1 s2.0 S1387181123000823 Main

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Microporous and Mesoporous Materials 353 (2023) 112511

Contents lists available at ScienceDirect

Microporous and Mesoporous Materials


journal homepage: www.elsevier.com/locate/micromeso

Synthesis of PHI-type zeolite harmotome using barium as an inorganic


structure-directing agent inspired from natural mineral compositions
Yu Liang a, Allan J. Jacobson a, b, **, Jeffrey D. Rimer a, c, *
a
Department of Chemistry, University of Houston, Houston, TX, 77204, USA
b
Texas Center for Superconductivity, University of Houston, Houston, TX, 77204, USA
c
Department of Chemical and Biomolecular Engineering, University of Houston, Houston, TX, 77204, USA

A B S T R A C T

Zeolites are nanoporous aluminosilicates widely used as commercial adsorbents, heterogeneous catalysts, and ion-exchange materials due to their unique porosity,
acidity, and thermal stability. The vast majority of synthetically-realized zeolite structures are prepared under hydrothermal conditions in alkaline media, and often
in the presence of an organic structure-directing agent that facilitates the crystallization of diverse porous networks. Due to economic and environmental disad­
vantages of organic-based syntheses, it is often desirable to produce zeolites in organic-free media using alkali metal ions as the most commonly employed inorganic
structure-directing agents (ISDAs). In this study, we were inspired by the elemental compositions of natural zeolite minerals to examine the use of an alkaline earth
metal ion as the ISDA to prepare zeolites with the PHI framework type. Two natural minerals exhibiting this crystal structure are phillipsite and harmotome. Both
structures contain sodium and potassium as extra-framework cations, but only harmotome contains barium cations. Here we examined how the combination of alkali
and alkaline earth ISDAs promote the formation of PHI-type zeolites. Our findings reveal that each isostructure is prepared under different synthesis conditions where
simply adding barium to a growth mixture does not always lead to harmotome formation. Harmotome is formed in more dilute and alkaline growth media; however,
time-resolved analyses of zeolite crystallization reveal that harmotome can be synthesized at low temperatures (ca. 65 ◦ C) with rates that are 5-times faster than
syntheses of phillipsite. This seemingly indicates that barium reduces the energetic barrier for PHI-type zeolite nucleation in ways that are not fully understood.
Moreover, harmotome crystals exhibit an average particle size of ca. 60 nm, which is nearly 40-fold smaller than phillipsite crystals prepared under similar con­
ditions. Collectively, these findings identify routes to control crystallization of PHI-type zeolites and may provide a broader strategy for using the composition of
natural zeolite minerals to optimize the preparation of synthetic analogues.

1. Introduction [13–18]. To avoid the negative environmental impacts of calcining or­


ganics in zeolites post-synthesis, and to minimize the costs associated
For decades, a wide range of zeolite structures have been intensely with producing/using OSDAs, it is desirable to develop organic-free
investigated on the basis of their diverse porosity, acidity, and (hydro) methods for zeolite synthesis. This active research area has numerous
thermal stability [1–5]. From the mid-18th century, zeolites were challenges for developing low-cost routes to prepare zeolites with
recognized as a category of naturally occurring minerals [6]. In the desired physicochemical properties without the aid of OSDAs [19,20].
1940’s, Barrer pioneered the study of zeolite synthesis, which led to the This study focuses on zeolites with the PHI framework type. In the
first preparation of zeolite framework types without natural analogues 1960’s, crystallographic investigations of naturally occurring zeolite
[7–9]. Since these pioneering studies there have been significant efforts minerals uncovered two structures with a PHI-type framework: har­
to optimize zeolite crystallization through diverse methods, often motome and phillipsite [21,22]. Kuhl successfully synthesized phil­
guided by the elemental composition and geothermal conditions for lipsite using growth mixtures with ratios Si/Al = 2–8 and Na2O/(Na2O
natural zeolite mineralization [10–12]. State-of-the-art syntheses of + K2O) = 0.3–0.85, which closely resemble the composition of natural
zeolites for commercial applications often require the use of costly phillipsite [23]. Synthetic harmotome was first prepared by Perrotta in a
organic molecules that function as organic structure-directing agents growth mixture with molar composition 5.25 SiO2: 1 Al2O3: 9.5 Na2O:
(OSDAs); however, a majority of commercial zeolites (e.g., faujasite, 0.24 BaO: 1528 H2O after 96 h of heating at 95 ◦ C [24]. Both phillipsite
mordenite, and chabazite) can be crystallized in the absence of organics and harmotome share the same topology, while the elemental compo­
using alkali metals as inorganic structure-directing agents (ISDAs) sitions of extra-framework species characterize their structural

* Corresponding author. Department of Chemical and Biomolecular Engineering, University of Houston, Houston, TX, 77204, USA.
** Corresponding author. Department of Chemistry, University of Houston, Houston, TX, 77204, USA.
E-mail addresses: ajjacob@central.uh.edu (A.J. Jacobson), jrimer@central.uh.edu (J.D. Rimer).

