Chapter 6 Six-fold clock-anisotropy in honeycomb titanates CoTiO3and FeTiO3

Abstract

Recently published neutron scattering experiments have shown that the easy-plane honeycomb cobaltate CoTiO3shows incredibly rich low temperature physics including a six-fold in-plane anisotropy proposed to originate in bond dependent exchange and quantum zero-point fluctuation mechanism. This chapter reports single-crystal piezocantilever torque magnetometry experiments that aim to directly probe the magnetic anisotropy and test the proposed scenario of in-plane six-fold anisotropy. Measurements are reported for both the easy-plane honeycomb magnet CoTiO3, as well as the isostructural easy-axis magnet FeTiO3. Clear evidence of a six-fold clock anisotropy is observed in both materials.

§ 6.1 Six-Fold Clock Anisotropy in CoTiO3

§ 6.1.1 Crystal Structure of CoTiO3

Figure 6.1: Model showing crystal structure of CoTiO3in the a) hexagonal b) rhombohedral settings. Dark blue spheres indicate Co2+ions, light blue spheres indicate Ti4+ ions, red spheres indicate O2 ions. The thin black lines indicate the structural unit cell in each setting, and the triad of arrows in the bottom left corner of each panel are the associated lattice vectors for that setting.

CoTiO3has an Ilmenite crystal structure comprised of stacked layers of honeycombs with an R3¯ space group. Figure 6.1a shows that the layers alternate between containing honeycombs of edge sharing CoO6 octahedra and edge sharing TiO6 octahedra. The honeycomb layers are not quite in the 𝒂, 𝒃 plane, but the two sites of the honeycomb are displaced relative to one another along the 𝒄 axis, leading to some buckling of the planes.

Figure 6.1b shows an alternative setting for the structure, the rhombohedral setting. This setting is the primitive structural unit cell and contains two Co2+sites, two Ti4+ sites, and six O2 sites. The hexagonal basis vectors (𝒂,𝒃,𝒄) can be written it terms of the primitive vectors (𝑨,𝑩,𝑪) as

(𝒂𝒃𝒄)=(011101111)(𝑨𝑩𝑪) (6.1)
Table 6.1: Structure parameters of CoTiO3in standard setting (Domain A in Fig. 6.2)
Space group: R3¯H (148)
Hexagonal setting
Z = 6
Cell parameters:
a,b,c (Å) 5.066 5.066 13.918
α,β,γ (°) 90 90 120
Volume (Å3) 309.365
Atomic fractional coordinates:
Atom Wyckoff Site Site Symmetry x y z
Co2+ 6c 3 0 0 0.35511
Ti4+ 6c 3 0 0 0.14558
O2 18f 1 0.31623 0.02091 0.24588

Previous studies have reported the crystal structure to a high quality standard [52, 75]. Table 6.1 shows a summary of the structural parameters and atomic fractional coordinates of the material’s ions, with data taken from [52].

Figure 6.2: Calculated Bragg peak intensities in the (hk1) plane for two structural domains of CoTiO3, related by a 2-fold rotation around (110). Highlighted are trios of peaks which change intensities strongly between the two domains.

In addition, previous reports [26, 27] pointed out commonly occurring twins in single crystals. These are the A domain, with parameters consistent with [52], shown in Table 6.1, and the B domain, which is rotated 180 ° around the (110) axis in the hexagonal setting. This rotation leaves the positions of the Co2+and Ti4+ ions unchanged, but does change the O2 ion positions. Because of this change in oxygen positions, one can determine which domains are present in a sample through X-ray diffraction. This is possible by finding Bragg peaks where the structure factor changes significantly in magnitude between the two domains. In this case, because the transition metal ion positions are invariant between the domains, this change in structure factor will be large if the contribution from the O2 ions changes sign between the two domains.

Figure 6.2 shows results of calculations of Bragg peak intensities in the (hk1) crystallographic plane for domains A and B. On these are highlighted a trio of peaks whose intensity varies strongly between the domains. X-ray diffraction was performed at room temperature with an Oxford Diffraction Supernova diffractometer on all CoTiO3crystals used in this chapter to verify sample quality. By comparing measured Bragg peak intensities to the calculated values, the domain populations can be estimated. Each crystal used for the torque measurements is an untwinned A-domain sample.

§ 6.1.2 Single Ion Physics of Co2+in CoTiO3

The magnetic ions in CoTiO3are 3d7 Co2+ions in an octahedron of oxygen ions, this octahedral crystal environment is trigonally distorted along the 𝒄 direction. The Hund’s rules ground state for this electron configuration is L=3, S=32. The crystal field Hamiltonian can be written

̂CEF=23B4(O40+202O43)+B20O20+B40O40 (6.2)

using the Stevens operators Onk in the Cartesian basis with x,y,z along the a,b,c axes respectively. B4>0 parametrises the octahedral cubic crystal field environment, and the additional terms characterise a trigonal distortion along the 𝒄 axis. With this, spin-orbit coupling is included to give a single-ion Hamiltonian

̂=λ𝑳̂𝑺̂+̂CEF (6.3)

where λ>0 because the shell is more than half filled.

This can be approached perturbatively by assuming that the octahedral cubic crystal field (OCF)

̂OCF=23B4(O40+202O43) (6.4)

is the largest energy scale.

In the |Lz basis, the OCF gives a ground state T1u triplet at energy 480B4, an excited T2u triplet at +120B4, and a further excited excited A2u singlet at +600B4 [1]. The ground state has basis vectors

|0~ =49|0+5912[|+3|3] (6.5a)
|±1~ =±56|2+16|±1 (6.5b)

where the states are labelled by the effective angular momentum operator |lz with the equivalence relation 𝑳̂32𝒍̂, valid on the projected subspace of this triplet of states. The spin-orbit and trigonal distortion terms can then be treated as a perturbation on this ground state manifold.

Figure 6.3: a) Schematic diagram showing the level splitting of isolated Co2+ion in an octahedral crystal field in the presence of trigonal distortion and spin-orbit coupling. Labels on states show good quantum numbers for those levels. b, c) show evolution of energy levels with spin-orbit coupling strength λ and trigonal distortion strength δ starting from the limit of fixed δ and fixed λ respectively. Vertical lines show values proposed in [27].

The spin-orbit coupling operator ̂SO=λ𝑳̂𝑺̂ can be rewritten in terms of the effective orbital angular momentum operator ̂SO=32λ𝒍̂𝑺̂. By itself, this couples the spin and orbital parts of the angular momentum into a total effective angular momentum Jeff, with values Jeff=1/2,3/2,5/2.

