Вы находитесь на странице: 1из 6

Journal of Physics and Chemistry of Solids 65 (2004) 11–16

www.elsevier.com/locate/jpcs

Synthesis and characterization of Co7(OH)12(C2H4S2O6)(H2O)2—a single


crystal structural study of a ferrimagnetic layered cobalt hydroxide
Paul M. Forster, Maya M. Tafoya, Anthony K. Cheetham*
Materials Research Laboratory, University of California, Santa Barbara, CA 93106-5121, USA
Received 26 June 2003; revised 22 August 2003; accepted 28 August 2003
Dedicated to Prof. Peter Day on the occasion of his 65th birthday

Abstract
A novel layered hydrotalcite-like material, Co7(H2O)2(OH)12(C2H4S2O6), has been prepared hydrothermally and the structure determined
using single crystal X-ray diffraction (a ¼ 6.2752(19) Å, b ¼ 8.361(3) Å, c ¼ 9.642(3) Å, a ¼ 96.613(5)8, b ¼ 98.230(5)8,
g ¼ 100.673(5)8, R1 ¼ 0:0551). The structure consists of brucite-like sheets where 1/6 of the octahedral sites are replaced by two
tetrahedrally coordinated Co(II) above and below the plane of the layer. Ethanedisulfonate anions occupy the space between layers and
provide charge balance for the positively charged layers. The compound is ferrimagnetic, with a Curie temperature of 33 K, Curie –Weiss u
of 2 31 K, and a coercive field of 881 Oe at 5 K.
q 2003 Published by Elsevier Ltd.
Keywords: C. X-ray diffraction; D. Crystal structure

1. Introduction Despite having many important properties [8], reliable


structural information on layered hydroxides is relatively
Brucite, the structure adopted by many divalent metal scarce. Diffraction studies are typically limited by disorder,
hydroxides including cobalt [1], may be described as sheets stacking faults, and severe broadening associated with the
with two hexagonal close-packed layers of hydroxide where small particle size typical of these hydroxides. Sample
all of the octahedral sites within the sheet are occupied by inhomogeneity is also a problem, particularly for materials
cations. Since ferrimagnetic ordering occurs within the rapidly precipitated by the addition of base [9]. Accurate
sheets in many of the parent transition metal hydroxides [2], structural information is important to better interpret
layered hydroxides based on transition metals are of interest magnetic exchange interactions between layers or to
as magnetic materials [3]. In addition, because no strong understand the preference of a material for particular
bonding exists between these sheets, ion exchange proper- anions. For the majority of layered hydroxides, especially
ties may be introduced by modifying the sheets to carry a those containing organic molecules, only broad 00l reflec-
charge that is compensated by non-coordinating ions tions are observed. While these peaks provide information
between the layers. Anion exchange is of particular on the space between sheets, the geometry of interlayer
importance in this class of materials as it is rare in water and organic molecules must normally be inferred
comparison to cation exchange, which is well-known in from geometric considerations or spectroscopy.
many materials such as zeolites [4]. Because a wide range of There are several methods by which charge may be
anion sizes may be accommodated, these materials are also
introduced into brucite-like sheets. The most convention-
promising for selective separations of a variety of negatively
al materials, layered double hydroxides, are made by the
charged species [5]. Other work has demonstrated that
substitution of trivalent cations, such as Al3þ, Cr3þ,
catalytically active materials [6] may be made by using, for
Mn3þ, Fe3þ, Ga3þ, Co3þ, or Ni3þ into brucite sheets
example, oxo-anion pillars to form a porous structure [7].
[10]. Secondly, hydroxides from the brucite layer may be
* Corresponding author. Fax: þ 1-805-893-3797. partially replaced by coordinating molecules from anions,
E-mail address: cheetham@mrl.ucsb.edu (A.K. Cheetham). as has been crystallographically observed for nitrates,
0022-3697/$ - see front matter q 2003 Published by Elsevier Ltd.
doi:10.1016/j.jpcs.2003.08.016
12 P.M. Forster et al. / Journal of Physics and Chemistry of Solids 65 (2004) 11–16

