introduction · various heteroatoms into the framework of different zeo-lites has received...

8
Introduction 1) Isomorphous substitution of silicon and/or aluminum by various heteroatoms into the framework of different zeo- lites has received considerable attention in the utilization of zeolites as catalysts. These elements can impart cata- lytic properties to the zeolite for many reactions other than those catalyzed by acid sites while offering the ad- vantages of atomic dispersion of the elements, which is difficult to achieve with traditional catalyst preparation methods. In this respect, isomorphous substitution of tita- nium, zirconium, or vanadium into the framework of high-silica zeolites has expanded the scope of application of these materials. These isomorphously substituted zeo- lites are reported to be very promising in the liquid-phase catalytic oxidation of hydrocarbons. Taramasso and cow- To whom all correspondence should be addressed. (e-mail: [email protected]) orkers [1] first reported the hydrothermal synthesis of a titanium-containing zeolite, denoted as titanium silica- lite-1 (TS-1). Titanium silicalite-1 was found to be active in the oxidation of a variety of organic substrates in the presence of aqueous hydrogen peroxide as oxidizing agent. Much the same as titanium-substituted molecular sieves, zirconium or vanadium silicalite-1 (ZS-1 or VS- 1) can be synthesized by isomorphous substitution of Zr 4+ or V 5+ for Si 4+ in the MFI structure framework, and has very interesting properties towards catalytic oxida- tion [2-7]. The oxidative property of vanadium could be combined with the shape-selective characteristics im- parted by the micropore structure in which the vanadium is located [8]. Niobium - which belongs to the same group in the Periodic Table as vanadium (Group V) - has been re- ported to have the capability of the photocatalytic and se- lective oxidations [9-11]. Although electronegativity and ionic radius between niobium and its neighbors (V, Mo, Yong Sig Ko , Hyun Tae Jang , and Wha Seung Ahn Department of Advanced Material Chemistry, Shinsung College, Dangjin-gun 343-860, Korea *Department of Chemical Engineering, Hanseo University, Seosan 356-706, Korea **Department of Chemical Engineering, College of Engineering, Inha University, Inchon 402-751, Korea Received January 29, 2007; Accepted May 3, 2007 Abstract: Niobium silicalite-1 (NbS-1) molecular sieves with MFI structure were prepared using the hydro- thermal synthesis method. X-ray diffraction, scanning and transmission electron microscopies, Fourier transform infrared spectroscopy, ultraviolet-visible (UV-vis) diffuse reflectance spectroscopy, 29 Si magic-angle spinning nuclear magnetic resonance spectroscopy, ammonia temperature-programmed desorption (NH 3 -TPD), physical adsorption of nitrogen and elemental analysis were then performed to evaluate their physico-chemical properties and to provide evidences for the incorporation of Nb 5+ into the zeolite framework. The unit cell volume of niobi- um silicalite-1 increased linearly with increasing niobium content, suggesting isomorphous substitution of Si 4+ by Nb 5+ in the lattice framework. The framework infrared spectrum showed a characteristic absorption band at 963 cm -1 , probably due to Si-O-Nb linkages. In the UV-vis spectrum, the absorption assigned to the presence of Nb 5+ in the zeolite framework was observed at around 200 nm. In addition, NH 3 -TPD studies revealed weak acidity in niobium silicalite-1 due possibly to [Nb (OH)-Si] sites in the framework. ․․․ Keywords: niobium, niobium silicalite-1, hydrothermal synthesis, NH 3 -TPD, isomorphous substitution, charac- terization, FT-IR, UV-vis DRS

Upload: others

Post on 21-Oct-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

  • Introduction1)

    Isomorphous substitution of silicon and/or aluminum by

    various heteroatoms into the framework of different zeo-

    lites has received considerable attention in the utilization

    of zeolites as catalysts. These elements can impart cata-

    lytic properties to the zeolite for many reactions other

    than those catalyzed by acid sites while offering the ad-

    vantages of atomic dispersion of the elements, which is

    difficult to achieve with traditional catalyst preparation

    methods. In this respect, isomorphous substitution of tita-

    nium, zirconium, or vanadium into the framework of

    high-silica zeolites has expanded the scope of application

    of these materials. These isomorphously substituted zeo-

    lites are reported to be very promising in the liquid-phase

    catalytic oxidation of hydrocarbons. Taramasso and cow-

    †To whom all correspondence should be addressed.