https://doi.org/10.1016/j.micromeso.2023.112511
Received 15 January 2023; Received in revised form 13 February 2023; Accepted 23 February 2023
Available online 25 February 2023
1387-1811/© 2023 Elsevier Inc. All rights reserved.
Y. Liang et al. Microporous and Mesoporous Materials 353 (2023) 112511

differences. Rinaldi and coworkers [25] investigated the structures of (FTIR) was measured with a Thermo Scientific Nicolet iS20 FTIR
phillipsite and harmotome from refinement of X-ray data from single Spectrometer.
crystals of natural minerals. Their findings revealed the same cationic
sites and water molecule distributions with slight distortions; however, 3. Results and discussion
syntheses of PHI-type zeolites have been rarely studied to elucidate the
role of growth medium composition, particularly those mimicking the Syntheses of zeolites in the absence of organic-structure-directing
chemical compositions of their respective natural minerals. agents (OSDAs) typically use alkali metals as inorganic structure-
Herein, we investigate the formation of PHI-type zeolite frameworks directing agents (ISDAs) [27–31]. Relatively few studies have explored
and explore the synthetic conditions that promote the crystallization of the use of multivalent ions as ISDAs, either as replacements of alkali
either harmotome or phillipsite zeolites. We systematically examined metals or in combination with monovalent ions as mixtures. A survey of
the impact of multiple ISDAs on zeolite structure and its physicochem­ the database of natural zeolites and their elemental compositions reveal
ical properties. Our methodology of zeolite synthesis was guided by the that divalent ions are common in zeolite minerals [32]. Table 1 lists
composition of natural zeolites wherein we use a combination of alkali several naturally occurring zeolites that have been synthesized using
and alkaline earth metals to tailor the final zeolite structure while organic-free reactions. Many of these zeolites (e.g., faujasite, chabazite
simultaneously reducing crystallization time compared to conventional and phillipsite) contain more than one extra-framework metal ion,
methods of preparation. Our findings reveal that the exploration of owing to variations in natural synthesis conditions (i.e. abundance of
OSDA-free zeolite syntheses could benefit from careful consideration of natural minerals in various regions of the world) and presumably the
natural zeolites in order to provide new opportunities for zeolite crys­ similarity of atomic radii of alkali and alkaline earth metals. We observe
tallization via green routes. only a marginal difference in the silicon-to-aluminum ratio (Si/Al)
among natural minerals with different monovalent/divalent cation
2. Experimental methods compositions. Considering the high abundance of alkaline earth metals
in the earth’s crust, it is reasonable to anticipate the elemental compo­
Materials. The following reagents were purchased from Sigma- sition of natural zeolites reflects the local environment of mineraliza­
Aldrich: sodium aluminate (technical grade), sodium hydroxide tion. This calls into question whether the cations are merely spectators
(NaOH, 98%), potassium hydroxide solution (KOH, 40%), LUDOX AS-40 that function as charge-balancing extra-framework species, or if their
colloidal silica (SiO2, 40%), and barium hydroxide octahydrate (Ba role is that of an ISDA that impacts either polymorph selection or the
(OH)2⋅8H2O, 98%). All reagents were used as received without further rate of zeolite crystallization. This question has not been thoroughly
purification. Deionized (DI) water was produced with an Aqua Solutions examined in the literature for laboratory syntheses employing mixtures
RODI-C-12A purification system (18.2 MΩ). of alkali and alkaline earth metals, which is the subject of this study.
Zeolite synthesis. Phillipsite zeolite was prepared by modifying a Taking inspiration from the compositions of natural zeolites listed in
reported protocol [26] using a growth mixture with molar composition 1 Table 1, we examine herein the impact of substituting alkali metals with
SiO2: 0.24 Al2O3: 0.10 K2O: 0.31 Na2O: 16.48 H2O. In a typical syn­ barium in conventional Na,K-zeolite syntheses.
thesis, potassium hydroxide solution and sodium hydroxide were mixed Combinations of Na, K, and Ba for the molar compositions and
in DI water until the sodium hydroxide was fully dissolved. To this so­ conditions selected for this study (see Table 2) resulted in four different
lution was added sodium aluminate, followed by continuous stirring zeolite structures: LTA, FAU, GIS, and PHI. Herein, we focus on the role
until the mixture was homogenous. LUDOX AS-40 colloidal silica was of cation mixtures to promote PHI-type zeolite, which has a 3-dimen­
then added as the silica source. sional pore network with 8-member ring (8 MR) pore apertures
Harmotome zeolite containing barium ions was synthesized using a (Fig. 1A). We report the synthesis of two materials exhibiting the same
growth mixture with molar composition 1 SiO2: 0.50 Al2O3: 2.75 Na2O: PHI topology: phillipsite and harmotome. The framework atoms (i.e., Si,
0.25 BaO: 95 H2O. First, sodium hydroxide pellets and barium hydrox­
ide octahydrate crystals were dissolved in DI water. To this solution was Table 1
added sodium aluminate, which was then mixed until homogenous, Examples of natural zeolites that have been synthetically realized in organic-free
followed by the addition of LUDOX AS-40 colloidal silica. syntheses [32].
Growth mixtures were aged for 24 h at room temperature prior to
Natural Framework Composition
being transferred to a 60 mL polypropylene (PP) laboratory bottle Zeolite Typea
(VWR) for syntheses at 65 ◦ C. For higher temperatures, the growth
Chabazite CHA K |(K,Na,Ca0.5)x(H2O)12| [Alx Si12-x O24] x =
mixture was placed in a 23-mL Teflon-lined metal autoclave (Parr In­ 3.0–4.5
struments). The synthesis temperature (65 or 100 ◦ C) was maintained Na |(Na,K,Ca0.5)x(H2O)12| [Alx Si12-x O24] x =
using a ThermoFisher Precision oven and reaction vessels were heated 2.5–4.8
under static conditions (i.e. without rotation). Containers were removed Sr |(Sr0.5,Ca0.5,K)4(H2O)12| [Al4 Si8 O24]
Faujasite FAU Na |(Na,Ca0.5,Mg0.5,K)x(H2O)16|[AlxSi12-xO24]
after heating times ranging from 1 h to 7 d. Upon removal from the oven,
x = 3.2–4.3
the container was quenched to room temperature in a water bath. Solids Ca |(Ca0.5,Na,Mg0.5,K)x(H2O)16|[AlxSi12-xO24]
were recovered by three cycles of centrifugation, each at 5 ◦ C and x = 3.3–3.9
13,000 rpm for 5 min using a Beckman Coulter Avanti J-E centrifuge and Mg |(Mg0.5,Ca0.5,Na,K)3.5(H2O)16|[Al3.5Si8.5
intermediate washings with DI water. The recovered products were O24]
Merlinoite MER |(K,Ca0.5,Ba0.5,Na)10 (H2O)22| [Al10Si22O64]
dried in air at 50 ◦ C overnight prior to characterization. Mordenite MOR |Na2,Ca,K2)4(H2O)28| [Al8Si40O96]
Characterization. The crystalline phase purity of each zeolite sam­ Offretite OFF |Ca,K,Mg)(H2O)16|[Al5Si13O36]
ple was determined by powder X-ray diffraction (XRD) using a Rigaku Gismondine GIS |Ca4 (H2O)18| [Al8Si8O32]
diffractometer with Cu Kα radiation (40 kV, 44 mA). Scanning electron Harmotome PHI |(Ba0.5,Ca0.5,K,Na)5(H2O)12| [Al5Si11O32]
Phillipsite PHI |(Na,K,Ca0.5)x(H2O)12| [AlxSi16-xO32] x =
microscopy (SEM) was conducted on a FEI-235 Dual-Beam Focused Ion
3.7–6.7
Beam instrument operated at 15 kV and a 5 mm working distance. |(K,Na,Ca0.5)x(H2O)12| [AlxSi16-xO32] x =
Samples for SEM analysis were prepared by affixing dried crystals to a 3.8–6.4
SEM sample holder with carbon tape. Elemental analysis was performed |(Ca0.5,Na,K)x(H2O)12| [AlxSi16-xO32] x =
by energy dispersive X-ray spectroscopy (EDX) using a JEOL JSM 6330F 4.1–6.8

