ArticlePDF Available

ESCA Studies of Framework Silicates With the Sodalite Structure: 1. Comparison of Purely Siliceous Sodalite and Aluminosilicate Sodalite

Authors:

Abstract

The first detailed X-ray photoelectron spectroscopy (ESCA) study of materials with the sodalite structure shows how the purely siliceous sodalite prepared in a nonaqueous medium differs from the conventional aluminosilicate sodalite and from various zeolites and clay minerals. The purely siliceous sodalite gives a unique set of binding energies and valence band patterns similar to those for silica, while the ESCA pattern of an aluminosilicate sodalite is similar to that of zeolite Na-A. The small differences in binding energies for aluminosilicate sodalite are due to the presence of some K+ and Mg2+ and the fact that the framework Si/Al ratio is slightly higher than unity. While the binding energies for zeolite Na-A and aluminosilicate sodalite form a common set, their ESCA patterns differ substantially from that for the layered aluminosilicate kaolinite. ESCA is thus shown to be sensitive to structure as well as composition.
1302
J.
Phys. Chem. 1994,98,
1302-1305
ESCA Studies
of
Framework Silicates with the Sodalite Structure.
1.
Comparison
of
Purely
Siliceous Sodalite and Aluminosilicate Sodalite
Bruno Herreros, Heyong He, Tery
L.
Barr,*’+ and Jacek
Klinowski’
Department
of
Chemistry, University
of
Cambridge, Lensfield Road, Cambridge CBl2
IEW,
U.K.
Received: June 29, 1993;
In
Final Form: October 29, 1993’
The first detailed X-ray photoelectron spectroscopy (ESCA) study of materials with the sodalite structure shows
how the purely siliceous sodalite prepared in a nonaqueous medium differs from the conventional aluminosilicate
sodalite and from various zeolites and clay minerals. The purely siliceous sodalite gives a unique set of binding
energies and valence band patterns similar
to
those for silica, while the ESCA pattern
of
an aluminosilicate
sodalite is similar to that of zeolite Na-A. The small differences in binding energies for aluminosilicate sodalite
are due to the presence of some
K+
and Mg2+ and the fact that the framework Si/Al ratio is slightly higher
than unity. While the binding energies for zeolite Na-A and aluminosilicate sodalite form a common set, their
ESCA patterns differ substantially from that for the layered aluminosilicate kaolinite. ESCA is thus shown
to be sensitive to structure as well as composition.
Introduction
In 1887 Sir William Thomson (later Lord Kelvin) demon-
stratedl that the truncated octahedron (more correctly a tetra-
kaidodecahedron), the simplest space-filling polyhedron apart
from the cube, divides space with minimum partitional area. The
tetrakaidodecahedron is thus the three-dimensional equivalent
of a regular hexagon, the most economical divisor of a plane. The
frameworkof sodalite is a periodic array of such [4686] polyhedra
which are known in mineralogy as sodalite or
@
cages. The cages
are
in
turn built from corner-sharing SiO4”- and A10P tetrahedra.
The negative charge of the framework brought about by the
presence of the aluminate tetrahedra is neutralized by charge-
balancing cations, typically Na+. In addition, sodalite cages may
accomodate a variety of guest species, such as inorganic salts,
water, or organic molecules. Many natural and synthetic
materials with the sodalite structure and a variety of enclathrated
species have been described.2 Sodalite may have a range of
compositions2 but generally has almost equal amounts of
tetrahedral Si and A1 in its framework. A typical unit cell formula
is Na6A16Si6024*2NaCl. Because all tetrahedral sites in sodalite
are crystallographically equivalent and only one kind of cages is
present, sodalite is an archetypal molecular sieve.
Sodalite is normally synthesized from strongly basic media
under mild
hydro thermal condition^.^
However, by using ethylene
glycol as solvent, Bibby and Dale4 prepared a purely siliceous
sodalite. The unit cell composition of the product is Si] 2024-2C2H4-
(OH)2, and the glycol is encapsulated in the sodalite
cage^.^^^
We
report an X-ray photoelectron spectroscopy (ESCA) study of
sodalites including its purely siliceous analogue. Our aim was
(i) to carry out the first detailed ESCA study of sodalite, (ii) to
see
how the purely siliceous sodalite differs from its aluminosilicate
analogue, and
(iii)
to compare the results with those for various
zeolites and clays.6 This work is part of a concerted effort to
extend the range of application of the technique to cover all types
of silicates.
Experimental Section
Synthesis
of
Purely Siliceous Sodalite.
The starting materials
were ethylene glycol, Cab-0-Si1 fused silica, and sodium hydroxide
in the molar ratio 4 SiOz:NaOH:40 CzHd(OH)z, slightly different
from that originally de~cribed.~ NaOH was first added to the
Permanent address: Department
of
Materials and Laboratory for Surface
@
Abstract published in
Adoance
ACS
Abstracts,
January 1, 1994.
