This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Magnetic field-induced deformation of the spin-density wave microphases in Ca3Co2O6

Y. Kamiya School of Physics and Astronomy, Shanghai Jiao Tong University, Shanghai 200240, China
Abstract

The frustrated triangular Ising magnet Ca3Co2O6 has long been known for an intriguing combination of extremely slow spin dynamics and peculiar magnetic orders, such as the evenly-spaced non-equilibrium metamagnetic magnetization steps and the long-wavelength spin density wave (SDW) order, the latter of which is essentially an emergent crystal of solitons. Recently, an elaborate field-cooling protocol to bypass the low-field SDW phase was proposed to overcome the extraordinarily long timescale of spin relaxation that impeded previous experimental studies in equilibrium, which may point to a deep connection between the low-temperature slow relaxation and the cooling process passing through the low-field SDW phase. As the first step to elucidate the conjectured connection, we investigate the magnetic field-induced deformation of the SDW state and incommensurate-commensurate transitions, thereby mapping out the equilibrium in-field phase diagram for a realistic three-dimensional lattice spin model by using Monte Carlo simulations. We also discuss Ginzburg-Landau theory that includes several Umklapp terms as well as an effective sine-Gordon model, which can qualitatively explain the observed magnetic field-induced deformation of the SDW microphases.

I Introduction

Frustrated magnets can have a manifold of nearly degenerate low-energy states from which interesting phenomena may emerge, such as exotic magnetic and nonmagnetic orders, topological order, liquid-like or even glassy behavior, and so on, varying from one material to another.Lacroix2011 Even a classical system can host unconventional quasiparticles, such as skyrmions,Bogdanov2020 solitons, and monopoles in spin ice,Bramwell01; Paulsen2014; Bramwell20 and they may crystallize into novel spin textures like skyrmion crystals Okubo2012; Leonov2015 or soliton crystals.Bak82; Selke88 Such emergent crystalline states can often be sensitive to external perturbations, which makes them attractive as potential devices in some cases.Fert2017 They can also provide a platform to study far-from-equilibrium dynamics due to metastable states.Kudasov06; Paulsen2014

Refer to caption
Figure 1: (a) Magnetic lattice of CCO with the intrachain coupling J1J_{1} and the interchain couplings J2J_{2} and J3J_{3}, which is superimposed on the schematic crystal structure with sublattices AA, BB, and CC. (b) Another schematic picture for the magnetic lattice, where only those exchange interactions that connect to the central site are shown to illustrate them in a concise manner. (c) Effective one-dimensional lattice used in the mean field theory (see the text), where an arrow represents a magnetic moment of a given abab plane.

Since the late 90’s,Fjellvag96; Aasland97; Kageyama97; Kageyama97b the frustrated triangular Ising magnet Ca3Co2O6 (CCO) has long been known for an intriguing combination of extremely slow spin dynamics and peculiar magnetic phases, such as metamagnetic magnetization steps and an incommensurate spin-density wave (SDW) state,Agrestini08a; Mazzoli09; Agrestini08b; Moyoshi11; Fleck10; Paddison14; Motoya18 the latter of which can be seen as a soliton crystal.Kamiya12 In CCO, trigonal prismatic Co3+ S=2S=2 sites form ferromagnetic Ising chains running along the cc axis, while arranged in a triangular lattice in the abab plane coupled by weak antiferromagnetic interactions (Fig. 1).Leedahl19 Below spin-freezing temperature TSF5KT_{\mathrm{SF}}\simeq 5\,\mathrm{K}, CCO exhibits striking evenly-spaced metamagnetic magnetization steps,Kageyama97 whose origin has been a subject of long-time debates.Kageyama97; Kageyama97b; Maignan00; Hardy04b; Maignan04; Moyoshi11; Kim18 Interestingly, while the step heights are sensitive to protocol details such as sweep rate of the external magnetic field, the transition magnetic fields (\simeq1.2 T, 2.4 T, and 3.6 T, with additional steps at higher fields) are rather robust.Hardy04b Although some theory invoked an analogy with quantum tunneling in molecular magnets,Maignan04 an alternative scenario is that peculiar frustration in CCO causes a non-equilibrium phenomenon.Kudasov06 In the so-called “rigid chain” model,Yao06a; Yao06b; Kudasov08; Qin09; Soto09; Zukovic12 each ferromagnetic chain is replaced by an effective Ising spin on a two-dimensional antiferromagnetic triangular lattice. Based on this mapping, it was argued that the metamagnetic transition steps in CCO may arise from the same kind of degenerate manifold as in the two-dimensional triangular lattice Ising model.Kudasov06

However, the origin of the slow dynamics can be more intricate than suggested by the rigid chain picture. More recently, resonant x-ray Agrestini08a; Mazzoli09 and neutron spectroscopies Agrestini08b; Moyoshi11; Fleck10; Paddison14; Motoya18 revealed the SDW order below TSDWT_{\mathrm{SDW}}\simeq 25 K, which has a three-sublattice structure and a very long modulation wavelength λSDW103\lambda_{\mathrm{SDW}}\!\simeq\!10^{3}Å\AA (102\simeq\!\!10^{2} magnetic sites) along the cc axis. It was found that λSDW\lambda_{\mathrm{SDW}} increases as temperature TT is lowered and the corresponding relaxation time grows substantially. Eventually, the system starts to deviate from equilibrium below TT\lesssim 13 K,Moyoshi11 which is much higher than TSFT_{\mathrm{SF}} for the appearance of the metamagnetic magnetization steps. Since the spin chains are not ferromagnetically ordered in the SDW state, the interpretation of the rigid chain picture is, if not questionable, more subtle than originally proposed.

Indeed, the SDW phase may hold the key to understanding the peculiar slow dynamics at low temperatures. It was recently demonstrated that the slow spin dynamics can be bypassed by an elaborate field-cooling protocol, where every in-field measurement is performed after a separate cooling in the target magnetic field.Nekrashevic21 Remarkably, the protocol allowed for reaching the 1/3 magnetization plateau down to T=2K<TSFT=2\,\mathrm{K}<T_{\mathrm{SF}} without being suffered from metastable states, which was in good agreement with MC simulations in equilibrium. Since the SDW order is believed to disappear and replaced by a ferrimagnetic state at high magnetic fields, the new experiment may suggest that the spin relaxation at low temperatures may be influenced by the extent to which the system has been through the low-field SDW phase during the cooling. In fact, it is known that the SDW order is accompanied by short-range order indicative of spin disordering,Agrestini11 which could be due to the combination of the TT-dependent ordering wavevector and the slowness in the relaxation to follow the variation. Moyoshi11

As mentioned above, the observed SDW state is essentially a soliton crystal as in the axial next-nearest-neighbor Ising (ANNNI) model,Kamiya12 a prototypical model for spontaneous superstructures due to competition between nearest and next nearest neighbor Ising interactions in one direction of a square or cubic lattice.Bak82; Selke88 The TT-dependent change of λSDW\lambda_{\mathrm{SDW}} Moyoshi11 corresponds to different magnetic microphases,Kamiya12 similar to other self-organizing modulated phases in physical and chemical systems.Seul1995 Thus, CCO may provide a rare intersection where the ANNNI model phenomenology Bak82; Selke88 meets out-of-equilibrium physics in a solid state system with only short-range interactions. To elucidate this conjecture in CCO and similar materials such as Ca3Co2-xMnxO6,Zapf18; Kim18 Sr2Ca2CoMn2O9,Hardy18 and Ca3CoRhO6,Niitaka01 it is a matter of paramount importance to investigate the magnetic field-induced deformation of the SDW state and incommensurate-commensurate (IC-C) transitions under the condition much closer to equilibrium than ever reached before. The goal of this work to present an equilibrium in-field phase diagram for a realistic three-dimensional (3D) lattice spin model for CCO, thereby providing a theoretical guide for experiments. We address both model-specific and universal physics by combining mean-field theory (Sec. III), MC simulations (Sec. IV), and Ginzburg-Landau (GL) theory (Sec. V).

