4 in NaFe3(HPO3)2(H2PO3)6. All these exchanges are of the Fe-O…O-Fe type with the O…O contact making a O…P5+…O bridge. (b) 2D spin lattice of NaFe3(HPO3)2(H2PO3)6 made up of the spin exchanges J2, J3, J6 and J4, which are indicated by the labels 2, 3, 6 and 4, respectively.
Condensedmatter 09 00045 g043
Figure 44. (a) Magnetization of NaFe3(HPO3)2(H2PO3)6 in static field up to 9 T and pulsed field up to 32 T. The inset shows the temperature dependence of the specific heat Cp/T taken at various magnetic fields [65]. (Reproduced with permission from [65]). (b) Spin lattice of NaFe3(HPO3)2(H2PO3)6 in terms of triangular ferrimagnetic fragments. (c) Ferrimagnetic ground state of a layer made up of the exchange paths J2, J3, J6 and J4 in NaFe3(HPO3)2(H2PO3)6, which has diamond chains (defined by J2, J3 and J6) antiferromagnetically coupled (via J4). (d) Spin arrangement in the FM state reached magnetic saturation.
Figure 44. (a) Magnetization of NaFe3(HPO3)2(H2PO3)6 in static field up to 9 T and pulsed field up to 32 T. The inset shows the temperature dependence of the specific heat Cp/T taken at various magnetic fields [65]. (Reproduced with permission from [65]). (b) Spin lattice of NaFe3(HPO3)2(H2PO3)6 in terms of triangular ferrimagnetic fragments. (c) Ferrimagnetic ground state of a layer made up of the exchange paths J2, J3, J6 and J4 in NaFe3(HPO3)2(H2PO3)6, which has diamond chains (defined by J2, J3 and J6) antiferromagnetically coupled (via J4). (d) Spin arrangement in the FM state reached magnetic saturation.
Condensedmatter 09 00045 g044
Figure 45. (a) Spin exchanges J1 and J5 between adjacent layers of diamond chains linked by J4 in NaFe3(HPO3)2(H2PO3)6. The ferrimagnetic triangular clusters belonging to two different layers are marked with different colors. The red labels 2, 3, 6 and 4 refer to the spin exchanges J2, J3, J6 and J4 of one layer, respectively. The green labels 1 and 5 refer to the interlayer spin exchanges J1 and J5, respectively. (b) Interlayer FM coupling between adjacent ferrimagnetic layers leading to J5 magnetic bonds and J1 broken magnetic bonds. (c) Interlayer AFM coupling between adjacent ferrimagnetic layers leading to J5 broken magnetic bonds and J1 magnetic bonds.
Figure 45. (a) Spin exchanges J1 and J5 between adjacent layers of diamond chains linked by J4 in NaFe3(HPO3)2(H2PO3)6. The ferrimagnetic triangular clusters belonging to two different layers are marked with different colors. The red labels 2, 3, 6 and 4 refer to the spin exchanges J2, J3, J6 and J4 of one layer, respectively. The green labels 1 and 5 refer to the interlayer spin exchanges J1 and J5, respectively. (b) Interlayer FM coupling between adjacent ferrimagnetic layers leading to J5 magnetic bonds and J1 broken magnetic bonds. (c) Interlayer AFM coupling between adjacent ferrimagnetic layers leading to J5 broken magnetic bonds and J1 magnetic bonds.
Condensedmatter 09 00045 g045
Figure 46. (a) (left) A Cu5 ribbon made up of one Cu1 and four Cu2 atoms in azurite Cu3(CO3)2(OH)2, where the labels 1–3 refer to the spin exchange paths J1J3, respectively. In this ribbon, the four Cu2O4 square planes are nearly orthogonal to the Cu1O4 square plane (right). (b) One diamond chain made up of edge-sharing ribbons. (c) (left) Arrangement between two Cu5 ribbons leading to the inter-ribbon exchanges J4, which occurs through a CO3 bridge (right). (d) A layer of diamond chains parallel to the ab-plane made up of edge-sharing Cu5 ribbons.
Figure 46. (a) (left) A Cu5 ribbon made up of one Cu1 and four Cu2 atoms in azurite Cu3(CO3)2(OH)2, where the labels 1–3 refer to the spin exchange paths J1J3, respectively. In this ribbon, the four Cu2O4 square planes are nearly orthogonal to the Cu1O4 square plane (right). (b) One diamond chain made up of edge-sharing ribbons. (c) (left) Arrangement between two Cu5 ribbons leading to the inter-ribbon exchanges J4, which occurs through a CO3 bridge (right). (d) A layer of diamond chains parallel to the ab-plane made up of edge-sharing Cu5 ribbons.