Field Emission scanning electron microscope at a working distance of 15 a


Three-letter codes for each zeolite framework are assigned by the Interna­
mm and voltage of 15 kV. Fourier-transform infrared spectroscopy tional Zeolite Association (IZA) Structure Commission.

2
Y. Liang et al. Microporous and Mesoporous Materials 353 (2023) 112511

Table 2 MR) and 8 MR of (Si, Al)O4 tetrahedra. The 8 MR channel has an opening
Parametric study of zeolite crystal phase(s) with varying molar ratios of inor­ of 5.4 Å that is nearly parallel to the a-axis. The framework tetrahedral
ganic cations. (T) sites (T = Si or Al) in 8 MR units are highlighted by the yellow
SiO2 Al2O3 H2O ISDA(s) Zeolite Product(s) tetrahedra in Fig. 1B and C. The oxygen atoms bonded to the T atoms are
Na2O K2O BaO
shown as red spheres in these schemes. Extra-framework Ca cations
(grey) are located in the 8 MR cavity, coordinated by four water mole­
1 0.2 16 0.3 0.1 Phillipsite

cules (cyan). Calcium sites can be partially or completely occupied by
– 0.1 Phillipsite
0.02 0.08 Phillipsite Na, considering that Ca and Na have comparable ionic radii of 116 and
0.05 0.05 Phillipsite 114 p.m., respectively. In phillipsite, extra-framework K cations (green)
0.08 0.02 Phillipsite sit in cradle-shaped cavities within the elongated 8 MR cavity, coordi­
2 1 190 5.5 – 0.5 Harmotome nated by five framework O atoms and three water molecules (Fig. 1B).
0.1 0.4 Harmotome The anisotropic positioning of the K cation in the cavity is presumably
0.2 0.3 Harmotome due to the electrostatic force from the presence of the Ca cation(s) at
0.3 0.2 GIS
neighboring site(s). In harmotome, the Ba cation (purple) sits at the
0.4 0.1 GIS
0.5 – LTA + FAU same site as the K cation in phillipsite and has similar coordination to
neighboring Ca ions, Na ions, and water molecules (Fig. 1C). The
refinement data from previously reported crystallography measure­
ments of natural PHI-type zeolites (Table S1) shows a significant
decrease of all Ba–O bond lengths between the Ba cation and water
molecules or framework O atoms. We speculate that Ba2+ cations exhibit
stronger electrostatic interactions with O sites compared to alkali
metals, which putatively stabilizes the structure of harmotome.
To our knowledge, few studies have reported the synthetic prepa­
ration of harmotome [24,34]; therefore, we investigated methods to
synthesize harmotome in environments that deviate from harsh
geothermal conditions (i.e., elevated temperature and prolonged treat­
ment time) required to form its natural counterpart. We observed that
the composition of natural harmotome is similar to that of natural fau­
jasite (FAU); thus, we attempted to synthesize harmotome by selecting a
laboratory FAU synthesis protocol and replacing a portion of sodium
with barium (i.e. a putative ISDA for the natural PHI-type mineral). The
growth mixture had a molar composition of 1 SiO2: 0.50 Al2O3: 2.75
Na2O: 0.25 BaO: 95 H2O. The temperature of FAU syntheses in literature
varies from 30 to 100 ◦ C. In our previous study of Na-zeolite poly­
morphism [35] we used 65 ◦ C as an average condition, which we
emulated in this study of Na,Ba-zeolite synthesis. Time-resolved bulk
crystallization was performed wherein the solid-phase was extracted
from the growth mixture after various heating times, and the product
was characterized by powder X-ray diffraction (XRD). Analysis of solids
after 24 h of heating revealed a powder XRD pattern (Fig. 2A) that is
Fig. 1. (A) Crystal structure of the zeolite PHI framework oriented along the identical to the harmotome reference pattern. The sharp peaks at 2θ =
[001] direction, composed of composite building units dcc and phi. Color 10–15◦ indicate a high crystallinity of as-synthesized harmotome.
coding is based on the concept of natural tiling proposed by Blatov and co­
Scanning electron micrographs (Fig. 2B) showed that this synthesis
workers as a new type of zeolite taxonomy wherein natural tiles of zeolites are
protocol resulted in nanocrystalline spheroidal harmotome particles
defined as natural building units (NBUs) [33]. The natural tiling method offers
a categorization system to compare the organizational characteristics of zeolite
with an average size of 56 ± 8 nm. This is both the smallest particle size
structures. For PHI-type zeolites there are two NBUs: t-oto (blue) and t-phi and shortest crystallization time reported for synthetic harmotome.
(green). (B and C) Location of atoms in two PHI-type materials, (B) phillipsite Synthetic phillipsite was reported by Cichocki et al. [26] using a
and (C) harmotome, are shown in crystal structures viewed along the [001] growth mixture containing both sodium and potassium ions as binary
direction. Images were created with the software DIAMOND using structural ISDAs. Here we used a similar molar composition, 1 SiO2: 0.24 Al2O3:
data for the natural minerals provided by the International Zeolite Association. 0.10 K2O: 0.31 Na2O: 16.48 H2O, but decreased the nominal tempera­
Note that while calcium is present in natural minerals, this element was omitted ture from 100 to 65 ◦ C for direct comparison with the harmotome syn­
from growth mixtures used to synthesize PHI-type zeolites in this study in order thesis. To our knowledge, this is the lowest temperature used for
to focus on the effect(s) of barium. phillipsite synthesis where a fully-crystalline product was achieved after
ca. 5 days of heating. The powder XRD pattern of as-synthesized phil­
Al, and O) in both materials are arranged similarly, but positioning of lipsite (Fig. 2C) agrees well with the reference pattern for its natural
extra-framework cations is distinct in each PHI-type material. Fig. 1 il­ counterpart. Although phillipsite and harmotome share the same PHI
lustrates the similarities and differences in the crystal structures of framework topology, the different distributions of extra-framework
phillipsite (Fig. 1B) and harmotome (Fig. 1C) determined by single cations (Fig. 1B and C, respectively) leads to distinct powder XRD pat­
crystal X-ray diffraction (SCXRD) of natural minerals [25]. The com­ terns. The majority of peaks overlap for both structures, but there are
positions of natural zeolites used to extract this crystallographic data differences in the range 2θ = 15–20◦ where phillipsite has four peaks in
were K~2Cã1.5Nã0.4Al~5Si~10O32.12H2O for phillipsite and contrast to the single peak at 17.5◦ for harmotome. A critical difference
Bã2Cã0.5Al~5Si~11O32.12H2O for harmotome. Given the small size of in powder XRD patterns, however, is the relative peak intensities. In the
synthetic PHI-type zeolites (<10 μm), SCXRD data is not available; range 2θ = 10–15◦ the three peaks of harmotome have similar intensity
therefore, we use the natural forms as guides for structural interpreta­ whereas the peak at 12.5◦ for phillipsite is significantly stronger than its
tion of the data presented in this study. two neighboring peaks. There is also a clear difference in the particle
The PHI framework is comprised of layers of four-membered rings (4 size of both PHI-type zeolites. Scanning electron micrographs of