    (e-mail: [email protected])

    orkers [1] first reported the hydrothermal synthesis of a

    titanium-containing zeolite, denoted as titanium silica-

    lite-1 (TS-1). Titanium silicalite-1 was found to be active

    in the oxidation of a variety of organic substrates in the

    presence of aqueous hydrogen peroxide as oxidizing

    agent. Much the same as titanium-substituted molecular

    sieves, zirconium or vanadium silicalite-1 (ZS-1 or VS-

    1) can be synthesized by isomorphous substitution of

    Zr4+or V

    5+for Si

    4+in the MFI structure framework, and

    has very interesting properties towards catalytic oxida-

    tion [2-7]. The oxidative property of vanadium could be

    combined with the shape-selective characteristics im-

    parted by the micropore structure in which the vanadium

    is located [8].

    Niobium - which belongs to the same group in the

    Periodic Table as vanadium (Group V) - has been re-

    ported to have the capability of the photocatalytic and se-

    lective oxidations [9-11]. Although electronegativity and

    ionic radius between niobium and its neighbors (V, Mo,

    Yong Sig Ko†, Hyun Tae Jang , andWha Seung Ahn

    Department of Advanced Material Chemistry, Shinsung College, Dangjin-gun 343-860, Korea

    *Department of Chemical Engineering, Hanseo University, Seosan 356-706, Korea

    **Department of Chemical Engineering, College of Engineering, Inha University, Inchon 402-751, Korea

    Received January 29, 2007; Accepted May 3, 2007

    Abstract:Niobium silicalite-1 (NbS-1) molecular sieves with MFI structure were prepared using the hydro-

    thermal synthesis method. X-ray diffraction, scanning and transmission electron microscopies, Fourier transform

    infrared spectroscopy, ultraviolet-visible (UV-vis) diffuse reflectance spectroscopy,29Si magic-angle spinning

    nuclear magnetic resonance spectroscopy, ammonia temperature-programmed desorption (NH3-TPD), physical

    adsorption of nitrogen and elemental analysis were then performed to evaluate their physico-chemical properties

    and to provide evidences for the incorporation of Nb5+into the zeolite framework. The unit cell volume of niobi-

    um silicalite-1 increased linearly with increasing niobium content, suggesting isomorphous substitution of Si4+

    by Nb5+in the lattice framework. The framework infrared spectrum showed a characteristic absorption band at

    963 cm-1, probably due to Si-O-Nb linkages. In the UV-vis spectrum, the absorption assigned to the presence of

    Nb5+

    in the zeolite framework was observed at around 200 nm. In addition, NH3-TPD studies revealed weak

    acidity in niobium silicalite-1 due possibly to [Nb (OH)-Si] sites in the framework.․․․

    Keywords: niobium, niobium silicalite-1, hydrothermal synthesis, NH3-TPD, isomorphous substitution, charac-

    terization, FT-IR, UV-vis DRS

  • Hydrothermal Synthesis and Characterization of Niobium-containing Silicalite-1 Molecular Sieves with MFI Structure 765

    Ta, and Zr) in the Periodic Table are different, it is re-

    ported that the promoter effect, support effect and acidic

    nature of niobium compounds are quite different from

    those of compounds of the surrounding elements [12].

    Niobium compounds and niobium materials have re-

    cently been shown to enhance catalytic properties when

    used as a component of catalysts or when small amounts

    are added to known catalysts. Depending on the catalyst

    composition and structure, they are used in the processes

    involving acidic active centers as well as redox character.

    These niobium-containing catalysts have been used for

    dehydration and dehydrogenation of alcohols [13-15],

    oxidative dehydrogenation of alkanes [16], alkylation of

    benzene [17], NO reduction by NH3 [18], dimerization

    and oligomerization of olefins [19], oxidation of meth-

    anol [20,21] and polycondensation reactions [22]. The

    presence of niobium in molecular sieves has the potential

    of generating catalysts with shape-selective properties

    and acidic or redox characteristics. Only a few reports

    are available on porous niobium silicates; they include

    work on the incorporation of niobium in micro/meso-

    porous materials, ETS-10 [23] denoted as ETNbS-10, to

    MFI [24] designated as NbS-1, and also to MCM-41

    [25].

    Though synthesis methods of niobium silicate molec-

    ular sieves have been reported [24,26], further inves-

    tigation concerning the hydrothermal synthesis of niobi-

    um-containing silicalite molecular sieves with more de-

    tailed characterization would contribute towards a better

    understanding of the system.

