Multiple Incommensurate Magnetic States in the Kagome Antiferromagnet Na2Mn3Cl8
Abstract
The kagome lattice can host exotic magnetic phases arising from frustrated and competing magnetic interactions. However, relatively few insulating kagome materials exhibit incommensurate magnetic ordering. Here, we present a study of the magnetic structures and interactions of antiferromagnetic Na2Mn3Cl8 with an undistorted Mn2+ kagome network. Using neutron-diffraction and bulk magnetic measurements, we show that Na2Mn3Cl8 hosts two different incommensurate magnetic states, which develop at K and K. Magnetic Rietveld refinements indicate magnetic propagation vectors of the form , and our neutron-diffraction data can be well described by cycloidal magnetic structures. By optimizing exchange parameters against magnetic diffuse-scattering data, we show that the spin Hamiltonian contains ferromagnetic nearest-neighbor and antiferromagnetic third-neighbor Heisenberg interactions, with a significant contribution from long-ranged dipolar coupling. This experimentally-determined interaction model is compared with density-functional-theory simulations. Using classical Monte Carlo simulations, we show that these competing interactions explain the experimental observation of multiple incommensurate magnetic phases and may stabilize multi-q states. Our results expand the known range of magnetic behavior on the kagome lattice.
I Introduction
Geometrical frustration—the inability of a system to satisfy all of its pairwise interactions simultaneously—can suppress conventional magnetic ordering and promote exotic magnetic states (Balents_2010, ). A focus of frustrated-magnetism research has been insulating materials in which magnetic ions occupy a kagome lattice of corner-sharing triangles, where strong frustration effects can occur if the interactions are antiferromagnetic. For example, if antiferromagnetic Heisenberg interactions couple neighboring spins only, a spin-liquid state is stable down to extremely low temperatures even in the classical limit (Chalker_1992, ), before eventually undergoing octupolar magnetic ordering (Zhitomirsky_2008, ). There is a continuing search for real materials that are candidates to realize frustrated kagome magnetism (Norman_2016, ). In the quantum () limit, notable candidates include herbertsmithite (Vries_2009, ; Han_2012, ) and barlowite (Han_2014, ; Pasco_2018, ; Tustain_2018, ). In the classical (large-) limit, probably the most studied candidates are iron-containing jarosite minerals (Wills_1996, ; Inami_2000, ; Nishiyama_2003, ), which are often off-stoichiometric (Janas_2020, ; Bisson_2008, ). Therefore, an important goal is to identify and characterize other structure types containing kagome lattices, particularly those with antiferromagnetic interactions, and where the kagome lattice is structurally undistorted.
Kagome antiferromagnets that exhibit long-range magnetic ordering may still show strong effects of geometrical frustration. In particular, the inclusion of further-neighbor interactions can stabilize several unusual magnetic states instead of conventional collinear antiferromagnetism. These states include noncollinear order as well as many noncoplanar states, which are more stable than collinear antiferromagnets in large regions of interaction space (Messio_2011, ). For certain exchange interactions, incommensurate magnetic ordering can also be stabilized; however, the nature of the incommensurate phase is difficult to determine from simulations (Grison_2020, ; Li_2022, ). Experimental studies of materials that occupy this part of the interaction space are therefore important to advance our understanding of kagome magnetism.

In this context, we identified Na2Mn3Cl8 as a promising material for frustrated magnetism on the kagome lattice. This material was first reported in the 1970s (Loon_1975, ) and is likely electrically insulating (Devlin_2022, ), but its magnetic structure and interactions have not previously been studied. Nevertheless, a recent materials survey highlighted Na2Mn3Cl8 as a candidate frustrated antiferromagnet (Meschke_2021, ). Due to the large magnetic moment of Mn2+with and the absence of an orbital contribution (), its behavior is expected to be predominantly classical. The reported crystal structure (Loon_1975, ) is shown in Figure 1(a); its trigonal symmetry (space group ) ensures that the kagome planes are undistorted. A recent investigation of its bulk magnetic properties showed multiple magnetic phase transitions below K, and the possibility of a low-temperature structural phase transition was suggested due to the observation of a broad specific-heat anomaly around 6 K (Devlin_2022, ). Notably, a structural transition to a trimerized polar phase is observed in the related kagome magnet Na2Ti3Cl8 (Hinz_1995, ; Hanni_2017, ; Kelly_2019, ; Paul_2020, ; Khomskii_2021, ).
In this paper, we report magnetic characterization and powder neutron-diffraction experiments on Na2Mn3Cl8. In agreement with a recent report (Devlin_2022, ), we observe that this material undergoes two magnetic phase transitions with decreasing temperature. However, our data do not indicate a measurable crystallographic distortion at temperatures down to K, indicating that the undistorted kagome lattice is preserved. Our powder neutron-diffraction measurements show that, unusually, the two ordered magnetic states both have incommensurate magnetic propagation vectors. These data are consistent with single- helical magnetic ordering, with an antiferromagnetic stacking of kagome layers. We show that the development of multiple incommensurate phases can be explained by a model including Heisenberg exchange interactions up to third-nearest neighbors and the long-ranged dipolar interaction, and we estimate the values of the exchange interactions by analyzing the magnetic diffuse scattering measured above . Our results place Na2Mn3Cl8 in a complex region of the kagome phase space, in which incommensurate ordering is stabilized by a competition between short-range ferromagnetic and longer-range antiferromagnetic interactions. Our interaction model also suggests that Na2Mn3Cl8 deserves further investigation as a potential host of multi- spin textures in zero applied magnetic field.
