ArticlePDF Available

Single-Crystalline cooperite (PtS): Crystal-Chemical characterization, ESR spectroscopy, and 195Pt NMR spectroscopy

Authors:

Abstract and Figures

Single-crystalline cooperite (PtS) with a nearly stoichiometric composition was characterized in detail by X-ray diffraction, electron-probe X-ray microanalysis, and high-resolution scanning electron microscopy. For the first time it was demonstrated that 195Pt static and MAS NMR spectroscopy can be used for studying natural platinum minerals. The 195Pt chemical-shift tensor of cooperite was found to be consistent with the axial symmetry and is characterized by the following principal values: δ xx = −5920 ppm, δ yy = −3734 ppm, δ zz = +4023 ppm, and δiso = −1850 ppm. According to the ESR data, the samples of cooperite contain copper(II), which is adsorbed on the surface during the layer-by-layer crystal growth and is not involved in the crystal lattice.
Content may be subject to copyright.
ISSN 1063-7745, Crystallography Reports, 2008, Vol. 53, No. 3, pp. 391–397. © Pleiades Publishing, Inc., 2008.
Original Russian Text © V.I. Rozhdestvina, A.V. Ivanov, M.A. Zaremba, O.N. Antsutkin, W. Forsling, 2008, published in Kristallografiya, 2008, Vol. 53, No. 3, pp. 423–430.
391
INTRODUCTION
For platinum as a mineral-forming element, 64 min-
erals were found. Eleven platinum minerals do not have
their own names [1]. Eighteen minerals belong to sul-
fides, in which Pt is associated with Pd, Ir, Rh, Cu, Ni,
Fe, Co, and Pb. Of the most abundant minerals, two sul-
fide minerals, cooperite (PtS) and braggite ((Pt,Pd)S,
belong to the same isomorphous series, in which the
platinum content gradually decreases, the palladium
content gradually increases, and a nickel impurity is
often present [2]. However, the discontinuity in the
metal content is indicative of the discreteness of this
series. Cooperite, which is one of the main sources of
platinum in copper–nickel sulfide ores, generally crys-
tallizes in the following two morphological modifica-
tions: irregular grains and, more rarely, micron- and
submicron-size prismatic crystals in association with
iron varieties of platinum and osmium–iridium miner-
als. Single-crystalline mineral grains of cooperite are
very rare in occurrence. Such crystals were found in
deposits in the Far-East.
In the present study, single-crystalline samples of
cooperites, which were found in placers from the Yuna-
Dan’sk gold-ore cluster of the platinum-bearing prov-
ince in the Sea of Okhotsk region of the Ma
œ
makansk
zone in the Far East, were studied by electron micros-
copy, electron-probe X-ray microanalysis, X-ray dif-
fraction, and magnetic resonance spectroscopy (ESR
and
195
Pt static and MAS NMR spectroscopy). This
gold-ore cluster is characterized by the presence of
large amounts of idiomorphic single-crystalline min-
eral grains of cooperite in association with ferrous vari-
eties of platinum, osmium–iridium minerals, and palla-
dium stibnites and arsenides. The paragenesis of coo-
perite single crystals and high-temperature platinum-
group minerals could be formed under high-tempera-
ture conditions and, consequently, under conditions
where sulfur exhibits high chemical activity. Xenomor-
phic inclusions of cooperite are prone to be located at
the periphery of mineral grains, thus filling microcavi-
ties and microcracks, which is evidence that cooperite
is a secondary mineral. This is confirmed by the data
[3, 4] on the chemical modeling of the formation of
cooperite even under rather mild conditions. In contrast
to other platinum-bearing ore clusters in the Far East
(containing predominantly minerals, which are inter-
metallic ordered and disordered varieties of platinum–
ferrous and osmium–iridium–ruthenium solid solutions
[5], and sometimes platinum arsenides, such as sper-
rylite (PtAs
2
) [6]), this ore cluster is characterized by
the pronounced platinum sulfide and stibiopalladinite
mineral reduction. The latter is indicative of specific
(even within the platinum-bearing province in the Sea
of Okhotsk region) local conditions of the genesis of
platinum-group minerals for the ore cluster under con-
sideration.
Single-Crystalline Cooperite (PtS): Crystal-Chemical
Characterization, ESR Spectroscopy,
and
195
Pt NMR Spectroscopy
V. I. Rozhdestvina
a
, A. V. Ivanov
a
, M. A. Zaremba
a
,
O. N. Antsutkin
b
, and W. Forsling
b
a
Institute of Geology and Nature Management, ul. B. Khmel’nitskogo 2,
Far East Division, Russian Academy of Sciences, Blagoveshchensk, 675000 Russia
e-mail: veronika@ascnet.ru
b
Luléa University of Technology, SE-971 87 Luléa, Sweden
Received December 11, 2006; in final form, December 3, 2007
Abstract
—Single-crystalline cooperite (PtS) with a nearly stoichiometric composition was characterized in
detail by X-ray diffraction, electron-probe X-ray microanalysis, and high-resolution scanning electron micros-
copy. For the first time it was demonstrated that
195
Pt static and MAS NMR spectroscopy can be used for study-
ing natural platinum minerals. The
195
Pt chemical-shift tensor of cooperite was found to be consistent with the
axial symmetry and is characterized by the following principal values:
δ
xx
= –5920 ppm,
δ
yy
= –3734 ppm,
δ
zz
=
+4023 ppm, and
δ
iso
= –1850 ppm. According to the ESR data, the samples of cooperite contain copper(II),
which is adsorbed on the surface during the layer-by-layer crystal growth and is not involved in the crystal lat-
tice.
PACS numbers:
61.50.
Nw
, 61.66.-
f
DOI:
10.1134/S106377450803005X
DIFFRACTION AND SCATTERING
OF IONIZING RADIATIONS
392
CRYSTALLOGRAPHY REPORTS
Vol. 53
No. 3
2008
ROZHDESTVINA et al.
EXPERIMENTAL
The morphological and microstructural features of
cooperite samples, the phase microinhomogeneity, and
the chemical composition were studied by electron
microscopy and electron-probe X-ray microanalysis.
The qualitative determinations of elements based on the
phases were performed by electron-probe X-ray
microanalysis on a RONTEC energy-dispersive spec-
trometer equipped with a LEO-1420 scanning electron
microscope. The quantitative determinations were car-
ried out by the ZAF-correction method on a 35-SDS
wavelength-dispersive spectrometer equipped with a
JSM-35C JEOL scanning electron microscope with the
use of chemically pure platinum and standardized sul-
fides (PtS and PbS) as reference samples. The morphol-
ogy, microstructures, and the phase microinhomogene-
ity were studied using different modes for the detection
of secondary and reflected electrons. The accelerating
voltage and the current passing through the samples
were chosen according to the problem at hand.