Condensedmatter 09 00045 g046
Figure 47. (a) Field dependence of magnetization of Cu3(CO3)2(OH)2 for H||b. (b) Field dependence of magnetization in Cu3(CO3)2(OH)2 for Hb [71]. (Reproduced with permission from [71]).
Figure 47. (a) Field dependence of magnetization of Cu3(CO3)2(OH)2 for H||b. (b) Field dependence of magnetization in Cu3(CO3)2(OH)2 for Hb [71]. (Reproduced with permission from [71]).
Condensedmatter 09 00045 g047
Figure 48. (a) Stacking of 2D layers made up of interlinked diamond chains in Cu3(CO3)2(OH)2. (b) Two 2D layers with interlayer spin exchange paths J5 and J6. (c) Interlayer exchange paths J5 between Cu1 and Cu2 atoms. (d) Interlayer exchange paths J6 between two Cu2 atoms. The labels 4–6 refer to the spin exchange paths J4J6, respectively.
Figure 48. (a) Stacking of 2D layers made up of interlinked diamond chains in Cu3(CO3)2(OH)2. (b) Two 2D layers with interlayer spin exchange paths J5 and J6. (c) Interlayer exchange paths J5 between Cu1 and Cu2 atoms. (d) Interlayer exchange paths J6 between two Cu2 atoms. The labels 4–6 refer to the spin exchange paths J4J6, respectively.
Condensedmatter 09 00045 g048
Figure 49. (a) One Cu12+ ion making 2J1 + 2J3 + 2J4 exchange bonds with six adjacent Cu22+ in azurite. (b) Idealized description associated with the 2J1 + 2J3 + 2J4 exchange bonds. The idealized Cu1O4 (shaded) and Cu2O4 units (unshaded) units are treated as ideal square planes with four-fold rotation symmetry, with the edges of Cu1O4 parallel to the x- and z-axes, and those of Cu2O4 parallel to the x- and y-axes. (c) The g-factors of and the amount of unquenched orbital moment on the Cu2+ ion in the Cu1O4 and Cu2O4 square planes.
Figure 49. (a) One Cu12+ ion making 2J1 + 2J3 + 2J4 exchange bonds with six adjacent Cu22+ in azurite. (b) Idealized description associated with the 2J1 + 2J3 + 2J4 exchange bonds. The idealized Cu1O4 (shaded) and Cu2O4 units (unshaded) units are treated as ideal square planes with four-fold rotation symmetry, with the edges of Cu1O4 parallel to the x- and z-axes, and those of Cu2O4 parallel to the x- and y-axes. (c) The g-factors of and the amount of unquenched orbital moment on the Cu2+ ion in the Cu1O4 and Cu2O4 square planes.
Condensedmatter 09 00045 g049
Figure 50. Fragmentation of (a) a kagomé and (b) a trigonal spin lattice into non-overlapping ferrimagnetic triangles. (c) Three possible spin arrangements of a ferrimagnetic triangle, where the up-spin and down-spin sites are indicated by unshaded and shaded circles, respectively.
Figure 50. Fragmentation of (a) a kagomé and (b) a trigonal spin lattice into non-overlapping ferrimagnetic triangles. (c) Three possible spin arrangements of a ferrimagnetic triangle, where the up-spin and down-spin sites are indicated by unshaded and shaded circles, respectively.
Condensedmatter 09 00045 g050
Figure 51. Three ordered spin arrangements representing the 1/3-magnetization plateau state of a kagomé spin lattice.
Figure 51. Three ordered spin arrangements representing the 1/3-magnetization plateau state of a kagomé spin lattice.
Condensedmatter 09 00045 g051
Figure 52. Three ordered spin arrangements representing the 1/3-magnetization plateau state of a trigonal spin lattice. The spin arrangements of the two nonequivalent ferrimagnetic triangles are encircled for clarity.
Figure 52. Three ordered spin arrangements representing the 1/3-magnetization plateau state of a trigonal spin lattice. The spin arrangements of the two nonequivalent ferrimagnetic triangles are encircled for clarity.
Condensedmatter 09 00045 g052
Figure 53. (a) Arrangement of three unshaded triangles around a ferrimagnetic triangle, indicated by shading, in a kagomé spin lattice. (b) The most stable arrangement of six ferrimagnetic triangles around one ferrimagnetic fragment. (c) The least stable arrangement of six ferrimagnetic triangles around one ferrimagnetic fragment. (d) An arrangement that makes four broken bonds and requires the breaking of two bonds for the (↑↓↑) to (↑↑↑) spin flipping. The red circles indicate the down-spin site to go through (↑↓↑) to (↑↑↑) spin flip.
