ArticlePDF Available

A hydride-ligated dysprosium single-molecule magnet

Authors:

Abstract and Figures

An experimental and ab initio computational study of an unsymmetrical, hydride-bridged di-dysprosium single-molecule magnet is reported.
Content may be subject to copyright.
This journal is
c
The Royal Society of Chemistry 2013 Chem. Commun., 2013, 49, 901--903 901
Cite this: Chem. Commun., 2013,
49, 901
A hydride-ligated dysprosium single-molecule
magnet
Ajay Venugopal,
a
Floriana Tuna,
bc
Thomas P. Spaniol,
a
Liviu Ungur,
d
Liviu F. Chibotaru,
d
Jun Okuda*
a
and Richard A. Layfield*
b
An experimental and ab initio computational study of an unsym-
metrical, hydride-bridged di-dysprosium single-molecule magnet
is reported.
Lanthanide coordination compounds have accounted for some of
the most important developments in studies of single-molecule
magnets (SMMs), a family of molecules that exhibit magnetic
memory effects.
1
The temperatures at which Ln-SMMs function
far exceed the limits previously set by transition metal SMMs,
2
and
the benchmark SMM blocking temperature was recently re-defined
by [Tb
2
{N(SiMe
3
)
2
}
4
(thf)
2
(m-N
2
)]
, where magnetic hysteresis up to
14 K was measured.
3
Indeed, energy barriers to reversal of the
magnetization (the anisotropy barrier, U
eff
) in Ln-SMMs often exceed
100 cm
1
,
4
which has created renewed confidence that such
materials may eventually be developed for device applications.
5
Single-ion effects, such as the symmetry and electrostatic
potential of the crystal field, strongly influence the relaxation of
the magnetization in SMMs,
1e
and exploring ligands that are new to
Ln-SMMs could push the field in new directions.
6
The hydride
ligand which has never been applied in SMM studies offers an
interesting opportunity because its strong ligand-field effects, and
potential to promote strong exchange, could influence the relaxa-
tion in Ln-SMMs in completely different ways to oxygen-donor
ligands, which are ubiquitous in SMM studies.
1
We now report the
hydride-bridged compounds [Ln(Me
5
trenCH
2
)(m-H)
3
Ln(Me
6
tren)]
[B{C
6
H
3
(CF
3
)
2
}
4
]
2
, where Ln = Gd(III)is[1][X]
2
,Ln=Dy(III)is
[2][X]
2
,andMe
6
tren = tris{2-(dimethylamino)ethylamine. Com-
pound [2][X]
2
is the first hydride-ligated SMM.
Adding [Et
3
NH][B{C
6
H
3
(CF
3
)
2
}
4
]to[Ln(CH
2
SiMe
3
)
3
(thf)
3
]inether
at 40 1C, followed by Me
6
tren and then by dihydrogen produced
[1][X]
2
2Et
2
O and [2][X]
2
2Et
2
O (Scheme 1). Both [1][X]
2
2Et
2
O
(Fig. S1, ESI†) and [2][X]
2
2Et
2
O (Fig. 1) crystallize in the space
group P
%
1, with the Ln centres being related by crystallographic
inversion symmetry. Disorder within the ligands was resolved
with split positions for the carbons. Taking the disorder into
account, the eight-coordinate Ln(
III) ion is complexed by three
m-hydrides, four nitrogens and a carbon atom of a C–H-activated
[Me
5
trenCH
2
]
ligand. The seven-coordinate Ln(III)ioniscom-
plexed by the four nitrogens of the Me
6
tren ligand in addition to
the m-hydrides. Key parameters are presented in Tables S1 and S2
(ESI†). The coordination geometries of the Ln centres in [1]
2+
and
[2]
2+
are similar: the seven-coordinate environment of Ln(1A) can
be roughly described as mono-capped octahedral, whereas the
eight-coordinate Ln(1) occupies a very low symmetry site.
Scheme 1 Synthesis of [1][X]
2
(Ln = Gd) and [ 2][X]
2
(Ln = Dy).
Fig. 1 Structure of [2]
2+
(50% thermal ellipsoids). Only one split positi on of
disordered atoms shown. Hydrogen atoms omitted, except m-hydrides.
a
Institute of Inorganic Chemistry, RWTH Aachen University, Landoltweg 1, 52056
Aachen, Germany. E-mail: jun.okuda@ac.rwth-aachen.de
b
School of Chemistry, The University of Manchester, Oxford Road, Manchester,
M13 9PL, UK. E-mail: Richard.Layfield@manchester.ac.uk
c
EPSRC National UK EPR Facility, Photon Science Institute, The University of
Manchester, Oxford Road, Manchester, M13 9PL, UK
d
Division of Quantum and Physical Chemistry, Katholieke Universiteit Leuven,
Celestijenlaan 200F, 3001 Leuven, Belgium
Electronic supplementary information (ESI) available: Synthetic, analytical,
crystallographic details. Magnetism graphs, computational details. CCDC
901745 and 901746. For ESI and crystallographic data in CIF or other electronic
format see DOI: 10.1039/c2cc38036f
Received 6th November 2012,
Accepted 7th December 2012
DOI: 10.1039/c2cc38036f
www.rsc.org/chemcomm
ChemComm
COMMUNICATION
Published on 11 December 2012. Downloaded by KU Leuven University Library on 23/03/2015 13:46:33.