Our paper is structured as follows. We first introduce the crystal structure and potential magnetic exchange pathways of Na2Mn3Cl8, and present thermomagnetic measurements of the bulk magnetic properties. We then discuss our powder neutron-diffraction data and symmetry-informed Rietveld analysis, from which the likely single- magnetic structures are determined. Magnetic diffuse-scattering analysis is employed to parametrize the magnetic interactions that stabilize incommensurate ordering. We compare and contrast our experimental results with density-functional-theory calculations. Finally, we discuss the extent to which our experimental observations can be rationalized using field-theoretical and Monte Carlo simulations, and conclude by summarizing our results and highlighting opportunities for future research.
II Methods
II.1 Sample synthesis
A polycrystalline sample (mass g) of Na2Mn3Cl8 was prepared by sealing a stoichiometric mixture MnCl2 and NaCl in SiO2 after heating at 250 ∘C under dynamic vacuum overnight. The mixture was heated to 750 ∘C for several hours and quenched by removing from the furnace. The sample was ground, sealed with -atm argon, and annealed at ∘C for at total of h with an additional intermediate grinding. All handling of this very air-sensitive sample was conducted in an inert-atmosphere glovebox, and the samples were kept under inert atmosphere when they were transferred to the vacuum lines for sealing of the silica tubes.
II.2 Experimental measurements
Magnetization measurements were performed using Quantum Design magnetometers with data below K collected using a 3He insert. The samples were loaded into measurement straws in an inert-atmosphere glovebox with grease to protect the powders from air during the rapid loading process.
Neutron-diffraction measurements were performed using the HB-2A powder diffractometer at the High Flux Isotope Reactor of Oak Ridge National Laboratory. The incident neutron wavelength . Our powder sample of mass g was loaded into a 4-mm-diameter cylindrical Cu container in a He glovebox. The sample was cooled using a cryostat with a 3He insert, affording a base temperature of K. Counting times were hr at 0.3, 0.8, 2.0, 5.0, and 40 K, and hr at other temperatures below . The data were corrected for neutron absorption by the sample (Hewat_1979, ).
II.3 Magnetic diffuse scattering refinements and field theory
Magnetic diffuse-scattering refinements were performed using the Spinteract program to refine the values of the exchange interactions (Paddison_2022a, ). The spin Hamiltonian included Heisenberg exchange interactions and the magnetic dipolar interaction (see Section III.5). The input data were collected at K and K and were placed in absolute intensity units (barn sr-1 Mn-1) by normalization to the nuclear Bragg profile. A high-temperature ( K) data set was subtracted from these data.
The magnetic diffuse scattering and bulk susceptibility were calculated using Onsager reaction-field theory (Paddison_2022a, ; Logan_1995, ; Brout_1967, ), and a K calculation was subtracted from the calculated . In this approach, the Fourier transform of the magnetic interactions is calculated as , where denote Cartesian spin components, denote sites within the primitive unit cell, and is the vector connecting unit cells containing sites and . The interaction matrix formed by the is diagonalized on a grid of up to points in the first Brillouin zone to determine its eigenvalues and eigenvector components ,
where indexes the normal modes. The long-range dipolar interaction is included using Ewald summation (Enjalran_2004, ). Within a reciprocal-space mean-field approximation, the magnetic propagation vector of the first ordered state is the wavevector at which reaches its maximal value. The and are given in terms of the and , as described in Ref. (Paddison_2022a, ).
During the refinements, we minimized the function
(1) |
where subscript “expt” and “calc” indicate measured and calculated diffuse scattering patterns, respectively, is an experimental uncertainty, and is a refined overall scale factor common to the neutron-scattering data and the magnetic susceptibility . The minimization was performed using the Minuit program (James_1975, ; James_1994, ). To identify local minima in , we performed refinements for each model, with different randomly-chosen initial parameter values in each case.
II.4 Rietveld refinements
Rietveld refinements were performed using the Fullprof software (Rodriguez-Carvajal_1993, ; Rodriguez-Carvajal_1993a, ). A crystal-structure refinement was first performed at K (). In addition to the crystallographic parameters given in Table 1, we refined the intensity scale factor, zero-offset, peak-shape, and background parameters. The peak shape was modeled using a pseudo-Voigt function initialized with the instrument resolution parameters, with , , and parameters subsequently refined. The background was fitted using Chebychev polynomials.
Magnetic Rietveld refinements were performed against K and K data from which the K data had been subtracted. This subtraction isolates the magnetic Bragg signal by subtracting the nuclear and background contributions, which are essentially unchanged between and K. In the magnetic refinements, asymmetry, Chebychev background, and magnetic-structure parameters were refined, as described in Section III.4; all other parameters were fixed at the values obtained from the K refinement. Magnetic-structure figures were prepared using the Vesta program (Momma_2008, ).
II.5 Density-functional-theory calculations
Density functional theory calculations were performed using the all-electron-density functional code Wien2K (Sjostedt_2000, ; Blaha_2001, ). The linearized augmented plane wave method (Singh_2006, ) and the generalized-gradient approximation of Perdew, Burke, and Ernzerhof (Perdew_1996, ) were utilized. The generated by the smallest linearized augmented plane wave sphere radius () and the interstitial plane-wave cutoff () was set as 7.0 for good convergence. The muffin-tin radii of Na, Cl, and Mn atoms were a.u., a.u., and a.u., respectively. The number of -points in the full Brillouin zone was . Lattice parameters of Na2Mn3Cl8 were fixed to the experimental values of Å and Å. Then, the internal atomic coordinates were relaxed until forces on all of the atoms were less than mRy/bohr, with non-magnetic, ferromagnetic, and interlayer antiferromagnetic states. It turns out the relaxed crystal structure with a ferromagnetic state is highly similar to the experimental crystal structure. However, in the non-magnetic state, the atomic coordination changes significantly, as Cl atoms moves towards the Mn layers. The relaxed antiferromagnetic crystal structure is same as the ferromagnetic crystal structure, but exhibits lower energy and smaller forces. The energy difference between ferromagnetic and antiferromagnetic states is meV/f.u.. We therefore used the relaxed antiferromagnetic structure to calculate the intralayer and interlayer magnetic couplings.