Figure 53. (a) Arrangement of three unshaded triangles around a ferrimagnetic triangle, indicated by shading, in a kagomé spin lattice. (b) The most stable arrangement of six ferrimagnetic triangles around one ferrimagnetic fragment. (c) The least stable arrangement of six ferrimagnetic triangles around one ferrimagnetic fragment. (d) An arrangement that makes four broken bonds and requires the breaking of two bonds for the (↑↓↑) to (↑↑↑) spin flipping. The red circles indicate the down-spin site to go through (↑↓↑) to (↑↑↑) spin flip.
Condensedmatter 09 00045 g053
Figure 54. (a) Arrangement of 12 unshaded triangles around a ferrimagnetic triangle, indicated by shading, in a trigonal spin lattice. (b) The most stable arrangement of six ferrimagnetic triangles around one ferrimagnetic fragment. (c) The least stable arrangement of six ferrimagnetic triangles around one ferrimagnetic fragment. The red circles indicate the down-spin site to go through (↑↓↑) to (↑↑↑) spin flip.
Figure 54. (a) Arrangement of 12 unshaded triangles around a ferrimagnetic triangle, indicated by shading, in a trigonal spin lattice. (b) The most stable arrangement of six ferrimagnetic triangles around one ferrimagnetic fragment. (c) The least stable arrangement of six ferrimagnetic triangles around one ferrimagnetic fragment. The red circles indicate the down-spin site to go through (↑↓↑) to (↑↑↑) spin flip.
Condensedmatter 09 00045 g054
Figure 55. (a) Trigonal layer of MO6 (M = Fe, Co, Ni) octahedra found in RbFe(MoO4)2, Ba3CoSb2O9 and Ba2LaNiTe2O12. (b,c) Two views of how a trigonal layer of FeO6 octahedra is capped by MoO4 tetrahedra. (d) Views of a MoO4 tetrahedron, a Sb2O9 double octahedron and a TeO6 octahedron capping the trigonal layers of FeO6, CoO6 and NiO6 octahedra.
Figure 55. (a) Trigonal layer of MO6 (M = Fe, Co, Ni) octahedra found in RbFe(MoO4)2, Ba3CoSb2O9 and Ba2LaNiTe2O12. (b,c) Two views of how a trigonal layer of FeO6 octahedra is capped by MoO4 tetrahedra. (d) Views of a MoO4 tetrahedron, a Sb2O9 double octahedron and a TeO6 octahedron capping the trigonal layers of FeO6, CoO6 and NiO6 octahedra.
Condensedmatter 09 00045 g055
Figure 56. Field dependence of magnetization observed for (a) RbFe(MoO4)2 [87]. (Reproduced with permission from [87]). (b) Ba3CoSb2O9 in the ab-plane at T  =  0.6  K [89]. (c) Ba2LaNiTe2O12 at 1.3 K [82]. (Reproduced with permission from [82]).
Figure 56. Field dependence of magnetization observed for (a) RbFe(MoO4)2 [87]. (Reproduced with permission from [87]). (b) Ba3CoSb2O9 in the ab-plane at T  =  0.6  K [89]. (c) Ba2LaNiTe2O12 at 1.3 K [82]. (Reproduced with permission from [82]).
Condensedmatter 09 00045 g056
Figure 57. (a) The M-O…O-M spin exchange path J in a trigonal layer of MO6 octahedra in RbFe(MoO4)2, Ba3CoSb2O9 and Ba2LaNiTe2O12. The dotted lines represent the O…O contacts, and each exchange path consists of two O…O contacts. (b) The O…O distance, the observed width Δ(μ0H) of the 1/3-plateau, and the calculated energies JS2 of the nearest-neighbor magnetic bonds in RbFe(MoO4)2, Ba3CoSb2O9 and Ba2LaNiTe2O12.
Figure 57. (a) The M-O…O-M spin exchange path J in a trigonal layer of MO6 octahedra in RbFe(MoO4)2, Ba3CoSb2O9 and Ba2LaNiTe2O12. The dotted lines represent the O…O contacts, and each exchange path consists of two O…O contacts. (b) The O…O distance, the observed width Δ(μ0H) of the 1/3-plateau, and the calculated energies JS2 of the nearest-neighbor magnetic bonds in RbFe(MoO4)2, Ba3CoSb2O9 and Ba2LaNiTe2O12.