II Model

In CCO, S=2S=2 spins have large easy-axis anisotropy,Kageyama97 which permits a description by an effective classical Ising model,

H^=ν=1,2,3ijνJνσizσjzhiσiz,\displaystyle\hat{H}=\sum_{\nu=1,2,3}\sum_{\langle{ij}\rangle_{\nu}}J_{\nu}\sigma^{z}_{i}\sigma^{z}_{j}-h\sum_{i}\sigma^{z}_{i}, (1)

where σiz=±1\sigma^{z}_{i}=\pm 1, h=gμBSHh=g\mu_{B}SH with gg, μB\mu_{B}, and HH being the gg-factor, the Bohr magneton, and a magnetic field, respectively, and ijν\langle{ij}\rangle_{\nu} denotes neighboring sites connected by JνJ_{\nu}, ν{1,2,3}\nu\in\{1,2,3\}. J1<0J_{1}<0 is the intrachain ferromagnetic interaction and J2J_{2} (J3J_{3}) is the antiferromagnetic interchain interaction shifted by 1/3 (2/3) lattice parameters along the cc axis (Fig. 1). An ab initio study suggested |J1|J2J3\lvert{J_{1}}\rvert\gg J_{2}\simeq J_{3} Fresard04 and J1=23.9(2)J_{1}=-23.9(2) K and J2+J3=2.3(2)J_{2}+J_{3}=2.3(2) K was reported by an NMR experiment, which further suggested J2=1.1J_{2}=1.1 K and J3=1.2J_{3}=1.2 K to explain the SDW ordering wavevector.Allodi14 Below, for simplicity, we assume J2/|J1|=J3/|J1|J_{2}/\lvert{J_{1}}\rvert=J_{3}/\lvert{J_{1}}\rvert and denote the ratio by κ\kappa; in relation with CCO, κ0.048\kappa\simeq 0.048.

Refer to caption
Figure 2: (a) Mean field (κ,T)(\kappa,T) phase diagram for h=0h=0, where κJ2/|J1|=J3/|J1|\kappa\equiv J_{2}/\lvert{J_{1}}\rvert=J_{3}/\lvert{J_{1}}\rvert. The mean field ground state is \uparrow\uparrow\downarrow or \downarrow\downarrow\uparrow (\uparrow\uparrow\downarrow\downarrow) along the cc axis for κ<1\kappa<1 (κ>1\kappa>1) in the effective one-dimensional description, the ordering wavevector of which is Q3/(2π)=1Q_{3}/(2\pi)=1 (3/43/4), respectively. (b) (TT, hh) phase diagram obtained by MC simulations for κ=0.1\kappa=0.1 based on the data for T0.45|J1|T\gtrsim 0.45\lvert{J_{1}}\rvert. The TT-sweep data for TcT_{c} or TICCT_{\mathrm{IC-C}} is determined by analyzing the Binder parameter as a function of TT, whereas the field-sweep data for hch_{c} is determined from a peak in χ=dM/dh\chi=dM/dh as a function of hh.
Refer to caption
Figure 3: TT-dependence of the third component of the effective ordering wavevector QL,3effQ^{\mathrm{eff}}_{L,3} (upper panels) and specific heat CC (lower panels) obtained by MC simulations for κ=0.1\kappa=0.1 and (a) h=0h=0, (b) h/|J1|=0.05h/\lvert{J_{1}}\rvert=0.05, and (c) h/|J1|=0.2h/\lvert{J_{1}}\rvert=0.2. The additional subpeaks in CC are considered as spurious ones (see the text).

III Mean field theory

In CCO, J2J_{2} and J3J_{3} compete with J1J_{1} after a few steps along a spiral path due to the vertical shifts of the interchain interactions.Chapon09 This spiral structure is the key to realize the same kind of geometrical frustration as in the ANNNI model Bak82; Selke88 despite the apparent structural differences. We first briefly discuss a heuristic mean-field theory in zero field by assuming a ferromagnetic order in each abab plane, which are separated by 1/3 lattice parameters from each other along the cc axis (Fig. 1).Kamiya12 The mean field equation for the magnetization mlm_{l} of layer ll is

ml=tanhβhl,\displaystyle\left\langle{m_{l}}\right\rangle=\tanh\beta h_{l}, (2)

where hl=J1(ml+3+ml3)3J2(ml+1+ml1)3J3(ml+2+ml2)h_{l}=-J_{1}(\left\langle{m_{l+3}}\right\rangle+\left\langle{m_{l-3}}\right\rangle)-3J_{2}(\left\langle{m_{l+1}}\right\rangle+\left\langle{m_{l-1}}\right\rangle)-3J_{3}(\left\langle{m_{l+2}}\right\rangle+\left\langle{m_{l-2}}\right\rangle). In this quasi-one-dimensional description, J1J_{1}, J2J_{2}, and J3J_{3} serve as the effective third, first, and second neighbor interactions, respectively, realizing a very similar situation as in the prototypical ANNNI model [Fig. 1(c)]. The reason for assuming an in-plane ferromagnetic order, even though the interchain interactions J2J_{2} and J3J_{3} are much smaller than the intrachain interaction J1J_{1}, is that the energy scale associated with the competition between SDW states with different wavelengths along the cc axis can be even smaller, as will be discussed by using a sine-Gordon model. In Fig. 2(a), we show the mean field (κ,T)(\kappa,T) phase diagram, by extending the previous work.Kamiya12 Below the SDW transition temperature TSDW(κ)T_{\mathrm{SDW}}(\kappa), the ordering wavevector 𝐐=(0,0,Q3)\mathbf{Q}=(0,0,Q_{3}) varies quasi-continuously. Eventually, there is a lock-in IC-C transition at T=TICCT=T_{\mathrm{IC-C}}, below which the magnetic unit cell of the mean field solution is \uparrow\uparrow\downarrow or \downarrow\downarrow\uparrow for κ<1\kappa<1 and \uparrow\uparrow\downarrow\downarrow for κ>1\kappa>1 in the effective one-dimensional description in Fig. 1(c). We find that quasi-continuous changes of Q3(κ,T)Q_{3}(\kappa,T) dominate the overall phase diagram, corresponding to numerous microphases of soliton lattice states, especially for relatively small κ\kappa.Kamiya12 Meanwhile, distinct discontinuous changes of Q3(κ,T)Q_{3}(\kappa,T) are also seen in the region with relatively large κ\kappa, where a few relatively extended commensurate states are found. However, the latter case has little significance in relation with CCO, where κ\kappa has been suggested to be very small.