carbonates, and carboxylates [11]. Finally, a fraction of 2. Experimental


the octahedral sites may be replaced by two divalent
tetrahedrally coordinated atoms above and below the Synthesis. Plate-like blue – green crystals of [1], along
sheets. This type of substitution was first observed in with a small amount of purple crystals, were prepared by
Zn5(OH)8(NO3)2·2H2O in 1970 [12]. A nearly isostruc- reaction of 0.5 g Co(OH)2 (Alfa Aesar) and 1.02 g 1,2
tural cadmium analogue was subsequently reported [13]. ethanedisulfonic acid (TCI) in 5 g deionized water (ratio
In this structure, two tetrahedra replace one out of every 1:1:52) at 200 8C in a Teflon-lined autoclave for two days.
four octahedral sites. The three oxygen atoms forming The purple compound, along with several other materials in
the base of these tetrahedra are from the brucite layer, this system, will be described separately [16]. The products
and the fourth is a coordinated water molecule. This type were separated by filtration, rinsed with water, and dried in
of substitution has only small effects on the arrangement air. Pure [1] was subsequently obtained by lowering the
of the hydroxide anions, but introduces an additional amount of ethanedisulfonic acid added to otherwise similar
divalent cation for each substitution without changing the reactions (such as 1:0.5:52), although the crystal size
number of hydroxide anions in the structure, resulting in appears to decrease under such conditions. Powder X-ray
a net positive charge for the layers. diffraction (XRD) indicated that [1] was free of visible
Some sulfates contain both tetrahedral cations and cobalt hydroxide peaks, and sharp peaks over the range
substitute SO22 oxygen atoms for hydroxide anions scanned (908 2u, Cu Ka) demonstrated good crystallinity
4
(Fig. 1).
within brucite sheets. Recently, several fascinating
A suitable single crystal was selected under a polarizing
structures of this type have been characterized from
microscope, glued to a glass fiber, and mounted on a
single crystals by Rosseinsky and co-workers; Co4(SO4)
Siemens SMART CCD diffractometer equipped with
(OH)6(H2NC2H3NH2)0.5·3H2O [14] and Co4(SO4)(OH)6
Mo Ka radiation (0.71073 Å). Following data collection at
(C6N2H12)0.5·H2O [15] contain sheets similar to those
room temperature, the structure was solved using direct
seen in the basic zinc sulfate structure. In both structures,
methods and refined against lF 2 l using the SHELX-TL suite
the diamine acts as a pillar by coordinating through N
[17]. SADABS was used to correct for absorption.
atoms with two tetrahedra from adjacent layers. Sulfate Hydrogen atoms were placed on the acid molecules using
groups, which do not appear to be exchangeable, the riding model. Hydrogen atoms on the hydroxide
coordinate to the sheets between tetrahedra by replacing molecules were located in the Fourier map, and their
a hydroxide anion with a sulfate oxygen atom. Both temperature factors were constrained to equal 1.2 times the
compounds exhibit reversible hydration/dehydration and isotropic temperature factor of the oxygen atom to which
show metamagnetism at low temperatures, as is typical they were bound. The hydrogen atoms on hydroxide atoms
for systems with spacing below 10 Å between sheets. within the layers did not refine to a reasonable O –H
Here, we present the synthesis, structure determination, distances, and were constrained to be equal to 0.86 Å within
and magnetic behavior of a novel hydrotalcite-like material, a standard deviation of 0.02 Å. They subsequently refined to
Co7(H2O)2(OH)12(C2H4S2O6) [1]. This material contains chemically reasonable positions. H atoms were not located
layers with a previously unknown arrangement of tetra- on the water molecule, presumably to the greater thermal
hedral sites above and below brucite-like layers. [1] also motion of the water molecule compared to the hydroxide
orders ferrimagnetically, with a Curie temperature of 33 K. oxygen atoms and the relative positional flexibility of these

Fig. 1. Powder XRD pattern for [1]. Data past 308 2u have been multiplied by 20. The pattern is dominated by 00l reflections, in part due to preferred
orientation.
P.M. Forster et al. / Journal of Physics and Chemistry of Solids 65 (2004) 11–16 13

