Local and Average Structural Changes in Zeolite a Upon Ion Exchange
Local and Average Structural Changes in Zeolite a Upon Ion Exchange
Article
Local and Average Structural Changes in Zeolite A
upon Ion Exchange
Lisa Price 1 , Ka Ming Leung 2 and Asel Sartbaeva 1, * ID
Abstract: The infamous ‘structure–property relationship’ is a long-standing problem for the design,
study and development of novel functional materials. Most conventional characterization methods,
including diffraction and crystallography, give us a good description of long-range order within
crystalline materials. In recent decades, methods such as Solid State NMR (SS NMR) are more widely
used for characterization of crystalline solids, in order to reveal local structure, which could be
different from long-range order and sometimes hidden from long-range order probes. In particular
for zeolites, this opens a great avenue for characterization through studies of the local environments
around Si and Al units within their crystalline frameworks. In this paper, we show that some
structural modifications occur after partially exchanging the extraframework Na+ ions with
monovalent, Li+ , K+ , Rb+ and NH4 + and divalent, Ca2+ cations. Solid state NMR is deployed to
study the local structure of exchanged materials, while average stricture changes can be observed by
powder diffraction (PXRD). To corroborate our findings, we also employ Fourier Transform Infrared
spectroscopy (FT-IR), and further characterization of some samples was done using Scanning Electron
Microscopy (SEM) and Energy-Dispersive X-ray spectroscopy (EDX).
1. Introduction
Zeolites are aluminosilicate porous minerals. Many zeolites occur in nature as aluminosilicate
minerals. To date, we can make more than 200 synthetic zeolites in the laboratory [1,2]. They are
classed as porous materials as they possess cages, channels and open void spaces within their highly
crystalline frameworks. Each zeolite framework has a unique structure, and because there is such a
variety of zeolite structures, there is also a very wide diversity of zeolite applications. Synthetic zeolites
are used as green, re-usable catalysts in industrial processes as heterogeneous catalysts for processes
that involve hydro-cracking, acrylation, oxidation and reforming. Most zeolite syntheses employ
Organic Structure Directing Agents (OSDAs), such as TMA-OH (Tetramethylammonium Hydroxide)
or crown-ether, which act as templates to guide the formation of particular types of zeolite pores and
channels [3–6]. This reduces the chance of producing competing zeolite phases. However, due to
the high manufacturing costs of producing these organic materials, which cannot be recovered after
calcination, current research is becoming more concerned with optimising synthesis conditions in
order to produce pure zeolites in the absence of OSDAs [3,7]. Here, we performed a low temperature
synthesis of small zeolite A (Na-A) crystals without the use of OSDAs, and the corresponding LTA
framework (Linde Type A) is shown in Figure 1.
Figure 1. Linde Type A (LTA) framework with the α cage as the unit cell in the middle (black square).
Red spheres: oxygens; blue: Si/Al atoms.
Na-A is a commercially important zeolite used in industry for catalysis, adsorption and industrial
gas separations [8]. More recently, its sustainability as a drug delivery system has been investigated [9].
One of its greatest applications, however, is ion exchange; in particular, rapid Na+ /Ca2+ exchange.
Consequently, Na-A is very effective in water softening, and one of its main functions is in washing
powders as a detergent builder [10].
Ion exchange can also produce zeolites with different properties. For example, K-A is commonly
used in the ethanol drying processes [11] and partially-exchanged K/Na-A is used to separate CO2 from
CO2 /N2 dry mixtures [12]. Ca-A zeolites are important in industry, where they selectively adsorb linear
alkanes from a mixture of branched alkanes [13], and Na+ /NH4 + exchange is useful in minimising
environmental pollution and eutrophication [14–17]. Li-A was proposed as a possible delivery material
for pharmacological studies [18]. Li-exchanged zeolites are also used for the separation of nitrogen
from air [19]. In this work, we carried out aqueous ion exchange of monovalent alkali metals: Li+ ,
K+ , Rb+ and NH4 + and divalent Ca2+ into zeolite A crystals produced from a low temperature and
organic template-free synthesis and performed analysis using solid state NMR and other methods to
determine the structural effects of the exchange. Previously, solid state NMR has been used to study
other zeolites [20–28].