View Article Online
View Journal
| View Issue
902 Chem. Commun., 2013, 49, 901--903 This journal is
c
The Royal Society of Chemistry 2013
The considerable interest in molecular rare earth hydr ides stems
from their applications in synthesis and catalysis, and from their
potential to activate dihydrogen.
7,8
Solid-state rare-earth hydrides are
valued for their magnetic, optical and semiconductor properties.
9
However, [2][X]
2
provides the first opportunity to investigate the
SMM properties of a rare-earth hydride. Mindful of the crucial role of
crystal field symmetry in SMMs, [2]
2+
also enables an investigation of
how two different, low-symmetry environments within the same
hydride species can impact upon dynamic magnetic behaviour.
Compounds [1][X]
2
and [2][X]
2
were investigated by SQUID
magnetometry in the temperature range 1.8–300 K. At 300 K in a
field of H
dc
= 1000 G, the w
M
T product for [1][X]
2
(w
M
is the molar
magnetic susceptibility) i s 15.2 cm
3
Kmol
1
(Fig. S2, ESI†), which is
consistent with the value of 15.76 cm
3
Kmol
1
predicted for two
non-interacting Gd(
III)ions(
8
S
7/2
, g =2).Oncooling,w
M
T decreases
gradually down to 50 K, and then rapidly to reach 1.22 cm
3
Kmol
1
at 1.8 K, which indicates antiferromagnetic e xchange. In agreement
with this, the magnetisation (M) versus field (H) isotherms are well
below the calculated Brillouin function of two uncoupled S =7/2
ions (Fig. S3, ESI†). The maximum M value at 2 K and 70 kG is
9.3 m
B
, which is much smaller than the calculated value of 15.0 m
B
for
two uncoupled Gd(
III) centres with g = 2. The magnetism of [1][X]
2
was modelled with the Hamiltonian H = JS
Gd1
S
Gd2
,whereJ is the
exchange coupling constant and S
Gd1
and S
Gd2
represent the spins
on the Gd(
III) ions. A good fit of the w
M
T(T) data was obtained with
g =1.995(expectedforGd
3+
)andJ = 1.22 cm
1
, thus indicating
weak antiferromagnetic coupling. Studies of hydride-mediated
exchange are hitherto unknown in lanthanide chemistry, however
the value of J determined for [1][X]
2
is similar to those measured in
other exchange-coupled gadolinium compounds.
3b
The w
M
T data for [2][X]
2
(Fig. S2, ESI†) reveal a value of
28.17 cm
3
K mol
1
at 300 K, in good agreement with the
theoretical value of 28.34 cm
3
K mol
1
for two uncoupled
Dy(
III) ions (
6
H
15/2
, g = 4/3), suggesting that all the m
J
sub-levels
within the electronic ground term of Dy(
III) are populated
at 300 K. At lower temperatures, w
M
T decreases gradually to
24.7 cm
3
K mol
1
at 25 K, and then more rapidly below 25 K, to
reach a value of 16.57 cm
3
K mol
1
at 2 K. The decrease of
w
M
T on cooling can be assigned to depopulation of excited m
J
sub-levels and weak antiferromagnetic exchange. The M(H)
data of 2, acquired at several temperatures between 1.8 and
10 K (Fig. S4, ESI†), reveal a rapid increase of the magnetisation
at small fields, followed by a more gradual increase at high
fields, without reaching saturation. Isothermal M(H/T) curves
do not superimpose, confirming the presence of significant
magnetic anisotropy and/or low-lying excited states. The value
of M at 1.8 K and 70 kG is 10.71 m
B
, (5.35 m
B
per Dy), as typically
observed in other Dy
2
SMMs.
6
The dynamic magnetism of [2]X
2
was investigated using
alternating current (ac) magnetic susceptibility measurements
as a function of temperature and of frequency. In zero dc field,
an ac field of H
ac
= 1.55 G and ac frequencies (n) oscillating at
1–1200 Hz, the in-phase (w
0
M
) (Fig. S5 and S6, ESI†) and the out-
of-phase (w
00
M
) (Fig. 2 and 3) components of the magnetic
susceptibility are temperature and frequency dependent below
22 K. The maximum in w
00
M
ðnÞ gradually shifts to lower frequen-
cies as the temperature is lowered from 13 K to 1.85 K,
indicating that [2]
2+
is an SMM. The temperature dependence
of w
00
M
in [2][X]
2
shows unsymmetrical maxima at several fre-
quencies up to 1200 Hz in zero applied field (Fig. 3). Deconvo-
lution of the overlapping maxima (Fig. S8, ESI†) enabled two
Lorenzian curves to be fitted, and two relaxation processes to be
identified. At fixed temperatures in the range 1.85–12 K, semi-
circular Cole–Cole plots of w
0
M
vs. w
00
M
were obtained, and were
fitted to a generalized Debye model with a = 0.27–0.33 (Fig. S9,
ESI†). The relatively high values of the a parameters imply that
more than one relaxation process is occurring, which is con-
sistent with the unsymmetrical w
00
M
ðT Þ curves, and the presence
of two distinct Dy sites in [2]
2+
.
The relaxation of the magnetization in SMMs can be characteri zed
by a relaxation time, t.