The trigonal distortion operator can be written in terms of effective orbital angular momentum operator as ̂trig=δ(l̂z223)=13δO20. Acting on the octahedral crystal field l=1,S=3/2 ground state, this splits the 12-fold degenerate manifold into an a 8-fold degenerate lz=±1 manifold of states and a 4-fold degenerate lz=0 manifold. Acting as on the Jeff states from spin-orbit coupling, it mixes the states with same Jeffz, as seen in the level splitting diagram in Figure 6.3. The result is a ground state Kramers doublet with Jeffz=±1/2, and the next excited level a Jeffz=±3/2 doublet at 15 meV. These levels are determined by diagonalising the perturbation Hamiltonian, treating the spin-orbit coupling and trigonal distortion at the same level. The effects of each of these terms on the energy levels can be seen in Figure 6.3.

The values of λ and δ used were determined by previous reports by fitting the energy levels to an inelastic neutron scattering energy scan [26, 27].

§ 6.1.3 Magnetic Structure of CoTiO3

Figure 6.4: a) Model showing magnetic unit cell of CoTiO3in hexagonal setting. Solid lines indicate the structural unit cell in the hexagonal setting. Arrows overlaid on Co2+ion sites indicate spin alignment at that site. b), projection of single honeycomb layer onto 𝒂𝒃 plane. Label ϕ shows angle between ordered spin orientation and 𝒂 axis.

CoTiO3has been shown previously to order at 38 K into an antiferromagnetic arrangement [38, 26, 27, 94, 6]. The structure consists of spins ferromagnetically aligned within each honeycomb layer, the layers alternate so that each layer has opposite spin alignment to the previous. The primitive magnetic unit cell has lattice vectors 𝑴a,𝑴b,𝑴c which can be written in terms of the structural lattice vectors as

(𝑴a𝑴b𝑴c)=(011101110)(𝑨𝑩𝑪)=13(122212112)(𝒂𝒃𝒄) (6.6)

The primitive magnetic unit cell is a doubling of the primitive (rhombohedral) structural unit cell. The magnetic cell unit vectors (𝑴a,𝑴b,𝑴c) are rotated relative to (𝑨,𝑩,𝑪) by π3 around the rhombohedral (111) axis. The cell is also stretched along the same axis, doubling the unit cell volume relative to the primitive structural unit cell. In total there are four Co2+sites in the magnetic unit cell.

Alternatively, Figure 6.4a) shows how the magnetic unit cell can be constructed in the hexagonal setting. In this setting, the magnetic unit cell is a simple doubling of the hexagonal structural unit cell along the 𝒄 direction. This doubling occurs because there are three layers per structural unit cell in this setting; the layers alternate in magnetisation so six layers, or two structural cells, are needed for the magnetic structure to repeat.

§ 6.1.4 Exchange Interactions and Magnetic Anisotropy

No single-ion anisotropy is allowed in the ground state, because the ground state doublet is described by an effective S=1/2 and all even polynomials of Sx2,Sy2,Sz2 are constants. The anisotropy effect, an energy preference of spins to point in particular directions over another, must be caused by exchange anisotropy. An inelastic neutron scattering study fit a minimal XXZ exchange model including nearest-neighbour intralayer exchange J1 and next-nearest-neighbour interlayer Heisenberg exchanges J4, J6 [111]. These parameters can be seen in Table 6.2. By using these parameters in a classical model, which is solved as discussed in Chapter 3, the magnetic ground state described above is stabilised up to a spontaneous orientation of spins in the 𝒂𝒃 plane.

Table 6.2: XXZ exchange parameters in a minimal model for CoTiO3from [111]

. J1x=4.44meV J1z=0.0meV J4x=J4z=0.6meV J6x=J6z=0.6meV

The XXZ Hamiltonian as described has no dependence on the orientation of spins in the 𝒂𝒃 plane. This is because it has a continuous U(1) symmetry, higher than the discrete 3¯ point group symmetry of the material. If a U(1) symmetry was realised, one would expect an associated gapless Goldstone mode in the excitation spectrum [30], whereas measurements observe a clear spin-gap [26]. The model developed by Elliot et.al [26] lifted this in-plane degeneracy by including a bond-dependent anisotropic exchange term on the nearest-neighbour honeycomb bonds through a parameter η defined such that

η=JyJx (6.7)

in the the local coordinates of the each nearest-neighbour bond. In these coordinates, 𝒛 is the 𝒄 axis, 𝒙 is the projection of the bond direction into the 𝒂𝒃 plane, and 𝒚=𝒛×𝒙. As such the bond’s Hamiltonian is written as

̂ij=JxSixSjx+(Jx+η)SiySjy+JzSizSjz (6.8)

when η is non-zero, it introduced a dependence of the spin-wave spectrum on the azimuthal angle ϕ, defined as the angle between the ordered spin direction and the 𝒂 axis, as seen in Fig.6.4. As a consequence, the zero-point contribution to the ground state energy acquires a dependence on ϕ and discrete spin orientations in-plane are selected, with a 60 ° periodicity.

The classical, mean field calculation developed in this thesis cannot capture this effect. This can be demonstrated by explicitly computing the effective interaction between sublattices discussed in Chapter 3.

The most general bilinear exchange matrix can be written as

𝒥=(JxΓz+DzΓyDyΓzDzJyΓx+DxΓy+DyΓxDxJz) (6.9)

where D terms parametrise a Dzyaloshinskii–Moriya (DM) interaction, and Γ terms parametrise symmetric off-diagonal exchange. In this coordinate system, 𝒛 along the local 3-fold axis, 𝒙 is the 𝒂 direction, and 𝒚=𝒛×𝒙. For each bond not parallel to the 𝒛 axis, there is another symmetry equivalent bond obtained by a 3z+ rotation with exchange matrix

𝒥=z(2π3)𝒥z(2π3) (6.10)

where z(±2π3) are matrix representations of the the 3-fold rotation about the 𝒛 axis for spins in these coordinates. This has components given in Table 6.3.