Condensedmatter 09 00045 g057
Figure 58. (a) Zigzag chain of trimer and dimer units in sodium copper pyrosilicate, Na2Cu5(Si2O7)2. (b) Ferrimagnetic state of a chain obtained by an AFM coupling between ferrimagnetic trimers and FM dimers. (c) Field dependence of magnetization in Na2Cu5(Si2O7)2 measured at 2 K [93]. (d) Ferrimagnetic state of a chain obtained by an FM coupling between ferrimagnetic trimers and FM dimers.
Figure 58. (a) Zigzag chain of trimer and dimer units in sodium copper pyrosilicate, Na2Cu5(Si2O7)2. (b) Ferrimagnetic state of a chain obtained by an AFM coupling between ferrimagnetic trimers and FM dimers. (c) Field dependence of magnetization in Na2Cu5(Si2O7)2 measured at 2 K [93]. (d) Ferrimagnetic state of a chain obtained by an FM coupling between ferrimagnetic trimers and FM dimers.
Condensedmatter 09 00045 g058
Figure 59. (ac) Building blocks of Y2Cu7(TeO3)6Cl6(OH)2. A trimer unit in (a), two diamond units in (b), and a heptamer formed by a trimer with two dimers using the Cu-O…O-Cu exchange paths in (c). (d) Field dependence of magnetization in Y2Cu7(TeO3)6Cl6(OH)2 showing a 3/7-magnetization plateau [94]. (Reproduced with permission from [94]).
Figure 59. (ac) Building blocks of Y2Cu7(TeO3)6Cl6(OH)2. A trimer unit in (a), two diamond units in (b), and a heptamer formed by a trimer with two dimers using the Cu-O…O-Cu exchange paths in (c). (d) Field dependence of magnetization in Y2Cu7(TeO3)6Cl6(OH)2 showing a 3/7-magnetization plateau [94]. (Reproduced with permission from [94]).
Condensedmatter 09 00045 g059
Figure 60. (a) Values of the spin exchanges in K. (b) A (5↑2↓) spin arrangement of a linear heptamer in Y2Cu7(TeO3)6Cl6(OH)2. (c) Heptamers interacting through the J5 spin exchanges. (d) AFM arrangements between adjacent heptamers leading to a heptamer chain.
Figure 60. (a) Values of the spin exchanges in K. (b) A (5↑2↓) spin arrangement of a linear heptamer in Y2Cu7(TeO3)6Cl6(OH)2. (c) Heptamers interacting through the J5 spin exchanges. (d) AFM arrangements between adjacent heptamers leading to a heptamer chain.
Condensedmatter 09 00045 g060
Figure 61. (a) A layer of Cu5(VO4)2(OH)4 made up of three nonequivalent CuO4 planes by corner- and edge-sharing. (b) Pattern of the Cu2+ ion arrangement showing chains of edge-sharing hexagons composed of six triangles. (c) Magnetization curve in Cu5(VO4)2(OH)4 at 2 K [96], where the magnetization of a paramagnetic S = 1/2 ion (red curve) was added for comparison. (Reproduced with permission from [96]).
Figure 61. (a) A layer of Cu5(VO4)2(OH)4 made up of three nonequivalent CuO4 planes by corner- and edge-sharing. (b) Pattern of the Cu2+ ion arrangement showing chains of edge-sharing hexagons composed of six triangles. (c) Magnetization curve in Cu5(VO4)2(OH)4 at 2 K [96], where the magnetization of a paramagnetic S = 1/2 ion (red curve) was added for comparison. (Reproduced with permission from [96]).
Condensedmatter 09 00045 g061
Figure 62. (a) Seven spin exchange paths in Cu5(VO4)2(OH)4 defined with respect to the crystal structure given in Figure 61. (b) Different arrangements of the two CuO4 planes associated with the J5 and J6 spin exchange paths. (c) Values of the calculated spin exchanges and the Cu…Cu distances associated with the spin exchange paths. (d) A zigzag ferrimagnetic fragment of (3↑2↓) spin configuration.
Figure 62. (a) Seven spin exchange paths in Cu5(VO4)2(OH)4 defined with respect to the crystal structure given in Figure 61. (b) Different arrangements of the two CuO4 planes associated with the J5 and J6 spin exchange paths. (c) Values of the calculated spin exchanges and the Cu…Cu distances associated with the spin exchange paths. (d) A zigzag ferrimagnetic fragment of (3↑2↓) spin configuration.
Condensedmatter 09 00045 g062
Figure 63. (a) AFM arrangement of two adjacent ferrimagnetic fragments within a hexagon chain. (b) FM arrangement of two adjacent ferrimagnetic fragments within a hexagon chain. (c) AFM arrangement of adjacent ferrimagnetic fragments within and between hexagon chains. (d) FM arrangement of adjacent ferrimagnetic fragments within and between hexagon chains.