1b
Plots of ln t versus 1/T for [2]
2+
in zero dc field
using w
00
M
ðT Þ and w
00
M
ðnÞ data are linear above 11 K (Fig. S10, ESI†),
suggesting that the magnetization relaxes via an Orbach process
involving the ground and first-excited m
J
levels. The Arrhenius
relationship t = t
0
exp (U
eff
/k
B
T) allowed the anisotropy barrier for
the slower relax ation process to be esti mated as U
eff
=65cm
1
,with
t
0
=1.04 10
7
s. Be low 11 K, a gradual c rossover to a t emperature-
independent regime is observed, suggesting relaxation via quantum
tunnelling of the magnetization (QTM). From the w
00
M
ðnÞ data, the
second relaxation process was estimated to have U
eff
=15cm
1
.
Application of a static field to [2]X
2
reduces the QTM rate. Field-
dependence studies at 4 K showed a rapid increase in t upon
increasing the field from zero to 800 G, and w
00
max
shifts to lower
frequencies. Above 800 G, t decreases rapidly, passing through a
minimum at 2.5 kG (Fig. S11, ESI†). This indicates an additional
relaxation pathway that shortcuts the energy barrier, as predicted at
level crossings when resonant QTM occurs. Application of the
optimum field of 800 G to [2][X]
2
thus shifts the maxima in w
00
M
ðnÞ
to lower frequencies (Fig. 2), and additional maxima at lower
frequencies were observed in w
00
M
ðT Þ (Fig. 3). At 800 G, the Arrhenius
analysis produced U
eff
=85cm
1
and t
0
=5.7 10
9
s (Fig. S10, ESI†).
To gain more insight into the two relaxation processes in [2]
2+
,
ab initio calculations were carried out using MOLCAS 7.6.
10
Fig. 2 w
00
M
ðnÞ isotherms in [2]X
2
with H
ac
= 1.55 G.
Fig. 3 w
00
M
ðT Þ in [2]X
2
with H
ac
= 1.55 G applied at various frequencies.
Communication ChemComm
Published on 11 December 2012. Downloaded by KU Leuven University Library on 23/03/2015 13:46:33.
View Article Online
This journal is
c
The Royal Society of Chemistry 2013 Chem. Commun., 2013, 49, 901--903 903
The calculations revealed that the ground Kramers’ doublet for
each of the two Dy(
III) centres is well separated from the first
excited state, but that the energy separation is 231 cm
1
in the
eight-coordinate Dy(1) and 94 cm
1
for t he seven-coordinate
Dy(1A) (Table 2 and Table S4–S8, ESI†). The calculated g values
of the ground Kramers’ doublets for Dy(1) are g
x
= 0.0187, g
y
=
0.0298 and g
z
= 19.7236; those for Dy(1A) are g
x
= 0.0615, g
y
= 0.1374
and g
z
= 19.2638. Thus, the transverse g-values are approximately
four times smaller for Dy(1) than for Dy(1A), and the different
symmetry of the Dy coordinati on environments influences the
energies and axiality of the Kramers’ doublets.
11
SMMs in which
the Ln centres are weakly exchange coupled relative to the measure-
ment temperature, as in [2]
2+
(see below), are likely to have their
slowly relaxing magnetization associated with individual lanthanide
ions. Thus, despite the high axiality for both dysprosiums, the eig ht-
coordinate Dy(1) will be characterized by a stronger suppression of
QTM, which can explain the observation of only one blocking centre
in the ac susceptibility measurements. In polymetallic Ln-SMMs that
featuremorethanonerelaxationprocess, assigning each processes
to a specific lanthanide is not straightforward using only bulk
susceptibility measurements,
12
so here the calculations are crucial.
The U
eff
value of 65 cm
1
in [2]
2+
inzerofieldismarkedlylessthan
the calculated energies of both first excited Kramers’ doublets. The
discrepancies can be explained by considering the possible relaxation
mechanisms, i.e. an Orbach process via the first excited Kramers’
doublet, or directly from the ground state via QTM and/or a Raman
process. The deviation from linearity of the Arrhenius plot for [2]
2+
in
zero field (Fig. S10, ESI†) shows that the QTM regime is entered below
about T = 11 K, hence the poor match between the experimental U
eff
value and the energies of the first excited Kramers’ doublets. However,
in the optimum dc field of 800 G the QTM is effectively suppressed,
and the close match of the theoretical (94 cm
1
) and experi-
mental (85 cm
1
) energy barriers in Dy(1A) suggests that relaxa-
tion probably occurs via the first excited Kramers’ doublet.
The magnetic moment s of the dysprosiums in the ground
Kramers’ doublets of [2]
2+
intersect at an angle of 4.851 (Fig. 4).
The small values of g
x
and g
y
for Dy(1) and Dy(1A) suggest that an
Ising-type exchange interaction occurs in [2]
2+
. The total magnetic
interaction (J
tot
) is a sum of the magnetic di pole–dipole contribution
(J
dip
) and the exchange part (J
exch
). The magnetic interaction between
the ground Kramers’ doublets on the two dysprosiums is described
by
^
H ¼ðJ
dip
þ J
exch
Þ
^
~
s
1;z
^
~
s
2;z
,wheres
˜
i,z
= 1/2 is the projection of the
pseudospin corresponding to the lowest Kramers doublet on Dy
i
on
the main anisotropy axis, z. Using the calculated g tensors for the
ground Kra mers’ doub lets, J
dip
was calculated exactly as +5.16 cm
1
(Table S11, ESI†). A fit of the experimental dat a allowed
J
exch
= 7.62 cm
1
to be calculated (Fig. S13–S18, Table S11, ESI†).