The structures of natural samples of platinum sul-
fide were studied by X-ray diffraction on an URS-2.0
instrument (Ni-filtered Cu-
K
α
radiation) by rotating a
single crystal about the principal axes. The layer lines
in the X-ray diffraction patterns were processed with
the use of a computer program based on the extraction
of the data on the optical density of X-ray films fol-
lowed by their storage as the full-profile two-coordinate
representation
I
(2
θ
)
. The algorithm works in the high-
precision range of the determination of the crystal-lat-
tice parameters with separated diffraction maxima cor-
responding to the
K
α
1
and
K
α
2
spectral lines. The unit-
cell parameters were calculated by the least-squares
method using the autoindexing program. The X-ray
powder diffraction patterns were measured on a
DRON-3 diffractometer (Ni-filtered Cu-
K
α
radiation)
from a rotating sample using the step-scan technique
with a
τ
-scan step
τ
= 0.01° 2
θ
. The exposure time per
step was 10 s. The parameters were determined and the
overlapping peaks were separated by the approxima-
tion method using graphical editors and analyzers for
X-ray diffraction patterns. The autoindexing and refine-
ment of the parameters were carried out with the use of
the PDWin program package (NPP Burevestnik). The
theoretical X-ray diffraction pattern was calculated
with the use of the POWDER CELL program [7].
195
Pt static and MAS NMR spectra
were recorded on
a CMX-360 pulse spectrometer (Varian/Chemagnetics
InfinityPlus, United States) in the operating-frequency
range of 76.909–77.709 MHz equipped with a superon-
ducting magnet (
B
0
= 8.46 T) and with the use of the
Fourier-transform technique. The magic-angle-spin-
ning (MAS) spectra were recorded using the excitation
pulse of
67.5°
with a delay time of 1.5
µ
s, which corre-
sponds to the excitation bandwidth of 125 kHz. The
total width of the experimental
195
Pt MAS NMR spec-
trum (on the frequency scale) was ca. 0.8 MHz. Since it
was impossible to achieve the simultaneous excitation
of the total NMR spectrum, the spectrum was recorded
in fragments with a frequency step of 100 kHz followed
by the summation of individual fragments. The
195
Pt
static and MAS NMR spectra were recorded from coo-
perite samples with a weight of ca. 310 mg, which were
placed in zirconium dioxide rotors with a diameter of
4.0 mm (in the former case, because of high structural
ordering, the samples were thoroughly ground; in the
latter case, natural single-crystalline mineral grains
were used). The samples were spun at the magic angle
at a spinning frequency of 8000(1) Hz. For each region
of the MAS spectrum, the number of acquisitions was
300. The time delay between the excitation pulses was
4.0 s. The static spectra were obtained in fragments
with a step of 100 kHz using a spin-echo pulse sequence
(
90°
τ
80°
τ
) with a pulse length of 2.0 and 4.8
µ
s,
respectively, and a delay time
τ
= 160
µ
s. For each
region of the static spectrum, the number of acquisi-
tions was 4512, and the time delay between the excita-
tion pulses was 4 s. Since the beginning of the interfer-
ence patterns (free-induction decay, FID) for the MAS
spectra was distorted by the attenuated excitation pulse
(due to the ring-down), the distorted region was sub-
tracted. Then, to avoid phase distortions of the MAS
NMR spectra, the interference pattern was shifted
toward the attenuation region to the required number of
points before performing the Fourier transform. Since
the interference patterns included 15 spin-echo pulses,
the subtraction of the initial fragment did not lead to
critical distortions of the spectra. The
195
Pt isotropic
chemical shift and the principal values of the
195
Pt
chemical-shift tensor are given with respect to a 0.1 M
aqueous H
2
[
PtCl
6
]
solution (Merck), 0 ppm [8], which
corresponds to the resonance frequency (77.3778 MHz)
of
195
Pt nuclei. The homogeneity of the magnetic field
was monitored based on the width of the reference line
of crystalline adamantane at
δ
(
13
C
)
= 38.56 ppm
(2.4 Hz). The parameters of the
195
Pt NMR spectrum
were corrected for the drift of the magnetic-field
strength in the course of experiments, whose frequency
equivalent for
195
Pt nuclei was 0.044 Hz/h. The magic
angle was set at the resonance frequency of the
79
Br
nuclide (90.189 MHz) according to a standard proce-
dure with the use of crystalline KBr. The additional
adjustment was performed to achieve the minimum lin-
ewidths in the
195
Pt MAS NMR spectrum of cooperite.
ESR spectra
were recorded on a 70-02 XD/1
radiospectrometer (MP CZ, Minsk) operating at
ca. 9.5 GHz at
~295
K. The operating frequency was
measured with a ChZ-46 microwave frequency meter.
The
g
factors were calculated relative to diphenylpic-
rylhydrazyl (DPPH). The errors of the determination of
the
g
factors and the hyperfine-structure constants
(given in Oe) are
±
0.002
and
±
2%
, respectively. To
reveal fine details, the double differentiation of the
experimental spectra was carried out. The ESR spectra
were simulated within the second-order perturbation
theory with the WIN-EPR SimFonia program (version
CRYSTALLOGRAPHY REPORTS
Vol. 53
No. 3
2008
SINGLE-CRYSTALLINE COOPERITE (PtS): CRYSTAL-CHEMICAL CHARACTERIZATION 393
1.2, implemented in the Bruker software). In the course
of fitting of the model spectra to the experimental spec-
tra, the
g
factors, the hyperfine- structure constants, the
resonance linewidths, and the Lorentz and Gaussian
contributions to the line shape were varied.
RESULTS AND DISCUSSION
Electron-Microscopic and Electron-Probe X-ray
Microanalysis Data
Cooperite from the Yuna-Dan’sk gold-ore cluster
exists primarily as idiomorphic black single-crystalline
mineral grains with metallic lustre and conchoidal frac-
ture. The grains are crystal-faceted, their faces are
hatched, and the grains are often characterized by the
rhombic geometry (Fig. 1a). Single crystals are often
chipped. Individual grains with smoothed edges and
traces of dissolution were found (Figs. 1b and 1c). The
average size of single-crystalline mineral grains is 200–
300
µ
m. Individual grains with sizes larger than 1 mm
were found (Fig. 1d). This variety of cooperite corre-
sponds in chemical composition to platinum sulfide
(Fig. 2) having the following stoichiometric composi-
tion: Pt, 85.826 wt %; S, 13.917 wt %. Insignificant iron
impurities were found in some samples. The amounts
of Pd, Ni, and some other elements are outside the sen-
sitivity limits of electron-probe X-ray microanalysis
(a wavelength-dispersive spectrometer). In addition, it
should be noted that no inclusions of impurity phases in
cooperite single crystals were revealed at the micron
and submicron level. It was found that the composition
of cooperite differs from the stoichiometric composi-
tion in that certain samples are either deficient or rich in
platinum (Table 1). The Pt-to-S ratio varies in the range
of 5.18–7.54, compared to 6.08 for the cooperite refer-
ence sample. The generalized crystal-chemical formula
for the cooperite samples under study can be written as
Pt
1 –
x
S
1 +
x
(
0.1
x
+0.1)
.
