Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Microporous and Mesoporous Materials 288 (2019) 109594

Contents lists available at ScienceDirect

Microporous and Mesoporous Materials


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

Synthesis and characterization of MFI-type zincosilicate zeolites with high T


zinc content using mechanochemically treated Si–Zn oxide composite
Peidong Hua, Kenta Iyokia, Hiroki Yamadaa, Yutaka Yanabab, Koji Oharac, Naonobu Katadad,
Toru Wakiharaa,*
a
Department of Chemical System Engineering, The University of Tokyo, 7-3-1 Hongo, Bunkyo-ku, Tokyo, 113-8656, Japan
b
Institute of Industrial Science, The University of Tokyo, 4-6-1 Komaba, Meguro-ku, Tokyo, 153-8505, Japan
c
Japan Synchrotron Radiation Research Institute/SPring-8, 1-1-1 Kouto, Sayo-cho, Sayo-gun, Hyogo, 679-5198, Japan
d
Center for Research on Green Sustainable Chemistry, Tottori University, 4-101 Koyama-cho Minami, Tottori, 680-8552, Japan

A R T I C LE I N FO A B S T R A C T

Keywords: Modification of zeolites by the incorporation of divalent Zn in the silicate framework is attracting extensive
Zincosilicate attention because it results in higher anionic charge density and creates new active acid sites. However, it is a
Zeolite great challenge to incorporate Zn into the framework of highly siliceous MFI-type zeolites without Al. In this
MFI-type study, ball milling, a mechanochemical method, was employed to prepare Si–Zn oxide composites from mixtures
Mechanochemical treatment
of fumed silica and ZnO as the starting materials for the zeolite synthesis. The dispersion of Zn on an atomic level
Ion exchange
inside the silica matrix was realized by mechanical forces. Then, the Si–Zn oxide composites were subjected to
hydrothermal treatment in the presence of an additional Si source and a structure-directing agent to synthesize
MFI-type zincosilicate zeolites with high Zn contents (Si/Zn = 36.9–13.4) but negligible Al impurity (Si/
Al > 900). The successful incorporation of Zn into the zeolite framework was confirmed by several character-
ization techniques. Ion exchange experiments showed the superior selectivity and capacity of MFI-type zinco-
silicate zeolites for divalent cations (Co2+ as a model cation) as compared with those of aluminosilicate analogs.
The mechanochemically treated Si–Zn oxide composite plays a crucial role in the synthesis of MFI-type zinco-
silicate zeolites with high Zn contents and minor extra-framework Zn species.

1. Introduction large number of organic molecules under mild conditions using H2O2 as
an oxidant, e.g., epoxidation of olefins, oxidation of alkane, and hy-
Zeolites are crystalline silicate-based materials constructed by TO4 droxylation of aromatics [10]. Stannosilicates have strong Lewis acidity
(T = Si, Al, P, etc.) tetrahedra that share corner oxygen atoms [1]. The due to the electron-deficient active sites created by isolated tin atoms in
unique properties of zeolites, including high thermal stability, large the silicate framework and are highly active catalysts for the selective
specific surface area, ordered and confined microporous structure, and conversion of biomass to chemicals [11]. Of note, the aforementioned
strong intrinsic acidity, make them promising candidates for various transition metals are tetravalent or trivalent, and do not introduce more
industrial applications as catalysts, adsorbents and separators [2–4]. charge imbalance than conventional aluminosilicate zeolites. In con-
Modification of zeolites by incorporating heteroatoms (T atoms other trast, zeolites with Zn incorporated in the silicate framework (i.e.,
than Si, Al and P) in the silicate framework is attracting extensive at- zincosilicate zeolites) can provide cation exchange capacities that ex-
tention because it can change the anionic charge density and allow the ceed those of aluminosilicate analogs because the isomorphous sub-
creation of new active acid sites (Brønsted or Lewis acid sites), which stitution of one Zn(II) for Si(IV) produces two negative charges per Zn
are closely related to the ion exchange capacity and catalytic activity atom [12]. Moreover, Zn in the silicate framework exhibits unique
[5]. Lewis acid properties that are different from those of framework Sn, Ti
Several transition metals (Ti, Sn, Fe, Mn, etc.) have been success- or Zr atoms, and can catalyze the direct production of dimethyl ter-
fully incorporated into the zeolite framework, and excellent catalytic ephthalate by Diels–Alder reactions [13]. Although Zn-rich zincosili-
performance has been verified [6–9]. For example, titanium-substituted cate zeolites with various topological structures, including VSV, VNI,
MFI-type zeolites (TS-1) are particularly effective in the oxidation of a and RSN, have already been developed, they contain only small pores

*
Corresponding author.
E-mail address: wakihara@chemsys.t.u-tokyo.ac.jp (T. Wakihara).

https://doi.org/10.1016/j.micromeso.2019.109594
Received 8 May 2019; Received in revised form 21 June 2019; Accepted 5 July 2019
Available online 05 July 2019
1387-1811/ © 2019 Elsevier Inc. All rights reserved.
P. Hu, et al. Microporous and Mesoporous Materials 288 (2019) 109594