IV Monte Carlo simulation

IV.1 Set up

Next, to demonstrate the ANNNI-like physics in an unbiased way, we present the results of our MC simulations. We consider a lattice of size L×L×LcL\times L\times L_{c} with periodic boundary conditions, where the total number of spins is Nspin=3L2LcN_{\text{spin}}=3L^{2}L_{c}. We combine single-spin updates, intra-chain cluster updates Kamiya12, and replica exchanges Hukushima96 included every 10 MC steps. Several hundreds of replicas are needed for largest lattices to maintain a reasonable exchange acceptance rate to guarantee efficient sampling at low temperatures (e.g., 400 replicas for 8×8×3208\times 8\times 320 for h0.2|J1|h\lesssim 0.2\lvert{J_{1}}\rvert). By fixing κ=0.1\kappa=0.1 and Lc/L=40L_{c}/L=40 in most cases shown below, we performed simulations for 2×2×802\times 2\times 808×8×3208\times 8\times 320. Here, although κ=0.1\kappa=0.1 is larger than κ0.048\kappa\simeq 0.048 estimated for CCO, no qualitatively different physics for smaller κ\kappa is suggested in our mean field phase diagram, as long as the SDW order is concerned [Fig. 2(a)].

The aspect ratio Lc/L=40L_{c}/L=40 is chosen to simulate long-wavelength SDW states with as little finite-size tension as possible while not making the system excessively anisotropic to address thermodynamic behaviors in 3D. As discussed by using a GL theory, the ordering wavevector 𝐐\mathbf{Q} at T=TSDWT=T_{\mathrm{SDW}} is expected to be the minima ±𝐪min\pm\mathbf{q}_{\text{min}} of the Fourier transform J(𝐪)J(\mathbf{q}) of the exchange interactions, which we find 𝐪min=(0,0,2π+ϵ)\mathbf{q}_{\text{min}}=(0,0,2\pi+\epsilon) with ϵ0.006×(2π)\epsilon\approx-0.006\times(2\pi) for κ0.048\kappa\simeq 0.048 and ϵ0.013×(2π)\epsilon\approx-0.013\times(2\pi) for κ=0.1\kappa=0.1. Because 𝐪min\mathbf{q}_{\text{min}} is very close to the three-sublattice commensurate wavevector 𝐪com=(0,0,2π)\mathbf{q}_{\text{com}}=(0,0,2\pi), even a single periodicity of the spin modulation requires a large number of unit cells along the cc axis. (Here, 𝐪=3𝐪com=(0,0,6π)\mathbf{q}=3\mathbf{q}_{\text{com}}=(0,0,6\pi) is equivalent to 𝐪=0\mathbf{q}=0 in our notation, but 𝐪=𝐪com\mathbf{q}=\mathbf{q}_{\text{com}} is not.) The minimum size thus required for κ0.048\kappa\simeq 0.048 is Lcmin=2π/|qmin,3qcom,3|160L_{c}^{\text{min}}=2\pi/\lvert{q_{\mathrm{min,3}}-q_{\mathrm{com,3}}}\rvert\approx 160, which can be a bit problematic. For κ=0.1\kappa=0.1, we find Lcmin80L_{c}^{\text{min}}\approx 80, which is also quite anisotropic but within the acceptable range. The aspect ratio Lc/L=40L_{c}/L=40 is determined on this basis.

IV.2 Modified Binder parameter method

To determine the transition point of a second-order phase transition into a commensurate ordering, say with a wavevector 𝐐\mathbf{Q}, a standard method is to analyze the order parameter M𝐐=Nspin1iσizexp(i𝐐𝐫i)M_{\mathbf{Q}}=N_{\text{spin}}^{-1}\sum_{i}\sigma^{z}_{i}\exp(-i\mathbf{Q}\cdot\mathbf{r}_{i}) and the corresponding Binder parameter U𝐐=|M𝐐|4/|M𝐐|22U_{\mathbf{Q}}=\langle{\lvert{M_{\mathbf{Q}}}\rvert^{4}}\rangle/\langle{\lvert{M_{\mathbf{Q}}}\rvert^{2}}\rangle^{2}.LandauBinder However, the possible TT-dependent variation as well as incommensurability of the ordering wavevector poses a challenge in numerical studies of the ANNNI model and its variants,Gendiar2005; KaiZhang2010; KaiZhang2011 demanding a modified approach. In such a model, a finite-size system with periodic boundary conditions is expected to develop a spin correlation whose dominant wavevector is necessarily commensurate but near the true ordering wavevector 𝐐(T)\mathbf{Q}(T) in thermodynamic limit. Such an effective ordering wavevector, 𝐐Leff(T)\mathbf{Q}^{\mathrm{eff}}_{L}(T), can be detected as a peak in the finite-size spin structure factor 𝒮𝐪=Nspin(|M𝐪|2|M𝐪|2)\mathcal{S}_{\mathbf{q}}=N_{\text{spin}}(\langle{\lvert{M_{\mathbf{q}}}\rvert^{2}}\rangle-\lvert{\langle{M_{\mathbf{q}}}\rangle}\rvert^{2}) and we expect limL𝐐Leff(T)=𝐐(T)\lim_{L\to\infty}\mathbf{Q}^{\mathrm{eff}}_{L}(T)=\mathbf{Q}(T). Indeed, the observed behavior of 𝐐Leff(T)\mathbf{Q}^{\mathrm{eff}}_{L}(T) shows relatively small variance with respect to LL, supporting this expectation (Fig. 3). Based on the estimate of 𝐐Leff(T)\mathbf{Q}^{\mathrm{eff}}_{L}(T), we evaluate |M𝐐Leff(T)|2\langle{\lvert{M_{\mathbf{Q}^{\mathrm{eff}}_{L}(T)}}\rvert^{2}} and |M𝐐Leff(T)|4\langle{\lvert{M_{\mathbf{Q}^{\mathrm{eff}}_{L}(T)}}\rvert^{4}}, thereby

U𝐐Leff=|M𝐐Leff(T)|4|M𝐐Leff(T)|22,\displaystyle U_{\mathbf{Q}^{\mathrm{eff}}_{L}}=\frac{\langle{\lvert{M_{\mathbf{Q}^{\mathrm{eff}}_{L}(T)}}\rvert^{4}}\rangle}{\langle{\lvert{M_{\mathbf{Q}^{\mathrm{eff}}_{L}(T)}}\rvert^{2}}\rangle^{2}}, (3)

which is a Binder parameter defined at 𝐪=𝐐Leff(T)\mathbf{q}=\mathbf{Q}^{\mathrm{eff}}_{L}(T). As in the usual usage of the Binder parameter,LandauBinder we look into the crossing point of U𝐐Leff(T)U_{\mathbf{Q}^{\mathrm{eff}}_{L}(T)} for different system sizes to evaluate TSDWT_{\mathrm{SDW}}.

