Maximizing Diesel Production through Oligomerization: A Landmark

Dec 31, 2014 - Maximizing Diesel Production through Oligomerization: A Landmark ... The global demand for diesel oil and gas oil, also known as middle...
0 downloads 0 Views 885KB Size
Review pubs.acs.org/IECR

Maximizing Diesel Production through Oligomerization: A Landmark Opportunity for Zeolite Research Oki Muraza* Chemical Engineering Department and Center of Research Excellence in Nanotechnology, King Fahd University of Petroleum & Minerals, Dhahran 30261, Saudi Arabia S Supporting Information *

ABSTRACT: High demand for diesel fuel has attracted researchers to revisit some refining processes. Hydrocracking and oligomerization are two targeted processes to produce diesel intensively. In this minireview, challenges in the acid-catalyzed oligomerization of olefins are briefly discussed. One-dimensional pore zeolites have been identified as potential catalytic materials for oligomerization. However, one-dimensional pore zeolites mostly suffer from a long diffusion path due to high crystal aspect ratio (length/width). Mass-transfer limitation are resolved by the development of more efficient catalysts; nanosized and hierarchical zeolites with the one-dimensional pore system such as TON, MTT, and MTW.

1. INTRODUCTION In the coming years, diesel consumption in US will increase while gasoline (C5−C12) demand is projected to decline (Figure 1).1 A similar trend was also observed in Europe and

Figure 2. Global consumption of different fuels [data taken from World Oil Outlook, 2013, Organisation of Petroleum Exporting Countries].

rating. Typical values of cetane number are in the range of 40 to 55. By increasing the cetane number, the speed and efficiency of the engine increase as well.8,9 In petroleum refining, diesel can be synthesized by the following processes (Figure 3): (1) atmospheric distillation of crude oil, (2) catalytic hydrocracking of atmospheric gas oil, and (3) oligomerization of light-cracked naphtha (LCN). Therefore, to maximize the diesel production rate, strategies are pursued such as hydrocracking of cheap gas oil from a vacuum distillation unit (VDU) and oligomerization of light-cracked naphtha (LCN), a byproduct from fluidized catalytic cracking (FCC) unit. Hydrocracking is an industrial catalytic hydrogenation process in which heavy feedstocks such as vacuum gas oil and other residual oils are cracked and hydrogenated to lower molecular weight products over bifunctional catalysts.

Figure 1. Projected diesel consumption in US (International Energy Agency).

Japan. The global demand for diesel oil and gas oil, also known as middle distillates, is expected to grow by 12 million barrel from 2012 to 2035 (Figure 2).2 The needs for higher diesel production stimulated energy providers to produce more diesel from different sources, be it from fossil fuel3 or from renewable sources (biodiesel) from algae,4 palm oil,5 and many others.6,7 Diesel fuel (C11−C24) is defined as a mixture of long hydrocarbon molecules with boiling points between 150 and 380 °C, obtained by fractionation of crude oil in the refinery or customized catalytic processes. Diesel is used as a fuel in truck engines, aircrafts, and locomotive engines. It is also used as drilling fluids and miscellaneous applications. Among others, the important properties of diesel are cetane number (C16H34), also known as cetane index, viscosity, density, and sulfur content.8 Cetane rating is the most important property as the performance of a diesel engine strongly depends on the cetane © 2014 American Chemical Society

Received: Revised: Accepted: Published: 781

October 19, 2014 December 26, 2014 December 31, 2014 December 31, 2014 DOI: 10.1021/ie5041226 Ind. Eng. Chem. Res. 2015, 54, 781−789

Review

Industrial & Engineering Chemistry Research

Figure 3. Potential streams for maximizing diesel production in refinery.

The typical catalysts consist of a metallic part to promote hydrogenation and an acidic material to induce cracking. Hydrogenation has a hydrotreatment role to remove impurities in the feedstock such as sulfur, nitrogen, and metals. Cracking breaks the carbon bonds in heavy hydrocarbons resulting in unsaturated products, which are sequentially converted into saturated paraffins. In addition to olefinic streams in petroleum refineries, olefins can also be produced from different sources as shown in Figure 4. The methanol-to-olefins process, for instance, is a technology to produce C2−C4 olefins from methanol, a flexible intermediate in the chemical industry. This option is famous in countries with abundant reserves of natural gas and coal. Olefins are also the typical product in the Fischer−Tropsch plant.10,11 The renewable route from biomass is also potentially integrated with petro-diesel as glycerol can be fed to oligomerization units.4,7,12−15 While, the biodiesel can be blended with petrodiesel.4,7,16 Previously, some interesting reviews were compiled elsewhere.17−21 In this work, the discussion is limited to oligomerization in a refinery using zeolites as catalyst. Special attention was given to the topology and pore system and their effects on oligomerization.

Figure 4. Role of oligomerization in diesel production from different natural sources.

2. OLIGOMERIZATION OF LIGHT CRACKED NAPHTHA AND OTHER OLEFINIC STREAMS An oligomer is defined as a complex molecule that consists of a few units of monomer. For instance, dimers, trimers, and tetramers are oligomers with two, three, and four monomers, respectively. Oligomerization is a chemical process where the monomer units are converted to oligomer complexes through a finite degree of polymerization. 782

DOI: 10.1021/ie5041226 Ind. Eng. Chem. Res. 2015, 54, 781−789

Review

Industrial & Engineering Chemistry Research

°C, typically 200−220 °C).44,45 When the temperature was high (above 300 °C), a higher degree of branching was observed. Meanwhile, gasoline range products are favored by low pressure (ambient pressure) and high temperature, where short hydrocarbon compounds and aromatics dominate the product distribution. High temperature is unwanted in oligomerization as cracking starts to happen above 217 °C over H-ZSM-5 catalysts as reported by van Den Berg et al.46 In general, the reaction rate of alkene oligomerization such as propene oligomerization increases by the increase of (a) temperature, (b) pressure, and (c) concentration of propylene in the reactant phase.41 The other crucial process parameter to product composition is space velocity (sometimes it is called as the weight hourly space velocity, WHSV).47,48 Different space velocities from 0.1 to 50 h−1 were explored for zeolite catalysts. Belussi et al. reported that the conversion of alkenes over MFI (H-ZSM-5) decreased with the increase of space velocity above 1 h−1.42 Different types of reactors have been reported in oligomerization.49 It can be performed in a fixed bed reactor, moving-bed, or in fluidized bed reactor. As oligomerization is an exothermic reaction, heat removal strategies should be incorporated in reactor design. There are different strategies such as (a) dilution with the feed, (b) installing heat-exchange in between two consecutive beds, and (c) heat removal through the reactor wall.41 After leaving the reactor, the diesel ranged products have to be recovered in fractionation using distillation. Oligomerization is a unimolecular reaction with typically first rate order dependence on the monomer concentration. The oligomerization reaction rate, especially when ethylene and propylene are used as feedstock, is limited by the mass transfer (diffusion) rather than the kinetics. The temperature of the reaction ranges from 150 to 500 °C with operating pressure ranges between 1 and 50 bar. The variance in the reaction conditions depends strongly on the type of the catalyst used in this process. Oligomerization is a function of temperature. If the reaction temperature increases, the reaction rate increases significantly. Deactivation is another serious challenge in oligomerization. There are different causes of catalyst deactivation in LCN oligomerization such as coke deposition and sulfur or nitrogen compounds contained in the feed. The latter deactivation problem due to poisoning can be overcome by installing a hydrotreating unit to remove sulfur and nitrogen from the feed stream. The deactivation rate in LCN oligomerization depends on the type of the catalyst, composition of LCN, and process severity. Hauge et al. reported that dimerization of isobutene over zeolites in the liquid phase at low temperature in the range of 30 to 70 °C suffered from rapid deactivation.31 Deactivation is a function of catalyst size; larger extrudate sizes result in a higher deactivation rate.41 Corma et al. reported that the crystal size was an important property for lower deactivation.33 The large crystals of ZSM-5 favored the formation of coke and induced faster deactivation. Coelho et al. recently reported that the narrow pore size affected oligomerization positively as the smaller pore hindered severe coke formation.50 The deactivation rate of 1-butene oligomerization decreased with the increasing contact time (1/WHSV).50 Deactivation occurred in the first 10 min of reaction time in 1-butene oligomerization and later stabilized at longer times on stream. The mechanisms of deactivation basically can be divided into two different types: (1) deposition of cokes at the active sites (either Lewis or