3
Y. Liang et al. Microporous and Mesoporous Materials 353 (2023) 112511

Fig. 2. Powder X-ray diffraction patterns of (A) harmotome and (C) phillipsite compared to reference patterns of corresponding natural PHI-type zeolites obtained
from the International Zeolite Association (IZA) Structure Database. Two overlapping peaks denoted by the asterisks (*) are assigned to crystalline BaCO3 impurity.
Scanning electron micrographs of (B) harmotome and (D) phillipsite prepared by the partial substitution of Na+ ions with Ba2+ and K+ ions, respectively.

as-synthesized phillipsite (Fig. 2D) show a spheroidal morphology with performed a parametric study wherein the alkalinity, temperature
a much larger average crystal size (2.4 ± 0.3 μm) compared to (65 ◦ C), and time (1 and 5 days for harmotome and phillipsite, respec­
harmotome. tively) for each synthesis was fixed while the ratios of Na, K, and Ba ions
Comparison of phillipsite and harmotome syntheses at the same were varied. As shown in Table 2, introduction of Ba2+ to phillipsite
temperature reveal that the latter is nearly 5-times faster. Both syntheses syntheses had no impact on the final product, and was unable to pro­
used identical sources of silicon and aluminum, but a key difference mote the formation of harmotome. It appears that phillipsite crystalli­
between the two synthesis compositions is a much higher water content zation is more robust to changes in the ISDA composition with full
in the growth mixture of harmotome (Table 2), which seems to indicate crystallinity obtained when (K2O + BaO)/Na2O ≥ 3. Elemental analysis
the more dilute system exhibits a faster rate of crystallization. The role of of all phillipsite products did not detect the presence of barium, indi­
ISDAs in both syntheses is difficult to rationalize. For instance, we cating this divalent cation is a bystander that functions neither as an

Fig. 3. (A) Powder X-ray diffraction patterns of as-


synthesized harmotome (bottom pattern) and the
same sample after post-synthesis treatment with 0.1
M HCl for 24 h (top pattern). The dashed lines are
peaks associated with BaCO3 as a minor impurity. (B)
FTIR patterns of four different samples: reference
BaCO3, solids extracted from a harmotome growth
mixture after 24-h aging at room temperature (pre­
cursor), as-synthesized harmotome (7-day synthesis),
and the acid-treated form of the latter. The dashed
lines at 693, 855, and 1417 cm− 1 confirm the pres­
ence of BaCO3 impurity. The two lowest wave­
numbers correspond to O–C–O bending vibrations in
CO2−
3 while the highest wavenumber corresponds to
an asymmetric C–O stretching vibration.