and no cavities, which makes them difficult for use in industrial ap- 2. Experimental
plications. In contrast, research on zincosilicate zeolites with other to-
pological structures similar to aluminosilicate analogs, such as CHA 2.1. Materials
[14], SOD [15], and *BEA [16], is limited.
ZSM-5 zeolite with an MFI topology has been widely used in the Zinc oxide (ZnO, Sigma-Aldrich), fumed silica (Cab-O-Sil, M5,
petroleum industry as a heterogeneous catalyst due to its distinctive Cabot), sodium metasilicate nonahydrate (Na2SiO3·9H2O, Wako), tet-
two-dimensional channel structure, excellent shape-selectivity, and rapropylammonium bromide (TPABr, Sigma-Aldrich), and ammonium
high thermal and hydrothermal stabilities. Nevertheless, this highly chloride (NH4Cl, Wako) were obtained commercially and utilized
siliceous zeolite has a higher Si/Al molar ratio compared with MOR-, without further purification.
FAU- and LTA-type zeolites, and consequently, fewer acid sites for
catalysis and ion exchange sites for stabilizing guest metal ions. Efforts 2.2. Preparation of milled composites
have also been made to incorporate Zn into the framework of MFI-type
zeolites. Although the preparation of zincosilicate zeolite with an MFI- A milled Si–Zn oxide composite was prepared by mechanically
type structure was reported decades ago, the Zn species were inclined to grinding fumed silica and ZnO in a certain ratio using a planetary ball
remain in the non-framework form [17]. MFI-type zincoaluminosilicate mill (Fritsch P6) equipped with silicon nitride balls and pot (Si3N4
zeolites with both Al and Zn in the framework were also recently de- ceramics sintered with Y2O3 and Al2O3 as additives). The milling was
veloped [18,19]. In this case, Al is still necessary to assist the crystal- conducted by repeated rotation at 600 rpm for 15 min, with 15 min
lization of the MFI-type zeolites, which decreases the amount of Zn intervals to avoid overheating of the apparatus and reagents. The total
incorporated in the framework and may lead to undesired side reactions milling time was fixed to 24 h (excluding break time). The milled Si–Zn
during the catalytic processes [12]. oxide composites were named as M-5, M-8, and M-12 (nominally called
It is always a challenge to introduce transition metals into a zeolite M-x), corresponding to molar percentages of Zn/(Si + Zn) of 4.8, 8.1,
framework because of the highly basic conditions required to trigger and 12.5 mol%, respectively. The mixtures of fumed silica and ZnO that
the crystallization of zeolites, causing the precipitation of metal ions as were mixed by hand were designated as H-x accordingly.
hydroxides. Chelating agents, such as citric acid [20], ammonia [21],
and ethylenediaminetetraacetic acid (EDTA) [22], are usually em- 2.3. Synthesis of zeolites
ployed in zeolite syntheses to stabilize the metal ions. However, extra-
framework metal oxide species are still present in the final products, The synthesis procedures for the MFI-type zincosilicate zeolites
and the amount of metal incorporated in the lattice is limited. The were similar to a reported method [9]. Briefly, Na2SiO3·9H2O (2.64 g),
fluoride-assisted method is an alternative strategy to prepare het- NH4Cl (0.88 g), TPABr (0.48 g), and M-x (0.66 g) were hand mixed in
eroatom-substituted zeolites owing to the excellent mineralization the presence of a small amount of water in a mortar for 10 min. The
ability of fluoride ions in neutral or even slightly acidic media [23]. batch compositions were tuned to the following molar ratio: T (Si and
Unfortunately, the hazardous effects of fluoride ions on the environ- Zn): NH4Cl: TPABr: H2O = 1: 0.8: 0.09: 5. Then, the mixture was
ment and humans restrict the practical applications of this method transferred into a Teflon-lined stainless steel autoclave, which was
[24]. sealed and kept at 170 °C statically. After several days, the reactor was
In recent decades, ball milling or bead milling, a kind of mechan- quenched with cold water, and the samples were collected by vacuum
ochemical method, has been employed for the preparation of zeolites filtration, thoroughly washed with deionized water, and dried at 80 °C.
[25,26]. In this method, mechanical energy is imposed on the materials Further calcination of the as-made samples was conducted at 500 °C for
by collision and/or friction of the milling media, and hence, it can be 5 h in a muffle furnace connected to an air pump. The final samples
used as a top-down approach to downsize zeolite crystals to the nano- were designated as Zn-MFI(M-x).
size regime [27,28]. More importantly, chemical reactions are also For comparison, a mixture of fumed silica and ZnO mixed by hand
believed to occur during this mechanochemical process because solid (H-8, Zn/(Si + Zn) = 8.2 mol%) was also used to prepare zeolites (Zn-
materials are mixed at the atomic level and chemical bonds are si- MFI(H-8)). Siliceous MFI-type zeolite without Zn (Si-MFI) was synthe-
multaneously rearranged [29,30]. Based on this idea, composite pre- sized through the above-mentioned procedures using mechanically
cursors with heteroatoms incorporated in the silica matrix were pre- milled fumed silica instead of the Si–Zn oxide composite.
pared by ball milling of silica and metal oxides, which were further
crystallized to prepare metal-containing zeotype materials without the 2.4. Characterization
problem of metal ion precipitation. During the milling process, im-
purities derived from the milling media may be left in the composites X-ray diffraction (XRD) was conducted on a Rigaku Ultima IV dif-
according to the operational conditions and material used as the milling fractometer with a Cu Kα radiation source (40 kV, 40 mA). The relative
media [31]. Using this method, titanosilicate zeolites with *BEA- and crystallinity was calculated as the ratio of integral areas of the dif-
MFI-type topologies were successfully synthesized [26,32], and its ap- fraction patterns between 22.5° and 25° of Zn-MFI(M-x) and Si-MFI.
plicability was expanded to other metallosilicate zeolites such as Sn- Diffuse reflectance ultraviolet–visible (DR UV–Vis) spectra were re-
Beta [33] and Mn-Silicalite-1 [9]. corded using a JASCO V-670 spectrometer in the wavelength range of
In this study, for the first time, MFI-type zincosilicate zeolites with 190–800 nm. Fourier transform infrared (FT-IR) spectra were obtained
high Zn contents were synthesized using the Si–Zn oxide composites on a JASCO FT/IR-6100 spectrometer using the potassium bromide
obtained by ball milling of mixtures of fumed silica and zinc oxide as pellet technique. X-ray photoelectron spectroscopy (XPS) measure-
the starting material. Using this mechanochemical process, Zn was ments were performed on a JEOL JPS-90 instrument using monochro-
dispersed in the silica matrix before hydrothermal treatment and was matic Mg Kα (1253.6 eV) X-ray radiation. The XPS spectra were cali-
finally incorporated in the framework of the products, as confirmed by brated using the C1s line (284.8 eV) as a reference. Nitrogen
several characterization techniques. Co2+ ion exchange experiments adsorption/desorption analyses were performed on a Quantachrome
showed that MFI-type zinosilicate zeolites exhibited superior selectivity Autosorb-iQ2-MP at liquid nitrogen temperature. Before the measure-
and capacity for divalent cations compared with those of aluminosili- ments, the samples were outgassed at 325 °C for 4 h under vacuum.
cate analogs. Field-emission scanning electron microscope (FE-SEM) images were
taken using a JSM-7000F (JOEL). The chemical compositions of the
samples were analyzed by inductively coupled plasma atomic emission
spectroscopy (ICP-AES, Thermo iCAP 6300). The samples were

