You are on page 1of 10

pubs.acs.

org/IC Article

Water Molecule-Induced Reversible Magnetic Switching in a Bis-


Terpyridine Cobalt(II) Complex Exhibiting Coexistence of Spin
Crossover and Orbital Transition Behaviors
Fumiya Kobayashi, Yuki Komatsumaru, Ryohei Akiyoshi, Masaaki Nakamura, Yingjie Zhang,
Leonard F. Lindoy, and Shinya Hayami*
Cite This: https://dx.doi.org/10.1021/acs.inorgchem.0c00818 Read Online
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ACCESS
Downloaded via AUCKLAND UNIV OF TECHNOLOGY on October 4, 2020 at 08:45:16 (UTC).

Metrics & More Article Recommendations *


sı Supporting Information

ABSTRACT: The development of molecule-based switchable


materials remains an important challenge in the field of molecular
science. Achievement of a structural phase transition induced by
adsorption/desorption of guest molecules in spin crossover (SCO)
Co(II) compounds is of significant interest because of the
possibility that the spin state of the magnetic anisotropic high-
spin (HS, S = 3/2) and low-spin (LS, S = 1/2) states can be
switched via the induced changes in associated intermolecular
interactions. In this study, we demonstrated a reversible magnetic
switching associated with spin state conversion, along with a single-
crystal to single-crystal (SCSC) phase transition induced by
dehydration/rehydration. [Co(terpy)2](BF4)2·H2O (1·H2O; terpy
= 2,2′:6′,2′′-terpyridine) assembles in the solid state via π−π and
CH−π interactions involving adjacent terpyridine cores along the ab direction to form two-dimensional (2D) layered domains. 1·
H2O exhibits gradual and incomplete SCO, from fully HS to ca. 0.5 HS, and the field-induced single-molecule magnet (SMM)
behavior attributed to the presence of the anisotropic partial high-spin Co(II) species. 1·H2O undergoes a SCSC transformation
accompanied by a change from the tetragonal space group I41/a to P42/n via a dehydration process. Dehydrated 1 exhibits a reverse
thermal hysteresis behavior (T1/2↑ = 287 K; T1/2↓ = 270 K) in the gradual SCO region from fully HS to ca. 0.5 HS, followed by an
ordinary thermal hysteresis (T′1/2↑ = 195 K; T′1/2↓ = 155 K) to fully LS Co(II). A temperature-dependent single-crystal X-ray
structural analysis revealed that the reverse hysteresis can be attributed to an order/disorder structural phase transition of the BF4−
anions involving a symmetry breaking to yield the monoclinic space group P21/n and orbital (angular momentum) transition (LT).
Both the SCSC phase transition and magnetic behavior are switchable by dehydration/rehydration processes; thus 1 again adsorbs
water at room temperature to give both the original structure and its magnetic behavior.

■ INTRODUCTION
Numerous bistable ferromagnetic and ferroelectric compounds
cooperativity and influence whether abrupt or thermal
hysteresis occurs.11,12
have now been reported,1,2 with interest in such compounds The acquirement of the structural dimensionality is one of
arising from their potential application in a number of areas the appropriate approaches leading to exhibit the cooperative
that include molecular sensing and information storage. Spin SCO phenomena. Increasing dimensionality in the complex
transition compounds may exhibit multistabilities under two or structure normally leads to increased cooperativity, giving rise
more conditions. Spin crossover (SCO) is characterized by a to multistable phenomena in some cases. Thus, multidimen-
transformation between high-spin (HS) and low-spin (LS) sional compounds like flexible MOFs that are able to absorb
states induced by external stimuli such as temperature, guest molecules often display multistable states that depend on
pressure, or light.3−10 Since spin state transitions in metal the presence of the guests.13−17 While multistep SCO is most
complexes also affect the latter’s optical, magnetic, and
dielectric properties, the control of spin state can be employed Received: March 18, 2020
to switch the physical properties. The SCO phenomenon is
influenced not only by the primary molecular structure of the
complex but also by the lattice packing mode and the presence
of intermolecular interactions (such as, hydrogen bonding,
CH−π and π−π interactions), which act to enhance

© XXXX American Chemical Society https://dx.doi.org/10.1021/acs.inorgchem.0c00818


A Inorg. Chem. XXXX, XXX, XXX−XXX
Inorganic Chemistry pubs.acs.org/IC Article

Figure 1. Single-crystal structures (including disordered BF4− anions) of (a) 1·H2O and (b) 1. The light blue dashed lines represent hydrogen
bonds. (c) Packing view of 2D layered structure along the c axis of 1·H2O and 1. In the packing view, all hydrogen atoms, counteranions, and water
molecules are omitted for clarity. Color code: H, white; B, pink; C, gray; N, blue; O, red; F, yellow; Co, magenta.

likely observed for multidimensional compounds, it also occurs Although the SMM behavior is not usually observed in six-
for mononuclear complexes displaying enhanced cooperative- coordinated Co(II) SCO complexes because the LS Co(II)
ness through the presence of strong intermolecular interactions species lack the magnetic anisotropy, in the present study, the
with adjacent molecules.18−22 Unlike MOFs, mononuclear switch of SMM behavior has been achieved by the spin state
complexes exhibiting multifunctionality tend to be less conversion between HS and LS states induced by the reversible
responsive to the reversible adsorption/desorption of guest adsorption/desorption of the lattice water molecules. Mono-
species with keeping their crystallinity due to their generally nuclear Co(II) complexes are unusual in that they character-
more rigid structures. Few mononuclear complexes are known istically show high responsivity to a slight change in their
that are reversibly converted to multistable states with single- intermolecular interactions that can be attributed to their lower
crystal to single-crystal (SCSC) transition in response to the spin entropy (ΔSspin = R[ln(2S + 1)HS − ln(2S + 1)LS] = 5.8 J
uptake of guest species.23−30 K−1 mol−1),32 occasionally leading to them exhibiting unique
Herein, we report new examples of water molecule-induced SCO behavior such as a reverse spin transition.33−35 In
reversible switching of multistep magnetic behavior along with addition, a high orbital angular momentum (L) contribution in
a “cross loop” type hysteresis that involves both reverse and HS Co(II) complexes can lead to orbital quenching and
normal thermal hysteresis. This occurs for the mononuclear unique hysteretic behaviors such as reported by Juhász et al.36
Co(II) complex [Co(terpy)2](BF4)2·H2O (1·H2O, terpy = As opposed to SCO driven by entropy change, such transition
2,2′:6′,2′′-terpyridine), which also exhibits single-molecule of the orbital angular momentum (orbital transition, LT) is
magnet (SMM) behavior, as well as for its desolvated mainly caused by structural transition that includes the
[Co(terpy)2](BF4)2 (1) complex. Although these properties symmetry change of the coordination sphere. However, the
suggest the prospect of forming new switching materials that, examples of LT are quite limited because the orbital angular
for example, switch between SCO and SMM behaviors, there momentum of most first-row transition metal complexes is
appears to be only one reported example of this so far.31 usually small, and in most cases, the detailed mechanism of
B https://dx.doi.org/10.1021/acs.inorgchem.0c00818
Inorg. Chem. XXXX, XXX, XXX−XXX
Inorganic Chemistry pubs.acs.org/IC Article

such unique LT and hysteretic behaviors remains unclear. In K, respectively (Table S1). 1·H2O crystallized in the tetragonal
contrast, in the present study, we have succeeded in elucidating space group I41/a. Several polymorphs of 1·H2O were
the origin of the cross loop hysteresis attributed to an LT obtained; however, the use of an acetone/water mixture for
triggered by an order/disorder structural phase transition of crystallization led to isolation of a single morphology (see the
the BF4− anions by analyzing the crystal structures before and Experimental Section and Figure S1). In contrast to the
after the observed phase transition. present case, Kilner and Halcrow have reported that the use of