Studies, University
of
Wisconsin-Milwaukee, Milwaukee, WI
53201.
0022-365419412098-1302$04.50/0
glycol, and the mixture was stirred for
6
h until the base dissolved.
Fused silica was then added under stirring, producing a thick gel
which was transferred into a Teflon-lined stainless steel autoclave
and heated under autogeneous pressure at 180 OC for 3 days. The
crystalline product was finally filtered, washed withdistilled water,
and dried at room temperature. It crystallizes in the cubic space
group Im3m, with a cell parameter
a
=
8.83
A
(as compared with
8.94
A
for a conventional aluminosilicate sodalite).
A synthetic aluminosilicate sodalite, made using standard
procedure^,^
was also examined in the as-prepared state. Other
results referred to in this study were described in earlier
publications.”l
I
Solid-state
NMR.
Magic-angle-spinning (MAS) NMR
spec-
tra were acquired at 9.4 T using a Chemagnetics CMX-400
spectrometer with rotors 4 mm in diameter spun in nitrogen gas
at 8-10 kHz. 29Si spectra were measured at 79.5 MHz with 45O
radiofrequency pulses and 30-s recycle and A1 spectra
at 104.3 MHz with very short,
0.6-ps
(less than loo), radiof-
requency pulses and 0.3-s recycle delays. IH-W cross-polar-
ization (CP) MAS NMR spectra were measured at 100.63 MHz
with 4-ms contact times. The length of the IH and I3C u/2
pulses was 3
ps,
recycle delay
10
s,
and MAS rate
8
kHz. IH-
29Si CP/MAS spectra were recorded at 7.05
T
using a Bruker
MSL-300 spectrometer with a single contact pulse sequence,
8-ms
contact time, 4.8-ps IH 90° pulse, and 10-s recycle delay. The
Hartmann-Hahn condition for ‘H-27A1 and lHJ9Si CP/MAS
was established using kaolinite12 and for IH-I3C CP/MAS using
hexamethylbenzene. Chemical shifts of 27Al are given in parts
per million from external Al(H2O)p and the shifts for 29Si and
I3C from external tetramethylsilane (TMS).
ESCA
Measurements.
ESCA measurements were carried out
using a Vacuum Generators ESCALAB system at the University
of Wisconsin-Milwaukee in the conventional mode with A1
Ka
X-rays. Powdered samples were either sprinkled onto double-
sided adhesive Scotch tape and the excess shaken off or pressed
as a wafer onto a thin indium foil. The tape or the foil was then
mounted onto a stainless steel specimen stub.
No
problems with
differential charging were encountered
(see
below). In order to
remove charging shifts and deal with Fermi edge coupling
problems, binding energies were scaled against the C( 1s) peak
(set to 284.6 eV) of the alkane part of the ethylene glycol retained
in the sodalite cage and also against any adventitious car-
bon.6,10.11
,I
3
0
1994 American Chemical Society
Framework Silicates with Sodalite Structure
The Journal
of
Physical Chemistry, Vol.
98,
No.
4, 1994
1303
TABLE 1: Core-Level Binding Energies and
Line
Widths
(in
Parentheses) in eV, Based on C(ls) at
284.6
eV.
Also
Included are Select Values
for
the Width at Half Maximum
of Various Valence Bands
’H
-
29Si
CP
1
MAS
purely
siliceous sodalite
-40
-60
-80
-100 -120 -140 -160
(b)
spinning
sample
-40
-60
-60
-100
-120
-140
-160
ppm
from
TMS
Figure
1.
1H-29Si CP spectra
of
purely siliceous sodalite: (a) static
sample, (b) with MAS.
1H-13CCPIMAS
purely
siliceous
sodalite
64
19
ppm
1
65
80
75
70
65 60 55
50
45
40
ppm
from
TMS
Figure
2.
lH-l3C CP/MAS spectrum
of
purely siliceous sodalite.
Results
and Discussion
Purely
Siceous
Sodalite.
The 29Si CP/MAS spectrum of
purely siliceous sodalite
consists
of a single very sharp single peak
at -1 17.3 ppm from TMS (see Figure lb). Even the spectrum
of a static sample (Figure la) is relatively narrow and symmetric,
reflecting the strictly cubic environment of the silicon atoms. The
IH-l3C CP/MAS spectrum (Figure
2)
consists of a single sharp
peak at 64.19 ppm corresponding to the encapsulated ethylene
glycol, indicating that all glycol molecules are located
on
equivalent
sites in the structure.