2
P. Hu, et al. Microporous and Mesoporous Materials 288 (2019) 109594

dissolved in hydrofluoric acid before the measurements. Solid-state


magic angle spinning nuclear magnetic resonance (MAS NMR) mea-
surements of 29Si were conducted using a JNM-ECA 500 (JEOL) at
99.3 MHz with a pulse length of 5 μs, a recycle delay of 60 s, and a
spinning frequency of 10 kHz.
High-energy X-ray total scattering (HEXTS) measurements were
performed on the BL04B2 high-energy X-ray diffraction beamline
(SPring-8, Japan) with a horizontal two-axis diffractometer at room
temperature. Powdered samples were placed in kapton tubes. The en-
ergy of incident X-ray was 61.4 keV (λ = 0.202 Å). The maximum Q
(Q = 4π sin θ / λ ) collected here was 25 Å−1. The obtained data were
processed by well-established analysis procedures, including absorp-
tion, background, and Compton scattering corrections, and subse-
quently normalized to give the Faber–Ziman total structure factor, S(Q),
which was finally used to calculate the reduced pair distribution
function, G(r), according to Eq. (1), where ρ is the atomic number
density.
Qmax
2
G (r ) = 4πr [ρ (r ) − ρ0 ] =
π
∫ Q [S (Q) − 1]sin(Qr )dQ
Qmin (1)
Fig. 2. DR UV–Vis spectra of mechanochemically treated Si–Zn oxide compo-
sites (M-x) and hand-mixed mixture of fumed silica and ZnO (H-12).
2.5. Ion exchange experiments

The ion exchange performance of the synthesized MFI-type zinco- Card No. 65-3411). These peaks were also identified in the mixture of
silicate zeolites was evaluated through Co2+–Na+ ion exchange equi- fumed silica and ZnO without mechanochemical treatment (hand-
librium isotherms. Here, a commercial aluminosilicate ZSM-5 zeolite mixed H-12). After ball milling for 24 h, the peaks corresponding to
(Tosoh, HSZ-820NAA, Si/Al = 11.9, C-ZSM-5 (abbreviation)) was used ZnO in the XRD patterns of M-5, M-8 and M-12 all disappeared, and
as the reference sample. In this binary ion exchange system, Co only amorphous solids were detected. Furthermore, peaks representing
(NO3)2·6H2O and NaNO3 were dissolved in water to obtain the molar silicon nitride derived from the milling media were indistinguishable in
ratios of 2Co/(2Co + Na) from 0.1 to 1.0 at a consistent normality the XRD patterns. According to the results of the ICP-AES measure-
(0.1 N). In each batch, 0.04 g of the Na-form zeolite was added into ments, the Si/Al molar ratios of M-x were all above 500, which sug-
40 mL of the aforementioned solution, which was placed in a shaker at gested that the Al impurity introduced by the milling media was neg-
70 °C for 24 h. After ion exchange, the solid and solution phases were ligible.
separated and subjected to ICP-AES measurements, respectively. The DR UV–Vis spectrum (Fig. 2) of hand-mixed H-12 showed a
significant absorption band at 360–380 nm due to the transition of
electrons from the valence band to the conduction band in ZnO crystals
3. Results and discussion
(O2−Zn2+→O−Zn+ ligand to metal charge transfer transition) [34].
However, this absorption band was drastically reduced in the DR
3.1. Mechanochemical treatment of Si–Zn oxide composites
UV–Vis spectra of M-5, M-8, and M-12, while the intensity of the ab-
sorption bands at 190–200 nm were gradually enhanced with in-
ZnO displayed an XRD pattern (Fig. 1) with peaks at 2θ = 31.8°,
creasing contents of Zn. It has been reported that the light absorption
34.4°, 36.3° and 47.5°, suggesting a hexagonal crystal structure (JCPDS
ability of ZnO is not weakened with decreased particle size [35].
Therefore, the reduced absorption at 360–380 nm of milled composites
could be assigned to the change in the coordination environment of Zn
and the dispersion of Zn atoms within the silica matrix.
To further investigate the structural change of the fumed silica/ZnO
composites before and after mechanochemical treatment, HEXTS
measurements were conducted. The reduced pair distribution function,
G(r), obtained in this measurement made it possible to derive certain
interatomic distances, r, in the solid samples. The G(r) of H-12 and M-
12, calculated by Fourier transformation of the Faber–Ziman total
structure factor, S(Q) (Fig. S1), and that of crystalline ZnO, calculated
by PDFgui software [36] (calculated G(r) of each correlation in ZnO
crystal using the structural model was appended in Fig. S2) are depicted
in Fig. 3. In the curve of H-12, the first and second peaks were assigned
to Si–O (ca. 1.6 Å) and Zn–O (ca. 2.0 Å) correlations, respectively [37].
The peak corresponding to O–[Si]–O correlation appeared at ca. 2.6 Å
[9]. The peak observed at ca. 3.2 Å could be regarded as the super-
position of Si–Si correlation in fumed silica and Zn–Zn correlation in the
ZnO crystal. However, after mechanochemical treatment, one peak at
ca. 3.1 Å derived from Si–Si(Zn) correlation was identified in the curve
of M-12 without the detection of the nearest neighbor Zn–Zn correla-
Fig. 1. XRD patterns of ZnO (0.25 times the original intensity), fumed silica,
tion, and the peaks at ca. 4.6 and 5.6 Å related to Zn–Zn correlation
mechanochemically treated Si–Zn oxide composites (M-5, M-8, and M-12, Zn/
(Si + Zn) = 4.8, 8.1, and 12.5 mol%, respectively), and hand-mixed mixture of
could no longer be found. These results corroborated the destruction of
fumed silica and ZnO (H-12, Zn/(Si + Zn) = 12.8 mol%). crystalline ZnO by mechanical forces and the dispersion of Zn at the