Table 1 a Quantum-Design MPMS-5S SQUID magnetometer. A


Crystallographic data summary pure sample, as verified by powder XRD, was obtained by
Formula Co7(H2O)2(OH)12(C2H4S2O6) separating the purple impurity phase manually. The sample
was cooled in a zero field and data then collected on
Space group P-1 warming in applied magnetic fields of 500 and 5000 Oe.
a (Å) 6.2752(19) The sample was warmed to 300 K between each step.
b (Å) 8.361(3) Susceptibility data between 200 and 300 K were used to
c (Å) 9.642(3)
determine the Curie – Weiss temperature. An additional scan
a (deg) 96.613(5)
b (deg) 98.230(5) was performed after heating to 300 K under zero applied
g (deg) 100.673(5) field in order to observe spontaneous magnetization
Volume (Å3) 486.8(3)
Z 1
Calc. density (mg/m3) 2.841
3. Results and discussion
Reflections collected 4917
Independent reflections completeness to 2250 Rint ¼ 0.0421
u ¼ 28:28 The synthesis of crystalline metal hydroxide materials is
Data/restraints/parameters 2250/6/160 difficult owing to the extreme insolubility of the products
Goodness-of-fit on F 2 1.007 under typical synthesis conditions. Our approach to the
Final R indices ½I . 2sI R1 ¼ 0:0551; wR2 ¼ 0:1396
crystallization of [1] clearly produces well-ordered material.
R indices (all data) R1 ¼ 0:0624; wR2 ¼ 0:1448
It is likely that the use of an extremely insoluble Co source,
cobalt hydroxide, limited the available Co2þ in solution and
protons compared to hydroxide protons. A summary of consequently suppressed nucleation. Hydrothermal reaction
refinement parameters appears in Table 1 and selected bond conditions and the presence of ethanedisulfonic acid
lengths in Table 2. Tables of atomic positions, complete undoubtedly increased the dissolution rate sufficiently to
bond lengths and angles have been deposited as supplemen- allow the reaction to proceed over a reasonable time scale.
tary materials. Crystallographic data (excluding structure However, the time scale is still far longer than typically
factors) have been deposited with the Cambridge Crystal- occurs with bench-top reactions where the addition of base
lographic Data Centre as number 218100. Copies of the data initiates the rapid precipitation of layered hydroxide
can be obtained free of charge on application to the CCDC, organics. It is unlikely a coincidence that the only two
12 Union Rd, Cambridge CB21EZ, UK (Fax: þ 44-1223- related materials containing organic anions were also
336-033; email: deposit@ccdc.cam.ac.uk). prepared hydrothermally [14,15]. We anticipate that the
Magnetic susceptibility data were measured on a strategy we adopted for preparing [1] may be generally
powdered sample in the temperature range 2– 300 K using useful for the preparation of other materials with similar
high crystallinity by using different anions. Cobalt hydrox-
Table 2 ide has been used previously as a reagent to synthesize
Selected bond lengths (Å) and angles (deg) for [1] hybrid metal –organic compounds [18]. However, hydro-
Co(1)–O(4) £ 2 2.058(3) thermal synthesis may also result in other phases with metal
Co(1)–O(1) £ 2 2.135(3) hydroxide sheets that differ significantly from brucite [19].
Co(1)–O(2) £ 2 2.149(3) The structure of [1] (Fig. 2) is based on brucite-like
Co(2)–O(3) 2.058(4) sheets separated by charged ethanedisulfonate anions.
Co(2)–O(4) 2.072(3)
These layers contain a novel arrangement of tetrahedral
Co(2)–O(5) 2.104(3)
Co(2)–O(3) 2.117(3) sites, shown in Fig. 2(b), which can be described as rows
Co(2)–O(6) 2.146(3)
Co(2)–O(2) 2.156(3)
Co(3)–O(4) 2.079(3)
Co(3)–O(3) 2.081(3)
Co(3)–O(5) 2.103(3)
Co(3)–O(5)#2 2.107(4)
Co(3)–O(6) 2.154(3)
Co(3)–O(1) 2.155(3)
Co(4)–O(6) 1.942(3)
Co(4)–O(2) 1.948(3)
Co(4)–O(1) 1.951(3)
Co(4)–O(7) 2.092(4)
C(1)–C(1) 1.533(10)
C(1)–S(1) 1.770(6)
S(1)–O(13) 1.450(4) Fig. 2. (a) View of [1] perpendicular to the a direction, with octahedral
S(1)–O(12) 1.456(4) cobalt represented as purple polyhedra, tetrahedral cobalt as blue
S(1)–O(11) 1.463(4) tetrahedra, carbon, sulfur, and oxygen as gray, yellow, and red,
respectively. (b) A single sheet viewed perpendicular to the c axis.
14 P.M. Forster et al. / Journal of Physics and Chemistry of Solids 65 (2004) 11–16