X-ray Diffraction Data
Cooperite (PtS) crystallizes in the tetragonal system
(sp. gr.
P
4
2
/
mmc
). Its structure can be described as a tet-
ragonally distorted (
c
/2
a
= 0.88) cubic packing formed
by S atoms (the unit-cell height is equal to the height of
two such pseudo-unit cells). The sulfur atoms in coo-
perite are surrounded by four platinum atoms occupy-
ing the vertices of a slightly distorted tetrahedron. The
Pt atoms occupy two crystallographically independent
sites and are arranged so that the opposite faces of two
pseudocubic unit cells comprising the cooperite struc-
ture are centered (Fig. 3a). In both cases, the metal
atoms are in the square-planar environment formed by
four sulfur atoms to give the chromophore [PtS
4
]
(Fig. 3b).
The X-ray diffraction pattern of single-crystalline
cooperite was, on the whole, indexed within the tetrag-
onal unit cell (sp. gr.
P
4
2
/
mmc
) with the most often
observed parameters
a
= 3.4695(2) Å,
c
= 6.1066(9) Å,
c
/2
a
= 0.088 Å,
V
= 73.508(5) Å
3
, and
ρ
= 10.263 g/cm
3
.
The experimental X-ray powder diffraction patterns
100 µm 30 µm
30 µm 100 µm
(a) (b)
(c) (d)
Fig. 1.
Crystal morphology of cooperite. (a) A single-crys-
talline grain with hatched faces. Individual grains with
(b) smoothed edges and (c) traces of dissolution. (d) Single
crystals of dimensions larger than 1 mm.
1000
0 4 8 12 16 20
E
, keV
2000
3000
4000
5000
I, rel. units
PtMα
SKα
PtLα
PtLβPtLγ
Fig. 2. Energy-dispersive spectrum of cooperite single crys-
tals.
394
CRYSTALLOGRAPHY REPORTS Vol. 53 No. 3 2008
ROZHDESTVINA et al.
(measured from 20 mineral grains ground to powder)
contain 42 reflections, which are in good agreement
with the calculated data (the POWDER CELL program
[7]). The set of interplanar spacings corresponds to the
PtS phase (PDF 18-972) characterized by 34 lines. The
tetragonal unit-cell parameters calculated from the
X-ray powder diffraction pattern are equal (within
experimental error) to the most often observed unit-cell
parameters for single-crystalline samples (the average
deviation of the calculated interplanar spacings from
the measured values is 0.003, and the De Wolf figure-
of-merit is 28).
As compared to PtS (PDF 18-972), the X-ray pow-
der diffraction pattern (Fig. 4) and single-crystal X-ray
diffraction patterns for most samples are characterized
by a shift of most of reflections to larger angles. The
X-ray diffraction patterns of some samples are charac-
Table 1. Chemical composition of PtS single crystals
Analysis wt % at % Crystal-chemical formula
Pt Fe S ΣPt Fe S
1 82.506 15.900 98.406 46.05 53.95 Pt0.921S1.079
2 83.815 15.118 98.933 47.70 52.30 Pt0.954S1.046
3 83.875 0.032 15.082 98.989 47.70 0.05 52.25 (Pt0.954Fe0.001)0.955S1.045
4 84.204 14.885 99.090 48.20 51.80 Pt0.964S1.036
5 84.301 0.044 14.828 99.173 48.20 0.10 51.70 (Pt0.964Fe0.002)0.966S1.034
6 84.325 14.813 99.138 48.35 51.65 Pt0.967S1.033
7 84.481 0.031 14.720 99.232 48.50 0.05 51.45 (Pt0.970Fe0.001)0.971S1.029
8 84.251 0.035 14.557 99.372 48.70 0.05 51.25 (Pt0.974Fe0.001)0.975S1.025
9 84.684 0.025 14.456 99.405 49.00 0.05 50.95 (Pt0.980Fe0.001)0.981S1.019
10 84.408 14.394 99.442 49.10 50.90 Pt0.982S1.018
11 85.115 14.341 99.456 49.40 50.60 Pt0.988S1.012
12 85.186 0.027 14.299 99.512 49.45 0.05 50.50 (Pt0.989Fe0.001)0.990S1.010
13 85.287 0.022 14.238 99.547 49.60 0.05 50.35 (Pt0.992Fe0.001)0.993S1.007
14 84.833 14.107 99.669 49.70 50.30 Pt0.994S1.006
15 85.461 14.134 99.596 49.85 50.15 Pt0.997S1.003
16 85.594 14.055 99.649 50.05 49.95 Pt1.001S0.999
17 85.826 13.917 99.742 50.35 49.65 Pt1.007S0.993
18 85.970 13.830 99.800 50.55 49.45 Pt1.011S0.989
19 85.986 13.821 99.807 50.55 49.45 Pt1.011S0.989
20 86.011 13.806 99.817 50.60 49.40 Pt1.012S0.988
21 86.011 13.806 99.817 50.60 49.40 Pt1.012S0.988
22 86.200 13.693 99.893 50.85 49.15 Pt1.017S0.983
23 86.424 13.559 99.983 51.20 48.80 Pt1.024S0.976
24 86.484 13.524 100.008 51.25 48.75 Pt1.025S0.975
25 86.536 13.492 100.028 51.30 48.70 Pt1.026S0.974
26 86.743 13.369 100.112 51.60 48.40 Pt1.032S0.968
27 86.630 13.317 99.947 51.70 48.30 Pt1.034S0.966
28 87.223 13.082 100.305 52.30 47.70 Pt1.046S0.954
29 87.238 13.073 100.311 52.30 47.70 Pt1.046S0.954
30 87.256 13.062 100.318 52.35 47.65 Pt1.047S0.953
31 87.419 12.965 100.384 52.60 47.40 Pt1.052S0.948
32 88.082 12.569 100.651 53.55 46.45 Pt1.071S0.929
33 89.285 11.850 101.135 55.35 44.65 Pt1.107S0.893
34 89.292 11.846 101.138 55.35 44.65 Pt1.107S0.893
Note: The sensitivity of the method with respect to iron is 0.0025 wt %.