3
P. Hu, et al. Microporous and Mesoporous Materials 288 (2019) 109594

coordinated into the framework of zeolite after hydrothermal treat-


ment, and few Zn atoms migrated from the lattice to form detectable
extra-framework Zn species after calcination. Of note, the absorption
band at approximately 200 nm increased proportionally with the Zn
contents in the zeolites. It is generally accepted that the absorption
band below 230 nm is the confirmation of isomorphous substitution of
metal atoms into the framework of zeolite [18].
The FT-IR spectra of Si-MFI and Zn-MFI(M-x) samples (Fig. S5)
depicted typical framework vibrations of MFI-type zeolites. The ab-
sorption peak at 550 cm−1 was attributed to the pentasil framework
vibration [40], while the absorption peaks at 450 and 800 cm−1 were
assigned to the bending vibration of the Si–O bond and the symmetrical
stretching vibration of the Si–O–Si bond, respectively [22]. The ab-
sorption peaks in the range of 1250–900 cm−1 were due to the asym-
metrical stretching vibration of the Si–O–Si bond. Different from the
peak of Si-MFI at approximately 1100 cm−1, the peaks of Zn-MFI(M-x)
showed obvious broadening toward lower frequencies, which could be
ascribed to the substitution of Si atom in the Si–O–Si bond with the
heavier Zn atom to form a Si–O–Zn bond in the zeolite framework
Fig. 3. Reduced pair distribution function, G(r), of ZnO (simulated), and hand- [41,42].
mixed (H-12) and mechanochemically treated (M-12) mixtures of fumed silica The XPS technique was applied to investigate the chemical en-
and ZnO. vironment of the Zn species in the zeolites. Zn2p core level XPS spectra
of ZnO, Zn-MFI(M-8), and Zn-MFI(M-12) are shown in Fig. 6A. All
atomic level inside the silica matrix to form Si–Zn oxide composites spectra depicted spin-orbital splitting with a doublet peak energy se-
[33]. paration of 23.1 eV, confirming the presence of divalent Zn species
[43]. However, an obvious chemical shift in binding energy was ob-
3.2. Synthesis of MFI-type zincosilicate zeolites served between bulk ZnO and zincosilicate zeolites, indicating different
electronic states of Zn. The Zn2p3/2 peaks of Zn-MFI(M-8) and Zn-MFI
The aforementioned mechanochemically treated Si–Zn oxide com- (M-12) were located at ca. 1023.4 eV, which were 1.6 eV higher than
posites were used as Si and Zn sources and underwent hydrothermal that of ZnO. This could be an evidence of the incorporation of Zn in the
treatment in the presence of an additional Si source (Na2SiO3) and a zeolite lattice because the lower electronegativity of Zn (1.65) com-
structure-directing agent (TPABr) to obtain the crystalline MFI-type pared with that of Si (1.90) could result in a lowered valence electron
zeolites. The high Zn contents in the products were confirmed by ICP- density of Zn in the Si–O–Zn bond and an increased binding energy
AES measurements (Table 1), while Al impurity could hardly be de- [44]. Furthermore, in the O1s spectra of Zn-MFI(M-8) and Zn-MFI(M-
tected (Si/Al > 900), corroborating that the products were zincosili- 12) (Fig. 6B), peaks at 533.3 and 531.1 eV were attributed to Si–O–Si
cates rather than zincoaluminosilicates. and Si–O–Zn bonds, respectively [45,46]. The peaks corresponding to
The as-made Zn-MFI(M-x) zeolites showed similar XRD patterns to ZnO, i.e., the Zn–O bond (529.9 eV) and O2− in the oxygen-deficient
that of the Zn-free Si-MFI sample (Fig. 4), indicating a crystalline MFI regions (531.6 eV) [47], were no longer observed.
structure. The diffraction peaks between 22.5° and 25° shifted to the The solid-state 29Si MAS NMR spectra of the Zn-MFI(M-x) samples
lower angle region after the incorporation of Zn, which was due to the are shown in Fig. S6, and two main signals were identified. As sug-
isomorphous substitution of tetrahedral Si atoms with Zn atoms pos- gested by previous papers regarding ZnO/SiO2 composites [48,49] and
sessing a larger radius and a longer Zn–O bond (1.95 Å) compared with zincosilicates [16], the peak at a resonance position of ca. −113 ppm
the Si–O bond (1.62 Å) for tetrahedral coordination [14,38]. Zn-MFI(M- could be attributed to Si(0Zn) silicon species, while the hump at ca.
12) with the highest Zn content required 3 days to obtain its highest −100 ppm could be assigned to Si(1Zn) silicon species, and its relative
crystallinity, which was longer than those for Zn-MFI(M-5) and Zn-MFI intensity was enhanced with the increase in Zn content. It has been
(M-8) (2 days) (Fig. S3). Moreover, the relative crystallinity also de- reported that the framework Si/Zn molar ratio can also be estimated
creased from 80% (Zn-MFI(M-5)) to 51% (Zn-MFI(M-12)) with the in- from the intensities of 29Si MAS NMR signals based on Eq. (2) as in the
crease in Zn content. This was reasonable because the nucleation rate case of aluminosilicate zeolites [15]. The values calculated for Zn-MFI
and crystal growth rate of MFI-type zeolite are susceptible to the nature (M-5), Zn-MFI(M-8), and Zn-MFI(M-12) were 39.0, 23.2, and 16.0, re-
of the heteroatoms present in the synthetic system [39]. The MFI to- spectively, and were comparable to the results of ICP-AES analyses
pological structure of Zn-MFI(M-x) zeolites was maintained after cal- (Table 1), suggesting that Zn was successfully incorporated into the
cination (Fig. S4). zeolite framework.
The DR UV–Vis spectra (Fig. 5) of Zn-MFI(M-x) did not show the
distinct absorption band derived from ZnO. Zn species were