Unlike the conventional approach, however, the wavevector 𝐪=𝐐Leff(T)\mathbf{q}=\mathbf{Q}^{\mathrm{eff}}_{L}(T) associated with the Binder parameter for one system size can be different from one for another size. Furthermore, it can also be TT-dependent. To analyze the effect of a small deviation of the wavevector from the true (incommensurate) ordering wavevector 𝐐\mathbf{Q}, we review the standard scaling theory for correlation functions.Cardy The two-point SDW correlation function G2(𝐫1𝐫2)=ϕc(𝐫1)ϕc(𝐫2)G_{2}(\mathbf{r}_{1}-\mathbf{r}_{2})=\langle{\phi_{c}(\mathbf{r}_{1})\phi_{c}(\mathbf{r}_{2})}\rangle, where ϕc(𝐫)σz(𝐫)ei𝐐c𝐫\phi_{c}(\mathbf{r})\sim\sigma^{z}(\mathbf{r})\,e^{-i\mathbf{Q}_{c}\cdot\mathbf{r}} is the coarse-grained SDW order parameter with the momentum 𝐐c𝐐(TSDW)\mathbf{Q}_{c}\equiv\mathbf{Q}(T_{\mathrm{SDW}}), is expected to have the following transformation under the scaling by a factor bb,

G2(𝐫,t)b2xSDWG2(b1𝐫,b1/νt),\displaystyle G_{2}(\mathbf{r},t)\sim b^{-2x_{\mathrm{SDW}}}G_{2}(b^{-1}\mathbf{r},b^{1/\nu}t), (4)

where t=(TTSDW)/TSDWt=(T-T_{\mathrm{SDW}})/T_{\mathrm{SDW}}, xSDWx_{\mathrm{SDW}} is the scaling dimension of the order parameter, and ν\nu is the critical exponent of correlation length ξ\xi. It follows that the correlation function has the universal finite-size scaling form of

G2(𝐫)ξ2xSDWΨ2(ξ1𝐫,ξ1L),\displaystyle G_{2}(\mathbf{r})\sim\xi^{-2x_{\mathrm{SDW}}}\Psi_{2}(\xi^{-1}\mathbf{r},\xi^{-1}L), (5)

from which we find

|M𝐪|2\displaystyle\left\langle{\left\lvert{M_{\mathbf{q}}}\right\rvert^{2}}\right\rangle 1Nspin2|σz(𝐫)ei𝐪𝐫𝑑𝐫|2\displaystyle\sim\frac{1}{N_{\text{spin}}^{2}}\left\langle{\left\lvert{\int\!\sigma^{z}(\mathbf{r})\,e^{-i\mathbf{q}\cdot\mathbf{r}}d\mathbf{r}\,}\right\rvert^{2}}\right\rangle
1NspinG2(𝐫)ei(𝐪𝐐c)𝐫𝑑𝐫\displaystyle\sim\frac{1}{N_{\text{spin}}}\int\!G_{2}(\mathbf{r})\,e^{-i(\mathbf{q}-\mathbf{Q}_{c})\cdot\mathbf{r}}d\mathbf{r}
L2xSDWΦ2(L(𝐪𝐐c),ξ1L).\displaystyle\sim L^{-2x_{\mathrm{SDW}}}\Phi_{2}(L(\mathbf{q}-\mathbf{Q}_{c}),\xi^{-1}L). (6)

A similar argument can also be made for the four-point correlation function G4(𝐫1,𝐫2,𝐫3,𝐫4)=ϕc(𝐫1)ϕc(𝐫2)ϕc(𝐫3)ϕc(𝐫4)G_{4}(\mathbf{r}_{1},\mathbf{r}_{2},\mathbf{r}_{3},\mathbf{r}_{4})=\langle{\phi_{c}(\mathbf{r}_{1})\phi_{c}(\mathbf{r}_{2})\phi_{c}(\mathbf{r}_{3})\phi_{c}(\mathbf{r}_{4})}\rangle, leading to

|M𝐪|4\displaystyle\left\langle{\left\lvert{M_{\mathbf{q}}}\right\rvert^{4}}\right\rangle L4xSDWΦ4(L(𝐪𝐐c),ξ1L),\displaystyle\sim L^{-4x_{\mathrm{SDW}}}\Phi_{4}(L(\mathbf{q}-\mathbf{Q}_{c}),\xi^{-1}L), (7)

Thus, near the critical point TTSDWT\approx T_{\mathrm{SDW}}, we find that the Binder parameter associated with the effective ordering wavevector behaves as

U𝐐LeffΦU(L(𝐐Leff𝐐c),ξ1L),\displaystyle U_{\mathbf{Q}_{L}^{\mathrm{eff}}}\sim\Phi_{U}(L(\mathbf{Q}_{L}^{\mathrm{eff}}-\mathbf{Q}_{c}),\xi^{-1}L), (8)

whereas U𝐐Leff2U_{\mathbf{Q}_{L}^{\mathrm{eff}}}\to 2 (corresponding to the Gaussian distribution for a one-component complex order parameter) and U𝐐Leff1U_{\mathbf{Q}_{L}^{\mathrm{eff}}}\to 1 in the high- and low-TT limits, respectively. Here, Ψ2\Psi_{2}, Φ2\Phi_{2}, Φ4\Phi_{4}, and ΦU\Phi_{U} are finite-size scaling functions.

Refer to caption
Figure 4: Monte Carlo results of the modified Binder parameter U𝐐LeffU_{\mathbf{Q}^{\mathrm{eff}}_{L}} [Eq. (3)] for (a) h/|J1|=0h/\lvert{J_{1}}\rvert=0, (b) h/|J1|=0.05h/\lvert{J_{1}}\rvert=0.05, and (c) h/|J1|=0.2h/\lvert{J_{1}}\rvert=0.2. The insets show enlarged views near T=TSDWT=T_{\mathrm{SDW}}.
Refer to caption
Figure 5: Monte Carlo results of the square of the commensurate order parameter (upper panels) and the C6C_{6} indicator to distinguish between the PDA state (C6<0C_{6}<0) and the FIM state (C6>0C_{6}>0) (lower panels) for (a) h/|J1|=0h/\lvert{J_{1}}\rvert=0, (b) h/|J1|=0.05h/\lvert{J_{1}}\rvert=0.05, and (c) h/|J1|=0.2h/\lvert{J_{1}}\rvert=0.2. The insets show enlarged views of the Binder parameter for the commensurate order parameter near T=TICCT=T_{\mathrm{IC-C}}.