Figure 2. PXRD pattern for Na-A synthesised at 40 ◦ C for 24 h, indexed as the LTA framework.
Figure 3. PXRD pattern for Na-A synthesised at 40 ◦ C for 48 h. Characteristic Faujasite (FAU)
framework peaks are indexed with black dots.
In this investigation, we decided to study all samples after one ion exchange only. The extent of
exchange was quantified from the filtrates using a sodium ion selective electrode (ISE). For each sample,
even after one exchange, we see good, although not complete, ion exchange. This is not surprising as
sometimes as many as 8–10 steps are required for a complete exchange to occur [32]. Results show that
the extent of exchange decreases with increasing cation size, Li+ > K+ > Rb+ . The steric restrictions
of the zeolite pores make full exchange difficult to obtain, particularly for those ions with large ionic
radii. Energy Dispersive X-ray (EDX) elemental analysis was also carried out. In all samples, it was
evident that partial exchange had taken place, as residual Na+ ions were detected.
PXRD and FT-IR analyses show that there is no significant alteration to either the long-range
crystal order or the local framework structure of Na-A after exchange with Li+ , K+ and Rb+ ions.
These monovalent alkali cations vary in their ionic radii, and changes in the PXRD peak intensities
are expected to occur as a result of these cations occupying slightly different sites in the pores.
Figure 4 shows the PXRD patterns for the alkali metal exchanged zeolites. For K-A, the (4,4,0) reflection
almost disappears, whereas the (4,2,2) and (8,0,0) peaks increase in intensity. These results are in
agreement with those observed by Lührs et al., where complete exchange with K and Ca was studied
using diffraction and structure refinement [33]. The PXRD pattern for Rb-A is also slightly different.
The characteristic intensities of the first four Na-A reflections are altered; most noticeable is the increase
in the intensity of the (2,2,0) reflection.
Magnetochemistry 2017, 3, 42 4 of 16
The unit cell parameters (a) for the exchanged zeolites were calculated from the PXRD peak
positions and Miller indices using the program UnitCell [34] and are shown in Table 1. All samples have
cubic symmetry, and it can be seen that the unit cell size decreases by about 1% on exchanging larger
Na+ (1.02 Å) for smaller Li+ (0.59 Å) ions and increases slightly on exchange with larger K+ (1.38 Å)
and Rb+ (1.49 Å) ions [35]. Correspondingly, an increase in the lattice parameter is progressive from
Li < Na < K < Rb-A, in accordance with the increasing ionic radii of the monovalent cations.
The FT-IR spectra for Li-A, Na-A, K-A and Rb-A all display the fundamental zeolite framework
vibration νmax /cm−1 at 959, 968, 972 and 969, respectively, corresponding to the asymmetric ν(O−T −O)
stretch, shown in Figure 5. In addition, the framework symmetric ν(O−T −O) stretch occurs at ν/cm−1
684, 664, 663 and 666 for Li-A, Na-A, K-A and Rb-A, respectively. The broad peaks at 3350–3450 cm−1
and the weaker peaks at 1650 cm−1 observed in all spectra correspond to the ν(O− H ) stretching
and δ( H −O− H ) bending vibrations of water molecules in the hydrated samples. The only noticeable
differences in the spectra are that the framework ν(O−T −O) asymmetric stretch shifts to slightly a
higher wavenumber and the ν(O−T −O) symmetric stretch shifts to slightly a lower wavenumber, with
increasing cation size, Li+ < Na+ < K+ .
Magnetochemistry 2017, 3, 42 5 of 16
Figure 5. FT-IR spectra for Na-A (blue line) and cation-exchanged Li-A (orange line); K-A (grey line)
and Rb-A (green line).