Table 6.3: Symmetry constrained exchange terms for bond obtained by a 3z+ rotation.
Jx 14Jx+34Jy+32Γz
Jy 34Jx+14Jy32Γz
Jz Jz
Dx 12(Dx3Dy)
Dy 12(3DxDy)
Dz Dz
Γx 12(Γx+3Γy)
Γy 12(3ΓxΓy)
Γz 12Γz34Jx+34Jy

In the magnetic structure of CoTiO3, the spins in the honeycomb plane are all parallel, so the total exchange energy between sites 𝑺A and 𝑺B for these three bonds can be written as the sum

=𝑺AT𝒥𝑺B+𝑺ATz(2π3)𝒥z(2π3)𝑺B+𝑺ATz(2π3)𝒥z(2π3)𝑺B (6.11a)
=𝑺AT𝒥T𝑺B (6.11b)

where 𝒥T can be written in terms of the single bond exchange

𝒥T=3(12(Jx+Jy)Dz0Dz12(Jx+Jy)000Jz) (6.12)

The off-diagonal terms show a DM contribution which also cancels. This is because all the bonds with multiplicity of three, such as nearest-neighbour bonds, have an inversion centre at their midpoint so Dz=0 by symmetry. Other bonds with a multiplicity of six, such as next-nearest neighbour, in these the other trio of bonds must have opposite Dz such that the total is zero. The diagonal terms show no dependence on orientation in the 𝒂𝒃 plane so at the mean field level, no bilinear exchange term can introduce an energy dependence on the alignment of spins in the xy plane.

Calculations of the zero-point contribution to the ground state energy find a clear energy dependence on the spin orientation in the 𝒂𝒃 plane of the form

E(ϕ)=Λcos6ϕ (6.13)

where Λη3 to leading order [26]. This 6-fold periodicity is a requirement of the 3¯ point group symmetry of the crystal combined with time reversal. The energy of the system cannot change if the system is transformed under a symmetry operation of the Hamiltonian, for example if all spin operators are reversed.

Figure 6.5: Angular dependence of effective site anisotropy energy, top labels indicate the corresponding terms in Eqn. 6.14. Blue colours show negative energies, red areas show positive energies. Here, the 𝗓 is the projection of the spin onto the crystallographic 𝒄 axis, 𝗑𝒂, 𝗒𝒃

In the mean field approximation, spin operators are replaced with their expectation values, which in shorthand can be written as 𝗑:=Sx,𝗒:=Sy,𝗓:=Sz. In order to preserve the 3 point group site symmetry of the Co2+ion, the basis functions must be unchanged under the symmetry operations of that group. Additionally, time reversal maps 𝑺𝑺, so the basis functions must also be unchanged under inversion. Up to 6thorder polynomials of 𝗑,𝗒,𝗓, this energy is

EA(𝗑,𝗒,𝗓)=A𝗓2+B𝗓4+C𝗓(𝗑33𝗑𝗒2)+C𝗓(3𝗑2𝗒𝗒3)+D𝗓3(𝗑33𝗑𝗒2)+D𝗓3(3𝗑2𝗒𝗒3)+E(𝗑615𝗑4𝗒2+15𝗑2𝗒4𝗒6)+E(6𝗑5𝗒20𝗑3𝗒3+6𝗑𝗒5) (6.14)

These functions are plotted over the unit sphere in Figure 6.5, labelled by their coefficient, which shows the negative energies (easy axes) in blue and positive energies (hard axes) in red. For spins in the 𝒂𝒃 plane 𝗓=0, so only the E and E functions are non-zero, and the anisotropy energy can be written

EA =Λ0sin6θcos6(ϕϕ6) (6.15a)
=Λcos6(ϕϕ6) (6.15b)

with ϕ the azimuthal angle measured from the 𝒂 axis, tan6ϕ6=EE, and |Λ|=E2+E2. When ϕ6=0, Λ>0 the easy axes are at ϕ=π2+nπ3 and change to ϕ=nπ3 for Λ<0. This behaviour matches that of the quantum-fluctuation induced anisotropy energy in the presence of bond-dependent anisotropy η [26].

It is worth highlighting that the E and E terms cannot come from single-ion anisotropy in CoTiO3. This is because the quantum form of the single-ion anisotropy terms involves Stevens operators. In order to get the E and E terms, the O6±6 operators are needed which have the form

O6+6 =12[L̂+6+L̂6] (6.16a)
=L̂x615L̂x4L̂y2+15L̂x2L̂y4L̂y6 (6.16b)
O66 =12i[L̂+6L̂6] (6.16c)
=6L̂x5L̂y20L̂x3L̂y3+6L̂xL̂y5 (6.16d)

These operators give no energy change for Co2+ions. This is because the L=3 quantum state in Co2+is made up of d orbital electrons, so the highest order Stevens operator with a non-zero eigenvalue is O4±n, where n can be 0, 2, or 4. The anisotropy being modelled originates from interactions; it is not a single-ion effect.

§ 6.1.5 Single Crystal Vibrating Sample Magnetometry studies of CoTiO3

In order to determine the high temperature magnetic properties of CoTiO3, the magnetic susceptibility of single crystals was measured using vibrating sample magnetometry. Figure 6.6a) shows the evolution of the magnetic susceptibility between 2 K and 300 K, measured at 0.1 T for orthogonal field orientations relative to the crystallographic axes. The curves in green and blue show the susceptibility along the 𝒂 and 𝒃 directions respectively, while the curve in red shows the susceptibility along the 𝒄 direction. There is a remarkably strong anisotropy between the directions parallel and normal to the 𝒄 axis, showing an overall easy-plane arrangement. There is a strong suppression of susceptibility for temperatures below  38 K in the 𝒂𝒃 plane, while the suppression along the normal direction is much smaller. This confirms that CoTiO3orders into an antiferromagnet with a Néel temperature of TN=38 K. There is no significant difference in susceptibilities for fields along the two directions in the 𝒂𝒃 plane across the entire temperature range, suggesting that CoTiO3behaves very close to uniaxially in its magnetic behaviour, i.e. the anisotropy in the 𝒂𝒃 plane is very small.

Figure 6.6: Temperature dependence of a) susceptibility and b) inverse susceptibility in CoTiO3for orthogonal sample orientations 𝑯𝒂, 𝑯𝒃, 𝑯𝒄. In a) dot-dash lines show reference data from [38]. In b) dotted line shows Curie-Weiss fit.

Temperature dependent susceptibility data (Fig. 6.6) can be fitted to a Curie-Weiss relationship

χ=χ0+CT+θCW (6.17)

over the temperature range 50 K-200 K.