The feedstocks of oligomerization are typically alkene streams. Propylene and other longer olefins, in particular C4 and C5 olefins, are preferred over ethylene as higher temperature is required to activate ethylene.22 In a typical industrial refinery, light cracked naphtha (LCN) is one of the side products produced in a FCC unit as shown in Figure 3. LCN comprises many unsaturated alkenes such as butylene (C4H8), pentene (C5), and hexene (C6) and others. This olefinic stream can be processed further by an oligomerization process to produce diesel fuel with a high cetane rating. In addition to LCN, the FCC unit also produces light cycle oil (LCO), which has a boiling point within the diesel boiling range. The LCO is rich in aromatics and has a low cetane number (CN: 15−24).23,24 To meet diesel specification (CN: 40−60), LCO must be sent to the hydrotreating unit.25 Figure 5 shows a simplified mechanism of olefin oligomerization. For instance, the olefins are n-propene (C3H6) and n-

Figure 5. Mechanism of oligomerization.

butene (C4H8). Sometimes, longer olefins such as n-heptene (C7H14) and n-octene (C8H16) can be used as feedstock as well. Notice that, in all oligomerization reactions, the molecular weight of the reactants increased. These typical reactions can be applied in the same way using the light cracked naphtha to produce diesel fuel. Theoretically, different olefins from propylene to heptene can be used as the feedstock for diesel production through oligomerization. Oligomerization has been explored by using different types of catalysts such as classical solid phosphoric acids,26−29 ionexchange resins,30−32 aluminosilicates (zeolites) (for instance33,34), and metal oxides.35−37 Relatively new materials such as ionic liquids37,38 and metal organic frameworks39,40 were also explored as catalysts for oligomerization of alkenes. For diesel production, the following components are required: (i) carbenium ions, (ii) steric hindrance to control the oligomeric products, and (iii) suppression of isomerization.34 The oligomerization of alkenes to diesel fuels happens through consecutive reaction steps to finally produce middle distillate fuels (diesel oil and gas oil). The oligomerization of light olefins has been successfully applied in refining industries. At least, there were two industrial oligomerization processes by using zeolites: (a) Mobil olefins to gasoline and distillate (MOGD) developed by Mobil Oil (now ExxonMobil) and (b) conversion of olefins to diesel, a process developed by Lurgi. There is also another classical processes using solid phosphoric acid (SPA) supported on kieselguhr (also known as the Catpoly process).11,41 Detailed discussion on the process parameters such as temperature and pressure are not discussed in details in this compilation. Oligomerization is a highly exothermic process (approximately 20 kcal/mol per double bond42 or 1.046−1.381 kJ/g of reacted olefins43). It is well-known that the selectivity to diesel in oligomerization is favored by high pressure (above 30 bar, for instance 50 bar) and low temperature (lower than 300 783

DOI: 10.1021/ie5041226 Ind. Eng. Chem. Res. 2015, 54, 781−789

Review

Industrial & Engineering Chemistry Research Table 1. Catalytic Performances of Selected Zeolites in Oligomerization with Optimized Conditions framework

zeolite

pore sizeb

Si/Al

TON TON

Theta-1 ZSM-22

M M

25 30

FER FER

ferrierite ferrierite

M M

66 6

MOR OFF MTW MFI MFI MFI

mordenite offretite ZSM-12 ZSM-5 ZSM-5 ZSM-5

L L L M M M

30 9 100 30 15 15

T [°C]

feed 60:40 propene/propane molar mixture 2-butene (16 wt %), propane (78 wt %), and pentane (6 wt %) 1-butene 2-butene (16 wt %), propane (78 wt %), and pentane (6 wt %) 1-butene 1-butene 1-butene 1-butene 60:40 propene/propane molar mixture 60:40 1-pentene/heptane molar mixture

P [bar]

200 150

40 60

190 140

46 60

230 250 200 210 200 200

44 47 41 42 40 40

WHSV [h−1] a

0.17 W/F = 115 kg s mol 2.03 W/F = 115 kg s mol 2.01 0.55 3.05 1.50 0.17a 0.17a

selc [%]

X [%] 92 82

ref

99.9 66

77 37.9 wt % to C12 and higherc 66.9 21.4

51 52 53 52

99.8 99.9 99.9 100 82 97

62.9 68.3 74.8 71.8 75 55

53 53 53 53 54 54

a Contact time (τ) is defined as 1/WHSV [h]. WHSV is weight hourly space velocity, a ratio of mass flow to catalyst mass. bPore size for 10- and 12membered rings are can be classified as medium (M) for 10-MR and large (L) for 12-MR. cSelectivity to oligomers or diesel range products. Sometimes, the products are classified according to the boiling point cut. Naphtha is defined for products with a boiling cut point up to 177 °C. Middle distillate is defined for products by boiling range cut point from 177 to 343 °C. Above 343 °C (up to 524 °C) is classified as heavy distillate. Diesel, together with heating oils are classified as middle distillate fuels.