    In this work, we report the hydrothermal synthesis and

    characterization of crystalline niobium silicalite molec-

    ular sieves with the MFI structure prepared using silicon

    and niobium alkoxides. Results from X-ray diffraction

    (XRD), scanning electron microscopy (SEM), transmi-

    ssion electron microscopy (TEM), framework infrared

    (IR) spectroscopy, ultraviolet-visible (UV-vis) diffuse re-

    flectance spectroscopy (DRS),29Si magic-angle spinning

    (MAS) nuclear magnetic resonance spectroscopy

    (NMR), temperature-programmed desorption of ammo-

    nia (NH3-TPD), nitrogen adsorption and elemental analy-

    sis are discussed to verify the incorporation of niobium

    in the zeolite lattice framework.

    Experimental

    Synthesis of niobium silicalite-1 catalyst

    NbS-1 samples were prepared from substrates having

    the following composition ratio: SiO2/Nb2O5 = 32.7∼

    99, TPA+/SiO2 = 0.46, H2O/SiO2 = 35. In a typical prepa-

    ration, a Pyrex beaker was placed in a glove box and ni-

    trogen was flushed over to minimize the adverse effect of

    moisture in air. After 27.01 g of tetraethylorthosilicate

    Figure 1. Schematic diagram for the hydrothermal synthesis of

    niobium silicalite-1.

    (Aldrich, 99.999 %) was transferred to the Pyrex beaker

    and vigorously stirred, 0.43 g of niobium(V) ethoxide

    (Aldrich, 99.95 %) was carefully introduced into this sol-

    ution, and the mixture was cooled to about 273 K. After

    a few minutes, 60.65 g of tetrapropylammonium hydrox-

    ide (TPAOH) aqueous solution (Aldrich, 20 %), also

    cooled to 273 K, was slowly added dropwise into the

    mixture. Stirring and cooling were maintained during this

    process. After addition of all TPAOH the synthesis mix-

    ture was kept for 5 6 h at 343 353 K in order to accel∼ ∼ -

    erate hydrolysis and to evaporate the ethyl alcohol

    produced. Deionized water was then added to increase

    the volume of the mixture to its original value. This clear

    homogeneous solution was then transferred to a Teflon-

    lined stainless steel autoclave and kept in a convection

    oven at 448 K under autogenous pressure. In order to in-

    vestigate the crystallization process, autoclaves were tak-

    en out from the oven at different time intervals and were

    quenched immediately in cold water for sample identifi-

    cation. The solid products were separated by means of

    suction-filtration or centrifugation, washed several times

    with hot deionized water and dried in an air oven at 383

    K overnight. The products were finally calcined at 823 K

    for 6 h. The synthesis procedure of niobium silicalite-1

    catalyst can be represented schematically as in Figure 1.

    A reference sample of silicalite-1 (MFI) was prepared

    following the same procedure without adding niobium

    ethoxide to the substrate mixture. In addition, a sample

  • Yong Sig Ko, Hyun Tae Jang, and Wha Seung Ahn766

    Figure 2. XRD patterns of niobium silicalite-1 and silicalite-

    1 (a) silicalite-1, (b) ZSM-5, (c) NbS-1 (3 mol% niobium),:

    and (d) Nb2O5 powder.

    of about 3 mol% niobium-impregnated silicalite-1 [Nb/

    silicalite-1 (IMP)] was also prepared for comparison pur-

    poses by the incipient wetness method.

    Characterization

    The samples synthesized were analyzed by X-ray dif-

    fraction (XRD) for both qualitative and quantitative

    phase identification. The unit used was a powder X-ray

    diffractometer (Philips, PW-1700) with a scintillation

    counter and a graphite monochromator attachment, uti-

    lizing nickel-filtered CuK radiation (40 kV, 25 mA).α

    Unit cell parameters were obtained by a least-squares fit

    to the interplanar spacings measured in the 5 45° (2 )∼ θ

    angular region, using silicon as an internal standard. The

    crystal size and morphology of the crystalline samples

    were examined using a scanning electron microscope

    (Hitachi, X-650) after coating with a Au-Pd evaporated

    film. For TEM analysis, samples were dispersed ultra-

    sonically in ethanol, and a drop of the suspension was

    deposited on a holey carbon copper grid. Micrographs

    were recorded with a 1024 × 1024 Gatan CCD camera on

    a Philips CM20 microscope operated at 200 kV. Frame-

    work infrared (IR) spectra of samples were recorded in

    air at room temperature on a Perkin Elmer 221 spec-

    trometer (in the range of 400 4000 cm∼-1) with wafers of

    zeolites mixed with dry KBr. The elemental analyses of

    crystalline samples were performed with an inductively

    coupled plasma (ICP) spectrometer (Jobin Yuon, JY-38

    VHR) and X-ray fluorescence (XRF; Rigaku, 3070).