It is reasonable to assume that these scaling functions are sufficiently isotropic with respect to small |𝐐Leff𝐐c|\lvert{\mathbf{Q}_{L}^{\mathrm{eff}}-\mathbf{Q}_{c}}\rvert, with an appropriate rescaling along the principle axes if needed.Cardy In fact, Z2Z_{2} reflection symmetry along the cc-axis concerning the sign of QL,3effQ3Q^{\mathrm{eff}}_{L,3}-Q_{3} is enough for the following discussion. As a wavevector best approximating the true ordering wavevector for a given system size LL, we expect |𝐐Leff𝐐c|O(L1)\lvert{\mathbf{Q}_{L}^{\mathrm{eff}}-\mathbf{Q}_{c}}\rvert\sim O(L^{-1}). Since, as usually the case, the universal scaling functions can be expected as a sufficiently slowly-varying function near the critical point TTSDWT\approx T_{\mathrm{SDW}}, the right hand side of Eq. (8) is almost a size-independent constant. Therefore, the Binder parameter at the size-dependent effective ordering wavevectors to exhibit a crossing behavior at around T=TSDWT=T_{\mathrm{SDW}}, as the temperature is lowered from the paramagnetic phase.

In the meantime, to determine TICCT_{\mathrm{IC-C}} for the lock-in IC-C transition into a commensurate phase in low fields, or TcT_{c} for a direct transition into the same phase in high fields, we use the ordinary the Binder parameter U𝐪com=|M𝐪com|4/|M𝐪com|22U_{\mathbf{q}_{\text{com}}}=\langle{\lvert{M_{\mathbf{q}_{\text{com}}}}\rvert^{4}}\rangle/\langle{\lvert{M_{\mathbf{q}_{\text{com}}}}\rvert^{2}}\rangle^{2} at the corresponding commensurate wavevector 𝐪=𝐪com=(0,0,2π)\mathbf{q}=\mathbf{q}_{\text{com}}=(0,0,2\pi).

IV.3 Ordering wavevector and the phase diagram

Below, we discuss the details of the phase diagram for κ=0.1\kappa=0.1 [Fig. 2(b)], obtained at T0.45|J1|T\gtrsim 0.45\lvert{J_{1}}\rvert, focusing on the behavior of the ordering wavevector (Fig. 3). At low fields, we observe that the ordering wavevector 𝐐Leff=(0,0,QL,3eff)\mathbf{Q}^{\mathrm{eff}}_{L}=(0,0,Q^{\mathrm{eff}}_{L,3}) at T=TSDWT=T_{\mathrm{SDW}} slightly, but clearly, deviates from 𝐪com=(0,0,2π)\mathbf{q}_{\text{com}}=(0,0,2\pi). Below TSDWT_{\mathrm{SDW}}, 𝐐Leff\mathbf{Q}^{\mathrm{eff}}_{L} slowly drifts towards 𝐪com=(0,0,2π)\mathbf{q}_{\text{com}}=(0,0,2\pi) as further lowering TT. Roughly speaking, the ordering wavevector changes more rapidly as the magnetic field is increased. The observed step-like behavior of 𝐐Leff\mathbf{Q}^{\mathrm{eff}}_{L} is simply due to the finite-size discretization of the wavevector (e.g., ΔQL,3eff/(2π)=0.003125\Delta Q^{\mathrm{eff}}_{L,3}/(2\pi)=0.003125 for Lc=320L_{c}=320), which also causes spurious peaks in CC at wildly size-dependent temperatures (Fig. 3). Considering the wide range of the system sizes we investigated, the most natural interpretation is that the change of the wavevector in thermodynamic limit, 𝐐(T)=limL𝐐Leff(T)\mathbf{Q}(T)=\lim_{L\to\infty}\mathbf{Q}^{\mathrm{eff}}_{L}(T), is (quasi-)continuous towards 𝐪com\mathbf{q}_{\text{com}}. When 𝐐Leff(T)\mathbf{Q}^{\mathrm{eff}}_{L}(T) changes from one value to another as a function of TT in a finite-size system, a disordering effect appears at large distance due to the mismatch between an ideal wavevector and the system size. The mismatch-induced disordering effect causes a spurious spiky peak in U𝐐LeffU_{\mathbf{Q}^{\mathrm{eff}}_{L}} (Fig. 4) and interfere with high-precision determination of TSDWT_{\mathrm{SDW}} as the behavior of U𝐐LeffU_{\mathbf{Q}^{\mathrm{eff}}_{L}} near the crossing point is affected and becomes less systematic. Therefore, we treat the first crossing points of U𝐐LeffU_{\mathbf{Q}^{\mathrm{eff}}_{L}} for (Lc,2Lc)=(80,160),(120,240),and(160,320)(L_{c},2L_{c})=(80,160),\,(120,240),\,\text{and}\,(160,320) simply on an equal footing, which nonetheless allows us to determine the transition point of the SDW ordering to the satisfactory precision, e.g., TSDW/|J1|=1.407(5), 1.406(1),and 1.41(2)T_{\mathrm{SDW}}/\lvert{J_{1}}\rvert=1.407(5),\,1.406(1),\,\text{and}\,1.41(2) for h/|J1|=0,0.05,and 0.2h/\lvert{J_{1}}\rvert=0,0.05,\,\text{and}\,0.2, respectively (Fig. 4). The estimated TSDWT_{\mathrm{SDW}} roughly coincides with the highest-TT peak in the specific heat (see Fig. 3).

At low temperatures, the ordering wavevector is pinned at 𝐪com=(0,0,2π)\mathbf{q}_{\text{com}}=(0,0,2\pi), corresponding to a three-sublattice ordered phase. At low fields, the low-TT phase appears through a lock-in IC-C transition, where the translational symmetry along the cc axis, which is broken in the SDW state, is restored. By analyzing the behavior of U𝐪comU_{\mathbf{q}_{\text{com}}}, we find that TICCT_{\mathrm{IC-C}} increases rapidly with hh (see Fig. 2), e.g., TICC/|J1|=0.52(1), 1.15(3),and 1.397(3)T_{\mathrm{IC-C}}/\lvert{J_{1}}\rvert=0.52(1),\,1.15(3),\,\text{and}\,1.397(3) for h/|J1|=0,0.05,and 0.2h/\lvert{J_{1}}\rvert=0,0.05,\,\text{and}\,0.2, respectively. Since TSDWT_{\mathrm{SDW}} is nearly constant in hh, the SDW phase shrinks rather rapidly with increasing hh, and above a magnetic field-induced multicritical point hLP0.2|J1|0.17hsath_{\mathrm{LP}}\simeq 0.2\lvert{J_{1}}\rvert\simeq 0.17h_{\mathrm{sat}}, where hsat=6(J2+J3)h_{\mathrm{sat}}=6(J_{2}+J_{3}) is the saturation field, no incommensurate phase exists. The estimated IC-C transition temperatures coincide with the temperatures at which the finite-size ordering wavevector reaches 𝐐Leff=𝐪com\mathbf{Q}^{\mathrm{eff}}_{L}=\mathbf{q}_{\text{com}} at each field, as expected. Also, |M𝐪com|2\langle{\lvert{M_{\mathbf{q}_{\text{com}}}}\rvert^{2}}\rangle increases rapidly around the estimated temperatures and becomes almost size-independent at lower temperatures (Fig. 5). While the IC-C transitions studied in the present MC work exhibit typical features of a first-order transition, such as the correction in the finite-size transition temperature varying as 1/L3\sim 1/L^{3}, they may be a spurious behavior caused by discrete wavevectors in finite size systems, as also found in specific heat as the spurious spiky peaks.

