Article
pubs.acs.org/cm
Cr-Doped TiSe2 − A Layered Dichalcogenide Spin Glass
Huixia Luo,*,† Jason W. Krizan,† Elizabeth M. Seibel,† Weiwei Xie,† Girija S. Sahasrabudhe,†
Susanna L. Bergman,† Brendan F. Phelan,† Jing Tao,‡ Zhen Wang,§ Jiandi Zhang,§ and R. J. Cava*,†
†
Department of Chemistry, Princeton University, Princeton, New Jersey 08544, United States
Department of Condensed Matter Physics and Materials Science, Brookhaven National Laboratory, Upton, New York 11973, United
States
§
Department of Physics and Astronomy, Louisiana State University, Baton Rouge, Louisiana 70803, United States
‡
ABSTRACT: We report the magnetic characterization of the
Cr-doped layered dichalcogenide TiSe2. The temperature
dependent magnetic susceptibilities are typical of those seen
in geometrically frustrated insulating antiferromagnets. The Cr
moment is close to the spin-only value, and the Curie−Weiss
temperatures (θcw) are between −90 and −230 K. Freezing of
the spin system, which is glassy, characterized by peaks in the
ac and dc susceptibility and specific heat, does not occur until
below T/θcw = 0.05. The CDW transition seen in the resistivity
for pure TiSe2 is still present for 3% Cr substitution but is
absent by 10% substitution, above which the materials are
metallic and p-type. Structural refinements, magnetic characterization, and chemical considerations indicate that the materials are of the type Ti1−xCrxSe2‑x/2 for 0 ≤ x ≤ 0.6.
occupy the interstitial octahedral sites in the van der Waals
layers, but at higher concentrations they occupy both the
octahedral interstitial sites and sites in the Ti layers; when they
do that they then displace some Ti into the interstitial positions
and the system becomes quite disordered. (The materials
system is complex: previous studies25,26 of Cr-intercalated
TiSe2 yielded magnetic properties that are significantly different
from those that are observed here, implying that the magnetic
behavior of the system may be dependent on synthetic
conditions.) Cr ions most frequently are ionic and have the
oxidation state Cr3+ in solids due to their very strong Hundsrule coupling.27 In a material like TiSe2, which is a small band
gap semiconductor with a conduction band made from empty
Ti d states and a valence band made from filled Se p states (i.e.,
consisting of Ti4+ and Se2−), the presence of Cr3+ along with
Ti4+ and Se2− requires a decrease in the Se to metal ratio to
below 2:1 and a formula Ti1−xCrxSe2‑x/2 to maintain charge
neutrality. Our XPS analysis supports these formal oxidation
state assignments. This picture for Cr-doped TiSe2 is further
supported by the materials synthesis, the diffraction data, the
magnetic data, and the presence of two types of magnetic spins
clearly seen in the ac susceptibility for higher x materials. The
DC magnetic susceptibilities confirm that the Cr moments in
Ti1−xCrxSe2‑x/2 are within experimental error of the expected
Cr3+ spin-only value.28 The antiferromagnetic Curie−Weiss
temperatures are large, between −90 and −230 K. The freezing
INTRODUCTION
Frustration of magnetic ordering arises in both atomically
disordered systems and systems where the magnetic
interactions are not compatible with the underlying structural
symmetry.1−5 Classical spin glasses and systems with magnetic
ions on triangular or tetrahedral lattices display such frustration.
The most-often-studied materials with strong geometric
magnetic frustration are electrically insulating or at best
strongly semiconducting,5−16 and classical spin glass systems
can be either metallic or semiconducting.17 Typically, frustrated
magnetic materials are characterized by a Curie−Weiss theta
θcw that is significantly greater than the spin freezing
temperature Tf. The frustration index f = θcw/Tf is often
taken as a general characterization of the degree of frustration.
The MX2 layered transition-metal dichalcogenides (TMDCs,
M = Mo, W, V, Nb, Ta, Ti, Zr, Hf, and Re, and X = Se, S, or
Te) are a large family of solids with layered triangular metal
lattices and have long been of interest due to the rich electronic
properties that arise from their low dimensionality (see, e.g. refs
18−22). Within the TMDCs, 1T-TiSe2 has attracted special
attention due to the presence of a Charge Density Wave
(CDW) that onsets at 200 K.22,23 The material has trigonal
symmetry,24 with TiSe6 octahedra sharing edges in triangular
geometry TiSe2 layers that are bonded to each other by Se−Se
van der Waals forces (Figure 1a). Here we study the effect of
Cr substitution for Ti in 1T-TiSe2.
We find that as Cr atoms are substituted for Ti, they induce
metal occupancy in the van der Waals gap between the TiSe2
layers (Figure 1a). We interpret our data to indicate that at
concentrations less than about 20%, the Cr ions primarily
■
© 2015 American Chemical Society
Received: August 11, 2015
Revised: September 16, 2015
Published: September 17, 2015
6810
DOI: 10.1021/acs.chemmater.5b03091
Chem. Mater. 2015, 27, 6810−6817
Chemistry of Materials
Article
Figure 1. Structural and spectroscopic characterization of 1T-Ti1‑xCrxSe2‑x/2. (a and b) The overall crystal structure of 1T-TiSe2 (a), (b)
Composition dependence of the room temperature lattice parameters a and c for Ti1−xCrxSe2‑x/2 (0 ≤ x ≤ 0.6), (c) and (d) refined powder Xray
diffraction data for the two structural models for Ti0.4Cr0.6Se1.7. The superioriy of the model for metal interstitials (Model 2) over selenium vacancies
(Model 1) is seen through comparison of the difference plots. Regions of interest are marked by circles. (e-h) XPS spectra of the Ti 2p, Cr 2p, and Se
3d regions of selected Ti1−xCrxSe2‑x/2 (x = 0.07, 0.6) materials. Vertical red lines indicate the positions for the Ti4+, Se2−, and Cr3+ species in the bulk
material. For comparison, the Se 3d spectrum for single crystal Bi2Te2Se is included in (h).
and heated in sealed evacuated quartz glass tubes at 700 °C for 72 h.