hydroxides coordinate with one tetrahedral and two


octahedral cobalt atoms. The remaining three have very
similar coordination to hydroxides in brucite.
Unlike many other layered hydroxide materials contain-
ing organic anions, [1] has no uncoordinated water between
layers as evidenced by Platon [21], which did not find any
solvent accessible space. This is presumably a result of very
efficient filling of the interlayer space by ethanedisulfonate
molecules. Significant hydrogen bonding occurs between
hydroxyl groups on the surface of the brucite-like layers and
the oxygen atoms in the ethanedisulfonate anions between
the layers, with four O – O distances between 2.8 and 3 Å.
Fig. 3. Asymmetric unit of [1]. Crystallographically equivalent hydroxide Although we were unable to locate hydrogen atoms on the
anions have been added to complete the coordination spheres of the cobalt water molecule O7 that coordinates to the tetrahedral cobalt
atoms. atom, it is clearly involved in several hydrogen bonding
interactions with crystallographically equivalent water
where alternately 1/3 or none of the octahedra have been molecules in adjacent layers (O– O separation 2.987(8) Å)
replaced by two tetrahedral cobalt sites (Fig. 2). To our and a sulfonate oxygen atom (O – O separation 2.787(9) Å).
knowledge, the only sheets characterized crystallographi- The efficient packing and the extensive hydrogen-bonding
cally to date contain 1/4 or 1/7 tetrahedral substitutions. network in this particular structure may contribute to our
Sheets of similar composition, based on elemental analysis, ability to grow highly crystalline samples hydrothermally.
have been observed for other cobalt alkylsulfonate materials Magnetic susceptibility measurements, performed on a
[20]. Although these monosulfonate materials could not be powdered sample, reveal a ferromagnetic transition between
characterized crystallographically, it is probable that their 30 and 35 K (Fig. 4). Fitting the paramagnetic region
sheets are similar given their compositional resemblance to between 200 and 300 K gives a Curie – Weiss u of 2 31 K,
[1]. Bond distances (Table 2) and angles fall within a range and a magnetic moment of 4.14mB per Co(II) atom. A scan
similar to those observed for other layered cobalt hydrox- conducted at zero field verifies spontaneous magnetization,
ides [15]. The asymmetric unit (Fig. 3) contains four with a sharp onset beginning at 33 K. A hysteresis loop,
independent cobalt atoms, six hydroxides, one water measured at 5 K, is well-defined and characteristic of ferro-
molecule, and 1/2 of an ethanedisulfonate molecule. As or ferrimagnetic ordering (Fig. 5). The coercive field for [1]
for brucite, each hydroxide anion coordinates with three at 5 K may explain the anomaly apparent in the low
cobalt centers. However, half of the six independent temperature region of Fig. 4. Since the applied field is lower

Fig. 4. Inverse magnetic susceptibility of [1] under zero field cooling conditions.
P.M. Forster et al. / Journal of Physics and Chemistry of Solids 65 (2004) 11–16 15

Fig. 5. Hysteresis loop for [1], measured at 5 K.