Whereas the dipolar interaction is ferromagnetic, the exchange
interaction is antiferromagnetic and stronger than the dipolar
contribution, hence the total interaction in [2]
2+
is antiferromagnetic
with J
tot
= 2.46 cm
1
, which is consistent with the exchange
coupling in [1]
2+
.
In summary, the hydride-bridged species [Ln(Me
5
trenCH
2
)-
(m-H)
3
Ln(Me
6
tren)]
2+
, (Ln = Gd, Dy) contain seven-coordinate and
eight-coordinate Ln(
III) centres. Two relaxation processes in the
SMM [2]
2+
were identified from ac susceptibility studies, with
anisotropy barriers of U
eff
=65cm
1
and 15 cm
1
,respectively.
Ab initio calculation of the g-values for Dy(1) and Dy(1A) in [2]
2+
allowed the process with U
eff
=65cm
1
to be assigned to the
eight-coordinate Dy(1), while fast QTM within seven-coordinate
Dy(1A) prevents detection of a blocking process.
We acknowledge the support of the EPSRC (RAL, FT), The
Humboldt Foundation (AV, RAL), the DFG (JO), the FWO-
Vlaanderen (LU), and the KU-Leuven (LU, LFC).
Notes and references
1(a) J. Luzon and R. Sessoli, Dalton Trans., 2012, 41, 13556;
(b) Y.-N. Guo, G.-F. Xu, Y. Guo and J. Tang, Dalton Trans., 2011,
40, 9953; (c) N. Ishikawa, Struct. Bonding, 2010, 135, 211;
(d) L. Sorace, C. Benelli and D. Gatteschi, Chem. Soc. Rev., 2011,
40, 3092; (e) J. D. Rinehart and J. R. Long, Chem. Sci., 2011, 2, 2078.
2 G. Aromı
´
and E. K. Brechin, Struct. Bonding, 2006, 122,1.
3(a) J. D. Rinehart, M. Fang, W. J. Evans and J. R. Long, J. Am. Chem.
Soc., 2011, 133, 14236; (b) J. D. Rinehart, M. Fang, W. J. Evans and
J. R. Long, Nat. Chem., 2011, 3, 538.
4(a)R.J.Blagg,C.A.Muryn,E.J.L.McInnes,F.TunaandR.E.P.
Winpenny, An gew. Chem., Int. Ed., 2011, 50,6530;(b)F.Branzoli,
P. Carretta, M. Filibian, G. Zoppellaro,M.J.Graf,J.R.Galan-Mascaros,
O.Fuhr,S.BrinkandM.Ruben,J. Am. Chem. Soc., 2009, 131, 4387;
(c)N.Ishikawa,M.Sugita,T.Ishikawa, S. Y. Koshihara and Y. Kaizu,
J. Am. Chem . Soc., 200 3, 125, 8694.
5 R. Vincent, S. Klyatskaya, M. Ruben, W. Wernsdorfer and
F. Balestro, Nature, 2012, 488, 357.
6(a) R. A. Layfield, J. J. W. McDouall, S. A Sulway, F. Tuna, D. Collison
andR.E.P.Winpenny,Chem.–Eur. J., 2010, 16, 4442; (b)S.A.Sulway,
R. A. Layfield, F. Tuna, W. Wernsdorfer and R. E. P. Winpenny, Chem.
Commun.,2012,47
, 1508; (c) M. Jeletic, P.-H. Lin, J. J. Le Roy,
I. Korobkov, S. I. Gorelsky and M. Murugesu, J. Am. Chem. Soc.,
2011, 133, 19286; (d) S.-D. Jiang, B.-W. Wang, H. L. Sun, Z. M. Wang
and S. Gao, J. Am. Chem. Soc.,2011,133,4730;(e) S.-D. Jiang, S.-S. Liu,
L.-N. Zhou, B.-W. Wang, Z.-M. Wang and S. Gao, Inorg. Chem., 2012,
51,3079;(f ) F. Tuna, C. A. Smith, M. Bodensteiner, L. Ungur,
L. F. Chibotaru, E. J. L. McInnes, R. E. P. Winpenny, D. Collison
and R. A. Layfield, Angew. Chem., Int. Ed.,2012,51, 6976.
7(a) M. Nishiura and Z. Hou, Nat. Chem., 2010, 2, 257; (b) J. Cheng,
K. Saliu, M. J. Ferguson, R. McDonald and J. Takats, J. Organomet.
Chem., 2010, 695, 2696; (c) M. Konkol and J. Okuda, Coord. Chem.
Rev., 2008, 252, 1577; (d) M. Ephritikine, Chem. Rev., 1997, 97, 2193.
8 A. Venugopal, W. Fegler, T. P. Spaniol, L. Maron and J. Okuda, J. Am.
Chem. Soc., 2011, 133, 17574.