4
Y. Liang et al. Microporous and Mesoporous Materials 353 (2023) 112511

ISDA nor as a charge-compensating extra-framework species under which is generated by a carbonation reaction of CO2 from air in contact
synthesis conditions favoring phillipsite formation. In contrast, harmo­ with the highly alkaline (pH > 14) harmotome growth mixture. The
tome can only form in the presence of Ba2+ at the condition BaO/Na2O BaCO3 peaks in powder XRD patterns are labelled in Fig. 3A, which was
= 0.05–0.09 (Table 2). When this ratio exceeds 0.09, we obtain an confirmed with a reference pattern (Fig. S1) where peaks at 2θ ≈ 24◦
amorphous product (Table S2); when the ratio falls below 0.05, we correspond to the (111) and (021) reflections of BaCO3. Removal of the
obtain GIS; and when no barium was added to the system we produce a impurity post-synthesis is possible by treating the sample with a mild
mixture of LTA and FAU. Collectively, these studies indicate that har­ acid (0.1 M HCl) solution. Analysis of the acid-treated sample by powder
motome synthesis is sensitive to the molar composition of the growth XRD (Fig. 3A) shows complete removal of the impurity. To further
mixture as well as the presence of Ba2+ in ways that are not fully confirm the formation and removal of BaCO3 impurity, we performed
understood. FTIR measurements (Fig. 3B). Comparison of FTIR spectra for a BaCO3
An undesirable outcome of harmotome synthesis is the presence of reference with the amorphous precursor extracted from a harmotome
an impurity, which is evident by the appearance of two prominent growth mixture after 24-h room temperature aging reveals nearly
overlapping peaks at 2θ ≈ 24◦ in the powder XRD pattern (asterisks in identical signatures. Notably, there is a strong absorption peak at 1417
Fig. 3A). The impurity has a distinct crystal size and habit (vide infra), cm− 1, which is assigned to the stretching of CO2−
3 [36]. This peak is also
thus the inability to detect this phase in electron micrographs of fully present in as-synthesized harmotome, but is expectedly absent from the
crystalline zeolite indicates its weight percentage of the total product is sample after acid treatment (Fig. 3B). Based on these analyses, we
low. The impurity forms prior to zeolite nucleation, even after aging the confirmed that the impurity in harmotome syntheses is BaCO3 and that
growth mixture at room temperature for 24 h where the powder XRD its presence can be removed through a facile post-synthesis treatment to
pattern of extracted solids shows intense peaks relative to the amor­ produce a pure zeolite product.
phous background (see Fig. 4A). We identified the impurity as BaCO3, Time-resolved syntheses were performed to determine the duration

Fig. 4. (A) Powder X-ray diffraction patterns of as-


synthesized harmotome samples prepared at 65 ◦ C
and extracted after various time periods of heating.
Two overlapping peaks denoted by the asterisks (*)
are assigned to crystalline BaCO3. (B - E) Scanning
electron micrographs of solids extracted from syn­
thesis mixtures used to prepare harmotome. (B)
Sample after aging the growth mixture for 24 h at
room temperature (RT) where both crystalline BaCO3
(arrow) and partially crystalline harmotome are pre­
sent. Additional images of extracted solids after
heating times of (C) 1 h, (D) 4 h, and (E) 24 h are
provided. All scale bars equal 500 nm.

5
Y. Liang et al. Microporous and Mesoporous Materials 353 (2023) 112511

of heating that is required to achieve fully-crystalline harmotome and very little difference in particle size and shape (Fig. 4C–E), and also
phillipsite. The initial appearance of BaCO3 impurity observed in pow­ reveals the absence of rod-like BaCO3 particles that are clearly evident in
der XRD patterns of harmotome precursors after room temperature (RT) the powder XRD patterns. This suggests that the initial precursors may
aging (Fig. 4A, bottom pattern) results in needle-shaped crystals undergo a solid-state transformation to harmotome, although direct
(Fig. 4B, arrow) within a solid phase that is predominantly comprised of evidence of this mechanism is lacking from our ex situ analysis. Such a
amorphous aluminosilicates. The habit of BaCO3 crystals agrees with process was recently proposed for zeolite FAU using in situ atomic force
those reported in studies of witherite nanorods [37]. The presence of the microscopy to show that the supernatant solution had no observable role
amorphous phase is not detected in the powder XRD pattern owing to in crystal growth [38].
the strong scattering of Ba, which is a much heavier element than Si and Time-resolved analysis of the phillipsite synthesis reveals additional
Al. The fact that amorphous precursors exhibit a spheroidal morphology details of its crystallization. Solids extracted after 7 h of heating appear
that matches that of the final harmotome product also makes it difficult to be completely amorphous by XRD analysis (Fig. 5A). The amorphous
to determine when crystallization is complete. After 1 h of heating at peaks are more clearly observed in comparison to those of harmotome
65 ◦ C, all Bragg peaks of the harmotome crystal were present in the precursors due to the absence of barium. The first Bragg peaks appear
powder XRD pattern (Fig. 4A). Heating for 4 h resulted in very little after ca. 24 h of heating; however, these peaks do not match PHI-type
change to the peak intensities, whereas the high intensities after 24-h of zeolite, but are in excellent agreement with zeolite LTA (Fig. S2). It is
heating signify a crystalline product. Comparison of scanning electron well known that Na+ ions promote LTA-type zeolite at these synthesis
microscopy (SEM) images throughout the duration of heating shows conditions (65 ◦ C, Si/Al = 2.5) [35]. Complete substitution of Na with K

Fig. 5. (A) Powder X-ray diffraction patterns of as-synthesized phillipsite samples prepared at 65 ◦ C and extracted after various time periods of heating: 7, 24, 72,
and 120 h. (B–E) Scanning electron micrographs of solids extracted from growth mixtures used to prepare phillipsite after heating for (B) 7 h, (C) 24 h, (D) 72 h, and
(E) 120 h. Scale bars in all insets are equal to 500 nm.