    UV-vis diffuse reflectance spectroscopy was performed

    on a Varian CARY 3E double-beam spectrometer and

    dehydrated MgO was used as a reference in the range

    190 500 nm. The solid-state NMR spectra were ob∼ -

    tained with a Bruker AM 300 spectrometer at a fre-

    quency of 59.6 MHz and spinning rate of 3.5 kHz with a

    pulse width of 3 µs, a relaxation delay of 5 s and 100∼

    200 acquisitions. The chemical shift was referenced with

    respect to the29Si signal of tetramethylsilane (TMS). The

    nitrogen-adsorption isotherms and specific surface areas

    were determined by nitrogen physisorption with the

    Brunauer-Emmett-Teller (BET) method at liquid-nitro-

    gen temperature using a Micromeritics ASAP 2010 auto-

    matic analyzer. Ammonia temperature-programmed de-

    sorption (NH3-TPD) tests were carried out in a quartz mi-

    croreactor with continuous analysis of the released NH3concentration via a TCD detector using a Micromeritics

    TPD/TPR 2900 analyzer. Before the tests, the sample

    was pretreated in situ with a flow of argon at 773 K in or-

    der to remove adsorbed species, and the adsorption was

    conducted after the sample was cooled to 373 K. The

    physisorbed NH3 was flushed by a dry argon flow at the

    same temperature for 3 h. The temperature was then in-

    creased at the rate of 10 K min-1.

    Results and Discussion

    The X-ray diffraction patterns of calcined NbS-1, ZSM-

    5, silicalite-1, and N2O5 powder are shown in Figure 2.

    The X-ray diffraction pattern of the calcined NbS-1 mo-

    lecular sieve was similar to that of ZSM-5 zeolite with

    MFI structure. Almost identical diffractograms of NbS-1

    were obtained irrespective of the niobium contents of the

    sample up to 3 mol% and no detectable niobium oxide

    peaks were observed. The symmetry of the calcined

    NbS-1 was orthorhombic, comprising single reflections

    around 2 at 24.5θoand 29.2

    o. As expected, calcined sili-

    calite-1 had monoclinic symmetry which presents double

    reflections at the same 2θ. The persistence of the ortho-

    rhombic symmetry in the calcined state of the niobium

    silicalites can be considered as a result of niobium ions

    present in the framework positions in NbS-1.

    The isomorphous substitution of Si4+

    by larger Nb5+

    ions in the NbS-1 framework causes a slight expansion

    of the unit cell parameters and volume. Figure 3 shows

    the variation of unit cell volume calculated for calcined

    NbS-1 samples with different niobium contents. The unit

    cell volume increases linearly with increasing niobium

    content of the reaction mixture up to 3 mol%. The in-

    crease in unit cell volume is due to the introduction of

    larger Nb5+

    ion in the framework lattice. A similar ex-

    pansion of the lattice has also been reported for iso-

    morphous substitution of Si4+

    by Ti4+, Zr

    4+or V

    5+in the

    silicalite framework lattice [1,2,5]. Again, linear in-

    creases in the unit cell volume, V, with increases in the

    amount (mol%) of niobium suggest the presence of

  • Hydrothermal Synthesis and Characterization of Niobium-containing Silicalite-1 Molecular Sieves with MFI Structure 767

    Figure 3. Unit cell volume of NbS-1 as a function of niobium

    content.

    Figure 4. Comparison between the crystallization rates of TS-1,

    ZrS-1 and NbS-1.

    niobium in the framework. Earlier studies [27,28] have

    shown that the expansion of the unit cell volume is re-

    lated to the concentration of the framework vanadium in

    the V5+containing zeolite samples.

    The overall crystallization rates at constant gel compo-

    sition and temperature for TS-1, Zr-containing zeolite

    ZrS-1 [29] and NbS-1 are compared in Figure 4. While

    the crystallization time was kept to 3 or 4 days to obtain

    the maximum crystallinity for TS-1 or ZrS-1, it was nec-

    essary for NbS-1 sample to be kept to 10 days in order to

    obtain the same degree crystallinity. The fact that a rela-

    tively longer crystallization time was necessary to main-

    tain the same degree of crystallinity for NbS-1 may be a

    consequence of the radius of niobium atoms being larger

    than that of titanium or zirconium atoms, and conse-

    quently it is believed to be more difficult for niobium to

    Figure 5. Relationship between the niobium content of the

    NbS-1 samples and the gel mixtures.

    enter the silicalite framework than titanium or zirconium.