■ EXPERIMENTAL SECTION
Synthesis. All chemicals were purchased from commercial sources
nitrobenzene as a solvent during crystallization leads to
isolation of only the nonsolvated compound [Co(terpy)2]-
(BF4)2.37 The molecular structure of this product is shown in
and used without further purification. Figure 1a. Two tridentate meridional terpyridine ligands yield
[Co(terpy)2](BF4)2·H2O (1·H2O; terpy = 2,2′:6′,2′′-terpyri- an octahedral coordination environment, with one lattice water
dine). Co(BF4)2·6H2O (74 mg, 22 mmol) in acetone (10 mL) was molecule and two BF4− counterions being present. The
added to a solution of 2,2′:6′,2′′-terpyridine (100 mg, 43 mmol) in respective Co−N distances are Co−N1 = 2.124 Å, Co−N2
acetone (10 mL). The solution immediately turned brown. Water (3 = 1.989 Å, Co−N3 = 2.118 Å, and Co−N4 = 1.990 Å. The
mL) was added, and the solution was allowed to slowly evaporate to
bond lengths are consistent with those of typical HS Co(II)
give the product as dark brown crystals. Elemental analysis for
C30H24B2CoF8N6O: Calc. C, 50.25; H, 3.37; N, 11.72%. Found. C, compounds.7,31,36 Selected bond lengths for 1·H2O are
50.50; H, 3.57; N, 11.76%. The TGA curve showed 2.6% weight loss summarized in Table S2. The distortion parameters Σ for
at 400 K. This shows good agreement with the water molecule the [CoN6] core were calculated; Σ is the sum of |90−α| for
content expected for 1·H2O (H2O = 2.5%). IR (νmax, cm−1) 3645, the 12 cis-N−Co−N angles.38,39 Σ for 1·H2O is 108.48 (Table
3568, 3114, 1601, 1473, 1452, 1400, 1321, 1288, 1243, 1163, 1103, S3), indicating high distortion at low temperature compared to
1052, 1025, 767, 648, 520. other reported LS state bis-terpyridine Co(II) compounds.37,40
Dehydrated [Co(terpy)2](BF4)2 (1). For the single-crystal Each water molecule in the crystal packing structure forms
structural analysis, a single crystal of 1·H2O was dehydrated to give hydrogen bonds with two neighboring BF4− ions (Figure S2).
(dehydrated) 1 by heating at 333 K for 1 h in the SCXRD system. For
the other physical measurements, the crystalline sample of 1·H2O was
The respective intermolecular hydrogen bond distances are B−
dehydrated to give dehydrated 1 by heating at 400 K for 1 h in the F1A···O8 = 2.927 Å, B−F2A′···O8 = 2.584 Å, B−F3′···O8 =
measuring instruments (such as the SQUID or DSC system). 3.266 Å, and B−F4′···O8 = 3.062 Å. These hydrogen bonds
Rehydrated [Co(terpy)2](BF4)2·H2O (1·reh). The crystalline result in anion layers along the ab direction, with the lattice
sample of 1 was left to stand for 1 day in air at room temperature water molecules being located in the resulting cavities (Figure
(ca. 25 °C) to give the rehydrated product. Elemental analysis for S3a,c). The two terpyridine moieties in 1·H2O are assembled
C30H24B2CoF8N6O: Calc. C, 50.25; H, 3.37; N, 11.72%. Found. C, with π−π (N1−C5; N3−C13: centroid−centroid distance
50.42; H, 3.45; N, 11.61%. The TGA curve showed 2.7% weight loss 3.679 Å) and CH−π (H2, H10; N1−C5, N3−C13: H−
at 400 K. centroid distance 2.953 Å, 2.974 Å) interactions with their
Physical Measurements. Single-crystal X-ray data for 1 and 1·
H2O were recorded with an Oxford Gemini Ultra diffractometer adjacent cores oriented along the ab direction; these are
employing graphite monochromated Mo−Kα radiation generated layered in the ab plane with infinite π−π and CH−π stacking
from a sealed 261 tube (λ = 0.7107 Å). Data integration and interactions present. Each layer is separated by BF4− anions;
reduction were undertaken with CrysAlisPro. Using Olex2, the there are no direct interactions between each layer (Figure 1c).
structure was solved with the SHELXT structure solution program 1·H2O was dehydrated to give 1 by heating at 333 K for 1 h;
using Direct Methods and refined with the SHELXL refinement the hydration process was also followed by thermogravimetric
package using Least Squares minimization. Hydrogen atoms were analysis (TGA) (Figure S4a). On letting dehydrated 1 stand in
included in idealized positions and refined using a riding model. The the atmosphere, the water molecules were spontaneously
single-crystal X-ray diffraction measurements for 1 at 223 and 100 K
reabsorbed at room temperature to produce the original 1·
were carried out on the MX2 beamline at the Australian Synchrotron.
Diffraction data were collected using Si ⟨111⟩ monochromated H2O. Multiple repetition of the dehydration/rehydration
synchrotron X-ray radiation (λ = 0.71074) at 100(2) K with BlueIce process using TGA was also in accord with this result (Figure
software and were corrected for Lorentz and polarization effects using S4b).
the XDS software. The structure was solved with SHELXT, and the VT-SCXRD measurement on the dehydrated product 1 was
full-matrix least-squares refinements were carried out using SHELXL- conducted at 333 K in order to compare its structure with that
2014 via the Olex2 interface. All non-hydrogen atoms with of 1·H2O. In the former, the space group had changed to the
occupancies over 0.5 were located from the electron density maps tetragonal space group P42/n at 333 K; the crystal structure is
and refined anisotropically. Hydrogen atoms bound to carbon, shown in Figure 1b. The BF4− anions in 1 are heavily
nitrogen, and oxygen atoms were added in the ideal positions and
refined using a riding model. Temperature-dependent direct-current
disordered due to the absence of structural restriction by
(dc) magnetic susceptibilities for crystalline samples of 1·H2O and 1 hydrogen bonding compared with those present in 1·H2O
were measured on a Superconducting Quantum Interference Device (Figure S2). The Co−N distances are Co−N1 = 2.127 Å and
(SQUID) magnetometer at a field strength of 5000 Oe with a sweep Co−N2 = 2.015 Å, with these bond lengths being consistent
mode of 2 K min−1 in the temperature range 5 to 400 K. Elemental with those for typical HS Co(II) compounds. The Σ value for
analyses (C, H, N) were carried out at the Instrumental Analysis 1 is 118.4, indicating higher distortion than that for 1·H2O. 1 is
Centre of Kumamoto University. Differential scanning calorimetry also assembled in the crystal lattice via intermolecular π−π
(DSC) thermal analysis for 1 was carried out at 10 K min−1 on a (N1−C5; N1′−C5′: centroid−centroid distance 3.736 Å) and
SHIMADZU DSC50.