Subsequently, the as-prepared powders were reground, repelletized,
and sintered again at 700 °C for 48 h. Powder X-ray diffraction
(PXRD, Bruker D8 Advance ECO, Cu Kα radiation) was initially used
to structurally characterize the samples. Excess elemental selenium was
present in all the synthesis tubes and easily separated from the sample
pellets by sublimation. The presence of the excess Se in the system
indicated that the Se:M ratio of the obtained products was less than
2:1, as described further below. Single phase powder samples (after the
excess Se was distilled away) were studied by synchrotron X-ray
diffraction at the Advanced Photon Source at Argonne National
Laboratory on beamline 11-BM. All diffraction patterns were refined
using the Rietveld method in the Fullprof software suite.29
The DC magnetization (M) as a function of applied magnetic field
(H) was linear for all samples up to applied fields of μ0H = 1 T above
the spin glass ordering temperature, and thus the magnetic
susceptibility χ was defined as χ = M/μ0H at 1 T. Zero-field cooled
of the spin system, characterized by peaks in the ac and dc
susceptibility, occurs between about 2 and 12 K. Resistivity and
hall effect measurements on polycrystalline pellets show the
suppression of the CDW state in TiSe2 for low values of Cr
substitution and metallic, p-type conductivity at higher levels of
substitution. The complexity of the material system precludes
the determination of a detailed model for the magnetism at this
stage, but the general features are well described by a picture of
magnetic disorder on a triangular lattice.
■
EXPERIMENTAL SECTION
Polycrystalline samples of target composition Ti1−xCrxSe2 were
synthesized by solid state reaction. Mixtures of high-purity fine
powders of Ti (99.9%), Cr (99.95%), and Se (99.999%) in the
appropriate stoichiometric ratios were thoroughly ground, pelletized,
6811
DOI: 10.1021/acs.chemmater.5b03091
Chem. Mater. 2015, 27, 6810−6817
Chemistry of Materials
Article
(ZFC) dc magnetization measurements to obtain χ were performed on
heating from 1.8 to 300 K in a magnetic field of 1 T in a Quantum
Design superconducting quantum interference device (SQUID)
equipped Magnetometer (MPMS-XL-5). Temperature dependent
magnetizations for selected high Cr content samples were also
measured in the MPMS with an applied field of 200 Oe to allow the
observation of the bifurcation of the zero field cooled and field cooled
magnetizations at the spin freezing transition. Measurements of the
temperature dependence of the electrical resistivity, ac magnetic
susceptibility, and heat capacity were performed in a Quantum Design
Physical Property Measurement System (PPMS). Resistivities and heat
capacities below 2 K for selected compositions were measured in the
PPMS equipped with a 3He cryostat. The nonmagnetic analogue TiSe2
was synthesized and used for the subtraction of the phonon
contribution to estimate the magnetic contribution to specific to the
specific heat for the Cr-doped TiSe2 materials.
X-ray Photoelectron Spectroscopy (XPS) characterization was
performed with a VG ESCA Lab Mk.II instrument. All spectra were
obtained using Mg Kα radiation (1284 eV) and 20 eV pass energy.
The sample powder was placed on carbon tape attached to the metal
sample holder. To avoid charging effects during XPS, a positive bias of
10 V was applied to the holder during spectral acquisition.30,31 Spectra
were taken for polycrystals of Ti0.93Cr0.07Se1.965, Ti0.75Cr0.25Se1.875,
Ti0.4Cr0.6Se1.7, and single crystal Bi2Te2Se (BTS). The Se 3d XPS for
single crystal BTS was used for calibration purposes. All scans were
taken with a 0.05 eV step size and 0.5 s dwell time. The obtained scans
were fit with the Casa XPS software using a Shirley background; area
and positions were constrained using standard values.
Specimens for transmission electron microscopy (TEM) were
obtained from synthesized samples crushed in a drybox and
transported to the microscope in ultrahigh vacuum. Scanning TEM
(STEM) in high-angle annular dark-field (HAADF) mode, energydispersive X-ray spectroscopy (EDXS), electron energy-loss spectroscopy (EELS), selected area electron diffraction (SAED), and highresolution TEM (HRTEM) were conducted in a JEOL 2100F
microscope equipped with a liquid-helium cooled sample holder at
Brookhaven National Laboratory.
intercalation corresponds to the value appropriate to the
material composition. The superiority of this model for
describing the structures of the Ti1−xCrxSe2‑x/2 materials is
clearly seen by comparison of Figures 1c and 1d. The fits to the
synchrotron X-ray pattern for Ti0.6Cr0.4Se1.8 for the interstitial
occupancy model (full metal site occupancy in the Ti layer by
both Ti and Cr, interstitial ions, and no selenium vacancies)
and the alternative Se vacancy model model (full metal site
occupancy in the Ti layer by both Ti and Cr, no interstitial ions,
and selenium vacancies) are compared. At higher Cr contents
than those studied here, the related compound Cr4TiSe8 has
been reported to have two magnetic transitions, at Tf = 50 K
and TN = 120 K.34 In our magnetic measurements, we did not
see those transitions in any samples (see Figure 3) indicating
that this compound is not present as an impurity in our
materials. The diffraction data analysis, the requirements of
charge neutrality, and the presence of excess Se in the synthetic
system indicate that the best way to represent the composition
of the fabricated compounds is Ti1−xCrxSe2‑x/2, which we use
here. (An equivalent, though more complex formula reflecting
the observed crystal structure, which shows the presence of
both interstitial metal atoms and full Se site occupancy, would
be Cry/3(Ti1−yCry)Se2, where the leading Cr is intercalated and
the metal atoms in parentheses are those in the normal
dichalcogenide layers; the relation between this and the simpler
formula employed here is [y + y/3]/[1+y/3] = x.)