III Experimental Results
III.1 Crystal structure refinement
Na2Mn3Cl8, K | ||
---|---|---|
, | ||
Site | Wyckoff | |
Na | ||
Mn | ) | |
Cl1 | ||
Cl2 | ) |
The crystal structure of Na2Mn3Cl8 is shown in Figure 1(a), and comprises of triangular Na+ layers separating kagome Mn2+ layers (Loon_1975, ; Devlin_2022, ). We performed Rietveld refinements against our K and K neutron-diffraction data to investigate the possibility of a crystallographic distortion from the published structure (space group ). Good agreement was obtained with the published structural model (Loon_1975, ) at both temperatures, except for two very weak peaks at and that were not accounted for, and were unchanged between and K. Since these peaks could not be explained by simple multiples of the crystallographic unit cell, or by possible impurity phases (NaCl, MnCl2, or NaMnCl3), we concluded that the sample or its environment contained a small fraction of unknown impurity. Our results do not show evidence for any structural phase transition between K and K, indicating that the broad K specific-heat anomaly reported previously (Devlin_2022, ) is probably due to magnetic ordering of a minor NaMnCl3 impurity phase (a possibility noted in Ref. (Devlin_2022, )).
Each nearest-neighbor Mn–Mn bond is bridged by a Cl1 ion and a Cl2 ion, which provide the nearest-neighbor superexchange pathways. The Mn–Cl1–Mn and Mn–Cl2–Mn bond angles are and , respectively. Since these values are close to , the Goodenough-Kanamori rules predict weak ferromagnetic nearest-neighbor exchange interactions. Further-neighbor interactions have more complicated pathways and, consequently, are difficult to predict. In particular, there are two inequivalent third-neighbor exchange pathways with the same interatomic separation [Figure 1(c)].
III.2 Thermomagnetic measurements

Our high-temperature bulk magnetic susceptibility measurements and Curie-Weiss fits are shown in Figure 2(a). They reveal an effective magnetic moment of , close to the spin-only value of for Mn2+, and a Weiss temperature of K, indicating net antiferromagnetic interactions. An anomaly is observed at K in our low-temperature magnetic susceptibility data, which is suppressed to lower temperature with increasing applied field, consistent with a magnetic ordering transition [Figure 2(b)]. The “frustration parameter”, , indicates a relatively small degree of frustration. Since the nearest-neighbor kagome antiferromagnet is highly frustrated, this result hints at the presence of significant further-neighbor couplings or anisotropies; however, the nature of these couplings cannot be determined from bulk characterization data alone. Interestingly, and consistent with Ref. (Devlin_2022, ), we also observe a second magnetic-susceptibility anomaly at K [Fig. 2(b)], suggesting a multi-stage magnetic ordering process. Such behavior is unusual and hints that, despite the relatively small value of , the frustrated topology of the kagome lattice may cause several magnetic structures to be nearly degenerate. Figure. 2(c) shows the low temperature field dependence of the magnetization, which does not follow the Brillouin function, in qualitative agreement for theoretical predictions for the kagome lattice (Nakano_2015, ). Figure 1(a) shows the field derivative of the magnetization, . Several anomalies are observed in below at small applied fields, suggesting that the magnetic ground state is fragile to external perturbations.
III.3 Overview of neutron data

We performed neutron-diffraction measurements to obtain microscopic insight into the magnetic interactions and structures of Na2Mn3Cl8 (see Methods). An overview of the temperature dependence of our neutron data is shown in Figure 3(a). Several new Bragg peaks appear below K, most prominently at wavevectors of approximately and . We identify these as magnetic Bragg peaks arising from the onset of long-range magnetic ordering, since they appear at the same temperature as the magnetic-susceptibility anomaly at . Interestingly, the positions of the magnetic Bragg peaks suddenly shift at K, revealing that the second phase transition involves a change in magnetic propagation vector. At temperatures above , broad magnetic diffuse scattering features can be seen, indicating the development of short-range magnetic correlations as is approached from above. We discuss the Bragg and diffuse magnetic scattering in Section III.4 and III.5, respectively.
III.4 Magnetic structures from Rietveld refinements
We first discuss possible ordered magnetic structures of Na2Mn3Cl8, as determined by analyzing the magnetic Bragg profiles obtained at temperatures below .
We used the program KSearch of the Fullprof suite (Rodriguez-Carvajal_1993, ; Rodriguez-Carvajal_1993a, ) to identify possible propagation vectors at K () and K (). The positions of 10 magnetic peaks (at K) and 15 magnetic peaks (at K) were provided as input, and a systematic search of candidate propagation vectors was performed, starting with those that lie on a symmetry point, line, or plane of the Brillouin zone. However, none of the high-symmetry propagation vectors was compatible with the observed Bragg positions, at either temperature. The best-fit propagation vectors were instead of the form , with . We obtain at K, and at K; precise values are given in Table 2. These propagation vectors lie on a general position, but they are close to the high-symmetry plane.