The ESCA results are given in Table 1 and Figures 3-7. Also
given in Table 1 is a representative set of results for several
material Si(2p)
O(
1s)
valence band A1 (2p)
Si02 (silica) 103.5 532.9 10.4
(1.8) (1.9)
purely siliceous sodalite 103.2 532.45 10.4
(2.6) (2.2)
zeolite
EM-5
103.1 532.45 10.4 14.5
zeolite Na-Y 102.55 532.0 10.2 14.4
zeolite Na-X 102.0 531.1 9.8 13.95
(1.8) (1.65) (1.6)
zeolite Na-A 101.1 530.5 9.1 13.5
(1.7) (1.75) (1.6)
montmorillonite 102.75 532.0 10.2 74.8
(2.05) (2.4)
(2.0)
kaolinite 102.45 531.5 10.2 14.3
(2.3) (2.45) (2.2)
(1.7) (2.05) (1.55)
aluminosilicate sodalite 101.5 530.9
a
13.6
Line width obscured by interference from
In(4d).
(b)
aluminosilicate sodalite
350
300 250
200
150
100
50
0
(a)
purely
siliceous sodalite
200
160
120
80
40
0
Binding energy (eV)
Figure
3.
(a) ESCA survey scan of purely siliceous sodalite; (b) ESCA
survey scan
of
aluminosilicate sodalite confirming a significant presence
of
C,
K,
and Mg.
reference materials. A “survey” ESCA scan for the purely
siliceous sodalite is shown in Figure 3a, and Figure 4 documents
the C( 1
s)
region, demonstrating that the principal elemental peaks
detected are due to Si and
0
and that substantial amounts of C
and moderate amounts of Na are also present. The carbon signal
comes from the glycol molecules and the adventitious carbon
species which are always present
on
the outer surface of air-
exposed material~.~J~ The intense peak at
288.5
eV is the C(1s)
signal primarily from the hydrocarbon parts of these species (with
a charging shift of 3.9 eV). The size and shape of the shoulders
in this C( 1s) spectrum indicate the presence of various types of
carbon-xygen units, most of them C-OH groups of the ethylene
glycol. Thevarious sodiumpeaks indicate that someof the sodium
1304
The Journal
of
Physical Chemistry,
VoI.
98.
No.
4, 1994
(b)
aluminosilicate sodalite
c
(1s)
.-
0
E
2
(a) purely siliceous sodalite
-
295 290 285 280 275
Binding energy (eV)
Figure
4.
C(1s) region
of
a high-resolution ESCA spectrum
of
purely
siliceous sodalite.
I\
(b)
aluminosilicate
sodalite
108 106 104 102
100
98 96 94 92
90
Binding energy (eV)
Figure
5.
Key Si(2p) peak binding energies adjusted to C(1s)
=
284.6
eV.
(which is abundant in the aluminosilicate sodalite) is retained in
the purely siliceous material. The sizes of these peaks indicate
that the amount of sodium is small. On the basis of the binding
energies, we attribute the majority of this sodium to residual
NaOH. This is supported by the absence
of
any other anions.I4
The Si(2p) and
O(1s)
spectra in Figures
5
and
6
are the
ESCA
“signatures” of the purely siliceous sodalite. This conclusion is
based on (i) the relative narrowness and the Lorentzian-Gaussian
shape
of
the lines, (ii) their relative intensities, (iii) the binding
Herreros et al.
r
18
536 534 532 530 528 526 524 522 520
Binding energy (eV)
Figure
6.
Key
O(1s)
peak binding energies adjusted to C(1s)
=
284.6
eV
.
20
15
io
b
0
Binding
energy
(eV)
Figure 7.
Valence band ESCA spectra: (a) Purely siliceous sodalite;
(b)
a-silica; (c) aluminosilicate sodalite; (d) zeolite Na-A. The dotted line
removes the In(4d) satellite peak, providing a more accurate rendition
of
the true valence band spectrum.
energies (see below), and (iv) the nature of the corresponding
valence band features (see Figure
7).
The widths and shapes of
the peaks are those of a singular species. The one exception is
in the
O(1s)
spectra because of the presence
of
a small amount
of NaOH. However, NaOH is not detected in the
O(1s)
ESCA
spectrum because its
O(
1s)
binding energy
is
apparently almost
identical to those for the purely siliceous sodalite, resulting in a
O(
1s)
line which is narrow (2.2 eV) and symmetric.
Framework Silicates with Sodalite Structure
The Journal
of
Physical Chemistry, Vol.
98,
No.
4,
1994
1305
aluminosilicate sodalite is not unexpected since the %/A1 ratio
of the latter is higher. As expected, the shift is to higher Si(2p)
and Al(2p) binding energies for the sodalite (see Table
1
and
Figures
5
and
6).
Theother notable feature is that the core-level binding energies
for aluminosilicate sodalite are not at all similar to those for
kaolinite (see Table
1).
In this case, ESCA is apparently sensitive
to thestructuraldifferences (frameworkversus sheet) ofthe species
as well as to the Si/Al ratio.