X-ray photoelectron spectroscopy (XPS) was performed on
Ti1−xCrxSe2‑x/2 (x = 0.07, 0.25, and 0.6) to compare the effects
of Cr doping on the oxidation state of Ti and Cr. The XPS of
Ti 2p (Figure 1e) shows the presence of two Ti species. The Ti
2p3/2 peak at 458.7 eV, corresponding to TiO235,36 is an
indication of the oxidation of the surface of the polycrystals
from exposure to ambient conditions, while the Ti 2p3/2 peak at
455.5 eV originates from the bulk material.36,37 The remaining
two peaks (Ti 2p1/2), at 464.9 and 461.7 eV, are the result of
spin orbit coupling. Similarly, the XPS of Cr 2p (Figure 1f)
shows two Cr 2p3/2 peaks: The peak at 576.0 eV corresponds to
Cr2O3 formed due to air oxidation,38,39 while the second peak
at 573.7 eV is representative of the bulk material.40 The 2p1/2
peaks are located at 586.2 and 584.0 eV. The Se 3d XPS (Figure
1g), when deconvoluted using standard restriction parameters,
indicates the presence of two major Se species. The Se 3d peak
at 53.2 eV originates from the bulk material, while the peak at
54.3 eV is the result of surface oxidation in air. The binding
energies of Ti, Cr, and Se for the bulk material were calibrated
using the Se XPS of single crystal BTS (Figure 1h). Due to the
layered structure of single crystal BTS,35 the Se is never
exposed to air which prevents oxidation. As the XPS of single
crystal BTS exhibits only one Se species with a well-defined Se
3d peak (fwhm =1.6 eV) at 53.2 eV, it makes a reliable
calibration standard. We find that the binding energies of Ti
and Cr in the bulk do not change with the chromium doping
level. This observation corroborates well with other measurements. Thus, the XPS characterization of the Ti1−xCrxSe2‑x/2
materials over a wide range of x spectroscopically confirms the
formal Cr3+, Ti4+, and Se2− oxidation state assignments.
For more detailed structural characterization, in order to
check for possible Cr clustering, deviation from random
subsititution, short-range Cr−Ti ordering, or nanoscale
precipitates in Cr doped TiSe2, which would impact the
magnetism of the system, high-resolution transmission electron
microscopy (HRTEM) was used to examine one of the high Cr
content materials, Ti0.6Cr0.4Se1.8. Figure 2a shows the crystal
RESULTS AND DISCUSSION
Figure 1b shows the composition dependence of the room
temperature lattice parameters for Ti1−xCrxSe2‑x/2 (0 ≤ x ≤
0.6). x = 0.6 is the high Cr composition limit of the intercalated
TiSe2-like solid solution. Surprisingly, for 0 ≤ x ≤ 0.2, a
decreases slightly and c decreases slightly. For x ≥ 0.2, in
contrast, the unit cell parameter a increases with Cr content,
and c decreases more quickly. Thus, for Cr-doped TiSe2, the
composition-dependent unit cell parameters do not follow
Vegard’s law. This is the fundamental characteristic showing
that there is a change in the structure of the system near x =
0.2. The complexity is due to the way that the Cr atoms are
accommodated in the structure−on Cr substitution extra metal
atoms are found in the van der Waals gap between in the layers,
as described below.
At low Cr doping levels, the powder diffraction data were not
very sensitive to the presence of intercalated metals,32 but for
the higher Cr doping levels, quantitative tests of different
structural models were possible. In these tests we observed
clearly that the Se sites are fully occupied. This is consistent
with what has been found for all layered dichalcogenides of the
type MxMX2 where the X:M ratio is less than 2:1; on doping,
metal interstitials are found in the van der Waals gap rather
than vacancies in the close packed chalcogen planes.33 Further,
free refinement of the occupancy of the interstitial metal sites
led to occupancies that were within the standard error of
expectations for the nominal compositions. Thus, we conclude
that Se sites are fully occupied and that the level of metal
■
6812
DOI: 10.1021/acs.chemmater.5b03091
Chem. Mater. 2015, 27, 6810−6817
Chemistry of Materials
Article
Figure 2. High-resolution transmission electron microscopy (HETEM) characterization of 1T-Ti0.6Cr0.4Se1.8 showing the random Ti−Cr Solid
Solution (a) crystal model in the [100] zone; (b) and (c) HAADF-STEM image in the [100] zone at different magnifications; (d) HAADF-STEM
image in the [001] zone; (e)-(i) Electron energy loss spectroscopy (EELS) spectra (g-i) obtained at each pixel during scanning in the boxed area
(e,f). Intensities of the core-loss edge of each element were integrated and mapped in the scanning area to show the elemental distribution at the
atomic level. No segregation or short-range ordering of Cr was found.
Ti0.6Cr0.4Se1.8, even at the atomic level. Therefore, there is a
homogeneous random distribution of Cr within the structure.
Figure 3a shows the zero-field cooled temperature (T)
dependence, from 1.8 to 300 K, of the dc magnetic
susceptibility (χ) for the Ti1−xCrxSe2‑x/2 polycrystalline samples.
The plot of χ versus T in Figure 3a shows increasing χ with
decreasing T for all samples, and the inset shows that the
magnetic susceptibility increases linearly with Cr concentration
at a fixed temperature; this is a qualitative indication of the fact
that the Cr has a constant magnetic moment across the series.