For the parent Na-A zeolites, 29 Si and 27 Al MASNMR spectra contain one peak at δ −89.54 ppm
and δ 58.41 ppm, respectively, confirming that the Si/Al ratio of the framework is one. This is in
agreement with data reported by Thomas et al. [24] for a single silicon, Si(OAl)4 , and aluminium
environment, Al(OSi)4 , in Linde Type A zeolites. 27 Al MAS NMR spectra for the exchanged zeolites
also display one sharp peak between δ 57 ppm and 60 ppm, pertaining to Al(4Si) units. Likewise,
29 Si spectra for Li-A, K-A and Rb-A are also dominated by a single sharp peak at δ −87.23 ppm,
−89.83 ppm and −89.87 ppm, respectively. These peaks all lie within the chemical shift range for
which Si(4Al) units can occur (δ −80.0 ppm–−90.5 ppm from TMS) [28]. Some small, low intensity
peaks at δ −84.02 ppm, −87.96 ppm and −85.03 ppm are observed in the Li, K and Rb-A samples,
respectively. These are identified as silanol peaks. From deconvolution of the Li, K and Rb-A 29 Si NMR
spectra as shown in Figure 6, the Si/Al ratios were calculated to be one using Equation (1) [22,28]:
Si ∑nn= 4
=0 I Si (nAl )
= n =4 n (1)
Al ∑n=0 ( 4 ) I Si(nAl )
Figure 6. Cont.
Magnetochemistry 2017, 3, 42 6 of 16
Figure 6. 29 Si NMR spectrum of the as-synthesised Na-A and monovalent cation-exchanged Li-A, K-A
and Rb-A.
It is also interesting to note that the Si(4Al) peak for Li-A is significantly shifted to a lower field,
centred at δ −87.23 ppm, in comparison to that of the parent Na-A peak, at δ −89.54 ppm. The linear
relationship between the average Si-O-T framework bond angles, (α) and 29 Si chemical shifts can offer
an explanation for this difference. Table 2 shows the Si/Al ratios and average T-O-T bond angles that
were calculated from the deconvoluted 29 Si NMR spectra using Equation (2) [36].
Table 2. 29 Si NMR data: Si/Al ratios and average T-O-T bond angles for Li, Na, K and Rb-A.
The angular values are estimated based on NMR data using the regression relationship from [36], and
we estimate the accuracy on angles up to ±2 degrees.
Na-A zeolites, synthesised at 40 ◦ C for 24 h, are shown to have good cation exchange ability
with Li+ , K+ and Rb+ ions. The only noticeable differences between these exchanged zeolites are the
sizes of the unit cells and average framework T-O-T bond angles, which increase accordingly with
increasing cation size Li < Na < K < Rb.
2.2. NH4 -A
Exchange with NH4 + ions compromises some of the long-range order of the zeolite A crystals,
inferred by the broad PXRD peaks and low signal to noise ratio, as seen in Figure 7. NMR and
FT-IR spectra also indicate that the local framework environment is affected. Figure 8 shows a weak
peak at ν( NH ) /cm−1 1453 in the FT-IR spectrum, which confirms that exchange has taken place.
The asymmetric ν(O−T −O) stretch is, however, weaker and broader than that of the parent spectrum
and is shifted toward a higher wavenumber, occurring at νmax /cm−1 987 [37,38]. As Si-O bonds (1.64 Å)
are shorter than Al-O bonds (1.73 Å) [35], the force constant is higher for the former. Therefore, the shift
to higher frequencies indicates that a loss of some aluminium from the framework has occurred.
Figure 8. FT-IR spectra for the parent Na-A zeolite (orange line) and NH4 -A (blue line).
Magnetochemistry 2017, 3, 42 8 of 16
Furthermore, deconvolution of the 29 Si MAS NMR, as shown in Figure 9, confirms the presence
of both Si(3Al) and Si(4Al) environments in the framework, with peaks occurring at characteristic
chemical shifts of δ −91.41 ppm and −89.08 ppm, respectively. Another peak at δ −85.04 ppm is
present in the spectrum due to silanol species. This is in line with the appearance of the ν(T −O− H )
stretch at 868 cm−1 in the FT-IR spectrum. Using Equation (1), the Si/Al ratio of NH4 -A was calculated
to be 1.04. The loss of some aluminium from the framework is further confirmed in the 27 Al MAS
NMR spectrum, which displays a resonance signal at δ 1.29 ppm, attributed to extraframework
octahedrally-coordinated aluminium [23,28]. This peak is broad and overlaps with the main signal
at δ 58.89 ppm (for a tetrahedral Al coordination). The broadening and overlapping of these peaks
can be attributed to severely distorted six coordinated and four coordinated aluminium environments.