than the coercive field, the sample does not saturate until Acknowledgements
sufficient temperatures are reached. Saturation values of
slightly more than 0.6mB per Co(II) atom indicate a The authors thank R. Seshadri, A.R. Burbank, G. Lawes,
ferrimagnetic rather than ferromagnetic structure, as has and J. Eckert for valuable discussions. This work was
been observed in pure cobalt hydroxide6 and related supported by the MRL Program of the National Science
systems. Foundation under Award No. DMR00-80034.
It is interesting to note that layered hydroxides similar to
this are typically metamagnetic as long as the spacing
between the layers is less than 10 Å. While the spacing in References
[1] is less (9.64 Å), the coupling between adjacent layers is
apparently ferromagnetic. The availability of high quality [1] C. Mockenhaupt, T. Zeiske, H.D. Lutz, J. Mol. Struct. 443 (1998)
191–196.
crystallographic data may aid in interpreting these [2] P. Rabu, S. Angelov, L. Legoll, M. Belaiche, M. Drillon, Inorg. Chem.
interactions. 32 (1993) 2463–2468.
[3] M. Kurmoo, P. Day, A. Derory, C. Estournès, R. Poinsot, M.J. Stead,
C.J. Kepert, J. Solid State Chem. 145 (1999) 452–459.
[4] R.P. Townsend, E.N. Coker, in: H. van Bekkum, E.M. Flanigen, P.A.
4. Conclusions Jacobs, J.C. Jansen (Eds.), Studies in Surface Science and Catalysis,
vol. 137, Elsevier, Amsterdam, 2001, pp. 467–524.
[5] A.M. Fogg, V.M. Green, H.G. Harvey, D. O’Hare, Adv. Mater. 11
We have demonstrated that a simple direct reaction (1999) 1466– 1469.
between cobalt hydroxide and ethanedisulfonic acid [6] S. Newman, W. Jones, New J. Chem. (1998) 105–115.
[7] M.A. Drezdzon, Inorg. Chem. 27 (1988) 4628–4632.
under hydrothermal conditions yields a highly crystalline
[8] F. Cavani, F. Trifirò, A. Vaccari, Catal. Today 11 (1991) 173–301.
layered hydroxide-organic. The crystal structure contains [9] C. Delmas, Y. Borthomieu, J Solid State Chem. 104 (1993) 345–352.
a novel arrangement of tetrahedral substitutions on a [10] A.I. Khan, D. O’Hare, J. Mater. Chem. 12 (2002) 3191–3198.
brucite-like sheet, with ethanedisulfonate anions packed [11] S. Ghose, Acta Crystallogr. 17 (1964) 1051–1057. R. Allmann,
Z. Kristallogr. 126 (1968) 417–426. P.M. Louër, D. Grandjean, D.
tightly between the layers. The high crystallinity,
Weigel, Acta Crystallogr. B29 (1973) 1703–1706.
ferromagnetic coupling between sheets, and relative [12] W. Stählin, H.R. Oswald, Acta Crystallogr. B26 (1970) 860–863.
structural simplicity of [1] allow us the unique [13] P. Benard, M. Louër, J.P. Auffredic, D. Louër, J. Solid State Chem. 91
opportunity to study the magnetic ordering within the (1991) 296–305.
layers. We are currently preparing deuterated samples for [14] A. Rujiwatra, C.J. Kepert, M.J. Rosseinsky, Chem. Commun. (1999)
2307– 2308.
neutron powder diffraction in the anticipation of gaining [15] A. Rujiwatra, C.J. Kepert, J.B. Claridge, M.J. Rosseinsky, H.
a better understanding of the magnetic structure within Kumagai, M. Kurmoo, J. Am. Chem. Soc. 123 (2001) 10584–10594.
the layers. [16] P.M. Forster, M.M. Tafoya, A.K. Cheetham, to be submitted.
16 P.M. Forster et al. / Journal of Physics and Chemistry of Solids 65 (2004) 11–16

[17] G.M. Sheldrick, SHELXS -97, Program for Crystal Structure Refine- M. Kurmoo, H. Kumagai, M.A. Green, B.W. Lovett, S.J. Blundell,
ment, Universität Göttingen, Germany, 1997. A. Ardavan, J. Singleton, J. Solid State Chem. 159 (2001)
[18] H. Kumagai, C.J. Kepert, M. Kurmoo, Inorg. Chem. 41 (2002) 343 – 351. P.M. Forster, M.M. Tafoya, A.R. Burbank, A.K.
3410–3422. Cheetham, to be submitted.
[19] Z.-L. Huang, M. Drillon, N. Mascoicchi, A. Sironi, J.-T. Zhao, P. [20] V. Laget, C. Hornick, P. Rabu, M. Drillon, R. Ziessel, Coord. Chem.
Rabu, P. Panissod, Chem. Mater. 12 (2000) 2805–2812. M. Rev. 178–180 (1998) 1533–1553.
Kurmoo, H. Kumagai, Mol. Cryst. Liq. Cryst. 376 (2002) 55–565. [21] P.V.D. Sluis, A.L. Spek, Acta Crystallogr. A46 (1990) 194 –201.

Вам также может понравиться