Table 1
Chemical and textural properties of Si-MFI and Zn-MFI(M-x).
Si/Zn (ICP) Si/Zn (29Si NMR) Si/Zn (XPS) Zn contenta (wt%) SBET (m2g−1) Vmicrob (cm3g−1)

Si-MFI – – – – 390 0.13


Zn-MFI(M-5) 36.9 39.0 n.d.c 2.8 350 0.12
Zn-MFI(M-8) 21.4 23.2 122 4.7 290 0.10
Zn-MFI(M-12) 13.4 16.0 52 7.2 220 0.08

a
Based on ICP-AES results.
b
Calculated by t-plot method.
c
Not determined.

4
P. Hu, et al. Microporous and Mesoporous Materials 288 (2019) 109594

Fig. 4. XRD patterns of as-made Si-MFI and Zn-MFI(M-x) without calcination, and enlarged spectra in the 2 theta range of 22°–25.5°.

reflecting the microporous features of the samples. The specific surface


areas and micropore volumes decreased with the increase in Zn content
in the zeolites (Table 1). These results indicated that the incorporation
of Zn into the framework of MFI-type zeolites had a profound effect on
their textural properties, which caused declined crystallinity, as re-
vealed by the XRD patterns, and distortion of the channel structure.
Similar trends were also found in previous reports regarding Cu-, Zn-
and Ni-substituted ZSM-5 zeolites [21,22,42]. Although Zhang et al.
[50] claimed that incorporation of Sn in the ZSM-5 zeolite had a weak
influence on its physical structure, the Sn content was not mentioned,
and the presence of extra-pore Sn cluster species was proposed. It
should be noted that the heteroatom (Zn) contents in our products were
much higher than those in previously reported heteroatom-substituted
MFI-type zeolites (Table S1), and most of the Zn was present in the
framework according to the UV–Vis, FT-IR, XPS and solid-state 29Si
MAS NMR characterization results.

3.3. Effectiveness of Si–Zn oxide composites on synthesis of zincosilicate


zeolites

Fig. 5. DR UV–Vis spectra of Si-MFI and Zn-MFI(M-x). The crystallization of MFI-type zincosilicate zeolites using the me-
chanochemically treated Si–Zn oxide composite as a starting material
4 4 was further investigated. It was found that when Si–Zn oxide composite
nSi
nZn
= ∑ ISi(nZn)/ ∑ 0.25n⋅ISi(nZn) was used as the only silica source, a peak appeared at approximately 6°
n=0 n=0 (2) in the XRD pattern, which was probably attributed to the layered
structure caused by Zn [14], and the crystallization process was slow
The nitrogen adsorption/desorption isotherms of Si-MFI and Zn-MFI
(Fig. S8). Here, Na2SiO3 was employed as an additional Si source to
(M-x) (Fig. S7) could be designated as type I according to the classifi-
modulate the composition of the reactant and help build the zeolite
cation of International Union of Pure and Applied Chemistry (IUPAC),
skeleton. Furthermore, Na2SiO3 could provide alkalinity necessary for

Fig. 6. (A) Zn2p and (B) O1s core level XPS spectra of ZnO, Zn-MFI(M-8) and Zn-MFI(M-12). The background was subtracted by Shirley algorithm, and the spectra
were fitted with a Gauss-Lorentz function.

5
P. Hu, et al. Microporous and Mesoporous Materials 288 (2019) 109594

Fig. 7. FE-SEM images of (A) mechanochemically treated Si–Zn oxide composite M-12 and the products obtained during the hydrothermal synthesis of Zn-MFI(M-12)
at (B) 1 d, (C) 2 d and (D) 3 d.