Table 6.4: Fitted Curie-Weiss parameters for CoTiO3temperature-dependent susceptibility data.
χ0 (102 emu mol-1) θCW (K) C (emu Kmol1) μeff (μB)
𝑯𝒂 1.049(4) 13.3(1) 5.98(1) 6.9
𝑯𝒃 0.936(2) 11.9(1) 5.89(1) 6.9
𝑯𝒄 1.056(4) 8.7(2) 0.73(1) 2.4

The fitted parameters can be seen in Table 6.4. These data in some areas agree well with previously published data [38, 94, 103]: the low-temperature shapes of the curves and absolute scale of the susceptibilities, and the strong anisotropy of susceptibilities that is seen at all temperatures. The high-temperature data does not agree well with the previously published results, here a larger value of χ is observed than previously reported. This discrepancy is likely caused by a magnetic background from the brass sample holder, which has not been subtracted. Where susceptibility data is used for calibration in later sections, the reference data [38] is used. While this high temperature discrepancy may also account for disagreement between the fitted and published Curie-Weiss parameters [103], if the fitting process is repeated on more recently published data in [38], the fitted moment μeff agrees relatively well. This data gives μeff= 2.5μB, 7.3μB along the 𝒄 and 𝒂 directions respectively. This is because a background term χ0 is captured in the fit. The poor fit of Curie-Weiss behaviour at high temperature is expected from the relatively small energy gap of 15 meV175 K to the next excited doublet. For temperatures greater than this value, the excited doublet states are populated and the CW treatment fails. A more thorough treatment including all of the energy levels predicted from the level splitting would capture this effect, but as the crystal field parameters and λ have already been established, this is not necessary.

Figure 6.7: Field dependence of a) magnetisation b) differential susceptibility in CoTiO3for a range of temperatures from 2K to 40K with 𝑯𝒂.

Figure 6.7a) shows the evolution of magnetisation with applied magnetic field up to 4 T for temperatures between 2 K and 40 K. Above TN, the relationship is linear. As the temperature falls below the Néel temperature, the linear relationship no longer holds. For fields below B2.5 T and temperatures below TN, the differential susceptibility is suppressed significantly, as can be seen in Fig 6.7b). This effect was reported previously [38], where the peak in MH at approximately 2.5 T was labelled B. This can be attributed to the in-plane anisotropy proposed previously, as magnetic domains exist in the material at low field, with ordered moments in the 𝒂𝒃 plane. This suggests that a single domain is selected for fields μ0HB.

§ 6.1.6 Single Crystal Torque Magnetometry Studies of CoTiO3

As has been discussed in Chapter 2.2, torque magnetometry is a powerful tool for exploring the anisotropy of a material. In this section I discuss how this technique was used to study the in-plane anisotropy of CoTiO3.

Figure 6.8: Photographs of CoTiO3single crystals mounted on piezoresistive cantilevers for torque measurements. In b) the crystal is attached ‘edge on’ and is just visible on the top side of the lever.

While I was able to cut large crystals of CoTiO3into arbitrary shapes to align the field along the different crystallographic axes for the VSM experiments described above, the size of the cantilevers requires very small single crystals of order 100 µm to be able to measure torque. To prepare these crystals I crushed small amounts of the larger CoTiO3crystals in grease. This allowed small crystallites to be collected from the resulting material. As CoTiO3is a layered material, it preferentially broke between layers, leading to large numbers of plate-like single crystals with a normal vector parallel to 𝒄. An example of such a crystal can be seen in Figure 6.8a), which shows a plate-like single crystal of CoTiO3with a diameter of 100 µm, mounted ready for torque measurements in the 𝒂𝒄 plane. This plate-like arrangement made mounting the crystal for torque in the 𝒂𝒃 plane much more challenging, as the plane normal 𝒄 had to be perpendicular to the lever surface normal. In order to measure the crystal in such an orientation, I mounted it on the side of the lever, as seen in Figure 6.8b), supported by a small quantity grease on the surface. This made precise alignment very difficult. In this image, the crystal is seen ‘edge-on’. As the in-plane susceptibility is much larger than the out-of-plane susceptibility χχ even at high temperature, subjecting the crystal to a large magnetic field causes a torque on the sample which would tend to improve the alignment of the crystal. That is, the energy is minimised if the field is in the 𝒂𝒃 plane. This effect was used to improve the crystal alignment in the 𝒂𝒃 plane by rotating the crystals in a 16 T field at 300 K, such that the grease was a unfrozen. The crystal was then cooled in field to base temperature.

Figure 6.9: a) Waterfall plot showing magnetic torque vs rotator angle for temperatures between 2 K and 200 K in ac plane. Top axis labels indicate angles where the magnetic field direction is aligned parallel to a crystallographic axis. Traces are raw torque piezocantilever voltages converted to absolute units as described in the text.

Unlike in the other systems studies in this thesis, I have been able to normalise the torque signal in CoTiO3in absolute units. This is because of the availability of high quality published data of anisotropic susceptibilities along different crystal axes in absolute units, as discussed above. Above the Néel temperature, CoTiO3behaves as a uniaxial paramagnet, so the torque at high temperatures can be calculated from the susceptibility data (Eqn. 2.36) using the susceptibility measured parallel to the 𝒄 axis χ, and measured in the 𝒂𝒃 plane χ:

τ(θ,T) =12B2(χ(T)χ(T))sin2θ (6.18a)
=τ2(T)sin2θ (6.18b)

where θ is the angle between the field and the 𝒄 axis. As the anisotropic susceptibilities are known in absolute units, torque in the 𝒂𝒄 plane can be measured for a large temperature range above the TN and the results compared to the predicted values to determine the conversion factor between the measured signal and the physical torque, which will capture both the sensitivity of the lever and the mass of the sample.

Figure 6.9 shows measured magnetic torque for rotations in the 𝒂𝒄 plane at 1 T and temperatures between 10 K and 200 K. These show an overall sin2θ dependence, with a positive gradient when fields are parallel to the 𝒄 axis and a negative gradient when fields are parallel to the 𝒂 axis at all temperatures, indicating a smaller susceptibility in-plane than out of plane. This is in agreement with easy-plane type antiferromagnet model determined from the VSM data reported previously. These piezocantilever voltages, in units of mV were fitted to the form

V=V0sin2(θθ0) (6.19)

and the values V0 compared to the predicted values

τ2(T)=12B2(χ(T)χ(T)) (6.20)

for temperatures T>50 K. Figure 6.10 shows the plot of V0 vs τ2 for all temperatures measured, the green line shows a linear fit to these data. The dashed red line shows a line with the same gradient and zero intercept. This calibrated scaling factor comes to

τ=V5.938(10)×103meVrad1/Co2+mV (6.21)
Figure 6.10: a) Plot showing fitted torque amplitudes in 𝒂𝒄 plane against predicted torque amplitudes from published susceptibility data [38] with linear fit (green line), and linear fit with zero intercept (red dashed line). b) Plot showing evolution of (cyan points) fitted torque amplitudes, (red curve) predicted torque amplitudes from published susceptibility data [38] for rotations in 𝒂𝒄 plane against temperature.

which can be used to calibrate all other measurements on the same sample-lever pair. It should be noted that these samples are too small for their mass to be measured using standard equipment, so this calibration factor is not able to calibrate other samples on the same lever. For this reason, all torque data reported in this chapter has been measured using the same calibrated sample shown in Figure 6.8. Figure 6.10 shows how this calibration was verified by comparing the temperature dependence of the experimentally observed torque amplitudes, scaled by this calibration factor, against the predicted temperature dependence for the paramagnetic phase from (6.20). These show good agreement over the high temperature region where (6.20) is valid, indicating that this calibration procedure was effective.