kinetic diameter of 0.56. The multibranched alkane such as 2,2,3-trimethylbutane has a larger kinetic diameter (0.62 nm) than most of 10-membered ring zeolites. Selectivity to diesel range products was higher for medium pore zeolites (ZSM-5 and ZSM-12).61 ZSM-12 was stable for a longer operating time up to 300 h in oligomerization. Apart from medium-sized zeolites with 10-MR rings, large zeolites such as ITQ-39 were reported for on-purpose diesel production.62 ITQ-39 is a 3D-pore zeolite consisting of a 12MR channel (0.65−0.79 nm) and 10-MR channels (0.55 nm × 0.57 nm, 0.54 nm × 0.60 nm, and 0.45 nm × 0.57 nm). Some patents and open literature also reported that the onedimensional pore zeolites with 10 MR channels have the ability to reduce the branching and to suppress the formation of aromatics.34,63,64 Therefore, one-dimensional pore zeolites such as TON, MTT, and EUO structures have potential for maximizing diesel production in oligomerization reactions. However, the crystal morphologies of one-dimensional pore zeolites are typically long rods and needles with mass-transfer problems. As the pores are normally parallel to the crystal length, diffusion limitations hinder catalyst effectiveness. In addition to low catalyst utilization issue, one-dimensional pore systems suffer from pore-mouth poisoning, in which the blocking of the channels happens very rapidly and deactivation rate is very high. The strategies to improve the effectiveness of one-dimensional pore zeolites by using different approaches65 are shown in Figure 6. The first approach is by creating mesopores to improve diffusion and accessibility to the active sites. Several studies on hierarchical one-dimensional pore zeolites for different frameworks have been reported such as for TON (ZSM-22),66,67 MTT (ZSM-23),68 MOR (mordenite),69,70 and MTW (ZSM-12).71,72 Mesoporous ZSM-12 zeolite was developed by alkaline treatment with different concentrations of NaOH. Acidity of the hierarchical zeolites was not directly related with desilication severity. The second approach to obtain higher catalyst effectiveness of one-dimensional pore zeolites is by creating submicrometer and nanosized crystals. The nanocrystalline zeolites have been explored for numerous zeolite frameworks,73 including some frameworks with one-dimensional pore systems such as TON (ZSM-22),74 MTT (ZSM-23),75 MTW (ZSM-12),76 and EUO

Brønsted acids) and (2) deposition at the pore mouth (pore plugging). Conversion and selectivity are other important aspects in the LCN oligomerization and they are strongly related to each other. The conversion of LCN to diesel is a function of numerous parameters such as process conditions, catalyst types, catalyst deactivation, and feed properties. The typical conversion of LCN in an oligomerization process is above 70%. The conversion of butylene oligomerization process with ZSM-5 (Si/Al = 30) exceeded 98%.50 Also, the selectivity toward diesel fuel depends strongly on catalyst types, reaction conditions, and the composition of the LCN. Table 1 and Supporting Information, Table S1 summarize oligomerization of olefinic streams over porous materials, especially zeolites, including conversion of feedstock and diesel selectivity.

3. CATALYST DESIGN IN OLIGOMERIZATION 3.1. Effect of Topology (1-D versus 3-D). Pore topology is very important in catalyst design for oligomerization. The basic idea involves the restriction of isomerization (the tendency of branching), which depends on the pore size of the zeolites and the channel system; either 3-dimensional (3-D) or 1-dimensional (1-D) pore system. The typical medium pore zeolites with 10-membered rings (10 MR) such as ZSM-5 have been reported as catalysts to restrict the degree of branching.55 Other medium pore zeolites such as ZSM-11 (MEL) and ZSM12 (MTW) were also reported as potential materials for diesel production.55,56 Different frameworks such as FER,57 ERI,58 MFS,52 NES, and EUO,59 were reported for diesel production in patents and the open literature. In general, zeolite with 10MR and 12- MR channels are potential catalysts for oligomerization. Datema et al. reported a fundamental study using nuclear magnetic resonance (NMR) on the effect of pore size.60 The level of branching in ethylene oligomerization was increased with the increase of pore sizes. The findings are in agreement with the pore size of 10-MR zeolites; for instance, ZSM-5 has a pore diameter in the range of 0.55 and 0.56 nm. The kinetic diameter of a short alkane such as n-heptane is smaller than 0.5 nm (approximately 0.43 nm for n-heptane). The monobranched alkanes such as 2methyl hexane have a kinetic diameter of 0.5 nm. A dibranched alkane such as 2,3-dimethyl pentane is slightly larger with a 784

DOI: 10.1021/ie5041226 Ind. Eng. Chem. Res. 2015, 54, 781−789

Review

Industrial & Engineering Chemistry Research

The most important factor in propylene oligomerization over H-ZSM-5 was the concentration of Bronsted acid sites that were available on the inner surface of zeolite pores.33 With the reaction inside the pores, the beauty of the shape-selectivity concept, especially the product shape selectivity, can be reached. The yields of the branched alkanes and aromatics were suppressed.42,61,83 The reaction outside the pore led to higher levels of branching. Therefore, there were numerous efforts to deactivate the external surface of the zeolite crystals, in such a way that the branched products were reduced. The presence of external acid sites can be removed post-treatment using oxalic acid (dealumination) as reported by Corma et al.33 The external surface of zeolites was also deactivated or modified using dicarboxylic acid, 2,4,6-collodine,88 hexamethyl-disilazane,89 and 4-methylquinoline.78 To selectively deactivate the external surface of zeolites, large molecules are normally used. For instance, 2,4,6-collodine or 2,4,6-trimethylpyridine was used to deactivate the outer surface of ZSM-23 (MTT). ZSM23 comprises 10-MR channels with pore size less than 0.6 nm (0.45 nm × 0.52 nm). The kinetic diameter of 2,4,6trimethylpyridine (0.74 nm)90 is certainly larger than the pore of H-ZSM-23. Another bulky pyridine-based modifier is 2,6-di-tert-butylpyridine, which has a larger kinetic diameter of 1.05 nm.90 Wilshier reported silylated ZSM-5 for oligomerization of propylene in a fixed bed reactor with a WHSV of 0.7 h−1.89 The zeolite was silylated with hexamethyl disilazane (HMDS, (CH3)3SiNHSi(CH3)3) at 20 °C for 6 h. The silylated ZSM-5 showed better shape-selectivity properties than the unmodified zeolite. HMDS was used to deactivate the active sites on the external surface of zeolite, but it was too bulky to enter the pore of ZSM-5. The surface of the hydroxyl groups was converted into trimethylsiloxyl groups. The nitrogencontaining heterocyclic aromatic compound such as 4methylquinoline (C10H9N) was also used for similar purposes. The kinetic diameter of 4-methyl quinoline (0.73 nm) was too large to enter the pore and to modify the inner surface of HZSM-5 (0.51 nm × 0.55 nm).91 When the external surface of H-ZSM-5 was deactivated, the structures of the C20+ products were mostly linear alkanes with a low level of branching.84 Oligomerization over the external surface of H-ZSM-5 resulted in more branching products. The nature of acidity (Brønsted or Lewis) is also crucial for diesel selectivity in oligomerization. Brønsted acidity is preferred for oligomerization of olefins to diesel. The presence of strong Lewis sites may initiate other side reactions such as polymerization, which produces branched products, aromatics, and finally cokes.78 Zeolite composition is often reported as the silicon to aluminum (Si/Al) ratio. The typical values of Si/Al of zeolites used in oligomerization are between 45 to 150. The higher is the Si/Al ratio, the stronger is the acid strength of zeolite. While, acid site density is inversely proportional to Si/ Al ratio.87 Mlinar et al. reported that as the Si/Al ratio decreased, the activity of H-ZSM-5 decreased, and the selectivity to shorter growth (dimerization) increased.44 Smaller Si/Al ratio had a positive effect on oligomerization of propylene over H-ZSM-5.92 A higher Si/Al ratio of H-ZSM-5 yielded slightly lower selectivity to diesel products (C12−C20). The catalysts for oligomerization of olefins for on-purpose diesel production are mostly protonated zeolites such as HZSM-5, H-ZSM-22, and H-mordenite. However, there are significant efforts to develop metal-exchanged zeolites with metals from Group IA and IIA such as Ni, Ga, Zn, et cetera.40,86,93,94 Diesel yields increased when Ni ion was

Figure 6. Strategies to improve one-dimensional pore zeolites for oligomerization.