products for different hydrothermal periods were close to the initial Si/
Zn molar ratio of the reactant, which was calculated as 14.1 (Fig. S9).
However, the surface Si/Zn molar ratio of the product with the highest
crystallinity (3 days) was 52 according to the XPS measurement, which
was much higher than its bulk composition (Table 1). This uneven
distribution of Zn, i.e., less Zn on the outer surface of the zeolite par-
ticles, was different from the case of aluminosilicate ZSM-5 zeolites
synthesized by the conventional TPA+-assisted hydrothermal method,
where Al comes from the liquid phase and is usually concentrated in the
rim portion of the crystal particles [51]. The FE-SEM image of M-12
(Fig. 7A) depicted that fine particles were obtained after the mechan-
ochemical treatment of the mixture of fumed silica and ZnO. Then, this
Si–Zn oxide composite was mixed well in the mortar with the pre-
cipitated silicic acid (H2SiO3), which formed from Na2SiO3 when pH
was 12, and became aggregated amorphous solids as observed in
Fig. 7B. During the hydrothermal treatment, the aggregated amorphous
solids underwent a condensation process together to build the frame-
work of Zn incorporated zeolite, and finally fully crystallized zeolite
particles with a silica-rich particle surface were obtained (Fig. 7C and
D).
Fig. 8. 2Na+→Co2+ exchange isotherms for C-ZSM-5, Zn-MFI(M-8), and Zn- Zincosilicate zeolite (Zn-MFI(H-8)) was also prepared from a SiO2/
MFI(M-12) at 70 °C.
ZnO mixture mixed by hand (H-8, Zn/(Si + Zn) = 8.2 mol%) instead of
from a mechanochemically treated Si–Zn oxide composite. The XRD
crystallization simultaneously. The pH of the reactant was adjusted to pattern (Fig. S10) showed that the diffraction peaks corresponding to
12 by the addition of NH4Cl. ZnO crystals were detected during hydrothermal treatment. Moreover,
In the case of Zn-MFI(M-12), the bulk Si/Zn molar ratios of the

Table 2
Chemical compositions of the products with the maximum Co exchange capacity.
Si/Zna Si/Ala 2Co/(2Zn + Al)a,b (2Co + Na)/(2Zn + Al)a,b Co content (wt%)

Zn-MFI(M-8) 23.1 (21.4)c


– 0.98 0.98 3.9
Zn-MFI(M-12) 14.3 (13.4)c – 0.99 1.07 6.2
C-ZSM-5 – 11.3 (11.9)c 0.89 0.89 3.5
(Zn, Al)-MFI 38.5 (25.2)c 16.6 (17.4)c 0.94 0.94 4.7

a
Determined by ICP-AES.
b
For Zn-MFI samples, Al = 0; while for C-ZSM-5, Zn = 0.
c
The value in the parenthesis was obtained before ion exchange.

6
P. Hu, et al. Microporous and Mesoporous Materials 288 (2019) 109594

Scheme 1. Ion exchange process on zincosilicate zeolites in the Co2+–Na+ binary ion exchange system. Zincosilicate zeolites display higher selectivity for Co2+
when both divalent Co2+ and monovalent Na+ are present.

only weak diffraction peaks attributed to the MFI topological structure than aluminosilicate analogs. In the case of Zn-MFI(M-12) (ρ = 0.14),
along with a hump derived from amorphous solids appeared after hy- Na+ cations (approximately 8% of ion exchange sites) initially present
drothermal treatment for 24 h. The crystallization process was appar- in the zeolites were not completely exchanged by Co2+ cations, prob-
ently hampered compared with that of Zn-MFI(M-8) because of the ably because of the distorted channel structure indicated by the ni-
existence of bulk ZnO. Although the crystallized product was obtained trogen sorption measurement, which was difficult for Co2+ cations to
after 48 h without detectable diffraction peaks of impurities in the XRD diffuse through to get to the ion exchange sites. Nevertheless, Zn-MFI
pattern, the presence of ZnO and extra-framework Zn2+ species was (M-12) still displayed selectivity for Co2+ in the low SCo range, and the
ascertained by the DR UV–Vis spectrum (Fig. S11) [19]. However, these maximum Co content of 6.2 wt% was exchanged into the zeolites,
absorption bands were not observed in the spectrum of Zn-MFI(M-12) which was much higher than that of C-ZSM-5 (3.5 wt%) (Table 2). The
that had a higher Zn content. These results strongly confirmed the Si/Zn molar ratios of Zn-MFI(M-8) and Zn-MFI(M-12) remained con-
crucial role of mechanochemically treated Si–Zn oxide composites in stant before and after Co2+ ion exchange, which manifested the good
the synthesis of MFI-type zincosilicate zeolites with high Zn contents stability of Zn in the framework. In contrast, Zn leakage occurred
and minor extra-framework Zn species. during Co2+ ion exchange when the reference sample of (Zn, Al)-MFI