Figure 3b shows the character of the AFM transition under the
relatively low applied field of 200 Oe. The antiferromagnetic
(AFM) transition temperature increases with increasing Cr
content for x ≤ 0.4. For higher Cr doping, however, the AFM
transition, while still seen, becomes broad and poorly defined.
Figure 4a shows the temperature dependent inverse
susceptibilities, 1/(χ − χ0) vs T, constructed from the data
shown in Figure 3a. In the high temperature region, above 200
K, all magnetic susceptibilities (χ) can be fit to χ − χ0 = C/(T −
structure model for the [100] zone for this material, where the
Ti/Cr atoms (gray balls) are in octahedral coordination with Se
(yellow balls). Figures 2b and 2c show the HAADF-STEM
image for the [100] zone at different magnifications, and the
inset of Figure 2b, top, is the selected area electron diffraction
(SEAD) pattern along a [100] zone axis. Figure 2d shows the
HAADF-STEM image for the [001] zone, and the inset of
Figure 2d on top is the SAED pattern in the [001] zone. The
STEM-HAADF micrographs in Figure 2b-d as well as the
SAED patterns reveal a homogeneous Cr distribution in
Ti0.6Cr0.4Se1.8 at the atomic level; no clusters, short-range order,
or nanoscale Cr-precipitates were observed. Figures 2d-i show
the EELS spectra obtained at each pixel during scanning in the
boxed area. Intensities of the core-loss edge of each element
were integrated and mapped in the scanning area to show the
Ti, Cr, and Se elemental distribution at the atomic level. The
element maps images further confirm that there is no
segregation, short-range ordering, or clustering of Cr in
6813
DOI: 10.1021/acs.chemmater.5b03091
Chem. Mater. 2015, 27, 6810−6817
Chemistry of Materials
Article
Figure 3. General magnetic characterization of Ti1‑xCrxSe2‑x/2 (a)
Magnetic susceptibility (χ) versus temperature for Ti1−xCrxSe2‑x/2 (x =
0, 0.07, 0.15, 0.2, 0.25, 0.3, 0.4, 0.5, 0.6). Inset χ at 200 K vs x for
Ti1−xCrxSe2‑x/2. (b) Normalized χ/χmax versus temperature in a 200 Oe
field for Ti1−xCrxSe2‑x/2 (x = 0.25, 0.3, 0.4, 0.5, 0.6).
Figure 4. Curie−Weiss plots for Ti1‑xCrxSe2‑x/2. (a) Inverse
susceptibility (1/χ−χ0) versus temperature in a 1 T field for
Ti1−xCrxSe2‑x/2. Inset: effective moment and Curie−Weiss temperatures. (b) Plot of C/|θcw|(χ−χ0) versus T/|θcw| for Ti1−xCrxSe2‑x/2 (x =
0.03, 0.07, 0.15, 0.2, 0.25, 0.3, 0.4, 0.5). This plot shows that these
compounds exhibit consistent Curie behavior at high temperatures.
θCW), where C is the Curie constant, θCW is the Curie−Weiss
temperature, and χ0 is the temperature independent contribution to the susceptibility. The fits were performed in the
temperature range between 200 and 300 K (as is commonly
done for frustrated magnetic materials, the fits are best
performed for temperatures above θCW); linear relationships
(shown as solid lines in Figure 4a) were found for all
Ti1−xCrxSe2‑x/2 compounds above 200 K. χ0 is negligibly small
and set to 0 for all samples with the exception of the sample
with the lowest Cr content studied, x = 0.03, where it is
0.00005. The effective magnetic moment (Peff) per Cr ion can
be obtained by using Peff = (8C)1/2. The Curie−Weiss
temperature is an estimate of the net magnetic interaction
strength; the θcws are larger in magnitude than −90 K for all
materials. In Ti0.75Cr0.25Se1.875, for example, θcw is −174 K, and
no magnetic transition is apparent until approximately 2.5 K,
giving this material a frustration index f of ∼70, indicating that
Ti0.75Cr0.25Se1.875 is strongly frustrated. The effective moment
per Cr, Peff, is ∼4 μB and is observed to be only weakly
dependent on Cr concentration. The θCW values, however,
increase in magnitude with increasing Cr content (see the inset
of Figure 4a, top).
To better compare the magnetic characteristics of all
members of the Ti1−xCrxSe2‑x/2 (0.03 ≤ x ≤ 0.6) family, we
rearrange the Curie−Weiss Law to the normalized form C/
(χ|θcw|) = T/|θcw|−1. Plots of the magnetic data in this form are
especially useful in comparing the general behavior of
geometrically frustrated magnets.1,2,41 The result is a
dimensionless plot of the normalized inverse susceptibility
against normalized temperature for all samples - the
susceptibility is scaled by the magnitude of the moments (C)
and the temperature is normalized by the strengths of the
magnetic interactions (|θcw|).33 Ideal antiferromagnets would
have a slope of 1 and intercept of the y axis at 1 in this
representation (shown as a solid green line y = x − 1 in Figure
4b), with indications of magnetic ordering on the order of T/
θcw in the range of 0.5 to 1. Antiferromagnetic or ferromagnetic
correlations at lower temperatures in excess of those expected
for simple Curie−Weiss behavior are manifested as positive or
negative deviations respectively from the green solid line y = x
− 1, respectively. This allows differences in the nature of
correlations above TN to be identified and easily compared
among different materials. It can be seen in the figure that the
Ti1−xCrxSe2‑x/2 system maintains nearly ideal Curie−Weiss
behavior to normalized temperatures near 1.2 T/θcw before
exhibiting increased ferromagnetic fluctuations just before its
spin freezing transition, which occurs well below a T/θcw of
0.05. The plot shows that the higher Cr doping compositions
show larger relative ferromagnetic deviations from the green
line. The ferromagnetic deviations from the antiferromagnetic
6814
DOI: 10.1021/acs.chemmater.5b03091
Chem. Mater. 2015, 27, 6810−6817
Chemistry of Materials
Article
Figure 5. Spin glass characterization of Ti1‑xCrxSe2‑x/2. (a) Temperature dependence of the dc susceptibility in an applied field of 200 Oe for
Ti0.5Cr0.5Se1.75. (b) Temperature dependence of the ac susceptibility in an applied field of 20 Oe for Ti0.5Cr0.5Se1.75 as a function of frequency. (c) MH curve for Ti0.5Cr0.5Se1.75 at 2 and 20 K. (d) The behavior parametrized in a fit to the Volger-Fulcher law.