CH−π (H4; N1−C5: H−centroid distance 3.003 Å)
interactions as occur in 1·H2O (Figure S3b,d). Reflecting the
RESULTS AND DISCUSSION strong lattice interactions present, the crystal packing structure
Structural Characterization. The crystal structures of 1· maintained the crystal void observed in the unit cell of 1·H2O,
H2O and 1 were solved by variable-temperature single-crystal with the dehydrated derivative giving an increased value Δ =
X-ray diffraction (VT-SCXRD) measurements at 123 and 333 19.65 Å3 per Co(II) cation (Figure S5). The absorption/
C https://dx.doi.org/10.1021/acs.inorgchem.0c00818
Inorg. Chem. XXXX, XXX, XXX−XXX
Inorganic Chemistry pubs.acs.org/IC Article

desorption process observed for the water molecule in 5000 Oe (Figure S8). The in-phase (χm′) and out-of-phase
crystalline 1·H2O is also synchronized with the reversible (χm″) components exhibited no frequency dependence under
single-crystal to single-crystal (SCSC) phase transition. The zero applied dc field. The absence of slow magnetic relaxation
powder X-ray diffraction (PXRD) patterns for 1·H2O and under zero applied field is in accord with the occurrence of
dehydrated 1 as well as for the rehydrated complex from 1 (1· quantum tunneling of the magnetization (QTM) as occurs
reh) are shown in Figure S6. The PXRD patterns for 1·H2O with many other mononuclear Co(II) SMMs. The frequency
and dehydrated 1 are slightly different; reversible changes in dependence of the χm′ and χm″ components becomes
the RXRD patterns between 1 and 1·reh were also observed. significantly more intense with the application of an external
Magnetic Behavior of 1·H2O. The temperature depend- dc field, revealing slow magnetic relaxation behavior as
ence of magnetic susceptibility for crystalline samples of 1·H2O expected for SMMs. As the applied field is increased, one
was measured employing a Superconducting Quantum relaxation pathway begins to appear at above ∼1500 Hz and
Interference Device (SQUID). The corresponding χmT versus increases in intensity until 1000 Oe, after which it begins to
T curves, where χm is the molar magnetic susceptibility and T decrease. A second relaxation pathway appears between 1 and
is the temperature, are shown in Figure 2. At 350 K, the χmT 10 Hz under an applied dc field of 1500 Oe; it appears initially
as a shoulder and then reaches a full peak as the applied field
increases. A peak in χm″ was evident under the applied external
fields, with a maximum shift to low frequencies being observed
for the 5000 Oe (dc) field, which we consider as the optimal
field for observing slow magnetic relaxation. Under applied
fields of 500 and 5000 Oe (where only one relaxation pathway
is clearly dominant), the χm″ versus frequency plots reveal clear
temperature-dependent peaks down to 1.9 K (Figures 3, S8,
S9, and S11). The occurrence of two such relaxation pathways
in other mononuclear Co(II) SMMs has frequently been
observed42−45 and attributed to the presence of the spin−spin
interactions caused by intermolecular short contacts, such as
hydrogen bonds, π−π stacking, and CH−π interactions−
resulting in much slower spin relaxations than occur in pure
mononuclear entities. The origin of the second main group of
such a spin relaxation process has been investigated by using
magnetic dilution of a seven-coordinate mononuclear Co(II)
Figure 2. χmT versus T plots for 1·H2O (black) and 1 (red, blue) complex, [Co(DAPBH)(NO3)(H2O)](NO3) (DAPBH = 2,6-
collected under an applied field of 5000 Oe. diacetylpyridinebis(2′-pyridylhydrazone)) which shows the
field-induced SMM behavior, with the diamagnetic Zn(II)
value of 2.11 cm3 K mol−1 is higher than the spin-only χmT analogue by Habib and co-workers.46 The observed second
value for one high-spin Co(II) ion (S = 3/2, 1.875 cm3 K relaxation step decreases with the dilution. Such a magnetic
mol−1), indicating an orbital angular momentum contribution. dilution study using a diamagnetic ion reveals the influence of
On lowering the temperature, the χmT value for 1·H2O the intermolecular interactions on the relaxation processes in
gradually decreased from 2.11 at 350 K to 1.14 cm3 K mol−1 at mononuclear SMMs. Ceglarska and co-workers also have
5 K, which represents incomplete SCO behavior for Co(II) in reported very recently that the creation of the two frequency
an octahedral environment and reflects the highly distorted domains observed in a six-coordinated Co(II) SMM
structure present at low temperature. This SCO behavior [CoBr2(pyridine)2] is of intermolecular origin by using a
corresponds to the presence of 1/2 HS (S = 3/2, 1.875 cm3 K similar systematic study of magnetic dilution.47 These studies
mol−1) and 1/2 LS (S = 1/2, 0.375 cm3 K mol−1) species, i.e., a clearly indicate that the occurrence of multiple relaxation
[HS-LS] phase. On heating the sample from 5 to 350 K, the pathways is attributed to the presence of intermolecular
same SCO behavior was observed. interactions.
Bis-terpyridine Co(II) complexes having highly distorted In order to elucidate the relative contributions of the two
octahedral high-spin Co(II) centers sometimes exhibit slow relaxation pathways, the temperature-dependent relaxation
magnetic relaxation behavior as is expected for SMMs due to times for applied fields of 500 and 5000 Oe were determined
their large magnetic anisotropy.31,41 For 1·H2O, variable-field by fitting the out-of-phase susceptibility using the generalized
magnetization plots show saturation behavior at 1.73 μβ under Debye model (Figures S10 and S12).48,49 For the Cole−Cole
an applied field of 5 T at 2 K (Figure S7). The expected analysis, the α values are considerably smaller, falling in the
magnetization value for 0.5 HS and 0.5 LS of a spin-only range of 0.001−0.06 for 5000 Oe and 0.22−0.24 for 5000 Oe,
Co(II) ion is 2 μβ, with derivation from this value arising from suggesting the presence of a single relaxation process (Table
the presence of the magnetic anisotropy of the HS species. S4). To extract the relaxation time and spin reversal barrier,
That saturation behavior in 1·H2O is observed at a lower value the maximums of the out-of-phase χm″ values from the
than 2 μβ indicates the presence of a highly anisotropic ground temperature-dependent data were used to construct the
state. Their large magnetic anisotropy leading to slow magnetic Arrhenius plots given in Figure S13. In these ln(τ) vs T−1
relaxation were demonstrated by also variable-field alternating plots, linear data corresponding to the Arrhenius law τ = τ0
current (ac) susceptibility measurements. In order to exp(Ueff/kBT) were obtained, with an effective energy barrier
investigate possible slow magnetic relaxation behavior in 1· of Ueff = 4.2 K (τ0 = 1.1 × 10−5 s) and 3.6 K (τ0 = 1.1 × 10−2 s)
H2O, variable-field ac susceptibility measurements for 1·H2O for the fast and slow relaxation processes, respectively. These
were performed at 1.9 K under applied fields of between 0 and values are similar to those obtained for other reported Co(II)
D https://dx.doi.org/10.1021/acs.inorgchem.0c00818
Inorg. Chem. XXXX, XXX, XXX−XXX
Inorganic Chemistry pubs.acs.org/IC Article