6
Y. Liang et al. Microporous and Mesoporous Materials 353 (2023) 112511

would result in an entirely different zeolite product [39]. The phillipsite subject to interzeolite transformation when increasing the synthesis
growth mixture has a ratio of Na/K = 3, which does not suppress LTA temperature to 100 ◦ C and extending the synthesis time to 7 days. This is
nucleation. As the heating time increases, however, we observe a commonly observed for other zeolites, such as FAU, which transforms to
LTA-to-PHI interzeolite transformation with Bragg peaks for phillipsite GIS when the temperature is increased from 65 to 100 ◦ C and the syn­
observed after ca. 72 h of heating (Fig. 5A). We do not observe a fully thesis time is extended [41]. The ternary reaction diagram in Fig. 7A,
crystalline phillipsite product until around 120 h of heating. Comparison however, shows that higher temperature reduces many (partially-)
of SEM images at periodic stages of synthesis (Fig. 5B–E) shows pro­ amorphous products and preserves the harmotome structure. We begin
gressive changes in particle size of each phase: amorphous precursors to see a reduction in the harmotome region once the water content is
(Fig. 5B), LTA crystals (Fig. 5C), and phillipsite crystals (Fig. 5E). The increased (Fig. 7A, top diagram). Surprisingly, a more dilute synthesis
morphology of each phase is much less defined with all products mixture results in the formation of zeolite LTL (or zeolite L) in a region of
appearing to be spheroidal. the ternary reaction diagram that is highly unusual for this zeolite
We examined the sensitivity of harmotome synthesis with respect to structure [42]. To our knowledge, organic-free syntheses of zeolite L
changes in growth mixture composition, using the three principle have only been reported in media where K+ ions are the sole ISDA. Here,
components (Si, Al, OH molar fractions) as variables and holding all we observe zeolite L crystallization in a binary Na+ and Ba2+ growth
other conditions fixed with molar ratios y SiO2: z Al2O3: 2.75 Na2O: 0.25 mixture. Moreover, zeolite L syntheses are typically reported at much
BaO: 95 H2O. The shaded areas within the ternary reaction diagram higher temperatures, suggesting this ISDA combination lowers the bar­
(Fig. 6A) depict regions of a single crystalline phase: yellow corresponds rier for LTL nucleation.
to harmotome (PHI type) and green corresponds to gismondine (GIS The potential applications for harmotome and any performance ad­
type). Pure harmotome products are limited to a region spanning 1 < Si/ vantages that it may have over phillipsite have yet to be realized. Here
Al < 4 with bordering regions leading to higher fractions of amorphous we assess differences in their relative hydrophilicity using thermogra­
product. Interestingly, zeolite GIS is shown to crystallize at 65 ◦ C vimetric analysis (TGA) to compare the water loss from as-synthesized
compared to higher temperatures (e.g. 100 ◦ C) reported in literature samples. Fig. 7B shows that harmotome more strongly binds water
[40]. Energy dispersive spectroscopy (EDS) elemental analysis of the GIS based on a delayed weight reduction (e.g. +46 ◦ C difference). On the
products show evidence of extra-framework Ba2+. The latter could not contrary, the presence of extra-framework Ba2+ ions in harmotome re­
be ion exchanged post-synthesis with NH+ 4 at 80 C, which was

duces its water capacity with phillipsite adsorbing 1.1% more water by
confirmed by the identical compositions and powder XRD patterns of weight. Extension of these analyses to gas adsorption using CO2 as a
samples before and after three cycles of ion exchange. We posit that Ba2+ probe molecule (Fig. S4) showed no differences between the two PHI-
functions as an ISDA for the GIS framework and lowers its barrier for type zeolites; however, future studies may uncover other differences
nucleation compared to when Na+ is the sole inorganic cation. The latter in physicochemical properties that impact their performance in diverse
is a conventional ISDA for GIS syntheses but requires a higher temper­ applications.
ature (i.e., >80 ◦ C). Zeolite GIS is a similar zeolite framework comprised
of 3-dimensional pores with 8 MR apertures and two composite building 4. Conclusion
units: gis cages and dcc chains (Fig. S3).
EDS analysis of harmotome also reveals that Na+ and Ba2+ ions are In summary, we have used a combination of both alkali and alkaline
incorporated in the crystalline product as extra-framework cations (with earth metals as inorganic structure-directing agents to synthesize PHI-
Na/Ba = 1.7). All previous experiments of harmotome in this study used type zeolites. The selection of Ba2+ was inspired by the natural min­
a fixed barium concentration; however, we also investigated the impact eral harmotome, which contains this alkaline earth metal as one of
of Ba2+ content on harmotome crystallization by selecting growth several extra-framework cations. The isostructural counterpart, phil­
mixtures with molar compositions of 1 SiO2: 0.50 Al2O3: 2.75 Na2O: x lipsite, is prepared with K+ and Na+ as ISDAs. Our findings uncovered a
BaO: 95 H2O (with x = 0.1–2.0). Fig. 6B reveals a non-monotonic low-temperature route to prepare harmotome (at 65 ◦ C) wherein the
relationship between percent crystallinity and the Al/Ba molar ratio of presence of Ba2+ leads to crystallization 5-times faster than the synthesis
the synthesis mixture. Under the conditions tested, harmotome exhibits of phillipsite. Our study indicates that harmotome is not formed by
the highest percent crystallinity when Al/Ba = 4 with reduced harmo­ simply replacing a fraction of K+ with Ba2+ cations. On the contrary, the
tome yields achieved at lower and higher Al/Ba ratios. This indicates growth mixture required to synthesize harmotome is more dilute and
that the concentration of barium is critical for harmotome alkaline than those used to prepare phillipsite. Moreover, there is a
crystallization. relatively narrow Al-to-Ba molar ratio of the starting gel that is required
We assessed whether harmotome is a metastable zeolite that is to achieve a purely crystalline harmotome product, whereas purely

