Sunteți pe pagina 1din 11

Microporous and Mesoporous Materials 261 (2018) 18–28

Contents lists available at ScienceDirect

Microporous and Mesoporous Materials


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

Nickel impregnated mesoporous USY zeolites for hydrodeoxygenation of T


anisole
David P. Gamliela,b, Brian P. Bailliea,b, Evan Augustinea, Jason Halla, George M. Bollasa,b,
Julia A. Vallaa,b,∗
a
Department of Chemical & Biomolecular Engineering, University of Connecticut, Storrs, 191 Auditorium Road, Unit 3222, Storrs, CT 06269-4602, USA
b
Center for Clean Energy Engineering, University of Connecticut, Storrs, 44 Weaver Road, Storrs, CT 06269-3222, USA

A R T I C L E I N F O A B S T R A C T

Keywords: Ni was impregnated on various mesoporous USY zeolites, which were characterized and tested for catalytic
Mesoporous zeolite hydrodeoxygenation (HDO) of anisole. Parent USY zeolite was modified using the mild surfactant-assisted
Ultra-stable Y method, and more severe desilication and desilication with pore directed agent techniques. Ni impregnation on
Impregnation the zeolites with higher micropore volume produced small, easily reduced, internal Ni crystallites, whereas the
Nickel
larger mesopores promoted the formation of bulky external Ni particles. Catalysts with larger external Ni par-
Hydrodeoxygenation
ticles were more effective for HDO of anisole. HDO with the optimal mesoporous USY supported catalyst pro-
ceeded at more than double the reaction rate, with greater yield to fully deoxygenated products, compared to
HDO using Ni supported on the parent USY. Kinetic modeling using a simple lumped mechanism demonstrated
that this improvement was a result of greater rates of initial Ni-catalyzed hydrogenation reactions, most likely
due to greater access to the Ni surface. The most successful technique was further modified to produce a final
mesoporous USY supported Ni zeolite with superior catalytic activity for anisole HDO.

1. Introduction inaccessible directly from the surface [12]. Hierarchical USY zeolites,
with secondary mesoporosity, are effective for a number of catalytic
Zeolites are commonly used as active catalysts and catalyst supports applications, especially for cracking of large molecules such as cumene
in the petrochemical industry. Excellent hydrothermal stability, shape [13] and α-pinene [14].
selective pore structure and Brønsted acidity make zeolites effective for While exhaustive research has been performed on synthesis
catalytic cracking applications [1]. High surface area and shape se- methods, catalytic properties and molecular transport advantages of
lectivity arise from the microporous nature of the zeolite [2]. This hierarchical USY zeolites, limited research has been performed on how
characteristic high surface area allows zeolites to host dispersed metal those materials act as a support for metals. De Jong et al. [15] im-
species and form highly active bifunctional catalysts [3,4]. However, pregnated 0.3 wt% Pt on desilicated USY zeolite for n-hexadecane hy-
the presence of micropores may result in reduced transport of reactant drocracking and showed the material was more active than a com-
to zeolite acid sites located within the micropore network, reducing mercial USY catalyst. They concluded that this was a result of high mass
catalyst effectiveness factor. Subsequently, many research studies have transfer rates of hydrocracked products from the micropores of the
been performed to demonstrate improved access to catalyst active sites zeolite to the surface, reducing secondary cracking. Many groups
via intracrystalline mesopores [5–8]. [16–18] have studied mesoporous zeolites as catalyst supports, but the
Faujisite-type zeolites are commonly used for refinery applications, role of zeolite support mesoporosity on metal dispersion and acid cat-
and among them Y zeolite is often cited for excellent hydrothermal alyzed reactions is not fully understood.
stability and tuneable acidity [9]. Ultra-stable Y (USY) is created by Transition metals supported on zeolites have been considered as
dealumination of the Y zeolite in order to improve stability and gen- alternative bio-oil hydrodeoxygenation (HDO) catalysts. Typical HDO
erate mesoporosity. While USY catalysts are more effective than tradi- catalysts such as NiMo and CoMo require sulfidation and deactivate
tional Y zeolites for reactions such as hydrocarbon cracking [10] and easily in the presence of water, while other proposed catalysts, such as
methanol dehydration [11], the mesopores formed from dealumination supported precious metals, are costly [19–21]. A zeolite supported,
have been shown to be encapsulated within the zeolite crystal and transition metal catalyst offers an inexpensive and effective route for


Corresponding author. Department of Chemical & Biomolecular Engineering, University of Connecticut, Storrs, 191 Auditorium Road, Unit 3222, Storrs, CT 06269-4602, USA.
E-mail address: valla@engr.uconn.edu (J.A. Valla).

http://dx.doi.org/10.1016/j.micromeso.2017.10.027
Received 11 August 2017; Received in revised form 10 October 2017; Accepted 17 October 2017
Available online 21 October 2017
1387-1811/ © 2017 Elsevier Inc. All rights reserved.
D.P. Gamliel et al. Microporous and Mesoporous Materials 261 (2018) 18–28