The presence of six coordinated extraframework Al species in NH4 -exchanged zeolite A has been
previously reported by Klinowski et al. [22,24], Sartbaeva et al. [15] and M. Dyballa et al. [39]. It is
evident that some dealumination of the zeolite A framework occurs upon exchange with NH4 + ions.
Figure 9. 29 Si (top) and 27 Al (bottom) NMR spectrum of the monovalent cation-exchanged NH4 -A.
2.3. Ca-A
There is no considerable alteration to the long-range crystal order of zeolite A after exchange
with Ca2+ . The increase in intensity of the (4,0,0) reflection and the decrease of the (4,4,0) reflection
is consistent with the results obtained by Lührs et al. [33]. The peak in the PXRD pattern at 2θ 29.57◦ ,
as shown in Figure 10, is characteristic of calcite (CaCO3 ), which must have formed from the Ca(OH)2
exchange solution. Further characterisation with 13 C Cross-Polarisation (CP) NMR (Figure 11) shows a
distinct peak centred at δ 168 ppm, which confirms the presence of CO3 2− [26]. Furthermore, the split
Magnetochemistry 2017, 3, 42 9 of 16
bands in the FT-IR spectrum at 1413 cm−1 and 1459 cm−1 identify the carbonate as a monodentate
species, as shown in Figure 12 [40].
Figure 10. PXRD pattern for Ca-A. The characteristic calcite peak is identified.
Figure 12. FT-IR spectra for the parent Na-A zeolite (orange) and Ca-A (blue).
Magnetochemistry 2017, 3, 42 10 of 16
Exchange with Ca2+ has a small effect on the local framework environment of zeolite A.
The asymmetric ν(O−T −O) stretch in the FT-IR spectrum is less intense and is shifted toward a higher
wavenumber, occurring at νmax /cm−1 999 [37,38]. The 29 Si MAS NMR spectrum, as shown in Figure 13,
is dominated by an intense peak at δ −91.41 ppm, due to the expected Si(4Al) units of the framework.
However, deconvolution of the spectrum also confirms the presence of Si(3Al), with a very small
peak occurring at δ −96.35 ppm. Another peak at δ −85.55 ppm is also present due to some silanol
species. Using Equation (1), the Si/Al ratio of Ca-A was calculated to be 1.01. Similar to the exchange
with NH4 + ions, it is evident that some dealumination of the zeolite A framework has occurred
upon exchange with Ca2+ ions. This loss of aluminium is confirmed in the 27 Al NMR spectrum,
which displays a broad peak centred at δ 50.11 ppm, characteristic of Al(4Si) units and also a slight
peak centred at δ 9.10 ppm, which can be attributed to a distorted octahedral aluminium environment.
The broadness of the peaks in the spectra can be attributed to quadrupolar interactions of 27 Al nuclei
in these distorted environments.
Figure 13. 29 Si (top) and 27 Al (bottom) NMR spectrum of the divalent cation-exchanged Ca-A.
2.4. SEM
The surface morphologies of the zeolite crystals were observed using Field Emission Scanning
Electron Microscopy (FESEM). Particle sizes were calculated from the scale bars on the FESEM
micrographs, using the ImageJ processing and analysis program. The crystallite sizes were also
calculated from the broadening of the most intense PXRD peaks, in this case the (6,2,2) and (6,4,4)
reflections, using the Scherrer Equation (3), where L is the crystallite size, β (hkl ) is the full width at half
Magnetochemistry 2017, 3, 42 11 of 16
maximum for the major peak of the PXRD pattern subtracting the instrumental contribution to the
broadening, K is the Scherrer constant, which is 0.9, λ is the wavelength of the X-rays in nm and θ is
the Bragg angle of the incident X-rays [41].