the precise values for other relaxation mechanisms for both


relaxation pathways.
Magnetic Behavior of 1. The dehydrated compound 1
was obtained by annealing polycrystalline samples of 1·H2O at
400 K for 1 h in the SQUID system. On lowering the
temperature, the χmT value of 1 gradually decreased from 2.22
at 400 K to 1.58 cm3 K mol−1 at 274 K, indicating thermal
SCO behavior, and it then slightly increased to 1.66 cm3 K
mol−1 at 264 K (T′1/2↓ = 270 K). This increase may be
rationalized in terms of a reverse spin transition or an angular
momentum quenching as reported for several Co(II) and
Fe(II) complexes.33−35,54−56 Upon further cooling, χmT
abruptly decreased from 0.89 at 161 K to 0.47 cm3 K mol−1
at 142 K (T1/2↓ = 155 K) following the initial linear decrease of
the χmT value. The χmT value for 1, 0.40 cm3 K mol−1 at 5 K, is
consistent with the presence of LS Co(II) (S = 1/2) in an
octahedral environment. Upon heating, the χmT value
gradually increased from 0.44 at 146 K to 0.69 cm3 K mol−1
at 192 K and then abruptly increased to 1.15 cm3 K mol−1 at
203 K (T1/2↑ = 195 K). On further heating, the χmT value
decreased slightly from 1.83 at 283 K to 1.81 cm3 K mol−1 at
297 K (T′1/2↑ = 287 K). The values of T1/2↑ and T1/2↓ and
T′1/2↑ and T′1/2↓ for the heating and cooling cycles correspond
to thermal hysteresis loops with ΔT = 40 K and ΔT′ = 17 K,
respectively. Although such a “cross loop” hysteresis, including
the reverse and ordinary thermal hysteresis loops, has been
reported previously by Agusti et al. for [Co(4-terpyridone)2]-
(CF3SO3)2·H2O, the mechanism involved still remains
unclear.57 As a scan rate is increasing, both reverse and
normal thermal hystereses become wider. The observed
thermal hysteresis loops at a scan rate of 5 K min−1 were
ΔT = 51 K and ΔT′ = 34 K, respectively (Figure S14, Table
S5). Although this is usually reflected in the thermalization
process of the samples, it seems that the observed reverse
thermal hysteresis is more sensitive to the scan rate compared
Figure 3. Variable-frequency out-of-phase (χm″) components of the with the normal thermal hysteresis at low temperature. In
ac susceptibilities for a crystalline sample of 1·H2O collected under an order to confirm the presence of slow magnetic relaxation,
applied dc field of (a) 500 Oe and (b) 5000 Oe. The solid lines variable-field ac susceptibility measurements for 1 were also
represent the best fits of the experimental results for the generalized performed at 1.9 K under applied field strengths between 0 and
Debye model. 3000 Oe (Figure S15). The out-of-phase (χm″) components
exhibited no frequency dependence even under an applied dc
field of 3000 Oe. The observed magnetic behavior indicates no
complexes.50,51 The relaxation mechanisms for a mononuclear presence of slow magnetic relaxation for 1, reflecting the
single-molecule magnet system are usually very complex and presence of LS Co(II) species which lack the magnetic
involve multiple relaxation pathways. Although up to four anisotropy exhibited by HS Co(II) species. Dehydrated 1
possible relaxation mechanisms (QTM, direct, Raman, and reabsorbed water at room temperature to again yield 1·H2O
Orbach processes) may usually occur, the QTM should, in (which exhibited the same SCO and SMM behaviors as the
principle, be suppressed in the case of measurements under an original hydrated product).
applied dc field. When a relaxation process other than the Differential Scanning Calorimetry and Crystallo-
Orbach process may play a role, or that there is more than one graphic Study of 1. Differential scanning calorimetry
thermally activated relaxation process occurring, then the ln(τ) (DSC) measurement and further crystallographic studies at
vs T−1 plots will show a departure from Arrhenius behavior. 223 and 100 K were performed in order to further probe the
For the type of Kramers ion incorporating Co(II), the main two hysteric behaviors observed for 1. Following the initial
contribution of such a slow magnetic relaxation is a thermal treatment of polycrystalline samples of 1·H2O at 400
combination of a direct mechanism and the Raman K for 1 h to produce 1, DSC curves were obtained for the
process.52,53 In the present study, a clear departure from product in the range 120 to 400 K (Figure S16, Table S6).
Arrhenius behavior was not observed in the measured Two exothermic peaks were observed at ca. 265 and 170 K,
temperature region, indicating that a Raman process is not a respectively, during the cooling process; these correspond to
main contributor to the slow magnetic relaxation observed for T′1/2↓ and T1/2↓ for the spin transition. Two endothermic peaks
this system. However, in the present study we need to point were observed at around 200 and 275 K thus confirming the
out that the obtained τ values are only rough estimations. reversible nature of the phase transition in 1.
Further measurements spanning a higher frequency range (ν > VT-SCXRD measurements for 1 at 223 and 100 K were
1500 Hz) or a lower temperature region are required to obtain performed at the Australian Synchrotron (Figure 4). During
E https://dx.doi.org/10.1021/acs.inorgchem.0c00818
Inorg. Chem. XXXX, XXX, XXX−XXX
Inorganic Chemistry pubs.acs.org/IC Article

Figure 4. Asymmetric unit and comparison of the [CoN6] coordination bond lengths for 1 at (a) 333 K and (b) 223 K. The BF4− anion is
disordered at 333 K.

Figure 5. Schematic representation of the order/disorder structural phase transitions in 1. The disordered BF4− anions at 333 K are ordered at 223
K.

the cooling process from 333 K, 1 gave rise to the structural reflections at 223 K yielded a doubling of the lattice parameter
phase transition with the exothermic peak at ca. 265 K. At 223 along the stacking direction, and consequently, the b
K, it is in the intermediate state as shown by the χmT versus T parameter, 38.398(8) Å at 233 K, is almost double the value
curves (and associated with the reverse χmT changes reflecting obtained at 333 K, namely 19.507(2) Å; that is, the volume of
the structural phase transition). Symmetry breaking was the unit cell changes from 1517.5(3) to 2981.2(10) Å3 (Table
observed along with the space group change to the monoclinic S7). The phase transition accompanying the change in the cell
space group P21/n; the crystal structure and packing structure parameters in this intermediate phase is in accord with the
are shown in Figures S17 and S18. The observed satellite occurrence of an order/disorder phase transition, induced by
F https://dx.doi.org/10.1021/acs.inorgchem.0c00818
Inorg. Chem. XXXX, XXX, XXX−XXX
Inorganic Chemistry pubs.acs.org/IC Article