Having determined possible propagation vectors, we used the program Sarah (Wills_2000, ) to identify symmetry-allowed magnetic structures. The primitive unit cell contains three Mn2+ sites, with fractional coordinates , , and with respect to the conventional axes , , . Each site has three magnetic degrees of freedom, which are not further constrained by symmetry. We choose these as basis-vector components along orthonormal axes , , , where is parallel to the -axis, is parallel to the projection of in the -plane, and is perpendicular to and c. These structures are amplitude-modulated spin-density waves (sine structures), with different spin magnitudes and orientations for each site,
(2) |
where denotes the propagation vector, denotes a lattice vector, labels sites within the unit cell, and are basis-vector components. Alternatively, it is possible to construct helical structures such as
(3) |
where, in this case, the spin plane is perpendicular to the -axis. The ordered magnetic-moment length can be identical on all sites in the crystal in a helical structure, for example if in Eq. (3).
Due to the relatively large number of variable parameters and the limitations of powder data, we make two assumptions when testing candidate magnetic structures. First, we only consider structures that order with a single propagation vector (single- structures). While multi- structures are possible, they cannot generally be distinguished from single- structures by powder diffraction Kouvel_1963 . Second, we initially assume that the basis vectors at sites , and are parallel; this assumption is reasonable because the interactions between nearest and next-nearest neighbors are ferromagnetic, as we will show in Section III.5. Magnetic-structure models were tested against the magnetic Bragg profile using Rietveld refinement (see Methods).

Na2Mn3Cl8, magnetic | |||||
at K | |||||
at K | |||||
(K) | Structure | () | () | () | (%) |
sine | |||||
sine | |||||
(K) | Structure | () | (°) | (°) | (%) |
-helix | |||||
-helix | |||||
-helix | |||||
-helix | |||||
helix | |||||
-helix | |||||
-helix | |||||
-helix | |||||
-helix | |||||
helix |
We first considered amplitude-modulated sine structures. The assumption of parallel basis vectors reduces the number of refined parameters from 9 to 3. Sine structures yield excellent agreement with our data at both 0.3 and 0.8 K, as shown in Figure 4(a) and (e), respectively. The magnetic moment is predominantly oriented along for the corresponding structures, which are shown in Figure 4(c) and (g), respectively. The refined parameter values and goodness-of-fit metric are given in Table 2. To determine if sine structures are physically reasonable, we calculated the maximum value of the ordered magnetic moment, . For a spin-only ion, this value should not normally exceed ( for Mn2+). This expectation is confirmed by the low-temperature magnetization of Na2Mn3Cl8, which saturates to approximately per Mn2+ [Figure 2(d)]. Unfortunately, we find for the refined sine structures: at K, and at K. These values are physically unreasonable, suggesting that the correct structures of Na2Mn3Cl8 are not single- sine structures.
Circular helices are promising alternative structures, since all sites have equal magnetic moment lengths. Initially, we consider circular helices with magnetic moments in either the plane, the plane, or the plane (equivalent to the plane). At both and K, the best fit is obtained for the -helix, with slightly worse agreement for the -helix [Table 2]. The -helix yields much worse agreement than the other structures, so we do not consider it further. The fits for -helices at and K are shown in Figure 4(b) and (f), respectively, and the corresponding structures are shown in Figure 4(d) and (h). The agreement with the data is very good, although marginally worse than for the corresponding sine structures. Importantly, however, the refined values of are now physically reasonable, with a maximum value of at K. This result favors the helical structures.
We tested two variations of the helical structures in an effort to improve the fit quality. First, we considered the -helix and relaxed our previous assumption of parallel basis vectors, by refining a clockwise rotation of the basis vector at position about the -axis. The optimal fit is obtained for relatively small at K, and substantial at K. Second, we maintain the assumption that the basis vectors are parallel, but vary the spin plane as . At K, a minimum in occurs for , whereas at K, fit quality is essentially unchanged for all . Each of these variations yields a similar or slightly improved fit compared to the simple -helix (see Table 2).

Figure 5 shows the temperature evolution of the refined parameter values for the -helix with collinear basis vectors. A discontinuity in the propagation vector is apparent at [Figure 5(a)]. No such anomaly is apparent in the refined value of the ordered magnetic moment, which increases smoothly on cooling the sample below [Figure 5(b)]. The temperature dependence of this order parameter at low temperatures ( K) is consistent with the phenomenological form for a three-dimensional magnet with half-integer spin Kobler_2003 . Its temperature dependence for K is consistent with critical form for a three-dimensional Heisenberg magnet, , although the small number of data points precludes fitting the critical exponent.
In conclusion, our powder-diffraction data are well explained by circular helical magnetic structures. Basis vectors are close to the plane and nearly collinear at K, with a possibility of greater noncollinearity at K. We emphasize, however, that the possibility of multi- structures cannot be ruled out, and we discuss this further in Section IV.2.
III.5 Magnetic interactions from diffuse scattering
We seek to parametrize the spin Hamiltonian of Na2Mn3Cl8 by analyzing the diffuse magnetic scattering measured above . This approach is an alternative to spin-wave analysis of inelastic neutron-scattering data, and has recently been applied to several frustrated antiferromagnets (Samarakoon_2020, ; Scheie_2021, ; Welch_2022, ). The magnetic diffuse scattering measured at K and K (with K data subtracted) is shown in Figure 6. Diffuse magnetic peaks are sharper at K than at K, consistent with an increase in the magnetic correlation length on cooling the sample. The bulk magnetic susceptibility expressed as is also shown in Figure 6 and confirms the development of antiferromagnetic correlations.