Figure 63. (a) AFM arrangement of two adjacent ferrimagnetic fragments within a hexagon chain. (b) FM arrangement of two adjacent ferrimagnetic fragments within a hexagon chain. (c) AFM arrangement of adjacent ferrimagnetic fragments within and between hexagon chains. (d) FM arrangement of adjacent ferrimagnetic fragments within and between hexagon chains.
Condensedmatter 09 00045 g063
Figure 64. (a) A (Cu2+)7 heptamer made up of two corner-sharing tetrahedra in Pb2Cu10O4(SeO3)4Cl7. (b) Magnetization curve observed for Pb2Cu10O4(SeO3)4Cl7 at 2 K [97]. (Reproduced with permission from [97]). (c) A (5↑2↓) spin configuration of a heptamer. (d) Bridging mode between heptamers to form a chain. (e) AFM coupling between heptamers (f) FM coupling between heptamers.
Figure 64. (a) A (Cu2+)7 heptamer made up of two corner-sharing tetrahedra in Pb2Cu10O4(SeO3)4Cl7. (b) Magnetization curve observed for Pb2Cu10O4(SeO3)4Cl7 at 2 K [97]. (Reproduced with permission from [97]). (c) A (5↑2↓) spin configuration of a heptamer. (d) Bridging mode between heptamers to form a chain. (e) AFM coupling between heptamers (f) FM coupling between heptamers.
Condensedmatter 09 00045 g064
Figure 65. (a) Magnetization curve in Na2Cu7(SeO3)4O2Cl4 at 2 K [98]. (Reproduced with permission from [98]). (b) A (Cu2+)7 heptamer made up of two corner-sharing tetrahedra. (c) Bridging mode between heptamers to form a chain. (d) Spin exchanges (in K) determined using DFT+U calculations. (e) A (5↑2↓) spin configuration of a heptamer.
Figure 65. (a) Magnetization curve in Na2Cu7(SeO3)4O2Cl4 at 2 K [98]. (Reproduced with permission from [98]). (b) A (Cu2+)7 heptamer made up of two corner-sharing tetrahedra. (c) Bridging mode between heptamers to form a chain. (d) Spin exchanges (in K) determined using DFT+U calculations. (e) A (5↑2↓) spin configuration of a heptamer.
Condensedmatter 09 00045 g065
Figure 66. (a) A (Cu2+)7 heptamer of Pb2Cu10O4(SeO3)4Cl7 composed of five CuO4 square planes and two CuO4Cl trigonal bipyramids. (b) The x2-y2 magnetic orbital of the central CuO4 square plane (SP) interacting with the xy magnetic orbitals of the two trigonal bipyramids (TBPs), leading to a strong (↑↓↑) spin coupling of the three Cu2+ ions of the linear TBP-SP-TBP paths. (c) A (5↑2↓) spin configuration of a heptamer.
Figure 66. (a) A (Cu2+)7 heptamer of Pb2Cu10O4(SeO3)4Cl7 composed of five CuO4 square planes and two CuO4Cl trigonal bipyramids. (b) The x2-y2 magnetic orbital of the central CuO4 square plane (SP) interacting with the xy magnetic orbitals of the two trigonal bipyramids (TBPs), leading to a strong (↑↓↑) spin coupling of the three Cu2+ ions of the linear TBP-SP-TBP paths. (c) A (5↑2↓) spin configuration of a heptamer.
Condensedmatter 09 00045 g066
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Whangbo, M.-H.; Koo, H.-J.; Kremer, R.K.; Vasiliev, A.N. Magnetization Plateaus by the Field-Induced Partitioning of Spin Lattices. Condens. Matter 2024, 9, 45. https://doi.org/10.3390/condmat9040045

AMA Style

Whangbo M-H, Koo H-J, Kremer RK, Vasiliev AN. Magnetization Plateaus by the Field-Induced Partitioning of Spin Lattices. Condensed Matter. 2024; 9(4):45. https://doi.org/10.3390/condmat9040045

Chicago/Turabian Style

Whangbo, Myung-Hwan, Hyun-Joo Koo, Reinhard K. Kremer, and Alexander N. Vasiliev. 2024. "Magnetization Plateaus by the Field-Induced Partitioning of Spin Lattices" Condensed Matter 9, no. 4: 45. https://doi.org/10.3390/condmat9040045

APA Style

Whangbo, M.-H., Koo, H.-J., Kremer, R. K., & Vasiliev, A. N. (2024). Magnetization Plateaus by the Field-Induced Partitioning of Spin Lattices. Condensed Matter, 9(4), 45. https://doi.org/10.3390/condmat9040045

Article Metrics

Back to TopTop