(EU-1).77 Apart from one-dimensional pore systems and threedimensional pore systems, two-dimensional pores such as ZSM-57 (MFS) also showed interesting catalytic activity in oligomerization.52 3.2. Effect of Porosity on Mass-Transfer Properties and Deactivation. The most widely used catalysts in oligomerization are medium pore size zeolites that suppress the quantity of branched products and increase the cetane rating. Large-pore zeolite catalysts favor the formation of branched products. For instance, zeolite omega (MAZ) (12 MR: 0.74 nm × 0.74 nm, 8 MR: 0.31 × 0.31), a large-pore zeolite, favored highly branched products.22,78−80 On the other hand, smaller pore sizes (micropores) cause easy poisoning and eventually lead to catalyst deactivation. Porosity is another important catalyst characterization along with pore size and it affects the catalyst lifetime. It has been found that the best performances in terms of activity and oligomerization selectivity were obtained over mesoporous catalyst.81 Zeolites with the presence of mesopores in addition to unique micropores are known as hierarchical zeolites.82 Under the same reaction conditions, the larger mesoporous TON catalyst is, the higher is the catalytic activity.34 However, if the porosity is too large, there will be less shape-selectivity effect and consequently aromatics will be formed. Peratello et al. reported the application of mesoporous SiO2−Al2O3 with the clear presence of mesopores (diameter of 4 nm).41 They proposed the Langmuir-Hinswelwood-HougenWatson kinetics of propylene oligomerization over mesoporous SiO2−Al2O3 and reported that oligomerization suffered from mass-transfer limitation. 3.3. Effect of Acidity. Theoretically, when the reaction takes place inside the pore (at Brønsted acidic sites), low branched diesels (linear products) with a high cetane number are produced. On the other hand, if the series of reactions takes place over external acid sites, high branched diesel fuels with low cetane number are the dominant products.83 Numerous studies have reported the oligomerization reaction happened at Brønsted acid sites inside the pores.30,84−86 Reduction in the density of the Brønsted acid sites affected lower selectivity to diesel range products,33 even though both Brønsted acid (proton donor) and Lewis acid (electron acceptor) sites play crucial roles in carbocation chemistry.18,87 785

DOI: 10.1021/ie5041226 Ind. Eng. Chem. Res. 2015, 54, 781−789

Review

Industrial & Engineering Chemistry Research Table 2. Some Selected One-Dimensional (1D) Pore Zeolites with 8- and 10-Membered Rings109 zeolite

framework

Si/Al [−]

pore size [nm]

ring

remarks

ref

crystal shape

MAPO-39

ATN

0.40 × 0.40

8

|H+n| [MgnAl8‑n P8 O32], P(50), Al(40), Mg(9)

110

AIPO-41/SAPO-41

AFO

0.43 × 0.70

10

Al10P10O40, Al(51), P(46), Si(3)

111, 112

AIPO4−11

AEL

0.4 × 0.65

10

Al(50): P(50)

113−115

AIPO-H2 EU-1

AHT EUO

0.33 × 0.68 0.41 × 0.54

10 10

ZSM-23

MTT

32.3

0.45 × 0.52

10

Theta-1/ZSM-22

TON

19

0.46 × 0.57

10

|(H2O)8| [Al6P6O24] |Na+n (H2O)26| [AlnSi112‑n O224], n < 19, typically n ≈ 3.6 Si(96.5), Ga(3.5) |Na+n (H2O)4| [AlnSi24‑n O48], n < 2. submicron particles were synthesized using microwaveassisted synthesis |Na+n (H2O)4| [AlnSi24‑n O48], n < 2

small irregular platelike particles, 2.5 μm × 3.3 μm hexagonal prisms (with diameter ca. 2 μm) and 6 μm long). different morphologies such as spheres and rods were reported needle-like crystals spheres (2−3 μm)

exchanged with Na cation for zeolite Y (FAU).93 Different heteroatoms (M) were also incorporated into the mesoporous silica SBA-15 structure, where M was Al, Fe, B, and Cr for oligomerization of 1-hexene.95,96 In general, the reaction rate is a function of heteroatom content. Cr-exchanged SBA-15 and Al-exchanged SBA-15 were more active than B- and Femesoporous silica, as Cr and Al have higher acid strength. However, selectivity to heavy oiligomers was higher for B-SBA15.96 The Al-SBA-15 contained both tetrahedral and octahedral Al center. The acid-strength increased with the aluminum content from 0.1 mequiv NH3 g−1 (for Si/Al of 86) to 0.56 mequiv NH3 g−1 (for Si/Al of 12). The higher the Al content is, the higher is the ratio of octahedral/tetrahedral. Rare earth elements such as lanthanum and cerium are also potentially applied for oligomerization. The metals or cations can be inserted by using well-established ion exchange procedures.

116 77

68, 103 98, 99, 117

bundles of needles needles

lifetime.34 The keys of catalyst stability are the preserved Brønsted acidity and better mass-transfer properties. Table 2 and Supporting Information, Table S2 present different types of one-dimensional pore zeolites with different frameworks. There is plenty room for improvement of the morphology and the porosity of one-dimensional pore zeolites as most of them are less studied. In summary, the oligomerization performance of nanosized and hierarchical one-dimensional zeolites might be improved as shape-selectivity was maintained while the mass transfer properties were enhanced significantly. The nanocrystals and the hierarchical structures also stabilize the zeolite catalysts against deactivation.

5. CONCLUSIONS Oligomerization of olefinic streams is a very important process in oil refining to maximize diesel production. The olefinic streams are mostly available from the FCC unit. Plenty of activities are required to systematically screen and to optimize catalysts for production of diesel fuel through oligomerization. One of the most potential catalysts for this process is microporous zeolites. The most important factor in oligomerization of olefinic streams over H-ZSM-5 is the preserved Bronsted acid sites on the internal surface of zeolite pores coupled with better mass transfer properties. The medium pore zeolites have been reported as the best choice. Among the medium-pore 10-MR zeolites, one-dimensional pore zeolites such as TON and MTT suppressed the formation of branched alkanes and maintained high selectivity to oligomerization toward diesel compounds. However, the mass transfer properties of one-dimensional pore zeolites need to be improved either by the creation of mesopores or by the fabrication of nanosized zeolites. Similar studies of numerous zeolites with the one-dimensional pore system will open a plethora of dedicated research for the oligomerization over effective catalysts of olefinic streams to diesel.