For T<TICCT<T_{\mathrm{IC-C}}, the two main candidates for the three-sublattice order are the ferrimagnetic (FIM) state and the partially-disordered antiferromagnetic (PDA) state, similar to the case in triangular lattice antiferromagnetic Ising models.Blankschtein84; Coppersmith85; Isakov03; Heinonen89; Bunker93; Isakov03; Lin14 In the FIM state, each spin chain has a ferromagnetic order and different spin chains takes the three-sublattice \uparrow\uparrow\downarrow structure. In the PDA phase, the spin chains in the first and the second sublattices have a ferromagnetic order with spin-\uparrow and \downarrow, respectively, with the spin chains in the third sublattice disordered. A convenient indicator for a finite-size calculation is

C6=M𝐪com6|M𝐪com|6,\displaystyle C_{6}=\frac{\langle{M^{6}_{\mathbf{q}_{\text{com}}}}\rangle}{\bigl{\langle}{\lvert{M_{\mathbf{q}_{\text{com}}}}\rvert^{6}}\bigr{\rangle}}, (9)

which takes C6>0C_{6}>0 (C6<0C_{6}<0) for the FIM (PDA) state.Isakov03 For h=0h=0, we confirm the PDA state below TICCT_{\mathrm{IC-C}} (Fig. 5), though the result should be distinguished from the previous claim of the same state in CCO below \simeq 25 K.Kageyama97 We find no evidence of an additional order-order transition at low TT as reported in CCO,Agrestini11 at least for T0.45|J1|T\gtrsim 0.45\lvert{J_{1}}\rvert. For h0h\neq 0, we find the FIM state at low TT down to h/|J1|=0.025h/\lvert{J_{1}}\rvert=0.025, implying that the observed PDA state is extremely fragile against the magnetic field. The FIM phase yields the 1/3 magnetization plateau. Nekrashevic21 At high fields, we find a direct transition into the FIM phase without the intervening SDW phase. We find that the transition is of the first order, which is consistent with the Z3Z_{3} symmetry breaking in d=3d=3, as in the three-state Potts model.Blote1979; Janke97

V Ginzburg-Landau theory

Finally, for a more universal description of the commensurate and incommensurate phases in CCO and similar materials, we consider a GL theory. As will be shown below, the theory is essentially in the same form as that for the ANNNI model.Aharony1981; Bak80 Interestingly, however, the GL theory for ANNNI-like models in magnetic fields Yokoi81 has yet been thoroughly discussed in the literature, despite its experimental relevance.

We use a complex order parameter ψ(𝐫)ei𝐪com𝐫σz(𝐫)\psi(\mathbf{r})\sim e^{-i\mathbf{q}_{\text{com}}\cdot\mathbf{r}}\sigma^{z}(\mathbf{r}), 𝐪com=(0,0,2π)\mathbf{q}_{\text{com}}=(0,0,2\pi), which can be formally introduced by using the Hubbard-Stratonovich transformation. ψ(𝐫)\psi(\mathbf{r}) describes the local three-sublattice order in a coarse-grained way, which can be either the FIM order or the PDA order depending on the phase factor. For h=0h=0, we find the following GL Hamiltonian,

h=0\displaystyle\mathscr{H}_{\,h=0} =d3𝐫[12|(iϵ)ψ(𝐫)|2+t|ψ(𝐫)|2+u4|ψ(𝐫)|4\displaystyle=\int\!\mathrm{d}^{3}\mathbf{r}\,\Bigl{[}\,\frac{1}{2}\bigl{\lvert}{\left(\nabla-i\bm{\epsilon}\right)\psi(\mathbf{r})}\bigr{\rvert}^{2}+t\lvert{\psi(\mathbf{r})}\rvert^{2}+u_{4}\lvert{\psi(\mathbf{r})}\rvert^{4}
+u6|ψ(𝐫)|6+v6(ψ(𝐫)6+ψ(𝐫)6)],\displaystyle\hskip 60.0pt+u_{6}\lvert{\psi(\mathbf{r})}\rvert^{6}+v_{6}\bigl{(}\psi(\mathbf{r})^{6}+\psi^{\ast}(\mathbf{r})^{6}\bigr{)}\,\Bigr{]}, (10)

where tt, u4u_{4}, u6u_{6}, and v6v_{6} are GL coefficients and ϵ=(0,0,ϵ)\bm{\epsilon}=(0,0,\epsilon). A crucial point of the present theory is that the wavevector 𝐪com\mathbf{q}_{\text{com}} of the local order described by ψ(𝐫)\psi(\mathbf{r}) is slightly shifted from the minima 𝐪=±𝐪min=±(𝐪com+ϵ)\mathbf{q}=\pm\mathbf{q}_{\text{min}}=\pm(\mathbf{q}_{\text{com}}+\bm{\epsilon}) of the Fourier transform J(𝐪)J(\mathbf{q}) of the exchange interactions, as briefly mentioned before. For this reason, h=0\mathscr{H}_{\,h=0} includes the vector potential-like (but constant) contribution iϵ-i\bm{\epsilon} in the gradient term. h=0\mathscr{H}_{\,h=0} also includes the six-order term (v6v_{6}) as the leading Umklapp term, the order of which is determined by the size of the magnetic sublattice of the commensurate order and time reversal symmetry; a factor of three comes from 3𝐪com03\mathbf{q}_{\text{com}}\equiv 0 and time reversal symmetry requires another factor of two.

As is clear from the origin, the gradient term acts as adding a momentum ϵ\bm{\epsilon} to ψ\psi, in favor of the three-sublattice long-wavelength SDW state ψ(𝐫)(const.)×eiϵ𝐫\langle{\psi(\mathbf{r})}\rangle\sim\text{(const.)}\times e^{i\bm{\epsilon}\cdot\mathbf{r}}. Although the SDW state thus appears to benefit from the exchange energy, the sinusoidal modulation made of localized Ising-like moments also requires entropy contribution to the free energy. In contrast, because of the shift of 𝐪com\mathbf{q}_{\text{com}} from the minima of J(𝐪)J(\mathbf{q}), the commensurate three-sublattice ordered state, ψ(𝐫)const.\langle{\psi(\mathbf{r})}\rangle\sim\text{const.}, may appear not to acquire the full energy gain of the exchange interaction. However, the state is quite compatible with the Ising anisotropy. In the GL theory (10), the Umklapp term plays the role of the entropy contribution related with the Ising anisotropy. While the commensurate state may be favored by the Umklapp term by adjusting its constant phase factor, the incommensurate SDW states generally gain no corresponding contribution because of the phase cancellation in the integral over the space. In fact, we could rederive the GL theory in terms of ϕc(𝐫)σz(𝐫)ei𝐪min𝐫\phi_{c}(\mathbf{r})\sim\sigma^{z}(\mathbf{r})\,e^{-i\mathbf{q}_{\text{min}}\cdot\mathbf{r}} instead of ψ(𝐫)σz(𝐫)ei𝐪com𝐫\psi(\mathbf{r})\sim\sigma^{z}(\mathbf{r})\,e^{-i\mathbf{q}_{\text{com}}\cdot\mathbf{r}}, and the result is an ordinary ϕc4\phi_{c}^{4}-theory for the one-component complex order parameter without Umklapp terms if 𝐪min\mathbf{q}_{\text{min}} is incommensurate. Hence, the key role in the GL theory (10) is played by the competition between the gradient and the Umklapp terms favoring incommensuration and commensuration, respectively. Thus, although somewhat different in appearance, the Hamiltonian of this system (1) realizes essentially the same kind of situation as the classic ANNNI model.Bak80