HDO. Ni is an attractive metal due to high availability, excellent hy- dissolved in water and added dropwise to the calcined zeolite until the
drogenation capability and good stability [22]. Dry impregnation is a pores were filled. If filling occurred prior to expiration of the precursor
common and scalable metal incorporation technique [23], where pre- solution, the material was dried in static air for 2 h at 80 °C, and the
cursor solution fills the zeolite pores enabling high metal dispersion procedure was repeated until no solution was left. The final material
[24]. Sankaranarayanan et al. [25] showed that Ni incorporated on the was dried for 12 h at 80 °C in static air and then calcined. Each material
mesoporous Al-SBA-15 is effective for HDO of anisole. Hunns et al. [26] was reduced for 4 h at 450 °C in pure H2 prior to catalyst testing. A
demonstrated that mesoporous ZSM-5 impregnated with 1 wt% Pd is color change from light-grey to very dark grey was observed after re-
effective for HDO of m-cresol. They concluded this was a result of high duction of each material.
Pd dispersion in combination with increased mass transport to catalyst The naming convention in this study follows the format: “AA-BB
active sites. Liu et al. [27] studied Ni impregnation on hierarchical Y (XX)-USY”. “AA” indicates whether Ni is present in the reduced form
zeolite, and found qualitatively that dispersion and Ni crystal size (Ni) or the oxidized form (NiO). “BB” indicates the preparation tech-
uniformity increases with optimal mesoporosity. This translated to in- nique (DS, SA, DS-PDA), and “XX” indicates the alkaline treatment
creased rates of furfural hydrogenation. strength in mmol/L (50, 25 or 10). For example, the material “Ni-DS-
The objective of this work is to determine how impregnated Ni is PDA(50)-USY” was prepared with the DS-PDA method, using 50 mM
dispersed on the USY zeolite, how a mesoporous USY support affects NaOH, and contains Ni in the reduced form.
this dispersion and how the factors of Ni dispersion and zeolite meso-
porosity affect the HDO of bio-oil. Hierarchical pore structure was in- 2.2. Catalyst characterization
troduced in a commercial USY zeolite using three established post-
secondary methods, and each method was confirmed to create a unique Gas sorption was conducted using a Micromeritics ASAP 2020C
mesopore structure. Ni was incorporated via the dry impregnation Sorption Analyzer. N2 physisorption isotherms were gathered at 77 K.
technique, and catalyst characterization was performed to determine Prior to analysis, 0.1 g of zeolite was degassed at 250 °C under vacuum
the Ni state and dispersion. Each catalyst was tested for liquid phase for 12 h. Surface area was calculated using the BET equation, and pore
HDO of anisole; a reaction which requires Ni sites for hydrogenation size distribution was determined using the BJH method applied to the
and Brønsted acid sites (B.A.S.) for dehydration. Kinetic modeling of adsorption branch of each isotherm. H2 chemisorption was conducted
intermediate and final product evolution was used to quantify the ef- using the same instrument, configured with a furnace and special gas
fectiveness of both Ni and acid sites. Support mesoporosity was shown inlet. Sample was degassed and then reduced in-situ in pure H2 at 450 °C
to affect Ni dispersion and particle size, which subsequently played an for 4 h. Isotherms were collected at 35 °C, and sample metal dispersion
important role on catalyst effectiveness. was calculated from the difference in two consecutive isotherms.
X-ray diffraction (XRD) measurements were performed using a
2. Experimental Bruker D8 Advance powder diffractometer with CuKα radiation source.
Average Ni and NiO crystallite sizes were calculated by XRD line
2.1. Material preparation broadening using the accompanying EVA software.
Temperature programmed reduction (TPR) was conducted grav-
Commercial USY zeolite with Si/Al ratio of 40 was purchased from imetrically using a Pyris 1 Thermogravimetric Analyzer (TGA). About
Zeolyst International (CBV780). The parent USY was modified using 35 mg of sample was loaded into a Pt crucible. The first stage of the
three different techniques: desilication (DS) [28], desilication with a program, intended to oxidize any reduced Ni and remove any con-
pore-directing agent (DS-PDA) [29] and the surfactant-assisted method taminants from the surface, occurred at 550 °C for 30 min in air. The
(SA) [30]. All chemicals were purchased from Sigma-Aldrich, unless second stage ramped the temperature from 100 °C to 900 °C at 5 °C/min
otherwise noted. Both parent and modified USY were calcined in air at in 3% H2 diluted in Ar. Bulk Ni loading was determined by inductively
550 °C for 6 h prior to use. All materials were stored in tightly sealed coupled plasma optical emission spectroscopy (ICP-OES), which was
vials and used directly after calcination to avoid adsorption of atmo- performed at an external laboratory. The amount of reducible Ni was
spheric moisture. calculated by determining the weight loss during reduction, attributing
Desilication by alkaline treatment was accomplished by stirring 3 g lost weight to atomic oxygen, and normalizing to the bulk Ni loading
of zeolite in 90 mL of 0.05 M NaOH for 30 min at 60 °C, followed by determined by ICP-OES.
washing and drying. The material was then ion-exchanged in 300 mL of Diffuse reflectance Fourier transform infrared spectroscopy
2.4 M NH4NO3 for 6 h, followed by washing, drying and two more (DRIFTS-FTIR) pyridine titration (Py-IR) was used to identify the
repetitions. The final material was obtained in the H- form through Brønsted and Lewis acid sites present in each material. The instrument
calcination, as described above. The exact same procedure was fol- consisted of a Thermo Nicolet 6700 FTIR with a Harrick Praying Mantis
lowed to obtain the DS-PDA material, with the exception that 0.8 g DRIFTS accessory. About 30 mg of sample was weighed and then out-
tetrapropylammonium bromide (TPABr) was added to the initial alka- gassed in-situ at 550 °C under vacuum. A spectrum was recorded at this
line treatment. The effects of alkaline treatment strength on the DS-PDA temperature with a KBr background. The sample was then cooled to
method were also studied by reducing the NaOH concentration from 130 °C, a background was collected, and then pyridine was swept from
0.05 M to 0.025 and 0.01 M. Material yield for the DS method varies a saturator to the cell by N2 carrier gas. After sample saturation, phy-
between 45 wt% and 75 wt% [28], while the yield of the DS-PDA sisorbed pyridine was removed by slowly heating the cell to 230 °C
method typically produces greater material yield (between 56 wt% and under high vacuum. The final spectrum was recorded at 130 °C. The
86 wt%) [29,31]. Brønsted and Lewis acidity were calculated from the area under the
The SA method is a more refined way to produce uniform secondary peaks at 1550 cm−1 and 1450 cm−1, respectively, and corrected using
mesopores, while maintaining material crystallinity [30,32]. 3 g of the extinction coefficients found by Emeis [33].
zeolite was stirred in 20 mL of DI water and 1.2 g of hexadecyl- Transmission electron microscopy (TEM) was performed on a
trimethlyammonium bromide (CTAB) and heated to 80 °C. Then, 6 mL Tecnai T12 scanning transmission electron microscope (STEM), at a
of 4.4 M NH4OH was added, and the treatment proceeded for 24 h, voltage of 120 kV. Dark-field STEM images were collected on a FEI
followed by washing, drying and calcination. The material yield of the Talos F200X FEG STEM, operating at a potential of 200 kV, and the
SA method is typically greater than 90 wt% [30]. presence of Ni was confirmed by energy dispersive X-ray spectroscopy
Ni was deposited on each material using the dry impregnation (EDX).
technique [24]. The theoretical Ni loading of each material prepared X-Ray photoelectron spectroscopy (XPS) was performed using a Phi
was 5 wt%. The corresponding amount of Ni(NO3)2●6H2O was 510 X-ray photoelectron spectrometer. Surface Si/Al ratio was