Kλ
L= (3)
β(hkl ) cos θ
Table 3 shows the average zeolite particle sizes. For Na-A and Ca-A zeolites, there is good
consistency between the sizes calculated from the Scherrer equation and FESEM data.
Figure 14 shows the FESEM micrograph for the parent Na-A zeolite. A variety of crystallite
shapes and sizes can be observed, somewhere in between cubic and spherical and ranging from
122–354 nm. The visibly larger crystals with well-defined edges display the typical cubic morphology
of LTA zeolites [42,43]. However, the low temperature conditions employed in this synthesis appear
to slow down crystal growth, instead favouring nucleation in the initial stages [44]. Particles that
are more rounded in shape and significantly smaller in diameter are evident in the FESEM images,
indicating that some of the crystals have not had enough time to form completely. These results are
in agreement with those reported by Dimitrov et al. [45] and Smaihi et al. [46]. The introduction of
Ca2+ into the zeolites is accompanied by noticeable changes in the surface morphology, as shown in
Figure 15.
Table 3. Crystallite sizes (nm) calculated from FESEM data using Equation (3) compared to crystal
size observed by FESEM. Statistical analysis shows a large spread of values from the Scherrer equation
up to ±28 nm; for FESEM data, the standard deviation is up to ±14 nm.
Na-A Ca-A
Scherrer Equation (nm) 242 270
FESEM (nm) 241 290
4. Conclusions
Solid state NMR revealed changes to the local structure of the LTA framework upon ion exchange
with NH4 + and Ca2+ . Exchange with Li+ , K+ and Rb+ ions does not significantly affect the long-range
crystal order. Exchange with NH4 + ions compromises some of the long-range order of the zeolite A
crystals due to the loss of some framework aluminium as can be seen from X-ray data. Exchange with
divalent Ca2+ ions introduces some monodentate carbonate species into the framework, but no
alteration to the long-range crystal order is observed. This study confirms that using a local probe such
as SS NMR alongside PXRD and other long-range methods to study zeolites can reveal an extra level
of information about the structure of those useful minerals, which will further their use as potential
catalysts and ion exchange materials.
Acknowledgments: A.S. would like to acknowledge the Royal Society for funding of URF. We thank John Mitchel
and Ursula Potter for help with collecting SEM data and Fraser Markwell for collecting SS NMR data at the
EPSRC-funded SS NMR facility at Durham University.
Author Contributions: A.S. conceived the study. K.M.L. designed experiments. L.P. and K.M.L. have collected
data and performed data analysis. All authors contributed to writing and editing the manuscript.
Magnetochemistry 2017, 3, 42 14 of 16
References
1. Baerlocher, C.; McCusker, L.B.; Olson, D.H. Atlas of Zeolite Framework Types; Elsevier:
Amsterdam, The Netherlands, 2017.
2. Sartbaeva, A.; Wells, S.A.; Treacy, M.M.J.; Thorpe, M.F. The flexibility window in zeolites. Nat. Mat.
2006, 5, 962–965, doi:10.1038/nmat1784.
3. Conato, M.T.; Oleksiak, M.D.; McGrail, B.P.; Motkuri, R.K.; Rimer, J.D. Framework Stabilization of Si-Rich
LTA Zeolite Prepared in Organic-Free Media. Chem. Commun. 2014, 51, 269–272, doi:10.1039/C4CC07396G.
4. Leung, K.M.; Edwards, P.P.; Jones, E.; Sartbaeva, A. Microwave synthesis of LTN framework zeolite with no
organic structure directing agents. RSC Adv. 2015, 5, 35580–35585, doi:10.1039/C4RA16583G.
5. Nearchou, A.; Raithby, P.; Sartbaeva, A. Systematic approaches towards template-free synthesis of EMT-type
zeolites. Microporous Mesoporous Mater. 2018, 255, 261–270, doi:10.1016/j.micromeso.2017.08.036.
