Sie sind auf Seite 1von 13

PHYSICAL REVIEW 8 VOLUME 43, NUMBER 16 1 JUNE 1991

Crystal structure of and low-temperature phase transitions in isostructural


RbSm(SO4)2 4Hzo and NH4Sm(SO4)2. 4H20

S. Jasty
Department of Physics and Molecular Science Program, Southern Illinois University at Carbondale,
Carbondale, Illinois 62901-4401

P. D. Robinson
Department of Geology, Southern Illinois University at Carbondale, Carbondale, Illinois 62901-4401

V. M. Malhotra
Department of Physics and Molecular Science Program, Southern IllinoisUniversity at Carbondale,
Carbondale, Illinois 62901-4401
(Received 6 August 1990; revised manuscript received 26 December 1990)
The crystal structure of RbSm(SO4), .4H20 was determined at 296 K using the x-ray-
crystallographic-diffraction technique. The unit cell is monoclinic with P2l/c space group. The
structure can be envisioned as consisting of the following constituents: a nine-coordinated Sm po-
lyhedron, two crystallographically independent sulfate tetrahedra, an interlayer water, and a caged
Rb cation. Both inequivalent sulfates bridge the nine-coordinated Sm polyhedra, forming
crisscrossing networks of chains along the [100j and [001] directions, which result in rippled layers
in the structure parallel to the (010) plane. It is argued that the monovalent cation Rb is caged by
11 sulfate oxygens and 2 water molecules, and the Rb+ ion rattles in a cavitylike geometry. The
crystal structure of RbSm(SO4)2-4H~O has been compared with the published structure of the
NH4Sm(SO4)2 4H20 lattice, and the two structures are isostructural in the conventional sense.
However, from our proposed hydrogen-bonding scheme for water molecules and from a compara-
tive analysis of the two structures, it is suggested that there are significant differences in the mono-
valent cation coordination in the two lattices. Also, the hydrogen-bond strengths, which are crucial
to the cohesion of the lattice, show significant difFerences in these two isostructural lattices. In-
frared and thermal dehydration probing further support our contention that there are three distinct
types of water in lanthanide double sulfates. From the specific-heat measurements at 100 & T(303
K, it is shown that the RbSm(SO4)2. 4820 lattice undergoes a single A. transition at 232+1 K, unlike
NH4Sm(SO4)2 4H20, which has two structural transitions in the above temperature range. The A,
anomaly reflects logarithmic dependence near the transition temperature, thus suggesting the transi-
tion is order-disorder in origin.

I. INTRODUCTION structural materials NH4L(SO4)2 4H2 O(L =La, Ce, Pr,


Nd, Sm, Eu, Gd, Tb, and Dy), which show a unique trend
Structural phase transitions have been investigated ex- in phase-transition behavior, has been reported. Unlike
tensively in a wide variety of materials, both experimen- other isostructural lanthanide series, the phase-transition
tally and theoretically. ' The major scientific thrust has temperatures in this series show no straightforward
been directed toward understanding the universality of correlation either with the unit-cell parameters or with
behavior at critical regions. Some progress has been the host lanthanide ion's ionic radius.
made in the study of structural phase transitions in three For several ammonium compounds in which phase
dimensions through computer simulations on idealized transitions occur, the ammonium ion has been found to
systems such as hard-sphere models. ' However, the ex- be responsible for driving the transition. It has been
act conditions under which a real material will undergo a observed that the ammonium ion forms a number of com-
structural phase transition still remain an open question. pounds which are isostructural with the corresponding
In order to predict the mechanism that will drive a alkali salts, especially rubidium, since NH4+ and Rb+
structural phase transition, a complete knowledge of the have similar, effective ionic radii. ' '"
Hence, it would be
microscopic force constants is required. In the absence interesting to study the effect of the replacement of the
of such detailed information, substantial headway can be monovalent cation, ammonium by rubidium, on the crys-
made by studying a sequence of isostructural materials, tal structure of NH4L(SO4)2 4HzO (ALSTH) and, conse-
which contain members displaying phase transitions. quently, compare the phase-transition behaviors of the
Perovskites' have been extensively investigated due to ammonium and rubidium compounds. Since the crystal
their simple structure. Recently, another series of iso- structure of ammonium samarium sulfate tetrahydrate

43 13 215 1991 The American Physical Society


13 216 S. JASTY, P. D. ROBINSON, AND V. M. MALHOTRA 43

(ASmSTH) has been reported by Eriksson et al. , ' the MITHRIL, together with an automatic procedure DIR-
corresponding rubidium compound, i.e. , rubidium DIF, for phase extension and refinement of the
samarium sulfate tetrahydrate (RSmSTH), was chosen for difference structure factors. Neutral atom scattering fac-
this study. tors were taken from Cromer and Waber. ' Anomalous
In the present paper, we investigate the crystal struc- dispersion effects were included in the calculation of F
ture of RSmSTH by x-ray diffraction and report the re- (=F, ). ' The values for the scattering factors Df' and
sults of the specific-heat measursement on RSmSTH. Df" required for the treatment of the anomalous disper-
Since the hydrogen-atom positions could not be obtained sion were those of Cromer. ' The full-matrix least-
from the x-ray-diffraction data, a hydrogen-bonding squares refinement was based on F, and the function min-
scheme for the water molecules in RSmSTH and imized was
ASmSTH is proposed based on geometrical considera-
tions derived from crystallographic data. The water's vi-
brational spectra have been used to test the validity of the
proposed hydrogen-bonding model. Thermal data, in
The weights w were taken as 4F, /[o (F, )'], where iF,
terms of the dehydration behavior of RSmSTH, have also l
i

been applied to the same end. It will be demonstrated and iF, are the observed and calculated structure factor
i

that a single A, -type phase transition occurs in RSmSTH, amplitudes, respectively.


quite unlike the two successive phase transitions reported A full-matrix least-squares refinement was performed
for ASmSTH. Further, it will be shown that the with 146 variables, and the final unweighted and weight-
significant differences between the hydrogen-bonding net- ed A factors were calculated from
works in the RSmSTH and ASmSTH lattices indicate
that the waters of hydration play a key role in driving the
structural phase transition in RSmSTH. =0.033,
II. EXPERIMENTAL TECHNIQUES
A. X-ray di8'raction measurements
=0.051 .
Single crystals of RSmSTH were grown by isothermal ta;F, l