Zn-containing MFI-type zeolite (zincoaluminosilicate (Zn, Al)-MFI) was used. The Si/Zn molar ratio of the zeolite noticeably increased from
was fabricated as a reference sample using the Zn complex ([Zn 25.2 to 38.5, while the Si/Al molar ratio remained unchanged
(NH3)4]2+) as the Zn source according to a conventional seed-assisted (Table 2), which was also an evidence of incomplete incorporation of
hydrothermal route (detailed synthetic procedures are described in Text Zn in the zeolite framework.
S1) [18]. It is noteworthy that the MFI-type zincosilicate zeolite could
not be obtained through this method without the addition of Al source
(Fig. S12). The DR UV–Vis spectrum signified the existence of extra- 4. Conclusions
framework Zn2+ species in the (Zn, Al)-MFI sample (Fig. S11). In the
hydrothermal process, soluble Zn species gradually reacted with early MFI-type zincosilicate zeolites with a negligible amount of Al were
formed aluminosilicates to form Si–O–Zn units and concomitantly in- prepared by using mechanochemically treated composites of fumed
corporated into the framework, during which part of the Zn species silica and ZnO as the starting materials. Zn was dispersed and in-
remained in the solution or became extra-framework species. In con- corporated into the silica matrix by mechanical forces before hydro-
trast, in this study, Zn was incorporated into the silica matrix by me- thermal treatment, and this strategy played a pivotal role in main-
chanochemical treatment before the hydrothermal process, which was taining Zn in the solid phase throughout the hydrothermal process and
beneficial for Zn to remain in the solid and form the zincosilicate fra- forming zincosilicate frameworks. The products had high Zn contents
mework. (up to 7.2 wt%) but minor amounts of extra-framework Zn species,
which were superior to those prepared by the hand-mixed SiO2/ZnO
mixture or the conventional hydrothermal method using a soluble Zn
3.4. Ion exchange performance
source. Ion exchange experiments revealed that MFI-type zincosilicate
zeolites exhibited superior selectivity and capacity for divalent cations,
Ion exchange equilibrium isotherms (Fig. 8) were plotted in terms of
like Co2+, compared with aluminosilicate analogs. This study success-
the equivalent fraction of the Co2+ cation in the solution
fully realized the incorporation of high amounts of Zn into highly si-
(SCo = 2[Co]s/(2[Co]s+[Na]s)) against that in the zeolite
liceous MFI-type zeolites, and our approach can be extended to the
(ZCo = 2[Co]z/(2[Co]z+[Na]z)). By assuming that all Zn atoms in the
fabrication of other heteroatom-substituted zeolites with improved
zeolites generated two negative charges per Zn atom, the charge den-
flexibility of the framework composition.
sities, ρ, defined as 2Zn/Si or Al/Si in molar ratio, for Zn-MFI(M-8)
(0.086) and the commercial aluminosilicate ZSM-5 (C-ZSM-5, 0.088)
were similar to each other. However, it was easier for the Co2+ cation Acknowledgements
to exchange Na+ on Zn-MFI(M-8) than on C-ZSM-5 when the Co con-
centration in solution was low (SCo < 0.4). Moreover, the maximum This study was supported in part by JST CREST Grant Number
Co2+ ion exchange capacity of Zn-MFI(M-8) reached 98%, while that of JPMJCR17P1, Japan. The high-energy X-ray total scattering experi-
C-ZSM-5 was 89%, even though Na+ was not detected in the zeolites ments at SPring-8 were approved by the Japan Synchrotron Radiation
(Table 2). These results implied that zincosilicate zeolites exhibited Research Institute (Proposal No. 2018A0155). P. Hu also thanks to
superior selectivity and ion exchange capacity for divalent cations China Scholarship Council for the financial support (File No.
compared with aluminosilicate analogs (Scheme 1). Because one Al 201706230216).
atom in the framework produces one negative charge, divalent cations
can only be stabilized by “Al pairs” (local Al–O–(Si–O)1,2–Al se-
quences), which are relatively difficult to form in Si-rich MFI-type Appendix A. Supplementary data
zeolites [52]. In contrast, one Zn in the framework with two negative
charges can hold one divalent cation. Therefore, zincosilicate zeolites Supplementary data to this article can be found online at https://
can have a higher ion exchange capacity at a lower substitution level doi.org/10.1016/j.micromeso.2019.109594.