6. Nearchou, A.; Sartbaeva, A. Influence of alkali metal cations on the formation of zeolites under
hydrothermal conditions with no organic structure directing agents. CrystEngComm 2008, 17, 2496–2503,
doi:10.1039/C4CE02119C.
7. Maldonado, M.; Oleksiak, M.D.; Chinta, S.; Rimer, J.D. Controlling Crystal Polymorphism in Organic-Free
Synthesis of Na-Zeolites. J. Am. Chem. Soc. 2012, 135, 2641–2652, doi:10.1021/ja3105939.
8. Nicholas, C.P. Zeolites in Industrial Separation and Catalysis; Wiley VCH: Weinheim, Germany, 2010;
pp. 355–402.
9. Amorim, R.; Vilaça, N.; Martinho, O.; Reis, R.M.; Sardo, M.; Rocha, J.; Fonseca, A.M.; Baltazar, F.; Neves, I.C.
Zeolite Structures Loading with an Anticancer Compound As Drug Delivery Systems. J. Phys. Chem. C
2012, 116, 25642–25650, doi:10.1021/jp3093868.
10. Chaves, T.F.; Soares, F.; Cardoso, D.; Carneiro, R.L. Monitoring of the crystallization of Zeolite LTA using
Raman and Chemometric tools. Analyst 2015, 140, 854–859, doi:10.1039/C4AN00913D.
11. Lalik, E.; Mirek, R.; Rakoczy, J.; Groszek, A. Microcalorimetric study of sorption of water and ethanol in
zeolites 3A and 5A. Catal. Today 2006, 114, 242–247.
12. Liu, Q.; Mace, A.; Bacsik, Z.; Sun, J.; Laaksonen, A.; Hedin, N. NaKA sorbents with high CO2 -over-N2
selectivity and high capacity to adsorb CO2 . Chem. Comm. 2010, 46, 4502–4504, doi:10.1039/C000900H.
13. Sun, H.; Wu, D.; Guo, X.; Shen, B.; Liu, J.; Navrotsky, A. Energetics of Confinement of n-Hexane in Ca-Na
Ion Exchanged Zeolite A. J. Phys. Chem. C 2014, 118, 25590–25596, doi:10.1021/jp508514e.
14. Kwakye-Awuah, B.; Labik, L.K.; Nkrumah, I.; Williams, C. Removal of ammonium ions by laboratory
synthesized zeolite linde type A adsorption from water samples affected by mining activities in Ghana.
J. Water Health 2014, 12, 151–160, doi:10.2166/wh.2013.093.
15. Sartbaeva, A.; Rees, N.H.; Edwards, P.P.; Ramirez-Cuesta, A.J.; Barney, E. Local probes show that
framework modification in zeolites occurs on ammonium exchange without calcination. J. Mater. Chem. A
2013, 1, 7415–7421, doi:10.1039/C3TA10243B.
16. Seel, A.G.; Sartbaeva, A.; Edwards, P.P.; Rammirez-Cuesta, A.J. Inelastic neutron scattering of Na-zeolite
A with in situ ammoniation: An examination of initial coordination. Phys. Chem. Chem. Phys.
2010, 12, 9661–9666.
17. Watanabe, Y.; Yamada, H.; Tanaka, J.; Komatsu, Y.; Moriyoshi, Y. Ammonium Ion Exchange of Synthetic
Zeolites: The Effect of Their Open Window Sizes, Pore Structures, and Cation Exchange Capacities.
Sep. Sci. Technol. 2005, 39, 2091–2104, doi:10.1081/SS-120039306.
18. Navarrete-Casas, R.; Navarrete-Guijosa, A.; Valenzuela-Calahorro, C.; López-González, J.D.;
García-Rodríguez, A. Study of lithium ion exchange by two synthetic zeolites: Kinetics and equilibrium.