evaporation at room temperature from an aqueous solu-


The goodness-of-fit indicator
tion of RbzSO4 and Sm2(SO4)3 8HzO mixed in a 4:1 ratio
under conditions described by Jasty and Malhotra. A
clear prismatic crystal of RbSm(SO~)2 4HzO, having ap-
proximate dimensions 0.24 X 0.21 X 0. 13 mm, was
mounted on a Rigaku diffractometer with graphite mono-
chromated Mo Ke radiation. The cell constants and an was 1.62, where X, was the number of observations and
orientation matrix for data collection were obtained from N, was the number of variables.
a least-squares refinement, using the setting angles of 23
carefully centered reflections in the range C. Hydrogen-atom positions
49. 55' & 20 &49.95 . The intensity data were collected at We attempted to locate the hydrogen atoms from the
296 K, using the co-20 technique at a speed of 4.0'/min in final-difference Fourier map. However, the maximum
co. The weak rejections [I (10. 0o. (I)] were rescanned
and minimum peak densities, corresponding to
(maximum of two rescans) and the counts were accumu- (1.05X10 ) e/nm and ( —1.66X10 ) e/nm, respective-
lated to ensure good counting statistics. Stationary back-
ly, were found within 0. 11 nm of the samarium atom.
ground counts were recorded on each side of the This is not surprising since the residual peaks of electron
reflection. Of the 2143 reflections collected, 1966 were
density, corresponding to hydrogen-atom positions, are
unique. The intensities of three representative reflections, generally obscured' in the presence of heavy atoms like
which were measured after every 150 reflections, Sm and Rb. Consequently, the density profiles could not
remained constant throughout data collection, indicating
be used to determine the hydrogen-atom positions. In-
crystal stability. The data were corrected for Lorentz stead, we adopted the following procedure. Based on the
and polarization effects. The calculations were per- positions of the nonhydrogen atoms and the known
formed using the TEXSAN (Ref. 13) crystallographic
hydrogen-bonding geometries, initial, approximate posi-
software package of Molecular Structure Corporation. tions of the hydrogen atoms were calculated. Then, a
full-matrix least-squares refinement of the 146 variable
B. Solution and refinement of the structure nonhydrogen parameters was performed with hydrogen
atoms included in the atoms' list. Successive recalcula-
The analysis of the RSmSTH crystal structure was tion of hydrogen-atom positions and least-squares
based on 1547 observed unique reflections with intensities refinement was carried out until the maximum shift in the
I & 3.00o(I). Since the space group of the crystal was refinement parameters was negligible.
determined to be centrosymmetric, the structure was To determine the approximate positions of the hydro-
solved by an integrated direct method computer program gen atoms, the following criteria were applied first to
43 CRYSTAL STRUCTURE OF AND LOW-TEMPERATURE PHASE. .. 13 217

decide if a particular O(W)-O(S) or O(W)-O(W') contact The weighted R factor dropped by 0. 1%%uo from the value
[where O(W) and O(W') denote water oxygens and O(S) obtained before the inclusion of hydrogen atoms, yielding
denotes a sulfate oxygen] corresponds to a hydrogen a final R =0.050 and an unweighted R factor equal to
bond. 0.032. The goodness of fit indicator improved by l~o,
(1) Distance criterion. The O(W)-0 distance should be with a final value equal to 1.60. The statistics for both
less than 0.31 nm. This follows from a simplified average models, without and with the inclusion of hydrogen
description of molecular packing in terms of the Van der atoms, i.e., models I and II, respectively, are summarized
baal's contact radii. in Table I. The Hamilton signific'ance test was ap-
(2) Angle criterion. (a) The hydrogen atoms of a water plied to the weighted R factors to determine if the de-
molecule are expected to orient away from the electron- crease in R, due to the inclusion of hydrogen atoms, is
acceptor neighbors X. For the present structure, X could significant. The computed R factor ratio,
be Sm, Rb, or an 0-H group. Therefore, the R„(I)/R (II)=1.020. From Table I, the dimension of
X —O(W) —O(acceptor) angle should be larger than 75 . the test is (178 —146) =32, and the degrees of freedom
This also precludes the possibility of a hydrogen bond are ( 1547 —178 ) = 1369. Interpolation in Hamilton's
along an edge of a coordination polyhedron about a cat- tables for +=0.01 indicates a ratio of about 1.019.
ion. (b) The angle subtended by the two hydrogen-bond Since the computed R factor ratio was slightly greater
acceptors at the oxygen of a single water molecule should than this value, we conclude that the decrease is
not deviate by more than 35' from the tetrahedral angle. significant at 1% level.
Only eight of the O(W)-0 contacts satisfied the above
criteria simultaneously, which corresponds exactly to the
number of hydrogen atoms available. Thus, the position-
al parameters for the hydrogen atoms could be calculated D. Specific-heat measurements
unambiguously by making the following assumptions: (1)
The O(W)-H distance was assumed to be equal to 0.97 The specific-heat data at 100& T &303 K were ob-
A. (2) The H-O(W)-H angle was assumed to be 107.2', tained by the Differential Scanning Calorimetry (DSC)
which corresponds to an average angle of crystalline hy- technique using a Perkin-Elmer DSC7 system. The
drates that has oxygen atoms as acceptors of the hydro- calorimeter was interfaced to a Perkin-Elmer 7700 com-
gen bonds donated by the water molecules. (3) The plane puter through a thermal controller for data collection
of the water molecule was assumed to coincide with the and analysis. Details of the temperature and enthalpy
plane defined by O(W) and the two hydrogen-bond accep- calibration and the experimental run conditions in the
tors since the acceptors tend to be located close to the subambient temperature region have been given else-
water plane. where. By comparing the C values of the A1203 discs,
An isotropic temperature factor, equal to 1.2 times the obtained on our instrument, with those published by Gin-
isotropic temperature factor of the water oxygen O(W), ' nings and Furukawa, the accuracy of the specific-heat
was assigned for each hydrogen atom. The positional capacity at 130 & T & 303 K is estimated to be better than
and thermal parameters of the hydrogen atoms were held 2%.
fixed. A full-matrix least-squares refinement, based on The dehydration behavior of RSmSTH was also deter-
the 146 variable parameters given above, was performed. mined using the Perkin-Elmer DSC7 system. Fine crys-
The positional and thermal parameters of the hydrogen tals of RSmSTH were selectively picked and crimped in
atoms were recalculated and reassigned using the updat- nonvolatile Al pans. Holes were punched in the lids of
ed parameters of the nonhydrogen atoms. A final full- the pans to lower the mechanical barrier to the dehydra-
matrix least-squares refinement converged after one cycle tion process. High-purity N2 was used as the purge gas.
with a largest-parameter shift of 0.003 times its e.s.d. The heating runs were made at a scan rate of 5 K/min.

TABLE I. Crystal data and statistical summary for RbSm(SO4)z. 4H20.


Model I' Model IIb

a =0.6565(2) nm Formula weight: 500.04 No. of reAections 1547 1547


used in refinement
b =1.8913(6) nm Space group: Monoclinic P2, /c No. of parameters 146 178
(fixed+ variable) (146+0) (146+32)
c =0. 8728{1) nm Radiation: MO Ko. , A, =0.07107 nm R 0.033 0.032
P = 96. 26(2)' T =296 K R 0.051 0.050
V=1.0772(5) nm pea)
= 3.083 kg/m' Goodness of fit indicator 1.62 1.61
@=102.7 cm
Transmission coefficient:
0.56—1.00
' Refinement excluding hydrogen atoms.
Refinement including hydrogen atoms.
13 218 S. JASTY, P. D. ROBINSON, AND V. M. MALHOTRA 43

TABLE II. Positional and equivalent isotropic thermal parameters for the non-H atoms and their
e.s.d. 's.
Beq =(81T /3)( U|1+ U2p+ U33+2U(pcosp+2U)3cosP+2U23coscx) .