19
D.P. Gamliel et al. Microporous and Mesoporous Materials 261 (2018) 18–28

calculated from the relative area under the XPS Si and Al regions col- hysteresis, a result of mesopores created by dealumination. The SA
lected at a pass energy of 58.7 eV and dwell time of 50 ms/(sweep*- method produced a mesoporous USY with a mixed type I and IV iso-
step). A similar analysis was performed to find the surface Ni/Si ratio, therm, with a more pronounced hysteresis loop closer to H1, indicative
but the pass energy was reduced to 29.35 eV. Peak locations were of ordered mesopores. The materials created with the DS and DS-PDA
corrected using the adventitious carbon peak located at 284.7 as a re- methods have a mixture of type I and II isotherms, like the parent USY,
ference. but with more pronounced H4 hysteresis loops. Low N2 uptake at P/
P0 < 0.1 indicated loss of microporosity and crystallinity for the DS
2.3. Hydrodeoxygenation of anisole and DS-PDA zeolites.
Table 1 contains the BET surface area, external surface area, mi-
HDO of anisole was performed in a 100 mL stirred stainless steel cropore volume and mesopore volume of each material. From the BJH
autoclave reactor (Parr Instruments), fitted with a gas entrainment pore size distribution presented in Fig. 1(B), the SA method created
impeller. All experiments were performed at a temperature of 200 °C, mesopores with a uniform diameter of 30 Å. The DS method produced a
pressure of 750 psi, and a stir rate of 1100 rpm. Initially, the reactor very broad range of mesopores, with the diameter from 20 to 500 Å.
contained 49 mL of 6 wt% anisole diluted in decane, 0.1 g of reduced The DS-PDA zeolite pore size distribution was concentrated at 45 Å,
catalyst and 1 mL of octane, which was used as an internal standard with some larger mesopores also present. The above observations are in
throughout the experiment. agreement with previously reported results [28,29,31]. Ni was im-
Prior to heating, the reactor was purged with Ar, and then H2. The pregnated on each zeolite using a standard dry impregnation technique.
reactor was pressurized to check for leaks, then purged and heated to Additional N2 sorption isotherms were gathered after Ni impregnation.
200 °C. Once the reactor reached the target temperature (heating As shown in Table 1, Ni impregnation resulted in decreased BET surface
time = 16 min), it was pressurized, the impeller was engaged and the area and pore volume for all materials, most likely due to pore blockage
reaction clock started. A constant feed of H2 was supplied to maintain [37].
reactor pressure at 750 psi for the duration of the experiment. Liquid Table 2 shows the physicochemical properties of each Ni-in-
samples (2 mL) were collected from the reactor using a custom-made corporated zeolite. The Ni content of each material was experimentally
sampling system. The sampling system was purged with Ar prior to quantified by ICP-OES, and was within 10% of the theoretical value of
each sample to avoid exposing the reactor contents to air. Liquid 5.0 wt%. Table 2 also contains the Brønsted and Lewis acidity of each
samples were collected every 20 min, diluted in isopropanol and fil- zeolite prior to Ni impregnation, determined by pyridine titration
tered to remove catalyst. The total duration of the reaction was monitored by DRIFTS. The DRIFTS spectra are available in Electronic
140 min. Reproducibility was quantified by repeating the catalytic HDO Supporting Information (ESI) Fig. S1. The parent USY zeolite was highly
experiments with the Ni-USY, Ni-SA-USY and Ni-DS-PDA(25)-USY three acidic, and Brønsted acidity decreased after alkaline treatment. Desili-
times. Dispersion bars represent the standard deviation of three ex- cation alone removed almost all acidity, consistent with previous lit-
periments. erature [29]. Desilication with DS-PDA and the SA methods retained
Liquid analysis was performed by gas chromatography mass spec- significant Brønsted acidity, and increased the amount of Lewis acidity,
trometry (GC-MS, Agilent 6890 GC with 5673 MS). The inlet split ratio possibly due to the redeposition of framework aluminium as extra-
was 100:1 and the GC was equipped with a HP-5 column. The tem- framework aluminium [38].
perature program held isothermal at 40 °C for 2 min, followed by a Determination of zeolite crystallinity, Ni crystallite size and reduc-
1 °C/min ramp to 75 °C and then a 20 °C/min ramp to 270 °C followed tion extent were accomplished by XRD. ESI Fig. S2 shows X-ray dif-
by another 2 min isothermal step. The system was calibrated externally fractograms for each material (A) before Ni impregnation, (B) after Ni
using neat cyclohexane, octane, cyclohexanone, anisole, phenol and impregnation and calcination and (C) after reduction in pure H2 at
guaiacol. Quantification of compounds without direct calibration 450 °C. The SA method preserved the majority of the material crystal-
standards was accomplished by the semi-quantification method, based linity, while the DS-USY method retained very little crystallinity. This is
on molecular weight and carbon number of the specific analyte. Anisole a result of the difference in alkaline treatment strength. NaOH, a strong
conversion and product selectivity were calculated using equations (1) base, was used for the DS and DS-PDA methods, whereas the SA method
and (2), respectively: utilized a weak base (NH4OH). The DS-PDA material retained some
crystallinity, confirming that the PDA does protect some of the crystal
CA structure, which was first reported by Verboekend et al. [28,29].
XA = 1 −
CA0 (1) XRD reflections at 2Θ of 37°, 43°, 63°, 75° and 79° are representative
moli of NiO crystal phases, and are apparent in Fig. S2(B) for all materials
Si = ∗100 after Ni impregnation and calcination. In Fig. S2(C), these phases dis-
∑ molproducts (2)
appear in favour of metallic Ni peaks located at 2Θ of 44°, 51° and 76°,
Due to the large excess of hydrogen, the initial reaction rate was indicating that the reduction of NiO in H2 was successful. Table 2
assumed to be first-order in anisole concentration and zero-order in H2 contains the average NiO and Ni crystallite size of each material, de-
concentration [34,35]. The overall rate constant was calculated from termined using the Scherrer equation. The average NiO crystallite size
the slope of a plot of ln(CA/CA0) vs. time. In order to determine the after impregnation and calcination on all materials was approximately
relative rates of Ni and Brønsted acid catalyzed reactions, a lumped 30 nm. This crystallite size is larger than the size of most mesopores,
kinetic model was proposed and fitted to the reaction data using a least- and far larger than the Y-zeolite supercage (12 Å). Thus, the majority of
squares algorithm developed with Gams IDE software. the NiO was formed on the exterior surface of the zeolite. After re-
duction, the crystallite size was reduced to 13.7 nm for the Ni-USY and
3. Results and discussion 19.5 for the Ni-SA-USY. The average Ni crystallite size for the Ni-DS
(50)-USY and Ni-DS-PDA(50)-USY was similar, but both were sig-
3.1. Ni-impregnated hierarchical USY zeolites nificantly greater than that of the Ni-USY. The significant decreases in
Ni crystallite size after reduction on the Ni-USY and Ni-SA-USY coin-
Hierarchical USY zeolites were created using three different cides with the presence of micropores and high zeolite surface area. H2
methods: SA, DS and DS-PDA. N2 sorption isotherms and the BJH pore chemisorption was used to find the Ni dispersion on each material. The
size distribution of a material created by each method is available in Ni-USY had almost 10% Ni dispersion, whereas the mesoporous USY
Fig. 1(A) and (B), respectively. The parent USY exhibits a characteristic materials exhibited about half that dispersion, consistent with the
type I and II mixed isotherm (IUPAC standard) [36], with minor H4 larger crystallite size observed from XRD. Ni may have migrated into

20
D.P. Gamliel et al. Microporous and Mesoporous Materials 261 (2018) 18–28

Fig. 1. (A) N2 Sorption isotherms of materials created using SA, DS and DS-PDA methods. (B) BJH pore size distribution of USY zeolites created by each method.

Table 1
Data gathered from N2 sorption isotherms of materials created using SA, DS and DS-PDA
methods, before and after Ni impregnation.

Material BET External Total Pore Micropore Mesopore


Surface Surface Volumeb Volumea Volumec
Area Area1

m2/g m2/g cm3/g cm3/g cm3/g

USY 719 297 0.479 0.195 0.284


Ni-USY 661 248 0.445 0.188 0.257
SA-USY 703 336 0.514 0.170 0.344
Ni-SA-USY 692 362 0.507 0.167 0.340
DS(50)-USY 243 221 0.409 0.007 0.402
Ni-DS(50)- 222 208 0.385 0.004 0.381
USY
DS-PDA(50)- 519 343 0.534 0.081 0.453
USY
Ni-DS- 485 335 0.517 0.069 0.448
PD-
A(50)-
USY

a
t-plot method.
b
Single point below P/Po = 0.99.
c
Total Pore Volume - Micropore Volume.
Fig. 2. Temperature programmed reduction (TPR) profiles for Ni supported on parent and
mesoporous USY zeolites created by SA, DS and DS-PDA methods. Profiles are offset by
the greater number of micropores present in the Ni-USY, resulting in
0.005%/min.
small intracrystalline Ni species, reducing the average Ni crystallite size
for that material. The reduction in crystal size was only observed after
reduction, indicating that this migration most likely occurs during profile of Ni-USY exhibits a low-temperature reduction peak, at 305 °C,
treatment at 450 °C in H2. and a larger higher-temperature peak at 360 °C. The profile of Ni-SA-
Reduction behavior and structural changes of the Ni-impregnated USY also exhibited a bi-modal distribution, but both peaks were shifted
USY zeolites were further elucidated by TPR, shown in Fig. 2. The TPR to slightly lower temperatures, and the higher temperature peak was

Table 2
Physicochemical properties of Ni-supported on parent and mesoporous zeolites prepared by SA, DS and DS-PDA methods.

Material Bulk Ni Loadinga Average NiO Crystallite Sizeb Average Ni Crystallite Sizec Reducibilityd Dispersione Brønsted Acid Sitesf Lewis Acid Sitesf

wt.% nm nm % % umol/g umol/g

Ni-USY 5.33 28.8 13.7 73 9.7 528 133


Ni-SA-USY 4.51 27.5 19.5 96 4.1 369 392
Ni-DS(50)-USY 4.62 28.7 21.3 94 5.8 46 1
Ni-DS-PDA(50)-USY 4.87 31.6 24.8 88 5.2 145 357

a
From ICP-OES.
b
Calculated from application of Scherrer equation to peaks at 2Θ of 37°, 43°, 63°, 75° and 79°.
c
Calculated from application of Scherrer equation to peaks at 2Θ of 44°, 51° and 76°.
d
From TPR.
e
From H2 Chemisorption.
f
From Pyridine titration of material prior to Ni impregnation monitored by FTIR.