Some of the other features of these ESCA results become
apparent from the analysis of the binding energies of the core
lines and of the features of the valence band.8 Thus, we see from
Table 1 and Figures 5 and
6
that the binding energies of Si(2p)
and
O(
1s)
lines from the purely siliceous sodalite are almost the
same as from pure silica.6 The slightly lower values for sodalite
may be due to the different structures of the two materials. It has
been our contention that there are readily detectible shifts in the
binding energy positions for different and that these
shifts are primarily due to the nature of the immediate chemical
bonding environment adjacent to the Si-0 bonds.9 Thus, for
example, as more A1-0 units are inserted into a silica-type
framework, the binding energy of the Si(2p) peak progressively
decreases to reflect the enhanced covalency induced into the Si-0
bonds by the more ionic A1-0 unit~.~J~ As a result, the Si(2p)
and
O(
1s)
binding energies in the purely siliceous sodalite and
silica are higher than those in ZSM-5.6,8 We note from Table
1
and Figure 7a that the structure and character of the valence
band for the purely siliceous sodalite are very similar to those for
silica (see Figure 7b).
Aluminosilicate Sodalite.
Studies of many zeolites and clays
allow
us
to suggest that the binding energies and valence band
patterns for aluminosilicate sodalite should be very similar to
those for zeolite Na-A (see Figure
7
and Table 1). This is based
on
the shift of ESCA binding energies and valence bands for
similar system~.~J~ These shifts are found to reflect the localized
bonding chemistry in each structural type. Thus the ESCA
features of aluminosilicate sodalite and zeolite Na-A, two
frameworksilicates with Si/Al
=
1
which obey the “Loewenstein
rule” (which forbids AI-0-A1 linkages), should be very similar,
while kaolinite, a clay silicate also with Si/A1
=
1
but built of
octahedral aluminate and tetrahedral silicate units, should give
substantially different binding energies (see Table
l).6,7,9
The
structural difference between zeolite Na-A and sodalite may not
be resolved by ESCA, except perhaps as subtle shifts at very high
resolution. At the same time, we note that if all the aluminate
units in framework silicates are located in tetrahedral environ-
ments, essentially substituted for some of the silicate units (and
obeying the Loewenstein rule), then both the A1 and Si binding
energies will progressively increase as the Si/Al ratio is
in~reased.~-~ Structural variations have been found to influence
these binding energies, but far less than the change from silicate
to aluminate bonding chemistry within the same structural unk6
The binding energies for aluminosilicate sodalite should therefore
be similar to those of zeolite Na-A, differing only slightly due to
the structural differences, just as is the apparent result in
comparing the ESCA results for the purely siliceous sodalite with
silica.
The sample of aluminosilicate sodalite contained OH- instead
of the C1- which is normally present in the sodalitecages.2 ESCA
detected
no
evidence of the presence of chlorine. The type of
anion has
no
effect
on
the sodalite framework2 and thus does not
influence the critical ESCA shifts. We note that the relative
quantification of the Si(2p) to Al(2p) and the Na(2s) to the
Al(2p) in aluminosilicate sodalite also agrees with the result of
chemical analysis (Si/Al
=
1.05)
and suggests that the Na/A1
ratio is slightly lower than unity. The latter feature is explained
by the presence, detected by ESCA, of
K+
and Mg2+ cations (see
Figure 3b).
The core-level binding energies andvalence band patterns from
aluminosilicate sodalite (see Figures 5-7c and Table
1)
shift
dramatically with respect to the corresponding features for the
purely siliceous structural analogue. Substantial shifts of the
core lines of the aluminosilicate sodalite also occur with respect
to the silica, whereas the key binding energies for the former are
almost equal to those for zeolite Na-A (see Table
1
and Figures
5
and
6).6
In addition, thevalence band pattern for aluminosilicate
sodalite is also almost identical to that for zeolite Na-A (see
Figure 7c and d).8 The fact that there are small differences
between the binding energy values for zeolite Na-A and
Conclusions
ESCA core-level shifts and valence band patterns of sodalite-
type framework silicates are summarized in Table
1
and Figures
5-7.
These show the following:
(1)
The purely siliceous sodalite gives a unique set of binding
energies and valence band patterns which resemble those for silica.
The small variations may be indicative of the detection by the
ESCA of the structural differences.
(2) The ESCA core-level binding energies and the valence
band pattern for the purely siliceous sodalite and the alumino-
silicate sodalite are very different.
(3) The ESCA patterns of aluminosilicate sodalite and zeolite
Na-A are similar. The small difference in binding energies may
be due to the presence of
K+
and Mg2+ and the higher Si/Al ratio
of the aluminosilicate sodalite.
(4)
In
some cases ESCA is as sensitive to structure as it is to
composition. Thus, while zeolite Na-A and aluminosilicate
sodalite form
a
common binding energy set (with Si/Al-
1
and
all tetrahedral Si-0 and A14 alternating according to the
Loewenstein rule), their ESCA patterns differ substantially from
that of kaolinite which also has Si/Al=
1
but in which all A1-0
units are in octahedrally-bonded subsheets linked via apical
oxygens to tetrahedral Si-0 subsheets.13
This study of the surface/near-surface chemistry of sodalites
leaves many aspects unexamined. Although we have demon-
strated that ESCA may differentiate between substantially
different members of this group, there are other members of the
sodalite-ultramarine family the differences between which are
much more subtlee2 An ESCA study of these materials will be
described in a future paper.