4. POTENTIAL DEVELOPMENT OF IMPROVED 1D PORE ZEOLITES To improve mass transfer properties of zeolites, especially the one-dimensional pore zeolites in oligomerization and in related acid-catalyzed reactions, two approaches have been explored such as (a) creation of mesopores97 and (b) reduction of crystal size73 as illustrated in Figure 6. Different strategies have been explored to modify the size of one-dimensional pore zeolites by controlling the nucleation and crystallization for one-dimensional pore zeolites such as ZSM-12 (MTW),76 EU-1 (EUO),77 ZSM-22 (TON),98−102 ZSM-23 (MTT),68,103 and mordenite (MOR).104,105 Size and morphology are a function of Si/Al ratio, alkalinity and aging time.76,77,103 The challenge is with the crystal habit; most one-dimensional pore zeolites are nanoneedles with high agglomeration rate.99 Mesopores were also created in zeolites with the onedimensional pore system to improve stability against coke deposition. Among others, some mesoporous zeolites derived from one-dimensional pore systems were studied such as ZSM12 (MTW),71,72 ZSM-22 (TON),66,67,106 NU-10 (TON),107 ZSM-23 (MTT),68,103 and mordenite (MOR).70,108 The posttreatment of aone-dimensional system is quite unique as more severe conditions are required to remove Si atoms.70 Mild dealumination was required to remove partial blocking due to deposition of Al species during alkaline treatment.66 Posttreatment of H-ZSM-22, a one-dimensional pore zeolite, was applied in propylene oligomerization to prolong catalyst



ASSOCIATED CONTENT

* Supporting Information S

Additional catalytic activities of several zeolites in oligomerization with reaction parameters and some selected onedimensional (1D) pore zeolites with 10-, 12-, and 14membered rings. This material is available free of charge via the Internet at http://pubs.acs.org. 786

DOI: 10.1021/ie5041226 Ind. Eng. Chem. Res. 2015, 54, 781−789

Review

Industrial & Engineering Chemistry Research



(19) O’Connor, C. T.; Kojima, M. Alkene oligomerization. Catal. Today 1990, 6, 329−349. (20) Schwarz, S.; Kojima, M.; O’Connor, C. T. Appl. Catal. 1989, 56, 263. (21) Corma, A.; Iborra, S. Oligomerization of Alkenes. In Catalysts for Fine Chemical Synthysis; John Wiley & Sons, Ltd: Chichester, England, 2006; pp 125−140. (22) Kofke, T. J. G.; Gorte, R. J. A temperature-programmed desorption study of olefin oligomerization in H·ZSM-5. J. Catal. 1989, 115, 233−243. (23) Corma, A.; Martínez, C.; Sauvanaud, L. New materials as FCC active matrix components for maximizing diesel (light cycle oil, LCO) and minimizing its aromatic content. Catal. Today 2007, 127, 3−16. (24) Sharafutdinov, I.; Stratiev, D.; Shishkova, I.; Dinkov, R.; Petkov, P. Industrial investigation on feasibility to raise near zero sulfur diesel production by increasing fluid catalytic cracking light cycle oil production. Fuel Process. Technol. 2012, 104, 211−218. (25) Tailleur, R. G. Low-emission diesel production by upgrading LCO plus SR diesel fractions. Catal. Today 2008, 130, 492−500. (26) de Klerk, A. Distillate Production by Oligomerization of Fischer−Tropsch Olefins over Solid Phosphoric Acid. Energy Fuels 2006, 20, 439−445. (27) Coetzee, J. H.; Mashapa, T. N.; Prinsloo, N. M.; Rademan, J. D. An improved solid phosphoric acid catalyst for alkene oligomerization in a Fischer−Tropsch refinery. Appl. Catal., A 2006, 308, 204−209. (28) Prinsloo, N. M. Solid phosphoric acid oligomerisation: Manipulating diesel selectivity by controlling catalyst hydration. Fuel Process. Technol. 2006, 87, 437−442. (29) Busca, G., Other Solid Acid and Basic Catalytic Materials. In Heterogeneous Catalytic Materials; Busca, G., Ed.; Elsevier: Amsterdam, 2014; Chapter 8, pp 251−296. (30) Bringué, R.; Cadenas, M.; Fité, C.; Iborra, M.; Cunill, F. Study of the oligomerization of 1-octene catalyzed by macroreticular ionexchange resins. Chem. Eng. J. 2012, 207−208, 226−234. (31) Hauge, K.; Bergene, E.; Chen, D.; Fredriksen, G. R.; Holmen, A. Oligomerization of isobutene over solid acid catalysts. Catal. Today 2005, 100, 463−466. (32) Kriván, E.; Valkai, I.; Hancsók, J. Investigation of production of motor fuel components on heterogeneous catalyst with oligomerization. Top. Catal. 2013, 56, 831−838. (33) Corma, A.; Martínez, C.; Doskocil, E. Designing MFI-based catalysts with improved catalyst life for oligomerization to high-quality liquid fuels. J. Catal. 2013, 300, 183−196. (34) Martínez, C.; Doskocil, E. J.; Corma, A. Improved THETA-1 for Light Olefins Oligomerization to Diesel: Influence of Textural and Acidic Properties. Top. Catal. 2014, 57, 668−682. (35) Chang, C. D.; Hellring, S. D.; Marler, D. O.; Santiesteban, J. G.; Vartuli, J. C., Process for producing low aromatic diesel fuel with high cetane index. US Patent US5639931 A, 1997. (36) Tzompantzi, F.; Mantilla, A.; Del Angel, G.; Padilla, J. M.; Fernández, J. L.; Díaz-Góngora, J. A. I.; Gómez, R. NiO−W2O3/Al2O3 catalysts for the production of ecological gasoline: Effect of both NiO and the preparation method on the isobutene oligomerization selectivity. Catal. Today 2009, 143, 132−136. (37) Feher, C.; Krivan, E.; Hancsok, J.; Skoda-Foldes, R. Oligomerisation of isobutene with silica supported ionic liquid catalysts. Green Chem. 2012, 14, 403−409. (38) Gu, Y.; Shi, F.; Deng, Y. SO3H-functionalized ionic liquid as efficient, green and reusable acidic catalyst system for oligomerization of olefins. Catal. Commun. 2003, 4, 597−601. (39) Mlinar, A. N.; Keitz, B. K.; Gygi, D.; Bloch, E. D.; Long, J. R.; Bell, A. T. Selective Propene Oligomerization with Nickel(II)-Based Metal−Organic Frameworks. ACS Catal. 2014, 4, 717−721. (40) Mlinar, A. N.; Shylesh, S.; Ho, O. C.; Bell, A. T. Propene oligomerization using alkali metal- and nickel-exchanged mesoporous aluminosilicate catalysts. ACS Catal. 2013, 4, 337−343. (41) Peratello, S.; Molinari, M.; Bellussi, G.; Perego, C. Olefins oligomerization: Thermodynamics and kinetics over a mesoporous silica−alumina. Catal. Today 1999, 52, 271−277.

AUTHOR INFORMATION

Corresponding Author

*Tel.:+966 13 860 7612. E-mail: [email protected]. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The author acknowledges the support provided by King AbdulAziz City for Science and Technology (KACST) through the Science & Technology Unit at King Fahd University of Petroleum & Minerals (KFUPM) for funding this work through Project 11-NAN2166-04 as part of the National Science, Technology and Innovation Plan. O.M. acknowledges a prestigious grant from Saudi Aramco on the development of zeolites for oil upgrading.