21
D.P. Gamliel et al. Microporous and Mesoporous Materials 261 (2018) 18–28

much more pronounced relative to the lower temperature peak. The Ni- to have a broad Ni particle size distribution, in agreement with the
DS(50)-USY exhibited a broad reduction region, with a high tempera- observations from TPR. As shown in Fig. 4(D), the Ni-DS-PDA(50)-USY
ture shoulder at 515 °C. Finally, the Ni-DS-PDA(50)-USY reduction exhibits many bulk surface Ni particles. Small Ni particles also appear
profile displayed a broad peak centred at 350 °C. to radiate inward from the larger surface Ni particles. The Ni-USY has
TPR provides information about metal location, the strength of high surface area and micropore volume, which provides sufficient
support-metal interaction and the reducibility of a metal within the room for Ni migration, resulting in the creation of external bulk Ni and
zeolite [39]. Stronger interaction between the metal and support results small internal Ni particles. Conversely, the Ni-DS(50)-USY has lower
in metal that is harder to reduce, and thus requires a higher reduction surface area and microporosity, so there is less dispersion. The Ni-DS-
temperature [40]. Larger NiO crystallites also require a higher reduc- PDA(50)-USY appears to be a blend between the two; there is some
tion temperature. Reduction temperatures below 400 °C indicate that surface area for some Ni migration.
most of the initial NiO is on the exterior of the zeolite micropores, In order to gain further insight on the reduction behavior and lo-
minor ion-exchange has occurred and the NiO does not interact strongly cation of surface Ni, XPS spectra for the Ni-USY, NiO-DS-PDA(50)-USY
with the support [41]. The two peaks of the Ni-USY and Ni-SA-USY and Ni-DS-PDA(50)-USY materials were collected. The surface Ni/Si
imply a bimodal distribution of Ni particle size. The broader peak of the ratio of the NiO-DS-PDA(50)-USY was found to be 0.131, while the Ni/
Ni-DS(50)-USY and Ni-DS-PDA(50)-USY suggests a broader crystal size Si ratio of the Ni-DS-PDA(50)-USY was calculated to be 0.042, in-
distribution of the NiO precursor. Moreover, the Ni-DS(50)-USY mate- dicating that significant Ni migration from the surface to the bulk of the
rial contains mesopores that are larger in size than the NiO crystallite material occurs during reduction, consistent with the observations from
size, so some interaction between external acid sites and Ni may be STEM. The surface Ni/Si ratio of Ni-USY was 0.037, slightly less than
reflected in the more pronounced shoulder at 515 °C [42]. the Ni-DS-PDA(50)-USY, implying that more Ni remains on the surface
As shown in Table 2, the reducibility of each mesoporous Ni-USY when it is impregnated on the DS-PDA(50)-USY, compared to the
material was high, evidence that the majority of NiO was accessible to parent. Further analysis of the Ni XPS spectrum can be found in the ESI.
hydrogen and not trapped within zeolite micropores. Reducibility was To confirm the presence of small and large Ni particles, high re-
slightly lower for the parent Ni-USY than the mesoporous Ni-USY cat- solution TEM images of the Ni-USY were captured and are displayed in
alysts, indicating that some of the bulk NiO may be inaccessible or Fig. 4(E–F). Fig. 4(E) shows a large Ni particle on the surface of the USY
highly coordinated with the zeolite. To summarize, the majority of NiO zeolite. The particle has well defined edges on the exterior surface.
prior to reduction was most likely on the exterior surface and did not Fig. 4(F) shows the coexistence of smaller Ni particles on the same
strongly interact with the support. sample. These particles are about 3–4 nm in size, 4 times the size of the
Further qualitative understanding of the zeolite topology and Ni zeolite crystal lattice dimension. Given the location and particle
speciation was gained by electron microscopy. Fig. 3 shows TEM mi- structure, it's possible that the Ni complexed with the zeolite to form Ni
crographs of the Ni-USY, Ni-SA-USY and Ni-DS-PDA(50)-USY after re- silicate. This behavior has been shown for Ni impregnated on SiO2 [43],
duction. The dark particles are Ni crystals, and the lighter areas are beta zeolite [44] and silicalite 1 [45], and would explain the decreased
zeolite. The zeolite area of the Ni-USY shown in Fig. 3(A) is uniform and reducibility of Ni supported on parent USY. Conversely, no peaks re-
dark. A few interstitial mesopores exist, a result of the dealumination presentative of Ni silicate were observed in any of the XRD patterns,
treatment. The bulk Ni particles appear to be between 20 and 50 nm in presented in ESI Fig. S2. Moreover, these particles were not observed on
diameter. The Ni-SA-USY zeolite in Fig. 3(B) is far more opaque than the Ni-DS(50)-USY material. It is apparent that the Ni-USY exhibits
the parent USY, with small pinpoint mesopores, consistent with the improved dispersion of metal particles, which often leads to improved
pore diameter of 30 Å measured by N2 sorption. Few of the pores ap- catalytic activity of a material [46].
pear to originate at the edges of the zeolite crystal. Finally, the Ni-DS- In summary, four Ni-impregnated materials were created with
PDA(50)-USY material shown in Fig. 3(C) contains visibly larger, varying degrees of mesoporosity. Three established post-secondary
random mesopores. It appears that several mesopores originate at the methods were used to produce a USY material with mild (SA) and se-
edge of the material. vere (DS) mesopores of varying pore diameter. The characterization
Dark field STEM was used to further investigate the nature of Ni shows that the type of porosity and the surface area of the support have
supported on parent and mesoporous USY, shown in Fig. 4(A–D). The a profound effect on the nature of impregnated Ni. High surface area
bright areas of each micrograph indicate Ni particles, and the grey areas and microporosity result in high dispersion of Ni particles after reduc-
are zeolite. This was confirmed by elemental mapping of the same tion and migration into the pore network. The introduction of meso-
images, available in ESI Fig. S3. Bulk Ni is clearly visible on the surface porosity by DS and DS-PDA methods result in lower microporosity and
of each material. The Ni-USY displayed in Fig. 4(A) contains many BET surface area. This translates to larger Ni crystallite sizes, mostly
smaller Ni crystals and threads within the zeolite. These smaller Ni present on the external surface of the zeolite, and lower Ni dispersion.
species are distributed throughout the pore network. The Ni-SA-USY in The following section discusses how the aforementioned properties of
Fig. 4(B) also contains very small Ni particles in addition to larger more these materials affect catalytic activity for the HDO of anisole.
uniform surface Ni particles. The Ni-DS(50)-USY from Fig. 4(C) appears

Fig. 3. TEM micrographs of (A) Ni-USY, (B) Ni-


SA-USY and (C) Ni-DS-PDA(50)-USY.

22
D.P. Gamliel et al. Microporous and Mesoporous Materials 261 (2018) 18–28

Fig. 4. Dark field scanning transmission electron micrographs of


(A) Ni-USY, (B) Ni-SA-USY, (C) Ni-DS(50)-USY and (D) Ni-DS-
PDA(50)-USY. High resolution TEM images of (E) large and (F)
small Ni crystals on USY material.

3.2. Hydrodeoxygenation of anisole calculated after 140 min of reaction time for anisole HDO over each
catalyst. Initial reaction rate was calculated by multiplying a first order
Pyrolysis oil from lignocellulosic biomass contains many different rate constant, kapp, by the initial concentration. The rate constant was
oxygenated compounds. HDO is a reaction that removes this atomic determined from a linear fit of the first-order disappearance of reactant
oxygen, improving bio-oil heating value and stability [47]. Due to the anisole, presented in Fig. 5(B). The small dispersion bars from the Ni-
complex nature of bio-oil, model compounds are employed to study USY and Ni-SA-USY experiments in Fig. 5(B) indicate that the experi-
catalyst effectiveness and reaction pathway [48,49]. Anisole is a model ment was highly reproducible.
compound, formed by the decomposition of lignin [25]. At moderate From Fig. 5(A), the final conversion of anisole mirrored the reaction
temperature and high H2 pressure, anisole HDO proceeds in two broad rate. The most active catalyst for anisole conversion was Ni-DS-
steps: Ni-catalyzed saturation of the aromatic ring, followed by PDA(50)-USY, and the Ni-SA-USY catalyst was the least active. Initial
Brønsted acid-catalyzed dehydration to eliminate the oxygen atom. As a reaction rate over Ni-USY was approximately 100 mmol/(gcat-hr), and
result, this reaction is ideally suited to evaluate the behavior of both the was outperformed by the Ni-DS(50)-USY and the Ni-DS-PDA(50)-USY
Ni and zeolite active sites discussed in Section 3.1. catalysts.
Catalytic HDO of anisole was performed at 200 °C and a pressure of The final product selectivity after 140 min of reaction for anisole
750 psi. Reactor contents were sampled every 20 min as the reaction HDO with each catalyst is available in Table 3. Ni-USY was the most
proceeded to determine the initial rate of reaction and product evolu- selective catalyst, producing the highest product selectivity to cyclo-
tion. Fig. 5(A) shows the initial rates of reaction and final conversion hexane and low selectivity to oxygenated intermediates. Several larger

23
D.P. Gamliel et al. Microporous and Mesoporous Materials 261 (2018) 18–28

Fig. 5. (A) Initial reaction rate (bars, left-axis) and final conversion (line, right axis) for anisole HDO over Ni impregnated on parent and mesoporous USY zeolites. (B) First order kinetic
fit for rate of anisole HDO over each catalyst. Reaction Conditions: temperature: 200 °C, pressure: 750 psi, reaction time: 140 min.