Figure 6. Comparison of the layered structure at 100 K of (a) 1 and (b) polymorph [Co(terpy)2](BF4)2.37

two independent BF4− counteranions in the interlayer space, actions, and 1 exhibits an abrupt spin transition with wide
leading to crystallographically distinct anion layers (Figure 5). thermal hysteresis. The average Co−N distance and the Σ
The resulting structural ordering is characteristic of the value at 223 K are 2.058 Å and 104, respectively. These are in
intermediate state of mononuclear complexes and 2D MOFs the middle range between HS and LS, reflecting a spin state
exhibiting multistep SCO behaviors.14,19,57−62 corresponding to the presence of 1/2 HS and 1/2 LS species.
The changes in the cell parameters generated for the Although the temperature range from 400 to 170 K for the
intermediate phase were observed to remain (at ∼170 K) even cooling process corresponds to the gradual SCO region, the
after the phase transition has occurred during the cooling structural phase transition occurred at 265 K resulting in the
process. The cell parameters at 100 K are doubled relative to observed change of the χmT value (Figure S20). This χmT
those present at 333 K, and two different lattice layers are change seems best attributed to angular momentum quenching
present, even for the complete LS state (Figure S18a). This rather than to a reverse spin transition. It is well-known that
seems to imply that the formation of the two crystallo- the contribution of the orbital angular momentum is quite
graphically independent alternating anion layers in the large in HS Co(II) complexes and that it is reflected in the
intermediate state led to abrupt spin transition to the LS structural symmetry. As reported by Titiš et al., the magnetic
species. The thermal variation of the cell parameters for 1 anisotropy in hexacoordinated Co(II) complexes correlates
derived from quick scans resulted in a small change in the in- with the structural distortion of the coordination polyhedron.63
plane direction, while the length of the out-of-plane direction The reduction in the structural symmetry in HS Co(II)
increased with the increasing temperature (Figure S19). The complexes leads to an increase in the contribution of the
rigidity of the layer structure in the in-plane is aided by the orbital angular momentum involving an increase of the χmT
presence of π−π and CH−π interactions between [Co- value.36 In the present case, the symmetry reduction involving
(terpy)2]2+ units, while the interlayer structure is influenced the change of asymmetric unit would be expected to lead to an
to some degree by the thermal vibrations of the BF4− anions. increase in the contribution of the orbital angular momentum
The cell parameters observed at high temperature were and in the χmT value. The entropy changes (|ΔS|) at 265 and
restored during the heating process, once again confirming 275 K obtained by DSC measurements are 3.356 J K−1 mol−1
the reversibility of the structural phase transition. and 3.467 J K−1 mol−1, respectively. These values correspond
The selected bond lengths and the distortion parameters Σ to the reported entropy change (ΔS = 3.622 J K−1 mol−1) for
for 1 at 223 and 100 K are summarized in Tables S8 and S9, an order/disorder type transition for a BF4− anion.64 These
respectively. At 100 K, although the average Co−N distance of reversible phase transitions cause the changes that occur in the
2.040 Å and the Σ value of 95.6 are higher than those of other intermolecular interactions between the [Co(terpy)2]2+ cations
LS Co(II) complexes, the ΔΣ value of 22.8 for 1 between HS and BF4− anions, resulting in the observed structural symmetry
and LS is a typical value observed for SCO Co(II) complexes. changes and unique cross loop hysteresis. In fact, several
As reported in our previous study, the structural distortion of examples of spin transitions, including reverse spin transitions,
the [CoN6] core is a dominant factor leading to abrupt spin associated with structural transitions in metal complexes
transition in hexacoordinated Co(II) complexes.40 Reflecting incorporating BF4− anions have been reported previously.35,65
the highly distorted structure and the in-plane high The formation of hydrogen bonding between water molecules
cooperativity attributed to the strong intermolecular inter- and BF4− anions induced by the hydration process clearly
G https://dx.doi.org/10.1021/acs.inorgchem.0c00818
Inorg. Chem. XXXX, XXX, XXX−XXX
Inorganic Chemistry pubs.acs.org/IC Article

hinders such order/disorder transition for these anions, leading Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44
to the gradual SCO behavior observed, including the small 1223 336033.
stepwise changes of the χmT value found for 1·H2O. Although
it seems that the similar crossed hysteretic behavior reported
by Agusti et al. for [Co(4-terpyridone)2](CF3SO3)2·H2O57 can
■ AUTHOR INFORMATION
Corresponding Author
also be attributed to LT associated with the structural Shinya Hayami − Department of Chemistry, Graduate School of
transitions of the CF3SO3− anions rather than to a reverse Science and Technology and Institute of Pulsed Power Science
spin transition, a more detailed structural study such as one (IPPS), Kumamoto University, Kumamoto 860-8555, Japan;
employing VTSCXRD would be required to determine the orcid.org/0000-0001-8392-2382; Email: hayami@
origin. In contrast to the observed cross loop hysteresis kumamoto-u.ac.jp
occurring for 1, the polymorph [Co(terpy)2](BF4)2 reported
by Kilner and Halcrow37 exhibits gradual SCO without Authors
structural transitions. This appears to be a reflection of the Fumiya Kobayashi − Department of Chemistry, Graduate
difference arising from the presence of the stacked layered School of Science and Technology, Kumamoto University,
structures rather than being the result of the intralayer Kumamoto 860-8555, Japan; orcid.org/0000-0002-6923-
interactions. Although the cationic layers in 1 are stacked in 6332
phase, those of the above polymorph are stacked out of phase Yuki Komatsumaru − Department of Chemistry, Graduate
and associated with a different interlayer distance (Figure 6). School of Science and Technology, Kumamoto University,
The different interlayer structural arrangements, in turn, lead Kumamoto 860-8555, Japan
to different intermolecular interactions between the [Co- Ryohei Akiyoshi − Department of Chemistry, Graduate School
(terpy)2]2+ cations and BF4− anions, which are the important of Science and Technology, Kumamoto University, Kumamoto
key to lead the cross loop hysteretic behaviors. 860-8555, Japan


Masaaki Nakamura − Department of Chemistry, Graduate
CONCLUSION School of Science and Technology, Kumamoto University,
Kumamoto 860-8555, Japan
In summary, we have presented an example of water molecule- Yingjie Zhang − Australian Nuclear Science and Technology
induced reversible magnetic switching in a mononuclear Organization, Kirrawee DC, New South Wales 2232, Australia
Co(II) complex. Thus, 1·H2O undergoes a reversible SCSC Leonard F. Lindoy − School of Chemistry, The University of
transformation with desorption of water molecules, leading to Sydney, Sydney, New South Wales 2006, Australia
a change in its magnetic behavior. Monohydrate 1·H2O
exhibits gradual SCO and field-induced SMM behavior, Complete contact information is available at:
whereas dehydrated 1 exhibits multistep cross loop hysteretic https://pubs.acs.org/10.1021/acs.inorgchem.0c00818
behavior. This cross loop hysteresis is attributed to the abrupt
spin crossover (SCO) in the low temperature region and Author Contributions
orbital transition (LT) in the high temperature region. In this The manuscript was written through contributions of all
flexible guest-responsive system, the presence of intermolecular authors. All authors have given approval to the final version of
interactions, such as π−π and CH−π interactions, creates the manuscript.
pseudodimensionality. In addition, investigation of the Notes
The authors declare no competing financial interest.


influence of the angular momentum on the order/disorder
phase transition involving the BF4− counterions leads to a
deeper understanding of similar SCO systems. In addition, it ACKNOWLEDGMENTS
was unexpected that the observed unusual multifunctional This work was supported by a KAKENHI Grant-in-Aid for
properties were exhibited by the “classical” bis-terpyridine Scientific Research (A) JP17H01200. F.K. is grateful to JSPS
Co(II) complex. The high responsivity of the Co(II) ion, Research Fellowships for Young Scientists (No. JP19J11651).
attributed to its spin entropy, clearly warrants more attention This research was undertaken in part using the MX2 beamline
in future studies of the present general type. In addition, the at the Australian Synchrotron, part of ANSTO, and made use
results presented provide the basis for new design strategies for of the Australian Cancer Research Foundation (ACRF)
generating guest-responsive, multifunctional switching materi- detector.
als as well as other unique SCO and LT systems.