Atom ~eq (10


Sm(1) 0. 153 06(6) 0. 124 48(3) 0.782 48(5) 1.01(2)
Rb(1) 0.649 6(2) 0.220 89(7) 0. 114 3(1) 3. 10(5)
S{1) 0. 126 0(4) 0.278 6(1) 0.905 5(2) 1.46(9)
S{2) 0.695 6(3) 0.076 1{1) 0.773 5(3) 1.7(1)
O(1) 0.087(1) 0. 157 6(4) 0.314 4(8) 2.4(3)
O(2) 0. 144(1) 0.296 9(4) 0.070 2(8) 2.9(3)
O(3) 0.958(1) 0.273 2(4) 0.372 5(7) 1.7(3)
O(4) 0.314(1) 0.257 4(4) 0.369 9(7) 1.9(3)
O(5) 0.516(1) 0.379 5(4) 0.285(1) 2.7(3)
O(6) 0.639(1) 0.003 0(5) 0.746(1) 4. 1(4)
O(7) 0.806(1) 0.393 8(5) 0. 149 3{8) 2.6(3)
O(8) 0.845(1) 0.417 4(4) 0.413 8(8) 2.3(3)
O(W(1) ) 0. 187(1) 0.443 7(4) 0.039 3(8) 3.0(3)
O(W(2)) 0.259(1) 0. 113 8(4) 0.059 8(8) 2. 5(3)
O(W(3) ) 0.635{1) 0. 105 8(5) 0.344(1) 4. 1(4)
O(W(4) ) 0.807(1) 0.000 4(4) 0. 168 1(8) 2.7(3)

E. FTIR measurements tween 0.2372 and 0.2556 nm, which is about the same
range observed in ASmSTH. ' The average Sm-sulfate
The infrared vibrational spectra of the RSmSTH and oxygen distance (0.247 nm) is approximately the same as
ASmSTH crystals were recorded at 296 K with the help the average distance from Sm to the water oxygen (0.246
of an IBM IR-32 Fourier transform infrared (FTIR) spec- nm). Based on the data presented in Table IV and that
trometer equipped with an IBM 9001 computer. The listed in Ref. 12, it is reasonable to argue that, on replac-
fiuorolube mull technique was employed. The sample ing the NH4+ ion with Rb+, the geometry around the
mull was sandwiched between polished KBr discs. Fifty lanthanide ion remains unaltered.
interferograms acquired at a resolution of 4 cm ' were The framework of the RSmSTH structure consists of
signal averaged to obtain the final spectrum for each sam- chains of Sm coordination polyhedra, linked in a parallel
ple. manner to the c axis by S(1)04 groups and to the a axis by
S(2)04 groups, as shown in Fig. 1. Due to the c glide
III. RESULTS AND DISCUSSION plane being perpendicular to the b axis, the adjacent Sm
polyhedra parallel to the c axis are displaced along the
A. Structure description: RbSm(SO4)2-4HzO [010] direction by about +0. 25b, with alternating signs.
This results in continuous zigzag chains of Sm
The unit-cell parameters for RSmSTH, along with polyhedra-S(1)O~ tetrahedra parallel to the c axis. On the
structural refinement parameters, are reproduced in other hand, adjacent Sm polyhedra along the a axis are a
Table I. The positional parameters of the final structural distance "a" apart so that the continuous chains of Sm
model for the nonhydrogen and hydrogen atoms in polyhedra-S(2)04 tetrahedra, parallel to the a axis are
RSmSTH are listed in Tables II and III, respectively. linear. Thus, the crisscrossing network of chains along
The isotropic Debye-Wailer temperature factors for the the [100] and [001] directions forms rippled layers in the
nonhydrogen atoms B,„are also given in Table II. The
atom numbering system used for this investigation is
identical to that reported by Eriksson et al. ' for
NH~Sm(SO4)z. 4HzO in order to facilitate a comparison TABLE III. Calculated positional parameters for the H atoms.
between the structural data of RSmSTH and ASmSTH. Atom
The RSmSTH structure may be envisioned as consist-
ing of the following components: a Sm polyhedron, two H(11) 0. 1356 0.4019 —0.0168
crystallographically independent sulfate tetrahedra, an H{12) 0.2533 0.4729 —0.0319
interlayer water molecule, and a caged Rb cation. The H(21) 0. 1893 0. 1388 0. 1359
Sm ion is coordinated to nine oxygens of which six belong H(22) 0.3086 0.0696 0. 1063
to the two independent sulfates and three to the waters of H(31) 0.6883 0. 1023 0.4521
hydration W(1), W(2), and W(4). The bond lengths and H(32) 0.5419 0.0665 0.3217
angles for the Sm polyhedron are given in Table IV. In H(41) 0.8259 0.0257 0.0740
this coordination sphere, the Sm-O distances vary be- H(42) 0.7479 0.0334 0.2359
CRYSTAL STRUCTURE OF AND LOW-TEMPERATURE PHASE. . . 13 219

TABLE IV. Coordination around samarium.


Bond distances (nm)
Sm(1) 0(2) 0.2372{7) Sm{1) O(3) 0.2496(6)
Sm(1) 0(5) 0.2383(7) Srn(1) O(W(1)) 0.25 14(7)
Sm(1) 0(W(4)) 0.2411(7) Srn(1) O{4) 0.2551(7)
Sm(1) O(W(2) ) 0.2453(7) Sm(1) 0(8) 0.2556{7)
Sm(1) O(7) 0.2465(7)
Bond angles (deg)
(2) Sm(1) O{5) 88.2{3) O(W(4) ) Sm(l) O(W(1)) 68.3{2)
0(2) Sm(1) 0(W(4)) 138.4(2) 0(W(4) ) Sm(1) O(4) 140.5(2)
0(2) Sm(1) 0(W(2)) 143.4(3) 0(W{4)) Sm{1) 0(8) 72.0(3)
0(2) Sm(1) 0(7) 77. 1(3) 0(W{2)) Sm(1) 0(7) 127.0(2)
0(2) Sm(1) 0(3) 77.8(2) O(W(2)) Sm{1) O(3) 81.6(2)
O(2) Sm(1) 0(W(l)) 70. 1(3) 0(W(2)) Sm(1) 0(W(1)) 138.5(2)
0(2) Sm(1) 0(4) 70. 1(2) O(W(2) ) Sm{1) 0(4) 73.3{2)
0(2) Sm(1) O(8) 126.2(3) 0(W(2) ) Sm(1) O(8) 71.6(2)
0(5) Sm(1) O(W(4) ) 83.0(3) 0(7) Sm(1) 0(3) 77. 1(2)
0(5) Sm(1) 0(W(2) ) 79. 1(3) O(7) Sm(1) O(W(1) ) 72. 1(3)
0(5) Sm(1) O(7) 150.7(3) 0(7) Sm(1) 0(4) 126.4(2)
0(5) Sm(1) 0(3) 124.7(2) 0{7) Sm(1) 0(8) 55.5(2)
0(5) Sm(1) 0(W(1) ) 79.2(3) 0(3) Sm{1) O(W(1)) 39.4(2)
0(5) Sm(1) 0(4) 69. 1(2) 0(3) Sm(1) O(4) 55.7(2)
0(5) Sm(1) 0(8) 145.5(3) 0(3) Sm(1) 0(8) 68.9(2)
0(W(4)) Sm(1) 0(W(2)) 74.3(2) 0(W(1) ) Sm(1) O(4) 129.0(2)
0(W(4) ) Sm(1) O{7) 91.3(3) 0(W(1)) Sm(1) 0(8) 111.5(3)
0(W(4) ) Sm(1) O(3) 138.8(2) 0(4) Sm(1) O(8) 117.3(2)

structure, parallel to the (010) plane. The regions be- "


oxides and halides ranges from 6 to 14. In more com-
tween the layers contain the water molecule W(3) and the plex Rb structures, such as the alums, the Rb ion is at
monovalent cation Rb. As will be discussed later, the least octahedrally coordinated. This is not surprising
layers are held together by hydrogen bonds. The intera- considering the large size of the ion. Even though the ru-
tomic distances and angles of the two crystallographically bidium and ammonium ions have similar effective ra-
independent sulfate groups are given in Table V. As can dii, ' '"a lower coordination number generally has been
be seen from this table, the S(2)04 sulfate is relatively observed for the NH4+ ion in ammonium salts' because
more distorted than the S(1)04 sulfate. of its tendency to form four, more or less, straight
For the monovalent cation Rb, the atoms comprising N— 0
H — hydrogen bonds. For a Rb+ ion, to our
its coordination sphere of radius 0.38 nm are listed in knowledge, the coordination number has never been less
Table VI. Rb has four close contacts ((0.
31 nm) in the than 6. Therefore, we believe, based on the data listed in
RSmSTH structure similar to the coordination reported Table VI, that it has 13 nearest neighbors. This large
for the ammonium ion in ASmSTH. However, the coor- number of nearest neighbors for the monovalent cation in
dination number listed for the rubidium ion in various RSmSTH appears to be the consequence of the rippled

TABLE V. Tetrahedral coordination around sulfur.