ideal glass temperature”.43,44 After rearranging the VolgerFulcher law, a simple relation between Tf and f with the
presentation of Tf = T0 − (Ea/kb)(1/(ln(τ0 f))) can be used.
Figure 5d shows the resulting fits. The intrinsic relaxation time
(τ0) cannot be fitted for the current data, and thus selected
values were used varying from 1 × 10−7 s (superparamagnets,
cluster glasses)45 to 1 × 10−13 s (conventional spin glasses).46
When τ0 was set to a value of 1 × 10−7 s, we obtain Ea = 7.17 ×
10−4 eV and T0 = 8.48 K, yielding Ea/kB = 2.14. Setting the
intrinsic relaxation time τ0 to be the smallest value, 1 × 10−13 s,
we obtain Ea = 4.57 × 10−3 eV, T0 = 7.10 K, and the value of
Ea/kB = 53. Finally, if τ0 = 1 × 10−12 s, a midrange value, is
assumed, we obtain Ea = 3.70 × 10−3 eV, T0 = 7.33 K. and the
ratio Ea/kB = 43.
Further characterization of the freezing of the spins was
performed via heat capacity measurements. Figure 6 shows in
the main panel the raw heat capacity data for Ti1−xCrxSe2‑x/2 (x
= 0, 0.25, 0.5), presented as Cp/T vs T for T between 2 and
120 K. Estimates of the magnetic specific heat can be obtained
by performing a subtraction of the TiSe2 data from that of the
Cr doped materials after normalization by 2% such that the
Cp/T values match at temperatues of 100 K and higher. This
subtracted data is shown in the inset to Figure 6. A sharp
ordering feature is not seen in the heat capacity, consistent with
a picture where the spins freeze in a random configuration.47
Rough integration of the magnetic heat capacities indicates that
the total integrated entropy values are low compared to those
expected for two-state or S = 3/2 Heisenberg systems, (R ln(2)
or R ln(2S+1)) and thus that the spin freezing observed in the
magnetic susceptibility does involve all of the spins; there may
be considerable residual magnetic entropy in the Ti1−xCrxSe2‑x/2
system below 2 K.
Finally, the temperature dependence of the electrical
resistivities, plotted as the ρ/ρ300 K ratios for polycrystalline
Ti1−xCrxSe2‑x/2 (0.03 ≤ x ≤ 0.6) are shown in Figure 7. All the
Cr-doped samples have resistivities below 6 mOhm cm at 300
Curie−Weiss law in the frustrated regime, at T/θcw = 0.1, are
illustrated in the inset of Figure 4b.
The low temperature dc and ac susceptibilities of
Ti0.5Cr0.5Se1.75 near the magnetic freezing transition are
presented in Figure 5. Figure 5a shows the dc susceptibility
under an applied 200 Oe field, which indicates that the
transition temperature is 8−9 K. Figure 5b presents the ac
susceptibility under an applied field of 20 Oe at different
frequencies. On this plot, we can see that the transition
temperature Tf shifts to higher temperature as the frequency of
the ac field increases (10 Hz data omitted for clarity), which is a
characteristic trait of spin glasses.
Small hysteresis is observed in the M(H) data at 2 K, but no
hysteresis is observed at 20 K recorded during increasing and
decreasing of magnetic field. In addition, it can be seen that the
data shows slight curvature but is nowhere near saturation
within the accessible field range. A broad peak is seen in the ac
susceptibility data in the 7−12 K temperature range, characteristic of spin freezing. Figure 5c shows the field-dependent
magnetization (as M vs μ0H) at 2 K. The ratio of the shift in
transition temperature (ΔTf) to the transition temperature (Tf)
times the log of the change in frequency (Δlogf) used in the
expression of ΔTf/(TfΔlogf), which parametrizes the dependence Tf on f, can be used to characterize spin glasses and spin
glass like materials.42 Based on the ac susceptibilities, taking the
transition temperature as the maximum of χ′ in Figure 5b, we
obtain p = ΔTf/(TfΔlogf) = 0.004, which is much smaller, for
example, than that of insulating pyrochlore NaCaCo2F7 (where
p = 0.029); however, it is very similar to what is seen for
metallic alloy spin glasses such as MMn (M = Cu, Au, Ag,
around 0.005).34 To further parametrize the spin glass behavior,
the frequency (f) dependence of Tf can be fitted by the
empirical Volger-Fulcher law with the following equation: τ =
1/f = τ0 exp(Ea/(kb(Tf − T0))), where Tf is the freezing
temperature, f is the frequency, τ0 is the intrinsic relaxation
time, Ea is the activation energy of the process, and T0 is “the
6815
DOI: 10.1021/acs.chemmater.5b03091
Chem. Mater. 2015, 27, 6810−6817
Chemistry of Materials
Article
clusters or Cr-rich particles in this system. XPS characterization
supports oxidation state assignments of Cr3+, Ti4+, and Se2− for
these materials. The X-ray diffraction refinements indicate that
Cr goes into the van der Waals gap positions. Based on the
partial substitution of Cr for Ti in Ti1−xCrxSe2‑x/2, Cr3+ on the
triangular-planar geometry lattice leads to the frustration of the
magnetic ordering: we have observed no long-range magnetic
ordering at low temperatures. The Curie−Weiss temperatures
θcw were much greater than the freezing temperatures Tfs,
confirming that the system is highly frustrated. The system
appears to be similar to systems such as Sr1−xEuxS,49 where
both disorder-induced and geometry-induced frustration of the
magnetic ordering are present, with the exception that the
present system is metallic with low carrier concentration. More
detailed studies of the charge transport and magnetism in this
complex system would be of further interest.