    In addition, while ZSM-5 (which has the same MFI

    structure) can be prepared in 1 day, it takes substantially

    longer when titanium, zirconium or niobium is sub-

    stituted into the framework. Such a trend is generally ob-

    served in the synthesis of metal-substituted silicate mo-

    lecular sieves [29-31], but it can be also considered as a

    consequence of the mineralizing agent NaOH/KOH be-

    ing absent in the reaction mixture in such cases.

    We have synthesized a series of NbS-1 samples by

    varying the amount of niobium in the synthesis gel.

    Figure 5 shows the bulk niobium contents of NbS-1 crys-

    tals plotted against the niobium content of their reaction

    mixtures. It is seen that the niobium content of the NbS-1

    crystals obtained is linearly correlated with the niobium

    content of the reaction mixtures. This indicates that vary-

    ing amounts of niobium can be incorporated in the crys-

    tal by changing the niobium content at the stage of gel

    preparation, and most of the niobium present in the sub-

    strate can be incorporated into the zeolite framework

    within the range of less than about 3 mol%.

    A scanning electron micrograph and transmission elec-

    tron micrograph of NbS-1 (3 mol% niobium) sample are

    presented in Figure 6. The niobium content in the sub-

    strate mixture did not lead to a significant modification

    in the morphology of NbS-1. All NbS-1 samples were

    made up of uniform crystals of size about 0.13 µm and

    hexagonal shape. The transmission electron micrograph

    of the NbS-1 sample shows the absence of amorphous

    matters outside the crystals of NbS-1.

    IR spectra of the calcined silicalite-1 and NbS-1 are

    compared in Figure 7. As is known, IR spectra in the

    mid-infrared region can provide evidence of the iso-

    morphous substitution of large heteroatoms into the

    framework of zeolites [32,33]. The band around 1100

  • Yong Sig Ko, Hyun Tae Jang, and Wha Seung Ahn768

    (a) (b)

    Figure 6. SEM photograph (a) and TEM image (bar = 100 nm) (b) of the calcined niobium silicalite-1 (3 mol% niobium).

    Figure 7. IR spectra of the calcined (a) silicalite-1 and (b)

    NbS-1 (3 mol%).

    cm-1corresponds to asymmetric vibrations of internal tet-

    rahedra, and the symmetric stretching of T-O is found

    around 820 780 cm∼-1[34]. The IR spectrum of NbS-1

    showed a characteristic absorption band at 963 cm-1,

    which may be ascribed to Si-O-Nb stretching vibrations.

    This band was not observed in the IR spectra of calcined

    ZSM-5, pure silicalite-1 or in Nb2O5 powder. An absorp-

    tion band near 950 970 cm∼-1is usually known as char-

    acteristic of metal-oxygen stretching vibrations [27,

    35,36] or it has been interpreted to be a consequence of

    the Si-O-vibrations [32] perturbed by metal ions nearby

    in the zeolite framework. Similar IR results have been re-

    ported earlier for isomorphous substitution of Si4+

    by

    Ti4+, Zr

    4+, and V

    5+in the silicalite framework [1,2,8] ; in

    all these cases, an IR band around 963 or 967 cm-1, not

    Figure 8. UV-vis diffuse reflectance spectra of (a) silicalite-1,

    (b) NbS-1, (c) Nb2O5 powder and (d) Nb/silicalite-1 (IMP).

    present in the titanium-, zirconium- or vanadium-free an-

    alog was observed.