S(1)04 S(2)0
Bond distances (nm)
S(1) 0(1) 0. 1453(8) S(2) O(6) 0. 1446(9)
S(1) 0(4) 0. 1470(7) S(2) 0(5) 0. 1459(7)
S(1) 0(2) 0. 1471(7) S(2) Q(7) 0. 1482(7)
S(1) 0{3) 0. 1481(7) S(2) 0(8) 0. 1489(7)
Bond angles (deg)
0(1) S(1) 0(4) 111.7(4) 0(6) S(2) O(5) 111.5(5)
0(1) S(1) 0(2) 109.3{4) 0(6) S(2) 0(7) 112.7(5)
0(1) S(1) 0(3) 111.1(4) O(6) S(2) 0(8) 110.3(5)
0(4) S(1) O(2) 109.6(4) O(5) S(2) O(7) 107.2(5)
0(4) S(1) 0(3) 106.1(4) 0(5) S(2) O(8) 111.1(5)
0(2) S(1) 0(3) 108.9(4) 0(7) S(2) 0(8) 103.8(4)
13 220 S. JASTY, P. D. ROBINSON, AND V. M. MALHOTRA 43

TABLE VI. Bond distances (nm) for Rb and N (recalculated based on atomic positions given in
Table II of Ref. 12 with standard deviations from Table VI of Ref. 12).
Rb(1) O(4) 0.2925(6) N(1) O(4) 0.2890(15)
Rb(1) O(W(3)) 0.2969(9) N(1) O(W(3)) 0.2951(17)
Rb(1) O(3) 0.3025(7) N(1) O(3) 0.2986(16)
Rb(1) O(3) 0.3080(6) N(1) O(3) 0.3004'
Rb(1) O(W(2)) 0.3261(8)
Rb(1) O(4) 0.3375(7)
Rb(1) O(1) 0.3413(7)
Rb(1) O(7) 0.3432(7)
Rb(1) O(8) 0.3471(8)
Rb(1) O(5) 0.3472(8)
Rb(1) O(5) 0.3505(8)
Rb(1) O(2) 0.3599(9)
Rb(1) O(2) 0.3607(9)
'Not given in Ref. 12.

layer framework formed by the samarium-sulfate chains Figure 2 shows the 13 nearest neighbors connected by
in the structure, discussed above. With the interlayer wa- bonds to the caged rubidium ion. A comparison of the
ter oxygen O(W(3)), acting as a cap, a cavity is formed first four nearest-neighbor interatomic distances M-0 be-
within which the Rb ion is caged. tween RSmSTH (M =Rb) and ASmSTH (M =NH4),
given in Table VI, indicates that the corresponding dis-
tances are consistently longer in the rubidium structure
than in the ammonium analog. It is worthwhile to point
out that Khan and Baur' have determined that the
effective ionic radius is larger for the ammonium ion than
for the rubidium ion. Since the N —H . hydrogen
0
bond is expected to be weaker than the Rb-0 covalent
bond, the opposite trend in interatomic M-0 distances d
would be expected, i.e. , d(Rb-0)(d(N-0). In addition,
the Rb-0 distances for the fifth and higher nearest neigh-
bors, given in Table VI, are longer than the sum of the
largest effective ionic radii for rubidium (0. 183 nm) and
O(3)- oxygen (0. 142 nm). " This observation suggests fairly
S{1
0{

p{SI 0{@

FIG. 1. Drawing shows the structure of RbSm(SO&)2. 4H20


at 296 K. The a axis is horizontal, the b axis is vertical, and the
c axis points into the paper. W(1), W(2), and W(4) are the oxy-
gens of water which are coordinated to the trivalent Sm ion.
W(3) represents the oxygen atom of the interlayer water. The
two inequivalent sulfates, forming bridges between the Sm po- FICx. 2. The drawing displays the coordination of the rubidi-
lyhedra, are shown by S(1) and S(2). um ion in RbSm(SO4)2. 4H20.
43 CRYSTAL STRUCTURE OF AND LOW-TEMPERATURE PHASE. .. 13 221

weak bonding between the monovalent cation and its


nearest neighbors. Hence, the model, wherein the mono-
valent cation Rb is envisaged as rattling in a cage bound
by 11 sulfate oxygens and two water oxygens, seems to be
valid. The substantially large thermal factor obtained for
Rb (0.0031 nm ), see Table II, further supports this con-
tention. Due to the odd number of nearest neighbors
(vertices), the cage bounding the rubidium ion does not
appear to have any regular polyhedral geometry.

B. Hydrogen-bonding scheme

The interatomic distances for the hydrogen bond in-


volving water molecules are reproduced in Table VII. As
pointed out earlier, the layers parallel to the (010) plane
are held together primarily by hydrogen bonds. A fur- FIG. 3. The proposed hydrogen-bonding scheme for water
ther illustration of the role played by the hydrogen- molecules in the RbSm(SO4)2. 4H20 structure is presented. The
bonding network in the cohesion of the structure, both horizontal dotted lines in the figure signify various layers. The
across and within the (010) layers, is shown in Fig. 3. Of dashed lines symbolize hydrogen bonds with arrows pointing to-
the eight hydrogen bonds donated by the four waters of ward acceptors. Similar bonding is also valid for isostructural
hydration, the majority, in this case six, are involved in NH4Sm(SO4), .4H20. The bond distances for
linking the layers. This is in agreement with the observed RbSm(SO4)2. 4H20 and NH4Sm(SO&)2 4H20 are reproduced in
peculiarity of the W . 0 hydrogen bonds in crystalline Table VII.
hydrates to link layers corresponding to cleavage sur-
faces. Only two of the eight hydrogen bonds
O(W(1))~O(1) and O(W2))~O(1) provide in-chain
coupling in addition to the bridging S(1)O(4) sulfate. This O(W(4))~O(8), and O(W(4))~O(W(3))] denoted by
contributes to the stability of the zigzag chains parallel to the water molecules W(1), W(2), and W(4) whose oxygens
the c axis. Of the six hydrogen bonds involved in linking are part of the Sm coordination polyhedron. The range
the layers, two belong to the interlayer water molecule of hydrogen-bond lengths donated by the water molecules
W(3) which has its hydrogen-bond acceptors O(7) and is 0.034 nm and spans from 0.273 to 0.307 nm, as given in
O(6) in opposite layers (see Fig. 3). Thus, the interlayer Table VII. This fairly small spread in hydrogen-bond
water molecule W(3) acts as a true bridge across the lay- lengths is typical of hydrated systems in which hydro-
ers. Additional cross-layer linking is provided by the gen bonds play an important role in the cohesion of the
four hydrogen bonds [O(W( 1 ) ) ~O(6), O(W(2) ) ~O(6), structure.