Acknowledgment.
We are grateful to Dr.
P.
J.
Barrie,
University College, London, for acquiring an NMR spectrum,
the Fulbright Commission for a Professorial Fellowship for T.L.B.,
the Oppenheimer Fund for a Research Studentship for
H.H.,
and Unilever Research, Port Sunlight, for a Research Studentship
for B.H.
References and Notes
(1)
Thomson, Sir William.
The London, Edinburgh, and Dublin
Philosophical Magazine and Journal
of
Science
1881,
24 (5th series),
503.
(2)
Deer, W. A.; Howie,
R.
A.; Zussman,
J.
An Introduction to the
Rock-Forming Minerals,
2nd
ed.; Longman: London,
1992;
pp
496-502.
(3)
Barrer, R. M.
Hydrothermal Chemistry
of
Zeolites;
Academic
Press: London,
1982.
(4)
Bibby, D. M.; Dale, P.
Nature
1985,
317,
157.
(5)
Richardson,
J.
W.;
Pluth,
J.
J.;
Smith,
J.
V.;
Dytrych,
W.
J.;
Bibby,
D. M.
J. Phys. Chem.
1988,
92,
243.
(6)
Barr, T. L.
In
Practical Surface Analysis,
2nd
ed., Briggs, D., Seah,
M. P., Eds.; John Wiley: Chichester, U.K.,
1990;
Chapter
8.
(7)
Barr,
T.
L.
Appl. Surface Sci.
1983,
15,
1.
(8)
Barr,T. L.:Lishka, M.A.J.
Am. Chem.Soc.
1986,108,3178.
Barr,
T. L.; Chen, L. M.; Mohsenian, M.; Lishka, M. A.
J. Am. Chem. SOC.
1988,
110, 7962.
(9)
Barr, T. L.
Zeolites
1990,
10,
760.
(10)
Barr, T.
L.;
Klinowski, J.; He,
H.;
Alberti, K.; MOller,
G.;
Lercher,
(11)
He, H.; Alberti, K.; Barr, T. L.; Klinowski,
J.
J. Phys. Chem.,
in
(12)
Rocha,
J.;
Klinowski,
J.
J. Magn. Reson.
1990,
90,
567.
(13)
Barr, T.
L.
Modern ESCA: The Principles and Practice
of
X-ray
Photoelectron Spectroscopy;
CRC Press: Boca Raton,
FL,
in press.
(14)
Briggs, D., Seah, M. P.,
Eds.
Practical Surface Analysis,
2nd
ed.;
John Wiley: Chichester, U.K.,
1990.
(15)
Barr,
T.
L.
J. Vac. Sci. Technol.
1991,
A9,
1793.
J.
A.
Nature
1993,
365,
429.
press.
... An increase in the BE of Al 2p from 73.7 eV to 74.5 eV, with a simultaneous decrease of the Si 2p BE from 103.7 eV to 101.9 eV, indicates the formation of the Al−O−Si bonds. Since Al is more electropositive than Si, 48 presence of Al−O units in the silica network decreases the BE of Si 2p and increases that of Al 2p. 49 In fact, the measured BEs are very close to the BEs of aluminosilicates, 50 further confirming the formation of the Al−O−Si bonds. ...
... The decrements in O1s intensity and increment in Al2p intensity also indicate the effective desilication or decrease in Si/Al ratio (Barr, 1990). The progressive decrements in binding energy for all elements also signify that aluminate units are increasing in tetrahedral environments ( Herreros et al., 1994 (2e4 h) in the beginning of the run. After that period, quinoline, TOC and H 2 O 2 conversions reach to steadystate. ...
... Although no specific XPS study has been performed on iodosodalite to the best of our knowledge, studies on zeolites have shown that different structures resulted in shift in binding energies for Si-O. 61 The measured binding energies of sodalite samples (Na, Al, Si, O, I) compared with published data are summarized in Table 2 (spectra are shown in Supplementary). The measured XPS result is valid for confirming the oxidation state of iodine as there is a significant difference (~5 eV) for the binding energies between iodide and iodate. ...