CRYSTALLOGRAPHY REPORTS Vol. 53 No. 3 2008
SINGLE-CRYSTALLINE COOPERITE (PtS): CRYSTAL-CHEMICAL CHARACTERIZATION 395
terized by a decrease in the interplanar spacing d100 to
3.446 Å (Fig. 5). An increase in the deficiency of plati-
num in cooperite and the appearance of an iron impu-
rity isomorphously replacing platinum causes a
decrease in all unit-cell parameters (Table 2). An
increase in the intensities of most of reflections, up to
the appearance of the 201 reflection (1.66–1.67 Å),
whose intensity in the theoretical X-ray diffraction pat-
tern is equal to zero, is apparently attributed to the
imperfections of the crystal structure caused by the
deviation from the stoichiometric composition of coo-
perite.
195Pt NMR Spectroscopic Data
In many cases, MAS NMR spectroscopy provides
unique information on the structural and electronic
state of metals in different compounds (see, for exam-
ple, [9–11]). Hence, we applied this method to this
method to the investigation of cooperite. The resulting
195Pt MAS NMR spectrum of single-crystalline cooper-
ite grains obtained by the summation of nine fragments
(recorded with a frequency step of 100 kHz) is pre-
sented in Fig. 6a. The width and shape of the MAS
NMR spectrum provide evidence that the 195Pt chemi-
cal-shift tensor is similar to that corresponding to the
axial symmetry. The anisotropy of the 195Pt chemical
shift is so large that the spectrum contains more than
100 spinning harmonics even at the sample-spinning
frequency of 8000(1) Hz. Because of this, to reveal the
195Pt resonance signal in the center of gravity of the
spectrum, which is characterized by the isotropic chem-
ical shift, we recorded the MAS NMR spectrum using
a continuous swinging of the sample-spinning fre-
quency. In this case, the amplitude of the signal (or sig-
nals) in the center of gravity of the spectrum remains
unchanged, whereas the spinning sidebands are cata-
strophically broadened (Fig. 6b). The above-described
data show that two crystallographically independent
platinum sites in the crystal lattice of cooperite are
structurally equivalent and are characterized by the iso-
tropic chemical shift δ(195Pt) = –1850 ppm (the width of
the resonance signal is 375 Hz).
To determine the principal values of the 195Pt chem-
ical-shift tensor, the static NMR spectrum of a thor-
oughly ground powder of cooperite was recorded in
fragments (Fig. 7). Three singularities corresponding to
δxx = –5920 ± 20 ppm, δyy = –3734 ± 5 ppm, and δzz =
+4023 ± 20 ppm were revealed. In addition, the anisot-
ropy of the 195Pt chemical shift, which was specified as
δaniso = δzzδiso, and the asymmetry parameter η = 0.37
{η = (δyyδxx)/(δzzδiso)}, were calculated from the
experimental data; 195Pt δaniso = 5873 ppm and η = 0.37.
It should be noted that η = 0 corresponds to the axially
symmetric chemical-shift tensor. An increase in η in the
range from 0 to 1 reflects an increase in the contribution
of the orthorhombic component.
Pt(+2)
S(–2)
c
b
a
(a) (b)
0.4
0 35 50 65 80 95 110 125 140 155 170
0
2θ, deg
0.2
0.6
0.8
1.0
I, rel. units
100 002
110
102
112
103
200
201
211
104
203
213
204
310
312
006
106 215
116
304
314
206400
216 410 411
330
332
226 413
420
1.0
0.8
0.6
0.4
0
I, rel. units
0.2
100
002
101
25 30 2θ, deg
Fig. 3. (a) Crystal lattice and (b) the mutual spatial orienta-
tion of the platinum and sulfur polyhedra in cooperite (mod-
els were constructed with the use of the TOPOS-4.0 pro-
gram).
Fig. 4. Experimental X-ray powder diffraction pattern of
cooperite.
Fig. 5. Fragment of the X-ray powder diffraction pattern of
cooperite (the profiles and intensities of simulated reflec-
tions are shown by a dashed line).
396
CRYSTALLOGRAPHY REPORTS Vol. 53 No. 3 2008
ROZHDESTVINA et al.
There is a good probability that the least shielded
direction for the platinum nucleus (the z axis of the
chemical-shift tensor) coincides with the fourfold sym-
metry axis of the square-planar chromophore [PtS4].
Two other directions cannot coincide with the x and y
molecular axes, because the very small difference
(0.0001 Å) in the length of the nonequivalent Pt–S
bonds cannot account for considerable differences in
the degree of electronic shielding. Most likely, these
directions coincide with the bisectors of the S–Pt–S
angles. However, in the case of a more shielded posi-
tion of the platinum nucleus, the axis passes inside the
small-size four-membered metallocycle [Pt2S2]. For the
less shielded position, the axis passes through the
extended eight-membered ring [Pt4S4].
Thoroughly ground cooperite grains give a multi-
component anisotropic ESR spectrum superimposed
onto a broad envelope (Figs. 8a, 8b). The double differ-
entiation (Fig. 8c) allowed us to reveal the correspon-
dence of the spectrum to the axial symmetry (gx = gy < gz),
the presence of quartets of the components of the
hyperfine structure in the parallel and perpendicular
orientations, and the additional absorption line [12, 13]
at high field. The well-resolved quartet structure of
anisotropic ESR spectra is typical of magnetically
dilute copper(II) compounds. The ESR parameters
refined based on the results of the computer simulation
(g|| = 2.358, = 117 Oe, g = 2.116, = 29 Oe)
are also consistent with interactions between the
unpaired electron (the ground state AO) and
the 63,65Cu nucleus (I = 3/2). However, these results pro-
vide no evidence that copper(II) is included in the crys-
tal lattice of cooperite. The fact is that the square-planar
chromophores [CuS4], for example, in coordination
copper(II) compounds with dithiooxamide and its
dialkyl-substituted derivatives [14], the N,N-dialky-
ldithiocarbamate ligand [15], and the O,O'-dialky-
ldithiophosphate ligand [16], are characterized by sub-
stantially smaller g factors and larger hyperfine-struc-
ture constants (g|| = 2.085, = 164 Oe, g = 2.024,
ACu
|| ACu
3dx2y2
ACu
||
Table 2. Unit-cell parameters and volume of PtS single crystals
Crystallochemical formula a, Å c, Å c/2aV, Å3
Pt1.012S0.988 3.4710(2) 6.1084(5) 0.88 73.595(5)
Pt1.001S0.999 3.4703(1) 6.1089(5) 0.88 73.569(5)
Pt0.997S1.003 3.4702(3) 6.1089(7) 0.88 73.569(6)
(Pt0.992Fe0.001)0.993S1.007 3.4690(1) 6.1018(5) 0.879 73.431(5)
(Pt0.989Fe0.001)0.990S1.010 3.4686(1) 6.1024(5) 0.88 73.419(6)
Pt0.982S1.018 3.4642(1) 6.0931(7) 0.879 73.125(6)
(Pt0.980Fe0.001)0.981S1.019 3.4637(1) 6.0606(8) 0.875 72.712(7)
(Pt0.974Fe0.001)0.975S1.025 3.4635(2) 6.0562(5) 0.874 72.649(5)
(Pt0.970Fe0.001)0.971S1.029 3.4630(3) 6.0514(7) 0.874 72.573(6)
(Pt0.964Fe0.002)0.966S1.034 3.4634(1) 6.0494(7) 0.873 72.567(6)
(Pt0.954Fe0.001)0.955S1.045 3.4639(1) 6.0398(5) 0.872 72.47(5)
2500 0 2500 –5000
δ, ppm
(a)
(b)
Fig. 6. (a) Resulting 195Pt MAS NMR spectrum of single-
crystalline grains of cooperite (the upper vertical mark indi-
cates the resonance signal in the center of gravity of the
spectrum). The sample-spinning frequency was 8000(1)
Hz. (b) The spectrum measured with the use of continuous
swinging of the spinning frequency.