Figure 6.11: Waterfall plot showing magnetic torque vs rotator angle ϕ for applied magnetic fields between 1 T and 16 T in the 𝒂𝒃 plane, measured at the experimental base temperature of 2 K. a) shows raw data, while b) shows the same data with background subtracted. ϕ is measured from the nominal 𝒂 axis. Top labels above a) indicate angles where the magnetic field direction is aligned parallel to a crystallographic axis. Dashed vertical line corresponds to field along the 𝒂 axis. Traces are torque piezocantilever voltages converted to absolute units as described in the text, and vertically offset between subsequent fields for clarity.

Figure 6.11a) shows the evolution of magnetic torque with field in the 𝒂𝒃 plane, measured at the experimental base temperature of 2 K, far below the ordering temperature for this material. These torque profiles show complex behaviour at all fields, unlike the simple sin2θ behaviour seen in the 𝒂𝒄 orientations. Despite the complexity, six peaks and six troughs can be clearly seen in all fields in the ϕ=0°360° angle range. This suggests a significant contribution from a six-fold energy term, however the signal also contains other components. Some of these will be contributed by the lever background, as the resistance of the bridge will vary slightly with applied magnetic field. Others will be caused by a small projection of the 𝒄 axis onto the rotation plane, as the anisotropy between the 𝒂𝒃 plane and the 𝒄 axis is much stronger than any anisotropy within the 𝒂𝒃 plane, only a small projection of this axis is needed for the signal to be significant compared to the very small contribution from in-plane anisotropy.

Figure 6.12: Plot showing evolution of fitted components of a) Eqn. (6.22a) b) Eqn. (6.22b) with applied magnetic field between 1 T and 16 T for torque measurements in the 𝒂𝒃 plane of CoTiO3measured at 2 K. Dashed vertical lines indicate fields where τ6 components are expected to change sign as described in the text.

In order to better show the six-fold contribution to the torque signal, I have performed a simple background subtraction by removing low order Fourier components. To do this I fit the signal to the form

τ =C0+C1ϕ+n=1mAnsinnϕ+Bncosnϕ (6.22a)
=C0+C1ϕ+n=1mτnsinn(ϕϕn) (6.22b)

where all An,Bn,Cn are free fitting parameters, and m=12 is chosen to be sufficiently large that increasing it does not significantly improve the fit. Here τn2=An2+Bn2, and tannϕn=AnBn. This form is motivated from the experimental design: C0 is non-zero if the electronic Wheatstone bridge balancing is not perfect. C1 can account for a linear drift in the lever balancing over the course of the experiment, which occurs as the temperature stability is not ideal. The choice of the Fourier series is because the rotation is a physical rotation of the levers in the magnetic field. Physically, the torque applied to the lever must be periodic in ϕ, unless it is hysteretic. I then take the fitted curve with parameters n<5 as the ‘background’ to be removed, which is subtracted from the raw signal. Note that Fourier components n=1,3,4 are very small, but above the noise level, likely coming from non-linearities and background in the lever assembly itself.

Figure 6.13: Angular dependence of magnetic torque in 𝒂𝒃 plane for a selection of applied magnetic fields 3, 8, 14, 16 T. All measured at 2 K. Top labels indicate angles where the magnetic field direction is aligned parallel to a crystallographic axis. Vertical dashed line indicates 𝑯𝒂. Traces are raw torque piezocantilever voltages converted to absolute units and with background subtracted as described in the text.

Figure 6.11b) shows the same signal measured at 2 K with the background subtracted as described above. This shows an overall pattern of a six-fold torque signal, which increases in magnitude as the strength of the applied magnetic field increases from 1 T to 16 T. Additionally the phase and amplitude of the signal is modulated with the magnetic field strength. In order to explore these effects, Figure 6.12 shows the evolution of the fitted parameters in Eqn. (6.22a) as a function of applied magnetic field. This shows that the six-fold components are comparable in magnitude to the two-fold components. Additionally it shows the modulation of the A6 and B6 components, which increase in magnitude with applied magnetic field and change sign at approximately equal magnetic fields. This effect can be more clearly seen in Figure 6.13, which shows select torque traces, with background subtracted, by themselves. These clearly show the change in sign of the signal as the applied magnetic field increases.

Figure 6.14: Field dependence of calculated total anisotropy energy per magnetic unit cell (Eqn. 6.15) for 𝑩𝒂 between 0 T and 16 T for the case of Λ<0 and ϕ=0. All exchange and g-factor parameters are as in the text (Table 6.2) and Λ<0 is assumed to be independent of field strength. Blue/red arrows show the alignment of spins in adjacent layers in the 𝒂𝒃 plane while colour in the annular region indicates the corresponding local energy environment of the spins.

I have drawn a vertical line at 0 ° on Figure 6.11b) to indicate where the magnetic field is approximately parallel to the 𝒂 axis. This is close to the zeros of the six-fold torque signal. In low fields, below 4 T, the gradient through this point is positive, indicating an unstable equilibrium. At large fields (above 14 T), the gradient is negative indicating a stable equilibrium. At intermediate fields the gradient changes three times, at 4 T, 10 T, and 14 T. The gradient is negative between 4 T and 10 T and positive between 10 T and 14 T.

These changes can be naturally explained by considering the changes in the anisotropy energy as the material becomes magnetised. At the largest field of 16 T in-plane the magnetisation is almost saturated; the spins are approximately parallel with the magnetic field. An overall stable equilibrium with the field along 𝒂 suggests that the local anisotropy for the spins has a stable equilibrium at the same position. In small magnetic fields below 4 T the spins are approximately antiparallel, and orthogonal to the field direction in order to minimise energy. With 𝑩𝒂, this constrains the spins to be collinear with 𝒃 (𝒂). The overall system is in an unstable equilibrium, indicating that spins along this orientation is an energy maximum, determined by the positive torque gradient τϕ>0 when the 𝑩𝒂 (see Fig. 6.13). The intermediate field behaviour can be understood as caused by the modulations in local energy of the spins as they rotate towards the field. I illustrate this is Figure 6.14, which shows the evolution in the anisotropy energy as field is increased along the 𝒂 direction. With the magnetic field below sin15°Bsat, the system is in an unstable equilibrium. For sin15°Bsat<B<sin45°Bsat the system is stable and so on. The numbers on the x-axis indicate the fields when these switches will occur for a saturation field of 16 T. These fields are overlaid as vertical lines on Figure 6.12, which show good agreement with the minima in the fitted amplitudes of the τ6 components.