Figure 6. Heat capacity characterization of Ti1‑xCrxSe2‑x/2. Main panel:
heat capacity of Ti1−xCrxSe2 (x = 0, 0.25, 0.5) in the form of Cp/T
over a wide temperature range. The presence of extra entropy at low
temperarture for the Cr-doped materials is clearly seen. Inset: The
temperature dependence of the excess heat capacity of Ti1−xCrxSe2 (x
= 0.25, 0.5), determined by the subtraction of the heat capacity of
TiSe2. This heat capacity must be a reflection of the spin freezing in
the system.
■
AUTHOR INFORMATION
Corresponding Authors
*E-mail:
[email protected] (R.J.C.).
*E-mail:
[email protected] (H.X.L.).
Notes
The authors declare no competing financial interest.
ACKNOWLEDGMENTS
The synthesis and magnetic characterization of the materials
was supported by the DOE grant FG02-98ER45706. The DOE
supported the powder diffraction work of J.K. through grant DE
FG02-08ER46544. The electron diffraction study at Brookhaven National Laboratory was supported by the DOE BES,
by the Materials Sciences and Engineering Division under
contract DE-AC02-98CH10886, and through the use of the
Center for Functional Nanomaterials. Z.W. was supported by
U.S. DOE under Grant No. DOE DE-SC0002136.
■
■
REFERENCES
(1) Greedan, J. E. Geometrically frustrated magnetic materials. J.
Mater. Chem. 2001, 11, 37−53.
(2) Moessner, R.; Ramirez, A. R. Geometrical frustration. Phys. Today
2006, 59, 24−29.
(3) Collins, M. F.; Petrenko, O. A. Triangular antiferromagnets. Can.
J. Phys. 1997, 75, 605−655.
(4) Gardner, J. S.; Gingras, M. J. P.; Greedan, J. E. Magnetic
pyrochlore oxides. Rev. Mod. Phys. 2010, 82, 53−107.
(5) Ramirez, A. P. Strongly Geometrically Frustrated Magnets. Annu.
Rev. Mater. Sci. 1994, 24, 453−480.
(6) Bramwell, S. T.; Gingras, M. J. P. Spin Ice State in Frustrated
Magnetic Pyrochlore Materials. Science 2001, 294, 1495−1501.
(7) Fennell, T.; Deen, P. P.; Wildes, A. R.; Schmalzl, K.; Prabhakaran,
D.; Boothroyd, A. T.; Aldus, R. J.; McMorrow, D. F.; Bramwell, S. T.
Magnetic Coulomb Phase in the Spin Ice Ho2Ti2O7. Science 2009, 326,
415−417.
(8) Morris, D. J. P.; Tennant, D. A.; Grigera, S. A.; Klemke, B.;
Castelnovo, C.; Moessner, R.; Czternasty, C.; Meissner, M.; Rule, K.
C.; Hoffmann, J. U.; Kiefer, K.; Gerischer, S.; Slobinsky, D.; Perry, R.
S. Dirac Strings and Magnetic Monopoles in the Spin Ice Dy2Ti2O7.
Science 2009, 326, 411−414.
(9) Krizan, J. W.; Cava, R. J. NaCaCo2F7: A single-crystal hightemperature pyrochlore antiferromagnet. Phys. Rev. B: Condens. Matter
Mater. Phys. 2014, 89, 214401−214405.
(10) Kemei, M. C.; Moffitt, S. L.; Darago, L. E.; Seshadri, R.;
Suchomel, M. R.; Shoemaker, D. P.; Page, K.; Siewenie, J. Structural
ground states of (A,A′)Cr2O4 (A = Mg, Zn; A′ = Co, Cu) spinel solid
solutions: Spin-Jahn-Teller and Jahn-Teller effects. Phys. Rev. B:
Condens. Matter Mater. Phys. 2014, 89, 174410−174424.
Figure 7. Characterization of the resistivity of Ti1‑xCrxSe2‑x/2. The
temperature dependence of the ratios (ρ/ρ300 K) for polycrystalline
pellets of Ti1−xCrxSe2‑x/2 (0 ≤ x ≤ 0.6).