Article
The effects of six process variables were investigated on the hydrothermal growth of iodosodalite, Na8Al6Si6O24I2: pH (NaOH concentration), aging time, temperature, Al/Si ratio, precursors used (ie, zeolite 4A, kaolinite, meta-kaolin, colloidal silica, and sodium aluminate), and precursor concentration. Powder X-ray diffraction (XRD) with Rietveld refinements, X-ray photoelectron spectroscopy (XPS), Fourier transform infrared spectroscopy (FTIR), and scanning electron microscopy (SEM) were performed to characterize the structures, phase fractions, chemical state, and surface morphology of the synthesized products. Iodosodalite yield increased as aging time and pH increased. The crystallization of iodosodalite was favored in the temperature range 140°C-180°C. Decreasing the Al/Si ratio by half increased the crystallization of basic cancrinite. Lowering the precursor concentration by adding water revealed the crystallization of nepheline hydrate I and a decrease in the sodalite fraction. Among the tested precursors, zeolite 4A yielded the highest mass fraction of iodosodalite in the synthesized powders. From the aging time and temperature variation experiments, the phase transformation of zeolite Asodalitecancrinite was observed. XPS and FTIR results showed the presence of only iodide but not iodate in the synthesized product. The crystallization of various minerals suggests that mechanisms for transport of the ions and formation of the aluminosilicate frameworks vary with hydrothermal conditions.
Article
More efficient soil remediation technologies are highly anticipated to treat large quantities of heavy metal-polluted urban sites nowadays. Herein, a novel hydrothermal technology of converting heavy metal-polluted soils into zeolites for in-situ immobilizing heavy metals was proposed. The zeolites (analcime and cancrinite) could be synthesized hydrothermally with certain Na/Si and Al/Si ratios. The formed zeolites could manage to change their species and structure during zeolitization to accommodate different heavy metals in soil according to their size and charge. Since smaller-size Cu²⁺ was introduced, analcime and some cancrinite possessing small cages could be formed adaptively to immobilize the Cu²⁺ by replacing Na⁺ and forming Cu²⁺-OH and Cu²⁺-O. Whereas, cancrinite with large channels managed to form to immobilize the larger-size Cd²⁺ by forming Cd²⁺-O. Interplanar spacing variation of zeolites also corresponded to their structural change for accommodating different heavy metals. Leaching results showed the amounts of Cu and Cd leached from the synthesized zeolites were reduced to 0.005% and 0.05% respectively, reflecting a more stable immobilization of smaller heavy metals by small cages, in agreement with the results of distribution coefficient (Kd). Negligible effect of pH environment on the leaching rates further confirmed the stable structural immobilization of heavy metals by zeolites.
Article
Full-text available
Highly ordered short-channel mesoporous silica SBA-15 with large pores (11.2 nm) has been synthesized by using tetramethyl orthosilicate (TMOS) as silica source, the amphiphilic block copolymer Pluronic PE-10400 as structure-directing...
Article
Developing new sodium‐ion battery (SIB) cathode materials with low cost and high capacity remains a challenge. Here, NaVSi2O6 was successfully synthesized by a simple sol‐gel method for the first time. This material was used as a new SIB cathode material. At a current density of 20 mA g−1, the first discharge specific capacity is 80 mA h g−1, which exhibits a higher capacity. The mechanism of the cycle stability of NaVSi2O6 was elucidated by the theoretical calculations. The volume and formation energy change of different content of Na+ insertion/extraction, as well as the energy band and density of states were calculated combined with the density functional theory (DFT). Our research results provide a reference for further exploring the NaMSi2O6 (M = Fe, Cr, and V) compounds as cathode materials of SIB. NaVSi2O6 was successfully synthesized by a simple sol‐gel method for the first time. This material was first used as a new SIB cathode material. Our research results provide a reference for further exploring NaMSi2O6 (M = Fe, Cr, and V) system as cathode materials of SIB.
Article
Full-text available
Gold nanoclusters and isolated gold atoms have been produced in a two-liquid phase procedure that involves a solution of gold in aqua regia and rosemary essential oil as organic layer. These gold entities have been immobilized on the ordered mesoporous silica material SBA-15 functionalized with different amounts of aminopropyl groups. The resulting materials have been characterized by XRD, N2 adsorption, chemical analysis, TGA, 29Si MAS NMR, 13C CP/MAS NMR, UV-vis spectroscopy, XPS, and STEM. The Au-containing materials retain the ordering and porosity of the pristine support. Gold content varies in the range of 0.07–0.7 wt% as a function of the specific immobilization conditions, while STEM evidences the presence of isolated gold atoms. XPS shows a shift of the Au 4f BE toward values lower than those of metallic gold. The catalytic activity in the oxidation of cyclohexene with molecular oxygen at atmospheric pressure parallels the Au content of the aminopropyl-SBA-15 supports. This activity is higher than that of analogous Au entities immobilized on SBA-15 functionalized with thiol or sulfonate groups, the activity decreasing in the order Au-NH2 > Au-SO3− > Au-SH. This behavior has been attributed to differences in the interaction strength between the functional group and the Au entities, which is optimum for the aminopropyl groups.
Article
Full-text available
Alkali–silica reaction (ASR) is an ongoing problem that causes damage to concrete constructions and reduces their durability. Therefore, minimizing this undesired reaction is of great interest for both safety and economic reasons. Additives containing high aluminium content are very effective in reducing the release of silica and enhancing the durability of concrete; however, the mechanism for this effect is still under discussion. In this study, an enrichment of aluminium in the near‐surface region was observed for natural quartzite rock after storage in Al (OH)3 and metakaolin as aluminium sources, from which we conclude that the formation of aluminosilicate sheets of a few nanometres inhibits the silica release; this hypothesis is supported by high‐resolution spectra of Al 2p, Si 2p and O 1s.