Fig. 6. (A) Ternary reaction diagram plotting the


zeolite phase(s) obtained from syntheses at 65 ◦ C
using different molar fractions of Si, Al, and OH− .
Growth mixtures were prepared with an initial molar
composition of x Si: y Al: 11 MOH: 173 H2O (where
M = Na+ and Ba2+ with Na+/Ba2+ = 22). Symbols
correspond to the solid phase(s) detected by powder
XRD after extraction from 168 h of heating at 65 ◦ C.
The circles represent areas of single phases: harmo­
tome (yellow) and GIS (green). Amorphous products
are represented as filled black symbols whereas
partially-amorphous samples are represented by
symbols with varying degrees of black shading
depending on its relative percentage in powder XRD
patterns. (B) Analysis of the nominal harmotome
synthesis mixture prepared with different Al-to-Ba
ratios with the percent crystallinity from powder
XRD patterns showing an optimum Ba concentration
around Al/Ba = 4.

7
Y. Liang et al. Microporous and Mesoporous Materials 353 (2023) 112511

Fig. 7. (A) Ternary reaction diagrams plotting the


zeolite phase(s) obtained from syntheses using
different molar fractions of Si, Al, and OH− . Growth
mixtures were prepared with two different initial
molar ratios: (top) x Si: y Al: 11 MOH: 590 H2O and
(bottom) x Si: y Al: 11 MOH: 173 H2O (where M =
Na+ and Ba2+ with Na+/Ba2+ = 22). Symbols corre­
spond to the solid phase(s) detected in powder XRD
patterns after extraction from their growth mixtures
after 7 days heating at 100 ◦ C. (B) Thermogravimetric
analysis of as-synthesized harmotome (orange) and
phillipsite (black) percent weight loss with increasing
temperature at a heating rate of 5 ◦ C min− 1.

crystalline phillipsite can be prepared using a broader range of synthesis suggest that combinations of ISDAs mimicking natural zeolites could
conditions. provide new opportunities to modify the crystal structure and/or
The physicochemical properties of the two PHI-type zeolites pre­ physicochemical properties of synthetic zeolites.
pared in this study are distinct, despite having similar crystal structures.
Harmotome forms nanosized crystals (56 ± 8 nm) that are nearly 40- CRediT authorship contribution statement
times smaller than the size of phillipsite (2.4 ± 0.3 μm). Both zeolites
have similar Si/Al ratios, but harmotome is slightly more hydrophilic. Yu Liang: Writing – review & editing, Writing – original draft,
From the collective findings in this study, we posit that the composition Investigation, Formal analysis, Data curation, Conceptualization. Allan
(s) of natural zeolite minerals may serve as a guideline for the design of J. Jacobson: Writing – review & editing, Supervision, Funding acqui­
synthetic analogues; however, the conditions required to successfully sition, Formal analysis, Conceptualization. Jeffrey D. Rimer: Writing –
prepare each zeolite may be narrow, as was observed here for harmo­ review & editing, Writing – original draft, Supervision, Project admin­
tome. Given that natural zeolites exhibit a variety of elemental com­ istration, Funding acquisition, Conceptualization.
positions that frequently include alkaline earth metals, it is reasonable to