(K) | (K) | (K) | (K) | (K) | (K) | (K) | ||||
---|---|---|---|---|---|---|---|---|---|---|
(a) | ||||||||||
(b) | ||||||||||
(c) | ||||||||||
(d) |
To model these data, we consider a Hamiltonian that includes Heisenberg exchange interactions and the long-range magnetic dipolar interaction,
(4) |
where is modeled as a classical vector of magnitude , is the spin quantum number of Mn2+, is a unit vector parallel to the separation of spins and , and is the nearest-neighbor distance. The exchange interactions include the nearest-neighbor exchange , the inter-layer coupling , and the further-neighbor couplings shown in Figure 1(c), so that . The magnitude of the dipolar interaction at the nearest-neighbor distance, K, is determined by the crystal structure. The values of the exchange interactions were optimized against our and data (see Methods).
We tested interaction models against our data in order of increasing number of exchange parameters, as follows. Unless otherwise noted, the dipolar interaction was fixed at K. For each exchange model, the best fit to and data is shown in Figure 6. The refined values of the exchange interactions are shown in Table 3, along with the goodness of fit metric , and two quantities estimated from the Onsager-reaction-field calculation that may indicate model quality: the predicted magnetic ordering temperature and the predicted magnetic propagation vector .
First, we considered a minimal model in which only and were refined [model (a)]. This model does not represent our neutron data well [Figure 6(a)], and the predicted propagation vector is commensurate, in contrast to the incommensurate propagation vector observed experimentally. Second, we refined , , and parameters [model (b)]. This model yields a substantially improved fit, but some misfit is still evident in the and, especially, the data [Figure 6(b)]. The calculated propagation vector is now incommensurate, but different to the experimental one. Third, we refined , , , , and parameters [model (c)]. This model yields an excellent fit to both the and data [Figure 6(c)]. Moreover, the calculated propagation vector, , is close to the experimental value of in the first ordered state at K, and the calculated K agrees with the measured value. This refinement was stable despite the relatively large number of free parameters; no large parameter covariances () were noted, and initializing the refinement with different parameter values yielded only one possible local minimum, which had significantly worse and .
Our results suggest that model (c) represents well the interactions of Na2Mn3Cl8. This model has weak ferromagnetic , consistent with the Goodenough-Kanamori rules. The inter-layer coupling is antiferromagnetic, consistent with the antiferromagnetic layer stacking observed below . The third-neighbor couplings and are antiferromagnetic and significantly larger than . Hence, Na2Mn3Cl8 is an unusual system where strong antiferromagnetic third-neighbor interactions compete with ferromagnetic nearest-neighbor interactions. To the best of our knowledge, a similar - competition has been identified in only one other kagome material, vesignieite Boldrin_2018 . However, this material differs from Na2Mn3Cl8 in its magnetic properties as well as its chemistry, as it has and shows commensurate magnetic ordering Boldrin_2018 .
Finally, we considered the relevance of the long-ranged dipolar interaction by performing a fourth refinement in which , , , , and parameters were varied, while was fixed at zero [model (d)]. This refinement yielded worse agreement with and data, and significantly underestimates the value of [Table 3]. This result shows that the dipolar interaction has a significant effect on the magnetic properties, as expected since is of comparable magnitude to the exchange interactions. However, the refined values of all parameters except are equivalent (within ) for models (c) and (d), suggesting that the effect of the dipolar term on these refinements is largely confined to nearest neighbors.
IV Theory and Modeling
IV.1 Magnetic interactions from first principles
To gain insight into the exchange interactions, we performed first-principles calculations using density-functional theory (see Methods). The values of the interactions calculated using DFT are given in Table 4 for different values of the Hubbard between and eV. Based on other materials, we anticipate that is likely between and eV.
The first-principles exchange interactions show similarities with the experimentally-determined values, but also substantial differences. On the one hand, the first-principles values of and are ferromagnetic and antiferromagnetic, respectively, consistent with the values fitted to experimental data. The magnitudes of and for eV are also comparable to the experimentally-determined magnitudes, in contrast to a previous DFT study that reported interactions larger than K . On the other hand, the first-principles values of , , and are opposite to the experimentally-determined values; moreover, the calculated magnitudes of these interactions are very large compared to the other interactions.
We carefully checked whether the first-principles results could be consistent with our experimental data. Taking eV, we calculate the Weiss temperature as K. Hence, DFT predicts a ferromagnetic Weiss temperature, which is not consistent with the antiferromagnetic value ( K) measured experimentally. We also estimate the magnetic ordering temperature to be K, which is much larger than the experimental value of K. Finally, we performed additional refinements to neutron and data as described in Section III.5, except we constrained the signs of the exchange interactions to be the same as those from DFT, while allowing their magnitudes to refine freely. These refinements yielded , essentially reproducing the results of model (b) in Section III.5.
We therefore conclude that the DFT results are not fully consistent with our experimental data, making Na2Mn3Cl8 a model material for benchmarking developments in first-principles calculations. The reason for the inaccuracy of the DFT exchange interactions beyond nearest-neighbors is not yet clear; an interesting possibility is that it may relate to the neglect of the Stoner coupling on the Cl ligand sites, as recently proposed in the related material NaMnCl3 (Solovyev_2022, ).
(eV) | (K) | (K) | (K) | (K) | (K) |
---|---|---|---|---|---|
IV.2 Origin of incommensurate ordering

In this section, we discuss the origin of the multiple incommensurate ordering transitions in Na2Mn3Cl8, using a combination of field-theoretic and Monte Carlo simulations.