Figure 6.15: a) Reproduced from Figure 6.11 for comparison. b) Waterfall plot showing calculated magnetic torque vs rotator angle ϕ for applied magnetic fields between 1 T and 16 T in the 𝒂𝒃 plane, with ϕ measured from the 𝒂 axis. Top labels above a) indicate angles where the magnetic field direction is aligned parallel to a crystallographic axis. Dashed vertical line corresponds to field along the 𝒂 axis. Calculation details are in the text.

From this model and the data in Figure 6.11, I can determine the anisotropy parameters of interest. In particular, at 16 T, the angular dependence of torque can be fit to the form

τ(ϕ)=τ6sin6(ϕϕ6) (6.23)

with τ6=8.5(2)×104meVrad1/Co2+ and ϕ6+2°. Note that the absence of mirror planes containing the 𝒄-axis means that ϕ6 is not symmetry constrained and could in principle take any value. As the material should be (approximately) fully polarised at this magnetic field strength, this torque should be coming from the full scale of the anisotropy energy

EA =Λcos6(ϕϕ6) (6.24)
τ6 =6Λsin6(ϕϕ6) (6.25)

so the local anisotropy energy can be read off simply as Λ=1.4(1)×104meV/Co2+.

The model for a quantum-fluctuation induced six-fold clock anisotropy [26] suggests that at leading order

Λ=18aη3 (6.26)

where a is a constant estimated at a=6.13(7)×105meV2 and η, defined in Eqn. (6.7), 1.7 meV, suggesting a value of Λ3.8×105meV. This is of the same order of magnitude as the value estimated from torque.

Practical considerations with the mounting mean that precisely aligning crystallographic axes within the plane is not possible, so the phase ϕ6 should not be taken to be a precise offset. However, it is clear from the data that the stable equilibrium is close to the 𝒂 axis and the unstable equilibrium close to the 𝒃 axis, tentatively placed within 5 ° of these. In the quantum fluctuations model, the negative η<0 favoured spins parallel to the bond direction, which is parallel to 𝒂. Together with the order of magnitude quantitative agreement of the size of the magnetic anisotropy this agreement on the locations of the in-plane easy axes provides strong support for the interpretation that quantum fluctuations are the origin of the anisotropy.

Figure 6.15 shows a numerical calculation for torque in the 𝒂𝒃 plane for fields between 2 T and 16 T with the anisotropy energy taken as above. This calculation captures the modulations in torque strength and phase change with magnetic field strength. The calculation predicts a larger magnitude of torque at small fields than is measured, this is likely a combination of two main factors. It is clear from the suppression of susceptibility in the 𝒂𝒃 plane that CoTiO3has magnetic domains which remain in small magnetic fields, the pinning of these domains is not modelled by the calculation and their presence is likely to suppress the energy modulation as the field is rotated. If the anisotropy amplitude Λ originates from zero-point quantum fluctuations, one would expect a variation with field strength. Most straightforwardly the quantum fluctuations would be expected to be suppressed at high field, inconsistent with what is seen experimentally. It should be noted however that the energy scale here is gμBBJx, so quantum fluctuations should still be significant.

§ 6.2 Six-Fold Clock Anisotropy in FeTiO3

A contrasting case to CoTiO3is the isostructural material FeTiO3, which has an identical magnetic structure of stacked ferromagnetic layers, where adjacent layers are antiferromagnetically coupled, however FeTiO3is an easy-axis system, rather than an easy-plane like CoTiO3. The structural parameters of FeTiO3can be seen in Table 6.5.

Table 6.5: Refined 300 K crystal structure parameters of FeTiO3from analysis of powder diffraction data in [27].
Space group: R3¯H (148)
Hexagonal setting
Z = 6
Cell parameters:
a,b,c (Å) 5.088 5.088 14.085
α,β,γ (°) 90 90 120
Volume (Å3) 309.365
Atomic fractional coordinates:
Atom Wyckoff Site Site Symmetry x y z
Fe2+ 6c 3 0 0 0.3562(6)
Ti4+ 6c 3 0 0 0.1469(6)
O2 18f 1 0.3561(13) 0.0397(18) 0.2384(7)

FeTiO3orders at 58 K as an easy-axis antiferromagnet [50], as can be seen clearly by the susceptibility-temperature curves shown in Figure 6.16. These show nearly isotropic susceptibilities in the high-temperature limit and a sharp peak in susceptibility along the 𝒄 axis (cyan) at the ordering temperature. Below the Néel temperature, the susceptibility along the 𝒄 is strongly suppressed while the susceptibility in the 𝒂𝒃 plane is approximately temperature independent. The magnetic structure of FeTiO3is identical to that of CoTiO3. Layers of Fe2+are ferromagnetically aligned, with alternating layers antiferromagnetic. In FeTiO3, the spins are almost parallel to the 𝒄 axis, although neutron diffraction studies report a clear experimental signature of a small but finite 1.6 ° tilt away from the 𝒄 axis [48, 49, 109].

Figure 6.16: Temperature dependence of anisotropic susceptibilities in plane (red) and out of plane (blue) for FeTiO3between 2 K and 300 K measured in an applied magnetic field of μ0H=1.38 T adapted from [50]

§ 6.2.1 Single ion physics of Fe2+ions in FeTiO3

The magnetic ions in FeTiO3are 3d6 Fe2+ions in an octahedral crystal environment with trigonal distortion. The Hund’s rules ground state for this electronic configuration is L=2, S=2. The octahedral crystal field Hamiltonian can be written

̂CEF=23B4(O40+202O43)+B20O20+B40O40 (6.27)

using the Stevens operators Onk in the Cartesian basis with x,y,z along the 𝒂,𝒃,𝒄 axes respectively [1]. B4>0 parametrises the octahedral cubic crystal field environment, and the additional terms characterise a trigonal distortion along the 𝒄 axis. With this, spin-orbit coupling is included to give a single-ion Hamiltonian

̂=λ𝑳̂𝑺̂+̂CEF (6.28)

where λ>0 because the shell is more than half filled.

This can be approached perturbatively by assuming that the cubic octahedral crystal field ̂OCF is the largest energy scale in ̂.