Table 3 final products [51]. However, the micropore opening diameter of USY
Final product molar selectivity for anisole HDO over Ni impregnated on parent and is relatively large, and can accommodate bulky molecules, such as
mesoporous USY zeolites.
glucose and 1,5-dimethylnaphthalene [52]. Here, the micropores most
Compound Selectivity (mol%)
likely stabilize the bulky C-C coupling products. This is a potentially
desirable reaction pathway, useful for promoting diesel range hydro-
Ni-USY Ni-SA- Ni-DS Ni-DS- carbons over gasoline range hydrocarbons.
USY (50)-USY PDA(50)-USY There are many mechanisms for anisole HDO published in the lit-
Hydrocarbons
erature [25,34,35,53]. Every mechanism shows cyclohexane as the final
Cyclohexane 82.6 77.7 21.4 71.9 product, formed from hydrogenated single-ring oxygenate inter-
Cyclohexene 0.7 – – 0.1 mediates. Zhao et al. [54] have demonstrated that phenol HDO cata-
Oxygenates lyzed by Ni-zeolite proceeds in a cascade fashion: Ni catalyzed hydro-
1-methoxycyclohexane 0.3 – 74.7 21.7
genation of the aromatic ring, followed by Brønsted acid catalyzed
Cyclohexanol – – 2.3 2.7
Cyclohexanone 0.6 0.4 – – dehydration to remove the oxygen atom as water.
C-C Coupling Here, we propose a simple lumped model for anisole HDO, pre-
1,1′ Bicyclohexyl 5.2 9.7 0.5 1.5 sented in Scheme 1. Anisole undergoes Ni catalyzed hydrogenation to
Cyclohexene, 1-Cyclohexyl 0.6 1.0 – – produce oxygenated intermediates. The model does not distinguish
1,1′ bicyclohexanone 0.8 2.9 1.3 –
Phenol, 2-cyclohexyl-4-methyl 2.6 3.2 – 1.2
between the oxygen functional group. It is worth noting that cyclo-
Benzene, 1-Cyclohexyl-4- 6.6 5.1 – 0.8 hexanone was primarily observed over the more acidic materials and 1-
methyl methoxycyclohexane was observed over the less acidic materials. The
oxygenated lump further undergoes Brønsted acid catalyzed dehydra-
tion to form the final cyclohexane product. This mechanism and kinetic
carbon-carbon (C-C) coupling products were also observed. Of the C-C analysis relies on several assumptions. First, no direct deoxygenation
coupling products, 1-1′ bicyclohexyl was the most abundant. HDO with from anisole to cyclohexane via benzene occurs. To address this, the
Ni-SA-USY also produced high selectivity to cyclohexane, and even reaction conditions (200 °C, 750 psi) were specifically chosen to make
higher selectivity to C-C coupling products, compared to Ni-USY.
Moreover, few oxygenated intermediates were observed in the products
from HDO with Ni-SA-USY. Anisole HDO over Ni-DS(50)-USY resulted
in far more oxygenated intermediates, less cyclohexane and far fewer C-
C coupling products. Finally, HDO of anisole with Ni-DS-PDA(50)-USY
catalyst produced about 72 mol% cyclohexane selectivity, with a bal-
ance of mostly 1-methoxycyclohexane. While selectivity to cyclohexane
was greater for anisole HDO with Ni-USY (82.6 mol%) compared to Ni-
DS-PDA(50)-USY (71.9 mol%), molar yield to cyclohexane was greater
for the Ni-DS-PDA(50)-USY (65 mol%) compared to the Ni-USY (61 mol
%). This is because overall anisole conversion was greater when the Ni-
DS-PDA(50)-USY catalyst was used. A full accounting of molar product
yields can be found in ESI Table S1.
C-C coupling product selectivity was much greater over the catalysts
with higher microporosity and acidity. This was not expected con-
sidering that previous studies have shown that application of a meso-
porous zeolite results in a loss of shape selectivity and formation of
larger products [38,50]. Shape selective product formation primarily
arises when pore size restricts intermediate formation and stabilizes Scheme 1. Simplified lumped reaction mechanism for anisole HDO.

24
D.P. Gamliel et al. Microporous and Mesoporous Materials 261 (2018) 18–28

benzene formation thermodynamically unfavourable [55]. Second, Table 4


hydrogenation of the first aromatic ring double bond occurs slowly Kinetic parameters from model fits for HDO of anisole over each catalyst, normalized to
the number of Ni (k1) and Brønsted acid (k2) active catalytic sites.
compared to hydrogenation of the second and third aromatic double
bond. This is reasonable, because no partially hydrogenated oxygenates Catalyst Ni Sitesa B.A.S.b k1 k2
were observed in any of the products at any point. Third, isomerization
and demethoxylation within the oxygenate (O) lump occur rapidly μmol/gcat μmol/gcat (1/s)/mol Ni (1/s)/mol B.A.S.
compared to acid-catalyzed dehydration. A similar assumption re-
Ni-USY 88 528 14.5 127.1
garding HDO of these compounds has been made in the literature [34]. Ni-SA-USY 32 369 10.1 83.9
Equations (3)–(5) are the governing differential equations for the Ni-DS(50)-USY 46 46 37.5 12.2
mechanism proposed in Scheme 1, for the changes in anisole (A), Ni-DS-PDA(50)-USY 43 145 46.6 22.7
oxygenate (O) and cyclohexane (C) concentration with respect to time. a
H2 Chemisorption.
b
dCA Pyridine FTIR.
= −k1 CA
dt (3)

dCO further justifying our assumptions.


= k1 CA − k2 CO In order to make a level comparison between each catalyst, para-
dt (4)
meters k1 and k2 are presented in Table 4, normalized to the number of
dCC Ni and Brønsted acid sites, respectively. The number of surface Ni sites
= k2 CO
dt (5) was determined from the dispersion of Ni on each material calculated
These equations were solved analytically [56], and those solutions from H2 chemisorption, and the number of Brønsted acid sites was
are available in ESI equations S1–S3. determined from Py-IR. Fitted rate constants were normalized to cata-
Parameters k1, k2 and CA0 were determined from a least squares fit lyst active sites as a concentration-dependent turnover frequency, so
to the time evolution of the concentrations of anisole, intermediate the effectiveness of each catalyst on a Ni and acid site basis can be
oxygenates and cyclohexane over each catalyst. Although the initial compared.
concentration of the starting reaction solution was known, it may have As shown in Table 4, HDO of anisole with Ni-USY proceeded at a
slightly changed during reactor purging and heating, so CA0 was in- moderate rate for hydrogenation and high rate for dehydration. HDO
cluded as a model parameter. Fig. 6 shows the reaction data and model with Ni-SA-USY progressed at a high rate constant for dehydration, but
fit for anisole HDO over each catalyst. The model fits the data well, not hydrogenation. Conversely, HDO with Ni-DS(50)-USY progressed at

Fig. 6. Reactant, intermediate and product concentration evolution with time (points) and kinetic model fit (lines) for anisole HDO over (A) Ni-USY, (B) Ni-SA-USY, (C) Ni-DS(50)-USY,
(D) Ni-DS-PDA(50)-USY.