J. Colloid Interface Sci. 2007, 306, 345–353, doi:10.1016/j.jcis.2006.10.002.
19. Kim, H.S.; Choi, S.Y.; Lim W.T. Complete Li+ exchange into zeolite X (FAU, Si\Al = 1.09) from undried
methanol solution. J. Porous Mater. 2013, 20, 1449–1456, doi:10.1007/s10934-013-9731-1.
20. Bignami, G.P.; Dawson, D.M.; Seymour, V.R.; Wheatley, P.S.; Morris, R.E.; Ashbrook, S.E. Synthesis, Isotopic
Enrichment, and Solid-State NMR Characterization of Zeolites Derived from the Assembly, Disassembly,
Organization, Reassembly Process. J. Am. Chem. Soc. 2017, 139, 5140–5148.
Magnetochemistry 2017, 3, 42 15 of 16
21. Brouwer, D.H.; Darton, R.J.; Morris, R.E.; Levitt, M.H. A solid-state NMR method for solution of zeolite
crystal structures. J. Am. Chem. Soc. 2005, 127, 10365–10370, doi:10.1021/ja052306h.
22. Fyfe, C.A.; Feng, Y.; Grondey, H.; Kokotailo, G.T.; Gies, H. One- and two-dimensional high-resolution
solid-state NMR studies of zeolite lattice structures. Chem. Rev. 1991, 91, 1525–1543, doi:10.1021/cr00007a013.
23. Fyfe, C.A.; Gobbi, G.C.; Hartman, J.S.; Klinowski, J.; Thomas, J.M. Solid-state magic-angle spinning.
Aluminum-27 nuclear magnetic resonance studies of zeolites using a 400-MHz high-resolution spectrometer.
J. Phys. Chem. 1982, 86, 1247–1250, doi:10.1021/j100397a006.
24. Fyfe, C.A.; Thomas, J.M.; Klinowski, J.; Gobbi, G.C. Magic-Angle-Spinning NMR (MAS-NMR) Spectroscopy
and the Structure of Zeolites. Angew. Chem. Int. Ed. Engl. 1983, 22, 259–275, doi:10.1002/anie.198302593.
25. Lippmaa, A.; Maedi, M.; Samoson, A.; Tarmak, M.; Engelhardt, G. Investigation of the structure of zeolites
by solid-state high-resolution silicon-29 NMR spectroscopy. J. Am. Chem. Soc. 1981, 103, 4992–4996,
doi:10.1021/ja00407a002.
26. Nebel, H.; Neumann, M.; Mayer, C.; Epple, M. On the Structure of Amorphous Calcium Carbonate–A
Detailed Study by Solid-State NMR Spectroscopy. Inorg. Chem. 2008, 47, 7874–7879, doi:10.1021/ic8007409.
27. Park, M.B.; Vicente, A.; Fernandez, C.; Hong, S.H. Solid-state NMR study of various mono- and
divalent cation forms of the natural zeolite natrolite. Phys. Chem. Chem. Phys. 2013, 15, 7604–7612;
doi:10.1039/C3CP44421J.
28. Thomas, J.M.; Klinowsky, J.; Ramadas, S.; Anderson, M.W.; Fyfe, C.A.; Gobbi, G.C. New Approaches
to the Structural Characterization of Zeolites: Magic-Angle Spinning NMR (MASNMR). In Intrazeolite
Chemistry; Galen, D.S., Francis, G.D., Eds.; ACS Publications: Washington, DC, USA, 1983; pp. 159–180,
ISBN 9780841207745.
29. Gramlich, V.; Meier, W.M. The crystal structure of hydrated NaA: A detailed refinement of a
pseudosymmetric zeolite structure. Z. Krist. 1971, 10, 134–149, doi:10.1524/zkri.1971.133.133.134.
30. Treacy, M.M.J.; Higgins, J.B. Collection of Simulated XRD Powder Patterns for Zeolites; Elsevier: Amsterdam,
The Netherlands, 2007; pp. 252–253.
31. Alfaro, S.; Rodriguez, C.; Valenzuela, M.A.; Bosch, P. Aging time effect on the synthesis of small crystal LTA
zeolites in the absence of organic template. Mater. Lett. 2007, 61, 4655–4658, doi:10.1016/j.matlet.2007.03.009.