In the |Lz basis, the octahedral field gives a ground state T2g triplet at energy 48B4, and an excited Eg doublet at +72B4 [1]. The ground state has basis vectors

|0~ =|0 (6.29a)
|±1~ =±23|213|±1 (6.29b)

where the states are labelled by the effective angular momentum operator |lz (z along the trigonal 𝒄 axis) with the equivalence relation 𝑳̂𝒍̂, valid on the projected subspace of this triplet of states. The spin-orbit and trigonal distortion terms can then be treated as a perturbation on this ground state manifold.

Figure 6.17: a) Schematic diagram showing the level splitting of isolated Fe2+ion octahedral crystal field ground state energy levels as a function of trigonal distortion and spin-orbit coupling. Labels on states show good quantum numbers for those levels. b, c) show evolution of energy levels with spin-orbit coupling strength λ with fixed δ and against trigonal distortion strength δ at fixed λ respectively on the 15-fold degenerate manifold of l=1, S=2 states. Dashed vertical lines indicate the values closest to the experimental situation according to [27].

The spin-orbit coupling operator ̂SO=λ𝑳̂𝑺̂ can be rewritten in terms of the effective orbital angular momentum operator ̂SO=λ𝒍̂𝑺̂. By itself, this couples the spin and orbital parts of the angular momentum into a total effective angular momentum Jeff, with values Jeff=1,2,3.

The trigonal distortion operator can be written in terms of the effective orbital angular momentum operator as ̂trig=δ(l̂z223)=13δO20. Acting on the 15-fold degenerate octahedral crystal field l̂=1,S=2 ground state manifold, this splits the state into an a 10-fold degenerate l̂z=±1,S=2 manifold and a 5-fold degenerate lz=0,S=2 manifold. Acting on the Jeff states from spin-orbit coupling, it mixes the states with same Jeffz, as seen in the level splitting diagram in Figure 6.17. The result is a ground state doublet with Jeffz=±1, and the next excited level a Jeffz=0 singlet. These levels are determined by diagonalising the perturbation Hamiltonian, treating the spin-orbit coupling and trigonal distortion together. The effects of each of these terms on the energy levels can be seen in Figure 6.17.

In the presence of this δ<0 trigonal distortion, as appears to be the experimental case [27], the ground state is the easy-axis situation Jeffz=±1, with next excited Jeffz=0 singlet at 3 meV. By applying a small field to split the ground state doublet, the moment along the z axis can be determined as μz=4.43μB. In qafm, the classical spin is taken as Jeff=1 with a single-ion anisotropy term A=D(Jeffz)2, where D>0 because the system is easy-axis. The energy D can be read off as as the gap between the ground state doublet and excited singlet as D3 meV, if the values from [27] are used. This easy-axis situation is confirmed by the metamagnetic spin-flip transition seen at 8 T for fields along the 𝒄 axis. These transitions are only present for large D(Jeffz)2-type anisotropy, while XXZ easy-axis antiferromagnetism will always favour a spin-flop transition.

Strictly, the above single-ion anisotropy of a octahedral cubic environment with trigonal distortion assumes a 3¯m point group for the Fe2+ion. The full symmetry allowed anisotropy for the 3 point group is however [48, 74]

B20O20+B40O40+B43O43+B43O43 (6.30)

§ 6.2.2 Torque magnetometry studies of FeTiO3

Figure 6.18: Photograph of FeTiO3single crystals mounted on piezoresistive cantilever for torque measurements. Overlaid symbols and arrows show crystallographic 𝒂,𝒃,𝒄 axes.

Piezoresistive cantilever torque magnetometry was performed on single crystals of FeTiO3in order to explore the anisotropy of the material. A photo of the sample reported throughout this section can be seen in Figure 6.18.

Figure 6.19 shows how the torque profiles in the 𝒂𝒄 plane change with temperature through the magnetic phase transition. At all temperatures the profile is a good fit to

τ(θ)=τ2sin2(θθ0) (6.31)

where θ is measured from the 𝒄 axis towards the a axis and θ0 is interpreted as an offset between the nominal 𝒄 orientation and the true orientation. In the high-temperature paramagnetic region, the gradient through the 𝒄 axis is negative, indicating that it is a stable equilibrium. This is expected as the susceptibility measured along the 𝒄 axis χ is larger than the in-plane susceptibility χ.

Figure 6.19: Waterfall plot showing magnetic torque vs rotator angle in the 𝒂𝒄 plane for temperatures between 2 K and 120 K and an applied magnetic field of 1 T. Top labels indicate angles where the magnetic field direction is aligned parallel to a crystallographic axis. Dashed line indicates 𝑯𝒄. Traces are raw torque piezocantilever voltages converted to absolute units as described in the text.
Figure 6.20: a) Plot showing fitted torque amplitudes in 𝒂𝒄 plane against predicted torque amplitudes from published susceptibility data [50] with linear fit (green line), and linear fit with zero intercept (red dashed line). b) Plot showing evolution of (cyan points) fitted torque amplitudes, (red curve) predicted torque amplitudes from published susceptibility data [50] for rotations in 𝒂𝒄 plane against temperature.

Figure 6.20 shows the plot of fitted piezocantilever voltage amplitude V2(T) against predicted torque 12B2(χ(T)χ(T)) for temperature range 60 K<T<110 K, the green line shows a linear fit to these data. The dashed red line shows a line with the same gradient and zero intercept, this fit was used to calibrate the absolute value of the torque as described in the previous section. At the phase transition the difference χχ is maximised, which is seen as a peak in the amplitude in the torque at 58 K in Figure 6.20b). As the temperature cools further below the transition the torque amplitude drops as the out-of-plane susceptibility is suppressed by the antiferromagnetic order. At approximately 39 K the torque in minimised at χχ. This temperature is lower than the 46 K expected from the susceptibility data. The most likely cause of this disagreement is local sample heating; as domain reorientation while the sample is rotated in field could cause the sample to be heated above the bath temperature. At lower temperatures than this the stability of the axes changes, such that the system in unstable for fields close to 𝒄, and stable for fields in the plane. This is because the out-of-plane susceptibility χ has been suppressed below the in-plane value χ. The evolution of this torque in the paramagnetic phase T>TN was used to calibrate the magnitude of the torque for these crystals by comparing the measured torque to the predicted value from 12B2(χχ)sin2θ as was described in the previous section, all further results in this chapter are on the same lever sample pair, although the results were verified on a second sample.