    The UV-vis diffuse reflectance spectra of the calcined

    NbS-1, Nb/silicalite-1 (IMP), pure silicalite-1 and Nb2O5powder are shown in Figure 8. Quite similar spectra were

    obtained among NbS-1 samples with niobium content up

    to 3 mol%. The absorption band of the NbS-1 sample

    was significantly different from the band pattern of

    Nb/silicalite-1 (IMP) or Nb2O5 powder. The UV-vis dif-

    fuse reflectance spectrum of NbS-1 exhibited a strong

    transition band around 200 nm, similar to the band at ∼

    210 nm of titanium- or zirconium-containing silicalite

    [4,32]. The band around 210 nm is a chargetransfer band

    arising from excitation of an oxygen 2p electron in the

    valance band to the empty d orbitals of titanium [6,36] or

    zirconium [4] ions in the framework of zeolites. In the ti-

    tanium or zirconium silicalite, the band around 210 nm

    has been assigned to isolated titanium or zirconium in

    zeolite lattice framework [4,32]. The pure silicalite-1

    sample showed no corresponding signals. Thus, we sug-

  • Hydrothermal Synthesis and Characterization of Niobium-containing Silicalite-1 Molecular Sieves with MFI Structure 769

    Figure 9.29Si MAS NMR spectra of the calcined (a) silicalite-1

    and (b) NbS-1.

    gest that the absorption band around 200 nm for NbS-1 is

    due to Si-O-Nb units in the framework. The main absorp-

    tion band of Nb2O5 powder was observed at 310 nm,∼

    and the absorption band of Nb/silicalite-1 (IMP) was

    very similar to that of the Nb2O5 powder. The absence of

    this band at 310 nm in NbS-1 samples prepared further∼

    confirms the absence of occluded Nb2O5 in the frame-

    work.

    Figure 9 shows the solid state29Si MAS NMR spectra

    of the calcined NbS-1 and pure silicalite-1 samples. The29Si MAS NMR spectrum of the NbS-1 was very similar

    to that of TS-1, ZrS-1 or VS-1 [37-39]. The NMR spec-

    trum of the NbS-1 displays a resonance at about -103

    ppm, a main signal at -113 ppm and a shoulder at about

    -116 ppm. Following the discussion in literature [40,41]

    the main peak at -113 ppm can be assigned to the central

    Si atom (in bold type) in Si(OSi)4 (Q4) sites based on the

    chemical shift. The29Si MAS NMR spectrum of NbS-1

    showed a characteristic shoulder around -116 ppm which

    did not appear in the case of pure silicalite-1. In an ear-

    lier study on TS-1, this signal was assigned to (TiO)Si

    (OSi)3 units and cited as evidence for framework sub-

    stitution of titanium [37]. Thus, we suggest that the sh-

    oulder at about -116 ppm of the NbS-1 is presumably at-

    tributed to the distorted silicon environment in tetrahe-

    dral due to Si-O-Nb bond. Axon and Klinowski [41] re-

    ported that the signal at about -103 ppm is indicative of

    Si(OSi)3O-(Q

    3unit) framework defects. This seems rea-

    sonable in particular in the case of niobium silicalite-1

    where at least a partial incorporation of the large Nb5+

    ions should lead to structural distortions and defects of

    Figure 10. Nitrogen-adsorption isotherms of the calcined silica-

    lite-1 and NbS-1.

    Figure 11. Specific surface area (BET) of NbS-1 with different

    niobium contents.

    the silicalite structure.

    Figure 10 shows the nitrogen adsorption-desorption iso-

    therms of NbS-1 and pure silicalite-1 samples calcined at

    823 K in N2. Both isotherms were of type I frequently

    found in zeolitic sorbents. As can be seen, the nitrogen-

    adsorption capacity of NbS-1 was very similar to that of

    the pure silicalite-1 and indicated the absence of oc-

    cluded matter like Nb2O5 within the pore system. The

    shapes of the two isotherms were also similar except for

    the presence of a large high-pressure hysteresis loop in

    silicalite-1, which results in a lower micropore volume.

    Figure 11 shows the specific surface area (BET) of

    NbS-1 with different amounts (mol%) of niobium in the

    reaction mixture. The specific surface area of NbS-1 was

    about 504.5 m2g-1and remained almost constant up to 3

    mol%. Therefore, it could be speculated that no occluded

    niobium species are present in the samples prepared.

    From the results of N2-adsorption measurements on

    NbS-1 samples, it could be seen that the majority of the

    niobium up to 3 mol% in the reaction mixture was sub-

    stituted isomorphously into the zeolite framework.