25
D.P. Gamliel et al. Microporous and Mesoporous Materials 261 (2018) 18–28

Fig. 7. (A) Effects of NaOH concentration on micropore volume (left axis), XRD crystallinity and Brønsted acidity (right axes). (B) BJH mesopore size distribution for materials created
with DS-PDA method with different alkaline treatment strength.

high rates of hydrogenation, but not dehydration. Finally, the HDO hierarchical USY supported Ni catalysts.
with Ni-DS-PDA(50)-USY resulted in moderate rate constants of both
hydrogenation and dehydration. Rate constants for hydrogenation were 3.3. Improving performance of Ni-DS-PDA catalyst
greater over the DS- and DS-PDA catalysts, whereas the dehydration
rate constant was greater over the Ni-USY and Ni-SA-USY catalysts. The previous section highlights the benefits of a mesoporous USY
The overall rate of reaction was greater for the Ni-DS(50)-USY and support with preserved acidity. To further improve catalyst efficacy, we
Ni-DS-PDA(50)-USY materials. As revealed in Table 4, the rate con- focused on improving catalyst surface area and acidity without sacri-
stants of hydrogenation were higher for these two materials. This was ficing mesopore volume. Verboekend et al. [29] have shown that in-
unexpected because Ni dispersion was lower on these materials com- creasing PDA concentration during alkaline treatment is one way to
pared to the Ni-USY, as determined from H2 chemisorption. The smaller retain acidity and crystallinity. It is also well known that reducing de-
Ni crystallites present in Ni-USY may not be easily accessible, resulting silication (DS) alkaline treatment strength improves acidity of the final
in fewer turnovers per Ni atom. This is consistent with Mortensen et al. material at the expense of mesopore formation [38,59]. Here, alkaline
[57], who have shown that Ni/SiO2 catalyzed hydrogenation of phenol treatment strength of the DS-PDA method was reduced from 50 mM to
occurs more rapidly over larger Ni particles (> 10 nm). They attributed 25 mM and 10 mM of NaOH, denoted PDA(50), PDA(25) and PDA(10),
this to the presence of larger Ni facets, resulting in greater reactant respectively.
coordination. The hydrogenation rate constant was particularly low for Fig. 7(A) shows the effects of alkaline treatment strength on catalyst
anisole HDO with the Ni-SA-USY material. We hypothesize that the Brønsted acidity, micropore volume and XRD crystallinity. Py-IR
small uniform internal mesopores of the SA material trapped the Ni spectra, N2 sorption isotherms and XRD diffractograms of each material
particles. This resulted in enclosed Ni crystals with limited access to the are available in ESI Figures S1, S4 and S5. As expected, micropore
surface and less exposed metal surface area. volume, crystallinity and Brønsted acidity all increase as alkaline
The rate of the acid-catalyzed dehydration reaction was at least an treatment concentration decreases.
order of magnitude greater than the rate of hydrogenation for HDO The mesopore size distribution for each material created with the
with all catalysts. The catalysts with a greater number of acid sites had DS-PDA method is available in Fig. 7(B). There are two mesopore size
much higher rates of dehydration per acid site. This is most likely a regimes, a narrower pore diameter centred at 75 Å, and a broad me-
result of the proximity between Ni sites and acid sites. Efficient HDO soporosity between 100 and 500 Å. Stronger alkaline treatment favors
requires metal and acid sites in close proximity, such that both hy- the production of smaller, more well-defined mesopores, and weaker
drogenation and C-O bond activation can occur in quick succession alkaline treatment favors large broader mesopores. This is likely a result
[58]. The Ni-DS(50)-USY material contains relatively few acid sites, so of the base concentration gradient within the zeolite micropores. At low
efficient dehydration of intermediates could not occur, which resulted alkaline strength, significant Si leaching occurs from the external sur-
in high selectivity to intermediate oxygenates (Table 3). In contrast, face of the zeolite crystal, whereas at higher alkaline treatment strength
HDO with the Ni-DS-PDA(50)-USY catalyst proceeded at a moderate more of the internal Si is removed. To confirm this, surface Si/Al ratio
dehydration rate per acid site and high hydrogenation rate per Ni site. was determined using XPS. The surface Si/Al ratio of the USY, DS-
This balance made the Ni-DS-PDA(50)-USY the most effective catalyst, PDA(50)-USY and DS-PDA(10)-USY materials was found to be 88, 81
which outperformed the Ni-USY in terms of total cyclohexane yield and 60, respectively. The lower surface Si/Al ratio of the material
65 mol% vs. 61 mol%. created with lower alkaline treatment strength indicates that more Si
Summarizing, anisole HDO proceeded faster over the Ni-DS- has been removed from the surface, resulting in the broader pore size
PDA(50)-USY catalyst. Greater support mesoporosity benefited the distribution observed in Fig. 7(B). Binding energy and shape of the Si 2p
rates of the Ni-catalyzed hydrogenation reactions, while the inclusion of peak for each material was similar, and the XPS spectra can be found in
PDA protected some of the acidity, leading to acceptable rates of acid- ESI Fig. S6.
catalyzed dehydration. Smaller Ni crystallites on the Ni-USY did not Approximately 5 wt% Ni was impregnated on each material. The
benefit hydrogenation reaction rates, whereas the Ni-DS-USY suffered bulk Ni loading measured from ICP-OES was 4.78 wt% and 4.65 wt%
from lack of acidity. The characterization and reaction data suggest that for the Ni-DS-PDA(25)-USY and Ni-DS-PDA(10)-USY, respectively. The
large mesopores are beneficial for metal impregnation and reaction, as TPR profile of each Ni incorporated material is available in ESI Fig. S8.
long as substantial acidity is not sacrificed during synthesis. In the The TPR profiles of both new materials were very similar to the Ni-DS-
following section, we discuss the reduction of DS-PDA alkaline treat- PDA(50)-USY material presented in Fig. 3. Interestingly, the TPR of the
ment strength as a means to further improve the effectiveness of Ni-DS-PDA(10)-USY exhibited a shoulder at low reduction temperature
(305 °C). This is the same location as a prominent peak exhibited by the

26
D.P. Gamliel et al. Microporous and Mesoporous Materials 261 (2018) 18–28

Fig. 8. (A) Initial reaction rate (bars, left-axis), final conversion and overall mass balance (lines, right axis) for anisole HDO over Ni impregnated on parent and USY zeolites prepared with
DS-PDA method. (B) First order kinetic fit for apparent rate of anisole HDO over each catalyst. Reaction Conditions: Temperature: 200 °C, pressure: 750 psi, reaction time: 140 min.