Figure 6.21: a) Waterfall plot showing magnetic torque vs rotator angle ϕ for applied magnetic fields between 1 T and 6 T in the 𝒂𝒃 plane of FeTiO3, measured at the experimental base temperature of 2 K. Top labels indicate angles where the magnetic field direction is aligned parallel to a crystallographic axis. Traces are raw torque piezocantilever voltages converted to absolute units as described in the text. b) Waterfall plot showing calculated magnetic torque vs rotator angle ϕ for applied magnetic fields between 1 T and 16 T in the 𝒂𝒃 plane, with ϕ measured from the 𝒂 axis. Dashed vertical line corresponds to field along the 𝒂 axis. Calculation details are in the text.
Figure 6.22: Field dependence of fitted torque amplitudes from Eqn. 6.22a in the range 1 Tμ0H6 T. Dashed lines indicate fit to the form (μ0H)αn, with α21.8 and α63.5.

In order to explore the existence of in-plane anisotropy in FeTiO3, angular dependence of torque was measured for rotations in the 𝒂𝒃 plane. Figure 6.21 shows how the magnetic torque measured in the 𝒂𝒃 plane varies with applied magnetic field at 2 K. A dashed vertical line indicates where field is parallel to the 𝒂 direction. These show a very clear 6-fold torque signal such that the system is in a stable equilibrium for fields close to the 𝒂 direction and an unstable equilibrium for fields close to 𝒃. This suggests that the local energy environment for the spins in minimised for spins parallel to 𝒂, that is parallel to the bond directions. This matches the case for CoTiO3. Unlike in the case for CoTiO3, this six-fold torque signal shows no change in the phase of the signal as field increases, and the magnitude increases monotonically. This is because in FeTiO3, the spins are canting from the 𝒄 axis, so their projection in the 𝒂𝒃 plane is directly towards the magnetic field, it is not dependent on field magnitude.

The largest field I was able to test with this setup was 6 T, this is because the growing torques became a danger of breaking the magnetometer levers. At this field, the amplitude of the 6-fold component of the torque is τ6=4.2×103meVrad1/Fe2+, approximately 5 times larger than the largest value measured in CoTiO3. Although the moments do not lie in-plane, so the full scale of the anisotropy is not exerting a torque on the sample, the observed size of the τ6 component of the torque allows a lower bound to be placed that |Λ0|>7×104meV/Fe2+. However, it should be noted that the value for the energy scale in CoTiO3was determined at 16 T, when the magnetisation was nearly saturated. At 6 T, far below the saturation field of the material, only a small part of the anisotropy energy will be seen, suggesting the six-fold clock anisotropy energy in FeTiO3is much larger than in CoTiO3.

The field scaling of this torque parameter was determined by plotting the amplitude of the six-fold torque τ6 against the magnetic field, as can be seen in Figure 6.22. These amplitudes were fitted to the form

τn(μ0H)αn (6.32)

with the curves for the fitted parameters overlaid on the figure. The exponent for the six-fold anisotropy α63.5. If the origin of the anisotropy was single-ion, then it should have the form from Eqn. (6.15)

EA =Λ0sin6θcos6(ϕϕ6) (6.33a)
=Λ0(MxyMsat)6cos6(ϕϕ6) (6.33b)

so a value of α6=6 would be expected as Mxy=χμ0Hxy. This is larger than the value found experimentally, providing further support that the anisotropy does not originate in single-ion effects.

Figure 6.23: a) Angular and temperature dependence of magnetic torque the 𝒂𝒃 plane between 2 K and 120 K with an applied magnetic field of 6 T. Top labels indicate angles where the magnetic field direction is aligned parallel to a crystallographic axis. Traces are raw torque piezocantilever voltages converted to absolute units as described in the text. b) Evolution of fitted torque amplitudes from Eqn. 6.22a against temperature in the range 2 KT100 K.

Figure 6.21b) shows the results of a calculation, performed as in the case of CoTiO3, with exchange parameters taken from [27]. To this I added an anisotropy energy of the form in Equation 6.14, with Λ<0,ϕ6=0, such that individual spins minimise their energy by pointing along nearest-neighbour bond directions. These show qualitative agreement with the data, with a six-fold contribution to the torque of the same sign and phase as is seen in the data. The calculations results also predict a torque that increases with the magnetic field faster than seen experimentally with exponent αn=6, larger than the fitted value of 3.5.

Figure 6.23 shows the temperature evolution of the magnetic torque measured in the 𝒂𝒃 plane measured at 6 T, between 100 K and 2 K. In the high temperature regime, a 2-fold torque signal can be seen. This 2-fold signal is symmetry disallowed for the 𝒂𝒃 plane because of the rhombohedral symmetry of the crystal. For this reason I attribute this signal primarily to a small misalignment, such that the rotation plane is skewed from the 𝒂𝒃 plane. Fits to the 2-fold component of the signal seen in Figure 6.23b) show a peak, associated with the onset of magnetic order. This can be understood as a projection of this much larger easy-axis anisotropy into the measurement plane. As the temperature cools below the transition, a large 6-fold component of torque can be seen, increasing monotonically as temperature decreases. Figure 6.23b) shows the evolution with temperature of the fitted components of the torque signal. As can be seen, the six-fold component can be detected only below the magnetic phase transition.

The lowest order single-ion anisotropy term which could induce a cos6ϕ energy profile is the O6±6 Stevens operator. The d-orbital valence electrons of Fe2+do not permit any high order Onq,n>4 terms in the CEF Hamiltonian. Because of this, the 6-fold torque profile seen at low temperatures in FeTiO3cannot have a single-ion anisotropy origin. This is further evidenced by the temperature dependence of the 6-fold component of the torque, which has onset alongside magnetic order, as can be seen in Figure 6.23. From this evidence I suggest that the anisotropy seen must originate from magnetic interactions and set in alongside the onset of the magnetic order, through quantum fluctuations inducing an angular dependence of the anisotropy energy or a spin-orbital exchange mechanism as proposed in CoTiO3[26].

§ 6.3 Conclusions

Despite their differences in spin orientations, both of the two systems discussed in this chapter, easy plane CoTiO3and easy axis FeTiO3show clear evidence of a 6-fold modulation in energy in the 𝒂𝒃 plane visible in the low-temperature piezocantilever torque magnetometry measurements. In both cases, the six-fold torque is only detected in the ordered phase. I have been able to rule out the most straightforward origins of the energy profile; single ion anisotropy of the required form is not allowed, and exchange anisotropy up to bilinear terms will not cause such a modulation in energy at the mean field limit. What remains as explanations are more exotic phenomena such as quantum fluctuations lifting the degeneracy of spin orientations in the 𝒂𝒃 plane, or spin-orbital exchange. The data presented in this chapter should provide some additional clues for future theoretical work which seeks to understand the origin of the magnetic anisotropy in these materials, which may in turn improve our understanding of the origins of anisotropy in magnetic systems more broadly.