TABLE VII. Effect of cation on the hydrogen-bond geometries in isostructural RbSm(SO4)2 4H20
and NH4Sm(SO4)2 4H20.
RbSm(SO4)2 4H20 NH4Sm( SO4) 2-4HqO'
Donor Acceptor Distances (nm)

O(W(1) ) O(1) 0.277(1) 0.2761(14)


O(W(1) ) O(6) 0.307(1) 0.3049
O(W(2) ) O(1) 0.273(1) 0.2710(14)
O(W(2) ) O(6) 0.282(1) 0.2884(16)
O(W(3)) O(7) 0.278(1) 0.2781(15)
O(W(3)) O(6) 0.279(1) 0.287(16)
O(W(4) ) O(8) 0.274(1) 0.2704(13)
O(W(4)) O(W(3) ) 0.283(1) 0.2777(15)
Angles (deg)

O(1) —0( W( 1)—O(6) 75.9(3) 78


O(1) —O(W(2) —O(6) 80.9(3) 82
O(7) —0( W( 3 ) —0( 6) 117.1(4) 113
O(8) —O(W(4) —O(W(3) ) 96.8(3) 99
' Standard deviations taken from Table VI of Ref. 12.
Data for NH4Sm(SO4)2 4H20 has been calculated based on atomic positions given in Table II of Ref.
12. Hence, the standard deviations are not given.
13 222 S. JASTY, P. D. ROBINSON, AND V. M. MALHOTRA 43

0
In RSmSTH, there are two O(W) — distances longer scheme. However, there are consequential differences in
than, but close to, the upper limit (0.31 nm) of hydrogen- the hydrogen-bond strength in these two isostructural lat-
bond lengths in the scheme discussed above. These dis- tices. The differences are best manifested by comparing
tances, corresponding to [O(W(1) ), O(5)] =0. 312 nm the corresponding distances listed in Table VII, and will
and [O(W(3)), O(1)]=0.316 nm, raise the possibility of be discussed along with the vibrational results in the next
bifurcated hydrogen bonds for W(1) and W(3). Based on section. The hydrogen-bond strength difFerences can
the hydrogen-bonding scheme illustrated in Fig. 3 and have a substantial effect on the thermodynamic stability
Table VII, the O(W) — O(acceptor) bond lengths and the of these two isostructural lattices since hydrogen bonding
O(acceptor, ) —O(W) —O(acceptor2) angles fall within the is crucial to their cohesion.
acceptable limits for hydrogen-bond geometry. Conse-
quently, O(W(1)) —O(5) and O(W(3)) —O(1) really can- C. Vibrational results
not be considered as true hydrogen bonds. Instead, it is
preferable to say that they correspond to very weak con- It is well known that the vibrational spectrum provides
tacts such as assigned in the MgHPO4 3H20 lattice.
~ critical information about the structure of water and
The hydrogen-bond acceptors are all sulfate oxygens water's hydrogen-bonding behavior. We believe the wa-
except for O(W(3)), shown in Fig. 3. Three oxygens of ter, and consequently the hydrogen bonding, play an im-
the S(2)04 group O(6), O(7), and O(8) participate in hy- portant role in triggering the structural phase transitions
drogen bonding, while only one of the S(1)04 oxygens in MLSTH systems. Therefore, we studied the FTIR
O(1) is involved. The coordination around the atoms spectra of RSmSTH and ASmSTH in the frequency range
of 3800—1550 cm ', where water's stretching and bend-
O(6), O(7), and O(8) of S(2)O~ is different. Around O(6)
ing modes are located. Figure 4 reproduces the observed
the coordination is approximately tetragonal, while
vibrational spectra of MSmSTH. In the water's stretch-
around O(7) and O(8) it is approximately trigonal. Con-
ing region for ASrnSTH, five overlapping bands were ob-
sequently, the distribution of S-O bond lengths about the
served at 3527 (shoulder), 3320 (severely overlapped
mean value is expected to be broader for S(2)04 than for
shoulder), 3270 (sharp but weak band), 3063 (sharp shoul-
S(1)04. This is precisely what was observed by compar- der), and 2857 cm . Three distinct vibrations at 1682,
ing the standard deviations in the S-0 bond lengths to the
1636, and 1433 cm ' were observed for ASmSTH in the
mean values for the S(1)04 and S(2)04 groups, as dis-
bending region of water. Based on published results on
cussed earlier. The standard deviation of the S-0 bond ammonium halides, 1433-, 2857-, 3063-, and 3270-cm
length from its mean value rejects the distortion of the
SO4 "tetrahedron.
" We therefore conclude that the
bands for ASmSTH can be assigned to the triply degen-
erate bending mode (v4), first overtone (2v~), symmetric
different role played by the two inequivalent SO4 stretch (v& ), and asymmetric stretch (v3), respectively, of
"tetrahedra" in the hydrogen-bonding network is one the ammonium ion. According to the crystal structure,
reason for the increased linear distortion for the S(2)04 the ammonium ion occupies C& site symmetry. Under
tetrahedron as compared to the S(1)04 tetrahedron. The such a low symmetry, the crystalline-field effects are ex-
other reason is the difference in the nature of the bridging pected to lift the degeneracy of the v4 mode. As can be
performed by the two sulfates. seen from Fig. 4, no three distinct vibrations are observed
The interatomic distances and angles around the water
rnolecules in the isostructural ASmSTH were recalculat-
ed based on the positional and thermal parameters given
in Table II of Ref. 12. A careful scrutiny of these results
revealed that the hydrogen-bonding scheme for ASmSTH
is identical to that given in Fig. 3 for RSmSTH. We have
also listed the computed hydrogen-bond lengths for iso-
structural ASmSTH in Table VII. Based on the data
presented in Table VII and the foresaid discussion, it is
suggested that the isostructural MSmSTH (M =NH4 or
Rb) lattices have three distinct types of water, i.e., W(1), A
distorted geometry; W(2) and W(4); and W(3), layer
geometry.
From the comparative analysis of the crystal structure
of RbSm(SO4)2 4H20 and NH~Sm(SO4)2 4H20, it is ar-
gued that these two systems are isostructural in the con-
ventional sense. However, there are significant
l l I I

differences in the coordination around the monovalent 3500 3000 2500 2000 1 500
cation and the hydrogen-bonding behavior. The am-
WAVE NL) M HER (crn )
monium ion in ASmSTH has four close contacts to which
hydrogen atoms of the ammonium ion are hydrogen
bonded, while the Rb cation in RSmSTH is held in a FICi. 4. The observed infrared spectra of isostructural
cagelike structure. The water molecules in RSrnSTH and MStn(SO4)z. 4HzO [M =NH4 (a) and Rb (b)] crystals show the
ASmSTH participate in an identical hydrogen-bonding similarities and dissimilarities between the two lattices.
43 CRYSTAL STRUCTURE OF AND LOW-TEMPERATURE PHASE. .. 13 223

at around 1433 cm '. We speculated from our specific- RbSrn(SO4) 2 4H20


heat measurements that ammonium ions rotate in the
ALSTH (L =Ce, Pr, Nd, Sm, Eu, Gd, Tb, and Dy) lat-
tice at T &200 K. If such is the case, the rotational I h II

motion of NH&+ ions at 296 K will effectively modulate ih


ii
ih I hii
the crystalline-electric-field effects. This will result in a ihlhi
narrowed, single-vibrational bending mode, as observed
in the present case. Thus, we believe that ammonium NH4Srn(SO4)2 4H20
ions rotate in ASmSTH at 296 K. This observation is
further supported by the absence of the combination
band v4+v6, which appears at around 1800 cm ' if the
ammonium ions do not rotate in the lattice. No such
band was observed for ASmSTH at 296 K.
I