Incommensurate magnetic structures are relatively uncommon in kagome antiferromagnets. For example, to the best of our knowledge, all known jarosite minerals that exhibit long-range order have either or propagation vectors (see (Mendels_2011, ) and references therein). Similarly, commensurate states are observed for many other insulating materials in which the kagome lattice is undistorted or slightly distorted; for example, MgFe3(OH)6Cl2 with (Fujihala_2017, ), centennialite CaCu3(OH)6Cl0.6H2O (Iida_2020, ), CdCu3(OH)6(NO3)0.6H2O (Ihara_2022, ), Nd3Sb3Mg2O14 (Scheie_2016, ), and Sr-vesignieite SrCu3V2O8(OH)2 with , -Cu3Mg(OH)6Br2 (Wei_2019, ) and YCu3(OH)6Cl3 with (Zorko_2019, ), and Ba-vesignieite BaCu3V2O8(OH)2 with (Boldrin_2018, ; Okamoto_2009, ). By contrast, the distorted-kagome material Ba2Mn3F11 is one of the only insulating kagome materials with incommensurate magnetic ordering (Hayashida_2018, ). Incommensurate modulations are more frequently observed in metallic kagome systems, such as Tb3Ru4Al12 (Rayaprol_2019, ) and YMn6Sn6, the latter of which undergoes an incommensurate-to-commensurate transition on cooling (Ghimire_2020, ).
To understand the preference for kagome magnets to form commensurate structures, and the conditions where incommensurate structures may appear, we use the reciprocal-space mean-field approximation introduced in Section II.3 to investigate the stability of different phases as a function of the interactions , , , and [Figure 1(c)]. Throughout large regions of this interaction space, the classical ground state is one of the commensurate “regular magnetic orders” described in Ref. (Messio_2011, ). Of the models previously investigated theoretically, the most relevant one to Na2Mn3Cl8 is the - Heisenberg model studied in Refs. (Grison_2020, ; Li_2022, ). The phase diagram for this model is shown in Figure 7(a), and contains five phases: ferromagnetic layers with antiferromagnetic stacking [], antiferromagnet, antiferromagnet [], three-sublattice antiferromagnet [], and an incommensurate region. This result reproduces the result of Ref. (Grison_2020, ) for isolated kagome planes, except that we include a small antiferromagnetic inter-layer coupling to stabilize three-dimensional ordering.
While the - phase diagram is relatively complicated, it is nevertheless simpler than our model for Na2Mn3Cl8, which also includes significant , , and dipolar couplings. We therefore extended the - phase diagram to consider the effects of these additional couplings, which are needed for a full description of our Na2Mn3Cl8 data. Notably, for all models, antiferromagnetic is necessary to stabilize incommensurate ordering with . In Figure 7(b), we fix ferromagnetic and consider the phase diagram in the - plane for antiferromagnetic . Nonzero has a dramatic effect on the phase diagram; in particular, including antiferromagnetic extends the stability region of the incommensurate phases observed for antiferromagnetic . Figure 7(b)–(d) show the effect of increasing the magnitude of , the antiferromagnetic interlayer coupling (, , and , respectively, in the same units as ). The effect of increasing is to increase further the region of phase space in which incommensurate order is stable within the mean-field approximation. Finally, in Figure 7(e), we show the - phase diagram including the long-range dipolar interaction appropriate for Na2Mn3Cl8. The inclusion of has a relatively small effect on the positions of the phase boundaries.

The reciprocal-space mean-field theory provides a useful overview of the phase space, but has several important limitations. First, for a non-Bravais lattice such as kagome, it only determines a lower bound on the energy of the ground state. As discussed in Ref. (Grison_2020, ), for the incommensurate region of the - phase diagram, a physical spin configuration could not be identified that reached this lower bound; hence, the actual magnetic ground state is uncertain in this region. Second, since this theory considers instabilities of the paramagnetic phase, it predicts only the propagation vector of the first ordered state that develops on cooling; it provides no information about the possibility of multiple phase transitions, as are observed experimentally in Na2Mn3Cl8.
We performed classical Monte Carlo simulations to address these limitations. Since the periodicity of an incommensurate magnetic structure does not “fit” within any finite-sized configuration, finite-size artifacts are encountered, which can be reduced by studying relatively large system sizes. However, the long-ranged nature of the magnetic dipolar interaction makes large system sizes computationally expensive. We therefore consider first an approximation to the full Hamiltonian, Eq. (4), where we simulate the parameters that best describe our diffuse-scattering data [model (c) in Table 3], but truncate the dipolar interaction at the nearest-neighbor distance; we will call this the “nearest-neighbor dipolar model”. For comparison, we also simulated the same model (c) except with . To identify finite-size effects, we considered different system dimensions from hexagonal unit cells ( spins) to hexagonal unit cells ( spins). For the and simulations only, we slightly adjusted the model (c) interaction parameters to stabilize ordering, which is commensurate with the system size; this was achieved by multiplying the best-fit values of and by . To investigate the effect of a different system geometry, we defined a orthogonal unit cell with axes , , and , and performed simulations of and orthogonal unit cells ( and spins, respectively). Simulations were run for up to moves per spin at low temperatures, where a single move involved one microcanonical (over-relaxation) update followed by a proposed spin rotation of a randomly-chosen spin, which was accepted or rejected according to the Metropolis criterion. Measurements of the autocorrelation function showed that these conditions allowed the system to decorrelate at all temperatures above K. Simulations including the long-ranged dipolar interaction, implemented using Ewald summation (Wang_2001, ), were also performed for a small system size of hexagonal unit cells, without over-relaxation updates.