■ ASSOCIATED CONTENT
■ REFERENCES
(1) Shi, P. P.; Tang, Y. Y.; Li, P. F.; Liao, W. Q.; Wang, Z. X.; Ye, Q.;
*
sı Supporting Information
Xiong, R. G. Symmetry breaking in molecular ferroelectrics. Chem.
Soc. Rev. 2016, 45, 3811−3827.
The Supporting Information is available free of charge at (2) Fiebig, M. The evolution of multiferroics. Nat. Rev. Mater. 2016,
https://pubs.acs.org/doi/10.1021/acs.inorgchem.0c00818. 1, 16046.
Crystallographic data, PXRD, TGA, magnetic properties, (3) Hayami, S.; Gu, Z. Z.; Shiro, M.; Einaga, Y.; Fujishima, A.; Sato,
O. First Observation of Light-Induced Excited Spin State Trapping for
and DSC (PDF)
an Iron(III) Complex. J. Am. Chem. Soc. 2000, 122, 7126−7127.
Accession Codes (4) Gütlich, P.; Garcia, Y.; Goodwin, H. A. Spin crossover
phenomena in Fe(II) complexes. Chem. Soc. Rev. 2000, 29, 419−427.
CCDC 1903790−1903791 and 1980590−1980591 contain (5) Letard, J. F. Photomagnetism of iron(II) spin crossover
the supplementary crystallographic data for this paper. These complexesthe T(LIESST) approach. J. Mater. Chem. 2006, 16,
data can be obtained free of charge via www.ccdc.cam.ac.uk/ 2550−2559.
data_request/cif, or by emailing data_request@ccdc.cam.ac. (6) Li, B.; Wei, R.-J.; Tao, J.; Huang, R.-B.; Zheng, L.-S. Pressure
uk, or by contacting The Cambridge Crystallographic Data Effects on a Spin-Crossover Monomeric Compound [Fe(Pmea)-

H https://dx.doi.org/10.1021/acs.inorgchem.0c00818
Inorg. Chem. XXXX, XXX, XXX−XXX
Inorganic Chemistry pubs.acs.org/IC Article