Article
This article reports a green strategy to synthesize Ti-ZSM-5 zeolites directly from a titanium-containing natural rectorite mineral, with particular emphasis on the transformation and crystallization behaviors of titanium species during the rectorite activation and the followed crystallization of the activated rectorites. The results show that: when taking the alkali-fusion and submolten salt activated rectorites as the starting materials, Ti-ZSM-5 zeolites with moderate acidity and superior hydroisomerization and aromatization performance are successfully obtained, in which part of Ti are isomorphically incorporated into the framework of the resulting ZSM-5 zeolites; whereas when taking the raw and thermally activated rectorites as the starting materials, ZSM-5 zeolites with TiO2 existing as impurities rather than as framework Ti atoms are obtained. This difference is attributed to the different chemical states of the titanium species in different activated rectorites. The approach reported here points a green avenue to directly synthesize heteroatom-containing zeolites from natural clay minerals.
Article
In this work we report the results of the crystal chemical, structural and surface characterization of erionite-K fibres from Rome (Oregon, USA) after interaction with Fe (III) chloride solutions at different concentrations. In addition, Fe (III) loaded samples were investigated after incubation in ascorbic acid in order to monitor the mobility of reduced Fe (II) and to highlight its possible incorporation as EF cation through ion exchange. Comparison between released and acquired charges under the form of Fe confirms, in perfect agreement with previous studies that Fe (III) is mainly fixed at the fibre surface. Nevertheless, in very diluted Fe (III) solutions (below 50 μM FeCl3) a significant fraction of Fe (III) is segregated by an ion-exchange mechanism in the erionite cavity at the Ca3 site, albeit with a significantly lower efficiency with respect to Fe (II). It is worth mentioning that, as a result of the catalytic properties of zeolites, the location of iron in well-defined crystallographic positions is the prerequisite for behaving as a very active site in the generation of reactive oxygen species. Incubation in ascorbic acid revealed that only Fe (III) residing at the fibres surface and characterized by low nuclearity is significantly reduced, whereas this reaction does not occur (or possibly occurs very marginally) in the case of the ion-exchanged metal. Considering that the total iron in lung fluids occurs at very low concentration (ca. 0.21 μM), our results strongly suggest that the physiological environment unfortunately represents the optimum condition for iron to behave as a very active site.
Article
The zeolites not only provide an outstanding example of geometrical patterns among the framework silicates, but also in their permanently porous and rigid nature are of great commercial interest. This book collates the current knowledge ragarding zeolite synthesis, isomorphous substitution and chemical transformation. The principal features of hydrothermal systems and reactions are outlined, and the extension of zeolite synthesis to compositions unknown in nature and with unusual topologies is described.-R.A.H.
Article
This paper is focused upon XPS studies of silicon in a number of diverse environments. First, the XPS analysis of Si in its “clean” zero valent form is reported. Novel observations of the corresponding XPS loss spectra are presented. Second, the changes in these XPS spectra during oxidation of Si0 to SiO2 are described. In addition to loss spectra, charge shift and valence band analyses are reported. Third, the changes experienced by silica are documented when metal aluminates are injected to create such common adsorbents as clays and zeolites. Contrary to previous suppositions these systems do exhibit “chemical shifts”, but these shifts are apparently motivated by group, rather than elemental, changes. As a fourth example, unique XPS analyses are reported of the changes experienced when silica is employed as a support for a metal oxide catalyst during on-line use situations. Finally, the study returns to elemental silicon by describing thin ( ∼ 500 Å) films of SiO2 on Si0. Contrary to previous contentions, it is shown that by utilizing loss spectra and charge shift analyses, a description of the SiO2 film, the resulting interface, and the Si0 underlayer may be achieved employing common ion-sputter etching. These studies demonstrate the interconnection of XPS techniques and results over fields that often seem totally divorced from one another.
Article
ZEOLITES derive their catalytic activity from the strong acidity of protons attached to the negatively charged aluminosilicate frame-work, which makes the materials excellent proton donors. Unlike zeolites, the aluminophosphate molecular sieves1,2 are built from alternating AlO− 4 and PO+ 4 tetrahedra and are thus electrically neutral. Much attention has therefore been devoted to the generation of Bronsted acidity in these materials by introducing heteroatoms, such as Si, Mg, Fe, Co or Zn, to produce negatively charged frameworks3–6. Similar arguments apply to gallophosphate molecular sieves7–10, of which cloverite9,10 is a remarkable example. This extra-large-pore material contains pore openings in the form of a four-leafed clover, defined by a ring of 20 gallium and phosphorus atoms, some of which are linked to terminal hydroxyl groups. Here we use NMR, X-ray photoelectron spectroscopy (ESCA) and infrared spectroscopy to show that the P–OH groups in cloverite are localized versions of those in solid phosphoric acid, H3PO4. Cloverite is thus a strong Bronsted acid even though no heteroatoms are present in its framework.