4000 3000 2000 1000 –3000 – 4000 –5000 –6000
δ, ppm
Fig. 7. Fragment of 195Pt static NMR spectrum of a coo-
perite powder.
CRYSTALLOGRAPHY REPORTS Vol. 53 No. 3 2008
SINGLE-CRYSTALLINE COOPERITE (PtS): CRYSTAL-CHEMICAL CHARACTERIZATION 397
= 40 Oe; the most typical parameters are given).
The analysis of the ESR data for copper(II) in polyhe-
dra with different compositions and geometric parame-
ters [17] leads to the conclusion that the ESR spectrum
observed for cooperite is determined by copper(II) in
the octahedral environment formed by oxygen atoms.
These data suggest that changes in the hydrothermal
conditions in the course of crystallogenesis of cooperite
were accompanied by adsorption of copper(II) on the
surface as oxides, hydroxides, or hydroxycarbonate
compounds (in the case of the layer-by-layer crystal
growth, which is dominant for minerals).
CONCLUSIONS
The generalized crystal-chemical formula for sin-
gle-crystalline samples of cooperite, which were found
in placers from the Yuna-Dan’sk gold-ore cluster of the
platinum-bearing province in the Sea of Okhotsk region
of the Maœmakansk zone in the Far East, can be written
as Pt1 – xS1 + x (–0.1 x +0.1). According to the ESR
data, single-crystalline cooperite samples contain cop-
per(II), which is not included in the crystal lattice of
cooperite. Apparently, the presence of copper(II) in
cooperite, as well as a deficient or excess amount of
platinum, are responsible for the appearance of defor-
mation defects in the structure, which are reflected in
X-ray diffraction patterns. It was demonstrated that
195Pt static and MAS NMR spectroscopy can, in princi-
ple, be used for studying natural platinum minerals.
The principal values of the 195Pt chemical-shift tensor
for cooperite were determined.
ACu
ACKNOWLEDGMENTS
This study was supported by the Russian Founda-
tion for Basic Research and the Far East Division of the
Russian Academy of Sciences (Program “Far East,
project no. 06-03-96009) and the Presidium of the Far
East Division of the Russian Academy of Sciences
(project no. 06-III-A-08-339).
A.V. Ivanov acknowledges the support of the Agri-
cola Research Centre at the Luléa University of Tech-
nology (Luléa, Sweden).
REFERENCES
1. V. V. Ivanov, Ecological Geochemistry of Elements:
Handbook in 6 Vols. Vol. 5: Rare d Elements, Ed. by
É. K. Burenkov (Ékologiya, Moscow, 1997) [in Rus-
sian].
2. O. E. Yushko-Zakharova, V. V. Ivanov, L. N. Soboleva,
et al., Handbook of Minerals of Noble Metals (Nedra,
Moscow, 1986) [in Russian].
3. L. P. Plyusnina, G. G. Likhoœdov, and A. I. Khanchuk,
Dokl. Akad. Nauk 405 (1), 105 (2005).
4. L. P. Plyusnina, G. G. Likhoœdov, and I. Ya. Nekrasov,
Dokl. Akad. Nauk 370 (1), 99 (2000).
5. V. G. Moiseenko, V. A. Stepanov, L. V. Éœrish, et al.,
Platinoferous Fractions of the Russian Far East
(Dal’nauka, Vladivostok, 2004) [in Russian].
6. P. P. Safronov and V. G. Moiseenko, Mineral Associa-
tions of Platinoids from Gold-Bearing Placers of the
Zeya-Selemja Area (Priamurie). Geodynamics and Met-
allogeny (Dal’nauka, Vladivostok, 1999), p. 157 [in Rus-
sian].
7. W. Kraus and G. Nolze, J. Appl. Crystallogr. 29 (3), 301
(1996).
8. G. A. Kirakosyan, Koord. Khim. 19 (7), 507 (1993).
9. A. V. Ivanov, A. V. Gerasimenko, O. N. Antzutkin, and
W. Forsling, Inorg. Chim. Acta 358 (9), 2585 (2005).
10. A.-C. Larsson, A. V. Ivanov, K. J. Pike, et al., J. Magn.
Reson. 177 (1), 56 (2005).
11. A. V. Ivanov, A. V. Gerasimenko, A. A. Konzelko, et al.,
Inorg. Chim. Acta 359 (12), 3855 (2006).
12. I. V. Ovchinnikov and V. N. Konstantinov, J. Magn.
Reson. 32, 179 (1978).
13. Ph. H. Rieger, Electron Spin Resonance (Senior
Reporter Symons M.C.R.) (Athenaeum, Newcastle upon
Tyne, 1993), Vol. 13B, p. 178.
14. P. M. Solozhenkin, A. V. Ivanov, and E. V. Rakitina, Zh.
Neorg. Khim. 28 (12), 3081 (1983).
15. A. V. Ivanov and P. M. Solozhenkin, Zh. Neorg. Khim.
35 (6), 1537 (1990).
16. A. V. Ivanov, A.-K. Larsson, N. A. Rodionova, et al., Zh.
Neorg. Khim. 49 (3), 423 (2004).
17. S. A. Al’tshuler and B. M. Kozyrev, Electron Spin Reso-
nance in Compounds of Intermediate-Group Elements
(Nauka, Moscow, 1972) [in Russian].
Translated by T. Safonova
116 Oe
H
262 Oe
H116 Oe
H
DPPH DPPH
DPPH
(a)
(b)
(c) 1
2
Fig. 8. ESR spectra of a cooperite powder. (a) The overall
view. (b) The anisotropic spectrum of copper(II), the first
derivative. (c) The third derivatives; 1, experiment; 2,
model.
... The computational XRD clearly shows that the highest intensities above 60 emanate from the (101), (112), and (211) planes, which is in agreement with the reported experimental XRD patterns. 8−11 Interestingly, both (002) and (101) peak intensities are at around 2θ = 29°, which has been found experimentally, 11 which makes them not easily distinguishable. However, the (002) peak has low intensity compared to the (101) peak, which reached 100 intensity, as shown in Table S1, demonstrating that the (101) plane is the most dominant for cooperite. ...