K. The ratios (ρ/ρ300 K) for Ti1−xCrxSe2‑x/2 (0.03 ≤ x ≤ 0.6)
were less than 1 (Figure 7a), except for the lowest doping (x =
0.03) case. For x = 0.03, an increase and then decrease in the
resistivity is seen on cooling. This behavior is similar in
character to what is observed for undoped TiSe2 (see Figure 7)
but with the increase in resistivity on cooling from 300 K,
which has been associated with the charge density wave in
TiSe2,22 significantly depressed. This observation is similar to
that seen for Ti1−xPtxSe2‑y at low Pt doping levels.48 At higher
Cr doping the Ti1−xCrxSe2‑x/2 compounds become more
metallic, which is different from the behavior of the
Ti1−xPtxSe2‑y compounds with higher Pt doping−they become
more insulating.48 The Cr-doped TiSe2 samples are found to be
p-type by Hall coefficient measurements (not shown) with
carrier concentrations in the range of (2−7) × 1020 cm−3,
typical of those found for heavily doped small band gap
semiconductors, as expected.49
CONCLUSIONS
Cr-substituted TiSe2 materials of the form Ti1−xCrxSe2‑x/2 (0.03
≤ x ≤ 0.6) have been synthesized successfully via a solid state
reaction method. TEM studies confirm that there are no Cr
■
6816
DOI: 10.1021/acs.chemmater.5b03091
Chem. Mater. 2015, 27, 6810−6817
Chemistry of Materials
Article
(11) Aeppli, G.; Broholm, C.; Ramirez, A.; Espinosa, G. P.; Cooper,
A. S. Broken spin rotation symmetry without magnetic Bragg peaks in
Kagome antiferromagnets. J. Magn. Magn. Mater. 1990, 90−91, 255−
259.
(12) Ramirez, A. P.; Espinosa, G. P.; Cooper, A. S. Strong frustration
and dilution-enhanced order in a quasi-2D spin glass. Phys. Rev. Lett.
1990, 64, 2070−2073.
(13) Obradors, X.; Labarta, A.; Isalgué, A.; Tejada, J.; Rodriguez, J.;
Pernet, M. Magnetic frustration and lattice dimensionality in
SrCr8Ga4O19. Solid State Commun. 1988, 65, 189−192.
(14) Rigol, M.; Singh, R. R. P. Magnetic Susceptibility of the Kagome
Antiferromagnet ZnCu3(OH)6Cl2. Phys. Rev. Lett. 2007, 98, 207204.
(15) Helton, J. S.; Matan, K.; Shores, M. P.; Nytko, E. A.; Bartlett, B.
M.; Yoshida, Y.; Takano, Y.; Suslov, A.; Qiu, Y.; Chung, J. H.; Nocera,
D. G.; Lee, Y. S. Spin Dynamics of the Spin-1/2 Kagomé Lattice
Antiferromagnet ZnCu3(OH)6Cl2. Phys. Rev. Lett. 2007, 98, 107204.
(16) Shores, M. P.; Nytko, E. A.; Bartlett, B. M.; Nocera, D. G. A
Structurally Perfect S = 1/2 Kagomé Antiferromagnet. J. Am. Chem.
Soc. 2005, 127, 13462−13463.
(17) Guo, Y.; Dai, J.; Zhao, J. Y.; Wu, C. Z.; Li, D. Q.; Zhang, L. D.;
Ning, W.; Tian, M. L.; Zeng, X. C.; Xie, Y. Large Negative
Magnetoresistance Induced by Anionic Solid Solutions in TwoDimensional Spin-Frustrated Transition Metal Chalcogenides. Phys.
Rev. Lett. 2014, 113, 157202.
(18) Wilson, J. A.; Yoffe, A. D. The transition metal dichalcogenides
discussion and interpretation of the observed optical, electrical and
structural properties. Adv. Phys. 1969, 18, 193−335.
(19) Ugeda, M. M.; Bradley, A. J.; Shi, S. F.; da Jornada, F. H.; Zhang,
Y.; Qiu, D. Y.; Ruan, W.; Mo, S. K.; Hussain, Z.; Shen, Z. X.; Wang, F.;
Louie, S. G.; Crommie, M. F. Giant bandgap renormalization and
excitonic effects in a monolayer transition metal dichalcogenide
semiconductor. Nat. Mater. 2014, 13, 1091−1095.
(20) Soumyanarayanan, A.; Yee, M. M.; He, Y.; van Wezel, J.; Rahn,
D. J.; Rossnagel, K.; Hudson, E. W.; Norman, M. R.; Hoffman, J. E.
Quantum phase transition from triangular to stripe charge order in
NbSe2. Proc. Natl. Acad. Sci. U. S. A. 2013, 110, 1623−1627.
(21) Luo, H. X.; Xie, W. W.; Tao, J.; Inoue, H.; Gyenis, A.; Krizan, J.
W.; Yazdani, A.; Zhu, Y. M.; Cava, R. J. Polytypism, polymorphism,
and superconductivity in TaSe2‑xTex. Proc. Natl. Acad. Sci. U. S. A.
2015, 112 (11), E1174−E1180.
(22) Di Salvo, F. J.; Moncton, D. E.; Waszczak, J. V. Electronic
properties and superlattice formation in the semimetal TiSe2. Phys.
Rev. B 1976, 14, 4321−4328.
(23) Stoffel, N. G.; Kevan, S. D.; Smith, N. V. Experimental band
structure of 1T-TiSe2 in the normal and charge-density-wave phases.
Phys. Rev. B: Condens. Matter Mater. Phys. 1985, 31, 8049−8055.
(24) Oftedal, I. Roentgenographische Untersuchungen von SnS2,
TiS2, TiSe2, TiTe2. Z. Phys. Chem. 1928, 134, 301−310.
(25) Pleschov, V. G.; Baranov, N. V.; Titov, A. N.; Inoue, K.;
Bartashevich, M. I.; Goto, T. Magnetic Properties of Cr-intercalated
TiSe2. J. Alloys Compd. 2001, 320, 13−17.
(26) Shkvarin, A. S.; Merentsov, A. I.; Yarmoshenko, Y. M.; Skorikov,
N. A.; Titov, A. N. Electronic Structure and Magnetic State of TiSe2
Doped with Cr by Means of Intercalation or Substitution of Ti. Solid
State Phenom. 2010, 168−169, 380−383.
(27) Georges, A.; Medici, L. d.; Mravlje, J. Strong Correlations from
Hund’s Coupling. Annu. Rev. Condens. Matter Phys. 2013, 4, 137−178.