8
Y. Liang et al. Microporous and Mesoporous Materials 353 (2023) 112511

Declaration of competing interest [16] G.T. Kerr, Chemistry of crystalline aluminosilicates. I. Factors affecting the
formation of zeolite A, J. Phys. Chem. 70 (4) (1966) 1047–1050.
[17] B.O. Hincapie, et al., Synthesis of mordenite nanocrystals, Microporous
The authors declare that they have no known competing financial Mesoporous Mater. 67 (1) (2004) 19–26.
interests or personal relationships that could have appeared to influence [18] Y. Liang, A.J. Jacobson, J.D. Rimer, Strontium ions function as both an accelerant
the work reported in this paper. and structure-directing agent of chabazite crystallization, ACS Materials Letters 3
(2) (2020) 187–192.
[19] M.D. Oleksiak, J.D. Rimer, Synthesis of zeolites in the absence of organic structure-
Data availability directing agents: Factors governing crystal selection and polymorphism, Rev.
Chem. Eng. 30 (1) (2014) 1–49.
[20] D. Parmar, et al., Manipulation of amorphous precursors to enhance zeolite
Data will be made available on request. nucleation, Faraday Discuss 235 (2022) 322–342.
[21] R. Sadanaga, F. Marumo, Y. Takeuchi, The crystal structure of harmotome,
Acknowledgments Ba2Al4Si12O32. 12H2O, Acta Crystallogr. 14 (11) (1961) 1153–1163.
[22] H. Steinfink, The crystal structure of the zeolite, phillipsite, Acta Crystallogr. 15 (7)
(1962) 644–651.
JDR acknowledges support primarily from the U.S. Department of [23] G. Kühl, Synthetic phillipsite, Am. Mineral.: Journal of Earth and Planetary
Energy Office of Basic Energy Sciences (Award DE-SC0021384). JDR Materials 54 (11–12) (1969) 1607–1612.
[24] A. Perrotta, A low-temperature synthesis of a harmotome-type zeolite, Am.
and AJJ acknowledge financial support from The Welch Foundation Mineral. 61 (5–6) (1976) 495–496.
(Awards E-1794 and E-0024, respectively). [25] R. Rinaldi, J. Pluth, J. Smith, Zeolites of the phillipsite family. Refinement of the
crystal structures of phillipsite and harmotome, Acta Crystallogr. Sect. B Struct.
Crystallogr. Cryst. Chem. 30 (10) (1974) 2426–2433.
Appendix A. Supplementary data
[26] A. Cichocki, P. Kościelniak, Experimental designs applied to hydrothermal synthesis of
zeolite ERI+ OFF (T) in the Na2O–K2O–Al2O3–SiO2–H2O system: part 2. Regular study,
Supplementary data to this article can be found online at https://doi. Microporous Mesoporous Mater. 29 (3) (1999) 369–382.
org/10.1016/j.micromeso.2023.112511. [27] X. Meng, F.S. Xiao, Green routes for synthesis of zeolites, Chem. Rev. 114 (2)
(2014) 1521–1543.
[28] M. Alkan, et al., The effect of alkali concentration and solid/liquid ratio on the
References hydrothermal synthesis of zeolite NaA from natural kaolinite, Microporous
Mesoporous Mater. 86 (1–3) (2005) 176–184.
[1] C.S. Cundy, P.A. Cox, The hydrothermal synthesis of zeolites: precursors, [29] A.J. Mallette, S. Seo, J.D. Rimer, Synthesis strategies and design principles for
intermediates and reaction mechanism, Microporous Mesoporous Mater. 82 (1–2) nanosized and hierarchical zeolites, Nature Synthesis 1 (7) (2022) 521–534.
(2005) 1–78. [30] R.F. Lobo, S.I. Zones, M.E. Davis, Structure-direction in zeolite synthesis, in:
[2] M. Koizumi, R. Roy, Zeolite studies. I. Synthesis and stability of the calcium Journal of Inclusion Phenomena and Molecular Recognition in Chemistry,
zeolites, J. Geol. 68 (1) (1960) 41–53. Springer, Dordrecht, 1995, pp. 47–78.
[3] C.L. Angell, W.H. Flank, Mechanism of Zeolite A Synthesis, ACS Publications, 1977. [31] J. Yu, Synthesis of zeolites, Introduction to zeolite science and practice 168 (2007)
[4] L.B. Sand, Zeolite synthesis and crystallization, Pure Appl. Chem. 52 (9) (1980) 39.
2105–2113. [32] C. Baerlocher, L.B. McCusker, Database of Zeolite Structures, International Zeolite
[5] P.V. Shertukde, et al., Acidity of H-Y zeolites: role of extralattice aluminum, Association, 2001.
J. Catal. 139 (2) (1993) 468–481. [33] N.A. Anurova, et al., Natural tilings for zeolite-type frameworks, J. Phys. Chem. C
[6] V. Bartow, Axel Fredrick Cronstedt, J. Chem. Educ. 30 (5) (1953) 247. 114 (22) (2010) 10160–10170.
[7] R.M. Barrer, 33. Synthesis of a zeolitic mineral with chabazite-like sorptive [34] O. Chiyoda, M.E. Davis, Hydrothermal conversion of Y-zeolite using alkaline-earth
properties, J. Chem. Soc. (1948) 127–132. cations, Microporous Mesoporous Mater. 32 (3) (1999) 257–264.
[8] R.M. Barrer, J.W. Baynham, The hydrothermal chemistry of the silicates. Part VII. [35] M. Maldonado, et al., Controlling crystal polymorphism in organic-free synthesis of
Synthetic potassium aluminosilicates, J. Chem. Soc. (1956) 2882–2891. 562. Na-zeolites, J. Am. Chem. Soc. 135 (7) (2013) 2641–2652.
[9] R.M. Barrer, Zeolites and their synthesis, Zeolites 1 (3) (1981) 130–140. [36] R.T. Downs, M. Hall-Wallace, The American Mineralogist crystal structure
[10] H. Lechert, New routes in zeolite synthesis, in: Studies in Surface Science and database, Am. Mineral. 88 (1) (2003) 247–250.
Catalysis, Elsevier, 1984, pp. 107–123. [37] Q.-Z. Yao, et al., Witherite nanorods form mesocrystals: a direct experimental
[11] T. Bauer, W.H. Baur, Structural changes in the natural zeolite gismondine (GIS) examination of a dipole-driven self-assembly model, Eur. J. Mineral 24 (3) (2012)
induced by cation exchagne with Ag, Cs, Ba, Li, Na, K and Rb, Eur. J. Mineral 10 (1) 519–526.
(1998) 133–147. [38] R. Jain, et al., In situ imaging of faujasite surface growth reveals unique pathways
[12] H.G. Karge, H. Robson, Verified Syntheses of Zeolitic Materials. Characterization by of zeolite crystallization, J. Am. Chem. Soc. 145 (2) (2023) 1155–1164.
IR Spectroscopy, Fritz-Haber-Institut der Max-Planck-Gesellschaft, Berlin, [39] A. Chawla, et al., Crystallization of potassium-zeolites in organic-free media,
Germany, 2001. Microporous Mesoporous Mater. 341 (2022), 112026.
[13] R.F. Lobo, S.I. Zones, M.E. Davis, Structure-direction in zeolite synthesis, [40] M.D. Oleksiak, et al., Synthesis strategies for ultrastable zeolite GIS polymorphs as
J. Inclusion Phenom. Mol. Recognit. Chem. 21 (1–4) (1995) 47–78. sorbents for selective separations, Chem. Eur J. 22 (45) (2016) 16078–16088.
[14] A. Chawla, et al., Cooperative effects of inorganic and organic structure-directing [41] M.D. Oleksiak, et al., Synthesis strategies for ultrastable zeolite GIS polymorphs as
agents in ZSM-5 crystallization, Molecular Systems Design & Engineering 3 (1) sorbents for selective separations, Chem. Eur J. 22 (45) (2016) 16078–16088.
(2018) 159–170. [42] R. Li, et al., Ultrasmall zeolite L crystals prepared from highly interdispersed alkali-
[15] H. Awala, et al., Template-free nanosized faujasite-type zeolites, Nat. Mater. 14 (4) silicate precursors, Angew. Chem. Int. Ed. 57 (35) (2018) 11283–11288.
(2015) 447–451.

You might also like