... A determined using X-ray powder diffraction. [9][10][11]51 The surfaces were obtained from a relaxed bulk model of cooperite and modeled using periodic supercells, consisting of eight layers for (001), (100), and (110) surface slabs. For (101), (111), (112), and (211) surface slabs, 12 layers, 18 layers, 5 layers, and 21 layers were adopted, respectively. ...
Article
Full-text available
Cooperite (PtS) is one of the main sources of platinum in the world and has not been given much attention, in particular from the computational aspect. Besides, the surface stability of cooperite is not fully understood, in particular the preferred surface cleavage. In the current study, we employed computer modeling methods within the plane-wave framework of density functional theory with dispersion correction and the U parameter to correctly predict the bulk and surface properties. We reconstructed and calculated the geometries and surface energies of (001), (100), (101), (112), (110), (111), and (211) cooperite surfaces of stoichiometric planes. The Pt d-orbitals with U = 4.5 eV and S p-orbitals with U = 5.5 eV were found optimum to correctly predict a band gap of 1.408 eV for the bulk cooperite model, which agreed with an experimental value of 1.41 eV. The PtS-, Pt-, and S-terminated surfaces were investigated. The structural and electronic properties of the reconstructed surfaces were discussed in detail. We observed one major mechanism of relaxation of cooperite surface reconstructions that emerged from this study, which was the formation of Pt-Pt bonds. It emanated that the (110) and (111) cooperite surfaces underwent significant reconstruction in which the Pt2+ cation relaxed into the surface, forming new Pt-Pt (Pt2 2+) bonds. Similar behavior was perceived for (101) and (211) surfaces, where the Pt2+ cation relaxed inward and sideways on the surface, forming new Pt-Pt (Pt2 2+) bonds. The surface stability decreased in the order (101) > (100) ≈ (112) > (211) > (111) > (110) > (001), indicating that the (101) surface was the most stable, leading to an octahedron cooperite crystal morphology with truncated corners under equilibrium conditions. However, the electronic structures indicated that the chemical reactivity stability of the surfaces would be determined by band gaps. It was found that the (112) surface had a larger band gap than the other surfaces and thus was a chemical stability competitor to the (101) surface. In addition, it was established that the surfaces had different reactivities, which largely depended on the atomic coordination and charge state based on population atomic charges. This study has shown that cooperite has many planes/surface cleavages as determined by the computed crystal morphology, which is in agreement with experimental X-ray diffraction (XRD) pattern findings and the formation of irregular morphology shapes.
... Crystal structure of cooperite PtS (Fig. 8C) belongs to the tetragonal system (sp gr 131 P42/m m c; Rozhdestvina et al., 2008). Atoms of Pt are in the position 2c; atoms of S are in the position 2e. ...
Article
Full-text available
Pyrite (FeS2) is a typical container of Pt in ores of magmatic and hydrothermal origin and in some carbon-rich ores of sedimentary-diagenetic origin. Knowledge of the state of Pt disseminated in the matrix of pyrite, including local atomic environment (type of atoms in the nearest and distant coordination shells, coordination numbers, interatomic distances) and oxidation state, is necessary for physical-chemical modeling of platinum group element mineralization and for the improvement of Pt ore extraction and processing technologies. Here we report results of an investigation of local atomic structure of synthetic Pt-bearing pyrites by means of X-ray absorption spectroscopy (XAS). Synthesis experiments, performed at 580° and 590°C in a Pt-saturated system by means of salt-flux method, yielded crystals of pyrite with concentrations of Pt up to 4 wt %. Scanning electron microscopy (SEM) and electron probe microanalysis (EPMA) showed that the distribution of Pt within the pyrite grains is of zonal character, but within the distinct zones Pt is distributed homogeneously. Negative correlation between the concentrations of Pt and Fe was observed in the synthesized pyrite grains. The slope of the correlation line corresponds to the formation of the solid solution in the Pt-Fe-S system and/or to the formation of PtS2. The XAS experiments revealed the existence of two forms of Pt in pyrite. The main form is the solid solution Pt(IV), which isomorphically substitutes for Fe. The Pt-S distance in pyrite is ~0.1 Å longer than that of Fe-S in pure pyrite. The distortion of the pyrite crystal structure disappears at R >2.5 Å. The second Pt-rich form was identified by means of high-resolution transmission electron microscopy (HRTEM) as nanosized inclusions of PtS2. Heating experiments with in situ registration of X-ray absorption spectra resulted in partial decomposition (dissolution) of PtS2 nanosized inclusions with the formation of the solid solution (Fe1-xPtx)S2. Therefore, the PtS2 nanosized particles can be considered as a quench product. Our data demonstrate that both Pt solid solution and PtS2 nanosized inclusions (at high Pt content) can exist in natural Pt-bearing pyrites.
Article
A re-evaluation of data relating to the crystal structures, phase equilibria, and ideal chemical formulae of the key minerals belonging to the PtS–PdS binary, namely vysotskite, braggite, and cooperite, has led to redefinitions of the ideal chemical formulae for vysotskite and braggite. Results show that vysotskite and braggite are isostructural and crystallize in the space group P42/m. Their crystal structures contain three Me sites and a single S site, giving rise to the general crystal-chemical formula Me(1)2Me(2)2Me(3)4S8. Ordering of Pt and Pd leads to the revised ideal formula of Pd2Pd2Pd4S8 (Z = 1) for vysotskite, which can be simplified to PdS (Z = 8), and a revised ideal formula of Pd2Pt2Pt4S8 (Z = 1) for braggite, which can be simplified to PdPt3S4 (Z = 2). Cooperite, ideally PtS (Z = 8), crystallizes in a related, but different, space group, P42/mmc. Its crystal structure contains a single Me and S site. Both Pt and Pd are disordered over the Me site but with Pt > Pd . Oversimplification of the ideal chemical formulae for all these minerals and failure to consider crystal-structure implications has led to confusion in the literature. Consideration of experimental data combined with that for minerals from world-wide occurrences suggests complete solid solution between vysotskite and braggite, but of a more limited degree between cooperite and braggite, consistent with what would be expected when their respective crystal structures are considered. The maximum Pt possible in braggite should correspond to (Pd0.51Pt0.49)Σ1Pt3S4 (or Pd0.51Pt3.49S4); Pt:Pd = 0.87, which requires Pd 6.29, Pt 78.86, S 14.85, total 100.00 wt.%) and the maximum Pd content, to Pd(Pt0.505Pd0.495)Σ1(Pt1.505Pd0.495)Σ2S4 (or Pd1.99Pt2.01S4); Pt:Pd = 0.50, which requires Pd 28.93, Pt 53.56, S 17.52, total 100.00 wt.%). In light of the complete solid solution between vysotskite and braggite, the 50% rule should be applied to compositions between them. Thus, within the PtS–PdS binary, it is recommended that (1) vysotskite be used for those minerals having compositions ranging from 100 to 56.5 mol.% PdS; (2) braggite, for those having compositions that range from 56.5 to 13 mol.% PdS; and (3) cooperite, for those having <13 mol.%.