(28) Myers, H. P. Introductory Solid State Physics; Taylor and Francis:
London, 1997; p 336.
(29) Rodríguez-Carvajal, J. Recent developments of the program
FULLPROF. Comm. Powder Diffr. 2001, 26, 12−19.
(30) Roudebush, J. H.; Sahasrabudhe, G.; Bergman, S. L.; Cava, R. J.
Rhombohedral Polytypes of the Layered Honeycomb Delafossites
with Optical Brilliance in the Visible. Inorg. Chem. 2015, 54, 3203−
3210.
(31) Suzer, S. Differential Charging in X-ray Photoelectron
Spectroscopy: A Nuisance or a Useful Tool? Anal. Chem. 2003, 75,
7026−7029.
(32) A powder X-ray refinement cannot differentiate unambiguously
between Cr and Ti as their scattering factors are very similar. Although
the structure is well determined, the ratio of Ti to Cr in the two metal
sites (i.e., in the van der Waals gap and in the Ti-plane) could not be
specified from the structural refinement of the diffraction data.
(33) Beeby, J. L.; Bhattacharya, P. K.; Gravelle, P. Ch.; Koch, F.;
Lockwood, D. J. Condensed Systems of Low Dimensionality; Springer:
1991; Vol. 253, pp 677−678.
(34) Bensch, W.; Sander, B.; Näther, C.; Kremer, R. K.; Ritter, C.
Synthesis, crystal structure, magnetic properties and spin glass
behaviour of the new ternary compound Cr4TiSe8. Solid State Sci.
2001, 3, 559−568.
(35) Thomas, C. R.; Sahasrabudhe, G.; Kushwaha, S. K.; Xiong, J.;
Cava, R. J.; Schwartz, J. Topological surface states of Bi2Te2Se are
robust against surface chemical modification. Phys. Status Solidi RRL
2014, 8, 997−1002.
(36) Bhatt, R.; Bhattacharya, S.; Basu, R.; Ahmad, S.; Chauhan, A. K.;
Okram, G. S.; Bhatt, P.; Roy, M.; Navaneethan, M.; Hayakawa, Y.;
Debnath, A. K.; Singh, A.; Aswal, D. K.; Gupta, S. K. Enhanced
Thermoelectric Properties of Selenium-Deficient Layered TiSe2−x: A
Charge-Density-Wave Material. ACS Appl. Mater. Interfaces 2014, 6,
18619−18625.
(37) Shkvarin, A. S.; Yarmoshenko, Y. M.; Yablonskikh, M. V.;
Skorikov, N. A.; Merentsov, A. I.; Kuznetsov, M. V.; Titov, A. N.
MxTiSe2 (M = Cr, Mn, Cu) electronic structure study by methods of
resonance X-ray photoemission spectroscopy and X-ray absorption
spectroscopy. J. Phys. Chem. Solids 2012, 73, 1562−1565.
(38) Ikemoto, I.; Ishii, K.; Kinoshita, S.; Kuroda, H.; Alario Franco,
M. A.; Thomas, J. M. X-ray photoelectron spectroscopic studies of
CrO2 and some related chromium compounds. J. Solid State Chem.
1976, 17, 425−430.
(39) Xu, C.; Hassel, M.; Kuhlenbeck, H.; Freund, H. J. Adsorption
and reaction on oxide surfaces: NO, NO2 on Cr2O3(111)/Cr(110).
Surf. Sci. 1991, 258, 23−24.
(40) Agostinelli, E.; Battistoni, C.; Fiorani, D.; Mattogno, G.;
Nogues, M. An XPS Study of the electronic structure of the
ZnxCd1‑xCr2X4 (X = S, Se) spinel system. J. Phys. Chem. Solids 1989,
50, 269−272.
(41) Dutton, S. E.; Hanson, E. D.; Broholm, C. L.; Slusky, J. S.; Cava,
R. J. Magnetic properties of hole-doped SCGO, SrCr8Ga4−xMxO19 (M
= Zn, Mg, Cu). J. Phys.: Condens. Matter 2011, 23, 386001.
(42) Mydosh, J. A.; Ebrary, I. Spin Glasses: An Experimental
Introduction; Taylor & Francis: London, 1993.
(43) Shtrikman, S.; Wohlfarth, E. P. The theory of the Vogel-Fulcher
law of spin glasses. Phys. Lett. A 1981, 85, 467−470.
(44) Aharoni, A. The Vogel-Fulcher law of spin glasses. Phys. Lett. A
1983, 99, 458−460.
(45) Bhowmik, R. N.; Ranganathan, R. Anomaly in cluster glass
behaviour of Co0.2Zn0.8Fe2O4 spinel oxide. J. Magn. Magn. Mater. 2002,
248, 101−111.
(46) Tholence, J. L. On the frequency dependence of the transition
temperature in spin glasses. Solid State Commun. 1980, 35, 113−117.
(47) Ramirez, A. P.; Hayashi, A.; Cava, R. J.; Siddharthan, R.; Shastry,
B. S. Zero-point entropy in ’spin ice’. Nature 1999, 399, 333−339.
(48) Chen, J. S.; Wang, J. K.; Carr, S. V.; Vogel, S. C.; Gourdon, O.;
Dai, P.; Morosan, E. Chemical tuning of electrical transport in
Ti1−xPtxSe2−y. Phys. Rev. B: Condens. Matter Mater. Phys. 2015, 91,
045125.
(49) Maletta, H. Spin glass properties of (Eu,Sr)S. J. Phys. Colloq.
1980, 41, C5-115−C5-125.
6817
DOI: 10.1021/acs.chemmater.5b03091
Chem. Mater. 2015, 27, 6810−6817