9 A. Simon, Angew. Chem., Int. Ed., 2012, 51, 4280.
10 (a)F.Aquilante,L.DeVico,N.Ferre,G.Ghigo,P.A.Malmqvist,
P.Neogrady,T.B.Pedersen,M.Pitonak,M.Reiher,B.O.Roos,
L.Serrano-Andres,M.Urban,V.VeryazoandR.Lindh,J. Compu t. Chem.,
2010, 31, 224; (b)L.F.ChibotaruandL.Ungur,J. Chem. Phys., 2012,
137, 064112.
11 (a) H. L. C. Feltham, Y. H. Lan, F. Klower, L. Ungur, L. F. Chibotaru,
A. K. Powell and S. Brooker, Chem.–Eur. J., 2011, 17, 4362; (b)L.Ungur
and L. F. Chibotaru, Phys. Chem. Chem. Phys.,2011,13, 20086.
12 (a
) P.-H. Lin, W.-B. Sun, M.-F. Yu, G .-M. Li, P.-F. Yan and M. Murugesu,
Chem. Commun., 2011, 47, 10993; (b)Y.N.Guo,G.F.Xu,W.Wernsdorfer,
L.Ungur,Y.Guo,J.Tang,H.J.Zhang,L.F.ChibotaruandA.K.Powell,
J. Am. Chem. Soc., 2011, 133,5319.Seealso:(c)F.Habib,P.H.Lin,J.Long,
I.Korobkov,W.WernsdorferandM.Murugesu,J. Am. Chem. Soc., 2011,
133, 8830.
Fig. 4 Orientation of the magnetic moments in the ground Kramers’ doublet of
[2]
2+
(dashed lines). The arrows show the antiferromagnetic coupling. Pink
atoms = hydrides, blue = N, grey = C.
ChemComm Communication
Published on 11 December 2012. Downloaded by KU Leuven University Library on 23/03/2015 13:46:33.
View Article Online
... A series of binuclear dicationic trihydride complexes of rare-earth metals [(Me 6 TREN)Ln(μ-H) 3 Ln(CH 2 TACDMe 3 )][BR 4 ] 2 (Ar = C 6 H 3 (CF 3 ) 2 -3,5; Ln = Y (103), Gd (104), Dy (105), Lu (106); Scheme 28), analogous to dicationic Lu complex 102, were obtained upon hydrogenolysis of the corresponding bis(alkyl) complexes (Me 6 TREN)Ln(CH 2 SiMe 3 ) 2 , bearing neutral tris(dimethylaminoethyl)amine (N(CH 2 CH 2 NMe 2 ) 3 = Me 6 TREN) as a stabilizing ligand, with molecular H 2 [67,68]. The interaction of bis(alkyl) complex (Me 6 TREN)Ln(CH 2 SiMe 3 ) 2 with H 2 was accompanied by the activation of the С sp3 −H bond of one of the methyl groups of the NMe 2 moieties, giving rise to binuclear complexes 103-106 in which one of the metal centers is coordinated by the neutral ligand (Me 6 TREN), whereas the second one is bound with the monoanionic ligand (Me 5 TRENCH 2 ) − . ...
Article
Comprehensive Summary One dihydride‐bridged dimeric Dy(III) guanidinate complex, formulated as [{(Me 3 Si) 2 NC(N ⁱ Pr) 2 } 2 Dy( μ ‐H)] 2 ( 1Dy ), was successfully isolated and the introduction of hydride bridges significantly reduces the intramolecular Dy(III)···Dy(III) distance to only 3.688(1) Å. To investigate the effect of such a short Dy(III)···Dy(III) distance for magnetism, we also prepared its dibromide‐bridged analogue [{(Me 3 Si) 2 NC(N ⁱ Pr) 2 } 2 Dy( μ ‐Br)] 2 ( 2Dy ), which has a much longer Dy(III)···Dy(III) distance of 4.605(4) Å. Surprisingly, 2Dy demonstrates much larger effective energy barrier for magnetization reversal ( U eff ) and higher blocking temperature ( T B ). The worse performance of 1Dy is attributed to the concerted effect of strong antiferromagnetic interactions between Dy(III) ions ( J total = ‐2.683 cm ‐1 ) and the unparallel arrangement of magnetic principle axes of the Dy(III) ions for 1Dy . This article is protected by copyright. All rights reserved.
Article
The reduction of a bimetallic yttrium ansa-metallocene hydride was examined to explore the possible formation of Y-Y bonds with 4d1 Y(II) ions. The precursor [CpAnY(μ-H)(THF)]2 (CpAn = Me2Si[C5H3(SiMe3)-3]2) was synthesized by hydrogenolysis of the allyl complex CpAnY(η3-C3H5)(THF), which was prepared from (C3H5)MgCl and [CpAnY(μ-Cl)]2. Treatment of [CpAnY(μ-H)(THF)]2 with excess KC8 in the presence of one equivalent of 2.2.2-cryptand (crypt) generates an intensely colored red-brown product crystallographically identified as [K(crypt)][(μ-CpAn)Y(μ-H)]2. The two rings of each CpAn ligand in the reduced anion [(μ-CpAn)Y(μ-H)]21- are attached to two yttrium centers in a "flyover" configuration. The 3.3992(6) and 3.4022(7) Å Y···Y distances between the equivalent metal centers within two crystallographically independent complexes are the shortest Y···Y distances observed to date. Ultraviolet-visible (UV-visible)/near infrared (IR) and electron paramagnetic resonance (EPR) spectroscopy support the presence of Y(II), and theoretical analysis describes the singly occupied molecular orbital (SOMO) as an Y-Y bonding orbital composed of metal 4d orbitals mixed with metallocene ligand orbitals. A dysprosium analogue, [K(18-crown-6)(THF)2][(μ-CpAn)Dy(μ-H)]2, was also synthesized, crystallographically characterized, and studied by variable temperature magnetic susceptibility. The magnetic data are best modeled with the presence of one 4f9 Dy(III) center and one 4f9(5dz2)1 Dy(II) center with no coupling between them. CASSCF calculations are consistent with magnetic measurements supporting the absence of coupling between the Dy centers.