Water molecules in the RSmSTH and ASmSTH lat- 0.250 0.275 0.300 0.325
tices occupy C, site symmetry. According to our HYDROGEN BOND LENGTH (nrn)
hydrogen-bonding scheme, there are three distinct water
molecules in the lattice, i.e. , group I, W(1); group II, W(2)
and W(4); and group III, W(3). The hydrogen-bond dis- FIG. 5. This bar graph shows the distribution of hydrogen-
tances for water, listed in Table VII, predict a complex bond lengths in RbSm(SO4)2 4H20 and NH4Sm(SO4)2 4H20.
vibrational spectrum in the water's stretching and bend-
ing regions due to the range of distribution of distances.
The well-resolved shoulder at 3553 cm ' for RSmSTH suggested earlier, the above-discussed differences could
and 3527 cm ' for ASmSTH occurs at an unusually high significantly alter the thermodynamic behavior of the two
water-stretching frequency for crystalline hydrates. lattices.
This unusually high frequency suggests that the band ob- bending region of water (see Fig. 4), both
served at ) 3500 cm ' originates from a water molecule
which is very weakly hydrogen bonded. ' The water mol-
In the
RSmSTH and ASmSTH show two distinct vibrations at
1636 and 1682 cm '. Eriksson et al. ' for ASmSTH and
ecule (W(1)) participates in a very weak hydrogen bond Malhotra et QI. for ACeSTH also observed a similar
since the bond length O(W(1)) —H(12) . . O(6)=0. 307 doublet in the bending region, and they assigned the
nm lies close to the upper limit for a bond of this type. lower-frequency oscillator to the bending mode of water
The other hydrogen bond for W(1), due to W(1), W(2), and W(4), and the higher-frequency
O(W(1) —H(ll) O(1)=0.277 nm is of intermediate oscillator to the bending mode of W(3). According to our
strength, thus resulting in asymmetric hydrogen bonding hydrogen-bonding scheme, we expected to observe three
for W(1). The asymmetric force field for water (W(1)) distinct bending oscillators. We attempted to curve
should break down the coupling between the symmetric resolve the bending region using the procedure outlined
(v, ) and antisymmetric (v3) stretch oscillations for W(l) in Ref. 31. Only two bands could be fitted to the ob-
in RSmSTH and ASmSTH, producing two separate served spectrum. The hydrogen-bonding effects on wa-
bands. Thus, we assign the shoulders at 3553 (RSmSTH) ter, though, are manifested in the bending region, but
and 3527 cm ' (ASmSTH) to the O(W(1)) —H(12) they are not as sensitive as in the stretching region. Also,
oscillator. The hydrogen-bond distance it has been argued by Falk that the bending frequency is
O(W(l)) —H(12) . O(6) for ASmSTH is smaller than strongly inlluenced by the cation charge (to which water
the corresponding distance for RSmSTH. Therefore, the is coordinated), i.e. , the higher the cation charge, the
observed lower frequency for the O(W(1)) —H(12) oscil- lower the frequency of the bending oscillator. Since the
lator in ASmSTH is consistent with our hydrogen- water molecules W(1), W(2), and W(4) are coordinated to
bonding scheme since, the stronger the hydrogen bond, the trivalent Sm ion and these water molecules are also
the lower the stretching frequency of the 0-H vibra- participating in hydrogen bonding, the observed bending
tion. The other hydrogen-bond distances for various mode frequency will be dictated by the complex combina-
waters in RSmSTH and ASmSTH are centered at around tion of the two foresaid effects.
0.278 nm (see Table VII). This fact should result in oscil-
lators other than O(W(1) —H(12) producing a broad, sin- D. Thermal behavior at 303 & T & 473 K
gle overlapping band at around 3320 cm '. Figure 4
rejects this. However, there is striking difference in Although the infrared results, presented above, are not
these intermediate-strength hydrogen bonds between inconsistent with our proposed hydrogen-bonding
RSmSTH and ASmSTH as revealed by the bar graph in scheme for MSmSTH, they do not provide conclusive
Fig. 5. In RSmSTH, the bond distances are closely evidence in support of the scheme. Consequently, we un-
clustered about the mean (0.278 nm) with a standard de- dertook thermal (DSC) measurements on RSmSTH at
viation of +0.004 nm, while, in ASmSTH, the distribu- 303 & T &473 K to see if three distinct activation ener-
tion about the mean is broader with a deviation of gies could be observed during the process of dehydration.
+0.007 nm. This implies that waters' hydrogen bonding Figure 5 shows the observed dehydration behavior of
plays a more dominant role in the cohesion of the RSmSTH at 303 & T &473 K. The measured weight loss
RSmSTH structure than of the ASmSTH structure. As at the end of the DSC run was found to be equal (within
13 224 S. JASTY, P. D. ROBINSON, AND V. M. MALHOTRA 43

the limits of experimental error) to the calculated weight


fraction for the four waters of hydration per formula unit
in the RSmSTH sample before the run. This indicates
that dehydration was completed on heating to the fairly
low temperature of 473 K, not 1323 K as reported by 0.8
Erametsa and Niinisto. Figure 5 displays two major en-
dothermic peaks in the DSC curve, a complex asym-
metric, overlapped peak at 407 K and a low enthalpy ~ 0. 6
peak at 432 K. It should be noted that, after dehydration
has occurred, the water molecules released from the lat- CL

tice can still be trapped in the pores of the anhydrous 0.4—


powder crimped in the perforated DSC sample pan. Ad-
I

ditional energy must be supplied for the water molecules


I I I

150 200 250 300


to overcome the mechanical barrier due to the product TEMPERATURE (K)
layer ' in order to be completely removed from the sys-
tem. A distribution of barrier heights as a function of
particle size, packing density, and other geometrical pa- FIG. 7. This graph reproduces the observed specific heat for
rameters is expected. It was suspected that the broad, isostructural (a) NH4Sm(SO4)~. 4H20 and (b) RbSm(SO4) 2 4H20
low enthalpy peak at 432 K was due to the release of such crystals at 100 & T & 305 K.
mechanically trapped water molecules rather than anoth-
er stage of the dehydration reaction. To ascertain the
origin of the endothermic peak at 432 K, the DSC run after heating to 420 K, confirming our suspicion that the
was terminated at 420 K and the FTIR spectrum of the
endothermic peak at 432 K in Fig. 6 is due to the release
powder retrieved from the DSC pan was recorded. A sin- of mechanically trapped water. Hence, from Fig. 6, it
' was observed in the
gle, low-intensity band at 1626 cm can be seen that the dehydration of RSmSTH occurs over
H-0-H bending region of the ir spectrum instead of the the temperature range 355& T &420 K, although com-
doublet shown in Fig. 4. This band can be assigned to
plete removal of the water molecules from the system re-
weakly hydrogen-bonded surface water. ' The FTIR
quires higher temperatures. A similar, two-step mecha-
spectrum of the RSmSTH sample, after being heated to nism for the dehydroxylation of clay minerals was sug-
473 K, was also recorded. An extremely weak band at
gested by Malhotra and Ogloza.
exactly the same frequency 1626 cm ' was again ob- For DSC data, the peak in the dehydration curve
served in the H-0-H bending region. Therefore, it is ar- occurs at a temperature at which there is maximum wa-
gued that the type of water present in the sample after ter loss. The peak temperature can be used as a reference
heating to 420 K is identical to the traces of the type of temperature to calculate the activation energy for the
water present in the anhydrous sample. This implies that dehydration of the water. As seen from Fig. 6, three
there are no waters of hydration present in the sample peaks were observed in the dehydration
overlapping
curve for RSmSTH at approximately 400, 407, and 410
K. The presence of three peaks indicates that there are
three types of water with three different activation energy
barriers. This observation is consistent with the
classification of the four waters of hydration into three
groups, based on the crystallographic data and the pro-
posed hydrogen-bonding scheme. The dehydration
DJ 4— enthalpy required for the liberation of the four waters of
C)
CL hydration from the RSmSTH structure was determined
to be 99 kJ/mol. This value compares favorably with the
range of activation energies 46—170 kJ/mol, reported for
QJ other crystalline hydrates.