    In order to study the potential development of acidity in

    the niobium silicalite-1 sample, NH3-TPD tests were car-

  • Yong Sig Ko, Hyun Tae Jang, and Wha Seung Ahn770

    Figure 12. NH3-TPD profiles for ammonia desorption from (a)

    silicalite-1, (b) Nb/silicalite-1 (IMP), and (c) NbS-1.

    ried out on the pure silicalite-1, Nb/silicalite-1 (IMP) and

    NbS-1. The results obtained are shown in Figure 12. As

    expected, acid sites were almost absent in pure silica-

    lite-1. A broad NH3 desorption peak positioned between

    400 and 600 K was observed in the NbS-1 sample. This

    corresponds to the weak acid sites, which may be attrib-

    uted to silanol groups located near to niobium sites [Nb․․․

    (OH)-Si] or due to defect site created within the zeolite

    structure as Nb-OH. On the other hand, the Nb/silica-

    lite-1 (IMP) produced a far smaller desorption peak cen-

    tered near 470 K, despite almost the same niobium∼

    content. This desorption peak is probably associated with

    Nb-OH of the extraframework niobium oxides in the

    sample. The NH3-TPD results further prove that Nb5+

    species in the niobium silicalite-1 sample were incorpo-

    rated into the silicalite framework.

    Conclusion

    Niobium silicalite-1 (NbS-1) molecular sieves with the

    MFI structure have been synthesized hydrothermally,

    isomorphously substituting Nb5+

    for the framework Si4+

    of silicalite. Uniform-sized (0.13 µm) and hexagonal-

    shaped crystals were obtained, and the unit cell volume

    of NbS-1 lattice was found to increase linearly with in-

    creasing niobium content of the substrate. Framework IR

    spectrum showed a characteristic absorption band at

    around 963 cm-1, probably due to Si-O-Nb linkages. In

    the UV-vis spectrum, the absorption around 200 nm,

    which was assigned to the presence of Nb5+in the zeolite

    framework, was observed for the NbS-1. The29Si MAS

    NMR spectrum of NbS-1 exhibited a well-defined shoul-

    der around -116 ppm due to the distorted Si environment.

    Finally, NH3-TPD studies revealed a development of

    mild acid sites in NbS-1. All the diverse characterization

    methods employed in this study confirmed the structural

    incorporation of niobium in NbS-1.

    Acknowledgment

    This research was supported by a grant (code DD2-101)

    from Carbon Dioxide Reduction & Sequestration Re-

    search Center, one of the 21st Century Frontier Programs

    funded by the Ministry of Science and Technology of

    Korean government.

    References

    1. U.S. Patent 4, 410, 501 (1983).

    2. M. K. Dongare, P. Singh, P. P Moghe, and P.

    Ratnasamy, Zeolites, 11, 690 (1991).

    3. R. Fricke, H. Kosslick, V. A. Tuan, I. Grohmann, W.

    Pilz, W. Storek, and G. Walther, Zeolites and

    Microporous Crystals, Studies in Surface Science

    and Catalysis, T. Hattori and T. Yashima Eds., 83,

    p. 57, Elsevier, Amsterdam (1993).

    4. G.-R. Wang, X.-Q. Wang, X.-S. Wang, and S.-X.

    Yu, Zeolites and Microporous Crystals, Studies in

    Surface Science and Catalysis, T. Hattori and T.

    Yashima Eds., 83, p. 67, Elsevier, Amsterdam

    (1993).

    5. P. R. H. Prasad Rao, A. A. Belhekar, S. G. Hegde,

    A. V. Ramaswamy, and P. Ratnasamy, J. Catal.,

    141, 595 (1993).

    6. A. Tuel and Y. Ben Taarit, Zeolites, 14, 18 (1994).

    7. (a) T. Sen, M. Chatterjee, and S. Sivasanker, J.

    Chem. Soc., Commun., 207 (1995); (b) H. J. Jo, B. J.

    Ahn, M. S. Gang, S. B. Ko, and W. Chang, J. Ind.

    Eng. Chem., 12, 689 (2006); (c) M. Heon and M.

    Kang, J. Ind. Eng. Chem., 11, 540 (2005).

    8. K. R. Reddy, A. V. Ramaswamy, and P. Ratnasamy,

    J. Catal., 143, 275 (1993).

    9. S. Yoshida, T. Tanaka, M. Okada, and T. Funabiki,

    J. Chem. Soc., Faraday Trans. 1, 80, 119 (1984).

    10. T. Tanaka, M. Ooe, T. Funabiki, and S. Yoshida, J.

    Chem. Soc., Faraday Trans. 1, 82, 35 (1986).

    11. T. Tanabe, Y. Nishimura, S. Kawasaki, T. Funabiki,

    and S. Yoshida, J. Chem. Soc., Chem. Commun., 506

    (1987).