Chapter
This chapter summarizes a decade of research in the field of organometallic single-molecule magnets (SMMs), with a focus on SMMs containing lanthanides or actinides. Dysprosium organometallics are prominent in the chapter, particularly metallocene SMMs, which account for some of the most impressive systems yet known. The relationship between the electronic structure and magnetism of certain other key Ln3 + ions and their crystal field interactions with organometallic ligands such as cyclopentadienyl and cyclooctatetraenyl are described in detail. The use of organometallic chemistry as a complement to classical coordination chemistry in the design of SMMs is presented, and shown to be a highly effective approach to improving key SMM performance metrics, including the effective energy barrier to reversal of the magnetization and the magnetic hysteresis properties.
Article
Full-text available
While the blocking barrier (Ueff) and blocking temperature (TB) for “Dysprocenium” SIMs were increased beyond liquid N2 temperature, device fabrication of these molecules remains a challenge as low coordinate Ln3+...
Article
Full-text available
Hydrogenolysis of the half-sandwich penta-arylcycopentadienyl-supported rare-earth metal dibenzyl complexes [(Cp Ar5 )Ln( p -CH 2 -C 6 H 4 -Me) 2 (THF)] (Cp Ar5 = C 5 Ar 5 , Ar = 3,5- ⁱ Pr 2 -C 6 H 3 ; Ln = Sc, La) afforded bimetallic scandium complex [(Cp Ar5 )Sc(H)(μ-OC 4 H 9 )] 2 ( 2 ) with two terminal hydrido...
Chapter
This article provides broad coverage of the key developments in lanthanide coordination chemistry during the period 2003–19. The term lanthanide covers the elements in the series from cerium to lutetium, but does not include very detailed discussion of coordination compounds of the other three rare earth elements, namely scandium, yttrium and lanthanum. The article places some emphasis on unconventional ligand types that have not previously been included in the articles on the lanthanides earlier editions of Comprehensive Coordination Chemistry, notably hydride and ligands in which the donor atoms include p-block metalloid and metallic elements, such as the heavier Group 13 elements and the heavier Group 14 elements. Bonds between lanthanides and transition metals are also reviewed. Complexes that can readily be identified as belonging to the ‘organometallic’ family are mentioned in cases where the chemistry represents an important development for the lanthanide elements in general, notably the discovery of new oxidation states, single-molecule magnets and small-molecule activation. More traditional Werner-type coordination compounds form the bulk of the article, including vast numbers of N- and O-donor ligands, including many beautiful lanthanide complexes of multidentate and macrocyclic ligands. The applications of such complexes in luminescence and as materials for imaging are summarized, as are their molecular magnetic properties. Whilst every effort has been made to provide comprehensive coverage of topics in lanthanide chemistry, the sheer volume of research activity in this vibrant field makes citation of every study practically impossible. Many more fascinating studies in lanthanide chemistry can be accessed via the extensive list of review articles cited in the following pages.
Article
The aggregation and formation of heterometallic nanoclusters usually involves a variety of complex self-assembly processes; thus, the exploration of their assembly mechanisms through process tracking is more challenging than that for homometallic nanoclusters. We explored here the effect of solvent on the formation of heterometallic clusters, which gave two heterometallic nanoclusters, [Dy2Co8(μ3-OCH3)2(L)4(HL)2(OAc)2(NO3)2(CH3CN)2]·CH3CN·H2O (1) and [Dy4Co6(L)4(HL)2(OAc)6(OCH2CH2OH)2(HOCH2CH2OH)(H2O)]·9CH3CN (2), with the H3L ligand formed from the in situ condensation reaction of 3-amino-1,2-propanediol with 2-hydroxy-1-naphthaldehyde in the presence of Co(OAc)2·4H2O and Dy(NO)3·6H2O. It is worth noting that the skeleton of cluster 1 has a high stability under high-resolution electrospray ionization mass spectrometry (HRESI-MS) conditions with a gradually increasing energy of the ion source. Cluster 2 underwent a multistep fragmentation even under a zero ion-source voltage for the measurement of HRESI-MS. Further analysis showed that cluster 2 underwent a possible fragmentation mechanism of Dy4Co6L6 → Dy2Co6L5/DyL → DyCo2L3/DyCo2L → DyL/Co2L2. Most notably, the species emerging in the formation process of cluster 1 were tracked using time-dependent HRESI-MS, from which we proposed its possible formation mechanism of H2L → Co2L2 → Co2DyL2/Co3L2 → Co3DyL2 → Co4DyL2 → Co5Dy2L4 → Co8Dy2L6. As far as we know, it is the first time to track the formation process of Dy-Co heterometallic clusters through HRESI-MS with the proposed assembly mechanism. The magnetic properties of the two titled DyIII x CoII10-x (x = 2, 4) clusters were studied. Both of them exhibit slow magnetic relaxation, and 1 is a single-molecule magnet at zero direct-current field.