REFERENCES

(1) Agency, I. E. The IEA World Energy Outlook. https://www.iea. org/stats (accessed Oct. 2014). (2) OPEC World Energy Outlook. http://www.opec.org/ (accessed Oct. 2014). (3) Stanislaus, A.; Marafi, A.; Rana, M. S. Recent advances in the science and technology of ultra low sulfur diesel (ULSD) production. Catal. Today 2010, 153, 1−68. (4) Galadima, A.; Muraza, O., Biodiesel production from algae by using heterogeneous catalysts: A critical review. Energy. (5) Mekhilef, S.; Siga, S.; Saidur, R. A review on palm oil biodiesel as a source of renewable fuel. Renew. Sustainable Energy Rev. 2011, 15, 1937−1949. (6) Bezergianni, S.; Dimitriadis, A. Comparison between different types of renewable diesel. Renew Sust Energ Rev. 2013, 21, 110−116. (7) Knothe, G. Biodiesel and renewable diesel: A comparison. Prog. Energy Combust. Sci. 2010, 36, 364−373. (8) Yang, H.; Ring, Z.; Briker, Y.; McLean, N.; Friesen, W.; Fairbridge, C. Neural network prediction of cetane number and density of diesel fuel from its chemical composition determined by LC and GC−MS. Fuel 2002, 81, 65−74. (9) Hashimoto, K.; Ikeda, M.; Arai, M.; Tamura, M. Cetane number improvement of diesel fuel by autoxidation. Energy Fuels 1996, 10, 1147−1149. (10) de Klerk, A. Properties of synthetic fuels from H-ZSM-5 oligomerization of Fischer−Tropsch type feed materials. Energy Fuels 2007, 21, 3084−3089. (11) Klerk, A. d. Fischer−Tropsch Refining; Wiley: New York, 2011; p 620. (12) Hoang, T. Q.; Zhu, X.; Danuthai, T.; Lobban, L. L.; Resasco, D. E.; Mallinson, R. G. Conversion of Glycerol to Alkyl-Aromatics over Zeolites. Energy Fuels 2010, 24, 3804−3809. (13) Krisnandi, Y. K.; Eckelt, R.; Schneider, M.; Martin, A.; Richter, M. Glycerol upgrading over zeolites by batch-reactor liquid-phase oligomerization: Heterogeneous versus homogeneous reaction. ChemSusChem 2008, 1, 835−844. (14) Rahmat, N.; Abdullah, A. Z.; Mohamed, A. R. Recent progress on innovative and potential technologies for glycerol transformation into fuel additives: A critical review. Renew. Sustainable Energy Rev. 2010, 14, 987−1000. (15) Alonso, D. M.; Bond, J. Q.; Serrano-Ruiz, J. C.; Dumesic, J. A. Production of liquid hydrocarbon transportation fuels by oligomerization of biomass-derived C9 alkenes. Green Chem. 2010, 12, 992−999. (16) Atadashi, I. M.; Aroua, M. K.; Abdul Aziz, A. R.; Sulaiman, N. M. N. The effects of catalysts in biodiesel production: A review. J. Ind. Eng. Chem. 2013, 19, 14−26. (17) Sanati, M.; C. H, Jaras, S.G., The oligomerization of alkanes by heterogeneous catalysts. In Catalysis; Spivey, J. J., Ed.; The Royal Society of Chemistry: London, 1999; Vol. 14, p 236. (18) Corma, A. Inorganic solid acids and their use in acid-catalyzed hydrocarbon reactions. Chem. Rev. 1995, 95, 559−614. 787

DOI: 10.1021/ie5041226 Ind. Eng. Chem. Res. 2015, 54, 781−789

Review

Industrial & Engineering Chemistry Research

(65) Muraza, O. In Nanosized zeolites with one-dimensional (1-D) pore systems and their applications in catalytic cracking, The 22nd Saudi-Japan Annual Symposium, Dhahran, November 25−26, 2012. (66) Verboekend, D.; Chabaneix, A. M.; Thomas, K.; Gilson, J.-P.; Perez-Ramirez, J. Mesoporous ZSM-22 zeolite obtained by desilication: Peculiarities associated with crystal morphology and aluminium distribution. CrystEngComm 2011, 13, 3408−3416. (67) Verboekend, D.; Thomas, K.; Milina, M.; Mitchell, S.; PerezRamirez, J.; Gilson, J.-P. Towards more efficient monodimensional zeolite catalysts: n-Alkane hydro-isomerisation on hierarchical ZSM22. Catal. Sci. Technol. 2011, 1, 1331−1335. (68) Muraza, O.; Bakare, I. A.; Tago, T.; Konno, H.; Taniguchi, T.; Al-Amer, A. M.; Yamani, Z. H.; Nakasaka, Y.; Masuda, T. Selective catalytic cracking of n-hexane to propylene over hierarchical MTT zeolite. Fuel 2014, 135, 105−111. (69) Zhu, J.; Meng, X.; Xiao, F. Mesoporous zeolites as efficient catalysts for oil refining and natural gas conversion. Front Chem. Sci. Eng. 2013, 7, 233−248. (70) Groen, J. C.; Peffer, L. A. A.; Moulijn, J. A.; Pérez, R.; amp, x; rez, J. On the introduction of intracrystalline mesoporosity in zeolites upon desilication in alkaline medium. Microporous Mesoporous Mater. 2004, 69, 29−34. (71) Gil, B.; Mokrzycki, Ł.; Sulikowski, B.; Olejniczak, Z.; Walas, S. Desilication of ZSM-5 and ZSM-12 zeolites: Impact on textural, acidic and catalytic properties. Catal. Today 2010, 152, 24−32. (72) Mokrzycki, L.; Sulikowski, B.; Antoine Gédéon, P. M.; Florence, B. Desilication of ZSM-12 and MCM-22 type zeolites and their performance in isomerization of α-pinene. Stud. Surf. Sci. Catal. B 2008, 174, 1231−1234. (73) Mintova, S.; Gilson, J.-P.; Valtchev, V. Advances in nanosized zeolites. Nanoscale 2013, 5, 6693−6703. (74) Hayasaka, K.; Liang, D.; Huybrechts, W.; De Waele, B. R.; Houthoofd, K. J.; Eloy, P.; Gaigneaux, E. M.; Van Tendeloo, G.; Thybaut, J. W.; Marin, G. B. Formation of ZSM-22 zeolite catalytic particles by fusion of elementary nanorods. Chem.Eur. J. 2007, 13, 10070−10077. (75) Muraza, O.; Bakare, I. A.; Tago, T.; Konno, H.; Adedigba, A.-l.; Al-Amer, A. M.; Yamani, Z. H.; Masuda, T. Controlled and rapid growth of MTT zeolite crystals with low-aspect-ratio in a microwave reactor. Chem. Eng. J. 2013, 226, 367−376. (76) Sanhoob, M.; Muraza, O.; Yamani, Z. H.; Al-Mutairi, E. M.; Tago, T.; Merzougui, B.; Masuda, T. Synthesis of ZSM-12 (MTW) with different Al-source: Towards understanding the effects of crystallization parameters. Microporous Mesoporous Mater. 2014, 194, 31−37. (77) Ahmed, M. H. M.; Muraza, O.; Al Amer, A. M. Effect of synthesis parameters and ion exchange on crystallinity and morphology of EU-1 zeolite. J. Alloys Compd. 2014, 617, 408−412. (78) Sanati, M.; Hornell, C.; Jaras, S. G., The oligomerization of alkenes by heterogeneous catalysts. In Catalysis; The Royal Society of Chemistry: London, 1999; Vol. 14, pp 236−288. (79) Galya, L. G.; Occelli, M. L.; Hsu, J. T.; Young, D. C. Propylene oligomerization over molecular sieves: Part II. 1H NMR and 13C NMR characterization of reaction products. J. Mol. Catal. 1985, 32, 391−403. (80) Occelli, M. L.; Hsu, J. T.; Galaya, L. G. Propylene oligomerization over molecular sieves: Part I. zeolite effects on reactivity and liquid product selectivities. J. Mol. Catal. 1985, 32, 377− 390. (81) Finiels, A.; Fajula, F.; Hulea, V. Nickel-based solid catalysts for ethylene oligomerizationA review. Catal. Sci. Technol. 2014, 4, 2412−2426. (82) Verboekend, D.; Pérez-Ramírez, J. Towards a sustainable manufacture of hierarchical zeolites. ChemSusChem 2014, 7, 651−651. (83) Bellussi, G.; Carati, A.; Clerici, M. G.; Esposito, A.; Poncelet, G.; J. P. G, P. A.; Delmon, B. Double substitution in silicalite by direct synthesis: A new route to crystalline porous bifunctional catalysts. Stud. Surf. Sci. Catal. B 1991, 63, 421−429.