C3 E. Phase transitions and speci6c heat


2—
Jasty and Malhotra, from their specific-heat measure-
350 400 450 ments on ASmSTH, showed that this system undergoes
TEMPERATURE (K) two structural phase transitions at 110& T &300 K. In
addition, they argued from the observed specific heat of
FIG. 6. Differential Scanning Calorimetry (DSC) thermo- ASmSTH that NH&+ ions initiate rotation at around 200
gram depicts the dehydration and desorption of water from K in this lattice. In view of the overall similarity of the
RbSm(SO&)&-4H&O. The broad peak in the curve at 432 K crystal structure of ASmSTH and RSmSTH lattices, but
represents the desorption of mechanically held water, while the yet having discernable differences in the hydrogen-
three overlapped peaks are associated with the dehydration pro- bonding strengths and in the coordination of monovalent
cess. cations, it was felt it would be interesting to determine
43 CRYSTAL STRUCTURE OF AND LOW-TEMPERATURE PHASE. . . 13 225

how the replacement of the NH4+ ions by the Rb ions between the nearest paramagnetic Sm + ions are larger
affects the phase-transition behavior in the MSmSTH lat- than 0.6 nm (calculated from Table II). If it is argued
tice. Consequently, the specific heat (C ) of RSmSTH that the A, shape implies the onset of molecular rotation
was recorded at 100& T &310 K and is reproduced in in the RSmSTH lattice, then the possible candidates are
Fig. 7. Also depicted in Fig. 7, for the sake of compar- sulfate ions and/or water molecules. The Rb+ ions are
ison, is the observed specific heat of ASmSTH. Unlike held in a distorted cavitylike structure in the lattice and,
the two anomalies observed for ASmSTH, a single anom- therefore, will rattle rather than rotate. Since both
aly was observed in the specific-heat curve of RSmSTH at S(1)Oz and S(2)O& ions in the RSmSTH lattice are in-
100 & T & 303 K. The shape of the anomaly with its long, volved in bridging the samarium polyhedra, it is difficult
low-temperature tail indicates that the transition occur- to see how the lattice will remain stable if the sulfate ions
ring in the RSmSTH lattice is partly gradual, but it is rotate. However, it is possible that the water molecule
completed isothermally at 232 K. This results in a k- dipoles initiate Auctuations in the RSmSTH lattice at 210
shaped anomaly in the specific heat which has been asso- K& T & To, similar to those reported for alums. The
ciated with order-disorder transition, magnetic transi- specific-heat measurements alone cannot distinguish be-
tion, onset of molecular rotation, and/or ordering of axis tween an order-disorder transition and the transition in-
orientation of molecular rotation. duced as a consequence of fluctuations of water dipoles.
The total entropy (AS) for the observed second-order For order-disorder transition, a logarithmic diver-
transition for RSmSTH was determined from gence ' ' in C is expected near the transition tem-
perature, i.e. ,
C = 2+8 In(e), (6)
(3)

To determine AS from Eq. (3), a graph representing where e= (T/To) —l~. Using the procedure outlined in
C /T for each T at 180 & T & 250 K was generated, and Refs. 39 and 43, we attempted to fit our specific-heat data
near the transition temperature, i.e. , 215 K & T & To and
the total entropy was simply calculated by determining
the area under the curve between the given limits. The )
240 K) T To. The logarithmic divergence of specific
total entropy AS associated with the second-order transi- heat for RSmSTH at 215 K & T & To can be described by
tion for RSmSTH was determined to be 0.0096 J/g K. It
T —1
is worth mentioning here that the observed anomaly for C = 0. 550 —0. 403 ln (J/g K),
RSmSTH is not as sharp or energetically intense as typi- 232
cal A, transitions observed ' in NH&cl or (NH&)2SO&. 0. 006 & e'& 0. 095, (7)
However, the transition in RSmSTH is almost ten times
more intense than the two transitions observed in and at 240 K & T & T, as
ASmSTH.
T —1
As can be seen from Fig. 7, the specific heat of
Cp
= 0. 535 —0. 035 ln (J/g K),
RSmSTH above the transition has a temperature depen-
dence which is different from the one observed below the
transition temperature ( To =232+1 K). The observed C 0. 003 & e &0.033 . (8)
(in J/g K) at T & 215 K & To can be described by The regression coe%cients for Eqs. (7) and (8) were 0.994
and 0.959, respectively. Owens, from his EPR measure-
C~ =(0.221+8. 664X10 T+1.QZQX1Q
~T2
ments on KCr(SO~)z 12H20 alum, showed that the
—2. 357X10 T ), 110& T &210 K, (4) linewidth of Cr + ions diverges as the transition tempera-
while the C~ (in )
J/g K) at T 245 K) To is represented
ture of 157 K is approached. It is interesting to note the
divergence in linewidth can be explained by a logarithmic
by function. If a similar fluctuation of some polar lattice
C =(0.267+1.618X10 T), 245&T &303 K. mode is in effect in RSmSTH near 215 K, and if Aucta-
tions are slow enough with respect to the specific-heat-
On extrapolating Eq. (4) to 245 K, the computed C„does measurement time, then a logarithmic divergence in
not join Eq. (5) but lies 0.036 J/g K above it. Similar be- specific heat is expected.
havior in C for ferroelectric gadolinium molybdate has Usually in isostructural lattices, the replacement of a
been reported by Cheung and Ullman. They suggested monoatomic cation such as Rb+ by a polyatomic cation
this behavior may be due to differences in anharmonicity such as NH4+ has been known to induce a phase transi-
between the T & To and T & To phases. tion. For example, RbzSO4 does not undergo any
As described earlier, the observed A, transition for structural phase transitions at T &290 K. On the other
RSmSTH indicates that the second-order transition is ei- hand, it is well known that (NH&)2SO& exhibits a fer-
ther an order-disorder effect or is the onset of molecular roelectric transition at 223 K. It has been argued that
rotation in the lattice. We exclude the possibility that the the ammonium ion is critically involved in triggering the
transition in RSmSTH is a magnetic transition, especially transition. If one extends the foresaid analogy to the
because the observed phase transitions in isostructural double sulfates, i.e. , RSmSTH and ASmSTH, then it may
ASmSTH are structural in nature. Also, the distances appear that the 273-K transition in ASmSTH is triggered
13 226 S. JASTY, P. D. ROBINSON, AND V. M. MALHOTRA 43