At T=TSDWT=T_{\mathrm{SDW}}, critical fluctuations renders ψ(𝐫)\psi(\mathbf{r}) nonzero with the additional momentum ϵ\bm{\epsilon}, resulting in the SDW state with the ordering wavevector 𝐐=𝐪com+ϵ=𝐪min\mathbf{Q}=\mathbf{q}_{\text{com}}+\bm{\epsilon}=\mathbf{q}_{\text{min}}. In other words, we expect a condensation of the softest mode ϕc(𝐫)σz(𝐫)ei𝐪min𝐫\phi_{c}(\mathbf{r})\sim\sigma^{z}(\mathbf{r})\,e^{-i\mathbf{q}_{\text{min}}\cdot\mathbf{r}} rather than ψ(𝐫)σz(𝐫)ei𝐪com𝐫\psi(\mathbf{r})\sim\sigma^{z}(\mathbf{r})\,e^{-i\mathbf{q}_{\text{com}}\cdot\mathbf{r}}. Since the Umklapp v6v_{6} term has no effect for the incommensurate SDW states, the transition, which breaks translation symmetry along the cc axis, will be in the 3D XY universality class (emergent U(1) symmetry). The observed main peak of the specific heat, which exhibits a sign of smearing [Fig. 3(a)], is consistent with the negative exponent α=0.0146(8)<0\alpha=-0.0146(8)<0 for the XY universality class.Campostrini2001

For T<TSDWT<T_{\mathrm{SDW}}, the competition between the gradient term and the Umklapp term sets in, which affects the phase factor of ψ(𝐫)\langle{\psi(\mathbf{r})}\rangle, thereby 𝐐(T<TSDW)\mathbf{Q}(T<T_{\mathrm{SDW}}). To see this, we may write ψ(𝐫)=A(𝐫)eiθ(𝐫)\psi(\mathbf{r})=A(\mathbf{r})e^{i\theta(\mathbf{r})} and apply a mean-field decoupling for the massive amplitude fluctuation (“Higgs”) mode δA(𝐫)=A(𝐫)A\delta A(\mathbf{r})=A(\mathbf{r})-\langle{A}\rangle and the phase mode θ(𝐫)\theta(\mathbf{r}).Bak80 The result is the following sine-Gordon model for θ(𝐫)\theta(\mathbf{r}),

h=0,θ=A2d3𝐫[12(θ(𝐫)ϵ)2+2A4v6cos6θ(𝐫)].\displaystyle\mathscr{H}^{\prime}_{\,h=0,\,\theta}=\langle{A}\rangle^{2}\!\int\!\mathrm{d}^{3}\mathbf{r}\,\Bigl{[}\frac{1}{2}\left(\nabla\theta(\mathbf{r})-\bm{\epsilon}\right)^{2}+2\langle{A}\rangle^{4}v_{6}\cos 6\theta(\mathbf{r})\Bigr{]}. (11)

The gradient term tends to drift the phase, which can lead to a plethora of soliton lattices through the competition against the cosine term,Bak80; Chaikin-Lubensky1995 in good agreement with our mean field and MC studies. In the meantime, because the order parameter amplitude A\langle{A}\rangle increases as TT is lowered below TSDWT_{\mathrm{SDW}}, the strength of the cosine term is enhanced. Consequently, the model at low TT is expected to undergo a lock-in transition eventually. For v6>0v_{6}>0 (v6<0v_{6}<0), the phase is locked-in at θ=2nπ/6\theta=2n\pi/6 [θ=(2n+1)π/6\theta=(2n+1)\pi/6] with an integer 0n<60\leq n<6, corresponding to the FIM (\uparrow\uparrow\downarrow or \downarrow\downarrow\uparrow) state and the PDA state,Blankschtein84; Coppersmith85; Isakov03; Heinonen89; Bunker93; Lin14 respectively. Although our mean field calculation implies v6>0v_{6}>0 while our unbiased MC simulation suggests v6<0v_{6}<0 in zero field, the subtle discrepancy does not require a serious attention as v6v_{6} is generated through fluctuations.

For h0h\neq 0, the uniform 𝐪=0\mathbf{q}=0 component m(𝐫)m(\mathbf{r}) is allowed by symmetry and may be induced by the magnetic field. Consequently, in addition to v6v_{6}, a lower-order Umklapp term appears in the GL Hamiltonian. We find

Δ 0𝐪com\displaystyle\Delta\mathscr{H}_{\,0\mathrm{-}\mathbf{q}_{\text{com}}} d3𝐫w4m(𝐫)(ψ(𝐫)3+ψ(𝐫)3)\displaystyle\simeq\int\mathrm{d}^{3}\mathbf{r}\,w_{4}m(\mathbf{r})\left(\psi(\mathbf{r})^{3}+\psi^{\ast}(\mathbf{r})^{3}\right) (12)

as the leading-order contribution with the new coupling constant w4w_{4}. The total effective GL Hamiltonian for h0h\neq 0 is

h0,ψ,m=h=0,ψ+𝐪=0,m+Δ 0𝐪com,\displaystyle\mathscr{H}_{\,h\neq 0,\,\psi,\,m}=\mathscr{H}_{\,h=0,\,\psi}+\mathscr{H}_{\,\mathbf{q}=0,\,m}+\Delta\mathscr{H}_{\,0\mathrm{-}\mathbf{q}_{\text{com}}}, (13)

where 𝐪=0,m=d3𝐫[12(cm(𝐫))2+μ2m(𝐫)2hm(𝐫)]\mathscr{H}_{\,\mathbf{q}=0,\,m}=\int\!\mathrm{d}^{3}\mathbf{r}\,\bigl{[}\frac{1}{2}\bigl{(}c\nabla m(\mathbf{r})\bigr{)}^{2}+\mu^{2}m(\mathbf{r})^{2}-hm(\mathbf{r})\bigr{]} is the noninteracting part for m(𝐫)m(\mathbf{r}) with c>0c>0 and μ2>0\mu^{2}>0 being the gapped spin wave parameters near 𝐪=0\mathbf{q}=0. By a similar mean field decoupling as in the h=0h=0 case, we find a new term in the sine-Gordon model,