32. Carey, T.; Tang, C.C.; Hriljac, J.A.; Anderson, P.A. Chemical Control of Thermal Expansion in
Cation-Exchanged Zeolite A. Chem. Mater. 2014, 26, 1561–1566, doi:10.1021/cm403312q.
33. Lührs, H.; Derr, J.; Fischer, R.X. K and Ca exchange behavior of zeolite A. Microporous Mesoporous Mater.
2012, 151, 457–465, doi:10.1016/j.micromeso.2011.09.025.
34. Holland, T.J.B.; Redfern, S.A.T. Unit cell refinement from powder diffraction data–the use of regression
diagnostics. Miner. Mag. 1997, 61, 65–77.
35. O’Keeffe, M.; Navrotsky, A.; Some Aspects of the Ionic Model of Crystals. In Structure and Bonding in Crystals;
O’Keeffe, M., Ed.; Academic Press: Cambridge, MA, USA, 1981; pp. 299–322.
36. Ramdas, S.; Klinowski, J. A simple correlation between isotropic 29Si-NMR chemical shifts and T-O-T angles
in zeolite frameworks. Nature 1984, 308, 521–523, doi:10.1038/308521a0.
37. Edith, M.F.; Hassan, K.; Herman, A.S. Infrared Structural Studies of Zeolite Frameworks. In Molecular
Sieve Zeolites-I; Edith M.F., Leonard B.S., Eds.; ACS Publications: Washington, DC, USA, 1974; pp. 201–229,
ISBN 9780841201149.
38. Milkey, R.G. Infrared Spectra of some Tectosilicates. Am. Minerol. 1960, 45, 990–1007.
39. Dyballa, M.; Obenaus, U.; Lang, S.; Gehring, B.; Traa, Y.; Koller, H.; Hunger, M. Brønsted sites and
structural stabilization effect of acidic low-silica zeolite A prepared by partial ammonium exchange.
Microporous Mesoporous Mater. 2015, 212, 110–116, doi:10.1016/j.micromeso.2015.03.030.
40. Ucun, F. An Infrared Study of the CaA Zeolite Reacted with CO2 . Z. Naturforsch. 2002, 57, 283.
41. Holzwarth, U.; Gibson, N. The Scherrer equation versus the ’Debye-Scherrer equation’. Nat. Nano 2011, 6, 534.
42. Cubillas, P.; Gebbie, J.T.; Stevens, S.M.; Blake, N.; Umemura, A.; Terasaki, O.; Anderson, M.W. Atomic Force
Microscopy and High Resolution Scanning Electron Microscopy Investigation of Zeolite A Crystal Growth.
Part 2: In Presence of Organic Additives. J. Phys. Chem. C 2014, 118, 23092–23099, doi:10.1021/jp506222y.
43. Kliewer, C.E. Electron Microscopy and Imaging. In Zeolite Characterization and Catalysis; Chester, A.W.,
Derouane, E.G., Eds.; Springer: Dordrecht, The Netherlands, 2009; pp. 169–196.
44. Mintova, S.; Olson, N.H.; Valtchev, V.; Bein, T. Mechanism of Zeolite A Nanocrystal Growth from Colloids at
Room Temperature. Science 1999, 283, 958–960, doi:10.1126/science.283.5404.958.
Magnetochemistry 2017, 3, 42 16 of 16
45. Dimitrov, L.; Valtchev, V.; Nihtianova, D.; Kalvachev, Y. Submicrometer Zeolite A Crystals Formation:
Low-Temperature Crystallization Versus Vapor Phase Gel Transformation. Cryst. Growth Des.
2011, 11, 4958–4962, doi:10.1021/cg2008667.
46. Smaihi, M.; Barida, O.; Valtchev, V. Investigation of the Crystallization Stages of LTA-Type Zeolite by
Complementary Characterization Techniques. Eur. J. Inorg. Chem. 2003, 2003, 4370–4377, doi:10.1002/
ejic.200300154.
c 2017 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).