parent Ni-USY, indicating that the smaller Ni crystallite precursors may Table 5
be present during reduction. This is correlated with greater micropore Final product molar selectivity for anisole HDO over Ni impregnated on parent and
mesoporous USY zeolites created using DS-PDA method with different NaOH con-
volume present in the DS-PDA(10)-USY support. The reducibility of
centrations. Selectivity is defined as moles of specific product formed over total moles of
each material remained high, greater than 90%, in all cases. The NiO product formed.
average crystallite size, measured by XRD line broadening, was com-
parable to that of all the other materials, between 27 and 35 nm. After Compound Selectivity (mol%)
reduction, the Ni crystallite size was 21.9 nm and 17.1 nm for the Ni-
Ni-DS- Ni-DS- Ni-DS-
DS-PDA(25)-USY and Ni-DS-PDA(10)-USY, respectively. The effects of
PDA(50)-USY PDA(25)-USY PDA(10)-USY
the increased microporosity by reduction of alkaline treatment strength
on the Ni morphology follows a trend. As micropore volume increased, Hydrocarbons
Ni migrated into the micropore network during reduction producing Cyclohexane 71.9 89.0 87.3
Cyclohexene 0.1 – –
smaller Ni crystallites, similar to those observed on the Ni-USY material
Oxygenates
in Fig. 5(F). From the conclusions reached in the previous section, it is 1-methoxycyclohexane 21.7 6.0 7.3
expected that the presence of smaller Ni crystals in the Ni-DS-PDA(10)- Cyclohexanol 2.7 0.1 0.2
USY catalyst may reduce the relative rate of Ni-catalyzed hydrogena- Cyclohexanone – 0.1 0.2
tion, reducing the overall rate of anisole HDO. C-C Coupling
1,1′ Bicyclohexyl 1.5 2.6 2.6
The mesoporous Ni materials were tested for the same anisole HDO Cyclohexene, 1-Cyclohexyl – – –
reaction. Fig. 8(A) shows the initial rates of reaction and final conver- 1,1′ bicyclohexanone – – –
sion, calculated after 140 min of reaction time. As expected, improving Phenol, 2-cyclohexyl-4-methyl 1.2 1.4 1.4
microporosity and acidity by reducing alkaline treatment strength im- Benzene, 1-Cyclohexyl-4- 0.8 0.8 0.9
methyl
proved the initial reaction rate. Improved reaction rate resulted in
greater anisole conversion. From Fig. 8(A) and (B), initial reaction rate
was higher over the Ni-DS-PDA(25)-USY catalyst, and fell slightly when Table 6
the Ni-DS-PDA(10)-USY catalyst was used. Table 5 shows the product Kinetic parameters from model fits for HDO of anisole over catalysts created using DS-
selectivity for HDO of anisole over each Ni-DS-PDA-USY catalyst. PDA method with different NaOH concentrations, normalized to the number of Ni (k1)
The rate constants of Ni-catalyzed hydrogenation (k1) and Brønsted and Brønsted acid (k2) active catalytic sites.
acid-catalyzed dehydration (k2), based on the model in Scheme 1, are
Catalyst Ni Sitesa B.A.S.b k1 k2
displayed in Table 6. The plot of each fit is available in ESI Fig. S9. As
presented in Table 6, the rate constant for hydrogenation is highly umol/gcat umol/gcat (1/s)/mol Ni (1/s)/mol B.A.S.
dependent on catalyst support. Hydrogenation was fastest with the Ni-
Ni-DS-PDA(50)-USY 43 145 46.6 22.7
DS-PDA(25)-USY material, correlating with the greater overall initial
Ni-DS-PDA(25)-USY 37 268 80.1 26.3
rate of reaction from Fig. 8(A). Hydrogenation was significantly slower Ni-DS-PDA(10)-USY 38 334 60.9 24.5
when the Ni-DS-PDA(10)-USY was used, compared to the Ni-DS-
a
PDA(25)-USY. This reduced Ni hydrogenation activity is related to the H2 Chemisorption.
b
decreased Ni crystallite size observed from XRD line broadening, which Pyridine FTIR.
suggests that some of the Ni has migrated into the pore network, be-
coming inaccessible for reaction (vide supra). The rate constants of acid- USY material. The pseudo-first order rate constant was calculated to be
catalyzed dehydration per active site were similar for the three cata- 0.99 L/(gcat*hr). This is an order of magnitude lower than the reaction
lysts, indicating that while alkaline treatment strength affects acidity, rate constant measured for anisole HDO over Pt/Al2O3 measured by
the differences in acid site location were not great enough to affect the Runnebaum et al. [53] (19 L/(gcat*hr)), and over CoMo/Al2O3 mea-
rates of acid-catalyzed reaction. sured by Hurff et al. [60](2.2–10 L/(gcat*hr)). A direct comparison
Anisole HDO proceeded most rapidly using the Ni-DS(250)-PDA- between these materials is not possible due to discrepancies in reactor
setup, reaction temperature and operating pressure.