Δ 0𝐪com,θ=2w4mA3d3𝐫cos3θ(𝐫).\displaystyle\Delta\mathscr{H}^{\prime}_{\,0\mathrm{-}\mathbf{q}_{\text{com}},\,\theta}=2w_{4}\langle{m}\rangle\langle{A}\rangle^{3}\!\int\mathrm{d}^{3}\mathbf{r}\,\cos 3\theta(\mathbf{r}). (14)

The field-induced Umklapp term has the following consequences. Firstly, because of the reduced symmetry in the θ\theta space, only a subset of the FIM states is favored by Δ 0𝐪com,θ\Delta\mathscr{H}^{\prime}_{\,0\mathrm{-}\mathbf{q}_{\text{com}},\,\theta} among the three-sublattice ordered states. Depending on the sign of w4w_{4}, the favored states are \uparrow\uparrow\downarrow or \downarrow\downarrow\uparrow, each of which is three-fold degenerate, though \uparrow\uparrow\downarrow is naturally anticipated for h>0h>0. Secondly, as the prefactor is A3\propto\langle{A}\rangle^{3}, as opposed to A6\propto\langle{A}\rangle^{6} in zero field, the strength of the field-induced Umklapp term is expected to grow faster for T<TSDWT<T_{\mathrm{SDW}}. Moreover, since the prefactor is m\propto\langle{m}\rangle, we expect that this trend is further enhanced for larger hh. Therefore, the region of the incommensurate SDW microphases is expected to become narrower for larger hh, in excellent agreement with our MC results (Fig. 2). Thirdly, considering that our MC simulation shows the PDA state at h=0h=0 below TICCT_{\mathrm{IC-C}}, the zero- and the field-induced Umklapp terms (cos6θ,cos3θ\sim\cos 6\theta,\,\cos 3\theta, respectively) must compete against each other in the present system. The competition opens a possibility of a kind of mixed phase below TICCT_{\mathrm{IC-C}}, though our MC results shows no evidence down to an extremely low field, h/|J1|=0.025h/\lvert{J_{1}}\rvert=0.025. Finally, near T=TSDWT=T_{\mathrm{SDW}}, the present GL theory suggests the condensation of the softest mode ϕc(𝐫)σz(𝐫)ei𝐪min𝐫\phi_{c}(\mathbf{r})\sim\sigma^{z}(\mathbf{r})\,e^{-i\mathbf{q}_{\text{min}}\cdot\mathbf{r}} rather than ψ(𝐫)σz(𝐫)ei𝐪com𝐫\psi(\mathbf{r})\sim\sigma^{z}(\mathbf{r})\,e^{-i\mathbf{q}_{\text{com}}\cdot\mathbf{r}}, as in the case of zero magnetic field. Hence, we expect the emergent U(1) symmetry at T=TSDWT=T_{\mathrm{SDW}} also for h0h\neq 0, because Umklapp terms disappear from the GL theory in terms of the incommensurate critical mode ϕc\phi_{c}. The theory thus predicts 𝐐(T=TSDW)=𝐪min\mathbf{Q}(T=T_{\mathrm{SDW}})=\mathbf{q}_{\text{min}} for any magnetic field, which is indeed consistent with our MC results at low magnetic fields. However, the simulation closer to the magnetic field-induced multicritical point hLP0.2|J1|h_{\mathrm{LP}}\simeq 0.2\lvert{J_{1}}\rvert might point to a deviation from this behavior, suggesting that 𝐐(T=TSDW)\mathbf{Q}(T=T_{\mathrm{SDW}}) approaches towards 𝐪com\mathbf{q}_{\text{com}} for the larger systems we investigated [Fig. 3(c)]. This observation may be an indication that the multicritical point is a Lifshitz point induced by a magnetic field.

VI Summary and outlook

To summarize, we presented the magnetic phase diagram in equilibrium of the 3D spin model for the frustrated quasi-one-dimensional triangular Ising antiferromagnet Ca3Co2O6. We identified the region of incommensurate SDW microphases in a magnetic field (Fig. 2). We found the deformation of SDW microphases as a function of TT, characterized by the temperature dependence of the ordering wavevector 𝐐(T)\mathbf{Q}(T), occurring much more rapidly in a magnetic field than in zero field (Fig. 3). The deformation eventually leads to the IC-C transition into the PDA (FIM) state for h=0h=0 (h0h\neq 0). Between the PDA and the FIM phases, there may be a mixed phase in an extremely low-field regime, though not confirmed in this work. The GL theory we derived includes different symmetry-allowed Umklapp terms for h=0h=0 and h0h\neq 0. The GL theory allowed for further deriving an effective sine-Gordon model that provides a qualitative explanation of the observed magnetic field-induced deformation of the SDW microphases. Moreover, these effective theories demonstrate that the present system can be seen as an incarnation of the classic ANNNI model,Bak80 despite different appearance of the lattice structure and the complicated network of the exchange interactions.

Finally, we discuss the relation between the theoretical phase diagram in this work and the previous experiments. As mentioned in Introduction, the material is known for the intriguing combination of the slow relaxation phenomena and the long-wavelength SDW order. As the recent field-cooling study suggested,Nekrashevic21 the slow relaxation at low temperature may be greatly influenced by the cooling process passing through the low-field SDW phase at intermediate temperature. To verify the conjectured relation experimentally, the challenge is that experiments under equilibrium conditions are known to be notoriously difficult for Ca3Co2O6. For example, a resonant X-ray experiments reported a field-induced IC-C transition at 5 K,Mazzoli09 which is unfortunately most likely non-equilibrium because the temperature is too low. However, at intermediate temperatures above TSFT_{\mathrm{SF}}, there are some experiments that seem to capture the desired physics of the field-induced deformation of the SDW phase and the IC-C transition. For example, a μ\muSR measurements at 20 K reported a magnetic field-induced phase transition at around 0.4 T.Takeshita2007 Although the original interpretation of the result suggested a PDA-FIM transition, the obtained phase boundary is very similar to the SDW-FIM transition line shown in the present work. Since the SDW phase was not confirmed back then, it is quite possible that the anomaly in the μ\muSR experiment is the sign of the IC-C transition induced by the magnetic field. We also note that a similar phase diagram was obtained also by weak anomaly in the magnetic entropy change.Lampen14 We thus believe that further experiments studying the field-induced IC-C transition in Ca3Co2O6, such as neutron scattering and other spectroscopies focusing on the low-field regime, will be very promising, especially when combined with the recently proposed field-cooling protocol.Nekrashevic21 Such experiments may provide further insights in the peculiar slow dynamics and out-of-equilibrium behaviors in Ca3Co2O6 and related materials.

Acknowledgements.
The author is grateful to valuable discussions with Vivien Zapf, Xiaxin Ding, Ivan Nekrashevich, and Cristian Batista. The author acknowledges the support by the NSFC (Nos. 11950410507, 12074246, and U2032213) and MOST (No. 2016YFA0300501) research programs.

References