Article
Pyrrhotite Fe 1– x S is the main sulfide component of platinum-group element (PGE) ores and commonly contains from a few tenths to a few dozen ppm of disseminated Pt. Here we report an X-ray absorption spectroscopy investigation into the state of Pt in synthetic pyrrhotite in combination with theoretical spectra modelling. The pyrrhotite crystals were obtained by means of the salt flux technique, using a eutectic mixture of alkali metal halides as the transport media. Analysis of the chemical composition of synthesised crystals showed that an increase of the temperature and sulfur fugacity yields higher concentrations of Pt in pyrrhotite. The Pt content reached 0.6 wt.% at the maximum temperature and sulfur fugacity ( t = 720°C, log $f_{{\rm S}_ 2}$ = –0.1) achieved in a Pt-saturated system. X-ray absorption near-edge structure (XANES) analysis of Pt L 3 -edge spectra revealed that Pt is present in pyrrhotite in the 4+ and 2+ ‘formal’ oxidation states. Theoretical modelling of XANES and interpretation of extended X-ray absorption fine structure (EXAFS) spectra showed that Pt ⁴⁺ substitutes for Fe in the crystal lattice of pyrrhotite, whereas Pt ²⁺ forms PtS-like clusters disseminated in the pyrrhotite matrix. Atoms of isomorphous Pt ⁴⁺ are surrounded by 6 S atoms at a distance of 2.39 ± 0.02 Å. According to theoretical simulations using the FDMNES program, the second coordination sphere of the solid-solution Pt contains one vacancy in the Fe sublattice within the Fe-layer. The Pt ²⁺ S-like clusters can be considered as a quench product. High sulfur fugacity stabilises the solid-solution Pt and prevents the formation of the PtS-like clusters during cooling. The maximum content of the solid-solution Pt in pyrrhotite is ca. 50 times lower than in pyrite and can be approximated by a straight line in the log C (Pt) vs . 1/ T plot, it increases from 1 ppm at 350°C to 3 wt.% at 900°C.
Article
A manufacturing-compatible 300 mm chamber reactor for atomic layer deposition or chemical vapour deposition, and pre-fitted with H2/H2S gases that can be uniformly delivered to the wafer surface, is employed to thermally convert Pt to PtS in a H2/H2S gaseous atmosphere for 7 hours at a chamber temperature of 550 °C. Prior to conversion, platinum layers 5 nm thick are uniformly deposited by electron beam evaporation onto ∼30 nm of amorphous aluminium sesquioxide deposited by atomic layer deposition on, (a) p-type silicon, and (b) c-plane sapphire. Structural characterisation is performed by high-resolution cross-sectional transmission-electron microscopy, scanning-electron microscopy and Raman spectroscopy, confirming the formation of continuous films of polycrystalline platinum monosulfide (PtS) with a ∼15 nm thickness. Electrical characterisation is performed by 4-point resistivity and Hall-effect transport measurements on van der Pauw structures of PtS on aluminium sesquioxide on c-plane sapphire, and by back-gate junctionless MOSFET device measurements for PtS on aluminium sesquioxide on p-type silicon, showing that PtS behaves as a semiconductor with a mobility of ∼16 cm²/V.s and with an n-type carrier concentration of ∼1.2 × 10¹⁵ cm⁻³. Advanced commercial-grade Sentaurus simulations, alongside density-functional theory calculations, agree well with the experimental observations and suggest a large bandgap of ∼1.58 eV may be possible that could lead to a low off-current and a high Ion/Ioff ratio, suggesting that PtS may be an advanced material candidate for future device integration with CMOS and for 3D integration applications in Beyond-CMOS and More-than-Moore technologies.
Article
Cooperite, or platinum sulfide (PtS), is a rare mineral that generally exists as microscale, irregularly-shaped crystallites. The presence of impurities, in both naturally-occurring and synthesized samples, has hindered the study of its optical properties in the past. In this work, we prepare large-scale, uniform PtS films in bulk to 2D form through the thermally assisted conversion (TAC) method. An abnormal trend is observed in linear spectral studies whereby the optical bandgap narrows as the film thickness decreases. A model based on the continuous distribution of carriers in real space, which can be regarded as a quantum well normal to the plane is used to describe the thickness-dependent carrier recombination phenomenon. In the nonlinear optical measurements, PtS exhibits ultrafast saturable absorption and self-defocusing properties in the visible region, which are dominated by the resonant electronic nonlinearities.
Article
Polycrystalline bis(dialkyldithiophosphato)Pt(II) complexes of the form Pt{S2P(OR)2}2 (R = ethyl, iso-propyl, iso-butyl, sec-butyl or cyclo-hexyl group) were studied using solid-state 31P and 195Pt NMR spectroscopy to determine the influence of the alkyl substituents to the structure of the central chromophore. Measured anisotropic chemical shift (CS) parameters for 31P and 195Pt tensors afford more detailed chemical and structural information as compared to isotropic NMR parameters, such as chemical shifts and J-couplings, alone. Advanced theoretical modeling at hybrid DFT level including both crystal lattice and relativistic spin—orbit effects qualitatively reproduced the experimentally observed CS tensors, thus supporting the experimental analysis, as well as providing extensive orientational information for the tensors in the molecular frame. The results for 195Pt CS tensors demonstrated that differences in the alkyl substitutes of the dialkyldithiophosphate ligands in the complexes studied have an insignificant effect on the distorted-square form structure of the central PtS4 core. However, the principal values of 31P chemical shift tensors in the dialkyldithiophosphate ligands do differ significantly, which may be used to distinguish between the different complexes. Relativistic effects (both scalar and spin—orbit) were shown to be important for the NMR parameters of both 31P and 195Pt nuclei. The effects due to the periodic crystal lattice were found to be non-negligible, especially for the CS tensor of the heavy-metal 195Pt isotope. A particular correction model for incorporating lattice effects was adopted to avoid a severe deterioration of the anisotropic parameters of 195Pt due to the high requirements posed on the pseudopotential quality in such calculations. The pseudopotentials available in standard software may be inadequate for periodic calculations of anisotropic NMR parameters of heavy nuclei.
Article
Full-text available
The main component of this program is a simultaneous representation of the unit cell and the calculated powder pattern. It allows the manipulation of the crystal structure by moving selected atoms of the asymmetric unit. The resulting powder pattern can be directly compared to experimental data in order to obtain reliable starting values for further computations in refinement programs.