27
D.P. Gamliel et al. Microporous and Mesoporous Materials 261 (2018) 18–28

In summary, reducing DS-PDA alkaline treatment concentration [8] K. Li, J. Valla, J. Garcia-martinez, ChemCatChem 6 (2014) 46–66.
[9] A.W. Chester, E.G. Derouane, Zeolite Characterization and Catalysis, fifth ed.,
increases the Brønsted acidity, crystallinity and micropore volume of Springer, New York, 2009.
the material. This increased micropore volume results in lower Ni [10] B.A. Williams, S.M. Babitz, J.T. Miller, R.Q. Snurr, H.H. Kung, Appl. Catal. A Gen.
crystallite size upon impregnation and reduction. Anisole HDO reaction 177 (1999) 161–175.
[11] S.J. DeCanio, J.R. Sohn, P.O. Fritz, J.H. Lunsford, J. Catal. 101 (1986) 132–141.
rate was higher using the Ni-impregnated DS-PDA materials compared [12] J. Zečević, C.J. Gommes, H. Friedrich, P.E. Dejongh, K.P. Dejong, Angew. Chem. -
to the parent Ni-USY. HDO of anisole proceeded faster over the material Int. Ed. 51 (2012) 4213–4217.
prepared with moderate alkaline treatment. Lower Ni crystallite size led [13] Z. Qin, B. Shen, X. Gao, F. Lin, B. Wang, C. Xu, J. Catal. 278 (2011) 266–275.
[14] N. Nuttens, D. Verboekend, A. Deneyer, J. Van Aelst, B.F. Sels, ChemSusChem 8
to reduced rates of the Ni-catalyzed hydrogenation reactions. The re- (2015) 1197–1205.
sults suggest that there must be a balance between surface-accessible Ni [15] K.P. de Jong, J. Zečević, H. Friedrich, P.E. de Jongh, M. Bulut, S. van Donk,
and sufficient acid sites present in the catalyst for efficient hydro- R. Kenmogne, A. Finiels, V. Hulea, F. Fajula, Angew. Chem. 122 (2010)
10272–10276.
genation and subsequent dehydration to take place.
[16] D.P. Gamliel, L. Wilcox, J.A. Valla, Energy Fuels 31 (2017) 679–687.
[17] S. Sartipi, K. Parashar, M.J. Valero-Romero, V.P. Santos, B. Van Der Linden,
4. Conclusions M. Makkee, F. Kapteijn, J. Gascon, J. Catal. 305 (2013) 179–190.
[18] G.T. Neumann, J.C. Hicks, ACS Catal. 2 (2012) 642–646.
[19] P.M. Mortensen, J.-D. Grunwaldt, P.A. Jensen, K.G. Knudsen, A.D. Jensen, Appl.
Mesoporous USY-supported Ni catalysts were synthesized, char- Catal. A, Gen. 407 (2011) 1–19.
acterized and studied for the HDO of anisole. Mild SA and stronger DS [20] W. Mu, H. Ben, A. Ragauskas, Y. Deng, Bioenergy Res. 6 (2013) 1183–1204.
and DS-PDA methods were used to produce varying mesopore struc- [21] A.V. Bridgwater, Biomass Bioenergy 38 (2012) 68–94.
[22] X. Zhang, T. Wang, L. Ma, Q. Zhang, T. Jiang, Bioresour. Technol. 127 (2013)
tures. Impregnation resulted in uniformly sized NiO particles. After 306–311.
reduction, Ni migrated from the surface of the USY supports with [23] M. Campanati, G. Fornasari, A. Vaccari, Catal. Today 77 (2003) 299–314.
greater micropore volume to produce smaller internal Ni particles, [24] A.J. Maia, B. Louis, Y.L. Lam, M.M. Pereira, J. Catal. 269 (2010) 103–109.
[25] T.M. Sankaranarayanan, A. Berenguer, C. Ochoa-Hernández, I. Moreno, P. Jana,
which were not completely reduced. Ni impregnated on mesoporous J.M. Coronado, D.P. Serrano, P. Pizarro, Catal. Today 243 (2015) 163–172.
USY supports produced larger Ni particles. [26] J.A. Hunns, M. Arroyo, A.F. Lee, J.M. Escola, D. Serrano, K. Wilson, Catal. Sci.
Mesoporous USY with bulk Ni particles on the surface were more Technol. 6 (2016) 2560–2564.
[27] C.Y. Liu, R.P. Wei, G.L. Geng, M.H. Zhou, L.J. Gao, G.M. Xiao, Fuel Process.
active for anisole conversion under HDO conditions. Kinetic modeling
Technol. 134 (2015) 168–174.
revealed that this increase in catalytic activity was attributed to the [28] D. Verboekend, G. Vilé, J. Pérez-Ramírez, Adv. Funct. Mater 22 (2012) 916–928.
higher rates of hydrogenation reactions per Ni atom. Acid-catalyzed [29] D. Verboekend, G. Vile, J. Perez-Ramirez, Cryst. Growth Des. 12 (2012) 3123–3132.
[30] J. García-Martínez, M. Johnson, J. Valla, K. Li, J.Y. Ying, Catal. Sci. Technol. 2
dehydration reaction constants were significantly lower for anisole
(2012) 987.
HDO over the Ni impregnated mesoporous materials, compared to the [31] D. Verboekend, J. Pérez-Ramírez, Chem. - A Eur. J. 17 (2011) 1137–1147.
parent, most likely due to limited zeolite acid site proximity. [32] K.X. Lee, J.A. Valla, Appl. Catal. B Environ. 201 (2017) 359–369.
The most efficient HDO catalyst support was created using the DS- [33] C.A. Emeis, J. Catal. 141 (1993) 347–354.
[34] S.A. Khromova, A.A. Smirnov, O.A. Bulavchenko, A.A. Saraev, V.V. Kaichev,
PDA method. Mild alkaline treatment strength resulted in retained S.I. Reshetnikov, V.A. Yakovlev, Appl. Catal. A Gen. 470 (2014) 261–270.
acidity, crystallinity and microporosity. Increasing the microporosity [35] R.C. Runnebaum, T. Nimmanwudipong, D.E. Block, B.C. Gates, Catal. Sci. Technol.
resulted in smaller Ni crystallites, leading to an eventual reduction in 2 (2012) 113.
[36] M. Thommes, K. Kaneko, A.V. Neimark, J.P. Olivier, F. Rodriguez-Reinoso,
the rate of Ni-catalyzed hydrogenation. Ni supported on mesoporous J. Rouquerol, K.S.W. Sing, Pure Appl. Chem. 87 (2015) 1051–1069.
USY zeolite with proper mesopore structure and acidity was shown to [37] E.F. Iliopoulou, S.D. Stefanidis, K.G. Kalogiannis, A. Delimitis, A.A. Lappas,
increase the rates of Ni-catalyzed hydrogenation reactions while pro- K.S. Triantafyllidis, Appl. Catal. B Environ. 127 (2012) 281–290.
[38] D.P. Gamliel, H.J. Cho, W. Fan, J.A. Valla, Appl. Catal. A Gen. 522 (2016) 109–119.
moting sufficient rates of acid-catalyzed dehydration. [39] C. Wu, L. Dong, J. Onwudili, P.T. Williams, J. Huang, ACS Sustain. Chem. Eng. 1
(2013) 1083–1091.
Acknowledgements [40] A. Penkova, S. Dzwigaj, W.R. Kefirov, K. Hadjiivanov, M. Che, J. Phys. Chem. C 111
(2007) 8623–8631.
[41] B. Pawelec, L. Daza, J. Fierro, J. Anderson, Appl. Catal. A Gen. 145 (1996) 307–322.
The authors would like to acknowledge the University of [42] S. Reddy Yenumala, S.K. Maity, D. Shee, Catal. Sci. Technol. 6 (2016) 3156.
Connecticut Office of Vice President (OVPR) Research of Excellence [43] O. Clause, M. Kermarec, L. Bonnevio, F. Villain, M. Che, J. Am. Chem. Soc. 114
(1992) 4709–4717.
Program (REP) (Award # 4634760) for funding this work. We would
[44] R. Nares, J. Ramírez, A. Gutiérrez-Alejandre, C. Louis, T. Klimova, J. Phys. Chem. B
also like to acknowledge Dr. Lichun Zhang and UConn Institute for 106 (2002) 13287–13293.
Material Science for help with TEM, as well as Bailey Gannett for her [45] A.G. Popov, A.V. Smirnov, E.E. Knyazeva, V.V. Yuschenko, E.A. Kalistratova,
help with material preparation. XPS was performed by Dr. William K.V. Klementiev, W. Grrunert, I.I. Ivanova, Microporous Mesoporous Mater 134
(2010) 124–133.
Willis, and we are grateful for his help and expertise. Finally, we would [46] S. Qiu, Y. Xu, Y. Weng, L. Ma, T. Wang, Catalysts 6 (2016) 134.
like to thank the UConn Center for Environmental Science and [47] V.K. Venkatakrishnan, J.C. Degenstein, A.D. Smeltz, W.N. Delgass, R. Agrawal,
Engineering for performing the ICP-OES. F.H. Ribeiro, Green Chem. 16 (2014) 792.
[48] K. Wang, K.H. Kim, R.C. Brown, Green Chem. 16 (2014) 727.
[49] S. Du, D.P. Gamliel, M.V. Giotto, J.A. Valla, G.M. Bollas, Appl. Catal. A Gen. 513
Appendix A. Supplementary data (2016) 67–81.
[50] A.J. Foster, J. Jae, Y. Cheng, G.W. Huber, R.F. Lobo, Appl. Catal. A Gen. 423–424
(2012) 154–161.
Supplementary data related to this article can be found at http://dx. [51] Y. Yu, X. Li, L. Su, Y. Zhang, Y. Wang, H. Zhang, Appl. Catal. A Gen. 447–448 (2012)
doi.org/10.1016/j.micromeso.2017.10.027. 115–123.
[52] J. Jae, G.A. Tompsett, A.J. Foster, K.D. Hammond, S.M. Auerbach, R.F. Lobo,
G.W. Huber, J. Catal. 279 (2011) 257–268.
References [53] R.C. Runnebaum, R.J. Lobo-Lapidus, T. Nimmanwudipong, D.E. Block, B.C. Gates,
Energy Fuels 25 (2011) 4776–4785.
[1] A. Corma, G. Huber, L. Sauvanaud, P. O’connor, J. Catal. 247 (2007) 307–327. [54] C. Zhao, S. Kasakov, J. He, J.A. Lercher, J. Catal. 296 (2012) 12–23.
[2] C. Christensen, K. Johannsen, E. Tornqvist, I. Schmidt, H. Topsoe, Catal. Today 128 [55] O. Jan, R. Marchand, L.C.A. Anjos, G.V.S. Seufitelli, E. Nikolla, F.L.P. Resende,
(2007) 117–122. Energy Fuels 29 (2015) 1793–1800.
[3] M. Choi, Z. Wu, E. Iglesia, J. Am. Chem. Soc. 132 (2010) 9129–9137. [56] H.S. Fogler, Elements of Chemical Reaction Engineering, fourth ed., Pearson
[4] D.P. Gamliel, G.M. Bollas, J.A. Valla, Energy Technol. 4 (2016) 1–13. Education, Upper Saddle River, NJ, 2006.
[5] W. Fan, M. a Snyder, S. Kumar, P.-S. Lee, W.C. Yoo, A.V. McCormick, R. Lee Penn, [57] P.M. Mortensen, J.D. Grunwaldt, P.A. Jensen, A.D. Jensen, Catal. Today 259 (2016)
A. Stein, M. Tsapatsis, Nat. Mater 7 (2008) 984–991. 277–284.
[6] J. Pérez-Ramírez, S. Abelló, A. Bonilla, J.C. Groen, Adv. Funct. Mater 19 (2009) [58] A.M. Robinson, J.E. Hensley, J. Will Medlin, ACS Catal. 6 (2016) 5026–5043.
164–172. [59] D. Verboekend, S. Mitchell, M. Milina, J.C. Groen, J. Pérez-Ramírez, J. Phys. Chem.
[7] J.C. Groen, J.A. Moulijn, J. Pérez-Ramírez, Microporous Mesoporous Mater 87 C 115 (2011) 14193–14203.
(2005) 153–161. [60] S.J. Hurff, M.T. Klein, Ind. Eng. Chem. Fundam. 22 (1983) 426–430.

28

S-ar putea să vă placă și