(SCN)2] (Pmea = Bis[(2-Pyridyl)Methyl]-2-(2-Pyridyl)Ethylamine). (25) Huang, W.; Shen, F.; Zhang, M.; Wu, D.; Pan, F.; Sato, O.
Inorg. Chem. 2010, 49, 745−751. Room-temperature switching of magnetic hysteresis by reversible
(7) Hayami, S.; Komatsu, Y.; Shimizu, T.; Kamihata, H.; Lee, Y. H. single-crystal-to-single-crystal solvent exchange in imidazole-inspired
Spin-crossover in cobalt(II) compounds containing terpyridine and its Fe(II) complexes. Dalton. Trans. 2016, 45, 14911−14918.
derivatives. Coord. Chem. Rev. 2011, 255, 1981−1990. (26) Miller, R. G.; Brooker, S. Reversible quantitative guest sensing
(8) Brooker, S. Spin Crossover with Thermal Hysteresis: via spin crossover of an iron(II) triazole. Chem. Sci. 2016, 7, 2501−
Practicalities and Lessons Learnt. Chem. Soc. Rev. 2015, 44, 2880− 2505.
2892. (27) Costa, J. S.; Rodríguez-Jiménez, S.; Craig, G. A.; Barth, B.;
(9) Harding, D. J.; Harding, P.; Phonsri, W. Spin crossover in Beavers, C. M.; Teat, S. J.; Aromí, G. Three-Way Crystal-to-Crystal
iron(III) complexes. Coord. Chem. Rev. 2016, 313, 38−61. Reversible Transformation and Controlled Spin Switching by a
(10) Hogue, R. W.; Singh, S.; Brooker, S. Chem. Soc. Rev. 2018, 47, Nonporous Molecular Material. J. Am. Chem. Soc. 2014, 136, 3869−
7303−7338. 3874.
(11) Hayami, S.; Gu, Z.-Z.; Yoshiki, H.; Fujishima, A.; Sato, O. (28) Jornet-Mollá, V.; Duan, Y.; Giménez-Saiz, C.; Tang, Y.-Y.; Li,
Iron(III) Spin-Crossover Compounds with a Wide Apparent Thermal P.-F.; Romero, F. M.; Xiong, R.-G. A Ferroelectric Iron(II) Spin
Hysteresis around Room Temperature. J. Am. Chem. Soc. 2001, 123, Crossover Material. Angew. Chem., Int. Ed. 2017, 56, 14052−14056.
11644−11650. (29) Shao, D.; Shi, L.; Shen, F.-X.; Wei, X.-Q.; Sato, O.; Wang, X.-Y.
(12) Tsukiashi, A.; Nakaya, M.; Kobayashi, F.; Ohtani, R.; Reversible On-Off Switching of the Hysteretic Spin Crossover in a
Nakamura, M.; Harrowfield, J.; Kim, Y.; Hayami, S. Intermolecular Cobalt(II) Complex via Crystal to Crystal Transformation. Inorg.
Interaction Tuning of Spin-Crossover Iron(III) Complexes with Chem. 2019, 58, 11589−11598.
Aromatic Counteranions. Inorg. Chem. 2018, 57, 2834−2842. (30) Nakaya, M.; Kosaka, W.; Miyaska, H.; Komatsumaru, Y.;
(13) Schneemann, A.; Bon, V.; Schwedler, I.; Senkovska, I.; Kaskel, Kawaguchi, S.; Sugimoto, K.; Zhang, Y.; Nakamura, M.; Hayami, S.
S.; Fischer, R. A. Flexible metal-organic frameworks. Chem. Soc. Rev. CO2-Induced Spin-State Switching at Room Temperature in a
2014, 43, 6062−6096. Monomeric Cobalt(II) Complex with the Porous Nature. Angew.
(14) Murphy, M. J.; Zenere, K. A.; Ragon, F.; Southon, P. D.; Chem., Int. Ed. 2020, 59, 10658−10665.
Kepert, C. J.; Neville, S. M. Guest Programmable Multistep Spin (31) Shao, D.; Shi, L.; Yin, L.; Wang, B.-L.; Wang, Z.-X.; Zhang, Y.-
Crossover in a Porous 2-D Hofmann-Type Material. J. Am. Chem. Soc. Q.; Wang, X.-Y. Reversible on-off switching of both spin crossover
2017, 139, 1330−1335. and single-molecule magnet behaviours via a crystal-to-crystal
(15) Ni, Z. P.; Liu, J. L.; Hoque, M. N.; Liu, W.; Li, J. Y.; Chen, Y. transformation. Chem. Sci. 2018, 9, 7986−7991.
C.; Tong, M. L. Recent advances in guest effects on spin-crossover (32) Gütlichm, P.; Goodwin, H. A. Spin crossoveran overall
behavior in Hofmann-type metal-organic frameworks. Coord. Chem. perspective. Spin crossover in transition metal compounds I; Topics in
Rev. 2017, 335, 28−43. Current Chemistry, 2004; Vol. 233, pp 1−47,
(16) Ohtani, R.; Hayami, S. Guest-Dependent Spin-Transition (33) Hayami, S.; Shigeyoshi, Y.; Akita, M.; Inoue, K.; Kato, K.;
Behavior of Porous Coordination Polymers. Chem. - Eur. J. 2017, 23, Osaka, K.; Tanaka, M.; Kawajiri, R.; Mitani, T.; Maeda, Y. Reverse
2236−2248. Spin Transition Triggered by a Structural Phase Transition. Angew.
(17) Meng, Y.; Dong, Y. J.; Yan, Z.; Chen, Y. C.; Song, X. W.; Li, Q. Chem., Int. Ed. 2005, 44, 4899−4903.
W.; Zhang, C. L.; Ni, Z. P.; Tong, M. L. New Porous Three- (34) Hayami, S.; Murata, K.; Urakami, D.; Kojima, Y.; Akita, M.;
Dimensional Iron(II) Coordination Polymer with Solvent-Induced Inoue, K. Dynamic structural conversion in a spin-crossover cobalt(II)
Reversible Spin-Crossover Behavior. Cryst. Growth Des. 2018, 18, compound with long alkyl chains. Chem. Commun. 2008, 6510−6512.
5214−5219. (35) Hayami, S.; Nakaya, M.; Ohmagari, H.; Alao, A. S.; Nakamura,
(18) Sato, T.; Nishi, K.; Iijima, S.; Kojima, M.; Matsumoto, N. One- M.; Ohtani, R.; Yamaguchi, R.; Kuroda-Sowa, T.; Clegg, J. K. Spin-
Step and Two-Step Spin-Crossover Iron(II) Complexes of ((2- crossover behaviors in solvated cobalt(II) compounds. Dalton Trans.
Methylimidazol-4-yl)methylidene)histamine. Inorg. Chem. 2009, 48, 2015, 44, 9345−9348.
7211−7229. (36) Juhász, G.; Matsuda, R.; Kanegawa, S.; Inoue, K.; Sato, O.;
(19) Lennartson, A.; Bond, A. D.; Piligkos, S.; Mckenzie, C. J. Four- Yoshizawa, K. Bistability of Magnetization without Spin-Transition in
Site Cooperative Spin Crossover in a Mononuclear FeII Complex. a High-Spin Cobalt(II) Complex due to Angular Momentum
Angew. Chem., Int. Ed. 2012, 51, 11049−11052. Quenching. J. Am. Chem. Soc. 2009, 131, 4560−4561.
(20) Li, Z.-Y.; Dai, J.-W.; Gagnon, K. J.; Cai, H.-L.; Yamamoto, T.; (37) Kilner, C. A.; Halcrow, M. A. An unusual discontinuity in the
Einaga, Y.; Zhao, H.-H.; Kanegawa, S.; Sato, O.; Dunbar, K. R.; Xiong, thermal spin transition in [Co(terpy)2][BF4]2. Dalton Trans. 2010,
R.-G. A neutral Fe(III) compound exhibiting a two-step spin 39, 9008−9012.
transition and dielectric anomalies. Dalton Trans. 2013, 42, 14685− (38) Halcrow, M. A. Structure:function relationships in molecular
14688. spin-crossover complexes. Chem. Soc. Rev. 2011, 40, 4119−4142.
(21) Luan, J.; Zhou, J.; Liu, Z.; Zhu, B.; Wang, H.; Bao, X.; Liu, W.; (39) Kershaw Cook, L. J.; Mohammed, R.; Sherborne, G.; Roberts,
Tong, M.-L.; Peng, G.; Peng, H.; Salmon, L.; Bousseksou, A. T. D.; Alvarez, S.; Halcrow, M. A. Spin state behavior of iron(II)/
Polymorphism-Dependent Spin-Crossover: Hysteretic Two-Step Spin dipyrazolylpyridine complexes. New insights from crystallographic
Transition with an Ordered [HS-HS-LS] Intermediate Phase. Inorg. and solution measurements. Coord. Chem. Rev. 2015, 289-290, 2−12.
Chem. 2015, 54, 5145−5147. (40) Nakaya, M.; Ohtani, R.; Shin, J. W.; Nakamura, M.; Lindoy, L.
(22) Li, Z.-Y.; Ohtsu, H.; Kojima, T.; Dai, J.-W.; Yoshida, T.; F.; Hayami, S. Abrupt spin transition in a modified-terpyridine
Breedlove, B. K.; Zhang, W.-X.; Iguchi, H.; Sato, O.; Kawano, M.; cobalt(II) complex with a highly-distorted [CoN6] core. Dalton
Yamashita, M. Direct Observation of Ordered High-Spin-Low-Spin Trans. 2018, 47, 13809−13814.
Intermediate States of an Iron(III) Three-Step Spin-Crossover (41) Kobayashi, F.; Ohtani, R.; Nakamura, M.; Lindoy, L. F.;
Complex. Angew. Chem. 2016, 128, 5270−5275. Hayami, S. Slow Magnetic Relaxation Triggered by a Structural Phase
(23) Li, B.; Wei, R.-J.; Tao, J.; Huang, R.-B.; Zheng, L.-S.; Zheng, Z. Transition in Long-Chain Alkylated Cobalt(II) Single-Ion Magnets.
Solvent-Induced Transformation of Single Crystals of a Spin- Inorg. Chem. 2019, 58, 7409−7415.
Crossover (SCO) Compound to Single Crystals with Two Distinct (42) Habib, F.; Luca, O. R.; Vieru, V.; Shiddiq, M.; Korobkov, I.;
SCO Centers. J. Am. Chem. Soc. 2010, 132, 1558−1566. Gorelsky, S. I.; Takase, M. K.; Chibotaru, L. F.; Hill, S.; Crabtree, R.
(24) Wei, R.-J.; Tao, J.; Huang, R.-B.; Zheng, L.-S. Reversible and H.; Murugesu, M. Influence of the Ligand Field on Slow
Irreversible Vapor-Induced Guest Molecule Exchange in Spin- Magnetization Relaxation versus Spin Crossover in Mononuclear
Crossover Compounds. Inorg. Chem. 2011, 50, 8553−8564. Cobalt Complexes. Angew. Chem., Int. Ed. 2013, 52, 11290−11293.