Article
Based on a multidentate Schiff-base ligand, N,N′-bis(2-hydroxy-5-methyl-3-formylbenzyl)-N,N′-bis-(pyridin-2-ylmethyl)ethylenediamine (H2L), where two binuclear DyIII compounds, with formulas [Dy2(L)(NO3)3(CH3O)] (1) and [Dy2(L)(NO3)3(CH3CH2O)] (2), have been synthesized under different solvent systems. DyIII ions in 1 and 2 adopt monocapped square antiprism coordination geometries, while the different structural distortions can be observed. Two DyIII ions in 1 and 2 are bridged by two phenoxide atoms of one L2- ligand and one bridged CH3O-/CH3CH2O- oxygen node, leading to an approximate fusiform Dy2O3 core. The different DyIII−Obridged node distances, DyIII−Obridged node−DyIII angles and DyIII…DyIII distances can be observed. Magnetic studies reveal that 1 and 2 display slow magnetic relaxation behaviours under a zero direct-current field with the effective energy barriers (Ueff) of 114.17 K and 171.23 K, respectively. What’s more, compound 2 possesses the highest Ueff in nine-coordinated Dy2 compounds. The M versus H data exhibit weak butterfly shaped hysteresis loops at 2 K for 2. The rare CH3O-/CH3CH2O--bridged nine-coordinated binuclear DyIII single-molecule magnets (SMMs) significantly regulate and enhance Ueff of compounds 1 and 2. To deeply understand their different magnetic behaviours, the magnetic anisotropies and magnetic interactions of 1 and 2 were studied by ab initio calculations. These findings demonstrate an efficient approach to regulating and enhancing the magnetic anisotropy barriers through using bridged CH3O- anion or CH3CH2O- anion.
Article
Full-text available
Scientists have long employed lanthanide elements in the design of materials with extraordinary magnetic properties, including the strongest magnets known, SmCo5 and Nd2Fe14B. The properties of these materials are largely a product of fine-tuning the interaction between the lanthanide ion and the crystal lattice. Recently, synthetic chemists have begun to utilize f-elements-both lanthanides and actinides-for the construction of single-molecule magnets, resulting in a rapid expansion of the field. The desirable magnetic characteristics of the f-elements are contingent upon the interaction between the single-ion electron density and the crystal field environment in which it is placed. This interaction leads to the single-ion anisotropies requisite for strong single-molecule magnets. Therefore, it is of vital importance to understand the particular crystal field environments that could lead to maximization of the anisotropy for individual f-elements. Here, we summarize a qualitative method for predicting the ligand architectures that will generate magnetic anisotropy for a variety of f-element ions. It is hoped that this simple model will serve to guide the design of stronger single-molecule magnets incorporating the f-elements.
Article
Full-text available
A methodology for the rigorous nonperturbative derivation of magnetic pseudospin Hamiltonians of mononuclear complexes and fragments based on ab initio calculations of their electronic structure is described. It is supposed that the spin-orbit coupling and other relativistic effects are already taken fully into account at the stage of quantum chemistry calculations of complexes. The methodology is based on the establishment of the correspondence between the ab initio wave functions of the chosen manifold of multielectronic states and the pseudospin eigenfunctions, which allows to define the pseudospin Hamiltonians in the unique way. Working expressions are derived for the pseudospin Zeeman and zero-field splitting Hamiltonian corresponding to arbitrary pseudospins. The proposed calculation methodology, already implemented in the SINGLE_ANISO module of the MOLCAS-7.6 quantum chemistry package, is applied for a first-principles evaluation of pseudospin Hamiltonians of several complexes exhibiting weak, moderate, and very strong spin-orbit coupling effects.
Article
Full-text available
Quantum control of individual spins in condensed-matter devices is an emerging field with a wide range of applications, from nanospintronics to quantum computing. The electron, possessing spin and orbital degrees of freedom, is conventionally used as the carrier of quantum information in proposed devices. However, electrons couple strongly to the environment, and so have very short relaxation and coherence times. It is therefore extremely difficult to achieve quantum coherence and stable entanglement of electron spins. Alternative concepts propose nuclear spins as the building blocks for quantum computing, because such spins are extremely well isolated from the environment and less prone to decoherence. However, weak coupling comes at a price: it remains challenging to address and manipulate individual nuclear spins. Here we show that the nuclear spin of an individual metal atom embedded in a single-molecule magnet can be read out electronically. The observed long lifetimes (tens of seconds) and relaxation characteristics of nuclear spin at the single-atom scale open the way to a completely new world of devices in which quantum logic may be implemented.
Article
Due to their usual large magnetic moments and large magnetic anisotropy lanthanide ions are investigated for the search of Single Molecule Magnets with high blocking temperature. However, the low symmetry crystal environment, the complexity of the electronic states or the non-collinearity of the magnetic anisotropy easy-axes in polynuclear systems make the rationalization of the magnetic behaviour of lanthanide based molecular systems difficult. In this perspective article we expose a methodology in which the use of additional characterization techniques, like single crystal magnetic measurements or luminescence experiments, complemented by relativistic ab initio calculations and a suitable choice of spin Hamiltonian models, can be of great help in order to overcome such difficulties, representing an essential step for the rational design of lanthanide based Single Molecule Magnets with enhanced physical properties.