Chapter
Specialist Periodical Reports provide systematic and detailed review coverage of progress in the major areas of chemical research. Written by experts in their specialist fields the series creates a unique service for the active research chemist, supplying regular critical in-depth accounts of progress in particular areas of chemistry. For over 80 years the Royal Society of Chemistry and its predecessor, the Chemical Society, have been publishing reports charting developments in chemistry, which originally took the form of Annual Reports. However, by 1967 the whole spectrum of chemistry could no longer be contained within one volume and the series Specialist Periodical Reports was born. The Annual Reports themselves still existed but were divided into two, and subsequently three, volumes covering Inorganic, Organic and Physical Chemistry. For more general coverage of the highlights in chemistry they remain a 'must'. Since that time the SPR series has altered according to the fluctuating degree of activity in various fields of chemistry. Some titles have remained unchanged, while others have altered their emphasis along with their titles; some have been combined under a new name whereas others have had to be discontinued. The current list of Specialist Periodical Reports can be seen on the inside flap of this volume.
Article
Effect of As and Ni on platinum solubility in chloride solutions at 300-500°C and fluid pressure 1 kbar was investigated. It is shown that sulfur and arsenic present in natural hydrothermal systems favor platinum precipitation from aqueous chloride solutions via generation of intermediate metastable complexes disproportioning with platinum sulfide and arsenide precipitation. Presence of nickel in hydrothermal medium obstacles formation of individual platinum sulfides and arsenides. For the first experimental evidence is obtained for platinum inclusion in arsenopyrite and pyrrhotine that enables to regard the latter as platinum mineral-concentrators.
Article
The nickel(II) and copper(II) complexes [M{S2P(OR) 2}2] (M = Ni, 63Cu, 65Cu; R = C 2H5, C3H7, i-C3H 7, C4H9, i-C4H9, s-C 4H9, i-C5H11, c-C6H 11) with eight symmetric O,O′-dialkyl phosphorodithioates were synthesized, and their structures and spectral properties were studied by EPR and CP/MAS 13C and 31P NMR. As determined by EPR, the [CuS4] chromophores in the lattice of nickel(II) complexes have predominantly a square-planar structure. The chromophores in complexes with phosphorodithioate ligands incorporating bulky alkyl substituents have somewhat distorted geometries. NMR shows that the Dtph groups in most nickel(II) complexes are structurally equivalent. An exception is [Ni{S 2P(OC3H7)2}2], which gives rise to two (1:1) 31P NMR signals. For [Ni(S 2P(O-i-C4H9)2}2], the α and β modifications were obtained. In the α modification, the phosphorodithioate groups are nonequivalent, whereas the β modification has structurally equivalent ligands. The broadening of the 31P NMR signal of the [Ni{S2P(O-s-C4H9) 2}2] complex is rationalized by the existence of six optical isomers (due to two chiral centers in each of the ligands). For the Dtph groups acting as bidentate terminal ligands in the nickel(II) complexes, the 31P chemical shift anisotropy δaniso and asymmetry parameter η were calculated from experimental NMR spectra. Single-crystal X-ray diffraction shows that bis(O,O′-dipropyl phosphorodithioato) nickel(II) exists in the form of two centrosymmetric molecules. In the structure of α-bis(O,O′-dipropyl phosphorodithioato)nickel(II), the Dtph ligands in the molecule are nonequivalent. The experimental 31P NMR signals in the spectra of [Ni{S2P(OC3H7) 2}2] and α-[Ni{S2P(O-i-C 4H9)2}2] are assigned to the structural positions of phosphorus atoms in the corresponding molecular structures determined by X-ray diffraction.
Article
O,O′-Dipropyldithiophosphate and O,O′-dibutyldithiophosphate (Dtph) cadmium(II) complexes were prepared and studied by means of heteronuclear 31P, 113Cd, 31C CP/MAS NMR spectroscopy and single-crystal X-ray diffraction. Linear-chain polynuclear structures have been established for both cadmium(II) complexes, in which each pair of equivalent dithiophosphate groups, playing the same bridging structural function, asymmetrically links the neighbouring cadmium atoms. One remarkable structural feature of the synthesised cadmium(II) compounds is defined by the alternation of two types of conformationally different (‘chair’–‘saddle’) eight-membered rings [Cd2S4P2] in the polymeric chains. Therefore, in both 31P NMR and XRD data, the bridging dithiophosphate ligands exhibit structural inequivalence in pairs. The structural states of both Dtph ligands and cadmium atoms have been characterised by the 31P and 113Cd chemical shift tensors, which display a profound axially symmetric and mainly rhombic characters, respectively. All experimental 31P resonances were assigned to the phosphorus structural sites in both resolved structures.
Article
Crystalline N,N-cyclo-pentamethylenedithiocarbamate (PmDtc) cadmium(II) complex was prepared and studied by means of 15N, 113Cd CP/MAS NMR spectroscopy and single-crystal X-ray diffraction. The unit cell of the cadmium(II) compound comprises two centrosymmetric isomeric binuclear molecules [Cd2{S2CN(CH2)5}4], which display structural inequivalence in both 15N and 113Cd NMR and XRD data. There are pairs of the dithiocarbamate ligands exhibiting different structural functions in both isomeric molecules. Each of the terminal ligands is bidentately coordinated to the cadmium atom and forms a planar four-membered chelate ring [CdS2C]; whereas pairs of the tridentate bridging ligands combine two neighbouring cadmium atoms forming an extended eight-membered tricyclic moieties [Cd2S4C2], whose geometry can be approximated by a ‘chair’ conformation. The structural states of cadmium atoms were characterised by almost axially symmetric 113Cd chemical shift tensors. All experimental 15N resonance lines were assigned to the nitrogen structural sites in both isomeric binuclear molecules.
Article
Analytical solutions for locating extra absorption peaks in “polycrystalline” EPR spectra of systems with anisotropic g-factor and hyperfine structure are obtained, and general conditions for the existence of these peaks are formulated. The effect of second-order terms of perturbation theory on these conditions is revealed. As an example, the frequency intervals of the occurrence of extra absorption peaks and their positions in the EPR spectrum of a frozen solution of Cu acetylacetonate are calculated.
Article
A number of lead(II) O,O'-dialkyldithiophosphate complexes were studied by (13)C, (31)P, and (207)Pb MAS NMR. Simulations of (31)P chemical shift anisotropy using spinning sideband analysis reveal a linear relationship between the SPS bond angle and the principal values delta(22) and delta(33) of the (31)P chemical shift tensor. The (31)P CSA data were used to assign ligands with different structural functions. In the cases of diethyldithiophosphate and di-iso-butyldithiophosphate lead(II) complexes, (2)J((31)P, (207)Pb)-couplings were resolved and used to confirm the suggested assignment of the ligands. The SIMPSON computer program was used to calculate (31)P and (207)Pb spectral sideband patterns.