    12. K. Tanabe, CHEMTECH October, 628 (1991).

    13. P. Brandao, A. Philippou, J. Rocha, and M. W.

    Anderson, Catal. Lett., 80, 99 (2002).

    14. M. Nishimura, K. Asakura, and Y. Iwasawa, J.

    Chem. Soc., Chem. Commun., 1660 (1986).

    15. M. Shirai, N. Ichikumi, K. Asakura, and J. I.

    Wasawa, Catal. Today, 8, 57 (1990).

    16. O. Desponds, R. L. Keiski, and G. A. Somorjai,

    Catal. Lett., 19, 17 (1993).

    17. S. Okazaki and N. Wada, Catal. Today, 16, 349

    (1993).

    18. M. Ziolek, I. Sobczak, I. Nowak, P. Decyk, A.

  • Hydrothermal Synthesis and Characterization of Niobium-containing Silicalite-1 Molecular Sieves with MFI Structure 771

    Lewandowska, and J. Kujawa, Microporous and

    Mesoporous Materials, 35, 195 (2000).

    19. A. Morikawa and A. Togashi, Catal. Today, 16, 333

    (1993).

    20. A. Philippou, P. Brandao, A. Ghanbari-Siahkali, J.

    Dwyer, J. Rocha, and M. W. Anderson, Appl. Catal.

    A, 207, 229 (2001).

    21. J. M. Jehng and I. E. Wachs, Catal. Today, 16, 417

    (1993).

    22. T. Kushimoto, Y. Ozawa, A. Baba, and H. Matsuda,

    Catal. Today, 16, 571 (1993).

    23. J. Rocha, P. Brandao, J. D. Pedrosa de Jesus, A.

    Philippou, and M. W. Anderson, Chem. Commun.,

    471 (1999).

    24. A. M. Prakash and L. Kevan, J. Am. Chem. Soc.,

    120, 13148 (1998).

    25. M. Ziolek and I. Nowak, Zeolites, 18, 356 (1997).

    26. I. Sobczak, P. Decyk, M. Ziolek, M. Daturi, J. C.

    Lavalley, L. Kevan, and A. M. Prakash, J. Catal.,

    207, 101 (2002).

    27. P. R. Hari Prasad Rao, A. V. Ramaswamy, and P.

    Ratnasamy, J. Catal., 137, 225 (1992).

    28. K. R. Reddy, A. V. Ramaswamy, and P. Ratnasamy,

    J. Chem. Soc., Chem. Commun., 1613 (1992).

    29. Y. S. Ko and W. S. Ahn, Korean J. Chem. Eng., 15,

    423 (1998).

    30. M. Shibata and Z. Gabelica, Zeolites, 19, 246

    (1997).

    31. N. K. Mal, V. Ramaswamy, P. R. Rajamohanan, and

    A. V. Ramaswamy, Microporous Materials, 12, 331

    (1997).

    32. M. R. Boccuti, K. M. Rao, A. Zecchina, G. Leofanti,

    and G. Petrini, Structure and Reactivity of Surfaces,

    Studies in Surface Science and Catalysis, C. Mor-

    terra, A. Zecchina and G. Costa Eds., 48, p. 133,

    Elsevier, Amsterdam (1989).

    33. D. R. C. Huybrechts, I. Vaesen, H. X. Li, and P. A.

    Jacobs, Catal. Lett., 8, 237 (1991).

    34. D. W. Breck, Zeolite Molecular Sieves: Structure,

    Chemistry and Use, p. 415, Wiley, New York

    (1974).

    35. J. S. Reddy, R. Kumar, and P. Ratnasamy, Appl.

    Catal., 57, L1 (1990).

    36. A. Thangaraj, R. Kumar, S. P. Mirajkar, and P.

    Ratnasamy, J. Catal., 130, 1 (1991).

    37. A. J. H. P. van der Pol, A. J. Verduyn, and J. H. C.

    Hooff, Appl. Catal. A, 92, 113 (1992).

    38. B. Rakshe, V. Ramaswamy, S. G. Hegde, R.

    Vetrivel, and A. V. Ramaswamy, Catal. Lett., 45, 41

    (1997).

    39. Y. Ben Taarit, A. Tuel, N. Velmasco, and C. Na-

    ccache, Fr. Patent 9 203 364 (1992).

    40. A. Tuel and Y. Ben Taarit, J. Chem. Soc., Chem.

    Commun., 1578 (1992).

    41. S. A. Axon and Klinowski, Appl. Catal. A, 81, 27

    (1992).