Article
The aluminosilicate sodalite has been known since antiquity1. It occurs naturally as the polysulphide-containing mineral lapis lazuli1 and as a salt-bearing feldspathoid2 and has also been synthesized as the zeolite basic sodalite with the formula Na6(Al6Si6O24) · xNaOH · (8−2x)H2o (ref. 2). The structure is an open framework of relatively large cages which are accessible, to water molecules and some ions, through relatively small windows of ~2.1 Å free diameter. Large ions or molecules may be trapped in the cages during synthesis. In aluminosilicate sodalite, the cages contain the number of cations required to compensate for the lattice charge, together with varying amounts of H2o and salts such as NaOH and NaCl. Aluminosilicate sodalites usually have an Al/Si ratio of about unity although alumina analogues of sodalite have been synthesized3 and a silica-rich sodalite (TMA sodalite) has been prepared from aqueous hydrothermal systems4,5 in the presence of tetramethylammonium ions. We report here a new method for the synthesis of sodalite, either in a silica-rich aluminosilicate form or in a novel pure-silica form. Our method of synthesis of sodalite is unusual in that the solvent system is essentially non-aqueous. In the absence of alumina, a new material is produced, namely a pure-silica form of sodalite which, following previously published suggestions6, we call silica-sodalite. If the ‘solvent’ system is ethylene glycol then the unit cell composition approximates to Si12O24 · 2C2H4(OH)2.
Article
It has been demonstrated that the constituents on the surfaces of various pure zeolites, clays, silicas, and aluminas yield reproducible ESCA peaks with unique binding energies. The collective patterns realized by these different binding energies strongly suggest the registration of selective group rather than elemental chemical shifts. The two primary chemical groups identified in many of the zeolites seem to be a unit that resembles SiOâ and another that mimics N/sup + +/-AlâOâ/sup 2 -/ (where N symbolizes the cations, usually alkali or alkaline earth species that balance the aluminate charge), a feature supported by quantum calculations, relatively narrow line widths, and reproducible valence band and loss data. These results have been employed to determine the relative purity of the surface region of different conventionally prepared zeolite systems. High-resolution ESCA studies of mordenite, ZSM-5, a silicate surfaces always seem to exhibit Al(2p) spectra that are significantly broadened into patterns that suggest several Al-containing species. The nonzeolites present at the surface of these systems generally constitute more than 50% of the total aluminum. The primary impurity species have been identified as metal aluminates and aluminas in differing ratios apparently depending upon the cation (e.g., sodium) concentrations. The possible presence of silica on the surface of some freshly prepared zeolite systems is suggested, but unconfirmed.
Article
The first detailed, reproducible, experimental results are presented of the valence band spectra of a variety of commercially important zeolites. These were obtained using X-ray photoelectron spectroscopy (XPS or ESCA). A general chemical formulation of (SiO2)x ·(M+p1/pAlO-2)y·ZH2O applies to the materials examined, where, for most of the cases studied, M+p=Na+. Comparisons are made with well-known experimental and theoretical (MO and band structure) results for various silicas and aluminas. Numerous interrelationships are found involving most of the key subband features documented for the latter materials. Distinct shifts and truncations are found in the zeolite bands. These suggest that zeolites more closely depict an agglomeration than a persistent type mixture of the aforementioned precursors.
Article
The crystal structure of silica sodalite, including possible locations for encapsulated ethylene glycol, was determined at room temperature by using a combined single-crystal X-ray and powder neutron diffraction analysis. Unit cell composition: Si/sub 12/O/sub 24/ x 2C/sub 2/H/sub 4/(OH)/sub 2/, M/sub r/ = 845, cubic, Im3m, a = 8.830 (1) A (X-ray), a = 8.8273 (1) A (neutron). These refinements reveal that the correct space group for silica sodalite is Im3m (rather than I43m) and, therefore, that the sodalite framework is fully expanded. At room temperature, each sodalite cage contains one ethylene glycol molecule which has a range of geometrical positions. Intramolecular distances for the ethylene glycol molecule are (X-ray) C-C 1.78 (4), C-O 1.30 (6) A, (neutron) C-C 1.70 (3), C-O 1.25 (3) A. The shortest distances (3.4 A) between oxygens of the framework and molecule are consistent with weak hydrogen bonding. From the neutron diffraction data it was found that a sluggish phase change from cubic to lower symmetry occurs upon cooling below 200 K. At 10 K, silica sodalite appears to be monoclinic with approximate cell parameters a = 12.250 (8) A, b = 12.471 (8) A, c = 8.512 (6) A, ..beta.. = 91.37 (6)/sup 0/, based on an indexing of 12 peaks, but the precise symmetry is as yet unknown.