I https://dx.doi.org/10.1021/acs.inorgchem.0c00818
Inorg. Chem. XXXX, XXX, XXX−XXX
Inorganic Chemistry pubs.acs.org/IC Article

(43) Rajnák, C.; Varga, F.; Titiš, J.; Moncol,̌ J.; Boča, R. Field- (60) Bréfuel, N.; Collet, E.; Watanabe, H.; Kojima, M.; Matsumoto,
Supported Single-Molecule Magnets of Type [Co(bzimpy)X2]. Eur. J. N.; Toupet, L.; Tanaka, K.; Tuchagues, J. P. Nanoscale Self-Hosting
Inorg. Chem. 2017, 2017, 1915−1922. of Molecular Spin-States in the Intermediate Phase of a Spin-
(44) Buvaylo, E. A.; Kokozay, V. N.; Vassilyeva, O. Y.; Skelton, B. Crossover Material. Chem. - Eur. J. 2010, 16, 14060−14068.
W.; Ozarowski, A.; Titiš, J.; Vranovičová, B.; Boča, R. Field-Assisted (61) Kusz, J.; Nowak, M.; Gütlich, P. Crystal-Structure Studies of
Slow Magnetic Relaxation in a Six-Coordinate Co(II)-Co(III) Mononuclear Iron(II) Complexes with Two-Step Spin Crossover:
Complex with Large Negative Anisotropy. Inorg. Chem. 2017, 56, [Fe{5-NO2-sal-N(1,4,7,10)}] Revisited. Eur. J. Inorg. Chem. 2013,
6999−7009. 2013, 832−842.
(45) Świtlicka, A.; Machura, B.; Penkala, M.; Bieńko, A.; Bieńko, D. (62) Kulmaczewski, R.; Cespedes, O.; Halcrow, M. A. Gradual
C.; Titiš, J.; Rajnák, C.; Boča, R.; Ozarowski, A.; Ozerov, M. Slow Thermal Spin-Crossover Mediated by a Reentrant ZG = 1 → Z = 6 →
Magnetic Relaxation in Cobalt(II) Field-Induced Single-Ion Magnets Z = 1 Phase Transition. Inorg. Chem. 2017, 56, 3144−3148.
with Positive Large Anisotropy. Inorg. Chem. 2018, 57, 12740−12755. (63) Titiš, J.; Boča, R. Magnetostructural D Correlations in
(46) Habib, F.; Korobkov, I.; Murugesu, M. Exposing the Hexacoordinated Cobalt(II) Complexes. Inorg. Chem. 2011, 50,
intermolecular nature of the second relaxation pathway in a 11838−11845.
mononuclear cobalt(II) single-molecule magnet with positive (64) Ye, H.-Y.; Li, S.-H.; Zhang, Y.; Zhou, L.; Deng, F.; Xiong, R.-G.
anisotropy. Dalton Trans. 2015, 44, 6368−6373. Solid State Molecular Dynamic Investigation of An Inclusion
(47) Ceglarska, M.; Stefańczyk, O.; Ohkoshi, S.; Majcher-Fitas, A. Ferroelectric: [(2,6-Diisopropylanilinium)([18]crown-6)]BF4. J. Am.
M. Influence of magnetic dilution on relaxation processes in a solid Chem. Soc. 2014, 136, 10033−10040.
solution comprising tetrahedral Co/ZnII complexes. Dalton Trans. (65) Capel Berdiell, I.; Kulmaczewski, R. K.; Cespedes, O.; Halcrow,
M. A. An Incomplete Spin Transition Associated with a ZZ = 1→ZZ
2020, 49, 6807−6815.
= 24 Crystallographic Symmetry Breaking. Chem. - Eur. J. 2018, 24,
(48) Aubin, S. M. J.; Sun, Z.; Pardi, L.; Krzystek, J.; Folting, K.;
5055−5059.
Brunel, L.-C.; Rheingold, A. L.; Christou, G.; Hendrickson, D. N.
Reduced Anionic Mn12 Molecules with Half-Integer Ground States as
Single-Molecule Magnets. Inorg. Chem. 1999, 38, 5329−5340.
(49) Guo, Y.-N.; Xu, G.-F.; Guo, Y.; Tang, J. Relaxation dynamics of
dysprosium(III) single molecule magnets. Dalton Trans. 2011, 40,
9953−9963.
(50) Chandrasekhar, V.; Dey, A.; Mota, A. J.; Colacio, E. Slow
Magnetic Relaxation in Co(III)-Co(II) Mixed-Valence Dinuclear
Complexes with a CoIIO5X (X = Cl, Br, NO3) Distorted-Octahedral
Coordination Sphere. Inorg. Chem. 2013, 52, 4554−4561.
(51) Lou, H.; Yin, L.; Zhang, B.; Ouyang, Z.-W.; Li, B.; Wang, Z.
Series of Single-Ion and 1D Chain Complexes Based on Quinolinic
Derivative: Synthesis, Crystal Structures, HF-EPR, and Magnetic
Properties. Inorg. Chem. 2018, 57, 7757−7762.
(52) Gómez-Coca, S.; Urtizberea, A.; Cremades, E.; Alonso, P. J.;
Camón, A.; Ruiz, E.; Luis, F. Origin of slow magnetic relaxation in
Kramers ions with non-uniaxial anisotropy. Nat. Commun. 2014, 5,
4300.
(53) Sertphon, D.; Murray, K. S.; Phonsri, W.; Jover, J.; Ruiz, E.;
Telfer, S. G.; Alkaş, A.; Harding, P.; Harding, D. J. Slow relaxation of
magnetization in a bis-mer-tridentate octahedral Co(II) complex.
Dalton Trans. 2018, 47, 859−867.
(54) Seredyuk, B. M.; Gaspar, A. B.; Ksenofontov, V.;
Galyametdinov, Y.; Kusz, J.; Gütlich, P. Iron(II) Metallomesogens
Exhibiting Coupled Spin State and Liquid Crystal Phase Transitions
near Room Temperature. Adv. Funct. Mater. 2008, 18, 2089−2101.
(55) Zhao, X.-H.; Huang, X.-C.; Zhang, S.-L.; Shao, D.; Wei, H.-Y.;
Wang, X.-Y. Cation-Dependent Magnetic Ordering and Room-
Temperature Bistability in Azido-Bridged Perovskite-Type Com-
pounds. J. Am. Chem. Soc. 2013, 135, 16006−16009.
(56) Rosario-Amorin, D.; Dechambenoit, P.; Bentaleb, A.;
Rouzières, M.; Mathonière, C.; Clérac, R. Multistability at Room
Temperature in a Bent-Shaped Spin-Crossover Complex Decorated
with Long Alkyl Chains. J. Am. Chem. Soc. 2018, 140, 98−101.
(57) Agusti, G.; Bartuqal, C.; Martinez, V.; Muñoz-Lara, F. J.;
Gaspar, A. B.; Muñoz, M. C.; Real, J. A. Polymorphism and ‘’reverse’’
spin transition in the spin crossover system [Co(4-terpyridone)2]-
(CF3SO3)2·1H2O. New J. Chem. 2009, 33, 1262−1267.
(58) Bonnet, S.; Siegler, M. A.; Costa, J. S.; Molnár, G.; Bousseksou,
A.; Spek, A. L.; Gamez, P.; Reedijk, J. A two-step spin crossover
mononuclear iron(II) complex with a [HS-LS-LS] intermediate
phase. Chem. Commun. 2008, 5619−5621.
(59) Bréfuel, N.; Watanabe, H.; Toupet, L.; Come, J.; Matsumoto,
N.; Collet, E.; Tanaka, K.; Tuchagues, J.-P. Concerted Spin Crossover
and Symmetry Breaking Yield Three Thermally and One Light-
Induced Crystallographic Phases of a Molecular Material. Angew.
Chem., Int. Ed. 2009, 48, 9304−9307.

J https://dx.doi.org/10.1021/acs.inorgchem.0c00818
Inorg. Chem. XXXX, XXX, XXX−XXX

You might also like