(42) Bellussi, G.; Mizia, F.; Calemma, V.; Pollesel, P.; Millini, R. Oligomerization of olefins from light cracking naphtha over zeolitebased catalyst for the production of high quality diesel fuel. Microporous Mesoporous Mater. 2012, 164, 127−134. (43) McMahon, J. F.; Solomon, C. B. E. Polymerization of Olefins as a Refinery Process. Adv. Pet. Chem. 1963, VII. (44) Mlinar, A. N.; Zimmerman, P. M.; Celik, F. E.; Head-Gordon, M.; Bell, A. T. Effects of Brønsted-acid site proximity on the oligomerization of propene in H-MFI. J. Catal. 2012, 288, 65−73. (45) O’Connor, C. T.; Van Steen, E.; Dry, M. E., New catalytic applications of zeolites for petrochemicals. In Stud. Surf. Sci. Catal., H. Chon, S. I. W.; Park, S. E., Eds.; Elsevier: 1996; Vol. 102, pp 323−362. (46) van den Berg, J. P.; Wolthuizen, J. P.; van Hooff, J. H. C. Reaction of small olefins on zeolite H-ZSM-5. A thermogravimetric study at low and intermediate temperatures. J. Catal. 1983, 80, 139− 144. (47) Levenspiel, O. Experimental search for a simple rate equation to describe deactivating porous catalyst . J. Catal. 1972, 25, 265. (48) Fogler, H. S. Essentials of Chemical Reaction Engineering; Pearson Education: 2010. (49) Corma, A.; Martinez, C.; Doskocil, E. J.; Yaluris, G. Alkene oligomerization process. CA Patent CA2766415 A1, 2011; https://www. google.com/patents/CA2766415A1?cl=en. (50) Coelho, A.; Caeiro, G.; Lemos, M. A. N. D. A.; Lemos, F.; Ribeiro, F. R. 1-Butene oligomerization over ZSM-5 zeolite: Part 1 Effect of reaction conditions. Fuel 2013, 111, 449−460. (51) Martínez, C.; Doskocil, E. J.; Corma, A. Improved Theta-1 for light olefins oligomerization to diesel: Influence of textural and acidic properties. Top. Catal. 2014, 57, 668−682. (52) Martens, J. A.; Ravishankar, R.; Mishin, I. E.; Jacobs, P. A. Tailored alkene oligomerization with H-ZSM-57 zeolite. Angew. Chem. Int. Ed. 2000, 39, 4376−4379. (53) Flego, C.; Marchionna, M.; Perego, C. High quality diesel by olefin oligomerisation: New tailored catalysts. Stud. Surf. Sci. Catal. 2005, 158, 1271−1278. (54) Corma, A.; Martínez, C.; Doskocil, E. Designing MFI-based catalysts with improved catalyst life for oligomerization to high-quality liquid fuels. J. Catal. 2013, 300, 183−196. (55) Garwood, W. E.; Lee, W. Process for separating ethylene from light olefin mixtures while producing both gasoline and fuel oil. US Patent US4227992 A, 1980. (56) Garwood, W. E.; Lee, W., Process for the treatment of olefinic gasoline. EP Patent EP0020017 A1, 1980. (57) Petkovic, L. M.; Larsen, G. Oligomerization of n-butenes on the surface of a ferrierite catalyst: Real-time monitoring in a packed-bed reactor during isomerization to isobutene. Ind. Eng. Chem. Res. 1999, 38, 1822−1829. (58) Minachev, K. M.; Bondarenko, T. N.; Kon’drat’ev, D. A. Activity of different structural types of zeolites in oligomerization of C3−C4 olefins. Bull. Acad. Sci. Russ. 1987, 36, 1128−1132. (59) Tosin, G.; Verberckmoes, A.; Jaensch, H.; Mathys, G.; Mertens, M.; Saha, S.; Li, H.; Saxton, R. Olefin oligomerization process. WO Patent WO2013013886 A2, 2013. (60) Datema, K. P.; Nowak, A. K.; van Braam Houckgeest, J.; Wielers, A. F. H. In-situ 13C magic-angle-spinning NMR measurements of the conversion of ethene to aliphatic hydrocarbons over structurally different zeolites. Catal. Lett. 1991, 11, 267−276. (61) Flego, C.; Marchionna, M.; Perego, C.; J. Č ejka, N. Ž .; Nachtigall, P. High quality diesel by olefin oligomerisation: New tailored catalysts. Stud. Surf. Sci. Catal. B 2005, 158, 1271−1278. (62) Willhammar, T.; Sun, J.; Wan, W.; Oleynikov, P.; Zhang, D.; Zou, X.; Moliner, M.; Gonzalez, J.; Martínez, C.; Rey, F.; Corma, A. Structure and catalytic properties of the most complex intergrown zeolite ITQ-39 determined by electron crystallography. Nat. Chem. 2012, 4, 188−194. (63) Martens, L. R. M.; Verduijn, J. P. Zeolite and manufacturing process. WO Patent WO1995019945 A1, 1995. (64) Verrelst, W. H.; Martens, L. R. M. Oligomerization and catalysts therefor. WO1995022516 A1, 1995. 788