by ammonium ions. This especially seems plausible be- greater than that of the S(1)04 tetrahedron because the
cause NH4+ ions in ASmSTH initiate rotation at T 200 ) former acts as a strain absorber in the linear samarium-
K. Though the hydrogen-bonding networks in RSmSTH sulfate chains parallel to the a axis. (4) Unlike the am-
and ASmSTH are identical, the strength of the dynamic monium ion which can be hydrogen bonded to four oxy-
coupling between the coordinated water molecules and gens, the large size of the rubidium ion precludes the pos-
the interlayer water molecules differs appreciably be- sibility of only four nearest-neighbor oxygen atoms for
tween the two lattices. On the other hand, the role of the the Rb ion. Hence, for RSmSTH, the structural data in-
sulfate ions in the two structures is identical. Also, the dicate that the rubidium ion is located in a cavity bound-
sulfate ions do not exhibit any disorder similar to the ed by 11 sulfate oxygens and 2 water oxygens. (5) The
kind observed in some of the alums. Since the monoa- hydrogen-bonding scheme proposed on the basis of crys-
tomic rubidium ion is not expected to cause a phase tran- tallographic data, which suggests three distinct types of
sition, by the process of elimination it is possible that the water molecules in lanthanide double sulfates, was criti-
water molecules through hydrogen bonding hold the key cally evaluated and found to be consistent with our spec-
to the mechanism that drives the structural phase transi- troscopic and thermal data. (6) From the specific-heat
tion in RSmSTH. Temperature-dependent x-ray measurements at 100(T (303 K, the structural phase-
diffraction and/or neutron scattering and/or solid-state transition behaviors in RbSm(SO4)2. 4HzO and
NMR measurements are required to shed more light on NH~Sm(SO4)2 4HzO were found to be different.
the transition mechanism. RbSm(SO4)z 4H20 undergoes a single, A, -type transition
at 232 K, unlike the successive structural transitions ob-
served in NH4L(SO4)2 4H20 at 160 and 273 K. The I,—
IV. SUMMARY AND CONCLUSIONS shape anomaly can be fitted by a logarithmic function
near the transition temperature. This argues that the
Based on a comparative analysis of the room- transition in RbSm(SO„)2 4H20 is an order-disorder type.
temperature crystal structures, infrared spectra, and
Based on the significant differences between the
temperature-dependent (100(T (303 K) specific heats hydrogen-bonding networks in the two isostructural
of RbL(SO4)2 4H20 and NH4L(SO4)2 4H20 lattices, we
MSm(SO4)2 4H20 lattices, it is possible that the water
conclude the following: (1) RbSm(SO4)2 4H20 and molecules are critically involved in driving the structural
NH4Sm(SO4)2 4HzO are isostructural at room tempera- phase transition in RbL(SO~)2 4H20.
ture. (2) The coordination geometry around the samari-
um ion remains unaffected by a substitution of the mono-
valent cation, ammonium by rubidium. (3) In the V. SUPPLEMENTARY MATERIAL
RbSm(SO4)z 4H20 as well as NH&Sm(SO4)2 4H20 struc-
tures, the two crystallographically inequivalent sulfates A listing of the anisotropic temperature factors and
are distorted to different extents from the tetrahedral structure factor amplitudes for RbSm(SO4)z 4H20 are
symmetry. The distortion of the S(2)O~ tetrahedron is available from the authors.

~J. F. Scott, Rev. Mod. Phys. 46, 83 (1974).


S. Jasty, M. Al-Naghy, and M. de Llano, Phys. Rev. A 35, 1376
D. T. Cromer and J. T. Waber, International
Crystallography (Kynoch, Birmingham,
Tables for XRay-
England, 1974), Vol.
(1987). IV, Table 2.2A.
J. G. Berryman, Phys. Rev. A 27, 1053 (1983). J. A. Ibers and W. C. Hamilton, Acta Crystallogr. 17, 781
4M. Rousseau, J. Y. Gesland, J. Julliard, J. Nouet, J. Zarem- (1964).
bowitch, and A. Zarembowitch, Phys. Rev. B 12, 1279 (1975). 'sD. T. Cromer, International Tables for X Ray Crystallograp-hy
5S. Jasty and V. M. Malhotra (unpublished). (Kynoch, Birmingham, England, 1974), Vol. IV, Table 2.3. 1.
6N. G. Parsonage and L. A. K. Staveley, Disorder in Crystals G. H. Stout and L. H. Jensen, X-Ray Structure Determination
(Clarendon, Oxford, 1978). (Macmillan, New York, 1972).
7D. R. Taylor, Phys. Rev. B 40, 493 (1989). M. Falk and O. Knop, in Water, A Comprehensive Treatise,
SN. E. Schumaker and C. W. Garland, J. Chem. Phys. 53, 392 edited by F. Pranks (Plenum, New York, 1973), Vol. 2, Chap.
(1970). 2, p. 55.
E. O. Schlemper and W. C. Hamilton, J. Chem. Phys. 44, 4498 ~ H. D. Lutz, Struct. Bonding 69, 97 (1988).
(1966). W. H. Baur and A. A. Khan, Acta Crystallogr. B 26, 1584
A. A. Khan and W. H. Baur, Acta Crystallogr. B 28, 683 (1970).
(1972). G. C. Chiari and G. Ferraris, Acta Crystallogr. B 38, 2331
R. D. Shannon, Acta Crystallogr. A 32, 751 (1976) ~ (1982).
~B. Eriksson, L. O. Larsson, L. Niinisto, and U. Skoglund, ~4W. C. Hamilton, Acta Crystallogr. 18, 502 (1965).
Inorg. Chem. 13, 290 (1974). 25W. C. Hamilton, Statistics in Physical Science: Estimation,
TEXSAN-TEXRAY Structure Analysis Package, Molecular Hypothesis Testing, and Least Squares (Ronald, New York,
Structure Corporation (1985). 1964).
' C. J. Gilmore, J. Appl. Cryst. 17, 42 (1984). D. C. Ginnings and G. T. Furukawa, J. Amer. Chem. Soc. 75,
P. T. Beurskens, Crystallography Laboratory, Toernooiveld 522 (1953).
Report No. 1984/1 (unpublished). ~7A. C. Larson and D. T. Cromer, Acta Crystallogr. 22, 793
43 CRYSTAL STRUCTURE OF AND LOW-TEMPERATURE PHASE. . . 13 227

{1967). sevier, Amsterdam, 1980), Vol. 22, p. 115.


2~G. Ferraris and M. Franchini-Angela, Acta Crystallogr. B 28, V. M. Malhotra and A. A. Ogloza, Phys. Chem. Miner. 16,
3572 (1972). 386 (1989).
H. A. Buckmaster, V. M. Malhotra, and H. D. Hist, Can. J. L. L. Sparks, in Materials at I.om Temperatures, edited by R.
Phys. 59, 596 (1981). P. Reed and A. F. Clark (American Society for Metals, Met-
G. Brink and M. Falk, Can. J. Chem. 48, 2096 (1970). als Park, Ohio, 1983).
V. M. Malhotra, S. Jasty, and R. Mu, Appl. Spectrosc. 43, 638 K. M. Cheung and F. G. Ullman, Phys. Rev. B 10, 4760
(1989). (1974)~

3~V. M. Malhotra, H. A. Buckmaster, and H. D. Bist, Can. J. 4"F. J. Owens, in Magnetic Resonance of Phase Transitions, edit
Phys. 58, 1667 (1980). ed by F. J. Owens, C. P. Poole, and H. A. Farach (Academic,
J. Schiffer and D. F. Hornig, J. Chem. Phys. 49, 4150 (1968). New York, 1979).
~M. Falk, Spectrochim. Acta 40A, 43 (1984). " E. Pytte and H. Thomas, Phys. Rev. 175, 610 (1968).
50. Erametsa and L. Niniisto, Suom. Kemistil. B 44, 207 42I. Hatta and A. Ikushima, J. Phys. Chem. Solids 34, 57 (1973).
(1971). 43H. J. Fecht, Z. Fu, and W. L. Johnson, Phys. Rev. Lett. 65„
M. E. Brown, D. D. Dollimore, and A. K. Galway, in Chemi- 1753 (1990).
cal Kinetics, edited by C. H. Bamford and C. F. H. Tipper (El-

Das könnte Ihnen auch gefallen