Article
The synthesis and chemistry of monoligand lanthanide dihydrides, LLnH2, is a very active current area of research. Herein we summarize the status of our contributions utilizing various scorpionates, TpR,R′, as protective ancillary ligands and show that the nuclearity of the so obtained hydride clusters depends on the size of the scorpionate ligands and, in one instance, the solvent used in the synthesis. Following brief consideration of the synthesis of the various precursor dialkyl complexes, (TpR,R′)Ln(CH2SiMe3)2(THF)1/0, the solid-state structures of the hydride clusters, [(TpR,R′)LnH2]n (R, R′ = Me, Ln = Nd, Sm, Y, Yb and Lu, n = 4; R, R′ = H, Ln = Y, Yb and Lu, n = 6; R, R′ = iPr, Ln = Y and Lu, n = 3), obtained via hydrogenolysis, are described.Graphical abstractLanthanide polyhydrides of different sizes have been obtained by hydrogenolysis of various scorpionate anchored lanthanide dialkyl complexes, (TpR,R′)Ln(CH2SiMe3)2(THF). The nuclearity of the cluster hydrides can be controlled by judicious choice of the scorpionate ligand. The solid-state structures of the currently available [(TpR,R′)LnH2]n complexes are reviewed.
Chapter
Magnetism and electronic structure of two types of phthalocyanine-based magnets, “ferromagnets” and “single-molecule magnets,” both of which exhibit spontaneous magnetization but by different mechanism, are reviewed. KeywordsElectronic structure-Ferromagnet-Phthalocyanine-Single molecule magnet
Chapter
The phenomenon of single-molecule magnetism was established in the early 1990s with the cluster [Mn12O12(AcO)16(H2O)4]. Since then, alarge number of compounds displaying this behaviour have been synthesized. The vast majority of these consist of coordination clusters of 3d transition metals prepared by self assembly processes where the structure of the final product has not been predicted. The majority feature manganese ions, primarily in the MnIII oxidation state, but over the years the search for novel examples of such compounds has been extended to other metals of the 3rd row, including iron, nickel, vanadium and cobalt. Ahost of new approaches and synthetic methodologies have been successfully explored and developed in order to prepare SMMs with new and improved properties. This review covers the synthesis and magnetic properties of SMMs consisting of homometallic coordination clusters of 3d transition metals made via serendipitous self-assembly – by far the largest source of such species – together with abrief overview of emerging methods.
Article
The sulfur-bridged dimers [{Cp'(2)Ln(μ-SSiPh(3))}(2)] (Ln=Gd (1), Dy (2); Cp'=η(5)-C(5)H(4)Me) were synthesized by the transmetalation reactions between [Cp'(3)Ln] and Ph(3)SiSLi. Compound 2 is a single-molecule magnet with slow relaxation of magnetization up to 40 K and an anisotropy barrier of U(eff) =133 cm(-1). Insight into the SMM properties of 2 and closely related SMMs has been obtained using ab initio calculations.
Article
The experimental and theoretical study of the electron spin dynamics in the anionic form of a single-ion molecule magnet (SIMM), the bis-phthalocyaninato terbium (III) molecule [Pc2Tb]-[TBA]+, has been addressed by means of solid state 1H NMR spectroscopy. The magnetic properties of the caged Tb3+ metal center were investigated in a series of diamagnetically diluted preparations, where the excess of tetrabutylamonium bromide ([TBA]Br)n salt was used as diamagnetic matrix complement. We found that a high temperature activated spin dynamics characterizes the systems, which involved phonon-assisted transitions among the crystal field levels in qualitative agreements with literature results. However, the activation barriers in these processes range from 641 cm-1 for the diamagnetically diluted samples to 584 cm-1 for those undiluted; thus, they exhibit barriers 2-3 times larger than witnessed in earlier (230 cm-1) reports (e.g., Ishikawa, N.; Sugita, M.; Ishikawa, T.; Koshihara, S.; Kaizu, Y. J. Am. Chem. Soc. 2003, 125, 8694-8695). At cryogenic temperatures, fluctuations are driven by tunneling processes between the m ) +6 and -6 low-energy levels. We found that the barrier Δ and the tunneling rates change from sample to sample and especially the diamagnetically diluted [Pc2Tb]- molecules appear affected by the sample’s magneto/thermal history. These observations emphasize that matrix arrangements around [Pc2Tb]- can appreciably alter the splitting of the crystal field levels, its symmetry, and hence, the spin dynamics. Therefore, understanding how small differences in molecular surroundings (as for instance occurring by depositing on surfaces) can trigger substantial modifications in the SIMM property is of utmost importance for the effective operation of such molecules for single-molecule data storage, for example.
Article
The chemistry of the hydride complexes of the rare-earth metals has been dominated by complexes with metallocene and half-sandwich ligand frameworks. In contrast to hydride complexes supported by cyclopentadienyl ligands, only relatively few examples of non-metallocene hydride complexes have been reported in the literature. Such complexes are potentially of some interest in view of their chemical reactivity and catalytic applications. This review surveys the chemistry of non-metallocene hydride complexes up to mid-2007, some spectroscopic and structural aspects as well as their reactivity.