Results of our Monte Carlo simulations are shown in Figure 8. For the model with , a sharp anomaly indicating a single magnetic phase transition is observed at K; we do not consider the low-temperature state here. The nearest-neighbor dipolar model shows a more complex temperature evolution. In all our simulations, sharp specific heat anomaly is observed at K, with a second feature between and K that is resolved as either a single broadened peak or two peaks close in temperature, depending on system dimensions. Hence, unlike the Heisenberg model, the nearest-neighbor dipolar model shows at least two magnetic phase transitions, in qualitative agreement with the experimental data for Na2Mn3Cl8. Properties of the magnetic phases obtained for a model of orthogonal unit cells are shown at , , and K, in Figure 8(c), (d), and (e) respectively. The phases observed at and K are resolved for all other system sizes and geometries. However, the K phase is not resolved in the simulation, suggesting its appearance for some other system sizes may be a finite-size artifact. The calculated magnetic powder diffraction patterns show remarkably good agreement with our experimental powder-diffraction data, especially at and K [Figure 8(c)–(e)]. Calculations of the single-crystal magnetic diffraction patterns reveal magnetic Bragg peaks corresponding to a single incommensurate wavevector at K, indicating a single- magnetic structure at this temperature [Figure 8(c)]. Remarkably, however, the same calculation shows magnetic Bragg peaks corresponding to two wavevectors at and K. The intensity of each wavevector is approximately equal at K but significantly different at K [Figure 8(d) and (e)]. The same effect was observed across all our simulations at and K, suggesting this is likely not an artifact due to domain formation, but instead indicates the formation of double- states in the Monte Carlo simulations. Our simulations of the long-ranged dipolar model also suggest a possible change in magnetic structure below approximately K, although a second transition is not clearly resolved in the heat capacity for this small simulation size [Figure 8(b)]. For this model, the magnetic structure is clearly 2- only below K.
Our results suggest the enticing possibility that the ordered incommensurate states may, in fact, be multi- structures rather than single- helices. Given the good agreement of our microscopic model with powder-diffraction data and its correct prediction of multiple phase transitions, this scenario is certainly possible. Further theoretical studies including the long-ranged dipolar interaction would be useful to elucidate the relative stabilities of single- and multi- states, which may be close in energy.
V Conclusions
Our neutron-diffraction study reveals that Na2Mn3Cl8 shows novel magnetic behavior. Unusually for a kagome antiferromagnet, it shows incommensurate ordering; even more unusually, it exhibits multiple incommensurate magnetic phases, which form at and K. To the best of our knowledge, ordering wavevectors of the form , as observed in Na2Mn3Cl8, have not previously been observed in insulating kagome magnets. As such, Na2Mn3Cl8 significantly expands the known range of magnetic behavior on the kagome lattice.
We investigated the magnetic interactions that drive incommensurate ordering in Na2Mn3Cl8 using experiment-driven and first-principles approaches. By fitting the magnetic diffuse scattering measured above the magnetic ordering temperature, we showed that the magnetic interactions extend to third-nearest neighbors. Antiferromagnetic third-neighbor interactions and are the largest terms in the Hamiltonian, and compete with ferromagnetic nearest-neighbor interactions . Using a mean-field theory, we showed that antiferromagnetic , , and interlayer couplings extends the stability region of incommensurate ordering in a model with ferromagnetic . Our experimentally-determined interactions could not be fully reproduced by DFT calculations, which predict ferromagnetic and , inconsistent with our experimental data. This material may therefore be a useful test case for advancements in first-principles methodologies.
Using magnetic Rietveld refinement, we showed that the magnetic Bragg profiles of the two incommensurate magnetic phases are well described by single- helical structures. These are cycloidal helices, in which the spins and the propagation vector both have a component in the -plane. Due to the limitations of powder data, however, other structures can give equivalent or slightly better agreement with the experimental pattern. We showed that single- sine structures are highly unlikely at and K, since some sites would have unphysically large magnitudes of the ordered magnetic moment. However, we were not able to rule out multi- structures, which are generally indistinguishable from their single- analogs in powder diffraction measurements. This issue is especially relevant here, because Monte Carlo simulations of our experimentally-determined interaction model show multiple magnetic phases transitions, in qualitative agreement with the experimental data, and indicate that two of the phases obtained are 2- states. Further experiments would therefore be valuable to distinguish between single- and double- states. These experiments could include single-crystal neutron diffraction under applied magnetic field, or inelastic neutron scattering. The growth of large single crystals of Na2Mn3Cl8 would facilitate such measurements and potentially shed further light on the nature of the spin texture in Na2Mn3Cl8.
Acknowledgements.
We are grateful to Andrew Christianson (Oak Ridge National Laboratory) for valuable discussions. This work was supported by the U.S. Department of Energy (DOE), Office of Science, Basic Energy Sciences, Materials Sciences and Engineering Division, and used resources at the High Flux Isotope Reactor, a DOE Office of Science User Facility operated by the Oak Ridge National Laboratory.References
- (1) L. Balents, Nature 464, 199 (2010).
- (2) J. T. Chalker, P. C. W. Holdsworth, E. F. Shender, Phys. Rev. Lett. 68, 855 (1992).
- (3) M. E. Zhitomirsky, Phys. Rev. B 78, 094423 (2008).
- (4) M. R. Norman, Rev. Mod. Phys. 88, 041002 (2016).
- (5) M. A. de Vries, et al., Phys. Rev. Lett. 103, 237201 (2009).
- (6) T.-H. Han, et al., Nature 492, 406 (2012).
- (7) T.-H. Han, J. Singleton, J. A. Schlueter, Phys. Rev. Lett. 113, 227203 (2014).
- (8) C. M. Pasco, et al., Phys. Rev. Mater. 2, 044406 (2018).
- (9) K. Tustain, G. J. Nilsen, C. Ritter, I. da Silva, L. Clark, Phys. Rev. Mater. 2, 111405 (2018).
- (10) A. S. Wills, A. Harrison, J. Chem. Soc., Faraday Trans. 92, 2161 (1996).
- (11) T. Inami, M. Nishiyama, S. Maegawa, Y. Oka, Phys. Rev. B 61, 12181 (2000).