7
P. Hu, et al. Microporous and Mesoporous Materials 288 (2019) 109594

References [28] C. Anand, T. Kaneda, S. Inagaki, S. Okamura, H. Sakurai, K. Sodeyama,


T. Matsumoto, Y. Kubota, T. Okubo, T. Wakihara, New J. Chem. 40 (2016)
492–496.
[1] M.S. Holm, S. Svelle, F. Joensen, P. Beato, C.H. Christensen, S. Bordiga, M. Bjørgen, [29] K. Yamamoto, T. Ikeda, C. Ideta, M. Yasuda, Cryst. Growth Des. 12 (2012)
Appl. Catal. Gen. 356 (2009) 23–30. 1354–1361.
[2] M. Hartmann, A.G. Machoke, W. Schwieger, Chem. Soc. Rev. 45 (2016) 3313–3330. [30] C. Kosanović, A. Čižmek, B. Subotić, I. Šmit, M. Stubičar, A. Tonejc, Zeolites 15
[3] Q.Y. Guan, D.Y. Wu, Y. Lin, X.C. Chen, X.Z. Wang, C.J. Li, S.B. He, H.N. Kong, J. (1995) 51–57.
Hazard Mater. 167 (2009) 244–249. [31] S.E.B. García, K. Yamamoto, F. Saito, A. Muramatsu, J. Jpn. Pet. Inst. 50 (2007)
[4] R. Yang, Z. Xu, S. Yang, I. Michos, L.-F. Li, A.P. Angelopoulos, J. Dong, J. Membr. 53–60.
Sci. 450 (2014) 12–17. [32] S.E. Borjas Garcia, K. Yamamoto, A. Muramatsu, J. Mater. Sci. 43 (2007)
[5] A. Corma, L.T. Nemeth, M. Renz, S. Valencia, Nature 412 (2001) 423–425. 2367–2371.
[6] H. Xin, J. Zhao, S. Xu, J. Li, W. Zhang, X. Guo, E.J.M. Hensen, Q. Yang, C. Li, J. [33] T. Iida, A. Takagaki, S. Kohara, T. Okubo, T. Wakihara, Chem. Nano. Mat. 1 (2015)
Phys. Chem. C 114 (2010) 6553–6559. 155–158.
[7] B. Tang, W. Dai, G. Wu, N. Guan, L. Li, M. Hunger, ACS Catal. 4 (2014) 2801–2810. [34] S. Bordiga, C. Lamberti, G. Ricchiardi, L. Regli, F. Bonino, A. Damin, K.P. Lillerud,
[8] G.I. Panov, G.A. Sheveleva, A.S. Kharitonov, V.N. Romannikov, L.A. Vostrikova, M. Bjorgen, A. Zecchina, Chem. Commun. (2004) 2300–2301.
Appl. Catal. Gen. 82 (1992) 31–36. [35] J. Xie, Y. Li, W. Zhao, L. Bian, Y. Wei, Powder Technol. 207 (2011) 140–144.
[9] T. Iida, M. Sato, C. Numako, A. Nakahira, S. Kohara, T. Okubo, T. Wakihara, J. [36] C.L. Farrow, P. Juhas, J.W. Liu, D. Bryndin, E.S. Bozin, J. Bloch, T. Proffen,
Mater. Chem. 3 (2015) 6215–6222. S.J. Billinge, J. Phys. Condens. Matter 19 (2007) 335219.
[10] A. Tuel, S. Moussa-Khouzami, Y.B. Taarit, C. Naccache, J. Mol. Catal. 68 (1991) [37] T. Iida, K. Ohara, Y. Roman-Leshkov, T. Wakihara, Phys. Chem. Chem. Phys. 20
45–52. (2018) 7914–7919.
[11] C.M. Osmundsen, M.S. Holm, S. Dahl, E. Taarning, Proc. R. Soc. A 468 (2012) [38] M.A. Camblor, M.E. Davis, J. Phys. Chem. 98 (1994) 13151–13156.
2000–2016. [39] R. Xu, W. Pang, Stud. Surf. Sci. Catal. 24 (1986) 27–38.
[12] M.A. Deimund, J. Labinger, M.E. Davis, ACS Catal. 4 (2014) 4189–4195. [40] Y. Meng, H.C. Genuino, C.H. Kuo, H. Huang, S.Y. Chen, L. Zhang, A. Rossi, S.L. Suib,
[13] M. Orazov, M.E. Davis, Chem. Sci. 7 (2016) 2264–2274. J. Am. Chem. Soc. 135 (2013) 8594–8605.
[14] N. Koike, K. Iyoki, S.H. Keoh, W. Chaikittisilp, T. Okubo, Chem. Eur J. 24 (2018) [41] S.R. Tomlinson, T. McGown, J.R. Schlup, J.L. Anthony, Int. J. Spectrosc. (2013) 1–7
808–812. 2013.
[15] M.A. Camblor, R.F. Lobo, H. Koller, M.E. Davis, Chem. Mater. 6 (1994) 2193–2199. [42] G. Vitale, H. Molero, E. Hernandez, S. Aquino, V. Birss, P. Pereira-Almao, Appl.
[16] T. Takewaki, L.W. Beck, M.E. Davis, Top. Catal. 9 (1999) 35–42. Catal. Gen. 452 (2013) 75–87.
[17] W.J. Ball, S.A.I. Barrl, S. Cartlidge, B.M. Maunders, D.W. Walker, Stud. Surf. Sci. [43] H. Li, S. Jiao, J. Ren, H. Li, S. Gao, J. Wang, D. Wang, Q. Yu, Y. Zhang, L. Li, Phys.
Catal. 28 (1986) 951–956. Chem. Chem. Phys. 18 (2016) 4144–4153.
[18] L. Wang, S. Sang, S. Meng, Y. Zhang, Y. Qi, Z. Liu, Mater. Lett. 61 (2007) [44] J.-H. Hong, Y.-F. Wang, G. He, J.-X. Wang, J. Non-Cryst. Solids 356 (2010)
1675–1678. 2778–2780.
[19] N. Koike, K. Iyoki, B. Wang, Y. Yanaba, S.P. Elangovan, K. Itabashi, W. Chaikittisilp, [45] C.W. Lopes, P.H. Finger, M.L. Mignoni, D.J. Emmerich, F.M.T. Mendes, S. Amorim,
T. Okubo, Dalton Trans. 47 (2018) 9546–9553. S.B.C. Pergher, Microporous Mesoporous Mater. 213 (2015) 78–84.
[20] M. Dong, J. Wang, Y. Sun, Microporous Mesoporous Mater. 43 (2001) 237–243. [46] Y. Liu, J. Shen, Z. Chen, L. Yang, Y. Liu, Y. Han, Appl. Catal. Gen. 403 (2011)
[21] E. Yuan, W. Han, G. Zhang, K. Zhao, Z. Mo, G. Lu, Z. Tang, Catal. Surv. Asia 20 112–118.
(2016) 41–52. [47] A. Hastir, N. Kohli, R.C. Singh, J. Phys. Chem. Solids 105 (2017) 23–34.
[22] E. Yuan, K. Zhang, G. Lu, Z. Mo, Z. Tang, J. Ind. Eng. Chem. 42 (2016) 142–148. [48] J. Qu, C.-Y. Cao, Y.-L. Hong, C.-Q. Chen, P.-P. Zhu, W.-G. Song, Z.-Y. Wu, J. Mater.
[23] C.I. Round, C.D. Williams, K. Latham, C.V.A. Duke, Chem. Mater. 13 (2001) Chem. 22 (2012) 3562.
468–472. [49] X. Collard, M. El Hajj, B.-L. Su, C. Aprile, Microporous Mesoporous Mater. 184
[24] M. Mohapatra, S. Anand, B.K. Mishra, D.E. Giles, P. Singh, J. Environ. Manag. 91 (2014) 90–96.
(2009) 67–77. [50] Y. Zhang, Y. Zhou, L. Huang, M. Xue, S. Zhang, Ind. Eng. Chem. Res. 50 (2011)
[25] T. Inui, S.-B. Pu, J.-i. Kugai, Appl. Catal. Gen. 146 (1996) 285–296. 7896–7902.
[26] K. Yamamoto, S.E. Borjas García, A. Muramatsu, Microporous Mesoporous Mater. [51] R. von Ballmoos, W.M. Meier, Nature 289 (1981) 782–783.
101 (2007) 90–96. [52] J. Dědeček, D. Kaucký, B. Wichterlová, O. Gonsiorová, Phys. Chem. Chem. Phys. 4
[27] H. Yamada, T. Iida, Z. Liu, Y. Naraki, K. Ohara, S. Kohara, T. Okubo, T. Wakihara, (2002) 5406–5413.
Cryst. Growth Des. 16 (2016) 3389–3394.

You might also like