DOI: 10.1021/ie5041226 Ind. Eng. Chem. Res. 2015, 54, 781−789

Review

Industrial & Engineering Chemistry Research (84) Chen, C. S. H.; Bridger, R. F. Shape-selective oligomerization of alkenes to near-linear hydrocarbons by zeolite catalysis. J. Catal. 1996, 161, 687−693. (85) Dai, P.-S. E.; Sanderson, J. R.; Knifton, J. F.; Weitkamp, J.; K. H. P, H. G.; Hölderich, W. Catalytic Properties of Zeolites for Synlube Production by Olefin Oligomerization. Stud. Surf. Sci. Catal. B 1994, 84, 1701−1707. (86) Espinoza, R. L.; Snel, R.; Korf, C. J.; Nicolaides, C. P. Catalytic oligomerization of ethene over nickel-exchanged amorphous silicaaluminas; effect of the acid strength of the support. Appl. Catal 1987, 29, 295−303. (87) Haw, J. F. Zeolite acid strength and reaction mechanisms in catalysis. Phys. Chem. Chem. Phys. 2002, 4, 5431−5441. (88) Page, N. M.; Young, L. B. Olefin oligomerization with surface modified zeolite. US Patent US4855527 A, 1989. (89) Wilshier, K. G.; Smart, P.; Western, R.; Mole, T.; Behrsing, T. Oligomerization of propene over H-ZSM-5 zeolite. Appl. Catal 1987, 31, 339−359. (90) Na, J.; Liu, G.; Zhou, T.; Ding, G.; Hu, S.; Wang, L. Synthesis and catalytic performance of ZSM-5/MCM-41 zeolites with varying mesopore size by surfactant-directed recrystallization. Catal. Lett. 2013, 143, 267−275. (91) Zheng, S.; Heydenrych, H. R.; Jentys, A.; Lercher, J. A. Influence of surface modification on the acid site distribution of HZSM-5. J. Phys. Chem. B 2002, 106, 9552−9558. (92) Schwarz, S.; Kojima, M.; O’Connor, C. T. Effect of tetraalkylammonium, alcohol and amine templates on the synthesis and high pressure propene oligomerisation activity of ZSM-type zeolites. Appl. Catal 1991, 73, 313−330. (93) Heveling, J.; van der Beek, A.; de Pender, M. Oligomerization of ethene over nickel-exchanged zeolite y into a diesel-range product. Appl. Catal 1988, 42, 325−336. (94) Mlinar, A. N.; Baur, G. B.; Bong, G. G.; Getsoian, A. B.; Bell, A. T. Propene oligomerization over Ni-exchanged Na-X zeolites. J. Catal. 2012, 296, 156−164. (95) Bagshaw, S. A.; Jaenicke, S.; Khuan, C. G. Structure and properties of Al−MSU−S mesoporous catalysts: Structure modification with increasing Al content. Ind. Eng. Chem. Res. 2003, 42, 3989. (96) van Grieken, R.; Escola, J. M.; Moreno, J.; Rodríguez, R. Direct synthesis of mesoporous M-SBA-15 (M = Al, Fe, B, Cr) and application to 1-hexene oligomerization. Chem. Eng. J. 2009, 155, 442−450. (97) M?ller, K.; Bein, T. MesoporosityA new dimension for zeolites. Chem. Soc. Rev. 2013, 42, 3689−3707. (98) Ernst, S.; Weitkamp, J.; Martens, J. A.; Jacobs, P. A. Synthesis and shape-selective properties of ZSM-22. Appl. Catal 1989, 48, 137− 148. (99) Hayasaka, K.; Liang, D.; Huybrechts, W.; De Waele, B. R.; Houthoofd, K. J.; Eloy, P.; Gaigneaux, E. M.; van Tendeloo, G.; Thybaut, J. W.; Marin, G. B.; Denayer, J. F. M.; Baron, G. V.; Jacobs, P. A.; Kirschhock, C. E. A.; Martens, J. A. Formation of ZSM-22 zeolite catalytic particles by fusion of elementary nanorods. Chem.Eur. J. 2007, 13, 10070−10077. (100) Klocke, D. J.; Vartuli, J. C.; Kirker, G. W. Synthesis of zeolites ZSM-22 and ZSM-23. In EP Patent 0,220,893, 1987. (101) Valyocsik, E. W. Synthesis of zeolite ZSM-22. US Patent US4902406 A, 1990. (102) Valyocsik, E. W. Synthesis of zeolite ZSM-22 with a heterocyclic organic compound. US Patent US4481177 A, 1984. (103) Muraza, O.; Bakare, I. A.; Tago, T.; Konno, H.; Adedigba, A.-l.; Al-Amer, A. M.; Yamani, Z. H.; Masuda, T. Controlled and rapid growth of MTT zeolite crystals with low-aspect-ratio in a microwave reactor. Chem. Eng. J. 2013, 226, 367−376. (104) Li, G.; Hou, H.-m.; Lin, R.-s. Rapid synthesis of mordenite crystals by microwave heating. Solid State Sci. 2011, 13, 662−664. (105) Xue, H.; Huang, X.; Ditzel, E.; Zhan, E.; Ma, M.; Shen, W. Dimethyl ether carbonylation to methyl acetate over nanosized mordenites. Ind. Eng. Chem. Res. 2013, 52, 11510−11515.

(106) Martens, J. A.; Verboekend, D.; Thomas, K.; Vanbutsele, G.; Gilson, J. P.; Pérez-Ramírez, J. Hydroisomerization of Emerging Renewable Hydrocarbons using Hierarchical Pt/H-ZSM-22 Catalyst. ChemSusChem 2013, 6, 421−425. (107) Sá Couto, C.; Matias, P.; Santos, E. T.; Fernandes, A.; Graça, I.; Lopes, J. M.; Ribeiro, M. F. Towards a Deep Desilication/ Dealumination of NU-10 Zeolite: Shape-Selectivity Regulation. Eur. J. Inorg. Chem. 2012, 2012, 4190−4199. (108) Bonilla, A.; Baudouin, D.; Pérez-Ramírez, J. Desilication of ferrierite zeolite for porosity generation and improved effectiveness in polyethylene pyrolysis. J. Catal. 2009, 265, 170−180. (109) Baerlocher, C.; McCusker, L. B.; Olson, D. H. Atlas of Zeolite Framework Types; Elsevier: Amsterdam, 2007 (110) Washmon-Kriel, L.; Balkus, K. J., Jr. Preparation and characterization of oriented MAPO-39 membranes. Microporous Mesoporous Mater. 2000, 38, 107−121. (111) Kirchner, R. M.; Bennett, J. M. The structure of calcined ALPO4−41: A new framework topology containing one-dimensional 10-ring pores. Zeolites 1994, 14, 523−528. (112) Wang, X.; Dai, W.; Wu, G.; Li, L.; Guan, N.; Hunger, M. Verifying the dominant catalytic cycle of the methanol-to-hydrocarbon conversion over SAPO-41. Catal. Sci. Technol. 2014, 4, 688−696. (113) Bennett, J. M.; Richardson, J. W., Jr.; Pluth, J. J.; Smith, J. V. Aluminophosphate molecular sieve AlPO4-11: Partial refinement from powder data using a pulsed neutron source. Zeolites 1987, 7, 160−162. (114) Chen, Y.; Luo, X.; Chang, P.; Geng, S. Crystal morphology control of AlPO4−11 molecular sieves by microwave irradiation. Mater. Chem. Phys. 2009, 113, 899−904. (115) Ma, Y.; Li, N.; Ren, X.; Xiang, S.; Guan, N. Synthesis of SAPO41 from a new reproducible route using H3PO3 as the phosphorus source and its application in hydroisomerization of n-decane. J. Mol. Catal. A: Chem. 2006, 250, 9−14. (116) Li, H.-X.; Davis, M. E. Further studies on aluminophosphate molecular sieves. Part 2.-VPI-5 and related aluminophosphate materials. J. Chem. Soc, Faraday Trans 1993, 89, 957−964. (117) Jamil, A. K.; Muraza, O.; Sanhoob, M.; Tago, T.; Konno, H.; Nakasaka, Y.; Masuda, T., Controlling naphtha cracking using nanosized TON zeolite synthesized in the presence of polyoxyethylene surfactant. J. Anal Appl. Pyrolysis

789

DOI: 10.1021/ie5041226 Ind. Eng. Chem. Res. 2015, 54, 781−789