- (12) M. Nishiyama, S. Maegawa, T. Inami, Y. Oka, Phys. Rev. B 67, 224435 (2003).
- (13) S. Janas, et al., Phys. Chem. Chem. Phys. 22, 25001 (2020).
- (14) W. G. Bisson, A. S. Wills, J. Phys.: Condens. Matter 20, 452204 (2008).
- (15) L. Messio, C. Lhuillier, G. Misguich, Phys. Rev. B 83, 184401 (2011).
- (16) V. Grison, P. Viot, B. Bernu, L. Messio, Phys. Rev. B 102, 214424 (2020).
- (17) H. Li, T. Li, Phys. Rev. B 106, 035112 (2022).
- (18) C. J. J. Loon, D. J. W. Ijdo, Acta Crystallogr. B 31, 770 (1975).
- (19) K. P. Devlin, R. Cava, Solid State Commun. 342, 114598 (2022).
- (20) V. Meschke, P. Gorai, V. Stevanović, E. S. Toberer, Chem. Mater. 33, 4373 (2021).
- (21) D. J. Hinz, G. Meyer, T. Dedecke, W. Urland, Angewandte Chemie International Edition in English 34, 71 (1995).
- (22) N. Hänni, M. Frontzek, J. Hauser, D. Cheptiakov, K. Krämer, Z. Anorg. Allg. Chem. 643, 2063 (2017).
- (23) Z. A. Kelly, T. T. Tran, T. M. McQueen, Inorg. Chem. 58, 11941 (2019).
- (24) A. Paul, C.-M. Chung, T. Birol, H. J. Changlani, Phys. Rev. Lett. 124, 167203 (2020).
- (25) D. I. Khomskii, T. Mizokawa, S. V. Streltsov, Phys. Rev. Lett. 127, 049701 (2021).
- (26) A. W. Hewat, Acta Crystallogr. A 35, 248 (1979).
- (27) J. A. M. Paddison, arXiv:2210.09016 (2022).
- (28) D. E. Logan, Y. H. Szczech, M. A. Tusch, Europhys. Lett. (EPL) 30, 307 (1995).
- (29) R. Brout, H. Thomas, Physics Physique Fizika 3, 317 (1967).
- (30) M. Enjalran, M. J. P. Gingras, Phys. Rev. B 70, 174426 (2004).
- (31) F. James, M. Roos, Comp. Phys. Commun. 10, 343 (1975).
- (32) F. James, MINUIT Function Minimization and Error Analysis: Reference Manual Version 94.1, CERN (1994).
- (33) J. Rodriguez-Carvajal, Physica B 55, 192 (1993).
- (34) J. Rodríguez-Carvajal, Physica B 192, 55 (1993).
- (35) K. Momma, F. Izumi, J. Appl. Crystallogr. 41, 653 (2008).
- (36) E. Sjöstedt, L. Nordström, D. Singh, Solid State Communications 114, 15 (2000).
- (37) P. Blaha, K. Schwarz, G. K. H. Madsen, D. Kvasnicka, J. Luitz, WIEN2K, an augmented plane wave+local orbitals program for calculating crystal properties, Technische Universität Wien, Vienna (2001).
- (38) D. J. Singh, L. Nordstrom, Planewaves pseudopotentials and the LAPW method (Springer, Berlin, 2006).
- (39) J. P. Perdew, K. Burke, M. Ernzerhof, Phys. Rev. Lett. 77, 3865 (1996).
- (40) H. Nakano, T. Sakai, J. Phys. Soc. Jpn. 84, 063705 (2015).
- (41) A. Wills, Physica B: Condens. Matter 276-278, 680 (2000).
- (42) J. Kouvel, J. Kasper, J. Phys. Chem. Solids 24, 529 (1963).
- (43) U. Köbler, A. Hoser, D. Hupfeld, Physica B: Condens. Matter 328, 276 (2003).
- (44) A. M. Samarakoon, et al., Nat. Commun. 11, 892 (2020).
- (45) A. O. Scheie, et al., arXiv 2109.11527 (2021).
- (46) P. G. Welch, et al., Phys. Rev. B 105, 094402 (2022).
- (47) D. Boldrin, et al., Phys. Rev. Lett. 121, 107203 (2018).
- (48) I. V. Solovyev, A. V. Ushakov, S. V. Streltsov, Phys. Rev. B 106, L180401 (2022).
- (49) P. Mendels, A. S. Wills, Kagomé Antiferromagnets: Materials Vs. Spin Liquid Behaviors (Springer Berlin Heidelberg, Berlin, Heidelberg, 2011), pp. 207–238.
- (50) M. Fujihala, et al., Phys. Rev. B 96, 144111 (2017).
- (51) K. Iida, et al., Phys. Rev. B 101, 220408 (2020).
- (52) Y. Ihara, et al., Phys. Rev. B 106, 024401 (2022).
- (53) A. Scheie, et al., Phys. Rev. B 93, 180407 (2016).
- (54) Y. Wei, et al., Phys. Rev. B 100, 155129 (2019).
- (55) A. Zorko, et al., Phys. Rev. B 100, 144420 (2019).
- (56) Y. Okamoto, H. Yoshida, Z. Hiroi, J. Phys. Soc. Jpn. 78, 033701 (2009).
- (57) S. Hayashida, et al., Phys. Rev. B 97, 054411 (2018).
- (58) S. Rayaprol, A. Hoser, K. K. Iyer, S. K. Upadhyay, E. Sampathkumaran, J. Mag. Magn. Mater. 477, 83 (2019).
- (59) N. J. Ghimire, et al., Sci. Adv. 6, eabe2680 (2020).
- (60) Z. Wang, C. Holm, J. Chem. Phys. 115, 6351 (2001).